paper_id
stringlengths 9
16
| version
stringclasses 26
values | yymm
stringclasses 311
values | created
timestamp[s] | title
stringlengths 6
335
| secondary_subfield
sequencelengths 1
8
| abstract
stringlengths 25
3.93k
| primary_subfield
stringclasses 124
values | field
stringclasses 20
values | fulltext
stringlengths 0
2.84M
|
---|---|---|---|---|---|---|---|---|---|
1505.02968 | 1 | 1505 | 2015-05-12T11:43:05 | Certain actions of finitely generated groups on higher dimensional noncommutative tori | [
"math.OA"
] | We study certain actions of finitely generated abelian groups on higher dimensional noncommutative tori. Given a dimension $d$ and a finitely generated abelian group $G$, we apply a certain function to detect whether there is a simple noncommutative $d$-torus which admits a specific action by $G$. We also describe the condition of $G$ under which the associated crossed product is an AF algebra. | math.OA | math | CERTAIN ACTIONS OF FINITELY GENERATED GROUPS
ON HIGHER DIMENSIONAL NONCOMMUTATIVE TORI
ZHUOFENG HE
Abstract. We study certain actions of finitely generated abelian groups
on higher dimensional noncommutative tori. Given a dimension d and
a finitely generated abelian group G, we apply a certain function to de-
tect whether there is a simple noncommutative d-torus which admits a
specific action by G. We also describe the condition of G under which
the associated crossed product is an AF algebra.
.
A
O
h
t
a
m
[
1
v
8
6
9
2
0
.
5
0
5
1
:
v
i
X
r
a
Contents
1.
Introduction
2. Preliminaries
3. Realization of cyclic groups acting on noncomutative d-tori
4. Case d − W (n) = 1 and actions on odd-dimensional
noncommutative tori
5. Finitely generated abelian group actions
Acknowledgment
References
1
4
8
14
17
18
19
1. Introduction
Certain finite group actions on the rotation algebras have already been
systematically studied in [1] by S. Echterhoff, W. Luck, N. C. Phillips and
S. Walters. It is known that up to conjugacy, a finite subgroup F of SL2(Z)
satisfies F = Zk where k = 2, 3, 4 or 6. According to their paper, by realizing
certain generators for Zk, the action of Zk on a rotation algebra can be given
for each such k. We note that this action is the natural generalization of
the action of SL2(Z) on a torus T2 = R2/Z2.
If we require the rotation
algebra to be a simple one, i.e. for the irrational rotation algebra, denoted
by Aθ where θ is an irrational number, the crosed product Aθ ⋊ Zk is an AF-
algebra for k = 2, 3, 4 and 6. Moreover for each k, the K0-group K0(Aθ ⋊Zk)
∼= Aθ′ ⋊ Zk′ if and only if
has been calculated, which shows that Aθ ⋊ Zk
k = k′ and θ = θ′ mod Z, see [1, Theorem 0.1]. Also they show that higher
dimensional noncommutative tori admit flip actions of Z2, and for simple
ones the crossed products are AF algebras.
1
Certain actions of finitely generated groups on higher dimensional noncommutative tori
2
As stated in their paper, the proof of the above theorem breaks into the
following three steps.
(1) Computation of the K-theory of the crossed product.
(2) Proof that the crossed product satisfies the Universal Coefficient
Theorem.
(3) Proof that the action has the tracial Rohlin property.
Their argument and proof in step (2) and (3) are general enough so that
they are also effective for higher dimensional cases. Hence it is accessible if
one obtains a proper realization of some finite group, a proper generalization
of actions above to higher dimensional cases, and a way to compute the K-
theory of the crossed product. In [2], the authors manage to obtain results
in this way. They give a proper definition of such actions which generalize
the action above, which they call it by "canonical actions". Throughout this
paper we are interested in actions of this type. A detailed discussion of such
actions is in the Preliminaries. By realizing a cyclic group with the com-
panion matrix of a cyclotomic polynomial, the authors study such actions
of cyclic group on simple higher dimensional tori for certain dimensions. To
finish step (1), i.e. to compute the K-groups of the associated crossed prod-
uct, the authors first apply [1, Lemma 2.1] and [5, Theorem 4.1] to write the
crossed product in terms of another twisted group C ∗-algebra, then to com-
pute its K-theory, according to [1, Theorem 0.3], it is the same to compute
the K-theory of untwisted group C ∗-algebra, and finally by [4, Theorem
0.1] they manage to obtain desired the K-groups. The authors then show
that for certain dimension d and order n there is a simple noncommutative
d-torus on which Zn acts and the K1-group of the crossed product is not
trivial, see [2, Theorem 3.6].
We continue their work and study such actions of finitely generated abelian
groups on simple higher dimensional noncommutative tori. By combining
the tensor structure of higher dimensional noncommutatvie tori and the re-
alization method in [2] for cyclic groups and actions, we manage to realize
all of the possible actions under cyclic groups. Then for a given dimension d
by the structure of finite cyclic group of GLd(Zd), we apply a certain func-
tion to tell when we can realize the action of the cyclic group on a simple
noncommutative d-torus. Moreover we describe the condition under which
the crossed product is an AF algebra. The function W (n) is defined later.
Hereby we remark that through this paper we consider noncommutative
d-tori, where we usually suppose d > 1, since there is no 1-dimensional non-
commutative tori. However some of our results are actually compatible with
the case when we allow d = 1.
Theorem 1.1 (Theorem 3.1). Given a dimension d and an order n > 1,
there is an action of Zn on a simple noncommutative d-tori AΘ if either
d − W (n) > 1 or d − W (n) = 0. The crossed product AΘ ⋊ Zn is an AT
algebra, and it is an AF algebra if and only if d − W (n) = 0, and n either
admits a form of 2m where m = 3j5ipel
l , or admits a form of 2k3j5i where
Certain actions of finitely generated groups on higher dimensional noncommutative tori
3
k 6= 1, j ≤ 2, i ≤ 1 and el are all nonnegative integers, and pl > 5 is a
prime number.
Then we generalize the function W for finite abelian groups, and realize
actions with a similar argument. We obtain the following.
Theorem 1.2 (Theorem 3.6). Given a dimension d and an abelian finite
subgroup G ≤ GLd(Z) with d − W (G) > 1 or d − W (G) = 0, there is an
action of G on a simple noncommutative d-torus AΘ. The crossed product
AΘ ⋊G is an AT algebra, and it is an AF algebra if and only if d−W (G) = 0
and, if we write G as G ∼=Qt
j=1 Znj such that
W (G) =
W (Znl),
t
Xl=1
each nl either admits a form of 2m where m = 3j5iper
r , or admits a form of
2k3j5i where k 6= 1, j ≤ 2, i ≤ 1 and er are all nonnegative integers, and
pr > 5 is a prime number.
Especially we have Corollary 3.7 which states that the only possibility
of a finite abelian group F acting on a irrational rotation algebra is when
F = Zk where k = 2, 3, 4 and 6. In Section 3 we prove above theorems and
discuss related results.
Then we find out that the function W contains more information. We
have the following.
Theorem 1.3 (Theorem 4.3). Given a dimension d and an order n, if
d − W (n) = 1, then there is no simple noncommutative torus on which Zn
acts in the above way.
In Section 4 we give the proof of this theorem.
With enough knowledge of the function W when considering cyclic group
actions, we move to the topic of finitely generated abelian group acting
on simple higher dimensional noncommutative tori. The action of torsion
free part of the finitely generated group is defined to be adding dimensions
concerning the crossed product. Similarly we study the condition by which
the crossed product is an AF algebra.
Theorem 1.4 (Theorem 5.1). For a given dimension d and a finitely gen-
erated abelian group G ∼= (cid:16)Qs
i=1 Zp
ei
i (cid:17) × Zr, if d − W (n) = 1 and r > 0,
or d − W (n) 6= 1 then there is a noncommutative d-torus AΘ admitting an
action of G, denoted by α : G → Aut(AΘ). We require AΘ is pseudo-simple
in the first case and simple in the second case. Then the crossed product
AΘ ⋊α G is a simple AT algebra, and it is an AF algebra if and only if
r = 0, d − W (G) = 0 and G satisfies the last condition in Theorem 1.2.
A detailed discussion is in Section 5.
Certain actions of finitely generated groups on higher dimensional noncommutative tori
4
2. Preliminaries
For a second countable locally compact Hausdorff topological group G
with its modular function ∆G : G → (0, +∞) and a Borel 2-cocycle ω ∈
full) twisted group C ∗-
Z(G, T), we define the associated reduced (resp.
r (G, ω) (resp. C ∗(G, ω)) in the following way. Regard L1(G) as
algebras, C ∗
a vector space, and equip it with twisted convolution given by
f ∗ω g(t) =ZG
and the involution given by
f (s)g(s−1t)ω(s, s−1t)ds
f ∗(t) = ∆G(t−1)ω(t, t−1)f (t−1).
Then L1(G, ω) := (L1(G), ∗ω , ∗) becomes a ∗-algebra which is named the
twisted convolution algebra. As similar to the definition of group C ∗-algebras,
to consider nondegenerate representations of L1(G, ω), we turn to a twisted
analogue of unitary representations of G. More precisely, an ω-representation
of G on a Hilbert space H is a Borel map V : G → U (H) such that
V (t)V (s) = ω(t, s)V (ts)
where U (H) stands for the unitary group of H endowed with the strong
operator topology. For example we may define Lω : G → U (L2(G)), the
regular ω-representation of G, by
Lω(s)ξ(t) = ω(s, s−1t)ξ(s−1t)
for all ξ ∈ L2(G). For arbitrary ω-representation V : G → U (H), define
π : L1(G, ω) → B(H) via
π(f ) =ZG
f (s)V (s)ds
for f ∈ L1(G, ω). Then π is a contractive ∗-homomorphism. As in the
definition of a group C ∗-algebra, with abuse of notations we may also denote
π by V : L1(G, ω) → B(H). Hence we can define the reduced twisted group
C ∗-algebra as
C ∗
r (G, ω) = Lω(L1(G, ω)).
Moreover every nondegenerate representation of L1(G, ω) is induced by an
ω-representation of G. We can see this by taking a norm one approximate
identity in L1(G, ω). Thus we obtain the universal representation πu of
L1(G, ω), and define the full twisted group C ∗-algebra as
C ∗(G, ω) = πu(L1(G, ω)).
We remark that C ∗(G, ω) has the universal property and the ∗-homomorphism
Lω : C ∗(G, ω) → C ∗
C ∗-algebras coincide if G is amenable.
r (G, ω) is surjective. The reduced and full twisted group
We realize higher dimensional noncommutative tori with twisted group
C ∗-algebras. For a given dimension d, take G = Zd. Throughout this paper
we denote by Td(R) the set of all d × d skew symmetric real matrices. Then
Certain actions of finitely generated groups on higher dimensional noncommutative tori
5
for Θ = (θij) ∈ Td(R) define the induced 2-cocycle ωΘ : Zd × Zd → T by
formula
ωΘ(x, y) = exp(πihΘx, yi)
for x, y ∈ Zd. Then AΘ := C ∗(Zd, ωΘ) is called a noncommutative d-torus.
Since Zd is discrete and amenable, if we take the standard basis of it, say
ei for i = 1, . . . , d, and denote the regular ωΘ-representation by lΘ : Zd →
U (ℓ2(Zd)), then we have
AΘ := C ∗(Zd, ωΘ) = C ∗{lΘ(ei) i = 1, . . . , d}
where each lΘ(ei) is a unitary and they satisfy the commuting relations
lΘ(ei)lΘ(ej) = ωΘ(ei, ej)2lΘ(ej )lΘ(ei) = exp(2πiθji)lΘ(ej )lΘ(ei).
This is clearly a generalization of the rotation algebra to the d-dimensional
case. We refer to the rotation algebra which is the case when d = 2.
One may take the zero d × d matrix as Θ and then get AΘ = C(Td),
however this is not so much in our interests since it is not simple. Thus we
have the following.
Definition 2.1. The matrix Θ ∈ Td(R) is said to be nondegerate if when-
ever x ∈ Zd satisfies exp(2πihΘx, yi) = 1 for all y ∈ Zd, we have x = 0.
Theorem 2.2 ([6, Theorem 1.9]). Let Θ ∈ Td(R). Then the noncommuta-
tive d-torus AΘ is simple if and only if so is Θ.
Proposition 2.3. For Θi ∈ Tdi(R), i = 1, . . . , m, denote Θ := Lm
Td(R) where d := Pm
i=1 Θi ∈
i=1 di. Then Θ is nondegenerate if and only if each Θi
is nondegenerate.
Proof. By induction we only show the the case when m = 2. If Θ1 and Θ2 are
both nondegenerate, for x = (x1, x2) ∈ Zd satisfying exp(2πihΘx, yi) = 1
for all y ∈ Zd, by taking y1 ∈ Zd1 × {0} and y2 ∈ {0} × Zd2 we obtain
exp(2πihΘixi, yii) = 1 for all yi ∈ Zdi, i = 1, 2. Then by nondegeneracy of
Θ1 and Θ2, we obtain x = (x1, x2) = 0, in other words Θ is nondegenerate.
Conversely if Θ is nondegenerate, then for x1 ∈ Zd1 satisfying
exp(2πihΘ1x1, y1i) = 1
for all y1 ∈ Zd1, then we have for x = (x1, 0) ∈ Zd,
exp(2πihΘx, yi) = exp(2πihΘ1x1, y1i) = 1
for all y ∈ Zd. Thus x1 = 0, and similarly x2 = 0. Hence Θ1 and Θ2 are
both nondegenerate.
(cid:3)
Then we discuss actions on noncommutatitive tori. We denote by AΘ a
noncommutative d-torus as above, which is not necessarily simple. Through
the regular representation lΘ : ℓ1(Zd, ωΘ) → B(ℓ2(Zd)), we regard AΘ as a
C ∗-subalgebra in B(ℓ2(Zd)).
For a matrix A ∈ GLd(Z), we have a unitary uA in B(ℓ2(Zd)) given by
uAξ(x) = ξ(A−1x)
Certain actions of finitely generated groups on higher dimensional noncommutative tori
6
for ξ ∈ ℓ2(Zd) and x ∈ Zd. Thus we have Ad uA ∈ Aut(B(ℓ2(Zd))). Then
we have an action denoted by
Ad u• : GLd(Z) → Aut(B(ℓ2(Zd))).
Then for AΘ = C ∗{lΘ(ei) i = 1, . . . , d} = span{lΘ(x) x ∈ Zd}, we have
the fomula
Ad(uA)(lΘ(x))ξ(y) = l(A−1)tΘA−1(Ax)ξ(y)
for ξ ∈ ℓ2(Zd) and y ∈ Zd. Thus for A ∈ GLd(Z) such that Θ = (A−1)tΘA−1
we have Ad(uA) ∈ Aut(AΘ) by restriction. Then denote
GΘ := {A ∈ GLd(Z) Θ = AtΘA},
we obtain an action
Ad u• : GΘ → Aut(AΘ).
This is an action on AΘ which generalizes the action of SL2(Z) on the
rotation algebra Aθ discussed in [1]. Particularly, when the dimension d = 2,
we have GΘ = SL2(Z). We are especially interested in such actions of some
subgroup G of GΘ on AΘ.
Then we roughly state classification results for our cases, and refer to
[1], [2],[6] and [7] for details. A simple AΘ is known to be with a unique
tracial state and has tracial rank zero. For an action of a finite group on it,
which has the tracial Rokhlin property, say α : G → Aut(AΘ), the resulting
crossed product AΘ ⋊α G becomes also simple, with a unique tracial state
and with tracial rank zero. Particularly, for a finite subgroup G ≤ GΘ, the
action described in last paragraph of G on simple AΘ is known to have the
tracial Rokhlin property. Moreover the resulting crossed product AΘ ⋊ G
satisfies the Universal Coefficient Theorem. Thus the crossed products of
such actions becomes classifiable. We also need the following theorem.
Theorem 2.4 ([6, Preposition 3.7]). Let A be a simple infinite dimensional
separable unital nuclear C ∗-algebra with tracial rank zero and which satisfies
the Universal Coefficient Theorem. Then A is a simple AH algebra with real
rank zero and no dimension growth. If K∗(A) is torsion free, A is an AT
algebra. If, in addition, K1(A) = 0, then A is an AF algebra.
Then we introduce results of J. A. Jeong and J. H. Lee in [2]. We use
their realization method as a building block to obtain more realizations and
actions.
For a give dimension d and an order n with d = φ(n), where φ is the
Euler's totient function. The nth cyclotomic polynomial Φn(x) is defined by
Φn(x) = Y0<k≤n
gcd(k,n)=1
(x − exp(2πi
k
n
)).
Φn(x) is known to be a monic polynomial of degree d = φ(n), and with
integer coefficients. Thus set
Certain actions of finitely generated groups on higher dimensional noncommutative tori
7
0 0
1 0
0 1
...
...
0 0
0 0
0
0
0
. . .
0
0
· · ·
· · ·
· · ·
...
. . .
· · ·
0 −a0
0 −a1
0 −a2
...
...
0 −ad−2
1 −ad−1
,
Cn =
which is the companion matrix of Φn(x). Then we have Cn is in GLd(Z)
and of order n.
Theorem 2.5 ([2, Theorem 4.2]). Let n ≥ 3 and d := φ(n). Then there exist
simple d-dimensional tori on which the group Zn = hCni acts canonically.
We remark in this theorem the "canonical action" is the action we dis-
cussed above and what we are interested in. To compute the K-theory for
associated crossed product of such action, we need the following theorem.
Theorem 2.6 ([4, Theorem 0.1]). Let n, d ∈ N. Consider the extension of
groups
1 → Zd → Zd ⋊α Zn → Zn → 1
such that conjugation action α of Zn on Zd is free outside of the origin
0 ∈ Zd. Then K0(C ∗(Zd ⋊α Zn)) ∼= Zs0 for some s0 ∈ Z and
K1(C ∗(Zd ⋊α Zn)) ∼= Zs1,
where s1 =Pl≥0 rkZ((V2l+1 Zd)Zn ). If n is even, s1 = 0. If n > 2 is prime
and d = n − 1, then s1 = 2n−1−(n−1)2
2n
.
The authors in [2] apply this method and obtain that for certain dimension
d and order n, the resulting crossed product is not an AF algebra. The
theorem is stated as the following.
Theorem 2.7. [2, Theorem 3.6] Let n ≥ 7 be an odd integer and d := φ(n).
Consider the extension of groups 1 → Zd → Zd ⋊α Zn → Zn → 1 with
Zn = hCni. If 2d ≥ n + 5, then
K1(C ∗(Zd ⋊α Zn)) 6= 0.
We denote by N the smallest subcategory of the category of separable
C ∗-algebras, which contains separable Type I algebras and is closed under
taking ideals, quotients, extensions, inductive limits, stable isomorphisms
and crossed products by Z and R, then the following theorem holds.
Theorem 2.8 ([8, Kunneth theorem]). Let A and B be C ∗-algebras with A
in N. Then there is a natural short exact sequence
0 → K∗(A) ⊗ K∗(B) α−→ K∗(A ⊗ B)
β
−→ Tor(K∗(A), K∗(B)) → 0.
The sequence is Z2-graded with deg α = 1, deg β = 1, where Kp ⊗ Kq and
Tor(Kp ⊗ Kq) are given degree p + q for p, q ∈ Z2.
Certain actions of finitely generated groups on higher dimensional noncommutative tori
8
3. Realization of cyclic groups acting on noncomutative d-tori
Given skew symmetric real matrices Θ1 and Θ2, suppose that Θ = Θ1 ⊕
Θ2, then naturally we have
AΘ ∼= AΘ1 ⊗ AΘ2.
Then for any finite group actions described in the Preliminaries of G1 ≤ GΘ1
on AΘ1, and G2 ≤ GΘ2 on AΘ2, we could naturally get a finite group action
of G1 × G2 on AΘ, one can easily verify that G = G1 × G2 ≤ GΘ and the
resulting action is the restriction of Ad u• : GLd(Z) → Aut(B(ℓ2(Zd))),
where we denote by d the degree of Θ. Moreover we have
AΘ ⋊ G ∼= (AΘ1 ⋊ G1) ⊗ (AΘ2 ⋊ G2)
by an isomorphism defined by
(a ⊗ b, (g, h)) 7→ (a, g) ⊗ (b, h).
For a given dimension d, consider GLd(Z). By [3, Theroem 2.7], GLd(Z)
if and only W (n) ≤ d, where p1 <
i=1 pei
i
has an element of order n = Qt
· · · < pt are all prime numbers and
W (n) :=(Pt
Pt
i=1(pi − 1)pei−1
i=1(pi − 1)pei−1
i
i
− 1,
,
pe1
1 = 2;
otherwise,
thus we are able to obtain finitely many candidates for n, which is the
possible order of a realizable cyclic group in GLd(Z). For those n's with
φ(n) = d, where φ is the Euler's totient function, by [2, Theorem 4.2] we may
find a noncommutative d-torus AΘ which is simple, namely figure out the
form of the nondegenerate Θ, and realize an action of Zn on AΘ. Moreover
we could compute K∗(AΘ ⋊ Zn) and then know whether AΘ ⋊ Zn is AF
or not. However this way does not work for those n's with φ(n) 6= d,
paticularly when d is odd. We manage a way of realizing cyclic groups for
more candidates n to find actions in the following.
If pe1
1
i=1 pei
By Kuzmanovich's proof in [3], suppose n = Qt
such that W (n) ≤ d.
could construct a di × di matrix Ai of order pei
i ) = (pi − 1)pei−1
We note that φ(pei
W (n) × W (n) matrix with W (n) ≤ d.
i with p1 < · · · < pt
6= 2, by applying the companion matrix we
i where di = (pi − 1)pei−1
.
i=1 Ai, which is a
. Then set B := Lt
Thus by [2, Theorem 4.2], for each i, we could find nondegenerate Θi
acts on AΘi. If d − W (n) > 1, set A :=
in Tdi(R) such that hAii = Zp
B ⊕ Id−W (n), which is a d × d matrix of order n. It is possible to find a
nondegenerate Θ0. Then set
ei
i
i
i
Θ :=(cid:16)
t
Mi=1
Θi(cid:17) ⊕ Θ0,
Certain actions of finitely generated groups on higher dimensional noncommutative tori
9
which is nondegenerate, thus we have an action of hAi = Zn on AΘ. Moreover
by our former discussion one may check that
AΘ ⋊ Zn ∼=(cid:16)
(AΘi ⋊ Zp
ei
i
)(cid:17) ⊗ AΘ0.
t
Oi=1
By the method above we may realize more finite cyclic group actions on
higher dimensional noncommutative tori. To compute their K-groups, we
apply the Kunneth theorem in [8], stated as Theorem 2.8.
To verify the conditions for our problem, we have the following confirma-
tion. Firstly by [2, Proposition 3.4] each AΘi ⋊ Zp
is an AT algebra, and
so is AΘ0. Since C(T) ⊗ F is separable and of Type I, where F stands for
a finite dimensional C ∗-algebra, then every AΘi ⋊ Zp
and AΘ0 are in N.
Secondly by [2, Remark 3.2] and [4, Theorem 0.1] we know that for each i,
ei
i
ei
i
K∗(AΘi ⋊ Zp
ei
i
) ∼= K∗(C ∗(Zdi ⋊ Zp
ei
i
, ωΘi) ∼= K∗(C ∗(Zdi ⋊ Zp
)),
ei
i
which is torsion free.
Then applying Kunneth theorem, we obtain
K∗(AΘ ⋊ Zn) =(cid:16)
K∗(AΘi ⋊ Zp
ei
i
)(cid:17) ⊗ K∗(AΘ0).
t
Oi=1
By [2, Propersition 3.1] AΘ ⋊ Zn satisfies the Universal Coefficient Theorem.
As we have already stated before, AΘ ⋊Zn is then classifiable. We obtain the
K∗(AΘ ⋊ Zn) is torsion free, thus under this realization the crossed product
AΘ ⋊ Zn is an AT algebra. However, We note that d − W (n) > 1 and
K0(AΘ0) ∼= K1(AΘ0) ∼= Z2d−W (n)−1.
Thus in this case K1(AΘ ⋊ Zn) 6= 0, and AΘ ⋊ Zn is not an AF-algebra.
If d − W (n) = 0, the situation is essentially the same to our discussion
i=1 Ai, which is a d × d matrix of order
n. Then set
before. Instead we set A := B = Lt
Θ :=(cid:16)
t
Mi=1
Θi(cid:17),
which is nondegenerate, thus we have an action of hAi = Zn on AΘ. Then
we obtain
AΘ ⋊ Zn ∼=
(AΘi ⋊ Zp
)
ei
i
and again by the Kunneth theorem,
t
Oi=1
t
Oi=1
K∗(AΘ ⋊ Zn) =
K∗(AΘi ⋊ Zp
).
ei
i
Thus K1(AΘ ⋊ Zn) = 0 if and only if K1(AΘi ⋊ Zp
) = 0 for each i.
ei
i
Certain actions of finitely generated groups on higher dimensional noncommutative tori 10
Then our attention is driven to each AΘi ⋊ Zp
ei
i
. Since in the case when
) 6= 0 if
(pi − 2) ≥ 5. Thus the only candidates of K1(AΘi ⋊ Zp
pi is an odd prime, by [2, Theorem 3.6], we have K1(AΘi ⋊ Zp
pei−1
) = 0 are
i
pei
i = 3, 32, 5 or 2k where k > 1. To compute K1-groups, we apply [4,
Theorem 0.1], stated as Theorem 2.6.
ei
i
ei
i
Hence the corresponding K1-groups vanish when pei
k > 1. In the following we only check the case when pei
above enables us to only compute s1 where
i = 3, 5 and 2k where
i = 32. The theorem
s1 = rkZ((^1
Z6)Z9) + rkZ((^3
Z6)Z9 ) + rkZ((^5
We obtain rkZ((V1 Z6)Z9) = 0 for the action is free outside the origin 0 ∈ Z6.
According to our realization we have Z9 is the cyclic group generalized by
A in GL6(Z), where
Z6)Z9).
A =
0 0 0 0 0 −a0
1 0 0 0 0 −a1
0 1 0 0 0 −a2
0 0 1 0 0 −a3
0 0 0 1 0 −a4
0 0 0 0 1 −a5
Φ9(x) = Y0<k<9
gcd(k,9)=1
(x − ζ k) =
aixi, ζ = exp(2πi
1
9
).
5
Xi=0
and
Write
A = Q−1diag(ζ, ζ 2, ζ 4, ζ 5, ζ 7, ζ 8)Q,
where Q ∈ GL6(C). Notice that x a fixed point of A acting on V3 Z if and
only if Qx is a fixed point of diag(ζ, ζ 2, ζ 4, ζ 5, ζ 7, ζ 8) acting onV3 QZ. Then
for 1 ≤ i < j < k ≤ 6, if
diag(ζ, ζ 2, ζ 4, ζ 5, ζ 7, ζ 8)ei ∧ ej ∧ ek = ei ∧ ej ∧ ek,
it is the same to stating that there is a way of pick three number in 1, 2, 4, 5, 7, 8
such that the sum of them can be divided by 9, which is impossible. Thus
we obtain
Hence when pei
which Z9 acts in the way we are interested in, and
i = 32, we obtain a simple noncommutative 6-torus AΘ on
in other words AΘ ⋊ Z9 is an AF algebra.
K1(AΘ ⋊ Z9) = 0,
Similarly we also obtain
Z6)Z9 ) = 0.
Z6)Z9 ) = 0.
rkZ((^3
rkZ((^5
Certain actions of finitely generated groups on higher dimensional noncommutative tori 11
We leave the case when d − W (n) = 1 to next section. To complete the
proof, we discuss the case when given the dimension d and the order n such
i=1 pei
that W (n) ≤ d, and with pe1
i .
By proof in [3] notice that W ( n
2 , by applying
the companion matrix we could construct a di × di matrix Ai of order pei
i
where di = (pi − 1)pei−1
i=2 Ai, which is
a W (n) × W (n) matrix with W (n) ≤ d, and of order n
2 .
1 = 2 in the prime factorization of n =Qt
for i = 2, . . . , t. Then set B := Lt
2 ) = W (n) ≤ d. Then for n
i
Again we could find nondegenerate Θi in Tdi(R) such that hAii acts on AΘi
2 ) > 1,
2 . It is possible to
in the way we described above for i = 2, . . . , t. If d − W (n) = d − W ( n
set A := B ⊕ Id−W (n), which is a d × d matrix of order n
find a nondegenerate Θ0 since d − W (n) > 1. Then set
Θ :=(cid:16)
t
Mi=1
Θi(cid:17) ⊕ Θ0,
which is nondegenerate. Since
(−A)tΘ(−A) = AtΘA = Θ,
we have an action of h−Ai = Zn on AΘ. However, in this case we can only
write Zn as
Zn ∼=(cid:16) Y2≤i≤t
i6=j
Zp
ei
i (cid:17) × Z
2p
ej
j
where 2 ≤ j ≤ t. Thus instead of A, we consider A(j) as a generator of Zn
where A(j) is defined by
A(j) =
0
A1
0 A2
...
...
0
0
...
...
0
0
0
0
0
· · ·
0
· · ·
...
. . .
· · · −Aj
...
...
0
· · ·
· · ·
0
0
0
0
0
...
...
0
0
...
. . .
0 At
0
0
0
0
...
0
...
0
Id−W (n)
where 1 ≤ j ≤ t. Then we consider actions of Zn = hA(j)i on AΘ.
Similarly we have
AΘ ⋊ Zn ∼=(cid:16) O2≤i≤t
i6=j
(AΘi ⋊ Zp
ei
i
)(cid:17) ⊗ (AΘj ⋊ Z
) ⊗ AΘ0,
2p
ej
j
where 2 ≤ j ≤ t. By the Kunneth therorem K∗(AΘ ⋊ Zn) is the tensor of K∗
of each factor. Again since the part of K∗(AΘ0), K1(AΘ ⋊ Zn) 6= 0, which
is equivalently to say that AΘ ⋊ Zn is not an AF algebra.
Certain actions of finitely generated groups on higher dimensional noncommutative tori 12
It is similar to above when d−W (n) = d−W ( n
2 ) = 0. However we remark
that
K∗(AΘ ⋊ Zn) ∼=(cid:16) O2≤i≤t
i6=j
K∗(AΘi ⋊ Zp
ei
i
)(cid:17) ⊗ K∗(AΘj ⋊ Z
)
2p
ej
j
where 2 ≤ j ≤ t. Thus by [4, Theorem 0.1], we have K1(AΘj ⋊ Z2p
since 2pej
is even. Hence in this case, AΘ ⋊Zn is an AF-algebra exactly when
j
2 admits a form of 3j5ipel
n
l where j, i and el are all nonnegetive integers,
j ≤ 2, i ≤ 1 and pl > 5 is a prime number. As a summarize we have the
following theorem.
) = 0
ei
j
Theorem 3.1. Given a dimension d and an order n > 1, there is an action
of Zn on a simple noncommutative d-tori AΘ if either d − W (n) > 1 or
d − W (n) = 0. The crossed product AΘ ⋊ Zn is an AT algebra, and it is an
AF algebra if and only if d − W (n) = 0, and n either admits a form of 2m
where m = 3j5ipel
l , or admits a form of 2k3j5i where k 6= 1, j ≤ 2, i ≤ 1
and el are all nonnegative integers, and pl > 5 is a prime number.
Corollary 3.2. Given a dimension d which is even and order n, there is
an action of Zn on simple noncommutative d-tori if and only if W (n) ≤ d.
Proof. Since W (n) is always even when n ≥ 3.
(cid:3)
Note that we have the following example which is already mentioned in
[1].
Example 3.3. Consider an action of Zn on a simple 2-dimensional non-
commutative torus , say AΘ. Suppose that the action is of the above type.
By Corollary 3.2, it is equivalent to say that d − W (n) = 0 or d − W (n) = 2
where d = 2. Then by knowledge of W (n), we obtain n = 2, 3, 4 or 6.
Moreover by Theorem 3.1, each crossed product AΘ ⋊ Zn is an AF algebra.
Then generally for a finite abelian group G, we may generalize our method
of construct such actions of G on noncommutative tori. By the structure
theorem for finitely generated abelian groups we write G as
s
G ∼=
Yi=1
we cannot simply define W (G) := Ps
where p1 ≤ · · · ≤ ps are primes and ei ≤ ei+1 whenever pi = pi+1. However
i ). Since if so, suppose the
case when G = Z2 × Z2, in which we have W (G) = 0, but we can only
realize it in GL4(Z) as hA, Bi where
i=1 W (pei
Zp
ei
i
A =
0
−1
0 0
0 −1 0 0
0
1 0
0 1
0
0
0
and B =
0
1 0
0 1
0
0 0 −1
0 0
0
0
0
0 −1
.
Certain actions of finitely generated groups on higher dimensional noncommutative tori 13
1 = 2, and there is no other pei
i such that pei
i
as a normal subgroup, where Z2p
i=1 Zp
This is because if for a finite abelian group G ∼=Qs
, if Z2 is a normal
group of it, i.e. pe1
6= 2, then we
cannot write Z2 into Z2p
is a normal
subgroup of G. Thus we cannot realize Z2 as the signature part of the matrix
which need not a cost of a dimension, and since the determinant of realized
matrix is 1, the degree should be even. Thus instead we have to cost two of
dimensions to set a diagonal part as example above. Hence in this case we
should modify our test funtion to taking value 2 of Z2.
ei
i
ei
i
ei
i
Definition 3.4. For a cyclic group Zn, define the function W as follows.
W (Zn) :=(W (n), n 6= 2;
n = 2.
2,
Definition 3.5. For a finitely generated abelian group G, by the structure
theorem for finitely generated abelian groups, we have
G ∼=(cid:16)
s
Yi=1
Zp
ei
i (cid:17) × Zr,
where p1 ≤ · · · ≤ ps are primes and ei ≤ ei+1 whenever pi = pi+1. Let Gtor
denote the torsion part Gtor.
Define the function W of G as
t
W (G) := minn
W (Znl) Gtor ∼=
Xl=1
l
Yj=1
Znjo.
Theorem 3.6. For a given dimension d and an abelian finite subgroup
G ≤ GLd(Z) with d − W (G) > 1 or d − W (G) = 0, there is an action of G
on a simple noncommutative d-torus AΘ. The crossed product AΘ ⋊ G is
an AT algebra, and it is an AF algebra if and only if d − W (G) = 0 and, if
we write G as G ∼=Qt
j=1 Znj such that
t
W (G) =
W (Znl),
Xl=1
Xl=1
each nl either admits a form of 2m where m = 3j5iper
r , or admits a form of
2k3j5i where k 6= 1, j ≤ 2, i ≤ 1 and er are all nonnegative integers, and
pr > 5 is a prime number.
Proof. By definition we can write G as G ∼=Qt
t
W (G) =
W (Znl).
j=1 Znl such that
Thus we have d − W (Znl) > 1 or d − W (Znl) = 0 for all l. Then for each j
by Theorem 3.1 we obtain a simple noncommutative torus AΘl on which Znl
acts in the above way. We emphasis that if nl = 2, we obtain an irrational
rotation algebra. Then by the tensor product argument similar in section
Certain actions of finitely generated groups on higher dimensional noncommutative tori 14
3 we obtain a simple noncommutative d-torus AΘ on which G acts in the
above way.
The rest of the proof is similar to the one in section 3 and from Theorem
(cid:3)
3.1.
Corollary 3.7. The only actions of finite abelian group on irrational rota-
tion algebras of the above type are by Zk where k = 2, 3, 4 and 6.
Proof. Note that in the current case d = 2. suppose G is a finite abelian
group acting on the irrational rotation algebra Aθ in the above way. It is a
conflict if W (G) > 2 then by our statement above. Thus W (G) = 2. Then
by the definition of the function W we have G = Zk where k = 2, 3, 4 and
6.
(cid:3)
4. Case d − W (n) = 1 and actions on odd-dimensional
noncommutative tori
d − W (n) = 1 for a given dimension d and order n = Qt
As a remaining problem of the last section, we need to study the case
i with primes
p1 < · · · < pt. We start our discussion again from Kuzmanovich's result
which shows that for any A ∈ GLd(Z) of order n, there is Q ∈ GLd(Q) ⊆
GLd(R) such that
i=1 pei
0
0
...
0
· · ·
0
· · ·
...
. . .
· · · At 0
· · ·
1
0
A1
0
0 A2
...
...
0
0
0
0
where Ai is the companion matrix for pei
i .
Λ := QAQ−1 =
In general for a subgroup G ≤ GLd(Z) and an element Q ∈ GLd(R)
with QGQ−1 ≤ GLd(Z), it is routine to check that G ≤ GΘ1 if and only if
QGQ−1 ≤ GΘ2 where Θ2 := (Q−1)tΘ1Q−1.
Consider a category G defined as following: an object is a pair (G, ω)
where G is a discete group and ω is a 2-cocycle in Z(G, T). For objects
(G1, ω1) and (G2, ω2),
HomG((G1, ω1), (G2, ω2)) := {ϕ : G1 → G2 ω2 ◦ ϕ = ω1}.
If we denote by C the category of C ∗-algebras and define
C ∗(·, ·)(G, ω) := C ∗(G, ω),
C ∗(·, ·)ϕ : ℓ1(G2, ω2) → ℓ1(G1, ω1), f 7→ f ◦ ϕ.
By the universal property of a full twisted C ∗-algebra, one may check that
C ∗(·, ·) is a functor from G to C.
Thus for arbitrary Θ1 ∈ Td(R) and Q ∈ GLd(R), we have Θ2 := QtΘ1Q ∈
Td(R). By taking different generators we can obtain an isomorphism
Q−1 : Zd → Zd,
Certain actions of finitely generated groups on higher dimensional noncommutative tori 15
moreover
ωΘ2 ◦ Q−1(x, y) = exp(πihΘ2Q−1x, Q−1yi) = exp(πihΘ1x, yi) = ωΘ1(x, y).
Thus we obtain
AΘ1 = C ∗(Zd, ωΘ1) ∼= C ∗(Zd, ωΘ2) = AΘ2.
Since Td,A(R) = Td,Q−1ΛQ(R) = QtTd,Λ(R)Q, and we have the following
preposition in general.
Proposition 4.1. For Λ ∈ GLd(Z) where
· · ·
0
· · ·
0
...
. . .
· · · At
where each Ai is the companion matrix for pei
Td,Λ(R), then
A1
0
0 A2
...
...
0
0
Λ =
i of degree di, and for Θ ∈
Θ =
Θ11
−Θt
...
−Θt
Θ12
12 Θ22
...
1t −Θt
2t
· · · Θ1t
· · · Θ2t
...
. . .
· · · Θtt
where for i ≤ j, Θij is a di × dj matrix , and Θij = (θkl) such that θkl =
θ(k+1)(l+1) for all k, l.
Proof. Write Θ = (Θij) then by ΛtΘΛ = Θ we obtain
At
iΘijAj = Θij.
Then by a similar proof to the one in [2, Lemma 4.3], we maintain the form
of Θij.
(cid:3)
Proposition 4.2. Θ ∈ Td,Λ(R) has the following form,
for some Θ′ ∈ Td−1(R).
Θ =(cid:18) Θ′ 0
0 (cid:19)
0
Proof. We write Θdi
ΛtΘΛ = Θ, we have
:= (θdk) where Pi−1
l=1 dl < k ≤ Pi
ΘdiAi = Θdi.
l=1 dl. Then for
Thus it suffices to show when i = 1. Since A1 is the companion matrix for
pe1
1 , we have
(θd,1, θd,2, . . . , θd,d1−1, θd,d1) = (θd,2, θd,3, . . . , θd,d1 , −
d1−1
Xl=0
alθd,l+1)
Certain actions of finitely generated groups on higher dimensional noncommutative tori 16
where
d1−1
alxl = Φp
e1
1
(x) =
Xl=0
Hence we have θ := θd,1 = · · · = θd,d1 and
xip
ei −1
i
.
pi−1
Xk=0
thus θ = 0 and then Θdi = 0.
−(cid:16)
d1−1
Xl=0
al(cid:17)θ = θ,
(cid:3)
Thus for any d-dimensional noncommutative torus AΘ on which Zn = hAi
acts, we have Θ = QtΘ′′Q where
Then we have AΘ ∼= AΘ′′ ∼= AΘ′ ⊗ C(T), which is clearly not simple.
Θ′′ :=(cid:18) Θ′ 0
0 (cid:19) .
0
Hence we have the following theorem.
Theorem 4.3. For a given dimension d and order n, if d − W (n) = 1, then
there is no simple noncommutative tori on which Zn acts in the above way.
Corollary 4.4. For a given dimension d and order n, there is a action of Zn
on a simple noncommutative d-torus AΘ if and only if either d − W (n) > 1
or d − W (n) = 0.
Corollary 4.5. For a given dimension d > 3 which is odd and an order n,
then there is an action of Zn on simple d-dimensional noncommutative tori
if and only if there is an action of Zn of the above type on d − 3-dimensional
noncommutative tori.
Proof. Notice that d − W (n) is an odd integer then by Corollary 4.4 and
Corollary 3.2 we draw the conclusion.
(cid:3)
We mention that Theorem 4.3 is a generalization of [2, Theorem 5.2] by
the following example.
Example 4.6. Consider an action of Zn on a simple 3-dimensional noncom-
mutative torus. Then by Theorem 4.3 and by the fact that W (n) is always
even, it is equivalent to say that d − W (n) = 3. Thus we obtain n = 2, in
other words, the only action by a nontrivial finite cyclic group on a simple
3-dimensional torus of the above type is the flip action by Z2, as stated in
[2, Theorem 5.2].
We rewrite a fact mentioned in [6, Lemma 1.5] as following. In general
for a d-dimensional noncommutative torus AΘ where Θ ∈ Td(R), write
Θ =(cid:18) Θ′
−(θid)t
(θid)
0 (cid:19)
and the standard generator of AΘ as ui for i = 1, . . . , d. Then
AΘ′ = C ∗{ui i = 1, . . . , d − 1},
Certain actions of finitely generated groups on higher dimensional noncommutative tori 17
moreover we can define an action α : Z → Aut(AΘ′) by α(1)(ui) = exp(2πiθid)ui
which is homotopic to the identity and
AΘ ∼= AΘ′ ⋊α Z.
Return to our current case for the given dimension d and the order n
with d − W (n) = 1. We can construct a noncommutative d-torus AΘ where
Θ ∈ Td(R) and
Θ =(cid:18) Θ′ 0
0 (cid:19) ,
0
and AΘ admits an action of Zn. Although AΘ ∼= AΘ′ ⊗ C(T) is never
simple, or equivalently Θ is never nondegenerate, by our construction Θ′ is
nondegenerate, and the simple noncommutative d − 1-tori AΘ′ admits an
action of Zn which is denoted by α′. Hence write ui for i = 1, . . . , d the
standard generators of AΘ, we could define an action
by the following,
α : Z × Zn → Aut(AΘ)
α(1, I)(ui) = ui,
α(1, I)(ui) = exp(2πiθ)ui,
α(0, A)(ui) = α′(ui),
α(0, A)(ui) = ui,
i = 1, . . . , d − 1;
i = d;
i = 1, . . . , d − 1;
i = d
where θ is an irrational number. It is actually the product action of two
commuting actions. Then we have
AΘ ⋊α (Z × Zn) ∼= (AΘ ⋊α Z) ⋊ α Zn ∼= (AΘ′ ⊗ Aθ) ⋊α′ Zn.
Since
Θ′′ = Θ′ ⊕(cid:18) 0
−θ 0 (cid:19)
θ
is simple, and action α′ of Zn on AΘ′′ ∼= AΘ′ ⊗ Aθ is of the above type,
the the crossed product AΘ ⋊α (Z × Zn) is covered by a discussion in Sec-
tion 3. We comment that action α restricted to Z is adding dimension to
noncommutative torus AΘ and making it a simple one on which there is an
action of Zn of above type, in our context this should be the action of Z on
a noncommutative torus in which we are interested.
5. Finitely generated abelian group actions
Motived by the last part of the former section, we have the following
discussion.
For a given dimension d and a finitely generated abelian group G, by the
structure theorem we have
G ∼=(cid:16)
s
Yi=1
Zp
ei
i (cid:17) × Zr,
Certain actions of finitely generated groups on higher dimensional noncommutative tori 18
if d − W (G) = 0 or d − W (G) > 0, there is an action of Qs
on a
simple noncommutative d-torus AΘ by Theorem 3.6, say α′. If r > 0 we
may define an action α : G → Aut(AΘ) with a nondegenerate Θ0 ∈ Tr(R)
such that
i=1 Zp
ei
i
s
AΘ ⋊α G ∼= (AΘ ⊗ AΘ0) ⋊α′
Zp
ei
i
.
Yi=1
Note that there is also a similar action when r = 0 stated in Theorem 3.6.
If d−W (n) = 1, there is a noncommutative d-torus AΘ of form AΘ′⊗C(T),
or equivalently, with
Θ =(cid:18) Θ′ 0
0 (cid:19)
0
where Θ′ ∈ Td−1(R) is nondegenerate. Both AΘ and AΘ′ admit an action
of Zn. Denote the latter one by α′. If additionally r > 0, we can obtain
an action α : G → Aut(AΘ). In such case if r > 1 we obtain α with an
irrational number θ and a nondegenerate Θ0 ∈ Tr−1(R) such that
AΘ ⋊α G ∼= (AΘ′ ⊗ Aθ ⊗ AΘ0) ⋊α′
s
Yi=1
Zp
ei
i
.
Similarly if r = 1 we obtain α with only an irrational number θ such that
AΘ ⋊α G ∼= (AΘ′ ⊗ Aθ) ⋊α′
s
Yi=1
Zp
ei
i
.
Theorem 5.1. For a given dimension d and a finitely generated abelian
ei
i=1 Zp
group G ∼= (cid:16)Qs
i (cid:17) × Zr, If d − W (n) = 1 and r > 0, or d − W (n) 6=
1 then there is a noncommutative d-torus AΘ admitting an action of G,
denoted by α : G → Aut(AΘ). We require AΘ is pseudo-simple in the first
case and simple in the second case. Then the crossed product AΘ ⋊α G is a
simple AT algebra, and it is an AF algebra if and only if r = 0, d−W (G) = 0
and G satisfies the last condition in Theorem 3.6.
Proof. The existence is by our discussion in this section. Then by Theorem
3.6 we draw the conclusion.
(cid:3)
Acknowledgment
The author would like to express his deep gratitude to his supervisor,
Professor Yasuyuki Kawahigashi, for helpful suggestions, warm encourage-
ment and support during two years of his master course. The author would
like to thank Dr. Chen Jiang, Dr. Yuki Arano, Dr. Yosuke Kubota, Dr.
Shuhei Masumoto, Dr. Takuya Takeishi and Dr. Lu Xu for many useful
discussions. The author is grateful to the University of Tokyo for Special
Scholarship for International Students (Todai Fellowship).
Certain actions of finitely generated groups on higher dimensional noncommutative tori 19
References
[1] S. Echterhorff, W. Luck, N. C. Phillips, S. Walters, The structure of crossed products
of irrational rotation algebras by finite subgroups of SL2(Z), J. Reine Angew. Math.
639 (2010), 173 -- 221.
[2] J. A. Jeong, J. H. Lee, Finite groups acting on higher dimensonal noncommutative
tori, arXiv:1402.1826v2.
[3] J. Kuzmanovich, A. Pavlichenkov, Finite groups of matrices whose entires are interg-
ers, Amer. Math. Monthly 109 (2002), 173 -- 186.
[4] M. Langer, W. Luck, Topological K-theroy of the group C ∗-algebra of a semi-direct
product Zn ⋊ Z/m for a free conjugation action, J. Topol. Anal. 4 (2012), 121 -- 172.
[5] J. A. Packer, I. Raeburn, Twisted crossed products of C ∗-algebras, Math. Proc. Cam-
bridge Philos. Soc. 106 (1989), 293 -- 311.
[6] N. C. Phillips, Every
higher noncommutative Torus
is
an AT-algebra,
arXiv:math/0609783.
[7] N. C. Phillips, The tracial Rokhlin property for actions of finite groups on C ∗-algebras,
Amer. J. Math. 133 (2011), 581 -- 636.
[8] C. Schochet, Topological methods for C ∗-algebras II: Geometric resolutions and the
Kunneth formula, Pacific J. Math. 98 (1982), 443 -- 458.
Graduate School of Mathematical Sciences, the University of Tokyo, 3-8-1
Komaba, Meguro-ku, Tokyo 153-8914, Japan.
E-mail address: [email protected]
|
1711.08786 | 2 | 1711 | 2018-04-11T13:16:38 | On a class of determinant preserving maps for finite von Neumann algebras | [
"math.OA",
"math.FA"
] | Let $\mathscr{R}$ be a finite von Neumann algebra with a faithful tracial state $\tau $ and let $\Delta$ denote the associated Fuglede-Kadison determinant. In this paper, we characterize all unital bijective maps $\phi$ on the set of invertible positive elements in $\mathscr{R}$ which satisfy $$\Delta(\phi(A)+\phi(B)) = \Delta(A+B).$$ We show that any such map originates from a $\tau$-preserving Jordan $*$-automorphism of $\mathscr{R}$ (either $*$-automorphism or $*$-anti-automorphism in the more restrictive case of finite factors). In establishing the aforementioned result, we make crucial use of the solutions to the equation $\Delta(A + B) = \Delta(A) + \Delta(B)$ in the set of invertible positive operators in $\mathscr{R}$. To this end, we give a new proof of the inequality $$\Delta(A+B) \ge \Delta(A) + \Delta(B),$$ using a generalized version of the Hadamard determinant inequality and conclude that equality holds for invertible $B$ if and only if $A$ is a nonnegative scalar multiple of $B$. | math.OA | math | ON A CLASS OF DETERMINANT PRESERVING MAPS FOR FINITE
VON NEUMANN ALGEBRAS
MARCELL GA ´AL AND SOUMYASHANT NAYAK
Abstract. Let R be a finite von Neumann algebra with a faithful tracial state τ and let
∆ denote the associated Fuglede-Kadison determinant. In this paper, we characterize all
unital bijective maps φ on the set of invertible positive elements in R which satisfy
∆(φ(A) + φ(B)) = ∆(A + B).
We show that any such map originates from a τ -preserving Jordan ∗-automorphism of R
(either ∗-automorphism or ∗-anti-automorphism in the more restrictive case of finite factors).
In establishing the aforementioned result, we make crucial use of the solutions to the equation
∆(A + B) = ∆(A) + ∆(B) in the set of invertible positive operators in R. To this end, we
give a new proof of the inequality
using a generalized version of the Hadamard determinant inequality and conclude that
equality holds for invertible B if and only if A is a nonnegative scalar multiple of B.
∆(A + B) ≥ ∆(A) + ∆(B),
8
1
0
2
r
p
A
1
1
]
.
A
O
h
t
a
m
[
2
v
6
8
7
8
0
.
1
1
7
1
:
v
i
X
r
a
1. Introduction
In 1897 Frobenius [9] proved that if φ is a linear map on the matrix algebra Mn(C) of
n × n complex matrices preserving the determinant, then there are matrices M, N ∈ Mn(C)
such that det(MN) = 1 and φ can be written in one of the following forms:
a) φ(A) = MAN, A ∈ Mn(C);
b) φ(A) = MAtN, A ∈ Mn(C)
where (·)t denotes transposition of a matrix.
In the past decades this result of Frobenius has inspired many researchers to deal with
different sorts of preserver problems involving various notions of determinant [2, 5, 7, 13, 17,
20]. Among others, in [13] Huang et al. completely described all maps on the positive definite
cone Pn of Mn(C) which satisfy the sole property det(φ(A) + φ(B)) = det φ(I) · det(A + B)
for all A, B ∈ Pn. Note that when φ is a unital linear map on Pn, the above property simply
means that φ is det preserving. In this paper, we consider an identical operator algebraic
counterpart of this problem in the setting of finite von Neumann algebras.
Our approach to the solution is based on a generalization of the Minkowski determinant
inequality to the setting of von Neumann algebras. Note that the usual Minkowski determi-
nant inequality for matrices A, B ∈ Pn asserts that
npdet(A + B) ≥ npdet(A) + npdet(B),
with equality if and only if A, B are positive scalar multiples of each other. In [1, Corol-
lary 4.3.3 (i)], Arveson gives a variational proof of a version of the Minkowski determinant
1
2
MARCELL GA ´AL AND SOUMYASHANT NAYAK
inequality in finite von Neumann algebras involving the Fuglede-Kadison determinant. Re-
cently, this result has been subsumed by the study of the anti-norm property of a wide class
of functionals by Bourin and Hiai [4, Corollary 7.6]. The equality conditions are harder
to isolate from these proofs because of limiting arguments and are not explicitly docu-
mented. As that will play an important role in our results, we first need to establish when
∆(A+ B) = ∆(A) + ∆(B) holds for positive operators A, B in a finite von Neumann algebra.
To this end, we give a new proof of the inequality using a generalized version of the Hadamard
determinant inequality [11, 8] in a bootstrapping argument. Consequently, we are able to
solve the aforementioned preserver problem concerning Fuglede-Kadison determinants on
finite von Neumann algebras.
The paper is organized as follows. In § 2 we fix the notation and briefly review some facts
from the theory of determinants on von Neumann algebras. The precise formulations of our
corresponding results and their proofs are collected in § 3 and § 4.
2. Preliminaries
Throughout this paper, R denotes a finite von Neumann algebra acting on the complex
(separable) Hilbert space H and containing the identity operator I. Let τ be a faithful
tracial state on R, by which we mean a linear functional τ : R → C such that for all
A, B ∈ R, we have (i) τ (AB) = τ (BA), (ii) τ (A∗A) ≥ 0 with equality if and only if A = 0,
(iii) τ (I) = 1. The set of invertible operators in R is denoted by GL1(R). We denote the
cone of positive operators in R by R+ and use GL1(R)+ to denote the set of invertible
operators in R+.
For A ∈ GL1(R), the Fuglede-Kadison determinant ∆ associated with τ is defined as
∆(A) = exp(τ (log √A∗A)).
The dependence of ∆ on τ is suppressed in the notation and it is to be assumed that a choice
of a faithful tracial state has already been made. Although this concept of determinant was
developed in [10] in the context of type II1 factors, it naturally extends to finite von Neumann
algebras as above.
Example 2.1. The simplest examples of finite von Neumann algebras are given by Mn(C),
the full matrix algebra of n × n complex matrices. For A ∈ Mn(C), the Fuglede-Kadison
determinant ∆(A) is given by npdet(A) where det is the usual matrix determinant.
Example 2.2. On M2(C) the unique faithful tracial state is given by
tr2 : M2(C) → C,
tr2(A) =
a11 + a22
2
where aij ∈ C (1 ≤ i, j ≤ 2) denotes the (i, j)th entry of the matrix A in M2(C). Denote by
D2(C) the ∗-subalgebra of diagonal matrices in M2(C). The von Neumann algebra M2(R) ∼=
R ⊗ M2(C) (acting on H ⊕ H ) is also finite and the faithful tracial state on M2(R) is given
by τ2 = τ ⊗ tr2, that is, for an operator A in M2(R), we have
τ (A11) + τ (A22)
.
τ2(A) =
2
DETERMINANT PRESERVING MAPS
3
We denote by ∆2 the Fuglede-Kadison determinant on M2(R) corresponding to τ ⊗ tr2. For
operators A1, A2 in R, we define
diag(A1, A2) := (cid:20)A1
0 A2(cid:21) ∈ D2(R) ∼= R ⊗ D2(C) ⊂ M2(R).
0
It is straightforward to see that for invertible operators A1, A2 in R, the operator diag(A1, A2)
in M2(R) is invertible and ∆2(diag(A1, A2)) = p∆(A1) · ∆(A2).
Example 2.3. Let X be a compact (Hausdorff) topological space with a probability Radon
measure ν. The space of essentially bounded complex-valued functions on (X, ν), denoted
by L∞(X, ν), which acts via left multiplication on L2(X, ν), forms an abelian von Neumann
algebra. The involution operation is given by f ∗(x) := f (x). A faithful tracial state on
L∞(X, ν) is obtained by
τν(f ) = ZG
f (x) dν(x), for f ∈ L∞(X, ν)
and the corresponding Fuglede-Kadison determinant determinant is given by
∆ν(f ) = exp(cid:18)ZG
log(f (x))dν(x)(cid:19) .
Group von Neumann algebras provide another important class of examples of finite von
Neumann algebras, see e.g. [19, §3.2].
One of the most remarkable properties of ∆ is that it is a group homomorphism of GL1(R)
into the multiplicative group of positive real numbers. However, there may be several exten-
sions of ∆ from GL1(R) to the whole of R. From the proof of [10, Lemma 6], note that for a
projection E 6= I in R, we have ∆′(E) = 0 for any extension ∆′ of ∆. In this paper, we con-
sider only the analytic extension which is defined as follows. For A ∈ R, let σ(A) ⊂ [0,∞)
denote the spectrum of √A∗A and let µ be the probability measure supported on σ(A) and
induced by the tracial state τ . Then we define
∆(A) := exp(cid:18)Zσ(A)
log λ dµ(λ)(cid:19)
and denote this extension also by ∆.
with understanding that ∆(A) = 0 whenever Rσ(A) log λ dµ(λ) = −∞. We abuse notation
Below we summarize some properties of ∆ which we shall need in § 3.
(p1) ∆(U) = 1 for a unitary U in R;
(p2) ∆(AB) = ∆(A) · ∆(B) for A, B ∈ R;
(p3) ∆ is norm continuous on GL1(R);
(p4) ∆(λA) = λ∆(A) for λ ∈ C, A ∈ R;
(p5) limε→0+ ∆(A + εI) = ∆(A) for a positive operator A in R.
3. The Minkowski determinant inequality
In this section, we aim to establish the following version of the Minkowski determinant
inequality.
4
MARCELL GA ´AL AND SOUMYASHANT NAYAK
Theorem 3.1 (generalized Minkowski determinant inequality). For positive operators A, B
in R, we have
(3.1)
∆(A + B) ≥ ∆(A) + ∆(B).
Moreover, if B is invertible, then equality holds in (3.1) if and only if A is a nonnegative
scalar multiple of B.
We work towards the proof of Theorem 3.1 using several lemmas. Before turning to their
proof, let us explain the main ideas implemented in them.
A proof of the Minkowski determinant inequality (see [16, p. 115]) for matrices is based on
the 'traditional' Hadamard determinant inequality which states that for a positive definite
matrix A in Mn(C), the determinant of A is less than or equal to the product of its diagonal
entries and equality holds if and only if A is a diagonal matrix. For a given A ∈ Mn(C),
considering the positive semidefinite matrix √A∗A, one may derive from this inequality the
geometrically intuitive fact that the volume of an n-parallelepiped with prescribed lengths
of edges is maximized when the edges are mutually orthogonal. In this paper, we make use
of an 'abstract' Hadamard-type determinant inequality in our proof to reflect the geometric
origins of inequality (3.1).
Recall that if S is a von Neumann subalgebra of R, then by a conditional expectation
we mean a unital (identity preserving) positive linear map Φ : R → S which satisfies
Φ(SAT ) = SΦ(A)T for all A ∈ R and S, T ∈ S . Concerning τ -preserving conditional
expectations, in [18, Theorem 4.1] the second author has proved the following generalization
of the Hadamard determinant inequality:
Theorem 3.2. For a τ -preserving conditional expectation Φ on R and an invertible positive
operator A in R, we have that
∆(Φ(A−1)−1) ≤ ∆(A) ≤ ∆(Φ(A))
and equality holds in either of the above two inequalities (and hence in both inequalities) if
and only if Φ(A) = A.
We consider the map Φ2 : M2(R) → M2(R) defined by
Φ2(A) := diag(A11, A22) ∈ D2(R) ⊂ M2(R)
where Aij ∈ R(1 ≤ i, j ≤ 2) is the (i, j)th entry of A. Note that Φ2 is a (τ ⊗ tr2)-preserving
normal conditional expectation from M2(R) onto the von Neumann subalgebra D2(R). In
Lemma 3.3, we will use Theorem 3.2 in the context of the von Neumann algebra M2(R)
and the (τ ⊗ tr2)-preserving normal conditional expectation Φ2. More precisely, for positive
operators A1, A2 in R we first prove in Lemma 3.3 that
(3.2)
∆(A1) · ∆(A2) ≤ ∆(tA1 + (1 − t)A2) · ∆(tA2 + (1 − t)A1),
t ∈ [0, 1].
Choosing t = 1/2, we arrive at
∆(A1) · ∆(A2) ≤ ∆(cid:18)A1 + A2
2
(cid:19)2
DETERMINANT PRESERVING MAPS
5
which readily implies that 2p∆(A1) · ∆(A2) ≤ ∆(A1 +A2), which is weaker than the desired
inequality. We then use a "tensor power trick" to proceed with a bootstrapping argument
to prove the required inequality.
Now we are in a position to prove our first lemma.
Lemma 3.3. For positive operators A1, A2 in R and t ∈ [0, 1], the following inequality
holds:
(3.3)
∆(A1) · ∆(A2) ≤ ∆(tA1 + (1 − t)A2) · ∆(tA2 + (1 − t)A1).
Further if A1, A2 are invertible, then equality holds if and only if either t ∈ {0, 1} or A1 = A2.
Proof. Consider the unitary operator U in M2(R) given by
U := (cid:20) √tI
√1 − tI
√1 − tI −√tI (cid:21) .
Note that
U ∗diag(A1, A2)U = (cid:20) tA1 + (1 − t)A2 pt(1 − t)(A1 − A2)
tA2 + (1 − t)A1 (cid:21) .
pt(1 − t)(A1 − A2)
Clearly, Φ2(U ∗diag(A1, A2)U) = diag(tA1 + (1 − t)A2, tA2 + (1 − t)A1). Using Theorem 3.2
and property (p5) concerning ∆, we get that
p∆(A1) ·p∆(A2) = ∆2(diag(A1, A2)) = ∆2(U ∗diag(A1, A2)U)
≤ ∆2(Φ2(U ∗diag(A1, A2)U)) = p∆(tA1 + (1 − t)A2) ·p∆(tA2 + (1 − t)A1).
If A1, A2 are invertible, then U ∗diag(A1, A2)U is also invertible and equality holds if and
(cid:3)
only if pt(1 − t)(A1 − A2) = 0, that is, either t ∈ {0, 1} or A1 = A2.
Lemma 3.4. Let n be a positive integer. For positive operators A1, A2, . . . , An in R, the
following inequality holds:
(3.4)
∆(cid:16) A1 + · · · + An
n
(cid:17) ≥ (∆(A1) · · · ∆(An))1/n.
Further if A1, . . . , An are invertible, then equality holds if and only if A1 = A2 = . . . = An.
Proof. First for n = 2k with k ∈ N we prove by induction the inequality (3.4) along with
the equality condition and then employ a standard argument to establish it for all n ∈ N.
Choosing t = 1/2 in Lemma 3.3, we get for A1, A2 ∈ R+ that
(cid:17) ≥ (∆(A1) · ∆(A2))1/2.
(3.5)
∆(cid:16)A1 + A2
2
Further if A1, A2 are invertible, equality holds if and only if A1 = A2. This proves the case
when n = 2.
Now assume that inequality (3.4) holds for n = 2k−1 along with the equality condition.
For A1, A2, . . . , A2k ∈ R+, we define
B1 :=
A1 + · · · + A2k−1
2k−1
, B2 :=
A2k−1+1 + · · · + A2k
2k−1
.
6
MARCELL GA ´AL AND SOUMYASHANT NAYAK
From (3.5), we infer that
(3.6)
Furthermore, the induction hypothesis furnishes
∆(cid:16)B1 + B2
2
(cid:17) ≥ (∆(B1)∆(B2))1/2.
(3.7)
∆(B1) ≥ (∆(A1) · · · ∆(A2k−1))1/2k−1
, ∆(B2) ≥ (∆(A2k−1+1) · · · ∆(A2k))1/2k−1
.
Combining (3.6) and (3.7), we have
(3.8)
∆(cid:16)A1 + · · · + A2k
2k
(cid:17) ≥ (∆(A1) · · · ∆(A2k ))1/2k
.
If A1, . . . , A2k are invertible, then so are B1, B2, and equality holds if and only if B1 = B2
A1 = · · · = A2k−1 and A2k−1+1 = · · · = A2k or, in other words, if and only if A1 = · · · = A2k .
Thus by induction, for n a power of 2, we have established inequality (3.4) along with the
equality condition.
Next we consider an arbitrary positive integer m. Let k be a positive integer such that
2k−1 ≤ m < 2k. For positive operators A1, . . . , Am, we define a positive operator B by
B := (A1 + · · · + Am)/m. It follows that
∆(B) = ∆(cid:16) A1 + · · · + Am + (2k − m)B
2k
(cid:17) ≥ (∆(A1) · · · ∆(Am))1/2k
(∆(B))1−m/2k
and using property (p5), we conclude that
≥ (∆(A1) · · · ∆(Am))1/2k
∆(B)m/2k
=⇒ ∆(B) ≥ (∆(A1) · · · ∆(Am))1/m.
If A1, . . . , Am are invertible, then so is B and equality holds if and only if A1 = A2 = . . . =
Am = B.
(cid:3)
Theorem 3.5. For a positive operator A in R and t ≥ 0, the following inequality holds:
(3.9)
∆(tI + A) ≥ t + ∆(A)
with equality if and only if either t = 0 or A is a nonnegative scalar multiple of I.
Proof. Let A be an invertible positive operator such that ∆(A) = 1. For p, q ∈ N, an
application of Lemma 3.4 gives us that
∆(cid:18) pI + qA
p + q (cid:19) ≥ p+qp∆(I)p∆(A)q = 1.
Thus ∆((p/q)I + A) ≥ p/q + 1 with equality if and only if A = I. Approximating with
strictly positive rational numbers, we have by property (p5) for ∆ that
(3.10)
Note that for A ∈ GL1(R)+ the operator B := (1/∆(A))A is an invertible positive operator
satisfying ∆(B) = 1. As ∆(tI + B) ≥ t + 1, for t ≥ 0 substituting s = t∆(A), we get the
desired inequality
∆(tI + A) ≥ t + 1,
for t ≥ 0.
(3.11)
∆(sI + A) ≥ s + ∆(A),
for s ≥ 0.
Next we derive conditions for the case of equality in (3.11). Note that for a particular
value of s under consideration, if s/∆(A) is rational, then equality holds in (3.11) if and only
DETERMINANT PRESERVING MAPS
7
if B = I ⇔ A = ∆(A)I. For s > 0, if ∆(sI + A) = s + ∆(A), using (3.11) repeatedly along
with the multiplicativity of ∆, we get that
s2 + ∆(A)(2s + ∆(A)) = (s + ∆(A))2 = ∆(sI + A)2 = ∆(s2I + A(2sI + A))
Thus we conclude that ∆(2sI + A) = 2s + ∆(A). For r ∈]0, s[, using (3.11) we deduce that
≥ s2 + ∆(A)∆(2sI + A) ≥ s2 + ∆(A)(2s + ∆(A)).
2s + ∆(A) = ∆(2sI + A) = ∆((s + r)I + (s − r)I + A)
≥ (s + r) + ∆((s − r)I + A)
≥ (s + r) + (s − r) + ∆(A) = 2s + ∆(A).
As (s + r) + ∆((s − r)I + A) = 2s + ∆(A), we have that ∆((s − r)I + A) = s − r + ∆(A) for
all r ∈]0, s[. We may choose r such that (s − r)/∆(A) is rational and thus conclude that A
is a scalar multiple of the identity. Hence equality holds in (3.11) if and only if either s = 0
or A is a scalar multiple of the identity.
Next we consider the case when A is not necessarily invertible. For some t > 0 and
any s ∈]0, t], define As := sI + A. As documented in property (p5) for ∆, note that
limε→0+ ∆(Aε) = ∆(A). We have
∆(At) = ∆(cid:18) t
I + At/2(cid:19) ≥
Taking the limit k → ∞, we conclude that
∆(tI + A) = ∆(At) ≥
2
t
2
+ ∆(At/2) ≥
k
Xi=1
t
2i + ∆(At/2k ).
∞
Xi=1
t
2i + ∆(A) = t + ∆(A).
If A is a scalar multiple of the identity, equality trivially holds. If ∆(tI + A) = t + ∆(A), we
must have ∆(At) = t/2 + ∆(At/2) and thus At/2 is a scalar multiple of I implying that A is
a scalar multiple of I.
(cid:3)
Proof of Theorem 3.1. If B is invertible, then by (3.9) we infer that
2 ) ≥ 1 + ∆(B− 1
2 AB− 1
2 )
∆(I + B− 1
2 AB− 1
2 AB− 1
with equality if and only if B− 1
2 = λI with some λ ≥ 0. Using the multiplicative
property of the determinant ∆, we conclude that ∆(A + B) ≥ ∆(A) + ∆(B) with equality
if and only if A = λB with some λ ≥ 0.
If B is not invertible, we consider the invertible positive operator Bε := B + εI (ε > 0).
Then we have ∆(A + Bε) ≥ ∆(A) + ∆(Bε) and taking the limit ε → 0+, we see that
∆(A + B) ≥ ∆(A) + ∆(B), as required.
Remark 3.6. If A or B is invertible, then the condition of equality has a straightforward
form demanding a scaling relationship between the operators unless one of them is 0. But if
neither A nor B is invertible, then the conditions under which equality holds are less easy
to characterize. Assume that E, F are orthogonal projections in R such that E + F < I.
We then have ∆(E + F ) = 0 = ∆(E) + ∆(F ) and thus we cannot expect the operators in
(cid:3)
8
MARCELL GA ´AL AND SOUMYASHANT NAYAK
question to have any specific 'correlation', a term which we do not define but whose spirit is
captured in the above example.
4. A class of determinant preserving maps
In this section, we present our result concerning certain determinant preserving bijective
maps on GL1(R)+. Before we do so let us recall the concept of Jordan ∗-isomorphisms.
Definition 4.1. A linear map J : R → R is called a
morphism;
(i) Jordan homomorphism if J(A2) = J(A)2, for all A ∈ R;
(ii) Jordan ∗-homomorphism if J(A∗) = J(A)∗ for all A ∈ R and J is a Jordan homo-
(ii) Jordan ∗-isomorphism if J is a bijective Jordan ∗-homomorphism.
Whenever we use the terms ∗-homomorphism, ∗-isomorphisms and ∗-automorphism, it
is implicitly understood to refer to the C ∗-algebraic structure of R. A celebrated result
of Kadison [14, Corollary 5] states that a (linear) order automorphism of a C ∗-algebra is
necessarily implemented by a Jordan ∗-automorphism. Roughly speaking, this means that
in a C ∗-algebra the order and the Jordan structures are intimately connected and in fact,
determine each other. The mentioned result of Kadison was crucially used in [3, Lemma 8]
where the structure of additive bijective maps was determined on the cone GL1(R)+. We
apply this in the proof of the main result of the section which is as follows.
Theorem 4.2. Let φ : GL1(R)+ → GL1(R)+ be a bijection. Then
∆(φ(A) + φ(B)) = ∆(φ(I)) · ∆(A + B)
holds for all A, B in GL1(R)+ if and only if there is a τ -preserving Jordan ∗-isomorphism
J : R → R and a positive invertible element T ∈ GL1(R)+ such that
φ(A) = T J(A)T, A ∈ GL1(R)+.
The proof uses the main ideas in [13], however, we must adapt them to the much more
general setting of finite von Neumann algebras. Before turning to the proof of Theorem
4.2, we paraphrase an auxiliary lemma from [10] which makes use of the Riesz-Dunford
holomorphic functional calculus for Banach algebras [6] to derive a pertinent corollary.
Lemma 4.3 ([10, Lemma 2]). Let B be a complex Banach algebra with a norm-continuous
tracial linear functional T. Let f (λ) be a holomorphic function on a domain Λ ⊂ C bounded
by a curve Γ and let γ : [0, 1] → R be a differentiable family of operators in R, such that
the spectrum of each operator γ(t) lies in Λ. Then f (γ(t)) is differentiable with respect to t
and
T((f ◦ γ)′(t))) = T(f ′(γ(t)) · γ′(t)).
Corollary 4.4. For invertible positive operators A, B in R, the function g : [0, 1] → R
defined by g(t) = ∆(tA + (1− t)B) is differentiable at 0+ and g′(0+) = ∆(B)(τ (B−1A)− 1).
DETERMINANT PRESERVING MAPS
9
Proof. Let γ : [0, 1] → GL1(R)+ be the line segment γ(t) = tA + (1 − t)B. Clearly, γ is a
continuously differentiable curve with γ′(t) = A − B for all t ∈ [0, 1]. If ε > 0 is such that
εI ≤ A and εI ≤ B, we have that εI ≤ γ(t) for any t ∈ [0, 1]. Thus we may choose a domain
Λ ⊆ C not containing 0 that is bounded by a curve Γ which surrounds the spectra of γ(t)
for all t ∈ [0, 1] and does not wind around 0. On the domain Λ, f = log is a holomorphic
function. Define G(t) := τ (log(γ(t)))). Using Lemma 4.3, we get that
G′(0+) = τ (B−1(A − B)) = τ (B−1A − I) = τ (B−1A) − 1.
As g(t) = exp G(t), we conclude that g′(0+) = exp(G(0))·G′(0+) = ∆(B)(τ (B−1A)−1). (cid:3)
Proof of Theorem 4.2. Let us begin with the necessity part. Observe that the transformation
ψ : GL1(R)+ → GL1(R)+ defined by ψ(A) = φ(I)−1/2φ(A)φ(I)−1/2 is unital. From the
multiplicativity of ∆, we see that
∆(ψ(A) + ψ(B)) = ∆(A + B),
(4.1)
Plugging A = B into equation (4.1), we deduce that ∆(ψ(A)) = ∆(A) for every A ∈
GL1(R)+ and thus for a positive real number λ > 0 we obtain
for A, B ∈ GL1(R)+.
∆(ψ(A) + ψ(λA)) = ∆(A + λA) = ∆(A) + ∆(λA) = ∆(ψ(A)) + ∆(ψ(λA)).
An application of Theorem 3.1 entails that ψ(λA) = µψ(A) for some µ > 0. As noted
earlier, ∆(ψ(A)) = ∆(A) which implies λ = µ meaning that ψ is positive homogeneous. For
A, B ∈ GL1(R)+ and a number t ∈]0, 1[, we get that
∆(tψ(A) + (1 − t)ψ(B)) = ∆ (ψ(tA) + ψ((1 − t)B))
= ∆(tA + (1 − t)B).
For s ∈ {0, 1}, as ∆(ψ(A)) = ∆(A), ∆(ψ(B)) = ∆(B), it follows that ∆(sψ(A) + (1 −
s)ψ(B)) = ∆(sA+(1−s)B). In summary, we have for t ∈ [0, 1] that ∆(tψ(A)+(1−t)ψ(B)) =
∆(tA + (1 − t)B). Taking the derivative of both sides with respect to t at 0+, and using
Corollary 4.4, we obtain that
(4.2)
τ (ψ(B)−1ψ(A)) = τ (B−1A), for all A, B ∈ GL1(R)+.
The right hand side of (4.2) is additive in the variable A. As B runs through the whole of
GL1(R)+, substituting X = ψ(B)−1, it follows from (4.2) that for all A, C, X ∈ GL1(R)+,
we must have
τ (Xψ(A + C)) = τ (Xψ(A)) + τ (Xψ(C)),
or, equivalently,
τ (X[ψ(A + C) − (ψ(A) + ψ(C))]) = 0.
Since a self-adjoint operator X in R may be written as the difference of two invertible
positive operators X + (kXk + ε)I, (kXk + ε)I for ε > 0, we further have that τ (X[ψ(A +
C) − (ψ(A) + ψ(C))]) = 0 for all A, C ∈ GL1(R)+ and all self-adjoint operators X in R.
Choosing X = ψ(A + C) − (ψ(A) + ψ(C)) and using the faithfulness of the tracial state τ ,
we conclude that ψ(A + C) − (ψ(A) + ψ(C)) = 0 for all A, C ∈ GL1(R)+. Thus, ψ is an
additive bijection. The structure of such maps is described in [3]. According to [3, Lemma 8]
there is a Jordan ∗-isomorphism J : R → R such that ψ(A) = J(A) for all A ∈ GL1(R)+.
10
MARCELL GA ´AL AND SOUMYASHANT NAYAK
the necessity part.
The desired τ -preserving property also follows from (4.2). Setting T := pφ(I) completes
We next prove the sufficiency. It is well-known that for a Jordan ∗-homomorphism J on
R and a continuous function f defined on the spectrum of a self-adjoint operator A in R, we
have J(f (A)) = f (J(A)). As J is assumed to be τ -preserving, we observe that ∆(J(A)) =
∆(A), by taking f = log. With these considerations in mind and by the multiplicativity of
∆ we finally conclude that, if there is a τ -preserving Jordan ∗-homomorphism J : R → R
and an operator T in GL1(R)+ such that Φ(A) = T J(A)T for all A ∈ GL1(R)+, then we
must have
∆(φ(A + B)) = ∆(T )∆(J(A + B))∆(T ) = ∆(T 2)∆(A + B) = ∆(φ(I)) · ∆(A + B).
(cid:3)
In the particular case of finite factors (that is, von Neumann algebras with trivial center
CI), the structure of Jordan ∗-automorphisms is quite straightforward from [12, Theorem
I] as finite factors are simple rings (see [15, Corollary 6.8.4]). This helps us elucidate the
solution to the preserver problem considered in this paper in a simple manner. Note that
∗-automorphisms and ∗-anti-automorphisms of a finite factor are trace preserving.
Corollary 4.5. Let R be a finite factor and φ : GL1(R)+ → GL1(R)+ be a bijective map.
Then we have
∆(φ(A) + φ(B)) = ∆(φ(I)) · ∆(A + B), for all A, B ∈ GL1(R)+
if and only if there is a ∗-automorphism (or ∗-anti-automorphism) θ of R and a positive
invertible element T ∈ GL1(R)+ such that
φ(A) = T θ(A)T, A ∈ GL1(R)+.
5. Acknowledgement
The first author was supported by the National Research, Development and Innovation
Office NKFIH Reg. No. K-115383.
References
[1] W. Arveson, Analyticity in operator algebras, Amer. J. Math. 89 (1967), 578-642.
[2] B. Aupetit, Spectrum-preserving linear mappings between Banach algebras or Jordan-Banach algebras, J.
London Math. Soc. 62 (2000), 917 -- 924.
[3] R. Beneduci and L. Moln´ar, On the standard K-loop structure of positive invertible elements in a C ∗-
algebra, J. Math. Anal. Appl. 420 (2014), 551 -- 562.
[4] J. Bourin and F. Hiai, Anti-norms on finite von Neumann algebras, Publ. Res. Inst. Math. Sci. 51 (2015),
207 -- 235.
[5] G. Dolinar and P. Semrl, Determinant preserving maps on matrix algebras, Linear Algebra Appl. 348
(2002), 189 -- 192.
[6] N. Dunford, Spectral theory, Bull. Amer. Math. Soc. vol. 49 (1943), 637 -- 651.
[7] M. L. Eaton, On linear transformations which preserve the determinant, Illinois J. Math. 13 (1969)
722 -- 727.
[8] E. Fischer, Uber den Hadamardschen Determinentsatz, Arch. Math. u. Phys. (3), 13 (1907), 32 -- 40.
DETERMINANT PRESERVING MAPS
11
[9] G. Frobenius, Uber die Darstellung der endlichen Gruppen durch lineare Substitutionen, Si. (1897) 994 --
1015.
[10] B. Fuglede and R. V. Kadison, Determinant theory in finite factors, Annals of Math. 55 (1952), 520 -- 530.
[11] J. Hadamard, R´esolution dune question relative aux determinants, Bull. des sciences math. (17)(1893),
240 -- 246.
[12] I. N. Herstein, Jordan homomorphisms, Transactions of the American Mathematical Society, Vol. 81,
No. 2 (Mar., 1956), 331 -- 341.
[13] H. Huang, C. N. Liu, P. Szokol, M. C. Tsai and J. Zhang, Trace and determinant preserving maps of
matrices, Linear Algebra Appl. 507 (2016), 373 -- 388.
[14] R. V. Kadison, A Generalized Schwarz Inequality and Algebraic Invariants for Operator Algebras, Annals
of Mathematics, Second Series, Vol. 56, No. 3 (Nov., 1952), 494 -- 503.
[15] R. Kadison, J. Ringrose, Fundamentals of the Theory of Operator Algebras, Vol. I-II, Academic Press,
San Diego, 1986 (Reprinted by AMS in 1997).
[16] M. Marcus, H. Minc, A Survey of Matrix Theory and Matrix Inequalities, Dover, 1992.
[17] G. Nagy, Determinant preserving maps: an infinite dimensional version of a theorem of Frobenius,
Linear Multilinear Algebra 65 (2017), 351 -- 360.
[18] S. Nayak, The Hadamard determinant inequality - extensions to operators on a Hilbert space, Journal
of Functional Analysis, [DOI:10.1016/j.jfa.2017.10.009].
[19] A. Sinclair and R. Smith, Finite von Neumann Algebras and Masas, LMS Lecture Note Series (Book
351), Cambridge University Press (2008).
[20] V. Tan and F. Wang, On determinant preserver problems, Linear Algebra Appl. 369 (2003), 311 -- 317.
Bolyai Institute, Functional Analysis Research Group, University of Szeged, H-6720
Szeged, Aradi v´ertan´uk tere 1, Hungary
E-mail address: [email protected]
Smilow Center for Translational Research, University of Pennsylvania, Philadelphia,
PA 19104
E-mail address: [email protected]
URL: https://nsoum.github.io/
|
1603.00209 | 1 | 1603 | 2016-03-01T10:05:13 | Group C*-algebras without the completely bounded approximation property | [
"math.OA"
] | It is proved that:
(1) The Fourier algebra A(G) of a simple Lie group G of real rank at least 2 with finite center does not have a multiplier bounded approximate unit.
(2) The reduced C*-algebra of any lattice in a non-compact simple Lie group of real rank at least 2 with finite center does not have the completely bounded approximation property.
Hence, the results obtained by J. de Canniere and the author for SO(n,1), n at least 2, and by M. Cowling for SU(n,1) do not generalize to simple Lie groups of real rank at least 2. | math.OA | math |
GROUP C∗-ALGEBRAS WITHOUT THE COMPLETELY BOUNDED
APPROXIMATION PROPERTY
UFFE HAAGERUP
ABSTRACT. It is proved that:
(1) The Fourier algebra A(G) of a simple Lie group G of real rank at least 2 with finite
center does not have a multiplier bounded approximate unit.
(2) The reduced C ∗-algebra C ∗
r (Γ) of any lattice Γ in a non-compact simple Lie group
of real rank at least 2 with finite center does not have the completely bounded ap-
proximation property.
Hence, the results obtained by J. de Canniere and the author [dCH85] for SOe(n, 1),
n ≥ 2, and by M. Cowling [Cow83] for SU(n, 1) do not generalize to simple Lie groups
of real rank at least 2.
PREAMBLE BY ALAIN VALETTE
In spring 2015, I contacted Uffe Haagerup about his manuscript "Group
C∗-algebras without the completely bounded approximation property",
written in 1986, and never published. I mentioned that Journal of Lie
Theory might be a good place to publish it. Uffe Haagerup liked the idea
and said he was willing to update the paper after the summer.
By a sad twist of fate, Uffe Haagerup tragically passed away in July
2015. After his untimely death, Maria Ramirez Solano volunteered to
type the manuscript, and Søren Knudby accepted to write an introduc-
tion and update the bibliography. We heartily thank both of them for
their help in making the manuscript available to the community. We also
thank Søren Haagerup for giving us permission to publish his father's
paper.
INTRODUCTION
The Fourier algebra A(G) of a locally compact group, introduced by Eymard [Eym64],
consists of the matrix coefficients of the regular representation. The Fourier algebra is the
predual of the group von Neumann algebra M(G) generated by the regular representation.
The multipliers M A(G) of the Fourier algebra consists of those continuous functions ϕ
on G such that ϕψ ∈ A(G) for every ψ ∈ A(G). One identifies ϕ with the corresponding
operator mϕ on A(G), and the multiplier norm kϕkMA is the operator norm of mϕ. If the
transposed operator m∗
ϕ on M(G) is completely bounded, we say that ϕ is a completely
bounded multiplier. The space of completely bounded multipliers is denoted M0A(G),
and the completely bounded multiplier norm kϕkM0A is the completely bounded operator
norm of m∗
ϕ. We refer to the papers [Cow89, dCH85] for details.
Date: March 2, 2016.
1
GROUP C∗-ALGEBRAS WITHOUT THE COMPLETELY BOUNDED APPROXIMATION PROPERTY
2
Leptin [Lep68] showed that the Fourier algebra A(G) has an approximate unit bounded in
norm if and only if G is amenable. In [dCH85], de Cannière and the author showed that
the Fourier algebra of the non-amenable group SOe(n, 1), n ≥ 2, admits an approximate
unit bounded in the completely bounded multiplier norm. In [Cow83], Cowling obtained
similar results for SU(n, 1). In the first half of the paper, we show these results do not
generalize to simple Lie groups of real rank at least 2:
Theorem 1. The Fourier algebra A(G) of a simple Lie group G of real rank at least 2
with finite center does not have an approximate unit bounded in multiplier norm.
The second half of the paper is concerned with applications to operator algebras. It is
shown that the Fourier algebra A(Γ) of a lattice Γ in a second countable locally compact
group G has an approximate unit bounded in the completely bounded multiplier norm if
and only if the Fourier algebra A(G) of G has such an approximate unit (Theorem 2.3).
It is also shown that, for a discrete group Γ, the Fourier algebra A(Γ) has an approximate
unit bounded in the completely bounded multiplier norm if and only if the reduced group
r (Γ) has the completely bounded approximation property, if and only if the
C∗-algebra C∗
group von Neumann algebra M(Γ) has the (σ-weak) completely bounded approximation
property (Theorem 2.6). As a corollary, C∗
r (Γ) does not have the completely bounded
approximation property when Γ is a lattice in a simple Lie group of real rank at least 2
with finite center (Corollary 2.7).
The paper ends with an appendix containing a characterization, due to Bozejko and Fendler,
of completely bounded multipliers.
A preliminary version of this paper was completed in 1986 and has circulated among ex-
perts in the field. We now mention some of the developments related to this paper up until
its publication.
In 1989, Cowling and the author [CH89] showed that the Fourier algebra of a simple Lie
group with finite center and real rank 1 admits an approximate unit bounded in the com-
pletely bounded multiplier norm. This generalized the results of [Cow83, dCH85]. The
condition of finite center was subsequently removed by Hansen [Han90].
In 1996, Dorofaeff [Dor93, Dor96] removed the finite center condition from Theorem 1,
thus giving a complete characterization of simple Lie groups whose Fourier algebras ad-
mits multiplier bounded approximate units.
In 2005, Cowling, Dorofaeff, Seeger, and Wright [CDSW05] extended the previous results
to cover many non-simple Lie groups including all real algebraic linear groups.
In 2012, Ozawa gave a short proof of Theorem 1 (see [Oza12] and [Knu15]).
In 1994, a weaker approximation property (called the Approximation Property or simply
the AP) than the one considered in Theorem 1 was introduced by Kraus and the author in
[HK94]. In 2011–2013, it was shown by Lafforgue, de la Salle, de Laat and the author
[HdL13, LDlS11] that simple Lie groups G of real rank at least 2 with finite center do not
even have the AP, thus improving Theorem 1. The finite center condition was subsequently
removed in [HdL16].
1. PROOF OF THEOREM 1
Reducing the problem to SL(3, R) and Sp(2, R). In [Wan69], Wang proved that any
simple Lie group G with finite center and real rank at least 2 has Kazhdan's property (T ),
GROUP C∗-ALGEBRAS WITHOUT THE COMPLETELY BOUNDED APPROXIMATION PROPERTY
3
by using the fact that all these groups contain a closed subgroup G′ with finite center and
locally isomorphic to either SL(3, R) or SO(2, 3) (cf. [BT65, Theorem 7.2]). If G′ fails to
have a multiplier bounded approximate unit for its Fourier algebra, so does G (cf. [dCH85,
Proposition 1.12]). It is elementary to check, that if F is a finite normal subgroup of a
locally compact group H, then A(H) has a multiplier bounded approximate unit if and
only if A(H/F ) has a multiplier bounded approximate unit. Thus if G′ and G′′ are locally
isomorphic simple Lie groups with finite center, then A(G′) has a multiplier bounded
approximate unit if and only if A(G′′) has a multiplier bounded approximate unit. Hence,
to prove Theorem 1 it suffices to consider the two groups SL(3, R) and SO(2, 3) of real
rank 2. But SO(2, 3) is locally isomorphic to Sp(2, R) (cf. [Hel78, p. 519]). So we can as
well work with SL(3, R) and Sp(2, R).
Case SL(3, R). Consider the closed subgroup G0 of SL(3, R) consisting of the 3 × 3-
matrices of the form
A ∈ SL(2, R), b1, b2 ∈ R.
Note that G0 is isomorphic to the semidirect product SL(2, R)×α R2, where α is the usual
action of SL(2, R) on R2. We will show that already G0 fails to have a multiplier bounded
approximate unit. Put
A
0
0
0
0
1 ,
b1
b2
0
0
1
K0 = SO(2)
,
p1 + x2/4
z
1
0
1 x z + 1
2 xy
y
0
0
1
γ(x, y, z) = −x, −
(x, y, z) ∈ R3.
, yp1 + x2/4! .
and let N ⊆ G0 be the nilpotent group of upper triangular matrices with 1's in the diagonal
(N is the Heisenberg group). It is convenient to use the following coordinates for N :
Lemma A. Let γ be the diffeomorphism of N given by
If ϕ : N → C is the restriction of a K0-biinvariant functioneϕ on G0 to N then ϕ = ϕ ◦ γ.
Proof. Define u, v ∈ SO(2) by
1
u =
By direct computation one gets:
2
1
x
−1
2 (cid:19)
p1 + x2/4(cid:18) x
(cid:18) v
1 x z + 1
2 xy
y
0
0
1
1
0
0
1 (cid:19)
(cid:18) u 0
0
where (x′, y′, z′) = γ(x, y, z). This proves the lemma.
and
v = −u.
0
1 (cid:19) =
1 x′
1
0
0
0
z′ + 1
2 x′y′
y′
1
(cid:3)
GROUP C∗-ALGEBRAS WITHOUT THE COMPLETELY BOUNDED APPROXIMATION PROPERTY
4
Lemma B ("Failure of Fubini's Theorem"). Let ϕ ∈ C∞
integrals
J(ϕ) =ZRZR
I(ϕ) =ZRZR
y2 − z2 dz dy,
ϕ(y, z)
ϕ(y, z)
y2 − z2 dy dz
c (R2). Then the two double
exist, when the inner integrals are taken in the principal value sense (exclude symmetric
intervals around the zeroes of y2 − z2, and let the length of the intervals go to zero).
Moreover
I(ϕ) − J(ϕ) = π2ϕ(0, 0).
Proof (sketch). I, J are the distributions on R2, whose Fourier transforms are given by the
L∞-functions
bI(t, u) = π2χ{u2>t2},
bJ(t, u) = −π2χ{u2<t2},
where χ denotes the characteristic function, and D → bD is the natural extension of the
Fourier transform on R2, which on the L1-functions is given by
ei(ty+uz)f (y, z) dy dz.
bf (t, u) =ZZR2
Hence bI − bJ = π2, which is the Fourier transform of π2 times the Dirac measure at
(0, 0).
(cid:3)
Lemma C. The map
is a well-defined distribution K on R2, independent of the order of integration, and bK is
the L∞-function
ϕ ∈ C∞
c (R2)
ϕ(y, z)
1 + y2 − z2 dy dz,
ϕ 7→ZZR2
bK(t, u) =(cid:26) J0(√u2 − t2) u2 > t2
u2 < t2
0
where J0 is the zero-order Bessel function.
Proof (sketch). To compute bK, observe that K is SO(1, 1)-invariant. Hence bK can only
depend on u2−t2, so it is sufficient to compute bK(t, 0) and bK(0, u). But bK(0, u) = J0(u)
follows from the formula
J0(u) =Z 1
−1
1
√1 − x2
eiux dx.
Lemma D. Let D be the distribution on N given by
D(ϕ) =Z ∞
−∞Z ∞
−∞
ϕ(x, y, z)
(1 + x2/4)y2 − z2 dz dy dx,
−∞Z ∞
D(ϕ) ≤ 2π3kϕkA(N ) ∀ϕ ∈ C∞
c (N ).
Then
(cid:3)
ϕ ∈ C∞
c (N )
GROUP C∗-ALGEBRAS WITHOUT THE COMPLETELY BOUNDED APPROXIMATION PROPERTY
5
Proof (sketch). Note that one cannot permute the order of integrations dz, dy (cf. Lemma
B). However, it is not hard to check that dy and dx can be permuted. We have to prove that
D corresponds to an operator T ∈ M(N ), the von Neumann algebra associated to the left
regular representation of N , such that kTk ≤ 2π3 (cf. [Eym64]). The Heisenberg group
N is of type I, and the infinite dimensional irreducible representations of N are given by
(ρa)a∈R\{0} acting on L2(R) by:
(ρa(x, 0, 0)f )(t) = eiaxtf (t)
(ρa(0, y, 0)f )(t) = f (t − y)
(ρa(0, 0, z)f )(t) = eiazf (t)
where f ∈ L2(R). The remaining irreducible representations are all one-dimensional, and
they form a null set for the Plancherel measure on bN . For f, g ∈ C∞
hρa(x, y, x)f ), gi
(1 + x2/4)y2 − z2 dz dy dx
Z ∞
−∞Z ∞
−∞Z ∞
c (R) the integral
can be computed by permuting the order of integration dy, dx and applying Lemma C.
After some reduction, one finds that the integral is equal to
−∞
Since J0(x) ≤ 1 for all x ∈ R,
st < 0,
st > 0.
(1.1)
where
where
where
g(s)k(s, t)f (t) dt,
ZZR2
k(s, t) =( 2π2
s−t J0(a√−4st)
0
k(s, t) ≤ 2π2K(s, t)
K(s, t) =(cid:26) 1
s−t
st < 0,
st > 0.
0
π
2
(U H − HU ),
H(f )(s) =
1
πZ ∞
−∞
f (t)
s − t
dt
Moreover, K is the kernel of a bounded operator on L2(R) of norm ≤ π because it is the
kernel of the operator
is the Hilbert transform (unitary on L2(R)), and U is the unitary operator given by mul-
tiplication with sign(s), s ∈ R. Therefore (1.1) implies that k(s, t) is the kernel of an
integral operator on L2(R) with norm ≤ 2π3. Hence D corresponds to an operator in
M(N ) =R ⊕
Lemma E. If ϕ ∈ C∞
R\{0} ρa(N )′′ da of norm ≤ 2π3.
c (N ) and ϕ ◦ γ = ϕ then
(cid:3)
where γ is the diffeomorphism of N defined in Lemma A.
Z ∞
−∞
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
ϕ(x, 0, 0)
p1 + x2/4
dx(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ 4πkϕkA(N ),
GROUP C∗-ALGEBRAS WITHOUT THE COMPLETELY BOUNDED APPROXIMATION PROPERTY
6
Proof. Clearly,
D(ϕ ◦ γ) = Z ∞
= Z ∞
−∞Z ∞
−∞Z ∞
−∞Z ∞
−∞Z ∞
−∞
−∞
Hence, using Lemma B
, yp1 + x2/4(cid:19)
dz dy dx
ϕ(cid:18)−x,−
z√1+x2/4
(1 + x2/4)y2 − z2
z2 − (1 + x2/4)y2 dy dz dx
ϕ(x, y, z)
2D(ϕ) = D(ϕ) + D(ϕ ◦ γ)
−∞
= Z ∞
−∞Z ∞
−∞Z ∞
−Z ∞
−∞Z ∞
−∞Z ∞
= π2Z ∞
p1 + x2/4
ϕ(x, 0, 0)
−∞
−∞
ϕ(x, y, z)
(1 + x2/4)y2 − z2 dz dy dx
(1 + x2/4)y2 − z2 dy dz dx
ϕ(x, y, z)
dx.
(cid:3)
The lemma follows now from Lemma D.
Proof of Theorem 1 for SL(3, R). It is sufficient to show that the subgroup G0 does not
have a multiplier bounded approximate unit for A(G0):
Assume that there existseϕn ∈ A(G0), n ∈ N, such that supn keϕnkMA(G0) < ∞ and
for all ϕ ∈ A(G0).
Since C∞
c (G0) is dense in A(G0), we can choose the eϕn-functions in C∞
averaging by K0-elements from left and right, we can also obtain that that eϕn is K0-
biinvariant. Note that eϕn → 1 uniformly on compact subsets of G0. Put
c (G0), and by
lim
n→∞kϕeϕn − ϕkA(G0) = 0
Then
ϕn = eϕnN .
n we get
ϕn(x, 0, 0)2
kϕnkA(N ) = kϕnkMA(N ) ≤ keϕnkMA(G0).
The first equality holds because N is amenable. Thus supn kϕnkA(N ) < ∞, and by
Lemma A, ϕn = ϕn ◦ γ for all n ∈ N. By using Lemma E to ϕ2
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
dx(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ 4πkϕnk2
Z ∞
4πZ ∞
p1 + x2/4
p1 + x2/4
This gives a contradiction.
Thus by Fatou's Lemma,
n→∞ kϕnk2
lim inf
A(N ) ≥
= ∞.
A(N ).
dx
−∞
−∞
(cid:3)
1
GROUP C∗-ALGEBRAS WITHOUT THE COMPLETELY BOUNDED APPROXIMATION PROPERTY
7
Case Sp(2, R). The group Sp(2, R) is the set of g ∈ GL(4, R) that leaves invariant the
exterior form
x1x3 − x3x1 + x2x4 − x4x2,
(cf. [Hel78, p. 445]). For our purpose, it is convenient to permute the third and fourth
coordinate in R4, so we will instead consider the group G ∼= Sp(2, R) of invertible 4 × 4-
matrices that leave invariant the form:
x1x4 − x4x1 + x2x3 − x3x2.
G is a 10 dimensional connected Lie group with Lie algebra
a11 + a44 = a22 + a33 = a12 + a34 = a21 + a43 = 0,
a13 = a24, a31 = a42
(cid:27) .
Let g0 ⊆ g be the Lie algebra
g =(cid:26)(aij)i,j=1,...,4(cid:12)(cid:12)(cid:12)(cid:12)
g0 =
0 −c2
0 −b2
b3
0
0
0
c1
b1
b2
0
c3
c1
c2
0
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
b1, b2, b3, c1, c2, c3 ∈ R
.
Then g0 = g1⊕sg2 (semidirect sum), where g1 ∼= sl(2, R) is the Lie algebra corresponding
to c1 = c2 = c3 = 0, and g2 is the Lie algebra corresponding to b1 = b2 = b3 = 0. (Note
that g2 is isomorphic to the Lie algebra of the three dimensional Heisenberg group). Hence
exp(g0) generates a closed subgroup G0 of G, namely, the semidirect product of
and
c1
0
1
0
where the action α : G1 → Aut(G2) is given by
1 −c2
1
0
0
0
0
0
G1 =
G2 = exp(g2) =
α(a)
c1
c2
c3
1
0
0
0
0 0
a
0 0
0
0
0
1
= a
K0 =
1 0
0
0
0 0
0 0
SO(2)
0 0
0
0
0 1
,
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
,
c3
c1
c2
1
a ∈ SL(2, R)
c1, c2, c3 ∈ R
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
.
1
c1
c2
c3
0
0
Consider next the compact subgroup K0 of G0 given by
GROUP C∗-ALGEBRAS WITHOUT THE COMPLETELY BOUNDED APPROXIMATION PROPERTY
8
.
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
z
1 −y
1
0
0
0
0
0
(x, y, z, w) ∈ R4.
w
z + 1
2 xy
y
1
z − 1
2 xy
x
1
0
c3 − 1
6 b1c2
2
c1 + 1
2 b1c2
c2
1
Lemma F. Let γ′ be the diffeomorphism of N given by
b1, c1, c2, c3 ∈ R
The group N is isomorphic to the group Γ4 considered by Dixmier in [Dix58].
In the rest of this section, we will use the following coordinates for N :
and the nilpotent subgroup of G0 given by
0 −c2
c1
c3
0
0
b1
c1
0
0
0
c2
0
0
0
0
c1 − 1
2 b1c2
b1
1
0
1 −c2
1
0
0
0
0
0
N = exp
=
n(x, y, z, w) =
γ′(x, y, z, w) = −x,−
b1, c1, c2, c3 ∈ R
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
,
, yp1 + x2/4, w! .
If ϕ : N → C is the restriction of a K0-biinvariant functioneϕ on G0 to N , then ϕ = ϕ◦ γ′.
= n(γ′(x, y, z, w)).
D′(ϕ) =Z ∞
p1 + x2/4
Proof. Let u, v ∈ SO(2) be as in Lemma A. Then
1 0
0
0
0 0
0 0
0
0
0 1
n(x, y, z, w)
−∞Z ∞
−∞Z ∞
D′(ϕ) ≤ 2π3kϕkA(N )
−∞
(1 + x2/4)y2 − z2 dz dy dx,
for all ϕ ∈ C∞
Lemma G. Let D′ be the distribution on N given by
1 0
0
0
0 0
u
v
0 0
0
0
0 1
ϕ ∈ C∞
c (N ).
ϕ(x, y, z, 0)
(cid:3)
Then
Proof (sketch). N is the semidirect product of the abelian subgroup
c (N ).
isomorphic to R3, and the one-parameter group
1
0
0
0
B =
C =
0
z w
1 x z
0
0
0
1
1
0
1 −y
1
0
0
0
0
0
0
0
0
0
1 y
1
0
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
,
(x, z, w) ∈ R3
y ∈ R
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
,
GROUP C∗-ALGEBRAS WITHOUT THE COMPLETELY BOUNDED APPROXIMATION PROPERTY
9
i.e. N ∼= R3 ×θ R where the action θ is given by
Let (s, u, v) be the dual coordinates to (x, z, w). The transpose of θy is
θy(x, z, w) = (x, z − yx, w − 2yz + y2x).
bθy(s, u, v) = (s − yu + y2v, u − 2yv, v).
[Mac52], [Dix58]): The orbits forbθy are
(v 6= 0),
(v = 0, u 6= 0),
(v = 0, u = 0).
(1) parabolas
(2) straight lines
(3) single points
The irreducible representation of N can now be obtained by the Mackey machine (cf.
Since {(s, u, v) v = 0} is a null set in R3, the first type of orbits gives sufficiently many
irreducible representations to disintegrate the regular representation of N . Let ρa,b be
the irreducible representation coming from the parabolic orbit with vertex (b, 0, a), a ∈
R\{0}, b ∈ R. Then, ρa,b can be realized on L2(R) as follows:
(ρa,b(x, 0, 0, 0))(t) = ei(at2+b)xf (t)
(ρa,b(0, y, 0, 0))(t) = f (t − y)
(ρa,b(0, 0, z, 0))(t) = ei2atzf (t)
(ρa,b(0, 0, 0, w))(t) = eiawf (t),
where f ∈ L2(R). A computation similar to the one in Lemma D gives now,
where
−∞
−∞Z ∞
−∞Z ∞
Z ∞
k(s, t) =( 2π2
hρa,b(x, y, z, 0)f, gi
(1 + x2/4)y2 − z2 dz dy dx =ZZR2
s−t J0(p−(as2 + b)(at2 + b))
0
If ab ≥ 0 then k(s, t) = 0 almost everywhere in R2. If ab < 0 we put c =p−b/a. Then
where K(s, t) = 1
operator on L2(R) of norm ≤ π, namely, the operator
s−t χ{(s2−c2)(t2−c2)<0}. But K(s, t) is the kernel of a bounded integral
k(s, t) ≤ 2π2K(s, t),
g(s)k(s, t)f (t) ds dt,
(as2 + b)(at2 + b) < 0
(as2 + b)(at2 + b) > 0.
π
2
(U2HU1 − U1HU2),
where H is the Hilbert transform, and U1 and U2 are the unitary multiplication operators
on L2(R) given by the functions sign(t + c) and sign(t − c), respectively. Hence k(s, t) is
the kernel of a bounded integral operator of norm ≤ 2π3. Since
ρa,b(N )′′ da db,
M(N ) =ZZ ⊕
a6=0,b∈R
it follows that D′ corresponds to an element in M(N ) ∼= A(N )∗ of norm ≤ 2π3.
Proof of Theorem 1 for Sp(2, R). Exactly as in the proof of Lemma E, we get that if ϕ ∈
c (N ) and ϕ ◦ γ′ = ϕ, then
C∞
Z ∞
ϕ(x, 0, 0, 0)
−∞
(cid:3)
p1 + x2/4
dx(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ 4πkϕkA(N ),
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
GROUP C∗-ALGEBRAS WITHOUT THE COMPLETELY BOUNDED APPROXIMATION PROPERTY
10
and the proof for Sp(2, R) can be completed by the same arguments as we used for
SL(3, R).
(cid:3)
Remark. It follows from the above proofs that
R2 ×s SL(2, R)
and
HG ×s SL(2, R)
(the semidirect product of the Heisenberg group HG by SL(2, R) described above) both
fail to have a multiplier bounded approximate unit for their Fourier algebras, although the
Fourier algebras of R2, HG and SL(2, R) all have multiplier bounded approximate units.
(R2 and HG are amenable, for SL(2, R) see [dCH85]).
2. RESULTS ABOUT LATTICES IN LIE GROUPS
A lattice Γ in a locally compact group G is a closed discrete subgroup, for which G/Γ has a
bounded G-invariant measure. A locally compact group that admits a lattice is necessarily
unimodular (cf. Definition 1.8 and Remark 1.9 in [Rag72]).
In the following, Γ denotes a lattice in a second countable locally compact group G. In this
case, Γ is countable and the quotient map ρ : G → G/Γ has a a Borel cross section. Let Ω
be the range of a Borel cross section. Then
G = [γ∈Γ
Ωγ
(disjoint union).
Let µ be the Haar measure on G. Since Γ is countable, µ(Ω) > 0. Moreover, the quotient
map ρ is a bijection of Ω onto G/Γ, that carries µΩ onto a G-invariant measure on G/Γ.
Thus by the assumption that Γ is a lattice, µ(Ω) < ∞. In the following we will assume
that µ is normalized such that
µ(Ω) = 1.
Let µΓ be the counting measure on Γ. For every bounded function ϕ on Γ, the function
is a well-defined bounded continuous function on G because χΩ ∈ L1(G), G is unimodu-
lar, and because χΩ ∗ ϕµΓ is the bounded function given by
bϕ = χΩ ∗ ϕµΓ ∗eχΩ
Lemma 2.1.
(χΩ ∗ ϕµΓ)(ωγ) = ϕ(γ),
γ ∈ Γ, ω ∈ Ω.
kbϕkA(G) ≤ kϕkA(Γ)
kbϕkM0A(G) ≤ kϕkM0A(Γ).
(1) If ϕ ∈ A(Γ) then bϕ ∈ A(G) and
(2) If ϕ ∈ M0A(Γ) thenbϕ ∈ M0A(G) and
(1) If ϕ ∈ A(Γ), then there exists f, g ∈ ℓ2(Γ) such that ϕ = f ∗eg, kfk2kgk2 = kϕkA(Γ).
kϕkA(Γ), and f1 ∗eg1 = bϕ. This proves (1).
Put f1 = χΩ ∗ f µΓ, g1 = χΩ ∗ gµΓ. Then f1, g1 ∈ L2(G), kf1k2kg1k2 = kfk2kgk2 =
(2) Every z ∈ G has a unique decomposition
Proof.
z = ωγ,
ω ∈ Ω, γ ∈ Γ.
GROUP C∗-ALGEBRAS WITHOUT THE COMPLETELY BOUNDED APPROXIMATION PROPERTY
11
The γ-part in the decomposition will be denoted by γ(z). For x ∈ G, we let τx : Ω → Ω be
the map given by τx(ω) = ω′, where ω′ is the Ω-component of xw in the decomposition
xω = ω′γ′,
ω′ ∈ Ω, γ′ ∈ Γ.
Each τx is a Borel isomorphism of Ω because τx corresponds to left translation by x in
G/Γ with the Borel isomorphism Ω → G/Γ given by the quotient map. Since the latter
Borel isomorphism carries µΩ to an invariant measure on G/Γ, it follows that µΩ is τx-
invariant.
We rewrite the functionbϕ = χΩ ∗ ϕµΓ ∗eχΩ in a suitable way. Observe first that
(χΩ ∗ ϕµΓ)(ωγ) = ϕ(γ),
ω ∈ Ω, γ ∈ Γ,
or equivalently,
Therefore,
(χΩ ∗ ϕµΓ)(x) = ϕ(γ(x)),
x ∈ G.
bϕ(x) =ZG
=ZxΩ
=ZΩ
ϕ(γ(y))χΩ(x−1y) dµ(y)
ϕ(γ(y)) dy
ϕ(γ(xω)) dµ(ω).
For x, y ∈ G and ω ∈ Ω:
Thus,
xω = τx(ω)γ(xω)
yω = τy(ω)γ(yω).
yx−1τx(ω) = yω(xω)−1τx(ω)
= τy(ω)γ(yω)γ(xω)−1.
Since τx(ω) ∈ Ω and γ(yω)γ(xω)−1 ∈ Γ, it follows that
γ(yx−1τx(ω)) = γ(yω)γ(xω)−1.
Hence
ZΩ
ϕ(γ(yω)γ(xω)−1) dµ(ω) =ZΩ
However, since dµ(ω) is invariant under τx, the latter integral is equal to
ϕ(γ(yx−1τx(ω))) dµ(ω).
(2.1)
Hence, ∀x, y ∈ G:
ZΩ
ϕ(γ(yx−1ω)) dµ(ω) = bϕ(yx−1).
bϕ(yx−1) =ZΩ
ϕ(γ(yω)γ(xω)−1) dµ(ω).
We can now apply Bozejko-Fendler's result that M0A(G) coincides with the space of
Herz-Shur multipliers on G with same norm [BF84]: Since Γ is discrete, it implies, that
there exist a Hilbert space H and bounded maps ξ, η from Γ to H such that
and
ϕ(γ2γ−1
1 ) = hξ(γ1), η(γ2)i,
γ1, γ2 ∈ Γ,
kξk∞kηk∞ = kϕkM0A(Γ).
GROUP C∗-ALGEBRAS WITHOUT THE COMPLETELY BOUNDED APPROXIMATION PROPERTY
12
This follows from Gilbert's characterization of Herz-Schur multipliers [Gil]. Define now
sup
bξ(x)(ω) = ξ(γ(xω))
bη(x)(ω) = η(γ(xω)).
bξ,bη : Γ → L2(Ω, H, dµ) by
Thenbξ,bη are bounded Borel functions from G to L2(Ω, H, dµ) and
x∈Gkbξ(x)k2 ≤ kξk∞,
x∈Gkbη(x)k2 ≤ kηk∞,
bϕ(yx−1) =ZΩhξ(γ(xω)), η(γ(yω))i dµ(ω) = hbξ(x),bη(y)i.
kbϕkM0A(G) ≤ kbξk∞kbηk∞ ≤ kξk∞kηk∞ = kϕkM0A(Γ),
Since bϕ is continuous, it implies that bϕ ∈ M0A(G) and
and by (2.1),
sup
(cf. proof of [Cow83, Proposition 1.1] or [Gil]). This proves (2).
Lemma 2.2. Let G be a locally compact group and let k ≥ 1. Then the following condi-
tions are equivalent:
(cid:3)
σ(L∞, L1)-topology.
(1) There exists a net (ϕα) in A(G) such that supα kϕαkM0A(G) ≤ k and ϕα → 1 in
(2) There exists a net (ϕα) in A(G) such that supα kϕαkM0A(G) ≤ k and ϕα → 1
(3) There exists an approximate unit (ϕα) for A(G) such that supα kϕαkM0A(G) ≤ k.
uniformly on compact sets.
(By an approximate unit in A(G), we just mean a net (ϕα) such that
∀ψ ∈ A(G).
lim
α kϕαψ − ψkA(G) = 0
The net (ϕα) will in general be unbounded in A(G)-norm.)
Proof. (3) =⇒ (2) =⇒ (1) is clear.
(1) =⇒ (2): Assume that (ϕα) satisfies (1) and put
ϕ′
α = h ∗ ϕα,
where h ∈ Cc(G)+,R h dµ = 1. Then
α(x) =ZG
Let K ⊆ G be compact. Then the functions h(x · ), x ∈ K form a compact subset of
L1(G). Since ϕα → 1 in σ(L∞, L1), and since supα kϕαk∞ ≤ k, the convergence is
uniform on compact subsets of L1(G). Hence,
h(xy)ϕα(y−1)dy.
ϕ′
ϕ′
α(x) = h qϕα, h(x · )i
α is contained
converges to h1, h(x · )i = (h ∗ 1)(x) = 1 uniformly on K. Moreover, ϕ′
in the σ(L∞, L1)-closed convex hull of left translates of ϕα, and since the unit ball in
M0A(G) is σ(L∞, L1)-closed (cf. [dCH85]), we have
kϕ′
αkM0A(G) ≤ kϕαkM0A(G) ≤ k.
GROUP C∗-ALGEBRAS WITHOUT THE COMPLETELY BOUNDED APPROXIMATION PROPERTY
13
(2) =⇒ (3): Assume that (ϕα) satisfies (2). Choose h ∈ Cc(G)+,RG h dµ = 1, and put
αkM0A(G) ≤ k for all α. Let ψ ∈ A(G)∩ Cc(G), and
α = h∗ ϕα. As above, we have kϕ′
ϕ′
put
K = supp(h),
L = supp(ψ).
Moreover, set gα = ϕαχK−1L (where χE is the characteristic function of a set E). Then
for x ∈ L,
h(y)ϕα(y−1x) dy = (h ∗ gα)(x)
(2.2)
(h ∗ ϕα)(x) =ZK
Similarly,
(2.3)
By the assumption on ϕα, gα → χK−1L uniformly on K −1L. Moreover, gα vanishes
outside the compact set K −1L. Hence,
(h ∗ 1)(x) = (h ∗ χK−1L)(x),
x ∈ L.
ega → eχK−1L,
and since h ∈ L2(G), this implies that
h ∗ gα → h ∗ χK−1L
Hence also
in L2(G)-norm,
in A(G)-norm.
which by (2.2) and (2.3) is equivalent to
(h ∗ gα)ψ → (h ∗ χK−1L)ψ
in A(G)-norm,
ϕ′
αψ = (h ∗ ϕα)ψ → (h ∗ 1)ψ = ψ
in A(G)-norm.
Since A(G) ∩ Cc(G) is dense in A(G) and since
α kϕ′
αkMA(G) ≤ sup
α kϕ′
sup
αkM0A(G) < ∞,
it follows that
kϕ′
αψ − ψkA(G) → 0,
for all ψ ∈ A(G).
(cid:3)
Remark. The proof of (2) =⇒ (3) in Lemma 2.2 is due to Michael Cowling (private
communication). It substitutes a previous proof of ours, that was valid only for Lie groups.
Theorem 2.3. Let Γ be a lattice in a second countable locally compact group G, and let
k ∈ [1,∞[. then the following conditions are equivalent.
(1) A(G) has an approximate unit (ϕα), such that kϕαkM0A(G) ≤ k, for all α.
(2) A(Γ) has an approximate unit (ψα), such that kψαkM0A(Γ) ≤ k for all α.
Proof. (1) =⇒ (2) follows from [dCH85], and is valid for any closed subgroup Γ ⊆ G.
(2) =⇒ (1): Assume that (ψα) ⊆ A(Γ) satisfies (2), and put bψα = χΩ ∗ ψαµΓ∗eχΩ, where
Ω is chosen as in Lemma 2.1. Then bψα ∈ A(G) by Lemma 2.1. Since Γ is discrete, the
net (ψα) is uniformly bounded, and since ψα → 1 pointwise in Γ, also
It is easy to check that the map ϕ 7→ bϕ = χΩ ∗ ϕµΓ ∗ eχΩ from ℓ∞(Γ) to L∞(G) is
the transpose of a bounded map from L1(G) to ℓ1(Γ). Hence, bψα → 1G in σ(L∞, L1)-
topology. Moreover, kbψαkM0A(G) ≤ kψαkMA(Γ) by Lemma 2.1. Hence (1) =⇒ (2)
follows from (1) ⇐⇒ (3) in Lemma 2.2.
σ(ℓ∞(Γ), ℓ1(Γ)).
ψα → 1Γ,
(cid:3)
GROUP C∗-ALGEBRAS WITHOUT THE COMPLETELY BOUNDED APPROXIMATION PROPERTY
14
Corollary 2.4. Every lattice Γ in a simple Lie group of real rank at least 2 fails to have a
complete multiplier bounded approximate unit.
Proof. Follows from Theorem 2.3 and Theorem 1.
(cid:3)
Let Γ be a discrete group, and let M(Γ) be the von Neumann algebra associated with the
left regular representation of Γ on ℓ2(Γ), and let
δx(y) =(cid:26) 1, x = y
0, x 6= y
be the standard basis in ℓ2(Γ). Put
Then tr is a normal faithful trace on M(Γ).
tr(a) = haδe, δei,
a ∈ M(Γ).
Lemma 2.5. Let T be a bounded map from C∗
to itself. Let
r (Γ) to itself or a bounded map from M(Γ)
ϕT (x) = tr(λ(x)∗T λ(x)),
x ∈ Γ.
(1) If T is completely bounded, then ϕT ∈ M0A(G), and
kϕTkM0A(Γ) ≤ kTkCB.
(2) If T is of finite rank, then ϕT ∈ ℓ2(Γ).
Proof. (1): Since λ ⊗ λ is unitarily equivalent to λ ⊗ 1ℓ2(Γ) where 1ℓ2(Γ) is the trivial
[Dix77, 13.11.3] and [Fel62]), there exists a normal
representation of Γ on ℓ2(Γ) (cf.
*-isomorphism π of M = M(Γ) onto a von Neumann subalgebra N of M ⊗ M such that
π(λ(x)) = λ(x) ⊗ λ(x),
x ∈ G.
Let ε be the normal conditional expectation of M⊗ M onto N that leaves tr⊗ tr invariant.
Then, since ε is orthogonal with respect to the inner product given by tr ⊗ tr, one gets
easily that
ε(λ(x) ⊗ λ(y)) =(cid:26) λ(x) ⊗ λ(x), x = y,
x 6= y.
0,
Put ρ = π−1ε. Since
one has
and
norm
C∗
r (Γ) = span{λ(g) g ∈ Γ}
r (Γ) ⊗ C∗
π(C∗
r (Γ)) ⊆ C∗
r (Γ) into itself,
r (Γ)) ⊆ C∗
r (Γ) ⊗ C∗
ρ(C∗
r (Γ),
r (Γ).
Hence, if T is completely bounded of C∗
S = ρ ◦ (T ⊗ idC ∗
is a well-defined completely bounded map on C∗
r (Γ)) ◦ π
r (G) and kSkCB ≤ kTkCB. Now,
cx,yλ(y),
where cx,y ∈ C and the sum is convergent in the k k2-norm associated with tr. Hence
(T ⊗ idC ∗
cx,yλ(y) ⊗ λ(x).
T λ(x) =Xy∈G
r (Γ))π(λ(x)) =Xy∈G
GROUP C∗-ALGEBRAS WITHOUT THE COMPLETELY BOUNDED APPROXIMATION PROPERTY
15
By the orthogonality property of ε we have
Hence
But
ε(T ⊗ idC ∗
r (Γ))π(λ(x)) = cx,xλ(x) ⊗ λ(x).
Sλ(x) = cx,xλ(x),
x ∈ G.
cx,x = hT λ(x), λ(x)itr = ϕT (x).
ϕT , and hence ϕT ∈ M0A(Γ)
This shows that (with the notation of [dCH85]) S = m∗
and kϕTkM0A(G) = kSkCB ≤ kTkCB, because π, π−1 and ε are completely positive. If
instead T is a completely bounded map on M(Γ), then we let,
S = ρ ◦ (T ⊗ idM(Γ)) ◦ π,
and the same proof applies (cf. [dCH85, Section 1]).
(2): It is sufficient to consider the rank one case: A rank one map on C∗
r (Γ) is of the form
T (a) = f (a)b
Thus
where f ∈ C∗
r (Γ)∗, b ∈ C∗
r (Γ).
ϕT (x) = tr(λ(x)∗b)f (λ(x)).
But x 7→ tr(λ(x)∗b) is in ℓ2(Γ), because (λ(x))x∈Γ is an orthonormal family in L2(C∗
and x 7→ f (λ(x)) is bounded. This proves (2) in the C∗
by the same arguments.
Theorem 2.6. Let Γ be a discrete group and let k ≥ 1. Then the following three conditions
are equivalent.
r (Γ)-case. The M(Γ)-case follows
(cid:3)
r (Γ), tr),
(1) A(Γ) has an approximate unit (ϕα) , such that kϕαkM0A(Γ) ≤ k for all α.
(2) There exists a net (Tα) of finite rank operators on C∗
r (Γ) such that kTαkCB ≤ k
for all α, and such that
kTαx − xk → 0
for all x ∈ C∗
r (Γ).
(3) There exists a net (Tα) of σ-weakly continuous finite rank maps on M(Γ), such
that kTαkCB ≤ k for all α and
hϕ, Tαx − xi → 0,
for all x ∈ M(Γ), ϕ ∈ M(Γ)∗.
Proof. (1) =⇒ (2) and (1) =⇒ (3) follows from [dCH85].
(2) =⇒ (1): Assume Tα are finite rank operators of C∗
k and
r (Γ) into itself, such that supα kTαkCB ≤
Let ϕα = ϕTα be as defined in the preceding lemma. Then ϕα ∈ ℓ2(Γ) ⊆ A(Γ) and
kTαa − ak → 0
for all a ∈ C∗
r (Γ).
sup
α kϕαkM0A(Γ) < ∞,
and since ϕα(x) = tr(λ(x)∗Tα(λ(x))) → 1 for all x ∈ Γ:
kϕαψ − ψkA(Γ) → 0,
(2.4)
for all ψ ∈ A(Γ) with finite support, i.e. for all ψ ∈ A(Γ) ∩ Cc(Γ) which form a dense
subset of A(Γ). Using that kϕαkMA(Γ) ≤ kϕαkM0A(Γ) ≤ k for all α, one gets that (2.4)
holds for all ψ ∈ A(Γ).
GROUP C∗-ALGEBRAS WITHOUT THE COMPLETELY BOUNDED APPROXIMATION PROPERTY
16
(3) =⇒ (1): Let (Tα) be a net satisfying the conditions in (3). Since the functional,
is in M(Γ)∗ (x fixed) we have that
a 7→ tr(λ(x)∗a),
a ∈ M(Γ)
tr(λ(x)∗Tα(λ(x))) → 1,
∀x ∈ G.
The proof can now be completed exactly as in (2) =⇒ (1) by use of Lemma 2.5.
Corollary 2.7. Let Γ be a lattice in a simple Lie group of real rank at least 2 with finite
center. Then
(cid:3)
r (Γ) does not have the completely bounded approximation property.
(1) C∗
(2) M(Γ) does not have the (σ-weak) completely bounded approximation property.
3. APPENDIX: ON COMPLETELY BOUNDED MULTIPLIERS AND
HERZ-SHUR-MULTIPLIERS.
In Section 2, we used Bozejko and Fendler's results [BF84] that M0A(G) coincides iso-
metrically with the space B2(G) introduced by Herz in [Her74]. Their result relies heavily
on a characterization of B2(G) found by Gilbert ([Gil, Theorem 4.7]). However, Gilbert's
paper, has never been published, so for the convenience of the reader, we give below a
self-contained proof of the result needed in Section 2.
Let a∗ b denote the Schur product (a∗ b)ij = aij bij of complex n× n-matrices in Mn(C),
the operator norm of a when Mn(C) acts
and let kak be the C∗-norm on Mn(C), i.e.
on the Euclidean space ℓ2({1, . . . , n}). We let k kS denote the Schur multiplier norm on
Mn(C), i.e.
kakS = sup{ka ∗ bk b ∈ Mn(C), kbk ≤ 1}.
Let ≤ be the usual ordering in Mn(C) as a C∗-algebra, i.e. a ≥ 0 iff a = a∗ and all
eigenvalues of a are nonnegative. It is well known that
Hence, if a ≥ 0 then the operator Ma : b 7→ a ∗ b is positive on the C∗-algebra Mn(C),
and therefore,
a ≥ 0, b ≥ 0 =⇒ a ∗ b ≥ 0.
kakS = kMa(1)k =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
a1
0
...
0
an
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
= max{a1, . . . , an}
for every a ∈ Mn(C)+ (see for instance [RD66, Corollary 1] or [KR97, 10.5.10]). The
following lemma is a special case of [Gil, Theorem 4.7]. The proof given here is inspired
by [Pau84].
Lemma 3.1 ([Gil]). Let a ∈ Mn(C). The following three conditions are equivalent:
(1) kakS ≤ 1.
(2) ∃b, c ∈ Mn(C)+ such that(cid:18) b a∗
c (cid:19) ≥ 0 and bii ≤ 1 cii ≤ 1, i = 1, . . . , n.
(3) There exist a Hilbert space H and 2n vectors ξ1, . . . , ξn, η1, . . . , ηn in the closed
a
unit ball of H such that
aij = hξi, ηji,
i, j = 1, . . . , n.
GROUP C∗-ALGEBRAS WITHOUT THE COMPLETELY BOUNDED APPROXIMATION PROPERTY
17
Proof. (1) =⇒ (2): Consider the real subspace E of M2n(C)s.a. consisting of the matrices
of the form
x
(cid:18) y x∗
z (cid:19)
(yi + zi) + 2Re
nXi,j=1
where y, z are self-adjoint diagonal matrices y = diag(y1, . . . , yn), z = diag(z1, . . . , zn).
Let ϕ : E → C be the linear form given by
z (cid:19) =
ϕ(cid:18) y x∗
We will prove that ϕ ≥ 0, i.e. ϕ(w) is non-negative on positive hermitian matrices in E.
Note that 12n ∈ E. Since
nXi=1
xij aij .
(3.1)
x
ϕ(w) ≥ 0 ⇐⇒ ϕ(w + ε12n) ≥ 0,
∀ε > 0,
it is sufficient to check that ϕ(w) ≥ 0 for all strictly positive
w =(cid:18) y x∗
z (cid:19)
x
i.e. those w ∈ E+ for which yi > 0, zi > 0, i = 1, . . . , n. However, w ≥ 0 implies that
(cid:18)
1n
y−1/2x∗z−1/2
z−1/2xy−1/2
1n
(cid:19) ≥ 0
which is equivalent to kz−1/2xy−1/2k ≤ 1 (C∗-norm). Since kakS ≤ 1, the matrix,
e = (aij z−1/2
i
xij y−1/2
j
)i,j=1,...,n
has also C∗-norm ≤ 1. Hence
which implies that
eij y1/2
j
nXi,j=1
=(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
xij aij(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
nXi,j=1
ϕ(cid:18) y x∗
z (cid:19) ≥
nXj=1
x
yj +
nXi=1
zi!1/2
z1/2
i
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
≤ nXj=1
zi − 2 nXj=1
yj!1/2 nXi=1
zi!1/2
yj!1/2 nXi=1
≥ 0.
Hence ϕ is a positive functional on E ⊆ M2n(C).
Let eϕ be a Hahn-Banach extension of ϕ to M2n(C)s.a.. Then eϕ ≥ 0, because keϕk =
kϕk = ϕ(12n), and because
Hence, there exists a positive hermitian matrix
0 ≤ w ≤ 12n =⇒ eϕ(12n − w) ≤ keϕk =⇒ eϕ(w) ≥ 0.
such that
(dij )i,j=1,...,2n
2nXi,j=1
eϕ(w) =
dij wij ,
w ∈ M2n(C)s.a..
GROUP C∗-ALGEBRAS WITHOUT THE COMPLETELY BOUNDED APPROXIMATION PROPERTY
18
x
then by (3.1)
x ∈ Mn(C).
Therefore dii = 1, i = 1, . . . , 2n. Moreover, if w is of the form
For diagonal matrices w = diag(y1, . . . , yn, z1, . . . , zn), eϕ(w) =Pn
w =(cid:18) 0 x∗
0 (cid:19) ,
di,n+j(x∗)ij = 2Re
nXi,j=1
nXi,j=1
for all (xij ) ∈ Mn(C). Hence aij = dn+i,j, i.e.
d =(cid:18) b a∗
c (cid:19)
aij xij =
2Re
nXi,j=1
dn+i,jxij +
a
where bii = cii = 1, i = 1, . . . , n. This proves (2).
(2) =⇒ (3): Let
i=1(yi + zi) by (3.1).
nXi,j=1
dn+i,j xij
a
d =(cid:18) b a∗
c (cid:19)
dij =Xk
fikfjk
d =(cid:18) b a∗
c (cid:19) .
a
be as in (2). Let H = ℓ2({1, . . . , 2n}), and let η1, . . . , ηn, ξ1, . . . , ξn be the row vectors of
the operator f = d1/2. Since
we get aij = hξi, ηji, and kξik2 = dn+i,n+i ≤ 1, kηik2 = dii ≤ 1, for i = 1 . . . , n.
(3) =⇒ (2): Let ξi, ηj be as in (3), put
bij = hηi, ηji,
cij = hξi, ξji
i, j = 1 . . . , n,
and let
If we let ηn+i = ξi, i = 1, . . . , n. Then
dij = hηi, ηji,
i, j = 1, . . . , 2n.
2
≥ 0.
2nXi,j=1
This implies that d is positive hermitian, because, for λ1, . . . , λ2n ∈ C,
(2) =⇒ (1): If
dij λiλj =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
λiηi(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
2nXi=1
d =(cid:18) b a∗
c (cid:19) ≥ 0
and bii ≤ 1, cii ≤ 1, then the map e 7→ d ∗ e, e ∈ M2n(C) is positive and kd ∗ ek ≤
max1≤i≤2n diikek. Taking e of the form e = (cid:18) 0
x 0 (cid:19), x ∈ Mn(C), we see that
ka ∗ xk ≤ kxk, x ∈ Mn(C). This proves (2) =⇒ (1).
Theorem 3.2 ([BF84]). Let ϕ be a continuous function on a locally compact group G, and
let k ≥ 0. Then the following conditions are equivalent.
(cid:3)
a
0
(1) ϕ ∈ M0A(G) and kϕkM0A(G) ≤ k.
GROUP C∗-ALGEBRAS WITHOUT THE COMPLETELY BOUNDED APPROXIMATION PROPERTY
19
(2) For every finite set x1, . . . , xn in G,
kϕ(x−1
j xi)i,j=1,...,nkS ≤ k
(3) There exist a Hilbert space H and two bounded maps ξ, η : G → H such that
ϕ(y−1x) = hξ(x), η(y)i,
∀x, y ∈ G
and
(sup
x∈G kξ(x)k)(sup
y∈Gkη(y)k) ≤ k.
(4) There exist a Hilbert space H and two bounded maps ξ, η : G → H as in (3) with
the additional property that the coordinate functions ξi and ηi (with respect to
any orthonormal basis (ei)i∈I of H) are continuous and the families {ξi}i∈I and
{ηi}i∈I are locally countable.
i,j=1 be the matrix units
It is elementary to check that f ∗
spani,j{fij} is a subalgebra of M(G)⊗Mn(C) ∗-isomorphic to Mn(C). Since ∗-isomorphisms
preserve norms,
i=1 fii = 1, so N =
fij = λ(x−1
Proof. (1) =⇒ (2) (cf. [BF84]): Let x1, . . . , xn ∈ G and let (eij)n
in Mn(C). Put
i xj ) ⊗ eij ∈ M(G) ⊗ Mn(C).
ij = fji, fijfkl = δjkfil andPn
=(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
αij fij(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
nXi,j=1
αij ∈ C.
Let Mϕ : M(G) → M(G) be as in [dCH85, Section 1]. Then by (1),
αij eij(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
nXi,j=1
,
where in is the identity on Mn(C). But,
kMϕ ⊗ ink ≤ k,
Therefore
and hence for aij ∈ C,
Mϕλ(x) = ϕ(x)λ(x),
x ∈ G.
(Mϕ ⊗ in)fij = ϕ(x−1
i xj )fij,
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
nXi,j=1
ϕ(x−1
i xj)αij fij(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
≤ k(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
nXi,j=1
αijfij(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
.
Since fij in this inequality can be exchanged by the standard matrix units eij of Mn(C), it
follows that the matrix (ϕ(x−1
i xj ))i,j=1,...,n as well as its transpose has Schur multiplier
norm ≤ k.
(2) =⇒ (3): Let (F ,⊆) be the family of all finite subsets of G ordered by inclusion.
Assuming (2), we can for each F ∈ F find a Hilbert space HF and bounded maps
with
and
ξF , ηF : F → HF .
x∈F kξF (x)k ≤ k1/2,
sup
x∈F kηF (x)k ≤ k1/2,
sup
ϕ(y−1x) = hξ(x), η(y)i,
x, y ∈ F.
(3) follows now easily by a standard ultraproduct argument: Let namely U be a cofinal
ultrafilter on (F ,⊆) (i.e. an ultrafilter that contains all sets of the form {F ′ ∈ F F ′ ⊇
GROUP C∗-ALGEBRAS WITHOUT THE COMPLETELY BOUNDED APPROXIMATION PROPERTY
20
F}, where F ∈ F), and let HU be the ultraproduct of (HF )F ∈F corresponding to U. For
x ∈ G, we let ξ(x), and η(x) be the elements in HU with representing sequences
ξ(x) = (ξF (x))F ∈F
η(x) = (ηF (x))F ∈F
where we set ξF (x) = ηF (x) = 0 if x 6∈ F . Since for fixed x, y ∈ G, x, y ∈ F eventually,
we have
Moreover,
hξ(x), η(y)i = lim
U hξF (x), ηF (y)i = ϕ(y−1x),
x, y ∈ G.
x∈Gkξ(x)k ≤ k1/2,
sup
x∈Gkη(x)k ≤ k1/2,
sup
(3) =⇒ (4): Let ξ, η : G → H satisfy the conditions in (3). Let ξ′(x) = P ξ(x), where P
is the orthogonal projection on the closed linear span of
{η(x) x ∈ G},
and put η′(y) = Qη(y), where Q is the orthogonal projection on the closed linear span of
{ξ′(x) x ∈ G}. Then
ϕ(y−1x) = hξ′(x), η(y)i = hξ′(x), η′(y)i,
x, y ∈ H,
and both {ξ′(x)}x∈G and {η′(x)}x∈G are total sets in the Hilbert space H ′ = Q(H).
There is therefore no loss of generality to assume, that {ξ(x)}x∈G and {η(x)}x∈G are
total in H. Since ϕ is continuous, the map
(x, y) → hξ(x), η(y)i
is separately continuous, and by the uniform boundedness, and totality of {ξ(x)}x∈G and
{η(x)}x∈G, it follows that ξ, η : G → H are continuous from G to H with σ(H, H ∗)-
topology. Let {ξi(x)}i∈I and {ηi(x)}i∈I be the coordinates of ξ(x) and η(x) with respect
to a fixed basis (ei)i∈I in H. Then ξi and ηi are continuous complex valued functions.
Moreover, for any open relative compact set U in G,
Xi∈IZU ξi(x)2 dx ≤ZU kξ(x)k2 dx < ∞.
By the continuity of ξi and the fact, that no non-empty open set has Haar measure zero,
it follows that all except countably many of the ξi's vanish on the set U , i.e. {ξi}i∈I is
locally countable, and similarly, {ηi}i∈I is locally countable.
(4) =⇒ (1): For f ∈ L∞(G), we let m(f ) denote the multiplication operator g 7→ f g on
L2(G). For x ∈ G
where γx(f )(y) = f (x−1y). Put now,
λ(x)m(f )λ(x)−1 = m(γx(f )),
ai = m(qξi),
bi = m(qηi),
a∗
1/2
b∗
i bi are convergent and
where qg(x) = g(x−1). ThenPi∈I a∗
i ai andPi∈I b∗
i bi(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
1/2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Xi∈I
i ai(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Φ(s) =Xi∈I
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Xi∈I
≤ sup
x∈Gkη(x)k ≤ k.
Thus, we can define a bounded σ-weakly continuous map Φ on B(L2(G)) by
x∈Gkξ(x)k · sup
b∗
i sai.
GROUP C∗-ALGEBRAS WITHOUT THE COMPLETELY BOUNDED APPROXIMATION PROPERTY
21
Now
Φ(λ(x)) = Xi∈I
= Xi∈I
= m Xi∈I
m(qηi)∗λ(x)m(qξi)
m(qηi)∗m(γx(qξi))λ(x)
qηiγx(qξi)! λ(x).
Here we have used that {ξi} and {ηi} are locally countable, so pointwise convergence of
the sum implies σ(L∞, L1)-convergence. But for y ∈ G,
qηi(y)(γx(qξi)(y)) =Xi∈I
Xi∈I
ηi(y−1)ξi(y−1x) = hξ(y−1x), η(y−1)i = ϕ(y(y−1x)) = ϕ(x).
Hence Φ(λ(x)) = ϕ(x)λ(x). This implies that
Fores ∈ B(L2(G)) ⊗ Mn(C),
Hence,
Φ(M(G)) ⊆ M(G).
(Φ ⊗ in)(es) =Xi∈I
(bi ⊗ 1n)∗es(ai ⊗ 1n).
1/2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Xi∈I
i bi(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
kΦ ⊗ ink ≤(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Xi∈I
i ai(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
a∗
b∗
1/2
ϕ ∈ M0A(G)
and kϕkM0A(G) ≤ k.
≤ k.
Thus kΦkCB ≤ k, so by [dCH85, Proposition 1.2]
Corollary 3.3 ([BF84] and [Her74]). Let Gd be the group G with discrete topology. Then
and the M0A-norms on the two spaces coincide.
M0A(G) = M0A(Gd) ∩ C(G),
(cid:3)
Proof. Immediate from (1) ⇐⇒ (2).
Remark 3.4. It is not hard to see that the functions ξ, η : G → H in (3) of Theorem 3.2
can actually be chosen to be continuous:
From the proof of (3) =⇒ (4) it follows that ξ, η can be chosen such that they are contin-
uous from G to H with respect to the σ(H, H ∗)-topology on H, and such that for every
open relative compact set U ⊆ G, ξ(U ) and η(U ) are contained in a separable subspace of
H. This implies that ξ, η ∈ L∞(G, H). Let now f ∈ Cc(G) with kfk2 = 1, and define
(cid:3)
by
bξ(x),bη(x) ∈ L2(G, H)
bξ(x)(z) = f (z)ξ(zx)
bη(x)(z) = f (z)η(zx)
for all x ∈ G
z ∈ G,
z ∈ G.
GROUP C∗-ALGEBRAS WITHOUT THE COMPLETELY BOUNDED APPROXIMATION PROPERTY
22
Then it is clear that
Moreover,
x∈Gkη(x)k.
sup
x∈Gkbξ(x)k2 ≤ sup
sup
x∈Gkξ(x)k,
x∈Gkbη(x)k2 ≤ sup
hbξ(x),bη(y)i = Z f (z)2hξ(zx), η(zy)idz
= Z f (z)2ϕ(y−1x)dz
= ϕ(y−1x).
Finally, using the fact that the group of right translations (Rx)x∈G defined by
(Rxh)(z) = ∆1/2
G (x)h(zx),
h ∈ L2(G, H)
acts norm-continuously on L2(G, H), it is not not hard to check that x 7→ bξ(x) and y 7→
bη(y) are norm continuous from G to L2(G, H).
REFERENCES
[BF84]
[BT65]
Marek Bozejko and Gero Fendler. Herz-Schur multipliers and completely bounded multipliers of the
Fourier algebra of a locally compact group. Boll. Un. Mat. Ital. A (6), 3(2):297–302, 1984.
Armand Borel and Jacques Tits. Groupes réductifs. Inst. Hautes Études Sci. Publ. Math., (27):55–
150, 1965.
[CDSW05] Michael Cowling, Brian Dorofaeff, Andreas Seeger, and James Wright. A family of singular oscilla-
[CH89]
tory integral operators and failure of weak amenability. Duke Math. J., 127(3):429–486, 2005.
Michael Cowling and Uffe Haagerup. Completely bounded multipliers of the Fourier algebra of a
simple Lie group of real rank one. Invent. Math., 96(3):507–549, 1989.
[Cow83] Michael Cowling. Harmonic analysis on some nilpotent Lie groups (with application to the repre-
sentation theory of some semisimple Lie groups). In Topics in modern harmonic analysis, Vol. I, II
(Turin/Milan, 1982), pages 81–123. Ist. Naz. Alta Mat. Francesco Severi, Rome, 1983.
[Cow89] M. Cowling. Rigidity for lattices in semisimple Lie groups: von Neumann algebras and ergodic
[dCH85]
[Dix58]
[Dix77]
[Dor93]
[Dor96]
[Eym64]
[Fel62]
actions. Rend. Sem. Mat. Univ. Politec. Torino, 47(1):1–37 (1991), 1989.
Jean de Cannière and Uffe Haagerup. Multipliers of the Fourier algebras of some simple Lie groups
and their discrete subgroups. Amer. J. Math., 107(2):455–500, 1985.
Jacques Dixmier. Sur les représentations unitaires des groupes de Lie nilpotents. III. Canad. J. Math.,
10:321–348, 1958.
Jacques Dixmier. C ∗-algebras. North-Holland Publishing Co., Amsterdam, 1977. Translated from
the French by Francis Jellett, North-Holland Mathematical Library, Vol. 15.
Brian Dorofaeff. The Fourier algebra of SL(2, R) ⋊ Rn, n ≥ 2, has no multiplier bounded approx-
imate unit. Math. Ann., 297(4):707–724, 1993.
Brian Dorofaeff. Weak amenability and semidirect products in simple Lie groups. Math. Ann.,
306(4):737–742, 1996.
Pierre Eymard. L'algèbre de Fourier d'un groupe localement compact. Bull. Soc. Math. France,
92:181–236, 1964.
J. M. G. Fell. Weak containment and induced representations of groups. Canad. J. Math., 14:237–
268, 1962.
J.E. Gilbert. Convolution operators of Banach space tensor products III. Unpublished.
[Gil]
[Han90] Mogens Lemvig Hansen. Weak amenability of the universal covering group of SU(1, n). Math.
[HdL13]
[HdL16]
[Hel78]
Ann., 288(3):445–472, 1990.
Uffe Haagerup and Tim de Laat. Simple Lie groups without the Approximation Property. Duke Math.
J., 162(5):925–964, 2013.
Uffe Haagerup and Tim de Laat. Simple Lie groups without the Approximation Property II. Trans.
Amer. Math. Soc., 368(6):3777–3809, 2016.
Sigurdur Helgason. Differential geometry, Lie groups, and symmetric spaces, volume 80 of Pure
and Applied Mathematics. Academic Press Inc. [Harcourt Brace Jovanovich Publishers], New York,
1978.
GROUP C∗-ALGEBRAS WITHOUT THE COMPLETELY BOUNDED APPROXIMATION PROPERTY
23
[Her74]
[HK94]
[Knu15]
[KR97]
Carl Herz. Une généralisation de la notion de transformée de Fourier-Stieltjes. Ann. Inst. Fourier
(Grenoble), 24(3):xiii, 145–157, 1974.
Uffe Haagerup and Jon Kraus. Approximation properties for group C ∗-algebras and group von Neu-
mann algebras. Trans. Amer. Math. Soc., 344(2):667–699, 1994.
Søren Knudby. Weak amenability and simply connected Lie groups. Preprint, to appear in Kyoto J.
Math., 2015.
Richard V. Kadison and John R. Ringrose. Fundamentals of the theory of operator algebras. Vol. II,
volume 16 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI,
1997. Advanced theory, Corrected reprint of the 1986 original.
[LDlS11] Vincent Lafforgue and Mikael De la Salle. Noncommutative Lp-spaces without the completely
[Lep68]
[Mac52]
[Oza12]
[Pau84]
[Rag72]
[RD66]
[Wan69]
bounded approximation property. Duke Math. J., 160(1):71–116, 2011.
Horst Leptin. Sur l'algèbre de Fourier d'un groupe localement compact. C. R. Acad. Sci. Paris Sér.
A-B, 266:A1180–A1182, 1968.
George W. Mackey. Induced representations of locally compact groups. I. Ann. of Math. (2), 55:101–
139, 1952.
Narutaka Ozawa. Examples of groups which are not weakly amenable. Kyoto J. Math., 52(2):333–
344, 2012.
Vern I. Paulsen. Every completely polynomially bounded operator is similar to a contraction. J.
Funct. Anal., 55(1):1–17, 1984.
M. S. Raghunathan. Discrete subgroups of Lie groups. Springer-Verlag, New York-Heidelberg, 1972.
Ergebnisse der Mathematik und ihrer Grenzgebiete, Band 68.
B. Russo and H. A. Dye. A note on unitary operators in C ∗-algebras. Duke Math. J., 33:413–416,
1966.
S. P. Wang. The dual space of semi-simple Lie groups. Amer. J. Math., 91:921–937, 1969.
DEPARTMENT OF MATHEMATICS AND COMPUTER SCIENCE, UNIVERSITY OF SOUTHERN DENMARK
CAMPUSVEJ 55, 5230 ODENSE M, DENMARK
|
1302.5781 | 1 | 1302 | 2013-02-23T10:02:08 | Type III actions on boundaries of $\tilde A_n$ buildings | [
"math.OA"
] | Let $\Gamma$ be a group of type rotating automorphisms of a building $\fX$ of type $\tilde A_n$ and order $q$. Suppose that $\G$ acts freely and transitively on the vertex set of $\fX$. Then the action of $\Gamma$ on the boundary of $\fX$ is ergodic, of type $\tq$ or type $\tqs$ depending on whether $n$ is odd or even. | math.OA | math |
TYPE III ACTIONS ON BOUNDARIES OF An
BUILDINGS
PAUL CUTTING AND GUYAN ROBERTSON
Abstract. Let Γ be a group of type rotating automorphisms of a
building X of type An and order q. Suppose that Γ acts freely and
transitively on the vertex set of X. Then the action of Γ on the
boundary of X is ergodic, of type III1/q or type III1/q2 depending
on whether n is odd or even.
Introduction
Let M be a compact Riemannian manifold of negative sectional cur-
vature, and let Γ = π1(M). Then Γ acts on the sphere at infinity S of
the universal cover M of M. The main result of [S] is that the action
of Γ on S is ergodic, amenable and type III1. This applies in particular
to a cocompact Fuchsian group in G = PGL(2, R) acting on the circle.
A discrete analogue of this result was proved in [RR1]. Namely, let
Γ be a free group acting simply transitively on the vertices of a locally
finite homogeneous tree T of degree q + 1. Then T is the universal
covering space of a graph with fundamental group Γ.
It was shown
in [RR1] that the action of Γ on the boundary of the tree is ergodic,
amenable and of type III1/q.
Turning to higher rank spaces of nonpositive curvature, it is known
that if Γ is a lattice in G = PGL(n + 1, R) with n ≥ 1 and if Ω = G/B
where B is the Borel subgroup of upper triangular matrices in G, then
the action of Γ on Ω is ergodic of type III1. Here Ω is the maximal
boundary of Furstenberg [Mar, VI.1.7]. A similar result holds more
generally for a lattice Γ in any semisimple noncompact Lie group G
[Zi, 4.3.15].
The discrete analogue of this construction is obtained by replacing
R by a nonarchimedean local field F with residue field of order q. The
affine Bruhat-Tits building X of G = PGL(n+1, F) is a building of type
Date: January 22, 2001.
1991 Mathematics Subject Classification. Primary 46L80; secondary 58B34,
51E24, 20G25.
This research was supported by The University of Newcastle.
1
Typeset by AMS-LATEX.
2
PAUL CUTTING AND GUYAN ROBERTSON
An [St]. The vertex set of X may be identified with the homogeneous
space G/K, where K is a maximal compact subgroup of G, and G acts
on the boundary Ω = G/B, where B is a Borel subgroup of G.
The precise higher rank analogue of the setup in [RR1] is as follows.
Let Γ be a group of type rotating automorphisms of a building X of
type An, and suppose that Γ acts simply transitively on the vertices
In view of the fact that A1 buildings are trees, such groups
of X.
Γ should be regarded as higher rank analogues of free groups. Note
however that not every A2 building X is the Bruhat-Tits building of
PGL (3, K) where K is a local field [CMSZ, II §8]. Geometrically, an
An building X is an n-dimensional contractible simplicial complex in
which each codimension one simplex lies on q + 1 maximal simplices
(chambers). If n ≥ 2 then the number q is necessarily a prime power
and is referred to as the order of the building. The boundary Ω of X is
a totally disconnected compact Hausdorff space and is endowed with
a natural family of mutually absolutely continuous Borel probability
measures.
In [RR2] it was proved that, if n = 2 and q ≥ 3, then
the action of Γ on Ω is ergodic and of type III1/q2. The purpose of
the present article is to remove both these hypotheses and prove the
following general result.
Theorem 1. Let n ≥ 2 and let X be a locally finite thick An building
of order q. Let Γ be a group of type rotating automorphisms of X which
acts simply and transitively on the vertices of X. Then the action of Γ
on the boundary Ω of X is amenable, ergodic and of type IIIλ, where
λ =(1/q
1/q2
if n is odd,
if n is even.
The proof of this result will be completed in Section 3. In Section 4
we deal with freeness of the action, which is required in order to prove
that the associated von Neumann algebra is a factor.
In particular,
Section 4 removes a gap in the proof of freeness in [RR2]. We therefore
obtain the following consequence.
Corollary 1. Let Γ and Ω be as above. Then the crossed product von
Neumann algebra L∞(Ω) ⋊ Γ is the AFD factor of type IIIλ, where
λ = 1/q if n is odd, and λ = 1/q2 if n is even.
A simple variation on the arguments leading to Theorem 1 proves
the following result: see subsection 3.2.
Theorem 2. Let p ≥ 2 be a prime number, let n ≥ 1, and let Ω be
the boundary of the affine building of PGL(n + 1, Qp). That is Ω =
PGL(n + 1, Qp)/B, where B is the Borel subgroup of upper triangular
matrices. Then the action of PGL(n + 1, Q) on Ω is ergodic and of
type IIIλ, where
3
λ =(1/p
1/p2
if n is odd,
if n is even.
Similar results can be stated for linear groups over other local fields,
but this is perhaps the most striking case. Note that, in contrast to
Theorem 1, PGL(n + 1, Q) is not a lattice in PGL(n + 1, Qp), and its
action on the boundary is not amenable.
Given an An building X, there is a type map τ defined on the vertices
of X such that τ (v) ∈ Z/(n+1)Z for each vertex v ∈ X. Every chamber
of X has precisely one vertex of each type. An automorphism α of X
is said to be type-rotating if there exists i ∈ {0, 1, . . . , n} such that
τ (αv) = τ (v) + i for all vertices v ∈ X. An A1 building is a tree,
with two types of vertices, and every automorphism of the tree is type
rotating. We shall refer to a group Γ satisfying the hypotheses of
Theorem 1 as an An group.
In [Ca1] it was shown that there is a
1-1 correspondence between An groups and "triangle presentations".
The right Cayley graph of an An group Γ relative to a natural set of
generators is the 1-skeleton of the An building X. We shall frequently
refer to [Ca3], which lays much of the groundwork for dealing with the
higher rank An buildings.
Throughout the paper X will denote a thick, locally finite An build-
ing, and the vertices of the building will be denoted by X0.
If X is
associated with the An group Γ then the underlying set of the group Γ
will be identified with X0, and the action of Γ on the building will be by
left multiplication. The identity of Γ will be denoted by 1 throughout.
For x, y ∈ X0, d(x, y) will denote the graph distance between those
vertices in the 1-skeleton of X, and x = d(x, 1).
Further information on buildings can be found in [Ca2], [St], [Br]
and [R]. The first two of these references are introductory, while the
last two provide a fuller account of the theory of buildings.
1. Preliminaries
This section mainly recalls material from [Ca3], to which we refer for
a more complete discussion. An An building is a union of apartments.
An apartment is isomorphic to a Coxeter complex of type An. Let
Σ denote the Coxeter complex of type An. The n-simplices of Σ are
referred to as chambers and can be regarded as forming a tessellation of
Rn. The vertex set of Σ can be identified with Zn+1/Z(1, 1, . . . , 1). Two
4
PAUL CUTTING AND GUYAN ROBERTSON
vertices [a], [b] ∈ Σ, [a] = a + Z(1, 1, . . . , 1) and [b] = b + Z(1, 1, . . . , 1),
are adjacent if there exist representative vectors (a1, a2, . . . , an+1) ∈ [a]
and (b1, b2, . . . , bn+1) ∈ [b] such that ai ≤ bi ≤ ai + 1 for all 1 ≤ i ≤ n.
The type τ (x) ∈ Z/(n + 1)Z of a vertex [x] = [(x1, x2, . . . , xn+1)] ∈ Σ
is given by
τ (x) = Xi
xi! mod (n + 1) .
Each chamber of Σ has precisely one vertex of each type.
Let bi = (0, . . . , 0, 1, . . . , 1), where precisely i entries equal 1. Note
that each x ∈ Zn+1 can be written as
Hence there is a mapping Zn+1/Z(1, 1, . . . , 1) → Zn defined by
x = x1(1, 1, . . . , 1) +X(xi+1 − xi)bi.
[x] 7→ (x2 − x1, x3 − x2, . . . , xn+1 − xn).
This mapping is a canonical group homomorphism between Zn+1/Z(1, 1, . . . , 1)
and Zn, and by means of it the vertices of Σ can be coordinatized by
Zn. Throughout this paper, if k ∈ Zn, then ki denotes the ith entry of
k.
1.1. The Boundary of an An Building. Given an An building X,
one can define the boundary of X by means of equivalence classes of
sectors. The central concern of this paper is the boundary regarded
as a measure space. For a discussion of the geometric structure of the
boundary, the reader is referred to [R, Chapter 9, 10].
Let S0 be the simplicial cone in the An Coxeter complex Σ with
vertex set coordinatized by Zn
+. A subcomplex S of X is called a sector if
there is an apartment A containing S and a type-rotating isomorphism
φ : A → Σ such that φ(S) 7→ S0.
(Recall that the isometry φ is
said to be type rotating if there exists j ∈ Z/(n + 1)Z such that, for
each vertex v of S, τ (φ(v)) = τ (v) + j (mod n + 1). Note that if a,
b are vertices in a sector S of X, and φ : S → S0 is a type preserving
isomorphism such that φ(a) = (0, 0, . . . , 0) and φ(b) = (k1, k2, . . . , kn),
where ki ∈ Z+, then the ki do not depend on the particular apartment
A containing S [Ca3, Lemma 2.3]. Thus, for x ∈ X0 and k ∈ Zn
+,
one can define a set Sk(x) consisting of those elements y ∈ X0 such
that there exists a sector S containing x and y, and a type rotating
isomorphism φ : S → S0 such that φ(x) = (0, 0, . . . , 0) and φ(y) = k.
Given a sector S and a type rotating isomorphism with φ(S) = S0, the
basepoint of S is v = φ−1(0, 0, . . . , 0). If x, y ∈ X and y ∈ Sk(x), with
k = (k1, k2, . . . , kn) ∈ Zn
+, then the graph distance between x and y is
given by d(x, y) =Pi ki.
5
Two sectors S1, S2 are said to be equivalent if S1 ∩ S2 contains a
subsector. Let Ω be the set of all such equivalence classes of sectors.
Then Ω is called the boundary of X. Given ω ∈ Ω and x ∈ X0, there
exists an unique sector with basepoint x which is contained in the
equivalence class ω [R, Lemma 9.7]. Denote this sector by [x, ω). Also,
for m ∈ Zn
m(ω) be the unique element in the intersection Sm(x)∩
[x, ω). What Follows Is Based On [Ca3].
Lemma 1.1. Let ω ∈ Ω, and let x, y ∈ X0. Then there exists
m(x, y; ω) ∈ Zn such that
k′(ω)
where k′ = k + m(x, y; ω) ,
k(ω) = sy
sx
+, let sx
for all k ∈ Z+ such that ki + mi(x, y; ω) ≥ 0 for 1 ≤ i ≤ n.
Proof. (c.f. [CMS, Lemma 2.1].) Since [x, ω) is in the same equiva-
lence class as [y, ω), [x, ω) ∩ [y, ω) contains a subsector. Choose
u = sx
k(ω) = sy
k′(ω) ∈ [x, ω) ∩ [y, ω) .
+}, and T ′ = {sy
k+l(ω); l ∈ Zn
Let T = {sx
+}. Then T , T ′
are sectors in the equivalence class ω with a common base point, and
so by [R, Lemma 9.7], T = T ′.
k′+l(ω) are
both in Su
l (ω) ∩ T , and hence are equal. Thus m(x, y; ω) = k′ − k, and
m(x, y; ω) is clearly independent of the choice of u ∈ [x, ω) ∩ [y, ω). (cid:3)
It follows that sx
k′+l(ω); l ∈ Zn
k+l(ω), sy
Lemma 1.2. Let x be a vertex of X, and let C be a chamber containing
x. Then for ω0 ∈ Ω, there exists an apartment A which contains C and
the sector S = [x, ω0).
Proof. By [R, Lemma 9.4], given the chamber C and sector [x, ω0),
there exists an apartment A containing a subsector S ′ ⊂ [x, ω0) and
the chamber C. Note that as x ∈ C, one has x ∈ A.
Choose a sector S ′′ in A with base vertex x and parallel to S ′. Then
S ′′ is equivalent to [x, ω0) and so S ′′ = [x, ω0), by uniqueness of the
sector with base vertex x representing the boundary point ω0.
(cid:3)
The next lemma is a generalisation of [CMS, Corollary 2.3].
Proposition 1.3. For x, y ∈ X0, and ω ∈ Ω, one has
if ki ≥ d(x, y) for 1 ≤ i ≤ n.
sx
k(ω) ∈ [x, ω) ∩ [y, ω)
Proof. Set r = d(x, y), and let k = (r, r, . . . , r). An easy consequence
of Lemma 1.1 is that z ∈ [x, ω) implies [z, ω) ⊂ [x, ω), and so it is
sufficient to show that sx
k ∈ [x, ω) ∩ [y, ω).
To proceed inductively, the case d(x, y) = 1 is established first. By
Lemma 1.2, there exists an apartment A containing both y and S =
6
PAUL CUTTING AND GUYAN ROBERTSON
[x, ω). As S is a sector, there exists a type rotating isomorphism ϕ :
A → Σ such that ϕ(S) = S0 with ϕ(x) = 0.
Since x, y are adjacent in A, ϕ(y) = (y1, y2, . . . , yn), where yi ∈
{−1, 1, 0}, 0 ≤ i ≤ n. Next, define the type rotating isomorphism
φ : A → Σ by
φ(z) = ϕ(z) − ϕ(y) .
This map takes y to the origin in Σ, and (φ)−1(S0) is a sector. Moreover,
for z ∈ sx
k(ω), φ(z) = ϕ(z) − ϕ(y) = ((k1 − y1), (k2 − y2), . . . , (kn − yn)).
Thus φ(z) ∈ S0 if and only if ki ≥ yi for all 1 ≤ i ≤ n.
It follows that φ−1(S0) = [y, ω). Moreover, as (1, 1, . . . , 1) ≥ (y1, . . . , yn),
one has that sx
d(x, y) = 1.
(1,1,...,1)(ω) ∈ [x, ω) ∩ [y, ω). This proves the case for
In general, given s ∈ Z+, s > 1, suppose that the statement of the
lemma is true for all y′ such that d(x, y′) ≤ s−1, and let y ∈ Sk(x) with
d(x, y) = s. Without loss of generality, suppose that k1 ≥ 1 and set
k′ = (k1 − 1, k2, . . . , kn). Let z be the unique element in conv(x, y) ∩
Sk′(x) and note that d(z, y) = 1 and d(z, x) = d(y, x) − 1 = s − 1.
Hence by the inductive hypothesis
a = sx
(s−1,s−1,...,s−1)(ω) ∈ [x, ω) ∩ [z, ω) .
Then for some t = (t1, t2, . . . , tn) ∈ Zn
t (ω) and
mi(x, z; ω) = (ti − (s − 1)). As ti ∈ Z+, it follows that ti + 1 ≥
d(y, z) = 1. By the inductive hypothesis, this implies that
+, one has a = sz
sz
(t1+1,...,tn+1)(ω) ∈ [z, ω) ∩ [y, ω) .
Writing t′ = t − m(x, z; ω),
sz
(t1+1,t2+1,...,tn+1)(ω) = sx
(t′
1+1,t′
2+1,...,t′
n+1) = sx
(s,...,s)(ω) ∈ [x, ω) ∩ [y, ω) .
The result follows.
(cid:3)
Definition 1.4. Given y ∈ X0, the topology on Ω based at y is given
by the basis of open sets {Ωx
y}x∈X0, where
Ωx
y = {ω ∈ Ω; x ∈ [y, ω)} .
The topology so defined is independent of the choice of y. See below
+, the boundary Ω can be
for details. Note that for y ∈ X0, and k ∈ Zn
expressed as the disjoint union
Ω = [x∈Sk(y)
Ωx
y .
There is a natural class of Borel measures on Ω. Namely, for a fixed
y ∈ X0 and a basic open set Ωx
y with x ∈ Sk(y), let
7
νy(Ωx
y) =
1
Sk(y)
.
Sk = Sk(y) is independent of y and its actual value was determined in
[Ca3, Corollary 2.7]. Specifically, let q be the order of the An building.
Also, for k = (k1, k2, . . . , kn) ∈ Zn
+, index the non-zero entries of k by
{i : ki ≥ 1} = {j1, . . . , jt}, and set j0 = 0 and jt+1 = n + 1. Then
(1.1) Sk = q− Pt
v=1 jv(jv+1−jv)(cid:20)
n + 1
j1 − j0, . . . , jt+1 − jt(cid:21)q
· qPn
i=1 i(n+1−i)ki ,
where [· · · ]q = [n + 1]q/([j1 − j0]q · · · [jt+1 − jt]q), and [k]q = (qk −
1) · · · (q − 1).
Unlike the topology on Ω, the value of the measure νy is dependent
on the choice of y ∈ X0. However, as shown by the following lemmas,
the set of measures {νy}y∈X0 is absolutely continuous.
The next lemma is generalized from [CMS, Lemma 2.4].
Lemma 1.5. Let y ∈ Sk(x). Suppose that z ∈ Sl(x) ∩ Sl′(y), where
li ≥ d(x, y) for all i. Then Ωz
y. Moreover, if m(x, y; ω) =
i −li for all ω ∈ Ωz
(m1, . . . , mn), as in Lemma 1.1, then mi(x, y; ω) = l′
x
x ⊂ Ωz
Proof. Let ω ∈ Ωz
x. Then z = sx
[x, ω) ∩ [y, ω) by Lemma 1.3 and choice of l ∈ Zn
Moreover, by the proof of Lemma 1.1, mi(x, y; ω) = l′
l (ω), and so z is an element of
+. Thus ω ∈ Ωz
y.
(cid:3)
i − li.
Lemma 1.6. The topology on Ω does not depend on the vertex y ∈ X0
chosen in Definition 1.4. For any x, y ∈ X0, the measures νx, νy are
mutually absolutely continuous, and the Radon Nikodym derivative of
νy with respect to νx is given by
(1.2)
dνy
dνx
(ω) = q− Pn
i=1 i(n+1−i)mi ,
for ω ∈ Ω, where mi = mi(x, y; ω).
Proof. Let x, y ∈ X0. In view of the preceding results, the proof
that topology is independent of the base vertex y proceeds exactly as
in the case n = 2 [CMS, Lemma 2.5].
Now fix ω ∈ Ω. Choose k ∈ Zn
mi(x, y; ω) ≥ d(x, y). Set z = sx
x = Ωz
Lemma 1.5 implies that Ωz
i=1 i(n+1−i)miSk)−1 = q− Pn
x) = (Sk′)−1 = (qPn
νy(Ωz
+ such that ki ≥ d(x, y) and ki +
k(ω) = sy
k′(ω), where k′ = k + m(x, y; ω).
y. Moreover it follows from (1.1) that
i=1 i(n+1−i)miνx(Ωz
x) .
8
PAUL CUTTING AND GUYAN ROBERTSON
Since {Ωz
hoods of ω, the result follows.
x; z ∈ [x, ω), d(x, z) ≥ d(x, y)} is a basic family of neighbour-
(cid:3)
Remark 1.7. Equation (1.2) is precisely [Ca3, Equation (1.6)], and
its proof is outlined in [Ca3, Section 4].
1.2. An Groups. Let Π be a finite projective geometry of dimension
n and order q. If n > 2 then Π is the Desarguesian projective geometry
Π(V ), where V is a vector space of dimension n + 1 over a finite field of
order q. Let dim(u) denote the dimension of the subspace u of V . In
the Desarguesian case the points and lines of Π are the one- and two-
dimensional subspaces of V respectively. We shall extend this notation
to the non Desarguesian case, so that an element u of a projective plane
Π satisfies dim u = 1 if it is a point and dim u = 2 if it is a line. Let λ be
an involution of Π such that dim(λ(u)) = n + 1 − dim(u) mod (n + 1).
An An triangle presentation T compatible with λ is defined as follows.
Let T be a set of triples {(u, v, w) : u, v, w ∈ Π} which satisfy the
following properties.
(1) Given u, v ∈ Π, then (u, v, w) ∈ T for some w ∈ Π if and only
if λ(u) and v are distinct and incident.
(2) If (u, v, w) ∈ T , then (v, w, u) ∈ T .
(3) If (u, v, w1) ∈ T and (u, v, w2) ∈ T , then w1 = w2.
(4) If (u, v, w) ∈ T , then (λ(w), λ(v), λ(u)) ∈ T .
(5) If (u, v, w) ∈ T , then dim(u)+dim(v)+dim(w) ≡ 0 mod n+1.
The group associated with this triangle presentation is given by
ΓT =(cid:28){av}v∈Π(x)(cid:12)(cid:12)(cid:12)(cid:12)
(1)aλ(v) = a−1
v
(2)auavaw = 1
for all v ∈ Π
for all (u, v, w) ∈ T(cid:29) .
The Cayley graph of ΓT , with respect to the generators {au}u∈Π is
the 1-skeleton of an An building X and ΓT acts on the vertices of the
building in a type rotating manner. Conversely any group Γ acting
on an An building in this way arises as Γ = ΓT for some triangle
presentation T [Ca1, pp 45 -- 46]. Unless otherwise specified, a generator
au of Γ will be identified with the corresponding element u ∈ Π.
Remark 1.8. The type rotating hypothesis in the definition of an
An group has been removed and the appropriate notion of triangle
presentation studied in the Ph.D. thesis of T. Svenson [Sv], thereby
generalising the results of [Ca1].
For the rest of this article, the An group Γ will be assumed to act
on X by left translation with Γ being identified with the vertex set X0.
The identity element 1 of Γ is a preferred vertex of X of type 0, and
9
we write Sk = Sk(1) for k = (k1, k2, . . . , kn) ∈ Zn
naturally on the boundary Ω.
+. The group Γ acts
If u1, u2 are elements of Π we denote by u1 ∨ u2 their join; that is
their least upper bound in the lattice of subspaces of Π. If Π = Π(V ) is
Desarguesian then u1 ∨ u2 = Π means simply that u1 + u2 = V . On the
other hand, if Π is a non Desarguesian plane and u1 is a point and u2
is a line of Π, then u1 ∨ u2 = Π means that u1 and u2 are not incident.
By [Ca1, Lemma 2.2], every word in Γ can be expressed uniquely in
normal form
x = u1u2 . . . ul ,
where dim(ui) ≤ dim(ui+1) and u−1
where kj = {ui : dim(ui) = j}.
i ∨ ui+1 = Π. Moreover, x ∈ Sk,
Recall from [Ca1, Proof of Theorem 2.5] that if x ∈ X0 then the
projective geometry of neighbours of x is {xu : u ∈ Π} and τ (xu) =
τ (x) + dim u mod (n + 1). Moreover, xu and xu′ are adjacent vertices
if and only if u and u′ are incident in Π (that is, u ⊂ u′ or u′ ⊂ u). In
particular a chamber of X containing the vertex x has the form
{x, xu1, xu2, . . . , xun}
where dim ui = i and u1 ⊂ u2 ⊂ · · · ⊂ un is a complete flag in Π.
For more information on An groups, the reader is referred to [Ca1].
2. An Ergodic measure preserving subgroup of the full
group.
The action of an An group Γ on the boundary Ω of the corresponding
An building, is measure-theoretically ergodic with respect to each of
the measures νy, y ∈ X. For the classification of the action it will
be necessary to show that the full group [Γ] (defined below) contains a
countable measure preserving subgroup K0 ⊂ [Γ] which acts ergodically
on Ω.
The following two lemmas are straightforward generalisations of [RR2,
Lemma 4.6] and [RR2, Lemma 4.7] respectively.
Lemma 2.1. Let K be a group which acts on Ω. If K acts transitively
on the collection of sets {Ωx
+, then K acts
ergodically on Ω.
1 : x ∈ Sk} for every k ∈ Zn
Proof. Observe first that K preserves ν1 since ν1(Ωx
1) is independent
of x. Suppose that X0 ⊆ Ω is a Borel set which is invariant under K
and such that ν1(X0) > 0. It will be shown that ν1(Ω \ X0) = 0, thus
establishing the ergodicity of the action.
10
PAUL CUTTING AND GUYAN ROBERTSON
Define a new measure µ by µ(X) = ν1(X ∩ X0) for each Borel set
X ⊆ Ω. Now, for each g ∈ K,
µ(gX) = ν1(gX ∩ X0)
= ν1(X ∩ g−1X0)
≤ ν1(X ∩ X0) + ν1(X ∩ (g−1X0 \ X0))
= ν1(X ∩ X0)
= µ(X).
Similarly, µ(gX) ≤ µ(g−1gX) = µ(X). Therefore µ is K-invariant.
1 = Ωy
For each x, y ∈ Sk there exists a g ∈ K such that gΩx
1 by
1). Since Ω is the union of Sk disjoint
1) = µ(Ωy
transitivity. Thus µ(Ωx
sets Ωx
1, y ∈ Sk, each of which has equal measure, one has that
µ(Ωx
1) =
c
Sk
, for each x ∈ Sk,
where c = µ(X0) = ν1(X0) > 0. Thus µ(Ωx
x ∈ X.
1) = cν1(Ωx
1) for every vertex
Since the sets Ωx
1 generate the Borel σ-algebra, it follows that µ(X) =
cν1(X) for each Borel set X. Therefore
ν1(Ω \ X0) = c−1µ(Ω \ X0)
= c−1ν1((Ω \ X0) ∩ X0) = 0,
thus proving ergodicity.
(cid:3)
Lemma 2.2. Assume that K ≤ Aut(Ω) acts transitively on the collec-
tion of sets {Ωx
+. Then there is a countable
subgroup K0 of K which also acts transitively on the collection of sets
{Ωx
1 : x ∈ Sm} for every m ∈ Zn
1 : x ∈ Sm} for every Sm, m ∈ Zn
+.
1 = Ωx
Proof. For each pair x, y ∈ Sm, there exists an element k ∈ K such that
kΩy
1. Choose one such element k ∈ K and label it kx,y. Since Sm is
finite, there are a finite number of elements kx,y ∈ K for each Sm. There
are countably many sets Sm, so the set {kx,y : x, y ∈ Sm, m ∈ Zn
+} is
countable. Hence the group
is countable and satisfies the required condition.
K0 =(cid:10)kx,y; x, y ∈ Sm, m ∈ Zn
+(cid:11) ≤ K
(cid:3)
Definition 2.3. Given a group Γ acting on a measure space Ω, define
the full group, [Γ], of Γ by
[Γ] = {T ∈ Aut(Ω); T ω ∈ Γω for almost every ω ∈ Ω} .
The set [Γ]0 of measure preserving maps in [Γ] is then given by
[Γ]0 = {T ∈ [Γ]; νy ◦ T = νy, y ∈ X0} .
It will be shown that there is a countable group K0 of measure-
preserving automorphisms of Ω such that
11
(1) K0 acts ergodically on Ω.
(2) K0 ≤ [Γ].
By the Lemmas above and the definition of [Γ], it is enough to find
+ and x, y ∈ Sk, an automorphism g ∈ Aut(Ω) such that
1) = Ωy
Identify a simplex in X with its vertex set, and recall from section
1 and gω ∈ Γω for almost all ω ∈ Ω.
for each k ∈ Zn
g(Ωx
1.2 that a chamber of X containing the vertex x is of the form
{x, xu1, xu2, . . . , xun} .
where dim ui = i and u1 ⊂ u2 ⊂ · · · ⊂ un is a complete flag in Π.
Lemma 2.4. Let C = {1, p1, p2, . . . , pn} be a chamber in X with base
vertex the identity element 1 of Γ, where pi are generators of Γ and
dim pi = i. There are q chambers C ′ = {x, p1 . . . , pn} in X meeting
C in the face {p1, . . . , pn}. The vertex x opposite 1 in C ∪ C ′ has the
normal form x = p1un, where dim un = n and p−1
1 ∨ un = Π. Thus
x ∈ S(1,0,...,0,1).
Equivalently, x = pnu′
1, where dim u′
1 = 1 and pn ∨ (u′
1)−1 = Π.
pj
x
C ′
......................................................................................................................................................................................
......................................................................................................................................................................................
C
1
p1
Figure 1.
Proof. Consider the projective geometry of the neighbours of p1. For
2 ≤ i ≤ n there exists ui−1 ∈ Πi−1 such that
pi = p1ui−1
and
ui−1 ⊂ uj−1 for i ≤ j .
Now choose un ∈ Πn such that un−1 ⊂ un and un 6= p−1
1 . There exist q
such choices for un. One then has that for all 2 ≤ i ≤ n, ui−1 ⊂ un, and
hence pi = p1ui−1 is adjacent to p1un. Thus C ′ = {p1, p2, . . . , pn, p1un}
is a chamber of X and p1un is the vertex x opposite 1 in C ∪C ′. Clearly
p−1
1 ∨ un = Π, so x = p1un is the normal form expressing x as a word
of minimal length.
12
PAUL CUTTING AND GUYAN ROBERTSON
It now follows that x ∈ S(1,0,...,0,1), and a similar argument proves the
(cid:3)
final statement.
Lemma 2.5. Let x ∈ Sk and y ∈ Sk′, k, k′ ∈ Zn
(k1, . . . , kn) and k′ = (k′
ϕ of Ω such that
+, where k =
n). Then there exists an automorphism
1, . . . , k′
(1) ϕ ∈ [Γ], the full group of Γ;
(2) ϕ is almost everywhere a bijection from Ωx
(3) ϕ is the identity on Ω\(Ωx
1 ∪ Ωy
1).
1 onto Ωy
1;
Moreover, if k = k′ then ϕ is measure preserving.
Proof. Let δ = e1 + en = (1, 0, . . . , 0, 1) and consider the set of all
vertices x1 ∈ Sk+δ such that x ∈ conv{1, x1}. For such a vertex x1,
one has that x1 ∈ Sδ(x) and Ωx1
1 is a (disjoint) union
of sets of the form Ωx1
x , where x1 ∈ Sδ(x) ∩ Sk+δ is constructed
using the procedure of Lemma 2.4.
x . Thus Ωx
1 = Ωx1
1 = Ωx1
Similarly, Ωy
1 is a disjoint union of sets of the form Ωy1
1 = Ωy1
y , where
y1 ∈ Sδ(y) ∩ Sk′+δ. Refer to Figure 2 below.
x2vnpi
x2vn
y1zvnpi
x3
x2
Cx
C ′
x
..............................................................................................................................................................
..............................................................................................................................................................
...............................................................................
...............................................................................
...............................................................................
x
x1
...............................................................................
...............................................................................
y1zvn
y1z = y2
y3
Cy
C ′
y
..............................................................................................................................................................
..............................................................................................................................................................
...............................................................................
...............................................................................
...............................................................................
y
y1
...............................................................................
...............................................................................
Figure 2.
It is therefore enough to show that for every such x1, y1, there is a
y which coincides pointwise
x . That is, for almost
measure preserving bijection ϕ : Ωx1
with the action of Γ almost everywhere on Ωx1
each ω ∈ Ωx1
x , there exists g ∈ Γ such that ϕω = gω.
x → Ωy1
Fix such x1, y1. Choose x2 ∈ Sen(x1)∩Sk+δ+en. Also choose vn ∈ Sen
such that
x2vn ∈ S2en(x1) ∩ Sk+δ+2en .
Since y1 ∈ Sk′+δ, it has normal form
y1 = u1 . . . ul
where ul ∈ Sen .
We now show that there exists z ∈ Sen such that
(2.1)
u−1
l ∨ z = Π
and
z−1 ∨ vn = Π .
13
To prove the claim, it is necessary to make use of the identification
of the generators of Γ with elements of the finite projective space Π.
Set Πr = {x ∈ Π : dim x = r} = Ser .
Now, u−1
l ∈ Π1, and vn ∈ Πn. Therefore
(a): {z ∈ Πn : u−1
l ∨ z 6= Π} = {z ∈ Πn : u−1
l ⊂ z} =
1 + q + q2 + · · · + qn−1.
(b): {z ∈ Πn : z−1 ∨ vn 6= Π} = {z ∈ Πn : z−1 ⊂ vn} =
1 + q + q2 + · · · + qn−1.
Also,
Πn = 1 + q + q2 + · · · + qn > 2(1 + q + · · · + qn−1) .
Hence there exists z ∈ Πn such that (2.1) holds.
It follows that the word y1zvn = u1 . . . ulzvn is in normal form and
hence that
y1zvn ∈ Sk′+δ+2en .
Moreover, y2 = y1z ∈ Sk′+δ+en.
It will now be shown that the chambers Cx, Cy can be constructed
which lie as indicated in (the two dimensional) Figure 2. By this it
is meant, for example, that if ω ∈ Ω and Cx ⊂ Sx2(ω) then Sx2(ω) ⊂
Sx1(ω) ⊂ Sx(ω). In fact, Cx = x2C and Cy = y1zC, where C is the
chamber based at 1, as illustrated in Figure 3 (in two dimensions).
vnpi
w
..................................................................................................................................................................................................................................................................................................
..................................................................................................................................................................................................................................................................................................
C
1
vn
Figure 3.
The vertices x2vnpi, 2 ≤ i ≤ n, will be constructed from a flag
v−1
n ⊂ p2 ⊂ p3 ⊂ · · · ⊂ pn, where pi ∈ Πi. For the following argument,
note that if b ∈ Πr−1, where r ≥ 2, then
{a ∈ Πr : a ⊃ b} = 1 + q + · · · + qn−r+1 ≥ 1 + q .
14
PAUL CUTTING AND GUYAN ROBERTSON
There are at least 1 + q elements p2 ∈ Π2 such that p2 ⊃ v−1
n . By
reference to both Lemma 2.4 and the proof of Proposition 2.7 in [Ca3],
there is precisely one such p2 such that x2vnp2 < x2vn. (In fact, in
that case x2vnp2 ∈ Sk+δ+2en−en−1.)
Similarly, there is precisely one p2 ∈ Π2, p2 ⊃ v−1
n such that y1zvnp2 <
y1zvn.
One can therefore choose p2 ∈ Π2 with p2 ⊃ v−1
n such that x2vnp2 ≥
x2vn and y1zvnp2 ≥ y1zvn. Moreover, since vnp2 is then adjacent
to 1, these inequalities are in fact equalities.
This process is now continued. There are at least 1 + q elements
p3 ∈ Π3 such that p3 ⊃ p2 and at most two of them satisfy either
x2vnp3 < x2vn or y1zvnp2 < y1zvn. Thus we may choose p3 ⊃ p2
such that x2vnp3 = x2vn and y1zvnp3 = y1zvn.
Continue in this way to obtain a flag
v−1
n ⊂ p2 ⊂ p3 ⊂ · · · ⊂ pn
such that the vertex set of the chamber C is
{1, vn, vnp2, vnp3, . . . , vnpn} .
Then Cx = x2C and Cy = y2C.
Now choose, by Lemma 2.4, a vertex w (one of q possible) of a
chamber C ′
x which meets Cx in the face Cx\{1}. Thus w ∈ Sδ, and
x3 = x2w ∈ Sk+2δ+en. Also y3 = y2w ∈ Sk′+2δ+en. (Recall that, by
definition, y2 = y1z.) Moreover, y2x−1
2 Cx = Cy.
It has now been shown that
Ωx3
x2 ⊂ Ωx1
x ,
Ωy3
y2 ⊂ Ωy1
y
and
y2x−1
2 Ωx3
x2 = Ωy3
y2 .
Therefore one can define the map ϕ on Ωx3
x2 by
ϕω = y2x−1
2 ω .
Now recall that x ∈ Sk, y ∈ Sk′ and x1 ∈ Sk+δ, y1 ∈ Sk′+δ were
fixed, and that x2 ∈ Sen(x2) ∩ Sk+δ+en was chosen. The set Ωx1
x is a
disjoint union of sets of the form Ωx3
x2 where x3 ∈ Sδ(x2). Let K denote
the number of such sets. This number is independent of the choice of
x, x1 and k by [Ca3, Lemma 2.4], (or by the fact that Γ acts simply
transitively on X0).
The definition ϕω = y2x−1
2 ω in the above choice of Ωx3
x2 therefore
leaves ϕ undefined on a proportion (1 − 1
x . However, where ϕ
is defined it coincides with the action of an element of Γ, namely y2x−1
2 .
Now repeat the process on each of the K − 1 sets of the form Ωx3
x2
where ϕ has not been defined. As before, ϕ can be defined except on
K ) of Ωx1
15
a proportion (1 − 1
everywhere except on a proportion (1 − 1
K ) of each such set, and ϕ can therefore be defined
K )2 of the original set Ωx1
x .
Continuing in this manner, one sees that at the nth step, ϕ has been
defined everywhere except on a proportion (1 − 1
K )n of Ωx1
x .
Since (1 − 1
K )n → 0 as n → ∞, ϕ is defined almost everywhere on
x and satisfies the required properties. If k = k′ then it is clear from
(cid:3)
Ωx1
the construction that ϕ is measure preserving.
Remark 2.6. This result extends [RR2, Proposition 4.9]. Moreover,
for n = 2 this proof deals with the case q = 2, which was left open in
[RR2]. Thus the hypothesis q ≥ 3 in the main result Theorem 4.10 of
[RR2] is not in fact necessary.
We can now prove the following.
Proposition 2.7. There exists a countable subgroup K0 of [Γ] such
that
(1) K0 is measure preserving;
(2) K0 acts ergodically on Ω.
Proof. It suffices to take the group generated by the automorphisms
of the form ϕ defined in Lemma 2.5, with k = k′ then use Lemma 2.2
to extract a countable subgroup K0. Finally, Lemma 2.1 shows that
the action of K0 is ergodic.
(cid:3)
Lemma 2.8. Let Γ be a countable group acting on a measure space
(Ω, µ). Suppose that the action of the full group [Γ] is ergodic. Then
so is the action of Γ.
Proof. Let S be a measurable subset of Ω such that µ(gS\S) = 0 for
all g ∈ Γ. Let k ∈ [Γ]. It will be shown that µ(kS\S) = 0. For each
g ∈ Γ, let
Sg = {ω ∈ S : kω = gω} ,
which is a measurable subset of S. Since k ∈ [Γ] it follows that
S = S0 ∪ [g∈Γ
Sg ,
where S0 has measure zero.
16
PAUL CUTTING AND GUYAN ROBERTSON
Then
µ(kS\S) ≤ Xg∈Γ
= Xg∈Γ
≤ Xg∈Γ
µ(kSg\S)
µ(gSg\S)
µ(gS\S) = 0 .
Thus µ(kS\S) = 0 for all k ∈ [Γ]. Since the action of [Γ] is ergodic it
follows that S is either null or conull with respect to the measure µ.
Therefore the action of Γ is ergodic.
(cid:3)
Corollary 2.9. The action of Γ on Ω is ergodic.
Proof. This follows from Proposition 2.7 and Lemma 2.8.
(cid:3)
3. Classification of the action of Γ on Ω
Having shown that the action of an An group on its boundary Ω is
ergodic, we now show that it is type IIIλ, where 0 ≤ λ ≤ 1, and the
value of λ depends on the ratio set.
Definition 3.1. Let Γ be a countable group of automorphisms of the
measure space (Ω, ν). Following Krieger, define the ratio set r(Γ) to
be the set of λ ∈ [0, ∞) such that for every ǫ > 0 and Borel set E with
ν(E) > 0, there exists a g ∈ Γ and a Borel set F such that ν(F ) > 0,
F ∪ gF ⊂ E and
for all ω ∈ F .
< ǫ
dν ◦ g
dν
(cid:12)(cid:12)(cid:12)(cid:12)
(ω) − λ(cid:12)(cid:12)(cid:12)(cid:12)
Remark 3.2. The ratio set r(Γ) depends only on the quasi-equivalence
class of the measure ν [HO, section I-3, Lemma 14]. It also depends
only on the full group in the sense that
[H] = [G] =⇒ r(H) = r(G) .
Proposition 3.3. Let X be a locally finite, thick An building of order
q, and let Γ be a countable group of type rotating automorphisms of X.
Fix a vertex O ∈ X0 of type 0, and suppose that for each 0 ≤ i ≤ n
there exists an element gi ∈ Γ such that d(giO, O) = 1 and giO is a
17
vertex of type i. Also, suppose that there exists a countable subgroup K
of [Γ]0 whose action on Ω is ergodic. Then
r(Γ) =(cid:26) {qn : n ∈ Z} ∪ {0}
for n odd
{q2n : n ∈ Z} ∪ {0} for n even
.
Proof. By Remark 3.2, it is sufficient to prove the statement for some
group H such that [H] = [Γ]. In particular, since [Γ] = [hΓ, Ki] for
any subgroup K of [Γ]0, we may assume without loss of generality that
K ≤ Γ.
Let ν = νO. For gi ∈ Γ as in the statement of the lemma, let x = giO,
. If m(O, x; ω) = (m1, m2, . . . , mn) then by
and note that νx = ν ◦ g−1
Lemma 1.6,
i
dν ◦ g−1
i
(ω) =
(ω) = q− Pn
i=1 i(n+1−i)mi .
dνx
dν
dν
Then for ω ∈ Ωx
O, one has that m(O, x; ω) = (0, . . . , 0, −1, 0, . . . , 0),
where the −1 is in the ith place. Thus
(3.1)
dνx
dν
(ω) = qi(n+1−i)
for ω ∈ Ωx
O.
Let E ⊂ Ω be a Borel set with ν(E) > 0. Then by the ergodicity of K,
there exist k1, k2 ∈ K such that the set
F = {ω ∈ E : k1ω ∈ Ωx
O and k2g−1
i k1ω ∈ E}
has positive measure.
Next, let t = k2g−1
since K is measure preserving,
i k1 ∈ Γ. By construction, F ∪ tF ⊂ E. Moreover,
dν ◦ t
dν
(ω) =
dν ◦ g−1
i
dν
(k1ω) =
dνxi
dν
(k1ω) = qi(n+1−i)
for all ω ∈ F
by (3.1), and since kiω ∈ Ωx
O. Hence qi(n+1−i) ∈ r(Γ) for 1 ≤ i ≤ n.
Since the action of Γ on Ω is ergodic, r(Γ) − {0} forms a group. It
is now possible to determine the generator of r(Γ) − {0}.
Suppose that n is odd. Then for i ∈ {1, 2}, one has that qn, q2(n−1) ∈
r(Γ). As n, 2(n − 1) are coprime for n odd, and as r(Γ) − {0} forms a
group, it follows that q ∈ r(Γ), and hence
r(Γ) = {qn : n ∈ Z}
for n odd.
Suppose that n is even. As before, qn, q2(n−1) ∈ r(Γ). Moreover, as
highest common factor of n, 2(n − 1) is 2 for n even, and as r(Γ) forms
a group, it follows that q2 ∈ r(Γ). Finally, as i(n + 1 − i) is even for all
i if n is even, it follows that for g ∈ Γ, and x = g−1O,
dν ◦ g
dν
=
dνx
dν
= q− P i(n+1−i)mi ∈ {q2n : n ∈ Z} .
18
PAUL CUTTING AND GUYAN ROBERTSON
Thus
r(Γ) = {q2n : n ∈ Z}
for n even.
(cid:3)
Proposition 3.4. Let Γ be an An group. Then the action of Γ on Ω
is amenable.
Proof. This is a straightforward generalization of the case n = 2,
(cid:3)
proved in [RR2, Proposition 3.14].
3.1. Proof of Theorem 1. This follows from Proposition 2.7, Corol-
lary 2.9, Proposition 3.3 and Theorem 3.4.
(cid:3)
3.2. Proof of Theorem 2. The proof of Theorem 2 is now easy. Let
X be the affine building of G = PGL(n + 1, Qp). By [Br, Proposition
VI.9F], the boundary Ω of X is isomorphic to PGL(n + 1, Qp)/B as a
topological G-space. The measure µ on G/B is (up to equivalence) the
natural quasi-invariant Borel measure on G/B.
The vertex set X0 of X is identified with PGL(n + 1, Qp)/PGL(n +
1, Zp), where Zp is the ring of p-adic integers. Let O = PGL(n+1, Zp) ∈
X0.
It follows from [St, Proposition 3.1] that PGL(n + 1, Zp) acts
transitively on each set Sk(O). Since the vertex set X0 = PGL(n +
1, Qp)/PGL(n + 1, Zp) is a discrete space and Z is dense in Zp it follows
that PGL(n + 1, Z) also acts transitively on Sk(O) for each k ∈ Zn
+.
Moreover PGL(n + 1, Z) stabilizes O. Therefore PGL(n + 1, Z) acts
ergodically on Ω by Lemma 2.1. The group PGL(n + 1, Q) also acts
transitively on X0. By our previous computation of Radon-Nikodym
derivatives and the argument of [RR2, Proposition 4.4] the type of the
action is as stated.
Note that in this argument there is no need to consider the full group,
since PGL(n + 1, Z) is already a measure preserving ergodic subgroup
of PGL(n + 1, Q). Thus the proof is considerably simpler than the
proof of Theorem 1.
4. Freeness of the action on the boundary
A simple modification of [RR2, Proposition 4.12] shows that if F is a
(possibly non commutative) local field then the action of PGL(n+1, F)
on its Furstenberg boundary Ω is measure-theoretically free. If X is a
thick, locally finite affine building of type An, where n ≥ 3, then X is
the building of PGL(n + 1, F) for some such local field [R, p137]. All
known type-rotating An groups, with n ≥ 3, embed in PGL(n + 1, F)
and act upon the building of PGL(n + 1, F) in the canonical way. For
such groups Γ, the action on Ω is therefore measure theoretically free.
19
The case of A2 groups is more interesting because the associated A2
building may not be the affine building of a linear group. In fact this is
the case for the buildings of many of the groups constructed in [CMSZ].
The boundary action of Γ is nevertheless free.
Proposition 4.1. Let X be an A2 building, and let Γ be an A2 group.
Then the action of Γ on Ω is measure-theoretically free [RR2, c.f.
Proposition 3.10].
This result was stated in [RR2, Proposition 3.10]. The proof given
there contains a gap, because the proof of [RR2, Lemma 3.7, Case 2]
is not complete. Our purpose is to fill this gap (Lemma 4.3 below). A
similar argument applies to the An case, where n ≥ 3. In view of the
comments at the beginning of this section, we have confined ourselves
to the A2 case.
Let O be a fixed vertex in X and let n ∈ Z+. Any sector S based at
O has at its base a triangle Tn with apex O consisting of all vertices v
in S such that d(v, O) ≤ n. (See Figure 4 below).
.
.
3n-3
..........................................................................................................................................................................................................................................................................................
..........................................................................................................................................................................................................................................................................................
........................................................................................................................................................................................................................................................................................................................................................................................
..........................................................................................................................................................................................................................................................................................
..........................................................................................................................................................................................................................................................................................
..............................................................................................
..............................................................................................
...........................................................................................................................................................................................................................................................................................
...........................................................................................................................................................................................................................................................................................
...........................................................................................................................................................................................................................................................................................
...........................................................................................................................................................................................................................................................................................
...........................................................................................................................................................................................................................................................................................
........................................................................................................................................................................................................................................................................................... ........................................................................................................................................................................................................................................................................................... ........................................................................................................................................................................................................................................................................................... ........................................................................................................................................................................................................................................................................................... ........................................................................................................................................................................................................................................................................................... ........................................................................................................................................................................................................................................................................................... ........................................................................................................................................................................................................................................................................................... ...........................................................................................................................................................................................................................................................................................
..............................................................................................
..................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................
..............................................................................................
..............................................................................................
..............................................................................................
..............................................................................................
..........................................................................................................................................................................................................................................................................................
..........................................................................................................................................................................................................................................................................................
..........................................................................................................................................................................................................................................................................................
............................................................................................................................................................................................
..............................................................................................
........................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................
..........................................................................................................................................................................................................................................................................................
..........................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................
......................................................................................................................................................................................................................................................................................................................................................................................................................................................................................
..........................................................................................................................................................................................................................................................................................
..............................................................................................
..............................................................................................
..............................................................................................
..............................................................................................
2n-1
.
4
3
2
2n
2n-2
..............................................................................................
1
C
O
Figure 4.
Consider the set Sn of all such triangles Tn in X.
Lemma 4.2. Let Sn be as above. Then
Sn = (q2 + q + 1)(q + 1)q3n−3 .
Proof. First note that there are (q2 + q + 1)(q + 1) choices for the
base chamber C.
20
PAUL CUTTING AND GUYAN ROBERTSON
The reason for this is that one edge of C containing O is determined
by the number of points P in a finite projective plane Π of order q.
There are q2 + q + 1 such points P . See Figure 5 below.
L
P
C
.............................................................................................................................................................
.............................................................................................................................................................
O
Figure 5.
Having chosen the point P , there are precisely q + 1 possible lines L
in Π which are incident with P . There are therefore (q2 + q + 1)(q + 1)
choices of C.
Having chosen C, there are q choices for each of the chambers labeled
1, 2, 3, 4, . . . , (2n − 2) in the figure. Then choose the chamber labeled
(2n − 1) (q choices) in Figure 4. This choice then determines the whole
shaded region in the figure (which is contained in the convex hull of
the chambers already chosen, and hence is uniquely determined). Now
choose the chamber labeled 2n (q choices) and continue the process
until finally chamber 3n − 3 is chosen. This determines the triangle
completely and there are (q2 +q +1)(q +1)q3n−3 possibilities altogether.
(cid:3)
This demonstrates that for each positive integer n, the boundary Ω
of X is partitioned into (q2 + q + 1)(q + 1)q3n−3 sets {ΩT : T ∈ Sn},
where
ΩT = {ω ∈ Ω : T ⊂ [O, ω)} .
Moreover each of these sets has the same measure [CMS].
The proof of [RR2, Lemma 3.7, case 2.] can now be completed.
Lemma 4.3. Let W be a wall of X and let Σ denote the set of boundary
points ω ∈ Ω such that for some vertex v, the sector [v, ω) lies in an
apartment containing W . Then
νO(Σ) = 0 .
Proof. By translating to a parallel sector, one can assume that v = O.
Also, W is the union of two sector panels, which will be denoted by
[O, W +), [O, W −).
Given n ∈ Z+, let S +
triangles T +
W as illustrated below in Figure 6.
n , T −
n , T ⊥
n , S ⊥
n , S −
n denote the subsets of Sn consisting of
n respectively, lying in some apartment containing
..........
..........
..........
..........
..........
..........
..........
..........
..........
..........
..........
..........
.......................................................................................................................................................................................
T ⊥
.....................................................................................................................................................................................
.....................................................................................................................................................................................
..........
..........
......................................................................................................................................................................................
T +
T −
..........
..........
..........
..........
..........
..........
..........
..........
..........
..........
..........
..........
O
W
[O, W −)
21
W
[O, W +)
Figure 6.
Let S W
n = S +
n ∪ S −
n ∪ S ⊥
n . Then
(4.1)
Σ ⊂ [T ∈SW
n
ΩT .
The first step is to calculate the number of triangles in S W
n .
To do this, the number of possible choices for T + ∈ S +
n must be
determined. Refer to Figure 7 below.
cn
...............................................
...............................................
....................................................................................................................................................................................................................................................................................................................................
.............................................................................................................................................
.........................................................................................................................................................................................................................................................................................................................................
........................................................................................................................................................................................................................................................................................................................................................................................
...............................................
.........................................................................................................................................................................................................................................................................................................................................
...............................................
.........................................................................................................................................................................................................................................................................................................................................
.............................................................................................................................................
...............................................
.............................................................................................................................................
.............................................................................................................................................
.............................................................................................................................................
............................................................................................................................................................................................
.............................................................................................................................................
.............................................................................................................................................
...............................................
...............................................
...............................................
...............................................
.............................................................................................................................................
........................................................................................................................................................................................................................................................................................................................................................................................
...............................................
.............................................................................................................................................
...............................................
...............................................
...............................................
2 2 2 2 2 2 2 2
c2
...............................................
...............................................
.............................................................................................................................................
...............................................
.............................................................................................................................................
.............................................................................................................................................
.............................................................................................................................................
.............................................................................................................................................
...............................................
...............................................
.............................................................................................................................................
...............................................
...............................................
.............................................................................................................................................
1 1 1 1 1 1 1 1 1 1
c1
O
[O, W +)
Figure 7.
There are (q + 1) possible choices for the chamber c1. This choice
then determines all the other chambers labeled 1 which lie in the convex
hull of c1 and [0, W +). There are then q choices for the chamber c2.
This choice now determines all chambers labeled 2 which are in the
convex hull of c2 and all the other chambers previously determined.
Continue in this way until the chamber cn is reached, thereby deter-
n . There are thus (q + 1)qn−1 choices
mining the whole triangle T + ∈ S +
for T +.
Now each triangle T + ∈ S +
T − ∈ S −
convex hull of S +
n , T ⊥ ∈ S ⊥
n determines a unique pair of triangles
n subject to the condition that T −, T ⊥ lie in the
n ∪ W . Conversely such T − or T ⊥ determine T +
22
PAUL CUTTING AND GUYAN ROBERTSON
uniquely. Hence the sets S +
It follows that
n , S −
n , S ⊥
n have the same number of elements.
S W
n = 3(q + 1)qn−1 .
Since the sets ΩT , T ∈ Sn have equal measure and partition Ω, it
follows from Lemma 4.2 and equation (4.1) that
νO(Σ) ≤ νO
[T ∈SW
n
ΩT
3(q + 1)qn−1
=
(q2 + q + 1)(q + 1)q3n−3
→ 0
as n → ∞ .
Thus νO(Σ) = 0.
(cid:3)
References
[Br]
[Ca1]
[Ca2]
[Ca3]
K. Brown. Buildings. Springer-Verlag, New York 1989.
D. I. Cartwright, Groups acting simply transitively on the vertices of a
building of type An, Groups of Lie type and their Geometries, W. M.
Kantor and L. Di Martino, editors, 43 -- 76. Cambridge University Press,
1995.
D. I. Cartwright, A brief introduction to buildings, Harmonic Functions
on Trees and Buildings (New York 1995), 45 -- 77, Contemp. Math. 206,
Amer. Math. Soc., 1997.
D. I. Cartwright, Harmonic functions on buildings of type An, Random
Walks and Discrete Potential Theory (Cortona 1997), 104 -- 138, Symposia
Mathematica, Vol XXXIX, Cambridge University Press 1999.
[CMS] D. I. Cartwright, W. M lotkowski, and T. Steger. Property (T) and A2
groups, Ann. Inst. Fourier 44 (1994), 213 -- 248.
[CS]
[CMSZ] D. I. Cartwright, A. M. Mantero, T. Steger, and A. Zappa, Groups acting
simply transitively on the vertices of a building of type A2, I & II, Geom.
Ded. 47 (1993), 143 -- 166 and 167 -- 226.
D. I. Cartwright, and T. Steger, A family of An groups, Israel J. Math.
103 (1998), 125 -- 140.
A. Connes, Une classification des facteurs de type III., Ann. Scient. Ec.
Norm. Sup. 6 (1973), 133 -- 252.
T. Hamachi and M. Osikawa, Ergodic Groups of Automorphisms and
Krieger's Theorems, Seminar on Mathematical Sciences No. 3, Keio Uni-
versity, Japan, 1981.
[HO]
[C]
[Mar] G. A. Margulis, Discrete subgroups of semisimple Lie groups, Springer-
[RR1]
[RR2]
Verlag, Berlin, 1991.
J. Ramagge and G. Robertson, Factors from trees, Proc. Amer. Math. Soc.
125 (1997), 2051 -- 2055.
J. Ramagge and G. Robertson, Triangle buildings and actions of type
III1/q2 , J. Func. Anal. 140 (1996), 472 -- 504.
23
[R]
[S]
[St]
[Su]
[Sv]
[Zi]
M. A. Ronan, Lectures on Buildings, Perspectives in Math. 7, Academic
Press, London 1989.
R. J. Spatzier, An example of an amenable action from geometry, Ergod.
Th. & Dynam. Sys. 7 (1987), 289 -- 293.
T. Steger, Local fields and buildings, Harmonic Functions on Trees and
Buildings (New York 1995), 79 -- 107, Contemp. Math. 206, Amer. Math.
Soc., 1997.
V. S. Sunder, An invitation to von Neumann Algebras, Universitext,
Springer-Verlag, New York 1987.
T. Svenson, Groups acting simply transitively on An buildings, Ph.D.
Thesis, University of Sydney, 1998.
R. L. Zimmer, Ergodic Theory and Semisimple Groups, Birkhauser, Boston
1985.
Mathematics Department, University of Newcastle, Callaghan, NSW
2308, Australia
E-mail address: [email protected]
E-mail address: [email protected]
|
1303.4483 | 2 | 1303 | 2013-05-29T06:22:33 | Purely infinite partial crossed products | [
"math.OA"
] | Let (A,G,\alpha) be a partial dynamical system. We show that there is a bijective correspondence between G-invariant ideals of A and ideals in the partial crossed product A xr G provided the action is exact and residually topologically free. Assuming, in addition, a technical condition---automatic when A is abelian---we show that A xr G is purely infinite if and only if the positive nonzero elements in A are properly infinite in A xr G. As an application we verify pure infiniteness of various partial crossed products, including realisations of the Cuntz algebras O_n, O_A, O_N, and O_Z as partial crossed products. | math.OA | math |
PURELY INFINITE PARTIAL CROSSED PRODUCTS
THIERRY GIORDANO1 AND ADAM SIERAKOWSKI2
Abstract. Let (A, G, α) be a partial dynamical system. We show that
there is a bijective correspondence between G-invariant ideals of A and
ideals in the partial crossed product A ⋊α,r G provided the action is
exact and residually topologically free. Assuming, in addition, a techni-
cal condition -- automatic when A is abelian -- we show that A ⋊α,r G is
purely infinite if and only if the positive nonzero elements in A are prop-
erly infinite in A ⋊α,r G. As an application we verify pure infiniteness
of various partial crossed products, including realisations of the Cuntz
algebras On, OA, ON, and OZ as partial crossed products.
1. Introduction
In the theory of operator algebras, the crossed product construction has
been one of the most important and fruitful tools both to construct examples
and to describe the internal structure of operator algebras (in particular the
von Neumann algebras).
Partial actions of a discrete group on C∗-algebras and their associated
crossed products were gradually introduced in [14] and [35], and since then
developed by many authors. Several important classes of C∗-algebras have
been realised as crossed products by partial actions, including in particular
AF-algebras, Cuntz-Krieger algebras, Bunce-Deddens algebras, among oth-
ers (see for example [5, 12, 13, 15, 16, 19, 26]). The description of C∗-algebras
as partial crossed products has also proved useful for the computation of
their K-theory.
In this paper we pursue the study of partial C∗-dynamical systems and
their crossed products associated. We begin by recalling (in Section 2 and
Appendix A.1) the construction of the partial crossed product A ⋊α,r G
associated to a partial action α (a compatible collection of isomorphisms
αt : Dt−1 → Dt, t ∈ G of ideals in A).
In Section 3, we study the ideal structure of partial crossed products,
generalising the results on C∗-dynamical systems obtained by the second
author in [41]. Recall that a partial action α on a C∗-algebra A has the the
Date: May 29 2013.
1Research supported by NSERC, Canada.
2Communicating author. Research supported by Professor Collins, Handelman and
Giordano, Canada.
1
2
THIERRY GIORDANO AND ADAM SIERAKOWSKI
intersection property if every nontrivial ideal in A ⋊α,r G intersects A non-
trivially. Then (Definition 3.1) we say that a partial C∗-dynamical system
(A, G, α) has the residual intersection property if for every G-invariant ideal
I in A, the induced partial action of G on A/I has the intersection property.
We establish in Theorem 3.2 a one-to-one correspondence between ideals in
A ⋊α,r G and G-invariant ideals of A provided -- in fact if and only if -- the
partial action α is exact and has the residual intersection property. When
the partial action α is minimal (no nontrivial G-invariant ideals in A), then
the exactness (any nontrivial G-invariant ideal I in A induces a short exact
sequence at the level of reduced crossed products) is automatic.
In [19], having defined a topologically free partial action (by partial home-
omorphisms) on a locally compact space X, Exel, Laca, and Quigg proved
the simplicity of the partial crossed product C0(X) ⋊α,r G under the pres-
ence of minimality and topological freeness of α. In [33], Lebedev extended
the definition of topological freeness to non-commutative partial actions and
showed that a topologically free partial action has always the intersection
property. With Lebedev's result, we recover in Corollary 2.9 the theorem of
Exel, Laca, and Quigg.
Theorem 3.2 allows us also to extend Echterhoff and Laca's work on
crossed products: We show that the canonical map J 7→ J ∩ A between
ideals in A ⋊α,r G and G-invariant ideals of A restricts to a continuous map
from the space of prime ideals of A ⋊α,r G to the space of G-prime ideals in
A. This restriction is a homeomorphism provided α is exact and residually
topologically free, where residual topological freeness is an ideal related ver-
sion of topological freeness. When A is separable and abelian we show that
the space of prime ideals of A ⋊α,r G is homeomorphic to the quasi-orbit
space of Prim A.
In Section 4 we generalise some of the main results in [39], by Rørdam and
the second named author, to partial C∗-dynamical systems. In particular we
give sufficient conditions for a partial crossed products to be purely infinite
in the sense of Kirchberg and Rørdam. One of the keys assumptions of The-
orem 4.2 goes back to Elliott's notion of proper outerness (an automorphism
α of A is properly outer if kαI − βk = 2 for every α-invariant ideal I in
A and any inner automorphism β of I). Based on Kishimoto's work, Olsen
and Pedersen proved that an automorphism α of a separable C∗-algebra A
is properly outer if and only if inf{kx(aδ)xk : x ∈ B+, kxk = 1} = 0 for
every a ∈ A, and every nonzero hereditary C∗-algebra B in A, where δ is
the unitary implementing α, i.e, α(a) = δaδ∗ in A ⋊α Z. Moreover Archbold
and Spielberg proved that topological freeness of an action α ensures that
αt, t 6= e is properly outer. In Proposition 3.10, generalising these results,
we prove the equivalence of the following statements for a partial action α
on an abelian C∗-algebra A:
(i) α is topologically free,
(ii) kαtI − idI k = 2 for every αt-invariant ideal I in Dt−1 (and t 6= e),
PURELY INFINITE PARTIAL CROSSED PRODUCTS
3
(iii) inf{kx(aδt)xk : x ∈ B+, kxk = 1} = 0 for every a ∈ Dt, and every
nonzero hereditary C∗-algebra B in A (and t 6= e).
Without assuming that A is commutative and under the assumption of
property (iii) for actions of G on A/I for any G-invariant ideal I in A and
exactness of α we once again obtain a one-to-one correspondence between
ideals in A ⋊α,r G and G-invariant ideals in A. However, in this context
we can state in Theorem 4.2 sufficient and necessary conditions for pure
infiniteness of the partial crossed products: A ⋊α,r G is purely infinite if and
only if every nonzero positive element in A is properly infinite in A ⋊α,r G
(under presence of ideal property of A).
In the second part of Section 4, we study pure infiniteness of partial
crossed products A ⋊α,r G when A is abelian. We establish a geometrical
condition sufficient -- and sometimes also necessary -- to obtain pure infinite-
ness of partial crosses products. Specifically we show in Theorem 4.4 that
an exact and residually topologically free partial action on a totally dis-
connected locally compact Hausdorff space X gives a purely infinite partial
crossed product C0(X) ⋊α,r G provided that every compact and open subset
of X is (G, τX )-paradoxical (Definition 4.3).
In Section 5, we apply our results to the a variety of know examples of
partial crossed products:
(i) Hopenwasser constructs a partial action of the semidirect product
Qn ×δ Z of n-adic rationals by the integers Z on the Cantor set XC such that
the partial crossed product C(XC) ⋊α (Qn ×δ Z) is isomorphic to the Cuntz
algebra On. We verify that for this (exact and residually topologically free)
action the compact and open subset of XC are (Qn ×δ Z, τXC )-paradoxical.
(ii) For a {0, 1}-valued n by n matrix A = [aij] with no zero rows Exel,
j = 1,Pj aijsjs∗
Laca, and Quigg realised OA = C∗(s1, . . . , sn : Pj sjs∗
j =
s∗
i si) as a partial crossed product C(XA) ⋊α Fn, where Fn denotes the free
group of rank n. The algebra OA was defined by Astrid an Huef and Raeburn
as an universal analogue of the Cuntz-Krieger algebra. When α is exact and
residually topologically free we prove that C(XA) ⋊α Fn is purely infinite if
and only if the compact and open subset of XA are (Fn, τXA)-paradoxical.
We use graph C∗-algebraic results by Raeburn, Kumjian, Pask, Raeburn,
Renault, Mann, Sutherland, Hong and Szymanski among others.
(iii) Boava and Exel showed that for each integral domain R with finite
quotients R/(m), m 6= 0, the semidirect product K ⋊ K × of K by K\{0}
(where K is the field of fractions of R) acts on the Cantor set XR such
that the partial crossed product C(XR) ⋊α (K ⋊ K ×) is isomorphic to the
regular C∗-algebra A[R] of R. We verify that for this (exact and residually
topologically free) action the compact and open subset of XR are (K ⋊
K ×, τXR)-paradoxical.
In [1], Abadie described a class of partial crossed products C0(X) ⋊α,r G
Morita-Rieffel equivalent to ordinary crossed products.
In Section 6, we
prove that for such a partial crossed product C0(X) ⋊α,r G if the compact
4
THIERRY GIORDANO AND ADAM SIERAKOWSKI
and open subsets in the spectrum of the the original C∗-algebra of the partial
crossed product are paradoxical, then the same property must holds for the
corresponding (ordinary) crossed product.
Much of the work described in this paper was done while the second
author held a postdoctoral fellowship at the University of Ottawa. He wishes
to thank the members of the Department of Mathematics and Statistics at
the University of Ottawa for their warm hospitality.
2. Definition of partial crossed product.
Let A be a C∗-algebra and G be a discrete group. We recall the definition
of a partial dynamical system and the corresponding partial crossed products
(see Appendix A.1 for more details).
Definition 2.1. Let A be a C∗-algebra and G be a discrete group. A partial
action of G on A, denoted by α, is a collection (Dt)t∈G of closed two-sided
ideals of A and a collection (αt)t∈G of ∗-isomorphisms αt : Dt−1 → Dt such
that
(i) De = A, where e represents the identity element of G;
(ii) α−1
(iii) αt ◦ αs(x) = αts(x), ∀ x ∈ α−1
s (Ds ∩ Dt−1) ⊆ D(ts)−1 ;
s (Ds ∩ Dt−1).
The triple (A, G, α) is called a partial dynamical system. The equivalence
between this definition of partial action by Dokuchaev and Exel in [9] and
the original definition of McClanahan ([35]) was proven in [37]. For the case
when A is abelian we refer to [1, 12, 19].
Definition 2.2. Let (A, G, α) be a partial dynamical system. Let L be the
normed ∗-algebra of the finite formal sums Pt∈G atδt, where at ∈ Dt. The
operations and the norm in L are given by
(atδt)(asδs) = αt(αt−1 (at)as)δts,
kXt∈G
atδtk = Xt∈G
(atδt)∗ = αt−1(a∗
katk.
t )δt−1 ,
Let Bt denote the vector subspace Dtδt of L. The family (Bt)t∈G generates
a Fell bundle. The full crossed product A ⋊α G and the reduced crossed
product A ⋊α,r G are, respectively, the full and the reduced cross sectional
algebras of (Bt)t∈G. Both crossed products are completions of L with respect
to a certain C∗-norm. We recall the construction of these crossed products
in Appendix A.1.
We suppress the canonical inclusion map A → A ⋊α G, a 7→ aδe and view
A as a sub-C∗-algebra of A ⋊α G. All ideals (throughout out this paper) are
assumed to be closed and two-sided. The set τX denotes the topology of a
topological space X. The C ∗-algebra of continuous functions vanishing at
infinity on a locally compact Hausdorff space X is denoted by C0(X). Every
abelian C ∗-algebra arises in this form. When the algebra is unital then X
is compact and we emphasise this fact by writing it as C(X).
PURELY INFINITE PARTIAL CROSSED PRODUCTS
5
3. Ideal structure of partial crossed product.
Before stating our results on the ideal structure of a partial crossed prod-
uct generalising [41] we need a few definitions.
Definition 3.1. Let (A, G, α) be a partial dynamical system. Then
(i) A closed two-sided ideal I of A is G-invariant provided that αt(I ∩
Dt−1) ⊆ I for every t ∈ G.
(ii) The partial action is exact if every G-invariant ideal I of A induces
a short exact sequence
0
/ I ⋊α,r G
/ A ⋊α,r G
/ A/I ⋊α,r G
/ 0
at the level of reduced crossed products.
(iii) The partial action has the residual intersection property if for every
G-invariant ideal I of A the intersection of A/I with any nonzero
ideal in A/I ⋊α,r G is nonzero.
Theorem 3.2. Let (A, G, α) be a partial dynamical system. There is a one-
to-one correspondence between ideals in A ⋊α,r G and G-invariant ideals of
A if and only if the partial action is exact and has the residual intersection
property.
Proof. Sufficiency: Suppose that α is exact and has the residual intersection
property. Let EA : A ⋊α,r G → A denote the usual conditional expectation
on the crossed product (see Appendix A.1). Let Ideal[S] denote the smallest
ideal in A ⋊α,r G generated by S ⊆ A ⋊α,r G. Let ϕ denote the map
J 7→ J ∩ A from the ideals in A ⋊α,r G into G-invariant ideals in A. Using
that Ideal[I] ∩ A = I, for any G-invariant ideals I of A, cf. Proposition A.4,
we conclude that ϕ is surjective. To show ϕ is injective it is enough to show
that J = Ideal[EA(J )] for every ideal J in A ⋊α,r G. If we have two ideals
J 1, J 2 of A ⋊α,r G with the same intersection it then follows that
J 1 = Ideal[EA(J 1)∩A] ⊆ Ideal[J 1∩A] = Ideal[J 2∩A] ⊆ Ideal[EA(J 2)] = J 2,
using that EA(J 1) ⊆ Ideal[EA(J 1)] = J 1, and J 2∩A ⊆ EA(J 2) (cf. Propo-
sition A.4). Fix any ideal J of A ⋊α,r G.
(i) J ⊆ Ideal[EA(J )]: Let I denote the (smallest) G-invariant ideal in A
generated by EA(J ). By Proposition A.3 we have the commuting diagram
0
0
/ I ⋊α,r G ι
A ⋊α,r G
ρ
A/I ⋊α,r G
EI
/ I
EA
/ A
EA/I
/ A/I
0
/ 0
For each x ∈ J + we have EA(x) ∈ I and hence that EA/I (ρ(x)) = EA(x) +
I = 0. By exactness ker ρ ⊆ Ideal[I] (not true in general). Hence x ∈ ker ρ =
Ideal[I] = Ideal[EA(J )]. Since every element in J is a linear combination
of positive elements J ⊆ Ideal[EA(J )].
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
6
THIERRY GIORDANO AND ADAM SIERAKOWSKI
(ii) Ideal[EA(J )] ⊆ J : Let I denote the intersection J ∩ A. By exactness
we have a short exact sequence
0
/ I ⋊α,r G ι
/ A ⋊α,r G
ρ
/ A/I ⋊α,r G
/ 0
With K := I ⋊α,r G ⊆ J we can therefore make the following identifications:
ρ(J ) = J /K, A/I = A/(K ∩ A) = (A + K)/K.
Suppose that J /K and (A + K)/K has a nonzero intersection. Then there
exist j ∈ J and a ∈ A such that j + K = a + K 6= K. Since K ⊆ J it
follows that a ∈ J and hence that a ∈ J ∩ A = I ⊆ K. But then a + K = K
giving a contradiction. We conclude that ρ(J ) ∩ A/I = 0. The residual
intersection property implies that ρ(J ) = 0. By Proposition A.3 we have
the commuting diagram
0
0
/ I ⋊α,r G ι
A ⋊α,r G
ρ
A/I ⋊α,r G
EI
/ I
EA
/ A
EA/I
/ A/I
0
/ 0
For each x ∈ J we have that EA/I (ρ(x)) = EA(x) + I = 0. Hence EA(x) ∈
I = J ∩ A. We conclude that Ideal[EA(J )] ⊆ J .
Necessity: Suppose that ϕ is bijective, where ϕ denotes the map J 7→
J ∩ A from the ideals in A⋊α,r G into G-invariant ideals in A. As previously
let EA : A ⋊α,r G → A denote the conditional expectation on the crossed
product.
(iii) Exactness: Fix any G-invariant ideal I of A. By Proposition A.3 we
have the commuting diagram
0
0
/ I ⋊α,r G ι
A ⋊α,r G
ρ
A/I ⋊α,r G
EI
/ I
EA
/ A
EA/I
/ A/I
0
/ 0
By assumption J := ker ρ has the form (J ∩ A) ⋊α,r G, cf. Proposition A.4.
This implies that EA/I (ρ(J )) = EA(J ) + I = 0. Hence EA(J ) ⊆ I. By
Proposition A.4 we also have that J ∩ A ⊆ EA(J ). We conclude
ker ρ = (J ∩ A) ⋊α,r G ⊆ Ideal[EA(J )] ⊆ I ⋊α,r G.
(iv) Residual intersection property: Fix any G-invariant ideal I of A and
any ideal J of A/I ⋊α,r G with zero intersection with A/I. By Proposition
A.3 we have the commuting diagram
0
0
/ I ⋊α,r G ι
A ⋊α,r G
ρ
A/I ⋊α,r G
EI
/ I
EA
/ A
EA/I
/ A/I
0
/ 0
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
PURELY INFINITE PARTIAL CROSSED PRODUCTS
7
By assumption J 1 := ρ−1(J ) has the form (J 1 ∩ A) ⋊α,r G, cf. Proposition
A.4. Hence
EA/I (J ) = EA/I (ρ(J 1)) = EA(J 1) + I = (J 1 ∩ A) + I
= EA(J 1 ∩ A) + I = EA/I (ρ(J 1 ∩ A)) ⊆ EA/I (ρ(J 1) ∩ ρ(A))
= EA/I (J ∩ A/I) = J ∩ A/I = 0
By the faithfulness of the conditional expectation (on positive elements) we
conclude that J = 0.
(cid:3)
Remark 3.3. Exactness of a partial action is somehow mysterious because
there are no concrete examples of a partial action that is not exact (even
though existence has been established). Nevertheless, exactness plays an im-
portant role as we have seen above. Here is another result ([40, Proposition
2.2.4]), proved by the second named author, relying heavily on exactness:
Let (A, G, α) be a partial dynamical system. The action α is exact and
A ⋊α,r G ∼= A ⋊α G if, and only if, A/I ⋊α,r G ∼= A/I ⋊α G for every
G-invariant ideal I in A.
If A is a C∗-algebra, let Prim A denote its primitive ideal space and A it
spectrum. Moreover, if J is a closed two-sided ideal of A, then supp J will
denote the subset {I ∈ Prim A : J * I} and AJ = {[π] ∈ A : π(J ) 6= 0}.
Following Lebedev in [33], we define the map θt : supp Dt−1 → supp Dt as
follows; for any point x ∈ supp Dt−1 such that x = ker π, where [π] ∈ ADt−1 ,
we set
θt(x) = ker (π ◦ αt−1),
t ∈ G.
Notice that we then have that (θt)t∈G is an partial action of G by partial
homeomorphisms of Prim A (see Appendix A.2 for more details).
Definition 3.4. Let (A, G, α) be a partial dynamical system.
(i) The partial action α is topologically free if for any finite set F in
G \ {e} the union
[t∈F (cid:8)x ∈ supp Dt : θt(x) = x(cid:9)
has empty interior, cf. [33].
(ii) The partial action α is residually topologically free if the induced
action of G on A/I is topologically free for every G-invariant ideal
I of A.
In [33], Lebedev shows that a residually topologically free action has
always the residual intersection property (see Appendix A.2 for details).
Hence, we have
Corollary 3.5. Let (A, G, α) be a partial dynamical system. Suppose that
the action is exact and residually topologically free. Then there is a one-
to-one correspondence between ideals in A ⋊α,r G and G-invariant ideals in
A.
8
THIERRY GIORDANO AND ADAM SIERAKOWSKI
There is a useful consequence of Corollary 3.5 regarding G-prime ideals.
Recall that a G-invariant ideal I in a C∗-algebra A is called G-prime (resp.
prime if G is the trivial group) if for any pair of G-invariant ideals J , K of A
with J ∩ K ⊆ I we have J ⊆ I or K ⊆ I, cf. [10]. Let I(A) denote the set
of closed two-sided ideals in A. Imposing that the set {J ∈ I(A) : I 6⊆ J }
is open for any I ∈ I(A) defines a sub-basis for the Fell-topology on I(A).
This topology induces topologies on the set of prime and G-prime ideals in
I(A).
Corollary 3.6. Let (A, G, α) be a partial dynamical system. The map
J 7→ J ∩ A, J ∈ I(A ⋊α,r G),
restricts to a continuous map from the space of prime ideals of A ⋊α,r G to
the space of G-prime ideals in A. Moreover, if the action of G on A is exact
and residually topologically free this restriction is a homeomorphism.
Proof. The second statement is contained in [10] for ordinary crossed prod-
ucts and it is evident that the proof generalizes. The same hold for the fact
that the map J 7→ J ∩ A is continuous. We conclude that the restriction is
continuous (provided it is well defined).
The fact that it is well defined follows from: The ideal J ∩ A is G-prime
for any prime ideal J in A⋊α,r G. The proof is contained in [10] omitting one
detail. If I and K are G-invariant ideals in A then (I ⋊α,r G) ∩ (K ⋊α,r G) ⊆
(I ∩ K) ⋊α,r G. To see this we use that the intersection of two closed two-
sided ideals equals their product and an approximation argument. It is then
sufficient to show that if aδt ∈ I ⋊α,r G and bδt ∈ K ⋊α,r G then (aδt)(bδs) ∈
(IK) ⋊α,r G. But this is evident from (aδt)(bδs) = αt(αt−1 (a)b)δts.
(cid:3)
Remark 3.7. If a partial action of a discrete group G on a C∗-algebra A is
exact then Ti(I i ⋊α,r G) = (Ti Ii) ⋊α,r G for any family (I i) of G-invariant
ideals, cf. [40].
G on Prim A define Gx := St∈G for which x∈supp Dt−1
Let x be a fixed element in Prim A. Using the partial action (θt)t∈G of
{θt(x)}. The then quasi-
orbit space O(Prim A) of Prim A is defined as the quotient space Prim A/ ∼
by the equivalence relation
x ∼ y ⇔ Gx = Gy.
Recall that for A = C0(X) the partial action α is induced by a partial action
θ of G on X, i.e., a collection of open sets (Ut)t∈G and a collection (θt)t∈G
of homeomorphisms θt : Ut−1 → Ut such that Ue = X and θst extends θs ◦ θt,
cf. [35, 16, 37] and Appendix A.2. Then α is given by αt(f )(x) := f (θt−1(x)),
f ∈ C0(Ut−1 ). So, here the ideals are Dt = C0(Ut). As the canonical
homeomorphism from Prim A to X is G-equivariant we identify the actions
on Prim A and X. The next corollary and is proof is a generalisation of [10,
Lemma 2.5 and Corollary 2.6].
PURELY INFINITE PARTIAL CROSSED PRODUCTS
9
Corollary 3.8. Let (A, G, α) be a partial dynamical system with A abelian
and separable. Suppose that the action of G on A is exact and residually
topologically free. Then the space of prime ideals of A ⋊α,r G is homeomor-
phic to the quasi-orbit space of Prim A.
Proof. We show that the following map is a homeomorphism:
[x] 7→ Ix ⋊α,r G,
Ix := \I∈Gx
I,
[x] ∈ O(Prim A).
Well defined: Fix any [x] ∈ TI∈Gx I. It is evident that Ix is a closed
two-sided ideal in A. Let π : A → B(H) be the irreducible representation
corresponding to x, i.e. x = ker π. Since A = C0(X) for some locally
compact Hausdorff space [4, II.6.2.9] ensures that π(f ) = f (y) for some
y ∈ X, so x = C0(X \ {y}). Hence
θt(x) = ker(π ◦ αt−1) = {f : αt−1(f )(y) = 0} = C0(X \ {θt(y)}).
By continuity we obtain that Ix = C0(X \ Gy). This ideal is G-invariant
since Gy is G-invariant (and the complement of a G-invariant set is also G-
invariant). Let us verify Gy is G-invariant for completeness. Fix any s ∈ G
and any θt(y) ∈ Gy ∩ Us−1, where Ds = C0(Us). Since θt(y) ∈ Ut ∩ Us−1
[37, Lemma 1.2] ensures that y ∈ θt−1(Ut ∩ Us−1) = Ut−1 ∩ U(st)−1.
In
particular θs(θt(y)) = θst(y). We conclude θs(Gy ∩ Us−1) ⊆ Gy making Gy
invariant. The fact that Gy is invariant is a simple limit argument used on
Gy ∩ Us−1 ⊆ Gy ∩ Us−1. Let us verify that Ix is G-prime. Suppose that
U, V are closed G-invariant subsets of X. If C0(X \ U ) ∩ C0(X \ V ) ⊆ Ix it
follows that y ∈ U ∪ V , hence either C0(X \ U ) ⊆ Ix or C0(X \ V ) ⊆ Ix. It
now follows from Corollary 3.6 that Ix ⋊α,r G is prime.
Surjectivity ([10, Lemma 2.5]): Let J be any prime ideal in A ⋊α,r G.
By Corollary 3.6 the ideal I = J ∩ A is G-prime. As I is G-invariant
I = C0(X \ V ) for some nonempty closed G-invariant set V in X. Let
ρ : Prim A → T denote the quotient map into the quasi-orbit space T :=
O(Prim A). To show that V = Gx for some x ∈ X we verify that F := ρ(V )
is the closure of a single point in T . Notice that T is totally Baire (every
intersection of an open and a closed subset is a Baire space) and second
countable. For such a space a non-empty closed subset F is the closure of
a simple point if and only if F is not a union of two proper closed subsets,
cf. [24, Lemma p. 222]. However such two proper closed subsets give raise
to two closed G-invariant subsets U1, U2 ( V with U1 ∪ U2 = V and hence
two G-invariant ideals I1, I1 ) I with I 1 ∩ I2 = I. This contradicts G-
primeness of I.
Injectivity: Fix any two prime ideals Ix ⋊r,α G and Iy ⋊r,α G in A ⋊r,α G
such that Ix⋊r,α G = I y ⋊r,α G. By Proposition A.4 we obtain C0(X \Gx) =
Ix = Iy = C0(X \ Gy). We conclude that x ∼ y.
(cid:3)
10
THIERRY GIORDANO AND ADAM SIERAKOWSKI
Corollary 3.9 (Lebedev [33]). Let (A, G, α) be a partial dynamical system.
Suppose that the action is minimal (i.e, A does not contain any non-trivial
G-invariant ideals) and topologically free. Then the crossed product A⋊α,r G
is simple.
The notation of topological freeness for partial actions is well known. We
recall it in Appendix A.2 and show that the following equivalent conditions
hold:
Proposition 3.10. Let (A, G, α) be a partial dynamical system with A
abelian. Then the following properties are equivalent:
(i) α is topologically free
(ii) kαtK − idKk = 2 for every αt-invariant ideal K in Dt−1 (and t 6= e)
(iii) inf{kxαt(x)k : x ∈ K+, kxk = 1} = 0 for every nonzero ideal K in Dt−1
(and t 6= e)
(iv) inf{kx(aδt)xk : x ∈ B+, kxk = 1} = 0 for every a ∈ Dt, and every
nonzero hereditary C∗-algebra B in A (and t 6= e)
In [11], Elliott defines an automorphism α of a C∗-algebra A to be properly
outer if for any nonzero α-invariant ideal I of A and any inner automorphism
β of I,
kαI − βk = 2.
Proper outerness for dynamical systems has been vastly studied in [31, 36,
3, 40, 41] among others. We will address in one of our upcoming works how
one might generalise the notion proper outerness beyond ordinary crossed
products.
In the rest of the section, we generalize [19], Theorem 2.6, to the non-
abelian case using condition (iv) of Proposition 3.10. We first need the
following lemma.
Lemma 3.11. Let (A, G, α) be a partial dynamical system. Suppose that
for every t 6= e, every a ∈ Dt, and every nonzero hereditary C∗-algebra B in
A
inf{kx(aδt)xk : x ∈ B+, kxk = 1} = 0.
Then for every b ∈ (A ⋊α,r G)+ and every ε > 0 there exist a positive
contraction x ∈ A satisfying
kxEA(b)x − xbxk < ε,
kxEA(b)xk > kEA(b)k − ε.
Proof. Fix b ∈ (A ⋊α,r G)+ and ε > 0.
(i) We may assume b ∈ L: Since L is dense in A ⋊α,r G there exist
c ∈ (A ⋊α,r G)+ ∩ L such that kc − bk < ε. Find a positive contraction x ∈ A
PURELY INFINITE PARTIAL CROSSED PRODUCTS
11
satisfying kxEA(c)x − xcxk < ε, and kxEA(c)xk > kEA(c)k − ε. Then
kxEA(b)x − xbxk ≤ kxEA(b)x − xEA(c)xk + kxEA(c)x − xcxk
+ kxcx − xbxk < 3ε,
kEA(b)k < kEA(b − c)k + ε + kxEA(c − b)xk + kxEA(b)xk
< kxEA(b)xk + 3ε.
(ii) We may assume EA(b) has norm one: If b = 0 then any positive
kEA(b)k using EA is faithful.
kEA(b)k ,
contraction x ∈ A works. For b 6= 0 define c :=
Find a positive contraction x ∈ A satisfying kxEA(c)x − xcxk <
and kxEA(c)xk > kEA(c)k −
kEA(b)k . Then
ε
b
ε
kxEA(b)x − xbxk < ε,
kxEA(b)xk > kEA(b)k − ε.
(iii) It is enough to show that: For every ε > 0, every b0 ∈ A+, with
kb0k = 1, every finite set F ⊆ G \ {e}, and every sequence of elements
(bt)t∈F , with bt ∈ Dt, there exist x ∈ A+ such that
kxk = 1,
kxb0xk > kb0k − ε,
kx(btδt)xk < ε,
t ∈ F.
By (i) − (ii) we may assume that b = b0 + Pt∈F btδt for some finite set
F ⊆ G \ {e}, where b0 = EA(b) is positive (since EA is positive) and has
norm one. Using the assumption on b0 and the sequence (bt)t∈F choose
x ∈ A+ such that
kxk = 1,
kxb0xk > kb0k − ε,
kx(btδt)xk < ε,
t ∈ F.
We conclude that
kxEA(b)xk > kEA(b)k − ε,
kxEA(b)x − xbxk = kx(b − b0)xk ≤ Xt∈F
kx(btδt)xk < F ε.
(iv) Finishing the proof: Fix ε > 0, b0 ∈ A+, with kb0k = 1, a finite
set F ⊆ G \ {e}, and a sequence of elements (bt)t∈F , with bt ∈ Dt. Let
f : [0, 1] → [0, 1] be a continuous increasing function taking the value zero
on [0, 1 − ε] and one on [1 − ε
2 , 1]. Define x0 := f (b0) and set
B1 := {x ∈ A : xx0 = x0x = x}.
If g : [0, 1] → [0, 1] is a continuous increasing function equal to zero on [0, 1−
ε
2 ] and one on [1 − ε
4 , 1] then gf = f g = g and kg(b0)k = 1. We conclude
that B1 is nonzero. Using the C∗-norm identify it follows that L = {x ∈
A : x∗x ∈ B1} is a closed left ideal in A. Verifying that B1 = L ∩ L∗ we
obtain that B1 is hereditary. Write F := {t1, . . . , tn}. Since bt1 ∈ Dt1 and
B1 is nonzero hereditary in A
inf{kx(bt1 δt1 )xk : x ∈ (B1)+, kxk = 1} = 0.
Select x′
1k < ε. Let h : [0, 1] →
[0, 1] be a continuous increasing function equal to the identity on [0, 1 − ε]
1 ∈ (B1)+ such that kx′
1k = 1 and kx′
1(bt1 δt1)x′
12
THIERRY GIORDANO AND ADAM SIERAKOWSKI
and one on [1 − ε
2 , 1]. With x1 := h(x′
1) we have (since gh = hg = g) that
x1 ∈ (B1)+,
B2 := {x ∈ B1 : xx1 = x1x = x} ∋ g(x′
kx1k = 1,
1) 6= 0.
kx1(bt1 δt1 )x1k < (2 + kbt1 k)ε,
Since bt2 ∈ Dt2 and B2 is nonzero hereditary in A
inf{kx(bt2 δt2 )xk : x ∈ (B2)+, kxk = 1} = 0.
Repeating the procedure above we obtain a sequence of hereditary C∗-
algebras
B1 ⊇ B2 ⊇ · · · ⊇ Bn,
and positive elements of norm one in them, x1, x2, . . . , xn, fulfilling that
xi ∈ (Bi)+,
kxik = 1,
kxi(btiδti )xik < (2 + kbtik)ε.
With x := xn we have that xxi = xix = x for 0 ≤ i < n. It follows that
x ∈ A+,
kxk = 1,
kx(bti δti)xk < (2 + kbti k)ε,
i = 1, . . . , n.
Finally, since f (t)2t ≥ (1 − ε)f (t)2 for every t ∈ [0, 1], we get
xx0b0x0x ≥ (1 − ε)xx0x0x,
kxb0xk > kb0k − ε.
(cid:3)
Theorem 3.12. Let (A, G, α) be a partial dynamical system. Suppose that
for every t 6= e, every a ∈ Dt, and every nonzero hereditary C∗-algebra B in
A
inf{kx(aδt)xk : x ∈ B+, kxk = 1} = 0.
Then every nonzero ideal in A ⋊α,r G has a nonzero intersection with A.
Proof. The proof by Exel, Laca and Quigg in [19] for partial crossed product
with A abelian generalises to the non-abelian case by means of Lemma
3.11.
(cid:3)
Corollary 3.13. Let (A, G, α) be a partial dynamical system. Suppose that
the action is exact and that for every t 6= e, every G-invariant ideal I in A,
every a ∈ Dt/(Dt ∩ I), and every nonzero hereditary C∗-algebra B in A/I
inf{kx(aδt)xk : x ∈ B+, kxk = 1} = 0.
Then there is a one-to-one correspondence between ideals in A ⋊α,r G and
G-invariant ideals in A.
Proof. Using Theorem 3.12 we obtain that α has the residual intersection
property. The desired correspondence follows from Theorem 3.2.
(cid:3)
PURELY INFINITE PARTIAL CROSSED PRODUCTS
13
4. Pure infiniteness of partial crossed products
We recall the definition of purely infinite C∗-algebras with the ideal prop-
erty and the notion of paradoxical actions.
Let A be a C∗-algebra, and let a, b be positive elements in A. We say that
a is Cuntz below b, denoted a - b, if there exists a sequence (rn) in A such
that r∗
nbrn → a. More generally for a ∈ Mn(A)+ and b ∈ Mm(A)+ we write
a - b if there exists a sequence (rn) in Mm,n(A) with r∗
nbrn → a. For a ∈
Mn(A) and b ∈ Mm(A) let a ⊕ b denote the element diag(a, b) ∈ Mn+m(A).
A nonzero positive element a in A is properly infinite if a ⊕ a - a.
A C∗-algebra A is purely infinite if there are no characters on A and if
for every pair of positive elements a, b in A such that b belongs to the ideal
in A generated by a, one has b - a. Equivalently, a C∗-algebra A is purely
infinite if every non-zero positive element a in A is properly infinite, cf. [29,
Theorem 4.16].
Definition 4.1. A C∗-algebra A has the ideal property if projections in A
separate ideals in A, i.e. , whenever I, J are ideals in A such that I * J ,
then there is a projection in I \ (I ∩ J ).
The next result generalises the work in [39] on ordinary crossed products.
The proof is included for completeness.
Theorem 4.2. Let (A, G, α) be a partial dynamical system. Suppose that
the action is exact and that for every t 6= e, every G-invariant ideal I in A,
every a ∈ Dt/(Dt ∩ I), and every nonzero hereditary C∗-algebra B in A/I
inf{kx(aδt)xk : x ∈ B+, kxk = 1} = 0.
Suppose also that A has the ideal property. Then the following statements
are equivalent
(i) Every nonzero positive element in A is properly infinite in A ⋊α,r G.
(ii) The C∗-algebra A ⋊α,r G is purely infinite.
(iii) Every nonzero hereditary sub-C∗-algebra in any quotient of A ⋊α,r G
contains an infinite projection.
Proof. The implications (iii) ⇒ (ii) ⇒ (i) are valid for any partial dynami-
cal system, cf. [29]. For (i) ⇒ (iii) let J be any ideal in A ⋊α,r G and let B
be any nonzero hereditary sub-C∗-algebra in the quotient (A⋊α,r G)/J . We
show that B contains an infinite projection. By assumption on α and Corol-
lary 3.13 we have that (A ⋊α,r G)/J ∼= (A/I) ⋊α,r G for I = J ∩ A. Select
a nonzero positive element b in B such that kEA/I (b)k = 1. By Lemma 3.11
there exist a positive contraction x ∈ A/I satisfying
kxEA/I (b)x − xbxk < 1/4,
kxEA/I (b)xk > kEA/I (b)k − 1/4 = 3/4.
With a = (xEA/I (b)x − 1/2)+ we claim that 0 6= a - xbx - b. Indeed, the
element a is nonzero because kxEA/I (b)xk > 1/2, and a - xbx holds since
kxEA/I (b)x − xbxk < 1/2, cf. [38, Proposition 2.2]. By the assumption that
A has the ideal property we can find a projection q ∈ A that belongs to the
14
THIERRY GIORDANO AND ADAM SIERAKOWSKI
ideal in A generated by the preimage of a in A but not to I. Then q + I
belongs to the ideal in A/I generated by a, whence q+I - a - b in A/I ⋊r G
(because a is properly infinite by assumption), cf. [29, Proposition 3.5 (ii)].
From the comment after [29, Proposition 2.6] we can find z ∈ A/I ⋊r G
such that q + I = z∗bz. With v = b1/2z it follows that v∗v = q + I, whence
p := vv∗ = b1/2zz∗b1/2 is a projection in B, which is equivalent to q + I.
By the assumption q is properly infinite, and hence so is q + I (since the
relation a ⊗ a - a passes to quotients) and p.
(cid:3)
We now introduce a geometrical condition sufficient for the pure infinite-
ness of a large class of partial crossed product C∗-algebras. It is not clear if
this geometrical condition is also a necessary one.
Definition 4.3. Let (C0(X), G, α) be a partial dynamical system, together
with the corresponding collection of homeomorphisms (θt : Xt−1 → Xt)t∈G,
inducing α, cf. Appendix A.2. Let E denote a family of subsets of X.
A non-empty set V ⊆ X is called (G, E)-paradoxical
if there exist sets
V1, V2 . . . , Vn+m ∈ E and elements t1, t2, . . . , tn+m ∈ G such that
n
n+m
Vi =
[i=1
[i=n+1
Vi = V, Vi ⊆ Xt−1
i
, θti(Vi) ⊆ V, θtk (Vk)∩θtl(Vl) = ∅, k 6= l.
We let τX denote the topology of X.
Theorem 4.4. Let (C0(X), G, α) be a partial dynamical system. Suppose
that the action is exact and residually topologically free and that X is totally
disconnected. Suppose also that every compact and open subset of X is
(G, τX )-paradoxical. Then C0(X) ⋊α,r G is purely infinite.
Proof. The proof consist of two parts. First we show that for any (G, τX )-
paradoxical, compact and open subset U ⊆ X the projection 1U is properly
infinite in C0(X) ⋊α,r G. We then prove that proper infiniteness of such
projections is enough to ensure pure infiniteness of C0(X) ⋊α,r G. We do
not know if the second part follows from Theorem 4.2.
(i) Proper infiniteness: Recall that α is induced by a collection of open
sets (Xt)t∈G and a collection (θt)t∈G of homeomorphisms θt : Xt−1 → Xt
such that Xe = X and θst extends θs ◦ θt, cf. [35, 16, 37] and Appendix
A.2. The partial action α of G on C0(X) corresponding to θ is given by
αt(f )(x) := f (θt−1(x)), f ∈ C0(Xt−1 ). So, here the ideals are Dt = C0(Xt).
Let (Vi, ti)n+m
i=1 denote the system of open sets of U and elements in G wit-
i=1 and (hi)n+m
nessing the paradoxality of U . Find partitions of unity (hi)n
i=n+1
for U relative to the open covers (Vi)n
i=n+1, respectively. For each
i = 1, . . . , n + m we have that the (compact) support of hi lies in Vi ⊆ Xt−1
.
Hence hi ∈ Dt−1
), for i = 1, . . . , n + m, and
the elements
and we can define ai := αti(h1/2
i=1 and (Vi)n+m
i
i
i
n
n+m
x =
Xi=1
aiδti,
y =
Xi=n+1
aiδti,
p = 1U .
PURELY INFINITE PARTIAL CROSSED PRODUCTS
15
Using that (atδt)(asδs) = αt(αt−1(at)as)δts and (atδt)∗ = αt−1 (a∗
follows that
t )δt−1 it
i
(a∗
(aiδti)∗(ajδtj ) = (αt−1
i )δt−1
(αti (αt−1
= αt−1
(a∗
i aj)δt−1
= αt−1
tj
= (cid:26) hi
if i = j
if i 6= j.
0
i
i
i
i
i
)(ajδtj )
(a∗
i ))aj )δt−1
i gj
We obtain that x∗x = y∗y = p and y∗x = 0. Moreover, since
1U (aiδti) = αe(αe(1U )ai)δeti
= (1U ai)δti
= aiδti,
we have that px = x, py = y. This implies that xx∗ + yy∗ ≤ p, hence p is
properly infinite, cf. [29].
(ii) Pure infiniteness: This part follows by inspection of the proof of
Theorem 4.2. The difference is that we do not assume that every nonzero
element a ∈ A/I is properly infinite (in A/I ⋊α,r G) and we can therefore
not conclude that the selected nonzero projection q + I in the ideal in A/I
generated by a fulfils that q + I - a (in A/I ⋊r G) using [29, Proposition
3.5 (ii)]. Instead we find q ∈ A fulfilling 0 6= q + I - a as follows: Since X is
totally disconnected we get that A and A/I have real rank zero and every
projection in A/I lifts to a projection in A, cf. [6, Proposition 1.1, Theorem
3.14]. We can therefore select any nonzero projections in the hereditary sub-
C∗-algebra of A/I generated by a and lift it to a projections q ∈ A. Since
0 6= q + I is contained in the hereditary sub-C∗-algebra of A/I, q + I - a,
[29, Proposition 2.7].
(cid:3)
5. Examples
Example 5.1 (The Cuntz algebra On). The Cuntz algebra, denoted by On,
n ≥ 2 is the universal C∗-algebra generated by isometries s1, s2, . . . , sn sub-
ject to the relation
n
sis∗
i = 1
Xi=1
Let Qn be the group { p
nk : p ∈ Z, k ∈ N∪{0}} of n-adic rationals. We will
denote by Qn×δZ the semidirect product: Qn×δZ = {(r, k) : r ∈ Qn, k ∈ Z}.
The two group operations are given by: (s, j)(r, k) = ( r
nj + s, j + k) and
(r, k)−1 = (−nkr, −k). Hopenwasser constructs in [26] a partial action α of
Qn on the Cantor set X such that C(X) ⋊α,r G ∼= On.
Proposition 5.2. Let X be the Cantor set and let (C(X), Qn ×δ Z, α) be the
partial dynamical system described in [26]. Then every clopen subset of X
is (Qn ×δ Z, τX)-paradoxical.
16
THIERRY GIORDANO AND ADAM SIERAKOWSKI
Proof. Let X denote the Cantor set based on [0, 1] where each n-adic rational
r (exempt 0 and 1) is replaced by a pair r−, r+, such that r− is the immediate
predecessor of r+ (i.e. there are no non-trivial elements x in X such that
r− ≤ x ≤ r+). To simplify notation we will identify 0+ with 0 and 1− with
1.
nk )+, ( q
Let U be any nonempty clopen subset of X. Recall that sets of the
form [( p
nl )−] (with p, q ∈ N∪{0}, and k, l ∈ N) form a basis for the
topology on X consisting of compact and open sets. It follows that U is a
finite disjoint union of such sets (because we can cover U by finitely many
such subsets of U and then remove the parts where they intersect). Since
a disjoint union of compact open (Qn ×δ Z, τX)-paradoxical sets is again
(Qn ×δ Z, τX)-paradoxical we can assume that U = [( p
nl )−] for some
p, q ∈ N∪{0}, and k, l ∈ N. Rewriting the fractions allows us to assume
l = k. Since U is nonempty p < q. Suppressing the index ± we have
that U is a disjoint union of sets [ p
nk ]. We can therefore
assume that q = p + 1. If U ⊆ Xs−1 for some s ∈ Qn ×δ Z it follows that
U is (Qn ×δ Z, τX)-paradoxical if, and only if, θs(U ) ⊆ Xs is (Qn ×δ Z, τX)-
paradoxical. Using the relations
nk ], . . . , [ q−1
nk )+, ( q
nk , p+1
nk , q
X
(− 1
nk ,0)
−1 = [
θ(− 1
1
nk , 1],
1
nk(cid:3)
p + 1
p
nk ,0)(cid:0)(cid:2)
nk ,
θ(0,−k)(cid:0)(cid:2)0,
nk (cid:3)(cid:1) = (cid:2)
nk(cid:3)(cid:1) = X.
1
p − 1
nk ,
p
nk(cid:3) (if k 6= 0),
X(0,−k)−1 = (cid:2)0,
we can assume that U = X. It is evident that X is (Qn×δZ, τX)-paradoxical.
In fact just like the crossed product associated to the action of the Baumslang-
Solitar group on R, cf. [30], we also have here an action that mimics both
translation and scaling. By first shrinking two copies of X and then translat-
ing one of subsets away from the other, we naturally obtain the paradoxical
property of X.
(cid:3)
Example 5.3 (The Cuntz-Krieger algebra OA). Let A = [aij] be a {0, 1}-
valued (n × n)-matrix with no zero rows. We define the algebra OA to be
the universal C∗-algebra generated by n partial isometries {si}n
i=1 satisfying
that
sjs∗
j = 1,
Xj
and Xj
aijsjs∗
j = s∗
i si
for i = 1, 2, . . . , n.
We have chosen to define OA as a universal object, as in [2], because it
allows us to identify it with a partial crossed product.
If the matrix A
satisfies Condition (I) of Cuntz and Krieger, which implies the uniqueness
of the C ∗-algebra C ∗({si}) provided that si 6= 0 for every i, we obtain the
well known Cuntz-Krieger algebra introduced in [7].
Proposition 5.4. Let X be a compact totally disconnected space and G = Fn
be the free group on n generators {g1, g2, . . . , gn}. Let (C(X), G, α) denote
the partial dynamical system described by [19] such that
C(X) ⋊α,r G ∼= OA.
PURELY INFINITE PARTIAL CROSSED PRODUCTS
17
If A is such that the action is exact and residually topologically free, then
the following statements are equivalent
(i) Every compact and open subset of X is (G, τX )-paradoxical
(ii) The C∗-algebra C(X) ⋊α,r G is purely infinite.
Proof. The implication (i) ⇒ (ii) follows form Theorem 4.4. For the con-
verse implication (ii) ⇒ (i) we need to recall the construction of the par-
tial crossed product. An infinite admissible path is an infinite sequence
µ = µ1µ2 . . . of generators of G such that A(µj, µj+1) = 1 for every j ∈ N
(where we identify A(gi, gj ) with A(i, j)). Let X be the path space of in-
finite admissible paths with the relative topology inherited as a closed and
i=1{g1, . . . , gn}. Let
g±1
t (= k) denote the length of a reduced word t = g±1
in G. An
i2
i1
action θ of G on X is called semisaturated if θts = θt ◦ θs for every t, s ∈ G
with ts = t + s (i.e. when there is no reduction in the concatenation
of the reduced words t and s). By [19]1 we have that C(X) ⋊α,r G ∼= OA,
where the partial action is the unique semisaturated partial action of G on
X such that
hence compact subspace of the infinite product space Q∞
· · · g±1
ik
Xg−1
i
= {µ ∈ X : A(gi, µ1) = 1} = domain( θgi ),
θgi = µ 7→ giµ,
i
.
where giµ means concatenation of gi at the beginning of µ. Using the iso-
morphism between OA and C(X) ⋊α,r G, we have that si = 1Xgi
δgi, where
Xgi = θgi(Xg−1
), cf. [19, Theorem 7.4] and [17, Theorem 6.5]. In particular
sis∗
i = 1Xgi
Let U be any compact and open subset of X. Recall that the cylinder
sets in X (i.e. the sets consisting of infinite admissible paths with the same
finite initial sequence) form a basis for the topology on X. In particular U
is a finite disjoint union of cylinder sets. Since a disjoint union of compact
open (G, τX )-paradoxical sets is again (G, τX )-paradoxical we can assume
that U is a cylinder set. If U ⊆ Xs−1 for some s ∈ G it follows that U is
(G, τX )-paradoxical if, and only if, θs(U ) ⊆ Xs is (G, τX )-paradoxical. We
can therefore translate U until it has the form U = Xgi for some i = 1, . . . , n.
We recall the definition of the C ∗-algebra corresponding to a directed
graph, cf. [22]. Let E = (E0, E1, r, s) be a directed graph with count-
ably many vertices E0 and edges E1, and range and source functions r, s :
E1 → E0, respectively. The C ∗-algebra C ∗(E) is the universal C ∗-algebra
generated by families of projections {pv : v ∈ E0} and partial isometries
{se : e ∈ E1}, subject to the following relations:
(i) pvpw = 0 for v, w ∈ E0, v 6= w.
esf = 0 for e, f ∈ E1, e 6= f .
(ii) s∗
ese = pr(e) for e ∈ E1.
(iii) s∗
e ≤ ps(e) for e ∈ E1.
(iv) ses∗
1The version on arxiv.org have an additional section on this topic. See also [16, p. 55].
18
THIERRY GIORDANO AND ADAM SIERAKOWSKI
(v) pv = X{e∈E1: s(e)=v}
ses∗
e for v ∈ E0 such that 0 < s−1(v) < ∞.
Let E be the graph corresponding to the matrix A, where the genera-
tors for G are the vertices and where we draw an edge from gi to gj when
A(i, j) = 1. It follows form [32, Proposition 4.1] that C ∗(E) ∼= OA. Us-
ing the isomorphism to identify elements in OA and C ∗(E) we have that
si = P{e∈E1 : s(e)=gi} se and se = ss(e)sr(e)s∗
r(e), cf. [34]. This gives us that
1U = 1Xgi
= sis∗
i = X{e∈E1 : s(e)=gi}
ses∗
e = pgi.
Since C ∗(E) is purely infinite, then 1U is properly infinite (since every
nonzero positive element in a purely infinite C∗-algebra is properly infinite),
but we need a bit more work to obtain the (G, τX )-paradoxical property of
U .
We will now argue that U = Xgi is a disjoint union of sets of the form
Xµ1µ2...µk+1, where there is a loop based at the last vertex µk+1. By a loop
we mean a sequence ν1ν2 . . . νl+1, l ≥ 1 of elements in {g1, . . . , gn} such that
A(νi, νi+1) = 1 and νl+1 = ν1). If there is a loop based at gi we are done
(k = 0). Otherwise we consider all gj in {g1, . . . , gn} such that A(i, j) = 1.
As A has no zero rows we have the union Xgi = S{j : A(i,j)=1} Xgigj is non-
empty. We now look at each gj to see if there is a loop based at gj.
If
there is a loop based at gj we keep Xgigj as it is. If not, we consider all gk ∈
{g1, . . . , gn} such that A(j, k) = 1 and rewrite Xgigj as S{k : A(j,k)=1} Xgigjgk .
We now look at each gk to see if there is a loop based at gk. If there is a
If not, we rewrite Xgigjgk as
loop based at gk we keep Xgigjgk as it is.
S{l : A(k,l)=1} Xgigjgkgl. We continue this process until U is rewritten into
the desired form. This process is finite because we pick a new element in
{g1, . . . , gn} every time we do a rewriting (if a vertex appears a second time
when doing a rewriting then there must be a loop at that vertex, hence we
never started the rewriting at that particular vertex to begin with).
Knowing that U is a disjoint union of sets of the form Xµ1µ2...µk+1, where
there is a loop based the last vertex µk+1 (and that disjoint union of para-
doxical sets is paradoxical), we can assume U = Xµ1µ2...µk+1, with a loop
based as at µk+1. With t = (µ1µ2 . . . µk)−1 we have that
k
(θµ−1
(Xµkµk+1) = θµ−1
Xµk+1 = θµ−1
= θµ−1
= θ(µk−2µk−1µk)−1(Xµk−2µk−1µkµk+1)
= θt(Xµ1...µkµk+1) = θt(U ).
k µ−1
k−1
k
k−1
(Xµk−1µkµk+1))
(Xµk−1µkµk+1) = θ(µk−1µk)−1(Xµk−1µkµk+1)
In particular we can assume that U = Xgi, gi ∈ {g1, . . . , gn} with a loop
based at gi. As C ∗(E) is purely infinite then cf. [25, Theorem 2.3], the
graph E satisfies condition (K) (i.e., no vertex v ∈ E0 lies on a loop, or
there are two loops β′, β′′ based at v such that neither β′ nor β′′ is an initial
PURELY INFINITE PARTIAL CROSSED PRODUCTS
19
subpart of the other, cf. [32]). In particular there are two distinct finite loops
β′, β′′ based at gi. This implies that Xβ′ ∪ Xβ′′ ⊆ Xgi, where the union is
disjoint, and where Xβ denotes the set of infinite admissible path starting
with the finite sequence β. If we follow the two loops β′ and β′′ we come
back to gi, implying the existence of t1, t2 ∈ G such that θt1(Xβ′) = Xgi and
θt2(Xβ′) = Xgi. We conclude that U = Xgi is (G, τX )-paradoxical.
(cid:3)
Example 5.5 (C∗-algebras of integral domains). Let R be an integral domain
with the property that the quotient R/(m) is finite, for all m 6= 0 in R. Set
R× := R\{0}. Following Boava and Exel [5] we define the regular C∗-algebra
A[R] of R as the universal C∗-algebra generated by isometries {sm : m ∈ R×}
and unitaries {un : n ∈ R} subject to the relations
smsm′ = smm′, unun′
= un+n′
, smun = umnsm, Xl+(m)∈R/(m)
ulsms∗
mu−l = 1,
for m, m′ ∈ R× and n, n′ ∈ R.
Following Boava and Exel [5], let K denote the field of fractions of R, and
K × the set K\{0}. Let G be the semidirect product K ⋊ K × = {(u, w) : u ∈
K, w ∈ K ×} equipped with the following operations (u, w)(u′, w′) = (u +
u′w, ww′) and (u, w)−1 = (−u/w, 1/w). As in [5] we define a partial order
on K × given by w ≤ w′ if w′ = wr for some r ∈ R. Let (w) denote the ideal
wR ⊆ K.
Let X be the space of all sequences (uw + (w))w∈K × in Qw∈K ×(R +
(w))/(w) fulfilling that uw′ + (w) = uw + (w) if w ≤ w′. Then in [5], Boava
and Exel prove (see Appendix A.3 for a more detailed description) that
where the partial action θ of G on X is defined by
C(X) ⋊α,r G ∼= A[R] ,
X(u,w) = {(uw′ + (w′))w′ ∈ X : uw + (w) = u + (w)},
θ(u,w) = (uw′ + (w′))w′ 7→ (u + wuw−1w′ + (w′))w′.
Note that the sets (Xt)t∈G are compact, open and form a basis for the
topology on X, cf. [5].
Proposition 5.6. If (C(X), G, α) denotes the partial dynamical system de-
scribed in [5] and recalled above, then every compact and open subset of X
is (G, τX )-paradoxical. In particular OZ ∼= A[Z] is purely infinite.
Proof. Let U be any compact and open subset of X. For w ∈ K × and
Cw ⊆ (R + (w))/(w) set
w := {(uw′ + (w′))w′ ∈ X : uw + (w) ∈ Cw}.
V Cw
Boava and Exel showed in [5] that the family of sets (V Cw
w )w∈K × is closed
under complement, intersections and finite unions. It follows that U = V Cw
for some w ∈ K × and Cw ⊆ (R+(w))/(w). Since a disjoint union of compact
open (G, τX )-paradoxical sets is again (G, τX )-paradoxical we can assume
that Cw contains precisely one element, i.e. U = Xt for some t = (u, w) ∈ G.
w
20
THIERRY GIORDANO AND ADAM SIERAKOWSKI
Since X(u,w) = ∅ ⇔ u /∈ R + (w), cf. [5, Proposition 4.4], we can assume
that u ∈ R + (w). Using that X(u,w) = X(u+w,w) we can assume u ∈ R.
Since Cw 6= ∅ it follows from an argument prior [5, Proposition 4.5] that
, with Cwr containing at last two elements for some2 r in R. In
w = V Cwr
V Cw
wr
particular
U = [s+(rw)∈Crw
X(s,rw),
Crw > 1.
Using the relations (with s ∈ R×)
X(0,s)−1 = X
X(s,1)−1 = X
θ(0,s)(X(u,w)) = X(su,sw),
θ(s,1)(X(u,w)) = X(s+u,w).
it follows that there exist a finite number of elements t1, . . . , tn in G and a
finite number of open, pairwise disjoint subsets U1, . . . , Un of U such that
U = [i∈{1,...,n}
Un,
n > 1,
θtj (U ) = Uj,
j ∈ {1, . . . , n}.
This shows that U is (G, τX )-paradoxical. Since G is solvable both the group
G and the action α is exact. It was shown in [5, Proposition 4.5] that the
action is residually topologically free. By Theorem 4.4 OZ ∼= A[Z] is purely
infinite.
(cid:3)
Corollary 5.7. Let (C(X), G, α) be the partial dynamical system as above
with R = Z and G = Q ⋊ Q×. Then every compact and open subset of X is
+. In particular ON ∼= C(X) ⋊α,r H is
(H, τX)-paradoxical, with H = Q ⋊ Q×
purely infinite.
6. Connection to crossed products
Abadie showed in [1] that certain partial crossed products are Morita-
Rieffel equivalent to ordinary crossed products. Since pure infiniteness is
preserved under stable isomorphism (see [29]) one might also expect that
the mentioned Morita-Rieffel equivalence maps paradoxical sets in the realm
of partial actions into paradoxical sets in the realm of ordinary action. This
is precisely the case.
Let us first recall the result of Abadie in [1].
Definition 6.1. Let (C0(X), G, α) be a partial dynamical system, together
with the corresponding collection of homeomorphisms (θt : Xt−1 → Xt)t∈G,
inducing α. The envelope space, denoted by X e, is the topological quotient
space (G×X)
∼ , where ∼ is the equivalence relation given by
(r, x) ∼ (s, y) ⇔ x ∈ Xr−1s and θs−1r(x) = y.
The envelope action, denoted by he, is the action induced in X e by the
action he
s(t, x) 7→ (st, x).
2In [5] it is stated that any non-invertible r in R will work, which is not true.
PURELY INFINITE PARTIAL CROSSED PRODUCTS
21
Theorem 6.2 (Abadie [1]). Let (C0(X), G, α) be a partial dynamical system
such that X e is Hausdorff. Let αe denote the action of G on C0(X e) induced
by the envelope action. Then C0(X) ⋊α,r G is Morita-Rieffel equivalent to
C0(Xe) ⋊αe,r G.
We then have:
Theorem 6.3. Let (C0(X), G, α) be a partial dynamical system such that
X e Hausdorff. If every compact open subset of X is (G, τX )-paradoxical,
then every compact open subset of X e is (G, τX e)-paradoxical.
Proof. Let U e be a compact open subset of X e. Find a cover of U e consisting
of open subsets U e of the form {t}×Ut
∼ , with Ut ∈ τX and t ∈ G. By
compactness we can assume the cover is finite. Moreover, we can assume
that
(i) The union is disjoint: Let ρ denote the canonical surjection G × X 7→
X e, and let X c
t , t ∈ G, denote the complement of Xt in X. Fix t ∈ G. For
each s ∈ G and x ∈ Xt−1s we have that (t, x) ∼ (s, θs−1t(x)). In particular
∼ ) = Ss∈G{s} × θs−1t(Xt−1s) is open in X, implying that {t}×X
ρ−1( {t}×X
is open in X e. Since X e is Hausdorff the sets (Xt)t∈G are clopen in X,
s−1t is open in X, and {t}×X
se can, by compactness,
∼ )t∈F . By passing
∼ of
t ∩ U e we
t , where the union is disjoint. For each f ∈ F define
t } ∈ τX, where ρt : X → X e denotes the (continues)
cf. [12, Proposition 3.1]. Hence Ss∈G{s} × X c
is therefore clopen in X e. Since X e = St∈G
cover U e by finitely many clopen set of the form ( {t}×X
to subsets of the sets ( {t}×X
clopen subsets of Xe fulfilling that V e
∼ )t∈F we obtain a partition (V e
∼ . With U e
t
have that U e = St∈F U e
Ut := {x ∈ X : ρt(x) ∈ U e
composition of x 7→ (t, x) and ρ. It follows easily that U e
t )t∈F of F ×X
t ⊆ {t}×X
:= V e
{t}×X
∼
∼
∼
t = {t}×Ut
∼ .
that Ut ⊆ Sm
of open subsets of X. Notice that each V e
i
i ∩ V e
t consisting of open subsets of X e and, together with (V e
(ii) Each set Ut is compact: Fix t ∈ F . Let (Vi)i be a cover of Ut consisting
is open in X e (since
t )i is a cover
of U e
s )s∈F \{t}, a
cover of U e. By compactness we can assume i = 1, . . . , m for some m ∈ N.
Intersecting with V e
i . It follows
Ss∈G{s} × θs−1t(Vi ∩ Xt−1s) is open in X). The sets (V e
t ⊆ Sm
i=1 Vi. Hence Ut is compact.
t we have that U e
t ⊆ Sm
:= {t}×Vi
i ∩ V e
i=1 V e
i=1 V e
∼
∼
∩ {e}×Vl
(iii) U e is paradoxical: Each set of the form {e}×U
∼ , with U compact and
open in X, is (G, τX e)-paradoxical: If (Vi, ti)n+m
i=1 witness the paradoxicality
of U and k 6= l are any natural numbers in 1, . . . , n + m if follows that
{e}×Vk
∼ , ti)n+m
i=1
witness the paradoxicality of {e}×U
∼ . A translate of a (G, τX e)-paradoxical set
is again (G, τX e )-paradoxical. Hence each {t}×Ut
is (G, τX e )-paradoxical. A
finite disjoint union of (G, τX e )-paradoxical set is again (G, τX e )-paradoxical.
We conclude that U e is (G, τX e )-paradoxical.
(cid:3)
∼ = ∅ because Vk ∩ Vl = ∅. We obtain that ( {e}×Vi
∼
22
THIERRY GIORDANO AND ADAM SIERAKOWSKI
Remark 6.4. In a very interesting recent preprint [28], Kellerhals, Monod,
and Rørdam proved that a countable group is non-supramenable if and only
if it admits a free, minimal, purely infinite action on the locally compact
non-compact Cantor set. Then they use this characterisation to associate
to such dynamical systems stable Kirchberg C*-algebras in the UCT class.
Based on the examples presented in Section 5, where the opens sets
(Xt)t∈G we used to construct the partial crossed products were clopen (en-
suring X e is Hausdorff, see [12, Proposition 3.1]), and Proposition 6.3, we
obtain another class of examples of dynamical systems whose associated
crossed products are stable Kirchberg C*-algebras.
Appendix A.
A.1. Basic definitions. Recall that a Fell bundle over a discrete group G
is a collection B = (Bt)t∈G of closed subspaces of a C∗-algebra B, indexed
by a discrete group G, satisfying B∗
t = Bt and BtBs ⊆ Bts for all t and s in
G (cf. [16]). A section of B is a function ξ : G → St∈G Bt with the property
that ξ(t) ∈ Bt for all t ∈ Bt. Let l1(B) denote the Banach ∗-algebra (cf. [35,
Proposition 2.1]) consisting of all sections ξ of B with a finite l1-norm. The
operations and the norm in l1(B) are given by
ξη(t) = Xs∈G
ξ(s)η(s−1t),
ξ∗(t) = ξ(t−1)∗,
kξk1 = Xt∈G
kξ(t)k.
We define the full cross sectional algebra of B, denoted C∗(B), to be the
enveloping C∗-algebra of l1(B) (cf. [21, Section VIII.17.2]). Recall that a
right Hilbert A-module X is a right A-module X equipped with a map h·, ·iA :
X × X → A that is linear in the second component and for x, y ∈ X, a ∈ A,
(i) hx, xiA ≥ 0 with equality only if x = 0;
(ii) hx, y · aiA = hx, yiAa;
(iii) hx, yiA = hy, xi∗
A; and
(iv) X is complete in the norm defined by kxk2
As in [27, Section 1.1.7], let l2(B) denote the right Hilbert Be-module con-
A = khx, xiAk.
converges unconditionally (i.e. for any ε > 0 there exist a finite set F ⊆
G such that for every finite subsets H ⊆ G containing F one has that
sisting of all cross sections ξ of B fulfilling that the series x = Pt∈G ξ(t)∗ξ(t)
kx −Pt∈H ξ(t)∗ξ(t)k < ε). We equip l2(B) with the inner product
hξ, ηi = Xt∈G
ξ(t)∗η(t),
ξ, η ∈ l2(B).
Let L(l2(B)) denote the C∗-algebra of all adjointable operators on l2(B)
(i.e. linear maps T : l2(B) → l2(B) with a linear map T ∗ such that hξ, T ηi =
PURELY INFINITE PARTIAL CROSSED PRODUCTS
23
hT ∗ξ, ηi, for all ξ, η ∈ l2(B)). With the ∗-homomorphism ΛB : l1(B) →
L(l2(B)) defined (in [20, Proposition 2.6]) by
(ΛB(ξ)η)(t) = Xs∈G
ξ(s)η(s−1t),
ξ ∈ l1(B), η ∈ l2(B), t ∈ G,
we define the reduced cross sectional algebra of B, denoted C∗
r(B), to be the
sub-C∗-algebra of L(l2(B)) generated by the range of ΛB. Let δs,t, s, t ∈ G
denote the Kronecker symbol. Each Bt is a right Hilbert Be-module with
the inner product hb, ci = b∗c. Let jt ∈ L(Bt, l2(B)) denote the adjointable
operator defined by (jt(bt))(s) = δs,tbt, for bt ∈ Bt and s ∈ G. The adjoint
j∗
t ∈ L(l2(B), Bt) is simply the map j∗
t (ξ) = ξ(t), ξ ∈ l2(B). Recall, cf. [16,
r(B) and t ∈ G the tth Fourier coefficient of x is
Definition 2.7], that for x ∈ C∗
the unique element x(t) ∈ Bt such that (j∗
t ◦ x ◦ je)(a) = x(t)a for all a ∈ Be.
The map E : C∗
r(B) → Be, given by x 7→ x(e) is a positive, contractive,
conditional expectation. Moreover, E is faithful on positive elements since
r(B)
(cf. [16, Proposition 2.12]). Let Cc(B) ⊆ l1(B) denote the set of finitely
supported sections of B.
E(x∗x) = Pt∈G x(t)∗ x(t) (unconditional convergence) for each x ∈ C∗
Let A be a C∗-algebra and G be a discrete group. Recall (see [9]) that a
partial action of G on A, denoted by α, is a collection (Dt)t∈G of closed two-
sided ideals of A and a collection (αt)t∈G of ∗-isomorphisms αt : Dt−1 → Dt
such that
s (Ds ∩ Dt−1) ⊆ D(ts)−1 ;
(i) De = A, where e represents the identity element of G;
(ii) α−1
(iii) αt ◦ αs(x) = αts(x), ∀ x ∈ α−1
The triple (A, G, α) is called a partial dynamical system. Fix a partial dy-
namical system (A, G, α). Recall the corresponding Fell bundle B as defined
in Definition 2.2: We let L be the normed ∗-algebra of the finite formal sums
s (Ds ∩ Dt−1).
The family (Bt)t∈G generates the Fell bundle B. It follows that Cc(B) = L.
Pt∈G atδt, where at ∈ Dt. We let Bt denote the vector subspace Dtδt of L.
Moreover, for x = Pt∈G atδt ∈ Cc(B), we have that E(ΛB(x)) = ae.
Lemma A.1 (McClanahan). Let (A, G, α) be a partial dynamical system.
For any G-invariant ideal I of A we have a canonical embedding of I ⋊α,r G
in A ⋊α,r G.
Remark A.2. To best of our knowledge our proof of Lemma A.1 is new.
For a different proof using covariant representation we refer to the work in
[35]. Since our proof applies to general Fell bundles (and not only the one
defining crossed products) we have included it for completeness.
Proof. Fix an approximate unit (en) for I, and the Fell bundle
E = (Et)t∈G, Et = (Dt ∩ I)δt,
equipped with the operations and norm coming from B. For each section
ξ ∈ l2(B) let ξn : G → St∈G Et denote the map t 7→ ξ(t)en contained in
24
THIERRY GIORDANO AND ADAM SIERAKOWSKI
l2(E). Let ϕ : L(l2(E)) → L(l2(B)) denote the map ϕ(T )ξ = limn T (ξn).
For x ∈ Cc(E), ϕ(ΛE(x)) = ΛB(x). We have that ϕ(ΛE(x)) = ΛB(x) and
computation
kϕ(ΛE(x))k = kΛE(x)k, for any element x = Pt∈G atδt in Cc(E), by direct
kϕ(ΛE(x))k2 =
kϕ(ΛE(x))ξk2
sup
ξ∈Cc(B),kξk≤1
=
=
=
sup
ξ∈Cc(B),kξk=1
lim
n
k Xt,s,r∈G(cid:0)(atδt)ξn(t−1r)(cid:1)∗((asδs)ξn(s−1r))k
sup
η∈Cc(E),kηk≤1
sup
η∈Cc(E),kηk≤1
k Xt,s,r∈G(cid:0)(atδt)η(t−1r)(cid:1)∗((asδs)η(s−1r))k
kΛE(x)ηk2
= kΛE(x)k2
The first and last equality above use the fact that the finitely supported
sections are dense in l2. The second and fourth equality follows from the
definition of the maps ϕ and ΛE. For the third equality we obviously
have ≥ after taking the limit. To get the ≤ notice that if we remove
"supξ∈Cc(B),kξk≤1 limn" from the left hand side we have ≤ since ξn ∈ {η ∈
Cc(E), kηk ≤ 1} for any n and any ξ ∈ Cc(B) with kξk ≤ 1. The inequality ≤
remains valid when we take the limit and the supremum. I particular we ob-
tain that the canonical embedding of Cc(E) into Cc(B), given by atδt 7→ atδt,
extends to an embedding ι : I ⋊α,r G → A ⋊α,r G.
(cid:3)
Let I be a G-invariant ideal of A. In order to introduce the map ρ : A ⋊α,r G →
A/I ⋊α,r G we can again turn to [16]. Let F, denote the Fell bundle
(Ft)t∈G, Ft = (Dt/(I ∩ Dt))δt,
equipped with the operations and norm coming from B. The canonical
surjection of Cc(B) onto Cc(F), given by atδt 7→ (at + Dt ∩ I)δt, extends to a
surjective ∗-homomorphism ρ : A ⋊α,r G → A/I ⋊α,r G (cf. [16, Proposition
3.11]).
In particular, by continuity of ι, ρ, and E, we have the following
result:
Proposition A.3 (Exel, McClanahan). Let (A, G, α) be a partial dynamical
system. For any G-invariant ideal I of A we have the commuting diagram
0
0
/ I ⋊α,r G ι
A ⋊α,r G
ρ
A/I ⋊α,r G
EI
/ I
EA
/ A
EA/I
/ A/I
0
/ 0
Using that EA is idempotent and identifying I ⋊α,r G with a subset of
A ⋊α,r G we have the additional properties:
/
/
/
/
/
/
/
/
/
/
/
PURELY INFINITE PARTIAL CROSSED PRODUCTS
25
Proposition A.4 (Exel, McClanahan). Let (A, G, α) be a partial dynamical
system. For any G-invariant ideal I of A and any ideal J of in A ⋊α,r G
we have the following identities:
(I ⋊α,r G) ∩ A = I, J ∩ A ⊆ EA(J ),
Ideal[I] = I ⋊r G,
where Ideal[S] denotes the smallest ideal in A ⋊α,r G generated by S ⊆
A ⋊α,r G.
A.2. Topological freeness. Let (A, G, α) be a partial dynamical system
with ideals (Dt)t∈G and ∗-isomorphisms (αt)t∈G. If we replace the C∗-algebra
A by a locally compact Hausdorff space X, the ideals Dt by open sets Xt
and the ∗-isomorphisms αt by homeomorphisms θt : Xt−1 → Xt, we obtain
a partial action θ of G on the space X. A partial action of a group G on a
space X induces naturally a partial action α of G on C0(X). The ideals are
C0(Xt) and αt(f ) = f ◦ θt−1. The converse (still in the abelian case) is also
true (cf. [1, 23]).
Let Prim A denote the primitive ideal space of A with respect to the
Jacobson topology. Let A denote the spectrum of A with respect to the
topology induced by the surjection κ : A → Prim A, κ([π]) = ker π. Fol-
lowing [33] recall how α defines a partial action of G on Prim A and on
A:
For any ideal J of A we let supp J denote the subset {x ∈ Prim A : J *
x}. It is known (see [8, Section 3.2.1] or [40, Section 1.4]) that the mapping
x 7→ x ∩ J establishes a homeomorphism supp J ↔ Prim J and supp J is
an open set in Prim A. Set also AJ = {[π] ∈ A : π(J ) 6= 0}. Then the
mapping [π] 7→ [πJ ] establishes a homeomorphism AJ ↔ J and AJ is an
open set in A (see [8, Section 3.2.1] or [40, Section 1.4]). Let us define the
mapping τt : ADt−1 → ADt in the following way: For any [π] ∈ ADt−1 we set
τt([π]) = [π ◦ αt−1],
t ∈ G.
The foregoing observations tell us that τt is a homeomorphism. Let us also
define the mapping θt : supp Dt−1 → supp Dt in the following way: For any
point x ∈ supp Dt−1 such that x = ker π where [π] ∈ ADt−1 we set
θt(x) = ker (π ◦ αt−1),
t ∈ G.
Clearly θt is a homeomorphism. For τt and θt defined in the above described
way we have that (τt)t∈G defines a partial action of G on A and (θt)t∈G
defines a partial action of G of Prim A. Recall that the action α is called
topologically free if for any finite set {t1, . . . , tn}, ti 6= e the set
n
[i=1(cid:8)x ∈ supp Dt−1
i
: θti(x) = x(cid:9)
has empty interior. This condition can be also formulated in the follow-
ing way: For any finite set {t1, ...tk} ⊆ G and any nonempty open set U
there exists a point x ∈ U such that all the points θti(x), i = 1, . . . , k that
26
THIERRY GIORDANO AND ADAM SIERAKOWSKI
are defined (⇔ x ∈ supp Dt−1
) are distinct. Recall that the action α has
the intersection property if every nonzero ideal in A ⋊α,r G has a nonzero
intersection with A.
i
Theorem A.5 (Lebedev). Let (A, G, α) be a partial dynamical system.
Suppose that the action is topologically free. Then α has the intersection
property.
Theorem A.6 (Exel, Laca, Quigg). Let (A, G, α) be a partial dynamical
system with A abelian. Suppose that the action is topologically free. Then
for every b ∈ A ⋊α,r G and every ε > 0 there exist a positive contraction
x ∈ A satisfying
kxEA(b)x − xbxk < ε,
kxEA(b)xk > kEA(b)k − ε.
Proposition A.7. Let (A, G, α) be a partial dynamical system with A abelian.
Then the following properties are equivalent:
(i) α is topologically free,
(i′) Ft = {x ∈ Ut−1 : θt(x) = x} has empty interior (t 6= e),
(ii) kαtK − idKk = 2 for every αt-invariant ideal K in Dt−1 (and t 6= e),
(ii′) kαtK − idKk 6= 0 for every αt-invariant ideal K in Dt−1 (and t 6= e),
(iii) inf{kxαt(x)k : x ∈ K+, kxk = 1} = 0 for every nonzero ideal K in Dt−1
(and t 6= e),
(iv) inf{kx(btδt)xk : x ∈ B+, kxk = 1} = 0 for every bt ∈ Dt, and every
nonzero hereditary C∗-algebra B in A (and t 6= e),
(iv′) inf{kx(btδt)xk : x ∈ K+, kxk = 1} = 0 for every bt ∈ Dt, and every
nonzero ideal K in A (t 6= e).
Proof. The implications (i) ⇔ (i′) follows from [19, Definition 2.1, Lemma
2.2], and the implications (iv) ⇔ (iv′) and (ii) ⇒ (ii′) are evident.
Pick any x ∈ C, where C := {y ∈ V : αt(f )(y) − f (y) > 0}.
(ii′) ⇒ (ii): Find some function f ∈ K+ = C0(V )+ such that αt(f ) 6= f .
Notice x 6= θt−1(x) in V . (cid:0)If f (x) 6= 0 then -- for h := αt−1(f ) ∈ K --
we have that h(θt−1 (x)) = f (x) 6= 0, recalling that αt(f )(x) = f (θt−1(x)).
Hence θt−1(x) ∈ V . If f (x) = 0 then αt(f )(x) = f (θt−1(x)) 6= 0. Hence
θt−1(x) ∈ V .
If x = θt−1(x) then 0 = f (θt−1(x)) − f (x) = αt(f )(x) −
h(x) = 2. Hence kαtK − idKk = 2.
f (x) > 0.(cid:1) One can now easily find a function h ∈ K such that h(θt−1 (x))−
(i′) ⇒ (iii): Assume not (iii). Find some t 6= e and K = C0(V ) in
C0(Ut−1) such that inf{kxβ(x)k : x ∈ K+, kxk = 1} > 0, where β = αt.
Suppose (i′). Then the Ut−1-open set F c
t = {x ∈ Ut−1 : θt(x) 6= x} is
dense in Ut−1. As V is Ut−1-open it has a non-empty intersection with F c
t .
Therefore we can find a non-empty Ut−1-open set Y in V such that θt(Y ) is
disjoint from Y . Hence there is a function x in K+ of norm one such that
xβ(x) = 0. Contradiction, hence not (i′).
(iii) ⇒ (ii′): Assume not (ii′). Then there exists a nonzero αt-invariant
ideal K = C0(V ) in C0(Ut−1 ) (for some t 6= e) such that kαtK − idKk = 0.
PURELY INFINITE PARTIAL CROSSED PRODUCTS
27
We conclude that αt(x) = x for all x ∈ K. Hence inf{kxαt(x)k : x ∈
K+, kxk = 1} is nonzero implying not (iii).
(ii′) ⇒ (i′): Assume not (i′). Find t 6= e and a Ut−1-open nonempty
subset V in Ft (recall that Ft is Ut−1-closed but has nonempty interior).
Since θt(Ut−1 ) = Ut we see that each x ∈ V ⊆ Ut−1 also belongs to Ut and
θt−1(x) = x, x ∈ V . Hence K := C0(V ) is αt-invariant (i.e. αt(f ) = f =
αt−1(f ), f ∈ K). We obtain that kαtK − idKk = 0. Hence not (ii′).
(iii) ⇒ (iv′): Fix any t 6= e, any bt ∈ Dt, and any nonzero ideal K in
A. Define I := K ∩ Dt−1. Suppose I = 0. Fix any x ∈ K. Using that
xbt ∈ Dt we get αt−1(xbt) ∈ Dt−1 and αt−1(xbt)x = 0. Hence x(btδt)x =
αt(αt−1 (xbt)x)δt = 0 and
inf{kxbtδt)xk : x ∈ K+, kxk = 1} = 0.
Suppose I 6= 0. Then I is an ideal in Dt−1 and
kx(btδt)xk ≤ kαt(αt−1(xbt))αt(x)k ≤ kbtkkxαt(x)k,
x ∈ I +.
It follows from (iii) that
inf{kx(btδt)xk : x ∈ K+, kxk = 1} = 0.
(iv′) ⇒ (iii): Fix any t 6= e and nonzero ideal K in Dt−1 . Define I :=
Dt ∩ K. Suppose I = 0. Fix any x ∈ K. Then x ∈ Dt−1, αt(x) ∈ Dt,
αt(x)x ∈ Dt ∩ K, and αt(x)x = 0. Hence
inf{kxαt(x)k : x ∈ K+, kxk = 1k} = 0.
Suppose I 6= 0. Then I := C0(V ) is a nonzero ideal in A. Pick any x ∈ V .
Choose an open set U with compact closure such that
x ∈ U ⊆ ¯U ⊆ V.
With K := ¯U ⊆ V there exist by Urysohn's Lemma a function h ∈ C0(V )
such that
0 ≤ h ≤ 1,
hK = 1.
Using (iv) on bt := h ∈ Dt and the nonzero ideal J := C0(U ) in A (and the
equality x(btδt)x = xbtαt(x)δt valid for every x ∈ J ⊆ I ⊆ K ⊆ Dt−1) we
obtain
inf{kxbtαt(x)k : x ∈ J +, kxk = 1k} = 0.
Since hU = 1 we have that xbt = x for every x ∈ J+, hence
inf{kxαt(x)k : x ∈ J +, kxk = 1k} = 0.
Since J ⊆ I ⊆ K also
inf{kxαt(x)k : x ∈ K+, kxk = 1k} = 0.
(cid:3)
28
THIERRY GIORDANO AND ADAM SIERAKOWSKI
A.3. C∗-algebras of an Integral Domain. Let R be an integral domain
(i.e. a commutative unital ring without zero divisor). Set R× := R \ {0}.
Following Boava and Exel [5] we impose that the quotient R/(m) is finite,
for all m 6= 0 in R and let A[R] denote the regular C∗-algebra of R, i.e., the
universal C∗-algebra generated by isometries {sm : m ∈ R×} and unitaries
{un : n ∈ R} subject to the relations
smsm′ = smm′, unun′
= un+n′
, smun = umnsm, Xl+(m)∈R/(m)
ulsms∗
mu−l = 1,
for every m, m′ ∈ R× and n, n′ ∈ R. As in [5] we let K denote the field
of fractions of R and let K × denote the set K\{0}. To form a group one
equip the semidirect product K ⋊ K × = {(u, w) : u ∈ K, w ∈ K ×} with the
following two group operations
(u, w)(u′, w′) = (u + u′w, ww′),
(u, w)−1 = (−u/w, 1/w).
Boava and Exel showed in [5] that the regular C∗-algebra of R is a partial
crossed product by first showing that the algebra A[R] is ∗-isomorphic to a
partial group algebra of K ⋊ K × (with certain relations R), and then use
that every partial group algebra is a partial crossed product. We will not
describe these isomorphisms but will instead focus on the the description of
the partial crossed product.
We have a partial order on K × given by w ≤ w′ if w′ = wr for some
r ∈ R. For w ∈ K let (w) denote the ideal wR ⊆ K. For each pair w ≤ w′
in K × we have a canonical map pw,w′ : (R + (w′))/(w′) → (R + (w))/(w)
given by
pw,w′(uw′ + (w′)) = uw′ + (w).
Boava and Exel proved that the inverse limit lim←−{(R + (w))/(w), pw,w′}
is isomorphic to the space X of all sequences
(uw + (w))w∈K × ∈ Yw∈K ×
(R + (w))/(w)
fulfilling that uw′ + (w) = uw + (w) if w ≤ w′. When (R + (w))/(w) is given
the discrete topology and (R + (w))/(w) the product topology X becomes
a compact topological space. Moreover, cf. [5], there is a partial action θ of
G on X defined by
X(u,w) = {(uw′ + (w′))w′ ∈ X : uw + (w) = u + (w)},
θ(u,w) = (uw′ + (w′))w′ 7→ (u + wuw−1w′ + (w′))w′,
for any (u, w) ∈ K ⋊ K ×. The partial actions θ induces a partial dynam-
ical system (C(X), K ⋊ K ×, α), and hence also a partial crossed product
C(X) ⋊α,r (K ⋊ K ×).
Theorem A.8 (Boava-Exel). Let R be a integral domain with finite quo-
tients R/(m), m ∈ R×. Suppose that R is not a filed. Then the maps
un 7→ 1X δ(n,1),
sm 7→ 1X(0,m) δ(0,m)
PURELY INFINITE PARTIAL CROSSED PRODUCTS
29
induce a ∗-isomorphism between A[R] and C(X) ⋊α,r (K ⋊ K ×).
References
1. Fernando Abadie, Enveloping actions and Takai duality for partial actions, J. Funct.
Anal. 197 (2003), no. 1, 14 -- 67. MR 1957674 (2004c:46130)
2. Astrid an Huef and Iain Raeburn, The ideal structure of Cuntz-Krieger algebras, Er-
godic Theory Dynam. Systems 17 (1997), no. 3, 611 -- 624. MR 1452183 (98k:46098)
3. R. J. Archbold and J. S. Spielberg, Topologically free actions and ideals in discrete
C ∗-dynamical systems, Proc. Edinburgh Math. Soc. (2) 37 (1994), no. 1, 119 -- 124.
MR 1258035 (94m:46101)
4. B. Blackadar, Operator algebras, Encyclopaedia of Mathematical Sciences, vol. 122,
Springer-Verlag, Berlin, 2006, Theory of C ∗-algebras and von Neumann algebras,
Operator Algebras and Non-commutative Geometry, III. MR 2188261 (2006k:46082)
5. Giuliano Boava and Ruy Exel, Partial crossed product description of the C ∗-algebras
associated with integral domains, 2012.
6. Lawrence G. Brown and Gert K. Pedersen, C ∗-algebras of real rank zero, J. Funct.
Anal. 99 (1991), no. 1, 131 -- 149. MR MR1120918 (92m:46086)
7. Joachim Cuntz and Wolfgang Krieger, A class of C ∗-algebras and topological Markov
chains, Invent. Math. 56 (1980), no. 3, 251 -- 268. MR 561974 (82f:46073a)
8. Jacques Dixmier, Les C ∗-alg`ebres et leurs repr´esentations, Deuxi`eme ´edition. Cahiers
Scientifiques, Fasc. XXIX, Gauthier-Villars ´Editeur, Paris, 1969. MR 0246136 (39
#7442)
9. M. Dokuchaev and R. Exel, Associativity of crossed products by partial actions, en-
veloping actions and partial representations, Trans. Amer. Math. Soc. 357 (2005),
no. 5, 1931 -- 1952. MR 2115083 (2005i:16066)
10. Siegfried Echterhoff and Marcelo Laca, The primitive ideal space of the C ∗-algebra of
the affine semigroup of algebraic integers, Math. Proc. Cambridge Philos. Soc. 154
(2013), 119 -- 126.
11. George A. Elliott, Some simple C ∗-algebras constructed as crossed products with dis-
crete outer automorphism groups, Publ. Res. Inst. Math. Sci. 16 (1980), no. 1, 299 -- 311.
MR 574038 (81j:46102)
12. R. Exel, T. Giordano, and D. Gon¸calves, Enveloping algebras of partial actions as
groupoid C ∗-algebras, J. Operator Theory 65 (2011), no. 1, 197 -- 210. MR 2765764
(2012f:46096)
13. Ruy Exel, The Bunce-Deddens algebras as crossed products by partial automorphisms,
Bol. Soc. Brasil. Mat. (N.S.) 25 (1994), no. 2, 173 -- 179. MR 1306559 (95m:46091)
14. Ruy Exel, Circle actions on C ∗-algebras, partial automorphisms, and a generalized
Pimsner-Voiculescu exact sequence, J. Funct. Anal. 122 (1994), no. 2, 361 -- 401.
MR 1276163 (95g:46122)
15.
16.
17.
18.
, Approximately finite C ∗-algebras and partial automorphisms, Math. Scand.
77 (1995), no. 2, 281 -- 288. MR 1379271 (97e:46085)
, Amenability for Fell bundles, J. Reine Angew. Math. 492 (1997), 41 -- 73.
MR 1488064 (99a:46131)
, Partial actions of groups and actions of inverse semigroups, Proc. Amer.
Math. Soc. 126 (1998), no. 12, 3481 -- 3494. MR 1469405 (99b:46102)
, Exact groups and Fell bundles, Math. Ann. 323 (2002), no. 2, 259 -- 266.
MR 1913042 (2003e:46095)
19. Ruy Exel, Marcelo Laca, and John Quigg, Partial dynamical systems and C ∗-algebras
generated by partial isometries, J. Operator Theory 47 (2002), no. 1, 169 -- 186.
MR 1905819 (2003f:46108)
20. Ruy Exel and Chi-Keung Ng, Approximation property of C ∗-algebraic bundles, Math.
Proc. Cambridge Philos. Soc. 132 (2002), no. 3, 509 -- 522. MR 1891686 (2002k:46189)
30
THIERRY GIORDANO AND ADAM SIERAKOWSKI
21. J. M. G. Fell and R. S. Doran, Representations of ∗-algebras, locally compact groups,
and Banach ∗-algebraic bundles. Vol. 2, Pure and Applied Mathematics, vol. 126,
Academic Press Inc., Boston, MA, 1988, Banach ∗-algebraic bundles, induced repre-
sentations, and the generalized Mackey analysis. MR 936629 (90c:46002)
22. Neal J. Fowler, Marcelo Laca, and Iain Raeburn, The C ∗-algebras of infinite graphs,
Proc. Amer. Math. Soc. 128 (2000), no. 8, 2319 -- 2327. MR 1670363 (2000k:46079)
23. Daniel Gon¸calves, Produtos cruzados, 2001, Tese de mestrado, Departamento de
Matematica, UFSC.
24. Philip Green, The local structure of twisted covariance algebras, Acta Math. 140
(1978), no. 3-4, 191 -- 250. MR 0493349 (58 #12376)
25. Jeong Hee Hong and Wojciech Szyma´nski, Purely infinite Cuntz-Krieger algebras of
directed graphs, Bull. London Math. Soc. 35 (2003), no. 5, 689 -- 696. MR 1989499
(2005c:46097)
26. Alan Hopenwasser, Partial crossed product presentations for On and Mk(On) us-
ing amenable groups, Houston J. Math. 33 (2007), no. 3, 861 -- 876. MR 2335740
(2008k:46161)
27. Kjeld Knudsen Jensen and Klaus Thomsen, Elements of KK-theory, Mathematics:
Theory & Applications, Birkhauser Boston Inc., Boston, MA, 1991. MR 1124848
(94b:19008)
28. Julian Kellerhals, Nicolas Monod, and Mikael Rørdam, Non-supramenable groups act-
ing on locally compact spaces, preprint (2013)
29. Eberhard Kirchberg and Mikael Rørdam, Non-simple purely infinite C ∗-algebras,
Amer. J. Math. 122 (2000), no. 3, 637 -- 666. MR 1759891 (2001k:46088)
30. Eberhard Kirchberg and Adam Sierakowski, Strong pure infiniteness of crossed prod-
ucts.
31. Akitaka Kishimoto, Outer automorphisms and reduced crossed products of simple C ∗-
algebras, Comm. Math. Phys. 81 (1981), no. 3, 429 -- 435. MR 634163 (83c:46061)
32. Alex Kumjian, David Pask, Iain Raeburn, and Jean Renault, Graphs, groupoids, and
Cuntz-Krieger algebras, J. Funct. Anal. 144 (1997), no. 2, 505 -- 541. MR 1432596
(98g:46083)
33. A. V. Lebedev, Topologically free partial actions and faithful representations of crossed
products, Funktsional. Anal. i Prilozhen. 39 (2005), no. 3, 54 -- 63, 96. MR 2174606
(2006f:46068)
34. M. H. Mann, Iain Raeburn, and C. E. Sutherland, Representations of finite groups
and Cuntz-Krieger algebras, Bull. Austral. Math. Soc. 46 (1992), no. 2, 225 -- 243.
MR 1183780 (93k:46046)
35. Kevin McClanahan, K-theory for partial crossed products by discrete groups, J. Funct.
Anal. 130 (1995), no. 1, 77 -- 117. MR 1331978 (96i:46083)
36. Dorte Olesen and Gert K. Pedersen, Applications of the Connes spectrum to C ∗-
dynamical systems. III, J. Funct. Anal. 45 (1982), no. 3, 357 -- 390. MR 650187
(83i:46080)
37. John Quigg and Iain Raeburn, Characterisations of crossed products by partial actions,
J. Operator Theory 37 (1997), no. 2, 311 -- 340. MR 1452280 (99a:46121)
38. Mikael Rørdam, On the structure of simple C ∗-algebras tensored with a UHF-algebra.
II, J. Funct. Anal. 107 (1992), no. 2, 255 -- 269. MR MR1172023 (93f:46094)
39. Mikael Rørdam and Adam Sierakowski, Purely infinite C ∗-algebras arising from
crossed products, Ergodic Theory Dynam. Systems 32 (2012), no. 1, 273 -- 293.
MR 2873171 (2012m:46063)
40. Adam Sierakowski, Discrete crossed product C ∗-algebras, 2009, PhD thesis published
by PhD Series, ISBN 978-87-91927-40-9, University of Copenhagen.
41.
, The ideal structure of reduced crossed products, Munster J. Math. 3 (2010),
237 -- 261. MR 2775364 (2012g:46103)
PURELY INFINITE PARTIAL CROSSED PRODUCTS
31
Department of Mathematics and Statistics, University of Ottawa, 585 King
Edward Ave, K1N6P1 Ottawa, Canada
E-mail address: [email protected]
The School of Mathematics & Applied Statistics, University of Wollon-
gong, Northfields Ave, 2522 NSW, Australia
E-mail address: [email protected]
|
1804.03455 | 1 | 1804 | 2018-04-10T11:16:19 | Monic representations of finite higher-rank graphs | [
"math.OA",
"math.DS",
"math.FA"
] | In this paper we define the notion of monic representation for the $C^*$-algebras of finite higher-rank graphs with no sources, and undertake a comprehensive study of them. Monic representations are the representations that, when restricted to the commutative $C^*$-algebra of the continuous functions on the infinite path space, admit a cyclic vector. We link monic representations to the $\Lambda$-semibranching representations previously studied by Farsi, Gillaspy, Kang, and Packer, and also provide a universal representation model for nonnegative monic representations. | math.OA | math |
Monic representations of finite higher-rank graphs
Carla Farsi, Elizabeth Gillaspy, Palle Jorgensen, Sooran Kang, and Judith Packer
April 11, 2018
Abstract
In this paper we define the notion of monic representation for the C∗-algebras
of finite higher-rank graphs with no sources, and undertake a comprehensive study
of them. Monic representations are the representations that, when restricted to the
commutative C∗-algebra of the continuous functions on the infinite path space, admit
a cyclic vector. We link monic representations to the Λ-semibranching representations
previously studied by Farsi, Gillaspy, Kang, and Packer, and also provide a universal
representation model for nonnegative monic representations.
2010 Mathematics Subject Classification: 46L05, 46L55, 46K10
Keywords and phrases: C∗-algebras, monic representations, higher-rank graphs, k-graphs,
Λ-semibranching function systems, coding map, Markov measures, projective systems.
Contents
1 Introduction
2 Foundational material
2.1 Higher-rank graphs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2 Λ-semibranching function systems and their representations
. . . . . . . . .
3 Representations of higher-rank graph C∗-algebras: first analysis
3.1 Λ-projective systems and representations . . . . . . . . . . . . . . . . . . . .
3.2 Projection valued measures
. . . . . . . . . . . . . . . . . . . . . . . . . . .
3.3 Disjoint and irreducible representations . . . . . . . . . . . . . . . . . . . . .
4 Monic representations of finite k-graph algebras
4.1 Λ-semibranching function systems and monic representations . . . . . . . . .
5 A universal representation for non-negative Λ-projective systems
2
4
4
6
8
8
11
13
14
19
23
1
1
Introduction
Higher-rank graphs Λ – also known as k-graphs – and their C∗-algebras C∗(Λ) were in-
troduced by Kumjian and Pask in [37], building on the work of Robertson and Steger
[45, 46]. Generalizations of the Cuntz–Krieger C∗-algebras associated to directed graphs
(cf. [11, 12, 23, 38]), k-graph C∗-algebras share many of the important properties of Cuntz
and Cuntz–Krieger C∗-algebras, including Cuntz–Krieger uniqueness theorems and realiza-
tions as groupoid C∗-algebras. Moreover, the C∗-algebras of higher-rank graphs are closely
linked with orbit equivalence for shift spaces [10] and with symbolic dynamics more generally
[43, 47, 44], as well as with fractals and self-similar structures [25, 26]. More links between
higher-rank graphs and symbolic dynamics can be seen via [3, 4] and the references cited
therein.
The research presented in the pages that follow develops a non-commutative harmonic
analysis for finite higher-rank graphs with no sources. More precisely, we introduce monic
representations for the C∗-algebras associated to finite higher-rank graphs with no sources;
undertake a detailed theoretical analysis of such representations; and present a variety of
examples.
Like the Cuntz–Krieger algebras, k-graph C∗-algebras often fall in a class of non-type I,
and in fact purely infinite C∗-algebras. The significance of this for representation theory is
that the unitary equivalence classes of irreducible representations of k-graph C∗-algebras do
not arise as Borel cross sections [29, 30, 15, 21, 22]. In short, for these C∗-algebras, only
subfamilies of irreducible representations admit "reasonable" parametrizations.
Various specific subclasses of representations of Cuntz and Cuntz–Krieger C∗-algebras
have been extensively studied by many researchers, who were motivated by their applicability
to a wide variety of fields. In addition to connections with wavelets (cf. [17, 18, 40, 27, 28,
26]), representations of Cuntz–Krieger algebras have been linked to fractals and Cantor
sets [48, 35, 25, 26] and to the endomorphism group of a Hilbert space [8, 39]. Indeed, the
astonishing goal of identifying both discrete and continuous series of representations of Cuntz
(and to some extent Cuntz–Krieger) C∗-algebras, was accomplished in [19, 20, 5], building
on the pioneering results of [7].
In the setting of higher-rank graphs, however, the representation theory of these C∗-
algebras is in its infancy. Although the primitive ideal space of higher-rank graph C∗-algebras
is well understood [9, 34], representations of k-graph C∗-algebras have only been systemati-
cally studied in the one-vertex case [13, 50, 14]. This motivated us to undertake the present
detailed study of monic representations of k-graph C∗-algebras and their unitary equivalence
classes. Despite the similarities between the Cuntz algebras and k-graph C∗-algebras which
we have highlighted above, there are fundamental structural differences between them: for
example, k-graph C∗-algebras need not be simple, nor is their K-theory known in general.
Thus, the extension of results on representations for Cuntz algebras to the k-graph context
is not automatic, and we are pleasantly surprised to have obtained such extensions in the
pages that follow.
The monic representations that we focus on in this paper were inspired in part by the
wavelet theory for higher-rank graphs which was developed in [27]. These wavelets relied on
the concept of Λ-semibranching function systems, introduced in [27] and further studied in
[24]. In this paper, we refine the Λ-semibranching function systems into a crucial technical
2
tool for studying monic representations, namely the Λ-projective systems of Definition 3.1.
Monic representations also have strong connections to Markov measures [19, 5] and Nelson's
universal representation of an abelian algebra [42]. Indeed, studying monic representations
enables us to convert questions about the representation theory of higher-rank graphs into
measure-theoretic questions (see Theorems 3.11 and 3.13 below).
This paper is organized as follows. We begin with an introductory section which reviews
the basic notation and terminology for higher-rank graphs, as well as the Λ-semibranching
function systems from [27]. Before turning our attention to a theoretic and systematic
analysis of the monic representations of finite k-graph C∗-algebras, Section 3 develops the
technical tools we will need for this analysis. The Λ-semibranching function systems of [27]
are refined in Section 3.1 into Λ-projective systems, and Section 3.2 analyzes the projection-
valued measure P = Pπ on the infinite path space Λ∞ which arises from a representation π
of C∗(Λ).
Section 3 ends by addressing the question of when representations of k-graph C∗-algebras
are disjoint or irreducible, see Theorems 3.11 and 3.13. To be precise, Theorem 3.11 shows
that for representations of C∗(Λ) arising from Λ-projective systems, the task of checking
when two representations are equivalent reduces to a measure-theoretical problem. Theorem
3.13 characterizes the commutant of such representations, enabling a precise description of
when a representation arising from a Λ-projective system is irreducible.
Having laid the necessary technical groundwork, we undertake the promised analysis
of monic representations of C∗(Λ) in Section 4. This section contains two of the main
results of this paper as well as a number of examples of monic representations. Theorem
4.2 establishes that, when Λ is a finite and source-free k-graph, monic representations of
C∗(Λ) are always unitarily equivalent to a Λ-projective representation on Λ∞. Theorem 4.5
gives an alternative, measure-theoretic characterization of when a Λ-semibranching function
system gives rise to a monic representation. More precisely, Theorem 4.5 proves that a Λ-
semibranching representation is monic if and only if the measure-theoretic subsets specified
by the Λ-semibranching function system (see Definition 2.7 below) generate the σ-algebra.
The final section, Section 5, relates monic representations to Nelson's universal Hilbert
space, which we denote H(Λ∞). Theorem 5.7 shows that every monic representation whose
associated Λ-projective system consists of positive functions is unitarily equivalent to a sub-
representation of the so-called "universal representation" of C∗(Λ) on H(Λ∞) which is de-
scribed in Proposition 5.3. In particular, this Theorem establishes that every representation
of C∗(Λ) which arises from a Λ-semibranching function system, as in [27], is unitarily equiv-
alent to a sub-representation of the universal representation.
Acknowledgments
The authors would like to thank Daniel Gon¸calves, Janos Englander and Alex Kumjian for
helpful discussions. E.G. was partially supported by the Deutsches Forschungsgemeinschaft
via the SFB 878 "Groups, Geometry, and Actions" of the Universitat Munster. S.K. was
supported by Basic Science Research Program through the National Research Foundation
of Korea (NRF) funded by the Ministry of Education (#2017R1D1A1B03034697). C.F. and
J.P. were partially supported by two individual grants from the Simons Foundation (C.F.
#523991; J.P. #316981). P.J. thanks his colleagues in the Math Department at the Uni-
3
versity of Colorado, for making a week-long visit there possible, for support, and for kind
hospitality. Progress towards the completion of this manuscript was made by the first three
named co-authors while in attendance at the Fields Institute (Toronto) and the Mathemat-
ical Congress of the Americas (Montreal) in 2017; we are grateful for their support of our
collaboration. C.F. also thanks IMPAN for hospitality during her visits to IMPAN, War-
saw, Poland, where some of this work was carried out (grant #3542/H2020/2016/2). This
paper was partially supported by the grant H2020-MSCA-RISE-2015-691246-QUANTUM
DYNAMICS.
2 Foundational material
2.1 Higher-rank graphs
We recall the definition of higher-rank graphs and their C∗-algebras from [37].
Let N = {0, 1, 2, . . .} denote the monoid of natural numbers under addition, and let
k ∈ N with k ≥ 1. We write e1, . . . ek for the standard basis vectors of Nk, where ei is the
vector of Nk with 1 in the i-th position and 0 everywhere else.
Definition 2.1. [37, Definition 1.1] A higher-rank graph or k-graph is a countable small
category1 Λ with a degree functor d : Λ → Nk satisfying the factorization property: for any
morphism λ ∈ Λ and any m, n ∈ Nk such that d(λ) = m + n ∈ Nk, there exist unique
morphisms µ, ν ∈ Λ such that λ = µν and d(µ) = m, d(ν) = n.
When discussing k-graphs, we use the arrows-only picture of category theory; thus, ob-
jects in Λ are identified with identity morphisms, and the notation λ ∈ Λ means λ is a
morphism in Λ. We often regard k-graphs as a generalization of directed graphs, so we call
morphisms λ ∈ Λ paths in Λ, and the objects (identity morphisms) are often called vertices.
For n ∈ Nk, we write
(1)
Λn := {λ ∈ Λ : d(λ) = n}
With this notation, note that Λ0 is the set of objects (vertices) of Λ, and we will call
elements of Λei (for any i) edges. We write r, s : Λ → Λ0 for the range and source maps in
Λ respectively. For vertices v, w ∈ Λ0, we define
vΛw := {λ ∈ Λ : r(λ) = v, s(λ) = w} and vΛn := {λ ∈ Λ : r(λ) = v, d(λ) = n}.
Our focus in this paper is on finite k-graphs with no sources. A k-graph Λ is finite if Λn
is a finite set for all n ∈ Nk. We say that Λ has no sources or is source-free if vΛn 6= ∅ for all
v ∈ Λ0 and n ∈ Nk. It is well known that this is equivalent to the condition that vΛei 6= ∅
for all v ∈ Λ and all basis vectors ei of Nk.
For m, n ∈ Nk, we write m ∨ n for the coordinatewise maximum of m and n. Given
λ, η ∈ Λ, we write
Λmin(λ, η) := {(α, β) ∈ Λ × Λ : λα = ηβ, d(λα) = d(λ) ∨ d(η)}.
(2)
1Recall that a small category is one in which the collection of arrows is a set.
4
If k = 1, then Λmin(λ, η) will have at most one element; this need not be true if k > 1.
For finite source-free k-graphs Λ, for each 1 ≤ i ≤ k, we can define the ith vertex matrix
(3)
Ai ∈ MΛ0(N) by
Ai(v, w) = vΛeiw.
Observe that the factorization property implies that AiAj = AjAi for 1 ≤ i, j ≤ k.
We now describe two fundamental examples of higher-rank graphs which were first men-
tioned in the foundational paper [37]. More examples of higher-rank graphs can be found in
Section 4.1 below.
Example 2.2.
(a) For any directed graph E, let ΛE be the category whose objects are the
vertices of E and whose morphisms are the finite paths in E. Then ΛE is a 1-graph
whose degree functor d : ΛE → N is given by d(η) = η (the number of edges in η).
(b) For k ≥ 1, let Ωk be the small category with
Obj(Ωk) = Nk,
and Mor(Ωk) = {(p, q) ∈ Nk × Nk : p ≤ q}.
Again, we can also view elements of Obj(Ωk) as identity morphisms, via the map
Obj(Ωk) ∋ p 7→ (p, p) ∈ Mor(Ωk). The range and source maps r, s : Mor(Ωk) →
If we define d : Ωk → Nk by
Obj(Ωk) are given by r(p, q) = p and s(p, q) = q.
d(p, q) = q − p, then one can check that Ωk is a k-graph with degree functor d.
Definition 2.3 ([37] Definitions 2.1). Let Λ be a k-graph. An infinite path in Λ is a k-graph
morphism (degree-preserving functor) x : Ωk → Λ, and we write Λ∞ for the set of infinite
paths in Λ. Since Ωk has a terminal object (namely 0 ∈ Nk) but no initial object, we think
of our infinite paths as having a range r(x) := x(0) but no source. For each m ∈ Nk, we
have a shift map σm : Λ∞ → Λ∞ given by
σm(x)(p, q) = x(p + m, q + m)
(4)
for x ∈ Λ∞ and (p, q) ∈ Ωk.
It is well-known that the collection of cylinder sets
Z(λ) = {x ∈ Λ∞ : x(0, d(λ)) = λ},
for λ ∈ Λ, form a compact open basis for a locally compact Hausdorff topology on Λ∞, under
reasonable hypotheses on Λ (in particular, when Λ is row-finite: see Section 2 of [37]). If Λ
is finite, then Λ∞ is compact in this topology.
We also have a partially defined "prefixing map" σλ : Z(r(λ)) → Z(λ) for each λ ∈ Λ:
λ(p, q),
x(p − d(λ), q − d(λ)),
λ(p, d(λ))x(0, q − d(λ)), p < d(λ) < q
q ≤ d(λ)
p ≥ d(λ)
σλ(x) = λx =(p, q) 7→
5
Remark 2.4. The factorization rule implies an important property of infinite paths: for any
x ∈ Λ∞ and m ∈ Nk, we have
x = x(0, m)σm(x).
Taking m = pej for an arbitrary p ∈ N reveals that every infinite path must contain infinitely
many edges of each color. Moreover, if we take m = (n, n, . . . , n) ∈ Nk for some n ≥ 1, the
factorization rule tells us that x(0, m) can be written uniquely as a "rainbow sequence" of
edges:
where d(f j
i ) = ei.
x(0, m) = f 1
1 f 1
2 · · · f 1
k f 2
1 · · · f 2
k f 3
1 · · · f n
k ,
For example, suppose Λ is a 2-graph. We can visualize Λ as arising from a 2-colored
graph (red and blue edges). Moreover, each infinite path x ∈ Λ∞ can be uniquely identified
with an infinite string of alternating blue and red edges (setting blue to be "color 1" and red
to be "color 2").
We stress that even finite k-graphs may have nontrivial infinite paths; in an infinite path,
the same edge may occur multiple times and even infinitely many times.
Now we introduce the C∗-algebra associated to a finite, source-free k-graph Λ.
Definition 2.5. Let Λ be a finite k-graph with no sources. A Cuntz–Krieger Λ-family is a
collection {tλ : λ ∈ Λ} of partial isometries in a C∗-algebra satisfying
(CK1) {tv : v ∈ Λ0} is a family of mutually orthogonal projections,
(CK2) tλtη = tλη if s(λ) = r(η),
(CK3) t∗λtλ = ts(λ) for all λ ∈ Λ,
(CK4) for all v ∈ Λ and n ∈ Nk, we have tv =Pλ∈vΛn tλt∗λ.
The Cuntz–Krieger C∗-algebra C∗(Λ) associated to Λ is the universal C∗-algebra generated
by a Cuntz–Krieger Λ-family.
The condition (CK4) implies that for all λ, η ∈ Λ, we have
t∗λtη = X(α,β)∈Λmin(λ,η)
tαt∗β.
(5)
It follows that C∗(Λ) = span{tαt∗β : α, β ∈ Λ, s(α) = s(β)}.
2.2 Λ-semibranching function systems and their representations
In [27], separable representations of C∗(Λ) were constructed by using Λ-semibranching func-
tion systems on measure spaces. A Λ-semibranching function system is a generalization of
the semibranching function systems studied by Marcolli and Paolucci in [40]. As established
in [40, 27], Λ-semibranching function systems (and their one-dimensional counterparts) give
rise to representations of C∗(Λ), and we provide examples of such representations in Section
4.1 below. Indeed, we build upon the notion of Λ-semibranching function systems in Sections
3 and 4 below to characterize the monic representations of higher-rank graphs.
Definition 2.6. [40, Definition 2.1] Let (X, µ) be a measure space. Suppose that, for each
1 ≤ i ≤ N, we have a measurable map σi : Di → X, for some measurable subsets Di ⊂ X.
The family {σi}N
i=1 is a semibranching function system if the following holds:
6
(a) Setting Ri = σi(Di), we have
µ(X \ ∪iRi) = 0,
µ(Ri ∩ Rj) = 0 for i 6= j.
(b) For each i, the Radon-Nikodym derivative
Φσi =
d(µ ◦ σi)
dµ
satisfies Φσi > 0, µ-almost everywhere on Di.
A measurable map σ : X → X is called a coding map for the family {σi}N
for all x ∈ Di.
Definition 2.7. [27, Definition 3.2] Let Λ be a finite k-graph and let (X, µ) be a measure
space. A Λ-semibranching function system on (X, µ) is a collection {Dλ}λ∈Λ of measurable
subsets of X, together with a family of prefixing maps {τλ : Dλ → X}λ∈Λ, and a family of
coding maps {τ m : X → X}m∈Nk, such that
(a) For each m ∈ Nk, the family {τλ : d(λ) = m} is a semibranching function system, with
i=1 if σ ◦ σi(x) = x
coding map τ m.
(b) If v ∈ Λ0, then τv = id, and µ(Dv) > 0.
(c) Let Rλ = τλ(Dλ). For each λ ∈ Λ, ν ∈ s(λ)Λ, we have Rν ⊆ Dλ (up to a set of measure
0), and
τλτν = τλν a.e.
(Note that this implies that up to a set of measure 0, Dλν = Dν whenever s(λ) = r(ν)).
(d) The coding maps satisfy τ m ◦ τ n = τ m+n for any m, n ∈ Nk. (Note that this implies
that the coding maps pairwise commute.)
Remark 2.8. We pause to note that condition (c) of Definition 2.7 above implies that
Dλ = Ds(λ) and Rλ ⊂ Rr(λ) for λ ∈ Λ. Also, when Λ is a finite 1-graph, the definition
In particular,
of a Λ-semibranching function system is not equivalent to Definition 2.6.
Definition 2.7(b) implies that the domain sets {Dv : v ∈ Λ0} must satisfy µ(Dv ∩ Dw) =
µ(Rv ∩ Rw) = 0 for v 6= w ∈ Λ0, but Definition 2.6 does not require that the domain sets Di
be mutually disjoint µ-a.e. In fact, Definition 2.7 implies what is called condition (C-K) in
Section 2.4 of [5]: up to a measure zero set,
Dv = [λ∈vΛm
Rλ
(6)
for all v ∈ Λ0 and m ∈ N, since Rv = τv(Dv) = id(Dv) = Dv. Also notice that in the above
decomposition the intersections Rλ ∩ Rλ′, Rλ 6= λ′, have measure zero. This condition is
crucial to making sense of the representation of C∗(Λ) associated to the Λ-semibranching
function system (see Theorem 2.10 below). As established in Theorem 2.22 of [5], in order to
obtain a representation of a 1-graph algebra C∗(Λ) from a semibranching function system,
one must also assume that the semibranching function system satisfies condition (C-K).
7
We pause to enumerate some properties of Λ-semibranching function systems, which can
be proved by routine computations.
Remark 2.9.
1. For any n ∈ Nk and any measurable E ⊆ X, we have
τλ(E)
and consequently
µ ◦ (τ n)−1 << µ
(7)
(τ n)−1(E) = [λ∈Λn
in any Λ-semibranching function system.
2. On Rλ, we have (τλ)−1 = τ n. Therefore, Condition (b) of Definition 2.6 implies that
µ ◦ (τλ)−1 << µ on Rλ, and d(µ◦(τλ)−1)
dµ
is nonzero a.e. on Rλ.
As established in [27], any Λ-semibranching function system gives rise to a representa-
tion of C∗(Λ) via 'prefixing' and 'chopping off' operators that satisfy the Cuntz-Krieger
relations. Intuitively, a Λ-semibranching function system is a way of encoding the Cuntz-
Krieger relations at the measure-space level: the prefixing map τλ corresponds to the partial
isometry sλ ∈ C∗(Λ). For the convenience of the reader, we recall the formula for these
Λ-semibranching representations of C∗(Λ).
Theorem 2.10. [27, Theorem 3.5] Let Λ be a finite k-graph with no sources and suppose
that we have a Λ-semibranching function system on a measure space (X, µ) with prefixing
maps {τλ : λ ∈ Λ} and coding maps {τ m : m ∈ Nk}. For each λ ∈ Λ, define an operator Sλ
on L2(X, µ) by
Sλξ(x) = χRλ(x)(Φτλ(τ d(λ)(x)))−1/2ξ(τ d(λ)(x)).
Then the operators {Sλ : λ ∈ Λ} generate a representation π of C∗(Λ) on L2(X, µ).
3 Representations of higher-rank graph C∗-algebras:
first analysis
We begin this section by developing the technical tools which we will rely on throughout
the paper: Λ-projective systems and projection valued measures. These tools enable us to
describe when certain representations of k-graph C∗-algebras are disjoint or irreducible, see
Theorems 3.11 and 3.13.
3.1 Λ-projective systems and representations
The definition of a Λ-projective system generalizes to the k-graph setting the definition of
a monic system in [19] (for the Cuntz algebras ON ) and [5] (in the case of Cuntz–Krieger
algebras OA). We have decided to change the name because even for OA, not every monic
system gives rise to a monic representation of OA. The word "projective" refers to the
cocycle-like Condition (b) of Definition 3.1.
Definition 3.1. Let Λ be a finite k-graph with no sources. A Λ-projective system on a
measure space (X, µ) is a Λ-semibranching function system on (X, µ), with prefixing maps
{τλ : Dλ → Rλ}λ∈Λ and coding maps {τ n : n ∈ Nk} together with a family of functions
{fλ}λ∈Λ ⊆ L2(X, µ) satisfying the following conditions:
8
dµ
(a) For any λ ∈ Λ, we have 0 6= d(µ◦(τλ)−1)
= fλ2;
(b) For any λ, ν ∈ Λ, we have fλ · (fν ◦ τ d(λ)) = fλν.
Thus, a Λ-projective system on (X, µ) consists of a Λ-semibranching function system plus
some extra information (encoded in the functions fλ). We have a certain amount of choice
for the functions fλ; we can take positive or negative (or imaginary!) roots of d(µ◦(τλ)−1)
for
fλ, as long as they satisfy the multiplicativity Condition (b) above.
dµ
Example 3.2. For any Λ-semibranching function system on (X, µ), there is a natural choice
of an associated Λ-projective system; namely, for λ ∈ Λn we define
fλ(x) := Φλ(τ n(x))−1/2χRλ(x).
(8)
Condition (a) is satisfied because of the hypothesis that the Radon–Nikodym derivatives be
strictly positive µ-a.e. on their domain of definition. Since the operators Sλ ∈ B(L2(X, µ))
of Theorem 2.10 are given by
and [27, Theorem 3.5] establishes that {Sλ}λ∈Λ is a Cuntz–Krieger family, Proposition 3.4
below shows that Equation (8) indeed describes a Λ-projective system.
Sλ(f ) = fλ · (f ◦ τ n),
Remark 3.3. Observe that Condition (a) of Definition 3.1 forces fλ(x) = 0 a.e. outside of
Rλ, since d(µ◦(τλ)−1)
is supported only on Rλ.
dµ
Condition (b) of Definition 3.1 is needed to associate a representation of C∗(Λ) to a
Λ-projective system. To be precise, we have:
Proposition 3.4. Let Λ be a finite, source-free k-graph. Suppose that a measure space (X, µ)
admits a Λ-semibranching function system with prefixing maps {τλ : λ ∈ Λ} and coding maps
{τ n : n ∈ Nk}. Suppose that {fλ}λ∈Λ is a collection of functions satisfying Condition (a) of
Definition 3.1. Then the maps {τλ}, {τ n} and {fλ}λ form a Λ-projective system on (X, µ)
if and only if the operators Tλ ∈ B(L2(X, µ)) given by
Tλ(f ) = fλ · (f ◦ τ d(λ))
(9)
form a Cuntz–Krieger Λ-family with each Tλ nonzero (and hence give a representation of
C∗(Λ)).
Proof. If the operators Tλ of Equation (9) form a nontrivial Cuntz–Krieger Λ-family, then it
is easily checked that the functions {fλ}λ∈Λ satisfy the hypotheses of Definition 3.1.
On the other hand, suppose that (X, µ) admits a Λ-projective system with prefixing maps
{τλ}λ∈Λ, coding maps {τ n}n∈Nk, and functions {fλ}λ∈Λ. We will show that the operators
{Tλ} of Equation (9) satisfy Conditions (CK1)–(CK4).
For (CK1), observe that if v ∈ Λ0, Tv(f ) = fv · (f ◦ τ 0) is supported on Dv = Rv by
Condition (a) of Definition 3.1. Moreover, since v = v2 for any v ∈ Λ0, and τv = idDv = τ 0,
Condition (b) of Definition 3.1 implies that
fv = fv · (fv ◦ τ 0) = f 2
v ⇒ fv = χDv.
9
Consequently, Tv(f ) = χDv · f . Since the sets {Dv = Rv}v∈Λ0 are disjoint (up to a set of
measure zero), it follows that {Tv : v ∈ Λ0} is a set of mutually orthogonal projections; in
other words, (CK1) holds.
For (CK2), fix λ, ν ∈ Λ with s(λ) = r(ν). Since fν(x) = 0 unless x ∈ Rν, we see that
TλTν(f )(x) =(0,
x 6∈ τλ(Rν)
fλ(τλ ◦ τν(y)) · fν(τν(y)) · f (y), x = τλτν(y).
On the other hand, Condition (b) of Definition 3.1 implies that if x = τλτν(y),
fλ(τλ ◦ τν(y)) · fν(τν(y)) = fλ(x) · fν(τ d(λ)(x)) = fλν(y).
This implies that TλTν = Tλν as claimed.
To check (CK3), we first compute that T ∗λ f = f ◦ τλ · fλ ◦ τλ · Φλ. Alternatively,
T ∗λ f =
χDλ · (f ◦ τλ)
fλ ◦ τλ
.
(10)
Condition (CK3), and the fact that the operators Tλ are partial isometries, now follow from
straightforward calculations. Finally, an easy computation establishes that TλT ∗λ (f ) = χRλ·f
for any λ ∈ Λ, from which (CK4) follows.
We call the representation given in Equation (9) a Λ-projective representation.
The following Proposition enables us to translate a Λ-projective system on (X, µ) to a
Λ-projective system on (X, µ′) for any measure µ′ which is equivalent to µ.
Proposition 3.5. Let Λ be a finite k-graph with no sources. Suppose we are given a Λ-
projective system {τλ : λ ∈ Λ}, {τ n : n ∈ Nk} and {fλ : λ ∈ Λ} on a measure space (X, µ).
Let µ′ be a measure equivalent to µ, and set g1(x) = dµ′
dµ′ (x). If we define
{ fλ}λ∈Λ by
dµ (x) and g2(x) = dµ
· fλ(x), λ ∈ Λ,
(11)
fλ(x) = pg1 ◦ τ d(λ)(x)
pg1(x)
then {τλ : λ ∈ Λ}, {τ n : n ∈ Nk} and { fλ}λ∈Λ give a Λ-projective system on (X, µ′).
Moreover, the associated representations {Tλ : λ ∈ Λ} and { Tλ : λ ∈ Λ} of C∗(Λ) on
L2(X, µ) and L2(X, µ′) given by Equation (9) of Proposition 3.4 are unitarily equivalent via
the unitary U given by
U(f )(x) = s dµ
dµ′
(x) · f (x), f ∈ L2(X, µ), U−1(h)(x) = sdµ′
dµ
(x) · h(x), h ∈ L2(X, µ′).
Proof. We leave the verification of this proposition to the reader.
Proposition 3.6 below is the analog of Proposition 2.11 of [19] for Λ-projective systems.
10
Proposition 3.6. Let Λ be a finite k-graph with no sources. Suppose we are given two
Λ-projective systems on X, with the same prefixing and coding maps {τλ : λ ∈ Λ}, {τ n :
n ∈ Nk}, but with different measures µ, µ′ and Λ-projective functions {fλ}λ∈Λ for (X, µ) and
{f′λ}λ∈Λ for (X, µ′).
Let dµ′ = h2dµ + dν be the Lebesgue-Radon-Nikodym decomposition, with h ≥ 0 and ν
singular with respect to µ. Then there is a partition of X into Borel sets X = A ∪ B such
that:
(a) The function h is supported on A, ν is supported on B, and µ(B) = 0, ν(A) = 0.
(b) The sets A, B are invariant under τ n for all n ∈ Nk, i.e.,
(τ n)−1(A) = A,
and (τ n)−1(B) = B.
λ << ν and kλ :=q d(ν◦τ −1
dν
λ )
(c) We have ν ◦ τ−1
(d) f′λ · h = fλ · (h ◦ τ d(λ)) µ-a.e. on A and f′λ = kλ ν-a.e. on B.
Proof. We start by proving (a) and (b) together. Let B be the support of ν, and observe
that µ( B) = 0. We observe that the definitions of Λ-semibranching function systems and
is supported on B.
Λ-projective systems, together with the fact that (τ n)−1( B) =Sλ∈Λn τλ( B), imply that
have µ-measure zero. Therefore we can take the orbit B of B under the functions {τ n :
n ∈ Nk} and {τλ : λ ∈ Λ}, and B will then have µ-measure zero. Let A := X\B. Then A
contains the support of µ, and we can choose h to be supported on A. Moreover, ν(A) = 0.
By construction, A and B are invariant under τ n. This establishes (a) and (b). To prove
(c), let E be a Borel set with ν(E) = 0. Then ν(E ∩ B) = 0, so the fact that µ vanishes on
B implies that µ′(E ∩ B) = 0. We consequently have µ′(τ−1
λ (E ∩ B)) = 0, which means that
µ′(τ−1
λ are
supported on B, it follows that kλ is supported on B. To see (d), let f be a bounded Borel
function supported on A. Then we have
λ (E)) = 0. Since B is invariant under τ−1
λ (E)∩ B) = 0, so ν(τ−1
λ and ν and ν ◦ τ−1
(τ n)−1( B) and (τλ)−1( B)
f
ZA f′λ2 f h2 dµ =ZA f′λ2 f dµ′ =ZA
(f ◦ τλ) h2 dµ =ZX
f (h2 ◦ τ d(λ))fλ2dµ,
=ZA
=ZX
dµ′ =ZA
d(µ′ ◦ τ−1
λ )
(f ◦ τλ) dµ′
d(µ′)
(f ◦ τλ) (h2 ◦ τ d(λ) ◦ τλ) dµ =ZX
f (h2 ◦ τ d(λ)) d(µ ◦ τ−1
λ )
which implies the first relation. The second relation follows from the fact that µ′B = ν.
3.2 Projection valued measures
The second technical tool which underpins our analysis of the monic representations of C∗(Λ)
is the projection valued measure associated to a representation of C∗(Λ). Our work in this
section is inspired by Dutkay, Haussermann, and Jorgensen [19, 20].
11
Definition 3.7. Let Λ be a finite k-graph with no sources. Given a representation {tλ}λ∈Λ
of a k-graph C∗-algebra C∗(Λ) on a Hilbert space H, we define a projection valued function
P on Λ∞ by
P (Z(λ)) = tλt∗λ
for all λ ∈ Λ.
In the proof (Proposition 3.9) that P indeed defines a projection-valued measure on Λ∞,
we rely on the following well-known Lemma. Thus, in our application, X = Λ∞ and Fn will
be the σ-algebra generated by the cylinder sets Z(λ) with d(λ) = (n, . . . , n).
Lemma 3.8 (Kolmogorov Extension Theorem, [36, 49]). Let (X,Fn, νn)n∈N be a sequence
of probability measures (νn)n∈N on the same space X, each associated with a σ-algebra Fn;
further assume that (X,Fn, νn)n∈N form a projective system, i.e., an inverse limit. Suppose
that Kolmogorov's consistency condition holds:
νn+1Fn = νn.
Then there is a unique extension ν of the measures (νn)n∈N to the σ-algebraWn∈N Fn gener-
ated bySn∈N Fn.
Proposition 3.9. Let Λ be a finite k-graph with no sources. Given a representation {tλ}λ∈Λ
of a k-graph C∗-algebra C∗(Λ) on a Hilbert space H, the function P of Definition 3.7 extends
to a projection valued measure on the Borel σ-algebra Bo(Λ∞) of the infinite path space Λ∞.
Proof. Recall from the proof of [27, Lemma 4.1] that
{Z(λ) : d(λ) = (n, n, . . . , n) for some n ∈ N}
generates the topology on Λ∞. Thus, Bo(Λ∞) = lim−→Fn. By Lemma 3.8, it therefore suffices
to show that
P (Z(λη))
P (Z(λ)) = Xη∈s(λ)Λ(1,...,1)
P (Z(λ)) = tλt∗λ = tλ Xη∈s(λ)Λ(1,...,1)
tηt∗η t∗λ = Xη∈s(λ)Λ(1,...,1)
whenever d(λ) = (n, . . . , n) for some n ∈ N. However, this follows immediately from (CK4):
P (Z(λη).
(12)
We now record some properties of P which we will rely on in the sequel. The equations
below are the analogues for k-graphs of Equations (2.7) and (2.8) and (2.13) of [20].
Proposition 3.10. Let Λ be a row-finite, source-free k-graph, and fix a representation {tλ :
λ ∈ Λ} of C∗(Λ).
(a) For λ, η ∈ Λ with s(λ) = r(η), we have tλP (Z(η))t∗λ = P (σλ(Z(η))), where σλ is the
prefixing map on Λ∞ given in Equation (4).
12
(b) For any fixed n ∈ Nk, we have
Xλ∈f (η)Λn
tλP (σ−1
λ (Z(η)))t∗λ = P (Z(η));
(c) For any λ, η ∈ Λ with r(λ) = r(η), we have tλP (σ−1
(d) When λ ∈ Λn, we have tλP (Z(η)) = P ((σn)−1(Z(η)))tλ.
Proof. Straightforward calculation.
λ (Z(η))) = P (Z(η))tλ;
3.3 Disjoint and irreducible representations
In this section we will derive from the technical results in Section 3.1 important consequences
that detail when representations of k-graph C∗-algebras are disjoint or irreducible. In par-
ticular, Theorem 3.11 suggests the importance of dealing with Λ-projective systems with
non-negative functions fλ. We will focus more exclusively on such Λ-projective systems in
Section 5 below.
Theorem 3.11. (C.f. Theorem 2.12 of [19]) Let Λ be a finite k-graph with no sources.
Suppose we are given two Λ-projective systems on the infinite path space Λ∞ with the standard
prefixing and coding maps {σλ : λ ∈ Λ}, {σn : n ∈ Nk}, but associated to different measures
µ, µ′ and different Λ-projective families of non-negative functions {fλ}λ∈Λ on (Λ∞, µ), and
{f′λ}λ∈Λ on (Λ∞, µ′). Then the two associated representations {Tλ : λ ∈ Λ} and {T ′λ : λ ∈ Λ}
of C∗(Λ) given by Equation (9) of Proposition 3.4 are disjoint if and only if the measures µ
and µ′ are mutually singular.
Proof. If the representations are not disjoint, there exist subspaces Hµ ⊆ L2(Λ∞, µ) and
Hµ′ ⊆ L2(Λ∞, µ′), preserved by their respective representations, and a unitary W : Hµ →
Hµ′ such that
W TλHµ = T ′λHµ′ W,
W T ∗λHµ = (T ′λ)∗Hµ′ W.
The fact that each operator T ∗λ also preserves Hµ implies that
W TλT ∗λHµ = W TλHµT ∗λHµ = T ′λHµ′ W T ∗λHµ
= T ′λ(T ′λ)∗Hµ′ W.
Moreover, it follows easily from the formulas for Tλ and T ∗λ in Equations (9) and (10) that
TλT ∗λ = MχZ(λ) = T ′λ(T ′λ)∗.
In other words, the representations of C(Λ∞) given by χZ(λ) 7→ TλT ∗λ and χZ(λ) 7→ T ′λ(T ′λ)∗
(on L2(Λ∞, µ) and L2(Λ∞, µ′) respectively) are multiplication representations. Since W im-
plements a unitary equivalence between their subrepresentations on Hµ and Hµ′ respectively,
Theorem 2.2.2 of [2] tells us that the measures µ, µ′ cannot be mutually singular.
For the converse, assume that the representations are disjoint and that the measures are
not mutually singular. Then, use Proposition 3.6 and decompose dµ′ = h2dµ + dν, with the
13
subsets A, B as in Proposition 3.6. Define the operator W on L2(Λ∞, µ′) by W (f ) = f · h if
f ∈ L2(A, µ′), and W (f ) = 0 on the orthogonal complement of L2(A, µ′) ⊆ L2(Λ∞, µ′). Since
A is invariant under τ n for all n, L2(A, µ′) is an invariant subspace for the representation.
To check that W is intertwining, we use part (d) of Proposition 3.6 and the non-negativity
condition on {fλ} and {f′λ} to obtain the almost-everywhere equalities
TλW (f ) = fλ(h ◦ τ d(λ))(f ◦ τ d(λ)) = f′λ h (f ◦ τ d(λ)) = W T ′λ(f ).
Since W intertwines the representations {Tλ}λ∈Λ,{T ′λ}λ∈Λ of C∗(Λ), we must have W = 0;
hence h = 0, so µ, µ′ are mutually singular.
Remark 3.12. As a Corollary of Theorem 3.11, we see that the examples of Markov measures
introduced in [24, Section 4.2] generate representations of C∗(Λ) disjoint from the represen-
tation of [27, Theorem 3.5]; see also Example 4.10 below. In fact, these Markov measures
are mutually singular with the Perron–Frobenius measure [19], [33].
Theorem 3.13. (C.f. Theorem 2.13 of [19]) Let Λ be a finite k-graph with no sources.
Suppose that the infinite path space Λ∞ admits a Λ-projective system on (Λ∞, µ) for some
measure µ, with the standard prefixing maps {σλ : λ ∈ Λ} and coding maps {σn : n ∈ Nk}
of Definition 2.3. Let {Tλ : λ ∈ Λ} be the associated representation of C∗(Λ). Then:
(a) The commutant of the operators {Tλ : λ ∈ Λ} consists of multiplication operators by
functions h with h ◦ σn = h, µ-a.e for all n ∈ Nk.
(b) The representation given by {Tλ : λ ∈ Λ} is irreducible if and only if the coding maps
σn are jointly ergodic with respect to the measure µ, i.e., the only Borel sets A ⊂ Λ∞
with (σn)−1(A) = A for all n are sets of measure zero, or of full measure.
Proof. We first observe that the commutant of {Tλ}λ∈Λ is contained in C(Λ∞)′ ⊆ B(L2(Λ∞, µ))
and hence consists of multiplication operators. The proof of part (a) is then a straightforward
calculation, and part (b) follows from part (a) and the definition of ergodicity.
4 Monic representations of finite k-graph algebras
The first main result of this section, Theorem 4.2, establishes that every monic representation
of a finite, strongly connected k-graph algebra C∗(Λ) is unitarily equivalent to a Λ-projective
representation of C∗(Λ) on L2(Λ∞, µπ), where the measure µπ arises from the representation.
(See Definition 4.1 and Equation (14) below for details.) After proving Theorem 4.2, we
examine a variety of examples of representations of C∗(Λ), and identify which representations
are monic. This analysis requires our second main result, Theorem 4.5, which provides a
measure-theoretic characterization of when a Λ-semibranching representation is monic.
Definition 4.1. Let Λ be a finite k-graph with no sources. A representation {tλ : λ ∈ Λ}
of a k-graph on a Hilbert space H is called monic if tλ 6= 0 for all λ ∈ Λ, and there exists a
vector ξ ∈ H such that
spanλ∈Λ{tλt∗λξ} = H.
14
From the projection valued measure P associated to {tλ : λ ∈ Λ} as in Theorem 3.9, we
obtain a representation π : C(Λ∞) → B(H):
π(f ) =ZΛ∞
f (x)dP (x),
which gives, for λ ∈ Λ,
π(χZ(λ)) =ZΛ∞
χZ(λ) (x) dP (x) = P (Z(λ)) = tλt∗λ.
(13)
Since we can view C(Λ∞) as a subalgebra of C∗(Λ) via the embedding χZ(λ) 7→ tλt∗λ, the
representation π is often understood as the restriction of the representation {tλ}λ∈Λ to the
"diagonal subalgebra" span{tλt∗λ}λ∈Λ.
If the representation {tλ}λ is monic, there is a cyclic vector ξ ∈ H for π. This induces
we obtain a Borel measure µπ on Λ∞ given by
µπ(Z(λ)) = hξ, P (Z(λ))ξi = hξ, tλt∗λξi.
(14)
Theorem 4.2. Let Λ be a finite k-graph with no sources. If {tλ}λ∈Λ is a monic representation
of C∗(Λ) on a Hilbert space H, then {tλ}λ∈Λ is unitarily equivalent to a representation
{Sλ}λ∈Λ associated to a Λ-projective system on (Λ∞, µπ), which is associated to the standard
coding and prefixing maps σn, σλ of Definition 2.3.
Conversely, if we have a representation of C∗(Λ) on L2(Λ∞, µ) which arises from a Λ-
projective system associated to the standard coding and prefixing maps σn, σλ, then the rep-
resentation is monic.
By Example 3.2, this implies that a Λ-semibranching function system on (Λ∞, µ), for any
Borel measure µ, gives rise to a monic representation of C∗(Λ).
Proof. Suppose that the representation {tλ}λ∈Λ of C∗(Λ) is monic, and let ξ ∈ H be a cyclic
vector for C(Λ∞). Note that the map W : C(Λ∞) → H given by
W (f ) = π(f )ξ
is linear. Moreover, if we think of C(Λ∞) as a dense subspace of L2(Λ∞, µπ), the operator
W is isometric:
L2 =ZΛ∞ f2 dµπ = hξ, π(f2)ξi = kπ(f )ξk2 = kW (f )k2.
kfk2
Therefore W extends to an isometry from L2(Λ∞, µπ) to H. Since W is also onto (because
the representation is monic), W is a surjective isometry; that is, W is a unitary.
Moreover, for any f ∈ C(Λ∞) and any ϕ ∈ L2(Λ∞, µπ), we have
π(f )W (ϕ) = π(f )π(ϕ)ξ = π(f · ϕ)ξ = W (f · ϕ).
Thus, unitarity of W implies that W ∗π(f )W acts on L2(Λ∞, µπ) by multiplication by f :
W ∗π(f )W = Mf and W Mf W ∗ = π(f ).
(15)
15
Now define the operator Sλ = W ∗tλW for λ ∈ Λ. By construction, the operators {Sλ}λ∈Λ
also give a representation of C∗(Λ). Moreover, since W is a unitary,
SλS∗λ(f ) = W ∗tλt∗λW (f ) = W ∗π(χZ(λ))π(f )ξ = W ∗π(χZ(λ) · f )ξ
= W ∗W (χZ(λ) · f ) = χZ(λ) · f.
(16)
Let 1 denote the characteristic function of Λ∞, and define a function fλ ∈ L2(Λ∞, µπ) by
fλ = Sλ1 = W ∗tλξ.
We will now show that the functions fλ, combined with the usual coding and prefixing maps
{σn, σλ}n,λ on Λ∞, form a Λ-projective system on (Λ∞,Bo(Λ∞), µπ). To that end, we will
invoke Proposition 3.10. Since
P (Z(ν)) = π(χZ(ν))
for any ν ∈ Λ, and the proof of [27, Lemma 4.1] shows that characteristic functions of
cylinder sets densely span L2(Λ∞, µπ), the equalities established in Proposition 3.10 still
hold if we replace P (Z(ν)) by π(f ) for any f ∈ L2(Λ∞, µπ). In particular, noting that
χ(σn)−1(Z(ν)) = χZ(ν) ◦ σn
and
χσ−1
λ (Z(ν)) = χZ(ν) ◦ σλ
Part (d) of Proposition 3.10 implies that if d(λ) = n,
tλπ(f ) = π(f ◦ σn)tλ
(17)
and Part (c) implies that
(18)
Let f ∈ L2(Λ∞, µπ) and let n = d(λ). By using Part (d) of Proposition 3.10, Equation
t∗λπ(f ) = π(f ◦ σλ)t∗λ.
(15), and the fact that W is a unitary, we obtain
Sλ(f ) = W ∗tλW (f ) = W ∗tλπ(f )ξ
= W ∗π(f ◦ σn)tλξ = W ∗π(f ◦ σn)W W ∗tλξ
= (f ◦ σn) · fλ.
In order to show that {Sλ}λ∈Λ is a Λ-projective representation, then, Proposition 3.4 tells
us that it remains to check that the standard prefixing and coding maps make (Λ∞, µπ) into
a Λ-semibranching function system, and that Condition (a) of Definition 3.1 holds for the
functions fλ.
To establish Condition (a), we work indirectly. Since W is a unitary, we have (for any
f ∈ L2(Λ∞, µπ) and any λ ∈ Λn)
ZΛ∞ fλ2 · f dµπ = hSλ(1), Sλ(1) · fiL2 = hW ∗tλ(ξ), Mf W ∗tλ(ξ)iL2
= htλξ, W Mf W ∗(tλξ)iH = hξ, t∗λπ(f )tλξiH = hξ, π(f ◦ σλ)ξiH
=ZΛ∞
f ◦ σλ dµπ =ZΛ∞
f d(µπ ◦ σ−1
λ ).
16
If E ⊆ Λ∞ is any set for which µπ(E) = 0, then taking f = χE above shows that
µπ ◦ σ−1
λ (E) = 0 also – in other words,
The uniqueness of Radon-Nikodym derivatives then implies that
µπ ◦ σ−1
λ << µπ.
(19)
fλ2 =
d(µπ ◦ σ−1
λ )
d(µπ)
.
In other words, Condition (a) of Definition 3.1 holds.
We now show that fλ 6= 0 a.e. on Z(λ). Define Eλ ⊆ Z(λ) by
Eλ := {x ∈ Z(λ) : fλ(x) = 0}.
λ (Eλ) = π(χσ−1
Then 0 =REλ fλ2 dµπ = µπ ◦ σ−1
λ (Eλ)) = t∗λπ(χEλ)tλ by Proposition 3.10.
By hypothesis, tλ 6= 0, so there exists ζ ∈ H such that tλ(ζ) 6= 0. However, for any ζ,
htλ(ζ), π(χEλ)tλ(ζ)i = htλ(ζ), π(χEλ)2π(χZ(λ))tλ(ζ)i = ht∗λπ(χEλ)tλ(ζ), t∗λπ(χEλ)tλ(ζ)i = 0
by the Cuntz–Krieger relations and the fact that π(C(Λ∞)) is abelian.
In other words,
π(χEλ) is orthogonal to the range projection π(χZ(λ)) of tλ.
On the other hand, χEλχZ(λ) = χEλ since Eλ ⊆ Z(λ).
It follows that π(χEλ) = 0;
equivalently, µπ(Eλ) = 0. In other words, the set Eλ ⊆ Z(λ) of points where fλ = 0 has
µπ-measure zero, as claimed.
Similarly, for any set F ⊆ Z(s(λ)) such that µπ(F ) = 0, taking f = χσλ(F ) reveals that
0 = µπ(F ) = µπ ◦ σ−1
λ (σλ(F )) =Zσλ(F ) fλ2 dµπ.
Since fλ2 > 0 a.e. on Z(λ) ⊇ σλ(F ), we must have µπ ◦ σλ(F ) = 0 and hence µπ ◦ σλ << µπ.
Furthermore, the Radon-Nikodym derivative d(µπ◦σλ)
is nonzero µπ-a.e. on Z(s(λ)). To see
d(µπ )
this, we set
λ )(σλ(E)) = 0.
Equation (19) therefore implies that µπ(E) = (µπ ◦ σ−1
[24, Proposition 4.1] now implies that the standard prefixing and coding maps make
(Λ∞, µ) into a Λ-semibranching function system. Consequently, the functions fλ make
{Sλ}λ∈Λ into a Λ-projective representation, which is unitarily equivalent to our initial monic
representation by construction.
For the converse, suppose that {tλ}λ∈Λ is a representation of C∗(Λ) on L2(Λ∞, µ), for
some Borel measure µ, which arises from a Λ-projective system {fλ}λ∈Λ associated to the
standard coding and prefixing maps {σn, σλ}n,λ. The computations from Proposition 3.4
establish that tλt∗λ is given by multiplication by χRλ = χZ(λ). Consequently, 1 = χΛ∞ is a
cyclic vector for C(Λ∞) ⊆ C∗(Λ). Thus, {tλ}λ∈Λ is monic.
17
and observe that
E =(cid:26)x ∈ Z(s(λ)) :
µπ(σλ(E)) =ZE
d(µπ ◦ σλ)
d(µπ)
= 0(cid:27)
d(µπ ◦ σλ)
d(µπ)
dµπ = 0.
Remark 4.3. In the final section of their paper [5], Bezuglyi and Jorgensen studied the
relationship between semibranching function systems and monic representations of Cuntz–
Krieger algebras (1-graph C∗-algebras). Theorem 5.6 of [5] establishes that within a specific
class of semibranching function systems, which the authors term monic systems, those for
which the underlying space is the infinite path space Λ∞ are precisely the systems which give
rise to monic representations of the Cuntz–Krieger algebra. The Λ-projective systems studied
in Section 3.1 constitute our extension to k-graphs of the monic systems for Cuntz–Krieger
algebras. Thus, even in the case of 1-graph algebras (Cuntz–Krieger algebras), our Theorem
4.2 is substantially stronger than Theorem 5.6 of [5]: our Theorem 4.2 gives a complete
characterization of monic representations, without the hypothesis that such representations
arise from a monic or Λ-projective system.
Theorem 4.4. Let Λ be a finite, source-free k-graph, and let {Sλ}λ∈Λ, {Tλ}λ∈Λ be two
monic representations of C∗(Λ). Let µS, µT be the measures on Λ∞ associated to these
representations as in (14). The representations {Sλ}λ∈Λ, {Tλ}λ∈Λ are equivalent if and only
if the measures µS and µT are equivalent and there exists a function h on Λ∞ such that
dµS
dµT
= h2
and
f S
λ =
h ◦ σn
h
f T
λ
for all λ ∈ Λ with d(λ) = n.
(20)
(21)
Proof. Suppose {Sλ}λ∈Λ, {Tλ}λ∈Λ are equivalent representations of C∗(Λ). From Theorem
3.11, it follows that the associated measures µS, µT are equivalent. Let W : L2(Λ∞, µS) →
L2(Λ∞, µT ) be the intertwining unitary for them. Then the two representations are also
equivalent when restricted to the diagonal subalgebra C∗({tλt∗λ : λ ∈ Λ}). By linearity, we
can extend the formula from Equation (16) to all of C(Λ∞). It follows that πS, πT are both
given on C(Λ∞) by multiplication:
πS(φ) = Mφ and πT (φ) = Mφ ∀ φ ∈ C(Λ∞).
Since W intertwines a dense subalgebra – namely πS(C(Λ∞)) – of the maximal abelian sub-
algebra L∞(Λ∞, µS) ⊆ B(L2(Λ∞, µS)) which consists of multiplication operators, with the
dense subalgebra πT (C(Λ∞)) ⊆ L∞(Λ∞, µT ), the unitary W must be given by multiplication
by some nowhere-vanishing function h on Λ∞: W (f ) = hf. Moreover, since W is a unitary,
ZΛ∞ W (f )2dµT =ZΛ∞ f2h2dµT =ZΛ∞ f2dµS
for all f ∈ L2(Λ∞, µS),
which implies (20).
From the intertwining property Tλ W = W Sλ we obtain, for any f ∈ L2(Λ∞, µS) and
any λ with d(λ) = n, that
Tλ W (f ) = W Sλ(f ), that is, f T
λ (h ◦ σn)(f ◦ σn) = hf S
λ (f ◦ σn).
Take f = 1 and we obtain that
f T
λ
h ◦ σn
h
= f S
λ
18
as claimed in (21).
For the converse, suppose that the measures µS, µT are equivalent and there is a function
h on Λ∞ satisfying (20) and (21). Then define W : L2(Λ∞, µS) → L2(Λ∞, µT ) by
W f = hf ;
it is then straightforward to check that W Sλ = TλW and that W is a unitary.
4.1 Λ-semibranching function systems and monic representations
In this section, we discuss several examples of Λ-semibranching function systems and identify
which of them give rise to monic representations of C∗(Λ) – or, equivalently, which are
unitarily equivalent to Λ-semibranching function systems on the infinite path space. First,
we provide another characterization of monic representations. The next theorem shows that
a Λ-semibranching system on (X, µ) induces a monic representation of C∗(Λ) if and only if
its associated range sets generate the σ-algebra of X. To state our result more precisely, we
will denote by (X,F , µ) the measure space associated to L2(X, µ); in particular, F is the
standard σ-algebra associated to L2(X, µ).
Theorem 4.5. Let Λ be a finite, source-free k-graph and let {tλ}λ∈Λ be a Λ-semibranching
representation of C∗(Λ) on L2(X,F , µ) with µ(X) < ∞. Let R be the collection of sets which
are modifications of range sets Rλ by sets of measure zero; that is, each element X ∈ R has
the form
X = Rλ ∪ S
or X = Rλ\S
for some set S of measure zero. Let σ(R) be the σ-algebra generated by R. The represen-
tation {tλ}λ∈Λ is monic, with cyclic vector χX ∈ L2(X,F , µ), if and only if σ(R) = F . In
particular, for a monic representation {tλ}λ∈Λ, the set
S :=n nXi=1
aitλit∗λiχX n ∈ N, λi ∈ Λ, ai ∈ Co =n nXi=1
aiχRλi n ∈ N, λi ∈ Λ, ai ∈ Co
is dense in L2(X,F , µ).
Proof. Suppose first that the representation {tλ}λ∈Λ is monic and that χX is a cyclic vector
for the representation. As computed in the proof of Theorem 3.4 of [27], we have
tλt∗λ(χX) = χRλ.
Therefore, our hypothesis that χX is a cyclic vector implies that for any f ∈ L2(X,F , µ),
there is a sequence (fj)j, with fj ∈ span{χRλ : λ ∈ Λ}, such that
j→∞ZX fj − f2 dµ = 0.
lim
In particular, (fj) → f in measure.
19
For any σ-algebra T , standard measure-theoretic results [41, Proposition 6] imply that
since µ(X) < ∞, convergence in measure among T -a.e. finite measurable functions on
(X,T , µ) is metrized by the distance
dT (f, g) :=ZΩ
f − g
1 + f − g
dµ.
Moreover, dT makes the space of S-a.e. finite measurable functions into a complete metric
space (this can be seen, for example, by combining Proposition 1 and Corollary 7 of [41]).
The fact that (fj)j → f in measure in (X,F , µ), and that fj ∈ L2(X, σ(R), µ) for all j,
implies that (fj)j is a Cauchy sequence with respect to both dF and dσ(R). Consequently,
the limit f of (fj)j must also be a σ(R)-a.e. finite measurable function.
In other words,
every f ∈ L2(X,F , µ) is in fact in L2(X, σ(R), µ). Since R ⊆ F by construction we must
have σ(R) = F , as desired.
For the converse, assume σ(R) = F . We begin by observing that
R := {finite unions of elements in R}
aiχBi n ∈ N, Bi ∈ σ( R), ai ∈ Co
is a subalgebra of P(X) – that is, closed under finite unions and complements. Closure under
finite unions follows from the definition, while the second claim follows from Equation (6).
Moreover, σ(R) = σ( R) = F , so
S :=n nXi
is dense in L2(X,F , µ). Therefore, the Carath´eodory/Kolmogorov extension theorem implies
restricted to R induces a unique (extended) measure on F = σ(R),
that the measure µ R
which we still call µ. (This is indeed the original measure on L2(X,F , µ) by the uniqueness
of the extension.)
To show that the vector χX is monic, equivalently that the set S is dense in L2(X,F , µ)
equipped with the usual metric dL2(X,F ,µ) coming from the L2 norm, we invoke a standard
if (Q, dQ) is a metric space, and if Σ ⊆ Q is a dense subset of
fact about metric spaces:
(Q, dQ), then any other subset Σ ⊆ Q having the property
∀ǫ > 0, ∀x ∈ Σ, ∃ xǫ ∈ Σ with dQ(x, xǫ) < ǫ
is also dense in (Q, dQ). We wish to apply this fact in the setting where
(Q, dQ) = (L2(X,F , µ), dL2(X,F ,µ)), with Σ = S, Σ = S.
Choose s ∈ S and fix ǫ > 0. Without loss of generality we can assume s =Pn
i aiχBi for
some n ∈ N, Bi ∈ σ( R), ai 6= 0. Define A :=Pn
i ai ∈ R. The Carath´eodory/Kolmogorov
extension theorem2 also guarantees that for any B ∈ σ( R) = F and for any ǫ > 0, there
exists A B
ǫ(cid:17) < ǫ,
µ(cid:16) B∆A B
ǫ ∈ R with
2See [16] Page 452, Appendix: Measure theory, Exercise 3.1.
20
where ∆ denotes symmetric difference. In other words, for each i, there exists Aǫ
that
i ∈ R such
µ(Bi∆Aǫ
i) <
ǫ2
A2 , or, equivalently,(cid:16)µ(Bi∆Aǫ
i)(cid:17)1/2
<
ǫ
A
.
Thus, setting sǫ :=Pn
as desired.
i aiχAǫ
i and using the triangle inequality yields that dL2(X,F ,µ)(s, sǫ) < ǫ,
Remark 4.6. Using the characterization of monic representations from Theorem 4.5, it is
straightforward to check that the Λ-semibranching function systems detailed in [24, Example
3.5 and Section 4] generate monic representations of C∗(Λ). Similarly, suppose Λ = Λ1×Λ2 is
a product k-graph and we have Λi-semibranching function systems on measure spaces (Xi, µi)
for i = 1, 2, such that the associated Λi-semibranching representations are monic. Then
Theorem 4.5 combines with [24, Proposition 3.4] to tell us that the product Λ-semibranching
function system on (X1 × X2, µ1 × µ2) also gives rise to a monic representation of C∗(Λ).
We now proceed to analyze several other examples of representations arising from Λ-
semibranching function systems and establish which ones are monic representations.
Example 4.7. We present here an example of a Λ-semibranching representation on a 1-graph
that is not monic. The 1-graph Λ has two vertices v1 and v2 and three edges f1, f2 and f3.
f2
f1
v1
v2
f3
Let X be the closed unit interval [0, 1] of R with the usual Lebesgue σ-algebra and measure
µ. For v1 and v2, let Dv1 = [0, 1
2, 1]. Also for each edge f ∈ Λ, let Df = Ds(f ),
and hence Df1 = Dv1 = [0, 1
2, 1]. Now define
prefixing maps for f1, f2 and f3 by
2] and Dv2 = ( 1
2], Df2 = Dv2 = ( 1
2 , 1] and Df3 = Dv2 = ( 1
τf1(x) = −
τf2(x) = −
1
2
1
2
x +
x +
1
2
1
2
τf3(x) = x
2
1
2(cid:3),
for x ∈ Df1 =(cid:2)0,
, 1(cid:3),
for x ∈ Df2 =(cid:0) 1
for x ∈ Df3 =(cid:0) 1
, 1(cid:3).
2, 1(cid:3). Then the ranges of the prefixing maps are
x∈E⊆Dfi( 1
i = 1, 2,
i = 3
2 µ(E)
µ(E) ,
µ(E)
µ(E) ,
= inf
2
Then Rf1 =(cid:2) 1
4 , 1
2(cid:3), Rf2 =(cid:2)0, 1
4(cid:1) and Rf3 =(cid:0) 1
mutually disjoint and X = Rf1 ∪ Rf2 ∪ Rf3. For each fi, since Lebesgue measure is regular,
the Radon-Nikodym derivative of τfi is given by
Φfi(x) = inf
x∈E⊆Dfi
(µ ◦ τfi)(E)
µ(E)
=(cid:26) 1
2,
1,
i = 1, 2
i = 3.
21
Now define τ 1 : X → X by
τ 1(x) =
τ−1
f1 (x)
τ−1
f2 (x)
τ−1
f3 (x)
for x ∈ Rf1
for x ∈ Rf2
for x ∈ Rf3
Since the sets Rfi are mutually disjoint, τ 1 is well defined on X. Then τ 1 is the coding map
satisfying τ 1 ◦ τfi(x) = x for all x ∈ Dfi.
It is a straightforward calculation to check that {τfi : Dfi → Rfi, i = 1, 2, 3} is a semi-
branching function system for (X, µ). To see that this this Λ-semibranching function system
does not give rise to a monic representation, we argue by contradiction. First, observe that
the only finite paths with range v2 are of the form f3f3 · · · f3; and since τf3(x) = x on
D3 = (1/2, 1], we have
Rf3 = Rf3f3···f3 = (1/2, 1].
Every other finite path λ, having range v1, will satisfy Rλ ⊆ Dv1 = [0, 1/2].
Consequently, R = {Rλ}λ∈Λ does not generate the usual Lebesgue σ-algebra on [0, 1],
even after modification by sets of measure zero, since the restriction of R to (1/2, 1] contains
no nontrivial measurable sets. Theorem 4.5 therefore implies that the representation of
C∗(Λ) associated to this Λ-semibranching function system is not monic, and hence is not
equivalent to any representation on L2(Λ∞, µ) arising from a Λ-projective system.
Remark 4.8. We observe that since monic representations are multiplicity free, it is easy
to construct further examples of non-monic representations by using direct sums of monic
representations, see [2] page 54.
In order to describe the following example of a Λ-semibranching representation which is
monic, we review the concept of a Markov measure (see [19, Section 3.1] or [24, Section 4.2]
for more details, or [5] for Markov measures in a more general context).
Definition 4.9 (Definition 3.1 of [19]). A Markov measure on the infinite path space Λ∞
ON
Λ∞
ON =
∞Yi=1
ZN = {(i1i2 . . . ) : in ∈ ZN , n = 1, 2, . . .}.
of the Cuntz algebra ON is defined by a vector λ = (λ0, . . . , λN−1) and an N × N matrix T
such that λi > 0, Ti,j > 0 for all i, j ∈ ZN , and if e = (1, 1, . . . , 1)t then λT = λ and T e = e.
The Carath´eodory/Kolmogorov extension theorem then implies that there exists a unique
Borel measure µ on Λ∞
ON extending the measure µC defined on cylinder sets by:
µC(Z(I)) := λi1Ti1,i2 · · · Tin−1,in, if I = i1 . . . in.
(22)
The extension µ is called a Markov measure on Λ∞
ON .
For N = 2, given a number x ∈ (0, 1), we can take T = Tx =(cid:18) x
(1 − x)
(1 − x)
x (cid:19), and
λ = (1, 1). The resulting measure will in this case be called µx. Moreover, if x 6= x′, Theorem
3.9 of [19] guarantees that µx, µx′ are mutually singular.
22
Example 4.10. We now consider an example of Λ-semibranching function system which does
give rise to a monic representation. Let Λ be the 2-graph below.
f2
f1
e
v
Recall from Remark 2.4 that every infinite path in a 2-graph can be uniquely written as an
infinite string of composable edges which alternate in color: red, blue, red, . . . . It follows
that the infinite path space of the above 2-graph is homeomorphic to Λ∞
Z2 via the
identification
O2 ∼=Q∞i=1
efj1efj2efj3 · · · 7→ j1j2j3 · · · .
Therefore, the measure µx described above can be viewed as a measure on Λ∞.
It is straightforward to check that, as operators on L2(Λ∞, µx), the prefixing operators
σe, σf1, σf2 have positive Radon-Nikodym derivatives at any point z ∈ Λ∞. Consequently,
the standard prefixing and coding maps make (Λ∞, µx) into a Λ-semibranching function
system. The associated representation of C∗(Λ) is therefore monic, by Theorems 4.2 and
Theorem 2.10, and Example 3.2.
5 A universal representation for non-negative Λ-projective
systems
The focus of this section is the construction of a 'universal representation' of C∗(Λ), general-
izing the work of [19, Section 4] for the Cuntz algebra setting, such that every non-negative
monic representation of C∗(Λ) is a sub-representation of the universal representation. The
Hilbert space H(Λ∞) on which our universal representation is defined is the 'universal space'
for representations of C(Λ∞), see [42], and also [19, 6, 1, 32]. For the case of ON , this space
was also shown to be the 'universal representation space' for monic representations in [19].
We recall the construction of H(Λ∞) below.
Definition 5.1. Let Λ be a finite k-graph with no sources, and let Λ∞ be the infinite path
space of Λ, endowed with the topology generated by the cylinder sets and the Borel σ-algebra
associated to it.
Consider the collection of pairs (f, µ), where µ is a Borel measure on Λ∞, and f ∈
L2(Λ∞, µ). We say that two pairs (f, µ) and (g, ν) are equivalent, denoted by (f, µ) ∼ (g, ν),
if there exists a finite Borel measure m on Λ∞ such that
µ << m, ν << m, and fr dµ
dm
= gr dν
dm
We write f √dµ for the equivalence class of (f, µ).
in L2(Λ∞, m).
23
Proposition 8.3 of [6] establishes that H(Λ∞) is a Hilbert space, with the vector space
structure given by scalar multiplication and
and the inner product given by
d(µ + ν)
fpdµ + g √dν :=(cid:16)fs dµ
hfpdµ, g √dνi :=ZΛ∞
+ gs dν
d(µ + ν)(cid:17)pd(µ + ν),
f g(cid:16)s dµ
d(µ + ν) s dν
d(µ + ν)(cid:17)d(µ + ν).
(23)
We call H(Λ∞) the universal Hilbert space for Λ∞.
The following fundamental property of H(Λ∞) justifies the name 'universal Hilbert space.'
Proposition 5.2. ([32, Theorem 3.1], [19], [1]) Let Λ be a finite k-graph with no sources.
For every finite Borel measure µ on Λ∞, define Wµ : L2(Λ∞, µ) → H(Λ∞) by Wµ(f ) = f√dµ.
Then Wµ is an isometry of L2(Λ∞, µ) onto a subspace of H(Λ∞), which we call L2(µ).
We are now ready to present the universal representation πuniv of C∗(Λ) on H(Λ∞).
Proposition 5.3. Let Λ be a finite k-graph with no sources. Fix (f, µ) ∈ H(Λ∞). For each
λ ∈ Λn, define Suniv ∈ B(H(Λ∞)) by
Suniv
λ
(fpdµ) := (f ◦ σn)qd(µ ◦ σ−1
λ ),
where σλ and σn are the standard prefixing and coding maps of Definition 2.3. Then:
(a) The adjoint of Suniv
λ
is given by (Suniv
λ
)∗(f √dµ) := (f ◦ σλ)pd(µ ◦ σλ).
(b) The operators {Suniv
λ
: λ ∈ Λ} generate a representation πuniv of C∗(Λ) on H(Λ∞),
which we call the 'universal representation'.
(c) The projection valued measure P on Λ∞ given in Definition 3.7 associated to the uni-
versal representation πuniv is given by:
where A is a Borel set of the Borel σ-algebra generated by the cylinder sets.
P (A)(fpdµ) = (χA · f )pdµ,
(24)
Proof. The proof is similar to that of Proposition 4.2 of [19], although the details are more
involved because of the more complicated k-graph structure. To simplify the notation, in
this proof we will drop the superscript univ from Suniv
}λ∈Λ gives a
representation of C∗(Λ), first we observe that the operators Sλ are well defined; in other
λ ) for all
words, if f√dµ = g√dν, we must have (f ◦ σn)qd(µ ◦ σ−1
λ ) = (g ◦ σn)qd(ν ◦ σ−1
. To check that {Suniv
λ
λ
λ ∈ Λ.
24
λ =
λ = σd(λ) on Z(λ) now implies that, if n := d(λ),
is zero off Z(λ), and µ ◦ σ−1
λ (cid:1) on Z(λ). The fact that σ−1
Suppose that f√dµ = g√dν. Observe that µ ◦ σ−1
(cid:0)mZ(s(λ)) ◦ σ−1
(f ◦ σn)qd(µ ◦ σ−1
=(cid:16)g√dν(cid:17)Z(s(λ)) ◦ σ−1
λ ) =(cid:18)(f ◦ σn)qd(µ ◦ σ−1
λ =(cid:18)(g ◦ σn)qd(ν ◦ σ−1
λ
It follows that Sλ is well defined.
λ )(cid:19)Z(λ) =(cid:16)fpdµ(cid:17)Z(s(λ)) ◦ σ−1
λ )(cid:19)Z(λ) = (g ◦ σn)qd(ν ◦ σ−1
λ
λ ).
To check the formula for S∗λ given in the statement of the proposition, we compute:
hS∗λ(fpdµ), g√dνi = hfpdµ, Sλg√dνi = hfpdµ, (g ◦ σn)qd(ν ◦ σ−1
=ZΛ∞
This integral vanishes off Z(λ), since σ−1
λ
fact that (ν ◦ σ−1
λ )Z(λ) = νZ(s(λ)) ◦ σ−1
λ ))s d(ν ◦ σ−1
λ )
d(µ + (ν ◦ σ−1
λ ))
(and consequently d(ν ◦ σ−1
to rewrite
f (x)(g ◦ σn)(x)s
d(µ + (ν ◦ σ−1
dµ
λ
λ )i
d(µ + (ν ◦ σ−1
λ )).
λ )) do. We thus use the
f ◦ σλ(x)g(x)s d(µ ◦ σλ)
hS∗λ(fpdµ), g√dνi =ZZ(s(λ))
Hence S∗λ(f√dµ) = (f ◦ σλ)pd(µ ◦ σλ), which proves (a).
d((µ ◦ σλ) + ν)s
straightforward computation, analogous to the proof of Proposition 3.4.
Checking condition (b), that the operators {Sλ} give a representation of C∗(Λ), is a
To see (c), note that Equation (24) follows from the observation that SνS∗ν acts by
multiplication by χZ(ν); the fact that disjoint unions of cylinder sets Z(ν) generate the σ-
algebra up to sets of measure zero [27, Lemma 4.1] therefore enables us to compute P (A)
by linearity, for any Borel set A.
dν
d((µ ◦ σλ) + ν)
d((µ ◦ σλ) + ν).
The following two Propositions, which detail additional technical properties of the pro-
jection valued measure associated to πuniv, will be used in the proof of Theorem 5.7, the
main result of this Section.
Proposition 5.4. Let Λ be a finite k-graph with no sources and let H(Λ∞) be the Hilbert
space described in Definition 5.1, and let πuniv = {Suniv
: λ ∈ Λ} be the universal represen-
tation of C∗(Λ) on H(Λ∞) given in Proposition 5.3.
(a) For y ∈ H(Λ∞), define a function νy on Λ∞ by
νy(Z(λ)) := h(Suniv
)∗)y, yi,
(Suniv
λ
λ
λ
where h·,·i is the inner product given on H(Λ∞) in Equation (23). Then νy gives a
measure on Λ∞.
(b) Let T be a bounded operator on H(Λ∞). If T commutes with πunivC(Λ∞), then for any
x ∈ H(Λ∞) we have
νT (x) << νx.
25
(c) For every vector f√dµ ∈ H(Λ∞), we have νf√dµ = f2µ.
Proof. As in Equation (14), it is straightforward to see (a). For (b), fix x ∈ H(Λ∞). Then
since T commutes with Suniv
, we have
λ
νT (x)(Z(λ)) = h(Suniv
)∗)x, T ∗T (x)i.
is a partial isometry, the Cauchy-Schwarz inequality then gives
)∗)T (x), T (x)i = h(Suniv
(Suniv
(Suniv
λ
Since each Suniv
λ
λ
λ
λ
νT (x)(Z(λ))2 ≤ k(Suniv
λ
(Suniv
λ
)∗)xk2 kT ∗T xk2 = νx(Z(λ))2 kT ∗T xk2,
which gives that νT (x) << νx.
For (c), Equation (24) implies that, for any cylinder set Z(η),
νf√dµ(Z(η)) = h(χZ(η)f )pdµ, fpdµi =Z χZ(η) · f2dµ =ZZ(η) f2 dµ.
This gives the desired result.
We now present an important result which will allow us to derive, in Theorem 5.7, the
desired universal property of the representation.
Theorem 5.5. Let Λ be a finite k-graph with no sources. Let H(Λ∞) be the universal Hilbert
space for Λ∞ and πuniv be the universal representation of C∗(Λ) on H(Λ∞). Then:
(a) An operator T ∈ B(H(Λ∞)) commutes with πunivC(Λ∞) if and only if for each finite
Borel measure µ on Λ∞ which arises from a monic representation of C∗(Λ) as in
Equation (14), there exists a function Fµ in L∞(Λ∞, µ) such that:
(i) sup{kFµk : µ arises from a monic representation } < ∞.
(ii) If µ << λ then Fµ = Fλ, µ-a.e.
(iii) T (f√dµ) = Fµf√dµ for all f√dµ ∈ H(Λ∞)
(b) Let H denote the subspace of H(Λ∞) spanned by vectors of the form f√dµ where
µ arises from a monic representation. An operator T ∈ B(H(Λ∞)) commutes with
πunivH if and only if for every finite Borel measure µ on Λ∞ arising from a monic
representation of C∗(Λ), and for each λ ∈ Λ, we have
Fµ = Fµ◦σ−1
λ ◦ σλ, µ − a.e.
Proof. Recall from Proposition 5.2 the isometry Wµ of L2(Λ∞, µ) onto L2(µ). Throughout
the proof, we will assume that the finite Borel measure µ arises from a monic representation.
We first claim that if T commutes with πunivC(Λ∞), then T maps L2(µ) into itself. To prove
this, let x = f√dµ be in L2(µ), and let T (x) = g√dζ for (g, ζ) ∈ H(Λ∞). Then Proposition
5.4 (b) implies that νT (x) << νx. By Proposition 5.4 (c), we have
νx = f2µ,
and νT (x) = g2ζ.
26
Therefore g2ζ << µ, so by the Radon–Nikodym theorem there exists h ≥ 0 in L1(Λ∞, µ)
such that g2 dζ = h dµ. Then
and g gpdζ = g √hpdµ.
If g = 0 on some Borel set A, then √h√dµ(A) = 0 also. Therefore,
√dµ ∈ L2(µ),
gpdζ = √hpdµ,
gpdζ =( g√h
g 6= 0
g = 0
g
0,
which shows that T maps L2(µ) into itself.
We now make some computations regarding the relationship between an arbitrary monic
representation π and the universal representation πuniv. Note that Equation (24) implies
that πuniv(ψ)(f√dµπ) = (ψ · f )√dµπ for any f ∈ L2(Λ∞, µπ).
On the other hand, since π is a monic representation, π(χZ(λ))f = χZ(λ) · f ∈ L2(Λ∞, µπ)
by Equation (16). Therefore,
πuniv(ψ)(fpdµπ) = [π(ψ)(f )]pdµπ.
By hypothesis, T commutes with πunivC(Λ∞). Since T preserves L2(µ) for each measure
µ arising from a monic representation, there must exist g ∈ L2(µπ) such that T (f√dµπ) =
g√dµπ. Consequently,
T [π(ψ)f ]pdµπ = T πuniv(ψ)(fpdµπ) = πuniv(ψ)T (fpdµπ)
= πuniv(ψ)(gpdµπ) = [π(ψ)(g)]pdµπ = π(ψ)T(cid:16)fpdµπ(cid:17) ,
so (identifying L2(µπ) ⊆ H(Λ∞) with L2(Λ∞, µπ)) we see that T commutes with π(ψ) for
all ψ ∈ C(Λ∞).
the multiplication operators {Mf : f ∈ C(Λ∞)}. The fact (cf. [31]) that the maximal abelian
subalgebra of B(L2(Λ∞, µ)), for any finite Borel measure µ, is the sub-algebra L∞(Λ∞, µ)
too. In other words, there exists a function Fµ in L∞(Λ∞, µ) such that
Therefore, we can pull-back T to an operator eT on L2(Λ∞, µ) that commutes with all of
consisting of multiplication operators now implies that eT must be a multiplication operator
(25)
T (fpdµ) = Fµ fpdµ
kFµkL∞(µ) ≤ kTk
for all f ∈ L2(Λ∞, µ), establishing (iii ). It remains to check the properties of the functions
Fµ. One immediately observes that
and this implies (i ). To check (ii ), suppose µ << λ. Then, for all f ∈ L2(Λ∞, µ), we have
f√dµ = fpdµ/dλ√dλ, and hence
T (fpdµ) = T frdµ
dλ
√dλ! =⇒ Fµfpdµ = Fλfrdµ
dλ
√dλ
27
Thus, as elements of L2(λ), Fµfpdµ/dλ = Fλfpdµ/dλ for any f ∈ L2(Λ∞, µ), which
implies Fµdµ/dλ = Fλdµ/dλ (λ − a.e.). It follows that, for any Borel set A,
ZA
(Fµ − Fλ) dµ =ZA
(Fµ − Fλ)
dµ
dλ
dλ = 0,
so Fµ = Fλ, µ-a.e. This proves (ii ).
For the converse, assume that T is given on L2(µ) by a function Fµ ∈ L∞(Λ∞, µ) satisfying
i.e. T (f√dµ) = Fµf√dµ for all f√dµ ∈ H(Λ∞) such that µ arises from a
(i ), (ii ), (iii ),
monic representation. Then (i) implies that T is bounded with kTk ≤ supµ{kFµkL∞(µ)}.
Since T acts as a multiplication operator on each L2(µ), Part (c) of Proposition 5.3 implies
that T commutes with P (A) for all Borel subsets A and therefore T commutes with the
restricted universal representation, πunivC(Λ∞), which proves (a).
To prove (b), note that if an operator T ∈ B(H(Λ∞)) commutes with the universal
representation πuniv of C∗(Λ) on H, then in particular T commutes with πunivC(Λ∞) on
H, and hence T (f√dµ) = Fµf√dµ is a multiplication operator on each L2(µ) when the
measure µ arises from a monic representation. In particular, T is normal (when restricted to
H). Therefore, by the Fuglede–Putnam theorem, TH commutes with πuniv iff T Suniv
H =
Suniv
λ TH for all λ ∈ Λ. Using the formulas for Suniv
from Theorem 5.5, we see that TH
commutes with πunivH if and only if, for each f√dµ ∈ H and λ ∈ Λn,
λ = (Fµ ◦ σn)(f ◦ σn)qdµ ◦ σ−1
Fµ◦σ−1
or equivalently, Fµ◦σ−1
from monic representations of C∗(Λ). Composing with σλ gives the desired result of (b).
Definition 5.6. A monic representation {tλ}λ∈Λ of a finite, source-free k-graph Λ is said
to be nonnegative if the functions {fλ}λ∈Λ of the associated Λ-projective system on Λ∞ are
nonnegative a.e.
(f ◦ σn)qdµ ◦ σ−1
= (Fµ ◦ σn) for λ ∈ Λn, (µ ◦ σ−1
λ ) − a.e. for all measures µ arising
λ ,
λ
λ
λ
λ
The following result, a consequence of Theorem 5.5, proves that every nonnegative monic
}λ∈Λ, justifying the name 'uni-
representation is equivalent to a sub-representation of {Suniv
versal representation' for {Suniv
Theorem 5.7. Let Λ be a finite k-graph with no sources. Let {tλ}λ∈Λ be a nonnegative
monic representation of C∗(Λ) on L2(Λ∞, µπ). Let W be the isometry from L2(Λ∞, µπ) onto
L2(µπ) given in Proposition 5.2, so that W f = f√dµπ. Then W intertwines {tλ}λ∈Λ with
the sub-representation {Suniv
}λ.
Proof. By Theorem 4.2 and Proposition 3.4, we can assume that tλ is of the form
L2(µπ )}λ∈Λ of the universal representation {Suniv
}λ∈Λ.
λ
λ
λ
λ
tλ(f ) = fλ · (f ◦ σd(λ)),
rem 5.5,
where, since {tλ}λ∈Λ is assumed nonnegative, we may assume fλ =q d(µπ◦(σλ)−1)
W (tλf ) = W (fλ(f ◦ σd(λ))) = fλ(f ◦ σd(λ))pdµπ = (f ◦ σd(λ))pfλ2dµπ
λ W (f ).
dµπ
= (f ◦ σd(λ))pd[µπ ◦ (σλ)−1] = Suniv
λ
(fpdµπ) = Suniv
L2(µπ )}λ∈Λ, as claimed.
In other words, W intertwines {tλ}λ∈Λ and {Suniv
λ
. By Theo-
28
References
[1] D. Alpay, P.E.T. Jorgensen, and I. Lewkowicz, Markov measures, transfer operators,
wavelets and multiresolutions. arXiv:1606.07692.
[2] W. Arveson, An invitation to C∗-algebras. Graduate Texts in Mathematics, No. 39.
Springer-Verlag, New York-Heidelberg, 1976. x+106 pp.
[3] S. Bezuglyi, J. Kwiatkowski and K. Medynets, Aperiodic substitution systems and their
Bratteli diagrams. Ergodic Theory Dynam. Systems 29 (2009), 37–72.
[4] S. Bezuglyi, J. Kwiatkowski, K. Medynets, and B. Solomyak, Finite rank Bratteli di-
agrams: structure of invariant measures. Trans. Amer. Math. Soc. 365 (2013), 2637–
2679.
[5] S. Bezuglyi and P.E.T. Jorgensen, Representations of Cuntz-Krieger relations, dynamics
on Bratteli diagrams, and path-space measures. Trends in harmonic analysis and its
applications, Contemp. Math. 650 (2015), Amer. Math. Soc., Providence, RI, 57–88.
[6] S. Bezuglyi and P.E.T. Jorgensen, Infinite-dimensional transfer operators, endomor-
phisms, and measurable partitions. arXiv:1702.02657.
[7] O. Bratteli and P.E.T. Jorgensen, Iterated function systems and permutation represen-
tations of the Cuntz algebra. Mem. Amer. Math. Soc. 139 (1999). x+89 pp.
[8] O. Bratteli, P.E.T. Jorgensen, and J. Price, Endomorphisms of B(H). Quantization,
nonlinear partial differential equations, and operator algebra, Proc. Sympos. Pure Math.
59 (Cambridge, MA, 1994), Amer. Math. Soc., Providence, RI, 93–138.
[9] T.M. Carlsen, S. Kang, J. Shotwell and A. Sims, The primitive ideals of the Cuntz-
Krieger algebra of a row-finite higher-rank graph with no sources. J. Funct. Anal. 266
(2014), 2570–2589.
[10] T.M. Carlsen, E. Ruiz, and A. Sims, Equivalence and stable isomorphism of groupoids,
and diagonal-preserving stable isomorphisms of graph C∗-algebras and Leavitt path al-
gebras. Proc. Amer. Math. Soc. 145 (2017), 1581–1592.
[11] J. Cuntz, Simple C∗-algebras generated by isometries. Comm. Math. Phys. 57 (1977),
173–185.
[12] J. Cuntz and W. Krieger, A class of C∗-algebras and topological Markov chains. Invent.
Math. 56 (1980), 251–268.
[13] K.R. Davidson, S.C. Power and D. Yang, Atomic representations of rank 2 graph alge-
bras. J. Funct. Anal. 255 (2008), 819–853.
[14] K.R. Davidson and D. Yang, Representations of higher rank graph algebras. New York
J. Math. 15 (2009), 169–198.
29
[15] J. Dixmier, Les C∗-alg`ebres et leurs repr´esentations. Cahiers Scientifiques, Fasc. XXIX.
´Editions Jacques Gabay, Paris, 1964. xi+382 pp.
[16] R. Durrett, Probability: Theory And Examples, second edition. Cambridge University
Press, Cambridge, MA, 1996. 440 pp.
[17] D.E. Dutkay and P.E.T. Jorgensen, Martingales, endomorphisms, and covariant systems
of operators in Hilbert space. J. Operator Theory 58 (2007), 269–310.
[18] D.E. Dutkay and P.E.T. Jorgensen, Fourier series on fractals: a parallel with wavelet
theory. Radon Transforms, Geometry, and Wavelets, Contemp. Math 464 (2008), Amer.
Math. Soc., Providence, RI, 75–101.
[19] D.E. Dutkay and P.E.T. Jorgensen, Monic representations of the Cuntz algebra and
Markov measures. J. Funct. Anal. 267 (2014), 1011–1034.
[20] D.E. Dutkay, J. Haussermann, and P.E.T. Jorgensen, Atomic representations of Cuntz
algebras. J. Math. Anal. Appl. 421 (2015), 215–243.
[21] E. Effros, The Borel space of von Neumann algebras on a separable Hilbert space. Pacific
J. Math. 15 (1965), 1153–1164.
[22] E. Effros, Transformation groups and C∗-algebras. Ann. of Math. (2) 81 (1965), 38–55.
[23] M. Enomoto and Y. Watatani, A graph theory for C∗-algebras. Math. Japon. 25 (1980),
435–442.
[24] C. Farsi, E. Gillaspy, P.E.T. Jorgensen, S. Kang, and J. Packer, Representations
of higher-rank graph C∗-algebras associated to Λ-semibranching function systems.
arXiv:1803.08779.
[25] C. Farsi, E. Gillaspy, A. Julien, S. Kang, and J. Packer, Wavelets and spectral triples
for fractal representations of Cuntz algebras. Contemp. Math. 687 (2017), Amer. Math.
Soc., Providence, RI, 103–133.
[26] C. Farsi, E. Gillaspy, A. Julien, S. Kang, and J. Packer, Spectral tripes and wavelets for
higher-rank graphs, arXiv:1803.09304.
[27] C. Farsi, E. Gillaspy, S. Kang, and J. Packer, Separable representations, KMS states,
and wavelets for higher-rank graphs. J. Math. Anal. Appl. 434 (2015), 241–270.
[28] C. Farsi, E. Gillaspy, S. Kang, and J. Packer, Wavelets and graph C∗-algebras. Excur-
sions in Harmonic Analysis 5 (2016), The February Fourier Talks at the Norbert Wiener
Center. Edited by R. Balan, M. Begu´e, J. J. Benedetto, W. Czaja and K. A. Okoudjou.
[29] J. Glimm, Locally compact transformation groups. Trans. Amer. Math. Soc. 101 (1961),
124–138.
[30] J. Glimm, Families of induced representations. Pacific J. Math. 12 (1962), 885–911.
30
[31] V. Jones, Von Neumann algebras in mathematics and physics. Introduction to Modern
Mathematics, 285–321, Adv. Lect. Math. (ALM) 33, Int. Press, Somerville, MA, 2015.
[32] P.E.T. Jorgensen, Iterated function systems, representations, and Hilbert space. Inter-
nat. J. Math. 15 (2004), no. 8, 813–832.
[33] S. Kakutani, On equivalence of infinite product measures. Annals of Math. 49 (1948),
214–224.
[34] S. Kang and D. Pask, Aperiodicity and primitive ideals of row-finite k-graphs. Internat.
J. Math. 25 (2014), 1450022, 25 pp.
[35] K. Kawamura, Pure states on Cuntz algebras arising from geometric progressions. Al-
gebr. Represent. Theory 19 (2016), 1297–1319.
[36] A.N. Kolmogorov, Foundations of the theory of probability. Chelsea Publishing Com-
pany, New York, NY, 1950.
[37] A. Kumjian and D. Pask, Higher-rank graph C∗-algebras. New York J. Math. 6 (2000),
1–20.
[38] A. Kumjian, D. Pask and I. Raeburn, Cuntz-Krieger algebras of directed graphs. Pacific
J. Math. 184 (1998), 161–174.
[39] M. Laca, Endomorphisms of B(H) and Cuntz algebras. J. Operator Theory 30 (1993),
85–108.
[40] M. Marcolli and A.M. Paolucci, Cuntz-Krieger algebras and wavelets on fractals. Com-
plex Anal. Oper. Theory 5 (2011), 41–81.
[41] C.C. Moore, Group extensions and cohomology for locally compact groups. III. Trans.
Amer. Math. Soc. 221 (1976), 1–33.
[42] E. Nelson, Topics in Dynamics. I: Flows. Math. Notes, Princeton University Press,
Princeton, NJ, 1969.
[43] D. Pask, I. Raeburn, and N. Weaver, A family of 2-graphs arising from two-dimensional
subshifts. Ergodic Theory Dynam. Systems 29 (2009), 1613–1639.
[44] D. Pask, A. Sierakowski, and A. Sims, Twisted k-graph algebras associated to Bratteli
diagrams. Integral Equations Operator Theory 81 (2015), 375–408.
[45] G. Robertson and T. Steger, C∗-algebras arising from group actions on the boundary of
a triangle building, Proc. London Math. Soc. 72 (1996), 613–637.
[46] G. Robertson and T. Steger, Affine buildings, tiling systems and higher rank Cuntz-
Krieger algebras, J. Reine Angew. Math. 513 (1999), 115–144.
[47] A. Skalski and J. Zacharias, Entropy of shifts on higher-rank graph C∗-algebras. Houston
J. Math. 34 (2008), no. 1, 269–282.
31
[48] R.S. Strichartz, Besicovitch meets Wiener–Fourier expansions and fractal measures.
Bull. Amer. Math. Soc. (N.S.) 20 (1989), 54–59.
[49] R. Tumulka, A Kolmogorov extension theorem for POVMs. Letters in Math. Phys. 84
(2008), 4–46.
[50] D. Yang, Endomorphisms and modular theory of 2-graph C∗-algebras. Indiana Univ.
Math. J. 59 (2010), 495–520.
Carla Farsi, Judith Packer : Department of Mathematics, University of Col-
orado at Boulder, Boulder, CO 80309-0395, USA.
E-mail address: [email protected], [email protected]
Elizabeth Gillaspy : Department of Mathematics, University of Montana, 32
Campus Drive #0864, Missoula, MT 59812-0864, USA.
E-mail address: [email protected]
Palle Jorgensen : Department of Mathematics, 14 MLH, University of Iowa,
Iowa City, IA 52242-1419, USA.
E-mail address: [email protected]
Sooran Kang : College of General Education, Chung-Ang University, 84 Heukseok-
ro, Dongjak-gu, Seoul, Republic of Korea.
E-mail address, [email protected]
32
|
1005.1305 | 1 | 1005 | 2010-05-07T22:29:21 | Spectra self-similarity for almost Mathieu operators | [
"math.OA",
"math-ph",
"math.FA",
"math-ph"
] | We determine numerically the self-similarity maps for spectra of the almost Mathieu operators, a two-dimensional fractal-like structure known as the Hofstadter butterfly. The similarity maps each have a horizontal component determined by certain algebraic maps, and vertical component determined by a Mobius transformation, indexed by a semigroup of the matrix group $GL_2(\Z)$. Based on the numerical evidence, we state and prove a continuity result for the similarity maps. We note a connection between the indexing of the similarity maps and Morita equivalence of rotation algebras $A_\theta$, a continuous field of C*-algebras. | math.OA | math | SPECTRA SELF-SIMILARITY FOR ALMOST MATHIEU
OPERATORS
MICHAEL P. LAMOUREUX, JAMES A. MINGO, AND SYDNEY R. PACHMANN
Abstract. We determine numerically the self-similarity maps for spectra of the
almost Mathieu operators, a two-dimensional fractal-like structure known as the
Hofstadter butterfly. The similarity maps each have a horizontal component deter-
mined by certain algebraic maps, and vertical component determined by a Mobius
transformation, indexed by a semigroup of the matrix group GL2(Z). Based on the
numerical evidence, we state and prove a continuity result for the similarity maps.
We note a connection between the indexing of the similarity maps and Morita
equivalence of rotation algebras Aθ, a continuous field of C*algebras.
0
1
0
2
y
a
M
7
]
.
A
O
h
t
a
m
[
1
v
5
0
3
1
.
5
0
0
1
:
v
i
X
r
a
Contents
Introduction
1.
2. Numerical evidence: Similarity maps
3. Numerical evidence: Vertical similarities from GL2(Z)
4. Numerical evidence: A symmetry with break
5. Generating the GL2(Z) symmetries
6. Horizontal similarity: Interval maps
7. Horizontal similarity: Cubic case 1 (cid:55)→ 1/3
8. Horizontal similarity: Quintic case 1/3 (cid:55)→ 1/5
9. Horizontal similarity: Algebraic curves
10. The similarity maps: General case
11. The similarity maps: Gap labelling
12. The similarity maps: Proof of continuity
13. Three generators for the butterfly similarities
14. Conclusions
Acknowledgments
References
Appendix 1
Appendix 2
Appendix 3
Date: May 7, 2010.
1
2
4
6
10
11
12
14
16
17
20
22
23
28
29
29
30
31
33
34
2
MICHAEL P. LAMOUREUX, JAMES A. MINGO, AND SYDNEY R. PACHMANN
Figure 1. The Hofstadter butterfly, a fractal-like structure of line spectra.
1. Introduction
Looking at the image presented in Figure 1, one immediately notices the striking,
repetitive pattern of "butterfly wings" that march off towards the vertical horizons at
the top and bottom of the image. This rendering of the so-called Hofstadter butterfly,
drawn to high-resolution using a combination of numerical algorithms and PostScript
graphics programming, reveals some beautiful symmetries of a fundamentally math-
ematical object. The goal of this paper is to specify exactly the symmetries of this
image and prove continuity results of the corresponding similarity maps, motivated
by numerical evidence collected in our study of the butterfly.
SPECTRA SELF-SIMILARITY FOR ALMOST MATHIEU OPERATORS
3
Although the Hofstadter butterfly looks like a fractal, the image is not created
using the usual iterative equations or recursive graphical methods designed to visually
render typical fractals (Mandelbrot, 1990). Rather, this image is created from an
explicit numerical computation of spectral values of a family of linear operators on
Hilbert space.
More precisely, this image is constructed as a layered suite of horizontal lines
which arise as the spectra of almost Mathieu operators. Each of these operators is
succinctly represented as the self-adjoint element
(1)
hθ = u + u∗ + v + v∗
in the C*-algebra Aθ generated by universal unitaries u, v satisfying the commutation
rule
(2)
In Figure 1, the vertical axis is spanned by the parameter θ, with 0 ≤ θ ≤ 1, while
the horizontal axis corresponds to spectral values x in the range −4 ≤ x ≤ 4.
vu = e2πiθuv.
Only rational values of θ are used in the construction of the image in Figure 1. For
this reason, the image is properly called the rational Hofstadter butterfly. Extensions
to irrational values of θ has been a theme in the long history of study these operators
and the physical problems that motivated it, going back at least to a mathematical
analysis of Bloch electrons (Brown, 1964). A selection of relevant studies over the
years is given in the references, including (Avila and Jitomirskaya, 2006; Bellissard
and Simon, 1982; Bellissard, 1990; Choi et al., 1990; Connes, 1994; Goldman, 2009;
Hofstadter, 1976; Kaminker and Putnam, 2003; Last, 1994; Puig, 2004; Ypma, 2007).
Unlike the irrational case as considered in (Arveson, 1994), rational values of
θ = p/q lead to the computation of spectral lines based on finite dimensional eigen-
value calculations using tridiagonal q × q matrices (Lamoureux, 1997). In particular,
the endpoints of the spectral lines are specified by eigenvalues of these matrices.
Numerical algorithms for the tridiagonal eigenvalue computations are rapid and ac-
curate, making calculations of the rational butterfly ideal for numerical experiments.
Based on these experiments, we deduce a systematic catalogue of the symmetries of
the butterfly.
Repetition of a geometric object fading out to infinity is characteristic of hyperbolic
geometry (Coxeter, 1942), hence it is perhaps no surprise that the matrix group
GL2(Z) arises in the observed symmetries. Sections 2 through 5 reveal a semigroup
of linear fractional transformations represented by elements of GL2(Z) which act on
the vertical parameter θ to generate the family of similarity maps.
Equally important are the algebraic curves specifying the horizontal component of
the similarities, acting on the horizontal parameter x. Sections 6 through 9 present
numerical evidence of the continuous maps taking one level of horizontal spectra
4
MICHAEL P. LAMOUREUX, JAMES A. MINGO, AND SYDNEY R. PACHMANN
to another, using an indexing of intervals and a correspondence of image points
under polynomial maps. These are precisely the characteristic polynomials of the
aforementioned finite dimensional q × q matrices giving the spectra.
Uniting the vertical and horizontal components leads to an identification of the
general form of the similarity maps, as presented in Section 10. The similarity extends
to gap-labelling, a specific method for indexing the butterfly wings based on Chern
characters, which is discussed in Section 11. Proof of continuity of the similarity maps
is given in Section 12 and generators for the semigroup of similarities is presented in
Section 13.
"Xenocides, who is ugly, makes ugly poetry," said Aristophanes. With the Hofs-
tadter butterfly, we see pretty mathematics making pretty pictures.
2. Numerical evidence: Similarity maps
Mirror image symmetry in the horizontal direction suggests an obvious self-similarity
of the image in Figure 1 given by a reflection about the line x = 0, namely:
(3)
(x, θ) (cid:55)→ (−x, θ).
This map is continuous, and does map the butterfly to itself, as the spectrum of each
hθ is symmetric.
In the vertical direction, another mirror image symmetry in the figure suggests a
second self-similarity map given by reflection about the line θ = 1/2, namely:
(4)
(x, θ) (cid:55)→ (x, 1 − θ).
Again, this is a continuous map, and again maps spectra properly since the algebra Aθ
is isomorphic to A1−θ, with operator hθ mapping onto h1−θ under the isomorphism.
Now, a more interesting symmetry is observed mapping the large central butterfly
onto the next largest butterfly in the bottom half of the image in Figure 1. Zooming
in on this butterfly, as shown in Figure 2, one sees the top of the butterfly at θ = 1/3,
the bottom at θ = 0, and the centre of the butterfly at θ = 1/4. This suggest we
must find a self-similarity map that, on vertical parameter values θ, will map θ (cid:55)→ θ(cid:48)
as
(5)
0 (cid:55)→ 0,
1/2 (cid:55)→ 1/4,
1 (cid:55)→ 1/3.
No linear map will do, but a linear fractional transformation1 does work, using the
map
(6)
1i.e. a Mobius transformation
θ (cid:55)→ θ(cid:48) =
θ
2θ + 1
.
SPECTRA SELF-SIMILARITY FOR ALMOST MATHIEU OPERATORS
5
Figure 2. Similarity to the bottom third of the butterfly.
It is reassuring to notice that other obvious lines in the full image map correctly
under this map. For instance, the line at θ = 1/3 in the full butterfly in Figure 1
should map to the top of the second largest butterfly in Figure 2, which is at θ(cid:48) = 1/5.
And indeed, the LFT here does just that. Similarly we can check for that the line
at 1/4 maps to 1/6, the line at 1/5 maps to 1/7, and so on. Checking numerically
many of these lines assures us that the linear fractional transformation is the proper
choice.
Going to a numerical experiment based on these observations, we construct a map
from the full image in Figure 1 to subset like Figure 2, using the LFT in the vertical
component, and contracting linearly in the horizontal component. Specifically, we
map according to the rule
(7)
(x, θ) (cid:55)→ (x(cid:48), θ(cid:48)) =
(1 − .82 ∗ θ)x,
θ
2θ + 1
(cid:18)
(cid:19)
.
The coefficient 0.82 was chosen to get the correct width for the top horizontal line.
The result is the image shown in Figure 3, which is very much like the lower butterfly
image in Figure 2.
Observing the slight differences between the image in Figure 1 and the result of
the numerical experiment shown in Figure 3, (the difference in the curvature of the
- 4401/31/41/51/61/7 1 0 2 16
MICHAEL P. LAMOUREUX, JAMES A. MINGO, AND SYDNEY R. PACHMANN
Figure 3. A rendering of an approximate similarity to the bottom
third, using a linear fractional transformation vertically, linear scaling
horizontally.
wingtips for instance), we can conclude the whatever the similarity map is, it is
only approximately linear. Nevertheless, the basic structure and position of the tiny
butterflies are preserved. This serves as numerical evidence that the vertical maps
are indeed given by linear fractional transformations. The horizontal component is
more complicated, and is discussed in Sections 7, 8, 9.
3. Numerical evidence: Vertical similarities from GL2(Z)
Perhaps it is suggestive that two of the self-similarity maps identified so far (verti-
cal flip, and the bottom third map) have vertical components give by Mobius trans-
formations,
(8)
Both transformations are specified by a 2 × 2 matrix in GL2(Z), in the form
2θ + 1
1
θ (cid:55)→ 1 − θ
and θ (cid:55)→ θ
.
(9)
M =
(cid:20) a b
c d
(cid:21)
,
yielding a corresponding linear fractional transformation
(10)
θ (cid:55)→ θ(cid:48) =
aθ + b
cθ + d
.
It will be convenient to label these maps by the corresponding matrix, and we note
that composition of the LFTs corresponds to matrix multiplication in GL2(Z).
A leap of faith suggests looking for symmetries in the butterfly indexed by elements
of GL2(Z). Examining Figure 4, we see the bottom of the central core of butterflies,
and we can compute a list of matrices that implement the self-similarity map on the
SPECTRA SELF-SIMILARITY FOR ALMOST MATHIEU OPERATORS
7
Figure 4. A sequence of similarities in the lower central core of the
butterfly, and matrices in SL2(Z) that implement the vertical compo-
nent of the similarity.
vertical component. For instance, we can map the whole butterfly onto the central
bottom butterfly that tops out at label θ = 1/5. For this, we need an LFT that
maps 0 (cid:55)→ 0, 1/2 (cid:55)→ 1/6, 1 (cid:55)→ 1/5. Solving for coefficients a, b, c, d in the LFT, we
find the matrix
(cid:21)
(cid:20) 1 0
4 1
(11)
(12)
will work to implement an LFT that maps the whole figure into this central bottom
butterfly.
Continuing down the central core, we see an infinite sequence of butterflies extend-
ing to the "horizon" at θ = 0. Some simple calculations analogous to those above
produce a sequence of similarity maps indexed by matrices of the form
= A2n, for n ≥ 1.
(cid:20) 1
0
2n 1
=
(cid:21)
(cid:21)2n
(cid:20) 1 0
1 1
We will see later that this matrix A is a key generator of the similarities.
Hopping to the top of butterfly, in Figure 5 we see a central core of butterflies
extending to the upper "horizon" at level θ = 1. Here, the self-similarity maps are
- 44010211041106110811/31/41/51/61/71/81/98
MICHAEL P. LAMOUREUX, JAMES A. MINGO, AND SYDNEY R. PACHMANN
Figure 5. A sequence of similarities in the upper central core of the
butterfly, and matrices in GL2(Z) that implement the vertical compo-
nent of the similarity.
obtained using matrices in GL2(Z) of the form
(cid:20) 1 − 2n 2n
−2n
2n + 1
(cid:21)
, for n ≥ 1.
Matrices from symmetries on top (Eqn 13) are related to matrices for the symme-
tries on the bottom (Eqn 12), via the conjugation
(cid:20) 1 − 2n 1
−2n
2n + 1
=
(cid:21)
(cid:20) −1 1
(cid:21)(cid:20) 1
(cid:21)
(cid:20) −1 1
0 1
(cid:21)(cid:20) −1 1
(cid:21)
.
0 1
0
2n 1
This is just conjugation with the flip symmetry map of the butterfly, as the matrix
(15)
induces the linear fractional transformation mapping θ to 1− θ, turning the butterfly
upside down.
B =
0 1
Attending to some of the butterflies on the side of the image, we find the relevant
linear fractional transformations are represented by matrices of the form
(13)
(14)
(16)
(cid:20)
(cid:21)
0 1−1 n
, for n ≤ 3,
1 -1 2 -2 3 -3 4 -4 5 -5 6 -6 7 -7 8 -8 92/33/44/55/66/77/8SPECTRA SELF-SIMILARITY FOR ALMOST MATHIEU OPERATORS
9
Figure 6. A sequence of similarities on the edge of the butterfly, and
matrices in GL2(Z) that implement the similarity.
which generate the symmetries on the bottom half of the image shown in Figure 6.
On the top of the image, we obtain transformations indexed by matrices of the form
(cid:21)
1 n − 1
(cid:20) 1 n − 2
(cid:20) −1 1
(cid:21)
=
0 1
(cid:21)(cid:20)
(cid:20) 1 n − 2
1 n − 1
(17)
(18)
Again, we observe there is a relation between these two forms of transformation,
through conjugation with the flip, as we have
, for n ≥ 3.
(cid:21)(cid:20) −1 1
(cid:21)
.
0 1
0 1−1 n
1 1 1 2 1 2 1 3 1 3 1 4 1 4 1 5 0 1-13 0 1-14 0 1-15 0 1-161/61/51/41/31/22/33/44/55/610
MICHAEL P. LAMOUREUX, JAMES A. MINGO, AND SYDNEY R. PACHMANN
Note, however, that these self-similarity maps have a certain discontinuity.
In
particular, the centre of the large butterfly gets mapped to a "broken" butterfly on
these side maps. This tells us the self-symmetry maps need not be continuous. In
particular there may be a discontinuity in the horizontal direction as we cross the
x = 0 spectral value.
Nevertheless, this is a mild discontinuity and is easy to understand. In the con-
struction of the Hofstadter butterfly, for θ = p/q with q even, the spectrum consists
of q intervals on the real line, two of which touch at the point zero. (Choi et al., 1990)
It is these "touching intervals" that are getting broken apart in the above similarity
maps. So although there is an apparent discontinuity, it is perhaps better to describe
it as the breaking apart of two touching spectral lines. The similarity map should
take the single, common endpoint, and map it to two distinct endpoints of two non-
overlapping intervals. The details of this double-valued map will be discussed in
Section 10.
4. Numerical evidence: A symmetry with break
As noted in the previous section, under the similarity map sometimes the butterfly
breaks. This suggests we can look for more similarities if we are a bit more open to
what a broken butterfly looks like.
In Figure 7, we have an example of a really broken butterfly. The butterfly fits
and the similarity map is given by the linear fractional transformation with matrix
into the region
(19)
(20)
0 ≤ θ ≤ 1/2,
(cid:20) 1 0
1 1
(cid:21)
.
(cid:20) 1 0
(cid:21)
This break (at θ = 1/3) looks pretty bad, but in fact we know from gap labelling
theorems that when one plots the gaps an inverse slope 2, there is a discontinuity
exactly at θ = 1/3 (Lamoureux, 2010). So in fact this is just confirming the fact
that the butterflies are coming from gap labelling.
By including this symmetry, we include the matrix
A =
(21)
in our collection of matrices in GL2(Z) generating similarity maps.
We observe that all the LFTs seen so far map the θ interval [0, 1] into itself, and
are indexed by certain elements of of GL2(Z). This evidence suggests a certain
semigroup in GL2(Z) specifies the possible self-similarity maps. A precise statement
identifying the semigroup is given in the next section.
1 1
SPECTRA SELF-SIMILARITY FOR ALMOST MATHIEU OPERATORS
11
Figure 7. An unusual, cracked similarity. Note the basic butterfly
structure spanning the bottom half of the full image.
5. Generating the GL2(Z) symmetries
The following technical result states that the set of linear fractional transforma-
tions which map the interval [0, 1] into itself forms a semigroup, indexed by a 2-
generated semigroup of matrices in GL2(Z).
Theorem 1. Let G be the semigroup of linear fractional transformations of the form
(22)
θ (cid:55)→ θ(cid:48) =
aθ + b
cθ + d
,
for some
∈ GL2(Z)/{±I},
which map the interval [0, 1] into itself. Then G is generated by the two maps φ(θ) =
θ/(θ + 1) and ψ(θ) = 1 − θ, represented by matrices
c d
(cid:21)
(cid:20) a b
(cid:20) −1 1
0 1
(cid:21)
.
(cid:21)
(cid:20) 1 0
1 1
(23)
A =
, and B =
The proof of this technical result is given in Appendix 1, and is based on the
Euclidean algoritm. We make a few observations about this theorem.
refers to the quotient group GL2(Z)/{±I}.
The LFTs do not notice a change of sign in the representing matrix, so the theorem
- 440 1 0 1 11/21/31/412
MICHAEL P. LAMOUREUX, JAMES A. MINGO, AND SYDNEY R. PACHMANN
Figure 8. Multiple similarities, indexed by the same matrix in SL2(Z).
These two generators correspond to two symmetries for the Hofstadter butterfly.
The first generates the map that takes the butterfly to its lower half, mentioned in
Section 4. The second is the vertical flip symmetry, mentioned in Section 2. By
including the vertical flip in our symmetries, we are able to generate all (vertical
components) of the symmetries using only two generating maps.
The fact that the Euclidean algorithm is involved suggests that determining the
details of a symmetry map will be rather involved -- the factorization into generators
is not simple.
It is worth mentioning a deep result from C*-algebras, that states two rotation
algebras Aθ, Aθ(cid:48) are Morita equivalent if and only if there is a linear fractional trans-
formation mapping
(24)
(cid:20) a b
c d
θ (cid:55)→ θ(cid:48) =
aθ + b
cθ + d
,
(cid:21)
for some matrix
in GL2(Z) (Effros and Shen, 1980). We have no idea what
connection this might have with the self-similarity maps we have above, which act
on spectra of operators in the algebras Aθ, Aθ(cid:48), not on the algebras themselves.
6. Horizontal similarity: Interval maps
Considering now the horizontal component of the similarity maps, a close exami-
nation of the apparent symmetries suggests the intervals I1, I2, . . . , Iq in the spectrum
at level θ = p/q are mapped onto a subcollection of the intervals I(cid:48)
1, I(cid:48)
q(cid:48) in the
spectrum at level θ(cid:48) = p(cid:48)/q(cid:48). We note here which subcollections appear.
2, . . . I(cid:48)
2/53/81/32/53/81/32/53/81/3 1 1 2 3 1 1 2 3 1 1 2 3SPECTRA SELF-SIMILARITY FOR ALMOST MATHIEU OPERATORS
13
Examining Figure 8, notice for the single vertical map with linear fractional trans-
formation given by matrix
(25)
(cid:20) 1 1
2 3
(cid:21)
,
there are in fact three choices of horizontal maps. Start first at initial θ = 1, which
maps to θ(cid:48) = 2/5 under this LFT. The spectrum at level θ = 1 has one interval,
while at level θ(cid:48) = 2/5, there are five intervals that we can map to. Only three are
obtained. In the left part of the image in Figure 8, the initial interval I1 maps onto
interval I(cid:48)
3, and the right image maps onto
interval I(cid:48)
(26)
1; the middle image maps onto interval I(cid:48)
5. Writing p(cid:48)/q(cid:48) = 2/5, we see that the three interval maps are given as
I1 (cid:55)→ I(cid:48)
I1 (cid:55)→ I(cid:48)
I1 (cid:55)→ I(cid:48)
5 = I(cid:48)
1 = I(cid:48)
Checking now at the level θ = 1/2 mapping to θ(cid:48) = 3/8, the two intervals I1, I2 at
5 in the middle;
8 on the right. In this case, we have p(cid:48)/q(cid:48) = 3/8 and we can summarize the
2 on the left; to the two interval I(cid:48)
θ = 1/2 map to two interval I(cid:48)
and I(cid:48)
left, middle and right maps as
1, I(cid:48)
7, I(cid:48)
4, I(cid:48)
3 = I(cid:48)
1p(cid:48)+1,
0p(cid:48)+1,
2p(cid:48)+1.
(27)
Ik (cid:55)→ I(cid:48)
0p(cid:48)+k,
Ik (cid:55)→ I(cid:48)
1p(cid:48)+k,
Ik (cid:55)→ I(cid:48)
2p(cid:48)+k,
for k = 1, 2. Similarly, at level θ = 0, the single interval I1 maps as
5 = I(cid:48)
(28)
where p(cid:48)/q(cid:48) = 1/3. The form of the interval map is very consistent.
I1 (cid:55)→ I(cid:48)
I1 (cid:55)→ I(cid:48)
I1 (cid:55)→ I(cid:48)
3 = I(cid:48)
1 = I(cid:48)
0p(cid:48)+1,
1p(cid:48)+1,
2p(cid:48)+1,
We go back and check the other similarity maps. In Figure 4, the interval maps
are apparently of the form
(29)
for the vertical similarity given by
(30)
Ik (cid:55)→ I(cid:48)
n·p+k = I(cid:48)
n·p(cid:48)+k,
(cid:20) 1 0
2n 1
(cid:21)
.
We can see this by observing the q intervals at level θ = p/q map to the central q
intervals of the set of q(cid:48) = 2np + q intervals at level θ(cid:48) = p(cid:48)/q.
(31)
In Figure 5, the interval maps are of the form
n·(q−p)+k = I(cid:48)
Ik (cid:55)→ I(cid:48)
for the vertical similarity given by
n·(q(cid:48)−p(cid:48))+k,
(cid:20) 1 − 2n
2n
−2n 2n + 1
(cid:21)
.
(32)
Again, the q intervals at level θ = p/q map to the central q intervals of the set of
q(cid:48) = 2n(q − p) + q intervals at level θ(cid:48) = p(cid:48)/q.
14
MICHAEL P. LAMOUREUX, JAMES A. MINGO, AND SYDNEY R. PACHMANN
In Figure 6, a similarity in the top half, say with matrix
(33)
(35)
(cid:20) 1 1
1 2
(cid:21)
,
(cid:20)
(cid:21)
,
0 1−1 3
gives interval maps of the form
(34)
Ik (cid:55)→ I(cid:48)
r·p(cid:48)+k,
with r = 1 on the right side of the butterfly (as illustrated), and r = 0 on the left
side of the butterfly (not shown in Figure 6). However, a similarity in the bottom
half of Figure 6, such as with matrix
gives interval maps of the form
(36)
Ik (cid:55)→ I(cid:48)
r·(q(cid:48)−p(cid:48))+k,
with r = 1 on the right map, r = 0 on the left map.
Generally speaking, it appears there are two types of interval maps, those of the
form
(37)
Ik (cid:55)→ I(cid:48)
r·p(cid:48)+k,
and those of the form
(38)
Ik (cid:55)→ I(cid:48)
for some choices of non-negative integer r.
r·(q(cid:48)−p(cid:48))+k,
In the next three sections we determine how the horizontal component of the
similarity map behaves within each interval Ik.
7. Horizontal similarity: Cubic case 1 (cid:55)→ 1/3
The vertical component of the self-similarity map is given by linear fractional
transformations, as described in Sections 2 through 4. The horizontal component
maps intervals to intervals in the spectra, and appears to be nearly linear, as we saw
in the construction of Figure 3. We investigate the details of this horizontal map.
Again referring to the butterfly list in Figure 1, we see that the spectral line at
height θ = 1 is approached by a dense cluster of spectral points a heights θ < 1.
Under the LFT mapping θ (cid:55)→ θ(cid:48) = θ/(2θ + 1), the point θ = 1 maps to θ(cid:48) = 1/3, and
this dense cluster of spectral points near 1 maps to a dense cluster near 1/3.
We would expect by continuity arguments that the horizontal map at θ = 1 should
be well-approximated by following how the nearby cluster points map. Setting ra-
tional p/q = θ < 1, we have p ≈ q, and under the LFT we have p(cid:48)/q(cid:48) = p/(2 ∗ p + q).
SPECTRA SELF-SIMILARITY FOR ALMOST MATHIEU OPERATORS
15
Figure 9. Map from θ = 1 line to 1/3 lines (middle interval).
There are q points near line 1, and they map to the middle third of the 2p + q ≈ 3q
points under the LFT.
So we do a numerical experiment: Map these cluster points near 1 to corresponding
points near 1/3, and plot the result, to obtain Figure 9. We note the curve is almost
linear, but not quite. Performing a best fit polynomial approximation2, we find the
map is in fact given by
(39)
x = 6y − y3.
We observe that these polynomial maps are known from earlier work (Choi et al.,
1990) as characteristic polynomials Pθ for matrices used to compute the spectra of
q× q matrices representing the operators hθ in the rotation algebra Aθ, with θ = p/q.
Thus, the map we observe is simply
(40)
where by Pp/q(x) we mean the q-th order characteristic polynomial of the q×q matrix
(41)
H = U + U∗ + V + V ∗,
P1(x) = −P1/3(y),
with U the cyclic permutation matrix and V a diagonal matrix with consecutive
powers of e2πip/q on the diagonal. Details are in (Choi et al., 1990; Lamoureux and
Mingo, 2007).
To see the full algebraic curve, we plot the full curve in Figure 10. Observe
the central portion of the curve corresponds to the nearly linear section we saw in
Figure 9.
2In fact, in our first numerical experiments, we inadvertently computed the inverse of a polyno-
mial, with expansion y = x/6 + x3/1296 + x5/93312 + ··· . The second author recognized this as
the inverse of a cubic!
-4-2024-2-101216
MICHAEL P. LAMOUREUX, JAMES A. MINGO, AND SYDNEY R. PACHMANN
Figure 10. Algebraic curve P1(x) = −P1/3(y) defines the 1 to 1/3 map.
Figure 11. Map from 1/3 lines to 1/5 lines (3 middle intervals).
8. Horizontal similarity: Quintic case 1/3 (cid:55)→ 1/5
We repeat the numerical calculation, using the same central symmetry with LFT
given by the matrix
(42)
(cid:20) 1 0
2 1
(cid:21)
.
But now we look how the three spectral lines at θ = 1/3 maps to the three central
spectral lines at level θ(cid:48) = 1/5. A numerical experiment as in the last section produces
the plot in Figure 11. Again, the three segments look nearly linear, but not quite.
An inspired guess suggest these segments come from a portion of the algebraic curve
(43)
P1/3(x) = −P1/5(y),
and a comparison of the two plots shows this is indeed the case.
-6-4-20246-4-2024-2-112-1-0.50.51SPECTRA SELF-SIMILARITY FOR ALMOST MATHIEU OPERATORS
17
Figure 12. Algebraic curve P1/3(x) = −P1/5(y) defines the 1/3 to 1/5
map. Note the three nearly linear segments across the centre diagonal,
matching Figure 11.
It is worth noting that we can restrict the algebraic curve to the segments shown
in Figure 12 by plotting only those points (x, y) such that P1/3(x) = −P1/5(y) and
P1/3(x) ≤ 4.
9. Horizontal similarity: Algebraic curves
We have seen in Sections 7 and 8 that the horizontal maps appear as nearly linear
sections of algebraic curves, in the form
Pθ(x) = ±Pθ(cid:48)(x(cid:48)) ≤ 4,
for the two characteristic polynomials Pθ and Pθ(cid:48), where θ, θ(cid:48) are rational. The full
algebraic curves are interesting in themselves, so in this section we present a few
plots of the curves and point out some obvious patterns.
Referring to Figure 13, we see the algebraic curves Pθ(x) = Pθ(cid:48)(y) for rationals
θ = 1/q (cid:55)→ θ(cid:48) = 1/(q + 2), with q odd, have a particularly simple graphs. There is a
single connected component, that spirals around the origin, with more spirals as the
denominator q increases. We see q nearly linear segments, which are the parts of the
graph that define the relevant horizontal map on the fractal. We also see the odd
symmetries of the curves, that point (x, y) is in the curve if and only if (−x,−y) is
in it too.
-4-2024-4-202418
MICHAEL P. LAMOUREUX, JAMES A. MINGO, AND SYDNEY R. PACHMANN
Figure 13. A sequence of related algebraic curves, for rationals: 1 (cid:55)→
1/3, 1/3 (cid:55)→ 1/5, 1/5 (cid:55)→ 1/7, 1/7 (cid:55)→ 1/9, 1/9 (cid:55)→ 1/11. Notice the odd
number of nearly linear segments down the diagonal of each map.
Referring to Figure 14, we see the maps for rationals 1/q (cid:55)→ 1/(q + 2), with q even,
have somewhat more complex graphs. There are q/2 connected components, that
form something like figure eights around the origin, with more figure eights as the
denominator q increases. We see q nearly linear segments, which are the parts of the
-6-4-20246-4-2024-4-2024-4-2024-4-2024-4-2024-4-2024-4-2024-4-2024-4-2024SPECTRA SELF-SIMILARITY FOR ALMOST MATHIEU OPERATORS
19
Figure 14. A sequence of related algebraic curves: 1/2 (cid:55)→ 1/4, 1/4 (cid:55)→
1/6, 1/6 (cid:55)→ 1/8, 1/8 (cid:55)→ 1/10. Notice the symmetry, and the even
number of nearly linear segments down the diagonal of each map.
graph that define the relevant horizontal map on the fractal. We also see the even
symmetries of the curves, that point (x, y) is in the curve if and only if the points
(±x,±y) are in it too.
Figure 15 examines the curves for rational θ of the form θ = p/3. In both cases
(p = 1/3, 2/3), there is odd symmetry, there are three nearly linear segments, and
some spiraling of the curve about the origin. But for the case θ = 2/3, there are also
two additional disconnected components to the curve, that seem to have nothing to
do with the linear segments that are involved in the fractal map.
In Appendix 3, we include more plots of these algebraic curves. We note with
even denominator q, we have even symmetry in the plots, there are q nearly linear
segments, and we see nested "figure eights" defining the key parts of the curve. For
p > 1, there are disconnected components of the curve that seem to have nothing to
do with the main part of the curve that defines the fractal map.
-4-2024-4-2024-4-2024-4-2024-4-2024-4-2024-4-2024-4-202420
MICHAEL P. LAMOUREUX, JAMES A. MINGO, AND SYDNEY R. PACHMANN
Figure 15. Algebraic curves with denominator 3: 1/3 (cid:55)→ 1/5, 2/3 (cid:55)→
2/7. Notice the the three nearly linear segments down the diagonal of
each map, and the disconnected components on the right graph.
We also observe the curves appear to get much more complex with increases in the
numerator p, for the parameter θ = p/q. There may be some connection between
the various straight line segments in curves of different θ = p/q with the same
denominator q, but these patterns are not entirely clear.
We leave discussion of these curves and their connection with symmetries to future
work.
10. The similarity maps: General case
The work in Sections 7 through 9 suggests an obvious candidate for the horizontal
maps within spectral intervals: namely, a correspondence of points determined by
two characteristic polynomials. Combining this observation with the earlier sections
discussing the vertical maps and interval mapping, we can now specify the form of
the general similarity maps of the butterfly, as follows:
Fix a matrix element M in GL2(Z), represented as
(cid:21)
(cid:20) a b
c d
(44)
Fix an integer r ≥ 0. The similarity map S = SM,r,+ on the rational butterfly is
described as
M =
.
(45)
(x, θ) (cid:55)→ S(x, θ) = (x(cid:48), θ(cid:48)),
-4-2024-4-2024-4-2024-4-2024SPECTRA SELF-SIMILARITY FOR ALMOST MATHIEU OPERATORS
21
where the vertical component of the map is given by
(46)
θ (cid:55)→ θ(cid:48) =
aθ + b
cθ + d
,
x ∈ Ik (cid:55)→ x(cid:48) ∈ I(cid:48)
r·p(cid:48)+k = Ik(cid:48) with (−1)q+kPθ(x) = (−1)q(cid:48)+k(cid:48)
while the horizontal component is given by the interval and polynomial correspon-
dences. Specifically, recall the spectrum at level θ = p/q is a set of q intervals Ik,
k = 1...q, while at level θ(cid:48) = p(cid:48)/q(cid:48) there are q(cid:48) intervals I(cid:48)
k(cid:48), k(cid:48) = 1...q(cid:48). The map from
x at level θ to x(cid:48) at level θ(cid:48) is determined by the interval and polynomial conditions
(47)
The sign choice for the polynomials is made so that ±Pθ(x) and ±Pθ(cid:48)(x(cid:48)) are both
monotonically increasing on Ik, I(cid:48)
k(cid:48), respectively. These two conditions determine the
image point x(cid:48) uniquely, except possibly in one special case. Except for this case
(described next), we have determined the similarity maps precisely.
The special case to consider is when x = 0, θ = p/q with q even. In this case, the
two intervals Iq/2, Iq/2+1 overlap at x = 0 and may map to two disjoint intervals I(cid:48)
k(cid:48)
and I(cid:48)
k(cid:48)+1. Here, we must decide where the point x = 0 maps to, either an endpoint
k(cid:48) or to an endpoint of I(cid:48)
of I(cid:48)
k(cid:48)+1. It is convenient to chose to map to both points,
making the function S double-valued at this point (x = 0, θ = p/q). We will see in
the next section that the double-valued function is continuous.
Pθ(cid:48)(x(cid:48)).
There is another set of similarity maps, S = SM,r,−, which differs from the above
example by the way the interval maps are selected. For this maps, we chose
(48)
x ∈ Ik (cid:55)→ x(cid:48) ∈ I(cid:48)
r·(q(cid:48)−p(cid:48))+k = Ik(cid:48),
which we observed in Section 6 as similarities that can occur.
Of course, not all such maps SM,r,± necessarily appear as similarities of the rational
butterfly. For instance, from Theorem 1, we know the matrix M must be chosen from
the semigroup G ⊂ GL2(Z)/±I described in Theorem 1. The integer r must be small
enough that
r · p(cid:48) + q ≤ q(cid:48)
r · (q(cid:48) − p(cid:48)) + q ≤ q(cid:48)
r ≤ min
θ
cθ + d − 1
aθ + b
r ≤ min
θ
cθ + d − 1
(c − a)θ + d − b
in the " -- " case, for all θ = p/q. Equivalently, we have
(49)
in the "+" case, and
(50)
(51)
in the "+" case, and
(52)
22
MICHAEL P. LAMOUREUX, JAMES A. MINGO, AND SYDNEY R. PACHMANN
in the " -- " case. But these are the only apparent restrictions on M, r,±.
11. The similarity maps: Gap labelling
The spectral gaps forming the butterfly are conveniently labeled using a Diophan-
tine equation, where integer parameters (s, t) are fixed, and the k-th gap in the
spectrum θ = p/q is identified using the formula
k = t ∗ p − s ∗ q.
(53)
This indexing is known as gap labelling, as described in (Bellissard, 1990; Goldman,
2009; Kaminker and Putnam, 2003; Ypma, 2007), where the parameters (s, t) are
related to Chern numbers. These parameters give a convenient way of labelling the
"wings" in the butterfly. Note there is a limited range on the indices: for t > 0 we
require
(54)
while for t < 0 we require
(55)
0 ≤ s ≤ t − 1,
t ≤ s ≤ −1.
It is easy to check that the similarity map S = SM,r,± maps labeled gaps to gaps,
as stated in the following:
Theorem 2. The similarity map S = SM,r,+ maps the gap labeled (s, t) to the gap
(s(cid:48), t(cid:48)) according to the formula
Also, the similarity map S = SM,r,− maps the gap labeled (s, t) to the gap (s(cid:48), t(cid:48))
according to the formula
(cid:21)
(cid:21)
t(cid:48)
(cid:20) s(cid:48)
(cid:20) s(cid:48)
t(cid:48)
(56)
(57)
r
(cid:20) 0
(cid:20) r
r
(cid:21)
(cid:21)
.
.
+
+
c d
(cid:20) a b
(cid:20) a b
(cid:20) a b
c d
t
(cid:21)(cid:20) s
(cid:21)(cid:20) s
(cid:21)(cid:20) p
t
(cid:21)
(cid:21)
(cid:21)
= (ad − bc)
= (ad − bc)
(cid:20) p(cid:48)
(cid:21)
Proof: We consider S = SM,r,+. Note the matrix M in the similarity determines
the map from angle θ = p/q to θ(cid:48) = p(cid:48)/q(cid:48) according to the matrix formula
(58)
In the case where the determinant ad − bc is one, we invert to find
(59)
p = (dp(cid:48) − bq(cid:48)) and q = (−cp(cid:48) + aq(cid:48)).
c d
=
q(cid:48)
q
.
SPECTRA SELF-SIMILARITY FOR ALMOST MATHIEU OPERATORS
23
k(cid:48) = r · p(cid:48) + k
The index r of the similarity tells us how intervals Ik are mapped to intervals Ik(cid:48),
where k(cid:48) = r · p(cid:48) + k. Thus the k-th gap is mapped to gap k(cid:48) = r · p(cid:48) + k. With gap
k = t ∗ p − s ∗ q, we write
(60)
(61)
(62)
(63)
(64)
from which we read off s(cid:48) = (as + bt) and t(cid:48) = (cs + dt + r), as desired.
The other cases (negative determinant, and S = SM,r,−) are similar.
= r · p(cid:48) + t ∗ p − s ∗ q
= r · p(cid:48) + t(dp(cid:48) − bq(cid:48)) − s(−cp(cid:48) + aq(cid:48))
= (cs + dt + r)p(cid:48) − (as + bt)q(cid:48)
= t(cid:48) ∗ p(cid:48) − s(cid:48) ∗ q(cid:48),
QED
12. The similarity maps: Proof of continuity
We show first continuity along the horizontal direction for the similarity maps
described Section 10:
Theorem 3. Given two rational parameters θ = p/q, θ(cid:48) = p(cid:48)/q(cid:48), and a correspon-
dence between points in pairs of intervals Ik, I(cid:48)
k(cid:48) in the spectra Spec(hθ), Spec(hθ(cid:48))
respectively, given by
(65)
x ∈ Ik (cid:55)→ x(cid:48) ∈ Ik(cid:48) = I(cid:48)
r·p(cid:48)+k with (−1)q+kPθ(x) = (−1)q(cid:48)+k(cid:48)
Pθ(cid:48)(x(cid:48)),
then the correspondence is a continuous bijection on each interval. In particular, this
yields a continuous map from the set of points x ∈ Spec(hθ) into x(cid:48) ∈ Spec(hθ(cid:48)), with
the possible exception at the point x = 0, when q is even.
Proof: The polynomial maps (−1)q+kPθ(x), (−1)q(cid:48)+k(cid:48)
Pθ(cid:48)(x(cid:48)) are both continuous
and monotonically increasing on the intervals in question, and thus have continuous
inverses. So on each interval, we have a continuous bijection. For q odd, the spectrum
Spec(hθ) is a disjoint union of the intervals I1...Iq, and so the collection extends to a
continuous map into Spec(hθ(cid:48)). In the case where q is even, two of the intervals Iq/2
and Iq/2+1 overlap at the point x = 0. So, except at this point, the correspondence
extends to a continuous map on the union of intervals -- that is, it is continuous at
Spec(hθ)\0. At the point x = 0, the correspondence maps the one point to (possibly)
two points, one at the right endpoint of I(cid:48)
r·p(cid:48)+q/2 and the other at the left endpoint
of I(cid:48)
r·p(cid:48)+q/2+1. This double-valued splitting (although not a singled-valued function)
QED
is nevertheless continuous.
Figure 16 indicates graphically how the bijection works on level θ = 1/3 to θ = 1/5.
We note that the correspondences of the form
(66)
x ∈ Ik (cid:55)→ x(cid:48) ∈ Ik(cid:48) = I(cid:48)
r·(q(cid:48)−p(cid:48))+k with (−1)q+kPθ(x) = (−1)q(cid:48)+k(cid:48)
Pθ(cid:48)(x(cid:48)),
24
MICHAEL P. LAMOUREUX, JAMES A. MINGO, AND SYDNEY R. PACHMANN
Figure 16. Plot of cubic polynomial P1/3(x) and quintic P1/5(x(cid:48)).
Match the zero crossing, move polynomials up and down, and observe
a continuous correspondence of the zero crossings. This gives the con-
tinuous correspondence between line spectra.
also are continuous, as in the theorem. This covers the similarity maps of the form
SM,r,−.
The numerical work suggests that piecing together these individual polynomial
maps, at the different θ levels, gives a continuous map on the whole butterfly. We
now state and prove this as a theorem:
Theorem 4. Suppose S = SM,r,± is a similarity of the rational butterfly, as described
in Section 10. Then S is a continuous (single-valued) map, except possibly at the
points (x = 0, θ = p/q) with q even, where the map is double-valued and continuous.
Proof: We consider the case S = SM,r,+. Fix a sequence of points (xn, θn) converg-
ing to point (x∗, θ∗) and set
(67)
By continuity of the linear fractional transformation, the sequence θ(cid:48)
n converges to
θ(cid:48)
∗ = (aθ∗ + b)/(cθ∗ + d). To show continuity of S, we only need to verify convergence
of the x(cid:48)
n, θ(cid:48)
n)
converges to the point (x(cid:48)
∗) = S(x∗, θ∗), which, by compactness, establishes conti-
nuity.
n. We will show every convergent subsequence of the image points (x(cid:48)
n) = S(xn, θn).
(x(cid:48)
n, θ(cid:48)
∗, θ(cid:48)
-4-224-100-5050-50SPECTRA SELF-SIMILARITY FOR ALMOST MATHIEU OPERATORS
25
n, θ(cid:48)
n converges to x(cid:48)
n) with limit (x(cid:48)
n = θ(cid:48)
Take a convergent subsequence of the (x(cid:48)
The other case to consider is when there are at most finitely many n with θ(cid:48)
o, θ(cid:48)
∗). Suppose in
this subsequence, there are infinity many points where θ(cid:48)
∗ (and consequently,
θn = θ∗). Applying Theorem 2, by continuity along the spectral lines at θ∗, we can
conclude that x(cid:48)
∗. (Except possibly at x∗ = 0, θ∗ = p/q, with q even.
Here, the map S may be doubly valued, as discussed earlier.)
n = θ(cid:48)
∗.
By restricting to the tail of the subsequence, and renumbering the subsequence,
WLOG we may assume θn (cid:54)= θ∗ for all n.
and points xn ∈ Ikn, x∗ ∈ Ik∗, while in the range we have θ(cid:48)
points x(cid:48)
and k(cid:48)
Set the notation consistently, so in the domain we have θn = pn/qn, θ∗ = p∗/q∗,
∗, and
n + kn
n, θ(cid:48)
∗ = p(cid:48)
∗/q(cid:48)
n = r· p(cid:48)
. From the definition of S, we know k(cid:48)
o is to be determined (it will equal k(cid:48)
∗).
n ∈ I(cid:48)
, x(cid:48)
k(cid:48)
∗ = r · p(cid:48)
∗ + k∗. The value of k(cid:48)
o ∈ I(cid:48)
k(cid:48)
∗ ∈ I(cid:48)
k(cid:48)∗, x(cid:48)
n = p(cid:48)
n/q(cid:48)
We apply the trace τθn to the spectral projection χ(−∞,xn)(hθn). By Lemma 7 in
n
o
Appendix 2, we have
(68)
τθn(χ(−∞,xn)(hθn)) =
kn − 1
qn
+
1
qn
F ((−1)kn+qnPθn(xn)),
where the function F is an integrated density of states on ho. By continuity of the
trace and spectral projection (see (Boca, 2001), Chap. 5), this quantity converges to
(69)
τθ∗(χ(−∞,x∗)(hθ∗)) =
k∗ − 1
q∗
+
1
q∗
F ((−1)k∗+q∗Pθ∗(x∗)).
Since the sequence of ratios pn/qn (cid:54)= p∗/q∗ converges to p∗/q∗, we have that qn
converges to infinity, so the terms above with 1/qn in them converge to zero. Thus,
from the limit of the traces, we conclude
(70)
convergent sequence x(cid:48)
lim
n→∞
kn
qn
n → x(cid:48)
k∗ − 1
q∗
=
F ((−1)k∗+q∗Pθ∗(x∗)).
+
1
q∗
Similarly, by applying the trace to the sequence of spectral projections on the
(71)
converges to
(72)
Thus
(73)
τθ(cid:48)
n(χ(−∞,x(cid:48)
n)(hθ(cid:48)
o, we find
n − 1
k(cid:48)
q(cid:48)
n)) =
n
+
τθ(cid:48)∗(χ(−∞,x(cid:48)
o)(hθ(cid:48)∗)) =
o − 1
k(cid:48)
q(cid:48)∗
+
F ((−1)k(cid:48)
n+q(cid:48)
n(x(cid:48)
n)),
nPθ(cid:48)
F ((−1)k(cid:48)
o+q(cid:48)∗Pθ(cid:48)∗(x(cid:48)
o)).
1
q(cid:48)
n
1
q(cid:48)∗
k(cid:48)
q(cid:48)
n
n
o − 1
k(cid:48)
q(cid:48)∗
+
1
q(cid:48)∗
=
lim
n→∞
F ((−1)k(cid:48)
o+q(cid:48)∗Pθ(cid:48)∗(x(cid:48)
o)).
26
MICHAEL P. LAMOUREUX, JAMES A. MINGO, AND SYDNEY R. PACHMANN
With
(74)
k(cid:48)
q(cid:48)
n
n
r · p(cid:48)
n + kn
q(cid:48)
n
=
=
r · (apn/qn + b) + kn/qn
cpn/qn + d
we take limits to obtain the identity
(75)
lim
k(cid:48)
q(cid:48)
n
n
= r · θ(cid:48)
∗ +
1
cθ∗ + d
lim
kn
qn
.
Replacing with the limiting values computed above, we obtain
(76)
o − 1
k(cid:48)
q(cid:48)∗
o)) = r·θ(cid:48)
∗+
o+q(cid:48)∗Pθ(cid:48)∗(x(cid:48)
F ((−1)k(cid:48)
cθ∗ + d
1
q(cid:48)∗
1
q∗
q∗
+
+
1
(cid:18) k∗ − 1
(cid:19)
F ((−1)k∗+q∗Pθ∗(x∗))
.
Multiplying through by q(cid:48)
o − 1 + F ((−1)k(cid:48)
k(cid:48)
(77)
∗ and simplifying yields
o+q(cid:48)∗Pθ(cid:48)∗(x(cid:48)
o)) = r · p(cid:48)
∗ + k∗ − 1 + F ((−1)k∗+q∗Pθ∗(x∗)).
Now, if x∗ is an interior point of the interval Ik∗, this last equation has F taking
a fractional value strictly between zero and one. Equating the integer part of the
equation gives
(78)
from which we conclude k(cid:48)
point x(cid:48)
∗ are in the same spectral interval I(cid:48)
k∗.
o − 1 = r · p(cid:48)
k(cid:48)
∗, which tells us that the limit point x(cid:48)
o = k(cid:48)
∗ + k∗ − 1,
Equating the fractional part of the equation gives
o and image
F ((−1)k(cid:48)∗+q(cid:48)∗Pθ(cid:48)∗(x(cid:48)
F ((−1)k(cid:48)∗+q(cid:48)∗Pθ(cid:48)∗(x(cid:48)
o)) = F ((−1)k∗+q∗Pθ∗(x∗)),
(79)
which, from the polynomial correspondence between x∗ and x(cid:48)
o)) = F ((−1)k(cid:48)∗+q(cid:48)∗Pθ(cid:48)∗(x(cid:48)
(80)
∗)).
Since F , composed with polynomial (−1)k(cid:48)∗+q(cid:48)∗Pθ(cid:48)∗(x) is strictly increasing on the
interval I(cid:48)
o must be the corresponding endpoint of interval I(cid:48)
limit point x(cid:48)
left endpoint of interval Ik: in this case, equation 77 reduces to
Now, when x∗ is at an endpoint of interval Ik, we use gap labeling to show the
k∗. Suppose x∗ is the
k∗, the equality implies x(cid:48)
∗, as desired.
o = x(cid:48)
∗, yields
o − 1 + F ((−1)k(cid:48)
k(cid:48)
(81)
so we know that k(cid:48)
o equals either k(cid:48)
o)) = r · p(cid:48)
∗. We just need to show it is equal to k(cid:48)
∗.
With the point x∗ (cid:54)= 0 at a left endpoint, there must be a gap to the immediate
left of x∗, which has some label (s, t). By continuity of gap labeling (Prop 11.11
in (Boca, 2001)), the gap with this label forms an open set, so for n sufficiently
large, the point xn is to the right of this gap. Hence the interval Ikn is also to the
o+q(cid:48)∗Pθ(cid:48)∗(x(cid:48)
∗ − 1 or k(cid:48)
∗ + k∗ − 1 + 0.
SPECTRA SELF-SIMILARITY FOR ALMOST MATHIEU OPERATORS
27
right of the gap. Since the interval number must be bigger than the gap number, we
have
kn > t ∗ pn − s ∗ qn.
kn and its index satisfies
(82)
kn is in interval I(cid:48)
The image point x(cid:48)
n = r · p(cid:48)
k(cid:48)
n + kn
n + t ∗ pn − s ∗ qn
> r · p(cid:48)
= r · p(cid:48)
n + t ∗ (dp(cid:48)
n − bq(cid:48)
n) − s ∗ (−cp(cid:48)
n − (as + bt) ∗ q(cid:48)
= (cs + dt + r) ∗ p(cid:48)
= t(cid:48) ∗ p(cid:48)
(83)
(84)
(85)
(86)
(87)
where the indices (s(cid:48), t(cid:48)) label the image gap, as in Theorem 2. Thus the interval I(cid:48)
k(cid:48)
is to the right of the gap (s(cid:48), t(cid:48)) and hence so is the point x(cid:48)
n. Since this gap is open,
the limit point x(cid:48)
ko containing point
x(cid:48)
o is to the right. Consequently, we have
o is to the right of the gap, and thus the interval I(cid:48)
n − s(cid:48) ∗ q(cid:48)
n
n + aq(cid:48)
n)
n
n
∗ − 1.
∗ − s(cid:48)q(cid:48)
k(cid:48)
o > t(cid:48)p(cid:48)
(88)
This eliminates the possibility that k(cid:48)
o = k(cid:48)
which we quickly conclude that the point x(cid:48)
thus is equal to x(cid:48)
∗.
common endpoint of two overlapping intervals shows the image sequence x(cid:48)
converge to endpoints of two different, disjoint intervals.
Handling a right endpoint is similar. A similar analysis of the case x∗ = 0 at the
n could
QED
∗ = k(cid:48)
∗ − 1, so we are left with k(cid:48)
o = k(cid:48)
o is the left endpoint of interval I(cid:48)
∗, from
k∗ and
It would seem the result on the rational butterfly should extend by continuity to
the full butterfly, including irrational values for vertical parameter θ. We state this
precisely as a conjecture.
Conjecture 5. Given a similarity S = SM,r,± of the rational butterfly, as described
in Section 10, there is a unique continuous extension of S to the full butterfly, possibly
double-valued at certain points along the vertical line x = 0.
We do not have a proof of this result. It seems a density argument and use of the
continuous trace on the field of rotation algebras Aθ should lead to the result.
It is curious to consider how the polynomial correspondences on the rational spec-
tral lines might extend to irrational values of θ, in which case there are no finite
polynomials to describe the mapping.
28
MICHAEL P. LAMOUREUX, JAMES A. MINGO, AND SYDNEY R. PACHMANN
13. Three generators for the butterfly similarities
There are three similarity maps of the butterfly which apparently generate all the
similarities discussed above. There is the horizontal flip H defined as the map
There is the vertical flip V defined as the map
(89)
(90)
(91)
with matrix B =
(cid:21)
(cid:20) −1 1
0
1
H(x, θ) = (−x, θ).
V (x, θ) = (x, 1 − θ),
V = SB,0,+,
which can be represented in the form discussed in Section 10 as
. Finally, there is the similarity map discussed in Sec-
tion 4, taking the butterfly to the bottom half given by
S(x, θ) = (x(cid:48),
θ
(92)
where x ∈ Ik in the k-th interval at level θ is mapped to x(cid:48) ∈ I(cid:48)
at level θ(cid:48). This map is represented as
(93)
S = SA,0,+,
θ + 1
),
(cid:20) 1 0
1 1
(cid:21)
.
for matrix A =
k in the k-th interval
Some elementary calculations show how these three similarities combine alge-
braically. First, there are the two obvious identities
V 2 = I,
H 2 = I,
(94)
and the commutation relation
(95)
HV = V H.
The horizontal flip almost commutes with similarity S; in fact it just gives a shift
in the interval indexing as follows:
(96)
HSA,0,+H = SA,1,+.
This shifting extends to powers of S, so we find
(97)
(HSA,0,+H)n = SAn,n,+.
By composing powers of S and HSH, we obtain all similarities of the form
(98)
SAn,r,+
for 0 ≤ r ≤ n.
We can also verify that such operators combine in the form
,r+r(cid:48),+.
,r(cid:48),+)(SAn,r,+) = SAn+n(cid:48)
(SAn(cid:48)
(99)
SPECTRA SELF-SIMILARITY FOR ALMOST MATHIEU OPERATORS
29
The vertical flip does not commute with similarity S, but instead introduces the
similarity SM,r,− in the mix. Again, a straightforward calculation shows
(100)
V SAn,r,+V = SBAnB,r,−.
As noted in Theorem 1, the matrices A, B generate a large semigroup of linear
fractional transformations. From the calculations noted above, the three similarities
of the butterfly generate enough similarities to cover this semigroup.
Theorem 6. The three similarities H, V, S of the rational butterfly generate a semi-
group of continuous (possibly double-valued) similarities of the butterfly, including
maps of the form
(101)
where the matrices M range over the semigroup of elements of GL2(Z)/ ± I repre-
senting linear fractional transformations mapping the interval [0, 1] into itself.
SM,r,+
and
SM,r,−
The double-valued character is, of course, from the interval splitting in the case of
θ = p/q with q even. There may be other similarities of the butterfly. The point is,
we can at least see this large semigroup from GL2(Z) appearing.
14. Conclusions
The Hofstadter butterfly, representing spectra of a continuous family of almost
Mathieu operators, shows obvious fractal-like symmetry. By investigating these sym-
metries numerically, we have catalogued the self-similarity maps of the butterfly using
a semigroup of Mobius transformations in the vertical direction, indexed by elements
in the matrix group GL2(Z). This semigroup is generated by two matrices. In the
horizontal direction, the self-similarity maps are given by algebraic curves determined
by characteristic polynomials. Properties of the algebraic curves are demonstrated
in a series of plots. These algebraic curves show nearly linear segments, demon-
strating an almost linear behaviour in the horizontal component of the self-similarity
maps. We proved continuity of the similarity maps on the rational butterfly, possibly
double-valued at points with horizontal parameter x = 0. The similarity maps are
generated by exactly three continuous symmetries. We conjecture the semigroup of
similarities on the rational butterfly extends to a family of continuous similarities on
the full butterfly.
Acknowledgments
This work was supported in part by NSERC Discovery grants of the first two
authors, and an NSERC Summer Research award of the third author. Numerical
calculations were done in MATLAB (TheMathWorks, 2008) and rendered directly
30
MICHAEL P. LAMOUREUX, JAMES A. MINGO, AND SYDNEY R. PACHMANN
in PostScript, following a method similar to those discussed in (Casselman, 2005).
Algebraic curves were rendered using Mathematica (WolframResearch, 2002).
We would like to thank the organizers of the 2006 BIRS Workshop on Operator
Methods in Fractal Analysis, Wavelets, and Dynamical Systems, for encouraging us
to present this work in an early form.
References
Arveson, W. 1994. C*-algebras and numerical linear algebra, J. Funct. Anal. 122, no. 2, 333 -- 360.
Avila, A. and S. Jitomirskaya. 2006. Solving the ten martini problem, Lect. Notes. in Physics 690,
5 -- 16.
Bellissard, J. 1990. Gap labelling theorems for Schrodinger operators, From number theory to
physics, pp. 140 -- 150.
Bellissard, J. and B. Simon. 1982. Cantor spectrum for the almost Mathieu equation, Journal of
Functional Analysis 48, no. 3, 408 -- 419.
Boca, F. 2001. Rotation c*-algebras and almost Mathieu operators, Fundatia Theta.
Brown, E. 1964. Bloch electrons in a uniform magnetic field, Phys. Rev. 133, no. 4A, A1038 --
A1044.
Casselman, B. 2005. Mathematical Illustrations, Cambridge University Press.
Choi, M. D., G. A. Elliott, and N. Yui. 1990. Gauss polynomials and the rotation algebra, Invent.
Math. 99, 225 -- 246.
Connes, A. 1994. Noncommutative geometry, Academic Press.
Coxeter, H.S.M. 1942. Non-euclidean geometry, University of Toronto Press, Toronto.
Effros, E. and C. Shen. 1980. Approximately finite C*-algebra and continued fractions, Indiana
Univ. Math. J. 12, no. 2, 191 -- 204.
Goldman, N. 2009. Characterizing the Hofstadter butterfly's outline with Chern numbers, J. Phys.
B: At. Mol. Opt. Phys. 42, no. 055302.
Hofstadter, D. R. 1976. Energy levels and wave functions of Bloch electrons in rational and irra-
tional magnetic fields, Physical Review B 14, no. 6, 2239 -- 2249.
Kaminker, J. and I. Putnam. 2003. A proof of the gap labeling conjecture, Michigan Math J. 51,
537 -- 546.
Lamoureux, M. P. 1997. Reflections on the almost Mathieu operator, Integr. equ. oper. theory 28,
no. 1, 45 -- 59.
. 2010. Drawing butterflies from the almost Mathieu operator. preprint.
Lamoureux, M. P. and J. A. Mingo. 2007. On the characteristic polynomial of the almost Mathieu
operator, Proc. AMS 135, 3205 -- 3215.
Last, Y. 1994. Almost everything about the almost Mathieu operator, Proceedings of the XI-th
International Congress of Mathematical physics.
Mandelbrot, Benoit B. 1990. The fractal geometry of nature, Spektrum Akademischer Verlag.
Puig, Joaquim. 2004. Cantor spectrum for the almost Mathieu operator, Comm. in Math Phys
244, no. 2, 297 -- 309.
TheMathWorks. 2008. MATLAB R2008a.
WolframResearch. 2002. Mathematica 4.2.
Ypma, F. 2007. K-theoretic gap labeling for quasicrystals, Contemporary mathematics: Geometric
and topological methods for quantum field theory., pp. 247.
(cid:20) a b
(cid:21)
(cid:20) 1 0
(cid:21)
SPECTRA SELF-SIMILARITY FOR ALMOST MATHIEU OPERATORS
31
Appendix 1
Proof of Theorem 1: We show the corresponding semigroup in GL(2, Z)/{±I}
is generated by the two matrices A and B. Suppose the matrix
(102)
is in the semigroup. In the case that b = 0, the determinant condition gives a = ±d =
±1; the condition that the LFT maps [0, 1] into [0, 1] reduces WLOG to a = d = 1
and c ≥ 0. This leave matrix M in the form
M =
c d
= Ac,
c 1
M =
(103)
with c ≥ 0. Hence the matrix M is generated by A alone.
In the case that b (cid:54)= 0, WLOG (by mod-ing out by ±I), we can assume d ≥
b > 0 as we know that 0 maps to b/d ∈ [0, 1] under the LFT. By the determinant
condition ad − bc = ±1, we have that the gcd of b, d is one, so we may apply the
Euclidean algorithm to obtain a sequence of strictly positive quotients q0, q1, q2, . . .
and remainders b = r0, r1, r2, . . . with
d = q0r0 + r1
b = q1r1 + r2
r1 = q2r2 + r3
. . .
rk−1 = qkrk + rk+1,
where rk = gcd(b, d) = 1 and rk+1 = 0. Applying these values to matrix M , we can
factor as
for some appropriate values of a0, a1, . . . , c0, c1, . . ..
Now, the matrix factors with the qj are generated by the A, B matrices, since
(104)
= Aq−1BA.
(cid:21)
(cid:21)(cid:20) a1 r2
(cid:20) 0
(cid:21)
c1 r1
···
(cid:21)
(cid:21)(cid:20) ak 0
(cid:21)
ck 1
,
1
1 qk
(cid:21)
(cid:20) a b
c d
=
=
=
(cid:20) 0
(cid:20) 0
(cid:20) 0
1
1 q0
1
1 q0
1
1 q0
(cid:21)(cid:20) a0 r1
(cid:21)(cid:20) 0
(cid:21)(cid:20) 0
(cid:21)
(cid:20) 0 1
c0 r0
1
1 q1
1
1 q1
1 q
32
MICHAEL P. LAMOUREUX, JAMES A. MINGO, AND SYDNEY R. PACHMANN
In the last factor,
(105)
we know that ck is non-negative, else the LFT has a pole in [0, 1], which is not
allowed. By the determinant condition, we have ak = ±1. In the case ak = 1, we
have that the last factor appears as
(106)
In the case where ak = −1, we may combine the last two factors to observe
ck 1
= Ack.
(cid:20) 0 1
(cid:21)(cid:20) −1 0
(cid:21)
1 q
c 1
(cid:20) ak 0
ck 1
(cid:21)
,
(cid:21)
(cid:20) 1 0
(cid:20) 0 1
(cid:20) 0
1 q
=
(cid:21)(cid:20) 1 −1
(cid:21)(cid:20) c − 1 1
(cid:21)(cid:20) c − 1 1
(cid:21)
c 1
0
1
(cid:21)
1
1 q − 1
=
= (Aq−2BA)(BAc).
c 1
(107)
(109)
For q ≥ 2, we are done: these last factors appear as generated by matrices A, B, as
desired.
In the case q = 1, it easy to check that we must have more than one q-type matrix
in the factorization, for if not, we get
(cid:21)
(cid:20) a b
c d
(cid:20) 0
=
1
1 q0
(cid:21)(cid:20) −1 0
(cid:21)
,
c0 1
where q0 = 1, giving an LFT mapping 1 to (c + 1)/c, which is outside the interval
[0, 1]. With a second q-type matrix in the factorization, we get
(108)
M = ··· (Aq−1BA)(A−1BA)(BAck) = ··· (Aq−1B)(AA−1)(BA)(BAck) = ··· AqBAck,
QED
which puts M in the semigroup generated by A, B.
Here is a nice example, to show that the ak = −1 case really does appear.
(cid:21)
(cid:20) 2 1
3 2
=
(cid:20) 0 1
(cid:21)(cid:20) −1 0
(cid:21)
1 2
2 1
The matrix on the left gives an LFT which maps 0 to 1/2, and 1 to 3/5. This simi-
larity map actually appears in the Hofstadter butterfly, as we can see by examining
Figure 1. We get a1 = −1 in the factorization, which perhaps was not expected.
Thus that little factor matrix on the right is not in the semigroup, since it maps 1
to -1/3, and thus does not map interval [0, 1] to itself.
SPECTRA SELF-SIMILARITY FOR ALMOST MATHIEU OPERATORS
33
However, the original matrix on the left is actually in the semigroup, as shown in
the proof.
Appendix 2
We prove a basic result concerning the trace of certain spectral projections in
the rotation algebra Aθ. The machinery for this result is standard, and we borrow
heavily from the results in (Boca, 2001).
Lemma 7. Let τθ and τo be the (framed) tracial states on rotation algebras Aθ, Ao,
respectively, with θ = p/q. Let [a, b] be the k-th interval Ik in the spectrum of the
operator hθ = u + u∗ + v + v∗ in algebra Aθ. Then for x ∈ [a, b],
(110)
τθ(χ[a,x)(hθ)) =
1
q
τo(χ[−4,(−1)k+qPθ(x))(ho)),
where χ[a,x) is the spectral projection onto interval [a, x). Consequently,
(111)
τθ(χ(−∞,x)(hθ)) =
k − 1
q
+
1
q
τo(χ[−4,(−1)k+qPθ(x))(ho)).
Note: the choice of the sign in front of polynomial Pθ(x) ensures that the map
(112)
x (cid:55)→ (−1)k+qPθ(x)
is monotonic increasing on the interval Ik. In the earlier sections of this paper, it is
convenient to write the trace-spectral projection function as the cumulative density
F (x) for the operator ho, so we may write
k − 1
(113)
τθ(χ(−∞,x)(hθ)) =
+
1
q
q
F ((−1)k+qPθ(x)),
where
(114)
is a continuous, strictly increasing function mapping [−4, 4] onto [0, 1].
F (x) = τo(χ[−4,x)(ho))
Proof: We do a direct calculation. The trace on Aθ is obtained from a representa-
tion of the algebra in C(T2) ⊗ Mq(C) with generators
u = ι1 ⊗ Uo, v = ι2 ⊗ Vo,
(115)
where the functions ι1, ι2 are the coordinate maps ι1(z1, z2) = z1, ι2(z1, z2) = z2, and
the q × q matrices Uo, Vo are the usual cyclic permutation and the diagonal with
powers of e2πiθ. The trace in this frame is just
(116)
τθ = µ2 ⊗ trq,
34
MICHAEL P. LAMOUREUX, JAMES A. MINGO, AND SYDNEY R. PACHMANN
where µ2 is the usual Haar measure on T2 and trq is the normalized trace on q × q
matrices. (See (Boca, 2001), page 11.)
Fix x in Ik. The matrix-valued function h = u + u∗ + v + v∗ has exactly one
eigenvalue in the interval Ik, which moves continuously for parameter values (z1, z2) ∈
T2. This eigenvalue contributes to the trace precisely when it lies in the interval
[a, x]. That is, we get a contribution precisely when the value of the polynomial
(−1)k+qPθ(x) is smaller than zq
q. We compute the trace in terms of
a characteristic function on the torus T2, normalized by q, so
q + zq
1 + z1
2 + z2
τθ(χ[a,x)(hθ)) =
(117)
q) dµ2.
With y = (−1)k+qPθ(x), and noting that the map (z1, z2) (cid:55)→ (zq
preserving inverse on the torus, we can write
χ((−1)k+qPθ(x)<zq
q+zq
2 +z2
1 +z1
T2
1, zq
2) has measure
(cid:90)
1
q
(cid:90)
(cid:90)
(cid:90)
T2
T2
T2
τθ(χ[a,x)(hθ)) =
=
=
=
1
q
1
q
1
q
1
q
χ(y<zq
1 +z1
q+zq
2 +z2
q) dµ2
χ(y<z1+z1+z2+z2) dµ2
χ(Po(y)<z1+z1+z2+z2) dµ2
τo(χ[−4,y)(ho)).
where we use the fact that Po(y) = y and so the last integral above yields the trace
in Ao applied to the spectral projection obtained by applying χ[−4,y] onto ho.
Using the relation y = (−1)k+qPθ(x), we obtain
τθ(χ[a,x)(hθ)) =
(118)
as desired. The trace evaluated on the full interval (−∞, x) picks up an additional
contribution of (k − 1)/q from the previous k − 1 intervals in the spectrum. QED
τo(χ[−4,(−1)k+qPθ(x))(ho)),
1
q
It is worth noting that this trace integral can be computed explicitly, the result
involves an integral of arccos(y/2 + cos(θ)).
Appendix 3
We conclude with some final plots of the algebraic curves which determine the
horizontal component of the similarity maps. Properties to notice are that the curves
have even symmetry for denominator q even, odd symmetry for q odd, and in all
cases there is a sequence of nearly linear segments along the diagonal line y = x. It
SPECTRA SELF-SIMILARITY FOR ALMOST MATHIEU OPERATORS
35
is also notably that as the numerator p increases, the algebraic curves become much
more complicated.
Figure 17. Algebraic curves with denominator 5: 1/5 (cid:55)→ 1/7, 2/5 (cid:55)→
2/9, 3/5 (cid:55)→ 3/11, 4/5 (cid:55)→ 4/13. In each graph, there are five nearly
linear segments down the diagonal of the graph. All but the first graph
have some disconnected components.
..
..
-4-2024-4-2024-4-2024-4-2024-4-2024-4-2024-4-2024-4-202436
MICHAEL P. LAMOUREUX, JAMES A. MINGO, AND SYDNEY R. PACHMANN
Figure 18. Algebraic curves with denominator 7: 1/7 (cid:55)→ 1/9, 2/7 (cid:55)→
2/11, 3/7 (cid:55)→ 3/13, 4/7 (cid:55)→ 4/15, 5/7 (cid:55)→ 5/17, 6/7 (cid:55)→ 4/19. In each
graph, there are seven nearly linear segments down the diagonal of the
graph. All but the first graph have some disconnected components.
..
..
-4-2024-4-2024-4-2024-4-2024-4-2024-4-2024-4-2024-4-2024-4-2024-4-2024-4-2024-4-2024SPECTRA SELF-SIMILARITY FOR ALMOST MATHIEU OPERATORS
37
Figure 19. Algebraic curve with denominator 2, the 1/2 (cid:55)→ 1/4 map.
Figure 20. Algebraic curves with denominator 4: 1/4 (cid:55)→ 1/6, 3/4 (cid:55)→
3/10. In each graph, there are four nearly linear segments down the
diagonal of the graph. The right graph has some components which
are disconnected from the linear segments.
..
..
-4-2024-4-2024-4-2024-4-2024-4-2024-4-202438
MICHAEL P. LAMOUREUX, JAMES A. MINGO, AND SYDNEY R. PACHMANN
Figure 21. Algebraic curves with denominator 6: 1/6 (cid:55)→ 1/8, 5/6 (cid:55)→
In each graph, there are six nearly linear segments down the
5/16.
diagonal of the graph. The right graph has some components which
are disconnected from the linear segments.
..
..
-4-2024-4-2024-4-2024-4-2024SPECTRA SELF-SIMILARITY FOR ALMOST MATHIEU OPERATORS
39
Figure 22. Algebraic curves with denominator 8: 1/8 (cid:55)→ 1/10,
3/8 (cid:55)→ 3/14, 5/8 (cid:55)→ 5/18, 7/8 (cid:55)→ 7/22.
In each graph, there are
eight nearly linear segments down the diagonal of the graph. All but
the first graph has some components which are disconnected from the
linear segments.
(M. Lamoureux) Dept. Mathematics and Statistics, University of Calgary, 2500
University Ave NW, Calgary AB T2N 1N4, Canada
E-mail address: [email protected]
(J. Mingo) Dept. Mathematics and Statistics, Queens University, Kingston, ON
Canada
E-mail address: [email protected]
(S. Pachmann) Faculty of Engineering, University of Calgary, Calgary AB, Canada
E-mail address: [email protected]
-4-2024-4-2024-4-2024-4-2024-4-2024-4-2024-4-2024-4-2024 |
1703.03605 | 1 | 1703 | 2017-03-10T10:06:50 | Infinite index extensions of local nets and defects | [
"math.OA",
"math-ph",
"math-ph",
"math.QA"
] | Subfactor theory provides a tool to analyze and construct extensions of Quantum Field Theories, once the latter are formulated as local nets of von Neumann algebras. We generalize some of the results of [LR95] to the case of extensions with infinite Jones index. This case naturally arises in physics, the canonical examples are given by global gauge theories with respect to a compact (non-finite) group of internal symmetries. Building on the works of Izumi, Longo, Popa [ILP98] and Fidaleo, Isola [FI99], we consider generalized Q-systems (of intertwiners) for a semidiscrete inclusion of properly infinite von Neumann algebras, which generalize ordinary Q-systems introduced by Longo [Lon94] to the infinite index case. We characterize inclusions which admit generalized Q-systems of intertwiners and define a braided product among the latter, hence we construct examples of QFTs with defects (phase boundaries) of infinite index, extending the family of boundaries in the grasp of [BKLR16]. | math.OA | math |
Infinite index extensions of local nets and defects
Simone Del Vecchio1 and Luca Giorgetti1
1Dipartimento di Matematica, Universit`a di Roma Tor Vergata,
Via della Ricerca Scientifica, 1, I-00133 Roma, Italy,
[email protected], [email protected]
Abstract
Subfactor theory provides a tool to analyze and construct extensions of Quantum
Field Theories, once the latter are formulated as local nets of von Neumann algebras. We
generalize some of the results of [LR95] to the case of extensions with infinite Jones index.
This case naturally arises in physics, the canonical examples are given by global gauge
theories with respect to a compact (non-finite) group of internal symmetries. Building on
the works of Izumi, Longo, Popa [ILP98] and Fidaleo, Isola [FI99], we consider generalized
Q-systems (of intertwiners) for a semidiscrete inclusion of properly infinite von Neumann
algebras, which generalize ordinary Q-systems introduced by Longo [Lon94] to the infinite
index case. We characterize inclusions which admit generalized Q-systems of intertwiners
and define a braided product among the latter, hence we construct examples of QFTs
with defects (phase boundaries) of infinite index, extending the family of boundaries in
the grasp of [BKLR16].
1 Introduction
The study of extensions in relativistic Quantum Field Theory (QFT) is well-motivated in
several respects. Gauge theory, for instance, provides examples of extensions where a theory
of (anti)commuting fields obeying Bose/Fermi statistics, equipped with a gauge group sym-
metry, contains a subtheory generated by gauge invariant (observable) fields. The former can
be viewed as an extension of the latter, and similarly any intermediate theory gives rise to a
smaller extension of the observable theory. Defects and boundaries can also be described by
extensions, where different types of bulk fields (depending on their relative spacetime localiza-
tion with respect to a certain "defect" line or hypersurface) generate extensions of a common
subtheory which contains, for example, the components of the stress-energy tensor that are
conserved across the boundary. Extensions also appear in classification instances of QFTs,
where all the theories belonging to a certain family share a common subtheory (dictated, e.g.,
by spacetime symmetry), hence the classification problem can be turned into a classification
of extensions. This is the case, for example, in chiral Conformal Field Theory (CFT) where
the Fourier modes of the conformal stress-energy tensor necessarily obey the commutation
relations of the Virasoro algebra at a fixed value of the central charge parameter. Lastly, the
analysis of extensions can be used to construct new examples of QFTs. Starting from some
1
theory, if one can write it as a non-trivial subtheory extended by a certain family of genera-
tors, then new theories can be constructed by suitably manipulating the generators and their
commutation relations, compatibly with locality, and leaving the subtheory untouched.
All of these different situations and problematics have a model-independent and mathe-
matically rigorous formulation in the Algebraic approach to QFT (AQFT ) due to Haag and
Kastler, see [Haa96].
Global gauge theories have been tackled since the early works of [DHR69a], [DHR69b],
[DR72], culminating in [DHR71], [DHR74] and [DR90] in particular, where it is shown that
every theory of local observables arises as gauge group fixed points of a bigger field theory,
obeying (anti)commutation relations and equipped with a (global) gauge group symmetry.
Both the gauge group and the field extension are intrinsically determined by the local ob-
servables (hence dictated by locality, i.e., Einstein's causality). Intermediate extensions of
gauge group fixed points (in 3+1 spacetime dimensions) have been studied in [CDR01]. In
the chiral CFT setting (1 spacetime dimension) gauge group fixed points (also called orbifold
CFTs) appear in [Xu00], [Xu05], [Mug05], and have been generalized to finite hypergroup
fixed points by [Bis17] (generalized orbifold CFTs), where the information about gauge in-
variance contained in the conditional expectation is expressed by an hypergroup action via
completely positive (CP) maps. Intermediate chiral extensions have been analysed by [Lon03]
and [Xu14]. Defects and boundaries have been studied with AQFT methods in recent works
by [BKL15], [BKLR16], [BR16], see also [BKLR15, Ch. 5] where the main mathematical
tools to construct and classify boundary conditions are developed. This analysis of defects
and boundaries in QFT has been our initial motivation for the work presented in this article.
Lastly, again using extensions, the classification of all chiral CFTs with central charge c < 1
has been achieved in [KL04].
In the Haag-Kastler formalism, a (local) quantum field theory is described by a net of
local algebras {O 7→ A(O)}, see [Reh15], [HM06] for self-contained introductions. Local
algebras A(O) are assumed to be von Neumann algebras on the vacuum Hilbert space, and
they typically turn out to be factors (hyperfinite and of type III 1 in the classification of
Connes [Con73], [Haa87]). Hence an extension of QFTs {A ⊂ B} is naturally described by a
family of subfactors indexed by spacetime regions O (e.g., double cones in Minkowski space or
bounded intervals on the line). It was in the work of Longo and Rehren [LR95] (indeed titled
"Nets of subfactors") that it became clear how to use subfactor theory as a tool to classify and
construct extensions in QFT. Their main idea is to exploit the notion of Q-system, due to
[Lon94], for nets of subfactors, in order to relate coherent families of conditional expectations
(which respect the net structure) to coherent families of (dual) canonical endomorphisms
[Lon87] (which turn out to be the restrictions to different spacetime regions O of a unique
global DHR endomorphism θ of {A}). The family of conditional expectations generalizes the
notion of global gauge symmetry, while the DHR endomorphism θ represents the (reducible)
vacuum representation of the bigger theory {B} once restricted to {A}.
Mathematically speaking, the theory of subfactors plays a prominent role in the panorama
of Operator Algebras since the work of Jones [Jon83]. He established a notion of index for
subfactors, which is an invariant (hence opened the way to classification questions) and
surprisingly quantized for values between 1 and 4 (Jones' rigidity theorem). Since then, the
major efforts have been devoted to the study of finite index (finite depth) subfactors and a
complete classification has been achieved for subfactors with index at most 5 + 1
4 [JMS14],
[AMP15], using techniques of [Pop95a] and [Jon99]. At the same time, the analyses of QFT
2
extensions [LR95] and of theories with defects and boundaries [BKLR16] cover the finite
index case only, both being based on the notion of Q-system (which is tightly connected to
the existence of conjugate morphisms ¯ι of the inclusion morphism ι : N ֒→ M for a subfactor
N ⊂ M, hence to the finiteness of the dimension of ι in the sense of [LR97]).
In this article, building on the notion of Pimsner-Popa basis for an inclusion of von
Neumann algebras, see [PP86], [Pop95b], and on the characterization of the canonical endo-
morphisms given by Fidaleo and Isola in [FI99], we reformulate the results on QFT extensions
of [LR95, Sec. 4] in the finite index case, and generalize them to infinite index extensions,
see Section 6. This case naturally appears in physical situations, e.g., if we consider global
gauge theories with respect to a compact non-finite group of internal symmetries.
In order to do so, we first adopt the notion of generalized Q-system, due to [FI99], see
Definition 3.1, and then consider more special generalized Q-systems of intertwiners, see Defi-
nition 3.7 and 5.8. The latter can be thought, roughly speaking, as C ∗ Frobenius algebra-like
objects with possibly infinitely many comultiplications, see Remark 3.8, 5.7 and cf. [BKLR16,
Sec. 3.1].
Any semidiscrete inclusion of (properly infinite, with separable predual) von Neumann
algebras N ⊂ M, i.e., an inclusion endowed with a faithful normal conditional expectation
E : M → N , admits a generalized Q-system. Vice versa, from any generalized Q-system one
can (re)construct the bigger algebra M and the conditional expectation E, see [FI99, Thm.
4.1]. An advantage of using generalized Q-systems (in the finite index case as well) is that
no factoriality or irreducibility assumption on the inclusion is needed along the way. This
enhanced flexibility is particularly desirable in the study of boundary conditions, see com-
ments after [BKLR16, Thm. 4.4], where non-irreducible, non-factorial extensions necessarily
appear. On the other hand, generalized Q-systems (in the infinite index case) dwell a bit
further away from the purely categorical setting of their finite index counterpart.
Given a semidiscrete inclusion of von Neumann algebras N ⊂ M (where N is an infinite
factor), we show that the existence of a generalized Q-system with the additional intertwining
property is actually equivalent to the discreteness of the inclusion in the sense of [ILP98] (but
admitting non-irreducible extensions), see Section 5. This characterization relies on strong
results of [ILP98] and [FI99], and can be physically interpreted by saying that a semidiscrete
extension is discrete if and only if it is generated by charged fields, in the sense of [DR72].
These are elements ψ ∈ M which generate from the vacuum a non-trivial (irreducible) sub-
sector ρ ≺ θ of the dual canonical endomorphism θ, i.e., ψn = ρ(n)ψ for every n ∈ N .
In Section 6 it is shown that generalized Q-systems of intertwiners indeed induce discrete
(finite or infinite index) extensions of QFTs in the sense of [LR95]. Two different ways to
obtain the construction are provided: one is a direct generalization of [LR95, Thm. 4.9], the
other one exploits an inductive procedure which is somewhat more suitable to be used for
the analysis of braided products and boundary conditions in the subsequent sections.
In Section 7, we give a general proof of covariance of QFT extensions constructed from
covariant nets of local observables. This fact is apparently well known to experts, and clear
in many examples, but we could not find a general statement in the literature (on finite
index extensions). The key ingredient in our proof is the equivariance of the action of the
spacetime symmetry group on the DHR category. More precisely, the mere existence of
covariance cocycles, see Definition 7.3, is not sufficient to guarantee covariance. One needs
in addition naturality and tensoriality properties of the cocycles.
Given two generalized Q-systems of intertwiners (in a C ∗ braided tensor category), one
3
can easily define their braided product in analogy with the case of ordinary Q-systems, see
Definition 4.1 and cf. [EP03, Sec. 3], [BKLR16, Sec. 4.9]. In Section 4, we prove the non-
trivial statement that the braided product of two generalized Q-systems of intertwiners is
again a generalized Q-system of intertwiners, i.e., that the analytical properties defining
a Pimsner-Popa basis (as a part of the definition of a generalized Q-system) behave well
with respect to the categorical notions of naturality and tensoriality of a braiding in a C ∗
tensor category. Thus, in the QFT setting, we can define the braided product of nets of
local observables and construct new examples of irreducible phase boundary conditions with
infinite index (infinitely many bulk fields) by taking the direct integral decomposition of the
braided product net with respect to its center, see Section 8 and 9. On the other hand,
we leave open the questions about universality of the braided product construction and the
classification of boundary conditions in the infinite index case, cf. [BKLR16, Sec. 5].
In Section 10, we work out examples of infinite index (discrete) extensions of the chiral
U (1)-current algebra [BMT88] and explicitly compute their braided products. These exam-
ples show an important difference with the analysis of boundary conditions in the finite index
case, namely the center of the braided product may be a continuous algebra, i.e., with no
non-trivial minimal projections, hence the irreducible boundary conditions constructed by
direct integral decomposition need not be representations of the braided product itself.
Notation-wise we work with nets of local algebras {O 7→ A(O)} indexed by partially
ordered and directed sets of spacetime regions K, in order to formulate our results, when
possible, for arbitrary spacetime dimensions, e.g., in 1D theories on the line, 1+1D or 3+1D
theories in Minkowski space.
2 Pimsner-Popa bases
E
Let N
⊂ M be a unital inclusion of von Neumann algebras with a normal faithful conditional
expectation E : M → N . Assume that M acts standardly on a separable Hilbert space H
and let N ⊂ M ⊂ M1 := hM, eNi 1 be the Jones basic construction [Jon83], see also [Pop95b,
Sec. 1.1.3], [LR95, Sec. 2.2]. Up to spatial isomorphism it can be characterized as follows. Let
Ω ∈ H be a cyclic and separating vector for M such that the induced (normal faithful) state
ω of M is invariant under E, i.e., ω ◦ E = ω, and set eN := [N Ω], the orthogonal projection
on the subspace H0 := N Ω ⊂ H. The projection eN ∈ N ′ ∩ M1 is the Jones projection of
N ⊂ M with respect to E, and implements E in the sense that E(m)eN = eN meN , m ∈ M.
Moreover, it is uniquely determined up to conjugation with unitaries in M′ [Kos89, App. I].
Definition 2.1. [PP86], [Pop95b]. A Pimsner-Popa basis for N
⊂ M is a family of
elements {Mi} ⊂ M, where i runs in some set of indices I, such that
E
(i) Pi := M ∗
i eN Mi are projections in M1 which are mutually orthogonal, i.e., PiP ∗
i = Pi
and PiPj = δi,jPi for every i, j ∈ I.
(ii) Pi Pi = 1, where the sum converges (unconditionally) in the strong operator topology.
For future reference, we mention the following equivalent characterization of the algebraic
properties of Pimsner-Popa bases, see [Pop95b, Sec. 1.1.4].
1Here hSi denotes the von Neumann algebra generated by a subset S ⊂ B(H). For a pair of subsets
S1, S2 ⊂ B(H) we also denote hS1, S2i by S1 ∨ S2.
4
Lemma 2.2. In the notation of Definition 2.1, the conditions (i) and (ii) are respectively
equivalent to
(i)′ qi := E(MiM ∗
i ) are projections in N (not necessarily mutually orthogonal) and E(MiM ∗
j ) =
0 for every i 6= j, i, j ∈ I.
i eNH = H in the Hilbert space topology.
(ii)′ Pi M ∗
Proposition 2.3. [Pop95b]. If {Mi} is a Pimsner-Popa basis for N
has the following expansion
m =Xi
M ∗
i E(Mim)
E
⊂ M then every m ∈ M
unconditionally convergent in the topology generated by the family of seminorms {k · kϕ : ϕ ∈
(M∗)+, ϕ = ϕ ◦ E}, with kmkϕ := ϕ(m∗m)1/2.
The expansion is unique if and only if E(MiM ∗
i ) = 1 for every i ∈ I.
Remark 2.4. In view of the proposition above, Pimsner-Popa bases {Mi}, or better their
adjoints {M ∗
i } can be seen as bases for M as a right pre-Hilbert N -module with the N -
valued inner product (M ∗
j ) := E(MiM ∗
j ).
i M ∗
The cardinality of a Pimsner-Popa basis {Mi} is not a invariant for N
⊂ M. Indeed, by
the following cutting and gluing procedures [Pop95b, Sec. 1.1.4] we obtain other Pimsner-
Popa bases:
E
i ∈ N such thatPj aj
i aj∗
i = E(MiM ∗
i ),
(1) If, for each i, we take a set of partial isometries aj
then {aj∗
i Mi} is also a basis.
j ) and E(MkM ∗
(2) If E(Mj M ∗
k ) are orthogonal, then we can replace the pair Mj, Mk in
{Mi} by Mj + Mk and we still get a basis.
The good notion of dimension of M as an N -module is given by the Jones index of the
inclusion N ⊂ M with respect to E, [Jon83], [Kos86]. This guiding idea is supported by
the following theorem due to [PP86, Prop. 1.3], [BDH88, Thm. 3.5], [Pop95b, Thm. 1.1.5,
1.1.6], which characterizes the finiteness of the index (and computes its value) by means of
Pimsner-Popa bases.
E
Theorem 2.5. [Pop95b]. N
⊂ M has finite Jones index if and only if it has a Pimsner-
Popa basis {Mi} such that Pi M ∗
i Mi is ultraweakly convergent in M. In this case, Pi M ∗
i Mi
belongs to the center of M, it holds
Xi
i Mi = Ind(N
⊂ M)
M ∗
E
E
where Ind(N
Pimsner-Popa basis.
⊂ M) denotes the Jones index of E, and the same is true for any other
If in addition N is properly infinite, then N
⊂ M has finite Jones index if and only if it
has a Pimsner-Popa basis made of one element {M}. Moreover, M can be chosen such that
E(M M ∗) = 1.
E
5
We are mainly interested in inclusions of properly infinite von Neumann algebras (with
separable predual), due to their appearance in QFT, see, e.g., [Kad63], [Lon79].
In this
setting, with no finite index or factoriality assumptions, it was shown by Fidaleo and Isola
[FI99, Thm. 3.5] that Pimsner-Popa bases made of elements of M always exist.
Proposition 2.6. [FI99]. Every inclusion N
⊂ M of properly infinite von Neumann algebras
with a normal faithful conditional expectation E : M → N admits a Pimsner-Popa basis
{Mi} ⊂ M in the sense of Definition 2.1.
E
3 Infinite index and generalized Q-systems (of intertwiners)
Q-systems were introduced by R. Longo in [Lon94, Sec. 6]. They provide a way to alge-
braically characterize infinite subfactors N ⊂ M with finite index together with a normal
faithful conditional expectation E : M → N by means of data pertaining to the smaller
factor N . The main technical tool to achieve this characterization is the notion of canonical
endomorphism [Lon87] for the inclusion N ⊂ M, namely the homomorphism γ : M → N
defined by γ := (jN jM)↾M, where jN := AdJN ,Φ, jM := AdJM,Φ and JN ,Φ, JM,Φ are respec-
tively the modular conjugations of N , M with respect to a cyclic and separating vector Φ for
N and M. From a categorical perspective, a Q-system is a special C ∗ Frobenius algebra in
a strict C ∗ tensor category C with simple unit, cf. [BKLR15, Def. 3.8]. In the more concrete
case of subfactors, the category is C = End0(N ), whose objects are the endomorphisms of
the factor N with finite dimension in the sense of [LR97].
Here we recall and analyze the more general notion of generalized Q-system, introduced
by F. Fidaleo and T. Isola in [FI99, Sec. 5] for a possibly infinite index (semidiscrete or
semicompact) inclusion of properly infinite von Neumann algebras. We then introduce the
more special notion of generalized Q-system of intertwiners that will be the fundamental
object in the subsequent sections, in particular for the applications to QFT.
Let N ⊂ M be a unital inclusion of properly infinite von Neumann algebras on a separable
Hilbert space H. Denote by C(M,N ) and E(M,N ) respectively the set of all normal and
normal faithful conditional expectations of M onto N . We call the inclusion N ⊂ M
semidiscrete if E(M,N ) 6= ∅, and semicompact if E(N ′,M′) 6= ∅, or equivalently if
E(M1,M) 6= ∅ or E(N ,N1) 6= ∅, where N1 ⊂ N ⊂ M ⊂ M1 denotes the tower of von
Neumann algebras obtained by canonical extension and restriction of the original inclusion
[LR95, Sec. 2.5 A]. The terminology is adopted from [FI99], [ILP98], [FI95], [HO89]. Recall
that a finite index inclusion is both semidiscrete and semicompact, see e.g. [Lon90, Prop.
4.4].
Let End(N ) be the collection of normal faithful unital *-endomorphisms of N . The
following notion is tailored to describe semidiscrete inclusions of von Neumann algebras
N ⊂ M with E ∈ E(M,N ), possibly of infinite index.
Definition 3.1. [FI99]. Let N be a properly infinite von Neumann algebra. A generalized
Q-system in C = End(N ) is a triple (θ, w,{mi}) consisting of an endomorphism θ ∈ End(N ),
an isometry w ∈ HomEnd(N )(id, θ) (i.e., wn = θ(n)w, n ∈ N ), and a family {mi} ⊂ N indexed
by i in some set I, such that
(i) pi := m∗
i ww∗mi are mutually orthogonal projections in N , i.e., pip∗
j = δi,jpi, such that
Pi pi = 1. ("Pimsner-Popa condition")
6
(ii) nw = 0 ⇒ n = 0 if n ∈ N1 := hθ(N ),{mi}i. ("faithfulness condition")
Remark 3.2. An analogous definition of generalized Q-system in End(N ), involving an isom-
etry x ∈ HomEnd(N )(θ, θ2) instead of w ∈ HomEnd(N )(id, θ), can be given in the semicompact
case, see [FI99, Sec. 5]. We shall however be interested in extensions N ⊂ M with a (nor-
mal faithful) conditional expectation E ∈ E(M,N ) as they arise in QFT when N = A(O),
M = B(O) are local algebras (relative to some spacetime region O) and {A ⊂ B} is an
extension of a net of local observables {A} by means of a "field net" {B}. Here E generalizes
the notion of an average over a global gauge group action on fields, giving the observables as
the gauge invariant part.
Theorem 3.3. [FI99]. Let N be a properly infinite von Neumann algebra with separable
predual and θ ∈ End(N ). Then the following are equivalent
(1) There is a von Neumann algebra N1 such that N1 ⊂ N with E′ ∈ E(N1,N2) 6= ∅, where
N2 := θ(N ) ⊂ N1, and θ is a canonical endomorphism for N1 ⊂ N .
(2) There is a von Neumann algebra M such that N ⊂ M with E ∈ E(M,N ) 6= ∅, and θ
is a dual canonical endomorphism for N ⊂ M, i.e., θ = γ↾N where γ ∈ End(M) is a
canonical endomorphism for N ⊂ M.
(3) The endomorphism θ is part of a generalized Q-system in End(N ), (θ, w,{mi}), see
Definition 3.1.
Proof. We may assume that N is in its standard representation on H. The equivalence of
(1) and (2) is then obtained by canonical extension and restriction [LR95, Sec. 2.5 A]. The
tower of von Neumann algebras reads
. . . ⊂ N2 = θ(N )
E′
⊂ N1 = hθ(N ),{mi}i
θ
⊂ N
E,γ
⊂ M ⊂ . . .
(1)
where N2 ⊂ N1 = γ(N ⊂ M) is a spatial isomorphism of inclusions and the relation
E′ ◦ γ = γ ◦ E on M gives a bijection between E(N1,N2) and E(M,N ).
The equivalence of (1) and (3) is due to [FI99, Thm. 4.1]. In particular, they show that
eN2 := ww∗ is a Jones projection for the inclusion N2 ⊂ N1 with respect to E′ := θ(w∗·w) and
that N = hN1, eN2i is the associated Jones extension. Hence the condition (i) in Definition
3.1 says that {mi} ⊂ N1 is a Pimsner-Popa basis for N2 ⊂ N1 with respect to E′. The
condition (ii) in Definition 3.1 is nothing but faithfulness of E′.
i w is a projection, which is equivalent to ww∗mim∗
Remark 3.4. The condition that the pi in Definition 3.1 are (mutually orthogonal) projections
j = δi,jpi, does not enter in the proof of (3) ⇒ (1) of Theorem 3.3, onlyPi pi = 1
in N , i.e., pip∗
i ww∗mi is a projection if and
is relevant there. We can however always assume it because m∗
i )ww∗ is a
only if w∗mim∗
i ) is a projection, because ww∗ = eN2 and n 7→ neN2 is an isomor-
projection, i.e., E′(mim∗
phism of N2 onto N2eN2. Hence we can apply a Gram-Schmidt orthogonalization procedure
to the {mi} with respect to the operator-valued inner product (mimj) := E′(mjm∗
i ) and
choose another basis { mi} such that E′( mj m∗
Remark 3.5. Notice that no factoriality Z(N ) = C1, Z(M) = C1, nor irreducibility N ′ ∩
M = C1 assumptions enter in the proof of Theorem 3.3, [FI99]. In the case of non-irreducible
finite index subfactors, E is not necessarily the minimal conditional expectation, see [Hia88],
[Lon89, Sec. 5].
i ww∗ = E′(mim∗
i ) = δi,j 1.
7
Proposition 3.6. [FI99]. Let N ⊂ M a semidiscrete inclusion of properly infinite von
Neumann algebras and let γ be a canonical endomorphism. The following are equivalent
(1) N ⊂ M is irreducible in the sense that N ′ ∩ M = Z(N ).
(2) E(M,N ) contains only one element.
(3) HomEnd(N )(id, θ) is cyclic as a Z(N )-module, where θ = γ↾N .
We now specialize the notion of generalized Q-system (Definition 3.1) by requiring an
additional intertwining property of the Pimsner-Popa elements.
Definition 3.7. Let N be a properly infinite von Neumann algebra. We call (θ, w,{mi}) a
generalized Q-system of intertwiners in C = End(N ) if, in addition to the properties of
Definition 3.1, it satisfies mi ∈ HomEnd(N )(θ, θ2) (i.e., miθ(n) = θ2(n)mi, n ∈ N ) for every
i ∈ I.
In this case we can use string diagrams to denote w and mi as follows
w =
id
•
θ
= •
,
mi =
θ
• i
, i ∈ I.
θ
θ
θ
At this point, a comparison between the notions of generalized Q-system and "ordinary"
Q-system in the finite index setting is due.
Remark 3.8. (The finite index case). An infinite subfactor N ⊂ M with E ∈ E(M,N )
can be characterized by an "ordinary" Q-system (θ, w, x) if and only if the Jones index of E is
finite, see [Lon94], [LR95, Sec. 2.7]. The algebraic relations defining a Q-system in End0(N )
read as follows: θ ∈ End0(N ), w ∈ HomEnd0(N )(id, θ), x ∈ HomEnd0(N )(θ, θ2) and
x∗x ∈ C1.
xx∗ = x∗θ(x) = θ(x∗)x,
w∗x = θ(w∗)x = 1,
x2 = θ(x)x,
The conditions in the line above are called respectively unit property, associativity, Frobenius
property and specialness, see [BKLR15, Def. 3.8]. It is known that the Frobenius property
is a consequence of the other properties [LR97], [BKLR15, Lem. 3.7] and that specialness is
not needed to construct the extension N2 = θ(N ) ⊂ N1, i.e., N ⊂ M [BKLR15, Rmk. 3.18].
Moreover, it is an easy exercise to check that ordinary Q-systems are also generalized
Q-system of intertwiners with {mi} = {x} (up to a normalization of w and x), in the sense
Indeed the Pimsner-Popa condition x∗ww∗x = 1 follows by w∗x = 1,
of Definition 3.7.
and the faithfulness condition nw = 0 ⇒ n = 0, n ∈ N1 = hθ(N ), xi follows because
hθ(N ), xi = θ(N )x = x∗θ(N ) hold, due to θ(w∗)x = 1, associativity and Frobenius property.
On the other hand, a finite index inclusion of infinite factors N2 ⊂ N1 with normal faithful
conditional expectation E′(·) = θ(w∗ · w), always has a Pimsner-Popa basis of one element,
m ∈ N1, such that E′(mm∗) = 1 by Theorem 2.5. The triple (θ, w, m) is a generalized
Q-system in the sense of Definition 3.1. The characterizing properties
m∗ww∗m = 1, w∗mm∗w = 1
8
are a weaker form of the unit property for ordinary Q-systems, and the Pimsner-Popa expan-
sion of Proposition 2.3 gives in particular
m2 = E′(m2m∗)m, mm∗ = E′(m(m∗)2)m.
If we assume the unit property w∗m = 1 to hold, we get back the associativity m2 = θ(m)m
and the Frobenius property mm∗ = θ(m∗)m.
If (θ, w,{mi}) is a generalized Q-system (of intertwiners) in C = End(N ), consider the
tower of von Neumann algebras
. . . ⊂ N2
E′
⊂ N1
θ
⊂ N
E,γ
⊂ M
γ1
⊂ M1 ⊂ . . .
as in equation (1), where the Jones extension M1 = hM, eNi of N ⊂ M with respect to
E coincides with the canonical extensions, namely hM, eNi = jM(N ′), see [LR95, Sec. 2.5
D], [Lon89, Sec. 3]. Here Ω is a cyclic and separating vector for M as in Section 2 and
jM = AdJM,Ω is the associated modular conjugation. Moreover, θ and γ1 are canonical
−1(eN2) = eN
endomorphisms dual to γ, hence θ = γ↾N , γ = γ1↾M. Then γ1
and {Mi := γ−1(mi)} ⊂ M clearly forms a Pimsner-Popa basis for N ⊂ M with respect to
E.
Definition 3.9. We call (γ, w,{Mi}) a generalized Q-system (of intertwiners) dual to
(θ, w,{mi}). The intertwining relation mi ∈ HomEnd(N )(θ, θ2) is equivalent to Min = θ(n)Mi,
n ∈ N .
−1(ww∗) = γ1
4 Braided products
Suppose additionally that two generalized Q-systems of intertwiners are composed of data
belonging to a certain braided tensor subcategory of End(N ), we can consider their braided
product as follows
Definition 4.1. Let N be a properly infinite von Neumann algebra and C ⊂ End(N ) a
C ∗ braided tensor subcategory of End(N ). Let (θA, wA,{mA
j }) two
generalized Q-systems of intertwiners in C (Definition 3.7), indexed respectively by i ∈ I and
j ∈ J. We call
i1}) and (θB, wB,{mB
the braided product of (θA, wA,{mA
where
ε mB
j
i ×±
mA
(θAθB, wAwB,{mA
i ×±
ε mB
j })
i }) and (θB, wB,{mB
j }), indexed by (i, j) ∈ I × J,
:= θA(ε±
i θA(mB
j )
θA,θB )mA
depending on the ± choice. Here ε+ = ε and ε− = εop denote respectively the braiding of C
and its opposite. Equivalently
wAwB = •
•
,
θA
θB
i ×+
mA
ε mB
j =
θA
θB
i
•
j
•
θA
θB
θA
θB
, (i, j) ∈ I × J
and similarly for mA
i ×−
ε mB
j .
9
Surprisingly, the analytic conditions dictated on generalized Q-systems by subfactor the-
ory (e.g. the property characterizing a Pimsner-Popa basis) turn out to be naturally compati-
ble with the categorical notion of braiding in a tensor category of endomorphisms. Indeed we
have the following proposition which extends the braided product construction, see [BKLR15,
Sec. 4.9], to the infinite index case.
Proposition 4.2. The braided product of two generalized Q-systems of intertwiners is again
a generalized Q-system of intertwiners.
Proof. The intertwining properties appearing in Definition 3.7 are easily checked once we
write the operators wAwB and mA
j , (i, j) ∈ I × J as tensor products and compositions
of arrows in the braided tensor category of endomorphisms C ⊂ End(N ), as in the case of
ordinary Q-systems [BKLR15, Def. 4.30].
The Pimsner-Popa condition (i) in Definition 3.1 is more lengthy to check. For each i ∈ I
i ×±
ε mB
and j ∈ J, let
= θA(mB∗
ε mB
:= (mA
pAB,±
i,j
i θA((ε±
j )mA∗
j )mA∗
= θA(mB∗
i ×±
j )∗wAwBwB∗wA∗(mA
i ×±
θA,θB )∗)θA(wB)wAwA∗θA(wB∗)θA(ε±
i θA(θA(wB))wAwA∗θA(θA(wB∗))mA
ε mB
j )
θA,θB )mA
i θA(mB
j )
i θA(mB
j )
θA,θB )∗wB = θA(wB)(ε±
because (ε±
tensor category C, or of its opposite ε− = εop, and ε±
i wAwA∗mA
j )θA(wB)mA∗
= θA(mB∗
θA,id = 1 by convention. Moreover
i θA(wB∗)θA(mB
j )
θA,id)∗ by naturality of the braiding ε+ = ε in the braided
= mA∗
i wAwA∗mA
i θA(mB∗
j wBwB∗mB
j )
hence we have shown
(2)
where pA
j , j ∈ J are the projections appearing in Definition 3.1 respectively
for the two generalized Q-systems. Equation (2) is much more effectively expressed using
graphical calculus
i , i ∈ I and pB
j ) = θA(pB
j )pA
i
pAB,±
i,j = pA
i θA(pB
θA
θB
•
•
i
•
•
•
• i
θA
Now one can easily check that pAB,±
1.
i,j
j
•
• j
θB
=
•
•
θA
i
•
• i
θA
θB
j
•
• j
θB
•
•
.
are mutually orthogonal projections which sum up to
1 ×±
ε N B
The faithfulness condition (ii) in Definition 3.1 follows from Lemma 4.3 below. Indeed, let
1 (see below), then nwAwB = nθA(wB)wA = 0 if and only if nθA(wB) = 0 since
θA,θB )∗wB = 0
1 we have that n = 0 and
n ∈ N A
1 ×±
N A
by naturality of the braiding. Since Ad(ε±
the proof is complete.
1 . Now nθA(wB) = 0 if and only if ε±
θA,θB )(N A
θA,θB nθA(wB) = ε±
1 ×±
ε N B
1 ⊂ N A
1 ) ⊂ N B
θA,θB n(ε±
ε N B
10
Lemma 4.3. In the notation of Definition 4.1, consider the two towers of von Neumann
1 ⊂ N ⊂ MB respectively associated to the
algebras N A
two generalized Q-systems (of intertwiners) as in Theorem 3.3. Let
1 ⊂ N ⊂ MA and N B
2 ⊂ N B
2 ⊂ N A
then
1 ×±
N A
ε N B
1 :=(cid:10)θAθB(N ),{mA
θA,θB )(N A
1 , Ad(ε±
ε mB
j }(cid:11)
1 ) ⊂ N B
ε N B
1 .
i ×±
1 ×±
1 ×±
N A
ε N B
1 ⊂ N A
Proof. The first inclusion follows from the very definitions. To show the second observe that
Ad(ε±
θA,θB )(θAθB(N )) = θBθA(N ) ⊂ θB(N ). Hence it is enough to check that
θA,θB θAθB(ε±
ε±
θA,θB )θA(ε±
θA,θB )mA
i θA(mB
j )(ε±
θA,θB )∗ = θB(ε∓
θB,θA)mB
j θB(mA
i ),
but this follows from repeated application of naturality and tensoriality of the braiding
mA
i θA(mB
j )(ε±
θA,θB )∗ = θAθA(mB
j )mA
i ε∓
θB ,θA
where ε∓
θBθB ,θAθA = θA(ε∓
θB,θA)θAθB(ε∓
= θAθA(mB
j )ε∓
θB ,θAθAθB(mA
θB ,θA)ε∓
i ) = ε∓
θB ,θAθB(ε∓
θB ,θA).
θBθB ,θAθAmB
j θB(mA
i )
Corollary 4.4. (of Proposition 4.2). θAθB ∈ End(N ) is a canonical endomorphism for the
inclusion N A
ε N B
1 ×±
1 ⊂ N . Moreover, the inclusion
1 ×±
θAθB(N ) ⊂ N A
ε N B
1
(3)
is semidiscrete 2 with (normal faithful) conditional expectation given by
EAB ′
:= θAθB(wB∗wA∗ · wAwB).
Denote by
MA ×±
the von Neumann algebra appearing in the tower
ε MB
. . . ⊂ θAθB(N ) ⊂ N A
1 ×±
ε N B
1
θAθB
⊂ N
γAB
⊂ MA ×±
ε MB ⊂ . . .
ε N B
ε MB dual to θAθB. By definition, γAB
obtained as in Theorem 3.3 from the braided product Q-system. We call it the braided
product of MA and MB. Here γAB denotes a canonical endomorphism for the inclusion
N ⊂ MA ×±
ε MB) =
1 ×±
1 . Similarly, γA and γB are respectively canonical endomorphisms dual to θA and
N A
θB.
In order to show that the braided product MA ×±
ε MB actually contains MA and MB
as subalgebras (see Proposition 4.5 below) we need to consider generalized Q-systems of
intertwiners with an additional property, which is a weaker version of the unit property in
ordinary Q-systems, namely θ(w∗)x = 1, cf. [BKLR16, Prop. 4.12]. We shall come back to
this property in the next section, see Proposition 5.5 and Definition 5.8.
↾N = θAθB and γAB(MA ×±
2By the results of the next section, the inclusion (3) is also discrete in the sense of Definition 5.1.
11
Proposition 4.5. In the notation of Definition 4.1, let (θA, wA,{mA
fulfill in addition θA(wA∗)mA
and 0 ∈ J. Then the maps
0 = 1 and θB(wB∗)mB
i }) and (θB, wB,{mB
j })
0 = 1 for two distinguished labels 0 ∈ I
A : MA → MA ×±
ε MB,
B : MB → MA ×±
A := (γAB)−1 ◦ Ad(ε±
θA,θB )∗ ◦ θB ◦ γA
ε MB,
B := (γAB)−1 ◦ θA ◦ γB
are embeddings respectively of MA and MB into MA ×±
ε MB. Call ιA : N → MA and
ιB : N → MB the embeddings of N into MA and MB respectively. Then the two embeddings
of N into the braided product coincide, i.e.
A ◦ ιA = B ◦ ιB,
i and M B
the commutation relations among M A
MA ×±
ε MB are given by
B(M B
i )B(M B
j ),
Moreover, MA and MB generate the braided product, i.e.
θA,θB A(M A
j )A(M A
i ) = ε±
i ∈ I, j ∈ J.
j , as in Definition 3.9, in the braided product
(4)
(5)
(6)
(7)
Proof. We show first that
MA ×±
Ad(ε±
ε MB =(cid:10)A(MA), B(MB)(cid:11) .
ε N B
θA,θB )∗ ◦ θB(N A
1 ) ⊂ N A
θA(N B
1 ) ⊂ N A
1 ×±
ε N B
1 ×±
1
1
(8)
ε MB. For the inclusion (8)
0 = 1, hence
from which it is clear that A and B are embeddings into MA ×±
it is enough to show that θA(mB
θA(mB
j ) = θA(wA∗)mA
j ) ∈ N A
1 ×±
0 θA(mB
1 . By assumption θA(wA∗)mA
0 θA(mB
j )
ε N B
j ) = θAθB(wA∗)θA(ε±
θA,θB )mA
0 ×±
1 ×±
ε N B
j ∈ N A
ε mB
1
= θAθB(wA∗)mA
using naturality of the braiding and ε±
id,θB = 1. For the inclusion (7), it is enough to observe
1 , cf. [BKLR15, Sec. 4.9],
that Adε∓
and consider the previous case interchanging A with B and ± with ∓. Now, (4) is clear. To
show the commutation relations among M A
j apply first γAB to equation (5). The
r.h.s. then reads
is an isomorphism between N A
1 and N B
i and M B
1 ×∓
ε N B
ε N A
1 ×±
θA,θB
θAθB(ε±
θA,θB )(ε±
θA,θB )∗θB(mA
i )ε±
= θAθB(ε±
θA,θB )(ε±
θA,θB )∗ε±
θAθA,θB mA
θA,θB θA(mB
j )
i θA(mB
j )
= θAθB(ε±
θA,θB )θA(ε±
θA,θB )mA
i θA(mB
j ).
Similarly, one can compute the l.h.s., namely
j )(ε±
θA(mB
θA,θB )∗θB(mA
By the intertwining property θAθA(mB
we have equation (5). In the previous computations we have shown in particular that
θA,θBθB )θAθA(mB
j ) and by tensoriality of the braiding
θA,θB ) = θA(ε±
i = mA
i θA(mB
j )mA
i .
i )(ε±
j )mA
A(M A
i )B(M B
j ) = (γAB)−1(mA
i ×±
ε mB
j )
from which equation (6) follows.
12
5 The case of discrete inclusions
Generalized Q-systems with the additional intertwining property mi ∈ HomEnd(N )(θ, θ2) as
in Definition 3.7 can be constructed whenever the inclusion N ⊂ M is discrete (see Definition
5.1 below, cf. [ILP98, Def. 3.7]). The main idea is to look first at elements ψρ ∈ M which
generate on N subendomorphisms ρ ≺ θ of the dual canonical endomorphism θ ∈ End(N )
of N ⊂ M from the vacuum (identity representation), namely such that
ψρn = ρ(n)ψρ, n ∈ N .
Such elements are called charged fields after the work of [DR72] in QFT. In the subfactor
setting they can be constructed as in [ILP98, Prop. 3.2]. We generalize the latter construction
to the case of non-irreducible, non-factorial extensions (as one needs in the study of defects in
QFT, see [BKLR16, Thm. 4.4]), and we show how charged fields can be used, in the discrete
case, to define generalized Q-systems of intertwiners. Moreover, we show that a semidiscrete
inclusion admitting a generalized Q-system of intertwiners is necessarily discrete.
Consider an inclusion N ⊂ M, where N is an infinite factor and M is a properly infinite
If E(M,N ) 6= ∅ denote by E ∈
von Neumann algebra on a separable Hilbert space H.
P (M1,M) the normal semifinite faithful operator-valued weight dual to E ∈ E(M,N ), see
[Kos86], [ILP98], [FI99].
Definition 5.1. [ILP98]. In the above notation, the inclusion N ⊂ M is called discrete
if E(M,N ) 6= ∅ (semidiscreteness) and E↾N ′∩M1 is semifinite for some (hence for all) E ∈
E(M,N ).
Proposition 5.2. Let N be an infinite factor with separable predual. Then a semidiscrete
extension N ⊂ M can be characterized as in Theorem 3.3 by a generalized Q-system of
intertwiners in C = End(N ) (Definition 3.7) if and only if it is discrete.
Proof. We begin with necessity. Let (θ, w,{mi}) be a generalized Q-system of intertwiners
in C = End(N ) and consider the dual generalized Q-system of intertwiners (γ, w,{Mi}) as in
Definition 3.9. By definition M ∗
i eN Mi are mutually orthogonal projections in M1 = hM, eNi
which Pi M ∗
i eN Mi = 1. On one hand, MeNM ⊂ m E, where m E denotes the domain of
E, because E(eN ) = 1 by [Kos86, Lem. 3.1]. On the other hand, M ∗
i eN Mi ∈ N ′ ∩ M1
by the intertwining property of the Mi on N . Hence M ∗
i eN Mi are mutually orthogonal
projections which sum up to 1 in the domain of E↾N ′∩M1 ∈ P (N ′ ∩ M1,N ′ ∩ M). This
is equivalent to semifiniteness of E↾N ′∩M1 by [FI99, Lem. 3.2], see also [HKZ91, Lem. 2.2],
hence to discreteness of N ⊂ M. The same is true if N ⊂ M is an arbitrary semidiscrete
inclusion of von Neumann algebras with separable predual.
The converse implication relies on deep results on the structure of N ′∩M1 due to [ILP98].
Consider a discrete inclusion N ⊂ M where N is a factor, M a von Neumann algebra, and
choose E ∈ E(M,N ). Then M1 = hM, eNi is a factor and N ⊂ M1 a subfactor. By the
same argument leading to [ILP98, Prop. 2.8] we get a decomposition of N ′ ∩ M1 as a direct
sum of four algebras, where only the first survives by discreteness assumption and because
AdJM,Ω(N ′∩M1) = N ′∩M1, cf. comments after [ILP98, Def. 3.7]. In particular N ′∩M1 is a
direct sum of type I factors and PN ⊂ PM1P has finite index for every finite rank projection
P ∈ N ′ ∩M1 by [ILP98, Lem. 2.7 (ii)]. Now, arguing as in the proof of [FI99, Thm. 3.5] and
using [FI99, Lem. 3.2], see also [ILP98, Prop. 3.2 (ii) ⇒ (i)], by discreteness we can write
13
1 =Pi Pi, i ∈ I, where Pi ∈ N ′ ∩ M1 are non-trivial mutually orthogonal projections such
that Pi ∈ m E. Each Pi gives rise to a subendomorphism of the dual canonical endomorphism
θ ∈ End(N ) of N ⊂ M. Indeed, Pi and eN are infinite projections in M1 because N is an
infinite factor, cf. [FI99, Lem. 3.1], hence we can choose partial isometries Wi ∈ M1 such
that W ∗
i Wi = Pi, WiW ∗
i = eN . Then
WinW ∗
i = ρi(n)eN , n ∈ N
defines ρi ∈ End(N ), ρi ≺ θ, because eNM1eN = N eN , cf. [ILP98, Lem. 3.1]. The
endomorphism ρi has finite index, i.e., finite dimension [LR97], whenever Pi has finite rank
in N ′ ∩ M1, indeed the inclusion PiN ⊂ PiM1Pi is isomorphic to ρi(N ) ⊂ N . Moreover,
θ = ⊕iρi. From Wi = eN WiPi we get WiPi ∈ n E because n E is a left ideal, and Wi ∈ m E =
n∗
n E. By the push-down lemma [ILP98, Lem. 2.2] generalized to non-factorial inclusions
E
[FI99, Lem. 3.3] we can write Wi = eN ψi, where ψi ∈ M, ψi := E(Wi). One can check that
ψi is a charged field for ρi, indeed
ψin = E(WiPin) = E(WinPi) = E(ρi(n)eN Wi) = ρi(n)ψi, n ∈ N
i ) = eN ψiψ∗
i eN = WiW ∗
i = eN , hence E(ψiψ∗
and that eN E(ψiψ∗
i ) = 1 because n 7→ neN
is an isomorphism of N onto N eN . Moreover Pi = ψ∗
i eN ψi, hence {ψi} is a Pimsner-Popa
basis for N ⊂ M with respect to E. In particular, M = hN ,{ψi}i as in the proof of [FI99,
Thm. 4.1], see also [Pop95b, Sec. 1.1.4], [ILP98, Lem. 3.8].
Now, chosen a canonical endomorphism γ for N ⊂ M, thanks to [Lon89, Prop. 5.1] there
is an isometry w ∈ N such that w ∈ HomEnd(N )(id, θ), where θ := γ↾N , E(m) = w∗γ(m)w
for every m ∈ M, and eN = γ−1
1 (ww∗). Define
wi := γ(ψ∗
i )w, Mi := wiψi,
i ∈ I
where wi ∈ N are such that wi ∈ HomEnd(N )(ρi, θ), cf. [LR95, Sec. 5], and Mi ∈ M have
the desired intertwining property with θ, namely Min = θ(n)Mi, n ∈ N , cf. Definition 3.9.
Observe that wi are non-trivial isometries w∗
i eN ψi =
M ∗
i eN Mi because eN ∈ N ′. As a consequence {Mi} is another Pimsner-Popa basis for
N ⊂ M with respect to E and E(MiM ∗
i ) = 1 and that Pi = ψ∗
i ) = wiE(ψiψ∗
i wi = E(ψiψ∗
i = wiw∗
i . Setting
i )w∗
mi := γ(Mi),
i ∈ I
we have that mi ∈ N1 := γ(M) = hθ(N ),{mi}i fulfill mi ∈ HomEnd(N )(θ, θ2),
pi := γ1(Pi) = m∗
i ww∗mi
are mutually orthogonal projections in N which Pi pi = 1, and nw = 0 ⇒ n = 0 for n ∈ N1
follows immediately from faithfulness of E′ := γ ◦ E ◦ γ−1. Hence (θ, w,{mi}) is a generalized
Q-system of intertwiners in C = End(N ) associated, in the sense of Theorem 3.3, to the
discrete inclusion N ⊂ M.
Remark 5.3. With these normalizations for w and ψi, i ∈ I, we have that ρi(n) = E(ψinψ∗
i ),
n ∈ N , i.e., ρi is implemented by a single charged field ψi via the conditional expectation E,
cf. [LR95, Sec. 5], [BKLR16, Sec. 4.4].
14
Proposition 5.4. Let N ⊂ M be a discrete inclusion as in Definition 5.1 and {ψi} ⊂ M
a Pimsner-Popa bases of charged fields as in the proof of Proposition 5.2. Then for every
m ∈ M, the coefficients E(ψim) ∈ N in the Pimsner-Popa expansion (Proposition 2.3)
m =Xi
ψ∗
i E(ψim)
are uniquely determined.
i ) = 1 for every
Proof. We have already checked in the proof of Proposition 5.2 that E(ψiψ∗
i ∈ I, hence we can apply Proposition 2.3.
Proposition 5.5. Let N be an infinite factor with separable predual and N ⊂ M a discrete
extension as in Proposition 5.2. Fix a conditional expectation E ∈ E(M,N ) and a canonical
endomorphism γ with dual canonical endomorphism θ = γ↾N . Then a generalized Q-system
of intertwiners (θ, w,{mi}) can be chosen such that the set of indices I labels the irreducible
subsectors [ρi] (necessarily with finite dimension) of [θ], counted with (arbitrary) multiplicity.
There is a distinguished label 0 ∈ I, corresponding to one occurrence of the identity sector
[id], such that
(9)
m0 = θ(w)
i.e.
θ
• 0
θ
=
• .
θ
θ
θ
θ
Proof. N ′ ∩ M1 is a direct sum of type I factors by discreteness assumption. Hence we
can refine the family of orthogonal projections Pi encountered in the proof of the previous
proposition such that each Pi is minimal in N ′ ∩ M1 and again in the domain of E, thus
each ρi ≺ θ, i ∈ I, is irreducible (with finite index). Every subsector of [θ] arises in this way
and θ = ⊕iρi.
The second statement follows by observing that the Jones projection eN is minimal in
N ′ ∩M1 if and only if N is a factor, if and only if id is irreducible as an object (tensor unit)
of End(N ). Now, by [HKZ91, Lem. 2.2, Prop. 2.4] we can assume that P0 = eN and choose
W0 = eN , hence ψ0 = E(eN ) = 1 and w0 = M0 = w, i.e., m0 = θ(w).
Remark 5.6. In the assumptions of Proposition 5.2, discreteness of the inclusion N ⊂ M
implies
[θ] = ⊕i[ρi]
where [ρi] are irreducible subsectors with finite dimension and counted with (arbitrary) mul-
tiplicity in the set of indices I, cf. comments after [ILP98, Def. 3.7].
If in addition N ⊂ M is irreducible, i.e., N ′ ∩ M = C1, then the multiplicity of each [ρi]
in [θ] is finite and bounded above by the square of the dimension of [ρi], see [ILP98, Thm.
3.3, App.].
15
Remark 5.7. The Pimsner-Popa elements {mi} ⊂ N , or equivalently {Mi} ⊂ M (Definition
3.9), constructed from discrete inclusions via charged fields as in Proposition 5.2 have the
following additional properties. Compute w∗mi = w∗γ(wiψi) = wiw∗γ(ψi) = wiw∗
i , hence
w∗mi = m∗
i w = wiw∗
i
(10)
i.e.
θ
• i
θ
•
= •
θ
• i
θ
θ
w∗
i
ρi ,
=
wi
θ
θ
θ
θ
• i
•
=:
• • i =
θ
θ
• • i
• • i
θ
for every i ∈ I. Consider the spatial isomorphism θ(N ) ⊂ N = γ1(N ⊂ M1) such that
θ(N )′ ∩ N = γ1(N ′ ∩ M1), where γ1 is the canonical endomorphism for M ⊂ M1 dual to γ.
From γ1(Pi) = wiw∗
i ) we conclude that
i = E(MiM ∗
pi = qi
where pi = γ1(Pi), and qi = E(MiM ∗
mutually orthogonal projections in N such that Pi qi = 1 as well.
i ) are defined in Lemma 2.2. In particular qi, i ∈ I, are
0 =
If we consider {Mi} constructed as in Proposition 5.5 we have in addition w∗m0 = w0w∗
ww∗ and
i.e.
(11)
θ(w∗)mi = δi,01
θ
• i
•
θ
= δi,0
θ
θ
for every i ∈ I. Moreover
E(Mi) = E(MiM ∗
0 )M0 = δi,0M0.
Definition 5.8. We say that a generalized Q-system of intertwiners (Definition 3.7) is unital,
if it satisfies in addition the analogue of equations (9), (10), (11), namely
m0 = θ(w), w∗mi = m∗
i w = (w∗mi)2,
θ(w∗)mi = δi,01
for every i ∈ I, and for a distinguished label 0 ∈ I.
One can easily check that the braided product (Definition 4.1) of two unital generalized
Q-systems of intertwiners is again unital.
16
6 Generalized Q-systems of intertwiners for local nets
Let {A} = {O ∈ K 7→ A(O)} be a net of infinite von Neumann factors (typically of type
III1) over a partially ordered by inclusion and directed set K of open bounded regions O of
spacetime (e.g., the set of open proper bounded intervals O = I ⊂ R, or double cones in
Minkowski space O ⊂ Rn+1, n ≥ 1). A net is called isotonous if O ⊂ O implies A(O) ⊂ A( O),
and local if A(O) and A( O) commute elementwise whenever O ⊂ O′, where O′ denotes the
space-like complement of O in Rn+1, n ≥ 1, or the interior of the complement of the interval
O = I in R.
Definition 6.1. A net {A} as above fulfilling isotony and locality is called a net of local
observables, also abbreviated as local net.
We refer to [Haa96], [LR95, Sec. 3], [CCG+04, Ch. 5] for more explanations and for the
physical motivations behind this notion.
Now, let {A} be realized on a separable Hilbert space H0 (vacuum space) and assume
the existence of a unit vector Ω0 ∈ H0 (vacuum vector) which is cyclic and separating for
each local algebra A(O). In this case, we say that {A} is a standard net on H0 with respect
to Ω0 and denote by ω0 := (Ω0 · Ω0) the vacuum state of the net. We say that Haag duality
holds for {A} in the vacuum space if
A(O′)′ = A(O)
for every O ∈ K, where A(O′) is the C ∗-algebra generated by all A( O), O ∈ K, O ⊂ O′.
Denote by DHR{A} ⊂ End(A) the category of DHR endomorphisms of the net,
see [DHR71], [DHR74], [FRS92], and by A the quasilocal algebra, i.e., the C ∗-algebra
generated by {A}. In the following we shall be interested in two distinguished subcategories
of the DHR category, namely
Definition 6.2. Denote by DHRf{A} and DHRd{A} the full subcategories of DHR{A}
whose objects are, respectively, finite-dimensional DHR endomorphisms and (possibly infi-
nite, countable) direct sums of those.
i ∈ ρ(A)′ ∩A(O) are mutually orthogonal projections and Pi wiw∗
j tPi wiw∗
More precisely the most general object ρ in DHRd{A} arises as follows. Let ρi be a
family of at most countably many irreducible finite-dimensional DHR endomorphisms which
can be localized in O ∈ K. Let {wi} be a (possibly infinite) Cuntz family of isometries in
A(O) such that wiw∗
i = 1.
Then ρ = Pi Adwi ρi, where ρi =: Adw∗
ρ and the sum converges elementwise in the strong
operator topology because ρ = Pi wiw∗
i . Similarly, the most
general arrow t between objects ρ, σ in DHRd{A} can be written as t =Pj vjv∗
i =
Pi,j vjtj,iw∗
j twi are
arrows from ρi to σj.
Remark 6.3. Observe that DHRf{A} ⊂ DHRd{A} ⊂ DHR{A} and each inclusion is full,
replete and stable under (finite) direct sums and subobjects. The first two categories are
semisimple in the sense that every object can be written as a (possibly infinite) direct sum
of irreducible finite-dimensional objects.
i where {wi}, {vj} are Cuntz families, respectively, for ρ, σ and tj,i := v∗
i
i ρ(·)Pj wjw∗
j = Pi wiρi(·)w∗
The following is the net-theoretic version of Definition 3.7, and generalizes the notion of
Q-system for local nets given in [LR95, Sec. 4].
17
Definition 6.4. Let {A} be a local net fulfilling Haag duality as above. A generalized net
Q-system of intertwiners in C = DHR{A} is a triple (θ, w,{mi}) consisting of a DHR
endomorphism θ in DHR{A}, an isometry w ∈ HomDHR{A}(id, θ), and a family {mi} ⊂ A
indexed by i in some set I, such that
(i) pi := m∗
i ww∗mi are mutually orthogonal projections in A, i.e., pip∗
j = δi,jpi, such that
Pi pi = 1.
of θ and for any other O ∈ K such that O ⊂ O.
(ii) aw = 0 ⇒ a = 0 if a ∈ A1(O) := hθ(A(O)),{mi}i for some localization region O ∈ K
(iii) mi ∈ HomDHR{A}(θ, θ2), i ∈ I.
Remark 6.5. By the localization property of θ and by Haag duality, (θ, w,{mi}) is a gener-
alized Q-system of intertwiners in End(A( O)) (Definition 3.7) for every O as above. Indeed,
DHR{A} sits into End(A( O)) via the restriction functor as a (full if local intertwiners are
global), replete and braided tensor subcategory for every such O, cf. [GR15, Sec. 3].
Remark 6.6. Condition (iii) in Definition 6.4, in view of Proposition 5.2, excludes many
interesting infinite index extensions of local nets. Notably the Virasoro net {Virc} in one
spacetime dimension, which sits in every conformal (diffeomorphism covariant) net, gives
often rise to infinite index semidiscrete but non-discrete extensions if c > 1, see [Reh94],
[Car04], [Xu05]. It is however fulfilled in many examples of chiral conformal embeddings with
infinite index, see Section 10, as in compact orbifold theories in low and higher dimensions,
see [Xu00] and [DR90], and of course in every finite index extension.
Definition 6.7. [LR95]. An inclusion of nets is defined by two isotonous nets of von
Neumann algebras {A}, {B} over the same partially ordered set of spacetime regions K and
realized on the same separable Hilbert space H such that A(O) ⊂ B(O) for every O ∈ K. In
this case, we write
{A ⊂ B}
and call {B} an extension of {A}. The inclusion is called irreducible if A(O)′ ∩ B(O) = C1
for every O ∈ K. The net {B} is relatively local with respect to {A} if B(O) ⊂ A(O′)′ for
every O ∈ K. If {A} is local, {B} will be always implicitly assumed to be relatively local
with respect to {A}.
The inclusion of nets is called standard if there is a vector Ω ∈ H which is standard for
{B} on H and for {A} on a subspace H0 ⊂ H. A normal faithful conditional expectation E
of {B} onto {A} is a family indexed by O ∈ K of normal faithful conditional expectations
EO ∈ E(B(O),A(O)) which respect inclusions, namely such that E O ↾B(O) = EO if O ⊂ O.
A normal faithful state ω of {B} is a conditional expectation of {B} onto the trivial net {C}
and E as above is called standard if it preserves the standard vector state ω := (Ω · Ω) of
the net, namely ωO ◦ EO = ωO for every O. We say that the extension {A ⊂ B} is discrete
if A(O) ⊂ B(O) is discrete (Definition 5.1) for every O ∈ K.
The following theorem extends the results of [LR95, Thm. 4.9] to the case of infinite index
discrete inclusions of nets of von Neumann algebras.
Theorem 6.8. Let {A} be a local net fulfilling Haag duality and standardly realized on H0 as
in the beginning of this section. Then a generalized net Q-system of intertwiners (θ, w,{mi})
18
in C = DHR{A} (Definition 6.4) which is also unital (Definition 5.8) gives an isotonous
net of von Neumann algebras {B} such that {A ⊂ B} is a discrete standard inclusion of nets
with a normal faithful standard conditional expectation. The net {B} is always relatively local
with respect to {A}, and it is itself local if and only if θ(εθ,θ)mimj = mjmi for every i, j ∈ I,
where ε denotes the DHR braiding.
Proof. Let O ∈ K be a localization region of the DHR endomorphism θ, call N := A(O) and
θ ≡ θ↾N ∈ End(N ) the restriction of θ to N , and observe that w ∈ N , mi ∈ N for every i ∈ I
by Haag duality. From Theorem 3.3 we get N2 ⊂ N1 ⊂ N with a normal faithful conditional
expectation E′ := θ(w∗ · w) ∈ E(N1,N2) and such that θ is a canonical endomorphism
for N1 ⊂ N . Now N acts standardly on H0 by assumption hence θ = AdΓ on N , where
Γ := JN1,ΦJN ,Φ and Φ ∈ H0 is cyclic and separating for N1 and N . Let M := AdΓ∗(N1) be
the corresponding canonical extension of N1 ⊂ N with canonical endomorphism γ := AdΓ↾M.
Lift accordingly the conditional expectation E := w∗γ(·)w ∈ E(M,N ) and consider the
normal faithful E-invariant state ϕ := ω0 ◦ E of M, where ω0 = (Ω0 · Ω0) is the vacuum
state of {A}. The operators Mi := γ−1(mi) ∈ M as in Definition 3.9 form a Pimsner-Popa
basis for N ⊂ M with respect to E and fulfill
Min = θ(n)Mi,
(12)
i ∈ I, n ∈ N .
Now consider the (normal faithful) GNS representation (Hϕ, πϕ, Ωϕ) of M with respect to
ϕ = ϕ ◦ E. The inclusion πϕ(N ) ⊂ πϕ(M) on Hϕ with conditional expectation Eϕ given by
Eϕ(πϕ(m)) := πϕ(E(m)), m ∈ M, is spatially isomorphic to N ⊂ M on H0 with respect
to E. Moreover, (ΩϕEϕ(πϕ(m))Ωϕ) = (Ωϕπϕ(m)Ωϕ), m ∈ M and eN := [πϕ(N )Ωϕ] is the
associated Jones projection. By spatial isomorphism we have that {πϕ(Mi)} is a Pimsner-
Popa basis for πϕ(N ) ⊂ πϕ(M) with respect to Eϕ, and γϕ given by γϕ(πϕ(n)) := πϕ(γ(n)),
n ∈ N , is a canonical endomorphism with dual canonical θϕ := γϕ↾πϕ(N ). In particular, we
have a direct sum decomposition
Hϕ =Xi
πϕ(M ∗
i )eNHϕ
where πϕ(M ∗
projections and we let
i )eN , i ∈ I, are partial isometries with mutually orthogonal range and domain
H0,N ,ϕ := eNHϕ.
i )ψi =Xi
Every n ∈ N ⊂ M acts by left multiplication on Hϕ, then
πϕ(M ∗
πϕ(n)Xi
where ψ =Pi πϕ(M ∗
i )ψi, with ψi := eN πϕ(Mi)ψ ∈ πϕ(qi)eNHϕ, is the generic vector of Hϕ.
As in the proof of [LR95, Thm. 4.9], this representation of N = A(O) extends to the whole
net. Indeed, the linear map
i )πϕ(θ(n))ψi
πϕ(M ∗
(13)
U0 : nΩ0 7→ πϕ(n)Ωϕ, n ∈ N
extends to a unitary operator from H0 onto H0,N ,ϕ, due to ϕ ≡ ω0 ◦ E and E(n) = n, n ∈ N ,
which implements πϕ↾N on the subspace H0,N ,ϕ via adjoint action. For every quasilocal
observable a ∈ A and ψi as above, define
πϕ(a)Xi
πϕ(M ∗
i )ψi :=Xi
πϕ(M ∗
i )U0θ(a)U ∗
0 ψi.
19
One can check that πϕ is a well-defined bounded and locally normal representation of A on
Hϕ which extends the GNS representation restricted to N due to equation (13).
In this
representation, the intertwining relation (12) extends to the net, namely
πϕ(Mi)πϕ(a) = πϕ(θ(a))πϕ(Mi),
i ∈ I,
a ∈ A.
(14)
To show this, we first check that in the representation on Hϕ we have that e N := [πϕ( N )Ωϕ],
where N := A( O), fulfills
eN = e N
for every O ∈ K (not necessarily O ⊂ O). Indeed, Ωϕ = πϕ(M ∗
0 w =
w∗w = 1 by unitality assumption, and the closed linear span in Hϕ of vectors of the form
0 )πϕ(w)Ωϕ because M ∗
πϕ(a)Ωϕ = πϕ(a)πϕ(M ∗
0 )πϕ(w)Ωϕ = πϕ(M ∗
0 )U0θ(a)wU ∗
0 Ωϕ
= πϕ(M ∗
0 )U0waΩ0 = U0aΩ0
does not depend on whether a ∈ N or a ∈ N by the intertwining property of w on A
and because Ω0 is cyclic for every local algebra on H0 by standardness assumption. Hence
e NHϕ = H0,N ,ϕ ⊂ Hϕ for every O ∈ K.
Now let ψ and ψi be as in equation (13), and assume that a ∈ A( O) for some O ⊂ O.
From the l.h.s. of (14) we get
πϕ(Mi)πϕ(a)ψ =Xj
πϕ(MiM ∗
j )U0θ(a)U ∗
0 ψj
because left multiplication is continuous in the GNS representation.
By Proposition 2.3 (valid for arbitrary semidiscrete inclusions) we can write
πϕ(Mi)πϕ(Mj )∗ =Xk
πϕ(Mk)∗πϕ(lki
j ),
lki
j
:= E(MkMiM ∗
j ) ∈ N
(15)
j = w∗γ(MkMiM ∗
j )w = w∗mkmim∗
j w and intertwines θ with θ2 on the whole net by
where lki
assumption, i.e.
lki
j ∈ HomDHR{A}(θ, θ2).
Recall that the convergence in the r.h.s. of equation (15) is given by the topology induced by
η = η(m∗m), m ∈ πϕ(M), with η any normal state on πϕ(M) such that
the seminorms kmk2
η = η ◦ Eϕ. Thus
Xj
0 ψj =Xj (cid:0)Xk
j )(cid:1)U0θ(a)U ∗
j )U0θ(a)U ∗
πϕ(MiM ∗
k )πϕ(lki
πϕ(M ∗
0 ψj
and since the vector U0θ(a)U ∗
πϕ(M), we get
=Xj,k
0 ψj ∈ e NHϕ = eNHϕ induces a normal Eϕ-invariant state on
πϕ(M ∗
k )U0θ2(a)lki
j U ∗
0 ψj = πϕ(θ(a))πϕ(Mi)ψ
which is the r.h.s. of (14), for every ψ ∈ Hϕ, thus the equation is proven.
We define
B(O) := πϕ(M) ≡ hπϕ(N ),{πϕ(Mi)}i,
20
the crucial point is to extend the construction to bigger regions and define accordingly a co-
herent family of normal faithful standard conditional expectations with respect to a common
cyclic and separating vector. Let O ∈ K be such that O ⊂ O and set
N := A( O),
B( O) := hπϕ( N ),{πϕ(Mi)}i,
clearly B(O) ⊂ B( O) holds by isotony of {A}.
onto πϕ( N )eN , and because of
Now, Ω0 is separating for every N , thus πϕ(a) 7→ πϕ(a)eN is an isomorphism of πϕ( N )
eNB( O)eN ⊂ eNhB( O), eNieN = eNhπϕ( N ), πϕ(M), eN ieN = πϕ( N )eN
provided O ⊂ O, we can define by
eN beN = Eϕ(b)eN ,
b ∈ B( O)
a conditional expectation of B( O) onto πϕ( N ) (for arbitrarily big region O) which extends
the one previously given on B(O) ≡ πϕ(M). Eϕ is clearly normal and fulfills (ΩϕEϕ(·)Ωϕ) =
(Ωϕ · Ωϕ), while faithfulness remains to be checked, together with the separating property
of Ωϕ for B( O) if O ⊂ O and cyclicity for B( O) if O ⊂ O, where B( O) in this second case is
defined below. For an arbitrary region O ∈ K, set
B( O) := hπϕ( N ),{πϕ(u)πϕ(Mi)}i,
(16)
where u ∈ HomDHR{A}(θ, θ) is a unitary charge transporter (in A) and θ is DHR localizable
in O, 3. In order to show the desired properties of Ωϕ with respect to these new local algebras
we need to introduce more GNS representations. Namely, let
N := A( O)
(θ ≡ θ↾ N , w := uw,{ mi := uθ(u)miu∗})
be a generalized Q-system of intertwiners in End( N ), see Remark 6.5, and perform the same
construction as above on the GNS Hilbert space H ϕ of some canonical extension N ⊂ M
with E ∈ E( M, N ) and state ϕ := ω0 ◦ E of M. Consider
)e NH ϕ
π ϕ( Mi
∗
H ϕ =Xi
and
extended as before to an isometric operator into H ϕ. Then the linear map defined by
∗
i )U0u∗φi, φi ∈ H0
U0 : nΩ0 7→ π ϕ(n)Ω ϕ, H0 → H0, N , ϕ ≡ e NH ϕ,
UXi
) U0φi :=Xi
πϕ(M ∗
π ϕ( Mi
n ∈ N
sends
U Ω ϕ = U π ϕ( M0
∗
) U0 wΩ0 = πϕ(M ∗
0 )U0u∗ wΩ0 = Ωϕ
because u∗ w = w, it extends to a unitary operator from H ϕ onto Hϕ, and fulfills
U π ϕ( M)U ∗ = hπϕ( N ),{πϕ(u)πϕ(Mi)}i ≡ B( O).
(17)
3Notice that we assume DHR sectors to be localizable in every region in K.
21
Indeed, if n ∈ N then
U π ϕ(n)U ∗Xi
=Xi
πϕ(M ∗
πϕ(M ∗
i )U0φi = U π ϕ(n)Xi
i )U0u∗ θ(n)uφi = πϕ(n)Xi
∗
π ϕ( Mi
) U0uφi
πϕ(M ∗
i )U0φi
because u∗ θ(n)u = θ(n), and
U π ϕ( Mi)U ∗Xj
πϕ(M ∗
j )U0φj =Xj,k
πϕ(M ∗
k )U0u∗lki
j uφj
where the coefficients lki
j
One can compute θ(u∗)u∗lki
j u = lki
j , hence
:= E( Mk Mi M ∗
j ) = w mk mi m∗
j w ∈ N are analogous to those in (15).
=Xj,k
πϕ(M ∗
k )U0θ(u)lki
j φj = πϕ(u)πϕ(Mi)Xj
πϕ(M ∗
j )U0φj
which proves (17). By considering this unitary intertwiner for every region O we obtain that
Ωϕ is cyclic and separating for every B( O) on Hϕ, in particular Eϕ is faithful over every O.
The extension {B} does not depend on the specific choice of unitary charge transporter u
made in equation (16). Indeed, by Haag duality any two of them u, v differ by uv∗ ∈ A( O).
Also, it depends on the choice of the initial localization region O for θ and of the extended
vacuum state ϕ only up to unitary isomorphism.
Relative locality of {B} with respect to {A} is always guaranteed by the localization
properties of θ while the statement about locality of {B} follows by the very definition of
the DHR braiding. Indeed, uMivMj = vMjuMi where u and v are unitaries in A which
transport the localization region of θ to two mutually space-like regions (respectively left and
right localized in low dimensions) if and only if εθ,θMiMj = θ(u∗)v∗uθ(v)MiMj = MjMi for
every i, j ∈ I, and the proof is complete.
Remark 6.9. If the Pimsner-Popa expansion appearing in equation (15) comes from an irre-
ducible subfactor, or from a finite index inclusion, and if the unital generalized Q-system of
intertwiners is defined from charged fields Mi = wiψi, i ∈ I, see Proposition 5.5 and Remark
5.6, then the sum over k (a priori convergent in the GNS topology) is finite by Frobenius
reciprocity among finite-dimensional endomorphisms of N , [LR97, Lem. 2.1]. Indeed, in this
case one has lki
In other words, we have a unital *-algebra of charged intertwiners with possibly infinitely
j ) ∈ HomEnd(N )(ρj, ρkρi).
j = wkρk(wi)E(ψkψiψ∗
j and E(ψkψiψ∗
j )w∗
many generators {Mi} but finite ("discrete") fusion rules, cf. [LR04, App. A].
In this section we assumed that a generalized net Q-system of intertwiners (Definition
6.4) was given and we have shown how to associate to it a relatively local net extension. In
Section 5, given a discrete inclusion of von Neumann algebras, we have seen how to construct
a generalized Q-system of intertwiners (Definition 3.7). But when do generalized Q-systems
of intertwiners exist for discrete relatively local net extensions?
In the finite index setting, in [GL92] it is proven that the DHR category restricted to
finite index endomorphisms of a chiral CFT is a full and replete subcategory of the category
of endomorphisms of a local algebra ("local intertwiners are global"). This of course single-
handedly carries over the theory of ordinary Q-systems to such nets. In [LR95] it is shown
22
in the presence of a coherent conditional expectation, with
that actually less is needed;
finite index arguments it can be shown that a Q-system in the DHR category does exist,
see [LR95, Cor. 3.8, Cor. 3.7 and Lem. 4.1]. On the other hand, when we consider also
infinite-dimensional irreducible sectors, in general it is not true that the category of DHR
endomorphisms is a full subcategory of the category of endomorphisms of a local algebra,
see [Wei08] for a counter-example using {Virc}, c > 2. In the absence of these features, we
can make some additional assumptions.
Proposition 6.10. Let {A ⊂ B} be a discrete relatively local inclusion of nets with standard
conditional expectation E (Definition 6.7) and let {A} be a Haag dual net of type III factors.
Let γ be a canonical endomorphism for {A ⊂ B} as in [LR95, Cor. 3.3] and θ ∈ DHR{A} be
its dual canonical endomorphism as in [LR95, Cor. 3.8] (localized in O0 ∈ K). Moreover, let
w ∈ A(O0) as in [LR95, Cor. 3.7] such that E = w∗γ(·)w.
Assume that one of the following two conditions is fulfilled
(i) {A ⊂ B} is irreducible and θ = ⊕iρi in DHR{A} where each ρi is an irreducible DHR
subendomorphism (localized in O0) and [ρi] has finite dimension in DHR{A}.
(ii) HomDHR{A}(θ, θ) = HomEnd(A(O))(θ, θ) for every O ∈ K, O0 ⊂ O, or equivalently
θ = ⊕iρi in DHR{A} and HomDHR{A}(ρi, ρj) = HomEnd(A(O))(ρi, ρj) for every i, j ∈ I.
Then there is a Pimsner-Popa basis {ψi} for the inclusion A(O0)
global charged fields, i.e.
E
⊂ B(O0) consisting of
ψia = ρi(a)ψi,
a ∈ A.
E
i = pi.
i )wψi we have another Pimsner-Popa basis for A(O0)
Setting Mi := γ(ψ∗
⊂ B(O0) (cf. Propo-
sition 5.2), and (θ, w,{mi}), where mi := γ(Mi) ∈ A(O0), is a unital generalized net Q-
system of intertwiners in DHR{A} (Definition 6.4 and 5.8).
Proof. Assume (i) and let pi ∈ HomDHR{A}(θ, θ) be the projections which determine the
decomposition θ = ⊕iρi, together with orthogonal isometries wi ∈ HomDHR{A}(ρi, θ)∩A(O0)
such that wiw∗
For O ∈ K, let eA = [A(O)Ω] be the Jones projection, and B1(O) = hB(O), eAi the Jones
extension for the inclusion A(O) ⊂ B(O). By standardness assumption the Jones projections
agree for every O ∈ K. Denote by EO ∈ P (B1(O),B(O)) the operator-valued weight dual to
be respectively its domain and definition ideal. By the same
E↾B(O) and let m EO
arguments leading to [LR95, Thm. 3.2, Cor. 3.3], the formula
and n EO
x ∈ B1(O)
γ1(x)eA = v1xv∗
1,
where v1 ∈ B1(O) is an isometry such that v1v∗
1 = eA, γ1(v1) = w, allows to extend the dual
canonical endomorphism γ1 : B1(O0) → A(O0) to a map between the quasilocal algebras B1
and A, such that γ1↾B1(O) is the dual canonical endomorphism for the inclusion A(O) ⊂ B(O)
for every O ∈ K, O0 ⊂ O (i.e., γ1↾B1(O) is a canonical endomorphism for B(O) ⊂ B1(O),
γ1↾B(O) = γ↾B(O) and γ1(B1(O)) = A(O)).
By the discreteness assumption and [ILP98, Prop. 2.8] we have that, for every O ∈ K,
B1(O)∩A(O)′ is a direct sum of type I factors, which by irreducibility of A(O) ⊂ B(O) and
[ILP98, Thm. 3.3] are finite dimensional.
23
Pi := γ−1
ρi↾A(O) has finite dimension by hypothesis, and thus Pi ∈ n EO
support of Pi in B1(O) ∩ A(O)′. EO↾zPi (B1(O)∩A(O)′)zPi
as a sum of minimal projections in n EO
dimensional. Now, let
1 (pi) is a finite sum of minimal projections in B1(O) ∩ A(O)′, O0 ⊂ O, since
. Indeed, let zPi be the central
is semifinite since zPi can be written
, and thus finite since zPi(B1(O) ∩ A(O)′)zPi is finite
1 (ww∗
. Let ψi,O := EO(Wi). Exactly as in
Since Wi = eAWiPi and Pi ∈ n EO
the proof of Proposition 5.2, and using wi = γ(ψ∗
i,O)w, we get that ψi,O is a charged field for
ρi on A(O), i.e., ψi,Oa = ρi(a)ψi,O, a ∈ A(O), and the collection {ψi,O} is a Pimsner-Popa
basis for the inclusion A(O) ⊂ B(O). For any O ∈ K, O0 ⊂ O, by the push-down lemma we
have
Wi := γ−1
, we have Wi ∈ m EO
i ) ∈ B1(O0).
Thus applying EO to the above formula, we obtain ψi,O = ψi,O0. The rest of the proof follows
exactly as in Proposition 5.2.
eAψi,O = Wi = eAψi,O0.
Assuming (ii), the proof proceeds along similar lines. By discreteness assumption, one
can take projections pi ∈ θ(A(O0))′ ∩ A(O0) which lay in n E′
= γ ◦ E↾B(O0) ◦
γ−1, and which give a local (in this case also global) decomposition of θ = ⊕iρi into DHR
subendomorphisms. Now, let wi such that pi = wiw∗
i and define Pi, Wi and ψi,O0 as before
such that Wi = eAψi,O0. To conclude it is enough to observe that for any O ∈ K, O0 ⊂ O,
eAψi,O0 by construction and ψi,O0 ∈ B(O0) ⊂ B(O),
we have Pi ∈ n EO
and also Pi ∈ B1(O) ∩ A(O)′, because θ(A(O0))′ ∩ A(O0) = θ(A(O))′ ∩ A(O) for every such
O.
, because Pi = ψ∗
, where E′
i,O0
O0
O0
These assumptions are verified, e.g., for compact group orbifolds [DR72], [Rob74, Thm.
4.3], and for theories with a good behaviour with respect to the scaling limit [DMV04, Cor.
6.2] in 3+1D, and of course for strongly additive CFTs in 1D, i.e., Haag dual nets on R
[GLW98, Lem. 1.3].
6.1 Construction of extensions: an alternative way
In this section we present an alternative proof of Theorem 6.8 which we feel somewhat more
intuitive and which lends itself to describing the braided product of nets in a more direct way.
The basic idea is that a generalized net Q-system of intertwiners (θ, w,{mi}) in DHR{A}
(Definition 6.4), assuming Haag duality of {A}, induces a family of Q-systems (one for every
local algebra) each one of which characterizes a local extension. The algebraic structure of
the extended net, including a distinguished conditional expectation, is then captured with a
coherent inductive procedure and the spatial features of the net are completely determined
by the vacuum state.
More precisely, let O0 ∈ K be a reference localization region for θ and choose a unitary
charge transporter uO ∈ HomDHR{A}(θ, θO) for every O ∈ K, where θO := AduO θ is localized
in O. For every O ∈ K, we obtain a generalized Q-system (of intertwiners) in End(A(O))
(Definition 3.7) by setting
(θO, wO,{miO}) := (AduO θ, uOw,{uOθ(uO)miu∗
O}).
24
The faithfulness condition appearing in Definition 3.1 is verified since, for every O ∈ K such
that O0 ∪ O ⊂ O, we have
where
A1(O) ⊂ uOhθ(A( O)), miiu∗
O = AduO (A1( O)),
A1(O) := hθO(A(O)), uOθ(uO)miu∗
Oi.
With this data, we now construct an inductive generalized sequence of net extensions
{{B O}, O ∈ K} of {A}, indexed by O ∈ K, and defined only on regions O ∈ K, O ⊂ O, which
we will then patch together. For a fixed O ∈ K, θ O = Adu O
θ is a canonical endomorphism
for A1( O) ⊂ A( O) by Theorem 3.3, and thus can be implemented on A( O) by a unitary Γ O,
namely
Similarly, for every O ⊂ O, we have
θ O(x) = Γ OxΓ∗
O ,
x ∈ A( O).
θO(x) = AduO (θ(x)) = AduOu∗
O
Γ O
(x) ,
x ∈ A( O).
Now, fixed O ∈ K, we define an isotonous net {B O} := {O ∈ K, O ⊃ O 7→ B O(O)} by setting
B O(O) := Ad(uOu∗
O
Γ O)∗(A1(O)) = hA(O),{uO AdΓ∗
O
u O
(mi)}i.
Remark 6.11. The dual canonical endomorphism θ for an extension of nets {A ⊂ B}, [LR95,
Cor. 3.8], is not implemented globally by unitaries. This is clear since by [LR95, Prop. 3.4]
the embedding homomorphism ι of {A} into {B} is equivalent to θ as a representation and
thus would imply the inclusion to be trivial. Of course it is possible to find unitaries which
implement θ locally but the choice of these unitaries in non-unique. The above coherent
choice guarantees the isotony of the net {B O}.
The next step is the construction of an inductive family of embeddings ι O1, O2
of {B O1}
into {B O2} with O1 ⊂ O2. This is straightforward.
Proposition 6.12. Let O1 ⊂ O2, O1, O2 ∈ K. The map
Γ O1
ι O1, O2
:= AdΓ∗
O2
u O2
u∗
O1
is an embedding of {B O1} into {B O2}, i.e., it sends local algebras B O1
B O2
Proof. Follows by easy direct computation.
(O) and acts as the identity map on A(O) for every O ∈ K, O ⊂ O1.
(O) onto local algebras
The collection of nets {{B O}, O ∈ K} and maps {ι O1, O2
, O1 ⊂ O1, O1, O2 ∈ K} forms an
inductive system. We can thus take the inductive limit of the C ∗-algebras B O( O) from which
we obtain a C ∗-algebra B. The subalgebras B(O) := ιO(BO(O)), where ιO is the embedding
of BO(O) into B, are W ∗-algebras since it is easy to see that they have a predual. Thus we
have obtained an isotonous net of W ∗-algebras, {B}. Now we see that from the data of the
Q-system we can also define a consistent family of conditional expectations.
Proposition 6.13. There is a normal faithful conditional expectation from {B} to {A}, i.e.,
EO : B(O) → A(O) for every O ∈ K, such that EO2 ↾B(O1) = EO1 if O1 ⊂ O2.
25
Proof. First define a coherent conditional expectation on {B O} for O ∈ K. By Theorem 3.3
we have a conditional expectation E′
O·uOw) for the inclusion θO(A(O)) ⊂ A1(O).
O can be lifted to a conditional expectation E OO for the inclusion A(O) ⊂ B O(O), O ⊂ O,
E′
since the two inclusions are isomorphic via AduOu∗
O
. Computing explicitly, we get
O := θO(w∗u∗
Γ O
E OO = w∗u∗
OΓ O · Γ∗
Ou Ow
which shows that we indeed have a consistent family of conditional expectations on {B O}.
Now, to show that these expectations lift to the inductive limit net {B}, it is enough to check
that E O2O ↾ι O1, O2
= E O1O for O1 ⊂ O2, but this is a trivial computation.
(B O1
(O))
If ω0 is the vacuum state of {A}, let ω := ω0 ◦ E, where E is the consistent conditional
expectation of the inclusion {A ⊂ B} defined above, lifted to the quasilocal C ∗-algebra B.
We call ω the vacuum state of {B} and the GNS representation induced by ω the vacuum
representation. We denote by {A ⊂ BQ} and {BQ} the extension constructed in this way, in
its vacuum representation.
Remark 6.14. It is not hard to check that the construction of the net {BQ} and its conditional
expectation EQ onto {A} does not depend on the choice of the family of unitary charge
transporters uO, nor on the choice of Γ O.
Up to now, we have seen that we can build (discrete, relatively local) extensions of nets
{A ⊂ BQ} associated to generalized net Q-systems of intertwiners in DHR{A}. A natural
question to ask is if this procedure insures that, if the Q-system comes from a given extension
{A ⊂ B}, the induced extension will be unitarily equivalent to the starting one. The answer
is affirmative when the generalized net Q-system is constructed as in Proposition 6.10.
Lemma 6.15. Let {A ⊂ B} be as in Proposition 6.10, assuming either (i) or (ii), and let
θ = ⊕imiρi be a decomposition of θ into irreducibles in DHR{A}, where [ρi] 6= [ρj], mi is the
multiplicity of [ρi] in [θ], and ρi, θ are localized in O ∈ K. Then
Hρi(O) := {ψ ∈ B(O), ψa = ρi(a)ψ, a ∈ A}
is isomorphic as a Hilbert space to HomDHR{A}(ρi, θ) via the map Φ : ψ 7→ γ(ψ∗)w.
Proof. Note that E(ψ1ψ∗
2) ∈ HomDHR{A}(ρi, ρi), ψ1, ψ2 ∈ Hρi(O), is an inner product for
Hρi(O) since ρi is irreducible in DHR{A}. We have seen in Proposition 6.10 that there is
a collection {ψj} ⊂ Hρi(O), j = 1, . . . , mi (with mi possibly infinite), which is orthonormal
with respect to the above inner product, and which is mapped via Φ : ψ 7→ γ(ψ∗)w onto an
2) = Φ(ψ1)∗Φ(ψ2), the map Φ : ψ 7→
orthonormal basis of HomDHR{A}(ρi, θ). Since E(ψ1ψ∗
γ(ψ∗)w is an isomorphism of Hρi(O) onto HomDHR{A}(ρi, θ).
Proposition 6.16. Let {A ⊂ B} and (θ, w,{mi}) be as in Proposition 6.10, assuming either
(i) or (ii). Then the inclusion {A ⊂ BQ} obtained from (θ, w,{mi}) is unitarily equivalent
to {A ⊂ B}.
Proof. We first show that
A1(O) = γO(B(O)), A1(O) := hθO(A(O)), uOθ(uO)miu∗
Oi
26
where γO := AduO γ is a canonical endomorphism for the inclusion A(O) ⊂ B(O). To see
∗uOwi
this, let wi ∈ A(O) be isometries such that wi wi
is a unitary and ρi,O := Ad wi
∗uOwi ρi is localized in O. Using Lemma 6.15, we have that
O. Then we have that wi
∗ = uOwiw∗
i u∗
Hρi(O0) = Hρi( O) = wi
∗uOwiHρi,O ( O) = wi
∗uOwiHρi,O (O)
for every O ∈ K with O0 ∪ O ⊂ O. Consequently
uOMi = uOwiψi = uOwiw∗
O)(uOwiψi)
i wiψi = (uOwiw∗
i u∗
∗)(uOwiψi) = wi ψi ∈ B(O)
= ( wi wi
from which the claim easily follows.
Thus it is clear that for a fixed O ∈ K, the map π O : B( O) → BQ
( O), π O := AdΓ∗
γ O, is
an isomorphism of von Neumann algebras, which maps B(O) onto BQ
(O) and which lifts to
a representation π of the net {B}. To show that this representation is unitarily equivalent to
the vacuum representation, it is enough to show by the GNS theorem that
O
O
O
EQ ◦ π = E
with EQ and E respectively the conditional expectations of {A ⊂ BQ} and {A ⊂ B}, but
this is clear using E = w∗γ(·)w.
Remark 6.17. Note that in our second construction of the net {BQ}, the only instance where
the intertwining property of the mi was used is to make sure that miO = uOθ(uO)miu∗
O ∈
A(O) (precisely by the intertwining property of the mi and Haag duality of {A}, cf. Remark
6.5). If a generalized Q-system does not have the intertwining property then the isotonous, rel-
atively local net extension {BQ} can still be defined in the same way, although the conditional
expectation EQ cannot be defined on regions O 6⊃ O0 since in general miO = uOθ(uO)miu∗
need not be in A(O), and thus (θO, wO,{miO}) is not a priori a generalized Q-system in
End(A(O)).
O
7 Covariance of extensions
In this section we show how spacetime covariance (e.g., Mobius covariance in 1D or Poincar´e
covariance in 3+1D) extends from {A} to {B}, where {A} is a local covariant net over a
directed set of spacetime regions K, and {A ⊂ B} is an extension with the properties implied
by Theorem 6.8. This fact is common knowledge among experts, cf. [KL04, Rmk. 4.3] for
irreducible extensions of {Virc}, c < 1, [BMT88, Sec. 3C] for time translation covariance
in extensions of the chiral U (1)-current, and [MTW16, Sec. 6] for more recent examples of
diffeomorphism covariant extensions of the U (1)-current in 1+1D. See also [DR90, Sec. 6],
[DR89, Thm. 8.4] for the covariance of canonical field extensions in 3+1D. In this section,
see Theorem 7.7, we give a general proof of covariance for extensions of nets, with finite or
infinite index (of discrete type as in Theorem 6.8). The proof essentially relies on tensoriality
and naturality properties of the action of the spacetime symmetry group (implemented by
covariance cocycles) on the DHR category. Hence we formulate it in a C ∗ tensor categorical
language, cf. [Tur10, App. 5] due to M. Muger. But first we need a few definitions.
Let P be a (pathwise) connected and simply connected group of spacetime symmetries
(e.g., P = gMob the universal covering of the Mobius group acting on R (actually on R =
27
R ∪ {∞}), or P = P↑
+ the universal covering of the proper orthochronous Poincar´e group
acting on R3+1). Assume that P contains a distinguished (n + 1)-parameter subgroup, n ≥ 0,
of "spacetime translations" (e.g., the rotations inside gMob, or the four-dimensional spacetime
translations inside P↑
+). The following definition describes Poincar´e covariant theories on
Minkowski space and Mobius covariant theories on the real line at the same time, cf. [GL92,
Sec. 8], [BGL93, Sec. 1], [CKL08, Sec. 3].
Definition 7.1. An isotonous net {A} of von Neumann algebras realized on H0 over a
directed set of spacetime regions K is called covariant with respect to P if there is a strongly
continuous unitary representation g 7→ U (g) of P on H0 such that
U (g)A(O)U (g)∗ = A(gO), O ∈ K,
g ∈ UO
where UO ⊂ P denotes the (pathwise) connected component of the identity e in P of the
set {g ∈ P : gO ∈ K}. We always assume that UO is a non-trivial neighborhood of e for
every O ∈ K (i.e., K is "locally stable" under the action of P), and that if O1,O2 ∈ K and
g ∈ UO1 ∩ UO2 then there is O ∈ K such that O1 ∪ O2 ⊂ O and g ∈ UO (i.e., K is "P-stably
directed").4
Concerning spectral properties, we assume that the generators of the spacetime transla-
tion subgroup (energy-momentum operators) have positive joint spectrum, and that there is a
P-invariant unit vector Ω0 ∈ H0 (vacuum vector) which is cyclic for hA(O), U (g) : O ∈ K, g ∈ Pi.
Remark 7.2. Assume first that P preserves K, i.e., gO ∈ K for every g ∈ P, O ∈ K (e.g., if K is
the set of all double cones in Minkowski space and P is the universal covering of the Poincar´e
group, or if K is the set of all open proper bounded intervals of R and P is the translation-
dilation subgroup of the Mobius group), or equivalently UO = P for every O ∈ K. Consider
then a local net {A} over K as in Definition 6.1, fulfilling Haag duality and covariant with
respect to P as in Definition 7.1. Denoted by αg := AdU (g) the adjoint action on B(H0), we
have an action of P on the net {A} (which extends to an action by normal *-automorphisms
of the quasilocal algebra A), and another action of P on DHR endomorphisms ρ in DHR{A}
given by gρ := αgρα−1
g . Observe that gρ is again DHR and localizable in gO if ρ is localizable
in O. Moreover, gt := αg(t) ∈ HomDHR{A}(gρ, gσ), g ∈ P, if t ∈ HomDHR{A}(ρ, σ).
In
other words, we have an action of P on the category DHR{A} (as a strict C ∗ braided
tensor category) by autoequivalences (actually automorphisms), which is also strict in the
terminology of [Tur10, App. 5]. Indeed, one can easily check that g(ρ × σ) = gρ × gσ, where
ρ × σ = ρσ (composition of endomorphisms of A), and g id = id for every g ∈ P and ρ, σ in
DHR{A}. Also, g(hρ) = ghρ and eρ = ρ if e is the identity in P.
On the other hand, if not every g ∈ P, O ∈ K fulfill gO ∈ K (e.g., if K is the set of all open
proper bounded intervals in R and P is the universal covering of the Mobius group) then αg,
g ∈ P, are not always automorphisms of the quasilocal algebra A and the previous global
statements have to be replaced with local ones by specifying local algebras and spacetime
regions. For instance, gρ = αgραg−1, for a fixed g ∈ P, is well defined on every A(O), O ∈ K,
such that g−1O ∈ K, and it is an endomorphisms of A(O) if ρ is and endomorphisms of
A(g−1O) (e.g., if ρ is DHR localizable in g−1O). Similarly, the intertwining relation for
4These assumptions are the abstraction of the geometric properties which are needed in this section. They
are fulfilled, e.g., by all the examples of spacetime symmetries P acting on directed sets of bounded regions
K mentioned above.
28
gt between gρ and gσ, if t ∈ HomDHR{A}(ρ, σ), must be intended locally.
generality we give the following definition, cf. [Lon97, Sec. 2, App. A], [Tur10, App. 5].
In this level of
Definition 7.3. Let {A} be a local net realized on H0 as in Definition 6.1, fulfilling Haag
duality and covariant with respect to a group of spacetime symmetries P as in Definition 7.1.
Let αg := AdU (g), g ∈ P, and let
C ⊂ DHR{A}
be a full and replete tensor subcategory, closed under finite direct sums and subobjects. We
say that P has an equivariant action on C (and write CP = C) if there is a map
where g ∈ P, ρ is an object of C, such that
g, ρ 7→ z(g, ρ)
(i) z(·, ρ) is a strongly continuous unitary valued map in B(H0), z(g, id) = 1 for every
g ∈ P, z(e, ρ) = 1 for every ρ in C, and
z(gh, ρ) = αg(z(h, ρ))z(g, ρ)
for every g, h ∈ P and ρ in C. ("cocycle identity")
(ii)
(iii)
(iv)
(v)
z(g, ρ)ρ(a)z(g, ρ)∗ = αgρ αg−1(a),
a ∈ A( O)
if ρ is DHR localizable in O0 ∈ K, g ∈ UO0 is such that g−1 ∈ UO0, and O ∈ K is such
that O0 ∪ gO0 ⊂ O, g−1 ∈ U O. 5 ("local intertwining property")
if ρ is DHR localizable in O0 ∈ K, g ∈ UO0, and O ∈ K is such that O0 ∪ gO0 ⊂ O.
z(g, ρ) ∈ A(O)
αg(t) = z(g, σ)tz(g, ρ)∗
if ρ and σ are DHR localizable respectively in O1 and O2 ∈ K, g ∈ UO1 ∩ UO2, and
t ∈ HomDHR{A}(ρ, σ). ("naturality of cocycles")
if ρ, σ, g are as in (iv). ("tensoriality of cocycles")
z(g, ρσ) = z(g, ρ)ρ(z(g, σ))
(vi) Adz(g,ρ) ρ is DHR localizable in gO0 ∈ K, if ρ, g are as in (iii). ("global localization
property")
5The existence of at least one O with these properties is guaranteed because K is P-stably directed by
assumption (Definition 7.1).
29
Remark 7.4. In the case that gO ∈ K for every g ∈ P, O ∈ K, we have a (global) action
of P on C ⊂ DHR{A} (as a strict C ∗ braided tensor category), see [Tur10, Def. 1.2]. Then
the equivariance of the action as in Definition 7.3, cf. [Tur10, Def. 2.1], says that the map z
defines a tensor natural transformation (isomorphism) between the trivial action ι of P on
C by autoequivalences and the action defined by α. Naturality is automatic because P is
considered as a discrete tensor category, i.e., the only morphisms are the identity morphisms,
while tensoriality is encoded in the cocycle identity (i). The properties (iv) and (v) above say
that z(g,·) is a natural tensor transformation (unitary isomorphism) between tensor functors
ιg and αg for every g ∈ P.
Lemma 7.5. In the assumptions of Definition 7.3, let z(·,·) be a map fulfilling the properties
(i) and (ii), then the following holds as well
(ii)′
z(g, ρ)ρ(αg (a))z(g, ρ)∗ = αg(ρ(a)),
a ∈ A(O),
g ∈ UO
if ρ is DHR localizable in O ∈ K. The same is true if a ∈ A( O) and g ∈ U O for any
O ∈ K.
In other words, the unitaries Uρ(g) := z(g, ρ)∗U (g), g ∈ P, implement the covariance of ρ
with respect to P (cf. [CKL08, Sec. 4.2]) and give a strongly continuous unitary representation
of P on H0.
Proof. Let a ∈ A(O), g ∈ UO and assume that O ∈ K is a localization region of ρ. Also,
let V ⊂ UO be a symmetric neighborhood of e, e.g., V := UO ∩ U −1
O . Consider the set of
all elements V(e) ⊂ UO that can be joined to e by a V-chain in UO, namely those g ∈ UO
such that there are x1, . . . , xn ∈ UO, n ≥ 1, with x1 = e, xn = g, and xj+1x−1
j ∈ V for
every j = 1, . . . , n − 1, cf. [BP01, Def. 19, 144]. By a standard argument, V(e) is open and
closed in UO, hence V(e) = UO by connectedness. Then every g ∈ UO can be written as
g = g1g2 ··· gn where gj ∈ V, and in addition gj ··· gn ∈ UO for every j = 1, . . . , n, n ≥ 1.
Just set gj := xj+1x−1
j
, j = 1, . . . , n − 1, and gn := e.
Now, by the cocycle identity (i) we have
z(g1g2 ··· gn, θ) = αg1···gn−1(z(gn, θ))··· αg1(z(g2, θ))z(g1, θ)
(18)
Thus, z(g1, θ)ρ(αg(a))z(g1, θ)∗ = αg1(ρ(αg2···gn(a))) because g1, g1
and we want to compute its adjoint action on ρ(αg(a)), a ∈ A(O).
−1 ∈ V ⊂ UO, hence
equality (ii) holds on every A( O), O ∈ K, such that O ∪ g1O ⊂ O and g−1
1 ∈ U O. Moreover,
1 g = g2 ··· gn ∈ UO, hence we can assume that gO ⊂ O,
g−1
1 ∈ UgO = UOg−1 because g−1
by enlarging O if necessary, because K is P-stably directed by assumption (Definition 7.1).
Continuing, z(g2, θ)ρ(αg2···gn(a))z(g2, θ)∗ = αg2(ρ(αg3···gn(a))) because g2, g2
−1 ∈ V ⊂ UO,
hence we can choose O ∈ K as in (ii) such that g−1
2 ∈ U O and again further assume that
g2 ··· gnO ⊂ O. Indeed, g−1
2 because g3 ··· gn ∈ UO. By finite
iteration we get the first claim.
By the cocycle identity (i), the unitaries Uρ(g) := z(g, ρ)∗U (g), g ∈ P, form a repre-
sentation of P. Indeed, Uρ(g)Uρ(h) = z(g, ρ)∗αg(z(h, ρ))∗U (gh) = Uρ(gh) and z(g−1, ρ) =
αg−1(z(g, ρ)∗), hence also Uρ(g−1) = Uρ(g)∗, follow from z(e, ρ) = 1. We want to show that
it implements the covariance of ρ.
2 ∈ Ug2···gnO = UOg−1
n ··· g−1
Let a ∈ A( O), g ∈ U O for an arbitrary O ∈ K. Define V := U O ∩ U −1
and consider
W := V ∩ V, or any other symmetric neighborhood W of e such that W ⊂ UO ∩ U O. By the
O
30
same argument as above, we have g ∈ U ( O) = W(e), i.e., we can write g = g1 ··· gn, where
gj ∈ W and gj ··· gn ∈ U O for every j = 1, . . . , n, n ≥ 1, and
Uρ(g)ρ(a)Uρ(g)∗ = Uρ(g1)··· Uρ(gn)ρ(a)Uρ(gn)∗ ··· Uρ(g1)∗.
Now, gn ∈ W ⊂ UO ∩U O hence there is O1 ∈ K such that O ∪ O ⊂ O1, gn ∈ UO1, moreover ρ
is localized in O1, a ∈ A(O1), thus by the first claim we get Uρ(gn)ρ(a)Uρ(gn)∗ = ρ(αgn(a)).
Continuing, gn−1 ∈ UO ∩Ugn O and we can repeat the previous argument on O2 ∈ K such that
O ∪ gn O ⊂ O2, gn−1 ∈ UO2, to get Uρ(gn−1)ρ(αgn(a))Uρ(g∗
n−1) = ρ(αgn−1gn(a)). By finite
iteration we get
Uρ(g)ρ(a)Uρ(g)∗ = ρ(αg(a))
where a ∈ A( O), g ∈ U O for an arbitrary O ∈ K, completing the proof.
With similar arguments one can extend naturality and tensoriality of cocycles to (almost
all) g ∈ P, namely
Lemma 7.6. In the assumptions of Definition 7.3, let z(·,·) be a map fulfilling the properties
(i) and (iv), then the following holds as well
(iv)′
αg(t) = z(g, σ)tz(g, ρ)∗,
g ∈ P.
If z(·,·) fulfills (i), (ii), (iii) and (v), then it fulfills also
(v)′
z(g, ρσ) = z(g, ρ)ρ(z(g, σ)),
where O2 ∈ K is a DHR localization region of σ.
g ∈ UO2,
Proof. Let O1,O2 ∈ K be respectively localization regions of ρ, σ. To prove the first state-
ment, write g ∈ P as g = g1 ··· gn, n ≥ 1, where gj ∈ UO1 ∩ UO2, j = 1, . . . , n. Then make
use of equation (18) and apply (iv) at each step.
To prove the second statement, write g ∈ UO2 as before, and assume in addition that
gj ··· gn ∈ UO2, j = 1, . . . , n, cf. proof of Lemma 7.5. Then make again use of equation
(18) for z(g, ρσ) and apply (v) for each gj ∈ UO1 ∩ UO2. Repeated use of Lemma 7.5 gives
the desired conclusion. Notice that z(g, σ) belongs to the quasilocal algebra A because of
assumption (iii), hence one can safely apply the endomorphisms ρ.
Now we show that the properties (mainly tensoriality and naturality) of covariance co-
cycles expressed by the equivariance of the action of spacetime symmetries on the DHR
category ensure covariance of the extended nets constructed as in Theorem 6.8.
Theorem 7.7. Let {A} be a local net fulfilling Haag duality, standardly realized on H0, and
covariant with respect to a group of spacetime symmetries P (Definition 7.1). Assume in
addition that either P acts transitively on K (i.e., for every O1,O2 ∈ K there is g ∈ UO1 such
that gO1 = O2), or P preserves K (i.e., gO ∈ K for every g ∈ P, O ∈ K), 6.
Then an extension {B} of {A} constructed as in Theorem 6.8 from a unital generalized
net Q-system of intertwiners (θ, w,{mi}) is automatically covariant, provided that P has an
equivariant action on a tensor subcategory C ⊂ DHR{A} (i.e., CP = C) which contains θ
(Definition 7.3).
6This assumption is needed to obtain covariance of {B} on all the regions in K, cf. footnote after equation
+ preserving double cones in R3+1.
(16). Examples are gMob acting transitively on bounded intervals in R, or P ↑
31
Proof. Let (θ, w,{mi}), i ∈ I be a generalized net Q-system of intertwiners in DHR{A} and
construct the extension {A ⊂ B} as in Theorem 6.8. Here O ∈ K is a fixed localization region
for θ and N = A(O). In the following we denote by H := Hϕ the Hilbert space of {B}, we
identify H0 = eNH and a ∈ A, Mi ∈ B(O) with their images under πϕ in B(H). Thus
i H0
where every ψ ∈ H can be written as ψ = Pi M ∗
i ψi, with ψi ∈ qiH0. Moreover, we have
Mia = θ(a)Mi for every a ∈ A, i ∈ I. Having full control of the Hilbert space thanks to the
Pimsner-Popa condition, we can set
H =Xi
M ∗
U (g)ψ :=Xi
M ∗
i z(g, θ)∗U (g)ψi
for every g ∈ P, ψ ∈ H, where U implements the covariance of {A} on H0 and z(·, θ) is the
covariance cocycle of θ given by equivariance. By definition of αg and by the cocycle identity
we have z(g, θ)∗U (g)z(h, θ)∗U (h) = z(g, θ)∗αg(z(h, θ))∗U (g)U (h) = z(gh, θ)∗U (gh), hence U
is a representation of P on H, which is strongly continuous and unitary as one can easily
check.
In order to show that U implements covariance of {B} with respect to P, take first
a ∈ A(O), where O is as above, take g ∈ UO, see Definition 7.1, ψ ∈ H, and compute
M ∗
U (g)a U (g)∗ψ =Xi
i z(g, θ)∗αg(θ(a))z(g, θ)ψi =Xi
M ∗
=Xi
i z(g, θ)∗U (g)θ(a)U (g)∗z(g, θ)ψi
M ∗
i θ(αg(a))ψi = αg(a)ψ,
where the third equality follows from Lemma 7.5. Take now Mi ∈ B(O), i ∈ I, then
U (g)Mi U (g)∗ψ =Xj,k
M ∗
k z(g, θ)∗U (g)lki
j U (g)∗z(g, θ)ψj
where the coefficients lki
cocycles, see the property (iv) in Definition 7.3, and because lki
have αg(lki
the property (v) in Definition 7.3, we have that z(g, θ2) = z(g, θ)θ(z(g, θ)), hence
j ∈ N = A(O) are those given in equation (15). By naturality of
j ∈ HomDHR{A}(θ, θ2), we
j z(g, θ)∗. Moreover, by tensoriality of cocycles, see
j U (g)∗ = z(g, θ2)lki
j ) ≡ U (g)lki
=Xj,k
M ∗
k θ(z(g, θ))lki
j ψj = z(g, θ)Miψ.
The global localization property (vi) in Definition 7.3 implies that z(g, θ) is a unitary charge
transporter for θ from O to gO ∈ K, hence z(g, θ)Mi ∈ B(gO) by definition (16) of the local
algebras, and we conclude
U (g)B(O) U (g)∗ = B(gO).
Now, covariance for arbitrary regions O ∈ K and g ∈ U O follows either by transitivity of P
on K (trivially), or because P preserves K, in which case U O = P and we can meaningfully
write AdUθ(g) θ(u) = θ(αg(u)), u ∈ HomDHR{A}(θ, θ) and θ(z(g, θ)) in A, cf. Lemma 7.5, 7.6.
32
0 wΩ0, indeed U (g)Ω = M ∗
Positivity of the energy-momentum spectrum holds because Uθ has positive spectrum,
indeed θ is a (possibly infinite) direct sum of covariant endomorphisms fulfilling the spectrum
condition, see [DHR74, Thm. 5.2]. The P-invariance of the vacuum vector Ω := Ωϕ follows
from Ω = M ∗
0 z(g, θ)∗αg(w)Ω0 = Ω by
naturality (iv) of the action of g on w ∈ HomDHR{A}(id, θ) and because z(g, id) = 1. Thus
the extended net {B} is covariant as in Definition 7.1.
Remark 7.8. If the quasilocal algebra A together with the elements of the Pimsner-Popa
basis Mi, i ∈ I, form a *-algebra of charged intertwiners in the sense of Remark 6.9, one can
try to define covariance of the extension B (at the *-algebra level) by postulating αg := αg
on A and αg(Mi) := z(g, θ)Mi. In this case as well, naturality and tensoriality of the cocycle
z guarantee that αg is *-multiplicative.
0 z(g, θ)∗U (g)wΩ0 = M ∗
Next, we show how equivariance holds, in the sense of Definition 7.3, for the action of
some typical spacetime symmetry groups on the DHR category in different dimensions. More
precisely, we consider here the subcategory C = DHRd{A} of DHR{A} (Definition 6.2) which
is relevant for finite index or infinite index discrete extensions treated in Theorem 6.8.
Example 7.9. (Mobius covariant nets in 1D). Let P = gMob the universal covering of the
Mobius group and K = {open proper bounded intervals I ⊂ R}. Consider a local P-covariant
net {A} over K as in Definition 6.1, 7.1, fulfilling Haag duality on R, namely A(I ′)′ = A(I),
I ∈ K, I ′ = R r ¯I. By locality and P-covariance we have U (Rot2π) = 1 [GL96, Thm.
1.1], hence {A} is automatically Mob-covariant and we can extend it to a net { A} over the
open proper intervals of S1, see [CKL08, Prop. 16, Cor. 17]. The extension coincides with
the one given by A(I) := A(S1 r ¯I)′ if I ⊂ S1 contains the point at infinity in its closure
and A(I) := A(I) otherwise, see [KLM01, Lem. 49]. Moreover, every endomorphism ρ in
DHR{A} extends to a representation {πI , I ⊂ S1} of { A} on H0 such that
πI = ρ↾A(I)
if I is identified to a bounded interval of R via the Cayley map, see [KLM01, Prop. 50]. The
Bisognano-Wichmann property [GL96, Prop. 1.1] and strong additivity [GLW98, Lem. 1.3]
ensure that finite-dimensional DHR endomorphisms are covariant (with positive energy) with
respect to P, see [GL92, Thm. 5.2]. For every ρ in DHRf{A}, following [GL92, Prop. 8.2],
we can define by equation (18) the cocycle z(·, ρ). The definition is well posed (in B(H0))
by [GL92, Eq. (8.5)] because any two chains in P are homotopic by simple connectedness
of P, see [BP01, Def. 45, Lem. 46] for more details. Thus the properties (i), (ii), (iii) of
Definition 7.3 hold, see also (ii)′ of Lemma 7.5. The property (vi) holds by additivity [FJ96,
Sec. 3] while (iv) and (v) can be derived from the results of [Lon97]. Indeed, let ρ and σ
in DHR{A} and choose a common localization interval I ∈ K. Let I1 ∈ K be such that
¯I ⊂ I1 and I2, I3 ∈ K such that {Ii, i = 1, 2, 3} is a partition of S1 obtained by removing
three distinct points and counterclockwise ordered. Let V ⊂ UI ∩U −1
I be an arbitrarily small
symmetric neighborhood of e in P, whose elements g can be written as products of dilations
ΛIi associated to Ii, i = 1, 2, 3 such that in addition Λj
Iij ··· Λn
map I inside I1 for
Iij
every j = 1, . . . , n and g = Λ1
. Thus at each step we can consider I as a subinterval
of either I1, or S1 r ¯I2, or S1 r ¯I3. Observe that the dilations with respect to any such
partition of S1 into three intervals generate P, see [GLW98, Lem. 1.1]. Now, with this choice
of V, cf. [Lon97, Lem. 2.2], for every g ∈ V we have
Ii1 ··· Λn
Iin
and Λj
Iin
αg(t) = z(g, σ)tz(g, ρ)∗
33
if t ∈ HomDHR{A}(ρ, σ), and
z(g, ρσ) = z(g, ρ)ρ(z(g, σ)).
Building suitable V-chains in P and reasoning as in the proof of Lemma 7.6, one can show
the properties (iv) and (v)7 in their global formulation (iv)′ and (v)′ of Lemma 7.6.
Now, if ρ is in DHRd{A} let {wi} be a (possibly infinite) Cuntz family of isometries
i ∈ ρ(A)′ ∩ A(I) are mutually orthogonal
ρ =: ρi are irreducible DHR endomorphisms of finite-
in A(I), for I ∈ K big enough, such that wiw∗
projections, Pi wiw∗
dimension. For every g ∈ P
i = 1 and Adw∗
i
z(g, ρ) :=Xi
αg(wi)z(g, ρi)w∗
i
converges in the strong operator topology and extends the definition given in DHRf{A} by
[Lon97, Prop. 1.3, Eq. (1.13)]. Let C := DHRd{A}, the unitaries z(g, ρ), g ∈ P, ρ in C form
again a cocycle map, as one can check directly on each direct summand of ρ = ⊕iρi, hence
the action of P on C is equivariant in the sense of Definition 7.3.
Example 7.10. (Poincar´e covariant nets in 3+1D). Let P = P↑
+ the universal covering
of the Poincar´e group and K = {double cones O ⊂ R3+1}. Consider a local P-covariant
net {A} over K as in Definition 6.1, 7.1, fulfilling Haag duality on R3+1. Assume that
{A} fulfills in addition the Bisognano-Wichmann property on wedges, see [BW75], and that
local intertwiners between finite-dimensional DHR endomorphisms are global intertwiners, cf.
[Rob74, Thm. 4.3], [DMV04, Cor. 6.2]. Due to the fact that Lorentz boosts with respect to
different wedges generate P = P↑
+, we can make again use of the results of [GL92], [Lon97],
in a different geometrical situation, to draw analogous conclusions. Namely, the action of P
of C := DHRd{A}, which in this case is globally defined, see Remark 7.2, is again equivariant
in the sense of Definition 7.3.
8 Braided product of nets
In this section we apply the braided product construction to nets of von Neumann algebras
and show that it enjoys some remarkable properties, in analogy to the finite index case, which
allows one to extract boundary quantum field theories as in [BKLR16]. Such field theories
with transparent boundaries will be discussed in the next section.
EL
ER
Denote by {A
⊂ BR} two discrete relatively local extensions of the same local
net {A} (Definition 6.7) constructed from unital generalized net Q-systems of intertwiners
(θL, wL,{mL
j }) in DHR{A} as in Theorem 6.8. By Proposition 4.2 and again
Theorem 6.8 we know that there is a braided product extension {A ⊂ BL ×±
ε BR} such that
⊂ BL}, {A
i }), (θR, wR,{mR
A
⊂ ⊂
⊂
⊂
BL
BR
BL ×±
ε BR
7Tensoriality also follows by observing that P = gMob is perfect, see, e.g., [Lon97, App. A], hence the
unitary representation Uρσ which implements covariance of ρσ in DHRf {A} is unique.
34
where ε is the DHR braiding. Of course we have the analogous of Proposition 4.5, which we
rewrite below to establish notation.
We prefer to think of the net {BL ×±
ε BR} as the one constructed in Section 6.1. Let
ε BR) O} be the inductive family of nets indexed by O ∈ K (see Section 6.1), and let
{(BL ×±
Γ O be a unitary that implements (θLθR) O := θL
O
on A( O), namely
θR
O
OθR
θL
O(x) = AduL
O
θL(uR
O
) θLθR(x) = AdΓ O
(x),
x ∈ A( O).
Proposition 8.1. The maps
L, O : BL( O) → (BL ×±
ε BR) O( O) ,
L, O := AdΓ∗
O ◦ Ad(ε±
)∗ ◦ θR
O ◦ γL
O
,θR
O
θL
O
O ◦ γR
lift to embeddings 8 of {BL} and {BR} into the braided product net {BL ×±
R, O : BR( O) → (BL ×±
ε BR) O( O) ,
R, O := AdΓ∗
O ◦ θL
O
ε BR}
L : BL → BL ×±
R : BR → BL ×±
ε BR
ε BR.
Proof. It is enough to show that L/R, O(BL/R(O)) ⊂ (BL×±
L/R, O1 = (L/R, O2)↾BL/R( O1) for O1 ⊂ O2, but these follow from elementary calculations.
Proposition 8.2. Let ELR denote the distinguished conditional expectation from {BL×±
to {A} obtained as in Proposition 6.13, then
ε BR) O(O) for O ⊂ O and ι O1, O2◦
ε BR}
ELR(L(bL)R(bR)) = EL(bL)ER(bR),
bL ∈ BL,
bR ∈ BR.
Proof. Using θL(wR∗
θL,θR = wR∗
)ε±∗
θR(wR∗
θLθR(wL∗
and wR∗
θL,θR = θR(wR∗
ε±
) we get
)ε±∗
θL,θRγRγL(bL)ε±
γRγL(bL)θL(wR∗
= θLθR(wL∗
θL,θRγLγR(bR)θL(wR)wL)
γR(bR)wR)wR)
= θLθR(EL(bLER(bR))) = θLθR(EL(bL)ER(bR))
from which the proposition follows.
For the rest of the section, we assume that {A ⊂ BL}, {A ⊂ BR} are as in Proposition 6.10,
so that the generalized Q-systems (θL, wL,{mL
j }) are induced by Pimsner-
σj} ⊂ BR(O), where ρi ≺ θL, σj ≺ θR
Popa bases of global charged fields {ψL
are respectively irreducible DHR subendomorphisms with finite dimension, localized in O ∈
K.
Proposition 8.3.
i }), (θR, wR,{mR
ρi} ⊂ BL(O),{ψR
{L({ψL
ρi})R({ψR
σj})}
is a Pimsner-Popa basis for ι(A(O))
expansion (Proposition 2.3).
ELR
⊂ (BL×±
ε BR)(O) and it gives a unique Pimsner-Popa
8The same is true in the representation employed in the first proof of Theorem 6.8, but it is more lengthy
to check.
35
Proof. The first statement is immediate. To prove the second statement, it is enough to
show that ELR(L(ψL
which follows directly from a
calculation analogous to the proof of Proposition 8.2.
σj )R(ψR∗
σ′
h
)L(ψL∗
ρ′
k
ρi )R(ψR
)) = δρi,ρ′
k
δσj ,σ′
h
In the following, with abuse of notation, we shall often suppress the above embeddings
ε BR as well.
L/R, and ι : A → BL ×±
Remark 8.4. In [BKLR16] it is shown that the extension associated to the braided product
of two "ordinary" Q-systems is characterized algebraically in the following way. Let N ⊂
MA and N ⊂ MB be two finite index inclusions and let (θA, wA, xA), (θB, wB, xB) be the
associated Q-systems. Denote by ιA/B the respective inclusion maps and by θA = ⊕iρi,
θB = ⊕jσj the irreducible decompositions of the dual canonical endomorphisms. Then it
is known [BKLR15, Thm. 3.11] that MA (resp. MB) is finitely generated by N and {ψA
ρi}
(resp. {ψB
ε MB can be completely characterized as the
*-algebra freely generated by MA and MB, modulo the relations
In this case, the braided product MA ×±
σj} are charged fields.
σj}), where {ψA
ρi}, {ψB
ιA(n) = ιB(n), n ∈ N ,
ψB
σj ψA
ρiψB
σj .
ρi = ε±
ρi,σj ψA
In the discrete (infinite index) case this is no longer true since the extensions are not finitely
generated by N and the charged fields. We have to settle for a weaker form of this result,
valid for pairs of irreducible extensions, that will nevertheless prove to be useful in Section
10.
Let BL ⊂ BL be the *-algebra generated by ιL(A) and the charged fields {ψL
ε BR and {R(ψR
ε BR be the *-algebra generated by L(ιL(N )) = R(ιR(N )), {L(ψL
ilarly, let BR ⊂ BR be the *-algebra generated by ιR(A) and the charged fields {ψR
Let BL×R ⊂ BL ×±
BL ×±
Lemma 8.5. Suppose in addition that {A ⊂ BL} and {A ⊂ BR} are irreducible exten-
sions. Then BL×R is isomorphic to the *-algebra freely generated by BL and BR, modulo
the relations
ε BR. Then we have the following
σj )} ⊂ BL ×±
ρi} ⊂ BL. Sim-
σj} ⊂ BR.
ρi)} ⊂
ιL(a) = ιR(a),
ρi = ε±
ψR
σj ψL
ρi,σj ψL
a ∈ A,
ρiψR
σj .
relations can be written as a finite sum x = P nρi,σj ψL
Proof. An arbitrary element x of the free *-algebra generated by BL and BR modulo these
σj in a unique way. The same is
true for any element in BL×R by Proposition 8.3 and Remark 6.9, thus the expansion yields
an isomorphism.
ρiψR
In [BKLR16] it was shown that the center of the braided product extension is an object
of great interest since it contains all the information on transparent boundary conditions
between the two starting quantum field theories. We here show that in the discrete case
some relevant structural features are retained, in particular that the center of the braided
product extension agrees with the relative commutant, which will be useful in the next section
for the construction of irreducible phase boundaries from the central decomposition of the
braided product.
The expansion in terms of the Pimsner-Popa basis of charged fields (Proposition 8.3) can
be used to characterize the relative commutant.
36
Lemma 8.6. For every x ∈ (BL ×±
ε BR)(O) we have
x ∈ (BL ×±
ε BR)(O) ∩ A(O)′ ⇔ x = Xρi,σj
σj x) ∈ HomEnd0(A(O))(id, ρiσj).
with rρi,σj := ELR(ψL
Proof. It is enough to use the uniqueness of the expansion in Proposition 8.3
ρiψR
ψR
σj
∗
∗
ψL
ρi
rρi,σj
∗
∗
ψL
ρi
ψR
σj
ρiψR
ELR(ψL
nx =Xi,j
=Xi,j
σj nx) =Xi,j
σj xn) =Xi,j
for every n ∈ A(O), thus rρi,σj n = ρi(σj(n)) rρi,σj .
ELR(ψL
ρiψR
ψR
σj
ψL
ρi
∗
∗
ψR
ρi
∗
ψL
σj
∗
ρiσj(n)ELR(ψL
ρiψR
σj x)
ψR
σj
∗
∗
ψL
ρi
ELR(ψL
ρiψR
σj x)n = xn
As in the finite index case, cf. [BKLR16, Prop. 4.19], the center of the braided product
of two local extensions coincides with the relative commutant of {A} in the braided product.
Proposition 8.7. Suppose in addition that {A ⊂ BL} and {A ⊂ BR} are local extensions,
then
(BL ×±
ε BR)(O) ∩ A(O)′ = (BL ×±
ε BR)(O)′.
Proof. Let us first verify that the von Neumann algebra Z generated by r∗
rρi,σj ∈ HomEnd0(A(O))(id, ρiσj), is contained in the center. For every n ∈ A(O) we have
ε BR)(O) ∩ (BL ×±
ρi,σj ψL
ρiψR
σj , with
r∗
ρi,σj ψL
ρiψR
σj nψL
ρ′
k
ψR
σ′
t
= nr∗
ρi,σj ρi(ε±
ρ′
k,σj
)ε±
ρ′
k,ρi
kρi(ε±
ρ′
σ′
t,σj
)ρ′
k(ε±
σ′
t,ρi
)ψL
ρ′
k
ψR
σ′
t
ψL
ρiψR
σj
by direct computation and using locality of {BL} and {BR}, i.e.
ψL
ρ′
k
ρi = ε±
ψL
ψL
ρiψL
ρ′
k
,
ρi,ρ′
k
ψR
σ′
t
σj = ε±
ψR
σj ,σ′
t
ψR
σj ψR
σ′
t
,
cf. Theorem 6.8. Now, it is easy to see that
ρi,σj ρi(ε±
r∗
ρ′
k,σj
)ε±
ρ′
k,ρi
kρi(ε±
ρ′
σ′
t,σj
)ρ′
k(ε±
σ′
t,ρi
) = ρ′
kσ′
t(r∗
ρi,σj )
from which r∗
ρi,σj ψL
ρiψR
σj is contained in the center of the braided product.
For brevity, in the following we denote B ∩ A′ = (BL ×±
ε BR)(O)′, and consider the inclusions
ε BR)(O) ∩ (BL ×±
(BL ×±
ε BR)(O) ∩ A(O)′ and B ∩ B′ =
Z ⊂ B ∩ B′ ⊂ B ∩ A′.
If we take the GNS representation of B ∩ A′ with respect to the vacuum Ω, we get a cyclic
and separating vector for B∩A′ and for B∩B′ as well, by Lemma 8.6. Now we check that the
canonical conjugations of B ∩ A′ and B ∩ B′ with respect to Ω agree. This holds because the
Tomita operator S of (B ∩ A′, Ω), i.e., the closure of the operator S0 : xΩ → x∗Ω, x ∈ B ∩ A′,
is an extension of the Tomita operator of (B ∩ B′, Ω). Since the latter is continuous and
defined on all the GNS Hilbert space (because B ∩ B′ is abelian), the two operators agree
and coincide with the respective canonical conjugations. Thus
J(B ∩ A′)J = (B ∩ A′)′ ⊂ (B ∩ B′)′ = J(B ∩ B′)J ⊂ J(B ∩ A′)J
from which the result follows.
37
Lastly, as an application of Theorem 7.7, we show covariance of the braided product net.
Proposition 8.8. Let {A} be a local net, covariant with respect to P as in the assumptions
of Theorem 7.7. Let {BL} and {BR} be two extensions of {A} constructed as in Theorem
6.8 from unital generalized net Q-systems of intertwiners (θL, wL,{mL
i }) and (θR, wR,{mR
j }).
Assume that P acts equivariantly on two tensor subcategories CL and CR of DHR{A} which
contain respectively θL and θR. Then the braided product net {BL ×±
ε BR} is also covariant
with respect to P. Moreover, the embeddings L and R given in Proposition 8.1 are covariant
as representations, namely L ◦ αL
g ◦ R(bR), where
g (bL) = αLR
bL/R ∈ BL/R(O), for every g ∈ UO and O ∈ K.
Proof. P acts equivariantly on θLθR and on the (full, replete) tensor subcategory D generated
in DHR{A} by θL and θR. Indeed, the cocycle given by z(g, θLθR) = z(g, θL)θL(z(g, θR)),
g ∈ P, is manifestly natural and tensor in D, hence we can apply Theorem 7.7.
i )) and αLR
j )) =
R(αR
i )) = L(αL
j )) by direct computation using naturality of cocycles.
g ◦ L(bL) and R ◦ αR
The second statement follows from αLR
g (M R
g
g (bR) = αLR
(L(M L
g (M L
(R(M R
g
9 Applications to phase boundaries in QFT
The main application in QFT for the braided product of ordinary Q-systems in [BKLR16] is
the construction and classification of phase boundary QFTs.
A boundary is simply a time-like hypersurface of codimension 1 in Minkowski spacetime
Rn+1, n ≥ 1, or a point in R. Perhaps the simplest type of boundary QFT is a system in a
one-sided box. Namely, on one side of the boundary (the side of the box) there is a physical
system described by bulk fields, while on the other side there is no physical content. This
situation is usually referred to as a hard boundary, or reflective boundary.
In the following we will be concerned with phase boundaries, also called transmissive
boundaries, which describe QFTs sharing some distinguished chiral fields across the boundary
(for example the stress-energy tensor) but the field content may in general be different on
the two opposite sides. If the common fields which are not affected by the presence of the
boundary include the stress-energy tensor, then the bulk fields may be defined by covariance
on all Minkowski spacetime. Of course they do not represent physically meaningful quantities
when they are transported to the opposite side of the boundary. In any case, this observation
is crucial for the meaningfulness of the following definition.
Let {A} be a local net and let {A ⊂ BL}, {A ⊂ BR} be two local extensions (see Definition
6.7). Let ιL and ιR be the corresponding embeddings. ML and MR denote the two portions
of Minkowski spacetime determined by the boundary.
Definition 9.1. A phase boundary condition (for short phase boundary) between two
local extensions {A ⊂ BL}, {A ⊂ BR} is a pair of locally normal representations πL and πR
of the nets {BL} and {BR}, respectively, on a common Hilbert space H, with the following
properties. They agree when restricted to the common subnet A, namely
πL ◦ ιL = πR ◦ ιR
and, for O1 ⊂ ML, O2 ⊂ MR, and O1,O2 in relative space-like position, πL(BL(O1)) and
πR(BR(O2)) commute, i.e., they respect locality across the boundary.
38
A phase boundary is called irreducible if the inclusions
πL/R ◦ ιL/R(A(O)) ⊂ πL(BL(O)) ∨ πL(BR(O))
are irreducible for every O ∈ K.
In the present setting, we show that the braided product can be decomposed over its
center (in general as a direct integral) and its components give rise to irreducible phase
boundaries, in analogy to the finite index case.
Remark 9.2. A prominent feature of the finite index case is that the phase boundaries found
within the braided product net by central decomposition do exhaust the set of all possible
irreducible phase boundaries modulo unitary equivalence. The proof of the latter heavily
relies on the finiteness of the index since this insures that the braided product construction
can be completely determined algebraically as the free ∗-algebra generated by the starting
nets {BL} and {BR} modulo relations as in Remark 8.4. This makes the braided product
a universal object in the sense that every irreducible phase boundary condition arises a
representation of the former [BKLR16, Prop. 5.1].
In the infinite index setting this is no
longer the case as we will see in Section 10.
For ease of exposition, we state the results for chiral CFTs (and thus phase boundaries
in 1D) although the analysis can be extended to greater generality without difficulty. More-
over, in order to avoid inconvenient technicalities with disintegration theory, we assume that
the starting local extensions have the split property, [DL84]. This assumption is not too
restrictive since most interesting models in QFT have this property, in particular all chiral
diffeomorphism covariant models [MTW16].
Let {A} be a local conformal net (Mobius covariant, see Definition 7.1) on R over a
separable Hilbert space and satisfying Haag duality on R. Exactly as in the notation of
[KLM01, Prop. 55], for I, I ∈ K (here K is the set of open proper bounded intervals of R),
I ⊂⊂ I means that ¯I ⊂ I. If {A} has the split property, then, for each pair of intervals
I ⊂⊂ I, there is an intermediate type I factor A(I) ⊂ N (I, I) ⊂ A( I) and we denote by
K(I, I) the compact operators of N (I, I). IQ is the set of intervals with rational endpoints
and A is the separable C ∗-subalgebra of A generated by all K(I, I) with I ⊂⊂ I, I, I ∈ IQ.
Proposition 9.3. [KLM01]. Let π be a locally normal representation of A. Then π↾A is a
representation of A and π↾K(I, I) is non-degenerate for every pair of intervals I ⊂⊂ I.
Conversely, if σ is a representation of A such that σ↾K(I, I) is non-degenerate for all
intervals I, I ∈ IQ, I ⊂⊂ I, there exists a unique locally normal representation σ of A that
extends σ. Moreover, equivalent representations of A correspond to equivalent representations
of A.
Now, let {BL} and {BR} be local conformal nets extending {A} as in Definition 6.7. As-
sume that {BL/R} have the split property and that {A ⊂ BL/R} are discrete irreducible ex-
tensions with corresponding unital generalized net Q-systems of intertwiners (θL, wL,{mL
i }),
(θR, wR,{mR
Define KL(I, I), KR(I, I), and the separable C ∗-algebras BL, BR as above. Using the
last proposition, we want to show that the embedding homomorphisms L and R into the
braided product (Proposition 8.1) can be decomposed as representations with respect to the
j }) given by global charged fields as in Proposition 6.10.
39
center of the braided product. Note that by Proposition 8.7 and 8.8 the centers of the local
algebras of the braided product agree, namely we have
Z((BL ×±
ε BR)(I)) = Z((BL ×±
ε BR)(J)) = Z(BL ×±
ε BR)
for every I, J ∈ K.
Proposition 9.4. Let
L
R
↾BL ∼=Z ⊕
↾BR ∼=Z ⊕
X
X
L
λ dµ(λ)
R
λ dµ(λ)
be the disintegration of the restrictions of the embeddings L, R to the separable C ∗-
ε BR) ∼=
λ lift to locally normal represen-
subalgebras BL and BR with respect to the center of the braided product Z(BL ×±
L∞(X, dµ). Then, for dµ-almost every λ ∈ X, the L
λ and R
tations of the quasilocal C ∗-algebras BL and BR respectively.
Proof. To prove the assertion, by the above proposition, it is enough to show that there is
a dµ-null set E such that L
λ ↾KR(I, I)) is non-degenerate for every λ /∈ E and
I, I ∈ IQ, I ⊂⊂ I. This is easily checked, because for fixed I, I, L
) is
non-degenerate by Proposition 9.3 and consequently L
λ ↾KL(I, I)) is also non-
degenerate for dµ-almost every λ ∈ X. Since I, I ∈ IQ, I ⊂⊂ I are countable, the statement
follows.
λ ↾KL(I, I) (resp. R
λ ↾KL(I, I) (resp. L
(resp. R
↾KR(I, I)
↾KL(I, I)
Proposition 9.5. Let λ ∈ X r E as above. Then
(1) R
λ ◦ ιR = L
λ (mL
λ (mR
λ ◦ ιL.
i ) = L
j )L
λ (ε±
(2) R
θL,θR)L
λ (mL
i )R
λ (mR
j ).
(3) If U (g) = RX
Uλ(g)dµ is the disintegration of the representation of the universal cov-
ering of the Mobius group (given by Proposition 8.8) with respect to the center of the
braided product, then
Uλ(g)L
Uλ(g)R
λ (BL(J)) Uλ(g)∗ = L
λ (BR(J)) Uλ(g)∗ = R
λ (BL(gJ))
λ (BR(gJ)).
(4) Let Ω =RX Ωλdµ, then Uλ(g)Ωλ = Ωλ and Ωλ is cyclic for
λ (BR(J)).
λ (BL(J)) ∨ R
L
_J
λ (BR(I)) is a factor.
λ (ιL(A(I))) ⊂ L
(5) L
λ (BL(I)) ∨ R
(6) The inclusion L
λ (BL(I)) ∨ R
λ (BR(I)) is irreducible.
Proof. Most of these assertions are trivial and follow from standard techniques in disintegra-
tion theory. Covariance, i.e., point (3), follows by Proposition 8.8, Example 7.9 and by the
fact that U (g) ∈ Z(BL ×±
ε BR)′ by the expansion in Lemma 8.6.
40
Remark 9.6. Proposition 9.5 shows that the braided product construction of two net exten-
sions {A ⊂ BL}, {A ⊂ BR} with the required properties induces, via central decomposition,
a family of irreducible phase boundaries (L
λ ) indexed by the spectrum X (up to a measure
zero set) of the center of BL ×±
Of course, depending on whether {BL} and {BR} are interpreted to be theories respec-
tively on the left and on the right of the boundary, or vice versa, one has to take the braided
product with the correct sign, namely with ε+ or ε−.
ε BR and living on the Hilbert space Hλ, λ ∈ X.
λ , R
10 An example with the U (1)-current
In this section we work out concretely the braided product between local extensions of the
U (1)-current net. We will see examples where the center of the braided product net is a
continuous algebra and therefore the direct integral representation as in Proposition 9.4 does
not reduce to a direct sum. This shows in particular that the braided product is not a
universal object in the sense of [BKLR16, Prop. 5.1]. This behaviour is expected, since, as in
the finite index case, phase boundary conditions for orbifold theories should be determined
by their gauge group, see [BKLR16, Sec. 6.2]. We will show the manifestation of this fact in
at least one example.
For the definition of the U (1)-current {AU (1)} we refer to [BMT88], [GLW98], [Lon08],
and to [DV17, Ch. 12] for more detailed calculations. Let I be a proper interval of S1 r {1}
and let f ∈ C ∞(S1, R) with support contained in I. Define the net representation {ρf,J}J
first on Weyl operators W (g) in the following way
ρf,J (W (g)) := ei R f (θ)g(θ) dθ
2π W (g)
for g ∈ C ∞(S1, R) with support in a proper interval J of S1 r{1}. These above defined maps
are locally unitarily implemented: let I0 be a proper interval of S1 r {1} disjoint from I and
J, and let f0 ∈ C ∞(S1, R) with support in I0 and such that RS1 f = RS1 f0. Define LI→I0 as
a primitive of f0 − f , namely L′
I→I0 = f0 − f . It is an easy calculation to show that
W (LI→I0)W (g)W (LI→I0)∗ = ρf,J (W (g)).
Thus the maps {ρf,J}J can be extended in a unique way to the local von Neumann algebras
and they determine a locally normal representation of {AU (1)}, which is clearly DHR. More-
over these representations are classified up to unitary equivalence by the value RS1 f which
is usually referred to as the charge, thus yielding a continuous family of irreducible DHR
sectors.
We now compute explicitly the braiding operator for the irreducible DHR representations
described above. Let ρf be localized in the interval I. If I is an interval disjoint from I and
I < I, take f ∈ C ∞(S1, R) with support in I and with same charge as f , i.e., RS1 f =RS1
f . If
we denote by u I := W (LI→ I) ∈ HomDHR{A}(ρf , ρ f ) the charge transporter between ρf and
ρ f , by definition the braiding operator ε+
ρf ,ρf is obtained by
Performing the computation we get
ρf ,ρf = u∗
ε+
ρf (u I ).
I
(u I ) = e−iπQ2
ρf ,ρf = ei R f LI→ I u∗
ε+
I
where Q is the charge of the DHR sector ρf . In particular ε+
with N ∈ N.
ρf ,ρf = 1 if and only if Q = √2πN
41
10.1 Buchholz-Mack-Todorov extensions
We here quickly review the local extensions of the U (1)-current net constructed in [BMT88].
Let ρf be a DHR automorphism of the U (1)-current net localized in the interval I as above,
such that ε+
ρf ,ρf = 1. To shorten notation denote ρ = ρf . Any such automorphism gives a
local extension of the net by a crossed product with the group Z which acts on the net as
powers of ρ. Let
H :=Mk∈Z
Hk
with Hk = H (= vacuum Hilbert space of the U (1)-current net) and let π be a representation
of the quasilocal C ∗-algebra AU (1) of the net restricted to R ∼= S1 r {1}, defined as
π : AU (1) → B( H)
π(a) :=Mk∈Z
ρk(a)
Denote by U the shift operator on H, i.e., U{ξk}k∈Z = {ξk+1}k∈Z for ξ ∈ H. It is clear
that the shift operator U implements the localized automorphism ρ in this representation
In other words U is a charged field for ρ.
U π(a)U ∗ = π(ρ(a)).
Definition 10.1. The BMT (Buchholz-Mack-Todorov) extension {Bρ} = {AU (1) ⋊ρ Z} is
the net given by
Bρ(I) := hπ(A(I)), Ui
Bρ(J) := hπ(A(J)), π(uJ )Ui .
It is an easy matter to check that this definition is well posed and the net is isotonous
(it follows directly from Haag duality of the U (1)-current net on R, i.e., strong additivity).
Locality of BMT extensions {Bρ} follows from ε+
ρ,ρ = 1, cf. Theorem 6.8. The inclusion
{A ⊂ Bρ} is clearly discrete and irreducible.
The DHR automorphisms of the U (1)-current extend to representations of the net {Bρ},
and the DHR sectors of BMT extensions were already classified in [BMT88]. We recall these
facts to establish the notation.
ρ,ρ = ε−
ρ,σ = ε−
Proposition 10.2. [BMT88]. For every DHR automorphism σ of AU (1) there are two locally
normal representations σ± of Bρ such that σ±(π(a)) = π(σ(a)), a ∈ AU (1). Moreover, σ+ =
σ− if and only if σ+ (or equivalently σ−) is a DHR representation of the net {Bρ}, if and only
ρ,σ. Otherwise σ± have solitonic localization (they are localizable in half-lines). In
if ε+
particular, there are 2N inequivalent DHR automorphisms of the net {Bρ}, where Q = √2πN
is the charge of ρ.
Proof. The automorphisms σ± can be defined by α-induction of σ for the extension {AU (1) ⊂
Bρ}, [LR95, Prop. 3.9], but we here describe them explicitly since we will need them in the
following. We first define the action of σ± on the *-algebra B generated by π(AU (1)) and
the shift U . Define
σ±(π(a)) := π(σ(a))
42
σ±(U n) := π(ε±
ρn,σ)U n
To check that this is a well defined endomorphism of the *-algebra it is enough to check that
σ±(U ∗) = σ±(U )∗,
σ±(U )σ±(π(a))σ±(U )∗ = σ±(π(ρ(a))).
The first relation is an immediate consequence of naturality of the braiding, for the second
we have
σ±(U )σ±(π(a))σ±(U )∗ = π(ε±
ρ,σ)π(ρ(σ(a)))π(ε±
= π(ε±
ρ,σ)U π(σ(a))U ∗π(ε±
ρ,σ)∗ = π(σ(ρ(a))).
ρ,σ)∗
Now, observe that for a fixed proper bounded interval J of R, the endomorphism (σ)±
restricted to Bρ(J) ∩ B is locally implemented by the unitary π(u I ) := π(W (LI→ I)) where I
is a proper bounded interval where I < J if we consider σ+ and J < I if we consider σ−, i.e.,
σ±(b) = Adπ(u I )(b) for every b ∈ Bρ(J) ∩ B. Since Bρ(J) ∩ B is ultraweakly dense in Bρ(J),
the endomorphism can be extended in a unique way consistently on every local algebra.
Regarding the localization of σ±, if J < I
σ+(π(uJ )U ) = π(σ(uJ ))π(ε+
ρ,σ)U
= π(σ(uJ ))π(σ(uJ ))∗π(uJ )U = π(uJ )U
Similarly for I < J we have the same result for σ−. Thus they are localizable in half-lines, a
priori, and also DHR if and only if ε+
ρ,σ = ε−
ρ,σ.
10.2 Braided product of BMT extensions
Let {BρL},{BρR} be two local BMT extensions of the U (1)-current net given by two DHR
automorphisms ρL and ρR as in the previous section. We would like to construct the braided
product of two such nets in a concrete fashion. Let
Ml∈Z
H = M(l,h)∈Z2 H
ιL : BρL →Ml∈Z
Let ιL be the solitonic representation of BρL defined on the above Hilbert space as follows
where H is the vacuum Hilbert space of the U (1)-current net, and ΩAU (1) is the vacuum
vector. We denote Ω = { Ωl,h}(l,h)∈Z2 , with Ωl,h := δl,0δh,0ΩAU (1).
B( H) ⊂ B( M(l,k)∈Z2 H)
ιL :=Ml∈Z
B( H) ⊂ B( M(l,k)∈Z2 H)
ιR :=Mh∈Z
ιR : BρR →Mh∈Z
and similarly for ιR
ρR
ρL
43
Define
ε : M(l,h)∈Z2 H → M(l,h)∈Z2 H
ε{ξl,h}(l,h)∈Z2 := {ε±
L,ρh
ρl
R
ξl,h}(l,h)∈Z2
and twist the representation ιR by ε
Observe that
ιR := Adε(ιR(·)).
ιL ◦ πL = ιR ◦ πR
where πL and πR are the inclusion maps of AU (1) into BρL and BρR respectively, explicitly
ρl
L(ρh
R(a))) = (ιR ◦ πR)(a)
(ιL ◦ πL)(a) = Mk,l∈Z
ρh
R(ρl
L(a)) = Adε(Mk,l∈Z
for every a ∈ AU (1). Let U ∈ BρL(I) and V ∈ BρR(I) be the charged fields for the DHR
automorphisms ρL and ρR respectively. Then
Proposition 10.3.
ιR(V )ιL(U ) = ιL(πL(ε±
ρL,ρR)) ιL(U )ιR(V )
Proof. By direct computation.
Proposition 10.4. Let {BρL} and {BρR} two local BMT extensions as above. The net of
von Neumann algebras defined by
B±(I) :=(cid:10)ιL ◦ πL(AU (1)(I)), ιL(U ), ιR(V )(cid:11) ,
B±(J) :=(cid:10)ιL ◦ πL(AU (1)(J)), ιL ◦ πL(uJ )ιL(U ), ιR ◦ πR(vJ )ιR(V )(cid:11) ,
where uJ and vJ are unitary charge transporters respectively for ρL and ρR between intervals I
and J (i.e. the endomorphisms AduJ ρL and AdvJ ρR are localized in J), is unitarily equivalent
to the braided product net, i.e.
{ B±} ∼= {BρL ×±
ε BρR}.
Proof. By Lemma 8.5, Proposition 10.3 and the relation ιL ◦ πL = ιR ◦ πR, we know that
there exists a surjective homomorphism of *-algebras
φ : BL×R → B
ε BρR is defined as in Lemma 8.5 and B ⊂ B± is the *-algebra generated
where BL×R ⊂ BρL×±
by ιL ◦ πL(AU (1)) and ιL(U ), ιR(V ). By the GNS theorem for *-algebras, see e.g. [KM15,
Sec. 1.3], in order to show that φ is implemented by a unitary it is enough to check that
ω0 ◦ ELR = ( Ωφ(·) Ω), where Ω is the vacuum vector of { B±}. This is clear since, for
x = Pi,j L/R(xi,j)L(U i)R(V j) ∈ BL×R, we have ω0 ◦ ELR(x) = (ΩAU (1), x0,0ΩAU (1)) =
( Ω, φ(·) Ω).
44
By considering the braided product of a local BMT extension with itself (as concretely
constructed in the previous proposition by taking ρL = ρR = ρ) we give examples where the
center of the braided product is a continuous algebra, more specifically L∞(S1, dµ).
Proposition 10.5. Let {Bρ} be the BMT extension obtained from a DHR automorphism
ε Bρ) ∼=
ρ and let {Bρ ×±
L∞(S1, dµ) with dµ the Lebesgue measure on the circle.
ε Bρ} be the braided product extension with itself. Then Z(Bρ ×±
Proof. Recall that the center of the braided product is the same as the relative commutant
Z(Bρ ×±
ε Bρ) = Z((Bρ ×±
ε Bρ)(J)) = (Bρ ×±
ε Bρ)(J)′ ∩ ι(AU (1)(J))
ε Bρ)
V −iU ixi
for any proper bounded interval J of R. Thus Lemma 8.6 provides an expansion for elements
x ∈ Z(Bρ ×±
x =Xi∈Z
with xi ∈ HomDHR{AU (1)}(id, ρiρ−i = id) ∼= C.
It is easy to see that there is an isomorphism between the *-algebra generated by the
{U iV −i}i and the *-algebra generated by the characters of the circle. This same map is
also an isomorphisms of pre-Hilbert spaces with inner product on one side induced by the
vacuum state ω = ω0 ◦ ELR, where ELR is the standard expectation of the braided product
net (Proposition 8.2) and ω0 the vacuum state for {AU (1)}, and on the other side the usual
L2(S1, dµ) inner product.
Thus let B denote the *-algebra generated by the {U iV −i}i, ¯Bk·kω its Hilbert completion
and let Char(S1) be the *-algebra generated by characters of the circle. ¯Bk·kω ∼= L2(S1, dµ)
as Hilbert spaces and let W be the unitary which implements the isomorphism. If πω is the
GNS representation of B induced by the state ω = ω0◦ E (on the Hilbert space ¯Bk·kω ), and if
πdµ is the GNS representation of Char(S1), we have AdW πω = πdµ. Hence the isomorphism
extends to the ultraweak closure, and
Z(Bρ ×±
ε Bρ)(I) ∼= πω(Z(Bρ ×±
ε Bρ)) = πω(B)′′ ∼= πdµ(Char(S1))′′ ∼= L∞(S1, dµ)
concluding the proof.
We thus have an example of an uncountable family of (one-dimensional) irreducible phase
boundaries, parametrized by S1, obtained from the braided product construction. This is
obviously in contrast with the finite index case, where the relative commutant is necessarily
finite-dimensional. But the difference from the finite index case is actually greater than this:
we have an example where the relative commutant is not a discrete algebra. This means that
the disintegration in Proposition 9.5 that yields irreducible phase boundaries is not a direct
sum. Moreover it is not true, in contrast with the finite index case, that every irreducible
phase boundary condition comes from a representation of the braided product extension, see
[BKLR16, Prop. 5.1, Cor. 5.3], due to the absence of non-trivial minimal central projections.
Similarly, one can construct examples where the braided product is itself an irreducible
extension and thus it yields a unique irreducible phase boundary. It is not hard to see that
this is the case for the braided product of two local BMT extensions of the U (1)-current
whose generating DHR automorphisms ρf1, ρf2 have charges RS1 f1 and RS1 f2 with irrational
quotient. The claim simply follows from the expansion given in Lemma 8.6 and observing
that, in this case, the dual canonical endomorphisms of the BMT extensions θ1 and θ2 have
only one irreducible subendomorphism in common: the identity.
45
11 Conclusions
Index theory provides an elegant and effective machinery to classify and construct extensions
of von Neumann algebras and local nets. When this framework is not fully applicable (infi-
nite index case), we have seen that under some physically meaningful structural hypotheses
(semidiscreteness, discreteness) some of these results can be suitably generalized. The price
to pay is abandoning the purely categorical setting of finite index Q-systems by the emer-
gence of analytical conditions. At the same time, these analytical conditions (convergence
of projections, faithfulness of expectations) provide a way to control infinite objects (gauge
groups, representation categories, sets of generating fields) exploiting techniques of Operator
Algebras in their application to QFT.
In particular, we have introduced the notion of generalized Q-system of intertwiners (in
the category of localizable superselection sectors DHR{A}) for a local net {A}, and we have
shown that from this data a net extension of {A}, in the spirit of [LR95], can be constructed.
At the level of properly infinite inclusions, we have seen that the existence of generalized Q-
systems of intertwiners is equivalent to the inclusion to be of discrete type. When passing from
subfactors to inclusions of local nets as in [LR95] this matter is more subtle, and we provided
sufficient conditions to guarantee the existence of generalized Q-systems of intertwiners for
nets, which cover most interesting examples in low and higher spacetime dimensions. We
leave open the question on whether these conditions are always verified by discrete QFT
extensions.
The notion of generalized Q-system of intertwiners lends itself to generalize the definition
of braided product between ordinary Q-systems. After proving that the analytic properties
of generalized Q-systems of intertwiners turn out to be compatible with the purely algebraic
definition of the braided product, we explore some properties of the resulting net extension,
showing that it retains some features of its finite index counterpart. In particular, in the case
of chiral CFTs, we have seen that its central decomposition can yield uncountable families
of irreducible phase boundaries with infinite index. An important issue left open is the
classification of all phase boundary conditions among two CFTs. In particular, one would
like to understand if, in analogy with [BKLR16], all the boundary conditions arise in the
disintegration of the center of the braided product.
Although the discrete case covers many physical examples, e.g., every orbifold construc-
tion by a compact group, the setting of greatest generality for irreducible inclusions of local
CFTs (at least assuming the existence of a vacuum vector) is semidiscreteness. Generalized
Q-systems do always exist for semidiscrete extensions of properly infinite von Neumann al-
gebras [FI99]. An issue that would be worth analyzing further is if methods similar to those
explored in this paper can be generalized to treat extensions of local nets which are semidis-
crete but not discrete [Car04], [Xu05]. It would also be interesting to extend the analysis of
discrete inclusions to the case of non-separable Hilbert spaces, given that good candidates
for such extensions in QFT already appear in [Cio09], [MTW16]. Lastly, we mention that
one can easily construct discrete non-finite local extensions which are not compact group
orbifolds by taking tensor products of local nets, 9.
It would also be worth investigating
which kind of extensions can arise from braided products of compact group orbifolds, given
that, by the arguments of our last section, one can construct extensions whose generating
fields have the commutation relations of non-commutative tori.
9We thank Y. Tanimoto for pointing out this interesting fact.
46
Acknowledgements. Supported by the European Research Council (ERC) through the
Advanced Grant QUEST "Quantum Algebraic Structures and Models", and by PRIN-MIUR.
We are indebted to R. Longo for proposing us the problem investigated in this work, and to
M. Bischoff and K.-H. Rehren for many discussions and suggestions, and for their motivating
interest. We also thank I. Khavkine and Y. Tanimoto for useful comments and criticism.
L.G. wishes to thank K.-H. Rehren for an invitation to Gottingen (Institut fur Theoretis-
che Physik, Georg-August-Universitat) and W. Yuan for an invitation to Beijing (Academy
of Mathematics and Systems Science, Chinese Academy of Sciences), where this work has
been presented, and thanks them for hospitality and for useful conversations in both these
occasions.
References
[AMP15] N. Afzaly, S. Morrison, and D. Penneys. The classification of subfactors with index at
most 5 1
4 . preprint arXiv:1509.00038, 2015.
[BDH88] M. Baillet, Y. Denizeau, and J.-F. Havet. Indice d'une esp´erance conditionnelle. Compo-
sitio Math., 66:199 -- 236, 1988.
[BGL93] R. Brunetti, D. Guido, and R. Longo. Modular structure and duality in conformal quantum
field theory. Comm. Math. Phys., 156:201 -- 219, 1993.
[Bis17]
M. Bischoff. Generalized orbifold construction for conformal nets. Rev. Math. Phys.,
29:1750002 1 -- 53, 2017.
[BKL15] M. Bischoff, Y. Kawahigashi, and R. Longo. Characterization of 2D rational local con-
formal nets and its boundary conditions: the maximal case. Doc. Math., 20:1137 -- 1184,
2015.
[BKLR15] M. Bischoff, Y. Kawahigashi, R. Longo, and K.-H. Rehren. Tensor categories and endo-
morphisms of von Neumann algebras. With applications to quantum field theory, Springer
Briefs in Mathematical Physics, Vol. 3. Springer, Cham, 2015.
[BKLR16] M. Bischoff, Y. Kawahigashi, R. Longo, and K.-H. Rehren. Phase Boundaries in Algebraic
Conformal QFT. Comm. Math. Phys., 342:1 -- 45, 2016.
[BMT88] D. Buchholz, G. Mack, and I. Todorov. The current algebra on the circle as a germ of
local field theories. Nucl. Phys., B, Proc. Suppl., 5:20 -- 56, 1988.
[BP01]
[BR16]
[BW75]
[Car04]
V. Berestovskii and C. Plaut. Covering group theory for topological groups. Topology
Appl., 114:141 -- 186, 2001.
M. Bischoff and K.-H. Rehren. The hypergroupoid of boundary conditions for local quan-
tum observables. preprint arXiv:1612.02972, 2016.
J. J. Bisognano and E. H. Wichmann. On the duality condition for a Hermitian scalar
field. J. Mathematical Phys., 16:985 -- 1007, 1975.
S. Carpi. On the representation theory of Virasoro nets. Comm. Math. Phys., 244:261 -- 284,
2004.
[CCG+04] A. Connes, J. Cuntz, E. Guentner, N. Higson, J. Kaminker, and J. E. Roberts. Non-
commutative geometry, Lecture Notes in Mathematics, Vol. 1831. Springer-Verlag, Berlin;
Centro Internazionale Matematico Estivo (C.I.M.E.), Florence, 2004. Lectures given at
the C.I.M.E. Summer School held in Martina Franca, September 3 -- 9, 2000, Edited by S.
Doplicher and R. Longo.
47
[CDR01] R. Conti, S. Doplicher, and J. E. Roberts. Superselection theory for subsystems. Comm.
Math. Phys., 218:263 -- 281, 2001.
[Cio09]
F. Ciolli. Massless scalar free field in 1 + 1 dimensions. I. Weyl algebras products and
superselection sectors. Rev. Math. Phys., 21:735 -- 780, 2009.
[CKL08]
[Con73]
S. Carpi, Y. Kawahigashi, and R. Longo. Structure and classification of superconformal
nets. Ann. Henri Poincar´e, 9:1069 -- 1121, 2008.
A. Connes. Une classification des facteurs de type III. Ann. Sci. ´Ecole Norm. Sup.,
6:133 -- 252, 1973.
[DHR69a] S. Doplicher, R. Haag, and J. E. Roberts. Fields, observables and gauge transformations.
I. Comm. Math. Phys., 13:1 -- 23, 1969.
[DHR69b] S. Doplicher, R. Haag, and J. E. Roberts. Fields, observables and gauge transformations.
II. Comm. Math. Phys., 15:173 -- 200, 1969.
[DHR71]
S. Doplicher, R. Haag, and J. E. Roberts. Local observables and particle statistics. I.
Comm. Math. Phys., 23:199 -- 230, 1971.
[DHR74]
S. Doplicher, R. Haag, and J. E. Roberts. Local observables and particle statistics. II.
Comm. Math. Phys., 35:49 -- 85, 1974.
[DL84]
S. Doplicher and R. Longo. Standard and split inclusions of von Neumann algebras. Invent.
Math., 75:493 -- 536, 1984.
[DMV04] C. D'Antoni, G. Morsella, and R. Verch. Scaling algebras for charged fields and short-
distance analysis for localizable and topological charges. Ann. Henri Poincar´e, 5:809 -- 870,
2004.
[DR72]
[DR89]
[DR90]
[DV17]
[EP03]
[FI95]
[FI99]
[FJ96]
S. Doplicher and J. E. Roberts. Fields, statistics and non-abelian gauge groups. Comm.
Math. Phys., 28:331 -- 348, 1972.
S. Doplicher and J. E. Roberts. Endomorphisms of C ∗-algebras, cross products and duality
for compact groups. Ann. of Math., 130:75 -- 119, 1989.
S. Doplicher and J. E. Roberts. Why there is a field algebra with a compact gauge group
describing the superselection structure in particle physics. Comm. Math. Phys., 131:51 --
107, 1990.
S. Del Vecchio. Extensions in Quantum Field Theory: Q-systems and defects for infinite
index inclusions. PhD thesis, Universit`a degli Studi di Roma Tor Vergata, Facolt`a di
Scienze Matematiche Fisiche e Naturali, 2017.
D. E. Evans and P. R. Pinto. Subfactor realisation of modular invariants. Comm. Math.
Phys., 237:309 -- 363, 2003.
F. Fidaleo and T. Isola. On the conjugate endomorphism in the infinite index case. Math.
Scand., 77:289 -- 300, 1995.
F. Fidaleo and T. Isola. The canonical endomorphism for infinite index inclusions. Z.
Anal. Anwendungen, 18:47 -- 66, 1999.
K. Fredenhagen and M. Jorss. Conformal Haag-Kastler nets, pointlike localized fields and
the existence of operator product expansions. Comm. Math. Phys., 176:541 -- 554, 1996.
[FRS92] K. Fredenhagen, K.-H. Rehren, and B. Schroer. Superselection sectors with braid group
statistics and exchange algebras. II. Geometric aspects and conformal covariance. Rev.
Math. Phys., SI1 (Special Issue):113 -- 157, 1992.
[GL92]
D. Guido and R. Longo. Relativistic invariance and charge conjugation in quantum field
theory. Comm. Math. Phys., 148:521 -- 551, 1992.
48
[GL95]
[GL96]
D. Guido and R. Longo. An algebraic spin and statistics theorem. Comm. Math. Phys.,
172:517 -- 533, 1995.
D. Guido and R. Longo. The conformal spin and statistics theorem. Comm. Math. Phys.,
181:11 -- 35, 1996.
[GLW98] D. Guido, R. Longo, and H.-W. Wiesbrock. Extensions of conformal nets and superselec-
tion structures. Comm. Math. Phys., 192:217 -- 244, 1998.
[GR15]
L. Giorgetti and K.-H. Rehren. Braided categories of endomorphisms as invariants for local
quantum field theories. preprint arXiv:1512.01995 (to appear in Comm. Math. Phys.),
2015.
[Haa87]
U. Haagerup. Connes' bicentralizer problem and uniqueness of the injective factor of type
III1. Acta Math., 158:95 -- 148, 1987.
[Haa96]
R. Haag. Local quantum physics. Springer Berlin, 1996.
[Hia88]
F. Hiai. Minimizing indices of conditional expectations onto a subfactor. Publ. Res. Inst.
Math. Sci., 24:673 -- 678, 1988.
[HKZ91] H. Halpern, V. Kaftal, and L. Zsid´o. Finite weight projections in von Neumann algebras.
[HM06]
[HO89]
Pacific J. Math., 147:81 -- 121, 1991.
H. Halvorson and M. Muger.
math-ph/0602036, 2006.
Algebraic quantum field theory.
preprint arXiv
R. H. Herman and A. Ocneanu. Index theory and Galois theory for infinite index inclusions
of factors. C. R. Acad. Sci. Paris S´er. I Math., 309:923 -- 927, 1989.
[ILP98] M. Izumi, R. Longo, and S. Popa. A Galois correspondence for compact groups of au-
tomorphisms of von Neumann algebras with a generalization to Kac algebras. J. Funct.
Anal., 155:25 -- 63, 1998.
[JMS14] V. F. R. Jones, S. Morrison, and N. Snyder. The classification of subfactors of index at
most 5. Bull. Amer. Math. Soc., 51:277 -- 327, 2014.
[Jon83]
V. F. R. Jones. Index for subfactors. Invent. Math., 72:1 -- 25, 1983.
[Jon99]
V. F. R. Jones. Planar Algebras. I. preprint arXiv math.QA/9909027, 1999.
[Kad63]
[KL04]
R. V. Kadison. Remarks on the type of von Neumann algebras of local observables in
quantum field theory. J. Mathematical Phys., 4:1511 -- 1516, 1963.
Y. Kawahigashi and R. Longo. Classification of local conformal nets. Case c < 1. Ann.
Math., 160:493 -- 522, 2004.
[KLM01] Y. Kawahigashi, R. Longo, and M. Muger. Multi-interval subfactors and modularity of
representations in conformal field theory. Comm. Math. Phys., 219:631 -- 669, 2001.
[KM15]
[Kos86]
[Kos89]
[Lon79]
I. Khavkine and V. Moretti. Algebraic QFT in curved spacetime and quasifree Hadamard
states: an introduction. In Advances in algebraic quantum field theory, Math. Phys. Stud.,
191 -- 251. Springer, Cham, 2015.
H. Kosaki. Extension of Jones' theory on index to arbitrary factors. J. Funct. Anal.,
66:123 -- 140, 1986.
H. Kosaki. Characterization of crossed product (properly infinite case). Pacific J. Math.,
137:159 -- 167, 1989.
R. Longo. Notes on algebraic invariants for noncommutative dynamical systems. Comm.
Math. Phys., 69:195 -- 207, 1979.
[Lon87]
R. Longo. Simple injective subfactors. Adv. Math., 63:152 -- 171, 1987.
49
[Lon89]
[Lon90]
[Lon94]
[Lon97]
[Lon03]
[Lon08]
R. Longo. Index of subfactors and statistics of quantum fields. I. Comm. Math. Phys.,
126:217 -- 247, 1989.
R. Longo. Index of subfactors and statistics of quantum fields. II. Correspondences, braid
group statistics and Jones polynomial. Comm. Math. Phys., 130:285 -- 309, 1990.
R. Longo. A duality for Hopf algebras and for subfactors. I. Comm. Math. Phys., 159:133 --
150, 1994.
R. Longo. An analogue of the Kac-Wakimoto formula and black hole conditional entropy.
Comm. Math. Phys., 186:451 -- 479, 1997.
R. Longo. Conformal subnets and intermediate subfactors. Comm. Math. Phys., 237:7 -- 30,
2003.
R. Longo. Lecture notes on conformal nets. Part II. Nets of von Neumann algebras. Lecture
notes, available at http://www.mat.uniroma2.it/longo/Home.html, 2008.
[LR95]
R. Longo and K.-H. Rehren. Nets of subfactors. Rev. Math. Phys., 7:567 -- 597, 1995.
[LR97]
R. Longo and J. E. Roberts. A theory of dimension. K-Theory, 11:103 -- 159, 1997.
[LR04]
R. Longo and K.-H. Rehren. Local fields in boundary conformal QFT. Rev. Math. Phys.,
16:909 -- 960, 2004.
[MTW16] V. Morinelli, Y. Tanimoto, and M. Weiner. Conformal covariance and the split property.
preprint arXiv:1609.02196 (to appear in Comm. Math. Phys.), 2016.
[Mug05] M. Muger. Conformal Orbifold Theories and Braided Crossed G-Categories. Comm. Math.
Phys., 260:727 -- 762, 2005.
[Pop95a]
S. Popa. An axiomatization of the lattice of higher relative commutants of a subfactor.
Invent. Math., 120:427 -- 445, 1995.
[Pop95b]
S. Popa. Classification of subfactors and their endomorphisms, CBMS Regional Conference
Series in Mathematics, Vol. 86. Amer. Math. Soc., Providence, RI, 1995.
[PP86]
M. Pimsner and S. Popa. Entropy and index for subfactors. Ann. Sci. Ecole Norm. Sup,
19:57 -- 106, 1986.
[Reh94]
K.-H. Rehren. A new view of the Virasoro algebra. Lett. Math. Phys., 30:125 -- 130, 1994.
[Reh15]
[Rob74]
[Tur10]
K.-H. Rehren. Algebraic conformal quantum field theory in perspective. In: Advances in
Algebraic Quantum Field Theory, R. Brunetti et al., eds., Mathematical Physics Studies,
331 -- 364. Springer International Publishing, 2015.
J. E. Roberts. Some applications of dilatation invariance to structural questions in the
theory of local observables. Comm. Math. Phys., 37:273 -- 286, 1974.
V. Turaev. Homotopy quantum field theory, EMS Tracts in Mathematics, Vol. 10. Eu-
ropean Mathematical Society (EMS), Zurich, 2010. Appendix 5 by Michael Muger and
Appendices 6 and 7 by Alexis Virelizier.
[Wei08] M. Weiner. Restricting positive energy representations of Diff +(S 1) to the stabilizer of n
points. Comm. Math. Phys., 277:555 -- 571, 2008.
[Xu00]
F. Xu. Algebraic orbifold conformal field theories. Proc. Nat. Acad. Sci. U.S.A., 97:14069,
2000.
[Xu05]
F. Xu. Strong additivity and conformal nets. Pacific J. Math., 221:167 -- 199, 2005.
[Xu14]
F. Xu. On intermediate conformal nets. J. Reine Angew. Math., 692:125 -- 151, 2014.
50
|
1802.09964 | 3 | 1802 | 2019-10-10T15:26:19 | On the vanishing cohomology problem for cocycle actions of groups on II$_1$ factors | [
"math.OA"
] | We prove that any free cocycle action of a countable amenable group $\Gamma$ on any II$_1$ factor $N$ can be perturbed by inner automorphisms to a genuine action. This {\em vanishing cohomology} property, that we call $\mathcal V\mathcal C$, is also closed to free products with amalgamation over finite groups. But beyond this no other examples of $\mathcal V\mathcal C$-groups are known. In turn, by considering special cocycle actions $\Gamma \curvearrowright N$ in the case $N$ is the hyperfinite II$_1$ factor $R$, respectively the free group factor $N=L(\Bbb F_\infty)$, we exclude many groups from being $\mathcal V\mathcal C$. We also show that any free action $\Gamma \curvearrowright R$ gives rise to a free cocycle $\Gamma$-action on the II$_1$ factor $R'\cap R^\omega$ whose vanishing cohomology is equivalent to Connes' Approximate Embedding property for the II$_1$ factor $R\rtimes \Gamma$. | math.OA | math |
ON THE VANISHING COHOMOLOGY PROBLEM
FOR COCYCLE ACTIONS OF GROUPS ON II1 FACTORS
SORIN POPA
University of California, Los Angeles
Abstract. We prove that any free cocycle action of a countable amenable group Γ
on any II1 factor N can be perturbed by inner automorphisms to a genuine action.
Besides being satisfied by all amenable groups, this universal vanishing cohomology
property, that we call V C, is also closed to free products with amalgamation over finite
groups. While no other examples of V C-groups are known, by considering special
cocycle actions Γ y N in the case N is the hyperfinite II1 factor R, respectively
the free group factor N = L(F∞), we exclude many groups from being V C. We also
show that any free action Γ y R gives rise to a free cocycle Γ-action on the II1
factor R′ ∩ Rω whose vanishing cohomology is equivalent to Connes' Approximate
Embedding property for the II1 factor R ⋊ Γ.
0. Introduction
A cocycle action of a group Γ on a II1 factor N is a map σ : Γ → Aut(N )
which is multiplicative modulo inner automorphisms of N , σgσh = Ad(vg,h)σgh,
∀g, h ∈ Γ, with the unitary elements vg,h ∈ U(N ) satisfying the cocycle relation
vg,hvgh,k = σg(vh,k)vg,hk, ∀g, h, k ∈ Γ.
If Γ is a free group Fn, for some 1 ≤ n ≤ ∞, then any cocycle Γ-action on any
II1 factor N can obviously be perturbed by inner automorphisms {Ad(wg)}g of N
so that to become a "genuine" action, i.e., such that σ′
g = Ad(wg)σg is a group
morphism, in fact so that the stronger condition vg,h = σg(w∗
g wgh, ∀g, h, holds
true. We obtain in this paper several results towards identifying the class VC of
all countable groups Γ that satisfy this universal vanishing cohomology property.
Thus, we first prove that any free product of amenable groups amalgamated over
a common finite subgroup is in the class VC. Then we show that if a group Γ has
h)w∗
Supported in part by NSF Grant DMS-1700344
1
Typeset by AMS-TEX
2
SORIN POPA
an infinite subgroup which either has relative property (T), or has non-amenable
centralizer, then Γ is not in VC. To prove that all amenable groups lie in VC we use
subfactor techniques to reduce the problem to the case N is the hyperfinite II1 factor
R, where vanishing cohomology holds due to results in ([Oc85]). To exclude groups
from being in VC we apply W∗-rigidity results to two types of cocycle actions that
are "hard to untwist": the ones arising from t-amplifications of Bernoulli actions
on N = R introduced in [P01]; and the ones considered in [CJ84], arising from
normal inclusions F∞ ֒→ Fn with Fn/F∞ = Γ, which give cocycle Γ-actions on
N = L(F∞).
Untwisting cocycle actions on II1 factors is a basic question in non-commutative
ergodic theory and very specific to this area. Besides its intrinsic interest, the prob-
lem occurs in the classification of group actions on II1 factors ([C74], [J80], [Oc85],
[P01a]) and, closely related to it, in the classification of factors through unique
crossed-product decomposition (as in [C74] for amenable factors, or [P01a], [P03],
[P06a], [IPeP05], [PV12] for non-amenable II1 factors). Another aspect, which goes
back to ([CJ84]) and is important in W∗-rigidity, relates non-vanishing cohomology
for certain cocycle Γ-actions on L(F∞) to non-embeddability of L(Γ) into L(Fn).
From an opposite angle, which offers a new point of view much emphasized here,
vanishing cohomology results for cocycle actions are relevant to embedding prob-
lems, such as finding unusual group factors that embed into L(F2) and Connes
Approximate Embedding conjecture.
To describe the results in this paper in more details we need some background
and notations. Let us first note that cocycle actions are more restrictive than outer
actions, which are maps σ : Γ → Aut(N ) that only require σgσhσ−1
gh ∈ Inn(N ),
∀g, h ∈ Γ. It has in fact been shown in ([NT59]) that there is a scalar 3-cocycle
νσ ∈ H 3(Γ) associated to an outer action σ. If σ is free, i.e., σg 6∈ Inn(N ), ∀g 6= e,
then νσ is trivial if and only if σ is a cocycle action. Thus, if we view the vanishing
cohomology problem as a question about lifting a 1 to 1 group morphism σ : Γ →
Out(N ) to a group morphism into Aut(N ), then the problem is not well posed
unless one requires νσ ≡ 1, i.e., that σ defines a cocycle action.
Like for genuine actions, one can associate to a cocycle action Γ yσ N a tracial
crossed product von Neumann algebra N ⋊ Γ, with the freeness of σ equivalent to
the condition N ′ ∩ N ⋊ Γ = C1. Thus, if σ is free then N ⊂ M = N ⋊ Γ is an
irreducible inclusion of II1 factors with the normalizer of N in M generating M as
a von Neumann algebra (N is regular in M ). Conversely, any irreducible regular
inclusion of II1 factors N ⊂ M arises this way, from a crossed product construction
involving a free cocycle action (cf. [J80]).
The crossed product framework allows an alternative formulation of vanishing
cohomology. Thus, if M = N ⋊ Γ denotes the crossed product II1 factor associated
VANISHING COHOMOLOGY
3
with the free cocycle action (σ, v) of Γ on N , and we let {Ug}g ⊂ M denote
the canonical unitaries implementing σ on N , then the existence of wg ∈ U(N )
such that vg,h = wgσg(wh)w∗
gh, ∀g, h (i.e., vanishing cohomology for v) amounts
to U ′
g),
g ∈ Γ, is a genuine action (i.e., weak vanishing cohomology for v) amounts to the
weaker condition that {U ′
g = wgUg being a Γ-representation. While the condition that σ′
g}g is a projective Γ-representation.
g = Ad(U ′
Given a II1 factor N , we denote by VC(N ) (respectively VCw(N )) the class of
countable groups Γ with the property that any free cocycle action of Γ on N satisfies
the strong form (respectively weak form) of the vanishing cohomology. Also, we
denote by VC (respectively VCw) the class of countable groups Γ with the property
that any free cocycle action of Γ on any II1 factor N satisfies the strong form
(respectively weak form) of the vanishing cohomology.
The class VC contains all finite groups by ([J80], [Su80]) and all groups with
polynomial growth by ([P89]). The first main result in this paper, which we prove
in Section 2, shows that in fact VC contains all countable amenable groups. Since
by [J80] all 1-cocycles for actions of finite groups are co-boundaries, this allows
to deduce that, more than just containing the free groups, all amalgamated free
products of amenable groups over finite groups belong to VC.
0.1. Theorem. The class VC contains all countable amenable groups. Also, if
{Γn}n is a sequence of groups in VC and K ⊂ Γn is a common finite subgroup,
n ≥ 1, then Γ1 ∗K Γ2 ∗K ... ∈ VC.
To prove the first part of this result we show that any cocycle action σ of a
countable amenable group Γ on a separable II1 factor N can be perturbed by inner
automorphisms to a cocycle action σ′ that leaves invariant an irreducible hyperfinite
−1 are implemented
subfactor R ⊂ N with the additional property that σ′
by unitaries in R, ∀g, h, with σ′ still free when restricted to R (see Theorem 2.1).
This reduces the vanishing cohomology problem to the case N = R, where one can
apply the vanishing cohomology result in ([Oc85]) to finish the proof.
hσ′
gh
gσ′
To prove the existence of a "large" R ⊂ N that's normalized by inner perturba-
tions of σ we use an idea introduced in ([P89]; cf. also 5.1.5 in [P91]), of translating
the problem into the question of whether there exists a sub-inclusion of hyperfinite
factors inside the "diagonal subfactor" N ≃ N σ ⊂ M σ associated with σ, so that
to have a non-degenerate commuting square satisfying a strong smoothness condi-
tion on higher relative commutants. This subfactor problem was solved in [P89]
in the case Γ is finitely generated with trivial Poisson boundary (e.g., with poly-
nomial growth; see [KV83]), by constructing R as a limit of relative commutants
P ′
n ∩ N of factors in a tunnel N ⊃ N1..., obtained by iterating the downward basic
construction (in the spirit of [P91], [P93]).
4
SORIN POPA
However, that construction depends crucially on the trivial Poisson boundary
condition on Γ. We use here the amenability of Γ alone to construct a more elab-
orate decreasing sequence of subfactors Pn ⊂ N with P ′
n ∩ N ր R "large" in N ,
obtained through reduction/induction in Jones tunnels. In fact, this method allows
us to obtain the existence of strongly smooth embedding of hyperfinite subfactors
into any finite index subfactor N ⊂ M with standard invariant GN⊂M amenable (in
the sense of [P91], i.e., with its graph ΓN⊂M satisfying the Kesten-type condition
kΓN⊂M k2 = [M : N ]; see also [P93], [P94a], [P97] for other equivalent definitions).
We in fact show that given any amenable C∗-category G of endomorphisms of a
II∞ factor N (viewed here as an outer action of an abstract rigid C∗-tensor cate-
gory), there exists a "large" approximately finite dimensional (AFD) II∞ subfactor
R⊗B(ℓ2N) in N that's normalized by G, modulo inner automorphisms (see Theorem
2.12).
In Section 5 we use the strong solidity of free group factors ([OP07]) to prove
that in order for a group Γ to satisfy the property that any of its actions on II1
factors normalizes a hyperfinite subfactor (modulo inner automorphisms), Γ must
necessarily be amenable. The problem of whether this dichotomy still holds for
subfactor standard invariants and rigid C∗-tensor categories, remains open.
In turn, in Sections 3 and 4 we obtain a series of obstruction criteria for groups
to belong to the classes VC(R), VC(L(F∞)), VC, summarized in the following:
0.2. Theorem. 1◦ If a countable group Γ has an infinite subgroup which either has
the relative property (T), or has non-amenable centralizer in Γ, then Γ 6∈ VCw(R).
2◦ Assume a countable group Γ satisfies one of the following: (a) Γ does not have
Haagerup property (e.g., it contains an infinite subset with relative property (T));
(b) The Cowling-Haagerup invariant Λ(Γ) is larger than 1; (c) Γ has an infinite
subgroup with non-amenable centralizer; (d) Γ has an infinite amenable subgroup
with non-amenable normalizer. Then Γ 6∈ VCw(L(F∞)), in particular Γ 6∈ VCw.
To prove the restrictions on VC(R) we use the t-amplifications of Bernoulli ac-
tions on R introduced in [P01a] and results obtained there and in ([P06a]) through
In turn, to get restrictions on VCw(L(F∞)), we
deformation-rigidity arguments.
use the Connes-Jones (CJ) cocycles associated with surjective group morphisms
π : FS → (Γ, S), extending the map assigning the free generators of FS to a set of
generators S ⊂ Γ. As shown in [CJ84], if Γ is infinite, non-free, then kerπ ≃ F∞
and the inclusion L(F∞) = N ⊂ M = L(FS) is of the form N ⊂ N ⋊(σπ ,vπ) Γ, for
a free cocycle action (σπ, vπ). The vanishing of the cocycle vπ implies that L(Γ)
embeds into L(FS), hence 2◦ above follows from results in ([CJ84], [P01b], [O03],
[P06b], [OP07]).
The CJ-cocycles seem the "most difficult to untwist", in the sense that if all such
VANISHING COHOMOLOGY
5
cocycle actions of Γ on L(F∞) untwist, then Γ ought to be in VC. In particular, this
would show that VC = VC(L(F∞)). Since untwisting a CJ-cocycle for Γ implies
that Γ is in the class W∗
leq(F2) of groups whose von Neumann algebra embeds into
L(F2), one has VC ⊂ VC(L(F∞)) ⊂ W∗
leq(F2). Very little is actually known about
the class W∗
leq(F2), which is extremely interesting by itself. Any success in proving
VC property for some "exotic" group Γ, would provide embeddings L(Γ) ֒→ L(F2).
In the final part of the paper we discuss a connection between vanishing coho-
mology phenomena and Connes Approximate Embedding conjecture, on whether
any separable II1 factor M embeds in the ultrapower Rω of the hyperfinite II1
factor R. Thus, we notice that any free action of a group Γ on R (such as the
"non-commutative" Bernoulli action Γ y R⊗Γ ≃ R), gives rise to a free cocycle
action of Γ on the centralizer Rω = R′ ∩ Rω of R in Rω. We deduce that this
cocycle untwists if and only M = R ⋊σ Γ satisfies the conjecture.
Acknowledgement.
I am very grateful to Damien Gaboriau, Vaughan Jones,
Jesse Peterson and Stefaan Vaes for many useful discussions related to this paper.
I am also grateful to the referee for his/her many pertinent questions that led to
what I believe to be a much improved final version.
1. Preliminaries and notations.
For general background on II1 factors we refer the reader to ([AP17]; also [T79],
[BrO08] for general theory of operator algebras and von Neumann algebras).
1.1. Cocycle actions and crossed products. Given a II1 factor N , we denote by
Aut(N ) the group of automorphisms of N . An automorphism θ of N is inner if there
exists u in the unitary group of N , U(N ), such that θ(x) = Adu(x) = uxu∗, ∀x ∈ N .
We denote by Inn(N ) ⊂ Aut(N ) the group of all such inner automorphisms and by
Out(N ) the quotient group Aut(N )/Inn(N ).
Given a discrete group Γ, an action of Γ on N is a group morphism σ : Γ →
Aut(N ). We will use the notation Γ yσ N to emphasize such an action.
More generally, a cocycle action σ of Γ on N is a map σ : Γ → Aut(N ) with the
property that there exists v : Γ × Γ → U(N ) such that:
(1.1.1)
(1.1.2)
σe = id and σgσh = Ad vg,hσgh, ∀g, h ∈ G
vg,hvgh,k = σg(vh,k)vg,hk, ∀g, h, k ∈ Γ.
The cocycle action σ is free if σg cannot be implemented by unitary elements in
N, ∀g 6= e, in other words if the factoring of σ through the quotient map Aut(N ) →
Out(N ) is 1 to 1. All cocycle actions (in particular all actions) that we will consider
in this paper are assumed to be free.
6
SORIN POPA
Following ([KT02]), a map σ : Γ → Aut(N ) that's a 1 to 1 group morphism when
factored through the quotient map Aut(N ) → Out(N ), is called an outer Γ-action
(an alternative terminology for such σ is Q-kernel, notably used in [J80], [Oc85]).
Thus, an outer action satisfies (1.1.1) above, but not necessarily (1.1.2). As shown
in ([NT59]), if σ is an outer Γ-action, then one can associate to it a scalar 3-cocycle
νσ ∈ H 3(Γ), with the property that νσ ≡ 1 if and only if σ is a cocycle action, and
which one calls the H 3(Γ)-obstruction of σ.
If σ is a cocycle action, then a map v satisfying (1.1.2) is called a 2-cocycle
for σ. The 2-cocycle is normalized if vg,e = ve,g = 1, ∀g ∈ G. Note that any
2-cocycle satisfies ve,e ∈ C. Thus any 2-cocycle v can be normalized by replacing
it, if necessary, by v′
e,e vg,h, g, h ∈ Γ. All 2-cocycles considered from now on
will be normalized.
g,h = v∗
Also, when given a cocycle action σ, we will sometimes specify from the beginning
the 2-cocycle it comes with, thus considering it as a pair (σ, v).
Note that the 2-cocycle v is unique modulo perturbation by a scalar 2-cocycle µ.
More precisely, v′ : Γ × Γ → U(N ), with v′
e,e = 1, satisfies conditions (1.1.1), (1.1.2)
if and only if v′ = µv for some scalar valued function µ : Γ × Γ → T satisfying
µe,e = 1 and
(1.1.3)
µg,hµgh,k = µh,kµg,hk,
∀g, h, k ∈ Γ
Let us recall the definition of the crossed product von Neumann algebra associated
with a cocycle action (σ, v) of Γ on N , denoted N ⋊(σ,v) Γ (or simply N ⋊σ Γ if σ
is a genuine action). So let H denote the Hilbert space ⊕h(L2N )h ≃ ℓ2(Γ, L2N ),
which we view as the space of ℓ2-summable formal series Ph Uhξh, where {Ug}g∈Γ
are here "indeterminates" (labels) and the "coefficients" ξh belong to L2N .
We define a ∗-algebra structure on the subspace H0 ⊂ H of finitely supported
sums with "bounded" coefficients ξg = xg ∈ N , and at the same time a Hilbert
H0-bimodule structure on H, as follows. First, we let Ue act on both left and
right on H as the identity idH and identify N with UeN acting left-right on H by
x(Ph Uhξh)y = Ph Uh(σ−1
h (x)ξhy). Then we let Ug · Ph Uhξh = Ph vg,hUghξh,
which by the change of variables h′ = gh and "moving" vg,h from left to right ac-
cording to the above multiplication by N rule, is equal to Ph Uhσ−1
h (vg−1h,h)ξg−1h.
We also let (Ph Uhξh) · Ug = Ph vh,gUhgσ−1
g (ξh) which by similar considerations
is equal to Ph Uhσ−1
g,g−1 and
(Ugx)∗ = x∗U ∗
g,g−1σg(x∗). The ∗-algebra H0 has a trace given by
τ (Ph Uhxh) = τN (xe) which recovers for elements in H0 the H-scalar product,
i.e., if X, Y ∈ H0 then hX, Y iH = τ (Y ∗X).
g (ξhg−1). Finally, we let U ∗
g = Ug−1 v∗
h (vhg−1,g)σ−1
g = Ug−1v∗
It is easy to verify that the left multiplication by Ug give unitary operators on
H, the left multiplication by N = N Ue ⊂ H0 on H gives a representation of N as
VANISHING COHOMOLOGY
7
a von Neumann algebra, and altogether left multiplication by elements in H0 are
bounded operators on H that give a ∗-representation M0 of H0 in B(H), with the
trace τ being implemented by the vector state given by 1 = Ue1 ∈ H.
The crossed product von Neumann algebra N ⋊(σ,v) Γ is by definition the weak
operator closure of M0 in B(H). It is a finite von Neumann algebra with faithful
normal state τ (X) = hX1, 1iH, ∀X ∈ M . One clearly has a natural identification
between the standard representation (or standard Hilbert M -bimodule) L2(M, τ )
and H.
The cocycle action (σ, v) is free if σg is an outer automorphism, ∀g 6= e. It is well
known (and immediate to check) that (σ, v) is free if and only if N ′∩N ⋊(σ,v)Γ = C1.
So in this case M = N ⋊σ Γ is a II1 factor with the normalizer NM (N ) = {u ∈
U(M ) uN u∗ = N } of N in M generating M (i.e., with N regular in M ).
Conversely, if N ⊂ M is a regular, irreducible inclusion of II1 factors and we
denote Γ = NM (N )/U(N ), with Ug ∈ NM (N ), g ∈ Γ, a lifting of Γ, Ue = 1, and
we denote σg = Ad(Ug)N , vg,h = UghU ∗
g , then (σ, v) is a free cocycle action of
Γ on N , with N ⊂ N ⋊(σ,v) Γ naturally isomorphic to N ⊂ M (see e.g., [J80]).
1.2. Cocycle conjugacy of cocycle actions. The cocycle actions (σi, vi) of Γ
on Ni, i = 1, 2, are cocycle conjugate if there exists an isomorphism θ : N1 ≃ N2
and a map w : Γ → U(N2) such that the following conditions are satisfied:
h U ∗
(1.2.1)
(1.2.2)
θσ1(g) θ−1 = Ad(wg)σ2(g),
∀g.
θ(v1(g, h)) = wgσ2(g)(wh)v2(g, h) w∗
gh, ∀g, h.
The cocycle actions σ1, σ2 are outer conjugate (or weakly cocycle conjugate) if
condition (1.2.1) is satisfied. Note that this is equivalent to σ1, σ2 composed with
the quotient map Aut(N ) → Out(N ) being conjugate by an element in Out(N ).
Similarly, two outer actions σ1, σ2 → Aut(N ) are outer conjugate, if there exists
θ ∈ Aut(N ) such that σ1(g) = θσ2(g)θ−1 modulo Inn(N ), for all g ∈ Γ.
The actions σ1, σ2 are conjugate if there exists an isomorphism θ : N1 ≃ N2 such
that conditions (1.2.1) is satisfied with w = 1. We then write σ1 ∼ σ2.
We recall here the following well known observation (see e.g., [J80]), which trans-
lates cocycle conjugacy of free cocycle actions into the isomorphism of the associated
crossed-product inclusions of factors.
Proposition. Let (σi, vi) be a cocycle action of the discrete group Γi on the II1
factor Ni, i = 1, 2. If there exists a ∗-isomorphism Φ : N1⋊(σ1,v1)Γ1 ≃ N2⋊(σ2,v2)Γ2
such that Φ(N1) = N2, then σ1 and σ2 are cocycle conjugate. More precisely, there
8
SORIN POPA
exists a group isomorphism γ : Γ1 → Γ2, and unitaries wg ∈ U(N2), for all g ∈ Γ1,
such that:
(i) Φσ1(g)Φ−1 = Adwg σ2(γ(g)), for all g ∈ Γ1,
(ii) Φ(v1(g, h)) = wgσ2(γ(g)(wh)v2(γ(g), γ(h))w∗
Conversely, if Φ : N1 ≃ N2 is a ∗-isomorphism, γ : Γ1 ≃ Γ2 is a group isomor-
phism, and there exist unitaries wg ∈ U(N2) for all g ∈ G1 such that (i), (ii) are
satisfied, then Φ can be extended to an isomorphism N1 ⋊(σ1,v1) Γ1 ≃ N2 ⋊(σ2,v2) Γ2
(hence, to an isomorphism of the associated inclusions).
gh, for all g, h ∈ Γ1.
1.3. 1-cocycles for actions. Assume σ is a genuine action of Γ on the II1 factor
N . A map w : Γ → U(N ) satisfying condition
(1.3.1)
wgσg(wh) = wgh, ∀g, h
is called a 1-cocycle for σ. Such a 1-cocycle for σ is a coboundary (or it is trivial)
if there exists a unitary element v ∈ U(N ) such that wg = v∗σg(v), ∀g. (Clearly,
such maps wg do satisfy the 1-cocycle condition (1.3.1)).
The map w is called a weak 1-cocycle if it satisfies the relation (1.2.1) modulo
the scalars, i.e.,
(1.3.1')
wgσg(wh)w∗
gh ∈ T1, ∀g, h ∈ Γ
Note that this is equivalent to Ad(wg)σg being an action. Note also that if w
is a weak 1-cocycle then µg,h = wgσg(wh)w∗
gh is a scalar 2-cocycle for Γ, i.e.,
µ ∈ H 2(Γ). Also, cocycle conjugacy of two (genuine) actions σi : Γ → Aut(Ni),
i = 1, 2, amounts to conjugacy of σ1 and σ′
2(g) = Ad(wg)σ2(g), g ∈ Γ,
for some 1-cocycle w for σ2.
2, where σ′
A (weak) 1-cocycle w is weakly trivial (or weak cobouboundary) if there exists a
unitary element v ∈ U(N ) such that vwgσg(v)∗ ∈ T1, ∀g.
Two (weak) 1-cocycles w, w′ of the action σ are equivalent if there exists a unitary
element v ∈ N such that w′
g = vwgσg(v)∗, ∀g ∈ Γ (resp. modulo scalars). Thus, a
weak 1-cocycle is weakly trivial iff it is equivalent to a scalar valued weak 1-cocycle
(N.B.: these are just plain scalar functions on Γ). Note that the scalar valued
genuine 1-cocycles are just characters of Γ.
Two free actions σ1, σ2 of Γ on N are cocycle conjugate iff σ1 is conjugate to σ′
2,
where σ′
2(g) = Ad(wg)σ2(g), ∀g ∈ Γ, for some 1-cocycle w for σ2.
We also mention here a well known result from [J80], showing that any 1-cocycle
of an action of a finite group Γ is co-boundary. This property is actually specific
to finite groups: we use a result in [P01a] to deduce that if Γ is infinite, then there
exist free ergodic actions Γ y R which admit non-trivial 1-cocycles. (N.B. In the
particular case when Γ is amenable, this fact can be derived from [Oc] as well).
VANISHING COHOMOLOGY
9
Proposition. 1◦ Let Γ yσ N be a free action of a finite group Γ on a II1 factor
N . If w is a 1-cocyle for σ, then there exists u ∈ U(N ) such that wg = uσg(u∗),
∀g ∈ Γ.
1+λ , λ
2◦ Let (N0, ϕ0) be a copy of the 2 by 2 matrix algebra with the state given by
weights { 1
1+λ }, for some 0 < λ < 1, and Γ be an infinite group. Let Γ y
(N , ϕ) = ⊗g(N0, ϕ0)g be the Bernoulli Γ-action with base (N0, ϕ0). Let Γ yσ N =
Nϕ ≃ R be the corresponding Connes-Størmer Bernoulli action. Let B ⊂ N be an
atomic von Neumann subalgebra of the form ⊕nBn, with Bn ≃ Mkn×kn (C) having
minimal projections of trace λmn , with m1 < m2 < ..... Then there exists a 1-
cocycle w for σ such that σ′
g = Ad(wg)σg, g ∈ Γ, has B as its fixed point algebra.
If B 6= C, then any such 1-cocycle is not a co-boundary.
Proof. 1◦ This is (Corollary 2.16 in [CT76; see also [J80]). We include here the
proof, for completeness, which is based on Connes well known "2 by 2 matrix trick".
Thus, let σ be the action of Γ on N = M2×2(N ) = N ⊗ M2×2(C) given by
If {eij 1 ≤ i, j ≤ 2} is a matrix unit for M2×2(C) ⊂ N , then
σg = σg ⊗ id.
If Q ⊂ N denotes the fixed point algebra
wg = e11 + wge22 is a cocycle for σ.
of the action σ′
g = Ad( wg)σ, then e11, e22 ∈ Q and the existence of a unitary
element u ∈ N satisfying wg = uσg(u∗), ∀g, is equivalent to the fact that e11, e22
are equivalent projections in Q. But the fixed point algebra of any free action of a
finite group on a II1 factor is a II1 factor. Thus, e11, e22 are equivalent in Q and w
follows co-boundary.
n }1≤j≤kn ∈ N be isometries such that V j
= N and {V i
∗ ∈ N ,
∗ 1 ≤ i, j ≤ kn} be the matrix units
τ (V j
of Bn. Let also πn be the trivial representation of Γ of multiplicity kn. Then by
n)∗,
(Theorem 3.2 in [P01a]), wg = Pn Pi,j πn(g)i,jV i
g ∈ Γ, defines a 1-cocycle for σ and σ′
g = Ad(wg)σg has B as its fixed point
algebra.
n )∗ = Pn Pi V i
nσg(V j
nσg(V i
2◦ For each n ≥ 1, let {V j
n V j
n
) = λmn , V j
n N V j
n
∗
∗
nV j
n
n V j
n
Since the fixed point algebra of an action is a conjugacy invariant of the action
and σ is mixing (thus ergodic), it follows that σ, σ′ are not conjugate, in particular
there exists no u ∈ U(N ) such that σ′
g = Ad(u)σgAd(u∗), ∀g, a relation that
amounts to wg = uσg(u∗) modulo scalars, ∀g.
(cid:3)
1.4. Vanishing cohomology and property VC. The 2-cocycle v for the cocycle
action σ vanishes (or it is a coboundary) if there exists a map w : Γ → U(N ) such
that we = 1 and v = ∂w, i.e.:
(1.4.1)
vg,h = (∂w)g,h
def= σg(w∗
h)w∗
gwgh, ∀g, h ∈ Γ.
The 2-cocycle v weakly-vanishes (or it is a weak coboundary) if there exists w :
10
SORIN POPA
Γ → U(N ) such that we = 1 and v = ∂w modulo scalars, i.e.:
(1.4.2)
wgσg(wh)vg,hw∗
gh ∈ C1, ∀g, h ∈ Γ.
Note that this is equivalent to
(1.4.2′)
(Ad wgσg) (Ad whσh) = Ad wghσgh, ∀g, h
i.e., to σ′
g
def= Ad wgσg being a "genuine" action.
{wg}g ⊂ N such that U ′
U ′
gh, ∀g, h ∈ Γ).
In turn, the vanishing of v is equivalent to the existence of unitary elements
g = wgUg ∈ M = N ⋊(σ,v) Γ give a representation of Γ (i.e.,
h = U ′
gU ′
Given a II1 factor N , we denote by VC(N ) the class of countable groups Γ for
which any free cocycle action (σ, v) of Γ on N has the property that the 2-cocycle
v vanishes (or is coboundary) and by VCw(N ) the class of groups Γ for which any
free cocycle action (σ, v) of Γ on N has the property that v is a weak-coboundary.
We denote by VC (respectively VCw) the class of countable groups Γ with the
property that Γ ∈ VC(N ) (resp. Γ ∈ VCw(N )) for any II1 factor N . If Γ ∈ VC then
we also say that Γ has property VC or that it is a VC-group.
We are especially interested in identifying the VC and VCw groups, i.e., groups
that have the most "universal" vanishing cohomology property. Other classes of
interest will be VC(N ) for N equal to the hyperfinite II1 factor R and for N equal to
the free group factor L(F∞). This is because R and L(F∞) are the most interesting
"non-commutative probability spaces". Also, any countable group Γ has "many"
free actions on these factors, in fact both of them have a lot of generalized sym-
metries (notably L(F∞), on which by [PS01] any "group-like" object admits free
actions). Also, both factors admit many cocycle actions that are "hard to untwist"
(cf. [CJ84], [P01a] and Section 3 and 4 below). (N.B. It should be noticed that by
the way we have defined VC(N ), if a factor N has only inner automorphisms, i.e.,
Out(N ) = {1}, like the examples in [IPeP05], then any Γ belongs to VC(N )!)
We'll now show that the class VC is closed to amalgamated free products over
finite subgroups and that vanishing cohomology for cocycle actions of countable
groups is essentially a "separability" property:
1.5. Proposition. 1◦ if {Γn}n≥0 ⊂ VC(N ) (respectively VCw(N )) for some II1
factor N and K ⊂ Γn is a common finite subgroup, n ≥ 0, then Γ0 ∗K Γ1 ∗K
Γ2 ∗K ... ∈ VC(N ) (respectively VCw(N )). Also, if {Γn}n ⊂ VC (resp. VCw), then
Γ0 ∗K Γ1 ∗K .... ∈ VC (resp. VCw).
2◦ Let N be a II1 factor, Γ ⊂ Out(N ) a countable group with a lifting {σg}g∈Γ ⊂
Aut(N ) and denote vg,h ∈ U(N ) a set of unitaries satisfying Ad(vg,h) = σgσhσ−1
gh ,
VANISHING COHOMOLOGY
11
g, h ∈ Γ. There exists a separable II1 subfactor Q ⊂ N that contains the countable
set {vg,h g, h ∈ Γ} and is normalized by σ, with σg outer, ∀g ∈ Γ.
3◦ VC (respectively VCw) coincides with the class of countable groups Γ with the
property that Γ ∈ VC(N ) (resp. Γ ∈ VCw(N )) for any separable II1 factor N .
Proof. 1◦ Assume Γn ∈ VC(N ). Let (σ, v) be a free cocycle action of G = ∗KΓn on
N and denote M = N ⋊σ G with Ug, g ∈ G the corresponding canonical unitaries.
Since Γn ∈ VC(N ), there exist unitaries {wn
g =
wn
gU 0
g ,
g ∈ Γ0, we may assume w0
g g ∈ Γn} in N such that U n
g Ug, g ∈ Γn, give left regular representations of Γn. Replacing Ug by w0
g = 1, ∀g ∈ Γ0.
k = wn
But then for each n ≥ 1, U n
k Uk, k ∈ K, for some 1-cocyles wn : K → U(N )
for the restriction to K of the Γn-action σn implemented by U n
g , g ∈ Γn. By
([J80]; see Proposition 1.3 above) any 1-cocycle of a free action of a finite group
vanishes. Hence, there exists vn ∈ U(N ) (with v0 = 1) such that wn
n),
n, k ∈ K. But then the unitaries {v∗
equivalently U n
g σg(vn) g ∈
Γn, n ≥ 0} generate inside M a copy of the left regular representation of G = ∗K Γn
implementing a G-action on N that gives an inner perturbation of the initial cocycle
G-action σ.
k = vnσn(k)(v∗
nU n
k = vnUkv∗
2◦ We construct recursively an increasing sequence of separable von Neumann
subalgebras Qn, n ≥ 0, such that Q0 ⊃ {vg,h g, h ∈ Γ} and for each m ≥ 1 we
have
m∩N (x) = τ (x)1, ∀x ∈ (Qm−1)1;
m∩N ⋊Γ(Ug) = 0, ∀g 6= e;
(a) EQ′
(b) EQ′
(c) Qm ⊃ ∪gσg(Qm−1),
where Ug ∈ N ⋊ Γ are the canonical unitaries implementing σ.
Assume we have constructed these algebras up to m = n. Since N ′ ∩ N ⋊ Γ = C,
by using (Theorem 0.1 in [P13]) we can get a Haar unitary v = (vk)k ∈ N ω that's
free independent to Qn−1 ∪ {Ug}g. Thus, if we take Q0
n to be the von Neumann
algebra generated by Qn−1 and {vk}k, then we already have (a) and (b) satisfied for
Qn = Q0
n by the von Neumann algebra
generated by Qn = ∪gσg(Q0
n, and then we can replace this "initial" Q0
n), to have (c) satisfied as well.
w
Finally, if we define Q = ∪nQn
, then Q is clearly separable, condition (c)
insures that Q is normalized by σ, condition (a) shows that EQ′∩N (∪nQn) ∈ C1,
implying that Q is a factor, while condition (b) shows that σg is outer on Q, ∀g 6= e.
3◦ This part is now trivial by 2◦.
1.6. Remarks 1◦ As we will see in Sections 3 and 4, it is in general not true
that if Γi are in VC then their amalgamated free product over a common (infinite)
(cid:3)
12
SORIN POPA
amenable subgroup H ⊂ Γi, Γ = Γ1 ∗H Γ2 is in VC. For instance, Z2 ⋊ SL(2, Z)
does not even belong to VCw(R) (see Theorem 3.2).
2◦ The classes VC may satisfy other general permanence properties. For instance,
it may be true that Γ ∈ VC implies Γ0 ∈ VC for any subgroup Γ0 ⊂ Γ (or at least
when [Γ : Γ0] < ∞). However, the obvious idea for a proof, which is to "co-induce"
a given cocycle action Γ0 yσ0 N0 to a set of automorphisms {σg g ∈ Γ} on
N = N ⊗Γ/Γ0
doesn't work when [Γ : Γ0] = ∞, because an infinite tensor product
of inner automorphisms may become outer, so the σg's may in fact not give a
cocycle action of Γ. When the index of Γ0 in Γ is finite, then σ defined this way
does give a cocycle action of Γ on N , but it is not immediate of how to "bring down
to N0" the vanishing of the cohomology for σ to the vanishing of the cohomology
for the initial Γ0 yσ0 N0.
0
2. Groups with the property VC
In this section we prove a vanishing cohomology result for arbitrary free cocycle
actions of countable amenable groups on arbitrary II1 factors.
To do this, we'll first show that any amenable subgroup Γ ⊂ Aut(N )/Inn(N ) can
be lifted to a set {σg g ∈ Γ} ⊂ Aut(N ) normalizing a "large" hyperfinite subfactor
of N (see Theorem 2.1). As it happens, this property, which is interesting by itself,
characterizes the amenability of the group Γ. Indeed, we will show in Section 5
that any non-amenable group admits a free action on N = L(F∞) that cannot be
perturbed to a cocycle action that normalizes a hyperfinite subfactor of N .
Once we prove that any cocycle action σ of an amenable group Γ on N normalizes
(modulo inner perturbation) a hyperfinite II1 factor R ⊂ N , we reduce the vanishing
cohomology problem to the case N = R, where by a well known result of Ocneanu
[Oc85] free cocycle actions of amenable groups can indeed be "untwisted" to genuine
actions. The fact that R is "large in N " assures that by untwisting σ on R we have
untwisted it as an action on N as well.
To show that σ normalizes up to Inn(N ) a "large hyperfinite subfactor of N ", we
reduce the problem to a statement about commuting squares of subfactors, as fol-
lows. Assume Γ is generated by a finite set e ∈ F = F −1 ⊂ Γ and consider the locally
trivial subfactor obtained by the diagonal embedding N σ,F := {Pg∈F σg(x)egg
x ∈ N } ⊂ MF ×F (N ) =: M σ,F , where {egh}g,h∈F ⊂ MF ×F (C) are the matrix
If Q ⊂ R is an inclusion of factors with Q ⊂ N σ,F ,
units (see 5.1.5 in [P91]).
R ⊂ M σ,F , egg ∈ R, ∀g ∈ F , and (Q ⊂ R) ⊂ (N σ,F ⊂ M σ,F ) makes a non-
degenerate commuting square, then Q ⊂ R is itself locally trivial and there exist
unitary elements wg ∈ N such that wgeeg ∈ R. If one denotes Q0 ⊂ N the image
of Q under the isomorphism N σ,F ≃ N , then this amounts to Q0 being invariant
VANISHING COHOMOLOGY
13
g = Adwg ◦ σg, ∀g ∈ F . Moreover, if Q′ ∩ R = N σ,F ′
to σ′
iff σg is outer, ∀g ∈ F . This observation applied to N σ,F ⊂ M σ,F
are
the factors in the tower for N σ,F ⊂ M σ,F ), in combination with (5.1.5 in [P91]),
shows that if all the higher relative commutants in the Jones towers for Q ⊂ R and
N σ,F ⊂ M σ,F coincide, then σ′ implements an outer action of Γ on Q0.
∩ R, then σ′
(where M σ,F
n
n
is outer
gQ0
So all we need to do is to produce an inclusion of hyperfinite factors Q ⊂ R
inside N σ,F ⊂ M σ,F , making a commuting square and having same higher relative
commutants.
We will solve this commuting square problem by only using that N σ,F ⊂ M σ,F
has amenable graph. Thus, we will in fact prove that any finite index inclusion of
separable II1 factors N ⊂ M with amenable standard invariant GN⊂M contains an
inclusion of hyperfinite factors (Q ⊂ R) ⊂ (N ⊂ M ), that makes a non-degenerate
commuting square and has identical higher relative commutants in the associated
Jones tower (in particular same standard invariant), in fact even satisfies the strong
smoothness condition Q′ ∩ Rn = Q′ ∩ Mn = N ′ ∩ Mn, ∀n (see Theorem 2.10 below).
We'll obtain Q ⊂ R as an inductive limit of relative commutants P ′
n ∩ M
of a decreasing sequence of finite index subfactors Pn ⊂ N that come from repeated
downward basic constructions of subfactors M p′ ⊂ p′Mnp′ obtained by appropriate
induction/reduction in the Jones tower N ⊂ M ⊂ M1 ⊂ ..., with choices "dictated"
by the local characterization of the amenability of ΓN⊂M in ([P97], Theorem 6.1).
n ∩ N ⊂ P ′
2.1. Theorem. Let N be a II1 factor and σ : Γ → Aut(N ) an outer action of
a countable amenable group Γ on N , with H 3(Γ)-obstruction νσ and with vg,h ∈
U(N ) satisfying σgσh = Ad(vg,h)σgh, ∀g, h ∈ Γ. Then there exist {wg}g ⊂ U(N )
and a hyperfinite subfactor R ⊂ N such that if we denote σ′
g = Ad(wg)σg and
g,h = wgσg(wh)vg,hw∗
v′
gh, g, h ∈ Γ, then we have:
g(R) = R and v′
g,h ∈ R, ∀g, h ∈ Γ.
(2.1.1) σ′
(2.1.2) {σ′
gR
}g is an outer action of Γ on R with same H 3(Γ)-obstruction as σ.
If in addition (σ, v) is a free cocycle action of Γ on N , then {wg}g can be chosen
R, v′) is a free cocycle action of Γ on R. Moreover, if N is separable,
so that (σ′
then one can choose σ′, v′, R so that to also have R′ ∩ N = C.
Let us show right away how Theorem 2.1 combined with Ocneanu's Theorem in
[Oc85] can be used to derive the vanishing cohomology result for cocycle actions of
arbitrary amenable groups:
14
SORIN POPA
2.2. Theorem. Let N be a II1 factor, Γ a countable amenable group and (σ, v) a
free cocycle action of Γ on N . Then there exist unitary elements {wg ∈ U(N ) g ∈
Γ} such that σ′
g = Ad(wg) ◦ σg, g ∈ Γ, is a genuine action of Γ on N , in fact such
that moreover we have vg,h = σg(w∗
gwgh, ∀g, h ∈ Γ.
h)w∗
Proof of 2.2. By Theorem 2.1, there exist unitary elements w0
hyperfinite II1 subfactor R ⊂ N such that:
g ∈ N , g ∈ Γ, and a
(2.2.1) σ0
(2.2.2) v0
(2.2.3) σ0
g := Ad(w0
g,h := wgσg(wh)vg,hw∗
gR is outer, ∀g 6= e.
g)σg leaves R invariant, ∀g ∈ Γ;
gh belongs to R, ∀g, h ∈ Γ.
But then, (σ0
R, v0) implements a free cocycle action of the countable amenable
group Γ on R, so by Ocneanu's Theorem [Oc85] the 2-cocycle v0 is co-boundary
on R, i.e., there exist unitary elements w1
gh,
∀g, h ∈ Γ. This shows that wg = w1
(cid:3)
g,h = σ0
g satisfy the required condition.
g ∈ R such that v0
∗)w1
g
g(w1
h
∗w1
gw0
2.3. Corollary. The class VC contains all countable amenable groups and is closed
to free products with amalgamation over finite subgroups, i.e., if {Γn}n ⊂ VC and
K ⊂ Γn is a common finite subgroup, then Γ0 ∗K Γ1 ∗K .... ∈ VC.
For the rest of this section, we will use concepts and notations from [J83] (such
as the basic construction, the Jones tower of factors, etc), as well as from ([PiP84],
[P91], [P93], [94a], [94b], [P97]). In particular, we will often use as framework the
symmetric enveloping (abbreviated SE) inclusion M ⊗M
of N ⊂ M ,
M
op
op
⊂ M ⊠
eN
introduced in [P94b] (cf. also the extended version of the paper, [P97]).
We begin by recalling some properties relating Jones tower/tunnel of a subfactor
with its symmetric enveloping inclusion. It will be useful for the reader to keep
in mind that if M ⊂ M1 ⊂ ... is the Jones tower of factors associated with a
subfactor of finite index N ⊂ M , then M L2(Mn)M =M L2(M1)⊗M ....⊗M L2(M1)M
(n-times tensor/M product). Also, if one denotes by {Hk}k∈K the list of irreducible
Hilbert M -bimodules appearing in {L2(Mn)}n, then for any m ≥ 1 and any nonzero
1 ∩ Mm we have M L2(M1)M ⊂M L2(p′Mmp′)M , and thus we
projection p′ ∈ M ′
recover all {Hk}k in the tensor powers M L2(p′Mmp′)n⊗M
M , n ≥ 1. Equivalently, if
M ֒→ p′Mmp′ = M 0 and M 0 ⊂ M 0
1 ⊂ ... is its Jones tower, then {Hk}k coincides
with the list of irreducible Hilbert M -bimodules appearing in L2(M 0
n), n ≥ 1 (see
[P91] and [Bi97] for basics of subfactor theory).
2.4. Lemma. Let N ⊂ M be an extremal inclusion of II1 factors of index [M :
N ] = λ−1 < ∞, T = M ⊗M
= S its SE inclusion of II1 factors
M
op
op
⊂ M ⊠
eN
VANISHING COHOMOLOGY
15
and ... ⊂ Nm ⊂ ... ⊂ N ⊂e−1 M ⊂eN =e0 M1 ⊂e1 ...Mm ⊂em .... a tunnel-tower for
N ⊂ M inside S.
1◦ If en
−n ∈ Mn+1 ⊂ S denotes the projection of trace [M : N ]−n−1 obtained
−n
−n) = S. Thus, the map that acts as
−n gives a natural identification between
as a scalar multiple of the word of maximal length in e−n, ..., e0, ...., en, then en
op
implements EM
Nn , EM
N op
n
the identity on M ∨M
M ∨ M
M
and one has vN (T, en
and takes eNn to en
and T ⊂ S.
op
op
op
⊂ M ⊠
eNn
2◦ Let p ∈ P(N ′
op
op
n ∩ M ) and p
of S. Then Nn+1pp
tiautomorphism
sic construction for (V ⊂ U ) = (Nn+1pp
f = τ (p)−1pp
= τ (p)−1pen
op
op
en
−npp
op
op
⊂ pM pp
∈ M ′ ∩ Mn+1 ⊂ S its image under the an-
is a ba-
), with Jones projection
. Moreover U ⊠
is
eV
⊂ pM pp
f p
−np = τ (p)−1p
Mn+1pp
⊂ pp
U
op
op
op
op
op
op
op
op
op
op
op
M
naturally embedded into S as the von Neumann subalgebra generated by pM pp
pp
p
equal to pp
and the amplification by τ (p)2 of (T ⊂ S), with eV 7→ f .
,
n ∩ N , then this latter algebra is actually
)
and f . If in addition p ∈ N ′
Spp
, thus giving a natural identification between (U ⊗U
⊂ U ⊠
eV
U
op
op
op
op
3◦ Let p ∈ P(N ′
n ∩ N ) be as in the last part of 2◦. Let P ⊂ N be a subfactor
op
= T ⊂ S1 = M ⊠
eP
such that P ⊂ M is a downward basic construction for M ≃ M p
and denote M ⊗M
an orthonormal basis of N over P , then e = Pj mjeP m∗
is a
projection of trace λ = [M : N ]−1 in S1 that implements both EM
N op and
satisfies vN (T, e) = S1, thus giving an identification between T ⊂ S and T ⊂ S1,
with eN 7→ e.
op
eP m
j
op
N and EM
its SE inclusion. If {mj}j ⊂ N is
j = Pj m
Mn+1p
⊂ p
op
j
M
op
∗
op
op
op
Proof. Part 1◦ is (Proposition 2.9(a) in [P97]), the first part of 2◦ is (Lemma 2.8.(c)
and 2.9(c) in [P97]), while the first part of 3◦ is (Proposition 2.10 in [P97]).
To prove the last part of 2◦, note that with the notation U = pM pp
op
op
op
op
op
U
Spp
⊂ pp
k⊗H′
we have U L2(S0)U ⊂ U L2(pp
S0 = U ⊠
eV
bimodules. Then notice that by (Theorem 4.5 in [P97]), U L2(pp
)U =
⊕k∈KH′
of all distinct ir-
reducible Hilbert M -bimodules in ∪nL2(Mn). Since the list of irreducible U -
bimodules in the Jones tower for Nnp ⊂ pM p contains all {H′
k}k∈K (because
p ∈ N ′
)U ⊂ U L2(S0)U as well. Thus, the
inclusion U L2(S0)U ⊂ U L2(pp
k}k∈K denotes the reduction by pp
)U is an equality, forcing S0 = S as well.
n ∩ N ), it follows that U L2(pp
k , where {H′
Spp
Spp
Spp
Spp
op
op
op
op
op
op
op
and
)U as Hilbert
op
op
The last part of 3◦ follows by taking again into account (Lemma 2.8. (c) in [P97]),
part 2◦ above and the remark before the statement of the lemma, then using the
16
SORIN POPA
same "exhaustion by bimodules" argument used above (based on Theorem 4.5 in
[P97]).
(cid:3)
2.5. Definition. Let N ⊂ M be an inclusion of II1 factors with finite index and
denote by N ⊂ M ⊂ M1 ⊂ M2 ⊂ ... its Jones tower. A subfactor P in N is said to
be (N ⊂ M )-compatible if there exist n ≥ 1 and a non-zero projection p′ ∈ M ′
1 ∩Mn
such that P p′ ⊂ M p′ ⊂ p′Mnp′ is a basic construction. Let us note right away that
this property is in some sense "hereditary":
2.6. Lemma. With N ⊂ M as above, assume P ⊂ N is (N ⊂ M )-compatible. If
a subfactor Q ⊂ P is (P ⊂ M )-compatible, then Q ⊂ N is (N ⊂ M )-compatible.
Proof. Let P ⊂ N ⊂ M ≃ M p′ ⊂ M1p′ ⊂ p′Mnp′ be a basic construction, for
some n ≥ 1 and a non-zero projection p′ ∈ M ′
1 ∩ Mn. Note that if we denote
V = M p′ ⊂ p′Mnp′ = U then, given any m ≥ 1, its associated Jones tower of
factors up to m, V1 ≃ P ⊂ M ≃ V ⊂ U ⊂ U1 ⊂ ...Um, can be realized (up to
isomorphism) by inducing/reducing in the initial tower M ⊂ M1 ⊂ M2... ⊂ Mk,
for some large enough k, with the projections involved p′
ij ∩ Mij+1 , where
i0 = p′, i1 = n, and i0 < i1 < .... This shows that if Q ⊂ P is (P ⊂ M )-
i0 = 1, p′
compatible, then one obtains a basic construction Q ⊂ M ≃ M q′ ⊂ q′Mk0 q′, for
some appropriate k0 and q′ ∈ M ′
ij . But this
means that Q ⊂ N is (N ⊂ M )-compatible.
(cid:3)
1 ∩ Mk0 obtained as a product of such p′
ij ∈ M ′
For the reader's convenience, we recall here two of the equivalent definitions of
amenability for "group-like" objects arising from subfactors, that we have intro-
duced and studied in ([P91], [P93], [P94b], [P97]), and that we need hereafter.
The standard invariant GN⊂M of an extremal inclusion of factors with finite
Jones index N ⊂ M is amenable if its principal graph ΓN⊂M satisfies the Kesten-
type condition kΓN⊂M k2 = [M : N ].
This very first definition of amenability was proposed in [P91], and we will also
take it to be the definition of amenability for the various abstractizations of standard
invariants: a standard λ-lattice G as in [P04b] (or a planar algebra as in [J99]) is
amenable if its graph ΓG satisfies the condition kΓGk2 = λ−1.
An alternative notion of amenability was introduced in [P93], by requiring the
following Følner-type condition on G = GN⊂M :
let {dk}k∈K denote the standard
weights of ΓN⊂M (resp. ΓG ), obtained for instance as dim(M HkM )1/2, where
{Hk}k∈K is the list of irreducible M − M Hilbert bimodules appearing at even
levels in G, indexed by the set K of left vertices of the bipartite graph ΓG = ΓN⊂M ;
G satisfies the Følner condition if for any ε > 0 there exists a finite set F ⊂ K
such that if one denotes by ∂F = {k ∈ K \ F ∃k0 ∈ F with (ΓGΓt
G)kk0 6= 0} (the
boundary of F ) then P
k∈∂F
d2
k < ε P
k∈F
d2
k.
VANISHING COHOMOLOGY
17
These two conditions were shown equivalent in ([P97] Theorem 5.3; the result had
already been announced in [P93] and [P94b]). Several other equivalent amenabil-
ity conditions were in fact introduced and studied in [P97], notably a local finite
dimensional approximation property ([P97], Theorem 6.1) which will be crucial for
the proof of Theorem 2.10 below.
The Kesten and the Følner-type conditions provide in particular equivalent def-
initions of amenability for a finitely generated rigid C∗-tensor category of Hilbert
bimodules G (as defined for instance in [PV14]), having {Hk}k∈K as irreducible
morphisms. One then defines amenability for an arbitrary (possibly infinitely gen-
erated) rigid C∗-tensor category by requiring that any finitely generated subcate-
gory is amenable (see the detailed definitions in the paragraphs preceding Theorem
2.12).
Due to its various equivalent characterizations, amenability in this context is
⊂
= S associated with an extremal inclusion N ⊂ M coincides with the
a very "robust" property. For instance, since the SE inclusion T = M ⊗M
M ⊠
eN
M
op
op
op
op
M
⊂ M ⊠
eNn
associated with Nn ⊂ M , for any n ≥ 0
SE inclusion M ⊗M
(e.g., by Lemma 2.4.1◦ above), it follows from (Theorem 5.3 in [P97]) that GN⊂M
is amenable iff GMi⊂Mj
is amenable for some (and thus all) i < j. Note that
this can also be deduced from the fact that ΓMi⊂Mj
is a alternate product of
ΓN⊂M and its transpose, or of ΓM⊂M1 and its transpose, j − i times, which shows
that one always have kΓMi⊂Mj k = kΓN⊂M kj−i (cf. 1.3.5 in [P91]), implying that
kΓN⊂M k2 = [M : N ] iff kΓMi⊂Mj k2 = [Mj : Mi].
Moreover, if p′ is a non-zero projection in M ′
i ∩ Mj, for some i < j in Z, then by
(Coroally 6.6 (ii) in [P97]) ΓMi⊂Mj amenable (which we already know is equivalent
to ΓN⊂M being amenable) implies V = Mip′ ⊂ p′Mjp′ = U has amenable graph
as well. Also, note that if j ≥ i + 1 and p′ ∈ M ′
i+1 ∩ Mj, then by Lemma 2.4 and
the above argument, one conversely has that V ⊂ U amenable implies N ⊂ M
amenable.
We record all these facts in the following:
2.7. Proposition. Let N ⊂ M be an extremal inclusion of factors with finite
index and {Mi}i∈Z be a tunnel/tower of factors for N ⊂ M . If GN⊂M is amenable,
then for any i < j in Z and any non-zero projection p′ ∈ M ′
i ∩ Mj, the inclusion
Mip′ ⊂ p′Mjp′ has amenable standard invariant.
If in addition j ≥ i + 1 and
p′ ∈ M ′
i+1 ∩ Mj, then conversely GMip′⊂p′Mj p′ amenable implies GN⊂M amenable.
2.8. Lemma. Let N ⊂ M be a finite index extremal inclusion of II1 factors with
. Given any (N ⊂ M )-
amenable standard invariant and SE factor S = M ⊠
eN
M
op
18
SORIN POPA
compatible subfactor P ⊂ N and any ε > 0, there exists a (P ⊂ M )-compatible
subfactor Q ⊂ P such that kE(Q′∩M )′∩S(x) − EM ′∩S(x)k2 ≤ εkxk, ∀x ∈ P ′ ∩ S.
Proof. Let us first note a few Facts needed in the proof.
Fact 1. It is sufficient to show that there exists a compatible subfactor Q ⊂ P
with the property that kEQ′∩M )′∩S(f ) − τ (f )1k2 ≤ ε/([M : P ] + 1)5/2, where f ∈ S
is the Jones projection for P ⊂ M ⊂ hM, P i, viewed inside S (cf. 2.4.3◦ above).
Indeed, because if {mj}j ⊂ M is an orthonormal basis of M over P with [M :
j f }j is
, f i is of
P ]+1 many elements of norm ≤ [M : P ]1/2 (cf. [PiP84]), then {[M : P ]1/2m
an orthonormal basis of P ′∩S over M ′∩S = M
the form x = Pj[M : P ]1/2m
j , where y
has operator norm majorized by [M : P ]1/2kxkkmjf k = [M : P ]1/2kxk, thus giving
the estimates
and any x ∈ P ′∩S = hM
∗
x) ∈ M
op
op
j = [M : P ]1/2EM op (f m
j
op
op
op
j f y
op
op
op
kE(Q′∩M )′∩S(x) − EM ′∩S(x)k2
= k X
[M : P ]1/2m
op
j (E(Q′∩M )′∩S(f ) − τ (f )1)y
op
j k2
j
≤ [M : P ]1/2 X
kmjkky
op
j kkE(Q′∩M )′∩S(f ) − τ (f )1k2
j
≤ [M : P ]3/2([M : P ] + 1)kxkε/([M : P ] + 1)5/2 < εkxk.
Fact 2. By Proposition 2.7 above, GN⊂M amenable implies GP ⊂M amenable.
Fact 3. By (Theorem 6.1 in [P97]), if P ⊂ M is an extremal inclusion of factors
with amenable standard invariant then for any δ > 0 there exists n ≥ 1 and a
projection p ∈ P ′
2 < δτ (p),
where ... ⊂ Pn ⊂ Pn−1 ⊂ ...P ⊂ M ⊂f0=eP hM, P i denotes a Jones tunnel and
basic construction for P ⊂ M .
n∩M )p′∩phM,P ip(f0p) − τ (f0)pk2
n ∩ P such that kEp(P ′
Let us now proceed with the proof of the lemma. By Fact 2, P ⊂ M is amenable
so we can apply Fact 3 to (P ⊂ M ⊂f0 hM, P i), to get an n ≥ 1 and a projection
p ∈ P ′
n ∩ P such that
(2.8.1)
kEp(P ′
n∩M )p′∩phM,P ip(f0p) − τ (f0)pk2 < εkpk2/([M : P ] + 1)2
By amplifying by α = τ (p)−1 the inclusions of factors
(2.8.2)
Pnp ⊂ pP p ⊂ pM p ⊂f0p phM, P ip
VANISHING COHOMOLOGY
19
using partial isometries in P , we obtain inclusions of factors
(2.8.3)
(Pnp)α = Q ⊂ P ⊂ M ⊂f0 hM, P i
having same relative commutants as (2.8.2). Thus, by (2.8.1), it follows that Q
satisfies
(2.8.4)
kEQ′∩hM,P i(f0) − τ (f0)k2 < ε/([M : P ] + 1)2,
By the way it is defined, Q ⊂ P is an (P ⊂ M )-compatible subfactor, and thus
(N ⊂ M )-compatible as well, while by Fact 1 and (2.8.4) we have kE(Q′∩M )′∩S(x)−
EM ′∩S(x)k2 ≤ εkxk, for all x ∈ P ′ ∩ S.
(cid:3)
2.9. Lemma. Let N ⊂ M be a finite index extremal inclusion of II1 factors with
N ⊂ M ⊂ M1.... ր M∞ its Jones tower of factors and S its SE factor. If B ⊂ M
is a diffuse von Neumann subalgebra, then B 6≺M∞ M ′ ∩ M∞ and B 6≺S M
Proof. By [P03], in order to prove B 6≺M∞ M ′ ∩ M∞, it is sufficient to prove that
for any finite set F in a given total subset X of M∞, there exist un ∈ U(B) such
that limn kEM ′∩M∞(y∗unx)k2 = 0, ∀x, y ∈ F . Taking X = ∪mMm, it is sufficient
to show this for any m and any finite F ⊂ Mm. But by [P03] this amounts to
M 6≺ M ′ ∩Mm, which is trivial since B is diffuse and M ′ ∩Mm is finite dimensional.
,
which is total in S by (Section 4 in [P97]). Thus, if F ⊂ X is finite then we
2 ∈ F , with
may assume F ⊂ MmM
x1, y1 ∈ Mm, x2, y2 ∈ M , and we take un ∈ U(B), then we get the estimate
we use the same criterion, but with X = ∪mMmM
for some large m so if x = x1x
To prove B 6≺S M
op
.
op
op
2 , y = y1y
op
op
op
kEM op (y
∗
op
2
y1
∗unx1x
op
2 )k2 ≤ kx2kky2kkEM ′∩Mm(y∗
1unx1)k2.
This shows that it is actually sufficient to check the criterion for F ⊂ Mm, which
(cid:3)
amounts again to M 6≺ M ′ ∩ Mm as before.
2.10. Theorem. Let N ⊂ M be a finite index extremal inclusion of separable II1
factors with amenable standard invariant. There exists a sub-inclusion of hyperfinite
factors (Q ⊂ R) ⊂ (N ⊂ M ) making a non-degenerate commuting square that's
strongly smooth, i.e., Q′ ∩ Rn = Q′ ∩ Mn = N ′ ∩ Mn and R′ ∩ Rn = R′ ∩ Mn =
M ′ ∩ Mn, ∀n, where N ⊂ M = M0 ⊂e0 M1 ⊂ ... is the Jones tower for N ⊂ M
and Rn = vN (R, e0, ..., en−1), n ≥ 1, the tower for Q ⊂ R. Moreover, Q ⊂ R can
be obtained as Q = ∪nP ′
n ∩ M = R, for some decreasing sequence of
(N ⊂ M )-compatible subfactors M ⊃ N ⊃ P0 ⊃ P1....
n ∩ N ⊂ ∪nP ′
20
SORIN POPA
Proof. We split this proof into several parts.
Fact 1. There exists a sequence of (N ⊂ M )-compatible subfactors .... ⊂ Pn ⊂
...P0 ⊂ N ⊂ M such that if we define Q = ∪mP ′
m ∩ M = R and Rn =
∪mP ′
m ∩ Mn, n ≥ 1, then (Q ⊂ R) ⊂ (N ⊂ M ) is a non degenerate commuting
square of II1 factors, Q ⊂ R ⊂e0 R1 ⊂e1 R2... is its Jones tower and Q′ ∩ Rn =
N ′ ∩ Mn, R′ ∩ Rn = M ′ ∩ Mn, ∀n.
m ∩ N ⊂ ∪mP ′
To see this, let M ∨ M
op
⊂ M ⊠
eN
op
M
= S be the SE inclusion of factors
associated with N ⊂ M . By applying recursively Lemma 2.8, we obtain a sequence
of subfactors M ⊃ N = P0 ⊃ P1... such that for each n ≥ 1 we have
op
n−1 ∩ S.
n ∩ N , R = ∪nP ′
n ∩ M . If we denote S0 = ∪nP ′
(a) Pn ⊂ Pn−1 is (Pn−1 ⊂ M )-compatible (thus also (N ⊂ M )-compatible).
(b) kE(P ′
Let Q = ∪nP ′
n∩M )′∩S(x) − EM ′∩S(x)k2 ≤ 2−nkxk, ∀x ∈ P ′
n ∩ S , then by
property (b) above, it follows that R′∩S0 = M
. In particular, R is a II1 factor. By
the definitions of Q, R, it follows that (Q ⊂ R) ⊂ (N ⊂ M ) is a commuting square,
with e0 = eN implementing the conditional expectation of R onto Q and Q =
{e0}′ ∩ R. From (b), one also gets ER′∩S(e0) = λ1. This implies that the algebra
R0
1 := spRe0R has support 1 and thus any orthonormal basis {mj}j of R over Q
must "fill up the identity", i.e., Pj mje0m∗
j = 1. Hence, (Q ⊂ R) ⊂ (N ⊂ M ) is in
′ ∩ S0 = {e0}′ ∩ M
fact a non-degenerate commuting square. Moreover, R0
,
1
implying that R0
1e0 is a II1 factor as well,
and R0
n := vN (R, e0, ..., en−1), n ≥ 1, is the Jones tower for Q ⊂ R.
1 is a II1 factor. Thus, Q ≃ Qe0 = e0R0
= N
At the same time, if for each n ≥ 1 we define Rn = ∪mP ′
m ∩ Mn, then both this
1 ⊂ ... make (non-degenerate) commuting
n = Rn and Rn,
n−1 be each other's commutant in S0, ∀n ≥ 1. Note that this also implies that
sequence and the sequence Q ⊂ R ⊂ R0
squares with N ⊂ M ⊂ M1 ⊂ ..., with R0
N
for the downward continuation of these towers we have Q′ ∩ S0 = M op
1 .
n ⊂ Rn. This forces R0
op
op
op
So for the higher relative commutants, we have the equalities
Q′ ∩ Rn = (Q′ ∩ S0) ∩ Rn = (M
op
1 ∩ Mn) ∩ Rn
= (N ′ ∩ Mn) ∩ Rn ⊂ Q′ ∩ Rn = M
finishing the proof of Fact 1.
op
1 ∩ Mn ∩ (N
op
n−1)′ = N ′ ∩ Mn,
Fact 2. Assume ... ⊂ P1 ⊂ P0 ⊂ N ⊂ M are as in Fact 1. If un ∈ U(Pn), n ≥ 0,
n is an (N ⊂ M )-compatible subfactor
0 ⊂ N ⊂ M is a sequence of factors still
and we define P n
in P n−1
satisfying the conditions in the statement of Fact 1.
n...u∗
n−1 ⊂ ...P 1
0, then P n
1 ⊂ P 0
n = u0...unPnu∗
n−1 and ...P n
n ⊂ P n−1
VANISHING COHOMOLOGY
21
′ ∩ N ⊂ P n
n
n ∩ Mk}n, {P n
n
Indeed, for each k the systems of commuting squares of algebras {P ′
n ∩ N ⊂
′ ∩ Mk}n (standard λ-lattices
P ′
n ∩ M ⊂ ...P ′
n....u∗
in the sense of [P94a]) are isomorphic via the map Φ(x) = limn u0...unxu∗
0,
n ∩ Mk. Thus, Φ implements an isomorphism between R′ ∩ Ri ⊂ Q′ ∩ Ri
x ∈ ∪nP ′
and R0′ ∩ R0
′ ∩ Mi, i ≤ n. Since the isomorphism
Φ leaves N ′ ∩ Mi = Q′ ∩ Ri and M ′ ∩ Mi = R′ ∩ Ri fixed, by equality of dimensions
via Φ it follows that N ′ ∩ Mi = Q′ ∩ Ri, M ′ ∩ Mi = R′ ∩ Ri, ∀i.
i ⊂ Q0′ ∩ R0
′ ∩ M ⊂ ...P n
n
i , where R0
i = ∪nP n
n
Fact 3. Assume ... ⊂ P1 ⊂ P0 ⊂ N ⊂ M are as in Fact 1. Then there exist
integers k0 = 0 < k1 < ... and unitaries vn ∈ U(Pkn−1 ), n ≥ 0, such that if we
′ ∩ M = R, then
define P n
kn
Q′ ∩ Mm = Q′ ∩ Rm, R′ ∩ Mm = R′ ∩ Rm, ∀m.
1 and let Q = ∪nP n
kn
′ ∩ N ⊂ ∪nP n
kn
= v1...vnPkn v∗
n...v∗
To show this, let {bk}k ⊂ (∪nMn)1 be a k k2-dense sequence. We choose recur-
sively km > km−1, vm ∈ U(Pkm−1 ) such that
(F3)
kE(P m
km
′∩P m−1
km−1
)′∩Mm
(bj) − EP m−1
km−1
′
∩Mm
(bj)k2 ≤ 2−m, ∀j ≤ m.
Assume we made this construction up to m = n. Due to Lemma 2.9, (Theorem 0.1
(a) in [P13]) implies that if A0 ⊂ P n
is a finite dimensional abelian von Neumann
kn
subalgebra with all minimal projections of sufficiently small trace, then there exists
′∩Mn+1 (bj)k2 < 2−n−1, ∀j ≤ n + 1.
u ∈ U(P n
kn
For each m ≥ kn denote P n
0. Since B = ∪jP n
is diffuse
j
(because it contains a Jones sequence of λ = [M : N ]−1 projections, which generate
a copy of the hyperfinite II1 factor by [J83]), we may assume A0 ⊂ B, and hence
) such that kEuA0u′∩Mn+1 (bj) − EP n
n...v∗
m = v0...vnPmv∗
′ ∩ P n
kn
kn
kEuBu′∩Mn+1(x) − EP n
kn
′∩Mn+1 (bj)k2
< kEuA0u′∩Mn+1(bj) − EP n
kn
′∩Mn+1 (bj)k2 < 2−n−1.
Since B is a "limit" of P n
j
′ ∩ P n
kn
, for j sufficiently large we'll still have
kE(uP n
j u∗ ′∩P n
kn
)′∩Mn+1 (bj) − EP n
kn
′∩Mn+1 (bj)k2 < 2−n−1, ∀j ≤ n + 1.
We choose such a large j and put kn+1 = j. Letting vn+1 = v∗
P n+1
kn+1
n + 1.
1uv1...vn ∈ P n
,
kn
u∗, we see that (F3) is satisfied for m =
= v1...vn+1Pn+1v∗
1 = uP n
n+1...v∗
n...v∗
If we now define Q = ∪nP n
kn
′ ∩ M = R, then condition (F3)
implies Q′ ∩ Mn ⊂ Q′ ∩ Rn, ∀n, while Fact 2 implies we have Q′ ∩ Rn = N ′ ∩ Mn.
′ ∩ N ⊂ ∪nP n
kn
kn+1
22
SORIN POPA
The calculations for the relative commutants of R are similar, thus finishing the
proof.
(cid:3)
Note that the case "Γ finitely generated" of Theorem 2.1 already follows from
Theorem 2.10 above, due to the observation we made just before stating Theorem
2.1. But deriving from this the case "Γ infinitely generated" is problematic, as
applying it for "larger and larger" finitely generated subgroups of Γ would involve
in the limit multiplying infinitely many perturbing unitaries. To deal with this
problem we'll use a similar trick, but with a "diagonal embedding" of N ≃ N σ into
the algebra of matrices "of size Γ" over N , M σ ≃ N ⊗B(ℓ2Γ).
Lemma 2.11. Let {σg}g∈Γ ⊂ Aut(N ) be an outer action of a group Γ on a II1
factor N , with σe = idN , and vg,h ∈ U(N ) satisfying σgσh = Ad(vg,h)σgh, ve,h =
ve,g = 1, ∀g, h ∈ Γ. Define M σ = N ⊗B(ℓ2Γ) and N σ = {Pg σg(x)e0
gg x ∈ N } ⊂
M σ, where {e0
gg ′}g,g ′∈Γ ⊂ B(ℓ2Γ) are the usual matrix units. Let R ⊂ M σ be a
subfactor, with the property that {e0
gg}g ⊂ R and let Q ⊂ R ∩ N σ be a common
subfactor such that Qe0
gg = e0
ggRe0
gg, ∀g ∈ Γ.
1◦ Let Q0 ⊂ N be the unique subfactor with Q0eee = Qeee. There exist wg ∈
U(N ), we = 1, such that σ′
g := Ad(wg) ◦ σg satisfy σ′
g(Q0) = Q0, ∀g ∈ Γ.
2◦ If in addition Q′ ∩ R = N σ ′ ∩ R = {e0
is outer, ∀g ∈ Γ, g 6= e,
and v′
g,h = wgσg(wh)vg,hw∗
gh normalize Q0, ∀g, h.
gg}′′
g , then σ′
gQ0
3◦ Let e ∈ Fi = F −1
e0
gg. Then the inclusions of II1 factors (Qq0
i ր Γ be the net of finite symmetric subsets, and denote q0
i =
i M σq0
Pg∈Fi
i )
make a non-degenerate commuting square (with respect to the trace preserving con-
ditional expectations), ∀i.
i ) ⊂ (N σq0
i ⊂ q0
i ⊂ q0
i Rq0
4◦ Denote M σ
1
:= M σ⊗B(ℓ2Γ) ≃ M ⊗B(ℓ2Γ)⊗B(ℓ2Γ) and consider the em-
gg, where σ = σ ⊗ 1 and
e1
gg, then
1 given by j1(x) = P−1
g σg(x)e1
gg ′}g,g ′ are the matrix units of the 2nd copy of B(ℓ2Γ). If q1
i = Pg∈Fi
bedding j1 : M σ ֒→ M σ
{e1
N σ ⊂ q0
i )q1
i M σq0
i M σq0
i ≃ j1(q0
i ⊂ q0
i q1
i M σ
1 q0
i q1
i is a basic construction.
5◦ With the above notations let Ri
i q1
1 ⊂ q0
i q1
i M σ
i )′ ∩ Ri
1 = (N σq0
1 is a basic construction. If (Qq0
Ri
{σ′
gQ0
0 ∩ N = C), then v′
Q′
H 3(Γ)-obstruction as σ.
i Rq0
i ֒→j1
i , ∀i, then
}g is an outer action of Γ on Q0. If in addition Q′ ∩ N σ = C (equivalently,
}g ⊂ Out(Q0) has same
g,h ∈ Q0, ∀g, h ∈ Γ, and {σ′
i be so that Qq0
i q1
i M σ
i ⊂ q0
1 q0
i q1
i )′ ∩ q0
1 q0
i q1
gQ0
Proof. 1◦ Identifying N = N ⊗ 1, let j0 : N ≃ N σ be the isomorphism defined by
the property j0(x)e0
ee, x ∈ N . Thus, Q0 = j−1
ee = xe0
0 (Q).
VANISHING COHOMOLOGY
23
ee = e0
Since R is a (necessarily semifinite) factor and e0
trace, there exist partial isometries e′
∀g, and e′
elements wg ∈ N such that e′
Aut(N ), g ∈ Γ. Note that for any x ∈ N we have σ′
addition x ∈ Q0, then e′
Thus, σ′
gg belong to R and have same
gg,
ee. From the way M σ is defined, it follows that there exist unitary
g = Ad(wg) ◦ σg ∈
∗. If in
ee.
eg, ∀g, with we = 1. Let σ′
eg ∈ R with left support e0
ee which is equal to Qe0
ee, right support e0
∗ lies in e0
ee = Q0e0
eg = wge0
egj(x)e′
eg
egj(x)e′
eg
ee = e′
g(x)e0
eeRe0
gσ′
h = Ad(v′
g,h)σ′
gh, ∀g, h ∈ Γ. Since σ′
g normalize Q0, it follows
g(Q0) = Q0.
We clearly have σ′
that v′
g,h normalize Q0, ∀g, h.
2◦ Note that if g, g′ ∈ Γ, then e0
is equivalent to σ′
g 6= σ′
this to g′ = e, proves 2◦.
g ′ in Out(N ) (resp. σ′
gg(N σ′∩M σ)e0
gQ0
g ′g ′ = 0 (resp. e0
gg(Q′∩R)e0
g ′g ′ = 0)
in Out(Q0)). Applying
6= σ′
g ′ Q0
3◦ The fact that the inclusions make a commuting square is obvious from 1◦
above.
4◦ This is in (Section 5.1.5 in [P91]).
5◦ Note that (Qq0
i )′ ∩ q0
q0
i Rq0
already know that σ′
i q1
i M σq0
gQ0
i = (N σq0
i )′ ∩ Ri
is outer, ∀g 6= e.
1 = (N σq0
i )′ ∩
i , ∀i, which implies Q′ ∩ R = N σ ′ ∩ M σ. So by 2◦, we
i , ∀i, implies (Qq0
i )′ ∩ q0
i M σ
1 q0
i q1
i q1
i q1
i Rq0
i
֒→ Ri
i ⊂ q0
i ⊂ q0
Fix g, h ∈ Γ and let Fi be sufficiently large so that g, h, gh ∈ Fi. By taking into
0 and
1, as well as (Section 5.1.5 in [P97]), we see like in the proof
gh from the
gh in Out(N ) iff
g,h) implements
g,h ∈ Q0,
have the same scalar
(cid:3)
account the form of the basic constructions N σq0
Qq0
of 2◦ above that the equality of relative commutants multiplied by e0
left and by e0
σ′
gQ0
inner automorphisms on Q0. If in addition Q′
∀g, h ∈ Γ. But if {v′
H 3(Γ)-obstruction.
h not equal to σ′
in Out(Q0). This shows that Ad(v′
g,h}g,h ⊂ Q0, then σ′ (thus σ) and σ′
hhfrom the right implies that σ′
0 ∩ N = C, then this forces v′
gge1
not equal to σ′
i ֒→ q0
i M σq0
0M σq0
σ′
hQ0
eee1
ghQ0
i q1
i q1
gσ′
Q0
Proof of Theorem 2.1. By Proposition 1.5.2◦, we may assume N is separable. Also,
without loss of generality, we may assume σe = idN .
Like in Lemma 2.11, we denote {e0
and let M σ := N ⊗B(ℓ2Γ), j0 : N ֒→ M σ given by j0(x) = Pg σg(x)e0
j0(N ) ⊂ M σ . Choose an increasing sequence of finite sets e ∈ Fn = F −1
exhaust Γ and denote q0
gh}g,h∈Γ ⊂ B(ℓ2Γ) the canonical matrix units
gg, N σ =
n ⊂ Γ that
n = Pg∈Fn
e0
gg.
As in Lemma 2.11.5◦, and with the notations established there, the embedding
of factors N ≃j0 N σ ⊂ M σ implements (by induction/reduction) a sequence of
n ≃ N Fn whose basic construction
inclusions of II1 factors N ≃ N σq0
nM σq0
n ⊂ q0
24
SORIN POPA
identifies, via j1(·)q1
n, with q0
nq1
nM σ
1 q0
nq1
n.
n ⊂ q0
nM σq0
For each n, the subfactor N σq0
n is a locally trivial extremal inclusion
of II1 factors with standard graph given by the Cayley graph of the subgroup Γn =
hFni ⊂ Γ (see 5.1.5 in [P91]), which is amenable. Using this, we apply Theorem
2.10 to construct recursively a sequence of subfactors of finite index N ⊃ P0 ⊃
P1 ⊃ ... such that if we denote Q0 = ∪nP ′
n ∩ N , R = ∪nj0(Pn)′ ∩ M σ, then R is an
irreducible type II subfactor in M σ containing the finite projections {e0
gg}g, with
Q := j0(Q0) ⊂ N σ ∩ R a II1 subfactor satisfying the condition Q′ ∩ R = Q′ ∩ M σ =
N σ ′ ∩ M σ = {e0
ggRegg, ∀g. By Lemma 2.11.5◦
the inclusion (Qq0
n) is a non-degenerate commuting
square and our construction will show that Rn
n is the
basic construction algebra for the subfactor (Qq0
n ⊂ q0
n), ∀n, where j = j1 ◦ j0.
The desired conclusions will then follow from 2.11.5◦.
gg g ∈ Γ}′′ and with Qe0
n ⊂ q0
n ⊂ q0
1 := ∪mj(Pm)′ ∩ q0
gg = e0
nM σq0
n) ⊂ (N σq0
nRq0
nRq0
nM σ
1 q0
nq1
nq1
Let T r denote the semifinite trace τ ⊗ T rB(ℓ2Γ) ⊗ T rB(ℓ2Γ) on M σ
1 and k k2,T r the
corresponding L2-norm. Let Y = {yn}n ⊂ (M σ
1 , T r) be a k k2,T r-dense
sequence and denote by Yn = {j1(q0
j ykj1(q0
j 1 ≤ j, k ≤ n}. We construct
the decreasing sequence of subfactors Pm ⊂ N recursively, such that P0 = N and
for m ≥ 1
1 )1 ∪ L2(M σ
j )q1
j )q1
1
m)′∩M σ
mM σq0
m)-compatible.
(y) − Ej1(N σ )′∩M σ
m is (N σqm ⊂ q0
(i) (j0(Pm) ⊂ N σ)q0
(y))k2,T r < 2−m/Fm, ∀y ∈ Ym.
(ii) kEj1((j0(Pm)′∩N σ )q0
If we made this construction up to m = n, then by applying Theorem 2.10 to
the inclusion of II1 factors j0(Pn)q0
n+1 (which has amenable graph
by Proposition 2.7, but this is trivial here, because this subfactor is in fact locally
trivial, with standard graph given by a Cayley graph of an amenable subgroup of
Γ, see 5.1.5 in [P97]) we get a subfactor Pn+1 ⊂ Pn so that its image via j0( · )q0
is (N σq0
m = n + 1.
n+1
n+1)-compatible and such that (ii) above is satisfied for
n+1 ⊂ q0
n+1M σq0
n+1M σq0
n+1 ⊂ q0
1
With the sequence {Pn}n this way constructed, define Q0, Q and R as explained
gg, ∀g, by construction.
gg}g and satisfies Qe0
above. Then R contains {e0
gg = e0
ggRe0
Moreover, condition (ii) shows that
j(Q)j1(q0
n)q1
n
′
∩ j1(q0
n)q1
nM σ
1 j1(q0
n)q1
n
= j1(q0
n)q1
n(j1(N σ)′ ∩ M σ
1 )j1(q0
n)q1
n = j(Q)′ ∩ Rn
1 .
From all this, it also follows that Q′
isomorphic to R by ([MvN43]).
0 ∩ N = C and that Q0 is AFD, and thus
Thus, 2.11.5◦ applies, to conclude that there exist {wg}g ⊂ U(N ) such that
σ′
g = Ad(wg)σg normalizes the irreducible hyperfinite II1 subfactor Q0 ⊂ N , ∀g,
VANISHING COHOMOLOGY
25
v′
g,h = wgσg(wh)vg,hw∗
obstruction.
gh belong to Q0, ∀g, h, and σ, σ′
Q0
have the same H 3(Γ)
(cid:3)
We end this section by noticing that the same argument we used in the proof
of Theorem 2.1 above shows that any "amenable family" G of endomorphisms
(or Hilbert bimodules) over a II∞ factor N normalizes a "large" AFD subfactor
R⊗B(ℓ2N) ⊂ N , on which it "acts faithfully". To state this result, we need to clar-
ify the terminology and fix some notations. Also, the reader should recall that for
a properly infinite factor N (of type II∞ in our case) one has Connes' well known
correspondence between a Hilbert N -bimodule N HN and a N -endomorphisms θH
(see [C80]; see also Sec. 1.1 in [P86]). Via this correspondence, the adjoint opera-
tion on endomorphisms, θ 7→ ¯θ, is given by ¯θH = θ ¯H and tensor product H⊗N H′
corresponds to composition θH ◦ θH′ (see e.g., 1.3 in [P86]). It is this "nice" cor-
respondence between Hilbert-bimodules and endomorphisms for properly infinite
factors that imposes using the framework of II∞ (i.e., infinite amplification of a
II1 factor), rather than II1 factors, as algebras on which a category "acts" (the II1
framework would instead require considering morphisms between amplifications of
the II1 factor, see 1.1 in [P86]).
For us here, if N is a given II∞ factor, a concrete C∗-tensor category G of endo-
morphisms on N is a family of classes (mod perturbation by inner automorphisms
of N ) of unital endomorphisms of N that contains the idN (thus the class of in-
ner automorphisms) and satisfies the following properties: (i) it is closed to the
if θ ∈ G then ¯θ ∈ G; (ii) it is closed under composition,
adjoint operation, i.e.
i.e, if θ, θ′ ∈ G then θ ◦ θ′ ∈ G; (iii) Each θ ∈ G dilates the trace T r = T rN
by a finite scalar 1 ≤ d(θ) < ∞, i.e., T r ◦ θ = d(θ)T r, with the image subfactor
θ(N ) ⊂ N having Jones index given by the formula [N : θ(N )] = d(θ)2; (iv) it
is closed to "direct sum and subtraction", in the following sense: (a) if θ, θ′ ∈ G
and v, v′ ∈ N are isometries such that vv∗ + v′v′∗ = 1 then the endomorphism
N ∋ x 7→ θ ⊕ θ′(x) := Ad(v)θ(x) + Ad(v′)θ′(x) ∈ N belongs to G; (b) if θ ∈ G,
0 6= p ∈ P(θ(N )′ ∩ N )) and v ∈ N is an isometry with range p, then Ad(v∗)θ ∈ G.
We denote Irr(G) = {θ ∈ G θ(N )′ ∩ N = C}, the family of irreducible elements
in G, which we label by the set K = KG, as {θk}k∈K, with e ∈ K so that θe = idN
and with the adjoint operation implemented by k 7→ ¯k (thus θ¯k = ¯θk). It is easy
to see from the above properties of G that any θ ∈ G decomposes as a finitely
supported direct sum ⊕k∈K nk θk of irreducible endomorphisms in G, with finite
multiplicities nk ≥ 0 (thus 0 < Pk nk < ∞).
A subset e ∈ F0 = ¯F0 ⊂ K generates G, if any θ ∈ G can be obtained from
F0 by applying consecutively finitely many times the operations (i), (ii), (iv). If
G is generated by a finite such set F0 ⊂ K, one denotes by ΓG,F0 (or simply ΓG
26
SORIN POPA
when F0 is clear from the context) the Cayley-type bipartite graph (or matrix with
non-negative integer entries) (akk′ )k,k′∈K with akk′ equal to the multiplicity of θk′
in (Pi∈F0
θi) ◦ θk.
We say that such a family G is amenable if it satisfies the Følner-type condition
we mentioned before Proposition 2.7, i.e., if given any finite e ∈ F0 = ¯F0 ⊂ K, and
any ε > 0, there exists a finite subset F in the subcategory hF0i ⊂ G generated
d2
by F0 ⊂ K, such that P
k, where ∂F0 F is the boundary of F in
hF0i, i.e., the set of all k′ ∈ KhF0i \ F such that ak′k 6= 0 for some k ∈ F , where
(akk′)k,k′ = ΓhF0i is the graph of hF0i. By (Theorem 5.3 in [P97]), this condition is
equivalent to a Kesten-type condition, requiring that kΓhF0ik2 = (Pk∈F0
dk)2 for
any finite e ∈ F0 = ¯F0 ⊂ K.
d2
k′ < ε P
k∈F
k′∈∂F0 F
If G is a concrete C∗-tensor category of endomorphisms on N and Q0 ⊂ N is a
II∞ subfactor, then we say that G faithfully normalizes Q0, if the following condi-
tions are satisfied: (1) the trace T r is semi finite on Q0; (2) for each endomorphism
θ ∈ G there exists θ′ ∈ G in the same class as θ such that θ′(Q0) ⊂ Q0, and given
any other θ′′ ∈ G in the same class as θ that satisfies θ′′(Q0) ⊂ Q0, there exists
w0 ∈ U(Q0) with Ad(w0) ◦ θ′′
; (3) If θ ∈ G is so that θ(Q0) ⊂ Q0, then
[Q0 : θ(Q0)] = d(θ)2 = [N : θ(N )] and θ is irreducible non-inner on N iff θQ0
irreducible and non-inner on Q0.
= θ′
Q0
Q0
Note that a "concrete C∗-tensor category G of endomorphisms on a factor N "
is an analogue of a "group of outer automorphisms Γ ⊂ Out(N )".
If one still
denotes by G the underlying abstract C∗-tensor category, as defined for instance
in [NeTu13], then such an object can as well be viewed as an "outer action of
G by endomorphisms on N ". With this interpretation, if Q0 ⊂ N is faithfully
normalized by G, then the restriction map θ 7→ θQ0 is an isomorphism of the
underlying abstract rigid C∗-tensor categories. We say that such a restriction is
strongly smooth if in addition θ(Q0)′ ∩ N = θ(N )′ ∩ N , ∀θ ∈ G.
2.12. Theorem. Let N be a II∞ factor and G an amenable countably generated
concrete C∗-tensor category of endomorphisms on N , as defined above. Then N
contains an AFD II∞ subfactor R ⊂ N that's faithfully normalized by G, in the
above sense. Moreover, if N is separable, then R can be chosen so that to satisfy
the strong smoothness condition θ(N )′ ∩ N = θ(R)′ ∩ N = θ(R)′ ∩ R, for all θ ∈ G
with θ(R) ⊂ R.
Proof. The proof uses the same ideas and follows exactly the same steps as the
proof of Theorem 2.1. Note first that, as in the proof of Theorem 2.1, it is clearly
sufficient to prove the case when N is separable.
Next, note that we have a version for endomorphisms of Lemma 2.11.1◦, as
VANISHING COHOMOLOGY
27
follows:
θk(x)e0
Fact 1. Let N be a properly infinite von Neumann factor and θ = {θk}k∈K0 be a
set of endomorphisms of N , with e ∈ K0 and θe = idN . Define Mθ = N ⊗B(ℓ2K0)
and N θ = {Pk∈K0
kk′}k,k′∈K0 ⊂ B(ℓ2K0)
are the usual matrix units. Let R ⊂ Mθ be a subfactor, with the property that
{e0
kk are mutually equivalent infinite projections in R. Let Q ⊂
ee = e0
R ∩ N θ be a common subfactor such that Qe0
ee. Then there exists a
unique subfactor Q0 ⊂ N such that Q0e0
ee = Qe0
ee and there exist unitary elements
wk ∈ N , k ∈ K0, we = 1, such that θ′
kk x ∈ N } ⊂ Mθ, where {e0
k = Ad(wk) ◦ θk satisfies θ′
kk}k ⊂ R and e0
k(Q0) ⊂ Q0, ∀k.
eeRe0
The proof if the same so we omit it. Let j0 : N ≃ N θ ⊂ Mθ denote the
isomorphism satisfying j0(x)e0
ee = we0
ee, x ∈ N . Thus, Q0 = j−1
0 (Q).
In order to make the ideas more transparent, let us first give an argument for the
case when G is generated by a finite subset of irreducible endomorphisms {θk}k∈K0,
with e ∈ K0 = ¯K0 ⊂ K, by exploiting the remark before the statement of Theorem
2.1. So let N be a II1 factor such that N ∞ = N , let M θ,K0 be the amplification of
dk, {pk}k∈K0 ⊂ M θ,K0 a partition of 1 with projections having traces
N by Pk∈K0
τ (pk) proportional to dk and consider its subfactor N θ,K0 = {Pk∈K0
θk(x)pk
x ∈ N ≃ peM θ,K0pe}, where θk are endomorphisms chosen in their class so that
θk(pe) = pk, k ∈ K0. Note that N θ,K0 ⊂ M θ,K0 is an extremal inclusion of II1
dk)2 and relative commutant generated by {pk}k. We
factors with index (Pk∈K0
denote j0 : N = peN pe ≃ N θ,K0 the identification.
Fact 2. With the above notation, assume (Q ⊂ R) ⊂ (N θ,K0 ⊂ M θ,K0) is a non-
degenerate commuting square of factors such that {pk}k ⊂ R and peRpe = Qpe.
Let (R ⊂ R1 ⊂ ...) ⊂ (M θ,K0 ⊂ M1 ⊂ ...) be the associated tower of commuting
squares, obtained by iterating the basic construction. If Q′ ∩ Rn = N θ,K0
∩ Mn,
∀n, and one denotes by Q0 = j−1
0 ⊂ N ∞ = N is faithfully
normalized by G.
0 (Q) ⊂ N , then Q0 = Q∞
′
Indeed, by Fact 1 we know that θk can be taken so that Q = {Pk∈K0
θk(x)pk
x ∈ Q0} and thus so that θk(Q0) ⊂ Q0. After ∞-amplification of the tower of com-
muting squares of factors, the iterated basic construction gives rise to consecutive
products of endomorphisms of N = N ∞, θkn ◦ θkn−1 ◦ ...θk1, for ki ∈ K0, which
all take Q0 = Q∞
0 ⊂ N ∞ = N into itself. The condition on the compatibility of
higher relative commutants amounts to Q0 being faithfully normalized by G.
Now, if G is amenable and finitely generated by its irreducible endomorphisms
indexed by K0 ⊂ K as above, then N θ,K0 ⊂ M θ,K0 has amenable graph, so The-
orem 2.10 applies to provide a strongly smooth non-degenerate commuting square
(Q ⊂ R) ⊂ (N θ,K0 ⊂ M θ,K0), with Q and R hyperfinite II1 factors. By Fact 2, Q
28
SORIN POPA
gives rise to a subfactor Q0 ≃ Q∞ that's faithfully normalized by G and satisfies
the required strong smoothness condition. This proves the statement in the case G
is finitely generated.
To deal with the infinitely generated case, we take the inclusion N θ ⊂ Mθ in Fact
1 with K0 = K = KG and {θk}k∈K = Irr(G). We further embed Mθ = N ⊗B(ℓ2K)
into Mθ
kk′ }k,k′∈K
are the matrix units in the 2nd copy of B(ℓ2K). Note that N ∋ x ∈ j1(j0(x)) ∈ Mθ
1
is given by j1(j0(x)) = Pk,k′∈K θk(θk′(x))e1
1 := N ⊗B(ℓ2K)⊗B(ℓ2K) by x 7→ j1(x) = Pk θk(x)e1
kk, where θk = θk ⊗ 1.
kk, where {e1
Note that Fact 1 already implies that if (Q ⊂ R) ⊂ (N θ ⊂ Mθ) is a subinclusion
ee, then there are
0 (Q) satisfies θk(Q0) ⊂ Q0, ∀k ∈ K. Then note
of factors with {e0
representants θk such that Q0 = j−1
the following:
∩ Mθ ⊂ Q′ ∩ R and Qe0
k = N θ′
ee = e0
eeRe0
k′k′ e0
kk}′′
Fact 3. Assume the above inclusion (Q ⊂ R) ⊂ (N θ ⊂ Mθ) is so that:
1 = N θ′
∩ Mθ
i = Pk∈Ki
1; (b) there exist finite subsets e ∈ Ki = ¯Ki ր K such that
(a) Q′ ∩ Mθ
if we denote q0
i ) is a non-
degenerate commuting square (with respect to the T r-preserving expectations).
Then Q0 = j−1
0 (Q) ⊂ N is faithfully normalized by G, with its restriction to Q0
strongly smooth.
e0
kk then (Qq0
i ) ⊂ (N θq0
i ⊂ q0
i ⊂ q0
i Rq0
i Rq0
The proof of this fact is identical to the proof of 2.11.5◦ above and we leave it
as an exercise.
θk(x)pk; as well as the embedding M θ,Kn ֒→jn
From this point on, the proof of Theorem 2.12 is similar to the proof of Theorem
2.1. Thus, we let e ∈ Kn = ¯Kn ր K be finite sets, denote Dn = Pk∈Kn
dk,
let {pk}k∈K ⊂ N be mutually orthogonal projections of trace T r(pk) = dk such
that 1 − Pk∈K is an infinite projection. By perturbing if necessary each θk by an
inner automorphism, we may assume θk(pe) = pk, ∀k. We also choose θ′
k to be in
the same class as θk so that θ′
k′ (Pk∈K pk), k′ ∈ K, are mutually orthogonal. For
each n consider the embedding N = peN pe ֒→jn
0 (x) =
:= (M θ,Kn)Dn ≃
Pk∈Kn
N D2
n , by x 7→ jn
Note that these are extremal embeddings of II1 factors with finite index = D2
n
and that jn
is a basic construction, ∀n. Moreover,
for each n > m, the inclusion for m is obtained from the one for n by induc-
tion/reduction. We can thus apply the same iterative construction of subfactors of
finite index N ⊃ P1 ⊃ P2..... as in the proof of Theorem 2.1, so that if one denotes
Q0 = ∪mP ′
, then
(jn
) is a basic
construction of non-degenerate commuting squares that are strongly smooth, ∀n.
0 M θ,Kn := N Dn by x 7→ jn
1 (M θ,Kn) ⊂ M θ,Kn
m ∩ N , Rn
1 (Rn
0 = ∪mjn
0 ) ⊂ Rn
1 (M θ,Kn) ⊂ M θ,Kn
0 (Pm)′ ∩ M θ,n, Rn
0 (Pm))′ ∩ M θ,n
1 (x) = Pk∈Kn
1
pk and pn
1 (Q0)) ⊂ jn
e = Pk∈Kn
0 (N )) ⊂ jn
k , where pn
0 (N )) ⊂ jn
1 = ∪mjn
1 M θ,Kn
1 ) ⊂ (jn
1 (jn
θ′
k(x)pn
k = θ′
k(pn
e ).
1 (jn
1
1
1 (jn
1
1 (jn
VANISHING COHOMOLOGY
29
But then, it is easy to see that after appropriate amplifications, Q0 ⊂ N gives
rise to a II∞ subfactor in Q0 ⊂ N that satisfies the conditions in Fact 3, and is
thus faithfully normalized by G, with the strongly smooth condition satisfied. (cid:3)
2.13. Remarks. 1◦ The first result about normalizing a "large" hyperfinite sub-
factor R modulo inner perturbations was obtained in [P83], in the case Γ = Z, with
R being constructed "by hand", using Rokhlin towers and an iterative procedure.
Shortly after, the question of whether any cocycle action of Z2 on an arbitrary II1
factor can be untwisted was asked in [CJ84]. While we realized at that time that if
a similar normalization result could be proved for Γ = Z2 then the problem would
reduce to the case N = R, where vanishing cohomology was just shown in [Oc85],
we could not extend the arguments in [P83] from Z to Z2, despite much effort.
Several years later in [P89], we were able to solve this problem by using tools from
subfactor theory. However, the argument in [P89] could only cover groups Γ that
have a finite set of generators S ⊂ Γ with respect to which Γ has trivial Poisson
boundary (e.g., groups with polynomial growth, in particular Z2), depending cru-
cially on this condition. In retrospect, it is quite surprising that in fact any outer
action of any amenable group Γ on any II1 factor N normalizes (modulo Inn(N ))
an irreducible hyperfinite subfactor, a property which turns out to characterize the
amenablility of Γ.
2◦ Both in the case of outer action σ of a group Γ on N (or more generally of
an outer action of a category G on N ) there may be an AFD subalgebra R ⊂ N
(resp. R ⊂ N ) that's normalized by Γ (resp. G) but on which the resulting "outer
action" has either more relations among generators (resulting into an outer action
of a quotient of Γ, resp. G), or fewer relations (resulting into an outer action of
some Γ, resp. G, whose quotient is Γ, resp. G). For instance, let Fn y R be a free
action of the free group with n ≥ 2 generators on the hyperfinite II1 factor and
Fn → Γ a surjective map on a (non-free) group Γ with n generators (e.g. Γ = Zn)
and H its kernel. If we let N = R ⋊ H then this gives rise to a free cocycle action
(σ, v) of Γ on N , which normalizes R, but on which it generates a free action of Fn.
Similarly, one can take Γ y N = N0⊗R an action that's trivial on R but free on
N . This normalizes R but it implements on it the action of the trivial group.
3. Non-vanishing cohomology for amplifications of actions
3.1. Definition ([P01a]). Let Γ yσ N be a free action of a group Γ on a type II1
factor N . Let p ∈ N be a projection and for each g ∈ Γ choose a partial isometry
wg ∈ N such that wgw∗
g = p, w∗
g ∈ Aut(pN p)
g (x) = wgσg(x)w∗
g, x ∈ pN p. Then σp is a free cocycle action of Γ on pN p,
by σp
with 2-cocycle vp
gh, g, h ∈ Γ. Moreover, up to cocycle conjugacy
g wg = σg(p) and we = p. Define σp
g,h = wgσg(wh)w∗
30
SORIN POPA
(σp, vp, pN p) only depends on τ (p) = t, thus defining a free cocycle action (σt, vt)
of Γ on N t, called the amplification of σ by t.
If in addition {wg}g ⊂ N satisfy wgσg(wh) = wgh, ∀g, h ∈ Γ, then w is called a
generalized 1-cocycle (of support t) for σ, while if this equality holds modulo scalars
then it is called a weak generalized 1-cocycle for σ (of support t).
Note that the vanishing (resp. weak-vanishing) of the cocycle vt amounts to
being able to chose the partial isometries {wg g ∈ Γ} ⊂ N so that w be a
generalized 1-cocycle (resp. weak generalized 1-cocycle) for σ.
3.2. Theorem. Let Γ be a countable group and Γ yσ R = R⊗Γ
the non-
commutative Bernoulli Γ-action with base R0 ≃ R. Let 0 < t < 1 and denote
by σt the free cocycle action obtained by amplifying σ by t, with its 2-cocycle de-
noted vt. Assume one of the following properties holds true: (a) Γ contains an
infinite subgroup with the relative property (T); (b) Γ contains an infinite subgroup
with non-amenable centralizer. Then the cocycle vt is not weak-vanishing.
0
Proof. Assume by contradiction that the cocycle vt is weak-vanishing.
If we are under assumption (a), i.e., if Γ has an infinite subgroup with the
relative property (T), then (Corollary 4.10 in [P01a]) shows that the support t of
the generalized weak 1-cocycle wg must be 1, a contradiction.
If in turn we are under assumption (b), with H ⊂ Γ an infinite subgroup with
centralizer H ′ = {g ∈ Γ gh = hg, ∀h ∈ H} non amenable, then let (α, β) be
the s-malleable deformation of R⊗R that commutes with the double action σ =
σ ⊗ σ : Γ y R = R⊗R = (R0⊗R0)⊗Γ, as in ([P01a]). Denote M = R ⊗ Γ,
M0 = R ⋊ H, M = R⊗R ⋊σ Γ and M0 = R⊗R ⋊σ H, with M identified as
the subfactor R ⊗ 1 ∨ {Ug}g in M , where {Ug}g ∈ M are the canonical unitaries
implementing σ.
Since H ′ is non-amenable, the action Ad(wgUg) of H ′ on (p ⊗ 1) M0(p ⊗ 1) has
spectral gap relative to (p ⊗ 1)M0(p ⊗ 1). Arguing like in ([P06a]), this implies that
whUh and αs(wh)Uh, h ∈ H, are uniformly k k2-close, for s ∈ R with s small.
Equivalently, the map ξ 7→ (wg ⊗ 1)σg(ξ)αs(w∗
g ⊗ 1) gives a unitary representation
πs of the group H on the Hilbert space (p⊗1)L2(R ⊗R)αs(p⊗1) which for s small
has the vector ξ0 = (p ⊗ 1)αs(p ⊗ 1) almost fixed by πs(h), uniformly in h ∈ H.
We are thus in exactly the same situation as in the case when H ⊂ Γ is rigid.
So the proof of (Corollary 4.10 in [P01a]) applies to conclude that the support t of
the generalized weak 1-cocycle {wh h ∈ H} for σ must necessarily be equal to 1,
a contradiction.
(cid:3)
3.3. Corollary. If Γ is a group that contains an infinite subgroup which either
has relative property (T), or has non-amenable centralizer in Γ, then Γ 6∈ VCw(R).
VANISHING COHOMOLOGY
31
3.4. Theorem. Let Γ be a countable group and Γ yρ L(FΓ) the action imple-
mented by the translation from the left by elements h ∈ Γ on the set {ag}g∈Γ of
generators of the free group FΓ, ρh(ag) = ahg, ∀h, g ∈ Γ. Let 1 > t > 0 and de-
note by ρt the free cocycle action obtained by amplifying ρ by t, with its 2-cocycle
denoted vt. Assume one of the following properties is satisfied: (a) Γ contains an
infinite subgroup with the relative property (T); (b) Γ contains an infinite subgroup
with non-amenable centralizer; (c) Γ contains an infinite amenable group with non-
amenable normalizer. Then the cocycle vt is not weak-vanishing.
Proof. The assumptions (a) and (b) lead to exactly the same argument as in the
proof of Theorem 3.2 above, by using the "free s-malleable deformation" of L(FΓ),
like in the proof of (Theorem 6.1 in [P01a]).
So let us assume we are under the assumption (c) and let H ⊂ Γ be an infinite
amenable subgroup with its normalizer G ⊂ Γ non-amenable. Note that FΓ ⋊ Γ is
naturally isomorphic to Γ ∗ Z. Denote N = L(FΓ), M = N ⋊ Γ = L(FΓ) ⋊ Γ =
L(Γ ∗ Z). Let p ∈ P(N ) be a projection of trace t < 1 and assume (ρp, vp, pN p) has
weak vanishing cohomology.
Thus, if {Ug}g∈Γ ⊂ M denote the canonical unitaries implementing ρ on N =
L(FΓ), then there exist partial isometries wg ∈ N of left support p such that
U ′
g = wgUg ⊂ U(pN p) gives a projective representation of Γ with scalar 2-cocycle
g g ∈ G}′′ gives an embedding of Lµ(H) ⊂ Lµ(G) into pM p.
µ. In particular, {U ′
Note that Lµ(H) is amenable diffuse, Lµ(G) has no amenable direct summand
and Lµ(H) is regular in Lµ(G). Thus, by (Corollary 1.7 in [I13]) it follows that
Lµ(G) ≺M L(Γ).
But Lµ(G) ≺M L(Γ) implies that the "free" malleable deformation (α, β) of M ⊂
M = L(FΓ ∗ FΓ) ⋊ Γ is uniform on {U ′
g g ∈ G} (because L(Γ) is in the fixed point
algebra of the malleable deformation and Lµ(G) is subordinated to L(Γ)). Thus,
for each s ∈ R, the G-representation ξ 7→ (wg ∗ 1)σg(ξ)αs(w∗
g ∗ 1) gives a unitary
representation πs of the group G on the Hilbert space (p ⊗ 1)L2(N ∗ N )αs(p ⊗ 1)
which for s small has the vector ξ0 = (p ∗ 1)αs(p ∗ 1) almost fixed by πs(g),
uniformly in g ∈ G. As in the proof of (Theorem 4.1 in [P01a]) this shows that we
must necessarily have t = 1, a contradiction.
(cid:3)
3.5. Remark. Let (N0, τ0) be an arbitrary tracial von Neumann algebra 6= C
and Γ a group that satisfies one of the conditions (a) or (b) in Theorem 3.2. If
Γ y N = N ⊗Γ
is the Bernoulli action with base N0 and 0 < t < 1, then by
using the s-malleable deformation of N = N ⊗Γ
in ([I06]) in combination with
the argument in the proof of Theorem 3.2, one obtains that the t-amplification
(σt, vt, N t) of σ has non-vanishing cohomology. Similarly, by using the s-malleable
0
0
32
SORIN POPA
deformation in [IPeP05] one can obtain a generalization of Theorem 3.4 for the
free-Bernoulli action Γ yρ N = N ∗Γ
0 .
4. Non-vanishing cohomology for cocycle actions on L(F∞)
4.1. Definition ([CJ84]). Let Γ be an infinite countable group with a set of
generators S ⊂ Γ and assume (Γ, S) 6= (FS, S). Let π : FS → Γ be the unique
group morphism taking the free generators S of FS onto S ⊂ Γ (so our assumption
is equivalent to π not being 1 to 1). Then kerπ ≃ F∞ and FS has infinite conjugacy
classes relative to kerπ. Thus, the inclusion of II1 factors L(kerπ) = N ⊂ M =
L(FS) is irreducible and regular, with NM (N )/U(N ) = Γ. Consequently, N ⊂ M
is a crossed product inclusion of the form L(F∞) ⊂ L(F∞) ⋊(σπ ,vπ) Γ, for some
free cocycle action (σπ, vπ) of Γ on L(F∞), that we'll call the Connes-Jones cocycle
action associated with π, with vπ its 2-cocycle, called the CJ-cocycle.
We summarize here some straightforward consequences of results from ([CJ84],
[P01a], [O03], [P06b], [OP07]), but which are stated in a manner pertaining to our
vanishing cohomology problem:
4.2. Theorem. Let (Γ, S), π : FS → Γ, Γ y(σπ,vπ) L(F∞) be as in 4.1. Assume
Γ satisfies one of the following conditions: (a) it does not have Haagerup property
(e.g., it contains an infinite subset with relative property (T)); (b) it has Cowling-
Haagerup invariant Λ(Γ) larger than 1; (c) it has an infinite subgroup with non-
amenable centralizer; (d) it has an infinite amenable subgroup with non-amenable
normalizer. Then the CJ-cocycle vπ is not weak-vanishing. Thus, if Γ satisfies any
of these properties then Γ 6∈ VCw(L(F∞)).
Proof. If vπ is weak-vanishing, then we can choose representatives {Ug g ∈ Γ} in
NM (N ) such that UgUh = µg,hUgh, ∀g, h ∈ Γ, for some scalar 2-cocycle µ ∈ H 2(Γ).
g ∈ L(FS)op ≃ L(FS) satisfy
Thus, Lµ(Γ) ⊂ M = L(FS). Also, the unitaries U op
UgUh = µg,hUgh, implying that {Ug ⊗ U op
g ≃ L(Γ) embeds into L(FS)⊗L(FS).
g }′′
Since FS has Haagerup property and Λ(FS) = 1 (cf [H79]), the II1 factor L(FS)
has Haagerup property and Λ(L(FS)) = 1, implying that L(FS)⊗L(FS) and all
its von Neumann subalgebras have these properties as well (see e.g.
[P01b] and
[OP07]). In particular, L(Γ) has these properties, thus Γ has Haagerup property
and Λ(Γ) = 1.
If in turn Γ has an infinite subgroup H with non-amenable centralizer H ′, then by
(Remark 3◦ in §4 of [P06b]) Lµ(Γ) would contain a diffuse von Neumann subalgebra
B with non-amenable centralizer. But then B ⊂ M = L(FS) has non-amenable
centralizer, contradicting the solidity of free group factors ([O03]).
If Γ has an infinite amenable subgroup H ⊂ Γ with non-amenable normalizer
G ⊂ Γ, then the normalizer of B = Lµ(H) in M = L(FS) contains all {Ug}g∈G.
VANISHING COHOMOLOGY
33
Since {Ug}′′
group factors ([OP07]).
g ≃ Lµ(G) is non-amenable, this contradicts the strong solidity of free
(cid:3)
4.3. Notations. Given two discrete groups Γ, Λ, one writes Γ ≤W ∗ Λ whenever
L(Γ) can be embedded (tracially) into L(Λ). Denote W∗
leq(Λ) the class of all groups
Γ that can be subordinated this way to Λ, i.e., W∗
leq(Λ) = {Γ Γ ≤W ∗ Λ}).
4.4. Corollary. One has VC ⊂ VC(L(F∞)) ⊂ W∗
leq(F2). Thus, if Γ ∈ VC, then Γ
has Haagerup property, Cowling-Haagerup constant equal to 1, any infinite subgroup
of Γ has amenable centralizer, and any infinite amenable subgroup has amenable
normalizer.
4.5. Remarks 1◦ The above criteria for Γ not to be in W∗
leq(F2) (and thus not in
VC(L(F∞)) ⊃ VC) are in fact not stated in their optimal form. Thus, the results
in [OP07] show that in order for Γ not to be in W∗
leq(F2), it is enough that L(Γ)
is not strongly solid, i.e., that L(Γ) merely has a diffuse amenable von Neumann
subalgebra whose normalizer in L(Γ) generates a non-amenable von Neumann al-
gebra. The list in 4.2 is also not exhaustive. For instance, by [Pe07] it follows
that if Γ ∈ W∗
leq(F2), then L(Γ) needs to be L2-rigid, while a result of Ozawa
(see e.g.
[BrO08]) shows that Γ needs to be exact. One should also note that in
the proof of Theorem 4.2 above we showed that an embedding Lµ(Γ) ֒→ L(F2),
for some µ ∈ H 2(Γ), gives rise to an embedding L(Γ) ֒→ L(F2 × F2), by simply
doubling the canonical unitaries {ug g ∈ Γ} ⊂ Lµ(Γ) ⊂ L(F2), i.e., by taking
L(Γ) ≃ {ug ⊗ uop
g g ∈ Γ}′′ ⊂ L(F2)⊗L(F2) = L(F2 × F2).
2◦ It is reasonable to expect that VC = VC(L(F∞)), and more specifically that
the CJ-cocycles are in some sense the "worse possible", ie., if any such cocycle
vanishes for some group Γ, then Γ ∈ VC. We also believe that VC = VCw.
3◦ Given a group Λ, denote by ME(Λ) the class of groups Γ that are measure
equivalent (ME) to Λ and by MEleq(Λ) the class of groups Γ that have a free m.p.
action which can be realized as a sub equivalence relation of a free ergodic m.p.
Λ-action. It would be interesting to explore the possible correlations between the
classes VC, W∗
leq(F2), MEleq(F2), etc. In this respect, one should point out that
while W∗
leq(F2), MEleq(F2) are obviously "hereditary" classes (i.e., if Γ belongs
to any of them, then all subgroups of Γ belong too), we could not prove such
hereditarity for VC (cf. Remark 1.6.2◦). See also Section 7 in [PeT10] for more
comments on MEleq(F2) and its relations to W∗
leq(F2). One should also note that
MEleq(F2) consists of groups Γ that are ME to either Z = F1, F2, or F∞, i.e.,
MEleq(F2) = ME(Z) ∪ ME(F2) ∪ ME(F∞) (cf. [G04], [Hj04]).
34
SORIN POPA
4◦ We do not know of any examples of groups in VC, W∗
leq(F2) other than
amalgamated free products of amenable groups over finite groups. The approach in
4.2 indicates that these two classes may coincide (perhaps with MEleq(F2) as well).
An intriguing class of groups that are known from [G04] to belong to MEleq(F2)
(in fact, even to ME(F2)), are the free products of finitely many copies of F2 with
amalgamation over the subgroup Z ⊂ F2 generated by the commutator aba−1b−1
(where a, b are the generators of F2). Gaboriau conjectured that in fact any amal-
gamated free product of Fki , 2 ≤ ki ≤ ∞, i ∈ I (with I finite or I = N), over some
Z ֒→ Fki which is maximal abelian in the corresponding Fki, ∀i, is in MEleq(F2).
Thus, according to the above speculations, the groups Fk1 ∗Z Fk2 ∗Z ..., with Z
maximal abelian in each Fki , should belong to VC as well. However, we were not
able to prove this for any such example, except of course the case when any subgroup
Z is freely complemented in Fki . Related to this, we pose here the following
Question: Let Fn yσ N be a free action of a free group of rank n on a II1 factor
N and let W denote the set of all 1-cocycles w for σ (i.e., maps w : Fn → U(N )
satisfying wgσg(wh) = wgh, ∀g, h ∈ Fn). Let Z ⊂ Fn be a maximal abelian subgroup,
generated by some element g ∈ Fn. Is it then true that the set {wg w ∈ W}
coincides with the unitary group U(N )?
Taking into account the way the 1-cocycles w ∈ W are constructed, from n-
tuples of unitaries in N that are taken as perturbations of the canonical unitaries
U1, ..., Un ∈ N ⋊σ Fn that implement σa1 , ...., σan (where a1, ..., an are the generators
of Fn), it immediately follows that this question has an affirmative answer in the
case Z is freely complemented in Fn. It is also trivial to see that if the statement
holds true, then all of the above Gaboriau groups Γ = Fk1 ∗Z Fk2 ∗Z ... lie in VC.
It is not known whether these groups are in W∗
leq(F2) either. In fact, deciding
that a group Γ satisfies L(Γ) ֒→ L(F2) is at least as interesting as deciding that it
has property VC. So the fact that VC ⊂ W∗
leq(F2) gives another strong motivation
for proving the universal vanishing cohomology property for various groups.
The above question can also be stated for free measure preserving actions on the
probability measure space, Fn y (X, µ), by simply replacing N by A = L∞(X, µ)
throughout that statement. Besides answering this question, it would be interesting
to know if an affirmative answer would imply Gaboriau's conjecture that the groups
Fk1 ∗Z Fk2 ∗Z ... belong to MEleq(F2).
5◦ It has been conjectured by Peterson and Thom (see end of Sec. 7 in [PeT10])
that if two amenable von Neumann subalgebras B1, B2 of the free group factor
L(F2) have diffuse intersection, then B1 ∨ B2 should follow amenable. There has
been an accumulation of evidence towards this fact being true (e.g., [P81], [Ju06],
[Pe07]). For us here, this would imply that if Γ ∈ W∗
leq(F2) is generated by amenable
VANISHING COHOMOLOGY
35
subgroups Γ1, Γ2 ⊂ Γ with H = Γ1 ∩ Γ2 infinite, then Γ must be amenable. Thus,
if Γ = Γ1 ∗H Γ2 then Γ 6∈ W∗
leq(F2), unless either H is finite, or [Γ1 : H] ≤ 2, [Γ2 :
H] ≤ 2. In particular, VC(L(F∞)) should not contain such groups either.
6◦ We expect that VC(R) is equal to VC(L(F∞)). This fact suggests various new
statements in deformation-rigidity for factors arising from Bernoulli actions. For
instance, it should be possible to prove that if a group Γ does not have Haagerup
property, or if it has an infinite amenable subgroup with non-amenable normalizer,
then some of the W∗-rigidity results in [P01a], [P03], [P06a] should be true. Com-
bining this conjecture with the Peterson-Thom conjecture and remark 5◦ above,
this also suggests that VC(R) doesn't contain any non-amenable group Γ that can
be generated by amenable subgroups Γ1, Γ2 with H = Γ1 ∩ Γ2 infinite (in particular
Γ = Γ1 ∗H Γ2). But the obstruction in this case should be of a completely different
nature. The II1 factors arising from Bernoulli actions of such groups (and more
generally from non-amenable groups Γ generated by n ≥ 2 amenable groups with
infinite intersection) may actually have additional W∗-rigidity properties, providing
a new class of factors on which deformation-rigidity techniques should be tested, in
the spirit of ([P01a], [P03], [P06a], [IPeP05], [PV12]).
7◦ Note that the W ∗-algebra version of von Neumann's conjecture on whether
any non-amenable group Λ contains a copy of F2, amounts to whether for any non-
amenable Λ one has F2 ≤W∗ Λ. Note also that by the Gaboriau-Lyons result in
[GL07] one indeed has F2 ≤W∗ Z ≀ Λ for any non-amenable Λ, while by [OP07] it
follows that if Z ≀ Λ ≤W∗ F2 then Λ must be amenable.
5. A related characterization of amenability
We prove in this section that the "normalization" property for cocycle Γ-actions
in Theorem 2.1 can only be true when the group Γ is amenable. More precisely, for
any non-amenable Γ we exhibit examples of embeddings Γ ⊂ Out(L(F∞)) which
admit no lifting to Aut(L(F∞)) that normalizes a hyperfinite subfactor of L(F∞).
5.1. Theorem. Let Γ be a countable group and Γ yσ L(FΓ) the action imple-
mented by the translation from the left by elements h ∈ Γ on the set {ag}g∈Γ of
generators of the free group FΓ. Let M = L(FΓ) ⋊ Γ with {Ug g ∈ Γ} ⊂ M the
canonical unitaries implementing σ. The following conditions are equivalent:
(a) Γ is amenable.
(b) L(FΓ) contains a hyperfinite subfactor R with R′ ∩ M = C for which there
g = wgUg, g ∈ Γ, normalize R,
exist {wg}g∈Γ ⊂ U(L(FΓ)) with the property that U ′
implement a free Γ-action on it and satisfy U ′
gh, ∀g, h ∈ Γ.
gU ′
h = U ′
36
SORIN POPA
(c) L(FΓ) contains a diffuse AFD von Neumann subalgebra B for which there
exist {wg}g∈Γ ⊂ U(L(FΓ)) with the property that Ad(wg) ◦ σg normalize B, ∀g.
Proof. By Theorem 2.2 we have (a) ⇒ (b), while (b) ⇒ (c) is trivial.
To see that (c) ⇒ (a), note first that one has a natural identification between
M = L(FΓ) ⋊σ Γ and L(Γ) ∗ L(Z). Then note that by [OP07], the von Neumann
algebra B1 generated by the normalizer of B in L(FΓ) then B1 is amenable, thus
AFD by [C75]. As in part (b), denote U ′
g = wgUg ∈ M , g ∈ Γ. Note that
{U ′
g}g normalize B1 and implement a cocycle action of Γ on B1, i.e., one has
U ′
gh(U ′
We first show that Γ non-amenable implies P = B1 ∨ {U ′
g g ∈ Γ} is non-
amenable. We will prove this by contradiction, by showing that if P is amenable,
then Γ follows amenable.
h)−1 ∈ B1, ∀g, h ∈ Γ.
gU ′
geBU ′
g
To see this, let P ⊂ hP, eB1i be the basic construction for B1 ⊂ P and note that
∗ ∈ hP, eBi, g ∈ Γ, are mutually orthogonal projections summing up
fg := U ′
to 1 and generating an atomic abelian von Neumann subalgebra A ⊂ hP, eB1i natu-
rally isomorphic to ℓ∞Γ. Moreover, {Ad(U ′
g)}g normalizes A ≃ ℓ∞Γ implementing
on it the Γ-action by left translation. If P is amenable then there exists a state ϕ on
B(L2P ) that has P in its centralizer. Then ψ = ϕA is a state on A ≃ ℓ∞Γ that's
invariant to {AdU ′
g}g, i.e., to left translations by Γ, showing that Γ is amenable.
Now, if P ⊂ M ≃ L(Γ) ∗ L(Z) is non-amenable, then by (Corollary 1.7 in [I13]) it
follows that P ≺M L(Γ), in particular B ≺M L(Γ). But by applying (Corollary 2.3
in [P03]), it is trivial to see that L(FΓ) (an algebra on which {Ug}g∈Γ acts) has no
diffuse von Neumann subalgebra that can be subordinated ≺M to L(Γ) = {Ug}′′
g ,
a contradiction.
(cid:3)
5.2. Theorem. Let Γ be a countable group with a set of generators S ⊂ Γ and
corresponding surjective morphism π : FS → Γ, with kernel kerπ ≃ F∞. Let
L(F∞) = N ⊂ M = L(FS) be the associated irreducible, regular inclusion of free
group factors, which satisfies NM (N )/U(N ) ≃ Γ, and denote by Γ yσπ N the
corresponding cocycle Γ-action. The following conditions are equivalent:
(a) Γ is amenable.
(b) N contains a hyperfinite subfactor R ⊂ N with R′ ∩ M = C for which
g = wgUg, g ∈ Γ, normalize R,
gh, ∀g, h ∈ Γ.
there exist {wg}g ⊂ U(N ) with the property that U ′
h = U ′
implement a free action on it, and satisfy U ′
gU ′
(c) N contains a diffuse AFD von Neumann subalgebra B such that ∀g ∈ Γ,
∃wg ∈ U(N ) with the property that Adwg ◦ σg normalizes B.
Proof. Theorem 2.2 shows that (a) ⇒ (b) and (b) ⇒ (c) is trivial. If (c) holds but
we assume Γ is non-amenable, then by (Theorem 3.2.4 in [P86]; see also the direct
VANISHING COHOMOLOGY
37
argument in the proof of Theorem 5.2 above) the von Neumann algebra generated
by B and its normalizer in M = L(FS) is non-amenable, contradicting the strong
solidity of the free group factors ([OP07]).
(cid:3)
5.3. Remark. The dichotomy amenable/non-amenable in the above results can
probably be extended to cover the converse to Theorem 2.10 as well. Thus, it
should be true that if G is a non-amenable standard λ-lattice, then there exists an
extremal inclusion of separable II1 factors N ⊂ M with standard invariant equal
to G in which one cannot embed with non-degenerate strongly smooth commuting
squares any inclusion of hyperfinite II1 factors Q ⊂ R (as before, strongly smooth
commuting square inclusion (Q ⊂ R) ⊂ (N ⊂ M ) means that it is non-degenerate
and satisfies Q′ ∩ Rn = Q′ ∩ Mn = N ′ ∩ Mn, R′ ∩ Rn = R′ ∩ Mn = M ′ ∩ Mn, ∀n).
Similarly, it should be true that if G is a non-amenable countable rigid C∗-tensor
category, then G admits an action (by endomorhisms) on a II∞ factor M ∞ which has
no hyperfinite II∞ subfactors R∞ with normal expectation that are left invariant
by G (modulo inner perturbations) and on which G acts outerly as a C∗-tensor
category.
In both statements, the obvious candidate for a proof is the canonical inclusion
N G(L(F∞)) ⊂ M G(L(F∞)), from ([P94a]), which has L(F∞) as "initial data" and
G as standard invariant. Note that the resulting factors N = N G(L(F∞)), M =
M G(L(F∞)) were in fact shown to be isomorphic to L(F∞) in ([PS01]).
If one
assumes by contradiction that there does exist a hyperfinite inclusion Q ⊂ R with
G as standard invariant and which can be embedded with strongly smooth com-
muting square into N ⊂ M and one takes the associated SE inclusions, then de-
formation/rigidity arguments in the style of the proofs of 5.1 and 5.2 above should
contradict the non-amenability of G. A study case is when G is the Temperley-Lieb-
Jones λ-lattice Gλ of index λ−1 > 4. One difficulty in proving such a result is that
so far (relative) strong solidity results can only say something about normalizers
of diffuse AFD von Neumann subalgebras, while in the case of an acting standard
λ-lattice G (or of an acting rigid C∗-tensor category) one generally has to deal with
quasi-normalizers (see [BHV15] for related results).
6. Vanishing cohomology and Connes Embedding conjecture
In this section, we'll show that Connes Approximate Embedding (CAE) conjec-
ture for factors of the form R ⋊ Γ can be reformulated as a vanishing cohomology
problem for a certain cocycle action of Γ. Thus, let ω be an (arbitrary) non-principal
ultrafilter on N and denote by Rω the corresponding ω-ultrapower II1 factor, with
R ⊂ Rω viewed as constant sequences.
6.1. Proposition. 1◦ Rω := R′ ∩ Rω is a II1 factor whose centralizer in Rω is
38
SORIN POPA
equal to R, i.e., R′
ω ∩ Rω = R.
2◦ Given any θ ∈ Aut(R) there exists a unitary element Uθ ∈ NRω (R) such that
θ)R = θ, then
θ = vUθ = Uθv′ for some v, v′ ∈ U(Rω). Moreover, if U ∈ NRω (R) and one
Ad(Uθ)R = θ. If U ′
U ′
denotes θ = Ad(U )R ∈ Aut(R), then U ∈ U(Rω)Uθ.
θ ∈ NRω (R) is another unitary satisfying Ad(U ′
3◦ If θ, Uθ are as in 2◦ above, then Ad(Uθ)Rω implements an element θω ∈
Out(Rω) and an element θω = Ad(Uθ)R∨Rω ∈ Out(R ∨ Rω), with θ ∈ Aut(R)
outer iff θω outer and iff θω outer.
4◦ The application Out(R) ∋ θ 7→ θω ∈ Out(R ∨ Rω) is a 1 to 1 group morphism
whose image has trivial scalar 3-cocycle, with corresponding cocycle crossed product
II1 factor (R ∨ Rω) ⋊ Out(R) equal to the von Neumann algebra generated in Rω
by R ∨ Rω and {Uθ θ ∈ Aut(R)} (thus equal to NRω (R)′′ as well).
5◦ Any free action Γ yσ R gives rise to a free cocycle action σω of Γ on R∨Rω, by
ω : Γ × Γ → U(Rω).
σω(g) = Ad(Uσ(g))R∨Rω , g ∈ Γ, with corresponding 2-cocycle vσ
Proof. 1◦ This is a particular case of (Theorem 2.1 in [P13]).
2◦ This is well known (see e.g., [C74]) and is due to the fact that any automor-
phism of R is approximately inner.
3◦ Since Uθ normalizes R, it also normalizes its relative commutant R′∩Rω = Rω,
and therefore R ∨ Rω as well. If the automorphism θω it implements on Rω is inner,
say implemented by some v ∈ U(Rω), then v∗Uθ ∈ R′
ω ∩ Rω = R, implying that
Ad(Uθ) is inner on R, i.e., θ is inner. Similarly, if Ad(Uθ) is inner on R, then it
is inner on Rω. Since R ∨ Rω ≃ R⊗Rω with Ad(Uθ) splitting as a tensor product
of its restrictions to R, Rω, one also has that this automorphism is inner iff both
restrictions are inner.
4◦ The II1 factor R∨Rω has trivial relative commutant in Rω and so if we denote
by N the unitaries in its normalizer that leave R (and thus also Rω) invariant, then
G = N /U(R)U(Rω) is a discrete group implementing a cocycle action on R ∨ Rω,
with (R ∨ Rω) ∨ N ≃ (R ∨ Rω) ⋊ G. Also, from the construction of the map
of Aut(R) ∋ θ 7→ Uθ ∈ N and part 3◦, we see that this map implements an
isomorphism Out(R) ≃ G.
5◦ This part is trivial from 3◦ above.
(cid:3)
6.2. Definition. A II1 factor M (respectively a group Γ) has the CAE property if
it can be embedded into Rω (respectively into the unitary group of Rω). Note that
by a result in [R06], Γ has a faithful representation into U(Rω) iff Rω contains a
copy of the left regular representation of Γ, equivalently L(Γ) ֒→ Rω. Thus, Γ has
the CAE property iff L(Γ) has the CAE property.
VANISHING COHOMOLOGY
39
6.3. Theorem. Let Γ yσ R be a free action of a countable group Γ on the
hyperfinite II1 factor R. The II1 factor R ⋊σ Γ has the CAE property if and only if
the U(Rω)-valued 2-cocycle vσ
ω vanishes, i.e., iff there exist unitary elements {Ug
g ∈ Γ} ⊂ NRω (R) that implement σ on R and satisfy UgUh = Ugh, ∀g, h ∈ Γ.
Proof. Let M = R ⋊σ Γ with {Ug g ∈ Γ} denoting the canonical unitaries
If M is embeddable into Rω, then by using the fact that any
implementing σ.
two copies of the hyperfinite II1 factor in Rω are conjugated by a unitary element
in Rω, it follows that we may assume the hyperfinite subfactor R in M = R ⋊ Γ
coincides with the algebra of constant sequences in Rω, with the action σ on it
being implemented by {Ug}g ⊂ M ⊂ Rω. By Proposition 6.1.5◦ above, this means
the 2-cocycle vσ
ω vanishes.
Conversely, if vσ
ω vanishes, then we clearly have R ⋊σ Γ ֒→ Rω.
(cid:3)
6.4. Corollary. Let Γ be a countable group and Γ yσ R⊗Γ the non-commutative
Bernoulli Γ-action with base R. Let H be an ICC amenable group (such as the group
S∞ of finitely supported permutations of N, or the lamp-lighter group Z/2Z ≀ Z).
Then H ≀ Γ is a CAE group iff vσ
ω vanishes.
6.5. Remarks. 1◦ It has been shown in [HaS16] that if two groups H, Γ are sofic,
then their wreath product H ≀ Γ is sofic as well, so in particular it is CAE. Taking
H to be an (arbitrary) amenable ICC group H, for which by Connes Theorem one
has L(H) ≃ R, it follows that the crossed product II1 factor R ⋊σ Γ = L(H ≀ Γ) is
CAE, where Γ yσ R⊗Γ ≃ R is the non-commutative Bernoulli Γ-action with base
≃ R as in 6.4. Thus, the corresponding cocycle vσ
ω vanishes. Equivalently, one can
choose Uσ(g) ∈ NRω (R) so that to be a representation of Γ. One can in fact show
that these unitaries can be taken so that to also normalize the ultrapower of the
Cartan subalgebra, i.e., {Ug}g ⊂ NRω (R) ∩ NRω (Dω), where D = D⊗Γ
, D0 being
the Cartan subalgebra of the base.
0
2◦ Given any Γ ∈ VC, the wreath product group S∞ ≀ Γ is CAE by Corollary 6.4
above, and thus Γ is a CAE group. However, one already knows this, since we have
seen in Section 4 that VC is contained in W∗
leq(F2), and L(F2) ֒→ Rω. But while
the class VC has a lot of restrictions on it (cf. Sections 3 and 4 in this paper), the
class of CAE groups is manifestly huge, in fact it may well be that all groups are
CAE.
3◦ Part 4◦ of Proposition 6.1 above naturally leads to the following:
def
= θω of Out(R) on R ∨ Rω have
6.6. Problem. Does the cocycle action θ 7→ π(θ)
vanishing 2-cohomology? Can π be perturbed by inner automorphisms to a genuine
action? Is it true that H 2(Out(R)) = 1 ?
40
SORIN POPA
Due to its "huge size" and properties (e.g., all torsion free elements are conju-
gate in Out(R), by [C74]), one should have H 3(Out(R)) = 1. If this is the case,
then perturbing the cocycle action π by inner automorphisms to a genuine action
would be equivalent to untwisting its 2-cocyle. If the CAE conjecture turns out to
hold true, then Theorem 6.3 implies that the restriction of π above to any count-
able subgroup Γ ⊂ Out(R) that implements a genuine action on R has vanishing
cohomology.
Nevertheless, even if this is the case, the entire cocyle action π cannot probably
be untwisted to a genuine action. One way to prove this would be to show that its
restriction to a certain countable subgroup cannot be untwisted. This amounts to
saying that some free cocycle action (θ, vθ) of a countable group Γ on R cannot be
untwisted when viewed as an action on R ∨ Rω. So to start with, this means (θ, vθ)
cannot be untwisted as a cocycle action on R.
But the only exemples of cocycle actions (θ, vθ) on R that we know to be "un-
twistable" are the ones provided by Theorem 3.2, which are amplifications (σt, vt) of
the Bernoulli Γ-actions Γ yσ R⊗Γ (as defined in [P01a]). But if π can be untwisted
on σ(Γ), then all these cocycle actions can actually be untwisted on R∨Rω. Indeed,
this is because any countable subgroup in the normalizer of R ∨ Rω in Rω has a
"huge" non-separable type II1 fixed point algebra. As noted in [P01a], if a genuine
action has II1 fixed point algebra, then all its amplifications can be untwisted.
Thus, if π can be untwisted when restricted to σ(Γ) ⊂ Out(R) then it can also be
untwisted when restricted to σt(Γ) ⊂ Out(R).
All this shows that the Problem 6.6. may lead to some interesting logic-related
considerations, especially when taken together with the CAE conjecture.
References
[AP17] C. Anantharaman, S. Popa: "An introduction to II1 factors",
www.math.ucla.edu/∼popa/Books/
[Bi97] D. Bisch: Bimodules, higher relative commutants and the fusion algebra associ-
ated to a subfactor, The Fields Institute for Research in Mathematical Sciences
Communications Series, Vol. 13 (1997), 13-63.
[BoHV15] R. Boutonnet, C. Houdayer, S. Vaes: Strong solidity of free Araki-Woods factors,
to appear in Amer. J. Math, arXiv:1512.04820
[BrO08] N. Brown, N. Ozawa: "C∗-algebras and finite dimensional-approximations",
Amer. Math. Soc. Grad. Studies in Math. 88, 2008.
[C74] A. Connes: Outer conjugacy classes of automorphisms of factors, Ann. Ecole
Norm. Sup., 8 (1975), 383-419.
[C75] A. Connes: Classification of injective factors, Ann. of Math., 104 (1976), 73-115.
VANISHING COHOMOLOGY
41
[C80] A. Connes: Correspondences, handwritten notes, 1980.
[CJ84] A. Connes, V.F.R. Jones: Property T for von Neumann algebras, Bull. London
Math. Soc. 17 (1985), 57-62.
[CT76] A. Connes, M. Takesaki: The flow of weights of factors of type III, Tohoku Math.
Journ. 29 (1977), 473-575.
[D92] K. Dykema: Interpolated free group factors, Pac. J. Math. 163 (1994), 123-135.
[G04] D. Gaboriau: Examples of groups that are measure equivalent to the free group,
Ergodic Theory Dynam. Systems 25 (2005), 1809-1827.
[GL07] D. Gaboriau, R. Lyons: A Measurable-Group-Theoretic Solution to von Neu-
mann's Problem, Invent. Math., 177 (2009), 533-540.
[H79] U. Haagerup: An example of a non-nuclear C∗-algebra which has the metric
approximation property, Invent. Math. 50 (1979), 279-293.
[HaS16] B. Hayes, A. Sale: The wreath product of two sofic groups is sofic, arXiv:1601.
03286
[Hj04] G. Hjorth: A lemma for cost attained, Ann. Pure Applied Logic bf 143 (2006),
87-102.
[I06] A. Ioana: Rigidity results for wreath product II1 factors, J. Funct. Anal. 252
(2007), 763-791.
[I13] A. Ioana: Cartan subalgebras of amalgamated free product II1 factors. (Appendix
by A. Ioana and S. Vaes), Ann. Sci. Ec. Norm. Super. 48 (2015), 71-130.
[IPeP05] A. Ioana, J. Peterson, S. Popa: Amalgamated Free Products of w-Rigid Factors
and Calculation of their Symmetry Groups, Acta Math. 200 (2008), 85-153.
[J80] V. F. R. Jones: Actions of finite groups on the hyperfinite type II1 factor, Mem.
Amer. Math. Soc., 237, 1980.
[J81] V. F. R. Jones: A converse to Ocneanu's theorem, Journal of Operator Theory
10 (1983), 61-64.
[J83] V.F.R. Jones: Index for subfactors, Invent. Math. 72 (1983), 1-25.
[J99] V.F.R. Jones: Planar Algebras, math.OA/9909027
[Ju06] K. Jung: Strongly 1-bounded von Neumann algebras, Geom. Funct. Anal. 17
(2007), 1180-1200.
[KV83] V. A. Kaimanovich, A. M. Vershik: Random Walks on Discrete Groups: Bound-
ary and Entropy, Annals of Probability, 11 (1983), 457-490.
[KT02] Y. Katayama, M. Takesaki: Outer actions of a countable discrete amenable group
on an AFD factor, in "Advances in quantum dynamics" (South Hadley, MA,
2002), 163-171, Contemp. Math., 335, AMS, Providence, RI, 2003.
[NT59] M. Nakamura, Z. Takeda: On the extension of finite factors I, II, Proc. Japan
Acad., 44 (1959), 149-154, 215-220.
[NeTu13] S. Neshveyev, L. Tuset: "Compact Quantum Groups and their Representation
Categories", Cours Sp´ecialis´e, Vol 20, Soci´et´e Math. de France, Paris (2013).
42
SORIN POPA
[Oc85] A. Ocneanu: "Actions of discrete amenable groups on factors", Springer Lecture
Notes No. 1138, Berlin-Heidelberg-New York 1985.
[O03] N. Ozawa: Solid von Neumann algebras, Acta Math. 192 (2004), 111-117.
[OP07] N. Ozawa, S. Popa: On a class of II1 factors with at most one Cartan subalgebra,
Annals of Mathematics 172 (2010), 101-137 (math.OA/0706.3623)
[Pe07] J. Peterson: L2-rigidity in von Neumann algebras, Invent. math. 175 (2009),
417-433.
[PeT10] J. Peterson, A. Thom: Group cocycles and the ring of affiliated operators, Invent.
Math. 185 (2011), 561-592.
[PiP84] M. Pimsner, S. Popa: Entropy and index for subfactors, Ann. Sci. Ec. Norm.
Super. 19 (1986), 57-106.
[PiP88] M. Pimsner, S. Popa: Iterating the basic construction, Trans. AMS, 310 (1988),
127-134.
[P81] S. Popa: Maximal injective subalgebras in factors associated with free groups,
Advances in Math., 50 (1983), 27-48.
[P83] S. Popa, Hyperfinite subalgebras normalized by a given automorphism and related
problems, in "Proceedings of the Conference in Op. Alg. and Erg. Theory"
Busteni 1983, Lect. Notes in Math., Springer-Verlag, 1132, 1984, pp 421-433.
[P86] S. Popa: Correspondences, INCREST Preprint 56/1986, www.math.ucla.edu/
∼popa/preprints.html
[P89] S. Popa: Sous-facteurs, actions des groupes et cohomologie, C.R. Acad. Sci.
Paris 309 (1989), 771-776.
[P91] S. Popa: Classification of amenable subfactors of type II, Acta Mathematica,
172 (1994), 163-255.
[P93] S. Popa: Approximate innerness and central freeness for subfactors: A classifi-
cation result, in "Subfactors", (Proc. Tanegouchi Symposium in Operator Alge-
bras), Araki-Kawahigashi-Kosaki Editors, World Scientific 1994, pp 274-293.
[P94a] S. Popa: An axiomatization of the lattice of higher relative commutants of a
subfactor, Invent. Math., 120 (1995), 427-445.
[P94b] S. Popa, Symmetric enveloping algebras, amenability and AFD properties for
subfactors, Math. Res. Letters, 1 (1994), 409-425.
[P97] S. Popa: Some properties of the symmetric enveloping algebras with applications
to amenability and property T, Documenta Mathematica, 4 (1999), 665-744.
[P01a] S. Popa: Some rigidity results for non-commutative Bernoulli shifts, J. Fnal.
Analysis 230 (2006), 273-328 (MSRI preprint No. 2001-005).
[P01b] S. Popa: On a class of type II1 factors with Betti numbers invariants, Ann. of
Math 163 (2006), 809-899 (math.OA/0209310; MSRI preprint 2001-024).
[P03] S. Popa: Strong Rigidity of
II1 Factors Arising from Malleable Actions of w-
Rigid Groups I, Invent. Math., 165 (2006), 369-408.
VANISHING COHOMOLOGY
43
[P06a] S. Popa: On the superrigidity of malleable actions with spectral gap, J. Amer.
Math. Soc. 21 (2008), 981-1000 (math.GR/0608429).
[P06b] S. Popa: On Ozawa's Property for Free Group Factors, Int. Math. Res. No-
tices (2007) Vol. 2007, article ID rnm036, 10 pages, doi:10.1093/imrn/rnm036
published on June 22, 2007 (math.OA/0608451)
[P13] S. Popa: Independence properties in subalgebras of ultraproduct II1 factors, Jour-
nal of Functional Analysis 266 (2014), 5818-5846 (math.OA/1308.3982)
[PS01] S. Popa, D. Shlyakhtenko: Universal properties of L(F∞) in subfactor theory,
Acta Mathematica, 191 (2003), 225-257.
[PSV15] S. Popa, D. Shlyakhtenko, S. Vaes: Cohomology and L2-Betti numbers for sub-
factors and quasi-regular inclusions, math.OA/1511.07329 , to appear in the
International Mathematical Research Notices.
[PV12] S. Popa, S. Vaes: Unique Cartan decomposition for II1 factors arising from
arbitrary actions of free groups, Acta Mathematica, 194 (2014), 237-284
[PV14] S. Popa, S. Vaes: Representation theory for subfactors, λ-lattices and C∗-tensor
categories, Commun. Math. Phys. 250 (2015), 1239-1280.
[R91] F. Radulescu: The weak closure of the group algebras associated to free groups
are stably isomorphic, Comm. Math. Physics 156 (1993), 17-36.
[R06] F. Radulescu: The von Neumann algebra of the non-residually finite Baumslag
group ha, b ab3a−1 = b2i embeds into Rω. Hot topics in operator theory, 173-
185, Theta Ser. Adv. Math., 9, Theta, Bucharest, 2008.
[Su80] C.E. Sutherland: Cohomology and extensions of von Neumann algebras , Publ.
RIMS, 16 (1980), 135-176.
[T79] M. Takesaki: "Theory of operator algebras" I., Springer-Verlag, New York-
Heidelberg, 1979.
Math.Dept., UCLA, Los Angeles, CA 90095-1555, [email protected]
|
1612.05549 | 1 | 1612 | 2016-12-16T16:47:22 | Construction of a new class of quantum Markov fields | [
"math.OA",
"math.FA",
"math.PR"
] | In the present paper, we propose a new construction of quantum Markov fields on arbitrary connected, infinite, locally finite graphs. The construction is based on a specific tessellation on the considered graph, that allows us to express the Markov property for the local structure of the graph. Our main result concerns the existence and uniqueness of quantum Markov field over such graphs. | math.OA | math |
CONSTRUCTION OF A NEW CLASS OF QUANTUM
MARKOV FIELDS
LUIGI ACCARDI,1 FARRUKH MUKHAMEDOV,2 ∗ and ABDESSATAR SOUISSI3
Abstract. In the present paper, we propose a new construction of quantum
Markov fields on arbitrary connected, infinite, locally finite graphs. The con-
struction is based on a specific tessellation on the considered graph, that allows
us to express the Markov property for the local structure of the graph. Our
main result concerns the existence and uniqueness of quantum Markov field
over such graphs.
1. Introduction
One of the basic open problem in quantum probability is to develop a theory of
quantum Markov fields, which are conventionally quantum Markov processes with
multi-dimensional index set. Here Quantum Markov fields are noncommutative
extensions of the classical Markov fields (see [4, 8, 11]). On the other hand, these
quantum fields can be considered as extensions of quantum Markov chains [1, 7]
to general graphs.
In [3, 10] the first attempts to construct quantum analogues of classical Markov
chains have been carried out. In [3] quantum Markov fields were considered over
integer lattices, unfortunately there was not given any non trivial examples of
such fields. In [5, 6], quantum Markov chains (fields) on the tree like graphs (like
Cayley tree) have been constructed and investigated, but the proposed construc-
tion does not work for general graphs.
A main aim of the present paper is to provide a construction of new class of
quantum Markov fields on arbitrary connected, infinite, locally finite graphs. The
construction is based on a specific tessellation on the considered graphs, it allows
us to express the Markov property for the local structure of the graph. Our main
result is the existence and uniqueness of quantum Markov field over such graphs.
We note that even in the classical case, the proposed construction gives a new
ways to define Markov fields (see [13, 14]).
2. Graphs
Let G = (V, E) be a ( non-oriented simple ) graph, that is, L is a nonempty
set and E is identified as a subset of an ordered pairs of V , i.e.
E ⊂ {{x, y} x, y ∈ E, x 6= y}
Copyright 2016 by the Tusi Mathematical Research Group.
Date: Received: xxxxxx; Revised: yyyyyy; Accepted: zzzzzz.
∗Corresponding author.
2010 Mathematics Subject Classification. Primary 46L53; Secondary 60J99, 46L60, 60G50,
82B10.
Key words and phrases. Quantum Markov field, graph, tessellation, construction.
1
2
L. ACCARDI, F. MUKHAMEDOV and A. SOUISSI
Elements of V and E are called, respectively, vertices and edges. Two vertices
x and y are said to be nearest neighbors if there exist an edge joining them (i.e.
{x, y} ∈ E) and we denote them by x ∼ y. For any vertex y ∈ V we denote its
nearest neighbors by
Ny := {x ∈ V y ∼ x}
(2.1)
Notice that x /∈ Nx. The set {y}∪Ny is said to be interaction domain or plaquette
at y. If for every x ∈ V one has Nx < ∞ then the graph is called locally finite.
An edge path or walk joining two vertices x and y is a finite sequence of edges
x = x0 ∼ x1 ∼ . . . xd−1 ∼ xd = y. In this case d is the length of the edge path.
The graph is said to be connected if every two disjoint vertices can be joined by
an edge path. In the sequel, we assume that the graph G is infinite, connected
and locally finite. Note that in this case the set V is automatically countable.
Now for any nonempty Λ ⊂ V we associate its following parts:
• complement:
• boundary:
• interior :
Λc := V \ Λ
∂Λ := {x ∈ Λ ∃y ∈ Λc;
x ∼ y}
◦
Λ := Λ \ ∂Λ
• external boundary:
~∂Λ := {y ∈ Λc
∃x ∈ Λ;
x ∼ y}
• closure:
Λ := Λ ∪ ~∂Λ
By F we denote a net of all finite subsets of V , i.e.
F := {Λ ⊂ V V < ∞}
where · denotes the cardinality of a set.
(2.2)
(2.3)
(2.4)
(2.5)
(2.6)
(2.7)
3. Tessellations on graphs
In this section we propose a tessellation on the considered graphs, which will
play a key role in the construction. Therefore, the resulting quantum Markov
field will depend also on the tessellation.
Fix a "root" y1 ∈ V and define by induction the following sets:
Having defined V0,n, put
V0,1 := {y1}
Vn := [y∈V0,n
({y} ∪ Ny)
V0,n+1 := V0,n ∪ ~∂Vn
Define the following set of vertices:
V0 := [n≥1
V0,n.
(3.1)
(3.2)
(3.3)
(3.4)
QUANTUM MARKOV FIELDS
3
From now on, elements of V0 will be called vertices, any other element of V
belongs to some plaquette at a certain element of V0. Notice that with in this
construction, for every n, the inner boundary ∂Vn of each Vn contain no vertex:
∂Vn ∩ V0 = ∅
Since V = +∞ and, by assumption, V is connected, one has
Vn+1 ≥ V n + 1 ≥ Vn + 2,
V0,n+1 ≥ V0,n + 1
It follows that, if Λ is any finite set, there exists N ∈ N such that
Λ ⊆ VN
Therefore {Vn} is an exhaustive sequence of finite subsets recovering the all the
vertices set V .
One can check that
and
~∂Vn
V0 := {y1} ∪ [n≥1
V = [y∈V0
{y} ∪ Ny
(3.5)
(3.6)
Remark 3.1.
(i) For each x ∈ V \ V0, there exists at least one y ∈ V0 such
that x belongs to the plaquette at y.
(ii) Each y ∈ V0 belongs to its plaquette (i.e. the plaquette {y} ∪ Ny) but no
other one with center in V0.
The set V0 given by (3.5) (or equivalently the family {V0,n; n = 1, 2, · · · } ) is
called tessellation on the graph G.
4. Quantum Markov Fields
In this section we propose a definition for backward Markov fields, for the same
graph G = (V, E) with the given tessellation {V0,n :
n = 1, 2, · · · }.
The map
(4.1)
defines a bundle on V whose fiber is a finite dimensional Hilbert space Hx. Denote
Ax := B(Hx), x ∈ V . Define for any finite subset Λ ⊂ V the algebra
x ∈ V −→ Hx " state space on x "
then one get on a canonical way, the quasi-local algebra
defined as the closure of the local algebra
AΛ.
where F is given by (2.7).
Ax.
Ax
AΛ :=Ox∈Λ
AV :=Ox∈V
AV,loc := [Λ∈F
(4.2)
(4.3)
(4.4)
4
L. ACCARDI, F. MUKHAMEDOV and A. SOUISSI
Analogously, one can define for any subset Λ′ ⊂ V , the algebra AΛ′ :=Nx∈Λ′ Ax.
Notice that for Λ ⊂ Λ′ ⊂ V one can see AΛ as C ∗-subalgebra of AΛ′ through the
following embedding
AΛ ≡ AΛ ⊗ 1IΛ′\Λ ⊂ AΛ′
(4.5)
Definition 4.1. Consider a triplet C ⊂ B ⊂ A of unital C ∗-algebras. Recall [2]
that a quasi-conditional expectation with respect to the given triplet is a com-
pletely positive (CP), unital linear map E : A → B such that E(ca) = cE(a), for
all a ∈ A, c ∈ C.
We give the definition of general of backward quantum Markov field which is
independent of the tessellation.
Definition 4.2. A state ϕ on AV is said to be backward quantum Markov field
if for any sequence {Λn}∞
n=0 of finite subsets of V satisfying Λn ⊂⊂ Λn+1, there
exists a pair (ϕΛ0, {EΛn,Λn+1}∞
n=0}) with ϕΛ0 is a state on AΛ0 and EΛn,Λn+1 is
a quasi-conditional expectation with respect to the triplet AΛn ⊂ A ¯Λn ⊂ AΛn+1
such that
ϕ = lim
n→∞
ϕΛ0 ◦ EΛ0,Λ1 ◦ · · · ◦ EΛn,Λn+1
(4.6)
where the limit is taken in the weak-*-topology.
Remark 4.3. In Definition 4.2, the condition Λn ⊂⊂ Λn+1 for every n ∈ N implies
that Λn ↑ V and the limit state obtained by the right side of the equation (4.6)
is defined on the full algebra AV ,
If ϕ is a backward quantum Markov field in the sense of Definition 4.2, then it
satisfy Definition 4.2 of [5] for any increasing sequence {Λn}∞
n=0 of finite subsets
of V such that ¯Λn = Λn+1, to get such a sequence of subsets, we consider Λ0 ∈ F ,
and for ∈ n ≥ 1 put
Λn = ¯Λn−1.
Clearly one has Λn ⊂⊂ Λn+1 and Λn ↑ V .
Now we introduce a class of backward quantum Markov field that depends on
the tessellation {V0,n, n = 1, 2, · · · }
Definition 4.4. A state ϕ on AV is said to be backward quantum Markov field
w.r.t. the tessellation {V0,n, n = 1, 2, · · · }, ( or V0-backward quantum Markov
field ) if for any sequence {Λn}∞
n=0 of finite subsets satisfying
Λn ⊂⊂ Λn+1,
(4.7)
there exists a pair (ϕΛ0, {EΛn,Λn+1}∞
n=0}) with ϕΛ0 is a state on AΛ0 and EΛn,Λn+1
is a quasi-conditional expectation with respect to the triplet AΛn ⊂ A ¯Λn ⊂ AΛn+1
such that
~∂Λ ∩ V0 = ∅
where the limit is taken in the weak-*-topology.
ϕ = lim
n→∞
ϕΛ0 ◦ EΛ0,Λ1 ◦ · · · ◦ EΛn,Λn+1
Now we fix the following product state
ϕ0 :=Ox∈V
ϕ0
x
(4.8)
(4.9)
QUANTUM MARKOV FIELDS
5
on the algebra AV , where ϕ0
x is a state on Ax for every x ∈ V . Denote for Λ ⊂ V ,
ϕ0
Λ :=Ox∈Λ
ϕ0
x
(4.10)
which is the restriction of the state ϕ0
V to AΛ.
We aim to construct a quantum Markov field on the algebra AV through a
perturbation of the product state ϕ0
V .
5. Construction of conditional density amplitudes
It is well known [2] that quasi-conditional expectations are more convenient
than Umegaki conditional expectations (see definition (5.1)) to express the non-
commutative Markov property. In what follows, we will perturb ϕ-conditional
expectations (see [2]) to get quasi-conditional expectations using a commuting
set of operators with the considered tessellation.
For any ordered pair y ∈ V0 and x ∈ Ny, let be given an operator
K(x,y) ∈ A{x,y}
such that it is invertible and the C ∗-subalgebra
K = { K ∗
{x,y} , K{x,y} : y ∈ V0, x ∈ Ny}
C ∗
(5.1)
is commutative.
Definition 5.1. A Umegaki conditional expectation is a norm one projection from
a C ∗-algebra onto its C ∗-subalgebra.
Definition 5.2. Let A1, A2 be two C ∗-algebras with units respectively I1 and I2
and let A = A1 ⊗A2. An element K ∈ A is called a conditional density amplitude
w.r.t. a state ϕ on I1 ⊗ A2, if one has
Eϕ(K ∗K) = I1
(5.2)
where Eϕ is the Umegaki conditional expectation from A onto A1 ⊗ I2 defined
by the linear extension of
Eϕ(a1 ⊗ a2) = ϕ(I1 ⊗ a2)a1 ⊗ I2.
(5.3)
An operator K is also called a conditional density amplitude for the ϕ-conditional
expectation Eϕ.
For each x ∈ V by E0
the algebra A onto the algebra A{x}c defined on localized elements a =Nz∈V az =
{x}c we denote the Umegaki conditional expectation from
ax ⊗ a{x}c by:
{x}c(ax ⊗ a{x}c) = ϕ0
E0
x(ax)a{x}c.
One can prove the following facts.
Lemma 5.3. For every pair of vertices (x, y) ∈ V 2 one has
[E0
{x}c, E0
{y}c] = 0
(5.4)
(5.5)
6
L. ACCARDI, F. MUKHAMEDOV and A. SOUISSI
For any Λ ∈ F by virtue of Lemma 5.3 we define
E0
Λc := Yx∈Λ
E0
{x}c.
(5.6)
Lemma 5.4. For any Λ ∈ F the map E0
expectation from AV onto AΛc such that for aΛ ∈ AΛaΛc ∈ AΛc one has
Λc given by (5.6) is a Umegaki conditional
Λc(aΛ ⊗ aΛc) = ϕ0
E0
Λ(aΛ)aΛc
(5.7)
Remark 5.5. The map E0
Λc can be defined, through the equation (5.7), for an
arbitrary part (not necessarily finite) Λ of V and it is still a Umegaki conditional
expectation from AV onto AΛc
Proposition 5.6. Let y ∈ V0, the operator
is invertible.
BNy := E0
{y}c
(cid:12)(cid:12)(cid:12)(cid:12) Yx∈Ny
Proof. Let us consider B{y}∪Ny :=(cid:12)(cid:12)(cid:12)Qx∈Ny
ator B{y}∪Ny is positive definite, then σ(B{y}∪Ny ) ⊂]0, keK{y}∪Ny k]. Since the spec-
∈ A{y}∪Ny and denote its spec-
trum by σ(B{y}∪Ny ), which is a closed subset of the complex field. Since the oper-
2
K{x,y}(cid:12)(cid:12)(cid:12)(cid:12)
∈ ANy
K{x,y}(cid:12)(cid:12)(cid:12)
trum does not contain zero then there is ε > 0 such that σ(B{y}∪Ny ) ⊂ [ε, kBk],
therefore B{y}∪Ny ≥ ε1I. This yields E0
{y}c is
invertible.
(cid:3)
{y}c(B) ≥ ε1I, which means BNy = E0
(5.8)
2
In the sequel, we assume that for every y ∈ V0 the operator BNy belongs to
the commutant K′ (w.r.t. AV ) of the algebra K (see (5.1)). Note that under this
condition the operators B±1/2
also belong to K′.
Ny
Lemma 5.7. The operator
is a ϕ0
{y}-conditional density amplitude in the algebra A{y}∪Ny .
Proof. Using the commutativity of the algebra K we obtain
E0
{y}c(cid:16)K ∗
{y}∪Ny
K{y}∪Ny(cid:17) = E0
K{y}∪Ny :=
Yx∈Ny eK{x,y}
{y}c
Ny
Yx∈Ny
B−1/2
{y}c
Yx∈Ny
)∗E0
Ny
)∗BNy B−1/2
Ny
= (B−1/2
= (B−1/2
= 1I.
Ny
Ny
B−1/2
(5.9)
K{x,y}
Ny
B−1/2
K{x,y}
∗
Yy∈Ny
K{x,y}
B−1/2
K ∗
{x,y}
Ny
This completes the proof.
(cid:3)
QUANTUM MARKOV FIELDS
7
Now, for each Λ ∈ F , we define
~∂0Λ := [y∈∂Λ∩V0
Ny
(5.10)
By construction the family
{K ∗
{y}∪Ny
, K{y}∪Ny
: x ∼ y ∈ V0}
is commutative, therefore the following operator is well defined
KΛ∪~∂0Λ := Yy∈Λ∩V0
K{y}∪Ny ∈ AΛ∪~∂0Λ ⊆ A ¯Λ
for every Λ ∈ F .
(5.11)
Remark 5.8.
1. In general, it is possible that AΛ∪~∂0Λ is a proper sub-algebra
of A ¯Λ. Since, by construction of the tessellation, the set Λ ∪ ~∂0Λ cannot
contain elements of V0.
2. If Λ ∩ V0 = ∅, we convent that KΛ∪~∂0Λ = 1I.
Theorem 5.9. For any Λ ∈ F , the operator KΛ∪~∂0Λ defined by (5.11) is a con-
ditional density amplitude for the Umegaki conditional expectation E0
(Λ∩V0)c.
Proof. Since the family {K{y}∪Ny , K ∗
one can write
{y}∪Ny
K ∗
Λ∪~∂0ΛKΛ∪~∂0Λ = Yy∈Λ∩V0
y ∈ Λ ∩ V0} is commutative, then
K ∗
{y}∪Ny
K{y}∪Ny
(5.12)
and using the following property of the tessellation: for disjoint elements y and z
of V0 the plaquette at y does not contain z, we conclude that K{y}∪Ny is localized
in {z}c. Then by Lemma 5.4 one gets
{z}c(K ∗
E0
{y}∪Ny
K{y}∪Ny ) = K ∗
{y}∪Ny
K{y}∪Ny
then after a small iteration, we obtain
E0
(Λ∩V0)c(cid:16)K ∗
Λ∪~∂0ΛKΛ∪~∂0Λ(cid:17) = Yy∈Λ∩V0
E0
{y}c(K ∗
{y}∪Ny
K{y}∪Ny ).
By Lemma 5.7, one has E0
{y}c (K ∗
{y}∪Ny
K{y}∪Ny ) = 1I, hence one gets
E0
(Λ∩V0)c(cid:16)K ∗
Λ∪~∂0ΛKΛ∪~∂0Λ(cid:17) = 1I.
The following auxiliary results can be easily proved.
Lemma 5.10. For every Λ1 ⊂ Λ2 ⊂ V , one has:
E0
Λ1 ◦ E0
Λ2 = E0
Λ1
Lemma 5.11. For Λ, Λ′ ⊂f in V with ¯Λ ∩ Λ′ = ∅, one has:
K(Λ∪Λ′)∪~∂(Λ∪Λ′) = KΛ∪ ~∂0ΛKΛ′∪~∂0Λ′
Theorem 5.12. For Λ0 ⊆ ¯Λ0 ⊂ Λ, one has:
(cid:3)
(5.13)
(5.14)
8
L. ACCARDI, F. MUKHAMEDOV and A. SOUISSI
(i) For z ∈ V0 ∩ (Λ \ ¯Λ0)
{z}c(K ∗
E0
Λ∪~∂0ΛaΛ0KΛ∪~∂0Λ) = K ∗
(Λ\{z})∪~∂0(Λ\{z})aΛ0K(Λ\{z})∪~∂0(Λ\{z})
for every aΛ0 ∈ AΛ0;
(ii)
(Λ\ ¯Λ0)∩V0(K ∗
E0
Λ∪~∂0ΛaΛ0KΛ∪~∂0Λ) = K ∗
¯Λ0∪~∂0(¯Λ0)aΛ0K ¯Λ0∪~∂0(¯Λ0)
(5.15)
(5.16)
for every aΛ0 ∈ AΛ0.
Proof. (i) For a general Λ0, if z ∈ (Λ\ ¯Λ0)∩V0, then Nz can intersect ~∂Λ0, but not
Λ0. Therefore, K{z}∪Nz and aΛ0 are localized on disjoint parts, so they commute.
Due to the commutativity of {K{y}∪Ny , K ∗
y ∈ Λ ∩ V0} it follows from
(5.14) that
{y}∪Ny
E0
{z}c(K ∗
Λ∪~∂0ΛaΛ0KΛ∪~∂0Λ)
K ∗
{y}∪Ny
aΛ0 Yy∈Λ∩V0
K{y}∪Ny )
= E0
= E0
= E0
{z}c( Yy∈Λ∩V0
{z}c(cid:16)(K ∗
{z}c(cid:0)K ∗
{z}∪Nz
and by Lemma 5.7 one has
{z}∪Nz
K{y}∪Ny ) × (K ∗
(Λ\{z})∪~∂0(Λ\{z})aΛ0K(Λ\{z})∪~∂0(Λ\{z}))(cid:17)
(Λ\{z})∪~∂0(Λ\{z})aΛ0K(Λ\{z})∪~∂0(Λ\{z})
K{y}∪Ny(cid:1) K ∗
{z}c(cid:0)K ∗
E0
{z}∪Nz
K{y}∪Ny(cid:1) = 1I
Hence, we get
{z}c(K ∗
E0
Λ∪~∂0ΛaΛ0KΛ∪~∂0Λ) = K ∗
(Λ\{z})∪~∂0(Λ\{z})aΛ0K(Λ\{z})∪~∂0(Λ\{z})
(ii) Iterating the procedure of (5.15) to cover all z ∈ (Λ \ ¯Λ0) ∩ V0 one finds
E0
(Λ\Λ0)∩V0(cid:16)K ∗
Λ∪~∂0ΛaΛ0KΛ∪~∂0Λ(cid:17) = (cid:18) Yz∈(Λ\ ¯Λ0)∩V0
E0
{z}c(cid:19)(cid:16)K ∗
Λ∪~∂0ΛaΛ0KΛ∪~∂0Λ(cid:17)
= K ∗
¯Λ0∪~∂0(¯Λ0)aΛ0K ¯Λ0∪~∂0(¯Λ0)
This completes the proof.
(cid:3)
Remark 5.13. Keeping the notations of Theorem 5.12, if ~∂Λ0 ∩ V0 = ∅ then using
the same argument, one gets
E0
Λ\Λ0(K ∗
Λ∪~∂0ΛaΛ0KΛ∪~∂0Λ) = K ∗
Λ0∪~∂0Λ0
aΛ0KΛ0∪~∂0Λ0
for every a0 ∈ AΛ0.
6. Main result
In this section, we prove a main result of the paper. First we need an auxiliary
result.
QUANTUM MARKOV FIELDS
9
Proposition 6.1. Let Λ1, Λ2 ∈ F with Λ1 ⊂⊂ Λ2. Define
EΛ1,Λ2(a) = E0
(Λ2\ ¯Λ1)c(cid:16)K ∗
(Λ2\ ¯Λ1)∪~∂0(Λ2\ ¯Λ1)aK(Λ2\ ¯Λ1)∪~∂0(Λ2\ ¯Λ1)(cid:17)
for a ∈ AV . Then EΛ1,Λ2 is a quasi-conditional expectation w.r.t.
AΛ1 ⊂ A ¯Λ1 ⊂ AΛ2.
(6.1)
the triplet
Proof. The map EΛ1,Λ2 is clearly linear and valued in A ¯Λ1.
Unitality: using commutativity of the family {E{z}c z ∈ (Λ2 \ Λ1) ∩ V0} (by
Lemma 5.3), one can write
0 )c ◦ E0
and using Theorem 5.9 for Λ = Λ2 \ ¯Λ1 we obtain
(Λ2\Λ1)c = E0
E0
((Λ2\ ¯Λ1)∩V c
((Λ2\Λ1)∩V0)c
then using (6.2) one finds
E0
(Λ2\ ¯Λ1)∪~∂0(Λ2\ ¯Λ1)K(Λ2\ ¯Λ1)∪~∂0(Λ2\ ¯Λ1)(cid:17) = 1I
((Λ2\ ¯Λ1)∩V0)c(cid:16)K ∗
(Λ2\ ¯Λ1)∪~∂0(Λ2\ ¯Λ1)K(Λ2\ ¯Λ1)∪~∂0(Λ2\ ¯Λ1)(cid:17) = E0
((Λ2\ ¯Λ1)∩V c
E0
(Λ2\ ¯Λ1)c(cid:16)K ∗
0 )c(1I) = 1I
(6.2)
hence,
Complete positivity: One can check that for any y ∈ V0 the map
EΛ1,Λ2(1I) = 1I.
a 7→ E{y}c(a) := E0
{y}c (K ∗
{y}∪Ny
aK{y}∪Ny )
is completely positive. Now using the commutativity of the set {K{y}∪Ny , y ∈ V0}
one gets
EΛ1,Λ2 = E0
(Λ2\ ¯Λ1)∩V0 ◦(cid:18) Yy∈(Λ2\ ¯Λ1)∩V0
E{y}c(cid:19).
Hence EΛ1,Λ2 is the composition of completely positive maps, therefore, it is so.
∈ A¯(Λ2\Λ1) then it commutes
Let a ∈ AΛ2, c ∈ AΛ1, while K ∗
(Λ2\Λ1)∪~∂0(Λ2\Λ1)
with c, then using the fact that
(Λ2\Λ1)c(cd) = cE0
E0
(Λ2\Λ1)c(d)
for every d ∈ A, one gets
EΛ1,Λ2(ca) = E0
(Λ2\ ¯Λ1)c(cid:16)K ∗
(Λ2\ ¯Λ1)c(cid:16)cK ∗
(Λ2\ ¯Λ1)c(cid:16)K ∗
(Λ2\ ¯Λ1)∪~∂0(Λ2\ ¯Λ1)caK(Λ2\ ¯Λ1)∪~∂0(Λ2\ ¯Λ1)(cid:17)
(Λ2\ ¯Λ1)∪~∂0(Λ2\ ¯Λ1)aK(Λ2\ ¯Λ1)∪~∂0(Λ2\ ¯Λ1)(cid:17)
(Λ2\ ¯Λ1)∪~∂0(Λ2\ ¯Λ1)aK(Λ2\ ¯Λ1)∪~∂0(Λ2\ ¯Λ1)(cid:17)
= E0
= cE0
= cEΛ1,Λ2(a).
Hence, EΛ1,Λ2 is a quasi-conditional expectation w.r.t. the given triplet. This
completes the proof.
(cid:3)
Now we pass to our main result.
10
L. ACCARDI, F. MUKHAMEDOV and A. SOUISSI
Theorem 6.2. For each Λ ∈ F define the state eϕΛ on A by
Then the net {eϕΛ}Λ∈F converges in the weak-*-topology, moreover the limiting
state ϕ is a backward Markov field on AV w.r.t. the tessellation V0.
eϕΛ(a) := ϕ0(K ∗
Λ∪~∂0ΛaKΛ∪~∂0Λ)
Proof. First we prove the existence of the limit. Due to the density argument, it
is sufficient to establish the existence of the limit in the local algebra AV,loc.
(6.3)
Let a ∈ AV,loc then a ∈ AΛ0 for some Λ0 ∈ F . For Λ ∈ F with Λ0 ⊂⊂ Λ, we
and by Theorem 5.12 one gets
= ϕ0 ◦ E
eϕΛ(a) = ϕ0(cid:16)K ∗
(Λ\ ¯Λ0)c(cid:16)K ∗
Λ∪~∂0ΛaKΛ∪~∂0Λ(cid:17)
(Λ\ ¯Λ0)c(cid:16)K ∗
Λ∪~∂0ΛaKΛ∪~∂0Λ(cid:17) = K ∗
E
Λ∪~∂0ΛaKΛ∪~∂0Λ(cid:17)
¯Λ0∪~∂0 ¯Λ0
aK ¯Λ0∪~∂0 ¯Λ0,
have
so
¯Λ0∪~∂0 ¯Λ0
aK ¯Λ0∪~∂0 ¯Λ0
)
eϕΛ(a) = ϕ0(K ∗
= eϕ ¯Λ0(a).
As Λ → V , we find that Λ0 ⊂⊂ Λ up to some order, hence the net {eϕ(a)}Λ∈F ;Λ0⊂⊂Λ
is stationary. This means that
(6.4)
lim
Λ→V ;Λ0⊂⊂ΛeϕΛ(a) = ϕ ¯Λ0(a) =: ϕ(a)
therefor the limit exist on the local algebra, and yet it exists on the full algebra
AV .
Now we establish that the state ϕ is a quantum Markov field.
Let {Λn n ∈ N}n∈N be a family of subset of F satisfying
Λn ⊂⊂ Λn+1,
~∂Λn ∩ V0 = ∅
Let EΛn,Λn+1 be given by (6.1). Then, for a ∈ AVn, we have
eϕΛn ◦ EΛn,Λn+1(a)
Λn∪~∂0Λn
= ϕ0(cid:16)K ∗
= ϕ0(cid:16)K ∗
EΛn,Λn+1(a)KΛn∪~∂0Λn(cid:17)
(Λn+1\ ¯Λn)c(cid:16)K ∗
E0
Λn∪~∂0Λn
(Λn+1\ ¯Λn)∪~∂0(Λn+1\ ¯Λn)aK(Λn+1\ ¯Λn)∪~∂0(Λn+1\ ¯Λn)(cid:17) KΛn∪~∂0Λn(cid:17)
Since KΛn∪~∂0Λn
expectation from AV onto A(Λn+1\ ¯Λn)c then one finds
∈ A ¯Λn ⊂ A(Λn+1\ ¯Λn)c and E0
(Λn+1\ ¯Λn)c is a Umegaki conditional
eϕΛn ◦ EΛn,Λn+1(a)
(Λn+1\ ¯Λn)c(cid:16)K ∗
= ϕ0E0
Λn∪~∂0Λn
K ∗
(Λn+1\ ¯Λn)∪~∂0(Λn+1\ ¯Λn)aK(Λn+1\ ¯Λn)∪~∂0(Λn+1\ ¯Λn)KΛn∪~∂0Λn(cid:17)
and by the assumption (4.7) one has
~∂Λn ∩ V0 = ∅
QUANTUM MARKOV FIELDS
11
then ¯Λn ∩ V0 = Λn ∩ V0 and
KΛn∪~∂0Λn
= Yy∈Λn∩V0
K{y}∪Ny = Yy∈ ¯Λn∩V0
K{y}∪Ny = K ¯Λn∪~∂0 ¯Λn
.
From Lemma 5.14 it follows that
KΛn+1∪~∂0Λn+1
= K ¯Λn∪ ¯Λ1K(Λn+1\ ¯Λn)∪~∂0(Λn+1\ ¯Λn)
then we obtain
eϕΛn ◦ EΛn,Λn+1(a) = ϕ0 ◦ E0
Hence, by construction one gets
ϕ0
V = ϕ0
V ◦ E
(Λn+1\ ¯Λn)c
(Λn+1\ ¯Λn)c(K ∗
Λn+1∪~∂0Λn+1
aKΛn+1∪~∂0Λn+1
)
so
eϕΛn ◦ EΛn,Λn+1(a) = ϕ0
Now iterating the equation (6.5), we obtain
V (K ∗
Λn+1∪~∂0Λn+1
aKΛn+1∪~∂0Λn+1
) = eϕΛn+1(a)
(6.5)
eϕn = eϕΛ0 ◦ EΛ0,Λ1 ◦ · · · ◦ EΛn−1,Λn
ϕV = lim ϕΛ0 ◦ EΛ0,Λ1 ◦ · · · ◦ EΛn−1,Λn
. This completes the proof.
(cid:3)
therefore
where ϕΛ0 = eϕΛ0⌈AΛ0
The provided construction allows us to produce a lot of interesting examples
of quantum Markov fields on arbitrary connected, infinite, locally finite graphs.
Note that the construction is based on a specific tessellation on the considered
graph, it allows us to express the Markov property for the local structure of the
graph. We note that even in the classical case, the proposed construction gives
other ways to define Markov fields different to the existing ones (see [13]). This
construction opens new perspectives in the theory of phase transitions in the
scheme of quantum Markov fields (comp. [9]).
References
1. L. Accardi, On the noncommutative Markov property, Funct. Anal. Appl. 9 (1975) 1 -- 8.
2. L. Accardi, C. Cecchini, Conditional expectations in von Neumann algebras and a Theorem
of Takesaki, J. Funct. Anal. 45 (1982), 245 -- 273.
3. L. Accardi, F. Fidaleo, Quantum Markov fields, Infin. Dim. Analysis, Quantum Probab.
Related Topics 6 (2003) 123 -- 138.
4. R.L. Dobrushin, Description of Gibbsian Random Fields by means of conditional probabil-
ities, Probab. Theory and Appl. 13(1968) 201 -- 229.
5. L. Accardi, H. Ohno, F. Mukhamedov, Quantum Markov fields on graphs, Infin. Dim.
Analysis, Quantum Probab. Related Topics 13(2010), 165 -- 189.
6. M. Fannes, B. Nachtergaele, R.F. Werner, Ground states of VBS models on Cayley trees,
J. Stat. Phys. 66 (1992) 939 -- 973.
7. M. Fannes, B. Nachtergaele, R.F. Werner, Finitely correlated states on quantum spin chains,
Commun. Math. Phys. 144 (1992) 443 -- 490.
8. H.-O. Georgi, Gibbs measures and phase transitions, de Gruyter Studies in Mathematics
vol. 9, Walter de Gruyter, Berlin, 1988.
12
L. ACCARDI, F. MUKHAMEDOV and A. SOUISSI
9. F. Mukhamedov, A. Barhoumi, A. Souissi, Phase transitions for Quantum Markov Chains
associated with Ising type models on a Cayley tree, J. Stat. Phys. 163 (2016) 544 -- 567.
10. V. Liebscher, Markovianity of quantum random fields, Proceedings Burg Conference 15 -- 20
March 2001, W. Freudenberg (ed.), World Scientific, QP -- PQ Series 15 (2003) 151 -- 159
11. C. Preston, Gibbs states on countable sets, Cambridge University Press, London, 1974.
12. U.A. Rozikov, Gibbs measures on Cayley trees, World Scientific, Singappore, 2013.
13. A. Spataru, Construction of a Markov field on an infinite tree, Advance in Math. 81(1990),
105 -- 116.
14. S. Zachary, Countable state space Markov random fields and Markov chains on trees, Ann.
Prob. 11 (1983) 894 -- 903.
1 Centro Vito Volterra, Universita di Roma "Tor Vergata", Roma I-00133,
Italy
E-mail address: [email protected]
2Department of Mathematical Sciences, College of Science, The United Arab
Emirates University, P.O. Box 15551, Al Ain, Abu Dhabi, UAE
E-mail address: [email protected]
3.1 Department of Mathematics, Faculty of Sciences of Tunis, University of
Tunis El-Manar, 1060 Tunis, Tunisia.
3.2Preparatory Institute for Scientific and Technical Studies, La Marsa, Carthage
University, Tunisia0
E-mail address: [email protected]; [email protected]
|
1706.08833 | 2 | 1706 | 2018-05-04T12:46:25 | Quantum Symmetries of Graph C*-Algebras | [
"math.OA",
"math.FA"
] | The study of graph C*-algebras has a long history in operator algebras. Surprisingly, their quantum symmetries have never been computed so far. We close this gap by proving that the quantum automorphism group of a finite, directed graph without multiple edges acts maximally on the corresponding graph C*-algebra. This shows that the quantum symmetry of a graph coincides with the quantum symmetry of the graph C*-algebra. In our result, we use the definition of quantum automorphism groups of graphs as given by Banica in 2005. Note that Bichon gave a different definition in 2003; our action is inspired from his work. We review and compare these two definitions and we give a complete table of quantum automorphism groups (with respect to either of the two definitions) for undirected graphs on four vertices. | math.OA | math | QUANTUM SYMMETRIES OF GRAPH C ∗-ALGEBRAS
SIMON SCHMIDT AND MORITZ WEBER
Abstract. The study of graph C ∗-algebras has a long history in operator alge-
bras. Surprisingly, their quantum symmetries have never been computed so far.
We close this gap by proving that the quantum automorphism group of a finite,
directed graph without multiple edges acts maximally on the corresponding graph
C ∗-algebra. This shows that the quantum symmetry of a graph coincides with the
quantum symmetry of the graph C ∗-algebra. In our result, we use the definition
of quantum automorphism groups of graphs as given by Banica in 2005. Note
that Bichon gave a different definition in 2003; our action is inspired from his
work. We review and compare these two definitions and we give a complete table
of quantum automorphism groups (with respect to either of the two definitions)
for undirected graphs on four vertices.
8
1
0
2
y
a
M
4
]
.
A
O
h
t
a
m
[
2
v
3
3
8
8
0
.
6
0
7
1
:
v
i
X
r
a
Introduction
Symmetry constitutes one of the most important properties of a graph.
It is
captured by its automorphism group
Aut(Γ) := {σ ∈ Sn σε = εσ} ⊆ Sn,
where Γ = (V, E) is a finite graph with n vertices and no multiple edges, ε ∈
Mn({0, 1}) is its adjacency matrix, and Sn is the symmetric group.
In modern
mathematics, notably in operator algebras, symmetries are no longer described only
by groups, but by quantum groups. In 2005, Banica [1] gave a definition of a quan-
tum automorphism group of a finite graph within Woronowicz's theory of compact
matrix quantum groups [19]. In our notation, G+
aut(Γ) is based on the C ∗-algebra
C(G+
aut(Γ)) := C(S+
= C ∗(uij, i, j = 1, . . . , n uij = u∗
n )/huε = εui
ij = u2
ij,Xl
uil = 1 =Xl
ulj, RBan hold),
Date: May 7, 2018.
2010 Mathematics Subject Classification. 46LXX (Primary); 20B25, 05CXX (Secondary).
Key words and phrases. finite graphs, graph automorphisms, automorphism groups, quantum
automorphisms, graph C ∗-algebras, quantum groups, quantum symmetries.
The second author was partially funded by the ERC Advanced Grant NCDFP, held by Roland
Speicher. This work was part of the first author's Master's thesis. This work was also supported
by the DFG project Quantenautomorphismen von Graphen.
1
2
SIMON SCHMIDT AND MORITZ WEBER
where S+
n is Wang's quantum symmetric group [17] and RBan are the relations
Xk
uikεkj =Xk
εikukj.
Earlier, in 2003, Bichon [5] defined a quantum automorphism group G∗
aut(Γ) via
C(G∗
aut(Γ))
:= C ∗(uij, i, j = 1, . . . , n uij = u∗
ij = u2
ij,Xl
uil = 1 =Xl
ulj, RBic hold),
where RBic are the relations
Xk
uikεkj =Xk
εikukj,
us(e)s(f )ur(e)r(f ) = ur(e)r(f )us(e)s(f ) for e, f ∈ E,
and r : E → V and s : E → V are range and source maps respectively. We
immediately see that
Aut(Γ) ⊆ G∗
aut(Γ) ⊆ G+
aut(Γ)
holds, in the sense that there are surjective ∗-homomorphisms:
aut(Γ))
C(G+
uij
→
7→
aut(Γ))
C(G∗
uij
→
7→
C(Aut(Γ))
(σ 7→ σij)
Relatively little is known about these two quantum automorphism groups of graphs
and we refer to Section 3.4 for an overview on all published articles in this area.
Graph C ∗-algebras in turn are well-established objects in operator algebras. They
emerged from Cuntz and Krieger's work [8] in the 1980's and they developed to
be one of the most important classes of examples of C ∗-algebras, see for instance
Raeburn's book for an overview [14]. Given a finite graph Γ = (V, E) the associated
graph C ∗-algebra C ∗(Γ) is defined as
C ∗(Γ) := C ∗(pv, v ∈ V, se, e ∈ E pv = p∗
v = p2
v, pvpw = 0 for v 6= w,
s∗
ese = pr(e), Xe∈E
s(e)=v
ses∗
e = pv, if s−1(v) 6= ∅).
A natural question is then: What is the quantum symmetry group of the graph
C ∗-algebra and is it one of the above two quantum automorphism groups of the
underlying graphs? The answer is: It is given by the one defined by Banica. Note
however, that Bichon's definition has its justification in other contexts such as in
[4, 6] or in the recent work by Speicher and the second author [15]. Moreover,
Bichon's work [5] inspired us how to formulate our main theorem, see also Remark
4.2.
QUANTUM SYMMETRIES OF GRAPH C ∗-ALGEBRAS
3
1. Main result
Intuitively speaking, our main result is that the quantum symmetry of a finite
graph without multiple edges coincides with the quantum symmetry of the associ-
ated graph C ∗-algebra. In other words, the following diagram is commutative:
finite graphs
Γ7→C ∗(Γ)
/ graph C ∗-algebras
*❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯
aut(Γ)
Γ7→G+
s❤❤❤❤❤❤❤❤❤❤❤❤❤❤❤❤❤❤❤
C ∗(Γ)7→QSym(C ∗(Γ))
quantum symmetry groups
More precisely, we have the following result.
Main Theorem. Let Γ be a finite graph with n vertices V = {1, ..., n} and m edges
E = {e1, ..., em} having no multiple edges. The maps
α : C ∗(Γ) → C(G+
aut(Γ)) ⊗ C ∗(Γ),
pi 7→
sej 7→
n
m
Xk=1
Xl=1
uik ⊗ pk,
1 ≤ i ≤ n,
us(ej)s(el)ur(ej)r(el) ⊗ sel,
1 ≤ j ≤ m,
and
β : C ∗(Γ) → C(G+
aut(Γ)) ⊗ C ∗(Γ),
pi 7→
sej 7→
n
m
Xk=1
Xl=1
uki ⊗ pk,
1 ≤ i ≤ n,
us(el)s(ej )ur(el)r(ej ) ⊗ sel,
1 ≤ j ≤ m
define a left and a right action of G+
aut(Γ) on C ∗(Γ), respectively. Moreover, when-
ever G is a compact matrix quantum group acting on C ∗(Γ) in the above way, we
have G ⊆ G+
aut(Γ) of Γ is
the quantum symmetry group of C ∗(Γ), see also Remark 4.1.
aut(Γ). In this sense, the quantum automorphism group G+
We also provide some tools for comparing and dealing with the two definitions of
quantum automorphism groups of graphs, G+
aut(Γ), notably depending
on the complement Γc of Γ, see Section 3.5. Moreover, we provide a list of all Aut(Γ),
G+
aut(Γ) for undirected graphs Γ on four vertices, having no multiple
edges and no loops, see Section 3.6.
aut(Γ) and G∗
aut(Γ) and G∗
*
/
s
4
SIMON SCHMIDT AND MORITZ WEBER
2. Preliminaries
2.1. Graphs. We fix some notations for graphs used throughout this article. A
graph Γ = (V, E) is finite, if the set V of vertices and the set E of edges are finite.
We denote by r : E → V the range map and by s : E → V the source map. A graph
is undirected if for every e ∈ E there is a f ∈ E with s(f ) = r(e) and r(f ) = s(e);
it is directed otherwise. Elements e ∈ E with s(e) = r(e) are called loops. A graph
without multiple edges is a directed graph, where there are no e, f ∈ E, e 6= f , such
that s(e) = s(f ) and r(e) = r(f ). For a finite graph Γ = (V, E) with V = {1, . . . , n},
its adjacency matrix ε ∈ Mn(N0) is defined as εij := #{e ∈ E s(e) = i, r(e) = j}.
Here N0 = {0, 1, 2, . . .}. Throughout this article we restrict to finite graphs
having no multiple edges.
If Γ = (V, E) is a directed graph without multiple edges, we denote by Γc = (V, E′)
the complement of Γ, where E′ = (V × V )\E. Within the category of graphs having
no loops, the complement Γc is defined using E′ = (V × V )\(E ∪ {(i, i); i ∈ V }).
2.2. Automorphism groups of graphs. For a finite graph Γ = (V, E) without
multiple edges, a graph automorphism is a bijective map σ : V → V such that
(σ(i), σ(j)) ∈ E if and only if (i, j) ∈ E. In other words, εσ(i)σ(j) = 1 if and only if
εij = 1. The set of all graph automorphisms of Γ forms a group, the automorphism
group Aut(Γ). We can view Aut(Γ) as a subgroup of the symmetric group Sn, if Γ
has n vertices:
Aut(Γ) = {σ ∈ Sn σε = εσ} ⊆ Sn
2.3. Graph C ∗-algebras. The theory of Graph C ∗-algebras has its roots in Cuntz
and Krieger's work [8] in 1980. Nowadays, it forms a well-developed and very active
part of the theory of C ∗-algebras, see [14] for an overview or [9] for recent devel-
opments. For a finite, directed graph Γ = (V, E) without multiple edges, the graph
C ∗-algebra C ∗(Γ) is the universal C ∗-algebra generated by mutually orthogonal pro-
jections pv, v ∈ V and partial isometries se, e ∈ E such that
(i) s∗
ese = pr(e) for all e ∈ E
(ii) and pv =Pe∈E : s(e)=v ses∗
It follows immediately, that s∗
e for every v ∈ V with s−1(v) 6= ∅.
esf = 0 for e 6= f andPv∈V pv = 1 hold true in C ∗(Γ).
2.4. Compact matrix quantum groups. Compact matrix quantum groups were
defined by Woronowicz [18, 19] in 1987. They form a special class of compact
quantum groups, see [12, 16] for recent books. A compact matrix quantum group G
is a pair (C(G), u), where C(G) is a unital (not necessarily commutative) C ∗-algebra
which is generated by uij, 1 ≤ i, j ≤ n, the entries of a matrix u ∈ Mn(C(G)).
k=1 uik ⊗ ukj
Moreover, the *-homomorphism ∆ : C(G) → C(G) ⊗ C(G), uij 7→ Pn
must exist, and u and its transpose ut must be invertible.
QUANTUM SYMMETRIES OF GRAPH C ∗-ALGEBRAS
5
Example 2.1. As an example,
S+
n = (C(S+
group given by
quantum symmetric group
n ), u) as defined by Wang [17] in 1998. It is the compact matrix quantum
consider
the
n
n
C(S+
n ) := C ∗(uij uij = u∗
ij = u2
ij,
uil = 1 =
uli for all 1 ≤ i, j ≤ n).
Xl=1
Xl=1
One can show that the quotient of C(S+
n ) by the relations that all uij commute is
exactly C(Sn). Moreover, the symmetric group Sn may be viewed as a compact
matrix quantum group Sn = (C(Sn), u), where uij : Sn → C are the evaluation
maps of the matrix entries. This justifies the name "quantum symmetric group".
If G = (C(G), u) and H = (C(H), v) are compact matrix quantum groups with
u ∈ Mn(C(G)) and v ∈ Mn(C(H)), we say that G is a compact matrix quantum
subgroup of H, if there is a surjective *-homomorphism from C(H) to C(G) mapping
generators to generators. In this case we write G ⊆ H. As an example: Sn ⊆ S+
n .
The compact matrix quantum groups G and H are equal as compact matrix quantum
groups, writing G = H, if we have G ⊆ H and H ⊆ G.
2.5. Actions of quantum groups. Let G = (C(G), u) be a compact matrix
quantum group and let B be a C ∗-algebra. A left action of G on B is a unital
*-homomorphism α : B → C(G) ⊗ B such that
(i) (∆ ⊗ id) ◦ α = (id ⊗ α) ◦ α
(ii) and α(B)(C(G) ⊗ 1) is linearly dense in C(G) ⊗ B.
A right action is a unital *-homomorphism β : B → C(G) ⊗ B with
(i) ((F ◦ ∆) ⊗ id)) ◦ β = (id ⊗ β) ◦ β
(ii) and β(B)(C(G) ⊗ 1) is linearly dense in C(G) ⊗ B,
where F is the flip map F : C(G) ⊗ C(G) → C(G) ⊗ C(G), a ⊗ b 7→ b ⊗ a. Note
that in some articles (for instance in [17]), the property (ii) is replaced by
(ii') (ε ⊗ id) ◦ α = id
(iii') and there is a dense *-subalgebra of B, such that α restricts to a right
coaction of the Hopf *-algebra on the *-subalgebra.
One can show that (ii') and (iii') are equivalent to (ii), see [13].
2.6. Quantum symmetry group of n points. According to Wang's work [17],
we know that S+
n (from Example 2.1) is the quantum symmetry group of n points
in the sense that
(i) S+
n acts from left and right on
i = p2
C ∗(p1, . . . , pn pi = p∗
i ,Xl
k=1 uik ⊗ pk and β(pi) :=Pn
pl = 1)
k=1 uki ⊗ pk
by α(pi) :=Pn
6
SIMON SCHMIDT AND MORITZ WEBER
(ii) and S+
n is maximal with these actions, i.e. any other compact matrix quan-
tum group with actions defined as α and β is a compact matrix quantum
subgroup of S+
n .
See also [11] for similar questions around quantum symmetries.
3. Quantum automorphism groups of graphs
Wang's work in the 1990's was the starting point of the investigations of quan-
tum symmetry phenomena for discrete structures (within Woronowicz's framework).
Note that n points may be viewed as the totally disconnected graph on n vertices. A
decade later, Banica and Bichon extended Wang's approach to a theory of quantum
automorphism groups of finite graphs.
In the sequel, we restrict to finite graphs
having no multiple edges.
3.1. Bichon's quantum automorphism group of a graph. In 2003, Bichon [5]
defined a quantum automorphism group as follows.
Definition 3.1. Let Γ = (V, E) be a finite graph with n vertices V = {1, ..., n}
and m edges E = {e1, ..., em}. The quantum automorphism group G∗
aut(Γ) is the
compact matrix quantum group (C(G∗
aut(Γ)) is the universal
C ∗-algebra with generators uij, 1 ≤ i, j ≤ n and relations
aut(Γ)), u), where C(G∗
uij = u∗
ij,
n
uijuik = δjkuij,
ujiuki = δjkuji,
1 ≤ i, j, k ≤ n,
n
uil = 1 =
uli,
1 ≤ i ≤ n,
(3.1)
(3.2)
(3.3)
(3.4)
(3.5)
Xl=1
Xl=1
us(ej)iur(ej)k = ur(ej)kus(ej)i = 0,
uis(ej)ukr(ej) = ukr(ej)uis(ej) = 0,
us(ej)s(el)ur(ej)r(el) = ur(ej)r(el)us(ej)s(el),
ej ∈ E, (i, k) /∈ E,
ej ∈ E, (i, k) /∈ E,
ej, el ∈ E.
In the original definition of Bichon, there is actually another relation which is
implied by the others:
(3.6)
m
Xl=1
us(el)s(ej )ur(el)r(ej ) = 1 =
m
Xl=1
us(ej)s(el)ur(ej)r(el),
ej ∈ E
Indeed, Relations (3.6) are implied by Relations (3.2), (3.3) and (3.4):
us(el)s(ej )ur(el)r(ej ) =
m
Xl=1
n
Xi,k=1
uis(ej)ukr(ej) = n
Xi=1
uis(ej)! n
Xk=1
ukr(ej)! = 1
3.2. Banica's quantum automorphism group of a graph. Two years later,
Banica [1] gave the following definition.
QUANTUM SYMMETRIES OF GRAPH C ∗-ALGEBRAS
7
Definition 3.2. Let Γ = (V, E) be a finite graph with n vertices and adjacency ma-
trix ε ∈ Mn({0, 1}). The quantum automorphism group G+
aut(Γ) is the compact ma-
trix quantum group (C(G+
aut(Γ)) is the universal C ∗-algebra
with generators uij, 1 ≤ i, j ≤ n and Relations (3.1), (3.2) together with
aut(Γ)), u), where C(G+
(3.7)
uε = εu,
which is nothing but Pk uikεkj =Pk εikukj.
3.3. Link between the two definitions. It is easy to see ([10, Lemma 3.1.1] or
[15, Lemma 6.7]) that Banica's definition may be expressed as:
C(G+
aut(Γ)) = C ∗(uij Relations (3.1) – (3.4))
We thus have
Aut(Γ) ⊆ G∗
aut(Γ) ⊆ G+
aut(Γ)
in the sense of compact matrix quantum subgroups, see Section 2.4. Equality holds,
if C(G∗
aut(Γ)) are commutative. Moreover, note that (see Example
2.1):
aut(Γ)) and C(G+
C(S+
n ) = C ∗(uij Relations (3.1) and (3.2))
Example 3.3. As an example, let Γ be the complete graph (i.e. E = V ×V ). Then:
Aut(Γ) = G∗
aut(Γ) = Sn,
G+
aut(Γ) = S+
n
For its complement Γc (i.e. E = ∅), we have:
Aut(Γc) = Sn,
G∗
aut(Γc) = G+
aut(Γc) = S+
n
3.4. Review of the literature on quantum automorphism groups of graphs.
At the moment there are only few articles regarding quantum automorphism groups
of graphs. Some results are the following. In [6], Bichon defined the hyperoctahedral
quantum group and showed that this group is the quantum automorphism group
of some graph. Banica computed the Poincar´e series of G+
aut(Γ) for homogenous
graphs with less than eight vertices in [1]. Banica, Bichon and Chenevier considered
circulant graphs having p vertices for p prime in [3]. They proved G+
aut(Γ) = Aut(Γ)
if the graph Γ does fulfill certain properties. Banica and Bichon investigated G+
aut(Γ)
for vertex-transitive graphs of order less or equal to eleven in [2]. They also computed
G+
aut(Γ) for the direct product, the Cartesian product and the lexicographic product
of specific graphs. Chassaniol also studied the lexicographic product of graphs in
[7]. In her PhD thesis [10], Fulton studied undirected trees Γ such that Aut(Γ) =
Z2 × Z2 × ... × Z2, where we have k kopies of the cyclic group Z2 = Z/2Z. She
proved Aut(Γ) = G∗
aut(Γ) for
k ≥ 2. See also [4] for links to quantum isometry groups.
aut(Γ) for k = 1 and Aut(Γ) 6= G∗
aut(Γ) = G+
aut(Γ) = G+
8
SIMON SCHMIDT AND MORITZ WEBER
3.5. Comparing with the complement of the graph. As can be seen from Sec-
tion 3.4, the theory of quantum automorphism groups of graphs is still in its infancy.
We now provide some basic results on the link between G∗
aut(Γc). Note
that while we have
aut(Γ) and G∗
Aut(Γ) = Aut(Γc)
and
G+
aut(Γ) = G+
aut(Γc)
for all graphs Γ (using εΓc = A − εΓ for the adjacency matrices, with A ∈ Mn({1})
the matrix filled with units, and uA = A = Au by Relation (3.2)), we may have
for instance when Γ is the complete graph, see Example 3.3.
G∗
aut(Γ) 6= G∗
aut(Γc),
Lemma 3.4. If G∗
aut(Γ) ⊆ G∗
aut(Γc), then G∗
aut(Γ) = Aut(Γ).
Proof. Relation (3.5) in C(G∗
C(G∗
(3.4) and (3.5) in C(G∗
implies that uik and ujl commute in
aut(Γ)) whenever (i, j) /∈ E and (k, l) /∈ E. Together with Relations (3.3),
(cid:3)
aut(Γ)) this yields commutativity of all generators.
aut(Γc))
Lemma 3.5. If G∗
aut(Γc), then G∗
aut(Γ) = Aut(Γ).
aut(Γc) = G+
aut(Γ) ⊆ G+
Proof. We have G∗
aut(Γc) and apply Lemma 3.4. (cid:3)
The next lemma shows that the quantum automorphism groups of a graph without
aut(Γc) = G∗
aut(Γ) = G+
loops does not change if we add those.
Lemma 3.6. Let Γ = (V, E) be a finite graph without loops. Consider Γ′ = (V, E′)
with E′ = E ∪ {(i, i), i ∈ V }. It holds
(i) G+
(ii) G∗
aut(Γ) = G+
aut(Γ) = G∗
aut(Γ′),
aut(Γ′).
Proof. For (i), we use εΓ′ = 1 + εΓ, where 1 is the identity matrix in Mn({0, 1}).
Thus, uεΓ = εΓu is equivalent to uεΓ′ = εΓ′u.
For (ii), all we need to check is that uis(ej)uir(ej) = uir(ej)uis(ej) is fulfilled in
(cid:3)
aut(Γ)) for all i ∈ V , ej ∈ E, which is true due to Relation (3.1).
C(G∗
3.6. Quantum automorphism groups on four vertices. For a small number
of vertices of undirected graphs, a complete classification of G∗
aut(Γ) is
possible. For n ∈ {1, 2, 3}, we have C(S+
aut(Γ) =
G+
aut(Γ). For n = 4, we now provide a complete table for graphs having no loops.
We restrict to undirected graphs in order to keep it simple. We need the following
lemma to compute the quantum automorphism groups.
aut(Γ) and G+
n ) = C(Sn), hence Aut(Γ) = G∗
Lemma 3.7. Let Γ = (V, E) be a finite graph with V = {1, ..., n} and let ej ∈ E.
Let q ∈ V with s−1(q) = ∅. For the generators of C(G+
aut(Γ)) it holds
uqs(ej) = 0 = us(ej)q.
QUANTUM SYMMETRIES OF GRAPH C ∗-ALGEBRAS
Proof. By Relations (3.2) and (3.4), we get
uqs(ej) = uqs(ej) n
Xi=1
uir(ej)! =
n
Xi=1
uqs(ej)uir(ej) = 0,
because (q, i) /∈ E for all i ∈ V . Likewise, we get us(ej)q = 0.
In the following, D4 denotes the dihedral group defined as
9
(cid:3)
D4 := hx, y x2 = y2 = (xy)4 = ei,
H +
2 denotes the hyperoctahedral quantum group defined by Bichon in [6] and Z2
denotes the cyclic group Z/2Z. The quantum group \Z2 ∗ Z2 = (C ∗(Z2 ∗ Z2), u) is
understood as the compact matrix quantum group with matrix
p
1 − p
1 − p
0
0
p
0
0
0
0
q
0
0
1 − q
1 − q
q
where C ∗(Z2 ∗ Z2) is seen as the universal unital C ∗-algebra generated by two pro-
jections p and q. Recall that Aut(Γ) = Aut(Γc) and G+
aut(Γc), where Γc
is the complement of Γ within the category of graphs having no loops. Parts of the
following table were also computed in [2] and [6].
aut(Γ) = G+
Theorem 3.8. Let Γ be an undirected graph on four vertices having no loops and
no multiple edges. Then:
Γ
•
•
•
•
•
•
•
•
•
•
•
•
•
•
•
•
•
•
•
•
•
•
•
•
Γc
•
•
❅❅
•
•
•
•
❅❅
•
•
•
•
❅❅
•
•
•
•
•
•
•
•
❅❅
•
•
•
•
•
•
(1)
(2)
(3)
(4)
(5)
(6)
Aut(Γ)
G∗
aut(Γc) G∗
aut(Γ) G+
aut(Γ)
S4
S4
S+
4
S+
4
Z2 × Z2
Z2 × Z2
\Z2 ∗ Z2
\Z2 ∗ Z2
Z2
D4
S3
Z2
Z2
D4
S3
Z2
Z2
H +
2
S3
Z2
Z2
H +
2
S3
Z2
Proof. For every row of the table, we compute G+
G∗
graphs as follows:
aut(Γ). We then obtain G∗
aut(Γ) =
aut(Γc) by using Lemma 3.5. We label the points of the
aut(Γ) and we show G+
1
3
•
•
•
•
2
4
10
SIMON SCHMIDT AND MORITZ WEBER
(1) Obvious, see Example 3.3.
(2) Let (uij)1≤i,j≤4 be the generators of C(G+
aut(Γ)). Lemma 3.7 yields
u31 = u32 = u41 = u42 = u13 = u23 = u14 = u24 = 0.
With Relations (3.2) we deduce
u11
1 − u11
1 − u11
0
0
u11
0
0
0
0
0
0
u33
1 − u33
1 − u33
u33
.
u =
Thus
G+
aut(Γ) = \Z2 ∗ Z2.
Since uijukl = ukluij holds
aut(Γ) = G+
C(G+
aut(Γ)), we get G∗
for
aut(Γ).
(i, k), (j, l) ∈ {(1, 2), (2, 1)} in
(3) Lemma 3.7 yields
u14 = u24 = u34 = u41 = u42 = u43 = 0.
This implies
3 = S3,
aut(Γ) is commutative and hence G+
aut(Γ) ⊆ S+
G+
thus G+
(4) Let ∆ and ∆′ be the comultiplication maps of G+
aut(Γ) = Aut(Γ) = Z2.
2 , respectively.
We first show that these two quantum groups coincide as compact quantum
groups, i.e. there is a ∗-isomorphism
aut(Γ) and H +
aut(Γ) = G∗
ϕ : C(H +
2 ) → C(G+
aut(Γ))
such that ∆′ ◦ ϕ = (ϕ ⊗ ϕ) ◦ ∆.
Step 1: The map ϕ exists and we have ∆′ ◦ ϕ = (ϕ ⊗ ϕ) ◦ ∆.
From εu = uε we get
u11 u12 u13 u14
u12 u11 u14 u13
u31 u32 u33 u34
u32 u31 u34 u33
u =
.
Define v11 := u11 − u12, v12 := u13 − u14, v21 := u31 − u32 and v22 := u33 − u34.
One can compute that vij, i, j = 1, 2 fulfill the relations of C(H +
2 ) and with
the universal property we get a *-homomorphism ϕ : C(H +
aut(Γ)).
Since ∆′ ◦ ϕ = (ϕ ⊗ ϕ) ◦ ∆ also holds, we get that G+
aut(Γ) is a quantum
2 ) → C(G+
QUANTUM SYMMETRIES OF GRAPH C ∗-ALGEBRAS
11
subgroup of H +
2 .
Step 2: The map ϕ is a ∗-isomorphism.
Let (vij)i,j=1,2 be the generators of C(H +
2 ). Define
u11 := u22 :=
u13 := u24 :=
u31 := u42 :=
u33 := u44 :=
v2
11 + v11
2
v2
12 + v12
2
v2
21 + v21
2
v2
22 + v22
2
,
,
,
,
u12 := u21 :=
u14 := u23 :=
u41 := u32 :=
u34 := u43 :=
v2
11 − v11
2
v2
12 − v12
2
v2
21 − v21
2
v2
22 − v22
2
,
,
,
.
One can show that the (uij)1≤i,j≤4 fulfill the relations of C(G+
versal property now gives us a *-homomorphism ϕ′ : C(G+
and ϕ′ is the inverse of ϕ and vice versa.
aut(Γ)). The uni-
aut(Γ)) → C(H +
2 )
Step 3: We have G+
We have seen in Step 1, that
aut(Γ) = G∗
aut(Γ).
u11 = u22,
u31 = u42,
u12 = u21,
u32 = u41,
u13 = u24,
u33 = u44,
u14 = u23,
u34 = u43
and therefore we get
uijukl = u2
kl = ukluij
for all (i, k), (j, l) ∈ E. Thus G+
aut(Γ) = G∗
aut(Γ).
(5) We conclude as in (3).
(6) Some direct computations using εu = uε and Relations (3.2) show
u33
1 − u33
1 − u33
u =
0
0
u33
0
0
0
0
0
0
u33
1 − u33
1 − u33
u33
.
(cid:3)
Thus G+
aut(Γ) is commutative.
12
SIMON SCHMIDT AND MORITZ WEBER
4. Proof of the main result
We now prove the main result of this article (see Section 1) for a finite graph Γ
with vertices V = {1, . . . , n} and edges E = {e1, . . . , em} having no multiple edges.
Remark 4.1. We define the quantum symmetry group QSym(C ∗(Γ)) of C ∗(Γ) to
be the maximal compact matrix quantum group G acting on C ∗(Γ) by α : C ∗(Γ) →
C(G) ⊗ C ∗(Γ) and β : C ∗(Γ) → C(G) ⊗ C ∗(Γ) as defined in the statement of our
main theorem. We thus have to show that G+
aut(Γ) acts on C ∗(Γ) via α and β (see
Sections 4.1 and 4.2) and that it is maximal with these actions (see Section 4.3).
4.1. Existence of the maps α and β. In order to prove that
α : C ∗(Γ) → C(G+
aut(Γ)) ⊗ C ∗(Γ)
n
pi 7→ p′
i :=
uik ⊗ pk,
1 ≤ i ≤ n
m
Xk=1
Xl=1
sej 7→ s′
ej :=
us(ej)s(el)ur(ej)r(el) ⊗ sel,
1 ≤ j ≤ m
defines a ∗-homomorphism, all we have to show is that the relations of C ∗(Γ) hold
ej . We may then use the universal property of C ∗(Γ). The proof for the
for p′
existence of β is analogous.
i and s′
4.1.1. The p′
over, using pkpl = δklpk and Relations (3.1), we have
i are mutually orthogonal projections. Obviously, p′
i = (p′
i)∗ holds. More-
p′
ip′
j =
n
Xk,l=1
uikujl ⊗ pkpl =
n
Xk=1
uikujk ⊗ pk = δijp′
i.
are partial isometries with (s′
4.1.2. The s′
ej
(see Section 2.3) and Relations (3.1), we have
ej )∗s′
ej = p′
r(ej ). Using s∗
elsei = δilpr(ei)
m
(s′
ej )∗s′
ej =
=
ur(ej)r(el)us(ej)s(el)us(ej)s(ei)ur(ej)r(ei) ⊗ s∗
elsei
ur(ej)r(ei)us(ej)s(ei)ur(ej)r(ei) ⊗ pr(ei).
m
Xl,i=1
Xi=1
By Relations (3.3) we have ur(ej)j ′us(ej)i′ur(ej)j ′ = 0 for (i′, j′) /∈ E. This yields
ur(ej)r(ei)us(ej)s(ei)ur(ej)r(ei) ⊗ pr(ei) =
m
Xi=1
n
Xi′,j ′=1
ur(ej)j ′us(ej)i′ur(ej)j ′ ⊗ pj ′.
ej (s′
get for v ∈ V with s−1(v) 6= ∅:
4.1.3. We have Pj: s(ej )=v s′
Xj∈{1,...,m}
s(ej)=v
s′
ej (s′
ej )∗ = Xj∈{1,...,m}
s(ej )=v
m
=
Xl=1 Xi∈{1,...,m}
r(ei)=r(el)
uvs(el)ur(ej)r(el)ur(ej)r(ei)uvs(ei) ⊗ sels∗
ei
m
Xi,l=1
ur(ej)r(el)
uvs(ei) ⊗ sels∗
ei
uvs(el)
s(ej )=v
Xj∈{1,...,m}
ur(ej)r(el) = Xq∈V
(v,q)∈E
uqr(el)
QUANTUM SYMMETRIES OF GRAPH C ∗-ALGEBRAS
13
Using Relations (3.2), we obtain Pn
n
(s′
ej )∗s′
ej =
ur(ej)j ′us(ej)i′ur(ej)j ′ ⊗ pj ′ =
i=1 us(ej)i′ = 1 and thus
Xi′,j ′=1
ur(ej)j ′ ⊗ pj ′ = p′
r(ej).
n
Xj ′=1
ej )∗ = p′
v for s−1(v) 6= ∅. Using Relations (3.1), we
Now,
Xj∈{1,...,m}
s(ej)=v
and for q ∈ V with (v, q) /∈ E we have uvs(el)uqr(el) = 0 by Relations (3.4). Thus, for
any l ∈ {1, . . . , m}, we have using Relations (3.2)
and hence:
uvs(el) Xq∈V
(v,q)∈E
uqr(el) = uvs(el)Xq∈V
uqr(el) = uvs(el)
s′
ej (s′
ej )∗ =
Xj∈{1,...,m}
s(ej )=v
m
Xl=1 Xi∈{1,...,m}
r(ei)=r(el)
uvs(el)uvs(ei) ⊗ sels∗
ei
Since Γ has no multiple edges by assumption, r(ei) = r(el) and s(ei) = s(el) implies
ei = el. We thus infer using Relations (3.1):
s′
ej (s′
ej )∗ =
Xj∈{1,...,m}
s(ej )=v
m
Xl=1
uvs(el) ⊗ sels∗
el
Now, for V ′ := {q ∈ V s−1(q) 6= ∅}, we have, using the relations in C ∗(Γ):
m
Xl=1
uvs(el) ⊗ sels∗
el = Xq∈V ′ Xl∈{1,...,m}
s(el)=q
uvq ⊗ sels∗
el = Xq∈V ′
uvq ⊗ pq
14
SIMON SCHMIDT AND MORITZ WEBER
Since we know that uvq = 0 for q /∈ V ′ by Lemma 3.7, we finally conclude
s′
ej (s′
ej )∗ =
Xj∈{1,...,m}
s(ej)=v
n
Xq=1
uvq ⊗ pq = p′
v.
This settles the existence of α.
4.2. The map α is a left action and β is a right action. We only prove this
claim for α, the proof for β being analogous.
4.2.1. (∆ ⊗ id) ◦ α = (id ⊗ α) ◦ α holds and α is unital. Using Relations (3.3), this
is straightforward to check.
It remains to show that
S := span α(C ∗(Γ))(C(G+
aut(Γ)) ⊗ 1)
is dense in C(G+
aut(Γ)) ⊗ C ∗(Γ), which we will do in the sequel.
4.2.2. The elements 1 ⊗ pl, 1 ⊗ sel and 1 ⊗ s∗
el
(3.2) we infer:
are in S. Using Relations (3.1) and
S ∋
n
Xi=1
α(pi)(uil ⊗ 1) =
n
n
Xi=1
Xj=1
uijuil ⊗ pj =
n
Xi=1
uil ⊗ pl = 1 ⊗ pl
Moreover, for el ∈ E we get, using Relations (3.1) and V ′ := {v ∈ V s−1(v) 6= ∅}:
Xv∈V ′ Xj∈{1,...,m}
s(ej )=v
α(sej )(ur(ej)r(el)uvs(el) ⊗ 1)
= Xv∈V ′ Xj∈{1,...,m}
s(ej )=v
= Xv∈V ′
Xk∈{1,...,m}
r(ek)=r(el)
m
uvs(ek)ur(ej)r(ek)ur(ej)r(el)uvs(el) ⊗ sek!
Xk=1
uvs(ek)
uvs(el) ⊗ sek
ur(ej)r(el)
Xj∈{1,...,m}
s(ej)=v
We proceed similar to Step 4.1.3. By Relations (3.4), we know uqr(el)uvs(el) = 0 for
(v, q) /∈ E. Thus, by Relations (3.1) and (3.2) and using that Γ has no multiple
edges, we obtain:
Xv∈V ′ Xj∈{1,...,m}
s(ej )=v
α(sej )(ur(ej)r(el)uvs(el) ⊗ 1)
QUANTUM SYMMETRIES OF GRAPH C ∗-ALGEBRAS
15
r(ek)=r(el)
Xk∈{1,...,m}
Xk∈{1,...,m}
r(ek)=r(el)
uvs(el) ⊗ sel
uqr(el)! uvs(el) ⊗ sek
uvs(ek) n
Xq=1
uvs(ek)uvs(el) ⊗ sek
= Xv∈V ′
= Xv∈V ′
= Xv∈V ′
Finally, Lemma 3.7 yields uvs(el) = 0 for v /∈ V ′. Hence, using Relations (3.2):
S ∋ Xv∈V ′ Xj∈{1,...,m}
s(ej )=v
α(sej )(ur(ej)r(el)us(ej)s(el) ⊗ 1) =
uis(el) ⊗ sel = 1 ⊗ sel
n
Xi=1
Define V ′′ := {v ∈ V r−1(v) 6= ∅}. Similar to the computations above, we get
S ∋ Xv∈V ′′ Xj∈{1,...,m}
s(ej)=v
α(s∗
ej )(us(ej)s(el)ur(ej)r(el) ⊗ 1) = 1 ⊗ s∗
el.
4.2.3. If 1 ⊗ x, 1 ⊗ y ∈ S, then also 1 ⊗ xy ∈ S. The remainder of the proof of Step
4.2 consists in general facts for actions of compact matrix quantum groups.
We may write 1 ⊗ x ∈ S and 1 ⊗ y ∈ S as
l
1 ⊗ x =
α(zi)(wi ⊗ 1),
1 ⊗ y =
Xi=1
α(tj)(vj ⊗ 1)
k
Xj=1
for some zi, tj ∈ C ∗(Γ) and wi, vj ∈ C(G+
aut(Γ)). Therefore
1 ⊗ xy =
=
=
l
l
Xi=1
Xi=1
Xi=1
l
α(zi)(wi ⊗ 1)(1 ⊗ y)
α(zi)(1 ⊗ y)(wi ⊗ 1)
α(zitj)(vjwi ⊗ 1) ∈ S
k
Xj=1
4.2.4. S is dense in C(G+
all monomials w in pi, sej , s∗
aut(Γ)) ⊗ C ∗(Γ). Summarizing, we get that 1 ⊗ w ∈ S for
ej , 1 ≤ i ≤ n, 1 ≤ j ≤ m. Since α is unital, we also have:
C(G+
aut(Γ)) ⊗ 1 ⊆ α(C ∗(Γ))(C(G+
aut(Γ)) ⊗ 1) ⊆ S
Xk=1
Xi,l=1
Xi=1
Xi=1
16
SIMON SCHMIDT AND MORITZ WEBER
We conclude that S is dense in C(G+
aut(Γ)) ⊗ C ∗(Γ), which settles Step 4.2.
4.3. The quantum group G+
aut(Γ) acts maximally on C ∗(Γ). For proving the
maximality, let G = (C(G), u) be another compact matrix quantum group acting
on C ∗(Γ) by α′ : C ∗(Γ) → C(G) ⊗ C ∗(Γ) and β′ : C ∗(Γ) → C(G) ⊗ C ∗(Γ) in
the way G+
aut(Γ) acts on C ∗(Γ) via α and β. We want to show that there is a *-
homomorphism C(G+
aut(Γ)) → C(G) sending generators to generators. Thus, we
need to compute that the generators uij of C(G) fulfill the relations of C(G+
aut(Γ)).
4.3.1. The Relations (3.1) hold in C(G). The equation
n
n
uik ⊗ pk = α′(pi) = α′(pi)∗ =
u∗
ik ⊗ pk
Xk=1
yields uij = u∗
ij after multiplying from the left with 1 ⊗ pj. We also have
n
n
n
ujiuki ⊗ pi =
ujiukl ⊗ pipl = α′(pj)α′(pk) = δjkα′(pj) =
δjkuji ⊗ pi
from which we infer ujiuki = δjkuji. Using β′, we also obtain uijuik = δjkuij.
4.3.2. The Relations (3.2) hold in C(G). From
n
n
1 ⊗ pk = 1 ⊗ 1 = α′(1) =
Xk=1
Xi=1
i=1 uik = 1, and likewise Pn
we deduce Pn
4.3.3. The Relations (3.3) hold in C(G). Using s∗
and Relations (3.1) in C(G), we obtain for any j:
α′(pi) =
i=1 uki = 1 using β′.
n
Xk=1 n
Xi=1
uik! ⊗ pk
elset = δltpr(el) (see Section 2.3)
n
Xq=1
ur(ej)q ⊗ pq = α′(pr(ej))
= α′(s∗
ej sej )
m
ur(ej)r(el)us(ej)s(el)us(ej)s(et)ur(ej)r(et) ⊗ s∗
elset
ur(ej)r(el)us(ej)s(el)ur(ej)r(el) ⊗ pr(el)
=
=
m
Xl,t=1
Xl=1
Multiplication with 1 ⊗ pk yields:
ur(ej)k = Xl∈{1,...,m}
r(el)=k
ur(ej)kus(ej)s(el)ur(ej)k
If r−1(k) = ∅, then ur(ej)k = 0 and hence us(ej)iur(ej)k = ur(ej)kus(ej)i = 0 for all
i ∈ V .
QUANTUM SYMMETRIES OF GRAPH C ∗-ALGEBRAS
17
Otherwise, if r−1(k) 6= ∅, we use Relations (3.1) and (3.2) in C(G) and get
ur(ej)kus(ej)s(el)ur(ej)k = ur(ej)k = u2
r(ej )k =
ur(ej)kus(ej)iur(ej)k
n
Xi=1
ur(ej)kus(ej)iur(ej)k = 0.
Xi∈V
(i,k) /∈E
Xl∈{1,...,m}
r(el)=k
and therefore
Since
ur(ej)kus(ej)iur(ej)k = (us(ej)iur(ej)k)∗us(ej)iur(ej)k
holds, the above is a vanishing sum of positive elements – and hence each summand
vanishes. This yields us(ej)iur(ej)k = 0 for all (i, k) /∈ E.
4.3.4. The Relations (3.4) hold in C(G). The argument is analogous to the one for
proving Relations (3.3) when replacing α′ by β′.
The proof of the main theorem is complete.
Remark 4.2. Let Γ be a finite graph with n vertices V = {1, ..., n} and m edges
E = {e1, ..., em}.
aut(Γ) is the quantum symmetry
group of Γ in his sense, where
In [5], Bichon showed that G∗
βV : C(V ) → C(G∗
aut(Γ)) ⊗ C(V ), gi 7→
βE : C(E) → C(G∗
aut(Γ)) ⊗ C(E), fj 7→
uki ⊗ gk,
us(el)s(ej )ur(el)r(ej ) ⊗ fl,
n
m
Xk=1
Xl=1
define actions of G∗
aut(Γ) on C(V ) and C(E), respectively. Those actions inspired
us, how an action of a compact matrix quantum group on C ∗(Γ) should look like.
However, note that edges in the commutative C ∗-algebra C(E) of continuous func-
tions on E are represented as projections unlike in the case of C ∗(Γ). Therefore, the
quantum symmetry group of C ∗(Γ) is G+
aut(Γ). On the other
hand, if we consider the quotient of C ∗(Γ) by the relations se = s∗
e, its quantum
symmetry group is G∗
aut(Γ). Indeed, selfadjointness of se yields
aut(Γ) rather than G∗
us(ej)s(el)ur(ej)r(el) ⊗ sel = α(sej ) = α(sej )∗ =
m
Xl=1
ur(ej)r(el)us(ej)s(el) ⊗ sel,
m
Xl=1
from which we obtain Relations (3.5) by multiplication with (1 ⊗ s∗
ei) from the left.
18
SIMON SCHMIDT AND MORITZ WEBER
References
[1] Teodor Banica. Quantum automorphism groups of homogeneous graphs. J. Funct. Anal.,
224(2):243–280, 2005.
[2] Teodor Banica and Julien Bichon. Quantum automorphism groups of vertex-transitive graphs
of order ≤ 11. J. Algebraic Combin., 26(1):83–105, 2007.
[3] Teodor Banica, Julien Bichon, and Gaetan Chenevier. Graphs having no quantum symmetry.
Ann. Inst. Fourier, pages 955–971, 2007.
[4] Jyotishman Bhowmick, Debashish Goswami, and Adam Skalski. Quantum isometry groups of
0-dimensional manifolds. Trans. Amer. Math. Soc., 363(2):901–921, 2011.
[5] Julien Bichon. Quantum automorphism groups of finite graphs. Proc. Amer. Math. Soc.,
131(3):665–673, 2003.
[6] Julien Bichon. Free wreath product by the quantum permutation group. Algebr. Represent.
Theory, 7(4):343–362, 2004.
[7] Arthur Chassaniol. Quantum automorphism group of the lexicographic product of finite reg-
ular graphs. J. Algebra, pages 23–45, 2016.
[8] Joachim Cuntz and Wolfgang Krieger. A class of C ∗-algebras and topological Markov chains.
Invent. Math., 56(3):251–268, 1980.
[9] Søren Eilers, Gunnar Restorff, Efren Ruiz, and Adam Sørensen. The complete classification
of unital graph C ∗-algebras: Geometric and strong. arXiv:1611.07120, 2016.
[10] Melanie Fulton. The quantum automorphism group and undirected trees. PhD Thesis, Vir-
ginia, 2006.
[11] Debashish Goswami and Jyotishman Bhowmick. Quantum isometry groups. Infosys Science
Foundation Series. Springer, New Delhi, 2016. Infosys Science Foundation Series in Mathe-
matical Sciences.
[12] Sergey Neshveyev and Lars Tuset. Compact quantum groups and their representation cate-
gories, volume 20 of Cours Sp´ecialis´es [Specialized Courses]. Soci´et´e Math´ematique de France,
Paris, 2013.
[13] Piotr Podle´s. Symmetries of quantum spaces. Subgroups and quotient spaces of quantum
SU(2) and SO(3) groups. Comm. Math. Phys., 170(1):1–20, 1995.
[14] Iain Raeburn. Graph algebras, volume 103 of CBMS Regional Conference Series in Mathe-
matics. Published for the Conference Board of the Mathematical Sciences, Washington, DC;
by the American Mathematical Society, Providence, RI, 2005.
[15] Roland Speicher and Moritz Weber. Quantum groups with partial commutation relations.
arXiv:1603.09192, 2016.
[16] Thomas Timmermann. An invitation to quantum groups and duality. EMS Textbooks in Math-
ematics. European Mathematical Society (EMS), Zurich, 2008.
[17] Shuzhou Wang. Quantum symmetry groups of finite spaces. Comm. Math. Phys., 195(1):195–
211, 1998.
[18] S. L. Woronowicz. Compact matrix pseudogroups. Comm. Math. Phys., 111(4):613–665, 1987.
[19] S. L. Woronowicz. A remark on compact matrix quantum groups. Lett. Math. Phys., 21(1):35–
39, 1991.
Saarland University, Fachbereich Mathematik, 66041 Saarbrucken, Germany
E-mail address: [email protected], [email protected]
|
1710.06123 | 1 | 1710 | 2017-10-17T06:51:30 | A theorem for random Fourier series on compact quantum groups | [
"math.OA",
"math.FA"
] | Helgason showed that a given measure $f\in M(G)$ on a compact group $G$ should be in $L^2(G)$ automatically if all random Fourier series of $f$ are in $M(G)$. We explore a natural analogue of the theorem in the framework of compact quantum groups and apply the obtained results to study complete representability problem for convolution algebras of compact quantum groups as an operator algebra. | math.OA | math |
A THEOREM FOR RANDOM FOURIER SERIES ON COMPACT
QUANTUM GROUPS
SANG-GYUN YOUN
Abstract. Helgason showed that a given measure f P M pGq on a compact
group G should be in L2pGq automatically if all random Fourier series of f
are in M pGq. We explore a natural analogue of the theorem in the framework
of compact quantum groups and apply the obtained results to study complete
representability problem for convolution algebras of compact quantum groups
as an operator algebra.
1. Introduction
The random Fourier series theory has been extensively studied for a long period
of time [PZ32], [Kah93], [SZ54], [Hun51], [Dud67], [Fer75], [Mar78] and [MP78] and
most of the results turned out to be valid for Banach space-valued functions on
compact groups [FTR66], [Kwa76] and [MP81]. Moreover, it has been found that
these studies can be applied to various topics of harmonic analysis [Pis77], [Pis78]
and [Rid75].
A theorem of Helgason [Hel57], which is a subject of this paper, is one of the re-
sults of the theory of random Fourier series. It is an improvement of the Littlewood's
work on circle [Lit26].
Throughout this paper, we denote by IrrpGq a maximal family of mutually in-
equivalent irreducible unitary representations of a compact group G and by U the
Upnπq where Updq is the group of unitary matrices of size d
product group źπPIrrpGq
and nπ is the dimension of π P IrrpGq. Also, we denote by
f „ ÿπPIrrpGq
nπtrppfpπqπpxqq
Fourier coefficients of f P MpGq.
Theorem 1.1 (S.Helgason, [Hel57]). Let G be a compact group and fix a measure
the Fourier-Stieltjes series of f P MpGq, where ppfpπqqπPIrrpGq is the sequence of
f P MpGq. Suppose that there exists µU P MpGq such that µU „ ÿπPpG
nπtrpUπpfpπqπpxqq
for any U " pUπqπPpG P U. Then
ÿπPpG
nπtrppfpπqpfpπqq ă 8.
After the works of S.Woronowicz, which are [Wor87a], [Wor87b] for compact
quantum groups, the theory of topological quantum groups has been greatly ele-
vated [KV00], [Kus01], [KV03] and harmonic analysis on quantum groups has also
been vigorously studied in [Cas13], [FHL`17], [JMP14], [JPPP17] and [You18]. In
particular, progress has been made in research on random Fourier series [Wan17].
Key words and phrases. Random Fourier series, compact quantum group, non self-adjoint
operator algebra.
2010 Mathematics Subject Classification. Primary 46L89, Secondary 20G42, 43A30.
The author is supported by TJ Park Science Fellowship.
1
The main purpose of this paper is to generalize Theorem 1.1 in the framework
of compact quantum groups in a natural manner through mobilizing well-known
studies on stochastic behavior of some special vector-valued random variables(e.g.
Rademacher and Gaussian variables), namely the non-commutative Khintchine in-
equalities [LP86], [LPP91] and (co-)type studies [TJ74], [Fac87].
The Fourier analysis on compact quantum groups has been developed in respect
of traditional philosophy of the Fourier analysis on compact groups to large extent.
However, a visible difference appears in the Fourier expansion. In fact, our random
Fourier series of f P MpGq will be described as
fU „ ÿαPIrrpGq
dαtrpUαpfpαqQαuαq
for any U " pUαqαPIrrpGq P U. We will explain the details of the above expansion in
Section 2.
As in the classical setting, f P L2pGq will imply that all of random fourier series
fU are in L2pGq Ď L1pGq by the Plancherel identity on compact quantum groups
(Proposition 2.1). Accordingly what we have to do is to demonstrate the converse
direction and main results of this paper are as follows.
Theorem 1.2. In the above notation, let us suppose that fU P MpGq for all U "
pUαqαPIrrpGq P U.
(1) If G is of Kac type, then we have
(2) For general compact quantum groups, we have
ÿαPIrrpGq
ÿαPIrrpGq
dα
nα
nαtrppfpαqpfpαqq ă 8.
trpQαpfpαqpfpαqq ă 8.
Moreover, we will apply the obtained results to a problem determining whether
the associated convolution algebra L1pGq can be completely isomorphic to a closed
subalgebra of BpHq, in which H is a Hilbert space (See Section 4). In this case, we
will say that L1pGq is completely representable as an operator algebra.
Corollary 1.3. Let G " pA, ∆q be a compact quantum group. Then L1pGq is
completely representable as an operator algebra if and only if A is finite dimensional.
2. Preliminaries
2.1. Compact quantum group. A compact quantum group is a pair G " pA, ∆q
where A is a unital C-algebra and ∆ : A Ñ Abmin A is a unital ´homomorphism
satisfying
(1) p∆ b idq ∆ " pid b ∆q ∆ and
(2) spant∆paqpb b 1q : a, b P Au and spant∆paqp1 b bq : a, b P Au are dense in
A bmin A.
Let us describe the basic representation theory and the Schur's orthogonality re-
lation on compact quantum groups. A finite dimensional unitary representation of
ui,k b uk,j and puuqi,j "
G is u " pui,jq1ďi,jďn P MnpAq such that ∆pui,jq "
u
k,iuk,j " δi,j1A "
j,k " puuqi,j for all 1 ď i, j ď n. We say that a uni-
tary representation u P MnpAq is irreducible if tX P Mn : Xu " uXu " C Idn and
a maximal family of mutually inequivalent
ui,ku
nÿk"1
denote by uα " puα
nÿk"1
i,jq1ďi,jďnα(αPIrrpGq
unitary irreducible representations of G.
nÿk"1
2
Every compact quantum group allows the analogue of the Haar measure, namely
the Haar state, which is the unique state h on A such that pidbhqp∆paqq " hpaq1A "
ph b idqp∆paqq for all a P A and hp1Aq " 1.
is unitary and trpQαq " trpQ´1
i,jqq1ďi,jďnα . We call dα "
trpQαq " trpQ´1
α q the quantum dimension of α P IrrpGq. For any α, β P IrrpGq,
1 ď i, j ď nα and 1 ď s, t ď nβ, we have the Schur's orthogonality relation as
follows.
For each α P IrrpGq, there exists a unique mathrix Qα P Mnα such that Q
α q where uα :" ppuα
α uαQ
´ 1
α
1
2
2
hppuβ
s,tquα
i,jq "
δα,βδj,tpQ´1
trpQαq
α qi,s
and hpuβ
s,tpuα
i,jqq "
δα,βδi,spQαqj,t
trpQαq
.
Moreover, we may assume that the matrices Qα are diagonal [Daw10]. Also, we
say that G is of Kac type if Qα " Idnα for all α P IrrpGq or equivalently if the Haar
state h is tracial.
We define the space of matrix coefficients as
PolpGq " span uα
i,j : α P IrrpGq, 1 ď i, j ď nα(
and then the Haar state h is faithful on PolpGq. We denote by L2pGq the closure
of PolpGq with respect to the pre-inner product xa, byL2pGq " hpbaq for all a, b P
PolpGq and denote by Λ : PolpGq ãÑ L2pGq, a ÞÑ Λpaq the natural embedding.
Also, we define CrpGq as the norm-closure of PolpGq in BpL2pGqq under the GNS
representation rιpaqspΛpxqq :" Λpaxq for all a, x P PolpGq and L8pGq as weak -
closure of CrpGq in BpL2pGqq.
The haar state h extends to L8pGq as a normal faithful state and we will denote
it by h again. We consider the predual L1pGq of L8pGq and the dual MpGq of CrpGq
respectively. Then we have a contractive embedding L8pGq ãÑ L1pGq, x ÞÑ hpxq
and the isometric formal identity from L1pGq into MpGq. Note that PolpGq is dense
in L1pGq.
2.2. Fourier analysis on compact quantum groups. It is known that every
sup
Mnα :
αPIrrpGq}Xα} ă 8,.-
quantum group structure under the generalized Pontrjagin duality. Although we
will not mention its quantum group structure, we will specifically explain some of
compact quantum group admits the dual discrete quantum group pG with a natural
non-commutative ℓp-spaces on the discrete quantum group pG.
The associated von Neumann algebra of pG is given by
ℓ8ppGq "$&%pXαqαPIrrpGq P źαPIrrpGq
with the norm››pXαqαPIrrpGq››ℓ8ppGq " sup
We define the Fourier-Stieltjes transform by F : MpGq " CrpGq Ñ ℓ8ppGq,
µ ÞÑ ppµpαqqαPIrrpGq with pµpαq " pµppuα
nαÿi,j"1
dαtrppµpαqQαuαq " ÿαPIrrpGq
f " ÿαPIrrpGq
to follow the standard notations of Fourier analysis on compact groups. Then the
Fourier expansion of µ P MpGq is given by
dαppµpαqQαqi,j uα
Indeed, if f P PolpGq, then we have
µ „ ÿαPIrrpGq
αPIrrpGq}Xα}Mnα
j,iqqq1ďi,jďnα ,
dαtrppfpαqQαuαq.
3
.
j,i.
The associated ℓ2 and ℓ1 spaces on pG are
ℓ2ppGq "$&%pXαqαPIrrpGq P ℓ8ppGq : ÿαPIrrpGq
ℓ1ppGq "$&%pXαqαPIrrpGq P ℓ8ppGq : ÿαPIrrpGq
››pXαqαPIrrpGq››ℓ2ppGq "p ÿαPIrrpGq
››pXαqαPIrrpGq››ℓ1ppGq " ÿαPIrrpGq
with the norm structures
,
dαtrpQαX
αXαq ă 8,.-
dαtrpXαQαq ă 8,.-
dαtrpQαX
α Xαqq
1
2 ,
dαtrpXαQαq
respectively. When we restrict the domain of the Fourier-Stieltjes transform to
L2pGq, we have the isometry F : L2pGq Ñ ℓ2ppGq [Wan17], which is called the
Plancherel identity.
2.3. Random Fourier series. From now on, we call formal series
ÿαPIrrpGq
dαtrpUαpfpαqQαuαq
the random Fourier series of f P MpGq for U " pUαqαPIrrpGq P U. The main
question of this paper is to find out when all of the random Fourier series are in
MpGq simultaneously.
It can be seen that one direction on our question can be
solved simply from the Plancherel identity.
Proposition 2.1. If f P L2pGq, all of random fourier series fU are in L2pGq.
Proof.
}fU}2
L2pGq " ÿαPIrrpGq
" ÿαPIrrpGq
α Uαpfpαqq
dαtrpQαpfpαqU
dαtrpQαpfpαqpfpαqq " }f}2
L2pGq .
(cid:3)
Lemma 2.2. For µ „ ÿαPIrrpGq
dαtrppµpαqQαuαq P MpGq and f P PolpGq,
Proof. For f " ÿαPIrrpGq
xµ, f yMpGq,CrpGq " ÿαPIrrpGq
dαppfpαqQαqi,j uα
nαÿi,j"1
f " ÿαPIrrpGq
nαÿi,j"1
dαtrppµpαqQαpfpαqq.
j,i P PolpGq,
dαppfpαqQαqi,jpuα
j,iq
4
and
xµ, f yMpGq,CrpGq " ÿαPIrrpGq
" ÿαPIrrpGq
" ÿαPIrrpGq
" ÿαPIrrpGq
" ÿαPIrrpGq
j,iqyMpGq,CrpGq
dαppfpαqQαqi,jxµ,puα
dαppfpαqQαqi,jrpµpαqsi,j
dαppfpαqQαqi,jrpµpαqsj,i
nαÿi,j"1
nαÿi,j"1
nαÿi,j"1
dαtrppfpαqQαpµpαqq
dαtrppµpαqQαpfpαqq.
(cid:3)
for all U " pUαqαPIrrpGq P U if and only if
Proposition 2.3. Fix a family of matrices pXαqαPIrrpGq. Then
dαtrpUαXαQαuαq P MpGq
µU „ ÿαPIrrpGq
µB „ ÿαPIrrpGq
for all B " pBαqαPIrrpGq P ℓ8ppGq.
Φ : ℓ8ppGq Ñ MpGq, B " pBαqαPIrrpGq ÞÑ µB „ ÿαPIrrpGq
dαtrpBαXαQαuαq P MpGq
In this case, the map
dαtrpBαXαQαuαq,
is automatically a bounded map by the closed graph theorem.
Proof. One direction is trivial. On the other direction, pick B " pBαqαPIrrpGq P
Ballpℓ8ppGqq and take an arbitrary f P PolpGq. Then
xµB, f yMpGq,CrpGq " ÿαPIrrpGq
Xα " ř4
For each α P IrrpGq, there exist unitaries vj
j"1 vj
2
by considering the following decomposition
α
dαtrpBαXαQαpfpαqq.
α with j " 1, , 4 such that
Xα "
X
α ` Xα
2
` i
Xα ´ X
α
2i
": h1 ` ih2
and
hj ` ib1 ´ hj2
2
hj ´ ib1 ´ hj2
2
`
hj " p
q for j " 1, 2.
We define Vj " pvj
xµB, f yMpGq,CrpGq "
αqαPIrrpGq P U for j " 1, , 4 and then
1
1
2
2
dαtrpvj
Since PolpGq is dense in CrpGq, we get the conclusion
αAαQαpfpαqq "
4ÿj"1xµVj , f yMpGq,CrpGq.
4ÿj"1 ÿαPIrrpGq
µB " ř4
j"1 µVj
2
5
P MpGq.
2.4. Vector valued probabilistic methods. In this subsection, we gather some
probabilistic techniques, namely non-commutative Khintchine inequality and cotype
2 condition, to be used in this study.
First of all, let us write down the non-commutative Khintchine inequality in
the context of the present compact quantum group setting. For more details, see
[Pis12].
(cid:3)
For a sequence pxjqn
j"1 Ď L1pGq, we set
~pxjqn
j"1~1 " sup!ÿxxj, yjyL1pGq,L8pGq : yj P L8pGq, max!››pyjqn
j"1››C
,››pyjqn
2›››››L8pGq
j"1››R "›››››p
j"1››C "›››››p
nÿj"1
nÿj"1
j"1 be an independent and identically distributed(in short, i.i.d.) sequece
of real valued Gaussian random variables with mean zero and variance 1. Then a
part of the Khintchine inequality for non-commutative L1-spaces is as follows:
where››pyjqn
2›››››L8pGq
and››pyjqn
Let pgjqn
yjy
j q
y
j yjq
.
1
1
j"1››R) ď 1)
Theorem 2.4 (Theorem 9.1, [Pis12]). If G is of Kac type, there exists a universal
constant C ą 0 such that
1
C ~pxjqn
dPpwq ď ~pxjqn
j"1~1
j"1~1 ďż ›››››
j"1 Ď L1pGq.
nÿj"1
gjpwqxj›››››
for any finite set pxjqn
Also, we will make use of a cotype study on non-commutative L1-spaces.
In
general, we say that a Banach space is of cotype 2 if there is a constant K ą 0 such
that for all finite sets pxjqn
j"1 Ď B we have
ż ›››››
nÿj"1
ǫjpwqxj›››››B
dPpwq ě Kp
nÿj"1}xj}2q
1
2 ,
where pǫjqj is a family of i.i.d. Rademacher variables.
It is known that every
predual of von Neumann algebra is of cotype 2 [TJ74] and it is possible to replace
the Rademacher varialbes ǫj to the Gaussian variables gj [LT13].
Theorem 2.5. There exists a universal constant K ą 0 such that
ż ›››››
nÿj"1
gjpwqxj›››››L1pGq
j"1 Ď L1pGq.
for any finite sets pxjqn
dPpwq ě Kp
nÿj"1 }xj}2
1
2
L1pGqq
3. The main results
3.1. Affirmative answer on Kac cases. In this subsection, we will make use of
random matrices Gn " p
dent real valued Gaussian random variables with mean zero and variance 1.
i,j"1 P Mn where pgn
i,j"1 is a family of indepen-
i,jqn
gn
1
?n
i,jqn
Lemma 3.1 (Proposition 1.5, Chapter 5, [MP81]). There exist universal constants
C1, C2 ą 0, which are independent of n, such that
C1 ďż }Gnpwq} dPpwq ď C2,
where }} is the operator norm of matrices.
6
Theorem 3.2. Let G be a compact quantum group of Kac type and suppose that
all random fourier series fU „ ÿαPIrrpGq
for all U " pUαqαPIrrpGq P U. Then we have
nαtrpUαpfpαquαq of f P MpGq are in MpGq
nαtrppfpαqpfpαqq ă 8.
ÿαPIrrpGq
K }B}ℓ8ppGq ě››››››
Proof. By Proposition 2.3, there exists a universal constant K ą 0 such that
fB „ ÿαPIrrpGq
rally takes into account a family of random matrices pGnαqαPS " pp
nαtrpBαpfpαquαq››››››MpGq
for any B " pBαqαPIrrpGq P ℓ8ppGq. Now we fix a finite subset S Ď IrrpGq and natu-
in ℓ8ppGq. Then from the estimate above and Theorem 2.4, we have
i,jpwqppfpαquαqj,i›››››L1pGq
αPS }Gnα} dPpwq ěż ›››››ÿαPS
Kż sup
nαÿi,j"1
?nαgα
i,jqnα
gα
1
?nα
dPpwq
i,j"1qαPS
(3.1)
1
ě
"
C ~p?nαppfpαquαqj,iqα,i,j~1
i,j qα,i,jÿxxα
i,j, yα
1
C
sup
pyα
i,jyL1pGq,L8pGq
i,jqα,i,j Ď L8pGq
where xα
such that
i,j " ?nαppfpαquαqj,i and the supremum runs over all pyα
max$&%
›››››pÿαPS
,›››››pÿαPS
2›››››L8pGq
2›››››L8pGq
For arbitrary pApαqqαPS such that ÿαPS
nαÿi,j"1pyα
nαÿi,j"1
i,jqyα
i,jq
yα
i,jpyα
i,jqq
1
1
,.- ď 1.
nαtrpApαqApαqq ď 1, we can find that
the family pyα
i,jq " p?nαrpApαquαqj,isq satisfies the condition above. Indeed,
nαÿi,j"1pyα
ÿαPS
nαpApαquαqj,irpApαquαqj,is
i,jqyα
nαpApαquαqj,irpuαqApαqsi,j
i,j " ÿαPS
" ÿαPS
" ÿαPS
" ÿαPS
" ÿαPS
nαÿi,j"1
nαÿi,j"1
nαtrpApαquαpuαqApαqq
nαtrpApαqApαqq1A
nαtrpApαqApαqq1A
7
and
ÿαPS
nαÿi,j"1
i,jpyαq
yα
i,j " ÿαPS
" ÿαPS
" ÿαPS
" ÿαPS
" ÿαPS
so that we have
nαrpApαquαqj,ispApαquαqj,i
nαpApαquαqj,irpuαqtApαqtsi,j
nαÿi,j"1
nαÿi,j"1
nαtrpApαquαpuαqtApαqtq
nαtrpApαqApαqtq1A
nαtrpApαqApαqq1A,
max$&%
›››››pÿαPS
ď pÿαPS
nαÿi,j"1pyα
i,jqyα
i,jq
nαtrpApαqApαqqq
1
2›››››L8pGq
1
2 ď 1
,›››››pÿαPS
nαÿi,j"1
yα
i,jpyα
i,jqq
1
2›››››L8pGq
,.-
Finally, by Lemma 3.1, we have
KC2 ě
"
"
"
"
sup
nα
nαÿi,j"1x?nαppfpαquαqj,i,?nαrpApαquαqj,isyL1pGq,L8pGq
nαÿi,j"1
nαÿi,j
nαtrppfpαqApαqq
nαÿk,l"1 pfpαqj,kApαqj,lxuα
nαÿk,l"1
nα pfpαqj,kApαqj,l
l,iqyL1pGq,L8pGq
k,i,puα
δk,l
nα
1
sup
sup
1
C
1
C
1
C
pApαqqαPS ÿαPS
pApαqqαPS ÿαPS
pApαqqαPS ÿαPS
pApαqqαPS ÿαPS
C pÿαPS
nαtrppfpαqpfpαqqq
1
C
sup
1
2
since the supremum runs over all pApαqqαPS such that ÿαPS
Now we reach the conclusion since the finite set S is chosen to be arbitrary.
nαtrpApαqApαqq ď 1.
(cid:3)
Remark 3.3. In the category of compact groups, it seems that some experts already
have their own way to understand Theorem 1.1 through the Khintchine inequality.
Indeed, the arguments below (3.1) are simplified as follows.
8
dPpwq
dPpwqdx
?nπgπ
?nπgπ
i,jpwqppfpπqπpxqqj,i›››››L1pGq
ż ›››››ÿπPS
nπÿi,j"1
"żGż ÿπPS
i,jpwqppfpπqπpxqqj,i
nπÿi,j"1
ÁżGpÿπPS
nπppfpπqπpxqqj,i
nπÿi,j"1
"żGpÿπPS
nπtrppfpπqπpxqπpxqpfpπqqq
" pÿπPS
nπtrppfpπqpfpπqqq
2 dx
2 dx
2 .
q
2
1
1
1
Remark 3.4. Let G be a compact quantum group of Kac type. If we suppose that
ÿαPIrrpGq
nαtrpX
α Xαq ă 8, then the multiplier
is bounded for all 2 ď p ď 8 since LppGq Ď L2pGq for such p. The converse is also
true thanks to Theorem 3.2. Indeed, boundedness of Φ implies that the multiplier
Φ : LppGq Ñ ℓ1ppGq, f ÞÑ pXαpfpαqqαPIrrpGq,
Ψ : ℓ8ppGq Ñ MpGq,pAαqαPIrrpGq ÞÑ ÿαPIrrpGq
dαtrpAαX
αuαq
is bounded. Then, by Proposition 2.3 and Theorem 3.2, we can conclude that the
space of such multipliers from LppGq into ℓ1ppGq can be identified with L2pGq inde-
pendently of the choice of 2 ď p ď 8.
3.2. A partial answer on general cases. In spite of the complete answer on
the Kac case, the proof of Theorem 3.2 could not be applied for the general case
for now. However, using the cotype 2 property of L1pGq, we are able to achieve a
convincing conclusion for non-Kac cases.
Lemma 3.5.
(1) For any α P IrrpGq and any 1 ď i, j ď nα, we have
Aikpuα
α qi,k2q
αqj,jp
ď pQ
´ 1
2 .
1
2
2
1
(2) For any α P IrrpGq and any 1 ď i, j ď nα, we have
Proof.
(1) Since Q
is a unitary representation, we have
nαÿk"1pAαQ
nαÿk"1pBQ
p
´ 1
ě pQ
2
α qj,j
dα
1
2
αqi,k2q
1
2 .
2
2
1
´ 1
α
α uαQ
›››››
k,jq›››››L8pGq
nαÿk"1
›››››
k,j›››››L1pGq
nαÿk"1pBQαqi,kuα
α qj,j›››››
nαÿk"1
" ›››››
nαÿk"1pAQ
nαÿk"1 pAQ
nαÿk"1 pAQ
α qi,k2q
α qi,k2q
α qi,krQ
Ai,kpuα
ď p
pQ
"p
2 p
´ 1
´ 1
´ 1
´ 1
2 .
2
2
2
1
1
2
1
2
α uQ
´ 1
k,jq›››››L8pGq
α sk,j›››››L8pGq
nαÿk"1rQ
α uQ
´ 1
2
2
1
9
2
α s
k,jrQ
1
2
α uQ
´ 1
2
α sk,jq
1
2
(2) Define wi,k " pAQ
p1q, we can see that
´ 1
k,j›››››L1pGq
nαÿk"1pBQαqi,kuα
x
2 ď1
›››››
nαÿk"1pBQαqi,kuα
přnα
α qj,j
" pQ
k"1 wi,k2q
sup
ě
´ 1
2
1
" pQ
´ 1
2
α qj,j
2
1
2
´ 1
α qj,j
k,j,pQ
Ai,kpuα
nαÿk"1
nαÿk"1pBQαqi,kwi,kpQ
αqk,khppuα
nαÿk"1
2 " pQ
αqi,k
pBQ
dα
α qj,j
dα
wi,k
nαÿk"1 pBQ
´ 1
p
2
2
1
1
1
sup
k"1 wi,k2q
1
2 ď1
sup
1
1
2 ď1
k"1 wi,k2q
αqi,k2
pBQ
d2
α
2
přnα
přnα
nαÿk"1
k,jqyL1pGq,L8pGq
k,jquα
k,jq
2
α qi,k for all 1 ď i, k ď nα for convenience. Then, by
´ 1
2
q
" pQ
α qj,jp
αqi,k2q
Theorem 3.6. Suppose that all random Fourier series fU „ ÿαPIrrpGq
are in MpGq for all U " pUαqαPIrrpGq P U. Then we have
2
2
1
(cid:3)
dαtrpUαpfpαqQαuαq
Proof. As in the proof of Theorem 3.2, there exists a universal constant K ą 0,
which is independent of the choice of finite subset S Ď IrrpGq, such that
dPpwq
gα
j,ipwq
where pgα
zero and variance 1.
Let vα
Theorem 2.5, we have that
ÿαPIrrpGq
dα
nα
dα?nα
nαÿi,j"1
j,ipwqqαPIrrpGq,1ďi,jďnα is a i.i.d.
trpQαpfpαqpfpαqq ă 8.
k,j›››››L1pGq
K ěż ›››››ÿαPS
nαÿk"1ppfpαqQαqi,kuα
i be the i-th row vector of the matrix pfpαqQ
nα›››››
k,j›››››
nαÿi,j"1
nαÿk"1ppfpαqQαqi,kuα
nαÿi,j"1
i }2
α qj,j }vα
trpQαpfpαqpfpαqqq
K ě K 1pÿαPS
ě K 1pÿαPS
" pÿαPS
pQ´1
dα
nα
L1pGq
d2
α
nα
2 .
q
q
2
1
2
1
2
1
1
2
family of Gaussian variables with mean
α . Then, by Lemma 3.5 and
Since S Ď IrrpGq is arbitrary, we can conclude that
ÿαPIrrpGq
dα
nα
trpQαpfpαqpfpαqq ă 8.
(cid:3)
Remark 3.7. Although Theorem 3.6 does not give the complete conclusion for
non-Kac cases, we can expect that the quantity
gets very close to the desired
quantity of dα. At least in certain cases, nα is significantly smaller than the quantum
dimension dα. Indeed, if the function α ÞÑ nα has subexponential growth and if G
10
dα
nα
is of non-Kac type, then the function α ÞÑ dα is of exponential growth [DPR16].
In particular, the Drinfeld-Jimbo deformations Gq with 0 ă q ă 1 are important
non-Kac type compact quantum groups. In this case, the function α ÞÑ nα is of
subexponential growth.
Let us clarify what Remark 3.7 implies for concrete example G " SUqp2q.
Then we have
Corollary 3.8. Let G " SUqp2q with 0 ă q ă 1 and suppose that all random fourier
series fU „ ÿαPIrrpGq
dαtrpUαpfpαqQαuαq are in MpGq for all U " pUαqαPIrrpGq P U.
ÿαPIrrpGq
α trpQαpfpαqpfpαqq ă 8 f or each ǫ ą 0.
Proof. It is known that IrrpGq is identified with t0u Y N and nk " k ` 1 for all
k P IrrpGq " t0u Y N. Also, dk " q´k ` q´k`2 ` ` qk ě q´k for all k ě 0.
Therefore, for each ǫ ą 0,
d1´ǫ
d1´ǫ
k
ÿkě0
dk
nk
dǫ
nk
k
k ` 1
q´ǫk
trpQkpfpkqpfpkqq ď ÿkě0
ď ÿkě0
p1 ´ qǫq2 ÿkě0
dk
nk
ď
1
trpQkpfpkqpfpkqq
trpQkpfpkqpfpkqq
trpQkpfpkqpfpkqq ă 8.
dk
nk
(cid:3)
3.3. A remark on central forms. While Theorem 3.6 clearly contains insufficient
points to reach the conclusion, we conjecture that the ultimate conclusion might
turn out to be positive. The reason is that if the starting measure f P MpGq has a
specific form, the so-called central form, it can offset the deficit considerably.
Throughout this subsection, we will consider a sub von Neumann algebra N
generated by tχαuαPIrrpGq in L8pGq. Then the restriction of Haar state to N is a
finite tracial state since
hpχαχβq " hpχ
Lemma 3.9. Let G be a compact quantum group satisfying
, χαq " hpχβχαq for all α, β P IrrpGq.
hpχαq ą 0.
Then there exists a universal constant K satisfying the following: for any sequence
aα2 ď 1, there exists pTiqi Ď PolpGq X K BallpL8pGqq
αχβq " δα,β " δβ,α " hpχ
αPIrrpGq
inf
β
paαqαPIrrpGq with ÿαPIrrpGq
hpTiχ
such that lim
i
αq ě aα for all α P IrrpGq.
Proof. By [Theorem 5, [LP97]], there exists a universal constant K ą 0 and for any
cαχα
such sequence paαqαPIrrpGq, there exists T P K BallpNq such that T „ ÿαPIrrpGq
with cα ě aα for all α P IrrpGq.
Lastly, since A " PolpGq X N is a unital -algebra and dense in N under the
weak -topology, Ball(Aq is dense in BallpNq by the Kaplansky's density theorem.
Thus, we obtain a net pTiqi Ď PolpGq X K BallpL8pGqq satisfying
cα " hpT χ
αq " lim
i
hpTiχ
αq ě aα
for all α P IrrpGq.
11
(cid:3)
αPIrrpGq
hpχαq ą 0
Theorem 3.10. Let G be a compact quantum group such that
inf
ÿαPIrrpGq
MpGq for all U " pUαqαPIrrpGq P U. Then we have
dαtrpUαpfpαqQαuαq are in
aud suppose that all random Fourier series fU „ ÿαPIrrpGq
trppfpαqq2 ă 8.
Proof. Let paαqαPIrrpGq be a sequence with ÿαPIrrpGq
aα2 ď 1. Then we have a fam-
ily tTiui Ď PolpGq X K BallpL8pGqq coming from Lemma 3.9. Let us write Ti
as ÿαPIrrpGq
Idnα for all α P IrrpGq. Therefore, using
αχα. Then pTipαqQα "
Proposition 2.3, there exists a universal constant K 1 ą 0 such that
ci
α
dα
ci
8 ą K 1 ě sup
" sup
U
U
U
ci
sup
" sup
i yL1pGq,L8pGq
i xfU , T
sup
i ÿαPIrrpGq
dαtrpUαpfpαqQαpTipαqq
sup
i ÿαPIrrpGq
αtrpUαpfpαqq
α trppfpαqq
i ÿαPIrrpGqci
cα trppfpαqq
ě ÿαPIrrpGq
aα trppfpαqq.
ě ÿαPIrrpGq
" sup
The conclusion can be reached as stated above since paαq is arbitrary.
inf
Corollary 3.11. Let G be a compact quantum group such that
(cid:3)
hpχαq ą 0
cαχα P MpGq. If all random Fourier series
αPIrrpGq
and fix a central element f „ ÿαPIrrpGq
fU „ ÿαPIrrpGq
have
dαtrpUαpfpαqQαuαq are in MpGq for all U " pUαqαPIrrpGq P U, then we
ÿαPIrrpGq
dαtrpQαpfpαqpfpαqq " ÿαPIrrpGq
cα2 ă 8.
Proposition 3.12. The quantum groups listed in the following list are known to
satisfy the condition of Corollary 3.11:
‚ Free orthogonal quantum groups O`
‚ Quantum automorphism group GautpB, ψq with a δ-form ψ
‚ Drinfeld-Jimbo q deformations Gq with 0 ă q ă 1
F
Proof. Let G1 and G2 be compact quantum groups. If there exists a bijective map
Φ : IrrpG1q Ñ IrrpG2q such that
Φpπ1 b π2q " Φpπ1q b Φpπ2q and Φp'n
i"1πiq " 'n
i"1Φpπiq
for all πi P IrrpG1q and n ě 1, then the map Φ extends to a bijective -homomorphism
Φ : N1 Ñ N2 where Nj is the weak -closure of span χj
each j " 1, 2. Also, recall that the Haar states hj are tracial on Nj respectiely.
α(αPIrrpGj q in L8pGjq for
12
Moreover, Φ is trace-preserving by repeating the proofs of [Proposition 6.7, [Wan17]]
and [Lemma 4.7, [You18]]. Therefore, we can conclude that
On the other hand, in [Pri75], it is shown that
αq " h2pχ2
h1pχ1
Φpαqq for all α P IrrpGq.
inf
πPIrrpGq}χπ}L1pGq ą 0
for any compact connected Lie group.
The existence of the bijective map Φ : IrrpGq Ñ IrrpGq is known for pG, Gq "
N , SUp2qq, pGautpB, ψq, SOp3qq and pGq, Gq. Refer to [Bra13], [NT13] and [Tim08]
pO`
for details.
(cid:3)
4. Application to complete representability problem
In the category of Banach algebras, C-algebras are in a special position. One
of typical examples that are not C-algebras is the convolution algebra L1pGq of
a locally compact group. Moreover, it turned out that L1pGq is isomorphic to a
closed subalgebra of BpHq for some Hilbert H as Banach algebras if and only if G
is finite. This fact is based on the study of the Arens irregularity for convolution
algebras L1pGq [You73], which shows L1pGq is Arens regular only if G is finite.
In the framework of locally compact quantum groups, the Fourier algebra ApGq
is understood as a dual object of the convolution algebra L1pGq in view of the
generalized Pontrjagin duality. Very recently, it has shown that the Fourier algebra
is completely isomorphic to a closed subalgebra of BpHq for some Hilbert space if
and only if G is finite ([LY17] for discrete groups and [LSS16] for general cases).
Here, complete isomorphism is the natural isomorphism in the category of operator
spaces. Some terminologies of operator space theory required in this section will be
explained in the below.
To my knowledge, prior to [LY17], all conclusions in this direction were based
on Arens regularity studies. In the quantum group setting, the situation is similar.
Thanks to [HNR12] investigating Arens irregularity for convolution algebras L1pGq
of locally compact quantum groups, we know that the convolution algebra L1pGq is
completely isomorphic to a closed subalgebra of BpHq if and only if L8pGq is finite
dimensional for G a co-amenable compact quantum group.
Our contribution in this direction is the following fact, which removes the addi-
tional condition of co-amenability in the category of compact quantum groups.
Theorem 4.1. Let G " pA, ∆q be a compact quantum group. Then L1pGq is
completely representable as an operator algebra, i.e. L1pGq is completely isomorphic
to a closed subalgebra of BpHq for some Hilbert space H if and only if A is finite
dimensional.
An abstarct operator space is a vector space E equipped with a matrical norm
structure on MnpEq " Mn b E for which
(1) }v1 ' v2}Mm`npEq ď maxt}v1}m ,}v2}nu for all v1 P MmpEq and v2 P
for all a P Mm,n, b P Mn,m and
(2) }avb}MmpEq ď }a}Mm,n }v}MnpEq }b}Mn,m
MnpV q and
v P MnpEq.
We say that a linear map L : E Ñ F between the two operator spaces E and F
is completely bounded if sup
nPN }idn b L : MnpEq Ñ MnpFq} ă 8.
The two operator spaces that are noted in this section are the convolution algebra
L1pGq and the extended Haagerup tensor product space L8pGq beh L8pGq. Each
13
spaces has a natural operator space structure. Let us present equivalent description
of those operator spaces.
First of all, for F " pfi,jqn
i,j"1 P MnpL1pGqq, the natural matricial norm structure
is given by
}F}MnpL1pGqq " sup
mPN
sup
X ››pxfi,j, xk,lyL1pGq,L8pGqqn
i,j"1
,
m
k,l"1››Mmn
where the supremum runs over all X " pxk,lqm
identification MmpL8pGqq Ď MmpBpL2pGqqq -- BpL2pGq ' ' L2pGqq.
k,l"1 P BallpMmpL8pGqqq under the
Secondly,
ai b bi : ÿi
aia
i and ÿi
i bi converges+
b
with respect to the weak -topologies and the matricial norm structure comes from
the completely isometric embedding into CBσpBpL2pGqq, BpL2pGqqq. More pre-
i,j"1 P Mn b pL8pGq beh L8pGqq is
ai,j
k b bi,j
k qn
L8pGq beh L8pGq "#ÿi
cisely, the matricial norm for pÿk
s,t"1›››››pÿk
given by
T "pT s,tqm
sup
mPN
sup
ai,j
k T s,tbi,jqn
i,j"1
m
s,t"1›››››MmnpBpL2pGqqq
,
where the supremum runs over all T " pT s,tq1ďs,tďm P BallpMmpBpL2pGqqqq.
Note that the restricted comultiplication ∆PolpGq
∆ : L8pGq Ñ L8pGqbL8pGq and it induces the natural Banach algebraic structure
on L1pGq. For f1, f2 P L1pGq, we define the convolution product of f1 and f2 by
extends to a normal -homomorphism
xf1 f2, ayL1pGq,L8pGq " pf1 b f1qp∆paqq for all a P L8pGq.
Moreover, the convolution product extends to a completely contractive map m "
category of operator spaces. Refer to [Pis03] or [ER00] for details. The result of
[Ble95], which actually covers general completely contractive Banach algebras, is
written as follows in our setting:
∆ : L1pGqpbL1pGq Ñ L1pGq where pb is the projective tensor product in the
Proposition 4.2. The convolution algebra L1pGq is completely isomorphic to a
closed subaglebra of BpHq for some Hilbert space H if and only if the comultiplica-
tion ∆ : L8pGq Ñ L8pGq beh L8pGq is completely bounded.
Lemma 4.3. Given B " pBαqαPIrrpGq P Ballpℓ8ppGqq, define an operator TB P
p,jqpBαqp,i. Then TB is a
α puαqBαsj,i "
j,i ÞÑ rQ´1
α qj,jpuα
nαÿp"1pQ´1
BpL2pGqq by uα
contraction.
Proof. Take v " ÿαPIrrpGq
nαÿi,j"1
TBv " ÿαPIrrpGq
cα
i,j uα
j,i. Then
nαÿj,p"1r
nαÿi"1
i,jpQ´1
cα
α qj,jpBαqp,ispuα
p,jq.
14
Also, put Cα " pcα
i,jq1ďi,jďnα . Then
}TBv}2
L2pGq " ÿαPIrrpGq
" ÿαPIrrpGq
" ÿαPIrrpGq
ď ÿαPIrrpGq
" ÿαPIrrpGq
nαÿj,p"1
nαÿi"1
nαÿj,p,i,i1"1pQ´1
trpQ´1
α C
α B
i,jpQ´1
cα
α qj,jpBαqp,i2pQαqj,j
dα
α qj,j cα
i,jpBαqp,icα
i1,jpBαqp,i1
αBαCαq
1
dα
α Cαq
1
dα
α C
trpQ´1
nαÿi,j"1pQ´1
α qj,jcα
i,j2 1
dα " }v}2
L2pGq .
1
dα
(cid:3)
Proof of Theorem 4.1. One direction is trivial. For the other direction, suppose
that ∆ : L8pGq Ñ L8pGq beh L8pGq is completely bounded. Then, for any
j,i P PolpGq,
f " ÿαPIrrpGq
nαÿk"1
nαÿi,j"1
dαppfpαqQαqi,j uα
j,k b uα
nαÿi,j"1
dαppfpαqQαqi,j uα
∆pfq " ÿαPIrrpGq
For each pBαqαPIrrpGq P Ballpℓ8ppGqq, pick TB of Lemma 4.3. Then p∆pfqqpTBq
x :" ÿαPIrrpGq
k,i P L8pGq beh L8pGq.
α qk,kpBαqp,iuα
nαÿi,j,k,p"1
p,kq P L2pGq.
dαppfpαqQαqi,jpQ´1
j,kpuα
sends 1A P L2pGq to
Then
}∆}cb }f}L8pGq ě }x}L2pGq
ě hpxq
" ÿαPIrrpGq
" ÿαPIrrpGq
1
dα pBαqj,i
nαÿi,j,k"1
dαppfpαqQαqi,j
nαtrppfpαqQαBαq.
nα
p
mined by
Since Bα is arbitrarily chosen, we have a bounded map Φ : CrpGq Ñ ℓ1ppGq, f ÞÑ
dα pfpαqqαPIrrpGq and it induces the bounded dual map Φ : ℓ8ppGq Ñ MpGq deter-
xf, ΦpAqyCrpGq,MpGq " ÿαPIrrpGq
IdnαqQαuαq, is bounded since for any f P PolpGq we have
Now, we can find that the map Ψ : ℓ8ppGq Ñ MpGq, A " pApαqqαPIrrpGq ÞÑ µA „
ÿαPIrrpGq
nαtrppfpαqQαApαqq.
dαtrpAαp
nα
dα
xf , ΨpAqyCrpGq,MpGq " xf, ΦpAqyCrpGq,MpGq
by Lemma 2.2.
15
Finally, Proposition 2.3 and Theorem 3.6 say that
ÿαPIrrpGq
dα
nα
trpQα
n2
α
d2
α
Idnαq " ÿαPIrrpGq
nα ă 8,
so that A should be finite dimensional.
(cid:3)
Acknowledgement. The author is grateful to Professor Hun Hee Lee for his
encouragement and to Professor Gilles Pisier for his helpful comments, particularly
on Remark 3.3.
References
[Ble95]
David P Blecher. A completely bounded characterization of operator algebras. Mathe-
matische Annalen, 303(1):227 -- 239, 1995.
[Bra13] Michael Brannan. Reduced operator algebras of trace-perserving quantum automor-
phism groups. Doc. Math., 18:1349 -- 1402, 2013.
[Cas13] Martijn Caspers. The lp -fourier transform on locally compact quantum groups.
J.Operator Theory, 2013.
[Daw10] Matthew Daws. Operator biprojectivity of compact quantum groups. Proceedings of
the American Mathematical Society, 138(4):1349 -- 1359, 2010.
[DPR16] Alessandro D'Andrea, Claudia Pinzari, and Stefano Rossi. Polynomial growth for com-
pact quantum groups, topological dimension and -regularity of the fourier algebra.
Preprint, arXiv:1602.07496, 2016.
[Dud67] Richard M Dudley. The sizes of compact subsets of hilbert space and continuity of
[ER00]
[Fac87]
[Fer75]
gaussian processes. Journal of Functional Analysis, 1(3):290 -- 330, 1967.
Edward G.. Effros and Zhong-Jin Ruan. Operator spaces. Clarendon Press, 2000.
Thierry Fack. Type and cotype inequalities for non commutative l p-spaces. Journal of
Operator Theory, pages 255 -- 279, 1987.
Xavier Fernique. Des resultats nouveaux sur les processus gaussients. In S´eminaire de
Probabilit´es IX Universit´e de Strasbourg, pages 318 -- 335. Springer, 1975.
[FHL`17] Uwe Franz, Guixiang Hong, Francois Lemeux, Michael Ulrich, and Haonan Zhang.
Hypercontractivity of heat semigroups on free quantum groups. Journal of Operator
Theory, 77(1):61 -- 76, 2017.
[FTR66] Alessandro Fig`a-Talamanca and Daniel Rider. A theorem of littlewood and lacunary
[Hel57]
series for compact groups. Pacific Journal of Mathematics, 16(3):505 -- 514, 1966.
Sigurdur Helgason. Topologies of group algebras and a theorem of littlewood. Transac-
tions of the American Mathematical Society, 86(2):269 -- 283, 1957.
[HNR12] Zhiguo Hu, Matthias Neufang, and Zhong-Jin Ruan. Module maps over locally compact
quantum groups. Studia Mathematica, 211:111 -- 145, 2012.
[Hun51] GA Hunt. Random fourier transforms. Transactions of the American Mathematical
Society, 71(1):38 -- 69, 1951.
[JMP14] Marius Junge, Tao Mei, and Javier Parcet. Smooth fourier multipliers on group von
neumann algebras. Geom. Funct. Anal, 24(6):1913 -- 1980, 2014.
[JPPP17] Marius Junge, Carlos Palazuelos, Javier Parcet, and Mathilde Perrin. Hypercontractiv-
[Kah93]
[Kus01]
[KV00]
[KV03]
[Kwa76]
[Lit26]
[LP86]
[LP97]
ity in group von neumann algebras. 2017.
Jean-Pierre Kahane. Some random series of functions, volume 5. Cambridge University
Press, 1993.
Johan Kustermans. Locally compact quantum groups in the universal setting. Internat.
J. Math., 12(3):289 -- 338, 2001.
Johan Kustermans and Stefaan Vaes. Locally compact quantum groups. Ann. Sci. ´Ecole
Norm. Sup. (4), 33(6):837 -- 934, 2000.
Johan Kustermans and Stefaan Vaes. Locally compact quantum groups in the von
Neumann algebraic setting. Math. Scand., 92(1):68 -- 92, 2003.
S Kwapie´n. A theorem on the rademacher series with vector valued coefficients. Prob-
ability in Banach spaces, pages 157 -- 158, 1976.
JE Littlewood. On the mean values of power series. Proceedings of the London Mathe-
matical Society, 2(1):328 -- 337, 1926.
Fran¸coise Lust-Piquard. In´egalit´es de khintchine dans cp (1¡ p¡). CR Acad. Sci. Paris,
303:289 -- 292, 1986.
Fran¸coise Lust-Piquard. On the coefficient problem:
a version of the kahane --
katznelson -- de leeuw theorem for spaces of matrices. journal of functional analysis,
149(2):352 -- 376, 1997.
16
[LPP91] Fran¸coise Lust-Piquard and Gilles Pisier. Non commutative khintchine and paley in-
[LSS16]
[LT13]
[LY17]
equalities. Arkiv for matematik, 29(1):241 -- 260, 1991.
Hun Hee Lee, Ebrahim Samei, and Nico Spronk. Similarity degree of Fourier algebras.
J. Funct. Anal., 271(3):593 -- 609, 2016.
Michel Ledoux and Michel Talagrand. Probability in Banach Spaces: isoperimetry and
processes. Springer Science & Business Media, 2013.
Hun Hee Lee and Sang-Gyun Youn. New deformations of convolution algebras and
fourier algebras on locally compact groups. Canad.J.Math., 69(2):434 -- 452, 2017.
[Mar78] Michael B Marcus. Continuity and the central limit theorem for random trigonometric
series. Probability Theory and Related Fields, 42(1):35 -- 56, 1978.
[MP78] Michael B Marcus and Gilles Pisier. Necessary and sufficient conditions for the uniform
convergence of random trigonometric series. 1978.
[NT13]
[MP81] Michael B Marcus and Gilles Pisier. Random Fourier Series with Applications to Har-
monic Analysis.(AM-101), volume 101 of Annals of Mathematics Studies. Princeton
University Press; University of Tokyo Press, 1981.
Sergey Neshveyev and Lars Tuset. Compact quantum groups and their representation
categories, volume 20 of Cours Sp´ecialis´es [Specialized Courses]. Soci´et´e Math´ematique
de France, Paris, 2013.
G Pisier. Sur l'espace de banach des s´eries de fourier al´eatoires presque surement con-
tinues. S´eminaire Analyse fonctionnelle (dit, pages 1 -- 33, 1977.
Gilles Pisier. Ensembles de sidon et processus gaussiens. CR Acad. Sci. Paris,
286(15):A671 -- A674, 1978.
Gilles Pisier. Introduction to operator space theory, volume 294 of London Mathematical
Society Lecture Note Series. Cambridge University Press, Cambridge, 2003.
Gilles Pisier. Grothendiecks theorem, past and present. Bulletin of the American Math-
ematical Society, 49(2):237 -- 323, 2012.
JF Price. On local central lacunary sets for compact lie groups. Monatshefte fur Math-
ematik, 80(3):201 -- 204, 1975.
REAC Paley and A Zygmund. On some series of functions,(3). In Mathematical Pro-
ceedings of the Cambridge Philosophical Society, volume 28, pages 190 -- 205. Cambridge
University Press, 1932.
Daniel Rider. Randomly continuous functions and sidon sets. Duke Math. J, 42(4):759 --
764, 1975.
Raphael Salem and Antoni Zygmund. Some properties of trigonometric series whose
terms have random signs. Acta Mathematica, 91(1):245 -- 301, 1954.
[Rid75]
[Pri75]
[PZ32]
[Pis77]
[Pis78]
[Pis03]
[Pis12]
[SZ54]
[Tim08] Thomas Timmermann. An invitation to quantum groups and duality. EMS Textbooks
in Mathematics. European Mathematical Society (EMS), Zurich, 2008. From Hopf al-
gebras to multiplicative unitaries and beyond.
N Tomczak-Jaegermann. On the rademacher averages and the moduli of convexity and
smoothness of the schatten classes sp. Studia Math, 50:163 -- 182, 1974.
Simeng Wang. Lacunary fourier series for compact quantum groups. Communications
in Mathematical Physics, 349(3):895 -- 945, 2017.
[Wan17]
[TJ74]
[Wor87a] Stanis law L. Woronowicz. Compact matrix pseudogroups. Comm. Math. Phys.,
111(4):613 -- 665, 1987.
[Wor87b] Stanis law L. Woronowicz. Twisted SUp2q group. An example of a noncommutative
[You73]
[You18]
differential calculus. Publ. Res. Inst. Math. Sci., 23(1):117 -- 181, 1987.
NJ Young. The irregularity of multiplication in group algebras. The Quarterly Journal
of Mathematics, 24(1):59 -- 62, 1973.
Sang-Gyun Youn. Hardy-littlewood inequalities on compact quantum groups of Kac
type. Analysis & PDE, 11(1):237 -- 261, 2018. DOI 10.2140/apde.2018.11.237.
Sang-Gyun Youn : Department of Mathematical Sciences, Seoul National University,
San56-1 Shinrim-dong Kwanak-gu, Seoul 151-747, Republic of Korea
E-mail address: [email protected]
17
|
1208.5939 | 2 | 1208 | 2013-06-11T14:08:19 | Approximation properties for noncommutative Lp-spaces associated with lattices in Lie groups | [
"math.OA",
"math.FA",
"math.GR"
] | In 2010, Lafforgue and de la Salle gave examples of noncommutative Lp-spaces without the operator space approximation property (OAP) and, hence, without the completely bounded approximation property (CBAP). To this purpose, they introduced the property of completely bounded approximation by Schur multipliers on Sp and proved that for p < 4/3 and p > 4 the groups SL(n,Z), with n \geq 3, do not have it. Since for 1 < p < \infty the property of completely bounded approximation by Schur multipliers on Sp is weaker than the approximation property of Haagerup and Kraus (AP), these groups were also the first examples of exact groups without the AP. Recently, Haagerup and the author proved that also the group Sp(2,R) does not have the AP, without using the property of completely bounded approximation by Schur multipliers on Sp. In this paper, we prove that Sp(2,R) does not have the property of completely bounded approximation by Schur multipliers on Sp for p < 12/11 and p > 12. It follows that a large class of noncommutative Lp-spaces does not have the OAP or CBAP. | math.OA | math |
APPROXIMATION PROPERTIES FOR NONCOMMUTATIVE
Lp-SPACES ASSOCIATED WITH LATTICES IN LIE GROUPS
TIM DE LAAT
Abstract. In 2010, Lafforgue and de la Salle gave examples of noncommu-
tative Lp-spaces without the operator space approximation property (OAP)
and, hence, without the completely bounded approximation property (CBAP).
To this purpose, they introduced the property of completely bounded approx-
imation by Schur multipliers on S p, denoted APSchur
p,cb , and proved that for
p ∈ [1, 4
3 ) ∪ (4, ∞] the groups SL(n, Z), with n ≥ 3, do not have the APSchur
p,cb .
Since for p ∈ (1, ∞) the APSchur
p,cb is weaker than the approximation property of
Haagerup and Kraus (AP), these groups were also the first examples of exact
groups without the AP. Recently, Haagerup and the author proved that also the
group Sp(2, R) does not have the AP, without using the APSchur
p,cb . In this pa-
p,cb for p ∈ [1, 12
per, we prove that Sp(2, R) does not have the APSchur
11 ) ∪ (12, ∞].
It follows that a large class of noncommutative Lp-spaces does not have the
OAP or CBAP.
1. Introduction
Let M be a finite von Neumann algebra with normal faithful trace τ . For 1 ≤ p <
∞, the noncommutative Lp-space Lp(M, τ ) is defined as the completion of M with
p , and for p = ∞, we put L∞(M, τ ) = M
respect to the norm kxkp = τ ((x∗x)
with operator norm. In [23], Kosaki showed that noncommutative Lp-spaces can
be realized by interpolating between M and L1(M, τ ). This leads to an operator
space structure on them, as described by Pisier [27] (see also [20]).
2 )
p
1
An operator space E is said to have the completely bounded approximation
property (CBAP) if there exists a net Fα of finite-rank maps on E such that
supα kFαkcb < C for some C > 0, and limα kFαx − xk = 0 for every x ∈ E.
The infimum of all possible C's is denoted by Λ(E).
If Λ(E) = 1, we say that
E has the completely contractive approximation property (CCAP). An operator
space E is said to have the operator space approximation property (OAP) if there
exists a net Fα of finite-rank maps on E such that limα k(idK(ℓ2) ⊗Fα)x − xk = 0
for all x ∈ K(ℓ2) ⊗min E. Here K(ℓ2) denotes the space of compact operators on
the Hilbert space ℓ2. The CBAP goes back to De Canni`ere and Haagerup [5], and
the OAP was defined by Effros and Ruan [9]. By definition, the CCAP implies the
CBAP, which in turn implies the OAP.
Recall that a lattice in a Lie group G is a discrete subgroup Γ of G such that G/Γ
has finite invariant measure. In this paper, we consider noncommutative Lp-spaces
of the form Lp(L(Γ)), where L(Γ) is the group von Neumann algebra of a lattice Γ
in a connected simple Lie group G. Such a von Neumann algebra L(Γ) is finite and
2010 Mathematics Subject Classification: Primary: 46B28; Secondary: 22D25, 46L07.
The author is supported by the Danish National Research Foundation through the Centre for
Symmetry and Deformation.
1
2
TIM DE LAAT
has canonical trace τ : x 7→ hxδ1, δ1i, where δ1 ∈ ℓ2(Γ) is the characteristic function
of the unit element 1 ∈ Γ.
It was proved by Junge and Ruan [20, Proposition 3.5] that if Γ is a weakly
amenable (countable) discrete group, then for p ∈ (1,∞), the noncommutative Lp-
space Lp(L(Γ)) has the CBAP. Recall that connected simple Lie groups of real rank
zero are amenable. By the work of Cowling and Haagerup [6] and Hansen [17], all
connected simple Lie groups of real rank one are weakly amenable. This implies
that for every p ∈ (1,∞) and every lattice Γ in a connected simple Lie group G of
real rank zero or one, the noncommutative Lp-space Lp(L(Γ)) has the CBAP.
The existence of noncommutative Lp-spaces without the CBAP follows from the
work of Szankowski [29]. The first concrete examples were given recently by Laf-
forgue and de la Salle [24]. They proved that for all p ∈ [1, 4
3 )∪(4,∞] and all lattices
Γ in SL(n, R), where n ≥ 3, the space Lp(L(Γ)) does not have the OAP (or CBAP).
They also proved analogous results for lattices in Lie groups over nonarchimedean
fields. In their work, the failure of the OAP for the aforementioned noncommu-
tative Lp-spaces follows from the failure of a certain approximation property for
the groups SL(n, R). This property, called the property of completely bounded
approximation by Schur multipliers on Sp (see Section 2.6), denoted APSchur
p,cb , was
introduced by Lafforgue and de la Salle exactly to this purpose.
p,cb
is weaker than the AP. In this way, the APSchur
Other approximation properties for groups (see [3]), e.g., amenability, weak
amenability, and the approximation property of Haagerup and Kraus (AP) (see
[14]), are related to the APSchur
p,cb . It is well-known that amenability of a group G
(strictly) implies weak amenability, which in turn (strictly) implies the AP. For
p ∈ (1,∞), the APSchur
p,cb gave rise
to the first example of an exact group without the AP, namely SL(3, Z). Recently,
Haagerup and the author proved that also Sp(2, R) does not have the AP [15], in
a more direct way than Lafforgue and de la Salle did for SL(3, R).
Indeed, the
APSchur
p,cb was not used in the proof. On the other hand, as was mentioned earlier,
the method of Lafforgue and de la Salle also gives information about approximation
properties of certain noncommutative Lp-spaces. For this, it is actually crucial to
use the APSchur
p,cb . Haagerup and the author also proved that all connected simple
Lie groups with finite center and real rank greater than or equal to two do not have
the AP, building on the failure of the AP for both SL(3, R) and Sp(2, R).
The following are the main results of this article.
11 ) ∪ (12,∞], the group Sp(2, R) does not have the
Theorem 3.1. For p ∈ [1, 12
APSchur
p,cb .
Theorem 4.3. Let p ∈ [1, 12
11 ) ∪ (12,∞], and let Γ be a lattice in a connected
simple Lie group with finite center and real rank greater than or equal to two.
Then Lp(L(Γ)) does not have OAP (or CBAP).
The paper is organized as follows.
In Section 2, we recall some preliminary
results, and we make a study of Schur multipliers on Schatten classes corresponding
to (compact) Gelfand pairs, which provides us with suitable tools for our proof. In
Section 3, we prove Theorem 3.1, and in Section 4, we prove Theorem 4.3.
2. Preliminaries
2.1. Schur multipliers on Schatten classes. This section partly follows the
exposition of [24, Section 1]. More details can be found there.
APPROXIMATION PROPERTIES FOR NONCOMMUTATIVE Lp-SPACES
3
p
2 )
1
n = Sp(ℓ2
For p ∈ [1,∞] and a (separable) Hilbert space H, let Sp(H) denote the pth
Schatten class on H. Recall that S∞(H) is the Banach space K(H) of compact
operators (with operator norm) on H, and for p ∈ [1,∞), the space Sp(H) consists
of the operators T on H such that kTkp = Tr((T ∗T )
p < ∞, where Tr denotes the
(semifinite) trace on B(H). In this way, Sp(H) is a Banach space for all p ∈ [1,∞].
n) and Sp = Sp(ℓ2). Note that the space S2(H)
We use the notation Sp
corresponds to the Hilbert-Schmidt operators on H.
Schatten classes can be realized by interpolating between certain noncommuta-
tive Lp-spaces in the semifinite setting. Indeed, we have Sp(H) = Lp(B(H), Tr).
Noncommutative Lp-spaces in the semifinite setting can be defined analogously to
the finite case, which was described in Section 1. For details, see [28]. The natural
operator space structure on Sp(H) follows from [27]. For our purposes, the following
characterization of the completely bounded norm of a linear map between Schatten
classes is important. Recall that Sp(H)⊗ Sp(K) (algebraic tensor product) embeds
naturally into Sp(H⊗K) (Hilbert space tensor product). Let T : Sp(H) −→ Sp(H)
be a bounded linear map, and let K = ℓ2. Then T is completely bounded if
the map T ⊗ idSp extends to a bounded linear map on Sp(H ⊗ ℓ2), and we have
kTkcb = kT ⊗ idSp k = supn∈N kT ⊗ idSp
n k (see [28, Lemma 1.7]).
A linear map T : Mn(C) −→ Mn(C) of the form [xij ] 7→ [ψij xij ] for some matrix
ψ ∈ Mn(C) is called a Schur multiplier on Mn(C). More precisely, the operator T
is called the Schur multiplier on Mn(C) with symbol ψ, and it is also denoted by
Mψ. In what follows, we need more general notions of Schur multipliers.
Let (X, µ) be a σ-finite measure space. Let k ∈ L2(X × X, µ ⊗ µ).
It is
well-known that the map Tk : L2(X, µ) −→ L2(X, µ) defined by (Tkf )(x) =
is a Hilbert-Schmidt operator on L2(X, µ). Conversely, if
RX k(x, y)f (y)dµ(y),
T ∈ S2(L2(X, µ)), then T = Tk for some k ∈ L2(X × X, µ ⊗ µ).
In this way,
we can identify S2(L2(X, µ)) with L2(X × X, µ ⊗ µ), and we see that every Schur
multiplier on S2(L2(X, µ)) comes from a function ψ ∈ L∞(X × X, µ⊗ µ) acting by
multiplication on L2(X × X, µ ⊗ µ).
Definition 2.1. Let p ∈ [1,∞], and let ψ ∈ L∞(X×X, µ⊗µ). The Schur multiplier
with symbol ψ is said to be bounded (resp. completely bounded) on Sp(L2(X, µ))
if it maps Sp(L2(X, µ)) ∩ S2(L2(X, µ)) into Sp(L2(X, µ)) (by Tk 7→ Tψk), and if
this map extends (necessarily uniquely) to a bounded (resp. completely bounded)
map Mψ on Sp(L2(X, µ)).
The norm of such a bounded multiplier ψ is defined by kψkM Sp(L2(X,µ)) = kMψk,
and its completely bounded norm by kψkcbM Sp(L2(X,µ)) = kMψkcb. The spaces
of multipliers and completely bounded multipliers are denoted by M Sp(L2(X, µ))
and cbM Sp(L2(X, µ)), respectively. It follows that for every p ∈ [1,∞] and ψ ∈
L∞(X × X, µ ⊗ µ), we have kψk∞ ≤ kψkM Sp(L2(X,µ)) ≤ kψkcbM Sp(L2(X,µ)).
q = 1, we have kψkM Sp(L2(X,µ)) = kψkM Sq(L2(X,µ)). By interpolation
and duality we have that whenever 2 ≤ p ≤ q ≤ ∞, then kψkM Sp(L2(X,µ)) ≤
kψkM Sq(L2(X,µ)). These results also hold for the completely bounded norm.
p + 1
If 1
Lemma 2.2. ([24, Lemma 1.5 and Remark 1.6]) The Schur multiplier correspond-
ing to ψ ∈ L∞(X × X, µ⊗ µ) is completely bounded on Sp(L2(X, µ)) if and only if
the Schur multiplier corresponding to ψ(x, ξ, y, η) = ψ(x, y) is completely bounded
4
TIM DE LAAT
on Sp(L2(X × Ω, µ ⊗ ν)), where (Ω, ν) is a σ-finite measure space, and
kψkcbM Sp(L2(X,µ)) = k ψkcbM Sp(L2(X×Ω,µ⊗ν)).
If L2(Ω, ν) is infinite-dimensional, these norms equal k ψkM Sp(L2(X×Ω,µ⊗ν)).
Lemma 2.3. ([24, Theorem 1.19]) Let (X, µ) be a locally compact space with a
σ-finite Radon measure µ, and let ψ : X × X −→ C be a bounded continuous
function. Let 1 ≤ p ≤ ∞. The following are equivalent:
(1) we have ψ ∈ M Sp(L2(X, µ)) with kψkM Sp(L2(X,µ)) ≤ C,
(2) for every finite set F = {x1, . . . , xn} ⊂ X such that F ⊂ supp(µ), the
Schur multiplier given by (ψ(xi, xj ))i,j is bounded on Sp(ℓ2(F )) with norm
smaller than or equal to C.
In particular,
The analogous statement holds in the completely bounded case.
the norm and the completely bounded norm of the multiplier only depend on the
support of µ, and if this support does not have any isolated points, then the norm
and the completely bounded norm coincide.
2.2. Schur multipliers on locally compact groups. For a locally compact
group G and a function ϕ ∈ L∞(G), we define the function ϕ ∈ L∞(G × G)
by ϕ(g, h) = ϕ(g−1h). The notation ϕ will be used without further mentioning. In
what follows, we will consider continuous functions ϕ : G −→ C such that ϕ is a
(completely bounded) Schur multiplier on Sp(L2(G)).
2.3. KAK decomposition for Lie groups. Recall that every connected semi-
simple Lie group G with finite center can be decomposed as G = KAK, where K
is a maximal compact subgroup (unique up to conjugation) and A is an abelian
Lie group such that its Lie algebra a is a Cartan subspace of the Lie algebra g of
G. The dimension of a is called the real rank of G and is denoted by RankR(G).
The KAK decomposition is in general not unique. However, after choosing a set of
positive roots and restricting to the closure A+ of the positive Weyl chamber A+,
we still have G = KA+K. Moreover, if g = k1ak2, where k1, k2 ∈ K and a ∈ A+,
then a is unique. For more details, see [18], [21].
2.4. Gelfand pairs and spherical functions. Let G be a Lie group with compact
subgroup K. We denote the (left) Haar measure on G by dx and the normalized
Haar measure on K by dk. A function ϕ : G −→ C is said to be K-bi-invariant
if ϕ(k1gk2) = ϕ(g) for all g ∈ G and k1, k2 ∈ K. Note that for ϕ ∈ C(G),
the continuous function defined by ϕK(g) =RKRK ϕ(kgk′)dkdk′ is K-bi-invariant.
By abuse of notation, we denote the space of K-bi-invariant compactly supported
continuous functions on G by Cc(K\G/K). This space can be considered as a
subalgebra of the convolution algebra Cc(G). If this subalgebra is commutative,
then the pair (G, K) is said to be a Gelfand pair. Equivalently, if G is a Lie group
with compact subgroup K, then (G, K) is a Gelfand pair if and only if for every
irreducible unitary representation π of G on a Hilbert space Hπ, the space Hπe
consisting of K-invariant vectors, i.e., Hπe = {ξ ∈ H ∀k ∈ K : π(k)ξ = ξ}, is at
most one-dimensional. Also, the pair (G, K) is a Gelfand pair if and only if the
representation L2(G/K) is multiplicity free.
Let (G, K) be a Gelfand pair. A function h ∈ C(K\G/K) is called a spherical
function if the functional χ on Cc(K\G/K) given by χ(ϕ) = RG ϕ(x)h(x−1)dx
defines a nontrivial character, i.e., χ(ϕ ∗ ψ) = χ(ϕ)χ(ψ) for all ϕ, ψ ∈ Cc(K\G/
APPROXIMATION PROPERTIES FOR NONCOMMUTATIVE Lp-SPACES
5
K). Spherical functions arise as the matrix coefficients of K-invariant vectors in
irreducible representations of G.
It is possible to consider Gelfand pairs in more general settings than Lie groups,
e.g., in the setting of locally compact groups (see [7],[11]).
2.5. Schur multipliers on compact Gelfand pairs. Let G and K be Lie groups
such that (G, K) is a Gelfand pair, and let X = G/K denote the homogeneous
space (with quotient topology) corresponding with the canonical (transitive) action
of G.
It follows that K is the stabilizer subgroup of a certain element e0 ∈ X.
In this section we consider Schur multipliers on the Schatten classes Sp(H), where
H = L2(G) or L2(X). To this end, it is natural to look at multipliers on G that are
K-bi-invariant. Denote by D the space K\G/K as a topological space, and denote
by f : K\G/K −→ D, KgK 7→ ξ the corresponding homeomorphism. It follows
that every function ϕ in C(K\G/K) induces a continuous function ϕ0 on D such
that ϕ(g) = ϕ0(ξ) for all g ∈ G, where ξ is the image under the homeomorphism f .
A Gelfand pair (G, K) is called compact if G is a compact group. In this section,
all Gelfand pairs are assumed to be compact, unless explicitly stated otherwise.
For compact groups every representation on a Hilbert space is equivalent to a
unitary representation, every irreducible representation is finite-dimensional, and
every unitary representation is the direct sum of irreducible ones. For an irreducible
unitary representation π of G on a Hilbert space Hπ, let Pπ =RK π(k)dk denote the
projection onto Hπe (see Section 2.4), and let GK denote the space of equivalence
classes of the irreducible unitary representations π of G such that Pπ 6= 0.
Lemma 2.4. Let (G, K) be a compact Gelfand pair, and let X = G/K be the
corresponding (compact) homogeneous space. Then
L2(X) = ⊕π∈ GKHπ.
Let hπ denote the spherical function corresponding to the equivalence class π of
representations. Then for every ϕ ∈ L2(K\G/K) we have
This lemma follows from the Peter-Weyl theorem applied to a compact homo-
geneous space (see, e.g., [19, Section V.4]). The decomposition of ϕ (and hence ϕ0)
is stated explicitly in [32, Proposition 9.10.4].
Lemma 2.5. Let (G, K) be a (not necessarily compact) Gelfand pair, and let
X = G/K denote the corresponding homogeneous space. Choose e0 ∈ X so that
K is its stabilizer subgroup. Let ϕ ∈ C(K\G/K). Then there exists a continuous
function ψ : X × X −→ C such that for all g, h ∈ G,
ϕ(g−1h) = ψ(ge0, he0).
Proof. If ge0 = g′e0 for g, g′ ∈ G, then g−1g′ ∈ K, and hence g′ = gk for some
k ∈ K. Hence, by the K-bi-invariance of ϕ, we know that ϕ(g−1h) depends only on
the pair (ge0, he0) ∈ X × X, so there exists a function ψ : X × X −→ C such that
ϕ(g−1h) = ψ(ge0, he0). Since X = G/K is equipped with the quotient topology,
this function is continuous.
(cid:3)
ϕ = Xπ∈ GK
where cπ = hϕ, hπi. Moreover, denoting by h0
responding to hπ, we have ϕ0 =Pπ∈ GK
cπ dimHπhπ,
cπ(dimHπ)h0
π.
π the (spherical) function on D cor-
6
TIM DE LAAT
Lemma 2.6. Let (G, K) be a compact Gelfand pair. If ϕ : G −→ C is a continuous
K-bi-invariant function such that ϕ ∈ cbM Sp(L2(G)) (see Section 2.2) for some
p ∈ [1,∞], then kψkcbM Sp(L2(X)) = k ϕkcbM Sp(L2(G)), where ψ : X × X −→ C is
as defined in Lemma 2.5. If K is an infinite group, then these norms are equal to
k ϕkM Sp(L2(G)).
Proof. By [25, Lemma 1.1], the quotient map G −→ G/K has a Borel cross section.
Let Y denote the image of this cross section. The result now follows directly from
Lemma 2.2 by putting Ω = K, so that G = Y × K as a measure space by the map
(y, k) 7→ yk for y ∈ Y and k ∈ K.
(cid:3)
We can now prove a decomposition result for Schur multipliers on Sp(L2(G))
coming from K-bi-invariant functions.
Proposition 2.7. Let (G, K) be a compact Gelfand pair, suppose that K has
infinitely many elements, and let p ∈ [1,∞). Let ϕ : G −→ C be a continuous
K-bi-invariant function such that ϕ ∈ M Sp(L2(G)). Then
1
p
Xπ∈ GK
cπp(dimHπ)
≤ k ϕkM Sp(L2(G)),
1
It follows that kT1kSp(L2(X)) = 1.
where cπ and Hπ are as in Lemma 2.4.
Proof. As before, let (Tkf )(x) = RG k(x, y)f (y)dy. Then T1 is the projection on
C1 ∈ L2(X).
It is sufficient to prove that
(Pπ∈ GK cπp(dimHπ))
p ≤ kTψkSp(L2(X)), where ψ is as before. Indeed, we have
kTψkSp(L2(X)) = kTψkSp(L2 (X))
kT1kSp(L2 (X)) ≤ kψkM Sp(L2(X)), which is smaller than or equal to
kψkcbM Sp(L2(X)) = k ϕkM Sp(L2(G)) by Lemma 2.6 under the assumption that K is
an infinite group.
By Lemma 2.4, we have ϕ = Pπ∈ GK
cπ dim Hπhπ. By [19, Theorem V.4.3],
it follows that the operator PHπ = dimHπTh′
π is the projection onto Hπ, where
h′π : X × X −→ C denotes the function induced by hπ (see Lemma 2.5). Since
L2(X) decomposes as a direct sum of Hilbert spaces, we have
kTψkp
Sp(L2(X)) =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xπ∈ GK
= Xπ∈ GK
p
π(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
cπ dimHπTh′
cπp Tr(PHπp) = Xπ∈ GK
Sp(L2(X))
cπp dimHπ.
(cid:3)
Lemma 2.8. Let G be a locally compact group with compact subgroup K. For
p ∈ [1,∞], let ϕ ∈ C(G) be such that ϕ ∈ M Sp(L2(G)). Then the continuous
function ϕK defined by ϕK(g) = RKRK ϕ(kgk′)dkdk′ induces an element ϕK of
M Sp(L2(G)), and k ϕKkM Sp(L2(G)) ≤ k ϕkM Sp(L2(G)). The analogous statement
holds in the completely bounded case.
Proof. Let νn be a sequence of finitely supported probability measures on K point-
wise converging to the Haar measure µ. Let ϕn : G −→ C be defined by ϕn(g) =
RKRK ϕ(kgk′)dνn(k)dνn(k′). Each ϕn is a convex combination of functions kϕk′ of
the form kϕk′ (g) = ϕ(kgk′), where k, k′ ∈ K are fixed. Hence, ϕK is an element
APPROXIMATION PROPERTIES FOR NONCOMMUTATIVE Lp-SPACES
7
of the pointwise closure of conv{kϕk′ k, k′ ∈ K}. One easily checks that for all
k, k′ ∈ K, we have kk ϕk′kM Sp(L2(G)) = k ϕkM Sp(L2(G)). Hence, by Lemma 2.3, we
have ϕK ∈ M Sp(L2(G)), and k ϕKkM Sp(L2(G)) ≤ k ϕkM Sp(L2(G)). The result for
the completely bounded case follows in an analogous way.
(cid:3)
2.6. The property APSchur
p,cb . In this section we recall the definition of the APSchur
p,cb ,
as given by Lafforgue and de la Salle in [24]. First, recall that the Fourier algebra
A(G) (see [10]) consists of the coefficients of the left-regular representation of G.
More precisely, ϕ ∈ A(G) if and only if there exist ξ, η ∈ L2(G) such that for all
x ∈ G we have ϕ(x) = hλ(x)ξ, ηi. With the norm kϕkA(G) = min{kξkkηk ∀x ∈
G ϕ(x) = hλ(x)ξ, ηi}, it is a Banach space.
Definition 2.9. ([24, Definition 2.2]) Let G be a locally compact Hausdorff second
countable group, and let 1 ≤ p ≤ ∞. The group G is said to have the property of
completely bounded approximation by Schur multipliers on Sp, denoted APSchur
p,cb ,
if there exists a constant C > 0 and a net ϕα ∈ A(G) such that ϕα → 1 uniformly
on compacta and supα k ϕαkcbM Sp(L2(G)) ≤ C. The infimum of these C's is denoted
by ΛSchur
p,cb (G).
The following result is a key property of the APSchur
p,cb
(see [24, Theorem 2.5]).
Theorem 2.10. Let G be a locally compact Hausdorff group, and let Γ be a lattice
in G. Then for 1 ≤ p ≤ ∞, we have ΛSchur
p,cb (Γ) = ΛSchur
p,cb (G).
Lafforgue and de la Salle also proved that for a discrete group Γ and p ∈ (1,∞),
it follows that ΛSchur
p,cb (Γ) ∈ {1,∞}. Since a semisimple Lie group G has lattices [1],
we conclude by the above proposition that for such a group, it also follows that
ΛSchur
p,cb (G) ∈ {1,∞} for p ∈ (1,∞).
Proposition 2.11. Let G be a locally compact Hausdorff group. The APSchur
p,cb
satisfies the following properties:
(1) for p = ∞ (or p = 1, by the third statement of this proposition), the group G
p,cb (G) = Λ(G),
p,cb if and only if it is weakly amenable, and ΛSchur
has the APSchur
where Λ(G) denotes the Cowling-Haagerup constant of G;
p,cb (G) = ΛSchur
q,cb (G);
2,cb (G) = 1;
q = 1, then ΛSchur
q,cb (G);
(2) for every locally compact group, ΛSchur
(3) if p, q ∈ [1,∞] such that 1
p + 1
(4) if 2 ≤ p ≤ q ≤ ∞, then ΛSchur
p,cb (G) ≤ ΛSchur
(5) if H is a closed subgroup of G and 1 ≤ p ≤ ∞, then ΛSchur
p,cb (G);
(6) if G has a compact subgroup K, and if ϕα is a net in A(G) converging to
1 uniformly on compacta such that supα k ϕαkcbM Sp(L2(G)) ≤ C, then there
exists a net ϕα in A(G)∩C(K\G/K) such that supα k ϕαkcbM Sp(L2(G)) ≤ C
that converges to 1 uniformly on compacta.
p,cb (G) =
(7) if K is a compact normal subgroup of G and 1 ≤ p ≤ ∞, then ΛSchur
p,cb (H) ≤ ΛSchur
ΛSchur
p,cb (G/K);
(8) if G1 and G2 are locally isomorphic connected (semi)simple Lie groups with
finite centers, then for p ∈ [1,∞], we have ΛSchur
p,cb (G1) = ΛSchur
p,cb (G2);
Proof. The first statement is clear. The second through the fifth statement are
covered in [24, Section 2]. The sixth statement follows from Lemma 2.8. By com-
bining the sixth statement and Lemma 2.6, the seventh statement follows. The fact
that the net on the group converges uniformly on compacta if and only if the net
8
TIM DE LAAT
on the quotient does, is straightforward (see [6]). For the eighth statement, note
that the center is a normal subgroup of a group. Using the seventh statement and
the fact that the adjoint groups G1/Z(G1) and G2/Z(G2), where Z(Gi) denotes
the center of Gi, are isomorphic, we obtain the result.
(cid:3)
2.7. Approximation properties for noncommutative Lp-spaces. The oper-
ator space structure on a noncommutative Lp-space Lp(M, τ ) can be obtained by
considering this space as a certain interpolation space (see [23]). Indeed, the pair
of spaces (M, L1(M, τ )) becomes a compatible couple of operator spaces, and for
1 < p < ∞ we have the isometry Lp(M, τ ) ∼= [M, L1(M, τ )] 1
. By [28, Lemma 1.7],
we know that for a linear map T : Lp(M, τ ) −→ Lp(M, τ ), its completely bounded
n[Lp(M )]k. Us-
norm kTkcb corresponds to supn∈N k idSp
ing [28, Corollary 1.4] and the fact that S1
n ⊗ L1(M ) = L1(M ⊗ Mn), we obtain
that Sp
n[Lp(M )] = Lp(M ⊗ Mn), which implies that kTkcb = supn∈N kT ⊗ id :
Lp(M ⊗ Mn) −→ Lp(M ⊗ Mn)k.
In Section 1 of this article, we recalled the definition of the CBAP, CCAP and
OAP. It was shown by Junge and Ruan [20] that if Γ is a discrete group with the
AP (of Haagerup and Kraus), and if p ∈ (1,∞), then Lp(L(Γ)) has the OAP, where
L(Γ) denotes the group von Neumann algebra of Γ. Lafforgue and de la Salle related
the AP for groups and the OAP for noncommutative Lp-spaces to the APSchur
p,cb .
n[Lp(M )] −→ Sp
p
n ⊗ T : Sp
Lemma 2.12. ([24, Corollary 3.12]) If Γ is a countable discrete group with the
AP, and if p ∈ (1,∞), then ΛSchur
Lemma 2.13. ([24, Corollary 3.13]) If p ∈ (1,∞) and Γ is a countable discrete
group such that Lp(L(Γ)) has the OAP, then ΛSchur
p,cb (Γ) = 1.
p,cb (Γ) = 1.
One of the main results of Lafforgue and de la Salle is the following.
Theorem 2.14. ([24, Theorem E]) Let n ≥ 3. For p ∈ [1, 4
group SL(n, R) does not have the APSchur
p,cb .
3 ) ∪ (4,∞], the (exact)
As a consequence, the group SL(n, R) does not have the AP, and for p ∈ [1, 4
3 )∪
(4,∞] and a lattice Γ in SL(n, R), the noncommutative Lp-space Lp(L(Γ)) does not
have the OAP or CBAP.
3. The group Sp(2, R)
In this section, we prove the following theorem. The proof is along the same
lines as the proof of the failure of the AP for Sp(2, R) in [15] (and for some details
we will refer to that article), but obtaining sufficiently sharp estimates for Schur
multipliers on Schatten classes is technically more involved.
Theorem 3.1. For p ∈ [1, 12
APSchur
p,cb .
11 ) ∪ (12,∞], the group Sp(2, R) does not have the
In this section, we write G = Sp(2, R). Recall that G is defined as the Lie group
where
G := {g ∈ GL(4, R) gtJg = J},
J =(cid:18) 0
−I2
I2
0 (cid:19) .
APPROXIMATION PROPERTIES FOR NONCOMMUTATIVE Lp-SPACES
9
Here I2 denotes the 2 × 2 identity matrix. The maximal compact subgroup K of G
is isomorphic to U(2) and explicitly given by
K =(cid:26)(cid:18) A −B
A + iB ∈ U(2)(cid:27).
Let A+ = {D(α1, α2) = diag(eα1, eα2 , e−α1, e−α2) α1 ≥ α2 ≥ 0}.
G = KA+K.
B A (cid:19) ∈ M4(R)(cid:12)(cid:12)(cid:12)(cid:12)
It follows that
p,cb
For p = 1 and ∞, the APSchur
is equivalent to weak amenability (as mentioned
in Proposition 2.11), and the failure of weak amenability for G was proved in [13].
Therefore, we can restrict ourselves to the case p ∈ (1,∞). As follows from Pro-
position 2.11, it suffices to consider approximating nets consisting of K-bi-invariant
functions. The following result gives a certain asymptotic behaviour of continuous
K-bi-invariant functions ϕ for which the induced function ϕ is a Schur multiplier
on Sp(L2(G)). From this, it follows that the constant function 1 cannot be approx-
imated pointwise (and hence not uniformly on compacta) by a K-bi-invariant net
in A(G) in such a way that the net of associated multipliers is uniformly bounded
in the M Sp(L2(G))-norm. This implies Theorem 3.1.
Proposition 3.2. Let p > 12. There exist constants C1(p), C2(p) (depending on
p only) such that for all ϕ ∈ C(K\G/K) for which ϕ ∈ M Sp(L2(G)), the limit
ϕ∞ = limkαk→∞ ϕ(D(α1, α2)) exists, and for all α1 ≥ α2 ≥ 0,
ϕ(D(α1, α2)) − ϕ∞ ≤ C1(p)k ϕkM Sp(L2(G))e−C2(p)kak2 ,
1 + α2
2.
where kαk2 =pα2
Remark 3.3. Note that Proposition 3.2 is stated in terms of the M Sp(L2(G))-
norm rather than the cbM Sp(L2(G))-norm. However, we have k.kM Sp(L2(G)) ≤
k.kcbM Sp(L2(G)), which shows that Proposition 3.2 is indeed sufficient to prove The-
orem 3.1. Moreover, by [24, Theorem 1.18], the claims are equivalent for non-
discrete groups.
For the proof of Proposition 3.2, we will identify two Gelfand pairs in G and
describe certain properties of their spherical functions.
Consider the group U(2), which contains the circle group U(1) as a subgroup via
the embedding
U(1) ֒→(cid:18) 1
0 U(1) (cid:19) ⊂ U(2).
0
Let K1 denote the copy of U(1) in G under the identification of U(2) with K. It
goes back to Weyl [31] that (U(2), U(1)) is a Gelfand pair (see, e.g., [21, Theorem
IX.9.14]). The homogeneous space U(2)/ U(1) is homeomorphic to the complex
1-sphere S1
C ⊂ C2 and the double coset space U(1)\ U(2)/ U(1) is homeomorphic to
the closed unit disc D ⊂ C by the map
U(1)(cid:18) u11 u12
u21 u22 (cid:19) U(1) 7→ u11.
The spherical functions for (U(2), U(1)) can be found in [22]. By the homeomorph-
ism U(1)\ U(2)/ U(1) ∼= D, they can be considered as functions of one complex
variable in the closed unit disc. They are indexed by the integers l, m ≥ 0 and
explicitly given by
hl,m(cid:18) u11 u12
u21 u22 (cid:19) = h0
l,m(u11),
10
TIM DE LAAT
where in the point z ∈ D, the function h0
l,m(z) =( zl−mP (0,l−m)
zm−lP (0,m−l)
h0
m
l
l,m is explicitly given by
(2z2 − 1)
(2z2 − 1)
l ≥ m,
l < m.
n
Here P (α,β)
denotes the nth Jacobi polynomial. These spherical functions satisfy
a certain Holder continuity condition, as is stated in the following lemma (see [15,
Corollary 3.5]). The proof of this Lemma makes use of recent results by Haagerup
and Schlichtkrull [16].
Lemma 3.4. For all l, m ≥ 0, and for θ1, θ2 ∈ [0, 2π), we have
Here C > 0 is a uniform constant. Combining the two, we get
≤ C(l + m + 1)
3
4 θ1 − θ2,
≤ 2C(l + m + 1)− 1
4 .
≤ 2
3
4 Cθ1 − θ2
1
4 .
h0
h0
(cid:12)(cid:12)(cid:12)(cid:12)
l,m(cid:18) eiθ1
√2(cid:19) − h0
l,m(cid:18) eiθ1
(cid:12)(cid:12)(cid:12)(cid:12)
√2(cid:19) − h0
l,m(cid:18) eiθ1
(cid:12)(cid:12)(cid:12)(cid:12)
√2(cid:19) − h0
ϕ(u) = ϕ(cid:18) u11 u12
l,m(cid:18) eiθ2
√2(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)
l,m(cid:18) eiθ2
√2(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)
l,m(cid:18) eiθ2
√2(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)
u21 u22 (cid:19) = ϕ0(u11),
h0
Let ϕ : U(2) −→ C be a U(1)-bi-invariant continuous function. Then
u ∈ U(2), u11 ∈ D,
for some continuous function ϕ0 : D −→ C. By Lemma 2.4, we know that L2(X) =
⊕l,m≥0Hl,m, where X = U(2)/ U(1) ∼= S C
1 . It is known that dimHl,m = l + m + 1,
so, by Proposition 2.7, we get
ϕ0 =
cl,m(l + m + 1)h0
l,m,
∞
Xl,m=0
as above by ϕ(g, h) = ϕ(g−1h).
for certain cl,m ∈ C. Moreover, by the same proposition, we obtain that if p ∈
(1,∞), then (Pl,m≥0 cl,mp(l + m + 1))
p ≤ k ϕkM Sp(L2(U(2))), where ϕ is defined
Lemma 3.5. Let p > 12, and let ϕ : U(2) −→ C be a continuous U(1)-bi-invariant
function such that ϕ is an element of M Sp(L2(U(2))). Then ϕ0 satisfies
1
1
2p
8− 3
p + 1
≤ C(p)k ϕkM Sp(L2(U(2)))θ1 − θ2
√2(cid:19) − ϕ0(cid:18) eiθ2
ϕ0(cid:18) eiθ1
√2(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)(cid:12)
for θ1, θ2 ∈ [0, 2π). Here, C(p) is a constant depending only on p.
Proof. Let p, q ∈ (1,∞) be such that 1
√2(cid:19) − ϕ0(cid:18) eiθ2
ϕ0(cid:18) eiθ1
√2(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)
cl,m(l + m + 1)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)(cid:12)
= Xl,m≥0
q
≤
cl,mq(l + m + 1)
(l + m + 1)(cid:12)(cid:12)(cid:12)(cid:12)
Xl,m≥0
Xl,m≥0
≤ k ϕkM Sq(L2(U(2)))
l,m(cid:18) eiθ1
(l + m + 1)(cid:12)(cid:12)(cid:12)(cid:12)
√2(cid:19) − h0
q = 1. Then for θ1, θ2 ∈ [0, 2π),
l,m(cid:18) eiθ2
l,m(cid:18) eiθ1
√2(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)
√2(cid:19) − h0
p
l,m(cid:18) eiθ2
l,m(cid:18) eiθ1
√2(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)
√2(cid:19) − h0
p
l,m(cid:18) eiθ2
√2(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)
Xl,m≥0
h0
h0
h0
1
p
.
1
1
p
APPROXIMATION PROPERTIES FOR NONCOMMUTATIVE Lp-SPACES
11
Note that k ϕkM Sq(L2(U(2))) = k ϕkM Sp(L2(U(2))). If we look at the terms of the last
sum, we get, using Lemma 3.4 and the fact that min{x, y} ≤ xεy1−ε for x, y > 0
and ε ∈ (0, 1), that
h0
l,m(cid:18) eiθ1
(l + m + 1)(cid:12)(cid:12)(cid:12)(cid:12)
√2(cid:19) − h0
≤ min{C p(l + m + 1)1+ 3
≤ 2p(1−ε)C pθ1 − θ2pε(l + m + 1)1+pε− 1
for ε ∈ (0, 1). Hence, the sum converges for 0 < ε < 1
for p > 12. Hence, if p > 12, and putting ε = 1
4 − 3
2 ( 1
l,m(cid:18) eiθ2
√2(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)
4 pθ1 − θ2p, 2pC p(l + m + n)1− 1
4 p}
p
4 p
4 − 3
p ) = 1
p . Such an ε only exists
8 − 3
2p , then
ϕ0(cid:18) eiθ1
(cid:12)(cid:12)(cid:12)(cid:12)
√2(cid:19) − ϕ0(cid:18) eiθ2
√2(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)
for some constant C(p) depending only on p.
≤ C(p)k ϕkM Sp(L2(U(2))θ1 − θ2
1
8− 3
2p
(cid:3)
For α ∈ R consider the map K −→ G defined by k 7→ DαkDα, where Dα =
diag(eα, 1, e−α, 1).
Lemma 3.6. Let ϕ : G −→ C be a continuous K-bi-invariant function such that
ϕ ∈ M Sp(L2(G)) for some p ∈ (1,∞), and for α ∈ R, let ψα : K −→ C be defined
by ψα(k) = ϕ(DαkDα). Then ψα is K1-bi-invariant and satisfies
k ψαkM Sp(L2(U(2)) ≤ k ϕkM Sp(L2(G)).
Proof. Using the fact that the group elements Dα commute with K1, it follows that
for all k ∈ K and k1, k2 ∈ K1 ⊂ K2,
ψα(k1kk2) = ϕ(Dαk1kk2Dα) = ϕ(k1DαkDαk2) = ϕ(DαkDα) = ψα(k),
so ψα is K1-bi-invariant.
The second part follows by the fact that DαKDα is a subset of G and by applying
(cid:3)
Lemma 2.3.
From the fact that ψα is K1-bi-invariant, it follows that ψα(u) = ψ0
ψ0
α : D −→ C is a continuous function.
Suppose that α1 ≥ α2 ≥ 0, and let D(α1, α2) be as defined above. If we find an
element of the form DαkDα in KD(α1, α2)K, we can relate the value of a K-bi-
invariant multiplier ϕ to the value of the multiplier ψα that was just defined. This
only works for certain α1, α2 ≥ 0. It turns out to be sufficient to consider certain
candidates for k, namely the ones of the form
α(u11), where
(1)
u =(cid:18)
a + ib
√1 − a2 − b2
−√1 − a2 − b2
a − ib
(cid:19)
with a2 + b2 ≤ 1. For a proof of the following result, see [15, Lemma 3.9].
Lemma 3.7. Let α ≥ 0 and β ≥ γ ≥ 0. If u ∈ K is of the form (1) with respect
to the identification of K with U(2), then DαuDα ∈ KD(β, γ)K if and only if
(2)
( sinh β sinh γ = sinh2 α(1 − a2 − b2),
sinh β − sinh γ = sinh(2α)a.
12
TIM DE LAAT
Consider the second Gelfand pair sitting inside G, namely the pair of groups
(SU(2), SO(2)). Both groups are subgroups of U(2), so under the embedding
into G, they give rise to compact Lie subgroups of G. The subgroup corres-
ponding to SU(2) will be called K2, and the one corresponding to SO(2) will be
called K3. The group K3 commutes with the group generated by the elements
D′α = diag(eα, eα, e−α, e−α), where α ∈ R. The subgroup SU(2) ⊂ U(2) consists of
matrices of the form
c + id
u =(cid:18) a + ib −c + id
with a, b, c, d ∈ R such that a2 + b2 + c2 + d2 = 1.
By [4, Theorem 47.6], the pair (SU(2), SO(2)) is a Gelfand pair. This also follows
from [12, Chapter 9]. The homogeneous space SU(2)/ SO(2) is the sphere S2, and
the spherical functions on the double coset space [−1, 1] are indexed by n ≥ 0, and
given by the Legendre polynomials
a − ib (cid:19) ,
Pn(2(a2 + c2) − 1) = Pn(a2 − b2 + c2 − d2).
Note that the double cosets of SO(2) in SU(2) are labeled by a2 − b2 + c2 − d2. We
use the following estimate (see [15, Lemma 3.11]).
Lemma 3.8. For all non-negative integers n, and x, y ∈ [− 1
4
√n
,
Pn(x) − Pn(y) ≤ Pn(x) + Pn(y) ≤
2 , 1
2 ],
Combining the two, we get
Pn(x) − Pn(y) ≤(cid:12)(cid:12)(cid:12)(cid:12)
Z y
x
P ′n(t)dt(cid:12)(cid:12)(cid:12)(cid:12)
≤ 4√nx − y.
Pn(x) − Pn(y) ≤ 4x − y
1
2
for x, y ∈ [− 1
on [− 1
2 , 1
2 , 1
2 ] with exponent 1
2 .
2 ], i.e., the Legendre polynomials are uniformly Holder continuous
Let ϕ : SU(2) −→ C be a SO(2)-bi-invariant continuous function. Then
c + id a − ib (cid:19) = ϕ0(2(a2 + c2) − 1) = ϕ0(a2 − b2 + c2 − d2),
ϕ(u) = ϕ(cid:18) a + ib −c + id
where u ∈ U(2), u11 ∈ D, and where ϕ0 : D −→ C is some continuous function. By
Lemma 2.4, we know that L2(X) = ⊕n≥0Hn, where X = SU(2)/ SO(2) ∼= S2. It is
known that dimHn = 2n + 1, so, by Proposition 2.7, we get
ϕ0 =
cn(2n + 1)Pn,
∞
Xn=0
1
ϕ(g, h) = ϕ(g−1h).
for certain cn ∈ C. Moreover, by the same proposition, we obtain that if p ∈ (1,∞),
then (Pn≥0 cnp(2n + 1))
p ≤ k ϕkM Sp(L2(SU(2))), where ϕ is defined as above by
Lemma 3.9. Let p > 4, and let ϕ : SU(2) −→ C be a continuous SO(2)-bi-invariant
function such that ϕ ∈ M Sp(L2(SU(2))). Then ϕ0 satisfies
ϕ0(δ1) − ϕ0(δ2) ≤ C(p)kϕkM Sp(L2(SU(2))δ1 − δ2
2 ]. Here C(p) is a constant depending only on p.
2 , 1
1
4− 1
p
for δ1, δ2 ∈ [− 1
APPROXIMATION PROPERTIES FOR NONCOMMUTATIVE Lp-SPACES
13
Proof. Let p, q ∈ (1,∞) be such that 1
p + 1
q = 1, and let δ1, δ2 ∈ [− 1
2 , 1
2 ]. Then
1
ϕ0(δ1) − ϕ0(δ2) = Xn≥0
≤
cnq(2n + 1)
Xn≥0
≤ k ϕkM Sq(L2(SU(2)))
Xn≥0
(2n + 1)Pn(δ1) − Pn(δ2)p
cn(2n + 1)Pn(δ1) − Pn(δ2)
q
Xn≥0
(2n + 1)Pn(δ1) − Pn(δ2)p
1
p
.
1
p
Note that k ϕkM Sq(L2(SU(2))) = k ϕkM Sp(L2(SU(2))). If we look at the terms of the last
sum, we get, using Lemma 3.8 and the fact that min{x, y} ≤ xεy1−ε for x, y > 0
and ε ∈ (0, 1), that
(2n + 1)Pn(δ1) − Pn(δ2)p ≤ min{4p(2n + 1)n− p
2 , 4p(2n + 1)n
p
2 δ1 − δ2p}
≤ 4p(3n)1+pε− p
2 δ1 − δ2pε
2 − 2
p ) = 1
for ε ∈ (0, 1). Hence, the sum converges for ε ∈ (0, 1
for p > 4. Hence, if p > 4, and putting ε = 1
p ). Such an ε only exists
4 − 1
ϕ0(δ1) − ϕ0(δ2) ≤ C(p)k ϕkM SpL2(U(2))δ1 − δ2
p , we have
2 − 2
4 − 1
p ,
2 ( 1
1
where C(p) is a constant depending only on p.
(cid:3)
For α ∈ R consider the map K −→ G defined by k 7→ D′αkvD′α, where D′α =
diag(eα, eα, e−α, e−α) and v ∈ Z(K) is chosen to be the matrix in K that in the
U(2)-representation of K is given by
(3)
v = 1√2
(1 + i)
0
0
(1 + i) ! .
1√2
Given a K-bi-invariant multiplier on G, this map gives rise to a K3-bi-invariant
multiplier on K. We state the following result, but omit its proof, as it is similar
to the one of Lemma 3.6.
Lemma 3.10. Let ϕ : G −→ C be a continuous K-bi-invariant function such that
ϕ ∈ M Sp(L2(G)) for some p ∈ (1,∞), and for α ∈ R let χα : K −→ C be defined
by χα(k) = ϕ(D′αkvD′α). Then χα is K3-bi-invariant and satisfies
k χαkM Sp(L2(K)) ≤ k ϕkM Sp(L2(G)).
Consider the restriction χα = χαK2, which is a K3-bi-invariant multiplier on
α(a2− b2 + c2− d2), where u ∈ K2, and where a, b, c, d
K2. It follows that χα(u) = χ0
are as before, and k χαkM Sp(L2(K2)) ≤ k ϕkM Sp(L2(G)).
Suppose that α1 ≥ α2 ≥ 0 and let D(α1, α2) be as defined above. Again, if we
find an element of the form D′αuvD′α in KD(α1, α2)K, where now u has to be an
element of SU(2), we can relate the value of a K-bi-invariant multiplier ϕ to the
value of the multiplier χα. This again only works for certain α1, α2 ≥ 0. Consider
a general element of SU(2),
u =(cid:18) a + ib −c + id
a − ib (cid:19) ,
c + id
with a2 + b2 + c2 + d2 = 1. For a proof of the following, see [15, Lemma 3.15].
14
TIM DE LAAT
Lemma 3.11. Let α ≥ 0 and β ≥ γ ≥ 0, and let u, v ∈ K be of the form as in
(1) and (3) with respect to the identification of K with U(2). Then D′αuvD′α ∈
KD(β, γ)K if and only if
( sinh2 β + sinh2 γ = sinh2(2α),
2 sinh2(2α)r,
sinh β sinh γ = 1
where r = a2 − b2 + c2 − d2.
Now we can combine the results that we obtained for both Gelfand pairs.
Lemma 3.12. Let β ≥ γ ≥ 0. Then the equations
(4)
sinh2(2s) + sinh2 s = sinh2 β + sinh2 γ,
sinh(2t) sinh t = sinh β sinh γ
have unique solutions s = s(β, γ), t = t(β, γ) in the interval [0,∞). Moreover,
(5)
,
β
4
s ≥
γ
2
.
t ≥
A proof of this Lemma can be found in [15, Lemma 3.16].
α2
α1 = α2
(2s, s)
α1 = 2α2
(2t, t)
(β, γ)
α1
The figure above shows the relative position of (β, γ), (2s, s) and (2t, t) as in Lemma
3.13 and Lemma 3.14 below. Note that (β, γ) and (2s, s) lie on a path in the
(α1, α2)-plane of the form sinh2 α1 + sinh2 α2 = constant, and (β, γ) and (2t, t) lie
on a path of the form sinh α1 sinh α2 = constant.
Lemma 3.13. For p > 4, there exists a constant C3(p) > 0 (depending only on p)
such that whenever β ≥ γ ≥ 0 and s = s(β, γ) is chosen as in Lemma 3.12, then
for all ϕ ∈ C(K\G/K) for which ϕ ∈ M Sp(L2(G)),
ϕ(D(β, γ)) − ϕ(D(2s, s)) ≤ C3(p)e− β−γ
4 ( 1
p )k ϕkM Sp(L2(G)).
4− 1
Proof. Assume first that β − γ ≥ 8. Let α ∈ [0,∞) be the unique solution to
sinh2 β + sinh2 γ = sinh2(2α), and observe that 2α ≥ β ≥ 2, so in particular α > 0.
Define
r1 =
2 sinh β sinh γ
sinh2 β + sinh2 γ ∈ [0, 1],
APPROXIMATION PROPERTIES FOR NONCOMMUTATIVE Lp-SPACES
15
and a1 =(cid:0) 1+r1
2 (cid:1)
and let
2 . Furthermore, put
1
1
2 and b1 =(cid:0) 1−r1
2 (cid:1)
u1 =(cid:18) a1 + ib1
v = 1√2
0
0
(1 + i)
0
a1 − ib1 (cid:19) ∈ SU(2),
(1 + i) ! ,
1√2
0
as previously defined. We now have 2 sinh β sinh γ = sinh2(2α)r1, and a2
1 = r1,
so by Lemma 3.11, we have D′αu1vD′α ∈ KD(β, γ)K. Let s = s(β, γ) be as in
Lemma 3.12. Then s ≥ 0 and sinh2(2s) + sinh2 s = sinh2 β + sinh2 γ = sinh2(2α).
Put
1 − b2
0
0
r2 =
2 sinh(2s) sinh s
sinh2(2s) + sinh2 s ∈ [0, 1],
a2 − ib2 (cid:19) ∈ SU(2),
u2 =(cid:18) a2 + ib2
2 and b2 = (cid:0) 1−r2
2 (cid:1)
α(r2) ≤ C(p)r1 − r2
α(r1) − χ0
2 − b2
1
1
1
where a2 = (cid:0) 1+r2
2 (cid:1)
2 = r2, it follows again by
Lemma 3.11 that D′αu2vD′α ∈ KD(2s, s)K. Now, let χα(u) = ϕ(D′αuvD′α) for
u ∈ K2 ∼= SU(2). Then by Lemma 3.9 and Lemma 3.10, it follows that
2 . Since a2
and
χα(u1) − χα(u2) = χ0
provided that r1, r2 ≤ 1
of ϕ, we get
2 . Hence, under this assumption, using the K-bi-invariance
4 − 1
p k ϕkM Sp(L2(G)),
1
4− 1
sinh2 β
ϕ(D(β, γ)) − ϕ(D(2s, s)) ≤ C(p)r1 − r2
(6)
Note that r1 ≤ 2 sinh β sinh γ
r1 ≤ 2 eγ (1−e−2γ )
cosh s ≤ 2e−s. By Lemma 3.12, equation (5), we obtain that r2 ≤ 2e− β
2e−2 ≤ 1
have proved that
sinh β . Hence, using β ≥ γ + 8 ≥ γ, we get
sinh 2s =
4 ≤
, we
4 ≤ 2e
2 . In particular, (6) holds, and since r1 − r2 ≤ max{r1, r2} ≤ 2e
eβ (1−e−2β ) ≤ 2eγ−β. In particular, r1 ≤ 2e−8 ≤ 1
2 . Similarly, r2 ≤ 2 sinh s
p k ϕkM Sp(L2(G)).
= 2 sinh γ
γ−β
γ−β
1
4
ϕ(D(β, γ)) − ϕ(D(2s, s)) ≤ C(p)2
(7)
under the assumption that β ≥ γ + 8. If γ ≤ β < γ + 8, we get from kϕk∞ ≤
k ϕkM Sp(L2(G)) that ϕ(D(β, γ)) − ϕ(D(2s, s)) ≤ 2k ϕkM Sp(L2(G)). It follows that
p )k ϕkM Sp(L2(G))
p e
γ−β
4 ( 1
4− 1
1
4− 1
ϕ(D(β, γ)) − ϕ(D(2s, s)) ≤ C3(p)e
γ−β
4 ( 1
4− 1
p )k ϕkM Sp(L2(G))
for all (β, γ) with β ≥ γ ≥ 0, if for all p ∈ (1,∞), we put C3(p) = max{ C(p)2
1
4− 1
p , 2e
(cid:3)
1
2}.
Lemma 3.14. For p > 12, there exists a constant C4(p) > 0 (depending only on
p) such that whenever β ≥ γ ≥ 0 and t = t(β, γ) is chosen as in Lemma 3.12, then
for all ϕ ∈ C(K\G/K) for which ϕ ∈ M Sp(L2(G)),
ϕ(D(β, γ)) − ϕ(D(2t, t)) ≤ C4(p)e− γ
4 ( 1
p )k ϕkM Sp(L2(G)).
4− 3
Proof. Let β ≥ γ ≥ 0. Assume first that γ ≥ 2, and let α ≥ 0 be the unique
2 sinh2 α, and observe that α > 0,
solution in [0,∞) to the equation sinh β sinh γ = 1
because β ≥ γ ≥ 2. Put
sinh β − sinh γ
a1 =
sinh(2α)
≥ 0.
16
TIM DE LAAT
Since sinh(2α) = 2 sinh α cosh α ≥ 2 sinh2 α, we have
=
In particular, a1 ≤ 1
sinh β sinh γ = sinh2 α(1 − a2
4γ ≤ 1
.
1
4 sinh γ
sinh β
2 sinh2 α
sinh β
a1 ≤
sinh(2α) ≤
8 . Put now b1 =q 1
1 − b2
a1 − ib1 ! ∈ SU(2).
u1 = a1 + ib1
− 1√2
1√2
1 = 1
1) and sinh β − sinh γ = sinh(2α)a1. Let
1. Then 1 − a2
2 − a2
1 − b2
2 . Hence,
By Lemma 3.12, we have sinh(2t) sinh t = sinh β sinh γ = 1
By Lemma 3.7, we have Dαu1Dα ∈ KD(β, γ)K.
by (5), we have t ≥ γ
we get that the number
2 sinh2 α. Moreover,
2 ≥ 1. By replacing (β, γ) in the above calculation with (2t, t),
a2
sinh(2t) − sinh t
sinh(2α)
≥ 0,
Hence, we can put b2 =q 1
satisfies
Then
1
4 sinh 1 ≤
1
4
.
1
4 sinh t ≤
2 and
a2 ≤
2 − a2
u2 = a2 + ib2
1√2
a2 − ib2 ! .
− 1√2
sinh(2t) sinh t = sinh2 α(1 − a2
2 − b2
2),
sinh(2t) − sinh t = sinh(2α)a2,
and u2 ∈ SU(2). Hence, by Lemma 3.7, Dαu2Dα ∈ KD(2t, t)K. Put now θj =
arg(aj + ibj) = π
2 for j = 1, 2, and
since d
dt sin−1 t = 1√1−t2 ≤ √2 for t ∈ [0, 1√2
], it follows that
2 − sin−1(cid:16) aj√2(cid:17) for j = 1, 2. Since 0 ≤ aj ≤ 1
sin−1(cid:18) a1√2(cid:19) − sin−1(cid:18) a2√2(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)
θ1 − θ2 ≤(cid:12)(cid:12)(cid:12)(cid:12)
≤ a1 − a2
≤ max{a1, a2}
≤ max(cid:26) 1
≤
4 sinh t(cid:27)
4 sinh γ
1
1
,
,
4 sinh γ
2
because t ≥ γ
θ1 − θ2 ≤ e− γ
3.6, the function ψα(u) = ϕ(DαuDα), u ∈ U(2) ∼= K satisfies
2 . Since γ ≥ 2, we have sinh γ
2 . Note that aj = 1√2
2 . Hence,
eiθj for j = 1, 2, so by Lemma 3.5 and Lemma
2 (1 − e−γ) ≥ 1
4 e
2 = 1
2 e
γ
γ
(8)
ψα(u1) − ψα(u2) ≤ C(p)θ1 − θ2
4 − 3
4 ( 1
≤ C(p)e− γ
1
2p k ψαkM Sp(L2(K))
8− 3
p )k ϕkM Sp(L2(G)).
Since Dαu1Dα ∈ KD(β, γ)K and Dαu2Dα ∈ KD(2t, t)K, it follows that
ϕ(D(β, γ)) − ϕ(D(2t, t)) ≤ C(p)e− γ
4 ( 1
4− 3
p )k ϕkM Sp(L2(G))
APPROXIMATION PROPERTIES FOR NONCOMMUTATIVE Lp-SPACES
17
for all γ ≥ 2. For γ satisfying 0 < γ ≤ 2, we can instead use that kϕk∞ ≤
k ϕkM Sp(L2(G)). Hence, for all p ∈ (1,∞) putting C4(p) = max{ C(p), 2e
8}, we
obtain
1
ϕ(D(β, γ)) − ϕ(D(2t, t)) ≤ C4(p)e− γ
4 ( 1
4− 3
p )k ϕkM Sp(L2(G))
(cid:3)
for all β ≥ γ ≥ 0.
For a proof of the following lemma, see [15, Lemma 3.19].
Lemma 3.15. Let s ≥ t ≥ 0. Then the equations
(9)
sinh2 β + sinh2 γ = sinh2(2s) + sinh2 s,
sinh β sinh γ = sinh(2t) sinh t,
have a unique solution (β, γ) ∈ R2 for which β ≥ γ ≥ 0. Moreover, if 1 ≤ t ≤ s ≤ 3t
2 ,
then
(10)
β − 2s ≤ 1,
γ + 2s − 3t ≤ 1.
Lemma 3.16. For all p > 12, there exists a constant C5(p) > 0 such that whenever
s, t ≥ 0 satisfy 2 ≤ t ≤ s ≤ 6
5 t, then for all ϕ ∈ C(K\G/K) for which ϕ ∈
M Sp(L2(G)),
ϕ(D(2s, s)) − ϕ(D(2t, t)) ≤ C5(p)e− s
8 ( 1
4 − 3
p )k ϕkM Sp(L2(G)).
Proof. Choose β ≥ γ ≥ 0 as in Lemma 3.15. Then by Lemma 3.13 and Lemma
3.14, we have for p > 12,
ϕ(D(2s, s)) − ϕ(D(β, γ)) ≤ C3(p)e− β−γ
ϕ(D(2t, t)) − ϕ(D(β, γ)) ≤ C4(p)e− γ
4 ( 1
4− 1
p )k ϕkM Sp(L2(G)),
p )k ϕkM Sp(L2(G)).
4− 3
4 ( 1
Moreover, by (10),
β − γ ≥ (2s − 1) − (3t − 2s + 1) = 4s − 3t − 2 ≥ s − 2,
s − 2
2
Hence, since s ≥ 2, we have min{e−γ, e−(β−γ)} ≤ e− s−2
from Lemma 3.13 and Lemma 3.14 with C5(p) = e
γ ≥ 3t − 2s − 1 ≥
s − 2s − 1 =
5
2
.
2 . Thus, the lemma follows
(cid:3)
16 (C3(p) + C4(p)).
1
Lemma 3.17. For p > 12, there exists a constant C6(p) > 0 such that for all
ϕ ∈ C(K\G/K) for which ϕ ∈ M Sp(L2(G)), the limit c∞(ϕ) = limt→∞ ϕ(D(2t, t))
exists, and for all t ≥ 0,
ϕ(D(2t, t)) − c∞(ϕ) ≤ C6(p)e− t
8 ( 1
4− 3
p )k ϕkM Sp(L2(G)).
ϕ(D(2u, u)) − ϕ(D(2u + 2γ, u + γ)) ≤ C5(p)e− u
Proof. By Lemma 3.16, we have for u ≥ 5 and γ ∈ [0, 1], that
4− 3
(11)
since u ≤ u + γ. Let s ≥ t ≥ 5. Then s = t + n + δ, where n ≥ 0 is an integer
and δ ∈ [0, 1). Applying equation (11) to (u, γ) = (t + j, 1), j = 0, 1, . . . , n − 1 and
(u, γ) = (t + n, δ), we obtain
p )k ϕkM Sp(L2(G)),
8 ( 1
ϕ(D(2t, t)) − ϕ(D(2s, s)) ≤ C5(p)
Xj=0
≤ C5(p)′e− t
n
e− t+j
8 ( 1
4− 3
p )
k ϕkM Sp(L2(G))
8 ( 1
4− 3
p )k ϕkM Sp(L2(G)),
18
TIM DE LAAT
Therefore, c∞(ϕ) = limt→∞ ϕ(D(2t, t)) exists, and
where C′5(p) = C5(p)P∞j=0 e− j
ϕ(D(2t, t))−c∞(ϕ) = lim
8 ( 1
4− 3
p ). Hence, (ϕ(D(2t, t)))t≥5 is a Cauchy net.
s→∞ ϕ(D(2t, t))−ϕ(D(2s, s)) ≤ C′5(p)e− t
8 ( 1
4 − 3
p )k ϕkM Sp(L2(G))
for all t ≥ 5. Since kϕk∞ ≤ k ϕkM Sp(L2(G)), we have for all 0 ≤ t < 5,
ϕ(D(2t, t)) − c∞(ϕ) ≤ 2k ϕkM Sp(L2(G)).
5
Hence, the lemma follows with C6(p) = max{C′5(p), 2e
Proof of Proposition 3.2. Let ϕ ∈ C(K\G/K) be such that ϕ ∈ M Sp(L2(G)), and
let (α1, α2) = (β, γ), where β ≥ γ ≥ 0. Assume first β ≥ 2γ. Then β − γ ≥ β
2 , so
by Lemma 3.12 and Lemma 3.13, there exists an s ≥ β
4− 1
ϕ(D(β, γ)) − ϕ(D(2s, s)) ≤ C3(p)e− β
p )k ϕkM Sp(L2(G)).
4 such that
32}.
8 ( 1
(cid:3)
By Lemma 3.17,
ϕ(D(2s, s)) − c∞(ϕ) ≤ C6(p)e− s
≤ C6(p)e− β
8 ( 1
4 − 3
32 ( 1
p )k ϕkM Sp(L2(G))
4− 3
p )k ϕkM Sp(L2(G)).
Hence,
ϕ(D(β, γ)) − c∞(ϕ) ≤ (C3(p) + C6(p))e− β
32 ( 1
4− 3
p )k ϕkM Sp(L2(G)).
Assume now that β < 2γ. Then, by Lemma 3.12 and Lemma 3.14, we obtain that
there exists a t ≥ γ
4 such that
2 > β
ϕ(D(β, γ)) − ϕ(D(2t, t)) ≤ C4(p)e− β
8 ( 1
4− 3
p )k ϕkM Sp(L2(G)),
and again by Lemma 3.17,
ϕ(D(2t, t)) − c∞(ϕ) ≤ C6(p)e− t
≤ C6(p)e− β
8 ( 1
4− 3
32 ( 1
p )k ϕkM Sp(L2(G))
4− 3
p )k ϕkM Sp(L2(G)).
Hence,
ϕ(D(β, γ)) − c∞(ϕ) ≤ (C4(p) + C6(p))e− β
32 ( 1
4− 3
p )k ϕkM Sp(L2(G)).
Combining these results, and using that kαk2 = pβ2 + γ2 ≤ √2β, it follows that
for all β ≥ γ ≥ 0,
ϕ(D(β, γ)) − c∞(ϕ) ≤ C1(p)e−C2(p)kαk2k ϕkM Sp(L2(G)),
4 − 3
( 1
where C1(p) = max{C3(p) + C6(p), C4(p) + C6(p)} and C2(p) = 1
32√2
proves the proposition.
p ). This
(cid:3)
The values p ∈ [1, 12
11 ) ∪ (12,∞] give sufficient conditions for Sp(2, R) to fail the
p,cb . We would like to point out that the set of these values might be bigger.
APSchur
APPROXIMATION PROPERTIES FOR NONCOMMUTATIVE Lp-SPACES
19
4. Noncommutative Lp-spaces without the OAP
In the previous section we proved that Sp(2, R) does not have the APSchur
p,cb
for
11 ) ∪ (12,∞]. By Lemma 2.13, this directly implies the following theorem.
11 ) ∪ (12,∞], and let Γ be a lattice in Sp(2, R). Then
p ∈ [1, 12
Theorem 4.1. Let p ∈ [1, 12
the noncommutative Lp-space Lp(L(Γ)) does not have the OAP (or CBAP).
Combining Theorem 3.1 and Theorem 2.14, this implies the following result.
Theorem 4.2. Let p ∈ [1, 12
11 )∪ (12,∞], and let G be a connected simple Lie group
with finite center and real rank greater than or equal to two. Then G does not have
the APSchur
p,cb .
Proof. Let G be a connected simple Lie group with finite center and real rank
greater than or equal to two. By Wang's method [30], we may assume that G is
the adjoint group, so that G has a connected semisimple subgroup H with real
rank 2. Such a subgroup is closed, as was proved in [8]. It is known that H has
finite center and is locally isomorphic to either SL(3, R) or Sp(2, R) [2], [26]. Since
the APSchur
p,cb passes to closed subgroups and is preserved under local isomorphisms
(see Proposition 2.11), we conclude that G does not have the APSchur
for p ∈
p,cb
11 )∪ (12,∞], since both SL(3, R) and Sp(2, R) do not have the APSchur
[1, 12
p,cb for such
p.
(cid:3)
Combining this result with Proposition 2.10 and Lemma 2.13, we obtain the
main theorem of this article.
Theorem 4.3. Let p ∈ [1, 12
11 ) ∪ (12,∞], and let Γ be a lattice in a connected
simple Lie group with finite center and real rank greater than or equal to two.
Then Lp(L(Γ)) does not have OAP (or CBAP).
Acknowledgements
I thank Uffe Haagerup and Magdalena Musat for valuable discussions and useful
suggestions and remarks.
References
1. A. Borel, Harish-Chandra, Arithmetic subgroups of algebraic groups, Ann. of Math. (2) 76
(1962), 485 -- 535.
2. A. Borel, J. Tits, Groupes r´eductifs, Inst. Hautes ´Etudes Sci. Publ. Math. No. 27 (1965),
55 -- 150.
3. N.P. Brown, N. Ozawa, C ∗-Algebras and Finite-Dimensional Approximations, Graduate Stud-
ies in Mathematics, 88, American Mathematical Society, Providence, RI, 2008.
4. D. Bump, Lie Groups, Graduate Texts in Mathematics, 225, Springer-Verlag, New York, 2004.
5. J. de Canni`ere, U. Haagerup, Multipliers of the Fourier algebras of some simple Lie groups
and their discrete subgroups, Amer. J. Math. 107 (1985), no. 2, 455 -- 500.
6. M. Cowling, U. Haagerup, Completely bounded multipliers of the Fourier algebra of a simple
Lie group of real rank one, Invent. Math. 96 (1989), no. 3, 507 -- 549.
7. G. van Dijk, Introduction to Harmonic Analysis and Generalized Gelfand Pairs, Studies in
Mathematics, 36, de Gruyter, Berlin, 2009.
8. B. Dorofaeff, Weak amenability and semidirect products in simple Lie groups, Math. Ann.
306 (1996), no. 4, 737 -- 742.
9. E. Effros, Z.-J. Ruan, On approximation properties for operator spaces, Internat. J. Math. 1
(1990), no. 2, 163 -- 187.
20
TIM DE LAAT
10. P. Eymard, L'alg´ebre de Fourier d'un groupe localement compact, Bull. Soc. Math. France 92
(1964), 181 -- 236.
11. J. Faraut, Analyse harmonique sur les paires de Guelfand et les espaces hyperboliques, In: Ana-
lyse Harmonique, Les Cours du CIMPA, Nice, 1982, 315 -- 446.
12. J. Faraut, Analysis on Lie groups, Cambridge Studies in Advanced Mathematics, 110, Cam-
bridge University Press, Cambridge, 2008.
13. U. Haagerup, Group C ∗-algebras without the completely bounded approximation property,
unpublished manuscript (1986).
14. U. Haagerup, J. Kraus, Approximation properties for group C ∗-algebras and group von Neu-
mann algebras, Trans. Amer. Math. Soc. 344 (1994), no. 2, 667 -- 699.
15. U. Haagerup, T. de Laat, Simple Lie groups without the Approximation Property, Duke
Math. J. 162 (2013), no. 5, 925 -- 964.
16. U. Haagerup, H. Schlichtkrull, Inequalities for Jacobi polynomials, Ramanujan J. (to appear).
17. M.L. Hansen, Weak amenability of the universal covering group of SU(1,n), Math. Ann. 288
(1990), 445 -- 472.
18. S. Helgason, Differential Geometry, Lie Groups and Symmetric Spaces, Pure and Applied
Mathematics, 80, Academic Press, New York, 1978.
19. S. Helgason, Groups and Geometric Analysis, Pure and Applied Mathematics, 113, Academic
Press, Orlando, 1984.
20. M. Junge, Z.-J. Ruan, Approximation properties for noncommutative Lp-spaces associated
with discrete groups, Duke Math. J. 117 (2003), no. 2, 313 -- 341.
21. A.W. Knapp, Lie Groups Beyond an Introduction, Birkhauser, Boston, 1996.
22. T. H. Koornwinder, The addition formula for Jacobi polynomials II. The Laplace type in-
tegral representation and the product formula, Report TW 133/72, Mathematical Centre,
Amsterdam 1972.
23. H. Kosaki, Applications of the complex interpolation method to a von Neumann algebra:
non-commutative Lp-spaces, J. Funct. Anal. 56 (1984), 29 -- 78.
24. V. Lafforgue, M. de la Salle, Noncommutative Lp-spaces without the completely bounded ap-
proximation property, Duke. Math. J. 160 (2011), no. 1, 71 -- 116.
25. G.W. Mackey, Induced representations of locally compact groups. I., Ann. of Math. (2) 55
(1952), 101 -- 139.
26. G.A. Margulis, Discrete subgroups of semisimple Lie groups, Springer-Verlag, Berlin, 1991.
27. G. Pisier, The operator Hilbert space OH, complex interpolation and tensor norms,
Mem. Amer. Math. Soc. 122 (1996), no. 585.
28. G. Pisier, Non-commutative vector valued Lp-spaces and completely p-summing maps,
Ast´erisque, No. 247, Soci´et´e Math´ematique de France, Paris, 1998.
29. A. Szankowski, On the uniform approximation property in Banach spaces, Israel J. Math. 49
(1984), 343 -- 359.
30. S.P. Wang, The dual space of semi-simple Lie groups, Amer. J. Math. 91 (1969), 921 -- 937.
31. H. Weyl, The Theory of Groups and Quantum Mechanics, Methuen and Co., Ltd., London,
1931.
32. J.A. Wolf, Harmonic Analysis on Commutative Spaces, Mathematical Surveys and Mono-
graphs, no. 142, American Mathematical Society, Providence, RI, 2007.
Department of Mathematical Sciences, University of Copenhagen,
Universitetsparken 5, DK-2100 Copenhagen Ø, Denmark
E-mail address: [email protected]
|
1209.1864 | 1 | 1209 | 2012-09-10T02:00:37 | Conditions $C_p$, $C'_p$, and $C"_p$ for $p$-operator spaces | [
"math.OA"
] | Conditions $C$, $C'$, and $C"$ were introduced for operator spaces in an attempt to study local reflexivity and exactness of operator spaces (Effros and Ruan, 2000). For example, it is known that an operator space $W$ is locally reflexive if and only if $W$ satisfies condition $C"$ (Effros and Ruan, 2000) and an operator space $V$ is exact if and only if $V$ satisfies condition $C'$ (Effros and Ruan, 2000). It is also known that an operator space $V$ satisfies condition $C$ if and only if it satisfies conditions $C'$ and $C"$ (Effros and Ruan, 2000, and Han, 2007). In this paper, we define $p$-operator space analogues of these definitions, which will be called conditions $C_p$, $C'_p$, and $C"_p$, and show that a $p$-operator space on $L_p$ space satisfies condition $C_p$ if and only if it satisfies both conditions $C'_p$ and $C"_p$. The $p$-operator space injective tensor product of $p$-operator spaces will play a key role. | math.OA | math |
CONDITIONS Cp, C ′
p, AND C ′′
p FOR p-OPERATOR SPACES
JUNG-JIN LEE
Abstract. Conditions C, C ′, and C ′′ were introduced for operator spaces in an attempt to study local
reflexivity and exactness of operator spaces [ER00, Chapter 14]. For example, it is known that an
operator space W is locally reflexive if and only if W satisfies condition C ′′ [ER00, Theorem 14.3.1] and
an operator space V is exact if and only if V satisfies condition C ′ [ER00, Theorem 14.4.1]. It is also
known that an operator space V satisfies condition C if and only if it satisfies conditions C ′ and C ′′
[ER00, Lemma 14.2.1], [Han07, Theorem 5]. In this paper, we define p-operator space analogues of these
definitions, which will be called conditions Cp, C ′
p, and C ′′
p , and show that a p-operator space on Lp
space satisfies condition Cp if and only if it satisfies both conditions C ′
p and C ′′
p . The p-operator space
injective tensor product of p-operator spaces will play a key role.
1. Introduction to p-Operator Spaces
A concrete operator space V is defined to be a closed subspace of B(H), where B(H) denotes the space
of all bounded linear operators on a Hilbert space H. For each n ∈ N, the matrix algebra Mn(B(H)) with
entries in B(H) can be identified with B(H ⊕ · · · ⊕ H
) via matrix multiplication
Tij
h1
...
hn
=
}
j=1 T1jhj
n
{z
...
Pn
Pn
j=1 Tnjhj
,
[Tij] ∈ Mn(B(H)),
hj ∈ H,
and this gives rise to a norm k · kn on Mn(V ), which we denote by Mn(V ). It is then easy to verify that
= max{kukn, kvkm}.
the following two properties (called Ruan's axioms) hold:
D∞: for u ∈ Mn(V ) and v ∈ Mm(V ), we have (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
u 0
0
v
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)n+m
M: for u ∈ Mm(V ), α ∈ Mn,m(C), and β ∈ Mm,n(C), we have kαuβkn ≤ kαkkukmkβk, where kαk
is the norm of α as a member of B(ℓm
2 , ℓn
2 ), and similarly for β.
An abstract operator space is a Banach space X together with a family of norms k·kn defined on Mn(X)
satisfying the conditions D∞ and M above. In [Rua88], Ruan showed that these two concepts coincide
and after Ruan's characterization, operator space theory has really been taken off and quickly developed
Date: November 16, 2018.
The author was supported by Hutchcroft Fund, Department of Mathematics and Statistics, Mount Holyoke College.
1
2
JUNG-JIN LEE
into an active research area in modern analysis. Many important applications have been found in some
related areas. For example, let G be a locally compact group. It is well known that G is amenable if
and only if the convolution algebra L1(G) is amenable as a Banach algebra [Joh72]. We consider another
Banach algebra called the Fourier algebra A(G) which consists of all coefficient functions of the left regular
representation λ of G, i.e.,
A(G) = {ω(·) = hλ(·)ξ, ηi : ξ, η ∈ L2(G)}.
By [Eym64], A(G) is a commutative Banach algebra with respect to pointwise multiplication and can
be regarded as the predual of V N (G), the group von Neumann algebra of G. If G is abelian, then its
dual group G is also abelian and we have the isometric isomorphism A(G) ∼= L1( G), and this suggests a
relationship between the amenability of G and the amenability (as a Banach algebra) of A(G). Indeed,
if A(G) is amenable, then G is amenable. In the opposite direction, Johnson showed that the Banach
algebra A(G) fails to be amenable even in the case of very simple compact groups, such as SU (2, C)
[Joh94].
In [Rua95], Ruan studied the operator amenability of A(G) which can be regarded as the amenability
of A(G) in the category of operator spaces, and proved that a locally compact group G is amenable if and
only if A(G) is operator amenable. This suggests that A(G) is better viewed as an operator space, and
motivated by this observation, there has been some research [Daw10, ALR10] to study Fig`a-Talamanca-
Herz Algebra Ap(G), which can be regarded as an Lp space generalization of the Fourier algebra A(G) (The
reader is referred to [FT65, Her71] for more details on Ap(G)), in the framework of Lp space generalization
of operator spaces. This leads to the definition of p-operator spaces we will give below. Throughout this
paper, we let 1 < p < ∞.
Definition 1.1. Let SQp denote the collection of subspaces of quotients of Lp spaces. A Banach space
X is called a concrete p-operator space if X is a closed subspace of B(E) for some E ∈ SQp, where B(E)
denotes the space of all bounded linear operators on E.
Let Mn(X) denote the linear space of all n × n matrices with entries in X. For a concrete p-operator
space X ⊆ B(E) and for each n ∈ N, define a norm k · kn on Mn(X) by identifying Mn(X) as a subspace
of B(ℓn
p (E)), and let Mn(X) denote the corresponding normed space. The norms k · kn then satisfy
D∞: for u ∈ Mn(X) and v ∈ Mm(X), we have ku ⊕ vkn+m = max{kukn, kvkm}.
Mp: for u ∈ Mm(X), α ∈ Mn,m(C), and β ∈ Mm,n(C), we have kαuβkn ≤ kαkkukmkβk, where kαk
is the norm of α as a member of B(ℓm
p , ℓn
p ), and similarly for β.
Remark 1.2. When p = 2, these are Ruan's axioms and 2-operator spaces are simply operator spaces
because the SQ2 spaces are exactly Hilbert spaces.
As in operator spaces, we can also define abstract p-operator spaces.
CONDITIONS Cp, C ′
p, AND C ′′
p FOR p-OPERATOR SPACES
3
Definition 1.3. An abstract p-operator space is a Banach space X together with a sequence of norms
k · kn defined on Mn(X) satisfying the conditions D∞ and Mp above.
Thanks to the following theorem by Le Merdy, we do not distinguish between concrete p-operator
spaces and abstract p-operator spaces, so from now on we will merely speak of p-operator spaces.
Theorem 1.4. [LeM96, Theorem 4.1] An abstract p-operator space X can be isometrically embedded in
B(E) for some E ∈ SQp in such a way that the canonical norms on Mn(X) arising from this embedding
agree with the given norms.
Note that a linear map u : X → Y between p-operator spaces X and Y induces a map un : Mn(X) →
Mn(Y ) by applying u entrywise. We say that u is p-completely bounded if kukpcb := supn kunk < ∞.
Similarly, we define p-completely contractive, p-completely isometric, and p-completely quotient maps. We
write CBp(X, Y ) for the space of all p-completely bounded maps from X into Y , and to turn the mapping
space CBp(X, Y ) into a p-operator space, we define a norm on Mn(CBp(X, Y )) by identifying this space
with CBp(X, Mn(Y )). Using Le Merdy's theorem, one can show that CBp(X, Y ) itself is a p-operator
space. In particular, the p-operator dual space of X is defined to be CBp(X, C). The next lemma by Daws
shows that we may identify the Banach dual space X ′ of X with the p-operator dual space CBp(X, C) of
X.
Lemma 1.5. [Daw10, Lemma 4.2] Let X be a p-operator space, and let ϕ ∈ X ′, the Banach dual of X.
Then ϕ is p-completely bounded as a map to C. Moreover, kϕkpcb = kϕk.
If E = Lp(µ) for some measure µ and X ⊆ B(E) = B(Lp(µ)), then we say that X is a p-operator space
on Lp space. These p-operator spaces are often easier to work with. For example, let κX : X → X ′′
denote the canonical inclusion from a p-operator space X into its second dual. Contrary to operator
spaces, κX is not always p-completely isometric. Thanks to the following theorem by Daws, however,
we can easily characterize those p-operator spaces with the property that the canonical inclusion is p-
completely isometric.
Proposition 1.6. [Daw10, Proposition 4.4] Let X be a p-operator space. Then κX is a p-complete
contraction. Moreover, κX is a p-complete isometry if and only if X ⊆ B(Lp(µ)) p-completely isometrically
for some measure µ.
Conditions C, C ′, and C ′′ for operator spaces were introduced and studied in [ER00, Chapter 14]
and [Han07] and they play an important role in understanding local reflexivity and exactness of operator
spaces. For example, it is known that an operator space is locally reflexive if and only if it satisfies
condition C ′′ [ER00, Theorem 14.3.1]. It is also known that an operator space is exact if and only if
it satisfies condition C ′ [ER00, Theorem 14.4.1]. In this paper, we define p-operator space analogues of
4
JUNG-JIN LEE
these conditions, which will be called conditions Cp, C ′
space satisfies condition Cp if and only if it satisfies both conditions C ′
p, and C ′′
p , and show that a p-operator space on Lp
p and C ′′
p .
2. Tensor Product of p-Operator Spaces
In this section, we recall basic properties of tensor products on p-operator spaces studied in [Daw10,
ALR10]. We mainly focus on p-projective tensor product and p-injective tensor product.
Definition 2.1. Let X, Y be p-operator spaces. Let X ⊗ Y denote the algebraic tensor product of X and
Y . For u ∈ Mn(X ⊗ Y ), let
kuk∧p = inf{kαkkvkkwkkβk : u = α(v ⊗ w)β},
where the infimum is taken over r, s ∈ N, α ∈ Mn,r×s, v ∈ Mr(X), w ∈ Ms(Y ), and β ∈ Mr×s,n.
Daws defined and studied the p-projective tensor product [Daw10]. Note that k · k∧p gives the algebraic
tensor product X ⊗ Y a p-operator space structure [Daw10, Proposition 4.8]. Furthermore, k · k∧p is the
largest subcross p-operator space norm on X ⊗ Y in the sense that kx ⊗ yk ≤ kxkrkyks for all x ∈ Mr(X)
and all y ∈ Ms(Y ) [Daw10, Proposition 4.8]. The p-operator space projective tensor product is defined to
be the completion of X ⊗ Y with respect to this norm and is denoted by X
∧p
⊗ Y .
Remark 2.2.
(a) One can show that p-operator space projective tensor product is commutative, i.e., X
∧p
⊗ Y =
∧p
⊗ X p-completely isometrically.
Y
(b) By universality of the Banach space projective tensor product
π
⊗ [BLM04, A.3.3], we have
for all u ∈ X ⊗ Y .
kuk∧p ≤ kukπ
Let V, W , and Z be p-operator spaces, and let ψ : V × W → Z be a bilinear map. Define bilinear maps
ψr,s;t,u by
ψr,s;t,u : Mr,s(V ) × Mt,u(W ) → Mr×t,s×u(Z),
(v, w) 7→ (ψ(vi,j , wk,l)),
and let ψr;s = ψr,r;s,s. Finally define
kψkjpcb = sup{kψr;sk : r, s ∈ N}.
We say that ψ is jointly p-completely bounded (respectively, jointly p-completely contractive) if kψkjpcb <
∞ (respectively, kψkjpcb ≤ 1). The space of all jointly p-completely bounded maps from V × W to Z will
be denoted by CBp(V × W, Z) and this space can be turned into a p-operator space in the same way as
for CBp(V, W ). Here we collect some results on the p-projective tensor product for convenience.
CONDITIONS Cp, C ′
p, AND C ′′
p FOR p-OPERATOR SPACES
5
Proposition 2.3. [Daw10, Proposition 4.9] Let X, Y , and Z be p-operator spaces. Then we have natural
p-completely isometric identifications
CBp(X
∧p
⊗ Y, Z) = CBp(X × Y, Z) = CBp(X, CBp(Y, Z)).
In particular,
(X
∧p
⊗ Y )′ = CBp(X, Y ′).
As in operator spaces, the p-operator space projective tensor product is projective in the following
sense:
Proposition 2.4. [Daw10, Proposition 4.10] Let X, X1, Y , and Y1 be p-operator spaces. If u : X → X1
and v : Y → Y1 are p-complete quotient maps, then u ⊗ v extends to a p-complete quotient map u ⊗ v :
∧p
⊗ Y → X1
∧p
⊗ Y1.
X
We now briefly introduce the p-operator space injective tensor product.
Definition 2.5. Let X, Y be p-operator spaces. Regarding the algebraic tensor product X ⊗ Y as
a subspace of CBp(X ′, Y ), we define the p-operator space injective tensor product X
completion of X ⊗ Y in CBp(X ′, Y ).
∨p
⊗ Y to be the
To be precise, for u = [uij] ∈ Mn(X ⊗ Y ) with uij = PNij
tensor product norm kuk∨p is defined by
k=1 xij
k ⊗ yij
k , the p-operator space injective
kuk∨p =kukMn(CBp(X ′,Y )) = kukCBp(X ′,Mn(Y ))
(2.1)
= sup
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Nij
Xk=1
fst(xij
k
k )yij
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Mmn(Y )
: m ∈ N, f = [fst] ∈ Mm(X ′)1
,
where Mm(X ′)1 denotes the closed unit ball of Mm(X ′) = CBp(X, Mm).
Proposition 2.6. Suppose that X, X1, Y , and Y1 are p-operator spaces. Given p-complete contractions
ϕ : X → X1 and ψ : Y → Y1, the mapping
extends to a p-complete contraction
ϕ ⊗ ψ : X ⊗ Y → X1 ⊗ Y1
ϕ ⊗ ψ : X
∨p
⊗ Y → X1
∨p
⊗ Y1.
6
JUNG-JIN LEE
Proof. Since ϕ ⊗ ψ = (idX1 ⊗ ψ) ◦ (ϕ ⊗ idY ), it suffices to show that ϕ ⊗ idY and idX1 ⊗ ψ extend to
p-complete contractions. Let u = [uij] ∈ Mn(X ⊗ Y ). Let us write uij = PNij
Since
k xij
k ⊗ yij
k for each uij.
Nij
(ϕ ⊗ idY )n(u) =
Xk
ϕ(xij
k
k ) ⊗ yij
∈ Mn(X1 ⊗ Y ),
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Mmn(Y )
k
k ))yij
from (2.1) it follows that
k(ϕ ⊗ idY )n(u)k∨p = sup
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Nij
Xk=1
gst(ϕ(xij
: m ∈ N, g = [gst] ∈ Mm(X ′
.
1)1
Define hst = gst ◦ ϕ for 1 ≤ s, t ≤ m, then h = [hst] = g ◦ ϕ ∈ Mm(X ′)1 and we have
To show that idX1 ⊗ ψ is also p-completely contractive, let v = [vij ] ∈ Mn(X1 ⊗ Y ). Writing vij =
k(ϕ ⊗ idY )n(u)k∨p ≤ kuk∨p.
PNij
k wij
k ⊗ yij
k , we have
On the other hand,
Nij
kvk∨p = sup
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xk=1
k(idX1 ⊗ ψ)n(v)k∨p = sup
= sup
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
(2.2)
fst(wij
k
k )yij
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Mmn(Y )
: m ∈ N, f = [fst] ∈ Mm(X ′
1)1
.
: m ∈ N, f = [fst] ∈ Mm(X ′
fst(wij
k )
k )ψ(yij
Nij
Xk=1
ψmn
Nij
Xk=1
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Mmn(Y1)
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Mmn(Y1)
k
k )yij
≤kψkpcbkvk∨p.
fst(wij
: m ∈ N, f = [fst] ∈ Mm(X ′
1)1
1)1
(cid:3)
Remark 2.7.
(a) By definition of the Banach space injective tensor product
ǫ
⊗, we have
kukǫ = kukB(X ′,Y ) ≤ kukCBp(X ′,Y ) = kuk∨p
for every u ∈ X ⊗ Y .
(b) Let u ∈ Mn(X ⊗ Y ). If Y ⊆ B(Lp(ν)) for some measure ν, then by Definition 2.5 and [Daw10,
Theorem 4.3, Proposition 4.4]
kuk∨p = sup{kψ(ϕst(uij ))kMrmn : m, k ∈ N, ϕ = [ϕst] ∈ Mm(X ′)1, ψ ∈ Mk(Y ′)1}
CONDITIONS Cp, C ′
p, AND C ′′
p FOR p-OPERATOR SPACES
7
= sup{k(ϕ ⊗ ψ)n(u)k : m, k ∈ N, ϕ ∈ Mm(X ′)1, ψ ∈ Mk(Y ′)1}.
(c) Let F : X ⊗ Y → Y ⊗ X denote the "flip", that is, F (P xi ⊗ yi) = P yi ⊗ xi. If Y ⊆ B(Lp(ν)) for
some measure ν, then by (b) above, for every u ∈ Mn(X ⊗ Y ), we get
kuk∨p = sup{k(ϕ ⊗ ψ)n(u)k : m, k ∈ N, ϕ ∈ Mm(X ′)1, ψ ∈ Mk(Y ′)1}.
On the other hand, if X ⊆ B(Lp(µ)) for some measure µ as well, then
kFn(u)k∨p = sup{k(ψ ⊗ ϕ)n(Fn(u))k : m, k ∈ N, ϕ ∈ Mm(X ′)1, ψ ∈ Mk(Y ′)1}
and it follows that X
∨p
⊗ Y = Y
∨p
⊗ X p-completely isometrically.
(d) Mr
∨p
⊗ Ms is p-completely isometrically isomorphic to Mrs. This follows immediately from [ALR10,
Theorem 3.2].
At this moment, we do not know whether the p-operator space injective tensor product is injective,
that is, if u : X → X and v : Y → Y are p-completely isometric injections, then we do not know
∨p
⊗ Y . But if
whether u ⊗ v always extends to a p-completely isometric injection u ⊗ v : X
∨p
⊗ Y → X
we assume that all the p-operator spaces under consideration are on Lp space, then we can show that
∨p
⊗ Y is a p-complete isometry as in the following proposition. This fact supports
∨p
⊗ Y → X
u ⊗ v : X
that the terminology p-injective tensor product is still reasonable.
Proposition 2.8. Let µ1, µ2 be measures. For i = 1, 2, suppose Xi ⊆ Yi ⊆ B(Lp(µi)). Then
p-completely isometrically.
∨p
⊗ X2 ⊆ Y1
∨p
⊗ Y2
X1
Proof. For i = 1, 2, let ϕi : Xi ֒→ Yi denote the (p-completely isometric) inclusion. Since ϕ1 ⊗ ϕ2 =
(ϕ1 ⊗ idY2 ) ◦ (idX1 ⊗ ϕ2), by Remark 2.7 (c) above, it suffices to show that
idX1 ⊗ ϕ2 : X1
∨p
⊗ X2 → X1
∨p
⊗ Y2
is p-completely isometric. Note that the following diagram commutes:
∨p
⊗ X2
X1
idX1 ⊗ϕ2
/ X1
∨p
⊗ Y2
CBp(X ′
1, X2)
/ CBp(X ′
1, Y2)
Since X1
∨p
⊗ X2 ⊆ CBp(X ′
1, X2), X1
∨p
⊗ Y2 ⊆ CBp(X ′
isometrically, we conclude that idX1 ⊗ ϕ2 is p-completely isometric.
1, Y2), and CBp(X ′
1, X2) ⊆ CBp(X ′
1, Y2) p-completely
(cid:3)
_
/
_
/
8
JUNG-JIN LEE
3. Conditions C ′
p, C ′′
p , and Cp for p-Operator Spaces
In this section, we define conditions C ′
p, C ′′
p , and Cp for p-operator spaces and prove the main result.
Throughout the section, µ and ν will denote measures.
Lemma 3.1. Let V and W be p-operator spaces. Then the bilinear mapping
Ψ : V ′ × W ′ → (V
∨p
⊗ W )′,
(f, g) 7→ f ⊗ g
is jointly p-completely contractive and hence the canonical mapping Ψ : V ′
∧p
⊗ W ′ → (V
∨p
⊗ W )′ is p-
completely contractive.
Proof. We identify [fij] ∈ Mr(V ′) with an operator F ∈ CBp(V, Mr), and likewise [gkl] ∈ Ms(W ′) with
G ∈ CBp(W, Ms). We have the identification Mrs((V
∨p
⊗ W )′) = CBp(V
∨p
⊗ W, Mrs). Let H be the map
[fij ⊗ gkl] : V
∨p
⊗ W → Mrs. Then by Proposition 2.6 and Remark 2.7 (d) we have the commutative
diagram
∨p
⊗ W
V
F ⊗G
/ Mrs
H
;✇✇✇✇✇✇✇✇✇
∼=
∨p
⊗ Ms
Mr
with kF ⊗ Gkpcb ≤ kF kpcbkGkpcb, and it follows that k[fij ⊗ gkl]k = kHkpcb ≤ kF ⊗ Gkpcb ≤ kF kpcbkGkpcb
as required.
(cid:3)
Lemma 3.2. Let V and W be p-operator spaces. Then k · k∨p is a subcross matrix norm. In particular,
for every u ∈ Mn(V ⊗ W ), we have kuk∨p ≤ kuk∧p.
Proof. Just to fix notation, we identify Mr(V ) ⊗ Mq(W ) with Mrq(V ⊗ W ) by (vij ) ⊗ (wkl) 7→ (vij ⊗
wkl)(i,k),(j,l) where we have the ordering (1, 1) ≤ (1, 2) ≤ · · · ≤ (1, q) ≤ (2, 1) ≤ · · · ≤ (r, q). Hence Ir ⊗w ∈
Mr ⊗Mq(W ) = Mrq(W ) is identified with a block matrix in Mr(Mq(W )) which has r copies of w down the
diagonal and 0 elsewhere. Applying axiom D∞ repeatedly hence shows that kIr ⊗ wkrq = kwkq. Then, for
α ∈ Mr, the matrix α⊗w ∈ Mr⊗Mq(W ) = Mrq(W ) is the product (α⊗Iq)(Ir ⊗w) which has norm at most
∨p
⊗ W ). This
kαkrkwkq by axiom Mp. Now let v ∈ Mr(V ) and w ∈ Mq(W ), and consider v ⊗ w ∈ Mrq(V
tensor induces the operator T ∈ CBp(V ′, Mrq(W )) given by T (f ) = (f (vij )wkl)(i,k),(j,l) = (f (vij )) ⊗ w.
For f = (fab) ∈ Mn(V ′), we see that Tn(f ) = hhf, vii ⊗ w ∈ Mnrq(W ), which by the previous paragraph
has norm at most khhf, viiknrkwkq ≤ kf knkvkrkwkq. Hence kT kpcb ≤ kvkrkwkq as required.
(cid:3)
Let V and W be p-operator spaces and fix ϕ ∈ (V
∨p
⊗ W )′. For v0 ∈ V , we define a bounded linear
functional v0 ϕ on W by
v0 ϕ(w) = ϕ(v0 ⊗ w),
w ∈ W.
/
;
CONDITIONS Cp, C ′
p, AND C ′′
p FOR p-OPERATOR SPACES
9
In general, when v0 = [vij ] ∈ Mr(V ) and ϕ = [ϕkl] ∈ Mn((V
∨p
⊗ W )′), we define v0 ϕ = [vij ϕkl] ∈
Mrn(W ′). Similarly, for w0 ∈ W , we define ϕw0 ∈ V ′ by
ϕw0 (v) = ϕ(v ⊗ w0),
v ∈ V.
As in v0 ϕ above, we can extend the definition of ϕw0 for w0 ∈ Mr(W ) and ϕ ∈ Mn((V
∨p
⊗ W )′). Define
a linear map ΦR
V,W : V ⊗ W ′′ → (V
∨p
⊗ W )′′ by
ΦR
V,W (v ⊗ w′′)(ϕ) = hvϕ, w′′iW ′,W ′′ ,
v ∈ V, w′′ ∈ W ′′, ϕ ∈ (V
Similarly, define a linear map ΦL
V,W : V ′′ ⊗ W → (V
∨p
⊗ W )′′ by
ΦL
V,W (v′′ ⊗ w)(ϕ) = hϕw, v′′iV ′,V ′′ ,
v′′ ∈ V ′′, w ∈ W, ϕ ∈ (V
∨p
⊗ W )′.
∨p
⊗ W )′.
Lemma 3.3. The map ΦR
V,W ) defined above extends to a p-completely contractive
map ΦR
V,W : V
∧p
⊗ W ′′ → (V
V,W (respectively, ΦL
∨p
⊗ W )′′ (respectively, ΦL
V,W : V ′′
∧p
⊗ W → (V
∨p
⊗ W )′′).
Proof. Consider the bilinear map Φ : V × W ′′ → (V
∨p
⊗ W )′′ given by
(v, w′′) 7→ (ϕ 7→ hvϕ, w′′iW ′,W ′′ ),
then we get
and
Φr;s : Mr(V ) × Ms(W ′′) → Mrs((V
∨p
⊗ W )′′),
([vij ], [wkl
′′]) 7→ [Φ(vij , wkl
′′)]
k[Φ(vij, wkl
′′)] = sup
n (cid:26)khhΦr;s(v, w′′), ϕiik : ϕ ∈ Mn((V
∨p
⊗ W )′), kϕk ≤ 1(cid:27) .
Since hhΦr;s(v, w′′), ϕii = hhvϕ, w′′ii, we have
khhΦr;s(v, w′′), ϕiik = khhvϕ, w′′iik ≤ kvϕkMrn(W ′) · kw′′kMs(W ′′)
and the result follows because
∨p
⊗ is a subcross matrix norm and hence
kvϕkMrn(W ′) = supm {khhvϕ, wiikMrnm : w ∈ Mm(W ), kwk ≤ 1}
= supm {khhϕ, v ⊗ wiikMrnm : w ∈ Mm(W ), kwk ≤ 1}
≤ kϕk · kvk
≤ kvk.
Remark 3.4. Let α be a general subcross matrix norm.
(a) We have a natural p-complete contraction V
∧p
⊗ W → V ⊗α W and the adjoint gives a contraction
(V ⊗α W )′ → CBp(V, W ′) ⊆ B(V, W ′) given by
ϕ 7→ Lϕ,
hLϕ(v), wi = ϕ(v ⊗ w), ϕ ∈ (V ⊗α W )′
v ∈ V, w ∈ W.
(cid:3)
10
JUNG-JIN LEE
(b) Using the natural p-complete contraction V
regarded as a member in (V
∧p
⊗ W )′.
∧p
⊗ W → V ⊗α W , each member in (V ⊗α W )′ can be
(c) We can define ΦR
V,W : V ⊗ W ′′ → (V ⊗α W )′′ and ΦL
V,W : V ′′ ⊗ W → (V ⊗α W )′′ for a general
subcross norm α and Lemma 3.3 remains valid if
∨p
⊗ is replaced by ⊗α.
Let Ψ : V ′
∧p
⊗ W ′ → (V
∨p
⊗ W )′ denote the canonical map, and consider the following commutative
diagram
V ⊗ W ′′
ΦR
V,W
∨p
⊗ W )′′
/ (V
Ψ′
/ (V ′
∧p
⊗ W ′)′
,
CBσ
p,F (V ′, W ′′)
ι
/ CBp(V ′, W ′′)
where CBσ
from V ′ to W ′′ and ι denotes the inclusion map. This commutative diagram shows that ΦR
p,F (V ′, W ′′) denotes the space of all weak∗-continuous p-completely bounded finite rank maps
V,W is one-to-
one, so one can equip V ⊗ W ′′ with the p-operator space norm inherited from (V
∨p
⊗ W )′′, which will be
denoted by, following the notation in [ER00], V
p (or V has
p) if this induced norm coincides with the p-operator space injective tensor product norm for
property C ′
∨p
⊗ : W ′′. We say that V satisfies condition C ′
every W ⊆ B(Lp(ν)).
Similarly, the following diagram
V ′′ ⊗ W
ΦL
V,W
∨p
⊗ W )′′
/ (V
Ψ′
/ (V ′
∧p
⊗ W ′)′
CBσ
p,F (W ′, V ′′)
ι
/ CBp(W ′, V ′′)
is also commutative, ΦL
V,W is one-to-one, and one can hence equip V ′′ ⊗ W with the p-operator space
∨p
∨p
⊗W . We say that V satisfies condition C ′′
⊗ W )′′, which will be denoted by V ′′ :
p
p ) if this induced norm coincides with the injective tensor product norm for every
norm inherited from (V
(or V has property C ′′
W ⊆ B(Lp(ν)).
/
/
/
/
/
/
CONDITIONS Cp, C ′
p, AND C ′′
p FOR p-OPERATOR SPACES
11
In order to define condition Cp for p-operator spaces, we need the natural map from V ′′ ⊗ W ′′ to
∨p
⊗ W )′′. To do this, let α be a general subcross matrix norm on V ⊗ W and consider the diagram
(V
(3.1)
∧p
⊗ W ′′)′′
(V
(V ⊗α W )
′′′′
P
/ (V ⊗α W )′′
,
V ′′ ⊗ W ′′
ΦL
V,W ′′
8rrrrrrrrrrrr
&▲▲▲▲▲▲▲▲▲▲
V ′′ ,W &
ΦR
(ΦR
V,W )′′
&◆◆◆◆◆◆◆◆◆◆◆
8qqqqqqqqqq
V,W )′′
(ΦL
(V ′′
∧p
⊗ W )′′
where P is the restriction mapping and (ΦR
V,W )′′ and (ΦL
V,W )′′ are from Remark 3.4 (c).
Consider the following p-complete contraction:
∧p
⊗ W )′ ∼= CBp(V, W ′)
(V
adj
−−−−→ CBp(W ′′, V ′) ∼= (V
∧p
⊗ W ′′)′.
For ϕ ∈ (V
∧p
⊗ W )′, let ϕ∧ ∈ (V
∧p
⊗ W ′′)′ denote the image of ϕ under this map. Then we have
ϕ∧(v ⊗ w′′) = hvϕ, w′′iW ′,W ′′ = ΦR
V,W (v ⊗ w′′)(ϕ),
v ∈ V, w′′ ∈ W ′′.
Moreover, ϕ∧ is weak*-continuous in the second variable. Similarly, we also consider the p-complete
contraction
∧p
⊗ W )′ ∼= CBp(W, V ′)
(V
adj
−−−−→ CBp(V ′′, W ′) ∼= (V ′′
∧p
⊗ W )′
and define ∧ϕ, and then we get that
∧ϕ(v′′ ⊗ w) = hϕw, v′′iV ′,V ′′ = ΦL
V,W (v′′ ⊗ w)(ϕ),
v′′ ∈ V ′′, w ∈ W,
and that ∧ϕ is weak*-continuous in the first variable.
Remark 3.5. Let α be a general subcross matrix norm. By Remark 3.4 (b), we can still define ϕ∧ ∈
∧p
⊗ W ′′)′ for any ϕ ∈ (V ⊗α W )′. Similarly, we can define ∧ϕ ∈ (V ′′
∧p
⊗ W )′ for any ϕ ∈ (V ⊗α W )′.
(V
The next result follows by Remarks 3.4 and 3.5, and the same argument as in the proof of [Han07,
Theorem 1].
Theorem 3.6. Let V and W be p-operator spaces. Let α be a subcross matrix norm on V ⊗ W and denote
by V ⊗α W the resulting normed space. Then the following are equivalent.
(a) There exists a separately weak*-continuous extension
Φ : V ′′ ⊗ W ′′ → (V ⊗α W )′′
of the natural inclusion ι : V ⊗ W → (V ⊗α W )′′.
&
8
/
8
12
JUNG-JIN LEE
(b) The following diagram commutes
V ′′ ⊗ W ′′
ΦL
V,W ′′
8rrrrrrrrrrrr
&▲▲▲▲▲▲▲▲▲▲
V ′′ ,W &
ΦR
(ΦR
V,W )′′
&◆◆◆◆◆◆◆◆◆◆◆
8qqqqqqqqqq
V,W )′′
(ΦL
(V ⊗α W )
′′′′
P
/ (V ⊗α W )′′
.
∧p
⊗ W ′′)′′
(V
(V ′′
∧p
⊗ W )′′
(c) For every ϕ ∈ (V ⊗α W )′, two functionals (∧ϕ)∧ and ∧(ϕ∧) coincide on V ′′ ⊗ W ′′.
(d) For every ϕ ∈ (V ⊗α W )′, Lϕ : V → W ′ is weakly compact, where hLϕ(v), wi = ϕ(v ⊗ w), v ∈ V ,
w ∈ W .
Theorem 3.7. Let V ⊆ B(Lp(µ)) and W ⊆ B(Lp(ν)). For every ϕ ∈ (V
∨p
⊗ W )′, Lϕ is weakly compact,
where Lϕ is as in Theorem 3.6 (d).
Proof. Without loss of generality, we may assume kϕk(= kϕkpcb) ≤ 1. Let ΦV (respectively ΦW ) denote
the embedding ΦV : V ֒→ B(Lp(µ)) (respectively, ΦW : W ֒→ B(Lp(ν))). By Proposition 2.8 and [ALR10,
Theorem 3.2], we have p-completely isometric embeddings
∨p
⊗ W ֒→ B(Lp(µ))
∨p
⊗ B(Lp(ν)) ֒→ B(Lp(µ × ν)).
V
Consider the diagram below:
C
9s
s
s
s
s
∨p
⊗ W
V
ϕ
ΦV ⊗ΦW
B(Lp(µ))
∨p
⊗ B(Lp(ν))
s
s
s
s
B(Lp(µ × ν))
s
s
s
s
ϕ
s
s
By Hahn-Banach Theorem, ϕ extends to ϕ : B(Lp(µ × ν)) → C. Applying the same technique as in the
proof of [ALR10, Theorem 3.6], we can find a measure space (Ω, Σ, θ) together with two vectors ξ ∈ Lp(θ),
η ∈ Lp′(θ), and a unital p-completely contractive homomorphism π : B(Lp(µ × ν)) → B(Lp(θ)) such that
ϕ(·) = hπ(·)ξ, ηi.
Define T : B(Lp(µ)) → B(Lp(ν))′ by
hT (x), yi = ϕ(x ⊗ y),
x ∈ B(Lp(µ)),
y ∈ B(Lp(ν)).
&
8
/
8
/
/
_
_
9
CONDITIONS Cp, C ′
p, AND C ′′
p FOR p-OPERATOR SPACES
13
Then it is easy to check that the following diagram is commutative:
V
ΦV
Lϕ
W ′
(ΦW )′
B(Lp(µ))
T
/ B(Lp(ν))′
Define R : B(Lp(µ)) → Lp(θ) and S : B(Lp(ν)) → Lp′(θ) by
R(x) = π(x ⊗ 1)ξ,
x ∈ B(Lp(µ)),
and
S(y) = (π(1 ⊗ y))′η,
y ∈ B(Lp(ν)),
then the diagram
is commutative, because
B(Lp(µ))
T
$❏❏❏❏❏❏❏❏❏
R
Lp(θ)
B(Lp(ν))′
9ttttttttt
S ′
hS ′R(x), yi = hR(x), S(y)i = hπ(x ⊗ 1)ξ, (π(1 ⊗ y))′ηi = hπ(x ⊗ y)ξ, ηi = ϕ(x ⊗ y) = hT (x), yi.
Combining these two commutative diagrams, we finally have Lϕ = (ΦW )′S ′RΦV , that is, Lϕ is factorized
through a reflexive Banach space Lp(θ), so Lϕ is a weakly compact operator [Meg98, Propositions 3.5.4
and 3.5.11].
(cid:3)
Corollary 3.8. Let V, W be p-operator spaces on Lp space. Then there exists a (necessarily unique)
separately weak*-continuous extension
Φ : V ′′ ⊗ W ′′ → (V
∨p
⊗ W )′′
of the natural inclusion ι : V ⊗ W → (V
∨p
⊗ W )′′.
Proof. Combine Theorem 3.6 and Theorem 3.7. Uniqueness follows from separate weak*-continuity. (cid:3)
Now we are ready to define condition Cp for p-operator spaces. Let Φ be as in Corollary 3.8. The
following commutative diagram
V ′′ ⊗ W ′′
Φ
∨p
⊗ W )′′
/ (V
Ψ′
/ (V ′
∧p
⊗ W ′)′
CBσ
p,F (V ′, W ′′)
ι
/ CBp(V ′, W ′′)
shows that Φ is injective. Thus we can equip V ′′ ⊗ W ′′ with the p-operator space structure induced by Φ,
∨p
⊗ : W ′′. We say that V ⊆ B(Lp(µ)) satisfies condition Cp (or has property
which will be denoted by V ′′ :
Cp) if the map Φ is isometric with respect to the injective tensor product norm for every W ⊆ B(Lp(ν)).
/
/
_
/
O
O
/
/
$
9
/
/
/
14
JUNG-JIN LEE
Proposition 3.9. Suppose that V ⊆ B(Lp(µ)). Then V satisfies condition Cp if and only if V satisfies
both condition C ′
p and C ′′
p .
Proof. Suppose that V satisfies condition Cp and W ⊆ B(Lp(ν)). By Proposition 2.8 and [Daw10,
Theorem 4.3], we have a p-completely isometric embedding V
∨p
⊗ W ′′ ⊆ V ′′
∨p
⊗ W ′′ and the bottom row in
the following commutative diagram
∨p
⊗ : W ′′
V
∨p
⊗ W ′′
V
V ′′ :
∨p
⊗ : W ′′
/ V ′′
∨p
⊗ W ′′
is isometric. Therefore the top row is also isometric and hence V satisfies condition C ′
p. That V satisfies
condition C ′′
p can be proved using a similar argument.
On the other hand, if V satisfies condition C ′′
p , we get
V ′′
∨p
⊗ W ′′ = V ′′ :
∨p
⊗ : W ′′ ֒→ (V
∨p
⊗ W ′′)′′.
If V also satisfies condition C ′
p, then
∨p
⊗ W ′′ = V
∨p
⊗ : W ′′ ֒→ (V
∨p
⊗ W )′′,
V
and hence we have isometric inclusion
V ′′
∨p
⊗ W ′′ ֒→ (V
∨p
⊗ W )′′′′.
Since V ′′
∨p
⊗ W ′′ ⊂ (V
∨p
⊗ W )′′ and (V
∨p
⊗ W )′′ ֒→ (V
∨p
⊗ W )′′′′ isometrically, the inclusion V ′′
∨p
⊗ W ′′ ⊆ (V
∨p
⊗ W )′′
must be isometric.
(cid:3)
4. Acknowledgement
The author would like to thank the reviewer for his/her valuable comments, especially the ones that
led to Remarks 3.4 and 3.5, to improve the quality of the paper.
References
[ALR10] Guimei An, Jung-Jin Lee, and Zhong-Jin Ruan. On p-approximation properties for p-operator spaces. Journal of
Functional Analysis, 259:933 -- 974, 2010.
[BLM04] David P. Blecher and Christian Le Merdy. Operator algebras and their modules -- an operator space approach,
volume 30 of London Mathematical Society Monographs. New Series. The Clarendon Press Oxford University
Press, Oxford, 2004. Oxford Science Publications.
[Daw10] Matthew Daws. p-operator spaces and Fig`a-Talamanca-Herz algebras. J. Opeator Theory, 63:47 -- 83, 2010.
[ER00] E. Effros and Z.-J. Ruan. Operator Spaces. Oxford Science Publications, 2000.
[Eym64] P. Eymard. L'algebre de Fourier d'un groupe localement compact. Bull. Soc. Math. France, 92:181 -- 236, 1964.
/
/
_
_
/
CONDITIONS Cp, C ′
p, AND C ′′
p FOR p-OPERATOR SPACES
15
[FT65] A. Fig`a-Talamanca. Translation invariant operators in Lp. Duke Math. J., 32:495 -- 501, 1965.
[Han07] K. H. Han. An operator space approach to condition C. J. Math. Anal. Appl., 336:569 -- 576, 2007.
[Her71] C. Herz. The theory of p-spaces with an application to convolution operators. Trans. Amer. Math. Soc., 154:69 -- 82,
1971.
[Joh72] B. Johnson. Cohomology in Banach algebras. Mem. Amer. Math. Soc., 127, 1972.
[Joh94] B. Johnson. Non amenablity of the Fourier algebra of a compact group. J. London Math. Soc., 50:361 -- 374, 1994.
[LeM96] Christian LeMerdy. Factorization of p-completely bounded multilinear maps. Pacific Journal of Mathematics,
172:187 -- 213, 1996.
[Meg98] Robert E. Megginson. An Introduction to Banach Space Theory. Springer, 1998.
[Rua88] Z.-J. Ruan. Subspaces of C ∗-algebras. J. Funct. Anal., 76:217 -- 230, 1988.
[Rua95] Z.-J. Ruan. The operator amenability of A(G). Amer. J. Math., 117:1449 -- 1474, 1995.
Department of Mathematics and Statistics, Mount Holyoke College, South Hadley, MA 01075, USA
E-mail address, Jung-Jin Lee: [email protected]
|
1001.0730 | 1 | 1001 | 2010-01-05T16:20:31 | Operator algebras associated to integral domains | [
"math.OA"
] | We study operator algebras associated to integral domains. In particular, with respect to a set of natural identities we look at the possible nonselfadjoint operator algebras which encode the ring structure of an integral domain. We show that these algebras give a new class of examples of semicrossed products by discrete semigroups. We investigate the structure of these algebras together with a particular class of representations. | math.OA | math |
OPERATOR ALGEBRAS ASSOCIATED TO INTEGRAL
DOMAINS
BENTON L. DUNCAN
Abstract. We study operator algebras associated to integral do-
mains.
In particular, with respect to a set of natural identities
we look at the possible nonselfadjoint operator algebras which en-
code the ring structure of an integral domain. We show that these
algebras give a new class of examples of semicrossed products by
discrete semigroups. We investigate the structure of these algebras
together with a particular class of representations.
Recently, in [4] and [8] the notion of a regular C ∗-algebra associated
to an integral domain was introduced as a generalization of a con-
struction due to Cuntz, [3].
In these papers the authors associate a
C ∗-algebra by representing ℓ2(R) and viewing the operators given by
the regular representation R acting on ℓ2(R). In addition they show
that the C ∗-algebras so constructed are universal with respect to a col-
lection of identities that encode information about the integral domain.
Now one can view an integral domain as an additive group together
with an action on the additive group given by multiplication by nonzero
elements of the ring. This suggests that an important viewpoint for
studying operator algebras associated to integral domains is through
the use of crossed products. More importantly, since crossed products
are well understood many of the significant results can be made brief
through the technology of crossed products.
For this paper we wish to investigate the operator algebras with
slightly less restrictive identities imposed by only natural ring-theoretic
constraints. This gives rise to operator algebras with a more natural
crossed product structure. However since the multiplication in a ring
need not act as automorphisms on the additive group, crossed products
are not entirely appropriate. To avoid this we use the nonselfadjoint
operator algebras where possible. This goes back to a construction of
Arveson and Josephson [1] which was generalized by Peters in [10]. This
semicrossed product is a nonselfadjoint operator algebra which encodes
2000 Mathematics Subject Classification. 47L74, 47L40.
Key words and phrases. integral domains, semicrossed products.
1
2
BENTON L. DUNCAN
the same dynamics as the crossed product but does not require that
the action on a topological space be via homeomorphisms.
While one may worry that we lose too much information when we
lose the ∗-structure of the C ∗-algebra, in recent work [5] it was shown
that the semicrossed products of Peters in fact encode the action of a
continuous self map on a topological space in a manner which is unique
up to conjugacy of the map. This is even true when the map is not a
homeomorphism and hence unlike with C ∗-algebras the nonselfadjoint
operator algebras can be used as a topological invariant.
It is these motivating examples which have led us to study the semi-
crossed product algebras in the context of integral domains.
In this
paper we have defined the universal operator algebra associated to an
integral domain (note the different conditions we require from those of
Cuntz and Li). We then study the situation in the case that our inte-
gral domain is a field. Here the semicrossed product and the crossed
product coincide and we can use standard results for crossed products
to prove facts about the algebra. After viewing the case of the integral
domain being a field we focus on the situation where this may not be
true. Here the semicrossed product technology is necessary, however
similar results carry through. After defining the requisite notion of
semicrossed product and proving some first results in the context of in-
tegral domains we prove some results which show that the algebras thus
defined are distinct from the algebras studied by Peters. In the last
section we analyze what we call unitary representations of an integral
domain R. We show that every such representation factors through a
regular unitary representation.
We describe some standard notation we intend to use.
If R is an
integral domain we write Q(R) for the field of quotients. We write
(R, +) for the additive group on R. This group is a locally compact
an element a in (R, +).
discrete group. We denote the Pontryagin dual of this group by (dR, +)
and denote byba the element in the Pontryagin dual corresponding to
The author would like to thank Jim Coykendall and Sean Sather-
Wagstaff for helpful discussions about and examples of integral do-
mains.
1. Universal algebras of integral domains
Let R be an integral domain. Given a Hilbert space H we define an
isometric representation of R to be {Sr ∈ B(H) : r ∈ R×} a collection
of isometries together with unitaries {U n ∈ B(H) : n ∈ R} that satisfy
the relations:
OPERATOR ALGEBRAS ASSOCIATED TO INTEGRAL DOMAINS
3
(1) SrSt = Srt for all r, t ∈ R×,
(2) U nU m = U m+n for all m, n ∈ R, and
(3) U nSr = SrU rn for all r ∈ R×, n ∈ R.
If the Sr are unitaries for all r ∈ R× then we say that the represen-
tation is a unitary representation.
We present first two examples:
Example. (The regular representation of an integral domain)
Let H be equal to ℓ2(R), with eq denoting the characteristic function
of {q} ⊆ R. Define operators U n and Sr as follows:
U n Xq∈R
Sr Xq∈R
ζqeq! =Xq∈R
ζqeq! =Xq∈R
ζqeq+n
ζqerq.
It is not difficult to see that {U n} and {Sr} give rise to an isometric
representation of R.
Example. (The regular unitary representation of an integral domain)
Let K be equal to ℓ2(Q(R)), with eq denoting the characteristic func-
) = eq for every q ∈ Q(R)).
r
An important point to notice is that H ⊆ K and further H is
tion of {q} ⊆ Q(R). We use the same formulas to define fU n and eSr.
Notice this time however that for all r, eSr is onto and hence a unitary.
(To see this notice that eSr(e q
an invariant subspace for the collections {eSr} and {fU n}. Further
Sr = PHeSrH for all r and U n = PHfU nH. For this reason we call
this representation the regular unitary representation of R.
We let A(R) be the norm closed operator algebra generated by uni-
taries {un : n ∈ R} and isometries {sr : r ∈ R×} which is universal for
isometric representations of R. We will denote the elements of A(R)
with lower case letters to distinguish from a representation of R for
which we will use upper case letters.
We notice some initial facts about the algebra A(R).
Lemma 1. A(R) is unital with u0 = s1.
Proof. Let {U r, Sr} be an isometric representation of R. Then notice
that U0 is an idempotent unitary, hence 1 = U ∗
0 U0. Then U0 = 1 · U0 =
U ∗
0 U0 = 1. A similar argument yields the same result for S1.
Since this is true for an arbitrary isometric representation of R, the
result follows for A(R).
(cid:4)
0 = U ∗
0 U 2
4
BENTON L. DUNCAN
Lemma 2. If r is invertible then sr is a unitary with s∗
r = sr−1
Proof. This follows from the previous lemma since SrSr−1 = S1 =
Sr−1Sr, for any isometric representation of R.
(cid:4)
It follows that if R is a field then A(R) is a C ∗-algebra. In addition,
in this case, the regular representation is a unitary representation. In
fact we have the following:
Proposition 1. R is a field if and only if every isometric representa-
tion is a unitary representation.
Proof. This comes from the fact that for the regular representation
(Sr)∗ is in the algebra if and only if r is invertible.
(cid:4)
Finally we can see that A(R) is functorial for ring monomorphisms
since any isometric representation of a ring R2 will give rise to an
isometric representation of a ring R1 if there is a ring monomorphism
from R1 into R2..
Proposition 2. A(R) is functorial in the sense that if π : R1 → R2
is a unital ring monomorphism then there is an induced completely
contractive representation π : A(R1) → A(R2)
2. The universal C ∗-algebra for a field
We now analyze the case where R is a field. Here any isometric
representation is, in fact, a unitary representation. We let (R, +) de-
note the additive group in R. Notice that R× acts on C ∗(R, +) as
∗-automorphisms via the mapping αλ(U n) = U λn where U n is the uni-
tary in C ∗(R, +) corresponding to n ∈ (R, +). This allows us to rewrite
C ∗(R) as a crossed product.
Proposition 3. Let R be a field, then A(R) ∼= C ∗(R, +) ⋊ R×.
Proof. We begin by noting (see [2, II.10.3.10]) that since C ∗(R, +) is
unital and R× is discrete we have C ∗(R, +) ⊂ C ∗(R, +) ⋊ R× via a
representation π0 and there is a natural map ρ0 : R× → C ∗(R, ×) ⋊R×.
Together (π0, ρ0) give rise to a covariant representation of the triple
(C ∗(R, +), R×, α).
Now analyzing the covariance conditions that define C ∗(R, +) ⋊ R×
we see that ρ0(r)π0(n)ρ0(r)∗ = π0(rn). Now for each n, π0(n) is
a unitary and for each r, ρ0(r) is a unitary and hence the natural
covariant representation π0 ⋊ ρ0 gives rise to a unitary representa-
tion of R, and hence there is a completely contractive representation
ι : A(R) → C ∗(R, +) ⋊ R×.
OPERATOR ALGEBRAS ASSOCIATED TO INTEGRAL DOMAINS
5
Next notice that any unitary representation of A(R) gives rise to a
covariant representation of (C ∗(R, +), R×, α) and hence kxk ≤ kι(x)k
so that ι is faithful.
(cid:4)
Since C ∗(R, +) and R× are both abelian we have that the universal
norm on the crossed product C ∗(R, +) ⋊R× coincides with the reduced
norm [11, Theorem 7.13]. In addition, we have that the algebra A(R)
is nuclear [11, Corollary 7.18], when R is a field. We next analyze the
regular representation of A(R) where R is a field.
Proposition 4. Let R be a field, then the regular representation of
A(R) is faithful.
Proof. This follows from analyzing the regular representation of the
algebra C ∗(R, +) ⋊α R×, which is faithful since R× is amenable.
In
effect, we take the left regular representation of C ∗(R, +) acting on
L2(R, +) and add to this the action of R× via ∗-automorphisms. This
is exactly the construction of the regular representation of R and hence
the two coincide.
(cid:4)
Other facts about A(R) can also be explained using the crossed prod-
uct machinery.
Proposition 5. For a field R the algebra A(R) is not simple.
dual of the locally compact abelian group (R, +). Notice that the ∗-
Proof. Notice that C ∗(R, +) = C(dR, +) where (dR, +) is the Pontryagin
automorphisms αλ induce a homeomorphism cαλ on (dR, +) which has
a fixed point for each λ; in particular, cαλ(b0) =b0) for all λ. It follows
that there is a nontrivial invariant ideal in C ∗(R, +) for the action by
R× and hence, see [11, Section 3.5] there is a nontrivial induced ideal
in C ∗(R, +) ⋊ R×.
(cid:4)
We can, however completely describe the ideal structure of A(R) in
the case of a field.
Theorem 1. A(R) is ∗-isomorphic to C ⊕ A where A is a simple C ∗-
Proof. We use the nontrivial invariant ideal from the previous propo-
sition.
algebra. In fact, A is ∗-isomorphic to C0((dR, +) \ {b0}) ⋊ R×.
In particular, since C ∗(R, +) = C(dR, +) let π be the multi-
plicative linear function given on C(dR, +) by evaluation at b0. Fur-
ther, if we let cαλ be the induced homeomorphism on (dR, +) given
by the ∗-automorphism αλ for all λ ∈ R×. The induced represen-
tation, π is a multiplicative linear functional and hence has range
C. Hence, A(R) ∼= C ⊕ ker π. We now wish to describe π. So let
6
BENTON L. DUNCAN
σ : C0(dR, +) → C0((dR, +) \ {0}) be the restriction mapping. Further,
if λ ∈ R× then cαλ, the homeomorphism on dR, + induced by the au-
tomorphism αλ, then the range of σ is invariant under cαλ and hence
there is a map τ : C0((dR, +) \ {0}) ⋊ R× → ker π. But since R× acts
transitively on (dR, +) the crossed product C0((dR, +) \ {0}) ⋊ R× is
simple and hence the map τ must be an isomorphism.
(cid:4)
3. Semicrossed products for discrete semigroups
The preceding construction suggests that for non-field integral do-
mains the crossed product may be replaced by a semicrossed product.
We quickly outline the relevant construction referring to [10] for moti-
vation and to [7] for more information about this semicrossed product.
Given a compact Hausdorff space X we say that a semigroup S acts
on X via continuous maps if for each s ∈ S there is a continuous map
τs : X → X with τs ◦ τt = τst.
If S is unital with identity 0 we
will assume that τ0 is the identity map. Say that a pair (π, St) is an
isometric covariant representation of (X, S, τs) if π is a representation
of C(X) on a Hilbert space H and for each t ∈ S, St is an isometry in
B(H) such that Stπ(f (x)) = π(f (τt(x)))St for all x ∈ X.
It is not hard to see that given the triple (X, S, τs), there is a non-
trivial isometric covariant representation. The construction follows
in the same manner as in [10], we only outline the idea here. Let
H = ℓ2(X, S) where this latter Hilbert space is sequences indexed over
elements of S with entries from x, with canonical basis {es}. Define
π : C(X) → B(H) by π(f (x)) = (f (τs(x)))s∈S. Then set St(es) = est
and extend by linearity. Then (π, St) is an isometric covariant repre-
sentation of (X, S, τs).
We say that the universal operator algebra generated by all isometric
covariant representations of (X, S, τs) is the semicrossed product of X
by S via τ . We denote this algebra as C(X) ⋊τ S.
As examples notice that if α is a single endomorphism of a C ∗-algebra
then we are in the situation described in Peter's original work [10],
where the semigroup is Z+. For an example on the opposite end of the
spectrum we can view the semicrossed product of [6] as a semicrossed
product where the monoid is the free monoid on n generators.
Returning to an integral domain R, we let αr be the ∗-endomorphism
of C ∗(R, +) induced by the group endomorphism given by left multipli-
cation by r. This gives a map from R× into the set of ∗-endomorphisms
of C ∗(R, +). Notice that since we are in an integral domain each of
these endomorphisms is injective. However they are only surjective
when r is a unit.
OPERATOR ALGEBRAS ASSOCIATED TO INTEGRAL DOMAINS
7
Proposition 6. The algebra A(R) is completely isometrically isomor-
phic to C ∗(R, +) ⋊α R×.
Proof. We will show that any isometric representation of R gives rise
to an isometric covariant representation of the pair (C ∗(R, +) ⋊α R×)
and vice-versa, hence the two algebras will be completely isometrically
isomorphic.
So let {U n : n ∈ R} and {Sr ∈ R×} be an isometric representation
of R. Then the map n 7→ U n gives rise to a representation of C ∗(R, +),
call it π. Further the isometries Sr will satisfy SrU nS ∗
r = U rn and
hence the pair (π, {Sr}) will be an isometric covariant representation
of (C ∗(R, +), R×, α).
Let (π, {Sr}) be an isometric covariant representation of the alge-
bra (C ∗(R, +), R×, α). Then {π(un)} is a collection of unitaries and
{Sr} is a collection of isometries that trivially satisfy the first two con-
ditions of an isometric representation of R. Further SrU nS ∗
r = U rn
so that SrU n = U rnSr for all r ∈ R× so that we have an isometric
representation of R.
(cid:4)
Notice that in the case of the algebra A(R) for an integral domain
R the semigroup R× will always be commutative with no torsion. In
addition the semigroup R× can be viewed as a spanning cone for the
group Q(R)× (see [9, Page 60] for the definition of a spanning cone for
a group).
We can actually improve our characterization of A(R) as a semi-
crossed product by looking at a more tractable semigroup. Let U(R)
denote the group of units in R. Now R× is a commutative monoid
which contains U(R) as a normal submonoid. We let M(R) denote
the monoid R×/U(R). Notice that R× ⊆ Q(R)× and further that
M(R) ⊆ Q(R)×/U(R), this latter group we call G(R).
For u ∈ U(R) let αu : C ∗(R, +) → C ∗(R, +) be the ∗-automorphism
induced by the automorphism of (R, +) that corresponds to left multi-
plication by u. Next for r ∈ M(R) define a covariant representation of
the triple (C ∗(R, +), U(R), α) by βr(U n) = U nr, ρ(Su) = Su. This
covariant representation induces a ∗-endomorphism of C ∗(R, +) ⋊α
U(R). Hence, β gives rise to a map from M(R) into the set of ∗-
endomorphisms of C ∗(R, +) ⋊α U(R).
Theorem 2. The algebra A(R) is completely isometrically isomorphic
to (C ∗(R, +) ⋊α U(R)) ⋊β M(R), and the diagonal algebra A(R) ∩
A(R)∗ = (C ∗(R, +) ⋊α U(R)).
8
BENTON L. DUNCAN
Proof. We will show that any isometric representation of R gives rise
to an isometric covariant representation of the pair
(C ∗(R, +) ⋊α U(R), M(R))
and vice-versa, hence the two algebras will be completely isometrically
isomorphic.
So let {U n : n ∈ R} and {Sr : r ∈ R×} be an isometric representation
of R. Then define a covariant representation of (C ∗(R, +), U(R), α) by
n 7→ U n and r 7→ Sr for all r ∈ U(R). This yields a representation π of
C ∗(R, +) ⋊α U(R). Next notice that Sλπ(U n) = π(U λn)Sλ for all λ ∈
M(R), and SλSr = SrSλ for all r ∈ U(R), λ ∈ M(R). Hence we have
an isometric covariant representation of (C ∗(R, +) ⋊α U(R), M(R), β).
Finally we take an isometric covariant representation (π, S) of the
triple (C ∗(R, +) ⋊α U(R), M(R), β). Define U n = π(n) and Sr = π(Sr)
if r ∈ U(R), else Sr = Sr. This gives rise to an isometric representation
of R.
The last result follows from Corollary 2 of [7].
(cid:4)
Notice that if R is not a field U(R) does not act transitively on the
nonunital subalgebra of C ∗(R, +), as in the case of a field. In fact we
have the following fact.
Corollary 1. The diagonal is isomorphic to C ⊕ A where
Further, A is simple if and only if R is a field.
A ∼= C0((dG, +) \b0) × U(R).
Proof. That A is not simple follows from the fact that U(R) does not
act transitively on C0( \C ∗(G, +) \ 0) unless R× = U(R).
(cid:4)
We now prove some other facts about the relationship between the
integral domain R and the structure of the algebra A(R).
Proposition 7. A(R) ∼= C(X) ⋊ S, where S is a monoid with no
nontrivial invertible elements if and only if the identity of R is the only
unit. Further if U(R) = {1} then M(R) will not be finitely generated.
Proof. If the identity of R is the only unit, then we have
C ∗(R, +) × U(R) ∼= C ∗(R, +)
and M(R) has no nontrivial invertible elements, else M(R) ∩ U(R) 6=
{1}.
Notice that if A(R) ∼= C(X) × S where S is a monoid with no
nontrivial invertible elements then the diagonal algebra A(R)∩A(R)∗ =
C(X). However, if x ∈ U(R) with x 6= 1, then x 6∈ M(R) and hence
OPERATOR ALGEBRAS ASSOCIATED TO INTEGRAL DOMAINS
9
Sx ∈ C ∗(R, +) × U(R) = A(R) ∩ A(R)∗. But notice that SxU n 6= U nSx
unless x = 1 and hence C ∗(R, +) × U(R) is not commutative.
Assume now that U(R) = {1} and M(R) is finitely generated by the
set {x1, x2, · · · , xn}, then 1 + x1x2 · · · xn 6∈ U(S) so
2 · · · xkn
1 + x1x2 · · · xn = xk1
1 xk2
n
with at least one kj 6= 0. We will assume without loss of generality
that k1 6= 0. Then 1 = x1(x2x3 · · · xn − xk1−1
n ) which implies
1
that x1 is a unit yielding a contradiction.
(cid:4)
xk2
2 · · · xkn
As a corollary we have the following.
Proposition 8. If R is a unique factorization domain and A(R) ∼=
A × Z+ where A is a C ∗-algebra, then A is not commutative and U(R)
is not trivial.
Proof. We will assume that A is commutative and hence U(R) is trivial.
In particular A = C ∗(R, +) = C0(dR, +). Now let x1 and x2 be two
irreducible elements of M(R), then define two two-dimensional nest
representations of A(R) by
0
f (b1)(cid:21) , for f ∈ C0(dR, +)
πi(f ) =(cid:20)f (bxi)
πi(Sxi) =(cid:20)0 1
0 0(cid:21)
0
πi(Sr) = 0 for r 6= xi.
It follows from the description of the two-dimensional nest represen-
tations of A × Z+, see [5], that x1 = x2 which contradicts the fact that
M(R) must be infinitely generated.
(cid:4)
It follows that if R is a unique factorization domain A(R) is never
a semicrossed product in the sense of [10] and hence this collection of
algebras presents a unique type of semicrossed product.
4. Unitary representations of R
Finally we wish to analyze the unitary representations of A(R).
Lemma 3. There is a canonical completely contractive representation
i : A(R) → C ∗(Q(R)).
Proof. As R ⊆ Q(R) the inclusion map provides an isometric represen-
tation of R inside C ∗(Q(R)) and hence the induced map on A(R) is
completely contractive.
(cid:4)
10
BENTON L. DUNCAN
Proposition 9. Let π : A(R) → B(H) be a unitary representation.
There exists a ∗-representation τπ : C ∗(Q(R)) → C ∗(π(A(R)) which is
onto and satisfies τπ ◦ i(x) = π(x) for all x ∈ A(R).
Proof. Since π is a unitary representation we know that π(sr) = Tr is
a unitary for all r ∈ R×. For all h p
q . We also define π(cid:16)u[ p
q ](cid:17) = TqV pT ∗
TpT ∗
unitaries TpT ∗
the induced representation τπ will be the required ∗-representation.
q . We need only show that the
q satisfy the relations for Q(R) and hence
qi ∈ Q(R)× we define π(cid:16)s[ p
q ](cid:17) =
q and TqV pT ∗
Notice first that TpTq = TqTp and T ∗
p T ∗
q = T ∗
q T ∗
p since π is a unitary
representation of A(R). It then follows that
q TqTpT ∗
q TpTqT ∗
q Tp
q = T ∗
= T ∗
= T ∗
TpT ∗
q
q
for all q, p ∈ R×. It follows that π(cid:16)s[ p1
] and [ p2
q2
Next notice that V pTq = TqV pq and T ∗
] in Q(R)×.
[ p1
q1
q1
](cid:17) π(cid:16)s[ p2
q2
](cid:17) = π(cid:16)s[ p1p2
q1q2
](cid:17) for all
Tq1V p1T ∗
q1Tq2V p2T ∗
q V p = V pqT ∗
q2 = Tq1Tq2V p1q2V q1p2T ∗
= Tq1q2V p1q2+q1p2Tq1q2.
q and hence
q1T ∗
q2
q1
Next we have that
In other words π(cid:16)u[ p1
π(cid:16)u[ p1
](cid:17) π(cid:16)u[ p2
](cid:17) π(cid:16)s[ p2
q1
q2
q2
]+[ p2
q2
](cid:17).
q1Tp2T ∗
q2
q1
](cid:17) = π(cid:16)u[ p1
](cid:17) = Tq1V p1T ∗
= π(cid:16)s[ p2
= Tp2T ∗
= Tp2T ∗
q2
q2T ∗
q1
q1
q1q2
q2Tq1Tq2V p1p2T ∗
q2Tq1q2V p1p2T ∗
][ p2
q2
](cid:17) π(cid:16)u[ p1
](cid:17) .
1 ](cid:17) = TpT ∗
Hence the C ∗-algebra generated by the Tp and V n satisfies the relations
for Q(R). We call the induced representation τπ.
Finally we note that τπ ◦ i(sp) = τπ(cid:16)s[ p
p ∈ Q× and τπ ◦ i(un) = τπ(cid:0)u[ n
1 ](cid:1) = T1VnT ∗
hence τπ ◦ i(x) = π(x) for all x ∈ A(R).
1 = Tp for all
1 = Vn for all n ∈ R and
(cid:4)
It would follow that if every isometric representation of R dilated to a
unitary representation (as for example the regular representation does),
then we could identify the C ∗-envelope of A(R) as a crossed product,
since the canonical representation i would be completely isometric. We
OPERATOR ALGEBRAS ASSOCIATED TO INTEGRAL DOMAINS
11
do not think this is likely since this does not even work in the case of
C(X) ⋊α Z+ where α is a non-surjective continuous mapping, see [6].
References
[1] W. Arveson and K. Josephson. Operator algebras and measure preserving au-
tomorphisms. II J. Functional Analysis 4 (1969) 100 -- 134.
[2] B. Blackadar, Operator algebras. Theory of C ∗-algebras and von Neumann
algebras. Encyclopaedia of Mathematical Sciences, 122. Operator Algebras and
Non-commutative Geometry, III. Springer-Verlag, Berlin, 2006.
[3] J. Cuntz, C ∗-algebras associated with the ax + b-semigroup over N K-theory
and noncommutative geometry, 201 -- 215, EMS Ser. Congr. Rep. Eur. Math.
Soc. Zrich, 2008.
[4] J. Cuntz and X. Li, The regular C ∗-algebra of an integral domain, preprint,
2008.
[5] K. Davidson and E. Katsoulis, Isomorphisms between topological conjugacy
algebras, J. reine angew. Math. 621, (2008), 29 -- 51.
[6] K. Davidson and E. Katsoulis, Operator algebras for multivariable dynamics,
to appear in Mem. Amer. Math. Soc. , 2007.
[7] B. Duncan and J. Peters, Dynamics and semicrossed products by discrete
semigroups, preprint, 2010.
[8] X. Li, Ring C ∗-algebras, preprint, 2009.
[9] V. Paulsen, Completely bounded maps and operator algebras Cambridge Uni-
versity Press, Cambridge, 2002.
[10] J. Peters, Semicrossed products of C ∗-algebras, J. Funct. Anal. 59 (1984)
498 -- 534.
[11] D. Williams, Crossed products of C ∗-algebras American Math. Soc. Provi-
dence, 2007.
Department of Mathematics, North Dakota State University, Fargo,
North Dakota, USA
E-mail address: [email protected]
|
1902.02572 | 1 | 1902 | 2019-02-07T11:29:59 | Subfactors and Hecke groups | [
"math.OA",
"math.GT",
"math.NT"
] | We study a relation between the Hecke groups and the index of subfactors in a von Neumann algebra. Such a problem was raised by V. F. R. Jones. We solve the problem using the notion of a cluster C*-algebra. | math.OA | math |
SUBFACTORS AND HECKE GROUPS
ANDREY GLUBOKOV1 AND IGOR NIKOLAEV2
Abstract. We study a relation between the Hecke groups and the index
of subfactors in a von Neumann algebra. Such a problem was raised by
V. F. R. Jones. We solve the problem using the notion of a cluster C ∗-algebra.
The following problem can be found in [Jones 1991] [5, p.24]:
1. Introduction
"Consider the subgroup Gλ of SL2(R) generated by ( 1 λ
0 1 ) and (cid:0) 0 1
−1 0(cid:1).
For what values of λ > 0 is it discrete? Answer: λ = 2 cos(cid:0) π
n(cid:1) , n =
3, 4, . . . or λ ≥ 2. (...) We have been unable to find any direct connection
between this result and Theorem 3.1 (Jones Index Theorem).
It is a
tantalizing situation."
The aim of our note is to solve the problem in terms of the cluster C∗-algebras [6,
Section 4.4.3]. To give an idea, let D = {z = x + iy ∈ C r ≤ z ≤ R} be an
annulus in the complex plane. Consider the Schottky uniformization of D, i.e.
D ∼= CP 1/AZ,
(1.1)
where CP 1 := C ∪ {∞} is the Riemann sphere and A ∈ SL2(C)/ ± I is a matrix
acting on the CP 1 by the Mobius transformation. It follows from [Glubokov &
Nikolaev 2018] [3] and Section 2.2, that the index of subfactors in a von Neumann
algebra coincides with the square of trace of matrix A, i.e.
tr2 (A) ∈ [4,∞) [ {4 cos2(cid:16) π
n(cid:17) n ≥ 3}.
(1.2)
To solve the Jones Problem, we prove in Section 3 that D is a ramified double
cover of the orbifold H/Gλ, where H := {x + iy ∈ C y > 0} is the Lobachevsky
half-plane and the group Gλ acts on H by the linear fractional trasformations.
Since such a cover takes the square root of the moduli parameter tr2 (A) of D,
we conclude that λ = tr (A). In other words, the Jones Index Theorem (1.2) is
equivalent to the following well-known result:
Theorem 1.1. ([Hecke 1936] [4, Satz 1,2 & 6]) The Gλ is a discrete subgroup of
SL2(R) if and only if
λ ∈ [2,∞) [ {2 cos(cid:16) π
n(cid:17) n ≥ 3}.
(1.3)
2010 Mathematics Subject Classification. Primary 46L37; Secondary 20H10.
Key words and phrases. Hecke group, subfactors, cluster C ∗-algebra.
1
2
A.GLUBOKOV AND I.NIKOLAEV
Remark 1.2. The group Gλ appears in the study of the Riemann zeta function
[Hecke 1936] [4]. In particular, Hecke's Theorem says that the space of automorphic
functions corresponding to Gλ is (i) infinite-dimensional, if λ > 2 or (ii) finite-
dimensional, if λ ∈ {2 cos(cid:0) π
[Hecke 1936] [4]. On the other hand, cases (i) and (ii) follow from the Sherman-
Zelevinsky Theorem for the cluster C∗-algebras of rank 2, see [Glubokov & Nikolaev
2018] [3].
n(cid:1) n ≥ 3}. The proof of this fact is purely analytic
The article is organized as follows. Section 2 contains a brief review of prelimi-
nary results. Theorem 1.1 is proved in Section 3.
2. Preliminaries
The cluster C∗-algebras and their K-theory are covered in [6, Section 4.4.3].
A correspondence between the cluster C∗-algebra of an annulus D and the Jones
Index Theorem was established in [Glubokov & Nikolaev 2018] [3]. The Hecke
groups were introduced in [Hecke 1936] [4].
2.1. Cluster C∗-algebras. A cluster algebra A (x, B) of rank n is a subring of
the field of rational functions in n variables depending on a cluster of variables
x = (x1, . . . , xn) and a skew-symmetric matrix B = (bij) ∈ Mn(Z); the pair (x, B)
is called a seed. A new cluster x′ = (x1, . . . , x′k, . . . , xn) and a new skew-symmetric
matrix B′ = (b′ij) is obtained from (x, B) by the exchange relations:
n
xmax(bik ,0)
xkx′k =
i
Yi=1
b′ij = (−bij
xmax(−bik,0)
i
,
+
n
Yi=1
bij + bikbkj +bikbkj
2
if i = k or j = k
otherwise.
(2.1)
The seed (x′, B′) is said to be a mutation of (x, B) in direction k, where 1 ≤
k ≤ n; the algebra A (x, B) is generated by cluster variables {xi}∞i=1 obtained
from the initial seed (x, B) by the iteration of mutations in all possible directions
k. The Laurent phenomenon says that A (x, B) ⊂ Z[x±1], where Z[x±1] is the
ring of the Laurent polynomials in variables x = (x1, . . . , xn) depending on an
initial seed (x, B). The A (x, B) is a commutative algebra with an additive abelian
semigroup consisting of the Laurent polynomials with positive coefficients. Thus
the algebra A (x, B) is a countable abelian group with an order satisfying the Riesz
interpolation property, i.e. a dimension group. A cluster C∗-algebra A(x, B) is an
AF-algebra, such that
K0(A(x, B)) ∼= A (x, B),
(2.2)
where ∼= is an isomorphism of the dimension groups [6, Section 4.4.3].
2.2. Schottky uniformization of D. Consider the Riemann surface D (an an-
nulus) defined by the formula (1.1). We shall use the Schottky uniformization of D
by the loxodromic transformations. Namely, let CP 1 := C ∪ {∞} be the Riemann
sphere and consider the Mobius transformation of CP 1 given by the formula:
z 7−→ k z, where z ∈ CP 1
and k 6= 1.
(2.3)
SUBFACTORS AND HECKE GROUPS
3
It is easy to see, that (2.3) can be written in the matrix form:
A = √k
0
0
1√k! ∈ P SL2(C) := SL2(C)/ ± I.
It is well known, that (2.3) is a loxodromic transformation if and only if
∈ C \ [0, 4].
The Schottky uniformization of D is given by the formula:
k
tr2 (A) =
(k + 1)2
D ∼= CP 1/AZ.
(2.4)
(2.5)
(2.6)
2.3. Admissible values of tr2 (A). Recall that the moduli space of the annulus
D is given by the formula:
TD =(cid:26)t =
R
r t > 1(cid:27) .
(2.7)
We consider a cluster C∗-algebra A(D) associated to a canonical triangulation of D
[Fomin, Shapiro & Thurston 2008] [1, Example 4.4]. It follows from the Sherman-
Zelevinsky Theorem for the algebra A(D), that the admissible values of the "index"
(t+1)2
t must belong to the set:
see [Glubokov & Nikolaev 2018] [3] for the proof.
[4,∞) [ {4 cos2(cid:16) π
n(cid:17) n ≥ 3},
(2.8)
We set k = t in the formulas (2.3) - (2.5). Comparing (2.5) and (2.8), one gets
Remark 2.1. It follows from (2.5) that A is a loxodromic transformation if and only
tr2 (A) ∈ [4,∞) [ {4 cos2(cid:16) π
if tr2 (A) ∈ (4,∞). The case tr2 (A) ∈ {4 cos2(cid:0) π
elliptic transformation A of order n. Finally, that case tr2 (A) = 4 gives a parabolic
transformation A. Note that for the elliptic and parabolic transformations, the
values of parameter t in (2.7) are the n-th roots of unity.
n(cid:1) n ≥ 3} corresponds to an
n(cid:17) n ≥ 3}.
(2.9)
3. Proof of theorem 1.1
We split the proof in a series of lemmas.
Lemma 3.1. Let p1, p2 ∈ CP 1 be two points on the Riemann sphere, such that
p1 6= p2. Then there exists a double covering map
p : CP 1 → CP 1
(3.1)
ramified over the points p1 and p2.
Proof. (i) Recall that the necessary condition for the existence of p is given by the
Riemann-Hurwitz formula:
χ(CP 1) = 2χ(CP 1) −
2
Xi=1
(ei − 1),
(3.2)
4
A.GLUBOKOV AND I.NIKOLAEV
where χ(CP 1) is the Euler characteristic and ei is the degree of the map z 7→ zei
in the ramification point pi. Since χ(CP 1) = e1 = e2 = 2, we conclude that the
condition (3.2) is satisfied.
(ii) The sufficient condition for the existence of p can be verified directly using
[Gersten 1987] [2, Theorem 1.5]. We leave it as an exercise to the reader.
Lemma 3.1 is proved.
(cid:3)
Remark 3.2. We assume further that in lemma 3.1 we have p1 = 0 and p2 = ∞.
Our assumption is not restrictive, since any two points p1, p2 ∈ CP 1 can be put
into such a position by a Mobius transformation.
ρ
D(tr2(A))
∞
D
D(λ)
ep
H/Gλ
Figure 1. Covering map ρ : D → H/Gλ.
Lemma 3.3. Consider a Riemann surface:
(3.3)
where D0 is a disk containing point 0 ∈ CP 1. Then there exists a double covering
map
H := CP 1 \ {D0,∞},
(3.4)
Proof. Let p : CP 1 → CP 1 be the double covering map ramified at 0 and ∞, see
lemma 3.1 and remark 3.2. Recall that
ρ : D → H .
D ∼= CP 1 \ {D0, D∞},
(3.5)
where D∞ is a disk containing point ∞ ∈ CP 1. We use a homotopy to contract D∞
to the point ∞, and set the map ρ ≡ p. Comparing (3.3) and (3.5), we conclude
that ρ is the required double covering map. Lemma 3.3 is proved.
Lemma 3.4. H ∼= H/Gλ, where λ ∈ [2,∞) S {2 cos(cid:0) π
n(cid:1) n ≥ 3}.
(cid:3)
SUBFACTORS AND HECKE GROUPS
5
Proof. It is well known, that the Hecke orbifold {H/Gλ λ > 2} is a topological
sphere S2 with a hole D(λ) of radius λ/2, one elliptic fixed point e2 of order 2 and
one puncture c, see e.g. [Schmidt & Sheingorn 1995] [7, p. 255]. The elliptic point
e2 = i is a fixed point of the matrix(cid:0) 0 1
−1 0(cid:1) having order 2 in the group SL2(Z)/±I.
Since S2 ∼= CP 1, we use (3.3) to identify D(λ) ≡ D0 and e2 ≡ ∞. Thus one
gets H ∼= H/Gλ. Using lemma 3.3, we obtain a double covering map ρ ramified in
the points e2 and 0 ∈ D(λ), i.e.
ρ : D → H/Gλ.
Remark 3.5. Notice that ρ−1(D(λ)) is a disk and ρ−1(e2) is a regular point of D.
The corresponding ramification points are shown in Figure 1.
1
To determine admissible values of the moduli parameter λ, observe that λ =
π∂D(λ), where ∂D(λ) is the boundary of the disk D(λ). Observe that the local
map at the point 0 ∈ D(λ) is given by the formula z 7→ z2. Therefore using the
polar coordinates, we conclude that:
(3.6)
(3.7)
1
λ2 =
π (cid:12)(cid:12)ρ−1(∂D(λ))(cid:12)(cid:12)
is a moduli parameter of the Riemann surface D. But according to (1.2) any such
a parameter must coincide with the tr2 (A), where A is the matrix in the Schottky
uniformization (1.1). Taking positive values of the square root, one gets from (1.2)
λ = tr (A) ∈ [2,∞) [ {2 cos(cid:16) π
n(cid:17) n ≥ 3}.
Lemma 3.4 is proved.
(3.8)
(cid:3)
Remark 3.6. For the sake of brevity, lemma 3.4 is proved for the continuous moduli
λ ∈ (2,∞). The case of the discrete moduli is treated likewise, see remark 2.1.
Theorem 1.1 follows from lemma 3.4.
References
1. S. Fomin, M. Shapiro and D. Thurston, Cluster algebras and triangulated surfaces, I. Cluster
complexes, Acta Math. 201 (2008), 83-146.
2. S. M. Gersten, On branched covers of the 2-sphere by the 2-sphere, Proc. Amer. Math. Soc.
101 (1987), 761-766.
3. A. Glubokov and I. Nikolaev, Jones Index Theorem revisited, arXiv:1801.05510
4. E. Hecke, Uber die Bestimmung Dirichletscher Reihen durch ihre Funktionalgleichung, Math.
Annalen 112 (1936), 664-699.
5. V. F. R. Jones, Subfactors and Knots, CBMS Series 80, AMS, 1991.
6. I. Nikolaev, Noncommutative Geometry, De Gruyter Studies in Mathematics 66, De Gruyter,
Berlin, 2017.
7. T. A. Schmidt and M. Sheingorn, On the infinite volume Hecke surfaces, Compositio Math.
95 (1995), 247-262.
1 Department of Mathematics, Ave Maria University, 5050 Ave Maria Blvd., Ave
Maria, FL 34142, United States.
E-mail address: [email protected]
2 Department of Mathematics and Computer Science, St. John's University, 8000
Utopia Parkway, New York, NY 11439, United States.
E-mail address: [email protected]
|
1506.02308 | 2 | 1506 | 2015-07-14T22:37:23 | The classification of simple separable unital locally ASH algebras | [
"math.OA"
] | Let $A$ be a simple separable unital locally approximately subhomogeneous C*-algebra (locally ASH algebra). It is shown that $A\otimes Q$ can be tracially approximated by unital Elliott-Thomsen algebras with trivial $\textrm{K}_1$-group, where $Q$ is the universal UHF algebra. In particular, it follows that $A$ is classifiable by the Elliott invariant if $A$ is Jiang-Su stable. | math.OA | math |
THE CLASSIFICATION OF SIMPLE SEPARABLE UNITAL LOCALLY ASH
ALGEBRAS
GEORGE A. ELLIOTT, GUIHUA GONG, HUAXIN LIN, AND ZHUANG NIU
Abstract. Let A be a simple separable unital locally approximately subhomogeneous C*-
algebra (locally ASH algebra). It is shown that A ⊗ Q can be tracially approximated by unital
Elliott-Thomsen algebras with trivial K1-group, where Q is the universal UHF algebra. In par-
ticular, it follows that A is classifiable by the Elliott invariant if A is Jiang-Su stable.
1. Introduction
Recently, several major steps have been taken in the classification of what might be called "well
behaved" separable amenable simple unital C*-algebras. The phenomenon of well behavedness
itself was explicitly noticed only relatively recently, by Toms and Winter (see [7]) who conjectured
that within this class of C*-algebras several properties were equivalent, and that the algebras in
this robust subclass (and only these) were the ones that could be classified by means of what
might be called the naive Elliott invariants -- the ordered K0-group together with the class of the
unit, the simplex of tracial states paired naturally with it, and the K1-group. (Other invariants,
such as the Cuntz semigroup, might then be helpful for more general classes of amenable C*-
algebras.) Breakthroughs in the understanding of the robustness of this class were made by
Matui and Sato in [20] and [21].
Perhaps the most striking development in the direction of actually proving isomorphism has
been the technique -- sometimes referred to as the Winter deformation technique -- introduced
by Winter in [33] (with refinements by Lin in [17] and Lin and Niu in [19]), through which a
class of (separable, amenable, simple unital) C*-algebras which are well behaved in the sense
of absorbing tensorially the Jiang-Su algebra Z, and are also known to satisfy the UCT, can
be classified in terms of the (naive) Elliott invariant if this is true for the subclass of algebras
absorbing the universal Glimm UHF C*-algebra Q.
Using this, Gong, Lin, and Niu, in [9] -- following on earlier work in this direction (see e.g.
[14], [22], [23], [24], [15], [33], [17], [19], [16]) -- have achieved a classification of finite algebras in
this well behaved class which is close to being definitive -- it is simple to describe and it exhausts
completely the possible values of the invariant. (The complementary class, the infinite algebras
in the well behaved class under consideration, were dealt with some time ago by Kirchberg and
Phillips -- [11], [12], [25].)
Unfortunately, while it is believed that all well behaved finite separable amenable simple
unital C*-algebras may be ASH algebras (inductive limits of subalgebras of full matrix algebras
over commutative C*-algebras) -- and, indeed, that the algebra need not be assumed to be well
Date: July 16, 2018.
1
THE CLASSIFICATION OF SIMPLE SEPARABLE UNITAL LOCALLY ASH ALGEBRAS
2
behaved if in addition matrix algebras over it are also finite in the Murray-von Neumann sense
(i.e., if the algebra is stably finite) -- the class considered by Gong, Lin, and Niu does not on the
face of it include the class of all well behaved -- "Jiang-Su stable" -- simple unital ASH algebras.
Using the recent result of Santiago, Tikuisis, and the present authors ([5]) that any Jiang-Su
stable simple unital ASH algebra has finite nuclear dimension (one of the Toms-Winter well
behavedness properties -- the important concept of nuclear dimension was introduced by Winter
and Zacharias in [34]) -- and also using the notion of non-commutative cell complex introduced
in [5] in the proof of this result -- , the present note shows that indeed such an algebra (Jiang-Su
stable simple unital ASH) belongs to the class dealt with by Gong, Lin, and Niu. (Even if the
ASH algebra has slow dimension growth, so that by [28] and [31] it is indeed Jiang-Su stable,
it is not known to belong to the class studied by Gong, Lin, and Niu -- the class of rationally
tracially approximately point -- line algebras -- see below.)
Acknowledgements. The research of G. A. E. was supported by a Natural Sciences and En-
gineering Research Council of Canada (NSERC) Discovery Grant, the research of G. G. was
supported by an NSF Grant, the research of H. L. was supported by an NSF Grant, and the
research of Z. N. was supported by a Simons Foundation Collaboration Grant. Z. N. also thanks
Aaron Tikuisis for his comments on the early version of the paper.
2. Noncommutative cell complexes
Definition 2.1 (Definition 2.1 of [5]). The class of noncommutative cell complexes (NCCC) is
the smallest class C of C*-algebras such that
(1) every finite dimensional algebra is in C; and
(2) if B ∈ C, n ∈ N, ϕ : B → Mk(C(Sn−1)) is a unital homomorphism, and A is given by the
pullback diagram
A
Mk(C(Dn))
f →f Sn−1
B ϕ
/ Mk(C(Sn−1)),
then A ∈ C.
The reason we consider NCCCs is as follows:
Theorem 2.2 (Theorem 2.15 of [5]). Let A be a unital subhomogeneous C*-algebra. Then A
can be locally approximated by sub-C*-algebras which are NCCCs.
Definition 2.3. Let A be an NCCC, and fix an NC cell complex decomposition of A with
length l (in the sense that A is built from a finite dimensional C*-algebra A0 by attaching l
noncommutative cells). Assume that A0 = Ms1(C) ⊕ · · · ⊕ Msr (C) for some natural number r,
and denote the attached cell at the i-th step by Mki(C(Dni)).
Consider the spaces
{pt}, ..., {pt}
, Dn1, ..., Dnl,
}
r
{z
/
/
/
THE CLASSIFICATION OF SIMPLE SEPARABLE UNITAL LOCALLY ASH ALGEBRAS
3
and denote them by X1, X2, ..., Xm, where m = r + l. Denote the matrix orders of the corre-
sponding C*-algebras by
Then there is a standard embedding
d1, ..., dr, dr+1, ..., dm.
Π : A →
mMi=1
Mdi(C(Xi)).
Denote by Πi : A → Mdi(C(Xi)), i = 1, ..., m, the projection of Π onto the i-th direct summand.
Lemma 2.4. Let A = B ⊕Mk(C(Sn−1)) Mk(C(Dn)) be an NCCC, and let τ ∈ T(A). Then there is
a decomposition
τ (fB, fD) = ατB(fB) + βZDn\Sn−1
tr(fD(x))dµ(x),
(fB, fD) ∈ A,
where τD ∈ T(B), µ is a probability measure on Dn \ Sn−1, tr is the standard trace of Mk(C),
and α, β ∈ [0, 1] with α + β = 1. Moreover, α and β are unique and τB (or µ) is unique if α (or
β) is not zero.
Proof. The uniqueness part of the lemma is clear. Let us show the existence part.
Consider the restriction of τ to I := Mk(C0(Dn \ Sn−1)) ⊆ A. Then it is a trace with norm at
most one, and thus there is β ∈ [0, 1] and a probability measure µ on Dn \ Sn−1 such that
χη : [0, 1] ∋ x 7→
x ∈ [1 − η/2, 1],
1,
linear, x ∈ [1 − η, 1 − η/2],
0,
x ∈ [0, 1 − η];
and
Then a direct calculation shows that
(2.1)
τ (f, g) = τ (f, gη),
(f, g) ∈ A, η ∈ (0, 1).
gη(x) = g(x)χη(kxk).
It is clear that τ is self-adjoint; let us show that it is positive. Let (f, g) ∈ A be positive.
Define
If δ = 0, let us show that τ (f, g) = 0. Indeed, in this case, one has
δ = inf{τ (f, gη) : η ∈ (0, 1)}.
τ ((f, g)) = τ (f, gη) + τ (0, g − gη)
Define a linear map τ : A → C by
τ ((0, g)) = βZDn\Sn−1
τ ((f, g)) = τ (f, g) − βZDn\Sn−1
For each g ∈ Mk(C(Dn)) and any η ∈ (0, 1), define
tr(g(x))dµ(x),
g ∈ Mk(C0(Dn \ Sn−1)).
tr(g(x))dµ(x),
(f, g) ∈ A.
THE CLASSIFICATION OF SIMPLE SEPARABLE UNITAL LOCALLY ASH ALGEBRAS
4
and hence
(2.2) τ ((f, g)) = sup{τ (0, g − gη) : η ∈ (0, 1)} = sup{βZDn\Sn−1
trx(g − gη)dµ(x) : η ∈ (0, 1)}.
Note that for any η ∈ (0, 1).
τ ((f, g)) = τ (f, g) − βZDn\Sn−1
= τ (f, g) − βZDn\Sn−1
+βZDn\Sn−1
tr(gη(x))dµ(x),
tr(g(x))dµ(x)
trx(g − gη)dµ(x)
and since µ is a probability measure, the integral βRDn\Sn−1 tr(gη(x))dµ(x) is arbitrarily small if
η is small enough. By (2.2), one has that τ ((f, g)) = 0.
If δ > 0, since µ is a probability measure, there is η ∈ (0, 1) such that
βZDn\Sn−1
tr(gη(x))dµ(x) < δ/2,
and therefore
τ ((f, g)) = τ ((f, gη)) = τ (f, gη) − βZDn\Sn−1
tr(gη(x))dµ(x) ≥ −δ/2 = δ/2 > 0.
Therefore, one always has τ ((f, g)) ≥ 0, and so τ is a positive linear functional. Therefore τ
is a (positive) trace of A. Note that τ (I) = 0, and therefore τ factors through A/I ∼= B, and
hence in fact is a trace of B. Therefore, there are α ∈ [0, 1] and τB ∈ T(B) such that
τ (f, g) − βZDn\Sn−1
tr(g(x))dµ(x) = τ (f, g) = ατB(f ),
(a, b) ∈ A,
as desired.
(cid:3)
Corollary 2.5. Let A be an NCCC with a given decomposition with length l. Then any trace τ
of A has a decomposition
τ = α1τ1 + α1µ1 + · · · + αmµm,
where m = rank(K0(A0)) + l, µi is a probability measure on Dni \ Sni−1 if Xi = Dni, and µi is
the Dirac measure if Xi consists of a point, αi ∈ [0, 1] and α1 + α2 + · · · + αm = 1. Moreover,
the coefficients αi are unique.
Definition 2.6. Let A be an NCCC with a given decomposition, and let τ ∈ T(A). Referring
to Corollary 2.5, define
αi(τ ) = αi.
Lemma 2.7. Let A be a noncommutative cell complex (NCCC). Then the K-groups of A are
finitely generated (as abelian groups).
THE CLASSIFICATION OF SIMPLE SEPARABLE UNITAL LOCALLY ASH ALGEBRAS
5
Proof. The statement is true if A is finite dimensional. Assume the statement is true for non-
commutative complexes with length at most l.
Let A be a noncommutative complex with length l + 1. Write
where B is a noncommutative complex with length l. Then there is a short exact sequence
A = B ⊕Mk(C(Sn−1)) Mk(C(Dn)),
0
/ Mk(C0(Rn))
/ A
/ B
/ 0,
and the corresponding six-term exact sequence is
K0(C0(Rn))
/ K0(A)
/ K0(B)
If n is odd, one has
and
K1(B)
K1(A)
K1(C0(Rn)).
0
/ K0(A)
/ K0(B)
/ · · · ,
0
/ Z/mZ
/ K1(A)
/ K1(B)
/ 0,
for some positive integer m. By the inductive hypothesis, the groups K0(B) and K1(B) are
finitely generated, and therefore the groups K0(A) and K1(A) are finitely generated.
If n is even, a similar argument shows that K0(A) and K1(A) are finitely generated. Therefore,
the K-groups of A are always finitely generated. Hence by induction, the statement holds for all
noncommutative cell complexes.
(cid:3)
In general, the positive cone of K0(A) might not be finitely generated; for instance, the positive
cone of K0(C(S2)) is
which is not finitely generated. (On the other hand, consider the image
{(m, n) ∈ Z2 : m > 0} ∪ {(0, 0)},
G := ρ(K0(A)) ⊆ Aff(T(A)),
with respect to the canonical map ρ, with the induced order from Aff(T(A)) (i.e., an element
g ∈ G is positive if and only if g is positive in Aff(T(A))), is isomorphic to Z and so the positive
cone of G is finitely generated.)
The following lemma was stated and proved in [9] for the K0-group of an Elliott-Thomsen
algebra. The argument in fact shows the following (for the reader's convenience, we include the
proof). (In fact the ordered groups arising are the same.)
Lemma 2.8 (Theorem 3.14 of [9]). Consider (Zl, (Zl)+) with the standard (direct sum) order.
Let G be a subgroup of Zm, and put G+ = G ∩ (Zm)+. Then the positive cone G+ is finitely
generated (as a semigroup).
/
/
/
/
/
/
O
O
o
o
o
o
/
/
/
/
/
/
/
THE CLASSIFICATION OF SIMPLE SEPARABLE UNITAL LOCALLY ASH ALGEBRAS
6
Proof. Let us first show that G+ \ {0} has only finitely many minimal elements. Suppose,
otherwise, that {qn} is an infinite set of minimal elements of G+ \ {0}. Write
qn = (m(1, n), m(2, n), ..., m(j, n)) ∈ Zm
+ ,
where m(i, n) are positive integers (including zero), i = 1, 2, ..., m and n = 1, 2, .... If there is
an integer M ≥ 1 such that m(i, n) ≤ M for all i and n, then {qn} is a finite set. So we may
assume that {m(i, n)} is unbounded for some 1 ≤ i ≤ m. Passing to a subsequence of {nk}
such that limk→∞ m(i, nk) = +∞, we may assume that limn→∞ m(i, n) = +∞. We may assume
that, for some j, {m(j, n)} is bounded. Otherwise, by passing to a subsequence, we may assume
that limn→∞ m(i, n) = +∞ for all i ∈ {1, 2, ..., m}. Therefore limn→∞ m(i, n) − m(i, 1) = +∞.
It follows that, for some n ≥ 1, m(i, n) > m(i, 1) for all i ∈ {1, 2, ..., m}. Therefore, qn > q1,
which contradicts the fact that qn is minimal. By passing to a subsequence, we may write
{1, 2, ..., m} = N ⊔ B such that limn→∞ m(i, n) = +∞ if i ∈ N and {m(i, n)} is bounded if
i ∈ B. Therefore, {m(j, n)} has only finitely many different values if j ∈ B. Thus, by passing
to a subsequence again, we may assume that m(j, n) = m(j, 1) if j ∈ B. Therefore, for some
n > 1, m(i, n) > m(i, 1) for all n if i ∈ N, and m(j, n) = m(j, 1) for all n if j ∈ B. It follows
that qn ≥ q1. This is impossible, since qn is minimal. This shows that G+ has only finitely many
minimal elements.
To show that G+ is generated by these minimal elements, fix an element q ∈ G+ \ {0}. It
suffices to show that q is a finite sum of minimal elements in G+. If q is not minimal, consider
the set of all elements in G+ \ {0} which are (strictly) smaller than q. This set is finite. Choose
one which is minimal among them, say p1. Then p1 is minimal element in G+ \ {0}, as otherwise
there is one smaller than p1. Since q is not minimal, q 6= p1. Consider q − p1 ∈ G+ \ {0}. If q − p1
is minimal, then q = p1 + (q − p1). Otherwise, we repeat the same argument to obtain a minimal
element p2 ≤ q − p1. If q − p1 − p2 is minimal, then q = p1 + p2 + (q − p1 − p2). Otherwise we
repeat the same argument. This process is finite. Therefore q is a finite sum of minimal elements
in G+ \ {0}.
(cid:3)
With Lemma 2.8, one has
Lemma 2.9. Let A be an NCCC. Then the ordered group
(ρA(K0(A)), ρA(K0(A)) ∩ Aff +(T(A)))
is finitely generated (as an ordered group). (In other words, the positive cone is finitely generated
as a semigroup.)
Proof. With the fixed NC cell complex decomposition of A, consider the standard embedding
Π : A →
mMi=1
Mdi(C(Xi)).
Define
ρ : K0(A) ∋ [p] 7→ (rank(Π1(p)), ..., rank(Πm(p))) ∈ Zm.
Clearly, the map ρ is positive.
THE CLASSIFICATION OF SIMPLE SEPARABLE UNITAL LOCALLY ASH ALGEBRAS
7
Define
G = ρ(K0(A)) and G+ = ρ(K0(A)) ∩ (Zm)+.
It follows from Lemma 2.8 that the cone (G, G+) is a finitely generated ordered group. In order
to prove the lemma, one only has to show that there is an isomorphism
(G, G+) ∼= (ρA(K0(A)), ρA(K0(A)) ∩ Aff +(T(A))).
Define
θ : ρ(K0(A)) ∋ (x1, x2, ..., xm) 7→ (τ 7→
mXi=1
xi
di
αi(τ )) ∈ Aff(T(A)).
Then θ is injective. Note that for any τ ∈ T(A) with the decomposition
one has
and therefore
τ = α1µ0 + α2µ2 + · · · + αmµm,
ρA(p)(τ ) =
=
=
αiZXi
αiZXi
tri(Πi(p)(x))dµi(x)
rank(Πi(p))
di
dµi,j
αi
rank(Πi(p))
di
= θ(ρ(p)),
mXi=1
mXi=1
mXi=1
θ(ρ(K0(A))) = θ(G) = ρA(K0(A)).
then the affine map τ 7→ Pm
It is clear that θ(G+) ⊆ ρA(K0(A)) ∩ Aff +(T(A)). Moreover, if θ(x1, ..., xm) ∈ Aff +(T(A)),
αi(τ ) is positive, and hence each xi must be positive; that
is, θ induces an order isomorphism between (G, G+) and ρA(K0(A)), ρ(K0(A)) ∩ Aff +(T(A)), as
desired.
(cid:3)
xi
di
i=1
Lemma 2.10. Let A be an NCCC. Then, for any finite set F ⊆ Aff(T(A)) and any ε > 0, there
are positive continuous affine maps θ1 : Aff(T(A)) → Rs and θ2 : Rs → Aff(T(A)) for some
s ∈ N such that
kθ2 ◦ θ1(f ) − f k∞ < ε,
f ∈ F .
Proof. The statement clearly holds for A a finite dimensional C*-algebra. Assume that
A = B ⊕Mk(C(Sn−1)) Mk(C(Dn)),
and suppose that the statement holds for B.
Let (F , ε) be given. Note that Aff(T(A)) is the pullback of Aff(T(B)) and CR(Dn) in the same
manner as A. For each f ∈ F , write it as f = (fB, fD), where fB ∈ Aff(T(B)) and fD ∈ CR(Dn).
Denote by FB the set of fB's.
Since B satisfies the statement, there are continuous positive affine maps θB,1 : Aff(T(B)) →
RsB and θB,2 : RSB → Aff(T(B)) such that
(2.3)
kθB,2 ◦ θB,1(fB) − fBk∞ < ε/3,
f ∈ F .
THE CLASSIFICATION OF SIMPLE SEPARABLE UNITAL LOCALLY ASH ALGEBRAS
8
Choose γ ∈ (0, 1) such that
provided dist(x, y) < γ.
kfD(x) − fD(y)k < ε/3,
f ∈ F ,
Define
gγ : [0, 1] ∋ x 7→
x ∈ [0, 1 − γ],
1,
linear, x ∈ [1 − γ, 1 − γ/2],
0,
x ∈ [1 − γ/2, 1];
and consider the linear map
χγ : C(Sn−1) ∋ f 7→ (x 7→ (1 − gγ(kxk)) · f (
x
kxk
)) ∈ C(Dn).
Then χγ is a positive affine map from Aff(T(Mk(C(Sn−1)))) to Aff(T(Mk(C(Dn)))). Note that
(2.4) kfDgγ − (fD − χγ(fDSn−1))k∞ = sup{(1 − gγ(kxk))(fD(x) − fD(
x
kxk
)) : x ∈ Dn} < ε/3
for any f ∈ F .
Pick an open cover U of Dn such that the variation of the function fDgγ on any open subset
in U is at most ε/3 for any f ∈ F . Moreover, one requires that the diameter of each open set in
U should be at most γ/4. Choose {φU : U ∈ U} to be a partition of unity subordinated to U,
and fix xU ∈ U for each U ∈ U.
Put
Uγ = {Ui : U ∩ Sn−1 = Ø},
and write Uγ = {U1, U2, ..., UUγ }.
Since the diameter of each Ui ∈ U is at most γ/4, one has that if U /∈ Uγ, then gγ(xU ) = 0,
and hence
(2.5)
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
fDgγ − XU ∈Uγ
Define
and
(fDgγ)(xU )φU(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
= (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
= (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
fDgγ − XU ∈Uγ
fDgγ −XU ∈U
(fDgγ)(xU )φU − XU /∈Uγ
(fDgγ)(xU )φU(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
≤ ε/3.
(fDgγ)(xU )φU(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
θ1 : Aff(T(A)) ∋ (f, g) 7→ θB,1(f ) ⊕ ((ggγ)(xUi)) ∈ RsB ⊕ RUγ ,
θ2 : RsB ⊕ RUγ ∋ ((ξ1, ..., ξsB), (η1, ..., ηUγ ))
7→ (θB,2((ξ1, ..., ξsB)), χγ(ϕ(θB,2((ξ1, ..., ξsB)))) +
Uγ Xi=1
ηiφUi) ∈ Aff(T(A)).
Then, a straightforward calculation shows that
θ2 ◦ θ1((fB, fD)) = (θB,2(θB,1(fB)), χγ(ϕ(θB,2(θB,1(fB)))) +
Uγ Xi=1
(fDgγ)(xUi)φUi)
THE CLASSIFICATION OF SIMPLE SEPARABLE UNITAL LOCALLY ASH ALGEBRAS
9
By the inductive assumption, one has that for any f ∈ F ,
kfB − θB,2(θB,1(fB))k < ε/3 < ε,
and
Therefore
as desired.
fD − (χγ(fDSn−1) +
fD − (χγ(ϕ(θB,2(θB,1(fB)))) +
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Uγ Xi=1
(fDgγ)(xUi)φUi)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)∞
< (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
UγXi=1
(fDgγ)(xUi)φUi)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)∞
= (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Uγ Xi=1
≤ (cid:13)(cid:13)(cid:13)gγfD −X(fDgγ)(xp)φp(cid:13)(cid:13)(cid:13)∞
(fD − χγ(fDSn−1) −
(fDgγ)(xUi)φUi)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)∞
+ ε/3 (by (2.3))
+ ε/3
≤ ε
(by (2.5)).
+ 2ε/3 (by (2.4))
kθ2 ◦ θ1(f ) − f k < ε,
f ∈ F ,
(cid:3)
Remark 2.11. In fact, as shown in [4] (Lemma 2.6), the statement of Lemma 2.10 always holds
if T(A) is replaced by an arbitrary compact metrizable Choquet simplex.
3. An existence theorem
Let A be an NCCC. The main result in this section (Theorem 3.7) is that any almost compatible
pair (κ, γ) from an NCCC to Q can be lifted to an algebra homomorphism, where κ : K0(A) → Q
and γ : Aff(T(A)) → Aff(T(Q)) ∼= R.
Lemma 3.1. Let P be an m × n matrix with integer entries, and let ξ ∈ Rm with each entry a
positive number (including zero). Assume that each entry of P ξ is rational. Then, for any ε > 0,
there is ζ ∈ Qm with positive entries (including zero) such that
(1) P ξ = P ζ, and
(2) kξ − ζk∞ < ε.
Proof. By deleting the columns of P corresponding to the 0's of ξ, one may assume that each
entry of ξ is strictly positive.
Let us show that
ker P = ker P ∩ Qm.
It is clear that ker P ⊇ Qm ∩ ker P . Note that P is rational, so that one can choose a basis of
ker P (as a real vector space) consisting of rational vectors, from which it follows that
dimR(ker P ) ≤ dimQ(ker P ∩ Qm),
THE CLASSIFICATION OF SIMPLE SEPARABLE UNITAL LOCALLY ASH ALGEBRAS
10
and hence dimR(ker P ) ≤ dimR(ker P ∩ Qm). This forces ker P = Qm ∩ ker P .
Again by the rationality of P , there is a vector η ∈ Qm such that
and hence
P η = P ξ,
P −1({P ξ}) = P −1({P η}) = η + ker P = η + Qm ∩ ker P = η + Qm ∩ ker P .
It is clear that
ξ ∈ η + Qm ∩ ker P .
On noting that any entry of ξ is strictly positive, and all vectors in η + Qm ∩ ker P are rational,
for any ε > 0, it follows that there is ζ ∈ η + Qm ∩ ker P such that
kζ − ξk∞ < ε
and each entry of ζ is positive.
(cid:3)
Lemma 3.2. Let A be a unital subhomogeneous C*-algebra such that Primd(A) has finitely
many connected components for each d. Let (F , ε) be given. Then, for any compatible pair (κ, γ)
satisfying, where κ : K0(A) → Q = K0(Q) and γ : Aff(T(A)) → R = Aff(T(A)), there is a
homomorphism φ : A → Q such that
(1) [φ]0 = κ, and
(2) γ( f )(tr) − tr(φ(f )) < ε,
f ∈ F ,
where tr is the canonical trace of Q. Moreover, φ can be chosen to have finite dimensional range.
Proof. Without loss of generality, one may assume that F is inside the unit ball. Since Q has
unique trace, it is enough to show that for any κ : K0(A) → Q and τ ∈ T(A) with
there is a homomorphism φ : A → Q such that [φ]0 = κ and
τ (p) = κ(p),
p ∈ K0(A),
τ (f ) − tr(φ(f )) < ε,
f ∈ F .
By Corollary 2.5, there is a probability measure µi on each Xi, and αi ∈ [0, 1] with
αi = 1
mXi=1
αiZXi
such that
τ (a) =
mXi=1
where tri is the canonical trace of Mki(C).
tri(Πi(a))dµi,
a ∈ A,
THE CLASSIFICATION OF SIMPLE SEPARABLE UNITAL LOCALLY ASH ALGEBRAS
11
Note that for each projection p ∈ A ⊗ K, one has
(3.1)
τ (p) =
=
mXi=1
mXi=1
αiZXi
rank(Πi(p))
di
dµi
αi
rank(Πi(p))
di
.
Since K0(A) is finitely generated, there are p1, p2, ..., pr ∈ K0(A) which generate K0(A). Con-
sider the matrix
P =
rank(Π1(p1))
rank(Π1(p2))
...
rank(Π2(p1))
rank(Π2(p2))
...
rank(Π1(pr))
rank(Π2(pr))
d1
d2
· · ·
· · ·
...
· · ·
· · ·
rank(Πm(p1))
rank(Πm(p2))
...
rank(Πm(pr))
dm
,
and the vector
By (3.1), one has
ξ = (
α1
d1
,
α2
d2
, ...,
αm
dm
)T.
P ξ = (τ (p1), τ (p2), ..., τ (pr), 1)T = (κ(p1), κ(p2), ..., κ(pr), 1)T ∈ Qr+1.
Then, by Lemma 3.1, there is a positive rational vector ξ ∈ Qm such that
(3.2)
Write
P ξ = P ζ
and
kξ − ηk∞ < ε/2l(k1 + · · · + kl).
ζ = (
β1
d1
,
β2
d2
, ...,
βm
dm
)T.
By (3.2), one has that
(3.3)
αi − βi < ε/2m,
1 ≤ i ≤ m.
Since P ξ = P ζ, one has
mXi=1
βi
rank(πx(pj))
di
=
mXi=1
αi
rank(πx(pj))
di
,
j = 1, ..., r.
Since p1, p2, ..., pl generate K0(A), one has
(3.4)
mXi=1
βi
rank(πx(p))
di
=
mXi=1
αi
rank(πx(p))
di
,
p ∈ K0(A).
It also follows from P ξ = P ζ that β1 + · · · + βm = 1.
Consider τ ′ ∈ T(A) defined by
τ ′(a) =
βiZXi
mXi=1
tri(Πi(a))dµi,
a ∈ A.
THE CLASSIFICATION OF SIMPLE SEPARABLE UNITAL LOCALLY ASH ALGEBRAS
12
Then, by (3.4) and (3.3), one has
and
τ ′(p) = τ (p),
p ∈ K0(A),
τ (f ) − τ ′(f ) < ε/2,
f ∈ F .
Consider each probability measure µi, and choose a discrete measure µi such that
Write
tri(Πi(f ))dµi −ZXi
(cid:12)(cid:12)(cid:12)(cid:12)
ZXi
tri(Πi(f ))dµi(cid:12)(cid:12)(cid:12)(cid:12) < ε/2,
liXk=1
δxik,
1
li
µi =
f ∈ F .
for some xik ∈ Xi, where δxik is the Dirac measure concentrated at xik. Without loss of generality,
one may assume that all li are the same and equal to l for some l.
Define
One then has
and
τ (a) =
βiZXi
mXi=1
tri(Πi(a))dµi,
a ∈ A.
τ (p) = τ (p),
p ∈ K0(A),
Write βi = qi/q for natural numbers qi ≤ q. Since P βi = 1, one has that
τ (f ) − τ (f ) < ε,
f ∈ F .
X qi = q.
Define a unital homomorphism by
φ : A ∋ a 7→
mMi=1
lMk=1
where N = d1 · · · dmlq. Then
(πxik (a) ⊕ · · · ⊕ πxik(a)
) ∈ MN (C),
{z
d1···di−1di+1···dmqi
}
i=1Pl
tr(φ(a)) = Pm
mXi=1
mXi=1
qi
q
1
l
=
=
lXk=1
k=1(d1 · · · di−1di+1 · · · dmqi)Tr(πxik(a))
d1 · · · dnql
d1 · · · di−1di+1 · · · dmdi
d1 · · · dnq
qi
1
l
lXk=1
tri(πxik(a))
(
tri(πxik(a))) = τ (a).
Let ι : MN (C) → Q be an unital embedding. Then the homomorphism ι◦φ satisfies the condition
of the lemma.
(cid:3)
THE CLASSIFICATION OF SIMPLE SEPARABLE UNITAL LOCALLY ASH ALGEBRAS
13
Lemma 3.3. Let A be a NCCC. For any finite set F ⊆ A and any ε > 0, there is a finite set
H ⊆ A+ such that for any σ > 0, there is δ > 0 such that if τ ∈ T(A) satisfies
τ (h) > σ,
h ∈ H,
there exists τ ′ ∈ T(A) such that
(1) τ ′(p) = τ (p), p ∈ K0(A),
(2) τ ′(f ) − τ (f ) < ε, f ∈ F , and
(3) αi(τ ′) > δ, all i.
Moreover, one may require that δ depends only on σ.
Proof. If A = Mn1(C) ⊕ · · · ⊕ Mnk(C). Then H = {1Mni (C) : 1 ≤ i ≤ k} satisfies the lemma with
δ = σ for any given σ (in fact, one has that τ ′ = τ in this case).
Let
A = B ⊕Mk(C(Sn−1)) Mk(C(Dn)),
where B is NCCC, and assume that the lemma holds for B.
Let F ⊆ A, ε > 0 be given. For each f ∈ A, write f = (fB, fD) for fB ∈ B and fD ∈
Mk(C(Dn)). Since B is assumed to satisfy the lemma, there is HB such that for any σ, there is
δB(σ) > 0 such that if τB ∈ T(B) satisfies
there is τ ′
B ∈ T(A) such that
and
Choose γ > 0 such that
τB(h) > σ,
h ∈ HB,
τ ′
B(p) = τB(p),
p ∈ K0(B),
τ ′
B(fB) − τB(fB) < ε/2,
f ∈ F ,
αi(τ ′
B) > δB(σ),
all i.
kfD(x) − fD(y)k < ε/2,
f ∈ F ,
if dist(x, y) < γ.
Define
gγ : [0, 1] ∋ x 7→
x ∈ [1 − γ/2, 1],
1,
linear, x ∈ [1 − γ, 1 − γ/2],
0,
x ∈ [0, 1 − γ/2];
similarly, define gγ/2 and g2γ.
For each h ∈ HB define hD ∈ Mk(C(Dn)) by
hD(x) =
0,
4kxk+2γ−4
γ
kxk ∈ [0, 1 − γ/2],
ϕ(h)( x
kxk), kxk ∈ [1 − γ/2, 1 − γ/4],
ϕ(h)( x
kxk),
kxk ∈ [1 − γ/4, 1],
where ϕ : B → Mk(C(Sn−1)) is the gluing map.
For each h ∈ HB, define
h = (h, hD) ∈ A = B ⊕Mk(C(Sn−1)) Mk(C(Dn)).
THE CLASSIFICATION OF SIMPLE SEPARABLE UNITAL LOCALLY ASH ALGEBRAS
14
Also define g2γ = (0, g2γ1Mk(C)), gγ/2 = (0, gγ/21Mk(C)) ∈ A.
Then
H = {h, g2γ, 1 − gγ/2 : h ∈ HB}
satisfies the condition of the lemma.
Indeed, for any σ > 0, set
Let τ ∈ T(A) be such that
δ = σδB(σ).
τ (h) > σ,
h ∈ H.
Without loss of generality, one may assume that
τ ((f, g)) = ατB(f ) + βZDn\Sn−1
gdµ,
where µ is a discrete probability measure of Dn \ Sn−1. Since τ (gγ) > σ, one has
β > σ.
Write µ =Pi βiδxi. For each xi with kxik ≥ γ
Then define the trace
τB,xi(f ) =
2 , define τB,xi ∈ T(B) by
1
k
Tr(ϕ(f )(
xi
kxik
)).
τ ((f, g)) = ατB(f ) + β( Xkxik≥ γ
= α′τB(f ) + β′Z(Dn)◦
2
gdµ′,
βiτB,xi(f )) + β( Xkxik< γ
2
βi)Z(Dn)◦
gd(
1
Pkxik< γ
2
βi Xkxik< γ
2
βiδxi)
τB(f ) =
2
ατB(f ) + β(Pkxik≥ γ
α +Pkxik≥ γ
βi Xkxik< γ
Pkxik< γ
1
2
2
2
µ′ =
βiτB,xi(f ))
ββi
βiδxi,
ββi,
, α′ = α + Xkxik≥ γ
β′ = β( Xkxik< γ
βi).
2
2
where
and
Note that
(3.5)
(3.6)
Noting that
one has
By the choice of γ, one has
τ (p) = τ (p),
p ∈ K0(A).
τ (f ) − τ (f ) < ε/2,
f ∈ F .
τ (g2γ) = τ (g2γ) and τ (1 − gγ/2) ≥ τ (1 − gγ/2),
τ (g2γ) > σ and τ (1 − gγ/2) > σ.
THE CLASSIFICATION OF SIMPLE SEPARABLE UNITAL LOCALLY ASH ALGEBRAS
15
A straightforward calculation shows that
β′ ≥ τ (g2γ)
and α′ = τ (1 − gγ/2),
and therefore
(3.7)
β′ > σ and α′ > σ.
Also noting that for any h ∈ HB, one has
and
Therefore,
τ (h) ≥ τ (h) > σ,
τ (h) = α′τB(h).
By the inductive assumption, there is τ ′
B ∈ T(B) such that
τB(h) > σ/α′ > σ,
h ∈ HB.
(3.8)
(3.9)
and
(3.10)
Then the trace
τ ′
B(p) = τB(p),
p ∈ K0(B),
τ ′
B(fB) − τB(fB) < ε/2,
f ∈ F ,
αi(τ ′
B) > δB(σ),
all i.
τ ′(f, g) = α′τ ′
B(f ) + β′Z(Dn)◦
gdµ′
satisfies the condition of the lemma. By (3.5) and (3.8), one has
τ ′(p) = τ (p),
p ∈ K0(A),
Indeed, by (3.6) and (3.9), one has
(3.11)
τ ′(f ) − τ (f ) < ε,
f ∈ F .
By (3.7) and (3.10), one has
as desired.
αi(τ ′) > σδB(σ),
all i,
(cid:3)
Lemma 3.4. Let P be a m × n matrix. Assume that m ≥ n and rank(P ) = n. Let σ > 0 be
given. Then, for any ε > 0, there is δ > 0 such that for any vectors ξ ∈ Rm and κ ∈ Rn with
(1) ξi > σ, i = 1, ..., m, and
(2) kκ − P ξk∞ < δ,
there is ζ ∈ Rm with ζi ≥ 0, i = 1, ..., m such that
(3) P ζ = κ, and
(4) kξ − ζk∞ < ε.
THE CLASSIFICATION OF SIMPLE SEPARABLE UNITAL LOCALLY ASH ALGEBRAS
16
Proof. Regard P as a linear map from Rm → Rn. Since rank(P ) = n, one has that P is surjective.
Therefore, for the given ǫ < σ, there is δ > 0 such that
(3.12)
Bk·k∞(0, δ) ⊆ P (Bk·k∞(0, ε)).
Then the δ satisfies the condition of the lemma. Indeed, let ξ ∈ Rm and κ ∈ Rn be given, and
satisfy
Then
kκ − P ξk∞ < δ.
κ − P (ξ) ∈ Bk·k∞(0, δ) ⊆ Rn.
By (3.12), there is θ ∈ Bk·k∞(0, ε) ⊆ Rm such that
and hence
P (θ) = κ − P (ξ),
P (ξ + θ) = κ.
Since each entries of ξ is at least σ > ε, one has that
and therefore
is the desired vector.
ξ + θ ∈ (R+)m,
ζ = ξ + θ
(cid:3)
Remark 3.5. Let A be an NCCC, and let κ : K0(A) → Q be a positive map. Since A is of type
I, it is exact; therefore, κ (regarded as a state of K0(A)) is induced by a trace. That is, there is
τ ∈ T(A) such that
In particular, this implies that κ factors through ρA(K0(A)).
κ(p) = τ (p),
p ∈ K0(A).
Lemma 3.6. Let A be an NCCC. Let (F , ε) be given. Let p1, p2, ..., pn ∈ K0(A) be such that
{p1, ..., pn} is a set of generators for ρA(K0(A)) (as an abelian group) (still use the same notation
for the image of pi). Then, there is a finite set H ⊆ A+ such that for any σ > 0, there is δ > 0
such that if κ : K0(A) → Q and τ ∈ T(A) satisfy
(1) τ (h) > σ, h ∈ H, and
(2) τ (pi) − κ(pi) < δ, i = 1, ..., n,
then there is τ ′ ∈ T(A) such that
(3) τ ′(f ) − τ (f ) < ε, f ∈ F , and
(4) τ ′(pi) = κ(pi), i = 1, ..., n.
Proof. Without loss of generality, one may assume that F is contained in the unit ball of A. One
may also assume that p1, p1, ..., pn ∈ ρA(K0(A)) are Q-independent.
Let H ⊆ A+ be the subset of Lemma 3.3 with respect to A, F , and ε/2. Then H satisfies the
lemma.
THE CLASSIFICATION OF SIMPLE SEPARABLE UNITAL LOCALLY ASH ALGEBRAS
17
In fact, for any σ > 0, let δ′ be the constant of Lemma 3.3 with respect to σ. Consider the
m × n matrix
P =
rank(Π1(p1))
rank(Π2(p1))
d1
rank(Π1(p2))
d2
rank(Π2(p2))
d1
...
d2
...
rank(Π1(pn))
rank(Π2(pn))
d1
d2
· · ·
· · ·
...
· · ·
rank(Πm(p1))
dm
rank(Πm(p2))
dm
...
rank(Πm(pn))
dm
.
Since p1, p1, ..., pn ∈ ρA(K0(A)) are Q-independent, the matrix P has rank n and it satisfies the
assumptions of Lemma 3.4. Applying Lemma 3.4 to δ′ (in the place of σ) and ε/2m, one obtains
δ.
Let (κ, τ ) be a pair satisfies
and
(3.13)
τ (pi) − κ(pi) < δ,
i = 1, ..., n,
τ (h) > σ,
h ∈ H.
By (3.13) and Lemma 3.3, there is τ ′ ∈ T(A) such that
(3.14)
and
Set
τ ′(pi) = τ (pi),
τ ′(f ) − τ (f ) < ε/2,
i = 1, ..., n,
f ∈ F ,
αi(τ ′) > δ′,
∀i.
ξ = (α1(τ ′), α2(τ ′), ..., αm(τ ′))T
and κ = (κ(p1), κ(p2), ..., κ(pn))T,
and then one has
kP ξ − κk∞ = max{τ ′(pi) − κ(pi) : i = 1, ..., n}
= max{τ (pi) − κ(pi) : i = 1, ..., n} (by (3.14))
< δ
By Lemma 3.4, there is a positive vector ζ = (β1, ..., βm) such that
(3.15)
and
Consider the trace
It is clear that
By (3.15), one has
P ζ = (κ(p1), κ(p2), ..., κ(pn))T
αi − βi < ε/2m,
∀i.
τ ′′ =X βiµi.
τ ′′(f ) − τ (f ) < ε,
f ∈ F .
τ ′′(p) = κ(p),
p ∈ K0(A).
Moreover, the trace τ ′′ is indeed a state since τ ′′(1A) = κ([1A]) = 1, as desired.
(cid:3)
THE CLASSIFICATION OF SIMPLE SEPARABLE UNITAL LOCALLY ASH ALGEBRAS
18
Theorem 3.7. Let A be a unital NCCC, and let σ > 0. Let (F , ε) be given. Let p1, p2, ..., pn ∈
K0(A) be such that the set {p1, ..., pn} generates the group ρA(K0(A)) (let us still use the same
notation for the image of pi). Then, there is a finite set H ⊆ A+ such that for any σ > 0, there
is δ > 0 such that if κ : K0(A) → Q and τ ∈ T(A) satisfy
(1) τ (pi) − κ[pi] < δ, i = 1, ..., n, and
(2) τ (h) > σ, h ∈ H,
then there is a homomorphism φ : A → Q such that
(3) [φ]0 = κ, and
(4) τ (f ) − tr(φ(f )) < ε, f ∈ F ,
where tr is the canonical trace of Q. Moreover, φ can be chosen to have finite dimensional range.
Proof. Let H ⊆ A+ denote the finite set of Lemma 3.6 with respect to the data A, F , ε/2. Then
H satisfies the theorem.
Indeed, given σ > 0, consider the constant δ of Lemma 3.6 with respect to σ. Let (κ, τ ) be a
pair as above -- δ-compatible on pi, i = 1, ..., n, and such that
By Lemma 3.6, there is τ ′ ∈ T(A) such that the pair (κ, τ ′) is exactly compatible on the pi and
τ (h) > σ,
h ∈ H.
Then, by Lemma 3.2, there is a homomorphism φ : A → Q such that
τ ′(f ) − τ (f ) < ε/2,
f ∈ F .
and
[φ]0 = κ
tr(φ(f )) − τ ′(f ) < ε/2,
f ∈ F ,
and therefore satisfying the condition of the theorem.
(cid:3)
4. A decomposition theorem and a uniqueness theorem
Recall (see, for instance, [10])
Lemma 4.1 (The Marriage Lemma). Let A and B be two finite subsets of a metric space.
Suppose that for any set X ⊆ A, one has
#{y ∈ B : dist(y, X) < ε} ≥ #X,
then there is a set B′ ⊆ B and a one-to-one pairing between A and B′ such that the distance
between the points in each pair is at most ε.
The following statement is a generalization of the Marriage Lemma due to Gong in a unpub-
lished manuscript:
THE CLASSIFICATION OF SIMPLE SEPARABLE UNITAL LOCALLY ASH ALGEBRAS
19
Lemma 4.2. Let A, B be two finite subsets of a metric space with subsets A1 ⊂ A and B1 ⊂ B.
Suppose that for each X ⊂ A1, one has
and for each Y ⊂ B1, one has
#{y ∈ B : dist(y, X) < ε} ≥ #X
#{x ∈ A : dist(x, Y ) < ε} ≥ #Y.
Then there are sets A2 ⊆ A and B2 ⊆ B such that
A1 ⊆ A2
and B1 ⊆ B2,
and the elements of A2 and B2 can be (bijectively) paired to within ε.
Proof. The proof is similar to that of [10]. We spell out the details for reader's convenience. In
the case #(B1) = 0, i.e., B1 is empty, then this is the case of the classical Marriage Lemma
(Lemma 4.1) in the case of A1 and B. (For this case, X0 can be chosen to be A1.)
We define partial order for (m, n) ∈ Z+ × Z+ (where Z+ = {0, 1, 2, ...} is the set of positive
integers) by (m, n) ≤ (m1, n1) if m ≤ m1 and n ≤ n1. We denote (m, n) < (m1, n1) if (m, n) ≤
(m1, n1) and (m, n) 6= (m1, n1). We will prove the lemma by induction on #(A1) and #(B1).
That is, we assume that if the result is true for the case (cid:0)#(A1), #(B1)(cid:1) < (m, n) and prove the
lemma to be true for the case
#(A1) = m and #(B1) = n.
The rest of the proof divides into two cases.
Case 1. For any nonempty set X ⊆ A1,
#{y ∈ B; dist(y, X) < ε} ≥ #(X) + 1
and for any nonempty set Y ⊆ B1,
#{x ∈ A; dist(x, Y ) < ε} ≥ #(Y ) + 1.
Choose any a ∈ A1, there is a b ∈ B such that dist(a, b) < ε. We will pair a ∈ A1 with b ∈ B.
Then let A = A\{a} with A1 = A1\{a}, and B = B \{b} with B1 = B1 if b 6∈ B1 or B1 = B1\{b}
if b ∈ B1. It is easy to verify that A ⊇ A1 and B ⊇ B1 satisfy the condition of the lemma with
(cid:0)#( A1), #( B1)(cid:1) = either (m − 1, n) or (m − 1, n − 1). That is, (cid:0)#(A1), #(B1)(cid:1) < (m, n). By the
induction assumption, there is a subset A2 ⊇ A1 and B2 ⊇ B1 such that X0 can be paired one
by one within ε. Then the sets A2 = A2 ∪ {a} and B2 = B2 ∪ {b} satisfy the lemma.
Case 2. The conditions of Case 1 do not hold. Then either there is X ⊆ A1 or Y ⊆ B1 does
not satisfy the condition in Case 1. That is, either there is X ⊆ A1 such that
#{y ∈ B; dist(y, X) < ε} = #(X),
or there is Y ⊆ B1 such that
#{x ∈ A; dist(x, Y ) < ε} = #(Y ).
Without loss of generality, let us assume there is Y1 ⊆ B1 such that
#{x ∈ A; dist(x, Y1) < ε} = #(Y1).
THE CLASSIFICATION OF SIMPLE SEPARABLE UNITAL LOCALLY ASH ALGEBRAS
20
Let X1 = {x ∈ A; dist(x, Y1) < ε}. Then for any subset Z ⊆ Y1, {x ∈ A; dist(x, Z) < ε} ⊆ X1
and therefore {x ∈ A; dist(x, Z) < ε} = {x ∈ X1; dist(x, Z) < ε} which has at least #(Z)
elements. That is, Y1 and X1 satisfy the condition for the classical Pairing Lemma with Y1 =
A, X1 = B as in 2.10. Hence Y1 and X1 can be paired one by one to within ε.
Let B = B \ Y1 with subset B1 = B1 \ Y1, and let A = A \ X1 with subset A1 = A1 \ (X1 ∩ A1).
Let us verify that B ⊇ B1 and A ⊇ A1 satisfy the condition of the lemma. Let Y ⊆ B1. Then
{x ∈ A; dist(x, Y ∪ Y1) < ε} = {x ∈ A1; dist(x, Y ) < ε} ∪ X1.
Since #(X1) = #(Y1) and {x ∈ A; dist(x, Y ∪ Y1) < ε} ≥ #(Y ) + #(Y1) = #(Y ) + #(X1), we
have, {x ∈ A1; dist(x, Y ) < ε} ≥ #(Y ).
Let X ⊂ A1. Since for point y ∈ Y1, dist(y, X) ≥ ε, we have
#{y ∈ B; dist(y, X) < ε} = #{y ∈ B; dist(y, X) < ε} ≥ #(X).
So the conditions of our lemma hold for A ⊇ A1 and B ⊇ B1 with
#( B1) = #(B1) − #(Y1) < n and #( A1) ≤ #(A1) ≤ m.
So by the inductive assumption, there exist A2 ⊇ A1 and B2 ⊇ B1 such that A2 and B2 can be
paired element by element to within ε. Then the sets A2 = A2 ∪ X1 and B2 = B2 ∪ Y1 satisfy
the lemma.
(cid:3)
Definition 4.3 (See 2.2 of [9]). Let A be a unital C*-algebra with T(A) 6= Ø. Recall that each
self-adjoint a ∈ A induces a ∈ Aff(T(A)) by a(τ ) = τ (a), τ ∈ T(A). Denote this map by ρA.
1 the set of positive elements with norm at most 1, and then denote by A+
Denote by A+
1,q the
image of A+
1 in Aff(T(A)) under the canonical map ρA.
As a consequence of this generalized version of the Marriage Lemma, one has
Lemma 4.4. Let A = B ⊕Mk(C(Sn−1)) Mk(C(Dn)) be an NCCC with n ≥ 1. Let ∆ : A+
1,q \ {0} →
(0, 1) be an order preserving map. Let F ⊆ A be a finite set, and let ε > 0, 1 > γ > 0, and
M ∈ N. There are finite sets H1, H2 ⊆ A+ and δ > 0 such that for any unital homomorphisms
φ, ψ : A → Mm(C) such that
(1) τ (φ(h)), τ (ψ(h)) > ∆(h), h ∈ H1, and
(2) τ (φ(h)) − τ (ψ(h)) < δ, h ∈ H2,
there are unital homomorphisms φ′, ψ′ : A → Mm(C) such that
(3) kφ′(f ) − φ(f )k < ε, kψ′(f ) − ψ(f )k < ε, f ∈ F ,
(4) SP(φ′) ∩B(1 −γ) = SP(ψ′) ∩B(1 −γ), and each point in SP(φ′) ∩B(1 −γ) has multiplicity
at least M,
where B(1 − γ) ⊆ Dn is the closed ball with radius 1 − γ.
Proof. Let (F , ε) be given. For each f ∈ F , write f = (fB, fD) where fB ∈ B and fD ∈
Mk(C(Dn)). Set
FB = {fB : f ∈ F } and FD = {fD : f ∈ F }.
THE CLASSIFICATION OF SIMPLE SEPARABLE UNITAL LOCALLY ASH ALGEBRAS
21
Choose η > 0 such that for any x, y ∈ Dn with dist(x, y) < 4η, one has
(4.1)
kfD(x) − fD(y)k < ε,
f ∈ F .
Without loss of generality, one may assume that 5η < γ.
Fix an η-dense finite subset {x1, x2, ..., xl} ⊆ B(1 − γ). For each subset
define
{y1, y2, ..., yt} ⊆ {x1, x2, ..., xl},
h{y1,...,yt} = max{1 −
1
η
dist(x, {y1, ..., yt}η), 0},
where Yη denotes the η-neighborhood of Y if Y is a subset of a metric space. Using
{y1, y2, ..., yt}2η 6= {y1, y2, ..., yt}3η,
choose a positive function g{y1,...,yt} ∈ C(Dn) such that 0 < g{y1,...,yt} ≤ 1 and
For functions hy1,...,yt and gy1,...,yt, regard them as the elements
supp(g{y1,...,yt}) ⊆ {y1, ..., yt}3η \ {y1, ..., yt}2η.
(0, hy1,...,yt1Mk(C(Dn))), (0, gy1,...,yt1Mk(C(Dn))) ∈ A = B ⊕Mk(C(Sn−1)) Mk(C(Dn)),
and still denote them by hy1,...,yt and gy1,...,yt, respectively.
Put
and
(4.2)
H1 = {g{y1,...,yt} : {y1, ..., yt} ⊆ {x1, x2, ..., xl}},
H2 = {h{y1,...,yt} : {y1, ..., yt} ⊆ {x1, x2, ..., xl}},
δ = min{∆(g{y1,...,yt}) : {y1, ..., yt} ⊆ {x1, ..., xl}}}.
Also pick a finite open cover U of B(1 − γ) such that each U ∈ U has diameter at most η,
[U ∈U
U ⊆ B(1 − γ/2) and U \ [V 6=U
V 6= Ø.
Then choose continuous functions sU,1, ..., sU,M ∈ C(Dn) such that
supp(sU,i) ⊆ [V 6=U
V
and supp(sU,i) ∩ supp(sU,j) = Ø, i 6= j.
Regard each sU,i as an element of A, and put
S = {sU,i : U ∈ U, i = 1, ..., M}.
Then H1 := H1 ∪ S, H2 and δ have the property stated in the conclusion of the lemma.
Indeed, let φ, ψ : A → Mm(C) be unital homomorphisms satisfying
(4.3)
and
(4.4)
τ (φ(h)), τ (ψ(h)) > ∆(h),
h ∈ H1,
τ (φ(h)) − τ (ψ(h)) < δ,
h ∈ H2.
THE CLASSIFICATION OF SIMPLE SEPARABLE UNITAL LOCALLY ASH ALGEBRAS
22
Let Y ⊆ SP(φ) ∩ B(1 − γ) be an arbitrary subset. Let
denote the subset of the points y satisfying dist(y, Y ) < η. Then
{y1, y2, ..., yt} ⊆ {x1, x2, ..., xl}
#Y ≤ m · tr(φ(h{y1,...,yt}))
≤ m · tr(ψ(h{y1,...,yt})) + m · δ
≤ m · tr(ψ(h{y1,...,yt})) + m · ∆(gy1,...,yt)
≤ #(SP(ψ) ∩ {y1, ..., yt}2η) + m · ∆(gy1,...,yt)
≤ #(SP(ψ) ∩ {y1, ..., yt}3η)
(by (4.3))
(by (4.4))
(by (4.2))
≤ #(SP(ψ) ∩ Y4η).
The same argument shows that
#X ≤ #(SP(φ) ∩ X4η), X ⊆ SP(ψ) ∩ B(1 − γ).
Then, applying Lemma 4.2 with
and
A = SP(φ) ∩ B(1 − η), A1 = SP(φ) ∩ B(1 − γ)
B = SP(ψ) ∩ B(1 − η), B1 = SP(ψ) ∩ B(1 − γ),
one obtains A2 and B2 such that
(4.5)
and
(4.6)
SP(φ) ∩ B(1 − γ) ⊆ A2 ⊂ SP(φ) ∩ B(1 − η)
SP(ψ) ∩ B(1 − γ) ⊆ B2 ⊂ SP(ψ) ∩ B(1 − η),
and A2 and B2 can be paired up to 4η.
Write
A2 = {zφ,1, zφ,2, ..., zφ,s} and B2 = {zψ,1, zψ,2, ..., zψ,s}
for some s, where
(4.7)
dist(zφ,i, zψ,i) < 4η,
i = 1, ..., s.
Then, up to unitary equivalence, there are decompositions
φ = φ ⊕
sMi=1
πzφ,i
and ψ = ψ ⊕
sMi=1
πzψ,i,
where SP φ ∩ B(1 − γ) = SP ψ ∩ B(1 − γ) = Ø, and the homomorphisms
φ′ = φ = φ ⊕
sMi=1
πzφ,i
and ψ′ = ψ ⊕
sMi=1
πzφ,i
have the required properties except the requirement on multiplicity.
Indeed, by (4.7) and (4.1), one has
kφ(f ) − φ′(f )k = 0 < ε
and
kψ(f ) − ψ′(f )k < ε,
f ∈ F .
THE CLASSIFICATION OF SIMPLE SEPARABLE UNITAL LOCALLY ASH ALGEBRAS
23
By (4.5) and (4.6), one has that
SP(φ′) ∩ B(1 − γ) = {zφ,1, ..., zφ,s} ∩ B(1 − γ) = SP(ψ′) ∩ B(1 − γ).
To satisfy the multiplicity condition, one needs to perturb φ′ and ψ′ further. Since
φ(sU,i) > ∆(sU,i) > 0, U ∈ U, i = 1, ..., M,
one has that for any U ∈ U, at least M of {zφ1, ..., zφs} are inside U \ SV 6=U V (counting
multiplicity). Then there is a grouping of grouping {zφ1, ..., zφs} such that each group is insider
at most one open set U and has elements at least M if it is covered by an open set U. Therefore,
after a perturbation, one may assume that each point in {zφ1, ..., zφs} ∩ B(1 − γ) has multiplicity
at least M, and hence φ′ and ψ′ satisfy the desired multiplicity condition.
(cid:3)
Theorem 4.5. Let A = B ⊕Mk(C(Sn−1)) Mk(C(Dn)) be an NCCC, and let ∆ : A+
1,q \{0} → (0, 1) be
an order preserving map. Let F ⊆ A be a finite set. Let ε > 0, η > 0 and K ∈ N \ {0}. There are
finite sets H1, H2 ⊆ A+ and δ > 0 such that for any unital homomorphisms φ, ψ : A → Mm(C)
such that
(1) τ (φ(h)), τ (φ(h)) > ∆(h),
(2) τ (φ(h)) − τ (ψ(h)) < δ,
h ∈ H1, and
h ∈ H2,
there exist unital homomorphisms φ, ψ : A → Mm(C) such that
(3) (cid:13)(cid:13)(cid:13) φ(f ) − φ(f )(cid:13)(cid:13)(cid:13) < ε, (cid:13)(cid:13)(cid:13) ψ(f ) − ψ(f )(cid:13)(cid:13)(cid:13) < ε, f ∈ F ,
(4) φ and ψ have decompositions
φ = φ0 ⊕ φ1 ⊕ · · · ⊕ φ1
and
ψ = ψ0 ⊕ ψ1 ⊕ · · · ⊕ ψ1
such that φ1 and ψ1 are unitarily equivalent, and
{z
K
}
K
{z
}
tr( φ0(a)) ≤ η · tr( φ(a))
and tr( ψ0(a)) ≤ η · tr( ψ(a)),
a ∈ F .
Proof. The statement holds if A is finite dimensional. Assume that the statement holds for B,
and let us show that the statement holds for A.
Let (F , ε) be given. For each f ∈ F , write f = (fB, fD) where fB ∈ B and fD ∈ Mk(C(Dn)).
Set
FB = {fB : f ∈ F } and FD = {fD : f ∈ F }.
For each r > 0, define
gr : [0, 1] ∋ x 7→
x ∈ [0, 1 − r],
1,
linear, x ∈ [1 − r, 1 − r/2],
0,
x ∈ [1 − r/2, 1];
Also define gr = (0, gr(kxk)1Mk(C(Dn))) ∈ A.
Pick γ > 0 such that
(4.8)
kfD(x) − fD(y)k < ε/8,
f ∈ F , dist(x, y) < γ,
and consider gγ.
THE CLASSIFICATION OF SIMPLE SEPARABLE UNITAL LOCALLY ASH ALGEBRAS
24
Consider the linear map
χγ : Mk(C(Sn−1)) ∋ f 7→ (x 7→ (1 − gγ(kxk)) · f (
x
kxk
)) ∈ Mk(C(Dn))
Clearly, χγ is a positive linear map, and then the map
Γγ : B ∋ f 7→ (f, χγ(ϕ(f ))) ∈ A
is positive and injective, where ϕ : B → Mk(C(Sn−1)) is the gluing map.
If f ∈ B satisfies that τ (f ) = 0, τ ∈ T(B), then τ (χ(f )) = 0, τ ∈ T(Mk(C(Dn))), and
therefore
Hence the map
τ (Γγ(f )) = 0,
τ ∈ T(A).
∆B : B1
is well defined and is order preserving.
q ∋ f 7→ ∆(Γγ(f )) ∈ (0, 1)
Applying the inductive hypothesis to B with ∆B, FB, ε/2, K and η, one obtains HB,1, HB,2,
and δB.
Define
Put
HB,1 = {Γγ(h) : h ∈ HB,1} and
HB,2 = {Γγ(h) : h ∈ HB,2}.
Let γ be a positive number such that
σB = min{∆(h) : h ∈ HB,1}.
(4.9)
kχγ(h)(x) − χγ(h)(y)k < min{
∆( \1 − gγ/2)δB
4
,
σB
4
,
ε
8
},
h ∈ HB,1 ∪ HB,2 ∪ FD
for any x, y ∈ Dn satisfying dist(x, y) < γ.
Let HD,1 ⊆ A, HD,2 ⊆ A and δD denote the finite sets and constant of Lemma 4.4 with respect
to F ∪ HB,1 ∪ HB,2 (in the place of F ), min{ε/8, ∆( \1 − gγ/2)δB/8, σB/8} (in the place of ε), γ,
M = ⌊2K/η⌋ + 1, and ∆.
Then
and
H1 = HB,1 ∪ HD,1 ∪ {1 − gγ/2, gγ}, H2 = HB,2 ∪ HD,2,
δ = min{∆( \1 − gγ/2)δB/4, σB/4, δD}
satisfy the statement.
In fact, let φ, ψ : A → Mm(C) be unital homomorphisms satisfying
(4.10)
and
(4.11)
τ (φ(h)), τ (φ(h)) > ∆(h),
h ∈ H1
τ (φ(h)) − τ (ψ(h)) < δ,
h ∈ H2.
Since δ ≤ δD, by Lemma 4.4, there are homomorphisms
φ′, ψ′ : A → Mm(C)
THE CLASSIFICATION OF SIMPLE SEPARABLE UNITAL LOCALLY ASH ALGEBRAS
25
such that
(4.12)
(4.13)
and
kφ′(f ) − φ(f )k < min{ε/8, ∆( \1 − gγ/2)δB/8, σB/8},
f ∈ F ∪ HB,1 ∪ HB,2,
kψ′(f ) − ψ(f )k < min{ε/8, ∆( \1 − gγ/2)δB/8, σB/8},
f ∈ F ∪ HB,1 ∪ HB,2,
Sp(φ′) ∩ B(1 − γ) = Sp(ψ′) ∩ B(1 − γ) = {x1, x2, ..., xl},
for some x1, x2, ..., xl ∈ B(1 − γ) with multiplicity at least M. Therefore, up to unitary equiva-
lence, there are decompositions
(4.14)
(4.15)
φ′ = φ′
B ⊕ (
πxφ,i) ⊕ (
mφMi=1
mψMi=1
lMi=1
lMi=1
πxi),
πxi).
ψ′ = ψ′
B ⊕ (
πxψ,i) ⊕ (
for some mφ, mψ ∈ N and some xφ,1, ..., xφ,mφ, xψ,1, ..., xψ,mψ ∈ Dn with 1 > kxφ,ik ≥ γ and
1 > kxψ,jk ≥ γ.
By (4.10), (4.12) and (4.13), one has
(4.16)
tr(φ′(h)), tr(ψ′(h)) >
∆(h),
h ∈ HB,1.
7
8
It also follows from (4.11) (4.12) and (4.13) that for any h ∈ HB,1 ∪ HB,2,
(4.17)
tr(φ′(h)) − tr(ψ′(h)) < tr(φ′(h)) − tr(ψ′(h)) + ∆( \1 − gγ/2)δB/4
< δ + ∆( \1 − gγ/2)δB/4 ≤ ∆( \1 − gγ/2)δB/2.
Therefore, by the decompositions (4.14) and (4.15), one has
(4.18)
tr(φ′
B(h) ⊕ (
mφMi=1
h(xφ,i))) − tr(ψ′
B(h) ⊕ (
mψMi=1
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
h(xψ,i)))(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
< ∆( \1 − gγ/2)δB/2
For each point xφ,i (or xψ,j), replace the homomorphism πxφ,i (or πxψ,i) by the homomorphism
ψ,i = xψ,i/ kxψ,ik ∈ Sn−1). By the choice of
φ,i = xφ,i/ kxφ,ik ∈ Sn−1 (or x′
), where x′
(or πx′
πx′
γ (see (4.9)), one has
ψ,i
φ,i
and
(4.19) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
mφMi=1
(4.20) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
mψMi=1
Put
πx′
φ,i
(h) −
πx′
ψ,i
(h) −
mφMi=1
mψMi=1
πxφ,i(h)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
πxψ,i(h)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
mφMi=1
B ⊕ (
< min{
< min{
∆( \1 − gγ/2)δB
4
,
σB
4
,
ε
8
},
h ∈ HB1,1 ∪ HB1,2 ∪ FD,
∆( \1 − gγ/2)δB
4
,
σB
4
,
ε
8
},
h ∈ HB1,1 ∪ HB1,2 ∪ FD.
φ′′
B = φ′
πx′
φ,i
) and ψ′′
B = ψ′
B ⊕ (
mψMi=1
πx′
ψ,i
).
THE CLASSIFICATION OF SIMPLE SEPARABLE UNITAL LOCALLY ASH ALGEBRAS
26
Since x′
φ,i, x′
ψ,j ∈ Sn−1, the homomorphisms φ′′
B and ψ′′
B factor through B. Define
(4.21)
φ′′ = φ′′
B ⊕ (
lMi=1
πxi) and ψ′′ = ψ′′
B ⊕ (
lMi=1
πxi).
By (4.19), (4.20), one has
(4.22)
kφ′(f ) − φ′′(f )k < ε/8 and
kψ′′(f ) − ψ′′′(f )k < ε/8,
f ∈ F .
By (4.19), (4.20), and (4.16), one has
(4.23)
tr(φ′′(h)), tr(ψ′′(h)) >
∆(h),
h ∈ HB,1.
3
4
By (4.19), (4.20) and (4.18), one has
(4.24)
tr(φ′′(h)) − tr(ψ′′(h)) <
3
4
∆( \1 − gγ/2)δB,
h ∈ HB,2.
For each point xi with kxik > 1 − γ, replace the homomorphism πxi by πx′
i =
factors through B. Also note that for each such
, where x′
i
xi/ kxik ∈ Sn−1. Clearly, the homomorphism πx′
xi, one has
i
(4.25)
tr(πx′
i
(h)) > tr(πxi(h)),
h ∈ HB,1.
By the choice of γ (see (4.8)), one also has
(fD) − πxi(fD)(cid:13)(cid:13) < ε/8,
f ∈ F .
i
(cid:13)(cid:13)πx′
B ⊕ ( Mkxik≥1−γ
φ′′′
B = φ′′
πx′
i
)
and ψ′′′
B = ψ′′
B ⊕ ( Mkx′
ik≥1−γ
πx′
i
),
(4.26)
Set
and consider
(4.27)
φ′′′ = φ′′′
B ⊕ ( Mkxik<1−γ
πxi) and ψ′′ = ψ′′′
B ⊕ (
lMkxik<1−γ
πxi).
By (4.26), one has
(4.28)
kφ′′(f ) − φ′′′(f )k < ε/8 and
kψ′′(f ) − ψ′′′(f )k < ε/8,
f ∈ F .
By (4.25) and (4.23), one has
(4.29)
tr(φ′′′(h)), tr(ψ′′′(h)) >
∆(h),
h ∈ HB,1.
3
4
By (4.24), one has
(4.30)
tr(φ′′′(h)) − tr(ψ′′′(h)) <
3
4
∆( \1 − gγ/2)δB,
h ∈ HB,2.
Denote by
Then, by (4.29),
N = rank(φ′′′
B(1)) = rank(ψ′′′
B (1)).
N
m
≥ tr(φ′′′(1 − gγ/2)) >
3
4
∆( \1 − gγ/2),
THE CLASSIFICATION OF SIMPLE SEPARABLE UNITAL LOCALLY ASH ALGEBRAS
27
and
Therefore
(4.31)
m − N
m
≥ tr(φ′′′(gγ)) > ∆(d(gγ)).
< 1 − ∆(d(gγ)).
N
m
3
4
∆( \1 − gγ/2) <
Then, by considering the unital homomorphisms φ′′′
B, ψ′′′
B : B → MN (C), it follows from (4.30)
and (4.31) that
(4.32)
1
N
(Tr(φ′′′
B(h)) − Tr(φ′′′
B(h))) =
m
N
(tr(φ′′′(Γγ(h))) − tr(φ′′′(Γγ(h)))) < δB,
h ∈ HB,2.
It also follows from (4.29) that
and
1
N
1
N
Tr(φ′′′
B(h)) ≥ tr(φ′′′
B(h)) = tr(φ′′′(Γγ(h))) > ∆(h),
Tr(ψ′′′
B (h)) ≥ tr(ψ′′′
B (h)) = tr(ψ′′′(Γγ(h))) > ∆(h),
h ∈ HB,1,
h ∈ HB,1.
Then, it follows from the inductive hypothesis that there are homomorphisms φB, ψB : B →
MN (C) such that
(4.33)
(cid:13)(cid:13)(cid:13) φB(fB) − φ′′′
and there are decompositions
(cid:13)(cid:13)(cid:13) ψB(fB) − ψ′′′
B(fB)(cid:13)(cid:13)(cid:13) < ε/2 and
}
{z
K
B (fB)(cid:13)(cid:13)(cid:13) < ε/2,
{z
K
f ∈ F ,
}
φB = φB,0 ⊕ φB,1 ⊕ · · · ⊕ φB,1
and
ψB = ψB,0 ⊕ ψB,1 ⊕ · · · ⊕ ψB,1
with φB,1 and ψB,1 unitarily equivalent and
(4.34)
(4.35)
1
N
1
N
rank( φB,0(a)) ≤ η ·
rank( ψB,0(a)) ≤ η ·
1
N
1
N
rank( φB(a)),
a ∈ FB,
rank( ψB(a)),
a ∈ FB.
By (4.31) and (4.34), one has that for any a ∈ FB, one has
(4.36) tr( φB,0(a)) =
1
m
rank( φB,0(a)) =
N
m
1
N
rank( φB,0(a)) ≤ η ·
1
m
rank( φB(a)) = η · tr( φB(a)),
and for the same reason,
(4.37)
tr( ψB,0(a)) ≤ η · tr( ψB(a)),
a ∈ FB.
Consider the map Lkxik<1−γ πxi, there is a decomposition
Mkxik<1−γ
πxi = Mkxik<1−γ
πx′
i
πx′′
i
) ⊕ · · · ⊕ ( Mkxik<1−γ
⊕ ( Mkxik<1−γ
{z
K
πx′′
i
)
,
}
THE CLASSIFICATION OF SIMPLE SEPARABLE UNITAL LOCALLY ASH ALGEBRAS
28
where each x′
K. Since each xi has multiplicity at least M = ⌊2K/η⌋ + 1, one has that
i are same as xi but with different multiplies, and x′
i and x′′
i has multiplicity at most
(4.38)
Then set
and
i
πx′
tr( Mkxik<1−γ
φ0 = φB,0 ⊕ Mkxik<1−γ
φ1 = φB,1 ⊕ ( Mkxik<1−γ
(a)) ≤ η · tr( Mkxik<1−γ
πxi(a)),
a ∈ A.
πx′
i
and
πx′′
i
) and
πx′
i
,
ψ0 = ψB,0 ⊕ Mkxik<1−γ
ψ1 = ψB,1 ⊕ ( Mkxik<1−γ
πx′′
i
),
where φB,0, φB,1, ψB,0, ψB,1 are regarded as maps on A.
The homomorphisms
and these decompositions have the properties required in the statement of the theorem.
Indeed, by (4.12), (4.13), (4.22), (4.28), and (4.33), one has
φ = φ0 ⊕ φB,1 ⊕ · · · ⊕ φB,1
and
ψ = ψ0 ⊕ ψB,1 ⊕ · · · ⊕ ψB,1
K
}
{z
{z
}
(cid:13)(cid:13)(cid:13)φ(f ) − φ(f )(cid:13)(cid:13)(cid:13) < kφ(f ) − φ′′′(f )k +(cid:13)(cid:13)(cid:13)φ′′′(f ) − φ(f )(cid:13)(cid:13)(cid:13) < ε/2 + ε/2 = ε,
(cid:13)(cid:13)(cid:13)ψ(f ) − ψ(f )(cid:13)(cid:13)(cid:13) < kψ(f ) − ψ′′′(f )k +(cid:13)(cid:13)(cid:13)ψ′′′(f ) − ψ(f )(cid:13)(cid:13)(cid:13) < ε/2 + ε/2 = ε,
and
f ∈ F ,
f ∈ F .
It is also clear that φ1 and ψ1 are unitarily equivalent, and it follows from (4.36), (4.38) and
(4.38) that the maps φ0 and ψ0 satisfy the desired trace condition.
(cid:3)
Recall (from [9])
Lemma 4.6 (Lemma 4.13 of [9]). Let A be a unital separable nuclear residually finite dimensional
C*-algebra satisfying the UCT, and let ∆ : A+
1,q \ {0} → (0, 1) be an order preserving map. For
any finite set F ⊆ A and any ε > 0, there exist δ > 0, a finite set G ⊆ A, a finite set P ⊆ K(A),
a finite set H ⊆ A1
+ \ {0}, and an integer K ≥ 1 satisfying the following condition: For any two
unital G-δ-multiplicative linear maps φ1, φ2 : A → Mn(C) (for some integer n) and any unital
homomorphism ψ : A → Mm(C) with m ≥ n such that
(1) τ ◦ ψ(g) ≥ ∆(g), g ∈ H,
(2) [φ1]P = [φ2]P,
there exists a unitary U ∈ MKm+n(C) such that
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
φ1(f ) ⊕ ψ(f ) ⊕ · · · ⊕ ψ(f )
−u∗(φ2(f ) ⊕ ψ(f ) ⊕ · · · ⊕ ψ(f )
< ε,
f ∈ F .
{z
K
}
{z
K
)u(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
}
THE CLASSIFICATION OF SIMPLE SEPARABLE UNITAL LOCALLY ASH ALGEBRAS
29
Theorem 4.7. Let A be an NCCC. Let ∆ : A+
1,q \ {0} → (0, 1) be an order preserving map. Let
(F , ε) be given. Then there are finite sets G ⊆ A, H1 ⊆ A+, H2 ⊆ A+ and P ⊆ K(A), and
positive numbers δ, σ > 0 satisfying the following condition:
If φ, ψ : A → Mm(C) are unital G-δ-multiplicative maps such that
(1) [φ]P = [ψ]P,
(2) tr(φ(h)) > ∆(h) and tr(φ(h)) > ∆(h), h ∈ H1, and
(3) tr(φ)(h) − tr(ψ)(h) < σ, h ∈ H2,
then there is a unitary u ∈ Mm(C) such that
kφ(f ) − u∗ψ(f )uk < ε,
f ∈ F .
Proof. Applying Lemma 4.6 to the data A, ∆/4, and (F , ε/3), one obtains (G′, δ′, P, H, K)
satisfying the conditions of Lemma 4.6.
Applying Theorem 4.5 to the data A, ∆/2, F ∪H, min{ε/6, ∆(H)/4} (in place of ε), η = 1/2K,
and K, one obtains H1, H2, and 2σ (in place of δ) satisfying the conclusion of Theorem 4.5.
Without loss of generality, one may assume that H1 and H2 are in the unital ball of A.
Applying Corollary 5.5 of [9] to
F ∪ H1 ∪ H2, min{ε/6, ∆(H1)/2, ∆(H2)/2, σ/4}, min{1/2K, ∆(H1)/2, ∆(H2)/2, σ/4(1 + σ)},
one obtains (G, δ).
Then G, δ, σ, H1, H2, and P satisfy the conclusion of the theorem. Indeed, let φ, ψ : A →
Mm(C) be G-δ-multiplicative maps such that
(1) [φ]P = [ψ]P,
(2) tr(φ(h)) > ∆(h) and tr(φ(h)) > ∆(h), h ∈ H1, and
(3) tr(φ)(h) − tr(ψ)(h) < σ, h ∈ H2.
By Theorem 5.5 of [9], there are φ0, φ1, ψ0, ψ1 : A → Mm(C) such that φ0, ψ0 are G′-δ′-
multiplicative, φ1, ψ1 are homomorphisms, such that
(4.39)
kφ(a) − φ0(a) ⊕ φ1(a)k < min{ε/6, ∆(H1)/2, ∆(H2)/2, σ/4},
a ∈ F ∪ H1 ∪ H2,
(4.40)
kψ(a) − ψ0(a) ⊕ ψ1(a)k < min{ε/6, ∆(H1)/2, ∆(H2)/2, σ/4},
a ∈ F ∪ H1 ∪ H2,
and
tr(φ0(1)) = tr(ψ0(1)) < min{1/2K, ∆(H1)/2, ∆(H2)/2, σ/4(1 + σ)}.
Consider the unital homomorphisms φ1, ψ1 : A → pMm(C)p, where p = φ1(1) = ψ1(1). One
has that
(1) τ (φ1(h)), τ (φ1(h)) > ∆(h)/2,
(2) τ (φ1(h)) − τ (ψ1(h)) < 2σ,
h ∈ H1, and
h ∈ H2,
where τ ∈ T(pMm(C)p).
THE CLASSIFICATION OF SIMPLE SEPARABLE UNITAL LOCALLY ASH ALGEBRAS
30
By Theorem 4.5, up to unitary equivalence, there are homomorphisms φ′
1, ψ′
1, µ : A →
pMm(C)p such that
< min{ε/6, ∆(H)/4},
a ∈ F ∪ H,
< min{ε/6, ∆(H)/4},
a ∈ F ∪ H,
(4.41)
(4.42)
and
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
φ1(a) − φ′
1(a) ⊕ µ(a) ⊕ · · · ⊕ µ(a)
ψ1(a) − ψ′
1(a) ⊕ µ(a) ⊕ · · · ⊕ µ(a)
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
}
}
{z
K
{z
K
τ (φ′
1(a)) ≤
1
2K
· τ (φ1(a))
and τ (ψ′
1(a)) ≤
1
2K
Therefore, one has that
· τ (ψ1(a)),
a ∈ F ∪ H, τ ∈ T(pMm(C)p).
τ (µ(h)) >
∆(h),
1
4
h ∈ H, τ ∈ T(qMm(C)q),
where q = µ(1). Consider the map
and note that
(φ0 ⊕ φ′
1) ⊕ (µ ⊕ · · · ⊕ µ
) and (ψ0 ⊕ ψ′
1) ⊕ (µ ⊕ · · · ⊕ µ
),
{z
K
}
{z
K
}
tr(φ0(1) ⊕ φ′
1(1)) = tr(ψ0(1) ⊕ ψ′
1(1)) < tr(µ(1)).
It then follows from Lemma 4.6 that there is a unitary u ∈ Mm(C) such that for any a ∈ F ,
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
(φ0(a) ⊕ φ′
1(a)) ⊕ (µ(a) ⊕ · · · ⊕ µ(a)
) − u∗((ψ0(a) ⊕ ψ′
1(a)) ⊕ (µ(a) ⊕ · · · ⊕ µ(a)
{z
K
}
{z
K
It then follows from (4.39), (4.40), (4.41), and (4.42) that
as desired.
kφ(a) − u∗ψ(a)uk < ε,
a ∈ F ,
))u(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
}
<
ε
3
.
(cid:3)
Note that KL(A, Q) ∼= Hom(K0(A), K0(Q)), a straightforward consequence is
Corollary 4.8. Let A be an NCCC. Let ∆ : A+
1,q \ {0} → (0, 1) be an order preserving map. Let
(F , ε) be given. Then there are finite sets H1 ⊆ A+ and H2 ⊆ A+ and a positive number δ > 0
satisfying the following condition:
If φ, ψ : A → Q are unital homomorphisms such that
(1) [φ]0 = [ψ]0,
(2) tr(φ(h)) > ∆(h) and tr(φ(h)) > ∆(h), h ∈ H1, and
(3) tr(φ)(h) − tr(ψ)(h) < δ, h ∈ H2,
then there is a unitary u ∈ Q such that
kφ(f ) − u∗ψ(f )uk < ε,
f ∈ F .
THE CLASSIFICATION OF SIMPLE SEPARABLE UNITAL LOCALLY ASH ALGEBRAS
31
It is also worth pointing out the following corollary:
Corollary 4.9. Let A be a subhomogeneous C*-algebra. Let ∆ : A+
1,q \ {0} → (0, 1) be an order
preserving map. Let (F , ε) be given. Then there are finite sets G ⊆ A, H1 ⊆ A+, H2 ⊆ A+ and
P ⊆ K(A), and positive numbers δ, σ > 0 satisfying the following condition:
If φ, ψ : A → Mm(C) are unital G-δ-multiplicative maps such that
(1) [φ]P = [ψ]P,
(2) tr(φ(h)) > ∆(h) and tr(φ(h)) > ∆(h), h ∈ H1, and
(3) tr(φ)(h) − tr(ψ)(h) < σ, h ∈ H2,
then there is a unitary u ∈ Mm(C) such that
kφ(f ) − u∗ψ(f )uk < ε,
f ∈ F .
Proof. This follows the fact that any subhomogeneous C*-algebra can be locally approximated
by NCCCs (Theorem 2.15 of [5]).
(cid:3)
5. Tracial factorization and tracial approximation
Recall that
Definition 5.1 ([13], [2]). Let S be a class of unital C*-algebras. A C*-algebra A is said to
be tracially approximated by the C*-algebras in S, and one writes A ∈ TAS, if the following
condition holds: For any finite set F ⊆ A, any ε > 0, and any nonzero a ∈ A+, there is a nonzero
sub-C*-algebra S ⊆ A such that S ∈ S, and if p = 1S, then
(1) kpf − f pk < ε, f ∈ F ,
(2) pf p ∈ε S, f ∈ F , and
(3) 1 − p is Murray-von Neumann equivalent to a subprojection of aAa.
One particularly important class S of building blocks is the class of Elliott-Thomsen algebras.
Definition 5.2. ([6], [1]) A C*-algebra C is said to be an Elliott-Thomsen algebra if
C = {(a, f ) ∈ E ⊕ (F ⊗ C([0, 1])) : f (0) = 0(a), f (1) = 1(a)}
for some finite dimensional C*-algebras E, F , where 0, 1 : E → F are unital homomorphisms.
Denote by π∞ the standard quotient map
π∞ : A ∋ (a, f ) 7→ a ∈ E.
Let us denote the class of unital Elliott-Thomsen algebras by C, and denote the class of unital
Elliott-Thomsen algebras with trivial K1-group by C0.
Remark 5.3. In fact, by Corollary 29.3 of [9], one has TAC = TAC0.
Remark 5.4. In fact, the class of unital Elliott-Thomsen algebras is exactly the class of NCCCs
with dimensions of cells at most one; see [5].
For TAC0 algebras, one has the following classification theorem.
THE CLASSIFICATION OF SIMPLE SEPARABLE UNITAL LOCALLY ASH ALGEBRAS
32
Theorem 5.5 (Corollary 28.7 of [9]). Let A, B be unital separable amenable simple C*-algebras
satisfying the UCT. Assume that A, B are Jiang-Su stable, and assume that A ⊗ Q ∈ TAC0 and
B ⊗ Q ∈ TAC0. Then A ∼= B if and only if Ell(A) ∼= Ell(B).
In this section, let us show that for any separable simple unital locally ASH algebra A, one
has that A ⊗ Q ∈ TAC0.
Theorem 5.6. Let A be a unital simple separable locally approximately subhomogeneous (ASH)
C*-algebra satisfying A ∼= A ⊗ Q. Then, for any finite set F ⊆ A and any ε > 0, there exist an
Elliott-Thomsen algebra C with K1(C) = {0}, a unital completely positive linear map Φ : A → C,
and a unital embedding Ψ : C → A such that
(1) Φ is F -ε-multiplicative, and
(2) τ (f ) − τ (Ψ(Φ(f ))) < ε, f ∈ F , τ ∈ T(A).
Proof. Without loss of generality, one may assume that 1 ∈ F and each element of F is self-
adjoint and has norm at most 1.
Let A be a unital separable simple locally ASH algebra satisfying A ∼= A ⊗ Q. By Theorem
2.15 of [5], the C*-algebra A can be locally approximated by unital NCCCs. Therefore, without
loss of generality, one may assume that there is a sub-C*-algebra A1 ⊆ A such that A1 is a
NCCC and F ⊆ A1.
Put
GA1 = ρA1(K0(A1)), G+
A1 = ρA1(K0(A1)) ∩ Aff +(T(A1)), uA1 = ρA1([1]),
and fix a set {p1, ..., pn} which generates the group GA1. Note that by Lemma 2.9, the positive
cone G+
A1 is finitely generated.
For each h ∈ A+, define
∆(h) = inf{τ (ι(h)) : τ ∈ T(A)},
where ι : A1 → A is the embedding map. Since A is simple, the map ∆ induces a order preserving
map from A1,q
+ to (0, 1). Let us still denote this by ∆.
Let H1 ⊆ (A1)+, H2 ⊆ (A1)+ and δ0 > 0 be the finite sets and constant of Corollary 4.8 with
respect to F · F , ε/4, and ∆/4.
Let H be the subset of Theorem 3.7 with respect to A1, F ∪ H1 ∪ H2 (in the place of F ),
min{ε/4, δ0/4, ∆(h)/4 : h ∈ H1} (in the place of ε). Put
σ =
1
2
min{∆(h) : h ∈ H},
and denote by δ1 the constant of Theorem 3.7 with respect to σ.
Put
G = ρA(K0(A)), G+ = ρA(K0(A)) ∩ Aff +(T(A)), u = ρA([1]).
Then (G, G+, u) is a unperforated order-unit group, and there is a natural pairing between
(G, G+, u) and T(A) induced by ρA. Still denote this pairing map by ρA. Note that one has the
THE CLASSIFICATION OF SIMPLE SEPARABLE UNITAL LOCALLY ASH ALGEBRAS
33
following commutative diagram:
[ι]0
K0(A1)
K0(A)
ρA1
ρA
GA1
(ι)∗
/ G.
lim
−→
Consider ((G, G+, u), T(A), ρA). By Theorem 5.2.2 of [1], there is an inductive system C ′ =
(C ′
i, η′
i) with C ′
i ∈ C0 and ηi injective such that there is an isomorphism
Ξ : ((G, G+, u), T(A), ρA) → ((K0(C ′), K0(C ′), [1]), T(C ′), ρ′
C).
Consider C = C ′ ⊗ Q = lim
−→
(Ci, ηi), where Ci = C ′
i ⊗ Q and ηi = η′
i ⊗ idQ. One has
((G, G+, u), T(A), ρA) ∼= ((K0(C), K0(C), [1]), T(C), ρC),
and let us still denote by Ξ the isomorphism. In the remaining part of the paper, let us also use
Ξ to denote its restriction to G or to Aff(T(A)), depending on the context.
Consider the following diagram:
GA1
[ι]0
/ G
Ξ
K0(C1)
/ K0(C2)
/ · · ·
[η2]0
[η1]0
/ K0(C).
Since the positive cone of GA1 is finitely generated (Lemma 2.9), the positive map [ι]0 can be lifted
to a positive homomorphism GA1 → K0(Cn) for sufficiently large n. Without loss of generality,
one may assume that [ι]0 has a lifting κ : GA1 → K0(C1), making the diagram commutative:
[ι]0
GA1
κ
G
Ξ
K0(C1)
/ K0(C2)
/ · · ·
[η2]0
[η1]0
/ K0(C).
By Lemma 2.10, after a telescoping of the inductive system (Ci, ηi), there is also an approximate
lifting, making the diagram of affine function spaces,
Aff(T(A1))
γ
(ι)∗
Aff(T(A)))
Ξ
Aff(T(C1)))
/ Aff(T(C2)))
(η1)∗
(η2)∗
/ · · ·
/ Aff(T(C))),
approximately commutative, and such that
(5.1)
(5.2)
τ (κ([pi])) − γ([pi])(τ ) < δ1,
τ ∈ T(C1), 1 ≤ i ≤ n,
γ(h)(τ ) >
1
2
∆(h) > σ,
h ∈ H ∪ H1, τ ∈ T(C1),
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
and
(5.3)
Write
where
THE CLASSIFICATION OF SIMPLE SEPARABLE UNITAL LOCALLY ASH ALGEBRAS
34
(cid:12)(cid:12)(cid:12)(Ξ−1 ◦ (η1,∞)∗ ◦ γ)( f )(τ ) − (ι)∗( f )(τ )(cid:12)(cid:12)(cid:12) < ε/8,
τ ∈ T(A), f ∈ F .
C1 = C ′
1 ⊗ Q = {(a, f ) ∈ E ⊕ (F ⊗ C([0, 1])) : f (0) = 0(a), f (1) = 1(a)},
E =
pMi=1
Q, F =
lMj=1
Q,
for natural numbers p, l, and 0, 1 : E → F are unital homomorphisms.
On each interval [0, 1]j, choose a partition
0 = t0 < t1 < t2 < · · · < tk−1 < tk = 1
such that
(5.4)
(cid:12)(cid:12)(cid:12)γ( f )(τs) − γ( f )(τti)(cid:12)(cid:12)(cid:12) < min{ε/4, δ0/2},
where τs = tr ◦ πs and tr is the canonical trace of Q. One may assume that k is sufficiently large
that
s ∈ [ti, ti+1], f ∈ F ∪ H2,
For each 0 < i < k, define
2π/(k − 1) < ε/8.
and
κi = [πti]0 ◦ κ : K0(A1) → K0(Q) ∼= Q,
γi = (πti)∗ ◦ γ : Aff(T(A1)) → Aff(T(Q)) ∼= R.
By (5.1), each pair (κi, γi) is δ1-compatible on [pi], 1 ≤ i ≤ n. By (5.2), one has that
γi(h)(tr) > σ,
h ∈ H.
Therefore, by Theorem 3.7, there is a homomorphism φi : A1 → Q such that
[φi]0 = κi
and
(5.5)
(cid:12)(cid:12)(cid:12)γi(h)(tr) − tr(φi(h))(cid:12)(cid:12)(cid:12) < min{ε/4, δ0/4, ∆(h)/4 : h ∈ H1},
Together with (5.2), it then follows that
h ∈ F ∪ H1 ∪ H2.
(5.6)
tr(φi(h)) > γi(h)(tr) −
∆(h) >
1
4
1
4
∆(h),
h ∈ H1.
(5.7)
It also follows from (5.5) and (5.4) that for any h ∈ H2 and any 1 ≤ i ≤ k − 2,
tr(φi(h)) − tr(φi+1(h)) ≤ (cid:12)(cid:12)(cid:12)tr(φi(h)) − γi(h)(tr)(cid:12)(cid:12)(cid:12) +(cid:12)(cid:12)(cid:12)γi(h)(tr) − γi+1(h)(tr)(cid:12)(cid:12)(cid:12)
+(cid:12)(cid:12)(cid:12)tr(φi+1(h)) − γi+1(h)(tr)(cid:12)(cid:12)(cid:12)
< δ0/4 + δ0/2 + δ0/4 = δ0.
THE CLASSIFICATION OF SIMPLE SEPARABLE UNITAL LOCALLY ASH ALGEBRAS
35
Since πi is homotopic to πi+1, it is clear that [κi] = [κi+1], 1 ≤ i ≤ k − 2, and therefore
(5.8)
[φ]0 = [φi+1]0,
1 ≤ i ≤ k − 2.
Denote by π∞ : C1 → E the standard quotient map, and consider
κ∞ = [π∞]0 ◦ κ : K0(A1) → K0(E) and γ∞ = [π∞] ◦ γ : Aff(T(A1)) → Aff(T(E)).
The same argument as above shows that there is a homomorphism φ∞ : A1 → E such that
[φ∞]0 = κ∞,
(5.9)
and
(5.10)
(cid:12)(cid:12)(cid:12)γ∞(h)(τ ) − τ (φ∞(h))(cid:12)(cid:12)(cid:12) < min{ε/4, δ0/4, ∆(h)/4 : h ∈ H1},
Define
h ∈ H1 ∪ H2, τ ∈ T(E).
φ0 = 0 ◦ φ∞ and φk = 1 ◦ φ∞
and consider the restrictions of these maps to the j-th direct summand; still denote them by φ0
and φk respectively. It then follows from (5.9) that
(5.11)
[φ0]0 = [φk] = [φi]0,
1 ≤ i ≤ k − 1,
and it follows from (5.10) and (5.2) that
(5.12)
tr(φ0(h)), tr(φk(h)) >
∆(h),
h ∈ H1.
1
4
Moreover, with (5.10) and (5.4), the same argument as that of (5.7) shows that
(5.13)
tr(φ0(h)) − tr(φ1(h)) < δ0
and
tr(φk−1(h)) − tr(φk(h)) < δ0,
h ∈ H2.
Thus, with (5.11), (5.6), (5.12), (5.7) and (5.13), Corollary 4.8 implies that there are unitaries
u1, u2, ..., uk−1 ∈ Q such that
(cid:13)(cid:13)φi(f ) − u∗
i+1φi+1(f )ui+1(cid:13)(cid:13) < ε/4,
Define v0 = 1, and
f ∈ F · F , 0 ≤ i ≤ k − 2.
vi = uiui−1 · · · u1,
i = 1, ..., k − 1.
Then, for any 0 ≤ i ≤ k − 2 and any f ∈ F · F , one has
kAd(vi) ◦ φi(f ) − Ad(vi+1) ◦ φi+1(f )k
= k(ui · · · u1)∗φi(f )(ui · · · u1) − (ui+1 · · · u1)∗φi+1(f )(ui+1 · · · u1)k
= (cid:13)(cid:13)φi(f ) − u∗
i+1φi+1(f )ui+1(cid:13)(cid:13) < ε/4.
Replacing each homomorphism φi by Ad(vi) ◦ φi for i = 1, ..., k − 1, and still denoting it by φi,
one has
kφi(f ) − φi+1(f )k < ε/4,
f ∈ F · F , 0 ≤ i ≤ k − 2.
Note that the replacement of φi does not change the induced map on the invariant, and hence
one still has
[φk−1]0 = [φk]0,
THE CLASSIFICATION OF SIMPLE SEPARABLE UNITAL LOCALLY ASH ALGEBRAS
36
and
tr(φ0(h)), tr(φk(h)) >
∆(h),
h ∈ H1,
1
4
tr(φk−1(h)) − tr(ψk(h)) < δ0,
h ∈ H2.
Applying Corollary 4.8 again, we obtain a unitary w ∈ Q such that
kw∗φk−1(f )w − φk(f )k < ε/4,
f ∈ F · F .
Since Q is AF, and any unitary in a finite dimensional C*-algebra can be connected to the
identity along a path with length at most π (i.e., it has exponential length at most π), there are
unitaries
1 = w0, w1, ..., wk−2, wk−1 = w ∈ Q
kwi − wi−1k < 2π/(k − 1) < ε/8.
(cid:13)(cid:13)w∗
i φi(f )wi − w∗
i+1φi+1(f )wi+1(cid:13)(cid:13) < 3ε/8,
f ∈ F · F , 0 ≤ i ≤ k − 2.
Replace each homomorphism φi again by Ad(wi) ◦ φi, 1 ≤ i ≤ k − 1, and still denote it by φi.
One then has
(5.14)
kφi(f ) − φi+1(f )k < 3ε/8,
f ∈ F · F , 0 ≤ i ≤ k − 1.
Define the positive linear map φ : A1 → C([0, 1]j, Q) by
Φj(f )(t) =
ti+1 − t
ti+1 − ti
φi(f ) +
t − ti
ti+1 − ti
φi+1(f ),
if t ∈ [ti, ti+1].
By (5.14), the map Φj is F -ε-multiplicative. It also follows from (5.4) that if t ∈ [ti, ti+1], then,
for any f ∈ F ,
such that
Hence,
(5.15)
(cid:12)(cid:12)(cid:12)γ( f )(τt) − τt(Φj(f ))(cid:12)(cid:12)(cid:12)
≤ (cid:12)(cid:12)(cid:12)γ( f )(τti) − τt(Φj(f ))(cid:12)(cid:12)(cid:12) + ε/4 (by (5.4))
= (cid:12)(cid:12)(cid:12)(cid:12)γi( f )(tr) − (
≤ (cid:12)(cid:12)(cid:12)γi( f )(tr) − tr(φi(f ))(cid:12)(cid:12)(cid:12) + 5ε/8 (by (5.14))
< ε/4 + 5ε/8 = 7ε/8 (by (5.5)).
ti+1 − t
ti+1 − ti
tr(φi(f )) +
t − ti
ti+1 − ti
tr(φi+1(f )))(cid:12)(cid:12)(cid:12)(cid:12) + ε/4
Repeat this construction for all j = 1, ..., l, and note that the maps Φ1, Φ2, ..., Φl induce a map
Φ : A1 → C1. Since each Φj is F -ε multiplicative, the map Φ is F -ε-multiplicative. By (5.15),
one has
(5.16)
f ∈ F , τ ∈ T(C1).
(cid:12)(cid:12)(cid:12)γ( f )(τ ) − τ (Φ(f ))(cid:12)(cid:12)(cid:12) < 7ε/8,
Now, let us construct an embedding Ψ : C1 → A such that
τ (f ) − τ (Ψ(Φ(f ))) < ε,
f ∈ H, τ ∈ T(A).
THE CLASSIFICATION OF SIMPLE SEPARABLE UNITAL LOCALLY ASH ALGEBRAS
37
Since A ∼= A ⊗ Q, the subgroup H := ker ρA ⊆ K0(A) is divisible, and therefore the exact
sequence
splits. Pick a decomposition
0
/ H
/ K0(A)
ρA
/ G
/ 0
K0(A) = G ⊕ H.
Since K0(A) is weakly unperforated, the order on (G ⊕ H) is completely determined by that on
G.
Define
Then κ′ is a positive homomorphism, and the pair
κ′ : G ∋ g 7→ (g, 0) ∈ G ⊕ H = K0(A).
(κ′ ◦ (Ξ−1K0(C)), ΞAff(T(C)))
is compatible. It induces a positive homomorphism
θ : Cu∼(C) → Cu∼(A).
By Theorem 1 of [26], there is a homomorphism ψ : C → A such that the Cuntz map induced
by ψ is θ. In particular, one has that
(5.17)
Then the map
(ψ)∗ = Ξ−1Aff(T(C)).
Ψ = ψ ◦ η1,∞
satisfies the conclusion of the theorem (together with Φ). Indeed, since C is simple, the map ψ
is an embedding, and therefore Ψ is an embedding. Moreover, for any f ∈ F , one has
τ (ι(f )) − τ (Ψ ◦ Φ(f ))
= (cid:12)(cid:12)(cid:12)(ι)∗( f )(τ ) − (Ψ)∗((Φ)∗( f ))(τ )(cid:12)(cid:12)(cid:12)
= (cid:12)(cid:12)(cid:12)(ι)∗( f )(τ ) − (ψ)∗((ηi,∞ ◦ Φ)∗( f ))(τ )(cid:12)(cid:12)(cid:12)
= (cid:12)(cid:12)(cid:12)(ι)∗( f )(τ ) − (Ξ−1 ◦ (ηi,∞)∗ ◦ (Φ)∗)( f )(τ )(cid:12)(cid:12)(cid:12)
< (cid:12)(cid:12)(cid:12)(ι)∗( f )(τ ) − (Ξ−1 ◦ (ηi,∞)∗ ◦ γ)( f )(τ )(cid:12)(cid:12)(cid:12) + 7ε/8 (by (5.16))
≤ ε/8 + 7ε/8 = ε
(by (5.17))
(by (5.3)),
as desired.
(cid:3)
Remark 5.7. With a slight modification (a perturbation of the linear map γ), the same argu-
ment as Theorem 5.6 shows that the same statement holds for C*-algebras which are tracially
approximated by subhomogeneous C*-algebras.
The passage from Theorem 5.6 to the actual tracial approximation is an application of the
following very important theorem due to Winter:
/
/
/
/
THE CLASSIFICATION OF SIMPLE SEPARABLE UNITAL LOCALLY ASH ALGEBRAS
38
Theorem 5.8 (Theorem 2.2 of [32]). Let S be a class of separable unital C*-algebras which have
a finite presentation with weakly stable relations. Suppose further that S is closed under taking
direct sums and under taking tensor products with finite dimensional C*-algebras, and that S
contains all finite dimensional C*-algebras.
Let A be a separable, simple, unital C*-algebra with dimnucA < ∞ and T(A) 6= Ø, and let
be a system of maps with the following properties:
( A
σi
/ Bi
i
/ A )i∈N
(1) Bi ∈ S, i ∈ N,
(2) i is an embedding for each i ∈ N,
(3) σi is a completely positive contraction for each i ∈ N,
(4) ¯σ : A →Qi∈N Bi/Li∈N Bi induced by σi is a unital homomorphism
(5) sup{τ (iσi(a) − a) : τ ∈ T(A)} → 0, as i → ∞ for each a ∈ A.
Then A ⊗ Q ∈ TAS.
With this and Theorem 5.6, one has
Theorem 5.9. Let A be a unital simple separable locally ASH C*-algebra. Then A ⊗ Q ∈ TAC0,
where C0 is the class of unital Elliott-Thomsen algebras with trivial K1-group. In particular, if
A ∼= A ⊗ Z, where Z is the Jiang-Su algebra, then A is classifiable (by means of the naive Elliott
invariant). (The converse is also true.)
Proof. By Theorem 3.1 of [5], one has dr(A⊗Q) ≤ 2, and in particular, dimnuc(A⊗Q) ≤ 2 < +∞.
It then follows from Theorems 5.6 and 5.8 that A⊗Q ∈ TAC0. By the classification theorem of [9]
(based in particular on the deformation technique of [33] and [17] -- see also [19]), the C*-algebra
A is classifiable.
(cid:3)
Corollary 5.10. Let A be a simple separable unital locally ASH (respectively, locally AH) algebra.
Then A ⊗ Z is an ASH (respectively, AH) algebra.
Proof. By Theorem 5.9, the C*-algebra A ⊗ Z is classifiable by means of the Elliott invariant.
By [1] and [8], the Elliott invariant for separable, Jiang-Su stable, simple, unital, finite C*-
algebras (in particular, locally ASH algebras) is exhausted by ASH algebras (by Theorem 3 of
[8], finiteness implies stable finiteness in this setting). Furthermore, by [30], the Elliott invariant
for separable, Jiang-Su stable, simple, unital, locally AH algebras is exhausted by AH algebras.
(In both settings, the models have no dimension growth.)
(cid:3)
The classification of locally ASH algebras (Theorem 5.9) in fact allow us to recover the recent
classification result for the C*-algebra of a minimal homeomorphism -- assumed to have mean
dimension zero but not to be uniquely ergodic ([27], [18] -- the uniquely ergodic case was dealt
with in [3], or in [29] on the ease the space is finite dimensional):
Corollary 5.11 (Corollary 5.3 of [18]). Let X be a compact metrizable space, and let σ : X → X
be a minimal homeomorphism. Then the C*-algebra (C(X)⋊σ Z)⊗Z is classifiable. In particular,
if (X, σ) has mean dimension zero, the C*-algebra C(X) ⋊σ Z is classifiable.
/
/
THE CLASSIFICATION OF SIMPLE SEPARABLE UNITAL LOCALLY ASH ALGEBRAS
39
Proof. By Theorem 4.1 of [5], the C*-algebra (C(X) ⋊σ Z) ⊗ Q is locally ASH. By Theorem
5.9, the C*-algebra (C(X) ⋊σ Z) ⊗ Q belongs to the class TAC0, and hence the C*-algebra
(C(X) ⋊σ Z) ⊗ Z is classifiable.
If (X, σ) has mean dimension zero, then it follows from [3] that
C(X) ⋊σ Z ∼= (C(X) ⋊σ Z) ⊗ Z,
and so the C*-algebra C(X) ⋊σ Z is classifiable.
(cid:3)
References
[1] G. A. Elliott. An invariant for simple C*-algebras. Canadian Mathematical Society. 1945 -- 1995, 3:61 -- 90,
1996.
[2] G. A. Elliott and Z. Niu. On tracial approximation. J. Funct. Anal., 254(2):396 -- 440, 2008. URL:
http://dx.doi.org/10.1016/j.jfa.2007.08.005, doi:10.1016/j.jfa.2007.08.005.
[3] G. A. Elliott and Z. Niu. The C*-algebra of a minimal homeomorphism of zero mean dimension. 06 2014.
URL: http://arxiv.org/abs/1406.2382, arXiv:1406.2382.
[4] G. A. Elliott and Z. Niu. On the classification of simple unital C*-algebras with finite decomposition rank.
Preprint, 2015.
[5] G. A. Elliott, Z. Niu, L. Santiago, and A. Tikuisis. Decomposition rank of approximately subhomogeneous
C*-algebras. 05 2015. URL: http://arxiv.org/abs/1505.06100, arXiv:1505.06100.
[6] G. A. Elliott and K. Thomsen. The state space of the K0-group of a simple separable C*-algebra. Geom. Funct.
Anal., 4(5):522 -- 538, 1994. URL: http://dx.doi.org/10.1007/BF01896406, doi:10.1007/BF01896406.
[7] G. A. Elliott and A. S. Toms. Regularity properties
in the
classification program for
sep-
2008. URL:
amenable C*-algebras. Bull. Amer. Math.
arable
http://dx.doi.org/10.1090/S0273-0979-08-01199-3, doi:10.1090/S0273-0979-08-01199-3.
45(2):229 -- 245,
(N.S.),
Soc.
[8] G. Gong, X. Jiang,
Canad. Math. Bull.,
doi:10.4153/CMB-2000-050-1.
and H. Su. Obstructions
43(4):418 -- 426,
2000. URL:
to Z-stability for unital
simple C*-algebras.
http://dx.doi.org/10.4153/CMB-2000-050-1,
[9] G. Gong, H. Lin, and Z. Niu. Classification of finite simple amenable Z-stable C*-algebras. 01 2015. URL:
http://arxiv.org/abs/1501.00135, arXiv:1501.00135.
[10] P. R. Halmos and H. E. Vaughan. The marriage problem. Amer. J. Math., 72:214 -- 215, 1950.
[11] E. Kirchberg. The classification of purely infinite C*-algebras using Kasparov's theory. Preprint, 1994.
[12] E. Kirchberg
algebra
O2. J. Reine Angew. Math., 525:17 -- 53, 2000. URL: http://dx.doi.org/10.1515/crll.2000.065,
doi:10.1515/crll.2000.065.
and N. C. Phillips. Embedding
exact C*-algebras
in the Cuntz
of
[13] H. Lin. Tracially AF C*-algebras. Trans. Amer. Math. Soc., 353(2):693 -- 722, 2001.
[14] H. Lin. Simple nuclear C*-algebras of tracial topological rank one. J. Funct. Anal., 251(2):601 -- 679, 2007.
URL: http://dx.doi.org/10.1016/j.jfa.2007.06.016, doi:10.1016/j.jfa.2007.06.016.
[15] H. Lin. Asymptotically unitary equivalence and asymptotically inner automorphisms. Amer. J. Math.,
131(6):1589 -- 1677, 2009. URL: http://dx.doi.org/10.1353/ajm.0.0086, doi:10.1353/ajm.0.0086.
simple amenable C*-algebras.
[16] H. Lin. Asymptotic unitary equivalence and classification of
In-
http://dx.doi.org/10.1007/s00222-010-0280-9,
vent. Math.,
doi:10.1007/s00222-010-0280-9.
183(2):385 -- 450,
2011. URL:
[17] H. Lin. Localizing the Elliott conjecture at strongly self-absorbing C*-algebras, II. J. Reine Angew. Math.,
692:233 -- 243, 2014.
[18] H.
Lin.
Crossed
products
and
minimal
dynamical
systems.
02
2015.
URL:
http://arxiv.org/abs/1502.06658, arXiv:1502.06658.
THE CLASSIFICATION OF SIMPLE SEPARABLE UNITAL LOCALLY ASH ALGEBRAS
40
[19] H. Lin and Z. Niu. Lifting KK-elements, asymptotic unitary equivalence and classification of simple C*-
algebras. Adv. Math., 219(5):1729 -- 1769, 2008. URL: http://dx.doi.org/10.1016/j.aim.2008.07.011,
doi:10.1016/j.aim.2008.07.011.
[20] H. Matui and Y. Sato. Strict comparison and Z-absorption of nuclear C*-algebras. Acta Math., 209(1):179 --
196, 2012. URL: http://dx.doi.org/10.1007/s11511-012-0084-4, doi:10.1007/s11511-012-0084-4.
[21] H. Matui and Y. Sato. Decomposition rank of UHF-absorbing C*-algebras. Duke Math. J., 163(14):2687 --
2708, 2014. URL: http://dx.doi.org/10.1215/00127094-2826908, doi:10.1215/00127094-2826908.
[22] Z. Niu. A classification of tracially approximate splitting interval algebras. I. The building blocks and the
limit algebras. C. R. Math. Acad. Sci. Soc. R. Can., 35(1):1 -- 34, 2014.
[23] Z. Niu. A classification of tracially approximate splitting interval algebras. II. Existence theorem. C. R. Math.
Acad. Sci. Soc. R. Can., 37(1):1 -- 32, 2015.
[24] Z. Niu. A classification of tracially approximate splitting interval algebras. III. Uniqueness theorem and
isomorphism theorem. C. R. Math. Acad. Sci. Soc. R. Can., 37(2):41 -- 75, 2015.
[25] N. C. Phillips. A classification theorem for nuclear purely infinite simple C*-algebras. Doc. Math., 5:49 -- 114
(electronic), 2000.
[26] L. Robert. Classification of inductive limits of 1-dimensional NCCW complexes. Adv. Math., 231(5):2802 --
2836, 2012. URL: http://dx.doi.org/10.1016/j.aim.2012.07.010, doi:10.1016/j.aim.2012.07.010.
[27] K. R. Strung. C*-algebras of minimal dynamical systems of the product of a cantor set and an odd dimensional
sphere. 03 2014. URL: http://arxiv.org/abs/1403.3136, arXiv:1403.3136.
[28] A. S. Toms. K-theoretic rigidity and slow dimension growth. Invent. Math., 183(2):225 -- 244, 2011. URL:
http://dx.doi.org/10.1007/s00222-010-0273-8, doi:10.1007/s00222-010-0273-8.
[29] A. S. Toms and W. Winter. Minimal dynamics and the classification of C*-algebras. Proc. Natl.
Acad. Sci. USA, 106(40):16942 -- 16943, 2009. URL: http://dx.doi.org/10.1073/pnas.0903629106,
doi:10.1073/pnas.0903629106.
[30] J. Villadsen. The range of the Elliott invariant of the simple AH-algebras with slow dimension growth.
K-Theory, 15(1):1 -- 12, 1998.
[31] W. Winter. Nuclear dimension and Z-stability of pure C*-algebras. Invent. Math., 187(2):259 -- 342, 2012.
URL: http://dx.doi.org/10.1007/s00222-011-0334-7, doi:10.1007/s00222-011-0334-7.
[32] W. Winter. Classifying crossed product C*-algebras. 08 2013. URL: http://arxiv.org/abs/1308.5084,
arXiv:1308.5084.
[33] W. Winter. Localizing the Elliott conjecture at strongly self-absorbing C*-algebras. J. Reine Angew. Math.,
692:193 -- 231, 2014.
[34] W. Winter and J. Zacharias. The nuclear dimension of C*-algebras. Adv. Math., 224(2):461 -- 498, 2010. URL:
http://dx.doi.org/10.1016/j.aim.2009.12.005, doi:10.1016/j.aim.2009.12.005.
Department of Mathematics, University of Toronto, Toronto, Ontario, Canada M5S 2E4.
E-mail address: [email protected]
Department of Mathematics, University of Puerto Rico.
E-mail address: [email protected]
University of Oregon and East China Normal University.
E-mail address: [email protected]
Department of Mathematics, University of Wyoming, Laramie, WY, 82071, USA.
E-mail address: [email protected]
|
1106.2419 | 3 | 1106 | 2012-04-14T20:28:35 | On a quantitative operator K-theory | [
"math.OA",
"math.KT",
"math.MG"
] | In this paper, we develop a quantitative K-theory for filtered C*-algebras. Particularly interesting examples of filtered C*-algebras include group C*-algebras, crossed product C*-algebras and Roe algebras. We prove a quantitative version of the six term exact sequence and a quantitative Bott periodicity. We apply the quantitative K-theory to formulate a quantitative version of the Baum-Connes conjecture and prove that the quantitative Baum-Connes conjecture holds for a large class of groups. | math.OA | math | ON QUANTITATIVE OPERATOR K-THEORY
HERV´E OYONO-OYONO AND GUOLIANG YU
Abstract. In this paper, we develop a quantitative K-theory for filtered
C ∗-algebras. Particularly interesting examples of filtered C ∗-algebras include
group C ∗-algebras, crossed product C ∗-algebras and Roe algebras. We prove
a quantitative version of the six term exact sequence and a quantitative Bott
periodicity. We apply the quantitative K-theory to formulate a quantitative
version of the Baum-Connes conjecture and prove that the quantitative Baum-
Connes conjecture holds for a large class of groups.
2
1
0
2
r
p
A
4
1
]
.
A
O
h
t
a
m
[
3
v
9
1
4
2
.
6
0
1
1
:
v
i
X
r
a
Contents
Introduction
0.
1. Quantitative K-theory
1.1. Filtered C∗-algebras
1.2. Almost projections/unitaries
1.3. Definition of quantitative K-theory
1.4. Elementary properties of quantitative K-theory
1.5. Morita equivalence
1.6. Lipschitz homotopies
2. Controlled morphisms
2.1. Definition and main properties
2.2. Controlled exact sequences
3. Extensions of filtered C∗-algebras
3.1. Semi-split filtered extensions
3.2. Controlled boundary maps
3.3. Long exact sequence
3.4. The mapping cones
4. Controlled Bott periodicity
4.1. Tensorization in KK-theory
4.2. The controlled Bott isomorphism
4.3. The six term (λ, h)-exact sequence
5. Quantitative K-theory for crossed product C∗-algebras
5.1. Lengths and propagation
5.2. Kasparov transformation
5.3. Application to K-amenability
6. The quantitative Baum-Connes conjecture
6.1. The Rips complex
2
3
3
4
7
10
12
15
16
17
19
20
20
22
26
29
32
32
38
39
40
40
41
44
46
46
2000 Mathematics Subject Classification. 19K35,46L80,58J22.
Key words and phrases. Baum-Connes Conjecture, Coarse Geometry, Group and Crossed
product C ∗-algebras Novikov Conjecture, Operator Algebra K-theory, Roe Algebras.
Oyono-Oyono is partially supported by the ANR "Kind" and Yu is partially supported by a
grant from the U.S. National Science Foundation.
1
2
H. OYONO-OYONO AND G. YU
6.2. Quantitative statements
7. Further comments
References
48
52
52
0. Introduction
The receptacles of higher indices of elliptic differential operators are K-theory
of C∗-algebras which encode the (large scale) geometry of the underlying spaces.
The following examples are important for purpose of applications to geometry and
topology.
equivariant elliptic differential operators on covering space [1, 2, 5, 11];
• K-theory of group C∗-algebras is a receptacle for higher index theory of
• K-theory of crossed product C∗-algebras and more generally groupoid C∗-
algebras for foliations serve as receptacles for longitudinally elliptic oper-
ators [3, 4];
• the higher indices of elliptic operators on noncompact Riemannian mani-
folds live in K-theory of Roe algebras [15].
The local nature of differential operators implies that these higher indices can be
defined in term of idempotents and invertible elements with finite propagation.
Using homotopy invariance of the K-theory for C∗-algebras, these higher indices
give rise to topological invariants.
In the context of Roe algebras, a quantitative operator K-theory was introduced
to compute the higher indices of elliptic operators for noncompact spaces with fi-
nite asymptotic dimension [19]. The aim of this paper is to develop a quantitative
K-theory for general C∗-algebras equipped with a filtration. The filtration struc-
ture allows us to define the concept of propagation. Examples of C∗-algebras with
filtrations include group C∗-algebras, crossed product C∗-algebras and Roe alge-
bras. The quantitative K-theory for C∗-algebras with filtrations is then defined in
terms of homotopy of quasi-projections and quasi-unitaries with propagation and
norm controls. We introduce controlled morphisms to study quantitative operator
K-theory.
In particular, we derive a quantitative version of the six term exact
sequence. In the case of crossed product algebras, we also define a quantitative ver-
sion of the Kasparov transformation compatible with Kasparov product. We end
this paper by using the quantitative K-theory to formulate a quantitative version
of the Baum-Connes conjecture and prove it for a large class of groups.
This paper is organized as follows: In section 1, we collect a few notations and
definitions including the concept of filtered C∗-algebras. We use the concepts of
almost unitary and almost projection to define a quantitative K-theory for filtered
C∗-algebras and we study its elementary properties. In section 2, we introduce the
notion of controlled morphism in quantitative K-theory. Section 3 is devoted to
extensions of filtered C∗-algebras and to a controlled exact sequence for quantitative
K-theory. In section 4, we prove a controlled version of the Bott periodicity and
as a consequence, we obtain a controlled version of the six-term exact sequence in
K-theory. In section 5, we apply KK-theory to study the quantitative K-theory of
crossed product C∗-algebras and discuss its application to K-amenability. Finally
ON A QUANTITATIVE OPERATOR K-THEORY
3
in section 8, we formulate a quantitative Baum-Connes conjecture and prove the
quantitative Baum-Connes conjecture for a large class of groups.
1. Quantitative K-theory
In this section, we introduce a notion of quantitative K-theory for C∗-algebras
with a filtration. Let us fix first some notations about C∗-algebras we shall use
throughout this paper.
b1
. . .
bk
of Mn1+···+nk (B).
• If B is a C∗-algebra and if b1, . . . , bk are respectively elements of
Mn1(B), . . . , Mnk (B), we denote by diag(b1, . . . , bk) the block diagonal
matrix
• If X is a locally compact space and B is a C∗-algebra, we denote by
C0(X, B) the C∗-algebra of B-valued continuous functions on X vanishing
at infinity. The special cases of X = (0, 1], X = [0, 1), X = (0, 1) and
X = [0, 1], will be respectively denoted by CB, B[0, 1), SB and B[0, 1].
• For a separable Hilbert space H, we denote by K(H) the C∗-algebra of
• If A and B are C∗-algebras, we will denote by A ⊗ B their spatial tensor
compact operators on H.
product.
1.1. Filtered C∗-algebras.
Definition 1.1. A filtered C∗-algebra A is a C∗-algebra equipped with a family
(Ar)r>0 of linear subspaces indexed by positive numbers such that:
• Ar ⊂ Ar′ if r 6 r′;
• Ar is stable by involution;
• Ar · Ar′ ⊂ Ar+r′ ;
• the subalgebra [r>0
Ar is dense in A.
If A is unital, we also require that the identity 1 is an element of Ar for every
positive number r. The elements of Ar are said to have propagation r.
• Let A and A′ be respectively C∗-algebras filtered by (Ar)r>0 and (A′r)r>0.
A homomorphism of C∗ -algebras φ : A−→A′ is a filtered homomorphism
(or a homomorphism of filtered C∗-algebras) if φ(Ar) ⊂ A′r for any positive
number r.
• If A is a filtered C∗-algebra and X is a locally compact space, then
In particular the
C0(X, A) is a C∗-algebra filtered by (C0(X, Ar))r>0.
algebras CA, A[0, 1], A[0, 1) and SA are filtered C∗-algebras.
• If A is a non unital filtered C∗-algebra, then its unitarization eA is filtered
by (Ar + C)r>0. We define for A non-unital the homomorphism
for a ∈ A and z ∈ C.
ρA : eA → C; a + z 7→ z
Prominent examples of filtered C∗-algebra are provided by Roe algebras asso-
ciated to proper metric spaces, i.e. metric spaces such that closed balls of given
radius are compact. Recall that for such a metric space (X, d), a X-module is a
Hilbert space HX together with a ∗-representation ρX of C0(X) in HX (we shall
4
H. OYONO-OYONO AND G. YU
write f instead of ρX (f )). If the representation is non-degenerate, the X-module is
said to be non-degenerate. A X-module is called standard if no non-zero function
of C0(X) acts as a compact operator on HX .
The following concepts were introduced by Roe in his work on index theory of
elliptic operators on noncompact spaces [15].
Definition 1.2. Let HX be a standard non-degenerate X-module and let T be a
bounded operator on HX .
(i) The support of T is the complement of the open subset of X × X
{(x, y) ∈ X × X s.t. there exist f and g in C0(X) satisfying
f (x) 6= 0, g(y) 6= 0 and f · T · g = 0}.
(ii) The operator T is said to have finite propagation (in this case propagation
less than r) if there exists a real r such that for any x and y in X with
d(x, y) > r, then (x, y) is not in the support of T .
(iii) The operator T is said to be locally compact if f · T and T · f are compact
for any f in C0(X). We then define C[X] as the set of locally compact
and finite propagation bounded operators of HX , and for every r > 0, we
define C[X]r as the set of element of C[X] with propagation less than r.
We clearly have C[X]r · C[X]r′ ⊂ C[X]r+r′. We can check that up to (non-
canonical) isomorphism, C[X] does not depend on the choice of HX .
Definition 1.3. The Roe algebra C∗(X) is the norm closure of C[X] in the algebra
L(HX ) of bounded operators on HX . The Roe algebra in then filtered by (C[X]r)r>0.
Although C∗(X) is not canonically defined, it was proved in [9] that up to
canonical isomorphisms, its K-theory does not depend on the choice of a non-
degenerate standard X-module. Furthermore, K∗(C∗(X)) is the natural receptacle
for higher indices of elliptic operators with support on X [15].
If X has bounded geometry, then the Roe algebra admits a maximal version [7]
filtered by (C[X]r)r>0. Other important examples are reduced and maximal crossed
product of a C∗-algebra by an action of a discrete group by automorphisms. These
examples will be studied in detail in section 5.
1.2. Almost projections/unitaries. Let A be a unital filtered C∗-algebra. For
any positive numbers r and ε, we call
• an element u in A a ε-r-unitary if u belongs to Ar, ku∗ · u − 1k < ε and
ku · u∗ − 1k < ε. The set of ε-r-unitaries on A will be denoted by Uε,r(A).
• an element p in A a ε-r-projection if p belongs to Ar, p = p∗ and kp2−pk <
ε. The set of ε-r-projections on A will be denoted by Pε,r(A).
For n integer, we set Uε,r
n (A) = Uε,r(Mn(A)) and Pε,r
n (A) = Pε,r(Mn(A)).
For any unital filtered C∗-algebra A, any positive numbers ε and r and any
positive integer n, we consider inclusions
and
n (A) ֒→ Pε,r
Pε,r
n (A) ֒→ Uε,r
Uε,r
n+1(A); p 7→(cid:18)p 0
0(cid:19)
n+1(A); u 7→(cid:18)u 0
1(cid:19) .
0
0
ON A QUANTITATIVE OPERATOR K-THEORY
5
This allows us to define
and
Uε,r
∞ (A) = [n∈N
∞ (A) = [n∈N
Pε,r
Uε,r
n (A)
Pε,r
n (A).
Remark 1.4. Let r and ε be positive numbers with ε < 1/4;
(i) If p is an ε-r-projection in A, then the spectrum of p is included in
2
2
, 1+√1+4ε
(cid:17) ∪(cid:16) 1+√1−4ε
1 − ε < kuk < 1 + ε/2,
1 − ε/2 < ku−1k < 1 + ε,
ku∗ − u−1k < (1 + ε)ε.
(cid:16) 1−√1+4ε
2
, 1−√1−4ε
2
(ii) If u is an ε-r-unitary in A, then
(cid:17) and thus kpk < 1 + ε.
and κ0,ε(t) = 1 if t > 1+√1−4ε
(iii) Let κ0,ε : R → R be a continuous function such that κ0,ε(t) = 0 if t 6
1−√1−4ε
. If p is an ε-r-projection in A,
then κ0,ε(p) is a projection such that kp − κ0,ε(p)k < 2ε which moreover
does not depends on the choice of κ0,ε. From now on, we shall denote this
projection by κ0(p).
2
2
(iv) If u is an ε-r-unitary in A, set κ1(u) = u(u∗u)−1/2. Then κ1(u) is a
unitary such that ku − κ1(u)k < ε.
1 − 2ε, then κ0(p) and q are homotopic projections [18, Chapter 5].
(v) If p is an ε-r-projection in A and q is a projection in A such that kp− qk <
(vi) If u and v are ε-r-unitaries in A, then uv is an ε(2 + ε)-2r-unitary in A.
Definition 1.5. Let A be a C∗-algebra filtered by (Ar)r>0.
• Let p0 and p1 be ε-r-projections. We say that p0 and p1 are homotopic ε-
r-projections if there exists a ε-r-projection q in A[0, 1] such that q(0) = p0
and q(1) = p1. In this case, q is called a homotopy of ε-r-projections in A
and will be denoted by (qt)t∈[0,1].
• If A is unital, let u0 and u1 be ε-r-unitaries. We say that u0 and u1
are homotopic ε-r-unitaries if there exists an ε-r-unitary v in A[0, 1] such
that v(0) = u0 and v(1) = u1. In this case, v is called a homotopy of
ε-r-unitaries in A and will be denoted by (vt)t∈[0,1].
Example 1.6. Let p be a ε-projection in a filtered unital C∗-algebra A. Set ct =
cos πt/2 and st = sin πt/2 for t ∈ [0, 1] and let us considerer the homotopy of projec-
tions (ht)t∈[0,1] with ht =(cid:18) c2
t (cid:19) in M2(C) between diag(1, 0) and diag(0, 1).
Set (qt)t∈[0,1] = (diag(p, 0)+(1−p)⊗ht)t∈[0,1]. Since q2
t (p2−p)⊗I2, we see
that (qt)t∈[0,1] is a homotopy of ε-r-projections between diag(1, 0) and diag(p, 1− p)
in M2(A).
t −qt = s2
ctst
s2
t
ctst
Next result will be frequently used throughout the paper and is quite easy to
prove.
Lemma 1.7. Let A be a C∗-algebra filtered by (Ar)r>0.
6
H. OYONO-OYONO AND G. YU
4
(i) If p is an ε-r-projection in A and q is a self-adjoint element of Ar such
that kp − qk < ε−kp2−pk
, then q is ε-r-projection . In particular, if p is
an ε-r-projection in A and if q is a self-adjoint element in Ar such that
kp− qk < ε, then q is a 5ε-r-projection in A and p and q are connected by
a homotopy of 5ε-r-projections.
(ii) If A is unital and if u is an ε-r-unitary and v is an element of Ar such
that ku − vk < ε−ku∗u−1k
, then v is an ε-r-unitary . In particular, if u
is an ε-r-unitary and v is an element of Ar such that ku − vk < ε, then
v is an 4ε-r-unitary in A and u and v are connected by a homotopy of
4ε-r-unitaries.
3
Lemma 1.8. There exists a real λ > 4 such that for any positive number ε with
ε < 1/λ, any positive real r, any ε-r-projection p and ε-r-unitary W in a filtered
unital C∗-algebra A, the following assertions hold:
(i) W pW ∗ is a λε-3r-projection of A;
(ii) diag(W pW ∗, 1) and diag(p, 1) are homotopic λε-3r-projections.
Proof. The first point is straightforward to check from remark 1.4. For the second
point, with notations of example 1.6, use the homotopy of ε-r-unitaries
(cid:16) W c2
t +s2
t
(W−1)stct W s2
(W−1)stct
t +c2
t (cid:17)t∈[0,1]
=(cid:0)(cid:0) ct −st
st ct (cid:1) · diag(W, 1) · ( ct st
−st ct )(cid:1)t∈[0,1]
to connect by conjugation diag(W pW ∗, 1) to diag(p, W W ∗) and then connect to
diag(p, 1) by a ray.
(cid:3)
Recall that if two projections in a unital C∗-algebra are close enough in norm,
then there are conjugated by a canonical unitary. To state a similar result in term
of ε-r-projections and ε-r-unitaries, we will need the definition of a control pair.
Definition 1.9. A control pair is a pair (λ, h), where
• λ > 1;
• h : (0, 1
4λ ) → (0, +∞); ε 7→ hε is a map such that there exists a non-
increasing map g : (0, 1
4λ ) → (0, +∞), with h 6 g.
Lemma 1.10. There exists a control pair (λ, h) such that the following holds:
for every positive number r, any ε in (0, 1
4λ ) and any ε-r-projections p and q of a
filtered unital C∗-algebra A satisfying kp−qk < 1/16, there exists an λε-hεr-unitary
W in A such that kW pW ∗ − qk 6 λε.
Proof. We follow the proof of [18, Proposition 5.2.6]. If we set
• then
z = (2κ0(p) − 1)(2κ0(q) − 1) + 1,
kz − 2k 6 2kκ0(p) − κ0(q)k
6 8ε + 2kp − qk
and hence z is invertible for ε < 1/16.
• Moreover, if we set U = zz−1 and since zκ0(q) = κ0(p)z, then we have
κ0(q) = U κ0(p)U∗.
ON A QUANTITATIVE OPERATOR K-THEORY
7
Let us define z′ = (2p − 1)(2q − 1) + 1. Then we have kz − z′k 6 9ε and kz′k 6 3.
If ε is small enough, then kz′∗z′ − 4k 6 2 and hence the spectrum of z′∗z′ is in
[2, 6]. Let us consider the expansion in power serie Pk∈N aktk of t 7→ (1 + t)−1/2
on (0, 1) and let nε be the smallest integer such that Pnε6k ak/2k 6 ε. Let us
set then W = z′/2Pnε
)k. Then for a suitable λ (not depending on
A, p, q or ε), we get that W is a λε-(4nε + 2)r-unitary which satisfies the required
condition.
(cid:3)
k=0 ak( z′∗z′−4
4
Remark 1.11. The order of h when ε goes to zero in lemma 1.10 is Cε−3/2 for
some constant C.
1.3. Definition of quantitative K-theory. For a unital filtered C∗-algebra A,
we define the following equivalence relations on Pε,r
• if p and q are elements of Pε,r
∞ (A) × N and on Uε,r
(q, l′) if there exists a positive integer k and an element h of Pε,r
such that h(0) = diag(p, Ik+l′ ) and h(1) = diag(q, Ik+l).
∞ (A), l and l′ are positive integers, (p, l) ∼
∞ (A[0, 1])
∞ (A), u ∼ v if there exists an element h of
• if u and v are elements of Uε,r
Uε,r
∞ (A[0, 1]) such that h(0) = u and h(1) = v.
If p is an element of Pε,r
class of (p, l) modulo ∼ and if u is an element of Uε,r
equivalence class modulo ∼.
Definition 1.12. Let r and ε be positive numbers with ε < 1/4. We define:
∞ (A) and l is an integer, we denote by [p, l]ε,r the equivalence
∞ (A) we denote by [u]ε,r its
∞ (A):
∞ (A) × N/ ∼ for A unital and
(i) K ε,r
0 (A) = Pε,r
K ε,r
0 (A) = {[p, l]ε,r ∈ Pε,r( A) × N/ ∼ such that dim κ0(ρA(p)) = l}
for A non unital.
1 (A) = Uε,r
∞ ( A)/ ∼ (with A = A if A is already unital).
(ii) K ε,r
Remark 1.13. We shall see in lemma 1.24 that as it is the case for K-theory,
K ε,r
∗ (•) can indeed be defined in a uniform way for unital and non-unital filtered
C∗-algebras.
It is straightforward to check that for any unital filtered C∗-algebra A, if p is an
ε-r-projection in A and u is an ε-r-unitary in A, then diag(p, 0) and diag(0, p) are
homotopic ε-r-projections in M2(A) and diag(u, 1) and diag(1, u) are homotopic
ε-r-unitaries in M2(A). Thus we obtain the following:
Lemma 1.14. Let A be a filtered C∗-algebra. Then K ε,r
equipped with a structure of abelian semi-group such that
0 (A) and K ε,r
1 (A) are
and
[p, l]ε,r + [p′, l′]ε,r = [diag(p, p′), l + l′]ε,r
[u]ε,r + [u′]ε,r = [diag(u, v)]ε,r,
for any [p, l]ε,r and [p′, l′]ε,r in K ε,r
0 (A) and any [u]ε,r and [u′]ε,r in K ε,r
1 (A).
According to example 1.6, for every unital filtered C∗-algebra A, any ε-r-projection
p in Mn(A) and any integer l with n > l, we see that [In − p, n − l]ε,r is an inverse
for [p, l]ε,r. Hence we obtain:
Lemma 1.15. If A is a filtered C∗-algebra, then K ε,r
0 (A) is an abelian group.
Although K ε,r
1 (A) is not a group, it is very close to be one.
8
H. OYONO-OYONO AND G. YU
Lemma 1.16. Let A be a filtered C∗-algebra. Then for any ε-r-unitary u in Mn( A)
(with A = A if A is already unital), we have [u]3ε,2r + [u∗]3ε,2r = 0 in K 3ε,2r
(A).
1
Proof. If u is an ε-r-unitary in a unital filtered C∗-algebra A, then according to
−st ct )(cid:1)t∈[0,1]
point (vi) of remark 1.4, we see that(cid:0)diag(1, u)(cid:0) ct −st
is a homotopy of 3ε-2r-unitaries between diag(u, u∗) and diag(uu∗, 1). Since kuu∗−
1k < ε, we deduce from lemma 1.7 that uu∗ and 1 are homotopic 3ε-2r-unitaries
and hence we get the lemma.
st ct (cid:1) · diag(1, u∗) · ( ct st
(cid:3)
Remark 1.17. According to lemma 1.16, if we define the equivalence relation on
Uε,r
∞ (A) to be homotopy within U3ε,2r
1 (A) can be endowed with an
abelian group structure.
∞ (A), then K ε,r
We have for any filtered C∗-algebra A and any positive numbers r, r′, ε and ε′
with ε 6 ε′ < 1/4 and r 6 r′ natural semi-group homomorphisms
0
1
• ιε,r
: K ε,r
• ιε,r
: K ε,r
• ιε,r
∗ = ιε,r
• ιε,ε′,r,r′
• ιε,ε′,r,r′
• ιε,ε′,r,r′
0 (A)−→K0(A); [p, l]ε,r 7→ [κ0(p)] − [Il];
1 (A)−→K1(A); [u]ε,r 7→ [u];
0 ⊕ ιε,r
1 ;
0 (A)−→K ε′,r′
: K ε,r
1 (A)−→K ε′,r′
: K ε,r
⊕ ιε,ε′,r,r′
= ιε,ε′,r,r′
(A); [p, l]ε,r 7→ [p, l]ε′,r′;
(A); [u]ε,r 7→ [u]ε′,r′.
∗
1
1
1
0
1
0
If some of the indices r, r′ or ε, ε′ are equal, we shall not repeat it in ιε,ε′,r,r′
Remark 1.18. Let p0 and p1 be two ε-r-projections in a filtered C∗-algebra such
that κ0(p0) and κ0(p1) are homotopic projections. Then for any ε in (0, 1/4), this
homotopy can be approximated for some r′ by a ε-r′-projection. Hence, using point
(iii) of remark 1.4, there exists a homotopy (qt)t∈[0,1] of ε-r′ projections in A such
that kp0 − q0k < 3ε and kp1 − q1k < 3ε. We can indeed assume that r′ > r and thus
by lemma 1.7, we get that p0 and p1 are homotopic as 15ε-r′-projections. Proceeding
in the same way for the odd case we eventually obtain:
∗
.
there exists λ > 1 such that for any filtered C∗-algebra A, any ε ∈ (0, 1
any positive number r, the following holds:
4λ ) and
∗ (A) such that ιε,r
Let x and x′ be elements in K ε,r
∗ (x) = ιε,r
there exists a positive number r′ with r′ > r such that ιε,λε,r,r′
K λε,r′
∗
Lemma 1.19. Let p be a matrix in Mn(C) such that p = p∗ and kp2 − pk < ε for
some ε in (0, 1/4). Then there is a continuous path (pt)t∈[0,1] in Mn(C) such that
∗ (x′) in K∗(A), then
(x) = ιε,λε,r,r′
(x′) in
(A).
∗
∗
• p0 = p;
• p1 = Ik with k = dim κ0(p);
• p∗t = pt and kp2
t − ptk < ε for every t in [0, 1].
Proof. The selfadjoint matrix p satisfies kp2 − pk < ε if and only if the eigenvalues
of p satisfy the inequality
i.e.
−ε < λ2 − λ < ε,
λ ∈(cid:18) 1 − √1 + 4ε
2
,
1 − √1 − 4ε
2
(cid:19)[(cid:18)√1 − 4ε + 1
2
,
√1 + 4ε + 1
2
(cid:19) .
ON A QUANTITATIVE OPERATOR K-THEORY
9
Let λ1, . . . , λk be the eigenvalues of p lying in (cid:16) 1−√1+4ε
λk+1, . . . , λn be the eigenvalues of p lying in (cid:16) √1−4ε+1
t ∈ [0, 1]
2
2
,
√1+4ε+1
2
, 1−√1−4ε
2
(cid:17) and let
(cid:17). We set for
• λi,t = tλi for i = 1, . . . , k;
• λi,t = tλi + 1 − t for i = k + 1, . . . , n.
Since λ 7→ λ2 − λ is decreasing on (cid:16) 1−√1+4ε
(cid:16)√1−4ε+1
(cid:17) then,
√1+4ε+1
2
2
2
,
−ε < λ2
i,t − λi,t < ε
, 1−√1−4ε
2
(cid:17) and increasing on
for all t in [0, 1] and i = 1, . . . , n. If we set pt = u · diag(λ1,t, . . . , λn,t) · u∗ where u
is a unitary matrix of Mn(C) such that p = u · diag(λ1, . . . , λn) · u∗, then
• p0 = p;
• p1 = κ0(p);
• p∗t = pt and kp2
t − ptk < ε for every t in [0, 1].
Since there is a homotopy of projections in Mn(C) between κ0(p) and Ik with
k = dim κ0(p), we get the result.
(cid:3)
As a consequence we obtain:
Corollary 1.20. For any positive numbers r and ε with ε < 1/4, then
K ε,r
0 (C) → Z; [p, l]ε,r 7→ dim κ0(p) − l
is an isomorphism.
Lemma 1.21. Let u be a matrix in Mn(C) such that ku∗u − Ink < ε and kuu∗ −
Ink < ε for ε in (0, 1/4). Then there is a continuous path (ut)t∈[0,1] in Mn(C) such
that
• u0 = u;
• u1 = In;
• ku∗t ut − Ink < ε and kutu∗t − Ink < ε for every t in [0, 1].
Proof. Since u is invertible, u∗u and uu∗ have the same eigenvalues λ1, . . . , λn, and
thus ku∗t ut − Ink < ε and kutu∗t − Ink < ε if and only if λi ∈ (1 − ε, 1 + ε) for
i = 1, . . . , n. Let us set
1
n
, . . . , λ−t/2
) · w∗ where w is a unitary matrix of Mn(C)
• ht = w · diag(λ−t/2
such that u∗u = w · diag(λ1, . . . , λn) · w∗;
• vt = u · ht for all t ∈ [0, 1]. Then v∗t vt = w · diag(λ1−t
Since λ1−t
i −1 < ε for all all t ∈ [0, 1], we get that kv∗t vt−Ink < ε and kvtv∗t −Ink <
ε for every t in [0, 1]. The matrix v1 is unitary and the result then follows from
path-connectness of Un(C).
(cid:3)
n ) · w∗.
, . . . , λ1−t
1
As a consequence we obtain:
Corollary 1.22. For any positive numbers r and ε with ε < 1/4, then we have
K ε,r
1 (C) = {0}.
10
H. OYONO-OYONO AND G. YU
1.4. Elementary properties of quantitative K-theory. Let A1 and A2 be two
unital C∗-algebras respectively filtered by (A1,r)r>0 and (A2,r)r>0 and consider
∞ (A1 ⊕ A2) ∼=
A1 ⊕ A2 filtered by (A1,r ⊕ A2,r)r>0. Since we have identifications Pε,r
Pε,r
∞ (A1) × Pε,r
∞ (A2) induced by the
inclusions A1 ֒→ A1 ⊕ A2 and A2 ֒→ A1 ⊕ A2, we see that we have isomorphisms
K ε,r
0 (A1)⊕K ε,r
1 (A1⊕A2).
Lemma 1.23. Let A be a filtered non unital C∗-algebra and let ε and r be positive
numbers with ε < 1/4. We have a natural splitting
∞ (A1 ⊕ A2) ∼= Uε,r
0 (A1⊕A2) and K ε,r
∞ (A2) and Uε,r
0 (A2) ∼−→ K ε,r
∞ (A1) × Uε,r
1 (A1)⊕K ε,r
1 (A2) ∼−→ K ε,r
Proof. Viewing A as a subalgebra of A, the group homomorphisms
0 ( A) ∼=−→ K ε,r
K ε,r
0 (A) ⊕ Z.
K ε,r
0 ( A) −→ K ε,r
[p, l]ε,r
0 (A) ⊕ Z
7→ ([p, dim κ0(ρA(p))]ε,r, dim κ0(ρA(p)) − l)
and
0 (A) ⊕ Z −→ K ε,r
K ε,r
7→ (cid:20)(cid:18)p
([p, l]ε,r, k − k′)
0 ( A)
0
0 Ik(cid:19) , l + k′(cid:21)ε,r
are inverse one of the other.
(cid:3)
Let us set A+ = A ⊕ C equipped with the multiplication
(a, x) · (b, y) = (ab + xb + ya, xy)
for a and b in A and x and y in C. Notice that
• A+ is isomorphic to A ⊕ C with the algebra structure provided by the
direct sum if A is unital;
• A+ = A if A is not unital.
Let us define also ρA in the unital case by ρA : A+ → C; (a, x) 7→ x. We know
that in usual K-theory, we can equivalently define for A unital the Z2-graded group
K∗(A) as A+ by
K0(A) = ker ρA,∗ : K0(A+) → K0(C) ∼= Z
and
K1(A) = K1(A+).
ε,r(A). If
Let us check that this is also the case for our Z2-graded semi-groups K∗
the C∗-algebra A is filtered by (Ar)r>0, then A+ is filtered by (Ar + C)r>0. Let us
define for a unital filtered algebra A
K′0
ε,r(A) = {[p, l]ε,r ∈ Pε,r(A+) × N/ ∼ such that dim κ0(ρA(p)) = l}
and
K′1
ε,r(A) = Uε,r(A+)/ ∼ .
Proceeding as we did in the proof of lemma 1.23, we obtain a natural splitting
0 (A+) ∼=−→ K′0
K ε,r
ε,r(A) ⊕ Z.
But then, using the identification A+ ∼= A ⊕ C and in view of lemmas 1.19 and
1.21, we get
ON A QUANTITATIVE OPERATOR K-THEORY
11
Lemma 1.24. The Z2-graded semi-groups K ε,r
isomorphic.
∗ (A) and K′
∗
ε,r(A) are naturally
This allows us to state functoriallity properties for quantitative K-theory.
If
φ : A → B is a homomorphism of unital filtered C∗-algebras, then since φ preserve
ε-r-projections and ε-r-unitaries, it obviously induces for any positive number r
and any ε ∈ (0, 1/4) a semi-group homomorphism
∗ (A) −→ K ε,r
∗ (B).
: K ε,r
φε,r
∗
In the non unital case, we can extend any homomorphism φ : A → B to a homomor-
phism φ+ : A+ → B+ of unital filtered C∗-algebras and then we use lemmas 1.23
and 1.24 to define φε,r
∗ (B). Hence, for any positive number r and
∗
any ε ∈ (0, 1/4), we get that K ε,r
1 (•)) is a covariant additive functor
from the category of filtered C∗-algebras (together with filtered homomorphism) to
the category of abelian groups (resp. semi-groups).
0 (•) (resp. K ε,r
∗ (A) −→ K ε,r
: K ε,r
Definition 1.25.
(i) Let A and B be filtered C∗-algebras. Then two homomorphisms of filtered
C∗-algebras ψ0 : A → B and ψ1 : A → B are homotopic if there exists a
path of homomorphisms of filtered C∗-algebras ψt : A → B for 0 6 t 6 1
between ψ0 and ψ1 and such that t 7→ ψt is continuous for the pointwise
norm convergence.
(ii) A filtered C∗-algebra A is said to be contractible if the identity map and
the zero map of A are homotopic.
Example 1.26. If A is a filtered C∗-algebra A, then the cone of A
is a contractible filtered C∗-algebra.
CA = {f ∈ C([0, 1], A) such that f (0) = 0}
We have then the following obvious result:
Lemma 1.27. If φ : A → B and φ′ : A → B are two homotopic homomorphisms
of filtered C∗-algebras, then φε,r
for every positive numbers ε and r with
ε < 1/4. In particular, if A is a contractible filtered C∗-algebra, then K ε,r
∗ (A) = 0
for every positive numbers ε and r with ε < 1/4.
∗ = φ′ε,r
∗
Let A be a C∗-algebra filtered by (Ar)r>0 and let (Bk)k∈N be an increasing
sequence of C∗-subalgebras of A such that [k∈N
Bk is dense in A. Assume that
Sr>0 Bk ∩ Ar is dense in Bk for every integer k. Then for every integer k, the
C∗-algebra Bk is filtered by (Bk ∩ Ar)r>0. If A is unital, then Bk is unital for some
k, and thus we will assume without loss of generality that Bk is unital for every
integer k.
Proposition 1.28. Let A be a unital C∗-algebra filtered by (Ar)r>0 and let (Bk)k∈N
be an increasing sequence of C∗-subalgebra of A such that
(Bk ∩ Ar) is dense in Bk for every integer k,
(Bk ∩ Ar) is dense in Ar for every positive number r.
• [r>0
• [k∈N
Then the Z2-graded semi-groups K ε,r
∗ (A) and lim
k
K ε,r
∗ (Bk) are isomorphic.
12
H. OYONO-OYONO AND G. YU
Proof. In particular, we see that [k∈N
Bk is dense in A. Let us denote by
Υ∗,ε,r : lim
k
K ε,r
∗ (Bk) → K ε,r
∗ (A)
12 . Since [k∈N
the homomorphism of semi-group induced by the family of inclusions Bk ֒→ A
where k runs through integers. We give the proof in the even case, the odd case
being analogous. Let p be an element of Pε,r
n (A) and let δ = kp2 − pk > 0 and
choose α < ε−δ
(Bk ∩ Ar) is dense in Ar, there is an integer k and a
selfadjoint element q of Mn(Bk ∩ Ar) such that kp − qk < α. According to lemma
1.19, q is a ε-r projection. Let q′ be another selfadjoint element of Mn(Bk ∩ Ar)
such that kp − q′k < α. Then kq − q′k < 2α and if we set qt = (1 − t)q + tq′ for
t ∈ [0, 1], then
kq2
t − qtqk + kqtq − q2k + kq2 − qk + kq − qtk
t − qtk 6 kq2
6 kqt − qk(kqtk + kqk + 1) + 4α + δ
6 12α + δ
< ε,
and thus q and q′ are homotopic in Pε,r
some Mn(Bk ∩ Ar) satisfying kq − pk < kp2−pk
image of [q, l]ε,r in lim
k
inverse for Υ0,ε,r. We proceed similarly in the odd case.
n (A) and q in
, we define Υ′0,ε,r([p, l]ε,r) to be the
K ε,r
∗ (Bk). Then Υ′0,ε,r is a group homomorphism and is an
n (Bk). Therefore, for p ∈ Pε,r
(cid:3)
12
1.5. Morita equivalence. For any unital filtered algebra A, we get an identifi-
cation between Pε,r
∞ (Mk(A)) and
Pε,r
∞ (A). This identification gives rise to a natural group isomorphism between
K ε,r
0 (A) and K ε,r
0 (Mk(A)), and this isomorphism is induced by the inclusion of
C∗-algebras
nk (A) and therefore between Pε,r
n (Mk(A)) and Pε,r
ιA : A ֒→ Mk(A); a 7→ diag(a, 0).
Namely, if we set e1,1 = diag(1, 0, . . . , 0) ∈ Mk(C), definition of the functoriality
yields
ιε,r
A,∗
for any p in Pε,r
[p, l]ε,r = [p ⊗ e1,1 + Il ⊗ (Ik − e1,1), l]ε,r ∈ K ε,r
n (A) and any integer l with l 6 n. We can verify that
0 (Mk(A))
(ιε,r
A,∗
)−1[q, l]ε,r = [q, kl]ε,r
for any q in Pε,r
n (Mk(A)) and any integer l with l 6 n, where on the right hand
side of the equality, the matrix q of Mn(Mk(A)) is viewed as a matrix of Mnk(A).
∞ (Mk(A))
1 (A) and
1 (Mk(A)). This isomorphism is also induced by the inclusion ιA and we have
In a similar way, we obtain in the odd case an identification between Uε,r
∞ (A) providing a natural semi-group isomorphism between K ε,r
and Uε,r
K ε,r
for any x in Uε,r
ιA,∗[x]ε,r = [x ⊗ e1,1 + In ⊗ (Ik − e1,1)]ε,r ∈ K ε,r
n (A).
1 (Mk(A))
Let us deal now with the non-unital case. For usual K-theory, Morita equivalence
for non-unital C∗-algebra can be deduced from the unital case by using the six-
term exact sequence associated to the split extension 0 → A → A → C → 0.
But for quantitative K-theory this splitting only gives rise (in term of section 2.1)
to a controlled isomorphism (see corollary 4.9). In order to really have a genuine
ON A QUANTITATIVE OPERATOR K-THEORY
13
isomorphism, we have to go through the tedious following computation. If B is a
non-unital C∗-algebra, let us identify Mk( B) with Mk(B) ⊕ Mk(C) equipped with
the product
(b, λ) · (b′, λ′) = (bb′ + λb′ + bλ′, λλ′)
for b and b′ in Mk(B) and λ and λ′ in Mk(C). Under this identification, if A is not
unital, let us check that the semi-group homomorphism
Φ1 : K ε,r
1 ( A) → K ε,r
1 ( ^Mk(A)); [(x, λ)]ε,r 7→ [(x ⊗ e1,1, λ]ε,r
induced by the inclusion ιA is invertible with inverse given by the composition
Ψ1 : K ε,r
1 ( ^Mk(A)) → K ε,r
1 (Mk( A)) ∼=→ K ε,r
1 ( A),
where the first homomorphism of the composition is induced by the inclusion
^Mk(A) → Mk( A); (a, z) 7→ (a, zIk).
Let (x, λ) be an element of Uε,r
n ( A), with x ∈ Mn(A) and λ ∈ Mn(C). Then
Ψ1 ◦ Φ1[(x, λ)]ε,r = [(x ⊗ e1,1, λ ⊗ Ik)]ε,r,
where we use the identification Mnk(C) ∼= Mn(C)⊗ Mk(C) to see x⊗ e1,1 and λ⊗ Ik
respectively as matrices in Mnk(A) and Mnk(C). According to lemma 1.21, as a
ε-r-unitary of Mn(C), λ is homotopic to In. Hence
[(x ⊗ e1,1, λ ⊗ Ik)]ε,r = [(x ⊗ e1,1, λ ⊗ e1,1 + In⊗Ik−1)]
and from this we get that Ψ1 ◦ Φ1 is induced in K-theory by the inclusion map
A ֒→ Mk( A); a 7→ diag(a, 0) which is the identity homomorphism (according to the
unital case). Conversely, let (y, λ) be an element in Uε,r
n ( ^Mk(A)) with
and λ ∈ Mn(C). Then
y ∈ Mn(Mk(A)) ∼= Mn(A) ⊗ Mk(C)
Φ1 ◦ Ψ1[(y, λ)]ε,r = [(y ⊗ e1,1, λ ⊗ Ik)]ε,r,
where
Let
• y ⊗ e1,1 belongs to Mn(Mk(A))⊗ Mk(C) ∼= Mn(A)⊗ Mk(C)⊗ Mk(C) (the
first two factors provide the copy of Mn(Mk(A)) where y lies in and e1,1
lies in the last factor).
• λ⊗Ik belongs to the algebra Mn(Mk(C)) ∼= Mn(C)⊗Mk(C) that multiplies
Mn(A) ⊗ Mk(C) ⊗ Mk(C) on the first two factors.
σ : Mn(A) ⊗ Mk(C) ⊗ Mk(C) → Mn(A) ⊗ Mk(C) ⊗ Mk(C)
be the C∗-algebra homomorphism induced by the flip of Mk(C)⊗ Mk(C). This flip
can be realized by conjugation of a unitary U in Mk(C) ⊗ Mk(C) ∼= Mk2(C). Let
(Ut)t∈[0,1] be a homotopy in Uk2 (C) between U and Ik2 . Let us define
A = {(x, z⊗Ik); x ∈ Mn(A)⊗Mk(C)⊗Mk(C), z ∈ Mn(C)} ⊂ Mn( ^Mk(A))⊗Mk(C),
where z ⊗ Ik is viewed as z ⊗ Ik ⊗ Ik in
Mn( ^Mk(A)) ⊗ Mk(C) ∼= Mn(C) ⊗ ^Mk(A) ⊗ Mk(C).
Then for any t ∈ [0, 1],
A → A; (x, z ⊗ Ik) 7→ ((In ⊗ Ut) · x · (In ⊗ Ut)−1, z ⊗ Ik)
14
H. OYONO-OYONO AND G. YU
is an automorphism of C∗-algebra. Hence,
(cid:0)(In ⊗ Ut) · (y ⊗ e1,1) · (In ⊗ U−1
nk ( ^Mk(A)) between (y ⊗ e1,1, λ ⊗ Ik) and (σ(y ⊗ e1,1), λ ⊗ Ik). The
is a path in Uε,r
range of σ(y ⊗ e1,1) being in the range of the projection In ⊗ e1,1 ⊗ Ik, we have an
orthogonal sum decomposition
), λ ⊗ Ik(cid:1)t∈[0,1]
t
(σ(y ⊗ e1,1), λ ⊗ Ik) = (σ(y ⊗ e1,1), λ ⊗ e1,1) + (0, λ ⊗ (Ik − e1,1))
(recall that λ ⊗ e1,1 and λ ⊗ (Ik − e1,1) multiply Mn(A) ⊗ Mk(C) ⊗ Mk(C) on
the first two factors). By lemma 1.21, λ is homotopic to In in Uε,r
n (C) and thus
(σ(y ⊗ e1,1), λ ⊗ Ik) is homotopic to (σ(y ⊗ e1,1), λ ⊗ e1,1) + (0, In ⊗ (Ik − e1,1)) in
nk ( ^Mk(A))) which can be viewed as
Uε,r
diag((y, λ), (0, Ik(k−1))
in Mk(Mn( ^Mk(A)). From this we deduce that [(y, λ)]ε,r = [(y ⊗ e1,1, λ ⊗ Ik)]ε,r in
1 ( ^Mk(A)).
K ε,r
For the even case, by an analogous computation, we can check that the group
homomorphisms
K ε,r
0 ( A) → K ε,r
0 ( ^Mk(A)); [(p, q), l)]ε,r 7→ [(p ⊗ e1,1), q, l]ε,r
and
0 ( A); [(p, q), l)]ε,r 7→ [(p, q ⊗ Ik), kl]ε,r,
0 ( ^Mk(A)) → K ε,r
K ε,r
0 (Mk(A)) → K ε,r
respectively induce by restriction homomorphisms Φ0 : K ε,r
0 (Mk(A))
and Ψ0 : K ε,r
0 (A) which are inverse of each other, where in the
right hand side of the last formula, we have viewed p ∈ Mn(Mk(A)) as a matrix
in Mnk(A) and q ⊗ Ik ∈ Mn(C) ⊗ Mk(C) as a matrix in Mnk(C). Since Φ0 is
induced by ιA, we get from lemma 1.23 that ιε,r
0 (Mk(A)) is an
A,∗
isomorphism.
0 (A) → K ε,r
0 (A) → K ε,r
: K ε,r
Let A be a C∗-algebra filtered by (Ar)r>0. Then K(H)⊗ A is filtered by (K(H)⊗
Ar)r>0 and applying proposition 1.28 to the increasing family (Mk(A)+)k∈N of C∗-
subalgebras of ^K(H) ⊗ A, lemmas 1.23 and 1.24, and the discussion above, we
deduce the Morita equivalence for K ε,r
∗ (•).
Proposition 1.29. If A is a filtered algebra and H is a separable Hilbert space,
then the homomorphism
A → K(H) ⊗ A; a 7→
a
0
. . .
induces a (Z2-graded) semi-group isomorphism (the Morita equivalence)
∗ (A) → K ε,r
for any positive number r and any ε ∈ (0, 1/4).
A : K ε,r
Mε,r
∗ (K(H) ⊗ A)
ON A QUANTITATIVE OPERATOR K-THEORY
15
1.6. Lipschitz homotopies.
Definition 1.30. If A is a C∗-algebra and C a positive integer, then a map h =
[0, 1] → A is called C-Lipschitz if for every t and s in [0, 1], then kh(t) − h(s)k 6
Ct − s.
Proposition 1.31. There exists a number C such that for any unital filtered C∗-
algebra A and any positive number r and ε < 1/4 then :
(i) if p0 and p1 are homotopic in Pε,r
l and a C-Lipschitz homotopy in Pε,r
diag(p1, Ik, 0l).
(ii) if u0 and u1 are homotopic in Uε,r
n (A), then there exist integers k and
n+k+l(A) between diag(p0, Ik, 0l) and
C-Lipschitz homotopy in U3ε,2r
n (A) then there exist an integer k and a
n+k (A) between diag(u0, Ik) and diag(u1, Ik).
Proof.
(i) Notice first that if p is an ε-r-projection in A, then the homotopy of ε-r-
projections of M2(A) between (cid:18)1 0
Let (pt)t∈[0,1] be a homotopy between p0 and p1 in Pε,r
0 0(cid:19) and (cid:18)p
0 1 − p(cid:19) in example 1.6 is
2-Lipschitz.
0
ε−kp2
t−ptk
4
n (A). Set α =
inf t∈[0,1]
ant let t0 = 0 < t1 < . . . < tk = 1 be a partition of [0, 1]
such that kpti − pti−1k < α for i ∈ {1, . . . , k}. We construct a homotopy
of ε-r-projections with the required property between diag(p0, In(k−1), 0)
and diag(p1, In(k−1), 0) in Mn(2k−1)(A) as the composition of the following
homotopies.
• We can connect diag(pt0 , In(p−1), 0) and diag(pt0 , In, 0, . . . , In, 0) within
Pε,r
n(2k−1)(A) by a 2-Lipschitz homotopy.
• As we noticed at the beginning of the proof, we can connect
diag(pt0 , In, 0, . . . , In, 0) and diag(pt0 , In − pt1 , pt1, . . . , In − ptk , ptk )
within Pε,r
• The ε-r-projections diag(pt0 , In − pt1, pt1, . . . , In − ptk , ptk ) and
diag(pt0 , In − pt0, . . . , ptk−1, In − ptk−1 , ptk ) satisfy the norm estimate
of the assumption of lemma 1.7(i) and hence then can be connected
within Pε,r
n(2k−1)(A) by a ray which is clearly a 1-Lipschitz homotopy.
• Using once again the homotopy of example 1.6, we see that diag(pt0 , In−
pt0, . . . , ptk−1, In − ptk−1, ptk ) and diag(0, In, . . . , 0, In, ptk ) are con-
nected within Pε,r
• Eventually, diag(0, In, . . . , 0, In, ptk ) and diag(ptk , In(k−1), 0) are con-
n(2k−1)(A) by a 2-Lipschitz homotopy.
n(2k−1)(A) by a 2-Lipschitz homotopy.
n(2k−1)(A) by a 2-Lipschitz homotopy.
(ii) Let (ut)t∈[0,1] be a homotopy between u0 and u1 in Uε,r
n (A). Set α =
and let t0 = 0 < t1 < . . . < tp = 1 be a partition of
inf t∈[0,1]
[0, 1] such that kuti − uti−1k < α for i ∈ {1, . . . , p}. We construct a homo-
topy with the required property between diag(u0, I2np) and diag(u1, I2np)
within U3ε,2r
n(2p+1)(A) as the composition of the following homotopies.
nected within Pε,r
t ut−Ink
3
ε−ku∗
• Since Inp and diag(u∗t1ut1, . . . , u∗tputp ) satisfy the norm estimate of
the assumption of lemma 1.7(ii), then diag(ut0, Inp) is a 3ε-2r-unitary
that can be connected to diag(ut0, u∗t1ut1, . . . , u∗tp utp) in U3ε,2r
n(p+1)(A)
by a 1-Lipschitz homotopy.
16
H. OYONO-OYONO AND G. YU
• Proceeding as in the first point of
lemma 1.8, we see that
diag(In, u∗t1, . . . , u∗tp, Inp) and diag(u∗t1 , . . . , u∗tp , In(p+1)) can be con-
nected within Uε,r
n(2p+1)(A) by a 2-Lipschitz homotopy and thus, in
view of remark 1.4,
diag(ut0 , u∗t1ut1, . . . , u∗tp utp , Inp) =
diag(In, u∗t1, . . . , u∗tp, Inp) · diag(ut0, ut1, . . . , utp, Inp)
and
diag(u∗t1, . . . , u∗tp, In(p+1)) · diag(ut0 , ut1, . . . , utp, Inp) =
diag(u∗t1 ut0, . . . , u∗tp utp−1, utp, Inp)
can be connected within U3ε,2r
n(2p+1)(A) by a 4-Lipschitz homotopy.
• Since ku∗tiuti−1 − Ink < ε, we get by using once again lemma 1.7(ii)
that diag(u∗t1 ut0, . . . , u∗tp utp−1, utp, Inp) and diag(Inp, utp, Inp) can be
connected within U3ε,2r
n(2p+1)(A) by a 1-Lipschitz homotopy.
• Eventually, diag(Inp, utp , Inp) can be connected to diag(utp , I2np) within
U3ε,2r
(2p+1)n(A) by a 2-Lipschitz homotopy.
(cid:3)
Corollary 1.32. There exists a control pair (αh, kh) such that the following holds:
For any unital filtered C∗-algebra A, any positive numbers ε and r with ε < 1
4αh
n (A), then there is for some
and any homotopic ε-r-projections q0 and q1 in Pε,r
integers k and l an αhε-kh,εr-unitary W in Uαhε,kh,εr
(A) such that
n+k+l
k diag(q0, Ik, 0l) − W diag(q1, Ik, 0l)W ∗k < αhε.
Proof. According to proposition 1.31, we can assume that q0 and q1 are connected
by a C-Lipschitz homotopy (qt)t∈[0,1], for some universal constant C. Let t0 = 0 <
t1 < ··· < tp = 1 be a partition of [0, 1] such that 1/32C < ti − ti−1 < 1/16C.
With notation of lemma 1.10, pick for every integer i in {1, . . . , p} a λε-lε-unitary
Wi in A such that kWiqti−1 W ∗i − qtik < λε. If we set W = Wp ··· W1, then W is
a 3pλε-plεr-unitary such that kW q0W ∗ − q1k < 2pλε. Since p < 2C, we get the
result.
(cid:3)
2. Controlled morphisms
As we shall see in section 3, usual maps in K-theory such as boundary maps
factorize through semi-group homomorphism of quantitative K-theory groups with
expansion of norm control and propagation controlled by a control pair. This
motivates the notion of controlled morphisms for quantitative K-theory in this
section.
Recall that a controlled pair is a pair (λ, h), where
• λ > 1;
• h : (0, 1
4λ ) → (0, +∞); ε 7→ hε is a map such that there exists a non-
increasing map g : (0, 1
4λ ) → (0, +∞), with h 6 g.
The set of control pairs is equipped with a partial order: (λ, h) 6 (λ′, h′) if λ 6 λ′
and hε 6 h′ε for all ε ∈ (0, 1
4λ′ )
ON A QUANTITATIVE OPERATOR K-THEORY
17
2.1. Definition and main properties. For any filtered C∗-algebra A, let us de-
fine the families K0(A) = (K ε,r
1 (A))0<ε<1/4,r>0 and
K∗(A) = (K ε,r
∗ (A))0<ε<1/4,r>0.
Definition 2.1. Let (λ, h) be a controlled pair, let A and B be filtered C∗-algebras,
and let i, j be elements of {0, 1,∗}. A (λ, h)-controlled morphism
0 (A))0<ε<1/4,r>0, K1(A) = (K ε,r
is a family F = (F ε,r)0<ε< 1
F : Ki(A) → Kj(B)
4λ ,r>0 of semigroups homomorphisms
F ε,r : K ε,r
(B)
(A) → K λε,hεr
j
i
such that for any positive numbers ε, ε′, r and r′ with 0 < ε 6 ε′ < 1
hεr 6 hε′ r′, we have
4λ and
F ε′,r′
◦ ιε,ε′,r,r′
i
= ιλε,λε′,hεr,hε′ r′
j
◦ F ε,r.
controlled morphism.
If it is not necessary to specify the control pair, we will just say that F is a
Let A and B be filtered algebras. Then it is straightforward to check that if
F : Ki(A) → Kj (B) is a (λ, h)-controlled morphism, then there is group homo-
morphism F : Ki(A) → Kj(B) uniquely defined by F ◦ ιε,r
◦ F ε,r. The
homomorphism F will be called the (λ, h)-controlled homomorphism induced by
F . A homomorphism F : Ki(A) → Kj(B) is called (λ, h)-controlled if it is induced
by a (λ, h)-controlled morphism. If we don't need to specify the control pair (λ, h),
we will just say that F is a controlled homomorphism.
i = ιλε,hεr
j
Example 2.2.
(i) Let A = (Ar)r>0 and B = (Br)r>0 be two filtered C∗-algebras and let
f : A → B be a homomorphism. Assume that there exists d > 0 such
that f (Ar) ⊂ Bdr for all positive r. Then f gives rise to a bunch of semi-
group homomorphisms (cid:16)f ε,r
and hence
to a (1, d)-controlled morphism f∗ : K∗(A) → K∗(B).
(B)(cid:17)0<ε< 1
∗ (A) → K ε,dr
: K ε,r
4 ,r>0
∗
∗
(ii) The bunch of semi-group isomorphisms
4 ,r>0
of proposition 1.29 defines a (1, 1)-controlled morphism
∗ (K(H)⊗A))0<ε< 1
∗ (A) → K ε,r
A : K ε,r
(Mε,r
and
MA : K∗(A) → K∗(K(H)⊗A)
M−1
A : K∗(K(H)⊗A) → K∗(A)
inducing the Morita equivalence in K-theory.
If (λ, h) and (λ′, h′) are two control pairs, define
h ∗ h′ : (0,
1
4λλ′
) → (0, +∞); ε 7→ hλ′εh′ε.
Then (λλ′, h ∗ h′) is a control pair. Let A, B1 and B2 be filtered C∗-algebras, let
i, j and l be in {0, 1,∗} and let F = (F ε,r)0<ε< 1
,r>0 : Ki(A) → Kj (B1) be a
(αF , kF )-controlled morphism, let G = (Gε,r)0<ε< 1
,r>0 : Kj (B1) → Kl(B2) be a
(αG, kG)-controlled morphism. Then G◦F : Ki(A) → Kl(B2) is the (αGαF , kG∗kF )-
controlled morphism defined by the family (GαF ε,kF,εr ◦ F ε,r)0<ε< 1
,r>0.
4αF αG
4αF
4αG
18
H. OYONO-OYONO AND G. YU
Remark 2.3. The Morita equivalence for quantitative K-theory is natural, i.e
MB ◦ f = (IdK(H))⊗f ) ◦ MA
for any homomorphism f : A → B of filtered C∗-algebras.
Notation 2.4. Let A and B be filtered C∗-algebras, let (λ, h) be a control pair,
and let F = (F ε,r)0<ε< 1
,r>0) be
a (αF , kF )-controlled morphism (resp. a (αG, kG)-controlled morphism). Then we
write F
,r>0 : Ki(A) → Kj(B) (resp. G = (Gε,r)0<ε< 1
∼ G if
(λ,h)
4αF
4αG
• (αF , kF ) 6 (λ, h) and (αG, kG) 6 (λ, h).
• for every ε in (0, 1
4λ ) and r > 0, then
ιαF ε,λε,kF,εr,hεr
j
◦ F ε,r = ιαG ε,λε,kG,εr,hεr
j
◦ Gε,r.
(λ,h)
If F and G are controlled morphisms such that F
∼ G for a control pair (λ, h),
then F and G induce the same morphism in K-theory.
Remark 2.5. Let F : Ki(A2) → Kj(B1) (resp. F′ : Ki(A2) → Kj(B1)) be a
(αF , kF )-controlled (resp. a (αF ′, kF ′)-controlled) morphisms and let G : Ki′ (A1) →
Ki(A2) (resp. G′ : Kj(B1) → Kl(B2)) be a (αG, kG)-controlled (resp. a (αG′ , kG′ )-
controlled) morphism. Assume that F
∼ F′ for a control pair (λ, h), then
(λ,h)
• G′ ◦ F
• F ◦ G
(αG′ λ,kG′∗h)
(αG λ,h∗kG )
∼
∼
G′ ◦ F′;
F′ ◦ G.
the controlled morphism induced by IdA.
If i is an element in {0, 1,∗} and A a filtered C∗-algebra, we denote by IdKi(A)
Let F : Ki(A1) → Ki′ (B1), F′ : Kj(A2) → Kl(B2), G : Ki(A1) → Kj (A2) and
G′ : Ki′ (B1) → Kl(B2) be controlled morphisms and let (λ, h) be a control pair.
Then the diagram
Ki′ (B1)
−−−−→ Kl(B2)
G′
Ki(A1)
G−−−−→ Kj(A2)
xF ′
Fx
is called (λ, h)-commutative (or (λ, h)-commutes) if G′ ◦ F
Definition 2.6. Let (λ, h) be a control pair, and let F : Ki(A) → Kj(B) be a
(αF , kF )-controlled morphism with (αF , kF ) 6 (λ, h).
∼ F′ ◦ G.
• F is called left (λ, h)-invertible if there exists a controlled morphism
(λ,h)
G : Kj (B) → Ki(A)
(λ,h)
such that G ◦ F
a left (λ, h)-inverse for F . Notice that definition of
(αF αG, kF ∗ kG) 6 (λ, h).
∼ IdKi(A). The controlled morphism G is then called
∼ implies that
• F is called right (λ, h)-invertible if there exists a controlled morphism
(λ,h)
G : Kj (B) → Ki(A)
(λ,h)
such that F ◦ G
right (λ, h)-inverse for F .
∼ IdKi(B). The controlled morphism G is then called a
ON A QUANTITATIVE OPERATOR K-THEORY
19
• F is called (λ, h)-invertible or a (λ, h)-isomorphism if there exists a con-
trolled morphism
G : Kj (B) → Ki(A)
which is a left (λ, h)-inverse and a right (λ, h)-inverse for F . The con-
trolled morphism G is then called a (λ, h)-inverse for F (notice that we
have in this case necessarily (αG, kG) 6 (λ, h)).
We can check easily that indeed, if F is left (λ, h)-invertible and right (λ, h)-
invertible, then there exists a control pair (λ′, h′) with (λ, h) 6 (λ′, h′), depending
only on (λ, h) such that F is (λ′, h′)-invertible.
Definition 2.7. Let (λ, h) be a control pair and let F : Ki(A) → Kj(B) be a
(αF , kF )-controlled morphism.
• F is called (λ, h)-injective if (αF , kF ) 6 (λ, h) and for any 0 < ε < 1
(A), then F ε,r(x) = 0 in K αF ε,kF,εr
4λ , any
(B) implies
r > 0 and any x in K ε,r
that ιε,λε,r,hεr
(x) = 0 in K λε,hεr
i
i
i
(A);
• F is called (λ, h)-surjective, if for any 0 < ε < 1
4λαF
, any r > 0 and
(B), there exists an element x in K λε,hεr
(A) such that
j
i
any y in K ε,r
F λε,hλεr(x) = ιε,αF λε,r,kF,λεhεr
j
j
(y) in K αF λε,kF,λεhεr
j
(B).
Remark 2.8.
(i) If F : K1(A) → Ki(B) is a (λ, h)-injective controlled morphism. Then
according to lemma 1.16, there exists a control pair (λ′, h′) with (λ, h) 6
(λ′, h′) depending only on (λ, h) such that for any 0 < ε < 1
4λ′ , any r > 0
and any x and x′ in K ε,r
(B)
implies that ιε,λ′ε,r,h′
εr
1 (A) , then F ε,r(x) = F ε,r(x′) in K αF ε,kF,εr
(x) = ιε,λ′ε,r,h′
εr
(x′) in K λ′ε,h′
εr
(A);
i
1
1
1
(ii) It is straightforward to check that if F is left (λ, h)-invertible, then F is
(λ, h)-injective and that if F is right (λ, h)-invertible, then there exists a
control pair (λ′, h′) with (λ, h) 6 (λ′, h′), depending only on (λ, h) such
that F is (λ′, h′)-surjective.
(iii) On the other hand, if F is (λ, h)-injective and (λ, h)-surjective, then there
exists a control pair (λ′, h′) with (λ, h) 6 (λ′, h′), depending only on (λ, h)
such that F is a (λ′, h′)-isomorphism.
2.2. Controlled exact sequences.
Definition 2.9. Let (λ, h) be a control pair,
4αG
4αF
• Let F = (F ε,r)0<ε< 1
,r>0 : Ki(A) → Kj (B1) be a (αF , kF )-controlled
morphism, and let G = (Gε,r)0<ε< 1
,r>0 : Kj (B1) → Kl(B2) be a (αG, kG)-
controlled morphism, where i, j and l are in {0, 1,∗} and A, B1 and B2
are filtered C∗-algebras. Then the composition
Ki(A) F→ Kj(B1) G→ Kl(B2)
, any r > 0 and any y in K ε,r
is said to be (λ, h)-exact at Kj(B1) if G ◦ F = 0 and if for any 0 < ε <
(B1) such that Gε,r(y) = 0 in
4 max{λαF ,αG}
K αG ε,kG,εr
(B2), there exists an element x in K λε,hεr
(A) such that
1
j
j
i
F λε,hλεr(x) = ιε,αF λε,r,kF,λεhεr
j
(y)
20
H. OYONO-OYONO AND G. YU
in K αF λε,kF,λεhεr
j
(B1).
• A sequence of controlled morphisms
···Kik−1 (Ak−1) Fk−1→ Kik (Ak) Fk→ Kik+1 (Ak+1) Fk+1→ Kik+2 (Ak+2)···
is called (λ, h)-exact if for every k, the composition
Kik−1 (Ak−1) Fk−1→ Kik (Ak) Fk→ Kik+1 (Ak+1)
is (λ, h)-exact at Kik (Ak).
Remark 2.10. If the composition Ki(A) F→ K1(B1) G→ Kl(B2) is (λ, h)-exact, then
according to lemma 1.16, there exists a control pair (λ′, h′) with (λ, h) 6 (λ′, h′)
depending only on (λ, h), such that for any 0 < ε <
, any r > 0 and
any y and y′ in K ε,r
there exists an element x in K λ′ε,h′
εr
4 max{λ′αF ,αG}
1 (B1) , then Gε,r(y) = Gε,r(y′) in K αF ε,kF,εr
(B) implies that
(A) such that
1
j
i
ε,αF λ′ε,r,kF,λ′ εh′
ι
j
εr
ε,αF λ′ε,r,kF,λ′ εh′
(y′) = ι
j
εr
(y) + F λ′ε,h′
εr(x)
in K αF λ′ε,kF,λεh′
εr
1
(B1).
3. Extensions of filtered C∗-algebras
The aim of this section is to establish a controlled exact sequence for quantitative
K-theory with respect to extension of filtered C∗-algebras admitting a completely
positive cross section that preserves the filtration. We also prove that for these
extensions, the boundary maps are induced by controlled morphisms. As in K-
theory, one is a map of exponential type and the other is an index type map, and
the later in turn fits in a long (λ, h)-controlled exact sequence for some universal
control pair (λ, h).
3.1. Semi-split filtered extensions. Let A be a C∗-algebra filtered by (Ar)r>0
and let J be an ideal of A. Then A/J is filtered by ((A/J)r)r>0, where (A/J)r
is the image of Ar under the projection A → A/J. Assume that the C∗-algebra
extension
0 → J → A → A/J
q
→ 0
admits a contractive filtered cross-section s : A/J → A, i.e such that s((A/J)r)) ⊂
Ar for any positive number. For any x ∈ J and any number ε > 0 there exists
a positive number r and an element a of Ar such that kx − ak < ε. Let us set
y = a − s ◦ q(a). Then y belongs to Ar ∩ J and moreover
ky − xk = ka − x + s ◦ q(x − a)k
6 ka − xk + ks ◦ q(a − x)k
6 ka − xk + kq(x − a)k
6 2ε.
Hence, [r>0
(Ar ∩ J) is dense in J and therefore J is filtered by (Ar ∩ J)r>0.
Definition 3.1. Let A be a C∗-algebra filtered by (Ar)r>0 and let J be an ideal of
A. The extension of C∗-algebras
0 → J → A → A/J → 0
ON A QUANTITATIVE OPERATOR K-THEORY
21
is said to be filtered and semi-split (or a semi-split extension of filtered C∗-algebras)
if there exists a completely positive cross-section
such that
s : A/J → A
s((A/J)r)) ⊂ Ar
for any number r > 0. Such a cross-section is said to be semi-split and filtered.
We have the following analogous of the lifting property for unitaries of the neutral
component.
Lemma 3.2. There exists a control pair (αe, ke) such that for any semi-split ex-
tension of filtered C∗-algebras
0 −→ J −→ A
q
−→ A/J −→ 0
n+j
with A unital, the following holds: for every positive numbers r and ε with ε < 1
4αe
and any ε-r-unitary V homotopic to In in Uε,r
n (A/J), then for some integer j,
there exists a αeε-ke,εr-unitary W homotopic to In+j in Uαeε,ke,εr
(A) and such
that kq(W ) − diag(V, Ij )k < αeε.
Proof. According to proposition 1.31, we can assume that V and In are connected
by a C-Lipschitz homotopy (Vt)t∈[0,1], for some universal constant C. Let t0 = 0 <
t1 < ··· < tp = 1 be a partition of [0, 1] such that 1/16C < ti − ti−1 < 1/8C.
Then we get that kVi−1 − Vik < 1/8 and hence kVi−1V ∗i − Ink < 1/2. Let lε
be the smallest integer such that Pk>lε+1 2−k/k < ε and Pk>lε+1 logk 2/k! < ε
and let us consider the polynomial functions Pε(x) = Plε
k=0 xk/k! and Qε(x) =
−Plε
k=1 xk/k. We get then kVi−1V ∗i − Pε ◦ Qε(1 − Vi−1V ∗i )k 6 3ε. Choose a
completely positive section s(A/J) → A such that s(1) = 1 and let us set W t
i =
Pε(s(tQε(In − Vi−1V ∗i ))) for t in [0, 1] and i in {1, . . . , p}. Since Vi−1V ∗i
is closed
to the unitary Vi−1V ∗i (ViV ∗i−1Vi−1V ∗i )−1/2, then W t
i is uniformly (in t and i) closed
to exp ts(log(Vi−1V ∗i (ViV ∗i−1Vi−1V ∗i )−1/2)) which is unitary (the logarithm is well
defined since Vi−1V ∗i (ViV ∗i−1Vi−1V ∗i )−1/2 is closed to In) and hence W t
i is a αε-2lεr-
unitary for some universal α. Hence W 1
n (A) homotopic
i
i ) − Vi−1V ∗i k < 3αε. If we set now W = W 1
to In and such that kq(W 1
p and
since p 6 16C, then W satisfies the required property.
(cid:3)
is a αε-2lεr-unitary in Uε,r
1 ··· W 1
Lemma 3.3. There exists a control pair (α, k) such that for any semi-split exten-
sion of filtered C∗-algebras 0 → J → A → A/J → 0 with A unital, any semi-split
filtered cross section s : A/J → A with s(1) = 1 and any ε-r-projection p in A/J
with 0 < ε < 1
4λ , there exists an element yp in Jkεr such that k1+yp−e2ıπs(k0(p))k <
αε/3. In particular 1 + yp is a αε-kεr-unitary of J +;
Proof. Let lε be the smallest integer such that
+∞Xl=lε+1
10l/l! < ε.
22
H. OYONO-OYONO AND G. YU
(2ıπs(p))l
l!
Let us define zp =
lεXl=0
(cid:13)(cid:13)(cid:13)zp − e2ıπs(κ0(p))(cid:13)(cid:13)(cid:13) = (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
lεXl=0
6 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
lεXl=0
6 ks(p) − s(κ0(p))ke10 + ε
6 (2e10 + 1)ε.
. Then zp belongs to Mn(Alεr) and we have
(2ıπs(p))l
l!
+∞Xl=0
−
(2ıπs(κ0(p))l
l!
(2ıπs(p))l − (2ıπs(κ0(p)l)
l!
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
+(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
+∞Xl=lε+1
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
(2ıπs(κ0(p))l
l!
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
If we set yp = zp − s ◦ q(zp), then yp ∈ Mn(J ∩ Alεr) and
kzp − (1 + yp)k = ks ◦ q(zp) − 1k
6 (cid:13)(cid:13)(cid:13)q(cid:16)zp − e2ıπs(κ0(p))(cid:17)(cid:13)(cid:13)(cid:13)
< λε,
with λ = (2e10 + 1). Therefore we have k1 + yp − e2ıπs(κ0(p))k < 2λε. The end of
the statement is then a consequence of lemma 1.7.
(cid:3)
3.2. Controlled boundary maps. For any extension 0 → J → A → A/J → 0
of C∗-algebras we denote by ∂J,A : K∗(A/J) → K∗(J) the associated (odd degree)
boundary map.
Proposition 3.4. There exists a control pair (αD, kD) such that for any semi-split
extension of filtered C∗-algebras
−→ A/J −→ 0,
there exists a (αD, kD)-controlled morphism of odd degree
0 −→ J −→ A
q
DJ,A = (∂ε,r
J,A)0<ε 1
4αD
,r : K∗(A/J) → K∗(J)
which induces in K-theory ∂J,A : K∗(A/J) → K∗(J).
Proof. Let s : A/J → A be a semi-split filtered cross-section. Let us first prove the
result when A is unital.
(i) Let p be an element of Pε,r
n (A/J). Then ∂J,A([κ0(p)]) is the class of
e2ıπs(κ0(p)) in K1(J). Fix a control pair (α, k) as in lemma 3.3 and pick
any yp in Mn(Jkεr) such that k1 + yp− e2ıπs(k0(p))k < αε/3. Then 1 + yp is
an αε-kεr-unitary of Mn(J +), and according to lemma 1.7, any two such
αε-kεr-unitaries are homotopic in U 3αε,kεr
(J +). Applying lemma 3.3 to
A/J[0, 1], we see that the map
n
Pε,r
n (A/J) −→ U3αε,kεr
n
(J +); p 7→ 1 + yp
preserves homotopies and hence gives rise to a bunch of well defined semi-
group homomorphism
∂ε,r
J,A : K ε,r
0 (A/J) −→ K 3αε,kεr
1
(J); [p, l]ε,r 7→ [1 + yp]3αε,kεr
which in the even case satisfies the required properties for a controlled
homomorphism.
ON A QUANTITATIVE OPERATOR K-THEORY
23
n (A/J), pick any element v in some Uε,r
(ii) In the odd case, we follow the route of [18, Chapter 8]. For any element u
of Uε,r
j (A/J) such that diag(u, v)
is homotopic to In+j in U3ε,2r
n+j (A/J) (we can choose in view of lemma 1.16
v = u∗). According to lemma 3.2, and up to replace v by diag(v, Ik) for
some integer k, there exists an element w in U3αeε,2ke,3εr
(A) such that
kq(w) − diag(u, v)k 6 3αeε. Let us set x = w diag(In, 0)w∗. Then x is an
element in P6αeε,4ke,3εr
Let us set now h = x − diag(In, 0) − s ◦ q(x − diag(In, 0)). Then h is a
(A) such that kq(x) − diag(In, 0)k 6 9αeε.
n+j
n+j
self-adjoint element of M2n(A4ke,3εr ∩ J) such that
kx − diag(In, 0) − hk 6 9αeε,
and therefore h + diag(In, 0) belongs to P45αeε,4ke,3εr
n+j
(J). Define then
∂ε,r
J,A([u]ε,r) = [h + diag(In, 0), n]450αeε,4ke,3εr .
It is straightforward to check that (compare with [18, Chapter 8]).
• two choice of elements satisfying the conclusion of lemma 3.2 relatively
(J) (this
to diag(u, v) give rise to homotopic elements P450αeε,4ke,3εr
is a consequence of lemma 1.7).
n+j
• Replacing u by diag(u, Im) and v by diag(v, Ik) gives also rise to the
same element of K 450αeε,4ke,3εr
(J).
0
Applying now lemma 3.2 to the exact sequence
0 → J[0, 1] → A[0, 1] → A/J[0, 1] → 0,
J,A([u]ε,r)
we get that ∂ε,r
• only depends on the class of u in K ε,r
• does not depend on the choice of v such that diag(u, v) is connected
1 (A/J);
to In+j in Uε,r
n+j(A/J).
• If A is not unital, use the exact sequence
0 → J → eA → gA/J → 0
J,A as the composition
to define ∂ε,r
0 (A/J) ֒→ K ε,r
K ε,r
0 (gA/J)
∂ε,r
J, eA−→ K 450αeε,4ks,3εr
1
(J),
where the inclusion in the composition is induced by the inclusion A/J ֒→
gA/J ∼= eA/J.
• Since the set of filtered semi-split cross-section s : A/J → A such that
s((A/J)r) ⊂ Ar is convex, the definition of ∂ε,r
J,A actually does not depend
on the choice of such a section.
• Using lemma 1.7,
(αD, kD), then DJ,A = (∂ε,r
inducing the (odd degree) boundary map ∂J,A : K∗(A/J) → K∗(J).
it is plain to check that for a suitable control pair
,r is a (αD, kD)-controlled morphism
J,A)0<ε 1
4αD
(cid:3)
For a semi-split extension of filtered C∗-algebras
0 −→ J −→ A
q
−→ A/J −→ 0,
24
H. OYONO-OYONO AND G. YU
J,A : K0(A/J) → K1(J), for the restriction of DJ,A to K0(A/J) and
we set D0
D1
J,A : K1(A/J) → K0(J), for the restriction of DJ,A to K1(A/J).
Remark 3.5.
(i) Let A and B be two filtered C∗-algebras and let φ : A → B be a filtered
homomorphism. Let I and J be respectively ideals in A and B and assume
that
• φ(I) ⊂ J;
• there exists semi-split filtered cross-sections s : A/I → A and s′ :
B/J → J such that s′ ◦ φ = φ ◦ s, where φ : A/I → B/J is the
homomorphism induced by φ,
(ii) Let 0 −→ J −→ A
then DJ,B ◦ φ∗ = φ∗ ◦ DI,A.
−→ A/J −→ 0 be a split extension of filtered C∗-
algebras, i.e there exists a homomorphism of filtered C∗-algebras s : A/J →
A such that q ◦ s = IdA/J . Then we have DJ,A = 0.
q
For a filtered C∗-algebra A, we have defined the suspension and the cone re-
spectively as SA = C0((0, 1), A) and CA = C0((0, 1], A). Then SA and CA are
filtered C∗-algebras and evaluation at the value 1 gives rise to a semi-split filtered
extension of C∗-algebras
(1)
and in the even case, the corresponding boundary ∂SA,CA : K0(A) → K1(SA)
implements the suspension isomorphism and has the following easy description
when A is unital: if p is a projection, then ∂SA,CA[p] is the class in K1(SA) of the
path of unitaries
0 → SA → CA → A → 0
[0, 1] → Un(A); t 7→ pe2ıπt + 1 − p.
Let us show that we have an analogous description in term of almost projection.
Notice that if q is an ε-r-projection in A, then
zq : [0, 1] → A; t 7→ qe2ıπt + 1 − q
is a 5ε-r-unitary in fSA. Using this, we can define a (5, 1)-controlled morphism
ZA = (Z ε,r
A )0<ε<1/20,r>0 : K0(A) → K1(SA) in the following way:
n (A) and any integer k let us set
• for any q in Pε,r
Vq,k : [0, 1] → U5ε,r
• define then Z ε,r
n (fSA) : t 7→ diag(e−2kıπt, 1, . . . , 1) · (1 − q + qe2ıπt);
A ([q, k]ε,r) = [Vq,k]5ε,r.
(λ,h)
CA,SA
∼ ZA.
Proposition 3.6. There exists a control pair (λ, h) such that for any filtered C∗-
algebra A, then D0
Proof. Let [q, k]ε,r be an element of K ε,r
n (A) and k integer. We
can assume without loss of generality that n > k. Namely, up to replace n by 2n
and using a homotopy between diag(q, 0) and diag(0, q) in Pε,r
2n (A), we can indeed
assume that q and diag(Ik, 0) commute. As in the proof of proposition 3.4, define lε
as the smallest integer such that P∞l=lε+1 10l/l! < ε. Let us consider the following
0 (A), with q in Pε;r
paths in Mn(A)
z : [0, 1]−→Mn(A); t 7→
lεXl=0
(2ıπ(tq + (1 − t) diag(Ik, 0)))l/l!
ON A QUANTITATIVE OPERATOR K-THEORY
25
and
z′ : [0, 1]−→Mn(A); t 7→ exp(2ıπ diag(−tIk, 0))(1 − q + e2ıπtq).
Since q and Ik commutes, then
exp(2ıπ(diag(−tIk, 0) + tq)) = exp(2ıπ diag(−tIk, 0)) · exp(2ıπtq)
and hence
z(t) = exp(2ıπ diag(−tIk, 0)) exp(2ıπtq) −
∞Xl=lε+1
(2ıπ(tq + (1 − t) diag(Ik, 0)))l/l!.
We get therefore
kz(t) − z′(t)k 6 ε + kqe2ıπt + (1 − q) − exp 2ıπtqk
6 ε + 2kκ0(q) − qk + k exp 2ıπtκ0(q) − exp 2ıπtqk
6 ε(5 + 4e4π).
Let us set
y : [0, 1] : −→Mn(A); t 7→ z(t) − 1 − (1 − t) diag(Ik, 0)
For some αs > α∂ , we get then that 1 + y and z′ are homotopic elements in
U αsε,k∂,εr
for the extension of equation (1), we get in view of the proof of proposition 3.4,
(fSA). Using the semi-split filtered cross-section A → CA; a 7→ [t 7→ ta]
(2ıπ)l/l! − t
(2ıπq)l/l!.
n
lεXl=1
lεXl=1
ια∂ ε,αsε,k∂,εr
1
◦ ∂ε,r
SA,CA([q, k]ε,r) = [1 + y]αsε,k∂,εr,
and thus we deduce
ια∂ ε,αsε,k∂,εr
1
◦ ∂ε,r
SA,CA([q, k]ε,r) = [z′]αsε,k∂,εr.
We get the result by using a homotopy of unitaries in Mn(fSA) between
and t 7→ exp(2ıπ diag(−tIk, In−k)).
t 7→ diag(e−2kπt, 1, . . . , 1)
(cid:3)
The inverse of the suspension isomorphism is provided, up to Morita equivalence
let us consider the unilateral shift S on ℓ2(N), i.e the
by the Toeplitz extension:
operator defined on the canonical basis (en)n∈N of ℓ2(N) by S(en) = en+1 for all
integer n. Then the Toeplitz algebra T is the C∗-subalgebra of L(ℓ2(N)) generated
by S. The algebra of compact operators K(ℓ2(N)) is an ideal of T and we get an
extension of C∗-algebras
0 → K(ℓ2(N)) → T
ρ
→ C(S1) → 0,
called the Toeplitz extension, where S1 denote the unit circle. Let us define T0 =
ρ−1(C0((0, 1)), where C0(0, 1) is viewed as a subalgebra of C(S1). We obtain then
an extension of C∗-algebras
For any C∗-algebra A, we can tensorize this exact sequence to obtain an extension
0 → K(ℓ2(N)) → T0
→ C0(0, 1) → 0.
ρ
0 → K(ℓ2(N)) ⊗ A → T0 ⊗ A→SA → 0
26
H. OYONO-OYONO AND G. YU
which is filtered and semi-split when A is a filtered C∗-algebra.
Proposition 3.7. There exists a control pair (λ, h) such that
for any unital filtered C∗-algebra A.
D1
K(ℓ2(N))⊗A,T0⊗A ◦ ZA
(λ,h)
∼ MA
Proof. Let q be a ε-r-projection in Mn(A). We can assume indeed without loss of
generality that n = 1.The Toeplitz extension is semi-split by the section induced
by the completely positive map s : C(S1) −→ T ; f 7→ Mf , where if π0 stands for
the projection L2(S1) ∼= ℓ2(Z) → l2(N), then Mf is the composition
(f· being the pointwise multiplication by f ). Notice first that (cid:0) S 1−SS∗
(cid:1) is a
unitary lift of S1 → M2(C); z 7→ diag(z, ¯z) in M2(T ) under the homomorphism
induced by ρ : T → C(S1). Under the section induced by s, we see that zq lifts to
1 ⊗ (1 − q) + S ⊗ q, and hence
l2(N) ֒→ ℓ2(Z) ∼= L2(S1)
f·→ L2(S1)
π0→ l2(N),
S∗
0
W =(cid:18)S 1 − SS∗
S∗ (cid:19) ⊗ q + I2 ⊗ (1 − q)
0
is a lift in U5ε,r
W ∗ diag(1, 0)W is closed to
2
(T0 ⊗ A) of diag(zq, z∗q ). Since kq(1 − q)k < ε, we see that
(cid:18) S∗
1 − SS∗ S(cid:19)(cid:18)1
0
0
0
0(cid:19)(cid:18)S 1 − SS∗
S∗ (cid:19) ⊗ q2 +(cid:18)1 0
0 0(cid:19) ⊗ (1 − q)2.
0
Hence, W ∗ diag(1, 0)W is an element of P10ε,2r
SS∗) ⊗ q). Since
2
(T0⊗A) which is closed to diag(1, (1−
MA([q, 0]ε,r) = [diag(0, (1 − SS∗) ⊗ q)]ε,r,
we get the existence of a positive real αt such that the proposition holds.
(cid:3)
3.3. Long exact sequence. We follow the route of [18, Sections 6.3, 7.1 and
8.2] to state for semi-split extensions of filtered C∗-algebras (λ, h)-exact long exact
sequences in quantitative K-theory, for some universal control pair (λ, h).
Proposition 3.8. There exists a control pair (λ, h) such that for any semi-split
extension of filtered C∗-algebras
0 −→ J
−→ A/J −→ 0,
−→ A
q
the composition
K∗(J)
j∗→ K∗(A)
q∗→ K∗(A/J)
is (λ, h)-exact at K∗(A).
Proof. We can assume without loss of generality that A is unital.
In the even
0 (A) such that q∗(y) = 0 in K ε,r
case, let y be an element of K ε,r
0 (A/J), let e be
an ε-r-projection in Mn(A) and let l be a positive integer such that y = [e, k]ε,r.
Up to stabilization, we can assume that k 6 n and that q(e) is homotopic to
pk = diag(Ik, 0) as an ε-r-projection in Mn(A/J). According to corollary 1.32,
there exists up to stabilization a αhε-kh,εr-unitary W of Mn(A/J) such that
kW q(e)W ∗ − pkk 6 αhε.
ON A QUANTITATIVE OPERATOR K-THEORY
27
The 3αsε-2kh,εr-unitary diag(W, W ∗) of M2n(A/J) is homotopic to I2n. Let choose
as in lemma 3.2, a control pair (α, l), an integer j and a αε-lεr-unitary V of
M2n+j(A) such that
kq(V ) − diag(W, W ∗, Ik+j )k 6 αε.
0
If we set e′ = V diag(e, 0)V ∗, then e′ is a 4αε-2lεr-projection in M2n+j(A).
If
s : A/J → A is a semi-split filtered cross-section such that s(1) = 1, define f =
e′ − s ◦ q(e′ − diag(In, 0)). We see that f belongs to M2n+j(J +) and moreover,
since kf − e′k 6 (4α + αh)ε, then according to lemma 1.7, f is for a suitable λ a
λε-2lεr-projection of M2n+k(J +) homotopic to e′. Then x = [f, k]λε,2lεr defines a
class in K λε,2lεr
(J). As in the proof of (ii) of lemma 1.8 we can choose λ big enough
so that diag(e′, I2n+j ) and diag(e, 0, I2n+j) are homotopic λε-2kh,εr-projections of
M2n(A) and hence we get the result in the even case.
For the odd case, let y be an element in K ε,r
1 (A/J)
and let us choose an ε-r-unitary V in some Mn(A) such that y = [V ]ε,r. In view
of lemma 3.2 and up to enlarge the size of the matrix V , we can assume that
kq(V ) − q(W )k 6 αeε with W a αeε-ke,εr-unitary of Mn(A) homotopic to In.
Hence W ∗V and V are homotopic 3αeε-(ke,ε + 1)r-unitary of Mn(A). If we set
1 (A) such that q∗(y) = 0 in K ε,r
U = W ∗V + s ◦ q(In − W ∗V ),
then the coefficients of the matrix U − In lie in J. Moreover, since
kU − W ∗V k 6 (2αe + 1)ε,
we obtain that U is a λε-(lε + 1)r-unitary for some λ > 1. Hence, x = [U ]λε,(ke,ε+1)r
defines a class in K λε,(ke,ε+1)r
(cid:3)
(J) with the required property.
1
Proposition 3.9. There exists a control pair (λ, h) such that for any semi-split
extension of filtered C∗-algebras
0 −→ J
−→ A/J −→ 0,
−→ A
q
the composition
K1(A)
q∗→ K1(A/J) D1
J,A→ K0(J)
is (λ, h)-exact at K1(A/J).
Proof. We can assume without loss of generality that A is unital. Let y be an
element of K ε,r
(A/J) and let U be an
ε-r-unitary of Mn(A/J) such that y = [U ]ε,r. With notation of lemma 3.2, let j be
an integer and W be a 3αeε-2ke,3εr-unitary in M2n+j(A) such that
J,A(y) = 0 in K α∂ ε,k∂,εr
1 (A/J) such that ∂ε,r
0
kq(W ) − diag(U, U∗, Ij )k 6 αε.
Set x = W diag(In, 0)W ∗ and h = x − diag(In, 0) − s ◦ q(x − diag(In, 0) as in the
proof of proposition 3.4. Since ∂ε,r
J,A(y) = 0, we can up to take a larger n assume that
h + diag(In, 0) is homotopic to diag(In, 0) as an αDε-kD,εr-projection of M2n+j( J).
Since x is close to h+ diag(In, 0), we get from corollary 1.32 that up to take a larger
j, there exists for a control pair (α, l), depending only on the control pairs (αh, kh)
and (αD, kD) of corollary 1.32 and lemma 3.3, an αε-lεr-unitary V ′ in M2n+j(eJ)
such that
W diag(In, 0)W ∗ − V ′ diag(In, 0)V ′∗k 6 αε.
28
H. OYONO-OYONO AND G. YU
Then V = ρJ (V ′)V ′−1W ∗ is a 10(α + αe)ε-(lε + ke,ε)r-unitary in M2n+j(A) such
that
Since for a suitable constant α′ depending only on α we have
kq(V ) − diag(U, U∗, Ij)k 6 αε.
kρJ (V ′) diag(In, 0)ρJ (V ′∗) − diag(In, 0)k 6 α′ε,
kV diag(In, 0)V ∗ − diag(In, 0)k 6 α′′ε
we obtain that
and
kV ∗ diag(In, 0)V − diag(In, 0)k 6 α′′ε
for some constant α′′ depending only on α′ that we can choose indeed larger than
(10α + αe). Hence the n × n-left upper corner X of V is a α′′ε-(lε + l′ε)r-unitary in
Mn(A) such that kq(X) − Uk 6 α′′ε. Hence we get the result.
Proposition 3.10. There exists a control pair (λ, h) such that for any semi-split
extension of filtered C∗-algebras
0 −→ J
−→ A/J −→ 0,
−→ A
(cid:3)
q
the composition
K1(A/J) D1
J,A→ K0(J)
∗→ K0(A)
∗ (y) = 0 in K ε,r
is (λ, h)-exact in K0(J).
Proof. It is enough to prove the result for A unital. Let y be an element of K ε,r
0 (J)
such that ε,r
0 (A), let e be an ε-r-projection in Mn(J +) and k be
a positive integer such that y = [e, k]ε,r. If we set pk = diag(Ik, 0), we can indeed
assume without loss of generality that kq(e) − pkk 6 2ε (where J + is viewed as a
subalgebra of A). Up to stabilization, we can also assume that e is homotopic to
pk as an ε-r-projection in Mn(A). According to corollary 1.32, there exists up to
stabilization a αhε-kh,εr-unitary W of Mn(A) such that
Up to replace n by 2n, W by diag(W, W ∗) and e by diag(e, 0), we can assume that
W is a 3αhε-2kh,εr-unitary homotopic to In. Since
ke − W pkW ∗k 6 αhε.
kq(W )pkq(W ∗) − pkk 6 kq(W )pkq(W ∗) − q(e)k + kq(e) − pkk
< (2 + αh)ε,
then
kq(W ∗)pkq(W ) − pkk < (2 + 4αh)ε.
Hence for an α′ > 1 depending only on αh, the left-up n × n corner V1 and the
right bottom corner V2 of q(W ) are α′ε-2ke,εr-unitaries of Mn(A/J) such that
kq(W )q(W ∗) − diag(V1, V2) diag(V1, V2)∗k < (αh + α′)ε
and
kq(W ∗)q(W ) − diag(V1, V2)∗ diag(V1, V2)k < (αh + α′)ε.
Hence q(W ) is close to diag(V1, V2) and hence there is a λ > 1 depending only on
αe such that as a λε-2kh,εr-unitary of Mn(A/J), then diag(V1, V2) is homotopic
ON A QUANTITATIVE OPERATOR K-THEORY
29
to q(W ) and hence to In. We can indeed choose λ big enough such that if we set
x = [V1]λε,2ke,ε r, then
∂λε,2ke,εr
J,A
(x) = [e, k]λα∂ ε,k∂,αε2ke,εr
= ιε,r,λε,2ke,εr
∗
(y).
(cid:3)
From propositions 3.8, 3.9 and 3.10 we can derive the analogue of the long exact
sequence in K-theory.
Theorem 3.11. There exists a control pair (λ, h) such that for any semi-split
extension of filtered C∗-algebras
0 −→ J
−→ A/J −→ 0,
−→ A
q
the sequence
K1(J)
is (λ, h)-exact.
∗−→ K1(A)
q∗−→ K1(A/J) DJ,A−→ K0(J)
∗−→ K0(A)
q∗−→ K0(A/J)
As a consequence, using the exact sequence
(2)
0 → SA → CA → A → 0,
and in view of lemma 1.27 and point (iii) of remark 2.8, we deduce in the setting
of the semigroup K ε,r
∗ (•) the analogue of the suspension isomorphism in K-theory.
Corollary 3.12. Let D1
SA,CA : K1(A) → K0(SA) be the controlled boundary
morphism associated to the semi-split and filtered extension of equation (2) for a
filtered C∗-algebra A.
A = D1
A is (λ, h)-invertible.
• There exists a control pair (λ, h) such that for any filtered C∗-algebra A,
then D1
• Moreover, we can choose a (λ, h)-inverse which is natural: there exists a
control pair (αβ, kβ) and for any filtered C∗-algebra A a (λ, h)-controlled
morphism B0
,r>0 : K0(SA) → K1(A) which is an (λ, h)-
inverse for D1
B ◦ fS = f ◦ B0
A for any homomorphism
f : A → B of filtered C∗-algebras, where fS : SA → SB is the suspension
of the homomorphism f .
A = (βε,r
A and such that B0
A )0<ε< 1
4αβ
3.4. The mapping cones. We end this section by proving that the mapping cones
construction can be performed in the framework of quantitative K-theory. Let
be a filtered semi-split extension of C∗-algebras. Let us set A/J[0, 1) = C0([0, 1), A/J)
and define the mapping cone of q:
0 → J → A
→ A/J→0
q
Cq = {(x, f ) ∈ A ⊕ A/J[0, 1); such that f (0) = q(x)}.
Using a semi-split filtered cross-section for q, we see that Cq is filtered by
Let us set
(Cq ∩ (Ar ⊕ A/J[0, 1))r)r>0 .
eq : J → Cq; x 7→ (x, 0)
30
and
H. OYONO-OYONO AND G. YU
We have then a semi-split extension of filtered C∗-algebras
φq : SA/J → Cq; f 7→ (0, f ).
0 → J
ej→ Cq
π2→ A/J[0, 1) → 0,
where π2 is the projection on the second factor of A ⊕ A/J[0, 1).
Lemma 3.13. There exists a control pair (λ, h) such that eq,∗ is (λ, h)-invertible
for any semi-split extension of filtered C∗-algebras 0 → J → A
Proof. The even case is a consequence of theorem 3.11. We deduce the odd case
from the even one using corollary 3.12.
(cid:3)
→ A/J → 0.
q
It is a standard fact in K-theory that the boundary of an extension of C∗-algebras
0 → J → A
→ A/J → 0 can be obtain using the equality
q
eq,∗ ◦ ∂J,A = φq,∗ ◦ ∂A/J ,
where ∂A/J = ∂SA/J,CA/J stands for the boundary map of the extension
0 → SA/J → CA/J → A/J → 0
(corresponding to the evaluation at 1). We have a similar result in quantitative
K-theory:
Lemma 3.14. With above notations, we have eq,∗ ◦ DJ,A = φq,∗ ◦ DA/J , where
DA/J stands for DSA/J,CA/J .
Proof. We can assume without loss of generality that A is unital. Let us fix a
semi-split filtered cross-section s : A/J → A such that s(1) = 1. Let p be an ε-r
projection in A/J. Using the notations of the proof of proposition 3.3, define for t
in [0, 1]
• xt =
lεXl=1
(2ıπts(p))l − t(2ıπ)ls(pl)
l!
in A;
• ft : [0, 1] → A/J : σ 7→
l!
Then, (1 + (yt, ft))t∈[0,1] is a path of αε-kεr unitary in C+
Moreover,
lεXl=1
((2ıπ(1 − σ)t + σ)p)l − ((1 − σ)t + σ)(2ıπp)l
.
q with x0 = 0 and f1 = 0.
q
• x1 belongs to J and satisfies the conclusion of lemma 3.3 starting from
the ε-r-projection p and with respect to the semi-split extension of filtered
C∗-algebras 0 → J → A
→ A/J → 0 and to the semi-split filtered cross-
section s;
• f0 belongs to SA/J and satisfies the conclusion of lemma 3.3 starting from
the ε-r-projection p and with respect to the semi-split extension of filtered
C∗-algebras 0 → SA/J → CA/J→A/J → 0 corresponding to evaluation
at 1 and to the semi-split filtered cross-section A/J 7→ CA/J; a 7→ [t 7→ ta].
Hence, following the construction of proposition 3.4 in the even case, we obtain
that eq,∗ ◦ DJ,A and φq,∗ ◦ DA/J coincide on K0(A/J).
Let us check now the odd case. Let u be an ε-r-unitary in Mn(A/J). Pick
any ε-r-unitary in some Mj(A/J) such that diag(u, v) is homotopic to In+j in
U3ε,2r
n+j (A/J). According to lemma 3.2, and up to replace v by diag(v, Ik) for some
ON A QUANTITATIVE OPERATOR K-THEORY
31
integer k, there exists an element w in U3αeε,2ke,3εr
(A) homotopic to In+j as a 3αeε-
2ke,3εr-unitary and such that kq(w) − diag(u, v)k 6 3αeε. Let (wt)t∈[0,1] be a path
in U3αeε,2ke,3εr
(A) with w0 = In+j and w1 = w and set yt = q(wt) diag(In, 0)q(w∗t ).
As in the proof of proposition 3.4, we see that yt is an element in P12αeε,4ke,3εr
(A/J)
such that ky1 − diag(In, 0)k 6 9αeε. Define
n+j
n+j
n+j
g : [0, 1] → Mn+j(A/J); t 7→ yt − diag(In, 0) − t(y1 − diag(In, 0)).
n+j
Then g + diag(In, 0) is the element of P12αeε,4ke,3εr
(S+A/J) that we get from u
and v when we perform the construction of proposition 3.4 in the odd case with
respect to the extension 0 → SA/J → CA/J → A/J → 0. Let us set now
xt = wt diag(In, 0)w∗t and ht = xt − diag(In, 0) − ts ◦ q(x1 − diag(In, 0)) for t
in [0, 1]. Then diag(In, 0) + ht belongs to P12αeε,4ke,3εr
(A) and diag(In, 0) + h1
is the element of P12αeε,4ke,3εr
(J) that we get from u and v when we perform
the construction of proposition 3.4 in the odd case with respect to the extension
0 → J → A
→ A/J → 0. Eventually, if we define
n+j
n+j
q
Ht : [0, 1] → Mn+j(A/J); σ 7→ g(1−σ)t+σ,
then ((ht, Ht) + diag(In, 0))t∈[0,1] is a homotopy in P12αeε,4ke,3εr
and ((h1, 0) + diag(In, 0)). Thus we obtain the result in the odd case.
(C+
n+j
q ) between ((0, g) + diag(In, 0))
As a consequence, we get that the controlled suspension morphism is compatible
with the controlled boundary maps.
Proposition 3.15. There exists a control pair (λ, h) such that for any semi-split
extension of filtered C∗-algebras 0 → J → A → A/J → 0, the following diagrams
are (λ, h)-commutative:
(cid:3)
K0(A/J) DA/J−−−−→ K1(SA/J)
yDSJ,SA
DJ,Ay
DJ−−−−→ K0(SJ)
K1(J)
and
K1(A/J) DA/J−−−−→ K0(SA/J)
yDSJ,SA
DJ,Ay
DJ−−−−→ K1(SJ)
K0(J)
,
where DJ and DA/J stands respectively for the controlled suspension morphisms
DSJ,CJ and DSA/J,CA/J .
Proof. Let qs : SA → SA/J the suspension of the homomorphism q : A → A/J.
Applying lemma 3.14 to the extensions 0 → J → A → A/J → 0 and 0 → SJ →
SA → SA/J → 0 and using the naturality of controlled boundary maps mentioned
32
H. OYONO-OYONO AND G. YU
in remark 3.5, we get
eqs,∗ ◦ DSJ,SA ◦ DA/J = φqs,∗ ◦ DSA/J ◦ DA/J
= DSCq ◦ φq,∗ ◦ DA/J
= DSCq ◦ eq,∗ ◦ DJ,A
= eqs,∗ ◦ DJ ◦ DJ,A
The proposition is then a consequence of lemma 3.13.
(cid:3)
4. Controlled Bott periodicity
The aim of this section is to prove that there exists a control pair (λ, h) such that
given a filtered C∗-algebra A, then Bott periodicity K0(A) ∼=→ K0(S2A) is induced
in K-theory by a (λ, h)-isomorphism K0(A) → K0(S2A). As an application, we use
the controlled boundary morphism of proposition 3.4 to close the controlled exact
sequence of 3.11 into a six-term (λ, h)-exact sequence for some universal control
pair (λ, h). This will be achieved by using the full power of KK-theory.
4.1. Tensorization in KK-theory. Let A be a C∗-algebra and let B be a C∗-
algebra filtered by (Br)r>0. Within all this section, we will assume for sake of
simplicity that Br is closed for every positive number r (which is the case for Roe
algebras and crossed product algebras). Let us define A⊗Br as the closure in the
spatial tensor product A⊗B of the algebraic tensor product of A and Br. Then the
C∗-algebra A⊗B is filtered by (A⊗Br)r>0. Moreover, if J is a semi-split ideal of
A, i.e 0 → J → A → A/J → 0 is a semi-split extension of C∗algebras, then
0 → J⊗B → A⊗B → A/J⊗B → 0
is a semi-split extension of filtered C∗-algebras. Recall from [11] that for C∗-
algebras A1, A2 and D, G. Kasparov defined a tensorization map
τD : KK∗(A1, A2) → KK∗(A1⊗D, A2⊗D)
in the following way: let z be an element in KK∗(A1, A2) represented by a K-cycle
(π, T,E), where
• E is a right A2-Hilbert module;
• π is a representation of A1 into the algebra L(E) of adjointable operators
• T is a self-adjoint operator on E satisfying the K-cycle conditions, i.e.
of E;
[T, π(a)], π(a)(T 2 − IdE ) are compact operators on E for any a in A1.
Then τD(z) ∈ KK∗(A1⊗D, A2⊗D) is represented by the K-cycle (π⊗IdD, T⊗IdD,E⊗D).
In what follows, we show that if A1 and A2 are C∗-algebras, if B is a fil-
tered C∗-algebra and if z is an element in KK∗(A1, A2), then the homomorphism
K∗(A1⊗B) → K∗(A2⊗B) provided by left multiplication by τB(z) is induced by a
controlled morphism. Moreover, we have some compatibility results with respect
to Kasparov product. As an outcome, we obtain a controlled version of the Bott
periodicity that induces in K-theory the Bott periodicity.
Proposition 4.1. Let A1 and A2 be C∗-algebras, let B be a filtered C∗-algebra
and let z be an element in KK1(A1, A2). Then there exists an (αD, kD)-controlled
ON A QUANTITATIVE OPERATOR K-THEORY
33
morphism
TB(z) = (τ ε,r
B (z))0<ε< 1
4αD
,r>0 : K∗(A1⊗B) → K∗(A2⊗B)
of degree 1 inducing in K-theory the right multiplication by τB(z).
Proof. Recall that z can be indeed represented by a odd A1-A2-K-cycle (π, T,H⊗A2),
where H is a separable Hilbert space, π is a representation of A1 in the algebra
L(H⊗A2) of adjointable operators of H⊗A2 and T is a self-adjoint operator in
L(H⊗A2) satisfying the K-cycle conditions. Let us set PB = IdH⊗A2⊗B +T⊗IdB
,
πB = π⊗IdB and define the C∗-algebra
E(π,T ) = {(x, y) ∈ A1⊗BML(H⊗A2⊗B) such that PB·πB(x)·PB−y ∈ K(H)⊗A2⊗B}.
2
Since PB has no propagation, the C∗-algebra E(π,T ) is filtered by (E(π,T )
r
)r>0 with
E(π,T )
r
= {(x, PB · πB(x) · PB + y); x ∈ A1⊗Br and y ∈ K(H) ⊗ A2⊗Br}.
The extension of filtered C∗-algebras
(3)
0 −→ K(H) ⊗ A2⊗B −→ E(π,T ) −→ A1 ⊗ B −→ 0
is semi-split by the cross-section
s : A1⊗B → E(π,T ); x 7→ (x, PB · πB(x) · PB).
Let us show that the associated controlled boundary (degree one) map
DK(H)⊗A2⊗B,E(π,T ) : K∗(A1⊗B) → K∗(K(H)⊗A2⊗B)
only depends on the class z of (π, T,H ⊗ A2) in KK1(A1, A2). Assume that
(π, T,H ⊗ A2[0, 1]) is a A1-A2[0, 1]-K -cycle providing a homotopy between two
A1-A2-K-cycles (π0, T0,H ⊗ A2) and (π1, T1,H ⊗ A2). For t ∈ [0, 1] we denote by
• et : A2[0, 1] → A2 the evaluation at t;
• Ft ∈ L(H ⊗ A2) the fiber at t of an operator F ∈ L(H ⊗ A2[0, 1]);
• πt : A1 → L(H ⊗ A2) the representation induced by π at the fiber t;
(with P = T +1
• st : A2⊗B → E(πt,Tt); x 7→ (x, Pt,B · πt,B(x) · Pt,B)
2 );
Then the homomorphism E(π,T ) → E(πt,Tt); (x, y) 7→ (x, yt) satisfies the con-
ditions of remark 3.5 (with s : A2⊗B → E(π,T ); x 7→ (x, PB · πB(x) · PB) and
st : A2⊗B → E(π,Tt)) and thus we get that
(IdK(H) ⊗ et⊗IdB)∗ ◦ DK(H)⊗A1⊗B[0,1],E(π,T ) = DK(H)⊗A1⊗B,E(πt ,Tt),
and according to lemma 1.27, we deduce that
DK(H)⊗A1⊗B2,E(π0,T0) = DK(H)⊗A1⊗B,E(π1,T1) .
This shows that for a A1-A2-K-cycle (π, T,H ⊗ A2), then DK(H)⊗A1⊗B,E(π,T ) de-
pends only on the class z of (π, T,H ⊗ A2) in KK1(A1, A2). Finally we define
TB(z) = (τ ε,r
B (z))0<ε< 1
4αD
def== M−1
A2⊗B ◦ DK(H)⊗A1⊗B,E(π,T ) ,
where
• (π, T,H⊗A2) is any A1-A2-K-cycles representing z;
• MA2⊗B is the Morita equivalence (see example 2.2).
34
H. OYONO-OYONO AND G. YU
The result then follows from the observation that up to the Morita equivalence
K∗(K(H) ⊗ A2⊗B) ∼=→ K∗(A2⊗B),
the boundary ∂K(H)⊗A1⊗B,E(π,T ) corresponding to the exact sequence (3) is induced
by right multiplication by τB(z).
(cid:3)
Remark 4.2. Let B be a filtered C∗-algebra.
(i) For any C∗-algebras A1 and A2 and any elements z and z′ in KK1(A1, A2)
then
TB(z + z′) = TB(z) + TB(z′).
(ii) Let 0 → J → A → A/J → 0 be a semi-split extension of filtered C∗-
algebras and let [∂J,A] be the element of KK1(A/J, J) that implements the
boundary map ∂J,A. Then we have
TB([∂J,A]) = DJ⊗B,A⊗B.
(iii) For any C∗-algebras A1, A2 and D and any K-cycle (π, T,H⊗A2) for
KK1(A1, A2), we have a natural identification between E(πD ,TD) and E(π,T )⊗D.
Hence, for any element z in KK1(A1, A2) then TB(τD(z)) = TB⊗D(z).
For a a filtered C∗-algebra B and a homomorphism f : A1 → A2 of C∗-algebras,
we set fB : A1⊗B → A2⊗B for the filtered homomorphism induced by f .
Proposition 4.3. Let B be a filtered C∗-algebra and let A1 and A2 be two C∗-
algebras.
(i) For any C∗-algebra A′1, any homomorphism of C∗-algebras f : A1 → A′1
(ii) For any C∗-algebra A′2, any homomorphism of C∗-algebras g : A2 → A′2
and any z in KK1(A′1, A2), we have TB(f∗(z)) = TB(z) ◦ fB,∗;
and any z in KK1(A1, A2), we have TB(g∗(z)) = gB,∗ ◦ TB(z).
Proof.
(i) Let A′1 be a filtered C∗-algebra, let f : A1 → A′1 be a homomorphism of
C∗-algebras and let (π, T, H ⊗ A2) be an odd A′1-A2-K-cycle. With the
notations of the proof of proposition 4.1, the homomorphism
f E : Ef ∗(π,T ) → E(π,T ); (x, y) 7→ (fB(x), y)
fits in the commutative diagram
0 −−−−→ K(H) ⊗ A2⊗B −−−−→ Ef ∗(π,T ) −−−−→ A1⊗B −−−−→ 0
.
f Ey
yfB
=y
0 −−−−→ K(H) ⊗ A2⊗B −−−−→ E(π,T ) −−−−→ A′1⊗B −−−−→ 0
Moreover fB and f E intertwines the semi-split and filtered cross-sections
A1⊗B → Ef ∗(π,T ); x 7→ (x, PB · πB ◦ fB(x) · PB)
and
A′1⊗B → E(π,T ); x 7→ (x, PB · πB(x) · PB)
and thus, we get by remark 3.5 that
for all z in KK1(A′1, A2).
TB(f∗(z)) = TB(z) ◦ f∗
ON A QUANTITATIVE OPERATOR K-THEORY
35
(ii) Let A′2 be a C∗-algebra and let g : A2 → A′2 be a homomorphism of
C∗-algebras. For any element F in L(H ⊗ A2), let us denote by
F = F⊗A2IdA′
2 ∈ L(H ⊗ A2⊗A2 A′2).
Notice that H⊗ A2⊗A2A′2 can be viewed as a right A′2-Hilbert-submodule
of H⊗A′2 and under this identification, for any F in K(H) ⊗ A2, then F is
the restriction to H⊗A2⊗A2A′2 of the homomorphism (IdK(H)⊗g)(F ). Let
z be an element of KK1(A1, A2) represented by a K-cycle (π, T,H⊗A2).
Consider the A1-A2-K-cycle (π′, T ′,H′⊗A2) with H′ = H1 ⊕ H2 ⊕ H3,
where H1, H2 and H3 are three copies of H, π′ = 0 ⊕ 0 ⊕ π and T ′ =
IdH1⊗A2 ⊕ IdH2⊗A2 ⊕ T . Then (π′, T ′,H′⊗A2) is again a K-cycle repre-
senting z and g∗(z) is represented by the K-cycle (π′′, T ′′,E), where
• E = H1 ⊗ A′2LH2 ⊗ A′2LH3 ⊗ A2⊗A2A′2;
• π′′ = 0 ⊕ 0 ⊕ π;
• T ′′ = IdH1⊗A′
Using Kasparov stabilization theorem, we get that H2⊗A′2L H3⊗A2⊗A2A′2
is isomorphic as a right-A′2-Hilbert module to H⊗ A′2 and hence, using this
identification, we can represent g∗(z) using a standard right-A′2-Hillbert
module, as in the proof of proposition 4.1. Then, under the above identi-
fication H2 ⊗ A′2LH3 ⊗ A2⊗A2A′2 ∼= H ⊗ A′2,
gE : E(π,T ) → Eg∗(π,T )
2 ⊕ IdH2⊗A′
2 ⊕ T .
(x, y)
7→ (x, P ′′Bπ′′(x)P ′′B + (IdK(H′)⊗B⊗g)(y − P ′Bπ′(x)P ′B))
restricts to a homomorphism K(H1⊕H2⊕H3)⊗A2⊗B → K(H1⊕H)⊗A′2⊗B.
We get now a commutative diagram
0 −−−−→ K(H1 ⊕ H2 ⊕ H3) ⊗ A2⊗B −−−−→ E(π′,T ′) −−−−→ A1⊗B −−−−→ 0
.
−−−−→ E(π′′,T ′′) −−−−→ A1⊗B −−−−→ 0
0 −−−−→ K(H1 ⊕ H) ⊗ A′2⊗B
gEy
gEy
y=
Hence, we get by remark 3.5 that
DK(H)⊗A′
2⊗B,E(π′′ ,T ′′ ) = gE,∗ ◦ DK(H)⊗A2⊗B,E(π′,T ′ ).
But the restriction of gE to the corner K(H1)⊗A2⊗B of the C∗-algebra
K(H1⊕H2⊕H3)⊗A2⊗B is IdK(H1)⊗g⊗IdB. Since the Morita equivalence
MA′
2⊗B : K∗(A′2⊗B) ∼=→ K∗(K(H1 ⊕ H)⊗A′2⊗B)
can be implemented by an inclusion of A′2⊗B in a corner of K(H1)⊗A′2⊗B,
and similarly for the Morita equivalence
MA2⊗B : K∗(A2⊗B) ∼=→ K∗(K(H1 ⊕ H2 ⊕ H3)⊗A2⊗B),
we deduce that the two following compositions coincide:
K∗(A2⊗B))
and
gB,∗−→ K∗(A′2⊗B)
2
⊗B
MA′
−→ K∗(K(H1 ⊕ H)⊗(A′2⊗B))
K∗(A2⊗B) MA2⊗B−→ K∗(K(H1 ⊕ H2 ⊕ H3)⊗A2⊗B)
gE,∗−→ K∗(K(H1 ⊕ H)⊗A′2⊗B).
36
H. OYONO-OYONO AND G. YU
Hence we get
for any z in KK1(A1, A2).
TB(g∗(z)) = g∗ ◦ TB(z)
(cid:3)
Let us now extend the definition of TB to the even case. Consider for a suit-
able control pair (αB, kB) and any filtered C∗-algebra A the (αB, kB)-controlled
morphism of odd degree BA : K∗(SA) → K∗(A) defined
• by B0
• by M−1
A on K0(SA) as in corollary 3.12;
A ◦ DK(ℓ2(N))⊗A,T0⊗A on K1(SA) using the Toeplitz extension
0 → K(ℓ2(N)) ⊗ A → T0 ⊗ A→SA → 0
(see the discussion at the end of section 3.2).
Then, according to corollary 3.12 and proposition 3.7, there exists a control pair
(λ, h) such that BA is a right (λ, h)-inverse for DSA,CA for any filtered C∗-algebra
A. Let us set αT = λαB and kT = h ∗ kB.
Now, let B be a filtered C∗-algebra, let A1 and A2 be C∗-algebras, then define
for any z in KK0(A1, A2) the (αT , kT )-controlled morphism
TB(z) = (τ ε,r
B )0<ε< 1
4αT
,r>0 : K∗(A1⊗B) → K∗(A2⊗B)
by
where
TB(z) def==BA2⊗B ◦ TB(z ⊗A2 [∂A2 ])
• [∂A2 ] = [∂SA2,CA2] ∈ KK1(A2, SA2) corresponds to the boundary of the
• ⊗A2 stands for Kasparov product.
exact sequence 0 → SA2 → CA2 → A → 0;
Up to compose on the left with ιαD ε,αT ε,kD r,kT r
TB(•) also as an (αT , kT )-controlled morphism.
Theorem 4.4. Let B be a filtered C∗-algebra, let A1 and A2 be C∗-algebras
, we can in the odd case define
∗
(i) For any element z in KK∗(A1, A2), then TB(z) : K∗(A1⊗B) → K∗(A2⊗B)
is a (αT , kT )-controlled morphism with same degree as z that induces in
K-theory right multiplication by τB(z).
(ii) For any elements z and z′ in KK∗(A1, A2) then
TB(z + z′) = TB(z) + TB(z′).
(iii) Let A′1 be a filtered C∗-algebras and let f : A1 → A′1 be a homomorphism
of C∗-algebras, then TB(f∗(z)) = TB(z) ◦ fB,∗ for all z in KK∗(A′1, A2).
(iv) Let A′2 be a C∗-algebra and let g : A′2 → A2 be a homomorphism of C∗-
algebras then TB(g∗(z)) = gB,∗ ◦ TB(z) for any z in KK∗(A1, A′2).
(αT ,kT )
(v) TB([IdA1 ])
(vi) For any C∗-algebra D and any element z in KK∗(A1, A2), we have TB(τD(z)) =
TB⊗D(z).
∼ IdK∗(A1⊗B).
Proof. Since BA2⊗B is a right (λ, h)-inverse for DSA2⊗B,CA2⊗B, it induces in K-
theory a right inverse (indeed an inverse) for the (degree 1) boundary map
∂SA2⊗B,CA2⊗B : K∗(A2⊗B) → K∗(SA2⊗B).
ON A QUANTITATIVE OPERATOR K-THEORY
37
But since TB(z⊗A2[∂SA2⊗B,CA2⊗B]) induces in K-theory right multiplication by
z⊗A2 [∂SA2⊗B,CA2⊗B], we eventually get that TB(z⊗A2 [∂SA2⊗B,CA2⊗B]) induced in
K-theory the composition
K∗(A1⊗B) ⊗A1⊗B τB (z)
−→ K∗(A2⊗B)
and hence we get the first point.
∂SA2 ⊗B,CA2 ⊗B
−→
K∗(SA2⊗B)
Point
(ii) is a consequence of remark 4.2. Point
(iii) is a consequence of
proposition 4.3. Point (iv) is a consequence of proposition 4.3 and of the naturality
of B• (see remark 3.5 and corollary 3.12), point (v) holds by definition of B•. Point
(vi) is a consequence of point (iii) of remark 4.2.
(cid:3)
We end this section by proving the compatibility of TB with Kasparov product.
Theorem 4.5. There exists a control pair (λ, h) such that the following holds :
let A1, A2 and A3 be C∗-algebras and let B be a filtered C∗-algebra. Then for
any z in KK∗(A1, A2) and any z′ in KK∗(A2, A3), we have
TB(z⊗A2z′)
(λ,h)
∼ TB(z′) ◦ TB(z).
Proof. We first deal with the case z even. According to [12, Lemma 1.6.9], there
exists a C∗-algebra A4 and homomorphisms θ : A4 → A1 and η : A4 → A2 such
that
• the element [θ] of KK∗(A4, A1) induced by θ is invertible.
• z = η∗([θ]−1).
Since θ∗([θ]−1) = [IdA1 ] in KK∗(A1, A1), we get in view of remark 2.5 and of points
(iii), (iv) and (v) of theorem 4.4 that
(λ,h)
∼ TB(θ∗(z⊗A2z′)) ◦ TB([θ]−1),
TB(z⊗A2 z′)
with (λ, h) = (α2
, kT ∗ kT ). But by bi-functoriality of KK-theory, we have
T
θ∗(z⊗A2 z′) = η∗(z′) and then the result is a consequence of points (iii) and (iv)
of theorem 4.4. We can proceed similarly when z′ is even. Let us prove now the
result when z and z′ are odd. Then [∂A2 ] = [∂SA2,CA2] is an invertible element
in KK1(A2, SA2) and z⊗A2z′ = z⊗A2 [∂A2 ]⊗SA2 [∂A2]−1⊗A2z′ and hence using the
even case, we get that
(4)
But
TB(z⊗A2z′)
(λ,h)
∼ TB([∂A2 ]−1⊗A2 z′) ◦ TB(z⊗A2 [∂A2]).
TB([∂A2 ]−1⊗A2 z′)
(λ′,h′)
=
∼
BA3⊗B ◦ TB([∂A2 ]−1⊗A2z′⊗A3 [∂A3])
BA3⊗B ◦ TB(z′⊗A3 [∂A3]) ◦ TB([∂A2]−1)
(5)
for some control pair (λ′, h′), depending only on (λ, h) and (αT , kT ), where equation
(5) holds by the even case applied to z′⊗A3[∂A3 ] and [∂A2]−1. Hence, for a control
pair (λ′′, h′′)-depending only on (λ, h), we get applying the even case to [∂A2]−1
and z⊗A2[∂A2 ] that
(6)
∼ BA3⊗B ◦ TB(z′⊗A3 [∂A3]) ◦ TB(z).
TB(z⊗A2 z′)
(λ′′,h′′)
In view of this equation, we deduce the odd case from the controlled Bott period-
if we set [∂] = [∂C0(0,1),C0(0,1]] ∈
icity, which will be proved in the next lemma:
KK1(C, C0(0, 1)), then there exists a controlled (α, k) such that TA([∂]−1) is an
38
H. OYONO-OYONO AND G. YU
(α, k)-inverse for DA for any filtered C∗-algebra A. Indeed, from this claim and
since for some control pair (α′, k′), the (αB, kB)-controlled morphism BA is for every
filtered C∗-algebra A a right (α′, k′)-inverse for TA([∂]), we get that
TA([∂]−1)
(α′′,k′′)
∼ BA
for some controlled pair (α′′, k′′) depending only on (α′, k′) and (αT , kT ). Noticing
by using point (vi) of theorem 4.4, that TA3⊗B([∂]−1) = TB([∂A3 ]−1), the proof of
the theorem in the odd case is then by equation (6) a consequence of the even case
applied to [∂A3]−1 and z′⊗A3[∂A3 ]
(cid:3)
4.2. The controlled Bott isomorphism. We prove in this subsection a con-
trolled version of Bott periodicity. The proof use the even case of theorem 4.5 and
is needed for the proof of the odd case. Let A = (Ar)r>0 be a filtered C∗-algebra
and let us assume that Ar is closed for every positive number r. Let us denote for
short as before DSA,CA by DA and [∂SA,CA] by [∂A] for any filtered C∗-algebra A
and let us set [∂] = [∂C].
(λ,h)
Theorem 4.6. There exists a control pair (α, k) such that for every filtered C∗-
algebra A, then TA([∂]−1) is an (α, k)-inverse for DA.
Proof. Consider the even element z = [∂]⊗S[∂S] of KK∗(C, S2), where S = C0(0, 1)
and S2 = SS. The lemma is a consequence of the following claim: there exists a
∼ TA(z) for any C∗-algebra A. Before
control pair (λ, h) such that DSA ◦ DA
proving the claim, let us see how it implies the lemma. Notice first that by point
(ii) of remark 4.2, we have DA = TA([∂]). Since by associativity of Kasparov
product [∂]−1⊗Cz = [∂S], we get from theorem 4.5 applied to the even case, that
there exists a control pair (λ′, h′) such that for any filtered C∗-algebra A, then
TA(z) ◦ TA([∂]−1) ◦ DA
∼ DSA ◦ DA. Using the claim and since z is an invertible
element of KK∗(C, S2), we obtain from theorem 4.5 applied to the even case that
there exists a control pair (α, k) such that TA([∂]−1) is a left (α, k)-inverse for DA.
Using associativity of the Kasparov product, we see that [∂] = z⊗S2 [∂S]−1. Then
applying twice theorem 4.5, on one hand to [∂]−1 and z⊗S2[∂S]−1 and on the other
hand to [∂]−1⊗z and [∂S]−1, we get that there exists a control pair (α′, k′) such
that TA([∂]) ◦ TA([∂]−1)
∼ TSA([∂]−1) ◦ TSA([∂]). But according to what we
have seen before, TSA([∂]−1) ◦ TSA([∂])
Let us now prove the claim. It is known that up to Morita equivalence, [∂A]−1 is
the element of KK1(SA, A) corresponding to the boundary element of the Toeplitz
extension
∼ IdK∗(SA).
(α′,k′)
(α,k)
(λ,h)
0 → K(ℓ2(N)) ⊗ A → T0 ⊗ A→SA → 0.
A : K0(A) → K1(SA) and D1
Let us respectively denote by D0
A : K1(A) → K0(SA)
the restriction of DA to K0(A) and K1(A). According to proposition 3.7, there
exists a control pair (λ′, h′) such that, on even elements
(λ′,h′)
TA([∂]−1) ◦ D0
(7)
Since [∂S] = [∂]−1⊗z, we get by left composition by TA(z) in equation (7) and by us-
ing theorem 4.5 in the even case that there exists a control pair (λ, h) depending only
on (λ′, h′) and such that that D1
A (z) : K0(A) → K0(S2A)
A (z) (here T 0
∼ IdK0(A).
SA◦D0
∼ T 0
(λ,h)
A
A
ON A QUANTITATIVE OPERATOR K-THEORY
39
stands for the restriction of TA(z) to K0(A)). For the odd case, we know from
corollary 3.12 that there exists a control pair (λ′′, h′′) such that D1
S2A : K1(S2A) →
K0(S3A) is (λ′′, h′′)-invertible. Using the previous case, and since by associativity
of the Kasparov product, we have [∂A]⊗SAτSA(z) = τA(z)⊗[∂S2A], we get by ap-
plying twice theorem 4.5 in the even case that there exists a control pair (λ′′′, h′′′)
such that D1
A (z) : K1(A) → K1(S2A)
A(z), where T 1
S2A ◦T 1
S2A : K1(S2A) → K0(S3A) is (λ′′, h′′)-
is the restriction of TA(z) to K1(A). Since D1
invertible, we get the result by remark 2.5.
S2A ◦D0
∼ D1
SA ◦D1
(λ′′′,h′′′)
(cid:3)
A
4.3. The six term (λ, h)-exact sequence. Recall from proposition 3.15 that
there exists a control pair (λ, h) such that for any semi-split extension of filtered C∗-
algebras 0 → J → A → A/J → 0, the following diagrams are (λ, h)-commutative:
and
K0(A/J) DA/J−−−−→ K1(SA/J)
DJ,Ay
yDSJ,SA
DJ−−−−→ K0(SJ)
K1(J)
K1(A/J) DA/J−−−−→ K0(SA/J)
yDSJ,SA
DJ,Ay
DJ−−−−→ K1(SJ)
K0(J)
As a consequence, by using theorem 4.6 and proposition 3.11, we get
Theorem 4.7. There exists a control pair (λ, h) such that for any semi-split ex-
tension of filtered C∗-algebras
0 −→ J
−→ A
q
−→ A/J −→ 0,
with Ar closed for every positive number r, then the following six-term sequence is
(λ, h)-exact
K0(J)
∗−−−−→ K0(A)
q∗−−−−→ K0(A/J)
DJ,Ax
K1(A/J)
q∗←−−−− K1(A)
∗←−−−− K1(J)
DJ,Ay
Remark 4.8. Let us consider with notations of section 3.4 the semi-split extension
of filtered C∗-algebras
φq→ Cq
0 → SA/J
(8)
where π1 : Cq → A is the projection on the first factor of Cq. Since we have a
π2→ A/J[0, 1) → 0, and since
semi-split extension of filtered algebras 0 → J
A/J[0, 1) is a contractible filtered C∗-algebra, we see in view of theorem 4.7 that
ej,∗ : K∗(J) → K∗(Cq) is a controlled isomorphism. It is then plain to check that
up to the controlled isomorphism ej,∗ and DA/J : K∗(SA/J) → K∗(A/J), we get
from the semi-split extension of filtered C∗-algebras of equation (8) (for a possibly
different control pair) the controlled six-term exact sequence of theorem 4.7.
π1→ A → 0,
ej→ Cq
If we apply theorem 4.7 to a filtered and split extension, we get:
40
H. OYONO-OYONO AND G. YU
Corollary 4.9. There exists a control pair (λ, h) such that for every split extension
of filtered C∗-algebra 0 → J → A → A/J → 0, with Ar closed for every positive
number r and any filtered split cross-section s : A/J → A, then
K∗(J) ⊕ K∗(A/J) −→ K∗(A); (x, y) 7→ ∗(x) + s∗(y)
is (λ, h)-invertible.
5. Quantitative K-theory for crossed product C∗-algebras
In this section, we study quantitative K-theory for crossed product C∗-algebras
and discuss its applications to K-amenability.
Let Γ be a finitely generated group. A Γ-C∗-algebra is a separable C∗-algebra
equipped with an action of Γ by automorphisms. Recall that the convolution alge-
bra Cc(Γ, A) of finitely supported A-valued functions on Γ admits two canonical C∗-
completions, the reduced crossed product A⋊redΓ and the maximal crossed product
A⋊maxΓ. Moreover, there is a canonical epimorphism λΓ,A : A⋊maxΓ → A⋊redΓ
which is the identity on Cc(Γ, A).
5.1. Lengths and propagation. Recall that a length on Γ is a map ℓ : Γ → R+
such that
• ℓ(γ) = 0 if and only if γ is the identity element e of Γ;
• ℓ(γγ′) 6 ℓ(γ) + ℓ(γ′) for all element γ and γ′ of Γ.
• ℓ(γ) = ℓ(γ−1).
In what follows, we will assume that ℓ is a word length arising from a finite generat-
ing symmetric set S, i.e ℓ(γ) = inf{d such that γ = γ1 ··· γd with γ1, . . . , γd in S}.
Let us denote by B(e, r) the ball centered at the neutral element of Γ with radius
r, i.e B(e, r) = {γ ∈ Γ such that ℓ(γ) 6 r}. For any positive number r, we set
(A⋊redΓ)r
def=={f ∈ Cc(Γ, A) with support in B(e, r)}.
Then the C∗-algebra A⋊redΓ is filtered by ((A⋊redΓ)r)r>0.
def=={f ∈ Cc(Γ, A) with support in B(e, r)}, then the C∗-algebra
setting (A⋊maxΓ)r
A⋊maxΓ is filtered by ((A⋊maxΓ)r)r>0 (notice that as sets, (A⋊redΓ)r = (A⋊maxΓ)r).
It is straightforward to check that two word lengths give rise for A⋊redΓ (resp. for
A⋊maxΓ) to quantitative K-theories related by a (1, c)-controlled isomorphism for
a constant c.
In the same way,
For a homomorphism f : A → B of Γ-C∗-algebras, we denote respectively by
fΓ,red : A⋊redΓ → B⋊redΓ and fΓ,max : A⋊maxΓ → B⋊maxΓ the homomorphisms
respectively induced by f on the reduced and on the maximal crossed product.
−→ A/J −→ 0, we
For any semi-split extension of Γ-C∗-algebras 0 −→ J
have semi-split extensions of filtered C∗-algebras
q
−→ A
0 −→ J ⋊redΓ
Γ,red−→ A⋊redΓ
qΓ,red−→ A/J ⋊redΓ −→ 0
and
0 −→ J ⋊maxΓ
Γ,max−→ A⋊maxΓ
qΓ,max−→ A/J ⋊maxΓ −→ 0
and hence, by theorem 4.7, we get:
ON A QUANTITATIVE OPERATOR K-THEORY
41
Proposition 5.1. There exists a control pair (λ, h) such that for any semi-split
extension of Γ-C∗-algebras
0 −→ J
−→ A
q
−→ A/J −→ 0,
the following six-term sequences are (λ, h)-exact
−−−−−→ K0(A⋊redΓ)
K0(J ⋊redΓ)
Γ,red,∗
DJ⋊red Γ,A⋊red Γx
and
DJ⋊red Γ,A⋊max Γx
K1(A/J ⋊redΓ)
qΓ,red,∗←−−−−− K1(A⋊redΓ)
K0(J ⋊maxΓ)
Γ,max,∗
−−−−−→ K0(A⋊maxΓ)
K1(A/J ⋊maxΓ)
qΓ,max,∗←−−−−− K1(A⋊maxΓ)
qΓ,red,∗
−−−−−→ K0(A/J ⋊redΓ)
DJ⋊red Γ,A⋊red Γy
Γ,red,∗←−−−−− K1(J ⋊redΓ)
qΓ,max,∗
−−−−−→ K0(A/J ⋊maxΓ)
DJ⋊max Γ,A⋊max Γy
Γ,max,∗←−−−−− K1(J ⋊maxΓ)
5.2. Kasparov transformation. In this subsection we see how a slight modifica-
tion of the argument used in section 4.1 allowed to define a controlled version of
the Kasparov transformation compatible with Kasparov product.
Notice first that every element z of KK Γ
∗ (A, B) can be represented by a K-cycle,
(π, T,H ⊗ B), where
• H is a separable Hilbert space;
• the right Hilbert B-module H ⊗ B is acted upon by Γ;
• π is an equivariant representation of A in the algebra L(H⊗B) of ad-
jointable operators on H ⊗ B;
• T is a self-adjoint operator on H⊗ B satisfying the K-cycle conditions, i.e.
[T, π(a)], π(a)(T 2 −IdH⊗B) and π(a)(γ(T )− T ) belongs to K(H)⊗ B, for
every a in A and γ ∈ Γ.
Let TΓ = T ⊗B IdB⋊redΓ be the adjointable element of (H ⊗ B) ⊗B B⋊redΓ ∼=
H ⊗ B⋊redΓ induced by T and let πΓ be the representation of A⋊redΓ in the
algebra L(H⊗B⋊redΓ) of adjointable operators of H⊗B⋊redΓ induced by π. Then
(πΓ, TΓ,H ⊗ B⋊redΓ) is a A⋊redΓ-B⋊redΓ-K-cycle and the Kasparov transform
Γ (z) of this K-cycle in KK∗(A⋊redΓ, B⋊redΓ). In the odd
[11] of z is the class J red
case, let us set P = IdH⊗B+T
. Then P induces an adjointable operator PΓ =
P ⊗B IdB⋊redΓ of (H ⊗ B) ⊗B B⋊redΓ ∼= H ⊗ B⋊redΓ. Let us define
E(π,T ) = {(x, y) ∈ A⋊redΓ⊕L(H⊗B⋊redΓ) such that PΓ·πΓ(x)·PΓ−y ∈ K(H)⊗B⋊redΓ}.
Since PΓ has no propagation, the C∗-algebra E(π,T ) is filtered by (E(π,T )
E(π,T )
)r>0 with
= {(x, PΓ · πΓ(x) · PΓ + y); x ∈ (A⋊redΓ)r and y ∈ K(H) ⊗ (B⋊redΓ)r}.
2
r
r
The extension of C∗-algebras
0 −→ K(H) ⊗ B⋊redΓ −→ E(π,T ) −→ A⋊redΓ −→ 0
is filtered semi-split by the cross-section
s : A⋊redΓ → E(π,T ); x 7→ (x, PΓ · πΓ(x) · PΓ).
Let us show that DK(H)⊗B⋊redΓ,E(π,T ) only depends on the class of (π, T,H⊗ B) in
KK Γ
1 (A, B). Assume that (π, T,H ⊗ B[0, 1]) is a Γ-equivariant A-B[0, 1]-K-cycle
42
H. OYONO-OYONO AND G. YU
providing a homotopy between two Γ-equivariant A-B-K-cycles (π0, T0,H⊗ B) and
(π1, T1,H ⊗ B). For t ∈ [0, 1] we denote by
• et : B[0, 1]⋊redΓ → B⋊redΓ the evaluation at t;
• Ft ∈ L(H⊗B⋊redΓ) the fiber at t of an operator F ∈ L(H⊗B[0, 1]⋊redΓ);
• πΓ,t the representation of A ⋊red Γ induced by πΓ at the fiber t;
• st : A ⋊red Γ → E(πt,Tt); x 7→ (x, PΓ,t · πΓ,t · PΓ,t)
(with P = T +1
2 );
Then the homomorphism E(π,T ) → E(πt,Tt); (x, y) 7→ (x, yt) satisfies the con-
ditions of remark 3.5 (with s : A⋊redΓ → E(π,T ); x 7→ (x, PΓ · πΓ(x) · PΓ) and
st : A ⋊red Γ → E(πt,Tt)) and thus we get that
(IdK(H) ⊗ et)∗ ◦ DK(H)⊗B[0,1]⋊redΓ,E(π,T ) = DK(H)⊗B⋊redΓ,E(πt ,Tt),
and according to lemma 1.27, we deduce that
DK(H)⊗B⋊redΓ,E(π0,T0) = DK(H)⊗B⋊redΓ,E(π1,T1) .
This shows that for a Γ-equivariant A-B-K-cycles (π, T,H⊗B), then DK(H)⊗B⋊redΓ,E(π,T )
depends only on the class z of (π, T,H⊗ B) in KK Γ
1 (A, B). Eventually, if we define
Γ (z) = M−1
J red
B⋊redΓ ◦ DK(H)⊗B⋊redΓ,E(π,T ),
where
• (π, T,H ⊗ B) is any Γ-equivariant A-B-K-cycles representing z;
• MB⋊redΓ is the Morita equivalence (see example 2.2).
we get as in section 4.1
Proposition 5.2. Let A and B be Γ-C∗-algebras. Then for any element z of
KK Γ
1 (A, B), there is a odd degree (αD, kD)-controlled morphism
Γ (z) = (J red,ε,r
J red
Γ
(z))0<ε< 1
4αD
,r>0 : K∗(A ⋊red Γ) → K∗(B⋊redΓ)
such that
(i) J red
(ii) J red
Γ (x) induces in K-theory the right multiplication by J red
Γ
Γ (z);
is additive, i.e
J red
Γ (z + z′) = J red
Γ (z) + J red
Γ (z′).
(iii) Let A′ be a Γ-C∗-algebra and let f : A → A′ be a homomorphism Γ-C∗-
algebras, then
for any z in KK Γ
Γ (f∗(z)) = J red
J red
1 (A′, B).
Γ (z) ◦ fΓ,red,∗
(iv) Let B′ be a Γ-C∗-algebra and let g : B → B′ be a homomorphism of
Γ-C∗-algebras, then
J red
Γ (g∗(z)) = gΓ,red,∗ ◦ J red
1 (A, B).
for any z in KK Γ
Γ (z)
(v) If
0 → J → A → A/J → 0
is a semi-split exact sequence of Γ-C∗-algebras, let [∂J,A] be the element of
KK Γ
1 (A/J, J) that implements the boundary map ∂J,A. Then we have
J red
Γ ([∂J,A]) = DJ ⋊redΓ,A⋊redΓ.
ON A QUANTITATIVE OPERATOR K-THEORY
43
We can now define J red
and kJ = kT ∗kD. If A and B are Γ-C∗-algebra and if z is an element in KK Γ
then we set with notation of section 4.1
for even element in the following way. Set αJ = αT αD
0 (A, B),
Γ
Γ (z) = (J red,ε,r
J red
Γ
(z))0<ε< 1
4αT
,r
def==TB⋊redΓ([∂]−1) ◦ J red
Γ (z ⊗B [∂SB]).
According to theorem 4.6, there exists a control pair (λ, h) such that for any Γ-C∗-
algebra A, then J red
we can assume indeed that J red
morphism. As for theorem 4.4, we get.
∼ IdK∗(A⋊redΓ). Up to compose with ιαD ε,αJ ε,kD,εr,kJ ,εr
Γ (•) is also, in the odd case a (αJ , kJ )-controlled
Γ ([IdA])
(λ,h)
∗
,
Theorem 5.3. Let A and B be Γ-C∗-algebras.
∗ (A, B), then
(i) For any element z of KK Γ
J red
Γ (z) : K∗(A ⋊red Γ) → K∗(B⋊redΓ)
is a (αJ , kJ )-controlled morphism of same degree as z that induces in
K-theory right multiplication by J red
∗ (A, B), then
(ii) For any z and z′ in KK Γ
Γ (z).
J red
Γ (z + z′) = J red
Γ (z) + J red
Γ (z′).
and any z in KK Γ
(iii) For any Γ-C∗-algebra A′, any homomorphism f : A → A′ of Γ-C∗-algebras
(iv) For any Γ-C∗-algebra B′, any homomorphism g : B → B′ of Γ-C∗-algebras
Γ (f∗(z)) = J red
Γ (g∗(z)) = gΓ,∗ ◦ J red
Γ (z) ◦ fΓ,∗.
Γ (z).
∗ (A′, B), then J red
∗ (A, B), then J red
and any z in KK Γ
Using the same argument as in the proof of theorem 4.5, we see that J red
Γ
is
compatible with Kasparov products.
Theorem 5.4. There exists a control pair (λ, h) such that the following holds:
for every Γ-C∗-algebras A, B and D, any elements z in KK Γ
∗ (A, B) and z′ in
KK Γ
∗ (B, D), then
J red
Γ (z ⊗B z′)
(λ,h)
∼ J red
Γ (z′) ◦ J red
Γ (z).
We can perform a similar construction for maximal cross products.
Theorem 5.5. Let A and B be Γ-C∗-algebras.
(i) For any element z of KK Γ
∗ (A, B), there exists a (αJ , kJ )-controlled mor-
Γ
(z) = (J max,ε,r
phism
J max
with same degree as z that induces in K-theory right multiplication by
J max
Γ
,r : K∗(A⋊maxΓ) → K∗(B⋊maxΓ)
(z))0<ε< 1
4αJ
Γ
(ii) For any z and z′ in KK Γ
(z) and such that λΓ,B,∗ ◦ J max
∗ (A, B), then
(z + z′) = J max
(z) = J red
(z) + J max
J max
Γ
Γ
Γ
Γ
(z′).
Γ (z) ◦ λΓ,A,∗.
and any z in KK Γ
(iii) For any Γ-C∗-algebra A′, any homomorphism f : A → A′ of Γ-C∗-algebras
(iv) For any Γ-C∗-algebra B′, any homomorphism g : B → B′ of Γ-C∗-algebras
(f∗(z)) = J max
(g∗(z)) = gΓ,max,∗ ◦ J max
(z) ◦ fΓ,max,∗.
(z).
∗ (A′, B), then J max
∗ (A, B), then J max
and any z in KK Γ
Γ
Γ
Γ
Γ
Moreover, there exists a controlled pair (λ, h) such that,
• for any Γ algebra A, then J max
Γ
([IdA])
(λ,h)
∼ IdK∗(A⋊maxΓ);
44
H. OYONO-OYONO AND G. YU
• For any semi-split extension of Γ algebras 0 → J → A → A/J → 0, then
J max
Γ
([∂J,A])
(λ,h)
∼ DJ,A.
Theorem 5.6. There exists a control pair (λ, h) such that the following holds:
for every Γ-C∗-algebras A, B and D, any elements z in KK Γ
∗ (A, B) and z′ in
KK Γ
∗ (B, D), then
J max
Γ
(z ⊗B z′)
(λ,h)
∼ J max
Γ
(z′) ◦ J max
Γ
(z).
5.3. Application to K-amenability. The original definition of K-amenability is
due to J. Cuntz [6]. For our purpose, it is more convenient to use the equivalent
definition given by P. Julg and A. Valette in [10]. If Γ is a discrete group, let us
0 (C, C) of the K-cycle (IdC, 0, C), where C is provided
denote by 1Γ the class in KK Γ
with the trivial action on Γ.
Definition 5.7. Let Γ be a discrete group. Then Γ is K-amenable if 1Γ can be
represented by a K-cycle such that the action of Γ on the underlying Hilbert space
is weakly contained in the regular representation.
(The previous definition indeed also makes sense for locally compact groups.)
Example 5.8. Amenable groups are obviously K-amenable. Typical example on
non-amenable K-amenable groups are free groups [6]. More generally, J. L. Tu
proved in [17] that group which satisfies the strong Baum-Connes conjecture (i.e
with γ = 1) are K-amenable. Examples of such group are groups with the Haagerup
property [8] and fundamental groups of compact and oriented 3-manifolds [13].
For a Γ-C∗-algebra B and an element T of L(H⊗B), where H is a separable
Hilbert space, let us set TΓ,max = T⊗BIdB⋊maxΓ and TΓ,red = T⊗BIdB⋊redΓ. If
A is a Γ-C∗-algebra and π : A → L(H⊗B) is a Γ-equivariant representation, let
πΓ,red : A⋊redΓ → L(H⊗B⋊redΓ) and πΓ,max : A⋊maxΓ → L(H⊗B⋊maxΓ) be
respectively the reduced and the maximal representation induced by π. Then, we
have the following (compare with the proof of [10, proposition 3.4]).
Proposition 5.9. Let Γ be a K-amenable discrete group and let A and B be Γ-
C∗-algebras. Then any elements of KK Γ
∗ (A, B) can be represented by a K-cycle
(π, T,H⊗B) such that the homomorphism πΓ,max : A⋊maxΓ → L(H⊗B⋊maxΓ)
factorises through the homomorphism λΓ,A : A⋊maxΓ → A⋊redΓ, i.e there exists a
homomorphism
such that
πΓ,red,max : A⋊redΓ → L(H⊗B⋊maxΓ)
As a consequence, for any Γ-C∗-algebra A, then
πΓ,max = πΓ,red,max ◦ λΓ,A.
is an isomorphism [6].
λΓ,A,∗ : K∗(A⋊maxΓ) → K∗(A⋊redΓ)
We have the following analogous result for quantitative K-theory.
Theorem 5.10. There exists a control pair (λ, h) such that
λΓ,A,∗ : K∗(A⋊maxΓ) → K∗(A⋊redΓ)
is a (λ, h)-isomorphism for every Γ-C∗-algebra A.
ON A QUANTITATIVE OPERATOR K-THEORY
45
Proof. Let (π, T,H⊗SA) be a Γ-equivariant K-cycle as in proposition 5.9 repre-
senting the element [∂A] of KK Γ
1 (A, SA) corresponding to the extension
0 → SA → CA → A → 0.
Let then choose πΓ,A,red,max : A⋊redΓ → L(H⊗B⋊maxΓ) such that πΓ,max =
πΓ,red,max ◦ λΓ,A. Let us set P = T +IdH⊗SA
and then define
2
E(π,T )
red = {(x, y) ∈ A⋊redΓ ⊕ L(H ⊗ SA⋊redΓ) such that
PΓ,red · πΓ,red(x) · PΓ,red − y ∈ K(H) ⊗ SA⋊redΓ},
E(π,T )
max = {(x, y) ∈ A⋊maxΓ ⊕ L(H ⊗ SA⋊maxΓ) such that
PΓ,max · πΓ,max(x) · PΓ,max − y ∈ K(H) ⊗ SA⋊maxΓ}
and
E(π,T )
red,max = {(x, y) ∈A⋊redΓ ⊕ L(H ⊗ SA⋊maxΓ) such that
PΓ,max · πΓ,red,max(x) · PΓ,max − y ∈ K(H) ⊗ A⋊maxΓ}
Then E(π,T )
red
, E(π,T )
max and E(π,T )
red,max are respectively filtered by
{(x, PΓ,red · πΓ,red(x) · PΓ,red + y); x ∈ A⋊redΓr and y ∈ K(H) ⊗ SA⋊redΓr},
{(x, PΓ,max · πΓ,max(x) · PΓ,max + y); x ∈ SA⋊maxΓr and y ∈ K(H) ⊗ SA⋊maxΓr}
and
{(x, PΓ,max·πΓ,red,max(x)·PΓ,max + y); x ∈ A⋊redΓr and y ∈ K(H)⊗ SA⋊maxΓr}.
Moreover, the extension of C∗-algebras
0 −→ K(H) ⊗ SA⋊redΓ −→ E(π,T )
0 −→ K(H) ⊗ SA⋊maxΓ −→ E(π,T )
red −→ A⋊redΓ −→ 0,
max −→ A⋊maxΓ −→ 0
and
0 −→ K(H) ⊗ SA⋊maxΓ −→ E(π,T )
red,max −→ A⋊redΓ −→ 0
provided by the projection on the first factor are respectively semi-split by the
filtered cross-sections
sred : A⋊redΓ → E(π,T )
smax : A⋊maxΓ → E(π,T )
red
; x 7→ (x, PΓ,red · πΓ,red(x) · PΓ,red),
max ; x 7→ (x, PΓ,max · πΓ,max(x) · PΓ,max)
and
sred,max : A⋊redΓ → E(π,T )
max ; x 7→ (x, PΓ,max · πΓ,red,max(x) · PΓ,max).
Let us set
and
f1 : E(π,T )
max → E(π,T )
red,max : (x, y) 7→ (λΓ,A,∗(x), y)
f2 : E(π,T )
red,max → E(π,T )
red
: (x, y) 7→ (x, y⊗A⋊maxΓIdA⋊redΓ).
=y
λΓ,K(H)⊗SAy
f1y
f2y
yλΓ,A
y=
46
H. OYONO-OYONO AND G. YU
The the three above extensions fit in a commutative diagram
0 −−−−→ K(H) ⊗ SA⋊maxΓ −−−−→ E(π,T )
max −−−−→ A⋊maxΓ −−−−→ 0
0 −−−−→ K(H) ⊗ SA⋊maxΓ −−−−→ E(π,T )
red,max −−−−→ A⋊redΓ −−−−→ 0
0 −−−−→ K(H) ⊗ SA⋊redΓ −−−−→ E(π,T )
red
−−−−→ A⋊redΓ −−−−→ 0
which satisfy the conditions of remark 3.5 relatively to sred, smax and sred,max, and
hence we deduce
(9)
and
(10)
DK(H)⊗SA⋊maxΓ,E(π,T )
red,max ◦ λA,Γ,∗ = DK(H)⊗SA⋊maxΓ,E(π,T )
max
λK(H)⊗SA,Γ,∗ ◦ DK(H)⊗SA⋊maxΓ,E(π,T )
red,max
= DK(H)⊗SA⋊redΓ,E(π,T )
red
Let us set then
D′A = M−1
SA⋊maxΓ ◦ DSA⋊maxΓ,E(π,T )
red,max
: K∗(A⋊redΓ) → K∗(SA⋊maxΓ).
Since we have by definition of the quantitative Kasparov transformation the equal-
ities
JΓ,red([∂A]) = M−1
SA⋊redΓ ◦ DSA⋊redΓ,E(π,T )
red
and
JΓ,max([∂A]) = M−1
SA⋊maxΓ ◦ DSA⋊maxΓ,E(π,T )
max
,
we deduce by using equations (9) and (10), theorems 5.3, 5.5, 5.4 and 5.6 and
naturality of Morita equivalence, that there exists a control pair (λ, h) such that
JΓ,max([∂A]−1) ◦ D′A is a (α, h)-inverse for λΓ,A,∗.
(cid:3)
6. The quantitative Baum-Connes conjecture
In this section, we formulate a quantitative version for the Baum-Connes con-
jecture and we prove it for a large class of groups.
6.1. The Rips complex. Let Γ be a finitely generated group equipped with a
lenght ℓ arising from a finite and symmetric generating set. Recall that for any
positive number d, then the d-Rips complex Pd(Γ) is the set of finitely supported
probability measures on Γ with support of diameter less than d for the distance
induced by ℓ. We equip Pd(Γ) with the distance induced by the norm khk =
sup{kh(γ)k; γ ∈ Γ} for h ∈ C0(Γ, C). Since ℓ is a proper function, i.e. B(e, r)
is finite for every positive number r, we see that Pd(Γ) is a finite dimension and
locally finite simplicial complexe and the action of Γ by left translations is simplicial,
proper and cocompact. Let us denote by
• Vd(Γ) the closed subset of elements of Pd(Γ) with support in B(e, d).
• Wd(Γ) the closed subset of elements of Pd(Γ) with support in B(e, 2d);
Then Vd(Γ) is a compact subset of Wd(Γ) and contains a fundamental domain for
the action of Γ on Pd(Γ).
Lemma 6.1. The compact Vd(Γ) is contained in the interior of Wd(Γ).
ON A QUANTITATIVE OPERATOR K-THEORY
47
Proof. Let h be an element in Vd(Γ) and choose an element γ in B(e, d) such that
h(γ) > 0. Then if g is an element of Pd(Γ) such that kg − hk < h(γ), we get that
g(γ) 6= 0 and thus every element γ′ of the support of g satisfies ℓ(γ−1γ′) < d. Hence
g belongs to Wd(Γ).
(cid:3)
Lemma 6.2. There is a continuous function φ : Pd(Γ) → [0, 1] compactly supported
in Wd(Γ) such that
γ(φ) = 1.
Xγ∈Γ
Proof. Let ψ : Pd(Γ) → [0, 1] a continuous function compactly supported in the
interior of Wd(Γ) and such that ψ(x) = 1 if x belongs to Vd(Γ). Since Vd(Γ) contains
a fundamental domain for the action of Γ on Pd(Γ), we get that Pγ∈Γ ψ(γx) > 0
for all x in Pd(Γ) (notice that the sum Pγ∈Γ ψ(γx) is locally finite). We define
ψ(x)
then φ(x) =
Pγ∈Γ ψ(γx) for any x in Pd(Γ).
(cid:3)
Let us define sΓ,d as the cardinality of the finite set
Then for any function φ as in lemma 6.2, the function
{γ ∈ Γ such that γWd(Γ) ∩ Wd(Γ) 6= ∅}.
eφ : Γ → C0(Pd(Γ)); γ 7→Xγ∈Γ
φ1/2γ(φ1/2)
is a projection of C0(Pd(Γ))⋊redΓ with propagation less than sΓ,d. Moreover, since
the set of function satisfying the condition of lemma 6.2 is an affine space, we get
that for any positive number ε and r with ε < 1/4 and r > sΓ,d, the class
[eφ, 0]ε,r ∈ K ε,r
0 (C0(Pd(Γ))⋊redΓ)
does not depend on the chosen function φ. Let us set then rΓ,d,ε = kJ ,ε/αJ sΓ,d.
Recall that kJ can be chosen non increasing and in this case, rΓ,d,ε is non decreasing
in d and non increasing in ε.
Definition 6.3. For any Γ-C∗-algebra A and any positive numbers ε, r and d with
ε < 1/4 and r > rΓ,d,ε, we define the quantitative assembly map
µε,r,d
Γ,A,∗
: KK Γ
∗ (C0(Pd(Γ)), A) → K ε,r
7→ (cid:0)J
z
Γ
∗ (A ⋊red Γ)
red, ε
αJ
,
r
kJ,ε/αJ
(z)(cid:1)(cid:18)[eφ, 0] ε
αJ
kJ,ε/αJ (cid:19) .
r
,
Then according to theorem 5.3, the map µε,r,d
Γ,A is a homomorphism of groups
(resp. semi-groups) in even (resp. odd) degree. For any positive numbers d and d′
such that d 6 d′, we denote by qd,d′ : C0(Pd′(Γ)) → C0(Pd(Γ)) the homomorphism
induced by the restriction from Pd′ (Γ) to Pd(Γ).
It is straightforward to check
that if d, d′ and r are positive numbers such that d 6 d′ and r > rΓ,d′,ε, then
Γ,A = µε,r,d′
µε,r,d
Γ,A ◦ qd,d′,∗. Moreover, for every positive numbers ε, ε′, d, r and r′ such
that ε 6 ε′ 6 1/4, rΓ,d,ε 6 r, rΓ,d,ε′ 6 r′, and r < r′, we get by definition of a
controlled morphism that
(11)
ιε,ε′,r,r′
∗
◦ µε,r,d
Γ,A,∗
= µε′,r′,d
Γ,A,∗
.
48
H. OYONO-OYONO AND G. YU
Furthermore, the quantitative assembly maps are natural in the Γ-C∗-algebra, i.e.
if A and B are Γ-C∗-algebras and if φ : A → B is a Γ-equivariant homomorphism,
then
φΓ,red,∗,ε,r ◦ µε,r,d
Γ,A,∗
= µε,r,d
Γ,B,∗ ◦ φ∗
for every positive numbers r and ε with r > rΓ,d,ε and ε < 1/4. These quantitative
assembly maps are related to the usual assembly maps in the following way: recall
from [2] that there is a bunch of assembly maps with coefficients in a Γ-C∗-algebra
A defined by
µd
Γ,A,∗ : KK Γ
∗ (C0(Pd(Γ)), A) → K∗(A ⋊red Γ)
z
7→ [eφ] ⊗C0(Pd(Γ))⋊Γ JΓ(z).
For every positive numbers r and ε with r > rΓ,d,ε and ε < 1/4, we have
(12)
∗ ◦ µε,r,d
ιε,r
Γ,A,∗
Γ,A,∗ ◦ qd,d′,∗ = µd
Γ,A,∗ for all positive numbers d and d′ with
d 6 d′, the family of assembly maps (µd
Γ,A)d>0 gives rise to a homomorphism
Recall that since µd′
Γ,A,∗.
= µd
µΓ,A,∗ : lim
d>0
KK Γ
∗ (C0(Pd(Γ)), A) −→ K∗(A ⋊red Γ)
called the Baum-Connes assembly map.
6.2. Quantitative statements. Let us consider for a Γ-C∗-algebra A and positive
numbers d, d′, r, r′, ε and ε′ with d 6 d′, ε′ 6 ε < 1/4, rΓ,d,ε 6 r and r′ 6 r the
following statements:
QIΓ,A,∗(d, d′, r, ε): for any element x in KK Γ
0 in K ε,r
∗ (A ⋊red Γ) implies that q∗d,d′(x) = 0 in KK Γ
∗ (C0(Pd(Γ)), A), then µε,r,d
Γ,A,∗
∗ (C0(Pd′ (Γ)), A).
(x) =
(A ⋊red Γ), there exists an element
QSΓ,A,∗(d, r, r′, ε, ε′): for every y in K ε′,r′
∗
x in KK Γ
∗ (C0(Pd(Γ)), A) such that
∗
Using equation (12) and remark 1.18 we get
µε,r,d
Γ,A,∗
(x) = ιε′,ε,r′,r
(y).
Proposition 6.4. Assume that for all positive number d there exists a positive
number ε with ε < 1/4 for which the following holds:
for any positive number r with r > rΓ,d,ε, there exists a positive number d′ with
d′ > d such that QIΓ,A(d, d′, r, ε) is satisfied.
Then µΓ,A,∗ is one-to-one.
We can also easily prove the following:
Proposition 6.5. Assume that there exists a positive number ε′ with ε′ < 1/4 such
that the following holds:
for any positive number r′ , there exist positive numbers ε, d and r with ε′ 6 ε <
1/4, rΓ,d,ε 6 r and r′ 6 r such that QSΓ,A(d, r, r′, ε, ε′) is true.
Then µΓ,A,∗ is onto.
The following results provide numerous examples of finitely generated groups
that satisfy the quantitative statements.
ON A QUANTITATIVE OPERATOR K-THEORY
49
Theorem 6.6. Let A be a Γ-C∗-algebra. Then the following assertions are equiv-
alent:
(i) µΓ,ℓ∞(N,K(H)⊗A),∗ is one-to-one,
(ii) For any positive numbers d, ε and r > rΓ,d,ε with ε < 1/4 and r > rΓ,d,
there exists a positive number d′ with d′ > d for which QIΓ,A(d, d′, r, ε) is
satisfied.
Proof. Assume that condition (ii) holds.
Let x be an element in some KK Γ
∗ (C0(Pd(Γ)), ℓ∞(N,K(H) ⊗ A)) such that
µd
Γ,ℓ∞(N,K(H)⊗A),∗(x) = 0.
Using equation (12), we get that ιε′,r′
(x)) = 0 for any ε′ in (0, 1/4) and
r′ > rΓ,d,ε′ and hence, by remark 1.18, we can find ε and r > rΓ,d,ε such that
µε,r,d
(x) = 0. Recall from [14, Proposition 3.4] that we have an iso-
Γ,ℓ∞(N,K(H)⊗A),∗
morphism
(µε′,r′,d
Γ,A,∗
∗
(13)
KK Γ
0 (C0(Pd(Γ)), ℓ∞(N,K(H) ⊗ A)) ∼=−→ KK Γ
0 (C0(Pd(Γ)), A)N
induced on the j th factor and up to the Morita equivalence
KK Γ
0 (C0(Pd(Γ)), A) ∼= KK Γ
0 (C0(Pd(Γ)),K(H) ⊗ A)
by the j th projection ℓ∞(N,K(H) ⊗ A) → K(H) ⊗ A. Let (xi)i∈N be the ele-
ment of KK Γ
0 (C0(Pd(Γ)), A)N corresponding to x under this identification and let
d′ > d be a number such that QIΓ,A(d, d′, r, ε) holds. Naturality of the quanti-
tative assembly maps implies that µε,r,d
(xi) = 0 and hence that qd,d′,∗(xi) = 0
Γ,A,∗
in KK Γ
∗ (C0(Pd′ (Γ)), A) for every integer i. Using once again the isomorphism of
∗ (C0(Pd′ (Γ)), ℓ∞(N,K(H)⊗ A) and
equation (13), we get that qd,d′,∗(x) = 0 in KK Γ
hence µΓ,ℓ∞(N,K(H)⊗A),∗ is one-to-one.
Let us prove the converse in the even case, the odd case being similar. As-
sume that there exists positive numbers d, ε and r with ε < 1/4 and r > rΓ,d,ε
and such that for all d′ > d, the condition QIΓ,A(d, d′, r, ε) does not hold. Let us
prove that µΓ,ℓ∞(N,K(H)⊗A),∗ is not one-to-one. Let (di)i∈N be an increasing and
unbounded sequence of positive numbers such that di > d for all integer i. For all
integer i, let xi be an element in KK Γ
(xi) = 0 in
K0(A ⋊red Γ) and qd,di,∗(xi) 6= 0 in KK Γ
0 (C0(Pdi (Γ)), A). Let x be the element of
0 (C0(Pd(Γ)), ℓ∞(N,K(H) ⊗ A)) corresponding to (xi)i∈N under the identifica-
KK Γ
tion of equation (13). Let (pi)i∈N be a family of ε-r-projections, with pi in some
Mli( ^A ⋊red Γ) and n an integer such that
0 (C0(Pd(Γ)), A) such that µε,r,d
Γ,A,∗
µε,r,d
Γ,ℓ∞(N,K(H)⊗A),∗
(x) = [(pi)i∈N, n]ε,r
0 (ℓ∞(N,K(H) ⊗ A)⋊redΓ). By naturality of µε,r,d
in K ε,r
, we get that [pi, n]ε,r = 0
Γ,•,∗
in K ε,r
0 (A ⋊red Γ) for all integer i. We see by using proposition 1.31 that then
ιε,r
∗ ([(pi)i∈N, n]) = 0 in K0(ℓ∞(N,K(H) ⊗ A)⋊redΓ). We eventually obtain that
Γ,A(x) = ιε,r
µd
Γ,A (x) = 0. Since qd,di,∗(x) 6= 0 for every integer i, we get that
µΓ,ℓ∞(N,K(H)⊗A),∗ is not one-to-one.
Theorem 6.7. There exists λ > 1 such that for any Γ-C∗-algebra, the following
assertions are equivalent:
∗ ◦ µε,r,d
(cid:3)
(i) µΓ,ℓ∞(N,K(H)⊗A),∗ is onto;
50
H. OYONO-OYONO AND G. YU
(ii) For any positive numbers ε and r′ with ε < 1
4λ , there exist positive num-
bers d and r with rΓ,d,ε 6 r and r′ 6 r for which QSΓ,A(d, r, r′, λε, ε) is
satisfied.
∗
(y) = z, with 0 < ε < 1
Proof. Choose λ as in remark 1.18. Assume that condition (ii) holds. Let z be an el-
(ℓ∞(N,K(H)⊗
4λ and r′ > 0. Let yi be the image
ement in K∗(ℓ∞(N,K(H)⊗A)⋊redΓ) and let y be an element in K ε,r′
∗
A)⋊redΓ) such that ιε,r′
of y under the composition
(14) K ε,r′
(A ⋊red Γ),
where the first map is induced by the evaluation ℓ∞(N,K(H) ⊗ A) −→ K(H) ⊗ A
at i and the second map is the Morita equivalence of proposition 1.29. Let d
and r be numbers with r > r′ and r > rΓ,d,ε and such that QSΓ,A(d, r, r′, λε, ε)
holds. Then for any integer i, there exists a xi in KK Γ
∗ (C0(Pd(Γ)), A) such that
µλε,r,d
Γ,A,∗
(ℓ∞(N,K(H)⊗A)⋊redΓ) → K ε,r′
∗
(K(H)⊗A ⋊red Γ) ∼=→ K ε,r′
∗
(xi) = ιε,λε,r′,r
∗
∗
(yi) in K ε,r
x ∈ KK Γ
∗ (A⋊redΓ). Let
∗ (C0(Pd(Γ)), ℓ∞(N,K(H) ⊗ A))
be the element corresponding to (xi)i∈N under the identification of equation (13).
By naturality of the quantitative assembly maps, we get according to proposition
1.31 and up to replace λ by 3λ (for the odd case) that
(x) = ιε,λε,r′,r
(y)
in K ε,r
µλε,r,d
Γ,ℓ∞(N,K(H)⊗A)),∗
∗
∗ (ℓ∞(N,K(H) ⊗ A)⋊redΓ). We have hence
Γ,ℓ∞(N,K(H)⊗A)),∗(x) = ιε,r′
µd
∗
(y) = z,
and therefore µΓ,ℓ∞(N,K(H)⊗A),∗ is onto.
Let us prove the converse in the even case, the odd case being similar. Assume
that there exist positive numbers ε and r′ with ε < 1
4λ such that for all positive
numbers r and d with r > r′ and r > rΓ,d,ε, then QSΓ,A(d, r, r′, λε, ε) does not
hold. Let us prove then that µΓ,ℓ∞(N,K(H)⊗A),∗ is not onto. Let (di)i∈N and (ri)i∈N
be increasing and unbounded sequences of positive numbers such that ri > rΓ,di,λε
and ri > r′. Let yi be an element in K ε,r′
(yi) is not
in the range of µλε,ri,di
(ℓ∞(N,K(H)⊗ A)⋊redΓ)
such that for every integer i, the image of y under the composition of equation (14)
0 (C0(Pd′(Γ)), ℓ∞(N,K(H)⊗ A))
is yi. Assume that for some d′, there is an x in KK Γ
such that ιε,r′
(x). Using remark 1.18, we see that there
exists a positive number r with r′ 6 r and rΓ,d′,λε 6 r and such that
. There exists an element y in K ε,r′
(A ⋊red Γ) such that ιε,λε,r′,ri
Γ,ℓ∞(N,K(H)⊗A),∗
(y) = µd′
Γ,A,∗
∗
∗
0
0
ιε,λε,r′,r
∗
◦ µε,r′,d′
Γ,ℓ∞(N,K(H)⊗A),∗
(x) = ιε,λε,r′,r
∗
(y).
But then, if we choose i such that ri > r and di > d′ we get by using naturality
of the assembly map and equation (11) that ιε,λε,r′,ri
(yi) belongs to the image of
µλε,ri,di
Γ,A,∗
, which contradicts our assumption.
∗
(cid:3)
Replacing in the proof of (ii) implies (i) of theorems 6.6 and 6.7 the algebra
ℓ∞(N,K(H) ⊗ A) by Qi∈N(K(H) ⊗ Ai) for a family (Ai)i∈N of Γ-C∗-algebras, we
can prove the following result.
ON A QUANTITATIVE OPERATOR K-THEORY
51
Theorem 6.8. Let Γ be a discrete group.
(i) Assume that for any Γ-C∗-algebra A, the assembly map µΓ,A,∗ is one-to-
one. Then for any positive numbers d, ε and r > rΓ,d,ε with ε < 1/4
and r > rΓ,d, there exists a positive number d′ with d′ > d such that
QIΓ,A(d, d′, r, ε) is satisfied for every Γ-C∗-algebra A;
(ii) Assume that for any Γ-C∗-algebra A, the assembly map µΓ,A,∗ is onto.
Then for some λ > 1 and for any positive numbers ε and r′ with ε < 1
4λ ,
there exist positive numbers d and r with rΓ,d,ε 6 r and r′ 6 r such that
QSΓ,A(d, r, r′, λε, ε) is satisfied for every Γ-C∗-algebra A.
In particular, if Γ satisfies the Baum-Connes conjecture with coefficients, then Γ
satisfies points (i) and (ii) above.
Recall from [16, 20] that if Γ coarsely embeds in a Hilbert space, then µΓ,A,∗ is
one-to-one for every Γ-C∗-algebra A. Hence we get:
Corollary 6.9. If Γ coarsely embeds in a Hilbert space, then for any positive num-
bers d, ε and r > rΓ,d,ε with ε < 1/4 and r > rΓ,d, there exists a positive number
d′ with d′ > d such that QIΓ,A(d, d′, r, ε) is satisfied for every Γ-C∗-algebra A;
The quantitative assembly maps admit maximal versions defined with notations
of definition 6.3 for any Γ-C∗-algebra A and any positive number ε, r and d with
ε < 1/4 and r > rΓ,d,ε, as
µε,r,d
Γ,A,max,∗
∗ (A⋊maxΓ)
: KK Γ
∗ (C0(Pd(Γ)), A) → K ε,r
7→ (cid:0)J
z
Γ
max, ε
αJ
,
r
kJ,ε/αJ
(z)(cid:1)(cid:18)[eφ, 0] ε
αJ
kJ,ε/αJ (cid:19) .
r
,
As in the reduced case, we have using the same notations
• for any positive number d and d′ such that d 6 d′, then
µε,r,d
Γ,A,max,∗
= µε,r,d′
Γ,A,max,∗ ◦ qd,d′,∗.
• for every positive numbers ε, ε′, d, r and r′ such that ε 6 ε′ 6 1/4, rΓ,d,ε 6
r, rΓ,d,ε′ 6 r′, and r < r′, then
◦ µε,r,d
ιε,ε′,r,r′
∗
Γ,A,max,∗
= µε′,r′,d
Γ,A,max,∗
.
• the maximal quantitative assembly maps are natural in the Γ-C∗-algebras.
Moreover, by theorem 5.5(i), the maximal quantitative assembly maps are compat-
ible with the reduced ones, i.e µε,r,d
. The surjectivity of the
Γ,A,max,∗
Γ,A,∗
Baum-Connes assembly map µΓ,A,∗ implies that the map
Γ,A,∗ ◦ µε,r,d
= λε,r
is onto. We have a similar statement in the setting of quantitative K-theory.
λΓ,A,∗ : K∗(A⋊maxΓ) → K∗(A⋊redΓ)
Theorem 6.10. There exists λ > 1 such the following holds : let Γ be a discrete
group and assume that for any Γ-C∗-algebra A, the assembly map µΓ,A,∗ is onto.
Then for any positive numbers ε and r, with ε < 1
4λ , there exists a positive number
r′ with r′ > r such that
• for any Γ-C∗-algebra A;
• for any x in K ε,r
there exists y in K λε,r′
∗
∗ (A⋊redΓ),
(A⋊maxΓ) such that λλε,r′
Γ,A,∗
(y) = ιε,λε,r,r′
∗
(x).
52
H. OYONO-OYONO AND G. YU
7. Further comments
The definition of quantitative K-theory can be extended to the framework of
filtered Banach algebras, i.e. Banach algebra A equipped with a family (Ar)r>0 of
linear subspaces indexed by positive numbers such that:
• Ar ⊂ Ar′ if r 6 r′;
• Ar · Ar′ ⊂ Ar+r′ ;
• the subalgebra [r>0
Ar is dense in A.
Since we no more have an involution, we need to introduce instead a norm control
for almost idempotents. Let ε be in (0, 1/4) and let r and N be positive numbers.
An element e of A is an ε-r-N -idempotent if
• e is in Ar;
• ke2 − ek < ε;
• kek < N ;
Similarly, if A is a unital, an element x in A is called ε-r-N -invertible if
• x is in Ar;
• kxk < N ;
• there exists an element y in Ar such that kyk < N , kxy − 1k < ε and
kyx − 1k < ε.
Quantitative K-theory can then be defined in the setting of ε-r-N -idempotents
and of ε-r-N -invertibles. We obtain in this way a bunch of abelian semi-groups
(K ε,r,N
∗
(A))ε∈(0,1/4),r>,N >1. Let us set for a fixed N > 1
(A))ε∈(0,1/4),r>0.
KN
∗ (A) = (K ε,r,N
∗
If A is a filtered C∗-algebra and e an ε-r-N -idempotent in A, then there is an obvious
0 (A). Approximating ((2e∗ − 1)(2e − 1) +
(1, 1)-controlled morphism K0(A) → KN
1)1/2e((2e∗ − 1)(e − 1) + 1)−1/2 by using a power serie (compare with the proof
of lemma 1.10), we get that for every N > 1, there exists a control pair (λN , hN )
such that K0(A) → KN
0 (A) is a (λN , hN )-controlled isomorphism. Using the polar
decomposition, we have a similar statement in the odd case.
References
[1] Michael Atiyah. Elliptic operators, discrete groups and von Neumann algebras. In Col-
loque "Analyse et Topologie" en l'Honneur de Henri Cartan (Orsay, 1974), pages 43 -- 72.
Ast´erisque, No. 32 -- 33. Soc. Math. France, Paris, 1976.
[2] Paul Baum, Alain Connes, and Nigel Higson. Classifying space for proper actions and K-
theory of group C ∗-algebras. In C ∗-algebras: 1943 -- 1993 (San Antonio, TX, 1993), volume
167 of Contemp. Math., pages 240 -- 291. Amer. Math. Soc., Providence, RI, 1994.
[3] Alain Connes. A survey of foliations and operator algebras. In Operator algebras and ap-
plications, Part I (Kingston, Ont., 1980), volume 38 of Proc. Sympos. Pure Math., pages
521 -- 628. Amer. Math. Soc., Providence, R.I., 1982.
[4] Alain Connes and Georges Skandalis. The longitudinal index theorem for foliations. Publ.
Res. Inst. Math. Sci., 20(6):1139 -- 1183, 1984.
[5] Alain Connes and Henri Moscovici. Cyclic cohomology, the Novikov conjecture and hyperbolic
groups. Topology, 29(3):345 -- 388, 1990.
[6] Joachim Cuntz. K-theoretic amenability for discrete groups. J. Reine Angew. Math., 344:180 --
195, 1983.
[7] Guihua Gong, Qin Wang, and Guoliang Yu. Geometrization of the strong Novikov conjecture
for residually finite groups. J. Reine Angew. Math., 621:159 -- 189, 2008.
ON A QUANTITATIVE OPERATOR K-THEORY
53
[8] Nigel Higson and Gennadi Kasparov. E-theory and KK-theory for groups which act properly
and isometrically on Hilbert space. Invent. Math., 144(1):23 -- 74, 2001.
[9] Nigel Higson, John Roe, and Guoliang Yu. A coarse Mayer-Vietoris principle. Math. Proc.
Cambridge Philos. Soc., 114(1):85 -- 97, 1993.
[10] Pierre Julg and Alain Valette. K-theoretic amenability for SL2(Qp), and the action on the
associated tree. J. Funct. Anal., 58(2):194 -- 215, 1984.
[11] Gennadi Kasparov. Equivariant KK-theory and the Novikov conjecture. Invent. Math.,
91(1):147 -- 201, 1988.
[12] Vincent Lafforgue. K-th´eorie bivariante pour les alg`ebres de Banach, groupoıdes et conjecture
de Baum-Connes. Avec un appendice d'Herv´e Oyono-Oyono. J. Inst. Math. Jussieu, 6(3):415 --
451, 2007.
[13] Michel Matthey, Herv´e Oyono-Oyono, and Wolfgang Pitsch. Homotopy invariance of higher
signatures and 3-manifold groups. Bull. Soc. Math. France, 136(1):1 -- 25, 2008.
[14] Herv´e Oyono-Oyono and Guoliang Yu. K-theory for the maximal Roe algebra of certain
expanders. J. Funct. Anal., 257(10):3239 -- 3292, 2009.
[15] John Roe. Coarse cohomology and index theory on complete Riemannian manifolds. Mem.
Amer. Math. Soc., 104(497), 1993.
[16] Georges Skandalis, Jean-Louis. Tu, and Guoliang Yu. The coarse Baum-Connes conjecture
and groupoids. Topology, 41(4):807 -- 834, 2002.
[17] Jean-Louis Tu. La conjecture de Baum-Connes pour les feuilletages moyennables. K-Theory,
17(3):215 -- 264, 1999.
[18] N. E. Wegge-Olsen. K-theory and C ∗-algebras. Oxford Science Publications. The Clarendon
Press Oxford University Press, New York, 1993. A friendly approach.
[19] Guoliang Yu. The Novikov conjecture for groups with finite asymptotic dimension. Ann. of
Math. (2), 147(2):325 -- 355, 1998.
[20] Guoliang Yu. The coarse Baum-Connes conjecture for spaces which admit a uniform embed-
ding into Hilbert space. Invent. Math., 139:201 -- 240, 2000.
Universit´e de Lorraine, Metz , France
E-mail address: [email protected]
Vanderbilt University, Nashville, USA
E-mail address: [email protected]
|
1409.8523 | 2 | 1409 | 2015-07-08T15:20:10 | Unbounded Operators on Hilbert $C^*$-Modules | [
"math.OA",
"math.FA"
] | Let $E$ and $F$ be Hilbert $C^*$-modules over a $C^*$-algebra $\CAlg{A}$. New classes of (possibly unbounded) operators $t:E\to F$ are introduced and investigated. Instead of the density of the domain $\Def(t)$ we only assume that $t$ is essentially defined, that is, $\Def(t)^\bot=\{0\}$. Then $t$ has a well-defined adjoint. We call an essentially defined operator $t$ graph regular if its graph $\Graph(t)$ is orthogonally complemented in $E\oplus F$ and orthogonally closed if $\Graph(t)^{\bot\bot}=\Graph(t)$. A theory of these operators is developed. Various characterizations of graph regular operators are given. A number of examples of graph regular operators are presented ($E=C_0(X)$, a fraction algebra related to the Weyl algebra, Toeplitz algebra, Heisenberg group). A new characterization of affiliated operators with a $C^*$-algebra in terms of resolvents is given. | math.OA | math |
UNBOUNDED OPERATORS ON HILBERT C∗-MODULES
REN´E GEBHARDT AND KONRAD SCHM UDGEN
Abstract. Let E and F be Hilbert C ∗-modules over a C ∗-algebra A. New
classes of (possibly unbounded) operators t : E → F are introduced and in-
vestigated. Instead of the density of the domain D(t) we only assume that t is
essentially defined, that is, D(t)⊥ = {0}. Then t has a well-defined adjoint. We
call an essentially defined operator t graph regular if its graph G(t) is orthog-
onally complemented in E ⊕ F and orthogonally closed if G(t)⊥⊥ = G(t). A
theory of these operators and related concepts is developed. Various character-
izations of graph regular operators are given. A number of examples of graph
regular operators are presented (E = C0(X), a fraction algebra related to the
Weyl algebra, Toeplitz algebra, Heisenberg group). A new characterization of
affiliated operators with a C ∗-algebra in terms of resolvents is given.
Contents
Introduction
1.
2. Orthogonally closed operators on Hilbert C∗-modules
3. Operators on the commutative C∗-algebra C0(X)
4. Graph regular operators
4.1. Definition and basics on graph regular operators
4.2. The (a, a∗, b)-transform
4.3. Quotients of adjointable operators
4.4. Absolute value
4.5. Bounded transform
4.6. Polar decomposition
4.7. Graph regular operators on C0(X)
4.8. Functional calculus of graph regular normal operators
5. Associated operators and affiliated operators
6. Examples
6.1. Matrices of commutative C∗-algebras and its multipliers
6.2. A fraction algebra related to the Weyl algebra
6.3. Unbounded Toeplitz operators
6.4. Heisenberg group
References
1
3
9
14
14
15
18
20
20
23
24
26
26
32
32
33
35
37
38
1. Introduction
Hilbert C∗-modules are a well established tool in the theory of C∗-algebras and
their applications. They have been invented by I. Kaplansky [K53] for commutative
C∗-algebras and by W. Paschke [P73] and M. Rieffel [R74] in the general case.
Standard textbooks are [L95] and [MT05]. Unbounded operators on Hilbert C∗-
modules play an important role for the study of noncompact quantum groups [W91],
in KK-theory [BJ83, K97] and in noncommutative geometry [GVF].
2010 Mathematics Subject Classification. Primary 46 L 08; Secondary 47 D 40, 47 L 05.
Key words and phrases. Hilbert C ∗-modules, unbounded operators, affiliated operators.
1
2
REN ´E GEBHARDT AND KONRAD SCHM UDGEN
Let E and F be Hilbert C∗-modules over a C∗-algebra A. By an operator from
E into F we mean an A-linear and C-linear mapping t of an A-submodule D(t)
of E into F . Such an operator t is called regular if t is closed, D(t) is dense in
E, D(t∗) is dense in F , and I + t∗t is a bijection of E. Regular operators form
the most important class of unbounded operators on Hilbert C∗-modules. A nice
presentation of their theory can be found in Chapters 9 and 10 of Lance' book
[L95]. Regular operators were invented by S. Baaj [B81, BJ83] and extensively
studied by S.L. Woronowicz in two seminal papers [W91, WN92]. Woronowicz
considered the case when E is the C∗-algebra A itself and called the corresponding
operators affiliated with A. That is, the affiliated operators are precisely the regular
operators on the Hilbert C∗-modules E = A. Unbounded operators on Hilbert C∗-
modules are studied in [H89], [K97], [K02], [FS10], [KL12], [Pal99], [Pie06]. A
generalization of regular operators are the semiregular operators introduced by A.
Pal [Pal99]. Note that semiregular operators are always densely defined.
The aim of the present paper is to introduce several new classes of unbounded
operators on C∗-modules and to develop the basics of their theory. Let us briefly
explain the main new concepts by avoiding technical subtleties. Precise definitions
and further explanations will be given in the corresponding sections of the text.
Suppose that t is an operator from E into F such that its domain D(t) is essential,
that is, D(t)⊥ = {0}. Here D(t)⊥ denotes the orthogonal complement of D(t)
with respect to the A-valued scalar product of E. Then the operator t has a
well-defined adjoint operator t∗. An operator t is called orthogonally closed if its
graph G(t) = {(x, tx); t ∈ D(t)} is orthogonally closed, that is, if G(t)⊥⊥ = G(t) in
E ⊕ F . An orthogonally closed operator t is called graph regular if its graph G(t)
is orthogonally complemented in E ⊕ F . Graph regularity is the most important
new concept of operators appearing in this paper. This notion is also of interest in
the case when the operator t is bounded and only essentially defined.
It should be emphasized that all these operators are not necessarily densely
defined! Instead we assume only that their domains are essential, that is, they have
trivial orthogonal complements, and we replace the closure of the graph G(t) in
E ⊕ F by its double orthogonal complement G(t)⊥⊥.
The difference between graph regularity and regularity can be nicely illustrated
by the Hilbert C∗-module C0(X) for the C∗-algebra C0(X), where X is a locally
compact Hausdorff space. Then both classes of operators are given by multiplica-
tion operators: the regular operators by functions of C(X), but the graph regular
operators by functions which are only continuous up to a no-where dense set and
for which the modulus goes to infinity in neighbourhoods of discontinuities (see
Theorem 15). This example shows another essential difference: While a regular
operator can be transported into a densely defined closed operator by each induced
representation, for a graph regular operator this is only possible for certain rep-
resentations. To include such phenomena was one motivation for studying graph
regular operators. We will elaborate this elsewhere more in detail.
This paper is organized as follows. Section 2 contains basic definitions and facts
on essentially defined operators, their adjoints and orthogonally closed operators.
Graph regular operators are introduced and studied in Section 4. Basic proper-
ties of these operators are derived (Theorems 1 and 2) and two characterizations of
graph regular operators in terms of bounded operators are obtained. The first one,
the (a, a∗, b)-transform, gives a purely algebraic description of graph regular opera-
tors by a triple of adjointable operators (Theorem 7). The second characterization
(Theorems 11 and 12) concerns the bounded transform. This transform played
already a crucial role in Woronowicz' approach to affiliated operators. To each
graph regular operator t : E → F we associate a regular operator t0 : E0 → F0
UNBOUNDED OPERATORS ON HILBERT C ∗-MODULES
3
acting between essential submodules E0 and F0 of E and F , respectively. As a
byproduct of these considerations we use the notion of essentially defined partial
isometries to prove a general result on the polar decomposition of adjointable op-
erators (Theorem 13) and of graph regular operators (Theorem 14). Further, we
show that quotients of adjointable operators provide a large source of examples of
graph regular operators (Theorems 8 and 9). The functional calculus of normal
regular operators is extended to normal graph regular operators (Theorem 16).
In Section 5 we specialize to the case when E is the C∗-algebra A itself and A is
faithfully realized on a Hilbert space H. Then the regular operators on the Hilbert
C∗-module E = A are precisely Woronowicz' affiliated operators. Suppose that t
is a densely defined closed operator on H. We shall say that t is associated with
A if the operators at = (I + t∗t)−1, at∗ = (I + tt∗)−1, and bt = t(I + t∗t)−1 are in
the multiplier algebra M(A). A number of results as well as examples and counter-
examples on associated and affiliated operators are derived. One of the main new
results (Theorem 17) provides a characterization of affiliated operators in terms of
resolvents: If λ ∈ ρ(t), then t is affiliated with A if and only if (t−λI)−1 ∈ M(A) and
(t − λI)−1A and (t∗ − λ I)−1A are dense in A. This seems to be a useful criterion
for proving that operators are affiliated, since the resolvent is better understood in
operator theory than the bounded transform.
In Sections 3 and 6 we develop various classes of examples. A particular em-
phasis is on graph regular operators that are not regular on the corresponding
C∗-modules. Section 3 and Subsection 4.7 contain a careful treatment of the com-
mutative case E = C0(X), where X is a locally compact Hausdorff space. Among
others, essentially defined orthogonally closed operators and graph regular oper-
ators are characterized in this case.
In Subsection 6.1 we consider some simple
examples for matrices over commutative C∗-algebras. In Subsection 6.2 the posi-
tion operator Q = x and the momentum operator P = −i d
dx become graph regular
operators on a C∗-algebra obtained from a fraction algebra.
In Subsection 6.3
some unbounded Toeplitz operators are described as graph regular operators on
the Toeplitz C∗-algebra, while in Subsection 6.4 a graph regular operator on the
C∗-algebra of the Heisenberg group is constructed.
Finally, let us fix some notation that will be used in this paper. Throughout,
A denotes a C∗-algebra, Ah := {a ∈ A : a∗ = a} is its hermitian part, and
A+ := {a ∈ Aa ≥ 0} is its cone of positive elements. If H is a Hilbert space, we
denote by B(H) the bounded operators on H, by K(H) the compact operators on
H, and by C(H) the set of densely defined closed operators on H. For an operator
t ∈ C(H), let D(t) be its domain on H, N (t) its null space, R(t) its range and ρ(t)
its resolvent set.
2. Orthogonally closed operators on Hilbert C∗-modules
First we recall a standard definition.
Definition 1. A (right) pre-Hilbert C∗-module E over A is a right A-module E
equipped with a map h., .i : E × E → A satisfying the following conditions:
λ(xa) = (λx)a = x(λa),
λ ∈ C, x ∈ E, a ∈ A,
hαx + βy, zi = α hx, zi + β hy, zi , α, β ∈ C, x, y, z ∈ E,
a ∈ A, x, y ∈ E,
x, y ∈ E,
hx, yai = hx, yi a,
hx, yi = hy, xi∗ ,
hx, xi ≥ 0,
x ∈ E,
hx, xi = 0 =⇒ x = 0,
x ∈ E.
4
REN ´E GEBHARDT AND KONRAD SCHM UDGEN
A pre-Hilbert C∗-module E over A is called a Hilbert C∗-module over A, briefly a
Hilbert A-module, if (E,k.kE) is complete, where k.kE is the norm on E given by
kxkE := k hx, xi k1/2
A
,
x ∈ E.
Example 1. The C∗-algebra A itself is a Hilbert A-module over A by taking the
multiplication as right action and the A-valued scalar product ha, bi := a∗b, a, b ∈ E.
In this case kakE = k ha, aik1/2
A
= kakA for a ∈ E.
Suppose that E is a Hilbert A-module. If x, y ∈ E and hx, yi = 0, we write x⊥y.
For a subset F of E, the set
F ⊥ := {x ∈ E∀y ∈ F : hx, yi = 0}
is called the orthogonal complement of F . Clearly, F ⊥ is a submodule of E and
x⊥y is equivalent to y⊥x, since hx, yi = hy, xi∗. If F, G are subsets of E, then
F ⊆ F ⊥⊥, F ⊆ G =⇒ G⊥ ⊆ F ⊥, F ⊥ = F ⊥⊥⊥.
We only verify the last equality F ⊥ = F ⊥⊥⊥. The first inclusion yields F ⊥ ⊆
(F ⊥)⊥⊥. On the other hand, the relation F ⊆ F ⊥⊥ implies that (F ⊥⊥)⊥ ⊆ F ⊥.
Therefore, F ⊥ = F ⊥⊥⊥.
Let F and G be subsets of E. We set F + G := {f + gf ∈ F, g ∈ G} and write
F ⊕ G := F + G if F ⊆ G⊥. Since F ⊆ G⊥ implies G ⊆ G⊥⊥ ⊆ F ⊥, we always
have F ⊕ G = G ⊕ F . Since F ⊆ (F ⊥)⊥, it is justified to write F ⊕ F ⊥.
Definition 2. A subset F of E is said to be essential if F ⊥ = {0}. A submodule
F of E is called orthogonally closed if F = F ⊥⊥ and orthogonally complemented
if F ⊕ F ⊥ = E.
Since F ⊥ is always closed in E, each orthogonally closed submodule is closed.
Example 2. For the C∗-algebra C0(X) of continuous functions on a locally com-
pact Hausdorff X space vanishing at infinity each closed ideal is of the form
IO := {f ∈ C0(X)∀x ∈ X \ O : f (x) = 0}
for some open subset O ⊆ X. The mapping O → IO is a bijection of the open
subsets of X onto the closed ideals of C0(X). For x ∈ X we have x ∈ O if and
only if there exists f ∈ IO with f (x) 6= 0. The following facts are easily verified:
= I(X\O)◦.
• I⊥
O
• IO is essential in C0(X) if and only if O is dense in X.
• IO is orthogonally closed if and only if O coincides with the interior of its
• IO is orthogonally complemented if O is closed.
closure.
The following simple facts will be often used:
If F, G are submodules of E,
then (F ∩ G)⊥ ⊇ (F ⊥ + G⊥)⊥⊥ and (F + G)⊥ = F ⊥ ∩ G⊥. Further, if F, G are
orthogonally closed, then (F ∩ G)⊥ = (F ⊥ + G⊥)⊥⊥.
Lemma 1. For any subset F of E, the set F ⊕ F ⊥ is essential.
Proof. If x ∈ (F ⊕ F ⊥)⊥ = F ⊥ ∩ F ⊥⊥, then hx, xi = 0 and hence x = 0.
(cid:3)
Lemma 2. If F, G are submodules of E with F ⊆ G and F ⊥ ∩ G = {0}, then
F ⊥ ∩ (G ⊕ G⊥) = G⊥.
Proof. Clearly, F ⊥ ⊇ G⊥, so F ⊥ ∩ (G ⊕ G⊥) ⊇ F ⊥ ∩ G⊥ = G⊥. Now assume that
x = g + g⊥ ∈ F ⊥ with g ∈ G, g⊥ ∈ G⊥. But then x − g⊥ = g ∈ G ∩ F ⊥, so g = 0
and x = g⊥ ∈ G⊥.
(cid:3)
UNBOUNDED OPERATORS ON HILBERT C ∗-MODULES
5
The direct sum E ⊕ F of two Hilbert A-modules E and F is a right A-module.
If we define a mapping h., .iE⊕F : (E ⊕ F ) × (E ⊕ F ) → A by
h(e1, f1), (e2, f2)iE⊕F := he1, e2iE + hf1, f2iF ,
then this module becomes also a Hilbert A-module.
Now we turn to operators on Hilbert A-modules. By an operator t from E into
F we mean a C-linear and A-linear mapping defined on a right A-submodule D(t)
of E, called the domain of t. The symbol t : E → F always denotes an operator
from E into F . The C-linearity and A-linearity of t mean that
e1, e2 ∈ E, f1, f2 ∈ F,
t(λx) = λt(x)
and t(xa) = t(x)a for λ ∈ C, x ∈ D(t), a ∈ A.
For an operator t : E → F , its null space N (t) := {x ∈ Etx = 0} is a right
A-submodule of E, its range R(t) := {txx ∈ D(t)} is a right A-submodule of F
and its graph G(t) := {(x, tx)x ∈ D(t)} is a right A-submodule of E ⊕ F . As in
the case of ordinary Hilbert space operators we say t is closed if G(t) is closed in
E ⊕ F and t is closable if there exists an operator s which is a closed extension of
t. In this case there exists a unique closed operator, denoted by t and called the
closure of t, such that G(t) = G(t).
An operator t : E → E is called positive if htx, xi ≥ 0 for all x ∈ D(t).
Definition 3. An operator t : E → F is called essentially defined if D(t) is an
essential submodule of E.
Suppose that t : E → F is an essentially defined operator. Set
D(t∗) := {y ∈ F∃z ∈ E : ∀x ∈ D(t) : htx, yiF = hx, ziE}.
Since D(t) is an essential submodule, the element z ∈ E is uniquely determined by
y. We define t∗y := z. Then t∗ : F → E is an operator, called the adjoint of t, and
htx, yi = hx, t∗yi
for x ∈ D(t), y ∈ D(t∗).
The operators of the set
L(E, F ) := {t : E → FD(t) = E,D(t∗) = F}
are called adjointable. Note that L(E) := L(E, E) is a unital C∗-algebra.
Definition 4. An essentially defined operator t : E → E is called symmetric if
t ⊆ t∗, and self-adjoint if t = t∗.
Let v : E ⊕ F → F ⊕ E denote the unitary operator (x, y) 7→ (−y, x).
Proposition 1. Suppose that t : E → F is essentially defined. Then:
(1) G(t∗) = vG(t)⊥.
(2) N (t∗) = R(t)⊥.
(3) If t is injective and R(t)⊥ = {0}, then t∗ is injective and (t∗)−1 = (t−1)∗.
(4) G(t) ⊕ vG(t∗) is essential.
The proof of these statements is similar to the Hilbert space case; we omit the
details.
Lemma 3. Let r : E → F be essentially defined. Suppose that D(r∗) = F and
R(r) is essential. If D ⊆ D(r) is essential, then R(r ↾D) is essential.
Proof. Let x ∈ F be such that hx, ryi = 0 for all y ∈ D. Then hr∗x, yi = 0 for
all y ∈ D. Since D⊥ = {0} by assumption, we conclude that r∗x = 0. That is,
x ∈ N (r∗) = R(r)⊥ = {0}.
(cid:3)
Proposition 2. Let t, t1, t2 be essentially defined operators from E into F and let
s be an essentially defined operator from F into G. Then:
6
REN ´E GEBHARDT AND KONRAD SCHM UDGEN
(1) t1 ⊆ t2 implies t∗1 ⊇ t∗2.
(2) If t1 + t2 is essentially defined, then (t1 + t2)∗ ⊇ t∗1 + t∗2. If D(t2) ⊆ D(t1)
and D(t∗1) = E, then t1 + t2 is essentially defined and (t1 + t2)∗ = t∗1 + t∗2.
(3) If st is essentially defined, then (st)∗ ⊇ t∗s∗. If R(t) ⊆ D(s) and D(s∗) =
G, then st is essentially defined and (st)∗ = t∗s∗.
(4) If t is injective, D(s) ⊆ R(t) and D((t−1)∗) = F , then st is essentially
defined and (st)∗ = t∗s∗.
Proof. Assertions (1) -- (3) are shown by simple computations.
We prove (4). By Lemma 3, the domain D(st) = t−1D(s) is essential. For
x ∈ D((st)∗) and y ∈ D(s) we derive
hsy, xi =(cid:10)(st)t−1y, x(cid:11) =(cid:10)t−1y, (st)∗x(cid:11) =(cid:10)y, (t−1)∗(st)∗x(cid:11) .
Therefore, x ∈ D(s∗) and s∗x = (t−1)∗(st)∗x = (t∗)−1(st)∗x ∈ D(t∗). That is,
D((st)∗) ⊆ D(t∗s∗). Now (3) completes the proof.
(cid:3)
Definition 5. An essentially defined operator p is a projection if p = p2 = p∗.
The next proposition characterizes projections and shows that they are in one-
to-one correspondence to orthogonally closed submodules.
Proposition 3. For a submodule G of E we define an operator pG : E → E by
D(pG) := G ⊕ G⊥,
pG(x + y) := x,
x ∈ G, y ∈ G⊥.
Then pG is essentially defined and pG = p2
(pG)∗ is a projection if and only if G is orthogonally closed.
G ⊆ p∗G = pG⊥⊥. In particular, pG =
Conversely, if p is a projection, then p = pG for some orthogonally closed sub-
module G of E.
Proof. By Lemma 1, p := pG is essentially defined. Obviously, p2 = p ⊆ p∗. Let
z ∈ D(p∗). Then there exists w ∈ E such that
hx, zi = hp(x + y), zi = hx + y, wi ,
x ∈ G, y ∈ G⊥.
Equivalently, hx, z − wi = hy, wi for all x ∈ G, y ∈ G⊥, so both sides are zero. The
latter is equivalent to z− w ∈ G⊥ and w ∈ G⊥⊥, that is, z ∈ G⊥⊕ G⊥⊥. Therefore,
since w = p∗z, the operator p∗ is given by
D(p∗) = G⊥⊥ ⊕ G⊥,
p∗(x + y) = x,
x ∈ G⊥⊥, y ∈ G⊥.
This completes the proof of the first half.
Now assume that p = p2 = p∗ and let G := R(p) ⊆ D(p). For x = py ∈ G we
have px = p2y = py = x. For x ∈ G⊥, we obtain hx, pyi = 0 = h0, yi for y ∈ D(p).
Therefore, x ∈ D(p∗) = D(p) and px = p∗x = 0. Thus, we have shown that pG ⊆ p.
Further, the assumption p = p2 implies that R(p) = N (1 − p). Hence
G = R(p) = N (1 − p) = N (1 − p∗) = R(1 − p)⊥.
Therefore, G = G⊥⊥ and pG is self-adjoint. Since p is also self-adjoint and pG ⊆ p,
(cid:3)
we conclude that p = pG.
Definition 6. Let t : E → F be an operator. A submodule D of D(t) is called an
essential core for t if G(t ↾D)⊥ = G(t)⊥.
By Lemma 1 this is equivalent to the relation (t ↾D)∗ = t∗ if D is essential.
Clearly, D is an essential core for t if and only if G(t ↾D)⊥ ⊆ G(t)⊥, or equivalently,
if the sum G(t ↾D)⊥ + G(t) is an orthogonal sum G(t ↾D)⊥ ⊕ G(t).
Example 3. Let b ∈ L(E, F ) and let D be an essential submodule of E. Then D
is an essential core for b, since (b ↾D)∗ ⊇ b∗ and b∗ is everywhere defined.
UNBOUNDED OPERATORS ON HILBERT C ∗-MODULES
7
Definition 7. A (not necessarily essentially defined) operator t : E → F is or-
thogonally closed if G(t) is orthogonally closed, that is, if G(t)⊥⊥ = G(t), and
orthogonally closable if G(t)⊥⊥ = G(s) for some (orthogonally closed) operator s.
By Proposition 1 the adjoint is always orthogonally closed.
Theorem 1. Let t : E → F be essentially defined. Then D(t∗) is an essential
submodule of F if and only if t is orthogonally closable.
Proof. Suppose that D(t∗) is an essential submodule. Then t∗∗ exists and it follows
easily that t ⊆ t∗∗. Applying Lemma 1 twice, first to t and then to t∗, we obtain
G(t∗∗) = G(t)⊥⊥. Hence t∗∗ is orthogonally closed and t is orthogonally closable.
Now suppose that G(t)⊥⊥ = G(s) for some (essentially defined) operator s. Let
z ∈ D(t∗)⊥. Then h(−z, 0), (y, t∗y)i = − hz, yi + h0, t∗yi = 0 for all y ∈ D(t∗), so
(z, 0) ∈ vG(t∗)⊥ = G(t)⊥⊥ = G(s) and hence z = 0, since s is an operator.
(cid:3)
Let us define the sets
C′o(E, F ) := {t : E → FD(t)⊥ = {0},D(t∗)⊥ = {0}},
Co(E, F ) := {t ∈ C′o(E, F )t is orthogonally closed},
C′o(E) := C′o(E, E),
Co(E) := Co(E, E).
Then, by Theorem 1, C′o(E, F ) is the set of essentially defined operators t : E → F
that are orthogonally closable.
Theorem 2.
(1) Suppose that t ∈ C′o(E, F ). Then we have:
(a) t ⊆ t∗∗, t∗ = t∗∗∗, and G(t)⊥⊥ = G(t∗∗).
(b) t = t∗∗ if and only if t ∈ Co(E, F ).
(c) N (t∗∗) is orthogonally closed.
(d) If R(t) and R(t∗) are essential, then t and t∗∗ are injective, t−1 is
essentially defined and orthogonally closable, and (t∗∗)−1 = (t−1)∗∗.
(2) Suppose that t ∈ Co(E, F ). Then:
(a) N (t) is orthogonally closed.
(b) If R(t) and R(t∗) are essential, then t is injective and t−1 is essentially
defined and orthogonally closed.
Proof. (1a): By Theorem 1, t∗∗ exists. Applying Proposition 1 several times we
obtain G(t∗∗) = vG(t∗)⊥ = G(t)⊥⊥, so that t ⊆ t∗∗. This implies that t∗∗∗ exists.
Using Proposition 1 once again we get G(t∗∗∗) = vG(t)⊥⊥⊥ = vG(t)⊥ = G(t∗),
hence t∗∗∗ = t∗.
(1b) follows at once from the equality G(t)⊥⊥ = G(t∗∗).
(1c): We derive N (t∗∗)⊥⊥ = R(t∗)⊥⊥⊥ = R(t∗)⊥ = N (t∗∗).
(1d): By Proposition 1, t∗ is injective and (t∗)−1 = (t−1)∗. By assumption,
R(t∗) is essential. Therefore, using Proposition 1 again we conclude that t∗∗ is
injective and (t∗∗)−1 = ((t∗)−1)∗ = (t−1)∗∗.
(cid:3)
(2a) and (2b) follow from (1c) and (1d), respectively.
The operator t∗∗ is called the orthogonal closure of the orthogonally closable
operator t.
Lemma 4. Let t ∈ Co(F, G) and b ∈ L(E, F ). Suppose that the operator tb is
essentially defined. Then tb ∈ Co(E, G) and (tb)∗ = (b∗t∗)∗∗.
Proof. From Proposition 2 it follows that (tb)∗ ⊇ b∗t∗ and the latter is essentially
defined. Therefore, tb ⊆ (tb)∗∗ ⊆ (b∗t∗)∗ = t∗∗b∗∗ = tb. This implies that (tb)∗ =
(b∗t∗)∗∗.
(cid:3)
Definition 8. An operator t ∈ C′o(E) is called essentially self-adjoint if t∗ = t∗∗.
An operator t ∈ Co(E, F ) is called normal if t∗t = tt∗.
8
REN ´E GEBHARDT AND KONRAD SCHM UDGEN
Proposition 4. Let t ∈ C′o(E, F ) and let D be a submodule of D(t). If D is an
essential core for t, then D is an essential submodule of E and
(t ↾D)∗ = t∗,
t ⊆ (t ↾D)∗∗ = t∗∗.
In particular, D is an essential core for t∗∗ and if t is orthogonally closed, then
t = (t ↾D)∗∗.
Proof. Define an operator pE : E ⊕ F → E by pE(x, y) := x for x ∈ E, y ∈ F .
It is straightforward to check that the adjoint operator p∗E : E → E ⊕ F acts
by p∗E(x) = (x, 0) for x ∈ E. Thus, pE ∈ L(E ⊕ F, E). Now, using Proposition
2,(3) for the second equality, Proposition 3 for the third equality, and the relations
G(t ↾D)⊥⊥ = G(t)⊥⊥ = G(t∗∗) for the fourth equality, we derive
(1)
G(t↾D )p∗E) = N (pG(t↾D )⊥⊥p∗E) = N (pG(t∗∗)p∗E).
D⊥ = R(pEpG(t↾D ))⊥ = N (p∗
Let x ∈ N (pG(t∗∗)p∗E). Then x ∈ D(pG(t∗∗)p∗E) and (x, 0) ∈ G(t∗∗) ⊕ vG(t∗). Hence
there exist elements y ∈ D(t∗∗) and z ∈ D(t∗) such that x = y+t∗z and 0 = t∗∗y−z.
Since (0, 0) = pG(t∗∗)p∗Ex = pG(t∗∗)(x, 0) = (y, t∗∗y), we get y = 0, so z = 0
and x = 0. Thus, N (pG(t∗∗)p∗E) = {0}. Hence D is essential by (1). The other
(cid:3)
statements are easily verified.
Some algebraic properties on orthogonally closable operators are collected in the
next proposition. The proofs are straightforward and we omit the details.
Proposition 5. Let t ∈ C′o(E, F ).
(1) If s ∈ L(E, F ), then t + s ∈ C′o(E, F ).
(2) If s ∈ L(F, G) is injective with s−1 ∈ L(G, F ), then st ∈ C′o(E, G).
(3) If s ∈ L(G, E) is injective with s−1 ∈ L(E, G), then ts ∈ C′o(G, F ).
All these statements remain valid if C′o is replaced by Co.
Remark 1. In this Remark we want to emphasize that the theory developed so far
is valid in a much more general setting. For this we introduce some definitions.
Let A be a complex ∗-algebra. A ∗-bimodule for A is a bimodule X for A
equipped with an involution (that is, an antilinear mapping x → x∗ of X such that
(x∗)∗ = x for x ∈ X) satisfying (ax)∗ = x∗a∗ and (xa)∗ = a∗x∗ for a ∈ A, x ∈ X.
Definition 9. A quadratic ∗-space over A is a triple (E, X,h., .i) of a right A-
module E, a ∗-bimodule X for A and a map h., .i : E × E → X such that for
α, β ∈ C, x, y, z, x1, . . . , xn ∈ E, a ∈ A:
hαx + βy, zi = αhx, zi + β hy, zi ,
hx, yai = hx, yi a,
hx, yi = hy, xi∗ ,
j=1 hxj , xji = 0
implies x1 = ··· = xn = 0.
Xn
All preceding notions and results on essentially defined operators and their ad-
joints, on orthogonally closed operators and on graph regular operators remain valid
for quadratic ∗-spaces rather than Hilbert A-modules. Indeed, an inspection of the
definitions and proofs shows that only the axioms from Definition 9 are needed. The
same is true for Theorem 6, Theorem 7 and Proposition 7. However, from now on
C∗-algebra properties are essentially used.
UNBOUNDED OPERATORS ON HILBERT C ∗-MODULES
9
3. Operators on the commutative C∗-algebra C0(X)
Throughout this section we suppose that X is a locally compact Hausdorff space.
We consider the C∗-algebra C0(X) of continuous functions on X vanishing at infin-
ity as a Hilbert C∗-module E (see Example 1) and study operators t : E → E. The
main aim of this section is to describe orthogonally closed operators on E and to
show that Co(E) consists of multiplication operators. For this reason we investigate
multiplication operators in detail. To state the results we introduce some notation
that is inspired by [KL12, Section 6].
For a function m : X → C we set
reg(m) :={x ∈ Xm is continuous in a neighborhood of x},
regb(m) :=nx ∈ ∂reg(m)∃U ⊆ reg(m) open, m : U → C continuous,
reg∞(m) :=nx ∈ ∂reg(m)∃U ⊆ reg(m) open, m : U → C continuous,
with x ∈ U, m ≡ m on U ∩ reg(m)} ,
with x ∈ U, m ≡ m on U ∩ reg(m), m(x) = ∞} ,
sing-suppr(m) :=∂reg(m) \ (regb(m) ∪ reg∞(m)).
Clearly, reg(m) is the largest open set on which m is continuous. Further,
reg(m) ∪ regb(m) is the largest open set contained in reg(m) on which m re-
stricted to reg(m) has a (indeed unique) continuous C-valued extension. Finally,
reg(m)∪ regb(m)∪ reg∞
(m) is the largest open set contained in reg(m) on which
m restricted to reg(m) has a (unique) continuous C-valued extension. In partic-
(m) is contained in (reg(m))◦. The space X is a
ular, reg(m) ∪ regb(m) ∪ reg∞
disjoint union
X = reg(m) ∪ regb(m) ∪ reg∞
(m) ∪ sing-suppr(m)
∂reg(m)
{z
∪ (X \ reg(m))
(X\reg(m))◦
.
}
{z
}
By F0(X) we denote the set of functions on X vanishing at infinity. We note
E = C0(X) = {f ∈ F0(X)reg(f ) = X}.
Let us briefly summarize the results that will be obtained in this section. For an
arbitrary function m on X a multiplication operator tm is defined and we introduce
an equivalence relation to characterize those functions giving the same operator
(Lemma 7). Then we show that tm is essentially defined if and only if reg(m) is
dense in X and that in this case tm is already orthogonally closed and its adjoint
is tm (Theorem 3). Further, we prove that t ∈ Co(E) if and only if t = tm for some
function m : X → C (Theorem 4). In Example 4 we give a normal operator t for
which the domains D(t) and D(t∗) are different.
For m, m : X → C we write
m ≃ m ⇔ reg(m) = reg( m) and m ≡ m on reg(m) ∩ reg( m).
In Lemma 5(2) it will be shown that ≃ is an equivalence relation. To simplify the
notation we associate to each function m : X → C a function m on X defined by
m(x) =(m(x)
m(x)
, x ∈ X \ regb(m)
, x ∈ regb(m)
,
where m is one of the functions appearing in the definition of regb(m). In fact,
any two of those functions have the same values at x, since x is in the boundary
of reg(m) and the two continuous functions coincide on this set. Hence m is well-
defined. Lemma 5(3) shows that m ≃ m.
10
REN ´E GEBHARDT AND KONRAD SCHM UDGEN
Lemma 5.
(1) Let m1, m2 : X → C. If m1 ≃ m2, then reg∞(m1) = reg∞(m2)
and sing-suppr(m1) = sing-suppr(m2).
(2) ≃ is an equivalence relation on the set of functions from X to C.
(3) If m : X → C, then m ≃ m, reg( m) = reg(m) ∪ regb(m), regb( m) = ∅.
Proof. Note that the intersection of two open and dense sets is again open and
dense.
(m2). This proves (1).
(1): From R := reg(m1) = reg(m2) we conclude that reg(m1) ∩ reg(m2) is
dense in R. Since m1 and m2 are equal and continuous on this set, any x in R
is in reg(m1) ∪ regb(m1) if and only if it is in reg(m2) ∪ regb(m2). By the same
argument any x in R is in reg(m1) ∪ regb(m1) ∪ reg∞
(m1) if and only if it is in
reg(m2) ∪ regb(m2) ∪ reg∞
(2): Obviously ≃ is reflexive and symmetric, so it remains to show transitivity.
Let m1 ≃ m2 and m2 ≃ m3. Clearly, reg(m1) = reg(m2) = reg(m3). Arguing as
above, reg(m1) ∩ reg(m2) ∩ reg(m3) is open and dense in the latter set. Again by
continuity of m1 and m3 on reg(m1) ∩ reg(m3), these functions coincides on this
set, since they do on reg(m1) ∩ reg(m2) ∩ reg(m3). That is, m1 ≃ m3.
(3): Clearly, m is continuous on reg(m) ∪ regb(m), so the latter is contained
in reg( m). Since m and m coincide on the open set X \ reg(m), we even have
reg( m) ⊆ reg(m). In particular, reg(m) = reg( m) and since m and m are equal
on reg(m)∩reg( m) = reg(m), it follows that m ≃ m. By (1) the proof is complete,
since reg(m) ∪ regb(m) = reg( m) ∪ regb( m) implies that regb( m) = ∅.
(cid:3)
(m) ∪ sing-suppr(m) does not change any of
In particular, changing m on reg∞
Now we define the multiplication operators tm.
the sets reg(m), regb(m), reg∞
(m), sing-suppr(m), and X \ reg(m).
Definition 10. For a function m : X → C let
f ∈ D(tm).
It is straightforward to check that tm is indeed an operator on E.
D(tm) := {f ∈ Edmf ∈ E},
tmf :=dmf ,
Lemma 6. Let m, f : X → C and x ∈ X. If f is continuous at x and f (x) 6= 0,
then x ∈ reg(m) if and only if x ∈ reg(mf ).
Proof. Since f (x) 6= 0, there is an open set Uf containing x such that f (x′) 6= 0
for all x′ ∈ Uf . By definition, x /∈ reg(m) if for any neighbourhood U of x, there
is an x′ ∈ U such that m is discontinuous at x′. This holds if and only if for any
neighbourhood U of x, there exists x′ ∈ U such that mf is discontinuous at x′.
This means that x /∈ reg(mf ).
(cid:3)
We now show that the operator tm depends only on the equivalence class of m.
Lemma 7. Let m : X → C be given. Then
D(tm) = {f ∈ Ef ≡ 0 on X \ reg( m), ∂reg( m) ⊆ reg(dmf ),bmf ∈ F0(X)},
f ∈ D(tm), x ∈ X \ reg∞
(tmf )(x) = m(x)f (x),
(m).
For x ∈ X there is an f ∈ D(tm) such that f (x) 6= 0 if and only if x ∈ reg( m). If
m, m : X → C, then tm = t m if and only if m ≃ m. In particular, tm = t m.
Proof. By Definition 10 D(tm) consists of those f ∈ E for which reg(dmf ) = X and
dmf ∈ F0(X). Since reg( m) ⊆ reg(dmf ) for f ∈ E, D(tm) is the set of all f ∈ E
such that X \ reg( m) ⊆ reg(dmf ) anddmf ∈ F0(X).
Let f ∈ E with bmf ∈ F0(X). Suppose that ∂reg( m) ⊆ reg(dmf ) and f ≡ 0
X\reg( m) = X\reg( m)∪∂reg( m) is contained in reg(dmf ). To showdmf ∈ F0(X),
on X \ reg( m). Then mf ≡ 0 on X \ reg( m) and so on X \ reg( m). Hence
UNBOUNDED OPERATORS ON HILBERT C ∗-MODULES
11
For f ∈ D(tm) it is obvious that tmf ≡ mf on reg( m). Since f ≡ 0 on
Now we suppose that f ∈ D(tm). In particular, reg(mf ) is dense in X, hence
continuity of f there exists an open set U ⊆ X \ reg( m) such that f (y) 6= 0 for
y ∈ U . From Lemma 6 it follows that U is even contained in X \ reg( mf ). But this
contradicts the density of reg(mf ) ⊆ reg( mf ) in X, hence f (x) = 0. In particular
let ǫ > 0. Since bmf ∈ F0(X), there exists an compact set K ⊆ X such that bmf ≤ ǫ
on X\K. By continuity ofdmf on X the same is true for this function, since bmf and
dmf coincide on the dense set X \ ∂reg(m). That isdmf ∈ F0(X), finally f ∈ D(tm).
∂reg( m) ⊆ X = reg(dmf ). Assume that x ∈ X \ reg( m) and f (x) 6= 0. By the
dmf coincides with bmf on reg( m) and bmf ≡ 0 on X \ reg( m). So dmf ∈ F0(X)
implies bmf ∈ F0(X) and the description of D(tm) is proven.
X \ reg( m), it remains to show dmf ≡ 0 on (X \ reg( m)) ∪ sing-suppr(m). For
this it suffices to prove the following: If x ∈ X with f (x) = 0 and dmf (x) 6= 0,
(m). Indeed, dmf does not vanish on a neighbourhood U of x. Let
m : U → C denote the functiondmf /f on U , where α/0 := ∞ for α ∈ C. Then m is
continuous, sincedmf does not vanish on U , and m coincides with m on U ∩ reg(m).
Further, since f (x) = 0 anddmf (x) 6= 0, we have m(x) = ∞. Hence x ∈ reg∞(m).
Let x ∈ X. By the preceding, if x /∈ reg( m), then f (x) = 0 for all f ∈ D(tm).
If x ∈ reg( m), there exists a function f : X → C with compact support contained
in reg( m) and f (x) 6= 0, since X is locally compact. Then f ∈ D(tm), since mf is
continuous everywhere and has compact support.
Next we prove that tm = t m if and only if m ≃ m. Assume first that tm = t m.
then x ∈ reg∞
Then, by the preceding,
x ∈ reg( m) ⇔ ∃f ∈ D(tm) : f (x) 6= 0 ⇔ ∃f ∈ D(t m) : f (x) 6= 0 ⇔ x ∈ reg( m).
Thus reg( m) = reg( m) which implies that reg(m) = reg( m) = reg( m) = reg( m).
For x ∈ reg(m) ∩ reg( m) we choose f ∈ D(tm) such that f (x) 6= 0. Then
reg(m) ∩ reg( m) and the latter set is dense in reg( m), since m ≃ m. Further,
m(x)f (x) =dmf (x) =dmf (x) = m(x)f (x), so that m(x) = m(x). Hence m ≃ m.
Conversely, assume that m ≃ m. Let f ∈ D(tm). Then dmf ≡ mf ≡ mf on
f ≡ 0 on X \ reg( m) by f ∈ D(tm) and hence dmf ≡ 0 ≡ mf on X \ reg( m).
Thus,dmf ≡ mf on a dense set. Therefore, sincedmf is continuous on X, we have
dmf = dmf . Thus, f ∈ D(t m) and t mf = tmf . This proves that tm ⊆ t m. By
Note that the sets reg(m), regb(m), reg∞
(m) and sing-suppr(m) remain un-
changed if m is replaced by its complex conjugate function m. Moreover, m = m.
Theorem 3. Let m : X → C. The operator tm is essentially defined if and only if
reg(m) is dense in X. In this case we have t∗m = tm and tm ∈ Co(E).
Proof. By Lemma 7 we can assume without loss of generality that m = m. Assume
that reg(m) is dense in X. Let g⊥D(tm). For x ∈ reg(m) there exists f ∈ D(tm)
such that f (x) 6= 0 by Theorem 7. From g(x)f (x) = hg, fi (x) = 0 we conclude
that g(x) = 0. Hence, by density of reg(m), g = 0. That is, D(tm)⊥ = {0}.
Conversely, suppose that reg(m) is not dense. Since X is a locally compact
Hausdorff space, there is a function g ∈ C0(X), g 6= 0, with support contained in
X \ reg(m). Then hf, gi = 0 for f ∈ D(tm) by Theorem 7 and so D(tm)⊥ 6= {0}.
Now assume that reg(m) is dense. We show that t∗m = tm. Let f ∈ D(tm) and
g ∈ D(tm). For x ∈ reg(m) = reg(m) we get
symmetry, tm = t m.
(cid:3)
htmf, gi (x) = m(x)f (x)g(x) = hf, tmgi (x).
12
REN ´E GEBHARDT AND KONRAD SCHM UDGEN
Since reg(m) is dense, we conclude that htmf, gi = hf, tmgi. Thus, tm ⊆ t∗m.
Let f ∈ D(t∗m). There exists h ∈ E such that
cmg(x)f (x) = htmg, fi (x) = hg, hi (x) = g(x)h(x),
reg(m). Since reg(m) is dense, we getdmf = h ∈ E. Thus f ∈ D(tm).
For x ∈ reg(m) we choose g ∈ D(tm) such that g(x) 6= 0. Hence h ≡ mf on
Finally, from reg(m) = reg(m) we easily derive t∗∗m = (tm)∗ = tm ∈ Co(E). (cid:3)
The domain D(tm) can be trivial if m is continuous on an dense set, since reg(m)
g ∈ D(tm), x ∈ X.
is empty if there is no open set of continuous points of m.
Lemma 8. Let m, n : X → C be two functions. Then:
(1) tm + tn ⊆ tm+n and tmtn ⊆ tmn.
(2) If reg(m) and reg(n) are dense in X, then tm + tn, tmtn ∈ C′o(E) and
(tm + tn)∗∗ = tm+n, (tmtn)∗∗ = tmn.
Proof. We verify the statements about the product; the proof for the sum is similar.
(1): Let f ∈ D(tmtn). Then reg(nf ) and reg(mcnf ) are dense and open in X,
so is their intersection which is contained in reg(mnf ). On this set, mnf ≡ mcnf .
Hence \(mn)f = [mcnf ]∧ ∈ E, that is, f ∈ D(tmn) and tmtnf = tmnf .
(2): First we prove that D(tmtn)⊥ = {0}. Let g⊥D(tmtn) and x ∈ reg(n) ∩
reg(m). Since the latter set is open and X is a locally compact Hausdorff space,
there exists f ∈ E such that its support is contained in this set and f (x) 6= 0.
Clearly, nf ∈ E and mnf ∈ E, hence f ∈ D(tmtn). But f (x)g(x) = hf, gi (x) = 0,
so g(x) = 0. Since reg(n) ∩ reg(m) is dense, g = 0. Thus, tmtn is essentially
defined. By (1), (tmtn)∗ ⊇ t∗mn = tmn, and tmn is essentially defined by Theorem
3, since reg(m) ∩ reg(n) ⊆ reg(mn) is dense.
Now we show that D((tmtn)∗) ⊆ D(tmn) which in turn implies the last assertion
(tmtn)∗∗ = t∗mn = tmn. Let f ∈ D((tmtn)∗). Then there exists g ∈ E such that
hf, tmtnhi = hg, hi for all h ∈ D(tmtn). Arguing as above, for x ∈ reg(n) ∩ reg(m)
there exists h ∈ D(tmtn) with h(x) 6= 0. Thereore, f mn ≡ g on the dense set
reg(m) ∩ reg(n). Hence [mnf = g ∈ E, that is, f ∈ D(tmn).
(cid:3)
Remark 2. Set m(x) := ei/x on X = [0, 1]. Then t∗mtm ( tm2.
Theorem 4. Let m : X → C. Suppose that reg(m) is dense in X. Then:
(1) tm is normal and t∗mtm is essentially self-adjoint.
(2) D(t∗mtm) is an essential core for tm.
(3) R(1 + t∗mtm) = {g ∈ E∀x ∈ sing-suppr(m) : g(x) = 0}. In particular,
R(1 + t∗mtm) is essential.
Proof. By Lemma 7 we can assume that m = m. Then tm ∈ Co(E) and t∗m = tm
by Theorem 3.
(1): Using Lemma 8(2) we get (t∗mtm)∗∗ = (tmtm)∗∗ = tm2. Since the latter is
Now we prove that tmtm = tmtm . Since both operators are restrictions of tm2,
it suffices to show that their domains coincide. By symmetry it is enough to prove
self-adjoint, t∗mtm is essentially self-adjoint.
that D(tmtm) ⊆ D(tmtm). Let f ∈ D(tmtm), that is, dmf ∈ E and [mdmf ]∧ ∈ E.
We have to show that dmf ∈ E and [mdmf ]∧ ∈ E. Clearly, [mdmf ]∧ = [m2f ]∧ =
[mdmf ]∧ ∈ E, since these functions coincide on reg(m). Further,
dmf2 = htmf, tmfi = htmtmf, fi = [m2f ]∧f.
UNBOUNDED OPERATORS ON HILBERT C ∗-MODULES
13
Hence [m2f ]∧(x) = 0 implies thatdmf (x) = 0 for x ∈ X. Using this fact it follows
that the function
2
h(x) :=dmf (x)(cid:16)\m2f (x)(cid:17)2
0
/(cid:12)(cid:12)(cid:12)\m2f (x)(cid:12)(cid:12)(cid:12)
, if \m2f (x) 6= 0,
, if \m2f (x) = 0
(2): We show that G(tm ↾D(t∗
belongs to E and coincides with mf on reg(m). Thus [mf ]∧ ∈ E.
mtm))⊥ ⊆ G(tm)⊥. Assume that h(g, h), (f, tmf )i =
0 for all f ∈ D(tmtm). Let x ∈ reg(m). Since X is a locally compact Hausdorff
space, there exists fx ∈ C0(X) with support contained in reg(m) and fx(x) 6= 0.
Clearly, fx ∈ D(tmtm). Then g(x)fx(x) = −h(x)tmfx(x) = −h(x)m(x)fx(x),
hence g ≡ −hm on reg(m). Now for all f ∈ D(tm) we have h(g, h), (f, tmf )i ≡
gf + hmf ≡ 0 on reg(m), that is, (g, h)⊥G(tm), since reg(m) is dense.
(3): Let x ∈ sing-suppr(m) = sing-suppr(m) and f ∈ D(tmtm). By Theorem
7, f (x) = 0, (tmf )(x) = 0 and (tmtmf )(x) = 0. Thus ((1 + t∗mtm)f )(x) = 0 which
proves one inclusion. Conversely, let g ∈ E with g(x) = 0 for all x ∈ sing-suppr(m).
The functions [1/(1 + m2)]∧, [m/(1 + m2)]∧ and [m2/(1 + m2)]∧ are bounded
and continuous on reg(m)∪ reg∞
∞(m). Hence, setting f := [g/(1 +m2)]∧, we have
gm
f ∈ E, dmf =(cid:20)
1 + m2(cid:21)∧
∈ E,
1 + m2(cid:21)∧
[mdmf ]∧ =(cid:20) gm2
∈ E,
by using that g ≡ 0 on sing-suppr(m). That is, f ∈ D(t∗mtm) and (1 + t∗mtm)f ≡ g
on reg(m), so that g ∈ R(1 + t∗mtm), since reg(m) is dense in X.
(cid:3)
The next theorem is one of our main results in this Section. It says that all es-
sentially defined orthogonally closable operators on E are multiplication operators.
Theorem 5. If t ∈ C′o(E), then there is a function m : X → C such that t = tm.
Proof. Let t ∈ C′o(E). We abbreviate D := D(t) and D∗ := D(t∗). We set
O := ∪f∈DOf , O∗ := ∪f∈D∗Of with Of := {x ∈ Xf (x) 6= 0}.
Further, since D and D∗ are essential, O and O∗ are dense in X. Hence O′ := O∩O∗
is also dense. For x ∈ X, we have
g(x)(tf )(x) = hg, tfi (x) = ht∗g, fi (x) = (t∗g)(x)f (x)
f ∈ D, g ∈ D∗.
If x ∈ O′, there are f ∈ D and g ∈ D∗ such that f (x) 6= 0 and g(x) 6= 0. Then
for
m(x) := (tf )(x)/f (x) = (t∗g)(x)/g(x).
In particular, this shows that m(x) is independent of the chosen functions f and
g and that m is continuous on O′. Now let f ∈ D and x ∈ O′. Then there is a
g ∈ D∗ such that g(x) 6= 0, so (tf )(x) = m(x)f (x). Similarly, (t∗g)(x) = m(x)g(x)
for g ∈ D∗ and x ∈ O′.
We now extend m arbitrarily to a function defined on the whole set X. It follows
that dmf ∈ E for f ∈ D and cmg ∈ E for g ∈ D∗. Then we have f ∈ D(tm) and
tmf = tf for f ∈ D. Likewise, g ∈ D(tm) and tmg = t∗g for g ∈ D∗. Thus, t ⊆ tm
and t∗ ⊆ tm. Therefore, tm = (tm)∗ ⊆ t∗∗ = t ⊆ tm, that is, tm = t.
(cid:3)
The next example gives a normal operator tm such that D(tm) 6= D(tm).
Example 4. Set X := [0, 1] and define m, f : X → C by
m(x) :=(ei/x/x , x 6= 0
, x = 0
0
,
f (x) :=(e−i/xx , x 6= 0
, x = 0
0
.
14
REN ´E GEBHARDT AND KONRAD SCHM UDGEN
Then reg(m) = (0, 1] and f ∈ C0(X). Further, dmf = 1, so f ∈ D(tm). On the
other side, (mf )(x) = e−2i/x for x ∈ (0, 1], so reg(dmf ) = (0, 1] and f /∈ D(tm).
This proves that D(tm) 6= D(tm). By Theorem 4, the operator tm is normal.
4. Graph regular operators
4.1. Definition and basics on graph regular operators. Graph regular oper-
ators are the most important new class of operators introduced in this paper.
Definition 11. An operator t : E → F is called
• graph regular if t is essentially defined and orthogonally closed and its graph G(t)
is orthogonally complemented in E ⊕ F ,
• regular if t is closed, D(t) is dense in E, D(t∗) is dense in F , and R(1 + t∗t) is
dense in E.
The preceding is the definition of a regular operator given in [L95, p. 96]. Each
regular operator is graph regular by [L95, Theorem 9.3].
By an equivalent definition, an operator t : E → F is graph regular if t is
closed, D(t) and D(t∗) are essential in E and F , respectively, and R(1 + t∗t) and
R(1 + tt∗) are dense in E and F , respectively. (The equivalence to Definition 11 is
easily verified by using some arguments from the proof of Theorem 6(1) below.)
We denote by Rgr(E, F ) the set of all essentially defined graph regular operators
and by R(E, F ) the set of regular operators from E into F . Let us abbreviate
Rgr(E) := Rgr(E, E) and R(E) := R(E, E).
A number of basic properties of graph regular operators are collected in the
following theorem.
Theorem 6.
(1) For t ∈ C′o(E, F ) the following statements are equivalent:
(a) t ∈ Rgr(E, F ).
(b) G(t) ⊕ vG(t∗) = E ⊕ F .
(c) R(1 + t∗t) = E and R(1 + tt∗) = F .
(2) If t ∈ Co(E, F ) and t∗ ∈ Rgr(F, E), then t ∈ Rgr(E, F ).
(3) If t ∈ Rgr(E, F ), then
(a) D(t∗t) is an essential core for t.
(b) t∗ ∈ Rgr(F, E).
Proof. (1a) ⇒ (1b) follows from the relation vG(t∗) = G(t)⊥ by Proposition 1.
Hence t is orthogonally closed and graph regular.
(1b) ⇒ (1a): Since G(t) ⊆ G(t)⊥⊥, (1b) clearly implies that G(t) = G(t)⊥⊥.
(1b) ⇒ (1c): If x ∈ E, then (x, 0) ∈ G(t)⊕vG(t∗), so there are y ∈ D(t), z ∈ D(t∗)
such that x = y + t∗z, 0 = ty − z. Then y ∈ D(t∗t) and x = (1 + t∗t)y ∈ R(1 + t∗t).
In the same way one shows that F = R(1 + tt∗).
(1c) ⇒ (1b): Since R(1+t∗t) = E, we have E⊕0 = R(1+t∗t)⊕0 ⊆ G(t)⊕vG(t∗).
For x ∈ D(t∗t), we set y := tx ∈ D(t∗). Then
((1 + t∗t)x, 0) = (x + t∗y, tx − y) ∈ G(t) ⊕ vG(t∗).
Similarly, 0 ⊕ F = 0 ⊕ R(1 + tt∗) ⊆ G(t) ⊕ vG(t∗). Thus E ⊕ F ⊆ G(t) ⊕ vG(t∗)
which yields G(t) ⊕ vG(t∗) = E ⊕ F .
(2) is obtained from t = t∗∗ and (1).
(3a): We have hx, yi + htx, tyi = hx, (1 + t∗t)yi for x ∈ D(t) and y ∈ D(t∗t), so
that G(t ↾D(t∗t))⊥ ∩ G(t) = {0}. With Lemma 2 it follows
G(t)⊥ = G(t ↾D(t∗t))⊥ ∩ (G(t) ⊕ vG(t∗)) = G(t ↾D(t∗t))⊥ ∩ E = G(t ↾D(t∗t))⊥,
so D(t∗t) is an essential core for t.
(3b): This follows from (2), since t and t∗ are orthogonally closed.
(cid:3)
UNBOUNDED OPERATORS ON HILBERT C ∗-MODULES
15
A special situation is treated in the following example.
Example 5. Let H be a separable Hilbert space, A a C∗-algebra of compact oper-
ators acting on H, and E a Hilbert A-module. Then we have
Co(E) = Rgr(E) = R(E).
Indeed, since R(E) ⊆ Rgr(E) ⊆ Co(E) by definition, it suffices to prove that
Co(E) ⊆ R(E). Let t ∈ Co(E). Since all closed submodules of any Hilbert C∗-
module over a C∗-algebra of compact operators are orthogonally complemented
[Mg97], all essential submodules are dense. Hence t and t∗ are densely defined
on E. Further, since t is closed, t is semiregular in the sense of [Pal99]. As shown
in [Pal99, Proposition 5.1], for any Hilbert A-module E of a C∗-algebra of compact
operators semiregular operators are always regular. Thus, t ∈ R(E).
In the very special case E = A = K(H) we have R(E) = C(H), since then the
regular operators on E are the affiliated operators with A and C(H) is the set of
affiliated operators with K(H) as noted in [W91].
Proposition 6. Let t ∈ Rgr(E, F ) and q ∈ L(E, F ). Suppose that r ∈ L(G, E)
and s ∈ L(F, G) are invertible with r−1 ∈ L(E, G) and s−1 ∈ L(G, F ). Then the
operators t + q, tr and st are essentially defined and graph regular.
Proof. Let pE and pF denote the projections from E ⊕ F onto E and F , respec-
tively. Clearly, t + q, tr, and st are essentially defined and orthogonally closed
by Proposition 2. In particular, their graphs are closed. Since t is graph regular,
G(t) is orthogonally complemented, so there is a projection p ∈ L(E ⊕ F ) with
R(p) = G(t). We now obtain
G(t + q) = {(x, tx + qx)x ∈ D(t)} = {(pEpv, pF pv + qpEpv)v ∈ E ⊕ F}
= R((pE, pF + qpE)p),
= R((r−1pE, pF )p),
= R((pE, spF )p).
G(tr) = {(r−1x, tx)x ∈ D(t)} = {(r−1pEpv, pF pv)v ∈ E ⊕ F}
G(st) = {(x, stx)x ∈ D(t)} = {(pEv, spF pv)v ∈ E ⊕ F}
Thus the closed subspaces G(t + q), G(tr), and G(st) are ranges of adjointable
operators, hence they are orthogonally complemented by [L95, Theorem 3.2]. (cid:3)
The next lemma describes a cases where graph regularity implies regularity.
Lemma 9. If t ∈ Rgr(E, F ), R(t) ⊆ D(t∗) and R(t∗) ⊆ D(t) , then t ∈ R(E, F ).
Proof. We have to prove that D(t) and D(t∗) are dense in E and F , respectively.
For D(t) this follows from the relations
E = R(1 + t∗t) ⊆ D(t) + R(t∗) ⊆ D(t).
Since t∗ ∈ Rgr(F, E) by Theorem 6, we can replace t by t∗ in the preceding and
obtain the density of D(t∗).
(cid:3)
4.2. The (a, a∗, b)-transform. In this section we establish a one-to-one correspon-
dence between graph regular operators and certain triples of adjointable operators.
As noted in Remark 1 this works in a purely algebraic setting and it neither requires
the C∗-condition nor even a norm.
Definition 12. For Hilbert A-modules E and F , let AB(E, F ) denote the set of
all triples (a, a∗, b) of operators a ∈ L(E), a∗ ∈ L(F ), b ∈ L(E, F ) such that a and
a∗ are self-adjoint, N (a) = {0}, N (a∗) = {0}, and
bb∗ = a∗ − a2
∗,
b∗b = a − a2,
ab∗ = b∗a∗.
16
REN ´E GEBHARDT AND KONRAD SCHM UDGEN
In particular 0 ≤ a ≤ I and 0 ≤ a∗ ≤ I in this case; further kbk ≤ 1.
We call the map t → (at, at∗ , bt) described in Theorem 7 the (a, a∗, b)-transform.
Theorem 7. If t ∈ Rgr(E, F ), then (at, at∗ , bt) ∈ AB(E, F ), where
at := (1 + t∗t)−1,
at∗ := (1 + tt∗)−1,
bt := t(1 + t∗t)−1.
Further, N (bt) = N (t), bt∗ = b∗t , and the projection onto the graph of t is given by
p =(cid:18)at
bt
b∗t
1 − at∗(cid:19) ∈ L(E ⊕ F, E ⊕ F ).
If (a, a∗, b) ∈ AB(E, F ), then ta,a∗,b ∈ Rgr(E, F ), where
∗ )∗,
ta,a∗,b := (ba−1)∗∗ = (b∗a−1
and we have t∗a,a∗,b = ta∗,a,b∗ . The map t 7→ (at, at∗ , bt) is a bijection from
Rgr(E, F ) onto AB(E, F ) with inverse (a, a∗, b) 7→ ta,a∗,b.
Proof. First we suppose that t ∈ Rgr(E, F ). Then R(1 + t∗t) = E, so at is defined
on the whole module E. It is straightforward to verify that 1 + t∗t is positive and
injective for each essentially defined operator t. Therefore at is positive and has
a trivial kernel. Analogous statements hold for at∗ . Further, bt is defined on E,
since R(at) ⊆ D(t). Similarly, bt∗ is defined on F . For x := (1 + t∗t)x′ ∈ E and
y := (1 + tt∗)y′ ∈ F , where x′ ∈ E, y′ ∈ F , we compute
hbtx, yi = htx′, (1 + tt∗)y′i = htx′, y′i + htx′, tt∗y′i = htx′, y′i + ht∗tx′, t∗y′i
= h(1 + t∗t)x′, t∗y′i = hx, bt∗ yi .
Hence bt = (bt∗ )∗ ∈ L(E, F ). From bt∗ = (bt)∗ = (tat)∗ we obtain
bt∗bt ⊇ att∗tat = at(1 − at).
t . Further,
bt∗ at∗ = t∗a2
Since at(1− at) is defined on the whole E, the latter yields bt∗ bt = at− a2
(1 + t∗t)t∗ = t∗(1 + tt∗) and R(a2
t∗ = 1 ↾D(t∗t) t∗a2
t∗ ) = D(tt∗tt∗) ⊆ D(t∗tt∗) imply that
t∗ = at(1 + t∗t)t∗a2
t∗ = att∗(1 + tt∗)a2
The preceding proves that (at, at∗ , bt) ∈ AB(E, F ).
Clearly, R(bt∗ ) ⊆ R(t∗), so N (t) ⊆ N (bt). Suppose that btx = 0 for some x ∈ E.
Then (at − a2
t )x = b∗t btx = 0, so x = atx ∈ D(t∗t) ⊆ D(t). Further, (1 + t∗t)x = x,
so t∗tx = 0 and from htx, txi = ht∗tx, xi = 0 it follows that x ∈ N (t). Thus,
N (t) = N (bt). The statement concerning the projection is easily verified.
Conversely, we now assume that (a, a∗, b) ∈ AB(E, F ). We define t := ba−1 and
∗ . Since D(t)⊥ = R(a)⊥ = N (a) = {0}, t is essentially defined. Similarly,
s := b∗a−1
s is essentially defined. For x ∈ E, y ∈ F we have
t∗ = atbt∗ .
ht(ax), a∗yi = hbx, a∗yi = ha∗bx, yi = hbax, yi = hax, b∗yi = hax, s(a∗y)i ,
Our next aim is to prove that R(a) is an essential core for s∗. Since s∗ = a−1
so t ⊆ s∗ and s ⊆ t∗. In particular, t ∈ C′o(E, F ).
Proposition 2, it suffices to show that G(ba−1)⊥ ⊆ G(a−1
Then h(r, s), (ax, bx)i = 0 for all x ∈ E, so ar + b∗s = 0. Further, we have
∗ b by
∗ b)⊥. Let (r, s) ∈ G(ba−1)⊥.
a∗(br + (1 − a∗)s) = a∗br + (a∗ − a2
∗)s = bar + bb∗s = b(ar + b∗s) = 0.
Since a∗ is injective, this yields s = a∗s − br.
Using the assumption b∗a∗ = ab∗ we obtain
Let x ∈ D(a−1
∗ b). Then there exists a (unique) element z ∈ F with bx = a∗z.
b∗z = a−1b∗a∗z = a−1b∗bx = a−1(a − a2)x = (1 − a)x.
UNBOUNDED OPERATORS ON HILBERT C ∗-MODULES
17
Now we compute
(cid:10)(r, s), (x, a−1
∗ bx)(cid:11) = hr, xi +(cid:10)s, a−1
∗ bx(cid:11) = hr, xi + hs, zi
= hr, xi + ha∗s − br, zi = hr, xi + hs, a∗zi − hr, b∗zi
= hr, xi + hs, bxi − hr, (1 − a)xi = hb∗s, xi + har, xi = 0.
∗ b). This proves that R(a) is an essential core for s∗.
Therefore, (r, s)⊥G(a−1
so that t∗∗ = s∗, that is, (ba−1)∗∗ = (b∗a−1
Since t ⊆ s∗ and D(t) = R(a) is an essential core for s∗, we have G(t)⊥⊥ = G(s∗),
∗ )∗. Finally, we derive
1 + t∗t∗∗ ⊇ 1 + t∗t = 1 + a−1b∗ba−1 = 1 + a−1(a − a2)a−1 = a−1,
1 + t∗∗t∗ ⊇ 1 + s∗s = 1 + a−1
∗ = a−1
∗ .
∗ = 1 + a−1
∗ (a∗ − a2
∗ bb∗a−1
∗)a−1
Hence at∗∗ = a ∈ L(E) and at∗ = a∗ ∈ L(F ), so that t ∈ Rgr(E, F ). Further, we
have bt∗∗ = t∗∗at∗∗ = t∗∗a ⊇ ta = b ∈ L(E, F ) and so bt∗∗ = b.
(cid:3)
t∗ bt = btan
From at ∈ L(E) it follows at once that the operator 1 + t∗t is self-adjoint for any
t ∈ Rgr(E, F ) by Proposition 1.
Lemma 10. If t ∈ Rgr(E, F ), then f (at∗ )bt = btf (at) for all f ∈ C([0, 1]).
Proof. The operators at, at∗ , bt are adjointable and at∗ bt = btat by Theorem 7.
Hence an
t for all n ∈ N, so f (at∗)bt = btf (at) for all polynomials f . Since
(cid:3)
the polynomials are uniformly dense in C([0, 1]), the assertion follows.
Lemma 11. Let t ∈ Co(E, F ) and suppose that t and t∗ are bounded. Then we
have t ∈ Rgr(E, F ) if and only if t ∈ L(E, F ).
Proof. The if direction is trivial. To prove the only if part assume that t is graph
regular. Since t is orthogonally closed, t is closed. Because t is closed and bounded,
the domain D(t) is closed in E. By Theorem 6(3b), t∗ is also graph regular. There-
fore, replacing t by t∗, it follows that D(t∗) is also closed. Hence D(t∗t) is closed.
Since t is graph regular, we have at ∈ L(E). Therefore, by [L95, Theorem 3.2],
R(at) = D(t∗t) is orthogonally complemented. But R(at)⊥ = N (at) = {0}, so
R(at) = E. In particular, E = D(t∗t) ⊆ D(t). Hence D(t) = E. By a similar
reasoning we obtain D(t∗) = F . Therefore, t ∈ L(E, F ).
(cid:3)
Corollary 1. Let t ∈ Rgr(E). Then t is normal if and only if at = at∗ . In this
case bt is normal and the operators at and bt commute.
Proof. Since t ∈ Rgr(E), we have (at, at∗ , bt) ∈ AB(E) by Theorem 7, so
b∗t bt = at − a2
t ,
btb∗t = at∗ − a2
t∗ ,
btat = at∗ bt.
The first statement is clear and in this case is b∗t bt = btb∗t and btat = atbt.
(cid:3)
In the next proposition E and F are Hilbert C∗-modules of (possibly different
!) C∗-algebras.
Proposition 7. Suppose that t ∈ Rgr(E) and φ ∈ Hom(L(E),L(F )). Then there
exists an orthogonally closed operator φ(t) : F → F such that φ(at)F is an essential
core for φ(t) and
φ(t)(φ(at)x) = φ(bt)x,
x ∈ F.
Moreover, hφ(t)x, yi = hx, φ(t∗)yi for x ∈ D(φ(t)), y ∈ D(φ(t∗)). If N (φ(at)) and
N (φ(at∗ )) are trivial, then φ(t) ∈ Rgr(F ), φ(t)∗ = φ(t∗), and
aφ(t) = φ(at),
aφ(t)∗ = φ(at∗ ),
bφ(t) = φ(bt).
18
REN ´E GEBHARDT AND KONRAD SCHM UDGEN
Proof. Clearly, 0 ≤ φ(at) ≤ I, 0 ≤ φ(at∗ ) ≤ I, and
φ(bt)∗φ(bt) = φ(at) − φ(at)2, φ(bt)φ(bt)∗ = φ(at∗ ) − φ(at∗ )2,
φ(at)φ(bt)∗ = φ(bt)∗φ(at∗ ),
since φ is a ∗-homomorphism. If x ∈ N (φ(at)), then
hφ(bt)x, φ(bt)xi = hx, φ(b∗t bt)xi =(cid:10)x, φ(at)x − φ(at)2x(cid:11) = 0,
so x ∈ N (φ(bt)). Therefore, the map φ(t)0 : φ(at)x 7→ φ(bt)x (x ∈ F ) is well-
defined. Similarly, the kernel of φ(at∗ ) is contained in the kernel of φ(b∗t ). Further,
it is easy to see that
G(φ(t∗)0) ⊆ vG(φ(t)0)⊥ = {(x, y) ∈ F ⊕ Fφ(b∗t )x = φ(at)y, y ∈ N (φ(at))⊥}
and the latter is the graph of an operator. Hence φ(t)0 and φ(t∗)0 are orthogonally
closable. If we denote the corresponding orthogonal closures by φ(t) and φ(t∗), the
first half of the proposition is shown. If the kernels of φ(at) and φ(at∗ ) are trivial,
then (φ(at), φ(at∗ ), φ(bt)) ∈ AB(F ) and all statements follow from Theorem 7. (cid:3)
If the kernel of φ(at) is not trivial, it can happen that the domain of the operator
φ(t) is only {0}, see Example 7 below.
Corollary 2. Let A be a (non-degenerated) concrete C∗-algebra on H. Let φ be
the embedding of L(A) = M(A) into B(H) = L(H).
such that φ(t) = T .
(1) For any T ∈ C(H) with aT , aT ∗ , bT ∈ M(A) there exists a unique t ∈ Rgr(A)
(2) If A contains the compact operators, then we have T := φ(t) ∈ C(H) and
aT , aT ∗ , bT ∈ M(A) for t ∈ Rgr(A). In particular, Rgr(A) can be identified
with those T ∈ C(H) for which aT , aT ∗ , bT ∈ M(A).
Proof. (1): Since C(H) = Rgr(H), we have (aT , aT ∗ , bT ) ∈ AB(H) by Theorem 7.
By assumption aT , aT ∗ , and bT are elements of M(A). To show that (aT , aT ∗ , bT ) ∈
AB(A) it suffices to prove that aT and aT ∗ are injective as operators on A. Clearly,
they are injective as operators on H. Assume that aT a = 0 for some a ∈ A. Then,
aT aξ = 0 for ξ ∈ H. Hence aξ = 0 for all ξ ∈ H, so that a = 0. Thus, aT is
injective on A. Similarly, aT ∗ is injective on A. Using once more Theorem 7 it
follows that there exists an operator t ∈ Rgr(A) such that at = aT , at∗ = aT ∗ ,
bt = bT . Further, φ(t) = T , since
T (φ(at)ξ) = T (atξ) = T aT ξ = bT ξ = btξ = φ(bt)ξ,
ξ ∈ H,
and R(aT ) = D(T ∗T ) is a core for T .
(2): If t ∈ Rgr(A), then (at, at∗ , bt) ∈ AB(A). We show that the kernels of φ(at)
and φ(at∗ ) are trivial. Assume that φ(at)ξ = 0 for some nonzero vector ξ ∈ H.
Since A contains all compact operators, the rank one projection pξ onto C·ξ is in
A. Therefore, atpξ = φ(at)pξ = 0 which contradicts the injectivity of at as operator
on A. Hence at, similarly at∗, is injective on H. Therefore, T ∈ Rgr(H) = C(H)
(cid:3)
by Proposition 7.
Example 6. From Corollary 2 it follows immediately that Rgr(K(H)) = C(H) and
Rgr(B(H)) = C(H), since M(K(H)) = M(B(H)) = B(H).
4.3. Quotients of adjointable operators. A large class of examples of un-
bounded graph regular operators can be obtained as quotients of adjointable oper-
ators.
UNBOUNDED OPERATORS ON HILBERT C ∗-MODULES
19
x ∈ G,
D(t) = R(a),
Theorem 8. Let a ∈ L(G, E) and b ∈ L(G, F ). Suppose that N (a) ⊆ N (b) and
N (a∗) = {0}. If the operator t : E → F defined by
t(ax) := bx,
is closed, then t ∈ Rgr(E, F ) and t∗ = (a∗)−1b∗.
Proof. Since N (a) ⊆ N (b) and R(a)⊥ = N (a∗) = {0}, t is well-defined and essen-
tially defined. Since the graph of t is the set {(ax, bx)x ∈ G}, it is the range of the
adjointable operator q : G → E ⊕ F defined by q(x) := (ax, bx). Since this range
is closed by assumption this range, [L95, Theorem 3.2] applies and shows that the
range is orthogonally complemented. Hence t is graph regular. The adjoint of t is
(cid:3)
then easily computed; we omit the details.
Corollary 3. Let x ∈ L(F, E) and assume that N (x) = N (x∗) = {0}. Then
x−1 ∈ Rgr(F, E) and (x−1)∗ = (x∗)−1.
Proof. Since x−1 is closed, the assertion follows from Theorem 8 by letting b the
(cid:3)
identity on F .
Corollary 4. Let a ∈ L(E) and let p, q ∈ C[X] be relatively prime. Assume that
R(q(a)) is essential and N (q(a)) ⊆ N (p(a)). Let t : E → E be the operator defined
by
D(t) := R(q(a)),
t(q(a)x) := p(a)x,
x ∈ E.
Then t is graph regular.
Proof. In order to apply Theorem 8 we only have to prove that t is closed. Let
(xn)n∈N be a sequence in E such that p(a)xn → xp ∈ E and q(a)xn → xq ∈ E.
Since p and q are relative prime, there are polynomials p, q ∈ C[X] such that
pp + qq = 1. Then
xn = (p(a)p(a) + q(a)q(a))xn → p(a)xp + q(a)xq =: xr ∈ E,
so xp = p(a)xr and xq = q(a)xr. That is, t is closed.
Theorem 9. Let a ∈ L(G, E), b ∈ L(G, F ), a∗ ∈ L(H, F ), and b∗ ∈ L(H, E) be
such that b∗a∗ = a∗b∗. Assume that N (a∗) = N (a∗
∗) = {0}. Then N (a) ⊆ N (b)
and N (a∗) ⊆ N (b∗). The operators t and t′ defined by
(cid:3)
D(t) := R(a),
D(t′) := R(a∗),
t(ax) := bx,
t′(a∗y) := b∗y,
x ∈ G,
y ∈ H,
are essentially defined, orthogonally closable and they satisfy (t′)∗∗ ⊆ t∗, t∗∗ ⊆ (t′)∗.
If in addition t and t′ are closed and ab∗ = b∗a∗
∗, then t∗ = t′ and t ∈ Rgr(E, F ).
Proof. Suppose that a∗x = 0 for some x ∈ H. Then 0 = b∗a∗a = a∗b∗x and hence
b∗x = 0, since N (a∗) = {0}. This shows that N (a∗) ⊆ N (b∗). In a similar manner,
the assumption N (a∗
∗) = {0} implies that N (a) ⊆ N (b). Hence the operators t and
t′ are well-defined. It is obvious that t and t′ are essentially defined.
From the relations (a∗)−1b∗a∗ = b∗ and t∗y = (a∗)−1b∗y, y ∈ G, we get t′ ⊆ t∗.
Since t′ ⊆ t∗ and t′ is essentially defined, so is t∗. Since t is also essentially defined,
Interchanging the role of t and t′ we conclude that
t is orthogonally closable.
t ⊆ (t′)∗ and t′ is orthogonally closable. Applying the involution to the relations
t ⊆ (t′)∗ and t′ ⊆ t∗ we obtain (t′)∗∗ ⊆ t∗ and t∗∗ ⊆ (t′)∗ which proves the first half
of the proposition.
∗. Since t and t′ are closed,
Now suppose that t and t′ are closed and ab∗ = b∗a∗
G := G(t) ⊕ vG(t′) = {(ax + b∗y, bx − a∗y)x ∈ E, y ∈ F}
20
REN ´E GEBHARDT AND KONRAD SCHM UDGEN
is a closed submodule of E ⊕ F . We define q(x, y) := (ax + b∗y, bx − a∗y) for
(a, b) ∈ E ⊕ F . Then q ∈ L(E ⊕ F ) and R(q) = G. By [L95, Theorem 3.2], G is
orthogonally complemented. It is easily calculated that
q∗(x′, y′) = (a∗x′ + b∗y′, b∗
∗x′ − a∗
∗y′)
for
(x′, y′) ∈ E ⊕ F.
We show that N (q∗) = {0}. Suppose that q∗(x′, y′) = 0. Then a∗x′ + b∗y′ = 0 and
a∗
∗y′ − b∗
∗x′ = 0, so we obtain
0 = aa∗x′ + ab∗y′ = aa∗x′ + b∗a∗
∗x′, b∗
∗y′ = aa∗x′ + b∗b∗
∗x′i = 0. Thus, a∗x′ = 0 and b∗
The latter implies that ha∗x′, a∗x′i + hb∗
∗x′ =
∗y′ = 0. Therefore, x′ = 0 and y′ = 0 by the assumption N (a∗) = N (a∗
a∗
∗) = {0}.
That is, N (q∗) = {0}. Hence, G⊥ = R(q)⊥ = N (q∗) = {0}. Therefore, since G is
orthogonally complemented, we have G = E ⊕ F . This proves that t′ = t∗.
(cid:3)
4.4. Absolute value. The next theorem is concerned with the absolute value of
graph regular operators.
Theorem 10. Suppose that t ∈ Rgr(E) and define
∗x′.
D(t) := R(a1/2
t
),
t(a1/2
t x) := (1 − at)1/2x,
x ∈ E.
Then t ∈ Rgr(E) is self-adjoint and positive. Further, we have t2 = t∗t, at = at,
and bt = bt.
Proof. Clearly, t is essentially defined. We prove that t is closed. Let (xn) be
a sequence in E such that a1/2
t xn → x ∈ E and (1 − at)1/2xn → y ∈ E. Then
xn = (1− at)xn + atxn → a1/2
t x + (1− at)1/2y =: x′, so that x = a1/2
t x′. This shows
that t is closed. Therefore, t is graph regular by Proposition 8.
and b = b∗ = (1 − at)1/2 we conclude
that t = t∗, since the relations b∗a∗ = a∗b∗ and ab∗ = b∗
∗a∗ are fulfilled. Further,
D(t2) ⊇ R(at) = D(t∗t). It is easily checked that (1 + t2)at = 1 = (1 + t∗t)at,
so t2 ⊇ t∗t. Since t∗t is self-adjoint and t2 = t∗t is symmetric, we obtain
t2 = t∗t. We derive
Applying Theorem 9 with a = a∗ = a1/2
t
t xE =D(1 − at)1/2, a1/2
t x), a1/2
Dt(a1/2
bt = tat = ta1/2
t xE =D(at − a2
t = (at − a2
t )1/2x, xE ≥ 0 for x ∈ E,
t )1/2 = (b∗t bt)1/2 = bt.
so t is positive. Clearly, we have at = at. Finally, we compute
t a1/2
t = (1 − at)1/2a1/2
(cid:3)
Remark 3. In contrast to the Hilbert space case the domains D(t) and D(t) do
not coincide in general, even more, neither D(t) ⊆ D(t) nor D(t) ⊇ D(t) holds.
Indeed, let E = A := C([0, 1]) and set m(x) := x−1ei/x for x ∈ (0, 1]. Then,
by Theorem 15 below, the operator tm is graph regular, since reg(m) = (0, 1] and
sing-suppr(m) = ∅. It is easily verified that tm = tm. Define f (x) := xe−i/x for
x ∈ (0, 1] and f (0) = 0. Then f ∈ D(tm), but f /∈ D(tm). The function g(x) = x
is in D(tm), but not in D(tm).
4.5. Bounded transform. The main results of this section give a characterization
of graph regular operators in terms of the bounded transform.
Let us begin with some important notation. Let Z(E, F ) denote the set of all
z ∈ L(E, F ) such that kzk ≤ 1 and N (I − z∗z) = {0} and let Z d(E, F ) be the set
of those z ∈ Z(E, F ) for which R(I − z∗z) is dense in E.
In [L95, Lemma 10.3] it was shown that z ∈ Z d(E, F ) implies that z∗ ∈ Z d(F, E).
The following lemma contains the analogues statement for Z(E, F ).
UNBOUNDED OPERATORS ON HILBERT C ∗-MODULES
21
Lemma 12. If z ∈ L(E, F ), then N (I−z∗z) = {0} if and only if N (I−zz∗) = {0}.
In particular, z ∈ Z(E, F ) if and only if z∗ ∈ Z(F, E).
Proof. It suffices to show one direction, since z can be replaced by z∗. Assume that
x ∈ N (I − zz∗) \ {0}. Then kz∗xk2 = hx, zz∗xi = hx, xi = kxk2. Hence z∗x 6= 0.
But (I − z∗z)z∗x = z∗(I − zz∗)x = 0, so z∗x ∈ N (I − z∗z) 6= {0}.
(cid:3)
For z ∈ Z(E, F ) we define an operator tz : E → F by
D(tz) := (I − z∗z)1/2E,
tz(I − z∗z)1/2x := zx,
x ∈ E.
Since R((I − z∗z)1/2)⊥ = N ((I − z∗z)1/2) = N (I − z∗z) = {0}, tz is essentially
defined. Clearly, R(z) = R(tz).
Theorem 11. If z ∈ Z(E, F ), then tz ∈ Rgr(E, F ). Further, the mapping z → tz
is injective from Z(E, F ) into Rgr(E, F ). We have t∗z = tz∗ and
atz = I − z∗z,
z = tza1/2
tz .
Proof. First we prove that the operator tz is closed. Let (xn) be a sequence of D(z)
such that ((I − z∗z)∗)1/2xn → x and zxn → y. Then
xn = (I − z∗z)∗xn + z∗zxn → ((I − z∗z)∗)1/2x + z∗y,
so x ∈ D(tz) and tzx = y. Thus, tz is closed and tz ∈ Rgr(E, F ) by Proposition 8.
Set a = (1 − z∗z)1/2, b = z, a∗ = (1 − zz∗)1/2, b∗ = z∗. Since z∗(1 − zz∗)1/2 =
(1− z∗z)1/2z∗, we then have b∗a∗ = a∗b∗ and ab∗ = b∗a∗
∗. Hence Theorem 9 applies
and yields z∗t = zt∗. Finally, by Proposition 2(4), t∗z = ((I − z∗z)1/2)−1z∗, so
t∗ztz = ((I − z∗z)1/2)−1z∗z((I − z∗z)1/2)−1
= ((I − z∗z)1/2)−1(I − z∗z)((I − z∗z)1/2)−1 − (((I − z∗z)1/2)−1)2
= I − ((I − z∗z)1/2)−1((I − z∗z)1/2)−1.
Therefore, (I + t∗ztz)−1 = I − z∗z. In particular, D(t∗ztz)⊥ = R(I − z∗z)⊥ = {0}
and ((I + t∗ztz)−1)∗ = (I + t∗ztz)−1. This also implies that t∗ztz is self-adjoint and
tz((I + t∗ztz)−1)1/2 = z.
(cid:3)
According to [L95, Theorem 10.4], the mapping z 7→ tz is a bijection from the
set Z d(E, F ) onto the set R(E, F ) of regular operators. For the extended mapping
acting on Z(E, F ) the situation is more subtle. It is still an injective mapping into
the set Rgr(E, F ) of graph regular operators, but it is not sujective as shown by
Example 7 below.
Lemma 13. If z ∈ Z(E, F ), then z ∈ Z(E) and tz = tz. Further, we have
N (z) = N (tz) = N (z) = N (tz).
Proof. Since 1 − z∗z = 1 − z2 and kzk = kzk ≤ 1, the first statement is clear.
Further, atz = (1 + t∗ztz)−1 = 1 − z∗z, so
D(tz) = R(a1/2
tz ) = R((1 − z∗z)1/2) = R(1 − z2) = D(tz),
tz(1 − z∗z)x = (1 − atz )1/2x = (z∗z)1/2x = zx = tz(1 − z2)x,
x ∈ E,
that is, tz = tz. Since kernels of orthogonally closed operators are orthogonally
closed, we obtain N (z) = R(z∗)⊥ = R(tz∗ )⊥ = R(t∗z)⊥ = N (tz). Because N (z) =
N (z), this completes the proof.
(cid:3)
The following lemma restates [L95, Proposition 3.7]. It will be used several times
in the proof of Theorem 12 below.
Lemma 14. If a ∈ L(E)+, then R(a) = R(aε) for any ε > 0.
22
REN ´E GEBHARDT AND KONRAD SCHM UDGEN
t
) is contained in its domain. Its adjoint is an extension of a1/2
Now suppose that t ∈ Rgr(E, F ). Recall that at ∈ L(E)+, at∗ ∈ L(F )+, and
is essentially defined, since
t∗, hence it
. Hence
bt = b∗t∗ ∈ L(E, F ) by Theorem 7. The operator ta1/2
R(a1/2
is also essentially defined. Moreover, by Proposition 2(3), (a1/2
ta1/2
Definition 13. The bounded transform zt of t ∈ Rgr(E, F ) is defined by
is orthogonally closed; in particular, ta1/2
t∗)∗ = ta1/2
is closed.
t
t
t
t
t
t
zt := ta1/2
t
↾
D(t∗t) .
Since t ∈ Rgr(E, F ), D(t∗t) is essential in E and D(tt∗) is essential in F .
Theorem 12. Suppose t ∈ Rgr(E, F ). Set E0 := D(t∗t) and F0 := D(tt∗). Then
zt ∈ Z d(E0, F0) and z∗t = zt∗, where the adjoint z∗t is taken in L(E0, F0). Further,
t∗ t ↾
zt = a1/2
at = (1 − zt∗ zt)∗,
D(t∗t), N (zt) = N (t), R(zt) ⊆ R(t),
bt = zta1/2
at∗ = (1 − ztzt∗ )∗,
t
,
and treg := tzt ↾E0 is a regular operator from E0 to F0 satisfying treg ⊆ t = (tzt )∗∗,
where t∗∗zt is the biadjoint of the operator tzt : E → F .
Proof. From Theorem 7 we already know that t = a−1
t∗ bt. Using this we derive
(2)
Here the second equality follows from Lemma 10 applied with f (x) = √x.
t∗ bt ⊆ a−1
t∗ bt = a−1
t∗ bta1/2
t = ta1/2
a1/2
t∗ t = a1/2
t∗ a−1
t∗ a1/2
.
t
Now we prove that a1/2
t∗ t is bounded with norm not exceeding 1. Let x ∈ D(t).
Using that t∗at∗ t = bt∗t = b∗t t ⊆ (t∗bt)∗ = (t∗tat)∗ = (1 − at)∗ = 1 − at, we derive
t∗ txk2 = kDa1/2
ka1/2
t∗ tx, a1/2
t∗ txEk = k ht∗at∗ tx, xi k
= k h(1 − at)x, txi k ≤ k1 − atkkxk2 ≤ kxk2.
t∗ t and ta1/2
t
concides on D(t) and so on its subspace
is closed, we conclude
By (2) the operators a1/2
D(t∗t). Since both operators are bounded on D(t∗t) and ta1/2
that zt ≡ ta1/2
D(t∗t) = a1/2
D(t∗t) .
t∗ t ↾
↾
t
t
The latter equality implies that R(zt) is contained in R(a1/2
t∗ ). Lemma 14 yields
R(a1/2
t∗ ) = R(at∗ ) = D(tt∗) = F0, so that R(zt) ⊆ F0. Hence zt becomes a bounded
operator from E0 into F0. Analogously, zt∗ is a bounded operator from F0 into E0.
We show that z∗t = zt∗. Let z∗E
denote the adjoint of zt considered as an
t
operator from E into F . Clearly, z∗E
t∗ ⊇ zt∗. This implies that
z∗t = z∗E
t
t ⊇ (a1/2
↾F0= zt∗ . Therefore, zt ∈ L(E0, F0).
t∗ t)∗ ⊇ t∗a1/2
Next we verify the formulas for at, at∗ , and bt. First we note that
t∗
↾F0 a1/2
zt∗ zt ↾D(t∗t) = t∗a1/2
(3)
Hence at = (1 − zt∗ zt)∗∗ by Example 3, so at = a∗t = (1 − zt∗zt)∗∗∗ = (1 − zt∗zt)∗.
The formula for at∗ is proven in a similar manner. Since R(a1/2
) ⊆ E0, we obtain
zta1/2
t∗ t ↾D(t∗t)= t∗at∗ t ↾D(t∗t)= 1 − at ↾D(t∗t) .
t
t
↾E0 a1/2
t = tat = bt.
t = ta1/2
Using relation (3) and again Lemma 14 we get
D(t∗t)) = R(at ↾
R(1 − zt∗ zt) = R(at ↾
= R(a2
t ) = R(at) = D(t∗t) = E0.
R(at)) ⊇ R(at ↾R(at))
Hence zt ∈ Z d(E0, F0). Therefore, treg = tzt is a regular operator from E0 to F0.
Clearly, treg ⊆ t.
UNBOUNDED OPERATORS ON HILBERT C ∗-MODULES
23
Now we prove that t = (tzt )∗∗. Since tzt ⊆ t and
D(tzt ) = R((1 − z∗t zt)1/2) = R(at ↾
D(t∗t)) = R(at ↾
R(at)) ⊇ R(a2
t ),
it suffices to show that R(a2
for some x ∈ D(t). Then, for all y ∈ E,
t ) is an essential core for t. Assume (x, tx)⊥G(t ↾R(a2
t ))
0 =(cid:10)(x, tx), (a2
t y, ta2
t y)(cid:11) =(cid:10)x, a2
t y(cid:11) = hatx, atyi + htx, btatyi
= hatx, atyi + hb∗t tx, atyi = hatx + b∗t tx, atyi
t y(cid:11) +(cid:10)tx, ta2
Therefore, since R(at) is essential, 0 = (at + b∗t t)x = (at + 1 − at)x = x. That is,
t ))⊥ = {0}. Since t is graph regular and hence G(t) ⊕ G(t)⊥ = E, it
G(t) ∩ G(t ↾R(a2
follows from Lemma 2 that
G(t ↾R(a2
t ))⊥ = G(t ↾R(a2
t ))⊥ ∩ (G(t) ⊕ G(t)⊥) = G(t)⊥.
t ) is an essential core for t which completes the proof of the equality
Thus R(a2
t = (tzt)∗∗.
Clearly, R(zt) ⊆ R(t). Finally, we show that N (zt) = N (t). Let x ∈ N (t) ⊆
D(t∗t). Since zt ⊇ a1/2
t∗ t ↾D(t∗t), we have x ∈ D(zt) and ztx = 0, so N (t) ⊆ N (zt).
Conversely, let x ∈ N (zt). Then (1 − at)x = zt∗ ztx = 0 and x = atx ∈ D(t∗t) ⊆
D(t). We obtain (1 + t∗t)x = x and t∗tx = 0. From htx, txi = ht∗tx, xi = 0 we get
tx = 0. Hence N (zt) ⊆ N (t).
(cid:3)
Definition 14. The operator treg := tzt ∈ R(E0, F0) from Theorem 12 is called
the regular operator associated with the graph regular operator t ∈ Rgr(E, F ).
Remark 4. There are two other possibilities to define the bounded transform for a
graph regular operator t and both of them are natural in some sense. Define
z′t := ta1/2
t
and z′′t := (bt(a1/2
t
)−1)∗∗.
t
)−1 is the restriction of z′t = ta1/2
Note that z′t ∈ Co(E, F ) and zt is the restriction of z′t to D(t∗t). It is easily seen
that bt(a1/2
)−1 is
essentially defined and orthogonally closable and its orthogonal closure may also be
taken as a bounded transform z′′t of t. Then z′t = z′′t
) is an
essential core for z′t. We were not able to prove or disprove this. But one can show
that
if and only if R(a1/2
). Hence bt(a1/2
to R(a1/2
t
t
t
t
zt ⊆ z′′t ⊆ z′t,
(z′′t )∗ = z′t∗ ≡ t∗a1/2
t∗ ,
(z′t)∗ = z′′t∗ ≡ (bt∗ (a1/2
t∗ )−1)∗∗.
Note that the operator t can be recovered from both transforms z′t and z′′t as well.
We dont know wether or not the equality (ta1/2
Remark 5. The relationship between t ∈ Rgr(E, F ), treg ∈ R(E0, F0), and their
bounded tranforms should be studied be further in detail. Despite of this it seems
that graph regular operators form an important notion in its own, because they act
on the given Hilbert C∗-module E. Though each symmetric operator on a Hilbert
space is a restriction of a self-adjoint operator in possibly larger space, symmetric
operators are a basic concept.
)∗ = t∗a1/2
t∗ holds.
t
4.6. Polar decomposition.
Definition 15. Let E′ ⊆ E and F ′ ⊆ F be orthogonally closed. An operator
v ∈ Co(E, F ) is called a partial isometry with initial space E′ and final space F ′ if
v∗v is the projection pE′ and vv∗ is the projection pF ′.
In this case v∗ is also a partial isometry with initial space F ′ and final space
E′. Using this general notion of essentially defined partial isometries there is the
following theorem on a polar decomposition of adjointable operators.
24
REN ´E GEBHARDT AND KONRAD SCHM UDGEN
Theorem 13. Let t ∈ L(E, F ). Then there is a partial isometry v ∈ Co(E, F )
with initial space R(t∗) and final space R(t) such that t = vt, t = v∗t and
R(v) = R(t), R(v∗) = R(t∗), N (v) = N (t), and N (v∗) = N (t∗) if and only if
R(t) and R(t∗) are orthogonally closed.
Proof. The only if direction follows easily from the definition of a partial isometry.
To prove the if part we assume that R(t) and R(t∗) are orthogonally closed.
Define a map w : R(t) → R(t) by w(tx) := tx for x ∈ E. Then w is well-defined
and isometric, since htx, txi = htx,txi for x ∈ E. The continuous extension of w
to a map from R(t) onto R(t) is also an isometry which is denoted again by w.
Now we define v : E → F by
v(x + y) := wx,
x ∈ R(t), y ∈ N (t).
Clearly, D(v)⊥ = (N (t) ⊕ R(t))⊥ = N (t)⊥ ∩ R(t)⊥ = N (t)⊥ ∩ N (t) = {0}, so
v is essentially defined. Further, t = vt, N (v) = N (t) and R(v) = R(t). Let
v′(x + y) := w−1x,
x ∈ R(t), y ∈ N (t∗).
(cid:10)v(x + x⊥), y(cid:11) =(cid:10)x + x⊥, z(cid:11) for all x ∈ R(t) and x⊥ ∈ N (t). Choosing x = 0, we
As above, v′ is essentially defined, t = v′t, N (v′) = N (t∗) and R(v′) = R(t). It is
easily seen that v′ ⊆ v∗ and v ⊆ (v′)∗. Therefore, since v′ is essentially defined, so
is v∗. Hence v is orthogonally closable by Theorem 1.
We show that v∗ = v′. Let y ∈ D(v∗). Then there is an element z ∈ E such that
conclude that z ∈ N (t)⊥ = R(t∗)⊥⊥ = R(t)⊥⊥ = R(t). Thus, R(v∗) ⊆ R(v′).
Putting now x⊥ = 0, we get htx′, yi = hvtx′, yi = htx′, zi for all x′ ∈ E. Hence
t∗y = tz = tv∗y and N (v∗) ⊆ N (t∗) = N (v). Now v′ ⊆ v∗, R(v∗) ⊆ R(v′) and
N (v∗) ⊆ N (v′) clearly imply that v′ = v∗. In a similar manner it is shown that
v′∗ = v. Obviously, v∗v is the projection onto the orthogonally closed submodule
R(t∗) and vv∗ the projection onto R(t).
(cid:3)
Theorem 14. Let z ∈ Z(E, F ). There exists a partial isometry v ∈ Co(E, F ) with
initial space R(tz) and final space R(tz) such that
tz = v∗tz,
tz = vtz,
R(v) = R(tz), R(v∗) = R(t∗z), N (v) = N (tz), and N (v∗) = N (t∗z) if and only if
R(z) and R(z∗) are orthogonally closed. In this case, z = vz.
Proof. Since R(z) = R(tz), R(z∗) = R(t∗z), N (z) = N (tz), and N (z∗) = N (t∗z),
one direction follows at once from the definition of a partial isometry.
Conversely, assume that R(z∗) = R(z) and R(z) are orthogonally closed. By
Proposition 13, there is a partial isometry v ∈ Co(E, F ) with z = vz, z = v∗z
and R(v) = R(z), R(v∗) = R(z∗), N (v) = N (z) and N (v∗) = N (z∗). Finally,
vtz(1 − z2)1/2x = vzx = zx = tz(1 − z2)1/2x for x ∈ E, so that vtz = tz.
Similarly, v∗tz = tz.
(cid:3)
Note that the partial isometry belongs to L(E, F ) if and only if R(z) and R(z∗)
are orthogonal complements: D(v) = R(z∗) ⊕ N (z) and D(v∗) = R(z) ⊕ N (z∗).
4.7. Graph regular operators on C0(X). Before we turn to the functional cal-
culus of graph regular normals we reconsider the commutative case from Section 3.
By Theorem 5, Co(C0(X)) consists of multiplication operators. Theorem 15 charac-
terizes the graph regular operators among them as those for which sing-suppr(m)
is empty.
Lemma 15. For any function m : X → C the following are equivalent:
UNBOUNDED OPERATORS ON HILBERT C ∗-MODULES
25
(1) tm is injective if and only if {x ∈ reg(m)m(x) 6= 0} is dense in reg(m).
(2) If tm is injective and m does not vanish on X, then t−1
m = t1/m.
Proof. (1): Set N := {x ∈ reg(m)m(x) = 0}. Assume that N contains a nonempty
open set U . Since X is locally compact and Hausdorff, there is a non-zero function
f ∈ E with support contained in U . Then mf = 0, f ∈ D(tm) and tmf = 0. Hence
tm is not injective. On the other hand, assume that f ∈ D(tm) and tmf = 0. Then
m(x)f (x) = (tmf )(x) = 0 for x ∈ reg(m) . Thus f ≡ 0 on reg(m) \ N . If the
latter is dense in reg(m), then f ≡ 0 on reg(m) by the continuity of f . Futher,
we have f ≡ 0 on X \ reg(m) by Lemma 7, since f ∈ D(tm). That is, f = 0.
(2): We have reg(1/m) = reg(m). In particular, t1/m is injective. From Lemma
8(1) it follows that tmt1/m and t1/mtm are restrictions of the identity. Therefore,
t1/m ⊆ t−1
m ⊆ t1/m, so equality is
(cid:3)
proven.
Theorem 15. If m : X → C, then tm ∈ Rgr(C0(X)) if and only if reg(m) is
dense in X and sing-suppr(m) is empty. In this case we have t∗mtm = tm2 and
1/m. The last inclusion gives t−1
m and tm ⊆ t−1
Proof. The first assertion is clearly a corollary of Theorem 4.
atm = t
,
1
1+m2
btm = t m
1+m2 .
Suppose that tm ∈ Rgr(E). Then atm is self-adjoint, hence t∗mtm is self-adjoint.
Further, by Lemma 8, t∗mtm is contained in the self-adjoint operator tm2. Hence
t∗mtm = tm2. Finally, using Lemma 8(2) and Lemma 15(2), we compute
,
atm = (1 + t∗mtm)−1 = (1 + tm2)−1 = (t1+m2)−1 = t
btm = b∗∗tm = (tmatm)∗∗ = (tmt
)∗∗ = t m
1+m2
1
1
1+m2 .
1+m2
(cid:3)
Corollary 5. Suppose that m : X → C is bounded and reg(m) dense in X.
In particular, we have
Then tm is graph regular if and only if reg( m) = X.
tm ∈ L(C0(X)) = Cb(X) in this case.
Proof. Since m is bounded, reg∞
from Theorem 15.
(m) is empty. Hence the statement follows directly
(cid:3)
The next example shows that not all operators t ∈ Rgr(C0(R)) are of the form
tz for some z ∈ Z(C0(R)). Moreover there is a representation π of L(C0(R)) such
that the domain of π(t) consists only of the zero element.
Example 7. Consider the operator t := tm on C0(R), where m(x) = 1/x for x 6= 0.
Then t is self-adjoint and graph regular by Theorem 15.
We show that there is no z ∈ Z(C0(R)) such that t = tz. Assume to the
by Theorem 11. Further,
t = t1/(√1+m2 ). Choose g ∈ C0(R)
x xg(x)
√1+x2 is
t g /∈ D(t).
Let π be the ∗-representation of L(C0(R)) = {tnn ∈ Cb(R)} given by π(tn) =
contrary that t = tz for z ∈ Z(C0(R)). Then z = ta1/2
we have at = t1/(1+m2) and therefore a1/2
t g(x) = xg(x)/(√1 + x2). Since the function 1
with g(0) 6= 0. Then a1/2
continuous on R\{0} and has no continuous extension to R, we have a1/2
Hence ta1/2
is not everwhere defined which contradicts the equality z = ta1/2
n(0). Then π(at) = 0, so the domain of π(t) is {0} (compare Proposition 7).
Example 8 (Continuing Example 4). Recall that the operator tm from Example
4 is normal, but D(tm) 6= D(t∗m). Since reg(m) = (0, 1] and 0 ∈ reg∞
(m), tm
is graph regular by Theorem 15. That is, even for graph regular operators t the
statements
.
t
t
t
26
REN ´E GEBHARDT AND KONRAD SCHM UDGEN
(1) t∗t = tt∗ (that is, t is normal),
(2) D(t) = D(t∗) and ktfk = kt∗fk for all f ∈ D(t),
are not equivalent!
4.8. Functional calculus of graph regular normal operators. Let A and B be
C∗-algebras, where A is non-unital and B is unital. Clearly, each ∗-homomorphism
φ : A → B extends uniquely to a ∗-homomorphism of the unitization A∼ := A ⊕ C
by φ(a + α) := φ(a) + α1 for a ∈ A, α ∈ C.
Let ζ denote the identity map of C. Considered as an operator on C0(C), ζ is a
regular operator, but on the unitization C0(C)∼ the operator ζ is no longer regular.
On the other hand, since aζ = aζ = (1 + ζ2)−1 ∈ C0(C) the operator ζ is graph
regular according to Theorem 7. Further, bζ = ζ(1 + ζ2)−1 ∈ C0(C).
Theorem 16. Let E be a Hilbert A-module and let t ∈ Rgr(E) be normal. Then
there exists a unique φt ∈ Hom(C0(C)∼,L(E)) with N (φt(aζ)) = {0} and φt(ζ) = t.
Proof. Set
D := {z ∈ C : z ≤ 1/2}, F := {(z1, z2) ∈ [0, 1] × D : z22 = z1 − z2
1} ⊆ [0, 1] × D.
Then ∂F = {(0, 0)} and F = F \∂F . By Corollary 1, aζ is self-adjoint, bζ is normal,
and aζ and bζ commute. Their joint spectrum σ(aζ, bζ) is contained in F . Similar
statements hold for at and bt.
Uniqueness: Let φ ∈ Hom(C0(C)∼,L(E)) be such that N (φt(aζ)) = {0} and
φ(ζ) = t. Then, by Proposition 7, φ(aζ ) = aφ(ζ) = at and φ(bζ) = bφ(ζ) = bt. For
f ∈ C(F ) the functional calculus of commuting bounded normal operators yields
(4)
Each function g + β ∈ C0(C)∼ is of the form f (aζ, bζ) for some function f ∈ C(F )
with f ↾∂F≡ β. Indeed, we have
φ(f (aζ, bζ)) = f (φ(aζ ), φ(bζ )) = f (at, bt).
g(z) = g(aζ(z)−1bζ(z)) = f (aζ, bζ)(z),
where
f (z1, z2) :=(g(z2/z1) + β , (z1, z2) ∈ F
β
, (z1, z2) ∈ ∂F
.
To show that f is continuous at ∂F , assume (z1, z2) → (0, 0). From z22 = z1 − z2
g vanishes at infinity. This proves the uniqueness assertion.
it follows z2/z1) =p1/z1 − 1 → ∞ since z1 → 0. Therefore g(z2/z1) → 0, since
Existence: Equation (4) defines a ∗-homomorphism from C0(C)∼ into L(E).
Inserting f (z1, z2) := z1 into (4) it follows that N (φ(aζ )) = N (at). Note the latter
is trivial. Similarly, φ(bζ) = bt. Frm Proposition 7 we get aφ(ζ) = φ(aζ) = at and
(cid:3)
bφ(ζ) = φ(bζ ) = bt. From Theorem 7 we finally conclude that φ(ζ) = t.
1
5. Associated operators and affiliated operators
Throughout this section we assume that the Hilbert A-module E is the C∗-
algebra A itself equipped with the A-valued scalar product ha, bi := a∗b, a, b ∈ A,
and that A is realized as a nondegenerate C∗-algebra on a Hilbert space H.
Then L(E) is the multiplier algebra M(A) = {x ∈ B(H) : xA ⊆ A,Ax ⊆ A}.
and R(E) is the set Aη of affiliated operators in the sense of Woronowicz [W91].
Recall that Aη is the set of operators t ∈ C(H) for which at = (I + t∗t)−1 ∈ M(A),
bt = t(I + t∗t)−1 ∈ M(A), and atA is dense in A. We write tηA if t ∈ Aη. Note
that t = bta−1
Definition 16. We say that an operator t ∈ C(H) is associated with A and write
tµA if t ∈ Rgr(E). The set of associated operators with A is denoted by Aµ.
for tηA.
t
UNBOUNDED OPERATORS ON HILBERT C ∗-MODULES
27
That is, by Theorem 7, t ∈ C(H) is in Aµ if and only if at, at∗ , bt ∈ M(A). Note
that the density of the set atA in A is not required for tµA. Obviously, Aµ ⊆ Aη.
Further, t ∈ Aµ is in Aη if and only if atA is dense in A.
Lemma 16. M(A) = {tµAt ∈ B(H)}.
Proof. If t ∈ M(A), then I + t∗t ∈ M(A) and so at ∈ M(A) and bt = tat ∈ M(A),
hence tµA. Conversely, suppose that t is bounded. Then I + t∗t is bounded and
at ∈ M(A), so that a−1
(cid:3)
t = I + t∗t ∈ M(A). Therefore, t = bta−1
t ∈ M(A).
For tµA all three operators at, at∗ , bt have to be in the multiplier algebra M(A),
while for tηA it is only required that at, bt ∈ M(A) (and the density of atA). From
tηA it follows that at∗ ∈ M(A). Therefore, it is natural to ask whether or not
at ∈ M(A) and bt ∈ M(A) already imply that tµA, that is, at∗ ∈ M(A). This is true
if t ∈ C(H) is normal, since then at = at∗ . Proposition 8 below contains an number
of other sufficient conditions. In Example 11 we will show that this is not true in
general. The following simple relations appeared already in Definition 12.
Lemma 17. Let t ∈ C(H). Then:
t∗ = btbt∗ and at∗ bt = btat.
t∗ = btan−1
Proof. (1): We have at∗ − a2
by a similar reasoning starting with the operator att∗at.
bt∗ for n ∈ N.
t∗ = tt∗a2
t∗ = ta2
t
t t∗ = btbt∗. The second equality follows
(1) at∗ − a2
(2) b∗t = bt∗ .
t∗ − an+1
(3) an
(2): Let x, y ∈ H. Then at∗ y ∈ D(tt∗) ⊆ D(t∗) and using (1) we obtain
hbtx, at∗ yi = hatx, t∗at∗yi = hx, atbt∗yi = hx, bt∗ at∗ yi .
Therefore b∗t = bt∗, since R(at∗ ) is dense and bt and bt∗ are bounded.
(3) is easily derived from (1).
(cid:3)
Proposition 8. Suppose that at, bt ∈ M(A). Each of the following conditions imply
that at∗ ∈ M(A) and so tµA.
(1) 0 ∈ ρ(t).
(2) kat∗k < 1, or equivalently, tt∗ ≥ ε for some ǫ > 0.
(3) M(A)sa is closed under strong convergence of monotone sequences.
(4) tt∗ = qt∗t for some q > 0.
Proof. Clearly, from (1) it follows that 0 ∈ ρ(t∗) which in turn implies (2).
(2), (3): By Lemma 17 and the assumptions at, bt ∈ M(A) we have
at∗ − an+1
t∗ = bt(I + . . . + an−1
If (2) is fulfilled, then an+1
t∗ → 0 in M(A), hence at∗ ∈ M(A). On the other side,
an+1
t∗ ∈ M(A)sa is monotone decreasing and strongly converging. Hence again by
assumption (3) it follows at∗ ∈ M(A). (4) finally follows from the relations
at∗ = (I + tt∗)−1 = (I + qt∗t)−1 = q−1(I + (q−1 − 1)at)−1at ∈ M(A).
)b∗t ∈ M(A)sa.
t
(cid:3)
Proposition 9. Suppose that t ∈ C(H) and 0 ∈ ρ(t). Then tµA if and only if
t−1 ∈ M(A).
Proof. Since 0 ∈ ρ(t), (t∗)−1 = (t−1)∗ ∈ B(H). Simple computations show that
I − at∗ = (I + (t−1)∗t−1)−1.
I − at = (I + t−1(t−1)∗)−1,
From these identities we conclude that t−1 ∈ M(A), so (t−1)∗ ∈ M(A), implies that
at, bt, at∗ ∈ M(A), that is, tµA.
bt = (t−1)∗(I − at),
28
REN ´E GEBHARDT AND KONRAD SCHM UDGEN
Conversely, suppose that tµA. Then, by Lemma 17,(1), we have bt∗ = (bt)∗ ∈
(cid:3)
M(A) and at∗ ∈ M(A). Therefore, t−1 = bt∗ (I − at∗ )−1 ∈ M(A).
Corollary 6. If t ∈ C(H) and tµA, then (I + t∗t)µA.
Proof. Since tµA, we have (I + t∗t)−1 = at ∈ M(A). Since 0 ∈ ρ(1 + t∗t) and
(1 + t∗t)−1 = at ∈ M(A), we obtain (1 + t∗t)µA by Proposition 9.
(cid:3)
Corollary 7. Suppose that tµA and sµA.
(1) If 0 ∈ ρ(t) and λ ∈ ρ(t) with 0 < λ < 1/kt−1k, then (t − λ)µA.
(2) If 0 ∈ ρ(t) ∩ ρ(s), then tsµA.
Proof. Both assertions follow immediately from Proposition 9. For (1) we use the
equality (t − λI)−1 = λ−1t−1(λ−1 − t−1)−1 ∈ M(A), while for (2) we note that
0 ∈ ρ(ts) and (ts)−1 = s−1t−1 ∈ M(A).
(cid:3)
Next we investigate affiliated operators and their resolvents. Before we turn to
the main result we prove two simple lemmas.
Lemma 18. Let A be a C∗-algebra acting on a Hilbert space H. Let s ∈ B(H)
and x, y ∈ M(A). Suppose that xA and yA are dense in A.
If sx ∈ M(A) and
s∗y ∈ M(A), then s ∈ M(A).
Proof. Let a ∈ A. Since xA is dense in A, there are elements an ∈ A, n ∈ N ,
such that xan → a in A. Hence sxan → sa in A. Since sx ∈ M(A) by assumption,
sxan ∈ A and so sa ∈ A. Replacing x by y and s by s∗ it follows that s∗a ∈ A.
Therefore, s ∈ M(A).
(cid:3)
Lemma 19. Let A be a C∗-algebra and x, y ∈ M(A). Suppose that λy ≥ xx∗ for
some λ > 0. If xA is dense in A, so is yA. In particular, xA is dense in A if
and only if xx∗A is.
Proof. Assume to the contrary that yA 6= xA = A. Then the closure of (yA)∗ is a
proper left ideal. Hence there exists a state ω of A that annihilates (yA)∗ (see e.g.
[Dix77, Lemma 2.9.4]). Let πω be the GNS representation of A associated with the
state ω and let ϕω be the corresponding cyclic vector ϕω. We denote the extension
of πω to the multiplier algebra M(A) also by the symbol πω. Then
(5)
0 = ω((ya)∗) = hπω(a∗y)ϕω, ϕωi = hπω(y)ϕω, πω(a)ϕωi
for all a ∈ A, so that πω(y)ϕω = 0. Therefore,
ω(xa)2 = hπω(a)ϕω, πω(x∗)ϕωi2 = kπω(a)ϕωk2kπω(x∗)ϕωk2
= kπω(a)ϕωk2hπω(xx∗)ϕω, ϕωi ≤ kπω(a)ϕωk2λhπω(y)ϕω, ϕωi = 0
for a ∈ A, that is, ω annihilates xA. Hence xA is not dense in A which is a
contradiction, since we assumed that xA = A.
Applying this to the case y = xx∗ we conclude that xx∗A is dense provided that
xA is dense. Since the converse implication is trivial, it follows that xx∗A is dense
if and only if xA is dense.
(cid:3)
The following theorem appeared in [Sch05].
Theorem 17. Suppose A is a C∗-algebra acting on a Hilbert space H and t is a
densely defined closed operator on H with non-empty resolvent set. Let λ ∈ ρ(t).
Then tηA if and only if (t − λI)−1 ∈ M(A) and (t − λI)−1A and (t∗ − λI)−1A are
dense in A.
UNBOUNDED OPERATORS ON HILBERT C ∗-MODULES
29
Proof. Since tηA is equivalent to (t − λI)ηA (see [W91], p. 412, Example 1), we
can assume without restriction of generality that λ = 0. Then t−1 and (t∗)−1 are
in B(H).
First we suppose that tηA. Set x := (I + (tt∗)−1)−1 and s := t−1. Since tηA
implies t∗ηA, it follows that zt∗ = t∗(I + tt∗)−1/2 = (zt)∗ ∈ M(A). Therefore, we
obtain (I +tt∗)−1 = I−ztz∗t ∈ M(A) and hence (I +tt∗)−1/2 ∈ M(A). These relations
imply that
sx ≡ t−1(I + (tt∗)−1)−1 = t∗(tt∗)−1(I + (tt∗)−1)−1
= t∗(I + tt∗)−1 = zt∗ (I + tt∗)−1/2 ∈ M(A).
(6)
Since x := (I +(tt∗)−1)−1 = I−(I +tt∗)−1 ∈ M(A) and x−1 is also bounded, we have
x−1 ∈ M(A) and hence xA = A. Recall that sx ∈ M(A) by (6). Now we interchange
the roles of t and t∗ and set y := (I + (t∗t)−1)−1. By a similar reasoning as in (6)
we derive s∗y ∈ M(A). Further, y ∈ M(A) and yA = A. Hence the assumptions of
Lemma 18 are satisfied, so we obtain t−1 = s ∈ M(A).
Recall that (I + t∗t)−1A is dense in A, because tηA. Therefore, since
(I + t∗t)−1A = (t∗t)−1(I + (t∗t)−1)−1A ⊆ (t∗t)−1A = t−1(t∗)−1A ⊆ t−1A,
t−1A is dense in A. Replacing t by t∗, it follows that (t∗)−1A is dense in A. This
completes the proof of the only if part.
Conversely, let us assume that t−1 ∈ M(A) and that t−1A and (t∗)−1A are
dense in A. Then I − z∗t zt = (I + t∗t)−1 = t−1(t−1)∗(I + t−1(t−1)∗)−1 ∈ M(A)
and zt(I − z∗t zt)1/2 = t(I + t∗t)−1 = (t−1)∗(I + t−1(t−1)∗)−1 ∈ M(A). Therefore,
setting x := (I − z∗t zt)1/2 and s := zt, we have x ∈ M(A) and sx ∈ M(A). Since
t has a bounded inverse, there exists ǫ ∈ (0, 1/4) such that t∗t ≥ 2ǫI. Then
I + t∗t ≤ 1
ǫ t∗t and hence (I + t∗t)−1 ≥ ǫt−1(t−1)∗. Therefore, since
t−1A is dense in A by assumption, (I + t∗t)−1A = (I − z∗t zt)A = x2A is dense in A
by Lemma 19. Since x ≥ 0, xA dense in A again by Lemma 19. By the assumptions
we can interchange the roles of t and t∗. Then we obtain y := (I − ztz∗t )1/2 ∈ M(A)
and s∗y = z∗t y ∈ M(A). Further, (I + tt∗)−1A = (I − ztz∗t )A = y2A in A and hence
yA are dense in A. Thus, zt ∈ M(A) by Lemma 18 and hence tηA.
(cid:3)
2ǫ t∗t + t∗t ≤ 1
The preceding theorem has a number of interesting corollaries. For these results
we assume that A is a C∗-algebra acting on a Hilbert space H and t and s are
densely defined closed operators on H.
Corollary 8. Suppose that t, sηA, λ ∈ ρ(t) and µ ∈ ρ(s). Then we have −λµ ∈
ρ(ts − λs − µt) and (ts − λs − µt)ηA.
Proof. By some straightforward arguments one verifies that
(7)
(ts − λs − µt + λµI)−1 = (s − µI)−1(t − λI)−1,
((ts − λs − µt)∗ + λµI)−1 = (t∗ − λI)−1(s∗ − λI)−1.
(8)
Hence −λµ ∈ ρ(ts − λs − µt). From the only if part of Theorem 17 it follows that
the operators in (7) and in (8) belong to M(A) and that they maps A densely into
A. Therefore, by the if part of Theorem 17, (ts − λs − µt)ηA.
(cid:3)
Proposition 10. Suppose that λ ∈ ρ(t), s(t− λI)−1 ∈ M(A) and ks(t− λI)−1k < 1.
Then (t + s)ηA.
Proof. By Theorem 17, (t − λI)−1 ∈ M(A) and (t − λI)−1A and (t∗ − λI)−1A are
dense in A. By the assumption we have r := s(t − λI)−1 ∈ M(A) and krk < 1.
Therefore (I + r)−1 is bounded and an element of M(A), since I + r ∈ M(A).
Further, since t − λI and I + r are bijective and (t + s − λI)ϕ = (I + r)(t − λI)ϕ
for ϕ ∈ D(t) ⊆ D(s), the map t + s− λI : D(t) → H is bijective. Hence λ ∈ ρ(t + s)
30
REN ´E GEBHARDT AND KONRAD SCHM UDGEN
and (t + s − λI)−1 = (t − λI)−1(I + r)−1 ∈ M(A). Because I + r is an invertible
element of M(A), the density of (t − λI)−1A implies the density of (t + s − λI)−1A
in A. Finally, (t∗ + s∗ − λI)−1 = (I + r∗)−1(t∗ − λI)−1 maps A densely into A,
since kr∗k < 1. Now applying again Theorem 17 we obtain (t + s)ηA.
(cid:3)
Corollary 9. Let A ⊆ B(H) a C∗-algebra and suppose that t, sηA. If λ ∈ ρ(t),
0 ∈ ρ(s), and kλ(t − λI)−1k < 1, then tsηA.
Proof. By Corollary 8 we have 0 ∈ ρ(ts − λs) and (ts − λs)ηA. Since tηA and
λ ∈ ρ(t), it follows from Theorem 17 that (t − λ)−1 ∈ M(A). Therefore,
λs(ts − λs)−1 = λs((t − λI)s)−1 = λ(t − λI)−1 ∈ M(A).
Hence, since kλ(t − λ)−1k < 1 by assumption, Proposition 10 applies to the opera-
tors t := ts − λs and s := λs and implies that t + s = tsηA.
(cid:3)
Corollary 10. Let A ⊆ B(H) a C∗-algebra. Suppose that tηA and λ, µ ∈ ρ(t). For
an operator s on H we have s(t − λI)−1 ∈ M(A) if and only if s(t − µI)−1 ∈ M(A).
Proof. Since (t− λI)−1 ∈ M(A) and (t− µI)−1 ∈ M(A) by Theorem 17, the assertion
follows from the identity
s(t−λI)−1−s(t−µI)−1 = (λ−µ)s(t−µI)−1(t−λI)−1 = (λ−µ)s(t−λI)−1(t−µI)−1.
(cid:3)
Corollary 11. Let A ⊆ B(H) a C∗-algebra. Suppose that tηA is a self-adjoint
operator and s is a symmetric t-bounded operator on H with t-bound less than 1.
If s(t − λI)−1 ∈ M(A) for some λ ∈ ρ(t), then (t + s)ηA and t + s is self-adjoint.
Proof. The proof can be given by repeating the standard proof of the Kato-Rellich
(cid:3)
theorem and using Lemma 10 and Corollary 10.
It is natural to ask whether or not the second density assumption in Theorem 17
can be omitted, that is, when does the density of (t− λI)−1A for (t− λI)−1 ∈ M(A)
imply the density of (t∗ − λI)−1A in A? A counterexample is provided by Example
9 below. The next proposition shows that the answer is affirmative if the distance
of (t − λI)−1 to the set A−1 of invertible elements of A is zero.
Proposition 11. Let t be a densely defined closed operator and A a C∗-algebra
acting on a Hilbert space H. Suppose that t has a bounded inverse t−1 contained in
M(A). Assume that dist(t−1,A−1) = 0. If t−1A dense in A, so is (t∗)−1A.
Proof. Set x := t−1. Then (t∗)−1 = (t−1)∗ = x∗. Assume to the contrary that
x∗A = (t∗)−1A is not dense in A. Then, as in the proof of Lemma 19, there
is a state ω on A that annihilates (x∗A)∗. Arguing as in line (5) it follows that
πω(x)ϕω = 0.
Let x = vx be the polar decomposition of the operator x. Since x ∈ M(A), we
have x = (x∗x)1/2 ∈ M(A). For ε > 0 let fε denote a continuous function on R
such that fε(τ ) = 0 on [0, ε
2 , ε] and fε(τ ) = u on [ε, +∞). By
[Ped98] Theorem 6.1, applied to the multiplier algebra M(A), there exists a unitary
operator uε ∈ M(A) such that
2 ], fε(τ ) ≤ ε on [ ε
vfε(x) = uεfε(x) ∈ M(A).
Clearly, x = limε→+0 fε(x) in M(A). Therefore,
0 = πω(x)ϕω = πω(vx)ϕω = lim
ε→+0
πω(vfε(x))ϕω = lim
ε→+0
πω(uε)πω(fε(x))ϕω,
UNBOUNDED OPERATORS ON HILBERT C ∗-MODULES
31
so that 0 = limε→+0 πω(fε(x))ϕω = πω(x)ϕω. For a ∈ A we now obtain
0 = hπω(x)2ϕω, πω(a)ϕωi = hπω(x∗x)ϕω πω(a)ϕωi
= hϕω, πω(x∗xa)ϕωi = ω((x∗xa)∗).
This implies that x∗xA is not dense in A. Hence xA is not dense by Lemma 19
(cid:3)
which is the desired contradiction.
Example 9. Let H be the Hilbert space l2(N2) and let A be the C∗-algebra
The multiplier algebra of A is
A =(cid:18)K(H) K(H)
K(H) B(H)(cid:19) .
M(A) =(cid:18)B(H) K(H)
K(H) B(H)(cid:19) .
(9)
(10)
(11)
Let {ekl}k,l∈N0 be the standard orthonormal basis of H. Let s ∈ B(H) be the shift
operator given by sekl = ek+1,l and let P0 be the orthogonal projection onto N (s∗).
Clearly, {e0,l}l∈N0 is an orthonormal basis of P0H. Further, let {λkl}k,l∈N0 be a
double sequence of positive numbers such that limk,l→∞ λkl = 0. Define a self-
adjoint compact operator on H by rekl := λklekl, k, l ∈ N0.
Let x ∈ B(H ⊕ H) defined by the operator matrix
x :=(cid:18)s
0 s∗(cid:19) .
r
Since λkl > 0 for all k, l, the compression P0r ↾ P0H of r to P0H has trivial kernel
and dense range. Using this fact it is easily seen that N (x) = {0} and R(x) is
dense in H ⊕ H. Hence t := x−1 is a densely defined closed operator on the Hilbert
space H ⊕ H. By (10), we have t−1 = x ∈ M(A).
Statement: t−1A = xA is dense in A, while (t∗)−1A = x∗A is not dense in A.
Proof. Let y be an element of A. Then y given by an operator matrix
a, b, c ∈ K(H) and d ∈ B(H), and have
(12)
y :=(cid:18)a b
c d(cid:19) ,
xy =(cid:18)sa + rc
s∗c
sb + rd
s∗d (cid:19) .
Since K(H) = s∗sK(H) ⊆ s∗K(H) ⊆ K(H), we have s∗K(H) = K(H). Similarly,
s∗B(H) = B(H). Since the range of r contains all rank one operators ekl ⊗ enm,
sK(H) + rK(H) is dense in K(H). Therefore, by (9) and (12), xA is dense in A.
Next we prove the second assertion. First we note that
P0(rK(H) + sB(H)) = P0(rK(H)) ⊆ P0K(H).
This implies that rK(H) + sB(H) is not dense in B(H). Therefore, since
x∗y =(cid:18) s∗a
ra + sc
s∗b
rb + sd(cid:19) ,
it follows from (9) that the set x∗A is not dense in A.
(cid:3)
32
REN ´E GEBHARDT AND KONRAD SCHM UDGEN
6. Examples
6.1. Matrices of commutative C∗-algebras and its multipliers. In this sub-
section we use matrices over commutative C∗-algebras to construct simple examples
of operators that help to delimit the general theory.
Let A be a C∗-algebra. If Aij ⊆ A for i, j ∈ {1, 2} set
Let X be a locally compact non-compact Hausdorff space and set
A21 A22 (cid:19) :=(cid:26)(cid:18) a11 a12
(cid:18) A11 A22
A0 :=(cid:18) C0(X) C0(X)
a21 a22 (cid:19)aij ∈ Aij for i, j ∈ {1, 2}(cid:27) .
C0(X) C0(X)∼ (cid:19) ,
C0(X) C0(X) (cid:19) , A :=(cid:18) C0(X) C0(X)
where C0(X)∼ := C0(X) + C · 1. A straightforward computation shows that the
left multiplier algebra LM(A) and the multiplier algebra M(A) are given by
From this we can read off that all elements of the form
LM(A) =(cid:18) Cb(X) C0(X)
M(A) =(cid:18) Cb(X) C0(X)
C0(X) C0(X)∼ (cid:19) .
Cb(X) C0(X)∼ (cid:19) ,
(cid:18) ∗
∗ (cid:19) ∈ LM(A) with f ∈ Cb(X) \ C0(X)
∗
f
act as operators t on A defined on the whole space such that the adjoints are not
defined on the whole space. From now on let X = R.
1
0
Example 10. Let f = 1 and set the other matrix entries zero. Then t acts as
D(t) = A,
t =(cid:18) 0
Hence t∗ is essentially defined. Further, it is easily checked that t∗∗ = t and
0 (cid:19) .
0 (cid:19) ,
D(1 + t∗t) = A, 1 + t∗t =(cid:18) 2 0
0 1 (cid:19) , D(1 + tt∗) = A0, 1 + tt∗ =(cid:18) 1 0
0 2 (cid:19) .
and D(t∗) = A0,
t∗ =(cid:18) 0
From these formuals we read off that R(1 + t∗t) = A and R(1 + tt∗) = A0 ( A.
Hence at is adjointable, while at∗ is not. In fact, bt is also not adjointable, since
1
0
D(bt) = A,
bt =(cid:18) 0
1/2 0 (cid:19) /∈ M(A).
0
In the following slighty more sophisticated example at and bt are both ad-
jointable, while at∗ is not adjointable.
c
For the adjoint t∗ we obtain
Example 11. Let f, g ∈ C(R) be functions given by f (x) := xq1 + sin2(x) and
g(x) := xp1 + cos2(x). Then f (x)2 + g(x)2 = 3x2. Define t : A → A by
t =(cid:18) 0 f
D(t) :=(cid:26)(cid:18) a b
g (cid:19) .
t =(cid:18) 0
g (cid:19) .
D(t∗) :=(cid:26)(cid:18) a b
1+f2+g2 ! ∈ A.
at =(cid:18) 1
d (cid:19) ∈ A f c, f d, gc ∈ C0(R), gd ∈ C0(R)∼(cid:27),
d (cid:19) ∈ A f a + gc ∈ C0(R), f b + gd ∈ C0(R)∼(cid:27),
1+f2+g2 (cid:19) ∈ A,
It is now easily verified that 1 + t∗t is surjective and
bt = 0
1+f2+g2
0
1
0
0
0
0
f
f
g
c
UNBOUNDED OPERATORS ON HILBERT C ∗-MODULES
33
The operator at∗ is computed as
D(at∗ ) = A0,
at∗ =
1
1 + f2 + g2(cid:18) 1 + g2
−gf
−f g
1 + f2 (cid:19) .
That is, at ∈ A and bt ∈ A are adjointable, but at∗ /∈ M(A) is not adjointable.
Example 12. Now we consider the operator t given by
D(t) = A,
t =(cid:18) 0
1
0
1 (cid:19) ,
and D(t∗) = A0,
t∗ =(cid:18) 0
0
1
1 (cid:19) .
Then one easily proves that R(t) * D(t∗) and D(t∗t) = A0 ( A = D(t).
particular, D(t∗t) is not dense in the domain D(t)!
6.2. A fraction algebra related to the Weyl algebra. Let P = −i d
dt and Q = t
be the momentum and position operators acting as self-adjoint operators on the
Hilbert space L2(R). Fix α, β ∈ R\{0} and define bounded operators x and y by
In
x := (Q − αiI)−1
and y := (P − βiI)−1.
It is not difficult to verify that these operators satisfy the commutation relations
(13)
(14)
x − x∗ = 2αix∗x = 2αixx∗, y − y∗ = 2βiy∗y = 2βiyy∗,
xy − yx = −ixy2x = −iyx2y, xy∗ − y∗x = −ix(y∗)2x = −iy∗x2y∗.
J0 := yxX = xyX = X yx = X xy
Let X be the unital ∗-subalgebra of B(L2(R)) generated by x and y. Since the
operators x, x∗, y, y∗ are bijections of the Schwartz space S(R), so are their inverses.
Hence τx(·) := x·x−1 and τy(·) := y·y−1 are automorphisms of the algebra L(S(R))
of linear operators on the Schwartz space S(R). From the relations (13) and (14)
we conclude that τx and τy leave the algebra X invariant, so they are algebra
automorphisms of X . Hence
(15)
and J0 is a two-sided ∗-ideal of the ∗-algebra X .
Let Fx be the unital ∗-subalgebra of X generated by x, that is, Fx is the com-
mutative ∗-algebra of polynomials f (x, x∗) in x and x∗ with complex coefficients.
Likewise, Fy denotes the unital ∗-subalgebra of X generated by y. Note that
Fx ∩ Fy = C · 1. From the relations it follows easily that X is the direct sum of
vector spaces Fx + Fy and J0. Hence each element a ∈ X can be written as
(16)
where f1, g1, f2, g2 are polynomials with g1(0, 0) = f2(0, 0) = 0 and b1, b2 ∈ X .
Moreover, these triples {f1, g1, b1} and {f2, g2, b2} are uniquely determined by a.
Let ε > 0 and λ ∈ R. We denote by χε the characteristic function of the interval
[0, ε]. Put ωε,λ(t) := 1√ε χε(t − λ).
Lemma 20. For polynomials f ∈ C[x, x∗] and g ∈ C[y, y∗], where g(0, 0) = 0, and
b ∈ X we have
(17)
a = f1(x, x∗) + g1(y, y∗) + yxb1 = f2(x, x∗) + g2(y, y∗) + xyb2,
lim
(18)
(19)
lim
lim
ε→+0 hf (x, x∗) ωε,λ, ωε,λi = f ((λ − αi)−1, (λ + αi)−1),
ε→+0 hg(y, y∗) ωε,λ, ωε,λi = 0,
ε→+0 hyxb ωε,λ, ωε,λi = 0.
hϕωε,λ, ωε,λi = ϕ(λ) +Z λ+ε
ε−1(ϕ(t) − ϕ(λ))dt → ϕ(λ)
λ
Proof. Suppose that ϕ ∈ C1(R). Then we have
which in turn implies (17).
34
REN ´E GEBHARDT AND KONRAD SCHM UDGEN
Next we prove (18). The crucial step for this is to show that
(20)
lim
ε→+0
yωε,λ = lim
ε→+0
y∗ωε,λ = 0
in L2(R). Without loss of generality we can assume that β < 0. Since β < 0, for
the resolvent of P = −i d
dt we obtain
(yϕ)(t) = ((P − βiI)−1ϕ)(t) = −iZ ∞
t
Hence we compute (yωε,λ)(t) = 0 for t ≥ λ + ε,
eβ(s−t)ϕ(s)ds, ϕ ∈ L2(R).
(yωε,λ)(t) = −i
√ε
eβs dt = −i
β√ε
(eβ(λ+ε−t) − 1)
for λ ≤ t ≤ λ + ε, and
(yωε,λ)(t) = −i
√ε
t
e−βtZ λ+ε
e−βtZ λ+ε
λ
eβs dt = −i
β√ε
e−βt(eβ(λ+ε) − eβ(λ−ε))
for t ≤ λ. From these formulas we easily derive that limε→+0 yωε,λ = 0. Replacing
β by −β a similar reasoning yields limε→+0 y∗ωε,λ = 0. Since the operator y is
bounded and g(0, 0) = 0, (20) implies (18). Since kωε,λk = 1, it follows from (20)
that
hyxb ωε,λ, ωε,λi = hxb ωε,λ, y∗ωε,λi ≤ kxbk ky∗ωε,λk → 0
which proves (19).
(cid:3)
Next we define two circles Kα and Kβ intersecting in the origin by
Kα := {(z, 0) ∈ C2 : z − z = 2αiz2}, Kβ := {(0, w) ∈ C2 : w − w = 2βiw2}.
Let R = R ∪ {∞} be the one point compactification of the real line. The maps
λ → (zλ, 0) := ((λ − αi)−1, 0) and λ → (0, wλ) := (0, (λ − βi)−1), where z∞ := 0
and w∞ := 0, are bijection of R onto Kα and Kβ, respectively. Let z ∈ Kα and
w ∈ Kβ. For a as in (16) we define
πx,z(a) = f1(z, z),
πy,w(a) = g2(w, w).
Lemma 21. For all z ∈ Kα and w ∈ Kβ, πx,z and πy,w are one dimensional
∗-representations of the ∗-algebra X such that
(21)
Moreover, for f ∈ Fx and g ∈ Fy we have
(22)
and πy,w(a) ≤ kak
πx,z(a) ≤ kak
for a ∈ X .
kfk = sup
z∈Kα f (z, z)
and kgk = sup
w∈Kβ g(w, w).
Proof. A simple computation based on the relations (13) and (14) shows that πx,z
and πy,w are well-defined ∗-homomorphisms of X .
Let λ ∈ R. From the formulas (17), (18) and (19) we infer that
πx,zλ(a) = f1((λ − αi)−1, (λ + αi)−1) = lim
ε→+0 haωε,λ, ωε,λi.
Therefore, since kωε,λk = 1, we obtain πx,zλ(a) ≤ kak. Passing to the limit
λ → ∞, we get πx,z0(a) ≤ kak. This proves the first inequality of (21) for all
z ∈ Kα. The inequality πy,w(a) ≤ kak follows in a similar manner by interchanging
the role of x and y and using the counterparts of formulas (17), (18) and (19).
Finally, we prove (22). One inequality of the equality (22) was shown by (21).
The reversed inequality follows at once from the fact that the spectra of the self-
(cid:3)
adjoint operators P and Q are equal to R.
UNBOUNDED OPERATORS ON HILBERT C ∗-MODULES
35
It was noted in [Sch10] (and is easily verified by using (13)), the ∗-algebra X is
algebraically bounded, that is, given a ∈ X there exist γa > 0 and finitely many
elements ai ∈ X such that γa = a∗a +Pi a∗i ai. Hence
kakun := sup
π kπ(a)k,
a ∈ X ,
a
defines a C∗-norm on X , where the supremum is taken over all ∗-representation of
X , and the completion Xun of (X ;k · kun) is called the universal C∗-algebra of X .
(The supremum is finite, since kπ(a)k ≤ γ1/2
for all a ∈ X .) By decomposition
theory it suffices to take all irreducible ∗-representations π. As proved in [Sch10],
the irreducible ∗-representation of X are the one-dimensional representations πx,z
and πy,w,where z ∈ Xx and w ∈ Kβ, and the identity representation π0 acting on
the Hilbert space L2(R). From (21) it follows that the universal C∗-norm k · kun
coincides with the operator norm k · k on L2(R). Hence the universal C∗-algebra
Xun is just the closure X of X in B(L2(R)). By (22) the closures of Fx and Fy
in B(L2(R)) are the commutative C∗-subalgebras C(Kα) and C(Kβ), respectively,
and the closure J of J0 in B(L2(R)) is a two-sided ∗-ideal of the C∗-algebra
Xun = X .
Lemma 22. J0 is a two-sided essential ideal of Xun.
Proof. Let a ∈ Xun be such that aJ0 = {0}. Then axy = 0 in L2(R). Since x and
(cid:3)
y are bijection, this implies a = 0.
The operator x∗y∗yx is an integral operator on L2(R) with kernel
K(t, s) : (2β)−1(t + αi)−1(s − αi)−1e−βt−s.
Since K ∈ L2(R2), the operator x∗y∗yx = yx2 is compact, so are yx and hence
yx. Hence J0 ⊆ K(L2(R)) by (15) and therefore J = K(L2(R)).
Now we define two operators, denoted by Q and P , on the C∗-algebra Xun by
Q := αiI + x−1, D(Q) := xXun
and P := βi + y−1, D(P ) := yXun,
that is, Q(xa) = αixa + a and P (ya) = βiya + a, where a ∈ Xun.
Theorem 18. Q and P are graph regular self-adjoint operators on the C∗-algebra
Xun = X .
Proof. We carry out the proof for Q; a similar reasoning yields the assertions for
P . Since xXun and x∗Xun contain the essential ideal J0 (by Lemma 22), xXun
and x∗Xun are essential in the C∗-algebra Xun. Therefore, by Theorem 3, x−1
and (x∗)−1 are graph regular operators on Xun, so Q and P are graph regular by
Proposition Corollary 6.
Further, (x−1)∗ = (x∗)−1 and hence Q∗ = −αiI + (x∗)−1 by Theorem 3 and
Proposition 2 (2). From the first relation of (13) it follows that −αiI + (x∗)−1 =
αiI + x−1. Hence Q = Q∗, that is, Q is self-adjoint.
(cid:3)
The operators Q and P are not regular on Xun, since neither xXun nor yXun is
dense in Xun. Note that the corresponding restrictions of Q and P are affiliated
with the essential ideal J = K(L2(R)) of the C∗-algebra Xun.
6.3. Unbounded Toeplitz operators. Let L2(T) be the Hilbert space of square
integrable functions on the unit circle T with scalar product
hf, gi :=Z 1
0
f (e2πit)g(e2πit) dt
f, g ∈ L2(T),
and let P denote the projection of L2(T) on the closed subspace H 2(T) generated
by {zn := e2πitnn ∈ N0}. For φ ∈ L∞(T) the Toeplitz operator Tφ is the bounded
operator on the Hilbert space H 2(T) is defined by Tφf := P φf , f ∈ H 2(T).
36
REN ´E GEBHARDT AND KONRAD SCHM UDGEN
The C∗-algebra generated by the unilateral shift S := Tz is the Toeplitz algebra
T := {Tφφ ∈ C(T)} ∔ K(H 2(T)).
Our aim is to construct a class of examples of graph regular (unbounded) Toeplitz
operators on the C∗-algebra T . Let p, q ∈ C[z] be relatively prime polynomials such
that q has no zeros in the open unit disc D. Then the Toeplitz operator with rational
symbol p/q is defined by
D(Tp/q) := {f ∈ H2(T)
p
q
f ∈ H2(T)},
Tp/qf :=
p
q
f
(f ∈ D(Tp/q)),
Since Tp/q is a multiplication operator, Tp/q is a closed densely defined operator on
the Hilbert space H2(T).
Theorem 19. Suppose that p, q are relatively prime polynomials such that q has
no zero in the open unit disc. Then the Toeplitz operator Tp/q is associated with
the Toeplitz algebra T . Further, Tp/q is affiliated with the Toeplitz algebra if and
only if in addition q has no zero on the unit circle.
Proof. Since q has no zero in D, q is an outer function (see e.g. [RR85]).
Now we use an argument from [Sar08, Section 3]. Since p and q are relatively
prime, we have p2 + q2 > 0 on the closed unit disc D. Therefore, by the Riezs-
Fej´er Theorem [RR85], there exists a polynomial r ∈ C[z] such that r has no zero
in D and p2 + q2 = r2 on T. Let f := q/r and g := p/r. Then f and g are
continuous and in the unit ball of H∞(T), f is outer, f2 + g2 = 1 on T. Upon
multiplying r by some constant of modulus one we can assume that f (0) > 0.
From [Sar94, Proposition 5.3] it follows that D(Tp/q) = f H 2(T) and Tp/q =
TgT −1
f
. Moreover, T ∗p/q = (T −1
f
)∗T ∗g = T −1
f
Tg. Using these facts we compute
TgTgTf = T −1
1 + T ∗p/qTp/q = 1 + T −1
f
T −1
f = (Tf Tf )−1,
= T −1
f
f
(Tf Tf + TgTg)T −1
f = T −1
f
(Tf2 + Tg2)T −1
f
1 + Tp/qT ∗p/q = 1 + TgT −1
f T −1
f
Tg = 1 + Tg(Tf Tf )−1Tg = 1 + Tg(1 − TgTg)−1Tg
= 1 + (1 − TgTg)−1TgTg = (1 − TgTg)−1.
Hence aTp/q = Tf Tf and aT ∗
p/q
= I − TgTg are in T . Further,
f Tf Tf = TgTf ∈ T .
bTp/q = Tp/qATp/q = TgT −1
, bTp/q ∈ T , Tp/q is associated with the C∗-algebra T .
Since aTp/q , aT ∗
p/q
Suppose now q has a zero at some λ ∈ T. Then a has a zero at λ as well. For
z ∈ T let ωz be the character on T given by
(23)
If Tφ + K ∈ T , then Tf Tf (Tφ + K) = Tf2φ + K for some K ∈ K(H2(T)). Hence
(φ ∈ C(T), K ∈ K(H 2(T)).
ωz(Tφ + K) = φ(z)
ωλ(aTp/q (Tφ + K)) = ωλ(Tf2φ + K) = f (λ)2φ(λ) = 0.
Therefore, aTp/qT is not dense in T and hence Tp/q is not affiliated with T .
so in particular, Tp/q is affiliated with T .
On the other hand, if q has no zero on T, then p/q ∈ C(T) and hence Tp/q ∈ T ,
(cid:3)
The simplest interesting example is the following.
Example 13. Set p(z) = 1 and q(z) = 1 − z, so that p/q = 1/(1 − z). Then, by
Theorem 19, T1/(1−z) associated with T , but T1/(1−z) is not affiliated with T . In
fact, T1/(1−z) = (I − S)−1.
UNBOUNDED OPERATORS ON HILBERT C ∗-MODULES
37
6.4. Heisenberg group. Let H be the 3-dimensional Heisenberg group, that is,
H is the Lie group whose differential manifold is the vector space R3 and whose
multiplication is given by
(x1, x2, x3)(x′1, x′2, x′3) := (x1 + x′1, x2 + x′2, x3 + x′3 +
1
2
(x1x′2 − x′1x2)).
The C∗-algebra C∗(H) of the Lie group H was described in [LT11]. We briefly
repeat this result. First we recall that C∗(H) is defined as the completion of L1(H)
with respect to the norm
kfk = sup{kπU (f )k : U unitary representation of H}.
where πU is the ∗-representation of L1(H) associated with U , that is,
U (x1, x2, x3)f (x1, x2, x3) dx1dx2dx3,
f ∈ L1(H).
πU (f ) :=ZR3
The irreducible unitary representations of H consist of a series Uλ, λ ∈ R×, of
infinite dimensional representations acting on L2(R) and of a series Ua, a ∈ R2, of
one dimensional representations. For (x1, x2, x3) ∈ H, these representations act as
(Uλ(x1, x2, x3)ξ)(s) = e−2πiλ(x3+ 1
2 x1x2+sx2)ξ(s − x1),
ξ ∈ L2(R), λ ∈ R×,
Ua(x1, x2, x3) = e−2πi(a1x1+a2x2),
a = (a1, a2) ∈ R2.
The Lie algebra of H has a basis {X, Y, Z} with commutation relations
[X, Y ] = Z,
[X, Z] = [Y, Z] = 0
and we have dUλ(iZ) = 2πλI and dUa(iZ) = 0.
Now let F be the C∗-algebra of all operator fields F = (F (λ); λ ∈ R) satifying
the following conditions:
(i) F (λ) is a compact operator on L2(R) for each λ ∈ R×,
(ii) F (0) ∈ C0(R2),
(iii) R× ∋ λ → F (λ) ∈ B(L2(R) is norm continuous,
(iv) limλ→∞ kF (λ)k = 0.
Let η be a fixed function of the Schwartz space S(R) of norm one in L2(R). For
ξ ∈ L2(R), let Pξ denote the projection on the one dimensional subspace C · ξ.
Then for h ∈ C0(R2) and λ ∈ R× := R\{0}, the operator νλ(h) is defined by
(24)
h(x1, x2)Pη(λ;x1,x2)λ−1dx1dx2,
νλ(h) :=ZR2
where h denotes the Fourier transform of h and
By Proposition 2.14 in [LT11], we have
η(λ; x1, x2)(s) := λ1/4e2πix1s η(cid:0)λ1/2(s + x2λ−1)(cid:1),
λ→0 kνλ(h)k = hk∞ for h ∈ C0(R2).
lim
(25)
x1, x2, s ∈ R.
Then, according to Theorem 2.16 in [LT11], the C∗-algebra C∗(H) is the C∗-
subalgebra of C∗(H) formed by all operator fields F ∈ F such that
(26)
lim
λ→0 kF (λ) − νλ(F (0))k = 0,
where νλ : C0(R2) → F is defined by (24), and for c ∈ C∗(H), we have F (c)(λ) =
πUλ , λ ∈ R×, and F (c)(0)(a) = πUa (c), a ∈ R2.
On the other hand, it was proved in [WN92] that the Lie algebra generators
X, Y, Z act as skew-adjoint regular operators on the C∗-algebra C∗(H).
We show that the range R(iZ) is essential in C∗(H). Assume that G(λ) ∈ C∗(H)
and G(λ) ∈ R(iZ)⊥. Since dUλ(iZ) = 2πλI for λ ∈ R×, R(iZ) contains all vector
fields F (λ) ∈ F of compact support contained in R×. This implies that G(λ) = 0
38
REN ´E GEBHARDT AND KONRAD SCHM UDGEN
on R×. Therefore, limλ→0 νλ(G(0)) = 0 by (26) and hence [G(0) = 0 by (25), so
G(0) ∈ C0(R2) is zero. Thus G = 0 in C∗(H) which proves that R(iZ) is essential.
Further, iZ is self-adjoint, so (iZ)−1 is by Proposition 1.
Since iZ is graph regular, so is (iZ)−1 by Proposition 3. Note that (iZ)−1 is not
regular, because dUa(iZ) = 0 for a ∈ R2 and hence (iZ)−1 is not densely defined.
Theorem 20. (iZ)−1 is a graph regular self-adjoint operator on the C∗-algebra
C∗(H).
References
[B81] S. Baaj, Multiplicateurs non bornes, Thesis, Universite Pierre et Marie Curie 37(1981),
1 -- 44.
[BJ83] S. Baaj, P. Julg, Theorie bivariante de Kasparov et operateurs non bornes dans les C*-
modules hilbertiens, C. R. Acad. Sci. Paris Sr. I Math. 296 (1983), no. 21, 875-878.
[Dix77] J. Dixmier, C*-algebras, North-Holland, Amsterdam, 1977.
[FS10] M. Frank, K. Sharifi, Generalized inverses and polar decomposition of unbounded regular
operators on Hilbert C ∗-modules, J. Operator Theory 64(2010), 377-386.
[GVF] J.M. Gracia-Bondia, J. Varilly, H. Figueroa, Elements of Noncommutative Geometry,
Birkhauser, Boston, 2001.
[H89] M. Hilsum, Fonctorialite en K-Theorie bevariante pour les varietes lipschitziennes, K-
Theory 3(1987), 401-440.
[KL12] J. Kaad, M. Lesch, A local global principle for regular operators in Hilbert C ∗-modules,
J. Funct. Anal. 262(2012), 4540-4569.
[K53] I. Kaplansky, Moduls over operator algebras, Amer. J. Math. 75(1953), 839-858.
[Kas80] G.G. Kasparov, Hilbert C ∗-modules: Theorems of Stinespring and Voiculescu, J. Oper-
ator Theory 4(1980), 133 -- 150.
[K97] D. Kucerovsky, The KK-product of unbounded modules, K-Theory 11(1997), 17 -- 34.
[K02] D. Kucerovsky, Functional calculus and representations of C0(X) on a Hilbert module,
Quart. J. Math. 53(2002), 467 -- 477.
[L95] E. C. Lance, Hilbert C ∗-modules - A toolkit for operator algebraists, Cambridge Univ.
Press, 1995.
[LT11] J. Ludwig, L. Turowska, The C ∗-algebras of the Heisenberg group and of thread-like Lie
groups, Math. Z. 268 (2011), 897-930.
[Mg97] B. Maganja, Hilbert C ∗-modules in which all closed submodules are complemented, Proc.
Math. Soc. 125(1997), 849 -- 852.
[MT05] V. M. Manuilov, E. V. Troitsky, Hilbert C ∗-Modules, Amer. Math. Soc, RI, 2005.
[Pal99] A. Pal, Regular operators on Hilbert C ∗-modules, J. Operator Theory 42 (1999), 331-350.
[P73] W. L. Paschke, Inner product moduls over B∗-algebras, Trans. Amer. Math. Soc. 182(1973),
443-468.
[Ped98] G. K. Pedersen, Factorization in C*-algebras, Expo. Math. 16(1998), 145-156.
[Pie06] F. Pierrot, Op´erateurs r´eguliers dans les C ∗-modules et structure des C ∗-alg`ebres de
groups de Lie semisimples complexes simplement connexes, J. Lie Theory 16 (2006), 651-
689.
[R74] M. A. Rieffel, Induced representations of C ∗-algebras, Adv. Math. 13(1974), 176-257.
[RR85] M. Rosenblum, J. Rovnyak, Hardy classes and operator theory, Oxford Univ. Press, 1985.
[Sar94] D. Sarason, Sub-Hardy Hilbert Spaces in the Unit Disc, John Wiley Sons, Inc., New York,
1994.
[Sar08] D. Sarason, Unbounded Toeplitz Operators, Integr. Equ. Oper. Theory 61 (2008), 281-
298.
[Sch05] K. Schmudgen, Unbounded operators affiliated with C ∗-algebras, Preprint, Leipzig, 2005.
[Sch10] K. Schmudgen, Algebras of fractions and strict Positivstellensatze for ∗-algebras, J. reine
angew. Math. 647(2010), 57-88.
[W91] S. L. Woronowicz, Unbounded elements affiliated with C ∗-algebras and non-compact quan-
tum groups, Commun. Math. Phys. 136(1991), 399-432.
[WN92] S. L. Woronowicz, K. Napi´orkowski, Operator theory in the C ∗-algebra framework, Re-
ports Math. Phys. 31(1992), 353-371.
K. Schmudgen; Universitat Leipzig, Mathematisches Institut, Augustusplatz
10/11, D-04109 Leipzig, Germany; E-Mail: [email protected]
UNBOUNDED OPERATORS ON HILBERT C ∗-MODULES
39
R. Gebhardt; Max Planck Institute for Mathematics in the Sciences, Inselstrasse
22, D-04103 Leipzig, Germany; E-Mail: [email protected]
|
1605.04900 | 2 | 1605 | 2016-06-06T18:08:06 | C*-algebras have a quantitative version of Pelczynski's property (V) | [
"math.OA",
"math.FA"
] | A Banach space X has Pelczynski's property (V) if for every Banach space Y every unconditionally converging operator T: X -> Y is weakly compact. H. Pfitzner proved that C*-algebras have Pelczynski's property (V). In the preprint "H. Krulisova: Quantification of Pelczynski's property (V)" the author explores possible quantifications of the property (V) and shows that C(K) spaces for a compact Hausdorff space K enjoy a quantitative version of the property (V). In this paper we generalize this result by quantifying Pfitzner's theorem. Moreover, we prove that in dual Banach spaces a quantitative version of the property (V) implies a quantitative version of the Grothendieck property. | math.OA | math |
C*-ALGEBRAS HAVE A QUANTITATIVE VERSION OF
PE LCZY ´NSKI'S PROPERTY (V)
HANA KRULISOV ´A
Abstract. A Banach space X has Pe lczy´nski's property (V) if for every Ba-
nach space Y every unconditionally converging operator T : X → Y is weakly
compact. H. Pfitzner proved that C ∗-algebras have Pe lczy´nski's property (V).
In the preprint [8] the author explores possible quantifications of the prop-
erty (V) and shows that C(K) spaces for a compact Hausdorff space K enjoy
a quantitative version of the property (V). In this paper we generalize this
result by quantifying Pfitzner's theorem. Moreover, we prove that in dual Ba-
nach spaces a quantitative version of the property (V) implies a quantitative
version of the Grothendieck property.
1. Introduction
In 1994, H. Pfitzner proved that C∗-algebras have Pe lczy´nski's property (V) (see
[10]). The aim of this paper is to prove a quantitative version of Pfitzner's result.
In this way we continue the study of quantitative versions of Pe lczy´nski's property
(V) presented in the preprint [8].
Section 2 summarizes all essential definitions and basic facts contained mostly in
the preprint [8]. In Section 3 we slightly improve Behrends's quantitative version
of Rosenthal's ℓ1 -- theorem [2, Section 3], which we use to prove the main theorem
in Section 4. Section 5 is devoted to the relationship of quantitative versions of
Pe lczy´nski's property (V) and the Grothendieck property in dual Banach spaces.
2. Preliminaries
We follow the notation of [8] with one exception. Because we deal also with
C∗-algebras, we write X ′ (instead of X ∗) for a dual to a Banach space X, since
the ∗ in C∗-algebras is already reserved for the involution. All Banach spaces are
considered either real or complex, unless stated otherwise. The closed unit ball of
a Banach space X is denoted by BX .
2.1. Pe lczy´nski's property (V) and its quantification. Let us recall some
essential definitions and facts (explained in more detail in [8] with many comments).
2010 Mathematics Subject Classification. 46B04,46L05,47B10.
Key words and phrases. Pe lczy´nski's property (V); C ∗-algebra; Grothendieck property.
The research was supported by the Grant No. 142213/B-MAT/MFF of the Grant Agency of
the Charles University in Prague and by the Research grant GA CR P201/12/0290.
1
A series P∞
n=1 xn in a Banach space X is said to be
is a bounded sequence of scalars,
• unconditionally convergent if the seriesP∞
• weakly unconditionally Cauchy (wuC ) if
P∞
n=1 x′(xn) converges.
n=1 tnxn converges whenever (tn)
for all x′ ∈ X ′ the series
2
HANA KRULISOV ´A
A bounded linear operator T : X → Y between Banach spaces X and Y is called
unconditionally converging if Pn T xn is an unconditionally convergent series in Y
whenever Pn xn is a weakly unconditionally Cauchy series in X. It is not difficult
to show that T is unconditionally converging if and only if for every series Pn xn
in X with
∞
sup
x′∈BX ′
Xn=1
x′(xn) < ∞
the series Pn T xn converges. We say that a Banach space X has Pe lczy´nski's
property (V) if for every Banach space Y every unconditionally converging operator
T : X → Y is weakly compact.
To quantify the property (V) means to replace the implication
(1)
T is unconditionally converging ⇒ T is weakly compact
by an inequality
measure of weak non-compactness of T
≤ C · measure of T not being unconditionally converging,
where C is some positive constant depending only on X, and the two measures
are positive numbers for each operator T and are equal to zero if and only if T
is weakly compact or unconditionally converging, respectively. This inequality is
a strengthening of the original implication (1).
For this purpose we use the following quantities. For a bounded sequence (xn)
in a Banach space X we define
ca(cid:0)(xn)(cid:1) = inf
n∈N
sup{kxk − xlk : k, l ∈ N, k, l ≥ n}.
It is a measure of non-Cauchyness of a sequence (xn), hence in Banach spaces it
measures non-convergence. Let T : X → Y be a bounded linear operator between
Banach spaces X and Y . We set
uc(T ) = sup(ca (cid:18) n
Xi=1
T xi(cid:19)n! : (xn) ⊂ X,
sup
x′∈BX ′
x′(xn) ≤ 1) .
∞
Xn=1
Then uc(T ) measures how far is the operator T from being unconditionally con-
verging.
Let A be a bounded subset of a Banach space X. The de Blasi measure of weak
non-compactness of the set A is defined by
ω(A) = inf{d(A, K) : ∅ 6= K ⊂ X is weakly compact},
where
d(A, K) = sup{dist(a, K) : a ∈ A}.
De Blasi has proved that ω(A) = 0 if and only if A is relatively weakly compact
(see [4]). Other quantities which measure relative weak non-compactness are for
example
γ(A) = sup{ lim
n
lim
m
x′
m(xn) − lim
m
lim
n
x′
m(xn) : (xn) is a sequence in A,
(x′
m) is a sequence in BX ′ , and the limits exist}
or
wckX (A) = sup{dist(clust(X ′′,w∗)(xn), X) : (xn) is a sequence in A},
C*-ALGEBRAS HAVE A QUANTITATIVE VERSION OF PE LCZY ´NSKI'S PROPERTY (V) 3
where clust(X ′′,w∗)(xn) stands for the set of all w∗-cluster points of the sequence
(xn) in X ′′. The quantities γ(A) and wckX (A) are equivalent by [1, Theorem 2.3]
in the following sense:
(2)
wckX (A) ≤ γ(A) ≤ 2 wckX (A).
However, the quantity ω(A) is not equivalent to the other two (see [1, Corollary
3.4]). We have only
(3)
by [1, Theorem 2.3].
wckX (A) ≤ ω(A)
For measuring weak non-compactness of a bounded linear operator T : X → Y
between Banach spaces X and Y we use the quantities ω(T (BX)), γ(T (BX)), and
wckY (T (BX )), which we denote simply by ω(T ), γ(T ), and wkY (T ).
We say that a Banach space X has a quantitative version of Pe lczy´nski's property
(V) -- we denote it by (Vq) -- if there is a constant C > 0 such that for every Banach
space Y and every operator T : X → Y
(4)
γ(T ) ≤ C · uc(T ).
If it is possible to replace γ(T ) in (4) with ω(T ), we say that X has the property
(Vq)ω. If γ(T ) in (4) is replaced by ω(T ′), where T ′ : Y ′ → X ′ denotes the dual
operator to T , we say that X has the property (Vq)∗
ω.
In [8, Proposition 3.2] it is proved that a Banach space X has the property (Vq)
if and only if there exists a constant C > 0 such that for each bounded subset K
of the dual space X ′
where
γ(K) ≤ C · η(K),
∞
η(K) = supn lim sup
n
x′∈K x′(xn) : (xn) ⊂ X,
sup
sup
x′∈BX ′
Xn=1
x′(xn) ≤ 1o.
Using the above-described characterization we will prove in Section 4 that
C∗-algebras have the property (Vq).
Note that the quantity η is translation-invariant, that is,
η(K) = η(K + z′), K ⊂ X ′, z′ ∈ X ′.
(5)
This follows from the fact that (xn) weakly null whenever P xn is a wuC series in
X.
2.2. Measures of weak and weak∗ non-Cauchyness of sequences in Banach
spaces. In sections 4 and 5 we will use the following standard quantities, analogous
to the quantity ca, which measure how far is a sequence in a (dual) Banach space
from being weakly (weak∗) Cauchy.
Let X be a Banach space and let (xn) be a bounded sequence X. We set
δ(xn) = sup
x′∈BX ′
lim
n→∞
k,l≥nx′(x′
sup
k) − x′(x′
l).
This quantity is a measure of weak non-Cauchyness of the sequence (xn). Further-
more, let us set
δ(xn) = inf {δ(xnk ) : (xnk ) is a subsequence of (xn)} .
It measures how close can subsequences of (xn) be to be weakly Cauchy.
4
HANA KRULISOV ´A
If (x′
n) is a bounded sequence in X ′, we set
k,l≥n x′
sup
n) = sup
x∈BX
δw∗(x′
lim
n→∞
k(x) − x′
l(x).
The last quantity is a measure of weak∗ non-Cauchyness of the sequence (x′
n).
The quantity δ(xn) equals 0 if and only if the sequence (xn) is weakly Cauchy.
n) is weak∗ Cauchy. If δ(xn) = 0, it it
Analogously, δw∗ (x′
not clear whether (xn) admits a weakly Cauchy subsequence.
n) = 0 if and only if (x′
2.3. Selfadjoint elements and selfadjoint functionals. Let A be a C∗-algebra.
Let us denote by Asa the selfadjoint elements of A, that is Asa = {a ∈ A : a = a∗}.
If f is a bounded linear
Then Asa is a real Banach space and A = Asa + iAsa.
functional on A, f ∗ is the functional defined by f ∗(x) = f (x∗), x ∈ A. Let (A′)sa
denote the set {f ∈ A′ : f = f ∗} of selfadjoint functionals on A. Then (A′)sa is
a real Banach space, and is isometrically isomorphic to (Asa)′. We will write A′
sa
for both these spaces. Every functional x′ ∈ A′ can be decomposed as x′ = f + ig
where f, g ∈ A′
sa. It suffices to set f = (x′ + (x′)∗)/2, g = (x′ − (x′)∗)/(2i).
3. A quantitative version of Rosenthal's ℓ1 -- theorem
For proving the main result we need the quantitative version of Rosenthal's
ℓ1 -- theorem proved by E. Behrends in [2, Section 3]. In this section we revise his
theorem, because it turns out that one of the estimates there can be easily improved.
We will then use this improved version.
Let us remind Behrends's definition [2, 3.1].
Definition. Let (xn) be a bounded sequence in a Banach space X and ε > 0.
We say that (xn) admits ε -- ℓ1 -- blocks if for every infinite M ⊂ N there are scalars
a1, . . . , ar with Par = 1 and i1, . . . , ir in M such that (cid:13)(cid:13)P aρxiρ(cid:13)(cid:13) ≤ ε.
The revised version of the quantitative Rosenthal's ℓ1 -- theorem for complex Ba-
nach spaces is the following.
Theorem 3.1. Let X be a complex Banach space X and ε > 0. Let (xn) be
a sequence in X which admits ε -- ℓ1 -- blocks. Then there is a subsequence (xnk ) of
(xn) such that for every x′ ∈ X ′ with kx′k = 1 the diameter of the set of cluster
points of the sequence (x′(xnk ))k is at most πε.
Remark. In the original Behrends' theorem [2, Theorem 3.3] there is a larger
constant 8/√2 in place of π. A similar result with the better constant π has been
obtained (in a different way) by I. Gasparis [5].
Sketch of the proof of Theorem 3.1. The proof is essentially the same as the original
one. Suppose that the conclusion were not true. We can find δ > 0 such that the
number
sup
x′∈SX ′ {diameter of the set of accumulation points of (x′(xnk ))k}
is greater than πε + δ for any subsequence (xnk ) of (xn). Fix τ ∈ (0, 1) such that
(2 + supn kxnk)τ < δ
π .
Similarly to the one in the proof of [2, Theorem 3.3 (or 3.2)] we can prove the
following lemma.
C*-ALGEBRAS HAVE A QUANTITATIVE VERSION OF PE LCZY ´NSKI'S PROPERTY (V) 5
Lemma. The sequence (xn) admits a subsequence (without loss of generality still
denoted by (xn)) which satisfies the following conditions:
(i) Whenever C and D are disjoint finite subsets of N, there are z0, w0 ∈ C with
w0 ≥ πε + δ and x′ ∈ X ′ with kx′k = 1 such that x′(xn)− z0 ≤ τ for n ∈ C
and x′(xn) − (z0 + w0) ≤ τ for n ∈ D.
(ii) There are i1 < ··· < ir in N and a1, . . . , ar ∈ C which satisfy
Finally, the time has come for the modification. By [11, Lemma 6.3] we find
Set C = {1, . . . , r} \ D. For these sets C and D we find z0, w0, and x′ from (i) of
the lemma. It follows that
r
Xρ=1
D ⊂ {1, . . . , r} such that
aρ = 1, (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
aρ(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xρ∈D
ε ≥(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
aρxiρ(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
≥(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
aρx′(xiρ )(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xρ=1
Xρ=1
≥(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
aρ(z0 + w0)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xρ∈C
aρz0 + Xρ∈D
≥ w0(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
aρ(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
aρ(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
− z0(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xρ=1
Xρ∈D
δ
π − (1 + z0)τ ≥ ε +
= ε +
r
r
r
r
Xρ=1
aρ(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
≤ ε.
r
Xρ=1
aρxiρ(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
≤ τ, (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xρ=1
r
1
π
≥
aρ =
1
π
.
=(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
r
aρx′(xiρ ) + Xρ∈D
Xρ∈C
Xρ=1
− τ ≥ w0
π − z0τ − τ ≥
aρ =(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xρ∈D
− τ
r
aρx′(xiρ )(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
aρz0(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
πε + δ
− τ
Xρ=1
π − (1 + z0)τ
aρw0 +
δ
π − (2 + sup
n kxnk)τ > ε,
which is a contradiction.
(cid:3)
4. Main theorem
This section is devoted to our main result -- a quantitative version of Pfitzner's
theorem (Theorem 4.1 below). We also prove a "real version" of this theorem
(Theorem 4.2).
Theorem 4.1. Let A be a C∗-algebra. Then for every bounded K ⊂ A′
(6)
wckA′(K) ≤ π · η(K).
Therefore A has the property (Vq).
Proof. The quantities γ(K) and wckA′ (K) are equivalent by [1, Theorem 2.3], more
specifically, the inequality (6) implies γ(K) ≤ 2π · η(K).
If this holds for each
bounded K ⊂ A′, Proposition [8, 3.2] mentioned also in Section 2 gives that A has
the property (Vq). Let us show the inequality (6).
Let K ⊂ A′ be bounded. The case wckA′ (K) = 0 is trivial. Suppose that
wckA′(K) > 0 and fix an arbitrary λ ∈ (0, wckA′(K)). By the definition of the
quantity wckA′(K) we find a sequence (x′
n) in K such that
dist(cid:0)clust(A′′′,w∗)(x′
n), A′(cid:1) > λ.
6
HANA KRULISOV ´A
Since every dual of a C∗-algebra is a predual of a von Neumann algebra, we deduce
from [13, Theorem III.2.14] (see also [6, Example IV.1.1(b)]) that A′ is L-embedded
-- it means that A′ is complemented in A′′′ by a projection P satisfying
kx′′′k = kP x′′′k + kx′′′ − P x′′′k,
x′′′ ∈ A′′′.
Consequently, from [7, Theorem 1] we have
δ(x′
nk ) : (x′
n) = inf{δ(x′
≥ 2 dist(cid:0)clust(A′′′,w∗)(x′
n), A′(cid:1) > 2λ.
nk ) is a subsequence of (x′
k)}
Fix an arbitrary ε > 0. We now prove the following claim.
Claim. There is a sequence of self-adjoint elements (xk) in BA satisfying xixj = 0,
i, j ∈ N, i 6= j, and a subsequence (x′
nk ) of the sequence (x′
n) such that
nk (xk)(cid:12)(cid:12) > (1 − ε)2 λ
(cid:12)(cid:12)x′
π
,
k ∈ N.
n is canonically decomposed in the following way: x′
Proof. Each x′
where fn, gn ∈ A′ are selfadjoint functionals. It suffices to find (xk) and (x′
that
n = fn + ign,
nk ) such
fnk (xk) > (1 − ε)2 λ
π
or
gnk (xk) > (1 − ε)2 λ
π
.
Indeed, since selfadjoint functionals attain real values on selfadjoint elements of A,
we have
nk (xk)(cid:12)(cid:12) = fnk (xk) + ignk (xk) ≥((cid:12)(cid:12) Re(fnk (xk) + ignk(xk))(cid:12)(cid:12) = fnk (xk)
(cid:12)(cid:12)x′
(cid:12)(cid:12) Im(fnk (xk) + ignk (xk))(cid:12)(cid:12) = gnk (xk)
We begin by proving that there is a strictly increasing sequence of indices (nk)
such that δ(fnk ) > λ or δ(gnk ) > λ. If δ(fn) > λ, the proof is over, so suppose that
n) > 2λ + 2τ . By the definition of δ(fn)
δ(fn) ≤ λ. Let us find τ > 0 satisfying δ(x′
there is a subsequence (fnk ) of the sequence (fn) with δ(fnk ) < λ + τ . We claim
that the corresponding subsequence (gnk ) of (gn) satisfies δ(gnk ) > λ. To obtain
a contradiction, suppose that δ(gnk ) ≤ λ. Using the definition of δ(gnk ) we find
a strictly increasing sequence of indices (kl) such that δ(gnkl
) < λ + τ . Then
.
δ(x′
nkl
) = δ(fnkl
= sup
+ ignkl
)
x′′∈BA′′
x′′∈BA′′
≤ sup
≤ sup
x′′∈BA′′
lim
l→∞
lim
l→∞
lim
l→∞
+ sup
x′′∈BA′′
sup
sup
p,q≥l(cid:12)(cid:12)x′′(fnkp + ignkp ) − x′′(fnkq + ignkq )(cid:12)(cid:12)
p,q≥l(cid:16)(cid:12)(cid:12)x′′(fnkp ) − x′′(fnkq )(cid:12)(cid:12) +(cid:12)(cid:12)x′′(gnkp ) − x′′(gnkq )(cid:12)(cid:12)(cid:17)
p,q≥l(cid:12)(cid:12)x′′(fnkp ) − x′′(fnkq )(cid:12)(cid:12)
p,q≥l(cid:12)(cid:12)x′′(gnkp ) − x′′(gnkq )(cid:12)(cid:12)
) < λ + τ + λ + τ = 2λ + 2τ,
sup
sup
lim
l→∞
= δ(fnkl
) + δ(gnkl
which contradicts the fact that δ(x′
n) > 2λ + 2τ .
Without loss of generality we may assume that we have found a subsequence
(fnk ) of the sequence (fn) with δ(fnk ) > λ and such that (fnk ) = (fn). By passing
C*-ALGEBRAS HAVE A QUANTITATIVE VERSION OF PE LCZY ´NSKI'S PROPERTY (V) 7
to a further subsequence we can also ensure that
inf n∈N kfnk
supn∈N kfnk
> 1 − ε.
Indeed, the sequence (fn) is bounded, hence we can find its subsequence (fnk ) such
that the limk→∞ kfnkk exists. This limit is nonzero, because otherwise we would
have δ(fn) = 0. We thus obtain the desired subsequence by omitting finitely many
members of (fnk ).
The inequality δ(fn) > λ says that for every subsequence (fnk ) of (fn) there is
some x′′ ∈ A′′ with kx′′k = 1 such that the diameter of the set of accumulation
points of the sequence (x′′(fnk ))k is greater than λ. By Theorem 3.1 the sequence
(fn) does not admit λ
π -- ℓ1 -- blocks, i.e. there is an infinite M ⊂ N such that whenever
i=1 ai = 1, and n1 < ··· < nr are indices in M , we have
π . Hence there is a subsequence (fnk ) of (fn) such that for each
a1, . . . , ar ∈ C satisfy Pr
(cid:13)(cid:13)Pr
i=1 aifni(cid:13)(cid:13) > λ
nonzero (αk) ∈ ℓ1 and N ∈ N large enough
N
∞
>
λ
π
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xk=1
By letting N → ∞ we obtain
λ
π
fnk(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
αk
PN
k=1 αk
αk ≤(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
αkfnk(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xk=1
Xk=1
Therefore we have for each (αk) ∈ ℓ1
kfnkk ≤(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
λ
Xk=1
Xk=1
π
π supk∈N kfnkk
αk ≤
Xk=1
Let us set
αk
λ
∞
∞
∞
∞
.
.
≤
∞
Xk=1
αk.
αk
fnk
kfnkk(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
r =
λ
π supk∈N kfnkk
and
θ = (1 − ε) r inf
k∈Nkfnkk.
Then
λ
π
λ
π
θ = (1 − ε)
inf k∈N kfnkk
supk∈N kfnkk ≥ (1 − ε)
inf n∈N kfnk
supn∈N kfnk ≥ (1 − ε)2 λ
Without loss of generality we can assume that (fnk ) = (fn). Then(cid:0) fn
(ak) ∈ ℓ1,
sequence consisting of selfadjoint elements which satisfies
αk,
≤
αk
π
∞
∞
∞
r
.
kfnk(cid:1)n is a basic
Xk=1
αk ≤(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xk=1
fk
kfkk(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xk=1
that is (36) of [10] (where a′
k = fk). By Pfitzner's proof of [10, Theorem 1] we
obtain a sequence (xk) in A and a subsequence (fnk ) of (fn) for which (35) of
[10] is valid (where a′
nk = fnk ), i.e. xk are selfadjoint elements in BA such that
xixj = 0, i, j ∈ N, i 6= j, and fnk (xk) > θ ≥ (1 − ε)2 λ
π , k ∈ N. This completes the
proof of the claim.
8
HANA KRULISOV ´A
Let (xk) and (x′
nk ) be sequences obtained by the claim. Since x′
nk (xk) >
(1 − ε)2 λ
π , k ∈ N, we have
lim sup
k→∞
sup
x′∈K x′(xk) ≥ (1 − ε)2 λ
π
.
N
N
N
N
Xk=1
µ(xk) =
But P xk is a wuC series in A satisfying supx′∈BA′ P x′(xk) ≤ 1. Indeed, all xk
are contained in a commutative subalgebra B of A, which can be identified with
the space C0(Ω) for some Ω by the Gelfand representatiton. Then xk, k ∈ N, are
real continuous functions on Ω with kxkk = supξ∈Ω xk(ξ) ≤ 1 and {xi 6= 0} ∩
{xj 6= 0} = ∅, i 6= j. Let x′ ∈ A′, and let us set µ = x′ ↾B∈ B′ = C0(Ω)′ = M(Ω).
For each N ∈ N we get
ZΩ
Xk=1
x′(xk) =
≤ZΩ
1 dµ = kµk ≤ kx′k.
Therefore supx′∈BA′ P∞
k=1 x′(xk) ≤ 1.
We thus obtain η(K) ≥ (1 − ε)2 λ
arbitrarily, it follows that η(K) ≥ 1
Remark. It is not clear whether C∗-algebras have also the property (Vq)∗
ω.
From [8, Theorem 4.1] it follows that the answer is affirmative for commutative
C∗-algebras. In fact we do not know any example of a Banach space with the prop-
erty (Vq) but not (Vq)∗
ω. Regarding the property (Vq)ω, we know from [8, Propo-
sition 4.3] that some (commutative) C∗-algebras enjoy this property and some do
not.
π . Since ε > 0 and λ < wckA′ (K) were chosen
π wckA′(K), which completes the proof.
(cid:3)
Xk=1Z{xk6=0} xk dµ
xk dµ(cid:12)(cid:12)(cid:12)(cid:12)
Xk=1(cid:12)(cid:12)(cid:12)(cid:12)
≤
The following theorem is a kind of "real version" of Theorem 4.1.
Theorem 4.2. Let A be a C∗-algebra. Then the space Asa has the property (Vq),
more precisely, for every bounded K ⊂ A′
(7)
sa
wckA′(K) ≤ η(K).
(0, wckA′
sa (K)). We find (fn) in K such that
Proof. The proof is analogous to the previous one, it suffices to use real versions of
the theorems that have allowed us to prove Theorem 4.1. Let us sketch it briefly.
sa(K) > 0 and an arbitrary λ ∈
sa(cid:1) > λ.
Consider a bounded set K ⊂ A′
dist(cid:0)clust((A′
Since A′ is L-embedded, the real version of A′ (let us denote it by (A′)R) is also
L-embedded. But (A′)sa is a 1-complemented subspace of (A′)R and is therefore
L-embedded by [6, Proposition IV.1.5]. We thus get
sa)′′,w∗)(fn), A′
sa with wckA′
from [7, Theorem 1]. Let us fix ε > 0. By passing to a subsequence we arrange
that
δ(fn) > 2λ
inf n∈N kfnk
supn∈N kfnk
> 1 − ε.
C*-ALGEBRAS HAVE A QUANTITATIVE VERSION OF PE LCZY ´NSKI'S PROPERTY (V) 9
By the real version of the quantitative Rosenthal's ℓ1 -- theorem [2, Theorem 3.2] the
sequence (fn) admits λ -- ℓ1 -- blocks, which yields a subsequence (fnk ) of the sequence
(fn) that for every (αn) ∈ ℓ1 satisfies
λ
supk∈N kfnkk
∞
Xk=1
∞
αk ≤(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xk=1
αk
fnk
kfnkk(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
≤
∞
Xk=1
αk.
Then we proceed exactly as in the proof of Theorem 4.1 to obtain the desired
conclusion.
(cid:3)
5. Relation to the Grothendieck property
Let us remind that a Banach space X has the Grothendieck property if every
weak∗ convergent sequence in its dual is weakly convergent. It is well known that
for dual Banach spaces the property (V) implies the Grothendieck property. In this
section we prove that this implication holds even for suitable quantitative versions
of these properties.
One possible quantification of the Grothendieck property has already been stud-
ied in [3] and [9]. Let us remind the definition: Let c > 0. A Banach space X is
c-Grothendieck if
(8)
whenever (x′
n) is a bounded sequence in X ′.
δ(x′
n) ≤ c · δw∗(x′
n)
A Banach space X has the Grothendieck property if and only if for every sequence
n) in X ′ the following implication holds:
(x′
n) is weak∗ Cauchy ⇒ (x′
(x′
n) is weakly Cauchy.
The inequality (8) quantifies this implication. But we can look at the Grothendieck
property also in another way: X has the Grothendieck property if and only if every
sequence (x′
n) in X ′ satisfies the implication
If we replace this implication by an inequality
n) is weak∗ Cauchy ⇒ {x′
(x′
wckX ′(cid:0){x′
n : n ∈ N} is a relatively weakly compact set.
n : n ∈ N}(cid:1) ≤ c · δw∗(x′
n)
where c > 0 is some constant not depending on (x′
n), we obtain another quantitative
version of the Grothendieck property. We will prove that all dual Banach spaces
with the property (Vq) have this kind of quantitative Grothendieck property (see
Corollary 5.2). We do not know whether the latter quantitative Grothendieck
property implies the former one (with a larger constant).
Theorem 5.1. Let X be a Banach space. Then for every bounded sequence (x′′
n)
in X ′′
2 δw∗ (x′′
n).
Proof. Let (x′′
n : n ∈ N}(cid:1) ≤ 1
η(cid:0){x′′
n) be a bounded sequence in X ′′. The case η(cid:0){x′′
n : n ∈ N}(cid:1) = 0 is
n : n ∈ N}(cid:1) > 0 and fix δ ∈(cid:0)0, η(cid:0){x′′
n : n ∈ N}(cid:1)(cid:1). Let us
n : n ∈ N}(cid:1) > δ + ε. By the definition of the quantity η we
k in X ′ with supx′′∈BX ′′ P∞
k=1 x′
k=1 x′′(x′
k) ≤ 1 such that
k) > δ + ε. Since (x′
k) is a weakly null sequence, there are
n) and (y′
n) >
n) is a bounded sequence in X ′′,
k) which for all n ∈ N satisfy y′′
trivial. Suppose that η(cid:0){x′′
find ε > 0 such that η(cid:0){x′′
can find a wuC series P∞
lim supk→∞ supn∈N x′′
subsequences of (y′′
δ + ε. The sequence (y′
n) is weakly null in X ′ and (y′′
n(x′
n) of (x′′
k) of (x′
n(y′
10
HANA KRULISOV ´A
hence by Simons' extraction lemma [12, Theorem 1] there is a strictly increasing
sequence of indices (nk) such that for all k ∈ N
nk (y′
y′′
nm ) < ε.
Xm∈N
m6=k
Let us define
αk =((−1)k sgn−1(cid:0)y′′
0,
nk (y′
nk )(cid:1),
nk (y′
y′′
nk (y′
y′′
nk ) 6= 0,
nk ) = 0,
k ∈ N,
where sgn denotes the complex signum function, i.e. sgn(z) = z
z , z ∈ C \ {0}. Set
x′ = w∗- lim
N→∞
N
Xk=1
αky′
nk ∈ X ′.
∞
Then x′ ∈ BX ′ because for all x ∈ BX
αkz′
∞
Xk=1
x′(x) =(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
For each k ∈ N even
nk (x)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
∞
≤
Re y′′
nk (x′) = αky′′
nk (y′
z′
nk (x) ≤
Xk=1
Xn=1
nk ) + Re Xm∈N
m6=k/2
x′
n(x) ≤ sup
x′′∈BX ′′
∞
Xn=1
x′′(x′
n) ≤ 1.
y′′
nk (α2my′
n2m)!
y′′
nk (α2m−1y′
nk (y′
y′′
nk (y′
≥ y′′
− Re Xm∈N
nk ) − Xm∈N
nk ) − Xm∈N
y′′
nk (y′
> (δ + ε) − ε = δ.
= y′′
nk (y′
m6=k/2
m6=k
n2m−1 )!
n2m ) − Xm∈N
nm)
nk (y′
y′′
n2m−1 )
Analogously, for each k ∈ N odd
nk (x′) = αky′′
Re y′′
nk (y′
nk ) + Re Xm∈N
y′′
nk (α2my′
y′′
nk (α2m−1y′
n2m)!
n2m−1)!
m6=(k+1)/2
− Re Xm∈N
nk ) + Xm∈N
< −(δ + ε) + ε = −δ.
≤ −y′′
nk (y′
m6=k
y′′
nk (y′
nm )
Therefore
inf
n∈N
k,l≥ny′′
sup
nk (x′) − y′′
nl(x′) ≥ inf
n∈N
sup
k,l≥n(cid:12)(cid:12) Re(cid:0)y′′
nk (x′) − y′′
nl(x′)(cid:1)(cid:12)(cid:12) ≥ 2δ.
C*-ALGEBRAS HAVE A QUANTITATIVE VERSION OF PE LCZY ´NSKI'S PROPERTY (V) 11
It follows that δw∗ (x′′
arbitrarily, we obtain the desired inequality.
n) ≥ δw∗ (y′′
nk ) ≥ 2δ. Since δ < η(cid:0){x′′
n : n ∈ N}(cid:1) was chosen
(cid:3)
Corollary 5.2. Let X be a Banach space and C > 0. Suppose that each bounded
K ⊂ X ′′ satisfy
(9)
(i.e. X ′ enjoys the property (Vq)). Then for every bounded sequence (x′′
n) in X ′′
wckX ′′ (K) ≤ C · η(K)
n : n ∈ N}(cid:1) ≤ 1
wckX ′′(cid:0){x′′
2 C · δw∗(x′′
n).
n : n ∈ N}.
Proof. It suffices to combine the previous theorem with the inequality (9) applied
to K = {x′′
(cid:3)
Corollary 5.3. Let A be a von Neumann algebra. Then A has a quantitative
version of the Grothendieck property -- more precisely, for every bounded sequence
(x′
n) in A′
Proof. Since every von Neumann algebra is a C∗-algebra and a dual Banach space,
the assertion follows from Theorem 4.1 and Corollary 5.2.
(cid:3)
wckA′(cid:0){x′
n : n ∈ N}(cid:1) ≤ 1
2 π δw∗ (x′
n).
Acknowledgement
The author wishes to express her gratitude to the referee, who suggested how to
improve Theorem 5.1 and shorten its proof. She also acknowledges many sugges-
tions and comments of her supervisor Ondrej Kalenda during the preparation of
the paper.
References
[1] C. Angosto and B. Cascales. Measures of weak noncompactness in Banach spaces. Topology
Appl., 156(7):1412 -- 1421, 2009.
[2] Ehrhard Behrends. New proofs of Rosenthal's l1-theorem and the Josefson-Nissenzweig the-
orem. Bull. Polish Acad. Sci. Math., 43(4):283 -- 295 (1996), 1995.
[3] Hana Bendov´a. Quantitative Grothendieck property. J. Math. Anal. Appl., 412(2):1097 -- 1104,
2014.
[4] Francesco S. De Blasi. On a property of the unit sphere in a Banach space. Bull. Math. Soc.
Sci. Math. R. S. Roumanie (N.S.), 21(69)(3-4):259 -- 262, 1977.
[5] I. Gasparis. ǫ-weak Cauchy sequences and a quantitative version of Rosenthal's ℓ1-theorem.
J. Math. Anal. Appl., 434(2):1160 -- 1165, 2016.
[6] P. Harmand, D. Werner, and W. Werner. M -ideals in Banach spaces and Banach algebras,
volume 1547 of Lecture Notes in Mathematics. Springer-Verlag, Berlin, 1993.
[7] O. F. K. Kalenda, H. Pfitzner, and J. Spurn´y. On quantification of weak sequential complete-
ness. J. Funct. Anal., 260(10):2986 -- 2996, 2011.
[8] Hana Krulisov´a. Quantification of Pe lczy´nski's property (V). Preprint is available at
http://arxiv.org/abs/1509.06610.
[9] Jindrich
Lechner.
1-Grothendieck
C(K)
spaces.
Preprint
is
available
at
http://arxiv.org/abs/1511.02202.
[10] H. Pfitzner. Weak compactness in the dual of a C ∗-algebra is determined commutatively.
Math. Ann., 298(2):349 -- 371, 1994.
[11] Walter Rudin. Real and complex analysis. McGraw-Hill Book Co., New York, third edition,
1987.
[12] S. Simons. On the Dunford-Pettis property and Banach spaces that contain c0. Math. Ann.,
216(3):225 -- 231, 1975.
[13] M. Takesaki. Theory of operator algebras. I, volume 124 of Encyclopaedia of Mathematical
Sciences. Springer-Verlag, Berlin, 2002. Reprint of the first (1979) edition, Operator Algebras
and Non-commutative Geometry, 5.
12
HANA KRULISOV ´A
Department of Mathematical Analysis, Faculty of Mathematics and Physics, Charles
University in Prague, Sokolovsk´a 83, 186 75, Praha 8, Czech Republic
E-mail address: [email protected]
|
1108.2598 | 1 | 1108 | 2011-08-12T10:14:23 | Traces on symmetrically normed operator ideals | [
"math.OA"
] | For every symmetrically normed ideal $\mathcal{E}$ of compact operators, we give a criterion for the existence of a continuous singular trace on $\mathcal{E}$. We also give a criterion for the existence of a continuous singular trace on $\mathcal{E}$ which respects Hardy-Littlewood majorization. We prove that the class of all continuous singular traces on $\mathcal{E}$ is strictly wider than the class of continuous singular traces which respect Hardy-Littlewood majorization. We establish a canonical bijection between the set of all traces on $\mathcal{E}$ and the set of all symmetric functionals on the corresponding sequence ideal. Similar results are also proved in the setting of semifinite von Neumann algebras. | math.OA | math |
TRACES ON SYMMETRICALLY NORMED OPERATOR IDEALS
F. SUKOCHEV AND D. ZANIN
We dedicate this paper to the memory of Nigel Kalton whose influence on us both has been
profound. Without his collaboration this paper would never have been written.
Abstract. For every symmetrically normed ideal E of compact operators, we
give a criterion for the existence of a continuous singular trace on E. We also
give a criterion for the existence of a continuous singular trace on E which
respects Hardy-Littlewood majorization. We prove that the class of all contin-
uous singular traces on E is strictly wider than the class of continuous singular
traces which respect Hardy-Littlewood majorization. We establish a canonical
bijection between the set of all traces on E and the set of all symmetric func-
tionals on the corresponding sequence ideal. Similar results are also proved in
the setting of semifinite von Neumann algebras.
1. Introduction
In his groundbreaking paper [6], J. Dixmier proved the existence of positive
singular traces (that is, linear positive unitarily invariant functionals which vanish
on all finite dimensional operators) on the algebra B(H) of all bounded linear
operators acting on infinite-dimensional separable Hilbert space H. Namely, if ψ :
R+ → R+ is a concave increasing function such that
(1)
lim
t→∞
ψ(2t)
ψ(t)
= 1,
then there is a singular trace τω, defined for every positive compact operator A ∈
B(H) by setting
(2)
τω(A) = ω(
1
ψ(n)
n
Xk=1
sk(A)).
Here, {sk(A)}k∈N is the sequence of singular values of the compact operator A ∈
B(H) taken in the descending order and ω is an arbitrary dilation invariant gen-
eralised limit on the algebra l∞ of all bounded sequences. This trace is finite
on 0 ≤ A ∈ B(H) if and only if A belongs to the Marcinkiewicz ideal (see e.g.
[14],[15],[27])
Mψ := {A ∈ B(H) : sup
n∈N
1
ψ(n)
n
Xk=1
sk(A) < ∞}.
2000 Mathematics Subject Classification. 47L20, 47B10, 46L52.
Key words and phrases. Symmetric functionals, singular traces.
1
2
F. SUKOCHEV AND D. ZANIN
In [18], Dixmier's result was extended to an arbitrary Marcinkiewicz ideal Mψ with
the following condition on ψ
(3)
lim inf
t→∞
ψ(2t)
ψ(t)
= 1.
All the traces defined above by formula (2) vanish on the ideal L1 consisting of
all compact operators A ∈ B(H) such that P∞
An ideal E of algebra B(H) is said to be symmetrically normed if {sk(B)}k∈N ≤
{sk(A)}k∈N and A ∈ E implies that kBkE ≤ kAkE (see [14], [15], [29]1, [28], [20]).
Since the ideal Mψ is just a special example of symmetrically normed operator
ideal, the following question (suggested in [18], [16], [17], [7]) arises naturally.
k=1 sk(A) < ∞.
Question 1. Which symmetrically normed operator ideals admit a nontrivial sin-
gular trace2?
In analyzing Dixmier's proof of the linearity of τω given by (1), it was observed
in [18] (see also [3]) that τω possesses the following fundamental property, namely
if 0 ≤ A, B ∈ Mψ are such that
(4)
sk(B) ≤
sk(A),
∀n ∈ N,
n
Xk=1
n
Xk=1
then τω(B) ≤ τω(A). Such a class of traces was termed "fully symmetric"in [20],
[30] (see also earlier papers [8],[25], where the term "symmetric"was used). It is
natural to consider such traces only on fully symmetrically normed operator ideals
E (that is, on symmetrically normed operator ideals E satisfying the condition:
if A, B satisfy (4) and A ∈ E, then B ∈ E and kBkE ≤ kAkE).
In fact, it was
established in [8] that every Marcinkiewicz ideal Mψ with ψ satisfying the condition
(3) possesses fully symmetric traces. Furthermore, in the recent paper [18], the
following unexpected result was established. If ψ satisfies the condition (3), then
every fully symmetric trace on Mψ is a Dixmier trace τω for some ω.
The following question ( also suggested in [18], [7], [16], [17]) arises naturally.
Question 2. Which fully symmetrically normed operator ideals admit a nontrivial
singular trace which is fully symmetric?
In papers [16],[17] the following two problems (closely related to Question 1 and
Question 2) were also suggested.
Question 3. Which fully symmetrically normed operator ideals admit a trace which
is not fully symmetric?
Let us fix an orthonormal basis {en}n∈N in H. An operator A ∈ B(H) is called
diagonal if (Aen, em) = 0 for every n 6= m.
Question 4. Let the mapping ϕ : E → C be unitarily invariant. Suppose that ϕ
is linear on the subset of all diagonal operators from E. Does it imply that ϕ is a
trace on E?
1 We have to caution the reader that in Theorem 1.16 of [29] the assertion (b) does not hold for
the norm of an arbitrary symmetrically normed ideal E (see e.g. corresponding counterexamples
in [19, p. 83]).
2 In this paper, we exclusively deal with positive traces
TRACES ON SYMMETRICALLY NORMED OPERATOR IDEALS
3
In some very special cases (for principal ideals contained in L1, which are, strictly
speaking, not symmetrically normed ideals), Question 3 was answered in the affir-
mative3 in [33]. In [20], question 3 was answered in the affirmative for the special
case of Marcinkiewicz ideals under the assumption (1). It should be pointed out
that the method used in [20] cannot be extended to an arbitrary Marcinkiewicz ideal
Mψ and, furthermore, cannot be extended to a general symmetrically normed op-
erator ideal. Question 4 was answered in [20] in full generality using deep results
from [11, 10] (see also [9]).
The following theorem is the main result of this paper.
It yields answers to
Questions 1 -- 3. In the course of the proof of Theorem 5, we also present a new (and
very simple) proof answering Question 4. Prior to stating Theorem 5, we make a
few preliminary observations, for which we are grateful to the referee.
Any trace ϕ : E → C obeys the condition
1
m
ϕ(A⊕m) = ϕ(A), A ∈ E, m ≥ 1.
Here, the direct sum A⊕m is formed with respect to some arbitrary Hilbert space
isomorphism H ⊕m ≃ H. Thus, traces are closely related to the following convex
(see Lemma 11 below) functional on E.
1
m
kA⊕mkE, A ∈ E.
π : A → lim
m→∞
The non-triviality of the functional π : E → R is an obvious necessary condition for
the existence of a trace.
Theorem 5. Let E be a symmetrically normed operator ideal. Consider the fol-
lowing conditions.
(1) There exist nontrivial singular traces on E.
(2) There exist nontrivial singular traces on E, which are fully symmetric.
(3) There exist nontrivial singular traces on E, which are not fully symmetric.
(4) E 6= L1 and there exist an operator A ∈ E such that
(5)
lim
m→∞
1
m
kA⊕mkE > 0.
(i) The conditions (1) and (4) are equivalent for every symmetrically normed
operator ideal E.
(ii) The conditions (1), (2) and (4) are equivalent for every fully symmetrically
normed operator ideal E.
(iii) The conditions (1) − (4) are equivalent for every fully symmetrically normed
operator ideal E equipped with a Fatou norm.
Recall that the norm on a symmetrically normed operator ideal E is called a
Fatou norm if the unit ball of E is closed with respect to strong (or, equivalently,
weak) operator convergence. Observe that classical ideals (such as Schatten-von
Neumann ideals Lp, Marcinkiewicz, Orlicz and Lorentz ideals [14], [15], [29]) have
a Fatou norm. In fact, in some standard references on the subject (e.g. Simon's
book [29]), the requirement that symmetrically normed operator ideal has a Fatou
norm appears to be a part of the definition. Similarly, in the book [24], devoted
to the study of symmetric4 function spaces (which are a commutative counterpart
3We are grateful to the referee for this remark.
4termed there "rearrangement invariant".
4
F. SUKOCHEV AND D. ZANIN
of symmetrically normed operator ideals), an assumption that the norm is a Fatou
norm is incorporated into the definition [24, p. 118].
The proof of Theorem 5 is given in Section 7.
In fact, in this paper we will
prove a more general result for symmetric spaces associated with semifinite von
Neumann algebras. The precise statements are given in Section 4 (see Theorems
23, 28, 29), Section 5 (see Theorems 33, 35, 36) and Section 6 (see Theorems 47,
48). The appendix contains the proof of important technical results for which we
were unable to find a suitable reference. We also present a new and short proof of
the Figiel-Kalton theorem from [13].
Finally, we say a few words about our proof and its relation to the previous
results in the literature. Our strategy is based on the approach from recent papers
[30] and [21], where condition (5) was connected to the geometry of E (see also [2]).
The condition (5) is easy to verify in concrete situations. For example, the following
corollary of Theorem 5 strengthens the main result of [20] and complements earlier
results of J. Varga [32].
Corollary 6. Every Marcinkiewicz ideal Mψ with ψ satisfying the condition (3)
admits a trace which is not fully symmetric.
Indeed, it is proved in [1, Proposition 2.3] that the condition (4) of Theorem 5
is equivalent to the condition (3) for the Marcinkiewicz ideal Mψ. Some examples
of symmetrically normed operator ideals, which are not Marcinkiewicz ideals, pos-
sessing symmetric traces were presented in [7]. These results are also an immediate
corollary of Theorem 5.
For completeness, we note that the assertion (ii) in Theorem 5 holds for a wider
class of relatively fully symmetrically normed operator ideals. The latter class is
defined as follows:
if A, B ∈ E are such that (4) holds, then kBkE ≤ kAkE . It
coincides with the class of all symmetrically normed subspaces of a fully symmetric
operator ideal (see [19])
2. Definitions and preliminaries
The theory of singular traces on symmetric operator ideals rests on some classical
analysis which we now review for completeness.
As usual, L∞(0, ∞) is the set of all bounded Lebesgue measurable functions on
the semi-axis equipped with the uniform norm. Given a function x ∈ L∞(0, ∞),
one defines its decreasing rearrangement t → µ(t, x) by the formula (see e.g. [22])
µ(t, x) = inf{s ≥ 0 : m({x > s}) ≤ t}.
Let H be a Hilbert space and let B(H) be the algebra of all bounded operators on
H equipped with the uniform norm.
Let M ⊂ B(H) be a semi-finite von Neumann algebra equipped with a fixed
faithful and normal semi-finite trace τ. M is said to be atomic (see [31, Definition
5.9]) if every nonzero projection in M contains a nonzero minimal projection. M
is said to be atomless if there is no minimal projections in M.
For every A ∈ M, the generalised singular value function t → µ(t, A) is defined
by the formula (see e.g. [12])
µ(t, A) = inf{kApk : τ (1 − p) ≤ t}.
If, in particular, M = B(H), then µ(A) is a step function and, therefore, can be
identified with the sequence of singular numbers of the operators A (the singular
TRACES ON SYMMETRICALLY NORMED OPERATOR IDEALS
5
values are the eigenvalues of the operator A = (A∗A)1/2 arranged with multiplicity
in decreasing order).
Equivalently, µ(A) can be defined in terms of the distribution function dA of A.
That is, setting
we obtain
dA(s) = τ (EA(s, ∞)),
s ≥ 0,
µ(t, A) = inf{s ≥ 0 : dA(s) ≤ t},
t > 0.
Here, EA denotes the spectral measure of the operator A.
Using the Jordan decomposition, every operator A ∈ B(H) can be uniquely
written as
A = (ℜA)+ − (ℜA)− + i(ℑA)+ − i(ℑA)−.
Here, ℜA := 1/2(A + A∗) (respectively, ℑA := 1/2i(A − A∗)) for any operator
A ∈ B(H) and B+ = BEB(0, ∞) ( respectively, B = −BEB(−∞, 0)) for any
self-adjoint operator B ∈ B(H). Recall that ℜA, ℑA ∈ M for every A ∈ M and
B+, B− ∈ M for every self-adjoint B ∈ M.
Further, we need to recall the important notion of Hardy -- Littlewood majoriza-
tion. Let A, B ∈ (L1 + L∞)(M). The operator B is said to be majorized by A and
written B ≺≺ A if and only if
Z t
0
µ(s, B)ds ≤Z t
0
µ(s, A)ds,
t ≥ 0.
We have (see [12])
A + B ≺≺ µ(A) + µ(B) ≺≺ 2σ1/2µ(A + B)
for every positive operators A, B ∈ (L1 + L∞)(M).
If s > 0, the dilation operator σs is defined by setting
(σs(x))(t) = x(
t
s
),
t > 0
in the case of the semi-axis. In the case of the interval (0, 1), the operator σs is
defined by
(σsx)(t) =(x(t/s),
0,
t ≤ min{1, s}
s < t ≤ 1.
Similarly, in the sequence case, we define an operator σn by setting
and an operator σ1/2 by setting
n times
n times
σn(a1, a2, · · · ) = (a1, · · · , a1
, a2, · · · , a2
, · · · )
{z
}
{z
}
σ1/2 : (a1, a2, a3, a4, · · · ) → (
a1 + a2
2
,
a3 + a4
2
, · · · ).
Definition 7. The Banach space E(M, τ ) ⊂ (L1 + L∞)(M) is said to be a sym-
metric operator space if the following conditions hold.
(1) Given A ∈ E(M, τ ) and B ∈ (L1 + L∞)(M) with µ(B) = µ(A), we have
B ∈ E(M, τ ) and kBkE = kAkE.
(2) Given 0 ≤ A ∈ E(M, τ ) and 0 ≤ B ∈ (L1 + L∞)(M) with B ≤ A, we have
B ∈ E(M, τ ) and kBkE ≤ kAkE.
6
F. SUKOCHEV AND D. ZANIN
The space E(M, τ ) is called fully symmetric if for every A ∈ E(M, τ ) and every
B ∈ (L1 + L∞)(M) with B ≺≺ A, we have B ∈ E(M, τ ) and kBkE ≤ kAkE.
The norm on a symmetric space E(M, τ ) is a Fatou norm if the unit ball of
E(M, τ ) is closed with respect to strong (or, equivalently, weak) operator conver-
gence. Every symmetric space equipped with a Fatou norm is necessarily fully
symmetric.
A linear functional ϕ : E(M, τ ) → C is said to be symmetric if ϕ(B) = ϕ(A)
for every positive A, B ∈ E(M, τ ) such that µ(B) = µ(A). A linear functional
ϕ : E(M, τ ) → C is said to be fully symmetric if ϕ(B) ≤ ϕ(A) for every positive
A, B ∈ E(M, τ ) such that B ≺≺ A. Every fully symmetric functional is symmetric
and bounded. The converse fails [20].
A functional ϕ : E(M, τ ) → C is called singular if ϕ = 0 on (L1 ∩ L∞)(M). If
E(M, τ ) 6⊂ L1(M), then every symmetric functional is singular.
If E = E(0, ∞) and if ϕ : E → R is a symmetric functional, then sϕ(x) = ϕ(σsx)
for every x ∈ E. If E = E(0, 1) and if ϕ : E → R is a singular symmetric functional,
then sϕ(x) = ϕ(σsx) for every x = µ(x) ∈ E.
Let E be a fully symmetric Banach space either on the interval (0, 1) or on the
semi-axis. We need the notion of an expectation operator (see [2]).
Let A = {Ak} be a (finite or infinite) sequence of disjoint sets of finite measure
and denote by A the collection of all such sequences. Denote by A∞ the complement
of ∪kAk.
The expectation operator E(·A) : L1 + L∞ → L1 + L∞ is defined by setting
E(xA) =Xk
1
m(Ak)
(ZAk
x(s)ds)χAk .
Note that we do not require A∞ to have finite measure.
Every expectation operator is a contraction both in L1 and L∞. Therefore,
E(xA) ≺≺ x,
x ∈ L1 + L∞.
It follows that E(·A) is also contraction in E.
It will be convenient to introduce the following notation.
If A is a discrete
subset of the semi-axis (i.e. a subset without limit points inside (0, ∞)), then the
elements of A ∪ {0} partition the semi-axis. This partition consists of a (finite
or infinite) sequence of sets of finite measure. We identify this partition with the
set A. Elements of A will be called nodes of the partition A. The corresponding
averaging operator will be denoted by E(·A).
Let E be a symmetric Banach space either on the interval (0, 1) or on the semi-
axis. Define the sets
DE = Lin({x ∈ E : x = µ(x)}) = {µ(a) − µ(b), a, b ∈ E},
ZE = Lin({x1 − x2 : 0 ≤ x1, x2 ∈ E, µ(x1) = µ(x2)}).
Let C be a Hardy operator defined by setting
(Cx)(t) =
1
t Z t
0
x(s)ds.
The following theorem was proved in [13]. For convenience of the reader, we give
a new and simple proof in the appendix.
TRACES ON SYMMETRICALLY NORMED OPERATOR IDEALS
7
Theorem 8. Let E be a symmetric space on the semi-axis and let x ∈ DE. We
have x ∈ ZE if and only if Cx ∈ E. A similar assertion is also valid for the interval
(0, 1) provided that R 1
0 x(s)ds = 0.
The following uniform submajorization was introduced by Kalton and Sukochev
in [19].
Let x, y ∈ L1(0, 1) (or x, y ∈ (L1 + L∞)(0, ∞)). We say that y ⊳ x if there exists
m ∈ N such that
(6)
Z b
ma
µ(s, y)ds ≤Z b
a
µ(s, x)ds,
∀ma ≤ b.
Let x, y ∈ l∞. We say that y ⊳ x if there exists m ∈ N such that
(7)
b
Xk=ma+1
µ(k, y) ≤
b
Xk=a+1
µ(k, x) ∀ma + 1 ≤ b.
The following important theorem was proved in [19] (see Theorem 5.4 and The-
orem 6.3 there).
Theorem 9. Let x, y ∈ L1(0, 1) or x, y ∈ (L1 + L∞)(0, ∞) or x, y ∈ l∞ be such
that y ⊳ x. For every ε > 0, the function (1 − ε)y belongs to a convex hull of the set
{z : µ(z) ≤ µ(x)}.
This theorem led to the following fundamental result (see [19]).
Theorem 10. Let E = E(0, 1) (or E = E(0, ∞) or E = E(N)) be a symmetric
Banach space either on the interval (0, 1) or on the semi-axis or on N. It follows
that the corresponding set E(M, τ ) is a symmetric Banach space.
Also, the uniform submajorization permits us to prove the convexity of the
functional π : E → R defined in Section 1.
Lemma 11. The functional π : E → R is convex on every symmetrically normed
operator ideal E.
Proof. Let E be the corresponding symmetrically normed ideal of l∞. For every
A, B ∈ E, it follows from Proposition 8.6 of [19] that µ(A + B) ⊳ µ(A) + µ(B).
Hence, σmµ(A + B) ⊳ σm(µ(A) + µ(B)). By Theorem 9, we have
kσmµ(A + B)kE ≤ kσm(µ(A) + µ(B))kE ≤ kσmµ(A)kE + kσmµ(B)kE.
Note that kA⊕mkE = kσmµ(A)kE. Dividing by m and letting m → ∞, we obtain
π(A + B) ≤ π(A) + π(B).
(cid:3)
3. Lifting of symmetric functionals
In this section, we explain a canonical bijection between symmetric functionals
and traces. In what follows, we require that a semifinite von Neumann algebra M
be either atomless or atomic with traces of all atoms being 1.
For an atomless von Neumann algebra M, we have (see e.g. [12])
Z t
0
µ(s, A)ds = sup{τ (pA) : p ∈ P (M), τ (p) = t}, A ∈ M.
8
F. SUKOCHEV AND D. ZANIN
For a atomic von Neumann algebra M, we have (see e.g. [12])
µ(k, A) = sup{τ (pA) : p ∈ P (M), τ (p) = m}, A ∈ M.
m
Xk=1
In either case, this implies a remarkable inequality (see e.g. [12])
(8) µ(A + B) ≺≺ µ(A) + µ(B) ≺≺ 2σ1/2µ(A + B),
0 ≤ A, B ∈ (L1 + L∞)(M).
Lemma 12. Let E = E(0, 1) (or E = E(0, ∞) or E = E(N)) be a symmetric
Banach space either on the interval (0, 1) or on the semi-axis or on N. If x, y ∈ E+
are such that y ⊳ x, then ϕ(y) ≤ ϕ(x) for every positive symmetric functional ϕ on
E.
Proof. Fix ε > 0. By Theorem 9, there exist zk ∈ E, 1 ≤ k ≤ n, and positive
numbers λk, 1 ≤ k ≤ n, such that µ(zk) ≤ µ(x) for every 1 ≤ k ≤ n and
(1 − ε)y =
λkzk,
n
Xk=1
λk = 1.
n
Xk=1
Since ϕ is positive and symmetric, it follows that
ϕ(zk) ≤ ϕ(zk) = ϕ(µ(zk)) ≤ ϕ(µ(x)) = ϕ(x).
Therefore, (1 − ε)ϕ(y) ≤ ϕ(x). Since ε > 0 is arbitrarily small, the assertion follows.
(cid:3)
The following assertion is essentially known. However, we provide the full proof
for readers convenience.
Lemma 13. Let M be a semifinite atomless von Neumann algebra and let A, B ∈
(L1 + L∞)(M, τ ) be positive operators.
2a
µ(s, A + B)ds ≤Z b
Z b
(µ(s, A) + µ(s, B))ds ≤Z 2b
Z b
2a
2a
a
(µ(s, A) + µ(s, B))ds,
∀2a ≤ b,
µ(s, A + B)ds,
∀2a ≤ b.
Similar assertion is valid for atomic von Neumann algebra M.
Proof. Applying inequality (8) to the operators A, B, we obtain that
and
Subtracting this inequalities, we obtain
0
Z b
Z 2a
Z b
0
0
µ(s, A + B)ds ≤Z b
µ(s, A + B)ds ≥Z a
µ(s, A + B)ds ≤Z b
0
(µ(s, A) + µ(s, B))ds
(µ(s, A) + µ(s, B))ds.
(µ(s, A) + µ(s, B))ds.
a
Proof of the second inequality is identical.
2a
(cid:3)
The following theorem answers Question 4 in the affirmative, as also does [20,
Theorem 5.2]. The proof below is very simple and based on a completely different
approach.
TRACES ON SYMMETRICALLY NORMED OPERATOR IDEALS
9
Theorem 14. Let E = E(0, 1) (or E = E(0, ∞) or E = E(N)) be a symmetric
Banach space either on the interval (0, 1) or on the semi-axis or on N and let
E(M, τ ) be the corresponding symmetric Banach operator space.
(1) If ϕ is a positive symmetric functional on E, then there exists a positive
symmetric functional L(ϕ) on E(M, τ ) such that ϕ(x) = L(ϕ)(A) for all
positive x ∈ E and A ∈ E(M, τ ) such that µ(A) = µ(x).
(2) If ϕ is a positive symmetric functional on E(M, τ ), then there exists a
positive symmetric functional L−1(ϕ) on E such that ϕ(A) = L−1(ϕ)(x)
for all positive x ∈ E and A ∈ E(M, τ ) such that µ(A) = µ(x).
Proof. We will only prove (1). Proof of (2) is identical.
Let A, B ∈ E+(M, τ ). It follows from Lemma 13 that
µ(A + B) ⊳ µ(A) + µ(B) ⊳ 2σ1/2µ(A + B).
It follows from Lemma 12 that
ϕ(µ(A + B)) ≤ ϕ(µ(A) + µ(B)) ≤ ϕ(2σ1/2µ(A + B)) = ϕ(µ(A + B)).
It follows that L(ϕ) is additive on E+(M, τ ). We than extend it to E(M, τ ) by
linearity.
(cid:3)
Theorem 14 provides a very natural bijection between the set of all symmetric
functionals on E and that on E(M, τ ), observed first for the case of fully symmetric
functionals in [8]. Next corollary follows immediately.
Corollary 15. Let E and E(M, τ ) be as in Theorem 14. The functional ϕ is fully
symmetric on E if and only if L(ϕ) is a fully symmetric functional on E(M, τ ).
We also need a lifting between sequence and function spaces. The following space
was introduced in [21].
Let A = {[n − 1, n]}n∈N be a partition of the semi-axis. Clearly, E(·A) maps
L1 + L∞ into the set of step functions which can be identified with sequences.
Proposition 16. Let E be a symmetric Banach sequence space and let F be the
linear space of all such functions x ∈ L∞ for which E(µ(x)A) ∈ E. The space F
equipped with the norm
is a symmetric Banach function space.
kxkF = kxk∞ + kE(µ(x)A)kE
The fact that the space F is a Banach space is non-trivial. Proof of this fact was
missing in both [19] and [21]. We include it in the appendix.
Below, we assume that E is embedded into F.
Theorem 17. Let E = E(N) be a symmetric Banach sequence space and let F be
the corresponding function space.
(1) If ϕ is a positive symmetric functional on E, then there exists a positive
symmetric functional L(ϕ) on F such that ϕ(E(µ(x)A)) = L(ϕ)(x) for all
positive x ∈ F.
(2) If ϕ is a positive symmetric functional on F, then its restriction on E is a
positive symmetric functional. This restriction is an inverse operation for
the L in (1).
10
F. SUKOCHEV AND D. ZANIN
Proof. Let us prove (1)
ϕ(σ1/2a) = 1/2ϕ(a1, a3, · · · ) + 1/2ϕ(a2, a4, · · · ) =
= 1/2ϕ(a1, 0, a2, 0, · · · ) + 1/2ϕ(0, a2, 0, a4, · · · ) = 1/2ϕ(a)
for every a ∈ E.
Let x, y ∈ F be positive. It follows from Lemma 50 that
E(µ(x + y)A) ⊳ E(µ(x) + µ(y)A) ⊳ 2σ1/2E(µ(x + y)A).
It follows from Lemma 12 that
ϕ(E(µ(x + y)A)) = ϕ(E(µ(x) + µ(y)A))
and (1) follows.
The first assertion of (2) is trivial. Clearly, µ(x) − E(µ(x)A) ∈ (L1 ∩ L∞)(0, ∞).
If E 6= l1, then ϕ(y) = 0 for every y ∈ (L1 ∩ L∞)(0, ∞) and every symmetric
functional ϕ on F. If E = l1, then F = (L1 ∩ L∞)(0, ∞) and the only symmetric
functional on both spaces is an integral. The second assertion of (2) follows.
(cid:3)
4. Existence of symmetric functionals
In this section, we present results concerning existence of symmetric functionals
on symmetric function spaces. The main results of this section are Theorem 23,
Theorem 28 and Theorem 29.
We need the following variation of the Hahn-Banach theorem.
Lemma 18. Let E be a partially ordered linear space and let p : E → R be convex
and monotone functional. For every x0 ∈ E, there exists a positive linear functional
ϕ : E → R such that ϕ ≤ p and ϕ(x0) = p(x0).
Proof. The existence of ϕ follows from the Hahn-Banach theorem. We only have
to prove that ϕ ≥ 0. If z ≥ 0, then ϕ(x0 − z) ≤ p(x0 − z). Therefore,
ϕ(z) ≥ ϕ(x0) − p(x0 − z) = p(x0) − p(x0 − z) ≥ 0
due to the fact that z ≥ 0 and p is monotone.
(cid:3)
Define operators Mm : (L1 +L∞)(0, ∞) → (L1+L∞)(0, ∞) (or, Mm : L1(0, 1) →
L1(0, 1)) by setting
(Mmx)(t) =
1
t log(m)Z t
t/m
x(s), m ≥ 2.
Lemma 19. If 0 ≤ x ∈ L1 + L∞ (or, 0 ≤ x ∈ L1(0, 1)), then
Z b/m
a
x(s)ds ≤Z b
a
(Mmx)(s)ds ≤Z b
a/m
x(s)ds
provided that ma ≤ b. In particular, m−1σmx ⊳ Mmx ⊳ x provided that x = µ(x).
Proof. Clearly,
(Mmx)(s)ds =
a
Z b
a/mZ min{ms,b}
max{a,s}
=
1
log(m)Z b
x(s) log(
min{ms, b}
max{a, s}
)ds.
x(s)ds
dt
t
=
1
t/m
log(m)Z b
a Z t
log(m)Z b
1
a/m
dt
t
x(s)ds =
TRACES ON SYMMETRICALLY NORMED OPERATOR IDEALS
11
The integrand does not exceed x(s) log(m) and the second inequality follows im-
mediately. The integrand is positive and is equal to x(s) log(m) for s ∈ (a, b/m).
The first inequality follows.
(cid:3)
Corollary 20. If E is a symmetric Banach function space either on the interval
(0, 1) or on the semi-axis, then Mm : E → E is a contraction for m ∈ N.
Proof. Let x = µ(x) ∈ E. It follows from Lemma 19 that Mmx ⊳ x. It follows from
theorem 9 that, for every ε > 0, the function (1 − ε)Mmx belongs to a convex hull
of the set {z : µ(z) ≤ µ(x)}. Therefore, Mmx ∈ E and (1 − ε)kMmxkE ≤ kxkE.
Since ε is arbitrarily small, the assertion follows.
(cid:3)
Lemma 21. Let E be a symmetric Banach space either on the interval (0, 1) or
on the semi-axis. Let p : DE → R be convex and monotone functional. If p = 0 on
ZE ∩ DE, then p extends to a convex monotone functional p : E → R by setting
Also, p(x) = 0 for every x ∈ ZE.
p(x) = p(µ(x+) − µ(x−)).
Proof. If x ∈ DE, then x − µ(x+) + µ(x−) ∈ ZE ∩ DE. Therefore, p(x − µ(x+) +
µ(x−)) = 0 and, due to the convexity of p, p(x) = p(µ(x+) − µ(x−)). This proves
the correctness of the definition.
For x, y ∈ E, we have
µ((x + y)+) − µ((x + y)−) − µ(x+) + µ(x−) − µ(y+) + µ(y−) ∈ ZE ∩ DE.
It follows that
p(µ((x + y)+) − µ((x + y)−) − µ(x+) + µ(x−) − µ(y+) + µ(y−)) = 0
and
p(x + y) = p(µ((x + y)+) − µ((x + y)−)) =
= p(µ(x+) − µ(x−) + µ(y+) − µ(y−)) ≤ p(x) + p(y).
Since p is monotone on DE, then p(y) ≤ 0 for every 0 ≥ y ∈ DE. It follows that
p(y) = p(−µ(y)) ≤ 0 for 0 ≥ y ∈ E. Therefore, p(x + y) ≤ p(x) + p(y) ≤ p(x) for
every 0 ≥ y ∈ E.
(cid:3)
Lemma 22. Let E be a symmetric Banach space either on the interval (0, 1) or
on the semi-axis. The functional
p : x → lim sup
m→∞
k(Mmx)+kE,
x ∈ DE
satisfies the assumptions of Lemma 21. Also, for every x ∈ DE, we have p(x) ≤
kxkE.
Proof. It follows from Corollary 20 that
k(Mmx)+kE ≤ kMmxkE ≤ kxkE,
x ∈ E.
It follows that
p(x) = lim sup
m→∞
k(Mmx)+kE ≤ kxkE,
x ∈ DE.
Clearly, the mappings x → (Mmx)+ are convex and monotone. So are the mappings
x → k(Mmx)+kE. Therefore, p : DE → R is a convex and monotone functional.
12
F. SUKOCHEV AND D. ZANIN
If x ∈ ZE ∩ DE, then by Theorem 8 Cx ∈ E. Therefore,
(Mmx)(t) ≤
1
log(m)
(
≤
1
log(m)
(
1
m
1
t Z t/m
0
x(s)ds +
1
t Z t
0
x(s)ds) ≤
σmCx + Cx)(t).
Since kσmkE→E ≤ m (see [22, Theorem II.4.5]), it follows that
k(Mmx)+kE ≤
2
log(m)
kCxkE
and p(x) = 0.
(cid:3)
Theorem 23. Let E = E(0, ∞) be a symmetric Banach space on the semi-axis.
For a given 0 ≤ x ∈ E, there exists a symmetric linear functional ϕ : E → R such
that
ϕ(x) = lim
m→∞
kσm(µ(x))kE.
1
m
Proof. Without loss of generality, x = µ(x). Let p be the convex monotone func-
tional constructed in Lemma 22.
It follows from Lemma 18 that there exist a
positive linear functional ϕ on E such that ϕ ≤ p and ϕ(x) = p(x). Since p(z) = 0
for every z ∈ ZE, it follows that ϕ(z) = 0 for every z ∈ ZE. Therefore, ϕ is a
symmetric functional.
Since ϕ(z) ≤ p(z) ≤ kzkE for every z = µ(z) ∈ E, it follows that kϕkE ∗ ≤ 1.
Therefore,
ϕ(x) = ϕ(
1
m
σmx) ≤
1
m
kσmxkE.
Passing m → ∞, we obtain
ϕ(x) ≤ lim
m→∞
1
m
kσmµ(x)kE.
On the other hand, It follows from Lemma 19 that m−1σmx ⊳ Mmx. Therefore,
p(x) = lim sup
m→∞
kMmxkE ≥ lim
m→∞
1
m
kσmµ(x)kE.
The assertion follows immediately.
(cid:3)
Consider the functional π : E → E (identical to the one defined in Section 1).
(9)
π(x) = lim
m→∞
1
m
kx⊕mkE,
x ∈ E.
Note that π(−x) = π(x) for every x ∈ E. If p is a functionals defined in Lemma 22,
then p(−x) = 0 for positive x ∈ E. Therefore, p 6= π. However, the assertion below
follows from Theorem 23.
Lemma 24. Let E = E(0, ∞) be a symmetric Banach space on the semi-axis. Let
p and π be the convex functionals on E defined in Lemma 22 and (9), respectively.
For every positive x ∈ E, we have p(x) = π(x).
Proof. For every x ∈ E, consider the functional ϕ constructed in Theorem 23. By
construction, we have ϕ(x) = p(x) = π(x).
(cid:3)
If E 6⊂ L1(0, ∞), then the functional ϕ constructed in Theorem 23 is necessarily
singular. The case E ⊂ L1 requires more detailed treatment.
TRACES ON SYMMETRICALLY NORMED OPERATOR IDEALS
13
Lemma 25. Let E be a symmetric (respectively, fully symmetric) Banach function
space either on the interval (0, 1) or on the semi-axis. Let {ϕi}i∈I ∈ E∗ be a net
and let ϕ ∈ E∗ be such that ϕi → ϕ ∗−weakly.
(1) If every ϕi is symmetric, then ϕ is symmetric.
(2) If every ϕi is fully symmetric, then ϕ is fully symmetric.
Proof. Let each ϕi be symmetric. If 0 ≤ x1, x2 ∈ E are such that µ(x1) = µ(x2),
then
ϕ(x1) = lim
i∈I
ϕi(x1) = lim
i∈I
ϕi(x2) = ϕ(x2).
Hence, ϕ is symmetric.
Let each ϕi be fully symmetric. Thus, ϕi(x) ≤ 0 for every x ∈ DE such that
Cx ≤ 0. Therefore, ϕ(x) = limi∈I ϕi(x) ≤ 0 for every x ∈ DE such that Cx ≤ 0.
Let x1, x2 ∈ E be positive elements such that x1 ≺≺ x2. Therefore, z = µ(x1) −
µ(x2) ∈ DE and Cz ≤ 0. It follows from above that ϕ(z) ≤ 0. Hence, ϕ is a fully
symmetric functional.
(cid:3)
Lemma 26. Let E be a symmetric (respectively, fully symmetric) Banach function
space either on the interval (0, 1) or on the semi-axis and let ϕ be a symmetric
(respectively, fully symmetric) functional on E. The formula
ϕsing(x) = lim
n→∞
ϕ(µ(x)χ(0,1/n)),
0 ≤ x ∈ E.
defines a singular symmetric (respectivley, fully symmetric) linear functional on E.
Proof. If x, y ∈ E are positive functions, then
µ(x + y)χ(0,1/n) ⊳ (µ(x) + µ(y))χ(0,1/n) ⊳ 2σ1/2µ(x + y)χ(0,1/n).
Taking the limit as n → ∞, we derive from Lemma 12 that
ϕsing(µ(x + y)) = ϕsing(µ(x) + µ(y)).
Since ϕ is symmetric, it follows that
ϕsing(x + y) = ϕsing(µ(x + y)) = ϕsing(µ(x) + µ(y)) = ϕsing(x) + ϕsing(y).
Hence, ϕsing is an additive functional on E+. Therefore, it extends to a linear
functional on E. Clearly, ϕsing is symmetric. Second assertion is trivial.
(cid:3)
In fact, the construction in Lemma 26 gives a singular part of the functional ϕ
as defined by Yosida-Hewitt theorem.
Lemma 27. Let E = E(0, ∞) ⊂ L1(0, ∞) be a symmetric Banach function space
on the semi-axis and let ϕ be a symmetric functional on E. If ϕsing is a functional
constructed in Lemma 26, then ϕ−ϕsing is a normal functional (that is, an integral).
Proof. It is clear that
0 ≤ ϕsing(z) ≤ kzk∞ lim
n→∞
ϕ(χ(0,1/n)) = 0
for every positive z ∈ (L1 ∩ L∞)(0, ∞). It follows that ϕsing(µ(x)χ(1/n,∞)) = 0 for
every x ∈ E. Therefore,
(10)
(ϕ − ϕsing)(x) = lim
n→∞
ϕ(µ(x)χ(1/n,∞)) = lim
n→∞
(ϕ − ϕsing)(µ(x)χ(1/n,∞)).
14
F. SUKOCHEV AND D. ZANIN
On the other hand, for every positive z ∈ (L1 ∩ L∞)(0, ∞) with kzk∞ = 1, we have
z ≺ χ(0,kzk1). It is proved in [30, Theorem 23] that z belongs to the closure (in the
topology of L1 ∩ L∞) of the set {u ≥ 0 : µ(u) = χ(0,kzk1)}. Thus,
(ϕ − ϕsing)(z) = (ϕ − ϕsing)(χ(0,kzk1)) = kzk1(ϕ − ϕsing)(χ(0,1)).
By linearity,
(11)
(ϕ − ϕsing)(z) = (ϕ − ϕsing)(χ(0,1)) ·Z ∞
0
It follows that from (10) and (11) that
z(s)ds,
∀z ∈ (L1 ∩ L∞)(0, ∞).
(ϕ−ϕsing)(x) = lim
n→∞Z ∞
1/n
µ(s, x)ds·(ϕ−ϕsing )(0, 1) =Z 1
0
x(s)ds·(ϕ−ϕsing )(χ(0,1))
for every positive function x ∈ E. The assertion follows immediately.
(cid:3)
Theorem 28. Let E ⊂ L1(0, ∞) be a symmetric Banach space on the semi-axis.
For a given 0 ≤ x ∈ E, there exists a singular symmetric linear functional ϕsing
such that
ϕsing(x) = lim
m→∞
1
m
kσm(µ(x))χ(0,1)kE.
Proof. Apply Theorem 23 to the function µ(x)χ(0,1/n). It follows that there exists
a symmetric linear functional ϕn such that kϕnkE ∗ ≤ 1 and
ϕn(µ(x)χ(0,1/n)) = lim
m→∞
1
m
kσm(µ(x)χ(0,1/n))kE ≥ lim
m→∞
1
m
kσm(µ(x))χ(0,1)kE.
Since the unit ball in E∗ is ∗−weakly compact (Banach-Alaoglu theorem), there
exists a convergent subnet ψi = ϕF (i), i ∈ I, of the sequence ϕn, n ∈ N. Let ψi → ϕ.
It follows from Lemma 25 that ϕ is a symmetric functional.
By the definition of a subnet (see [26, Section IV.2]), for every fixed n ∈ N, there
exists in ∈ I such that F (i) > n for every i > in. Thus, for every i > in, we have
ψi(µ(x)χ(0,1/n)) ≥ ϕF (i)(µ(x)χ(0,1/F (i))) ≥ lim
m→∞
1
m
kσm(µ(x))χ(0,1)kE.
The subnet ψi, in < i ∈ I converges to the same limit ϕ. Therefore,
ϕ(µ(x)χ(0,1/n)) ≥ lim
m→∞
1
m
kσm(µ(x))χ(0,1)kE.
Now, taking the limit as n → ∞, we obtain the inequality
ϕsing(x) ≥ lim
m→∞
1
m
kσm(µ(x))χ(0,1)kE,
where ϕsing is a singular symmetric functional defined in Lemma 26. The opposite
inequality is trivial.
(cid:3)
Theorem 29. Let E be a symmetric Banach space on the interval (0, 1). For a
given 0 ≤ x ∈ E, there exists a singular symmetric linear functional ϕsing such
that
ϕsing(x) = lim
m→∞
1
m
kσm(µ(x))kE.
TRACES ON SYMMETRICALLY NORMED OPERATOR IDEALS
15
Proof. Let F be a symmetric Banach space on the semi-axis with a norm given by
the formula
kxkF = kµ(x)χ(0,1)kE + kxk1,
∀x ∈ F.
Clearly, F ⊂ L1(0, ∞). Applying Theorem 28, we obtain a symmetric singular
functional ϕ on F such that
ϕ(x) = lim
m→∞
1
m
kσm(µ(x))χ(0,1)kF = lim
m→∞
1
m
kσm(µ(x))kE.
(cid:3)
5. Existence of fully symmetric functionals
In this section, we present results concerning existence of fully symmetric func-
tionals on fully symmetric function spaces. The main results of this section are
Theorem 33, Theorem 35 and Theorem 36.
Lemma 30. Let E be a symmetric Banach function space either on the interval
(0, 1) or on the semi-axis. If x, z ∈ DE are such that Cx ≤ Cz, then CMmx ≤
CMmz.
Proof. Let x = µ(a) − µ(b) and z = µ(c) − µ(d) with a, b, c, d ∈ E. It follows
from assumption Cx ≤ Cz that C(µ(a) + µ(d)) ≤ C(µ(b) + µ(c)) or, equivalently,
µ(a) + µ(d) ≺≺ µ(b) + µ(c).
Arguing as in Lemma 19, we have
with
It is now clear that
0
0
Z t
(Mmz)(s)ds =Z t
h(s, t) =
log(t/s)
log(m)
1,
,
z(s)h(s, t)ds
0 ≤ s ≤ t/m
t/m ≤ s ≤ t
0
Z t
Z t
0
Mm(µ(a) + µ(d))(s)ds =Z t
Mm(µ(b) + µ(c))(s)ds =Z t
0
0
(µ(s, a) + µ(s, d))h(s, t)ds,
(µ(s, b) + µ(s, c))h(s, t)ds.
Clearly, h is positive and decreasing with respect to s. It follows from [22, Equality
2.36] that
Mm(µ(a) + µ(d)) ≺≺ Mm(µ(b) + µ(c))
and the assertion follows.
(cid:3)
Lemma 31. Let E be a fully symmetric Banach function space either on the in-
terval (0, 1) or on the semi-axis and let x = µ(x) ∈ E. If z ∈ DE is such that
Cx ≤ Cz, then p(x) ≤ p(z).
Proof. Since Mmx is decreasing, it follows from Lemma 30 that
Z t
0
µ(s, Mmx)ds =Z t
0
(Mmx)(s)ds ≤Z t
0
(Mmz)+(s)ds ≤Z t
0
µ(s, (Mmz)+)ds.
Therefore, (Mmx)+ = Mmx ≺≺ (Mmz)+. The assertion follows now from the
definition of the functional p.
(cid:3)
16
F. SUKOCHEV AND D. ZANIN
Lemma 32. Let E be a fully symmetric Banach function space either on the in-
terval (0, 1) or on the semi-axis. Let p be the functional constructed in Lemma 22.
The functional
q(x) = inf{p(z) : z ∈ DE, Cx ≤ Cz},
x ∈ DE
satisfies the assumptions of Lemma 21.
Proof. It is clear from the definition of q that q ≤ p and that q is a positive
functional.
We claim that q is convex on DE. Let x1, x2 ∈ DE. Fix ε > 0 and select z1, z2 ∈
DE such that Cxi ≤ Czi and p(zi) ≤ q(xi) + ε for i = 1, 2. Thus, C(x1 + x2) ≤
C(z1 + z2) and
q(x1 + x2) ≤ p(z1 + z2) ≤ p(z1) + p(z2) ≤ q(x1) + q(x2) + 2ε.
Since ε is arbitrarily small, the claim follows.
We claim that q is monotone on DE. Let x1, x2 ∈ DE be such that x1 ≤ x2.
Fix ε > 0 and select z ∈ DE such that Cx2 ≤ Cz and p(z) ≤ q(x2) + ε. Thus,
Cx1 ≤ Cx2 ≤ Cz and q(x1) ≤ p(z) ≤ q(x2) + ε. Since ε is arbitrarily small, the
claim follows.
For x ∈ ZE ∩ DE, we have 0 ≤ q(x) ≤ p(x) = 0 and, therefore, q(x) = 0. s
(cid:3)
The following theorem is the first main result of this section.
Theorem 33. Let E = E(0, ∞) be a fully symmetric Banach space on the semi-
axis. For a given 0 ≤ x ∈ E, there exists a fully symmetric linear functional
ϕ : E → R such that
ϕ(x) = lim
m→∞
kσm(µ(x))kE.
1
m
Proof. Without loss of generality, x = µ(x). Let q be the convex monotone func-
tional constructed in Lemma 32.
It follows from Lemma 18 that there exist a
positive linear functional ϕ on E such that ϕ ≤ q and ϕ(x) = q(x).
It is clear that ϕ ≤ q ≤ p. Since p(z) = 0 for every z ∈ ZE, it follows that
ϕ(z) = 0 for every z ∈ ZE. Therefore, ϕ is a symmetric functional. For every
z ∈ DE with Cz ≤ 0, we have ϕ(z) ≤ q(z) ≤ p(0) = 0.
Let x1, x2 ∈ E be positive elements such that x1 ≺≺ x2. Therefore, z = µ(x1) −
µ(x2) ∈ DE and Cz ≤ 0. It follows from above that ϕ(z) ≤ 0. Hence, ϕ is a fully
symmetric functional.
Since ϕ(z) ≤ q(z) ≤ p(z) ≤ kzkE for every z = µ(z) ∈ E, it follows that
kϕkE ∗ ≤ 1. Therefore,
ϕ(x) = ϕ(
1
m
σmx) ≤
1
m
kσmxkE.
Passing m → ∞, we obtain
ϕ(x) ≤ lim
m→∞
1
m
kσmµ(x)kE.
On the other hand, q(x) = p(x) by Lemma 31. By Lemma 19, we have m−1σmx ⊳
Mmx. Therefore,
ϕ(x) = q(x) = p(x) = lim sup
m→∞
kMmxkE ≥ lim
m→∞
1
m
kσmµ(x)kE.
The assertion follows immediately.
(cid:3)
TRACES ON SYMMETRICALLY NORMED OPERATOR IDEALS
17
If π : E → E is a convex functional defined in (9), then π(−x) = π(x) for every
x ∈ E. If q is a functional defined in Lemma 32, then q(−x) = 0 for positive x ∈ E.
Therefore, q 6= π. However, the assertion below follows from Theorem 33.
Lemma 34. Let E = E(0, ∞) be a fully symmetric Banach space on the semi-
axis. Let q and π be the convex functionals on E defined in Lemma 32 and (9),
respectively. For every positive x ∈ E, we have q(x) = π(x).
Proof. For every x ∈ E, consider the functional ϕ constructed in Theorem 33. By
construction, we have ϕ(x) = q(x) = p(x) = π(x).
(cid:3)
The proofs of the two following theorems are very similar to that of Theorem 28
(respectively, Theorem 29) and are, therefore, omitted. The only difference is that
the reference to Theorem 23 (respectively, Theorem 28) has to be replaced with the
reference to Theorem 33 (respectively, Theorem 35).
Theorem 35. Let E ⊂ L1(0, ∞) be a fully symmetric Banach space on the semi-
axis. For a given 0 ≤ x ∈ E, there exists a singular fully symmetric linear functional
ϕsing such that
ϕsing(x) = lim
m→∞
1
m
kσm(µ(x))χ(0,1)kE.
Theorem 36. Let E ⊂ L1(0, 1) be a fully symmetric Banach space on the inter-
val (0, 1). For a given 0 ≤ x ∈ E, there exists a singular fully symmetric linear
functional ϕsing such that
ϕsing(x) = lim
m→∞
1
m
kσm(µ(x))kE.
6. The sets of symmetric and fully symmetric functionals are
different
In this section, we demonstrate that the sets of symmetric and fully symmetric
functionals on a given fully symmetric space E are distinct (provided that one of
these sets is non-empty). The main results are Theorem 47 and Theorem 48.
Let x = µ(x) ∈ (L1 + L∞)(0, ∞) (or x = µ(x) ∈ L1(0, 1)) and let X(t) =
R t
0 x(s)ds. For every θ > 0, let an(θ) be such that
X(an(θ)) = (3/2)nθ
for every n ∈ Z such that an(θ) does exist. Given a sequence κ = {κn}n∈Z ∈
(N ∪ {∞})Z, let
Bκ,θ = {κna3n(θ), where n ∈ Z is such that κ2
na3n(θ) < a3n+1(θ)}.
If κn = m for all n ∈ N, we write Bm,θ instead of Bκ,θ. Also, set
Am = {man(1) : m2an(1) < an+1(1), n ∈ Z}.
Lemma 37. If x = µ(x) ∈ L1 + L∞ and if Ci, 1 ≤ i ≤ k, are discrete sets, then
E(x ∪k
i=1 Ci) ≺≺
E(xCi).
k
Xi=1
18
F. SUKOCHEV AND D. ZANIN
Proof. It is sufficient to verify
Z t
0
E(x ∪k
i=1 Ci)(s)ds ≤
only at the nodes of E(x ∪k
However, if t ∈ Ci for some i, then
k
Xi=1Z t
0
E(xCi)(s)ds
i=1 Ci), that is at the nodes of E(xCi) for every i.
Z t
0
and we are done.
E(x ∪k
i=1 Ci)(s)ds = X(t) =Z t
0
E(xCi)(s)ds
(cid:3)
We will need the following lemma.
Lemma 38. If x = µ(x) ∈ L1 + L∞ and if κ ≥ κ′ (that is κn ≥ κ′
then
n for every n),
(12)
E(xBκ,θ) ≺≺
3
2
E(xBκ′,θ).
Proof. Let n ∈ Z be such that κ2
a3n+1(θ). Therefore,
na3n(θ) < a3n+1(θ). It follows that κ′2
n a3n(θ) <
Z κna3n(θ)
0
0
na3n(θ)
E(xBκ,θ)(s)ds ≤Z a3n+1(θ)
x(s)ds = 3/2Z κ′
E(xBκ,θ)(s)ds ≤ 3/2Z t
≤ 3/2Z κ′
Z t
0
0
0
0
Hence, we have
(13)
x(s)ds = 3/2Z a3n(θ)
0
na3n(θ)
x(s)ds ≤
E(xBκ′,θ)(s)ds.
E(xBκ′,θ)(s)ds
for every t being a node of the partition Bκ,θ. Thus, (13) holds for every t > 0. (cid:3)
Remark 39. The inequality (12) holds if κn ≥ κ′
the inequality κ2
na3n(θ) < a3n+1(θ).
n only for such n ∈ Z that satisfy
Lemma 40. Let E be a fully symmetric Banach function space either on the in-
terval (0, 1) or on the semi-axis. Let x = µ(x) ∈ E and y = µ(y) ∈ E be such
that ϕ(y) ≤ ϕ(x) for every positive symmetric functional ϕ ∈ E∗. There exists
0 ≤ um ∈ E such that um → 0 in E and
Z b
ma
y(s)ds ≤Z mb
a
(x + um)(s)ds,
∀ma ≤ b.
Proof. Let p be a convex positive functional considered in Lemma 22. By Lemma
18, there exists a positive functional ϕ ∈ E∗ such that ϕ ≤ p and ϕ(y−x) = p(y−x).
We have p(z) = 0 for every z ∈ ZE and, therefore, ϕ(z) = 0 for every z ∈ ZE.
Therefore, ϕ is a positive symmetric linear functional on E.
By the assumption, ϕ(y − x) ≤ 0 and, therefore, p(y − x) = 0. Hence, by the
definition of p, we have um = (Mm(y − x))+ → 0 in E. Clearly, Mmy ≤ Mmx + um.
It follows from Lemma 19 that
Z b
ma
y(s)ds ≤Z mb
ma
(Mmy)(s)ds ≤Z mb
ma
(Mmx + um)(s)ds ≤Z mb
a
(x + um)(s)ds.
(cid:3)
TRACES ON SYMMETRICALLY NORMED OPERATOR IDEALS
19
For each sequence κ and λ > 0, we define the sequence κλ by setting
κλ
n =(κn,
∞,
κn ≥ λ
κn < λ.
Lemma 41. If m ∈ N, x = µ(x) ∈ L1 + L∞ and 0 ≤ u ∈ L1 + L∞ are such that
Z b
ma
E(xBκ,θ)(s)ds ≤Z mb
a
(x + u)(s)ds,
∀ma ≤ b ∈ R,
m−1σmE(xBκ100m,θ) ≺≺ 30µ(u).
then
(14)
Proof. If κ100m
trivial.
n
= ∞ for every n ∈ Z, then E(xBκ100m,θ) = 0 and the assertion is
Let n ∈ Z be such that κ2
na3n(θ) < a3n+1(θ) and κn ≥ 100m. It follows that
(15)
Z mκna3n(θ)
0
u(s)ds ≥Z mκna3n(θ)
a3n(θ)
(x + u)(s)ds −Z mκna3n(θ)
a3n(θ)
x(s)ds.
By the assumption, we have
(16)
Z mκna3n(θ)
a3n(θ)
(x + u)(s)ds ≥Z κna3n(θ)
ma3n(θ)
E(xBκ,θ)(s)ds.
Note that mκna3n(θ) < a3n+1(θ). It follows from (15) and (16) that
(17)
Z mκna3n(θ)
0
u(s)ds ≥Z κna3n(θ)
ma3n(θ)
E(xBκ,θ)(s)ds −Z a3n+1(θ)
a3n(θ)
x(s)ds.
Let n′ be the maximal integer number such that n′ < n and κ2
It is clear that
n′ a3n′(θ) < a3n′+1(θ).
κ2
n′a3n′ (θ) < a3n′+1(θ) ≤ a3n−2(θ) < ma3n(θ)
and
(18) E(xBκ,θ) =
X(κna3n(θ)) − X(κn′a3n′ (θ))
κna3n(θ) − κn′ a3n′(θ)
≥
X(a3n(θ)) − X(a3n−2(θ))
κna3n(θ)
on the interval (ma3n(θ), κna3n(θ)).
If κ2
n′a3n′ (θ) ≥ a3n+1(θ) for every n′ < n, then
X(κna3n(θ))
(19)
E(xBκ,θ) =
κna3n(θ)
≥
X(a3n(θ))
κna3n(θ)
on the interval (ma3n(θ), κna3n(θ)).
It follows from (17) and (18) (or (19)) that
Z mκna3n(θ)
0
u(s)ds ≥
κn − m
κn
· (1 −
4
9
)X(a3n(θ)) −
1
2
X(a3n(θ)).
Since κn ≥ 100m, it follows that
Z mκna3n(θ)
0
u(s)ds ≥ ((1 −
1
100
)(1 −
4
9
) −
1
2
)X(a3n(θ)) =
1
20
X(a3n(θ)) =
=
1
30
X(a3n+1(θ)) ≥
1
30
X(κna3n(θ)) =
1
30Z κna3n(θ)
0
E(xBκ100m,θ)(s)ds.
for every t being a node of the partition Bκ100m,θ. Therefore,
u(s)ds ≤ 30Z mt
0
µ(s, u)ds
20
F. SUKOCHEV AND D. ZANIN
It follows immediately that
(20)
Z t
0
or, equivalently,
0
E(xBκ100m,θ)(s)ds ≤ 30Z mt
Z t
Z t/m
E(xBκ100m,θ)(s)ds ≤ 30Z mt
E(xBκ100m,θ)(s)ds ≤ 30Z t
0
0
0
0
µ(s, u)ds,
t > 0
µ(s, u)ds,
t > 0.
(cid:3)
The assertion follows immediately.
Lemma 42. Let E be a fully symmetric Banach function space either on the in-
terval (0, 1) or on the semi-axis.
If x = µ(x) ∈ E is such that ϕ(y) ≤ ϕ(x)
for every positive symmetric functional ϕ on E and every 0 ≤ y ≺≺ x, then
λ−1σλE(xBκλ,θ) → 0 as λ → ∞.
Proof. Since E(xBκ,θ) ≺≺ x, it follows from the assumption and Lemma 40 that
there exists 0 ≤ um → 0 such that
Z b
ma
E(xBκ,θ)(s)ds ≤Z mb
a
(x + um)(s)ds,
∀ma ≤ b ∈ R.
For every λ ≥ 100m, we have κ100m ≤ κλ. It follows from Lemma 41 that
1
λ
σλE(xBκλ,θ) ≺≺
1
m
σmE(xBκλ,θ)
Lemma38
≺≺
3
2m
σmE(xBκ100m,θ)
(14)
≺≺ 45µ(um).
The assertion now follows immediately.
(cid:3)
Proposition 43. Let E be a fully symmetric Banach function space either on the
interval (0, 1) or on the semi-axis equipped with a Fatou norm. If x = µ(x) ∈ E is
such that ϕ(y) ≤ ϕ(x) for every positive symmetric functional ϕ on E and every
0 ≤ y ≺≺ x, then m−1σmE(xBm,θ) → 0 as m → ∞.
Proof. For every m, r ∈ N, set
κm,r
n =(m
0 ≤ n < r
∞ r ≤ n
and κm,r = {κm,r
r → ∞. It follows from the definition of Fatou norm that
n }n∈Z. Clearly, E(xBκm,r,θ) → E(xBm,θ) almost everywhere when
lim
r→∞
kσmE(xBκm,r,θ)kE = kσmE(xBm,θ)kE.
Select rm so large that
(21)
1
m
kσmE(xBκm,rm ,θ)kE >
1
2m
kσmE(xBm,θ)kE.
Now define the sequence κ = {κn}n∈Z by setting
κn = inf
m≥1
κm,rm
n
= inf
rm>n
m, n ∈ Z.
TRACES ON SYMMETRICALLY NORMED OPERATOR IDEALS
21
Clearly, rκn ≥ n and, therefore, κn → ∞ as n → ∞. In particular, the set
{n : κn < λ} is finite for every λ ∈ N. Set
M (λ) = max{λ, max
(
κn<λ
a3n+1(θ)
a3n(θ)
)1/2}.
If m > M (λ), then m2a3n(θ) ≥ a3n+1(θ) whenever κn < λ. Thus, κn ≥ λ when-
ever m2a3n(θ) < a3n+1(θ). Hence, κλ
)2a3n(θ) < a3n+1(θ).
)2a3n(θ) < a3n+1(θ).
Therefore, κλ
According to Remark 39, it follows that
n = κn whenever (κm,rm
for every n ∈ Z such that (κm,rm
n ≤ κm,rm
n
n
n
E(xBκm,rm ,θ) ≺≺
3
2
E(xBκλ,θ).
Since m ≥ λ, it follows that
(22)
1
m
σmE(xBκm,rm ,θ) ≺≺
3
2λ
σλE(xBκλ,θ).
By Lemma 42, for every ε > 0, there exists λ such that
(23)
It follows that
1
λ
kσλE(xBκλ,θ)kE <
1
3
ε.
1
m
kσmE(xBm,θ)kE
(21)
≤
2
m
kσmE(xBκm,rm ,θ)kE
(22)
≤
3
λ
kσλE(xBκλ,θ)kE
(23)
< ε
for every m > M (λ). Since ε > 0 is arbitrarily small, the assertion follows.
(cid:3)
Lemma 44. Let E be a fully symmetric Banach space either on the interval (0, 1)
or on the semi-axis equipped with a Fatou norm. If x = µ(x) ∈ E is such that
ϕ(y) ≤ ϕ(x) for every positive symmetric functional ϕ on E and every 0 ≤ y ≺≺ x,
then m−1σmE(xAm) → 0 as m → ∞.
Proof. It is clear that ak(3/2) = ak+1(1) and ak((3/2)2) = ak+2(1) for every k ∈ N.
It follows that
Therefore, by Lemma 37, we have
Bm,1 ∪ Bm,3/2 ∪ Bm,(3/2)2 = Am.
(24)
E(xAm) ≺≺ E(xBm,1) + E(xBm,3/2) + E(xBm,(3/2)2).
The assertion follows now from Proposition 43.
(cid:3)
Lemma 45. Let x = µ(x) ∈ L1 + L∞(0, ∞) be a function on the semi-axis. If
x /∈ L1(0, ∞), then, for every t > 0 and every m ∈ N, we have
(25)
X(t) ≤
2
3
X(m4t) +
3
2Z m4t
0
E(xAm)(s)ds.
Proof. For a given t > 0, there exists n ∈ Z such that t ∈ [an(1), an+1(1)]. If
an+1(1) > m2an(1), then
If an+1(1) ≤ m2an(1) and an+2(1) > m2an+1(1), then
0
Z m4t
Z m4t
E(xAm)(s)ds ≥Z man(1)
E(xAm)(s)ds ≥Z man+1(1)
0
0
0
E(xAm)(s)ds = X(man(1)) ≥
2
3
X(t).
E(xAm)(s)ds = X(man+1(1)) ≥ X(t).
22
F. SUKOCHEV AND D. ZANIN
If an+2(1) ≤ m2an+1(1) and an+1(1) ≤ m2an(1), then
X(m4t) ≥ X(an+2(1)) =
3
2
X(an+1(1)) ≥
3
2
X(t)
and the assertion follows.
(cid:3)
The situation in the case that x ∈ L1 is slightly more complicated.
Lemma 46. If x = µ(x) ∈ L1(0, 1) or x ∈ L1(0, ∞), then there exists constant C
such that for every t > 0
(26)
X(t) ≤
2
3
X(m4t) +
3
2Z m4t
0
E(xAm)(s)ds + CZ m4t
0
χ[0,1](s)ds.
Proof. Consider first the case of the semi-axis. Fix n0 such that X(an0) ≤ 4/9X(∞).
For a givne t ∈ [a, an0], there exists n ∈ Z such that n < n0 and t ∈ [an, an+1].
Then, the argument in Lemma 45 applies mutatis mutandi. For every t ≥ an0 we
have
X(t) ≤
X(∞)
min{an0, 1}
min{m4t, 1} =
X(∞)
min{an0, 1}Z m4t
0
χ[0,1](s)ds.
Setting C = X(∞)/ min{an0, 1}, we obtain the assertion.
The same argument applies in the case of the interval (0, 1) by replacing X(∞)
(cid:3)
by X(1).
The following two theorems are crucial for the proof of the implication (3) ⇔ (4)
in Theorem 5.
Theorem 47. Let E be a fully symmetric Banach space either on the interval (0, 1)
or on the semi-axis and let x ∈ E. Suppose that the norm on E is a Fatou norm. If
ϕ(y) ≤ ϕ(x) for every positive symmetric functional on E and every 0 ≤ y ≺≺ x,
then
(27)
lim
m→∞
1
m
kσm(µ(x))kE = 0
provided that one of the following conditions is satisfied
(1) E = E(0, 1) is a space on the interval (0, 1).
(2) E = E(0, ∞) is a space on the semi-axis and E(0, ∞) 6⊂ L1(0, ∞).
Proof. Without loss of generality, x = µ(x). If x /∈ L1, then by Lemma 45,
Z t/m4
0
or, equivalently,
3
2Z t
0
E(xAm)(s)ds,
∀t > 0
x(s)ds ≤
x(s)ds +
0
2
3Z t
1
m4 σm4 x ≺≺
3
2
Applying m−1σm to the both parts, we obtain
3
2
1
m5 σm5 x ≺≺
σmx +
1
m
x +
2
3
2
3
E(xAm).
1
m
σmE(xAm).
Take norms and let m → ∞. It follows from Lemma 44 that
lim
m→∞
1
m
kσmxkE ≤
2
3
lim
m→∞
1
m
kσmxkE.
This proves (27).
TRACES ON SYMMETRICALLY NORMED OPERATOR IDEALS
23
If x ∈ L1 and C are as in Lemma 46, then it follows from Lemma 46 that
Z t/m4
0
x(s)ds ≤
or, equivalently,
2
3Z t
0
x(s)ds +
3
2Z t
0
E(xAm)(s)ds + CZ t
0
χ[0,1](s)ds,
∀t > 0
1
m4 σm4 x ≺≺
2
3
x +
3
2
E(xAm) + Cχ(0,1).
Applying m−1σm to the both parts, we obtain
1
m5 σm5 x ≺≺
2
3
1
m
σmx +
3
2
1
m
σmE(xAm) + C
1
m
σmχ(0,1).
Take norms and let m → ∞. For every symmetric space E on the interval (0, 1)
and for every symmetric space E on the semi-axis such that E 6⊂ L1(0, ∞) we have
m−1kσmχ(0,1)kE → 0. It follows from Lemma 44 that
lim
m→∞
1
m
kσmxkE ≤
2
3
lim
m→∞
1
m
kσmxkE
and again (27) follows.
(cid:3)
Theorem 48. Let E = E(0, ∞) be a fully symmetric Banach space on the semi-
axis equipped with a Fatou norm such that E(0, ∞) ⊂ L1(0, ∞). If ϕ(y) ≤ ϕ(x) for
every positive symmetric functional on E and every 0 ≤ y ≺≺ x, then
(28)
lim
m→∞
1
m
kσm(µ(x))χ(0,1)kE = 0.
Proof. Fully symmetric Banach space F on the interval (0, 1) consists of those z ∈ E
supported on the interval (0, 1). Let x1 = µ(x)χ(0,1) ∈ F. Suppose that
It clearly follows that
lim
m→∞
1
m
kσm(µ(x))χ(0,1)kE > 0.
lim
m→∞
1
m
kσm(µ(x1))kF > 0.
By Theorem 47, there exists 0 ≤ y1 ≺≺ x1 and a positive symmetric functional
ϕ ∈ F ∗ such that ϕ(y1) > ϕ(x1). Let ϕsing be a singular part of the functional ϕ
constructed in Lemma 26. It follows from Lemma 26 that ϕsing is symmetric. By
Lemma 27, the difference ϕ − ϕsing is a symmetric normal functional on F (that
is, an integral). Therefore, ϕsing(y1) > ϕsing(x1).
Now we show that the functional ϕsing can be extended from F to E by setting
ϕsing(z) = lim
n→∞
ϕsing(µ(z)χ(0,1/n)),
0 ≤ z ∈ E.
Repeating the argument in Lemma 26, we prove that the extension above is additive
on E+. Thus, the functional ϕsing ∈ E∗ is positive and symmetric. Since y1 ≺≺ x
and ϕsing(y1) > ϕsing(x1) = ϕsing(x), the assertion follows.
(cid:3)
24
F. SUKOCHEV AND D. ZANIN
7. Proof of Theorem 5
In this section, we prove an assertion more general then that of Theorem 5. The
assertion of Theorem 5 follows from that of Theorem 49 by setting M = B(H).
In what follows, the semifinite von Neumann algebra M is either atomless or
atomic so that the trace of every atom is 1.
Theorem 49. Let E(M, τ ) be a symmetric operator space. Consider the following
conditions.
(1) There exist nontrivial positive singular symmetric functionals on E(M, τ ).
(2) There exist nontrivial singular fully symmetric functionals on E(M, τ ).
(3) There exist positive symmetric symmetric functional on E(M, τ ) which are
not fully symmetric.
(4) If E(M, τ ) 6⊂ L1(M, τ ), then there exists an operator A ∈ E(M, τ ) such
that
(29)
(30)
lim
m→∞
1
m
kσmµ(A)kE > 0.
If E(M, τ ) ⊂ L1(M, τ ), then there exists an operator A ∈ E(M, τ ) such
that
lim
m→∞
1
m
k(σmµ(A))χ(0,1)kE > 0.
(i) The conditions (1) and (4) are equivalent for every symmetric operator space
E(M, τ ).
(ii) The conditions (1), (2) and (4) are equivalent for every fully symmetric op-
erator space E(M, τ ).
(iii) The conditions (1)-(4) are equivalent for every fully symmetric operator space
E(M, τ ) equipped with a Fatou norm.
Proof. Implications (2) ⇒ (1) and (3) ⇒ (1) are trivial.
(1) ⇒ (4) Let E(M, τ ) be a symmetric operator space with a singular symmetric
functional ϕ. Let A ∈ E(M, τ ) be an operator such that ϕ(A) 6= 0. Without loss
of generality, A ≥ 0.
If E(M, τ ) 6⊂ L1(M, τ ), then
ϕ(A) =
1
m
ϕ(A ⊕ · · · ⊕ A
) ≤ kϕkE ∗(M,τ ) ·
1
m
kσmµ(A)kE.
Passing m → ∞, we obtain the required inequality (29).
Let now E(M, τ ) ⊂ L1(M, τ ). If M is atomic, then E(M, τ ) = L1(M, τ ) and
the assertion is trivial. Let M be atomless. Since ϕ is a singular functional and
A − AEA(µ(
1
m
, A), ∞) ∈ (L1 ∩ L∞)(M, τ ),
∀m ∈ N,
m times
{z
}
we infer that
=
1
m
≤ kϕkE ∗(M,τ )·
1
m
ϕ(A) = ϕ(AEA(µ(
1
m
, A), ∞)) =
ϕ(AEA(µ(
, A), ∞) ⊕ · · · ⊕ AEA(µ(
, A), ∞)
) ≤
1
m
1
m
1
m
m times
{z
kσmµ(AEA(µ(
, A), ∞))kE ≤ kϕkE ∗(M,τ )·
}
k(σmµ(A))χ(0,1)kE.
1
m
TRACES ON SYMMETRICALLY NORMED OPERATOR IDEALS
25
Passing m → ∞, we obtain the required inequality (30).
(4) ⇒ (1) Firstly, we assume that the algebra M is finite. Without loss of
generality, τ (1) = 1. Let E(M, τ ) be a symmetric operator space and let E(0, 1)
be the corresponding symmetric function space. By the assumption, there exists
an element x = µ(A) ∈ E(0, 1) such that m−1σmx 6→ 0 in E(0, 1). By Theorem 29,
there exists a positive singular symmetric functional 0 6= ϕ ∈ E(0, 1)∗. Let L(ϕ)
be a functional on E(M, τ ) defined in Theorem 14. Clearly, L(ϕ) is a nontrivial
positive symmetric functional on E(M, τ ).
The case when M is an infinite atomless von Neumann algebra can be treated
in a similar manner. The only difference is that the reference to Theorem 29 has
to be replaced with the reference to either Theorem 28 or Theorem 23.
Let E(M, τ ) be a symmetric operator space on a atomic von Neumann algebra
M and let E(N) be the corresponding symmetric sequence space. It follows from
the assumption that E(M, τ ) 6= L1(M, τ ) or, equivalently, E(N) 6= l1. By the
assumption, there exists an element x = µ(A) ∈ E such that m−1σmx 6→ 0 in E.
Let F (0, ∞) be a symmetric function space constructed in Proposition 16. Since
E(N) 6= l1, it follows that F (0, ∞) 6⊂ L1(0, ∞). Recall that the space E(N) is
naturally embedded into the space F (0, ∞) and that the norms k · kE and k · kF are
equivalent on E(N). We have x ∈ F and m−1σmx 6→ 0 in F (0, ∞). By Theorem
23, there exists a positive symmetric functional 0 ≤ ϕ ∈ F (0, ∞)∗. The restriction
of the functional ϕ to E(N) is a nontrivial positive symmetric functional on E(N).
Let L(ϕ) be a functional on E(M, τ ) defined in Theorem 14. Clearly, L(ϕ) is a
nontrivial positive symmetric functional on E(M, τ ).
(4) ⇒ (2) The proof is very similar to that of the implication (4) ⇒ (1) and is,
therefore, omitted. The only difference is that references to Theorem 29, Theorem
28 or Theorem 23 have to be replaced with references to Theorem 36, Theorem 35
or Theorem 33, respectively.
(4) ⇒ (3) Firstly, we assume that the algebra M is finite. Without loss of
generality, τ (1) = 1. Let E(M, τ ) be a symmetric operator space and let E(0, 1)
be the corresponding symmetric function space. By the assumption, there exists
an element x = µ(A) ∈ E(0, 1) such that m−1σmx 6→ 0 in E(0, 1). By Theorem 47,
there exists a positive symmetric but not fully symmetric functional ϕ ∈ E(0, 1)∗.
Let L(ϕ) be a functional on E(M, τ ) defined in Theorem 14. Clearly, L(ϕ) is a
symmetric but not fully symmetric functional on E(M, τ ).
The case when M is an infinite atomless von Neumann algebra can be treated
in a similar manner. The only difference is that the reference to Theorem 47 has
to be replaced with the reference to either Theorem 47 or Theorem 48.
Let E(M, τ ) be a symmetric operator space on a atomic von Neumann algebra
M and let E(N) be the corresponding symmetric sequence space. It follows from
the assumption that E(M, τ ) 6= L1(M, τ ) or, equivalently, E(N) 6= l1. By the
assumption, there exists an element x = µ(A) ∈ E such that m−1σmx 6→ 0 in E.
Let F (0, ∞) be a symmetric function space constructed in Proposition 16. Since
E(N) 6= l1, it follows that F (0, ∞) 6⊂ L1(0, ∞). Recall that the space E(N) is
naturally embedded into the space F (0, ∞) and that the norms k · kE and k · kF are
equivalent on E(N). We have x ∈ F and m−1σmx 6→ 0 in F (0, ∞). By Theorem 47,
there exists a positive symmetric functional ϕ ∈ F (0, ∞)∗ and a function 0 ≤ y ≺≺
x such that ϕ(y) > ϕ(x). Set z = E(µ(y){(n − 1, n)}n∈N). Clearly, z ∈ E(N) and
ϕ(z) = ϕ(y) > ϕ(x). Hence, the restriction of the functional ϕ to E(N) is a positive
26
F. SUKOCHEV AND D. ZANIN
symmetric but not fully symmetric functional on E(N). Let L(ϕ) be a functional
on E(M, τ ) defined in Theorem 14. Clearly, L(ϕ) is a positive symmetric but not
fully symmetric functional on E(M, τ ).
(cid:3)
In this appendix, we set A = {(n − 1, n)}n∈N.
8. Appendix
Lemma 50. If x, y ∈ (L1 + L∞)(0, ∞) are positive functions, then
E(µ(x + y)A) ⊳ E(µ(x)A) + E(µ(y)A) ⊳ 2σ1/2E(µ(x + y)A).
Proof. Recall that
µ(x + y) ≺≺ µ(x) + µ(y) ≺≺ 2σ1/2µ(x + y).
It follows that
Let now a, b be positive integers. Subtracting the above inequalities, we obtain
0
0
0
0
Z b
µ(s, x + y)ds ≤Z b
µ(s, x + y)ds ≥Z a
Z 2a
E(µ(x + y)A)(s)ds =Z b
Z b
(µ(s, x) + µ(s, y))ds =Z b
≤Z b
E(µ(x) + µ(y)A)(s)ds ≤Z 2b
Z b
2a
2a
2a
a
(µ(s, x) + µ(s, y))ds,
(µ(s, x) + µ(s, y))ds.
µ(s, x + y)ds ≤
2a
E(µ(x) + µ(y)A)(s)ds.
E(µ(x + y)A)(s)ds.
a
Similarly, we have
(cid:3)
(cid:3)
Corollary 51. The quasi-norm in Construction 16 is a norm.
Proof. It follows from Lemma 50 that
provided that x, y are positive functions. By Theorem 9,
E(µ(x + y)A) ⊳ E(µ(x) + µ(y)A)
kE(µ(x + y)A)kE ≤ kE(µ(x)A)kE + kE(µ(y)A)kE.
Lemma 52. Let y = µ(y) ∈ (L1 + L∞)(0, ∞). It follows that
Z b
2−kλa
λ
λ − 1Z b
a
y(s)ds ≤
(σ2k y)(s)ds
provided that b ≥ λa.
Proof. Let α be the average value of y on the interval [2−kλa, 2−kb]. Clearly, y ≤ α
on the interval [2−kλa, b] and y ≥ α on the interval [2−ka, 2−kb]. Thus, σ2k y ≥ α
on the interval [a, b]. Therefore,
Z b
2−kλa
y(s)ds ≤ (b − 2−kλa)α ≤
λ
λ − 1
(b − a)α ≤
λ
λ − 1Z b
a
(σ2k y)(s)ds.
(cid:3)
TRACES ON SYMMETRICALLY NORMED OPERATOR IDEALS
27
Theorem 53. If {xn}n∈N be a Cauchy sequence in F, then there exists x ∈ F such
that xn → x in F.
Proof. For every k > 0, there exists mk such that kxm − xmk kF ≤ 4−k for m ≥ mk.
Set yk = xmk+1 − xmk . Clearly, kykkF ≤ 4−k for every k ∈ N. In particular, the
seriesP∞
Set zn =P∞
k=1 yk converges in L∞(0, ∞).
k=n σ2k µ(yk). We claim that zn ∈ F and zn → 0 in F. Indeed,
µ(yk) ≤ kykk∞χ(0,1) + T E(µ(yk)A).
Here, T is a shift to the right. It follows that
E(µ(zn)A) ≤
∞
Xk=n
σ2k (kykk∞χ(0,1) + T E(µ(yk)A)).
Therefore,
kznkF ≤ kznk∞ +
∞
Xk=n
2kkkykk∞χ(0,1) + T E(µ(yk)A)kE ≤
≤ kznk∞ +
2k+1kykkF ≤
1
3
∞
Xk=n
· 41−n + 22−n = o(1).
It follows from Lemma 8.5 of [19] that
Z b
λa
µ(s,
∞
Xk=n
yk)ds ≤
∞
Xk=nZ b
2−kλa
µ(s, yk)ds.
It follows from Lemma 52 that
Z b
2−kλa
µ(s, yk)ds ≤
λ
λ − 1Z b
a
(σ2k µ(yk))(s)ds.
Therefore,
Hence,
Z b
λa
µ(s,
∞
Xk=n
yk)ds ≤
λ
λ − 1Z b
a
zn(s)ds.
yk ⊳
λ
λ − 1
zn.
∞
Xk=n
Since λ > 1 is arbitrarily large, it follows from Theorem 9 that
k
∞
Xk=n
ykkF ≤ kznkF → 0.
Thus, the seriesP∞
k=1 yk does converge in F. The assertion follows immediately.
(cid:3)
28
F. SUKOCHEV AND D. ZANIN
9. Proof of Figiel-Kalton theorem
The proof of Theorem 8 follows from the combinations of Lemmas below.
Lemma 54. Let E be a symmetric Banach space either on the interval (0, 1) or
on the semi-axis. If x ∈ ZE, then C(µ(x+) − µ(x−)) ∈ E.
k=1(xk − yk) with xk, yk ∈ E+ and µ(xk) = µ(yk), 1 ≤ k ≤ n.
Proof. Let x = Pn
Set
z = x+ +
yk = x− +
n
Xk=1
xk.
n
Xk=1
It follows from the definition of C and (8) that
Cµ(z) ≤ C(x+) +
Cµ(yk) = C(µ(x+) − µ(x−)) + Cµ(x−) +
Using the second inequality in (8), we obtain
Cµ(xk).
n
Xk=1
n
Xk=1
µ(s, xk))ds ≤Z (n+1)t
0
Z t
0
(µ(s, x−) +
Therefore,
n
Xk=1
µ(s, z)ds ≤Z t
0
µ(s, z)ds + ntµ(t, z).
Cµ(z) ≤ Cµ(z) + C(µ(x+) − µ(x−)) + nµ(z).
It follows that C(µ(x−) − µ(x+)) ≤ nµ(z). Similarly, C(µ(x+) − µ(x−)) ≤ nµ(z)
and the assertion follows.
(cid:3)
Lemma 55. Let E be a symmetric Banach space either on the interval (0, 1) or
on the semi-axis. If x ∈ DE, then C(µ(x+) − µ(x−)) ∈ Cx + E.
Proof. Since x ∈ DE, it follows that x = µ(a) − µ(b) with a, b ∈ E. Set u =
µ(a) − x+ ≥ 0. Clearly, µ(a) = u + x+ and µ(b) = u + x−. It follows from the
definition of C and (8) that
Cµ(a) ≤ Cµ(u) + Cµ(x+) = C(µ(x+) − µ(x−)) + Cµ(u) + Cµ(x−).
Using the second inequality in (8), we obtain
It follows that
Similarly,
Cµ(x−) + Cµ(u) ≤ Cµ(b) + µ(b).
Cx ≤ C(µ(x+) − µ(x−)) + µ(b).
Cx ≥ C(µ(x+) − µ(x−)) − µ(a)
and the assertion follows.
(cid:3)
Lemma 56. Let E = E(0, ∞) be a symmetric space on the semi-axis. If x ∈ DE
is such that Cx ∈ E, then x ∈ ZE.
Proof. Define a partition A = {(2n, 2n+1)}n∈Z and set x1 = E(xA). If x = µ(a) −
µ(b) with a, b ∈ E, then x1 = E(µ(a)A) − E(µ(b)A). Clearly,
E(µ(a)A) ≤ σ2µ(a) ∈ E, E(µ(b)A) ≤ σ2µ(b) ∈ E
are decreasing functions. It follows that x1 ∈ DE. It is easy to see that
Cx1 − Cx ≤ 2σ2(µ(a) + µ(b)).
TRACES ON SYMMETRICALLY NORMED OPERATOR IDEALS
29
Therefore, Cx1 ∈ E. Define a function z ∈ E by setting
z(t) = (Cx1)(2n+1),
t ∈ (2n, 2n+1).
Clearly, x1 = 2z − σ2z ∈ ZE.
Consider the function x − x1 on the interval (2n, 2n+1). By Kwapien theorem
[23], there exist positive equimeasurable functions y1n, y2n supported on (2n, 2n+1)
such that
µ(y1n) = µ(y2n),
ky1nk∞, ky2nk∞ ≤ 6k(x − x1)χ(2n,2n+1)k∞.
Set y1 =Pn∈N y1n and y2n =Pn∈N y2n. It follows that y1, y2 ∈ E+. Since x − x1 =
y1 − y2 and µ(y1) = µ(y2), it follows that x − x1 ∈ ZE. The assertion follows
immediately.
(cid:3)
References
[1] S. Astashkin, F. Sukochev, Banach-Saks property in Marcinkiewicz spaces. J. Math. Anal.
Appl. 336 (2007) 1231 -- 1258.
[2] M. Braverman, A. Mekler, The Hardy-Littlewood property for symmetric spaces, Siberian
Math. J. 18 (1977), 371 -- 385.
[3] A. Carey, F. Sukochev, Dixmier traces and some applications to noncommutative geometry.
(Russian) Uspekhi Mat. Nauk 61 (2006), no. 6 (372), 45 -- 110; translation in Russian Math.
Surveys 61 (2006), no. 6, 1039 -- 1099.
[4] A. Connes, Noncommutative Geometry Academic Press, San Diego, 1994.
[5] A. Connes, The action functional in noncommutative geometry. Comm. Math. Phys. 117
(1988), no. 4, 673 -- 683.
[6] J. Dixmier, Existence de traces non normales (French) C. R. Acad. Sci. Paris Ser. A-B 262
1966 A1107 -- A1108.
[7] P. Dodds, B. de Pagter, A. Sedaev, E. Semenov, F. Sukochev, Singular symmetric func-
tionals and Banach limits with additional invariance properties. (Russian) Izv. Ross. Akad.
Nauk Ser. Mat. 67 (2003), no. 6, 111 -- 136; translation in Izv. Math. 67 (2003), no. 6,
1187 -- 1212.
[8] P. Dodds, B. de Pagter, E. Semenov, F. Sukochev, Symmetric functionals and singular
traces Positivity 2 (1998), no. 1, 47 -- 75.
[9] K. Dykema, T. Figiel, G. Weiss, M. Wodzicki, Commutator structure of operator ideals.
IMADA preprint PP-1997 -- 22, May 13, 1997, Odense.
[10] K. Dykema, T. Figiel, G. Weiss, M. Wodzicki, Commutator structure of operator ideals.
Adv. Math. 185 (2004), no. 1, 1 -- 79.
[11] K. Dykema, N. Kalton, Sums of commutators in ideals and modules of type II factors. Ann.
Inst. Fourier (Grenoble) 55 (2005), no. 3, 931 -- 971.
[12] T. Fack, H. Kosaki, Generalized s-numbers of τ -measurable operators Pacific J. Math. 123
(1986), no. 2, 269 -- 300.
[13] T. Figiel, N. Kalton, Symmetric linear functionals on function spaces Function Spaces,
Interpolation Theory and Related Topics. Lund, 2000
[14] I. Gohberg, M. Krein, Introduction to the theory of linear nonselfadjoint operators. Trans-
lations of Mathematical Monographs, Vol. 18 American Mathematical Society, Providence,
R.I. 1969.
[15] I. Gohberg, M. Krein, Theory and applications of Volterra operators in Hilbert space. Trans-
lations of Mathematical Monographs, Vol. 24 American Mathematical Society, Providence,
R.I. 1970.
[16] D. Guido, T. Isola, Singular traces on semifinite von Neumann algebras. J. Funct. Anal.
134 (1995), no. 2, 451 -- 485.
[17] D. Guido, T. Isola, Dimensions and singular traces for spectral triples, with applications to
fractals. J. Funct. Anal. 203 (2003), no. 2, 362 -- 400.
[18] N. Kalton, A. Sedaev, F. Sukochev, Fully symmetric functionals on a Marcinkiewicz space
are Dixmier traces Adv. Math. 226 (2011) 3540 -- 3549.
[19] N. Kalton, F. Sukochev, Symmetric norms and spaces of operators J.reine angew. Math.
(2008), 1 -- 41
30
F. SUKOCHEV AND D. ZANIN
[20] N. Kalton, F. Sukochev, Rearrangement-invariant functionals with applications to traces on
symmetrically normed ideals Canad. Math. Bull. 51 (2008), 67 -- 80.
[21] N. Kalton, F. Sukochev, D. Zanin, Orbits in symmetric spaces. II Studia Math. 197 (2010),
no. 3, 257 -- 274.
[22] S. Krein, Ju. Petunin, E. Semenov, Interpolation of linear operators, Nauka, Moscow, 1978
(in Russian); English translation in Translations of Math. Monographs, Vol. 54, Amer.
Math. Soc., Providence, RI, 1982.
[23] S. Kwapien, Linear functionals invariant under measure preserving transformations Math.
Nachr. 119 (1984), 175 -- 179.
[24] J. Lindenstrauss, L. Tzafriri, Classical Banach Spaces II: Function Spaces, Springer, 1996.
[25] S. Lord, A. Sedaev, F. Sukochev, Dixmier traces as singular symmetric functionals and
applications to measurable operators. J. Funct. Anal. 224 (2005), no. 1, 72 -- 106.
[26] M. Reed, B. Simon, Methods of modern mathematical physics. I. Functional analysis. Second
edition. Academic Press, Inc. [Harcourt Brace Jovanovich, Publishers], New York, 1980.
[27] A. Pietsch, About the Banach envelope of l1
,
∞. Rev. Mat. Complut. 22 (2009), no. 1, 209 --
226.
[28] R. Schatten, Norm ideals of completely continuous operators. Second printing. Ergebnisse
der Mathematik und ihrer Grenzgebiete, Band 27 Springer-Verlag, Berlin-New York 1970.
[29] B. Simon, Trace ideals and their applications. Second edition. Mathematical Surveys and
Monographs, 120. American Mathematical Society, Providence, RI, 2005.
[30] F. Sukochev, D. Zanin, Orbits in symmetric spaces J.Funct.Anal. 257 (2009), no.1, 194-218.
[31] M. Takesaki, Theory of operator algebras. I. Reprint of the first (1979) edition. Encyclopae-
dia of Mathematical Sciences, 124. Operator Algebras and Non-commutative Geometry, 5.
Springer-Verlag, Berlin, 2002.
[32] J. Varga, Traces on irregular ideals. Proc. Amer. Math. Soc. 107 (1989), no. 3, 715 -- 723.
[33] M. Wodzicki, Vestigia investiganda. Mosc. Math. J. 2 (2002), no. 4, 769 -- 798, 806.
School of Mathematics and Statistics, University of New South Wales, Sydney, 2052,
Australia.
E-mail address: [email protected]
School of Mathematics and Statistics, University of New South Wales, Sydney, 2052,
Australia.
E-mail address: [email protected]
|
1811.00100 | 4 | 1811 | 2019-06-02T17:40:58 | Pullbacks of graph C*-algebras from admissible pushouts of graphs | [
"math.OA",
"math.QA"
] | We define an admissible decomposition of a graph $E$ into subgraphs $F_1$ and $F_2$, and consider the intersection graph $F_1\cap F_2$ as a subgraph of both $F_1$ and $F_2$. We prove that, if the graph $E$ is row finite and its decomposition into the subgraphs $F_1$ and $F_2$ is admissible, then the graph C*-algebra $C^*(E)$ of $E$ is the pullback C*-algebra of the canonical surjections from $C^*(F_1)$ and $C^*(F_2)$ onto $C^*(F_1\cap F_2)$. | math.OA | math |
PULLBACKS OF GRAPH C*-ALGEBRAS
FROM ADMISSIBLE PUSHOUTS OF GRAPHS
PIOTR M. HAJAC
Instytut Matematyczny, Polska Akademia Nauk
´Sniadeckich 8, 00-656 Warszawa, Poland
E-mail: [email protected]
and
Department of Mathematics
University of Colorado Boulder
2300 Colorado Avenue Boulder, CO 80309-0395 USA
SARAH REZNIKOFF
Mathematics Department, Kansas State University
138 Cardwell Hall, Manhattan, KS 66502 US
E-mail: [email protected]
MARIUSZ TOBOLSKI
Instytut Matematyczny, Polska Akademia Nauk
´Sniadeckich 8, 00-656 Warszawa, Poland
E-mail: [email protected]
Abstract. We define an admissible decomposition of a graph E into subgraphs F1
and F2, and consider the intersection graph F1 ∩ F2 as a subgraph of both F1 and F2.
We prove that, if the graph E is row finite and its decomposition into the subgraphs F1
and F2 is admissible, then the graph C*-algebra C∗(E) of E is the pullback C*-algebra
of the canonical surjections from C∗(F1) and C∗(F2) onto C∗(F1 ∩ F2).
1. Introduction and preliminaries
Pushouts of graphs have proven to be very useful in the theory of free groups [11]. We
hope that our approach to pullbacks of graph algebras through pushouts of underlying
graphs will also turn out to be beneficial.
A graph C*-algebra is the universal C*-algebra associated to a directed graph. If one
considers a specific class of morphisms of directed graphs (e.g., see [1, Definition 1.6.2]),
then the graph C*-algebra construction yields a covariant functor from the category of di-
rected graphs to the category of C*-algebras. On the other hand, Hong and Szyma´nski [7]
showed that a pushout diagram in the category of directed graphs can lead to a pullback
of C*-algebras. The purpose of this paper is to find conditions on the pushout diagram of
2010 Mathematics Subject Classification. 46L45, 46L55, 46L85.
Key words and phrases. graph algebra, pushout and pullback, gauge action, quantum spaces: spheres,
balls, lens spaces, weighted projective spaces.
1
graphs that give rise to the pullback diagram of the associated graph C*-algebras. This
leads to a notion of an admissible decomposition of a directed graph, which we present
in Section 2. The main result is contained in Section 3 and examples are in Section 4.
Our result is closely related to [8, Corollary 3.4], where it is proven, in an appro-
priate form, for k-graphs without sinks but not necessarily row finite. Herein, we focus
our attention on row-finite 1-graphs but possibly with sinks. Thus our results are com-
plementary and lead to the following question: Is it possible to get rid of both of these
assumtions ("row finite" and "no sinks") at the same time to prove a more general
pushout-to-pullback theorem?
In this paper, by a graph E we will always mean a directed graph, i.e. a quadruple
(E0, E1, sE, rE), where E0 is the set of vertices, E1 is the set of edges, sE : E1 → E0 is
the source map and rE : E1 → E0 is the range map. A graph E is called row finite if
each vertex emits only a finite number of edges. Next, E is called finite if both E0 and
E1 are finite. A vertex is called a sink if it does not emit any edge. By a path µ in E
of length µ = k > 0 we mean a sequence of composable edges µ = e1e2 . . . ek. We treat
vertices as paths of length zero. The set of all finite paths for a graph E is denoted by
Path(E). One extends the source and the range maps to Path(E) in a natural way. We
denote the extended source and range maps by sPE and rPE, respectively.
Definition 1.1. The graph C*-algebra C∗(E) of a row-finite graph E is the universal
C*-algebra generated by mutually orthogonal projections P :=(cid:8)Pv v ∈ E0(cid:9) and partial
isometries S :=(cid:8)Se e ∈ E1(cid:9) satisfying the Cuntz -- Krieger relations [4]:
(cid:88)
S∗
e Se = PrE (e)
SeS∗
e = Pv
for all e ∈ E1, and
for all v ∈ E0 that are not sinks.
(CK1)
(CK2)
e∈s−1
E (v)
The datum {S, P} is called a Cuntz -- Krieger E-family.
One can show that the above relations imply the standard path-algebraic relations:
(1.1)
S∗
f Se = 0 for
e (cid:54)= f, PsE (e)Se = Se = SePrE (e) .
Any graph C*-algebra C∗(E) can be endowed with a natural circle action
(1.2)
α : U (1) −→ Aut(C∗(E))
defined by its values on the generators:
(1.3)
αλ(Pv) = Pv ,
αλ(Se) = λSe , where λ ∈ U (1),
v ∈ E0,
e ∈ E1.
The thus defined circle action is called the gauge action.
A subset H of E0 is called hereditary iff, for any v ∈ H such that there is a path
starting at v and ending at w ∈ E0, we have w ∈ H. (Note that we can equivalently
define the property of being hereditary by replacing "path" with "edge".) A subset H of
E0 is called saturated iff there does not exist a vertex v /∈ H such that
(1.4)
0 < s−1
E (v) < ∞ and
2
rE(s−1
E (v)) ⊆ H.
Saturated hereditary subsets play a fundamental role in the theory of gauge-invariant
It follows from [2, Lemma 4.3] that, for any hereditary
ideals of graph C*-algebras.
subset H, the algebraic ideal generated by {Pv v ∈ H} is of the form
IE(H) = span(cid:8)SxS∗
y x, y ∈ Path(E) , rPE(x) = rPE(y) ∈ H(cid:9) .
(1.5)
Here, for any path µ = e1 . . . ek, we adopt the notation Sµ := Se1 . . . Sek. Furthermore, if
µ is a vertex, then Sµ := Pµ.
By [2, Theorem 4.1 (b)], quotients by closed ideals generated by saturated hereditary
subsets can also be realised as graph C*-algebras by constructing a quotient graph. Given
a hereditary subset H of E0, the quotient graph E/H is given by
(E/H)0 := E0 \ H
(E/H)1 := E1 \ r−1
(1.6)
Note that the restriction-corestriction of the range map rE to (E/H)1 → (E/H)0 makes
sense for any H, but the same restriction-corestriction of the source map sE exists because
H is hereditary.
E (H).
and
Moreover, if H is also saturated, we obtain the *-isomorphism
(1.7)
C∗(E)/IE(H) ∼= C∗(E/H),
where IE(H) is the norm closure of IE(H).
2. Admissible decompositions of graphs
Given two graphs E = (E0, E1, sE, rE) and G = (G0, G1, sG, rG), one can define a
graph morphism f : E → G as a pair of mappings f 0 : E0 → G0 and f 1 : E1 → G1
satisfying
(2.1)
sG ◦ f 1 = f 0 ◦ sE
and rG ◦ f 1 = f 0 ◦ rE .
We call the thus obtained category the category of directed graphs.
A subgraph of a graph E = (E0, E1, sE, rE) is a graph F = (F 0, F 1, sF , rF ) such that
(2.2)
F 0 ⊆ E0, F 1 ⊆ E1,
∀ e ∈ F 1 : sF (e) = sE(e) and rF (e) = rE(e).
Next, let F1 and F2 be two subgraphs of a graph E. We define their intersection and
union as follows:
F1 ∩ F2 := (F 0
1 ∩ F 0
2 , F 1
1 ∩ F 1
2 , s∩, r∩),
∀ e ∈ F 1
1 ∩ F 1
2 : s∩(e) := sE(e), r∩(e) := rE(e),
F1 ∪ F2 := (F 0
1 ∪ F 0
2 , F 1
1 ∪ F 1
2 , s∪, r∪),
∀ e ∈ F 1
1 ∪ F 1
2 : s∪(e) := sE(e), r∪(e) := rE(e).
(2.3)
To consider pushout diagrams in the category of directed graphs, we follow the con-
vention used in [5]. If a graph E has two subgraphs F1 and F2 such that E = F1 ∪ F2,
3
then the following diagram
E
(2.4)
F1
F2
F1 ∩ F2
is automatically a pushout diagram. Let us illustrate the concept of a pushout diagram
of graphs with the following example:
(2.5)
We are now ready to define an admissible decomposition of a row-finite graph:
Definition 2.1. An unordered pair {F1, F2} of subgraphs of a row-finite graph E is called
an admissible decomposition of E iff the following conditions are satisfied:
.
(1) E = F1 ∪ F2 ,
(2) if v is a sink in F1 ∩ F2 , then v is a sink in Fi , i = 1, 2,
2 = r−1
(3) F 1
2 ), i = 1, 2.
1 ∩ F 0
1 ∩ F 1
(F 0
Fi
Note that, by (1) in Definition 2.1, E is a pushout of F1 and F2 over their intersection.
Observe also that Diagram (2.5) gives an example of an admissible decomposition of a
graph.
Definition 2.1 prompts the following two lemmas.
Lemma 2.2. Let {F1, F2} be an admissible decomposition of a row-finite graph E. Then
F1 ∩ F2 = Fi/(F 0
i ), for i = 1, 2.
1 ∩ F 0
i \ (F 0
2 )) and Fi = E/(E0 \ F 0
i \ (F 0
2 implies e /∈ F 1
1 ∩ F 0
1 ∩F 1
Proof. First, note that F 0
sFi(e) /∈ F 0
1 ∩F 0
2 ) is hereditary in Fi. Indeed, take e ∈ F 1
2 . Hence, by Definition 2.1(3), we have rFi(e) /∈ F 0
i . Then
1 ∩F 0
2 .
4
;
;
c
c
c
c
;
;
=
=
a
a
:
:
d
d
Therefore, we can define Fi/(F 0
nition 2.1(3).
i \ (F 0
1 ∩ F 0
2 )), which coincides with F1 ∩ F2 due to Defi-
Next, note that
i = (F 0
E0 \ F 0
i = F 0
(2.6)
where j (cid:54)= i and j = 1, 2, so we already know that E0 \ F 0
that it is hereditary in E, we only need to exclude edges starting in E0 \ F 0
in E0 \ F 0
j . They do not exist because E1 = F 1
j , so E0 \ F 0
i ∩ F 0
j ),
is hereditary in Fj. To see
i and ending
is hereditary in E.
j ) \ F 0
j \ (F 0
i ∪ F 0
i ∪ F 1
j \ F 0
i = F 0
i
i
E (F 0
i ). To this end, taking advantage of the admis-
It remains to verify that F 1
i = r−1
sibility of (Fi ∩ Fj) ⊆ Fi, we compute
i ) \ F 1
(2.7)
r−1
E (F 0
(F 0
i = r−1
(F 0
i ) \ F 1
i ⊆ r−1
i = r−1
E (F 0
Fj
i ), we conclude that F 1
Fj
(cid:4)
Therefore, as F 1
Lemma 2.3. Let {F1, F2} be an admissible decomposition of a row-finite graph E. Then
i \ (F 0
F 0
1 ∩ F 0
i ∩ F 0
i = r−1
j ) \ F 1
E (F 0
i = (F 1
i ∩ F 1
i ), as desired.
j ) \ F 1
i = ∅.
2 ) is saturated in Fi and in E for i = 1, 2.
i \ (F 0
1 ∩ F 0
2 ) were not saturated in Fi, then there would exist a vertex v in
i \ (F 0
F 0
i \ (F 0
1 ∩ F 0
2 )) = F 0
1 ∩ F 0
2
Proof. If F 0
(2.8)
such that
s−1
(v) (cid:54)= ∅ and rFi(s−1
(v)) ⊆ F 0
i \ (F 0
1 ∩ F 0
2 ).
(2.9)
Thus we would have a vertex in F1 ∩ F2 that is a sink in F1 ∩ F2 but not in Fi, which
contradicts Definition 2.1(2).
Fi
Fi
Much in the same way, if F 0
i \ (F 0
1 ∩ F 0
2 ) were not saturated in E, then there would
exist a vertex
w ∈ E0 \ (F 0
i \ (F1 ∩ F 0
(2.10)
where j (cid:54)= i and j = 1, 2, such that
E (w) (cid:54)= ∅ and rE(s−1
s−1
(2.11)
Hence, there is e ∈ s−1
E (w) such that rE(e) /∈ F 0
e ∈ F 1
again contradicts Definition 2.1(2).
i , so w = sE(e) ∈ F 0
E (w)) ⊆ F 0
i \ (F 0
1 ∩ F 0
2 ).
i ∪ F 1
j . As E1 = F 1
j , it follows that
i . Consequently, w is a sink in F1 ∩ F2 but not in Fi, which
(cid:4)
2 )) = F 0
j ,
3. Pullbacks of graph C*-algebras
Let {F1, F2} be an admissible decomposition of a row-finite graph E. Then, by
Lemma 2.2 and Lemma 2.3, we can take an advantage of the formula (1.7) to define the
canonical quotient maps:
(3.1)
(3.2)
(3.3)
π1 : C∗(E) −→ C∗(E)/IE(F 0
π2 : C∗(E) −→ C∗(E)/IE(F 0
1 \ (F 0
χ1 : C∗(F1) −→ C∗(F1)/IF1(F 0
2 \ F 0
1 \ F 0
1 ∩ F 0
1 ) ∼= C∗(F1),
2 ) ∼= C∗(F2),
2 )) ∼= C∗(F1 ∩ F2),
5
(3.4)
χ2 : C∗(F2) −→ C∗(F2)/IF2(F 0
2 \ (F 0
1 ∩ F 0
2 )) ∼= C∗(F1 ∩ F2).
Note that quotient maps are automatically U (1)-equivariant for the gauge action.
This brings us to the main theorem:
Theorem 3.1. Let {F1, F2} be an admissible decomposition of a row-finite graph E.
Then there exist canonical quotient gauge-equivariant ∗-homomorphisms rendering the
following diagram
(3.5)
C∗(F1)
π1
χ1
C∗(E)
C∗(F1 ∩ F2)
π2
χ2
C∗(F2)
commutative. Moreover, this is a pullback diagram of U (1)-C*-algebras.
Proof. Note first that all the canonical surjections in the diagram are well defined due to
the admissibility conditions of the decomposition of the graph E (see the discussion at the
beginning of this section). The commutativity of the diagram is obvious as all maps are
canonical surjections. Finally, using [9, Proposition 3.1] and the surjectivity of χ1 and χ2,
to prove that (3.5) is a pullback diagram, it suffices to show that ker π1 ∩ ker π2 = {0}
and that π2(ker π1) ⊆ ker χ2.
Since ker π1 and ker π2 are closed ideals in a C*-algebra, we know that
(3.6)
ker π1 ∩ ker π2 = ker π1 ker π2.
1 \ F 0
Next, as F 0
that
2 and F 0
2 \ F 0
1 are saturated hereditary subsets of E0, it follows from (1.7)
(3.7)
ker π1 = IE(F 0
2 \ F 0
1 )
and
ker π2 = IE(F 0
1 \ F 0
2 ).
Furthermore, using the characterization (1.5) of ideals generated by hereditary subsets,
we know that an arbitrary element of ker π1 ker π2 is in the closed linear span of elements
of the form SαS∗
βSγS∗
δ , where α, β ∈ Path(E) with
rPE(α) = rPE(β) ∈ F 0
2 \ F 0
1 ,
(3.8)
and γ, δ ∈ Path(E) with
(3.9)
The conclusion ker π1 ∩ ker π2 = {0} follows from the analysis of all possible paths satis-
fying the above conditions.
rPE(γ) = rPE(δ) ∈ F 0
1 \ F 0
2 .
Indeed, it follows from (1.1) that S∗
2 , rPE(β) ∈ F 0
2 \ F 0
βSγ (cid:54)= 0 is possible only if sPE(β) = sPE(γ). As
2 , if β = e1 . . . em and γ = f1 . . . fn,
1 \ F 0
1 and rPE(γ) ∈ F 0
1 ∪ F 1
E1 = F 1
we infer that
(3.10)
rE(em−1) = sE(em) ∈ F 0
2
and rE(fn−1) = sE(fn) ∈ F 0
1 .
6
x
x
&
&
&
&
x
x
1 ∩ F 0
2 or rE(em−1) ∈ F 0
Hence rE(em−1) ∈ F 0
1 . Now, we continue by induction using
Definition 2.1(3) for the intersection case of the alternative. This brings us to conclusion
that sPE(β) ∈ F 0
1 . It follows that
sPE(β) = sPE(γ) ∈ F 0
1 \ F 0
2 ,
we conclude that β (cid:54)= γ, so there exists the smallest index i such that ei (cid:54)= fi. Now,
remembering the relations (CK1) and (1.1), we compute
2 . Much in the same way, we argue that sPE(γ) ∈ F 0
2 . Furthermore, as rPE(β) ∈ F 0
1 and rPE(γ) ∈ F 0
1 ∩ F 0
2 \ F 0
2 \ F 0
S∗
βSγ = S∗
= S∗
ei+1...emS∗
ei+1...emS∗
ei
ei
S∗
ei−1
. . . S∗
e1Se1 . . . Sei−1SfiSfi+1...fn
SfiSfi+1...fn
(3.11)
= 0.
Finally, if β or γ is a path of length zero, i.e. a vertex, then it is straightforward to
conclude that S∗
βSγ = 0.
Next, taking again an advantage of (1.5) and (1.7), we obtain
ker χ2 = IF2(F 0
β α, β ∈ Path(F2), rPF2(α) = rPF2(β) ∈ F 0
2 \ F 0
1 ) = span{SαS∗
2 \ F 0
Any element of IF2(F 0
1 ) is an element of IE(F 0
2 \ F 0
α ∈ Path(F2). Hence π2(IE(F 0
we conclude that π2(ker π1) ⊆ ker χ2.
Remark 3.2. One can also prove Theorem 3.1 in the setting of Leavitt path algebras [1].
A proof of the Leavitt version of Theorem 3.1 is completely analogous due to [1, Corol-
lary 2.5.11].
2 \ F 0
1 }.
1 ), and π2(Sα) = Sα for all
1 ). Finally, from the continuity of π2,
(cid:4)
1 )) ⊆ IF2(F 0
2 \ F 0
2 \ F 0
4. Examples
We end the paper by providing motivating examples from noncommutative topology.
4.1. Even quantum spheres. Not only the graph at the top of the diagram (2.5)
representing the generic Podle´s quantum sphere [10] admits a natural admissible decom-
position, but also the finite graphs L2n [6, Section 5.1] representing, respectively, the
C*-algebras C(S2n
q ) of all even quantum spheres enjoy natural admissible decomposi-
tions {F 1
2n) = C∗(F 2
q ) of the
Hong-Szyma´nski quantum 2n-ball [7, Section 3.1], and C∗(F 1
2n) coincides [6, Ap-
pendix A] with the C*-algebra C(S2n−1
) of the boundary Vaksman-Soibelman quantum
odd sphere [12]. Thus we recover in terms of graphs the classical fact that an even sphere
is a gluing of even balls over the boundary odd sphere.
2n) coincides with the C*-algebra C(B2n
2n}. Here C∗(F 1
2n ∩ F 2
2n, F 2
q
As Theorem 3.1 applies, we infer that the diagram
(4.1)
C(B2n
q )
π1
χ1
C(S2n
q )
C(S2n−1
q
)
7
π2
χ2
C(B2n
q )
y
y
%
%
%
%
y
y
is a pullback diagram. This fact was already proved in [7, Proposition 5.1] by direct
considerations of generators and relations.
The case n = 3 is illustrated by the diagram:
(4.2)
.
4.2. Quantum lens space L3
space L3
q(l; 1, l) can be viewed as the graph C*-algebra (e.g., see [3]) of the graph L3
l :
q(l; 1, l). The C*-algebra C(L3
q(l; 1, l)) of the quantum lens
v0
0
···
v1
0
v1
1
v1
l−2
v1
l−1
.
(4.3)
The graph L3
yielding, by Theorem 3.1, the pullback diagram:
l enjoys an admissible decomposition {L3
l−k}, where k ∈ {1, . . . l − 1},
k, L3
(4.4)
C(L3
q(k; 1, k))
π1
χ1
C(L3
q(l; 1, l))
π2
C(L3
q(l − k; 1, l − k))
χ2
C(S1) .
8
:
:
d
d
9
9
e
e
w
w
)
)
(
(
u
u
Recall that C∗(L3
1) ∼= C(S3
q ), so for l = 2 we obtain the following pullback diagram:
(4.5)
C(S3
q )
C(L3
q(2; 1, 2))
π1
χ1
C(S1) .
π2
χ2
C(S3
q )
Since the above diagram is U (1)-equivariant,
U (1)-fixed-point subalgebras:
it induces a pullback diagram for
(4.6)
C(CP 1
q )
C(WP 1
q (1, 2))
π1
χ1
wC .
π2
χ2
C(CP 1
q )
q ) and C(WP 1
Here C(CP 1
q (1, 2)) denote the quantum complex projective space (see [6,
Section 2.3]) and the quantum weighted projective space (see [3, Section 3]), respectively.
Interestingly, the C*-algebras in the above diagram can be viewed as graph C*-algebras,
and an infinite graph representing C(WP 1
q (1, 2)) is a pushout of infinite graphs repre-
q ) over the graph consisting of one vertex and no edges representing C
senting C(CP 1
(see Diagram (4.7) below). However, this example is beyond the scope of Theorem 3.1,
because the above diagram is no longer U (1)-equivariant and the infinite graphs are not
row finite.
(∞)
(∞)
(4.7)
(∞)
(∞)
Here edges with (∞) denote countably infinitely many edges.
9
x
x
&
&
&
&
x
x
w
w
'
'
'
'
w
6
6
h
h
5
5
i
i
Acknowledgements
The work on this project was partially supported by NCN-grant 2015/19/B/ST1/03098
(Piotr M. Hajac, Mariusz Tobolski) and by a Simons Foundation Collaboration Grant
(Sarah Reznikoff). Piotr M. Hajac is very grateful to Kansas State University for its hos-
pitality and financial support provided by this Simons Foundation Collaboration Grant.
It is a pleasure to thank Carla Farsi for drawing our attention to using pushouts of graphs
in the theory of free groups, Tatiana Gateva -- Ivanova for a helpful discussion concerning
admissible decompositions of graphs, and Aidan Sims for making us aware on how to
prove our main result in a different context.
References
[1] G. Abrams, P. Ara, M. Siles Molina, Leavitt path algebras. Lecture Notes in Mathematics, 2191.
Springer, London, 2017.
[2] T. Bates, D. Pask, I. Raeburn, W. Szyma´nski, The C*-algebras of row-finite graphs, New York
J. Math., 6 (2000), 307 -- 324.
[3] T. Brzezi´nski, W. Szyma´nski, The C*-algebras of quantum lens and weighted projective spaces,
J. Noncomm. Geom., 12 (2018), 195 -- 215.
[4] J. Cuntz, W. Krieger, A class of C∗-algebras and topological Markov chains, Invent. Math., 56
(1980), 251 -- 268.
[5] H. Ehrig, H.-J. Kreowski, Pushout-properties: an analysis of gluing constructions for graphs. Math.
Nachr., 91 (1979), 135 -- 149.
[6] J. H. Hong, W. Szyma´nski, Quantum spheres and projective spaces as graph algebras, Comm.
Math. Phys., 232 (2002), 157 -- 188.
[7] J. H. Hong, W. Szyma´nski, Noncommutative balls and mirror quantum spheres, J. Lond. Math.
Soc., 77 (2008), 607 -- 626.
[8] A. Kumjian, D. Pask, A. Sims, M. F. Whittaker. Topological spaces associated to higher-rank
graphs. J. Combin. Theory Ser., A 143 (2016), 19 -- 41.
[9] G. K. Pedersen, Pullback and Pushout Constructions in C*-Algebra Theory, J. Funct. Anal., 167
(1999), 243 -- 344.
[10] P. Podle´s, Quantum spheres, Lett. Math. Phys. 14 (1987), no. 3, 193 -- 202.
[11] J. R. Stallings. Topology of finite graphs. Invent. Math., 71 (1983), no. 3, 551 -- 565.
[12] L. L. Vaksman, Ya. S. Soibelman, Algebra of functions on the quantum group SU(n + 1) and
odd-dimensional quantum spheres, Algebra i Analiz, 2 (1990), 101 -- 120.
10
|
1209.4094 | 3 | 1209 | 2016-06-09T20:17:12 | Construction of enveloping actions | [
"math.OA",
"math.DS"
] | We study the problem of constructing a globalization for partial actions on *-algebras, C*-algebras and Hilbert modules. For the first ones we give a necessary condition for the existence of a globalization and we prove this conditions is necessary and sufficient for C*-algebras. Using the linking algebra of a Hilbert module we translate this condition to the realm of partial action on Hilbert modules. | math.OA | math |
CONSTRUCTION OF ENVELOPING ACTIONS
DAMI ´AN FERRARO
Abstract. We study the problem of constructing a globalization for partial
actions on *-algebras, C*-algebras and Hilbert modules. For the first ones we
give a necessary condition for the existence of a globalization and we prove
this conditions is necessary and sufficient for C*-algebras. Using the linking
algebra of a Hilbert module we translate this condition to the realm of partial
action on Hilbert modules.
1. Introduction
Among the simplest examples of partial actions on C*-algebras [5, 6, 7] we find
restrictions of actions (also called global actions) to non invariant C*-ideals [1].
Many aspects of these examples, as the representations and crossed products, can be
studied using techniques developed for actions on C*-algebras. Then it is interesting
to know which partial actions can be globalized, that is: described as the restriction
of a global action (the globalization). This problem was stated in [1], where it was
solved in case the C*-algebra is commutative.
Partial actions can be defined in other categories, as topological spaces, rings and
Hilbert modules [1, 2, 3, 4]. Here we will work with *-algebras, C*-algebras and
Hilbert modules, mainly because C*-algebras are, at once, *-algebras and Hilbert
modules.
This work is organized as follows.
In the first section we study the problem
of globalizing partial actions on *-algebras (*-partial actions). Our intention is
to study continuous partial action on C*-algebras (C*-partial actions) from a *-
algebraic point of view, which amounts to dump all the topological structure keep-
ing the *-algebraic properties we need. Hence, when necessary, we make some
assumptions on the *-algebras which are known to hold for C*-algebras. Under
these assumptions we give a necessary and sufficient condition for the existence of a
globalization. Then we turn to consider all the structure of C*-partial actions. At
this point we show a C*-partial action can be globalized to a C*-partial action if and
only if it can be globalized to a *-partial action, independently of the topological
structure. Finally, in the last section, we give a necessary and sufficient condition
for the existence of a globalization of a partial action on a Hilbert bimodule (Hb-
partial action). We specifically show that a Hb-partial action has a globalization if
and only if it's linking partial action [1] has a globalization.
Date: September 3, 2018.
2010 Mathematics Subject Classification. Primary 46L55. Secondary 46L40, 46L05.
Key words and phrases. Partial actions, Enveloping actions.
This work was partially supported by Mathamsud network U11-MATH05 (partially funded by
ANII, Uruguay) and started during my visit to the Universidade Federal de Santa Catarina. I
thank Professors Alcides Buss and Ruy Exel for receiving me there.
1
2
DAMI ´AN FERRARO
2. Partial action on *-algebras
Each one of the types of partial actions considered here has it's own notion of
globalization. Besides, a partial action on a C*-algebra is a partial action in a *-
algebra and in a Hilbert module. For this reason we introduce the terms "*-partial
action", "C*-partial action" and "Hb-partial action".
In case the partial action
under consideration is global we substitute the term partial for global.
We start by recalling some definitions and facts.
2.1. Algebras with involution. A *-algebra is an algebra A over the complex
field together with a conjugate linear function (the involution) A → A, a 7→ a∗,
satisfying (ab)∗ = b∗a∗ and a∗∗ = a.
Assume A is a *-algebra. By a *-ideal of A we mean a subspace I ⊂ A such
that I ∗ = I and IA ⊂ A. A function between *-algebras, φ : A → B, is a *-
homomorphism if it is linear, multiplicative and φ(a∗) = φ(a)∗. An automorphism
of A is a bijective *-homomorphism from A to A and the set of automorphism of
A will be denoted Aut(A).
An element a ∈ A is a right (left) annihilator if aA = {0} (Aa = {0}). Note a is
a right annihilator if and only if a∗ is a left annihilator. We say A is non-degenerate
if it does not have a right (or left) annihilator different from 0.
Remark 2.1. If A is non-degenerate then for all a, b ∈ A the following conditions
are equivalent: (1) a = b, (2) for all c ∈ A, ac = bc (3) for all c ∈ A, ca = cb and
(4) for all c, d ∈ A, cad = cbd.
A double centraliser of A is a pair µ = (L, R), where L, R : A → A are linear
functions, L(ab) = L(a)b, R(ab) = aR(b) and aL(b) = R(a)b (for all a, b ∈ A). It
is usual to write µa := L(a) and aµ := R(a). The *-algebra structure of M (A) is
given by point wise vector space operations, product (L, R)(M, S) := (L◦ M, S◦ R)
and involution (L, R)∗ := (L′, R′), where L′(a) = R(a∗)∗ and R′(a) = L(a∗)∗. The
function τ : A → M (A), where τ (a)b = ab and bτ (a) = ba, is *-homomorphism
which is injective if and only if A is non-degenerate.
Remark 2.2. Double centralisers of *-algebras can be treated as adjointable opera-
tors of Hilbert modules. More precisely, if A is non-degenerate then
(1) Given a function T : A → A there exists (L, R) ∈ M (A) such that T = L if
and only if there exists a function T ∗ : A → A, with T (a)∗b = a∗T ∗(b) for
all a, b ∈ A.
(2) Given (L1, R1), (L2, R2) ∈ M (A) we have L1 = L2 if and only if R1 = R2.
Take any group G and name AG the *-algebra of functions from G to A (point
wise operations). Given t ∈ G name θt the automorphism of AG defined by
θt(f )r := frt. The canonical action of G on M (AG) is Θ : G → Aut(M (AG))
where Θt(L, R) = (θt ◦ L ◦ θt−1, θt ◦ R ◦ θt−1).
2.2. Partial actions and globalizations. Recall [7, 6] that a partial action of a
group G on a set X is a pair σ := ({Xt}t∈G,{σt}t∈G) where
(1) For all t ∈ G, Xt is a subset of X and σt : Xt−1 → Xt a function.
(2) σe is the identity of X.
(3) For all s, t ∈ G, if x ∈ Xt−1 and σt(x) ∈ Xs−1 , then x ∈ Xt−1s−1 and
σs(σt(x)) = σst(x).
CONSTRUCTION OF ENVELOPING ACTIONS
3
In case Xt = X for all t ∈ G, σ is said to be global and it is just a common action
of G on X.
Let us assume that σ and τ are partial actions of G on the sets X and Y,
respectively. We say f : σ → τ is a morphism of partial actions on sets if f is a
function from X to Y and for all t ∈ G : f (Xt) ⊂ Yt and f (σt(x)) = τt(f (x))
(for all x ∈ Xt−1). The identity morphism associated to σ is idσ := idX and the
composition of morphism is just the composition of functions.
With σ as before take a set U ⊂ X. Given t ∈ G define Ut := U ∩ σt(Xt−1 ∩ U ).
It is obvious that Ut−1 ⊂ Xt−1 and σt(Ut−1) ⊂ Ut; hence it makes sense to define
κt : Ut−1 → Ut as κt(x) = σt(x). Straightforward arguments imply the restriction
of σ to U , defined as σU := ({Ut}t∈G,{κt}t∈G) , is partial action of G on U. From
[1] we know every partial action on a set is isomorphic to the restriction of a global
action on a set.
A set U ⊂ X is said to be σ−invariant if for all t ∈ G, σt(Xt−1 ∩ U ) ⊂ U.
Remark 2.3. If U ⊂ V ⊂ X then σU = σV U . Besides, if σ is global, σU is global
if and only if U is σ−invariant.
Definition 2.4. A *-partial action of the (discrete) group G on the *-algebra A is
a set theoretic partial action of G on A, α = ({At}t∈G,{αt}t∈G) , such that At is a
*-ideal of A and αt a *-homomorphism (for all t ∈ G).
Example 2.5. Let α be a *-global action of G on the *-algebra A and let I be a
*-ideal of A. The restriction of α to I, αI := (cid:0){I ∩ αt(I)}t∈G,{αtI∩αt(I)}t∈G(cid:1) , is
a *-partial action of G on I.
Example 2.6. Let A be a ∗−algebra and I a *-ideal of it. The *-partial action of
Z2 = {0, 1} on A determined by I is αAI := ({A0, A1},{α0, α1}) where A0 = A,
A1 = I, α0 = idA and α1 = idI .
Morphisms of *-partial actions are just morphisms of partial actions on set which
are also *-homomorphisms. The composition and identity are the natural ones.
Definition 2.7 ([1, 3]). A *-globalization the *-partial action α (of G on A) is a
4-tuple Ξ = (B, β, I, ι) where: B is a *-algebra, β is a *-global action of G on B, I
is a *-ideal of B and ι : α → βI is an isomorphism.
we say Ξ is non degenerate if B is non degenerate.
We say Ξ is minimal if [I] := span{βt(I) : t ∈ G} equals B and, for convenience,
Do not confuse our concept of minimality with topological minimality, here β
may have many open invariant sets1. Note that Ξ is minimal iff the only β−invariant
In case Ξ is not minimal, ([I], β[I], I, ι) is
*-ideal of B containing I is B itself.
minimal because [I] is invariant and β global (Remark 2.3).
Example 2.8. Given a complex vector space V and a conjugate linear bijection
T : V → V with T 2 = idV , let V T be the *-algebra obtained by considering on
V the null product (uv = 0) and T as involution. Here we consider the entry
wise conjugation T : Cn → Cn. Let α be the partial action of Z3 = {0, 1, 2} (with
additive notation) on CT such that α0 = idCT and α1 = α2 = id{0}. There are two
completely different globalizations for α, (C2T , β, I, ι) and (C3T , γ, J, κ), that we
1Note that there is no topology involved.
4
DAMI ´AN FERRARO
now describe. Just set I = {(u, 0) : u ∈ C}, J = {(u, 0, 0) : u ∈ C}, ι(u) = (u, 0),
κ(u) = (u, 0, 0),
β1(u, v) = −
1
2
(u + √3v,−
√3u + v) and γ1(u, v, w) = (w, u, v).
Note β1 is the rotation by angle 2π/3 and a simple dimension argument implies
that β is not isomorphic to γ. In fact it can be shown that these are the unique
globalizations of α (up to isomorphism).
Example 2.9. Consider the partial action of Example 2.6. Assume there exists a *-
ideal J of A such that A = I⊕J. Now form the (external) direct sum B := I⊕J ⊕J,
considered as a *-algebra with entry wise operations. Let β be the action of Z2 on
B given by β1(a ⊕ b ⊕ c) = a ⊕ c ⊕ b and ι : A → B the unique linear map such
that ι(a + b) = a ⊕ b, for all a ∈ I and b ∈ J. Then (B, β, I ⊕ J ⊕ 0, ι) is a minimal
*-globalization of αAI .
The existence of the direct complement J is not necessary for the existence of a
globalization, as we show with the next Example. Moreover, with it we also show
that there are *-partial actions on non degenerate *-algebras with a *-globalization
but without a non degenerate *-globalization.
Example 2.10. Let U ∈ M4(C) be the matrix corresponding to the permutation
(1 4)(2 3) (written as a product of cycles). Define an involution in M4(C) by the
formula a∗ := uatu, where a 7→ a is the entry wise complex conjugation and a 7→ at
is the usual matrix transposition2. Note that a∗ is obtained from a by performing
a symmetry with respect to the anti-diagonal3.
Name A the ∗−subalgebra of M4(C) formed by the matrices of the form
a11 a12
0
0
0
0
0
0
a13 a14
a23 a24
a34
0
0
a44
With I := {a ∈ A : a23 = 0} we have I = I ∗ and AA ⊂ I. Then I is a *-ideal of A
and it can be shown that A and I are non-degenerate. Now let αAI be the *-partial
action described in Example 2.6.
Assume Ξ = (B, β, J, ι) is a *-globalization of αAJ . For convenience we think
J = A, ι = idA and α = βJ . Since B = A + β(A) and dim(A∩ β(A)) = dim(I) = 7,
we know dim(B) = dim(A) + 1 = 9. Let u ∈ A be the matrix with 1 in the entry
2− 3 and 0 elsewhere. Then A = I ⊕ Cu and β1(u) /∈ A because u /∈ I = A∩ β1(A).
Moreover, B = A⊕ Cβ1(u). As a vector space B is isomorphic to the external direct
sum A⊕ C and, with this notation, the involution of B is (a⊕ λ)∗ = a∗⊕ λ (because
u∗ = u).
To describe the product of B we compute
(a + λβ1(u))(b + µβ1(u)) = ab + µaβ1(u) + λβ1(u)b + λµβ1(u2).
Firstly note that u2 = 0, so β1(u2) = 0. Secondly, aβ1(u) ∈ I and for all c ∈ I
we have (aβ1(u) − au)c = aβ1(u)α1(c) − auc = aα1(uc) − auc = 0. Since I is non-
degenerate and au ∈ I, aβ1(u) = au. In the same way we deduce that β1(u)b = ub.
2With this structure M4(C) is not a C*-algebra because a∗a = 0 if ai,j = δ1(i)δ1(j).
3From lower left corner to upper right corner.
CONSTRUCTION OF ENVELOPING ACTIONS
5
Now the product of B takes the form
(a + λβ1(u))(b + µβ1(u)) = ab + µau + λub.
Then the formula for the product of A ⊕ C should be
(a ⊕ λ)(b ⊕ µ) := ab + µau + λub ⊕ 0.
In fact A ⊕ C is a *-algebra with this product and the involution described in
the previous paragraph. Moreover, it is degenerate because u ⊕ −1 is a bilateral
annihilator.
The final step is to determine β, which amounts to give an expression for β1.
Assume b = a+λu+µβ1(u), with a ∈ I and λ, µ ∈ C. Then β1(b) = β1(a)+λβ1(u)+
µu = a + µu + λβ1(u). In terms of A ⊕ C β1 should be given by β1((a + λu) ⊕ µ) =
(a ⊕ µu) ⊕ λ. Now the reader can verify that β1 is in fact a *-homomorphism of
A ⊕ C with β2
1 = idA⊕C and β1A = α. If we set ι : A → A ⊕ C as the natural
inclusion then (A ⊕ C, β, A ⊕ 0, ι) is a minimal *-globalization of α. Furthermore,
it is essentially the unique *-globalization of α.
The algebras from Example 2.8 are those in which all the elements are annihila-
tors. Our source of inspiration are partial action on C*-algebras and these algebras
are non-degenerate because on such algebras the identity aa∗ = 0 implies a = 0.
Then we will assume our *-algebras are non degenerate when needed.
Now our idea is to take a *-partial action with a *-globalization and try to
construct another *-globalization using just the *-partial action. In this way we will
be sure that the *-partial action has a *-globalization every time the construction
can be performed. This new *-globalization may be completely different from the
initial one.
Fix, for the rest of this section, a ∗−partial action α of G on A and a minimal
*-globalization Ξ = (B, β, I, ι) of α. The canonical morphism associated to Ξ,
denoted π or πΞ in case is necessary to mention Ξ, is the *-homomorphism π : B →
M (AG), π(b)fr = ι−1(βr(b)ι(fr)). A simple computation shows that π : β → Θ
is a morphism, where Θ is the canonical action of G on M (AG). Then π(B) is a
Θ-invariant *-subalgebra and π(A) is a *-ideal of it. Moreover, Θπ(B) is a *-global
action and Θπ(B)π(I) = Θπ(I). The quadruple ΞΘ := (π(B), Θπ(B), π(I), π ◦ ι)
is a minimal *-globalization of α if and only if π ◦ ι is an injective function and
π(It) = π(I) ∩ Θt(π(I)) (for all t ∈ G). Unfortunately these two conditions seem
to be unrelated in general.
Example 2.11. Let αAI be the *-partial action from Example 2.10 and (A⊕C, β, A⊕
0, ι) it's *-globalization. Then AZ2 = A⊕A and π(a⊕λ)(b⊕c) = (a+λu)b⊕(a+λu)c.
Since π(a⊕ λ) = π(a+ λu⊕ 0), we have π(A⊕ C) = π(A) and π is not injective. But
π ◦ ι is injective because A is non degenerate. In this case the kernel of π, ker(π),
is exactly the space generated by u ⊕ −1, which is the set of right annihilators of
A ⊕ C, AnnR(A ⊕ C). This last condition is not accidental, as we now show.
Lemma 2.12. Let Ξ = (B, β, I, ι) be a *-globalization of the *-partial action α (of
G on A). If π is the canonical morphism associated to Ξ then ker(π) = AnnR(B).
Moreover, if A is non-degenerate then π◦ ι is injective and π(B) is non degenerate.
Proof. Assume b ∈ ker(π). Given a ∈ A and r ∈ G denote bδr the element of
AG taking the value b at r and 0 elsewhere. Then 0 = ι(π(b)aδrr) = βr(b)ι(a).
This implies bB = span bβr(ι(A)) = 0, so b ∈ AnnR(B). For the converse assume
6
DAMI ´AN FERRARO
b ∈ AnnR(B) and take f ∈ AG. Then for all r ∈ G we have βr(b) ∈ AnnR(B), thus
π(b)fr = ι−1(βr(b)ι(fr)) = 0 and b ∈ ker(π).
In case A is non-degenerate and a ∈ ker(π ◦ ι), we have ι(a) ∈ AnnR(B). This
implies a ∈ AnnR(A) and, so, a = 0.
Finally assume π(b) ∈ AnnR(π(B)). Since π(B) is Θ−invariant, for all r ∈ G
we have π(βr(b)) = Θr(π(b)) ∈ AnnR(π(B)). Then, for all r ∈ G and c, d ∈ A,
0 = π(βr(b))π(ι(c))dδee = ι−1(βr(b)ι(c)ι(d)). This implies 0 = βr(b)ι(c) for all
r ∈ G and c ∈ A. Using the definition of π(b) we conclude that π(b) = 0.
(cid:3)
To give a sufficient condition for π ◦ ι to be an isomorphism we introduce as-
similative ideals. A subset S of the *-algebra C is assimilative (in C) if given
c ∈ C such that cC ⊂ S, then c ∈ J. It is immediate that every subset of a unital
*-algebra is assimilative. Furthermore, a *-ideal J of C is assimilative iff C/J is
non-degenerate.
Remark 2.13. Recall [4] that a ring R is s-unital if for all a ∈ R there exists b ∈ R
such that ba = a. Evidently every s-unital *-algebra is non-degenerate and any
subset of a s-unital *-algebra is assimilative in the algebra. Any sum of s-unital
ideals is s-unital [4, Remark 2.5].
The next Proposition resumes all we have to say about uniqueness and non
degeneracy of globalizations.
Proposition 2.14. Let α be a *-partial action of G on A with A non-degenerate
and At assimilative in A, for all t ∈ G. Then
(1) α has a *-globalization if and only if α has a non-degenerate and minimal
*-globalization.
(2) Given minimal *-globalizations of α, Ξ = (B, β, I, ι) and Σ = (C, γ, J, κ), with
Σ non-degenerate, there exists a unique morphism ΞρΣ : β → γ such that ΞρΣ ◦
ι = κ. Moreover, ΞρΣ is an isomorphism iff Ξ is non-degenerate iff πΞ is
injective.
(3) In case for all n ∈ N and t1, . . . , tn ∈ G the ideal At1 +··· + Atn is assimilative
in A, every minimal *-globalization of α is non-degenerate.
Proof. Assume α has a ∗−globalization Ξ = (B, β, I, ι), which we can assume is
minimal (see the comments after Definition 2.7). Let π be the canonical morphism
associated to Ξ. We claim that π(It) = π(I) ∩ Θt(π(I)). Indeed, note that π(It) =
π(I ∩ βt(I)) ⊂ π(I) ∩ Θt(π(I)). For the converse inclusion take a, b ∈ I and t ∈ G
such that π(a) = Θt(π(b)). Given c ∈ A, aι(c) = ι(π(a)δc
ee) = ι(Θt(π(a))δc
ee) =
βt(a)ι(c) ∈ It. As It is assimilative in I, a ∈ It and π(It) = π(I) ∩ Θt(π(I)).
It is straightforward to show that π◦ι(A) = π(I) is *-ideal of π(B), [π(I)] = π(B).
From the last paragraph and Remark 2.1 we know π(B) is non-degenerate and π ◦
ι : α → Θπ(I) is an isomorphism. As Θπ(I) = Θπ(B)π(I), (π(B), Θπ(B), π(I), π◦ ι)
is a non-degenerate and minimal *-globalization of α. Thus we have proved (1).
Lets prove the existence claim of statement (2). Let πΞ be the canonical mor-
phism associated to Ξ. To show that the image of πΞ, Im(πΞ), is Im(πΣ) note that
t (a, b) := ι−1(βt(ι(a))ι(b)) is the unique element
given (t, a, b) ∈ G, the element uα
of At such that αt(c)uα
t (a, b) = αt(ca)b (for all c ∈ At−1 ). Indeed, as ι : α → βI is
CONSTRUCTION OF ENVELOPING ACTIONS
7
an isomorphism and ι(At) = I ∩ βt(I) an ideal of B :
αt(c)uα
t (a, b) = ι−1(βt(ι(c))βt(ι(a))ι(b)) = ι−1(βt(ι(c))βt(ι(a))ι(b))
= ι−1(βt(ι(ca))ι(b)) = αt(ca)b.
Uniqueness follows from the non degeneracy of At = αt(At−1 ). Then uniqueness
implies ι−1(βt(ι(a))ι(b)) = uα
t (a, b) = κ−1(γt(κ(a))κ(b)). This last equality implies
that πΞ ◦ ι = πΣ ◦ κ. Then πΞ(I) = πΞ ◦ ι(A) = πΣ ◦ κ(A) = πΣ(J) and Im(πΞ) =
[πΞ(I)] = [πΣ(J)] = Im(πΣ).
From Lemma 2.12 we know ker(πΣ) = AnnR(C). Since C is non-degenerate this
implies πΣ is injective. Now define ΣρΞ := π−1
Σ ◦ πΞ. In case Ξ is non degenerate,
πΞ is injective and π−1
Ξ ◦ πΣ is the inverse of ΣρΞ. To close the circle note that
in case ΣρΞ is an isomorphism, Ξ is non degenerate because Σ is non degenerate.
Uniqueness of ΣρΞ follows from tree facts:
it is a morphism of *-global actions,
ΣρΞI = κ ◦ ι−1 and span{βt(I) : t ∈ G} = B.
To prove (3) assume Ξ is minimal, we will show that Ξ is non-degenerate. From
the hypothesis we get that, for all n ∈ N and t1, . . . , tn ∈ G, the ideal It1 +···+Itn is
assimilative in I. Assume b ∈ B satisfies Bb = {0}. Given that [I] = B, there exits
n ∈ N, t1, . . . , tn ∈ G and b1, . . . , bn ∈ I such that b = Pn
j=1 βtj (bj). We show b = 0
by induction in n. If n = 1 it follows that Ib1 = {0} and this implies b = 0 because
I is non-degenerate. For n > 1 we have βr(b)c = βr(bβr−1(c)) = 0, for all c ∈ I and
r ∈ G. Then Pn
(bj)c ∈
Pn−1
(j = 1, . . . , n − 1) such that
n tj
bn = Pn−1
j=1 b′
(I) = βtj (I), so there
j ∈ I such that βtn (bn) = Pn−1
j ).
(cid:3)
n we get bnc = −Pn−1
j ∈ It−1
) ⊂ βtn ◦ βt−1
j ). This implies b = Pn−1
exists b′′
By induction it follows that b = 0.
j=1 βrtj (bj)c = 0. With r = t−1
, for all c ∈ I. Thus there exists b′
j. Besides, βtn (b′
j) ∈ βtn (It−1
n tj
j=1 βtj (b′′
j=1 βtj (bj + b′′
j=1 It−1
j=1 βt−1
n tj
n tj
n tj
Before giving a sufficient condition for the existence of a *-globalization we prove
the following
Lemma 2.15. Let A be a *-algebra, I and J *-ideals of A, a ∈ I and b ∈ J such
that: I ∩ J and I + J are non-degenerate, Ia ⊂ J, Jb ⊂ I and ca = cb for all
c ∈ I ∩ J. Then a = b.
Proof. For all z ∈ I ∩ J and x ∈ I ∪ J we have xa, xb ∈ I ∩ J and zxa = zxb. Thus,
for all x ∈ I + J, xa = xb and this implies a = b.
(cid:3)
The next result is a combination of [3, Theorem 4.5] and [4, Theorem 3.1].
Theorem 2.16. Let α = ({At}t∈G,{αt}t∈G) be a *-partial action of G on A and
consider the conditions
(1) α has a *-globalization.
(2) For all (t, a, b) ∈ G × A × A there exists u ∈ At such that, for all c ∈ At−1 ,
αt(c)u = αt(ca)b.
Then (1) implies (2). In case for all s, t ∈ G the *-subalgebras At + As and At ∩ As
are non-degenerate and At is assimilative in A, (2) implies (1). Moreover, in this
last case α has a non-degenerate and minimal *-globalization.
Proof. The implication (1) ⇒ (2) is part of the proof of Proposition 2.14. For the
converse note that the element u in (2) is uniquely determined by (t, a, b) because
8
DAMI ´AN FERRARO
At is non-degenerate and αt(At−1 ) = At. The expression ut(a, b) will be used to
denote the element u given for (t, a, b) in (2).
Let AG be the *-algebra of function from G to A with point wise operations
and M (AG) it's multiplier algebra. Note A = Ae is non-degenerate, so AG is
non-degenerate and we can appeal to Remark 2.2, what we do without explicit
mention. We claim that there exists a unique injective homomorphism of *-algebras
ρ : A → M (AG) such that ρ(a)ft = ut(a, ft).
Given a ∈ A, to show ρ(a) is a multiplier with adjoint ρ(a∗) it suffices to show
that [ρ(a)f ]∗g = f ∗[ρ(a∗)g], for all f, g ∈ AG. The equality holds because for all
f, g ∈ AG, t ∈ G and c, d ∈ At we have [ρ(a)f ]∗gt, f ∗[ρ(a∗)g]t ∈ At and
d[ρ(a)f ]∗gtc = dut(a, ft)∗gtc = d[αt(αt−1 (c∗g∗t)a)ft]∗
= df ∗tαt(a∗αt−1 (gtc)) = αt(αt−1 (df ∗t)a∗αt−1(gtc))
= αt(αt−1 (df ∗t)a∗)gtc = df ∗[ρ(a∗)g]tc.
The uniqueness of elements ut(a, b) implies a 7→ ut(a, b) is linear, so ρ is linear.
To show ρ is multiplicative it suffices to prove, for all a, b, c ∈ A and t ∈ G, the
identity ut(a, ut(b, c)) = ut(ab, c). Note ut(a, ut(b, c)), ut(ab, c) ∈ At and for all
d ∈ At :
dut(a, ut(b, c)) = αt(αt−1 (d)a)ut(b, c) = αt(αt−1 (d)ab)c = dut(ab, c).
Thus ut(a, ut(b, c)) = ut(ab, c).
Assume ρ(a) = 0. Given b ∈ A let δb
on e and 0 elsewhere. Then 0 = ρ(a)δb
0 = cue(a, b) = αe(ca)b = cab. As A is non-degenerate, a = 0 and ρ is injective.
e ∈ AG be the function taking the value b
ee = ue(a, b). Given c ∈ A = Ae we have
With Θ being the canonical action of G on M (AG) set
B := span{Θt(ρ(A)) : t ∈ G}.
So B is a Θ-invariant *-sub algebra of M (AG) and the restriction of Θ to B, β, is
a *-global action. The proof will be completed once we show ρ(A) is an ideal of B,
that ρ is an isomorphism between α and βρ(A) and that B is non-degenerate.
βt(ρ(a))ρ(b) = ρ(ut(a, b)). From the definitions of β and ρ we obtain
Lets show that βt(ρ(A))ρ(A) ⊂ ρ(At), for all t ∈ G. It suffices to show that
βt(ρ(a))ρ(b)fr = urt(a, ur(b, fr)) and ρ(ut(a, b))fr = ur(ut(a, b), fr).
Putting c := fr it suffices to show urt(a, ur(b, c)) = ur(ut(a, b), c).
Note urt(a, ur(b, c)) ∈ Art, ur(ut(a, b), c) ∈ Ar,
Arturt(a, ur(b, c) = αrt (At−1r−1a) ur(b, c) ∈ Art ∩ Ar and
Arur(ut(a, b), c) = αr (Ar−1ut(a, b)) c ∈ αr(Ar−1 ∩ At) = Ar ∩ Art.
Besides, for z ∈ Art ∩ Ar :
zurt(a, ur(b, c)) = αrt(αt−1r−1(z)a)ur(b, c) = αr (αt(αt−1r−1(z)a)b) c
= αr (αr−1(z)ut(a, b)) c = zur (ut(a, b), c) .
Then Lemma 2.15 implies urt(a, ur(b, c)) = ur(ut(a, b), c).
Now the inclusion βt(ρ(A))ρ(A) ⊂ ρ(At) (valid ∀ t ∈ G) implies ρ(A) is an ideal
of B because
Bρ(A) = X
t∈G
βt(ρ(A))ρ(A) ⊂ ρ(A) and ρ(A)B = (Bρ(A))∗ ⊂ ρ(A).
CONSTRUCTION OF ENVELOPING ACTIONS
9
Take t ∈ G and a ∈ At−1 . To prove ρ(αt(a)) = βt(ρ(a)) is equivalent to prove
ur(αt(a), b) = urt(a, b), for all b ∈ A and r ∈ G. As before note that ur(αt(a), b) ∈
Ar, urt(a, b) ∈ Art,
Arur(αt(a), b) = αr (Ar−1αt(a)) b ∈ αr(Ar−1 ∩ At) = Ar ∩ Art and
Arturt(a, b) = αrt (At−1r−1a) b ∈ αrt (At−1r−1At−1 ) = Art ∩ Ar.
For z ∈ Ar ∩ Art :
zur(αt(a), b) = αr(αr−1(z)αt(a))b = αr(αt(αt−1r−1(z)a))b
= αrt(αt−1r−1(z)a)b = zurt(a, b).
Then ur(αt(a), b) = urt(a, b) by Lemma 2.15.
To finish the proof all we need to show is that ρ(At) = ρ(A) ∩ βt(ρ(A)). Note
ρ(At) = ρ(αt(At−1 )) = βt(ρ(At−1 )) ⊂ ρ(A) ∩ βt(ρ(A)). Now take T ∈ ρ(A) ∩
βt(ρ(A)). Then there exist a, b ∈ A such that T = ρ(a) = βt(ρ(b)). Thus, for all
f ∈ AG and r ∈ G :
ur(a, fr) = ρ(a)fr = βt(ρ(a))fr = urt(a, fr).
By replacing f with δb
e and r with e we obtain ue(a, b) = ut(a, b), for all b ∈ A. But
ue(a, b) = ab, then aA ⊂ At and this implies a ∈ At.
Up to this point we have shown that Ξ := (B, ΘB, ρ(A), ρ) is a minimal *-
globalization of α. Note that the canonical morphism associated to Ξ, πΞ : B →
M (AG), is just the natural inclusion B ⊂ M (AG) because πΞ ◦ ρ = ρ. Then Lemma
2.12 implies B is non-degenerate.
(cid:3)
With the previous Theorem we can extend Example 2.9 considerably. Let A be
a *-algebra and I an *-ideal of it. Recall that an orthogonal complement for I is a
*-ideal J ⊂ A such that A = I ⊕ J. In this situation IJ = 0 because IJ ⊂ I ∩ J.
Several remarks are in order. Firstly, if I is non degenerate the it has at most
one orthogonal complement. Secondly, if J is an orthogonal complement for I,
then A and I are non degenerate ⇔ I and J are non degenerate ⇔ A and J
are non degenerate. Thirdly, I is assimilative in A every time A and I are non
degenerate and I has an orthogonal complement. Finally, there are non degenerate
assimilative ideals without an orthogonal complement. As an example take the
complex sequences vanishing except on a finite set, considered as a *-ideal of the
complex sequences converging to zero.
Corollary 2.17. Let α = ({At}t∈G,{αt}t∈G) be a *-partial action of G on A.
If for all s, t ∈ G the *-subalgebras At + As and At ∩ As are non-degenerate and
At has an orthogonal complement in A, then α has a non degenerate and minimal
*-globalization.
Proof. From the hyphoteses it follows that A = Ae + Ae and At = At + At are non
degenerate. Besides At has a unique orthogonal complement in A, so it is assim-
ilative in A. Then it suffices to verify that α satisfies condition (2) from Theorem
2.16.
Take (t, a, b) ∈ G× A× A. Since At−1 has an orthogonal complement in A, there
exists a′ ∈ At−1 such that ca = ca′, for all c ∈ At−1. With u := αt(a′)b ∈ At we
have, for all c ∈ At−1 : αt(c)u = αt(c)αt(a′)b = αt(ca′)b = αt(ca)b.
(cid:3)
10
DAMI ´AN FERRARO
It may happen that a *-partial action has a minimal *-globalization with many
ideals At without an orthogonal complement. To exhibit an example consider let
C0(R) be the *-algebra of continuous function from R to C vanishing at ±∞. Let
β be the action of R on C0(R) given by βt(f )(r) = f (r − t). As an ideal of C0(R)
take A = C0(0, +∞) = {f ∈ C0(R) : fR\(0,+∞) ≡ 0} and let α be the restriction
of β to A. Then we can construct a non degenerate and minimal *-globalization of
α using β. Every ideal of A is non degenerate because in A the identity aa∗ = 0
implies a = 0, but the only *-ideals of A with an orthogonal complement are the
trivial ones. For example, with t > 0 the ideal At = A ∩ βt(A) = C0(t, +∞) does
not have an orthogonal complement in A.
2.2.1. About *-partial actions on commutative algebras. We close this section with
an adaptation of [1, Proposition 2.1] to *-algebras.
Proposition 2.18. Assume α is a *-partial action of G on the commutative *-
algebra A and σ = (B, β, I, ι) a minimal *-globalization. If At is non degenerate
for all t ∈ G then B is commutative.
Proof. Since B = span{βr(a) : r ∈ G, a ∈ I}, it suffices to prove that βr(a)βs(b) =
βs(b)βr(a), for all a, b ∈ I and r, s ∈ G. Fix a, b ∈ A and r, s ∈ G. With t := r−1s
we have βr(aβt(b)) = βr(a)βs(b) and βr(βt(b)a) = βs(b)βr(a). Then the proof will
be completed once we show that aβt(b) = βt(b)a.
Note that I is commutative, It := ι(At) non degenerate and aβt(b), βt(b)a ∈
I ∩ βt(I) = It. Then aβt(b) = βt(b)a if and only if caβt(b) = cβt(b)a, for all c ∈ It.
Take c ∈ It, using that I is commutative and a, b, c, βt−1(ac), βt−1 (c), βt(b)a ∈ I,
we obtain caβt(b) = βt(βt−1 (ac)b) = βt(bβt−1 (ac)) = βt(b)ac = cβt(b)a.
(cid:3)
3. Partial actions on C*-algebras
Assume A is a C*-algebra and G a topological group (we do not require it to be
locally compact nor Hausdorff).
Definition 3.1. The pair α = ({At}t∈G,{αt}t∈G) is a C*-partial action (of G on
A) if:
• It is a *-partial action of G on A.
• For all t ∈ G, At is closed in A.
• {At}t∈G is a continuous family, that is: for every open set U ⊂ A the set
• The function {(t, a) ∈ G×A : a ∈ At−1} → A, (t, a) 7→ αt(a), is continuous.
Morphisms between C*-partial actions are just morphism between *-partial ac-
tions. Recall from [1, Example 2.1] that the restriction of a C*-partial action to a
C*-ideal is a C*-partial action.
GU := {t ∈ G : At ∩ U 6= ∅} is open in G.
Definition 3.2. A C*-globalization of the C*-partial action α, of G on A, is a
4-tuple Ξ = (B, β, I, ι) where: B is a C*-algebra, β is a C*-global action of G on
B, I is a C*-ideal of B and ι : α → βI is an isomorphism of *-partial actions.
Note we do not require ι to be a homeomorphism, this is automatic because
every *-homomorphism between C*-algebras is contractive and has closed range.
In case α is a C*-partial action, it is also a *-partial action and every C*-
globalization of α is a *-globalization of α. Although not every *-globalization
of α is a C*-globalization, the existence of a *-globalization implies the existe
CONSTRUCTION OF ENVELOPING ACTIONS
11
of a C*-globalization. We will prove this claim in two steps: first we show the
group's topology is irrelevant to globalize C*-partial actions, then we construct a
C*-globalization from a *-globalization.
In the next statement Gdis denotes the group G with the discrete topology.
Lemma 3.3. Let G be a topological group, β a C*-global action of Gdis on B and
I an ideal of B such that [I] := span{βt(I) : t ∈ G} is dense in B. Set α := βI ,
which is a C*-partial action of Gdis on I. Then β is a C*-partial action of G if and
only if α is a C*-partial action of G.
Proof. The direct implication is [1, Example 2.1]. For the converse note that (as
each βr is an isometry) G × B → B, (t, b) 7→ βt(b), is continuous if and only if for
every b ∈ B the function evb : G → B, t 7→ βt(b), is continuous at e. Moreover,
U := {b ∈ B : evb is continuous at e} is a closed β-invariant subspace of B, thus all
we need to show is that I ⊂ U.
Fix a ∈ I. From [1, Lemma 2.1] (using the family of ideals {βt(I)}t∈G) we get,
for all t ∈ G, that
kβt(a) − ak = sup{k(βt(a) − a)βr(b)k : r ∈ G, b ∈ I, kbk ≤ 1}
= sup{k(βr−1t(a) − βr−1(a))bk : r ∈ G, b ∈ I, kbk ≤ 1}
= sup{k(βrt(a) − βr(a))bk : r ∈ G, b ∈ I, kbk ≤ 1}.
Fix r ∈ G and b ∈ I with kbk ≤ 1. Note that (βrt(a)−βr(a))b = βrt(a)b−βr(a)b ∈
βrt(I)I + βr(I)I = Irt + Ir. Thus, if {vλ}λ∈Λ is an approximate unit of Irt + Ir, we
have k(βrt(a) − βr(a))bk = limλ k(βrt(a) − βr(a))bvλk.
r such that
vλ = c + d. This implies kck,kdk ≤ kvλk ≤ 1. One one hand, if {wµ}µ∈M is an
approximate unit of It, we have
Given λ ∈ Λ, vλ ∈ (Irt + Ir)+ and there exists c ∈ I +
rt and d ∈ I +
k(βrt(a) − βr(a))bck = kaβt−1r−1(bc) − βt−1(a)βt−1r−1(bc)k
= lim
µ kaβt−1r−1(bc) − βt−1(wµ)βt−1(a)βt−1r−1(bc)k
≤ lim sup
µ
ka − βt−1 (wµa)k = lim sup
µ
ka − αt−1 (wµa)k.
To construct a particular approximate unit consider Mt := {c ∈ I +
with it's natural order, then {µ}µ∈Mt is an approximate unit of It and
t : kak < 1}
k(βrt(a) − βr(a))bck ≤ lim
With s = rt by symmetry we get
µ∈Mt
sup{ka − αt−1 (νa)k : ν ∈ Mt, ν ≥ µ} =: C(t).
k(βrt(a) − βr(a))bdk = k(βst−1(a) − βs(a))bdk ≤ C(t−1).
Putting all together we obtain kβt(a) − ak ≤ C(t) + C(t−1). All we need to show is
that limt→e C(t) = 0.
Take ε > 0. As α is a partial action of G on I there are neighbourhoods V ⊂ G
and W ⊂ B of e and a, respectively, such that for s ∈ V and b ∈ Is−1 ∩ W we
have kαs(b) − ak < ε/2. Take δ > 0 such that B(a, δ) ⊂ W. Then U := {r ∈
G : B(a, δ/2) ∩ Ir 6= ∅} is an open set containing e because {It}t∈G is a continuous
family. For r ∈ U ∩ V −1 there exists b ∈ Ir ∩ B(a, δ/2), so limµ∈Mr ka − µak =
dist(a, Ir) ≤ δ/2. Take µr ∈ Mr such that ka− νak < δ for all ν ∈ Mr with ν ≥ µr.
12
DAMI ´AN FERRARO
Then for r ∈ U ∩ V −1 we have C(r) < ε because for all ν ∈ Mr with ν ≥ µr the
inequality ka − αr−1(νa)k ≤ ε/2 holds and implies
sup{ka − αr−1 (ν′a)k : ν′ ∈ Mr, ν′ ≥ ν} ≤ ε/2 < ε.
(cid:3)
Corollary 3.4. Let α = ({At}t∈G,{αt}t∈G) be a C*-partial action and use the
expression αdis to denote α as a C*-partial action of Gdis. Then α has a C*-
globalization if and only if αdis has a C*-globalization.
Proof. As mentioned before the direct implication in immediate. For the converse
assume αdis has a C*-globalization Ξ = (B, β, I, ι). Without loss of generality we
may assume Ξ is minimal (an enveloping action in the sense of F. Abadie [1]) this is
[I] = B. Note βI is a C*-partial action of G because it is isomorphic (as a *-partial
action) to α. Then the previous lemma implies β is a C*-global action of G. Thus
Ξ is a C*-globalization of α.
(cid:3)
As promised before we now construct a C*-globalization from a *-globalization.
Theorem 3.5. Let α = ({At}t∈G,{αt}t∈G) be a C*-partial action. Then the
following are equivalent:
(1) α has a C*-globalization.
(2) α has a *-globalization.
(3) For all (t, a, b) ∈ G × A × A there exists u ∈ At such that, for all c ∈ At−1 ,
αt(c)u = αt(ca)b.
Proof. We already know (1) implies (2). On C*-algebras the condition a∗a = 0
implies a = 0, thus every C*-ideal of a C*-algebra is non-degenerate. The existence
of approximate units implies every closed ideal is assimilative. Besides the sum4 of
two C*-ideals of a C*-algebra is, again, a C*-ideal. Then (2) and (3) are equivalent
by Theorem 2.16.
To show (2) implies (1), without loss of generality, we assume G is discrete
(Corollary 3.4). From the previous paragraph and Proposition 2.14 we know α
has a non degenerate and enveloping *-globalization Ξ = (B, β, I, ι). Note I is a
C*-algebra because it is isomorphic (as a *-algebra) to a C*-algebra. Moreover,
βI is a C*-partial action isomorphic to α. Then it suffices to show that βI has a
C*-globalization and me may think I = A, ι is inclusion of A in B and α = βI .
Let π be the canonical morphism associated to Ξ and Θ the canonical action
of G on M (AG). From Proposition 2.14 we know π is injective and, as we think
A ⊂ B, π(b)fr = βt(b)fr.
structure inherited from AG and the supremum norm. Define
b ) ⊂ AG
b }.
The set of bounded functions from G to A, AG
C := {T ∈ M (AG) : T (AG
b , is a C*-algebra with the *-algebra
b ) ∪ T ∗(AG
Note C is Θ invariant and that to show π(B) ⊂ C it suffices to prove that
b and r ∈ G and
π(A) ⊂ C. This last inclusion holds because given a ∈ A, f ∈ AG
an approximate unit of Ar−1, {vλ}λ∈Λ, we have βr(a)fr ∈ Ar and
kπ(a)frk = kβr(a)frk = lim
λ kαr(vλ)βr(a)frk = lim
λ kαr(vλa)frk ≤ kakkfk.
4Without closure.
CONSTRUCTION OF ENVELOPING ACTIONS
13
b → AG
Consider M (AG
b ) as a C*-algebra and let ρ : π(B) → M (AG
b ) be defined as
ρ(T )f = T f. To show ρ is injective it suffices to show that ρ ◦ π is injective. Take
b ∈ B such that ρ ◦ π(b) = 0. There are a1, . . . , an ∈ A and t1, . . . , tn ∈ G such
that b = Pn
j=1 βtj (aj ). Given r ∈ G and c ∈ A we have bβr(c) = βr(βr−1(b)c) =
βr(ρ ◦ π(b)δc
r−1r−1 ) = 0. Thus bB = {0} and this implies b = 0.
b ) → M (AG
b as ψt(f )r = frt and Ψt : M (AG
Given t ∈ G set ψt : AG
b ) as
Ψt(T ) = ψt ◦ T ◦ ψt−1 . Then Ψ is a C*-global action of G on M (AG
b ); ρ ◦ π : β → Ψ
is a morphism of C*-partial actions and the closure of ρ◦ π(B), D, is a Ψ-invariant
C*-subalgebra of M (AG
b ). Name γ the restriction of Ψ to D. Note J := ρ(π(A))
is a C*-ideal of D because ρ ◦ πA has closed range and J is an ideal of ρ ◦ π(B).
From Remark 2.3 we get γJ = ΨDρ◦π(A) = Ψρ◦π(A) = Ψρ◦π(B)ρ◦π(A). Besides,
ρ : Θπ(B) → Ψρ◦π(B) is an isomorphism. Then ρ◦ πA : α → ΨJ is an isomorphism
and, consequently, (D, γ, J, ρ ◦ πA) is a C*-globalization of α.
(cid:3)
Condition (3) asserts the existence of a certain element u for each (t, a, b) ∈
G × A × A. That element is the limit of the net given in (2) below.
Proposition 3.6. Let α = ({At}t∈G,{αt}t∈G) a C*-partial action. Then the fol-
lowing are equivalent.
(1) α has a C*-globalization.
(2) There exists U, V ⊂ A such that: (i) span AU = span V A = A and (ii) for all
(t, a, b) ∈ G × U × V there exists an approximate unit of At−1 , {vλ}λ∈Λ, such
that {αt(vλa)b}λ∈Λ is a Cauchy net (or converges).
Proof. In case (1) holds set U = V = A and take (t, a, b) ∈ G × A × A. Let u be
the element given in (2) of the previous theorem and {vλ}λ∈Λ and approximate
unit of At−1 . Then {αt(vλ)}λ∈Λ is an approximate unit of At and {αt(vλ)u}λ∈Λ =
{αt(vλa)b}λ∈Λ converges to u.
Now assume (2) is true. Take (t, a, b) ∈ G × A × A for which there exists an
approximate unit, {vλ}λ∈Λ, such that {αt(vλa)b}λ∈Λ converges to u. We claim that
given any other approximate unit of At−1 , {wµ}µ∈M , the net {αt(wµa)b}µ∈M =
{αt(wµ)u}µ∈M converges to u. Indeed
lim
λ
αt(vλa)b = lim
µ
αt(wµ) lim
λ
αt(vλa)b = lim
µ
lim
λ
αt(wµvλa)b = lim
µ
αt(wµa)b.
Then u is determined by (t, a, b), so we denote it ut(a, b). Moreover, if c ∈ At−1
then
αt(c)ut(a, b) = lim
λ
αt(c)αt(vλa)b = lim
λ
αt(cvλa)b = αt(ca)b.
Take a ∈ U, b ∈ V and c, d ∈ A and any approximate unit of At−1 , {vλ}λ∈Λ.
Then {αt(vλca)bd}λ∈Λ converges to αt(cαt−1 (ut(a, b)d)) because, for all λ ∈ Λ,
αt(vλca)bd = αt(vλc)ut(a, b)d = αt(vλcαt−1(ut(a, b)d))
and limλ αt(vλcαt−1 (ut(a, b)d)) = αt(cαt−1 (ut(a, b)d)).
The conclusions of the previous paragraphs imply that (2) holds if we replace U
for AU and V for V A. From now on we assume AU = U and V A = V.
Given a, b ∈ U, c ∈ V and λ ∈ C note that ut(a, c) + λut(b, c) ∈ At and for
all d ∈ At−1 : αt(d)(ut(a, b) + λut(b, c)) = αt(d(a + λb))c. In case a ∈ U and
b, c ∈ V : αt(d)(ut(a, b) + λut(a, c)) = αt(da)(b + λc). Then, for all t ∈ G, there
exists a unique bilinear function ut : span U × span V → At such that for all a ∈
span U, b ∈ span V and c ∈ At−1 : αt(c)ut(a, b) = αt(ca)b. Note that kut(a, b)k2 =
14
DAMI ´AN FERRARO
kαt(αt−1 (ut(a, b)∗)a)bk ≤ kut(a, b)kkakkbk, then kut(a, b)k ≤ kakkbk. So there is a
unique continuous bilinear function vt : A × A → At extending ut, for every t ∈ G.
Given (t, a, b) ∈ G × A × A and c ∈ At−1 take sequences {an}n∈N ⊂ span U and
{bn}n∈N ⊂ span V converging to a and b, respectively. Then
αt(c)vt(a, b) = lim
n
αt(c)vt(an, bn) = lim
n
αt(can)bn = αt(ca)b.
This shows (2) implies (3) of the previous theorem, so α admits a C*-globalization.
(cid:3)
Now we use our criterion to give some conditions were a C*-partial action can
be globalized.
On a unital C*-algebra, A, for every unital ideal I there exists a unique ideal
J such that A = I ⊕ J and IJ = {0}. In general, even for non unital A, we say
a C*-ideal I has an orthogonal complement if there exists a C*-ideal J such that
A = I ⊕ J and IJ = {0}.
Corollary 3.7. Let α = ({At}t∈G,{αt}t∈G) be a C*-partial action.
t ∈ G, At has an orthogonal complement then α has a C*-globalization.
Proof. Given (t, a, b) ∈ G × A × A let J be the orthogonal complement of At−1 .
Then there are a1 ∈ At−1 and a2 ∈ J such that a = a1 + a2. Note for all c ∈ At−1 ,
ca = ca1. Thus u := αt(a1)b ∈ At and αt(c)u = αt(ca1)b = αt(ca)b.
If, for all
(cid:3)
The next corollary is a version for C*-algebras of [3, Theorem 4.5]. The proof is
omitted because it follows form the cited theorem and the previous corollary.
Corollary 3.8. Let α = ({At}t∈G,{αt}t∈G) a C*-partial action with A unital.
Then α has a C*-globalization if and only if At is a unital algebra, for all t ∈ G.
There are certain *-homomorphisms that force the existence of C*-enveloping
actions. For partial actions on locally compact and Hausdorff spaces the next
result can be showed using [1, Proposition 2.1] instead of Theorem 3.5 (in that
case the homomorphism φ defines a continuous function between the spectra of the
algebras).
Corollary 3.9. Let α = ({At}t∈G,{αt}t∈G) and β = ({Bt}t∈G,{βt}t∈G) be C*-
partial actions. If α has a C*-globalization and there exists a *-algebra's homomor-
phism φ : A → M (B) such that:
(1) for all t ∈ G, span{φ(a)b : a ∈ At, b ∈ B} = Bt and
(2) for all t ∈ G, a ∈ At−1 and b ∈ Bt−1, φ(αt(a))βt(b) = βt(φ(a)b);
then β has a C*-globalization.
Proof. By Cohen-Hewitt's Theorem Bt = {φ(a)b : a ∈ At, b ∈ B}, for all t ∈ G. The
same theorem implies At = {ab : a, b ∈ At}, then Bt = {φ(a)b : a ∈ At, b ∈ Bt}.
1)c∗
1 = φ(a∗
1
and b2 = φ(a2)c2. Now let u be the element given for (t, a1, a2) by condition (3) of
Theorem 3.5. Then c1φ(u)c2 ∈ Bt because u ∈ At.
φ(e∗)d′∗. Then
Given d ∈ Bt−1 choose d′ ∈ Bt−1 and e ∈ At−1 such that βt−1(c∗
Given (t, b1, b2) ∈ G×B×B take a1, a2 ∈ A and c1, c2 ∈ B such that b∗
1βt(d∗)) =
βt(d)c1φ(u)c2 = βt(d′)φ(αt(e)u)c2 = βt(d′)φ(αt(ea1)a2)c2
= βt(d′)φ(αt(ea1))φ(a2)c2 = βt(d′φ(e)φ(a1))b2 = βt(db1)b2.
The rest follows directly from the previous theorem.
(cid:3)
CONSTRUCTION OF ENVELOPING ACTIONS
15
To close this section we give an example to show condition (1) from the previous
statement can not we weakened to (1') Bt = span{φ(a)b : a ∈ At, b ∈ Bt}, for all
t ∈ G.
Let G = Z2 = {1,−1} (with multiplicative notation) A = B = C[0, 1]. As
α consider the trivial global action of G on A. To give a partial action of G on
B, β, we just need to specify β−1, which will be the identity on C0([0, 1)). For
φ : C[0, 1] → M (C[0, 1]) = C[0, 1] just take the identity. Note (1') and (2) are
satisfied, α has a C*-globalization but β does not.
4. Partial action on equivalence bimodules
In [1] F. Abadie defines Morita equivalence of C*-partial actions using partial ac-
tions on positive C*-trings. Recall that a positive C*-trings are exactly equivalence
bimodules [1, 9, 8]. Here we adopt the terminology of equivalence bimodules.
In this last section we give a necessary and sufficient condition for the existence of
a globalization of a partial action on an equivalence bimodule and, as a consequence,
we obtain the uniqueness of Morita enveloping actions (as was shown in [1]).
We adopt the terminology of [8] and agree that "AXB is an equivalence bimodule"
means "X is an A-B-equivalence bimodule". Take equivalence bimodules AXB and
a CYD. A function φ : X → Y is an Hb-homomorphism if it is linear and for
all x, y, z ∈ X , φ(xhy, ziB) = φ(x)hφ(y), φ(z)iD . Such functions are contractive
and, with the previous notation, there exist unique *-homomorphisms lφ : A →
C and φr : B → D such that, for all x, y ∈ X , φl(Ahx, yi) = Chφ(x), φ(y)i and
φr(hx, yiB) = hφ(x), φ(y)iD . The proof of these facts can be found in [1].
Given a C*-ideal I of A, by Cohen-Hewitt IX := {ax : a ∈ I, x ∈ X} is a closed
submodule of X . We denote I B the C*-ideal of B induced by I (or IX ) through X ,
that is I B = span{hu, viB : u, v ∈ IX}. In a similar way we define, for a C*-ideal
J of B, X J and AJ . Recall that AI B = I and J = AJ B.
For a closed subspace Z of X the following are equivalent: (i) there exists a
C*-ideal I of A such that Z = IX , (ii) there exists a C*-ideal J of B such that
Z = X J, (iii) XhZ,XiB ⊂ Z and (iv) AhX ,ZiX ⊂ Z. If these conditions are
satisfied then AZ := span AhZ,Zi and ZB are C*-ideals, Z = AZX = X Z B and
we say Z is an ideal of X . Every ideal Z of X is an AZ − ZB-equivalence bimodule.
Definition 4.1. The pair γ = ({γt}t∈G,{Xt}t∈G) is a Hb-partial action if:
• X is an (A-B-)equivalence bimodule and G a topological group.
• γ is a set theoretic partial action of G on X .
• {Xt}t∈G is a continuous family of ideals of X .
• For all t ∈ G, γt : Xt−1 → Xt is an Hb-homomorphism.
• The function {(t, x) ∈ G×X : x ∈ Xt−1} → X , (t, x) 7→ γt(x), is continuous.
Assume δ is a Hb-partial action G on CYD. We say φ : γ → δ is a Hb-morphism
if it is a Hb-homomorphism from X to Y which is also a morphism of set theoretic
partial actions.
From [1] we know there are unique C*-partial actions, α = ({αt}t∈G,{At}t∈G)
and β = ({βt}t∈G,{Bt}t∈G) , such that
• For all t ∈ G, At = AXt and Bt = XtB.
• For all t ∈ G and x, y ∈ Xt−1 :
αt(Ahx, yi) = Ahγt(x), γt(y)i and βt(hx, yiB) = hγt(x), γt(y)iB.
16
DAMI ´AN FERRARO
We will call α the left side of γ and β the right side of γ and will denote them
lγ and γr, respectively.
Example 4.2. Every C*-partial action, α, on a C*-algebra A is a Hb-partial action
on AAA. Besides α = lα = αr.
Example 4.3. Given a Hb-global action of G on AXB, γ, and an ideal Y of X ; the
restriction γY is a Hb-partial action on the AY -Y B-equivalence bimodule Y. In this
case l(γY ) = (lγ)AY and (γY )r = γrY B.
With restrictions of global actions, on one hand, and isomorphisms on the other
we define globalizations.
Definition 4.4. Let γ be an Hb-partial action of G on X . A globalization of γ is a
4−tuple Ξ = (Y, δ,Z, ι) such that: Y is an equivalence bimodule, δ is an Hb-global
action of G on Y, Z is an ideal of Y and ι : γ → δZ is an isomorphism of Hb-partial
actions. In case [Z] := span{δt(Z) : t ∈ G} is dense in Y we say Ξ is a minimal
globalization.
The nexus between Hb-partial actions and C*-partial actions is the linking par-
tial action [1]. To describe this action we start with an Hb-partial action of a group
G on AXB, γ, and set α := lγ and β := γr. The linking algebra of X is the algebra of
generalized compact operators of the A-Hilbert module X ⊕ A, L(X ) = K(X ⊕ A).
In matrix representation L(X ) = (cid:0) A X
eX B(cid:1) , with eX the B-A-equivalence bimodule
adjoint to X .
The linking partial action of γ, L(γ) = ({L(γ)t}t∈G,{L(X )t}t∈G) , is the unique
C*-partial action such that, for all t ∈, L(X )t = L(Xt) and
ey b ) = (cid:16) αt(a) γt(x)
]γt(y) βt(b)(cid:17) ∀ x, y ∈ Xt−1 , a ∈ At−1 and b ∈ Bt−1.
L(γ)t ( a x
In case δ is an Hb-partial action of G on Y and π : γ → δ is an isomorphism, the
morphism L(π) : L(γ) → L(δ) defined as
L(π) ( a x
ey b ) = (cid:16) lπ(a) π(x)
gπ(y) πr(b)(cid:17) ,
is an isomorfphism with inverse L(π−1).
Proposition 4.5. If Ξ = (CYD, δ,Z, ι) is an enveloping globalization of γ then
• (L(Y), L(δ), L(Z), L(ι)) is a C*-enveloping globalization of L(γ).
• (C, lδ, CZ, lι) is a C*-enveloping globalization of lγ.
• (D, δr, Z D, ιr) is a C*-enveloping globalization of γr.
Proof. It is straightforward and is left to the reader.
(cid:3)
In the same way equivalence bimodules are constructed form C*-algebras and
projections, Hb-partial actions are constructed form C*-partial actions and equi-
variant projections.
Theorem 4.6. Let α be a C*-partial action of G on A and p ∈ M (A) a projection
such that span ApA = A and αt(pa) = pαt(a), for all t ∈ G and a ∈ At−1 . If
X := (1 − p)Ap, C := (1 − p)A(1 − p) and D := pAp then
(1) X is a C-D-equivalence bimodule,
(2) X is α-invariant,
(3) the restriction of α to X , γ, is an Hb-partial action.
CONSTRUCTION OF ENVELOPING ACTIONS
17
(4) C and D are α-invariant, lγ = αC and γr = αD.
Moreover, α is isomorphic (as a C*-partial action) to L(γ) and α has a C*-
globalization if and only if γ has an Hb-globalization.
Proof. The proof of claims (1)-(4) are left to the reader. Besides, the usual iden-
tification of A with L(X ) is an isomorphism of C*-partial actions between α and
L(γ). Then, in case γ has an Hb-globalization, L(γ) has a C*-globalization and this
implies α has a C*-globalization.
To prove the converse assume α has a C*-globalization. To prove γ has a glob-
alization we can assume, without loss of generality, that there exists a C*-partial
action, β of G on B, such that A is an ideal of B, α = βA and [A] = B.
The key claim is that there exists a unique projection p ∈ M (B) such that
βt(pb) = pβt(b) and pa = pa, for all t ∈ G, b ∈ B and a ∈ A. To prove the claim it
suffices to show that for all t1, . . . , tn ∈ G and a1, . . . , an ∈ A
nX
j=1
k
nX
j=1
βtj (paj)k ≤ k
βtj (aj)k.
(1)
From [1, Lemma 2.1] we conclude that it is enough to show that
nX
j=1
k
βtj (paj)βr(b)k ≤ k
nX
j=1
βtj (aj)k,
for all r ∈ G and b ∈ A with kbk < 1.
Take r ∈ G and b ∈ B as before. Note pajβt−1
nX
βtj (paj)βr(b)k = k
βtj (pajβt−1
j r ∈ At−1
nX
j r(b))k = k
k
j=1
j r so that
βr−1tj (pajβt−1
j r(b))k
j=1
j=1
j=1
nX
nX
nX
nX
j=1
j=1
= k
≤ k
≤ k
αr−1tj (pajβt−1
pαr−1tj (aj βt−1
j r(b))k
αr−1tj (aj βt−1
βr−1tj (ajβt−1
j r(b))k
nX
j r(b))k = k
nX
j r(b))k = k
j=1
j=1
βtj (aj)k.
Set Y := (1 − p)Ap ⊂ B, note that X ⊂ Y and define ι : X → Y as the canonical
inclusion. Remark 2.3 implies βYX = βX = βAX = αX = γ, then (Y, βY ,X , ι)
is a globalization of γ.
(cid:3)
Corollary 4.7. An Hb-partial action has a globalization if and only if it's linking
partial action has a C*-globalization.
Proof. Let γ be a partial action of G on the A-B-equivalence bimodule X . The
thesis follows easily from the previous theorem with α = L(γ) and p = ( 1 0
0 0 ) because
L(γ)pL(X )(1−p) is isomorphic to γ.
(cid:3)
As a consequence of the Corollary we get that the group's topology does not
affects the existence of Hb-globalizations because it does nor affects the existence
of C*-globalizations. This conclusion can be equally derived from our last result,
18
DAMI ´AN FERRARO
which also implies that a C*-partial action has a C*-globalization if and only if it
has a Hb-globalization.
Theorem 4.8. An Hb-partial action has an Hb-globalization if and only if it's left
and right sides have C*-globalizations.
Proof. The direct implication follows from Proposition 4.5. For the converse assume
γ is an Hb-partial action of the topological group G on AXB. Assume α := lγ and
β := γr have C*-globalizations. Given t ∈ G set At := AXt and Bt := XtB.
It suffices to show L(γ) has a C*-globalization, for which we use Proposition
3.6 with U = V = {(cid:0) 0 x
ey 0(cid:1) : x, y ∈ X}. Take (t, ξ, η) ∈ G × U × V. Let {vλ}λ∈Λ
be an approximate unit of At−1 and {wµ}µ∈M one of Bt−1. Consider K = Λ × M
with the ordered (λ, µ) ≤ (λ′, µ′) iff λ ≤ λ′ and µ ≤ µ′. Given κ = (λ, µ) ∈ K set
dκ := (cid:16) vλ 0
For κ = (λ, µ), ξ = (cid:0) 0 x
0 wµ(cid:17) . Then {dκ}κ∈K is an approximate unit of L(X )t.
ev 0 ) we have
ey 0(cid:1) and η = ( 0 u
L(γ)t(dκξ)η = (cid:16) hγt(vλx),vil
0
hγt(ywµ),uir(cid:17) .
0
By Cohen-Hewitt there exists b, c ∈ B and z, w ∈ X such that x = zb and
v = wc. As β has a C*-globalization {βt(wµb)c∗}µ∈M converges to an element
p ∈ Bt−1. Then
hγt(vλx), vil = hγt(vλazb), wcil = lim
µ
ν hγt(vλzwµfνb)c∗, wil
lim
= lim
µ
= lim
ν hγt(vλzwµ)βt(fν b)c∗, wil = lim
lim
µ hγt(vλzwµ)p, wil
µ hγt(vλzwµβt−1(p)), wil = hγt(vλzβt−1(p)), wil.
Note zβt−1(p) ∈ Xt−1 , so that limλhγt(vλx), vil = hγt(zβt−1 (p)), wil.
vergent.
By symmetry {hγt(ywµ), uir}µ∈M is convergent. Then {L(γ)t(dκξ)η}κ∈K is con-
(cid:3)
References
1. Fernando Abadie, Enveloping actions and Takai duality for partial actions, J. Funct. Anal.
197 (2003), no. 1, 14 -- 67.
2. Wagner Cortes and Miguel Ferrero, Globalization of partial actions on semiprime rings, Con-
temporary Mathematics 499 (2009), 27.
3. M. Dokuchaev and R. Exel, Associativity of crossed products by partial actions, enveloping
actions and partial representations, Trans. Amer. Math. Soc. 357 (2005), no. 5, 1931 -- 1952.
MR 2115083 (2005i:16066)
4. Michael Dokuchaev, ´Angel Del R´ıo, and Juan Sim´on, Globalizations of partial actions on
nonunital rings, Proceedings of the American Mathematical Society 135 (2007), no. 2, 343 --
352.
5. Ruy Exel, Circle actions on C*-algebras, partial automorphisms, and a generalized Pimsner-
Voiculescu exact sequence, J. Funct. Anal. 122 (1994), no. 2, 361 -- 401. MR 1276163 (95g:46122)
, Twisted partial actions: a classification of regular C*-algebraic bundles, Proc. London
6.
Math. Soc. (3) 74 (1997), no. 2, 417 -- 443.
7. Kevin McClanahan, K-theory for partial crossed products by discrete groups, J. Funct. Anal.
130 (1995), no. 1, 77 -- 117. MR 1331978 (96i:46083)
8. Marc A. Rieffel, Morita equivalence for operator algebras, Operator algebras and applications,
Part I (Kingston, Ont., 1980), Proc. Sympos. Pure Math., vol. 38, Amer. Math. Soc., Provi-
dence, R.I., 1982, pp. 285 -- 298.
9. Heinrich Zettl, A characterization of ternary rings of operators, Advances in Mathematics 48
(1983), no. 2, 117 -- 143.
CONSTRUCTION OF ENVELOPING ACTIONS
19
Departamento de Matem´atica y Estad´ıstica del Litoral, Universidad de la Rep´ublica,
Gral. Rivera 1350. Salto. Uruguay.
E-mail address: [email protected]
|
1102.5430 | 3 | 1102 | 2013-09-17T05:43:34 | Cyclic Hilbert spaces and Connes' embedding problem | [
"math.OA",
"math.FA"
] | Let $M$ be a $II_1$-factor with trace $\tau$, the linear subspaces of $L^2(M,\tau)$ are not just common Hilbert spaces, but they have additional structure. We introduce the notion of a cyclic linear space by taking those properties as axioms. In Sec.2 we formulate the following problem: "does every cyclic Hilbert space embed into $L^2(M,\tau)$, for some $M$?". An affirmative answer would imply the existence of an algorithm to check Connes' embedding Conjecture. In Sec.3 we make a first step towards the answer of the previous question. | math.OA | math |
CYCLIC HILBERT SPACES AND CONNES'
EMBEDDING PROBLEM
Valerio CAPRARO and Florin R ADULESCU
May 5, 2018
Abstract. Let M be a II1-factor with trace τ , the linear subspaces of L2(M, τ ) are
not just common Hilbert spaces, but they have additional structure. We introduce the
notion of a cyclic linear space by taking those properties as axioms. In Sec.2 we formulate
the following problem: "does every cyclic Hilbert space embed into L2(M, τ ), for some
M ?". An affirmative answer would imply the existence of an algorithm to check Connes'
embedding Conjecture. In Sec.3 we make a first step towards the answer of the previous
question.
Contents
1 Cyclic Hilbert spaces
2 Relation with Connes' embedding conjecture
3 Extension of cyclic vector spaces
4 Problems we were not able to solve
5 Acknowledgement
1 Cyclic Hilbert spaces
1
3
4
10
11
Let M be a finite factor with unique normalized trace τ and let L2(M, τ ) be the
Hilbert space obtained by taking the closure of the vector space M with respect to
1
the inner product (x, y) = τ (y∗x)
2 . Consider a finite-dimensional real Hilbert subspace
H ⊆ Msa ⊆ L2(M, τ ) containing the identity. Observe that H is not just a common
Hilbert space, but it has additional structure.
1
Proposition 1. The mapping ≪ a ⊗ b, c ⊗ d ≫
form on (H ⊗ H) ⊗ C and the following properties are satisfied
.
= τ (abdc) is a bilinear hermitian positive
1. the mappings v → v ⊗ 1 and v → 1 ⊗ v are isometric embeddings, i.e.
(v, v) =≪ v ⊗ 1, v ⊗ 1 ≫=≪ 1 ⊗ v, 1 ⊗ v ≫
2. ≪ ·,· ≫ is cyclic in the following sense
≪ a ⊗ b, c ⊗ d ≫=≪ c ⊗ a, d ⊗ b ≫
3. ≪ ·,· ≫ is self-adjoint in the following sense
≪ a ⊗ b, c ⊗ d ≫= ≪ b ⊗ a, d ⊗ c ≫
4. ≪ ·,· ≫ verifies the following property
≪ a ⊗ b, c ⊗ d ≫=≪ b ⊗ d, a ⊗ c ≫
5. The mapping JH : (H ⊗ H)⊗ C → (H ⊗ H)⊗ C defined by setting JH (a⊗ b) = b⊗ a
is an isometric involution, i.e.
(a) JH (JH (a ⊗ b)) = a ⊗ b
(b) ≪ JH(a ⊗ b), JH (a ⊗ b) ≫=≪ a ⊗ b, a ⊗ b ≫
In this article we want to consider Hilbert spaces which verify these five additional
properties. Before giving the definition let us observe that properties 2. and 3. together
imply properties 4. and 5. Indeed
Lemma 2. Let (H, (·,·),≪ ·,· ≫) be a Hilbert space equipped with a bilinear positive
hermitian form ≪ ·,· ≫ on (H ⊗ H) ⊗ C. If ≪ ·,· ≫ verifies 2. and 3. then it verifies
also 4. and 5.
2
Proof. Suppose 2. and 3. are verified. Applying hermitianity, 2. and hermitianity again
we get 4. Indeed
≪ a ⊗ b, c ⊗ d ≫= ≪ c ⊗ d, a ⊗ b ≫ = ≪ a ⊗ c, b ⊗ d ≫ =
On the other hand, applying 3. and hermitianity, we get 5. Indeed
=≪ b ⊗ d, a ⊗ c ≫
≪ JH(a ⊗ b), JH (a ⊗ b) ≫=≪ b ⊗ a, b ⊗ a ≫= ≪ a ⊗ b, a ⊗ b ≫ =
=≪ a ⊗ b, a ⊗ b ≫
Notice that we have not used the completeness with respect to (·,·). Thus we can give
the following
Definition 3. A cyclic pre-Hilbert space is a quadruple (V, (·,·), 1,≪ ·,· ≫), where
(V, (·,·)) is a real pre-Hilbert space, 1 ∈ V is a pointed vector such that (1, 1) = 12 = 1
and ≪ ·,· ≫ is a bilinear complex-valued, hermitian positive form on (V ⊗V )⊗ C verifying
properties 1.,2. and 3. (and, consequently, 4. and 5.).
2 Relation with Connes' embedding conjecture
We have begun studying cyclic spaces motivated by Connes' embedding conjecture. Before
explaining how they are related to each other, let us briefly recall Connes' embedding
conjecture. Let R be the hyperfinite II1 factor (with unique trace denoted by τ ) and
let ω ∈ β(N) \ N be a free ultrafilter on the natural number. One can construct the
ultrapower Rω in the following way: first consider l∞(R) = {(xn)n ⊆ R : supnxn < ∞};
then consider its ideal Iω = {(xn)n ∈ l∞(Rω) : limn→ωτ (x∗
2 = 0}; finally consider the
quotient Rω = l∞(R)/Iω. It turns out to be a non weakly separable II1 factor with trace
τRω (x + Iω) = limn→ωτ (xn), where (xn) is any representative sequence for x. Connes'
embedding conjecture states that any II1-factor with separable predual embeds into Rω
nxn)
1
([Co]). This conjecture has become more and more interesting in recent years, since many
authors have found lots of equivalent conditions showing that this conjecture is linked to
3
several branches of mathematics (like group theory and metric geometry), besides being
transversal to most of the sub-specializations of Operator Algebras (see [Br], [Br2], [Ca-Pa],
[Co-Dy], [El-Sz], [Ha-Wi2], [Ki], [Ne-Th], [Oz], [Pe], [Ra1], [Ra2], [Ra3], [Vo2], [Vo3] for
some reference). Here is the problem we want to focus
Problem 4. Does every separable cyclic space embed into some II1-factor with separable
predual?
We are interested in this problem because an affirmative answer would imply the
existence of an algorithm to check Connes' embedding conjecture. Indeed
1. Take a II1-factor with separable predual M . If Prob.4 has affirmative answer, then
we could theoretically enumerate all the inequalities verified by the moments of order
3 and 4 in M . They are positive definite polynomials of degree less than or equal to
4, that are quite easy to understand, being exactly the inequalities of a cyclic space.
2. Take these polynomials and calculate their own infimum on positive matrices of order
n. Let εn ≥ 0 be such an infimum. Observe that Connes' embedding conjecture is
Indeed Connes'
true if and only if εn converges to 0, when n goes to infinity.
embedding problem has an affirmative answer if and only if one can approximate
the moments of order 3 and 4 (see [Ra3]).
If the Connes embedding conjecture is true then the algorithm is infinite for at least one
M . On the other hand, if it finite for some M , that is, it stops after a finite time, then the
Connes embedding conjecture might be true or false. In this case, the algorithm could be
used as a tool for constructing possible counter-example.
3 Extension of cyclic vector spaces
The idea to answer Prob.4 is the following: suppose we have an orthonormal basis {xn}
for L2(M, τ ), then we would have
xixj = X
αn
ijxn
n
and thus the first requirement is that an element of V ⊗V should be actually an element of
V ⊗ 1. It means that the first step is to extend the cyclic structure by adjoining elements.
4
More precisely we have to extend the cyclic structure on V to a cyclic structure on a space
W of the shape V ⊕ RY , where Y is an indeterminate, in order to reconstruct step by step
the product. We mean that, chosen arbitrarily y ∈ (V ⊗ V ) ⊗ C, y has to be represented
as the indeterminate Y , i.e. y = 1 ⊗ Y = Y ⊗ 1. This is why extending the cyclic scalar
product ≪ ·,· ≫ to one over ((V ⊕ RY )⊗ (V ⊕ RY ))⊗ C (where Y represents an arbitrary
element in V ⊗ V ) means exactly that we are extending the scalar product on V to get
the fixed product y of elements in V . Such a purpose forces some necessary assumptions
on y:
1. y must be self-adjoint, in the sense that JV y = y.
2. y must have norm 1, i.e. ≪ y, y ≫= 1.
3. y is not an element of V ⊗ 1 or 1 ⊗ V (otherwise we would have trivial product). In
this case we say that y is a non-trivial element in V ⊗ V .
Unfortunately we are not able to extend the structure exactly, but just approximately.
Definition 5. Let V be a finite dimensional cyclic vector space with orthonormal basis
{x1, ...xn}. An ε-perturbation of the original scalar product ≪ ·,· ≫ is another scalar
product ≪ ·,· ≫ε such that
≪ xi ⊗ xj, xk ⊗ xl ≫ − ≪ xi ⊗ xj, xk ⊗ xl ≫ε < ε
∀i, j, k, l ∈ {1, ...n}
Proposition 6. Let V be a finite dimensional cyclic vector space and y a self-adjoint
and non-trivial element in (V ⊗ V ) ⊗ C with norm 1. For every ε > 0, there exists an
ε-perturbation of ≪ ·,· ≫ which extends to a cyclic structure on W := V ⊕ RY with the
property y = 1 ⊗ Y = Y ⊗ 1.
Proof of this proposition is quite technical, so we will divide it in several steps. Indeed,
let 1, x2...xn an orthonormal basis of V (1 is the pointed vector on V ), we need to define
the products
≪ Y ⊗ xi, xj ⊗ xl ≫ε
≪ Y ⊗ xi, Y ⊗ xj ≫ε
≪ Y ⊗ Y, xi ⊗ xj ≫ε
≪ Y ⊗ xi, xj ⊗ Y ≫ε
≪ Y ⊗ xi, Y ⊗ Y ≫ε
≪ Y ⊗ Y, Y ⊗ Y ≫ε
5
The remaining products ≪ xi ⊗ xj, xk ⊗ xl ≫ε will be defined in the course of the proof,
when we find the suitable ε-perturbation of ≪ ·,· ≫. The most technical part of the proof
is the definition of ≪ Y ⊗ xi, xj ⊗ xk ≫ε, which will be the first and second step. In the
first step we follow a sequence of necessary conditions in order to construct a linear system
whose solutions allow us to define such products; in the second step we solve this linear
system. Before going into the first step, let us state some preliminary notions.
By the fifth property in Prop.1, JV behaves on (V ⊗ V ) ⊗ C like an involution, so it
is natural to fix the following terminology.
Definition 7. An element x ∈ (V ⊗ V ) ⊗ C is called self-adjoint if JV x = x. For an
element x ∈ (V ⊗ V ) ⊗ C which is not self-adjoint, its real part is Re(x) = x+JV x
and the
imaginary part is Im(x) = x−JV x
.
2
2i
Step 1
Let P be the projection of (V ⊗ V ) ⊗ C onto the first (V ⊗ C)⊥. Observe that
≪ Y ⊗ xi, xj ⊗ xk ≫ε=≪ Y ⊗ xi, P (xj ⊗ xk) + (1 − P )(xj ⊗ xk) ≫ε=
=≪ Y ⊗ xi, P (xj ⊗ xk) ≫ε + ≪ Y ⊗ xi, (1 − P )(xj ⊗ xk) ≫ε
Consider the second summand
λk
ij =≪ Y ⊗ xi, (1 − P )(xj ⊗ xk) ≫ε=≪ Y ⊗ xi, 1 ⊗ xk ≫ε=≪ 1 ⊗ Y, xk ⊗ xi ≫W
=≪ y, xk ⊗ xi ≫ε=≪ y, xk ⊗ xi ≫
So we can think of the numbers λk
ij as being pre-determined. Let us focus on the first
summand: we are going to find a suitable perturbation in order to determine those
numbers. Let ξi be the projection of the vector Y ⊗ xi on (V ⊗ V )⊗ C and ηi = P ξi. Then
≪ Y ⊗ xi, P (xj ⊗ xk) ≫ε=≪ ξi, P (xj ⊗ xk) ≫ε=≪ ηi, xj ⊗ xk ≫ε
So we would solve our problem if we found suitable ηi's. Now observe that they should
verify the following
≪ ηi, xj ⊗ xk ≫ε − ≪ Jεηj, xk ⊗ xi ≫ε= λji
k
6
ij =: θk
ij
− λk
This is just a linear system. Before attempting to solve it, let us write it separately for
the real and the imaginary part. We get
Re(θk
ij) = Re ≪ ηi, xj ⊗ xk ≫ε −Re ≪ Jεηj, xk ⊗ xi ≫ε=
1
2
=
(≪ ηi, xj ⊗ xk ≫ε +≪ ηi, xj ⊗ xk ≫ε−
− ≪ Jεηj, xk ⊗ xi ≫ε −≪ Jεxj, xk ⊗ xi ≫ε) =
1
(≪ ηi, xj ⊗ xk ≫ε + ≪ Jεηi, xk ⊗ xj ≫ε −
=
2
− ≪ Jεηj, xk ⊗ xi ≫ε − ≪ ηj, xi ⊗ xk ≫ε) =
xj ⊗ xk + xk ⊗ xj
ηj + Jεηj
,
≫ε − ≪
2
2
xk ⊗ xi + xi ⊗ xk
2
ηi + Jεηi
2
,
ηi − Jεηi
,
xj ⊗ xk − xk ⊗ xj
ηj − Jεηj
,
xk ⊗ xi − xi ⊗ xk
2i
2i
=≪ Re(ηi), Re(xj ⊗ xk) ≫ε − ≪ Im(ηi), Im(xj ⊗ xk) ≫ε +
− ≪ Re(ηj), Re(xk ⊗ xi) ≫ε − ≪ Im(ηj), Im(xk ⊗ xi ≫ε
2i
2i
≫ε − ≪
=≪
− ≪
≫ε +
≫ε +
By an analogous calculation we get
Im(θk
ij) =≪ Im(ηi), Re(xj ⊗ xk) ≫ε + ≪ Re(ηi), Im(xj ⊗ xk) ≫ε +
+ ≪ Im(ηj), Re(xk ⊗ xi) ≫ε − ≪ Re(ηj), Im(xk ⊗ xi ≫ε
Thus we have to solve the equations
< (...Re(ηi), Im(ηi), ..., Re(ηj ), Im(ηj )...), vk
ij >= Re(θk
ij)
and
being
and
< (...Re(ηi), Im(ηi), ..., Re(ηj ), Im(ηj )...), wk
ij >= Im(θk
ij)
vk
ij = (0, ...Re(xj ⊗ xk),−Im(xj ⊗ xk), ...0...,−Re(xk ⊗ xi),−Im(xk ⊗ xi), ...)
wk
ij = (0, ...Im(xj ⊗ xk), Re(xj ⊗ xk), ...0... − Im(xk ⊗ xi), Re(xk ⊗ xi)...)
where the non-zero components are exactly the ones corresponding to i and j.
Now observe that i, j are switchable everywhere and the case i = j is trivial. So we
7
can suppose i < j. Moreover, since the solvability of a linear system neither depend on
permutations of the columns nor on multiplication by non-zero numbers, we can replace
vk
ij and wk
ij by the following
vk
ij = (0, ...Re(xj ⊗ xk), Im(xj ⊗ xk), ...0... − Re(xk ⊗ xi), Im(xk ⊗ xi), ...)
wk
ij = (0, ... − Im(xj ⊗ xk), Re(xj ⊗ xk), ...0...Im(xk ⊗ xi), Re(xk ⊗ xi), ...)
Such a re-writing concludes the first step.
Step 2.
The purpose of this step is to find a deformation of the xi's (namely: a perturbation of
the scalar product) such that the new vk
ij's become linearly independent so that we
ij's, wk
can solve the equations in Step 1.
Suppose we have a linear combination which gives 0:
X
αk
ijvk
ij + X
βk
ijwk
ij = 0
2≤i<j≤n,2≤k≤n
2≤i<j≤n,2≤k≤n
Now fix i and look at this relation in the i-th component. We have of course the case
i < j, but also a contribution that can be obtained from some j′ < i. So we can split the
previous condition in the following ones:
X
(αk
ijRe(xj ⊗ xk) − βk
ijIm(xj ⊗ xk))+
2≤i<j≤n,2≤k≤n
and
+ X
2≤j ′<i≤n,2≤k′≤n
X
2≤i<j≤n,2≤k≤n
+ X
2≤j ′<i≤n,2≤k′≤n
(−αk′
j ′iRe(xk′ ⊗ xj ′) + βk′
j ′iIm(xk′ ⊗ xj ′))
(αk
ijIm(xj ⊗ xk) + βk
ijRe(xj ⊗ xk))+
(αk′
j ′iIm(xk′ ⊗ xj ′) + βk′
j ′iRe(xk′ ⊗ xj ′))
Now let si semicircular (see [Vo]) and ε′ > 0 small enough. Semicircularity guarantees
√ε′si = xi are still an orthonormal basis, for any ε′ > 0. The choice of
that √1 − ε′xi ⊕
ε′ small enough guarantees that the scalar product
≪ xi ⊗ xj, xk ⊗ xl ≫ε:=≪ xi ⊗ xj, xk ⊗ xl ≫
8
is an ε-deformation. Moreover observe that in this deformation xi ⊗ xk are linearly
In particular Re(xi ⊗ xj) and Im(xi ⊗ xj)
independent and independent from xi ⊗ 1.
are linearly independent over the real numbers. It follows that we can separate real and
imaginary part in the previous conditions and get
X
2≤i<j≤n,k=2,...n
ijRe(xj ⊗ xk) − X
αk
2≤j ′<i≤n,k′=2,...n
αk′
j ′iRe(xk′ ⊗ xj ′) = 0
X
2≤i<j≤n,k=2,...n
ijIm(xj ⊗ xk) + X
αk
2≤j ′<i≤n,k′=2,...n
αk′
j ′iIm(xk′ ⊗ xj ′) = 0
− X
2≤i<j≤n,k=2,...n
ijIm(xj ⊗ xk) + X
βk
2≤j ′<i≤n,k′=2,...n
βk′
j ′iIm(xk′ ⊗ xj ′) = 0
X
2≤i<j≤n,k=2,...n
ijRe(xj ⊗ xk) + X
βk
2≤j ′<i≤n,k′=2,...n
βk′
j ′iRe(xk′ ⊗ xj ′) = 0
Now, let us consider the first two conditions. If in the first sum i < k or in the second
sum i > k′, the respective terms cannot cancel each other, so their coefficients must be
zero. So one can have a term in the first sum equal to one in the second sum only in case
i > k, i < k′, k corresponds to j′ in the second sum and k′ corresponds to j in the first
sum. In this case one has αk
j ′i = 0 from
the second one. It follows that these coefficients must be zero. Similarly we obtain that
ij − αk′
j ′i = 0 from the first condition and αk
ij + αk′
the β's are equal to zero.
Step 3.
Here we want to define the scalar product ≪ ·,· ≫ε whenever Y appears twice. Recalling
that the following properties have to be satisfied
1. ≪ Y ⊗ xi, Y ⊗ xj ≫ε= ≪ xi ⊗ Y, xj ⊗ Y ≫ε
2. ≪ Y ⊗ xi, Y ⊗ xj ≫ε= ≪ Y ⊗ Y, xi ⊗ xj ≫ε
it follows that it will be enough to define the numbers ≪ Y ⊗ xi, Y ⊗ xj ≫ε and
≪ Y ⊗ xi, xj ⊗ Y ≫ε. So, we can define the matrix (≪ Y ⊗ xi, Y ⊗ xj ≫ε) as any
positive matrix (Indeed the perturbation in the second step causes xi ⊗ xj to be linearly
independent with respect to ≪ ·,· ≫ε and then the second of the previous conditions gives
no further constrictions). Finally we can set ≪ Y ⊗ xi, xj ⊗ Y ≫ε= 0.
9
Forth Step:
We can complete the proof very easily. Indeed we can set ≪ Y ⊗ xi, Y ⊗ Y ≫ε= 0, without
contradictions. Finally, Bessel's inequality forces
≪ Y ⊗ Y, Y ⊗ Y ≫ε≥ X
i,j
≪ Y ⊗ Y, xi ⊗ xj ≫ε 2
Also in this case there are no contradictions: it is enough to choose ≪ Y ⊗ Y, Y ⊗ Y ≫
large enough. (What is the smallest possible value?)
Corollary 8. Let V be a cyclic finite dimensional space with orthonormal basis x1, ...xn.
Then for every ε > 0 there exists a countably generated cyclic space Wε, with cyclic
structure ≪ ·,· ≫ε, that verifies the following properties
1. W extends V as a vector space,
i.e.
{x1, ...xn, xn+1, ...} of W .
the set {x1, ...xn} extends to a basis
2. ≪ ·,· ≫ε V is an ε-deformation of the cyclic structure on V .
3. dε(xi ⊗ xj, Wε ⊗ 1) = 0 for every i, j ∈ N, where dε(x, y) = √≪ x − y, x − y ≫ε
Proof. It is enough to iterate the previous lemma, taking ε/2n at step n.
4 Problems we were not able to solve
Let V be a separable cyclic vector space with orthonormal basis {xn}. We can think
about xi as the operator on V defined by setting xi(xj) = xi ⊗ xj. Prop.6 guarantees that
this operator is well defined by linearity, but the problem is that it could be unbounded.
Indeed xi2 = Pn αn
iixn2 could be infinite. So we have several open questions
1. Is the set of operators obtained in such a way a tracial algebra or at least as
unbounded algebra of operators of type II in the sense of Inoue (see [In])?
2. Can we modify the proof of Prop.6 in such a way that we get bounded operators?
3. What is the relation between this construction and that of Netzer and Thom (see
[Ne-Th]), who seem to obtain similar objects?
10
5 Acknowledgement
The authors are grateful to Robin Hillier for reading the draft of the paper and for
suggesting a correction.
References
[Br] N.P. Brown, Connes' embedding problem and Lance's WEP, Int.Math.Rev.Notices
(2004) 10, 501-510.
[Br2] , N.P.Brown, Topological Dynamical Systems Associated to II1-factors, preprint
arXiv:1010.1214
[Ca-Pa] V. Capraro - L. Paunescu, Product between ultrafilter and applications to the
Connes' embedding problem, preprint arXiv:math/0911.4978. Accepted by J. Oper.
Theory.
[Co] A.Connes, Classification of injective factors. Cases II1, II∞. III1, Ann. of Math.
(2) 104 (1976) no. 1, 73-115.
[Co-Dy] B. Collins and K. Dykema, Linearization of Connes' embedding problem, New
York J. Math. 14 (2008) 617-641.
[El-Sz] G. Elek - E. Szab´o, Hyperlinearity, essentially free actions and L2-invariants. The
sofic property, Math. Ann. 332 (2005) n.2, 421-441.
[Ha-Wi2] U. Haagerup and C. Winslow, The Effros-Marechal topology in the space of von
Neumann algebras, II, J. Math. Anal. 171 (2000), 401-431.
[Ki] E. Kirchberg, On semisplit extensions, tensor products and exactness of group C ∗-
algebras, Inv. Math. 112 (1993), 449-489.
[In] A. Inoue, Tomita-Takesaki theory in algebras of unbounded operators, Lecture notes
in Mathematics 1699, Springer Verlag.
[Ne-Th] T.Netzer and A. Thom, Tracial algebras and an embedding
theorem,
arxiv:math/1005.0823.
11
[Oz] N. Ozawa, About the QWEP conjecture, Intern. J. Math. 15 (2004), 501-530.
[Pe] V. Pestov, Hyperlinear and sofic groups: a brief guide, The Bulletin of Symbolic
Logic, vol. 14, number 4 (2008).
[Ra1] F. Radulescu, The von Neumann algebras of the non-residually finite Baumslag
group < a, bab3a−1 = b2 > embeds into Rω, Theta Ser. Adv. Math., vol. 9, Theta,
Bucharest, 2008.
[Ra2] F. Radulescu, A non-commutative, analytic version of Hilbert's 17-th problem in
type II1 von Neumann algebras, To appear in Proceedings Theta Foundation.
[Ra3] F. Radulescu, Convex sets associated with von Neumann algebras and Connes'
approximate embedding problem, Math. Res. Lett. 6 (1999), no.2, 229-236.
[Vo] D.Voiculescu, Circular and semicircular system and free product factor, Progress in
Math., vol.92, Birkhauser, 1990.
[Vo2] D. Voiculescu, The analogues of entropy and of Fischer's information measure in
free probability theory, I, Commun. Math. Phys. 155 (1993), 71-92.
[Vo3] D. Voiculescu, The analogues of entropy and of Fischer's information measure in
free probability theory, II, Invent. Mat. 118 (1994), 411-440.
Valerio CAPRARO - University of Rome "Tor Vergata" - [email protected]
or [email protected]
Florin R ADULESCU - University of Rome "Tor Vergata" - [email protected]
12
|
1811.05789 | 4 | 1811 | 2019-10-31T13:00:41 | Dilations of markovian semigroups of Fourier multipliers on locally compact groups | [
"math.OA"
] | We prove that any weak* continuous semigroup $(T_t)_{t \geq 0}$ of Markov Fourier multipliers acting on a group von Neumann algebra $\mathrm{VN}(G)$ associated to a locally compact group $G$ can be dilated by a weak* continuous group of Markov $*$-automorphisms on a bigger von Neumann algebra. Our construction relies on probabilistic tools and is even new for the group $\mathbb{R}^n$. Our results imply the boundedness of the McIntosh's $\mathrm{H}^\infty$ functional calculus of the generators of these semigroups on the associated noncommutative $\mathrm{L}^p$-spaces. | math.OA | math |
Dilations of markovian semigroups of Fourier
multipliers on locally compact groups
Cédric Arhancet
Abstract
We prove that any weak* continuous semigroup (Tt)t>0 of Markov Fourier multipliers
acting on a group von Neumann algebra VN(G) associated to a locally compact group G
can be dilated by a weak* continuous group of Markov ∗-automorphisms on a bigger von
Neumann algebra. Our construction relies on probabilistic tools and is even new for the
group Rn. Our results imply the boundedness of the McIntosh's H∞ functional calculus of
the generators of these semigroups on the associated noncommutative Lp-spaces.
1 Introduction
The study of dilations of operators is of central importance in operator theory and has a long
tradition in functional analysis.
Indeed, dilations are powerful tools which allow to reduce
general studies of operators to more tractable ones.
Suppose 1 < p < ∞.
In the spirit of Sz.-Nagy's dilation theorem for contractions on
Hilbert spaces, Fendler [Fen1] proved a dilation result for any strongly continuous semigroup
(Tt)t>0 of positive contractions on an Lp-space Lp(Ω). More precisely, this theorem says that
there exists a (bigger) measure space Ω′, two positive contractions J : Lp(Ω) → Lp(Ω′) and
P : Lp(Ω′) → Lp(Ω) and a strongly continuous group of positive invertible isometries (Ut)t∈R
on Lp(Ω′) such that
(1.1)
Tt = P UtJ
for any t > 0, see also [Fen2]. Note that in this situation, J : Lp(Ω) → Lp(Ω′) is an isometric
embedding whereas J P : Lp(Ω′) → Lp(Ω′) is a contractive projection.
In the noncommutative setting, measure spaces and Lp-spaces are replaced by von Neumann
algebras and noncommutative Lp-spaces and positive maps by completely positive maps. In
their remarkable paper [JLM], Junge and Le Merdy essentially1 showed that there is no hope to
have a "reasonable" analog of Fendler's result for semigroups of completely positive contractions
acting on noncommutative Lp-spaces.
It is a striking difference with the world of classical
(=commutative) Lp-spaces of measure spaces.
The semigroups of selfadjoint Markov Fourier multipliers plays a fundamental role in non-
commutative harmonic analysis and operator algebras, see e.g. [Haa1]. In this paper, our main
2010 Mathematics subject classification: Primary 47A20, 47D03, 46L51 ; Secondary 47D07.
Key words and phrases: semigroups, dilations, Markov operators, von Neumann algebras, noncommutative
Lp-spaces, functional calculus, Fourier multipliers.
1. The authors prove that there exists no "reasonable" analog of a variant of Fendler's result for a discrete
semigroup (T k)k>0 of completely positive contractions.
1
result (Theorem 3.1) gives a dilation of weak* continuous semigroups of Markov Fourier mul-
tipliers acting on the group von Neumann algebra VN(G) of a locally compact group G in the
spirit of (1.1) but at the level p = ∞. Our construction induces an isometric dilation similar to
the one of Fendler's theorem for the strongly continuous semigroup induced by the semigroup
(Tt)t>0 on the associated noncommutative Lp-space Lp(VN(G)) for any 1 6 p < ∞. Note that
our paper [Arh4] gives a nonconstructive and complicated proof of such a dilation and effective
only for discrete groups. Here our approach is very different, direct and short. In addition, it
can also be used with non-discrete locally compact groups. We refer to [Arh1], [Arh2], [Arh4],
[ALM], [AFM], [HaM] and [Ric] for strongly related things.
One of the important consequences of Fendler's theorem is the boundedness, for the genera-
tor of a strongly continuous semigroup (Tt)t>0 of positive contractions, of a bounded H∞ func-
tional calculus which is a fundamental tool in various areas: harmonic analysis of semigroups,
multiplier theory, Kato's square root problem, maximal regularity in parabolic equations, con-
trol theory, etc. For detailed information, we refer the reader to [Haa], [JMX], [KW], to the
survey [LeM1] and to the recent book [HvNVW2] and references therein. Our theorem also
gives a similar result on H∞ functional calculus in the noncommutative context as explained
in [JMX, Proposition 3.12] and [JMX, Proposition 5.8] in the case of semigroups of Fourier
multipliers.
The paper is organized as follows. The next Section 2 gives background. In particular, we
give some information on crossed products since our construction relies on this notion. We also
prove some elementary results which will be used in the sequel. Section 3 gives a proof of our
main result of dilation of semigroups of Markov Fourier multipliers. In the last section 4, we
describe some applications of our results to functional calculus.
2 Preliminaries
Isonormal processes Let H be a real Hilbert space. An H-isonormal process on a probability
space (Ω, µ) [Nua1, Definition 1.1.1] [Neer1, Definition 6.5] is a linear mapping W : H → L0(Ω)
with the following properties:
(2.1)
for any h ∈ H the random variable W(h) is a centered real Gaussian,
(2.2)
(2.3)
for any h1, h2 ∈ H we have E(cid:0)W(h1)W(h2)(cid:1) = hh1, h2iH .
The linear span of the products W(h1)W(h2) · · · W(hm), with m > 0 and h1, . . . , hm
in H, is dense in the real Hilbert space L2
R(Ω).
Here L0(Ω) denote the space of measurable functions on Ω and we make the convention that
the empty product, corresponding to m = 0 in (2.3), is the constant function 1.
If (ei)i∈I is an orthonormal basis of H and if (γi)i∈I is a family of independent standard
Gaussian random variables on a probability space Ω then for any h ∈ H, the family (γihh, eii)i∈I
is summable in L2(Ω) and
(2.4)
W(h)
def
= Xi∈I
γihh, eiiH ,
h ∈ H
define an H-isonormal process.
Recall that the span of elements eiW(h) is weak* dense in L∞(Ω) by [Jan1, Remark 2.15].
Using [HvNVW2, Proposition E.2.2], we see that
(2.5)
E(cid:0)eitW(h)(cid:1) = e− t
2khk2
H ,
t ∈ R, h ∈ H.
2
If u : H → H is a contraction, we denote by Γ∞(u) : L∞(Ω) → L∞(Ω) the (symmetric)
second quantization of u acting on the complex Banach space L∞(Ω). Recall that the map
Γ∞(u) : L∞(Ω) → L∞(Ω) preserves the integral2. If u is an isometry we have
(2.6)
h ∈ H
and Γ∞(u) : L∞(Ω) → L∞(Ω) is a ∗-automorphism of the von Neumann algebra L∞(Ω). Fur-
thermore, the second quantization functor Γ satisfies the following. In the part 1, we suppose
that the construction3 is given by the concrete representation (2.4). We will only use this
observation in the non-discrete case.
Γ∞(u)(cid:0)eiW(h)(cid:1) = eiW(u(h)),
Lemma 2.1
1. If L∞(Ω) is equipped with the weak* topology then the map H → L∞(Ω),
h 7→ eiW(h) is continuous.
2. If π : G → B(H) is a strongly continuous orthogonal representation of a locally compact
group, then G → B(L∞(Ω)), s 7→ Γ∞(πs) is a weak* continuous4 representation on the
Banach space L∞(Ω).
Proof
: 1. Suppose that the sequence (hn) converges to h in H. Note that there exists an at
most countable subset J of I such that i ∈ I − J implies hhn, eiiH = 0 and hh, eiiH = 0. By
(2.4), note that W(hn) =Pi∈J γihhn, eiiH and W(h) =Pi∈J γihh, eiiH in L2(Ω), hence almost
everywhere by [HvNVW2, Corollary 6.4.4]. For any i ∈ J, the sequence (hhn, eiiH ) converges
to hh, eiiH . Since J is at most countable, W(hn) converges almost everywhere to W(h). It is
not difficult to conclude using the dominated convergence theorem.
2. The second part can be proved with the arguments of the proof of [JMX, Lemma 9.3].
Von Neumann algebras Let M be a von Neumann algebra. We denote by U(M ) the group
of all unitaries of M . Suppose that M is equipped with a semifinite normal faithful weight ψ.
We denote by nψ the left ideal of all x ∈ M such that ψ(x∗x) < ∞.
Let M and N be von Neumann algebras equipped with faithful normal semifinite weights
ψM and ψN . We say that a positive linear map T : M → N is weight preserving if for any
x ∈ m
we have ψN (T (x)) = ψM (x).
Note that a locally convex space X is said to be quasi-complete if every bounded Cauchy
net of X converges [Osb1, Definition 4.23]. Recall that if H is a Hilbert space and if M is a von
Neumann algebra acting on H then the σ-strong* topology on M is defined by the seminorms
+
ψM
x 7→ (cid:0)P∞n=1(kx(ξn)k2 + kx∗(ξn)k2)(cid:1)
2 where ξn ∈ H and P∞n=1 kξnk2 < ∞. The following
lemma is folklore. Unable to locate this statement in the literature, we give a short argument.
1
Lemma 2.2 Let M be a von Neumann algebra acting on a Hilbert space H. Then M equipped
with the canonical locally convex structure which gives the σ-strong* topology is quasi-complete.
Proof
: First, we prove that M equipped with its canonical locally convex structure which
gives the strong topology is quasi-complete. Consider a bounded Cauchy net (xi) of elements
of M for this structure. Then, for any h ∈ H, (xi(h)) is a Cauchy net of elements of H, hence
converges. So (xi) is a net which converges for the strong topology and which is norm-bounded
2. That means that for any f ∈ L∞(Ω) we haveRΩ Γ∞(u)f dµ =RΩ
3. The existence of a proof of Lemma (2.1) without (2.4) is unclear.
4. That means that B(L∞(Ω)) is equipped with the point weak* topology.
f dµ.
3
by the uniform boundedness principle [KaR1, Theorem 1.8.9]. It is well-known that this implies
that (xi) converges strongly to some bounded linear operator x. Since each xi belongs to M .
We conclude that x belongs to M , we conclude that the strong structure is quasi-complete.
Recall that a net (xi) is Cauchy in the strong* topology if and only if (xi) and (x∗i ) are
Cauchy in the strong operator topology and that (xi) converges to x in the strong* topology
if and only if (xi) and (x∗i ) converge to x and x∗ in the strong operator topology. Using this
fact, it is not difficult to see that the canonical locally convex structure which gives the strong*
topology is quasi-complete. Finally, recall that by [Tak1, Lemma 2.5] the σ-strong* topology
coincide with the strong* topology on bounded subsets. So the proof is complete.
Group von Neumann algebras Let G be a locally compact group equipped with a fixed
left invariant Haar measure µG. Let VN(G) be the von Neumann algebra of G generated by
the set (cid:8)λ(g) : g ∈ L1(G)(cid:9). It is called the group von Neumann algebra of G and is equal to
the von Neumann algebra generated by the set {λs : s ∈ G} where λs : L2(G) → L2(G), f 7−→
(t 7→ f (s−1t)) is the left translation by s.
Crossed products We refer to [Haa2], [Haa3], [Str1], [Sun] and [Tak2]. Let M be a von
Neumann algebra acting on a Hilbert space H. Let G be a locally compact group equipped
with some left Haar measure µG. Let α : G → M be a representation of G on M which is
weak* continuous, i.e. for any x ∈ M and any y ∈ M∗, the map G → M , s 7→ hαs(x), yiM,M∗
is continuous. For any x ∈ M , we define the operators π(x) : L2(G, H) → L2(G, H) [Str1, (2)
page 263] by
(2.7)
def
= α−1
s (x)ξ(s),
ξ ∈ L2(G, H), s ∈ G.
(cid:0)π(x)ξ(cid:1)(s)
These operators satisfy the following commutation relation [Str1, (2) page 292]:
(2.8)
(λs ⊗ IdH )π(x)(λs ⊗ IdH )∗ = π(αs(x)),
x ∈ M, s ∈ G.
Recall that the crossed product of M and G with respect to α is the von Neumann algebra
M ⋊α G = (π(M ) ∪ {λs ⊗ IdH : s ∈ G})′′ on the Hilbert space L2(G, H) generated by the
operators π(x) and λs ⊗ IdH where x ∈ M and s ∈ G. By [Str1, page 263] or [Dae1, Proposition
2.5], π is a normal injective ∗-homomorphism from M into M ⋊αG (hence σ-strong* continuous).
We denote by K(G, M ) the space of σ-strong* continuous function f : G → M , s 7→ fs
with compact support. If f ∈ K(G, M ) then f (G) is a σ-strong* compact subset of M , hence
by [Osb1, Proposition 2.7 d)] a σ-strong* bounded subset of M . Hence it is a strong bounded
subset and finally a norm-bounded subset of M by the principle of uniform boundedness [KaR1,
Theorem 1.8.9]. Note that by [Str1, Proposition page 186] and [Str1, page 41], the bounded
function G → M , s 7→ λs ⊗ IdH is σ-strong* continuous and the norm-bounded function
s 7→ π(fs) is also σ-strong* continuous. Recall that the product of M is σ-strong* continuous
on bounded subsets by [BrR1, Proposition 2.4.5]. We infer5 that the function G → M ⋊β G,
s 7→ π(fs)(λs ⊗ IdH ) is σ-strong* continuous with compact support. So, by Lemma 2.2 and
[Bou1, Corollary 2, III page 38] we can define the element RG fs ⋊ λs dµG(s) of the crossed
product M ⋊α G by
(2.9)
fs ⋊ λs dµG(s)
ZG
def
= ZG
π(fs)(λs ⊗ IdH ) dµG(s).
5. In the book [Str1], the author considers weak* continuous functions, it is problematic since the product of M
is not weak* continuous even on bounded sets by [KaR1, Exercise 5.7.9] (indeed this latter fact is equivalent to
the weak continuity of the product on bounded sets by [Tak1, Lemma 2.5]).
4
If f, g ∈ K(G, M ) then essentially by the proof of [Haa2, Lemma 2.3] and [Str1, page 289]
the function g ∗ f : G → M , t →RG g(s)f (s−1t) dµG(s) is σ-strong* continuous and we have
(g ∗ f )(s) ⋊ λs dµG(s).
(2.10)
(cid:18)ZG
gs ⋊ λs dµG(s)(cid:19)(cid:18)ZG
fs ⋊ λs dµG(s)(cid:19) =ZG
weak* dense subalgebra of M ⋊α G.
It is very easy to see that the space of elements RG fs ⋊ λs dµG(s) for f ∈ K(G, M ) is a
The following is a particular case6 of [Tak3, Proposition 3.5] and its proof, see also [Tak2,
Theorem 1.7 (ii) p. 241]. Note that M is abelian in the statement. With [Bou1, Proposition 2,
III page 35], the last part is an easy computation left to the reader.
Proposition 2.3 Let M be an abelian von Neumann algebra acting on a Hilbert space H
equipped with a continuous action α of a locally compact group G. Suppose that there exists a
strongly continuous function u : G → U(M ) such that
(2.11)
u(sr) = u(s)αs(u(r)),
s, r ∈ G.
Then V : L2(G, H) → L2(G, H), ξ 7→ (s 7→ u(s−1)(ξ(s))) is a unitary and we have an ∗-
isomorphism U : M ⋊α G → M ⋊α G, x 7→ V xV ∗ such that
(2.12)
U (λs ⊗ IdH ) = π(u(s)∗)(λs ⊗ IdH ) and U (π(x)) = π(x),
s ∈ G, x ∈ M.
Moreover, for any f ∈ K(G, M ), we have
(2.13)
U(cid:18)ZG
fs ⋊ λs dµG(s)(cid:19) =ZG
u(s)∗fs ⋊ λs dµG(s).
Now, we suppose that M is finite and equipped with a normal finite faithful trace τ . By
[Haa2, Lemma 3.3] [Str1, Theorem p. 301] [Tak2, Theorem 1.17], there exists a unique normal
semifinite faithful weight ϕ⋊ on M ⋊β G which satisfies for any f, g ∈ K(G, M ) the fundamental
"noncommutative Plancherel formula"
(2.14)
ϕ⋊(cid:18)(cid:18)ZG
fs ⋊ λs dµG(s)(cid:19)∗(cid:18)ZG
gs ⋊ λs dµG(s)(cid:19)(cid:19) =ZG
τ (f∗s gs) dµG(s)
and the relations σϕ⋊
t
(π(x)) = π(x) where x ∈ M and t ∈ R and
σϕ⋊
t
(λs ⊗ IdH ) = ∆it
G(s)(λs ⊗ IdH )π([D(τ ◦ αs) : Dτ ]t),
s ∈ G, t ∈ R.
If M = C, we recover the Plancherel weight ϕG on VN(G).
preserving, we obtain in particular
If each αs : M → M is trace
σϕ⋊
t
(λs ⊗ IdH ) = ∆it
G(s)(λs ⊗ IdH),
s ∈ G, t ∈ R.
Using [Bou1, Proposition 2, III page 35], we deduce that
(2.15)
σϕ⋊
t (cid:18)ZG
fs ⋊ λs dµG(s)(cid:19) =ZG
6. The function u : G → U(M ) is a α-1-cocycle.
∆it
G(s)fs ⋊ λs dµG(s),
f ∈ K(G, M ), t ∈ R.
5
Fourier multipliers Let G be a locally compact group. We say that a weak* continuous
operator T : VN(G)) → VN(G) is a Fourier multiplier if there exists a continuous function
φ : G → C such that for any s ∈ G we have T (λs) = φ(s)λs. In this case φ is bounded and for
any f ∈ L1(G) the element RG φ(s)f (s)λs dµG(s) belongs to VN(G) and
(2.16)
φ(s)f (s)λs dµG(s),
i.e. T (λ(f )) = λ(φf ).
T(cid:18)ZG
f (s)λs dµG(s)(cid:19) =ZG
In this case, we let T = Mφ and we say that φ is the symbol of T . We refer to the book
[KaL1] and references therein for more information and to [ArK1] for Fourier multipliers on
noncommutative Lp-spaces.
Semigroups of Fourier multipliers Consider a locally compact group G with identity
element e. Let (Tt)t>0 be a weak* continuous semigroup of selfadjoint unital completely positive
Fourier multipliers. For any t > 0, let φt : G → C be the continuous symbol of Tt. Since each
Tt is selfadjoint, φt is real-valued. By [DCH, Proposition 4.2], the function φt is of positive
type. Moreover, for any t, t′ > 0 and any s ∈ G, the relation TtTt′(λs) = Tt+t′ (λs) gives
φt(s)φt′ (s) = φt+t′ (s). Furthermore, for any s ∈ G, any x ∈ VN(G)∗ and any t > 0, we have
φt(s)hλs, xiVN(G),VN(G)∗ = hTt(λs), xiVN(G),VN(G)∗ −−−→
t→0
hλs, xiVN(G),VN(G)∗ .
We infer that limt→0 φt(s) = 1. Consequently, there exists a uniquely determined real number
ψ(s) such that
φt(s) = e−tψ(s),
t > 0, s ∈ G.
Since each Tt is unital, we have ψ(e) = 0. It is easy to check that ψ : G → R is continuous. By
Schoenberg's theorem [BHV, Corollary C.4.19], we deduce that the function ψ is conditionally
of negative type.
By [BHV, Proposition 2.10.2], there exist a real Hilbert space H and an affine isometric
action β : G → Isom(H), s 7→ βs such that the linear span of {βs(0) : s ∈ G} is dense in H and
such that
(2.17)
ψ(s) = kβs(0)k2
H,
s ∈ G.
By [BHV, page 75], we can consider the associated orthogonal representation π : G → B(H),
s 7→ πs of G on H. By [BHV, Lemma 2.2.1], there exists a 1-cocycle bψ : G → H with respect
to π, i.e. a continuous function satisfying the cocycle law
(2.18)
πs(bψ(r)) = bψ(sr) − bψ(s),
i.e.
bψ(sr) = bψ(s) + πs(bψ(r)),
s, r ∈ G,
such that
(2.19)
βs(h) = πs(h) + bψ(s),
h ∈ H.
For any s ∈ G, we deduce that
(2.20)
ψ(s)
(2.17)
= kβs(0)k2
H
6
(2.19)
= kbψ(s)k2
H .
3 Dilations of semigroups of Fourier multipliers
The following theorem is our main result.
Theorem 3.1 Let G be a locally compact group. Let (Tt)t>0 be a weak* continuous semigroup
of selfadjoint unital completely positive Fourier multipliers on the group von Neumann algebra
VN(G). Then, there exist a von Neumann algebra M equipped with a semifinite normal faithful
weight ϕM , a weak* continuous group (Ut)t∈R of weight preserving ∗-automorphisms of M , a
unital weight preserving injective normal ∗-homomorphism J : VN(G) → M such that
(3.1)
σϕM
t
◦ J = J ◦ σϕG
t
for any t ∈ R and satisfying
(3.2)
Tt = EUtJ
for any t > 0, where E : M → VN(G) is the canonical faithful normal weight preserving condi-
tional expectation associated with J. Moreover, we have the following properties.
1. If G is discrete then ϕM is a normal finite faithful trace.
2. If G is unimodular then ϕM is a normal semifinite faithful trace.
3. If G is amenable then the von Neumann algebra M is injective.
Proof
: Here we use the continuous function ψ : C → R, the 1-cocycle bψ : G → H and the
orthogonal representation π : G → B(H), s 7→ πs of G on H of Section 2. Let W : H → L0(Ω)
be an H-isonormal process on a probability space (Ω, µ), see again Section 2. For any s ∈ G, we
= Γ∞(πs) : L∞(Ω) → L∞(Ω) which is integral preserving.
will use the second quantization αs
In particular, if r, s ∈ G and if t ∈ R, we have
def
(3.3)
αs(cid:0)e−√2itW(bψ(r))(cid:1) = Γ∞(πs)(cid:0)e−√2itW(bψ(r))(cid:1) (2.6)
= e−√2itW(πs(bψ(r))).
Since π is strongly continuous, by Lemma 2.1, we obtain a continuous action α : G → Aut(L∞(Ω)).
def
= L∞(Ω) ⋊α G equipped with its canonical normal
So we can consider the crossed product M
def
= ϕ⋊. We denote by J : VN(G) → L∞(Ω) ⋊α G, λs 7→ 1 ⋊ λs
semifinite faithful weight ϕM
the canonical unital normal injective ∗-homomorphism. Using [Bou1, Proposition 2, III page
35], for any f ∈ K(G), we see that
(3.4)
J(cid:18)ZG
f (s)λs dµG(s)(cid:19) =ZG
f (s)1 ⋊ λs dµG(s).
Since each αs : L∞(Ω) → L∞(Ω) is integral preserving, we have for any t ∈ R and any
f ∈ K(G)
σϕ⋊
t
◦ J(cid:18)ZG
= J(cid:18)ZG
f (s)λs dµG(s)(cid:19) (3.4)
t (cid:18)ZG
G(s)f (s)λs dµG(s)(cid:19) (2.15)
= σϕ⋊
f (s)1 ⋊ λs dµG(s)(cid:19) (2.15)
= ZG
f (s)λs dµG(s)(cid:19).
t (cid:18)ZG
= J ◦ σϕG
(3.4)
∆it
∆it
G(s)f (s)1 ⋊ λs dµG(s)
By density, we obtain (3.1).
7
Lemma 3.2 We have ϕ⋊ ◦ J = ϕG.
: We will use [Str1, Theorem 6.2] with the weights ϕG and ϕ⋊ ◦ J. Note7 that λ(K(G))
Proof
is a ∗-subalgebra which is σϕG -invariant, weak* dense in VN(G) and included in nϕG.
For any f ∈ K(G), we have f (·)1 ∈ K(G, L∞(Ω)). So J(λ(f ))
= RG f (s)1 ⋊ λs dµG(s)
belongs to nϕ⋊. Hence ϕ⋊ ◦ J(cid:0)λ(f )∗λ(f )(cid:1) = ϕ⋊(cid:0)J(λ(f ))∗J(λ(f ))(cid:1) < ∞. Consequently, we
have λ(f ) ∈ nϕ⋊◦J . We infer that nϕ⋊◦J is weak* dense in VN(G). By [Str1, page 19], we
conclude that the weight ϕ⋊ ◦ J is semifinite. It is obvious that ϕ⋊ ◦ J is normal. For any t ∈ R,
note that
(3.4)
ϕ⋊ ◦ J ◦ σϕG
t
(3.1)
= ϕ⋊ ◦ σϕ⋊
t
◦ J = ϕ⋊ ◦ J.
So the weights ϕ⋊ ◦ J and ϕG commutes by [Str1, pages 67-68]. Now, if f ∈ K(G), we have
(3.4)
ϕ⋊ ◦ J(cid:0)λ(f )∗λ(f )(cid:1) = ϕ⋊(cid:0)J(λ(f ))∗J(λ(f ))(cid:1)
f (s)1 ⋊ λs dµG(s)(cid:19)∗ZG
f (s)21 dµ(cid:19) dµG(s) =ZG
= ϕ⋊(cid:18)(cid:18)ZG
= ZG(cid:18)ZΩ
(2.14)
f (s)1 ⋊ λs dµG(s)(cid:19)
f (s)2 dµG(s)
(2.14)
= ϕG(cid:0)λ(f )∗λ(f )(cid:1).
We conclude with [Str1, Theorem 6.2] that ϕ⋊ ◦ J = ϕG.
For any t ∈ R, we consider the function ut : G → U(L∞(Ω)), s 7→ e−√2itW(bψ (s)). The map
bψ : G → H is continuous. By the point 1 of Lemma 2.1, the map H → L∞(Ω), h 7→ eiW(h)
is continuous if L∞(Ω) is equipped with the weak* topology, hence with the weak operator
topology when we consider that L∞(Ω) acts on L2(Ω). By composition, the function ut is
continuous. Recall that by [KaR1, Exercice 5.7.5] or [Str1, page 41] the weak operator topology
and the strong operator topology coincide on the unitary group U(L∞(Ω)). So the function ut
is continuous if U(L∞(Ω)) is equipped with the strong operator topology. For any t ∈ R and
any r, s ∈ G, note that
ut(sr) = e−√2itW(bψ(sr)) (2.18)
= e−√2itW(bψ(s))e−√2itW(πs(bψ(r)))
(3.3)
= ut(s)αs(cid:0)e−√2itW(bψ(r))(cid:1) = ut(s)αs(ut(r)).
Hence (2.11) is satisfied. By Proposition 2.3, for any t ∈ R, we have a unitary Vt : L2(G, L2(Ω)) →
L2(G, L2(Ω)), ξ 7→ (s 7→ ut(s−1)ξ(s)) and a ∗-isomorphism Ut : L∞(Ω) ⋊α G → L∞(Ω) ⋊α G,
x 7→ VtxV ∗t such that
(3.5)
Ut(cid:18)ZG
fs ⋊ λs dµG(s)(cid:19) (2.13)
= ZG
√2itW(bψ(s))fs ⋊ λs dµG(s)
e
for any f ∈ K(G, M ). For any t, t′ ∈ R any ξ ∈ L2(G, L2(Ω)), note that almost everywhere
(cid:0)VtVt′ (ξ)(cid:1)(s) = ut(cid:0)s−1(cid:1)(ut′ (s−1)ξ(s) = e−√2itW(bψ(s−1))e−√2it′W(bψ(s−1))ξ(s)
= e−√2i(t+t′)W(bψ(s−1))ξ(s) = ut+t′ (s−1)ξ(s) =(cid:0)Vt+t′ (ξ)(cid:1)(s).
f (s)λs dµG(s).
7. Recall that λ(f ) =RG
8
We conclude that VtVt′ = Vt+t′ . Moreover, for any ξ, η ∈ L2(G, L2(Ω)) = L2(G × Ω), using
dominated convergence theorem, we obtain
(cid:10)Vt(ξ), η(cid:11)L2(G,L2(Ω)) =ZG×Ω
Vt(ξ)(s)(ω)η(s, ω) dµG(s) dµ(ω)
(ut(s−1)ξ(s))(ω)η(s, ω) dµG(s) dµ(ω)
√2itW(bψ(s−1))(ω)ξ(s, ω)η(s, ω) dµG(s) dµ(ω).
ξ(s, ω)η(s, ω) dµG(s) dµ(ω) = hξ, ηiL2(G,L2(Ω)).
=ZG×Ω
=ZG×Ω
t→0 ZG×Ω
−−−→
e
So (Vt)t∈R is a weakly continuous group of unitaries hence a strongly continuous group by [Str1,
Lemma 13.4] or [Tak2, page 239]. By [Tak2, page 238], we conclude that (Ut)t∈R is a weak*
continuous group of ∗-automorphisms.
Lemma 3.3 Each ∗-automorphism Ut preserves the weight ϕ⋊.
Proof
: We will use [Str1, Theorem 6.2] with the weights ϕ⋊ and ϕ⋊ ◦ Ut. Note that the space
of elements RG fs ⋊ λs dµG(s) for f ∈ K(G, L∞(Ω)) is a ∗-subalgebra which is σϕ⋊ -invariant,
weak* dense in L∞(Ω) ⋊α G and included in nϕ⋊. The formulas (2.15) and (3.5) show that each
Ut and σϕ⋊
commute. So, we have
t
ϕ⋊ ◦ Ut ◦ σϕ⋊
t = ϕ⋊ ◦ σϕ⋊
t
◦ Ut = ϕ⋊ ◦ Ut.
So the weights ϕ⋊ ◦ Ut and ϕ⋊ commutes by [Str1, pages 67-68]. It is (really) easy to check
that the weight ϕ⋊ ◦ Ut is normal and semifinite. If f ∈ K(G, L∞(Ω)), we have
ϕ⋊ ◦ Ut(cid:18)(cid:18)ZG
fs ⋊ λs dµG(s)(cid:19)∗(cid:18)ZG
fs ⋊ λs dµG(s)(cid:19)(cid:19)
e
(3.5)
= ϕ⋊ (cid:18)Ut(cid:18)ZG
= ϕ⋊ (cid:18)ZG
= ZGZΩ
= ϕ⋊(cid:18)(cid:18)ZG
(2.14)
(2.14)
Ut(cid:18)ZG
fs ⋊ λs dµG(s)(cid:19)(cid:19)∗
√2itW(bψ(s))fs ⋊ λs dµG(s)(cid:19)∗(cid:18)ZG
fs ⋊ λs dµG(s)(cid:19)!
e
√2itW(bψ(s))fs ⋊ λs dµG(s)(cid:19)!
√2itW(bψ(s))fs dµ dµG(s) =ZGZΩ
f∗s fs dµ dµG(s)
e−√2itW(bψ(s))f∗s e
fs ⋊ λs dµG(s)(cid:19)∗(cid:18)ZG
fs ⋊ λs dµG(s)(cid:19)(cid:19).
We conclude with [Str1, Theorem 6.2] that ϕ⋊ ◦ Ut = ϕ⋊ for any t > 0.
With [Str1, Theorem 10.1], we introduce the canonical faithful normal weight preserving
conditional expectation E : L∞(Ω) ⋊α G → VN(G). Using [AcC, Theorem 7.5], it is entirely left
to the reader to check that the conditional expectation is given by
(3.6)
E(cid:18)ZG
gs ⋊ λs dµG(s)(cid:19) =ZG
E(gs)λs dµG(s).
9
For any t > 0, we have
EUtJ(cid:18)ZG
f (s)λs dµG(s)(cid:19) (3.4)
= EUt(cid:18)ZG
f (s)1 ⋊ λ(s) dµG(s)(cid:19)
(3.5)
(3.6)
(2.5)
e
= E(cid:18)ZG
= ZG
E(cid:0)e
= ZG
= ZG
e− t
√2itW(bψ(s))f (s) ⋊ λs dµG(s)(cid:19)
√2itW(bψ(s))f (s)(cid:1)λs dµG(s) =ZG
H f (s)λs dµG(s) =ZG
= Tt(cid:18)ZG
(2.16)
2 2kbψ (s)k2
e−tψ(s)f (s)λs dµG(s)
(2.20)
√2itW(bψ(s))(cid:1)f (s)λs dµG(s)
E(cid:0)e
e−tkbψ (s)k2
H f (s)λs dµG(s)
f (s)λs dµG(s)(cid:19).
Thus by density, for any t > 0, we obtain (3.2). Now, we prove the last assertions. Note each
αs : L∞(Ω) → L∞(Ω) is integral-preserving. The first is well-known, e. g.
[Ped1, Corollary
7.11.8]. The second is folklore. The third is [Ana1, Proposition page 301].
Remark 3.4 In the case of the Heat semigroup (Ht)t>0 on L∞(Rn), we can take H = Rn,
bψ = IdRn and the action α is trivial. Here, we does not need a noncommutative von Neumann
algebra for the dilation since M is a tensor product of commutative von Neumann algebras.
This observation is a complement to [Arh4, Secton 4].
4 Functional calculus
We start with a little background on sectoriality and H∞ functional calculus. We refer to [Haa],
[KW], [JMX], [HvNVW2] and [Arh2] for details and complements. Let X be a Banach space.
A closed densely defined linear operator A : dom A ⊂ X → X is called sectorial of type ω if its
def
= {z ∈ C∗ : arg z < ω}, and for
spectrum σ(A) is included in the closed sector Σω where Σω
any angle ω < θ < π, there is a positive constant Kθ such that
(cid:13)(cid:13)(λ − A)−1(cid:13)(cid:13)X→X
6
Kθ
λ
,
λ ∈ C − Σθ.
If −A is the negative generator of a bounded strongly continuous semigroup on a X then A is
sectorial of type π
2 coincide
with negative generators of bounded analytic semigroups.
2 . By [HvNVW2, Example 10.1.3], sectorial operators of type < π
For any 0 < θ < π, let H∞(Σθ) be the algebra of all bounded analytic functions f : Σθ → C,
equipped with the supremum norm kf kH∞(Σθ ) = sup(cid:8)f (z) : z ∈ Σθ(cid:9). Let H∞0 (Σθ) ⊂ H∞(Σθ)
be the subalgebra of bounded analytic functions f : Σθ → C for which there exist s, c > 0 such
that f (z) 6
czs
(1+z)2s for any z ∈ Σθ.
Given a sectorial operator A of type 0 < ω < π, a bigger angle ω < θ < π, and a function
f ∈ H∞0 (Σθ), one may define a bounded operator f (A) by means of a Cauchy integral (see e.g.
[Haa, Section 2.3] or [KW, Section 9]). The resulting mapping H∞0 (Σθ) → B(X) taking f to
f (A) is an algebra homomorphism. By definition, A has a bounded H∞(Σθ) functional calculus
provided that this homomorphism is bounded, that is if there exists a positive constant C such
6 Ckf kH∞(Σθ ) for any f ∈ H∞0 (Σθ). In the case when A has a dense range,
the latter boundedness condition allows a natural extension of f 7→ f (A) to the full algebra
H∞(Σθ).
that (cid:13)(cid:13)f (A)(cid:13)(cid:13)X→X
10
Using the connection between the existence of dilations in UMD spaces and H∞ functional
calculus together with the well-known angle reduction principle of Kalton-Weis relying on R-
sectoriality, Theorem 3.1 allows us to obtain the following result (see [JMX, Proposition 3.12],
[JMX, Proposition 5.8] and [KW, Corollary 10.9]). Here, we confine ourselves to unimodular
groups by simplicity. Non-unimodular extensions and vector-valued versions of this result are
left to the reader. Other applications will be given in subsequent papers, e.g. [ArK2].
Theorem 4.1 Let G be unimodular locally compact group. Let (Tt)t>0 be a weak* continuous
semigroup of selfadjoint unital completely positive Fourier multipliers on VN(G). Suppose 1 <
p < ∞. We let −Ap be the generator of the induced strongly continuous semigroup (Tt,p)t>0 on
the Banach space Lp(VN(G)). Then for any θ > π 1
2 , the operator Ap has a completely
bounded H∞(Σθ) functional calculus.
p − 1
Acknowledgements. The author acknowledges support by the grant ANR-18-CE40-0021
(project HASCON) of the French National Research Agency ANR. Finally, I would like to thank
the referee for useful remarks.
References
[AcC] L. Accardi and C. Cecchini. Conditional expectations in von Neumann algebras and a theorem
of Takesaki. J. Funct. Anal. 45 (1982), 245 -- 273.
[Ana1] C. Anantharaman-Delaroche. Action moyennable d'un groupe localement compact sur une
algèbre de von Neumann. (French). Math. Scand. 45 (1979), no. 2, 289 -- 304.
[Arh1] C. Arhancet. On Matsaev's conjecture for contractions on noncommutative Lp-spaces. Journal
of Operator Theory 69 (2013), no. 2, 387 -- 421.
[Arh2] C. Arhancet. Analytic semigroups on vector valued noncommutative Lp-spaces. Studia Math.
216 (2013), no. 3, 271 -- 290.
[Arh4] C. Arhancet. Dilations of semigroups on von Neumann algebras and noncommutative Lp-spaces.
J. Funct. Anal. 276 (2019), no. 7, 2279 -- 2314.
[ALM] C. Arhancet and C. Le Merdy. Dilation of Ritt operators on Lp-spaces. Israel J. Math. 201
(2014), no. 1, 373 -- 414.
[AFM] C. Arhancet, S. Fackler and C. Le Merdy. Isometric dilations and H ∞ calculus for bounded
analytic semigroups and Ritt operators. Transactions of the American Mathematical Society 369
(2017), 6899 -- 6933.
[ArK1] C. Arhancet and C. Kriegler. Projections, multipliers and decomposable maps on noncommu-
tative Lp-spaces. Preprint, arXiv:1707.05591.
[ArK2] C. Arhancet and C. Kriegler. Riesz transforms, Hodge-Dirac operators and functional calculus
for Fourier multipliers. Preprint.
[BHV] B. Bekka, P. de la Harpe, Pierre and A. Valette. Kazhdan's property (T). New Mathematical
Monographs, 11. Cambridge University Press, Cambridge, 2008.
[Bou1] N. Bourbaki. Integration. I. Chapters 1 -- 6. Translated from the 1959, 1965 and 1967 French
originals by Sterling K. Berberian. Elements of Mathematics. Springer-Verlag, Berlin (2004).
[BrR1] O. Bratteli and D. W. Robinson. Operator algebras and quantum statistical mechanics. 1.
C ∗- and W ∗-algebras, symmetry groups, decomposition of states. Second edition. Texts and
Monographs in Physics. Springer-Verlag, New York, 1987.
[DCH] J. de Cannière and U. Haagerup. Multipliers of the Fourier algebras of some simple Lie groups
and their discrete subgroups. Amer. J. Math. 107 (1985), no. 2, 455 -- 500.
11
[Dae1] A. van Daele. Continuous crossed products and type III von Neumann algebras. London
Mathematical Society Lecture Note Series, 31. Cambridge University Press, Cambridge-New York,
1978.
[Fen1] G. Fendler. Dilations of one parameter semigroups of positive contractions on Lp spaces. Canad.
J. Math. 49 (1997), no. 4, 736 -- 748.
[Fen2] G. Fendler. On dilations and Transference for Continuous One-Parameter Semigroups of Positive
Contractions on Lp-spaces. Ann. Univ. Sarav. Ser. Math. 9 (1998), no. 1.
[Haa] M. Haase. The functional calculus for sectorial operators. Operator Theory: Advances and
Applications, 169. Birkhäuser Verlag (2006).
[Haa1] U. Haagerup. An Example of a nonnuclear C*-Algebra, which has the metric approximation
property. Invent. Math. 50 (1978/79), no. 3, 279 -- 293.
[Haa2] U. Haagerup. On the dual weights for crossed products of von Neumann algebras. I. Removing
separability conditions. Math. Scand. 43 (1978/79), no. 1, 99 -- 118.
[Haa3] U. Haagerup. On the dual weights for crossed products of von Neumann algebras. II. Application
of operator-valued weights. Math. Scand. 43 (1978/79), no. 1, 119 -- 140.
[HaM] U. Haagerup and M. Musat. Factorization and dilation problems for completely positive maps
on von Neumann algebras. Comm. Math. Phys. 303 (2011), no. 2, 555 -- 594.
[HvNVW2] T. Hytönen, J. van Neerven, M. Veraar and L. Weis. Analysis on Banach spaces, Volume
II: Probabilistic Methods and Operator Theory. Springer, 2018.
[Jan1] S. Janson. Gaussian Hilbert spaces. Cambridge Tracts in Mathematics, 129. Cambridge Uni-
versity Press, Cambridge, 1997.
[JLM] M. Junge, C. Le Merdy. Dilations and rigid factorisations on noncommutative Lp-spaces. J.
Funct. Anal. 249 (2007), 220 -- 252.
[JMX] M. Junge, C. Le Merdy and Q. Xu. H ∞ functional calculus and square functions on noncom-
mutative Lp-spaces. Astérisque No. 305 (2006).
[JMP2] M. Junge, T. Mei and J. Parcet. Noncommutative Riesz transforms-dimension free bounds
and Fourier multipliers. J. Eur. Math. Soc. (JEMS) 20 (2018), no. 3, 529 -- 595.
[KaL1] E. Kaniuth and A. T.-M. Lau. Fourier and Fourier-Stieltjes algebras on locally compact groups.
Mathematical Surveys and Monographs, 231. American Mathematical Society, Providence, RI,
2018.
[KaR1] R. V. Kadison and J. R. Ringrose. Fundamentals of the theory of operator algebras. Vol. I.
Elementary theory. Reprint of the 1983 original. Graduate Studies in Mathematics, 15. American
Mathematical Society, Providence, RI, 1997.
[KW] P. C. Kunstmann and L. Weis. Maximal Lp-regularity for parabolic equations, Fourier multiplier
theorems and H ∞-functional calculus. Functional analytic methods for evolution equations, Lect.
Notes in Math. 1855, 65 -- 311, in Springer (2004).
[LeM1] C. Le Merdy. Square functions, bounded analytic semigroups, and applications. Perspectives in
operator theory, 191 -- 220, Banach Center Publ., 75, Polish Acad. Sci. Inst. Math., Warsaw, 2007.
[Neer1] J. v. Neerven. Stochastic Evolution Equations. ISEM Lecture Notes 2007/08.
[Nua1] D. Nualart. The Malliavin calculus and related topics. Second edition. Springer-Verlag, Berlin,
2006.
[Osb1] M. S. Osborne. Locally convex spaces. Graduate Texts in Mathematics, 269. Springer, Cham,
2014.
[Ped1] G. K. Pedersen. C ∗-algebras and their automorphism groups. London Mathematical Society
Monographs, 14. Academic Press, Inc. [Harcourt Brace Jovanovich, Publishers], London-New York,
1979.
12
[PiX] G. Pisier and Q. Xu. Non-commutative Lp-spaces. 1459 -- 1517 in Handbook of the Geometry of
Banach Spaces, Vol. II, edited by W.B. Johnson and J. Lindenstrauss, Elsevier (2003).
[Ric] É. Ricard. A Markov dilation for self-adjoint Schur multipliers. Proc. Amer. Math. Soc. 136
(2008), no. 12, 4365 -- 4372.
[Str1] S. Stratila. Modular theory in operator algebras. Translated from the Romanian by the author.
Editura Academiei Republicii Socialiste România, Bucharest; Abacus Press, Tunbridge Wells,
1981.
[Sun] V. Sunder. An invitation to von Neumann algebras. Springer-Verlag, New York, 1987.
[Tak1] M. Takesaki. Theory of operator algebras. I. Reprint of the first (1979) edition. Encyclopaedia
of Mathematical Sciences, 124. Operator Algebras and Non-commutative Geometry, 5. Springer-
Verlag, Berlin, 2002.
[Tak2] M. Takesaki. Theory of operator algebras. II. Encyclopaedia of Mathematical Sciences, 125.
Operator Algebras and Non-commutative Geometry, 6. Springer-Verlag, Berlin, 2003.
[Tak3] M. Takesaki. Duality for crossed products and the structure of von Neumann algebras of type
III. Acta Math. 131 (1973), 249 -- 310.
Cédric Arhancet
13 rue Didier Daurat, 81000 Albi, France
URL: https://sites.google.com/site/cedricarhancet
[email protected]
13
|
1104.1216 | 2 | 1104 | 2011-05-28T04:30:31 | Residually finite actions and crossed products | [
"math.OA",
"math.DS"
] | We study a notion of residual finiteness for continuous actions of discrete groups on compact Hausdorff spaces and how it relates to the existence of norm microstates for the reduced crossed product. Our main result asserts that an action of a free group on a zero-dimensional compact metrizable space is residually finite if and only if its reduced crossed product admits norm microstates, i.e., is an MF algebra. | math.OA | math |
RESIDUALLY FINITE ACTIONS AND CROSSED PRODUCTS
DAVID KERR AND PIOTR W. NOWAK
Abstract. We study a notion of residual finiteness for continuous actions of discrete groups on
compact Hausdorff spaces and how it relates to the existence of norm microstates for the reduced
crossed product. Our main result asserts that an action of a free group on a zero-dimensional
compact metrizable space is residually finite if and only if its reduced crossed product admits
norm microstates, i.e., is an MF algebra.
1. Introduction
Finite-dimensional approximation is a ubiquitous notion in the structure theory of C ∗-algebras,
and it appears in a variety of different ways through properties like nuclearity, exactness, qua-
sidiagonality, and the existence of norm microstates (i.e., being an MF algebra) [11]. Given that
reduced crossed products by actions of countable discrete groups on compact metrizable spaces
play an important role as examples and motivation in the study of C ∗-algebras, one would like
to understand the extent to which various forms of finite-dimensional approximation in such
crossed products are reflections of finite approximation properties at the level of the dynamics.
Nuclearity and exactness are essentially measure-theoretic concepts and do not reflect any-
thing intrinsically topological in the dynamics. Indeed nuclearity of the reduced crossed prod-
uct is equivalent to the amenability of the action, which can be expressed in purely measure-
dynamical terms [1], while exactness of the reduced crossed product is equivalent to the exactness
of the acting group [11, Thm. 10.2.9], which can be characterized by the existence of an amenable
action of the group on a compact metrizable space.
On the other hand, quasidiagonality and the existence of norm microstates both involve
the matricial approximation of multiplicative structure and are thus topological in nature, and
their precise relation to the dynamics is for the most part poorly understood. Actually much
of the difficulty stems from the group itself, as anything but the simplest geometry in the
Cayley graph can cause severe complications for an operator analysis based on perturbations,
and obtaining a topological understanding of the reduced crossed product amounts in part to
knowning something about the reduced group C ∗-algebra, which sits inside the reduced crossed
product in a canonical way. For free groups on two or more generators it is a deep theorem
of Haagerup and Thorbjørnsen that the reduced group C ∗-algebra is MF [21]. For countable
discrete groups G, if the reduced group C ∗-algebra C ∗
λ(G) is quasidiagonal then G is amenable
[22], while C ∗
λ(G) is residually finite-dimensional if and only if G is residually finite and amenable
[14, Cor. 4]. Not much else seems to be known however about quasidiagonality and the existence
of norm microstates for reduced group C ∗-algebras.
One important dynamical setting in which we do have a complete understanding of quasidi-
agonality is that of integer actions. Pimsner showed in [31] that for a self-homeomorphism of
Date: April 5, 2011.
1
2
DAVID KERR AND PIOTR W. NOWAK
a compact metrizable space the following are equivalent: (i) the homeomorphism is pseudo-
nonwandering (i.e., chain recurrent), (ii) the crossed product is embeddable into an AF algebra,
(iii) the crossed product is quasidiagonal. Already though for Z2-actions it is not clear what
dynamical condition should correspond to quasidiagonality. Lin succeeded however in showing
that, for actions of Zd where d ≥ 1, embeddability of the crossed product into a simple AF
algebra is equivalent to the existence of an invariant Borel probability measure of full support
[26].
It is also known that almost periodic actions of certain amenable groups produce qua-
sidiagonal crossed products [30]. Note that in all of these cases the reduced and full crossed
products coincide and are nuclear due to the amenability of the acting group, and that a nuclear
C ∗-algebra is quasidiagonal if and only if it admits norm microstates, i.e., is an MF algebra.
The goal of the present paper is to initiate a dynamical study of the MF property in reduced
crossed products that is primarily targeted at actions of free groups. Such a program owes
its very possibility to the Haagerup-Thorbjørnsen result cited above which establishes the MF
property of the reduced group C ∗-algebras of free groups on two or more generators, a conclusion
whose validity is still unknown for the vast majority of nonamenable groups. For actions of a
countable discrete group G on a compact metric space X we define a dynamical version of
residual finiteness for groups, which we again call residual finiteness (Definition 2.1), and our
main result asserts that, when G is free and X is zero-dimensional, the action is residually finite
if and only if the reduced crossed product is an MF algebra (Theorem 3.10).
For G = Z residual finiteness is the same as chain recurrence and the situation reduces to
a special case of the equivalence of (i) and (iii) in the statement of Pimsner's theorem above.
Pimnser showed in [31, Lemma 2] that, for a Z-action, chain recurrence is equivalent to every
open set being topologically incompressible, and he establish the implication (iii)⇒(i) by using
compressibility to construct a nonunitary isometry in the crossed product as an obstruction
to quasidiagonality. For a free group Fr with r ≥ 2 we do not have an analogue of this non-
compressibility characterization for residual finiteness, and so our proof that an Fr-action on a
zero-dimensional space is residually finite if the reduced crossed product is MF must proceed by
different means. The idea is to directly extract the finite approximations of the action from the
existence of norm microstates. To carry out the required perturbation arguments we need to be
able to work with projections, which explains our hypothesis that X is zero-dimensional.
A residually finite action of G on X is, roughly speaking, one that approximately admits
extensions consisting of actions of G on finite sets, i.e., there are actions of G on finite sets which
map into X approximately equivariantly with approximately dense image (see Section 2). This
can be viewed as a topological analogue of soficity for actions preserving a Borel probability
measure, in which the approximations are measured in 2-norm and the maps into X must
approximately push forward the uniform measure to the given measure on X. One can count
the exponential growth of the number of such approximately equivariant maps up to some
observational error to define a notion of entropy (see [8, 25] and especially Sections 2 and 3 of
[24]) and this is one motivation for the study of finite dynamical approximations. It can be shown
that all measure-preserving actions of Fr on a standard probability space are sofic (see [7]), and
that the von Neumann algebra crossed product of such actions admit tracial microstates, i.e.,
embed into the ultrapower Rω of the hyperfinite II1 factor (use [10]). So residual finiteness in our
topological context can be thought of as playing the role of measure-preservingness, while the
MF property (existence of norm microstates) is the analogue of embeddability into Rω (existence
of tracial microstates). Note however that in our setting one of the main points is to show that
RESIDUALLY FINITE ACTIONS
3
the MF property implies residual finiteness, while in the von Neumann algebra case there is
nothing to prove in this direction.
After introducing residually finite actions in Section 2, we show in the first part of Section 3
that such actions give rise to an MF reduced crossed product whenever the reduced group
C ∗-algebra of the acting group is MF. In fact we prove this more generally in Theorem 3.4 for
actions on arbitrary separable C ∗-algebras which are quasidiagonal in the sense of Definition 3.2.
Using this fact we then proceed in the second part of Section 3 to establish the equivalence of
residual finiteness and the MFness of the reduced crossed product in the case of a free group
acting on a zero-dimensional compact metrizable space (Theorem 3.10). Motivated by recent
work of Rørdam and Sierakowski on paradoxical decompositions in the context of continuous
actions on the Cantor set and purely infinite crossed products [34], we examine in Section 4 how
paradoxical decomposability fits into our discussion of residual finiteness at the other extreme.
Section 5 gives a list of characterizations of residual finiteness for minimal actions of free groups
which incorporates paradoxical decomposability among other phenomena. In Section 6 we use
spaces of probability measures to construct, for every nonamenable countable discrete group, an
example of an action which is not residually finite although its restriction to any cyclic subgroup
is residually finite. Finally, in Section 7 we revisit integer actions and observe in this case,
combining arguments and results from [22, 3, 16], that the following conditions are equivalent:
(i) the crossed product is a strong NF algebra, (ii) the OL∞ invariant of the crossed product is
1, and (iii) there is collection of transitive residually finite subsystems with dense union.
We remark that Margulis and Vinberg defined in [29] a nontopological notion of residual
finiteness for actions in which the maps in the finite modelling go in the other direction.
We round out the introduction with a few words about notational convention. Throughout
the paper G will always be a countable discrete group, with extra hypotheses added explicitly
whenever appropriate. Actions of G on a compact Hausdorff space X will often be unnamed,
in which case simply write G y X, and they will always be expressed using the concatenation
(s, x) 7→ sx for x ∈ X and s ∈ G. When necessary we will refer to actions by means of a
symbol such as α, in particular when we need to talk about the induced action on C(X), which
will actually be expressed using this symbol, i.e., αs(f )(x) = f (s−1x) for f ∈ C(X), x ∈ X,
and s ∈ G. We write λ for the left regular representation of G on ℓ2(G). The reduced crossed
product of a continuous action G y X will be written C(X) ⋊λ G and the reduced group C ∗-
algebra of G will be written C ∗
λ(G) associated
to a group element s will invariably be denoted by us. For background on crossed products and
group C ∗-algebras, especially from the kind of finite-dimensional approximation viewpoint of
this paper, we refer the reader to [11].
λ(G). The canonical unitary in C(X) ⋊λ G or C ∗
For a compact Hausdorff space we write MX for the space of Borel probability measures on
X equipped with the weak∗ topology, under which it is a compact Hausdorff space. Whenever
convenient we will simultaneously regard elements of MX as states on C(X). The unitary group
of a unital C ∗-algebra A will be written U(A).
Acknowledgements. The first author was partially supported by NSF grant DMS-0900938 and
the second author partially supported by NSF grant DMS-0900874. We thank Adam Sierakowski
for comments and corrections, and Hanfeng Li for the argument in Example 2.5 and the simple
proof of Lemma 5.1.
4
DAVID KERR AND PIOTR W. NOWAK
2. Residually finite actions
Definition 2.1. A continuous action of G on a compact Hausdorff space X is said to be
residually finite if for every finite set F ⊆ G and neighbourhood ε of the diagonal in X × X
there are a finite set E, an action of G on E, and a map ζ : E → X such that ζ(E) is ε-dense
in X and (ζ(sz), sζ(z)) ∈ ε for all z ∈ E and s ∈ F .
Note that if S is a generating set for G then to verify residual finiteness it is sufficient for F
in the definition to be quantified over all finite subsets of S by uniform continuity.
We will mostly be interested in actions on compact metrizable spaces, in which case residual
finiteness can be expressed in terms of a given compatible metric. Indeed if X is metrizable in
the above definition and d is a compatible metric on X then residual finiteness is the same as
saying that for every finite set F ⊆ G and ε > 0 there are a finite set E, an action of G on E,
and a map ζ : E → X such that ζ(E) is ε-dense in X and d(ζ(sz), sζ(z)) < ε for all z ∈ E and
s ∈ F .
Note that if G admits a free residually finite continuous action on a compact metrizable space,
then G must be a residually finite group, as is easy to see. Indeed this is the motivation for our
terminology in the dynamical setting.
In the case G = Z, residual finiteness is equivalent to chain recurrence, as explained at the
beginning of Section 7.
If X has no isolated points then a simple perturbation argument shows that in Definition 2.1
we can always take the set E to be a subset of X and the map ζ : E → X to be the inclusion.
However, for general X this is not possible, as the following example demonstrates. For each
i = 1, 2, 3 take Xi to be a copy of Z and consider the two-point compactification Xi ∪ {yi, zi}
where yi and zi are the points at −∞ and +∞. Let Z act on Xi ∪ {yi, zi} so that it translates
Xi via addition and fixes yi and zi. Now define X by taking the quotient disjoint union of the
sets Xi ∪ {yi, zi} for i = 1, 2, 3 which collapses each of the sets {z1, y2, y3} and {y1, z2, z3} to
single points.
The following proposition shows that in the definition of residual finiteness it is enough to
verify the ε-density condition in a local way.
Proposition 2.2. A continuous action of G on a compact Hausdorff space X is residually finite
if and only if for every x ∈ X, finite set F ⊆ G, and neighbourhood ε of the diagonal in X × X
there are a finite set E, an action of G on E, and a map ζ : E → X such that x lies in the
ε-neighbourhood of ζ(E) and (ζ(sz), sζ(z)) ∈ ε for all z ∈ E and s ∈ F .
Proof. For the nontrivial direction, given a finite set F ⊆ G and neighbourhood ε of the diagonal
in X × X take, for every x in some finite ε-dense set D ⊆ X, a finite set Ex and an action of
G on Ex, and a map ζx : Ex → X such that x lies in the ε-neighbourhood of ζx(Ex) and
(ζx(sz), sζx(z)) ∈ ε for all z ∈ Ex and s ∈ F . Now set E = `x∈D Ex and define ζ : E → X so
that it restricts to ζx on Ex for each x ∈ D. Then ζ(E) is ε-dense in X and (ζ(sz), sζ(z)) ∈ ε
for all z ∈ E and s ∈ F , verifying residual finiteness.
(cid:3)
Proposition 2.3. Let X be a compact Hausdorff space and G y X a residually finite continuous
action. Then there is a G-invariant Borel probability measure on X.
Proof. Let Λ be the net of pairs (F, ε) where F is a finite subset of G and ε is a neighbourhood of
the diagonal in X × X and the order relation (F ′, ε′) ≻ (F, ε) means that F ′ ⊇ F and ε′ ⊆ ε. For
RESIDUALLY FINITE ACTIONS
5
a given (F, ε) ∈ Λ there exist, by residual finiteness, a finite set D, an action of G on E, and a
map ζ : E → X such that (ζ(sz), sζ(z)) ∈ ε for all z ∈ E and s ∈ F , and we define µF,ε to be the
pullback under ζ of the uniform probability measure on E, i.e., µF,ε(f ) = E−1Pz∈E f (ζ(z))
for f ∈ C(X). Now take a weak∗ limit point of the net {µF,ε}(F,ε)∈Λ, which is easily verified to
be a G-invariant Borel probability measure on X.
(cid:3)
Example 2.4. Suppose that G is residually finite. Let X be a compact Hausdorff space. Then
the Bernoulli action G y X G given by s(xt)t∈G = (xs−1t)t∈G is easily seen to be residually finite.
Example 2.5. Let f be an element in the group ring ZG. Then G acts on ZG/ZGf by left
translation, and this gives rise to an action αf of G by automorphisms on the compact Abelian
dual group Xf := \ZG/ZGf . When f is equal to d times the unit for some d ∈ N we obtain the
Bernoulli action G y {1, . . . , d}G. Suppose now that G is residually finite and f is invertible as
an element in the full group C ∗-algebra C ∗(G). Then the action αf is residually finite. Indeed
in this case the points in Xf which are fixed by some finite-index normal subgroup of G are
dense in Xf , which can be seen as follows. Let H be a finite-index normal subgroup of G.
Then the canonical surjective ring homomorphism πH : ZG → Z(G/H) induces a surjective
ring homomorphism ZG/ZGf → Z(G/H)/Z(G/H)πH (f ) which in turn induces an injective
group homomorphism XπH (f ) ֒→ Xf . Note that XπH (f ), identified with its image in Xf , is equal
to the set of points in Xf which are fixed by H. It thus suffices to show that SH∈H XπH (f )
happens precisely when the natural homomorphism ZG/ZGf →QH∈H
is dense in Xf where H denotes the collection of finite-index normal subgroups of G. This
Z(G/H)/Z(G/H)πH (f )
is injective. To verify this injectivity, let g be an element in the kernel. Then for every H ∈ H
there is a wH ∈ Z(G/H) such that πH(g) = wH πH(f ). Since f is invertible in C ∗(G), wH
is unique and its ℓ2-norm is bounded above by some constant not depending on H. Now if
H1, H2 ∈ H and H1 ⊆ H2 then the image of wH1 in Z(G/H2) under the canonical map
Z(G/H1) → Z(G/H2) is equal to wH2, and so we can define the projective limit of the wH as an
element w in ZG. Then g = wf . Moreover, w lies in ℓ2(G) because the ℓ2-norms of the wH are
uniformly bounded, and so w has finite support since it takes integer values. Therefore g ∈ ZGf
and we obtain the desired injectivity.
We remark that when f is invertible in ℓ1(G) and {Gi}∞
subgroups of G with T∞
i=1 is a sequence of finite-index normal
i=j Gi = {e}, the measure entropy of αf with respect to the Haar
measure and the sofic approximation sequence Σ arising from {Gi}∞
i=1 is equal to the exponential
growth rate of the number of Gi-fixed points and to the logarithm of the Fuglede-Kadison
determinant of f in the group von Neumann algebra of G [8]. This is also true for the topological
entropy with respect to Σ more generally whenever f is invertible in the full group C ∗-algebra
[25].
j=1S∞
3. Residually finite actions and MF algebras
A separable C ∗-algebra is said to be an MF algebra if it can be expressed as the inductive
limit of a generalized inductive system of finite-dimensional C ∗-algebras [6, Defn. 3.2.1]. This is
n=1 in
N [6, Thm. 3.2.2], as well as to the existence of norm microstates (see Section 11.1 of [11]). The
following characterizations are tailored for our purposes.
equivalent to the embeddability of A into Q∞
n=1 Mkn for some sequence {kn}∞
n=1 Mkn/L∞
Proposition 3.1. Let A be a separable C ∗-algebra. The following are equivalent.
6
DAVID KERR AND PIOTR W. NOWAK
(1) A is an MF algebra,
(2) for every ε > 0 and finite set Ω ⊆ A there are a k ∈ N and a ∗-linear map ϕ : A → Mk
(3) there is a dense ∗-subalgebra A0 of A such that for every ε > 0 and finite set Ω ⊆ A0
there are a k ∈ N and a ∗-linear map ϕ : A0 → Mk such that kϕ(ab) − ϕ(a)ϕ(b)k < ε
such that kϕ(ab) − ϕ(a)ϕ(b)k < ε and (cid:12)(cid:12)kϕ(a)k − kak(cid:12)(cid:12) < ε for all a, b ∈ Ω,
and (cid:12)(cid:12)kϕ(a)k − kak(cid:12)(cid:12) < ε for all a, b ∈ Ω.
Moreover, if A is unital then the maps ϕ in (2) and (3) may be taken to be unital.
n=1 Mkn for some
n=1 in such a way that for each a ∈ A one has limn→∞ kank = kak for every
Proof. (1)⇒(2). By Proposition 2.2.3(iii) of [6], A embeds into Q∞
(an) ∈ π−1({a}) where π :Qλ Mkn →Qλ Mkn/Lλ Mkn is the quotient map. Take a linear map
ψ : A →Q∞
n=1 Mkn such that π ◦ ψ = Φ. We may assume ψ to be ∗-linear by replacing it with
a 7→ (ψ(a) + ψ(a∗)∗))/2 if necessary. For each n let ψn : A → Mkn be the composition of ψ with
the projection onto the nth factor. Then we have limn→∞ kψn(a)k = kak and limn→∞ kψn(ab) −
ψn(a)ψn(b)k = 0 for all a, b ∈ A, from which (2) follows.
n=1 Mkn/L∞
sequence {kn}∞
n=1 Ωn
all a, b ∈ Ωn
yielding (1).
n. Now define a map θ : A →Q∞
(2)⇒(3). Trivial.
(3)⇒(1). Since A is separable there is an increasing sequence Ω1 ⊆ Ω2 ⊆ · · · of finite subsets
n is a dense ∗-subalgebra A1 of A. For each n ∈ N take a kn ∈ N and a
of A0 such that S∞
∗-linear map ϕn : A0 → Mkn such that kϕn(ab) − ϕn(a)ϕn(b)k < ε and (cid:12)(cid:12)kϕn(a)k − kak(cid:12)(cid:12) < ε for
with the canonical projection mapQ∞
multiplicative, and hence extends to an injective ∗-homomorphism A →Q∞
to be unital, since we can lift the image of the unit under Φ to a projection (pn) in Qλ Mkn,
n=1 pnMknpn/L∞
producing an injective unital ∗-homomorphism from A toQ∞
n=1 pnMknpn. This
enables us to arrange ψ to be unital, so that each ψn is unital. It follows that the maps ϕ in
conditions (2) and (3) may be taken to be unital.
(cid:3)
n=1 Mkn by composing a 7→ (ϕn(a))∞
n=1
n=1 Mkn. Then θ is isometric and
k=1 Mkn,
In the case that A is unital, note that in the proof of (1)⇒(2) the embedding Φ may be taken
n=1 Mkn →Q∞
n=1 Mkn(cid:14)L∞
n=1 Mkn(cid:14)L∞
n=1 Mkn(cid:14)L∞
Note that the maps ϕ in Proposition 3.1 are typically not bounded.
A C ∗-algebra is said to be quasidiagonal if it admits a faithful representation whose image is a
quasidiagonal set of operators. Voiculescu showed that a separable C ∗-algebra A is quasidiagonal
if and only if for every ε > 0 and finite set Ω ⊆ A there are a n ∈ N and a contractive completely
positive map ϕ : A → Mn such that kϕ(ab) − ϕ(a)ϕ(b)k < ε and kϕ(a)k ≥ kak − ε for all a, b ∈ Ω
[39]. Blackadar and Kirchberg showed that a separable nuclear C ∗-algebra is an MF algebra if
and only if it is quasidiagonal [6, Thm. 5.2.2]. The reduced group C ∗-algebra of a free group on
two or more generators is an example of an MF algebra which is not quasidiagonal [21].
Voiculescu's abstract characterization of quasdiagonality motivates the following definition,
which extends the concept of residual finiteness to actions on noncommutative C ∗-algebras in
view of Proposition 3.3. We will formulate and prove Theorems 3.4 and 3.5 within this general
noncommutative framework.
Definition 3.2. Let α be an action of G on a separable C ∗-algebra A. We say that α is
quasidiagonal if for every finite set Ω ⊆ A, finite set F ⊆ G, and ε > 0 there are an d ∈ N, an
action γ of G on Md, and a unital completely positive map ϕ : A → Md such that
(1) kϕ(ab) − ϕ(a)ϕ(b)k < ε for all a, b ∈ Ω,
RESIDUALLY FINITE ACTIONS
7
(2) kϕ(a)k ≥ kak − ε for all a ∈ Ω, and
(3) kϕ(αs(a)) − γs(ϕ(a))k < ε for all a ∈ Ω and s ∈ F .
Note that the existence of a quasidiagonal action on A in the sense of the above definition
implies that A is quasidiagonal as a C ∗-algebra.
Proposition 3.3. Let α be a residually finite continuous action of G on a compact metrizable
space X. Then α is quasidiagonal as an action of G on C(X).
Proof. Fix a compatible metric d on X. Let ε > 0 and let Ω be a finite subset of C(X). By
uniform continuity there is a δ > 0 such that f (x) − f (y) < ε for all f ∈ Ω and all points
x, y ∈ X satisfying d(x, y) < δ. By residual finiteness there are a finite set E, an action γ of G
on E, and a map ζ : E → X such that ζ(E) is δ-dense in X and d(ζ(s−1z), s−1ζ(z)) < δ for all
z ∈ E and s ∈ F . Define ϕ : C(X) → C(E) by ϕ(f )(z) = f (ζ(z)) for all f ∈ C(X) and z ∈ E.
Then ϕ is a homomorphism, and for all f ∈ Ω and s ∈ F we have
and
kϕ(f )k = sup
f (x) ≥ kf k − ε
x∈ζ(E)
showing that α is quasidiagonal.
(cid:3)
kϕ(αs(f )) − γs(ϕ(f ))k = sup
z∈E(cid:12)(cid:12)f (s−1ζ(z)) − f (ζ(s−1z)(cid:12)(cid:12) < ε,
The proof of the following theorem is reminiscent of part of the proof of Theorem 7 in [31].
Theorem 3.4. Suppose that C ∗
a separable C ∗-algebra A. Then A ⋊λ G is an MF algebra.
λ(G) is an MF algebra. Let α be a quasidiagonal action of G on
Proof. We view A as acting on a separable Hilbert space H via some faithful essential represen-
tation and A ⋊λ G as acting on H ⊗ ℓ2(G) in the standard way, as determined by aus(ξ ⊗ δt) =
α−1
st (a)ξ ⊗ δst for a ∈ A, s, t ∈ G, and ξ ∈ H, where {δs : s ∈ G} is the canonical basis of ℓ2(G).
Since α is quasidiagonal there exist, for each n ∈ N, a positive integer kn, an action γn of G
on Mkn, and a unital completely positive map ϕn : A → Mkn so that
(1) limn→∞ kϕn(ab) − ϕn(a)ϕn(b)k = 0 for all a, b ∈ A,
(2) limn→∞ kϕn(a)k = kak for all a ∈ A, and
(3) limn→∞ kϕn(αs(a)) − γn,s(ϕn(a))k = 0 for all a ∈ A and s ∈ G.
Note that if we view Mkn as acting on L2(Mkn, τ ) via the GNS construction with respect to the
unique tracial state τ , then the action γ is implemented through conjugation by the unitaries ws
for s ∈ G defined by wsxη = γs(x)η where η is the canonical cyclic vector. Thus by replacing Mkn
with B(L2(Mkn, τ )) ∼= Mk2
and relabeling we may assume that there are unitary representations
wn : G → U(ℓkn
2 ) such that
n
lim
n→∞
kϕn(αs(a)) − wn,sϕn(a)w∗
n,sk = 0
for all a ∈ A and s ∈ G.
Let Ω be a finite subset of the algebraic crossed product A ⋊alg G, which by definition consists
of sums of the form Ps∈F asus where F is a finite subset of G and each as lies in A. Let
ε > 0. We will show the existence of a d ∈ N and ∗-linear map β : A ⋊alg G → Md such that
8
DAVID KERR AND PIOTR W. NOWAK
the proof by Proposition 3.1.
Regard Mkn as acting on ℓkn
2
in the standard way. For N ∈ N we define the unital completely
kβ(cc′) − β(c)β(c′)k < ε and (cid:12)(cid:12)kβ(c)k − kck(cid:12)(cid:12) < ε for all c, c′ ∈ Ω. This is sufficient to complete
positive map ΦN : A →L∞
wn,s ⊗ λs for s ∈ G. Define a ∗-linear map ΘN : A ⋊alg G →L∞
2 ⊗ ℓ2(G)) generated by Mkn ⊗ 1 and the operators
ΘN (aus) = (ϕN (a)wN,s ⊗ λs, ϕN +1(a)wN +1,s ⊗ λs, . . . )
Write Dn for the C ∗-subalgebra of B(ℓkn
ΦN (a) = (ϕN (a), ϕN +1(a), . . . ).
n=N Dn by setting
n=N Mkn by
for a ∈ A and s ∈ G and extending linearly, which we can do since the subspaces Aus for s ∈ G
are orthogonal with respect to the canonical conditional expectation from A ⋊λ G onto A. We
aim to show the existence of an N ∈ N such that
(1) kΘN (cc′) − ΘN (c)ΘN (c′)k < ε/2 for all c, c′ ∈ Ω, and
(2) (cid:12)(cid:12)kΘN (c)k − kck(cid:12)(cid:12) < ε/2 for all c ∈ Ω.
Since Ω is finite we can find an ε′ > 0 such that, for all n ∈ N and c ∈ Ω, if(cid:12)(cid:12)kΘn(c)k2 −kck2(cid:12)(cid:12) <
ε′ then (cid:12)(cid:12)kΘn(c)k − kck(cid:12)(cid:12) < ε/2. Take a finite set F ⊆ G such that for every c ∈ Ω we can write
c = Ps∈F ac,sus where ac,s ∈ A for each s ∈ F . Set M = maxc∈Ω,s∈F kac,sk. For each c ∈ Ω
take a unit vector ηc in H ⊗ ℓ2(G) such that kck ≤ kcηck + ε′/2 and ηc =Pt∈K ξc,t ⊗ δt for some
finite set K ⊆ G and vectors ξc,t ∈ H. For notational simplicity we may assume K to be the
same for all c ∈ Ω.
Take a δ > 0 such that (1 + 2F K)F 2δ2 ≤ ε′/2. Since the unital completely positive maps
ϕn are asymptotically multiplicative and asymptotically isometric, by the version of Voiculescu's
theorem which appears as Theorem 2.10 in [9] and stems from [35] we can find a N ∈ N and a
n=N ℓkn
2 → H such that
unitary U :L∞
for all c ∈ Ω, s ∈ F , and t ∈ K. For s ∈ G set ws = (wN,s, wN +1,s, . . . ) ∈ L∞
asymptotic equivariance of the maps ϕk, we may assume that N is large enough so that
st (ac,s))U −1 − α−1
δ
2
n=N Mkn. By the
(cid:13)(cid:13)U ΦN (α−1
(cid:13)(cid:13)ΦN (α−1
(cid:13)(cid:13)wstU −1α−1
st (ac,s)(cid:13)(cid:13) <
stΦN (ac,s)wst(cid:13)(cid:13) <
st (ac,s) − ΦN (ac,s)wstU −1(cid:13)(cid:13) < δ
st (ac,s)) − w∗
δ
2
and hence
for all c ∈ Ω, s ∈ F , and t ∈ K. We may moreover assume that N is large enough so that
kΘN (cc′) − ΘN (c)ΘN (c′)k < ε/2 for all c, c′ ∈ Ω, since for a, b ∈ A, and s, t ∈ G we have
kΘN (ausbut) − ΘN (aus)ΘN (but)k = sup
n≥N
≤ sup
n≥N
kϕn(aαs(b))wn,st − ϕn(a)wn,sϕn(b)wn,tk
k(ϕn(aαs(b)) − ϕn(a)ϕn(αs(b)))wn,stk
+ sup
n≥N
kϕn(a)(ϕ(αs(b))wn,s − wn,sϕn(b))wn,tk,
where the last two suprema tend to zero as N → ∞, and the subspaces Aus for s ∈ G linearly
span A ⋊alg G.
RESIDUALLY FINITE ACTIONS
9
Let c ∈ Ω. Write U for the unitary operator from (L∞
t ζ ⊗ δt. Since ηc is a unit vector, for s ∈ F we have
U (ζ ⊗ δt) = U w−1
n=N ℓkn
2 ) ⊗ ℓ2(G) to H ⊗ ℓ2(G) given by
(cid:13)(cid:13)(cid:13)(cid:13)Xt∈K(cid:0)wstU −1α−1
2
st (ac,s) − ΦN (ac,s)wstU −1(cid:1)ξc,t ⊗ δst(cid:13)(cid:13)(cid:13)(cid:13)
= Xt∈K(cid:13)(cid:13)(cid:0)wstU −1α−1
≤ Xt∈K(cid:13)(cid:13)wstU −1α−1
st (ac,s) − ΦN (ac,s)wstU −1(cid:1)ξc,t(cid:13)(cid:13)
st (ac,s) − ΦN (ac,s)wstU −1(cid:13)(cid:13)
< δ2.
2
2kξc,tk2
For any vectors x1, . . . , xn, y1, . . . , yn ∈ H we have
(∗)
we obtain
and so applying this inequality and the crude bound
2
n
n
2
n
n
n
2
n
(cid:13)(cid:13)(cid:13)(cid:13)
Xi=1
xi(cid:13)(cid:13)(cid:13)(cid:13)
(cid:19)2
≤(cid:18)(cid:13)(cid:13)(cid:13)(cid:13)
yi(cid:13)(cid:13)(cid:13)(cid:13)
+(cid:13)(cid:13)(cid:13)(cid:13)
(xi − yi)(cid:13)(cid:13)(cid:13)(cid:13)
Xi=1
Xi=1
+(cid:18)1 + 2(cid:13)(cid:13)(cid:13)(cid:13)
yi(cid:13)(cid:13)(cid:13)(cid:13)
(cid:19)(cid:13)(cid:13)(cid:13)(cid:13)
yi(cid:13)(cid:13)(cid:13)(cid:13)
≤(cid:13)(cid:13)(cid:13)(cid:13)
(xi − yi)(cid:13)(cid:13)(cid:13)(cid:13)
Xi=1
Xi=1
Xi=1
ΦN (ac,s)wstU −1ξc,t ⊗ δst(cid:13)(cid:13)(cid:13)(cid:13)
kΘN (c) U −1ηck =(cid:13)(cid:13)(cid:13)(cid:13)Xs∈F Xt∈K
≤Xs∈F Xt∈K
st (ac,s)ξc,t ⊗ δst(cid:13)(cid:13)(cid:13)(cid:13)
kcηck2 =(cid:13)(cid:13)(cid:13)(cid:13)
U −1Xs∈F Xt∈K
st (ac,s)ξc,t ⊗ δst(cid:13)(cid:13)(cid:13)(cid:13)
=(cid:13)(cid:13)(cid:13)(cid:13)Xs∈F Xt∈K
≤ kΘN (c) U −1ηck2 +(cid:0)1 + 2kΘN (c) U −1ηck(cid:1)
×(cid:18)Xs∈F(cid:13)(cid:13)(cid:13)(cid:13)Xt∈K(cid:0)wstU −1α−1
≤ kΘN (c)k2 + (1 + 2F K)F 2δ2
wstU −1α−1
α−1
2
2
kΦN (ac,s)k ≤ F K.
(cid:19)2
st (ac,s) − ΦN (ac,s)wstU −1(cid:1)ξc,t ⊗ δst(cid:13)(cid:13)(cid:13)(cid:13)
= kΘN (c)k2 +
ε′
2
.
2 ≥ kcηck2 − ε′/2 ≥ kck2 − ε′.
Consequently (cid:13)(cid:13)ΘN (c)(cid:13)(cid:13)
Next let us show that kck2 ≥ kΘN (c)k2 − ε′. Take a unit vector η in H ⊗ ℓ2(G) such that
kΘN (c)k ≤ kΘN (c) U −1ηk + ε′/2. The idea is to argue as in the above paragraph, reversing the
roles of c and ΘN (c) and replacing ηc with η. Notice that the only way the particular choice of
the vectors ηc entered into the above estimates, besides their being of norm one, was in obtaining
the bound (∗). Here however we can simply take k U −1cηk ≤ kck as the replacement for this
10
DAVID KERR AND PIOTR W. NOWAK
bound, in which case
kΘN (c) U −1ηk2 =(cid:13)(cid:13)(cid:13)(cid:13)Xs∈F Xt∈K
2
≤ k U −1cηk2 +(cid:0)1 + 2k U −1cηk(cid:1)
ΦN (ac,s)wstU −1ξc,t ⊗ δst(cid:13)(cid:13)(cid:13)(cid:13)
×(cid:18)Xs∈F(cid:13)(cid:13)(cid:13)(cid:13)Xt∈K(cid:0)ΦN (ac,s)wstU −1 − wstU −1α−1
≤ kck2 + 3F 2δ2
(cid:19)2
st (ac,s)(cid:1)ξc,t ⊗ δst(cid:13)(cid:13)(cid:13)(cid:13)
= kck2 +
ε′
2
and hence kck2 ≥ kΘN (c) U −1ηk2 − ε′/2 ≥ kΘN (c)k2 − ε′. By our choice of ε′ we conclude that
for a ∈ A and s ∈ G and extending linearly. In view of the properties of ΘN we have kθ(cc′) −
Now let M be an integer greater than or equal to N to be determined shortly. Set D =
n=N Dn. Define the ∗-linear map θ : A ⋊alg G → D by declaring that
θ(aus) = (ϕN (a)wn,s ⊗ λs, . . . , ϕM (a)wn,s ⊗ λs)
(cid:12)(cid:12)kΘN (c)k − kck(cid:12)(cid:12) < ε/2 for all c ∈ Ω, so that the map ΘN satisfies the desired properties.
LM
θ(c)θ(c′)k < ε/3 for all c, c′ ∈ Ω and, by taking M large enough, (cid:12)(cid:12)kθ(c)k − kck(cid:12)(cid:12) < ε/2 for all
Set k = PM
represented on (LM
LM
λ(G) as canonically
2 ) ⊗ ℓ2(G) where the latter is identified in the standard way with
2 ) with Mk we view D as a
λ(G), for
λ(G) is MF then
n=N kn. Note that D is a C ∗-subalgebra of B(LM
2 ⊗ ℓ2(G)). Via some fixed identification of B(LM
λ(G). By hypothesis C ∗
n=1 Mkn is an embedding witnessing the fact that C ∗
λ(G) is MF, and hence so is Mk ⊗ C ∗
n=N (ℓkn
n=N ℓkn
2 ) ⊗ C ∗
n=N ℓkn
n=N ℓkn
c ∈ Ω.
C ∗-subalgebra of Mk ⊗ C ∗
if C ∗
we obtain an embedding
λ(G) ֒→Q∞
n=1 Mkn(cid:14)L∞
λ(G) ֒→ Mk ⊗ ∞
Yn=1
Mk ⊗ C ∗
Mkn(cid:30) ∞
Mn=1
Mkn! ∼=
∞
Yn=1(cid:0)Mk ⊗ Mkn(cid:1)(cid:30) ∞
Mn=1(cid:0)Mk ⊗ Mkn(cid:1).
Thus by Proposition 3.1 there is a d ∈ N and a ∗-linear map ϕ : Mk ⊗ C ∗
for all c, c′ ∈ Ω,
λ(G) → Md such that,
(1) kϕ(θ(c)θ(c′)) − ϕ(θ(c))ϕ(θ(c′))k < ε/3,
(2) kϕ(θ(cc′) − θ(c)θ(c′))k < kθ(cc′) − θ(c)θ(c′)k + ε/3, and
Set β = ϕ ◦ θ. Then β is ∗-linear, and for all c, c′ ∈ Ω we have
(3) kϕ(θ(c))k − kθ(c)k(cid:12)(cid:12) < ε/2.
kβ(cc′) − β(c)β(c′)k ≤ kϕ(θ(cc′) − θ(c)θ(c′))k + kϕ(θ(c)θ(c′)) − ϕ(θ(c))ϕ(θ(c′))k
≤ kθ(cc′) − θ(c)θ(c′)k +
ε
3
+
ε
3
< ε
and
completing the proof.
(cid:12)(cid:12)kβ(c)k − kck(cid:12)(cid:12) ≤(cid:12)(cid:12)kϕ(θ(c))k − kθ(c)k(cid:12)(cid:12) +(cid:12)(cid:12)kθ(c)k − kck(cid:12)(cid:12) <
ε
2
+
ε
2
= ε,
(cid:3)
RESIDUALLY FINITE ACTIONS
11
λ(G) is quasidiagonal. Let α be a quasidiagonal action of G on
Theorem 3.5. Suppose that C ∗
a separable nuclear C ∗-algebra A. Then A ⋊λ G is quasidiagonal.
Proof. By a result of Rosenberg (see the appendix of [22]), the quasidiagonality of C ∗
λ(G) implies
that G is amenable. Since A is nuclear, it follows that the crossed product A ⋊λ G is nuclear
(see Section IV.3.5 of [2]). Since C ∗
λ(G) is quasidiagonal it is an MF algebra, and so A ⋊λ G
is an MF algebra by Theorem 3.4. Since separable nuclear MF algebras are quasidiagonal [6,
Thm. 5.2.2], we conclude that A ⋊λ G is quasidiagonal.
(cid:3)
An action of G on a compact Haudorff space X is said to be topologically free if the set of
points in X with trivial isotropy group is dense.
Corollary 3.6. Suppose that G is amenable. Let G y X be a topologically free residually finite
action on a compact metrizable space. Then C(X) ⋊λ G is quasidiagonal.
Proof. Because it admits a topologically free residually finite action, G must be a residually
finite group, as is easy to verify. Since the full and reduced group C ∗-algebras of G coincide by
amenability, it follows that C ∗
λ(G) is quasidiagonal [14, Cor. 4]. Since C(X) is nuclear we obtain
the conclusion by Theorem 3.5.
(cid:3)
Now we fix an r ∈ {2, 3, . . . , ∞} and concentrate on actions of the free group Fr for the remain-
der of the section. Our aim is to show that, for continuous actions of Fr on a zero-dimensional
compact metrizable space, residual finiteness is equivalent to the reduced crossed product being
an MF algebra. Note that such a crossed product cannot be quasidiagonal since it contains
C ∗
λ(Fr), which, by a result of Rosenberg [22], is not quasidiagonal since Fr is nonamenable. The
property of being an MF algebra is the appropriate substitute for quasidiagonality in the nona-
menable case. Indeed quasidiagonality and the property of being an MF algebra are equivalent
for separable nuclear C ∗-algebras [6, Thm. 5.2.2], and so quasidiagonality in the general group
action setting can viewed as an artifact of nuclearity.
The following two lemmas are standard types of perturbation results.
Lemma 3.7. Let η > 0. Then there is a δ > 0 such that whenever d ∈ N and p and q are
projections in Md satisfying kpqk < δ there exists a projection q′ ∈ (1 − p)Md(1 − p) such that
kq′ − qk < η.
Proof. Let δ be a strictly positive number less than η/6 to be further specified. Let p and q be
projections in some matrix algebra Md such that kpqk < δ. Set a = (1 − p)q(1 − p). Then
kq − ak = kpq + qp − pqpk ≤ 3kpqk < 3δ.
Since
ka2 − ak ≤ ka2 − q2k + kq − ak < k(a − q)ak + kq(a − q)k + 3δ < 9δ,
if δ is small enough as a function of η there exists, by the functional calculus, a projection
q′ ∈ (1−p)Md(1−p) such that kq′−ak < η/2. Since δ < η/6, we have kq′−qk ≤ kq′−ak+ka−qk <
η/2 + η/2 = η, as desired.
(cid:3)
Lemma 3.8. Let n ∈ N and ε > 0. Then there exists a δ > 0 such that whenever d ∈ N and
a1, . . . , an are n self-adjoint elements in Md satisfying
(1) ka2
i − aik < δ for all i = 1, . . . , n,
12
DAVID KERR AND PIOTR W. NOWAK
(2) kaiajk < δ for all distinct i, j = 1, . . . , n, and
i=1 ai = 1
(3) Pn
there exist pairwise orthogonal projections p1, . . . , pn ∈ Md such that Pn
ε for all i = 1, . . . , n.
i=1 pi = 1 and kpi −aik <
Proof. We may assume that ε ≤ 1. Using Lemma 3.7 we successively choose numbers δn >
δn−1 > · · · > δ1 > 0 such that δn = ε/2n and, for each k = 1, . . . , n − 1, whenever d ∈ N
and p and q are projections in Md satisfying kpqk < (k + 1)δk there exists a projection q′ ∈
(1 − p)Md(1 − p) such that kq′ − qk < δk+1. Let δ be a strictly positive number less than δ1/4n
to be further specified. Let a1, . . . , an be n self-adjoint elements in some matrix algebra Md such
that ka2
i=1 ai = 1.
By the functional calculus, we may take δ to be small enough as a function of ε in order to be
able to find projections q1, . . . , qn ∈ Md such that kai − qik < ε/2n for all i = 1, . . . , n. Then for
all distinct i, j = 1, . . . , n we have
i − aik < δ for all i = 1, . . . , n, kaiajk < δ for all distinct i, j = 1, . . . , n, andPn
kqiqjk ≤ k(qi − ai)qjk + kai(qj − aj)k + kaiajk
≤ kqi − aik + (kqik + kai − qik)kqj − ajk + δ < 4δ.
Using our choice of the numbers δn, . . . , δ1 as given by Lemma 3.7, we successively construct
pairwise orthogonal projections p1, . . . , pn ∈ Md such that kpi − qik < ε/2n for all i = 1, . . . , n.
i=1 pi and q = qk+1
More precisely, at the kth stage we use the fact that the projections p = Pk
satisfy
k
k
kpqk ≤
Xi=1
kpiqk+1k ≤
Xi=1(cid:0)kpi − qik + kqiqk+1k(cid:1)
≤ δ1 + · · · + δk + 4kδ < (k + 1)δk
to obtain a projection pk+1 with kpk+1 − qk+1k < δk+1. We then have kpi − aik < ε/n ≤ ε for
all i = 1, . . . , n, and
n
n
n
1 −
=(cid:13)(cid:13)(cid:13)(cid:13)
(ai − pi)(cid:13)(cid:13)(cid:13)(cid:13)
Xi=1
i=1 pi must be zero, i.e., Pn
which shows that the projection 1 −Pn
pi(cid:13)(cid:13)(cid:13)(cid:13)
Xi=1
Xi=1
(cid:13)(cid:13)(cid:13)(cid:13)
≤
kai − pik < n ·
ε
n
≤ 1,
i=1 pi = 1.
(cid:3)
Lemma 3.9. Let r ∈ {2, 3, . . . , ∞}. Let X be a zero-dimensional compact metrizable space and
Fr y X a continuous action. Suppose that C(X) ⋊λ Fr is an MF algebra. Then the action is
residually finite.
Proof. Since the MF property passes to C ∗-subalgebras and residual finiteness is a condition
that is witnessed on finitely many group elements at a time, we may assume that r is finite. We
may also assume that X is infinite, for otherwise the action is automatically residually finite.
Then by metrizability we can take an enumeration q1, q2, . . . of the projections in C(X). On X
we define the compatible metric
d(x, y) =
∞
Xi=1
2−jqj(x) − qj(y).
RESIDUALLY FINITE ACTIONS
13
Let ε > 0. Choose a n ∈ N such that 2−n+1 < ε. Take a partition of unity P = {p1, . . . , pm} ⊆
C(X) consisting of nonzero projections such that (i) each of the projections q1, . . . , qn is a sum of
projections in P, and (ii) the clopen subsets of X of which the projections in P are characteristic
functions all have diameter less than ε.
Write u1, . . . , ur for the canonical unitaries in the crossed product C(X) ⋊λ Fr corresponding
to the standard generators s1, . . . , sr of Fr. Write Ω for the set of all elements in C(X) ⋊λ Fr
of the form up or pu∗ where p ∈ P ∪ {1} and u ∈ {1, u1, . . . , ur}. Write A for the unital
commutative C ∗-subalgebra of C(X) ⋊λ Fr generated by elements of the form upu∗ where p ∈ P
and u ∈ {1, u1, . . . , ur}. Let δ > 0 be such that ((1+ δ)2 + (1+ δ)+ 1)δ < 1/4. Let δ′ be a strictly
positive number less than δ to be further specified. Since C(X) ⋊λ Fr is MF, by Proposition 3.1
there exist a d ∈ N and a unital ∗-linear map ϕ : A → Md such that
(1) (cid:12)(cid:12)kϕ(a)k − kak(cid:12)(cid:12) < δkak for all a ∈ A,
(2) kϕ(ukpiu∗
(3) kϕ(ukpiu∗
(4) kϕ(piu∗
(5) kϕ(uk)ϕ(uk)∗ − 1k < δ′ for all k = 1, . . . , r.
k) − ϕ(pi)ϕ(u∗
l ) − ϕ(ukpiu∗
kulpju∗
k) − ϕ(uk)ϕ(piu∗
k)k < δ for all i = 1, . . . , m and k = 1, . . . , r,
k)k < δ for all i = 1, . . . , m and k = 1, . . . , r,
k)ϕ(ulpju∗
l )k < δ′ for all i, j = 1, . . . , m and k, l = 1, . . . , r,
1 · · · urpir u∗
Since every minimal projection in A has the form pi0u1pi1u∗
r for some i0, . . . , ir ∈
{1, . . . , m}, by repeated use of (1) and (2) and the triangle inequality we see that by taking δ′
small enough we can arrange for kϕ(p)ϕ(p′)k and kϕ(p)2 − ϕ(p)k to be as small as we wish for all
distinct minimal projections p, p′ ∈ A. Thus assuming δ′ to be small enough we can construct
a unital homomorphism Ψ : A → Md such that kΨ(a) − ϕ(a)k < δkak for all a ∈ A by using
Lemma 3.8 to perturb the images of the minimal projections in A under ϕ to pairwise orthogonal
projections in Md and defining Ψ(p) for a minimal projection p ∈ A to be the perturbation of
ϕ(p). Take a maximal commutative subalgebra B of Md containing Ψ(A). By recoordinatizing
via an automorphism of Md, we may assume B to be the diagonal subalgebra of Md.
In view of condition (3) above, for each k = 1, . . . , r we can define the unitary vk =
ϕ(uk)/pϕ(uk)ϕ(uk)∗ in Md and we may arrange that kϕ(uk) − vkk < δ by taking δ′ to be
small enough. For i = 1, . . . , m and k = 1, . . . , r we have
kϕ(uk)ϕ(pi)ϕ(uk)∗ − vkΨ(pi)v∗
kk
≤ k(ϕ(uk) − vk)ϕ(pi)ϕ(uk)∗k + kvk(ϕ(pi) − Ψ(pi))ϕ(uk)∗k
+ kvkΨ(pi)(ϕ(uk) − vk)∗k
≤ (1 + δ)2kϕ(uk) − vkk + (1 + δ)kϕ(pi) − Ψ(pi)k + kϕ(uk) − vkk
≤ ((1 + δ)2 + (1 + δ) + 1)δ <
1
4
in which case
kΨ(ukpiu∗
k) − vkΨ(pi)v∗
kk
≤ kΨ(ukpiu∗
k) − ϕ(ukpiu∗
+ kϕ(uk)kkϕ(piu∗
k)k + kϕ(ukpiu∗
k) − ϕ(pi)ϕ(u∗
k) − ϕ(uk)ϕ(piu∗
k)k
k)k + kϕ(uk)ϕ(pi)ϕ(uk)∗ − vkΨ(pi)v∗
kk
< δ + δ + (1 + δ)δ +
1
4
< 1.
14
DAVID KERR AND PIOTR W. NOWAK
k) and vkΨ(pi)v∗
It follows that Ψ(ukpiu∗
trace. But the trace of vkΨ(pi)v∗
and k = 1, . . . , r the projections Ψ(ukpiu∗
each k = 1, . . . , r we can find a permutation matrix wk ∈ Mk such that wkΨ(pi)w∗
for all i = 1, . . . , m. Note that since wk is a permutation matrix we have wkBw∗
k are homotopic projections, and thus have the same
k is the same as the trace of Ψ(pi). Thus for all i = 1, . . . , m
k) and Ψ(pi) have the same trace. It follows that for
k = Ψ(ukpiu∗
k)
k = B.
Write P (B) for the pure state space of B, which has d elements. Let ω ∈ P (B). Since
A is a unital C ∗-subalgebra of C(X), by Gelfand theory there exists a point xω ∈ X which,
when viewed as a pure state xω on C(X), restricts to ω ◦ Ψ on A. Set ζ(ω) = xω. This
defines a map ζ : P (B) → X. Now for each i = 1, . . . , m the projection Φ(pi) is nonzero since
kΦ(pi)k ≥ kϕ(pi)k − δ ≥ kpik − 2δ > 0, and so there is an ω ∈ P (B) such that ω(Φ(pi)) = 1 and
hence pi(xω) = xω(pi) = 1. It follows that the image of ζ is ε-dense in X since the supports of
the projections p1, . . . , pm all have diameter less than ε.
Let Fr act on P (B) so that for each k = 1, . . . , r the generator sk acts as the bijection
ω 7→ ω ◦ Ad w∗
k. Then for ω ∈ P (B), k = 1, . . . , r, and i = 1, . . . , m we have
(sk xω)(pi) = xω(α−1
sk
(pi)) = xω(u∗
kpiuk)
= (ω ◦ Ψ)(u∗
= xskω(pi).
kpiuk) = (ω ◦ Ad w∗
k)(Ψ(pi))
so that (sk xω)(qi) = xskω(qi) for i = 1, . . . , n and hence
d(skζ(ω), ζ(skω)) = d(sk xω, xskω) =
2−i(sk xω)(qi) − xskω(qi) ≤
1
2n−1 < ε.
∞
Xi=1
Since ε was an arbitrary positive number we conclude that the action is residually finite.
(cid:3)
Theorem 3.10. Let r ∈ {2, 3, . . . , ∞}. Let X be a zero-dimensional compact metrizable space
and Fr y X a continuous action. Then the action is residually finite if and only if C(X) ⋊λ Fr
is an MF algebra.
Proof. Suppose that the action is residually finite. Then by Proposition 3.3 the action is qua-
sidiagonal, and since C ∗
λ(Fr) is an MF algebra [21] we infer by Theorem 3.4 that C(X) ⋊λ Fr is
an MF algebra. The other direction is Lemma 3.9.
(cid:3)
Note that for r = 1 the conclusion of Theorem 3.10 is valid without the zero-dimensionality
hypothesis by Pimsner's result from [31] (see Section 7).
4. Paradoxical decompositions
For a compact Hausdorff space X we write CX for the collection of clopen subsets of X and
BX for the collection of Borel subsets of X.
Definition 4.1. Suppose that G acts on a set X. Let S be a collection of subsets of X. Let k
and l be integers with k > l > 0. We say that a set A ⊆ X is (G, S , k, l)-paradoxical (or simply
(G, S )-paradoxical when k = 2 and l = 1) if there exist A1, . . . , An ∈ S and s1, . . . , sn ∈ G
i=1 1siAi ≤ l · 1A. The set A is said to be completely
(G, S )-nonparadoxical if it fails to be (G, S , k, l)-paradoxical for all integers k > l > 0.
such that Pn
i=1 1Ai ≥ k · 1A and Pn
RESIDUALLY FINITE ACTIONS
15
Remark 4.2. Suppose that S is actually a subalgebra of the power set PX , which will always
be the case in our applications. Then we can express the (G, S , k, l)-paradoxicality of a set
A in S by partitioning copies of A instead of merely counting multiplicities. More precisely,
A is (G, S , k, l)-paradoxical if and only if for each i = 1, . . . , k there exist an ni ∈ N and
j=1 Ai,j = A for
each i = 1, . . . , k and the sets si,jAi,j × {mi,j} ⊆ A × {1, . . . , l} for j = 1, . . . , ni and i = 1, . . . , k
are pairwise disjoint. For the nontrivial direction, observe that if A1, . . . , An and s1, . . . , sn are
as in the definition of (G, S , k, l)-paradoxicality then the sets of the form
Ai,1, . . . , Ai,ni ∈ S , si,1, . . . , si,ni ∈ G, and mi,1, . . . , mi,ni ∈ {1, . . . , l} so that Sni
A ∩(cid:18)(cid:18) \i∈P
Ai(cid:19) \ [i∈{1,...,n}\P
Ai(cid:19) ∩ s−1
j (cid:18)(cid:18) \i∈Q
siAi(cid:19) \ [i∈{1,...,n}\Q
siAi(cid:19),
where P and Q are nonempty subsets of {1, . . . , n} with P ≤ k and Q ≤ l and j ∈ P , can be
relabeled so as to produce the desired Ai,j.
Suppose that G acts on a set X. Let S be a G-invariant subalgebra of the power set PX of
X. The type semigroup S(X, G, S ) of the action with respect to S is the preordered semigroup
(cid:26)[i∈I
Ai × {i} : I is a finite subset of N and Ai ∈ S for each i ∈ I(cid:27). ∼
where ∼ is the equivalence relation under which P =Si∈I Ai×{i} is equivalent to Q =Si∈J Bi×
{i} if there exist ni, mi ∈ N, Ci ∈ S , and si ∈ G for i = 1, . . . , k such that P =Fk
and Q =Fk
Bi × {i}(cid:19)(cid:21),
(cid:20)[i∈I
Ai × {i}(cid:19) ∪(cid:18) [i∈J+max I
Bi × {i}(cid:21) =(cid:20)(cid:18)[i∈I
Ai × {i}(cid:21) +(cid:20)[i∈J
i=1 siCi × {mi}. Addition is defined by
i=1 Ci × {ni}
and for the preorder we declare that a ≤ b if b = a + c for some c.
The following is a standard observation.
Lemma 4.3. Let X be a compact metrizable space and G y X a continuous action. Let B be
a nonempty Borel subset of X. Suppose that there is a G-invariant Borel probability measure µ
on X with µ(B) > 0. Then B is completely (G, BX )-nonparadoxical.
Proof. Let µ be a G-invariant Borel probability measure on X with µ(B) > 0. Suppose that
B fails to be completely (G, BX )-nonparadoxical. Then there are k, l ∈ N with k > l for each
i = 1, . . . , k there are an ni ∈ N and Bi,1, . . . , Bi,ni ∈ BX, si,1, . . . , si,ni ∈ G, and mi,1, . . . , mi,ni ∈
j= Bi,j = B for each i = 1, . . . , k and the sets si,jBi,j ×{mi,j} ⊆ B ×{1, . . . , l}
for j = 1, . . . , ni and i = 1, . . . , k are pairwise disjoint. Since µ is G-invariant we have
and dividing by µ(B) yields k ≤ l, a contradiction. We conclude that B is completely (G, BX )-
nonparadoxical.
(cid:3)
Lemma 4.4. Let G y X be a continuous action on a zero-dimensional compact metrizable
space. Let A be a completely (G, CX )-nonparadoxical clopen subset of X such that G · A = X.
Then there exists a G-invariant Borel probability measure µ on X such that µ(A) > 0.
{1, . . . , l} so thatSni
Xj=1
kµ(B) ≤
k
Xi=1
n
k
n
µ(Bi) =
Xj=1
Xi=1
µ(si,jBi) = µ(cid:18) k
[j=1
n
[i=1
µ(si,jBi)(cid:19) ≤ lµ(B),
16
DAVID KERR AND PIOTR W. NOWAK
Proof. We claim that there is a state σ on S(X, G, CX ) such that σ([A]) > 0. Indeed suppose
that this is not the case. Since G · A = X, by compactness there is a finite set F ⊂ G such
that F · A = X and hence [X] ≤ F [A]. By the Goodearl-Handelman theorem [18, Lemma 4.1]
we can find n, m ∈ N such that m > nF and m[A] ≤ n[X], in which case m[A] ≤ nF [A],
contradicting the complete (G, CX )-nonparadoxicality of A. Thus the desired σ exists.
By Lemma 5.1 of [34] there is a G-invariant Borel measure ν on X such that ν(B) = σ([B])
for all clopen sets B ⊆ X. Since ν(X) ≥ ν(V ) > 0 and
ν(X) = ν(F · A) ≤ [s∈F
ν(sA) = F ν(A) < ∞,
we can set µ(·) = ν(·)/ν(X) to obtain a G-invariant Borel probability measure on X.
(cid:3)
Proposition 4.5. Let G y X be a minimal continuous action on a zero-dimensional compact
metrizable space. Then the following are equivalent:
(1) there is a G-invariant Borel probability measure on X,
(2) X is completely (G, CX )-nonparadoxical,
(3) there is a nonempty clopen subset of X which is completely (G, CX )-nonparadoxical.
Proof. Lemma 4.3 yields (1)⇒(2), while (2)⇒(3) is trivial. Since for any clopen set A ⊆ X we
have G · A = X by minimality, we obtain (3)⇒(1) from Lemma 4.4.
(cid:3)
Proposition 4.6. Let α be a continuous action of G on a zero-dimensional compact metrizable
space X such that C(X) ⋊λ G is stably finite. Then every nonempty clopen subset of X is
completely (α, CX )-nonparadoxical.
Proof. Write A = C(X) ⋊λ G for economy. Suppose that there is a nonempty clopen subset V
of X which fails to be completely (α, CX )-nonparadoxical. Then there are k, l ∈ N with k > l,
a clopen partition {V1, . . . , Vn} of V , and si,j ∈ G and mi,j ∈ {1, . . . , l} for i = 1, . . . , n and
j = 1, . . . , k such that the sets si,jVi × {mi,j} ⊆ V × {1, . . . , l} for i = 1, . . . , n and j = 1, . . . , k
are pairwise disjoint. For i = 1, . . . , n and j = 1, . . . , k set
ai,j = emi,j ,j ⊗ usi,j 1Vi ∈ Mk ⊗ A
where the first component in the elementary tensor is a standard matrix unit in Mk. Then aij
is a partial isometry with
Now since
k
in K0(A) we have Pk
Xj=1
j=1Pn
(ai,j)∗ai,j = ejj ⊗ 1Vi,
ai,j(ai,j)∗ = emi,j mi,j ⊗ 1si,j Vi.
(ai,j)∗ai,j =
n
Xi=1
k
Xj=1
ejj ⊗ 1V = 1 ⊗ 1V ,
i=1[(ai,j)∗ai,j] = k[1V ]. On the other hand,
k
Xj=1
n
Xi=1
ai,j(ai,j)∗ ≤(cid:18) l
Xj=1
ejj(cid:19) ⊗ 1V ,
RESIDUALLY FINITE ACTIONS
17
so that in K0(A) we have k[1V ] ≤ l[1V ] and hence −[1V ] ≥ (l − k − 1)[1V ] ≥ 0. Thus there exists
a projection p in some matrix algebra over A such that [p] + [1V ] = 0. We can thus find a d ∈ N
and pairwise orthogonal projections p′, q, r ∈ Md ⊗ A such that
(1) p′ and q are Murray-von Neumann equivalent to p and 1V , respectively, in some matrix
algebra over A (viewing 1V and p as sitting in arbitrarily large matrix algebras over A
as consistent with the definition of K0), and
(2) p′ + q + r is Murray-von Neumann equivalent to r in Md ⊗ A.
Since 1V 6= 0 the projection q is not equal to zero, and so p′ + q + r is an infinite projection.
This means that C(X) ⋊λ G fails to be stably finite, contradicting (1).
(cid:3)
Finally we discuss the relation between residual finiteness and complete nonparadoxicality in
the zero-dimensional setting.
Proposition 4.7. Let G y X be a residually finite continuous action on a zero-dimensional
compact metrizable space. Then every nonempty clopen subset of X is completely (G, CX )-
nonparadoxical.
Proof. Suppose to the contrary that there is a nonempty clopen set A ⊆ X which is (G, CX , k, l)-
nonparadoxical for some integers k > l > 0. Then there exist nonempty clopen sets A1, . . . , An ⊆
i=1 1siAi ≤ l · 1A. Let d be a compatible
metric on X. Take an ε > 0 which is smaller than the Hausdorff distance between Ai and X \ Ai
for each i = 1, . . . , n and smaller that the Hausdorff distance between siAi and X \ siAi for
each i = 1, . . . , n. By residual finiteness there is a finite set E, an action of G on E, and a map
ζ : E → X such that ζ(E) is ε-dense in X and d(ζ(siz), siζ(z)) < ε for all z ∈ E and i = 1, . . . , n.
Since ζ(E) is ε-dense in X, ζ −1(A) is nonempty. For each i = 1, . . . , n set Ei = ζ −1(Ai).
A and s1, . . . , sn ∈ G such thatPn
i=1 1Ai ≥ k · 1A andPn
Each element of ζ −1(A) is contained in ζ −1(siAi) for at most l values of i, and thus, since
ζ(siEi) ⊆ siAi for i = 1, . . . , n by our choice of ε,
n
n
n
lζ −1(A) ≥
ζ −1(siAi) ≥
Xi=1
Xi=1
siEi =
Ei.
Xi=1
that Pn
On the other hand, each element of ζ −1(A) is contained in ζ −1(Ai) for at least k values of i, so
(cid:3)
i=1 Ei ≥ kζ −1(A). Therefore l ≥ k, a contradiction.
Residual finiteness is not a necessary condition for the conclusion in Proposition 4.7, as can
be seen for G = Z as follows. Recall that the group G is said to be supramenable if for every
nonempty set A ⊆ G there is a finitely additive left-invariant measure µ : PX → [0, +∞] with
µ(A) = 1 [33] (see Chapter 10 of [40]). Every group of polynomial growth, in particular Z, is
supramenable [33]. Supramenability is easily seen to be equivalent to the property that for every
action of G on a set X and every nonempty set A ⊆ X there is a finitely additive G-invariant
measure µ : PX → [0, +∞] with µ(A) = 1.
It follows by a standard observation as in the
proof of Lemma 4.3 that, for every action of a supramenable G on a set X, every subset of X is
completely (G, PX )-nonparadoxical, where PX is the power set of X. On the other hand, for
Z-actions residual finiteness is the same as chain recurrence (see the beginning of Section 7), and
one can easily construct a continuous Z-action on the Cantor set which fails to be chain recurrent
using Lemma 2 of [31], which shows that if X a compact Haudorff space and T : X → X is a
homeomorphism then T is chain recurrent if and only if there does not exist an open set U ⊆ X
such that U \ T (U ) 6= ∅ and T (U ) ⊆ U .
18
DAVID KERR AND PIOTR W. NOWAK
5. Minimal actions of Fr
Here we give several characterizations of residual finiteness for minimal actions of the free
group Fr where r ∈ N ∪ {∞}. This is done by stringing together a variety of known results with
a few of the observations from the previous sections.
The following lemma is essentially extracted from the proof of Proposition 2.3 in [27]. Note
that in that proof it is claimed that, given a continuous action of Fr on compact metric space
X preserving a Borel probability measure, for every ε one can partition X into finitely many
measurable subsets of equal (and hence rational) measure and diameter less than ε. This is false
in general, as X could contain a clopen subset A of irrational measure and such an A would be
equal to a union of sets in the partition whenever ε is smaller than the distance between x and
y for every x ∈ A and y ∈ X \ A. The argument can be modified to take of this problem, but
we will instead give a simpler proof that was supplied to us by Hanfeng Li.
Lemma 5.1. Let X be a compact metrizable space and Fr y X a continuous action. Suppose
there exists an Fr-invariant Borel probability measure µ on X with full support. Then the action
is residually finite.
Proof. We may assume that r is finite in view of the definition of residual finiteness. Write S
for the standard generating set for Fr. Let ε > 0. Take a finite measurable partition P of X
whose elements have nonzero measure and diameter less than ε. Write Q for the collection of
sets in the join Ws∈S sP which have nonzero measure. Consider, for each P ∈ P and s ∈ S, the
homogeneous linear equation PQ∈Q,Q⊆P xQ = PQ∈Q,Q⊆sP xQ in the variables xQ for Q ∈ Q.
This system of equations has the solution xQ = µ(Q) for Q ∈ Q, and, since the rational solutions
are dense in the set of real solutions by virtue of the rationality of the coefficients, we can find a
solution consisting of rational xQ which are close enough to the corresponding quantities µ(Q)
to be all nonzero. Pick a positive integer M such that M xQ is an integer for every Q ∈ Q. For
each Q ∈ Q take a set EQ of cardinality M xQ and define E to be the disjoint union of these
sets. Take a map ζ : E → X which sends EQ into Q for each Q ∈ Q. Now for every P ∈ P
and s ∈ S the sets SQ∈Q,Q⊆P EQ and SQ∈Q,Q⊆sP EQ have the same cardinality and so we can
define an action of Fr on E by having a generator s send SQ∈Q,Q⊆P EQ to SQ∈Q,Q⊆sP EQ in
some arbitrarily chosen way for each P ∈ P. Then ζ and this action on E witness the definition
of residual finiteness with respect to ε and the generating set S.
(cid:3)
In [13] Cuntz and Pedersen introduced, for an action α of G on a C ∗-algebra A, a notion of
G-finiteness, which means that there do not exist distinct elements a, b ∈ A such that (i) a ≤ b
and (ii) a and b are G-equivalent in the sense that there exist a collection of elements ui,si ∈ A,
where i ranges in an arbitrary index set I and each si is an element of G, such that
u∗
i,siui,si
and
a =Xi∈I
αsi(ui,siu∗
i,si)
b =Xi∈I
where the sums are norm convergent. Theorem 8.1 of [13] asserts that A is G-finite if and only
if it has a separating family of G-invariant tracial states, which is equivalent to the existence of
a faithful G-invariant tracial state if A is separable. We will say that a continuous action G y
on a compact Hausdorff space X is G-finite (in the Cuntz-Pedersen sense) if the induced action
on C(X) is G-finite.
RESIDUALLY FINITE ACTIONS
19
Theorem 5.2. Let X be a compact metrizable space and Fr y X a minimal continuous action.
Then the following are equivalent:
(1) the action is residually finite,
(2) there is an Fr-invariant Borel probability measure on X,
(3) C(X) ⋊λ Fr is an MF algebra,
(4) C(X) ⋊λ Fr is stably finite,
(5) the action is Fr-finite in the Cuntz-Pedersen sense,
If moreover X is zero-dimensional then we can add the following conditions to the list:
(6) every nonempty clopen subset of X is completely (Fr, CX)-nonparadoxical.
(7) there exists a nonempty clopen subset of X which is completely (Fr, CX)-nonparadoxical.
Proof. (1)⇒(2). By Proposition 2.3.
(2)⇒(1). By minimality every G-invariant Borel probability measure on X has full support,
and so Lemma 5.1 applies.
(1)⇒(3). By Theorem 3.4.
(3)⇒(4). By Proposition 3.3.8 of [6].
(4)⇒(2). Stable finiteness implies the existence of a quasitrace (see Section V.2 of [2]) and
restricting a quasitrace on C(X) ⋊λ Fr to C(X) yields a G-invariant Borel probability measure
on X.
(2)⇔(5). Apply Theorem 8.1 of [13] as quoted above, using the fact that every Borel proba-
bility measure on X has full support by the minimality of the action.
Finally, in the case that X is zero-dimensional (2)⇔(6)⇔(7) is a special case of Proposition 4.5.
(cid:3)
The hypotheses in the above theorem actually imply conditions (1) to (5) in the case of Z,
i.e., when r = 1. When r ≥ 2 these conditions may fail, as the action on the Gromov boundary
(or any amenable minimal action) shows.
Problem 5.3. Suppose in the above theorem that the action is topologically free and X is
Cantor set. It then follows by Theorem 5.4 of [34] that if every nonempty clopen subset of X
is (Fr, CX)-paradoxical then C(X) ⋊λ Fr is purely infinite. We therefore ask whether there is
a dichotomy between the MF property and pure infiniteness within this class of actions. This
would be the case if for any such action (Fr, CX)-nonparadoxicality implies complete (Fr, CX)-
nonparadoxicality for every nonempty clopen subset of X, i.e., if the type semigroup S(X, G, CX )
is almost unperforated in the sense of Section 5 of [34].
Example 5.4. Let α be a minimal continuous action of Fr on T. Then the following are
equivalent.
(1) α is residually finite,
(2) there is a G-invariant Borel probability measure on T,
(3) C(T) ⋊λ Fn is an MF algebra,
(4) C(T) ⋊λ Fn is stably finite,
(5) αs is residually finite for every s ∈ G.
The equivalences (1)⇔(2)⇔(3)⇔(4) are by Theorem 5.2, (1)⇒(5) is trivial, and (5)⇒(2) is
a consequence of a result of Margulis [28]. Note that, by Theorem 2 of [28], if a minimal
action of a countable group on the circle has an invariant Borel probability measure then it is
20
DAVID KERR AND PIOTR W. NOWAK
conjugate to an isometric action with respect to the standard metric and thus factors through a
group containing a commutative subgroup of index at most two. Therefore there are no faithful
residually finite continuous actions of F2 on the circle.
6. Actions on spaces of probability measures
In this section we will use spaces of probability measures in order to construct an example,
for every nonamenable G, of a continuous action of G on a compact metrizable space such that
α is not residually finite but its restriction to every cyclic subgroup of G is residually finite.
The crossed product of such an action fails to be stably finite, but, unlike in the case of integer
actions, one cannot witness this failure by using the compression of an open set by a single group
element to construct a nonunitary isometry. See the discussion after Proposition 6.4.
For a compact Hausdorff space X we will view the space MX of Borel probability measures
on X with its weak∗ topology as a topological subspace of MX via the point mass identification.
Proposition 6.1. Let X be a compact metrizable space and G y X a residually finite continuous
action. Then the induced action G y MX is residually finite.
Proof. Let Ω be a finite subset of the unit ball of C(X), and equip MX with the continuous
pseudometric dΩ(σ, ω) = maxf ∈Ω σ(f ) − ω(f ). Let ε > 0. By residual finiteness there are a
finite set Y , an action of G on Y , and a map ζ : Y → X such that (i) dΩ(sζ(y), ζ(sy)) < ε
for all y ∈ Y and s ∈ F and (ii) ζ(Y ) is ε-dense in X with respect to dΩ. Let m ∈ N. Let E
be the finite subset of MY consisting of all convex combinations of the form Py∈Y cyδy where
cy ∈ {0, 1/m, 2/m, . . . , (m − 1)/m, 1} for each y ∈ Y . The the action of G on Y extends to an
action of G on E by setting
for s ∈ G. Define ζ : E → MX as the restriction of the push-forward map MY → MX associated
cyδy(cid:19) = Xy∈Y
to ζ. Then for z =Py∈Y cyδy ∈ E and s ∈ F we have
s(cid:18)Xy∈Y
dΩ(sζ(z), ζ(sz)) = sup
cyδsy
cysζ(y)(cid:19) − f(cid:18)Xy∈Y
cyζ(sy)(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)
≤ sup
cyf (sζ(y)) − f (ζ(sy)) < ε.
f ∈Ω(cid:12)(cid:12)(cid:12)(cid:12)
f(cid:18)Xy∈Y
f ∈ΩXy∈Y
Since the set of convex combinations of point masses is weak∗ dense in MX , given any η > 0 we
can take ε sufficiently small and m sufficiently large to guarantee that the image of ζ is η-dense
in MX with respect to dΩ. Since the pseudometrics of the form dΩ generate the uniformity on
MX , we conclude that the action G y MX is residually finite.
(cid:3)
Proposition 6.2. Let K be a compact convex subset of a locally convex topological vector space
and let T : K → K be a homeomorphism. Then the action of Z generated by T is residually
finite.
Proof. Let ε > 0. Let Ω be a finite set of continuous affine real-valued functions on K, and
equip K with the continuous pseudometric dΩ(x, y) = maxf ∈Ω f (x) − f (y). Let ε > 0. Take a
finite set V ⊆ K which is ε-dense for dΩ. Choose an m ∈ N such that 2m−1 maxf ∈Ω kf k < ε.
RESIDUALLY FINITE ACTIONS
21
Set E = {(v, k) : v ∈ V and k = −m, . . . , m} and define the bijection S : E → E by setting
S(v, k) = (v, k + 1) for v ∈ V and k = −m, . . . , m − 1 and S(v, m) = −m for v ∈ V .
Since Z is amenable there is a T -fixed point w ∈ K. Define a map ζ : E → K by setting, for
each v ∈ V and k = −m, −m + 1, . . . , m,
Then for v ∈ V and k = −m, . . . , m − 1 we have
ζ(v, k) =(cid:18)1 −
k
m(cid:19)T kv +
k
m
w.
dΩ(ζ(S(v, k)), T ζ(v, k)) = dΩ(ζ(v, k + 1), T ζ(v, k))
1
m(cid:12)(cid:12)f (T k+1v) − f (w)(cid:12)(cid:12)
kf k < ε,
= max
f ∈Ω
2
m
≤
max
f ∈Ω
while ζ(S(v, m)) = w = T ζ(v, m) so that dΩ(ζ(S(v, m)), T ζ(v, m)) = 0. Observe that by the
uniform continuity of T we can enlarge m to obtain the same kind of estimates for finitely many
powers of T at a time. Since the continuous pseudometrics of the form dΩ generate the topology
on K, we conclude that T is residually finite.
(cid:3)
Proposition 6.3. Let G y X be a continuous action on a compact metrizable space. Suppose
that the induced action G y MX is residually finite. Then there exists a G-invariant Borel
probability measure on X.
Proof. By Proposition 2.3 there is a G-invariant Borel probability measure µ on MX . Consider
the barycentre bµ of µ, i.e., the unique element of K satisfying bµ(f ) = µ(f ) for all f ∈ CR(MX ),
with f viewed on the right side of the equality as an affine function on MK under the canonical
identification. Then bµ is a fixed point for the action G y MX , that is, bµ is a G-invariant Borel
probability measure on X.
(cid:3)
Proposition 6.4. Suppose that G is nonamenable. Then there exists a continuous action α of
G on a compact metrizable space such that α is not residually finite but its restriction to every
cyclic subgroup of G is residually finite.
Proof. Since G is nonamenable there is a continuous action α of G on a compact metrizable
space X which does not admit an invariant Borel probability measure. By Proposition 6.3, the
induced action α of G on MX is not residually finite. However, for every cyclic subgroup H of
G the restriction of α to H is residually finite, which is obvious if H < ∞ and follows from
Proposition 6.2 otherwise.
(cid:3)
Note that if G is nonamenable and G y X is an action as in the statement of Proposition 6.4
then C(X)⋊λG fails to be stably finite, for otherwise it would admit a quasitrace (see Section V.2
of [2]) and restricting a quasitrace to C(X) would produce a G-invariant Borel probability
measure on X. Therefore Mk(C(X) ⋊λ G) contains a nonunitary isometry for some k ∈ N.
However one cannot verify the existence of such an isometry by using the compression of an
open set by a single group element as in the case of crossed products by Z [31] (see the discussion
preceding Theorem 7.2 in the next section).
Problem 6.5. For an action of the kind just described, exhibit a nonunitary isometry in some
matrix algebra over C(X) ⋊λ G.
22
DAVID KERR AND PIOTR W. NOWAK
Problem 6.6. In the statement of Proposition 6.4, can the action be taken to be minimal
and/or the space taken to be the Cantor set?
7. Z-systems
Here we specialize our discussion to Z-actions, for which the C ∗-algebraic notions of quasidi-
agonality, strong quasidiagonality, and strong NFness all admit direct dynamical interpretations
in terms of residual finiteness. Tacit use will be made of the fact that, due to the amenability of
Z, the reduced crossed product associated to a continuous action of Z on a compact Hausdorff
space is nuclear and coincides canonically with the full crossed product.
For terminological economy we will conceive of a Z-action as a pair (X, T ) where X is a
compact Hausdorff space and T : X → X is a homeomorphism associated to the generator 1,
and we will call such a pair a Z-system. We will thus speak of residually finite Z-systems. A
Z-system (X, T ) is said to be metrizable if X is metrizable. We write orb(x) for the orbit of the
point x under T , i.e., orb(x) = {T nx : n ∈ Z}.
For Z-systems, residual finiteness coincides with Conley's concept of chain recurrence [12],
defined as follows. Let (X, T ) be a Z-system. Let x, y ∈ X and let ε be a neighbourhood of
the diagonal in X × X. An ε-chain from x to y is a finite sequence {x1 = x, x2, . . . , xn = y}
in X such that n > 1 and (T xi, xi+1) ∈ ε for every i = 1, . . . , n − 1. The point x is said to be
chain recurrent if for every neighbourhood of the diagonal in X × X there is an ε-chain from x
to itself (if X is a metric space then we can quantify instead over all ε > 0, with ε-chain taking
its meaning in relation to the set {(x, y) ∈ X × X : d(x, y) < ε}). This is equivalent to x being
pseudo-nonwandering in the sense of [31]. We remark that the set of chain recurrent points is a
closed T -invariant subset of X. Finally, the system (X, T ) is said to be chain recurrent if every
point in X is chain recurrent. Note that a Z-system for which there is a dense set of recurrent
points is chain recurrent.
Proposition 7.1. A Z-system (X, T ) is residually finite if and only if it is chain recurrent.
Proof. Suppose first that (X, T ) is chain recurrent. Let F be a finite subset of Z. Let ε and ε′
be neighbourhoods of the diagonal in X × X. Take a finite ε-dense subset D of X. For each
x ∈ D pick an ε′-chain from x to itself and write Cx for the sequence obtained by omitting x at
the end of the ε′-chain. Set E = Fx∈D Cx, and let S : E → E be the bijection which cyclically
permutes the points of each Cx according to the sequential order. Then the map ζ : E → X
defined by taking the inclusion on each Cx has ε-dense image, and by uniform continuity it will
satisfy (ζ(Snz), T nζ(z)) ∈ ε for every n ∈ F if ε′ is taken fine enough. Thus (X, T ) is residually
finite.
Conversely, suppose that (X, T ) is residually finite. Let x ∈ X and let ε be a neighbourhood
of the diagonal in X × X. By residual finiteness there is a finite set E, a bijection S : E → E, a
map ζ : E → X such that (ζ(Sz), T ζ(z)) ∈ ε for all z ∈ E, and a z0 ∈ E such that (ζ(z0), x) ∈ ε.
By a simple perturbation argument involving a finer choice of ε, we may assume that ζ(z0) = x.
Then by applying ζ to the S-cycle in which z0 lies we obtain an ε-chain for x. Hence (X, T ) is
chain recurrent.
(cid:3)
In Lemma 2 of [31] (which is also valid in the nonmetrizable case) Pimsner showed that a
Z-system (X, T ) is chain recurrent if and only if there does not exist an open set U ⊆ X which
is compressed by T in the sense that U \ T (U ) 6= ∅ and T (U ) ⊆ U . From such an open set
RESIDUALLY FINITE ACTIONS
23
one can construct a nonunitary isometry in the crossed product C(X) ⋊λ Z [31, Prop. 8], which
establishes the implication (3)⇒(1) in the following theorem of Pimsner from [31].
Theorem 7.2. For a Z-system (X, T ) the following are equivalent:
(1) (X, T ) is pseudo-nonwandering (i.e., residually finite),
(2) C(X) ⋊λ Z is quasidiagonal,
(3) C(X) ⋊λ Z is stably finite.
Moreover, if X is metrizable then the following condition can be added to the list:
(4) C(X) ⋊λ Z can be embedded into an AF algebra.
Actually only the metrizable case of the above theorem was treated in [31], but since residual
finiteness passes to factors and quasidiagonality is a local property one can easily derive from
[31] the equivalence of conditions (1), (2), and (3) for general Z-actions by passing through (4)
as applied to metrizable factors. The most involved implication in the theorem is (1)⇒(4), while
(4)⇒(2) follows from the fact that AF algebras are quasidiagonal and quasidiagonality passes
to C ∗-subalgebras, and (2)⇒(3) is true for any unital C ∗-algebra (see Section V.4 of [2]). Note
that, since crossed products by actions of Z are always nuclear, for metrizable X the crossed
product C(X) ⋊λ Z is quasidiagonal if and only if it is an MF algebra [6, Thm. 5.2.2].
The property of quasidiagonality can be strengthened in certain natural ways, and in view
of the above theorem we may ask whether residual finiteness can also be used to dynamically
characterize the situation in which the crossed product satisfies such a stronger property. Recall
for example that a C ∗-algebra is said to be strongly quasidiagonal if all of its representations are
quasidiagonal. By Voiculescu's theorem [38], every simple separable quasidiagonal C ∗-algebra is
strongly quasidiagonal. Hadwin showed in [22, Thm. 25] that, for a compact metric space (X, d)
and a homeomorphism T : X → X, the crossed product C(X) ⋊λ Z is strongly quasidiagonal
if and only if for every x ∈ X there are integers m, n ≥ N such that d(T nx, T −mx) < ε. It is
easy to verify that the latter condition is equivalent to hereditary residual finiteness, by which
we mean that every subsystem of (X, T ) is residually finite. We also remark that C(X) ⋊λ Z is
residually finite dimensional (i.e., has a separating family of finite-dimensional representations)
if and only if the periodic points are dense in X [37, Thm. 4.6].
The proof of Theorem 25 in [22] can also be used to characterize when the crossed product of
a metrizable Z-system is strong NF in terms of residual finiteness. Following some preliminary
observations we will give this characterization in Theorem 7.5. Recall that a separable C ∗-
algebra A is said to be an NF algebra if it can be expressed as the inductive limit of a generalized
inductive system with contractive completely positive connecting maps, and a strong NF algebra
if the connecting maps can be chosen to be complete order embeddings [6]. Theorem 5.2.2 of
[6] gives various characterizations of NF algebras; in particular, the following are equivalent: (i)
A is an NF algebra, (2) A is a nuclear MF algebra, and (iii) A is nuclear and quasidiagonal. In
the strong NF case we have by [6, 4, 5] the equivalence of the following conditions:
(1) A is a strong NF algebra,
(2) for every Ω ∈ F A and ε > 0 there are a finite-dimensional C ∗-algebra B and a complete
order embedding ϕ : B → A such that for each x ∈ Ω there is a b ∈ B with kx−ϕ(b)k < ε,
(3) A is nuclear and inner quasidiagonal,
(4) A is nuclear and has a separating family of irreducible quasidiagonal representations.
24
DAVID KERR AND PIOTR W. NOWAK
For a λ > 1, we say that a C ∗-algebra A is an OL∞,λ space if for every finite set Ω ⊆ A and
ε > 0 there is a finite-dimensional C ∗-algebra B and an injective linear map ϕ : B → A with
Ω ⊆ B such that kϕkcbkϕ−1 : ϕ(B) → Bkcb < λ. We write OL∞(A) for the infimum over all
λ > 1 such that A is an OL∞,λ space. These notions were introduced in [23] as a means to
analyze and quantify the relationships between properties like nuclearity, quasidiagonality, inner
quasidiagonality, and stable finiteness using local operator space structure. If A is a strong NF
algebra then OL∞(A) = 1 by a straightforward perturbation argument. Whether the converse
is true is an open question, although in [16] it was shown to hold under the assumption that
A has a finite separating family of primitive ideals. As part of Theorem 7.5 we will verify that
the converse also holds in our situation, so that we obtain a dynamical characterization of both
strong NFness and the OL∞ invariant equalling 1 for crossed products of metrizable Z-systems.
Lemma 7.3. Let (X, T ) be a transitive metrizable Z-system and let d be a compatible metric
on X. Let x be a point in X such that orb(x) = X. Then the following are equivalent:
(1) (X, T ) is residually finite,
(2) {T kx}k≥0 ∩ {T kx}k<0 6= ∅,
(3) for every ε > 0 and N ∈ N there are integers m, n ≥ N such that d(T nx, T −mx) < ε,
(4) for every ε > 0 there are m, n ∈ N such that d(T nx, T −mx) < ε.
Proof. (1)⇒(2). If (2) does not hold then {T kx}k≥0 equals X \ {T kx}k<0 and hence is a clopen
set which is sent to a proper clopen subset of itself under T , so that (X, T ) fails to be residually
finite by Pimsner's characterization of chain recurrence in Lemma 2 of [31].
(2)⇒(3). Let ε > 0 and N ∈ N. Take a δ > 0 with δ ≤ ε such that if d(y, z) < δ then
d(T ky, T kz) < ε for every k = −N, −N + 1, . . . , N . Whether x is periodic or nonperiodic, it is
clear from (2) that there exist integers m, n ≥ 0 such that m + n ≥ 2N and d(T nx, T −mx) < δ.
Set r equal to N − n, m − N , or 0 according to whether n < N , m < N , or m, n ≥ N . Then
m − r, n + r ≥ N and d(T n+rx, T −m+rx) < ε, yielding (3).
(3)⇒(4). Trivial.
(4)⇒(1). Given m, n ∈ N, the sequence {x, T x, . . . , T n−1x, T −mx, T −m+1x, . . . , T −1x, x}
forms an ε-chain precisely when d(T nx, T −mx) < ε, and so x is chain recurrent. Since the
chain recurrent set is closed and Z-invariant, we obtain (1).
(cid:3)
The following lemma is a local version of Theorem 25 of [22] and essentially follows by the
same argument using Berg's technique as in the proof of Lemma 3.6 in [32]. We reproduce this
argument below. For background on induced representations see [41].
Lemma 7.4. Let (X, T ) be a transitive metrizable Z-system, and let x be a point in X such
that orb(x) = X. Then the following are equivalent:
(1) (X, T ) is residually finite,
(2) every irreducible representation of C(X) ⋊λ Z induced from the isotropy group of x is
quasidiagonal,
(3) C(X) ⋊λ Z has a faithful irreducible quasidiagonal representation.
Proof. If orb(x) has finite cardinality n then T is a cyclic permutation. As is well known,
the crossed product C(X) ⋊λ Z in this case is ∗-isomorphic to Mn ⊗ C(T), whose irreducible
representations are all evidently quasidiagonal and, up to unitary equivalence, induced from the
isotropy group of x. We may thus assume that orb(x) is infinite. Then the isotropy group of x
RESIDUALLY FINITE ACTIONS
25
is trivial and there is only one induced representation, namely the irreducible representation π :
C(X) ⋊λ Z → B(ℓ2(Z)) defined by π(f )ξn = f (T nx)ξn and π(u)ξn = ξn+1 for all f ∈ C(X) and
n ∈ Z, where u is the canonical unitary associated to T and {ξn}n∈Z is the canonical orthonormal
basis of ℓ2(Z). Since Z acts freely on orb(x) and subrepresentations of πC(X) correspond to
subsets of orb(x), it follows that πC(X) is G-almost free in the sense of Definition 1.12 of [42],
so that π is faithful by Corollary 4.19 of [42]. We thus have (2)⇒(3). Note also that condition
(3) implies that C(X) ⋊λ Z is a quasidiagonal C ∗-algebra and thus we get (3)⇒(1) in view of
Theorem 7.2. Let us then assume (1) and show that π is quasidiagonal in order to obtain (2).
Let Ω be a finite subset of C(X) and let n ∈ N. In view of the definition of a quasidiagonal
representation, we need only produce an orthogonal projection p in B(ℓ2(Z)) which dominates
the orthogonal projection onto the subspace spanned by {ξj
: −n ≤ j ≤ n} and satisfies
k[p, π(f )]k < 2/n for all f ∈ Ω and k[p, π(u)]k < 8/n. Fix a compatible metric d on X.
By uniform continuity there exists a δ > 0 such that if x1, x2 ∈ X and d(x1, x2) < δ then
f (T kx1) − f (T kx2) < 1/n for each f ∈ Ω and k = 0, . . . , n − 1. By Lemma 7.3 we can find
r > n and s < −2n such that d(T rx, T sx) < δ, so that f (T r+kx) − f (T s+kx) < 1/n for each
f ∈ Ω and k = 0, . . . , n − 1. For each f ∈ Ω we define a perturbation af ∈ B(ℓ2(W )) of π(f ) by
af ξn =(cid:26) f (T s+kx)ξn
f (T nx)ξn
if n = r + k for some k = 0, . . . , n − 1,
otherwise,
in which case we have kaf − π(f )k < 1/n.
Next we apply Berg's technique (see Section VI.4 of [15]) to produce orthogonal unit vectors
ζk, ηk ∈ span(ξr+k, ξs+k) for each k = 0, . . . , n − 1 and a unitary v ∈ B(ℓ2(Z)) such that
(1) vζk = ζk+1 and vηk = ηk+1 for k = 0, . . . , n − 2,
(2) vζn−1 = ξs+n and vηn−1 = ξr+n,
(3) v agrees with π(u) on the orthogonal complement of the span of the vectors ξr, ξr+1, . . . ,
ξr+n−1, ξs, ξs+1, . . . , ξs+n−1, and
(4) kπ(u) − vk < 4/n.
Let p be the orthogonal projection onto the span of the vectors ξr+n, ξr+n+1, . . . , ξs−1, η0, η1, . . . , ηn−1.
Since the unitary v cyclically permutes these vectors we have [p, v] = 0 and hence
k[p, π(u)]k ≤ kp(π(u) − v)k + k(v − π(u))pk ≤ 2kπ(u) − vk < 8/n.
Also, if f ∈ Ω then span(ξr+k, ξs+k) is an eigenspace for af for every k = 0, . . . , n − 1, so that
[p, af ] = 0 and hence
k[p, π(f )]k ≤ kp(π(f ) − af )k + k(af − π(f ))pk ≤ 2kaf − f k < 2/n,
completing the proof.
(cid:3)
Theorem 7.5. Let (X, T ) be a metrizable Z-system and let d be a compatible metric on X.
Then the following are equivalent:
(1) there is a collection {(Xi, T )}i∈I of transitive residually finite subsystems of (X, T ) such
(2) there is a dense set D ⊆ X such that for every x ∈ D, ε > 0, and N ∈ N there are
that Si∈I Xi is dense in X,
integers m, n ≥ N for which d(T nx, T −mx) < ε,
(3) C(X) ⋊λ Z is strong NF,
(4) OL∞(C(X) ⋊λ Z) = 1.
26
DAVID KERR AND PIOTR W. NOWAK
Proof. (1)⇔(2). This is a simple consequence of Lemma 7.3.
(2)⇒(3). Since the set W = Si∈I Xi is dense in X the canonical quotient maps C(X) ⋊λ
Z → C(Xi) ⋊λ Z for i ∈ I form a separating family. This can be seen by taking the faithful
representation π : C(X) ⋊λ Z → B(ℓ2(W ) ⊗ ℓ2(Z)) canonically induced from the multiplication
representation of C(X) on ℓ2(W ) and observing that for each i ∈ I the cut-down of π(C(X)⋊λ Z)
by the orthogonal projection onto ℓ2(Xi) ⊗ ℓ2(Z) is ∗-isomorphic to C(Xi) ⋊λ Z. Since for each
i ∈ I the crossed product C(Xi) ⋊λ Z has a faithful irreducible quasidiagonal representation by
Lemma 7.4, we thus conclude using Theorem 4.5 of [4] that C(X) ⋊λ Z is strong NF.
(3)⇒(4). This implication holds for general C ∗-algebras, as mentioned prior to the theorem
statement.
(4)⇒(1). By Theorem 5.4 of [16] C(X) ⋊λ Z has a separating family Π of irreducible represen-
tations whose images are stably finite C ∗-algebras. By [19] every primitive ideal of C(X) ⋊λ Z
is the kernel of an irreducible representation induced from the isotropy group of some point
x ∈ X. Now any two faithful irreducible representations of a separable prime (equivalently,
separable primitive) C ∗-algebra are approximately unitarily equivalent. In the antiliminal case
this follows from Voiculescu's theorem [38] as every faithful irreducible representation will be
essential, while in the non-antiliminal case the C ∗-algebra has an essential ideal ∗-isomorphic to
the compact operators and hence has only one faithful irreducible representation up to unitary
equivalence. Consequently we may assume that each π ∈ Π is induced from the isotropy group
of some xπ ∈ X.
If this point xπ has nontrivial isotropy group then it is periodic and the
system (orb(xπ), T ) is trivially residually finite.
If on the other hand xπ has trivial isotropy
group then, as indicated in the proof of Lemma 7.4, there is a unique induced representation
of C(orb(xπ)) ⋊λ Z and it is faithful, implying that C(orb(xπ)) ⋊λ Z is stably finite and hence
is dense in X, which results from the fact that π(f ) = 0 for all π ∈ Π and f ∈ C(X) whose
(cid:3)
by Theorem 7.2 that (orb(xπ), T ) is residually finite. It remains to observe that Sπ∈Π orb(xπ)
support is contained in the complement of Sπ∈Π orb(xπ).
Example 7.6. Using Theorem 7.5 we can produce by dynamical means many examples of NF
algebras which are not strong NF (cf. Examples 5.6 and 5.19 of [4]). For instance, take two copies
of translation on Z each compactified with two fixed points ±∞ and identify +∞ from each
copy with −∞ of the other copy. This system is residually finite, but the transitive subsystems
generated by each copy of Z fail to be residually finite, so that C(X) ⋊λ Z is NF but not strong
NF. This example is a dynamical analogue of Example 3.2 in [16].
Example 7.7. We can also use Theorem 7.5 to exhibit strong NF crossed products which are not
strongly quasidiagonal. For instance, take translation on Z and compactify it so that it spirals
around to the example from the previous paragraph in both the forward and backward directions,
spending longer and longer intervals near each of the two fixed points in each successive approach.
The crossed product is strong NF since the system is transitive and residually finite, but it is
not strongly quasidiagonal by the discussion in the second paragraph following Theorem 7.2. In
the backward direction we could instead have convergence to one of the fixed points, in which
case we have the additional feature that the backward and forward limit sets of the unique dense
orbit do not coincide.
RESIDUALLY FINITE ACTIONS
27
References
[1] C. Anantharaman-Delaroche and J. Renault. Amenable Groupoids. L'Enseignement Math´ematique, Geneva,
2000.
[2] B. Blackadar. Operator Algebras. Theory of C∗-Algebras and von Neumann Algebras. Encyclopaedia of Math-
ematical Sciences, 122. Operator Algebras and Non-commutative Geometry, III. Springer-Verlag, Berlin,
2006.
[3] B. Blackadar and E. Kirchberg.
Irreducible representations of
inner quasidiagonal C ∗-algebras.
arXiv:0711.4949.
[4] B. Blackadar and E. Kirchberg. Inner quasidiagonality and strong NF algebras. Pacific J. Math. 198 (2001),
307 -- 329.
[5] B. Blackadar and E. Kirchberg. Generalized inductive limits and quasidiagonality. In: C ∗-algebras (Munster,
1999), 23 -- 41, Springer, Berlin, 2000.
[6] B. Blackadar and E. Kirchberg. Generalized inductive limits of finite-dimensional C ∗-algebras. Math. Ann.
307 (1997), 343 -- 380.
[7] L. Bowen. The ergodic theory of free group actions: entropy and the f -invariant. Groups Geom. Dyn. 4
(2010), 419 -- 432.
[8] L. Bowen. Measure conjugacy invariants for actions of countable sofic groups. J. Amer. Math. Soc. 23
(2010), 217 -- 245.
[9] N. P. Brown. On quasidiagonal C ∗-algebras. In: Operator algebras and applications, 19 -- 64, Adv. Stud. Pure
Math., 38, Math. Soc. Japan, Tokyo, 2004.
[10] N. P. Brown, K. J. Dykema, and K. Jung. Free entropy dimension in amalgamated free products. Proc.
London Math. Soc. 97 (2008), 339 -- 367.
[11] N. P. Brown and N. Ozawa. C∗-Algebras and Finite-Dimensional Approximations. Graduate Studies in
Mathematics, 88. Amer. Math. Soc., Providence, RI, 2008.
[12] C. Conley. Isolated invariant sets and the Morse index. CBMS Regional Conference Series in Mathematics,
38. Amer. Math. Soc., Providence, RI, 1978.
[13] J. Cuntz and G. K. Pedersen. Equivalence and traces on C ∗-algebras. J. Funct. Anal. 33 (1979), 135 -- 164.
[14] M. Dadarlat. On the approximation of quasidiagonal C ∗-algebras. J. Funct. Anal. 167 (1999), 69 -- 78.
[15] K. R. Davidson. C ∗-algebras by Example. Fields Institute Monographs, 6. Amer. Math. Soc., Providence,
RI, 1996.
[16] C. Eckhardt. On OL∞ structure of nuclear, quasidiagonal C ∗-algebras. J. Funct. Anal. 258 (2010), 1 -- 19.
[17] G. Gong and H. Lin. Almost multiplicative morphisms and almost commuting matrices. J. Operator Theory
40 (1998), 217 -- 275.
[18] K. R. Goodearl and D. Handelman. Rank functions and K0 of regular rings. J. Pure Appl. Algebra bf 7
(1976), 195 -- 216.
[19] E. C. Gootman. Primitive ideals of C ∗-algebras associated with transformation groups. Trans. Amer. Math.
Soc. 170 (1972), 97 -- 108.
[20] E. C. Gootman and J. Rosenberg. The structure of crossed product C ∗-algebras: a proof of the generalized
Effros-Hahn conjecture. Invent. Math. 52 (1979), 283 -- 298.
[21] U. Haagerup and S. Thorbjørnsen. A new application of random matrices: Ext(C ∗
red(F2)) is not a group.
Ann. of Math. (2) 162 (2005), 711 -- 775.
[22] D. Hadwin. Strongly quasidiagonal C ∗-algebras. With an appendix by J. Rosenberg. J. Operator Theory 18
(1987), 3 -- 18.
[23] M. Junge, N. Ozawa, and Z.-J. Ruan. On OL∞ structures of nuclear C ∗-algebras. Math. Ann. 325 (2003),
449 -- 483.
[24] D. Kerr and H. Li. Soficity, amenability, and dynamical entropy. To appear in Amer. J. Math.
[25] D. Kerr and H. Li. Entropy and the variational principle for actions of sofic groups. To appear Invent. Math.
[26] H. Lin. AF-embeddings of the crossed products of AH-algebras by finitely generated abelian groups. Int.
Math. Res. Pap. IMRP 2008, Art. ID rpn007, 67 pages.
[27] A. Lubotzky and Y. Shalom. Finite representations in the unitary dual and Ramanujan groups. In: Discrete
geometric analysis, 173 -- 189, Contemp. Math., 347, Amer. Math. Soc., Providence, RI, 2004.
28
DAVID KERR AND PIOTR W. NOWAK
[28] G. Margulis. Free subgroups of the homeomorphism group of the circle. C. R. Acad. Sci. Paris S´er. I Math.
331 (2000), 669 -- 674.
[29] G. Margulis and E. B. Vinberg. Some linear groups virtually having a free quotient. J. Lie Theory 10 (2000),
171 -- 180.
[30] S. Orfanos. Quasidiagonality of crossed products. To appear in J. Operator Theory.
[31] M. V. Pimsner. Embedding some transformation group C ∗-algebras into AF-algebras. Ergod. Th. Dynam.
Sys. 3 (1983), 613 -- 626.
[32] M. Pimsner and D. Voiculescu. Exact sequences for K-groups and Ext-groups of certain cross-product
C ∗-algebras. J. Operator Theory 4 (1980), 93 -- 118.
[33] J. M. Rosenblatt. Invariant measures and growth conditions. Trans. Amer. Math. Soc. 193 (1974), 33 -- 53.
[34] M. Rørdam and A. Sierakowski. Purely infinite C ∗-algebras arising from crossed products. To appear in
Ergod. Th. Dynam. Sys.
[35] N. Salinas. Homotopy invariance of Ext(A). Duke Math. J. 44 (1977), 777 -- 794.
[36] J.-L. Sauvageot. Id´eaux primitifs de certains produits crois´es. Math. Ann. 231 (1977), 61 -- 76.
[37] J. Tomiyama. The Interplay between Topological Dynamics and Theory of C ∗-Algebras. Lecture Notes Series,
vol. 2, Res. Inst. Math., Seoul, 1992.
[38] D. Voiculescu. A noncommutative Weyl-von Neumann theorem. Rev. Roum. Math. Pures Appl. 21 (1976),
97 -- 113.
[39] D. Voiculescu. A note on quasidiagonal C ∗-algebras and homotopy. Duke Math. J. 62 (1991), 267 -- 271.
[40] S. Wagon. The Banach-Tarski Paradox. Cambridge University Press, Cambridge, 1993.
[41] D. P. Williams. Crossed Products of C ∗-Algebras. Mathematical Surveys and Monographs, 134. Amer. Math.
Soc., Providence, RI, 2007.
[42] G. Zeller-Meier. Produits crois´es d'une C ∗-alg`ebre par un groupe d'automorphismes. J. Math. Pures Appl.
(9) 49 (1968), 101 -- 239.
David Kerr, Department of Mathematics, Texas A&M University, College Station TX 77843-3368,
U.S.A.
E-mail address: [email protected]
Piotr W. Nowak, Department of Mathematics, Texas A&M University, College Station TX 77843-
3368, U.S.A.
E-mail address: [email protected]
|
1001.0424 | 1 | 1001 | 2010-01-04T01:42:44 | Families of Type {\rm III KMS} States on a Class of $C^*$-Algebras containing $O_n$ and $\mathcal{Q}_\N$ | [
"math.OA",
"math.FA"
] | We construct a family of purely infinite $C^*$-algebras, $\mathcal{Q}^\lambda$ for $\lambda\in (0,1)$ that are classified by their $K$-groups.
There is an action of the circle
$\T$ with a unique ${\rm KMS}$ state $\psi$ on each $\mathcal{Q}^\lambda.$ For $\lambda=1/n,$ $\mathcal{Q}^{1/n}\cong O_n$, with its usual $\T$ action and ${\rm KMS}$ state.
For $\lambda=p/q,$ rational in lowest terms, $\mathcal{Q}^\lambda\cong O_n$ ($n=q-p+1$) with UHF fixed point algebra of type $(pq)^\infty.$ For any $n>0,$ $\mathcal{Q}^\lambda\cong O_n$ for infinitely many $\lambda$ with distinct KMS states and UHF fixed-point algebras. For any $\lambda\in (0,1),$ $\mathcal{Q}^\lambda\neq O_\infty.$ For $\lambda$ irrational the fixed point algebras, are NOT AF and the $\mathcal{Q}^\lambda$ are usually NOT Cuntz algebras. For $\lambda$ transcendental, $K_1\cong K_0\cong\Z^\infty$, so that $\mathcal{Q}^\lambda$ is Cuntz' $\mathcal Q_{\N}$, \cite{Cu1}. If $\lambda^{\pm 1}$ are both algebraic integers, the {\bf only} $O_n$ which appear satisfy $n\equiv 3(mod 4).$ For each $\lambda$, the representation of $\mathcal{Q}^\lambda$ defined by the KMS state $\psi$ generates a type ${\rm III}_\lambda$ factor. These algebras fit into the framework of modular index (twisted cyclic) theory of \cite{CPR2,CRT} and \cite{CNNR}. | math.OA | math |
Families of Type III KMS States on a Class of C∗-Algebras containing On and QN
A.L. Careya, J. Phillipsb,c, I.F. Putnamb, A. Renniea
a Mathematical Sciences Institute, Australian National University, Canberra, ACT, AUSTRALIA
b Department of Mathematics and Statistics, University of Victoria, Victoria, BC, CANADA
c Corresponding Author. email: [email protected]; telephone: 250-721-7450; fax: 250-721-8962
Abstract
We construct a family of purely infinite C∗-algebras, Qλ for λ ∈ (0, 1) that are classified by their
K-groups. There is an action of the circle T with a unique KMS state ψ on each Qλ. For λ = 1/n,
Q1/n ∼= On, with its usual T action and KMS state. For λ = p/q, rational in lowest terms, Qλ ∼= On
(n = q − p + 1) with UHF fixed point algebra of type (pq)∞. For any n > 0, Qλ ∼= On for infinitely
many λ with distinct KMS states and UHF fixed-point algebras. For any λ ∈ (0, 1), Qλ 6= O∞. For λ
irrational the fixed point algebras, are NOT AF and the Qλ are usually NOT Cuntz algebras. For λ
transcendental, K1 ∼= K0 ∼= Z∞, so that Qλ is Cuntz' QN, [Cu1]. If λ±1 are both algebraic integers,
the only On which appear satisfy n ≡ 3(mod 4). For each λ, the representation of Qλ defined by the
KMS state ψ generates a type IIIλ factor. These algebras fit into the framework of modular index
(twisted cyclic) theory of [CPR2, CRT] and [CNNR].
Keywords: KMS state, IIIλ factor, modular index, twisted cyclic theory, K-Theory.
AMS Classification codes: 46L80, 58J22, 58J30.
1. Introduction
In this paper we introduce some new examples of KMS states on a large class of purely infinite C∗-
algebras that were motivated by the 'modular index theory' of [CPR2, CNNR]. We were aiming to
find examples of algebras that were not Cuntz-Krieger algebras (or the CAR algebra) and were not
previously known in order to explore the possibilities opened by [CNNR]. These algebras, denoted by
Qλ for 0 < λ < 1, are not constructed as graph algebras, but as "corner algebras" of certain crossed
product C∗-algebras. The Qλ have similar structural properties to the Cuntz algebras, however there
are important new features, such as
1) when λ = p/q is rational in lowest terms, then Qλ ∼= Oq−p+1 as mentioned in the Abstract,
2) when λ is algebraic, the K-groups depend on the minimal polynomial (and its coefficients) of λ,
3) when λ is transcendental, Qλ ∼= QN, Cuntz' algebra, [Cu1].
We prove in Section 3 that the Qλ are purely infinite, simple, separable, nuclear C∗-algebras, so there
is no nontrivial trace on them. Also in Section 3 we determine in many cases the K-groups of these
algebras and use classification theory to identify them when these algebras have the same K-groups
as others found previously (these facts are summarised in the Abstract). As each Qλ comes equipped
with a gauge action of the circle, our results thus give an uncountable family of distinct circle actions
on QN, each with its own unique KMS state. Indeed, for all 0 < λ < 1, we find a unique KMS state,
[BR2], for this gauge action, and we prove in Section 4 that the GNS representation of Qλ associated
to our KMS state generates a type IIIλ von Neumann algebra. The result of [CPR2] that motivated
this paper was the construction of a 'modular spectral triple' with which one may compute an index
pairing using the KMS state. In [CNNR] it was shown how modular spectral triples arise naturally for
1
2
KMS states of circle actions and lead to 'twisted residue cocycles' using a variation on the semifinite
residue cocycle of [CPRS2]. It is well known that such twisted cocycles can not pair with ordinary
K1. In [CPR2, CRT] a substitute was introduced which is called 'modular K1'. The correct definition
of modular K1 was found in [CNNR], and there is a general spectral flow formula which defines the
pairing of modular K1 with our 'twisted residue cocycle'.
There is a strong analogy with the local index formula of noncommutative geometry in the L1,∞-
summable case, however, there are important differences: the usual residue cocycle is replaced by
a twisted residue cocycle and the Dixmier trace arising in the standard situation is replaced by a
KMS-Dixmier functional. The common ground with [CPRS2] stems from the use of the spectral
flow formula of [CP2] to derive the twisted residue cocycle and this has the corollary that we have a
homotopy invariant. To illustrate the theory for these examples we compute, for particular modular
unitaries in matrix algebras over the algebras Qλ, the precise numerical values arising from the general
formalism.
2. The algebras Qλ for 0 < λ < 1.
2.1. The C∗-algebras C∗(Γλ) = C(Γλ) and their K-theory. We will construct our algebras Qλ
as "corner" algebras in certain crossed product C∗-algebras but first we need some preliminaries. For
0 < λ < 1, let Γλ be the countable additive abelian subgroup of R defined by:
Γλ =( k=NXk=−N
nkλk(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
N ≥ 0 and nk ∈ Z) .
Loosely speaking, Γλ consists of Laurent polynomials in λ and λ−1 with integer coefficients. It is not
only a dense subgroup of R, but is clearly a unital subring of R.
Proposition 2.1. Let 0 < λ < 1.
(1)
(2)
sum where d ≥ 2 is the degree of the minimal (monic) polynomial in Z[x] satisfied by λ.
(3)
(4)
internal direct sum.
(5)
If λ = p/q where 0 < p < q are integers in lowest terms, then Γλ = Z[1/n], where n = pq.
If λ and λ−1 are both algebraic integers, then Γλ = Z + Zλ + ··· + Zλd−1 is an internal direct
If λ is transcendental then, Γλ =Lk∈Z Zλk is an internal direct sum.
If λ = 1/√n with n ≥ 2 a square-free positive integer, then Γλ = Z[1/n] + Z[1/n] · √n is an
In general, if λ is algebraic with minimal polynomial, nλd + ··· + m = 0 over Z, then
Z ⊕ Zλ ⊕ ··· ⊕ Zλd−1 ⊆ Γλ ⊆ Z[
1
mn
] ⊕ Z[
1
mn
]λ ⊕ ··· ⊕ Z[
1
mn
]λd−1.
Hence, rank(Γλ) := dimQ(Γλ ⊗Z Q) = d.
Proof. In case (1), since gcd(p, q) = 1, there exist a, b ∈ Z so that 1 = ap + bq. Therefore, 1
aλ + b ∈ Γλ; and similarly, 1
have: m
q =
p ∈ Γλ. Since, Γλ is a commutative ring, for any k, m ∈ Z with k ≥ 1 we
q = ap+bq
nk = m
(pq)k is in Γλ. That is, Z[1/n] ⊆ Γλ. On the other hand, for k ≥ 1 we have
pk
qk = p2k
That is, Z[1/n] = Γλ.
nk ∈ Z[1/n] and λ−k =
(pq)k = p2k 1
nk ∈ Z[1/n].
qk
pk = q2k
1
(pq)k = q2k 1
λk =
1
In case (2), it is not hard to see the minimal polynomial of λ in Z[x] is not only monic, but also
has constant term = ±1; say, p(λ) = λd + aλd−1 + ··· ± 1 = 0. Clearly, λ ∈ Z + Zλ + ··· + Zλd−1.
Since λ−1p(λ) = 0, we also have λ−1 ∈ Z + Zλ + ··· + Zλd−1. By an easy induction, we have λk ∈
Z + Zλ + ··· + Zλd−1, for all k ∈ Z. Hence, Γλ = Z + Zλ + ··· + Zλd−1. The sum is direct by the
minimality of the degree of the minimal polynomial.
In case (3) the sum is direct because if λ satisfied a Laurent polynomial over Z, then by multipling by
a high power of λ it would also satisfy a genuine polynomial over Z.
3
Case (4) is an easy calculation which we leave to the reader. Case (5) is proved by similar methods
used in case (2). Again, the sum Z[ 1
mn ]λd−1 is direct by the minimality of the
degree of the minimal polynomial.
mn ]λ + ··· + Z[ 1
mn ] + Z[ 1
(cid:3)
Proposition 2.2. Let 0 < λ < 1.
(1)
If λ = p/q is rational in lowest terms so that Γλ = Z[1/n], where n = pq, then
K0(C(Γλ)) = Z[1Γλ)] and K1(C(Γλ)) = Z[1/n].
(2)
sum as above, then
If λ and λ−1 are both algebraic integers, so that Γλ = Z + Zλ + ··· + Zλd−1 is an internal direct
K0(C(Γλ)) =
even^ (Γλ) =
dMk=0,k even
k^(Γλ) and K1(C(Γλ)) =
odd^(Γλ) =
dMk=1,k odd
k^(Γλ).
(3)
If λ is transcendental then,
k^(Γλ) and K1(C(Γλ)) =
K0(C(Γλ)) =
If λ = 1/√n with n ≥ 2 a square-free positive integer, then
even^ (Γλ) =
∞Mk=0,k even
odd^(Γλ) =
∞Mk=1,k odd
k^(Γλ).
K0(C(Γλ)) ∼= Z ⊕ Z[1/n] and K1(C(Γλ)) ∼= Z[1/n] ⊕ Z[1/n]
In general, if λ is algebraic with nλd + ··· + m = 0 over Z then the composition of the inclusions
Z ⊕ Zλ ⊕ ··· ⊕ Zλd−1 ⊆ Γλ ⊆ Z[
1
mn
] ⊕ Z[
1
mn
]λ ⊕ ··· ⊕ Z[
1
mn
]λd−1
(4)
(5)
induces an inclusion on K-Theory, so that both of the following maps are one-to-one
odd^(Zd) ∼= K1(C∗(Z⊕··· Zλd−1)) ֒→ K1(C(Γλ)).
even^ (Zd) ∼= K0(C∗(Z⊕··· Zλd−1)) ֒→ K0(C(Γλ)) and
T. Since
Proof. In case (1), Γλ = lim−→
K0(C(T)) = Z[1] is generated by multiples of the trivial rank one bundle, the maps in the direct limit
K0(C(Γλ) = lim−→ K0(C(T)) are the identity map in each case, so that K0(C(Γλ)) = Z[1]. On the other
hand, K1(C(T)) is generated by the maps on C(T) z 7→ zk, and each map in the direct limit is the
same map induced by z 7→ zn. Thus, K1(C(Γλ)) = Z[1/n].
Cases (2) and (3) are well-known facts about the K-Theory of tori.
Z where each map is multiplication by n, so that Γλ = lim←−
Case (4): first one uses item (4) of the previous Proposition, then the proof of case (1) above in order
to apply Proposition 2.11 of [Sc]. The proof is finished off with the easily proved observation that
Z[1/n] ⊗ Z[1/n] = Z[1/n].
Case (5) the composed embedding is just containment:
mn ]λ ⊕ ··· ⊕ Z[ 1
Z ⊕ Zλ ⊕ ··· ⊕ Zλd−1 ⊆ Z[ 1
mn ]λd−1. Since we know that K∗(C∗(Z)) →
mn ] ⊕ Z[ 1
4
K∗(C∗(Z[1/mn])) is one-to-one (even an isomorphism after tensoring with Q), an application of C.
Schochet's Kunneth Theorem, [Sc] shows that the induced map on K-Theory:
K∗(C∗(Z ⊕ Zλ ⊕ ··· ⊕ Zλd−1)) −→ K∗(C∗(Z[
1
mn
] ⊕ Z[
1
mn
]λ ⊕ ··· ⊕ Z[
1
mn
]λd−1))
is one-to-one (even an isomorphism after tensoring with Q).
(cid:3)
Corollary 2.3. If λ is algebraic with minimal polynomial of degree d so that rank(Γλ) = d then
rank(K0(C(Γλ))) = rank(
even^ (Zd)) = 2d−1 = rank(
odd^(Zd)) = rank(K1(C(Γλ))).
Proof. For each N ≥ d − 1, let ΓN = Zλ−N + ··· + ZλN ⊆ Γλ. Then each ΓN is a finitely generated
torsion free (and hence free abelian) subgroup of Γλ. Moreover,
1
mn
Z ⊕ Zλ ⊕ ··· ⊕ Zλd−1 ⊆ ΓN ⊆ Γλ ⊆ Z[
]λ ⊕ ··· ⊕ Z[
] ⊕ Z[
1
mn
1
mn
]λd−1,
so that by tensoring with Q the induced inclusions are all equalities, and hence all are Q-vector spaces
of dimension d. Since ΓN is free abelian, ΓN ∼= Zd. Now,
K0(C∗(ΓN )) ∼= K0(C(Td)) ∼=
So, each Ki(C∗(ΓN )) ⊗Z Q is a Q-vector space of dimension 2d−1 and the map:
K∗(C∗(Z ⊕ Zλ ⊕ ··· ⊕ Zλd−1)) ⊗Z Q −→ K∗(C∗(ΓN )) ⊗Z Q
and K1(C∗(ΓN )) ∼= K1(C(Td)) ∼=
even^ (Zd) ∼= Z2d−1
odd^(Zd) ∼= Z2d−1
.
is one-to-one and hence an isomorphism of Q-vector spaces. Since the corresponding isomorphism
onto K∗(C∗(ΓN +1)) ⊗Z Q factors through K∗(C∗(ΓN )) ⊗Z Q the maps
K∗(C∗(ΓN )) ⊗Z Q → K∗(C∗(ΓN +1)) ⊗Z Q
are all isomorphisms. Now, C∗(Γλ) = limN C∗(ΓN ) and so Ki(C∗(Γλ)) = limN Ki(C∗(ΓN )), and
therefore,
for each i = 1, 2.
(cid:3)
Ki(C∗(Γλ)) ⊗Z Q = lim
N
Ki(C∗(ΓN )) ⊗Z Q ∼= Q2d−1
Now, let Gλ ⊃ G0
λ be the following countable discrete groups of matrices:
Of course, G0
λ is isomorphic to the additive group Γλ, and Gλ is semidirect product of Z acting on
λ ∼= Γλ. We let Gλ act on R as an "ax+b" group, noting that the action leaves Γλ invariant. That
G0
is,
Gλ =(cid:26)(cid:18) λn a
0
1 (cid:19)(cid:12)(cid:12)(cid:12)(cid:12) a ∈ Γλ, n ∈ Z(cid:27) ⊃ G0
λ =(cid:26)(cid:18) 1 a
0 1 (cid:19)(cid:12)(cid:12)(cid:12)(cid:12) a ∈ Γλ(cid:27) .
1 (cid:19) ∈ Gλ define g · t := λnt + a.
for t ∈ R and g =(cid:18) λn a
0
Notation. For such an element g ∈ Gλ we will use the notation g := [λn : a] in place of the matrix
for g and g := det(g) = λn for the determinant of g. Note: G0
We use this action on R to define the transpose action α of Gλ on L∞(R) :
αg(f )(t) = f (g−1t) for f ∈ L∞(R) and t ∈ R.
λ = {g ∈ Gλ g = 1} ⊳ Gλ.
Now let C λ
jections X[a,b) where a, b ∈ Γλ. That is,
0 (R) be the separable C∗-subalgebra of L∞(R) generated by the countable family of pro-
C λ
0 (R) = closure ( nXk=1
ckX[ak,bk)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
5
ck ∈ C; ak, bk ∈ Γλ)! .
We observe that C λ
0 (R) and since αg(X[a,b)) =
X[g(a),g(b)) both are invariant under the action α of Gλ. We define the separable C∗-algebras Aλ ⊃ Aλ
as the crossed products:
0 (R) is a commutative AF-algebra. Clearly, C0(R) ⊂ C λ
0
Aλ = Gλ ⋊α C λ
0 (R) = Z ⋊ (G0
λ ⋊α C λ
0 (R)) ⊃ Aλ
0 = G0
λ ⋊α C λ
0 (R).
c ⊂ l1
00(R) denote the dense ∗-subalgebra of C λ
λ are amenable these equal the reduced crossed products by [Ped, Theorem 7.7.7 ].
0 (R) consisting of finite linear combinations of the
0 (R)) ⊂ Aλ denote the dense ∗-subalgebra of Aλ
00(R). Similarly we define Aλ
Since Gλ and G0
Let C λ
generating projections, X[a,b), and let Aλ
α(Gλ, C λ
consisting of finitely supported functions x : Gλ → C λ
Proposition 2.4. For any λ ∈ (0, 1) Aλ
Proof. Since Aλ = Z ⋊ Aλ
0 is in Nnuc. By the proof of the previous Corollary,
we can write Γλ as an increasing union of finitely generated torsion-free abelian groups ΓN which are
free abelian group of finite rank so that Aλ
0 is the direct limit of crossed products of the separable
commutative C∗-algebra C λ
(cid:3)
0 and Aλ are in the bootstrap class Nnuc.
0, it suffices to see that Aλ
0 (R) by Zmi and hence is in Nnuc.
0,c ⊂ Aλ
0 .
Notation: We remind the reader of the crossed product operations in our setting (Definition 7.6.1 of
[Ped]) together with some particular notations we use. To this end, let x, y ∈ l1
0 (R)) then we
have the product and adjoint formulas:
α(Gλ, C λ
(x · y)(g) = Xh∈Gλ
x∗(g) = αg((x(g−1))∗) for g ∈ Gλ.
x(h)αh(y(h−1g)) for g ∈ Gλ;
c , then:
α(Gλ, C λ
c (respectively, Aλ
0,c) is dense in Aλ (respectively, Aλ
0 (R), then we write
0 ) we often do our calculations
0 (R)) is supported on the single element g ∈ Gλ and x(g) = f ∈ C λ
If x ∈ l1
x = f · δg. Since Aλ
with these elements and we have the following easily verified calculus for them.
Lemma 2.5. Let f1 · δg1, f2 · δg2, f · δg ∈ Aλ
(1)
(2)
(3)
(4)
(5)
(6) The product of partial isometries of the form X[a,b) · δg is a partial isometry of the same form.
(7) Consider the partial isometry, v = X[a,b) · δg. Any two of the following: vv∗, v∗v, g completely
determine the interval [a, b) and the element g.
Definition 2.6. Let e ∈ Aλ
(f1 · δg1) · (f2 · δg2) = f1αg1(f2) · δg1g2
(f · δg)∗ = αg−1( ¯f ) · δg−1.
f · δg is self-adjoint if and only if f is self-adjoint and g = 1.
f · δg is a projection if and only if f is a projection and g = 1.
f · δg is a partial isometry if and only if f is a projection.
0,c be the projection e = X[0,1)·δ1. We define the separable unital C∗-algebras:
Qλ := eAλe ⊃ eAλ
0 e =: F λ.
c e, and F λ
We will also have occasion to use the dense subalgebras Qλ
Proposition 2.7. The orthogonal family of projections en = X[n,n+1) · δ1 ∈ Aλ
equivalent by partial isometries in Aλ
c := eAλ
0 for n ∈ Z are mutually
0 of the form Vn,k := X[n,n+1) · δgn−k where gn−k = [1 : (n − k)].
c := eAλ
0,ce.
6
Moreover, the finite sums EN :=PN−1
that
n=−N en = X[−N,N ) · δ1 form an approximate identity for Aλ so
Aλ ∼= Qλ ⊗ K(l2(Z)) and Aλ
0 ∼= F λ ⊗ K(l2(Z)).
Proof. By Lemma 2.5, one easily calculates that:
for each pair n, k ∈ Z, Vn,kV ∗n,k = en and V ∗n,kVn,k = ek.
Now for each positive integer N if we have y ∈ Aλ
c that satisfies supp(yh) ⊆ [−N, N ) for all h, then
using Lemma 2.5 again we see that EN · y = y. Since the collection of all such elements y ∈ Aλ
c is
dense in Aλ, we see that the increasing sequence of projections {EN} form an approximate identity
for Aλ.
Corollary 2.8. It follows from Proposition 2.4.7 of [RS] and Proposition 2.4 that for any λ ∈ (0, 1),
Qλ and F λ are both in Nnuc.
Lemma 2.9. (cf. [PhR, Proposition 3.1, Lemma 3.6]) The algebra C λ
0 (R) is a commutative separable
AF algebra consisting of all functions f : R → C which vanish at ∞ and: are right continuous at
each x ∈ Γλ; have a finite left-hand limit at each x ∈ Γλ; and are continuous at each x ∈ (R \ Γλ).
Moreover, if φ ∈ \C λ
0 (R) → C) then there exists a
unique x0 ∈ R such that:
(1) if x0 ∈ (R \ Γλ) then φ(f ) = f (x0) for all f ∈ C λ
(2) if x0 ∈ Γλ then either(cid:26) φ(f ) = f (x0) for all f ∈ C λ
φ(f ) = f−(x0) = limx→x−
0 (R), (the space of all nonzero ∗-homomorphisms: C λ
0 (R), or
f (x) for all f ∈ C λ
0 (R).
0 (R),
(cid:3)
0
Proof. Since generating functions for C λ
0 (R) satisfy each of the properties above which are clearly
preserved by passing to uniform limits, we see that any function in C λ
0 (R) satisfies these properties.
Conversely, it is easy to show that any function satisfying these properties can be uniformly approxi-
mated by a finite linear combination of the generators. The remainder of the proof is given in [PhR,
Lemma 3.6].
(cid:3)
Notation. We denote the dual space, \C λ
that is the topology of pointwise convergence on C λ
space, and C λ
0 (R) ∼= C0(Rλ).
0 (R) by Rλ and endow it with the relative weak-∗ topology,
0 (R). Of course, Rλ is a locally compact Hausdorff
Proposition 2.10. The algebras Aλ and Aλ
over, Aλ is purely infinite and hence so is Qλ.
Proof. Now, both Gλ and G0
0 (R) as countable discrete groups of outer automorphisms.
Thus, we can apply Theorem 3.2 of [E] once we check that neither action has any nontrivial invariant
ideals in C λ
0 (and hence Qλ and F λ) are simple C∗-algebras. More-
0 (R) and that the actions are properly outer in the sense of Definition 2.1 of [E].
λ act on C λ
g
To do this we look at the induced action of Gλ and G0
λ on Rλ. So, for g ∈ Gλ we have g acting on Rλ
via g(φ) = φ◦ α−1
so that for φ = φx given by evaluation at x ∈ R, we have as expected g(φx) = φg(x).
Now, if x ∈ Γλ we use the notation φx− to denote the ∗-homomorphism φx−(f ) = f−(x) = f (x−) =
limy→x− f (y). One easily checks that since g(x) ∈ Γλ, we have g(φx−) = φg(x)−.
Next we claim that each of the sets {φm m ∈ Γλ} and {φm− m ∈ Γλ} is dense in Rλ in the relative
weak-∗ topology. For example, we show that the second set is dense. To approximate φx for some
x ∈ R we let {mn} be a sequence in Γλ converging to x from the right in R. Let f ∈ C λ
0 (R) so that f
is right continuous at x. One easily shows that φmn− (f ) − φx(f ) → 0; that is, the sequence {φmn−}
converges to φx in the relative weak-∗ topology.
7
λ and Gλ on C λ
It is easy to see that the action of G0
λ on Rλ has dense orbits, and so, of course, the action of Gλ has
dense orbits also. This implies that the actions of G0
0 (R) have no nontrivial invariant
ideals since the induced action on Rλ has no nontrivial invariant closed sets. We complete the proof
by showing that the action is properly outer in the sense of Definition 2.1 of [E]. Since there are
no nontrivial α-invariant ideals and C λ
0 (R) is commutative this is the condition that for each g 6= 1
and each nonzero closed two sided ideal I invariant under αg we have k(αg − Id)Ik = 2. Since I is
nonzero there is a nonempty open subset, O of Rλ so that I = O. But since g 6= 1 and O is not finite
there exists y ∈ O such that g(y) 6= y and g(y) ∈ O. Let x = g(y) ∈ O so that g−1(x) = y ∈ O and
x 6= g−1(x). So we can choose a continuous compactly supported real-valued function f on O with
f (x) = 1, f (g−1(x)) = −1 and kfk = 1. But then f ∈ I and
2 ≥ k(αg − Id)Ik ≥ k(αg − Id)(f )k = kαg(f ) − fk ≥ f (g−1(x)) − f (x) = 2.
Now that we know Aλ is simple, we can easily apply Theorem 9 of [LS] to conclude that Aλ satisfies
hypothesis (v) of Proposition 4.1.1 (page 66) of [RS]. For simple C∗-algebras, this is equivalent to being
purely infinite by Definition 4.1.2 of [RS]: the authors of [LS] had used one of the earlier definitions of
purely infinite in their paper (namely, hypothesis (v)). By Proposition 4.1.8 of [RS] Qλ is also purely
infinite.
(cid:3)
Corollary 2.11. It follows from Corollaries 8.2.2 and 8.4.1 (Kirchberg-Phillips) of [RS] and the fact
that Aλ is stable that for any λ ∈ (0, 1), Aλ is classified up to isomorphism (among Kirchberg algebras
in Nnuc) by its K-theory.
Since we need to calculate with elements of Qλ and F λ, we make the following observations.
Lemma 2.12. Now, Qλ (respectively, F λ) is the norm closure of finite linear combinations of the ele-
ments of the form e(X[a,b) · δg)e, where g ∈ Gλ (respectively, g ∈ G0
λ), henceforth called the generators.
Thus, we calculate
(1) If f · δg ∈ Aλ, (respectively, f · δg ∈ Aλ
0 (R), then
0 ) where f ∈ C λ
e(f · δg)e = X[a,b)f · δg where [a, b) = [0, 1) ∩ [g(0), g(1)).
(2) Thus, for g ∈ Gλ, (respectively, g ∈ G0
[g(0), g(1)). In particular, for g ∈ Gλ, (respectively, g ∈ G0
[a, b) ⊆ [0, 1) ∩ [g(0), g(1)).
λ) f · δg is in Qλ (respectively, F λ) iff supp(f ) ⊆ [0, 1) ∩
λ) X[a,b) · δg is in Qλ (respectively, F λ) iff
Proof. The first item is an easy calculation using part (1) of Lemma 2.5 and the fact that αg(X[a,b)) =
X[g(a),g(b)). The second item follows easily from the first.
Proposition 2.13. If λ is rational, then Aλ
0 and F λ are AF-algebras. In particular, if λ = p/q where
0 < p < q are in lowest terms, then F λ is the UHF algebra n∞ where n = pq. Moreover, the minimal
projections in the finite-dimensional subalgebras can all be chosen from the canonical commutative
subalgebra C λ
(cid:3)
0 (R) · δI .
Proof. We have shown in Proposition 2.1 that if λ = p/q where 0 < p < q are in lowest terms, then
Γλ = Z[1/n], where n = pq. Now, any element in Z[1/n] has the form m/nk = m(1/nk) where k ≥ 1.
Therefore any of the generating partial isometries X[a,b) · δ[1:c] ∈ Aλ
0 can (by bringing a, b and c to a
common denominator) be written (assuming c > 0) as a finite linear combination of partial isometries
of the form X[l/nk,(l+1)/nk)·δ[1:1/nk]. For partial isometries in F λ we would have to restrict 0 ≤ l ≤ nk−1
and such partial isometries generate an nk by nk matrix subalgebra of F λ. It should now be clear that
F λ is a UHF algebra of type n∞.
(cid:3)
8
At this point we define some special elements in Qλ which behave very much like the isometries
Sµ ∈ On, except for the fact that some of them are not isometries.
Definition 2.14. Fix 0 < λ < 1 and let k be a positive integer. Define mk to be the unique positive
integer satisfying: mkλk < 1 ≤ (mk + 1)λk. For 0 ≤ m ≤ mk define partial isometries Sk,m ∈ Qλ
via:
Sk,m = X[mλk,(m+1)λk ) · δgk,m where gk,m = [λk : mλk].
Note: for m < mk the Sk,m are actually isometries, and Sk,mk is an isometry iff 1 = (mk + 1)λk.
Remarks. The defining inequalities mkλk < 1 ≤ (mk + 1)λk for the positive integer mk are equivalent
to: 0 < λ−k − mk ≤ 1. In particular, these differences are positive and bounded above by 1. In the
case of Q1/n we have mk = nk − 1. Generally we have mk
Lemma 2.15. With the previously defined elements we have:
1 ≤ mk < 1 ≤ (mk + 1) ≤ (m1 + 1)k.
S∗k,m = X[0,1) · δg−1
k,m
and S∗k,mk
= X[0,λ−k−mk) · δg−1
k,mk
where for all m, g−1
k,m = [λ−k : −m].
Moreover, for 0 ≤ m < mk, S∗k,mSk,m = X[0,1) · δ1 = e while S∗k,mk
Finally, for 0 ≤ m < mk, Sk,mS∗k,m = X[mλk,(m+1)λk) · δ1 while Sk,mk S∗k,mk
Sk,mk = X[0,λ−k−mk) · δ1.
= X[mkλk,1) · δ1, so that
mkXm=0
Sk,mS∗k,m = X[0,1) · δ1 = e.
Proof. These are just straightforward calculations based on Lemma 2.5 which we leave to the reader.
(cid:3)
Theorem 2.16. For each λ with 0 < λ < 1, consider the partial isometries S1,m for m = 0, 1, ..., m1
m=0 S1,mS∗1,m = 1. For λ = 1/n,
where m1λ < 1 ≤ (m1 + 1)λ. For m < m1, S1,m is an isometry andPm1
m1 = n − 1, S1,m1 is also an isometry, and Q1/n ∼= On, the usual Cuntz algebra.
Proof. The first statement is clear. With λ = 1/n we have inside Q1/n, n isometries one for each
m = 0, 1, ..., (n − 1) defined by:
Sm = X[ m
Using Lemma 2.12, we easily see that for each m, Sm ∈ Q1/n.
Then, using item (1) of Lemma 2.5 we calculate:
n ) · δgm where gm = [1/n : m/n] and so S∗m = X[0,1) · δg−1
m = [n : −m].
n , m+1
where g−1
m
S∗mSm = X[0,1) · δ1 = e and SmS∗m = X[ m
n , m+1
n ) · δ1 and so
n−1Xm=0
SmS∗m = X[0,1) · δ1 = e.
Since e is the identity of Q1/n, we have constructed a unital copy of On inside Q1/n. Now one shows by
induction that for each k > 0 the product of exactly k of these n isometries has the form Sk,m where
Sk,m has the same defining equation as Sm above but with nk in place of n and m = 0, 1, ..., (nk − 1).
These new isometries have range projections Sk,mS∗k,m = X[ m
nk ) · δ1 which therefore lie in this copy
of On. By adding up some of these projections, we can get any projection of the form X[a,b) · δ1 where
0 ≤ a < b ≤ 1 and both a, b have the form m/nk. But any element a ∈ Γ1/n can be written as a = m
for a sufficiently large k ≥ 0 and some m ∈ Z depending on k, and any pair a, b can be brought to a
common denominator nk. Hence any projection of the form X[a,b) · δ1 in Q1/n is in this copy of On.
nk , m+1
nk
Now, a straightforward calculation gives us:
9
(1)
nk−1Xm=1
Sk,mS∗k,m−1 =
nk−1Xm=1
X[ m+1
nk , m+2
nk ) · δ[1 : 1/nk] = X[1/nk,1) · δ[1 : 1/nk] ∈ On.
Finally, let X[a,b) · δg ∈ Q1/n be an arbitrary generator. By taking adjoints if necessary we can assume
that g has the form g = [nk : ∗] where k ≥ 0. Since Sk,0 is an isometry in On it suffices to prove that
Sk,0(X[a,b) · δg) ∈ On. That is, we are reduced to the case g = [1 : c] and again by taking adjoints
if necessary we can assume that c ≥ 0. The case c = 0 is done and so we can assume that c > 0. So
(with possibly new a, b) we have X[a,b) · δ[1 : c] where 0 < c ≤ 1 and [a, b) ⊆ [0, 1) ∩ [c, c + 1) = [c, 1).
But, X[a,b)· δ[1 : c] = X[a,b)X[c,1)· δ[1 : c] = X[a,b)· δ1X[c,1)· δ[1 : c] and we already know that X[a,b) · δ1 ∈ On.
Therefore it suffices to see that X[c,1) · δ[1 : c] ∈ On. However, c = l/nk for some 0 < l < nk and so:
X[c,1) · δ[1 : c] = X[l/nk,1) · δ[1 : l/nk] =(cid:16)X[1/nk,1) · δ[1 : 1/nk](cid:17)l
which is in On by Equation 1. Since all generators for Q1/n are in On we're done.
(cid:3)
2.2. K-Theory of Qλ for λ rational. : Since Aλ
0 is stable and stably isomorphic to the UHF
algebra Fλ, each of its projections is equivalent to one in some finite-dimensional subalgebra and
0 (R), and in this case the trace induces an isomorphism from K0(Aλ
hence to some projection in C λ
0 )
onto Γλ = Z[1/(pq)] ⊂ R. This isomorphism carries the projection e = X[0,1) · δ1 which is the identity
of Qλ and F λ onto 1 ∈ Z[1/(pq)]. Now, since Aλ
0 we
can use the Pimsner-Voiculescu exact sequence to calculate K∗(Aλ) = K∗(Qλ). When we do this we
get:
0 ) = {0}, and since Aλ = Z ⋊λ Aλ
0 is AF, K1(Aλ
K1(Qλ) = {0}, and K0(Qλ) = Z[1/(pq)]/(1 − λ)Z[1/(pq)].
Proposition 2.17. For λ rational with λ = p/q in lowest terms, we have
K1(Qλ) = {0}, and K0(Qλ) ∼= Z[1/(pq)]/(1 − λ)Z[1/(pq)] ∼= Z(q−p).
Proof. By Proposition 2.1, Γλ = Z[1/(pq)], so we must show that
Z[1/(pq)]/(1 − (1/(pq))Z[1/(pq)] ∼= Z(q−p).
Since (q − p) = (1− p/q)q and every element of Z[1/(pq)] is of the form m/(pq)N , it is easy to see that
(q − p)Z[1/(pq)] = (1 − p/q)Z[1/(pq)]. Now, (q − p) and (pq)N are relatively prime for any N and so
there exist a, b ∈ Z so that 1 = a(q − p) + b(pq)N and hence m/(pq)N = (q − p)am/(pq)N + mb. That
is, m/(pq)N and mb represent the same element in the quotient. So, every element in the quotient has
an integer representative. Two integers c, d represent the same element in the quotient if and only if
c − d = (p − q)n/(pq)N , or (c − d)(pq)N = n(q − p). But then:
(c − d) = (c − d)[a(q − p) + b(pq)N ] = (c − d)a(q − p) + b(c − d)(pq)N = [(c − d)a + bn](q − p).
That is, c, d represent the same element in Z/(q − p)Z = Z(q−p). On the other hand if (c − d) is in
(q − p)Z then clearly, [c] = [d] in Z[1/(pq)]/(1 − (1/(pq))Z[1/(pq)] and we are done.
Corollary 2.18. If λ = p/q in lowest terms, then
(cid:3)
In particular, if λ = k
F λ = F p/q ∼= U HF ((pq)∞) and Qλ = Qp/q ∼= O(q−p+1).
k+1 then
F λ ∼= U HF ((k(k + 1))∞) and Qλ ∼= O2.
10
Proof. Since each Qλ is separable, nuclear, simple, purely infinite and in the bootstrap category Nnuc
once we show that the class of the identity e ∈ Qλ is a generator for K0(Qλ) = Z/(q − p)Z, the
Kirchberg-Phillips Classification Theorem, Theorem 8.4.1 of [RS], shows that Qλ ∼= O(q−p+1). To this
end we observe that since e is mapped to 1 in Z[1/pq], we must show that [1] is a generator for
K0(Qλ) = Z[1/pq]/(1 − (p/q))Z[1/pq]. Now, by the proof of the previous proposition, k[1] = [k · 1] =
0 ∈ Z[1/pq]/(1 − (p/q))Z[1/pq] if and only if [k · 1] = 0 ∈ Z/(q − p)Z if and only if k − 0 = m(q − p)
for some m ∈ Z if and only if k is a multiple of (q − p). That is, [1], [2 · 1], . . . , [(q − p − 1) · 1] are all
nonzero in K0(Qλ) = Z/(q − p)Z and hence [1] is a generator.
2.3. The K-Theory of the Algebras Aλ
0 for λ irrational. The case λ rational is much simpler,
and while it does fit into the following scheme, it does not need this deeper machinery. Initially, we
(and others) believed that the algebras Aλ
In fact we will
show that Aλ
0 is never AF when λ is irrational. We will set up our examples to fit the situation on
page 1487 of [Put2] so that we can apply the six-term exact sequence of Theorem 2.1 on page 1489 of
[Put2].
We let Γ = Γλ ∼= G0
λ. Thus, Γ ⊂ R is a countable dense subgroup of R which acts on R by translations.
Before looking at the crossed product of Γ acting on C λ
0 ) we first
consider the crossed product of Γ acting on C0(R). Since Γ acts on R by translation we can Fourier
transform to get an isomorphism:
0 were AF algebras when λ is irrational.
0 (R) = C0(Rλ) (which gives us Aλ
(cid:3)
Then, by Connes' Thom isomorphism we get for i = 0, 1:
Γ ⋊ C0(R) ∼= R ⋊ C(Γ).
Proposition 2.19. The composition:
Ki(Γ ⋊ C0(R)) ∼= Ki( R ⋊ C(Γ)) ∼= Ki+1(C(Γ)).
K1(C0(R))
i∗−→ K1(Γ ⋊ C0(R))
b
−→ K1( R ⋊ C(Γ)) ∼=−→ K0(C(Γ))
takes the generator [u] ∈ K1(C0(R)) = Z · [u]; where u is the Bott element in C0(R)1 defined by
u(t) = 1+it
1−it ; to [1Γ] where 1Γ is the identity function in C(Γ).
Proof. We first work on the right hand side of this sequence of maps. Let u(t) = 1 + ε(t), then by the
proof of Connes' Thom isomorphism from
K0(C(Γ)) ⊗Z K1(C0(R)) −→ K1(R ⋊ C(Γ))
we see that [1Γ] ⊗ [u] gets mapped to the class [1 + (convolution by ε · 1Γ)]. Now in this displayed
equation, K0(C(Γ)) ⊗Z K1(C0(R)) = K0(C(Γ)) ⊗Z Z · [u] = K0(C(Γ)) · [u] ∼= K0(C(Γ)). Thus, [1Γ] in
K0(C(Γ)) gets mapped to the class [1 + (convolution by ε · 1Γ)] by the Thom isomorphism.
On the other hand, the map K1((C0(R)1) −→ K1((Γ ⋊ C0(R))1) takes [u] 7−→ [δ0 · ε + 1] and by the
Fourier transform this goes to [(convolution by ε · 1Γ) + 1] in K1(R ⋊ C(Γ)). Combining these we get:
1 ∈ Z 7−→ [u] ∈ Z · [u] = K1((C0(R))1) = K1(C0(R)) 7−→ [1Γ] ∈ K0((C(Γ)).
(cid:3)
Now, by Proposition 2.1 we know Γ in many cases so that these last groups are quite computable. In
the notation of [Put2] we define the transformation groupoids:
G := Rλ ⋊ Γ, G′ := R ⋊ Γ, and H := Γ ⋊ Γ.
0 = C∗r (G) is the reduced C∗-algebra of G; Γ ⋊ C0(R) = C∗r (G′) is the reduced C∗-algebra of
Then, Aλ
G′; and K(l2(Γ)) is the reduced C∗-algebra of H. By the proof of Proposition 2.10 there is a continuous
proper surjective map: Rλ → R, where points in R which are not in Γ each have a single pre-image,
while points γ ∈ Γ have exactly two pre-images in Rλ, which we denote by γ− and γ+. Thus, there
are two disjoint embeddings of Γ in Rλ :
11
i0, i1 : Γ → Rλ :
i0(γ) = γ−,
i1(γ) = γ+.
Now in order to mesh with the notation of [Put2], we let Y := Γ with the equivalence relation,
"="; X := Rλ, with the equivalence relation (i0(γ) ∼ i1(γ)); and quotient π : X → X′ := R where
X′ = X/(i0(γ) ∼ i1(γ)) = R; while the "factor groupoid" of G = Rλ×Γ = X×Γ is G′ := R×Γ = X′×Γ.
We represent each of these three C∗-algebras on H := l2(Γ+) ⊕ l2(Γ−) where Γ± = {γ± γ ∈ Γ} in
the following way. First we denote the natural orthonormal basis elements of H by δa+ and δa− for
each a ∈ Γ. Now the unitary representation U of Γ on H is Uγ(δa±) = δ(a−γ)± . The actions of C0(Rλ),
C0(R), and C0(Γ) on H are as follows for f1 ∈ C0(Rλ), f2 ∈ C0(R), f3 ∈ C0(Γ), and δa± ∈ H
π1(f1)(δa± ) = f1(a±)δa± π2(f2)(δa± ) = f2(a)δa± π3(f3)(δa±) = f3(a)δa± .
These three covariant pairs of representations, (π1, U ), (π2, U ), and (π3, U ) define representations of
0 , C∗r (G′) = Γ ⋊ C0(R), and C∗r (H) = K(l2(Γ)) respectively on H. Since each of these
C∗r (G) = Aλ
C∗-algebras is simple these representations are faithful.
Now, one checks that the hypotheses of Theorem 2.1 of [Put2] are satisfied. As in [Put1, Put2] one
shows that the two mapping cone algebras of the inclusions:
C∗r (G′) = Γ ⋊ C0(R) −→ Aλ
0 = C∗r (G) and C∗r (H) −→ C∗r (H) ⊕ C∗r (H) :
( x 7→ (x, x) )
have isomorphic K-Theory. One then pastes these isomorphisms into the mapping cone long exact
0 = C∗r (G). Next one observes that for any C∗-algebra, B
sequence for C∗r (G′) = Γ ⋊ C0(R) −→ Aλ
the diagonal embedding B −→ B ⊕ B induces the diagonal embedding K∗(B) −→ K∗(B) ⊕ K∗(B)
with quotient isomorphic to K∗(B) (this is true for any abelian group). This implies that K∗(B) ∼=
K∗+1(M (B, B ⊕ B)) so that we get the six-term exact sequence from [Put2]:
K1(C∗r (H))
/ K0(C∗r (G′))
/ K0(C∗r (G))
In our set-up this becomes:
K1(C∗r (G))
K1(C∗r (G′))
K0(C∗r (H))
{0}
/ K0(Γ ⋊ C0(R))
/ K0(Γ ⋊ C0(Rλ))
K1(Γ ⋊ C0(Rλ))
K1(Γ ⋊ C0(R))
Z
Which by Connes' Thom isomorphism becomes:
{0}
/ K1(C(Γ))
/ K0(Aλ
0 )
K1(Aλ
0 )
K0(C(Γ))
Z
By Proposition 2.19, the nonzero element [1Γ] in K0(C0(Γ)) ∼= K1(Γ ⋊ C0(R)) is mapped to the
image of the class [u] in K1(Γ ⋊ C0(R)) by Connes' Thom isomorphism, and then the image of [1Γ] in
K1(Γ⋊C0(Rλ)) is the same as the image of [u] under the inclusion K1(Γ⋊C0(R)) −→ K1(Γ⋊C0(Rλ)).
However, this is clearly the same as the image of [u] under the inclusion K1(C0(R)) → K1(C0(Rλ)) →
K1(Γ ⋊ C0(Rλ)). This composition is 0 since C0(Rλ) is an AF-algebra. That is, the element [1Γ] in
/
/
O
O
o
o
o
o
/
/
O
O
o
o
o
o
/
/
O
O
o
o
o
o
12
0 ) and hence is in the image of the map Z −→ K0(C0(Γ)). Since
K0(C0(Γ)) is mapped to 0 in K1(Aλ
[1Γ] generates a copy of Z in K0(C0(Γ)), we have a nonzero homomorphism from Z to Z[1Γ] which is
onto and hence one-to-one. By the exactness, the map K0(Aλ
CONCLUSION : K0(Aλ
0 ) ∼= K1(C(Γλ)) and K1(Aλ
0 ) −→ Z is the zero map.
0 ) ∼= K0(C(Γλ))/[1Γλ
]Z.
Proposition 2.20. If λ is irrational, then K1(Aλ
0 ) 6= {0} so that Aλ
0 is not an AF-algebra.
Proof. By items (3) and (5) of Proposition 2.2 we see that when λ is irrational, K0(C(Γλ)) is not
singly generated so that K1(Aλ
(cid:3)
2.4. K-theory computations of particular Qλ for λ irrational. Example(s) λ = 1/√n:
0 ) ∼= K0(C(Γλ))/[1Γλ
]Z 6= {0}.
n > 1 a square-free integer. Using Proposition 2.1, we get:
for
K1(F λ) = K1(Aλ
K0(F λ) = K0(Aλ
0 ) = K1(C( Γλ)) = Z[1/n] ⊕ Z[1/n]
0 ) = (K0(C( Γλ)))(cid:14)Z[1] = (Z[1] ⊕ Z[1/n])(cid:14)Z[1] = Z[1/n].
To compute the K-theory of Qλ in this case using the Pimsner-Voiculescu exact sequence, one must
first compute the induced automorphism λ∗ on K1(C( Γλ)) and on K0(C( Γλ)) by a more detailed
analysis of the proof of [Sc, Proposition 2.11]. In the case of K1(C( Γλ)) we get a copy of the group
Γλ = Z[1/n] + Z[1/n]√n and the action on Γλ is just multiplication by λ = 1/√n. As an action
translated to the abstract group Z[1/n] ⊕ Z[1/n], the automorphism becomes λ∗(a, b) = (b, a/n).
Therefore, id∗ − λ∗ on K0(Aλ
0 ) = Z[1/n] ⊕ Z[1/n] to itself is clearly 1 : 1. Now it is an instructive
exercise to show that the kernel of the homomorphism
is exactly the range of the homomorphism
(a, b) ∈ Z[1/n] ⊕ Z[1/n] 7→ [a + b] ∈ Z[1/n](cid:14)(1 − 1/n)Z[1/n]
id∗ − λ∗ : Z[1/n] ⊕ Z[1/n] −→ Z[1/n] ⊕ Z[1/n].
Hence, we have the isomorphisms:
(Z[1/n] ⊕ Z[1/n])(cid:14)(id∗ − λ∗)(Z[1/n] ⊕ Z[1/n]) ∼= Z[1/n](cid:14)(1 − 1/n)Z[1/n] ∼= Z(cid:14)(n − 1)Z.
where the last isomorphism follows from the proof of Proposition 2.17 with p = 1 and q = n.
Once we have computed the action of λ∗ on K1(Aλ
Now, by Proposition 2.11 of [Sc] we have the isomorphism:
0 ) = Z[1/n] we will be ready to compute K∗(Qλ).
K0(C(Γλ)) ∼= (Z[1] ⊗Z Z[1]) ⊕ (Z[1/n] ⊗Z Z[1/n]) = Z[1] ⊕ (Z[1/n] ⊗Z Z[1/n]).
The action of λ∗ on Z[1] is of course the identity. However, the action of λ∗ on (Z[1/n] ⊗Z Z[1/n])
is just x ⊗ y 7→ y ⊗ x/n. If one combines this with the multiplication isomorphism x ⊗ y 7→ xy :
Z[1/n]⊗Z Z[1/n] −→ Z[1/n] we see that λ∗ acts as multiplication by 1/n on Z[1/n] = Z[1/n]⊗Z Z[1/n].
Thus, λ∗ on the quotient K1(Aλ
0 ) = Z[1/n] is just multiplication by 1/n. Therefore, id∗ − λ∗ becomes
multiplication by (1 − 1/n) on Z[1/n] which is clearly 1 : 1. Applying the Pimsner-Voiculescu exact
sequence and recalling that Ki(Qλ) = Ki(Aλ) we get the isomorphisms:
K0(Qλ) ∼= Z(cid:14)(n − 1)Z, and K1(Qλ) ∼= Z(cid:14)(n − 1)Z,
For n > 2 we get K1 6= 0 and so these are not Cuntz algebras, in fact not even Cuntz-Krieger algebras
since K1 has nonzero torsion. For λ = 1/√2 however we get K0 = 0 = K1 and by classification theory,
we must have Q1/√2 ∼= O2! However, even in this case the fixed point algebra, is NOT AF since it has
K1 = Z[1/2], the tape-measure group. So for the simplest irrational number 1/√2 we get the Cuntz
algebra, O2 with a strange gauge action of T.
for λ = 1/√n.
13
Remarks. In the examples below it is important to note that any polynomial of the form f (x) =
xn + axn−1 + ··· + bx ± 1 has at most n − 1 roots in the open interval (0, 1) because the product of
all the roots of f must equal ±1.
Example(s) quadratic integers and an algorithm: If both λ and λ−1 are quadratic integers
with λ ∈ (0, 1), then λ2 + aλ ± 1 = 0 where the integer polynomial f (x) = x2 + ax ± 1 is irreducible
over Q. With these restrictions there are two cases, either f (x) = x2 + ax − 1 where a > 0 and
λ = 1/2·(√a2 + 4−a) ∈ (0, 1); or f (x) = x2+ax+1 where a ≤ −3 and λ = 1/2·(−√a2 − 4−a) ∈ (0, 1).
In the first case, λ2 + aλ − 1 = 0, with a > 0, so that λ + a − λ−1 = 0 and λ−1 = a + λ. For this case
we outline an algorithm using the ideas of the Smith Normal Form and the Pimsner-Voiculescu exact
sequence to calculate the K-Theory. By Proposition 2.2, and the CONCLUSION before Proposition
0 ) ∼= K1(C(Γλ)) =V1(Γλ) = Γλ ∼= Z2. Giving Γλ its Z-basis {1, λ} we
2.20, Γλ = Z + Zλ and K0(Aλ
1 −a (cid:21) . So, (id − λ∗) =
see that the action of the automorphism λ∗ on K0(Aλ
(cid:20) 1
−1 (a + 1) (cid:21) := M. To compute the kernel and cokernel of this matrix mapping Z2 → Z2 we
D =(cid:20) 1 0
0 a (cid:21) . Hence, on K0(Aλ
ker(id − λ∗) = ker(M ) ∼= ker(D) = {0} and coker(id − λ∗) ∼= coker(D) = Z/aZ.
row and column-reduce M over Z to obtain matrices P, Q ∈ GL(2, Z) so that P M Q = D where
D is diagonal over Z. Then ker(M ) ∼= ker(D) and Z2/M (Z2) ∼= Z2/D(Z2). In this case, we get
0 ) ∼= Γλ has matrix: (cid:20) 0
0 ) we have
−1
1
Now we compute (id − λ∗) on
K1(Aλ
0 ) ∼= K0(C(Γλ))/Z · 1o = (Z · 1o ⊕ Z(1 ∧ λ))/Z · 1o = Z(1 ∧ λ).
0 ) ∼= Z. Therefore,
Now, λ∗(1∧λ) = λ∧λ2 = λ∧(1−aλ) = λ∧1 = (−1)1∧λ. That is, λ∗ = −id on K1(Aλ
(id− λ∗) = multiplication by 2 on Z(1∧ λ) which has ker(id− λ∗) = {0} and cokernel(id− λ∗) ∼= Z/2Z.
Applying these results to the Pimsner-Voiculescu exact sequence we obtain:
K0(Qλ) = Z/aZ and K1(Qλ) = Z/2Z,
for λ2 + aλ − 1 = 0, n ≥ 1.
λ = (1/2)(√5 − 1) is the inverse of the golden mean, we get K0 = {0} and K1 = Z/2Z.
None of these examples are Cuntz-Krieger algebras since K1 is not torsion-free. In particular, when
0 ) ∼= Γλ = Z + Zλ with Z-basis
In the second case, λ2 + aλ + 1 = 0, we have as above, K0(Aλ
{1, λ}; the diagonal version of (id − λ∗) is D = diag[1, (a + 2)] so that ker(id − λ∗) = {0} and
0 ) ∼= Z(1 ∧ λ) only now, λ∗ = id here so that
coker(id − λ∗) ∼= Z/(a + 2)Z. On the other hand, K1(Aλ
(id − λ∗) = 0 and hence ker(id − λ∗) ∼= Z while coker(id − λ∗) ∼= Z. By Pimsner-Voiculescu we get
K0(Qλ) = Z ⊕ (Z/(a + 2)Z) and K1(Qλ) = Z,
for λ2 + aλ + 1 = 0, a ≤ −3.
We note that in this case, Qλ has the correct K-theory to be a Cuntz-Krieger algebra (and is therefore
stably isomorphic to one), and that in the case a = −3 (i.e., λ = (1/2)(3−√5)) we have K0 = Z = K1.
Example cubic integers: If λ and λ−1 are cubic integers with λ ∈ (0, 1), then λ3 + aλ2 + bλ± 1 = 0
where the integer polynomial f (x) = x3 + ax2 + bx ± 1 is irreducible over Q. Such an f is irreducible
if and only if f (1) 6= 0 6= f (−1). There are two cases depending on the constant, ±1.
First, consider f (x) = x3 + ax2 + bx − 1 = 0 with f (1) = a + b 6= 0 and f (−1) = a − b − 2 6= 0 so that
f is irreducible. Now assume a + b is positive (but a 6= b + 2). Then f (0) = −1 and f (1) = a + b > 0
so that f has a unique root in (0, 1) since it is a cubic.
14
Next consider the same polynomial, f (x) = x3 + ax2 + bx− 1 = 0, with a + b negative (but a 6= b + 2).
Since both f (0) and f (1) are negative, in order to have a solution the function f must have a local
maximum on (0, 1). There are examples with no solutions in (0, 1); for example, f (x) = x3 − 3x − 1.
In order to have a unique solution, then considering f′(x), one would need 4a2 − 12b = 0 : while
this has many solutions, they all satisfy a ≤ b and so we can not have a + b < 0. So solutions are not
unique in this case. But, there are infinitely many cubics with two distinct solutions in (0, 1); eg.,
f (x) = x3 − (a + k)x2 + ax − 1 for a ≥ k + 4 and k ≥ 1 has two solutions in (0, 1), since f (.5) > 0.
We now calculate the K-theory of Qλ assuming that λ satisfies f (x) = x3 + ax2 + bx − 1 = 0, where
a+b 6= 0, and a−b 6= 2. Now, λ3+aλ2+bλ−1 = 0, so that λ3 = 1−aλ2−bλ and λ−1 = λ2+aλ+b. Then,
0 ) ∼= K1(C(Γλ)) =Vodd(Γλ) = Γλ⊕(Γλ∧Γλ∧Γλ) = Γλ⊕(Z(1∧λ∧λ2)) ∼=
Γλ = Z+Zλ+Zλ2 and K0(Aλ
0 ) ∼= Z4
Z4. Giving Γλ its natural Z-basis {1, λ, λ2} the induced homomorphism (id − λ∗) on K0(Aλ
yields the diagonal matrix, D = diag[1, 1, (a + b), 0] so that on K0(Aλ
0 ) we have
Now, K1(Aλ
computations we get for K1(Aλ
0 ) we have
ker(id − λ∗) ∼= ker(D) ∼= Z and coker(id − λ∗) ∼= coker(D) = (Z/(a + b)Z) ⊕ Z.
0 ) ∼= K0(C(Γλ))/Z · 1o = V2(Γλ) = Z(1 ∧ λ) + Z(1 ∧ λ2) + Z(λ ∧ λ2) ∼= Z3. By similar
ker(id − λ∗) ∼= ker(D) = {0} and coker(id − λ∗) ∼= coker(D) = Z/(a + b)Z.
0 ) ∼= Z3; the matrix D = diag[1, 1, (a + b)]. Hence, on K1(Aλ
Applying these results to the Pimsner-Voiculescu exact sequence we obtain:
K0(Qλ) = Z ⊕ (Z/(a + b)Z) and K1(Qλ) = Z ⊕ (Z/(a + b)Z) for λ3 + aλ2 + bλ − 1 = 0.
Remarks. In case a + b = 1 (which has infinitely many solutions corresponding to infinitely many
distinct invertible cubic integers λ ∈ (0, 1)) we get K0(Qλ) = Z = K1(Qλ), which as noted above
is also true for the invertible quadratic integer, λ = (1/2)(3 − √5). In the general cubic case with
constant term −1 we always have non-torsion elements in both K0 and K1 : this is the opposite of
the case where the constant term is +1, where we see below that K0 and K1 are both torsion groups.
A similar phenomenon occurs in the quadratic case above, except that there we get torsion in the
−1 case and non-torsion in the +1 case! That this may be a periodic phenomenon is supported by
a calculation of two quartic examples: first, the unique solution λ ∈ (0, 1) to the irreducible quartic
f (x) = x4 − 3x3 + 1 gives us K0 = Z and K1 = Z ⊕ (Z/3Z) ⊕ (Z/3Z); while, second, the unique
solution λ ∈ (0, 1) to the irreducible quartic f (x) = x4 + 3x3 − 1 gives us K0 = (Z/3Z) ⊕ (Z/3Z) and
K1 = (Z/9Z) ⊕ (Z/2Z), similar to the quadratic case. Proposition 2.21 is further evidence.
When an irreducible polynomial f (x) = xn + axn−1 + ··· ± 1 has two roots, λ1, λ2 ∈ (0, 1), then
Γλ1 ∼= Γλ2 as rings (but not as ordered rings, for that would imply equality). Still, Qλ1 ∼= Qλ2 (at
least stably) since the calculation of their K-groups are identical. Their KMS states are not equivalent
since the type III factors that they generate are not isomorphic, as we will see below.
Proposition 2.21. Suppose λ satisfies the irreducible (over Z) polynomial, f (x) = xn + ··· ± 1 = 0.
(1) For n odd, if f (x) = xn + ··· + 1 then K0(Qλ) has Z/2Z as a summand.
While, if f (x) = xn+···−1 then K0(Qλ) has Z as a summand (so, by the next Proposition, rank(K0) =
rank(K1) ≥ 1 in this case).
(2) For n even, if f (x) = xn +··· + 1 then K1(Qλ) has Z as a summand (so, by the next Proposition,
rank(K0) = rank(K1) ≥ 1 in this case).
While, if f (x) = xn + ··· − 1 then K1(Qλ) has Z/2Z as a summand.
0 ) there is a λ∗-invariant summand, (1∧ λ∧ λ2 ∧···∧ λn−1)Z. Depending on n(mod 2)
Proof. In K∗(Aλ
and the constant term ±1, the action of λ∗ on this summand is ±id. Hence, (id − λ∗) here is either 0
or 2(id). Applying Pimsner-Voiculescu gives a summand in K∗(Qλ) of either Z or Z/2Z.
(cid:3)
Proposition 2.22. Suppose λ is algebraic.
(1) Then, rank(K0(Qλ)) = rank(K1(Qλ)) so that Qλ is not stably isomorphic to O∞.
(2) If λ and λ−1 are both algebraic integers and Qλ is stably isomorphic to a Cuntz algebra On, then the
minimal polynomial of λ has odd degree and constant term +1. Moreover, n is congruent to 3(mod 4)
and all such Cuntz algebras appear this way.
Proof. To see part (1) we tensor the Pimsner-Voiculescu exact sequence by Q (which preserves exact-
ness) to obtain an exact hexagon of Q-vector spaces:
15
V1
θ1
/ V1
τ1 /
/ K Q
1
µ0
µ1
K Q
0
τ0
V0
V0
θ0
where Vi = Ki(Aλ
0 ) ⊗ Q, and K Q
i = Ki(Aλ) ⊗ Q. Then
dim(K Q
0 ) = rank(µ0) + nullity(µ0) = rank(µ0) + rank(τ0) and dim(K Q
1 ) = rank(µ1) + rank(τ1)
and
rank(τ0) + rank(θ0) = dim(V0) = rank(θ0) + rank(µ1) so that rank(τ0) = rank(µ1).
Similarly, rank(τ1) = rank(µ0), so that:
dim(K Q
0 ) = rank(µ0) + rank(τ0) = rank(µ1) + rank(τ1) = dim(K Q
1 ).
That is, rank(K0(Qλ)) = dim K0(Qλ) ⊗Z Q = dim K0(Aλ) ⊗Z Q = ··· rank(K1(Qλ)). By Proposition
2.21, if the minimal polynomial of λ has even degree, then K1(Qλ) 6= {0}, and so Qλ cannot be stably
isomorphic to a Cuntz algebra. If Qλ is stably isomorphic to On then n is finite by part (1)and by
Proposition 2.21, the order of K0(Qλ) must be even and therefore n must be odd. Furthermore, the
minimal polynomial must have constant term +1. In order for K0(Qλ) to be a finite cyclic group of
even order, it must be of the form Z/mZ ⊕ Z/2Z where m is odd since Z/2Z is a summand. Let
m = 2k + 1 then
as claimed.
n = ♯[Z/mZ ⊕ Z/2Z] + 1 = 2m + 1 = 4k + 3
In the examples below where λ3 + aλ2 + bλ+ 1 = 0, and either a− b = 1 and b ≤ −2 OR a = b = −1
and b ≤ −1, we obtain (stably, at least) all the Cuntz algebras On where n ≡ 3(mod 4).
Now consider the case of irreducible cubics of the form f (x) = x3+mx2+nx+1; so f (1) = m+n+2 6= 0
and f (−1) = m − n 6= 0. Since f (0) = 1, if we have f (1) = m + n + 2 < 0, then we have as above a
unique root in (0, 1).
(cid:3)
If f (1) = m + n + 2 > 0, we can have distinct roots. For example, if n = −4 and m = 3, then,
f (x) = x3 + 3x2 − 4x + 1 has two roots in (0, 1). If n << 0, we get several solutions m for each n: eg.,
n = −7 implies that any m with 6 ≤ m ≤ 9 will yield a polynomial with two roots in (0, 1).
We now calculate the K-Theory of Qλ assuming λ satisfies f (λ) = λ3 + mλ2 + nλ + 1 = 0. Again,
0 ) ∼= Z3, the matrix
K0(Aλ
D = diag[1, 1, (m − n)]. Both matrices are 1 : 1 since m + n + 2 6= 0 6= m − n. We get:
0 ) ∼= Z4, but now the diagonal matrix D = diag[1, 1, (m + n + 2), 2]. On K1(Aλ
K0(Qλ) = Z/(n + m + 2)Z ⊕ Z/2Z and K1(Qλ) = Z/(m − n)Z for λ3 + mλ2 + nλ + 1 = 0.
To obtain Cuntz algebras, we need m − n = ±1. It turns out f (1) > 0 can not occur, so we must
have f (1) = m + n + 2 < 0 hence there is a unique root λ in (0, 1). Combining this inequality with
m − n = ±1 we get exactly two infinite families of solutions; m = n + 1 for n ≤ −2, OR m = n − 1
/
O
O
o
o
o
o
16
for n ≤ −1. In either case, the sequence of numbers {m + n + 2} is the same: {2k + 1k ≥ 0}. For
this sequence we get the K0 groups: Z/(2k + 1)Z ⊕ Z/2Z ∼= Z/(4k + 2)Z. Since the K1 groups are all
{0}, by construction, the algebras Qλ are (at least stably) the Cuntz algebras, O4k+3 for k ≥ 0. That
is, O3, O7, O11, etc.
Example, λ transcendental:
Lemma 2.23. Let ϕ : Ln∈Z Z −→ Z be the surjective homomorphism, φ({an}) := Pn∈Z an; and
let S ∈ Aut(Ln∈Z Z) be the shift S({an}n∈Z) := {an−1}n∈Z. Then, (id − S) is 1 : 1 and ker(ϕ) =
Im(id − S), so that (Ln∈Z Z)/Im(id − S) ∼= Z.
Proof. As a model for Ln∈Z Z we use Z[x, x−1] =Ln∈Z Zxn, the ring of Laurent polynomials over
Z (i.e., the group ring over Z of the group {xnn ∈ Z}). Here, ϕ is the augmentation map, S is
multiplication by x, and (id − S) is multiplication by (1 − x) which is 1 : 1. Now,
anxn+N ∈ ker(ϕ).
an = 0 ⇔
NXn=−N
where initially q(x) ∈ Q[x]. Since p(x) ∈ Z[x] it is easy to see that in fact, q(x) ∈ Z[x] also. Then,
n=−N an = 0. Hence, p(x) factors: p(x) = (1 − x)q(x)
NXn=−N
anxn ∈ ker(ϕ) ⇔
NXn=−N
n=−N anxn+N ∈ Z[x] so p(1) =PN
Let p(x) =PN
NXn=−N
anxn = x−N p(x) = (1 − x)x−N q(x) ∈ (1 − x)Z[x, x−1] = Im(id − S).
That is, ker(ϕ) ⊆ Im(id − S), and the other containment is immediate.
Proposition 2.24. If λ is transcendental then
(cid:3)
K0(Qλ) ∼=
∞Mn=1
Z ∼= K1(Qλ).
Proof. In this case, by Proposition 2.2 and the CONCLUSION before Proposition 2.20 we have:
K0(Aλ
0 ) =
∞Mk=1
2k−1^ (Γλ) and K1(Aλ
0 ) =
∞Mk=1
2k^(Γλ) where Γλ =
∞Mn=−∞
Zλn.
integers, we have:
sum of (λ∗-invariant) examples of the previous lemma where the action of λ∗ is just the shift. The
Now, each individual summandVm(Γλ) is invariant under λ∗ and yields (for m > 1) an infinite direct
general case is notation-heavy, so we do the examples, V2 and V3 . Letting Z+ denote the positive
2^(Γλ) = Mk∈Z+ Mn∈Z
(λn ∧ λn+k1 ∧ λn+k1+k2)Z! .
The case m = 1 is just the group Γλ =Ln∈Z Zλn which yields a single instance of the lemma.
Applying the lemma we see that (id − λ∗) is 1 : 1 on both K0(Aλ
K0(Aλ
of copies of Z. An application of the Pimsner-Voiculescu exact sequence completes the proof.
0 ) and that both
0 )) are isomorphic to a countable direct sum
(cid:3)
(λn ∧ λn+k)Z! and
3^(Γλ) = M(k1,k2)∈Z2
0 )/(id − λ∗)(K1(Aλ
0 )/(id − λ∗)(K0(Aλ
Mn∈Z
0 )) and K1(Aλ
0 ) and K1(Aλ
+
Remark. The classification theory of Kirchberg algebras implies that for λ transcendental we have a
new realisation of the algebras found in [Cu1] and denoted QN there.
2.5. The dual action of T1 on Aλ and its restriction to the gauge action on Qλ. Recall,
G0
λ = {g ∈ Gλ g = 1} is a normal subgroup of Gλ. The subgroup of Gλ of elements of the form
[λn : 0] is isomorphic to Z and acts on G0
λ by conjugacy:
[λn : 0][1 : b][λ−n : 0] = [1 : λnb].
Thus Gλ = Z ⋊ G0
λ is a semidirect product and we can write Aλ as an iterated crossed product:
17
Aλ = Gλ ⋊α C λ
0 (R) = Z ⋊ (G0
λ ⋊α C λ
0 (R)) = Z ⋊ Aλ
0 .
The dual action γ of T1 on Aλ is relative to this latter crossed product so that for each z ∈ T1 and x
in the Banach ∗-algebra, l1
0 (R)) we have:
α(Gλ, C λ
γz(x)(g) = znx(g) if x ∈ l1
α(Gλ, C λ
0 (R)); g ∈ Gλ and g = λn.
Since Aλ is defined to be the completion of this Banach ∗-algebra in its universal representation, the
action γ extends uniquely to an action (also denoted by γ) of T1 as automorphisms of Aλ. The fixed
point subalgebra of the dual action is, of course, exactly Aλ
0 = G0
λ ⋊α C λ
0 (R).
0 , the action γ restricts to an action of T1 on Qλ = eAλe, which we will
Since the projection e is in Aλ
also denote by γ. We call this the gauge action of T1 on Qλ. Now, γ is clearly a strongly continuous
action of T1 on Qλ. Averaging over γ with respect to normalised Haar measure gives a positive,
faithful expectation Φ of Qλ onto the fixed-point algebra which is clearly F λ:
Φ(a) :=
1
2πZT1
γz(a) dθ for a ∈ Qλ, and z = eiθ.
Proposition 2.25. The fixed point algebra, F λ = eAλ
tions of elements of the form:
0 e is the norm closure of finite linear combina-
for a, b, c ∈ Γλ. Recall, Aλ
X[a,b) · δg where g = [1 : c] and [a, b) ⊆ [0, 1) ∩ [c, 1 + c),
0 ∼= K(l2(Z)) ⊗ F λ.
Proof. Applying the integral formula for Φ to a finite linear combination of the generators for Qλ we
see that the only terms that survive are those where g = 1 : that is, g has the above form. Then we
apply item (2) of Lemma 2.12 to obtain the condition on the interval [a, b).
Corollary 2.26. The stabilised algebra Qλ⊗K is a crossed product of the stabilised fixed-point algebra
F λ ⊗ K by an action of Z. For λ = 1/n this is a theorem of J. Cuntz.
(cid:3)
0 ∼= F λ ⊗ K, and Aλ ∼= Qλ ⊗ K. By the discussion at the beginning of
0 and the proof is complete. See [Cu, Section 2].
Proof. By Proposition 2.7, Aλ
subsection 3.1, Aλ ∼= Z ⋊ Aλ
Remarks. If we combine the previous observation that F λ is the fixed point subalgebra of Qλ under
the gauge action with Corollary 2.18 we get, for example, O2 ∼= Q2/3 with a gauge action whose fixed
point subalgebra F 2/3 is a UHF algebra of type 6∞. Interestingly, F 3/4 is UHF of type 12∞ = 6∞
which is therefore isomorphic to F 2/3. So we have two gauge actions on O2 with isomorphic UHF
fixed point subalgebras, with distinct, inequivalent KMS states: one where β = log(3/2) and the other
where β = log(4/3) by Proposition 2.30 below. Moreover, the two von Neumann algebras generated
by the GNS representations of O2 are not isomorphic as they are type IIIλ factors for λ equalling 2/3
and 3/4, respectively, by Theorem 2.35 below.
(cid:3)
18
2.6. The γ-invariant semifinite weight on Aλ and its restriction to Qλ. The aim of this
subsection is to exhibit the unique KMS states for the gauge action on Qλ. We first recall the
definition of KMS states.
Definition 2.27. Let A be a C∗-algebra with a continuous action γ : R → Aut(A). Let ψ be a state
on A and β ∈ R a real number. We define ψ to be a KMSβ state for the action γ if
ψ(x γiβ (y)) = ψ(yx)
for all x, y ∈ A a dense γ-invariant ∗-subalgebra of Aγ, the subalgebra of analytic elements for the
action γ. We refer to [BR1, Section 2.5] for basic information on the subalgebra of analytic elements,
Aγ and to [BR2, Section 5.3] for all the basic information on KMS states.
Since Gλ is discrete it is well-known that the map
0 (R)) → C λ
0 (R)
extends uniquely to a faithful conditional expectation E : Aλ → C λ
defined (norm) lower semicontinuous weight on C λ
(norm) lower semicontinuous weight on Aλ which we denote by ¯ψ. In particular, for x ∈ l1
we have:
0 (R). Composing E with the densely
0 (R) given by integration, gives us a densely defined
0 (R))
x 7→ x(1) : l1
α(Gλ, C λ
α(Gλ, C λ
¯ψ(xx∗) =ZR
xx∗(1)(t)dt =ZRXh∈Gλ
x(h)x(h) (t)dt = Xh∈Gλ(cid:18)ZR x(h)(t)2dt(cid:19) .
So that ¯ψ is faithful. We observe that ¯ψ is not a trace, since ¯ψ(x∗x) =Ph∈Gλ h−1RR x(h)(t)2dt.
Proposition 2.28. The weight ¯ψ on Aλ restricts to a faithful semifinite trace ¯τ on Aλ
restricts to a state denoted by ψ on Qλ satisfying:
(1) The gauge action γ of T1 on Qλ leaves the state ψ invariant.
(2) The state ψ restricted to the fixed point algebra, F λ is a faithful (finite) trace denoted by τ ; which
is, of course, the restriction of ¯τ on Aλ
(3) With Φ : Qλ → F λ the canonical expectation, we have ψ = τ ◦ Φ.
Proof. Since ¯ψ(e) =RR X[0,1)(t)dt = 1, we see that ¯ψ restricted to Qλ is a faithful state. To see item
0 (R)) leaves ¯ψ invariant. To this end, let x ≥ 0
(1), it suffices to see that the gauge action on l1
be in l1
0 and also
α(Gλ, C λ
0 to F λ.
α(Gλ, C λ
0 (R)), and let z ∈ T1. Then
and so
E(γz(x)) = γz(x)(1) = z0x(1) = x(1) = E(x)
¯ψ(γz(x)) =ZR
γz(x)(1)(t)dt =ZR
E(x)(t)dt = ¯ψ(x).
To see item (2) we use Proposition 2.25 and the above computation that shows that while ¯ψ is not
generally a trace, to see that it is a trace when the group elements all have determinant 1.
To see item (3), it suffices to see that for any x ∈ Qλ we have E(x) = E(Φ(x)), but this is the same
as x(1) = Φ(x)(1) which is clear since det(1) = 1.
(cid:3)
Now, since the state ψ is invariant under the action γ, this action is unitarily implemented on
L2(Qλ, ψ). For z ∈ T1 and x ∈ Qλ
c we define:
(uz(x))h = znxh for h ∈ Gλ with h = λn.
19
We define the spectral subspaces of this unitary group on L2(Qλ, ψ) in the usual way. For each k ∈ Z
let Φk be the operator on L2(Qλ, ψ) :
2πZT1
z−kuz(x)dθ, z = eiθ, x ∈ L2(Qλ, ψ).
We observe that if x = f · δg is a typical generator of Qλ considered as a vector in L2(Qλ, ψ) then we
have:
Φk(x) =
1
Φk(f · δg) =(cid:26) f · δg
0
if g = λk
otherwise
More generally, on H := L2(Qλ, ψ), we have Φk(H) = {x ∈ H uz(x) = zkx for all z ∈ T1}.
Lemma 2.29. For each k ∈ Z the subspace span{f · δg ∈ Qλ g = λk} is dense in the range of
Φk. The operators Φk are mutually orthogonal projections on H which sum to the identity operator
1 = π(e).
Proof. The proof of the first statement is similar to the proof of Proposition 2.25. The mutual orthog-
onality of the Φk follows from the fact that hf1 · δg1f2 · δg2iψ = 0 unless g1 = g2.
(cid:3)
Proposition 2.30. The dense ∗-subalgebra of Qλ consisting of finite linear combinations of the partial
isometries X[a,b) · δg is contained in the subset of entire elements, Qλ
γ , for the action γ considered as
an action of R : t 7→ γeit. Moreover, ψ is a KMSβ state for this action where β = log(λ−1). In fact, ψ
is the unique KMS state for this action (regardless of β).
Proof. Let y = X[a,b) · δg ∈ Qλ where det(g) = λk. Then, t 7→ γeit(y) = eikty; t ∈ R obviously extends
to the entire function w 7→ γeiw (y) = eikwy; w ∈ C. For w = log(λ−1)i, this equation becomes
γeiw (y) = γλ(y) = λky. Letting β = log(λ−1), we have γβi(y) = λky. Now, let x = X[c,d) · h so we want
to see that: λkψ(xy) = ψ(xγβi(y)) = ψ(yx). That is, we want λkψ(xy) = ψ(yx). Now both sides of
this equation are zero unless h = g−1. But, when h = g−1, we have
xy = X[c,d) · X[g−1(a),g−1(b)) · δI while yx = X[a,b) · X[g(c),g(d)) · δI .
Moreover,
s ∈ [c, d) ∩ [g−1(a), g−1(b)) ⇐⇒ g(s) ∈ [g(c), g(d)) ∩ [a, b).
Since det(g) = λk the transformation g increases the measure by a factor of λk and the result follows.
That is, ψ is a KMSβ state for the action γ of R for β = log(λ−1).
Now let φ be a KMS state on Qλ for the action γ. Since Qλ is purely infinite it has no nontrivial traces
and so φ must be KMS for some nonzero β. Hence by [BR2, Proposition 5.3.3], φ is invariant under
the action of γ. Now, if X[a,b) · δg ∈ Qλ with det(g) = λk, then we have for all z ∈ T:
φ(X[a,b) · δg) = φ(γz(X[a,b) · δg)) = zkφ(X[a,b) · δg).
That is, if det(g) 6= 1 we must have φ(X[a,b) · δg) = 0, and so φ is supported on F λ. Since F λ is
γ-invariant and φ is KMS for some nonzero β, φ is a trace on F λ by [BR2, 5.3.28].
Now, if x = X[a,b) · δg ∈ F λ and g 6= I, then we claim that φ(x) = 0. For suppose g = [1 : c] with
c > 0. Then there is a positive integer n such that a + nc < b ≤ a + (n + 1)c and so
x = X[a,b) · δg = X[a,c) · δg + X[a+c,a+2c) · δg + ··· + X[a+nc,b) · δg := v0 + v1 + ··· + vn.
Now each of these partial isometries vk satisfies v2
φ is a trace on F λ. Thus, φ(x) = 0 as claimed.
k = 0, and so φ(vk) = φ(vkv∗kvk) = φ(v2
kv∗k) = 0 since
20
Hence φ is supported on the commutative subalgebra
C := span{f · δI f ∈ C λ
0 (R) and supp(f ) ⊆ [0, 1)}.
0 ∼= F λ ⊗ K we can define a lower semicontinuous, densely defined trace, T r on Aλ
Morever, if f1, f2 are characteristic functions of subintervals of [0, 1) with endpoints in Γλ and having
the same length they give equivalent elements fi · δI in F λ and therefore have the same value under φ.
Now, since Aλ
0 via
T r = φ ⊗ T r, where T r is the trace on K. So, for X[a,b) · δI ∈ F λ we have T r(X[a,b) · δI ) = φ(X[a,b) · δI ).
Then, for k1 < k2 ∈ Z the element X[k1,k2) · δI is the sum of (k2 − k1) projections in Aλ
0 each equivalent
to X[0,1) · δI which has trace equal to 1; that is, T r(X[k1,k2) · δI ) = (k2 − k1). Now, for any a < b in Γλ,
we have X[a,b) · δI ∼ X[0,b−a) · δI and so T r(X[a,b) · δI ) = T r(X[0,b−a) · δI ), and these values are finite
since both these projections are dominated by X[−N,N ) · δI for a sufficiently large integer N. It now
suffices to prove the following. Claim: T r(X[a,b) · δI ) = b − a for a < b ∈ Γλ.
By the previous discussion we can assume that a = 0 so that b > 0. Given ε > 0 we choose positive
integers m, n such that
1
m ≤ ε and
n − 1
m ≤ b <
n
m
,
so that (n − 1) ≤ bm < n and (n − 1), bm, n ∈ Γλ. Hence
(n − 1) = T r(X[0,(n−1)) · δI ) ≤ T r(X[0,bm) · δI ) ≤ T r(X[0,n) · δI ) = n.
But,
and these projections are mutually equivalent in Aλ
X[0,bm) · δI = X[0,b) · δI + X[b,2b) · δI + ··· + X[(m−1)b,bm) · δI
(n − 1) ≤ m T r(X[0,b) · δI ) ≤ n so that
0 . That is,
n − 1
m ≤ T r(X[0,b) · δI ) ≤
n
m
.
m ≤ ε. That is, b = T r(X[0,b) · δI ) = φ(X[0,b) · δI ) and φ agrees with the
Hence, T r(X[0,b) · δI ) − b < 1
given trace τ on F λ and therefore φ agrees with ψ on Qλ.
Remarks. The above proof shows that the algebra F λ has a unique (faithful) tracial state τ, and
that Aλ
0 has a unique (faithful) lower semicontinuous, densely defined trace normalized so that it has
value 1 at e = X[0,1) · δI .
2.7. The von Neumann algebra π(Aλ)−wo acting on L2(Aλ, ¯ψ) is a type IIIλ factor. To prove
this we will show that it is unitarily equivalent to a version of the Murray-von Neumann "group-
measure space" construction of type III factors on l2(Gλ) ⊗ L2(R) : see [D, Chapter 1, Section 9]. We
conclude that it is a IIIλ factor by an appeal to Connes' thesis [C0]. In order to be consistent with
our use of right C∗-modules later, we will do our GNS constructions so that our inner products are
linear in the second variable.
Proposition 2.31. The ∗-algebra Aλ
c is a Tomita algebra with the inner product:
(cid:3)
hyxi ¯ψ = ¯ψ(y∗x) = Xh∈Gλ
h−1hxhyhiL2(R).
Here we denote xh in place of x(h) to simplify notation. In this setting we have for x ∈ Aλ
c :
(1) Sharp: S(x)h = αh(xh−1);
(2) Flat: F (x)h = hαh(xh−1);
(3) Delta: ∆(x)h = hxh.
21
Proof. We refer to [Ta] for Takesaki's version of the axioms for a Tomita algebra. Since Sharp is defined
to be the adjoint operation on the algebra, item (1) is immediate. A straightforward calculation shows
that for all x, y ∈ Aλ
c we have that the defining equation for F lat holds, namely:hS(y)xi ¯ψ = hF (x)yi ¯ψ
so that item (2) holds. By definition, ∆ = F S and so a simple calculation shows that ∆(x)h = hxh
and (3) holds. From this formula for ∆ we see that for each z ∈ C we have ∆z(x)h = hzxh and a
straightforward calculation shows that ∆z(x · y) = (∆z(x)) · (∆z(y)) so that each ∆z is an algebra
homomorphism of Aλ
c as required. That each left multiplication π(x) is bounded when x is supported
on a single group element is straightforward and the generalization to finitely supported elements is
then trivial. The fact that it is a ∗-representation holds as it does for the GNS representation for any
weight.
In order to see that products are dense we recall that we have local units. That is, for each positive
integer N we have defined EN = X[−N,N ) · δ1, and have noted that for each y ∈ Aλ
c that satisfies
supp(yh) ⊆ [−N, N ) for all h, we have EN · y = y. Axioms IV, V, VI in [Ta] are simple calculations
involving the definitions of S, F , and ∆ which we leave to the reader.
Since our inner products are linear in the second variable, we modify Tomita's Axiom VIII to
c . We easily calculate that hx∆z(y)i ¯ψ =
read: z 7→ hx∆x(y)i ¯ψ is analytic on C for all x, y ∈ Aλ
Ph hz−1hyhxhiL2(R). This function is analytic since the sum is finite.
c , ¯ψ) decomposes as the integrated form of a covariant
(cid:3)
Lemma 2.32. The representation of Aλ
pair of representations:
c on L2(Aλ
00(R) → B(L2(Aλ
(1) π : C λ
(2) U : Gλ → U (L2(Aλ
c , ¯ψ)), where : (π(f )(y))h = f · yh for f ∈ C λ
00(R) and y ∈ Aλ
c ;
c , ¯ψ)) where : (Ug(y))h = αg(yg−1h) for g ∈ Gλ and y ∈ Aλ
c .
Proof. It is straightforward to verify that U is a unitary representation of Gλ and that π is a ∗-
representation of C λ
00(R). To see the covariance condition:
(Ug π(f )Ug−1(y))h = ··· = αg(f · αg−1(ygg−1)) = αg(f ) · yh = (π(αg(f ))y)h.
That is, Ug π(f )Ug−1 = π(αg(f )).
Now, by Proposition 7.6.4 of [Ped] the integrated form of this covariant pair is the representation:
(π × U )(y) =Xh
Now, we evaluate this operator on the vector x ∈ Aλ
c :
[π(yh)Uh(x)]k =Xh
yhαh(xh−1k) = (y · x)(k) = (π(y)(x))k.
[((π × U )(y))(x)]k =Xh
π(yh)Uh for y ∈ Aλ
c .
That is, (π × U )(y) = π(y) the operator left multiplication by y.
(cid:3)
2.7.1. A representation of Aλ on l2(Gλ) ⊗ L2(R). We define a covariant pair of representations of
0 (R) and Gλ on l2(Gλ) ⊗ L2(R) as follows:
C λ
(1) for f ∈ C λ
0 (R) let π(f ) = 1 ⊗ Mf , and (2) for g ∈ Gλ let U g = Λ(g) ⊗ Vg
where Λ is the left regular representation of Gλ on l2(Gλ) :
(Λ(g)ξ)(h) = ξ(g−1h) for ξ ∈ l2(Gλ);
and V is the unitary action of Gλ on L2(R) induced by the action of Gλ on R :
(Vg(f ))(t) = g−1/2f (g−1t); for f ∈ L2(R).
22
Using these equations one easily checks the covariance condition for g ∈ Gλ and f ∈ C λ
0 (R) :
U gπ(f )U∗g = π(αg(f )).
Clearly the representation π extends uniquely by weak-operator continuity to the usual representation
1 ⊗ M of L∞(R) on l2(Gλ) ⊗ L2(R) and is covariant with the unitary representation U of Gλ for the
action α of Gλ on L∞(R). Clearly, the von Neumann algebra on l2(Gλ) ⊗ L2(R) generated by the
unitaries U g and the operators 1 ⊗ Mf for g ∈ Gλ and f ∈ C λ
00(R), is the same as the von Neumann
algebra generated by the unitaries U g and the operators 1 ⊗ Mf for g ∈ Gλ and f ∈ L∞(R). The
second item of the following Proposition is clear.
Proposition 2.33. (1) The representation π = (π × U ) of Aλ on L2(Aλ
the representation (π × U ) of Aλ on l2(Gλ) ⊗ L2(R).
(2) (π × U )(Aλ)′′ is the von Neumann crossed product (in the sense of the group-measure space con-
struction of Chapter 1 Section 9 of [D]) Gλ ⋊α L∞(R).
(3) This von Neumann algebra is a type III factor.
c , ¯ψ) is unitarily equivalent to
Proof. To see item (3) we use the proof of [D, Theorem 2, Section 9, Chapter 1] where instead of the
ax + b group G with a, b ∈ Q and a > 0 and its subgroup G0 (with a = 1), we use Gλ and its subgroup
G0
λ (with g = 1), to conclude that our von Neumann algebra is a type III factor.
To see item (1), we first define a unitary W : L2(Aλ
fi · δhi! =
c , ¯ψ) → l2(Gλ) ⊗ L2(R) as follows:
mXi=1
W mXi=1
hi−1/2δhi ⊗ fi.
On the left side of this equation we are using the formalism f · δh for singly supported elements in Aλ
with f ∈ C λ
00(R) and h ∈ Gλ. On the right of this equation we are using δh to denote the canonical
orthonormal basis elements in l2(Gλ) and regarding f ∈ C λ
00(R) ⊂ L2(R). Clearly, W is well-defined
and linear with dense range. One easily checks that: for all x, y ∈ Aλ
c we have
c
recalling that the inner product on Aλ
inverse (adjoint) defined at first on the elements in l2(Gλ) ⊗ L2(R) of the formPm
fi ∈ C λ
00(R), is given by:
hyxi ¯ψ = hW (x)W (y)il2⊗L2
c is linear in the second coordinate. Thus W is a unitary and its
i=1 δhi ⊗ fi with the
δhi ⊗ fi! =
One then verifies the following two equations for f ∈ C λ
W ∗ mXi=1
mXi=1
hi1/2fi · δhi.
00(R) and g ∈ Gλ :
(1) W π(f )W ∗ = 1 ⊗ Mf = π(f ) and (2) W UgW ∗ = Λ(g) ⊗ Vg = U g.
The second equation is more subtle and requires the observation: Ug(f · δh) = g1/2Vg(f ) · δgh.
This completes the proof of the proposition.
(cid:3)
c , ¯ψ) is type IIIλ. We work in the unitarily equivalent setting
2.7.2. The factor π(Aλ)′′ acting on L2(Aλ
of (π× U )(Aλ)′′ acting on l2(Gλ)⊗L2(R) afforded by Proposition 2.33. Recall that the subgroup of Gλ
of matrices of the form [λn : 0] is isomorphic to Z and acts on the normal subgroup G0
λ by conjugacy,
and so Gλ = Z ⋊ G0
λ is a semidirect product and we can write a canonical right coset decomposition
of Gλ :
Gλ = [n∈Z
λ · [λn : 0].
G0
23
This gives us an internal orthogonal decomposition of l2(Gλ) :
l2(Gλ) =Xn∈Z
⊕ l2(cid:0)G0
Here the latter isomorphism is given explicitly on basis elements by the map which takes the δ-function
at g · [λn : 0] to δn ⊗ δg for n ∈ Z and g ∈ G0
λ.
One checks that the restriction of the representation (π× U) of Aλ = Gλ ⋊ C λ
on l2(Gλ)⊗L2(R) is unitarily equivalent to the representation on l2(Z)⊗l2(G0
pair:
λ · [λn : 0](cid:1) ∼= l2(Z) ⊗ l2(G0
0 (R) to Aλ
0 (R)
λ)⊗L2(R) via the covariant
λ).
0 := G0
λ ⋊ C λ
1Z ⊗ Λ(h) ⊗ Vh = 1Z ⊗ U h for h ∈ G0
1Z ⊗ 1 ⊗ Mf = 1Z ⊗ π(f ) for f ∈ C λ
λ and
0 (R).
Therefore, the von Neumann subalgebra of (π × U )(Aλ)′′ generated by (π × U )(Aλ
the von Neumann algebra on l2(G0
1⊗Mf for f ∈ C λ
Λ(h) ⊗ Vh for h ∈ G0
II∞ by the methods of [D, Chapter 1, Section 9]. Thus, (π × U )(Aλ
type III factor, (π × U )(Aλ)′′. Moreover, the faithful normal semifinite trace on (π × U )(Aλ
by the restriction of ¯ψ.
0 ) is isomorphic to
λ and
0 (R). This is clearly the same as the von Neumann algebra generated by the operators
λ and 1 ⊗ Mf for f ∈ L∞(R), and this von Neumann algebra is a factor of type
0 )′′ is a type II∞ subfactor of the
0 )′′ is given
λ) ⊗ L2(R) generated by the operators Λ(h) ⊗ Vh for h ∈ G0
Finally, conjugation by the unitary, U g for g = [λ : 0], which lies in our type III factor, defines an
automorphism β of the type II∞ subfactor which scales the trace by λ. If N0 is our type II∞ factor
λ)⊗L2(R) is unitarily
acting on l2(G0
equivalent to the von Neumann crossed product Aλ ∼= Z ⋊β N0 and hence is a type IIIλ factor by [C0,
λ)⊗L2(R) then our type III factor, say Aλ acting on l2(Z)⊗ l2(G0
Theorem 4.4.1]. We have proved the following Proposition.
c , ¯ψ) is a type IIIλ factor.
Proposition 2.34. The von Neumann algebra π(Aλ)′′ acting on L2(Aλ
Moreover, it is unitarily equivalent to (π × U )(Aλ)′′ acting on l2(Gλ) ⊗ L2(R). The von Neumann
subalgebra of (π × U )(Aλ)′′ generated by (π × U )(Aλ
0 ) is a type II∞ factor. The space l2(Gλ) ⊗ L2(R)
λ)⊗L2(R) and with this factorization, our II∞ factor has the form N0 = 1Z⊗ N0
factors as l2(Z)⊗l2(G0
where N0 acts on l2(G0
λ)⊗L2(R). Thus, our type IIIλ factor is unitarily equivalent to the von Neumann
crossed product Z ⋊β N0 where the automorphism β of N0 is given by β = Ad(U g) where g = [λ : 0].
2.8. The von Neumann algebra, π0(Qλ)−wo acting on L2(Qλ, ψ) is type IIIλ.
Theorem 2.35. The von Neumann algebra, π0(Qλ)−wo acting on L2(Qλ, ψ) is type IIIλ. Moreover,
the von Neumann subalgebra, π0(F λ)−wo is a type II1 factor with unique faithful normal state given
by the restriction of the vector state, ψ which is the same as τ on F λ. By the general theory of type
III factors, π0(Qλ)−wo is isomorphic to π(Aλ)−wo acting on L2(Aλ
Proof. Recall that Qλ = eAλe where e = X[0,1) · δ1 ∈ Aλ. Then
c , ¯ψ).
π(e)(π(Aλ)−wo)π(e) = (π(e)π(Aλ)π(e))−wo = π(Qλ)−wo
and the cut-down of the type III factor π(Aλ)−wo (on its separable Hilbert space) by the nonzero
projection π(e) is isomorphic to π(Aλ)−wo since π(e) is Murray-von Neumann equivalent to the
identity operator. Of course the cut-down mapping by π(e) is not an isomorphism. Moreover,
by left Hilbert algebra theory, the operator right multiplication by e which is denoted by π′(e)
is in the commutant of π(Aλ)−wo acting on L2(Qλ, ¯ψ) and since we are in a factor the mapping
π(Aλ)−wo → π′(e)π(Aλ)−wo is an isomorphism by [D, Chapter 1, Section 2, Prop. 2]. Restricting
this isomorphism to π(Qλ)−wo gives us an isomorphism π(Qλ)−wo → π′(e)π(Qλ)−wo which acts on the
24
Hilbert space π′(e)π(e)(L2(Aλ, ¯ψ)), which has as a dense subspace π′(e)π(e)(Aλ
c e ⊂ eAλe = Qλ
with the inner product given by ¯ψ which is the same as the inner product on eAλ
c e given by the state
ψ. The completion of this space is, of course, L2(Qλ, ψ) with the action of Qλ being the GNS represen-
tation afforded by the state ψ. We denote this representation of Qλ on L2(Qλ, ψ) by π0 to distinguish
it from the representation π of Aλ on the larger space, L2(Aλ
Similar considerations applied to the type II∞ subfactor, π(Aλ
that:
c , ¯ψ).
0 )−wo ⊂ π(Aλ)−wo on L2(Aλ
c , ¯ψ), show
c ) = eAλ
π(e)(π(Aλ
0 )−wo)π(e) = (π(e)π(Aλ
0 )π(e))−wo = π(F λ)−wo.
0 )−wo of π(Aλ)−wo and has finite (ψ)
Now the projection π(e) is actually in the type II∞ subfactor π(Aλ
trace = 1 there. Therefore, π(F λ)−wo is a type II1 factor on L2(Qλ, ψ) with trace given by the vector
state ψ. We remark that this is clearly a larger space than the subspace, L2(F λ, τ ) ⊂ L2(Qλ, ψ). (cid:3)
c is a Tomita algebra with the inner product: hyxiψ = ψ(y∗x).
Proposition 2.36. The ∗-algebra Qλ
Again we denote xh in place of x(h) to simplify notation. In this setting we have for x ∈ Qλ
c :
(1) Sharp: S(x)h = αh(xh−1);
(2) Flat: F (x)h = hαh(xh−1);
(3) Delta: ∆(x)h = hxh.
Proof. This is really a corollary of Proposition 2.7, as Qλ
c is just a Tomita-subalgebra of Aλ
c .
(cid:3)
3. The modular spectral triple of the algebra Qλ
Having introduced the main features of the algebras Qλ, we now turn briefly to the modular index
theory of [CNNR, CPR2, CRT]. We begin with some semifinite preliminaries.
3.1. Semifinite noncommutative geometry. We need to explain some semifinite versions of stan-
dard definitions and results following [CPRS2]. Let φ be a fixed faithful, normal, semifinite trace on
a von Neumann algebra N . Let KN be the φ-compact operators in N (that is the norm closed ideal
generated by the projections E ∈ N with φ(E) < ∞).
Definition 3.1. A semifinite spectral triple (A,H,D) is given by a Hilbert space H, a ∗-algebra
A ⊂ N where N is a semifinite von Neumann algebra acting on H, and a densely defined unbounded
self-adjoint operator D affiliated to N such that [D, a] is densely defined and extends to a bounded
operator in N for all a ∈ A and (λ − D)−1 ∈ KN for all λ 6∈ R. The triple is said to be even if there
is Γ ∈ N such that Γ∗ = Γ, Γ2 = 1, aΓ = Γa for all a ∈ A and DΓ + ΓD = 0. Otherwise it is odd.
Note that if T ∈ N and [D, T ] is bounded, then [D, T ] ∈ N .
We recall from [FK] that if S ∈ N , the t-th generalized singular value of S for each real t > 0 is
given by
µt(S) = inf{kSEk : E is a projection in N with φ(1 − E) ≤ t}.
The ideal L1(N , φ) consists of those operators T ∈ N such that kTk1 := φ(T) < ∞ where T =
√T ∗T . In the Type I setting this is the usual trace class ideal. We will denote the norm on L1(N , φ)
by k · k1. An alternative definition in terms of singular values is that T ∈ L1(N , φ) if kTk1 :=
R ∞0 µt(T )dt < ∞. When N 6= B(H), L1(N , φ) need not be complete in this norm but it is complete in
the norm k · k1 + k · k∞. (where k · k∞ is the uniform norm). We use the notation
L(1,∞)(N , φ) =(cid:26)T ∈ N : kTkL(1,∞) := sup
t>0
1
log(1 + t)Z t
0
µs(T )ds < ∞(cid:27) .
25
measurable operators affiliated to N . Our notation is however consistent with that of [C] in the
special case N = B(H). With this convention the ideal of φ-compact operators, K(N ), consists of
The reader should note that L(1,∞)(N , φ) is often taken to mean an ideal in the algebra eN of φ-
those T ∈ N (as opposed to eN ) such that µ∞(T ) := limt→∞ µt(T ) = 0.
Definition 3.2. A semifinite spectral triple (A,H,D) relative to (N , φ) with A unital is (1,∞)-
summable if (D − λ)−1 ∈ L(1,∞)(N , φ) for all λ ∈ C \ R.
It follows that if (A,H,D) is (1,∞)-summable then it is n-summable (with respect to the trace φ) for
all n > 1. We next need to briefly discuss Dixmier traces. For more information on semifinite Dixmier
traces, see [CPS2, CRSS]. For T ∈ L(1,∞)(N , φ), T ≥ 0, the function
FT : t →
is bounded. There are certain ω ∈ L∞(R+
by setting
1
log(1 + t)Z t
µs(T )ds
0
∗ )∗, [CPS2, C], which define (Dixmier) traces on L(1,∞)(N , φ)
φω(T ) = ω(FT ), T ≥ 0
and extending to all of L(1,∞)(N , φ) by linearity. For each such ω we write φω for the associated
Dixmier trace. Each Dixmier trace φω vanishes on the ideal of trace class operators. Whenever the
function FT has a limit at infinity, all Dixmier traces return that limit as their value. This leads to
the notion of a measurable operator [C, LSS], that is, one on which all Dixmier traces take the same
value.
3.2. The Kasparov module and modular spectral triple. We have seen that the algebras Qλ
do not possess a faithful gauge invariant trace but that there is a KMSβ where β = − log(λ) for the
gauge action, γ, namely ψ := τ ◦ Φ : Qλ → C, where Φ : Qλ → F λ is the expectation and τ : F λ → C
In fact, ψ is the only KMS state for the gauge action (for any β),
is a faithful normalised trace.
by Proposition 2.30. We show below that the generator of the gauge action D acting on a suitable
C∗-F λ-module X gives us a Kasparov module (X,D) whose class lies in KK 1,T(Qλ, F λ). In some
examples, including the case λ ∈ Q, we have K1(Qλ) = {0} and so pairing with ordinary K1 would be
fruitless. However, following [CPR2, CNNR] we may compute a numerical pairing using a 'modular
spectral triple' constructed from the Kasparov module.
We now review this construction adapted to the present situation. Let H = L2(Qλ) be the GNS
Hilbert space given by the faithful state ψ with the inner product on Qλ defined by ha, bi = ψ(a∗b) =
(τ ◦ Φ)(a∗b). Then D is a self-adjoint unbounded operator on H, [CPR2]. The representation of Qλ
on H by left multiplication (which we now denote by π in place of π0) is bounded and nondegenerate:
the left action of an element a ∈ Qλ by π(a) satisfies π(a)b = ab for all b ∈ Qλ. This distinction
between elements of Qλ as vectors in L2(Qλ) and operators on L2(Qλ) is sometimes crucial. The
dense subalgebra Qλ
c e which is the finite span of elements in Qλ of the form X[a,b) · δg is in the
smooth domain of the derivation δ = ad(D). We remind the reader that the KMS condition on the
modular automorphism group of the state ψ, [Ta], (for t = i) is: ψ(xy) = ψ(σi(π(y))x) = ψ(σ(y)x)
for x, y ∈ π(Qλ), where σ(y) = ∆−1(y).
Lemma 3.3. The group of modular automorphisms of the von Neumann algebra π(Qλ)′′ is given on
the generators by
c := eAλ
(2)
σt(π(f · δg)) := ∆itπ(f · δg)∆−it = π(∆it(f · δg)) = gitπ(f · δg) = det(g)itπ(f · δg).
Proof. This is immediate from Lemma 2.8 if we note that g = det(g).
(cid:3)
26
Corollary 3.4. With Qλ acting on H := L2(Qλ) and with D the generator of the natural unitary
implementation of the gauge action of T1 on Qλ, we have ∆ = λD or eitD = ∆it/ log λ.
To simplify notation, we let A = Qλ and F = F λ = Aγ, the fixed point algebra for the T1 gauge
action, γ. For convenience we will suppress the notations D ⊗ 1k and so on. The algebras Ac, Fc are
defined as the finite linear span of the generators. Right multiplication makes A into a right F -module,
and similarly Ac is a right module over Fc. We define an F -valued inner product (··)R on both these
modules by (ab)R := Φ(a∗b).
Definition 3.5. Let X be the right F C∗-module obtained by completing A (or Ac) in the norm
kxk2
X := k(xx)RkF = kΦ(x∗x)kF .
The algebra A acting by left multiplication on X provides a representation of A as adjointable operators
on X. Let Xc be the copy of Ac ⊂ X. The T1 action on Xc is unitary and extends to X, [CNNR, PR].
For all k ∈ Z, the projection operator onto the k-th spectral subspace of the T1 action is also denoted
(somewhat carelessly) Φk on X:
Φk(x) =
z−kuz(x)dθ, z = eiθ, x ∈ X.
Observe that Φ0 restricts to Φ on A and on generators of Qλ we have
(3)
if g = λk
otherwise
1
2πZT1
Φk(f · δg) =(cid:26) f · δg
0
Of course L2(Qλ) and X have a common dense subspace Qλ
c on which these projections are identical.
Let Ak = Φk(A) and observe from (3) that A∗kAk = F = AkA∗k so that the gauge action γ on Qλ has
full spectral subspaces.
We quote the following result from [PR], the proof in our case is the same.
Lemma 3.6. The operators Φk are adjointable endomorphisms of the F -module X such that Φ∗k =
Φk = Φ2
k and ΦkΦl = δk,lΦk. If K ⊂ Z then the sumPk∈K Φk converges strictly to a projection in the
endomorphism algebra. The sum Pk∈Z Φk converges to the identity operator on X. For all x ∈ X,
the sum x =Pk∈Z Φkx =Pk∈Z xk converges in X.
The unbounded operator of the next proposition is of course the generator of the T1 action on X. We
refer to Lance's book, [L, Chapters 9,10], for information on unbounded operators on C∗-modules.
Proposition 3.7. [PR] Let X be the right C∗-F -module of Definition 3.5. Define D : XD ⊂ X to be
the linear space
XD = {x =Xk∈Z
xk ∈ X : kXk∈Z
k2(xkxk)Rk < ∞}.
For x ∈ XD define D(x) =Pk∈Z kxk. Then D : XD → X is a is self-adjoint, regular operator on X.
This should be compared to the following Hilbert space version.
Proposition 3.8. The generator D of the one-parameter unitary group {uz z ∈ T1} on L2(Qλ, ψ)
has eigenspaces given by the ranges of the Φk and D(x) = kx iff Φk(x) = x. In particular
dom(D) = {x =Xk
xk Φk(xk) = xk and Xk
k2kxkk2 < ∞},
and D(Pk xk) =Pk kxk.
27
X = kΦ(x∗x)kF λ. Introduce the rank one operators on X : ΘR
Remark. On generators in Qλ regarded as elements of either X or L2(Qλ, ψ) we have D(f · δg) =
(logλ(g))f · δg.
To continue, we recall the underlying right C∗-F λ-module, X, which is the completion of Qλ for the
norm kxk2
x,yz = x(yz)R. Then
using the operators Sk,m defined above, we obtain formulas for the projections Φk similar to those of
[PR, Lemma 4.7] with some important differences. First recall [CPR2, Lemma 3.5].
Lemma 3.9. Any F λ-linear endomorphism T of the module X which preserves the copy of Qλ inside
X, extends uniquely to a bounded operator on the Hilbert space H = L2(Qλ).
In particular, the finite rank endomorphisms of the pre-C∗ module Qλ
condition, and we denote the algebra of all these endomorphisms by End00
Lemma 3.10. Compare [PR, Lemma 4.7]. The following formulas hold in both L(X) and in B(H).
(1) For k ≥ 0, we have
c (acting on the left) satisfy this
F (Qλ
c ).
x,y by ΘR
Φ0 = ΘR
e,e while for k > 0, Φk =
ΘR
Sk,m,Sk,m.
mkXm=0
(2) For −k < 0, we have
k,m,S ∗
S ∗
Φ−k = ΘR
k,m
for any m = 0, 1, ..., mk − 1 and also for mk if λ−k = mk + 1.
Proof. Since both Φk and the finite rank endomorphisms satisfy the hypotheses of the previous lemma,
the first statement of this lemma will follow from calculations done on generators. The following
calculations are based on the formulas in Lemma 2.15.
(1) Let k > 0 and let x =Pl xl be a finite sum of generators, xl satisfying Φl(xl) = xl. Then
mkXm=0
Sk,m,Sk,m(xl) =Xl
Sk,mΦ(S∗k,mxl) =
mkXm=0
mkXm=0
Sk,mΦ(S∗k,mxk)
ΘR
ΘR
Sk,mS∗k,mxk = exk = xk = Φk(x).
mkXm=0
Sk,m,Sk,m(x) = Xl
mkXm=0
=
For k = 0 this is a similar but far easier calculation.
(2) Let −k < 0 and let x =Pl xl be a finite sum of generators as above. Then, for 0 ≤ m < mk
S∗k,mΦ(Sk,mxl) = S∗k,mΦ(Sk,mx−k)
ΘR
k,m,S ∗
S ∗
ΘR
k,m,S ∗
S ∗
k,m
k,m
(x) = Xl
(xl) =Xl
= S∗k,mSk,mx−k = ex−k = x−k = Φ−k(x).
(cid:3)
We recall the following result discussed in Section 3 of [CNNR] (a 'bare hands' proof can be given by
the method in [CPR2]).
c ))′′, where we take the commu-
Proposition 3.11. Let N be the von Neumann algebra N = (End00
tant inside B(H). Then N is semifinite, and there exists a faithful, semifinite, normal trace τ : N → C
such that for all rank one endomorphisms ΘR
F (Qλ
x,y of Qλ
c ,
x,y) = (τ ◦ Φ)(y∗x),
τ (ΘR
x, y ∈ Qλ
c .
In addition, D is affiliated to N and π(Qλ) is a subalgebra of N .
28
The fact that τ (Φk) = λ−k implies that with respect to the trace τ we can not expect D to satisfy a
finite summability criterion. We solve this problem exactly as in [CPR2].
Definition 3.12. We define a new weight on N +: let T ∈ N + then τ∆(T ) := supN τ (∆N T ) where
∆N = ∆(Pk≤N Φk).
Remarks. Since ∆N is τ -trace-class, we see that T 7→ τ (∆N T ) is a normal positive linear functional
on N and hence τ∆ is a normal weight on N + which is easily seen to be faithful and semifinite.
As in [CPR2], we now give another way to define τ∆ which is not only conceptually useful but also
makes a number of important properties straightforward to verify. Many proofs require only trivial
notation changes and the substitution of n± with λ∓.
Notation. Let M be the relative commutant in N of the operator ∆. Equivalently, M is the relative
commutant of the set of spectral projections {Φkk ∈ Z} of D. Clearly, M =Pk∈Z ΦkN Φk.
Definition 3.13. As τ restricted to each ΦkN Φk is a faithful finite trace with τ (Φk) = λ−k we define
bτk on ΦkN Φk to be λk times the restriction of τ . Then,bτ :=Pkbτk on M =Pk∈Z ΦkN Φk is a faithful
normal semifinite trace bτ with bτ (Φk) = 1 for all k.
We usebτ to give an alternative expression for τ∆ below
M. The map Ψ : N → M defined by Ψ(T ) =Pk ΦkT Φk is a conditional expectation onto M and
τ∆(T ) =bτ (Ψ(T )) for all T ∈ N +. That is, τ∆ =bτ ◦ Ψ so that bτ (T ) = τ∆(T ) for all T ∈ M+. Finally,
if one of A, B ∈ M is bτ -trace-class and T ∈ N then τ∆(AT B) = τ∆(AΨ(T )B) =bτ (AΨ(T )B).
Lemma 3.14. An element m ∈ N is in M if and only if it is in the fixed point algebra of the action,
στ∆
t (T ) = ∆itT ∆−it. Both π(F λ) and the projections Φk belong to
t
Proof. The proof is the same as the proof of [CPR2, Lemma 3.9] with λk in place of n−k.
Lemma 3.15. The modular automorphism group στ∆
The weight τ∆ is a KMS weight for the group στ∆
t
t of τ∆ is inner and given by στ∆
, and στ∆
on N defined for T ∈ N by στ∆
t (T ) = ∆itT ∆−it.
(cid:3)
t
Qλ = στ◦Φ
t
.
Proof. This follows from: [KR, Thm 9.2.38], which gives us the KMS properties of τ∆: the modular
group is inner since ∆ is affiliated to N . The final statement about the restriction of the modular
group to Qλ is clear.
We now have the key lemma:
Lemma 3.16. Suppose g is a function on R such that g(D) is τ∆ trace-class in M, then for all f ∈ F λ
we have
(cid:3)
τ∆(π(f )g(D)) = τ∆(g(D))τ (f ) = τ (f )Xk∈Z
g(k).
Proof. First note that τ∆(g(D)) =bτ (Pk∈Z g(k)Φk) =Pk∈Z g(k)bτ (Φk) =Pk∈Z g(k). We first do the
c so that all the sums are finite. Now,
computation for f ∈ F λ
g(k)bτ (π(f )Φk)
g(k)λk τ (π(f )Φk).
τ∆(π(f )g(D)) =bτ (π(f )Xk∈Z
=Xk∈Z
g(k)bτk(π(f )Φk) =Xk∈Z
g(k)Φk) =Xk∈Z
So it suffices to see for each k ∈ Z, we have τ (π(f )Φk) = λ−kτ (f ).
Now, by Theorem 2.35 π(F λ)′′ is a type II1 factor on H whose unique trace say T r (with norm
one) extends the trace τ on F λ in the sense that T r(π(f )) = τ (f ). Since the projection Φk is in the
commutant of the factor π(F λ)′′ the map
T ∈ π(F λ)′′ 7→ T Φk = ΦkT Φk
is a normal isomorphism by [D, Chapter 1, section 2, Prop. 2] and so it has a unique normalised trace
also given by T race(T Φk) = T r(T ). But τ (T Φk) is a trace on Φkπ(F λ)′′Φk = π(F λ)′′Φk and so must
be τ (Φk) = λ−k times the unique norm one trace. That is, we get the required formula:
τ (π(f )Φk) = λ−kT race(π(f )Φk) = λ−kT r(π(f )) = λ−kτ (f ).
29
So for f ∈ F λ
c , we have the formula:
τ∆(π(f )g(D)) = τ∆(g(D))τ (f ) =Xk∈Z
g(k)τ (f ).
Now, the right hand side is a norm-continuous function of f . To see that the left side is norm-continuous
we do it in more generality. Let T ∈ N , then sincebτ is a trace on M we get:
τ∆(T g(D)) = bτ (Ψ(T g(D)) = bτ (Ψ(T )g(D)) ≤ kΨ(T )kbτ (g(D)) ≤ kTkbτ ((g(D)) = kTkτ∆(g(D)).
That is the left hand side is norm-continuous in T and so we have the formula:
τ∆(π(f )g(D)) = τ∆(g(D))τ (f ) =Xk∈Z
g(k)τ (f )
1
2
for all f ∈ F λ.
Proposition 3.17. (i) We have (1 + D2)−1/2 ∈ L(1,∞)(M, τ∆). That is, τ∆((1 + D2)−s/2) < ∞ for
all s > 1. Moreover, for all f ∈ F λ
lim
s→1+
(s − 1)τ∆(π(f )(1 + D2)−s/2) = 2τ (f )
so that π(f )(1 + D2)−1/2 is a measurable operator in the sense of [C].
(ii) For π(a) ∈ π(Qλ) ⊂ N the following (ordinary) limit exists and
(cid:3)
bτω(π(a)) =
(s − 1)τ∆(π(a)(1 + D2)−s/2) = τ ◦ Φ(a),
lim
s→1+
the original KMS state ψ = τ ◦ Φ on Qλ.
Proof. (i) This proof is identical to [CPR2, Proposition 3.12].
(ii) This proof is the same as [CPR2, Proposition 3.14] with Qλ, F λ replacing On, F.
Definition 3.18. The triple (A,H,D) along with γ, ψ, N , τ∆ satisfying properties (0) to (3) below
is called the modular spectral triple of the dynamical system (Qλ, γ, ψ)
c of the algebra Qλ is faithfully represented in N with the latter acting on
0) The ∗-subalgebra A = Qλ
the Hilbert space H = L2(Qλ, ψ),
1) there is a faithful normal semifinite weight τ∆ on N such that the modular automorphism group of
τ∆ is an inner automorphism group σt (for t ∈ C) of (the Tomita algebra of ) N with σiA = σ in the
sense that σi(π(a)) = π(σ(a)), where σ is the automorphism σ(a) = ∆−1(a) on A,
2) τ∆ restricts to a faithful semifinite trace bτ on M = N σ, with a faithful normal projection Ψ : N →
M satisfying τ∆ =bτ ◦ Ψ on N ,
3) with D the generator of the one parameter group implementing the gauge action of T on H we
[D, π(a)] extends to a bounded operator (in N ) for all a ∈ A and for λ in the resolvent set of
have:
D, (λ − D)−1 ∈ K(M, τ∆), where K(M, τ∆) is the ideal of compact operators in M relative to τ∆. In
particular, D is affiliated to M.
(cid:3)
30
For matrix algebras A = Qλ
triple in the obvious fashion.
c ⊗ Mk over Qλ
c , (Qλ
c ⊗ Mk,H ⊗ Mk,D ⊗ Idk) is also a modular spectral
We need some technical lemmas for the discussion in the next Section. A function f from a complex
domain Ω into a Banach space X is called holomorphic if it is complex differentiable in norm on Ω.
The following is proved in [CPR2, Lemma 3.15].
Lemma 3.19. (1) Let B be a C∗-algebra and let T ∈ B+. The mapping z 7→ T z is holomorphic (in
operator norm) in the half-plane Re(z) > 0.
(2) Let B be a von Neumann algebra with faithful normal semifinite trace φ and let T ∈ B+ be in
L(1,∞)(B, φ). Then, the mapping z 7→ T z is holomorphic (in trace norm) in the half-plane Re(z) > 1.
(3) Let B, and T be as in item (2) and let A ∈ B then the mapping z 7→ φ(AT z) is holomorphic for
Re(z) > 1.
Lemma 3.20. In these modular spectral triples (A,H,D) for matrices over the algebras Qλ we have
(1+D2)−s/2 ∈ L1(M, τ∆) for all s > 1 and for x ∈ N , τ∆(x(1+D2)−r/2) is holomorphic for Re(r) > 1
and we have for a ∈ Qλ
Proof. We include a brief proof since there are some small but important differences from [CPR2,
Lemma 3.16]. Since the eigenvalues for D are precisely the set of integers, and the projection Φk on
the eigenspace with eigenvalue k satisfies τ∆(Φk) = 1, it is clear that (1 + D2)−s/2 ∈ L1(M, τ∆). Now,
lemma.
To see the last statement, we observe that τ∆([D, π(a)](1 + D2)−r/2) = τ∆(Ψ([D, π(a)])(1 + D2)−r/2),
so it suffices to see that Ψ([D, π(a)]) = 0 for a ∈ A = Qλ
c . To this end, let a = f · δg where det(g) = λn
is one of the linear generators of Qλ
c . Then by calculating the action of the operator Dπ(f · δg) on the
linear generators fi · δhi of the Hilbert space, H, we obtain:
τ∆(x(1 + D2)−r/2) = bτ (Ψ(x)(1 + D2)−r/2) is holomorphic for Re(r) > 1 by item (3) of the previous
c , τ∆([D, π(a)](1 + D2)−r/2) = 0, for Re(r) > 1.
Dπ(f · δg) = nπ(f · δg) + π(f · δg)D that is
[D, π(f · δg)] = logλ(g)π(f · δg).
More generally,
[D, π(
cifi · δhi)] =
ci(logλ(hi))π(fi · δhi).
mXi=1
mXi=1
c ) for all a ∈ Qλ
If we apply Ψ to this equation, we see that Ψ(π(fi · δhi)) = π(Φ(fi · δhi)) = 0 whenever logλ(hi) 6= 0,
and so the whole sum is 0. We also observe that [D, π(a)] ∈ π(Qλ
c . This is not too
surprising since D is the generator of the action γ of T on Qλ.
(cid:3)
3.3. Modular K1. We now make appropriate modifications to [CPR2, Section 4]) using [CNNR]
introducing elements of these modular spectral triples (A,H,D) (where A is a matrix algebra over
Qλ
we say that a unitary (invertible, projection,...) in the n × n matrices over Qλ for some n is a unitary
(invertible, projection,...) over Qλ. We write σt for the automorphism σt ⊗ Idn of A.
Definition 3.21. Let v be a partial isometry in the ∗-algebra A. We say that v satisfies the modular
condition with respect to σ if the operators vσt(v∗) and v∗σt(v) are in the fixed point algebra F ⊂ A
for all t ∈ R. Of course, any partial isometry in F is a modular partial isometry.
Lemma 3.22. ([CPR2, Lemma 4.8]) Let v ∈ A be a modular partial isometry. Then we have
c ) that will have a well defined pairing with our Dixmier functionalbτω. Let A = Qλ. Following [HR]
is a modular unitary over A. Moreover there is a modular homotopy uv ∼ uv∗ .
uv =(cid:18) 1 − v∗v
v
v∗
1 − vv∗ (cid:19)
Note that in [CPR2] we used a different approach which is implied by the one given here. In [CPR2]
we defined modular unitaries in terms of the regular automorphism:
π(σ(a)) = π(∆−1(a)) = ∆−1π(a)∆ = σi(π(a)).
31
That is we said that a unitary in A is modular if uσ(u∗) and u∗σ(u) are in the fixed point algebra.
Examples.
(1) For k, j > 0 recall Sk,m ∈ Qλ
X[mλk,(m+1)λk ) · δ1 which is in clearly F λ. Then for each {k, m},{j, n} we have a unitary
c with m < mk (see Definition 2.14) we write Pk,m = Sk,mS∗k,m =
u{k,m},{j,n} =(cid:18) 1 − Pk,m Sk,mS∗j,n
Sj,nS∗k,m 1 − Pj,n (cid:19) .
It is simple to check that this a self-adjoint unitary satisfying the modular condition, and that
τ (Pk,m) = λk and τ (Pj,n) = λj. These examples behave very much like the SµS∗ν examples of [CPR2].
(2) For k, j > 0 consider the "leftover" partial isometries Sk,mk and Sj,mj of Definition 3.13 which we
will denote by Sk and Sj to lighten the notation. We let vj,k = SjS∗k and calculate its range and initial
projections which are both in F λ:
Pj = SjS∗kSkS∗j = X[mj λj ,mjλj +λj(λ−k−mk)) · δ1, and
Pk = SkS∗j SjS∗k = X[mkλk,mkλk+λk(λ−j−mj )) · δ1.
We note for future reference that:
We also note that we have a modular unitary uj,k:
τ (Pj ) = λj(λ−k − mk) and τ (Pk) = λk(λ−j − mj).
uj,k =(cid:18) 1 − Pk
SjS∗k
SkS∗j
1 − Pj (cid:19) .
Define the modular K1 group as follows.
Definition 3.23. Let K1(A, σ) be the abelian group with one generator [v] for each partial isometry
v over A satisfying the modular condition and with the following relations:
1)
2)
3)
[v] = 0 if v is over F,
[v] + [w] = [v ⊕ w],
if vt, t ∈ [0, 1], is a continuous path of modular partial isometries in some matrix algebra over A
then [v0] = [v1].
One could use modular unitaries as in [CPR2] in place of these modular partial isometries.
The following can now be seen to hold.
Lemma 3.24. (Compare [CPR2, Lemma 4.9]) Let (A,H,D) be our modular spectral triple relative
to (N , τ∆) and set F = Aσ and σ : A → A. Let L∞(∆) = L∞(D) be the von Neumann algebra
generated by the spectral projections of ∆ then L∞(∆) ⊂ Z(M). Let v ∈ A be a partial isometry with
vv∗, v∗v ∈ F . Then π(v)Qπ(v∗) ∈ M and π(v∗)Qπ(v) ∈ M for all spectral projections Q of D, if and
only if v is modular. That is, π(v)∆π(v∗) and π(v∗)∆π(v) (or π(v)Dπ(v∗) and π(v∗)Dπ(v)) are both
affiliated to M if and only if v is modular.
32
Thus we see that modular partial isometries conjugate ∆ to an operator affiliated to M, and so v∆v∗
commutes with ∆ (and vDv∗ commutes with D).
We will next show that there is an analytic pairing between (part of) modular K1 and modular spectral
triples. To do this, we are going to use the analytic formulae for spectral flow in [CP2].
3.4. The mapping cone algebra. Our aim in the remainder is to calculate an index pairing explicitly
c of Qλ. In the following few pages we will
for the matrix algebras A over the smooth subalgebra Qλ
sometimes abuse notation and write a in place of π(a) for a ∈ A in order to make our formulae more
readable. Whenever we do this, however, we will use σi(·) = ∆−1(·)∆ the spatial version of the algebra
homomorphism, σ. We will generally use the spatial version σi when in the presence of operators not
in π(A).
We briefly review some results from [CNNR], that provide an interpretation of the modular index
pairing given by the spectral flow.
If F ⊂ A is a sub-C∗-algebra of the C∗-algebra A, then the mapping cone algebra for the inclusion is
M (F, A) = {f : R+ = [0,∞) → A : f is continuous and vanishes at infinity, f (0) ∈ F}.
When F is an ideal in A it is known that K0(M (F, A)) ∼= K0(A/F ), [Put1]. In general, K0(M (F, A))
is the set of homotopy classes of partial isometries v ∈ Mk(A) with range and source projections
vv∗, v∗v in Mk(F ), with operation the direct sum and inverse −[v] = [v∗]. All this is proved in [Put1].
It is shown in [CNNR] that there is a natural map that injects K1(A, σ) into K T
0 (M, F ), the equivariant
K-theory of the mapping cone algebra. Note that the T action on A lifts in the obvious way to the
mapping cone. Now, it was shown in [CPR1] that certain Kasparov A, F -modules extend to Kasparov
M (F, A), F -modules, and this was extended to the equivariant case in [CNNR]. Importantly the theory
applies to the equivariant Kasparov module coming from a circle action. The extension is explicit,
namely there is a pair ( X, D) which is a graded unbounded Kasparov module for the mapping cone
algebra M (F, A) constructed using a generalised APS construction, [APS3].
If v is a partial isometry in Mk(A), setting
ev(t) = 1 − vv∗
iv∗
1+t2
t
1+t2 −iv t
1+t2 ! ,
1+t2
v∗v
defines ev as a projection over M (F, A).
Then in [CNNR] we showed that if v ∈ A is a modular partial isometry we have
h[ev] −(cid:20)(cid:18) 1 0
0 0 (cid:19)(cid:21) , [( X, D)]i = −Index(P vP : v∗vP (X) → vv∗P (X)) ∈ K0(F )
= Index(P v∗P : vv∗P (X) → v∗vP (X)) ∈ K T
0 (F ).
(4)
We thus obtain an index map K1(A, σ) → K T
0 (F ). The latter may be thought of as the ring of
Laurent polynomials K0(F )(χ, χ−1) where we think of χ, χ−1 as generating the representation ring of
T. We may obtain a real valued invariant from this map by evaluating χ at e−β where β is the inverse
temperature of our KMS state and applying the trace to the resultant element of K0(F ). Then one of
the main results of [CNNR] is that the real valued invariant so obtained is identical with the spectral
flow invariant of the next subsection. However the general theory of [CNNR] does not tell us the range
of this index map and it is the latter that is of interest for these explicit calculations.
33
3.5. A local index formula for the algebras Qλ. Using the fact that we have full spectral subspaces
we know from [CNNR] that there is a formula for spectral flow which is analogous to the local index
formula in noncommutative geometry. We remind the reader that τ∆ =bτ ◦ Ψ where Ψ : N → M is
the canonical expectation, so that τ∆ restricted to M isbτ .
Theorem 3.25. (Compare [CPR2, Theorem 5.5]) Let (A,H,D) be the (1,∞)-summable, modular
spectral triple for the algebra Qλ we have constructed previously. Then for any modular partial isometry
v and for any Dixmier trace bτω associated to bτ , we have spectral flow as an actual limit
sfbτ (vv∗D, vDv∗) =
The functional on A⊗A defined by a0⊗ a1 7→ 1
b, B-cocycle (see the proof below for the definition).
2bτω(v[D, v∗](1 +D2)−1/2) = τ ◦ Φ(v[D, v∗]).
(s− 1)bτ (v[D, v∗](1 +D2)−s/2) =
2 lims→1+(s− 1)τ∆(a0[D, a1](1 +D2)−s/2) is a σ-twisted
lim
s→1+
1
2
1
Remark. Spectral flow in this setting is independent of the path joining the endpoints of unbounded
self adjoint operators affiliated to M however it is not obvious that this is enough to show that it is
constant on homotopy classes of modular unitaries. This latter fact is true but the proof is lengthy so
we refer to [CNNR].
Theorem 3.26. We let (Qλ
(1) Let u be a modular unitary defined in Section 5 of the form
c ⊗ M2).
c ⊗ M2,H ⊗ C2,D ⊗ 12) be the modular spectral triple of (Qλ
u{k,m},{j,n} =(cid:18) 1 − Pk,m Sk,mS∗j,n
Sj,nS∗k,m 1 − Pj,n (cid:19) .
Then the spectral flow is positive being given by
(2) Let u be a modular unitary defined in Section 5 of the form:
sfτ∆(D, uDu∗) = (k − j)(λj − λk) ∈ Z[λ] ⊂ Γλ.
uj,k =(cid:18) 1 − Pk
SjS∗k
SkS∗j
1 − Pj (cid:19) ,
where SkS∗j = Sk,mkS∗j,mj
spectral flow is given by
and Pk and Pj are its range and initial projections, respectively. Then the
sfτ∆(D, uDu∗) = (k − j)[λj(λ−k − mk) − λk(λ−j − mj)] ∈ Γλ.
Proof. We have already observed that these are, in fact modular unitaries. For the computations we
use a calculation from the proof of Lemma 3.20 to get in example (1):
u[D ⊗ 12, u] =(cid:18) 1 − Pk,m Sk,mS∗j,n
= (cid:18) 1 − Pk,m Sk,mS∗j,n
Sj,nS∗k,m 1 − Pj,n (cid:19)(cid:18)
So using Theorem 3.25 and our previous computation of the Dixmier trace, Proposition 3.17, and the
fact that Pk,m = Sk,mS∗k,m = X[mλk,(m+1)λk) · δ1 so that τ (Pk,m) = λk we have
Sj,nS∗k,m 1 − Pj,n (cid:19)(cid:18)
(j − k)Sj,nS∗k,m
(cid:19) = (k − j)(cid:18) −Pk,m
(k − j)Sk,mS∗j,n
Pj,n (cid:19) .
[D, Sk,mS∗j,n]
[D, Sj,nS∗k,m]
(cid:19)
0
0
0
0
0
0
sfτ∆(D, uk,mDuk,m) = (k − j)τ (Pj,n − Pk,m) = (k − j)(λj − λk).
This number is always positive as the reader may check, and is contained in Z[λ].
The computations in example (2) are similar and use the fact that Pk = X[mkλk,mkλk+λk(λ−j−mj )) · δ1,
so that τ (Pk) = λk(λ−j − mj) ∈ Γλ. In these examples, the spectral flow is not contained in the
smaller polynomial ring, Z[λ].
(cid:3)
34
Remarks. The observation of [CPR2] that the twisted residue cocycle formula for spectral flow
is calculating Araki's relative entropy of two KMS states [Ar] also applies to the examples in this
subsection.
Acknowledgements We would like to thank Nigel Higson, Ryszard Nest, Sergey Neshveyev, Marcelo
Laca, Iain Raeburn and Peter Dukes for advice and comments. The first and fourth named authors
were supported by the Australian Research Council. The second and third named authors acknowledge
the support of NSERC (Canada).
References
[Ar]
H. Araki, Relative entropy of states of von Neumann algebras, Publ. RIMS, Kyoto Univ., 11 (1976) 809 -- 833
and Relative entropy for states of von Neumann algebras II, Publ. RIMS, Kyoto Univ., 13 (1977) 173 -- 192.
[APS3] M.F. Atiyah, V.K. Patodi, I.M. Singer, Spectral asymmetry and Riemannian geometry. III, Math. Proc. Camb.
[BR1]
[BR2]
Phil. Soc. 79 (1976) 71 -- 99.
O. Bratteli, D. Robinson, Operator algebras and quantum statistical mechanics 1, Springer-Verlag, 2nd Ed,
1987.
O. Bratteli, D. Robinson, Operator algebras and quantum statistical mechanics 2, Springer-Verlag, 2nd Ed,
1987.
[CNNR] A.L. Carey, R. Nest, S. Neshveyev, A. Rennie, Twisted cyclic theory, equivariant KK-theory and KMS states,
[CP1]
[CP2]
to appear in J. reine angew. Math.
A. L. Carey, J. Phillips, Unbounded Fredholm modules and spectral flow, Canadian J. Math., 50 (4) (1998)
673 -- 718.
A. L. Carey, J. Phillips, Spectral flow in θ-summable Fredholm modules, eta invariants and the JLO cocycle,
K-Theory, 31 (2004) 135 -- 194.
[CPR1] A.L. Carey, J. Phillips, A. Rennie, A noncommutative Atiyah-Patodi-Singer index theorem in KK-theory, to
appear in J. reine angew. Math.
[CPR2] A.L. Carey, J. Phillips, A. Rennie, Twisted cyclic theory and an index theory for the gauge invariant KMS
state on Cuntz algebras, to appear in Journal of K-Theory.
A.L. Carey, J. Phillips, F. Sukochev, Spectral flow and Dixmier traces, Adv. Math, 173 (2003) 68 -- 113.
[CPS2]
[CPRS2] A.L. Carey, J. Phillips, A. Rennie, F. Sukochev, The local index formula in semifinite von Neumann algebras
I: Spectral Flow, Adv. in Math. 202 (2006) 451 -- 516.
[C0]
[C]
[Cu]
[Cu1]
[D]
[E]
[K]
[CRSS] A. L. Carey, A. Rennie, A. Sedaev, F. Sukochev, Dixmier traces and asymptotics of zeta functions, somewhere
A.L. Carey, A. Rennie, K. Tong, Spectral flow invariants and twisted cyclic theory from the Haar state on
[CRT]
SUq(2), to appear
A. Connes, Une classification des facteurs de type III, Annales Scientifiques de le'Ecole Norm. Sup., 4em serie
t. 6 (1973) 18 -- 252.
A. Connes, Noncommutative geometry, Academic Press, 1994.
J. Cuntz, Simple C ∗-algebras generated by isometries, Commun. Math. Phys, 57 (1977) 173 -- 189.
J. Cuntz, C ∗-algebras associated with the ax + b-semigroup over N inK-Theory and Noncommutative Geometry,
Eds, G. Cortinas, J. Cuntz, M Karoubi, R. Nest, C.A. Weibel, EMS Series of Congress Reports, Volume 2.
J. Dixmier, von Neumann algebras, North-Holland, 1981.
G. Elliott, Some simple C ∗-algebras constructed as crossed products with discrete outer automorphism groups,
Publ. RIMS Kyoto Univ., 16 (1980) 299-311.
T. Fack and H. Kosaki, Generalised s-numbers of τ -measurable operators, Pac. J. Math., 123 (1986) 269 -- 300.
N. Higson, J. Roe, Analytic K-homology, Oxford University Press, 2000.
J. Kaad, R. Nest, A. Rennie, KK-theory and spectral flow in von Neumann algebras, arXive:math.OA/0701326.
R.V. Kadison, J. R. Ringrose, Fundamentals of the theory of operator algebras, Vol. II: advanced theory,
Academic Press, 1986.
G. G. Kasparov, The operator K-functor and extensions of C ∗-algebras, Math. USSR. Izv. 16 No. 3 (1981)
513 -- 572.
J. Kustermans, G. Murphy, L. Tuset, Differential calculi over quantum groups and twisted cyclic cocycles, J.
Geom. Phys., 44 (2003) 570 -- 594.
E. C. Lance, Hilbert C ∗-modules, Cambridge University Press, Cambridge, 1995.
M. Laca, J. Spielberg, Purely infinite C ∗-algebras from boundary actions of discrete groups, Journal fur die
reine und ang. Mathematik, 480 (1996) 125 -- 139.
[FK]
[HR]
[KNR]
[KR]
[L]
[LS]
[KMT]
35
[LSS]
[PR]
[PhR]
[Ped]
[PT]
[Put1]
[Put2]
[RS]
[Sc]
[Ta]
[T]
S. Lord, A. Sedaev, F. A. Sukochev, Dixmier traces as singular symmetric functionals and applications to
measurable operators, Journal of Functional Analysis, 224 no.1 (2005) 72 -- 106.
D. Pask, A. Rennie, The noncommutative geometry of graph C ∗-algebras I: The index theorem, Journal of
Functional Analysis, 233 (2006) 92 -- 134.
J. Phillips, I. Raeburn, Semigroups of isometries, Toeplitz algebras and twisted crossed products, J. Int. Equat.
and Op. Th., 17 (1993) 579-602.
G. K. Pedersen, C ∗-algebras and their automorphism groups, London Math. Soc. monographs 14, Academic
Press, London 1979.
G. K. Pedersen, M. Takesaki, The Radon-Nikodym theorem for von Neumann algebras, Acta Math., 130 (1973)
53 -- 87.
I. Putnam, An excision theorem for the K-theory of C ∗-algebras, J. Operator Theory, 38 (1997) 151 -- 171.
I. Putnam, On the K-theory of C ∗-algebras of principal groupoids, Rocky Mountain J. Math., 28 no. 4 (1998)
1483 -- 1518.
M. Rørdam and E. Størmer, Classification of nuclear C ∗-algebras. Entropy in operator algebras, Encyclopedia
of Mathematical Sciences, 126 (2002), Springer, Berlin.
C. Schochet, Topological methods for C ∗-algebras II: geometric resolutions and the Kunneth formula, Pac. J.
Math., 98 No. 2 (1982) 443 -- 458.
M. Takesaki, Tomita's theory of modular Hilbert algebras and its applications, Lecture Notes in Mathematics,
128 (1970), Springer, Berlin.
J. Tomiyama, On the projection of norm one in W ∗-algebras, Proc. Japan Acad., 33 (1957) 608 -- 612.
E-mail
[email protected]
address:
[email protected], [email protected], [email protected],
|
1708.07496 | 3 | 1708 | 2019-02-25T13:48:59 | Non-classification of free Araki-Woods factors and $\tau$-invariants | [
"math.OA",
"math.LO"
] | We define the standard Borel space of free Araki-Woods factors and prove that their isomorphism relation is not classifiable by countable structures. We also prove that equality of $\tau$-topologies, arising as invariants of type III factors, as well as coycle and outer conjugacy of actions of abelian groups on free product factors are not classifiable by countable structures. | math.OA | math |
Non-classification of free Araki-Woods factors and τ -invariants
by Rom´an Sasyk1, Asger Tornquist2 and Stefaan Vaes3
Abstract
We define the standard Borel space of free Araki-Woods factors and prove that their iso-
morphism relation is not classifiable by countable structures. We also prove that equality
of τ -topologies, arising as invariants of type III factors, as well as cocycle and outer conju-
gacy of actions of abelian groups on free product factors are not classifiable by countable
structures.
1
Introduction
In his celebrated work, Connes [Co75] proved that all amenable factors acting on a separable
Hilbert space are hyperfinite and deduced that there is a unique amenable factor for each of
the types In with n ∈ N, II1, II∞ and IIIλ with λ ∈ (0, 1). The uniqueness of the hyperfinite
III1 factor was proved in Haagerup's fundamental paper [Ha85].
In [Co75, Co74b], Connes
also proved that hyperfinite type III0 factors are isomorphic to Krieger factors. So, using
[Kr74, CT76], they are completely classified by their flow of weights, which is a properly ergodic
action of R on a standard measure space. This last result can equally well be interpreted as a
non-classification theorem. Indeed, properly ergodic flows up to conjugacy cannot be classified
in any concrete way, and descriptive set theory provides a framework to give a precise meaning
to the complexity of such a classification problem.
In particular, using Hjorth's concept of
turbulence [Hj97], it follows from [FW03] that properly ergodic flows and therefore amenable
type III0 factors cannot be classified by countable structures like countable groups or other
combinatorial invariants. Even for the subclass of infinite tensor products of matrix algebras,
such a non-classification result holds, see [ST09].
For nonamenable factors, complete classification theorems for large families of such factors
were obtained in Popa's deformation/rigidity theory [Po06]. In particular, his work allows to
"embed" unclassifiable structures (like, for instance, probability measure preserving transfor-
mations) into the theory of II1 factors and prove that II1 factors are not classifiable by countable
structures, see [ST08].
Among the most natural and most studied nonamenable factors are the free Araki-Woods
factors of Shlyakhtenko [Sh96], associated with an orthogonal representation (Ut)t∈R of R on
a real Hilbert space HR of dimension at least 2. This construction generalizes Voiculescu's
free Gaussian functor. When Ut is the trivial representation, the associated free Araki-Woods
factor is of type II1 and isomorphic to the free group factor L(Fn) with n = dimR(HR). When
Ut is non trivial, one obtains a full factor of type III that may be viewed as a free probability
analog of the classical Araki-Woods factors.
The almost periodic free Araki-Woods factors, associated with almost periodic representations
Ut, were completely classified by Shlyakhtenko in [Sh96] in terms of Connes' Sd-invariant, which
1Universidad de Buenos Aires, Departamento de Matem´atica and Instituto Argentino de Matem´aticas-
CONICET, Buenos Aires (Argentina). E-mail: [email protected]. Supported in part by research grant PICT
2012-1292 (ANPCyT).
2University of Copenhagen, Department of Mathematics, Copenhagen (Denmark). E-mail: [email protected].
Supported in part by the DNRF Niels Bohr Professorship of Lars Hesselholt, and by DFFResearch Project
7014-00145B.
3KU Leuven, Department of Mathematics, Leuven (Belgium). E-mail: [email protected]. Supported
by European Research Council Consolidator Grant 614195 RIGIDITY, and by long term structural funding --
Methusalem grant of the Flemish Government.
1
is a countable subgroup of R+
∗ . While the complete classification of arbitrary free Araki-Woods
factors remains wide open, a first result in that direction, classifying a large family of non
almost periodic free Araki-Woods factors, was obtained in [HSV16]. As the first main result
of our paper, we turn the space of free Araki-Woods factors into a standard Borel space and
reinterpret one of the examples in [HSV16] as providing a Borel family Mµ of free Araki-Woods
factors indexed by arbitrary non-atomic probability measures on the Cantor set C such that
Mµ ∼= Mη if and only if the measures µ and η are mutually absolutely continuous; see Theorem
2.2. In combination with [KS99], it follows that free Araki-Woods factors are not classifiable
by countable structures.
The finest modular theory invariant for full type III factors is Connes τ -invariant [Co74a],
defined as the weakest topology on R that makes the canonical modular homomorphism R →
Aut(M )/ Inn(M ) continuous.
In [Sh02], Shlyakhtenko proved that there is a continuum of
possible τ -topologies and used this to construct a continuum of non almost periodic free Araki-
Woods factors. As our second main result, we prove in Theorem 3.1 that τ -topologies are not
even classifiable by countable structures. This provides another proof for the non-classification
of free Araki-Woods factors, but also applies to other classification questions where a τ -invariant
makes sense. As an example, we prove in Theorem 4.2 that cocycle or outer conjugacy of actions
of Z, R, or any other non compact, abelian, locally compact group G on a free group factor
L(Fn), or any other free product factor M1 ∗ M2, are not classifiable by countable structures.
Finally note that a similar non-classification theorem for conjugacy of a single automorphism
(i.e. an action of Z) was obtained in [KLP08] for the hyperfinite II1 factor R and in [KLP14]
for free product factors, using a spectral invariant. Such spectral invariants are however not
preserved under cocycle conjugacy. Moreover, by [Oc85], all outer actions of a discrete amenable
group G on R are cocycle conjugate, and this leads to a complete classification of actions of G
on R up to cocycle conjugacy.
Acknowledgment. Research for this paper was mainly carried out when R. Sasyk was at
the Hausdorff Research Institute for Mathematics in Bonn during the Trimester program on
"von Neumann Algebras". He wishes to thank the Hausdorff Institute for kind hospitality and
support. S. Vaes wants to thank D. Shlyakhtenko for discussions and suggestions that helped
to improve this article.
2 Non-classification of free Araki-Woods factors
Whenever K is a separable Hilbert space, we denote by vN(K) the set of all von Neumann
subalgebras of B(K) and we equip vN(K) with the Effros Borel structure defined in [Ef64]
as the smallest σ-algebra such that for every ϕ ∈ B(K)∗, the map vN(K) → R : M 7→
kϕM k is measurable. We start by defining a Borel subset FAW ⊂ vN(K) that contains, up
to isomorphism, all free Araki-Woods factors. The correct interpretation of [HSV16, Example
5.4] then provides a Borel map Θ : Probc(C) → FAW from the space of continuous probability
measures on the Cantor set C to the space of free Araki-Woods factors such that Θ(µ) ∼= Θ(µ′)
if and only µ ≈ µ′, where ≈ denotes the equivalence relation of mutual absolute continuity
on Prob(K). So, Θ is a Borel reduction from (Probc(K), ≈) to (FAW, ∼=) and it follows from
[KS99, Theorem 2.1] that free Araki-Woods factors are not classifiable by countable structures.
To define the Borel set FAW, fix a separable, infinite dimensional Hilbert space H. Define the
full Fock space K = F(H) given by
F(H) = CΩ ⊕
∞Mn=1(cid:0)H ⊗ · · · ⊗ H(cid:1)
}
{z
n times
2
.
(2.1)
We denote by Ba(H) the space of antilinear bounded operators from H to H and we equip
both Ba(H) and B(H) with the standard Borel structure induced by the strong topology. We
define the following parameter space for FAW ⊂ vN(K).
D =(cid:8)(J, T ) ∈ Ba(H) × B(H)(cid:12)(cid:12)J = J ∗, J 2 = 1, T = T ∗, 0 ≤ T ≤ 1, T + JT J = 1,
T and 1 − T have trivial kernel(cid:9) .
Since the operators with trivial kernel form a Borel subset of B(H), it is easy to see that D is a
Borel subset of Ba(H) × B(H). The formula S = J(T −1 − 1)1/2 defines a bijection between the
set D and the set of all densely defined, closed, antilinear operators S on H satisfying S2 ⊂ 1,
with the inverse being determined by the polar decomposition of S. To every such involution
S is associated the free Araki-Woods factor acting on K given by
(2.2)
Φ(J, T ) =(cid:8)ℓ(ξ) + ℓ(S(ξ))∗(cid:12)(cid:12) ξ ∈ D(S)(cid:9)′′ =(cid:8)ℓ(T 1/2ξ) + ℓ(J(1 − T )1/2ξ)∗(cid:12)(cid:12) ξ ∈ H(cid:9) .
It follows that Φ : D → vN(K) is a Borel map. Since the graph of S can be recovered from
M = Φ(J, T ) as the closure of the set
(2.3)
it follows that Φ is injective.
(cid:8)(xΩ, x∗Ω)(cid:12)(cid:12) x ∈ M, xΩ ∈ H, x∗Ω ∈ H(cid:9) ,
Definition 2.1. For K = F(H), we define the Borel set FAW ⊂ vN(K) as the image of the set
D defined in (2.2) under the injective Borel map Φ defined in (2.3).
By construction, every free Araki-Woods factor is isomorphic with an element of FAW. We
can now deduce from [HSV16, Example 5.4] the following result.
Theorem 2.2. Let C be the Cantor set. There exists a Borel map Θ : Probc(C) → FAW such
that Θ(µ) ∼= Θ(µ′) if and only if µ ≈ µ′, where ≈ denotes the equivalence relation of mutual
absolute continuity. The map Θ can be chosen such that all Θ(µ) have as τ -invariant the usual
topology on R.
In particular, the equivalence relation (FAW, ∼=) is not classifiable by countable structures.
Proof. Denote by λ the Lebesgue measure on the interval [0, 1]. Fix the separable Hilbert space
H = L2([0, 1], λ) ⊕ L2([0, 1], λ) ⊕ C2 .
Denote by e0, e1 ∈ C2 the two standard basis vectors. Define the anti-unitary involutions
J0 : C2 → C2 and J : H → H given by
J0(e1) = e2 , J0(e2) = e1
and J(ξ1 ⊕ ξ2 ⊕ ξ3) = ξ2 ⊕ ξ1 ⊕ J0(ξ3) .
Fix any q ∈ (0, 1). Using the functions ϕµ : [0, 1] → [0, 1] introduced in Lemma 2.3 below, we
define for every probability measure µ ∈ Prob([0, 1]), the operator Tµ ∈ L∞([0, 1])⊕L∞([0, 1])⊕
M2(C) ⊂ B(H), given by
Tµ =(cid:0)1 + exp(ϕµ)(cid:1)−1 ⊕(cid:0)1 + exp(−ϕµ)(cid:1)−1 ⊕(cid:18)(1 + q)−1
0
0
(1 + q−1)−1(cid:19) .
By construction, (J, Tµ) ∈ D and by Lemma 2.3, the map µ 7→ (J, Tµ) is Borel. We define the
Borel map
Θ1 : Prob([0, 1]) → FAW : Θ1(µ) = Φ(J, Tµ) .
3
Denoting by δa the Dirac measure in a ∈ R, we associate to any nonatomic probability measure
µ on [0, 1] the following probability measures on R : the opposite measure µop given by µop(U ) =
eµ =
in [HSV16].
1
4(cid:0)µ + µop + δlog q + δ− log q(cid:1) .
µ(−U ) and the symmetric probability measure eµ given by
Araki-Woods factor Φ(J, Tµ) is isomorphic with the free Araki-Woods factor denoted by Γ(eµ, 1)
Using the unitary V : L2([0, 1], µ) → L2([0, 1], λ) given by Lemma 2.3, it follows that the free
By [Ru62, Theorems 5.1.4 and 5.2.2] (see also [Ke95, 19.1 and 19.2]), we can choose a copy of
the Cantor set C ⊂ [0, 1] such that C is independent as a subset of R. Viewing Probc(C) as a
subset of Probc([0, 1]), we define
Θ : Probc(C) → FAW : Θ(µ) = Θ1((µ + λ)/2) .
By [HSV16, Example 5.4], we get that for all µ, µ′ ∈ Probc(C), the von Neumann algebras
Θ(µ) and Θ(µ′) are isomorphic if and only if the measures µ and µ′ are mutually absolutely
continuous, and that the τ -invariant of every Θ(µ) equals the usual topology on R.
By [KS99, Theorem 2.1], the equivalence relation ≈ of mutual absolute continuity on Probc(C)
is not classifiable by countable structures. A fortiori, the equivalence relation ∼= on FAW is not
classifiable by countable structures.
The following lemma is probably well known. It is for instance essentially contained in [Ke95,
Proof of Theorem 17.41]. For completeness, we provide a short proof.
Lemma 2.3. For every probability measure µ ∈ Prob([0, 1]), define the increasing function
ϕµ : [0, 1] → [0, 1] : ϕµ(y) = min{x ∈ [0, 1] µ([0, x]) ≥ y} .
Equip Prob([0, 1]) with the weak topology and denote by λ the Lebesgue measure on [0, 1]. Then,
the following holds.
1. Each ϕµ is a Borel function.
2. For every µ ∈ Prob([0, 1]), we have that (ϕµ)∗(λ) = µ.
3. The map Prob([0, 1]) → L1([0, 1], λ) : µ 7→ ϕµ is continuous.
4. If µ has no atoms, the map ϕµ is strictly increasing and
V : L2([0, 1], µ) → L2([0, 1], λ) : (V ξ)(y) = ξ(ϕµ(y))
is unitary.
Proof. 1. Since ϕµ is increasing, it is a Borel function.
2. For every a ∈ [0, 1], we have that ϕ−1
coincide on [0, a] for all a ∈ [0, 1]. This implies that (ϕµ)∗(λ) = µ.
3. A direct computation gives that for all µ, η ∈ Prob([0, 1]), we have
µ ([0, a]) = [0, µ([0, a])]. Therefore, (ϕµ)∗(λ) and µ
kϕµ − ϕηk1 =Z 1
0 (cid:12)(cid:12)µ([0, a]) − η([0, a])(cid:12)(cid:12) da .
It follows that Prob([0, 1]) → L1([0, 1]) : µ 7→ ϕµ is continuous.
4. When µ has no atoms, the function x 7→ µ([0, x]) is continuous and thus attains all values
in [0, 1].
It follows that ϕµ is strictly increasing. Defining Rµ = ϕµ([0, 1]), we have that
µ([0, 1] \ Rµ) = 0 and ϕµ : [0, 1] → Rµ is a bijection sending λ to µ. It follows that V is a well
defined unitary operator.
4
3 Non-classification of τ -topologies
Recall that a factor M is called full when Inn(M ) ⊂ Aut(M ) is closed. Then, Out(M ) has
a natural Polish group structure.
In [Co74a, Definition 5.1], Connes defined the invariant
τ (M ) as the weakest topology on R that makes the (canonical) modular homomorphism R →
Out(M ) : t 7→ σϕ
t continuous. Given a unitary representation π : R → U (H) on a separable
Hilbert space, in [Co74a, Theorem 5.2], a full factor M is constructed such that τ (M ) equals
the weakest topology on R that makes π continuous.
Let G be a second countable, locally compact, abelian group. Since unitary representations of
G on a separable Hilbert space are classified by their spectral measure and multiplicity function,
it is natural to define as follows the τ -topology τ (µ) for any Borel probability measure µ on
and denote by τ (µ) the weakest topology on G that makes the map g 7→ πµ(g) continuous
the Pontryagin dual bG. Consider the unitary representation
πµ : G → U (L2(bG, µ)) : (πµ(g)ξ)(ω) = ω(g) ξ(ω)
for all g ∈ G, ξ ∈ L2(bG, µ), ω ∈ bG (3.1)
from G to the unitary group equipped with the strong topology. We denote by bµ the Fourier
bµ : G → C :bµ(g) =Z bG
transform of µ defined as
ω(g) dµ(ω) .
Given a Polish space K, we always equip Prob(K) with the weak topology, i.e. the weakest
Note that gn → e in the topology τ (µ) if and only if bµ(gn) → 1.
topology making the maps µ 7→ RK F dµ continuous for all bounded continuous functions
F ∈ Cb(K).
Finally, denote by S∞ the Polish group of all permutations of N. Recall that an equivalence
relation E on a Polish space X is said to be generically S∞-ergodic if for any continuous action
of S∞ on a Polish space Y , any Baire measurable morphism f : X → Y from E to the orbit
equivalence relation of S∞ y Y must map a comeager subset of X to a single S∞-orbit. If
moreover E has meager orbits, it follows that E is not classifiable by countable structures.
In [Sh02, Theorem 2.3], it was proven that there is a continuum of different τ -topologies on R.
We now prove that equality of τ -topologies is not even classifiable by countable structures.
Theorem 3.1. Let G be a second countable, locally compact, noncompact, abelian group. Let
K ⊂ bG be a nonempty closed subset for which there exists a probability measure µ1 ∈ Prob(bG)
with support K such that cµ1 tends to zero at infinity.
Then, the equivalence relation on Prob(K) defined by µ ∼ µ′ iff τ (µ) = τ (µ′) has meager equiv-
alence classes and is generically S∞-ergodic. In particular, ∼ is not classifiable by countable
structures.
Note that the theorem applies wheneverbλ(K) > 0, wherebλ is the Haar measure on bG, because it
then suffices to choose µ1 ∈ Prob(K) in the same measure class asbλK . So, for G as in Theorem
3.1, the equivalence relation on Prob(bG) given by equality of τ -topologies is not classifiable by
countable structures. In particular, it is not smooth, meaning that it is impossible to define a
standard Borel space of topologies on G in such a way that µ 7→ τ (µ) becomes a Borel map.
Proof. Denote by ≈ the equivalence relation on Prob(K) given by mutual absolute continuity.
Note that ≈ ⊂ ∼. Since K is the support of the probability measure µ1 whose Fourier transform
cµ1 tends to zero at infinity, it follows in particular that µ1 has no atoms, so that K has no
5
isolated points. By [KS99, Theorem 2], the equivalence relation ≈ is generically S∞-ergodic.
A fortiori, the equivalence relation ∼ is generically S∞-ergodic.
To prove that ∼ has meager equivalence classes, choose an atomic µ0 ∈ Prob(K) such that the
set of atoms of µ0 is dense in K.
(3.2)
is comeager in Prob(K).
Claim. If gn ∈ G is such that gn → ∞ andcµ0(gn) → 1 and if a ∈ [0, 1], then
(cid:8)µ ∈ Prob(K)(cid:12)(cid:12) there exists a subsequence gnk such that bµ(gnk ) → a(cid:9)
P (k, n0) where P (k, n0) = [n≥n0(cid:8)µ ∈ Prob(K)(cid:12)(cid:12) bµ(gn) − a < 1/k(cid:9) .
\k≥1 \n0≥1
Note that the set in (3.2) equals
Since P (k, n0) is a union of open sets, to prove the claim, it suffices to prove that P (k, n0) is
dense in Prob(K). For i ∈ {0, 1}, denote by Pi ⊂ Prob(K) the set of probability measures
that are absolutely continuous w.r.t. µi. Since both µ0 and µ1 have support K, it follows that
Pi ⊂ Prob(K) is dense for every i ∈ {0, 1}. Choose ηi ∈ Pi and define η = aη0 + (1 − a)η1.
Since cµ0(gn) → 1, we get that πµ0(gn) → 1 strongly and thus, bη0(gn) → 1. Since cµ1 tends to
zero at infinity, the representation πµ1 is mixing and it follows that bη1(gn) → 0. So,bη(gn) → a
and η ∈ P (k, n0). Thus, aP0 + (1 − a)P1 ⊂ P (k, n0). Since Pi ⊂ Prob(K) is dense for every
i ∈ {0, 1}, it follows that P (k, n0) ⊂ Prob(K) is dense as well.
Assume that the equivalence class [µ]∼ of some µ ∈ Prob(K) is not meager. Since the closure
of πµ0(G) is compact while G is noncompact, we can choose as follows a sequence gn ∈ G
such that gn → ∞ and πµ0(gn) → 1 strongly. Starting with any sequence hn ∈ G such that
hn → ∞, after passage to a subsequence, we may assume that πµ0(hn) converges strongly.
Taking a sequence kn ∈ N that tends to infinity sufficiently fast, the sequence gn = hknh−1
n has
the required properties.
the claim above to the sequence (gnk ), there exists η1 ∈ [µ]∼ and a subsequence gnkl
In particular,cµ0(gn) → 1. Since the nonmeager set [µ]∼ intersects every comeager set, by the
claim above, there exists η0 ∈ [µ]∼ and a subsequence gnk such that limk bη0(gnk ) = 1. Applying
liml bη1(gnkl
such that
6→ e in the topology
τ (η1). Since τ (η0) = τ (µ) = τ (η1), we have reached a contradiction. So we have proven that
the equivalence relation ∼ has meager equivalence classes.
→ e in the topology τ (η0), while gnkl
) = 0. It follows that gnkl
Remark 3.2. Let M be the free Araki-Woods factor associated with an orthogonal represen-
tation (Ut)t∈R of R on a real Hilbert space HR of dimension at least 2. In [Sh97, Corollary 8.6],
it is proven that M is full and that τ (M ) coincides with the weakest topology on R that makes
the map R → O(HR) : t 7→ Ut continuous (see [Va04, Th´eor`eme 2.7] for the fully general case).
So also from Theorem 3.1, one can deduce that free Araki-Woods factors cannot be classified
by countable structures.
In the case where G = R, the non-classification of τ -topologies in Theorem 3.1 can be made
concrete as follows. Define the compact space K = {−1, 0, 1}N and the continuous map
θ : K → R : θ(x) =
xn2−n .
∞Xn=0
Consider the Polish space X = (0, 1/4)N and define for every a ∈ X, the probability measure
νa on K given by
νa =
∞Yn=1(cid:16)(1 − an)δ0 +
an
2
δ−1 +
an
2
δ1(cid:17) .
6
Note that a 7→ νa is a continuous map from X to Prob(K) equipped with the weak topology.
Pushing forward with θ, we define the probability measure µa on R given by µa = θ∗(νa). The
following proposition describes the τ -topology of the measure µa.
Proposition 3.3. For every a ∈ X, the topology τ (µa) is induced by the translation invariant
metric da on R defined by
da(t, s) = ∞Xn=0
an dZ(cid:0)2−n(t − s)(cid:1)2!1/2
,
(3.3)
where dZ denotes the distance of a real number to Z ⊂ R.
Proof. Fix a ∈ X. For every t ∈ R, we have
exp(2πi θ(x) t) dνa(x) =
∞Yn=0(cid:0)1 − an(1 − cos(2πt2−n))(cid:1) .
Since an ∈ (0, 1/4) for all n, it follows that for every sequence tk ∈ R, the following holds:
cµa(t) =ZK
cµa(tk) → 0 if and only if
This concludes the proof of the proposition.
an(1 − cos(2πtk2−n)) → 0
if and only if da(tk, 0) → 0 .
∞Xn=0
Combining Proposition 3.3 with the following result, we obtain a more concrete proof for the
non-classification of τ -topologies on R.
Proposition 3.4. Define the equivalence relation ∼ on X given by a ∼ b if and only if the
metrics da and db defined in (3.3) induce the same topology on R. Then ∼ has meager equiv-
alence classes and is generically S∞-ergodic. In particular, ∼ is not classifiable by countable
structures.
Proof. Define the equivalence relation ≈ on X given by a ≈ b if and only ifP∞
n=0 an −bn < ∞.
Let a ∈ X and tk ∈ R a sequence such that da(tk, 0) → 0. For every fixed n, we have that
an > 0 and thus that limk dZ(2−ntk) = 0. It follows that ≈ ⊂ ∼. Define the homeomorphism
ϕ : RN → X : ϕ(x)n = ϕ0(xn) where ϕ0(y) =
1
8
+
1
4π
arctan(y) .
Define the equivalence relation ≈1 on RN given by a ≈1 b if and only if a − b ∈ ℓ1
R(N). Since
ϕ0 is a contraction, when a ≈1 b, also ϕ(a) ≈ ϕ(b) and thus, ϕ(a) ∼ ϕ(b). Since ϕ is a
homeomorphism and since, by [Hj97, Proposition 3.25], ≈1 is generically S∞-ergodic, also ∼ is
generically S∞-ergodic.
Let U ⊂ X be a non meager set. By Lemma 3.5 below, we can take a ∈ U and a sequence
nk ∈ N such that nk → ∞ and da(2nk , 0) → 0. Also by Lemma 3.5 below, we can take b ∈ U
such that db(2nk , 0) 6→ 0. So the topologies on R induced by da and db are different. Since
a, b ∈ U and U was an arbitrary non meager set, it follows that ∼ has meager equivalence
classes.
Lemma 3.5. The set of a ∈ X for which there exists a sequence nk ∈ N such that nk → ∞
and da(2nk , 0) → 0 is comeager.
Given a sequence nk ∈ N with nk → ∞, the set of b ∈ X such that db(2nk , 0) 6→ 0 is comeager.
7
Proof. Note that for all a ∈ X and m ∈ N, we have that
da(2m, 0)2 =
am+k2−k .
∞Xk=1
Using that aj ∈ (0, 1/4) for all j ∈ N, it follows that for all a ∈ X and m, n ∈ N,
1
2
am+1 ≤ da(2m, 0)2 ≤ 2−n−2 +
am+k2−k .
nXk=1
(3.4)
Using the second inequality in (3.4), it follows that the first set in the lemma contains
∞\n=1
∞[m=nna ∈ X(cid:12)(cid:12)(cid:12)
nXk=1
am+k2−k < 2−no ,
which is a countable intersection of open dense sets.
Given a sequence nk ∈ N with nk → ∞, using the first inequality in (3.4), it follows that the
second set in the lemma contains
∞\k=1
∞[j=k(cid:8)b ∈ X(cid:12)(cid:12) bnj+1 > 1/8(cid:9) ,
which also is a countable intersection of open dense sets.
4 Non-classification of actions on free product II1 factors
Let G be a locally compact, second countable group and M a von Neumann algebra with
separable predual. An action of G on M is a continuous group homomorphism from G to the
Polish group Aut(M ). We denote this space of actions as Hom(G, Aut(M )). Equipped with
the topology of uniform convergence on compact sets, Hom(G, Aut(M )) is a Polish space.
There are several natural notions of equivalence of actions. Given α, β ∈ Hom(G, Aut(M )),
one says that
1. α and β are conjugate if there exists an element θ ∈ Aut(M ) such that θ ◦ βg ◦ θ−1 = αg for
all g ∈ G ;
2. α and β are cocycle conjugate if there exists a continuous map u : G → U (M ) and an
element θ ∈ Aut(M ) such that
θ ◦ βg ◦ θ−1 = (Ad ug) ◦ αg
and ugh = ugαg(uh)
for all g, h ∈ G ;
3. α and β are outer conjugate if there exists an element θ ∈ Aut(M ) such that for all g ∈ G,
θ ◦ βg ◦ θ−1 and αg have the same image in the outer automorphism group Out(M ) =
Aut(M )/ Inn(M ).
Clearly, conjugacy implies cocycle conjugacy, and cocycle conjugacy implies outer conjugacy,
but the converse implications need not hold.
We recall from [Sh97] the following definition of the τ -invariant of a group action on a full
factor, which is analogous to the τ -invariant for von Neumann algebras (see [Co74a] and the
beginning of Section 3).
8
Definition 4.1 ([Sh97, Definition 8.1]). Let G be a locally compact, second countable group
and M a full factor with separable predual. For every action α ∈ Hom(G, Aut(M )), the
topology τ (α) is defined as the weakest topology on G that makes the map G → Out(M ) :
g 7→ αg continuous.
Note that τ (α) is invariant under outer conjugacy. We can then combine Theorem 3.1 with
the construction of Gaussian actions to prove that for noncompact abelian groups G, cocycle
conjugacy and outer conjugacy of actions of G on the free group factors L(Fn), and more
generally on arbitrary free products (M, τ ) = (M1, τ1) ∗ (M2, τ2) of a diffuse and a nontrivial
tracial von Neumann algebra, are not classifiable by countable structures. In particular, outer
conjugacy of a single automorphism of L(Fn) or of such a free product (i.e. an action of G = Z)
is not classifiable by countable structures. Note that in [KLP14, Theorem 6.2], it was proven
that conjugacy of a single automorphism of a free product of II1 factors is not classifiable by
countable structures.
Theorem 4.2. Let G be a second countable, locally compact, noncompact, abelian group. Let
(M, τ ) = (M1, τ1) ∗ (M2, τ2) be the free product of two von Neumann algebras equipped with a
faithful normal tracial state. Assume that M1 is diffuse and that M2 6= C1. Then M is a full
II1 factor. None of the equivalence relations of conjugacy, cocycle conjugacy or outer conjugacy
on the space Hom(G, Aut(M )) of actions of G on M is classifiable by countable structures.
Before proving Theorem 4.2, we need the following lemma.
Lemma 4.3. Let (X, µ) be a standard nonatomic probability space and let G be a second
countable, locally compact, noncompact, abelian group. Denote by Autpmp(X, µ) the Polish
group of probability measure preserving automorphisms of (X, µ). There exists a continuous
map
with the following properties.
β : Probc(bG) → Hom(G, Autpmp(X, µ)) : ρ 7→ βρ
that βρ(g) = θ ◦ βη(g) ◦ θ−1 for all g ∈ G.
1. If ρ, η ∈ Probc(bG) are mutually absolutely continuous, there exists a θ ∈ Autpmp(X, µ) such
2. For every ρ ∈ Probc(bG), the weakest topology on G making the map G → Autpmp(X, µ) :
g 7→ βρ(g) continuous equals the topology τ (ρ).
Proof. Let H be a fixed separable Hilbert space. Since H can be viewed as a real Hilbert space,
we can view (X, µ) as the Gaussian probability space associated with H, see e.g. [AEG93, Proof
of Proposition 1.2]. We then obtain a continuous and injective homomorphism π : U (H) →
Autpmp(X, µ) with the property that for all un, u ∈ U (H), we have that un → u strongly if and
only if π(un) → π(u).
θ∗(γ) = λ, where λ is the Lebesgue measure on [0, 1]. Using the notation of Lemma 2.3, define
To prove the lemma, we choose a probability measure γ on bG such that γ ≈bλ, wherebλ is the
Haar measure on bG. Put H = L2(bG, γ). Choose a bijective Borel map θ : bG → [0, 1] such that
for every ρ ∈ Probc(bG), the Borel map
For every ρ ∈ Probc(bG), we define the unitary representation ζρ of G on H given by
ψρ : bG → bG : ψρ = θ−1 ◦ ϕθ∗(ρ) ◦ θ .
ζρ : G → U (L2(bG, γ)) : (ζρ(g)ξ)(ω) = (ψρ(ω))(g) ξ(ω) .
9
Using the unitary V in Lemma 2.3, it follows that the unitary representation ζρ is unitarily
equivalent to the unitary representation πρ defined in (3.1). In particular, if ρ, η ∈ Probc(bG) are
mutually absolutely continuous, the unitary representations ζρ and ζη are unitarily equivalent.
It follows that the map
has all the required properties.
β : Probc(bG) → Hom(G, Autpmp(X, µ)) : βρ(g) = π(ζρ(g))
Proof of Theorem 4.2. By [Ue10, Theorem 3.7], any free product (M, τ ) = (M1, τ1)∗(M2, τ2) of
tracial von Neumann algebras with M1 diffuse and M2 nontrivial is a full II1 factor. Moreover,
in the proof of [Ue10, Theorem 3.7], the following is shown:
if N1 ⊂ M1 is a diffuse von
Neumann subalgebra and if un ∈ M is a bounded sequence such that kunx − xunk2 → 0 for all
x ∈ N1 ∪ M2, then kun − τ (un)1k2 → 0.
We are in exactly one of the following two cases.
Case 1. M1 has a direct summand of type In.
Case 2. M1 is of type II1.
Proof in case 1. Take a nonzero central projection z ∈ Z(M1) such that M1z = A⊗Mn(C) for
some n ≥ 1 and some diffuse abelian von Neumann algebra A. Write A = L∞(X, µ) ⊗ B where
B is a diffuse abelian von Neumann algebra. Whenever β ∈ Autpmp(X, µ), we denote by Ψ(β) ∈
Aut(M1) the unique automorphism given by the identity on N1 = (B ⊗ Mn(C)) ⊕ M1(1 − z)
and given by β on L∞(X, µ) ⊂ M1z. Note that β 7→ Ψ(β) is a continuous homomorphism from
Autpmp(X, µ) to the Polish group of trace preserving automorphisms of (M1, τ1). Using the
notation of Lemma 4.3, define
Θ : Probc(bG) → Hom(G, Aut(M )) : ρ 7→ Θρ with Θρ(g) = Ψ(βρ(g)) ∗ id .
By Lemma 4.3, we get that the actions Θρ and Θη are conjugate whenever the probability
measures ρ, η ∈ Probc(bG) are mutually absolutely continuous.
We claim that the τ -invariant τ (Θρ) equals τ (ρ). To prove the claim, assume that Θρ(gn) → id
in Out(M ). We have to prove that Θρ(gn) → id in Aut(M ), because then βρ(gn) → id in
Autpmp(X, µ), so that gn → e in τ (ρ) by Lemma 4.3. Take un ∈ U (M ) such that (Ad un) ◦
Θρ(gn) → id in Aut(M ). Since the automorphisms Θρ(gn) act as the identity on N = N1 ∗ M2,
we get that kunx − xunk2 → 0 for all x ∈ N . By [Ue10, Theorem 3.7], as explained in the first
paragraph of the proof, we get that Ad un → id and the claim follows.
We can now complete the proof of the theorem in case 1. Assume that f : Hom(G, Aut(M )) →
Z is a Borel reduction of either conjugacy, cocycle conjugacy or outer conjugacy to the orbit
equivalence relation E of a continuous action of S∞ on a Polish space Z. Since G is noncompact,
Since Θρ and Θη are conjugate when ρ ≈ η, we get that (f (Θρ), f (Θη)) ∈ E whenever ρ ≈ η.
the dual group bG is nondiscrete and therefore, bG has no isolated points. By [KS99, Theorem 2],
the equivalence relation ≈ of mutual absolute continuity on Prob(bG) is generically S∞-ergodic.
By [KS99, Proposition 4.1], Probc(bG) ⊂ Prob(bG) is a dense Gδ-set. So ≈ is also generically
S∞-ergodic on Probc(bG).
So, there exists a comeager set U ⊂ Probc(bG) such that (f (Θρ), f (Θη)) ∈ E for all ρ, η ∈ U .
By Theorem 3.1, equality of τ -topologies on Prob(bG) has meager equivalence classes. So, we
find ρ, η ∈ U with τ (ρ) 6= τ (η). Since (f (Θρ), f (Θη)) ∈ E and f is a Borel reduction, we also
get that Θρ and Θη are outer conjugate. This implies that τ (Θρ) = τ (Θη). So, τ (ρ) = τ (η)
and we have reached a contradiction.
10
Proof in case 2. Fix a projection q ∈ M2 with 0 < τ (q) < 1. By [Dy92b, Proposition 3.2], we
can take a large enough integer n ≥ 1 such that the free product (Mn(C), τ ) ∗ (Cq + C(1 − q), τ )
is of type II1. A fortiori, (Mn(C), τ ) ∗ (M2, τ ) is of type II1. Since M1 is of type II1, we can
write M1 = N1 ⊗ Mn(C). Fix a minimal projection p ∈ Mn(C). Then, M ∼= pM p ⊗ Mn(C).
By [Dy92b, Theorem 1.2], we have pM p ∼= N1 ∗ N2 where N2 = p(Mn(C) ∗ M2)p. Now we can
use the same trick as in [KLP14, Theorem 6.2]. Since N1 and N2 are of type II1, we can write
Ni = Pi ⊗ M2(C). By [Dy92a, Theorem 3.5(iii)], we can identify
N1 ∗ N2 = (P1 ⊗ M2(C)) ∗ (P2 ⊗ M2(C)) = (L(F3) ∗ P1 ∗ P2) ⊗ M2(C) .
Further identifying L(F3) = L∞(X, µ) ∗ L(F2), we have found an isomorphism
θ : M → (L∞(X, µ) ∗ L(F2) ∗ P1 ∗ P2) ⊗ M2n(C) .
Defining
Θ : Probc(bG) → Hom(G, Aut(M )) : ρ 7→ Θρ with
Θρ(g) = θ−1 ◦(cid:0)(βρ(g) ∗ id ∗ id ∗ id) ⊗ id(cid:1) ◦ θ ,
the same reasoning as above concludes the proof of the theorem.
Remark 4.4. In the specific case of M = L(F∞), one can also consider the so-called free
Bogoljubov actions arising from Voiculescu's free Gaussian functor (see [Vo83]). When G is a
second countable, locally compact, noncompact, abelian group, the same arguments as in the
proof of Theorem 4.2 show that the equivalence relations of conjugacy, cocycle conjugacy and
outer conjugacy on the free Bogoljubov actions are not classifiable by countable structures.
References
[AEG93] S. Adams, G.A. Elliott and T. Giordano, Amenable actions of groups. Trans. Amer. Math.
Soc. 344 (1994), 803-822.
[Co74a] A. Connes, Almost periodic states and factors of type III1. J. Funct. Anal. 16 (1974), 415-445.
[Co74b] A. Connes, On hyperfinite factors of type III0 and Krieger's factors. J. Funct. Anal. 18 (1975),
318-327.
A. Connes, Classification of injective factors. Ann. of Math. 104 (1976), 73-115.
[Co75]
[CT76] A. Connes and M. Takesaki, The flow of weights on factors of type III. Tohoku Math. J. 29
(1977), 473-575.
[Dy92a] K. Dykema, Interpolated free group factors. Pacific J. Math. 163 (1994), 123-135.
[Dy92b] K. Dykema, Free products of hyperfinite von Neumann algebras and free dimension. Duke
[Ef64]
Math. J. 69 (1993), 97-119.
E.G. Effros, The Borel space of von Neumann algebras on a separable Hilbert space. Pacific
J. Math. 15 (1965), 1153-1164.
[FW03] M. Foreman and B. Weiss, An anti-classification theorem for ergodic measure preserving
[Ha85]
[Hj97]
transformations. J. Eur. Math. Soc. (JEMS) 6 (2004), 277-292.
U. Haagerup, Connes' bicentralizer problem and uniqueness of the injective factor of type
III1. Acta Math. 158 (1987), 95-148.
G. Hjorth, Classification and orbit equivalence relations. Mathematical Surveys and Mono-
graphs 75, American Mathematical Society, Providence, 2000.
[HSV16] C. Houdayer, D. Shlyakhtenko and S. Vaes, Classification of a family of non almost periodic
[Ke95]
free Araki-Woods factors. J. Eur. Math. Soc. (JEMS), to appear. arXiv:1605.06057
A.S. Kechris, Classical descriptive set theory. Graduate Texts in Mathematics 156, Springer-
Verlag, New York, 1995.
11
[KS99]
A.S. Kechris and N.E. Sofronidis, A strong generic ergodicity property of unitary and self-
adjoint operators. Ergodic Theory Dynam. Systems 21 (2001), 1459-1479.
[KLP08] D. Kerr, H. Li and M. Pichot, Turbulence, representations, and trace-preserving actions. Proc.
Lond. Math. Soc. 100 (2010), 459-484.
[KLP14] D. Kerr, M. Lupini and N.C. Phillips, Borel complexity and automorphisms of C ∗-algebras.
J. Funct. Anal. 268 (2015), 3767-3789.
[Kr74] W. Krieger, On ergodic flows and the isomorphism of factors. Math. Ann. 223 (1976), 19-70.
A. Ocneanu, Actions of discrete amenable groups on von Neumann algebras. Lecture Notes
[Oc85]
in Mathematics 1138, Springer-Verlag, 1985.
S. Popa, Deformation and rigidity for group actions and von Neumann algebras. In Pro-
ceedings of the International Congress of Mathematicians (Madrid, 2006), Vol. I, European
Mathematical Society Publishing House, 2007, pp. 445-477.
[Po06]
[Ru62] W. Rudin, Fourier analysis on groups. Interscience Tracts in Pure and Applied Mathematics
[Sh96]
[Sh97]
[Sh02]
[ST08]
[ST09]
[Ue10]
[Va04]
[Vo83]
12, John Wiley and Sons, New York, London, 1962.
D. Shlyakhtenko, Free quasi free-states. Pacific J. Math. 177 (1997), 329-368.
D. Shlyakhtenko, A-valued semicircular systems. J. Funct. Anal. 500 (1999), 1-47.
D. Shlyakhtenko, On the classification of full factors of type III. Trans. Amer. Math. Soc. 356
(2004), 4143-4159.
R. Sasyk and A. Tornquist. The classification problem of von Neumann factors, J. Funct.
Anal. 256 (2009), 2710-2724.
R. Sasyk and A. Tornquist, Turbulence and Araki-Woods factors. J. Funct. Anal. 259 (2010),
2238-2252.
Y. Ueda, Factoriality, type classification and fullness for free product von Neumann algebras.
Adv. Math. 228 (2011), 2647-2671.
S. Vaes, ´Etats quasi-libres libres et facteurs de type III (d'apr`es D. Shlyakhtenko). Ast´erisque
299, (2005), 329-350.
D. Voiculescu, Symmetries of some reduced free product C∗-algebras. In Operator algebras and
their connections with topology and ergodic theory (Busteni, 1983), Lecture Notes in Math.
1132, Springer, Berlin, 1985, pp. 556-588.
12
|
1107.2053 | 1 | 1107 | 2011-07-11T15:16:36 | Aligned CP-semigroups | [
"math.OA"
] | A CP-semigroup is aligned if its set of trivially maximal subordinates is totally ordered by subordination. We prove that aligned spatial E_0-semigroups are prime: they have no non-trivial tensor product decompositions up to cocycle conjugacy. As a consequence, we establish the existence of uncountably many non-cocycle conjugate E_0-semigroups of type II_0 which are prime. | math.OA | math |
ALIGNED CP-SEMIGROUPS
CHRISTOPHER JANKOWSKI, DANIEL MARKIEWICZ, AND ROBERT T. POWERS
Abstract. A CP-semigroup is aligned if its set of trivially maximal subordinates is totally
ordered by subordination. We prove that aligned spatial E0-semigroups are prime: they have
no non-trivial tensor product decompositions up to cocycle conjugacy. As a consequence, we
establish the existence of uncountably many non-cocycle conjugate E0-semigroups of type
II0 which are prime.
Let H be a Hilbert space, which we will always assume to be separable and infinite-
dimensional, and let B(H) denote the ∗-algebra of all bounded operators over H. A CP-
semigroup acting on B(H) is a point-σ-weakly continuous semigroup α = {αt : B(H) →
B(H)}t≥0 of normal completely positive contractions such that α0 = id. When αt is an en-
domorphism and αt(I) = I for all t ≥ 0, then α is called an E0-semigroup. In the special
case when H = K ⊗ L2(0, ∞), a CP-semigroup α acting on B(H) is called a CP-flow over K
if αt(A)St = StA for all A ∈ B(H), t ≥ 0 where {St : t ≥ 0} is the right shift semigroup. We
will say that a CP-semigroup β is a subordinate of α if it also acts on B(H) and αt − βt is
completely positive. If in addition β is a CP-flow, then it is called a flow subordinate of α.
We direct the reader to [Arv03] for a general reference on the theory of CP-semigroups and
to [Pow03b] for the basic theory of CP-flows.
In this paper we study a class of CP-semigroups which has a set of subordinates which
is minimal in a natural sense. We call such CP-semigroups aligned. This class is shown to
include examples considered previously in [Pow03a, APP06, Jan10, JMP11]. We prove that
aligned E0-semigroups have a notable property: they are prime in the sense that they have
no non-trivial tensor product decompositions up to cocycle conjugacy. As a consequence,
we establish the existence of uncountably many non-cocycle conjugate E0-semigroups of type
II0 which are prime. The previously known non-trivial examples of prime E0-semigroups
were obtained independently in [MP09] (type II1), [Tsi08] (type II1) and [Lie09] (type IIk for
k = 1, 2, . . . ).
Powers introduced in [Pow03a] the concept of q-purity which has been valuable for the
study of E0-semigroups (see for example [APP06, Jan10]). The notion of q-purity was refined
in [JMP11]: a CP-flow is q-pure if and only if its set of flow subordinates is totally ordered by
subordination. It is clear that an E0-semigroup which is in addition a q-pure CP-flow must
have index 0 or 1, and must be of type I1, II0 or II1. The case of type I0 is excluded because
an automorphism group cannot be a CP-flow.
The restriction to flow subordinates in the definition of q-pure CP-flows can obscure some
useful properties of these CP-semigroups. In order to circumvent this difficulty we consider
an alternative concept which is inspired by the notion of q-purity.
Definition 1. Let α be a CP-semigroup and let β be a CP-semigroup subordinate of α. We
will say that β is trivially maximal if the semigroup ektβt is not a subordinate of α for k > 0.
2000 Mathematics Subject Classification. Primary: 46L55, 46L57.
C.J. was partially supported by the Skirball Foundation via the Center for Advanced Studies in Mathematics
at Ben-Gurion University of the Negev.
D.M. and R.T.P. were partially supported by grant 2008295 from the U.S.-Israel Binational Science
Foundation.
1
2
CHRISTOPHER JANKOWSKI, DANIEL MARKIEWICZ, AND ROBERT T. POWERS
We will denote by S(α) the set of all trivially maximal subordinates of α partially ordered
by subordination. We will say α is aligned if S(α) is totally ordered.
Let β be a trivially maximal subordinate of α. We will denote by S(α; β) the set of all
trivially maximal CP-semigroup subordinates of α which dominate β, partially ordered by
subordination. We will say that α is aligned relative to β if S(α; β) is totally ordered.
Notice that if an aligned E0-semigroup is spatial, then it must have index zero.
Lemma 2. A unital CP-semigroup is aligned if and only if its minimal dilation is aligned.
Proof. Suppose that α is a unital CP-semigroup acting on H with minimal dilation αd acting
on B(H1), i.e. there exists an isometry W : H → H1 such that
αt(x) = W ∗αd
t (W xW ∗)W
and W W ∗ is increasing for αd. In order to prove the statement it suffices to show that there is
an order isomorphism between S(α) and S(αd). As proved in [Bha01] or by Theorem 3.5 of
[Pow03b], there exists an order isomorphism between the set of CP-semigroup subordinates of
α and the set of CP-semigroup subordinates of αd. This isomorphism is described as follows.
For every subordinate β of α there exists a unique subordinate β′ of αd such that
βt(x) = W ∗β′
(1)
It is clear that if β is trivially maximal (with respect to α), then β′ is trivially maximal (with
respect to αd). Conversely, suppose that β is not trivially maximal, so that there exists k > 0
such that ektβt is a subordinate of α. Then there exists γ a unique subordinate of αd such
that
t(W xW ∗)W.
ektβt(x) = W ∗γt(W xW ∗)W, ∀x ∈ B(H), t ≥ 0.
Therefore, by dividing by ekt we obtain that β is also the compression of e−ktγt which is
t = e−ktγt
obviously also a subordinate since k > 0. By uniqueness of the correspondence, β′
for all t. Hence β′ is not trivially maximal.
(cid:3)
CP-flow subordinates of a CP-flow are always trivially maximal, therefore if a CP-flow is
aligned then it is automatically q-pure. We also note that a CP-flow is q-pure if and only if it
is aligned with respect to the subordinate x 7→ StxS∗
t . We omit the elementary proof of this
fact.
We now show that unital CP-flows are aligned if and only if they are q-pure and induce E0-
semigroups of type II0. First let us approach the case when the CP-flow is in also a semigroup
of endomorphisms.
Proposition 3. Let α be an E0-semigroup which is in addition a CP-flow over K. If α has
type II0, then S(α) is precisely the set of flow subordinates of α. Therefore, α is aligned if
and only if it is type II0 and it is q-pure.
Proof. Suppose that α is an E0-semigroup of type II0 which is also a CP-flow over K. Note
that every flow subordinate of α is clearly an element of S(α), as flow subordinates are always
trivially maximal. Conversely, let β be a trivially maximal CP-semigroup subordinate of α.
By Theorem 3.4 of [Pow03b] there exists a family of operators (C(t))t≥0 in B(H) such that
βt(x) = C(t)αt(x),
where the family (C(t))t≥0 is a contractive positive local cocycle, i.e. C(t) ∈ αt(B(H))′ and
0 ≤ C(t) ≤ I for all t > 0, C(t + s) = C(t)αt(C(s)) for all t, s ≥ 0 and t 7→ C(t) is strongly
continuous for t ≥ 0 with C(t) → I as t → 0+.
Let St denote as usual the right shift semigroup on H = K ⊗ L2(0, ∞) and let Vt = C(t)St.
Note that Vt is strongly continuous and furthermore it is a semigroup: for all t, r ≥ 0,
VtVr = C(t)StC(r)Sr = C(t)αt(C(r))StSr = C(t + r)St+r = Vt+r.
ALIGNED CP-SEMIGROUPS
3
Moreover, Vt is a unit of αt, since for all x ∈ B(H),
αt(x)Vt = αt(x)C(t)St = C(t)αt(x)St = C(t)Stx = Vtx.
Notice, however, that α is type II0, therefore there exists κ ∈ C such that Vt = e−κtSt for all
t ≥ 0. Furthermore, since C(t) is a contraction, we have that Vt is a contraction for all t > 0,
hence we must have Re(κ) ≥ 0. We now show that in fact κ must be a real number. Recall
that βt is a CP-semigroup, and
0 ≤ S∗
t βt(I)St = S∗
t C(t)St = e−κtI,
hence κ is real and satisfies κ ≥ 0.
We will now prove that κ = 0. Let γt(x) = eκtβt(x). Notice that St is an intertwiner
semigroup for γ: for all t > 0 and x ∈ B(H),
γt(x)St = eκtβt(x)St = eκtC(t)αt(x)St = eκtC(t)Stx = Stx.
It follows that γ is a CPκ-flow in the sense of Definition 4.0 of [Pow03b], that is to say,
e−κtγt is a CP-semigroup and γ is intertwined by the shift. However, by Theorem 4.15 of
[Pow03b], every CPκ-flow must be a CP-flow. But a CP-flow is contractive, hence we must
have γt(I) ≤ I, thus
eκtβt(I) = eκtC(t) ≤ I
for all t > 0. Therefore, we have that for all positive t > 0,
ektβt(x) = eκtC(t)αt(x) ≤ αt(x).
Since β is trivially maximal, we must have that κ = 0. Thus we have shown that every
element of S(α) is a CP-flow.
Therefore, if α is q-pure and of type II0 then it is aligned. On the other hand, it is clear
that if α is an aligned E0-semigroup, then it is type II0 and q-pure as discussed before the
proposition.
(cid:3)
Theorem 4. Let α be a unital CP-flow over K. If the minimal dilation of α is type II0, then
S(α) is precisely the set of flow subordinates of α. Therefore, α is aligned if and only if it is
q-pure and its minimal dilation is type II0.
Proof. By Lemma 4.50 of [Pow03b], there exists a minimal dilation of α to an E0-semigroup
αd which is a CP-flow over the Hilbert space H1 = K1 ⊗ L2(0, ∞), i.e. there exists an isometry
W : H → H1 such that
αt(x) = W ∗αd
W W ∗ is increasing for αd and if Sd
t denotes the right shift semigroup on the space H1, then
W St = Sd
t W for all t > 0. We use the order isomorphism established in the proof of Lemma 2
that associates to each subordinate β ∈ S(α) a unique subordinate β′ ∈ S(αd) satisfying
(1). If αd is type II0 and β ∈ S(α), then it follows from the previous proposition that β′ is a
CP-flow. Hence for all t > 0,
t (W xW ∗)W,
βt(x)St = W ∗β′
t(W xW ∗)W St = W ∗β′
t(W xW ∗)Sd
t W = W ∗Sd
t W xW ∗W = Stx.
In other words, β is a CP-flow. Thus we proved that all elements of S(α) are CP-flows. On
the other hand, every flow subordinate of α is clearly an element of S(α).
Thus, if α is a unital q-pure CP-flow and its minimal dilation has type II0, then it is aligned.
Conversely, it is clear that if α is aligned then its minimal dilation αd as discussed above is
also aligned, thus it has index zero. Since αd is a CP-flow, it cannot be an automorphism
group, hence it has type II0. Moreover, since α is aligned, it is clearly q-pure.
(cid:3)
4
CHRISTOPHER JANKOWSKI, DANIEL MARKIEWICZ, AND ROBERT T. POWERS
Prime E0-semigroups
Definition 5. An E0-semigroup α is called prime if, whenever α is cocycle conjugate to β ⊗ γ
where β and γ are E0-semigroups, then β or γ is type I0.
It follows from the complete classification of E0-semigroups of type I in terms of the index,
and the additivity of the index with respect to tensoring, that a prime E0-semigroup of type
I is cocycle conjugate either to an automorphism group or to the CAR/CCR flow of index 1.
It is a corollary of the work on the gauge group of an E0-semigroup by Markiewicz-Powers
in [MP09] or Tsirelson in [Tsi08], that prime E0-semigroups of type II1 exist. And, as it has
belatedly come to our attention, earlier Liebscher had proven that prime E0-semigroups of
type IIk exist for k ≥ 1 (see Proposition 4.32 and Note 4.33 in [Lie09]). We establish that
there exist uncountably many prime E0-semigroups of type II0.
Theorem 6. Aligned spatial E0-semigroups are prime.
Proof. We prove the contrapositive. Suppose that α is an E0-semigroup and α = β⊗γ where β
and γ are two spatial E0-semigroups neither of which has type I0. Without loss of generality,
by applying an appropriate conjugacy, we may assume that both act on B(H) where H is
infinite-dimensional and separable. Let U and V be normalized units of β and γ, respectively.
Let us denote by θU and θV the semigroups given by θU
t (x) = VtxV ∗
t ,
which are E-semigroup subordinates of β and γ, respectively (notice these are not unital since
β and γ are not automorphism groups). Notice that σV = β ⊗ θV is a subordinate of α. We
show that it is trivially maximal. Suppose that k ≥ 0 and ektσV
is also a subordinate of α.
t
Then we have that for all x ∈ B(H),
t (x) = UtxU ∗
t and θV
ekt(I ⊗ VtV ∗
t ) = ekt(βt ⊗ θV
t )(I) = ektσV
t (I) ≤ αt(I) = I
However kVtk = 1, hence by taking norms on both sides we conclude ekt ≤ 1 for all t > 0.
Thus k = 0. Analogously, σU = θU ⊗ γ is trivially maximal.
We prove that σV and σU are incomparable. Suppose that σV ≥ σU . Notice that for all
x, y ∈ B(H) and t > 0,
σV
t (x ⊗ y)(Ut ⊗ I) = αt(x)Ut ⊗ θV
σU
t (x ⊗ y)(Ut ⊗ I) = θU
t (y))
t (x)Ut ⊗ γt(x) = (Ut ⊗ I)(x ⊗ γt(y))
t (x) = (Ut ⊗ I)(x ⊗ θV
Therefore we have that for all x, y ∈ B(H) and t > 0,
(Ut ⊗ I)∗hσV
t (x ⊗ y) − σU
t (x ⊗ y)i(Ut ⊗ I) = x ⊗ [θV
t (y) − γt(y)]
Thus in the special case when x = I, we have that the map y 7→ I ⊗ [θV
t (y) − γt(y)] is
completely positive for all t > 0, hence θV ≥ γ. However θV is a pure element in the cone of
completely positive maps, therefore we have that for every t > 0, γt is a multiple of θV
t , and
we have a contradiction since γ is not type I0. By symmetry, we obtain that σU and σV are
incomparable as asserted. Thus we proved that α is not aligned.
(cid:3)
Corollary 7. There exist uncountably many non-cocycle conjugate E0-semigroups of type II0
which are prime.
Proof. By Theorem 6, it suffices to exhibit an uncountable family of non-cocycle conjugate
aligned E0-semigroups. In order to do so, we consider a class of E0-semigroups arising from
boundary weight doubles as in [Jan10]. Let g(x) be a fixed complex-valued measurable
function such that g 6∈ L2(0, ∞) yet (1 − e−x)1/2g(x) ∈ L2(0, ∞), and for each t > 0 let
gt = χ(t,∞)g, which is an element of L2(0, ∞). Define the weight on B(L2(0, ∞)) given by
ν(A) = limt→0+(gt, Agt). For every 0 < λ < 1/2, let µλ : M2(C) → C be given by
µλ(X) = λx11 + (1 − λ)x22.
ALIGNED CP-SEMIGROUPS
5
Let us define the boundary weight map from M2(C)∗ to boundary weights on B(C2⊗L2(0, ∞))
given by ω(ρ)(A) = ρ(I)µλ(cid:0)(I ⊗ ν)(A)(cid:1) for all ρ ∈ M2(C) and A in the domain of finiteness of
I⊗ν (the so called null boundary algebra of definition 4.16 in [Pow03b]). Then by Corollary 3.3
of [Jan10] and the assumptions on g(x), we have that ω gives rise to an E0-semigroup of type
II0. Once one applies Theorem 4.4 of [JMP11] to reconcile the earlier definition of q-purity
with the one in this paper, we obtain from Lemma 4.3 and Proposition 5.2 of [Jan10] that
ω gives rise to a q-pure unital CP-flow. Therefore by Theorem 4 it gives to a aligned E0-
semigroup αλ. Finally, by Theorem 5.4 of [Jan10], given λ, ζ ∈ (0, 1/2), we have that αλ is
cocycle conjugate to αζ if and only if λ = ζ.
(cid:3)
We should point out that it is possible to obtain a different uncountable family of non-
cocycle conjugate E0-semigroups by using Theorem 3.22 of [Pow03a]. For details see the
discussion in the end of section III therein. This would be indeed be a different family from
the one exhibited above in the sense that, by Corollary 5.5 of [Jan10], the E0-semigroups
constituting both families are not cocycle conjugate.
References
[APP06] A. Alevras, R. T. Powers, and G. L. Price, Cocycles for one-parameter flows of B(H), J. Funct.
Anal. 230 (2006), no. 1, 1 -- 64.
[Arv03] W. Arveson, Noncommutative dynamics and E-semigroups, Springer Monographs in Mathematics,
Springer-Verlag, New York, 2003.
[Bha01] B. V. R. Bhat, Cocycles of CCR flows, Mem. Amer. Math. Soc. 149 (2001), no. 709, x+114.
[Jan10] C. Jankowski, On type II0 E0-semigroups induced by boundary weight doubles, J. Funct. Anal. 258
(2010), no. 10, 3413 -- 3451.
[JMP11] C. Jankowski, D. Markiewicz, and R. T. Powers, E0-semigroups and q-purity, preprint
[Lie09]
arXiv:1106.2304, 2011.
V. Liebscher, Random sets and invariants for (type II) continuous tensor product systems of Hilbert
spaces, Mem. Amer. Math. Soc. 199 (2009), no. 930, xiv+101.
[MP09] D. Markiewicz and R. T. Powers, Local unitary cocycles of spatial E0-semigroups, J. Funct. Anal.
256 (2009), no. 5, 1511 -- 1543.
[Pow03a] R. T. Powers, Construction of E0-semigroups of B(H) from CP-flows, Advances in quantum dy-
namics (South Hadley, MA, 2002), Contemp. Math., vol. 335, Amer. Math. Soc., Providence, RI,
2003, pp. 57 -- 97.
[Pow03b]
, Continuous spatial semigroups of completely positive maps of B(H), New York J. Math. 9
[Tsi08]
(2003), 165 -- 269 (electronic).
B. Tsirelson, On automorphisms of type II Arveson systems (probabilistic approach), New York J.
Math. 14 (2008), 539 -- 576.
Christopher Jankowski, Department of Mathematics, Ben-Gurion University of the Negev,
P.O.B. 653, Be'er Sheva 84105, Israel.
E-mail address: [email protected]
Daniel Markiewicz, Department of Mathematics, Ben-Gurion University of the Negev, P.O.B.
653, Be'er Sheva 84105, Israel.
E-mail address: [email protected]
Robert T. Powers, Department of Mathematics, David Rittenhouse Lab., 209 South 33rd St.,
Philadelphia, PA 19104-6395, U.S.A.
E-mail address: [email protected]
|
1307.6988 | 1 | 1307 | 2013-07-26T10:43:03 | Bounded elements of C*-inductive locally convex spaces | [
"math.OA",
"math.RA"
] | The notion of bounded element of C*-inductive locally convex spaces (or C*-inductive partial *-algebras) is introduced and discussed in two ways: the first one takes into account the inductive structure provided by certain families of C*-algebras; the second one is linked to natural order of these spaces. A particular attention is devoted to the relevant instance provided by the space of continuous linear maps acting in a rigged Hilbert space. | math.OA | math |
BOUNDED ELEMENTS OF C*-INDUCTIVE LOCALLY
CONVEX SPACES
GIORGIA BELLOMONTE, SALVATORE DI BELLA, AND CAMILLO TRAPANI
Abstract. The notion of bounded element of C*-inductive locally con-
vex spaces (or C*-inductive partial *-algebras) is introduced and dis-
cussed in two ways: the first one takes into account the inductive struc-
ture provided by certain families of C*-algebras; the second one is linked
to natural order of these spaces. A particular attention is devoted to
the relevant instance provided by the space of continuous linear maps
acting in a rigged Hilbert space.
1. Introduction
Some locally convex spaces exhibit an interesting feature: they contain a
large number of C*-algebras that often contribute to their topological struc-
ture, in the sense that these spaces can be thought as generalized inductive
limits of C*-algebras. These objects were called C*-inductive locally convex
spaces in [8] and their structure was examined in detail, also taking in mind
that they arise naturally when one considers the operators acting in the
joint topological limit of an inductive family of Hilbert spaces as described
in [7]. Indeed, a typical instance of this structure is obtained by considering
B(D, D×) of operators acting in the rigged Hilbert space canoni-
the space L
cally associated to an O*-algebra of unbounded operators acting on a dense
domain D of Hilbert space H. In [8] a series of features of this structure
were studied giving a particular attention to the order structure, positive
B(D, D×) contains
linear functionals and representation theory. The space L
a subspace isomorphic to the *-algebra B(H) of bounded operators in H
whose elements can be in natural way considered as the bounded elements of
B(D, D×). The notion of bounded element of a locally convex *-algebra A
L
was first introduced by Allan [1] with the aim of developing a spectral theory
for topological *-algebras: an element x of the topological *-algebra A[τ ] is
2010 Mathematics Subject Classification. 47L60, 47L40.
Key words and phrases. bounded elements, inductive limit of C*-algebras, partial *-
algebras.
1
2
GIORGIA BELLOMONTE, SALVATORE DI BELLA, AND CAMILLO TRAPANI
Allan bounded if there exists λ 6= 0 such that the set {(λ−1x)n; n = 1, 2, . . .}
is a bounded subset of A[τ ]. This definition was suggested by the successful
spectral analysis for closed operators in Hilbert space H: a complex number
λ is in the resolvent set ρ(T ) of a closed operator T if T − λI has an inverse
in the *-algebra B(H) of bounded operators.
There are, however, several other possibilities for defining bounded ele-
ments. For instance, one may say that x is bounded if π(x) is a bounded
operator, for every (continuous, in a certain sense) *-representation π de-
fined on a dense domain Dπ of some Hilbert space Hπ. This could be a
reasonable definition in itself, provided that A possesses sufficiently many
*-representations in Hilbert space.
Moreover some attempts to extend this notion to the larger set-up of
locally convex quasi *-algebras [18, 20, 21, 11] or locally convex partial *-
algebras [4, 5, 6] have been done. But in these cases, Allan's notion cannot
be adopted, since powers of a given element x need not be defined.
In the case of *-algebras, bounded elements in purely algebraic terms
have been considered by Vidav [23] and Schmudgen [17] with respect to
some (positive) wedge.
The aim of this paper is to extend the notion of bounded element to the
case of C*-inductive locally convex spaces A with defining family of C*-
algebras {Bα; α ∈ F} (F is an index set directed upward). There are also
in this case several possibilities: the first one consists in taking elements
that have representatives in every C*-algebra Bα of the family whose norms
are uniformly bounded; the second one consists into taking into account the
order structure of A, in the same spirit of the quoted papers of Vidav and
Schmudgen.
The paper is organized as follows. After some preliminaries (Section 2),
B(D, D×) can be derived
we study, in Section 3, how bounded elements of L
from its C*-inductive structure and from its order structure. We show that
these two notions are equivalent and that an element X is bounded if and
only if X maps D into H and X ∈ B(H). Finally, in Section 4, we consider
the same problem for abstract C*-inductive locally convex spaces and give
B(D, D×) maintain
conditions for some of the characterizations proved for L
their validity. Some of these results are then specialized to the case where
A is a C*-inductive locally convex partial *-algebra.
BOUNDED ELEMENTS OF C*-INDUCTIVE SPACES
3
2. Notations and preliminaries
For general aspects of the theory of partial *-algebras and of their repre-
sentations, we refer to the monograph [2]. For the convenience of the reader,
however, we repeat here the essential definitions.
A partial *-algebra A is a complex vector space with conjugate linear
involution ∗ and a distributive partial multiplication ·, defined on a subset
Γ ⊂ A× A, satisfying the property that (x, y) ∈ Γ if, and only if, (y∗, x∗) ∈ Γ
and (x · y)∗ = y∗ · x∗. From now on, we will write simply xy instead of x · y
whenever (x, y) ∈ Γ. For every y ∈ A, the set of left (resp. right) multipliers
of y is denoted by L(y) (resp. R(y)), i.e., L(y) = {x ∈ A : (x, y) ∈ Γ}, (resp.
R(y) = {x ∈ A : (y, x) ∈ Γ}). We denote by LA (resp. RA) the space of
universal left (resp. right) multipliers of A. In general, a partial *-algebra
is not associative.
The unit of partial *-algebra A, if any, is an element e ∈ A such that
e = e∗, e ∈ RA ∩ LA and xe = ex = x, for every x ∈ A.
Let H be a complex Hilbert space and D a dense subspace of H. We
denote by L†(D, H) the set of all (closable) linear operators X such that
D(X) = D, D(X*) ⊇ D. The map X → X † = X ∗
↾D defines an involution
on L†(D, H), which can be made into a partial *-algebra with respect to the
weak multiplication [2]; however, this fact will not be used in this paper.
Let L†(D) be the subspace of L†(D, H) consisting of all its elements which
leave, together with their adjoints, the domain D invariant. Then L†(D) is a
*-algebra with respect to the usual operations. A *-subalgebra M of L†(D)
is called an O*-algebra.
Let M be an O*-algebra. The graph topology tM on D is the locally convex
topology defined by the family {k · kA}A∈M, where
kξkA =pkξk2 + kAξk2 = k(I + A∗A)1/2ξk,
ξ ∈ D.
For A = 0, the null operator of L†(D), k · k0 is exactly the norm of H, thus
we will omit the 0 in the notation of the norm.
The topology tM is finer than the norm topology, unless M does consist
of bounded operators only.
If M is an O*-algebra, we write A (cid:22) B if kAξk ≤ kBξk, for every ξ ∈ D.
Then, M is directed upward with respect to this order relation.
If A ∈ M, we denote by HA the Hilbert space obtained by endowing D(A)
with the graph norm k · kA.
4
GIORGIA BELLOMONTE, SALVATORE DI BELLA, AND CAMILLO TRAPANI
If A, B ∈ M and A (cid:22) B, then UBA = (I + B∗B)−1/2(I + A∗A)1/2 is a
contractive map of HA into HB; i.e., kUBAξkB ≤ kξkA, for every ξ ∈ HA.
If the locally convex space D[tM ] is complete, then M is said to be closed.
If M = L†(D) then the corresponding graph topology is denoted by t†
instead of tL†(D).
As is known, a locally convex topology t on D finer than the topology
induced by the Hilbert norm defines, in standard fashion, a rigged Hilbert
space (RHS)
D[t] ֒→ H ֒→ D×[t×],
where D× is the vector space of all continuous conjugate linear functionals on
D[t], i.e., the conjugate dual of D[t], endowed with the strong dual topology
t× = β(D×, D) and ֒→ denotes a continuous embedding with dense range.
The Hilbert space H is identified (by considering the form which puts D
and D× as an extension of the inner product of D) with a dense subspace
of D×[t×].
Let L(D, D×) denote the vector space of all continuous linear maps from
D[t] into D×[t×]. In L(D, D×) an involution X 7→ X † can be introduced by
the equality
hXξ η i = hX †η ξ i,
∀ξ, η ∈ D.
Hence L(D, D×) is a *-invariant vector space.
To every X ∈ L(D, D×) there corresponds a separately continuous sesquilin-
ear form θX on D × D defined by
θX(ξ, η) = hXξ η i ,
ξ, η ∈ D.
The vector space of all jointly continuous sesquilinear forms on D × D will
B(D, D×) the subspace of all X ∈
be denoted with B(D, D). We denote by L
L(D, D×) such that θX ∈ B(D, D) and by L†(D) the *-algebra consisting
of all operators of L†(D), which together with their adjoints are continuous
If t = t†, then L†(D) = L†(D). We will refer to
from D[t] into D[t].
the rigged Hilbert space defined by endowing D with the topology t† as
to the canonical rigged Hilbert space defined by L†(D) on D. In this case
(L
B(D, D×), L†(D)) is a quasi *-algebra [2].
The spaces L(D, D×) and L
B(D, D×) have been studied at length by sev-
[12, 13, 14, 22]) and several pathologies concerning
eral authors (see, e.g.
their multiplicative structure have been considered (see also [2, 3] and ref-
erences therein). Recently some spectral properties of operators of these
classes have also been studied [10].
BOUNDED ELEMENTS OF C*-INDUCTIVE SPACES
5
3. Bounded elements of L
B(D, D×)
The inductive structure of L
B(D, D×), with D endowed with the graph to-
pology t†, has been discussed in [8, Section 5]. To keep the paper reasonably
self-contained, we sum the main features up.
By the definition itself, X ∈ L
B(D, D×) if, and only if, there exists γX > 0
and A ∈ L†(D) such that
(1)
θX(ξ, η) = hXξ η i ≤ γXkξkA kηkA,
∀ξ, η ∈ D.
Conversely, if θ ∈ B(D, D), there exists a unique X ∈ L
B(D, D×) such
that θ = θX.
Thus, the map
I : X ∈ L
B(D, D×) 7→ θX ∈ B(D, D)
is an isomorphism of vector spaces and I(θ∗) = X †, where θ∗(ξ, η) = θ(η, ξ),
for every ξ, η ∈ D.
We denote by BA(D, D) the subspace of B(D, D) consisting of all θ ∈
B(D, D) such that (1) holds for fixed A ∈ L†(D).
If θ ∈ BA(D, D), it extends to a bounded sesquilinear form on HA × HA
(we use the same symbol for this extension). Hence, there exists a unique
operator X θ
A ∈ B(HA) such that
θ(ξ, η) =DX θ
Aξ ηEA
,
∀ξ, η ∈ HA.
On the other hand, if XA ∈ B(HA), then the sesquilinear form θXA defined
by
θXA(ξ, η) = hXAξ η iA ,
ξ, η ∈ D,
is an element of BA(D, D) and the map
ΦA : XA ∈ B(HA) → θXA ∈ BA(D, D)
is a *-isomorphism of vector spaces with involution.
If B (cid:23) A, then, for ξ, η ∈ D,
θXA(ξ, η) = hXAξ η iA ≤ kXAkA,AkξkA kηkA ≤ kXAkA,AkξkB kηkB ,
where k · kA,A denotes the operator norm in B(HA). Hence, there exists a
unique XB ∈ B(HB) such that
hXAξ η iA = hXBξ η iB ,
∀ξ, η ∈ D.
So it is natural to define
JBA(XA) = XB,
∀XA ∈ B(HA).
6
GIORGIA BELLOMONTE, SALVATORE DI BELLA, AND CAMILLO TRAPANI
It is easily seen that JBA = Φ−1
B ΦA.
The space LA
B (D, D×) := I−1BA(D, D) is a Banach space, with norm
kXkA :=
sup
θX (ξ, η)
kξkA,kηkA≤1
and L
the family of subspaces {LA
B(D, D×) can be endowed with the inductive topology τind defined by
B (D, D×); A ∈ L†(D)} as in [16, Section 1.2. III].
In conclusion,
XA ∈ B(HA) ↔ θXA ∈ BA(D, D) ↔ X ∈ LA
B(D, D×)
are isometric *-isomorphisms of Banach spaces.
Hence, to every X ∈ L
B(D, D×) one can associate the net {XB; B ∈
L†(D); B (cid:23) A} of its representatives in each of the spaces HB.
Definition 3.1. We say that X ∈ L
L
B(D, D×) if X has a representative XA in every B(HA) and
B(D, D×) is a bounded element of
kXkb := sup
kXAkA,A < +∞.
A∈L†(D)
The space L
B(D, D×)b of all bounded elements of L
B(D, D×) is a Banach
space with norm k · kb.
Proposition 3.2. L
algebra of operators.
B(D, D×)b is *-isomorphic (as Banach space) to a C*-
Proof. Let H⊕ denote the Hilbert space direct sum of the HA, A ∈ L†(D);
i.e.,
HA
H⊕ := MA∈L†(D)
=(ξ⊕ = (ξA); ξA ∈ HA, ∀A ∈ L†(D) and XA
kξAk2
A < +∞) .
If {XA}A∈L†(D) is a net of operators XA ∈ B(HA), A ∈ L†(D), we define
X⊕ξ⊕ = {XAξA} provided that PA kXAξAk2 < +∞, ξA ∈ HA.
The space constructed in this way is QA
The operator X⊕ = {XA} is bounded if and only if supA kXAkA,A < +∞.
B(HA) = B(H⊕). To every X ∈
B(D, D×)b we can associate the net {XA} which we have defined above.
Clearly, {XA} ∈ B(H⊕). It is easily seen that the map
L
τ : X ∈ L
B(D, D×)b 7→ {XA} ∈ B(H⊕)
is isometric. Thus, the statement is proved.
(cid:3)
BOUNDED ELEMENTS OF C*-INDUCTIVE SPACES
7
B(D, D×) having a representative XA for
Remark 3.3. An element X ∈ L
every A ∈ L†(D) need not be bounded in the sense of Definition 3.1. The
spaces {HA; A ∈ L†(D)}, together with their conjugate duals make D× into
an indexed PIP-space [3, Ch.2]. In that language, operators having repre-
sentatives in every HA are called totally regular operators. For more details
on their behavior see [3, Sect. 3.3.3] where also a C*-agebra corresponding
to our bounded elements has been studied.
Our next goal is to characterize bounded elements of L
B(D, D×) in sev-
eral different ways. For doing this, we need to consider the natural order
structure of L
B(D, D×).
We say that X ∈ L
B(D, D×) is positive, and write X ≥ 0, if hXξ ξ i ≥ 0,
for every ξ ∈ D.
It is easy to see that, if X is positive, then it is symmetric; i.e., X = X †.
Proposition 3.4. The following conditions are equivalent.
(i) X ≥ 0.
(ii) There exists A ∈ L†(D) such that XB ≥ 0,
∀B (cid:23) A.
Proof. (i)⇒(ii): Since X ∈ L
such that
B(D, D×), there exists A ∈ L†(D) and γ > 0
hXξ η i ≤ γkξkBkηkB , B (cid:23) A.
If X ≥ 0, then, for every ξ ∈ D,
hXBξ ξ iB = hXξ ξ i ≥ 0,
∀B (cid:23) A.
Since D is dense in HB, we have hXBξ ξ iB ≥ 0, ∀ξ ∈ HB.
(ii)⇒(i): Let XB ≥ 0 for every B (cid:23) A. Then, for every ξ ∈ D, hXξ ξ i =
hXBξ ξ iB ≥ 0.
(cid:3)
Theorem 3.5. Let X ∈ L
lent.
B(D, D×). The following statements are equiva-
(i) X : D → H and X ∈ B(H).
(ii) X ∈ L
(iii) There exists λ > 0 such that
B(D, D×)b.
−λI ≤ ℜ(X) ≤ λI, −λI ≤ ℑ(X) ≤ λI
where ℜ(X) = X+X †
2
and ℑ(X) = X−X †
2i
.
Proof. (i)⇒(ii): If X : D → H and X is bounded, then, for every A ∈ L†(D),
(2)
hXξ η i ≤ kXkkξkkηk ≤ kXkkξkAkηkA.
8
GIORGIA BELLOMONTE, SALVATORE DI BELLA, AND CAMILLO TRAPANI
This means that X has a bounded representative XA in every B(HA). By
(2), kXAkA,A ≤ kXk, for every A ∈ L†(D), so supA∈L†(D) kXAkA,A < +∞.
(ii)⇒(i) Let X ∈ L
B(D, D×)b. Then, for every A ∈ L†(D)
∀ξ, η ∈ D.
hXξ η i ≤ kXAkA,A kξkA kηkA,
In particular, for A = 0,
(3)
hXξ η i ≤ kX0kkξkkηk,
∀ξ, η ∈ D.
By (3), for every ξ ∈ D, F (η) = hXξ η i is a bounded conjugate linear
functional on D, so by Riesz's lemma Xξ ∈ H. It is, finally easily seen that
X ∈ B(H).
(iii)⇒(i) Suppose first that X = X †. Note that the operator X sat-
is a positive operator and
isfies the following: 0 ≤ X+λI
2λ ≤ I; so X+λI
2λ
(cid:10) X+λI
2λ ξ ξ(cid:11) ≤ hξ ξ i , ∀ξ ∈ D; this implies that
and by Riesz's lemma there exists ζ ∈ H such that
(4)
(5)
2λ
(cid:12)(cid:12)(cid:12)(cid:12)
(cid:28) X + λI
(cid:28) X + λI
2λ
≤ kξk kηk,
ξ η(cid:29)(cid:12)(cid:12)(cid:12)(cid:12)
ξ η(cid:29) = hζ η i ,
∀ξ, η ∈ D
∀ξ, η ∈ D
and then X+λI
representative for every A ∈ L†(D). Indeed,
2λ ξ ∈ H. This implies that Xξ ∈ H too. Moreover, X has a
hXξ η i ≤ γkξkkηk ≤ γkξkAkηkA ∀A ∈ L†(D),
where γ > 0. From (4) it follows that X is bounded and X ∈ B(H). In the
very same way one can prove the boundedness of X if X † = −X. The result
for a general X follows easily.
(i)⇒ (iii): this is a standard result of the C*-algebras theory.
(cid:3)
4. Bounded elements of C*-inductive locally convex spaces
The results obtained in Section 3 have an abstract generalization to locally
convex spaces that are inductive limits of C*-algebras in a generalized sense.
These spaces were called C*-inductive locally convex spaces in [8]. We begin
with recalling the basic definitions.
Let A be a vector space over C. Let F be a set of indices directed upward
and consider, for every α ∈ F, a Banach space Aα ⊂ A such that:
(I.1) Aα ⊆ Aβ, if α ≤ β;
(I.2) A =Sα∈F
Aα;
BOUNDED ELEMENTS OF C*-INDUCTIVE SPACES
9
(I.3) ∀α ∈ F, there exists a C*-algebra Bα (with unit eα and norm k · kα)
and a norm-preserving isomorphism of vector spaces φα : Bα → Aα;
β , for every α, β ∈ F with β ≥ α.
β φα)(xα) ∈ B+
α ⇒ xβ = (φ−1
(I.4) xα ∈ B+
We put jβα = φ−1
If x ∈ A, there exist α ∈ F such that x ∈ Aα and (a unique) xβ ∈ Bβ
β φα, if α, β ∈ F, β ≥ α.
such that x = φβ(xβ), for all β ≥ α.
Then, we put
jβα(xα) := xβ
if α ≤ β.
By (I.4), it follows easily that jβα preserves the involution; i.e., jβα(x∗
α) =
(jβα(xα))∗.
The family {Bα, jβα, β ≥ α} is a directed system of C*-algebras, in the
sense that:
(J.1) for every α, β ∈ F, with β ≥ α, jβα : Bα → Bβ is a linear and
injective map; jαα is the identity of Bα,
(J.2) for every α, β ∈ F, with α ≤ β, φα = φβjβα.
(J.3) jγβjβα = jγα, α ≤ β ≤ γ.
We assume that, in addition, the jβα's are Schwarz maps (see, e.g.
i.e.,
[15]);
(sch) jβα(xα)∗jβα(xα) ≤ jβα(x∗
αxα),
∀xα ∈ Bα, α ≤ β.
For every α, β ∈ F, with α ≤ β, jβα is continuous [15] and, moreover,
kjβα(xα)kβ ≤ kxαkα,
∀xα ∈ Bα.
An involution in A is defined as follows.. Let x ∈ A. Then x ∈ Aα, for
α).
some α ∈ F, i.e., x = φα(xα), for a unique xα ∈ Bα. Put x∗ := φα(x∗
Then if β ≥ α, we have
φ−1
β (x∗) = φ−1
β (φα(x∗
α)) = jβα(x∗
α) = (jβα(xα))∗ = x∗
β.
It is easily seen that the map x 7→ x∗ is an involution in A. Moreover, by
the definition itself, it follows that every map φα preserves the involution;
i.e., φα(x∗
α) = (φα(xα))∗, for all xα ∈ Bα, α ∈ F.
Definition 4.1. A locally convex vector space A, with involution ∗, is called
a C*-inductive locally convex space if
(i) there exists a family {{Bα, φα}, α ∈ F}, where F is a direct set and,
for every α ∈ F, Bα is a C*-algebra and φα is a linear injective map
of Bα into A, satisfying the above conditions (I.1) - (I.4) and (sch),
with Aα = φα(Bα), α ∈ F;
10
GIORGIA BELLOMONTE, SALVATORE DI BELLA, AND CAMILLO TRAPANI
(ii) A is endowed with the locally convex inductive topology τind gener-
ated by the family {{Bα, φα}, α ∈ F}.
The family {{Bα, φα}, α ∈ F} is called the defining system of A. We
notice that the involution is automatically continuous in A[τind].
A C*-inductive locally convex space has a natural positive cone.
An element x ∈ A is called positive if there exists γ ∈ F such that φ−1
α , ∀α ≥ γ.
B+
We denote by A+ the set of all positive elements of A.
α (x) ∈
Then,
(i) Every positive element x ∈ A is hermitian; i.e., x ∈ Ah := {y ∈ A :
y∗ = y} .
(ii) A+ is a non empty convex pointed cone; i.e. A+ ∩ (−A+) = {0}.
(iii) If α ∈ F and xα ∈ B+
Moreover, every hermitian element x = x∗ is the difference of two positive
α , φα(xα) is positive.
elements, i.e. there exist x+, x− ∈ A+ such that x = x+ − x−.
A linear functional ω is said to be positive if ω(x) ≥ 0 for every x =
(xα) ∈ A. As shown in [8, Prop. 3.9, 3.10], ω is positive if, and only
if, ωα(xα) := ω(φα(xα)) ≥ 0 for every α ∈ F. We write, in this case,
ω = lim
−→
ωα.
4.1. Bounded elements.
Definition 4.2. Let A be a C*-inductive locally convex space. An element
x = (xα) ∈ A, with xα ∈ Bα, is called bounded if x ∈ Aα, for every α ∈ F
and supα∈F kxαkα < ∞. The set of bounded elements of A is denoted by
Ab.
Proposition 4.3. The set Ab is a Banach space under the norm kxkb =
supα∈F kxαkα.
Proof. We only prove the completeness. Let {xn} be a Cauchy sequence in
Ab. Then for every α ∈ F the sequence {xα
n := (xn)α, is Cauchy
in Bα so it converges to some xα ∈ Bα. Since the jβα's are continuous, one
easily proves that the family {xα} defines an element x = (xα) of A. From
the Cauchy condition, for every ǫ > 0, there exists nǫ ∈ N such that
n}, with xα
(6)
sup
α∈F
kxα
n − xα
mkα < ǫ
BOUNDED ELEMENTS OF C*-INDUCTIVE SPACES
11
If m > nǫ,
Hence,
Thus x ∈ Ab.
kxαkα ≤ kxα − xα
mkα.
mkα + kxα
mkα ≤ ǫ + kxα
sup
α∈F
kxαkα ≤ ǫ + sup
α∈F
kxα
mkα < ∞.
Fix now n > nǫ and let m → ∞ in (6). Then,
kxα
n − xαkα ≤ ǫ.
sup
α∈F
This proves that xn → x.
(cid:3)
In what follows we will consider *-representations of a C*-inductive locally
convex space. We recall the basic definitions.
Let F be a set directed upward by ≤. A family {Hα, Uβα, α, β ∈ F, β ≥ α},
where each Hα is a Hilbert space (with inner product h· · iα and norm k · kα)
and, for every α, β ∈ F, with β ≥ α, Uβα is a linear map from Hα into
Hβ, is called a directed contractive system of Hilbert spaces if the following
conditions are satisfied
(i) Uβα is injective;
(ii) kUβαξαkβ ≤ kξαkα,
(iii) Uαα = Iα, the identity of Hα;
(iv) Uγα = UγβUβα, α ≤ β ≤ γ.
∀ξα ∈ Hα;
A directed contractive system of Hilbert spaces defines a conjugate dual
pair (D×, D) which is called the joint topological limit [7] of the directed
contractive system {Hα, Uβα, α, β ∈ F, β ≥ α} of Hilbert spaces.
Definition 4.4. Let A be the C*-inductive locally convex space defined by
the system {{Bα, Φα}, α ∈ F} as in Definition 4.1.
For each α ∈ F, let πα be a *-representation of Bα in Hilbert space Hα.
The collection π := {πα} is said to be a *-representation of A if
(i) for every α, β ∈ F there exists a linear map Uβα : Hα → Hβ such
that the family {Hα, Uβα, α, β ∈ F, β ≥ α} is a directed contractive
system of Hilbert spaces;
(ii) the following equality holds
(7)
πβ(jβα(xα)) = Uβαπα(xα)U ∗
βα,
∀xα ∈ Bα, β ≥ α.
12
GIORGIA BELLOMONTE, SALVATORE DI BELLA, AND CAMILLO TRAPANI
In this case we write π(x) = lim
−→
π = lim
−→
πα.
πα(xα) for every x = (xα) ∈ A or, for short,
The *-representation π is said to be faithful if x ∈ A+ and π(x) = 0
imply x = 0 (of course, π(x) = 0 means that there exists γ ∈ F such that
πα(xα) = 0, for α ≥ γ).
Remark 4.5. With this definition (which is formally different from that
given in [8] but fully equivalent), π(x), x ∈ A, is not an operator but rather
a collection of operators. But as shown in [8], π(x) can be regarded as an
operator acting on the joint topological limit (D×, D) of {Hα, Uβα, α, β ∈
F, β ≥ α}. The corresponding space of operators was denoted by LB(D, D×);
B(D, D×) studied in Section
it behaves in the very same way as the space L
3 and reduces to it when the family of Hilbert spaces is exactly {HA; A ∈
L†(D)}. The main difference consists in the fact that the Hα's need not be
all subspaces of a certain Hilbert space H.
Lemma 4.6. Let π = lim
−→
every α ∈ F, πα is a faithful *-representation of Bα.
πα be a faithful *-representation of A. Then, for
Proof. Let xα ∈ B+
α with πα(xα) = 0. Let x ∈ A be the unique element of
A such that x = φα(xα). Then πβ(xβ) = πβ(jβα(xα)) = Uβαπα(xα)U ∗
βα = 0.
Hence π(x) = 0 and, therefore x = 0. Thus there exists γ ∈ F such that
xγ = 0, for γ ≥ γ. Let β ≥ α, γ. Then 0 = xβ = jβα(xα). Hence, by the
injectivity of jβα, xα = 0.
(cid:3)
As shown in [8, Proposition 3.16], if a C*-inductive locally convex space
A fulfills the following conditions
(r1) if xα ∈ Bα and jβα(xα) ≥ 0, β ≥ α, then xα ≥ 0;
(r2) eβ ∈ jβα(Bα),
(r3) every positive linear functional ω = lim
−→
∀α, β ∈ F, β ≥ α;
ing property
ωα on A satisfies the follow-
• if α ∈ F and ωβ(jβα(x∗
α)jβα(xα)) = 0, for some β > α and
xα ∈ Bα, then ωα(x∗
αxα) = 0;
then, A admits a faithful representation. The conditions (r1), (r2), in fact,
guarantee that A possesses sufficiently many positive linear functionals, in
the sense that for every x ∈ A+, x 6= 0 there exists a positive linear functional
ω on A such that ω(x) > 0 [8, Theorem 3.14].
Theorem 4.7. Let A be a C∗-inductive locally convex space and x = (xα) ∈
A. The following statements hold.
BOUNDED ELEMENTS OF C*-INDUCTIVE SPACES
13
(i) If x ∈ Ab, then, for every *-representation π = lim
−→
πα of A, one has
kπα(xα)kαα < ∞,
sup
α∈F
where k · kαα denote the norm of B(Hα).
(ii) Conversely, if A admits a faithful *-representation πf = lim
−→
πf
α and
kπf
α(xα)kαα < ∞,
sup
α∈F
then x ∈ Ab.
Proof. (i): For every α ∈ F, πα is a *-representation of the C*-algebra Bα.
Hence
kπα(xα)kαα ≤ kxαkα.
Thus if x ∈ Ab the statement follows immediately from the definition.
(ii): Let πf (x) = lim
−→
is a faithful representation of Bα. The *-representation πf
isomorphism of C∗-algebras, for all α ∈ F; hence
α(xα). Then, by Lemma 4.6, for every α ∈ F, πf
πf
α
α is an isometric
sup
α∈F
kxαkα = sup
α∈F
kπf
α(xα)kαα < ∞.
This proves that x is a bounded element of A.
(cid:3)
4.2. Order bounded elements. Let A be a C*-inductive locally convex
space. If x ∈ A, we put
ℜ(x) =
x + x∗
2
and ℑ(x) =
x − x∗
2i
.
Both ℜ(x) and ℑ(x) are symmetric elements of A.
Assume that A has an element u = u∗ such that kuαkα ≤ 1, for every
α ∈ F, and there exists γ ∈ F such that uβ = jβγ(eγ) ∀β ≥ γ, (eγ is the unit
of Bγ). For shortness we call the element u a pre-unit of A.
Remark 4.8. The pre-unit u ∈ A, if any, is unique. Indeed, let suppose
there is another v ∈ A satisfying the same properties as u. Then,
∃γ, γ′ ∈ F; uβ = jβγ(eγ), vβ′ = jβ′γ′(e′
γ),
∀β ≥ γ, β′ ≥ γ′
so, if δ ≥ γ, γ′, one has uλ = vλ, ∀λ ≥ δ.
Definition 4.9. Let A be a C*-inductive locally convex space with pre-unit
u. We say that x ∈ A is order bounded (with respect to u) if there exists
λ > 0 such that
−λu ≤ ℜ(x) ≤ λu
− λu ≤ ℑ(x) ≤ λu.
14
GIORGIA BELLOMONTE, SALVATORE DI BELLA, AND CAMILLO TRAPANI
Theorem 4.10. Let A be a C*-inductive locally convex space satisfying
condition (r1). Assume that A has a pre-unit u.
Then, x ∈ Ab if, and only if, x has a representative for every α ∈ F (i.e.
for every α ∈ F, there exists xα ∈ Bα such that x = φα(xα)) and x is order
bounded with respect u.
Proof. Let us assume that x = x∗ ∈ Ab. Then, x has a representative xα,
with x∗
α = xα, in every B(Hα) and λ := supα∈F kxαkα < ∞. Hence, we have
−λeα ≤ xα ≤ λeα,
∀α ∈ F,
where eα denotes the unit of Bα. By the definition of u, there exists γ ∈ F
such that uβ = jβγ(eγ ) for β ≥ γ. Hence, taking into account that the maps
jβα preserve the order, we have
−λuβ ≤ xβ ≤ λuβ,
∀β ≥ γ.
This implies that −λu ≤ x ≤ λu.
Now, let us suppose that for some λ > 0, −λu ≤ x ≤ λu. Then, there exists
γ ∈ F such that
(8)
− λuβ ≤ xβ ≤ λuβ,
∀β ≥ γ.
Let now α ∈ F. Then, there is δ ≥ α, γ such that (8) holds for δ ≥ α. Hence,
by using (r1), we conclude that
−λuα ≤ xα ≤ λuα ∀α ∈ F.
This implies that, kxαkα ≤ λ, for every α ∈ F. Thus, x ∈ Ab.
(cid:3)
From the proof of the previous theorem it follows easily that
Proposition 4.11. Let x = x∗ ∈ Ab and put
p(x) = inf{λ > 0; −λu ≤ x ≤ λu}.
Then, p(x) = kxkb.
5. C*-inductive partial *-algebras
As shown in [8], a partial multiplication in A can be defined by a family
w = {wα}, wα ∈ Bα. Let w = {wα} be a family of elements, such that each
wα ∈ B+
α and jβα(wα) = wβ, for all α, β ∈ F with β ≥ α.
Let x, y ∈ A. The partial multiplication x · y is defined by the conditions:
∃γ ∈ F : φβ(φ−1
x · y = φβ(φ−1
β (x)wβφ−1
β (y)) = φβ′(φ−1
β′ (x)wβ′ φ−1
β′ (y)), ∀β, β′ ≥ γ
β (x)wβφ−1
β (y)),
β ≥ γ.
BOUNDED ELEMENTS OF C*-INDUCTIVE SPACES
15
Then, A is an associative partial *-algebra with respect to the usual op-
erations and the above defined multiplication (see [2, Sect. 2.1.1] for the
definitions) and we will call it a C*-inductive partial *-algebra.
The partial *-algebra A has a unit e (that is, an element e which is a left-
and right universal multiplier such that x · e = e · x = x, for every x ∈ A) if,
and only if, every element wα of the family {wα} defining the multiplication
is invertible and
(9)
jβα(w−1
α ) = w−1
β ,
∀α, β ∈ F, β ≥ α.
In this case, e = φα(w−1
α ), independently of α ∈ F.
The element e is called a bounded unit if it is a bounded element of A and
kekb = 1.
Proposition 5.1. Let A be a C*-inductive partial *-algebra with the mul-
tiplication defined by a family {wα}. Assume that e = (w−1
α ) is a bounded
unit of A. Then Ab is a Banach partial *-algebra; that is, Ab[k · kb] is a
Banach space with isometric involution ∗ and there exists C ≥ 1 such that
the following inequality holds
(10)
kx · ykb ≤ Ckxkbkykb,
∀x, y ∈ Ab with x · y well-defined.
Remark 5.2. The constant C in (10) can be taken equal to 1 if w−1
α = eα,
for each α ∈ F, where eα is the unit of the C*-algebra Bα. Under the same
assumption, the norm of Ab satisfies the C*-property, which in our case reads
kx∗ · xkb = kxk2
b ,
∀x ∈ Ab with x∗ · x well-defined.
This is no longer true in the general case.
Remark 5.3. In Example 5.3 of [8] two of us tried to construct a family
{WA ∈ B(HA); A ∈ L†(D)} so that the partial multiplication defined in
B(D, D×) by the method mentioned above would reproduce the quasi *-
L
B(D, D×), L†(D)) (see Section 2). Unfortunately, the
algebra structure of (L
conclusion of that discussion is uncorrect (see, [9] for more details).
Let A be a C*-inductive partial *-algebra with the multiplication defined
by a family {wα} as above. The spaces RA and LA of the right-, respectively,
left universal multipliers (with respect to w) of A are algebras. Hence,
A0 := LA ∩ RA is a *-algebra and, thus,
(i) (A, A0) is a quasi *-algebra.
(ii) If A is endowed with τind, then the maps x 7→ x∗, x 7→ a · x, x 7→ x · b,
a, b ∈ A0 are continuous.
16
GIORGIA BELLOMONTE, SALVATORE DI BELLA, AND CAMILLO TRAPANI
It is easily seen from the very definition that, if a ∈ RA and x ∈ A+,
then a∗xa ∈ A+. Hence, if P(A) denotes the family of all positive linear
functionals on A, we have in particular ω(a∗xa) ≥ 0, for every ω ∈ P(A).
Theorem 5.4. Let A be a C*-inductive partial *-algebra with the multipli-
cation defined by a family {wα} and with pre-unit u. Assume, moreover,
that the following condition (P) holds:
(P) y ∈ A, ω(a∗ya) ≥ 0, ∀ω ∈ P(A) and a ∈ RA ⇒ y ∈ A+ ;
then, for x ∈ A, the following conditions are equivalent.
(i) x is order bounded with respect to u.
(ii) There exists γx > 0 such that
ω(a∗xa) ≤ γxω(a∗ua),
∀ω ∈ P(A),
∀a ∈ RA.
(iii) There exists γx > 0 such that
ω(b∗xa)2 ≤ γxω(a∗ua)ω(b∗ub),
∀ω ∈ P(A), ∀a, b ∈ RA.
Proof. It is sufficient to consider the case x = x∗;
(i)⇒(ii): Let ω ∈ P(A). By the hypothesis, −γu ≤ x ≤ γu, for some γ > 0;
then ω(γu − x) ≥ 0 and ω(a∗(γu − x)a) ≥ 0, ∀a ∈ RA. On the other hand,
similarly, one can show that ω(a∗(x − γu)a) ≥ 0.
(ii)⇒(i): Assume now that u is a pre-unit and there exists γx > 0 such that
ω(a∗xa) ≤ γxω(a∗ua),
∀ω ∈ P(A),
a ∈ RA.
Then
γxω(a∗ua) ± ω(a∗xa) ≥ 0 ⇒ ω(a∗(γxu ± x)a) ≥ 0,
∀ω ∈ P(A), a ∈ RA.
So, by (P), γxu ± x ≥ 0.
(i)⇒(iii): By the assumption, there exists γ > 0 such that −γu ≤ x ≤ γu.
Let ω ∈ P(A). Then, the linear functional ωa on A, defined by ωa(x) :=
ω(a∗xa), is positive. Hence, if x = x∗
i.e.,
−γωa(u) ≤ ωa(x) ≤ γωa(u);
ω(a∗xa) ≤ γω(a∗ua).
Now, let x ∈ A+, a, b ∈ RA. Let us define Ωx
ω(a, b) := ω(b∗xa). Then, it
ω is a positive sesquilinear form on RA × RA. Using
is easily checked that Ωx
BOUNDED ELEMENTS OF C*-INDUCTIVE SPACES
17
the the Cauchy-Schwartz inequality we obtain
ω(b∗xa) ≤ ω(a∗xa)1/2ω(b∗xb)1/2
≤ γω(a∗ua)1/2ω(b∗ub)1/2.
The extension to arbitrary x ∈ A goes through as in the proof of Proposition
4.3 of [8].
(iii)⇒(ii) It is trivial.
(cid:3)
The previous proof shows that if x = x∗ ∈ A is order bounded with
respect to u then
p(x) ≤ sup{ω(b∗xa); ω ∈ P(A); a, b ∈ RA; ω(a∗ua) = ω(b∗ub) = 1}.
where p(x) is the quantity defined in Proposition 4.11.
The following statement is an easy consequence of Theorem 4.11 and
Theorem 5.4.
Theorem 5.5. Let A be a C*-inductive partial *-algebra with the multi-
plication defined by a family {wα} and pre-unit u. Assume that conditions
(r1) and (P) are satisfied. For an element x ∈ A, having a representative in
every Bα, α ∈ F, the following statements are equivalent.
(i) x ∈ Ab.
(ii) x is order bounded with respect to u.
(iii) For every ω ∈ P(A)
ω(b∗xa)2 ≤ γxω(a∗ua)ω(b∗ub),
∀a, b ∈ RA.
References
[1] G.R. Allan, A spectral theory for locally convex algebras, Proc. London. Math. Soc.
15 (1965) 399 -- 421.
[2] J-P. Antoine, A. Inoue, C. Trapani, Partial *-algebras and their operator realizations,
Kluwer, Dordrecht, 2002.
[3] J-P. Antoine, C. Trapani, Partial Inner Product Spaces -- Theory and Applications,
Springer Lecture Notes in Mathematics, vol. 1986, Berlin, Heidelberg, 2009.
[4] J-P. Antoine, C. Trapani, F. Tschinke, Spectral properties of partial *-algebras
Mediterr. j. math. 7 (2010) 123 -- 142.
[5] J.-P. Antoine, C.Trapani and F. Tschinke, Bounded elements in certain topological
partial *-algebras, Studia Math. 203 (2011), 223-251.
[6] J-P. Antoine, G. Bellomonte, C. Trapani, Fully representable and *-semisimple topo-
logical partial *-algebras, Studia Mathematica, 208 (2012), 167-194.
18
GIORGIA BELLOMONTE, SALVATORE DI BELLA, AND CAMILLO TRAPANI
[7] G. Bellomonte, C. Trapani, Rigged Hilbert spaces and contractive families of Hilbert
spaces Monatshefte f. Math., 164, (2011) 271-285 (published on line on October
2010 DOI 10.1007/s00605-010-0249-1).
[8] G. Bellomonte, C. Trapani, Quasi *-algebras and generalized inductive limits of C*-
algebras, Studia Mathematica 202 (2011), 165-190.
[9] G. Bellomonte, C. Trapani, Erratum/Addendum to the paper "Quasi *-algebras and
generalized inductive limits of C*-algebras", [Studia Mathematica 202 (2011), 165-
190], to appear.
[10] G. Bellomonte, S. Di Bella, C. Trapani,Operators in Rigged Hilbert spaces: some
spectral properties, preprint, Palermo 2013.
[11] M. Fragoulopoulou, C. Trapani and S. Triolo, Locally convex quasi *-algebras with
sufficiently many *-representations, J. Math. Anal. Appl. 388 (2012), 1180-1193.
[12] K-D. Kursten, The completion of the maximal Op*-algebra on a Fr´echet domain,
Publ. Res. Inst. Math. Sci., Kyoto Univ. 22 (1986), 151 -- 175.
[13] K-D. Kursten, On algebraic properties of partial algebras, Rend. Circ. Mat. Palermo,
Ser.II, Suppl. 56 (1998), 111 -- 122.
[14] K-D. Kursten and M. Lauter, An extreme example concerning factorization products
on the Schwartz space S(Rn) Note Mat. 25 (2005/06), 31 -- 38.
[15] T.W. Palmer, Banach Algebras and the General Theory of *-Algebras, Volume 2,
Cambridge Univ. Press, Cambridge, 2001.
[16] K. Schmudgen, Unbounded operator algebras and representation theory, Birkhauser
Verlag, Basel, 1990.
[17] K. Schmudgen, A strict Positivstellensatz for theWeyl algebra, Math. Ann. 331
(2005) 779 -- 794.
[18] C. Trapani, Bounded elements and spectrum in Banach *-algebras, Studia Mathe-
matica 172 (2006) 249 -- 273.
[19] C.Trapani, Unbounded C*-seminorms, biweights and *-representations of partial *-
algebras: a review, International J. Math. Math.Sci., Volume 2006 (2006), Article
ID 79268, 34 pages.
[20] C. Trapani, Bounded and strongly bounded elements of Banach *-algebras, Contem-
porary Math. 427 (2007) 417 -- 424.
[21] C. Trapani, *-Representations, seminorms and structure properties of normed quasi
*-algebras, Studia Mathematica, 186 (2008), 47-75.
[22] C. Trapani and F. Tschinke, Partial multiplication of operators in rigged Hilbert
spaces, Integral Equations Operator Theory 51 (2005), 583 -- 600.
[23] I. Vidav, On some *regular rings, Acad. Serbe Sci. Publ. Inst. Math. 13 (1959)
73 -- 80.
Dipartimento di Matematica e Informatica, Universit`a di Palermo, I-90123
Palermo, Italy
E-mail address: [email protected]
E-mail address: [email protected]
E-mail address: [email protected]
|
1102.2012 | 2 | 1102 | 2011-12-15T16:07:46 | Mapping Cones are Operator Systems | [
"math.OA",
"quant-ph"
] | We investigate the relationship between mapping cones and matrix ordered *-vector spaces (i.e., abstract operator systems). We show that to every mapping cone there is an associated operator system on the space of n-by-n complex matrices, and furthermore we show that the associated operator system is unique and has a certain homogeneity property. Conversely, we show that the cone of completely positive maps on any operator system with that homogeneity property is a mapping cone. We also consider several related problems, such as characterizing cones that are closed under composition on the right by completely positive maps, and cones that are also semigroups, in terms of operator systems. | math.OA | math |
MAPPING CONES ARE OPERATOR SYSTEMS
NATHANIEL JOHNSTON AND ERLING STØRMER
Abstract. We investigate the relationship between mapping cones and matrix ordered
∗-vector spaces (i.e., abstract operator systems). We show that to every mapping cone
there is an associated operator system on the space of n-by-n complex matrices, and
furthermore we show that the associated operator system is unique and has a certain
homogeneity property. Conversely, we show that the cone of completely positive maps on
any operator system with that homogeneity property is a mapping cone. We also consider
several related problems, such as characterizing cones that are closed under composition
on the right by completely positive maps, and cones that are also semigroups, in terms of
operator systems.
Keywords: operator systems, mapping cones, dual cones, positive maps
AMS Subject Classifications: 15B48, 47D03, 47L99
1. Introduction
In operator theory, some of the most important families of linear maps are the positive
and k-positive maps, and their dual cones [1] of superpositive and k-superpositive maps.
These sets of maps are all specific examples of mapping cones [2], which are closed cones
of positive maps that are invariant under left and right composition by completely positive
maps -- a property of k-positive and k-superpositive maps that is easily verified.
It has recently been shown [3, 4, 5] that the k-positive and k-superpositive maps can be
seen as the completely positive maps on certain natural operator system structures. We
thus have two settings, seemingly very different, that give rise to the familiar cones of k-
positive and k-superpositive maps. A natural question that arises is whether this is simply
coincidence, or if there is indeed a fundamental link between mapping cones and operator
systems.
In this work, we show that there is indeed an extremely strong connection between
mapping cones and operator systems. In fact, we show that there is a bijection between
mapping cones and operator systems with a property that we refer to as super-homogeneity.
If we remove the super-homogeneity property, then the bijection is no longer with mapping
cones but rather with cones that are only closed under right (but not necessarily left)
composition by completely positive maps. We also answer some related questions involving
semigroup cones of positive maps.
The paper is organized as follows. In Section 2 we introduce much of our notation and
present the basics of cones of linear maps on complex matrices. In Section 3 we present
abstract operator systems and derive a simple uniqueness property in the finite-dimensional
setting that we are interested in. In Section 4 we present and prove our most general result
for right-CP-invariant cones, which shows their intimate link with operator systems, and in
Section 5 we investigate the special case of mapping cones and the kinds of operator systems
1
2
N. JOHNSTON AND E. STØRMER
that they give rise to. We close in Section 6 by exploring some properties of mapping cones
that have the additional property of being semigroups, and we see that they too can be
seen as arising from operator systems.
2. Cones of Positive Maps
If H is a (finite-dimensional) Hilbert space and L(H) is the space of linear maps on H,
then a map Φ : L(H) → L(H) is said to be positive if Φ(X) ∈ L(H)+ whenever X ∈ L(H)+,
k-positive if idk ⊗ Φ is positive, and completely positive if Φ is k-positive for all k ∈ N. If
A ∈ L(H) then the map AdA : L(H) → L(H) defined by AdA(X) ≡ A∗XA is completely
positive, and conversely every completely positive map can be written as a sum of maps of
the form AdAi for some {Ai} ∈ L(H) [6, 7].
i,j=1 eiej
∗ ⊗ eiej
∗ (ej
Given a fixed orthonormal basis {ei}n
L(H) ⊗ L(H), where E := Pn
i=1 of H, the Choi-Jamio lkowski isomorphism [8]
associates a linear map Φ : L(H) → L(H) with the operator CΦ := (idn ⊗ Φ)(E) ∈
∗ is the
outer product of ei and ej, and CΦ is called the Choi matrix of Φ). For us, it will be
useful to know that Φ is completely positive if and only if CΦ is positive, and Φ is positive
if and only if CΦ is block-positive -- i.e., (v∗ ⊗ w∗)CΦ(v ⊗ w) ≥ 0 for all v, w ∈ H. Given
a cone of positive maps C, we define CC := {CΦ : Φ ∈ C} and C† := {Φ† : Φ ∈ C}, where
Φ† : L(H) → L(H) is the adjoint map defined via the Hilbert-Schmidt inner product so
that Tr(Φ(X)Y ) = Tr(XΦ†(Y )) for all X, Y ∈ L(H).
∗ is the dual vector of ej, eiej
A mapping cone [2] is a nonzero closed cone C of positive maps from L(H) into itself
with the property that Φ ◦ Ω ◦ Ψ ∈ C whenever Ω ∈ C and Φ, Ψ : L(H) → L(H) are
completely positive. For the remainder of this work, we will generally assume all cones
to be convex, though we will still specify if the distinction is important or there is the
possibility of confusion. By linearity, it is enough that AdA ◦ Ω ◦ AdB ∈ C whenever Ω ∈ C
and A, B ∈ L(H) for a convex cone C to be a mapping cone. It will also sometimes be
useful for us to consider (not necessarily closed) cones C such that Ω ◦ Ψ ∈ C whenever
Ω ∈ C and Ψ : L(H) → L(H) is completely positive -- that is, cones that are closed under
right-composition, but not necessarily left-composition, by completely positive maps. We
will call such cones right-CP-invariant. Left-CP-invariant cones can be defined analogously,
and it is clear that C is right-CP-invariant if and only if C† is left-CP-invariant.
The dual cone C◦ of a cone C ⊆ L(H) of Hermitian operators is defined via the Hilbert-
Schmidt inner product as
C◦ := {Y ∈ L(H) : Tr(XY ) ≥ 0 for all X ∈ C}.
Similarly, the dual cone C◦ of a cone C of maps on L(H) is defined via the Choi-Jamio lkowski
isomorphism as C◦ := {Ψ : L(H) → L(H) : Tr(CΦCΨ) ≥ 0 for all Φ ∈ C}. We note that
for convex cones C ⊆ L(H), we have C◦◦ = C -- the closure of C. This fact is well-known
in convex analysis and follows easily from [9, Theorem 14.1] or [10, Theorem 3.4.3], for
example.
Throughout the rest of this work, we will associate the n-dimensional Hilbert space H
with Cn and L(H) with the space of n × n complex matrices Mn, both for simplicity and
to be consistent with standard operator system notation. Then L(Mn) denotes the set of
linear maps from Mn into itself, P(Mn) denotes the set of positive maps on Mn, Pk(Mn)
denotes the set of k-positive maps on Mn, and CP(Mn) the set of completely positive maps
MAPPING CONES ARE OPERATOR SYSTEMS
3
on Mn. We let Pn ⊆ Mn ⊗ Mn denote the cone of block-positive operators, and Sn := P ◦
n
is the cone of separable operators [1] -- operators X ∈ Mn ⊗ Mn that can be written in the
form
Yi ⊗ Zi
for positive semidefinite {Yi}, {Zi} ∈ Mn.
X =Xi
Associated to the cone of separable operators via the Choi-Jamio lkowski isomorphism is
the cone of superpositive maps S(Mn) (sometimes called entanglement-breaking maps [11]).
Similarly, the cone of k-superpositive maps is the dual cone of the cone of k-positive maps:
Sk(Mn) := Pk(Mn)◦.
We close this section with a simple lemma (which also appeared as [12, Lemma 3] with
a different proof) that allows us to relate the Choi matrices of Φ and Φ†. Note that
in the particularly important case Φ = AdA, the lemma says that (idn ⊗ AdA)(E) =
(AdAT ⊗ idn)(E), where T denotes the transpose map.
Lemma 1. Let Φ : Mn → Mn. Then (idn ⊗ Φ)(E) = ((T ◦ Φ† ◦ T ) ⊗ idn)(E).
Proof. Throughout this proof, by a vectorization vec(X) of a matrix X, we mean the vector
in Cn ⊗ Cn ∼= Cn2
obtained from X ∈ Mn by stacking the columns of X on top of each
other, starting with the leftmost column. Use the singular value decomposition to write
∗. It is easily verified that for any X ∈ Mn, vec(X T ) = F vec(X), where F
is the "swap" or "flip" operator that acts on elementary tensors as F (a ⊗ b) = b ⊗ a. The
result follows from recalling (see [7] or [13, Proposition 6.2], for example) that we can write
(cid:3)
i Y Ai, where vec(Ai) = vi and vec(Bi) = wi.
CΦ = Pi viwi
Φ(X) =Pi AiXB∗
i and Φ†(Y ) =Pi B∗
3. Operator Systems on Mn
An (abstract) operator system on Mn is a family of convex cones {Cm}∞
m=1 ⊆ Mm ⊗ Mn
that satisfy the following two properties:
• C1 = M +
• for each m1, m2 ∈ N and A ∈ Mm1,m2 we have (AdA ⊗ idn)(Cm1 ) ⊆ Cm2 .
n , the cone of positive semidefinite elements of Mn; and
Abstract operator systems can be defined more generally as matrix ordered ∗-vector spaces
on any Archimedean ∗-ordered vector space V , but the above definition with V = Mn is
much simpler and suited to our particular needs. The interested reader is directed to [14,
Chapter 13] for a more thorough treatment of general abstract operator systems. The fact
that matrix ordered ∗-vector spaces can be thought of as operator systems follows from the
work of Choi and Effros [15].
Remark 2. Abstract operator systems typically are defined with two additional require-
ments that we have not mentioned:
• Cm ∩ −Cm = {0} for each m ∈ N; and
• for every m ∈ N and X = X ∗ ∈ Mm ⊗Mn, there exists r > 0 such that rI +X ∈ Cm.
Both of these conditions follow for free from the fact that, in our setting, C1 = M +
n .
To see that the first property holds, notice that C1 ∩ −C1 = {0}, and suppose that
X ∈ Cm ∩ −Cm for some m ≥ 2. Then (AdA ⊗ idn)(X) ∈ C1 ∩ −C1 for any A ∈ Mm,1.
Because C1 ∩ −C1 = {0}, it follows that (v∗ ⊗ w∗)X(v ⊗ w) = 0 for all v ∈ Cm, w ∈ Cn. It
follows (via [16, Lemma 2.1], for example) that X = 0, so Cm ∩ −Cm = {0} for all m ∈ N.
4
N. JOHNSTON AND E. STØRMER
The second property holds because the smallest family of cones on Mn such that (AdA ⊗
idn)(Cm1) ⊆ Cm2 for all m1, m2 ∈ N are the cones of separable operators in Mm ⊗ Mn [5,
Theorem 5]. It is well-known that there always exists r > 0 such that rI + X is separable
[17], so the same r ensures that rI + X ∈ Cm.
One particularly important operator system is the one constructed by associating Mm ⊗
Mn with Mmn in the natural way and letting Cm ⊆ Mm ⊗ Mn be the cones of positive
semidefinite operators. We will denote this operator system simply by Mn, and it will be
clear from context whether we mean the operator system Mn or simply the set Mn without
regard to any family of cones. Other operator systems on Mn will be denoted like O(Mn)
in order to avoid confusion with the operator system Mn itself.
If O1(Mn) and O2(Mn) are two operator systems defined by the cones {Cm}∞
m=1 and
{Dm}∞
m=1 respectively, then a map Φ : Mn → Mn is said to be completely positive
from O1(Mn) to O2(Mn) if (idm ⊗ Φ)(Cm) ⊆ Dm for all m ∈ N. The set of maps that
are completely positive from O1(Mn) to O2(Mn) is denoted by CP(O1(Mn), O2(Mn)), or
simply CP(O(Mn)) if the target operator system equals the source operator system.
It
will often be useful for us to consider operator systems with the additional property that
(idm ⊗ AdB)(Cm) ⊆ Cm for each m ∈ N and B ∈ Mn -- a property that is equivalent to
the fact CP(Mn) ⊆ CP(O(Mn)). We will call operator systems with this property super-
homogeneous.
We now present a result that shows that operator systems on Mn are in fact characterized
completely by their nth cone. That is, there is a unique way to construct an operator system
given an appropriate cone Cn ⊆ Mn ⊗ Mn.
Proposition 3. Let Cn ⊆ Mn ⊗ Mn be a convex cone such that Sn ⊆ Cn ⊆ Pn and
(AdA ⊗idn)(Cn) ⊆ Cn for all A ∈ Mn. Then there exists a unique family of cones {Cm}m6=n
such that {Cm}∞
m=1 defines an operator system on Mn, given by
Cm :=(cid:8)Xi
(AdAi ⊗ idn)(X) : Ai ∈ Mn,m ∀ i, X ∈ Cn(cid:9).
Furthermore, the operator system is super-homogeneous if and only if (idn ⊗AdB)(Cn) ⊆ Cn
for all B ∈ Mn.
Proof. We first prove that the family of convex cones given by the proposition do indeed
define an operator system. We first show that (AdB ⊗ idn)(Y ) ∈ Cm2 for any m1, m2 ∈ N,
Y ∈ Cm1, and B ∈ Mm1,m2. This is true from the definition of Cm if m1 = n. If m1 6= n then
write Y = Pi(AdAi ⊗ idn)(X) for some X ∈ Cn and {Ai} ⊂ Mn,m1. Then AiB ∈ Mn,m2
for all i, so
We now show that C1 = M +
n , and similarly vv∗ ⊗ X ∈ Pn if and only if X ∈ M +
X ∈ M +
if and only if X ∈ M +
where we have identified R+ ⊗ M +
noting that if X ∈ C1 and v ∈ Cn then vv∗ ⊗ X ∈ Cn, so X ∈ M +
n .
C1 ⊆ M +
n . For any v ∈ Cn, note that vv∗ ⊗ X ∈ Sn if and only if
n . It follows that vv∗ ⊗ X ∈ Cn
n ,
n . The opposite inclusion follows simply from
It follows that
n . Then C1 ⊇ {(AdA ⊗ idn)(vv∗ ⊗ X) : A ∈ Mn,1, X ∈ M +
n(cid:9) = M +
n , so C1 = M +
n , so the cones {Cm}∞
m=1 define an operator system on Mn.
n with M +
(AdB ⊗ idn)(Y ) =Xi
(AdAiB ⊗ idn)(X) ∈ Cm2.
MAPPING CONES ARE OPERATOR SYSTEMS
5
To prove uniqueness, assume that there exists another family of cones {Dm}∞
m=1 that
define an operator system, such that Dn = Cn. It is clear that Cm ⊆ Dm for all m ∈ N, so
we only need to prove the other inclusion. If m ≤ n, let X ∈ Dm and let V : Cm → Cn
be an isometry (i.e., V ∗V = I). Then Y := (AdV ∗ ⊗ idn)(X) ∈ Dn = Cn, so X =
(AdV ⊗ idn)(Y ) ∈ Cm. Thus Dm ⊆ Cm, so Dm = Cm for m ≤ n. If m > n then we recall
from [4, Section 2.3] the k-super minimal and k-super maximal operator system structures.
In particular, it was shown that if two operator systems on Mn, defined by cones {Cm}∞
and {Dm}∞
m ∈ N. See also [5, Section 4].
m=1
m=1 respectively, are such that Cm = Dm for 1 ≤ m ≤ n, then Cm = Dm for all
The "only if" direction of the final claim is trivial from the definition of super-homogeneity,
and the "if" direction follows easily from the fact that (AdA ⊗ idn) and (idm ⊗ AdB) com-
mute. This completes the proof.
(cid:3)
We close this section with a result that shows that to determine complete positivity of a
map from one operator system on Mn to another, it is enough to look at the action of that
map on the nth cone of the operator systems.
Corollary 4. Let Φ : Mn → Mn and let O1(Mn) and O2(Mn) be operator systems defined
by families of cones {Cm}∞
m=1, respectively. Then Φ ∈ CP(O1(Mn), O2(Mn))
if and only if (idn ⊗ Φ)(Cn) ⊆ Dn.
m=1 and {Dm}∞
Proof. The "only if" implication follows trivially from the definition of CP(O1(Mn), O2(Mn)).
For the "if" implication, suppose (idn ⊗ Φ)(Cn) ⊆ Dn. Fixing m ∈ N arbitrarily and ap-
plying Pi AdAi ⊗ idn for {Ai} ∈ Mn,m to both sides then gives
(idm ⊗ Φ)(Cm) = [{Ai}∈Mn,m(Xi
⊆ [{Ai}∈Mn,m(Xi
(AdAi ⊗ Φ)(Cn))
(AdAi ⊗ idn)(Dn))
= Dm,
where both of the above equalities follow from the form of the cones {Cm}∞
guaranteed by Proposition 3.
proof.
m=1
It follows that Φ ∈ CP(O1(Mn), O2(Mn)), completing the
(cid:3)
m=1 and {Dm}∞
4. Right-CP-Invariant Cones as Operator Systems
In this section we establish a link between right-CP-invariant cones and operator systems.
Our first result is in the same vein as some known results on mapping cones such as [18,
Theorem 1] and [19, Theorem 1]. Here we prove an analogous statement for cones that are
just right-CP-invariant.
Proposition 5. Let C ⊆ L(Mn) be a right-CP-invariant cone. Then Ψ† ◦ Φ ∈ CP(Mn) for
all Φ ∈ C if and only if Ψ ∈ C◦.
6
N. JOHNSTON AND E. STØRMER
Proof. To prove the "only if" implication, suppose Ψ ∈ L(Mn) and Ψ† ◦ Φ ∈ CP(Mn) for
all Φ ∈ C. Then CΨ†◦Φ ∈ (Mn ⊗ Mn)+ so
0 ≤ Tr(ECΨ†◦Φ) = Tr(E(idn ⊗ (Ψ† ◦ Φ))(E))
= Tr((idn ⊗ Ψ)(E)(idn ⊗ Φ)(E)) = Tr(CΨCΦ) ∀ Φ ∈ C,
i,j=1 eiej
∗ ⊗ eiej
∗. It follows that Ψ ∈ C◦. It is perhaps worth
where we recall that E :=Pn
noting that the proof of this implication did not make use of right-CP-invariance of C.
To see why the "if" implication holds, assume Ψ ∈ C◦. Then, because C is right-CP-
invariant, it follows that for any Φ ∈ C and Ω ∈ CP(Mn) we have Φ ◦ Ω ∈ C so
0 ≤ Tr(CΨCΦ◦Ω) = Tr((idn ⊗ Ψ)(E)(idn ⊗ (Φ ◦ Ω))(E))
= Tr((idn ⊗ (Φ† ◦ Ψ))(E)(idn ⊗ Ω)(E)) = Tr(CΦ†◦ΨCΩ).
It follows via the Choi-Jamio lkowski isomorphism that CΦ†◦Ψ ∈ (Mn ⊗ Mn)+, so Φ† ◦ Ψ ∈
CP(Mn). Then (Φ† ◦ Ψ)† = Ψ† ◦ Φ ∈ CP(Mn), completing the proof.
(cid:3)
It is not difficult to verify that if O(Mn) is any operator system, then CP (Mn, O(Mn))
is a right-CP-invariant cone. Similarly, CP (O(Mn), Mn) is easily seen to be a closed left-
CP-invariant cone. The main result of this section shows that these properties actually
characterize the possible cones of completely positive maps to and from Mn, and furthermore
that these cones uniquely determine O(Mn).
Recall that P(Mn) denotes the cone of positive maps on Mn, S(Mn) denotes the cone
of superpositive maps on Mn, and CC denotes the cone of Choi matrices of maps from the
cone C.
Theorem 6. Let C ⊆ L(Mn) be a convex cone. The following are equivalent:
(1) C is right-CP-invariant with S(Mn) ⊆ C ⊆ P(Mn).
(2) There exists an operator system O1(Mn), defined by cones {Cm}∞
m=1, such that
CC = Cn.
(3) There exists an operator system O2(Mn) such that C = CP(Mn, O2(Mn)).
(4) There exists an operator system O3(Mn) such that (C◦)† = CP(O3(Mn), Mn).
Furthermore, O1(Mn) = O2(Mn) and is uniquely determined by C, and O3(Mn) is uniquely
determined by C and can be chosen so that O3(Mn) = O1(Mn).
Proof. We prove the result by showing that (1) ⇔ (2), (2) ⇔ (3), and (2) ⇔ (4).
To see that (1) ⇒ (2), define Cn := CC. If A ∈ Mn and Φ ∈ C then
(1)
(AdA ⊗ idn)(CΦ) = (AdA ⊗ Φ)(E) = (idn ⊗ (Φ ◦ AdAT ))(E) ∈ Cn,
where the second equality comes from Lemma 1 and the inclusion comes from the fact that
C is right-CP-invariant, so Φ ◦ AdAT ∈ C. The implication (1) ⇒ (2) and uniqueness of O1
then follow from Proposition 3. The reverse implication (2) ⇒ (1) also follows from the
string of equalities (1), but this time we use the fact that Cn is a cone defining an operator
system to get the inclusion. The fact that S(Mn) ⊆ C ⊆ P(Mn) follows from the fact that
for the minimal operator system on Mn, Cn is the cone of block-positive operators and for
the maximal operator system on Mn, Cn is the cone of separable operators [5, Theorem 5].
To see that (2) ⇒ (3), let O2(Mn) = O1(Mn). We then have to show that if CC = Cn,
then C = CP(Mn, O1(Mn)). We already showed that (2) ⇒ (1), so we know that C is
MAPPING CONES ARE OPERATOR SYSTEMS
7
right-CP-invariant. So if Φ ∈ C then for any X ∈ (Mn ⊗ Mn)+ there exists Ψ ∈ CP(Mn)
such that
(idn ⊗ Φ)(X) = (idn ⊗ (Φ ◦ Ψ))(E) ∈ Cn,
where the inclusion comes from C being right CP-invariant. It follows via Corollary 4 that
Φ ∈ CP(Mn, O1(Mn)), so C ⊆ CP(Mn, O1(Mn)). To see the opposite inclusion, simply note
that if Φ ∈ CP(Mn, O1(Mn)) then, because E ∈ (Mn ⊗ Mn)+, we have CΦ = (idn ⊗ Φ)(E) ∈
Cn = CC, so Φ ∈ C. It follows that C = CP(Mn, O1(Mn)).
To establish uniqueness of O2 (and simultaneously prove (3) ⇒ (2)), suppose that the
cones {Dm}∞
m=1 define an operator system O2(Mn) such that C = CP(Mn, O2(Mn)). Be-
cause E ∈ (Mn ⊗ Mn)+, we again have that (idn ⊗ Φ)(E) ∈ Dn for any Φ ∈ C, so CC ⊆ Dn.
On the other hand by the equivalence of (1) and (2), Dn = CC′ for some right-CP-invariant
cone C′. If Φ ∈ C′ then for any X ∈ (Mn ⊗ Mn)+ there exists Ψ ∈ CP(Mn) such that
(idn ⊗ Φ)(X) = (idn ⊗ (Φ ◦ Ψ))(E) ∈ Dn,
where the inclusion comes from C′ being right CP-invariant.
It follows via Corollary 4
that C′ ⊆ CP(Mn, O2(Mn)). Then C ⊆ C′ ⊆ CP(Mn, O2(Mn)) = C, so C = C′ and hence
Cn = Dn. Uniqueness now follows from Proposition 3.
The proof that (2) ⇔ (4) mimics the proof that (2) ⇔ (3) and makes use of the fact that
Ψ† ◦ Φ ∈ CP(Mn) for all Φ ∈ C if and only if Ψ ∈ C◦ (Proposition 5). To see that (2) ⇒ (4),
let O3(Mn) = O1(Mn). Then for any Ψ ∈ C◦ and Φ ∈ C we have Ψ† ◦ Φ ∈ CP(Mn), so
CΨ†◦Φ ∈ (Mn ⊗ Mn)+. It follows that (idn ⊗ Ψ†)(Cn) ⊆ (Mn ⊗ Mn)+. Corollary 4 implies
that Ψ† ∈ CP(O3(Mn), Mn), so (C◦)† ⊆ CP(O3(Mn), Mn). The opposite inclusion follows
by simply reversing this argument.
Uniqueness of O3 (up to closure) and the implication (4) ⇒ (2) follow similarly by the
fact that Ψ† ∈ CP(OC(Mn), Mn) if and only if Ψ† ◦ Φ ∈ CP(Mn) for all Φ ∈ C if and only if
Ψ ∈ C◦, where OC(Mn) is an operator system with its nth cone Cn := CC.
(cid:3)
The equivalence of conditions (1) and (2) in Theorem 6 can be seen as providing a bijection
between right-CP-invariant cones and operator systems on Mn. Given an operator system
O(Mn) defined by cones {Cm}∞
m=1, the associated right-CP-invariant cone is given via the
maps associated to Cn via the Choi-Jamio lkowski isomorphism. In the other direction, given
a right-CP-invariant cone, the associated operator system gets its nth cone from the Choi-
Jamio lkowski isomorphism and then gets its remaining cones via the construction given in
Proposition 3.
5. Mapping Cones as Operator Systems
Before introducing the main results of this section, we present a lemma that shows that
the largest cone of completely positive maps between any two operator systems on Mn is
the cone of positive maps -- a result that follows from recent work on minimal and maximal
operator systems [3, 5].
Lemma 7. Let O1(Mn) and O2(Mn) be operator systems. Then CP(O1(Mn), O2(Mn)) ⊆
P(Mn).
8
N. JOHNSTON AND E. STØRMER
Proof. Let O1(Mn) and O2(Mn) be defined by the families of cones {Cm}∞
m=1,
respectively. Let Φ ∈ L(Mn) be such that Φ /∈ P(Mn). Because the smallest family of cones
defining an operator system on Mn are the separable operators and the largest such family
of cones are the block-positive operators [5, Theorem 5], we know that I ⊗ X ∈ Cn for all
X ∈ M +
n such that
Φ(X) /∈ M +
n . It is then easily verified that I ⊗ Φ(X) /∈ Pn, so (idn ⊗ Φ)(I ⊗ X) /∈ Dn. It
follows that Φ /∈ CP(O1(Mn), O2(Mn)), so CP(O1(Mn), O2(Mn)) ⊆ P(Mn).
(cid:3)
n and Dn ⊆ Pn. Because Φ /∈ P(Mn), there exists a particular X ∈ M +
m=1 and {Dm}∞
The following result shows how the bijection inroduced by Theorem 6 works when the
right-CP-invariant cone is in fact a mapping cone -- in this situation the associated operator
system is super-homogeneous.
Corollary 8. Let C ⊆ L(Mn) be a closed, convex cone. The following are equivalent:
(1) C is a mapping cone.
(2) There exists a super-homogeneous operator system O1(Mn), defined by cones
{Cm}∞
m=1, such that CC = Cn.
(3) There exists a super-homogeneous operator system O2(Mn) such that
C = CP(Mn, O2(Mn)).
(4) There exists a super-homogeneous operator system O3(Mn) such that
(C◦)† = CP(O3(Mn), Mn).
(5) There exist super-homogeneous operator systems O4(Mn) and O5(Mn) such that
C = CP(O4(Mn), O5(Mn)).
Furthermore, O1(Mn) = O2(Mn) = O3(Mn) and is uniquely determined by C.
Proof. The equivalence of (1), (2), (3), and (4) (and uniqueness of the corresponding op-
erator systems) follows immediately from the corresponding statements of Theorem 6 and
the fact that C is left-CP-invariant if and only if (idn ⊗ AdB)(CC) ⊆ CC, which then gives
super-homogeneity of the corresponding operator system via Proposition 3.
Because Mn is a super-homogeneous operator system, it is clear that (3) ⇒ (5). All
that remains to do is prove that (5) ⇒ (1). To this end, let O4(Mn) and O5(Mn) be
super-homogeneous operator systems defined by families of cones {Cm}∞
m=1,
respectively. By the equivalence of conditions (1) and (2), we know that there exist mapping
cones C′ and C′′ such that CC′ = Cn and CC′′ = Dn. By Corollary 4, (idn ⊗ Φ)(Cn) ⊆ Dn
if and only if Φ ∈ CP(O4(Mn), O5(Mn)), so it follows that Φ ◦ Ψ ∈ C′′ for all Ψ ∈ C′ if
and only if Φ ∈ CP(O4(Mn), O5(Mn)). Right-CP-invariance of CP(O4(Mn), O5(Mn)) now
follows from left-CP-invariance of C′ and left-CP-invariance of CP(O4(Mn), O5(Mn)) follows
from left-CP-invariance of C′′. The fact that CP(O4(Mn), O5(Mn)) ⊆ P(Mn) follows from
Lemma 7.
(cid:3)
m=1 and {Dm}∞
It is natural at this point to consider well-known mapping cones and ask what are the
corresponding operator systems via the bijection of Corollary 8. The mapping cone of stan-
dard completely positive maps CP(Mn) of course corresponds to the "naive" operator system
with positive cones equal to the cones of positive semidefinite operators. It was shown in [3]
that S(Mn) = CP(Mn, OM AX(Mn)), where OM AX(Mn) is the maximal operator system
structure on Mn. It follows that the operator system associated with the mapping cone
S(Mn) is OM AX(Mn), and the cones that define OM AX(Mn) are exactly the cones of
separable operators. It was similarly shown that S(Mn) = CP(OM IN (Mn), Mn), where
MAPPING CONES ARE OPERATOR SYSTEMS
9
OM IN (Mn) is the minimal operator system structure on Mn. It follows from condition (4)
of Corollary 8 (and the fact that S(Mn) = (P(Mn)◦)†) that the operator system associated
with the mapping cone P(Mn) is OM IN (Mn), and the cones that define OM IN (Mn) are
the cones of block-positive operators.
It was shown in [5] that if OM INk(Mn) and OM AXk(Mn) denote the super k-minimal
and super k-maximal operator systems on Mn [4], respectively, then we have that Pk(Mn) =
CP(Mn, OM INk(Mn)) and Sk(Mn) = CP(Mn, OM AXk(Mn)). Thus the operator systems
associated with the mapping cones Pk(Mn) and Sk(Mn) are OM INk(Mn) and OM AXk(Mn),
respectively. Finally, consider the mapping cone of completely co-positive maps {Φ◦T : Φ ∈
CP(Mn)}. It is not difficult to see that the associated operator system is the one defined by
the cones of operators with positive partial transpose -- i.e., the operators X ∈ Mm ⊗ Mn
such that (idm ⊗ T )(X) ≥ 0.
We close this section by considering what Corollary 8 says in the case when the mapping
cone C is symmetric -- that is, when T ◦ Φ ◦ T ∈ C and Φ† ∈ C whenever Φ ∈ C. The concept
of symmetric mapping cones was seen to be important in [12], and it is worth noting that
all of the specific mapping cones considered so far, such as the cones of k-positive and
completely co-positive maps, are in fact symmetric. It will be useful for us to define a linear
operator F ∈ Mn ⊗ Mn by F (v ⊗ w) = w ⊗ v and extending linearly. The operator F is
sometimes called the swap or flip operator, and we observe that F = F T .
Theorem 9. Let C ⊆ L(Mn) be a convex mapping cone and let O(Mn) be the operator
system, defined by cones {Cm}∞
m=1, associated to C via the bijection of Corollary 8. Then C
is symmetric if and only if Cn is closed under the transpose map and the map X 7→ F XF .
Proof. The proof relies on Lemma 1 which tells us that CT ◦Φ†◦T = F CΦF , and [19, Lemma
4] which tells us that CT ◦Φ◦T = C T
Φ F .
It then follows immediately that T ◦ Φ ◦ T ∈ C whenever Φ ∈ C if and only if Cn (which
equals CC) is closed under the transpose map T . Similarly, Φ† ∈ C whenever Φ ∈ C if and
only if Cn is closed under the map X 7→ F X T F . The result follows.
(cid:3)
Φ . Combining these two results shows that CΦ† = F C T
6. Semigroup Cones as Operator Systems
Theorem 6 and Corollary 8 provide characterizations of completely positive maps to and
from Mn, and completely positive maps between two different super-homogeneous operator
systems on Mn. However, they say nothing about completely positive maps from a super-
homogeneous operator system back into itself. Toward deriving a characterization for this
situation, we will say that a cone C ⊆ L(Mn) is a semigroup if it is closed under composition
-- i.e., if Φ ◦ Ψ ∈ C for all Φ, Ψ ∈ C. Notice that many of the standard examples of mapping
cones, such as the k-positive maps and the k-superpositive maps, are semigroups (however,
the cone of completely co-positive maps is not).
The following proposition is a generalization of the fact that Φ is k-positive if and only if
Φ ◦ Ψ is k-superpositive for all k-superpositive Ψ [1, Theorem 3.8]. Note that it is similar to
Proposition 5, but by using the fact that C is a semigroup instead of just right-CP-invariant
or a mapping cone we are able to show that Φ† ◦ Ψ ∈ C◦, not just that Φ† ◦ Ψ ∈ CP(Mn).
Proposition 10. Let C ⊇ CP(Mn) be a closed convex cone semigroup. Then Φ ∈ C if and
only if Φ† ◦ Ψ ∈ C◦ for all Ψ ∈ C◦.
10
N. JOHNSTON AND E. STØRMER
Proof. To show the "only if" direction, it is enough to show that Tr(CΦ†◦ΨCΩ) ≥ 0 for all
Ω ∈ C. To this end, simply note that
Tr(CΦ†◦ΨCΩ) = Tr(CΨCΦ◦Ω) ≥ 0,
where the final inequality follows from the fact that Φ, Ω ∈ C so Φ ◦ Ω ∈ C.
To see the "if" direction, suppose Φ† ◦ Ψ ∈ C◦ for all Ψ ∈ C◦. Then, because idn ∈
CP(Mn) ⊆ C, we have
It follows that Φ ∈ C◦◦ = C.
(cid:3)
0 ≤ Tr(CΦ†◦ΨE) = Tr(CΨCΦ) ∀ Ψ ∈ C◦.
If O(Mn) is an operator system defined by cones {Cm}∞
m=1, then the dual cones {C ◦
m}∞
m=1
define an operator system as well, which we will denote O◦(Mn). For simplicity, we will only
consider this operator system as a family of dual cones, in keeping with our focus throughout
the preceding portion of the paper, and not the associated dual operator space structure.
The interested reader is directed to [20] for a more thorough treatment of dual operator
systems. It is easily verified that O(Mn) is super-homogeneous if and only if O◦(Mn) is
super-homogeneous, and the "naive" operator system on Mn is easily seen to be self-dual:
M ◦
n = Mn. By the duality of the cones of k-positive maps and k-superpositive maps we
know that OM IN ◦
k (Mn) = OM AXk(Mn) and OM AX ◦
k (Mn) = OM INk(Mn).
We now consider what types of cones can be completely positive from a super-homogeneous
operator system back into itself. By using [5, Theorem 5] and the fact that Pk(Mn) is a
semigroup, it is not difficult to see that CP(OM INk(Mn)) = Pk(Mn). By using [1, Theorem
3.8] we can similarly see that CP(OM AXk(Mn)) = Pk(Mn), so we can't possibly hope for
a uniqueness result as strong as that of Theorem 6 or Corollary 8 in this setting. Nonethe-
less, we have the following result, which shows that duality plays a strong role here and the
fact that CP(OM INk(Mn)) = CP(OM AXk(Mn)) follows from the duality of OM INk(Mn)
and OM AXk(Mn). Furthermore, there is a unique operator system that gives the cone
CP(O(Mn)) that is "large enough" to contain (Mn ⊗ Mn)+ as a subset of its nth cone -- in
this case it is OM INk(Mn).
Theorem 11. Let C ⊆ L(Mn) be a convex cone. The following are equivalent:
†
(1) C is a semigroup cone with CP(Mn) ⊆ C ⊆ P(Mn).
(2) There exists a super-homogeneous operator system O(Mn) such that C = CP(O(Mn)).
and O(Mn) can be chosen so that its nth cone
Additionally, CP(O◦(Mn)) = CP(O(Mn))
Cn = CC. Furthermore, O(Mn) is unique up to the condition (Mn ⊗ Mn)+ ⊆ Cn.
Proof. We first prove that (2) ⇒ (1). Let {Cm}∞
m=1 be the cones associated with the
operator system O(Mn). If X ∈ Cm and Φ, Ψ ∈ CP(O(Mn)) then (idm ⊗ Φ)(X) ∈ Cm. But
then applying idm ⊗ Ψ shows (idm ⊗ (Ψ ◦ Φ))(X) ∈ Cm as well, so it follows that Ψ ◦ Φ ∈
CP(O(Mn)) and thus CP(O(Mn)) is a semigroup. Because O(Mn) is super-homogeneous,
we know that AdB ∈ CP(O(Mn)) for all B ∈ Mn, and so CP(Mn) ⊆ CP(O(Mn)). To see
that CP(O(Mn)) ⊆ P(Mn), simply use Lemma 7.
To see that (1) ⇒ (2), we argue much as we did in Theorem 6.
It is clear, via the
Choi-Jamio lkowski isomorphism, that Sn ⊆ CC ⊆ Pn. Now note that C is left- and right-
CP-invariant (but perhaps not a mapping cone because it may not be closed) because
MAPPING CONES ARE OPERATOR SYSTEMS
11
Φ ◦ Ψ ∈ C for any Φ ∈ C and Ψ ∈ CP(Mn) ⊆ C (and similarly for composition on the left
by Ψ ∈ CP(Mn)). Thus, if A ∈ Mn and Φ ∈ C then
(AdA ⊗ AdB)(CΦ) = (AdA ⊗ (AdB ◦ Φ))(E) = (idn ⊗ (AdB ◦ Φ ◦ AdAT ))(E) ∈ CC,
where the second equality comes from Lemma 1. It follows from Proposition 3 that there
exists a super-homogeneous operator system O(Mn), defined by cones {Cm}∞
m=1, such that
Cn = CC. Because C is a semigroup, it follows that (idn ⊗ (Φ ◦ Ψ))(E) ∈ Cn for any
Φ, Ψ ∈ C. Then (idn ⊗ Φ)(CΨ) ∈ Cn, so (idn ⊗ Φ)(Cn) ⊆ Cn, which implies C ⊆ CP(O(Mn))
by Corollary 4. To see the other inclusion, note that idn ∈ CP(Mn), so idn ∈ C. It follows
that (idn ⊗ idn)(E) = E ∈ Cn. Thus, if Φ ∈ CP(O(Mn)) then (idn ⊗ Φ)(E) ∈ Cn = CC, so
Φ ∈ C, which implies that C = CP(O(Mn)).
To see the claim about CP(O◦(Mn)), suppose that CP(Mn) ⊆ C ⊆ P(Mn) is a closed con-
vex cone semigroup. Then for any Φ, Ψ ∈ C◦ ⊆ CP(Mn) and Ω ∈ C we have Tr(CΦ◦ΨCΩ) =
Tr(CΨCΦ†◦Ω). We know from Proposition 10 that Φ† ◦ Ω = (Ω† ◦ Φ)† ∈ (C◦)† ⊆ CP(Mn)† =
CP(Mn) ⊆ C. It follows that Tr(CΨCΦ†◦Ω) ≥ 0, so Φ ◦ Ψ ∈ C◦, which implies that C◦ is also
a semigroup.
†
†
†
Now by repeating our argument from earlier, we see from Proposition 3 that there is
an operator system on Mn defined by the cone Cn := CC◦ = C ◦
C, and this is the dual
operator system O◦(Mn) of the operator system defined by CC. For any Φ ∈ C◦◦, Ψ ∈ C◦,
we have (idn ⊗ Φ†)(CΨ) = CΦ†◦Ψ ∈ CC◦ by Proposition 10. It follows via Corollary 4 that
⊆ CP(O◦(Mn)). To see the other inclusion, suppose Φ ∈ CP(O◦(Mn)). Then
(C◦◦)† = C
Φ ◦ Ψ ∈ C◦ for all Ψ ∈ C◦, so Proposition 10 tells us that Φ ∈ (C◦◦)† = C
. It follows that
C
= CP(O◦(Mn)).
Finally, to see the uniqueness condition, suppose that the cones {Dm}∞
m=1 define an
operator system O2(Mn) such that C = CP(O(Mn)) = CP(O2(Mn)), where O(Mn) is the
operator system with nth cone Cn := CC already introduced. We furthermore require that
(Mn ⊗ Mn)+ ⊆ Dn, and in particular that E ∈ Dn. Then (idn ⊗ Φ)(E) ∈ Dn for any Φ ∈ C,
so CC ⊆ Dn. On the other hand by the equivalence of (1) and (2), Dn = CC′ for some
semigroup cone C′. If Φ ∈ C′ then for any X ∈ Dn there exists Ψ ∈ C′ such that
(idn ⊗ Φ)(X) = (idn ⊗ (Φ ◦ Ψ))(E) ∈ Dn,
where the inclusion comes from C′ being a semigroup.
It follows via Corollary 4 that
C′ ⊆ CP(O2(Mn)). Then C ⊆ C′ ⊆ CP(O2(Mn)) = C, so C = C′ and hence Cn = Dn.
Uniqueness now follows from Proposition 3.
(cid:3)
It is worth noting that if C is closed and condition (1) of Theorem 11 holds, then C
is necessarily a mapping cone. It follows that if O(Mn) is a super-homogeneous operator
system defined by closed cones then CP(O(Mn)) is always a mapping cone (which can also
be seen from Corollary 8), although the converse does not hold. That is, there exist mapping
cones C such that there is no operator system O(Mn) with C = CP(O(Mn)) -- the simplest
example being the mapping cone of completely co-positive maps.
Acknowledgements. Thanks are extended to Vern Paulsen for valuable suggestions and
comments on an early draft. N.J. was supported by an NSERC Canada Graduate Scholar-
ship and the University of Guelph Brock Scholarship.
12
N. JOHNSTON AND E. STØRMER
References
[1]
L. Skowronek, E. Størmer, and K.
Math. Phys. 50 (2009), 062106.
Zyczkowski, Cones of positive maps and their duality relations, J.
[2] E. Størmer, Extension of positive maps into B(H), J. Funct. Anal. 66 (1986), 235 -- 254.
[3] V. Paulsen, I. Todorov, M. Tomforde, Operator system structures on ordered spaces, Proc. Lond. Math.
Soc. (2010), doi:10.1112/plms/pdq011.
[4] B. Xhabli, Universal operator system structures on ordered spaces and their applications, Ph.D. Thesis,
University of Houston (2009).
[5] N. Johnston, D. W. Kribs, V. I. Paulsen, and R. Pereira, Minimal and maximal operator spaces and
operator systems in entanglement theory, J. Funct. Anal. 260 8 (2011), 2407 -- 2423.
[6] M. Nielsen and I. Chuang, Quantum computation and quantum information, Cambridge University
Press (2000).
[7] M.-D. Choi, Completely positive linear maps on complex matrices, Linear Algebra Appl. 10 (1975),
285-290.
[8] A. Jamio lkowski, Linear transformations which preserve trace and positive semidefiniteness of operators,
Rep. Math. Phys. 3 (1972).
[9] R. Rockafellar, Convex analysis, Princeton University Press (1997).
[10] C. J. Goh and X. Q. Yang, Duality in optimization and variational inequalities, Taylor & Francis,
London (2002).
[11] M. Horodecki, P. W. Shor, and M. B. Ruskai, General entanglement breaking channels, Rev. Math.
Phys. 15 (2003), 629 -- 641.
[12] E. Størmer, Mapping cones of positive maps, Math. Scand. 108 (2011), 223 -- 232.
[13] J. Watrous, Theory
lecture notes,
http://www.cs.uwaterloo.ca/~watrous/lecture-notes.html (2004).
of
quantum information
published electronically
at
[14] V. I. Paulsen, Completely bounded maps and operator algebras, Cambridge University Press, Cambridge
(2003).
[15] M.-D. Choi and E. G. Effros, Injectivity and operator spaces, J. Funct. Anal. 24 (1977), 156-209.
[16] N. Johnston, Characterizing operations preserving separability measures via linear preserver problems,
Linear Multilinear Algebra 59 (2011), 1171 -- 1187.
[17] L. Gurvits and H. Barnum, Size of the separable neighborhood of the maximally mixed bipartite quantum
[18]
state, Los Alamos National Laboratory unclassified technical report LAUR (2002) 02-2414.
L. Skowronek, Cones with a mapping cone symmetry in the finite-dimensional case, Linear Algebra
Appl. 435 (2011), 361 -- 370.
[19] E. Størmer, Duality of cones of positive maps, Munster J. Math. 2 (2009), 299 -- 310.
[20] D. Blecher and B. Magajna, Dual operator systems, Bull. London Math. Soc. (2010), doi:
10.1112/blms/bdq103.
Department of Mathematics and Statistics, University of Guelph, Guelph, Ontario N1G
2W1, Canada
E-mail address: [email protected]
Department of Mathematics, University of Oslo, P.O. Box 1053 Blindern, NO-0316 Oslo,
Norway
E-mail address: [email protected]
|
1306.0945 | 1 | 1306 | 2013-06-04T23:53:02 | Notes on extremality of the Choi map | [
"math.OA",
"quant-ph"
] | It is widely believed that the Choi map generates an extremal ray in the cone $\mathcal P(M_3)$ of all positive linear maps between $C^*$-algebra $M_3$ of all $n\times n$ matrices over the complex field. But the only proven fact is that the Choi map generates the extremal ray in the cone of all positive linear map preserving all real symmetric $3\times 3$ matrices. In this note, we show that the Choi map is indeed extremal in the cone $\mathcal P(M_3)$. We also clarify some misclaims about the correspondence between positive semi-definite biquadratic real forms and postive linear maps, and discuss possible positive linear maps which coincide with the Choi map on symmetric matrices. | math.OA | math | NOTES ON EXTREMALITY OF THE CHOI MAP
KIL-CHAN HA
Abstract. It is widely believed that the Choi map generates an extremal ray in the
cone P(M3) of all positive linear maps between C ∗-algebra M3 of all n × n matrices
over the complex field. But the only proven fact is that the Choi map generates the
extremal ray in the cone of all positive linear map preserving all real symmetric 3 × 3
matrices. In this note, we show that the Choi map is indeed extremal in the cone
P(M3). We also clarify some misclaims about the correspondence between positive
semi-definite biquadratic real forms and postive linear maps, and discuss possible
positive linear maps which coincide with the Choi map on symmetric matrices.
3
1
0
2
n
u
J
4
]
.
A
O
h
t
a
m
[
1
v
5
4
9
0
.
6
0
3
1
:
v
i
X
r
a
1. Introduction
Let Mn be the C ∗-algebra of all n × n matrices over the complex field. Because the
convex structure of the positive cone P(Mn) of all positive linear maps between Mn is
highly complicated even in lower dimensions, it would be very useful to find extreme
rays of this cone. Another approach to understand the convex structure of P(Mn) is
to considered the possibility of decomposition of P(Mn) into subcones. For example, a
positive linear map between matrix algebra is said to be decomposable if it is the sum
of a completely positive linear map and a completely copositive linear map.
In the sixties, it was shown that every positive linear map in P(M2) is decomposable,
and all extreme points of the convex set of unital positive linear maps in P(M2) had
been found [10]. The first example of indecomposable positive linear map was given
by Choi [1, 2]. This Choi map is defined by
Φ(X) =
−x21
−x31
x11 + x33
−x12
x22 + x11
−x13
−x23
,
−x32
x33 + x22
where X = (xij) ∈ M3. It is widely believed [8, 7, 6, 9] that the Choi map generates an
extremal ray in P(M3). It is not as trivial as one may think. We note that extremality
of some variants of Choi map can be confirmed from their exposedness [4, 5]. But the
only proven fact on the Choi map is that, for x = (x1, x2, x3)t, y = (y1, y2, y3)t ∈ R3,
the corresponding real form ytΦ(xxt)y is extremal in the convex cone of all positive
semi-definite biquadratic real forms [3]. Let Sn be the real vector space of all n × n real
Date: August 2, 2021.
1991 Mathematics Subject Classification. 15A30, 46L05.
Key words and phrases. Positive semi-definite biquadratic real form, Positive linear map, Convex
cone, Extremal.
partially supported by NRFK 2012-0002600.
1
symmetric matrices in Mn. Then we see that the restriction map ΦS3 of Φ is extremal
in the cone P(S3) of all positive linear map between S3.
We note that the extremality of ΦS3 in the cone P(S3) does not imply the ex-
tremality of Φ in P(M3). To explain this, we consider a positive linear map Ψ1 defined
by
(1)
Ψ1 =
1
2
(Φ + Φ ◦ t).
Since Φ(S3) ⊂ S3, we see that Ψ1(S3) ⊂ S3 and Ψ1S3 = ΦS3. Thus the restriction
map Ψ1S3 is extremal in P(S3). But Ψ1 is not extremal in P(Mn).
The purpose of this note is to clarify this situation. In the next section, we show
that the Choi map is indeed extremal in P(M3).
In section 3, we explain briefly
the correspondence between positive semi-definite real biquadratic forms and positive
linear maps, and clarify some misclaims about this correspondence in the literatures.
We also discuss possible extensions of ΦS3 to positive linear maps in P(M3).
Throughout this note, Mn(R) denotes the real vector space consisting of n × n real
matrices, and {Ekℓ} the usual matrix units in Mn. For a n × n matrix A, det(A)
denotes the determinant of A, and dk(A) denotes the determinant of the submatrix
formed by deleting the k-th row and k-th column of A.
2. Extremality of the Choi map
In this section, we show that the Choi map Φ is extremal in P(M3). Suppose that
(2)
Φ = φ1 + φ2
for φ1, φ2 ∈ P(Mn). For each φk (k = 1, 2), we define two linear maps φk1, φk2 by
φk1(X) =
1
2
(φk(X) + φk(X)), φk2(X) =
1
2i
(φk(X) − φk(X)), X ∈ M3
where φk(X) denotes the matrix whose entries are conjugates of the corresponding
entries of the matrix φk(X). Then, we see that
φkℓ(Mn(R)) ⊂ Mn(R) (k = 1, 2 and ℓ = 1, 2)
and φk1 is positive linear map for k = 1, 2. Therefore, both φ11 and φ21 are positive
linear maps preserving S3.
We note that
(3)
φk = φk1 + iφk2 (k = 1, 2)
and φ(S3) ⊂ S3. So we see that
φS3 = φ11S3 + φ21S3.
2
Thus, we can conclude that φ11S3 = λΦS3 and φ21S3 = (1 − λ)ΦS3 for some 0 ≤ λ ≤ 1
because Φ is extremal in P(S3). From this, we have that
(4)
φ11(E11) = λ(E11 + E22),
φ21(E11) = (1 − λ)(E11 + E22),
φ11(E22) = λ(E22 + E33),
φ21(E22) = (1 − λ)(E22 + E33),
φ11(E33) = λ(E33 + E11),
φ21(E33) = (1 − λ)(E33 + E11),
φ11(S12) = −λS12,
φ11(S23) = −λS23,
φ11(S31) = −λS31,
φ21(S12) = −(1 − λ)S12,
φ21(S23) = −(1 − λ)S23,
φ21(S31) = −(1 − λ)S31,
where Skℓ = Ekℓ + Eℓk for 1 ≤ k < ℓ ≤ 3.
For each 1 ≤ k < ℓ ≤ 3, we define hermitian matrices Hkℓ = (Ekℓ − Eℓk)i. Then M3
is generated by B = {E11, E22, E33, S12, S13, S23, H12, H13, H23}. We note that a positive
linear map in P(M3) is uniquely determined by its value on B. Now, we examine φ1(X)
and φ2(X) for each X ∈ B to determine two positive linear maps φ1, φ2. First, we
consider the positive semi-definite (PSD) matrices φk(E11), φk(E22) and φk(E33) for
k = 1, 2.
Lemma 2.1. φ1(Ekk) and φ2(Ekk) are of the following form
(5)
φ1(E11) =
φ1(E22) =
φ1(E33) =
−a1i
λ
0
a1i 0
0
λ
0
0
0
λ
0
0
0 −a2i
0
a2i
λ
λ
0
0 a3i
0
0
−a3i 0
λ
,
,
,
φ2(E11) =
φ2(E22) =
φ2(E33) =
1 − λ −a1i 0
1 − λ 0
a1i
0
0
0
0
0
0
0 1 − λ −a2i
0
1 − λ
a2i
1 − λ 0 −a3i
0
a3i
0
0
0 1 − λ
,
,
for real numbers a1, a2 and a3.
Proof. Since (3, 3)-entry of the PSD matrix Φ(E11) is equal to zero, (3, 3)-entries of
both φ1(E11) and φ2(E11) are also equal to zero from the equation (2) and the positivity
of φk. Again, the positivity of φk(E11) implies that φk(E11) is of the form in equation
(5). The rest can be checked similarly. (cid:3)
Note that φ1(Skℓ) and φ2(Skℓ) are hermitian matrices. So, all diagonal entries of
both φ12(Skℓ) and φ22(Skℓ) are equal to zero for all Skℓ ∈ B by the identity (3) since
φ12(Skℓ) and φ22(Skℓ) are real matrices. Thus, from (3) and (4), we see that φ1(Skℓ)
3
and φ2(Skℓ) are of the form
φ1(S12) =
φ1(S13) =
φ1(S23) =
(6)
(7)
0
−λ + b1i
0
−b3i
−λ − b1i
−b2i
0
−b4i
b4i −λ + b5i
0
−λ − b5i −b6i
0
b7i
0
−b7i
−b8i −λ − b9i
−λ + b9i
0
b2i
b3i
0
b6i
0
b8i
,
,
,
φ2(S12) =
φ2(S13) =
φ2(S23) =
0
−1 + λ + b1i
b2i
0
b4i
−1 + λ + b5i
−b7i
0
b7i
b8i −1 + λ + b9i
0
−1 + λ − b1i −b2i
−b3i
0
b3i
−b4i −1 + λ − b5i
0
b6i
−b6i
0
−b8i
−1 + λ − b9i
0
0
,
for real numbers b1, b2, · · · , b9.
Now, for any x = (x1, x2, x3)t ∈ C3, we define a rank one PSD matrix
(8)
X[x1, x2, x3] := xx∗ = (xkxℓ).
From the positivity of φk(X[x1, x2, x3]), we can correlate the variables bk's in (6) with
the variables aℓ's in (5).
Lemma 2.2. Let b1, b2, · · · , b9 be the variables in (6), and a1, a2, a3 be the variables in
(5). Then we have
(9)
b1 = 0,
b2 = −a2,
b3 = −a3,
b4 = a2,
b5 = 0,
b6 = a1,
b7 = −a3,
b8 = −a1,
b9 = 0.
Proof. For a real number t, we consider two PSD matrices
φk(X[1, t, 0]) = φ1(E11) + tφ1(S12) + t2φ1(E22)
(k = 1, 2).
We know that the principal minors dk (φℓ(X[1, t, 0]))'s are nonnegative. In particular,
we have
d3 (φ1(X[1, t, 0])) = −b2
1t2 − 2a1b1t + (λ2 − a2
1) ≥ 0
for all real number t. Thus we get b1 = 0. Similarily, by considering the principal
minors d2 (φ1(X[1, 0, t])) and d1 (φ1(X[0, 1, t])), we can show that b5 = 0 and b9 = 0
respectively.
4
We see that determinants det (φk(X[1, t, 0])) (k = 1, 2) are quartic polynomial in
t and divisible by t2. Therefore, the coefficients of t4 in det (φk(X[1, t, 0])) (k = 1, 2)
should be nonnegative. So, we have
−λ(a2 + b2)2 ≥ 0 and − (1 − λ)(a2 + b2)2 ≥ 0.
Therefore, we can conclude that b2 = −a2. By the same method, we can show that
b7 = −a3 by considering two quartic polynomials det (φk(X[0, 1, t])) (k = 1, 2) in
variable t.
Two quartic polynomials det (φℓ(X[1, 0, t])) in t are also divisible by t2. When
b5 = 0, the coefficients of t2 are
−λ(a1 − b6)2 and − (1 − λ)(a1 − b6)2.
Therefore, we get b6 = a1.
Up to now, we have shown that b1 = b5 = b9 = 0, b2 = −a2, b6 = a1 and b7 = −a3.
From the equations (5), (6) and (7), we can compute that
det (φ1(X[1, 1, 1])) + det (φ1(X[1, 1, −1])) + det (φ1(X[1, −1, 1]))
det (φ2(X[1, 1, 1])) + det (φ2(X[1, 1, −1])) + det (φ2(X[1, −1, 1]))
+ det (φ1(X[1, −1, −1])) = −8λ(cid:0)(a1 + b8)2 + (a2 − b4)2 + (a3 + b3)2(cid:1) ,
+ det (φ2(X[1, −1, −1])) = −8(1 − λ)(cid:0)(a1 + b8)2 + (a2 − b4)2 + (a3 + b3)2(cid:1) .
Since the above two values must be nonnegative, we can conclude that
b3 = −a3, b4 = a2 and b8 = −a1.
This completes the proof. (cid:3)
Now, we consider hermitian matrices φ1(Hkℓ) and φ2(Hkℓ) for Hkℓ ∈ B. For real
numbers ci's and complex numbers αi's, we may write φk(H12) (k = 1, 2) as
φ1(H12) =
c1 α1 α2
c2 α3
α1
α2 α3
c3
, φ2(H12) =
−c1 −i − α1 −α2
−α3
i − α1
−α2
−c3
−c2
−α3
.
We consider PSD matrices
φk(X[1, −ti, 0]) = φk(E11) + tφk(H12) + t2φk(E22)
(k = 1, 2).
We see that d3 (φk(X[1, −ti, 0])) (k = 1, 2) are cubic polynomial in t, and the coeffi-
cients of t3 are
Therefore, we have c1 = 0. Then we have the following:
λc1 and (λ − 1)c1.
d3 (φ1(X[1, −ti, 0])) =λ2 − a2
d3 (φ2(X[1, −ti, 0])) =(λ − 1)2 − a2
1 + [λc2 − 2a1Im(α1)]t + (λ2 − α12)t2 ≥ 0,
1 + [(λ − 1)c2 − 2a1(1 + Im(α1))]t
+ [(λ − 1)2 − Re(α1)2 − (1 + Im(α1))2]t2 ≥ 0,
5
for all t ∈ R, where α1 = Re(α1) + Im(α1)i. So we get the conditions
(10)
(11)
λ2 ≥ α12
(λ − 1)2 ≥ Re(α1)2 + (1 + Im(α1))2
Since 0 ≤ λ ≤ 1, we see that Im(α1) < 0 by (11). Then, we get that −λ ≤ Im(α1) < 0
from the condition (10), that is, 0 ≤ 1 − λ ≤ 1 + Im(α1) < 1. Therefore, we have that
(1 − λ)2 ≤ (1 + Im(α1))2 ≤ Re(α1)2 + (1 + Im(α1))2 ≤ (1 − λ)2.
Consequently, we conclue α1 = −λi. We note that this implies the coefficient of t in
d3 (φ1(X[1, −ti, 0])) should be zero. Therefore, we get c2 = −2a1.
When c1 = 0, d2 (φk(X[1, −ti, 0])) (k = 1, 2) are quadratic polynomials in t divisible
by t. Therefore, the coefficients of t should be zero. From this observation, we have
that
λc3 = 0 and (λ − 1)c3 = 0.
Consequently, we see that c3 = 0.
Finally, we show that α2 = a2 by considering the determinant of φk(X[1, −ti, 0]).
Under the conditions c1 = 0 and α1 = −λi, we can show that the determinants
det (φk(X[1, −ti, 0])) (k = 1, 2) are quartic polynomials in t, and the coefficients of t4
are
−λα2 − a22 and − (1 − λ)α2 − a22.
Since both coefficients should be nonnegative, we get α2 = a2.
To sum up, we have correlated all entries of φk(H12) except α3 with ai in (5). That
is, c1 = c3 = 0, c2 = −2a1, α1 = −λi and α2 = a2.
Lemma 2.3. Let a1, a2, a3 be the real variables in (6). Then φ1(Hkℓ) and φ2(Hkℓ) are
of the following forms
(12)
φ1(H12) =
φ1(H13) =
φ1(H23) =
0 −λi
a2
λi −2a1 α
0
a2
−2a3
α
β
λi −a1
0 −a3
β −λi
0 −a1
−a3
γ
−λi
0
λi −2a2
,
0
γ
φ2(H12) =
, φ2(H13) =
, φ2(H23) =
0
(1 − λ)i
−a2
2a3
−β
(1 − λ)i
(λ − 1)i −a2
−α
0
2a1
−α
−β (λ − 1)i
0
a1
a1
0
a3
0
0
a3
−γ (1 − λ)i
(λ − 1)i
−γ
2a2
,
,
.
Proof. For the case of φk(H12), we have done it with α3 = α. Through the same process
with principal minors d3 (φk(X[0, 1, −ti])) , d1 (φk(X[0, 1, −ti])) and the determinant
det (φk(X[0, 1, −ti])), we can show that φk(H23) is of the form in (12) for k = 1, 2.
6
For the case of φk(H13), it suffices to consider d3 (φk(X)) , d2 (φk(X)) and then
det (φk(X)) with X = X[1, 0, −ti]. But, in this case, each d2 (φk(X)) is quartic poly-
nomial in t divisible by t2, and we can determine (1, 3) and (1, 1) entries of φk(H13) by
examining the coefficients of t2 as in (10) and (11). (cid:3)
Now, we are ready to prove that the Choi map is indeed extremal in the cone
P(M3). We note that a positive linear map ψ satisfies
ψ(Ekℓ) =
1
2
(ψ(Skℓ) − iψ(Hkℓ)) , ψ(Eℓk) =
1
2
(ψ(Skℓ) + iψ(Hkℓ))
for 1 ≤ k < ℓ ≤ 3. Therefore, positive linear maps φ1 and φ2 in (2) are uniquely
determined by the Lemma 2.1, 2.2 and 2.3. For the convenience of readers, we make
up a list of entries of φ1(X) and φ2(X) for 3 × 3 matrix X = (xkℓ). In the following
list, [A]kℓ denotes the entry in the k-th row and ℓ-th column of a matrix A.
(13)
[φ1(X)]12 = −x12λ + a1x11i − a3x32i +
[φ1(X)]21 = −x21λ − a1x11i + a3x23i −
[φ1(X)]13 = −x13λ + a3x33i − a2x12i −
[φ1(X)]11 = (x11 + x33)λ + a3(x13 − x31)i,
1
2
1
2
1
2
[φ1(X)]22 = (x22 + x11)λ + a1(x12 − x21)i,
1
2
1
2
1
2
[φ1(X)]33 = (x33 + x22)λ + a2(x23 − x32)i,
[φ1(X)]32 = −x32λ − a1x31i − a2x22i +
[φ1(X)]23 = −x23λ + a1x13i + a2x22i −
[φ1(X)]31 = −x31λ − a3x33i + a2x21i +
(a2 − β)x13i +
(a1 + γ)x23i −
(a2 − β)x31i −
(a3 + α)x12i −
(a1 + γ)x32i +
(a3 + α)x21i +
1
2
1
2
1
2
1
2
1
2
1
2
(a2 + β)x31i,
(a1 − γ)x32i,
(a2 + β)x13i,
(a3 − α)x21i,
(a1 − γ)x23i,
(a3 − α)x12i,
[φ2(X)]21 = −x21(1 − λ) + a1x11i − a3x23i +
[φ2(X)]13 = −x13(1 − λ) − a3x33i + a2x12i +
[φ2(X)]12 = −x12(1 − λ) − a1x11i + a3x32i −
[φ2(X)]11 = (x11 + x33)(1 − λ) − a3(x13 − x31)i,
1
2
1
2
1
2
[φ2(X)]22 = (x22 + x11)(1 − λ) − a1(x12 − x21)i,
1
2
1
2
1
2
[φ1(X)]33 = (x33 + x22)(1 − λ) − a2(x23 − x32)i
[φ2(X)]31 = −x31(1 − λ) + a3x33i − a2x21i −
[φ2(X)]23 = −x23(1 − λ) − a1x13i − a2x22i +
[φ2(X)]32 = −x32(1 − λ) + a1x31i + a2x22i −
(a2 − β)x13i −
(a1 + γ)x23i +
(a2 − β)x31i +
(a3 + α)x12i +
(a1 + γ)x32i −
(a3 + α)x21i −
1
2
1
2
1
2
1
2
1
2
1
2
(a2 + β)x31i,
(a1 − γ)x32i,
(a2 + β)x13i,
(a3 − α)x21i,
(a1 − γ)x23i,
(a3 − α)x12i,
where a1, a2, a3 ∈ R, α, β, γ ∈ C and 0 ≤ λ ≤ 1.
7
First, we determine real variables a1, a2 and a3. To do this, we consider principal
minors d3 (φk(X[1, t, si])) (k = 1, 2) for real numbers t and s, . They are quadratic
polynomials in t with the following coefficients of t2.
4λa3s + (λ2 − a2
3)s2,
4(λ − 1)a3s +(cid:0)(λ − 1)2 − a2
3(cid:1) s2
These coefficients should be nonnegative for all s because principal minors of a PSD
matrix are nonnegative. Therefore, we have a3 = 0. And then, by considering
d2 (φk(X[t, 1, si])) as a polynomial in t for each k = 1, 2, we can show that a2 = 0
similarly. Finally, we get a1 = 0 from the condition that the coefficients of t2 in qua-
dratic polynomials d1(φk(X[si, 1, t])) (k = 1, 2) should be nonnegative. Consequently,
we have
(14)
a1 = a2 = a3 = 0.
Now, we show that all complex variables α, β and γ should be zero when a1 = a2 =
a3 = 0 in (13).
(Case 1: λ = 0)
If λ = 0, then we have PSD matrices
φ1(X[1, 1, i]) =
0 −β −γ
−β
0
−γ
0
0
0
, φ1(X[1, i, 1]) =
0
0
γ −α
0
γ
0 −α
0
from (13). Therefore, we have α = β = γ = 0 when λ = 0. In this case, we conclude
that φ2 is the Choi map Φ and φ1 is the zero map.
(Case 2: λ = 1)
In this case, by considering PSD matrices φ2(X[1, 1, i]) and φ2(X[1, i, 1]), we can
show that φ1 is the Choi map Φ and φ2 is the zero map as in (Case 1).
(Case 3: 0 < λ < 1)
Since the determinant of a PSD is nonnegative, we have the following inequalities:
(15)
= − 4λ(α2 + β2) + 12λ2Im(β) ≥ 0,
det(cid:0)φ1(X[1, e
det(cid:0)φ2(X[1, e
π
2 i, e
π
2 i, e
π
2 i])(cid:1) + det(cid:0)φ1(X[1, e−
2 i])(cid:1) + det(cid:0)φ2(X[1, e−
π
π
2 i, e−
π
2 i, e−
π
2 i])(cid:1)
2 i])(cid:1)
π
= − 4(1 − λ)(α2 + β2) − 12(1 − λ)2Im(β) ≥ 0,
These are equivalent to the inequalities
0 ≤
α2 + β2
3λ
≤ Im(β) ≤ −
α2 + β2
3(1 − λ)
≤ 0
because of 0 < λ < 1. Therefore, we see Im(β) = 0, and so α = β = 0.
Finally, we show that γ = 0 when a1 = a2 = a3 = 0 and α = β = 0 in (13). For a
PSD matrix X[1, 1, i], we get two inequalities
det (φ1(X[1, 1, i])) = −2λγ2 + 6λ2Im(γ) ≥ 0,
det (φ2(X[1, 1, i])) = −2(1 − λ)γ2 − 6(1 − λ)2Im(γ) ≥ 0.
8
Then we see that
0 ≤
γ2
3λ
≤ Im(γ) ≤ −
γ2
3(1 − λ)
≤ 0.
Thus, we get γ = 0. Consequently, we have shown that φ1 = λΦ and φ2 = (1 − λ)Φ.
In any cases, we have the following.
Theorem 2.4. The Choi map generates an extreme ray in the cone P(M3).
3. Correspondence between positive semi-definite biquadratic real
forms and positive linear maps
We note [2] that the Choi map Φ is originated from a real biquadratic form B(x, y)
with the relation
B(x, y) =(x2
1 + x2
3)y2
1 + (x2
2 + x2
1)y2
2 + (x2
3 + x2
2)y2
3
− 2(x1x2y1y2 + x2x3y2y3 + x3x1y3y1)
=ytΦ(xxt)y
for real column vector x = (x1, x2, x3)t, y = (y1, y2, y3)t ∈ R3.
In general, for any positive real linear map φ : Sn → Sn, we get the correspond-
ing positive semi-definite real biquadratic form Bφ(x, y) with x, y ∈ Rn defined by
Bφ(x, y) = ytφ(xxt)y.
On the other hand, let B(x, y) be a positive sem-definite real biquadratic form with
x, y ∈ Rn. Then, for any fixed x ∈ Rn, B(x, y) is a positive semi-definite real quadratic
form with respect to y ∈ Rn. So we can write B(x, y) in the form ytS(x)y for some
n × n symmetric matrix S(x). Thus we get a map which take any one dimensional
projection xxt to S(x). Consequently, we get the corresponding positive linear map
φ : Sn → Sn by linearity, which satisfy the relation Bφ(x, y) = B(x, y). Therefore, we
see that there is a one-to-one correspondence between the set of positive semi-definite
real biquadratic forms and the set P(Sn) consisting of positive real linear maps between
Sn.
We can find some misclaims on the above correspondence in the literatures [7, 9].
They claim that
(i) Let Ψ be a positive complex linear map with Ψ(Mn(R)) ⊂ Mn(R).
If the
corresponding real biquadratic form BΨ(x, y) is extremal in the cone of all
positive semi-definte real biquadratic forms, then Ψ is extremal in the cone
P(Mn).
(ii) Using linearity and hermicity, the above correspondence can be extended to the
correspondence between the set of positive semi-definite real biquadratic forms
and the set P(Mn) trivially.
For the first claim, we have shown there exists a counter example Ψ1 in (1). Here, we
clarify the claim (ii).
9
As before, we define n × n symmetric matrices Skℓ and antisymmetric matrices Akℓ
for 1 ≤ k < ℓ ≤ n by
Skℓ = Ekℓ + Eℓk, Akℓ = Ekℓ − Eℓk.
We note that Sn is generated by
G = {Ekk : 1 ≤ k ≤ n} ∪ {Skℓ : 1 ≤ k < ℓ ≤ n}
and real matrix algebra Mn(R) is generated by G ∪ {Akℓ : 1 ≤ k < ℓ ≤ n}. Thus, a
map φ ∈ P(Sn) can be extended to a positive linear map eφ in P(Mn(R)) by defining
eφ(Akℓ) for 1 ≤ k < ℓ ≤ n. It is worthy to note that the positivity of eφ is not affected
by eφ(Akℓ)'s. That is, the positivity of eφ is determined by eφSn = φ.
On the other hand, we can uniquely extend eφ ∈ P(Mn(R)) to the complex linear
map eφ between Mn by the linearity eφ(X + iY ) = eφ(X) + ieφ(Y ) for X, Y ∈ Mn(R).
But, in this extension, the positivity of the complex linear map eφ is not determined by
the positivity of the real linear map eφ.
In general, let Ψ be a positive linear map in P(Mn) with Ψ(Mn(R)) ⊂ Mn(R). So,
we can regard Ψ as an extension of real linear map. Then, it is well known that Ψ
preserve hermitian matrices. That is, Ψ(iAkℓ) must be a hermitian matrix.
Since any hermitian matrix H ∈ Mn can be written by H = S+iA with a symmetric
matrix S ∈ Mn(R) and an antisymmetric matrix A ∈ Mn(R), we observe the following.
Proposition 3.1. Let Ψ be a positive linear map in P(Mn) with Ψ(Mn(R)) ⊂ Mn(R).
Then Ψ preserve hermitian matrices if and only if Ψ preserve symmetric matrices and
antisymmetric matrices respectively.
Proof. For a symmetric matrix S and an antisymmetric matrix A, we have
Ψ(S + iA)∗ = Ψ(S)t − iΨ(A)t = Ψ(S) + iΨ(A) = Ψ(S + iA)
since Ψ preserve hermitian matrices and Ψ(Mn(R)) ⊂ Mn(R). Therefore, Ψ(S)t =
Ψ(S) and Ψ(A)t = −Ψ(A). This completes the proof. (cid:3)
We note that there exists non-positive linear map Ψ2 between M3, which satisfies
the following conditions
• Ψ2 preserves hermitian matrices.
• Ψ2M3(R) is a positive linear map between M3(R) and Ψ2S3 = ΦS3 for the Choi
map Φ.
From the condition Ψ1S3 = ΦS3, Ψ2 is determined by the values Ψ2(Akℓ). We put
Ψ2(A12) = −A12, Ψ2(A13) = −A13, Ψ2(A23) = −A12.
Then, Ψ2 is defined by
(16)
Ψ2(X) =
x11 + x33
−x21 + 1
2(x23 − x32)
−x31
−x12 − 1
2 (x23 − x32)
x11 + x22
2(x23 + x32)
− 1
10
−x13
− 1
2(x23 + x32)
x22 + x33
for X = (xkℓ) ∈ Mn. We know that Ψ2M3(R) is a positive linear map between M3(R)
and Ψ2S3 = ΦS3. But, this map is not positive map between M3 because det (Ψ2(X)) =
−25 for a PSD matrix X = X[1, 2 − i, −1 − i] in (8).
Now, we give an example of extremal positive linear map Ψ3 in P(M3), which is
not equal to the Choi map Φ but Ψ3S3 = ΦS3. This example explains the claim (ii)
is nonsense, and we can conclude that extremal extension of ΦS3 is not unique. We
define Ψ3 by
Ψ3(X) = Φ(X) for all X ∈ {E11, E22, E33, S12, S13, S23, A12, A13},
Ψ3(A23) = A23.
Then, we obtain a complex linear map Ψ3 preserving hermitian matrices by Proposi-
tion 3.1. We will show that this map is indeed positive. We recall that the positivity
of the Choi map is easily proven through the positivity of the real map ΦS3 with the
following relation:
Φ(X[x1, x2, x3]) = V Φ(X[r1, r2, r3])V ∗ with V =
eiθ1
0
0
0
eiθ2
0
0
0
eiθ3
where xi = rieiθi for i = 1, 2, 3. But, this method is not applicable to general cases. In
fact, it is easy to show that there exists no matrix V satisfying Ψ3(X) = V Ψ3(Y )V ∗
for rank one PSD matrices X = X[r1eiθ1, r2eiθ2, r3eiθ3] ∈ M3 and Y = X[r1, r2, r3] ∈
M3(R).
So, we will show that Ψ3(X) is a PSD matrix for any rank one PSD matrix X.
Since Φ is positive, we see that
dk (Ψ3(X)) = dk (Φ(X)) ≥ 0
for any rank one PSD matrix X = X[x1, x2, x3] and k = 1, 2, 3. By the arithmetic
mean-geometric mean inequality, we also see that
det (Ψ3(X)) =x12x34 + x22x4
1 + x32x24
− x12x22x32 − 2x12Re(x2
2x2
3)
≥2x12(cid:0)x22x32 − Re(x2
3)(cid:1) ≥ 0.
2x2
Therefore, Ψ3(X) is a PSD matrix for any rank one PSD matrix X = X[x1, x2, x3].
That is, Ψ3 is a positive linear map in P(M3). The extremality of Ψ3 in P(M3) is
similarly checked as that of the Choi map.
Finally, we discuss extremality for various convex cones of positive linear maps.
Let Ψ be a positive linear map preserving Mn(R) in P(Mn). Then ΨMn(R) belongs
to P(Mn(R)). Furthermore, we see that Ψ(Sn) ⊂ Sn from the positivity of Ψ and
Ψ(Mn(R)) ⊂ Mn(R), that is, ΨSn ∈ P(Sn). Now, let PR(Mn) be the set of all positive
complex linear maps preserving Mn(R). Then PR(Mn) is a convex subcone contained
in P(Mn). Since any real linear map between Mn(R) can be uniquely extended to the
complex linear map between Mn, we may think PR(Mn) = P(Mn) ∩ P(Mn(R)). So, we
11
may consider extremality of a map in each convex cones P(Mn), PR(Mn), P(Mn(R))
and P(Sn). We note that the Choi map is extremal in each of these convex cones.
Owing to the map Ψ1 in (1), we know that the extremality of ΨSn in P(Sn) implies
neither the extremality of ΨMn(R) in P(Mn(R)) nor the extremality of Ψ in P(Mn).
On the other hand, for Ψ ∈ PR(Mn), it is easy to see that the extremality of ΨMn(R)
in P(Mn(R)) implies the extremality of Ψ in PR(Mn). But, it seems that the converse
does not hold since there exists a non-positive complex linear map Ψ satisfying two
properties Ψ(Mn(R)) ⊂ Mn(R) and ΨMn(R) ∈ P(Mn(R)) as a map Ψ2 in (16). So, it
is interesting to find an extremal positive linear map Ψ in PR(Mn) whose restriction
ΨMn(R) is not extremal in P(Mn(R)).
References
[1] M.-D. Choi, Positive semidefinite biquadratic forms, Linear Algebra and Appl. 12 (1975), 95 -- 100.
[2] M.-D. Choi, Some assorted inequalties for positive linear maps on C ∗-algebras, J. Operator The-
ory 4 (1980), 271-285.
[3] M.-D. Choi, T.-Y. Lam, Extremal positive semidefinite forms, Math. Ann. 231 (1977), 1 -- 18.
[4] K.-C. Ha, S.-H. Kye, Entanglement witnesses arising from exposed positive linear maps, Open
Syst. Inf. Dyn. 18 (2011), 323 -- 337.
[5] K.-C. Ha, S.-H. Kye, Exposedness of Choi type entanglement witnesses and applications to lengths
of separable states, arXiv:1211.5675 (2013).
[6] H.-J. Kim, S.-H. Kye, Indecomposable extreme positive linear maps in matrix algebras, Bull.
London Math. Soc. 26 (1994), 575 -- 581.
[7] H. Osaka, A class of extremal positive maps in 3 × 3 matrix algebras, Publ. Res. Inst. Math. Sci.
28 (1992), 747 -- 756.
[8] A. G. Robertson, Positive projections on C ∗-algebras and an extremal positive map, J. London
Math. Soc. (2) 32 (1985), 133 -- 140.
[9] R. Sengupta, Arvind, Extremal extensions of entanglement witnesses: Finding new bound entan-
gled states, Phys. Rev. A 84 (2011), 032328
[10] E. Størmer, Positive linear maps of operator algebras, Acta Math. 110 (1963), 233 -- 278.
Faculty of Mathematics and Applied Statistics, Sejong University, Seoul 143-747,
Korea
12
|
1612.05127 | 1 | 1612 | 2016-12-15T16:18:54 | Graph products and the absence of property (AR) | [
"math.OA"
] | We discuss the internal structure of graph products of right LCM semigroups and prove that there is an abundance of examples without property (AR). Thereby we provide the first examples of right LCM semigroups lacking this seemingly common feature. The results are particularly sharp for right-angled Artin monoids. | math.OA | math |
GRAPH PRODUCTS AND THE ABSENCE OF PROPERTY (AR)
NICOLAI STAMMEIER
Abstract. We discuss the internal structure of graph products of right LCM semi-
groups and prove that there is an abundance of examples without property (AR).
Thereby we provide the first examples of right LCM semigroups lacking this seemingly
common feature. The results are particularly sharp for right-angled Artin monoids.
1. Introduction
The starting point of a number of recent breakthroughs in the theory of semigroup
C ∗-algebras is the seminal work [Li12, Li13], in which a universal C ∗-algebra C ∗(S) is
associated to every left cancellative monoid S. In the last years, a particular line of
research focused on left cancellative monoids for which the intersection of two principal
right ideals is either empty, or another principal right ideal again. Such monoids are
called right LCM semigroups, and they form an intriguing and tractable class of examples
in between positive cones in quasi-lattice ordered groups and general left cancellative
monoids, see [BLS17, Lemma 3.3 and Corollary 3.6] for details.
Inspired by the treatment of the quasi-lattice ordered case in [CL07], a boundary
quotient Q(S) of C ∗(S) was introduced for right LCM semigroups S in [BRRW14]. Soon
thereafter, Starling provided an in-depth analysis of Q(S) in [Star15], relying on major
advances in the understanding of the connections between inverse semigroups, groupoids,
and C ∗-algebras stemming from [EP17, EP16, Ste16]. In [BS16], it was shown that the
boundary quotient has a more accessible presentation if the right LCM semigroup has
the so-called accurate refinement property, henceforth abbreviated property (AR). This
property is an analogue of 0-dimensionality for topological spaces in the context of
semigroups, and is enjoyed by various examples, see [BS16, Section 2 and Corollary 3.11].
The presence of property (AR) was found to be useful in the construction of a bound-
ary quotient diagram for right LCM semigroups in the spirit of [BaHLR12], see [Sta].
This diagram sets the grounds for a unifying approach to the study of equilibrium states
on C ∗-algebras in [ABLS], where remarkable results on the structure of KMS-states on
C ∗(S) were obtained for right LCM semigroups satisfying an admissibility condition
which implies property (AR), see Subsection 2.1. Working with abstract right LCM
semigroups as opposed to explicit classes of examples allowed for a unification of the in-
spiring case studies [LR10, BaHLR12, LRR11, LRRW14, CaHR16], and also for coverage
of a substantial amount of new examples, most notably, algebraic dynamical systems.
Moreover, the techniques in [BS16, Sta, ABLS] raise several questions on the structure
of right LCM semigroups, perhaps most notably:
(a) Are there right LCM semigroups without property (AR)?
2010 Mathematics Subject Classification. 20M10 (Primary) 20F36, 46L05 (Secondary).
The author was supported by RCN through FRIPRO 240362.
1
2
NICOLAI STAMMEIER
(b) Which right LCM semigroups are admissible?
The aim of the present work is to investigate in how far graph products of right LCM
semigroups as considered in [VdC01, FK09] provide answers to these two questions. In
addition, we also address structural aspects related to the distinguished subsemigroups
S∗, Sc and Sci. We apply our results to the classical case of right-angled Artin monoids
A+
Γ given by an undirected graph Γ since many graph related phenomena can already
be witnessed here. Indeed, the explicit presentation of the boundary quotient in [CL07,
Corollary 8.5] involving only the vertex sets of the finite coconnected components of
the graph Γ may be regarded as an indication for a particularly accessible structure of
foundation sets. Another motivation comes from the elegant solution to the isomorphism
problem for C ∗(A+
Γ ), see [ELR16].
Since property (AR) is known for various kinds of right LCM semigroups, we were
struck by surprise to find that a right-angled Artin monoid A+
Γ has property (AR) if and
only if all of its finitely generated direct summands are free, see Corollary 4.6. In terms
of the Γ, this means that all finite coconnected components Γi do not contain any edges.
The result follows from more general graph product considerations in Corollary 4.5 that
rely on Theorem 4.3, where we show that graph products over infinite coconnected
graphs have no foundation set other than the obvious ones containing an invertible
element, while the analogous statement holds in the finite case for accurate foundation
sets.
The characterisation of property (AR) for right-angled Artin monoids A+
Γ in Corol-
lary 4.6 allows us to determine when A+
Γ is admissible in the sense of [ABLS]. It turns
out that admissibility and the existence of a generalised scale coincide for right-angled
Artin monoids, see Corollary 4.9 and Corollary 4.10. If existent, the generalised scale on
A+
Γ is unique and arises as the product of the unique generalised scales on its non-abelian
direct summands A+
Γi, see Proposition 4.8 and Corollary 4.9.
Thus we are lead to the conclusion that graph products of right LCM semigroups
mostly lack property (AR), and are therefore not admissible in the sense of [ABLS].
While this rules out the possibility of applying [ABLS] to graph products of right LCM
semigroups in great generality, we obtain a fairly detailed description of the behaviour
of graph products with respect to the subsemigroups Sc and Sci, see Theorem 3.4.
These result show that the graph product represents a useful tool to construct new, and
potentially very interesting examples of right LCM semigroups that are well-behaved
to some degree, but demand more sophisticated techniques then those applicable to
right LCM semigroups that have property (AR) or even a generalised scale. That is
why we feel that this work might stimulate further research in the direction of inverse
semigroups and groupoids related to (right LCM) semigroups and their C ∗-algebras.
Acknowledgements: The author thanks Nathan Brownlowe, Nadia Larsen, and Adam
Sørensen for helpful conversations.
2. Preliminaries
Here we provide the prerequisites we shall need concerning right LCM semigroups
and graph products.
2.1. Right LCM semigroups. A left cancellative semigroup S is called right LCM if
the intersection of two principal right ideals in S is either empty or a principal right
GRAPH PRODUCTS AND THE ABSENCE OF PROPERTY (AR)
3
ideal. For s, t ∈ S, we say that s and t are orthogonal and write s ⊥ t if sS ∩ tS = ∅.
Unless specified otherwise, we will always assume that a right LCM semigroup S has an
identity, i.e. S is a monoid.
Let us first discuss property (AR). A finite subset F ⊂ S is called a foundation set for
S if for every s ∈ S there is f ∈ F such that f 6⊥ s, see [BRRW14, Section 5]. A subset
F ⊂ S is accurate if f ⊥ f ′ for all f, f ′ ∈ F, f 6= f ′, see [BS16, Definition 2.1]. If F, F ′
are foundation sets such that F ′ ⊂ F S, then F ′ is called a refinement of F . We then
say that S has the accurate refinement property, or property (AR), if every foundation
set for S has an accurate refinement, see [BS16, Definition 2.3].
For a right LCM semigroup S, its subgroup of invertible elements shall be denoted
by S∗. This subgroup lies inside the core subsemigroup Sc := {a ∈ S a 6⊥ s for all s ∈
S}, which was first considered for right LCM semigroups in [Star15], but stems from
[CL07, Definition 5.4]. We remark that Sc is again a right LCM semigroup. Furthermore,
it induces an equivalence relation s ∼ t :⇔ sa = tb for some a, b ∈ Sc called the core
relation. In contrast to Sc, we also consider the subsemigroup Sci of core irreducible
elements, that is, the collection of all elements s ∈ S \ Sc for which every factorization
s = ta with t ∈ S, a ∈ Sc satisfies a ∈ S∗. While Sci does not have an identity by
construction, its unitisation S1
ci := Sci ∪ {1} and S′
ci := Sci ∪ S∗ do.
A right LCM semigroup S is called core factorable if S = S1
ciSc. We say that Sci ⊂ S is
∩-closed if sS ∩ tS = rS implies r ∈ Sci whenever s, t ∈ Sci. To provide some indication
why this property is of interest, let us mention that Sci ⊂ S is ∩-closed if and only if S′
ci
is right LCM and its inclusion into S is a homomorphism of right LCM semigroups, i.e. it
preserves intersections of principal right ideals, see [ABLS, Proposition 3.3]. Finally, a
nontrivial homomorphism N : S → N× is called a generalised scale if N −1(n)/∼ = n
and every minimal complete set of representatives for N −1(n)/∼ forms an accurate
foundation set for S for all n ∈ N(S). Every generalised scale N satisfies ker N = Sc
by [ABLS, Proposition 3.6(i)], and the existence of a generalised scale entails vital
information on the structure of S. For instance, it implies that the right LCM semigroup
has property (AR), see [ABLS, Proposition 3.6(v)].
Finally, we recall from [ABLS, Definition 3.1] that a right LCM semigroup S is called
admissible, if it is core factorable, Sci ⊂ S is ∩-closed, and S admits a generalised scale
N such that N(S) ⊂ N× is freely generated by its irreducible elements.
2.2. Graph products. Within this work, a graph will mean a countable, undirected
graph Γ = (V, E) without loops or multiple edges. The concept of a graph product of
groups emerged in [Gre90] as a generalization of graph groups, and has been transferred
to the setting of monoids in [VdC01]: For a graph Γ = (V, E) and a family of monoids
(Sv)v∈V , the graph product is the monoid SΓ obtained as the quotient of the direct sum
Lv∈V Sv by the congruence generated by the relation (st, ts) if s ∈ Sv, t ∈ Sw with
(v, w) ∈ E, see [VdC01, Section 2] and [FK09, Section 1]. Given a graph Γ, its right-
angled Artin monoid A+
Γ is the graph product with Sv = N for all v ∈ V . These monoids
have also been studied under the names of graph monoids, free partially commutative
monoids, and trace monoids, see for instance [Die90]. If one switches the vertex monoids
from the natural numbers to the integers, the resulting graph product is the right-angled
Artin group AΓ associated to Γ, see [Cha07] for more.
4
NICOLAI STAMMEIER
It was shown in [CL02] that the graph product is well-behaved with respect to quasi-
lattice orders. Invoking a characterization of the right LCM property via the inverse hull
semigroup, Fountain and Kambites showed that this can be generalised to right LCM
semigroups, see [FK09, Theorem 2.6], where we note that we can move back and forth
between right cancellative, left LCM semigroups (used in [FK09]) and left cancellative,
right LCM semigroups by passing to the opposite semigroup.
According to [FK09, Theorem 1.1], which is an adaptation of the corresponding result
in [Gre90], every element s in a graph product SΓ is represented by an essentially unique
reduced expression sv(1)sv(2) · · · sv(n), that is, sv(k) ∈ Sv(k), v(k) 6= v(k + 1), and whenever
there are 1 ≤ k < m ≤ n such that v(k) = v(m), then there exists k < ℓ < m such
that (v(k), v(ℓ)) /∈ E. The analogous result had been proven in the quasi-lattice ordered
case before, see [CL02, Theorem 2]. The reduced expression is unique in the sense that
any two reduced expressions for the same element are shuffle equivalent, i.e. we can
move from one to the other by a finte number of switches of neighbouring factors whose
vertices are adjacent in Γ. Thus there exists a subadditive function ℓ : SΓ → N that
assigns the length of any reduced expression to the element in question.
A graph Γ is said to be coconnected if there exists no partition V = V1 ⊔ V2 with
Vi 6= ∅ and V1 × V2 ⊂ E. Equivalently, Γ is coconnected if the opposite graph Γopp :=
(V, V × V \ (E ∪ {(v, v) v ∈ V })) is connected. The decomposition of Γ into its
coconnected components is the initial step in the analysis of SΓ, see for instance [ELR16],
where the synonym co-irreducible is used. Every graph Γ has a unique decomposition
into coconnected components, which we denote by (Γi)i∈I with Γi = (Vi, Ei). The
original graph can be recovered from (Γi)i∈I as V = Fi∈I Vi and E = {(v, w) ∈ V × V
(v, w) ∈ Ei or w /∈ Vi ∋ v for some i ∈ I}.
It follows from this observation that
SΓ coincides with the direct sum Li∈I SΓi over the graph products obtained from its
coconnected components.
A vertex v ∈ V is called isolated if v does not emit any edge, and universal if v is
connected to every other vertex in Γ. We note that the only coconnected graph with
a universal vertex v is V = {v}, and that any graph containing an isolated edge is
necessarily coconnected. For convenience, we let Vu denote the set of universal vertices,
and I2 := {i ∈ I Vi ≥ 2}.
We will make use of the following notion of a blocking path, that is actually a path
in the opposite graph.
Definition 2.1. Let Γ = (V, E) be a graph and C ⊂ V . A blocking path for C is a finite
sequence of vertices w(1), . . . , w(n) ∈ V such that
(a) w(1) /∈ C, (w(k), w(k + 1)) /∈ E for all 1 ≤ k ≤ n − 1, and
(b) for every u ∈ C there exists 1 ≤ k ≤ n such that (w(k), u) /∈ E.
It turns out that blocking paths are almost always available whenever the graph is
coconnected, and we will frequently make use of this elementary observation in the
course of this work.
Lemma 2.2. If Γ is a coconnected graph with at least two vertices, then every finite
proper subset C of V admits a blocking path ending in any prescribed vertex.
Proof. Let C = {v(1), . . . , v(m)} ⊂ V be finite and proper, that is, V \C 6= ∅. If (v, u) ∈
E for all v ∈ C, u ∈ V \ C, then we would get a contradiction to Γ being coconnected.
GRAPH PRODUCTS AND THE ABSENCE OF PROPERTY (AR)
5
Thus there exists w(1) ∈ V \ C such that (v(k), w(1)) /∈ E for some 1 ≤ k ≤ m.
Without loss of generality, we can assume k = 1. Since Γ is coconnected, we can choose
w′(k) ∈ V for 2 ≤ k ≤ m such that (v(k), w′(k)) /∈ E. Again by coconnectedness, there
exists a finite path w(1), . . . , w(n) in Γopp that visits every w′(k), 2 ≤ k ≤ n. This is a
blocking path for C, and since Γopp is connected, we can attach to this blocking path a
path leading to any prescribed vertex without loosing the blocking property for C. (cid:3)
Remark 2.3. Let Γ = (V, E) be a graph and (Sv)v∈V a family of right LCM semi-
groups. Suppose w(1), . . . , w(n) is a blocking path for some nonempty C, and we can
choose sn, tn ∈ Sw(n) \ S∗
w(n). Then for all s0, t0 whose reduced expressions only contain
parts from vertex semigroups of vertices in C, and all sk, tk ∈ Sw(k), 1 ≤ k < n, we
have ℓ(s0s1 · · · sn) = ℓ(s0) + n and s0s1 · · · sn ⊥ t0t1 · · · tn, unless sk = tk for 0 ≤ k <
n and sn 6⊥ tn. Thus blocking paths allow for the construction of shuffle inert elements
in graph products, which turns out to be quite useful.
3. The internal structure of graph products
In this section we show that many of the properties of SΓ that are of interest to
us, e.g. in connection with [ABLS], can be understood from a study of the corre-
sponding graph products for the coconnected components (Γi)i∈I of Γ. The reason
is SΓ = Li∈I SΓi and the following list of straightforward observations, where we write
s = ⊕i∈I si for s ∈ Li∈I Si:
Proposition 3.1. Let (Si)i∈I be a family of right LCM semigroups. Then S := Li∈I Si
has the following features:
i , Sc = Li∈I(Si)c, and S′
(i) S∗ = Li∈I S∗
(ii) s, t ∈ S are core related if and only if si and ti are core related in Si for all i ∈ I.
(iii) The following statements hold for S if and only if their analogues hold for all Si:
S is core factorable, Sci ⊂ S is ∩-closed, α : Sc y S/∼, a.[s] := [as] is faithful,
and S has finite propagation.
ci = Li∈I(Si)′
ci.
(iv) The action α : Sc y S/∼, a.[s] := [as] is almost free if and only if one of the
following conditions holds:
(a) Si is left reversible for all i ∈ I, that is, S = Sc so that S/ ∼ is a singleton.
(b) There exists a unique i ∈ I such that Si is not left reversible, αi : (Si)c y
Si/∼ is almost free, and Sj is left reversible for all j ∈ I \ {i}.
In view of the direct sum decomposition for graph products over the coconnected
components, we need to understand the behaviour of the graph product in the case of
a coconnected graph with at least two vertices. To do this, we will need to consider a
variant of the action α for S∗, i.e. α∗ : S∗ y S/S∗, x.[s] := [xs]. Also, we will assume
that all vertex semigroups Sv, v ∈ V are nontrivial in order to avoid pathological cases.
For instance, if Γ is the union of a complete graph and an isolated vertex v, and Sv
is trivial, then the graph product will be the direct sum of the right LCM semigroups
attached to the vertices of the complete graph, even though the original graph was larger
and coconnected.
Theorem 3.2. If Γ = (V, E) is coconnected, V ≥ 2, and (Sv)v∈V is a family of
nontrivial right LCM semigroups, then the following assertions hold:
6
NICOLAI STAMMEIER
Γ is the graph product of (S∗
(i) S∗
(ii) For s, t ∈ SΓ, s ∼ t is equivalent to s ∈ tS∗
Γ.
(iii) SΓ is core factorable and (SΓ)ci ⊂ SΓ is ∩-closed.
(iv) The action α : S∗
Γ
(v) The action α is almost free if and only if
v )v∈V , (SΓ)c = S∗
y SΓ/∼ is faithful if and only if SΓ is not a group.
Γ, and (SΓ)ci = SΓ \ S∗
Γ.
v : S∗
(a) α∗
(b) for every connected component U ⊂ V with U ≥ 2, either Su is a group
v is almost free for every isolated vertex v ∈ V , and
y Sv /S∗
v
for all u ∈ U or S∗
u is trivial for all u ∈ U.
v )v∈V to SΓ (resulting from the universal property) is bijective, so that S∗
Proof. For (i), let sv(1)sv(2) · · · sv(n) be a reduced expression for s ∈ SΓ. Clearly, s is
invertible in SΓ if and only if sv(k) ∈ S∗
v(k) for all k. The homomorphism from the graph
product of (S∗
Γ is
the graph product with respect to Γ and (S∗
v )v∈V . Now assume that there is 1 ≤ m ≤ n
such that sv(k) ∈ S∗
v(m). Since Γ is coconnected,
there is w ∈ V with w 6= v(m) and (v(m), w) /∈ E. For every t ∈ Sw \ {1}, we
thus have sv(m)sv(m+1) · · · sv(n) ⊥ tsv(m)sv(m+1) · · · sv(n). By left cancellation, this yields
s ⊥ sv(1)sv(2) · · · sv(m−1)tsv(m)sv(m+1) · · · sv(n), so that s /∈ (SΓ)c. This proves (SΓ)c = S∗
Γ,
and the claims (SΓ)ci = SΓ \ S∗
Γ, (ii), and (iii) are immediate consequences of this.
v(k) for 1 ≤ k < m but sv(m) /∈ S∗
v . Every x ∈ S∗
For (iv), we note that α is not faithful if Sv is a group for all v ∈ V because then
SΓ/∼ is a singleton while S∗
Γ = SΓ is nontrivial. So let us assume that there exists v ∈ V
with Sv 6= S∗
Γ \ {1} has a reduced expression xu(1)xu(2) · · · xu(m) with
xu(k) ∈ S∗
u(k) \ {1}. Since Γ is coconnected and V ≥ 2, there exists a blocking path
w(1), . . . , w(n) for {u(m)} with w(n) = v, see Lemma 2.2. Choose sw(k) ∈ Sw(k) \ {1} for
1 ≤ k < n and sw(n) ∈ Sw(n) \ S∗
w(n). Then s := sw(1)sw(2) · · · sw(n) ∈ SΓ satisfies xu(m)s ⊥
s. If 1 ≤ k ≤ m − 1 satisfies (u(k), u(ℓ)) ∈ E for all k < ℓ ≤ m, then (u(k), u(m)) ∈ E
in particular implies u(k) 6= v(1). For the same reason, (u(k), v(1)) ∈ E implies u(k) 6=
v(2), and so on. Thus xu(1)xu(2) · · · xu(m)sw(1)sw(2) · · · sw(n) is a reduced expression for xs
and we conclude that orthogonality is not destroyed by xu(1)xu(2) · · · xu(m−1), i.e. xs ⊥ s.
In particular, [xs] 6= [s] and therefore α is faithful.
To prove (v), we first observe that (a) is necessary for α to be almost free: If v ∈ V is
isolated, then [xs] = [s] for x ∈ S∗
v \ {1} and [s] ∈ SΓ implies s ∈ Sv. Suppose next that
(b) does not hold, i.e. there exists a connected component U ⊂ V of Γ with U ≥ 2
such that there are u, v ∈ U with Sv 6= S∗
u 6= {1}. If u = v, then we can pick
w ∈ U \ {v} with (v, w) ∈ E. If there is x ∈ S∗
w 6= {1}, then [xs] = [sx] = [s] for all
s ∈ Sv, and since Sv /S∗
v is infinite, α fails to be almost free for x. On the other hand,
Sw is nontrivial, so S∗
w is infinite, and then almost freeness
fails for every x ∈ S∗
w = {1} implies that Sw/S∗
v and S∗
v 6= {1}.
Now suppose u 6= v. As U is connected, we can find a path v(0) := u, v(1), . . . , v(n) :=
v from u to v inside U, i.e. (v(k), v(k + 1)) ∈ E for all 0 ≤ k < n. Then there exists
0 ≤ k < n such that S∗
v(k+1), and we can apply the above
argument to deduce that α is not almost free. We have thus proven that almost freeness
of α implies (a) and (b).
v(k) 6= {1} and Sv(k+1) 6= S∗
Conversely, assume that (a) and (b) hold. If SΓ is a group, then there is nothing to
show, so we may suppose that SΓ 6= S∗
Γ \ {1} be presented by a reduced
expression xu(1)xu(2) · · · xu(m) with xu(k) ∈ S∗
Γ with reduced
expression sv(1) · · · sv(n), sv(k) ∈ Sv(k). Let 1 ≤ j ≤ n be the smallest number such that
u(k) \ {1}. Fix s ∈ SΓ \ S∗
Γ. Let x ∈ S∗
GRAPH PRODUCTS AND THE ABSENCE OF PROPERTY (AR)
7
sv(j) /∈ S∗
v(j). By (b), j is invariant under shuffling and we know that v(j) does not
belong to the connected component of any u(k) that emits an edge. Therefore, xs ⊥ s
and then [xs] 6= [s], unless j = m = 1 and u(1) = v(1) = v for some isolated vertex
v ∈ V .
In this case, (a) says that there are only finitely many fixed points for x in
Sv /S∗
(cid:3)
Remark 3.3. The graph product SΓ for a coconnected graph Γ with V ≥ 2 has finite
propagation if S∗
v . Thus α is almost free if (and only if) (a) and (b) hold.
v is a finite group for all v ∈ V .
Let us now summarise what Proposition 3.1 and Theorem 3.2 imply for graph products
of right LCM semigroups.
Theorem 3.4. Let Γ = (V, E) be a graph and (Sv)v∈V a family of nontrivial right LCM
semigroups. Then:
Γ = Lv∈Vu S∗
v ⊕ Li∈I2 S∗
Γi.
(i) S∗
(ii) (SΓ)c = Lv∈Vu(Sv)c ⊕ Li∈I2 S∗
Γi.
(iii) (SΓ)′
ci ⊕ Li∈I2 SΓi.
(iv) Two elements s, t ∈ SΓ are core related if and only if sv ∼v tv for all v ∈ Vu and
ci = Lv∈Vu(Sv)′
si ∈ tiS∗
Γi for all i ∈ I2.
(v) SΓ is core factorable if and only if Sv is core factorable for every v ∈ Vu.
(vi) (SΓ)ci ⊂ SΓ is ∩-closed if and only if (Sv)ci ⊂ Sv is ∩-closed for every v ∈ Vu.
(vii) The action α : (SΓ)c y SΓ/∼ is faithful if and only if αv : (Sv)c y Sv/∼ is faithful
for every v ∈ Vu, and for every i ∈ I2 there exists w ∈ Vi such that Sw is not a
group.
(viii) The action α : (SΓ)c y SΓ/∼ is almost free if and only if one of the following
conditions holds:
(a) (Sv)c = {1} for all v ∈ Vu and S∗
(b) (Sv)c 6= {1} for a unique v ∈ Vu with αv : (Sv)c → Sv/∼ almost free, while
w = {1} for all w ∈ V \ Vu.
Sw = (Sw)c for all w ∈ Vu \ {v} and Sw′ = S∗
(c) S∗
6= {1} for a unique i ∈ I2 with αi : S∗
Γi
Sw = (Sw)c for all w ∈ Vu and SΓj = S∗
Γj for all j ∈ I2 \ {i}.
w′ for all w′ ∈ V \ Vu.
Γi → SΓi/∼ almost free, while
(ix) SΓ has finite propagation if Sv has finite propagation for every v ∈ Vu and S∗
w is
a finite group for all w ∈ V \ Vu.
The conditions for almost freeness in Theorem 3.4 correspond to (SΓ)c = {1}, (SΓ)c =
∼== SΓi/∼, respectively. Hence they are quite restrictive, and we
(Sv)c, and SΓ/∼
view this as an indication that finite propagation might be much more useful for graph
products than almost freeness of α, see [ABLS, Theorem 4.2(2)] for details.
When applied to right-angled Artin monoids, Theorem 3.4 takes a simpler form:
Corollary 3.5. For a graph Γ = (V, E), the right-angled Artin monid A+
Γ satisfies:
N, and (A+
Γ )1
ci = Li∈I2 A+
Γi.
Γ )c = Lv∈Vu
Γ )∗ = {1}, (A+
(i) (A+
(ii) Two elements s, t ∈ A+
(iii) A+
(iv) The action α : (A+
(v) The action α : (A+
Γ is core factorable, (A+
Γ )c y A+
Γ )c y A+
Γ is trivial or A+
core of A+
Γ are core related if and only if si = ti for all i ∈ I2.
Γ is ∩-closed, and A+
Γ )ci ⊂ A+
Γ has finite propagation.
Γ /∼ is faithful if and only if Γ has no universal vertex.
Γ /∼ is almost free if and only if Vu ∈ {∅, V }, i.e. the
Γ is the free abelian monoid in V .
8
NICOLAI STAMMEIER
4. The absence of property (AR)
In this section, we will prove that for many graph products of right LCM semigroups
SΓ, the only accurate foundation sets are given by elements of S∗
Γ. In particular, we
obtain the an abundance of right LCM semigroups that lack property (AR). Again, the
starting point is a basic observation for direct sums of right LCM semigroups, which
allows us to boil the analysis down to the coconnected case:
Proposition 4.1. Let (Si)i∈I be a family of right LCM semigroups.
property (AR), then Si has property (AR) for all i ∈ I.
Proof. Fix i ∈ I and let S := Li∈I Si. Every foundation set F for Si is a foundation set
for S. Suppose that F has an accurate refinement Fa in S. For s ∈ SΓ, we let s = si + si
with si ∈ Si and si ∈ Lj∈I\{i} Sj. If s ∈ Fa, then {fi ∈ Si f ∈ Fa : fi 6⊥ si} is an
accurate refinement for F inside Si.
Corollary 4.2. If a graph product SΓ has property (AR), then SΓi has property (AR)
for each coconnected component Γi of Γ.
If Li∈I Si has
(cid:3)
Theorem 4.3. Let Γ = (V, E) be a coconnected graph with at least two vertices and
suppose (Sv)v∈V is a family of nontrivial right LCM semigroups.
(i) If Γ is infinite, then every foundation set for SΓ contains an invertible element.
In particular, SΓ has property (AR) and C ∗(SΓ) = Q(SΓ).
(ii) If Γ is finite and E 6= ∅, then the accurate foundation sets for SΓ correspond
Γ. In particular, SΓ has property (AR) if and only if SΓ does not admit a
to S∗
foundation set without invertible elements.
Proof. Both (i) and (ii) hold for trivial reasons if SΓ is a group, so we can assume that
there exists w ∈ V with Sw 6= S∗
w. Let F ⊂ SΓ be a finite subset without invertible
elements. For every f ∈ F , we choose a reduced expression f = fv(1) · · · fv(mf ) with
mf ∈ N× and fv(k) ∈ Sv(k).
Suppose first that Γ is infinite. As f ∈ SΓ \ S∗
Γ, there is a least 1 ≤ kf ≤ mf such that
fv(kf ) /∈ S∗
v(kf ). Then C := {v ∈ V fv(k) ∈ Sv for some f ∈ F, 1 ≤ k ≤ kf }. Then C is
a finite set of vertices so that Lemma 2.2 grants us a blocking path w(1), . . . , w(n) for
C ending in w. If we choose any sk ∈ Sw(k) \ {1} for 1 ≤ k < n and sn ∈ Sw \ S∗
w, then
s1 · · · sn ⊥ f for all f ∈ F as s1 · · · sn ⊥ fv(1) · · · fv(kf ) by construction, see Remark 2.3.
Therefore F is not a foundation set. We conclude that every foundation set for SΓ
contains an invertible element x, which clearly gives an accurate refinement {x}. So
SΓ has property (AR), but if the only accurate foundation sets come from invertible
elements, then the boundary relation Pf ∈F ef SΓ = 1 becomes trivial so that C ∗(S) =
Q(S).
Now let Γ be finite, E 6= ∅, and assume F to be accurate as well. We need to show
that F is not a foundation set. Without loss of generality, we can require that fv(mf ) is
not invertible for all f ∈ F because invertible ends do not play a role when it comes to
intersections of right ideals. Since F does not contain any invertibles, we have ℓ(f ) ≥ 1
for all f ∈ F . Let L := maxf ∈F ℓ(F ), and choose f ∈ F with ℓ(f ) = L. Then we have
f = stv for some v ∈ V, tv ∈ Sv \ {1}, and s ∈ SΓ with ℓ(s) = L − 1. We will first show
that v is isolated, and then use this together with E 6= ∅ to conclude that F cannot be
a foundation set.
GRAPH PRODUCTS AND THE ABSENCE OF PROPERTY (AR)
9
If (v, u) ∈ E for some u ∈ V , we employ Lemma 2.2 to obtain a blocking path
w(1), . . . , w(n) for C := {u} ∪ Nu, and set w(0) := u. Next, choose bk ∈ Sw(k) \ S∗
w(k)
for each 1 ≤ k ≤ n, and let r ∈ Su \ {1}. It then follows that srb ⊥ f for b := b1 · · · bn.
Moreover, we have ℓ(srb) ≥ m + 1. This could be assumed by extending the path
w(0), . . . , w(n) in Γopp, but actually holds true in any case. It then follows that whenever
f ′ ∈ F satisfies f ′
6⊥ f 6= f ′ so that
F would not be accurate. The blocking path then forces f ′ = srb1 · · · bk for some
1 ≤ k ≤ n. However, we then get f ′ ⊥ sr′b for every r′ ∈ Su \ {r}. Since Su is a left
cancellative semigroup that is not a group, it is infinite. Thus there is r ∈ Su \ {1} such
that srb ⊥ f ′ for all f ′ ∈ F .
6⊥ srb, we have srb ∈ f ′SΓ. If sr ∈ f ′Γ, then f ′
We deduce from this that F cannot be a foundation set if there exists f ∈ F with
ℓ(f ) = L that does not end in a part from an isolated vertex. In particular, if Γ does
not have any isolated vertices, no accurate finite subset F without invertible elements
is a foundation set. Now suppose Γ has an isolated vertex v, and let
F ′ := {f ∈ F fv(k) ∈ Sv for some k ⇒ v is not isolated.},
that is, the subset of F consisting of those elements whose reduced expressions do not
contain any part coming from an isolated vertex. As E 6= ∅ and the vertex semigroups
are all nontrivial, the finite accurate set F ′ is also non-empty.
Suppose first that there is f ∈ F ′ with f ∈ Sv \ S∗
v for some v ∈ V . Since F ′ is
accurate and (v, u) ∈ E for some u ∈ V , we have s /∈ f ′SΓ for all s ∈ Su and f ′ ∈ F ′.
Thus we get str ⊥ f ′ for all f ′ ∈ F ′ whenever s ∈ Su, t ∈ Sv, and r ∈ Sw \ S∗
w, compare
Remark 2.3. For f ∈ F \ F ′, we have strtr ⊥ f unless f ∈ strtSΓ because v is isolated
and r is not invertible. Since F is finite while Sw \ S∗
w is infinite, we conclude that there
are s ∈ Su, t ∈ Sv, and r ∈ Sw \ S∗
w such that strtr ⊥ f for all f ∈ F . So F is not a
foundation set.
On the other hand, if we have ℓ(f ) ≥ 2 for every f ∈ F ′, we pick a vertex v that
emits an edge. Then s /∈ f SΓ for all s ∈ Sv, f ∈ F ′, and thus str ⊥ f for all f ∈ F ′
whenever s ∈ Sv, t ∈ Sv, and r ∈ Sw \ S∗
w. As in the previous case, there are s, t, r such
that strtr ⊥ f for all f ∈ F , and thus F is not a foundation set.
Finally, if F is a foundation set for SΓ with F ∩ S∗
F satisfies F ′ ∩ S∗
every foundation set F with x ∈ F ∩ S∗
Γ = ∅, then every refinement F ′ of
Γ = ∅ as well, and thus can never be accurate. On the other hand,
(cid:3)
Γ has an accurate refinement {x}.
We point out that the assumptions in Theorem 4.3 are modest means to avoid the
somewhat pathological cases: SΓ = Sv, the free product SΓ = ∗v∈V Sv, and the graph
product of groups.
Remark 4.4. By Theorem 4.3 (i), foundation sets of SΓ are governed by parts from
the finite coconnected components in the following sense: Let F be a foundation set
for SΓ such that no propert subset of F is a foundation set. If s = sv(1) · · · sv(n) ∈ F
with sv(k) ∈ Sv(k), then sv(k) /∈ S∗
v(k) implies that v(k) ∈ Vi for some finite coconnected
component Γi = (Vi, Ei) of Γ.
Corollary 4.5. Let Γ be a graph and (Sv)v∈V a family of nontrivial right LCM semi-
groups. If there is i ∈ I2 for which Γi = (Vi, Ei) is finite with Ei 6= ∅, Sv is not a group
10
NICOLAI STAMMEIER
for some v ∈ Vi, and there exists a foundation set F for SΓi without invertible elements,
then SΓ does not have property (AR).
Proof. The claim follows from combining Theorem 4.3 with Corollary 4.2.
(cid:3)
The previous results apply nicely to right-angled Artin monoids.
Corollary 4.6. For graph Γ, the following conditions are equivalent:
(1) Every finite coconnected component Γi of Γ is edge-free.
(2) Every finitely generated direct summand of A+
(3) The right-angled Artin monoid A+
Γ is free.
Γ has property (AR).
Proof. The equivalence of (1) and (2) is clear from the direct sum description of A+
Γ in
Subsection 2.2. From Remark 4.4 we infer that it suffices to obtain accurate refinements
of foundation sets F for A+
Γi. But if (2) holds,
then the latter is just a direct sum of finitely generated free monoids, and clearly admits
accurate refinements. So (2) implies (3). Finally, if (3) is valid and Γi = (Vi, Ei) is a
coconnected component of Γ with 2 ≤ Vi < ∞, then {av v ∈ Vi} is a foundation set
for A+
Γi without invertible elements, so Corollary 4.5 forces Ei = ∅, that is, (1) holds. (cid:3)
Γ with F ⊂ Lv∈Vu Sv ⊕ Li∈I2:Vi<∞ A+
By Corollary 4.6, there exist countably many mutually non-isomorphic, finitely gen-
erated right LCM semigroups without property (AR). As a final part of this section, we
address the existence of a generalised scale for right-angled Artin monoids associated
to finite graphs. The existence of a generalised scale turned out to be relevant for a
standardised approach to study KMS-states on the semigroup C ∗-algebra C ∗(A+
Γ ), see
[ABLS]. We first note that free monoids have a generalised scale only if they are finitely
generated and nonabelian, in which case it is unique:
Proposition 4.7. The free monoid F+
generalised scale N : F+
of w ∈ F+
m.
m in 2 ≤ m < ∞ generators admits a unique
m → N× given by N(w) = mℓ(w), where ℓ denotes the word length
Proof. The map N is a generalised scale. On the other hand, let N be a generalised
m = ha1, . . . , ami, and fix 1 ≤ i ≤ m. Then N (ai) > 1 as ai is not part of
scale on F+
m)c = {1}. By definition of N and since ∼ is trivial, the set N −1( N (ai)) is an accurate
(F+
m of cardinality N(ai) that contains ai. If there was 1 ≤ j ≤ m, j 6= i
foundation set for F+
such that N (aj) 6= N(ai), then the foundation set property would give a w ∈ N −1( N (ai))
m. As this forces N (w) ≥ N (ai) N (aj) > N (ai) = N(w), we arrive
such that w ∈ ajaiF+
at a contradiction. Thus N (aj) = N(ai) for all j 6= i. But as {a1, . . . , am} is an accurate
foundation set for F+
(cid:3)
m, we conclude that N (ai) = m for all 1 ≤ i ≤ m.
We call m ∈ Li∈I{k ∈ N 2 ≤ k < ∞} rationally independent if for all distinct
N, the supernatural numbers Qi∈I mki
k, k′ ∈ Li∈I
Proposition 4.8. Let M be a free abelian monoid, m ∈ Li∈I{k ∈ N 2 ≤ k < ∞} for
mi admits a generalised scale N : S → N× if
F+
some nonempty set I. Then S := M ⊕Li∈I
and only if m is rationally independent. In this case, N restricts to the unique generalised
scales Ni on F+
mi, and is therefore unique.
i and Qi∈I mk′
i
i are distinct.
GRAPH PRODUCTS AND THE ABSENCE OF PROPERTY (AR)
11
Proof. As M = Sc = ker N for any generalised scale N on S, see [ABLS, Proposi-
tion 3.6(i)], we can focus on (F+
mi is the free monoid in mi generators,
which we denote by ai,1, . . . , ai,mi. The strategy is to prove that
mi, and
(a) any generalised scale N on S restricts to Ni on F+
(b) the homomorphism N : S → N× arising from (Ni)i∈I is a generalised scale if and
mi)i∈I. Recall that F+
only if m is rationally independent.
For (a), suppose S admits a generalised scale N and fix i ∈ I, 1 ≤ k ≤ mi. Then
N(ai,k) > 1 and there are w1, . . . , wN (ai,k)−1 ∈ S such that {ai,k, w1, . . . , wN (ai,k)−1}
is an accurate foundation set for S contained in N −1(N(ai,k)). Let us decompose wℓ
F+
as wℓ = wℓ ⊕ wℓ ∈ F+
mj(cid:1). Then {ai,k, w1, . . . , wN (ai,k)−1} is a
foundation set for F+
mi with ai,k ⊥ wℓ and N( wℓ) ≤ N(ai,k) for all ℓ. This forces
{ai,k, w1, . . . , wN (ai,k)−1} ⊃ {ai,1, . . . , ai,mi}, and thus N(ai,ℓ) ≤ N(ai,k) for all 1 ≤ ℓ ≤
mi, just like in the proof of Proposition 4.7. As k was arbitrary, we deduce N(ai,k) =
mi = Ni(ai,k) for all i, k.
mi ⊕ (cid:0)Mn ⊕ Lj∈I\{i}
In view of (a), the question behing the main claim becomes: Under which condition
is the homomorphism N : S → N× arising from the family of generalised scales (Ni)i∈I
itself a generalised scale? If m is rationally independent, then every k ∈ N(S) has a
factorization k = Qi∈I mki
i with uniquely determined ki ∈ N. This implies that
N −1(k) = {t ⊕ L
i∈I
wi t ∈ M, wi ∈ F+
mi with ℓi(wi) = ki}.
N, k 6= k′ such that K := Qi∈I mki
Therefore, N −1(k)/∼ = k, and any transversal of N −1(k)/∼ is an accurate foundation
set for S, that is, N is a generalised scale. On the other hand, if there are k, k′ ∈
i , then both k and k′ yield a set of
Li∈I
K mutually orthogonal elements s1, . . . , sK ∈ S and t1, . . . , tK ∈ S, respectively, with
N(sj) = K = N(tj) for all j. Since there is i ∈ I with ki 6= k′
i, the i-th components
of sj and tj ′ have different length for all j, j′. Thus sj 6∼ tj ′ for all j, j′, and we get
N −1(K)/∼ ≥ 2K. Therefore N is not a generalised scale in this case.
(cid:3)
i = Qi∈I mk′
i
We can now state our conclusions for right-angled Artin monoids.
Corollary 4.9. For every graph Γ, the right-angled Artin monoid A+
Γ admits a gener-
alised scale N if and only if Vu 6= V , all coconnected components Γi = (Vi, Ei) are finite
and edge-free, and Li∈I2Vi is rationally independent. In this case, N is unique.
Proof. The condition Vu 6= V is equivalent to saying that A+
So if all coconnected components Γi = (Vi, Ei) are finite and edge-free, then A+
Γ
Lv∈Vu
scale N if and only if Li∈I2Vi is rationally independent.
Γ is non-abelian, i.e.I2 6= ∅.
∼=
Γ has a (unique) generalised
F+
Vi. Hence, Proposition 4.8 implies that A+
N⊕Li∈I2
Conversely, suppose A+
Γ admits a generalised scale N. Since N is a nontrivial ho-
momorphism with ker N = Lv∈Vu
N, we need to have Vu 6= V so that the set I2 is
non-empty. Moreover, A+
Γ has property (AR) by [ABLS, Proposition 3.6], so Corol-
lary 4.6 implies that all finite coconnected components Γi of Γ are edge-free. If there
was an infinite coconnected component Γi = (Vi, Ei), then 1 < N(av) < ∞ for all v ∈ Vi,
and the defining property of a generalised scale would yield an accurate foundation set
12
NICOLAI STAMMEIER
of the form {av, f1, . . . , fN (av )−1} for suitable fk ∈ A+
mark 4.4, and we conclude that Γi is finite for all i ∈ I2. But then A+
Proposition 4.8, and it follows that Li∈I2Vi is rationally independent.
Γ . However, this contradicts Re-
Γ is covered by
(cid:3)
Corollary 4.10. For every graph Γ, the right-angled Artin monoid A+
and only if it admits a generalised scale.
Γ is admissible if
Proof. According to Corollary 3.5 (iii), A+
Γ is ∩-
closed, no matter what Γ is. By Corollary 4.9, the conditions characterising the existence
of a generalised scale N include rational independence of Li∈I2Vi. This feature implies
Irr(N(A+
Γ ), which is the last
extra condition for admissibility.
(cid:3)
Γ )) = {Vi i ∈ I2} and that this set freely generates N(A+
Γ is core factorable and (A+
Γ )ci ⊂ A+
References
[ABLS] Zahra Afsar, Nathan Brownlowe, Nadia S. Larsen, and Nicolai Stammeier, Equilibrium
states on right LCM semigroup C ∗-algebras. preprint, arxiv:1611.01052.
[BaHLR12] Nathan Brownlowe, Astrid an Huef, Marcelo Laca, and Iain Raeburn, Boundary quotients
of the Toeplitz algebra of the affine semigroup over the natural numbers, Ergodic Theory
Dynam. Systems 32 (2012), no. 1, 35–62, DOI 10.1017/S0143385710000830.
[BLS17] Nathan Brownlowe, Nadia S. Larsen, and Nicolai Stammeier, On C ∗-algebras associ-
ated to right LCM semigroups, Trans. Amer. Math. Soc. 369 (2017), no. 1, 31–68, DOI
10.1090/tran/6638.
[BRRW14] Nathan Brownlowe, Jacqui Ramagge, David Robertson, and Michael F. Whittaker, Zappa-
Sz´ep products of semigroups and their C ∗-algebras, J. Funct. Anal. 266 (2014), no. 6,
3937–3967, DOI 10.1016/j.jfa.2013.12.025.
[BS16] Nathan Brownlowe and Nicolai Stammeier, The boundary quotient for algebraic dynamical
systems, J. Math. Anal. Appl. 438 (2016), no. 2, 772–789, DOI 10.1016/j.jmaa.2016.02.015.
[Cha07] Ruth Charney, An introduction to right-angled Artin groups, Geom. Dedicata 125 (2007),
141–158, DOI 10.1007/s10711-007-9148-6.
[CaHR16] Lisa Orloff Clark, Astrid an Huef, and Iain Raeburn, Phase transitions on the Toeplitz
algebras of Baumslag-Solitar semigroups, Indiana Univ. Math. J. 65 (2016), no. 6, 2137 –
2173, DOI 10.1512/iumj.2016.65.5934.
[VdC01] Ant´onio Veloso da Costa, Graph products of monoids, Semigroup Forum 63 (2001), no. 2,
247–277, DOI 10.1007/s002330010075.
[CL02] John Crisp and Marcelo Laca, On the Toeplitz algebras of right-angled and finite-type Artin
groups, J. Aust. Math. Soc. 72 (2002), no. 2, 223–245, DOI 10.1017/S1446788700003876.
, Boundary quotients and ideals of Toeplitz C ∗-algebras of Artin groups, J. Funct.
[CL07]
Anal. 242 (2007), no. 1, 127–156, DOI 10.1016/j.jfa.2006.08.001.
[Die90] Volker Diekert, Combinatorics on traces, Lecture Notes in Computer Science, vol. 454,
Springer-Verlag, Berlin, 1990. With a foreword by Wilfried Brauer.
[ELR16] Søren Eilers, Xin Li, and Efren Ruiz, The isomorphism problem for semigroup C ∗-algebras
of right-angled Artin monoids, Doc. Math. 21 (2016), 309–343. electronic version.
[EP17] Ruy Exel and Enrique Pardo, Self-similar graphs, a unified treatment of Katsura
and Nekrashevych C ∗-algebras, Adv. Math. 306 (2017), no. 1, 1046–1129, DOI
10.1016/j.aim.2016.10.030.
[EP16]
, The tight groupoid of an inverse semigroup, Semigroup Forum 92 (2016), no. 1,
274–303, DOI 10.1007/s00233-015-9758-5.
[FK09] John Fountain and Mark Kambites, Graph products of right cancellative monoids, J. Aust.
Math. Soc. 87 (2009), no. 2, 227–252, DOI 10.1017/S144678870900010X.
[Gre90] Elisabeth R. Green, Graph products of groups, University of Leeds, 1990.
GRAPH PRODUCTS AND THE ABSENCE OF PROPERTY (AR)
13
[LR10] Marcelo Laca and Iain Raeburn, Phase transition on the Toeplitz algebra of the affine
semigroup over the natural numbers, Adv. Math. 225 (2010), no. 2, 643–688, DOI
10.1016/j.aim.2010.03.007.
[LRR11] Marcelo Laca, Iain Raeburn, and Jacqui Ramagge, Phase transition on Exel crossed prod-
ucts associated to dilation matrices, J. Funct. Anal. 261 (2011), no. 12, 3633–3664, DOI
10.1016/j.jfa.2011.08.015.
[LRRW14] Marcelo Laca, Iain Raeburn, Jacqui Ramagge, and Michael F. Whittaker, Equilibrium
states on the Cuntz-Pimsner algebras of self-similar actions, J. Funct. Anal. 266 (2014),
no. 11, 6619–6661, DOI 10.1016/j.jfa.2014.03.003.
[Li12] Xin Li, Semigroup C ∗-algebras and amenability of semigroups, J. Funct. Anal. 262 (2012),
no. 10, 4302–4340, DOI 10.1016/j.jfa.2012.02.020.
[Li13]
, Nuclearity of semigroup C*-algebras and the connection to amenability, Adv. Math.
244 (2013), 626–662, DOI 10.1016/j.aim.2013.05.016.
[Sta] Nicolai Stammeier, A boundary quotient diagram for right LCM semigroups, to appear in
Semigroup Forum. arxiv:1604.03172.
[Star15] Charles Starling, Boundary quotients of C ∗-algebras of right LCM semigroups, J. Funct.
Anal. 268 (2015), no. 11, 3326–3356, DOI 10.1016/j.jfa.2015.01.001.
[Ste16] Benjamin Steinberg, Simplicity, primitivity and semiprimitivity of ´etale groupoid algebras
with applications to inverse semigroup algebras, J. Pure Appl. Algebra 220 (2016), no. 3,
1035–1054, DOI 10.1016/j.jpaa.2015.08.006.
Department of Mathematics, University of Oslo, P.O. Box 1053 Blindern, NO-0316
Oslo, Norway
E-mail address: [email protected]
|
1111.6927 | 3 | 1111 | 2012-11-15T20:43:55 | C*-algebras for categories of paths asociated to the Baumslag-Solitar groups | [
"math.OA"
] | In this paper we describe the C*-algebras associated to the Baumslag-Solitar groups with the ordering defined by the usual presentations. These are Morita equivalent to the crossed product C*-algebras obtained by letting the group act on its directed boundary. We use the method of categories of paths to define the algebras, and to deduce the presentation by generators and relations. We obtain a complete description of the Toeplitz algebras, and we compute the K-theory of the Cuntz-Kreiger algebras. | math.OA | math |
C ∗-ALGEBRAS FOR CATEGORIES OF PATHS ASSOCIATED TO THE
BAUMSLAG-SOLITAR GROUPS
JACK SPIELBERG
Abstract. In this paper we describe the C ∗-algebras associated to the Baumslag-Solitar groups with the
ordering defined by the usual presentations. These are Morita equivalent to the crossed product C ∗-algebras
obtained by letting the group act on its directed boundary. We use the method of categories of paths to
define the algebras, and to deduce the presentation by generators and relations. We obtain a complete
description of the Toeplitz algebras, and we compute the K-theory of the Cuntz-Kreiger algebras.
1. Introduction
In this paper we study the C ∗-algebras associated to the Baumslag-Solitar groups with the ordering
defined by the usual presentation. We show that in most cases this is a quasi-lattice ordering in the sense of
[9]. However we use the notion of category of paths of [13] to describe the C ∗-algebras. This is a construction
of Toeplitz and Cuntz-Krieger algebras in a very general setting that includes ordered groups, higher-rank
graphs, and many other examples of C ∗-algebras obtained from oriented combinatorial objects. The chief
virtue of this construction is that a presentation of the algebras by generators and relations is obtained
naturally from the category, eliminating the guesswork typically needed for their identification. For example,
if a free semigroup acts on its ℓ2-space, it is easy to observe that the generators define isometries having
pairwise orthogonal ranges spanning a codimension-one subspace. The rank-one projection onto this subspace
generates the compact operators when pushed around by the isometries, and in the quotient, the classes of
isometries have range projections adding to the identity. Cuntz's theorem ([4]) shows that this quotient
algebra is uniquely defined by this presentation. For other semigroups, however, it may not be obvious what
is the correct quotient, even if the compact operators are present in the algebra generated by the isometries
on ℓ2. Moreover, it may also be unclear what relations ought to be used to define the quotient. In the present
case, we obtain generators and relations for a C ∗-algebra Morita equivalent to the crossed product algebra
associated to the action of a Baumslag-Solitar group on its directed boundary. These presentations turn out
to coincide with certain examples obtained by Katsura in his work on topological graphs ([6]). Our method
gives a new approach to the description of these algebras by generators and relations, and also gives the
ideal structure of the Toeplitz versions. In the case of the solvable examples BS(1, d), this is a well-known
example with core isomorphic to a Bunce-Deddens algebra ([1]). The presentations of the Baumslag-Solitar
groups can be thought of as arising from their status as the fundamental group of a graph of groups: the
graph consists of one vertex and one edge, with infinite cyclic groups attached to both vertex and edge. The
Bass-Serre theory requires the choice of an orientation for this graph; this amounts to a direction on the
Cayley graph ([12]). In this case, the directed Cayley graph is a category of paths, and we use the theory
developed in [13] to study the action of the group on the boundary (see [13], Example 8.5).
The organization of this paper is as follows. We first use the HNN structure of a Baumslag-Solitar group
to prove that, as an ordered group (relative to the subsemigroup defined by the presentation), it is suitable
for our constructions. In the process we define an associated odometer-like action by the second generator.
We then identify the maximal directed hereditary subsets of the category, and hence the boundary of the
semigroup. We prove that the restriction of the groupoid to the boundary of the category is amenable. We
then identify all directed hereditary subsets, and prove that the entire groupoid is amenable. From this
we deduce the generators and relations for the Cuntz-Krieger algebra. We then compute the K-theory of
the algebra by first studying the fixed-point algebra by the gauge action, and then applying the Pimsner-
Voiculescu exact sequence. We mention that the degree functors for these categories are degenerate, and
Date: 28 November 2011.
2010 Mathematics Subject Classification. Primary 46L05; Secondary 46L80, 46L55.
Key words and phrases. Baumslag-Solitar group, Cuntz-Krieger algebra, Toeplitz Cuntz-Krieger algebra, K-theory.
1
2
JACK SPIELBERG
thus the fixed-point algebras are not AF. Finally we use theorems from [13] to establish the fundamental
structural properties of these algebras.
We conclude this introduction with a description of the definitions and results from [13] needed for the
rest of the paper. We identify the objects of a category with the identity morphisms in that category, and
we use juxtaposition to indicate composition of morphisms. Morphisms are referred to as paths, and objects
as vertices. We use s and r to denote the source and range of morphisms, and Λ0 for the vertices in the
category Λ. A category of paths is a small category satisfying
only consider (certain) finitely aligned examples.
(1) αβ = αγ implies β = γ (left-cancellation).
(2) βα = γα implies β = γ (right-cancellation).
(3) αβ = s(β) implies α = β = s(β) (no inverses).
For any α ∈ Λ we define the left shift σα : αΛ → s(α)Λ by σα(αβ) = β (σα is well-defined by left-
cancellation). The right shift map β ∈ s(α)Λ 7→ αβ ∈ r(α)Λ is the inverse of σα. We say that β extends
α if there exists α′ ∈ Λ such that β = αα′. (We may express this by writing β ∈ αΛ.) It follows from the
definition that this is a partial order on Λ. If β is an extension of α, we call α an initial segment of β. The
set of initial segments of β is denoted [β]. We write α ⋓ β (α meets β) if αΛ ∩ βΛ 6= ∅, and α ⊥ β (α is
disjoint from β) otherwise. We let α ∨ β denote the set of minimal common extensions of α and β, i.e. the
of the elements of F . We say that Λ is finitely aligned if for every pair of elements α, β ∈ Λ, there is a finite
minimal elements of αΛ ∩ βΛ. For a subset F ⊆ Λ we letW F denote the set of minimal common extensions
subset G of Λ such that αΛ ∩ βΛ =Sε∈G εΛ. It follows that we may take G = α ∨ β. In this paper we will
A subset C ⊆ Λ is directed if for all α, β ∈ C there is γ ∈ C extending both α and β. C is hereditary
if [α] ⊆ C for every α ∈ C. The collection of all directed hereditary subsets of Λ is denoted Λ∗; the set of
maximal elements of Λ∗ is denoted Λ∗∗. A directed hereditary set is finite if it contains a maximal element;
in this case it must be of the form [α] for some α ∈ Λ. Otherwise it is infinite. We define a topology on Λ∗
as follows. For α ∈ Λ, and β1, . . ., βn ∈ αΛ \ {α}, let E = αΛ \Sn
Then the bE form a base of compact-open sets for a locally compact Hausdorff topology. The boundary of Λ
bE = {C ∈ Λ∗ : E ⊇ C ∩ γΛ for some γ ∈ C}.
We define a groupoid with unit space Λ∗ as follows. First define a relation on Λ × Λ × Λ∗ by (α, β, x) ∼
(α′, β′, x′) if there are y ∈ Λ∗ and γ, γ′ ∈ Λ such that x = γy, x′ = γ′y, αγ = α′γ′, and βγ = β′γ′. Then ∼ is
an equivalence relation. The set of equivalence classes becomes a locally compact Hausdorff ´etale groupoid,
where the set of composable pairs is
is defined to be ∂Λ = Λ∗∗.
i=1 βiΛ. Set
G2 =(cid:8)(cid:0)[α, β, x], [γ, δ, y](cid:1) : βx = γy(cid:9),
y = ηz, and βξ = γη. Then
and inversion is given by [α, β, x]−1 = [β, α, x]. Multiplication G2 → G is given as follows. Let (cid:0)[α, β, x],
[γ, δ, y](cid:1) ∈ G2. Since βx = γy, it follows from Lemma 4.12 of [13] that there are z, ξ, and η such that x = ξz,
A base of compact-open sets for G is given by the sets [α, β, bE] = {[α, β, x] : x ∈ bE}.
The C ∗-algebras of Λ are defined as T C ∗(Λ) = C ∗(G) (the Toeplitz C ∗-algebra), and C ∗(Λ) = C ∗(G∂Λ)
(the Cuntz-Krieger algebra). We have the following theorem giving generators and relations for T C ∗(Λ)
([13], Theorem 6.3).
[α, β, x] [γ, δ, y] = [αξ, δη, z].
Theorem 1.1. Let Λ be a finitely aligned category of paths. The representations of T C ∗(Λ) are in one-to-one
correspondence with the families {Tα : α ∈ Λ} of Hilbert space operators satisfying the relations
αTα = Ts(α).
(1) T ∗
(2) TαTβ = Tαβ, if s(α) = r(β).
(3) TαT ∗
γ .
β =Wγ∈α∨β TγT ∗
αTβT ∗
In order to describe the analogous theorem for C ∗(Λ), we need the notion of exhaustive set. Let v ∈ Λ0.
A subset F ⊆ vΛ is exhaustive (at v) if for every α ∈ vΛ there exists β ∈ F such that α ⋓ β. The exhaustive
sets can be used to characterize the points of ∂Λ ([13], Theorem 7.8). We have the following theorem giving
generators and relations for C ∗(Λ) ([13], Theorem 8.2).
C ∗-ALGEBRAS FOR CATEGORIES OF PATHS ASSOCIATED TO THE BAUMSLAG-SOLITAR GROUPS
3
Theorem 1.2. Let Λ be a countable finitely aligned category of paths. Assume that G is amenable. The
representations of C ∗(Λ) are in one-to-one correspondence with the families {Sα : α ∈ Λ} of Hilbert space
operators satisfying the relations
αSα = Ss(α).
(1) S∗
(2) SαSβ = Sαβ, if s(α) = r(β).
(3) SαS∗
γ.
αSβS∗
β =Wγ∈α∨β SγS∗
(4) Sv =Wβ∈F SβS∗
β if F is finite exhaustive at v. (Equivalently, 0 =Qδ∈F (Sv − SδS∗
The last theorem requires amenability of the groupoid. The usual way of obtaining this is to decompose
C ∗(Λ) by means of a cocycle to a discrete abelian group (called a degree functor in [13], Definition 9.1). We
cite the following summary of the standard argument ([13], Proposition 9.3).
δ ).)
Proposition 1.3. Let G be a locally compact Hausdorff ´etale groupoid, Q a countable abelian group, and
c : G → Q a continuous homomorphism. Let Gc = c−1(0), also a locally compact Hausdorff ´etale groupoid.
Suppose that Gc is amenable. Then G is amenable.
An ordered group (Γ, Λ) is called quasi-lattice ordered if for each t ∈ ΛΛ−1 there is an element α ∈ tΛ ∩ Λ
such that tΛ ∩ Λ = αΛ ([9]). This idea was generalized in [13], Definition 8.6, as follows. The ordered group
(Γ, Λ) is called finitely aligned if for each t ∈ Γ there is a finite set F ⊆ tΛ ∩ Λ such that tΛ ∩ Λ =Sα∈F αΛ.
It is possible that Λ is a finitely aligned category of paths even if (Γ, Λ) is not a finitely aligned ordered
group. A weaker notion is given in [13], Definition 8.11: (Γ, Λ) is locally finitely exhaustible if for each t ∈ Γ,
there is a finite set F ⊆ tΛ ∩ Λ such that every element of tΛ ∩ Λ meets some element of F . In Lemma 8.10
of [13] it is shown how to define a locally compact Γ-space, ∂(Γ, Λ), which is the directed boundary of the
ordered group. The following appears in [13] as part of Theorem 8.13 and Corollary 8.17.
Theorem 1.4. Let (Γ, Λ) be a countable ordered group. Suppose that Λ is finitely aligned as a category of
paths. Then ∂(Γ, Λ) is locally compact Hausdorff if and only if (Γ, Λ) is locally finitely exhaustible. Moreover,
C ∗(Λ) is Morita equivalent to the crossed product algebra C0(∂(Γ, Λ)) × Γ.
2. The category of paths of a Baumslag-Solitar group
equivalent relation, we may as well assume that c and d are not both negative. We will consider separately
the cases cd > 0 and cd < 0. Thus we formulate the situation for a pair of positive integers c and d, and
let the group be defined by the relation abc = bda, or by the relation abc = b−da. The case cd > 0 will be
further divided accordingly as c ≥ d or c < d, giving three cases overall:
For nonzero integers c and d we consider the group Γ = ha, b(cid:12)(cid:12) abc = bdai ([3]). Since ab−c = b−da is an
Case (BS1): Γ = ha, b(cid:12)(cid:12) abc = bdai where c ≥ d ≥ 1.
Case (BS2): Γ = ha, b(cid:12)(cid:12) abc = bdai where d > c ≥ 1.
Case (BS3): Γ = ha, b(cid:12)(cid:12) abc = b−dai where c, d ≥ 1.
We denote by θ : G → Z the homomorphism given by θ(a) = 1 and θ(b) = 0. We sometimes refer to
θ(g) as the height of g. Let Λ be the submonoid generated by a and b, and let B = {bi : i ≥ 0} denote the
submonoid generated by b. The following may be found in [11].
Proposition 2.1. Each element of Γ has a unique represention in the form bi1 aε1 · · · bin aεn bq, where εµ ∈
{±1}, iµ ∈ [0, d) if εµ = +1, and iµ ∈ [0, c) if εµ = −1.
The standard from of the proposition is obtained by moving b's to the right via bkda = abkc and bkca−1 =
a−1bkd.
Corollary 2.2. Let t ∈ Γ have the form in Proposition 2.1. Then t ∈ Λ if and only if εµ = +1 for all µ,
and q ≥ 0 in cases (BS1) and (BS2), and in case (BS3) if n = 0.
We note the following proposition for later use.
It follows from the same kind of arguments as gives
Proposition 2.1 and its corollary.
Proposition 2.3. Each element α ∈ Λ has unique representations in the two forms
(L) α = bi0 abi1 a · · · bik−1 abp, iµ ∈ [0, d), p ∈ Z;
(R) α = bqabj1abj2 · · · abjk , jµ ∈ [0, c), q ∈ Z,
4
JACK SPIELBERG
where k = θ(α).
Remark 2.4. In cases (BS1) and (BS2), in forms (L) and (R) we have p, q ≥ 0. In case (BS3), if θ(α) = 0
we have p = q ≥ 0. It also follows that b has infinite order in G.
The following corollary follows easily.
Corollary 2.5. If 0 < k < θ(α), or if k = θ(α) > 0 and we are in case (BS1) or (BS2), then α has a unique
initial segment of height k in the form (L) of Proposition 2.3 with p = 0. If k = θ(α) > 0 and we are in case
(BS3), then there is a unique element α0 in the form (L) of Proposition 2.3 with p = 0, and p0 ∈ Z, such
that every initial segment of height k has form (L) equal to α0bq with q ≤ p0.
Lemma 2.6. In case (BS3), if α ∈ Λ and θ(α) > 0, then αbp ∈ Λ for all p ∈ Z. Moreover, B ⊆ [α].
Proof. We can write α = α0abm for α0 ∈ Λ and m ≥ 0. Choose n so that nc + m + p ≥ 0. Then using the
relation a = bdabc, we find that αbp = α0abm+p = α0bndabnc+m+p ∈ Λ. For the second statement, we write
α = biaα1 for some i ∈ N and α1 ∈ Λ. Then α = bi+mdabmcα1 for all m ∈ N.
(cid:3)
The following begins our study.
Lemma 2.7. Λ is a category of paths ([13], Example 8.5).
Proof. Note that θ(α) ≥ 0 for all α ∈ Λ. We note that if α, β ∈ Λ are such that αβ = e, then θ(α)+θ(β) = 0,
and hence θ(α) = θ(β) = 0. But then α, β ∈ B. But then we must have α = β = e. Hence Λ is a category
of paths.
(cid:3)
We will prove that Λ is finitely aligned, and that in cases (BS1) and (BS2), the ordered group (Γ, Λ) is
quasi-lattice ordered in the sense of [9] (Theorem 2.11 below). (The case (BS3) is slightly different -- see
Lemma 2.12.) The argument varies by case. We require lemmas for cases (BS1) and (BS2).
Lemma 2.8. Suppose that we are in case (BS2). Fix k ≥ 1. For i = (i0, . . . , ik) ∈ [0, c)k+1, let α(i) =
bi0abi1 a · · · bik a. There are maps ψ : [0, c)k+1 → [0, c)k+1 and r : [0, c)k+1 → Z+ such that
(1) bdr(i)α(ψ(i)) = α(i)bc and ψ(i)0 = i0.
(2) If q < c then bi0+1 6∈ [α(i)bq].
(3) If i0 < m ≤ i0 + dr(i), then α(i)bc = bdr(i)α(ψ(i)) is the unique minimal common extension of bm
and α(i).
(4) For h ≥ 0, if
then
i0 + d
h−1Xµ=0
r(ψµ(i)) < m ≤ i0 + d
r(ψµ(i))
hXµ=0
α(i)b(h+1)c = bd Ph
µ=0 r(ψµ(i))α(ψh+1(i))
is the unique minimal common extension of bm and α(i).
Proof. For (1), the maps ψ and r are defined by the unique form (R) of α(i)bc. That r(i) ≥ 1 follows from
the assumption c < d. It follows from the defining relation of G that if i0 = 0 then α(i)0 = 0. Hence by
the uniqueness of form (L), ψ(i)0 = i0 for all i. For (2), note that if q < c then α(i)bq is already in form
(R). We next prove (3). Let m be as in the statement. By (1) we know that α(i)bc ∈ bmΛ ∩ α(i)Λ. Let
β ∈ bmΛ ∩ α(i)Λ. Then β = α(i)γ for some γ. Write γ = bqabℓ1 · · · abℓn with ℓµ ∈ [0, c). Then the form (R)
of β equals (form (R) of α(i)bq) · abℓ1 · · · abℓn. Therefore bm ∈ [β] if and only if bm ∈ [α(i)bq]. This occurs if
and only if q ≥ c, by (2). Therefore β ∈ α(i)bcΛ. Finally, we prove (4), by induction on h. The case h = 0
is (3). Suppose it is true for h, and let
i0 + d
hXµ=0
r(ψµ(i)) < m ≤ i0 + d
r(ψµ(i)).
h+1Xµ=0
First we note that
Thus α(i)b(h+2)c is a common extension of bm and α(i).
α(i)b(h+2)c = bd Ph
µ=0 r(ψµ(i))α(ψh+1(i))bc = bd Ph+1
µ=0 r(ψµ(i))α(ψh+2(i)) ∈ bmΛ.
C ∗-ALGEBRAS FOR CATEGORIES OF PATHS ASSOCIATED TO THE BAUMSLAG-SOLITAR GROUPS
5
Now let β ∈ bmΛ ∩ α(i)Λ be any common extension. Then
β ∈ bm−dr(ψh(i))Λ ∩ α(i)Λ.
µ=0 r(ψµ(i)) < m − dr(ψh(i)), the inductive hypothesis implies that β ∈ α(i)b(h+1)cΛ. Define
Since i0 + dPh−1
β′ by
β = bd Ph
µ=0 r(ψµ(i))β′.
Then β′ ∈ bi0+1Λ ∩ α(ψ(h+1)(i))Λ. By (3), β′ ∈ α(ψh+1(i))bcΛ. Therefore β ∈ α(i)b(h+2)cΛ.
(cid:3)
We next consider case (BS1). We will use the following notation: for i ∈ [0, d)N and k ∈ N, let αk(i) =
bi0a · · · bik a.
Lemma 2.9. Suppose that we are in case (BS1). There are maps φ : [0, d)N → [0, d)N and r : [0, d)N → (Z+)N
such that for all k ∈ N,
(1) bdαk(φ(i)) = αk(i)bcr(i)k. φ and r are uniquely determined by this condition, and φ(i)k, r(i)k depend
only on i0, . . ., ik.
(2) If q < cr(i)k then bi0+1 6∈ [αk(i)bq].
(3) If i0 < m ≤ i0 + d then αk(i)bcr(i)k is the unique minimal common extension of αk(i) and bm.
(4) If i0 + (h − 1)d < m ≤ i0 + hd then αk(i)bc Ph−1
µ=0 r(φµ(i))k = bhdαk(φh(i)) is the unique minimal
common extension of αk(i) and bm.
(5) The map φ is a homeomorphism (for the product topology on [0, d)N).
Proof. (1): We define φ and r inductively. Put φ(i)0 = i0 and r(i)0 = 1. Then
bdα0(φ(i)) = bdbi0 a = bi0 abc = α0(i)bcr(i)0.
It is clear that φ(i)0 and r(i)0 are determined uniquely. Suppose that φ(i)µ and r(i)µ have been defined for
µ < k so that bdαµ(φ(i)) = αk(i)bcr(i)µ . Define φ(i)k ∈ [0, d) by
Then there is r(i)k ≥ 0 such that
(∗)
We note that
φ(i)k ≡ ik − cr(i)k−1
(mod d).
φ(i)k − dr(i)k = ik − cr(i)k−1.
and hence that r(i)k ≥ 1. Now,
dr(i)k = φ(i)k − ik + cr(i)k−1 ≥ φ(i)k − ik + c > φ(i)k,
bdαk(φ(i)) = bdαk−1(φ(i))bφ(i)k a = αk−1(i)bcr(i)k−1bφ(i)k a = αk−1(i)bik+dr(i)k a = αk(i)bcr(i)k .
It is clear that φ(i)k and r(i)k are uniquely determined by the equation bφ(i)k+cr(i)k−1a = bik abcr(i)k. More-
over, the construction shows that they depend only on i0, . . ., ik.
(2): We again prove this by induction on k. Let q < cr(i)0 = c. Then α0(i)bq = bi0 abq is in form (R),
and hence does not extend bi0+1. Suppose that the result is true for k − 1. If q < cr(i)k, write q = sc + t,
with t ∈ [0, c). Then s < r(i)k. We have
αk(i)bq = αk−1(i)bik abscbt = αk−1(i)bik bsdabt.
By (∗), ik + sd ≤ ik + (r(i)k − 1)d = ik + r(i)kd − d = φ(i)k + cr(i)k−1 − d < cr(i)k−1. Then the inductive
hypothesis implies that αk−1(i)bik+sd 6∈ bi0+1Λ. By uniqueness of form (R), it follows that αk(i)bq 6∈ bi0+1Λ.
(3): Let β ∈ αk(i)Λ. Write σαk(i)β = bqabe1 · · · abeℓ in form (R), so eµ ∈ [0, c). By uniqueness of form
(R), β ∈ bmΛ if and only if αk(i)bq ∈ bmΛ. By (1) and (2) this occurs if and only if q ≥ cr(i)k, that is, if
and only if β ∈ αbcr(i)k Λ.
(4): The case h = 1 is given by (3). Suppose the result is true for h − 1. Let i0 + (h − 1)d < m ≤ i0 + hd.
The inductive hypothesis implies that b(h−1)dαk(φh−1(i)) is the minimal common extension of bm−d and
αk(i). Let β ∈ bmΛ ∩ αk(i)Λ. Then β ∈ b(h−1)dΛ ∩ αk(i)Λ. Therefore β ∈ b(h−1)dαk(φh−1(i)). Writing
β = b(h−1)dβ′, we have that β′ ∈ bm−(h−1)dΛ ∩ αk(φh−1(i))Λ. Since i0 < m − (h − 1)d ≤ i0 + d, (3) implies
that β′ ∈ αk(φh−1(i))bcr(φh−1(i))k = bdαk(φh(i)). Therefore β ∈ bhdαk(φh(i)). Since bhdαk(φh(i)) extends
bm and αk(i), the result is proved.
6
JACK SPIELBERG
(5): Given j ∈ [0, d)N, the unique form (L) gives i ∈ [0, d)N and r ∈ [0, d)Z+
such that bdαk(j) = αk(i)bcrk .
It follows that j = φ(i), and so φ is onto. By the uniqueness of forms (L) and (R), j and i uniquely
determine each other, so that φ is bijective. The continuity of φ follows from the last part of (1). Since
[0, d)N is compact, φ is a homeomorphism.
(cid:3)
Proposition 2.10. Let α = be0 a · · · bes abes+1 and β = bf0a · · · bftabft+1 in form (L), i.e. with eµ, fν ∈ [0, d)
for µ ≤ s and ν ≤ t. Then α ⋓ β if and only if eµ = fµ for µ ≤ min{s, t}. In this case, α and β have a
unique minimal common extension. (In particular, Λ is finitely aligned.)
Proof. The "only if" statement follows from Corollary 2.5. For the converse, suppose that s ≤ t and eµ = fµ
for µ ≤ s. We first consider case (BS2). Let bfs+1a · · · bft+1 = bvabi1 · · · bik abq in form (R), i.e. with
q, iµ ∈ [0, c) for 1 ≤ µ ≤ k. If es+1 ≤ v, then β ∈ αΛ, and the result holds. Suppose instead that v < es+1.
Applying σbe0 a···bes abv
to α and β, and letting m = es+1 − v, we find that it is enough to show that the paths
bm and abi1a · · · bik abq meet, and have a unique minimal common extension. Choose h as in Lemma 2.8(4).
Then abi1 a · · · bik ab(h+1)c is the unique minimal common extension of bm and abi1 a · · · bik a. Since q < c, this
is also the unique minimal common extension of bm and abi1a · · · bik abq.
Now we consider case (BS1). If es+1 ≤ fs+1 then β ∈ αΛ, and the result holds. Suppose instead that
es+1 > fs+1. Applying σbe0 a···bes abfs+1 to α and β, and letting m = es+1 − fs+1, k = t − s − 1, iµ = fµ+s+1
for 1 ≤ µ ≤ k, and q = ft+1, it is enough to show that bm and abi1 · · · bik abq meet and have a unique minimal
common extension. Choose h such that (h − 1)d < m ≤ hd. By Lemma 2.9 there are j1, . . ., jk ∈ [0, d),
and r ≥ 0, such that bhdabj1 · · · bjk a = abi1 · · · bik abcr is the unique minimal common extension of bm and
abi1 · · · bik a. If q ≤ cr this is also the unique minimal common extension of bm and abi1 · · · bik abq. If q > cr,
then abi1 · · · bik abq is the unique minimal common extension.
Finally we consider case (BS3). If s < t, choose m ≥ 0 such that fs+1 − es+1 + md ≥ 0. Then
β = be0 a · · · besabfs+1a · · · = αbfs+1−es+1a · · · = αbfs+1−es+1+mdabmc · · · ∈ αΛ.
If s = t, then β ∈ αΛ if es+1 ≤ fs+1, and α ∈ βΛ if es+1 > fs+1. (Thus in case (BS3), α ⋓ β if and only if α
and β are comparable.)
(cid:3)
Theorem 2.11. In cases (BS1) and (BS2), the ordered group (Γ, Λ) is quasi-lattice ordered.
Proof. Let t ∈ ΛΛ−1. Let t = αβ−1 with α, β ∈ Λ. Write α = αk(i)bp and β = αℓ(j)bq in form (L). Then
t = αk(i)bp−qαℓ(j)−1. If p − q ≡ 0 (mod c), then abp−qa−1 = bn, with n ∈ Z. Then t = αk−1(i)bnαℓ−1(j)−1.
If n ≡ 0 (mod c), we may repeat this procedure. Continue this until no more such cancellation is possible.
Thus we may assume that t = αk(i)bnαℓ(j)−1 with n 6≡ 0 (mod c).
We first consider the case that n ≤ 0. Let µ ∈ tΛ ∩ Λ. There is ν ∈ Λ such that tν = µ; that
is, αk(i)bnαℓ(j)−1ν = µ. Since the reduced form of the left hand side is obtained by moving b's to the
right (see the remarks after Proposition 2.1), and since no cancellation is possible across bn, it follows that
bnαℓ(j)−1ν ∈ Λ. Then by Corollary 2.5 we have µ ∈ αk(i)Λ. Since αk(i) ∈ tΛ ∩ Λ, we have tΛ ∩ Λ = αk(i)Λ.
Finally, if n > 0 we have t−1 = αℓ(j)b−nαk(i)−1. By the previous argument we know that t−1Λ ∩ Λ =
αℓ(j)Λ. Let µ ∈ tΛ ∩ Λ, and let ν ∈ Λ with tν = µ. Then t−1µ = ν, so ν = αℓ(j)γ for some γ ∈ Λ. Then
µ = tαℓ(j)γ = αk(i)bnγ ∈ αk(i)bnΛ. Since αk(i)bn ∈ tΛ ∩ Λ, we have tΛ ∩ Λ = αk(i)bnΛ.
(cid:3)
In case (BS3), it turns out that the ordered group (Γ, Λ) is not usually finitely aligned, even though
Λ is finitely aligned (with unique minimal common extensions). However (Γ, Λ) is always locally finitely
exhaustible.
Lemma 2.12. Suppose we are in case (BS3). Then (Γ, Λ) is locally finitely exhaustible. It is finitely aligned
if and only if c = 1 (and in this case is quasi-lattice ordered).
Proof. Let t ∈ ΛΛ−1. As in the proof of Theorem 2.11, we may assume that t = αk(i)bnαℓ(j)−1, where
n 6≡ 0 (mod c). We first assume that ℓ ≥ 0, (so that θ(αℓ(j)) > 0). Then αk(i) = tαℓ(j)b−n ∈ tΛ ∩ Λ. If
µ ∈ tΛ ∩ Λ, and tν = µ with ν ∈ Λ, then we have αk(i)bnαℓ(j)−1ν = µ. By Corollary 2.5 there is p0 ∈ Z
such that αk(i)bq0 ∈ [µ]. Therefore αk(i) ⋓ µ. Now assume that ℓ = −1 (that is, that αℓ(j) ∈ B). If k ≥ 0,
then t ∈ αk(i)B ⊆ Λ. If k = −1, then t ∈ B ∪ B−1. Then tΛ ∩ Λ = tΛ, if t ∈ B, and = Λ if t ∈ B−1.
To see that (Γ, Λ) is not finitely aligned if c > 1, consider t = aba−1. Then tΛ ∩ Λ =Sn∈Z abnΛ, but for
all n ∈ Z, abn 6∈Sk>n abkΛ.
C ∗-ALGEBRAS FOR CATEGORIES OF PATHS ASSOCIATED TO THE BAUMSLAG-SOLITAR GROUPS
7
In the case that c = 1, it is easy to see that ΛΛ−1 = Λ ∪ Λ−1, so that (Γ, Λ) is totally ordered, hence
(cid:3)
quasi-lattice ordered.
We remark that the definition of C ∗(Λ), and hence the rest of the paper, requires only that Λ be a finitely
aligned category of paths (Proposition 2.10). The previous results show that the interpretation as an algebra
Morita equivalent to the crossed product of Γ acting on its directed boundary is valid.
Lemma 2.8 served the purpose of proving that (Γ, Λ) is quasi-lattice ordered in case (BS2). We next give
results analogous to Lemma 2.9 for cases (BS2) and (BS3).
Lemma 2.13. Assume that we are in case (BS2).
(1) There are maps φ : [0, d)N → [0, d)N and r : [0, d)N → {0, 1}N such that for all k ≥ 0 we have
bdαk(φ(i)) = αk(i)bcr(i)k.
(2) Let ℓ = inf{µ ≥ 1 : iµ ≥ c} (and ℓ = ∞ if iµ < c for µ ≥ 1). Then r(i)k = 1 if and only if k < ℓ.
(3) φ is a homeomorphism of [0, d)N for the product topology.
Proof. (1) and (2): We define φ(i) by
φ(i)µ =
i0,
iµ + d − c,
iℓ − c,
iµ,
if µ = 0
if 0 < µ < ℓ
if µ = ℓ
if µ > ℓ.
Then bdα0(φ(i)) = bdbi0 a = bi0abc = α0(i)bc. Inductively, for 0 < µ < ℓ, we have
bdαµ(φ(i)) = αµ−1(i)bcbd−c+iµa = αµ−1(i)biµ abc = αµ(i)bc.
bdαℓ(φ(i)) = αℓ−1(i)bcbiℓ−ca = αℓ(i),
If ℓ < ∞,
and for µ > ℓ,
(3): Let j ∈ [0, d)N. Define i ∈ [0, d)N and r ∈ NN by form (L) of bdαk(j) for all k:
bdαµ(φ(i)) = bdαℓ(φ(i))biℓ+1 a · · · biµ a = αµ(i).
bdαk(j) = αk(i)bcrk .
That iµ is well-defined independently of the choice of k ≥ µ follows from Corollary 2.5. We first show that
rk ∈ {0, 1} for all k. When k = 0 we have bdbj0 a = bj0 abc, so r0 = 1. Suppose rk−1 ∈ {0, 1}. Then
bdαk(j) = bdαk−1(j)bjk a = αk−1(i)bcrk−1+jk a.
If rk−1 = 0, then bdαk(j) = αk−1(i)bjk a, so that ik = jk and rk = 0.
If rk−1 = 1, then bdαk(j) =
αk−1(i)bc+jk a. Then, if c + jk ≥ d we have ik = jk + c − d and rk = 1, while if c + jk < d we have ik = jk + c
and rk = 0.
Notice that in the course of the last argument, we showed also that if rk−1 = 0 then rk = 0. Let rµ = 1
for µ ≤ k. Then
αk−1(i)bik abc = αk(i)bc = bdαk(j) = bdαk−1(j)bjk a = αk−1(i)bc+jk a.
Therefore bik+da = bik abc = bjk+ca, and hence ik + d = jk + c. Thus ik = c + jk − d < c. It follows that
j = φ(i), and thus φ is onto. To see that φ is one-to-one, let φ(i) = φ(i′). Then
αk(i′)bcr′
k = bdαk(φ(i′)) = bdαk(φ(i)) = αk(i)bcrk .
By the uniqueness of form (L) we have that i′ = i. From the definition of φ we see that φ(i)k is locally
constant, so that φ is continuous, and hence a homeomorphism.
(cid:3)
Lemma 2.14. Assume we are in case (BS3). There are maps φ : [0, d)N → [0, d)N and r : [0, d)N → NN
such that for all k ≥ 0 we have bdαk(φ(i)) = αk(i)b(−1)kcr(i)k . Moreover, φ is a homeomorphism of [0, d)N
for the product topology.
8
JACK SPIELBERG
Proof. For t ∈ R let f0(t) = ⌊t⌋ and f1(t) = ⌈t⌉. Let τ (k) = 1
r(i)0 = 1, φ(i)0 = i0 and
2(cid:0)1 − (−1)k(cid:1). We define r and φ recursively by
(cid:19)
r(i)k = fτ (k)(cid:18) (−1)kik + r(i)k−1c
φ(i)k = ik + (−1)k(cid:0)r(i)k−1c − r(i)kd(cid:1).
d
− ik
We first show that r(i)k ≥ 0 for all k. This is true for k = 0; suppose it is true for k − 1. If k is even,
d > −1.
We next show that φ(i)k ∈ [0, d) for all k. Again, it is true for k = 0. Suppose it is true for k−1. If k is even,
then r(i)k = (cid:4) 1
d(cid:0)ik + r(i)k−1c(cid:1)(cid:5) ≥ ⌊0⌋ = 0. If k is odd, r(i)k = (cid:6) 1
d(cid:7) ≥ 0, since
d(cid:0)−ik + r(i)k−1c(cid:1)(cid:7) ≥ (cid:6)− ik
r(i)k =(cid:4) 1
d(cid:0)ik +r(i)k−1c < r(i)k +1, and hence 0 ≤ ik +r(i)k−1c−r(i)kd < d,
d(cid:0)−ik + r(i)k−1c(cid:1)(cid:7), hence r(i)k − 1 <
If k is odd, r(i)k = (cid:6) 1
d(cid:0)−ik + r(i)k−1c(cid:1) ≤ r(i)k, and hence 0 ≤ ik −(cid:0)r(i)k−1c − r(i)kd(cid:1) < d, which is the statement that
d(cid:0)ik +r(i)k−1c(cid:1)(cid:5), hence r(ik) ≤ 1
which is the statement that φ(i)k ∈ [0, d).
1
Now we check the equation of the statement. It is true for k = 0. Suppose it is true for k − 1. Then
φ(i)k ∈ [0, d).
bdαk(φ(i)) = bdαk−1(φ(i))bφ(i)k a = αk−1(i)b(−1)k−1cr(i)k−1 bφ(i)k a.
Note that (−1)k−1cr(i)k−1 + φ(i)k = ik − (−1)kr(i)kd. Thus
bdαk(φ(i)) = αk−1(i)bik −(−1)kr(i)kda = αk−1(i)bik ab(−1)kcr(i)k = αk(i)b(−1)kcr(i)k .
From the definition of φ we see that φ(i)k is locally constant, so that φ is continuous. We show that φ
is bijective, and hence is a homeomorphism. For this, we construct the inverse of φ. We define maps
ψ : [0, d)N → [0, d)N and s : [0, d)N → {0, 1}N recursively by s(i)0 = 1, ψ(i)0 = i0 and
(cid:19)
s(i)k = fτ (k−1)(cid:18) (−1)k−1ik + s(i)k−1c
ψ(i)k = ik + (−1)k−1(cid:0)s(i)k−1c − s(i)kd(cid:1).
d
The proof that s(i)k ≥ 0 and ψ(i)k ∈ [0, d) for all k is similar to the proof of the analogous facts for r and
φ above. We show that r ◦ ψ = s and s ◦ φ = r. We have
r(ψ(i))k = fτ (k)(cid:18) (−1)kψ(i)k + r(ψ(i))k−1c
d
(cid:19)
= fτ (k) (−1)k(cid:0)ik + (−1)k−1s(i)k−1c + (−1)ks(i)kd(cid:1) + s(i)k−1c
= fτ (k)(cid:18) (−1)kik
(cid:19) + s(i)k
d
d
!
= s(i)k;
s(φ(i))k = fτ (k−1)(cid:18) (−1)k−1φ(i)k + s(φ(i))k−1c
d
(cid:19)
= fτ (k−1) (−1)k−1(cid:0)ik + (−1)kr(i)k−1c − (−1)kr(i)kd(cid:1) + ri())k−1c
= fτ (k−1)(cid:18) (−1)k−1ik
d
d
!
(cid:19) + r(i)k
since fτ (n)(cid:0)(−1)nt(cid:1) = 0 for 0 ≤ t < 1. Now we have
= r(i)k,
ψ(φ(i))k = φ(i)k + (−1)n−1(cid:0)s(φ(i))k−1c − s(φ(i))kd(cid:1) = φ(i)k + (−1)k−1(cid:0)r(i)k−1c − r(i)kd(cid:1) = ik,
and similarly, φ(ψ(i))k = ik.
(cid:3)
C ∗-ALGEBRAS FOR CATEGORIES OF PATHS ASSOCIATED TO THE BAUMSLAG-SOLITAR GROUPS
9
3. Directed hereditary subsets of Λ
We pause a moment to describe the standard picture of Λ. This is just the portion of the Cayley graph
of G given by Λ, with direction given by the generators. The set Λ/B then becomes a directed tree --
the portion of the Bass-Serre tree for G corresponding to Λ. We think of the Cayley graph as made up of
branching sheets, which we will describe precisely in a moment. The tree is then the side view, where each
sheet becomes an infinite path. (See e.g.
[5] for a sketch of the case (BS2), c = 1, d = 2.) The precise
description of these objects is found in the identification of the directed hereditary subsets of Λ, i.e. of Λ∗.
For α ∈ Λ we have the directed hereditary subset [α]. Moreover, since αΛ \ αaΛ \ αbΛ = {α}, it follows
that [α] is an open point of Λ∗. Thus Λ is a discrete subset of Λ∗ (where we identify Λ with the set of finite
directed hereditary subsets of Λ).
n=0[αbn]. It is clear that [αB] is hereditary and directed, hence is an
element of Λ∗. Letting α = bi0 a · · · bik a in form (L), it follows from Corollary 2.5 that α is uniquely determined
µ=0 αbµaΛ. Then
Next, for αB ∈ Λ/B we let [αB] =S∞
by [αB]. Thus we may identify {[αB] : α ∈ Λ} with Λ/B. For m ≥ 0 set Um = αbmΛ \Sd−1
Um = {αbj : j ≥ m} defines a neighborhood cUm of [αB] in Λ∗. Thus Λ/B is a relatively discrete subset of
Definition 3.1. An element C ∈ Λ∗ is of finite height if θ is bounded on C. An element of Λ∗ not of finite
height is said to be of infinite height.
Λ∗ \ Λ.
Lemma 3.2. The elements of Λ∗ of finite height are {[α] : α ∈ Λ} ∪ {[ξ] : ξ ∈ Λ/B}.
Proof. Let C ∈ Λ∗ have finite height. Let α ∈ C have maximal height. By the hereditary property of C
we may assume that α = bi0 a · · · bik a in form (L). By Corollary 2.5, every element of C of height k is of
the form αbp for some p. If {p : αbp ∈ C} has a maximum element q, then C = [αbq]. Otherwise, we have
C = [αB].
(cid:3)
Remark 3.3. The elements Λ/B form a directed tree when ordered by containment; this follows from
Corollary 2.5. For α = bi0 a · · · bik a, the immediate successors of [αB] in this tree are [αaB], [αbaB], . . .,
[αbd−1aB]. Thus if we orient the tree so as to be directed upward, each vertex has d upward edges (and,
apart from B, one downward edge).
We now consider the the directed hereditary subsets having infinite height.
Definition 3.4. Let i ∈ [0, d)N.
C0(i) =
C∞(i) =
∞[k=0
∞[k,p=0
[αk(i)]
[αk(i)bp].
Lemma 3.5.
(1) C0(i) and C∞(i) are directed hereditary subsets of Λ.
(2) C0(i) ⊆ C∞(i).
(3) (cid:8)C∞(i) : i ∈ [0, d)N(cid:9) = Λ∗∗.
(4) If C ∈ Λ∗ is of infinite height, there is a unique i ∈ [0, d)N such that C ⊆ C∞(i). Moreover C0(i) ⊆ C.
Proof. (1) C0(i) is directed since it is an increasing union of directed sets. C∞(i) is directed by Lemmas
2.8(4) and 2.9(4). Both are hereditary since they are unions of hereditary sets.
(2) is immediate.
(3) To see that C∞(i) is a maximal directed hereditary subset, let β ∈ Λ \ C∞(i). Write β = bj0 a · · · bjk abp
in form (L). Then there is ℓ ≤ k such that jℓ 6= iℓ. By Proposition 2.10 we have that β ⊥ αk(i), and hence
there cannot exist a directed hereditary subset containing β and C∞(i). Now we show that these are all
of the maximal elements of Λ∗∗. Let C ∈ Λ∗. Since any two elements of C have a common extension,
Proposition 2.10 implies that there is i ∈ [0, d)N such that C ⊆S∞
(4) If C ∈ Λ∗ is of infinite height, the sequence i in the proof of part (3) is uniquely determined by C.
(cid:3)
For each k there is p ≥ 0 such that αk(i)bp ∈ C. Therefore αk(i) ∈ C for all k, and hence C0(i) ⊆ C.
k,p=0[αk(i)bp] = C∞(i).
10
JACK SPIELBERG
We now describe precisely the sheets mentioned above. Namely, the sheets are the subsets C∞(i) of Λ,
for i ∈ [0, d)N. The cosets(cid:8)αk(i)B : k ∈ N(cid:9) define an infinite path in the tree Λ/B. Thus we may identify
the boundary of Λ with the boundary of the directed tree Λ/B.
Lemma 3.6. Let i ∈ [0, d)N and let C ∈ Λ∗ with C0(i) ⊆ C. Then C = C∞(i) if and only if B ⊆ C.
Proof. Of course, the hypothesis C0(i) ⊆ C is equivalent to assuming that C ⊆ C∞(i) and that C is of
infinite height. We first consider case (BS1). By Lemma 2.9(4) we have
Thus B ∨ C0(i) ⊇ C∞(i). Conversely, given h ≥ 0 let m = i0 + hd. Then
B ∨ C0(i) ⊇(cid:2)bm ∨ αk(i)(cid:3) =(cid:2)αk(i)bc Ph−1
bm ∈(cid:2)bm ∨ αk(i)(cid:3) =(cid:2)αk(i)bc Ph−1
µ=0 r(φµ(i))k(cid:3).
µ=0 r(φµ(i))k(cid:3) ⊆ C∞(i).
Hence B ⊆ C∞(i).
Now we consider case (BS2). Fix k. There are i′ ∈ [0, c)k and q ≥ 0 such that αk(i) = bqα(i′) in form
µ=0 r(ψµ(i)) (where ψ is as in Lemma 2.8). By Lemma 2.8(4) we
(R). Let h ≥ 0, and choose m = q + dPh
have
bm ∨ αk(i) = bq(cid:0)bm−q ∨ α(i′)(cid:1) = bqα(i′)b(h+1)c = αk(i)b(h+1)c.
If B ⊆ C, then since h was arbitrary, we see that C∞(i) ⊆ B ∨ C0(i) ⊆ C. Conversely, for all h we have
Thus B ⊆ C∞(i).
bhd ∈(cid:2)bd Ph−1
µ=0 r(ψµ(i))α(ψh(i))(cid:3) =(cid:2)α(i)bhc(cid:3) ⊆ C∞(i).
Finally we consider case (BS3). By Lemma 2.6, [αk+1(i)] = [αk(i)bik+1 a] ⊇ [αk(i)B]. Thus C0(i) = C∞(i),
(cid:3)
and of course, B ⊆ C∞(i).
Corollary 3.7. Λ∗∗ is a closed subset of Λ∗ (and thus ∂Λ = Λ∗∗). The map C∞(i) 7→ i of ∂Λ → [0, d)N
is a homeomorphism, equivariant for the maps bd· on ∂Λ, and φ−1 on [0, d)N (from Lemmas 2.9, 2.13, and
2.14).
Proof. It was pointed out in the remarks before Definition 3.1 that the subset of Λ∗ consisting of elements
of finite height is an open subset of Λ∗. Thus we must show that if C 6∈ Λ∗∗ is of infinite height, then C
has a neighborhood disjoint from Λ∗∗. By Lemma 3.6 we know that m = max{ℓ : bℓ ∈ C} is finite. Then
Λ∗ \ \bm+1Λ is a neighborhood of C disjoint from Λ∗∗.
It is clear that C∞(i) 7→ i is a bijection of ∂Λ with [0, d)N. Given (j0, . . . , jk) ∈ [0, d)k+1, the open set
(bj0 a · · · bjk aΛ)b corresponds to the cylinder set defined by (j0, . . . , jk), showing that the map is a homeomor-
phism. The equivariance is clear from the three lemmas mentioned.
(cid:3)
Remark 3.8. It follows from Corollary 3.7 that left-concatenation by b defines a homeomorphism of ∂Λ.
Definition 3.9. Let Σ ≡ Σ(Λ) = Λ∗ \(cid:0)Λ ∪ (Λ/B) ∪ Λ∗∗(cid:1).
Lemma 3.10. Σ is a relatively open subset of Λ∗ \ Λ.
Proof. Let C ∈ Σ. Then m = sup{j : bj ∈ C} < ∞. Then Λ∗ \ \bm+1Λ is a neighborhood of C disjoint from
(Λ/B) ∪ Λ∗∗.
(cid:3)
We now see that G0 = Λ∗ is the disjoint union of four invariant subsets: Λ∗ = Λ ⊔ (Λ/B) ⊔ ∂Λ ⊔ Σ, where
the meanings of the first two subsets were specified in the remarks before Definition 3.1 (in case (BS3), the
set Σ is empty). The subset Λ is discrete. It was shown in the remarks before Definition 3.1 that the subset
Λ/B is relatively discrete, hence open, in Λ∗ \ Λ. We now have seen that Σ is also open in Λ∗ \ Λ.
Let G = G(Λ) be the groupoid of Λ. We wish to prove that G is amenable. We note that since the maximal
degree functor θ is degenerate, the fixed-point groupoid is not AF. However, we may still use Proposition
1.3. The first step is to show that the restriction of G to ∂Λ is amenable.
Theorem 3.11. G∂Λ is amenable.
C ∗-ALGEBRAS FOR CATEGORIES OF PATHS ASSOCIATED TO THE BAUMSLAG-SOLITAR GROUPS
11
Proof. By Proposition 1.3, it suffices to show that Gθ∂Λ is amenable. Let H = Gθ∂Λ. Thus H =(cid:8)[α, β, x] :
θ(α) = θ(β), x ∈ ∂Λ(cid:9). For k ∈ N let Hk =(cid:8)[α, β, x] ∈ H : θ(α) = θ(β) = k(cid:9). Then H =S∞
show that Hk is a subgroupoid of H. It is clear that Hk is closed under inversion. Let(cid:0)[α, β, x], [γ, δ, y](cid:1) ∈
k . Since βx = γy, we have that in form (L), β = bi0a · · · βik abp and γ = bi0a · · · bik abq. Thus y = bp−qx
H (2)
(cf. Remark 3.8). Therefore, letting ε = bi0 a · · · bik a, we have
k=0 Hk. First we
[α, β, x][γ, δ, y] = [α, εbpx][εbq, δ, bp−qx] = [αbq, εbp+q, b−qx][εbp+q, δbp, b−qx] = [αbq, δbp, b−qx] ∈ Hk.
Next we show that Hk ⊆ Hk+1. Let θ(α) = θ(β) = k, and [α, β, x] ∈ Hk. Write x = C∞(i) and y = C∞(σ(i)),
where we use σ to denote the (noninvertible) left shift on [0, d)N. Then
[α, β, x] = [αbi0 a, βbi0 a, y] ∈ Hk+1.
Note also that
Hk = [i,j∈[0,d)k [p,q∈N
[bi0 a · · · bik−1 abp, bj0a · · · bjk−1 abq, ∂Λ]
is an open subgroupoid of H, hence also of Hk+1.
Now we observe that the map [bp, bq, x] 7→ (p − q, x) is an isomorphism of H0 onto the transformation
groupoid Z ⋉ ∂Λ. Thus H0 is amenable. Finally, if we write the multiplication in Hk as
[bi1 a · · · bik abp, bj1 a · · · bjk abq, x][bj1 a · · · bjk abq, bℓ1a · · · bℓk abr, x] = [bi1 a · · · bik abp, bℓ1a · · · bℓk abr, x],
we see that Hk is isomorphic to the product groupoid(cid:0)[0, d)k × [0, d)k(cid:1) × H0, which is amenable. Therefore
H is amenable (by [10], III.1).
(cid:3)
Before proving that G is amenable, we require a detailed description of the remaining elements of Λ∗
having infinite height, that is, the elements of Σ. We first consider case (BS1). Let i ∈ [0, d)N. Let n0, n1,
. . . ∈ N satisfy
(∗∗)
cnℓ−1 − iℓ
d
≤ nℓ <
c(nℓ−1 + 1) − iℓ
d
,
ℓ ≥ 1.
Note that since the outside terms of these inequalities differ by c
d ≥ 1, such sequences exist for any choice
of n0. We let n = (n0, n1, . . .), and set
Cn(i) =
[αℓ−1(i)bnℓd+iℓ].
∞[ℓ=0
Lemma 3.12. The sets in the above union increase, and hence Cn(i) is directed and hereditary.
Proof. Since cnℓ ≤ nℓ+1d + iℓ+1, we have
αℓ−1(i)bnℓd+iℓ ∈ [αℓ−1(i)bnℓd+iℓa] = [αℓ(i)bnℓc] ⊆ [αℓ(i)bnℓ+1d+iℓ+1].
(cid:3)
We remark that what we have denoted C0(i) equals the directed hereditary set defined as above for the
sequence n consisting entirely of zeros.
Lemma 3.13. αℓ−1(i)b(nℓ+1)d 6∈ Cn(i).
Proof. We will show that for m ≥ ℓ we have αℓ−1(i)b(nℓ+1)d 6∈ [αm−1(i)bnmd+im ]. Since iℓ < d, this is true
when m = ℓ. Let m ≥ ℓ, and suppose inductively that biℓa · · · bim−1 abnmd+im = bnℓd+iℓabjℓ+1 · · · abjm , where
jℓ+1, . . ., jm ∈ [0, c) (this is true vacuously when m = ℓ). Since cnm ≤ dnm+1 + im+1 < c(nm + 1), we have
that abnm+1d+im+1 = bnmdabjm+1, where jm+1 ∈ [0, c). Then
biℓa · · · bimabnm+1d+im+1 = biℓa · · · bim−1 abnmd+imabjm+1 = bnℓd+iℓabjℓ+1 · · · abjm+1.
Since this is in form (R), we see that b(nℓ+1)d 6∈ [biℓa · · · bnm+1d+im+1], since (nℓ + 1)d > nℓd + iℓ. Therefore
αℓ−1(i)b(nℓ+1)d 6∈ [αm(i)bnm+1d+im+1].
(cid:3)
Corollary 3.14. nℓ = max(cid:8)m : αℓ−1(i)bmd ∈ Cn(i)(cid:9).
Lemma 3.15. Let n and n′ both satisfy the inequalities (∗∗). Suppose that nk = n′
nℓ < n′
ℓ. Then Cn(i) ( Cn′ (i).
k for k < ℓ, and that
12
JACK SPIELBERG
Proof. By Corollary 3.14 we know that αℓ−1(i)bn′
n′
ℓ ≥ nℓ + 1, we have
ℓd ∈ Cn′ (i) \ Cn(i), and hence Cn(i) 6= Cn′ (i). Since
c(nℓ + 1) − iℓ+1
nℓ+1 <
≤
cn′
ℓ − iℓ+1
d
≤ n′
ℓ+1.
d
Inductively we find that nk < n′
k for k ≥ ℓ, and hence that Cn(i) ⊆ Cn′ (i).
(cid:3)
Lemma 3.16. Let C ∈ Λ∗ with C0(i) ⊆ C ( C∞(i). Then there exists n = (n0, n1, . . .) satisfying the
inequalities (∗∗) such that C = Cn(i).
Proof. Since C 6= C∞(i), Lemma 3.6 implies that B 6⊆ C. Thus we may define nℓ = max(cid:8)m : αℓ−1(i)bmd ∈
C(cid:9). Thus αℓ−1(i)bnℓd ∈ C and αℓ−1(i)b(nℓ+1)d 6∈ C. Let C ′ = σαℓ−1(i)C. Then bnℓd ∈ C ′ and biℓa ∈ C ′.
Therefore
bnℓd+iℓa = bnℓd ∨ biℓ a ∈ C ′.
Thus bnℓd+iℓ ∈ C ′, and hence αℓ−1(i)bnℓd+iℓ ∈ C.
We claim that the sequence n = (n0, n1, . . .) satisfies the inequalities (∗∗). For the strict inequality,
suppose that iℓ + dnℓ ≥ c(nℓ−1 + 1). Then αℓ−1(i)bc(nℓ−1+1) ∈ C, and hence
suppose otherwise; i.e.
αℓ−2(i)biℓ−1+d(nℓ−1+1) ∈ C. But this contradicts the definition of nℓ−1.
For the weak inequality, we already know that αℓ−2(i)biℓ−1+dnℓ−1 ∈ C. Thus biℓ−1+dnℓ−1, biℓ−1 a ∈
σαℓ−2(i)C. Therefore
Hence αℓ−1(i)bcnℓ−1 ∈ C.
On the other hand, we know that αℓ−1(i)biℓ+nℓd ∈ C. Moreover, by Lemma 2.9(4), we have that
biℓ−1 abcnℓ−1 = biℓ−1+dnℓ−1 ∨ biℓ−1 a ∈ σαℓ−2(i)C.
bnℓd+iℓ+1 ∨ biℓa = biℓabc(nℓ+1).
Hence if αℓ−1(i)bnℓd+iℓ+1 ∈ C, then αℓ−1(i)b(nℓ+1)d+iℓa = αℓ(i)bc(nℓ+1) ∈ C, contradicting the definition of
nℓ. Therefore nℓd + iℓ = max(cid:8)m : αℓ−1(i)bm ∈ C(cid:9). Thus we conclude that cnℓ−1 ≤ nℓd + iℓ. This finishes
the demonstration that n satisfies the inequalities (∗∗). By its definition, we have Cn(i) ⊆ C.
Finally, we claim that Cn(i) = C. For this consider a typical element β ∈ C. Then β = αℓ−1(i)bq for
(cid:3)
some ℓ and q. Then from the above argument we must have q ≤ nℓd + iℓ, and hence that β ∈ Cn(i).
Corollary 3.17. Σ = {Cn(i) : i ∈ [0, d)N, n satisfies the inequalities (∗∗) for i}.
We will now study the restriction of G to Σ.
Lemma 3.18. Recall the maps φ and r from Lemma 2.9. Let i ∈ [0, d)N.
(1) Let n satisfy the inequalities (∗∗) for φ(i). Set n′ = n + r(i). Then n′ satisfies the inequalities (∗∗)
for i. Moreover bdCn(φ(i)) = Cn′ (i).
(2) Let n′ satisfy the inequalities (∗∗) for i, and suppose that n′
0 ≥ 1. Then n′ ≥ r(i), and n = n′ − r(i)
satisfies the inequalities (∗∗) for φ(i).
Proof. (1): n satisfies the inequalities (∗∗) for φ(i) if and only if (for all ℓ,)
Using equation (∗) from Lemma 2.9, this is equivalent to
cnℓ−1 ≤ dnℓ + φ(i)ℓ ≤ cnℓ−1 + c.
cnℓ−1 ≤ dnℓ + iℓ + dr(i)ℓ − cr(i)ℓ−1 ≤ cnℓ−1 + c,
and hence to
Then we have
bdCn(φ(i)) =
=
=
cn′
ℓ−1 ≤ dn′
ℓ + iℓ ≤ cn′
ℓ−1 + c.
∞[ℓ=0(cid:2)αℓ−1(i)bnℓd+φ(i)ℓ+cr(i)ℓ−1(cid:3), by Lemma 2.9(1),
∞[ℓ=0(cid:2)bdαℓ−1(φ(i))bnℓd+φ(i)ℓ(cid:3) =
∞[ℓ=0(cid:2)αℓ−1(i)bnℓd+iℓ+dr(i)ℓ(cid:3), by equation (∗),
∞[ℓ=0(cid:2)αℓ−1(i)bn′
ℓd+iℓ(cid:3) = Cn′ (i).
C ∗-ALGEBRAS FOR CATEGORIES OF PATHS ASSOCIATED TO THE BAUMSLAG-SOLITAR GROUPS
13
(2): We show that n′
ℓ ≥ r(i)ℓ for all ℓ ∈ N. By assumption this holds for ℓ = 0. Suppose it holds for ℓ − 1.
Then using the inequalities (∗∗) for n′, and equation (∗), we have
dn′
ℓ ≥ cr(i)ℓ−1 − iℓ = dr(i)ℓ − φ(i)ℓ > dr(i)ℓ − d.
Therefore n′
satisfies the inequalities (∗∗) for φ(i).
ℓ > r(i)ℓ − 1, and hence n′
ℓ ≥ r(i)ℓ. Thus n = n′ − r(i) ≥ 0. The proof of (1) shows that n
(cid:3)
Theorem 3.19. Assume that we are in case (BS1). Then GΣ is amenable.
Proof. Since the action of b on Σ is not surjective, the argument differs in a few places from that of Theorem
3.11. By Proposition 1.3, it suffices to show that GθΣ is amenable. Let M = GθΣ = (cid:8)[β, γ, x] : x ∈
k=0 Mk, where Mk =(cid:8)[β, γ, x] ∈ M : θ(β) = θ(γ) = k(cid:9). Letting σ denote
Σ, θ(β) = θ(γ)(cid:9). Write M =S∞
the left shift on sequences (as well as the left shift on Λ), we have
σbi0 aCn(i) = σbi0 a
∞[ℓ=0(cid:2)αℓ−1(i)bnℓd+iℓ(cid:3) =
∞[ℓ=0(cid:2)bi1a · · · biℓabnℓ+1d+iℓ+1(cid:3)
∞[ℓ=0(cid:2)αℓ−1(σ(i))bσ(n)ℓd+σ(i)ℓ(cid:3) = Cσ(n)(σ(i)).
=
Thus if θ(β) = θ(γ) = k,
[β, γ, Cn(i)] =(cid:2)β, γ, bi0aCσ(n)(σ(i))(cid:3) =(cid:2)βbi0 a, γbi0a, Cσ(n)(σ(i))(cid:3) ∈ Mk+1.
It is clear that Mk is closed under inversion. If [α, β, Cm(i)] and [γ, δ, Cn(j)] are composable elements of
Mk, then βCm(i) = γCn(j) and θ(β) = θ(γ) = k. Then in form (L) we have β = bµ0 a · · · bµk−1 abp and
γ = bµ0 a · · · bµk−1 abq, with p, q ≥ 0. Without loss of generality, assume that p ≤ q. Then γ = βbq−p, and
Cm(i) = bq−pCn(j). Then
[α, β, Cm(i)] [γ, δ, Cn(j)] = [α, β, bq−pCn(j)] [βbq−p, δ, Cn(j)] = [αbq−p, δ, Cn(j)] ∈ Mk.
In the same way as in the proof of Theorem 3.11 we have that Mk is isomorphic to(cid:0)[0, d)k × [0, d)k(cid:1) × M0,
and also that Mk is an open subgroupoid of Mk+1. Thus it suffices to prove that M0 is amenable. We have
M0 =(cid:8)[bp, bq, x] : x ∈ Σ, p, q ≥ 0(cid:9).
We claim that [bp, bq, x] 7→ p − q is a continuous homomorphism from M0 to Z. Since the kernel is just the
diagonal {(x, x) : x ∈ Σ}, hence amenable, Proposition 1.3 will imply that M0 is amenable.
We have only to show that the homomorphism is well-defined. Let [bp, bq, Cm(i)] = [br, bs, Cn(j)]. Then
comparing sources gives bqCm(i) = bsCn(j). Without loss of generality we suppose that q ≤ s. Then
Cm(i) = bs−qCn(j). Comparing ranges, we then have brCn(j) = bpCm(i) = bp−q+sCn(j). Since the action
of b on Λ∗ is injective, we have that r = p − q + s, hence r − s = p − q.
(cid:3)
We now turn to case (BS2).
Lemma 3.20. Suppose that we are in case (BS2). Let i ∈ [0, d)N. For n ∈ N, let Cn(i) =S∞
Put s = lim supµ→∞ iµ.
k=0(cid:2)αk−1(i)bnc(cid:3).
(1) Cn(i) is directed and hereditary, and Cn(i) ⊆ Cn+1(i) ⊆ C∞(i) for n ∈ N.
(2) If s ≥ c then C0(i) = C∞(i).
(3) If s < c then C0(i) 6= C1(i). If C is a directed hereditary set with C0(i) ( C then C1(i) ⊆ C.
(4) Suppose s < c, and let m =(cid:6) c−s
d−c(cid:7). Then C0(i) ( C1(i) ( · · · ( Cm−1(i) ( C∞(i), and these are the
only directed hereditary sets of infinite height contained in C∞(i).
Proof. (1): We have
αk(i)bnc = αk−1(i)bik abnc = αk−1(i)bik+nda = αk−1(i)bnc · bik+n(d−c)a.
Therefore [αk−1(i)bnc] ⊆ [αk(i)bnc], and therefore the union defining Cn(i) is increasing. Hence Cn(i) is a
hereditary directed set. Since [αk(i)bnc] ⊆ [αk(i)b(n+1)c], it follows that Cn(i) ⊆ Cn+1(i). It is clear that
Cn(i) ⊆ C∞(i).
14
JACK SPIELBERG
(2): We note that for any β ∈ Λ we have bd ∈ [βabc] (e.g. from Lemma 2.13(1), since r(i)k ≤ 1 in that
lemma). Thus if 1 ≤ l1 < · · · < lm are such that iℓµ ≥ c for 1 ≤ µ ≤ m, then bmd ∈ [αℓm(i)]. If p ≥ 0 is
given, there is m such that md ≥ p. Since s ≥ c there are ℓµ as above, 1 ≤ µ ≤ m. Then
bp ∈ [bmd] ⊆ [αℓm(i)] ⊆ C0(i).
Therefore B ⊆ C0(i), and by Lemma 3.6 it follows that C0(i) = C∞(i).
(3): Since s < c there is ℓ such that iµ < c for µ ≥ ℓ. Let β = αℓ−1(i). Since biℓ a · · · bik a is in form (R)
for k ≥ ℓ, we have
On the other hand,
bd 6∈ [k≥ℓ
[biℓ a · · · bik a] = σβ [k≥ℓ
[αk(i)] = σβ(C0(i)).
βbd+iℓa = βbiℓabc = αℓ(i)bc ∈ C1(i),
so that bd ∈ σβ(C1(i)). Therefore C0(i) 6= C1(i).
Let C ) C0(i) be a directed hereditary set. Choose γ ∈ C \ C0(i). We may assume that k = θ(γ) ≥ ℓ.
Then γ = bi0 a · · · bik abq, where q > ik+1. Let i′ = σk+1(i) and C ′ = σαk(i)(C). Then C0(i′) ( C ′ and
bq ∈ C ′, where q > i′
0 = ik+1. By Lemma 2.8(3), for each µ,
Therefore C1(i′) ⊆ C ′. Hence
αµ(i′)bc ∈ [bq ∨ αµ(i′)] ⊆ C ′.
C1(i) =[µ
[αµ(i)bc] = αk(i)[µ
[αµ(i′)bc] ⊆ αk(i)C ′ = C.
(4): We have that s + (m − 1)(d − c) < c ≤ s + m(d − c). Let h be such that ik ≤ s for k ≥ h. Set
j = σh(i) ∈ [0, s]N. Then for each n, Cn(i) = αh−1(i)Cn(j). The definition of φ in the proof of Lemma
2.13 implies that φn(j)k = jk + n(d − c), k ≥ 1, for 0 ≤ n ≤ m, and that φn(j) ∈ [0, c)N for n < m, while
φm(j)k ≥ c for infinitely many k. Hence by Lemma 2.13(1), for 0 ≤ n < m we have
αk(j)bnc = bndαk(φn(j)), for k ≥ 1,
and hence by (3) that
σbnd
Cn(j) = C0(φn(j)) ( C1(φn(j)) = σbnd
Cn+1(j),
with no directed hereditary set strictly between them. Therefore Cn(j) ( Cn+1(j) with no directed hereditary
set strictly between them. Since φm(j)k ≥ c infinitely often, it follows from (2) that Cm(j) = C∞(j), finishing
the proof.
(cid:3)
Theorem 3.21. Suppose that we are in case (BS2). Then GΣ is amenable. (In fact, it is Morita equivalent
to the standard groupoid of Oc.)
Proof. We let Z = (cid:8)C0(i) : i ∈ [0, c)N(cid:9) ⊆ Σ. Note that for C ∈ Σ, C ∈ Z if and only if bd 6∈ C. Thus
Z = (dbdΛ)c is a compact-open subset of Σ. We claim that it is a transversal, in the sense of [8]. To see this,
let C ∈ Σ. Then m = sup{µ : bµ ∈ C} < ∞. Let C ′ = σbm
has source C and range in Z. It follows from [8] that GΣ is equivalent to GZ .
(C). Then bd 6∈ C ′, so C ′ ∈ Z. Then [e, bm, C ′]
To analyze GZ we first consider a pair β ∈ Λ and C ∈ Σ such that βC ∈ Z. Since bd 6∈ βC, we must
have bd 6∈ C, hence C ∈ Z. Moreover, writing β = bm0a · · · bmk abmk+1 in form (L), and C = C0(i) with
i ∈ [0, c)N, we must have m0, . . ., mk ∈ [0, c) and mk+1 ∈ [0, c − i0). Then βC = C0(f (m, i)), where
f (m, i) = (m0, . . . , mk, mk+1 + i0, i1, i2, . . .). Now, if [β, γ, C] ∈ GZ, let β and C be as above, and let
γ = bn0a · · · bnℓabnℓ+1 in form (L). Then the map [β, γ, C] 7→(cid:0)f (m, i), k − ℓ, f (n, i)(cid:1) defines an isomorphism
of GZ onto (cid:8)(x, p, y) ∈ [0, c)N × Z × [0, c)N : xµ−p = yµ for all large enough µ(cid:9), namely the standard
groupoid for the Cuntz algebra Oc. This groupoid is well-known to be amenable (e.g. [10]). Therefore GZ
is amenable.
(cid:3)
Theorem 3.22. G is amenable.
C ∗-ALGEBRAS FOR CATEGORIES OF PATHS ASSOCIATED TO THE BAUMSLAG-SOLITAR GROUPS
15
Proof. Since G∂Λ is amenable (Theorem 3.11), by [10] it suffices to prove that G(∂Λ)c is amenable. It is
easy to see that GΛ is amenable -- for any finitely aligned category of paths, GΛ is the direct sum over
v ∈ Λ0 of the elementary groupoids s−1(v) × s−1(v). Since Λ/B and Σ are disjoint relatively open subsets of
Λc, it remains to show that GΛ/B is amenable (GΣ was shown to be amenable in Theorems 3.19 and 3.21).
This follows since GΛ/B is a transitive groupoid, with isotropy isomorphic to Z (generated by b).
(cid:3)
We now apply Theorem 1.2 to give generators and relations for C ∗(Λ).
Theorem 3.23. Let Λ be the category of paths associated to the Baumslag-Solitar group G. The represen-
tations of C ∗(Λ) are in one-to-one correspondence with pairs {Sa, Sb} of Hilbert space operators satisfying
the relations
b Sa in cases (BS1) and (BS2); Sd
b SaSc
b = Sa in case (BS3).
(1) Sa and Sb are isometries.
(2) Sb is a unitary.
(3) SaSc
b = Sd
(4)
Si
bSaS∗
aS−i
b = 1.
d−1Xi=0
Moreover, in case (BS3), relation (2) is redundant.
Proof. By Theorem 1.2, we know that representations of C ∗(Λ) are in one-to-one correspondence with
families {Sα : α ∈ Λ} satisfying the relations (1) - (4) of Theorem 1.2. Given such a family, we check that
the operators Sa and Sb satisfy conditions (1) - (4). From Theorem 1.2(1) it follows that (1) holds, and
that Sb is an isometry. Then (2) follows from Theorem 1.2(4) applied to the finite exhaustive set {b}. From
Theorem 1.2(2) we conclude that (3) holds. Since bia ⊥ bja when 0 ≤ i 6= j < d, it follows from Theorem
1.2(3) that Si
b Sa have orthogonal ranges when 0 ≤ i 6= j < d. Theorem 1.2(4) applied to the finite
exhaustive set {bia : 0 ≤ i < d} verifies (4).
bSa and Sj
Conversely, suppose that Sa and Sb are given satisfying (1) - (4). We define Sα for all α ∈ Λ by setting
Se = 1, and for α = bi0a · · · bik abp in form (L), setting Sα = Si0
b . The proof (e.g. in [11]) of
Proposition 2.3 uses only the relation abc = b±da. Thus relation (3) (and relation (2) in case (BS3)) implies
that Theorem 1.2(2) holds. Relations (1) and (2) imply that Theorem 1.2(1) holds.
b Sa · · · Sik
b SaSp
Now we verify Theorem 1.2(3). Let α, β ∈ Λ.
If α ⊥ β, let them be written as in the statement of
Proposition 2.10. Then by that Proposition, there is ℓ ≤ min{s, t} such that eµ = fµ for µ < ℓ, and
eℓ 6= fℓ. Then we may compute: S∗
b Sa
have orthogonal ranges. Since α ⊥ β, this verifies Theorem 1.2(3) in this case. Suppose that α ⋓ β. If e.g.
β ∈ αΛ, then we find that SαS∗
β, and β = α ∨ β. Suppose instead that neither of α and β
extends the other. The proof of Proposition 2.10 shows that (without loss of generality) we may assume that
α = γbm and β = γabi1a · · · bik abq, and that α ∨ β = γabi1a · · · bik abp for some p. Since SbS∗
b = 1 by (2), the
final factors of Sb do not affect the computation of SS∗. Hence SαS∗
αSβ = · · · (beℓa)∗(bfℓ a) · · · = 0, since (4) implies that Seℓ
b Sa and Sfℓ
β = SβS∗
αSβS∗
β = SβS∗
β = Sα∨βS∗
αSβS∗
α∨β.
Finally we verify Theorem 1.2(4). Let F be a finite exhaustive set. If v ∈ F then Theorem 1.2(4) is
immediately satisfied. Suppose v 6∈ F . First we suppose that bm ∈ F for some m ≥ 1. Then the right-hand
side of Theorem 1.2(4) dominates SbmS∗
αb = SαS∗
α,
we may assume that elements of F have the form bi0a · · · bik a (in form (L)). Let us identify such elements
with cylinder sets in [0, d)N via the sequences (i0, . . . , ik). Moreover, because F is exhaustive we have that
bm = 1, by (2). Now suppose that F ∩ B = ∅. Since SαbS∗
these cylinder sets form a cover of [0, d)N. Thus we see thatWα∈F SαS∗
b = S∗
a(S∗
b )dSa. Now we have from (3):
it follows from (3) that Sc
α = 1.
For the final statement of the theorem, assume that we are in case (BS3). Since Sa and Sb are isometries,
and hence SaS∗
a = Sd
b SaS∗
a(S∗
b )dSaS∗
a. Thus SaS∗
b SaS∗
a(S∗
Sa = Sd
b SaSc
b SaS∗
a(S∗
b = Sd
a ≤ Sd
b )dSa,
b )d. Now using (4) gives
SbS∗
b =
dXi=1
This proves (2).
Si
bSaS∗
a(S∗
b )i =
d−1Xi=1
Si
bSaS∗
a(S∗
b )i + Sd
b SaS∗
a(S∗
b )d ≥
Si
bSaS∗
a(S∗
b )i + SaS∗
a = 1.
d−1Xi=1
(cid:3)
16
JACK SPIELBERG
Remark 3.24. The "Cuntz-Krieger" relation (Theorem 1.2(4)), defining the C ∗-algebra from the Toeplitz
C ∗-algebra of Λ, is represented by (2) and the equality (as opposed to ≤) in (4). We point out here that if
the group falls under case (BS1) or (BS2), then both of these relations are necessary. To see this, we consider
the representations of T C ∗(Λ) on ℓ2(Λ/B) and on ℓ2(Σ). The first of these satisfies (2) but not (4); the
second satisfies (4) but not (2).
Remark 3.25. The relations 1.2(1) - (4) are the same as those found by Katsura ([6], Example A.6).
Specifically, our C ∗(Λ) is isomorphic to Katsura's O(En,m) for m 6= 0, with the identifications d = n, and
c = m in case (BS1) or (BS2) when m > 0, and c = −m in case (BS3) when m < 0. We remark that in the
case where m < 0, our analysis shows that one of the relations for the C ∗-algebra turns out to be redundant.
(In the case that m = 0, the group is not one of those discussed in [3]. In fact, if c = 0, the group becomes
Z ∗ Z/dZ. If d > 1 then Λ is not a category of paths, since (3) fails. If d = 1, then G = Z, and Λ is the path
category of the directed graph having one vertex and one edge, giving the same result as [6].)
4. K-theory
Our next task is to compute the K-theory of C ∗(Λ) = C ∗(G∂Λ). We give a different calculation than
that of [6]. Thus we also compute the K-theory of the core algebra. Let A = C ∗(Λ). We let γ denote
the gauge action of T on A induced by the cocycle θ : Λ → Z. Then Takai-Takesaki duality implies that
A ⊗ K ∼= (A ×γ T) ×bγ Z.
Lemma 4.1. The fixed-point algebra Aγ is Morita equivalent to A ×γ T.
Proof. Let ζ ∈ C(T) be the function ζ(z) = z. Then the collection {ζnSαS∗
β : n ∈ Z, α, β ∈ Λ} ⊆ C(T, A) ⊆
A×γ T is a total set. A short calculation in the convolution algebra C(T, A) shows that (ζmSαS∗
ν ) =
β : θ(α) = θ(β)} is a total set in Aγ. Now let
δm,n−θ(µ)+θ(ν)ζnSαS∗
n ∈ Z and α, β ∈ Λ. Choose k ≥ 0 such that n + θ(β) + k ≥ 0, and let ν ∈ Λ with θ(ν) = n + θ(β) + k. Let
M = {bi1a · · · bik a : ij ∈ [0, d) for 1 ≤ j ≤ k}. Then SαS∗
ν . Of course, the collection {SαS∗
β)(ζnSµS∗
βSµS∗
βµ. We have that
ζnSαS∗
ζnSαµS∗
β = Xµ∈M
β =Pµ∈M SαµS∗
(SαµS∗
ν )(Sν S∗
ν )(ζnSν S∗
βµ)
βµ = Xµ∈M
is in the ideal generated by Aγ. Thus Aγ is a full hereditary subalgebra of A ×γ T.
(cid:3)
We next compute the K-theory of Aγ. We know that Aγ = C ∗(H), where H = Gγ∂Λ. Recall from the
n=0 Hn, that H0 ∼= Z ⋉ ∂Λ, and that Hn ∼= ([0, d)n × [0, d)n) × H0. So we
begin with the computation of K∗(C ∗(H0)). Since C ∗(H0) = C(∂Λ) ×b Z, and ∂Λ is totally disconnected,
we obtain from the Pimsner-Voiculescu exact sequence:
proof of Theorem 3.11 that H =S∞
0 −→ K1(C ∗(H0)) −→ K0(C(∂Λ))
id−b∗−−−−→ K0(C(∂Λ)) −→ K0(C ∗(H0)) −→ 0
Since K0(C(∂Λ)) ∼= C(∂Λ, Z), it follows that K1(C ∗(H0)) ∼= Z-span{χE : E ⊆ ∂Λ is compact-open and
b-invariant}, and K0(C ∗(H0)) ∼= C(∂Λ, Z)(cid:14)Z-span{χE − χbE : E ⊆ ∂Λ compact-open}. We recall the action
of b on ∂Λ = [0, d)N from Lemmas 2.9, 2.13 and 2.14: add 1 in the 0th coordinate; d in the jth coordinate
carries as ±c to the (j + 1)st coordinate, using +c in cases (BS1) and (BS2), and −c in case (BS3) (bd acts
as φ−1).
Lemma 4.2. Let e = (c, d) be the greatest common divisor of c and d and let c = c′e, d = d′e. For µ1,
. . ., µk ∈ [0, e), let U (µ1, . . . , µk) = {i ∈ [0, d)N : iℓ ≡ µℓ (mod e), 1 ≤ ℓ ≤ k}. Then U (µ1, . . . , µk) is
b-invariant, and every open b-invariant set is a union of such sets.
Proof. Since the complement of U (µ1, . . . , µk) is a union of such sets, it is enough to show that b·U (µ1, . . . , µk)
⊆ U (µ1, . . . , µk); moreover, since the 0th coordinate is unrestricted, it is enough to check that bd·U (µ1, . . . , µk)
⊆ U (µ1, . . . , µk). Let i ∈ U (µ1, . . . , µk), and let i′ = bd · i. Then for each ℓ there is kℓ such that i′
ℓ ≡ iℓ + kℓc
ℓ ≡ iℓ ≡ µℓ (mod e) for 1 ≤ ℓ ≤ k. Thus
(mod d). Since e divides both c and d,
i′ ∈ U (µ1, . . . , µk). Therefore U (µ1, . . . , µk) is b-invariant.
it follows that i′
Let us write Z(j0, . . . , jk) for the cylinder set {i ∈ ∂Λ : iℓ = jℓ for 0 ≤ ℓ ≤ k} in ∂Λ. We claim that for µ1,
. . ., µk ∈ [0, e), and for any j0, U (µ1, . . . , µk) =Sn∈Z bn · Z(j0, µ1, . . . , µk). The containment ⊇ follows from
the fact that U (µ1, . . . , µk) is invariant and contains Z(j0, µ1, . . . , µk). To see the containment ⊆, notice first
C ∗-ALGEBRAS FOR CATEGORIES OF PATHS ASSOCIATED TO THE BAUMSLAG-SOLITAR GROUPS
17
that we may adjust the 0th coordinate arbitrarily. Viewing the remaining coordinates as copies of Z/dZ,
note that addition by c has d′ orbits. Thus we may adjust the first coordinate to any element congruent to
µ1 modulo e. Then adding a multiple of cd′ will not further change the first coordinate. Since cd′ = c′d,
this amounts to adding cc′ in the second coordinate. Again, since (cc′, d) = e, we may adjust the second
coordinate to any element congruent to µ2 modulo e without changing the first coordinate. Repeating this
argument, we see that we may fill up U (µ1, . . . , µk) by applying b repeatedly to Z(j0, µ1, . . . , µk), proving
the claim.
Now let V be an open b-invariant set. We write V =Sp Z(p) as a union of cylinder sets, where the p are
tuples from [0, d). Then
V = [n∈Z
bn · V = [n∈Z[p
bn · Z(p) =[p
U (p′),
where if p = (j0, j1, . . . , jk) then p′ = (µ1, . . . , µk) for µℓ ∈ [0, e), µℓ ≡ jℓ (mod ℓ), 1 ≤ ℓ ≤ k.
(cid:3)
Definition 4.3. We will use the following notation. For j ∈ [0, d)k+1 let QZ(j0,...,jk) = Sbj0 a···bjk aS∗
bj0 a···bjk a
in C ∗(H0). If Z ⊆ ∂Λ is a compact-open subset, then Z is a finite disjoint union of cylinder sets. Since
Z(i0, . . . , ik) =Fd−1
j=0 Z(i0, . . . , ik, j), and QZ(i0,...,ik) =Pd−1
Z =Fj Zj for any finite disjoint collection of cylinder sets {Zj}.
j=0 QZ(i0,...,ik,j), we may define QZ =Pj QZj if
Note that if Z ⊆ ∂Λ is an invariant compact-open set, then QZ and Sb commute. In the next few items
, though it is only the coordinate-wise group structure that will be
we write Ze for (Z/eZ)Z+ ∼= [0, e)Z+
convenient (and we omit the 0th coordinate).
Corollary 4.4. K1(C ∗(H0)) ∼= C(Ze, Z).
We note that under this isomorphism, we have that [SbQU(µ1,...,µk)] corresponds to χZ(µ1+eZ,...,µk+eZ).
Lemma 4.5. K0(C ∗(H0)) ∼= C(Ze, Z[ 1
d′ ]).
Proof. We define a homomorphism C(∂Λ, Z) → C(Ze, Z[ 1
see that it is well-defined, we note that we have to check that the relation Z(j0, . . . , jk) =Sd−1
is respected. For this we compute
d′ ]) by χZ(j0,...,jk) 7→ (d′)−kχZ(j1+eZ,...,jk+eZ). To
ℓ=0 Z(j0, . . . , jk, ℓ)
(d′)−k−1χZ(j1+eZ,...,jk+eZ,ℓ+eZ)
(d′)−k−1d′χZ(j1+eZ,...,jk+eZ,ℓ′+eZ) = (d′)−kχZ(j1+eZ,...,jk+eZ).
χZ(j0,...,jk,ℓ) 7→
d−1Xℓ=0
=
d−1Xℓ=0
e−1Xℓ′=0
The map is clearly surjective, and its kernel contains χZ − χb·Z for every cylinder set Z. We claim that its
kernel is generated by the functions of the form χZ − χb·Z; this will conclude the proof. Let f be an element
of the kernel. We may choose k such that f is a linear combination of characteristic functions of cylinder sets
of length k: f =Pj njχZ(j), where j ranges over [0, d)k+1. Let [0, d)k+1 =Fp Ep be the equivalence classes
nj(cid:1)χZ(p).
defined by congruence modulo e in coordinates 1 through k. Then f maps to (d′)−kPp(cid:0)Pj∈Ep
njχZ(j)(cid:1), and it
Since f is in the kernel, we have thatPj∈Ep
nj = 0 for each p. But then f =Pp(cid:0)Pj∈Ep
is easy to see that each inner sum is in the span of the functions of the form χZ − χb·Z.
(cid:3)
The computation of the K-theory of Aγ uses the following elementary lemma.
Lemma 4.6. Let M and N be abelian groups, and η : M → M , ξ : N → N , and I : M → N homo-
that
morphisms, such that I ◦ η = ξ ◦ I. Let fM = lim
(1) ker(I) =Sn ker(ηn).
(2) N =Sn ξ−n(I(M )).
Then eI is an isomorphism.
−→
η
M , eN = lim
−→
ξ
N , and eI = lim
−→
I : fM → eN . Suppose
18
JACK SPIELBERG
with xn 7→ x. There is k such that ξk(I(xn)) = 0. Let xn+k = ηk(xn). Then xn+k ∈ ker(I), so by (1) there
is ℓ such that xn+k ∈ ker ηℓ. Then 0 = ηℓ(xn+k) 7→ x, so x = 0.
Proof. We let M (n) denote the nth copy of M in the inductive limit, etc. Let x ∈ ker(eI). Choose xn ∈ M (n)
Let y ∈ eN . Choose yn ∈ N (n) with yn 7→ y. By (2) there is k such that yn ∈ ξ−k(I(M )). Then there is
zn+k ∈ M (n+k) such that ξk(yn) = I(zn+k). Let zn+k 7→ z ∈ fM . Then yn 7→ eI(z), so that y = eI(z).
Theorem 4.7. The K-theory of Aγ is given by
(cid:3)
K0(Aγ) ∼= Z[ 1
d ]
and
K1(Aγ) ∼= Z[ 1
c ].
Moreover, the generator d−k is represented in K0 by [Sbj0 a···bjk aS∗
sented in K1 by [Sbi1 a···bik aSbS∗
Proof. Recall from the proof of Theorem 3.11 that C ∗(Hn) ∼= Mdn ⊗ C ∗(H0). Explicitly, we have that
C ∗(H0) ∼= C(∂Λ) ×b Z via
bj0 a···bjk a], and the generator c−k is repre-
bi1 a···bik a].
χ[e,e,α∂Λ] ←→ SαS∗
α
and
χ[b,e,∂Λ] ←→ Sb
(where we view C(∂Λ) ×b Z ⊆ C ∗(Λ) by means of the generators of Theorem 3.23. The inclusion C ∗(H0) ֒→
C ∗(H1) ∼= Md ⊗ C ∗(H0) is described on these generators as follows. Let α = bi0 a · · · bim a, and α′ =
bi1a · · · bima. Then
SαS∗
α ←→ χ[e,e,α∂Λ] = χ[bi0 a,bi0 a,α′∂Λ] ←→ ei0,i0 ⊗ Sα′ S∗
α′ ,
Sb ←→ χ[b,e,∂Λ] =
=
d−2Xℓ=0
χ[b,e,bℓa∂Λ] =
d−1Xℓ=0
d−1Xℓ=0
χ[bℓ+1a,bℓa,∂Λ]
χ[bℓ+1a,bℓa,∂Λ] + χ[ab±c,bd−1a,∂Λ] ←→
eℓ+1,ℓ ⊗ 1 + e0,d−1 ⊗ S±c
b
,
d−2Xℓ=0
where the exponent +c is used in cases (BS1) and (BS2), while −c is used in case (BS3). This dichomoty
In general, the inclusion Mdn ⊗ C ∗(H0) ∼= C ∗(Hn) ֒→ C ∗(Hn+1) ∼=
will continue throughout the proof.
Mdn+1 ⊗ C ∗(H0) is given by tensoring by Mdn on the left of the above inclusion.
For the computation of K0(Aγ), we consider the map on C(Ze, Z[ 1
The generator χZ(µ1+eZ,...,µk+eZ) corresponds to (d′)k[Sbj0 abµ1 a···bµk aS∗
The above inclusion sends this to
d′ ]) induced by the above inclusion.
bj0 abµ1 a···bµk a], for any choice of j0.
(d′)k[ej0,j0 ⊗ Sbµ1 a···bµk aS∗
bµ1 a···bµk a] = (d′)k[Sbµ1 a···bµk aS∗
bµ1 a···bµk a],
which corresponds to (d′)k(d′)−k+1χZ(µ2+eZ,...,µk+eZ) = d′χZ(µ2+eZ,...,µk+eZ). Thus we have
K0(Aγ) = lim
−→
η0
C(Ze, Z[ 1
d′ ]),
0 ). Let Ze have
the usual product measure, assigning measure e−k to the set Z(µ1 + eZ, . . . , µk + eZ). We claim that
K = {f ∈ C(Ze, Z[ 1
via the map η0 : χZ(µ1+eZ,...,µk+eZ) 7→ d′χZ(µ2+eZ,...,µk+eZ). We will find K = Sn ker(ηn
cylinder set of length k. Thus f =Pµ1,...,µk ∈[0,e) cµ1,...,µk χZ(µ1+eZ,...,µk+eZ). Then
d′ ]) : R f = 0}. To see this, first let R f = 0. Choose k so that f is constant on each
and hence
cµ1,...,µk ,
cµ1,...,µk = 0.
ηk
0 =Z f = e−k Xµ1,...,µk ∈[0,e)
0 (f ) = e−k(d′)k Xµ1,...,µk ∈[0,e)
f = Xµ1,...,µk∈[0,e)
Conversely, let f ∈ ker(ηn
Thus
0 ) for some n. Choose k ≥ n such that f is constant on cylinder sets of length k.
cµ1,...,µk χZ(µ1+eZ,...,µk+eZ),
C ∗-ALGEBRAS FOR CATEGORIES OF PATHS ASSOCIATED TO THE BAUMSLAG-SOLITAR GROUPS
19
and hence
0 = ηn
0 (f ) = (d′)n Xµ1,...,µk∈[0,e)
= (d′)n Xµn+1,...,µk∈[0,e)(cid:0) Xµ1,...,µn∈[0,e)
cµ1,...,µk χZ(µn+1+eZ,...,µk+eZ)
cµ1,...,µk(cid:1)χZ(µn+1+eZ,...,µk+eZ).
cµ1,...,µk(cid:1) = 0.
Thus the inner sum vanishes for each choice of µn+1, . . ., µk. Therefore
Z f = e−k Xµ1,...,µk ∈[0,e)
cµ1,...,µk = e−k Xµn+1,...,µk ∈[0,e)(cid:0) Xµ1,...,µn∈[0,e)
Thus Lemma 4.6(1) holds, where M = C(Ze, Z[ 1
ξ is surjective, Lemma 4.6(2) holds. Therefore Lemma 4.6 implies that K0(Aγ) ∼= Z[ 1
we have that [Sbj0 a···bjk aS∗
is (d′)−ke−k = d−k, thus identifying generators of K0(Aγ).
d ], I is integration, η = η0, and ξ = d·. Since
d ]. From Lemma 4.5
bj0 a···bjk a] = (d′)−kχZ(j1+eZ,...,jk+eZ) in K0(C ∗(H0)). The integral of this function
d′ ]), N = Z[ 1
For the computation of K1(Aγ) we consider the map on C(Ze, Z) induced by the inclusion of C ∗(Hn) into
C ∗(Hn+1). Recalling Definition 4.3, the inclusion gives
7→
ei0,i0 ⊗
QU(µ1,...,µk) =
d′−1Xj1,...,jk=0
d−1Xi0=0
d−1Xi0=0
d′−1Xj1,...,jk=0
where eU (µ1, . . . , µk) =Sd′−1
SbχU(µ1,...,µk) 7−→
= 1 ⊗
0
. . .
0
1
Hence
Sbi0 abµ1 +cj1 a···bµk +cjk aS∗
bi0 abµ1 +cj1 a···bµk +cjk a
d′−1Xj1,...,jk=0
Sbµ1 +cj1 a···bµk +cjk aS∗
bµ1 +cj1 a···bµk +cjk a
Sbµ1 +cj1 a···bµk +cjk aS∗
bµ1 +cj1 a···bµk +cjk a = 1 ⊗ QeU(µ1,...,µk),
S±c
b
. . .
1
0
QeU(µ1,...,µk)
. . .
. . .
QeU(µ1,...,µk)
.
j1,...,jk=0 Z(µ1 + cj1, . . . , µk + cjk). Thus the inclusion gives (in d × d matrices)
(SbQU(µ1,...,µk))d 7−→ 1 ⊗ S±c
b QeU(µ1,...,µk) = 1 ⊗ (Se
b QeU(µ1,...,µk))±c′
,
i=0 Ei = U (µ2, . . . , µk). We claim that [Se
this, we first define a *-homomorphism τ : Me(C) → Aγ by τ (eij ) = Si
b QeU(µ1,...,µk)]. Let Ei = bieU (µ1, . . . , µk) for 0 ≤ i < e. Then the Ei
b QeU(µ1,...,µk)] = [SbQU(µ2,...,µk)]. To see
bQE0S−j
i=0 ei+1,i + e0,e−1
. Let vt, 0 ≤ t ≤ 1 be a continuous path
b + QE0S−e+1
b QE0S−j
b
b
i=0 Si+1
and therefore [SbQU(µ1,...,µk)] 7−→ ±c′[Se
are pairwise disjoint, andSe−1
be the shift matrix. Then τ (v) =Pe−2
e−1Xi=0
of unitary matrices from 1 to v. We have
SbQU(µ2,...,µk) =
e−1Xi=0
SbQEi =
b QE0S−i
Si+1
b =
τ (ei+1,i) + Se
b τ (e0,e−1) = (Se
QEi)τ (v).
e−2Xi=0
Thus SbQU(µ2,...,µk)τ (vt)∗ is a continuous path from SbQU(µ2,...,µk) to Se
Now we find that
i=1 QEi, proving the claim.
χZ(µ1+eZ,...,µk+eZ) ←→ [SbQU(µ1,...,µk)]
7−→ ±c′[Se
b QeU(µ1,...,µk)] = ±c′[SbQU(µ2,...,µk)] ←→ ±c′χZ(µ2+eZ,...,µk+eZ).
Thus we have
K1(Aγ) = lim
−→
η1
C(Ze, Z),
. Let v =Pe−2
e−1Xi=1
b QE0 +
b QE0 +Pe−1
20
JACK SPIELBERG
via the map η1 : χZ(µ1+eZ,...,µk+eZ) 7→ c′χZ(µ2+eZ,...,µk+eZ). Note that
Z η1(χZ(µ1+eZ,...,µk+eZ)) =Z ±c′χZ(µ2+eZ,...,µk+eZ) = ±c′e−k+1 = ±ce−k = ±cZ χZ(µ1+eZ,...,µk+eZ).
Thus I ◦ η = ξ ◦ I, where M = C(Ze, Z), N = Z[ 1
c ], η = η1, ξ = ±c·, and I is given by integration. Essentially
the same computation as for K0 shows that Lemma 4.6(1) holds. Since Lemma 4.6(2) clearly holds, we have
that K1(Aγ) ∼= Z[ 1
c ]. Finally, we have M (n) = K1(C ∗(Hn)) ∼= C(Ze, Z), where
C(Ze, Z) ∋ 1 = [ei1,i1 ⊗ · · · ⊗ ein,in ⊗ Sb] = [Sbi1 a···bin aSbS∗
bi1 a···bin a] 7−→ ±c−n,
thus identifying the generators of K1(Aγ).
Theorem 4.8. The K-theory of C ∗(Λ) is given as follows.
(cid:3)
(1) If either c > 1 or we are in case (BS3), and if d > 1, then K0(C ∗(Λ)) ∼= Z/(d−1)Z and K1(C ∗(Λ)) ∼=
Z/(±c − 1)Z, where the minus sign is used in case (BS3).
(2) If either c > 1 or we are in case (BS3), and if d = 1, then K0(C ∗(Λ)) ∼= Z and K1(C ∗(Λ)) ∼=
Z/(±c − 1)Z ⊕ Z, where the minus sign is used in case (BS3).
(3) If c = 1 in case (BS2) (so d > 1), then K0(C ∗(Λ)) ∼= Z/(d − 1)Z ⊕ Z and K1(C ∗(Λ)) ∼= Z.
(4) If c = 1 in case (BS1) (so d = 1), then K0(C ∗(Λ)) ∼= Z ⊕ Z and K1(C ∗(Λ)) ∼= Z ⊕ Z.
In the first two cases, the class [1] of the identity is given by 1, while in the last two cases it is given by (1, 0).
Proof. Recall from Lemma 4.1 that Aγ is Morita equivalent to A ×γ T. In A ×γ T we have partial isometries
β) = ζn+1SαS∗
ζSbia, 0 ≤ i < d, with (ζSbia)∗(ζSbia) = ζ1 and (ζSbia)(ζSbia)∗ = SbiaS∗
β,
the partial isometries ζSbia, 0 ≤ i < d. We have
it follows thatbγ∗ is given by multiplication by d−1 in K0. To calculate the effect ofbγ∗ on K1, we consider
bia]; i.e. bγ∗(1) = (±c)−1. Thus bγ∗ is given on K1 by multiplication by
Thus bγ∗([Sb]) = [ζSb] = [SbiaSbS∗
(±c)−1. The Pimsner-Voiculescu exact sequence for A ∼ (A ×γ T) ×bγ Z gives
(ζSbi a)(ζSb)(ζSbia)∗ = (ζSbia)(ζSb)(S∗
bia) = SbiaSbS∗
bia.
bia. Sincebγ(ζnSαS∗
Z[ 1
c ]
↑
±c−1−−−→ Z[ 1
c ] −→ K1(A)
K0(A) ←− Z[ 1
d ]
↓
d−1←−− Z[ 1
d ]
The various cases of the theorem follow from this diagram. The identification of the class of the identity
(cid:3)
in K0 follows from the form of the generators given in Theorem 4.7.
We end by deriving the essential properties of C ∗(Λ) from properties of the groupoid G∂Λ.
Theorem 4.9.
(1) G∂Λ is minimal.
(2) G∂Λ is contractive if and only if d > 1.
(3) G∂Λ is topologically free if and only if d ∤ c.
Proof. (1): This follows from Theorem 10.14 of [13], since Λ has only one vertex.
(2): We use Theorem 10.16 of [13]. Since Λ has only one vertex, every nontrivial path is a cycle. Any
element which does not by itself form an exhaustive set will be a non-exhaustive cycle in Λ. If d > 1, then
a is such an element. Conversely, if d = 1, then the boundary of Λ reduces to a point, and then it is clear
that G∂Λ is not contractive (or even locally contractive).
(3): For the only if direction, note that if dc, then bda = abc = a(bd)(
c
d ). Hence for any γ ∈ Λ we have
c
bdγ = γbθ(γ)
d . Thus bdγ ⋓ γ for all γ, so that Λ has {bd, e}-periodicity (as in [13], Definition 10.8). By [13],
Theorem 10.10, G∂Λ is not topologically free. For the converse, suppose that d ∤ c. Let α 6= β. We must
find γ such that αγ ⊥ βγ (as in [13], Remark 10.11). If α ⊥ β, we may take γ = e. So suppose that α ⋓ β.
By Proposition 2.10 and left-cancellation we may assume that, say, θ(α) = 0. We treat three cases. (In the
following, when we write ±c we mean +c in cases (BS1) and (BS2), and −c in case (BS3).) First, suppose
that α = e and θ(β) > 0. Then β = biaβ′, where i ∈ [0, d). Let j ∈ [0, d) with j 6= i (since d 6= 1). Then
αbja = bja ⊥ biaβ′bja = βbja. Second, suppose that α = e and θ(β) = 0. Then β = bq with q > 0. Since
C ∗-ALGEBRAS FOR CATEGORIES OF PATHS ASSOCIATED TO THE BAUMSLAG-SOLITAR GROUPS
21
= j + rd, where 0 < j < d. We
have
d ∤ c, there is a least positive integer k such that q(cid:0) c
d(cid:1)k
d (cid:1)k−1
bqak = ak−1bq(cid:0) ±c
6∈ Z. Then q(cid:0) ±c
d (cid:1)k−1
a = ak−1bjabrc ⊥ ak.
Finally we suppose that α = bp and θ(β) > 0. Then β = aβ′ (otherwise we could cancel some element of
B on the left). Let p = i + rd with i ∈ [0, d). If i 6= 0, then αa = biab±rc ⊥ aβ′a = βa. If i = 0, then
αba = bab±rc ⊥ aβ′ba = βba.
(cid:3)
Corollary 4.10. C ∗(Λ) is a Kirchberg algebra if and only if d ∤ c.
References
[1] H. Al-Sulami, C ∗-algebras for boundary actions of solvable Baumslag-Solitar groups, Int. J. Math. Anal. (Ruse) 1 (2007),
no. 21-24, 10671080.
[2] C. Anantharaman-Delaroche, Purely infinite C ∗-algebras arising from dynamical systems, Bull. Soc. Math. France 125
(1997), no. 2, 199 -- 225.
[3] G. Baumslag and D. Solitar, Some two-generator one-relator non-Hopfian groups, Bull. Amer. Math. Soc. 68 1962 199 -- 201.
[4] J. Cuntz, Simple C ∗-algebras generated by isometries, Comm. Math. Phys. 57 (1977), no. 2, 173 -- 185.
[5] B. Farb, L. Mosher, A rigidity theorem for the solvable Baumslag-Solitar groups (with an appendix by Daryl Cooper),
Invent. Math. 131 (1998), no. 2, 419451.
[6] T. Katsura, A class of C ∗-algebras generalizing both graph algebras and homeomorphism C ∗-algebras IV, Pure infiniteness.
J. Funct. Anal. 254 (2008), no. 5, 1161 -- 1187.
[7] M. Laca and I. Raeburn, Semigroup crossed products and the Toeplitz algebras of nonabelian groups, J. Funct. Anal. 139
(1996), no. 2, 415 -- 440.
[8] P. Muhly, J. Renault and D. Williams, Equivalence and isomorphism for groupoid C ∗-algebras, J. Operator Theory 17
(1987), no. 1, 3 -- 22.
[9] A. Nica, C ∗-algebras generated by isometries and Wiener-Hopf operators, J. Operator Theory 27 (1992), no. 1, 17 -- 52.
[10] J. Renault, A groupoid approach to C ∗-algebras, Lecture Notes in Mathematics, 793, Springer, Berlin, 1980.
[11] P. Scott and T. Wall, Topological methods in group theory, Homological group theory (Proc. Sympos., Durham, 1977), pp.
137 -- 203, London Math. Soc. Lecture Note Ser., 36, Cambridge Univ. Press, Cambridge-New York, 1979.
[12] J-P. Serre, Trees, Springer-Verlag, Berlin Heidelberg, 1980.
[13] J. Spielberg, Groupoids and C ∗-algebras for categories of paths, preprint 2011, arXiv:1111.6924v2.
School of Mathematical and Statistical Sciences, Arizona State University, P.O. Box 871804, Tempe, AZ 85287-
1804
E-mail address: [email protected]
|
1001.3182 | 2 | 1001 | 2011-07-12T05:05:34 | The structure of an isometric tuple | [
"math.OA",
"math.FA"
] | An $n$-tuple of operators $(V_1,...,V_n)$ acting on a Hilbert space $H$ is said to be isometric if the operator $[V_1\...\ V_n]:H^n\to H$ is an isometry. We prove a decomposition for an isometric tuple of operators that generalizes the classical Lebesgue-von Neumann-Wold decomposition of an isometry into the direct sum of a unilateral shift, an absolutely continuous unitary and a singular unitary. We show that, as in the classical case, this decomposition determines the weakly closed algebra and the von Neumann algebra generated by the tuple. | math.OA | math |
THE STRUCTURE OF AN ISOMETRIC TUPLE
MATTHEW KENNEDY
Abstract. An n-tuple of operators (V1, . . . , Vn) acting on a Hilbert
space H is said to be isometric if the operator [V1 · · · Vn] : H n → H
is an isometry. We prove a decomposition for an isometric tuple of
operators that generalizes the classical Lebesgue-von Neumann-Wold
decomposition of an isometry into the direct sum of a unilateral shift,
an absolutely continuous unitary and a singular unitary. We show that,
as in the classical case, this decomposition determines the weakly closed
algebra and the von Neumann algebra generated by the tuple.
1. Introduction
This paper concerns the structure of an isometric tuple of operators, an
object that appears frequently in mathematics and mathematical physics.
From the perspective of an operator theorist, the notion of an isometric tuple
is a natural higher-dimensional generalization of the notion of an isometry.
An n-tuple of operators (V1, . . . , Vn) acting on a Hilbert space H is said to
be isometric if the row operator [V1 ··· Vn] : H n → H is an isometry. This
is equivalent to requiring that the operators V1, . . . , Vn satisfy the algebraic
relations
These relations are often referred to as the Cuntz relations.
The main result in this paper is a decomposition of an isometric tuple that
generalizes the classical Lebesgue-von Neumann-Wold decomposition of an
isometry into the direct sum of a unilateral shift, an absolutely continuous
unitary and a singular unitary. We show that, as in the classical case, this
decomposition determines the structure of the weakly closed algebra and the
von Neumann algebra generated by the tuple.
The existence of a higher-dimensional Lebesgue-von Neumann-Wold de-
composition was conjectured by Davidson, Li and Pitts in [DLP05]. They
2000 Mathematics Subject Classification. Primary 47A13; Secondary 47L55, 46L10.
Research partially supported by an NSERC Canada Graduate Scholarship.
1
V ∗
I
0
i Vj =
if i = j,
if i 6= j.
THE STRUCTURE OF AN ISOMETRIC TUPLE
2
observed that the measure-theoretic definition of an absolutely continuous
operator was equivalent to an operator-theoretic property of the functional
calculus for that operator. Since this property naturally extends to the
higher-dimensional setting, this allowed them to define the notion an abso-
lutely continuous isometric tuple.
The key technical result in this paper is a more effective operator-algebraic
characterization of an absolutely continuous isometric tuple. The lack of
such a characterization had been identified as the biggest obstruction to
establishing the conjecture in [DLP05] (see also [DY08]). As we will see, the
difficulty here can be attributed to the lack of a higher-dimensional analogue
of the spectral theorem.
In this paper, we overcome this difficulty by extending ideas from the
commutative theory of dual algebras to the noncommutative setting. A
similar approach was used in [Ken11] to prove that certain isometric tuples
are hyperreflexive. In the present paper, the assumptions on the isometric
tuples we consider are much weaker, and the problem is substantially more
difficult. The idea to use this approach was inspired by results of Bercovici
in [Ber98].
In Section 2, we review the Lebesgue-von Neumann-Wold decomposition
In Section 3,
of a single isometry that is the motivation for our results.
we provide a brief review of the requisite background material on higher-
dimensional operator theory, and we introduce the notions of absolute conti-
nuity and singularity. In Section 4, we prove an operator-algebraic character-
ization of an absolutely continuous isometric tuple. In Section 5, we prove an
operator-algebraic characterization of a singular isometric tuple. In Section
6, we prove the Lebesgue-von Neumann-Wold decomposition of an isometric
tuple, and we obtain some consequences of this result.
The present exposition was inspired by the perspective of Muhly and Solel
in [MS10], which appeared shortly after the first version of this paper. They
consider the notion of absolute continuity in a more general setting.
2. Motivation
The structure of a single isometry V is well understood. By the Wold
decomposition of an isometry, V can be decomposed as
V = Vu ⊕ U,
THE STRUCTURE OF AN ISOMETRIC TUPLE
3
where Vu is a unilateral shift of some multiplicity, and U is a unitary. By
the Lebesgue decomposition of a measure applied to the spectral measure of
U , we can decompose U as
U = Va ⊕ Vs,
where Va is an absolutely continuous unitary and Vs is a singular unitary, in
the sense that their spectral measures are absolutely continuous and singular
respectively with respect to Lebesgue measure. This allows us to further
decompose V as
V = Vu ⊕ Va ⊕ Vs.
We will refer to this as the Lebesgue-von Neumann-Wold decomposi-
tion of an isometry.
It will be convenient to consider the above notions of absolute continuity
and singularity from a different perspective. Let A(D) denote the classical
disk algebra of analytic functions on the complex unit disk D with continuous
extension to the boundary. An isometry V induces a contractive representa-
tion of A(D), namely the A(D) functional calculus for V , given by
f → f (V ),
f ∈ A(D).
Recall that the algebra A(D) is a weak-* dense subalgebra of the algebra H ∞
of bounded analytic functions on the complex unit disk. In certain cases, the
representation of A(D) induced by V is actually the restriction to A(D) of a
weak-* continuous representation of H ∞, namely the H ∞ functional calculus
for V , given by
f → f (V ),
f ∈ H ∞.
It follows from Theorem III.2.1 and Theorem III.2.3 of [SF70] that this occurs
if and only if Vs = 0 in the Lebesgue-von Neumann-Wold decomposition of
V . This motivates the following definitions.
Definition 2.1. Let V be an isometry. We will say that V is absolutely
continuous if the representation of A(D) induced by V extends to a weak-*
continuous representation of H ∞. If V has no absolutely continuous restric-
tion to an invariant subspace, then we will say that V is singular.
The importance of the Lebesgue-von Neumann-Wold decomposition of an
isometry V is that it determines the structure of the weakly closed algebra
W(V ) and the von Neumann algebra W∗(V ) generated by V . Recall that
THE STRUCTURE OF AN ISOMETRIC TUPLE
4
W(V ) is the weak closure of the polynomials in V , and W∗(V ) is the weak
closure of the polynomials in V and V ∗.
Let α denote the multiplicity of Vu as a unilateral shift, and let µa and
µs be scalar measures equivalent to the spectral measures of Va and Vs re-
spectively. Since a unilateral shift of multiplicity one is irreducible, W∗(V )
is given by
W∗(V ) ≃ B(ℓ2)α ⊕ L∞(Va) ⊕ L∞(µs)(Vs).
It was established by Wermer in [Wer52] that W(V ) can be self-adjoint,
depending on α and µa.
If α 6= 0 or if Lebesgue measure is absolutely
continuous with respect to µa, then W(V ) is given by
W(V ) ≃ H ∞(Vu ⊕ Va) ⊕ L∞(µs)(Vs).
Otherwise, if neither of these conditions holds, then W(V ) = W∗(V ).
The following example shows that it is possible for the weakly closed alge-
bra generated by an absolutely continuous isometry to be self-adjoint. We
will see later that there is no higher-dimensional analogue of this phenome-
non.
Example 2.2. Let U denote the operator of multiplication by the coordinate
function on L2(T, m), where m denotes Lebesgue measure. Let m1 and m2
denote Lebesgue measure on the upper and lower half of the unit circle
respectively, and let U1and U2 denote the operator of multiplication by the
coordinate function on L2(T, m1) and L2(T, m2) respectively.
Since the spectral measure of U ≃ U1 ⊕ U2 is equivalent to Lebesgue
measure, U is absolutely continuous. Thus U1 and U2 are also absolutely
continuous. From above,
W∗(U ) ≃ L∞(U ),
W(U ) ≃ H ∞(U ).
However, since Lebesgue measure is not absolutely continuous with respect
to m1 or m2,
W(Ui) = W∗(Ui) = L∞(Ui),
i = 1, 2.
In particular, the weakly closed algebras W(U1) and W(U2) generated by U1
and U2 respectively are self-adjoint.
THE STRUCTURE OF AN ISOMETRIC TUPLE
5
3. Background and preliminaries
3.1. The noncommutative function algebras. The noncommutative Hardy
space F 2
n is defined to be the full Fock-Hilbert space over Cn, i.e.
n = ⊕∞
F 2
k=0(Cn)⊗k,
where we will write ξ∅ to denote the vacuum vector, so that (Cn)⊗0 = Cξ∅.
Let ξ1, . . . , ξn be an orthonormal basis of Cn and let F∗
n denote the unital free
semigroup on n generators {1, . . . , n} with unit ∅. For a word w = w1 ··· wk
in F∗
n, it will be convenient to write ξw = ξw1 ⊗ ··· ⊗ ξwk . We can identify
F 2
n with the set of power series in n noncommuting variables ξ1, . . . , ξn with
square-summable coefficients, i.e.
F 2
n =
Xw∈F∗
n
awξw : Xw∈F∗
n
.
aw2 < ∞
In particular, we can identify the noncommutative Hardy space F 2
1 with the
classical Hardy space H 2 of analytic functions having power series expansions
with square-summable coefficients.
The left multiplication operators L1, . . . , Ln are defined on F 2
n by
Liξw = ξi ⊗ ξw = ξiw, w ∈ F∗
n.
It is clear that the n-tuple L = (L1, . . . , Ln) is isometric. We will call it the
unilateral n-shift since, for n = 1, L1 can be identified with the unilateral
shift on H 2. For a word w = w1 ··· wk in F∗
n, it will be convenient to write
Lw = Lw1 ··· Lwk .
The noncommutative disk algebra An is the norm closed unital alge-
bra generated by L1, . . . , Ln and the noncommutative analytic Toeplitz
algebra Ln is the weakly closed unital algebra generated by L1, . . . , Ln.
These algebras were introduced by Popescu in [Pop96], and have subse-
quently been studied by a number of authors (see for example [DP98] and
[DP99]).
The noncommutative disk algebra An and the noncommutative analytic
Toeplitz algebra Ln are higher-dimensional analogues of the classical disk
algebra A(D) and the classical algebra H ∞ of bounded analytic functions.
In particular, the algebra An is a proper weak-* dense subalgebra of the
algebra Ln. If we agree to identify functions in H ∞ with the corresponding
multiplication operators on H 2, then we can identify A(D) with A1 and H ∞
with L1.
THE STRUCTURE OF AN ISOMETRIC TUPLE
6
As in the classical case, an element A in Ln is uniquely determined by its
Fourier series
awLw,
A ∼ Xw∈F∗
n
where aw = (Aξ∅, ξw) for w in F∗
n. The Cesaro sums of this series converge
strongly to A, and it is often useful heuristically to work directly with this
representation.
We will also need to work with the right multiplication operators R1, . . . , Rn
defined on F 2
n by
Riξw = ξw ⊗ ξi = ξwi, w ∈ F∗
n.
The n-tuple R = (R1, . . . , Rn) is unitarily equivalent to L = (L1, . . . , Ln).
The unitary equivalence is implemented by the "unitary flip" on F 2
n that, for
a word w1 ··· wk in F∗
n, takes ξw1···wk to ξwk···w1. We will let Rn denote the
weakly closed algebra generated by R1, . . . , Rn.
3.2. Free semigroup algebras. Let V = (V1, . . . , Vn) be an isometric n-
tuple. The weakly closed unital algebra W(V ) generated by V1, . . . , Vn is
called the free semigroup algebra generated by V . As in Section 3.1, for
a word w = w1 ··· wk in the unital free semigroup F∗
n, it will be convenient
to write Vw = Vw1 ··· Vwk .
Example 3.1. The noncommutative analytic Toeplitz algebra Ln intro-
duced in Section 3.1 is a fundamental example of a free semigroup algebra.
We will see that it plays an important role in the general theory of free
semigroup algebras.
The study of free semigroup algebras was initiated by Davidson and Pitts
in [DP99]. They observed that information about the unitary invariants
of an isometric tuple can be detected in the algebraic structure of the free
semigroup algebra it generates, and used this fact to classify a large fam-
ily of representations of the Cuntz algebra. Free semigroup algebras have
subsequently received a great deal of interest (see for example [Dav01]).
It was shown in [DP98] that Ln has a great deal of structure that is
analogous to the analytic structure of H ∞. This motivates the following
definition.
Definition 3.2. An isometric n-tuple V = (V1, . . . , Vn) is said to be an-
alytic if the free semigroup algebra generated by V is isomorphic to the
noncommutative analytic Toeplitz algebra Ln.
THE STRUCTURE OF AN ISOMETRIC TUPLE
7
The notion of analyticity is of central importance in the theory of free
semigroup algebras. This is apparent from the work of Davidson, Katsoulis
and Pitts in [DKP01]. They proved the following general structure theorem.
Theorem 3.3 (Structure theorem for free semigroup algebras). Let V =
W(V ) be a free semigroup algebra. Then there is a projection P in V with
range invariant under V such that
(1) if P 6= 0, then the restriction of V to the range of P is an analytic
free semigroup algebra,
(2) the compression of V to the range of P ⊥ is a von Neumann algebra,
(3) V = PVP + (W∗(V ))P ⊥.
The analytic structure of a free semigroup algebra reveals itself in the form
of wandering vectors. Let V = (V1, . . . , Vn) be an isometric n-tuple acting
on a Hilbert space H. A vector x in H is said to be wandering for V if the
set of vectors {Vwx : w ∈ F∗
n} is orthonormal. In this case we will also say
that x is wandering for the free semigroup algebra generated by V .
The existence of wandering vectors for an analytic free semigroup algebra
was established in [Ken11], settling a conjecture first made in [DKP01] (see
also [DLP05] and [DY08]). Examples show that the structure of an analytic
free semigroup algebra can be quite complicated, making this result far from
obvious.
3.3. Dilation theory. Recall that an operator T is said to be contractive
if kTk ≤ 1. An n-tuple of operators T = (T1, . . . , Tn) acting on a Hilbert
space H is said to be contractive if the row operator [T1 ··· Tn] : H n → H
is contractive.
Sz.-Nagy showed that every contractive operator T acting on a Hilbert
space H has a unique minimal dilation to an isometry V , acting on a bigger
Hilbert space K (see for example [SF70]). This means that H ⊆ K, H is
cyclic for V and
T k = PHV k H ,
k ≥ 1.
Sz.-Nagy's dilation theorem was generalized in the work of Bunce, Frazho
and Popescu in [Bun84], [Fra82] and [Pop89a] respectively. They showed
that every contractive n-tuple of operators T = (T1, . . . , Tn) acting on a
Hilbert space H has a unique minimal dilation to an isometric n-tuple V =
(V1, . . . , Vn), acting on a bigger Hilbert space K. This means that H ⊆ K,
THE STRUCTURE OF AN ISOMETRIC TUPLE
8
H is cyclic for V1, . . . , Vn and
PH Vi1 ··· Vik H = Ti1 ··· Tik ,
i1, . . . , ik ∈ {1, . . . , n} and k ≥ 1.
3.4. The Wold decomposition. The classical Wold decomposition decom-
poses a single isometry into the direct sum of a unilateral shift of some
multiplicity and a unitary. In order to state the Wold decomposition of an
isometric tuple, we need to generalize these notions.
In Section 3.1, we introduced the unilateral n-shift L = (L1, . . . , Ln),
and we saw that it is the natural higher-dimensional generalization of the
classical unilateral shift. An isometric n-tuple is said to be a unilateral
shift of multiplicity α if it is unitarily equivalent to the ampliation L(α) =
(L(α)
1 , . . . , L(α)
The higher-dimensional generalization of a unitary is based on the fact
that a unitary is the same thing as a surjective isometry. An n-tuple of
operators U = (U1, . . . , Un) is said to be unitary if the operator [U1 ··· Un] :
H n → H is a surjective isometry. This is equivalent to requiring that the
operators U1, . . . , Un satisfy
n ), for some positive integer α.
n
UiU ∗
i = I.
Xi=1
Note that a unilateral shift is not unitary. This is because the "vacuum"
vector ξ∅ in F 2
n is not contained in the range of the unilateral n-shift L =
(L1, . . . , Ln).
In [DP99], Davidson and Pitts studied a family of "atomic" isometric tuples
that arise from certain infinite directed trees. As the following example
shows, this family contains a large number of unitary tuples.
Example 3.4. Fix an infinite directed n-ary tree B with vertex set V such
that every vertex has a parent. For a vertex v in V , let ci(v) denote the i-th
child of v. Let H = ℓ2(V ), so that the set {ev : v ∈ V } is an orthonormal
basis for H. Define operators S1, . . . , Sn on H by
Siev = eci(v),
1 ≤ i ≤ n.
It's clear that S1, . . . , Sn are isometries, and the fact that B is an infinite
directed n-ary tree implies that the range of Si and the range of Sj are
orthogonal for i 6= j. Thus S = (S1, . . . , Sn) is an isometric n-tuple. The
fact that every vertex has a parent implies that every basis vector is in the
range of some Si. Thus S is a unitary n-tuple.
THE STRUCTURE OF AN ISOMETRIC TUPLE
9
Let V = (V1, . . . , Vn) be an arbitrary isometric n-tuple. If V is unitary,
then the C∗-algebra C∗(V1, . . . , Vn) generated by V is isomorphic to the
Cuntz algebra On. Otherwise, it is isomorphic to the extended Cuntz al-
gebra En, the extension of the compacts by On. Since the only irreducible
*-representation of the compacts is the identity representation, and since On
is simple, a *-representation of En can be decomposed into a multiple of the
identity representation and a representation of On. The Wold decomposition
of an isometric n-tuple, which was proved by Popescu in [Pop89a], can be
obtained as a consequence of these C∗-algebraic facts, based on the observa-
tion that the C∗-algebra generated by a unilateral n-shift is isomorphic to
En.
Proposition 3.5 (The Wold decomposition). Let V = (V1, . . . , Vn) be an
isometric n-tuple. Then we can decompose V as
V = Vu ⊕ U,
where Vu is a unilateral n-shift and U is a unitary n-tuple.
3.5. Absolutely continuous and singular isometric tuples. As in the
classical case, an isometric n-tuple V = (V1, . . . , Vn) induces a contractive
representation of the noncommutative disk algebra An, called the An func-
tional calculus for V , determined by
Li1 ··· Lik → Vi1 ··· Vik ,
i1, . . . , ik ∈ {1, . . . , n} and k ≥ 1.
This is a consequence of Popescu's generalization of von Neumann's inequal-
ity in [Pop91].
Recall from Section 3.1 that An is a proper weak-* dense subalgebra of
the noncommutative analytic Toeplitz algebra Ln. The following definition
is the natural generalization of Definition 2.1.
Definition 3.6. Let V = (V1, . . . , Vn) be an isometric n-tuple. We will say
that V is absolutely continuous if the representation of An induced by V
is the restriction to An of a weak-* continuous representation of Ln. We will
say that V is singular if V has no absolutely continuous restriction to an
invariant subspace.
It is clear from Definition 3.2 and Definition 3.6 that an analytic iso-
metric tuple is absolutely continuous. In order to obtain the Lebesgue-von
Neumann-Wold decomposition of an isometric tuple, we will prove the con-
verse result that an absolutely continuous isometric tuple is analytic.
THE STRUCTURE OF AN ISOMETRIC TUPLE
10
4. Absolutely continuous isometric tuples
The main result in this section is an operator-algebraic characterization
of an absolutely continuous isometric tuple. Specifically, we will show that
for n ≥ 2, every absolutely continuous isometric n-tuple is analytic.
For n ≥ 2, fix an absolutely continuous isometric n-tuple S = (S1, . . . , Sn)
acting on a Hilbert space H. Let Φ denote the corresponding representation
of the noncommutative disk algebra An, given by
Φ(Lw) = Sw, w ∈ F∗
n.
Since S is absolutely continuous, Φ extends to a representation of Ln that is
weak-* continuous.
It was shown in Corollary 1.2 of [DY08] that Φ is actually a completely
isometric isomorphism and a weak-* homeomorphism from Ln to the weak-*
closed algebra generated by S1, . . . , Sn. This is equivalent to the fact that
an infinite ampliation of S is an analytic isometric tuple. Evidently, it is
much more difficult to show that S is analytic. As an explanation, we offer
the aphorism that things are generally much nicer in the presence of infinite
multiplicity.
Showing that S is analytic amounts to showing that the free semigroup
algebra (i.e.
the weakly closed algebra) W(S) generated by S1, . . . , Sn is
isomorphic to the noncommutative analytic Toeplitz algebra Ln. Since we
know from above that the weak-* closed algebra generated by S1, . . . , Sn is
isomorphic to Ln, our strategy will be to show that this algebra is actually
equal to W(S).
4.1. The noncommutative Toeplitz operators. Let S denote the weak-
* closed algebra generated by S1, . . . , Sn. The map Φ introduced at the
beginning of this section is a completely isometric isomorphism and a weak-*
homeomorphism from Ln to S. It will be useful for what follows to extend
Φ even further. Let Mn denote the weak-* closure of the operator system
Ln + L∗
n. We will call the elements of Mn the noncommutative Toeplitz
operators, because they are a natural higher-dimensional generalization of
the classical Toeplitz operators.
The noncommutative Toeplitz operators were introduced by Popescu in
[Pop89b]. It was shown in Corollary 1.3 of [Pop09] that A belongs to Mn if
THE STRUCTURE OF AN ISOMETRIC TUPLE
11
and only if
A if i = j,
0
otherwise,
R∗
i ARj =
where R1, . . . , Rn are the right multiplication operators introduced in Section
3.1. A short proof of this fact was also given in Lemma 3.2 of [Ken11]. It
follows from this characterization that Mn is weakly closed.
Let T denote the weak-* closure of the operator system S +S ∗. The proof
of the following proposition is nearly identical to the proof of Theorem 3.6
of [Ken11].
Proposition 4.1. Let S = (S1, . . . , Sn) be an absolutely continuous isomet-
ric n-tuple. The representation Φ of Ln induced by S extends to a completely
isometric and weak-* homeomorphic *-map from Mn to T .
We will need to exploit the fact that Mn and T are dual spaces. Let T∗
denote the predual of T , i.e. the set of weak-* continuous linear functionals
on T . Similarly, let Mn∗ denote the predual of Mn. Basic functional analysis
implies that the inverse map Φ−1 is the dual of an isometric isomorphism φ
from Mn∗ to T∗. Moreover, since Φ−1 is isometric, so is φ.
We can identify the predual of B(F 2
n ), with the set of trace class operators C 1(F 2
n ), i.e. the set of weak-* continuous
n ) on
linear functionals on B(F 2
F 2
n , where K in C 1(F 2
n ) corresponds to the linear functional
(T, K) = tr(T K), T ∈ B(F 2
n ).
If we let (Mn)⊥ denote the preannihilator of Mn, i.e.
(Mn)⊥ = {K ∈ C 1(F 2
n ) : tr(AK) = 0,
∀A ∈ Mn},
then we can identify the predual (Mn)∗ with the quotient space C 1(F 2
Similarly, we can identify the predual T∗ with the quotient space C 1(H)/T⊥.
n , it will be convenient to let [ξ ⊗ η]Mn denote the weak-*
For ξ and η in F 2
n )/(Mn)⊥.
continuous linear functional on Mn given by
(A, [ξ ⊗ η]Mn) = (Aξ, η), A ∈ Mn.
In other words, [ξ⊗η]Mn denotes the equivalence class of the rank one tensor
x ⊗ y in (Mn)∗. Similarly, for x and y in H, let [x ⊗ y]T denote the weak-*
continuous linear functional on T given by
(T, [x ⊗ y]T ) = (T x, y),
T ∈ T .
THE STRUCTURE OF AN ISOMETRIC TUPLE
12
4.2. Intertwining operators. An operator X : F 2
n → H is said to in-
tertwine the isometric n-tuple S = (S1, . . . , Sn) and the unilateral n-shift
L = (L1, . . . , Ln) if it satisfies
XLi = SiX,
1 ≤ i ≤ n.
Observe that if X intertwines S and L, then the operator JX∗XJ is a
noncommutative Toeplitz operator, where J is the unitary flip introduced
in Section 3.1. Indeed, using the fact that JRi = LiJ for 1 ≤ i ≤ n, we
compute
R∗
i JX∗XJRj = JL∗
i X∗XLjJ
i SjXJ
= JX∗S∗
JX∗XJ if i = j,
0
otherwise.
=
Since S is absolutely continuous, it follows from Theorem 2.7 of [DLP05]
that every vector x in H is in the range of an operator that intertwines S
and L.
4.3. Dual algebra theory. Recall that to prove the isometric n-tuple S =
(S1, . . . , Sn) is analytic, our strategy is to show that the weak-* closed al-
gebra S = W∗(S1, . . . , Sn) is actually equal to the weakly closed algebra
W(S1, . . . , Sn). This amounts to showing that S is already weakly closed.
However, instead of working directly with S, it will be necessary to work
with the operator system T . In fact, we will need to consider the general
structure of the predual of T .
In Section 4.1, we saw that an element in the predual T∗ of the operator
system T can be identified with an equivalence class of trace class operators.
We will show that T satisfies a very powerful predual "factorization" property,
in the sense that the equivalence class of an element in the predual T∗ always
contains "nice" representatives. We will see that S inherits this property from
T , and that this will imply the desired result.
The idea of studying factorization in the predual of an operator algebra
is the central idea in dual algebra theory, which has been applied with great
success to a number of problems in the commutative setting (see for ex-
ample [BFP85]). As we will see, many of the factorization properties that
were introduced in the commutative setting make sense even in the present
noncommutative setting.
THE STRUCTURE OF AN ISOMETRIC TUPLE
13
Definition 4.2. A weak-* closed subspace A of operators acting on a Hilbert
space H is said to have property A1(1) if, given a weak-* continuous linear
functional τ on A with kτk ≤ 1 and ǫ > 0, there are vectors x and y in H
such that kxk ≤ (1 + ǫ)1/2, kyk ≤ (1 + ǫ)1/2 and τ = [x ⊗ y]A.
If a weak-* closed subspace of B(H) has property A1(1), then the equiva-
lence class of any weak-* continuous linear functional on the subspace con-
tains an operator of rank one. Note that in this case, every weak-* continuous
linear functional on the subspace is actually weakly continuous. It was shown
in [DP99] that Ln has property A1(1), and the same proof also shows that
Mn has property A1(1).
Of course, the main difficulty with a predual factorization property like
property A1(1) is that it is often extremely difficult to show that it holds.
The next factorization property turns out to be much stronger than property
A1(1), but it is sometimes easier to show that it holds due to its approximate
nature.
Definition 4.3. A weak-* closed subspace A of operators acting on a Hilbert
space H is said to have property X0,1 if, given a weak-* continuous linear
functional τ on A with kτk ≤ 1, z1, ..., zq in H and ǫ > 0, there are vectors
x and y in H such that
(1) kxk ≤ 1 and kyk ≤ 1,
(2) k[x ⊗ zj]Ak < ǫ and k[zj ⊗ y]Ak < ǫ for 1 ≤ j ≤ q,
(3) kτ − [x ⊗ y]Ak < ǫ.
It's easy to see that the infinite ampliation of a weak-* closed subspace of
B(H) has property X0,1. Thus, intuitively, a weak-* closed subspace of B(H)
that has property X0,1 can be thought of as having "approximately infinite"
multiplicity. It was shown in [BFP85] that property X0,1 implies property
A1(1).
We will show that T has property X0,1. Since this property is inherited
by weak-* closed subspaces, it will follow that S has property X0,1, and
hence that S has property A1(1). It is easy to show that any weak-* closed
subspace of operators with property A1(1) is weakly closed (see for example
Proposition 59.2 of [Con00]). Thus this will imply the desired result that S
is weakly closed.
4.4. Approximate factorization.
THE STRUCTURE OF AN ISOMETRIC TUPLE
14
Lemma 4.4. Given unit vectors x, z1, ..., zq in H and ǫ > 0, there are vectors
ξ, ζ1, ..., ζq in F 2
n such that
(1) kξk < √q(1 + ǫ)1/2,
(2) kζik < (1 + ǫ)1/2 for 1 ≤ i ≤ q,
(3) [x ⊗ zi]T = φ([ξ ⊗ ζi]Mn) for 1 ≤ i ≤ q.
ik < (1+ǫ)1/2, kζ ′
q, ζ ′
ik < (1+ǫ)1/2 and [x⊗zi]T = φ([υ′
Proof. Since Mn has property A1(1), there are vectors υ′
n such that kυ′
F 2
for 1 ≤ i ≤ q.
Let Vi = R12k for 1 ≤ i ≤ q, so that V1, ..., Vq are isometries in Rn with
pairwise orthogonal ranges. Set ξ = Pq
i and ζi = Viζ ′
i for 1 ≤ i ≤ q.
Then kξk < √q(1 + ǫ)1/2, kζik < (1 + ǫ)1/2 and for T in T ,
1, ..., ζ ′
i⊗ζ ′
q in
i]Mn)
i=1 Viυ′
1, ..., υ′
(φ([ξ ⊗ ζi]Mn), T ) = (Φ−1(T )ξ, ζi)
Xj=1
= (Φ−1(T )
q
Vjυ′
j, Viζ ′
i)
i, ζ ′
i)
i]Mn), T )
= (Φ−1(T )υ′
= (φ([υ′
i ⊗ ζ ′
= ([x ⊗ zi]T , T ).
Hence [x ⊗ zi]T = φ([ξ ⊗ ζi]Mn).
Lemma 4.5. Let η be a unit vector contained in the algebraic span of {ξw :
w ∈ F∗
n}. Then there are words u and v in F∗
n such that
(cid:3)
LuRvη = Lξ∅ = Rξ∅,
where L is an isometry in Ln, and R is an isometry in Rn with range or-
thogonal to the range of R1.
Proof. Expand η as η = Pw≤m awξw for some m ≥ 0. Let u = 12m and
let v = 1m2. Then LuRvη = Pw≤m awξuwv. Set L = Pw≤m awLuwv and
R = Pw≤m awRuwv. Then LuRvη = Lξ∅ = Rξ∅, and it's clear that the
w ≤ m and w′ ≤ m,
It remains to show that L and R are isometries. For w and w′ in F+
range of R is orthogonal to the range of R1.
n with
L∗
vL∗
wLw′Lv =
I
0
if w = w′,
otherwise.
THE STRUCTURE OF AN ISOMETRIC TUPLE
15
This gives
awaw′L∗
uwvLuw′v
awaw′L∗
vL∗
wLw′Lv
L∗L = Xw≤m Xw′≤m
= Xw≤m Xw′≤m
= Xw≤m
aw2I
= I,
where the last equality follows from the fact that η is a unit vector. Thus
L is an isometry, and it follows from a similar computation that R is an
isometry.
(cid:3)
Lemma 4.6. Given unit vectors z1, ..., zq in H and ǫ > 0, there exists a unit
vector x in H and vectors ξ, ζ1, ..., ζq in F 2
n such that
(1) kξk < √q(1 + ǫ)1/2,
(2) kζik < (1 + ǫ)1/2 for 1 ≤ i ≤ q,
(3) ξ = kξkLξ∅ = kξkRξ∅, where L is an isometry in Ln, and R is an
isometry in Rn with range orthogonal to the range of R1,
(4) k[x ⊗ zi]T − φ([ξ ⊗ ζi]Mn)k < ǫ for 1 ≤ i ≤ q.
Proof. Let x′ be any unit vector in H. By Lemma 4.4, there are vectors
ξ′, ζ ′
q in F 2
1, ..., ζ ′
n such that
(1) kξ′k < √q(1 + ǫ)1/2,
(2) kζ ′
(3) [x′ ⊗ zi]T = φ([ξ′ ⊗ ζ ′
ik < (1 + ǫ)1/2 for 1 ≤ i ≤ q,
i]Mn) for 1 ≤ i ≤ q.
Let η be a vector contained in the algebraic span of {ξw : w ∈ F∗
kηk < √q(1 + ǫ)1/2 and kξ′ − ηk < ǫ/(1 + ǫ)1/2. Then
n} such that
k[x′ ⊗ zi]T − φ([η ⊗ ζ ′
i]Mn)k ≤ k[x′ ⊗ zi]T − φ([ξ′ ⊗ ζ ′
i]Mnk
+k[(ξ′ − η) ⊗ ζ ′
i]Mn)k
≤ kξ′ − ηkkζ ′
ik
< ǫ
for 1 ≤ i ≤ q.
By Lemma 4.5, there are words u and v in F+
n such that
LuRvη = kηkLξ∅ = kηkRξ∅,
THE STRUCTURE OF AN ISOMETRIC TUPLE
16
where L is an isometry in Ln, and R is an isometry in Rn with range or-
thogonal to the range of R1. Set x = Sux′, ξ = LuRvη and ζi = Rvζ ′
i for
1 ≤ i ≤ q. Then for T in T ,
([x ⊗ zi]T − φ([ξ ⊗ ζi]Mn), T ) = ([Sux′ ⊗ zi]T − φ([η ⊗ ζ ′
= ([x′ ⊗ zi]T − φ([η ⊗ ζ ′
≤ k[x′ ⊗ zi]T − φ([η ⊗ ζ ′
< ǫkTk.
i]Mn), T )
i]Mn), T Su)
i]Mn)kkT Suk
Hence k[x ⊗ zi]T − φ([ξ ⊗ ζi]Mn)k < ǫ.
The following result is implied by Lemma 1.2 in [Kri01].
(cid:3)
Lemma 4.7. Given a proper isometry R in Rn, vectors ζ1, ..., ζq in F 2
ǫ > 0, there exists k ≥ 1 such that k(R∗)kζik < ǫ for 1 ≤ i ≤ q.
Lemma 4.8. Given a proper isometry S in S, vectors u and v in H and
ǫ > 0, there exists k ≥ 1 such that k[u ⊗ (S∗)kv]Sk < ǫ.
Proof. Since Ln has property A1, there are vectors µ and ν in F 2
[u ⊗ v]S = φ([µ ⊗ ν]Ln). Thus for A in S,
n such that
n and
([u ⊗ (S∗)kv]S , A) = ([µ ⊗ (Φ−1(S)∗)kν]Ln, Φ−1(A))
= (Φ−1(A)µ, (Φ−1(S)∗)kν)
≤ kAkkµkk(Φ−1(S)∗)kνk,
which gives k[u⊗(S∗)kv]Sk ≤ kµkk(Φ−1(S)∗)kνk. Since Φ−1(S) is a a proper
isometry in Ln, and since Ln and Rn are unitarily equivalent, the result now
follows by Lemma 4.7.
(cid:3)
Lemma 4.9. Given unit vectors z1, ..., zq in H and ǫ > 0, there exists a unit
vector x in H and vectors ξ, ζ1, ..., ζq in F 2
n such that
(1) kξk < √q(1 + ǫ)1/2,
(2) kζik < (1 + ǫ)1/2 for 1 ≤ i ≤ q,
(3) ξ = kξkLξ∅ = kξkRξ∅, where L is an isometry in Ln, and R is an
isometry in Rn with range orthogonal to the range of R1,
(4) kR∗ζik < ǫ for 1 ≤ i ≤ q,
(5) (Φ(L)kx, x) < ǫ for k ≥ 1,
(6) k[x ⊗ zi]T − φ([ξ ⊗ ζi]Mn)k < ǫ for 1 ≤ i ≤ q.
Proof. By Lemma 4.6, there exists a unit vector x′ in H and vectors ξ′, ζ1, ..., ζq
in F 2
n such that
THE STRUCTURE OF AN ISOMETRIC TUPLE
17
(1) kξ′k < √q(1 + ǫ)1/2,
(2) kζik < (1 + ǫ)1/2 for 1 ≤ i ≤ q,
(3) ξ′ = kξ′kL′ξ∅ = kξ′kR′ξ∅, where L′ is an isometry in Ln and R′ is an
isometry in Rnwith the range of R′ orthogonal to the range of R1,
(4) k[x′ ⊗ zi]T − φ([ξ′ ⊗ ζi]Mn)k < ǫ for 1 ≤ i ≤ q.
1)m(R′)∗ζik <
By Lemma 4.7 and Lemma 4.8, there exists m ≥ 1 such that k(R∗
1 L′
ǫ for 1 ≤ i ≤ q and k[x′ ⊗ (S∗
1 )mΦ(L′)∗x′]Sk < ǫ. Set ξ = Lm
and R = R′Rm
1 . Then ξ = kξkLξ∅ = kξkRξ∅, L is an isometry in Ln, and R
is an isometry in Rn with range orthogonal to the range of R1. For 1 ≤ i ≤ q,
this gives kR∗ζik = k(R∗
1)m(R′)∗ζik < ǫ.
1 ξ′, L = Lm
Let x = Sm
1 x′. Then for k ≥ 1, we compute
(Φ(L)kx, x) = (Φ(Lm
1 L′)kSm
1 x′, Sm
1 )kx′, Sm
1 x′)
1 x′)
1 Φ(L′Lm
= (Sm
= (Φ(L′Lm
= (Φ(L′Lm
= ([x′ ⊗ (S∗
≤ k[x′ ⊗ (S∗
< ǫ.
1 )kx′, x′)
1 )k−1x′, (S∗
1 )mΦ(L′)∗x′)
1 )mΦ(L′)∗x′]S , Φ(L′Lm
1 )mΦ(L′)∗x′]Skk(L′Lm
1 )k−1)
1 )k−1k
Finally, for T in T we have
([x ⊗ zi]T − φ([ξ ⊗ ζi]Mn), T ) = ([x′ ⊗ zi]T − φ([ξ′ ⊗ ζi]Mn), T Sm
1 )
≤ k[x′ ⊗ zi]T − φ([ξ′ ⊗ ζi]MnkkT Sm
1 k
< ǫkTk.
Thus k[x ⊗ zi]T − φ([ξ ⊗ ζi]Mn)k < ǫ.
4.5. Approximately orthogonal vectors. The following lemma is extracted
from the proof of Theorem 4.3 in [Ber98].
(cid:3)
Lemma 4.10. Given two isometries R and R′ in Rn with orthogonal ranges
and vectors ξ and µ in F 2
n with µ in the kernel of R∗, define
Then
µk =
1
√k
k[ξ ⊗ µk]Mnk ≤
k
RjR′µ.
Xj=1
1
√kkµkkDk − 1k1,
THE STRUCTURE OF AN ISOMETRIC TUPLE
18
where Dk denotes the k-th Dirichlet kernel and k · k1 denotes the L1 norm.
Lemma 4.11. Given unit vectors z1, ..., zq in H and ǫ > 0, there exists a
unit vector x in H such that k[x ⊗ zi]T k < ǫ for 1 ≤ i ≤ q.
Proof. We may suppose that ǫ < 1. Using the fact that lim k−1/2kDkk1 = 0,
where Dk denotes the k-th Dirichlet kernel and k · k1 denotes the L1 norm,
choose k ≥ 1 such that 2(q/k)−1/2kDk − 1k1 < ǫ/(3(1 + ǫ)). Next choose
ǫ′ > 0 such that
ǫ′ < min(cid:26)1,
ǫ(1 − ǫ)
3√k
,
ǫ(1 − ǫ)
6√q
,
kǫ
k2 − k(cid:27) .
By Lemma 4.9, there exists a unit vector x′ in H and vectors ξ′, ζ1, ..., ζq in
F 2
n such that
(1) kξ′k < √q(1 + ǫ′)1/2,
(2) kζik < (1 + ǫ′)1/2 for 1 ≤ i ≤ q,
(3) ξ′ = kξ′kLξ∅ = kξ′kRξ∅, where L is an isometry in Ln, and R is an
isometry in Rn with range orthogonal to the range of R1,
(4) kR∗ζik < ǫ′ for 1 ≤ i ≤ q,
(5) (Φ(L)kx′, x′) < ǫ′ for k ≥ 1,
(6) k[x′ ⊗ zi]T − φ([ξ′ ⊗ ζi]Mn)k < ǫ′ for 1 ≤ i ≤ q.
By (4) we can write ζi = µi + νi, where µi is in the kernel of R∗ and kνik < ǫ′.
Let ξ = k−1/2Pk−1
j=0 L1Ljξ′. Then by (3) we can write ξ as
k−1
ξ =
L1Ljξ′
k−1
1
Xj=0
√k
= kξ′k√k
Xj=0
= kξ′k√k
Xj=1
k
L1Lj+1ξ∅
L1Ljξ∅,
THE STRUCTURE OF AN ISOMETRIC TUPLE
19
which implies kξk = kξ′k. Applying (3) again, we can also write ξ as
ξ =
k−1
L1Ljξ′
k−1
1
Xj=0
√k
= kξ′k√k
Xj=0
= kξ′k√k
Xj=0
= kξ′k√k
Xj=1
k−1
k
L1LjRξ∅
Rj+1R1ξ∅
RjR1ξ∅.
By Lemma 4.10 and the choice of k, this gives
k[ξ ⊗ µi]Mnk ≤
1
√kkξkkµikkDk − 1k1
(1 + ǫ′)1/2kµikkDk − 1k1
≤ r q
< ǫ(1 − ǫ)/3.
k
j=0 Φ(L1Lj). Then kSk ≤ √k, so for T
Let y = Sx′, where S = k−1/2Pk−1
in T ,
([y ⊗ zi]T − φ([ξ ⊗ ζi]Mn), T ) = ([x′ ⊗ zi]T − φ([ξ′ ⊗ ζi]Mn), T S)
≤ k[x′ ⊗ zi]T − φ([ξ′ ⊗ ζi]Mn)kkT Sk
< ǫ′√kkTk
< (ǫ(1 − ǫ)/3)kTk,
which gives k[y ⊗ zi]T − φ([ξ ⊗ ζi]Mn)k < ǫ(1 − ǫ)/3. Since
k[ξ ⊗ νi]Mnk ≤ kξkkνik < √q(1 + ǫ′)1/2ǫ′ < ǫ(1 − ǫ)/3,
this gives
k[y ⊗ zi]T k ≤ k[y ⊗ zi]T − φ([ξ ⊗ ζi]Mn)k + k[ξ ⊗ µi]Mnk + k[ξ ⊗ νi]Mnk
< ǫ(1 − ǫ).
THE STRUCTURE OF AN ISOMETRIC TUPLE
20
k−1
Finally, we compute
kyk2 = kSx′k2
Xj=0
k X0≤i<j≤k−1
= kx′k2 +
1
Φ(L1Lj)x′(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
= (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
1
√k
1
2
k X0≤i<j≤k−1
k2 − k
ǫ′.
k
≥ 1 −
≥ 1 −
> 1 − ǫ.
(x′, Φ(L)j−ix′) +
(Φ(L)i−jx′, x′)
(x′, Φ(L)j−ix′) −
(Φ(L)i−j x′, x′)
1
k X0≤j<i≤k−1
k X0≤j<i≤k−1
1
Hence taking x = (1 − ǫ)−1y, kxk ≥ 1 and k[x ⊗ zi]T k < ǫ for 1 ≤ i ≤ q. (cid:3)
Lemma 4.12. Given unit vectors z1, ..., zq in H and ǫ > 0, there exists an
intertwining operator X : F 2
n → H such that kXξ∅k = 1 and k[Xξ∅⊗zi]T k <
ǫ for 1 ≤ i ≤ q.
Proof. By Lemma 4.11, there exists a unit vector x in H such that k[x ⊗
zi]T k < ǫ for 1 ≤ i ≤ q. By Theorem 2.7 of [DLP05], x is in the range of
an intertwining operator X′ : F 2
n such
that X′ξ = x. The result now follows from the fact that the set of vectors
{Rξ∅ : R ∈ Rn} is dense in F 2
n , and the fact that for R in Rn, the operator
X′R is intertwining.
n → H. Hence there is a vector ξ in F 2
(cid:3)
Lemma 4.13. Let X : F 2
Then given ǫ > 0, there is a word v in F∗
n → H be an intertwining operator with kXξ∅k = 1.
n such that
k[XRvξ∅ ⊗ XRvξ∅]T − φ([ξ∅ ⊗ ξ∅]Mn)k < ǫ.
Proof. Since X∗X is an L-Toeplitz operator, by Lemma 4.5 of [Ken11], there
is a word v in F∗
vX∗XRv
is also an L-Toeplitz operator. Let ξ = (R∗
vX∗XRv−I)ξ∅, so that kξk < ǫ/2.
For w in F∗
vX∗XRvξ∅−ξ∅k < ǫ/2. Note that R∗
n such that kR∗
n, since (Lwξ, ξ∅) = 0 we can write
(SwXRvξ∅, XRvξ∅) = (Lwξ∅, R∗
vX∗XRvξ∅)
= (Lwξ∅, ξ∅) + (Lwξ∅, ξ) + (Lwξ, ξ∅).
THE STRUCTURE OF AN ISOMETRIC TUPLE
21
Similarly,
(S∗
wXRvξ∅, XRvξ∅) = (L∗
= (L∗
vX∗XRvξ∅, ξ∅)
wR∗
wξ∅, ξ∅) + (L∗
wξ∅, ξ) + (L∗
wξ, ξ∅).
This gives
[XRvξ∅ ⊗ XRvξ∅]T = φ([ξ∅ ⊗ ξ∅]Mn + [ξ ⊗ ξ∅]Mn + [ξ∅ ⊗ ξ]Mn),
so we conclude that
k[XRvξ∅ ⊗ XRvξ∅]T − φ([ξ∅ ⊗ ξ∅]Mn)k ≤ k[ξ ⊗ ξ∅]Mn + [ξ∅ ⊗ ξ]Mnk
≤ 2kξkkξ∅k
< ǫ,
as required.
(cid:3)
Lemma 4.14. Given unit vectors z1, ..., zq in H and ǫ > 0, there exists an
intertwining operator X : F 2
n → H such that kXξ∅k = 1, k[Xξ∅ ⊗ zi]T k < ǫ
for 1 ≤ i ≤ q and
k[Xξ∅ ⊗ Xξ∅]T − φ([ξ∅ ⊗ ξ∅]Mn)k < ǫ.
Proof. By Lemma 4.12, there exists an intertwining operator X′ : F 2
n → H
such that kX′ξ∅k = 1 and k[X′ξ∅ ⊗ zi]T k < ǫ for 1 ≤ i ≤ q. By Lemma
4.13, there is a word v in F∗
n such that
k[X′Rvξ∅ ⊗ X′Rvξ∅]T − φ([ξ∅ ⊗ ξ∅]Mn)k < ǫ.
Let X = X′Rv. Then
kXξ∅k = kX′Rvξ∅k = kX′Lvξ∅k = kSvX′ξ∅k = kX′ξ∅k = 1.
For T in T ,
([Xξ∅ ⊗ zi]T , T ) = ([X′Rvξ∅ ⊗ zi]T , T )
= ([X′Lvξ∅ ⊗ zi]T , T )
= ([SvX′ξ∅ ⊗ zi]T , T )
= ([X′ξ∅ ⊗ zi]T , T Sv)
≤ k[X′ξ∅ ⊗ zi]T kkTk.
Hence k[Xξ∅ ⊗ zi]T k < ǫ for 1 ≤ i ≤ q.
4.6. The strong factorization property.
(cid:3)
THE STRUCTURE OF AN ISOMETRIC TUPLE
22
Theorem 4.15. Given a weak-* continuous linear functional τ on T with
kτk ≤ 1, unit vectors z1, ..., zq in H and ǫ > 0, there are vectors x and y in
H such that
(1) kxk ≤ 1 and kyk ≤ 1,
(2) kτ − [x ⊗ y]T k < ǫ,
(3) k[x ⊗ zi]T k < ǫ and k[zi ⊗ y]T k < ǫ for 1 ≤ i ≤ q.
In other words, T has property X0,1.
Proof. Choose ǫ′ > 0 such that ǫ′ < ǫ and 1 − (1 + 2ǫ′)−2(1 − ǫ′) < ǫ. Since
n with kξk ≤ 1 + ǫ′/2
Mn has property A1(1), there are vectors ξ and υ in F 2
and kυk ≤ 1 + ǫ′/2 such that τ = φ([ξ ⊗ υ]Mn). Since ξ∅ is cyclic for
Ln, there are A and B in Ln such that kAξ∅ − ξk < ǫ′/(4(1 + ǫ′)) and
kBξ∅ − υk < ǫ′/(4(1 + ǫ′)). Then
kAξ∅k ≤ kAξ∅ − ξk + kξk < 1 + ǫ′,
and similarly kBξ∅k < 1 + ǫ′. This gives
k[Aξ∅ ⊗ Bξ∅]Mn − [ξ ⊗ υ]Mnk ≤ k[(Aξ∅ − ξ) ⊗ Bξ∅]k
+k[ξ ⊗ (Bξ∅ − υ)]Mnk
≤ kAξ∅ − ξkkBξ∅k + kξkkBξ∅ − υk
< ǫ′/2.
By Lemma 4.14, there is an intertwining operator X : F 2
n → H such
that kXξ∅k = 1, k[Xξ∅ ⊗ zi]T k < ǫ′/(kAk + kBk) for 1 ≤ i ≤ q and
k[Xξ∅ ⊗ Xξ∅]T − φ([ξ∅ ⊗ ξ∅]Mn)k < ǫ′/(2(kAk +kBk)2). Note that since T
is self-adjoint, we also have k[zi ⊗ Xξ∅]T k < ǫ′/(kAk + kBk) for 1 ≤ i ≤ q.
Define vectors x′ and y′ in H by x′ = Φ(A)Xξ∅ and y′ = Φ(B)Xξ∅. Then
kx′k2 = kΦ(A)Xξ∅k2
= kΦ(A)Xξ∅k2 − kAξ∅k2 + kAξ∅k2
= ([Xξ∅ ⊗ Xξ∅]T − φ([ξ∅ ⊗ ξ∅]Mn), Φ(A∗A)) + kAξ∅k2
≤ k[Xξ∅ ⊗ Xξ∅]T − φ([ξ∅ ⊗ ξ∅]Mn)kkAk2 + kAξ∅k2
< 1 + 2ǫ′,
and similarly, ky′k2 < 1 + 2ǫ′. For T in T ,
THE STRUCTURE OF AN ISOMETRIC TUPLE
23
([x′ ⊗ y′]T − φ([Aξ∅ ⊗ Bξ∅]Mn), T )
= ([Φ(A)Xξ∅ ⊗ Φ(B)Xξ∅]T − φ([Aξ∅ ⊗ Bξ∅]Mn), T )
= ([Xξ∅ ⊗ Xξ∅]T − φ([ξ∅ ⊗ ξ∅]Mn), Φ(A)∗T Φ(B))
≤ k[Xξ∅ ⊗ Xξ∅]T − φ([ξ∅ ⊗ ξ∅]Mn)kkAkkBkkTk
<
ǫ′
2 kTk,
which implies k[x′ ⊗ y′]T − φ([Aξ∅ ⊗ Bξ∅]Mn)k < ǫ′/2. Thus
k[x′ ⊗ y′]T − τk = k[x′ ⊗ y′]T − φ([ξ ⊗ υ]Mn)k
≤ k[x′ ⊗ y′]T − φ([Aξ∅ ⊗ Bξ∅]Mn)k
+k[Aξ∅ ⊗ Bξ∅]Mn − [ξ ⊗ υ]Mnk
< ǫ′.
For 1 ≤ i ≤ q,
k[x′ ⊗ zi]T k = k[AXξ∅ ⊗ zi]T k ≤ kAkk[Xξ∅ ⊗ zi]T k < ǫ′,
and similarly, k[zi ⊗ y′]T k < ǫ′.
Now take x = (1 + 2ǫ′)−1x′ and y = (1 + 2ǫ′)−1y′. Then by choice of ǫ′
we get kxk ≤ 1 and kyk ≤ 1. Similarly, k[x ⊗ zi]T k < ǫ and k[zi ⊗ y]T k < ǫ
for 1 ≤ i ≤ q. Finally, we have
k[x ⊗ y]T − τk ≤ (1 + 2ǫ′)−2k[x′ ⊗ y′]T − τk + (1 − (1 + 2ǫ′)−2)kτk
< 1 − (1 + 2ǫ′)−2(1 − ǫ′)
< ǫ,
as required.
(cid:3)
4.7. Absolute continuity and analyticity.
Theorem 4.16. For n ≥ 2, every absolutely continuous isometric n-tuple is
analytic.
Proof. For n ≥ 2, let S = (S1, . . . , Sn) be an absolutely continuous isomet-
ric n-tuple, and let S denote the weak-* closed unital algebra generated by
S1, . . . , Sn. By Corollary 1.2 of [DY08], S is isomorphic to the noncommu-
tative analytic Toeplitz algebra Ln. By Theorem 4.15, S has property X0,1,
and hence has property A1(1). Therefore, by the discussion in Section 4.3,
S is weakly closed, and hence S is actually the free semigroup algebra (i.e.
THE STRUCTURE OF AN ISOMETRIC TUPLE
24
the weakly closed algebra) generated by S1, . . . , Sn. Since S is isomorphic
to Ln, this implies that S is analytic.
(cid:3)
The next result follow from Theorem 4.12 of [Ken11].
Corollary 4.17. For n ≥ 2, let S = (S1, . . . , Sn) be an absolutely continuous
isometric n-tuple acting on a Hilbert space H. Then the wandering vectors
for S span H.
It was shown in Corollary 5.8 of [Ken11] that every analytic isometric
tuple is hyperreflexive with hyperreflexivity constant at most 3, but the next
result can also be proved directly using Theorem 4.15 of the present paper
and Theorem 3.1 of [Ber98].
Corollary 4.18. Absolutely continuous row isometries are hyperreflexive
with hyperreflexivity constant at most 3.
5. Singular isometric tuples
In Theorem 4.16, we showed that for n ≥ 2, an isometric n-tuple is abso-
lutely continuous if and only if it is analytic. With this operator-algebraic
characterization of an absolutely continuous isometric tuple, we are now able
to give an operator-algebraic characterization of a singular isometric tuple.
Theorem 5.1. For n ≥ 2, an isometric n-tuple is singular if and only if the
free semigroup algebra it generates is a von Neumann algebra.
Proof. Let V = (V1, . . . , Vn) be an isometric n-tuple, and let V denote the
free semigroup algebra (i.e. the weakly closed algebra) generated by V . If
V is a von Neumann algebra, then V has no absolutely continuous part
since, by Theorem 4.16, an absolutely continuous isometric tuple is analytic,
and the noncommutative analytic Toeplitz algebra Ln is not self-adjoint by
Corollary 1.5 of [DP99].
Conversely, if V is singular then it has no analytic restriction to an in-
variant subspace since, by Theorem 4.16, an absolutely continuous isometric
tuple is analytic. Thus by Theorem 3.3, V is a von Neumann algebra.
(cid:3)
Example 2.2 showed that it is possible for an absolutely continuous unitary
to generate a von Neumann algebra. Theorem 5.1 implies that there is no
higher-dimensional analogue of this phenomenon.
Recall that a family of operators is said to be reductive if every subspace
invariant for the family is also coinvariant.
THE STRUCTURE OF AN ISOMETRIC TUPLE
25
Corollary 5.2. For n ≥ 2, every reductive unitary n-tuple is singular.
Proof. Let V = (V1, . . . , Vn) be a reductive isometric n-tuple, and let V
denote the free semigroup algebra generated by V . By the dichotomy for free
semigroup algebras, Corollary 4.13 of [Ken11], if V is not a von Neumann
algebra, then there is a vector x that is wandering for V . Let V[x] denote the
cyclic invariant subspace generated by x. Then the subspace Pn
i=1 ViV[x]
is invariant for V but not coinvariant, which would contradict that V is
reductive. Thus V is a von Neumann algebra and V is singular by Theorem
5.1.
(cid:3)
Example 5.3. By Theorem 5.1, for n ≥ 2 an isometric n-tuple is singular
if and only if the free semigroup algebra it generates is a von Neumann
algebra. The existence of a self-adjoint free semigroup algebra on two or more
generators was conjectured in [DKP01], but it took some time for the first
example to be constructed. In [Read05], Read showed that B(ℓ2) is generated
as a free semigroup algebra on two generators. In [Dav06], Davidson gave an
exposition of Read's construction and showed that it could be generalized to
show that B(ℓ2) is generated as a free semigroup algebra on n generators for
every n ≥ 2. By our characterization of singularity, this gives an example of
a singular isometric n-tuple for every n ≥ 2.
6. The Lebesgue-von Neumann-Wold decomposition
In Theorem 4.16, we showed that for n ≥ 2, an isometric n-tuple is abso-
lutely continuous if and only if it is analytic. In Theorem 5.1, we showed that
for n ≥ 2, an isometric n-tuple is singular if and only if the free semigroup
algebra (i.e. the weakly closed algebra) it generates is a von Neumann alge-
bra. With these operator-algebraic characterizations of absolute continuity
and singularity, we will be able to prove the Lebesgue-von Neumann-Wold
decomposition of an isometric tuple.
In the classical case, the Lebesgue decomposition of a measure guarantees
that every unitary splits into absolutely continuous and singular parts. For
n ≥ 2, it turns out that it is possible for a unitary n-tuple to be irreducible
and neither absolutely continuous nor singular.
Definition 6.1. An isometric n-tuple V = (V1, . . . , Vn) is said to be of
dilation type if it has no summand that is absolutely continuous or singular.
Note that by the Wold decomposition of an isometric tuple, Proposition
3.5, an isometric n-tuple of dilation type is necessarily unitary. The next
THE STRUCTURE OF AN ISOMETRIC TUPLE
26
result provides a characterization of an isometric tuple of dilation type as a
minimal dilation, in the sense of Section 3.3.
Proposition 6.2. Let V = (V1, . . . , Vn) be an isometric n-tuple of dilation
type. Then there is a subspace H coinvariant under V such that H is cyclic
for V and the compression of V to H ⊥ is a unilateral n-shift. In other words,
V is the minimal isometric dilation of its compression to H.
Proof. Note that since V has no summand that is absolutely continuous,
by Proposition 3.5 V is necessarily a unitary n-tuple. Let V denote the
free semigroup algebra generated by V , and let P be the projection from
Theorem 3.3 applied to V. Let H be the range of P , so that H is coinvariant
under V .
Let K = (H +Pn
i=1 ViH) ⊖ H. Then K is wandering for the compres-
sion of V to H ⊥. If K = 0, then by Theorem 3.3, V can be decomposed
into the direct sum of a self-adjoint free semigroup algebra and an analytic
free semigroup algebra. By the characterization of singular isometric tuples,
Corollary 5.1, this would contradict that V is of dilation type. Thus K 6= 0.
The fact that K is cyclic follows from the fact that H is cyclic.
(cid:3)
Example 6.3 (An irreducible isometric tuple of dilation type). For n ≥ 2,
define isometries V1, . . . , Vn on ℓ2(N) by
Vkel = en(l−1)+k,
where {el}∞
each Vk is spanned by the orthonormal set
l=1 is the standard orthonormal basis of ℓ2(N). Then the range of
{Vkel}∞
l=1 = {en(l−1)+k}∞
l=1 = {el : l ≡ k mod n}.
Therefore, the operators V1, . . . , Vn are isometries with mutually orthogonal
ranges, meaning V = (V1, . . . , Vn) is an isometric tuple. We will show that
V is an irreducible isometric tuple of dilation type.
Since the vector e1 is fixed by V1, it is straightforward to check that the
k=1 is weakly convergent to the rank one projection e1e∗
1 }∞
sequence {V k
1. In
1 is contained in the von Neumann algebra W∗(V ) generated
particular, e1e∗
by V . (In fact, this is the projection provided by Theorem 3.3). Since the
vector e1 is cyclic for V , it follows that W∗(V ) = B(ℓ2(N)), and hence that
V is irreducible.
To see that V is of dilation type, it suffices to show that V is neither
singular nor absolutely continuous. Since V is irreducible from above, if V
THE STRUCTURE OF AN ISOMETRIC TUPLE
27
was singular then by Theorem 5.1, the free semigroup algebra W(V ) gen-
erated by V would be B(ℓ2(N). However, the vector e2 is wandering for
V and the vector e1 is orthogonal to the wandering subspace spanned by
{Vwe2 : w ∈ F∗
n}, which implies that W(V ) is not transitive, and hence
that W(V ) is properly contained in B(ℓ2(N). Thus V is not singular. The
fact that V is not absolutely continuous follows from Theorem 4.16 and the
observation made above that the sequence {V k
k=1 is weakly convergent to
the projection e1e∗
1.
1 }∞
Example 6.4 (A family of irreducible isometric tuples of dilation type). It
was shown in Corollary 6.6 of [DKS01] that the minimal isometric dilation of
a contractive n-tuple A = (A1, . . . , An) acting on a finite-dimensional space
i = I and
C∗(A) has a minimal coinvariant subspace that is cyclic for C∗(A). These
conditions are satisfied, for example, by the contractive tuple A = (A1, A2),
where
is an irreducible unitary n-tuple if and only if both Pn
i=1 AiA∗
A1 = 0 1
0 0 ! ,
A2 = 0 0
1 0 ! .
Thus the minimal isometric dilation of A is an example of an irreducible
isometric tuple of dilation type. A similar construction can be carried out
for all n ≥ 2.
Theorem 6.5 (Lebesgue-von Neumann-Wold Decomposition). Let V =
(V1, . . . , Vn) be an isometric n-tuple. Then V decomposes as
V = Vu ⊕ Va ⊕ Vs ⊕ Vd,
where Vu is a unilateral n-shift, Va is an absolutely continuous unitary n-
tuple, Vs is a singular unitary n-tuple, and Vd is a unitary n-tuple of dilation
type.
Proof. The case for n = 1 follows by the discussion in Section 2. Thus we
can suppose that n ≥ 2. By the Wold decomposition of an isometric tuple,
Proposition 3.5, we can decompose V as
where Vu is a unilateral n-shift and U is a unitary n-tuple.
V = Vu ⊕ U,
THE STRUCTURE OF AN ISOMETRIC TUPLE
28
By the characterization of an absolutely continuous isometric n-tuple as
analytic, Theorem 4.16, and the characterization of a singular isometric n-
tuple, Corollary 5.1, an isometric n-tuple cannot be both absolutely contin-
uous and singular. Therefore, we can decompose U as
U = Va ⊕ Vs ⊕ Vd,
where Va is an absolutely continuous isometric n-tuple, Vs is a singular iso-
metric n-tuple, and Vd is of dilation type. Thus we can further decompose
V as
V = Vu ⊕ Va ⊕ Vs ⊕ Vd,
as required.
(cid:3)
The next result follows from combining Proposition 6.2 and Theorem 3.3.
Proposition 6.6. Let V = (V1, . . . , Vn) be an isometric n-tuple of dilation
type acting on a Hilbert space H. Then there is a projection P and α ≥ 1
and such that the weakly closed algebra W(V ) generated by V is of the form
W(V1, . . . , Vn) = W∗(V )P + P ⊥W(V )P ⊥,
where P ⊥W(V1, . . . , Vn) P ⊥H≃ L(α)
n .
The next result follows from the Lebesgue-von Neumann-Wold decompo-
sition of an isometric tuple, Proposition 3.5, and Proposition 6.6.
Theorem 6.7. Let V = (V1, . . . , Vn) be an isometric n-tuple acting on a
Hilbert space H, and let V = Vu⊕Va⊕Vs⊕Vd be the Lebesgue-von Neumann-
Wold decomposition of V as in Theorem 6.5. Then there is a projection P
and α, β ≥ 0 such that the weakly closed algebra W(V ) generated by V is
W(V ) ≃ (Ln(Vu ⊕ Va))(α) ⊕ W∗(Vs) ⊕(cid:16)W∗(Vd)P + P ⊥W(Vd)P ⊥(cid:17) ,
n . The von Neumann algebra W∗(V1, . . . , Vn)
where P ⊥W(V1, . . . , Vn) P ⊥H≃ L(β)
generated by V is
W∗(V ) ≃ (B(ℓ2))(α) ⊕ W∗(Vs) ⊕ W∗(Vd).
Acknowledgement. The author is grateful to his advisor, Ken Davidson, for
his support and encouragement.
References
[Arv69]
W.B. Arveson, Subalgebras of C∗-algebras, Acta Mathematica 123 (1969),
141 -- 224.
[Arv75]
[Ber88]
[Ber98]
[BFP85]
[Bro78]
[Bun84]
[Cho74]
[Con00]
[Dav87]
[Dav01]
[Dav06]
[DKS01]
[DKP01]
[DLP05]
[DP98]
[DP99]
[DY08]
[Fra82]
[Ken11]
THE STRUCTURE OF AN ISOMETRIC TUPLE
29
W.B. Arveson, Interpolation problems in nest algebras, Journal of Functional
Analysis 20 (1975), No. 3, 208 -- 233.
H. Bercovici, Factorization theorems and the structure of operators on Hilbert
space, Annals of Mathematics 128 (1988), No. 2, 399 -- 413.
H. Bercovici, Hyper-reflexivity and the factorization of linear functionals,
Journal of Functional Analysis 158 (1998), No. 1, 242 -- 252.
H. Bercovici, C. Foias, C. Pearcy, Dual algebras with applications to invariant
subspaces and dilation theory, CBMS Regional Conference Series in Mathe-
matics 56 (1985), American Mathematical Society, Providence.
S.W. Brown, Some invariant subspaces for subnormal operators, Integral
Equations and Operator Theory 1 (1978), No. 3, 310 -- 333.
J. Bunce, Models for n-tuples of non-commuting operators, Journal of Func-
tional Analysis 57 (1984), No. 1, 21 -- 30.
M.D. Choi, A Schwarz inequality for positive linear maps on C∗-algebras,
Illinois Journal of Mathematics 18 (1974), No. 4, 565 -- 574.
J.B. Conway, A course in operator theory, Graduate Studies in Mathematics
21 (2000), American Mathematical Society, Providence.
K.R. Davidson, The distance to the analytic Toeplitz operators, Illinois Jour-
nal of Mathematics 31 (1987), No. 2, 265 -- 273.
K.R. Davidson, Free semigroup algebras: a survey, Operator Theory: Ad-
vances and Applications 129 (2000), Birkhauser, Bordeaux.
K.R. Davidson, B(H) is a free semigroup algebra, Proceedings of the Ameri-
can Mathematical Society 134 (2006), No. 2, 1753 -- 1757.
K.R. Davidson, D.W. Kribs, M.E. Shpigel, Isometric dilations of non-
commuting finite rank n-tuples, Canadian Journal of Mathematics 53 (2001),
506 -- 545.
K.R. Davidson, E. Katsoulis, D.R. Pitts, The structure of free semigroup
algebras, Journal für die reine und angewandte Mathematik 533 (2001), 99 --
125.
K.R. Davidson, J. Li, D.R. Pitts, Absolutely continuous representations and a
Kaplansky density theorem for free semigroup algebras, Journal of Functional
Analysis 224 (2005), No. 1, 160 -- 191.
K.R. Davidson, D.R. Pitts, The algebraic structure of noncommutative ana-
lytic Toeplitz algebras, Mathematische Annalen 311 (1998), 275 -- 303.
K.R. Davidson, D.R. Pitts, Invariant subspaces and hyper-reflexivity for free
semigroup algebras, Proceedings of the London Mathematical Society 78
(1999), No. 2, 401 -- 430.
K.R. Davidson, D. Yang, A note on absolute continuity in free semigroup
algebras, Houston Journal of Mathematics 34 (2008), 283 -- 288.
A. Frahzo, Models for non-commuting operators, Journal of Functional Anal-
ysis 48 (1982), No. 1, 1 -- 11.
M. Kennedy, Wandering vectors and the reflexivity of free semigroup algebras,
Journal für die reine und angewandte Mathematik 653 (2011), 47 -- 73.
[Kri01]
[LM78]
[LS75]
[MS10]
[Pop89a]
[Pop89b]
[Pop09]
[Pop91]
[Pop96]
[Read05]
[SF70]
[Wer52]
THE STRUCTURE OF AN ISOMETRIC TUPLE
30
D.W. Kribs, Factoring in non-commutative analytic Toeplitz algebras, Jour-
nal of Operator Theory 45 (2001), No. 1, 175 -- 193.
R.I. Loebl, P.S. Muhly, Analyticity and flows in von Neumann algebras, Jour-
nal of Functional Analysis 29 (1978), No. 2, 214 -- 252.
A.N. Loginov and V.S. Shulman, Hereditary and intermediate reflexivity of
W∗-algebras, Izvestiya Rossiiskoi Akademii Nauk Seriya Matematicheskaya
39 (1975), No. 6, 1260 -- 1273.
P. Muhly, B. Solel, Representations of Hardy algebras: absolute continuity,
intertwiners and superharmonic operators, preprint (2010), arXiv:1006.1398.
G. Popescu, Isometric dilations for infinite sequences of noncommuting oper-
ators, Transactions of the American Mathematical Society 316 (1989), No. 2,
523 -- 536.
G. Popescu, Multi-analytic operators and some factorization theorems, Indi-
ana University Mathematics Journal 38 (1989), No. 3, 693 -- 710.
G. Popescu, Noncommutative transforms and free pluriharmonic functions,
Advances in Mathematics 220 (2009), No. 3, 831 -- 893.
G. Popescu, von Neumann inequality for (B(H)n)1, Mathematica Scandinav-
ica 68 (1991), No. 2, 292 -- 304.
G. Popescu, Non-commutative disc algebras and their representations, Pro-
ceedings of the American Mathematical Society 124 (1996), No. 7, 2137 -- 2148.
C.J. Read, A large weak operator closure for the algebra generated by two
isometries, Journal of Operator Theory 54 (2005), No. 2, 305 -- 316.
B. Sz-Nagy, C. Foias, Harmonic analysis of operators on Hilbert space, Uni-
versitext (1970), Springer, North Holland.
J. Wermer, On invariant subspaces of normal operators, Proceedings of the
American Mathematical Society 3 (1952), 270 -- 277.
School of Mathematics and Statistics, Carleton University, 1125 Colonel
By Drive, Ottawa, Ontario K1S 5B6, Canada
E-mail address: [email protected]
|
1707.04552 | 2 | 1707 | 2018-11-24T20:54:49 | Relative commutant pictures of Roe algebras | [
"math.OA",
"math-ph",
"math-ph"
] | Let X be a proper metric space, which has finite asymptotic dimension in the sense of Gromov (or more generally, straight finite decomposition complexity of Dranishnikov and Zarichnyi). New descriptions are provided of the Roe algebra of X: (i) it consists exactly of operators which essentially commute with diagonal operators coming from Higson functions (that is, functions on X whose oscillation tends to 0 at infinity) and (ii) it consists exactly of quasi-local operators, that is, ones which have finite epsilon propogation (in the sense of Roe) for every epsilon>0. These descriptions hold both for the usual Roe algebra and for the uniform Roe algebra. | math.OA | math |
RELATIVE COMMUTANT PICTURES OF ROE
ALGEBRAS
J ´AN SPAKULA AND AARON TIKUISIS
Abstract. Let X be a proper metric space, which has finite
asymptotic dimension in the sense of Gromov (or more gener-
ally, straight finite decomposition complexity of Dranishnikov and
Zarichnyi). New descriptions are provided of the Roe algebra of
X: (i) it consists exactly of operators which essentially commute
with diagonal operators coming from Higson functions (that is,
functions on X whose oscillation tends to 0 at ∞), and (ii) it con-
sists exactly of quasi-local operators, that is, ones which have finite
ǫ-propogation (in the sense of Roe) for every ǫ > 0. These descrip-
tions hold both for the usual Roe algebra and for the uniform Roe
algebra.
1. Introduction
The Roe algebra is a C*-algebra constructed from a proper metric
space, which encodes "coarse" or "large-scale" properties of the space
(in the sense of Gromov). In typical applications, the space may be a
complete, non-compact Riemannian manifold with bounded geometry,
or a finitely generated group with the word metric. The origins of this
construction come from index theory, reflecting the insight that the Roe
algebra is large enough to contain indices of many operators with which
one wants to do index theory -- such as geometric differential operators --
yet small enough to have interesting and informative K-theory. It plays
a central role in the coarse Baum -- Connes conjecture, the study (and
partial confirmation) of which has been a fruitful endeavor, leading
to significant results concerning the Novikov conjecture and the scalar
curvature of Riemannian manifolds [3, 12, 15, 21, 22, 24, 25, 29, 30,
31, 32]. It furthermore appears in work on the essential spectrum of
Hamiltonian operators of quantum systems, Schrodinger operators, and
various other operators, which are affiliated to the appropriate versions
of Roe algebras [7, 8, 9, 19].
Roughly, the Roe algebra consists of bounded, locally compact op-
erators on something like L2(X) (where X is the underlying space)
which can be approximated by those with finite propogation. Here an
operator a has "finite propogation" if it is localized near to the diag-
onal; one way of making this precise is, that there exists R > 0 such
that for any f, f ′ ∈ Cb(X) (acting on L2(X) as diagonal operators -- by
multiplication), if the supports of f and f ′ are separated by a distance
1
2
J ´AN SPAKULA AND AARON TIKUISIS
of at least R, then f af ′ = 0. Operators in the Roe algebra are required
to be approximated in the operator norm by these finite propogation
operators.
The Roe algebra is an interesting object to study from an operator
algebraic perspective: that is, asking about its structure as an operator
algebra, and how this structure relates to the proper metric space from
which it is constructed. For example, Ozawa showed that exactness of
a group can be characterized by amenability of the corresponding uni-
form Roe algebra ([18]). The question of how much of the large -- scale
structure of a space is remembered by the Roe algebra, was partially
answered by JS and Rufus Willett: given two uniformly discrete proper
metric spaces with Yu's property A, their Roe algebras are ∗-isomorphic
if and only if the spaces are coarsely equivalent ([26]). In [28], Winter
and Zacharias showed an interesting one-way connection between the
asymptotic dimension of a metric space and the nuclear dimension of
the corresponding uniform Roe algebra; the latter is a numerical invari-
ant for amenable C*-algebras which is crucial in recent results in the
classification of amenable C*-algebras. Their result is that the nuclear
dimension of the Roe algebra is at most the asymptotic dimension of
the underlying uniformly discrete proper metric space, and they asked
the (still open) question of whether the reverse inequality also holds.
In this paper, we look at a fundamental question: exactly which
operators are in the Roe algebra? In [22], Roe defined the concept
of finite ǫ-propogation for an operator a on L2(X), as the following
variant of finite propogation: a has finite ǫ-propogation if there exists
R > 0 such that for any f, f ′ ∈ Cb(X), if the supports of f and
f ′ are separated by a distance of at least R, then kf af ′k ≤ ǫkf k ·
kf ′k. Operators with finite ǫ-propogation for all ǫ > 0 have also been
called quasi-local operators in the literature (originally from [21, Page
100]). It is a straightforward observation that, although limits of finite
propogation operators need not have finite propogation, limits of finite
ǫ-propogation operators have finite ǫ-propogation (that is, the set of
quasi-local operators is norm-closed). Therefore, all operators in the
Roe algebra are quasi-local.
The question we address is the converse: if an operator is quasi-local,
is it in the Roe algebra, i.e., is it approximated by operators with finite
propogation? We provide an affirmative answer in the situation that
the space has finite asymptotic dimension (as predicted by Roe), and
more generally under the hypothesis of straight finite decomposition
complexity of Dranishnikov and Zarichnyi [4]. The latter is a weaker
version of the "classical" finite decomposition complexity, as defined
by Guentner, Tessera, and Yu [13, 14].
A motivation for asking whether quasi-local operators are in the
Roe algebra, pointed out to the authors by Alexander Engel, is that
whereas indices of genuine differential operators are known to be in
RELATIVE COMMUTANT PICTURES OF ROE ALGEBRAS
3
the Roe algebra, corresponding arguments only shows that indices of
pseudo-differential operators (using their most natural representative)
are quasi-local (see [5, Section 2]). Since the Roe algebra is better
studied and understood than the C*-algebra of all quasi-local operators,
it is interesting and useful to know that a pseudo-differential operator
belongs to it; indeed, our result answers [5, Question 6.4] under the
assumption of straight finite decomposition complexity (this sort of
assumption is anticipated in the question).
Additionally, we expose that the above question can be reformulated
in terms of essential commutation with Higson functions, or in terms
of relative commutants.
So far we have been a bit vague about what we mean by the Roe
algebra (hiding behind the phrase "something like L2(X)"). This is
because in the literature there are two different versions of the Roe al-
gebra, the "Roe algebra" and the "uniform Roe algebra". Our results
apply to both of these C*-algebras, and indeed our main theorem is
formulated in a way that encompasses both, as well as the "uniform al-
gebra" UC ∗(X). The main result was proven by Lange and Rabinovich
for the uniform Roe algebra of Zd in [17]. Engel proved a special case
of the result, namely that for discrete groups G that are lattices in a
Riemannian manifold with bounded geometry and polynomial volume
growth, quasi-local operators that decay sufficiently quickly are in the
Roe algebra ([6, Corollary 2.33]).1
Let us now summarize the argument behind the main result: that
quasi-local operators are in the Roe algebra (assuming straight finite
decomposition complexity). Suppose for simplicity that X is a discrete
space with asymptotic dimension at most 1 -- for example a finitely
generated free group. This case is much more restricted than finite de-
composition complexity, but still difficult enough to allow us to convey
the main ideas. Let a be a quasi-local operator. Asymptotic dimension
at most 1 will allow us to decompose the space X into 2 pieces, X (0)
and X (1), each piece being a disjoint union of sets that are far apart
from each other and uniformly bounded in diameter. The characteristic
functions e(0), e(1) of these pieces produce a partition of unity, and di-
vides a into a sum of four pieces: e(i)ae(j) over i, j = 0, 1. Each e(i)ae(i)
looks roughly like an infinite block matrix, indexed by the pieces from
X (i). The hypothesis that a is quasi-local (finite ǫ-propogation) gives
1In fact, Engel proved the result for quasi-local operators that decay sufficiently
on any Riemannian manifold with bounded geometry and polynomial volume
growth. For groups, polynomial growth implies virtual nilpotency ([11]), which
in turn implies finite asymptotic dimension ([1, Corollary 68]). To our knowledge,
it is not known whether polynomial volume growth implies finite asymptotic di-
mension (or even (straight) finite decomposition complexity) outside of the case of
groups.
4
J ´AN SPAKULA AND AARON TIKUISIS
us a lot of control over the norm of the non-diagonal entries of this ma-
trix, and a conditional expectation argument allows us to conclude that
e(i)ae(i) is not far away from its "restriction" to the diagonal (provided
that the pieces in X (i) are sufficiently well separated), see Corollary
4.3. Since the pieces of the X (i) are uniformly bounded, the operator
we get by expecting onto the diagonal has genuinely finite propogation.
An algebraic trick allows us to view the asymmetric pieces e(i)ae(j) as
matrices in a similar way, so that we can likewise approximate each of
them by finite propogation operators. In this way, we approximate a
as a sum of four operators with finite propogation.
Outline. In Section 2 we introduce our general setup, with the Roe
algebra, the uniform Roe algebra, and the uniform algebra as exam-
ples. We then state the main result, Theorem 2.8, in the language of our
general setup. We give some background on asymptotic dimension and
(straight) finite decomposition complexity in Section 2.1. The equiva-
lence between quasi-locality and the relative commutant-type property
is fairly straightforward, and laid out in Section 3. We use a more tech-
nical formulation of quasi-locality as a stepping stone towards proving
that it implies being in the Roe algebra (assuming straight finite de-
composition complexity), a proof that is carried out in Section 4. In
Section 5, we prove that the relative commutant-type property is equiv-
alent to essential commutation with Higson functions. The final sec-
tion, Section 6, is concerned with the commutative (but non-separable)
C*-algebra VL∞(X) that arises in our relative commutant-type prop-
erty, looking at how well it determines X (up to coarse equivalence),
and at its nuclear dimension (roughly, the covering dimension of its
spectrum).
Acknowledgments. AT was supported by EPSRC EP/N00874X/1.
JS was supported by Marie Curie FP7-PEOPLE-2013-CIG Coarse Anal-
ysis (631945). We would like to thank Ulrich Bunke, Alexander Engel,
John Roe, Thomas Weighill, Stuart White, and Rufus Willett for com-
ments and discussion relating to this piece.
2. Definitions and the main result
Let A be a C*-algebra. We denote by A1 the closed unit ball of A.
For a, b ∈ A and ǫ > 0, we write a ≈ǫ b to mean ka − bk ≤ ǫ. Define
A∞ := l∞(N, A)/{(an)∞
n=1 ∈ l∞(N, A) : lim
n→∞
kank = 0},
which is a C*-algebra.
We now set up a general situation to which our main result applies,
encompassing both Roe algebras and uniform Roe algebras, as well as
uniform algebras (see Example 2.5). Subsequently, we will state our
main result in its full generality (Theorem 2.8)
RELATIVE COMMUTANT PICTURES OF ROE ALGEBRAS
5
Definition 2.1. Let X be a proper metric space. By an X-module,
we mean a Hilbert space H and an injective unital ∗-homomorphism
Cb(X) → B(H), which is strictly continuous when viewing Cb(X) and
B(H) as multiplier algebras of C0(X) and K(H) respectively. We shall
suppress the ∗-homomorphism Cb(X) → B(H), and treat Cb(X) as a
C*-subalgebra of B(H).
For R ≥ 0, an operator a ∈ B(H) has propogation at most R if
for any f, f ′ ∈ Cb(X), if the supports of f and f ′ are R-disjoint then
f af ′ = 0. For R ≥ 0 and ǫ > 0, an operator a ∈ B(H) has ǫ-
propogation at most R if for any f, f ′ ∈ Cb(X)1, if the supports of
f and f ′ are R-disjoint then kf af ′k < ǫ. An operator a ∈ B(H) is
quasi-local if for every ǫ > 0, it has finite ǫ-propogation.
Definition 2.2. Let X be a proper metric space and let H be an X-
module. Given an equicontinuous family (ej)j∈J of positive contractions
in Cb(X) with pairwise disjoint supports, define the block cutdown map
θ(ej )j∈J : B(H) → B(H) by
θ(ej )j∈J (a) :=Xj∈J
ejaej
(using disjointness of the supports and the fact that the family is con-
tractive, the right-hand sum converges in the strong operator topology).
Let B ⊆ B(H) be a C*-subalgebra such that Cb(X)BCb(X) = B. B is
closed under block cutdowns if θ(ej )j∈J (B) ⊆ B for every equicontinuous
family (ej)j∈J of positive contractions from Cb(X) with pairwise disjoint
supports.
For an equicontinuous family (ej)j∈J of positive contractions from
Cb(X) with pairwise disjoint supports, the block cutdown map θ(ej )j∈J
defined above is evidently completely positive and contractive (c.p.c.).
Note that multiplication by Cb(X) commutes with block cutdowns:
f θ(ej )j∈J (a) = θ(ej )j∈J (f a)
and θ(ej )j∈J (a)f = θ(ej )j∈J (af )
for f ∈ Cb(X) and a ∈ B(H). Also note that
(2.1)
kθ(ej )j∈J (a)k = sup
j∈J
kejaejk.
Note that, if (ej)j∈J is an equicontinuous family of positive contrac-
tions from Cb(X) with uniformly bounded, pairwise disjoint supports,
then θ(ej )j∈J (a) has finite propogation, for every a ∈ B(H).
Definition 2.3. Let X be a proper metric space, H an X-module, and
let B ⊆ B(H) be a C*-subalgebra such that Cb(X)BCb(X) = B, and
which is closed under block cutdowns. Define
(i) Roe(X, B) := {b ∈ B : b has finite propogation})
(ii) K(X, B) := C0(X)BC0(X).
k·k
, and
6
J ´AN SPAKULA AND AARON TIKUISIS
If, in addition, we have
(2.2)
[C0(X), B] ⊆ K(X, B),
we shall call Roe(X, B) a Roe-like algebra of X.
Remark 2.4. The condition (2.2) implies that K(X, B) is an ideal in
Roe(X, B).
It is automatically satisfied in all the examples below,
where in fact C0(X)B ⊆ K(X, B) (and K(X, B) turns out to be the
ideal of compact operators). Finally, it is not needed for the substantial
part of this piece, so we shall explicitly refer to it when needed.
Example 2.5. Let X be a uniformly discrete proper metric space. Let
H′ be an infinite dimensional, separable Hilbert space. Set Hu :=
l2(X) and H := l2(X, H′); Cb(X) acts on both of these by pointwise
multiplication, making them X-modules.
(i) With Bu := B(Hu), we see that Cb(X)BuCb(X) = Bu, and Bu is
closed under block cutdowns. In this case, Roe(X, Bu) = C ∗
u(X), the
uniform Roe algebra, and K(X, Bu) = K(Hu). Since C0(X) ⊆ K(Hu),
it follows that C0(X)Bu ⊆ K(Hu) = K(X, Bu).
(ii) Set B equal to the set of all b ∈ B(H) which are locally compact,
in the sense that for every f ∈ C0(X),
f b, bf ∈ K(H).
We see that Cb(X)BCb(X) = B, and B is closed under block cutdowns.
Then Roe(X, B) = C ∗(X), the Roe algebra, and K(X, B) = K(H).
(iii) Assume that X has bounded geometry. Set B0 equal to the
closure of the set of all b = (bx,y)x,y∈X ∈ B(H) for which the rank of
bx,y ∈ B(H′) is uniformly bounded. When b = (bx,y)x,y∈X ∈ B(H) has
entries with rank bounded by k, then so does any block cutdown map
applied to b. Since each block cutdown map is continuous, it follows
that B0 is closed under block cutdowns. Continuity of multiplication
ensures that
Cb(X)B0Cb(X) = B0.
When X has bounded geometry, then Roe(X, B0) = UC ∗(X), the uni-
form algebra of X, defined as the closure of finite propogation operators
b = (bx,y)x,y∈X ∈ B(H) for which the rank of bx,y is uniformly bounded.
To see this, it is clear that Roe(X, B0) contains UC ∗(X). To show
Roe(X, B0) ⊆ UC ∗(X), it suffices to check that every finite propoga-
tion operator a ∈ B0 is contained in UC ∗(X). For such a, say its
propogation is less than R > 0. Set
K := sup
x∈X
BR(x),
RELATIVE COMMUTANT PICTURES OF ROE ALGEBRAS
7
which is finite due to the hypothesis of bounded geometry. Define
ER : B(H) → B(H) by ER ((bx,y)x,y∈X) := (cx,y)x,y∈X where
cx,y :=(bx,y,
0,
d(x, y) < R;
d(x, y) ≥ R.
Note that kER ((bx,y)x,y∈X)k ≤ K k(bx,y)x,y∈Xk (this is a straightfor-
ward argument, see e.g., the proof of [28, Lemma 8.1]), so that in
particular, EK is continuous. Also note that EK(a) = a. Since a ∈ B0,
it is a limit of a sequence of operators bn = (bn
x,y)x,y∈X such that for each
n, there exists Kn bounding the rank of bn
x,y over all x, y ∈ X. Thus
the same bound Kn applies to EK(bn) so that EK(bn) ∈ UC ∗(X). By
continuity of EK, a = limn→∞ EK(bn) ∈ UC ∗(X).
In this example, we also have K(X, B0) = K(H), and since B0 ⊆ B
(from (ii)), C0(X)B ⊆ K(H) = K(X, B0)
(iv) Generalizing (ii), let X be any proper metric space and let H be
an adequate X-module in the sense of [22, Definition 3.4]. Recall that
an operator b ∈ B(H) is locally compact if C0(X)b, bC0(X) ⊆ K(H).
Set B equal to the set of all locally compact, bounded operators. One
can easily see that Cb(X)BCb(X) = B; it is also true that B is closed
under block cutdowns.
To see this, let b ∈ B(H) be locally compact with kbk ≤ 1, let
(ej)j∈J be an equicontinuous family of positive contractions in Cb(X)
with pairwise disjoint supports, and set b′ := θ(ej )j∈J (b), which we must
prove is locally compact. As K(H) is closed, it suffices to show that for
any f ∈ Cc(X) with kf k ≤ 1, f b′, b′f ∈ K(H). Given ǫ > 0, note that
b′ ≈2ǫ θ((ej −ǫ)+)j∈J (b),
where (ej − ǫ)+ ∈ Cb(X) is given by (ej − ǫ)+(x) := max{ej(x) − ǫ, 0}.
By equicontinuity and pairwise disjointness of the family (ej), we may
choose δ such that if d(x, y) < δ and j 6= j′, then at most one of ej(x)
or ej ′(y) can be nonzero. Thus if f ∈ Cc(X), then by compactness of
its support, there are only finitely many j for which f (ej − ǫ)+ 6= 0.
Consequently,
f b′ ≈2ǫ f θ((ej −ǫ)+)j∈J (b)
(ej − ǫ)+b(ej − ǫ)+
f (ej − ǫ)+b(ej − ǫ)+,
= fXj∈J
=Xj
and as this is a finite sum of elements of K(H), it is itself in K(H). As
K(H) is closed and ǫ > 0 is arbitrary, it follows that f b′ ∈ K(H). Like-
wise, b′f ∈ K(H), establishing that b′ is locally compact, and therefore
that B is closed under block cutdowns.
8
J ´AN SPAKULA AND AARON TIKUISIS
In this example, we get Roe(X, B) = C ∗(X), the Roe algebra, and
K(X, B) = K(H) = C0(X)B.
Definition 2.6. Let X be a metric space. A bounded sequence (fn)∞
n=1
from Cb (X) is very Lipschitz if, for every L > 0, there exists n0 such
that fn is L-Lipschitz for all n ≥ n0. Let VL (X) denote the set of all
very Lipschitz bounded sequences from Cb (X). Define
VL∞ (X) := VL (X) /{(fn)∞
n=1 ∈ VL (X) lim
n→∞
kfnk = 0}.
VL (X) is a C∗-subalgebra of l∞ (N, Cb (X)),2 and therefore the quo-
tient VL∞ (X) is a C*-subalgebra of (Cb(X))∞.
E.g., if X is a finitely generated group G with the word metric, then
VL∞ (X) can be identified with the fixed point algebra of l∞ (G)∞
under the action of G induced by left translation on l∞ (G).
Recall the following definition from [23].
Definition 2.7. Let X be a proper metric space. A function g ∈ Cb(X)
is a Higson function (also called a slowly oscillating function) if, for
every R > 0 and ǫ > 0, there exists a compact set A ⊆ X such that
for x, y ∈ X\A, if d(x, y) < R then g(x) − g(y) < ǫ. The set of all
Higson functions on X is denoted Ch(X).
E.g., if X is a finitely generated group G with the word metric,
then Ch(X) ⊆ l∞(X) is the preimage of the fixed point algebra of
l∞ (G) /c0(G) under the action of G induced by left translation on
l∞ (G).
In the following, H is an X-module, and we view both VL∞ (X)
and B ⊆ B (H) as C*-subalgebras of B (H)∞, and consider the relative
commutant
B ∩ VL∞ (X)′ .
It is easy to see (at least in the standard cases of Example 2.5) that any
finite propogation operator commutes with VL∞ (X), and by taking
limits it follows that
Roe(X, B) ⊆ B ∩ VL∞ (X)′ .
The main result is as follows. Recall that straight finite decomposi-
tion complexity, as introduced in [4], is a weakening of finite asymptotic
dimension ([14, Theorem 4.1]). Both properties are defined in the fol-
lowing subsection.
2To check that the product of two very Lipschitz sequences is itself very Lipschitz,
use the fact that if f, g are bounded functions, such that f is L-Lipschitz and g is
L′-Lipschitz, then f g is (kf kL′ + kgkL)-Lipschitz.
RELATIVE COMMUTANT PICTURES OF ROE ALGEBRAS
9
Theorem 2.8. Let X be a proper metric space, H an X-module, and
let B ⊆ B(H) be a C*-subalgebra such that Cb(X)BCb(X) = B, which
is closed under block cutdowns, and such that (2.2) holds. For b ∈ B,
the following are equivalent.
(i) [b, f ] = 0 for all f ∈ VL∞(X);
(ii) b is quasi-local (it has finite ǫ-propogation for every ǫ > 0);
(iii) [b, g] ∈ K(X, B) (i.e., b essentially commutes with g) for all
g ∈ Ch(X).
If X has straight finite decomposition complexity, then these are also
equivalent to
(iv) b ∈ Roe(X, B).
The equivalence of (i) and (ii) is fairly straightforward, and the equiv-
alence of these conditions with (iii) (at least in the standard cases of
Example 2.5) seems to be known by coarse geometers; we shall provide
a detailed proof for completeness. The implication (iv) =⇒ (ii) is
straightforward and holds in complete generality.
The implication (i) ⇒ (iv) was proven by Lange and Rabinovich for
the uniform Roe algebra of Zd (i.e., the case X = Zd, H = l2(X), and
B = B(H) as in Example 2.5 (i)) in [17] (see [19, Proposition 8] for a
proof in English).
The result (ii) ⇒ (iv) was claimed by Roe in a remark on page 20 of
[22] under a "finite dimensionality" assumption, but it was later found
that his supposed proof was incomplete ([20]). The present paper is to
the authors' knowledge the first complete proof of a more general case
(which is even more general than finite asymptotic dimension).
Question 2.9. Is there a uniformly discrete countable metric space
with bounded geometry, for which (i)-(iii) does not imply (iv) of Theo-
rem 2.8?
2.1. Coarse geometric notions. We collect some terminology from
[13, 14, 4].
Definition 2.10. Let X be a proper metric space, let Z, Z ′ ⊆ X, let X
and Y be metric families (i.e. at most countable sets of subsets of X),
and finally let R ≥ 0.
• We shall say that X is uniformly bounded, if supY ∈X diam(Y ) <
∞.
• We shall denote the metric neighbourhood of Z of radius R by
NR(Z) := {z ∈ X d(z, Z) ≤ R}. We further set
NR(X ) := {NR(Y ) : Y ∈ X }.
• The distance between Z and Z ′ is d(Z, Z ′) := inf{d(z, z′) : z ∈
Z, z′ ∈ Z ′}.
10
J ´AN SPAKULA AND AARON TIKUISIS
• A family (Yj)j∈J is R-disjoint if d(Yj, Yj ′) > R for all j 6= j′;
we write
Yj
GR-disjoint
for the union of the Yj to indicate that the family is R-disjoint.
• We say that Z R-decomposes over Y, if we can decompose
Z = X0 ∪ X1 and
Xi = GR-disjoint
Xij,
i = 0, 1,
such that Xij ∈ Y for all i, j.
• We say that X R-decomposes over Y, denoted X R−→ Y, if
every Y ∈ X R-decomposes over Y.
• We say that X has asymptotic dimension at most n, if for
every r ≥ 0, we can decompose X = X0 ∪ · · · ∪ Xn and
Xi = Gr-disjoint
Xij,
i = 0, . . . , n,
such that the metric family {Xij i, j} is uniformly bounded.
• We say that X has straight finite decomposition complexity,
if for any sequence 0 ≤ R1 < R2 < · · · , there exists m ∈ N and
Ri−−→ Xi
metric families {X} = X0, X1, . . . , Xm, such that Xi−1
for i = 1, . . . , m, and the family Xm is uniformly bounded.
The notion of straight finite decomposition complexity (sFDC) [4]
is apriori weaker than the original notion of finite decomposition com-
plexity of Guentner, Tessera and Yu [13, 14], see [4, Proposition 2.3].
The definition of finite decomposition complexity uses a certain "de-
composition game", which effectively means that the choices of Ri can
depend on the previous decompositions X1, . . . , Xi−1.
Already finite decomposition complexity is weaker than finite asymp-
totic dimension ([14, Theorem 4.1]).
3. Proof of (i) ⇔ (ii)
To prove the main result, we begin with a technical-looking charac-
terization of condition (ii).
Lemma 3.1. Let X be a proper metric space, let H be an X-module,
and let a ∈ B(H). Then k[a, f ]k < ǫ for every f ∈ VL∞ (X)1 if and
only if there exists L > 0 such that k[a, f ]k < ǫ whenever f ∈ Cb(X)1
is L-Lipschitz.
Remark 3.2. As we shall need to refer to the conclusion of the above
lemma later, we shall fix the following notation.
In the setup as in
the above lemma, we write a ∈ Commut(L, ǫ) if k[a, f ]k < ǫ whenever
f ∈ Cb(X)1 is L-Lipschitz.
RELATIVE COMMUTANT PICTURES OF ROE ALGEBRAS
11
Proof. The reverse implication is immediate from the definition of VL (X).
For the forward direction, we use a proof by contradiction. Suppose
for a contradiction that, for every n there exists fn ∈ Cb (X)1 that is
(1/n)-Lipchitz and k[a, fn]k ≥ ǫ.
Then evidently, (fn)∞
tradicts the hypothesis that k[a, VL∞ (X)1]k < ǫ.
n=1 ∈ VL (X) yet lim∞ k[a, fn]k ≥ ǫ. This con-
(cid:3)
Proof of Theorem 2.8 (i) ⇒ (ii). Suppose that [b, VL∞ (X)1] = 0 and
let ǫ > 0. By Lemma 3.1, let b ∈ Commut(L, ǫ) (in the notation of
Remark 3.2) for some L > 0.
We claim that b has ǫ-propogation at most L−1. Certainly, suppose
that f, f ′ ∈ Cb (X)1 have L−1-disjoint supports. We may define g ∈
Cb (X) such that gsuppf ≡ 1, gsuppf ′ ≡ 0 and g is L-Lipschitz. Hence,
k[b, g]k < ǫ. Consequently,
kf bf ′k = kf gbf ′k ≤ k[b, g]k + kf bgf ′k < ǫ + 0,
as required.
(cid:3)
Proof of Theorem 2.8 (ii) ⇒ (i). Suppose that b has finite ǫ-propogation
for all ǫ > 0. Assume that b is a contraction. We shall verify the con-
dition in Lemma 3.1. Therefore, let ǫ > 0 be given. Pick N such that
6/N < ǫ/2. By the hypothesis, let b have (ǫ/ (2N 2))-propogation at
most R > 0.
Let f ∈ Cb (X)1 be (2RN)−1-Lipschitz. We claim that k[b, f ]k < ǫ.
Surely, define sets
A1 := f −1(cid:0)[0, 1
N ](cid:1) , Ai := f −1(cid:0)(cid:0) i−1
N , i
N(cid:3)(cid:1) ,
These sets partition X and, for i − j > 1, Ai is (2R)-disjoint from
Aj. We may find a partition of unity e1, . . . , eN ∈ Cb(X) such that ei
is supported in NR/2(Ai). It follows that the supports of ei and ej are
R-disjoint when i − j > 1.
i = 2, . . . , N.
Thus,
(3.1)
Also,
(3.2)
keibejk < ǫ
2N 2 .
f ≈1/N
i
N ei
N
Xi=1
12
and so
k[f, b]k
(3.2)
≤ 2
J ´AN SPAKULA AND AARON TIKUISIS
N
= 2
i
N
N + k[
i
N ei, b]k
Xi=1
N +(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
N
N eib! N
Xi=1
Xj=1
N(cid:1) eibej(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
N +(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xi,j=1(cid:0) i
keibejk +(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xi−j≤1(cid:0) i
N + Xi−j>1
N − j
= 2
≤ 2
ej! − N
Xi=1
ei! N
Xj=1
j
N bej!(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
N − j
.
N(cid:1) eibej(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
ǫ
The terms of the first sum are each dominated by
2N 2 by (3.1), so
this entire sum is less than ǫ/2. The second sum can be broken into 4
sums with orthogonal terms (namely, note that when i = j, the terms
vanish; what remains is j = i + 1 and j = i − 1, and we break each of
these into even and odd parts). Each of the terms of the second sum
has norm at most 1/N; thus, we have
N + ǫ
k[f, b]k < 2
N < ǫ,
2 + 4
as required.
(cid:3)
4. Proof of (i) ⇒ (iv)
In this section, we prove that (i) ⇒ (iv) in Theorem 2.8. We begin
by establishing a few general functional analytic facts.
Recall that the strong* topology on B(H) is the one in which a net
α → a∗ in
(aα) converges to a ∈ B(H) if and only if both aα → a and a∗
the strong operator topology (i.e., kaαξ −aξk → 0 and ka∗
αξ −a∗ξk → 0
for every ξ ∈ H). A conditional expectation from C*-algebra A to a
C*-subalgebra B is a completely positive and contractive projection E
from A to B satisfying
for all b1, b2 ∈ B and a ∈ A.
E(b1ab2) = b1E(a)b2
Lemma 4.1. Let H be a Hilbert space and let G be a subgroup of the
group of unitary operators, which is compact in the strong∗ topology.
Then there is a unique conditional expectation EG : B(H) → G′ whose
restriction to the unit ball is weak operator topology continuous.
It
satisfies
(4.1)
kEG(a) − ak ≤ sup
u∈G
k[a, u]k ,
a ∈ B(H).
RELATIVE COMMUTANT PICTURES OF ROE ALGEBRAS
13
Proof. Let µG be the normalized Haar measure on G (under the strong∗
topology). Fix a ∈ B(H), and consider the map G → B(H) defined
by u 7→ u∗au. Then, with the strong∗ topology on the domain G and
the weak operator topology on the range B(H), this map is continuous.
We may therefore integrate, defining
EG(a) := WOT-ZG
u∗au dµG(u).
(Here, WOT-RG · dµG indicates the Pettis integral,
unique operator satisfying
i.e., EG(a) is the
hEG(a)ξ, ηi =ZG
hu∗auξ, ηi,
for all ξ, η ∈ H.) Using invariance of the Haar measure µG, one easily
sees that EG(a) commutes with all of G.
We now check (4.1); for this, set γ := supu∈G k[a, u]k. For η, ξ ∈ H,
h(EG(a) − a) η, ξi =(cid:12)(cid:12)(cid:12)(cid:12)
ZG
≤ZG
≤ZG
=ZG
≤ γ kηk kξk .
h(u∗au − a) η, ξi dµG(u)(cid:12)(cid:12)(cid:12)(cid:12)
h(u∗au − a) η, ξi dµG(u)
ku∗au − ak kηk kξk dµG(u)
k[u, a]k kηk kξk dµG(u)
Therefore, (4.1) follows.
In particular, we conclude that if a ∈ G′ then EG(a) = a. It is also
straightforward to see that the function EG is c.p.c., and therefore it is
a conditional expectation.
On the unit ball of B(H), the integral defining EG can be uniformly
approximated in the weak operator topology by (finite) Riemann sums,
which themselves are continuous in the weak operator topology.
It
follows that the restriction of EG to the unit ball is continuous using
the weak operator topology.
If E : B(H) → G′ is another conditional expectation whose restric-
tion to the unit ball is weak operator topology continuous, then for a
14
J ´AN SPAKULA AND AARON TIKUISIS
contraction a ∈ B(H),
EG(a) = E (EG(a))
(E fixes G′)
u∗au dµG(u)(cid:17)
E(u∗au) dµG(u)
u∗E(a)u dµG(u)
= E(cid:16)WOT-ZG
= WOT-ZG
= WOT-ZG
= EG(E(a))
= E(a)
(WOT-continuity of EB(H)1)
(E is a conditional expectation)
(EG fixes G′).
Thus, E = EG.
(cid:3)
Recall that an atomic abelian von Neumann algebra is a von Neu-
mann algebra isomorphic to l∞(X), for some set X. In the following,
when H = l2(X), then the conditional expectation B(l2(X)) → l∞(X)
consists simply of taking an operator to its diagonal.
Corollary 4.2. Let D ⊂ B(H) be an atomic abelian von Neumann al-
gebra. Then there is a unique conditional expectation ED : B(H) → D′
whose restriction to the unit ball is weak operator topology continuous.
It satisfies
(4.2)
kED(a) − ak ≤ sup
k[a, x]k ,
a ∈ B(H).
x∈D,kxk≤1
Proof. Without loss of generality, D contains the identity operator. D
is generated by a family of orthogonal projections (pj)j∈J, whose sum
converges strongly to 1. Define
G :=(Xj∈J
(−1)αj pj : (αj)j∈J ∈ (Z/2)J) .
This is a strong∗ compact subgroup of the unitary group of D (it is
homeomorphic to (Z/2)J with the product topology), so that Lemma
4.1 applies to it.
It is clear that it generates D as a von Neumann
algebra, so that G′ = D′. The conclusion follows from Lemma 4.1. (cid:3)
Corollary 4.3. Let X be a proper metric space, let H an X-module,
and let a ∈ B (H). Suppose a ∈ Commut(L, ǫ) for some L, ǫ > 0 (in the
notation of Remark 3.2). Let (ej)j∈J be a family of positive contractions
from Cb(X) with (2L−1)-disjoint supports, and define e := Pj∈J ej.
Then, with θ(ej )j∈J from Definition 2.2,
keae − θ(ej )j∈J (a)k ≤ ǫ.
RELATIVE COMMUTANT PICTURES OF ROE ALGEBRAS
15
Proof. Set Aj equal to the support of ej for each j ∈ J. We may find
pairwise disjoint projections pj ∈ B(H), for j ∈ J, such that pj acts as
a unit on ej. Define D to be the von Neumann subalgebra generated
the unique conditional expectation provided by Corollary 4.2. Then
by {pj : j ∈ J} (with unit 1D = Pj pj), and let ED : B(H) → D′ be
one finds that for x ∈ B(H), ED(x) = Pj∈J pjxpj (converging in the
strong operator topology), and therefore
ED(eae) = θ(ej )j∈J (a).
Using (2L−1)-disjointness of the family (Aj)j∈J , for f ∈ D1, there
exists a function f ∈ Cb (X)1 that is L-Lipschitz such that
Therefore,
f = 1D f .
f (eae) = f eae
= e f ae
≈ǫ ea f e
= (eae)f.
Hence, by Corollary 4.2,
eae ≈ǫ ED(eae) = θ(ej )j∈J (a),
as required.
(cid:3)
Definition 4.4. Let X be a proper metric space, H be an X-module,
a ∈ B(H) and let X be a metric family (of subsets of X). We say that
a is block diagonal with respect to X , if there exists an equicontinuous
family (ej)j∈J of positive contractions in Cb(X) with pairwise disjoint
supports, such that a = θ(ej )j∈J (a), and the support of each ej is con-
tained in some set Yj ∈ X . Furthermore, in this case we shall denote
aYj := ejaej and call these operators blocks of a.
The next lemma sets up the "induction step" to be applied in the
context of the decomposition game, in the proof of Theorem 2.8 (i) ⇒
(iv).
Lemma 4.5. Let X be a proper metric space, and let Y be a metric
family such that {X} 4L−1+4
−−−−−→ Y for some L > 0. Let H be an X-
module, and let a ∈ B (H). Let ǫ > 0 be such that a ∈ Commut(L, ǫ).
Then we can write
a ≈8ǫ a00 + a01 + a10 + a11,
where each aii′ is of the form θ(fk)k∈K (gag′) for some contractions g, g′ ∈
Cb(X) and some family (fk)k∈K of 1-Lipschitz positive contractions
16
J ´AN SPAKULA AND AARON TIKUISIS
in Cb(X) with disjoint supports, such that the support of each fk is
contained in a set in NL−1+1(Y).
Proof. By the decomposition assumption, we can write
X = X (0) ∪ X (1), X (i) =
Gj∈Ji
(4L−1+4)-disjoint
X (i)
j
,
i = 0, 1,
j ∈ Y for each i, j. We may find a partition of unity consisting
j ∈ Cb(X), for i = 0, 1 and j ∈ Ji,
j ). It follows that
with X (i)
of 1-Lipschitz positive contractions e(i)
such that the support of e(i)
j
for each i, the supports of (e(i)
is contained in N1(X (i)
j )j∈Ji are (4L−1 + 2)-disjoint.
For each i = 0, 1, define e(i) :=Pj∈Ji
a = e(0)ae(0) + e(0)ae(1) + e(1)ae(0) + e(1)ae(1),
e(i)
j . Since
it suffices to find aii′ ≈ e(i)ae(i′) for each i, i′ ∈ {0, 1} with the required
properties. (We will be precise about the degree of approximation -- in
short, it depends on whether i and i′ are equal.)
For the case i = i′, Corollary 4.3 shows that
e(i)ae(i) ≈ǫ θ(e(i)
j )j∈Ji
(e(i)ae(i)) =: aii.
(cid:16)NL−1+1(X (i)
j )(cid:17)j∈Ji
The latter operator is clearly block diagonal with respect to N1(Y)
(hence also with respect to NL−1+1(Y)).
Turning now to the case i 6= i′, note that for fixed i, the family
is (2L−1 + 2)-disjoint. For each i, j, there exists
e(i)
j ∈ Cb (X) that is L-Lipschitz, that acts as the identity on e(i)
j , and
is supported on NL−1+1(X (i)
e(i)
j . We
have
j ). For each i, define e(i) :=Pj∈Ji
e(i)ae(i′) = e(i)e(i)ae(i′)e(i′)
≈2ǫ e(i)e(i′)ae(i)e(i′).
For each j ∈ Ji and j′ ∈ Ji′, there exists a 1-Lipschitz positive con-
traction fj,j ′ ∈ Cb(X) that is 1 on NL−1+1(X (i)
j ′ ), and
is supported on the metric neighbourhood of this set of radius 1. In
particular, the support of each fj,j ′ is contained in a set in NL−1+2(Y),
the family of supports of the family (fj,j ′)j,j ′∈Ji′ is (2L−1)-disjoint, and
j ) ∩ NL−1+1(X (i′)
f := Pj,j ′ fj,j ′ acts as an identity on (both sides of) e(i)e(i′)ae(i)e(i′).
Applying Corollary 4.3, we obtain
f e(i)e(i′)ae(i)e(i′)f ≈ǫ θ(fj,j′ )j∈Ji, j′ ∈Ji′ (cid:16)e(i)e(i′)ae(i)e(i′)(cid:17) .
RELATIVE COMMUTANT PICTURES OF ROE ALGEBRAS
17
Thus,
e(i)ae(i′) ≈2ǫ e(i)e(i′)ae(i)e(i′)
= f e(i)e(i′)ae(i)e(i′)f
≈ǫ θ(fj,j′ )j∈Ji, j′ ∈Ji′ (cid:16)e(i)e(i′)ae(i)e(i′)(cid:17) =: aii′.
By construction, it is clear that aii′ is block diagonal with respect to
NL−1+2(Y).
Summarizing, we have a ≈ǫ+3ǫ+3ǫ+ǫ a00 + a10 + a01 + a11, and all the
(cid:3)
aii′ are of the right form.
We now strengthen the previous lemma by allowing an arbitrary met-
ric family in place of {X} and with a a correspondingly block diagonal
operator.
Lemma 4.6. Let X be a proper metric space, and let X and Y be
metric families, such that X 4L−1+4
−−−−−→ Y for some L > 0. Let H be an
X-module, and let a ∈ B (H) be block diagonal with respect to X . Let
ǫ > 0 be such that a ∈ Commut(L, ǫ). Then we can write
(4.3)
a ≈8ǫ a00 + a01 + a10 + a11,
where each aii′ is of the form θ(fk)k∈K (gag′) for some contractions g, g′ ∈
Cb(X) and some equicontinuous family (fk)k∈K of positive contractions
in Cb(X) with disjoint supports, such that the support of each fk is
contained in a set in NL−1+1(Y). In particular:
(i) each aii′ is block diagonal with respect to NL−1+1(Y),
(ii) if a ∈ Commut(L′, ǫ′) for some L′, ǫ′ > 0, then each aii′ is in
Commut(L′, ǫ′) as well, and
(iii) if B ⊆ B(H) is a C*-subalgebra such that Cb(X)BCb(X) = B
and B is closed under block cutdowns, and if a is in B, then
each aii′ is in B as well.
Proof. Without loss of generality, both X and Y are closed under tak-
ing subsets. Start by letting (ej) be an equicontinuous family of pos-
itive contractions in Cb(X) with disjoint supports, such that Yj
:=
supp(ej) ∈ X . Applying Lemma 4.5 to each aYj yields
aYj ≈8ǫ aj
00 + aj
01 + aj
10 + aj
11,
satisfying the conclusions of that lemma. Set
aj
ii′.
aii′ :=Xj
Then (4.3) follows from (2.1)
To see that each aii′ has the right form, fix i and i′. For each j, there
j ∈ Cb(X) and a family (fj,k)k∈Kj of 1-Lipschitz
exist contractions gj, g′
18
J ´AN SPAKULA AND AARON TIKUISIS
positive contractions in Cb(Yj) with disjoint supports, such that
aj
ii′ = θ(fj,k)k∈Kj
(gjag′
j).
and the support of each fj,k is of the form Y ∩ Z for some Z ∈
NL−1+1(Y).
Then observe that
aii′ = θ(e1/4
j
fj,k)j∈J,k∈Kj
((Xj
e1/4
j gj)a(Xj
e1/4
j gj)),
where the family appearing in this block-cutdown formula, namely
(e1/4
j fj,k)j∈J,k∈Kj
is equicontinuous and contained in NL−1+1(Y).
(i)-(ii) are immediate consequences of the form that aii′ takes. Since
multiplication by Cb(X) preserves block structure, and using (2.1), (iii)
can also be seen to be a consequence of the form that aii′ takes.
(cid:3)
We have seen in the previous lemma that we will need to work with
"thickened" metric families, so we record the following straightforward
observation.
Lemma 4.7. Let X and Y be metric families, R, S ≥ 0. Assume that
X R−→ Y and that R − 2S ≥ 0. Then NS(X ) R−2S−−−−→ NS(Y).
Proof of Theorem 2.8 (i) ⇒ (iv). Recall that we are given an operator
b on an X-module H satisfying [b, f ] = 0 for all f ∈ VL∞(X). Given
ǫ > 0, our task is to produce a finite propogation operator in B which
is ǫ-far from b. Lemma 3.1 provides us with Ln for every
ǫn := ǫ/(2 · 8n),
such that b ∈ Commut(Ln, ǫn). Let
n + 1) + 2(L−1
Rn := 4(L−1
n−1 + 1) + · · · + 2(L−1
1 + 1).
As X has straight finite decomposition complexity (see Definition 2.10),
Rn−−→
we obtain metric families {X} = X0, X1, . . . , Xm, such that Xn−1
Xn for n ∈ {1, . . . , m} and Xm is uniformly bounded. Note that Lemma
4.7 gives us that
N(L−1
n−1+1)+···+(L−1
1 +1)(Xn−1)
n +1)
4(L−1
−−−−−−→ N(L−1
n +1)+···+(L−1
1 +1)(Xn).
Thus, we can inductively apply Lemma 4.6, with Ln, ǫn, the operators
obtained in the previous iteration, and metric families from the above
display. After m steps, we will have approximated the operator b by an
operator b′ which is a sum of finitely many (4m to be precise) operators
in B which are block diagonal with respect to
N(L−1
m +1)+···+(L−1
1 +1)(Xm).
RELATIVE COMMUTANT PICTURES OF ROE ALGEBRAS
19
Since Xm is uniformly bounded, so is the above family, and therefore
operators which are block diagonal with respect to it have finite pro-
pogation; consequently, b′ ∈ Roe(B, X). Tracing through the estimates
given by Lemma 4.6, we compute that the distance from b to b′ is at
most
8ǫ1 + 4 (8ǫ2 + 4 (8ǫ3 + 4 (. . . ))) = ǫ(cid:0) 1
This finishes the proof.
2 + 1
4 + 1
8 + . . .(cid:1) = ǫ.
(cid:3)
Remark 4.8. When the asymptotic dimension of X is at most d <
∞, the induction component of the above proof can be removed: one
can use the idea of Lemma 4.5 with a decomposition of X into d + 1
(instead of 2) uniformly bounded, (4L−1 + 4)-disjoint families, and
correspondingly approximate a be a sum of (d + 1)2 block diagonal
operators.
5. Higson functions
To prepare to prove (i) ⇔ (iii) of Theorem 2.8, we begin by consid-
ering a special class of Higson functions which are more closely related
to our definition of VL∞(X).
Definition 5.1. Let X be a proper metric space. A function g ∈
Cb(X) is a Lipschitz -- Higson function if, for every L > 0, there exists
a compact set A ⊆ X such that gX\A is L-Lipschitz. The set of all
Lipschitz -- Higson functions on X is denoted Clh(X).
Fix a proper metric space X and a point x0 ∈ X. For R > 0, define
eR ∈ C0(X) by
eR(x) := max{0, 1 − d(x, BR(x0))/R}.
(5.1)
Observe that eR is R−1-Lipschitz, is 1 on ¯BR(x0), and vanishes outside
of B2R(x0).
Lemma 5.2. Ch(X) = Clh(X) + C0(X).
Proof. The inclusion ⊇ is straightforward. To go the other direction,
let g ∈ Ch(X). We shall produce f ∈ Clh(X) such that f − g ∈ C0(X).
Without loss of generality, g is a positive contraction. Fix a point
x0 ∈ X.
Recursively define R0 := 0 and Rn ≥ max{2(n+1), 2Rn−1} such that
2(n+1) .
if x, y ∈ X\BRn(x0) and d(x, y) < n + 1 then g(x) − g(y) < 1
Using eR from (5.1), set
g1 := eR1g,
gn :=(cid:0)eRn − eRn−1(cid:1) g, n ≥ 2.
20
J ´AN SPAKULA AND AARON TIKUISIS
Note that, for n ≥ 2, gn is such that if d(x, y) < n then gn(x)−gn(y) <
1
n , and that
∞
(5.2)
g =
gn,
Xn=1
converging pointwise (as at each point, at most two terms of the sum
are nonzero).
Define f1 := g1.
Fix n ≥ 2; we shall define a function fn which approximates gn, but
is more Lipschitz. Define
i = 1, . . . , n,
i = 1, . . . , n,
and define ci ∈ Cb(X) by
Ai := g−1
n (cid:2) i
n, 1(cid:3) ,
, 0o ,
ci(x) := maxn1 − d(x,Ai)
Xi=1
fn :=
1
n
ci,
n
which is (1/n)-Lipschitz. Set
n
which is also (1/n)-Lipschitz, as it is an average of such. Moreover,
kfn − gnk ≤ 1
n , and the support of fn is contained in the support of gn.
n=1 fn; as in (5.2), at each point, at most two terms of
fn is 2n−1
0 -
Lipschitz, for all n0 ≥ 2. Moreover, f agrees with this tail sum outside
of B2Rn0−1, which proves that f ∈ Clh(X).
Set f := P∞
the sum are nonzero. Using this fact, one sees that P∞
Similarly, since f − g agrees with the tail P∞
of B2Rn0−1, and this tail has norm at most 2
C0(X).
(fn − gn) outside
n , it follows that f − g ∈
(cid:3)
n=n0
n=n0
Remark 5.3. We note in passing that, due to the previous lemma, the
Higson corona νX (defined as the compact Hausdorff space satisfying
C(νX) ∼= Ch(X)/C0(X)) satisfies
C(νX) ∼= Clh(X)/C0(X).
Now we set out two constructions to be used, producing a Lipschitz --
Higson function from a very Lipschitz sequence and vice versa. Neither
construction is canonical: both depend on a number of choices.
Let (fk)∞
k=1 ∈ VL(X), let (fki)∞
i=0 ⊂
(0, ∞) be a sequence such that Ri+1 ≥ 6Ri for each i. From these, and
using eR from (5.1), define
i=1 be a subsequence, and let (Ri)∞
(5.3)
gx0,(fki )∞
i=1,(Ri)∞
i=0 :=
fki(eRi − e3Ri−1) ∈ Cb(X).
∞
Xi=1
RELATIVE COMMUTANT PICTURES OF ROE ALGEBRAS
21
Note that the supports of the functions in the summation are pair-
wise disjoint, so we can treat the series as converging pointwise. It is
straightforward to see that kgx0,(fki )∞
Lemma 5.4. Fix a proper metric space X and a point x0 ∈ X. For
(fk)∞
i=0 ⊂
be as defined
(0, ∞) such that Ri+1 ≥ 6Ri for each i, let gx0,(fki )∞
in (5.3). Then
k=1 ∈ VL(X), a subsequence (fki)∞
i=1, and a sequence (Ri)∞
k ≤ k(fk)∞
i=1,(Ri)∞
i=0
i=1,(Ri)∞
i=0
k=1k.
gx0,(fki )∞
i=1,(Ri)∞
i=0 ∈ Clh(X).
Proof. Without loss of generality, we may assume kfkk ≤ 1 for all k.
For ease of notation, we set
g := gx0,(fki )∞
i=1,(Ri)∞
i=0
.
As in the definition of a Lipschitz -- Higson function, let L > 0 be
given. Pick i0 such that Ri0−1 > 2/L and such that fi is (L/2)-Lipschitz
for all i ≥ i0. We will be done when we show that the restriction of g
to X\BRi0
(x0) is L-Lipschitz.
For i ≥ i0, eRi − e3Ri−1 is (L/2)-Lipschitz, so that the product
hi := fki(eRi − e3Ri−1)
is L-Lipschitz. For x, y ∈ X\BRi0
(x0), note that, by the definition of
eR and the condition Ri+1 ≥ 6Ri, that at least one of the following
conditions holds.
(i) d(x, y) ≥ Ri0, or
(ii) There exists i such that g(x) = hi(x) and g(y) = hi(y).
In the first case,
g(x) − g(y) ≤ 2kgk = 2 ≤ 3LRi0−1 ≤ 3LRi0 ≤ Ld(x, y).
In the second case, since hi is L-Lipschitz, it follows that g(x) −
(cid:3)
g(y) ≤ Ld(x, y).
Next, let g ∈ Ch(X) be given, along with a sequence (Rk) ∈ (0, ∞)
such that limk→∞ Rk = ∞. From this data, define (using eR from (5.1))
(5.4) Fx0,g,(Rk)∞
Lemma 5.5. Fix a proper metric space X and a point x0 ∈ X. For
a Higson function g ∈ Clh(X) and a sequence (Rk) ∈ (0, ∞) such that
limk→∞ Rk = ∞, let Fx0,g,(Rk)∞
k=1 ∈ l∞(N, l∞(X)) where fk := (1 − eRk )g.
be as defined in (5.4). Then
:= (fk)∞
k=1
k=1
Fx0,g,(Rk)∞
k=1
∈ VL(X).
Proof. Without loss of generality, we may assume kgk ≤ 1. As in the
definition of VL(X), let L > 0 be given. Pick M ≥ 2/L such that g
is (L/2)-Lipschitz on X\BM (x0). Pick k0 such that Rk ≥ M for all
k ≥ k0. For k ≥ k0, using the fact that (1 − eRk ) is (L/2)-Lipschitz
and vanishes on BM (x0), it is easy to see that fk = g(1 − eRk ) is
L-Lipschitz.
(cid:3)
22
J ´AN SPAKULA AND AARON TIKUISIS
Proof of Theorem 2.8 (i) ⇒ (iii). By Lemma 5.2, (iii) is equivalent to
[g, b] ∈ K(X, B) for all g ∈ Clh(X), which is the statement we will prove
assuming (i). Assume that [b, g] 6∈ K(X, B) for some g ∈ Clh(X). Set
ǫ := k[b, g]+K(X, B)k. Fix a point x0 ∈ X, let (Rk)∞
i=1 ⊂ (0, ∞) be any
sequence such that limi→∞ Ri = ∞, and define (fk)∞
k=1 := Fx0,g,(Rk)∞
as in (5.4), that is,
k=1
fk := (1 − eRk )g.
By Lemma 5.5, (fk)∞
the condition (2.2) we obtain
k=1 ∈ VL(X). Since eRk ∈ C0(X) for each k, using
[fk, b] = [g, b] − [eRk g, b] ∈ [g, b] + K(X, B)
and therefore by the definition of ǫ,
k[fk, b]k ≥ ǫ.
Consequently, limk→∞ k[fk, b]k ≥ ǫ, so that b does not commute with
the image of (fk)∞
(cid:3)
k=1 in VL∞(X).
Before we embark on the proof of Theorem 2.8 (iii) =⇒ (i), note
that as a consequence of the Spectral Theorem, we may extend the
X-module structure on B(H) from bounded continuous functions to
bounded Borel functions on X. This is convenient in the following
proof, since it allows us to easily "cut up" operators on H using charac-
teristic functions of Borel sets in X. (Of course, we cannot assume the
algebra B is closed under multiplication by these bounded Borel func-
tions.) We opt for this approach for the sake of readability, although
it is possible to modify the proof to only use continuous functions for
the price of more approximations.
Proof of Theorem 2.8 (iii) ⇒ (i). Since each of VL∞(X), Ch(X), and
K(X, B) is ∗-closed, it suffices to prove the theorem in the case that b
is self-adjoint. We henceforth assume that b = b∗. Fix a point x0 ∈ X,
and to shorten notation in this proof, set
BR := BR(x0).
For each R > 0, we will use χBR to denote the support projection of a
function whose cozero set is BR.
Assume that [b, f ] 6= 0 for some f ∈ VL∞(X). Then in fact, [b, f ] 6= 0
k=1 for which each fk is a self-adjoint contraction; we
for some f = (fk)∞
fix this sequence. Let 0 < ǫ < k[b, f ]k.
Consider now two cases.
Case 1. There exists R0 > 0 such that for all S > 0 there are
infinitely many k for which
kχBR0
[b, fk](1 − χBS )k >
ǫ
5
.
RELATIVE COMMUTANT PICTURES OF ROE ALGEBRAS
23
Roughly, for this case, we will construct some g (of the form (5.3))
such that the block-column of [b, g] corresponding to BR0 doesn't con-
verge to 0 at ∞.
Note that as k → ∞, fkBR0
tends towards being constant; so without
is constant. Adding a
loss of generality, we may assume that fkBR0
scalar to each fk, we arrive at another sequence(cid:16) fk(cid:17)∞
with the same
n=1 (that is, self-adjoint and satisfying the Case 1
[b, fk] =
b fk for all n, so that we have: for all S > 0 there exist infinitely
properties as (fk)∞
condition), such that fkBR0
χBR0
many k such that
≡ 0. From this it follows that χBR0
n=1
kχBR0
b fk(1 − χBS )k >
ǫ
5
.
Using R0 as above and k0 := 0, recursively choose k1 < k2 < · · ·
and R1 ≥ 6R0, R2 ≥ 6R1, . . . as follows. Having chosen Ri−1, pick
ki > ki−1 such that
kχBR0
a fki(1 − χB6Ri−1
)k >
ǫ
5
and k fkiχB6Ri−1
k ≤
ǫ
10kbk
.
(The second inequality can be arranged as fk converge to 0 on any
given bounded subset of X.) Then, since (eR)∞
R=1 converges strongly
to 1, pick Ri ≥ 6Ri−1 such that
kχBR0
b fki(1 − χB6Ri−1
)eRik >
ǫ
5
.
Since eRi − e3Ri−1 differs from (1 − χB6Ri−1
that
)eRi only on B6Ri−1, it follows
kχBR0
ǫ
10
Using these recursive choices, define g := gx0,( fki )∞
b fki(eRi − e3Ri−1)k >
∈ Clh(X) ⊆
Ch(X) by Lemma 5.4. If [b, g] ∈ K(X, B), then k[b, g](1 − χBS )k → 0
as S → ∞. However, given S ≥ R0, there exists i such that 3Ri−1 > S.
Then
i=1,(Ri)∞
i=0
.
k[b, g](1 − χBS )k ≥ kχBR0
= kχBR0
>
,
ǫ
10
[b, g](1 − χBS )(χB2Ri
b fki(eRi − e3Ri−1)k
− χB3Ri−1
)k
which is a contradiction. This concludes the proof in Case 1.
Case 2. For every R > 0, there exists S > 0 such that, for all but
finitely many k ∈ N,
kχBR[b, fk](1 − χBS )k ≤
ǫ
5
.
Without loss of generality, we may assume that S > R.
24
J ´AN SPAKULA AND AARON TIKUISIS
Roughly, for this case, we will construct some g (of the form (5.3))
such that the certain blocks on the diagonal of [b, g] don't converge to
0 at ∞.
In preparation for this, suppose we are given R > 0 and K ∈ N. Let
S be given by the Case 2 property. Then there exists k ≥ K such that
kχBR[b, fk](1 − χBS )k ≤
ǫ
5
,
and in addition, k[b, fk]k > ǫ and fkBS is ǫ
5-approximately constant.
From the latter property, it follows that there is a scalar γ such that
fkBS ≈ǫ/10 γ, so that
(5.5)
kχBS [b, fk]χBS k ≤ 2 ·
ǫ
10
kbk ≤
ǫ
5
.
Since b and fk are self-adjoint,
(5.6)
k(1 − χBS )[b, fk]χBR)k = kχBR[b, fk](1 − χBS )k ≤
ǫ
5
.
We now cut up the operator T = [b, fk] as follows:
❅
❅
BS
❅
BR
❅
❅
❅
BR
❅
BS
❅
χBRT χBS
χBRT (1 − χBS )
(χBS − χBR)T χBR
(1 − χBS )T χBR
(1 − χBR)T (1 − χBR)
That is, we use the equality
T = (1 − χBR)T (1 − χBR) + χBRT χBS + χBRT (1 − χBS )
+ (χBS − χBR)T χBR + (1 − χBS )T χBR
RELATIVE COMMUTANT PICTURES OF ROE ALGEBRAS
25
and the reverse triangle inequality to deduce
k(1 − χBR)[b, fk](1 − χBR)k
≥
(5.5),(5.6)
≥
>
(5.7)
=
k[b, fk]k −(cid:0)kχBR[b, fk]χBS k + kχBR[b, fk](1 − χBS )k
+k(χBS − χBR)[b, fk]χBRk + k(1 − χBS )[b, fk]χBRk(cid:1)
k[b, fk]k − 4 ·
ǫ
5
4ǫ
5
ǫ −
ǫ
5
.
In summary, we have shown that for every R > 0 and K ∈ N, there
exists k ≥ K such that
(5.8)
k(1 − χBR)[b, fk](1 − χBR)k >
ǫ
5
.
Now, start with R0 := 1 and k0 := 0, and (as in Case 1) we will
choose k1 < k2 < · · · and R1 ≥ 6R0, R2 ≥ 6R1, . . . recursively. Given
Ri−1, pick k > ki−1 satisfying (5.8) for R := 6Ri−1, and set ki equal to
this k. That is, ki > ki−1 satisfies
k(1 − χB6Ri−1
)[b, fki](1 − χB6Ri−1
)k >
ǫ
5
.
R=1 converges strongly to 1, there exists Ri ≥ 6Ri−1
Then, since (χBR)∞
such that
kχBRi
Note that χBRi
(1 − χB6Ri−1
(1 − χB6Ri−1
)[b, fki](1 − χB6Ri−1
.
) = χBRi
− χB6Ri−1
)χBRi
k >
ǫ
5
.
Using these recursive choices, define g := gx0,(fki )∞
∈ Clh(X) ⊆
Ch(X) by Lemma 5.4. If [b, g] ∈ K(X, B), then k[b, g](1 − χBS )k → 0
as S → ∞. However, given S > 0, there exists i such that 6Ri−1 > S.
Then
i=1,(Ri)∞
i=0
k[b, g] (1 − χBS )k ≥(cid:13)(cid:13)(cid:13)
=(cid:13)(cid:13)(cid:13)
>
ǫ
5
,
(χBRi
− χB6Ri−1
)[b, g](χBRi
− χB6Ri−1
(χBRi
− χB6Ri−1
)[b, fki](χBRi
− χB6Ri−1
)(cid:13)(cid:13)(cid:13)
)(cid:13)(cid:13)(cid:13)
again a contradiction. This concludes the proof.
(cid:3)
6. More about VL∞ (X)
6.1. To what extent does VL∞ (X) determine X? If X is a metric
space, then we say that a set E ⊆ X × X is uniformly bounded (also
called a metric entourage) if there exists R > 0 such that
E ⊆ {(x, y) ∈ X × X : d(x, y) ≤ R}.
26
J ´AN SPAKULA AND AARON TIKUISIS
Definition 6.1. Let X and Y be metric spaces. We say that a (not
necessarily continuous) function φ : X → Y is:
• bornologous, if for every R ≥ 0 there exists S ≥ 0, such that
for all x, y ∈ X, d(x, y) ≤ R implies d(φ(x), φ(y)) ≤ S;
• cobounded, if f −1(y) is bounded for every y ∈ Y ;
• coarse, if it is both cobounded and bornologous;
• a coarse equivalence, if it is bornologous, and there exists a
bornologous ψ : Y → X, such that both ψ ◦ φ and φ ◦ ψ are
uniformly close to the identity maps, i.e., their graphs are uni-
formly bounded subsets of X × X and Y × Y respectively. Note
that in this case both φ and ψ are automatically coarse;
• locally Lipschitz, if there exist δ > 0 and T ≥ 0, such that
d(x, y) ≤ δ implies d(f (x), f (y)) ≤ T d(x, y), x, y ∈ X.
• a Lip-coarse equivalence, if it is a coarse equivalence, it is lo-
cally Lipschitz, and in the definition of coarse equivalence, ψ
can be chosen to be locally Lipschitz as well.
Note that traditionally, coarse geometry does not concern itself with
local behaviour. However, as our main tool in this piece are Lipschitz
functions, we will insist that the maps involved are locally Lipschitz.
On the other hand, in the key setting in which the metric spaces in-
volved are uniformly discrete, this requirement is automatic, and thus
can be ignored.
Proposition 6.2. Let X be a proper metric space. VL∞ (X) is a
Lip-coarse invariant for X. More precisely, if X and Y are proper
metric spaces which are Lip-coarsely equivalent via a (locally Lipschitz)
map φ : X → Y , then composition by φ induces a ∗-isomorphism
VL∞ (Y ) → VL∞ (X).
Proof. Suppose that X, Y are Lip-coarsely equivalent, so that there are
locally Lipschitz coarse maps φ : X → Y and ψ : Y → X such that
the graphs of φ ◦ ψ and ψ ◦ φ are uniformly bounded. Denote by δ > 0
and T ≥ 0 the constants of local Lipschitzness of φ. As locally Lips-
chitz maps are continuous, φ and ψ induce maps φ∗ : l∞ (N, Cb(Y )) →
l∞ (N, Cb(X)) and ψ∗ : l∞ (N, Cb(X)) → l∞ (N, Cb(Y )).
First we show that φ∗ (VL (Y )) ⊆ VL (X). Surely, let (fn)∞
n=1 ∈
VL (Y ) with k(fn)∞
n=1k ≤ 1. Let L > 0. Since φ is a coarse map,
there exists S > 0 such that if x, y ∈ X satisfy d(x, y) ≤ 2/L then
d(φ(x), φ(y)) < S. Without loss of generality, we can assume that
S > δT . Since (fn)∞
n=1 ∈ VL(Y ), there exists n0 such that fn is (Lδ/S)-
Lipschitz for all n ≥ n0. For n ≥ n0, let us show that fn ◦ φ is
L-Lipschitz. Let x, y ∈ X, there are three cases:
• If d(x, y) > 2/L then fn(x) − fn(y) ≤ 2kfnk < Ld(x, y).
RELATIVE COMMUTANT PICTURES OF ROE ALGEBRAS
27
• If δ ≤ d(x, y) ≤ 2/L, then d(φ(x), φ(y)) < S. Since fn is
(Lδ/S)-Lipschitz,
kfn ◦ φ(x) − fn ◦ φ(y)k ≤ Lδ ≤ Ld(x, y).
• If d(x, y) < δ, then
kfn(φ(x)) − fn(φ(y))k ≤
Lδ
S
d(φ(x), φ(y)) ≤
LδT
S
d(x, y) < Ld(x, y).
Thus, φ∗ induces a map θφ : VL∞ (Y ) → VL∞ (X); likewise, ψ∗ induces
a map θψ : VL∞ (X) → VL∞ (Y ). Let us show that these maps are
inverses. By the symmetry of their definition, it suffices to show that
θψ ◦ θφ = idVL∞(Y ).
Let (fn)∞
n=1 ∈ VL∞ (Y ). Let Γ (φ ◦ ψ) ⊂ Y × Y denote the graph of
φ ◦ ψ. Since this is uniformly bounded, given ǫ > 0, we may find n0
such that, for n ≥ n0 and (x, y) ∈ Γ (φ ◦ ψ),
fn (x) − fn (y) ≤ ǫ.
In other words, for n ≥ n0, fn (x) − fn (φ (ψ (x))) ≤ ǫ, and thus,
kfn − ψ∗ ◦ φ∗ (fn) k ≤ ǫ.
Consequently, k(fn)∞
trary, θψ ◦ θφ ((fn)∞
n=1 − θψ ◦ θφ ((fn)∞
n=1) = (fn)∞
n=1 in VL∞(Y ).
n=1)kVL∞(Y ) ≤ ǫ. Since ǫ is arbi-
(cid:3)
Lemma 6.3. Let X be a metric space, and let E ⊆ X × X. The
following are equivalent:
(i) E is uniformly bounded.
(ii) For every (fn)∞
n=1 ∈ VL (X) and ǫ > 0, there exists n0 such
that, if n ≥ n0 and (x, y) ∈ E then fn (x) − fn (y) < ǫ.
(iii) For every (fn)∞
n=1 ∈ VL (X), there exists n1 such that, if (x, y) ∈
E then fn1 (x) − fn1 (y) < 1.
Proof. (i) ⇒ (ii) is by the definition of VL (X). (ii) ⇒ (iii) is trivial.
For (iii) ⇒ (i), let E ⊆ X × X be a set that isn't uniformly bounded,
n=1 ∈ VL (X) such that, for all n
and let us show that there exists (fn)∞
there exists (xn, yn)∞
n=1 ∈ E such that fn (x) − fn (y) = 1.
For each n, there exists (xn, yn) ∈ E such that d(x, y) > n. Thus
there exists a (1/n)-Lipschitz function fn : X → R such that fn(xn) =
0, and fn(yn) = 1. It follows that (fn)∞
n=1 ∈ VL (X), showing that (iii)
doesn't hold.
(cid:3)
Proposition 6.4. Let X, Y be metric spaces. Let φ : Y → X be a
function.
(i) If φ∗ (VL (X)) ⊆ VL (Y ) then φ is a bornologous map;
(ii) If, moreover, the induced map θφ : VL∞ (X) → VL∞ (Y ) is an
isomorphism, then φ is a coarse equivalence.
28
J ´AN SPAKULA AND AARON TIKUISIS
Proof. (i): Let R > 0. We must show that E := {(φ(x), φ(y)) : x, y ∈
X, d(x, y) < R} is uniformly bounded. To this end, we will verify
Lemma 6.3 (iii) for this set. Therefore, let (fn)∞
n=1 ∈ VL (X).
Since φ∗ ((fn)∞
n=1) ∈ VL (Y ), there exists n1 such that fn1 ◦ φ is
R−1-Lipschitz. Thus for (φ(x), φ(y)) ∈ E, i.e., d(x, y) < R,
fn1(φ(x)) − fn1(φ(x)) = fn1 ◦ φ(x) − fn1 ◦ φ(y) < 1,
as required.
(ii): We must show two things: (a) for every R > 0, the set {(x, y) ∈
Y × Y : d(φ(x), φ(y)) < R} is uniformly bounded, and (b) there ex-
ists R > 0 such that, for all x ∈ X, there exists y ∈ Y such that
d (x, φ (y)) < R.
(a): Let R ≥ 0 be given. We will verify Lemma 6.3 (iii) for E :=
n=1 ∈ VL (Y ).
{(x, y) ∈ Y × Y : d(φ(x), φ(y)) < R}. Therefore, let (fn)∞
Since θφ is surjective, there exists (gn)∞
n=1 ∈ VL (X) such that lim∞ kfn−
gn◦φk = 0. By Lemma 6.3 (i) ⇒ (ii), there exists n0 such that, if n ≥ n0
and w, z ∈ X satisfy d(w, z) < R then gn(w) ≈1/3 gn(z). Pick n ≥ n0
such that kfn − gn ◦ φk < 1
3 .
Now, let (x, y) ∈ E, i.e., d(φ(x), φ(y)) < R. Then
fn (x) ≈1/3 gn (φ (x)) ≈1/3 gn (φ (y)) ≈1/3 fn (y) ,
as required.
(b): Proof by contradiction. Suppose for a contradiction that, for
every n there exists xn ∈ X such that for all y ∈ Y ,
(xn, φ (y)) ≥ n.
Thus, there exists a (1/n)-Lipschitz function fn : X → [0, 1] such
that fn(xn) = 1 and fn(φ(y)) = 0 for all y ∈ Y . Putting these together,
we obtain (fn)∞
n=1 ∈ VL (X) and kfnk = 1 (since fn (xn) = 1), so that
k(fn)∞
n=1k = 1 in VL∞ (X). However, since fn (φ (y)) = 0 for all y ∈ Y ,
it follows that θφ ((fn)∞
n=1) = 0, which contradicts injectivity of θφ. (cid:3)
In other words, when VL∞ (X) ∼= VL∞ (Y ), and the isomorphism
comes from a map between Y and X, it follows that X and Y are
coarsely equivalent. Here is the more interesting question:
Question 6.5. Let X, Y be uniformly discrete metric spaces. If VL∞ (X) ∼=
VL∞ (Y ), must X and Y be coarsely equivalent?
6.2. The nuclear dimension of VL∞ (X). VL (X) and VL∞ (X) are
commutative unital C*-algebras, and therefore by Gelfand's Theorem,
each are algebras of continuous functions on a compact Hausdorff space,
namely the Gelfand spectrum of the respective algebras. As these C*-
algebras are nonseparable, their spectra are nonmetrizable. Here we
show a relationship between the asymptotic dimension of X and the
RELATIVE COMMUTANT PICTURES OF ROE ALGEBRAS
29
covering dimension (suitably interpreted) of these spectra. In fact we
use the nuclear dimension of the algebras, which (for the spectra) cor-
responds to a version of covering dimension which is slightly modified
(in this nonseparable case) from the original definition.
Definition 6.6. [28, Definition 2.1] Let A be a C*-algebra and let d ∈
N. We say that the nuclear dimension of A is at most d if there exists a
net (Fλ, ψλ, φλ) where Fλ is a finite dimensional C*-algebra, ψλ : A →
Fλ and φλ : Fλ → A are completely positive maps such that
lim
λ
ψλ(φλ(a)) = a,
a ∈ A,
ψλ is contractive, and Fλ decomposes into direct summands as Fλ =
λ ⊕ · · ·⊕ F (d)
F (0)
is contractive and order zero, for each
i.
such that φλF (i)
λ
λ
Since we will be considering nuclear dimension for the commutative
and nonseparable C*-algebras VL(X) and VL∞(X), let us explain ex-
actly the modification to covering dimension that is entailed by nuclear
dimension. Let Y be a locally compact Hausdorff space; call an open
set U ⊆ Y a preimage-open set if it is the preimage of an open subset of
R under a continuous function Y → R. Then the nuclear dimension of
C0(Y ) is equal to the smallest number d such that every finite cover of
Y consisting of preimage-open sets has a (d + 1)-colourable refinement
consisting of preimage-open sets (see the proof of [16, Proposition 3.3],
and [28, Proposition 2.4]; this fact is alluded to in the discussion be-
fore [28, Proposition 2.4]). (In the second countable situation, or more
generally when Y is a normal space, all open sets are preimage-open,
which is why nuclear dimension coincides with the usual definition of
covering dimension in this case.)
Definition 6.7. ([10, §1.E]) Let X be a metric space. Then the as-
ymptotic dimension of X is at most d ∈ N, written asdim(X) ≤ d, if
for every R > 0, there exists a cover of X of the form (U (i)
j )i=0,...,d; j∈J ,
such that for each i = 0, . . . , d, the family (U (i)
j )j∈J is R-disjoint and
uniformly bounded.
Proposition 6.8. Let X be a metric space. dimnuc VL (X) ≤ asdim(X)
and dimnuc VL∞ (X) ≤ asdim(X).
Proof. As the nuclear dimension decreases when passing to quotients,
it suffices to prove the first statement.
Set d := asdim(X). Let F ⊂ VL(X) be a finite set and let ǫ > 0 be
given.
Using the definition of asymptotic dimension in a fairly straightfor-
ward way, for each n ∈ N, we may find an infinite partition of unity
(e(i)
j (n))j∈J(n); i=0,...,d, such that:
(i) for each i, (e(i)
j (n))j∈J(n) are pairwise orthogonal,
30
J ´AN SPAKULA AND AARON TIKUISIS
(ii) each e(i)
j
(iii) there is a uniform bound, S(n), on the diameters of the sup-
is (1/n)-Lipschitz, and
ports of e(i)
j (n) (allowed to depend only on n).
Let us also pick a point x(i)
i, j, and n. For n = 0, define J(0) := X, e(0)
for i > 0, S(0) := 0, and x(i)
j (0) = j ∈ X.
j (n) inside the support of e(i)
j (0)(x) := δj,x, e(i)
j (n), for each
j (0) := 0
Define a ∗-homomorphism ψn = (ψ(0)
n , . . . , ψ(d)
coordinatewise by evaluation at x(i)
homomorphism φ(i)
n : l∞(J(n)) → Cb(X) by
j (n). For i = 0, . . . , n, define a ∗-
n ) : Cb(X) →Ld
i=0 l∞(J(n))
φ(i)
n (λ) = Xj∈J(n)
λ(j)e(i)
j (n),
with the sum converging pointwise, since in fact at each point x ∈ X,
at most one summand is nonzero (by condition (i)). Note that if f ∈
Cb(X) is (ǫ/S(n))-Lipschitz then
(6.1)
Let
φ(i)
n ◦ ψ(i)
n (f ) ≈ǫ f.
d
Xi=0
F = {(fi,k)∞
k=1 : i = 1, . . . , m}.
For each k, we may find some nk ≥ 0 such that fi,k has (ǫ/S(nk))-
Lipschitz for all i = 1, . . . , m. Since F ⊆ VL(X), we can pick these nk
such that they converge to ∞.
We now define a ∗-homomorphism
Ψ := (ψnk )∞
k=1 : l∞(N, l∞(X)) →Yk
d
Mi=0
l∞(J(nk)) ∼=
d
Mi=0 Yk
l∞(J(nk));
we may write Ψ = (Ψ(0), . . . , Ψ(d)). For i = 0, . . . , d, define a ∗-
homomorphism
Φ(i) := (φ(i)
nk)∞
k=1 :Yk
l∞(J(nk)) → l∞(N, l∞(X)).
Since nk → ∞, we see that the image of Φ(i) is in fact contained in
VL(X). Moreover, by (6.1) and our choice of nk, we find that
d
Φ(i) ◦ Ψ(i)(f ) ≈ǫ f
Xi=0
for f ∈ F . Since Qk l∞(J(nk)) has nuclear dimension zero, this is
sufficient to prove that VL(X) has nuclear dimension at most d.
(cid:3)
RELATIVE COMMUTANT PICTURES OF ROE ALGEBRAS
31
We have an argument to get inequalities in the other direction, under
the hypothesis that X has finite asymptotic dimension. For this, we
begin with the following lemma.
Lemma 6.9. Let X be a set and let η > 0. Let f1, . . . , fm, e1, . . . , en :
X → [0, ∞), and λi,j ∈ [0, ∞) for i = 1, . . . , n, j = 1, . . . , m. Suppose
j=1 λi,j = 1 for each i. Then
that fj ≈η Pn
i=1 λi,jei for each j and Pm
for each i, there exists j(i) such that
{x : ei(x) > mη} ⊆ {x : fj(i)(x) > 0}.
λi,j ≥ 1/m. For x ∈ X such that fj(x) = 0, it follows that
j=1 λi,j = 1, there exists some j = j(i) such that
Proof. Fix i. Since Pm
1
m
ei(x) ≤ λi,jei(x)
n
≤
λi′,jei′(x)
Xi′=1
≤ fj(x) + η = η.
This shows that ei(x) ≤ mη, as required.
(cid:3)
Theorem 6.10. If X has finite asymptotic dimension, then
asdim(X) = dimnuc VL(X) = dimnuc VL∞(X).
Proof. Set d := dimnuc VL∞(X). By Proposition 6.8, it suffices to show
that asdim(X) ≤ d. Let R > 0 be given, and we will partition X into
(d + 1) uniformly bounded, R-disjoint families.
By hypothesis, let asdim(X) ≤ m − 1. Then from this, there exists a
partition of unity g1, . . . , gm ∈ VL∞(X), such that gj = (gj,l)∞
l=1 where
for each j, l, the support of gj,l decomposes as an l-disjoint, uniformly
bounded family of subsets of X.
Set
η :=
1
3(d + 1)m
.
The only nonzero order zero maps from a matrix algebra into a com-
mutative algebra occur when the matrix algebra is one-dimensional
(this follows from [27, Proposition 3.2 (a)]). Hence, dimnuc VL(X) ≤ d
implies that there exists s ∈ N, a c.p.c. map ψ = (ψ(0), . . . , ψ(d)) :
C⊕s, and c.p.c. order zero maps φ(i) : Cs → VL∞(X)
i=0
such that
VL∞(X) →Ld
Xi=0
gj ≈η
m
φ(i) ◦ ψ(i)(gj) and 1VL∞(X) ≈1/2
φ(i) ◦ ψ(i)(1).
m
Xi=0
32
J ´AN SPAKULA AND AARON TIKUISIS
By rescaling, we may assume that ψ(i)(1VL∞(X)) = (1, . . . , 1). Write
ψ(i)(gj) = (λ(i,1),j, . . . , λ(i,s),j) ∈ [0, ∞)s (since ψ is positive). By linear-
ity, looking at the i′ component of ψ(i)(1), we have
m
λ(i,i′),j = 1.
Xj=1
Each map φ(i) lifts to a c.p.c. order zero map (φ(i)
l )∞
l=1 : Cs → VL(X),
by [16, Remark 2.4]. For all but finitely many l, we have
gj,l ≈η
φ(i)
l (λ(i,1),j, . . . , λ(i,s),j)
and
(6.2)
1Cb(X) ≈1/2
φ(i)
l (1, . . . , 1).
d
Xi=0
Xi=0
d
Xi=0
Xi′=1
Fix l ≥ R for which (6.2) holds, and such that the image of each mini-
is (mη/L)-Lipschitz. Write e(i,1), . . . , e(i,s)
l of the minimal projections in Cs, and write
mal projection in Cs under φ(i)
l
for these images under φ(i)
fj := gj,l; thus (6.2) becomes
(6.3)
fj ≈η
λ(i,i′),je(i,i′)
and 1 ≈1/2
e(i,i′).
d
s
d
s
Xi=0
Xi′=1
We now apply Lemma 6.9 with (i, i′), i = 0, . . . , d, i′ = 1, . . . , s in
place of the index i = 1, . . . , n. This tells us that for each i = 0, . . . , d
and i′ = 1, . . . , s, there exists some j(i, i′) such that
Bi,i′ := {x ∈ X : e(i,i′) > mη} ⊆ {x ∈ X : fj(i,i′)(x) > 0}.
Since the support of fj(i,i′) (= gj(i,i′),l) decomposes as a union of an
l-disjoint uniformly bounded family, and l ≥ R, we can partition Bi,i′
into an R-disjoint, uniformly bounded family, say
A(i)
i′,t
Bi,i′ =at∈T
Fixing i ∈ {0, . . . , d}, we now consider the family (A(i)
i′,t)i′=1,...,s, t∈T .
This family is a finite union of uniformly bounded families, whence it
is uniformly bounded. Let us check that it is R-disjoint. Since for fixed
i′ we already have R-disjointness of (A(i)
i′,t)t∈T , we need to show that for
i′
1 6= i′
is
at least R. In other words, we need to show that the minimal distance
between Bi,i′
2 and t1, t2 ∈ T , the minimal distance between A(i)
i′
1,t1
and A(i)
i′
2,t2
is at least R.
and Bi,i′
We have that ei,i′
1(x) 6= 0, it must be the case that ei,i′
and ei,i′
ei,i′
Bi,i′
that d(x, y) ≥ R, as required.
, so that ei,i′
(y) > mη. Since ei,i′
2
2
1
2
2
are orthogonal, so if x ∈ Bi,i′
then since
2(x) = 0. Consider now y ∈
is (mη/R)-Lipschitz, it follows
1
1
2
RELATIVE COMMUTANT PICTURES OF ROE ALGEBRAS
33
Finally let us show that
(A(i)
i′,t)i=0,...,d, i′=1,...,s, t∈T
is a cover of X, i.e., that X = Si,i′ Bi,i′. For x ∈ X, from the second
part of (6.3), we have
d
s
1/2 <
e(i,i′)(x).
Xi=0
Xi′=1
At most d+1 terms in this sum are nonzero, due to the pairwise orthog-
onality withing each family (e(i,i′))s
i′=1. Therefore, there exists some i, i′
such that e(i,i′)(x) > 1
2(d+1) = mη. Thus, x ∈ Bi,i′ (by definition) as
required.
(cid:3)
We have the following consequence.
Corollary 6.11. Suppose X is a metric space, and for all m ∈ N, X
contains a subspace Ym such that asdim(Ym) ∈ [m, ∞). Then
dimnuc VL(X) = dimnuc VL∞(X) = ∞.
Proof. It is not too hard to see that restriction to Ym produces a surjec-
tive ∗-homomorphism VL∞(X) → VL∞(Ym) (for surjectivity, the key
point is that an L-Lipschitz function on a closed subset of Ym extends
to an L-Lipschitz function on X). Hence we have
dimnuc VL∞(X) ≥ dimnuc VL∞(Ym) ≥ m,
using [28, Proposition 2.3(iv)] for the first inequality and Theorem 6.10
for the second.
(cid:3)
In [2, Theorem 7.2], it was shown that the asymptotic dimension of
X is equal to the covering dimension of the Higson corona νX, likewise
provided that asdim(X) < ∞.
Question 6.12. Is dimnucVL(X) = asdim(X) always? Is dimnucVL(X) =
dim(νX) always?
In light of Corollary 6.11 and [2, Theorem 7.2], the above question is
only open in the case of a metric space X of infinite dimension, which
does not contain subspaces of arbitrarily large finite dimension.
References
[1] G. Bell and A. Dranishnikov. Asymptotic dimension. Topology Appl.,
155(12):1265 -- 1296, 2008.
[2] A. N. Dranishnikov. Asymptotic topology. Uspekhi Mat. Nauk, 55(6(336)):71 --
116, 2000.
[3] Alexander Dranishnikov. On Gromov's positive scalar curvature conjecture for
duality groups. J. Topol. Anal., 6(3):397 -- 419, 2014.
34
J ´AN SPAKULA AND AARON TIKUISIS
[4] Alexander Dranishnikov and Michael Zarichnyi. Asymptotic dimension, de-
composition complexity, and Haver's property C. Topology Appl., 169:99 -- 107,
2014.
[5] Alexander Engel. Index theory of uniform pseudodifferential operators. arXiv
preprint 1502.00494v1.
[6] Alexander Engel. Rough index theory on spaces of polynomial growth and
contractibility. arXiv preprint 1505.03988v3.
[7] Vladimir Georgescu. On the structure of the essential spectrum of elliptic op-
erators on metric spaces. J. Funct. Anal., 260(6):1734 -- 1765, 2011.
[8] Vladimir Georgescu. On the essential spectrum of the operators in certain
crossed products, 2017. arXiv preprint 1705.00379.
[9] Vladimir Georgescu and Andrei Iftimovici. Localizations at infinity and essen-
tial spectrum of quantum Hamiltonians. I. General theory. Rev. Math. Phys.,
18(4):417 -- 483, 2006.
[10] M. Gromov. Asymptotic invariants of infinite groups. In Geometric group the-
ory, Vol. 2 (Sussex, 1991), volume 182 of London Math. Soc. Lecture Note
Ser., pages 1 -- 295. Cambridge Univ. Press, Cambridge, 1993.
[11] Mikhael Gromov. Groups of polynomial growth and expanding maps. Inst.
Hautes ´Etudes Sci. Publ. Math., (53):53 -- 73, 1981.
[12] Erik Guentner and Jerome Kaminker. Exactness and the Novikov conjecture.
Topology, 41(2):411 -- 418, 2002.
[13] Erik Guentner, Romain Tessera, and Guoliang Yu. A notion of geometric com-
plexity and its application to topological rigidity. Invent. Math., 189(2):315 --
357, 2012.
[14] Erik Guentner, Romain Tessera, and Guoliang Yu. Discrete groups with finite
decomposition complexity. Groups Geom. Dyn., 7(2):377 -- 402, 2013.
[15] Gennadi Kasparov and Guoliang Yu. The coarse geometric Novikov conjecture
and uniform convexity. Adv. Math., 206(1):1 -- 56, 2006.
[16] Eberhard Kirchberg and Wilhelm Winter. Covering dimension and quasidiag-
onality. Internat. J. Math., 15(1):63 -- 85, 2004.
[17] B. V. Lange and V. S. Rabinovich. Noethericity of multidimensional discrete
convolution operators. Mat. Zametki, 37(3):407 -- 421, 462, 1985.
[18] Narutaka Ozawa. Amenable actions and exactness for discrete groups. C. R.
Acad. Sci. Paris S´er. I Math., 330(8):691 -- 695, 2000.
[19] V. S. Rabinovich, S. Roch, and B. Silbermann. Fredholm theory and finite sec-
tion method for band-dominated operators. Integral Equations Operator The-
ory, 30(4):452 -- 495, 1998. Dedicated to the memory of Mark Grigorievich Krein
(1907 -- 1989).
[20] John Roe. Personal communication.
[21] John Roe. An index theorem on open manifolds. I. J. Differential Geom.,
27(1):87 -- 113, 1988.
[22] John Roe. Index theory, coarse geometry, and topology of manifolds, volume 90
of CBMS Regional Conference Series in Mathematics. Published for the Con-
ference Board of the Mathematical Sciences, Washington, DC; by the American
Mathematical Society, Providence, RI, 1996.
[23] John Roe. Lectures on coarse geometry, volume 31 of University Lecture Series.
American Mathematical Society, Providence, RI, 2003.
[24] Thomas Schick. The topology of positive scalar curvature. In Proceedings of
the International Congress of Mathematicians Seoul 2014, volume II, pages
1285 -- 1308. Kyung Moon SA Co. Ltd., 2014.
[25] G. Skandalis, J. L. Tu, and G. Yu. The coarse Baum-Connes conjecture and
groupoids. Topology, 41(4):807 -- 834, 2002.
RELATIVE COMMUTANT PICTURES OF ROE ALGEBRAS
35
[26] J´an Spakula and Rufus Willett. On rigidity of Roe algebras. Adv. Math.,
249:289 -- 310, 2013.
[27] Wilhelm Winter. Covering dimension for nuclear C*-algebras. J. Funct. Anal.,
199(2):535 -- 556, 2003.
[28] Wilhelm Winter and Joachim Zacharias. The nuclear dimension of C*-algebras.
Adv. Math., 224(2):461 -- 498, 2010.
[29] Nick Wright. C0 coarse geometry and scalar curvature. J. Funct. Anal.,
197(2):469 -- 488, 2003.
[30] Guoliang Yu. Zero-in-the-spectrum conjecture, positive scalar curvature and
asymptotic dimension. Invent. Math., 127(1):99 -- 126, 1997.
[31] Guoliang Yu. The Novikov conjecture for groups with finite asymptotic dimen-
sion. Ann. of Math. (2), 147(2):325 -- 355, 1998.
[32] Guoliang Yu. The coarse Baum-Connes conjecture for spaces which admit a
uniform embedding into Hilbert space. Invent. Math., 139(1):201 -- 240, 2000.
J´an Spakula, Mathematical Sciences, University of Southampton,
Highfield, Southampton, SO17 1BJ, United Kingdom
E-mail address: [email protected]
Aaron Tikuisis, Department of Mathematics and Statistics, Univer-
sity of Ottawa, Ottawa, Ontario, K1N 6N5, Canada
E-mail address: [email protected]
|
1512.03347 | 1 | 1512 | 2015-12-10T17:59:23 | Cohomology of Jordan triples via Lie algebras | [
"math.OA",
"math.RA"
] | We develop a cohomology theory for Jordan triples, including the infinite dimensional ones, by means of the cohomology of TKK Lie algebras. This enables us to apply Lie cohomological results to the setting of Jordan triples. Some preliminary results for von Neumann algebras are obtained. | math.OA | math | Cohomology of Jordan triples via Lie algebras
Cho-Ho Chu and Bernard Russo
Abstract. We develop a cohomology theory for Jordan triples, including the infinite dimensional
ones, by means of the cohomology of TKK Lie algebras. This enables us to apply Lie cohomological
results to the setting of Jordan triples. Some preliminary results for von Neumann algebras are
obtained.
.
A
O
h
t
a
m
[
1
v
7
4
3
3
0
.
2
1
5
1
:
v
i
X
r
a
Contents
Introduction
1.
2. Jordan triples and TKK Lie algebras
3. Cohomology of Lie algebras with involution
4. Cohomology of Jordan triples
5. Examples
6. Proofs of Theorems 2.3 and 5.5
References
1
3
6
7
13
16
26
1. Introduction
A veritable army of researchers took the theory of derivations of operator algebras to
dizzying heights -- producing a theory of cohomology of operator algebras as well as
much information about automorphisms of operator algebras -- Richard Kadison [19]
In addition to associative algebras, cohomology groups are defined for Lie algebras and, to some
extent, for Jordan algebras. Since the structures of Jordan derivations and Lie derivations on von
Neumann algebras are well understood, and in view of the above quotation, isn't it time to study
the higher dimensional non associative cohomology of a von Neumann algebra? The present paper is
motivated by this rhetorical question.
In this paper we develop a cohomology theory for Jordan triples, including the infinite dimensional
ones, by means of the cohomology of TKK Lie algebras. This enables us to apply Lie cohomological
results to the setting of Jordan triples. Several references, which will be mentioned below, use Lie
theory as a tool to study Jordan cohomology.
The outline of the paper is the following. In the rest of this introduction, we give an overview of
various cohomology theories, both classical and otherwise. (For a more detailed survey see [29].) In
section 2, the definitions of Jordan triple module and Lie algebra module, as well as the Tits-Kantor-
Koecher (TKK) construction are reviewed, basically following [5]. It is shown in Theorem 2.3 that
1991 Mathematics Subject Classification. Primary 17C65, 18G60; Secondary 46L70, 16W10.
Key words and phrases. Jordan triple, cohomology, TKK algebra, derivation, cocycle, structural transformation,
von Neumann algebra.
1
2
CHO-HO CHU AND BERNARD RUSSO
a Jordan triple module gives rise to a Lie module for the corresponding TKK algebra. The proof of
Theorem 2.3 is deferred to subsection 6.1.
After reviewing the cohomology of Lie algebras (with or without an involution) in section 3, two
infinite families of cohomology groups are defined for a Jordan triple system V in section 4, one using
the Lie cohomology of the TKK algebra of V and the other using the Lie cohomology of the TKK
algebra with its canonical involution θ. A complete analysis is given for the first cohomology groups
in Proposition 4.6, which shows that structural transformations on V correspond to derivations of the
TKK Lie algebra, and triple derivations on V correspond to the θ-invariant derivations.
Section 5 contains examples of Jordan cocycles and TKK algebras, and applications, including a
characterization of certain 3-cocycles in Theorem 5.5, the proof of which appears in subsection 6.2.
The applications to von Neumann algebras appear in Theorem 4.7 and Corollary 5.7.
1.1. Brief survey of cohomology theories. The starting point for the cohomology theory of
associative algebras is the paper of Hochschild from 1945 [12]. The standard reference of the theory
is [3]. Two other useful references are due to Weibel ([35],[36]).
Shortly after the introduction of cohomology for associative algebras, there appeared in [4] a cor-
responding theory for Lie algebras. We follow [16] for the definitions and initial results. Applications
can be found in [7] and [20].
The cohomology theory for Jordan algebras is less well developed than for associative and Lie
algebras. A starting point would seem to be the papers of Gerstenhaber in 1964 [8] and Glassman
in 1970 [10], which concern arbitrary nonassociative algebras. A study focussed primarily on Jordan
algebras is [9].
We next recall two fundamental results, namely, the Jordan analogs of the first and second
Whitehead lemmas as described in [15].
Theorem 1.1 (Jordan analog of first Whitehead lemma [14]). Let J be a finite dimensional
semisimple Jordan algebra over a field of characteristic 0 and let M be a J -module. Let f be a linear
mapping of J into M such that
f (ab) = f (a)b + af (b).
Then there exist vi ∈ M, bi ∈ J such that
f (a) =Xi
((via)b − vi(abi)).
Theorem 1.2 (Jordan analog of second Whitehead lemma [27]). Let J be a finite dimensional
separable1 Jordan algebra and let M be a J -module. Let f be a bilinear mapping of J × J into M
such that
and
f (a2, ab) + f (a, b)a2 + f (a, a)ab = f (a2b, a) + f (a2, b)a + (f (a, a)b)a
Then there exist a linear mapping g from J into M such that
f (a, b) = f (b, a)
f (a, b) = g(ab) − g(b)a − g(a)b
Two proofs of Theorem 1.2 are given in [17]. One of them, which uses the classification of finite
dimensional Jordan algebras, is outlined in [29, 4.3.1]. The other proof uses Lie algebras and is
contained in [17, pp. 324 -- 336].
A study of low dimensional cohomology for quadratic Jordan algebras is given in [24]. Since
quadratic Jordan algebras (which coincide with "linear" Jordan algebras over characteristic 0 fields)
can be considered a bridge from Jordan algebras to Jordan triple systems, this would seem to be a
good place to look for exploring cohomology theory for Jordan triples. Indeed, this is hinted at in
1Separable, in this context, means that the algebra remains semisimple with respect to all extensions of the ground
field. For algebraically closed fields, this is the same as being semisimple
COHOMOLOGY OF JORDAN TRIPLES VIA LIE ALGEBRAS
3
[25], since although [24] is about Jordan algebras, the concepts are phrased in terms of the associated
triple product {abc} = (ab)c + (cb)a − (ac)b.
However, both papers stop short of defining higher dimensional cohomology groups. The paper
[24], which is mostly concerned with representation theory, proves, for the only cohomology groups
defined, the linearity of the functor H n:
H n(J, ⊕iMi) = ⊕iH n(J, Mi), n = 1, 2.
The paper [25], which is mostly concerned with compatibility of tripotents in Jordan triple systems,
proves versions of the linearity of the functor H n, n = 1, 2, corresponding to the Jordan triple
structure.
The earliest work on cohomology of triple systems seems to be [11] (Lie triple systems), which is
discussed in section 3. Four decades later, the second paper on the cohomology of Lie triple systems
appeared [13].
The following is from the review [32] of [1] (associative triple systems).
"A cohomology for associative triple systems is defined, with the main purpose to get
quickly the cohomological triviality of finite-dimensional separable objects over fields
of characteristic 6= 2, i.e., in particular the Whitehead lemmas and the Wedderburn
principal theorem."
The authors of the present paper know of only two other references dealing with the Wedderburn
principal theorem in the context of triple systems, namely, [2] (alternative triple systems) and [23]
(Jordan triple systems). In the latter paper, the well-known Koecher-Tits-construction of a Lie algebra
from a Jordan algebra is generalized to Jordan pairs. The radical of this Lie algebra is calculated
in terms of the given Jordan pair and a Wedderburn decomposition theorem for Jordan pairs (and
triples) in the characteristic zero case is proved.
Finally, we mention that a more general approach to cohomology of algebras and triple systems
appears in the paper of Seibt [31].
2. Jordan triples and TKK Lie algebras
By a Jordan triple, we mean a real or complex vector space V , equipped with a Jordan triple
product {·, ·, ·} : V 3 → V which is linear and symmetric in the outer variables, conjugate linear in
the middle variable, and satisfies the Jordan triple identity
{x, y, {a, b, c}} = {{x, y, a}, b, c} − {a, {y, x, b}, c} + {a, b, {x, y, c}}
for a, b, c, x, y ∈ V . Given two elements a, b in a Jordan triple V , we define the box operator a b :
V → V by a b(·) = {a, b, ·}.
All Lie algebras in this paper are real or complex. We construct a cohomology theory of Jordan
triples using the Tits-Kantor-Koecher (TKK) Lie algebras associated with them. Although we could
develop the theory for all Jordan triples, we focus on the nondegenerate ones, which will be assumed
throughout, to avoid unnecessary complication. For degenerate Jordan triples, the construction is
exactly the same albeit more computation is involved. A Jordan triple is called nondegenerate if for
each a ∈ V , the condition {a, a, a} = 0 implies a = 0. Given that V is nondegenerate, one has
dk
ck
(aj, bj, ck, dk ∈ V )
Xj
aj
bj =Xk
ck
dk ⇒ Xk
bj aj =Xj
which facilitates a simple definition of the TKK Lie algebra L(V ) of V , with an invoultion θ (cf. [5,
p.45]), where
V0 = {Pj aj
(2.1)
L(V ) = V ⊕ V0 ⊕ V,
bj : aj, bj ∈ V }, the Lie product is defined by
[(x, h, y), (u, k, v)] = (hu − kx, [h, k] + x v − u y, k♮y − h♮v),
4
CHO-HO CHU AND BERNARD RUSSO
and for each h =Pi ai
bi in the Lie subalgebra V0 of L(V ), the map h♮ : V → V is well defined by
h♮ =Xi
bi ai.
The involution θ : L(V ) → L(V ) is given by
θ(x, h, y) = (y, −h♮, x)
((x, h, y) ∈ L(V )).
Identifying V with the subspace {(x, 0, 0) : x ∈ V } of L(V ), we have the following relationship between
the triple and Lie products:
{a, b, c} = [ [a, θ(b)], c]
(a, b, c ∈ V ).
If no confusion is likely, we often simplify the notation {a, b, c} to {abc}.
Given a Lie algebra L and a module X over L, we denote the action of L on X by
so that
(ℓ, x) ∈ L × X 7→ ℓ.x ∈ X
[ℓ, ℓ′].x = ℓ′.(ℓ.x) − ℓ.(ℓ′.x).
Definition 2.1. Let V be a Jordan triple. A vector space M over the same scalar field is called
a Jordan triple V -module (cf. [29]) if it is equipped with three mappings
{·, ·, ·}1 : M × V × V → M,
{·, ·, ·}2 : V × M × V → M, {·, ·, ·}3 : V × V × M → M
such that
(i) {a, b, c}1 = {c, b, a}3;
(ii) {·, ·, ·}1 is linear in the first two variables and conjugate linear in the last variable, {·, ·, ·}2
is conjugate linear in all variables;
(iii) denoting by {·, ·, ·} any of the products {·, ·, ·}j (j = 1, 2, 3), the identity
{a, b, {c, d, e}} = {{a, b, c}, d, e} − {c, {b, a, d}, e} + {c, d, {a, b, e}}
is satisfied whenever one of the above elements is in M and the rest in V .
For convenience, we shall omit the subscript j from {·, ·, ·}j in the sequel. A V -module M is
called nondegenerate if for each m ∈ M , each one of the conditions
{m, V, V } = {0}; {V, m, V } = {0}
implies m = 0. A nondegenerate Jordan triple V is a nondegenerate module over itself. For a JB*-
triple V , its dual V ∗ is a nondegenerate V -module. All Jordan triple modules throughout the paper
are assumed to be nondegenerate.
Given a, b ∈ V , the box operator a b : V → V can also be considered as a mapping from M to
M . Similarly, for u ∈ V and m ∈ M , the "box operators"
are defined in a natural way as v 7→ {u, m, v} and v 7→ {m, u, v} respectively.
Given a, b ∈ V , the identity (iii) in Definition 2.1 implies
u m, m u : V −→ M
and
[a b, u m] = {a, b, u} m − u {m, a, b}
[a b, m u] = {a, b, m} u − m {u, a, b}.
for u ∈ V and m ∈ M . We also have [u m, a b] = {u, m, a} b − a {b, u, m} and similar identity
for [m u, a b].
COHOMOLOGY OF JORDAN TRIPLES VIA LIE ALGEBRAS
5
(2.2)
Using similar arguments to the proof in [5, Lemma 1.3.7], one can show that
Xi
⇒ Xi
ui mi +Xj
mi ui +Xj
nj
vj =Xk
vj nj =Xk
k m′
u′
m′
k u′
k +Xℓ
k +Xℓ
n′
ℓ
v′
ℓ
ℓ n′
v′
ℓ
for ui, vj, u′
k, v′
ℓ ∈ V and mi, njm′
Let M0 be the linear span of
k, n′
ℓ ∈ M .
{u m, n v : u, v ∈ V, m, n ∈ M }
in the vector space L(V, M ) of linear maps from V to M . Then M0 is the space of inner structural
transformations Instrl (V, M ) (see [25, Section 7]) . Extending the above product by linearity, we can
define an action of V0 on M0 by
(h, ϕ) ∈ V0 × M0 7→ [h, ϕ] ∈ M0.
Lemma 2.2. M0 is a V0-module of the Lie algebra V0.
Proof. We are required to show that
(2.3)
[[h, k], ϕ] = [h, [k, ϕ]] − [k, [h, ϕ]].
We can assume that h = a b, k = c d and ϕ = w m or m w. We assume ϕ = w m, the other
case being similar. For the left side of (2.3), we have
[[a b, u v], w m] = [{abu} v − u {vab}, w m]
= {{abu}vw} m − w {m{abu}v}
− {u{vab}w} m + w {mu{vab}}
= ({{abu}vw} − {u{vab}w}) m
− w ({m{abu}v} − {mu{vab}}).
For the right side of (2.3), we have
[a b, [u v, w m]] − [u v, [a b, w m]] =
[a b, {uvw} m − w {muv}] − [u v, {abw} m − w {mab}]
= {ab{uvw}} w − {uvw} {mab} − {abw} {muv} + w {{muv}ab}
− {uv{abw}} m + {abw} {muv} + {uvw} {mab} − w {{mab}uv}
= ({ab{uvw}} − {uv{abw}}) m − w ({{mab}uv} − {{muv}ab}).
(2.3) now follows from the main identity for Jordan triples.
(cid:3)
Let V be a Jordan triple and L(V ) its TKK Lie algebra. Given a triple V -module M , we now
construct a corresponding Lie module L(M ) of the Lie algebra L(V ) as follows.
Let
and define the action
L(M ) = M ⊕ M0 ⊕ M
((a, h, b), (m, ϕ, n)) ∈ L(V ) × L(M ) 7→ (a, h, b).(m, ϕ, n) ∈ L(M )
by
(2.4)
(a, h, b).(m, ϕ, n) = (hm − ϕa, [h, ϕ] + a n − m b, ϕ♮b − h♮(n) ),
where, for h =Pi ai
bi and ϕ =Pi ui mi +Pj nj
{ui, mi, a} +Xj
{ai, bi, m}, ϕa =Xi
hm =Xi
vj, we have the following natural definitions
{nj, vj , a}, ϕ♮ =Xi
mi ui +Xj
vj nj
6
CHO-HO CHU AND BERNARD RUSSO
in which ϕ♮ is well-defined by (2.2).
Theorem 2.3. Let V be a Jordan triple and let L(V ) be its TKK Lie algebra. Let M be a triple
V -module. Then L(M ) is a Lie L(V )-module.
The proof of Theorem 2.3 consists of straightforward but tedious calculations. Details can be
found in subsection 6.1.
3. Cohomology of Lie algebras with involution
Let T be a Lie triple system. Harris [11, p. 155] has developed a cohomology theory for T in
which the cohomology groups are derived from the ones of its enveloping Lie algebra Lu = T + [T, T ]
where Lu is equipped with an involution θ and the cochains in the cohomology complex are invariant
under θ.
Our Jordan triple cohomology makes use of TKK Lie algebras which are involutive. To pave the
way, we review briefly the cohomology for Lie algebras, with or without an involution. Let L be a
(real or complex) Lie algebra with involution θ.
Definition 3.1. Given an involutive Lie algebra (L, θ), an (L, θ)-module is a (left) L-module M,
equipped with an involution eθ : M → M satisfying
eθ(ℓ.µ) = θ(ℓ).eθ(µ)
We also call M an involutive L-module if θ is understood.
(ℓ ∈ L, µ ∈ M).
For ℓ ∈ L and µ ∈ M, we define
[ℓ, µ] := ℓ.µ and [µ, ℓ] := −ℓ.µ.
L × · · · × L be the k-fold cartesian product of L. A k-linear map ψ : Lk → M is called
Let Lk =
θ-invariant if
z
k−times
{
}
ψ(θx1, · · · , θxk) = eθψ(x1, · · · , xk)
Let (L, θ) be an involutive Lie algebra and M an (L, θ)-module. We define A0(L, M) = M and
A0
θ(L, M) to be the 1-eigenspace of eθ: A0
For k = 1, 2, . . ., we let
θ(L, M) = {µ ∈ M : eθµ = µ}.
Ak(L, M) = {ψ : Lk → M ψ is k-linear and alternating}
and
for
(x1, · · · , xk) ∈ L × · · · × L.
For k = 0, 1, 2, . . ., we define the coboundary operator dk : Ak(L, M) → Ak+1(L, M) by d0m(x) =
Ak
θ (L, M) = {ψ ∈ Ak(L, M) : ψ is θ-invariant}.
x.m and for k ≥ 1,
(dkψ)(x1, . . . , xk+1) =
k+1Xℓ=1
(−1)ℓ+1xℓ.ψ(x1, . . . ,bxℓ, . . . , xk+1)
+ X1≤i<j≤k+1
(3.1)
(−1)i+jψ([xi, xj], . . . ,bxi, . . . ,bxj, . . . , xk+1)
where the symbol bz indicates the omission of z. The restriction of dk to the subspace Ak
θ (L, M),
θ (L, M) since a simple verification shows that dkψ is θ-invariant and
[20, p. 167]) and the two
still denoted by dk, has range Ak
alternating whenever ψ is. Also, we have dkdk−1 = 0 for k = 1, 2, . . . (cf.
cochain complexes
d0
A0(L, M) −→
d1
A1(L, M) −→
d2
A2(L, M) −→
· · ·
A0
d0
θ(L, M) −→
A1
d1
θ(L, M) −→
A2
d2
θ(L, M) −→
· · · .
COHOMOLOGY OF JORDAN TRIPLES VIA LIE ALGEBRAS
7
We often omit the subscript k from dk if there is no ambiguity.
As usual, we define the k-th cohomology group of L with coefficients in M to be the quotient
H k(L, M) = ker dk/dk−1(Ak−1(L, M)) = ker dk/im dk−1
for k = 1, 2, . . . and define H 0(L, M) = ker d0. We define the k-th involutive cohomology group of
(L, θ) with coefficients in an (L, θ)-module M to be the quotient
H k
θ (L, M) = ker dk/dk−1(Ak−1
θ
(L, M)) = ker dk/im dk−1
for k = 1, 2, . . . and define H 0
For k = 1, 2, . . ., the map
ψ + dk−1(Ak−1
θ
θ (L, M) = ker d0 ⊂ H 0(L, M).
(L, M)) ∈ H k
θ (L, M) 7→ ψ + dk−1(Ak−1(L, M)) ∈ H k(L, M)
identifies H k
θ (L, M) as a subgroup of H k(L, M).
4. Cohomology of Jordan triples
4.1. The cohomology groups. Let V be a Jordan triple and let L(V ) = V ⊕ V0 ⊕ V be its
TKK Lie algebra with the involution θ(a, h, b) = (b, −h♮, a). Given a V -module M , we have shown
in Theorem 2.3 that L(M ) = M ⊕ M0 ⊕ M is an L(V )-module. We define an induced involution
eθ : L(M ) → L(M ) by
for (m, ϕ, n) ∈ M ⊕ M0 ⊕ M.
eθ(m, ϕ, n) = (n, −ϕ♮, m)
Lemma 4.1. L(M ) is an (L(V ), θ)-module, that is, we have eθ(ℓ.µ) = θ(ℓ).eθ(µ) for ℓ ∈ L(V ) and
µ ∈ L(M ).
Let k(V ) = {(v, h, v) ∈ L(V ) : h♮ = −h} be the 1-eigenspace of the involution θ (see [5, p.48]),
which is a real Lie subalgebra of L(V ), and let k(M ) = {(m, ϕ, m) ∈ L(M ) : ϕ = −ϕ♮} be the 1-
eigenspace of eθ. Then k(M ) is a Lie module over the Lie algebra k(V ). We will construct cohomology
groups of a Jordan triple V with coefficients in a V -module M using the cohomology groups of L(V )
with coefficients L(M ). For a real Jordan triple V , one can also make use of the cohomology groups
of the real Lie algebra k(V ) with coefficients k(M ).
Let V be a Jordan triple. As usual, V is identified as the subspace
{(v, 0, 0) : v ∈ V }
of the TKK Lie algebra L(V ). For a triple V -module M , there is a natural embedding of M into
L(M ) = M ⊕ M0 ⊕ M given by
and we will identify M with ι(M ). We denote by ιp : L(M ) → ι(M ) the natural projection
ι : m ∈ M 7→ (m, 0, 0) ∈ L(M )
We define A0(V, M ) = M and for k = 1, 2, . . ., we denote by Ak(V, M ) the vector space of all
ιp(m, ϕ, n) = (m, 0, 0).
alternating k-linear maps ω : V k =
Given m ∈ M , we define
k−times
V × · · · × V → M .
z
}
{
and view L0(m) as an extension of m ∈ A0(V, M ) to an element in A0(L(V ), L(M )) = L(M ).
L0(m) = (m, 0, 0) ∈ L(M )
8
CHO-HO CHU AND BERNARD RUSSO
To motivate the definition of an extension Lk(ω) ∈ Ak(L(V ), L(M )) of an element ω ∈ Ak(V, M ),
for k ≥ 1, we first consider the case k = 1 and note that ω ∈ A1(V, M ) is a Jordan triple derivation
if and only if
ω ◦ (a b) − (a b) ◦ ω = ω(a)
b + a ω(b).
Let us call a linear transformation ω : V → M extendable if the following condition holds:
Xi
ai
bi = 0 ⇒Xi
(ω(ai)
bi + ai ω(bi)) = 0.
Thus a Jordan triple derivation is extendable, and if ω is any extendable transformation in A1(V, M ),
then the map
L1(ω)(x1 ⊕ a1
b1 ⊕ y1) := ( ω(x1), ω(a1)
b1 + a1 ω(b1), ω(y1) )
is well defined and extends linearly to an element L1(ω) ∈ A1(L(V ), L(M )), in which case we call
L1(ω) the Lie extension of ω on the Lie algebra L(V ).
Now for k > 1, given a k-linear mapping ω : V k → M , we say that ω is extendable if it satisfies
the following condition under the assumption Pi ui
vi = 0:
(ω(ui, a2, . . . , ak)
(vi + b2 + · · · + bk) + (ui + a2 + · · · + ak) ω(vi, b2, . . . , bk)) = 0,
Xi
for all a2, . . . , ak, b2, . . . , bk ∈ V .
For an extendable ω, we can unambiguously define a k-linear map Lk(ω) : L(V )k → L(M ) as the
linear extension (in each variable) of
(4.1)
Lk(ω)(x1 ⊕ a1
b1 ⊕ y1, x2 ⊕ a2
b2 ⊕ y2, · · · , xk ⊕ ak
bk ⊕ yk)
= ( ω(x1, . . . , xk),
kXj=1
ω(a1, . . . , ak)
bj +
kXj=1
aj ω(b1, . . . , bk), ω(y1, . . . , yk) ).
We call Lk(ω) the Lie extension of ω and often omit the subscript k if no confusion is likely. The
following lemma is easy to verify.
Lemma 4.2. Given an extendable ω ∈ Ak(V, M ), we have Lk(ω) ∈ Ak(L(V ), L(M )). Moreover,
Lk(ω) ∈ Ak
θ (L(V ), L(M )) if and only if k is odd.
This lemma enables us to define the following extension map on the subspace Ak(V, M )′ of
extendable maps in Ak(V, M ):
Lk : ω ∈ Ak(V, M )′ 7→ Lk(ω) ∈ Ak(L(V ), L(M )).
Conversely, given ψ ∈ Ak(L(V ), L(M )) for k = 1, 2, . . ., one can define an alternating map
J k(ψ) : V k → M
by
J k(ψ)(x1, . . . , xk) = ιpψ( (x1, 0, 0), . . . , (xk, 0, 0) )
for (x1, . . . , xk) ∈ V k. We define J 0 : L(M ) → ι(M ) ≈ M = A0(V, M ) by
((m, ϕ, n) ∈ L(M )).
J 0(m, ϕ, n) = (m, 0, 0)
We call J k(ψ) the Jordan restriction of ψ in Ak(V, M ) and sometimes write J for J k if the index k
is understood.
Example 4.3. Given a map ψ ∈ Ak
θ (L(V ), L(M )), we need not have ψ((x1, 0, 0), . . . , (xk, 0, 0)) ∈
ι(M ). Consider the inner derivation ad(m, ϕ, m) ∈ A1
θ(L(V ), L(M )) defined by
For x ∈ V , we have
ad(m, ϕ, m)(x ⊕ a b ⊕ y) = (x ⊕ a b ⊕ y).(m, ϕ, m).
ad(m, ϕ, m)(x, 0, 0) = (x, 0, 0) · (m, ϕ, m) = (−ϕ(x), x m, 0) /∈ ι(M ).
COHOMOLOGY OF JORDAN TRIPLES VIA LIE ALGEBRAS
9
With the identification of M and ι(M ), the map
J k : ψ ∈ Ak(L(V ), L(M )) 7→ Ak(V, M )
can be viewed as the left inverse of Lk : Ak(V, M )′ → Ak(L(V ), L(M )) since for an extendable ω, we
have
J kLk(ω)(x1, . . . , xk) = ιpL(ω)((x1, 0, 0), . . . , (xk, 0, 0))
= ω(x1, . . . , xk).
We can now define the cohomology groups for a Jordan triple V with coefficients M by means of
the cochain complexes for the Lie algebra L(V ) and the involutive Lie algebra (L(V ), θ):
d0
L(M ) = A0(L(V ), L(M )) −→
d1
A1(L(V ), L(M )) −→
d2
A2(L(V ), L(M )) −→
· · ·
↓ J 0
↓ J 1
↓ J 2
M = A0(V, M )
A1(V, M )
A2(V, M )
A0
d0
θ(L(V ), L(M )) −→
A1
d1
θ(L(V ), L(M )) −→
A2
d2
θ(L(V ), L(M )) −→
↓ J 0
↓ J 1
↓ J 2
A2(V, M )
For k = 0, 1, 2, . . ., the k-th cohomology groups H k(V, M ) are defined by
M = A0(V, M )
A1(V, M )
· · ·
· · ·
· · ·
· · ·
· · ·
H 0(V, M ) = J 0(ker d0) = J 0{(m, ϕ, n) : (u, h, v).(m, ϕ, n) = 0, ∀(u, h, v) ∈ L(V )}
= {m ∈ M : m v = 0, ∀v ∈ V } = {0}
H k(V, M ) = Z k(V, M )/Bk(V, M )
(k = 1, 2, . . .)
Z k(V, M ) = J k(Z k(L(V ), L(M ))), Z k(L(V ), L(M )) = ker dk
and
where
and
Bk(V, M ) = J k(Bk(L(V ), L(M ))), Bk(L(V ), L(M )) = im dk−1.
For k = 0, 1, 2 . . ., the k-th involutive cohomology groups H k
H 0(V, M ) = J 0(ker d0) = {0}
θ (V, M ) are defined by
and
where
and
H k
θ (V, M ) = Z k
θ (V, M )/Bk
θ (V, M )
(k = 1, 2, . . .)
Z k
θ (V, M ) = J k(Z k
θ (L(V ), L(M ))), Z k
θ (L(V ), L(M ))) = ker dkAk
θ (L(V ),L(M))
Bk
θ (V, M ) = J k(Bk
θ (L(V ), L(M ))), Bk
θ (L(V ), L(M )) = dk−1(Ak−1
θ
(L(V ), L(M ))).
We see that the map
ω + Bk
θ (V, M ) ∈ H k
θ (V, M ) 7→ ω + Bk(V, M ) ∈ H k(V, M )
identifies H k
k-cocycles, and the ones in H k
in Bk(V, M ) are called the coboundaries.
θ (V, M ) as a subgroup of H k(V, M ). We call elements in H k(V, M ) the Jordan triple
θ (V, M ) the involutive Jordan triple k-cocycles. Customarily, elements
10
CHO-HO CHU AND BERNARD RUSSO
4.2. Triple derivations.
Definition 4.4. Let V be a Jordan triple and M a triple V -module. A mapping ω : V → M is
called an inner triple derivation if it is of the form
ω =
kXi=1
(mi
vi − vi mi) ∈ M0
for some m1, . . . , mk ∈ M and v1, . . . , vk ∈ V . Note that ω♮ = −ω and (0, ω, 0) ∈ k(M ).
Let us compute the first involutive cohomology group H 1
θ (V, M ) = Z 1
θ (V, M )/B1
θ (V, M ). First,
we show that B1
θ (V, M ) coincides with the space of inner triple derivations from V to M .
Let ω be an inner triple drivation on V . We show that its Lie extension L(ω) is a Lie inner
derivation on the Lie algebra L(V ). Indeed, we have
L(ω)(x ⊕ a b ⊕ y) = (ω(x), ω(a)
b + a ω(b), ω(y))
= (x ⊕ a b ⊕ x).(0, −ω, 0).
Hence ω = J 1(L1(ω)) ∈ B1
A1
θ(L(V ), L(M )) be a Lie inner derivation. Then for x ∈ V , we have
θ (V, M ), where (0, −ω, 0) ∈ k(M ). Conversely, let ψ = ad(m, ϕ, n) ∈
J 1(ψ)(x) = ιpψ(x, 0, 0) = ιp(x, 0, 0).(m, ϕ, n)
= ιp(−ϕ(x), x n, 0) = −ϕ(x),
where eθ(ϕ) = ϕ implies that ϕ : V → M is an inner triple derivation.
θ (V, M ) coincides with the set of triple derivations of V .
We now show that Z 1
Lemma 4.5. Let ω : V → M be a triple derivation. Then L(ω) : L(V ) → L(M ) is a θ-invariant
Lie derivation.
Proof. For notation's sake we denote L(ω) by D. Thus
D(x, a b, y) = (ω(x), ω(a)
b + a ω(b), ω(y)),
and it is clear that D is θ-invariant. We need to verify
D[(x, a b, y), (u, c d, v)] = (x, a b, y) · D(u, c d, v) − (u, c d, v) · D(x, a b, y).
for (x, a b, y), (u, c d, v) ∈ L(V ). By writing
D[(x, a b, y), (u, c d, v)] = D[(x, 0, y), (u, 0, v)] + D[(0, a b, 0), (u, 0, v)]
+ D[(0, a b, 0), (0, c d, 0)] + D[(x, 0, y), (0, c d, 0)],
we only need to verify the three identities
(4.2)
(4.3)
and
D[(x, 0, y), (u, 0, v)] = (x, 0, y) · D(u, 0, v) − (u, 0, v) · D(x, 0, y),
D[(0, a b), (u, 0, v)] = (0, a b, 0) · D(u, 0, v) − (u, 0, v) · D(0, a b, 0),
(4.4)
D[(0, a b, 0), (0, c d, 0)] = (0, a b, 0) · D(0, c d, 0) − (0, c d, 0) · D(0, a b, 0).
These are easy consequences of the definitions. For completeness we include details. The left side of
(4.2) is
D[(x, 0, y), (u, 0, v)] = D(0, x v − u y, 0)
= (0, ω(x)
v + x ω(v) − ω(u) y − u ω(y), 0),
COHOMOLOGY OF JORDAN TRIPLES VIA LIE ALGEBRAS
11
and the right side is
(x, 0, y) · D(u, 0, v) − (u, 0, v) · D(x, 0, y) =
(x, 0, y) · (ω(u), 0, ω(v)) − (u, 0, v) · (ω(x), 0, ω(y))
= (0, x ω(v) − ω(u) y, 0) − (0, u ω(y) − ω(x)
v, 0),
proving (4.2). The left side of (4.3) is
D[(0, a b, 0), (u, 0, v)] = D({abu}, 0, −{bav}) = (ω{abu}, 0, −ω{bav})
and the right side is
(0, a b, 0) · D(u, 0, v) − (u, 0, v) · D(0, a b, 0) =
(0, a b, 0) · (ω(u), 0, ω(v)) − (u, 0, v) · (0, ω(a)
b + a ω(b), 0)
= ({abω(u)}, 0, −{baω(v)}) − ({−ω(a)bu} − {aω(b)u}, 0, {bω(a)v} + {ω(b)av}),
proving (4.3). The left side of (4.4) is
D[(0, a b, 0), (0, c d, 0)] = D(0, [a b, c d], 0) = D(0, {abc} d − c {dab}, 0)
= (0, ω{abc} d + {abc} ω(d) − ω(c) {dab} − c ω{dab}, 0)
= (0, {ω(a)bc} d + {aω(b)c} d + {abω(c)} d + {abc} ω(d)
−ω(c) {dab} − c {ω(d)ab} − c {dω(a)b} − c {daω(b)}, 0),
and the right side is
(0, a b, 0) · D(0, c d, 0) − (0, c d, 0) · D(0, a b, 0)
= (0, a b, 0) · (0, ω(c) d + c ω(d), 0) − (0, c d, 0) · (0, ω(a)
b + a ω(b), 0)
= (0, [a b, ω(c) d] + [a b, c ω(d)] − [c d, ω(a)
= (0, [a b, ω(c) d] + [a b, c ω(d)] + [ω(a)
b] − [c d, a ω(b)], 0)
b, c d] + [a ω(b), c d], 0
= (0, {abω(c)} d − ω(c) {dab} + {abc} ω(d) − c {ω(d)ab}
+{ω(a)bc} d − c {dω(a)b} + {aω(b)c} d − c {daω(b)}, 0)
proving (4.4).
(cid:3)
The previous lemma shows that all triple derivations ω on V are contained in Z 1
θ (V, M ). Con-
versely, given a Lie derivation ψ ∈ A1
θ(L(V ), L(M )), we show below that J(ψ) is a triple derivation
on V . This shows that every element in Z 1
θ (V, M ) is the
space of triple derivations modulo the inner triple derivations of V into M . This will be generalized
in the next subsection.
θ (V, M ) is a triple derivation and hence H 1
4.3. Structural Transformations. A (conjugate-) linear transformation S : V → M is said to
be a structural transformation if there exists a (conjugate-) linear transformation S ∗ : V → M such
that
and
S{xyx} + {x(S ∗y)x} = {xySx}
S ∗{xyx} + {x(Sy)x} = {xyS ∗x}.
A triple derivation D is a special case of a structural transformation with D∗ = −D. By polarization,
this property is equivalent to
and
S{xyz} + {x(S ∗y)z} = {zySx} + {xySz}
S ∗{xyz} + {x(Sy)z} = {zyS ∗x} + {xyS ∗z}.
12
CHO-HO CHU AND BERNARD RUSSO
As noted earlier, the space of inner structural transformations coincides, by definition, with the
space M0. Triple derivations which are inner structural transformations are inner triple derivations.
Also, if ω is a structural transformation, then ω −ω∗ is a triple derivation and if ω is a triple derivation,
then iω is a structural transformation which is inner if ω is inner.
Proposition 4.6. Let ψ be a Lie derivation of L(V ) into L(M ). Then
(i) J(ψ) : V → M is a structural transformation with (Jψ)∗ = −Jψ′ where ψ′ = eθψθ.
(ii) If ψ is θ-invariant, then ψ′ = ψ and Jψ is a triple derivation.
(iii) If ψ is an inner derivation then Jψ is an inner structural transformation. In particular, if
ψ is a θ-invariant inner derivation then Jψ is an inner triple derivation.
Conversely, let ω be a structural transformation.
(iv) The mapping D = 1
2 L1(ω − ω∗) : L(V ) → L(M ) defined by
D(x, a b, y) =
1
2
(ω(x) − ω∗(x), ω(a)
b − a ω∗(b) − ω∗(a)
b + a ω(b), ω(y) − ω∗(y))
is a derivation of the Lie algebra L(V ) into L(M ).
(v) D is θ-invariant if and only if ω is a triple derivation, that is, ω∗ = −ω.
(vi) If ω is an inner structural transformation then D is an inner derivation. In particular, if ω
is an inner triple derivation then D is a θ-invariant inner derivation.
Proof. Let ψ be a Lie derivation of L(V ) into L(M ). We show first that
(4.5)
Jψ{abc} = {abJψ(c)} + {a, Jψ′(b), c} + {Jψ(a)bc}.
Let us define n : V → M , and n1 : V → M by the formulas ψ(0, 0, x) = (m(x), ϕ(x), n(x)), and
ψ(x, 0, 0) = (Jψ(x), ϕ1(x), n1(x)). Then
(Jψ{abc}, ϕ1{abc}, n1{abc}) = ψ({abc}, 0, 0)
= ψ[[(a, 0, 0), (0, 0, b)], (c, 0, 0)]
= [(a, 0, 0), (0, 0, b)] · ψ(c, 0, 0) − (c, 0, 0) · ψ[(a, 0, 0), (0, 0, b)]
= (0, a b, 0) · (Jψ(c), ϕ1(c), n1(c)) − (c, 0, 0) · ((a, 0, 0) · ψ(0, 0, b) − (0, 0, b) · ψ(a, 0, 0))
= ({abJψ(c)}, [a b, ϕ1(c)], −{ban1(c)})
−(c, 0, 0) · ((a, 0, 0) · (m(b), ϕ(b), n(b)) − (0, 0, b) · (Jψ(a), ϕ1(a), n1(a)))
= ({abJψ(c)}, [a b, ϕ1(c)], −{ban1(c)})
−(c, 0, 0) · (−ϕ(b)(a), a n(b), 0) − (c, 0, 0) · (0, −Jψ(a)
b, ϕ1(a)♮b)
= ({abJψ(c)}, [a b, ϕ1(c)], −{ban1(c)})
−(−{an(b)c}, 0, 0)) − ({Jψ(a)bc}, c ϕ1(a)♮b, 0).
Note that
(m(b), ϕ(b), n(b)) = ψ(0, 0, b)
= eθψ′θ(0, 0, b)
= eθψ′(b, 0, 0)
= eθ(Jψ′(b), ϕ′(b), n′(b))
= (n′(b), −ϕ′♮(b), Jψ′(b)),
so that n(b) = Jψ′(b), proving (4.5).
Applying (4.5) to ψ′ = eθψθ, we have, since ψ′′ = ψ
Jψ′{abc} = {abJψ′(c)} + {a, Jψ(b), c} + {Jψ′(a)bc}
(4.6)
proving (i).
COHOMOLOGY OF JORDAN TRIPLES VIA LIE ALGEBRAS
13
If ψ is θ-invariant, then ψ′ = ψ so that Jψ is a triple derivation, proving (ii). Example 4.3
provides a proof of (iii).
(iv) is immediate from Lemma 4.5 since ω − ω∗ is a triple derivation. The definitions show that
eθDθ = D if and only if ω = (ω −ω∗)/2, proving (v). Finally, if ω is an inner structural transformation,
then ω − ω∗ is an inner triple derivation, so that L1(ω − ω∗) is an inner derivation, proving (vi). (cid:3)
The following theorem provides some significant infinite dimensional examples of Lie algebras in
which every derivation is inner. Its proof is in the spirit of [28].
Theorem 4.7. Let V be a von Neumann algebra considered as a Jordan triple system with the
triple product {xyz} = (xy∗z + zy∗x)/2. Then every structural transformation on V is an inner
structural transformation. Hence, every derivation of the TKK Lie algebra L(V ) is inner.
Proof. Let S be a structural transformation on the von Neumann algebra V and to avoid
cumbersome notation, denote S ∗ by S. From the defining equations, S(1) = S(1)∗, and if S(1) = 0,
then S is a Jordan derivation.
For an arbitrary structural transformation S, write S = S0 + S1 where S0 = S − 1 S(1) is
therefore a Jordan derivation and S1 = 1 S(1) is an inner structural transformation. By the theorem
of Sinclair [33], S0 is a derivation and by the theorems of Kadison and Sakai, [18, 30], S0 is an inner
derivation, say S0(x) = ax − xa for some a ∈ V . By well known structure of the span of commutators
in von Neumann algebras due to Pearcy-Topping, Halmos, Halpern, Fack-de la Harpe, and others
(see [28] for the references), a = z +P[ci, di], where ci, di ∈ V and z belongs to the center of V . It
follows that
S0 = 2Xi
ci d∗
i − 2Xi
di
c∗
i
and is therefore also an inner structural transformation. The second statement follows from Proposi-
tion 4.6.
(cid:3)
We determine the structure of L(V ) when V is a finite von Neumann algebra in Corollary 5.7
below.
5. Examples
We conclude the paper with some examples of TKK Lie algebras and some Jordan triple cocycles.
Let us first note the following immediate consequences of our construction.
Theorem 5.1. Let V be a Jordan triple with TKK Lie algebra (L(V ), θ). If the k-th Lie coho-
mology group H k(L(V ), L(M )) vanishes, then H k(V, M ) = {0} and H k
θ (V, M ) = {0}.
We have noted the one-to-one correspondence between the triple derivations of a Jordan triple
V and the θ-invariant Lie derivations of the TKK Lie algebra (L(V ), θ), as well as the one-to-one
correspondence between the Jordan inner derivations of V and the Lie inner derivations of (L(V ), θ).
Corollary 5.2. Let V be a finite dimensional Jordan triple with semisimple TKK Lie algebra
L(V ). Then for any finite dimensional V -module M , we have H 1(V, M ) = H 2(V, M ) = {0}. In
particular, every triple derivation from V to M is inner.
Proof. This follows from Whitehead's lemmas H 1(L(V ), L(M )) = H 2(L(V ), L(M )) = {0}. (cid:3)
In fact, in the above corollary, we have H k(L(V ), L(M )) = {0} for all k ≥ 3 if L(M ) is a nontrivial
irreducible module over L(V ). We refer to [37] for a converse of this result.
14
CHO-HO CHU AND BERNARD RUSSO
5.1. Examples of cocycles. Let V be a Jordan triple with TKK Lie algebra (L(V ), θ). We
discuss examples of Jordan triple cocycles in Z k(V, M ), where M is a triple V -module, and compare
them with the Lie cocycles in Z k(L(V ), L(M )). We have shown in the previous section that the space
of Jordan triple derivations is exactly the space of 1-cocycles Z 1
θ (L(V ), L(M ))), where
the θ-invariant Lie 1-cocycles Z 1
θ (L(V ), L(M )) are exactly the θ-invariant Lie derivations from L(V )
to L(M ). We have also shown that B1
θ (L(V ), L(M ))) is the space of triple inner
derivations on V , coming from the θ-invariant Lie inner derivations B1
θ (V, M ) = J 1(B1
θ (V, M ) = J 1(Z 1
θ (L(V ), L(M )).
Examples of triple 2-cocycles can be constructed from Jordan restrictions of Lie 2-cocycles.
Example 5.3. If ω ∈ A2(V, M ) is extendable with L2(ω) ∈ Z 2(L(V ), L(M )), then ω = 0.
Proof. For x, y, z ∈ V ,
0 = d2L2(ω)((x, 0, 0), (y, 0, 0), (0, 0, z))
= (x, 0, 0) · (L(ω)((y, 0, 0), (0, 0, z) − (y, 0, 0) · (L(ω)((x, 0, 0), (0, 0, z)
+(0, 0, z) · (L(ω)((x, 0, 0), (y, 0, 0) − L(ω)([(x, 0, 0), (y, 0, 0], (0, 0, z))
+L(ω)([(x, 0, 0), (0, 0, z], (y, 0, 0)) − L(ω)([(y, 0, 0), (0, 0, z], (x, 0, 0))
= −(0, ω(x, y)
z, 0),
hence ω(x, y)
z = 0 for all x, y, z and ω = 0.
(cid:3)
Example 5.4. Let ϕ ∈ M0 be an inner triple derivation, and let b ∈ V . Define a linear map
ψ : L(V ) → L(M ) by
ψ(z) = [ [z, (0, ϕ, 0))] , (0, 0, b)]
(z ∈ L(V )).
Observe that ψ is not θ-invariant. Indeed, it can be seen readily that θψ(x, 0, 0) = (0, b ϕ(x), 0) while
ψ(θ(x, 0, 0)) = 0. Nevertheless d1ψ ∈ B2(L(V ), L(M )) and the triple 2-coboundary Jd1ψ ∈ B2(V, M )
is given by
Jd1ψ(x, y) = ιpd1ψ((x, 0, 0), (y, 0, 0))
= ιp((x, 0, 0) · ψ(y, 0, 0) − (y, 0, 0) · ψ(x, 0, 0))
= ιp((x, 0, 0) · (0, −ϕ(y)
= ιp(({ϕ(y), b, x}, 0, 0) − ({ϕ(x), b, y}, 0, 0))
= {ϕ(y), b, x} − {ϕ(x), b, y}
b, 0) − (y, 0, 0) · (0, −ϕ(x)
b, 0))
showing that B2(V, M ) 6= 0. We note that d1ψ is not θ-invariant since
eθd1ψ((x, 0, 0), (y, 0, 0)) = (0, 0, {ϕ(y), b, x}) − (0, 0, {ϕ(x), b, y}),
d1ψ((0, 0, x), (0, 0, y)) = (0, 0 − {b, ϕ(y), x} + {b, ϕ(x), y}).
Also Jd1ψ need not be extendable. Let V = M2(C) be the Jordan triple of 2 × 2 complex matrices.
Let v =(cid:18)1
0
0
0(cid:19) and u =(cid:18)0
0
0
1(cid:19). Then we have u v = 0 and one can find a, c ∈ V such that
To see this, let c = v. Then it suffices to find a ∈ V such that Jd1ψ(u, a) 2v 6= 0, where
Jd1ψ(u, a)
(v + c) + (u + a) Jd1ψ(v, c) 6= 0.
Jd1ψ(u, a) = {ϕ(u), b, a} − {ϕ(a), b, u}.
Let ϕ = m v − v m ∈ M0 where m = (cid:18) 1 −1
v(x) =(cid:26)(cid:26)(cid:18)0
Now let b = v. Then we have
1 (cid:19). Then we have ϕ(u) = −{v, m, u} = (cid:18)0
0(cid:19) ,(cid:18)1
0(cid:19) , a(cid:27) , v, x(cid:27) .
Jd1ψ(u, a)
−1
1
1
0
1
0(cid:19).
1
0
COHOMOLOGY OF JORDAN TRIPLES VIA LIE ALGEBRAS
15
Finally let a = v, then
Jd1ψ(u, a)
v(v) =
1
4(cid:18)0 1
1 0(cid:19) 6= 0.
We have seen in Example 5.3 that there are no non-zero extendable elements ω ∈ Z 2(V, M ) with
L2(ω) ∈ Z 2(L(V ), L(M )). The next example examines this phenomenon for extendable ω ∈ A3(V, M )
with L3(ω) ∈ Z 3
θ (L(V ), L(M )). We state it now as a theorem, in the statement of which, for a, b ∈ V
and m ∈ M , [a, b] denotes a b − b a and [m, a] denotes m a − a m. The proof is provided in
subsection 6.2.
Theorem 5.5. Let ω be an extendable element of A3(V, M ). Then its Lie extension L3(ω) is a
Lie 3-cocycle in A3
(5.1)
θ(k(V ), k(M )) if and only if ω satisfies the following three conditions:
[a, b]ω(x, y, z) = ω([a, b]x, y, z) + ω(x, [a, b]y, z) + ω(x, y, [a, b]z)
for all a, b, x, y, z ∈ V ;
(5.2)
[ω(a, b, c), d] = [ω(d, b, c), a] = [ω(a, b, d), c] = [ω(a, d, c), b]
for all a, b, c, d ∈ V ; and
(5.3)
[ω(x, y, [a, b]z), c] = 0.
for all x, y, z, a, b, c ∈ V .
5.2. Examples of TKK algebras. We begin with the following construction from [26, Chapter
12], which has its genesis in [21, pp. 809 -- 810]. Let A be a unital associative algebra with Lie product
the commutator [x, y] = xy − yx, Jordan product the anti-commutator x ◦ y = (xy + yx)/2 and Jordan
triple product {xyz} = (xyz + zyx)/2 (or {xyz} = (xy∗z + zy∗x)/2 if A has an involution). Denote
by Z(A) the center of A and by [A, A] the set of finite sums of commutators.
Proposition 5.6. Let A be a unital associative algebra with or without an involution considered
as a Jordan triple system. If Z(A) ∩ [A, A] = {0}, then the mapping (x, a b, y) 7→ (cid:20) ab
y −ba (cid:21) is
x
an isomorphism of the TKK Lie algebra L(A) onto the Lie subalgebra
(5.4)
(cid:26)(cid:20) u +P[vi, wi]
y
x
−u +P[vi, wi] (cid:21) : u, x, y, vi, wi ∈ A(cid:27)
of the Lie algebra M2(A) with the commutator product.
Corollary 5.7. Let V be a finite von Neumann algebra. Then L(V ) is isomorphic to the Lie
algebra [M2(V ), M2(V )].
Proof. The center valued trace of V is zero on [V, V ] and the identity on Z(V ), so the theorem
applies. Since M2(V ) is also a finite von Neumann algebra, [M2(V ), M2(V )] coincides with the
elements of M2(V ) of central trace zero (by [6, Theoreme 3.2]), so it remains to show that every such
element has the form (5.4). For this one can use the argument from [26, pp. 129 -- 130] as follows: if
(cid:20) a b
c
d (cid:21) ∈ M2(V ) has central trace zero, then tr (a) = −tr (d) and
−b′ + c′ (cid:21) ,
d (cid:21) =(cid:20) b′ + c′
(cid:20) a b
c
c
b
where c′ = (a + d)/2 and b′ = (a − d)/2.
(cid:3)
In a properly infinite von Neumann algebra, the assumption Z(A) ∩ [A, A] = {0} fails since
A = [A, A]. This assumption also fails in the Murray-von Neumann algebra of measurable operators
affiliated with a factor of type II1 ([34]). For a finite factor of type In, Corollary 5.7 states that the
classical Lie algebras sl(2n, C) of type A are TKK Lie algebras. Similarly, the TKK Lie algebra of
16
CHO-HO CHU AND BERNARD RUSSO
a Cartan factor of type 3 on an n-dimensional Hilbert space is the classical Lie algebra sp(2n, C) of
type C ([26, Theorem 3,p. 131]). More examples of TKK Lie algebras can be found in [5, 1.4] and
[22, Chapter III].
6. Proofs of Theorems 2.3 and 5.5
6.1. Proof of Theorem 2.3.
For the convenience of the reader, we repeat the statement of Theorem 2.3.
Theorem. Let V be a Jordan triple and let L(V ) be its TKK Lie algebra. Let M be a triple
V -module. Then L(M ) is a Lie L(V )-module.
For the proof, we are required to show that
(6.1)
[(a, h, b), (c, k, d)] · (m, ϕ, n) = (a, h, b) · ((c, k, d) · (m, ϕ, n))
− (c, k, d) · ((a, h, b)) · (m, ϕ, n)).
Let L denote the left side of (6.1). Then
(6.2)
L = (hc − ka
, [h, k] + a d − c
, k♮b − h♮d
) · (m, ϕ, n)
= (Hm − ϕA
, [H, ϕ] + A n − m B
{z }A
{z
L1
}
H
{z
L2
{z
b
}
}
{z
}
B
, ϕ♮B − H ♮n
).
L3
{z
}
We can assume that h = x y, k = u v so that
• A = {xyc} − {uva}
• H = {xyu} v − u {vxy} + a d − c
• B = {vub} − {yxd}.
b
Let R denote the right side of (6.1). Then
(6.3)
R = (a, h, b) · (km − ϕc
, [k, ϕ] + c n − m d
, ϕ♮d − k♮n
)
− (c, k, d) · (hm − ϕa
, [h, ϕ] + a n − m b
, ϕ♮b − h♮n
)
= (hC − Φa
Φ′
, [h, Φ] + a D − C b
, Φ♮b − h♮D
)
C
{z
{z
C ′
}
}
}
}
{z
{z
R′
2
R2
R1
{z
{z
R′
1
}
}
Φ
{z
{z
}
}
}
}
R3
{z
R3
{z
D
{z
{z
D′
}
}
− (kC ′ − Φ′c
, [k, Φ′] + c D′ − C ′ d
, Φ′♮ − k♮D′
)
= (R1 − R′
1, R2 − R′
2, R3 − R′
3).
As above, with h = x y, k = u v and with ϕ = w p + q
z, with p, q ∈ M , we have
• C = {uvm} − {wpc} − {qzc}
• D = {pwd} + {zqd} − {vun}
• Φ = {uvw} p − w {puv} + {uvq} z − q
• C ′ = {xym} − {wpa} − {qza}
• D′ = {pwb} + {zqb} − {yxn}
• Φ′ = {xyw} p − w {pxy} + {xyq} z − q
{zuv} + c n − m d
{zxy} + a n − m b
COHOMOLOGY OF JORDAN TRIPLES VIA LIE ALGEBRAS
17
We now show that L1 = R1 − R′
1. We have from (6.2)
(6.4)
L1 = Hm − ϕA
= {{xyu}vm} − {u, vxy, m} + {adm} − {cbm} − {wpA}
−{qzA} = {{xyu}vm}
− {u, vxy, m}
+ {adm}
− {wp{xyc}}
+ {wp{uva}}
− {qz{xyc}}
+ {qz{uva}}
}
1
{z
}
{z
6
}
1
{z
}
3
{z
{z }5
}
− {cbm}
{z }8
}
{z
7
2
{z
and from (6.3)
(6.5)
R1 = hC − Φa
= {xy{uvm}}
− {xy{wpc}}
− {xy{qzc}}
− {cna}
+ {mda}
− {{uvw}pa}
+ {w, puv, a}
− {{uvq}za}
+ {q, zuv, a}
}
}
2
{z
{z
6
}
}
3
{z
{z
7
}
{z }4
7
{z
{z }5
}
}
1
{z
{z
6
= {uv{xym}}
− {uv{wpa}}
− {uv{qza}}
− {{xyw}pc}
+ {w, pxy, c}
− {{xyq}zc}
+ {q, zxy, c}
1
{z
{z
2
}
}
6
{z
{z
3
}
}
7
{z
{z
3
}
}
2
− {anc}
{z
{z }4
+ {mbc}
}
{z }8
and
(6.6)
R′
1 = kC ′ − Φ′c
From (6.4)-(6.6), we have L1 = R1 − R′
1. In (6.4)-(6.6) we have indicated which terms cancel. To
see that the terms labeled 6 cancel, replace {uv{wpa}} by {{uvw}pa} − {w, vup, a} + {wp{uva}}.
Similarly, to see that the terms labeled 7 cancel, replace {uv{qza}} by {{uvq}za} − {q, vuz, a} +
{qz{uva}}.
We next show that L2 = R2 − R′
2. We have from (6.2)
(6.7)
L2 = [H, ϕ] + a n − m B
= [H, w p + q
z] + {xyc} n − {uva} n − m {vub} + m {yxd}
= [{xyu} v, w p] − [u {vxy}, w p] + [a d, w p] − [c
z] − [c
z] − [u {vxy}, q
+[{xyu} v, q
z] + [a d, q
b, w p]
z]
b, q
and from (6.3)
(6.8)
+{xyc} n − {uva} n − m {vub} + m {yxd}
R2 = [h, Φ] + a D − C b
= [x y, [u v, w p + q
z]] + [x y, c n] − [x y, m d]
+a {pwd} + a {zqd} − a {vun} − {uvm} b
+{wpc} b + {qzc} b
= [x y, {uvw} p] − [x y, w {puv}] + [x y, {uvq} z]
−[x y, q
−a {vun} − {uvm} b + {wpc} b + {qzc} b,
{zuv}] + [x y, c n] − [x y, m d] + a {pwd}
18
and
(6.9)
CHO-HO CHU AND BERNARD RUSSO
R′
2 = [k, Φ′] + c D′ − C ′ d
= [u v, [x y, w p + q
z]] + [u v, a n] − [u v, m b]
+c {pwd} + c {zqb} − c {yxn} − {xym} d
+{wpa} d + {qza} d
= [u v, {xyw} p] − [u v, w {pxy}] + [u v, {xyq} z]
−[u v, q
{zxy}] + [u v, a n] − [u v, m b] + c {pwb}
+c {zqb} − {xym} d + {wpa} d + {qza} d − c {yxn}
From (6.7) we have
(6.10)
L2 = {{xyu}vw} p − w {v, xyu, p} − {u, vxy, w} p
+w {{vxy}up} + {adw} p − w {dap} − {cbw} p
+w {bcp} + {{xyu}vq} z − q
{v, xyu, z} − {u, vxy, q} z
{{vxy}uz} + {adq} z − q
+q
+{xyc} n − {uva} n − m {vub} + m {yxd}
{daz} − {cbq} z + q
{bcz}
= ({{xyu}vw} − {u, vxy, w}
+ {adw}
− {cbw}
) p
19
+w (−{v, xyu, p} + {{vxy}up}
1
{z
3
{z
{z }9
2
{z
4
{z
{z }11
}
}
+ {bcp}
)
{z}6
z
−{dap}
19
− {cbq}
)
{z }6
{z }
{z }
}
{z}7
{z }20
}
{z }
{z }13
−{daz}
+ {yxd}
+ {bcz}
)
{z}7
{z }12
).
+({{xyu}vq} − {u, vxy, q}
+ {adq}
+q
(−{v, xyu, z} + {{vxy}uz}
+({xyc}
− {uva}
20
) n + m (− {vub}
From (6.8) we have
(6.11)
R2 = {xy{uvw}}
p − {uvw} {yxp}
− {xyw} {puv}
− {xyq} {zuv}
+q
{yx{zuv}}
+w {yx{puv}}
+ {xy{uvq}}
z − {uvq} {yxz}
1
{z
}
{z
2
18
− {xym}
{z
{z }
{z }8
14
{vun}
−a
}
}
15
{z
{z
3
{z }12
}
{z
4
+a
}
+ {xyc}
}
{z }11
{z }
{pwd}
19
+a
d + m {yxd}
− {uvm}
b + {wpc}
b + {qzc}
{z }5
{z }6
{z}7
16
{z
{z
17
n − c
}
}
{zqd}
{z }20
b.
{yxn}
{z }
10
}
}
+w {vu{pxy}}
+ {uv{xyq}}
z − {xyq} {vuz}
1
{z
}
{z
2
}
16
{z
{z
3
}
− {uvq} {yxz}
+q
{vu{yxz}}
+ {uva}
n − a
17
− {uvm}
{z
{z }5
{z }
− {xym}
b + m {vub}
{z
4
+c
{z }13
{pwb}
}
{z }6
{z }20
{z }
d + {wpa}
d + {qza}
d − c
{vun}
{z }8
.
{z }9
+c
15
{z
{z
18
}
}
{zqb}
{yxn}
{z}7
{z }
10
COHOMOLOGY OF JORDAN TRIPLES VIA LIE ALGEBRAS
19
From (6.9) we have
(6.12)
R′
2 = {uv{xyw}}
p − {xyw} {vup}
− {uvw} {pxy}
19
From (6.7)-(6.9), we have L2 = R2 − R′
2. In (6.7)-(6.9) we have indicated which terms cancel.
The terms labeled 1 -- 4 cancel by the main identity. The terms labeled 5 and 8-18 cancel in pairs. The
terms labeled 6,7,19,20 all cancel because of the following identity:
14
{abc} d − c {bad} = a {dcb} − {cda} b.
which follows from the main identity
{ab{cde}} − {cd{abe}} = {{abc}de} − {c, bad, e}
by interchanging (a, b) with (c, d) and noticing that the left side changes sign.
It remains to show that L3 = R3 − R′
3. We leave this as an exercise for the reader.
6.2. Proof of Theorem 5.5.
For the reader's convenience, we repeat the statement of Theorem 5.5, recalling that we write [a, b]
for a b − b a and [m, a] for m a − a m for a, b ∈ V and m ∈ M .
Theorem. Let ω be an extendable element of A3(V, M ). Then its Lie extension L3(ω) is a Lie
3-cocycle in A3
θ(k(V ), k(M )) if and only if ω satisfies the following three conditions:
(6.13)
[a, b]ω(x, y, z) = ω([a, b]x, y, z) + ω(x, [a, b]y, z) + ω(x, y, [a, b]z)
for all a, b, x, y, z ∈ V ;
(6.14)
[ω(a, b, c), d] = [ω(d, b, c), a] = [ω(a, b, d), c] = [ω(a, d, c), b]
for all a, b, c, d ∈ V ; and
(6.15)
[ω(x, y, [a, b]z), c] = 0.
for all x, y, z, a, b, c ∈ V .
Let ω ∈ A3(V, M ) be extendable and let ψ = d3L3(ω) (ψ is θ-invariant since 3 is odd). Write
aj, xj) ∈ k(V ) as Xj = (xj , 0, xj) + (0, [aj, bj], 0). By the alternating character
Xj = (xj , aj
of ψ, it is a Lie 3-cocycle, that is, ψ(X1, X2, X3, X4) = 0 for Xj ∈ k(V ), if and only if the following
five equations hold for ai, bi, xi ∈ V .
bj − bj
(6.16)
(6.17)
(6.18)
(6.19)
(6.20)
ψ((x1, 0, x1), (x2, 0, x2), (x3, 0, x3), (x4, 0, x4)) = 0,
(4 variables)
ψ((x1, 0, x1), (x2, 0, x2), (x3, 0, x3), (0, [a4, b4], 0)) = 0,
(5 variables)
ψ((x1, 0, x1), (x2, 0, x2), (0, [a3, b3], 0), (0, [a4, b4], 0)) = 0,
(6 variables)
ψ((x1, 0, x1), (0, [a2, b2], 0), (0, [a3, b3], 0), (0, [a4, b4], 0)) = 0,
(7 variables)
ψ((0, [a1, b1], 0), (0, [a2, b2], 0), (0, [a3, b3], 0), (0, [a4, b4], 0)) = 0.
(8 variables)
20
CHO-HO CHU AND BERNARD RUSSO
Note that (6.13)-(6.15) involve 5,4 and 6 variables respectively, so there is an additional amount
of redundancy in (6.16)-(6.20). We shall begin by showing that (6.16)-(6.20) imply (6.13)-(6.15).
Straightforward calculation of (6.16), using (3.1) and (4.1), shows that it is equivalent to
(6.21)
[x1, ω(x2, x3, x4)] − [x2, ω(x1, x3, x4)]
+[x3, ω(x1, x2, x4)] − [x4, ω(x1, x2, x3)] = 0.
We shall see shortly that (6.21) is redundant since it will follow from the identity (6.14), which will
be proved using (6.18). However, (6.21) will be used later, in the proof that (6.13)-(6.15) imply
(6.16)-(6.20).
Similarly, (6.17) is equivalent to
−{a4b4ω(x1, x2, x3)} + {b4a4ω(x1, x2, x3)}
+ω({a4b4x1}, x2, x3) − ω({b4a4x1}, x2, x3)
−ω({a4b4x2}, x1, x3) + ω({b4a4x2}, x1, x3)
+ω({a4b4x3}, x1, x2) − ω({b4a4x3}, x1, x2) = 0.
which can be rewritten as
(6.22)
[a4, b4](ω(x1, x2, x3)) = ω([a4, b4]x1, x2, x3)
+ω(x1, [a4, b4]x2, x3) + ω(x1, x2, [a4, b4]x3),
proving (6.13) (assuming only (6.17)).
An interpretation of (6.22) is that the inner triple derivation [a, b] (for the triple product {·, ·, ·}
of V ) is also a "triple derivation" for the (ad hoc M -valued) triple product (x, y, z) 7→ ω(x, y, z) of V .
In order to proceed efficiently, it is convenient to state the following formulas. First, for ai and
bi in V , by (4.1),
(6.23)
where
(6.24)
L3(ω)((0, [a1, b1], 0), (0, [a2, b2], 0), (0, [a3, b3], 0)) = (0, Λ, 0)
Λ = [ω(a1, a2, a3), b1 + b2 + b3] − [ω(b1, a2, a3), a1 + b2 + b3]
+ [ω(b1, b2, a3), a1 + a2 + b3] − [ω(b1, b2, b3), a1 + a2 + a3]
+ [ω(b1, a2, b3), a1 + b2 + a3] + [ω(a1, b2, b3), b1 + a2 + a3]
− [ω(a1, b2, a3), b1 + a2 + b3] − [ω(a1, a2, b3), b1 + b2 + a3].
Second, for a, b, c ∈ V and m ∈ M , by (2.4),
(6.25)
(0, [a, b], 0) · (0, [m, c], 0) = (0, [[a, b]m, c] + [m, [a, b]c], 0),
and, for ai and bi in V , by (2.1),
(6.26)
[(0, [a1, b1], 0), (0, [a2, b2], 0)] = (0, [[a1, b1]a2, b2] + [[b1, a1]b2, a2], 0).
Returning to (6.18)-(6.20) and observing that
a straightforward calculation of (6.18) shows that it is equivalent to
L3(ω)((∗, 0, ∗), (0, ∗, 0), (∗, ∗, ∗)) = 0,
0 = −L3(ω)((0, [x1, x2], 0), (0, [a3, b3], 0), (0, [a4, b4], 0)),
which by (6.23) and (6.24) is equivalent to
(6.27)
0 = [ω(x1, a3, a4), x2 + b3 + b4] − [ω(x2, a3, a4), x1 + b3 + b4]
+ [ω(x2, b3, a4), x1 + a3 + b4] − [ω(x2, b3, b4), x1 + a3 + a4]
+ [ω(x2, a3, b4), x1 + b3 + a4] + [ω(x1, b3, b4), x2 + a3 + a4]
− [ω(x1, b3, a4), x2 + a3 + b4] − [ω(x1, a3, b4), x2 + b3 + a4].
COHOMOLOGY OF JORDAN TRIPLES VIA LIE ALGEBRAS
21
We shall now see that (6.27) simplifies considerably and gives the same information as (6.19), namely
(6.27) is equivalent to
(6.28)
[ω(a, b, c), d] = [ω(d, b, c), a] = [ω(a, b, d), c] = [ω(a, d, c), b],
which is (6.14). Assuming that this has been done, we will have proved that (6.17) is equivalent to
(6.13); and that (6.18), (6.19) and (6.14) are equivalent. We shall complete the proof by showing
that (6.20), together with (6.13) and (6.14), implies (6.15); and then proving that (6.13)-(6.15) imply
(6.16)-(6.20).
Note that (6.13), (6.15) and the alternating character of ω imply [[a, b]ω(x, y, z), c] = 0, and that
(6.14) and (6.15) imply
(6.29)
[ω(x, y, z), [a, b]c] = 0.
We continue the proof of Theorem 5.5 by showing that (6.28) follows from (6.27) and that (6.19) does
not contribute any new properties of ω. After that, we shall deal with (6.20).
Since we are assuming (6.18), we may set x1 = 0 and a3 = 0 in (6.27). The result is
[ω(x2, b3, a4), b4] − [ω(x2, b3, b4), a4] = 0.
If one repeats this process with (x1 = 0 and) a3 = 0 replaced successivly by a4 = 0, b3 = 0, b4 = 0, one
obtains three more such equations. Next, replace x1 = 0 by x2 = 0 to obtain four more such equations.
Finally, setting a3 = 0 and a4 = 0 in (6.27), and repeating with (a3, a4) replaced successively with
(a3, b4), (b3, a4), (b3, b4) results in four more such equations. By changing the names of the variables,
the resulting twelve equations reduce to (6.28) (which is (6.14)).
We next show that (6.19) yields the same information as (6.18). Straightforward calculation of
(6.19) shows that it is equivalent to
0 = (x1, 0, x1) · L3(ω)((0, [a2, b2], 0), (0, [a3, b3], 0), (0, [a4, b4], 0)),
which by (6.23) equals (x1, 0, x1) · (0, Λ, 0) = (−Λx1, 0, −Λx1) where
(6.30)
Λ = [ω(a2, a3, a4), b2 + b3 + b4] − [ω(b2, a3, a4), a2 + b3 + b4]
+ [ω(b2, b3, a4), a2 + a3 + b4] − [ω(b2, b3, b4), a2 + a3 + a4]
+ [ω(b2, a3, b4), a2 + b3 + a4] + [ω(a2, b3, b4), b2 + a3 + a4]
− [ω(a2, b3, a4), b2 + a3 + b4] − [ω(a2, a3, b4), b2 + b3 + a4],
Thus, (6.19) results in
(6.31)
Λx1 = 0.
where Λ is given by (6.30). Comparing this with (6.27) shows that (6.19) is equivalent to (6.18).
We now have that (6.18), (6.19), (6.27), (6.28) and (6.14) are equivalent, and that (6.17) and
(6.13) are equivalent. It remains, for this part of the proof, to establish (6.15) using (6.16)-(6.20).
This will take some perseverance!
In order to process (6.20) we shall adopt the following self-explanatory notation. For distinct
elements i, j, k, l ∈ {1, 2, 3, 4}, set
(6.32)
and
(6.33)
ijkl1 = (0, [ai, bi], 0) · L3(ω)((0, [aj, bj], 0), (0, [ak, bk], 0), (0, [al, bl], 0))
ijkl2 = L3(ω)([(0, [ai, bi], 0), (0, [aj, bj], 0)], (0, [ak, bk], 0), (0, [al, bl], 0)).
Then equation (6.20) for ψ = d3L3(ω) is restated as:
(6.34)
0 = 12341 − 21341 + 31241 − 41231
−12342 + 13242 − 14232 − 23142 + 24132 − 34122.
By (6.32), using (6.23)-(6.24),
ijkl1 = (0, [ai, bi], 0) · (0, Λj,k,l, 0)
22
where
(6.35)
and by (6.25),
CHO-HO CHU AND BERNARD RUSSO
Λj,k,l = [ω(aj, ak, al), bj + bk + bl] − [ω(bj, ak, al), aj + bk + bl]
+ [ω(bj, bk, al), aj + ak + bl] − [ω(bj, bk, bl), aj + ak + al]
+ [ω(bj, ak, bl), aj + bk + al] + [ω(aj, bk, bl), bj + ak + al]
− [ω(aj, bk, al), bj + ak + bl] − [ω(aj, ak, bl), bj + bk + al],
(6.36)
where
(6.37)
ijkl1 = (0, Γi,j,k,l, 0)
Γi,j,k,l = [[ai, bi]ω(aj, ak, al), bj + bk + bl] + [ω(aj, ak, al), [ai, bi](bj + bk + bl)]
− [[ai, bi]ω(bj, ak, al), aj + bk + bl] − [ω(bj, ak, al), [ai, bi](aj + bk + bl)]
+ [[ai, bi]ω(bj, bk, al), aj + ak + bl] + [ω(bj, bk, al), [ai, bi](aj + ak + bl)]
− [[ai, bi]ω(bj, bk, bl), aj + ak + al] − [ω(bj, bk, bl), [ai, bi](aj + ak + al)]
+ [[ai, bi]ω(bj, ak, bl), aj + bk + al] + [ω(bj, ak, bl), [ai, bi](aj + bk + al)]
+ [[ai, bi]ω(aj, bk, bl), bj + ak + al] + [ω(aj, bk, bl), [ai, bi](bj + ak + al)]
− [[ai, bi]ω(aj, bk, al), bj + ak + bl] − [ω(aj, bk, al), [ai, bi](bj + ak + bl)]
− [[ai, bi]ω(aj, ak, bl), bj + bk + al] − [ω(aj, ak, bl), [ai, bi](bj + bk + al)].
By (6.33), using (6.26) and (6.23)-(6.24),
(6.38)
where
(6.39)
ijkl2 = L3(ω)((0, [[ai, bi]aj, bj] + [[bi, ai]bj, aj], 0), (0, [ak, bk], 0), (0, [al, bl], 0))
= (0, ∆i,j,k,l, 0),
∆i,j,k,l = [ω([ai, bi]aj, ak, al), bj + bk + bl] + [ω([bi, ai]bj, ak, al), aj + bk + bl)]
− [ω(bj, ak, al), [ai, bi]aj + bk + bl] − [ω(aj, ak, al), [bi, ai](bj + bk + bl)]
+ [ω(bj, bk, al), [ai, bi]aj + ak + bl] + [ω(aj, bk, al), [bi, ai](bj + ak + bl)]
− [ω(bj, bk, bl), [ai, bi]aj + ak + al] − [ω(aj, bk, bl), [bi, ai](bj + ak + al)]
+ [ω(bj, ak, bl), [ai, bi]aj + bk + al] + [ω(aj, ak, bl), [bi, ai](bj + bk + al)]
+ [ω([ai, bi]aj, bk, bl), bj + ak + al] + [ω([bi, ai]bj, bk, bl), aj + ak + al]
− [ω([ai, bi]aj, bk, al), bj + ak + bl] − [ω([bi, ai]bj, bk, al), aj + ak + bl]
− [ω([ai, bi]aj, ak, bl), bj + bk + al] − [ω([bi, ai]bj, ak, bl), aj + bk + al].
We next analyze (6.37) and (6.39). First, applying (6.13) to the first bracket on each line of (6.37)
and applying (6.14) to the expansion of those brackets results in 72 terms, 24 of which cancel with
all of the terms in the second bracket on each line of (6.37). Thus the 96 terms in (6.37) are reduced
to the 48 terms in
(6.40)
Γi,j,k,l =
COHOMOLOGY OF JORDAN TRIPLES VIA LIE ALGEBRAS
23
[ω([ai, bi]aj, ak, al), (bk + bl)] + [ω(aj, [ai, bi]ak, al), (bj + bl)] + [ω(aj, ak, [ai, bi]al), (bj + bk)]
− [ω([ai, bi]bj, ak, al), (bk + bl)] − [ω(bj, [ai, bi]ak, al), (aj + bl)] − [ω(bj, ak, [ai, bi]al), (aj + bk)]
+ [ω([ai, bi]bj, bk, al), (ak + bl)] + [ω(bj, [ai, bi]bk, al), (aj + bl)] + [ω(bj, bk, [ai, bi]al), (aj + ak)]
− [ω([ai, bi]bj, bk, bl), (ak + al)] − [ω(bj, [ai, bi]bk, bl), (aj + al)] − [ω(bj, bk, [ai, bi]bl), (aj + ak)]
+ [ω([ai, bi]bj, ak, bl), (bk + al)] + [ω(bj, [ai, bi]ak, bl), (aj + al)] + [ω(bj, ak, [ai, bi]bl), (aj + bk)]
− [ω([ai, bi]aj, bk, bl), (ak + al)] − [ω(aj, [ai, bi]bk, bl), (bj + al)] − [ω(aj, bk, [ai, bi]bl), (bj + ak)]
+ [ω([ai, bi]aj, bk, al), (ak + bl)] + [ω(aj, [ai, bi]bk, al), (bj + bl)] + [ω(aj, bk, [ai, bi]al), (bj + ak)]
− [ω([ai, bi]aj, ak, bl), (bk + al)] − [ω(aj, [ai, bi]ak, bl), (bj + al)] − [ω(aj, ak, [ai, bi]bl), (bj + bk)].
Second, the 8 first brackets on the lines of (6.39) sum to zero, as can be seen by expanding and
noting that the resulting terms cancel in pairs by applying (6.14). Thus (6.39) reduces (initially) to
the sum of the 8 second brackets on the lines of (6.39), namely,
(6.41)
∆i,j,k,l = [ω([bi, ai]bj, ak, al), aj + bk + bl)]
− [ω(aj, ak, al), [bi, ai](bj + bk + bl)]
+ [ω(aj, bk, al), [bi, ai](bj + ak + bl)]
− [ω(aj, bk, bl), [bi, ai](bj + ak + al)]
+ [ω(aj, ak, bl), [bi, ai](bj + bk + al)]
+ [ω([bi, ai]bj, bk, bl), aj + ak + al]
− [ω([bi, ai]bj, bk, al), aj + ak + bl]
− [ω([bi, ai]bj, ak, bl), aj + bk + al].
However, there is still more cancellation in (6.41) using (6.14), and what remains is
(6.42)
∆i,j,k,l = −[ω(aj, ak, al), [bi, ai](bk + bl)]
+[ω(aj, bk, al), [bi, ai](ak + bl)]
−[ω(aj, bk, bl), [bi, ai](ak + al)]
+[ω(aj, ak, bl), [bi, ai](bk + al)]
.
The equation (6.34) is thus equivalent to
(6.43)
0 = Γ1234 − Γ2134 + Γ3124 − Γ4123
−∆1234 + ∆1324 − ∆1423 − ∆2314 + ∆2413 − ∆3412,
where Γijkl and ∆ijkl are given by (6.40) and (6.42).
We are now going to decompose each term in (6.43) into "irreducible pieces" as follows. First
some notation. Let Σ denote the right side of (6.43), let Γijkl(a1 = 0) denote the sum of the terms
of Γijkl which do not involve the variable a1, and Γijkl(a1 6= 0) the sum of the terms of Γijkl which
contain the variable a1, with similar notation for other variables, for more then one variable, and for
∆ijkl. With Σ(a1 = 0) denoting the sum of the terms of Σ not containing a1, etc., we have (and
this is the first of two underlying principles in what follows) Σ = 0 if and only if Σ(a1 = 0) = 0 and
Σ(a1 6= 0) = 0.
We shall use (6.40) to process the Γijkl in (6.43) and in parallel use (6.42) to process the ∆ijkl in
(6.43). Here we go! By (6.40),
(6.44)
(6.45)
Γi,j,k,l(ai = 0) = 0,
Γi,j,k,l(aj = 0) =
24
CHO-HO CHU AND BERNARD RUSSO
− [ω([ai, bi]bj, ak, al), (bk + bl)] − [ω(bj, [ai, bi]ak, al), bl] − [ω(bj, ak, [ai, bi]al), bk]
+ [ω([ai, bi]bj, bk, al), (ak + bl)] + [ω(bj, [ai, bi]bk, al), bl] + [ω(bj, bk, [ai, bi]al), ak]
− [ω([ai, bi]bj, bk, bl), (ak + al)] − [ω(bj, [ai, bi]bk, bl), al] − [ω(bj, bk, [ai, bi]bl), ak]
+ [ω([ai, bi]bj, ak, bl), (bk + al)] + [ω(bj, [ai, bi]ak, bl), al] + [ω(bj, ak, [ai, bi]bl), bk],
(6.46)
Γi,j,k,l(ak = 0) =
[ω([ai, bi]bj, bk, al), bl] + [ω(bj, [ai, bi]bk, al), (aj + bl)] + [ω(bj, bk, [ai, bi]al), aj]
− [ω([ai, bi]bj, bk, bl), al] − [ω(bj, [ai, bi]bk, bl), (aj + al)] − [ω(bj, bk, [ai, bi]bl), aj]
− [ω([ai, bi]aj, bk, bl), al] − [ω(aj, [ai, bi]bk, bl), (bj + al)] − [ω(aj, bk, [ai, bi]bl), bj]
+ [ω([ai, bi]aj, bk, al), bl] + [ω(aj, [ai, bi]bk, al), (bj + bl)] + [ω(aj, bk, [ai, bi]al), bj],
and
(6.47)
Γi,j,k,l(al = 0) =
− [ω([ai, bi]bj, bk, bl), ak] − [ω(bj, [ai, bi]bk, bl), aj] − [ω(bj, bk, [ai, bi]bl), (aj + ak)]
+ [ω([ai, bi]bj, ak, bl), bk] + [ω(bj, [ai, bi]ak, bl), aj] + [ω(bj, ak, [ai, bi]bl), (aj + bk)]
− [ω([ai, bi]aj, bk, bl), ak] − [ω(aj, [ai, bi]bk, bl), bj] − [ω(aj, bk, [ai, bi]bl), (bj + ak)]
− [ω([ai, bi]aj, ak, bl), bk] − [ω(aj, [ai, bi]ak, bl), bj] − [ω(aj, ak, [ai, bi]bl), (bj + bk)].
On the other hand, by (6.42),
(6.48)
(6.49)
(6.50)
and
(6.51)
∆i,j,k,l(ai = 0) = 0,
∆i,j,k,l(aj = 0) = 0,
∆i,j,k,l(ak = 0) = [ω(aj, bk, al), [bi, ai]bl]
− [ω(aj, bk, bl), [bi, ai]al],
∆i,j,k,l(al = 0) = −[ω(aj, bk, bl), [bi, ai]ak]
+[ω(aj, ak, bl), [bi, ai]bk].
Returning to (6.40), by (6.44)
(6.52)
By (6.45)
(6.53)
Γ1234(a1 = 0) = 0.
Γ2134(a1 = 0) =
− [ω([a2, b2]b1, a3, a4), (b3 + b4)] − [ω(b1, [a2, b2]a3, a4), b4] − [ω(b1, a3, [a2, b2]a4), b3]
+ [ω([a2, b2]b1, b3, a4), (a3 + b4)] + [ω(b1, [a2, b2]b3, a4), b4] + [ω(b1, b3, [a2, b2]a4), a3]
− [ω([a2, b2]b1, b3, b4), (a3 + a4)] − [ω(b1, [a2, b2]b3, b4), a4] − [ω(b1, b3, [a2, b2]b4), a3]
+ [ω([a2, b2]b1, a3, b4), (b3 + a4)] + [ω(b1, [a2, b2]a3, b4), a4] + [ω(b1, a3, [a2, b2]b4), b3],
(6.54)
Γ3124(a1 = 0) =
− [ω([a3, b3]b1, a2, a4), (b2 + b4)] − [ω(b1, [a3, b3]a2, a4), b4] − [ω(b1, a2, [a3, b3]a4), b2]
+ [ω([a3, b3]b1, b2, a4), (a2 + b4)] + [ω(b1, [a3, b3]b2, a4), b4] + [ω(b1, b2, [a3, b3]a4), a2]
− [ω([a3, b3]b1, b2, b4), (a2 + a4)] − [ω(b1, [a3, b3]b2, b4), a4] − [ω(b1, b2, [a3, b3]b4), a2]
+ [ω([a3, b3]b1, a2, b4), (b2 + a4)] + [ω(b1, [a3, b3]a2, b4), a4] + [ω(b1, a2, [a3, b3]b4), b2],
COHOMOLOGY OF JORDAN TRIPLES VIA LIE ALGEBRAS
25
and
(6.55)
Γ4123(a1 = 0) =
− [ω([a4, b4]b1, a2, a3), (b2 + b3)] − [ω(b1, [a4, b4]a2, a3), b3] − [ω(b1, a2, [a4, b4]a3), b2]
+ [ω([a4, b4]b1, b2, a3), (a2 + b3)] + [ω(b1, [a4, b4]b2, a3), b3] + [ω(b1, b2, [a4, b4]a3), a2]
− [ω([a4, b4]b1, b2, b3), (a2 + a3)] − [ω(b1, [a4, b4]b2, b3), a3] − [ω(b1, b2, [a4, b4]b3), a2]
+ [ω([a4, b4]b1, a2, b3), (b2 + a3)] + [ω(b1, [a4, b4]a2, b3), a3] + [ω(b1, a2, [a4, b4]b3), b2].
On the other hand, by (6.48)
(6.56)
By (6.50),
(6.57)
(6.58)
and
(6.59)
∆1234(a1 = 0) = 0, ∆1324(a1 = 0) = 0, ∆1423(a1 = 0) = 0.
∆2314(a1 = 0) = [ω(a3, b1, a4), [b2, a2]b4]
− [ω(a3, b1, b4), [b2, a2]a4],
∆2413(a1 = 0) = [ω(a4, b1, a3), [b2, a2]b3]
− [ω(a4, b1, b3), [b2, a2]a3],
∆2413(a1 = 0) = [ω(a4, b1, a2), [b3, a3]b2]
− [ω(a4, b1, b2), [b3, a3]a2].
By (6.43), and (6.52)-(6.59),
0 = Σ(a1 = 0) = −(6.53) + (6.54) − (6.55) − (6.57) + (6.58) − (6.59),
and each of the terms on the right side must be decomposed further. Here, we are using the notation
(6.53) to denote Γ2134(a1 = 0) and similarly for (6.54), etc.
We shall analyze (6.53) first. By (6.53),
(6.60)
and
(6.61)
Γ2134(a1 = 0, a3 = 0) =
+ [ω([a2, b2]b1, b3, a4), b4] + [ω(b1, [a2, b2]b3, a4), b4]
− [ω([a2, b2]b1, b3, b4), a4] − [ω(b1, [a2, b2]b3, b4), a4]
Γ2134(a1 = 0, a3 6= 0) =
− [ω([a2, b2]b1, a3, a4), (b3 + b4)] − [ω(b1, [a2, b2]a3, a4), b4] − [ω(b1, a3, [a2, b2]a4), b3]
+ [ω([a2, b2]b1, b3, a4), a3] + [ω(b1, b3, [a2, b2]a4), a3]
− [ω([a2, b2]b1, b3, b4), a3] − [ω(b1, b3, [a2, b2]b4), a3]
+ [ω([a2, b2]b1, a3, b4), (b3 + a4)] + [ω(b1, [a2, b2]a3, b4), a4] + [ω(b1, a3, [a2, b2]b4), b3].
The identity given by (6.60) is "irreducible" in the sense that if any of its variables is zero, then it
vanishes identically (This is the second of the two underlying principles mentioned earlier). However,
since it is a consequence of (6.14), it does not give any new identities and can be ignored. We proceed
to decompose (6.61) as follows.
(6.62)
(6.63)
Γ2134(a1 = 0, a3 6= 0, a4 = 0) =
− [ω([a2, b2]b1, b3, b4), a3] − [ω(b1, b3, [a2, b2]b4), a3]
+ [ω([a2, b2]b1, a3, b4), b3] + [ω(b1, a3, [a2, b2]b4), b3].
Γ2134(a1 = 0, a3 6= 0, a4 6= 0) =
26
CHO-HO CHU AND BERNARD RUSSO
− [ω([a2, b2]b1, a3, a4), (b3 + b4)] − [ω(b1, [a2, b2]a3, a4), b4] − [ω(b1, a3, [a2, b2]a4), b3]
+ [ω([a2, b2]b1, b3, a4), a3] + [ω(b1, b3, [a2, b2]a4), a3]
+ [ω([a2, b2]b1, a3, b4), a4] + [ω(b1, [a2, b2]a3, b4), a4]
The identity given by (6.62) is irreducible and can also be ignored, so we proceed to decompose (6.63)
as follows.
(6.64)
Γ2134(a1 = 0, a3 6= 0, a4 6= 0, b3 = 0) =
and
(6.65)
− [ω([a2, b2]b1, a3, a4), b4] − [ω(b1, [a2, b2]a3, a4), b4]
+ [ω(b1, b3, [a2, b2]a4), a3]
+ [ω([a2, b2]b1, a3, b4), a4] + [ω(b1, [a2, b2]a3, b4), a4],
Γ2134(a1 = 0, a3 6= 0, a4 6= 0, b3 6= 0) =
− [ω([a2, b2]b1, a3, a4), b3]
+ [ω([a2, b2]b1, b3, a4), a3] + [ω(b1, b3, [a2, b2]a4), a3].
By using (6.14), each of (6.64) and (6.65) gives the new identity
(6.66)
[ω(b1, b3, [a2, b2]a4), a3] = 0,
which establishes (6.15), and at the same time shows that (6.57), (6.58) and (6.59) produce no new
identities.
This completes the analysis of (6.53), which has produced (6.15). Since (6.54) is obtained from
(6.53) by interchanging the indices 3 and 2, no new information is provided by (6.54). Similarly, since
(6.55) is obtained from (6.54) by interchanging the indices 3 and 4, no new information is provided
by (6.55). Thus we have found all irreducible expressions which sum to Σ(a1 = 0), resulting in only
one identity, namely (6.15). This completes the proof that (6.16)-(6.20) imply (6.13)-(6.15). (See the
paragraph following (6.43).)
It is now a simple matter to prove that, conversely, (6.13)-(6.15) imply (6.16)-(6.20). Note that
by (6.15), (6.29), and (6.37),(6.39), Γijkl and ∆ijkl vanish, showing that Σ(a1 6= 0) = 0, hence
(6.13)-(6.15) imply (6.20). Since earlier arguments have shown that
• (6.14) ⇒ (6.21) ⇔ (6.16),
• (6.13) = (6.22) ⇔ (6.17),
• (6.14) ⇔ (6.27) ⇔ (6.18)⇔ (6.19),
this completes the proof that (6.13)-(6.15) imply (6.16)-(6.20), and hence the proof of Theorem 5.5.
References
[1] R. Carlsson, Cohomology of associative triple systems. Proc. Amer. Math. Soc. 60 (1976), 1 -- 7. Erratum and
supplement: Proc. Amer. Math. Soc. 67, no. 2 (1977), 361.
[2] R. Carlsson, Der Wedderburnsche Hauptsatz fur alternative Tripelsysteme und Paare. Math. Ann. 228, no. 3
(1977), 233 -- 248.
[3] H. Cartan and S. Eilenberg, Homological algebra. Princeton University Press, Princeton, N. J., 1956. xv+390 pp.
[4] Chevalley, Claude; Eilenberg, Samuel Cohomology theory of Lie groups and Lie algebras. Trans. Amer. Math. Soc.
63, (1948). 85 -- 124.
[5] C-H. Chu, Jordan structures in geometry and analysis. Cambridge Tracts in Math. 190, Cambridge Univ. Press,
Cambridge, 2012.
[6] Th. Fack and P. de la Harpe, Sommes de commutateurs dans les algbres de von Neumann finies continues. Ann.
Inst. Fourier (Grenoble) 30, no. 3 (1980) 49 -- 73.
[7] D. B. Fuks, Cohomology of infinite-dimensional Lie algebras. Translated from the Russian by A. B. Sosinski.
Contemporary Soviet Mathematics. Consultants Bureau, New York, 1986. xii+339 pp.
[8] M. Gerstenhaber, A uniform cohomology theory for algebras. Proc. Nat. Acad. Sci. U.S.A. 51 (1964), 626 -- 629.
[9] N. D. Glassman, Cohomology of Jordan algebras. J. Algebra 15 (1970), 167 -- 194.
COHOMOLOGY OF JORDAN TRIPLES VIA LIE ALGEBRAS
27
[10] N. D. Glassman, Cohomology of nonassociative algebras. Pacific J. Math. 33 (1970), 617 -- 634.
[11] B. Harris, Cohomology of Lie triple systems and Lie algebras with involution. Trans. Amer. Math. Soc. 98 (1961),
148 -- 162.
[12] G. Hochschild, On the cohomology groups of an associative algebra. Ann. of Math. (2) 46 (1945), 58 -- 67.
[13] T. L. Hodge and B. J. Parshall, On the representation theory of Lie triple systems. Trans. Amer. Math. Soc. 354,
no. 11 (2002), 4359 -- 4391.
[14] N. Jacobson, General representation theory of Jordan algebras. Trans. Amer. Math. Soc. 70 (1951), 509 -- 530.
[15] N. Jacobson, Jordan algebras. 1957 Report of a conference on linear algebras, June, 1956 pp 12 -- 19 National
Academy of Sciences National Research Council, Washington, Public. 502
[16] N. Jacobson, Lie algebras. Interscience Tracts in Pure and Applied Mathematics, No. 10 Interscience Publishers
(a division of John Wiley & Sons), New York-London 1962 ix+331 pp.
[17] N. Jacobson, Structure and representations of Jordan algebras. American Mathematical Society Colloquium Pub-
lications, Vol. XXXIX American Mathematical Society, Providence, R.I. 1968 x+453 pp.
[18] R, V. Kadison, Derivations of operator algebras. Ann. of Math. (2) 83 (1966), 280 -- 293.
[19] R. V. Kadison, Which Singer is that? Surveys in differential geometry, 347 -- 373, Surv. Differ. Geom., VII, Int.
Press, Somerville, MA, 2000.
[20] A. W. Knapp, Lie groups, Lie algebras, and cohomology. Mathematical Notes, 34. Princeton University Press,
Princeton, NJ, 1988. xii+510 pp.
[21] M. Koecher, Imbedding of Jordan algebras in Lie algebras. I, Amer. J. Math 89, no. 3 (1967), 787 -- 816.
[22] M. Koecher, An elementary approach to bounded symmetric domains, Lecture Notes, Rice University, 1969.
[23] O. Kuhn and A. Rosendahl, Wedderburnzerlegung fur Jordan-Paare. (German. English summary) Manuscripta
Math. 24 (1978), no. 4, 403 -- 435.
[24] K. McCrimmon, Representations of quadratic Jordan algebras. Trans. Amer. Math. Soc. 153 (1971), 279 -- 305.
[25] K. McCrimmon, Compatible Peirce decompositions of Jordan triple systems. Pacific J. Math. 103 (1982), no. 1,
57 -- 102.
[26] K. Meyberg, Lectures on algebras and triple systems. Notes on a course of lectures given during the academic year
1971 -- 1972. The University of Virginia, Charlottesville, Va., 1972. v+226 pp.
[27] A. J. Penico, The Wedderburn principal theorem for Jordan algebras. Trans. Amer. Math. Soc. 70 (1951). 404 -- 420.
[28] R. Pluta and B. Russo, Triple derivations on von Neumann algebras. Preprint 2014, arXiv:1309.3526
[29] B. Russo, Derivations and Projections on Jordan triples. Nonassociative algebra, continuous cohomology and
quantum functional analysis, Proceedings of V CIDAMA, Almeria, Spain, September 12-16, 2011. World Scientific,
to appear.
[30] S. Sakai, Derivations of W ∗-algebras. Ann. of Math. (2) 83 (1966), 273 -- 279.
[31] P. Seibt, Cohomology of algebras and triple systems. Comm. Algebra 3 (1975), no. 12, 1097 -- 1120.
[32] P. Seibt, Review of [1], Mathematical Reviews MR0430026.
[33] A. M. Sinclair, Jordan homomorphisms and derivations on semisimple Banach algebras. Proc. Amer. Math. Soc.
24 (1970), 209 -- 214.
[34] A. Thom, A note on commutators in the Murray-von Neumann algebra., Preprint 2013.
[35] C. A. Weibel, An introduction to homological algebra. Cambridge Studies in Advanced Mathematics, 38. Cambridge
University Press, Cambridge, 1994. xiv+450 pp.
[36] C. A. Weibel, History of homological algebra. History of topology, 797 -- 836, North-Holland, Amsterdam, 1999.
[37] P. Zusmanovich, A converse to the Whitehead theorem, J. Lie Theory 18 (2008), 811-815.
School of Mathematical Sciences, Queen Mary, University of London, London E1 4NS, UK
E-mail address: [email protected]
Department of mathematics, University of California, Irvine, USA
E-mail address: [email protected]
|
1107.0552 | 1 | 1107 | 2011-07-04T07:46:05 | Absolute continuity, Interpolation and the Lyapunov order | [
"math.OA"
] | We extend our Nevanlinna-Pick theorem for Hardy algebras and their representations to cover interpolation at the absolutely continuous points of the boundaries of their discs of representations. The Lyapunov order plays a crucial role in our analysis. | math.OA | math |
ABSOLUTE CONTINUITY, INTERPOLATION AND THE
LYAPUNOV ORDER
PAUL S. MUHLY AND BARUCH SOLEL
Abstract. We extend our Nevanlinna-Pick theorem for Hardy algebras and
their representations to cover interpolation at the absolutely continuous points
of the boundaries of their discs of representations. The Lyapunov order plays
a crucial role in our analysis.
1. Introduction
The celebrated theorem of Nevanlinna and Pick asserts that if n distinct points,
z1, z2, . . . , zn, are given in the open unit disc D and if n other complex numbers are
also given, w1, w2, . . . , wn, then there is a function f in the Hardy algebra H ∞(T),
with norm at most one, such that f (zi) = wi, i = 1, 2, . . . , n, if and only if the Pick
matrix
(cid:18) 1 − wiwj
1 − zizj (cid:19)n
i,j=1
is positive semidefinite. In [4, Theorem 5.3], we generalized this Nevanlinna-Pick
theorem to the setting of Hardy algebras over W ∗-correspondences. Here we intend
to push the work in [4] further, using tools developed in [2]. In a sense that we
shall make precise, we show that there is a condition similar to the positivity of
the Pick matrix that allows one to interpolate at "absolutely continuous" points of
the boundaries of the domains considered in [4]. Before stating our main theorem,
we want to see how one might try to extend the Nevanlinna-Pick theorem to cover
families of absolutely continuous contraction operators along the lines suggested by
[4, Theorem 6.1].
For this purpose, suppose H is a Hilbert space and Z := (Z1, Z2, . . . , Zn) is an
n-tuple of operators in B(H). Then Z defines a completely positive operator ΦZ
on the n × n matrices over B(H) via the formula
Z1
Z2
ΦZ((aij )) : =
=(Ziaij Z ∗
j ).
a11
a21
...
an1
a12
a22
an2
· · ·
a1n
. . .
· · ·
ann
. . .
Zn
Z1
Z2
∗
. . .
Zn
2000 Mathematics Subject Classification. Primary: 15A24, 46H25, 47L30, 47L55, Secondary:
46H25, 47L65.
Key words and phrases. Nevanlinna-Pick interpolation, representations, Hardy algebra, abso-
lute continuity, Lyapunov order.
The research of both authors was supported in part by a U.S.-Israel Binational Science Foun-
dation grant. The second author was also supported by the Technion V.P.R. Fund.
1
2
PAUL S. MUHLY AND BARUCH SOLEL
If all the Zi's have norm less than 1, then I − ΦZ is an invertible map on Mn(B(H))
and (I −ΦZ)−1 is also completely positive. The following theorem, then, is a special
case of [4, Theorem 6.1].
Theorem 1.1. Suppose Z1, Z2, · · · , Zn are n distinct operators in B(H), each
of norm less than 1, and suppose W1, W2, · · · , Wn are n operators in B(H). Then
there is an function f in H ∞(T), with supremum norm at most 1, such that f (Zi) =
Wi, i = 1, 2, . . . , n, where f (Zi) is defined through the Riesz functional calculus, if
and only if the Pick map
(1)
(I − ΦW ) ◦ (I − ΦZ )−1
defined on Mn(B(H)) is completely positive.
Observe that when H is one-dimensional this theorem recovers the classical the-
orem of Nevanlinna and Pick that we stated at the outset. Now Sz.-Nagy and Foiaş
have shown that the proper domain for their H ∞-functional calculus is the collec-
tion of all absolutely continuous contractions. One way to say that a contraction
T is absolutely continuous is to say that when T is decomposed as T = Tcnu + U ,
where Tcnu is completely non-unitary and U is unitary, then the spectral measure
of U is absolutely continuous with respect to Lebesgue measure on the circle. The
content of [10, Theorems III.2.1 and III.2.3] is that a contraction T is absolutely
continuous if and only if the H ∞(T)-functional calculus may be evaluated on T . It
is therefore of interest to modify the hypothesis of Theorem 1.1 and ask for con-
ditions that allow one to interpolate in the wider context where the variables Zi
are assumed to be merely absolutely continuous contractions. In that setting the
map I − ΦZ need no longer be invertible, and so it may not be possible to form
the generalized Pick operator (I − ΦW ) ◦ (I − ΦZ)−1, let alone determine whether
or not it is completely positive. However, there is a notion from matrix analysis,
called the "Lyapunov order", which suggests a replacement for the condition that
the Pick operator (I − ΦW ) ◦ (I − ΦZ)−1be completely positive. To formulate it
we require an idea from the theory of completely positive maps that we analyzed
in [2].
Definition 1.2. Let Φ be a completely positive map on a W ∗-algebra A. An
element a ∈ A is called superharmonic for Φ in case a ≥ 0 and Φ(a) ≤ a; a is called
pure superharmonic in case a is superharmonic and Φn(a) ց 0 as n → ∞.
The superharmonic elements for a completely positive map evidently form a
convex subset in the cone of all non-negative elements in the W ∗-algebra A.
Definition 1.3. Let B be a W ∗-algebra and suppose A is a sub-W ∗-algebra of B.
Suppose Φ : A → A is a completely positive map and that Ψ : B → B is also com-
pletely positive. Then we say Ψ completely dominates Φ in the sense of Lyapunov
in case every pure superharmonic element of Mn(A) for Φn is superharmonic for
Ψn, where Φn (resp. Ψn) is the usual promotion of Φ (resp. Ψ) to Mn(A) (resp.
Mn(B)).
The following proposition links the notion of complete Lyapunov domination to
the complete positivity of (1).
Proposition 1.4. Suppose that A is a sub-W ∗-algebra of a W ∗-algebra B and
suppose Φ and Ψ are completely positive maps on A and B, respectively. Assume
that kΦk < 1, so I − Φ is invertible. Then the Pick operator, P := (I − Ψ) ◦ (I −
ABSOLUTE CONTINUITY, INTERPOLATION AND THE LYAPUNOV ORDER
3
Φ)−1, is completely positive if and only if Ψ completely dominates Φ in the sense
of Lyapunov.
Proof. Note that the hypothesis that kΦk < 1 implies that every superharmonic
element of A is pure superharmonic. Also note that it suffices to prove that P is
positive if and only if {a ∈ A a ≥ 0, Φ(a) ≤ a} ⊆ {b ∈ B b ≥ 0, Ψ(b) ≤ b},
since the same argument will work for every n. Suppose, then, that P is positive
and suppose that a ≥ 0 and Φ(a) ≤ a. Then (I − Φ)(a) ≥ 0. Consequently,
0 ≤ P ((I − Φ)(a)) = (I − Ψ)(a), showing that a ≥ Ψ(a). Suppose, conversely, that
b ≥ 0. Then since kΦk < 1 and Φ is positive, (I − Φ)−1 = Pn≥0 Φn is positive.
Consequently, a = (I −Φ)−1(b) is positive. But also, since (I −Φ)(a) = b is positive,
a ≥ Ψ(a), by hypothesis. That is, P (b) = a − Ψ(a) ≥ 0, which is what we want to
show.
(cid:3)
Our extension of Theorem 1.1 can now be formulated as
Theorem 1.5. Suppose Z1, Z2, · · · , Zn are n distinct absolutely continuous con-
tractions on a Hilbert space H and suppose W1, W2, · · · , Wn are n contractions on
H, then there is a function f ∈ H ∞(T), of norm at most 1, such that f (Zi) = Wi,
i = 1, 2, . . . , n, if and only if ΦW completely dominates ΦZ in the sense of Lya-
punov.
The technology we use to prove Theorem 1.5 works in the more general context of
Hardy algebras over W ∗-correspondences, as we mentioned earlier. This is the arena
in which our analysis takes place. But first, we must provide some background from
[4, 2]. We shall follow terminology and most of the notation from [2]. In particular,
we shall cite the second section of [2] for further background because it gives a fairly
detailed birds-eye view of the theory as of 2010.
We are very grateful to Nir Cohen for introducing us to the Lyapunov order.
Acknowledgment
2. Background and the Main Theorem
Throughout this note, M will denote a fixed W ∗-algebra. We will treat M as
an abstract C ∗-algebra that is a dual space and we will not think of it as acting
concretely on Hilbert space except through representations that we will specify.
Also, E will denote a W ∗-correspondence over M . This means first that E is a
right Hilbert C ∗-module over M which is self-dual. Consequently, the algebra of
all bounded adjointable M -module maps on E, L(E), is all the bounded module
maps and L(E) is a W ∗-algebra. To say that E is a W ∗-correspondence over M
means, then, that there is a normal representation ϕ : M → L(E), making E a
left M -module [2, Paragraph 2.2]. To eliminate technical digressions we assume
that ϕ is faithful and unital. The tensor powers of E, E⊗n, will be the self-dual
completions of the usual C ∗-Hilbert module tensor powers, and the Fock space
F (E) will be the self-dual completion of the C ∗-direct sum of the E⊗n. Then
F (E) is a W ∗-correspondence over M and we denote by ϕ∞ the left action of M
on F (E) [2, Paragraph 2.7]. If ξ ∈ E, then Tξ will denote the creation operator
it determines: Tξη := ξ ⊗ η, η ∈ F (E). The norm-closed subalgebra generated
ϕ∞(M ) and {Tξ ξ ∈ E} is called the tensor algebra of E and will be denoted by
4
PAUL S. MUHLY AND BARUCH SOLEL
T+(E) [2, Paragraph 2.7]. The ultra weak closure of T+(E) in L(F (E)) is called
the Hardy algebra of E and is denoted H ∞(E) [2, Definition 2.1].
Suppose σ : M → B(Hσ) is a normal representation and let σE : L(E) →
B(E ⊗σ Hσ) be the induced representation of L(E) in the sense of Rieffel [7, 8]:
σE(T ) := T ⊗ I, T ∈ L(E). We write I(σE ◦ ϕ, σ) for the set of all operators
C : E ⊗σ Hσ → Hσ that satisfy the equation Cσ ◦ ϕ(a) = σ(a)C for all a ∈ M ; i.e.,
I(σE ◦ ϕ, σ) denotes all the intertwiners of σE ◦ ϕ and σ. Also, we write D(E, σ)
for the set of all elements of I(σE ◦ ϕ, σ) that have norm less than 1, and we write
D(E, σ) for its norm closure. In [6] we proved
Lemma 2.1. (See [2, Paragraph 2.8].) Given z ∈ D(E, σ) , define z × σ by z ×
σ(ϕ∞(a)) := σ(a) and z × σ(Tξ)(h) := z(ξ ⊗ h), a ∈ M , ξ ∈ E, and h ∈ Hσ.
Then z × σ extends to a completely contractive (c.c.) representation of T+(E) on
Hσ. Conversely, given a c.c. representation ρ of T+(E), then ρ = z × σ, where
σ := ρ ◦ ϕ∞ and z(ξ ⊗ h) := ρ(Tξ)h. Further, for F ∈ H ∞(E), the B(Hσ)-valued
function bFσ, defined on D(E, σ) by bFσ(z) := z × σ(F ), is bounded analytic and it
extends to be continuous on D(E, σ) when F ∈ T+(E).
Remark 2.2. We note here that our D(E, σ) is denoted D(Eσ)∗ in [2, Paragraph
2.8], where Eσ := I(σE ◦ ϕ, σ)∗ = I(σ, σE ◦ ϕ) is the σ-dual of E [2, Paragraph
2.6]. This dual space plays an important role in our theory, as we shall see, but we
have opted for the change of notation in order to eliminate numerous unnecessary
and often confusing adjoints from our formulas.
Definition 2.3. A point z ∈ D(E, σ) and the representation z × σ are called abso-
lutely continuous in case z × σ extends to be an ultra weakly continuous represen-
tation of H ∞(E) in B(Hσ). We write AC(E, σ) for all the absolutely continuous
points of D(E, σ).
Our choice of terminology is inspired by the fact that when M = E = C, then
z is absolutely continuous in our sense if and only if z, which is just an ordinary
contraction operator on Hσ, is absolutely continuous in the sense described in the
Introduction.
In general, D(E, σ) ⊆ AC(E, σ) ⊆ D(E, σ), and both inclusions are proper. If
M = E = C, and if σ is the one-dimensional representation of C on C, then D(E, σ)
is the open unit disc in the complex plane and D(E, σ) = AC(E, σ). In every other
setting of which we are aware, D(E, σ) ( AC(E, σ). Also, we know of no situation
where D(E, σ) = AC(E, σ). We have been able to identify AC(E, σ) explicitly in
numerous instances [2, Sections 4 and 5] and we know a lot about this space, but
there is still much that remains mysterious.
The σ-dual of E, Eσ := I(σ, σE ◦ ϕ), is important in this study for several
reasons. The first is that it is a W ∗-correspondence over σ(M )′ in a very natural
way. For ξ, η ∈ Eσ, hξ, ηi is defined to be ξ∗η - the product being the ordinary
operator product, which makes sense as an operator on Hσ since ξ and η both map
from Hσ to E ⊗σ Hσ. The actions of σ(M )′ on Eσ are given by the formula:
a · ξ · b := (IE ⊗ a)ξb,
a, b ∈ σ(M )′, ξ ∈ Eσ.
Again, the products on the right hand side of the equation are ordinary operator
products. The concept of the σ- dual of a W ∗-correspondence was formalized in
[4], but it appeared, implicitly, in a number of places. A key role that it will play
ABSOLUTE CONTINUITY, INTERPOLATION AND THE LYAPUNOV ORDER
5
here is in the identification of the commutant of an induced representation, which
we will describe in the next section. But here we can already see its relevance for
the present considerations by virtue of the following observation: Let z1, z2, . . . , zn
be points in D(E, σ). Then they define a map Φz on the n × n matrices over σ(M )′
by the formula
(2)
Φz((aij )) := (hzi, aij · zj i)
(aij ) ∈ Mn(σ(M )′).
A moment's reflection reveals that Φz is completely positive, as it is the composition
of manifestly completely positive maps.
Our objective in this note is the proof of the following theorem, which will occupy
the next section.
Theorem 2.4. Suppose E is a W ∗-correspondence over a W ∗-algebra M and that
σ is a faithful normal representation of M on the Hilbert space Hσ. Suppose, too,
that n distinct points z1, z2, . . . , zn ∈ AC(E, σ) are given and that n operators in
B(Hσ), W1, W2, · · · , Wn, are given. Define the map Φz on Mn(σ(M )′) by the
formula Φz((aij )) = (hzi, aij · zji) and define the map ΦW on Mn(B(Hσ)) by the
j(cid:1). Then there is an element F in H ∞(E), with
formula ΦW ((Tij)) := (cid:0)WiTijW ∗
kF k ≤ 1, such that bF (zi) = Wi, i = 1, 2, . . . , n, if and only if ΦW completely
dominates Φz in the sense of Lyapunov.
Proof of Theorem 1.5. That theorem is an immediate consequence of Theorem 2.4.
Indeed, in the setting of the former, M = E = C, and σ is just a multiple of the
identity representation, the multiple being the Hilbert space dimension of Hσ. Since
we may safely identify C ⊗σ Hσ with Hσ, D(E, σ) may be identified with the closed
unit ball in B(Hσ), i.e., with all contractions on Hσ. As we noted, the celebrated
theorems of Sz.-Nagy and Foiaş identify AC(C, σ) with the absolutely continuous
contractions on Hσ in the classical sense. When these identifications are made, Φz
of Theorem 2.4 becomes the ΦZ of Theorem 1.5 and, of course, the two ΦW 's are
the same.
(cid:3)
3. The Proof of Theorem 2.4
We will first show that if ΦW completely dominates Φz in the sense of Lyapunov,
then we can find an interpolating F ∈ H ∞(E) of norm at most 1. The route we
shall follow is similar, in certain respects, to the route followed in the proof of [4,
Theorem 5.3] and is based, ultimately, on the commutant lifting approach to the
classical Nevanlinna-Pick theorem pioneered by Sarason [9]. For this purpose, we
need another way to express Lyapunov dominance that reflects the fact that the
zi's involved all lie in AC(E, σ). The key tool in our approach is the notion of an
induced representation for T+(E) and the connection such representations have with
the concept of absolute continuity. They are defined as follows: Let τ be a normal
representation of M on the Hilbert space Hτ . Then we may induce τ to F (E),
obtaining a normal representation τ F (E) of L(F (E)) on the Hilbert space F (E) ⊗τ
Hτ . The restriction of τ F (E) to T+(E), then, is called the representation of T+(E)
induced by τ . It is clearly an absolutely continuous representation of T+(E), since
H ∞(E) is contained in L(F (E)) by definition and τ F (E) is ultraweakly continuous.
As we showed in [2], and will discuss in a moment, induced representations are the
architypical absolutely continuous representations. We continue to use the notation
τ F (E) for its restrictions to T+(E) and H ∞(E).
6
PAUL S. MUHLY AND BARUCH SOLEL
In [5] we develop at length the analogies between induced representations of
T+(E) and H ∞(E) and unilateral shifts. Indeed, a unilateral shift arises from an
induced representation of T+(E), where M = E = C.
Definition 3.1. Let π be a faithful representation of M on Hπ and assume that π
has infinite multiplicity. Then πF (E) is called the universal induced representation
of T+(E) and H ∞(E) determined by π.
Any two faithful π's with infinite multiplicity give unitarily equivalent induced
representations. Further, every induced representation of T+(E) is unitarily equiv-
alent to a subrepresentation of πF (E) obtained by restricting πF (E) to a subspace
of the form F (E) ⊗π K, where K is a subspace of Hπ that reduces π [2, Paragraphs
2.5 and 2.11]. This explains the terminology, allowing us to use the definite article.
The representation π and the induced representation πF (E) will be fixed for the
remainder of this note.
We also shall extend the notation I(σE ◦ ϕ, σ), and write I(ρ1, ρ2) for the set of
operators C : Hρ1 → Hρ2 that intertwine ρ1 and ρ2, where ρ1 and ρ2 are any two
completely contractive representations of T+(E). If ρi is written as zi × σi, i = 1, 2,
then it is easy to see that an operator C : Hρ1 → Hρ2 lies in I(ρ1, ρ2) if and only
if C ∈ I(σ1, σ2) and z2(IE ⊗ C) = C z1.
An important linkage among the universal induced representation, intertwiners,
and absolute continuity is the following theorem.
Theorem 3.2. [2, Theorem 4.7] A point z ∈ D(E, σ) is absolutely continuous if
and only if
_{Ran(c) c ∈ I(πF (E), z × σ)} = Hσ.
Further, for F ∈ H ∞(E), z ∈ AC(E, σ), and c ∈ I(πF (E), z × σ),
(3)
bF (z)c = cπF (E)(F ).
Proof. The first assertion is explicitly in [2] as Theorem 4.7. The second assertion
is easily checked on generators of H ∞(E) of the form ϕ∞(a), a ∈ M , and Tξ, ξ ∈ E.
That is all that is necessary to check.
(cid:3)
Recall that if z ∈ D(E, σ), then z
∗ lies in the W ∗-correspondence Eσ over σ(M )′.
It therefore defines a completely positive map Θz on σ(M )′ by the formula
Θz(a) = hz
∗, a · z
∗i = z(IE ⊗ a)z
∗,
a ∈ σ(M )′.
Indeed, Θz is just a special case of the map Φz in the statement of Theorem 2.4. We
are going to use the following theorem from [2] to obtain an alternate formulation
of the complete Lyapunov dominance assertion in that theorem.
Theorem 3.3. [2, Theorem 4.6] If z ∈ D(E, σ), then an operator q ∈ σ(M )′ is a
pure superharmonic operator for Θz if and only if q can be written as q = cc∗ for
an element c ∈ I(πF (E), z × σ).
Corollary 3.4. We adopt the notation of Theorem 2.4. The map ΦW completely
dominates Φz in the sense of Lyapunov if and only if the following condition is
satisfied: For every integer m ≥ 1, for every choice of function l : {1, 2, . . . , m} →
{1, 2, . . . , n}, and for any choice of m operators cj ∈ I(πF (E), zl(j) × σ) the operator
matrix inequality
(4)
(Wl(i)cic∗
j W ∗
l(j))m
i,j=1 ≤ (cic∗
j )m
i,j=1
ABSOLUTE CONTINUITY, INTERPOLATION AND THE LYAPUNOV ORDER
7
is satisfied.
σ
Proof. Fix m and a function l : {1, 2, . . . , m} → {1, 2, . . . , n}. Write H (m)
for the
direct sum of m copies of Hσ and let σm be the inflated representation of M on
H (m)
, i.e., σm(a) := diag{a, a, · · · , a}. Then σm(M )′ = Mm(σ(M )′). Also, write
E(m) for the direct sum of copies of E, which is also a W ∗-correspondence over
M in the obvious way, and set z := (zl(1), zl(2), · · · , zl(m)). Then we may view z
as a map from E(m) ⊗σm H (m)
, which clearly belongs
to I(σE(m)
◦ ϕ, σm). Consequently, z defines a completely positive map Θz on
σm(M )′ = Mm(σ(M )′), and it is easy to see that
σ = (E ⊗σ Hσ)(m) to H (m)
m
σ
σ
Θz((bl(i),l(j)))m
i,j=1 =(cid:16)zl(i)(IE ⊗ bl(i),l(j))z
∗
l(j)(cid:17) ,
(bl(i),l(j))m
i,j=1 ∈ Mm(σ(M )′).
Moreover, by Theorem 3.3, the pure superharmonic elements Mm(σ(M )′) for Θz
i,j=1, where ci ∈ I(πF (E), (zl(i) × σ)). On the other hand, the
are of the form (cic∗
Wi's may be used to define the completely positive map ΨW,l on Mm(σ(M )′) by
the formula
j )m
ΨW,l((bl(i),l(j))) :=(cid:16)Wl(i)bl(i),l(j)W ∗
l(j)(cid:17) ,
(bl(i),l(j))m
i,j=1 ∈ Mm(σ(M )′).
The inequality 4 is the statement that the condition of the corollary is equivalent
to the assertion that ΨW,l dominates Θz in the sense of Lyapunov for each choice
of m and l. It is now evident that the condition of the corollary implies that ΦW
completely dominates Φz in the sense of Lyapunov by choosing m and l judiciously.
On the other hand, if ΦW completely dominates Φz, then given m and l, one can
clearly choose a k so that the domination of (Φz)kby (ΦW )k in the sense of Lyapunov
gives the desired inequalities of the condition for that m and l.
(cid:3)
In order to follow the commutant lifting approach pioneered by Sarason, we
require the description of the commutant of πF (E)(H ∞(E)) that we developed in
[4]. The description there works for any induced representation, but we formulate
it here specifically for πF (E).
Theorem 3.5. [4, Theorem 3.9] Write ι for the identity representation of π(M )′
on Hπ, and let τ be the induced representation of L(Eπ) acting on F (Eπ) ⊗ι Hπ,
i.e., let τ = ιF (Eπ). Then the map U : F (Eπ) ⊗ι Hπ → F (E) ⊗π Hπ, defined by
the formula
(5) U (ξ1 ⊗ ξ2 ⊗ · · · ⊗ ξn ⊗ h) := (IE⊗(n−1) ⊗ ξ1)(IE⊗(n−2) ⊗ ξ2) · · · (IE ⊗ ξn−1)ξn(h),
ξ1 ⊗ ξ2 ⊗ · · · ⊗ ξn ⊗ h ∈ (Eπ)⊗n ⊗ι Hπ, is a Hilbert space isomorphism and
U τ (H ∞(Eπ))U ∗ = πF (E)(H ∞(E))′.
Likewise, U ∗πF (E)(H ∞(E))U = τ (H ∞(Eπ))′, and the double commutant relations
hold:
πF (E)(H ∞(E))′′ = πF (E)(H ∞(E)),
and
τ (H ∞(Eπ))′′ = τ (H ∞(Eπ)).
We are now ready to show how the complete domination of Φz in the sense of
Lyapunov by ΦW implies that we can interpolate the W 's at the z's in Theorem
2.4.
8
PAUL S. MUHLY AND BARUCH SOLEL
Lemma 3.6. Let
M = span{U ∗c∗h h ∈ Hσ, c ∈ I(πF (E), zi × σ), 1 ≤ i ≤ n}.
Then M is a closed subspace of F (Eπ) ⊗ι Hπ that is invariant under τ (H ∞(Eπ))∗.
Proof. For X ∈ τ (H ∞(Eπ)), U τ (X)U ∗ lies in the commutant of πF (E)(H ∞(E))
by Theorem (3.5). Consequently cU τ (X)U ∗ ∈ I(πF (E), zi × σ) for every c ∈
I(πF (E), zi × σ). But then τ (X)∗U ∗c∗h = U ∗(U τ (X)∗U ∗)c∗h = U ∗(cU τ (X)U ∗)∗h
lies in M for all U ∗c∗h ∈ M.
(cid:3)
i , c ∈ I(πF (E), zi × σ), de-
Lemma 3.7. The correspondence, U ∗c∗h → U ∗c∗W ∗
fined on the generators of M extends to a well-defined contraction operator on
M, say R, if and only if for every integer m ≥ 1, for every choice of func-
tion l : {1, 2, . . . , m} → {1, 2, . . . , n}, and for every choice of m operators cj ∈
I(πF (E), zl(j) × σ) the operator matrix inequality
(Wl(i)cic∗
j W ∗
l(j))m
i,j=1 ≤ (cic∗
j )m
i,j=1
is satisfied. In this event, R commutes with the restriction of τ (H ∞(Eπ))∗to M.
Proof. A linear combination of generators of M is a vector of the form k =
j=1 U ∗c∗
j hj, where cj ∈ I(πF (E), zl(j)×σ) for some m and function l : {1, 2, . . . , m} →
Pm
{1, 2, . . . n}. Since
hcic∗
j hj, hii,
kkk2 =Xj,i
l(j)hjk2 =Xi,j
U ∗c∗
j W ∗
hWi(l)cic∗
j W ∗
i(j)hj, hli,
while
k
mXj=1
the first assertion is immediate. But the second is also immediate since R is "right
i on a generator of the form U ∗c∗h, c ∈ I(πF (E), zi × σ), i.e.,
multiplication" by W ∗
i h, while the restriction of τ (X)∗ to M acts by left multiplication
RU ∗c∗h = U ∗c∗W ∗
for all X ∈ H ∞(Eπ): τ (X)∗U ∗c∗h = U ∗(U τ (X)∗U ∗)c∗h.
(cid:3)
Since M is invariant for τ (H ∞(Eπ))∗, we obtain an ultra weakly continuous com-
pletely contractive representation ρ of H ∞(Eπ) on M by compressing τ (H ∞(Eπ))
to M, i.e.,
ρ(X) := PMτ (X)M,
X ∈ H ∞(Eπ).
Since τ is isometric in the sense of [6] and since R∗ commutes with ρ(H ∞(Eπ)), we
may apply our commutant lifting theorem [6, Theorem 4.4] to conclude that there
is an operator Y ∈ B(F (Eπ) ⊗ι Hπ) of norm at most one such that PMY M = R∗,
Y M⊥ ⊆ M⊥, and Y commutes with τ (H ∞(Eπ)) (see [2, Theorem 2.6], also). By
Theorem 3.5, there is an F ∈ H ∞(E), kF k ≤ 1, such that Y = U ∗πF (E)(F )U . We
conclude from the properties of Y and the definition of R that
U ∗πF (E)(F )∗c∗h = (U ∗πF (E)(F )∗U )U ∗c∗h = Y ∗U ∗c∗h = RU ∗c∗h = U ∗c∗W ∗
i h
for all c ∈ I(πF (E), zi × σ). This, in turn, implies that
cπF (E)(F ) = Wic
for all such c. But cπF (E)(F ) = bF (zi)c for all c ∈ I(πF (E), zi × σ), by equation (3)
in Theorem 3.2. Therefore,
bF (zi)c = Wic
ABSOLUTE CONTINUITY, INTERPOLATION AND THE LYAPUNOV ORDER
9
for all i and all c ∈ I(πF (E), zi × σ). However, by hypothesis, all the zi lie in
AC(E, σ). Consequently, by the first assertion of Theorem 3.2, the closed span of
the ranges of the c's in I(πF (E), zi × σ) is all of Hσ, for every i. We conclude that
bF (zi) = Wi for every i. This completes the proof that if ΦW completely dominates
Φz in the sense of Lyapunov, then there is an F ∈ H ∞(E) that interpolates Wi at
zi.
Proof of the Converse. Part of the argument just given is reversible. Suppose F is
an element of H ∞(E) of norm at most one such that bF (zi) = Wi, i = 1, 2, · · · , n.
Then for each c ∈ I(πF (E), zi × σ)
by equation (3). But then
cπF (E)(F ) = bF (zi)c = Wic,
(U ∗πF (E)(F )∗U )U ∗c∗ = U ∗c∗W ∗
i
for all c ∈ I(πF (E), zi × σ). Since the norm of F is at most 1 we conclude
from Lemma 3.7 that for every integer m ≥ 1, for every choice of function l :
{1, 2, . . . , m} → {1, 2, . . . , n}, and for every choice of m operators cj ∈ I(πF (E), zl(j)×
σ) the operator matrix inequality
(Wl(i)cic∗
i,j=1 ≤ (cic∗
j )m
j W ∗
l(j))m
i,j=1
is satisfied. So by Corollary 3.4, we conclude that ΦW completely dominates Φz in
the sense of Lyapunov.
(cid:3)
References
[1] N. Cohen and I. Lewkowicz, The Lyapunov order for real matrices, Linear Algebra and Its
Applications, 430, (2009), 1849-1866.
[2] P. S. Muhly and B. Solel, Representations of Hardy algebras: absolute continuity, inter-
twiners, and superharmonic operators, Integral Equations and Operator Theory 70 (2011),
151-203 (arXiv:1006.1398).
[3] P. S. Muhly and B. Solel, Schur Class Operator Functions and Automorphisms of Hardy
Algebras, Documenta Math. 13 (2008), 365 -- 411.
[4] P. S. Muhly and B. Solel, Hardy algebras, W ∗-correspondences and interpolation theory,
Math. Ann. 330 (2004), 353-415.
[5] P. S. Muhly and B. Solel , Tensor algebras, induced representations, and the Wold decompo-
sition, Canad. J. Math. 51 (1999), 850-880.
[6] P. S. Muhly and B. Solel, Tensor algebras over C ∗-correspondences (Representations, dila-
tions, and C ∗-envelopes), J. Functional Anal. 158 (1998), 389 -- 457.
[7] M. Rieffel, Morita equivalence for C ∗-algebras and W ∗-algebras, J. Pure Appl. Alg. 5 (1974),
51 -- 96.
[8] M. Rieffel, Induced representations of C ∗-algebras, Adv. in Math. 13 (1974), 176 -- 257.
[9] D. Sarason, Generalized interpolation in H∞, Trans. Amer. Math. Soc. 127 (1967), 179 -- 203.
[10] B. Sz.-Nagy, C. Foiaş, H. Bercovici, and L. Kérchy, Harmonic Analysis of Operators on
Hilbert Space, Springer, New York, 2010.
Department of Mathematics, University of Iowa, Iowa City, IA 52242
E-mail address: [email protected]
Department of Mathematics, Technion, 32000 Haifa, Israel
E-mail address: [email protected]
|
1807.11425 | 3 | 1807 | 2019-11-26T04:55:06 | The non-selfadjoint approach to the Hao-Ng isomorphism | [
"math.OA",
"math.FA"
] | In an earlier work, the authors proposed a non-selfadjoint approach to the Hao-Ng isomorphism problem for the full crossed product, depending on the validity of two conjectures stated in the broader context of crossed products for operator algebras. By work of Harris and Kim, we now know that these conjectures in the generality stated may not always be valid. In this paper we show that in the context of hyperrigid tensor algebras of C*-correspondences, each one of these conjectures is equivalent to the Hao-Ng problem. This is accomplished by studying the representation theory of non-selfadjoint crossed products of C*-correspondence dynamical systems; in particular we show that there is an appropriate dilation theory. A large class of tensor algebras of C*-correspondences, including all regular ones, are shown to be hyperrigid. Using Hamana's injective envelope theory, we extend earlier results from the discrete group case to arbitrary locally compact groups; this includes a resolution of the Hao-Ng isomorphism for the reduced crossed product and all hyperrigid C*-correspondences. A culmination of these results is the resolution of the Hao-Ng isomorphism problem for the full crossed product and all row-finite graph correspondences; this extends a recent result of Bedos, Kaliszewski, Quigg and Spielberg. | math.OA | math |
THE NON-SELFADJOINT APPROACH TO THE
HAO-NG ISOMORPHISM
ELIAS G. KATSOULIS AND CHRISTOPHER RAMSEY
Abstract. In an earlier work, the authors proposed a non-self-
adjoint approach to the Hao-Ng isomorphism problem for the full
crossed product, depending on the validity of two conjectures stated
in the broader context of crossed products for operator algebras.
By work of Harris and Kim, we now know that these conjec-
tures in the generality stated may not always be valid.
In this
paper we show that in the context of hyperrigid tensor algebras
of C∗-correspondences, each one of these conjectures is equiva-
lent to the Hao-Ng problem. This is accomplished by studying
the representation theory of non-selfadjoint crossed products of
C∗-correspondence dynamical systems; in particular we show that
there is an appropriate dilation theory. A large class of tensor alge-
bras of C∗-correspondences, including all regular ones, are shown
to be hyperrigid. Using Hamana's injective envelope theory, we
extend earlier results from the discrete group case to arbitrary
locally compact groups; this includes a resolution of the Hao-Ng
isomorphism for the reduced crossed product and all hyperrigid C∗-
correspondences. A culmination of these results is the resolution of
the Hao-Ng isomorphism problem for the full crossed product and
all row-finite graph correspondences; this extends a recent result
of Bedos, Kaliszewski, Quigg and Spielberg.
1. Introduction
Let ((X,C),G, α) be a C∗-correspondence dynamical system where
G is a locally compact group and α is a generalized gauge action. This
action can be extended uniquely to the Cuntz-Pimsner algebra OX The
Hao-Ng isomorphism problem asks whether
OX ⋊α G ≃ OX⋊αG
in the reduced or full crossed products. This problem is named after
Hao and Ng who proved the validity of this formula when G is amenable
2010 Mathematics Subject Classification. 46L07, 46L08, 46L55, 47B49, 47L40,
47L65.
Key words and phrases: C∗-correspondence, crossed product, Cuntz-Pimsner
algebra, tensor algebra, Hao-Ng isomorphism, C∗-envelope, operator algebra.
1
2
E.G. KATSOULIS AND C. RAMSEY
[14, Theorem 2.10]. However, this formula was first studied by Abadie
in the context of Takai duality for equivalence bimodules. Indeed, in
Abadie's proof for the Takai duality, the Hao-Ng isomorphism forms the
crucial step of the proof and corresponds to the key isomorphism of [34,
Lemma 7.2] in the classical case. In general, the Hao-Ng isomorphism
has proved to be a significant stimulant to research as versions of it
appear in many different contexts, e.g.
in Schafhauser's work [33]
on AF-embedability, or in Deaconu's work [7, 8] on group actions on
graph C∗-algebras. In its full generality, the problem remains open and
under investigation by several authors [3, 17, 18, 24, 27].
The authors initiated a study in [22, 20, 23] of non-selfadjoint
crossed products of operator algebra dynamical systems (A,G, α) where
α acts by completely isometric isomorphisms of A. The main thrust
of [22, Chapter 7] and [20] is that the Hao-Ng isomorphism problem
can and should be thought of as a non-selfadjoint problem. For the
reduced crossed product this kind of approach has been and continues
to be quite successful. For instance, we now know that the Hao-Ng iso-
morphism for the reduced crossed product holds for all discrete groups
[20], a fact that resolves an open problem from [3] (and more is ac-
complished in this paper).
The Hao-Ng isomorphism for the full crossed product seems to be
In [22] we envisioned the following line of
a much harder problem.
attack. First one verifies
(1)
C∗
env(A ⋊α G) ≃ C∗
env(A) ⋊α G
in the full crossed product case for an arbitrary non-selfadjoint dynami-
cal system (A,G, α); this is Problem 1 in [22]. Subsequently, one solves
Problem 2 in [22] by showing that all relative crossed products coin-
cide. Assuming that both problems have been resolved in the positive,
now one specializes on tensor algebra dynamical systems and obtains
(2)
T +
X
⋊α G ≃ T +
X⋊αG
by invoking the solution of Problem 2 and the remarks following [22,
Theorem 7.13]. Recalling that C∗
X ) = OX , one recovers now the
Hao-Ng isomorphism by combining equations (1) and (2). Note that
even though a positive answer for both Problems 1 and 2 leads to a
positive resolution for the Hao-Ng isomorshism, the exact relation of
each one of these problems with the Hao-Ng isomorphism was never
clarified in [22].
env(T +
The central result of this paper, Theorem 4.9, clarifies that rela-
tion and shows that the Hao-Ng problem actually leads to equivalent
statements in non-selfadjoint operator algebra theory, whose validity
THE NON-SELFADJOINT APPROACH TO THE HAO-NG ISOMORPHISM 3
or refutation will therefore resolve the isomorphism. Specifically, for
a non-degenerate C∗-correspondence X we show that validity of (1)
for A = T +
X is equivalent to the validity of the Hao-Ng isomorphism
OX ⋊αG ≃ OX⋊αG. In addition, for a large class of C∗-correspondences,
including all regular ones, we show that the validity of the Hao-Ng iso-
morphism OX ⋊α G ≃ OX⋊αG is equivalent to the fact that all relative
crossed products for (T +
X ,G, α) coincide, where α is a generalized gauge
action. (For a general C∗-correspondence X this last statement is im-
plied by any of the previous two.)
Theorem 4.9 relates to exciting new work by Harris and Kim [15].
Indeed these authors have answered both Problems 1 and 2 from [22,
Chapter 7] by producing finite dimensional, hyperrigid dynamical sys-
tems (A,G, α) with distinct relative crossed products and failing (1).
However the examples of Harris and Kim [15] do not concern tensor
algebras of C∗-correspondences and so the Hao-Ng problem remains
open. Theorem 4.9 shows now that the resolution of the Hao-Ng prob-
lem will lead to or will follow from the existence or the absence of
Harris-Kim type examples but in the realm of tensor algebras. Need-
less to say that the quest for such examples, or the refutation of their
existence, becomes now a project of high priority.
To test our new results, we study the Hao-Ng isomorphism for a
class of C∗-correspondences that plays a central role in the theory:
graph C∗-correspondences. In Theorem 5.4 we show that the Hao-Ng
isomorphism problem is true in the case of row-finite graph correspon-
dences, thus showing that the crossed product of such a Cuntz-Krieger
algebra is the Cuntz-Pimsner algebra of a crossed product (of a graph)
correspondence. This is done by showing that in the case of a dynam-
ical system (A,G, α) where A is the tensor algebra of any graph, G
any locally compact group and α a generalized gauge action, all rela-
tive crossed products coincide. Then, Theorem 4.9 finishes the proof
for row-finite graphs. Note that in the special case where G is dis-
crete, Theorem 5.4 has also been obtained independently by Bedos,
Kaliszewski, Quigg and Spielberg using different methods [4, Corol-
lary 6.8 and Remark 6.10]. It is worth mentioning here that Theorem
5.4 is essentially obtained by dilating representations in a wholly con-
structive manner and should prove of much interest to those who study
the representation theory of C∗-correspondences. At the moment, the
lack of a constructive dilation proof of C∗
X ) = OX seems to be a
barrier to establishing (2) for the full crossed product in general.
On the way to proving the above theorems we obtain several results of
independent interest. First we resolve Problem 3 from our monograph
env(T +
4
E.G. KATSOULIS AND C. RAMSEY
X
[22, Chapter 8]. Specifically, we show if (X,C) is a non-degenerate C∗-
correspondence and α : G → (X,C) is the generalized gauge action of a
locally compact group, then T +
⋊α G is necessarily the tensor algebra
of some C∗-correspondence.
Another result of independent interest is Theorem 3.1, which identi-
fies a large class of hyperrigid C∗-correspondences, i.e., C∗-correspon-
dences whose tensor algebras are hyperrigid. Indeed our central Theo-
rem 4.9 actually applies to all hyperrigid C∗-correspondences. To make
that result usable, we show that any C∗-correspondence (X,C, ϕX ) with
ϕX(JX)X = X is hyperrigid (here JX denotes Katsura's ideal). This
includes all previous known examples of hyperrigid C∗-correspondences
and many more, e.g., all regular ones.
An interesting byproduct of our techniques on the full crossed prod-
uct version of the Hao-Ng problem, is the resolution of the same prob-
lem for the reduced crossed product and all hyperrigid C∗-correspon-
dences. In [20] the first named author verified the Hao-Ng isomorphism
for the reduced crossed product and all discrete groups. Because here
we are addressing locally compact groups which may not be discrete, we
have to use an approach different from that of [20]. In particular, the
algebra A does not embed in either A⋊α G or A⋊r
α G and so restricting
a maximal map of the crossed product on the core algebra A (as we
did in [20]) is no longer an option. Instead we use Hamana's injective
envelope theory, an approach towards the Hao-Ng isomorphism which
is used in this paper for the first time. This approach was adopted
after illuminating discussions with S. Echterhoff and we are grateful to
him for that.
We denote by N the set of positive integers, while Z+
0 = N∪{0}. We
denote by span{· · ·} the closure of the linear span of {· · ·}. An ideal
of a C∗-algebra always means a closed two-sided ideal.
2. Crossed products and C∗-covers
Let (A,G, α) be an operator algebra dynamical system, meaning
that A is an approximately unital operator algebra and G is a locally
compact (Hausdorff) group acting continuously on A by completely
isometric automorphisms, α : G → Aut(A). The aim of this section is
to better understand the relationship of α-admissible C∗-covers. Recall
that a C∗-cover (C, ι) of A is a C∗-algebra C and a complete isometry
ι : A → C such that C ∗(ι(A)) = C.
The two nicest C∗-covers of A are the "biggest" and the "smallest"
covers C∗
env(A). These are defined by their universal
properties. Namely, whenever (C, ι) is a C∗-cover there are (unique)
max(A) and C∗
THE NON-SELFADJOINT APPROACH TO THE HAO-NG ISOMORPHISM 5
surjective ∗-homomorphisms ϕ : C∗
env(A)
such that ϕ(a) = ι(a) and ψ(ι(a)) = a, for all a ∈ A.
From [22], a C∗-cover (C, ι) is called α-admissible if there exists a
group representation β : G → Aut(C) acting on C by ∗-automorphisms
such that
max(A) → C and ψ : C → C∗
βs(ι(a)) = ι(αs(a)),
∀s ∈ G, a ∈ A.
In [22, Lemma 3.3] we established that both C∗
max(A) are
always α-admissible. However, in [22] we did not provide any examples
of C∗-covers which fail to be α-admissible. We thank David Sherman
for bringing this to our attention and asking us whether such covers do
exist.
env(A) and C∗
Proposition 2.1. Not all C∗-covers are α-admissible.
Proof. Let C = C(T)⊕M2 and ι : A(D) → C be given by z 7→ z⊕[ 0 0
1 0 ].
By von Neumann's inequality it is straightforward that ι is a complete
isometry. Now
ι(z) − ι(z2)ι(z)∗ = (z ⊕ [ 0 0
1 0 ]) − (z2 ⊕ [ 0 0
= 0 ⊕ [ 0 0
1 0 ] .
0 0 ])(¯z ⊕ [ 0 1
0 0 ])
Thus, C ∗(ι(A(D))) = C and (C, ι) is a C∗-cover of A(D).
Consider the Mobius transformation ϕ(z) = z− 1
2
1− z
2
which gives ϕ ∈
Aut(D). From this define the dynamical system (A(D), α, Z) where
αn(f ) = f ◦ ϕn which is the same as z 7→ ϕn(z).
It is well known
that composition with a Mobius map is a completely isometric auto-
morphism of the disc algebra.
Suppose that there exists α : Z → Aut(C) such that αn(i(f )) =
i(αn(f )),∀f ∈ A(D). Calculating
=(cid:18)z −
= −
z − 1
1 − z
ϕ(z) =
1
2
2
2
+
1
2(cid:19)(cid:18)1 +
+
z2
z
2
4
z2 + · · ·
3
8
3
4
z +
+ · · ·(cid:19)
we get that
ι(ϕ(z)) = ϕ(z) ⊕h − 1
3
2
0
4 − 1
2i .
6
E.G. KATSOULIS AND C. RAMSEY
Hence,
α1 (0 ⊕ [ 0 0
1 0 ]) = α1(ι(z) − ι(z2)ι(z)∗)
2i −(cid:16)ϕ(z)2 ⊕h 1
4
− 3
4
0
1
4i(cid:17)(cid:16)ϕ(z) ⊕h − 1
3
2
4
0 − 1
2i(cid:17)
2
3
8 − 3
0
4 − 1
= ι(α1(z)) − ι(α1(z2))ι(α(z))∗
= ι(ϕ(z)) − ι(ϕ(z))2ι(ϕ(z))∗
= ϕ(z) ⊕h − 1
= 0 ⊕h − 3
16 i .
= √2(cid:13)(cid:13)(cid:13)h 3
16i(cid:13)(cid:13)(cid:13)
16 i(cid:13)(cid:13)(cid:13)
8 − 3
=
16
3
16
3
3
8
8
3
3√10
16
But then
0 ⊕h − 3
3
8
(cid:13)(cid:13)(cid:13)
< 1 = k0 ⊕ [ 0 0
1 0 ]k ,
a contradiction as ∗-automorphisms are isometric. Therefore, no such
α exists and (C, ι) is a non α-admissible C∗-cover of (A(D), α, Z).
If we do have an α-admissible cover then we can abuse the notation
and call the group representation α again because of the next result.
Lemma 2.2. Let (A,G, α) be an operator algebra dynamical system.
If (C, ι) is an α-admissible C∗-cover then there is a unique group rep-
resentation of G on C acting by ∗-automorphisms extending α.
Proof. Let β1 and β2 be two such extensions. Then for any s ∈ G we
have β1,s ◦ β−1
2,s = id on A and we just need to prove this is the identity
map. To this end assume that β : G → Aut(C) extends the identity
map on A. That is, βsA = idA, for all s ∈ G.
max(A) there is a unique surjective ∗-
homomorphism ϕ : C∗
max(A) → C such that ϕ(a) = ι(a), for all a ∈ A.
Notice that βs ◦ ϕ(a) = βs ◦ ι(a) = ι(a), for all a ∈ A. Thus, βs ◦ ϕ = ϕ
by uniqueness which implies that βs = id on C, for any s ∈ G.
By the universal property of C∗
Now we turn to crossed products of non-selfadjoint operator algebras.
Definition 2.3 ([22]). Let (A,G, α) be an operator algebra dynamical
system and let (C, ι) be an α-admissible C∗-cover. The relative reduced
and full crossed products are denoted by A ⋊r
(C,ι),α G and A ⋊(C,ι),α G
and are defined to be the closure of Cc(G,A) in C ⋊r
α G and C ⋊α G,
respectively.
All relative reduced crossed products are in fact completely isomet-
rically isomorphic [22, Theorem 3.12] and so we define the reduced
crossed product, denoted A ⋊r
α G, to be this unique object. Lastly, we
define the full crossed product to be
A ⋊α G := A ⋊C∗
max(A),α G.
THE NON-SELFADJOINT APPROACH TO THE HAO-NG ISOMORPHISM 7
In fact, this is the universal algebra for all covariant representations of
(A,G, α) [22, Proposition 3.7]. Finally, it should be noted that, as in
the selfadjoint case, if G is amenable then the full and reduced crossed
products coincide [22, Theorem 3.14].
Now we are able to state and prove the main theorem of this section.
Theorem 2.4. Let (A,G, α) be an operator algebra dynamical system.
Then for every α-admissible C∗-cover (C, ι) there are surjective com-
pletely contractive homomorphisms
A ⋊α G
env(A),α G
env(A), (C, ι)
max(A) respectively αenv, αC and αmax. By the universal proper-
qmax−−−→ A ⋊(C,ι),α G
qmin−−→ A ⋊C∗
such that they are just the identity on Cc(G,A).
Proof. Label the unique extensions of α to the C∗-covers C∗
and C∗
ties there exists surjective ∗-homomorphisms
env(A) and ϕmax : C∗
max(A) → C
such that ϕenv(ι(a)) = a and ϕmax(a) = ι(a), for all a ∈ A.
the following commutative diagram:
By uniqueness of the quotient maps and of the extensions we have
ϕenv : C → C∗
ϕmax−−−→ C
ϕenv−−−→ C∗
env(A)
ϕmax−−−→ C
ϕenv−−−→ C∗
env(A)
αenvy
C∗
max(A)
αmaxy
C∗
max(A)
αCy
Thus, ker ϕmax is an αmax-invariant ideal and ker ϕenv is an αC-invariant
ideal. By [34, Proposition 3.19], full C∗-crossed products preserve
exact sequences by α-invariant ideals. Hence, we have the following
surjective ∗-homomorphisms
max(A) ⋊αmax G ϕmax⋊id
C∗
−−−−−→ C ⋊αC G ϕenv⋊id
env(A) ⋊αenv G.
−−−−→ C∗
So
qmax = ϕmax ⋊ idA⋊αG and qmin = ϕenv ⋊ idA⋊(C,ι),αG
are completely contractive homomorphisms which amount to the iden-
tity on Cc(G,A).
The benefit of this theorem, as will be used later, is that one needs
only to show that the map qmin ◦ qmax is a completely isometric isomor-
phism to establish that all relative crossed products are the same.
8
E.G. KATSOULIS AND C. RAMSEY
3. Hyperrigidity and the Hao-Ng isomorphism
A not necessarily unital operator algebra A is said to be hyperrigid
if given any (non-degenerate) ∗-homomorphism
env(A) −→ B(H)
τ : C∗
then τ is the only completely positive, completely contractive extension
of the restricted map τA. By adding an injective direct summand if
necessary, it is easy to see that in order to verify hyperrigidity, one
needs to consider only injective ∗-representations τ but this need not
concern us here. The term hyperrigid was coined by Arveson in [2] but
the concept had been floating around in various forms before this, e.g.
[9].
Our definition is slightly weaker than that of Duncan's [9, Section 4]
as Duncan requests that τ be the only completely contractive extension
of the restricted map, i.e., no requirement of positivity in the non-unital
case. In any case [9, Proposition 4] shows that the graph algebra of any
row-finite graph is hyperrigid. Actually we are about to provide a much
stronger result but first we need to remind the reader the definition and
some of the basic notation regarding C∗-correspondences.
A C∗- correspondence (X,C, ϕX ), or just (X,C), consists of a C∗-
algebra C, a Hilbert C-module (X,h , i) and a (non-degenerate) ∗-
homomorphism ϕX : C → L(X) into the C∗-algebra of adjointable op-
erators on X. Equivalently, a (represented) C∗-correspondence (X,C)
consists of a C∗-algebra C ⊆ B(K), K a Hilbert space, and a norm-
closed C-bimodule X ⊆ B(K) satisfying X ∗X ⊆ C (this allows us to
define the inner product h , i) and span{CX} = X (this is the non-
degeneracy of the left action of C). The equivalence of the two defini-
tions follows from the fact that an abstract C∗-correspondence embeds
in the Toeplitz C∗-algebra TX that we will define below and therefore
can be represented on a Hilbert space.
A representation (ρ, t) of a C∗-correspondence into B(H), is a pair
consisting of a non-degenerate ∗-homomorphism ρ : C → B(H) and a
linear map t : X → B(H), such that
ρ(c)t(x) = t(ϕX(c)(x)),
for all c ∈ C and x ∈ X. It is called an isometric (Toeplitz) represen-
tation when
t(x)∗t(x′) = ρ(hx, x′i),
for all c ∈ C and x, x′ ∈ X.
THE NON-SELFADJOINT APPROACH TO THE HAO-NG ISOMORPHISM 9
By the relations above, the C∗-algebra generated by an isometric
representation (ρ, t) equals the closed linear span of
t(x1)· · · t(xn)t(y1)∗ · · · t(ym)∗,
xi, yj ∈ X.
For any isometric representation (ρ, t) there exists a ∗-homomorphism
ψt : K(X) → B, such that ψt(θx,y) = t(x)t(y)∗, where K(X) is the sub-
algebra of L(X) of so-called compact operators generated by θx,y(z) =
xhy, zi. (See [19, Chapter 3] for more details on this topic.)
There exists a universal Toeplitz representation, denoted as (ρ∞, t∞),
so that any other representation of (X,C) is equivalent to a direct
sum of sub-representations of (ρ∞, t∞). The Cuntz-Pimsner-Toeplitz
C∗-algebra TX is defined as the C∗-algebra generated by the image of
(ρ∞, t∞).
X of a C∗-correspondence [28] (X,C) is the
norm-closed subalgebra of TX generated by all elements of the form
ρ∞(c), t∞(x), c ∈ C, x ∈ X. The tensor algebra T +
X contains a faithful
copy of the C∗-correspondence (X,C). Thus X inherits an operator
space from T +
X ; we can now say that a representation (ρ, t) of (X,C)
is completely contractive whenever t is a completely contractive map
with respect to that operator space structure.
The tensor algebra T +
Consider the ideal
JX ≡ ϕ−1
X (K(X)) ∩ ker ϕ⊥
X .
(which we will call Katsura's ideal.) An isometric representation (ρ, t)
of (X,C, ϕX) is said to be covariant (Cuntz-Pimsner) if and only if
ψt(ϕX(c)) = ρ(c), for all c ∈ JX. The universal C∗-algebra for all iso-
metric covariant representations of (X,C) is the Cuntz-Pimsner algebra
OX. The algebra OX contains (a faithful copy of) C and (a unitarily
equivalent) copy of X.
The first author and Kribs [21, Lemma 3.5] have shown that the
non-selfadjoint algebra of OX generated by these copies of C and X is
completely isometrically isomorphic to T +
X ) ≃
OX. See [21, 28] for more details.
X . Furthermore, C∗
env(T +
Now to the hyperrigidity of tensor algebras.
X is a hyperrigid operator algebra.
Theorem 3.1. Let (X,C) be a C∗-correspondence.
If ϕX(JX) acts
non-degenerately on X, then (X,C) is a hyperrigid C∗-correspondence,
i.e., T +
Proof. Let τ : OX −→ B(H) be a ∗-homomorphism and let τ ′ : OX −→
B(H) be a completely contractive and completely positive map that
agrees with τ on T +
X . We are to prove that τ ′ is multiplicative and so
it agrees with τ . Since τ ′ is a completely contractive and completely
10
E.G. KATSOULIS AND C. RAMSEY
positive map, we can use multiplicative domain arguments [6, Propo-
sition 1.5.7].
Let (ρ, t) be the universal Cuntz-Pimsner representation of (X,C).
X is a C∗-algebra, the multiplicative domain of τ ′ con-
Since ρ(C) ⊆ T +
tains ρ(C). We claim that it also contains t(X).
Indeed, for any x ∈ X we have
τ ′(t(x))∗τ ′(t(x)) = τ (t(x))∗τ (t(x)) = τ (t(x)∗t(x))
(3)
= τ (ρ(hx, xi)) = τ ′(ρ(hx, xi))
= τ ′(t(x)∗t(x)),
where the equation on the second line holds because ρ(C) ⊆ T +
the two maps agree there.
X and
Let a ∈ JX and x ∈ X. Since ϕX(a) ∈ K(X), we have zm,k, wm,k ∈
X, m, k ∈ N, so that
(4)
ϕX(a) = lim
θzm,k,wm,k
m→∞Xk
is a limit of finite rank operators in K(X). Let X0 ⊆ X be the C-
submodule generated by x and all zm,k, wm,k ∈ X, m, k ∈ N. Since X0
is countably generated, Kasparov's Stabilization Theorem implies the
n=1 in X0 so that kPl
existence of {xn}∞
n=1 θxn,xnk ≤ 1, for all l ∈ N,
and
Xn=1
θxn,xn(ξ) = ξ, for all ξ ∈ X0.
From this, a standard approximation argument involving (4) shows
that
∞
∞
∞
(5)
θxn,xnϕX(a) =
ϕX(a)θxn,xn = ϕX (a),
Xn=1
Xn=1
with the convergence in the norm topology1. Then
ϕX(aa∗) = lim
k
=
By the Schwarz inequality
∞
Xn=1
ϕX(a)(cid:0)
k
Xn=1
θxn,xn(cid:1)ϕX (a)∗
θϕX (a)xn,ϕX (a)xn.
(6)
τ ′(cid:0)t(ϕX(a)xn)t(ϕX(a)xn)∗(cid:1) ≥ τ ′(t(ϕX (a)xn))τ ′(t(ϕX(a)xn))∗,
1This is exactly the same argument one uses on non-separable Hilbert space to
write any compact operator as a (perhaps infinite) sum of rank-one operators.
THE NON-SELFADJOINT APPROACH TO THE HAO-NG ISOMORPHISM 11
for all n ∈ N, and so
τ ′(ρ(aa∗)) = τ ′(ψt(ϕX(aa∗))
∞
=
≥
=
τ ′(cid:0)t(ϕX (a)xn)t(ϕX (a)xn)∗(cid:1)
τ ′(t(ϕX (a)xn))τ ′(t(ϕX(a)xn))∗
τ (t(ϕX (a)xn))τ (t(ϕX (a)xn))∗
∞
Xn=1
Xn=1
Xn=1
∞
= τ (ψt(ϕX(aa∗)) = τ (ρ(aa∗)) = τ ′(ρ(aa∗)).
Hence (6) is actually an equality. Combining this with (3), we conclude
that t(ϕX(a)xn) belongs to the multiplicative domain of τ ′, for all a ∈
JX and n ∈ N. Since ρ(C) is also contained in the multiplicative
domain of τ ′, we have that
∞
t(ϕX (a)x) =
t(ϕX(a)xn)ρ(hxn, xi)
Xn=1
belongs to the multiplicative domain of τ ′, for all a ∈ JX and x ∈ X.
Since ϕX(JX) acts non-degenerately on X, the multiplicative domain
of τ ′ contains t(X), as desired. This completes the proof.
Recall that a C∗-correspondence (X,C) is said to be regular iff C acts
faithfully on X by compact operators, i.e., JX = C. The following is
immediate.
Corollary 3.2. A regular C∗-correspondence is necessarily hyperrigid.
We are about to see that the assumption of injectivity cannot be
removed from the Corollary above. But first we need criterion for the
failure of hyperrigidity.
Proposition 3.3. Let (X,C) be a C∗-correspondence with JX = {0}.
Then (X,C) fails to be hyperrigid.
Proof. Let (π, t) be any isometric representation of (X,C) on a Hilbert
space H. If V1, V2 are the unilateral and bilateral (forward) shift re-
spectively, then the associations
(7)
C ∋ a −→ a ⊗ I,
X ∋ x −→ x ⊗ Vi, i = 1, 2,
12
E.G. KATSOULIS AND C. RAMSEY
determine isometric representations of X, which are neccesarilly Cuntz-
Pimsner covariant, since JX = {0}. Therefore they promote to repre-
sentations ϕ1 and ϕ2 of OX ≃ C∗
0 ) and H⊗ℓ2(Z) re-
spectively. Now notice that when ϕ2 is being compressed on H⊗ℓ2(Z+
0 ),
it produces a completely positive contractive map ϕ2 6= ϕ1, which how-
ever agrees with ϕ1 on T +
X . Hence (X,C) is not hyperrigid.
X ) on H⊗ℓ2(Z+
env(T +
Recall that if α is an endomorphism of a C∗-algebra A, then the
semicrossed product A⋊α Z+
0 (also denoted as A⋊α Z+ in the literature)
is simply the tensor algebra of the C∗-correspondence Aα, where the
left action on A is coming from α.
In the case where both A and
α are unital and α is injective, such algebras are always hyperrigid.
This has already been noted in the literature, eg. [16], but it is also
an immediate consequence of Corollary 3.2.
It is worth noting that
the requirement of α being injective cannot be dropped from neither
Corollary 3.2 nor the discussion above.
Example 3.4. Let X be a compact Hausdorff space which is not a
singleton and consider some x ∈ X which is not an isolated point. Let
ϕ : X → X with ϕ(y) = x, for all y ∈ X . Then the semicrossed
product C(X ) ⋊ϕ Z+
Indeed, in that case, the kernel of the right action equals C0(X\{x}).
Hence Katsura's ideal is trivial and Proposition 3.3 applies.
0 is not hyperrigid.
env(A), ι) where C∗
Finally recall that the C*-envelope of a non-unital operator alge-
bra can be computed from the C*-envelope of its unitization. More
precisely, as the pair (C∗
env(A) is the C*-subalgebra
env(A1), ι) of the (unique)
generated by ι(A) inside the C*-envelope (C∗
unitization A1 of A. By the proof of [5, Proposition 4.3.5] this C*-
envelope of an operator algebra A has the desired universal property,
that for any C*-cover (ι′,B′) of A, there exists a (necessarily unique
and surjective) ∗-homomorphism π : B′ → C∗
env(A), such that π◦ι′ = ι.
We start with an elementary result regarding crossed products.
Lemma 3.5. Let (C,G, α) be a C∗-dynamical system and let D ⊆ C be
the C∗-subalgebra of C generated by some selfadjoint approximate unit
for C. Then
(i) CCc(G,D) is dense in C ⋊α G.
(ii) If π : C ⋊αG → B(H) is a non-degenerate representation, then
its restriction on Cc(G,D) is also non-degenerate.
Proof. Let {ei}i∈I be the selfadjoint approximate unit generating D.
Then any elementary tensor h⊗c ∈ Cc(G,D), where (h⊗c)(s) = h(s)c,
THE NON-SELFADJOINT APPROACH TO THE HAO-NG ISOMORPHISM 13
h ∈ Cc(G), c ∈ C, can be written as
h ⊗ c = lim
i∈I
c (h ⊗ ei) ∈ CCc(G,D).
This implies (i). For (ii) notice that by taking adjoints in (i), Cc(G,D)C
is also dense in C ⋊α G. Hence
π (Cc(G,D))H = π (Cc(G,D)) π(C)H = π (Cc(G,D)C)H
which is dense in π(C)H = H and the conclusion follows.
Our next result has been established for all discrete groups in [20].
Here we extend it to arbitrary locally compact groups provided that
the pertinent algebras are hyperrigid. One of the key ingredients of the
proof is the use of injectivity for operator spaces. We briefly review
the key definitions and the results used in the proof. We follow [29] in
our presentation; most of the material first appeared in [13].
An operator space I is said to be injective provided that for any pair
of operator spaces E ⊆ F and completely contractive map ϕ : E → I,
there exists a completely contractive map ψ : F → I that extends ϕ.
Given an operator space F , we say that (E, κ) is an injective envelope
of F provided that
(i) E is injective,
(ii) κ : F → E is a complete isometry,
(iii) if E1 is injective with κ(F ) ⊆ E1 ⊆ E, then E1 = E.
Hamana essentially showed that every operator space F ⊆ B(H)
admits an injective envelope (E, κ), with E ⊆ B(H) and κ being the
inclusion map [29, Theorem 15.4]. The proof of [29, Theorem 15.4]
shows that E materializes as the range of a completely contractive
idempotent ϕ : B(H) → B(H). If B(H) ∋ I ∈ F then the completely
contractive idempotent ϕ is unital and therefore completely positive.
Hence the range of ϕ, i.e., E, is an operator system and the Choi-
setting a ◦ b = ϕ(ab)
Effros Theorem [29, Theorem 15.2] applies:
defines a multiplication on E = ϕ(B(H)), and E equipped with this
multiplication and its usual ∗-operation becomes a C∗-algebra. If on
top of beiing unital, F happens to be an operator algebra as well, then
the C∗-subalgebra of (E,◦) generated by F , gives the C∗-envelope of
F [29, Theorem 15.16].
Theorem 3.6. Let A be a hyperrigid operator algebra which possesses a
contractive approximate unit {ei}i∈I consisting of selfadjoint operators.
Let α : G → AutA be a continuous action of a locally compact group.
Then
(8)
C∗
env (A ⋊r
α G) ≃ C∗
env(A) ⋊r
α G
14
and
(9)
via canonical embeddings.
Proof. Let ρ : C∗
sentation and let
E.G. KATSOULIS AND C. RAMSEY
C∗
env(A),α G(cid:1) ≃ C∗
env(cid:0)A ⋊C∗
env(A) → B(H) be a faithful (non-degenerate) repre-
env(A) ⋊α G
ρ : C∗
env(A) −→ B(cid:0)H ⊗ L2(G)(cid:1)
u : G −→ B(cid:0)H ⊗ L2(G)(cid:1)
so that ρ ⋊ u (which we will denote as π) is the regular representation
induced by ρ. (See [34, Section 2.2] for notation and additional infor-
mation.) Since ρ is non-degenerate, [34, Lemma 2.17] implies that the
induced representation π = ρ ⋊ u is also non-degenerate.
Let
ϕ : B(cid:0)H ⊗ L2(G)(cid:1) −→ B(cid:0)H ⊗ L2(G)(cid:1)
be a completely contractive idempotent map whose range is the injec-
α G)1. Let D be the closed (selfadjoint) subalge-
tive envelope of π (A ⋊r
bra of A generated by {ei}i∈I. Then Cc(G,D) is a selfadjoint subalgebra
of A ⋊r
(10)
α G and so [5, 1.3.12] implies that
ϕ (Sπ(f )) = ϕ(S)π(f ),
for any S ∈ B(cid:0)H ⊗ L2(G)(cid:1) and f ∈ Cc(G,D). In particular
ϕ(cid:0)ρ(a)π(f )(cid:1) = ϕ(ρ(a))π(f ),
for all a ∈ A, f ∈ Cc(G,D). On the other hand,
ρ(a)π(f ) = π(af ) ∈ π(A ⋊r
α G)
and so ϕ(ρ(a)π(f )) = ρ(a)π(f ), a ∈ A, f ∈ Cc(G,D). Hence
(cid:0)ϕ(ρ(a)) − ρ(a)(cid:1)π(f ) = 0, for all f ∈ Cc(G,D).
By Lemma 3.5(ii), π(cid:0)Cc(G,D)(cid:1) acts non-degenerately on H ⊗ L2(G)
and so
(11)
ϕ(ρ(a)) = ρ(a), for all a ∈ A.
Hence the mapping ϕ is a completely positive and completely contrac-
tive extension of the identity map on ρ(A). However ρ(A) is hyper-
rigid and according to the discussion in the beginning of the section,
the identity map on ρ(A) is the only such completely contractive and
completely positive extension to C∗
env(A)(cid:1). Therefore
env(ρ(A)) = ρ(cid:0)C∗
env(A).
ϕ(ρ(c)) = ρ(c), for all c ∈ C∗
THE NON-SELFADJOINT APPROACH TO THE HAO-NG ISOMORPHISM 15
Appealing again to (10), with S = ρ(c), we obtain
for all c ∈ C∗
ϕ(π(cf )) = ϕ(cid:0)ρ(c)π(f )(cid:1) = ϕ(ρ(c))π(f ) = ρ(c)π(f ) = π(cf ),
env(A) and f ∈ Cc(G,D). By Lemma 3.5(i) we have
ϕ(S) = S, for all S ∈ π(cid:0)C∗
π(cid:0)C∗
α G(cid:1).
α G(cid:1)1 ⊆ ϕ(cid:0)B(cid:0)H ⊗ L2(G)(cid:1)(cid:1)
env(A) ⋊r
env(A) ⋊r
But this implies that the Choi-Effros multiplication on
. Hence
env(A) ⋊r
is actually the original one coming from B(cid:0)H ⊗ L2(G)(cid:1) and so the
C∗-algebra generated by
π(A ⋊r
α G(cid:1)1
env(cid:0)π(A ⋊r
env(A) ⋊r
α G)1 ⊆ ϕ(cid:0)B(cid:0)H ⊗ L2(G)(cid:1)(cid:1)
equals π(cid:0)C∗
α G(cid:1)1.
env(A) ⋊r
α G)1(cid:1) = π(cid:0)C∗
Furthermore, the C∗-algebra generated by π(A ⋊r
α G) ⊆ π(cid:0)C∗
G(cid:1)1
α G(cid:1). This establishes (8).
equals π(cid:0)C∗
In order to prove (9), let this time π := ρ ⋊ u, where (ρ, u) is the
env(A),G, α). With this π, a
universal covariant representation of (C∗
verbatim repetition of the proof of (8) establishes (9).
env(A) ⋊r
C∗
α
α G,C⋊r
α G) is the completion of Cc(G, X) and Cc(G,C) in T +
Now we turn to crossed product correspondences. Let G be a locally
compact group acting on a non-degenerate C∗-correspondence (X,C)
by a generalized gauge action α : G → Aut(TX ), i.e., αs(X) = X and
αs(C) = C, for all g ∈ G. The reduced crossed product correspondence
(X ⋊r
⋊r
α G,
which can be thought of as living in T +
⋊r
α G but equivalently can be
considered as living in OX ⋊r
α G. The left and right module actions are
given by multiplication and hS, Ti = S∗T , for S, T ∈ Cc(G, X).
In a similar manner, one defines the full crossed product corre-
spondence (X ⋊α G,C ⋊α G) by completing the spaces in TX ⋊α G.
This was shown to be unitarily equivalent to the abstract characteri-
zation of the full crossed product correspondence in [22, Remark 7.8].
Lastly, we recall the definition of the crossed product correspondence
(X ⋊α G,C ⋊α G) which is the completion of the spaces in OX ⋊α G. In
general, it is unknown whether these two correspondences are unitarily
equivalent or not.
X
X
Our next result has been established in [22] in the case where G is
discrete. In [22] it was also noted that the proof carries over to the
general locally compact case. We are about to explain how this is done.
16
E.G. KATSOULIS AND C. RAMSEY
In the proof we will use the language of product systems, which we now
discuss briefly.
Let G be a countable group with unit e ∈ G and let P ⊆ G be a
positive cone. A product system X = {Xp}p∈P over (G, P ) consists of a
C∗-algebra Xe ⊆ B(K), K a Hilbert space, and a family of (represented)
Xe-correspondences Xp ⊆ B(K), p ∈ P\{e}, satisfying the semigroup
rule span{XpXq} = Xpq, for all p, q ∈ P . For instance, if (X,C) is
a (represented) C∗-correspondence, then by taking G = Z, P = Z+
0 ,
X0 = C and Xn = span{X n}, n = 1, 2, . . . , we obtain a product
system {Xn}∞
0 ). As with C∗-correspondences, one can
define product systems abstractly but we will not do that here. See
[11] for more details.
n=0 over (Z, Z+
If X = {Xp}p∈P is a poduct system over (G, P ), then an isomet-
ric (Toeplitz) representation ψ = {ψp}p∈P of X on a Hilbert space
H consists of a ∗-representation ψe : Xe → B(H) and isometric C∗-
correspondence representations ψp : Xp → B(H), p ∈ P , so that
ψpq(xpxq) = ψp(xp)ψq(xq) for all xp ∈ Xp, xq ∈ Xq, p, q ∈ P .
If
(ρ, t) is a representation of a C∗-correspondence (X,C) on H, then by
setting ψ0 = ρ and
ψn : Xn −→ B(H); x1x2 . . . xn 7−→ t(x1)t(x2) . . . t(xn)
we obtain a representation of the product system {Xn}∞
n=0 discussed
in the previous paragraph. These representations are exactly the com-
pactly aligned, Nica covariant representations of {Xn}∞
n=0 in the lan-
guage of [11] and so the Toeplitz C∗-algebra of {Xn}∞
n=0 generated by
the theory of [11] coincides with our TX.
The use of the product system language in the proof of the next re-
sult allows us to import an important result from the theory of product
systems, Fowler's Theorem [11, Theorem 7.2]. This result has no ana-
logue within the theory of C∗-correspondences as it allows us to check
whether a representation of the Toeplitz algebra of a C∗-correspondence
is faithful without using gauge actions.
Theorem 3.7. Let G be a locally compact group acting by a generalized
gauge action α on a non-degenerate C∗-correspondence (X,C). Then
Therefore,
X
T +
env(cid:0)T +
X
C∗
⋊r
α G ≃ T +
X⋊r
α G.
⋊r
α G(cid:1) ≃ OX⋊r
α G.
Proof. Let ρ : TX → B(H) be some faithful ∗-representation and let
V ∈ B(ℓ2(Z+
0 )) be the forward shift. As we did earlier, set Xn := X n
THE NON-SELFADJOINT APPROACH TO THE HAO-NG ISOMORPHISM 17
for n ≥ 1, X0 := C and for the rest of the proof let X denote the
product system {Xn}∞
n=0. Then the map
Xn ∋ x 7−→ ρ(x) ⊗ V n ∈ B(H ⊗ l2(Z+
0 )), n ∈ Z+
0 ,
0 )).
Since X ⋊r
α G, n ∈ Z+
0 .
is a representation of X that satisfies the requirements of Fowler's The-
orem [11, Theorem 7.2]. Therefore it establishes a faithful representa-
tion π : TX → B(H ⊗ ℓ2(Z+
α G ⊆ TX ⋊r
α G, we may consider the regular representation
Indπ, when restricted on X ⋊r
α G, as a representation of the product
system X ⋊r
α G, which we denote as ψ. We furthermore write ψn :=
ψ Xn⋊r
We claim that ψ satisfies the requirements of Fowler's Theorem [11,
Theorem 7.2]. Indeed let Qn ∈ l2(Z+
0 ) be the projection on the one
dimensional subspace corresponding to the characteristic function of
0 and let Qn ≡ (I ⊗ Qn) ⊗ I ∈ B(cid:0)H ⊗ l2(Z+
n ∈ Z+
(constant) B(H⊗l2(Z+
to any s ∈ G. Then given any f ∈ Cc(G, Xn), n ≥ 1, we have
(cid:0) Q0Indπf(cid:1)h(t) = Q0ZG
t (f (s))(cid:1)h(s−1t)ds
0 ) ⊗ L2(G)(cid:1) be the
0 ))-valued function that assigns the value I ⊗Qn
π(cid:0)α−1
t (f (s))) ⊗ V n(cid:1)h(s−1t)ds
= (I ⊗ Q0) ⊗ IZG(cid:0)ρ(α−1
=ZG(cid:0)ρ(α−1
the product Qn∈F \{0}(I − P ψ
n , always dominates Q0 and so
factor of the form I − P ψ
(I − P ψ
(12)
n denotes the range space of ψ(Xn ⋊r
t (f (s))) ⊗ Q0V n(cid:1)h(s−1t)ds = 0
α G), n ∈ Z+
0 , then
n ), which in our case collapses to a single
Therefore if P ψ
n )(cid:13)(cid:13) ≥(cid:13)(cid:13)ψ0(f ) Q0(cid:13)(cid:13) =(cid:13)(cid:13)Indπ(f ) Q0(cid:13)(cid:13),
for any f ∈ Cc(G,C). However, each I ⊗ Qn reduces π C and therefore
(I ⊗ Qn)π(cid:1) C≃(cid:0) ⊕ (I ⊗ Q0)π(cid:1) C,
i.e., the restriction of π on C is unitarily equivalent to a direct sum
of countably many copies of (I ⊗ Q0)π restricted on C. From this we
obtain,
(cid:13)(cid:13)ψ0(f ) Yn∈F \{0}
π C≃(cid:0) ⊕n∈Z+
0
ψ0 = Indπ C⋊r
α G ≃ ⊕IndQ0π C⋊r
Combining the above with (12) we now obtain
α G .
α G ≃ ⊕ Q0Indπ C⋊r
n )(cid:13)(cid:13) =(cid:13)(cid:13)Indπ(f ) Q0(cid:13)(cid:13) = kψ0(f )k
(cid:13)(cid:13)ψ0(f ) Yn∈F \{0}
(I − P ψ
18
E.G. KATSOULIS AND C. RAMSEY
Since the claim is valid, Fowler's Theorem shows now that the in-
α G. Note
for any f ∈ Cc(G,C), which establishes the claim.
duced representation ψ∗ is a faithful representation of TX⋊r
now that ψ∗(T +
ψ∗(X ⋊r
α G is equal to the closed linear span of
α G) ≃ T +
α G) = [n∈Z+
⋊r
α G is also isomorphic to the closed linear span of
α G) = [n∈Z+
IndπCc(G, Xn).
However, T +
X
ψn(Xn ⋊r
X⋊r
X⋊r
0
0
IndπCc(G, Xn).
[n∈Z+
0
Hence, T +
X
Finally
⋊r
α G.
X⋊r
α G ≃ T +
C∗
env(cid:0)T +
X
⋊r
α G(cid:1) ≃ C∗
with the last identification following from [21, Theorem 3.7].
env(T +
X⋊r
α G) ≃ OX⋊r
α G.
Remark 3.8. The use of the language of product systems in the pre-
vious proof has an additional benefit. By switching from (Z, Z+
0 ) to an
arbitrary totally ordered group (G, P ), the same exact proof as above
establishes the more general result NT +
α G, where NT +
X
denotes the Nica tensor algebra of X. Actually the proof works for any
quasi-lattice ordered group (G, P ) provided that certain issues involv-
ing compact alignment and Nica covariance are worked out first. We
are recording this fact here for future reference.
α G ≃ NT +
X⋊r
⋊r
X
In [20] the Hao-Ng isomorphism problem was resolved for the re-
duced crossed product and all discrete groups. In the next result we
address the case of an arbitrary locally compact group and we resolve
the Hao-Ng problem for the reduced crossed product provided that the
C∗-correspondence is hyperrigid. Note that our result subsumes an
earlier result [24, Proposition 5.5] which was posted on the arXiv but
has not appeared in print.
Theorem 3.9. Let G be a locally compact group acting by a general-
ized gauge action α on a non-degenerate hyperrigid C∗-correspondence
(X,C), e.g. ϕX(JX)X = X. Then
OX ⋊r
Proof. By Theorem 3.7 we have C∗
other hand, Theorem 3.6 implies that C∗
Hence OX ⋊r
α G ≃ C∗
env(T +
X ) ⋊r
α G.
α G ≃ OX⋊r
env(cid:0)T +
env(cid:0)T +
α G ≃ OX⋊r
X
α G.
X
⋊r
⋊r
α G(cid:1) ≃ OX⋊r
α G(cid:1) ≃ C∗
α G. On the
env(T +
α G.
X )⋊r
THE NON-SELFADJOINT APPROACH TO THE HAO-NG ISOMORPHISM 19
It is important to us that an analogous results holds for the full
crossed product.
Theorem 3.10. Let G be a locally compact group acting by a general-
ized gauge action α on a non-degenerate hyperrigid C∗-correspondence
(X,C). Then
T +
X
OX ⋊α G ≃ OX ⋊α G.
Proof. In [22, Theorem 7.13] we proved that
env(cid:0)T +
env(T +
x ),α G(cid:1) ≃ C∗
⋊OX ,α G ≃ T +
env(cid:0)T +
Now (9) in Theorem 3.6 and the above imply that
and C∗
OX ⋊α G ≃ C∗
C∗
and the conclusion follows.
⋊
X ⋊α G
X
env(T +
X
⋊OX ,α G(cid:1) ≃ OX ⋊α G.
X ) ⋊α G ≃ OX ⋊α G
At this point one might think that the above theorem is the final word
regarding the Hao-Ng isomorphism for hyperrigid correspondences. As
it turns out, this couldn't be further from the truth. It is indeed the
case that we have expressed the crossed product OX ⋊αG as the Cuntz-
Pimsner algebra of a C∗-correspondence, namely X ⋊α G, but this is
not the C∗-correspondence that the authors of [3] ask for. Is this a big
deal? Most definitely yes, and we devote the next section explaining
the reasons why.
4. Isometric coextensions
X
The goal of this section is to answer Problem 3 in Chapter 8 of [22]:
Is T +
⋊α G the tensor algebra of some C∗-correspondence? The (affir-
mative) answer is one of the key ingredients in the proof of Theorem 4.9,
one of the central results of the paper.
Definition 4.1. Let (X,C) be a C∗-correspondence and G a locally
compact group. A generalized gauge action α : G → Aut((X,C)) is a
map from G into the completely isometric module automorphisms. In
particular, for each s ∈ G, αs is an isometric automorphism of X and
a ∗-automorphism of C such that
αs(ξc) = αs(ξ)αs(c), αs(ϕX(c)ξ) = ϕX(αs(c))αs(ξ)
for all ξ, η ∈ X and c ∈ C.
and αs(hξ, ηi) = hαs(ξ), αs(η)i
Earlier we said that α : G → AutTX forms a generalized gauge action
of TX if αs(X) = X and αs(C) = C, for all g ∈ G. The following result
says that the two definitions are equivalent and it was observed in [20,
pg 5760]. (See also [14, Lemma 2.6] for the analogous result with OX).
20
E.G. KATSOULIS AND C. RAMSEY
Proposition 4.2. Let (X,C) be a non-degenerate C∗-correspondence
and G a locally compact group. If α : G → Aut(TX) is a generalized
gauge action of TX then it restricts to a generalized gauge action of
(X,C). Conversely, a generalized gauge action α of (X,C) extends
uniquely to a generalized gauge action of TX .
The fundamental object of study in this section is the C∗-correspon-
dence dynamical system ((X,C),G, α) which is given by a non-degenerate
C∗-correspondence (X,C) and a generalized gauge action α of the lo-
cally compact group G acting on (X,C).
Definition 4.3. A representation of the C∗-correspondence dynami-
cal system ((X,C),G, α) is a quadruple (ρ, t, u,H) consisting of a com-
pletely contractive representation (ρ, t,H) of (X,C) and a strongly con-
tinuous unitary representation u : G → U(H) satisfying the covariance
relations
u(s)t(ξ) = t(αs(ξ))u(s) and u(s)ρ(c) = ρ(αs(c))u(s)
for all s ∈ G, ξ ∈ X and c ∈ C. Moreover, (ρ, t, u,H) is said to be
isometric if (ρ, t,H) is an isometric (Toeplitz) representation of (X,C).
The following theorem is an extension of [28, Theorem 2.12].
Theorem 4.4. Let ((X,C),G, α) be a C∗-correspondence dynamical
system. The isometric representations (ρ, t, u,H) of ((X,C),G, α) are
in bijective correspondence with the isometric representations (π, u,H)
of (TX ,G, α). Specifically, they are related by π = ρ ⋊ t.
Proof. If (π, u,H) is an isometric representation of (TX,G, α) then
[28, Theorem 2.12] proves that there exists an isometric representation
(ρ, t,H) of (X,C) such that ρ ⋊ t = π. Proposition 4.2 gives that α is a
generalized gauge action of G on (X,C) and so the covariance relations
between ρ, t and u are automatic. Hence, (ρ, t, u,H) is an isometric
representation of ((X,C),G, α).
Conversely, suppose (ρ, t, u,H) is an isometric representation of
((X,C),G, α). Again by [28, Theorem 2.12] this gives that (ρ ⋊ t,H)
is an isometric representation of TX and by Proposition 4.2 α ex-
tends uniquely to a generalized gauge action of G on TX. Because
u(s)(ρ ⋊ t(·))u(s)∗ and (ρ ◦ αs) ⋊ (t ◦ αs) = (ρ ⋊ t) ◦ αs agree on X and
C then the uniqueness of [28, Theorem 2.12] gives that
u(s)ρ ⋊ t(a) = ρ ⋊ t(αs(a))u(s)
for all s ∈ G and a ∈ TX. Therefore, (ρ ⋊ t, u,H) is an isometric
representation of (TX ,G, α).
THE NON-SELFADJOINT APPROACH TO THE HAO-NG ISOMORPHISM 21
If (ρ, t, u,H) and (ρ1, t1, u1,H1) are completely contractive represen-
tations of ((X,C),G, α) we say that the latter is a dilation of the former
if H ⊆ H1 and
(i) H reduces π1 and u1 with π1(c)H = π(c), c ∈ C and u1(g)H =
(ii) H is a semi-invariant subspace for t1 and PHt1(ξ)H = t(ξ), ξ ∈
u(g), g ∈ G, and
X.
We call such a dilation an extension if H is an invariant subspace for
t1 and a coextension if H is a coinvariant subspace for t1.
The dilation (ρ1, t1, u1,H1) is called minimal when H1 is the smallest
reducing subspace for t1 containing H.
Now we need a lemma relating to the step-by-step dilation tech-
niques of Muhly and Solel of [28, Section 3]. Their proof is modelled
after Popescu's step-by-step dilation technique [30] using the Schaeffer
matrix construction [32]. The following is also a slight simplification
of the original proof in [28].
To this end suppose (ρ, t, u,H) is a completely contractive represen-
tation of the C∗-correspondence dynamical system ((X,C),G, α). The
ultimate goal is to prove that every such representation dilates to an
isometric representation.
As in [28], define the Hilbert space HX = X ⊗ρ H
h·,·i
where
ξc ⊗ h = ξ ⊗ ρ(c)h and hξ ⊗ h, η ⊗ ki = hh, ρ(hξ, ηi)ki
for ξ, η ∈ X, c ∈ C and h, k ∈ H. As well, define σX : X → B(H,HX)
by σX (ξ)h = ξ ⊗ h and t : HX → H by t(ξ ⊗ h) = t(ξ)h. From here
one defines the one step dilation to H1 = H ⊕ HX given by
t1(ξ) =(cid:20)
and
t(ξ)
(I − t∗t)1/2σX(ξ) 0 (cid:21)
0
ρ1(c) =(cid:20) ρ(c)
0
0
ρ(c) (cid:21)
where ρ : C → B(HX ) is given by ρ(c)(ξ ⊗ h) = ϕX (c)ξ ⊗ h.
Lemma 4.5. Consider
u1(s) =(cid:20) u(s)
0
0
u(s) (cid:21)
22
E.G. KATSOULIS AND C. RAMSEY
where u : G → B(HX) is given by u(s)(ξ ⊗ h) = αs(ξ) ⊗ u(s)h, which
is well-defined. Then (ρ1, t1, u1,H1) is a completely contractive repre-
sentation of ((X,C),G, α) such that
t1(ξ)∗t1(η) =(cid:20) ρ(hξ, ηi) 0
0 (cid:21)
0
for all ξ, η ∈ H.
Proof. By [28, Lemma 3.7] (ρ1, t1) is a completely contractive repre-
sentation of (X,C) on H1 which satisfies the last two statements in the
lemma.
First note that u is in fact well-defined since it respects the internal
C-modularity of HX,
u(s)(ξc ⊗ h) = αs(ξc) ⊗ u(s)h
= αs(ξ)αs(c) ⊗ u(s)h
= αs(ξ) ⊗ ρ(αs(c))u(s)h
= αs(ξ) ⊗ u(s)ρ(c)h
= u(s)(ξ ⊗ ρ(c)h),
for all s ∈ G, c ∈ C, ξ ∈ X, h ∈ H.
and h, k ∈ H we have that
Additionally, observe that u(s) is unitary since for all s ∈ G, ξ, η ∈ X
hu(s)(ξ ⊗ h), u(s)(η ⊗ k)i = hαs(ξ) ⊗ u(s)h, αs(η) ⊗ u(s)ki
= hu(s)h, ρ(hαs(ξ), αs(η)i)u(s)ki
= hu(s)h, ρ(αs(hξ, ηi))u(s)ki
= hu(s)h, u(s)ρ(hξ, ηi)ki
= hh, ρ(hξ, ηi)ki
= hξ ⊗ h, η ⊗ ki.
Next we need to make the following calculations:
σX(αs(ξ))u(s)h = αs(ξ) ⊗ u(s)h
= u(s)(ξ ⊗ h)
= u(s)σX (ξ)h
THE NON-SELFADJOINT APPROACH TO THE HAO-NG ISOMORPHISM 23
and
ht∗tσX (αs(ξ))u(s)h, η ⊗ ki = ht(αs(ξ) ⊗ u(s)h), t(η ⊗ k)i
= ht(αs(ξ))u(s)h, t(η)ki
= hu(s)t(ξ)h, t(η)ki
= ht(ξ)h, t(αs−1(η))u(s)∗ki
= ht∗t(ξ ⊗ h), αs−1(η) ⊗ u(s−1)ki
= hu(s)t∗t(ξ ⊗ h), η ⊗ ki
= hu(s)t∗tσX (ξ)h, η ⊗ ki.
Combining these one gets
u(s)(I − t∗t)σX (ξ) = (I − t∗t)σX(αs(ξ))u(s)
= (I − t∗t)u(s)σX(ξ).
Hence,
u(s)(I − t∗t) = (I − t∗t)u(s)
and by a standard trick often attributed to Halmos
u(s)(I − t∗t)1/2 = (I − t∗t)1/2 u(s).
Now, we need to establish the covariance relations between (ρ1, t1)
and u1. From the previous paragraph we have that
u(s)(I − t∗t)1/2σX (ξ) = (I − t∗t)1/2σX(αs(ξ))u(s)
and thus u1(s)t1(ξ) = t1(αs(ξ))u1(s).
Second, it is much more straightforward to calculate that
u(s)ρ(c)(ξ ⊗ h) = u(s)(ϕX(c)ξ ⊗ h)
= αs(ϕX(c)ξ) ⊗ u(s)h
= ϕX (αs(c))αs(ξ) ⊗ u(s)h
= ρ(αs(c))u(s)(ξ ⊗ h).
Therefore, u1(s)ρ1(c) = ρ1(αs(c))u1(s) and the conclusion follows.
Theorem 4.6. Every completely contractive representation of the non-
degenerate C∗-correspondence dynamical system ((X,C),G, α) has a
minimal isometric coextension. Moreover, the minimal isometric coex-
tension is unique up to unitary equivalence.
Proof. Let (ρ, t, u,H) be a completely contractive representation of
((X,C),G, α). Following the proof of [28, Theorem 3.3] repeatedly use
Lemma 4.5 to get a sequence of completely contractive representations
24
E.G. KATSOULIS AND C. RAMSEY
(ρn, tn, un,Hn) of ((X,C),G, α) in the obvious manner: use (ρ, t, u,H)
to produce (ρ1, t1, u1,H1) and then recursively
(ρn, tn, un,Hn) = ((ρn−1)1, (tn−1)1, (un−1)1, (Hn−1)1)
from the previous lemma and the discussion preceding it.
Let H′ = ∪n≥1Hn and define ρ′ = lim−→ ρn, t′ = lim−→ tn and u′ = lim−→ un.
By [28, Theorem 3.3] (ρ′, t′,H′) is an isometric representation of (X,C).
Note that u′ is a strongly continuous unitary representation of G since
it is the direct sum of such representations.
Now to the covariance relations:
u′(s)ρ′(c)PHn = un(s)ρn(c)PHn
= ρn(αs(c))un(s)PHn
= ρ′(αs(c))u′(s)PHn
for all s ∈ G, c ∈ C and n ∈ N. As well,
u′(s)t′(ξ)PHn = un+1tn+1(ξ)PHn
= tn+1(αs(ξ))un+1(s)PHn
= tn+1(αs(ξ))un(s)PHn
= t′(αs(ξ))u′(s)PHn
for all s ∈ G, ξ ∈ X and n ∈ N. Therefore, the covariance relations are
satisfied and (ρ′, t′, u′,H′) is an isometric coextension of (ρ, t, u,H).
Now let
K = span{t′(ξ1)· · · t′(ξn)h : ξi ∈ X, h ∈ H, n ≥ 1} ⊂ H′.
Because of the covariance relations of (ρ′, t′, u′,H′) one can see that K is
a reducing subspace of (ρ′, t′, u′,H′) that contains H. Thus, (ρ′, t′, u′,K)
is a minimal isometric coextension of (ρ, t, u,H).
In regard to uniqueness, suppose that (ρ′′, t′′, u′′,H′′) is another iso-
metric coextension of (ρ, t, u,H).
[28, Proposition 3.2] proves that
there exists a unitary W : K → H′′ such that W ρ′(·) = ρ′′(·)W ,
W t′(·) = t′′(·)W and W h = h, for all h ∈ H. Now
W u′(s)t′(ξ1)· · · t′(ξn)h = W t′(αs(ξ1))· · · t′(αs(ξn))u′(s)h
= t′′(αs(ξ1))· · · t′′(αs(ξn))W u(s)h
= t′′(αs(ξ1))· · · t′′(αs(ξn))u(s)h
= t′′(αs(ξ1))· · · t′′(αs(ξn))u′′(s)W h
= u′′(s)t′′(ξ1)· · · t′′(ξn)W h
= u′′(s)W t′(ξ1)· · · t′(ξn)h
THE NON-SELFADJOINT APPROACH TO THE HAO-NG ISOMORPHISM 25
Therefore, by minimality W u′(·) = u′′(·)W and so (ρ′, t′, u′,K) and
(ρ′′, t′′, u′′,H′′) are unitarily equivalent.
Theorem 4.7. Let (X,C) be a non-degenerate C∗-correspondence and
let α be a generalized gauge action of a locally compact group G. Then
T +
X
X⋊αG
⋊α G ≃ T +
X
⋊TX ,α G ≃ T +
X
X
⋊TX ,α G ≃ T +
Towards proving the remaining isomorphism, let ϕ : T +
Proof. It is already proven in [22], in a discussion following Theorem
7.13, that T +
X⋊αG. (It also follows from [3, Theorem 3.1]
as the isomorphism Φ of that theorem maps generators to generators.)
⋊α G →
B(H) be a completely contractive representation. In the same way as in
the proof of [22, Theorem 4.1] one can assume that ϕ is nondegenerate.
Now by [22, Proposition 3.8] there exists a representation (π, u,H) of
(T +
X ,G, α) so that ϕ = π ⋊ u.
By [28, Theorem 3.10] there is a completely contractive representa-
tion, (ρ, t), of (X,C) such that π = ρ ⋊ t. Hence, in the same way as
the first part of the proof of Theorem 4.4, (ρ, t, u,H) is a completely
contractive representation of ((X,C),G, α). By Theorem 4.6 (ρ, t, u,H)
has a unique minimal isometric coextension (ρ′, t′, u′,H′) and thus by
Theorem 4.4 (ρ′ ⋊ t′, u′,H′) is an isometric representation of (TX,G, α).
As discussed in the proof of [28, Theorem 3.10] H ⊂ H′ is sem-
invariant for ρ′ and t′ and thus PHρ′ ⋊ t′H is a completely contractive
representation of T +
X . Moreover, that same theorem gives that
π = ρ ⋊ t = PHρ′ ⋊ t′H
Therefore, every completely contractive representation of T +
because ρ(c) = PHρ′(c)H and t(ξ) = PHt′(ξ)H for all c ∈ C and ξ ∈ X.
⋊α G
dilates to a completely contractive representation of T +
⋊TX ,α G and
thus they are completely isometrically isomorphic.
Corollary 4.8. Let ((X,A),G, α) be a C∗-correspondence dynamical
system and assume that JX = {0}. Then all relative crossed prod-
ucts for (T +
X ,G, α) are canonically isomorphic via completely isometric
maps.
X
X
In particular, the above applies to the non-commutative disc algebra
A∞ ⊆ O∞. To obtain the same conclusion for the non-commutative
disc algebras An, n < ∞, we need to work much harder. (See the next
section.)
For the moment, we can put together all previous results to obtain
the following, which summarizes our knowledge on the Hao-Ng isomor-
phism problem for the full crossed product.
26
E.G. KATSOULIS AND C. RAMSEY
Theorem 4.9. Let ((X,A),G, α) be a non-degenerate C∗-correspondence
dynamical system. Then the following two statements are equivalent
X
(i) C∗
env(T +
generators to generators,
⋊α G) ≃ OX ⋊α G via a ∗-isomorphism that sends
(ii) OX ⋊α G ≃ OX⋊α G via a ∗-isomorphism that sends generators
to generators, (Hao-Ng isomorphism)
and both imply
(iii) all relative crossed products for (T +
metrically isomorphic via canonical maps.
X ,G, α) are completely iso-
If (X,C) is hyperrigid, e.g., ϕX (JX)X = X, then all of the above
statements are equivalent.
Proof. Assume first that
(13)
env(T +
C∗
X
⋊α G) ≃ OX ⋊α G
canonically. Theorem 4.7 shows now that
T +
X
⋊α G ≃ T +
X
⋊TX ,α G ≃ T +
X⋊αG
canonically and so by taking C∗-envelopes we have a canonical isomor-
phism
(14)
C∗
X
env(cid:0)T +
⋊α G(cid:1) ≃ C∗
env(cid:0)T +
X
⋊TX ,α G(cid:1) ≃ OX⋊αG.
By "equating" the right sides of (13) and (14), we obtain (ii).
Conversely, assume that (ii) holds. Then by taking C∗-envelopes in
the isomorphisms of Theorem 4.7 we obtain
C∗
env(T +
env(T +
⋊TX ,αG) ≃ C∗
by (ii). Therefore (i) holds, as desired.
⋊αG) ≃ C∗
env(T +
X
X
X⋊αG) ≃ OX⋊αG ≃ OX ⋊α G
Assume now that (ii) is valid, i.e., the Hao-Ng isomorphism is im-
plemented via a canonical map. The same map establishes
(15)
T +
X
⋊OX ,α G ≃ T +
X⋊α G.
By Theorem 4.7 we also have
T +
(16)
From (15) and (16), we obtain T +
⋊α G ≃ T +
T +
X
X
X
X
⋊α G ≃ T +
X⋊α G.
⋊α G ≃ T +
env(T +
C∗
⋊
X
X ),α G
⋊OX ,α G, or,
canonically. By Theorem 2.4, all relative crossed products for (T +
are canonically isomorphic, which is (iii).
X ,G, α)
THE NON-SELFADJOINT APPROACH TO THE HAO-NG ISOMORPHISM 27
Assume now that (X,C) is hyperrigid and all relative crossed prod-
X ,G, α) are canonically isomorphic. Therefore (9) in Theo-
ucts for (T +
rem 3.6 implies
X
env(cid:0)T +
C∗
⋊OX ,α G ≃ T +
⋊OX ,α G(cid:1) ≃ OX ⋊α G.
⋊α G and so (i) is valid.
By assumption T +
with ϕX(JX)X = X is always hyperrigid.
X
X
Finally recall that Theorem 3.1 shows that a C∗-correspondence X
The importance of the previous result can not be understated. First,
note that condition (i) in Theorem 4.9 is just the equivalence
env(A ⋊α G) ≃ C∗
C∗
env(A) ⋊α G
of [22, Problem 1] with A = T +
X . Hence the equivalence of (i) and (ii) in
Theorem 4.9 shows that the Hao-Ng isomorphism and Problem 1 in [22]
are actually equivalent problems in the context of tensor algebras of C∗-
correspondences. Furthermore, if the Hao-Ng isomorphism holds, then
we automatically have from (iii) that OX⋊α G ≃ OX ⋊α G. Therefore a
positive resolution for the Hao-Ng isomorphism conjecture also implies
a positive resolution for the modified conjecture of [22, page 70].
Also notice that according to condition (iii), the verification of the
Hao-Ng isomorphism for hyperrigid C∗-correspondences depends on the
canonical identification of two non-selfadjoint operator algebras. We
pursue this direction successfully in the next section where we verify
the Hao-Ng isomorphism for all graph correspondences of row finite
graphs. Furthermore, unlike condition (ii) (Hao-Ng isomorphism), both
conditions (i) and (iii) are applicable to arbitrary dynamical systems
(T +
X ,G, α), i.e., α does not have to be a gauge action. Thus in a sense,
a generalization of the Hao-Ng isomorphism problem beyond the realm
of gauge actions is possible but only in the language of non-selfadjoint
operator algebras. In light of the recent results of Harris and Kim [15],
this seems to be a direction worth pursuing.
5. Graph correspondences
Following from the last section, we would like to prove that every iso-
metric (Toeplitz) representation of ((X,C),G, α) dilates to a covariant
(Cuntz-Pimsner) representation. However, the standard proofs that
the C∗-envelope of the tensor algebra is the Cuntz-Pimsner algebra
are non-constructive [21, 28] which at the moment is a barrier to our
method of proof. Significantly.
in the case of graph correspondences
such a constructive dilation proof is shown to exist.
28
E.G. KATSOULIS AND C. RAMSEY
Let (E, V, s, r) be a directed graph, where both E and V are sepa-
rable, with associated graph correspondence (X,C, ϕX ). Recall this is
where C = c0(V ), X is the completion of cc(E) under the right module
structure
0,
hαδe, βδfi =(cid:26) αβδs(e), e = f
δe · δv =(cid:26) δe, s(e) = v
otherwise
and the left action of C on X is given by
ϕX(δv)δe =(cid:26) δe, r(e) = v
otherwise
0,
otherwise
0,
,
.
When (X,C) is the graph correspondence for a directed graph (V, E)
then OX is ∗-isomorphic to the Cuntz-Krieger algebra of the graph.
We wish to find Cuntz-Pimsner representations of these graph cor-
respondences. As usual, the main concern is looking at which elements
of C are mapped into K(X) by ϕX.
In the case of a graph correspondence, Raeburn [31, Proposition 8.8]
gives that ϕX(δv) ∈ K(X) if and only if r−1(v) < ∞. Furthermore,
δv ∈ ker ϕX if and only if r−1(v) = ∅. Thus, let
Vfin = {v ∈ V : 1 ≤ r−1(v) < ∞}
be the set of vertices generating Katsura's ideal JX and let
K = {(v, w) ∃e ∈ E with s(e) = w, r(e) = v, v ∈ Vfin}.
For each pair (v, w) ∈ K, let E(v, w) be the collection of all edges
starting from w and ending on v and let [E((v, w)] = CE(v,w).
In
what follows we will identify the canonical basis of [E(v, w)] with the
elements of E(v, w) and use the same symbol for both.
Suppose α is a generalized gauge action of (X,C) then it is clear
that this induces a permutation of V and in particular that Vfin is
invariant under this permutation. By abuse of notation we call this
permutation α : V → V . Furthermore, the action α maps [E(v, w)]
Indeed, if E(v, w) = {e1, e2, . . . , en}
unitarily onto [E(α(v), α(w))].
THE NON-SELFADJOINT APPROACH TO THE HAO-NG ISOMORPHISM 29
and ξ =Pn
i=1 ciδei, then
n
deδej(cid:17)E
n
n
n
n
Xj=1
cjδejE(cid:17)
ciδei,
hα(ξ), α(ξ)i =Dα(cid:16)
ciδei(cid:17), α(cid:16)
Xi=1
= α(cid:16)D
Xi=1
Xj=1
= α(cid:16)
ci2δw(cid:17)
Xi=1
Xi=1
ci2δα(w) = hξ, ξi .
=
n
Proposition 5.1. Let (X,C) be the graph correspondence of (E, V )
and suppose (ρ, t, u,H) is a completely contractive representation of the
dynamical system ((X,C),G, α). There exists a dilation to a completely
contractive representation (ρ1, t1, u1,H1) such that for every v ∈ Vfin
ρ(δv) = Xe∈r−1(v)
t1(δe)t1(δe)∗
Proof. By [28, Lemma 3.5] because (ρ, t,H) is completely contractive
then for v ∈ VK we have the matrix inequality
[t(e)∗t(f )]e,f ∈r−1(v) ≤ [ρ(he, fi]e,f ∈r−1(v) = ⊕e∈r−1(v)ρ(δs(e))
and so [t(e) : e ∈ r−1(v)] is a row contraction. Hence,
and so we can define
∆v :=
t(δe)t(δe)∗
ρ(δv) ≥ Xe∈r−1(v)
pr−1(v)(cid:16)ρ(δv) − Xe∈r−1(v)
1
t(δe)t(δe)∗(cid:17)1/2
.
Let Hw = ρ(δw)H and so we can assume H = ⊕w∈V Hw. For each
pair (v′, w′) ∈ K let
Hv′,w′ := Hv′ ⊗ [E(v′, w′)].
Then for each w ∈ V we define
and
w ≡ M(v′,w)∈K
H+
Hv′,w = M(v′,w)∈K
Hw,1 = Hw ⊕ H+
w.
Hv′ ⊗ [E(v′, w)].
30
E.G. KATSOULIS AND C. RAMSEY
We combine these to define
Hence
H1 = ⊕w∈V Hw,1.
ρ1(δw) = IHw,1, w ∈ V,
extends to a ∗-homomorphism of C that dilates ρ.
We also have a continuous unitary representation u1 : G → B(H1)
dilating u : G → B(H) and defined as follows.
Given g ∈ G and h ∈ H, we let u1(g)h = u(g)h. Otherwise, on each
H+
w the operator u1(g) is defined by
H+
w ⊇ Hv′,w ∋ h ⊗ ξ 7−→ ug(h) ⊗ αg(ξ) ∈ Hαg (v′),αg(w) ⊆ H+
α(w).
It is easy to see that u1 : G → B(H1) is a continuous unitary represen-
tation dilating u.
We are ready to dilate t : X → B(H). If r(e) /∈ Vfin, then we let
t1(e) = t(e). Otherwise, if e ∈ E with s(e) = w and r(e) = v ∈ Vfin,
then t1(e) ∈ B(H1) has cokernel contained in
Hw ⊕ (. . . 0 ⊕ 0 ⊕ Hv,w ⊕ 0 . . . ) ⊆ Hw ⊕ H+
w ≡ Hw,1
range contained in Hv ⊕ 0 ⊆ Hv ⊕ H+
v and it is given by
t1(δe) = [t(δe) ∆vτ (e)] ∈ B(Hw ⊕ Hv,w,Hv),
where
τ (e) : Hv,w −→ Hv ; h ⊗ ξ 7−→ he, ξi h.
(In general, for ζ ∈ [E(v, w)], τ (ζ) will be given by τ (ζ)(h ⊗ ξ) =
hζ, ξi h.)
It is easy to see that τ (e)τ (e)∗ = IHv and so,
Xe∈r−1(v)
t1(δe)t1(δe)∗ = Xe∈r−1(v)
=(cid:16) Xe∈r−1(v)
= ρ(δv)
t(δe)t(δe)∗ + ∆vτ (e)τ (e)∗∆v
t(δe)t(δe)∗(cid:17) + r−1(v)∆2
v
This establishes one of the main conclusions of this proposition and
gives that
[t1(e)∗t1(f )]e,f ∈r−1(v) ≤ [ρ1(he, fi)]e,f ∈r−1(v)
as in the start of the proof. Thus, by [28, Lemma 3.5] again, (ρ1, t1,H1)
is a completely contractive representation of (X,C).
THE NON-SELFADJOINT APPROACH TO THE HAO-NG ISOMORPHISM 31
Lastly we must establish the covariance relations. For any g ∈ G
recall that αg acts as a unitary between [E(v, w)] = [{e1, e2, . . . , en}]
and [E(αg(v), αg(w))] = [{f1, f2, . . . , fn}]. This implies that
n
n
n
n
t(αg(δei))t(αg(δei))∗ =
Xi=1
=
=
and so
n
Xk=1
(αg)k,iδfk(cid:1)∗
Xi=1
Xj=1
t(cid:0)
(αg)j,iδfj(cid:1)t(cid:0)
Xj,k=1 n
(αg)j,i(αg)k,i! t(δfj )t(δfk )∗
Xi=1
Xj=1
t(δfj )t(δfj )∗
n
u(g)(ρ(δv) − Xe∈r−1(v)
t(δe)t(δe)∗)
t(αg(δe))t(αg(δe))∗(cid:1)u(g)
=(cid:0)ρ(δαg (v)) − Xe∈r−1(v)
=(cid:16)ρ(δαg (v)) − Xw∈s(r−1(v)) Xe∈E(v,w)
=(cid:16)ρ(δαg (v)) − Xw∈s(r−1(αg (v))) Xf ∈E(αg (v),w)
t(δf )t(δf )∗(cid:17)u(g).
=(cid:16)ρ(δαg (v)) − Xf ∈r−1(αg (v))
t(αg(δe))t(αg(δe))∗(cid:17)u(g)
t(δf )t(δf )∗(cid:17)u(g)
By a standard functional analysis trick,
u(g)∆v = ∆αg (v)u(g).
Furthermore, if h ⊗ ξ ∈ Hv,w, then
τ (αg(e))u1(g)(h ⊗ ξ) = τ (αg(e))(u(g)h ⊗ αg(ξ)
= hαg(e), αg(ξ)i u(g)h
= he, ξi u(g)h = u(g)τ (e)(h ⊗ ξ).
Hence,
w (cid:1)
t1(αg(δe))u1(g) =(cid:2)t(αg(δe)) ∆αg (v)τ (αg(e))(cid:3)(cid:0)u1(g) Hw⊕H+
=(cid:2)u(g)t(δe) ∆αg(v)u(g)τ (e)(cid:3)
= [u(g)t(δe) u(g)∆vu(g)τ (e)]
= u1(g)t1(δe).
32
E.G. KATSOULIS AND C. RAMSEY
It is also immediate that
u1(g)ρ1(δv) = ρ1(δαg(v))u1(g).
Therefore, (ρ1, t1, u1,H1) is a completely contractive representation of
((X,C),G, α).
Theorem 5.2. Let (X,C) be the graph correspondence of (E, V ). Every
completely contractive representation of ((X,C),G, α) can be dilated to
a Cuntz-Pimsner representation.
Proof. Let (ρ0, t0, u0,H0) be a completely contractive representation
of ((X,C),G, α). Recursively use the previous proposition to gener-
ate a sequence of completely contractive representations (ρn, tn, un,Hn)
such that for each n ≥ 1, Hn−1 ⊂ Hn, (ρn, tn, un,Hn) is a dilation of
(ρn−1, tn−1, un−1,Hn−1) and for every v ∈ V such that 1 ≤ r−1(v) < ∞
we have
Thus, define H′ = ∪∞
Xe∈r−1(v)
n=0Hn and
tn(δe)tn(δe)∗ = ρn−1(δv).
ρ′(c)Hn = ρn(c), t′(ξ)Hn = tn(ξ), and u′(g)Hn = un(g).
Hence, (ρ′, t′, u′,H′) is a dilation (ρ0, t0, u0,H0) to a completely con-
tractive representation such that for every v ∈ Vfin
n→∞ Xe∈r−1(v)
n→∞ Xe∈r−1(v)
t′(δe)t′(δe)∗ = sot − lim
= sot − lim
Xe∈r−1(v)
t′(δe)IHnt′(δe)∗
tn(δe)tn(δe)∗
= sot − lim
= ρ′(δv).
n→∞
ρn−1(δv)
According to [28, Definition 5.3] the representation (ρ′, t′,H′) is JX-
coisometric, i.e., Cuntz-Pimsner in the sense of Katsura but without
being isometric.
Lastly, by Theorem 4.6 there is a unique minimal isometric coex-
tension of (ρ′, t′, u′,H′) to (ρ′′, t′′, u′′,H′′). By [28, Corollary 5.21] this
coextension process preserves the property of being JX-coisometric.
Therefore, (ρ′′, t′′, u′′,H′′) is an isometric and JX-coisometric dilation of
(ρ0, t0, u0,H0), that is, a Cuntz-Pimsner representation of ((X,C),G, α).
Corollary 5.3. Let (X,C) be the graph correspondence of a directed
graph (E, V ) and ((X,C),G, α) a C∗-correspondence dynamical system,
THE NON-SELFADJOINT APPROACH TO THE HAO-NG ISOMORPHISM 33
that is, α is a generalized gauge action. Then all relative crossed prod-
ucts for (T +
X ,G, α) are canonically isomorphic via completely isometric
maps.
Proof. This is an immediate consequence of Theorem 5.2 and Theo-
rem 2.4.
With an added assumption this gives a positive solution to the Hao-
Ng isomorphism problem.
Corollary 5.4. Let (X,C) be the graph correspondence of a row-finite
directed graph (E, V ) and ((X,C),G, α) a C∗-correspondence dynamical
system. Then
OX⋊αG ≃ OX ⋊α G.
Proof. By the description of JX mentioned at the start of this section
and Theorem 3.1, the graph correspondence of a row-finite directed
graph (E, V ) is hyperrigid. The conclusion follows from Theorem 4.9.
Acknowledgement. Part of this research was carried out in the summer
of 2018 while EK was visiting Chongqing Normal University, Chongqing
University, the Chinese Academy of Science in Beijing and the Chern
Institute of Mathematics. EK would like to thank his hosts Professors
Liming Ge, Hanfeng Li, Chi-Keung Ng, Wenhua Qian, Wenmimg Wu
and Wei Yuan for the stimulating conversations and their hospitality
during his stay at their institutions.
References
[1] B. Abadie, Takai duality for crossed products by Hilbert C∗-bimodules, J. Op-
erator Theory 64 (2010), 19 -- 34.
[2] W. Arveson, The noncommutative Choquet boundary II: hyperrigidity, Israel
J. Math. 184 (2011), 349 -- 385.
[3] E. Bedos, S. Kaliszewski, J. Quigg and D. Robertson, A new look at crossed
product correspondences and associated C∗ -- algebras, J. Math. Anal. Appl. 426
(2015), 1080 -- 1098.
[4] E. Bedos, S. Kaliszewski, J. Quigg and J. Spielberg, On finitely aligned left
cancellative small categories, Zappa-Sz´ep products and Exel-Pardo algebras,
Thy. App. Categ. 33 (2018), 1346 -- 1406.
[5] D. Blecher and C. Le Merdy, Operator algebras and their modules-an operator
space approach, London Mathematical Society Monographs New Series 30,
Oxford University Press, 2004.
[6] N. Brown and N. Ozawa, C∗ -- algebras and finite-dimensional approximations,
Graduate Studies in Mathematics 88, American Mathematical Society, Prov-
idence, RI, 2008. xvi+509 pp.
34
E.G. KATSOULIS AND C. RAMSEY
[7] V. Deaconu, Group actions on graphs and C∗-correspondences, Houston J.
Math. 44 (2018), 147 -- 168.
[8] V. Deaconu, A. Kumjian and J. Quigg, Group actions on topological graphs,
Ergodic Theory Dynam. Systems 32 (2012),1527 -- 1566.
[9] B. Duncan, Certain free products of graph operator algebras, J. Math. Anal.
Appl. 364 (2010), 534 -- 543.
[10] R. Exel, Amenability for Fell bundles, J. Reine Angew. Math. 492 (1997),
41 -- 73.
[11] N. Fowler, Discrete product systems of Hilbert bimodules, Pacific J. Math. 204
(2002), 335 -- 375.
[12] E. Gootman and A. Lazar, Crossed products of type I AF C∗-algebras by abelian
groups, Isr. J. Math 56 (1986), 267 -- 279.
[13] M. Hamana, Injective envelopes of operator systems, Publ. Res. Inst. Math.
Sci. 15 (1979), 773 -- 785.
[14] G. Hao and C-K. Ng, Crossed products of C∗-correspondences by amenable
group actions, J. Math. Anal. Appl. 345 (2008), 702 -- 707.
[15] S. Harris and S. Kim, Crossed products of operator systems, J. Funct. Anal.
276 (2018), 2156 -- 2193.
[16] E.T.A. Kakariadis, The Dirichlet property for tensor algebras, Bull. Lond.
Math. Soc. 45 (2013), 1119 -- 1130.
[17] S. Kaliszewski, J. Quigg and D. Robertson, Functoriality of Cuntz-Pimsner
correspondence maps, J. Math. Anal. Appl. 405 (2013), 1 -- 11.
[18] S. Kaliszewski, J. Quigg and D. Robertson, Coactions on Cuntz-Pimsner al-
gebras, Math. Scand. 116 (2015), 222 -- 249.
[19] E. Katsoulis, Non-selfadjoint operator algebras: dynamics, classification and
C ∗-envelopes, Recent advances in operator theory and operator algebras, 27 --
81, CRC Press, Boca Raton, FL, 2018.
[20] E. Katsoulis, C∗-envelopes and the Hao-Ng isomorphism for discrete groups,
Inter. Math. Res. Not. (2017), 5751 -- 5768.
[21] E. Katsoulis and D. Kribs, Tensor algebras of C ∗-correspondences and their
C∗-envelopes, J. Funct. Anal. 234 (2006), 226 -- 233.
[22] E. Katsoulis and C. Ramsey, Crossed products of operator algebras, Memoirs
Amer. Math. Soc 258 (2019), no. 1240, vii+85 pp.
[23] E. Katsoulis and C. Ramsey, Crossed products of operator algebras: applica-
tions of Takai duality, J. Funct. Anal. 275 (2018), 1173 -- 1207.
[24] D.W. Kim, Coactions of Hopf C∗-algebras on Cuntz-Pimsner algebras, arXiv:
1407.6106 [math.OA].
[25] D. Larson and B. Solel, Structured triangular limit algebras Proc. London
Math. Soc. 75 (1997), 177 -- 193.
[26] M. McAsey and P. Muhly, Representations of nonselfadjoint crossed products,
Proc. London Math. Soc. 47 (1983), 128 -- 144.
[27] A. Morgan, Cuntz-Pimsner
algebras
and
twisted
tensor
products,
arXiv:1601.07826.
[28] P. Muhly and B. Solel, Tensor algebras over C ∗-correspondences: representa-
tions, dilations, and C ∗-envelopes, J. Funct. Anal. 158 (1998), 389 -- 457.
[29] V. Paulsen, Completely Bounded Maps and Operator Algebras, Cambridge
Studies in Advanced Mathematics 78, Cambridge University Press, 2002.
MR1976867
THE NON-SELFADJOINT APPROACH TO THE HAO-NG ISOMORPHISM 35
[30] G. Popescu, Isometric dilations for infinite sequences on noncommuting oper-
ators, Trans. Amer. Math. Soc. 316 (1989), 523 -- 536.
[31] I. Raeburn, Graph algebras, CBMS Regional Conference Series in Mathematics
103, American Mathematical Society, Providence, RI, 2005.
[32] J. Schaeffer, On unitary dilations of contractions, Proc. Amer. Math. Soc. 127
(1955), 322.
[33] C. Schafhauser, Cuntz-Pimsner algebras, crossed products and K-theory, J.
Funct. Anal. 269 (2015), 2927 -- 2946.
[34] D. Williams, Crossed products of C∗-algebras, Mathematical Surveys and
Monographs, Vol. 134, American Mathematical Society, 2007.
Department of Mathematics, East Carolina University, Greenville,
NC 27858, USA
E-mail address: [email protected]
Department of Mathematics and Statistics, MacEwan University,
Edmonton, AB, Canada
E-mail address: [email protected]
|
1704.06845 | 1 | 1704 | 2017-04-22T20:35:49 | Positive definite functions on Coxeter groups with applications to operator spaces and noncommutative probability | [
"math.OA",
"math.CO",
"math.GR"
] | A new class of positive definite functions related to colour-length function on arbitrary Coxeter group is introduced. Extensions of positive definite functions, called the Riesz-Coxeter product, from the Riesz product on the Rademacher (Abelian Coxeter) group to arbitrary Coxeter group is obtained. Applications to harmonic analysis, operator spaces and noncommutative probability is presented. Characterization of radial and colour-radial functions on dihedral groups and infinite permutation group are shown. | math.OA | math |
Positive definite functions on Coxeter groups
with applications to operator spaces and
noncommutative probability
by Marek Bo zejko, ´Swiatosław R. Gal, and Wojciech Młotkowski
Uffe Haagerup ( -- ) in Memoriam
Abstract. A new class of positive definite functions related to
colour-length function on arbitrary Coxeter group is introduced.
Extensions of positive definite functions, called the Riesz-Coxeter
product, from the Riesz product on the Rademacher (Abelian Cox-
eter) group to arbitrary Coxeter group is obtained. Applications
to harmonic analysis, operator spaces and noncommutative prob-
ability is presented. Characterization of radial and colour-radial
functions on dihedral groups and infinite permutation group are
shown.
Introduction
In Uffe Haagerup in his seminal paper [haa], essentially
proved the positive definitness, for ≤ q ≤ , of the function Pq(x) =
qx = exp(−tx), where · is the word lenght on a free Coxeter group
W = Z/∗···∗Z/. From this he deduced also Khinchine type inequal-
ities. He has shown that the regular C∗-algebra of W has bounded
approximation property and later [dch] the completely bounded
approximation property (cbap). These results of Uffe Haagerup have
had significant impact on harmonic analysis on free groups and,
more generally, on Coxeter groups; they also influenced free proba-
bility theory and other noncommutative probability theories.
In the paper [bjs] it was shown that the function Pq(x) = qx is
positive definite for q ∈ [−, ] and all Coxeter groups, where the
length · is the natural word length function on a Coxeter group
with repect to the set of its Coxeter generators. This fact implies
that infinite Coxeter groups have the Haagerup property and do not
have Kazhdan's propery (T).
Mathematics Subject Classification. Primary f, a, l, Secondary
a, a.
Key words and phrases. Coxeter group, positive definite functions, operator spaces,
Sidon sets, Khinchine inequality, length function, de Finetti theorem.
Marek Bo zejko, ´Swiatosław R. Gal, and Wojciech Młotkowski
Later, Januszkiewicz [jan] and Fendler [fenb] showed, in the
spirit of Haagerup proof, that w 7→ zw is a coefficient of a uniformely
bounded Hilbert representation of W for all z ∈ C such that z < .
As shown in a very short paper of Valette [val], this implies cbap.
See the book [bo] for futher extension of Uffe Haagerup results for
a big class of groups.
In the paper [bs] Bo zejko and Speicher considered the free prod-
uct (convolution) of classic normal distribution N(, ) and the new
length function on the permutation group Sn (i.e. the Coxeter group
of type a) was introduced, which we shall call the colour-length func-
tion k·k. It is defined as follows: for w ∈ Sn in the minimal (reduced)
representations w = s . . . sk, where each sj belong to the set S of trans-
posions of the form (i, i + ), we put kwk = #{s, s, . . . , sk}.
For our study one of the most important results of this paper is that
the function called Riesz-Coxeter product Rq defined on all Coxeter
groups (W, S) as
Rq(s) = qs, for s ∈ S, and Rq(xy) = Rq(x)Rq(y), if kxyk = kxk +kyk
is positive definite for ≤ qs ≤ .
This implies, in particular, that in an arbitrary Coxeter group the set
of its Coxeter generators is a weak Sidon set and also it is completely
bounded Λcb
p -set, see Theorems . and .. Equivalently, the span
of the linear operators {λ(s)s ∈ S} in the noncommutative Lp-space
Lp(W) is completely boundedly isomorphic to row and column op-
erator Hilbert space (see Theorem .).
Another interesting connection between the two length functions ·
and k · k appeared in [bs] in the formula for the moments of free
additive convolution power of the Bernoulli law µ− = (δ− + δ)/
(cf. Corollary in cited paper):
(−)πq−kπk,
mn(cid:16)µ
⊞q
−(cid:17) = qnXπ∈P(n)
for q ∈ N. (See also Section of the present paper.)
Also, in [bbls] the colour-length function on the permutation group
Sn was studied. Some of its extensions to pairpartitions appeared in
the presentation of the proof that classical normal law N(, ) is free
infinitely divisible under free additive convolution ⊞.
Since we have recent extensions of the free probability (which is re-
lated to type a Coxeter groups) to the free probability of type b Cox-
eter groups (see [beh]), it seems to be interesting to determine the
role of the colour-length functions for the Coxeter groups of type b
and d.
Positive definite functions on Coxeter groups. . .
The plan of the paper is as follows.
In Section we recall definitions of Coxeter groups and of the length
and the colour-length functions.
In Section we recall the definition of positive definite funcios and
discuss various classes of those, namely radial, colour-radial, and
colour-dependant.
In Section we discuss Abelian Coxeter groups.
In Section we show the following formula characterizing the ra-
dial normalised positive definite functions on these Coxeter groups
which contain the infinite Rademacher group L∞
i= Z/ as a para-
bolic subgroup (these include the infinite permutation group S∞):
every radial positive definite function ϕ is of the form
ϕ(w) =Z
−
qw µ(dq)
for a probability measure µ.
That characterisation is a variation on the classical de Finetti theo-
rem. A noncommutative version was shown by Kostler ans Speicher
[ks] (see also [leh]).
We also show in Theorem ., that the function exp(−twp) is posi-
tive definite for all t ≥ if and only if p ∈ [, ].
In Section we give a short proof of the equivalence of the two
known results concerning positive definite functions on finite Cox-
eter groups.
In Section we present the main properties of the colour-dependent
positive definite functions on Coxeter groups, in particular we show
in Proposition .. that on S∞ and some other Coxeter groups, the
function w 7→ rkwk is positive definite if and only if r ∈ [, ].
The Section gives characterization of all colour-length functions on
finite and infinite dihedral groups Dm, for m = , , . . . ,∞.
In Section we prove that the set S of Coxeter generators is a weak
Sidon set in arbitrary Coxeter groups (W, S) with constant and that
it is also a completely bounded Λ(p) set with contants as C√p, for
p > .
In Section we prove for arbitrary finitely generated Coxeter group
an identity involving both lengths · and k · k (see Proposition .).
We apply it to give a proof of Corollary from [bs], (see Equation
(.)) where the proof, involving probabilistic considerations, was
not presented in [bs].
Marek Bo zejko, ´Swiatosław R. Gal, and Wojciech Młotkowski
. Coxeter groups
In this part we recall the basic facts regarding Coxeter groups and
introduce notation which will be used throughout the rest of the
paper. For more details we refer to [bou, hum].
A group W is called a Coxeter group if it admits the following pre-
sentation:
W =(cid:28)S(cid:12)(cid:12)(cid:12)(cid:12)n(ss)m(s,s) = : s, s ∈ S, m(s, s) , ∞o(cid:29) ,
where S ⊂ W is a set and m is a function m : S×S → {, , , . . . ,∞} such
that m(s, s) = m(s, s) for all s, s ∈ S and m(s, s) = if and only
if s = s. The pair (W, S) is called a Coxeter system. In particular,
every generator s ∈ S has order two and every element w ∈ W can be
represented as
(.)
w = ss . . . sm
for some s, s, . . . , sm ∈ S. If the sequence (s, . . . , sm) ∈ Sm is chosen
in such a way that m is minimal then we write w = m and call it
the length of w. In such a case the right hand side of (.) is called
a reduced representation or reduced word of w. This is not unique in
general, but the set of involved generators is unique [bou, Ch. iv,
§, Prop. ], i.e.
if w = ss . . . sm = tt . . . tm are two reduced rep-
resentations of w ∈ W then {s, s, . . . , sm} = {t, t, . . . , tm}. This set
{s, s, . . . , sm} ⊆ S will be denoted Sw and called the colour of w.
Given a subset T ⊂ S by WT we denote the subgroup generated by T
and call it the parabolic subgroup associated with T. To see that Sw is
independent of the reduced representation of w notice that
(.)
s ∈ Sw ⇐⇒ w < WSr{s}.
We define the colour-length of w putting kwk = #Sw (the cardinality of
Sw). Both lengths satisfy the triangle inequality and we have kwk ≤
w.
In the case of the permutation group the colour-length has the fol-
lowing pictorial interpretation. If σ is a permutation in Sn+ then kσk
equals n minus the number of connected components of the diagram
representing σ. Notice, that σ equals to the number of crossings in
the diagram (the number of pairs of chords that cross).
σ
σ
kσk
e
Positive definite functions on Coxeter groups. . .
()
()()
()()()
It would be convenient to define
(.)
then, clearly, kwk =Ps∈Skwks.
kwks : =( if s < Sw,
if s ∈ Sw,
. Positive defined functions
Xx,y∈Γ
A complex function ϕ on a group Γ is called positive definite if we
have
ϕ(y−x)α(x)α(y) ≥
for every finitely supported function α : Γ → C.
A positive definite ϕ function is Hermitian and satisfies ϕ(x) ≤ ϕ(e)
for all x ∈ Γ. Usually it is assumed, that ϕ is normalised, i.e. that
ϕ(e) = .
In this and the following sections we discuss the radial functions on
Coxeter groups. These are functions which depend on w rather then
on w.
We call a function ϕ on (W, S) colour-dependent if ϕ(w) depends only
on Sw. We call it colour-radial if it depends only on kwk.
An Abelian Coxeter group generated by S is isomorphic to the direct
product ⊕s∈SZ/. On these groups the lengths · and k · k coincide
and all functions are colour dependent.
s
The main example of a positive definite function will be the Riesz --
Coxeter function. Given a sequence q = (qs)s∈S we define Rq(w) =
Qs∈S qkwks
=Qs∈Sw
qs. We will abuse notation and denote by Rq also
the associated operatorPw∈W Rq(w)w. That is
Rq = +Xs∈S
qsqsw + Xw:Sw={s,s,s}
qss + Xw:Sw ={s,s}
qsqsqsw + . . .
Marek Bo zejko, ´Swiatosław R. Gal, and Wojciech Młotkowski
In the case all qs = q we get Rq =P qkwkw.
This generalises the classical case of Rademacher -- Walsh functions in
the Rademacher group Radn. If we denote the generator of the i-th
factor Z/ of the latter by the symbol ri then, by definition, r
=
and ri rj = rj ri and
Rq =
nYi=
( + qiri ).
. Rademacher groups
In this section we are going to study positive definite radial func-
tions on the Abelian Coxeter groups, (W, S) = RadS. Since positive
definiteness is tested on functions with finite support, we can as-
sume that S is countable. If #S = n we will write Radn instead of
RadS. Given n ∈ N ∪ {∞}, we denote by Pn the class of all func-
tions f : {, , . . . , n} → R for n finite and f : N → R if n = ∞ such
that ϕ(w) = f (w) is a normalised positive definite on Radn.
The following observation is straightforward.
Proposition .. Assume that ≤ m < n ≤ ∞ and f ∈ Pn. Then the
restriction of f to {, . . . , m} belongs to Pm. A fuction f belogs to P∞ if
and only if all its restrictions to {, . . . , m} for any m ∈ N belong to Pm.
Theorem .. Assume n is finite. The set Pn form a simplex whose ver-
l−i(cid:1), where ≤ l ≤
i(cid:1)(cid:0)n−k
tices (extreme points) are f n
n. Equivalently, every normalised radial positive definite function on the
group Radn is of the form
i=(−)i(cid:0)k
l(cid:1)−Pl
l (k) =(cid:0)n
nXl=
λl f n
ϕ(x) =
l (x),
l= λl = .
l= is unique and satisfies
where the sequence of nonnegative numbers (λl)n
Pn
Proof. We can indentify the dual [Radn group of Radn with Radn
via the paring (x, y) = (−)Pn
i= xi yi . By Bochner's theorem every nor-
ϕ(x) =Z
malised positive definite function ϕ on Radn is of the form
(x, y) µ(dy),
for some probability measure µ. Clearly, such a function is radial if
and only if µ is invariant under the action of Sn.
dRad∞
Positive definite functions on Coxeter groups. . .
Among such measures extreme ones are measures µl for ≤ l ≤ n,
where µl is equally distributed among elements of length l. More-
over,
ϕ(x) =Z
(x, y) µl(dy) = f n
l (x)
dRad∞
(cid:3)
as claimed.
The following theorem is a version of the classical de Finetti Theo-
rem (see [fel, p. ]) for the infinite Rademacher group.
Theorem .. Assume that ϕ is a radial function on the Rademacher
group Rad∞. Then ϕ is a normalised positive definite if and only if there
exists a probability measure µ on [−, ] such that
ϕ(x) =Z
−
qxµ(dq).
This measure µ is unique.
Proof. Since the function qx is normalised positive definite for q ∈
[−, ], the "if" implication is obvious.
Assume that ϕ is normlised positive definite. The group Rad∞ is dis-
crete and Abelian and its dual is the compact group [Rad∞ =Q∞i= Z/.
By Bochner's theorem, there exists a probability measure η on dRad∞
such that
(x, y) dη(y),
ϕ(x) =Z
dRad∞
where for x = (x, x, . . .) ∈ Rad∞, y = (y, y, . . .) ∈ dRad∞ we put (x, y) =
(−)P∞i= xi yi . The radiality of ϕ is equivalent to the fact that for every
permutation σ ∈ S∞ we have ϕ(x) = ϕ(σ(x)), where σ(x) = (xσ(), xσ(), . . .).
This, in turn, implies that η is σ-invariant for every σ ∈ S∞, i.e. we
have η(A) = η(σ(A)) for every Borel subset A ⊂ dRad∞.
Cn(ǫ) = {y ∈ dRad∞yi = ǫi : ≤ i ≤ n},
i= ∈ {, }n we define Cn(ǫ) ⊆ dRad∞ by
in particular C(∅) = dRad∞. Then we have η(Cn(ǫ)) = η(Cn(ǫ′)) if
ǫ′i = ǫσ(i) for some σ ∈ Sn and every ≤ i ≤ n. For ε ∈ R we put
For a sequence ǫ = (ǫi)n
εn = ε, ε, . . . , ε
{z }n
and an = η(Cn(n)). Moreover, for n, k ≥ we define the difference
operators ∆kan by induction: ∆an = an and ∆k+an = ∆kan+ − ∆kan.
We claim that
(.)
(−)k∆kan = η(cid:16)Cn+k(cid:16)nk(cid:17)(cid:17) .
Marek Bo zejko, ´Swiatosław R. Gal, and Wojciech Młotkowski
Denoting the right hand side of (.) by c(n, k) we note that c(n, ) =
an and
Cn+k+(cid:16)nk(cid:17) ∪ Cn+k+(cid:16)nk(cid:17) = Cn+k(cid:16)nk(cid:17) ,
is a disjoint union. This implies
This formula, by induction on k, leads to (.).
c(n, k + ) = c(n, k)− c(n + , k).
From (.) we see that the sequence (an) is completely monotone, i.e.
that (−)k∆kan ≥ for all n, k ≥ . By the celebrated theorem of Haus-
dorff (see [hau, Satze ii und iii]), there exists a unique probability
measure ρ on [, ] such that
(.)
(−)k∆kan =Z
un(− u)k dρ(u).
(Note that Equation (.) for arbitrary k ≥ follows from the case
k = .)
For x = (n∞) ∈ Rad∞ so that x = n, we have
ϕ(x) =Z
(x, y) dη(y) =Z
dRad∞
= Xǫ∈{,}n
(−)Pn
k!(−)kZ
n
nXk=
=Z
qn dµ(q),
=
dRad∞
i= ǫi η(Cn(ǫ)) =
i= yi dη(y)
(−)Pn
k!(−)kη(cid:16)Cn(cid:16)kn−k(cid:17)(cid:17)
n
nXk=
(− u)n dρ(u)
uk(− u)n−k dρ(u) =Z
−
where µ is defined by µ(A) = ρ(cid:16)
+
A(cid:17) for a Borel set A ⊆ [−, ]. (cid:3)
. Remarks on radial positive definite functions on some infin-
itely generated Coxeter groups
In this Section we extend the last theorem of the previous section to
a certain class of Coxeter groups.
Theorem .. Assume that (W, S) is a Coxeter system and that there is
an infinite subset S ⊆ S such that st = ts for s, t ∈ S. Assume that ϕ is
a radial function on W with ϕ(e) = . Then ϕ is positive definite if and
only if there exists a probability measure µ on [−, ] such that
ϕ(σ) =Z
−
qσµ(dq).
This measure µ is unique.
Positive definite functions on Coxeter groups. . .
Proof. It is sufficient to note that the group generated by S is a par-
abolic subgroup isomorphic with Rad∞.
(cid:3)
Example. For W = S∞ we have S = {(n, n + ) : n ∈ N}. Then we can
take S = {(n − , n) : n ∈ N}. Similar S can be found in infinitely
generated groups of type b and d.
Problem .. When − ≤ q ≤ , q , is the positive definite function
qx on S∞ an extreme point in the set of normalised positive definite
functions?
Theorem .. The function ϕp(σ) = e−tσp
and only if ≤ p ≤ .
Proof. A contrario. Assume that for some p > and t > the function
ψp(σ) = e−tσp
For q = e−t, choosing σ such that σ = n we have q
for some probability measure µ on [−, ]. Since(cid:16)R
tends to max{qq ∈ supp µ} while(cid:18)q
=R
qn dµ(q)(cid:17)/n
→ , we conclude that
is positive definite on S∞ if
is positive definite on S∞.
µ is the Dirac measure at , which is a contradiction.
The "if" part is standard. We need to show that f (x) = e−txp
is the
Laplace transform of some probability measure supported on [,∞),
so f is a convex combination of functions of the form e−sx.
By characterisation of Laplace transforms (see [hau, Satz iii]) this
is equivalent to complete monotonicity, that is (−)nf (n) > for all
n = , , . . . . And indeed, by induction, (−)nf (n) is a positive linear
combination of positive functions of the form xpj−nf (x) for ≤ j ≤
n.
(cid:3)
(n)p
(cid:19)/n
qn dµ(q)
−
(n)p
−
The measures with Laplace transforms e−txp
for t ≥ and ≤ p ≤
are studied in detail in [yos, Ch. ix.] (see Propositions and
there).
Let us note that for such groups like Zk or Rk with the Euclidean
distance d the functions exp(−tdp) are positive definite for all t ≥
and ≤ p ≤ (the case p = corresponds to the Gaussian Law).
. The longest element
If a Coxeter group W is finite, then it contains the unique element
ω◦ which has the maximal length with respect to ·.
Marek Bo zejko, ´Swiatosław R. Gal, and Wojciech Młotkowski
From the definition it is clear, that a function ϕ on a group Γ with
values in the field of complex numbers C is positive definite if and
onlyPg∈Γ ϕ(g)g is a nonnegative (bounded if the group is finite) op-
erator on ℓΓ. (We will identify g ∈ Γ with λ(g) ∈ B(ℓΓ), where λ is
the left regular representation, for short.)
Let W be a finite Coxeter group. The following two statements are
well known.
(a) The function qw is positive definite for any ≤ q ≤ .
(b) The function ∆(w) = ω◦/−w is positive definite.
The first one was proven in [bjs] (even for infinite Coxeter groups
and also for − ≤ q ≤ ) while the second -- in [bs, Proposition ].
Here we give a short direct prove of the following.
Proposition .. The above statements (a) and (b) are equivalent.
Proof. Let q = e−t (with t ≥ , as we assume q ≤ ). The case (a) is
equivalent to Φt =Pw∈W et∆(w)w = etω◦/Pw∈W qww being nonnega-
tive.
Assume (a). Recall first, that ω◦w = ω◦ − w = wω◦. Therefore
ω◦/ − ω◦w = −(ω◦/ − w), ie. ∆(ω◦w) = −∆(w) and similarly,
∆(wω◦) = −∆(w).
The equality ∆(ω◦w) = ∆(wω◦) implies that ω◦ (and thus Q = ( −
ω◦)/) commutes with ∆ (and thus Φt). Since Q = Q is nonnegative
we conclude that
t
sinh(t∆(w))
et(ω◦/−w) − et(ω◦/−wω◦)
t−ΦtQ = Xw∈W
is nonnegative. Therefore, taking the limit as t → , we obtain that
Pw∈W ∆(w)w is nonnegative. Thus (b).
Assuming (b) and using the Schur lemma, which says that the (point-
wise) product of positive definite functions is positive definite, we
get that
w = Xw∈W
w.
t
Φt = Xw∈W
et∆(w)w =Xn≥
is nonnegative. Thus (a).
tn
n!Xw∈W
∆(w)nw
(cid:3)
. Colour-dependent positive definite functions on Coxeter groups
The question which colour-dependant or colour-radial functions are
positive functions on Coxeter groups is wide open. In this section
Positive definite functions on Coxeter groups. . .
we provide some sufficient conditions. In the next section we will
examine the dihedral groups in full details.
Lemma .. Let H be a subgroup of a group Γ of index d. Then the func-
tion ϕr defined by ϕr(x) = if x ∈ H and ϕr(x) = r otherwise is positive
definite on Γ if and only if r ∈ [−/(d − ), ], with natural convention
that if d = ∞ then −/(d − ) = .
Note, that if H = {} then d = Γ.
Proof. First assume that d is finite and let us enumerate the left cosets:
{g H : g ∈ Γ} = {H, H, . . . , Hd}.
Note, that for x ∈ Hi, y ∈ Hj we have y−x ∈ H if and only if i = j.
Therefore, for r = −/(d − ) and for a finitely supported complex
function f on Γ we have
−
ds − ≤ qs ≤ ,
where ds denotes the index if the parabolic subgroup generated by S r{s}
in W: ds = [W : WSr{s}]. Then the Riesz -- Coxeter Rq is positive definite
on W.
Proof. From Lemma . the function w 7→ qkwks
is positive definite
for s ∈ S and −/(ds − ) ≤ qs ≤ . Since the pointwise product of
positive definite functions is positive definite, the statement holds.
(cid:3)
s
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xx∈Hi
,
f (y)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
ϕr(y−x)f (x)f (y) =
Xx,y∈Γ
d − X≤i<j≤d
f (x)−Xy∈Hj
which proves that ϕr is positive definite. For r ∈ [−/(d − ), ] the
function ϕr is positive definite as a convex combination of ϕr and
the constant function ϕ.
On the other hand, if we choose xi ∈ Hi for each i ≤ d and define f
as the characteristic function of the set {x, . . . , xd} then
ϕr(y−x)f (x)f (y) = d + (d − d)r,
(.)
Xx,y∈Γ
which proves that r ≥ −/(d − ) is a necessary condition for positive
definiteness of ϕr.
If d = ∞ then r = and the function ϕ is positive definite as the
characteristic function of the subgroup H. For "only if" part we
chose an arbitrarily long sequence x, . . . , xd′ of elements from dif-
ferent left cosets and use (.) with d′ instead of d.
(cid:3)
Theorem .. Assume that for every s ∈ S we are given a number qs,
Marek Bo zejko, ´Swiatosław R. Gal, and Wojciech Młotkowski
Example. Take W = Sn, the permutation group on the set {, , . . . , n}.
It is generated by the transpositions S = {si = (i, i + ), ≤ i ≤ n −
}. For ≤ i ≤ n − the parabolic subgroup generated by S r {si} is
i(cid:1).
isomorphic with Si− × Sn−i−, so its index is i(cid:0)n
It would be interesting to determine for which r the function w 7→
rkwk is positive definite. By Proposition . this holds for r ∈ [−/(d −
), ], where d is the maximal index of the parabolic subgroups of
the form WSr{s}. We note a necessary condition.
Proposition .. Assume that we have distinct generators s, s, . . . , sn ∈
S such that ssk , sks (i.e. m(s, sk) > ) for ≤ k ≤ n. If the function
w 7→ rkwk is positive definite on W, then − /(n − ) ≤ r ≤ .
If there is an element s ∈ S for which there are infinitely many s ∈ S such
that ss , ss then rkwk is positive definite on W if and only if ≤ r ≤ .
Proof. Consider elements wk = ssks. Note, that for k , l we have
kw−
l wkk = . If ϕr is positive definite on W then we have
ϕr(x−
l xk) = n + (n − n)r,
≤
nXk,l=
which implies r ≥ − /(n − ).
Corollary .. The function w 7→ qkwk on S∞ is positive definite if and
only if ≤ q ≤ .
Problem .. Thus, it is valid to ask the following. Is it true that ev-
ery normalised positive definite colour-lenght-radial function φ : S∞ →
R is of the form φ(σ) =R
on [, ]?
qkσk dµ(q) for some probability measure µ
(cid:3)
. Dihedral groups
In this part we are going to examine the class of colour-dependent
positive definite functions on the case the simplest nontrivial Cox-
eter groups. Assume that W = Dn = hs, t(st)ni (i.e. the group of sym-
metries of a regular n-gon), and define a colour-dependent function
on W:
(.)
φ(w) =
if w = e,
p if w = s,
if w = t,
q
otherwise.
r
If p = q then φ is colour radial. We are going to determine for which
parameters p, q, r the function φ is positive definite on W. It is easy
Positive definite functions on Coxeter groups. . .
to observe necessary conditions: p, q, r ∈ [−, ]. Moreover, since hsti
is a cyclic subgroup of order n, Lemma ., implies a necessary con-
dition: −/(n − ) ≤ r ≤ .
Finite dihedral groups. Assume that W is a finite dihedral group,
W = Dn, so that (st)n = . We will use the following version of
Bochner's theorem: A function f on a compact group G is positive
definite if and only if its Fourier transform:
is a positive operator for every π ∈bG, wherebG denotes the dual ob-
ject of G, i.e. the family of all equivalency classes of unitary irre-
ducible representations of G, see [sim]. Then we have
bf (π) =ZG
f (x)π(x−)dx
f (x) =Xπ∈bG
dπ trhbf (π)π(x)i .
nXg∈G
bφ(π) =
φ(g)π(g−).
Therefore, for every irreducible representation π of Dn we are going
to find
We will identify s with (,−) and t with (,−). If n is odd then Dn
possesses two characters: χ+,+ such that χ+,+(w) = for every w ∈ Dn
and χ−,− such that χ−,−(s) = χ−,−(t) = −. If n is even then we have two
additional characters χ+,− and χ−,+ such that χ+,−(s) = χ−,+(t) = and
χ+,−(t) = χ−,+(s) = −. It is easy to check that
which gives
and, for n even,
which implies
nbφ(χ+,+) = + p + q + (n − )r,
nbφ(χ−,−) = − p − q + r,
−− (n − )r ≤ p + q ≤ + r
nbφ(χ+,−) = + p − q − r,
nbφ(χ−,+) = − p + q − r,
p − q ≤ − r.
Ua(k, ) = eπika/n
Ua(k,−) =
e−πika/n
e−πika/n! ,
! ,
eπika/n
We have also the family of two dimensional representations Ua:
Marek Bo zejko, ´Swiatosław R. Gal, and Wojciech Młotkowski
k. Then for the function given by (.) we have
where a = , , . . . ,j n−
=
p − r + (q − r)e−πia/n
nbφ(Ua) = (− r)Id + (p − r)Ua(,−) + (q − r)Ua(,−)
! .
p − r + (q − r)eπia/n
− r
− r
This matrix is positive definite if and only if r ≤ and
Therefore we have
(cid:12)(cid:12)(cid:12)p − r + (q − r)eπia/n(cid:12)(cid:12)(cid:12) ≤ − r.
Proposition .. The function φ given by (.) is positive definite on
Dn if and only if
+ p + q + (n − )r ≥ ,
− p − q + r ≥
(plus
whenever n is even) and
+ p − q − r ≥ ,
− p + q − r ≥
(cid:12)(cid:12)(cid:12)p − r + (q − r)eπia/n(cid:12)(cid:12)(cid:12) ≤ − r.
for a = , , . . . ,j n−
k.
Let us confine ourselves to colour-radial functions.
Corollary .. Assuming that p = q, the function φ defined by (.) is
positive definite on W = Dn if and only if
i.e. if and only if the point (p, r) belongs to the triangle whose vertices are
max(−p −
n −
, p − ) ≤ r ≤
+ (n − ) cos(π/n)! ,
− cos(π/n)
− n − cos(π/n)
+ (n − ) cos(π/n)
,
+ p cos(π/n)
+ cos(π/n)
,
(cid:18) n −
n −
, −
n − (cid:19) ,
(, ).
Proof. For p = q the conditions from Proposition . reduce to
p − ≤ r, −− p ≤ (n − )r,
and
cos(π/n)p − r ≤ − r.
It is sufficient to note that p− ≤ r implies cos(π/n)(p− r) ≤ − r for
p ≤ .
Example. For D we have the positive definiteness of φ is equivalent
to
(cid:3)
− +p + q ≤ r ≤ −p − q,
which means that the set of all possible (p, q, r) forms a tetrahedron
with vertices (−, ,−), (,−,−), (−,−, ), (, , ). For p = q the
condition reduces to p− ≤ r ≤ .
In the case of D Proposition . leads to the following conditions:
Positive definite functions on Coxeter groups. . .
− p − q + r ≥ ,
+ p + q + r ≥ ,
− r ≥pp + q + r − pq − pr − qr,
which can be expressed as
max(−− p − q
, p + q − ) ≤ r ≤
− p − q + pq
− p − q
.
The infinite dihedral group. Here we are going to study W = D∞.
Proposition .. The function φ given by (.) is positive definite on
W = D∞ if and only if ≤ r and p − r +q − r ≤ − r, i.e.
+ p + q
(.)
max{, p + q − } ≤ r ≤ min(−p − q,
) .
Proof. First we note that the set of (p, q, r) ∈ R satisfying (.) con-
stitutes a pyramid which is the convex hull of the points (±, , ),
(,±, ) and (, , ) (apex). For these particular parameters it is
easy to see that φ is positive definite: (, , ) corresponds to the
constant function , (, , ) to the characteristic function of the sub-
group hsi = {, s}, and (−, , ) to the character χ−,− times the char-
acteristic function of hsi. Similarly for (,±, ). This, by convexity,
proves that (.) is a sufficient condition.
On the other hand, we know already that r ≥ is a necessary con-
dition. Let us fix n and define W+(n) = {x ∈ W : sx < x ≤ n},
W−(n) = {x ∈ W : tx < x ≤ n} and
f (x) =(± if x ∈ W±(n),
otherwise.
For x, y ∈ W+(n) we have Sy−x = ∅ in n cases (namely, if x = y)
Sy−x = {s} in n − cases (namely if x = k, y = k + or vice-
versa, k = , . . . , n − ) Sy−x = {t} in n cases (namely if x = k, y =
k − or vice-versa, k = , . . . , n) and Sy−x = {s, t} in all the other
(n − )(n − ) cases. Similarly, for x, y ∈ W−(n) we have Sy−x = ∅
in n cases, Sy−x = {s} in n cases, Sy−x = {t} in n − cases and
Sy−x = {s, t} in (n − )(n − ) cases. If x ∈ W+(n), y ∈ W−(n) or vice-
versa then Sy−x = {s, t}. Summing up, we get
Xx,y∈W
φ(y−x)fn(x)fn(y)
= n + (n − )p + (n − )q + (n − )(n − )r − nr
= n + (n − )p + (n − )q − (n − )r.
Marek Bo zejko, ´Swiatosław R. Gal, and Wojciech Młotkowski
+(cid:18)−
Therefore for every n ∈ N we have a necessary condition
n(cid:19) r ≥ .
n(cid:19) p +(cid:18)−
Letting n → ∞ we get + p + q ≥ r.
Put xk = stst . . ., xk = k. Fix n and define
n(cid:19) q −(cid:18)−
g(x) =(χ−,+(x)
if x = xk for ≤ k ≤ n,
otherwise,
where, as before, χ−,+ is the character on W for which χ−,+(s) = −,
χ−,+(t) = . Then
φ(y−x)g(x)g(y) =
Xx,y∈W
nXk,l=
φ(x−
l xk)g(xk)g(xl).
Denote ck,l = φ(x−
l xk)g(xk)g(xl). Then we have ck,k = , ≤ k ≤ n,
ck,k− = q if k is even, ck,k− = −p if k is odd, ≤ k ≤ n and ck,l = cl,k
for all ≤ k, l ≤ n. If ≤ k, l ≤ n and k − l ≥ then ck,l = (−)j r,
where j is the total number of s appearing in xk and xl. Now it is not
difficult to check that
nXl=
ck,l =( + q − r
− p + q − r
if k = or k = n,
if < k < n,
which implies
Xx,y∈W
φ(y−x)g(x)g(y) = n − (n − )p + nq − (n + )r
and leads to necessary condition r ≤ − p +q. In a similar manner we
get r ≤ + p − q.
Finally, define a function h similarly like g, but now we use the char-
acter χ−,−:
h(x) =(χ−,−(x) = (−)k
if x = xk for ≤ k ≤ n,
otherwise.
Putting dk,l = φ(x−
l xk)h(xk)h(xl) we have dk,k = , dk,k− = −p if ≤
k ≤ n is even and dk,k− = −q if k is odd. Moreover, if k − l ≥ ,
≤ k, l ≤ n then dk,l = (−)k+l r. Now one can check that
if k = or k = n,
if < k < n,
dk,l =(− q
nXl=
− p − q + r
which yields − p − q + r ≥ and completes the proof that the condi-
tions (.) are necessary.
(cid:3)
. Weak Sidon sets and operator Khinchin inequality
Positive definite functions on Coxeter groups. . .
The aim of this section is to show that the set of Coxeter generators
S in an arbitrary Coxeter group W is a weak Sidon set, ie. an interpo-
lation set for the Fourier -- Stieltjes algebra B(W).
Given a group Γ, the Fourier -- Stieltjes algebra consists of linear com-
binations of positive definite functions on Γ, ie. every element of
B(Γ) is of the form f = ϕ − ϕ + i(ϕ − ϕ) for some positive definite
functions ϕi ( ≤ i ≤ ) on Γ. The norm on B(Γ) is defined as
kf kB(Γ) = infnXϕi(e)where f decomposes as aboveo
Theorem .. The set of Coxeter generators S in an arbitrary Coxeter
for every bounded function f : S →
group W is a weak Sidon set, ie.
[−, ] there exists positive definite functions ϕ±, such that f (s) = ϕ+(s)−
ϕ−(s) for any s ∈ S. One can take ϕ± = Rq± for a suitable choice of q±.
Moreover
kϕ+ − ϕ−kB(W) ≤
Proof. Put S±(f ) = {s ∈ S± f (s) > }. Set
q±s =(±f (s)
for s ∈ S±(f ),
otherwise.
Then f (s) = Rq+(s) − Rq−(s) as claimed. The rest of the statement
hold as the Riesz-Coxeter function at the identity element equals to
one.
(cid:3)
Given a matrix A ∈ Mn(C) and p ≥ the Schatten p-class norm kAkSp
is defined as kAkp
Sp
finite sum f =P cgλ(g) ∈ C[Γ] we define noncommutative Lp-norm
Let λ denote the left regular representation of a group Γ. Given a
= (trA)/p, where A = (A∗A)/.
Lp(Γ) =(cid:16)τ(cid:16)(f ∗ ∗ f )/(cid:17)(cid:17)/p
kf kp
where τ(f ) = ce is the von Neumann trace and Lp(Γ) is a completion
of C[Γ] with respect to the above norm.
We recall, that a scalar-valued map ϕ on a group Γ is called a com-
pletely bounded Fourier multiplier on Lp(Γ) if the associated operator
Mϕ(λ(g)) = ϕ(g)λ(g),
g ∈ Γ
extends to a completely bounded operator on Lp(Γ).
We let Mcb(Lp(Γ)) to be an algebra of completely bounded Fourier
multipliers equipped with the norm
kϕkMcb(Lp(Γ)) = kMϕ ⊗ idS p k.
Marek Bo zejko, ´Swiatosław R. Gal, and Wojciech Młotkowski
Following Pisier [pis], for as ∈ Mn(C), where s ∈ S, we define
For a set E ∈ Γ we define the completely bounded constant Λcb
infimum of C such that
p (E) as
,(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
(a∗sas)/(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)S p
Xs∈S
.
≤ Ck(as)s∈SkR∩ C
k(as)s∈SkR∩ C = max
Xs∈S
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
k(as)s∈SkR∩ C ≤(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xs∈S
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
(asa∗s)/(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)S p
Xs∈S
as ⊗ λ(s)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Lp(W)
as ⊗ λ(s)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Lp(W)
for all matrices as ∈ Mn(C) and n ∈ N.
Theorem .. If as ∈ Mn(C), then for all p ≥ and any Coxeter system
(W, S) we have
≤ A′√pk(as)s∈SkR∩ C .
Proof. It was shown by Harcharras [lp, Prop. .] that Λcb
p (E) if
finite if and only if E is an interpolation set for Mcb(Lp(Γ)), i.e. every
bounded function on E can be extended to a multiplier, and
Λcb
p (E) ≤ Λcb
p (R)µcb
p (E),
where R is the generating set in the Rademacher group Rad∞ and
µcb
p (E) is the interpolation constant.
As shown by Buchholz [buc, Thm. ] for p = n, and S te stan-
dart generating set in Rad∞, Λcb
absolute A. This was extended by Pisier [pis, Thm. ..] for any
p ≥ , i.e
n(R) = ((n − )!!)/n ≤ A√p for some
p (R) ≤ A′√p,
Λcb
for an absolute constant A′.
We have shown in Theorem . that in an arbitrary Coxeter group W
its Coxeter generating set S is a weak Sidon set, i.e. it is interpolation
set for the Fourier -- Stieltjes algebra B(W). Since for p ≥ , B(Γ) is a
subalgebra of Mcb(Lp(Γ)) and
kϕkMcb(Lp(Γ)) ≤ kϕkB(Γ),
p (S) ≤ . Thus Λcb
we see that µcb
of the right inequality.
The left inequality holds for any group Γ and any S ⊂ Γ (see [lp]).
p (S) ≤ A′√p. This finishes the proof
(cid:3)
Positive definite functions on Coxeter groups. . .
Remark .. Fendler [fena] has shown that if for all s, t ∈ S, s , t,
we have ms,t ≥ , then
Λcb
p (S) ≤ √.
See also [bo z] and [buc] for related results in the case of free
Coxeter groups. Also Haagerup and Pisier have shown that Λcb
∞ (S) =
, where Λcb
p (E) [hp]. See the paper of Haagerup
[haa] where the best constant was calculated for the set of Cox-
eter generators of the Rademacher group in case when as are scalars.
∞ (E) = supp≥ Λcb
. Chromatic length function for Coxeter groups and pairparti-
tions
Let [n] = {, . . . , n}. Let [n] denote the set of subsets of [n]. By
a partition of [n] we mean π ⊂ s[n] such that S π = [n] and if
π′, π′′ ∈ π then π′ = π′′ or π′ ∩ π′′ = ∅. We say, that partition is a
coarsening of a partition π if for any π′ ∈ π there exists ′ ∈ such
that π′ ⊂ ′.
A partition is called crossing if there exist ≤ a < b < c < d ≤ n and
π, π ∈ π with a, c ∈ π , π ∋ b, d; otherwise it is called noncrossing.
For any partition π there exists th the smallest noncrossing coars-
ening Φ(π) of π (ie. if is a noncrosing coarsening of π then it is a
coarsening of Φ(π)). We define kπk = n − #Φ(π). The notion for the
map Φ was introduced in [by].
We say that π is a pairpartition if every member of π has cardinality
two. The set of pairpartitions of [n] is denoted by P(n). Given
π ∈ P(n) we write π to denote the number of ordered quadruples
≤ a < b < c < d ≤ n such that both {a, c} and {b, d} belong to π. Note,
that π = precisely when π is noncrossing. The set of noncrossing
pairpartitions is denoted N C(n).
Given a noncrossing pairpartition we call {b, c} ∈ an inner block if
there exists {a, d} ∈ with a < b < c < d. The number of inner blocks
of we denote as inn().
In [bs, Cor. ] Bo zejko and Speicher observed the following iden-
tity.
(.)
Let fn(q) =P∈N C(n)(− q)inn(). It is elementary to derive
(− q)inn().
Xπ∈P(n)
(−)πqkπk = X∈N C(n)
n + (cid:12)(cid:12)(cid:12)(cid:12) q(cid:19) ,
fn(q) = Cn F(cid:18) n, − n
Marek Bo zejko, ´Swiatosław R. Gal, and Wojciech Młotkowski
j= tn
n(cid:1)−(cid:0) n
j qj, then the triangle (tn
n−(cid:1) = #N C(n) denote the n-th Catalan number
where Cn =(cid:0)n
and F is the classical hypergeometric funcion. If we write f (q) =
Pn−
j )≤j<n appears in [slo] as "sequence"
a). Since we are not going to use this formula, we leave
it as an exercise to the reader. For the expansion of fn( − t) and
the Delanoy triangle appearing there the reader may consult [bw,
Prop .].
In what follows, we prove a result about an arbitrary finitely gen-
erated Coxeter group, which for permutation groups implies the
above one. Given a permutation σ ∈ Sn, we construct a pairparti-
tion σ = {{i, n + − σ(i)} ≤ i ≤ n}. Note, that σ is equal to the
length of σ with respect to the Coxeter generators (, ), . . . , (n− , n)
of Sn. Therefore, denoting by w the Coxeter length of an element w
of some Coxeter group W will not lead into any confusion.
It is also clear that, with respect to the identification of permutations
with a subset of pairpartitions, the two definitions of k · k agree (see
Equations (.) and (.)).
By W(t) we denote a growth series of a finitely generated Coxeter
(Note, that the coefficient at t equals to #S. This explains why here
and in the rest of this section we consider only finitely generated
Coxeter groups. We will not repeat this assumption for short.) More-
group W (length function). That is, a power series W(t) =Pw∈W tw.
over, for X ⊂ W we write X(t) =Pw∈X tw.
Let us define a multivariable formal power series (chromatic length
function. For any X ⊂ W define
X(t, q) =Xw∈X
twYs∈Sw
qs.
In particular X(t) = X(t, ), where = ()s∈S.
Proposition .. The polynomial (or formal power series, if W is infi-
nite) W(t, q) satisfies
W(t, q) =XT⊂S
WT(t)Yr∈T
rs Ys∈SrT
(− qs).
Proof. Let W◦R denote the set of all elements of WR not contained
in any proper parabolic subgroup of WR, ie. W◦R : = WR −ST R WT.
Then, by inclusion-exclusion principle, W◦R(t) =PT⊂R(−)#(R−T)WT(t).
Positive definite functions on Coxeter groups. . .
Therefore,
qr
twYs∈Sw
W(t, q) = Xw∈W
W◦R(t)Yr∈R
qs =XR⊂S
WT(t)(−)#(R−T)Yr∈R
=XR⊂SXT⊂R
WT(t)Yr∈T
=XT⊂S
qr XT⊂R⊂S Ys∈RrT
qi Ys∈SrT
WT(t)Yr∈T
=XT⊂S
(− qs).
qr
(−qs)
(cid:3)
Corollary .. If W is a finite Coxeter group then
(.)
W(−, q) =Ys∈S
(− qs).
Proof. Choose s ∈ T and put W{s}T = {w ∈ W : w < ws}. Clearly,
WT = W{s}T WT therefore WT(t) = W{s}T (t)W{s}(t). Since W is a finite
group, W{s}T is a polynomial. Thus WT(−) = if T is nonempty (and
W∅(−) = ).
In order to prove Equation (.) we define the Wick map P(n) ∋
π 7→ :π: ∈ N C(n) (related to the normal order in quantum field
theory). Given a pairpartition π we define :π: by repetitive resolving
crossings. That is, we replace repetitively every crossing pair {a, c}
and {b, d} with a < b < c < d by {a, d} and {b, c}. In order to see that the
result is independent of the order of resolution we describe :π: in an
equivalent way.
(cid:3)
Let Φ(π) be the smallest noncrossing coarsening of π. For each
block β of Φ(π) define β+ = {y(∃x) x ∈ β, y > x, {x, y} ∈ π} and
β− = {x(∃y) y ∈ β, y > x, {x, y} ∈ π}. Order β+ = {y, . . . , yk} in in-
creasing way and β− = {x, . . . , xk} in decreasing way. Then all pairs
{xi, yi} will be parts of :π:.
Equation (.) will follow from a more refined statement.
Proposition .. For every ∈ N C(n)
(.)
(−)πqkπk = (− q)inn().
Xπ∈P(n)
:π:=
Proof. Let us first consider the case of = = {(i, n + − i) ≤ i ≤ n}.
Clearly, {π:π: = } = {σσ ∈ Sn}. And Equation (.) is equivalent to
Equation (.) (as all qs are set to q) for W = Sn.
Marek Bo zejko, ´Swiatosław R. Gal, and Wojciech Młotkowski
π
:π:
Φ(π)
Figure. Examples of π, :π:, and Φ(π).
In a general case observe, that Φ(π) is a coarsening of = :π:. Yet, not
every coarsening may appear. The obvious condition is that for each
block β of :π: the pair {min β, max β} belong to . For the purpose
of this proof we will call such a coarsening admissible.
Its clear,
that abmissible coarsenings ρ are in one to one correspondence with
subsets of containing all outer (not inner) parts of of the form
{{min ρ′, max ρ′}ρ′ ∈ ρ}.
Let us refine Equation (.) further. For every ∈ N C(n) and any
admissible coarsening η of we have
Xπ∈P(n)
:π:=, Φ(π)=ρ
(.)
(−)π = (−)#ρ.
Equation (.) follows from (.) by multiplying by qn−#η and sum-
ming over all admissible coarsenings η of .
Equation (.) is again equivalent to to Equation (.) (for all aper-
mutation groups and all qs set to q) as both sides factor as a product
over blocks of ρ.
(cid:3)
Question .. We have proven Equation . with the help of an em-
bedding Sn ∋ σ 7→ σ ∈ N C(n) (or several such embeddings, one for
each outer block of ). Corollary . holds for any Coxeter group.
Is there a corresponding formula concerning some generalization of
pairpartitions?
In the proof of Proposition . we have not assumed that W was
finite. Let us finish this section with a discussion of infinite Coxeter
groups. Recall, that − does not lie in the radius of convergence
on W(t) if W is not finite. Nevertheless, W(t) represents a rational
function as follows from the following result.
Positive definite functions on Coxeter groups. . .
Proposition .. ([ste],[ser, Prop. ]) Let (W, S) be an an infi-
nite Coxeter system. Then
(.)
W(t)
=XT∈F
(−)#T
WT(/t)
.
Where F denote the family of subsets T ⊂ S, such that the group WT gen-
erated by T is finite. In particular, W(t) is a series of a rational function
(i.e. a quotient of polynomials).
One may ask a question what is the class of (infinite) Coxeter groups
such that WT(−) = for any nonempty subset T of generators. A naıve
argument that
W(t) = W{s}(t)W{s}(t) = ( + t)W{s}(t)
shows, that the question if W(−) , is equivalent to whether W{s}(t)
can have a pole at t = −. On the other hand note, that if W is of type
a, ie. W is given by a presentation hsi : ≤ i ≤ s
i , (si sj) : ≤ i <
j ≤ i then, by Equation (.), W(t) = +t+t
More generally, it is known ([bou]) that in each coset of WT there
exists the unique shortest element. Let WT denote the set of those
shortest representatives. Moreover if w = wTwT with wT ∈ WT and
wT ∈ WT then w = wT + wT. Therefore WT(t)WT(t) = W(t). In par-
ticular, WT(t) represents a rational function, and it is legitimate to
ask about the value of WT(−).
In the case of finite Coxeter group W, Eng [eng] observed that
(−t) and W(−) = /.
WT(−) = #nw ∈ WTwww ∈ WTo ,
where w is the longest element in W. (Eng's proof was case-by-case.
Later, a general classification-free proof of Eng's theorem was given
in [rsw]).
Subsequently, Reiner [rei] has shown that if W is crystallographic
(ie. the Weyl group in a compact Lie group G), then both sides of
the above equality compute the signature of the corresponding flag
variety G/ QT, where QT is a parabolic subgroup associated to T.
What is the meaning of W(−) or WT(−) for infinite W?
We do not know if it possible for W(−) to be negative. If one takes
W = hsi : ≤ i ≤ s
i , (si sj) : ≤ i < j ≤ i. Then, by Equation (.),
W(t) = (+t)(+t+t)
t−t−t+ and Wsi (−) = − /.
Marek Bo zejko, ´Swiatosław R. Gal, and Wojciech Młotkowski
References
[bbls] Serban T. Belinschi, Marek Bo zejko, Franz Lehner, and
Roland Speicher. The normal distribution is ⊞-infinitely divisible.
Adv. Math., (): -- , .
[beh] Marek Bo zejko, Wiktor Ejsmont, and Takahiro Hasebe. Fock
J. Funct. Anal.,
type B.
space associated to Coxeter groups of
(): -- , .
[bjs] M. Bo zejko, T. Januszkiewicz, and R. J. Spatzier. Infinite Coxeter
groups do not have Kazhdan's property. J. Operator Theory, (): -- ,
.
Nathanial P. Brown and Narutaka Ozawa. C∗-algebras and finite-
dimensional approximations, volume of Graduate Studies in Mathemat-
ics. American Mathematical Society, Providence, RI, .
[bo]
[bou] N. Bourbaki. ´El´ements de math´ematique. Fasc. XXXIV. Groupes et
alg`ebres de Lie. Chapitre IV: Groupes de Coxeter et syst`emes de Tits.
Chapitre V: Groupes engendr´es par des r´eflexions. Chapitre VI: syst`emes
de racines. Actualit´es Scientifiques et Industrielles, No. . Hermann,
Paris, .
[bo z] Marek Bo zejko. On Λ(p) sets with minimal constant in discrete non-
commutative groups. Proc. Amer. Math. Soc., : -- , .
[bs] Marek Bo zejko and Roland Speicher. Interpolations between
bosonic and fermionic relations given by generalized Brownian mo-
tions. Math. Z., (): -- , .
[bs] Marek Bo zejko and Ryszard Szwarc. Algebraic length and
Poincar´e series on reflection groups with applications to representa-
tions theory. In Asymptotic combinatorics with applications to mathemati-
cal physics (St. Petersburg, ), volume of Lecture Notes in Math.,
pages -- . Springer, Berlin, .
[buc] Artur Buchholz. Norm of convolution by operator-valued functions
on free groups. Proc. Amer. Math. Soc., (): -- , .
[buc] Artur Buchholz. Optimal constants in Khintchine type inequalities
for fermions, Rademachers and q-Gaussian operators. Bull. Pol. Acad.
Sci. Math., (): -- , .
[bw] Marek Bo zejko and Janusz Wysocza ´nski. Remarks on t-
transformations of measures and convolutions. Ann. Inst. H. Poincar´e
Probab. Statist., (): -- , .
[by] Marek Bo zejko and Hiroaki Yoshida. Generalized q-deformed
Gaussian random variables. In Quantum probability, volume of Ba-
nach Center Publ., pages -- . Polish Acad. Sci. Inst. Math., Warsaw,
.
Jean De Canni `ere and Uffe Haagerup. Multipliers of the Fourier
algebras of some simple Lie groups and their discrete subgroups. Amer.
J. Math., (): -- , .
[dch]
[eng] Oliver D. Eng. Quotients of Poincar´e polynomials evaluated at −. J.
[fel] William Feller. An introduction to probability theory and its applica-
Algebraic Combin., (): -- , .
tions. Vol. II. Second edition. John Wiley & Sons Inc., New York, .
[fena] Gero Fendler. A note on L-sets. Colloq. Math., (): -- , .
[fenb] Gero Fendler. Weak amenability of Coxeter groups. arXiv:math/
, .
[haa] Uffe Haagerup. The best constants in the Khintchine inequality. Stu-
dia Math., (): -- (), .
Positive definite functions on Coxeter groups. . .
[haa] Uffe Haagerup. An example of a nonnuclear C∗-algebra, which
has the metric approximation property. Invent. Math., (): -- ,
/.
[hau] Felix Hausdorff. Summationsmethoden und Momentfolgen. II.
[hp]
Math. Z., (-): -- , .
Uffe Haagerup and Gilles Pisier. Bounded linear operators be-
tween C∗-algebras. Duke Math. J., (): -- , .
[hum] James E. Humphreys. Reflection groups and Coxeter groups, volume
of Cambridge Studies in Advanced Mathematics. Cambridge University
Press, Cambridge, .
[jan] Tadeusz Januszkiewicz. For Coxeter groups zg is a coefficient of a
[ks]
uniformly bounded representation. Fund. Math., (): -- , .
Claus K ostler and Roland Speicher. A noncommutative de Finetti
theorem: invariance under quantum permutations is equivalent to free-
ness with amalgamation. Comm. Math. Phys., (): -- , .
[lp]
[leh] Franz Lehner. Cumulants in noncommutative probability theory.
I. Noncommutative exchangeability systems. Math. Z., (): -- ,
.
Franc¸ oise Lust-Piquard. In´egalit´es de Khintchine dans Cp ( < p <
∞). C. R. Acad. Sci. Paris S´er. I Math., (): -- , .
Gilles Pisier. Introduction to operator space theory, volume of
London Mathematical Society Lecture Note Series. Cambridge University
Press, Cambridge, .
[pis]
[rei] Victor Reiner. Note on a theorem of Eng. Ann. Comb., (): -- ,
.
[rsw] V. Reiner, D. Stanton, and D. White. The cyclic sieving phenome-
[ser]
[sim]
non. J. Combin. Theory Ser. A, (): -- , .
Jean-Pierre Serre. Cohomologie des groupes discrets. In Prospects in
mathematics (Proc. Sympos., Princeton Univ., Princeton, N.J., ), pages
-- . Ann. of Math. Studies, No. . Princeton Univ. Press, Princeton,
N.J., .
Barry Simon. Representations of finite and compact groups, volume of
Graduate Studies in Mathematics. American Mathematical Society, Prov-
idence, RI, .
[slo] N. J. A. Sloane. The On-Line Encyclopedia of Integer Sequences.
A, Jul .
[ste] Robert Steinberg. Endomorphisms of linear algebraic groups. Memoirs
of the American Mathematical Society, No. . American Mathematical
Society, Providence, R.I., .
[val] Alain Valette. Weak amenability of right-angled Coxeter groups.
Proc. Amer. Math. Soc., (): -- , .
[yos] K osaku Yosida. Functional analysis, volume of Grundlehren der
Mathematischen Wissenschaften [Fundamental Principles of Mathematical
Sciences]. Springer-Verlag, Berlin-New York, sixth edition, .
Acknowledgemens
Marek Bo zejko was supported by ncn maestro grant dec-//
a/st/. Marek Bo zejko and Wojciech Młotkowski were sup-
ported by ncn grant //b/st/.
Marek Bo zejko, ´Swiatosław R. Gal, and Wojciech Młotkowski
The authors would like to thank Ryszard Szwarc and Janusz Wysocza-
ski for many discussions and help with the preparation of this paper.
The first author would like to thank professor F. Gotze for his kind
invitation to sfb and his hospitality in Bielefeld in .
M. Bo zejko -- Polska Akademia Nauk, ul. Kopernika , - Wrocław
E-mail address: [email protected]
´S. R. Gal -- Uniwersytet Wrocławski, pl. Grunwaldzki /, - Wrocław
E-mail address: [email protected]
W. Młotkowski -- Uniwersytet Wrocławski, pl. Grunwaldzki /, - Wrocław
E-mail address: [email protected]
|
1704.04403 | 1 | 1704 | 2017-04-14T12:00:36 | Morita embeddings for dual operator algebras and dual operator spaces | [
"math.OA"
] | We define a relation < for dual operator algebras. We say that B < A if there exists a projection p in A such that B and pAp are Morita equivalent in our sense. We show that < is transitive, and we investigate the following question: If A < B and B < A, then is it true that A and B are stably isomorphic? We propose an analogous relation < for dual operator spaces, and we present some properties of < in this case. | math.OA | math |
MORITA EMBEDDINGS FOR DUAL OPERATOR
ALGEBRAS AND DUAL OPERATOR SPACES
G. K. ELEFTHERAKIS
Abstract. We define a relation ⊂∆ for dual operator algebras. We say
that B ⊂∆ A if there exists a projection p ∈ A such that B and pAp are
Morita equivalent in our sense. We show that ⊂∆ is transitive, and we
investigate the following question: If A ⊂∆ B and B ⊂∆ A, then is it true
that A and B are stably isomorphic? We propose an analogous relation
⊂∆ for dual operator spaces, and we present some properties of ⊂∆ in this
case.
1. Introduction
An operator space X is said to be a dual operator space if X is completely
isometrically isomorphic to the operator space dual Y ∗ of an operator space
Y. If, in addition, X is an operator algebra, then we call it a dual operator
algebra. For example, Von Neumann algebras and nest algebras are dual
operator algebras. Blecher, Muhly and Paulsen introduced the notion of the
Morita equivalence of non-self-adjoint operator algebras [4]. Subsequently,
Blecher and Kashyap developed a parallel theory for dual operator algebras
[1], [14]. At the same time, the author of the present article proposed a
different notion of Morita equivalence for dual operator algebras, called ∆-
equivalence. Two unital dual operator algebras A and B are ∆-equivalent if
there exist faithful normal representations α : A → α(A),
β : B → β(B)
and a ternary ring of operators M (i.e., a space satisfying M M ∗M ⊆ M) such
that α(A) = [M ∗β(B)M]−w∗ and β(B) = [M α(A)M ∗]−w∗ [9]. In this case, we
write A ∼∆ B. An important property is that two algebras are ∆-equivalent
if and only if they are stably isomorphic, as was proved by Paulsen and
the present author in [12]. Subsequently, Paulsen, Todorov and the present
author defined a Morita-type equivalence ∼∆ for dual operator spaces [13].
This equivalence also has the property of being equivalent with the notion of
a stable isomorphism.
Key words and phrases. Dual operator algebras, Dual operator spaces, TRO, Stable
isomorphism, Morita equivalence.
2010 Mathematics Subject Classification. 47L05 (primary), 47L25, 47L35, 46L10, 16D90
(secondary).
1
2
G. K. ELEFTHERAKIS
In this paper, we define a weaker relation between dual operator algebras.
We say that the dual operator algebra B ∆-embeds into the dual operator
algebra A if there exists a projection p ∈ A such that B ∼∆ pAp. In this
case, we write B ⊂∆ A. We investigate the relation ⊂∆ between unital dual
operator algebras, and we prove that it is a transitive relation. In the case of
von Neumann algebras, it is a partial order relation. This means that it has
the additional property that if A, B are von Neumann algebras and A ⊂∆
B, B ⊂∆ A, then A ∼∆ B. We present a counterexample to demonstrate
In
that this does not always hold in the case of non-self-adjoint algebras.
Section 2, we also present a characterisation of the relation ⊂∆ in the terms
of reflexive lattices.
In Section 3, we present an analogous theory defining the relation ⊂∆ for
dual operator spaces. In this case, if X, Y are dual operator spaces such that
Y ⊂∆ X, then there exist projections p and q such that pX ⊆ X, Xq ⊆
X and Y ∼∆ pXq. We also define a weaker relation ⊂cb∆ . We say that
Y ⊂cb∆ X if there exist w∗-continuous completely bounded isomorphisms
φ : X → φ(X), ψ : Y → ψ(Y ) such that φ(X) ⊂∆ ψ(Y ). We present a
theorem describing ⊂cb∆ in the terms of stable isomorphisms ( Theorem 3.11),
and we investigate the problem of whether ⊂∆ is a transitive relation for dual
operator spaces ( Theorem 3.14).
In the following, we briefly describe the notions used in this paper. We
refer the reader to the books [3], [6], [7], [15] and [16] for further details. If
V is a linear space and S ⊆ V, then by [S] we denote the linear span of S.
If H, K are Hilbert spaces, then we write B(H, K) for the space of bounded
operators from H to K. We denote B(H, H) as B(H). If L is a subset of
B(H), then we write L′ for the commutant of L, and L′′ for (L′)′. If A is an
operator algebra, then by ∆(A) we denote its diagonal A∩ A∗. A ternary ring
of operators M, referred to as a TRO from this point, is a subspace of some
B(H, K) satisfying the following:
m1, m2, m3 ∈ M ⇒ m1m∗
2m3 ∈ M.
It is well known that in the case that M is norm closed,
it is equal to
[M M ∗M]−k·k. If X is a dual operator space and I is a cardinal, then we
write MI(X) for the set of I × I matrices whose finite submatrices have uni-
formly bounded norm. We underline that MI(X) is also a dual operator
space, and it is completely isometrically and w∗-homeomorphically isomor-
phic with X ¯⊗B(l2(I)). Here, ¯⊗ denotes the normal spatial tensor product.
We say that two dual operator spaces X and Y are stably isomorphic if there
exists a cardinal I and a w∗-continuous completely isometric map from MI (X)
onto MI (Y ). If L ⊆ B(H) is a lattice, then we write Alg(L) for the algebra
MORITA EMBEDDINGS FOR DUAL OPERATOR ALGEBRAS AND DUAL OPERATOR SPACES3
of operators x ∈ B(H) satisfying
If A ⊆ B(H) is an algebra, then we write Lat(A) for the lattice of projections
l ∈ B(H) satisfying
A lattice L is called reflexive if
l⊥xl = 0, ∀ l ∈ L.
l⊥xl = 0, ∀ x ∈ A.
L = Lat(Alg(L)).
A reflexive algebra is an algebra of the form Alg(L), for some lattice L. An
important example of a class of reflexive lattices is given by nests. A nest
N ⊆ B(H) is a totally ordered set of projections containing the zero and
identity operators, which is closed under arbitrary suprema and infima. The
corresponding algebra Alg(N ) is called a nest algebra. If A ⊆ B(H) is a w∗-
closed algebra and I is cardinal, then we write AI for the algebra of operators
x ∈ B(H ⊗ l2(I)) satisfying
x((ξi)i∈I) = (a(ξi))i∈I , ∀ (ξi)i∈I ∈ H ⊗ l2(I)
for some a ∈ A.
2. Morita embeddings for dual operator algebras
We consider the following known theorem concerning von Neumann alge-
bras.
Theorem 2.1. Let A, B be von Neumann algebras. Then, the following are
equivalent:
(i) There exist w∗-continuous, one-to-one, ∗-homomorphisms
α : A → B(H), β : B → B(K),
where H, K are Hilbert spaces such that the commutants α(A)′, β(B)′ are ∗-
isomorphic.
(ii) The algebras A, B are weakly Morita equivalent in the sense of Rieffel.
(iii) There exists a cardinal I and a ∗-isomorphism from MI (A) onto
MI(B).
Definition 2.1. [8] Let A, B be w∗-closed algebras acting on the Hilbert spaces
H and K, respectively. We call these weakly TRO-equivalent if there exists a
TRO M ⊆ B(H, K) such that
A = [M ∗BM]−w∗
, B = [M AM ∗]−w∗
.
In this case, we write A ∼T RO B.
The following defines our notion of weak Morita equivalence for dual oper-
ator algebras.
4
G. K. ELEFTHERAKIS
Definition 2.2. [9] Let A, B be dual operator algebras. We call these weakly
∆-equivalent if there exist w∗-continuous completely isometric homomorphisms
α and β, respectively, such that α(A) ∼T RO β(B). In this case, we write
A ∼∆ B.
The following theorem is a generalisation of Theorem 2.1 to the setting of
unital dual operator algebras:
Theorem 2.2. Let A, B be unital dual operator algebras. Then, the following
statements are equivalent:
(i) There exist reflexive lattices L1 and L2, w∗-continuous completely iso-
β : B → Alg(L2), and a
metric onto homomorphisms α : A → Alg(L1),
∗-isomorphism θ : ∆(A)′ = L′′
(ii) The algebras A, B are weakly ∆-equivalent.
(iii) There exists a cardinal I and a w∗-continuous completely isometric
1 → ∆(B)′ = L′′
2 such that θ(L1) = L2.
homomorphism from MI(A) onto MI(B).
The previous theorem has been proved in various papers.
In fact, if (i)
holds, then by Theorem 3.3 in [8] Alg(L1) ∼T RO Alg(L2), and thus A ∼∆ B.
Conversely, if (ii) holds, then by Theorems 2.7 and 2.8 in [10], by choosing
a w∗-continuous completely isometric homomorphism α : A → α(A) with
reflexive range, there exists a w∗-continuous completely isometric homomor-
phism β : B → β(B), also with a reflexive range, such that α(A) ∼T RO β(B).
Thus, by Theorem 3.3 in [8], (i) holds. The equivalence of (ii) and (iii) con-
stitutes the main result of [12].
Remark 2.3. In the remainder of this section, if A is a unital dual operator
algebra and p ∈ A is a projection, then pAp is also a dual operator algebra
with unit p. If A is a w∗-closed unital algebra acting on the Hilbert space H
and p ∈ A is a projection, then we identify pAp with the algebra pAp(H) ⊆
B(p(H)).
2.1. TRO-embeddings for dual operator algebras.
Definition 2.3. Let A, B be unital w∗-closed algebras acting on the Hilbert
spaces H and K, respectively. We say that B weakly TRO-embeds into A if
there exists a projection p ∈ A such that B ∼T RO pAp. In this case, we write
B ⊂T RO A.
Remark 2.4. The above definition is equivalent to the following statements.
Let A, B be unital w∗-closed algebras acting on the Hilbert spaces H and K,
respectively. Then,
(i) The algebra B weakly TRO-embeds into A if and only if there exists a
TRO M ⊆ B(H, K) such that B = [M ∗AM]−w∗
, M BM ∗ ⊆ A.
MORITA EMBEDDINGS FOR DUAL OPERATOR ALGEBRAS AND DUAL OPERATOR SPACES5
(ii) The algebra B weakly TRO-embeds into A if and only if there exists a
TRO M ⊆ B(H, K) such that B = [M ∗AM]−w∗, M M ∗ ⊆ A.
Remark 2.5. If A is a unital w∗-closed algebra and p ∈ A is a projection,
then pAp ⊂T RO A. For the proof, we can take the linear span of the element
p as a TRO.
Proposition 2.6. Suppose that A, B, C are unital w∗-closed algebras acting
on the Hilbert spaces H, K, and L, respectively. If C ⊂T RO B and B ⊂T RO A,
then C ⊂T RO A.
Proof. We may assume that there exist projections p ∈ ∆(B), q ∈ ∆(A) such
that
C ∼T RO pBp, B ∼T RO qAq.
By Proposition 2.8 in [8], there exists a TRO M such that
qAq = [M ∗BM]−w∗
,
, q∆(A)q = [M ∗M]−w∗
B = [M qAqM ∗]−w∗
∆(B) = [M M ∗]−w∗
Define N = pM. Then, we have that
,
.
(pM)(pM)∗(pM) = pM M ∗pM ⊆ pM M ∗∆(B)M ⊆ p∆(B)M ⊆ pM.
Thus, N is a TRO. Then, we have that
pBp = [pM qAqM ∗p]−w∗
= [N qAqN ∗]−w∗
= [N(N ∗N qAqN ∗N)N ∗]−w∗
,
and therefore
Thus,
[N ∗pBpN]−w∗
= [N ∗N qAqN ∗N]−w∗
.
Therefore,
pBp ∼T RO [N ∗N qAqN ∗N]−w∗
C ∼T RO [N ∗N qAqN ∗N]−w∗
.
We may assume that there exists a TRO L such that
.
,
= [L∗CL]−w∗
.
(2.1)
[N ∗N qAqN ∗N]−w∗
C = [LN ∗N qAqN ∗N L∗]−w∗
We make the following observations:
(2.2) [N q∆(A)qN ∗]−w∗
= [pM M ∗p]−w∗
Furthermore, N ∗N = M ∗pM ⊆ [M ∗M]−w∗ = q∆(A)q. Thus,
) ⊆ q∆(A)q.
(2.3)
Define D = [LN ∗N q]−w∗. We shall prove that D is a TRO. We have that
L∗L ⊆ ∆([N ∗N qAqN ∗N]−w∗
= [pM M ∗M M ∗p]−w∗
= [N N ∗]−w∗
.
(LN ∗N q)(qN ∗N L∗)(LN ∗N q) = LN ∗N qN ∗(N L∗LN ∗)N q.
By (2.3), it holds that
N L∗LN ∗ ⊆ N q∆(A)qN ∗ ⊆ N N ∗.
6
Thus,
G. K. ELEFTHERAKIS
By (2.2), we have that N qN ∗ ⊆ N N ∗. Thus,
DD∗D ⊆ [LN ∗N qN ∗N N ∗N q]−w∗
.
DD∗D ⊆ [LN ∗N N ∗N N ∗N q]−w∗
= D.
Thus, D is a TRO. By (2.1), we have that
Furthermore,
By (2.3), we have that
C = [DAD∗]−w∗
.
D∗D = qN ∗N L∗LN ∗N q.
In addition, by (2.2) we have that
D∗D ⊆ qN ∗N q∆(A)qN ∗N q.
D∗D ⊆ qN ∗N N ∗N q ⊆ qN ∗N q ⊆ q∆(A)q ⊆ A.
Thus, by Remark 2.4 (ii), we have that C ⊂T RO A.
Remark 2.7. In light of the above proposition, one could expect that the rela-
tion ⊂T RO is a partial order relation in the class of unital w∗-closed operator
algebras, if we identify those algebras that are TRO-equivalent. This means
that ⊂T RO has the additional property that
(cid:3)
A ⊂T RO B, B ⊂T RO A ⇒ A ∼T RO B.
This is true in the case of von Neumann algebras, as we will show in Section
1.3. However, it fails in the case of non-self-adjoint algebras, as we prove in
Section 1.4.
The following Lemma will be useful.
Lemma 2.8. Let B be a w∗-closed unital operator algebra acting on the
Hilbert space H, and let q ∈ B be a projection.
If p is the projection
onto ∆(B)(q(H)), then p is a central projection for the algebra ∆(B), and
qBq ∼T RO pBp.
Proof. Clearly, p is a central projection for ∆(B). We consider the TRO
M = ∆(B)q ⊆ B(q(H), p(H)).
M(q(H)) = p(H), M ∗(p(H)) = q(H),
We have that
and
Then, Proposition 2.1 in [8] implies that
M ∗pBpM ⊆ qBq, M qBqM ∗ ⊆ pBp.
[M ∗pBpM]−w∗
= qBq,
[M qBqM ∗]−w∗
= pBp.
MORITA EMBEDDINGS FOR DUAL OPERATOR ALGEBRAS AND DUAL OPERATOR SPACES7
(cid:3)
2.2. ∆-embeddings for dual operator algebras.
Definition 2.4. Let A, B be dual operator algebras. We say that B weakly
∆-embeds into A if there exist w∗-continuous completely isometric homomor-
phisms α : A → α(A),
β : B → β(B) such that β(B) ⊂T RO α(A). In this
case, we write B ⊂∆ A.
The following theorem is a generalisation of Theorem 2.2.
Theorem 2.9. Let A, B be unital dual operator algebras. Then, the following
are equivalent:
(i) There exist reflexive lattices L1,L2 acting on the Hilbert spaces H and
K, respectively, w∗-continuous completely isometric onto homomorphisms
α : A → Alg(L1), β : B → Alg(L2),
and an onto w∗-continuous ∗-homomorphism
θ : ∆(A)′ = L′′
1 → ∆(B)′ = L′′
2
such that θ(L1) = L2.
(ii) B ⊂∆ A.
(iii) There exists a cardinal I, a projection q ∈ A, and a w∗-continuous
completely isometric homomorphism from MI(B) onto MI(qAq).
Proof. The equivalence of (ii) and (iii) is a consequence of Definition 2.4 and
Theorem 2.2.
It suffices to prove that
(i) ⇒ (ii)
Define the lattice
and the spaces
Alg(L2) ⊂T RO AlgL1.
L = {(cid:18) l 0
0 θ(l) (cid:19) :
l ∈ L1}
U = {x : l⊥xθ(l) = 0 ∀ l ∈ L1}, V = {y : θ(l)⊥yl = 0 ∀ l ∈ L1}.
We can easily prove that
Because the map
V
Alg(L) =(cid:18) Alg(L1) U
1 → L′′, a →(cid:18) a 0
Alg(L2) (cid:19) .
0 θ(a) (cid:19)
ρ : L′′
8
G. K. ELEFTHERAKIS
is a ∗-isomorphism such that θ(L1) = L, from Theorem 3.3 in [8] we have
that
Define the TRO
Now, observe that
and
Thus,
Alg(L1) ∼T RO Alg(L).
M = (0 C).
MAlg(L)M ∗ = Alg(L2)
M ∗M ⊆ Alg(L).
Then, Proposition 2.6 implies that
Alg(L2) ⊂T RO AlgL.
Alg(L2) ⊂T RO AlgL1.
(iii) ⇒ (i)
Suppose that MI (B) and MI(qAq) are completely isometrically and w∗-
homeomorphically isomorphic. Then, B ∼∆ qAq. Every unital dual operator
algebra has a w∗-completely isometric representation whose image is reflexive.
Thus, we may assume that α : A → B(H) is a w∗-continuous completely
isometric homomorphism such that α(A) = Alg(L1) for a reflexive lattice L1.
If p = α(q), then
α(qAq) = Alg(L1p(H)).
Theorem 2.7 in [10] implies that there exists a reflexive lattice L2 and a
w∗-continuous completely isometric homomorphism β : B → Alg(L2) such
that
Alg(L2) ∼T RO Alg(L1p(H)).
Thus, by Theorem 3.3 in [8] there exists a ∗-isomorphism ρ : (L1p(H))′′ → L′′
such that ρ(L1p(H)) = L2. If τ : L′′
1p(H) : x → xp(H), then we write
θ = ρ ◦ τ. This is the required map.
Theorem 2.10. Let A, B, C be unital dual operator algebras such that
1 → L′′
2
(cid:3)
C ⊂∆ B, B ⊂∆ A.
Then, C ⊂∆ A.
Proof. We may assume that there exist projections p ∈ B, q ∈ A such that
C ∼T RO pBp, β(B) ∼T RO qAq,
where β : B → β(B) is a w∗-continuous completely isometric homomor-
phism. From Theorem 2.7 in [10], we know that for the representation
MORITA EMBEDDINGS FOR DUAL OPERATOR ALGEBRAS AND DUAL OPERATOR SPACES9
βpBp : pBp → β(p)β(B)β(p) there exists a w∗-continuous completely iso-
metric homomorphism γ : C → γ(C) such that
By Proposition 2.6, we have that
γ(C) ∼T RO β(p)β(B)β(p).
β(p)β(B)β(p) ⊂T RO qAq.
(cid:3)
Because qAq ⊂T RO A, we have that γ(C) ⊂T RO A, and thus C ⊂∆ A.
Remark 2.11. In view of Theorem 2.10, one should expect that weak ∆-
embedding is a partial order relation in the class of unital dual operator al-
gebras if we identify those unital dual operator algebras that are weakly ∆-
equivalent. Thus, one should expect that A ⊂∆ B, B ⊂∆ A ⇒ A ∼∆ B.
In Section 1.4 we shall see that this is not true. However, in the case of von
Neumann algebras, this is indeed true. For further details, see Section 2.3
below.
Example 2.12. Let A = Alg(N1), B = Alg(N2), where N1 is a continuous
nest, and N2 is a nest with at least one atom. We shall prove that it is
impossible that B ⊂∆ A.
Proof. Suppose on the contrary that B ⊂∆ A. Thus, there exists a projection
p ∈ ∆(A) such that B ∼∆ pAp. Because B and pAp are nest algebras, it
follows from Theorem 3.2 in [10] that B ∼T RO pAp. Thus, by Theorem 3.3 in
[8] there exists a homeomorphism θ : N2 → N1p. This is impossible, because
N2 contains an atom, and N1p is a continuous nest.
(cid:3)
2.3. The case of von Neumann algebras.
Lemma 2.13. Let A be a von Neumann algebra, and let p, q be central pro-
jections of A such that p ≤ q and A ∼T RO Ap. Then, A ∼T RO Aq.
Proof. By Theorem 3.3 in [8], there exists a ∗-isomorphism θ : A′ → A′p. We
need to prove that there exists a ∗-isomorphism ρ : A′ → A′q. Suppose that
e0 = IdA, e1 = q, e2 = p, en = θ(en−2), n = 2, 3, ...
Clearly (en)n is a decreasing sequence of central projections. Observe that
Thus, the map ρ : A′ → A′q sending
⊕(e2n − e2n+1) ⊕ (e2n+1 − e2n+2)! ⊕ ∧nen.
e0 = ∞
Xn=0
⊕a(e2n − e2n+1) ⊕ a(e2n+1 − e2n+2)! ⊕ (a ∧n en)
a = ∞
Xn=0
10
to
G. K. ELEFTHERAKIS
ρ(a) = ∞
Xn=0
is a ∗-isomorphism .
⊕θ(a)(e2n+2 − e2n+3) ⊕ a(e2n+1 − e2n+2)! ⊕ (a ∧n en)
(cid:3)
The above Lemma is based on the fact if p, q are central projections of the
von Neumann algebra A such that p ≤ q and A ∼= Ap, then A ∼= Aq, where
∼= is the ∗-isomorphism. We acknowledge that this was known to the authors
of [5] (see the proof of Lemma 6.2.3). In this Lemma, an alternative proof to
ours was provided.
Theorem 2.14. The relation ⊂T RO is a partial order relation for von Neu-
mann algebras, if we identify those von Neumann algebras that are TRO-
equivalent.
Proof. Let A, B be von Neumann algebras. It suffices to prove the implication
that
A ⊂T RO B, B ⊂T RO A ⇒ A ∼T RO B.
Let q0 ∈ B, p0 ∈ A be projections such that
A ∼T RO q0Bq0, B ∼T RO p0Ap0.
By Lemma 2.8, there exist central projections q ∈ B, p ∈ A such that
A ∼T RO Bq, B ∼T RO Ap.
Thus, there exist ∗-isomorphisms
θ : A′ → B′q, ρ : B′ → A′p.
We can easily see that there exists a central projection p ∈ A such that p ≤ p
and
ρ(B′q) = A′ p.
Therefore, we obtain a ∗-isomorphism from A′ onto A′ p. Because p ≤ p and
p, p are central, Lemma 2.13 implies that there exists a ∗-isomorphism from
A′ onto A′p. Thus, A ∼T RO Ap, which implies that A ∼T RO B.
Theorem 2.15. Let A, B be von Neumann algebras. Then, the following are
equivalent:
(cid:3)
continuous onto ∗-homomorphism θ : α(A)′ → β(B)′.
(i) There exist ∗-isomorphisms α : A → α(A), β : B → β(B) and a w∗-
(ii) B ⊂∆ A.
(iii) There exists a cardinal I and a w∗-continuous onto ∗-homomorphism
ρ : MI (A) → MI(B).
MORITA EMBEDDINGS FOR DUAL OPERATOR ALGEBRAS AND DUAL OPERATOR SPACES11
Proof. The equivalence of (i) and (ii) is a consequence of Theorem 2.9.
(ii) ⇒ (iii)
Suppose that B ⊂∆ A. By Theorem 2.9, we may assume that there exists a w∗-
continuous onto ∗-homomorphism θ : A′ → B′. Suppose that A′p⊥ = Kerθ
for a projection p in the centre of A. We also assume that A ⊆ B(H). Then,
the map
is a ∗-isomorphism. Because
A′p(H) → B′ : ap(H) → θ(a)
Theorem 2.1 implies that there exists a cardinal I and a w∗-continuous onto
∗-isomorphism
(Ap(H))′ = A′p(H),
MI (pAp) → MI (B).
Suppose that ρ : MI (A) → MI(B) is a w∗-continuous onto ∗-homomorphism.
Let p be a projection in the centre of MI(A) such that
(iii) ⇒ (ii)
MI(A)p⊥ = Kerρ.
Because Z(MI(A)) = Z(A)I , where Z(A) (resp. Z(MI(A))) is the centre of
A (resp. MI(A)), we may assume that p = qI for q ∈ Z(A). Thus, the map
MI(Aq) → MI(B) : (ai,jq) → ρ((ai,j))
(cid:3)
is a ∗-isomorphism. Then, Theorem 2.9 implies that B ⊂∆ A.
Theorem 2.16. The weak ∆-embedding is a partial order relation in the class
of von Neumann algebras, if we identify those von Neumann algebras that are
weakly Morita equivalent in the sense of Rieffel.
Proof. Claim: Let A be a von Neumann algebra, and let r ∈ A be a projec-
tion such that A ∼∆ rAr. If q is a projection in A such that r ≤ q, then it
also holds that A ∼∆ qAq.
Proof of the claim: There exists a cardinal I such that the algebras
MI(A) and MI(rAr) are ∗- isomorphic. We suppose that M = MI (A), N =
MI(qAq). Then, we have that M ∼= MI (rAr) = rI N rI and N ∼= qI M qI .
By Lemma 6.2.3 in [5], the von Neumann algebras M and N are stably
isomorphic. Thus, M ∼∆ N. However,
A ∼T RO M,
qAq ∼T RO N.
Therefore, A ∼∆ qAq, and the proof of the claim is complete.
To prove the theorem, it suffices to prove that if A, B are von Neumann
algebras such that A ⊂∆ B, B ⊂∆ A, then A ∼∆ B. We may assume that
12
G. K. ELEFTHERAKIS
there exist projections p ∈ B, q ∈ A and w∗-continuous completely isometric
homomorphisms α : A → α(A), β : B → β(B), such that
α(A) ∼T RO pBp, β(B) ∼T RO qAq.
For the representation βpBp : pBp → β(p)β(B)β(p), there exists a w∗-
continuous one-to-one ∗-homomorphism γ : α(A) → γ(α(A)) such that
γ(α(A)) ∼T RO β(p)β(B)β(p).
By Proposition 2.6, we have that
β(p)β(B)β(p) ⊂T RO qAq.
Therefore, there exists a projection r ≤ q such that
γ(α(A)) ∼T RO rAr ⇒ A ∼∆ rAr.
The claim implies that A ∼∆ qAq. However, qAq ∼∆ B. Thus A ∼∆ B. Thus,
the proof is complete.
(cid:3)
Corollary 2.17. Let A, B be von Neumann algebras, I, J be cardinals, and
θ : MI (A) → MI(B), ρ : MJ (B) → MJ (A)
be onto w∗-continuous homomorphisms. Then, A ∼∆ B.
Proof. By Theorem 2.15, A ⊂∆ B and B ⊂∆ A. The conclusion then follows
from the above theorem.
Corollary 2.18. Let A, B be unital dual operator algebras such that A ⊂∆ B
and B ⊂∆ A. Then, ∆(A) ∼∆ ∆(B).
Proof. We can easily see that ∆(A) ⊂∆ ∆(B) and ∆(B) ⊂∆ ∆(A). Now, we
can apply the above theorem.
(cid:3)
(cid:3)
Example 2.19. Let A be a factor, and B be a unital dual operator algebra
such that B ⊂∆ A. Then, B is a von Neumann algebra, and B ∼∆ A.
Proof. There exist a ∗-isomorphism α : A → α(A), a w∗- continuous com-
pletely isometric homomorphism β : B → β(B), and a TRO M such that if
p is the projection onto [M M ∗]−w∗, then
β(B) = [M ∗α(A)M]−w∗
,
pα(A)p = [M β(B)M ∗]−w∗
.
Define N = [α(A)pM]−w∗
. Because
it follows that
M M ∗pα(A) ⊆ pα(A) ⊆ α(A),
pM M ∗p ⊆ α(A) ⇒ [α(A)pM M ∗pα(A)]−w∗
= [N N ∗]−w∗
MORITA EMBEDDINGS FOR DUAL OPERATOR ALGEBRAS AND DUAL OPERATOR SPACES13
is an ideal of α(A). However, α(A) is a factor, and thus α(A) = [N N ∗]−w∗.
On the other hand,
[N ∗N]−w∗
= [M ∗pα(A)α(A)pM]−w∗
= [M ∗α(A)M]−w∗
= β(B).
Thus, A and B are weakly Morita equivalent in the sense of Rieffel. However,
in the case of von Neumann algebras, Rieffel's Morita equivalence is the same
as ∆-equivalence.
(cid:3)
2.4. A counterexample in non-self-adjoint operator algebras. Despite
the situation for von Neumann algebras, we shall prove that if A, B are uni-
tal non-self-adjoint dual operator algebras, it does not always hold that the
implication
(2.4)
A ⊂∆ B, B ⊂∆ A ⇒ A ∼∆ B.
By Theorem 3.12 in [10], if A, B are nest algebras, then
Because for every nest algebra B and every projection p ∈ B the algebra pBp
is a nest algebra, we can conclude that
A ∼∆ B ⇔ A ∼T RO B.
A ⊂∆ B ⇔ A ⊂T RO B.
Thus, in order to prove that (2.4) does not hold, it suffices to find nest algebras
A and B such that A ⊂T RO B, B ⊂T RO A and A is not TRO-equivalent
to B. Let m be the Lebesgue measure on the Borel sets of the interval [0, 1].
Suppose that Q is the set of rationals, and Q+ (resp. Q−) is the projection
onto l2(Q∩ [0, t]) (resp. l2(Q∩ [0, t)) ). Furthermore, let Nt be the projection
onto L2([0, t], m) for 0 ≤ t ≤ 1. We define the nest
N = {Q+
t ⊕ Nt, Q−
t ⊕ Nt,
0 ≤ t ≤ 1}.
By A = Alg(N ), we denote the corresponding nest algebra acting on the
Hilbert space
H = l2(Q ∩ [0, 1]) ⊕ L2([0, 1], m).
The above nest appeared in Example 7.18 in [6]. Suppose that f (t) = 1
[0, 1], and define
2 , t ∈
M = {Q+
We can define a unitary
f (t) ⊕ Nf (t), Q−
f (t) ⊕ Nf (t),
u2 : L2([0, 1]) → L2([0,
1
2
0 ≤ t ≤ 1}.
])
such that u2(χΩ) = √2χf (Ω) where χΩ is the characteristic function of the
Borel set Ω. This unitary maps Nt onto Nf (t) in the sense that
u2Ntu∗
2 = Nf (t), 0 ≤ t ≤ 1.
14
G. K. ELEFTHERAKIS
Furthermore, the map
t , Q−
{Q+
t onto Qj
t
: 0 ≤ t ≤ 1} −→ {Q+
f (t), Q−
f (t) : 0 ≤ t ≤ 1}
sending Qj
f (t) for j = +,− is a nest isomorphism. Because these nests
are multiplicity free (they generate a maximal abelian self-adjoint algebra,
referred to as an MASA from this point) and totally atomic, the above map
extends as a ∗-isomorphism between the corresponding MASAs. Thus, there
exists a unitary
such that
u1 : l2(Q ∩ [0, 1]) → l2(Q ∩ [0,
1
2
])
u2Q+
t u∗
2 = Q+
f (t), u2Q−
t u∗
2 = Q−
f (t), 0 ≤ t ≤ 1.
Therefore, the unitary u = u1 ⊕ u2, implies a unitary equivalence between N
and M.
Let s be the projection
s : l2(Q ∩ [0, 1]) → l2(Q ∩ [0,
1
2
])
and r be the projection
r : L2([0, 1], m) → L2([0,
1
2
], m).
If p = s⊕ r, then p ∈ A and pAp = Alg(M). By the above arguments, A and
pAp are unitarily equivalent, and thus they are TRO-equivalent. Suppose
that q0 is the projection
0 IL2([0,1])−r (cid:19) ∈ A
q0 =(cid:18) 0 0
and q = p + q0. Because p ≤ q ≤ IdA, we have that
pAp ⊂T RO qAq ⊂T RO A.
However, A ∼T RO pAp. This implies that A ⊂T RO qAq. Thus, if (2.4) holds,
then we should have that
A ∼T RO qAq.
Suppose that L is the nest Lat(qAq). By Theorem 3.3 in [8], there exists a
∗-isomorphism
θ : ∆(A)′ → (∆(A)q(H))′
such that θ(N ) = L. However, the algebras ∆(A), ∆(A)q(H) are MASAs.
Therefore, there exists a unitary w : q(H) → H such that
θ(x) = w∗xw, ∀x ∈ ∆(A) = ∆(A)′.
We have that
A = wqAqw∗.
MORITA EMBEDDINGS FOR DUAL OPERATOR ALGEBRAS AND DUAL OPERATOR SPACES15
We can easily see that L = L1 ∪ L2, where
t ⊕ Nt,
t ⊕ Nt, Q−
L1 = {Q+
0 ≤ t ≤
1
2}
and
L2 = {Q+
2 ⊕ Nt,
1
1
2 ≤ t ≤ 1}.
Observe that L1 ≤ L0 ≤ L2 for all Li ∈ Li, i = 1, 2 where L0 = Q+
2 ⊕ N 1
1
2
. If
M0 = wL0w∗, N1 = wL1w∗, N2 = wL2w∗,
then
M1 ≤ M0 ≤ M2
for all Mi ∈ Ni, i = 1, 2. Suppose that M0 = Q+
N2 = {Q+
t ⊕ Nt, Q−
t ⊕ Nt,
If
L2 = {(Q+
1
2 ⊕ Nt) − L0 :
we can consider L2 to be a nest acting on L2([ 1
t0 ⊕ Nt0. Then,
t0 ≤ t ≤ 1}.
1
2 ≤ t ≤ 1},
2 , 1], m). Furthermore, if
N2 = {(Q+
t ⊕ Nt) − M0, (Q−
t ⊕ Nt) − M0,
t0 ≤ t ≤ 1},
then L2 and N2 are isomorphic nests. However, this is impossible, because L2
is a continuous nest and N2 is a nest with atoms. This contradiction shows
that A and qAq are not TRO-equivalent.
Remark 2.20. Let A and q be as above. As we have seen, qAq ⊂∆ A, A ⊂∆
qAq but A and qAq are not ∆-equivalent. We can prove further that they are
not Morita equivalent even in the sense of Blecher and Kashyap [1], [14]. If
they were, then by [11] N and L would be isomorphic as nests. However, we
can see that this is impossible by applying the same arguments as above.
3. Morita embeddings for dual operator spaces
Definition 2.1 can be adapted to the setting of dual operator spaces as
follows.
Definition 3.1. [13] Let H1, H2, K1, K2 be Hilbert spaces, and let
X ⊆ B(H1, H2), Y ⊆ B(K1, K2)
be w∗-closed spaces. We call these weakly TRO-equivalent if there exist TROs
Mi ⊆ B(Hi, Ki), i = 1, 2 such that
X = [M ∗
2 Y M1]−w∗
, Y = [M2XM ∗
1 ]−w∗
.
In this case, we write X ∼T RO Y.
16
G. K. ELEFTHERAKIS
Remark 3.1. If W1, W2 are Hilbert spaces and Z is a subspace of B(W1, W2),
then we call it nondegenerate if Z(W1) = W2, Z ∗(W2) = W1. If H1, H2, K1,
K2, X, Y are as in the above definition, p2 (resp. q2) is the projection onto
X(H1) (resp. Y (K1)), and p1 (resp. q1) is the projection onto X ∗(H2) (resp.
Y ∗(K2)), then the spaces p2Xp1(H1), q2Y q1(K1) are nondegenerate, and also
weakly TRO-equivalent. This can be concluded from Proposition 2.2 in [13].
The following defines our notion of weak Morita equivalence for dual oper-
ator spaces.
Definition 3.2. [13] Let X, Y be dual operator spaces. We call these weakly
∆-equivalent if there exist w∗-continuous completely isometric maps φ, ψ, re-
spectively, such that φ(X) ∼T RO ψ(Y ). In this case, we write X ∼∆ Y.
The following theorem constitutes the main result of [13].
Theorem 3.2. Let X, Y be dual operator spaces. Then, the following are
equivalent:
(i) X ∼∆ Y.
(ii) There exists a cardinal I and a w∗-continuous completely isometric
map from MI(X) onto MI(Y ).
Remark 3.3. Throughout Section 3, we shall employ the following notation.
If H1, H2 are Hilbert spaces, X ⊆ B(H1, H2) is a w∗-closed subspace, and
q ∈ B(H1), p ∈ B(H2) are projections such that pX ⊆ X, Xq ⊆ X, then by
pXq we denote the space {pxq : x ∈ X} ⊆ B(H1, H2). This space is w∗-closed
and completely isometrically and w∗-homeomorphically isomorphic with the
space pXq(H1) ⊆ B(q(H1), p(H2)).
3.1. TRO-embeddings for dual operator spaces.
Definition 3.3. Let H1, H2, K1, K2 be Hilbert spaces, and let X ⊆ B(H1, H2),
Y ⊆ B(K1, K2) be w∗-closed spaces. We say that Y weakly TRO embeds into
X if there exist TROs M1 ⊆ B(H1, K1) and M2 ⊆ B(H2, K2) such that
Y = [M2XM ∗
1 M1 ⊆ X. In
this case, we write Y ⊂T RO X.
Remark 3.4. We can easily see that if Y ⊂T RO X, then there exist projec-
tions p, q such that pX ⊆ X, Xq ⊆ X and Y ∼T RO pXq.
Examples 3.5. (i) If X ∼T RO Y , then clearly X ⊂T RO Y and Y ⊂T RO X.
2 Y M1 ⊆ X and M ∗
2 M2X ⊆ X, XM ∗
1 ]−w∗
, M ∗
(ii) If Ki, Wi, i = 1, 2 are Hilbert spaces,
are w∗-closed spaces, and
Y ⊆ B(K1, K2), Z ⊆ B(W1, W2)
X = Y ⊕ Z ⊆ B(K1 ⊕ W1, K2 ⊕ W2),
MORITA EMBEDDINGS FOR DUAL OPERATOR ALGEBRAS AND DUAL OPERATOR SPACES17
then Y ⊂T RO X. For the proof, we apply the TROs M1 = (CIK1, 0), M2 =
(CIK2, 0).
(iii) If X ⊆ B(H1, H2) is a w∗-closed operator space and p ∈ B(H2), q ∈
B(H1) are projections such that pX ⊆ X, Xq ⊆ X, then pXq ⊂T RO X.
(iv) A generalisation of W ∗-modules over von Neumann algebras is given
by the projectively w∗-rigged modules over unital dual operator algebras. See
[2] for more details. Given a unital dual operator algebra A, a projectively
w∗-rigged module over A is a dual operator space Z that is completely isomet-
rically and w∗-homeomorphically isomorphic to a space Y = [M A]−w∗
, where
M is a TRO satisfying M ∗M ⊆ A. Observe that Y = [M AC]−w∗, M ∗Y C ⊆
A and M ∗M A ⊆ A, AC ⊆ A. Thus, Y ⊂T RO A. Therefore, for every pro-
jectively w∗-rigged module Z over a unital dual operator algebra A, we have
that Z ⊂∆ A. Here, ⊂∆ is the relation defined in Definition 3.4 below.
Proposition 3.6. Let X, Y, Z be w∗-closed operator spaces. If
Z ⊂T RO Y, Y ⊂T RO X,
then there exist projections p, q such that pX ⊆ X, Xq ⊆ X and Z ⊂T RO
pXq.
Proof. There exist projections p, q, r, s such that
pX ⊆ X, Xq ⊆ X, rY ⊆ Y, Y s ⊆ Y
and TROs Mi, Ni, i = 1, 2 such that
1 ]−w∗
1 ]−w∗
Y = [M2pXqM ∗
Z = [N2rY sN ∗
, pXq = [M ∗
, rY s = [N ∗
,
2 Y M1]−w∗
2 ZN1]−w∗
.
We may assume that
M2p = M2, M1q = M1, N2r = N2, N1s = N1.
Suppose that Di is the W ∗-algebra generated by the set
Define
{MiM ∗
i } ∪ {N ∗
i Ni}, i = 1, 2.
Because M1M ∗
1 ⊆ D1, N ∗
Li = [NiDiMi]−w∗
, i = 1, 2.
1 N1 ⊆ D1, it follows that
N1D1M1M ∗
1 D1N ∗
1 N1D1M1 ⊆ N1D1M1,
and thus
Therefore, L1, and similarly L2, are TROs. Now, we have that
L1L∗
1L1 ⊆ L1.
[L2pXqL∗
1]−w∗
= [N2D2M2pXqM ∗
1 D1N ∗
1 ]−w∗
= [N2D2Y D1N ∗
1 ]−w∗
.
18
Because
G. K. ELEFTHERAKIS
[M2M ∗
2 Y ]−w∗
= Y = [Y M1M ∗
1 ]−w∗
, N ∗
2 N2Y ⊆ Y, Y N ∗
1 N1 ⊆ Y,
we have that D2Y = Y = Y D1. Thus,
= [N2Y N ∗
[L2pXqL∗
(3.1)
1]−w∗
1 ]−w∗
Furthermore,
2ZL1 ⊆ [M ∗
L∗
2 D2N ∗
[M ∗
2 ZN1D1M1]−w∗
2 D2Y D1M1]−w∗
= [M ∗
2 Y M1]−w∗
.
= [N2rY sN ∗
1 ]−w∗
= Z.
= [M ∗
2 D2rY sD1M1]−w∗ ⊆
Thus,
(3.2)
On the other hand,
L∗
2ZL1 ⊆ pXq.
L∗
2L2pXq ⊆ [M ∗
[M ∗
2 D2N ∗
2 D2M2M ∗
2 N2D2M2pXq]−w∗ ⊆ [M ∗
2 D2M2pXq]−w∗
=
2 Y M1]−w∗ ⊆ [M ∗
2 Y M1]−w∗
.
2L2pXq ⊆ pXq, and similarly pXqL∗
Thus, L∗
tions (3.1) and (3.2) imply that Z ⊂T RO pXq.
Remark 3.7. From the above proof, we isolate the fact that if Z ⊂T RO Y
and Y ∼T RO X, then Z ⊂T RO X.
3.2. ∆-embeddings for dual operator spaces.
1L1 ⊆ pXq. Therefore, the rela-
(cid:3)
Definition 3.4. Let X, Y be dual operator spaces. We say that Y weakly
∆-embeds into X if there exist w∗-continuous completely isometric maps
φ : X → φ(X), ψ : Y → ψ(Y )
such that ψ(Y ) ⊂T RO φ(X). In this case, we write Y ⊂∆ X.
Definition 3.5. Let X, Y be dual operator spaces. A map φ : X → Y
that is one-to-one, w∗-continuous, and completely bounded with a completely
bounded inverse is called a w∗-c.b.
isomorphism, and the spaces X, Y are
called w∗-c.b. isomorphic. Under the above assumptions, the map φ−1 is also
w∗-continuous.
Definition 3.6. Let X, Y be dual operator spaces. We call these c.b. ∆-
equivalent if there exist w∗- c.b. isomorphisms φ : X → φ(X), ψ : Y → ψ(Y )
such that φ(X) ∼T RO ψ(Y ). In this case, we write X ∼cb∆ Y.
Definition 3.7. Let X, Y be dual operator spaces. We say that Y c.b. ∆-
embeds into X if there exist w∗-c.b. isomorphisms φ : X → φ(X), ψ : Y →
ψ(Y ) such that ψ(Y ) ⊂T RO φ(X). In this case, we write Y ⊂cb∆ X.
MORITA EMBEDDINGS FOR DUAL OPERATOR ALGEBRAS AND DUAL OPERATOR SPACES19
Remark 3.8. Observe the following:
(i) X ∼∆ Y ⇒ X ∼cb∆ Y
(ii) X ⊂∆ Y ⇒ X ⊂cb∆ Y
In what follows, if X is a dual operator space, then Ml(X) (resp. Mr(X))
denotes the algebra of left (resp. right) multipliers of X. In this case, Al(X) =
∆(Ml(X)), (resp. Ar(X) = ∆(Mr(X))) is a von Neumann algebra [3].
Lemma 3.9. Suppose that Z, Y are w∗-closed operator spaces satisfying Z ∼T RO
Y, H1, H2 are Hilbert spaces such that
Al(Y ) ⊆ B(H2), Ar(Y ) ⊆ B(H1),
and ψ : Y → B(H1, H2) is a w∗-continuous complete isometry such that
Al(Y )ψ(Y )Ar(Y ) ⊆ ψ(Y ).
Then, there exists a w∗-continuous complete isometry ζ : Z → ζ(Z) such that
ζ(Z) ∼T RO ψ(Y ).
Proof. Assume that M1, M2 are TROs such that
Z = [M ∗
2 Y M1]−w∗
, Y = [M2ZM ∗
1 ]−w∗
.
By Remark 3.1, we may assume that Z and Y are nondegenerate spaces. We
denote
A = [M ∗
2 M2]−w∗
, B = [M ∗
1 M1]−w∗
, C = [M2M ∗
2 ]−w∗
, D = [M1M ∗
1 ]−w∗
.
The algebras
Ω(Z) =(cid:18) A Z
0 B (cid:19) , Ω(Y ) =(cid:18) C Y
0 D (cid:19)
are weakly TRO-equivalent as algebras. Indeed,
Ω(Z) = [M ∗Ω(Y )M]−w∗
, Ω(Y ) = [MΩ(Z)M ∗]−w∗
,
where M is the TRO M2 ⊕ M1. If c ∈ C, then define
γ(c) : ψ(Y ) → ψ(Y ), γ(c)ψ(y) = ψ(cy).
We can easily see that γ(c) ∈ Al(Y ) and kγ(c)k ≤ 1. Thus, γ : C → Al(Y )
is a contractive homomorphism and hence a ∗-homomorphism. If γ(c) = 0,
then cY = 0. Because Y is nondegenerate, we conclude that c = 0. Thus,
γ is a one-to-one ∗-homomorphism. Similarly, there exists a one-to-one ∗-
homomorphism δ : D → Ar(Y ) such that ψ(y)δ(d) = ψ(yd), ∀ y. The map
π : Ω(Y ) → π(Ω(Y )), given by
π(cid:18)(cid:18) c y
0 d (cid:19)(cid:19) =(cid:18) γ(c) ψ(y)
δ(d) (cid:19) ,
is a w∗-continuous completely isometric homomorphism. This can be shown
by applying 3.6.1 in [3]. By Theorem 2.7 in [10], there exists a w∗-continuous
0
20
G. K. ELEFTHERAKIS
completely isometric homomorphism ρ : Ω(Z) → ρ(Ω(Z)) and a TRO N such
that
ρ(Ω(Z)) = [N ∗π(Ω(Y ))N]w∗
, π(Ω(Y )) = [N ρ(Ω(Z))N ∗]w∗
.
As in the discussion concerning the map Φ below Theorem 2.5 in [13], the
map ρ is given by
ρ(cid:18)(cid:18) a z
0 b (cid:19)(cid:19) =(cid:18) α(a) ζ(z)
β(b) (cid:19) ,
0
where α : A → α(A), ζ : Z → ζ(Z), β : B → β(B) are completely isometric
maps. By Lemma 2.8 in [13], the TRO N is of the form N = N2 ⊕ N1. Thus,
ζ(Z) = [N ∗
2 ψ(Y )N1]−w∗
, ψ(Y ) = [N2ζ(Z)N ∗
1 ]−w∗
.
(cid:3)
Lemma 3.10. Let Z, Ω, X be dual operator spaces. We assume that Z ∼T RO
Ω, and that ψ0 : Ω → ψ0(Ω) is a w∗-continuous complete isometry such that
ψ0(Ω) ⊂T RO X. Then, there exists a w∗-c.b. isomorphism φ : X → φ(X) and
a w∗-continuous complete isometry ζ : Z → ζ(Z) such that ζ(Z) ⊂T RO φ(X).
Thus, Z ⊂cb∆ X.
Proof. Suppose that
Al(Ω) ⊆ B(H2), Ar(Ω) ⊆ B(H1)
and ψ : Ω → B(H1, H2) is a w∗-continuous complete isometry such that
Al(Ω)ψ(Ω)Ar(Ω) ⊆ ψ(Ω).
By Lemma 3.9, there exists a w∗-continuous complete isometry ζ : Z → ζ(Z)
such that
We assume that p, q are projections such that pX ⊆ X, Xq ⊆ X and
ζ(Z) ∼T RO ψ(Ω).
ψ0(Ω) ∼T RO pXq.
Again by Lemma 3.9, there exists a w∗-continuous complete isometry φ :
pXq → φ(pXq) such that ψ(Ω) ∼T RO φ(pXq). Define
xq⊥
,
for all x ∈ X. Observe that φ is a w∗-continuous completely bounded and
one-to-one map. If φ∞ is the ∞×∞ amplification of φ, then φ∞ has a closed
range. Thus, by the open mapping theorem, φ∞ has a bounded inverse.
φ(x) =
p⊥xq
0
φ(pxq) 0
0
0
0
0
MORITA EMBEDDINGS FOR DUAL OPERATOR ALGEBRAS AND DUAL OPERATOR SPACES21
Therefore, φ−1 is completely bounded. We have that ζ(Z) ∼T RO φ(pXq).
Thus, there exist TROs M1, M2 such that
ζ(Z) = [M2φ(pXq)M ∗
1 ]−w∗
, φ(pXq) = [M ∗
2 ζ(Z)M1]−w∗
.
Define the TROs
We can see that
and
Ni =(cid:0) Mi 0 0 (cid:1) ,
i = 1, 2.
[N2 φ(X)N ∗
1 ]−w∗
= ζ(Z), N ∗
2 ζ(Z)N1 ⊆ φ(X)
1 N1 ⊆ φ(X).
2 N2 φ(X) ⊆ φ(X), φ(X)N ∗
N ∗
Thus, ζ(Z) ⊂T RO φ(X).
Theorem 3.11. Let X, Y be dual operator spaces. Then, the following are
equivalent.
(cid:3)
(i) Y ⊂cb∆ X.
(ii) There exist w∗-c.b.
φ : X → φ(X);
projections p, q such that pφ(X) ⊆ φ(X), φ(X)q ⊆ φ(X); a cardinal I; and
a completely isometric w∗-continuous onto map
isomorphisms ψ : Y → ψ(Y ),
Proof.
π : MI (ψ(Y )) → MI(pφ(X)q).
By definition, there exist w∗-c.b. isomorphisms ψ : Y → ψ(Y ), φ : X →
φ(X) such that
(i) ⇒ (ii)
ψ(Y ) ⊂T RO φ(X).
There exist projections p, q such that pφ(X) ⊆ φ(X),
φ(X)q ⊆ φ(X)
and ψ(Y ) ∼T RO pφ(X)q. By Theorem 3.2, there exists a cardinal I and a
completely isometric w∗-continuous onto map
π : MI (ψ(Y )) → MI(pφ(X)q).
Define
(ii) ⇒ (i)
pφ(x)q 0
0
0
p⊥φ(x)q
0
φ(x) =
0
0
φ(x)q⊥
,
for all x ∈ X. As in the proof of Lemma 3.10, we can see that φ is a w∗-c.b.
isomorphism. By Example 3.5 (ii), we have that
φ(X).
pφ(X)q ⊂T RO
22
G. K. ELEFTHERAKIS
By Theorem 3.2, it holds that ψ(Y ) ∼∆ pφ(X)q. Thus, there exist completely
isometric w∗-continuous maps
µ : ψ(Y ) → µ(ψ(Y )), χ : pφ(X)q → χ(pφ(X)q)
such that µ(ψ(Y )) ∼T RO χ(pφ(X)q). Now, apply Lemma 3.10 for
Z = µ(ψ(Y )), Ω = χ(pφ(X)q), ψ0 = χ−1 : Ω → pφ(X)q.
We have that
We conclude that
ψ0(Ω) ⊂T RO
φ(X).
µ(ψ(Y )) ⊂cb∆
φ(X) ⇒ Y ⊂cb∆ X.
(cid:3)
Lemma 3.12. Suppose that Z ⊆ B(W1, W2), Y ⊆ B(K1, K2) are w∗-closed
spaces such that Z ⊂T RO Y. We also assume that p2 is the projection onto
Y (K1), and p1 is the projection onto Y ∗(K2). Thus, Y0 = p2Y p1(K1) is a
nondegenerate space into B(p1(K1), p2(K2)). We shall prove that Z ⊂T RO Y0.
Proof. By definition, there exist TROs Ni ⊆ B(Wi, Ki), i = 1, 2 such that
Z = [N ∗
2 Y N1]−w∗
, N2ZN ∗
1 ⊆ Y, N2N ∗
2 Y ⊆ Y, Y N1N ∗
1 ⊆ Y.
Define Mi = piNi ⊆ B(Wi, pi(Ki)), i = 1, 2. We have that
p2y = y ⇒ p2m1m∗
2y = m1m∗
2y, ∀ y ∈ Y, m1, m2 ∈ N2.
Thus,
p2N2N ∗
2 p2 = N2N ∗
2 p2 ⇒ p2N2N ∗
2 p2 = p2N2N ∗
2 .
The above relation implies that
2 M2 = p2N2N ∗
M2M ∗
2 p2N2 = p2N2N ∗
2 N2 ⊆ p2N2 = M2.
Therefore, M2 is a TRO. Similarly, M1 is also a TRO. Now, we have that
[M ∗
2 Y0M1]−w∗
= [N ∗
2 p2Y p1N1]−w∗
= [N ∗
2 Y N1]−w∗
= Z
and
Furthermore,
M2ZM ∗
1 = p2N2ZN ∗
1 p1 ⊆ p2Y p1 = Y0.
M2M ∗
2 Y0 = p2N2N ∗
2 p2Y p1 = p2N2N ∗
2 Y p1 ⊆ p2Y p1 = Y0.
Therefore, Y0 is a nondegenerate subspace of B(p1(K1), p2(K2)), and Z ⊂T RO
Y0.
(cid:3)
The Lemma below is weaker than Lemma 3.9.
MORITA EMBEDDINGS FOR DUAL OPERATOR ALGEBRAS AND DUAL OPERATOR SPACES23
Lemma 3.13. Suppose that Z, Y are w∗-closed operator spaces satisfying
Z ⊂T RO Y , H1, H2 are Hilbert spaces such that
Al(Y ) ⊆ B(H2), Ar(Y ) ⊆ B(H1),
and ψ : Y → B(H1, H2) is a w∗-continuous complete isometry such that
Al(Y )ψ(Y )Ar(Y ) ⊆ ψ(Y ).
Then, there exists a w∗-continuous complete isometry ζ : Z → ζ(Z) such that
ζ(Z) ⊂T RO ψ(Y ).
Proof. Assume that M1, M2 are TROs such that
1 ⊆ Y, M2M ∗
2 Y ⊆ Y, Y M1M ∗
2 Y M1]−w∗
, M2ZM ∗
1 ⊆ Y.
Z = [M ∗
By Lemma 3.12, we may assume that Y is nondegenerate. Suppose that p is
the identity of [M2M ∗
. Then, we have
that pY ⊆ Y, Y q ⊆ Y. We denote A, B, C, D, Ω(Z) as in Lemma 3.9, and
2 ]−w∗ and q is the identity of [M1M ∗
1 ]−w∗
If c ∈ C, then define map
Ω(pY q) =(cid:18) C pY q
0 D (cid:19) .
ψ(Y ) → ψ(Y ) : ψ(y) → ψ(cpy).
Clearly, this map belongs to Al(Y ), and thus there exists γ(c) ∈ Al(Y ) satis-
fying
γ(c)ψ(y) = ψ(cpy) ∀ y ∈ Y.
Note that we can define a ∗-homomorphism γ : C → Al(Y ). If γ(c) = 0,
then cy = 0 for all y ∈ Y , and thus because Y is nondegenerate, it follows
that c = 0. Therefore, γ is one-to-one. Similarly, there exists a one-to-one
∗-homomorphism
such that
δ : D → Ar(Y )
ψ(y)δ(d) = ψ(yqd) ∀ y ∈ Y.
We can conclude that there exist projections p ∈ Al(Y ), q ∈ Ar(Y ) such that
ψ(py) = pψ(y), ψ(yq) = ψ(y)q, ∀ y ∈ Y.
The map π : Ω(pY q) → π(Ω(pY q)), given by
π(cid:18)(cid:18) c pyq
0 d (cid:19)(cid:19) =(cid:18) γ(c) ψ(pyq)
δ(d)
0
(cid:19) ,
is a w∗-continuous completely isometric homomorphism. Because
Ω(Z) ∼T RO Ω(pY q),
24
G. K. ELEFTHERAKIS
as in Lemma 3.9 we can find a w∗-continuous completely isometric map ζ :
Z → ζ(Z) and TROs N1, N2 such that
ζ(Z) = [N2ψ(pY q)N ∗
γ(C) = [N ∗
1 ]−w∗
2 N2]−w∗
,
, ψ(pY q) = [N ∗
2 ζ(Z)N1]−w∗ ⊆ ψ(Y ),
1 N1]−w∗
.
δ(D) = [N ∗
Because
N ∗
2 N2ψ(Y ) ⊆ γ(C)ψ(Y ) = ψ(CY ) ⊆ ψ(Y ),
1 N1 ⊆ ψ(Y ), we have that ζ(Z) ⊂T RO ψ(Y ).
and similarly ψ(Y )N ∗
Theorem 3.14. Let X, Y, Z be dual operator spaces such that Z ⊂∆ Y, Y ⊂∆
X. Then:
(i) There exist a w∗ continuous complete isometry χ : X → χ(X) and pro-
jections p, q such that pχ(X) ⊆ χ(X), χ(X)q ⊆ χ(X) and Z ⊂∆ pχ(X)q.
(ii) Z ⊂cb∆ X.
(cid:3)
Proof. Suppose that H1, H2 are Hilbert spaces such that
Al(Y ) ⊆ B(H2), Ar(Y ) ⊆ B(H1)
and ψ : Y → B(H1, H2) is a w∗-continuous complete isometry such that
By Lemma 3.13, there exists a w∗-continuous complete isometry ζ : Z → ζ(Z)
such that
Al(Y )ψ(Y )Ar(Y ) ⊆ ψ(Y ).
Because Y ⊂∆ X, there exist a w∗-continuous complete isometry χ : X →
χ(X) and projections p, q such that pχ(X) ⊆ χ(X), χ(X)q ⊆ χ(X) and
ζ(Z) ⊂T RO ψ(Y ).
Y ∼∆ pχ(X)q.
By Lemma 3.9, there exists a w∗-continuous complete isometry
such that
φ : pχ(X)q → φ(pχ(X)q)
Thus, by Remark 3.7 it holds that
ψ(Y ) ∼T RO φ(pχ(X)q).
There exist projections r, s and TROs M1, M2 such that
ζ(Z) ⊂T RO φ(pχ(X)q) ⇒ Z ⊂∆ pχ(X)q.
rφ(pχ(X)q) ⊆ φ(pχ(X)q), φ(pχ(X)q)s ⊆ φ(pχ(X)q)
and
ζ(Z) = [M2rφ(pχ(X)q)sM ∗
1 ]−w∗
, rφ(pχ(X)q)s = [M ∗
2 ζ(Z)M1]−w∗
.
For every x ∈ X, define
φ(x) = rφ(pχ(x)q)s ⊕ r⊥φ(pχ(x)q)s ⊕ φ(pχ(x)q)s⊥ ⊕ p⊥χ(x)q ⊕ χ(x)q⊥.
MORITA EMBEDDINGS FOR DUAL OPERATOR ALGEBRAS AND DUAL OPERATOR SPACES25
As in the proof of Lemma 3.10, we can see that φ is a w∗-c.b. isomorphism
from X onto φ(X). Define the TROs
Ni = (Mi 0 0 0 0),
i = 1, 2.
We can see that
[N2 φ(X)N ∗
1 ]−w∗
= ζ(Z), N ∗
2 N2 φ(X) ⊆ φ(X), φ(X)N ∗
N ∗
2 ζ(Z)N1 ⊆ φ(X)
1 N1 ⊆ φ(X).
and
Thus,
ζ(Z) ⊂T RO
φ(X) ⇒ Z ⊂cb∆ X.
(cid:3)
Example 3.15. Let Y be a dual operator space, and let H be a Hilbert space
such that Y ⊂∆ B(H). Then, Y ∼∆ B(H).
Proof. There exist w∗-continuous completely isometric maps ψ : Y → ψ(Y )
and projections p, q such that pφ(B(H)) ⊆ φ(B(H)), φ(B(H))q ⊆ φ(B(H))
and
ψ(Y ) ∼T RO pφ(B(H))q.
We define the map α : B(H) → B(H) given by α(x) = φ−1(pφ(x)). This is a
multiplier of B(H), and also a projection. Because Al(B(H)) = B(H), there
exists a projection p ∈ B(H) such that
φ−1(pφ(x)) = px ⇒ φ(px) = pφ(x) ∀ x ∈ B(H).
Similarly, there exists a projection q ∈ B(H) such that
φ(xq) = φ(x)q ∀ x ∈ B(H).
We have that
Because
φ−1(pφ(B(H))q) = pB(H)q.
Al(pB(H)q) = B(p(H)), Ar(pB(H)q) = B(q(H)),
it follows from Lemma 3.9 that there exists a w∗-continuous completely iso-
metric map ζ : ψ(Y ) → ζ(ψ(Y )) such that
ζ(ψ(Y )) ∼T RO pB(H)q.
Thus, there exist TROs M1, M2 such that
ζ(ψ(Y )) = [M2 pB(H)qM ∗
1 ]−w∗
.
Define
N ∗
2 = [M2 pB(H)]w∗
, N1 = [B(H)qM ∗
1 ]−w∗
.
26
G. K. ELEFTHERAKIS
Then, we have that
[N2N ∗
2 N2]w∗
= [B(H)pM ∗
2 M2 pB(H)pM ∗
2 ]w∗
= [B(H)pM ∗
2 ]−w∗
= N2.
Thus, N2 is a TRO. Similarly, N1 is a TRO. Now,
ζ(ψ(Y )) = [N ∗
2 B(H)N1]−w∗
and
Thus,
[N2B(H)N ∗
1 ]−w∗
= [B(H)pM ∗
2 B(H)M1 qB(H)]−w∗
= B(H).
ζ(ψ(Y )) ∼T RO B(H) ⇒ Y ∼∆ B(H).
(cid:3)
Example 3.16. Let H be a Hilbert space, Φ : B(H) → B(H) be a w∗-
continuous completely bounded idempotent map, and
Y = RanΦ, Z = RanΦ⊥.
Let φ : B(H) → Y ⊕ Z be the map given by φ(x) = Φ(x) ⊕ Φ⊥(x). We can
easily prove that φ is a w∗-c.b.
isomorphism onto Y ⊕ Z. By Example 3.5
(ii), we have that Y ⊂T RO φ(B(H)), and thus Y ⊂cb∆ B(H).
Example 3.17. Let H be a Hilbert space, and let e, f ∈ B(H) be nontrivial
projections. Define
Φ : B(H) → B(H), Φ(x) = exf + exf ⊥ + e⊥xf ⊥,
and denote Y = RanΦ. By Example 3.16, we have that Y ⊂cb∆ B(H). If
Y weakly ∆-embeds into B(H), then by Example 3.15 we should have that
Y ∼∆ B(H). However, this contradicts the fact that B(H) is a self-adjoint
algebra and Y is a non-self-adjoint algebra. Thus, the relation Y ⊂cb∆ X
does not always imply that Y ⊂∆ X holds.
Acknowledgement:
I would like to express appreciation to Dr Evgenios
Kakariadis for his helpful comments and suggestions during the preparation
of this work.
References
[1] D. P. Blecher and U. Kashyap, Morita equivalence of dual operator algebras, J. Pure
Appl. Algebra, 212 (2008), 2401-2412
[2] D. P. Blecher and J. E. Kraus, On a generalization of W*-modules, Banach Center
Publ. 91, (2010), 77-86
[3] D. P. Blecher and C. Le Merdy, Operator Algebras and Their Modules -- An Operator
Space Approach, Oxford University Press, 2004
[4] D. P. Blecher, P. S. Muhly, and V. I. Paulsen, Categories of operator modules -- -Morita
equivalence and projective modules, Memoirs of the A.M.S. 143 (2000) no. 681
MORITA EMBEDDINGS FOR DUAL OPERATOR ALGEBRAS AND DUAL OPERATOR SPACES27
[5] D. P. Blecher and V. Zarikian, The calculus of one sided M-ideals and multipliers in
operator spaces, Memoirs of the A.M.S. 179 (842), 2003
[6] K. R. Davidson, Nest Algebras, Longman Scientific & Technical, Harlow, 1988
[7] E. Effros and Z.-J. Ruan, Operator Spaces, Oxford University Press, 2000
[8] G.K. Eleftherakis, TRO equivalent algebras, Houston J. of Mathematics, 38:1 (2012),
153-175
[9] G.K. Eleftherakis, A Morita type equivalence for dual operator algebras, J. Pure Appl.
Algebra, 212:5 (2008), 1060-1071
[10] G.K. Eleftherakis, Morita type equivalences and reflexive algebras, J. Operator Theory,
64 (2010) no 1, 3-17
[11] G.K. Eleftherakis, Morita equivalence of nest algebras, Math. Scand., 113 (2013), no
1, 83-107
[12] G.K. Eleftherakis, V. I. Paulsen, Stably isomorphic dual operator algebras, Math.
Ann., 341:1 (2008), 99-112
[13] G. K. Eleftherakis, V. I. Paulsen, and I. G. Todorov, Stable isomorphism of dual
operator spaces, J. Funct. Anal. 258 (2010), 260 -- 278
[14] U. Kashyap, A Morita theorem for dual operator algebras, J. Funct. Analysis, 256
(2009) 3545-3567
[15] V. I. Paulsen, Completely Bounded Maps and Operator Algebras, Cambridge University
Press, 2002
[16] G. Pisier, Introduction to Operator Space Theory, Cambridge University Press, 2000
G. K. Eleftherakis, University of Patras, Faculty of Sciences, Depart-
ment of Mathematics, 265 00 Patras Greece
E-mail address: [email protected]
|
1812.07194 | 1 | 1812 | 2018-12-18T06:31:24 | Quotients of \'etale groupoids and the abelianizations of groupoid C*-algebras | [
"math.OA"
] | In this paper, we introduce quotients of \'etale groupoids. Using the notion of quotients, we describe the abelianizations of groupoid C*-algebras. As another application, we obtain a simple proof that effectiveness of an \'etale groupoid is implied by the full uniqueness property of its groupoid C*-algebra. | math.OA | math |
QUOTIENTS OF ´ETALE GROUPOIDS AND THE
ABELIANIZATIONS OF GROUPOID C*-ALGEBRAS
FUYUTA KOMURA
Abstract. In this paper, we introduce quotients of ´etale groupoids.
Using the notion of quotients, we describe the abelianizations of groupoid
C*-algebras. As another application, we obtain a simple proof that effec-
tiveness of an ´etale groupoid is implied by the full uniqueness property
of its groupoid C*-algebra.
0. Introduction
The study of C*-algebras associated to ´etale groupoids, groupoid C*-
algebras, was initiated by Renault in [7]. Since then, many researchers
have studied the relationship between ´etale groupoids and groupoid C*-
algebras.
In the previous studies, there are many results for C*-algebras
associated to Hausdorff ´etale groupoids. For a non-Hausdorff ´etale groupoid,
its C*-algebra seems not to be studied sufficiently. However, non-Hausdorff
groupoids naturally arise as mentioned in [3], [5] and so on. Exel pointed
out that some results known for Hausdorff ´etale groupoids do not necessarily
hold for non-Hausdorff groupoids in [4]. In [2], the authors treat simplicity
of groupoid C*-algebras associated to non-Hausdorff ´etale groupoids.
In this paper, we calculate the abelianization of a groupoid C*-algebra.
For a discrete group Γ, the abelianization C∗(Γ)ab of its group C*-algebra
C∗(Γ) is isomorphic to C∗(Γab), where Γab is the abelianization of Γ. Fur-
thermore, C∗(Γab) is isomorphic to C((cid:100)Γab), where (cid:100)Γab is the Pontryagin
(cid:100)Gab so that C∗(G)ab (cid:39) C∗(Gab) (cid:39) C0((cid:100)Gab) holds. In order to construct
dual of Γab. It is natural to consider an ´etale groupoid analogy. For an ´etale
groupoid G, we construct an ´etale groupoid Gab and a topological groupoid
Gab, we introduce the notion of quotient ´etale groupoids. A quotient ´etale
groupoid often becomes non-Hausdorff even if the original ´etale groupoid
is Hausdorff. Therefore, we treat not necessarily Hausdorff ´etale groupoids
and their C*-algebras, which are defined by Connes [3]. As a byproduct, we
obtain a simple proof that effectiveness of an ´etale groupoid is implied by the
full uniqueness property of its groupoid C*-algebra (see Corollary 2.2.5). We
remark that this result has been shown in [1] for Hausdorff ´etale groupoids
in a different way and the proof in [1] seems to work for non-Hausdorff ´etale
groupoids.
1
2
FUYUTA KOMURA
This paper is organized as follows.
In Section 1, we recall definitions
and basic facts about not necessarily Hausdorff ´etale groupoids and their
C*-algebras.
In Section 2, we introduce the notion of quotient ´etale groupoids and
show some applications. Using quotients, we obtain a simple proof that full
uniqueness property of a groupoid C*-algebra induces effectiveness of an
´etale groupoid.
In Section 3, for an ´etale groupoid G, we construct an ´etale abelian group
bundle Gab through quotients. Finally, we show that the abelianization of
a groupoid C*-algebra is isomorphic to the C*-algebra associated to Gab.
Since the abelianization of a C*-algebra is commutative, it is isomorphic to
C0(X) for some locally compact Hausdorff space X by the Gelfand-Naimark
theorem. We show that the abelianization of a groupoid C*-algebra C∗(G)
is isomorphic to C0((cid:100)Gab), where (cid:100)Gab is introduced in this paper.
We obtain some results by using quotients of ´etale groupoids, which are
not necessarily Hausdorff. We expect that this paper stimulates the study
of non-Hausdorff ´etale groupoids.
Acknowledgement. The author would like to thank his supervisor, Prof.
Takeshi Katsura, for fruitful discussions and guidance.
1. ´Etale groupoids and groupoid C*-algebras
In this section, we recall the notions of ´etale groupoids and groupoid
C*-algebras. We refer to [7], [6] and [9] for details.
1.1. ´Etale groupoids. A groupoid is a set G together with a distinguished
subset G(0) ⊂ G, source and range maps s, r : G → G(0) and a multiplication
G(2) ··= {(α, β) ∈ G × G s(α) = r(β)} (cid:51) (α, β) (cid:55)→ αβ ∈ G
such that
(1) for all x ∈ G(0), s(x) = x and r(x) = x hold,
(2) for all α ∈ G, αs(α) = r(α)α = α holds,
(3) for all (α, β) ∈ G(2), s(αβ) = s(β) and r(αβ) = r(α) hold,
(4) if (α, β), (β, γ) ∈ G(2), we have (αβ)γ = α(βγ),
(5) every γ ∈ G, there exists γ(cid:48) ∈ G which satisfies (γ(cid:48), γ), (γ, γ(cid:48)) ∈ G(2)
and s(γ) = γ(cid:48)γ and r(γ) = γγ(cid:48).
Since the element γ(cid:48) in (5) is uniquely determined by γ, γ(cid:48) is called the
inverse of γ and denoted by γ−1. We call G(0) the unit space of G. A
subgroupoid of G is a subset of G which is closed under the inversion and
multiplication. For U ⊂ G(0), we define GU ··= s−1(U ) and GU ··= r−1(U ).
We define also Gx ··= G{x} and Gx ··= G{x} for x ∈ G(0). The isotropy
bundle of G is denoted by Iso(G) ··= {γ ∈ G s(γ) = r(γ)}. If G satisfies
G = Iso(G), G is called a group bundle over G(0). A group bundle G is said
to be abelian if Gx is an abelian group for all x ∈ G(0).
THE ABELIANIZATIONS OF GROUPOID C*-ALGEBRAS
3
A topological groupoid is a groupoid equipped with a topology where the
multiplication and the inverse are continuous. Note that the source map
and range map of a topological groupoid are continuous.
Definition 1.1.1. A topological groupoid G is said to be ´etale if
(1) the unit space G(0) ⊂ G is a locally compact Hausdorff space with
respect to the relative topology of G,
(2) the source map s : G → G(0) is locally homeomorphic (i.e. for all
α ∈ G, there exists an open neighborhood U ⊂ G of α such that
s(U ) ⊂ G(0) is open and sU is a homeomorphism onto s(U )).
An ´etale topological groupoid is called an ´etale groupoid in short.
In
this paper, we assume that the unit space of an ´etale groupoid is a locally
compact Hausdorff space. We do not assume that an ´etale groupoid is a
Hausdorff space as a topological space.
Note that a local homeomorphism is an open map. If s is locally homeo-
morphic, then r is also locally homeomorphic since r(γ) = s(γ−1) holds for
all γ ∈ G. By Definition 1.1.1, the family of all locally compact Hausdorff
open subsets of G is an open basis for the topology of G. An ´etale groupoid
G is said to be effective if G(0) = Iso(G)◦, where Iso(G)◦ denotes the interior
of Iso(G).
In some papers, the condition that the source map s : G → G(0) is locally
homeomorphic in Definition 1.1.1 is replaced by the condition that the source
map s : G → G is locally homeomorphic. As in Proposition 1.1.2, these
definitions are equivalent.
Proposition 1.1.2. Let G be an ´etale groupoid. The unit space G(0) is an
open subset of G. Furthermore, the source and range maps s, r are local
homeomorphisms as maps from G to G.
Proof. First, we show that G(0) is an open subsets of G. Take x ∈ G(0)
arbitrarily. Then, there exists an open subset U ⊂ G with x ∈ U such that
s(U ) is an open subsets of G(0) and sU is a homeomorphism onto s(U ).
Define V ··= U ∩ s−1(G(0) ∩ U ). Observe that V is an open subset of G
with x ∈ V . Moreover, observe that V ⊂ G(0). Indeed, for every γ ∈ V ,
we have s(γ) ∈ U and s(γ) = s(s(γ)). Since sU is injective, it follows that
γ = s(γ) ∈ G(0). Therefore, we have shown V ⊂ G(0) and this implies
that G(0) is an open subset of G. Now, the second assertion can be easily
(cid:3)
checked.
Definition 1.1.3. Let G be an ´etale groupoid. A subset U ⊂ G is called a
bisection if both sU and rU are injective.
For an ´etale groupoid G, an open bisection of G is a locally compact
Hausdorff space because it is homeomorphic to open subset of G(0) and we
assume that G(0) is locally compact Hausdorff. Note that the set of all open
bisections composes an open basis of G.
4
FUYUTA KOMURA
Example 1.1.4. Every locally compact Hausdorff space is regarded as a
Hausdorff ´etale groupoid whose unit space coincides with the whole space.
Example 1.1.5. Every topological group is regarded as a topological groupoid
whose unit space is a singleton. A topological group is discrete if and only
if it is ´etale as a topological groupoid.
Example 1.1.6. Let X be a locally compact Hausdorff space and Γ be a
discrete group. We denote by Aut(X) the set of all homeomorphisms on
X, which is a group under the composition. An action of Γ on X is a
group homomorphism α : Γ (cid:51) s (cid:55)→ αs ∈ Aut(X) and written α : Γ (cid:121) X.
Now, we construct the groupoid associated to an action α : Γ (cid:121) X. Define
Γ (cid:110)α X ··= Γ × X as a topological space. The unit space of Γ (cid:110)α X is
X, which is identified with the subset of Γ (cid:110)α X via an inclusion X (cid:51)
x (cid:55)→ (e, x) ∈ Γ (cid:110)α X. The source map and range map are defined by
s((t, x)) = x and r((t, x)) = αt(x) respectively for (t, x) ∈ Γ (cid:110)α X. For a
pair (t1, y), (t2, x) ∈ Γ (cid:110)α X with y = αt2(x), their multiplication is defined
by (t1, y) · (t2, x) ··= (t1t2, x). An inverse is given by (t, x)−1 = (t−1, αt(x)).
Then, Γ (cid:110)α X is a Hausdorff ´etale groupoid.
Proposition 1.1.7 ([6, Proposition 2.2.4]). Let G be an ´etale groupoid and
U, V ⊂ G be open sets. Then, a set U V ··= {αβ ∈ G α ∈ U, β ∈ V, s(α) =
r(β)} ⊂ G is an open set. Furthermore, if U, V ⊂ G are open bisections,
U V is also an open bisection.
Proof. Take γ ∈ U V and an open bisection W ⊂ G with γ ∈ W . Then,
there exist α ∈ U and β ∈ V such that γ = αβ. By the continuity of the
multiplication of G, there exist open bisections U1, V1 ⊂ G with α ∈ U1 ⊂
U, β ∈ V1 ⊂ V and U1V1 ⊂ W . Note that r(U1V1) = r(U1 ∩ s−1(r(V1))) ⊂
G(0) is an open subset. Therefore, U1V1 = r−1(r(U1V1)) ∩ W ⊂ G is open.
Now, we have γ ∈ U1V1 ⊂ U V , so U V ⊂ G is an open subset.
Assume that U, V ⊂ G are open bisections. One can show that sU V and
rU V are injective. Therefore, U V is an open bisection.
(cid:3)
Definition 1.1.8. Let G be a groupoid. A subset F ⊂ G(0) is said to be
invariant if s(γ) ∈ F implies r(γ) ∈ F for all γ. A point x ∈ G(0) is called a
fixed point if {x} ⊂ G(0) is invariant.
Note that a set F ⊂ G(0) is invariant if and only if G(0) \ F is invariant.
If F ⊂ G(0) is invariant, then GF = GF ∩ GF ⊂ G is a subgroupoid whose
unit space is F .
Proposition 1.1.9. Let G be an ´etale groupoid. Then, the set of all fixed
points F ⊂ G(0) is a closed subset.
Proof. We show that G(0) \ F ⊂ G(0) is an open set. Take x ∈ G(0) \
F . Then, there exists γ ∈ G such that x = s(γ) and x (cid:54)= r(γ). Take
an open bisection U which contains γ. Let SU : s(U ) → r(U ) denote a
homeomorphism defined by SU (s(α)) = r(α) for each α ∈ U . Since G(0)
THE ABELIANIZATIONS OF GROUPOID C*-ALGEBRAS
5
is Hausdorff, there exist open sets U1, V1 ⊂ G(0) such that s(γ) ∈ U1 ,
r(γ) ∈ V1 and U1 ∩ V1 = ∅. By the continuity of SU , there exists an open set
U2 ⊂ U such that γ ∈ U2 and SU (U2) ⊂ V1. Now, one can see U2 ⊂ G(0)\ F .
Therefore, G(0) \ F ⊂ G(0) is an open set.
(cid:3)
We will use the next proposition for the set of all fixed points.
Proposition 1.1.10. Let G be an ´etale groupoid and U, F ⊂ G(0) be an
invariant open and closed subset respectively. Then, GU ⊂ G is an open
subgroupoid of G and an ´etale groupoid in the relative topology. Similarly,
GF ⊂ G is a closed subgroupoid of G and an ´etale groupoid in the relative
topology.
Proof. Observe that U and F are locally compact Hausdorff spaces in
the relative topology of G(0). Now, it is clear that GU and GF are ´etale
(cid:3)
groupoids.
In particular, if x ∈ G(0) is a fixed point, then Gx ⊂ G is a discrete group.
1.2. ´Etale groupoid C*-algebras. Following Connes's idea in [3], we as-
sociate a C*-algebra to an ´etale groupoid which is not necessarily Hausdorff.
Let G be an ´etale groupoid. For an open Hausdorff subset U ⊂ G, we
denote the set of all continuous functions with compact support on U by
Cc(U ). We regard an element in Cc(U ) as an element in Funct(G), the
vector space of all complex valued functions on G, by defining it to be 0
U Cc(U ) ⊂ Funct(G), where the
union is taken over all open Hausdorff subsets U ⊂ G.
If G is Hausdorff, then C(G) coincides with Cc(G). If G is not Hausdorff,
an element in C(G) may not be continuous.
Proposition 1.2.1. Let G be an ´etale groupoid. Take an open basis {Ui}i∈I
of G consisting of open Hausdorff subsets. Then, C(G) is the linear span of
U Cc(U ), where the
i∈I Cc(Ui). In particular, C(G) is the linear span of(cid:83)
outside of U . We define C(G) ··= span(cid:83)
(cid:83)
union is taken over all open bisections of G.
(cid:3)
Proof. This follows from a partition of unity argument.
Definition 1.2.2. Let G be an ´etale groupoid. Recall that C(G) is equipped
with a structure of C-vector space by pointwise addition and scalar multi-
plication. The multiplication f ∗ g ∈ C(G) and involution f∗ ∈ C(G) of
f, g ∈ C(G) are defined by
(cid:88)
β∈Gs(γ)
f ∗ g(γ) =
f (γβ−1)g(β), f∗(γ) = f (γ−1).
Then, C(G) is a *-algebra under these operations.
One can see that Cc(G(0)) ⊂ C(G) is a *-subalgebra.
Lemma 1.2.3. Let G be an ´etale groupoid and f ∈ C(G). Then, there
exists Cf ≥ 0 such that (cid:107)ρ(f )(cid:107) ≤ Cf for all Hilbert spaces H and *-
homomorphisms ρ : C(G) → B(H).
6
FUYUTA KOMURA
Proof. We may assume that f ∈ Cc(U ) for some open bisection U ⊂ G.
One can see that f∗∗f ∈ Cc(G(0)). Since Cc(G(0)) is a union of commutative
C*-algebras, we have (cid:107)ρ(h)(cid:107) ≤ supx∈G(0)h(x) for all h ∈ Cc(G(0)). Then,
we obtain (cid:107)ρ(f )(cid:107)2 = (cid:107)ρ(f∗ ∗ f )(cid:107) ≤ supx∈G(0)f∗ ∗ f (x) < ∞.
(cid:3)
The universal norm of f ∈ C(G) is defined by
(cid:107)f(cid:107) ··= sup{(cid:107)ρ(f )(cid:107) ρ : C(G) → B(H) is a *-representation}.
is spanned by(cid:83)
By Lemma 1.2.3, the universal norm takes values in [0,∞). Since the left
regular representations of C(G) induces a faithful *-representation of C(G),
the universal norm becomes a C*-norm (see [2, Section 4]). The completion
of C(G) by universal norm is denoted by C∗(G). We shall remark that every
*-representation of C(G) induces the *-representation of C∗(G). Note that
the inclusion Cc(G(0)) ⊂ C(G) extends to C0(G(0)) ⊂ C∗(G).
Proposition 1.2.4. Let G be an ´etale groupoid and F ⊂ G(0) be a closed
invariant set. Then, the restriction C(G) (cid:51) f (cid:55)→ fGF ∈ C(GF ) extends to
the surjective *-homomorphism C∗(G) → C∗(GF ).
Proof. First, we check that fGF ∈ C(GF ) for all f ∈ C(G). We may
assume that f ∈ Cc(U ) for some open Hausdorff subset U ⊂ G, since C(G)
U Cc(U ), where the union is taken over all open Hausdorff
subsets U ⊂ G. Defining V ··= GF ∩ U , V is a Hausdorff open subset of GF .
Then fGF is contained in Cc(V ) ⊂ C(GF ).
Direct calculations imply that the restriction C(G) (cid:51) f (cid:55)→ fGF ∈ C(GF )
is a *-homomorphism.
Next, we show that the restriction C(G) (cid:51) f (cid:55)→ fGF ∈ C(GF ) is surjec-
tive. Note that {GF ∩ U U ⊂ G is an open Hausdorff subset} is an open
basis of GF . Take an open Hausdorff subset U ⊂ GF and f ∈ Cc(GF ∩ U )
arbitrarily. Put V ··= GF ∩ U . Since V ⊂ U is a closed subset of U and
f ∈ Cc(V ), there exists f ∈ Cc(U ) such that fV = f by the Tietze exten-
sion theorem. Now, we obtain f ∈ C(G) such that fGF = f . By Propo-
U Cc(GF ∩ U ), where the union
is taken over all open Hausdorff subsets U ⊂ G. Therefore, the restriction
C(G) (cid:51) f (cid:55)→ fGF ∈ C(GF ) is surjective.
By the universality of C∗(G), the restriction C(G) (cid:51) f (cid:55)→ fGF ∈ C(GF )
extends to the *-homomorphism C∗(G) → C∗(GF ). Since the image of
C∗(G) is dense in C∗(GF ), C∗(G) → C∗(GF ) is surjective.
(cid:3)
2. Quotients of ´etale groupoids
sition 1.2.1, C(GF ) is the linear span of (cid:83)
In this section, we introduce the notion of quotient ´etale groupoids.
2.1. Quotients of ´etale groupoids.
Definition 2.1.1. Let G be a groupoid. A subgroupoid H ⊂ G is said to
be normal if
(1) G(0) ⊂ H ⊂ Iso(G) holds and
THE ABELIANIZATIONS OF GROUPOID C*-ALGEBRAS
7
(2) αHα−1 ⊂ H holds for all α ∈ G.
Definition 2.1.2. Let G be a groupoid and H ⊂ G be a normal sub-
groupoid. Then, we define an equivalence relation ∼ on G by declaring that
α ∼ β if s(α) = s(β) and αβ−1 ∈ H. We denote the quotient set G/∼ by
G/H.
Lemma 2.1.3. Let G be a groupoid and H ⊂ G be a normal subgroupoid.
Suppose that α, α(cid:48) ∈ G satisfy α ∼ α(cid:48). Then, we have s(α) = s(α(cid:48)) and
r(α) = r(α(cid:48)).
Proof. It follows that s(α) = s(α(cid:48)) from the definition of α ∼ α(cid:48). Since
αα(cid:48)−1 ∈ H ⊂ Iso(G), we have r(α) = r(αα(cid:48)−1) = s(αα(cid:48)−1) = r(α(cid:48)).
(cid:3)
Lemma 2.1.4. Let G be a groupoid and H ⊂ G be a normal subgroupoid.
Suppose that α, α(cid:48), β, β(cid:48) ∈ G satisfy α ∼ α(cid:48), β ∼ β(cid:48), s(α) = r(β). Then, we
have s(α(cid:48)) = r(β(cid:48)) and αβ ∼ α(cid:48)β(cid:48).
Proof. By Lemma 2.1.3, we have s(α) = s(α(cid:48)) and r(β) = r(β(cid:48)). Using
s(α) = r(β), we obtain s(α(cid:48)) = r(β(cid:48)).
s(αβ) = s(β) = s(β(cid:48)) = s(α(cid:48)β(cid:48)) and
The last assertion follows from a simple calculation.
Indeed, we have
αβ(α(cid:48)β(cid:48))−1 = αββ(cid:48)−1α(cid:48)−1 = (αββ(cid:48)−1α−1)(αα(cid:48)−1) ∈ H.
Note that αββ(cid:48)−1α−1 ∈ H, since H is normal.
(cid:3)
Definition 2.1.5. Let G be a groupoid, H ⊂ G be a normal subgroupoid
and q : G → G/H be the quotient map. A groupoid structure of G/H is
defined as the following;
• a unit space (G/H)(0) is q(G(0)), which can be identified with G(0)
via an injection qG(0),
• source and range maps s, r : G/H → G(0) are defined by s(q(γ)) ··=
q(s(γ)), r(q(γ)) ··= q(r(γ)) for γ ∈ G,
• a multiplication of G/H is defined by q(α)q(β) ··= q(αβ) for α, β ∈ G
with s(α) = r(β).
One can see that the inverse map of G/H satisfies q(γ)−1 = q(γ−1) for
γ ∈ G. Then, G/H is a groupoid under these operations.
Remark 2.1.6. The operations of G/H are well-defined by Lemma 2.1.3
and Lemma 2.1.4.
If G is a topological groupoid, then we consider the quotient topology as
a topology of G/H.
Lemma 2.1.7. Let G be an ´etale groupoid and H ⊂ G be an open nor-
mal subgroupoid. Then, the quotient map q : G → G/H is an open map.
Furthermore, q is a local homeomorphism.
Proof. Let U ⊂ G be an open subset. Then, q−1(q(U )) = U H is an open
subset of G by Proposition 1.1.7. Hence, q(U ) ⊂ G/H is an open subset by
the definition of the quotient topology.
8
FUYUTA KOMURA
Next, we show that the quotient map q : G → G/H is a local homeomor-
phism. Fix a γ ∈ G. Then, take an open bisection U ⊂ G with γ ∈ U . One
can see that qU is injective. Since q is an open map, qU is a homeomorphism
onto an open subset q(U ) ⊂ G. Hence, q is a local homeomorphism.
(cid:3)
Observe that qG(0) : G(0) → (G/H)(0) is homeomorphic.
Proposition 2.1.8. Let G be an ´etale groupoid and H ⊂ G be an open
normal subgroupoid. Then, G/H is an ´etale groupoid.
Proof. First, we show the continuity of the inverse G/H (cid:51) δ (cid:55)→ δ−1 ∈
G/H. One can see that a map G (cid:51) γ (cid:55)→ q(γ)−1 ∈ G/H is continuous, since
the following diagram is commutative;
G
q
inverse
G
q
G/H
inverse
G/H.
Next, we show that the multiplication of G/H is continuous. Take (q(α), q(β)) ∈
By the definition of the quotient topology, the inverse of G/H is continuous.
(G/H)(2) and an open set U ⊂ G/H such that q(α)q(β) ∈ U . Since
αβ ∈ q−1(U ) and q−1(U ) ⊂ G is open, there exist open sets V1, V2 ⊂ G
such that α ∈ V1, β ∈ V2 and V1V2 ⊂ q−1(U ). Subsets V1, V2 ⊂ G are
open, so q(V1), q(V2) ⊂ G/H are open. One can see that q(α) ∈ q(V1),
q(β) ∈ q(V2) and q(V1)q(V2) = q(V1V2) ⊂ U . Therefore, the multiplication
of G/H is continuous.
Finally, we show that G/H is ´etale. Since the restriction qG(0) gives
a homeomorphism from G(0) to (G/H)(0), (G/H)(0) is a locally compact
Hausdorff space. One can see that the source map s : G/H → (G/H)(0)
is a local homeomorphism, since we have Lemma 2.1.7 and the following
diagram is commutative for every open bisection U ⊂ G;
U
s
s(U )
q
q
q(U )
s
s(q(U )).
(cid:3)
Therefore, G/H is an ´etale groupoid.
Definition 2.1.9. Let G1 and G2 be groupoids. A map Φ : G1 → G2
is called a groupoid homomorphism if (Φ(α), Φ(β)) ∈ G(2)
and Φ(αβ) =
Φ(α)Φ(β) hold for all (α, β) ∈ G(2)
1 .
2
Now, we obtain the next theorem by Lemma 2.1.7 and Proposition 2.1.8,
Theorem 2.1.10. Let G be an ´etale groupoid and H ⊂ G be an open
normal subgroupoid. Then, the sequence of the ´etale groupoids
THE ABELIANIZATIONS OF GROUPOID C*-ALGEBRAS
9
inclusion
q
G/H
H
G
is exact, that is, q−1((G/H)(0)) = H.
Proposition 2.1.11. Let G be an ´etale groupoid and H ⊂ G be an open
normal subgroupoid. Then, G/H is Hausdorff if and only if H ⊂ G is closed.
Proof. Recall that an ´etale groupoid G is Hausdorff if and only if its unit
space G(0) is a closed subset of G (for example, see [9, Lemma 2.3.2]). If
G/H is Hausdorff, (G/H)(0) ⊂ G/H is closed. Hence, H = q−1((G/H)(0))
is a closed subset of G.
Suppose that H ⊂ G is closed. Since q is an open map, (G/H) \
(G/H)(0) = q(G \ H) ⊂ G/H is open. Hence, (G/H)(0) ⊂ G/H is closed,
(cid:3)
which implies that G/H is Hausdorff.
Proposition 2.1.12. Let G be an ´etale groupoid. Then, the interior of
isotropy Iso(G)◦ ⊂ Iso(G) is a normal subgroupoid.
Proof. We show that Iso(G)◦ is normal. By Proposition 1.1.2, G(0) is
contained in Iso(G)◦. Take α ∈ G and γ ∈ Iso(G)◦ with s(α) = r(γ). There
exist open bisections U, V ⊂ G with α ∈ U and γ ∈ V ⊂ Iso(G). Then,
by Proposition 1.1.7, U V U−1 ⊂ G is an open subset which contains αγα−1.
Since U is bisection and V ⊂ Iso(G), we have U V U−1 ⊂ Iso(G). Therefore,
αγα−1 ∈ Iso(G)◦ and Iso(G)◦ is an open normal subgroupoid.
(cid:3)
An ´etale groupoid G/ Iso(G)◦, which is a special case of quotient groupoids,
coincides with a groupoid of germs of the canonical action (see [8, Section
3]). One can see that G/ Iso(G)◦ is effective.
2.2. *-homomorphisms induced by quotients of ´etale groupoids.
For an ´etale groupoid G and an open normal subgroupoid H ⊂ G, we have
obtained the quotient ´etale groupoid G/H. Next, we see that the quotient
map q : G → G/H induces a *-homomorphism C∗(G) → C∗(G/H).
For f ∈ C(G), we define f : G/H → C by
(cid:88)
f (γ) ··=
f (α)
q(α)=γ
for γ ∈ G/H. Then, the following proposition holds.
Proposition 2.2.1. Let G be an ´etale groupoid and H ⊂ G be an open
normal subgroupoid. Then, C(G) (cid:51) f (cid:55)→ f ∈ C(G/H) is a surjective *-
homomorphism.
Proof. First, we show f ∈ C(G/H). We may assume that there exists
an open bisection U ⊂ G such that fU ∈ Cc(U ) and fG\U = 0. Then,
q(U ) ⊂ G/H is an open bisection and fq(U ) = f ◦ (qU )−1 ∈ Cc(q(U )),
since qU is a homeomorphism onto the image. Moreover, one can see that
f(G/H)\q(U ) = 0. Hence, f ∈ Cc(q(U )) ⊂ C(G/H).
10
FUYUTA KOMURA
We show that C(G) (cid:51) f (cid:55)→ f ∈ C(G/H) is a *-homomorphism. We only
check that C(G) (cid:51) f (cid:55)→ f ∈ C(G/H) preserves the multiplications, since it
is easy to check that this map is linear and preserves the involutions. For
all f, g ∈ C(G) and γ(cid:48) ∈ G/H, we have
(cid:88)
(cid:88)
(cid:88)
q(γ)=γ(cid:48)
α(cid:48)β(cid:48)=γ(cid:48)
(cid:93)f ∗ g(γ(cid:48)) =
f ∗ g(γ(cid:48)) =
=
(cid:88)
(cid:88)
(cid:88)
αβ=γ
q(γ)=γ(cid:48)
(cid:88)
f ∗ g(γ) =
f (α(cid:48))g(β(cid:48)) =
(cid:88)
(cid:88)
q(αβ)=γ(cid:48)
f (α)g(β)
α(cid:48)β(cid:48)=γ(cid:48)
q(α)=α(cid:48)
q(β)=β(cid:48)
f (α)g(β) =
f (α)g(β),
f (α)g(β).
q(αβ)=γ(cid:48)
Hence, C(G) (cid:51) f (cid:55)→ f ∈ C(G/H) is a surjective *-homomorphism.
Finally, we show that C(G) (cid:51) f (cid:55)→ f ∈ C(G/H) is surjective. Note
that {q(U ) ⊂ G/H U ⊂ G is an open bisection} is an open basis of G.
Let U ⊂ G be an open bisection and f ∈ Cc(q(U )). One can see that
Then, we have (cid:101)g = f . By Proposition 1.2.4, C(G/H) is the linear span
qU is a homeomorphism onto its image. Define g ··= f ◦ qU ∈ Cc(U ).
of (cid:83)
U Cc(q(U )), where the union is taken over all open bisections U ⊂ G.
(cid:3)
By the universality of C∗(G), the surjective *-homomorphism C(G) (cid:51)
f (cid:55)→ f ∈ C(G/H) induces the surjective *-homomorphism Q : C∗(G) →
C∗(G/H).
Proposition 2.2.2. Let Q : C∗(G) → C∗(G/H) be the *-homomorphism
as above. Then, ker Q ∩ C0(G(0)) = {0} holds.
Proof. Since the universal norm of a function in Cc(G(0)) coincides with the
supremum norm, QCc(G(0)) is isometric. Therefore, QC0(G(0)) is isometric
and ker Q ∩ C0(G(0)) = {0}.
(cid:3)
Definition 2.2.3. Let G be an ´etale groupoid. We say that G has the full
uniqueness property if every non-zero ideal I ⊂ C∗(G) satisfy I∩C0(G(0)) (cid:54)=
{0}.
The full uniqueness property of G is equivalent to the condition that a
*-representation π : C∗(G) → B(H) is injective if and only if πC0(G(0)) is
injective.
Lemma 2.2.4. Let G be an ´etale groupoid and H ⊂ G be an open normal
subgroupoid. Then, the *-homomorphism Q : C∗(G) → C∗(G/H) induced
by Proposition 2.2.1 is injective if and only if H = G(0).
Proof. It is clear that the *-homomorphism Q : C∗(G) → C∗(G/H) is
injective if H = G(0). Suppose that G(0) (cid:40) H and take γ0 ∈ H \ G(0). There
exists an open bisection U ⊂ G with γ0 ∈ U ⊂ H. By the Urysohn lemma,
THE ABELIANIZATIONS OF GROUPOID C*-ALGEBRAS
11
there exists f1 ∈ Cc(U ) with f1(γ0) = 1. Define f2 ∈ Cc(G(0)) by
f2(γ) =
f1 ◦ (sU )−1(γ)
0
(γ ∈ s(U ))
(γ ∈ G(0) \ s(U )).
(cid:40)
We have f ··= f1 − f2 (cid:54)= 0, since f (γ0) = 1. One can see that Q(f ) = 0,
(cid:3)
which means that Q is not injective.
Recall that an ´etale groupoid G is said to be effective if G(0) = Iso(G)◦.
Corollary 2.2.5 (cf. [1, Proposition 5.5]). Let G be an ´etale groupoid. If
G has the full uniqueness property, then G is effective.
Proof. By Proposition 2.1.12, Iso(G)◦ is a normal subgroupoid of G. Let-
ting Q : C∗(G) → C∗(G/ Iso(G)◦) be the *-homomorphism induced by Propo-
sition 2.2.1, we have ker Q ∩ C0(G(0)) = {0}. The full uniqueness property
implies that Q : C∗(G) → C∗(G/ Iso(G)◦) is injective. Therefore, we obtain
Iso(G)◦ = G(0) by Lemma 2.2.4.
(cid:3)
Remark 2.2.6. It had been known that Corollary 2.2.5 holds for Hausdorff
´etale groupoids as proved in [1, Proposition 5.5].
In [1, Proposition 5.5],
the authors use the augmentation representations, which seems to work for
non-Hausdorff ´etale groupoids.
As shown in Proposition 2.2.1, the quotient map G → G/ Iso(G)◦ of ´etale
groupoids induces the *-homomorphism C∗(G) → C∗(G/ Iso(G)◦). Using
this *-homomorphism, we obtain the proof of Corollary 2.2.5, which seems
to be more direct than the one in [1, Proposition 5.5]. We shall remark that
G/ Iso(G)◦ coincides with the groupoid of germs of the canonical action in
[8, Proposition 3.2].
The converse of Corollary 2.2.5 does not hold for non-Hausdorff ´etale
groupoids. Indeed, Exel showed that there exists an effective ´etale groupoid
G which does not have the full uniqueness property in [4] (cf. Example 2.2.7).
Example 2.2.7 ([4, Section 2]). Let X ··= ([−1, 1]×{0})∪ ({0}× [−1, 1]) ⊂
R2 and K ··= {e, s, t, st} be the Klein group, which is isomorphic to Z/2Z ⊕
Z/2Z. We define an action σ of K on X by
fot (x, y) ∈ X.
σs((x, y)) = (−x, y), σt((x, y)) = (x,−y), σst((x, y)) = (−x,−y)
Consider the transformation groupoid G ··= K (cid:110)σ X (see Example 1.1.6).
One can see that
Iso(G) = G(0) ∪ {(s, (0, y)) ∈ G y ∈ [−1, 1]}
∪ {(t, (x, 0)) ∈ G x ∈ [−1, 1]} ∪ {(st, (0, 0))}.
Moreover, we have Iso(G)◦ = Iso(G)\{(s, (0, 0)), (t, (0, 0)), (st, (0, 0))}. Since
Iso(G)◦ is not closed in G (for example, (s, (0, 0)) ∈ Iso(G)◦ \ Iso(G)◦), the
quotient ´etale groupoid G/ Iso(G)◦ is not Hausdorff by Proposition 2.1.11.
In [4], Exel shows that G/ Iso(G)◦ does not have the full uniqueness property,
although it is effective.
12
FUYUTA KOMURA
(0,1)
(−1,0)
(1,0)
(0,−1)
Figure 1. Picture of X in Example 2.2.7
Based on the work in [2], we shall give a necessary and sufficient condition
of the full uniqueness property. Let G be an ´etale groupoid. We denote
the left representation at x ∈ G(0) by λx : C(G) → B((cid:96)2(Gx)). By the
universality of C∗(G), the left representation extends to the *-representation
λx : C∗(G) → B((cid:96)2(Gx)). Following [2], we say that an element a ∈ C∗(G)
is singular if the interior of {γ ∈ G (cid:104)δγλs(γ)(a)δs(γ)(cid:105) (cid:54)= 0} is empty, where
δγ ∈ (cid:96)2(Gx) denotes the delta function at γ ∈ Gx.
We denote the reduced groupoid C*-algebra of G by C∗
λ(G). We denote
the canonical surjective *-homomorphism by Q : C∗(G) → C∗
λ(G). In [2],
the notion of a singular element is defined for elements in C∗
λ(G) in the same
way as elements in C∗(G). Since an element in ker Q is singular, C∗(G) has
no nonzero singular elements if and only if C∗
λ(G) has no nonzero singular
elements and C∗(G) (cid:39) C∗
λ(G) via the canonical *-homomorphism Q.
Proposition 2.2.8. Let G be a second countable ´etale groupoid. Then, G
has the full uniqueness property if and only if
• G is effective and
• C∗(G) has no nonzero singular elements.
3.6]. Therefore, letting π ··=(cid:76)
Proof. Assume that G has the full uniqueness property. Corollary 2.2.5
implies that G is effective. We show that C∗(G) has no nonzero singular
elements. Let a ∈ C∗(G) be a singular element. We define S ··= {x ∈
G(0) Gx ∩ Gx = {x}}, which is a dense subset of G(0) by [8, Proposition
x∈S λx, π is injective on C0(G(0)). The full
uniqueness property implies that π is injective. Since λx(a) = λx(Q(a))
holds for all x ∈ G(0), we have
s({γ ∈ G (cid:104)δγλs(γ)(a)δs(γ)(cid:105) (cid:54)= 0}) ⊂ G(0) \ S
by [2, Lemma 4.2]. Using this fact, one can see that π(a) = 0 if a ∈ C∗(G)
is singular. Hence, C∗(G) has no nonzero singular element.
Next, we show the converse. Note that Q is an isomorphism by the
assumpution that C∗(G) has no nonzero singular elements. Now, the full
(cid:3)
uniqueness property of G follows from [2, Theorem 4.4].
THE ABELIANIZATIONS OF GROUPOID C*-ALGEBRAS
13
3. The abelianizations of ´etale groupoid C*-algebras
In this section, we calculate the abelianizations of ´etale groupoid C*-
algebras. First, recall the abelianizations of C*-algebras. For a C*-algebra
A, its abelianization is defined by Aab = A/I, where I ⊂ A is the closed
two-sided ideal generated by {xy − yx ∈ A x, y ∈ A}. The abelianization
Aab is a commutative C*-algebra with the following universality; for all
commutative C*-algebra B and *-homomorphism π : A → B, there exists
the unique *-homomorphism π : Aab → B such that π◦q = π, where q : A →
Aab denotes the quotient map.
3.1. One dimensional representations of a groupoid C*-algebra.
For a C*-algebra A, we denote the set of all one dimensional nondegen-
erate representations of A by ∆(A). Namely, ∆(A) is the set of all nonzero
*-homomorphisms from A to C. We suppose that ∆(A) is equipped with
the pointwise convergence topology. If A is commutative, ∆(A) is known as
the Gelfand spectrum of A. First, we calculate ∆(C∗(G)).
Let G be an ´etale groupoid and x ∈ G(0) be a fixed point of G. Note
that Gx is a discrete group. We denote the surjection in Proposition 1.2.4
by Qx : C∗(G) → C∗(Gx) temporarily. Also, we denote the circle group
by T ··= {z ∈ C z = 1}. For a group homomorphism χ : Gx → T, a
χ(γ)f (γ) ∈ C is a *-homomorphism. This *-
homomorphism extends to the *-homomorphism C∗(Gx) → C, which we
also denote by χ : C∗(Gx) → C.
Definition 3.1.1. Let G be an ´etale groupoid, x ∈ G(0) be a fixed point and
χ : Gx → T be a group homomorphism. Then, we define a *-homomorphism
ϕx,χ : C∗(G) → C by ϕx,χ ··= χ ◦ Qx.
map Cc(Gx) (cid:51) f (cid:55)→ (cid:80)
γ∈Gx
We will show that all elements of ∆(C∗(G)) are obtained by this form
(Theorem 3.1.8).
Proposition 3.1.2. Let G be an ´etale groupoid and ϕ ∈ ∆(C∗(G)). Then,
there uniquely exists xϕ ∈ G(0) which satisfies ϕ(f ) = f (xϕ) for all f ∈
C0(G(0)).
Proof. We have ϕC0(G0) (cid:54)= 0 since C0(G(0)) has an approximate identity
of C∗(G). Therefore, ϕC0(G(0)) belongs to ∆(C0(G(0))). Now, the existence
and uniqueness of xϕ ∈ G(0) follow from the Gelfand-Naimark theorem. (cid:3)
Proposition 3.1.3. Let G be an ´etale groupoid and ϕ ∈ ∆(C∗(G)). Then,
xϕ ∈ G(0) is a fixed point.
Proof. Assume that γ ∈ G satisfies s(γ) = xϕ. We will show r(γ) = xϕ.
There exists an open bisection U ⊂ G with γ ∈ U . Take nγ ∈ Cc(U )
γ ∗ nγ ∈ Cc(G(0)) and
which satisfies nγ(γ) = 1. Note that we have n∗
γ ∗ nγ(xϕ) = nγ(γ)2 = 1. Fix f ∈ Cc(G(0)) arbitrarily.
n∗
14
FUYUTA KOMURA
γ∗f ∗nγ) = ϕ(n∗
γ)ϕ(f )ϕ(nγ) = ϕ(n∗
Direct calculations show that n∗
γ ∗ f ∗ nγ(xϕ) = nγ(γ)f (r(γ))nγ(γ) =
γ ∗ f ∗ nγ ∈ Cc(G(0)). Then,
f (r(γ)). On the other hand, one can see that n∗
we have
γ∗nγ)ϕ(f ) = f (xϕ).
γ∗f ∗nγ(xϕ) = ϕ(n∗
n∗
Therefore, f (r(γ)) = f (xϕ) holds for all f ∈ Cc(G(0)), which implies
r(γ) = xϕ. Hence, xϕ ∈ G(0) is a fixed point of G.
(cid:3)
Proposition 3.1.4. Let G be an ´etale groupoid, ϕ ∈ ∆(C∗(G)) and γ ∈
Gxϕ. Take an open bisection Uγ ⊂ G with γ ∈ Uγ and fγ ∈ Cc(Uγ) with
fγ(γ) = 1. Then, ϕ(fγ) is independent of the choice of Uγ and fγ. Moreover,
we have ϕ(fγ) ∈ T.
Proof. First, we show ϕ(fγ) ∈ T. Since f∗
γ ∗ fγ ∈ C0(G(0)), we have
ϕ(fγ)2 = ϕ(f∗
γ ∗ fγ) = f∗
γ ∗ fγ(xϕ) = fγ(γ)2 = 1.
Therefore, ϕ(fγ) ∈ T.
Second, we show that ϕ(fγ) is independent of the choice of Uγ and fγ.
Assume that fγ ∈ Cc(Uγ) and gγ ∈ Cc(Vγ) satisfies fγ(γ) = gγ(γ) = 1, where
Uγ and Vγ ⊂ G are open bisections. Find a function h ∈ Cc(s(Uγ ∩ Vγ)) ⊂
Cc(G(0)) such that h(s(γ)) = 1. Recall that s(γ) = r(γ) = xϕ since xϕ is
a fixed point. Also, note that ϕ(h) = h(xϕ) = 1. Putting (cid:101)fγ ··= fγ ∗ h and
(cid:101)gγ = gγ ∗ h, we have that (cid:101)fγ and (cid:101)gγ are contained in Cc(Uγ ∩ Vγ). Then, it
follows that (cid:101)fγ
∗ ∗ (cid:101)gγ ∈ C0(G(0)) and
ϕ(fγ)ϕ(gγ) = ϕ(h)ϕ(fγ)ϕ(gγ)ϕ(h) = ϕ((cid:101)fγ
∗ ∗ (cid:101)gγ)
∗ ∗ (cid:101)gγ(xϕ) = h(r(γ))fγ(γ)gγ(γ)h(s(γ)) = 1.
Now, we have ϕ(fγ) = ϕ(gγ) since ϕ(fγ) ∈ T.
(cid:3)
Proposition 3.1.5. Let G be an ´etale groupoid and ϕ ∈ ∆(C∗(G)). We
define χϕ : Gxϕ → T by χϕ(γ) ··= ϕ(fγ), where γ ∈ Gxϕ and fγ ∈ C(G) is a
function as in Proposition 3.1.4. Then, χϕ : Gxϕ → T is a group homomor-
phism.
Proof. Take α, β ∈ Gxϕ. We show χϕ(α)χϕ(β) = χϕ(αβ). Take fα, fβ ∈
C(G) as in Proposition 3.1.4. It follows that fα ∗ fβ ∈ Cc(U ) for some open
bisection U ⊂ G and fα ∗ fβ(αβ) = 1. Hence, we have
= (cid:101)fγ
χϕ(αβ) = ϕ(fα ∗ fβ) = ϕ(fα)ϕ(fβ) = χϕ(α)χϕ(β)
by the definition of χϕ.
Proposition 3.1.6. Let G be an ´etale groupoid. Then, we have ϕ = ϕxϕ,χϕ
for all ϕ ∈ ∆(C∗(G)).
Proof. Take f ∈ Cc(U ), where U ⊂ G is an open bisection.
It suffices
to show ϕ(f ) = ϕxϕ,χϕ(f ), since C∗(G) is generated by such functions.
If Gxϕ ∩ f−1(C \ {0}) = ∅, then we have
Note that f∗ ∗ f ∈ Cc(G(0)).
(cid:3)
THE ABELIANIZATIONS OF GROUPOID C*-ALGEBRAS
15
0 = f∗ ∗ f (xϕ) = ϕ(f )2. Since the restriction of fGxϕ is zero, it follows
that ϕxϕ,χϕ(f ) = 0 = ϕ(f ). If Gxϕ ∩ f−1(C \ {0}) (cid:54)= ∅, Gxϕ ∩ f−1(C \ {0})
is a singleton because f is supported on an open bisection. Let γ ∈ Gxϕ ∩
f−1(C \ {0}) be the unique element of Gxϕ ∩ f−1(C \ {0}). Observe that
F ··= f /f (γ) ∈ Cc(U ) satisfies F (γ) = 1. Now, we have
ϕxϕ,χϕ(f ) = f (γ)χϕ(γ) = f (γ)ϕ(F ) = ϕ(f ).
(cid:3)
Hence, we have ϕxϕ,χϕ = ϕ.
Proposition 3.1.7. Let G be an ´etale groupoid, x ∈ G(0) be a fixed point
and χ : Gx → T be a group homomorphism. Then, x = xϕx,χ and χ = χϕx,χ.
Proof. First, we show x = xϕx,χ. Take f ∈ Cc(G(0)) arbitrarily. Then, we
have
f (xϕx,χ) = ϕx,χ(f ) = f (x)χ(x) = f (x).
Hence, it follows x = xϕx.
Next, we show χ = χϕx,χ. Take γ ∈ Gx arbitrarily. There exist an open
bisection U ⊂ G with γ ∈ U and f ∈ Cc(U ) with f (γ) = 1. Then, we have
χϕx,χ(γ) = ϕx,χ(f ) = f (γ)χ(γ) = χ(γ).
Hence, we have shown x = xϕx,χ and χ = χϕx,χ.
(cid:3)
Combining Proposition 3.1.6 and Proposition 3.1.7, we obtain the next
theorem.
Theorem 3.1.8. Let G be an ´etale groupoid. Define a set
D ··= {(x, χ) x ∈ G(0) is a fixed point
and χ : Gx → T is a group homomorphism}.
Then, a map
is bijective.
D (cid:51) (x, χ) −→ ϕx,χ ∈ ∆(C∗(G))
tator subgroupoid of G by [G, G] ··=(cid:83)
3.2. Construction of an ´etale abelian group bundle Gab. For an
´etale groupoid G, we construct an ´etale abelian group bundle Gab so that
C∗(G)ab (cid:39) C∗(Gab) holds.
Proposition 3.2.1. Let G be an ´etale group bundle. We define the commu-
x∈G(0)[Gx, Gx], where [Gx, Gx] is the
commutator subgroup of Gx. Then, [G, G] is an open normal subgroupoid
of G.
Proof. It is obvious that [G, G] ⊂ G is a normal subgroupoid. We show
that [G, G] ⊂ G is open. Take γ ∈ [G, G]. By the definition of the commu-
tator subgroup, there exists {αj}k
j=1 ⊂ Gs(γ) such that
γ = α1β1α−1
1 β−1
··· αkβkα−1
k β−1
k .
j=1,{βj}k
2 β−1
1 α2β2α−1
2
16
FUYUTA KOMURA
1 V −1
1 V −1
1
Take open bisections Uj, Vj ⊂ G such that αj ∈ Uj and βj ∈ Vj for all
j = 1, 2, . . . , k. We show that U1V1U−1
1 ⊂ [G, G], where we define
U−1 ··= {γ−1 γ ∈ U} for U ⊂ G. Fix γ(cid:48) ∈ U1V1U−1
. Then, there exist
α, α(cid:48) ∈ U1 and β, β(cid:48) ∈ V1 which satisfy γ = αβα(cid:48)−1β(cid:48)−1. Since G is a group
bundle, we have s(α) = s(α(cid:48)) = s(β) = s(β(cid:48)). We obtain α = α(cid:48) and β = β(cid:48)
because U1 and V1 are bijections. Therefore, γ(cid:48) = αβα−1β−1 ∈ [G, G].
Similarly, one can show that U1V1U−1
k ⊂
[G, G].
1 V −1
is an
open set and this contains γ. Hence, [G, G] ⊂ G is an open normal sub-
(cid:3)
groupoid.
By Proposition 1.1.7, U1V1U−1
··· UkVkU−1
k V −1
··· UkVkU−1
1 V −1
1 U2V2U−1
1 U2V2U−1
2 V −1
k V −1
2 V −1
2
2
k
Let G be an ´etale groupoid. Recall that the set of all fixed points F ⊂ G(0)
is a closed subset of G(0) (Proposition 1.1.9). We define Gfix ··= GF , which
is an ´etale groupoid from Proposition 1.1.10. Since we have Gfix = Iso(Gfix),
Gfix is an ´etale group bundle.
Definition 3.2.2. Let G be an ´etale groupoid. We define the abelianization
of G by Gab ··= Gfix/[Gfix, Gfix].
Let G be an ´etale groupoid. Then, we have a *-homomorphism C∗(G) →
C∗(Gfix) induced by the restriction (Proposition 1.2.4). Composing with
the *-homomorphism C∗(Gfix) → C∗(Gab) in Proposition 2.2.1, we obtain a
*-homomorphism π : C∗(G) → C∗(Gab).
Note that C∗(G) is commutative if and only if G is an ´etale abelian group
bundle. In particular, C∗(Gab) is commutative.
Lemma 3.2.3. Let G be an ´etale groupoid. Then, a map Φ : ∆(C∗(Gab)) (cid:51)
χ (cid:55)→ χ ◦ π ∈ ∆(C∗(G)) is bijective.
Proof. Surjectivity of π implies that Φ is injective. We show that Φ is
surjective. Take ϕ ∈ ∆(C∗(G)). Then, we have the fixed point xϕ ∈ G(0)
and the group homomorphism χϕ which makes the following diagram com-
mutative;
C∗(G)
ϕ
C
q
C∗(Gxϕ),
χϕ
where q : C∗(G) → C∗(Gxϕ) is the *-homomorphism obtained in Proposition
1.2.4.
··= (Gxϕ)ab = (Gab)xϕ, we obtain the group
xϕ → T which makes the following diagram commu-
By the universality of Gab
xϕ
homomorphism ¯χϕ : Gab
tative;
THE ABELIANIZATIONS OF GROUPOID C*-ALGEBRAS
17
C∗(Gxϕ)
χϕ
C
q(cid:48)
C∗(Gab
xϕ),
¯χϕ
where q(cid:48) : C∗(Gxϕ) → C∗(Gab
xϕ) denotes the *-homomorphism induced by
the quotient map Gxϕ → Gab
xϕ.
Let res : C∗(Gab) → C∗(Gab
the restriction C(Gab) → C(Gab
following commutative diagram;
xϕ) denote the *-homomorphism obtained by
xϕ) (see Proposition 1.2.4). Now, we have the
ϕ
C∗(G)
q
C∗(Gxϕ)
χϕ
C
π
q(cid:48)
¯χϕ
C∗(Gab)
res
C∗(Gab
xϕ).
In particular, we have ϕ = ( ¯χϕ ◦ res)◦ π and ¯χϕ ◦ res ∈ ∆(C∗(Gab)). Hence,
(cid:3)
Φ is surjective.
Now, we are ready to calculate the abelianization of C∗(G).
Theorem 3.2.4. Let G be an ´etale groupoid. Then, C∗(G)ab is isomorphic
to C∗(Gab) via the unique isomorphism ¯π which makes the following diagram
commutative;
C∗(G)
π
C∗(Gab)
Q
C∗(G)ab,
¯π
where Q : C∗(G) → C∗(G)ab denotes the quotient map.
Proof. By the universality of C∗(G)ab, we obtain a *-homomorphism which
makes the following diagram commutative;
C∗(G)
π
C∗(Gab)
Q
C∗(G)ab.
¯π
It is clear that ¯π is surjective. We show that ¯π is injective. Suppose that a ∈
C∗(G) satisfies π(a) = 0. It suffices to show Q(a) = 0, which is equivalent
to ¯ϕ(Q(a)) = 0 for all ¯ϕ ∈ ∆(C∗(G)ab) since C∗(G)ab is commutative. Take
18
FUYUTA KOMURA
¯ϕ ∈ ∆(C∗(G)ab) and define ϕ ··= ¯ϕ◦ Q. Then, by Lemma 3.2.3, there exists
ϕ ∈ ∆(C∗(Gab)) which makes the following diagram commutative;
C∗(G)
π
C∗(Gab)
ϕ
ϕ
C.
Now, we have the following commutative diagram;
C∗(Gab)
C∗(G)
π
Q
C∗(G)ab
ϕ
¯ϕ
ϕ
C.
Hence, we have ¯ϕ(Q(a)) = ϕ(π(a)) = 0.
(cid:3)
3.3. Duals of ´etale abelian group bundles. Let G be an ´etale groupoid.
Since the abelianization of C∗(G) is a commutative C*-algebra, C∗(G)ab
is isomorphic to C0(∆(C∗(G)ab)) via the Gelfand transformation. In this
section, we calculate the Gelfand spectrum ∆(C∗(G)ab).
the set of all group homomorphisms from Γ to T, which is denoted by (cid:98)Γ.
Then,(cid:98)Γ is an abelian group with respect to the pointwise multiplication. It
is known that(cid:98)Γ is a compact abelian topological group with respect to the
For a discrete abelian group Γ, its Pontryagin dual group is defined as
topology of pointwise convergence.
Proposition 3.3.1. Let Γ be a discrete group and Q : C∗(Γ) → C∗(Γab) be
the *-homomorphism induced by the quotient map Γ → Γab. Then, a map
Φ :(cid:100)Γab (cid:51) χ (cid:55)→ χ ◦ Q ∈ ∆(C∗(Γ))
is a homeomorphism. Hence, C∗(Γ)ab is isomorphic to C((cid:100)Γab).
Proof. This follows from the universality of Γab and C∗(Γ).
(cid:3)
As seen in the previous proposition, the key to calculate ∆(C∗(G)) is the
Pontryagin dual.
Definition 3.3.2. Let G be an ´etale abelian group bundle. We define a
group bundle (cid:98)G ··= {(χ, x) x ∈ G(0), χ ∈ (cid:99)Gx} over G(0).
Note that (cid:98)G is a group bundle such that (cid:98)Gx = (cid:99)Gx × {x}((cid:39) (cid:99)Gx) for every
Let G be an ´etale abelian group bundle and (χ, x) ∈ (cid:98)G. Recall that we
C(G), we define evf : (cid:98)G → C by evf ((χ, x)) = ϕx,χ(f ), where (χ, x) ∈ (cid:98)G. We
x ∈ G(0).
obtain the *-homomorphism ϕx,χ ∈ ∆(C∗(G)) as in Definition 3.1.1.
Definition 3.3.3. Let G be an ´etale abelian group bundle. For each f ∈
19
THE ABELIANIZATIONS OF GROUPOID C*-ALGEBRAS
define a topology of (cid:98)G as the weakest topology in which evf is continuous
Ψ : ∆(C∗(G)) (cid:51) ϕ (cid:55)→ (χϕ, xϕ) ∈ (cid:98)G
for all f ∈ C(G).
Proposition 3.3.4. Let G be an ´etale abelian group bundle. Then, the
map1)
is a homeomorphism. Hence, C∗(G) is isomorphic to C0((cid:98)G)
by Ψ−1((χ, x)) = ϕx,χ for each (χ, x) ∈ (cid:98)G. For each f ∈ C(G), a map
Proof. Proposition 3.1.8 states that Ψ is a bijection and Ψ−1 is given
∆(C∗(G)) (cid:51) ϕ (cid:55)→ evf ((χϕ, xϕ)) = ϕ(f ) ∈ C is continuous. This means
that Ψ is continuous. The continuity of Ψ−1 follows from approximation
(cid:3)
arguments. Therefore, Ψ is a homeomorphism.
compact if and only if G(0) is compact.
Let G be an ´etale groupoid. Recall that Gab is an ´etale abelian group bundle.
Corollary 3.3.5. Let G be an ´etale groupoid. Then, C∗(G)ab is isomorphic
Proof. Recall that C∗(G)ab is isomorphic to C∗(Gab) by Theorem 3.2.4.
Since Gab is an ´etale abelian group bundle, Proposition 3.3.4 implies that
(cid:3)
to C0((cid:100)Gab).
C∗(Gab) is isomorphic to C0((cid:100)Gab).
Proposition 3.3.6. Let G be an ´etale abelian group bundle. Then, (cid:98)G is
a locally compact Hausdorff topological group bundle. Furthermore, (cid:98)G is
Proof. It is clear that (cid:98)G is locally compact Hausdorff, since (cid:98)G is homeo-
take f ∈ C(G) arbitrarily. Then, the map (cid:98)G(2) (cid:51) (χ1, χ2) (cid:55)→ evf (χ1χ2) =
evf (χ1) evf (χ2) ∈ C is continuous. Therefore, the multiplication of (cid:98)G(2) is
continuous. Similarly, one can show that the inverse is continuous. Hence, (cid:98)G
follows from the fact that G(0) is compact if and only if C∗(G) (cid:39) C0((cid:98)G) is
morphic to ∆(C∗(G)). In order to show the continuity of the operations,
is a locally compact Hausdorff topological group bundle. The last assertion
(cid:3)
unital.
Example 3.3.7. Let S3 = (cid:104)s, t s3 = t2 = e, st = ts2(cid:105) = {e, s, s2, t, ts, ts2}
be the symmetric group of degree 3 and A3 ··= {e, s, s2} ⊂ S3 be the sub-
group of even permutations. Let G ··= S3 × [0, 1] \ {(t, 1) t (cid:54)∈ A3} be an
´etale group bundle over [0, 1]. Then, G can be drawn as follows;
(ts2,0)
(s2,0)
(ts,0)
(s,0)
(t,0)
(e,0)
(s2,1)
(s,1)
(e,1)
1)See Proposition 3.1.2 and 3.1.5 for the definition of xϕ and χϕ.
20
FUYUTA KOMURA
One can see that [G, G] ⊂ G is not closed. By Proposition 2.1.11, Gab =
G/[G, G] is not Hausdorff. Indeed, letting q : G → Gab denote the quotient
map, Gab looks like the following;
q((t,0))
The dual (cid:100)Gab of Gab can be drawn as follows;
q((e,0))
q((e,1)),q((s,1)),q((s2,1))
Note that (cid:100)Gab is not ´etale.
References
[1] J. Brown, L. O. Clark, C. Farthing, and A. Sims. Simplicity of algebras associated to
´etale groupoids. Semigroup Forum, 88(2):433 -- 452, 2014.
[2] L. O. Clark, R. Exel, E. Pardo, A. Sims, and C Starling. Simplicity of algebras asso-
ciated to non-Hausdorff groupoids. arXiv:1806.04362, 2018.
[3] A. Connes. A survey of foliations and operator algebras. In Proc. Sympos. Pure, volume
38, pages 521 -- 628, 1982.
[4] R. Exel. Non-Hausdorff ´etale groupoids. Proceedings of the American Mathematical
Society, 139(3):897 -- 907, 2011.
[5] R. Exel and E. Pardo. The tight groupoid of an inverse semigroup. Semigroup Forum,
92(1):274 -- 303, 2016.
[6] A. Paterson. Groupoids, Inverse Semigroups, and their Operator Algebras. Progress in
Mathematics. Birkhauser Boston, 2012.
[7] J. Renault. A Groupoid Approach to C*-Algebras. Lecture Notes in Mathematics.
Springer-Verlag, 1980.
[8] J. Renault. Cartan subalgebras in C*-algebras. Irish Math. Soc. Bull., 61:29 -- 63, 2008.
[9] A. Sims. Hausdorff ´etale groupoids and their C*-algebras. arXiv:1710.10897v1, 2017.
Department of Mathematics, Faculty of Science and Technology, Keio Uni-
versity 3141 Hiyoshi, Kohoku-ku, Yokohama, 2238522, Japan
E-mail address: [email protected]
|
0905.0478 | 3 | 0905 | 2010-04-02T14:52:37 | Leavitt path algebras with coefficients in a commutative ring | [
"math.OA",
"math.RA"
] | Given a directed graph E we describe a method for constructing a Leavitt path algebra $L_R(E)$ whose coefficients are in a commutative unital ring R. We prove versions of the Graded Uniqueness Theorem and Cuntz-Krieger Uniqueness Theorem for these Leavitt path algebras, giving proofs that both generalize and simplify the classical results for Leavitt path algebras over fields. We also analyze the ideal structure of $L_R(E)$, and we prove that if $K$ is a field, then $L_K(E) \cong K \otimes_\Z L_\Z(E)$. | math.OA | math |
LEAVITT PATH ALGEBRAS WITH COEFFICIENTS IN A
COMMUTATIVE RING
MARK TOMFORDE
Abstract. Given a directed graph E we describe a method for con-
structing a Leavitt path algebra LR(E) whose coefficients are in a com-
mutative unital ring R. We prove versions of the Graded Unique-
ness Theorem and Cuntz-Krieger Uniqueness Theorem for these Leav-
itt path algebras, giving proofs that both generalize and simplify the
classical results for Leavitt path algebras over fields. We also analyze
the ideal structure of LR(E), and we prove that if K is a field, then
LK (E) ∼= K ⊗Z LZ(E).
1. Introduction
In [1] the authors introduced a class of algebras over fields, which they
constructed from directed graphs and called Leavitt path algebras.
(The
definition in [1] was given for row-finite directed graphs, but the authors
later extended the definition in [2] to all directed graphs.) These Leavitt
path algebras generalize the Leavitt algebras L(1, n) of [21], and also contain
many other interesting classes of algebras over fields. In addition, Leavitt
path algebras are intimately related to graph C ∗-algebras (see [23]), and
for any graph E the Leavitt path algebra LC(E) is ∗-isomorphic to a dense
∗-subalgebra of the graph C ∗-algebra C ∗(E) [26, Theorem 7.3].
In this paper we generalize the construction of Leavitt path algebras by
replacing the field K with a commutative unital ring R. We use the no-
tation LR(E) for our Leavitt path algebra, and prove that it is a Z-graded
R-algebra with characteristic equal to the characteristic of R. We also prove
versions of the Graded Uniqueness Theorem and the Cuntz-Krieger Unique-
ness Theorem, which are fundamental to the study of Leavitt path algebras.
The Graded Uniqueness Theorem for Leavitt path algebras over a field
says that a graded homomorphism φ : LK (E) → A is injective if φ(v) 6= 0 for
all v ∈ E0. For Leavitt path algebras over rings we need slightly different
hypotheses: We prove that a graded homomorphism φ : LR(E) → A is
injective if φ(rv) 6= 0 for all v ∈ E0 and for all r ∈ R \ {0}. Similarly,
the Cuntz-Krieger Uniqueness Theorem for Leavitt path algebras over a
field says that if every cycle in E has an exit, then a homomorphism φ :
Date: May 4, 2009.
2000 Mathematics Subject Classification. 16W50, 46L55.
Key words and phrases. graph algebras, rings, R-algebras, C ∗-algebras.
The author was supported by NSA Grant H98230-09-1-0036.
1
2
MARK TOMFORDE
LK(E) → A is injective if φ(v) 6= 0 for all v ∈ E0. Again, our hypotheses
for Leavitt path algebras over rings are slightly different: We prove that
if every cycle in E has an exit, then a homomorphism φ : LK(E) → A is
injective if φ(rv) 6= 0 for all v ∈ E0 and for all r ∈ R \ {0}.
Our proofs of the Uniqueness Theorems use techniques that are different
from those that have been used in the proofs for Leavitt path algebras over
fields. Consequently, this paper gives new proofs of each of the Uniqueness
Theorems in the case that R = K is a field. One of the main points of this
article is that our proofs of the Uniqueness Theorems are shorter than those
in the existing literature.
After proving our Uniqueness Theorems we continue by analyzing the
ideal structure of LR(E). For ease and clarity as we analyze ideals, we re-
strict our attention to the case when the graph E is row-finite. Because of
the hypothesis φ(rv) 6= 0 for all v ∈ E0 and for all r ∈ R \ {0}, the Unique-
ness Theorems only allow us to analyze what we call basic ideals: an ideal
I of LR(E) is basic if rv ∈ I for r ∈ R \ {0} implies that v ∈ I. In analogy
with Leavitt path algebras over fields, we prove in Theorem 7.9 that the
map H 7→ IH is a lattice isomorphism from the saturated hereditary subsets
of E onto the graded basic ideals of LR(E). We also prove in Theorem 7.17
that all basic ideals in LR(E) are graded if and only if E satisfies Condi-
tion (K). Finally, in Theorem 7.20 and Proposition 7.22 we derive conditions
for LR(E) to have no nontrivial proper basic ideals. These results are sim-
ilar to the classification of gauge-invariant ideals of graph C ∗-algebras and
Cuntz-Krieger C ∗-algebras, and we use similar techniques in this paper. We
refer the reader to Remark 7.23 for references to the corresponding results
for Cuntz-Krieger algebras, graph C ∗-algebras, and Leavitt path algebras
over fields.
In the final section, we discuss extending the coefficients of a Leavitt
path algebra by tensoring with a commutative unital ring. In particular, we
show that if K is a field, then LK(E) ∼= K ⊗Z LZ(E); and if K is a field
of characteristic p, then LK(E) ∼= K ⊗Zp LZp(E). This allows us to relate
properties of LZ(E) and LZp(E) to properties of LK(E).
This paper is organized as follows: After some preliminaries in §2, we
continue in §3 by constructing the Leavitt path algebra over a commutative
until ring, and prove that LR(E) exists and has the appropriate universal
property. In §4 we establish some basic properties of LR(E). In §5 we prove
the Graded Uniqueness Theorem for LR(E), and in §6 we prove the Cuntz-
Krieger Uniqueness Theorem for LR(E). In §7 we analyze the ideal structure
of LR(E). Finally, in §8 we discuss extending the coefficients of a Leavitt
path algebra by taking tensor products. We conclude with a discussion of
the significance of the rings LZ(E) and LZn(E).
LEAVITT PATH ALGEBRAS WITH COEFFICIENTS IN A COMMUTATIVE RING 3
2. Preliminaries
When we refer to a graph in this paper, we shall always mean a directed
graph E := (E0, E1, r, s) consisting of a countable set of vertices E0, a
countable set of edges E1, and maps r : E1 → E0 and s : E1 → E0
identifying the range and source of each edge.
Definition 2.1. Let E := (E0, E1, r, s) be a graph. We say that a vertex
v ∈ E0 is a sink if s−1(v) = ∅, and we say that a vertex v ∈ E0 is an
infinite emitter if s−1(v) = ∞. A singular vertex is a vertex that is either
a sink or an infinite emitter, and we denote the set of singular vertices by
E0
sing, and refer to the elements of E0
reg
as regular vertices; i.e., a vertex v ∈ E0 is a regular vertex if and only if
0 < s−1(v) < ∞.
sing. We also let E0
reg := E0 \ E0
in E0 to be paths of length zero. We also let E∗ := S∞
Definition 2.2. If E is a graph, a path is a sequence α := e1e2 . . . en of edges
with r(ei) = s(ei+1) for 1 ≤ i ≤ n−1. We say the path α has length α := n,
and we let En denote the set of paths of length n. We consider the vertices
n=0 En denote the
paths of finite length, and we extend the maps r and s to E∗ as follows: For
α := e1e2 . . . en ∈ En, we set r(α) = r(en) and s(α) = s(e1). A cycle in E
is a path α ∈ E∗ \ E0 with s(α) = r(α). If α := e1 . . . en, then an exit for
α is an edge f ∈ E1 such that s(f ) = s(ei) but f 6= ei for some 1 ≤ i ≤ n.
We say that a graph E satisfies Condition (L) if every cycle in E contains
an exit.
Definition 2.3. We let (E1)∗ denote the set of formal symbols {e∗ : e ∈ E1},
and for α = e1 . . . en ∈ En we define α∗ := e∗
1. We also define
v∗ = v for all v ∈ E0. We call the elements of E1 real edges and the
elements of (E1)∗ ghost edges.
n−1 . . . e∗
ne∗
Definition 2.4. Let E be a directed graph and let R be a ring. A collection
{v, e, e∗ : v ∈ E0, e ∈ E1} ⊆ R is a Leavitt E-family in R if {v : v ∈ E0}
consists of pairwise orthogonal idempotents and the following conditions are
satisfied:
(1) s(e)e = er(e) = e for all e ∈ E1
(2) r(e)e∗ = e∗s(e) = e∗ for all e ∈ E1
(3) e∗f = δe,f r(e) for all e, f ∈ E1
(4) v = X{e∈E 1:s(e)=v}
ee∗ whenever v ∈ E0
reg.
Definition 2.5. Let E be a directed graph, and let K be a field. The Leavitt
path algebra of E with coefficients in K, denoted LK(E), is the universal
K-algebra generated by a Leavitt E-family (see Definition 2.4).
Note that LK(E) is universal for Leavitt E-families in K-algebras; i.e.,
if A is a K-algebra and {av, be, ce∗ : v ∈ E0, e ∈ E1} is a Leavitt E-family
in A, then there exists a K-algebra homomorphism φ : LK(E) → A such
4
MARK TOMFORDE
that φ(v) = av, φ(e) = be, and φ(e∗) = ce∗ for all v ∈ E0 and e ∈ E1. It
is shown in [1, §1] and [2, §1] that for any graph E the generators {v, e, e∗ :
v ∈ E0, e ∈ E1} of LK(E) are all nonzero.
In any algebra generated by a Leavitt E-family {v, e, e∗ : v ∈ E0, e ∈ E1},
we see that
(2.1)
(αβ∗)(γδ∗) =
αγ′δ∗
αδ∗
αβ′∗δ∗
0
if γ = βγ′
if β = γ
if β = γβ′
otherwise.
2.1. Algebras over commutative rings. If R is a commutative ring with
unit 1, then an R-algebra is an abelian group A that has the structure of
both a ring and a (left) R-module in such a way that
(1) r · (xy) = (r · x)y = x(r · y) for all r ∈ R and x, y ∈ A; and
(2) 1 · x = x for all x ∈ A.
Note that as a ring, A is not necessarily commutative and A does not neces-
sarily contain a unit. By a homomorphism between R-algebras we mean an
R-linear ring homomorphism. If A and B are R-algebras, we let HomR(A, B)
denote the collection of R-linear ring homomorphisms from A to B. We ob-
serve that for any R-algebra A, the endomorphism ring HomR(A, A) is an
R-algebra in the obvious way.
If R is a commutative ring, the characteristic of R, denoted char(R), is
defined to be the smallest positive integer n such that nr = 0 for all r ∈ R,
if such an n exists, and 0 otherwise. It is a fact that if K is a field, then
char K is either equal to 0 or a prime p.
Any ring R may be viewed as a Z-algebra in the natural way, and if R has
characteristic n, then R may also be viewed as a Zn-algebra. Furthermore,
if A is an R-algebra and X ⊆ A, then we define
spanR X :=( n
Xi=1
rixi : ri ∈ R and xi ∈ X for all 1 ≤ i ≤ n)
to be the R-submodule of A generated by the set X.
3. Constructing Leavitt path algebras with coefficients in a
commutative ring with unit.
In this section we wish to extend the definition of a Leavitt path algebra
to allow for coefficients in an arbitrary commutative ring with unit.
Definition 3.1. Let E be a directed graph, and let R be a commutative ring
with unit. The Leavitt path algebra with coefficients in R, denoted LR(E), is
the universal R-algebra generated by a Leavitt E-family (see Definition 2.4).
Note that LR(E) is universal for Leavitt E-families in R-algebras; i.e., if
A is a R-algebra and {av, be, ce∗ : v ∈ E0, e ∈ E1} is a Leavitt E-family in
LEAVITT PATH ALGEBRAS WITH COEFFICIENTS IN A COMMUTATIVE RING 5
A, then there exists a R-algebra homomorphism φ : LR(E) → A such that
φ(v) = av, φ(e) = be, and φ(e∗) = ce∗ for all v ∈ E0 and e ∈ E1.
Recall that any ring is a Z-algebra and any ring of characteristic n is a
Zn-algebra. This motivates the following definitions.
Definition 3.2. If E is a graph, the Leavitt path ring of characteristic 0 is
the ring LZ(E), and for each n ∈ N the Leavitt path ring of characteristic n
is the ring LZn(E).
Remark 3.3. In the next proposition we show that the elements of {v, e, e∗ :
v ∈ E0, e ∈ E1} are all nonzero, and that rv 6= 0 for all v ∈ E0 and all
r ∈ R \ {0}. In Proposition 4.9, we are able to prove a stronger result: The
set of paths E∗ in LR(E) is linearly independent over R, and the set of ghost
paths {α∗ : α ∈ E∗} in LR(E) is linearly independent over R.
The construction in the next proposition is an R-algebra version of a sim-
ilar construction that has been done for graph C ∗-algebras (see [19, Theo-
rem 1.2]) and for Leavitt path algebras over fields (see [14, Lemma 1.5]).
Proposition 3.4. If E is a graph and R is a commutative ring with unit,
then the Leavitt path algebra LR(E) has the property that the elements of
the set {v, e, e∗ : v ∈ E0, e ∈ E1} are all nonzero. Moreover,
LR(E) = spanR{αβ∗ : α, β ∈ E∗ and r(α) = r(β)}
and rv 6= 0 for all v ∈ E0 and all r ∈ R \ {0}.
Proof. The fact that e∗f = δe,f r(e) allows us to write any word in the
generators {v, e, e∗ : v ∈ E0, e ∈ E1} as αβ∗ with α, β ∈ E∗. It follows that
LR(E) = spanR{αβ∗ : α, β ∈ E∗ and r(α) = r(β)}.
To see that the elements of the set {v, e, e∗ : v ∈ E0, e ∈ E1} ⊆ LR(E)
are all nonzero, it suffices (due to the universal property) to construct an
R-algebra generated by nonzero elements satisfying the relations described
in Definition 3.1. Define Z := R ⊕ R ⊕ . . . to be the direct sum of countably
many copies of R. For each e ∈ E1 let Ae := Z, and for each v ∈ E0 let
Av :=
Ae
Ms(e)=v
Z ⊕ Ms(e)=v
Z
if 0 < s−1(v) < ∞
Ae
if s−1(v) = ∞
if s−1(v) = 0.
Note that the Av's and Ae's are all mutually isomorphic since each is the
direct sum of countably many copies of R. Let A := Lv∈E 0 Av. For each
v ∈ E0 define Tv : Av → Av to be the identity map, and extend to a
homomorphism Tv : A → A by defining Tv to be zero on A ⊖ Av. Also, for
each e ∈ E1 choose an isomorphism Te : Ar(e) → Ae ⊆ As(e) and extend
to a homomorphism Te : A → A by defining Te to be zero on A ⊖ Ae.
Finally, we define Te∗ : A → A by taking the isomorphism T −1
: Ae ⊆
e
6
MARK TOMFORDE
As(e) → Ar(e) and extending to obtain a homomorphism Te∗ : A → A by
defining Te∗ to be zero on A ⊖ Ae. Let A be the subalgebra of HomR(A, A)
generated by {Tv, Te, Te∗ : v ∈ E0, e ∈ E1}. One can check that {Tv, Te, Te∗ :
v ∈ E0, e ∈ E1} is a collection of nonzero elements satisfying the relations
described in Definition 3.1. Thus the subalgebra of HomR(A, A) generated
by {Tv, Te, Te∗ : v ∈ E0, e ∈ E1} is the desired R-algebra.
Finally, we note that for any v we have Av = R ⊕ M for some R-module
M . Thus for any r ∈ R \ {0}, using the fact that R is unital we have
rTv(1, 0) = Tv(r, 0) = (r, 0) 6= 0. Hence rTv 6= 0. The universal property of
LR(E) then implies that rv 6= 0 for any v ∈ E0 and any r ∈ R \ {0}.
(cid:3)
Corollary 3.5. Let E be a graph and let R be a commutative ring with unit.
Then char LR(E) = char R.
Remark 3.6 (A realization of LR(E)). Suppose E is a graph and R is a
commutative ring with unit. The path algebra of E with coefficients in R
is the R-algebra generated by paths with the operation of path concatena-
tion. (Here vertices are considered as paths of length zero.) In other words,
n=0 En with the
AR(E) is the free R-algebra generated by the paths E∗ =S∞
following relations:
(i) vw = δv,wv for all v, w ∈ E0
(ii) e = er(e) = s(e)e for all e ∈ E1.
If E = (E0, E1, r, s) is a graph, we let E be the graph with vertex set
E0 := E0, edge set E1 := {e, e∗ : e ∈ E1}, and maps r and s extended to
E1 by r(e∗) := s(e) and s(e∗) = r(e) for all e ∈ E1. We see that LR(E)
may be realized as the quotient AR( E)/I, where AR( E) is the path algebra
of E with coefficients in R, and I is the ideal of AR( E) generated by the
elements
(3.1)
(cid:8)e∗f − δe,f r(e) : e, f ∈ E1(cid:9) ∪(cid:8)v − Xs(e)=v
ee∗ : v ∈ E0
reg(cid:9).
4. Properties of Leavitt Path Algebras
4.1. Involution and selfadjoint ideals. As we have seen, any element
k where αk, βk ∈ E∗ with
k=1 rkαkβ∗
x ∈ LR(E) may be written x = PN
r(αk) = r(βk) and rk ∈ R for 1 ≤ k ≤ N .
Remark 4.1. If E is a graph, R is a commutative ring with unit, and LR(E)
is the associated Leavitt path algebra, we may define an R-linear involution
k=1 rkβkα∗
k.
Note that this operation is R-linear, involutive ((x∗)∗ = x), and antimulti-
plicative ((xy)∗ = y∗x∗).
x 7→ x∗ on LR(E) as follows: If x =PN
k, then x∗ =PN
k=1 rkαkβ∗
Definition 4.2. If LR(E) is the Leavitt path algebra of a graph E with
coefficients in R, an ideal I of LR(E) is selfadjoint if I ∗ = I.
LEAVITT PATH ALGEBRAS WITH COEFFICIENTS IN A COMMUTATIVE RING 7
4.2. Enough idempotents and local units. A ring R has enough idempo-
tents if there exists a collection of pairwise orthogonal idempotents {eα}α∈Λ
such that R = Lα∈Λ eαR = Lα∈Λ Reα. A set of local units for a ring R
is a set Λ ⊆ R of commuting idempotents with the property that for any
x ∈ R there exists t ∈ Λ such that tx = xt = x.
If E is a graph, R is a commutative ring with unit, and LR(E) is the
associated Leavitt path algebra, then
LR(E) = Mv∈E 0
vLR(E) = Mv∈E 0
LR(E)v
so LR(E) is a ring with enough idempotents. Furthermore, if E0 is finite,
k=1 vk, then {tn}n∈N is a set of local units for LR(E).
Definition 4.3. A ring R is idempotent if R2 = R; that is, if every x ∈ R
not have a unit, but if we list the vertices of E as E0 = {v1, v2, . . .} and set
then 1 = Pv∈E 0 v is a unit for LR(E). If E0 is infinite, then LR(E) does
tn :=Pn
can be written as x =Pn
Remark 4.4. We see that if R is a ring with a set of local units, then R is
idempotent: If x ∈ R, then there exists an idempotent t ∈ R with x = tx.
Consequently, the Leavitt path algebra LR(E) is an idempotent ring.
k=1 akbk for a1, . . . an, b1, . . . , bn ∈ R.
4.3. Z-graded rings. We show that all Leavitt path algebras have a natural
Z-grading.
Definition 4.5. If R is a ring, we say R is Z-graded if there is a a collection
of additive subgroups {Rk}k∈Z of R with the following two properties:
(1) R =Lk∈Z Rk
(2) RjRj ⊆ Rj+k for all j, k ∈ Z.
The subgroup Rk is called the homogeneous component of R of degree k.
Definition 4.6. If R is a graded ring, then an ideal I of R is a Z-graded
ideal if I = Lk∈Z(I ∩ Rk). If φ : R → S is a ring homomorphism between
Z-graded rings, then φ is a graded ring homomorphism if φ(Rk) ⊆ Sk for all
n ∈ Z.
Note that the kernel of a Z-graded homomorphism is a Z-graded ideal.
Also, if I is a Z-graded ideal in a Z-graded ring R, then the quotient R/I
admits a natural Z-grading and the quotient map R → R/I is a Z-graded
homomorphism. In this paper we will be concerned only with Z-gradings,
and hence we will often omit the prefix Z and simply refer to rings, ideals,
homomorphisms, etc. as graded.
Proposition 4.7. If E is a graph and R is a commutative ring with unit,
then we may define a Z-grading on the associated Leavitt path algebra LR(E)
by setting
LR(E)k :=( N
Xi=1
riαiβ∗
i : αi, βi ∈ E∗, ri ∈ R, and αi − βi = k for all i) .
8
MARK TOMFORDE
Proof. Let A be the free R-algebra generated by E0 ∪ E1 ∪ (E1)∗. Then A
has a unique Z-grading for which the elements of E0, E1, and (E1)∗ have
degrees 0, 1, and −1, respectively. Let I be the ideal in A generated by
elements of the following type:
• vw − δv,wv for v, w ∈ E0
• e − er(e) for e ∈ E1
• e − s(e)e for e ∈ E1
• e∗f − δe,f r(e) for e, f ∈ E1
• v −Ps(e)=v ee∗ for v ∈ E0
reg.
Since the elements generating I are all homogeneous of degree zero, it follows
that I is a graded ideal. Furthermore, we see that A/I ∼= LR(E), so that
LR(E) is graded with the homogeneous elements of degree k equal to the set
of R-linear combinations of elements of the form αβ∗ with α − β = k. (cid:3)
Definition 4.8. If x ∈ LR(E), we say that x is a polynomial in real edges if
i=1 riαi for ri ∈ R \ {0} and αi ∈ E∗. In this case we also define the
x =Pn
degree of x to be
deg x = max{αi : 1 ≤ i ≤ n}.
Note that deg x is independent of how x is written.
Proposition 4.9. Let E be a graph and let R be a commutative ring with
unit. The set of paths E∗ in LR(E) is linearly independent over R. Likewise,
the set of ghost paths {α∗ : α ∈ E∗} in LR(E) is linearly independent over
R.
Proof. Suppose that α1, . . . , αn ∈ E∗, andPn
i=1 riαi = 0 for some r1, . . . , rn ∈
R. Using the Z-grading on LR(E) we may, without loss of generality, assume
that all the αi's have the same length. Then for any 1 ≤ j ≤ n we have
rj(αj) = α∗
i=1 riαi) = 0. Proposition 3.4 implies that ri = 0. It
follows that {α1, . . . , αn} is linearly independent over R. A similar argument
works for ghost paths.
(cid:3)
j (Pn
j αj = α∗
4.4. Morita equivalence. Throughout this paper we will need to discuss
Morita equivalence for rings that do not necessarily have an identity element.
We recall the necessary definitions and results here.
Definition 4.10. If R is a ring, we say that a left R-module M is unital
if RM = M . We also say that M is nondegenerate if for all m ∈ M
we have that Rm = 0 implies that m = 0. We let R-MOD denote the
full subcategory of the category of all R-modules whose objects are unital
nondegenerate R-modules. (Note that if R is unital, R-MOD is the usual
category of R-modules.) When R and S are rings, and RMS is a bimodule,
we say M is unital if RM = M and M S = M .
Definition 4.11. Let R and S be idempotent rings. A (surjective) Morita
context (R, S, M, N, ψ, φ) between R and S consists of unital bimodules RMS
LEAVITT PATH ALGEBRAS WITH COEFFICIENTS IN A COMMUTATIVE RING 9
and SNR, a surjective R-module homomorphism ψ : M ⊗S N → R, and a
surjective S-module homomorphism φ : N ⊗R M → S satisfying
φ(n ⊗ m)n′ = nψ(m ⊗ n′)
and
m′φ(n ⊗ m) = ψ(m′ ⊗ n)m
for every m, m′ ∈ M and n, n′ ∈ N . We say that R and S are Morita
equivalent in the case that there exists a Morita context.
It is proven in [13, Proposition 2.5] and [13, Proposition 2.7] that R-MOD
and S-MOD are equivalent categories if and only if there exists a Morita
context (R, S, M, N, ψ, φ).
In addition, the following result is obtained in
[13].
Proposition 4.12. [13, Proposition 3.5] Let R and S be Morita equivalent
idempotent rings, and let (R, S, M, N, ψ, φ) be a Morita context. If
LR := {I ⊆ R : I is an ideal and RIR = I}
and
LS := {I ⊆ S : I is an ideal and SIS = I},
then there is a lattice isomorphism from LR onto LS given by I 7→ φ(N I, M )
with inverse given by I 7→ ψ(M I, N ).
Remark 4.13. Note that when R is a ring with a set of local units, LR is the
lattice of ideals of R. Thus if each of R and S is a ring with a set of local
units, and if R and S are Morita equivalent, then the lattice of ideals of R
is isomorphic to the lattice of ideals of S.
Recall that in rings the property of being a ring ideal is not transitive;
i.e., if R is a ring, I is an ideal of R, and J is an ideal of I, then it is not
necessarily true that J is an ideal of R. Despite this fact, there is a special
case when the implication does hold, and this will be of use to us.
Lemma 4.14. Let R be a ring and let I be an ideal of R with the property
that I has a set of local units. If J is an ideal of I, then J is an ideal of R.
Proof. Let r ∈ R and x ∈ J. Since I has a set of local units, there exists
t ∈ I with tx = x. Because I is an ideal, we have that rt ∈ I. Hence
rx = r(tx) = (rt)x ∈ J. A similar argument shows that xr ∈ I.
(cid:3)
5. The Graded Uniqueness Theorem
Lemma 5.1. Let I be a graded ideal of LR(E). Then I is generated as an
ideal by the set I0 := I ∩ LR(E)0.
Proof. Let k > 0. Given x ∈ Ik := I ∩LR(E)k, we may write x =Pn
i=1 αixi,
where xi ∈ LR(E)0 and αi ∈ Ek for all 1 ≤ i ≤ n, and αi 6= αj for i 6= j.
Then for any 1 ≤ j ≤ n we have
xj = α∗
j n
Xi=1
αixi! = α∗
j x ∈ I.
10
MARK TOMFORDE
Thus xj ∈ I0 and Ik = LR(E)kI0. Similarly, I−k = I0LR(E)−k. Since I is a
(cid:3)
graded ideal, I =Lk∈Z Ik, and I is generated as an ideal by I0.
Lemma 5.2. Let E be a graph, and let R be a commutative ring with unit.
If x ∈ LR(E)0 and x 6= 0, then there exists α, β ∈ E∗ such that α∗xβ = rv
for some v ∈ E0 and some r ∈ R \ {0}.
Proof. Define GN := spanR{αβ∗ : α, β ∈ Em for 1 ≤ m ≤ N }. Then
N =0 GN . We will prove by induction on N that if x ∈ GN and
x 6= 0, then there exists α, β ∈ E∗ such that α∗xβ = rv for some v ∈ E0
i=1 rivi
for vi ∈ E0 and nonzero ri ∈ R with vi 6= vj for i 6= j. If we let α = β = v1,
then α∗xβ = r1v1.
LR(E)0 = S∞
and some r ∈ R \ {0}. In the base case we have N = 0, and x = Pn
In the inductive step, we assume that for all nonzero y ∈ GN −1 there exists
α′, β′ ∈ E∗ such that (α′)∗yβ′ = rv for some v ∈ E0 and some r ∈ R \ {0}.
Suppose that x ∈ GN and x 6= 0. Then we can write
x =
M
Xi=1
riαiβ∗
i +
sjvj,
P
Xj=1
for α, β ∈ E∗ with αi = βi ≥ 1, vj ∈ E0 with vj 6= vj ′ for j 6= j′, and
ri, sj ∈ R \ {0}.
If any vj is a sink, we may let α = β = vj, and then
α∗xβ = sjvj. If any vj is an infinite emitter, then we may choose an edge
e ∈ E1 with s(e) = vj and e not equal to any edge appearing in any of the
αi's. If we let α = β = e, then α∗xβ = e∗sjvje = sjr(e). The only other
case to consider is when every vj is a regular vertex (i.e., neither a sink nor
ee∗
to write x as a linear combination of elements γδ∗ where γ, δ ∈ E∗ with
γ = δ ≥ 1. By regrouping the elements in this linear combination, we
may write
an infinite emitter). In this case we may use the relation vj = Ps(e)=vj
P
Q
x =
Xi=1
Xj=1
eixi,jf ∗
j
j 6= 0 for all i, j. Since e1x1,1f ∗
6= ei′ for i 6= i′ and fj 6= fj ′ for j 6= j′; and
where ei, fi ∈ E1 with ei
xi,j ∈ GN −1 with eixi,jf ∗
1 6= 0, it follows that
r(e1)x1,1r(f1) 6= 0. Because r(e1)x1,1r(f1) 6= 0 and r(e1)x1,1r(f1) ∈ GN −1,
the inductive hypothesis implies that there exists α′, β′ ∈ E∗ such that
(α′)∗r(e1)x1,1r(f1)β′ = rv for some v ∈ E0 and some r ∈ R \ {0}. If we let
α := e1α′ and β := f1β′, then
α∗xβ = (α′)∗e∗
1xf1β′ = (α′)∗e∗
1 f1β′ = (α′)∗r(e1)x1,1r(f1)β′ = rv.
The Principle of Mathematical Induction shows that the claim holds for all
N , and hence the lemma holds for all nonzero x in LR(E)0.
(cid:3)
1e1x1,1f ∗
Theorem 5.3 (Graded Uniqueness Theorem). Let E be a graph, and let R
be a commutative ring with unit. If S is a graded ring and φ : LR(E) → S is
LEAVITT PATH ALGEBRAS WITH COEFFICIENTS IN A COMMUTATIVE RING 11
a graded ring homomorphism with the property that φ(rv) 6= 0 for all v ∈ E0
and for all r ∈ R \ {0}, then φ is injective.
Proof. Suppose that x ∈ LR(E)0 ∩ ker φ. If x is nonzero, then by Lemma 5.2
there exists α, β ∈ E∗ such that α∗xβ = rv for some v ∈ E0 and some
r ∈ R \ {0}. But then φ(rv) = φ(α∗xβ) = φ(α∗)φ(x)φ(β) = 0, which is a
contradiction. Hence x = 0, and LR(E)0 ∩ ker φ = {0}.
Since φ is a graded ring homomorphism, ker φ is a graded ideal of LR(E).
It follows from Lemma 5.1 that ker φ is generated as an ideal by LR(E)0 ∩
ker φ = {0}. Thus ker φ = {0}, and φ is injective.
(cid:3)
Corollary 5.4. Let E be a graph, and let K be a field. If S is a graded ring
and φ : LK(E) → S is a graded ring homomorphism with the property that
φ(v) 6= 0 for all v ∈ E0, then φ is injective.
Remark 5.5. The Graded Uniqueness Theorem for Leavitt path algebras
may be thought of as an analogue of the Gauge-Invariant Uniqueness The-
orem for graph C ∗-algebras, with the grading playing the role of the gauge
action.
In [8] Cuntz and Krieger showed that if A is a finite {0, 1}-matrix satisfy-
ing Condition (I), then there is a unique C ∗-algebra generated by a nonzero
Cuntz-Krieger A-family, which they denote by OA. Universal Cuntz-Krieger
algebras of finite {0, 1}-matrices were introduced in [15], and a Gauge-
Invariant Uniqueness Theorem for these algebras was proven in [15, The-
orem 2.3]. A Gauge-Invariant Uniqueness Theorem for C ∗-algebras of row-
finite graphs was obtained in [5, Theorem 2.1], and this was extended to
C ∗-algebras of non-row-finite graphs in [6, Theorem 2.1]. Furthermore, the
Gauge-Invariant Uniqueness Theorem was generalized to Cuntz-Krieger al-
gebras of infinite matrices in [24, Theorem 2.7] and to Cuntz-Pimnser alge-
bras in [11, Theorem 4.1] and [17, Theorem 6.4].
In the Leavitt path algebra setting, the gauge action is replaced by a Z-
grading -- in fact, if one views the Leavitt path algebra LC(E) as a dense
∗-subalgebra of the graph C ∗-algebra C ∗(E), then the gauge action on C ∗(E)
induces a Z-grading on LC(E) (see the proof of [26, Theorem 7.3] for details).
In [3, Theorem 5.1], Ara, Moreno, and Pardo proved the Graded Uniqueness
Theorem for LK(E), where K is a field and E is a row-finite graph. A proof
of the Graded Uniqueness Theorem for LK(E), where K is a field and E
is an arbitrary graph, was given by the author in [26, Theorem 4.8]. The
proof in Theorem 5.3 uses different techniques than [3] or [26].
6. The Cuntz-Krieger Uniqueness Theorem
Recall that a graph E is said to satisfy Condition (L) if every cycle in E
has an exit. (See Definition 2.2 for more details.)
Lemma 6.1. Suppose E is a graph satisfying Condition (L). If F is a finite
subset of E∗ \ E0 and v ∈ E0, then there exists a path α ∈ E∗ such that
s(α) = v and for every µ ∈ F we have α∗µα = 0.
12
MARK TOMFORDE
Proof. Given v ∈ E0 and a finite subset F ⊆ E∗, consider two cases.
Case I: There is a path from v to a sink in E.
In this case, let α be a
path with s(α) = v and r(α) a sink. For any µ ∈ F , we see that α∗µα is
nonzero if and only if there exists ν ∈ E∗ \ E0 such that µα = αν, which is
impossible since r(α) is a sink. Thus α∗µα = 0.
Case II: There is no path from v to a sink in E.
Let M = max{µ : µ ∈ F } + 1. If there is a path α = α1 . . . αM ∈ EM
with s(α) = v and no repeated vertices, then for any µ ∈ F we see that
α∗µα is nonzero if and only if there exists ν ∈ E∗ \ E0 such that µα = αν,
which is impossible since this would imply that s(α1) = s(αj) for some j ≥ 2
contradicting that α has no repeated vertices. Thus α∗µα = 0.
Otherwise, every path EM with s(λ) = v has repeated vertices, and there
exists a path from v to the base point of a cycle in E. Choose a path τ
of minimal length such that s(τ ) = v and r(τ ) is the base point of a cycle.
Choose a cycle β of minimal length based at r(τ ). Let f be an exit for β,
and let β′ be the segment of β from r(τ ) to s(f ). By the minimality of τ ,
the edge f is not equal to any of the edges in the path τ . Likewise, by the
minimality of β, the edge f is not equal to any of the edges on the cycle β
or the path β′. Thus the path α := τ ββ . . . ββ′f has the property that f is
not equal to any edge αi for 1 ≤ i ≤ α − 1. By choosing sufficiently many
repetitions of the cycle β we can ensure that α has length greater than or
equal to M (to avoid the possibility that α ∈ F ). Then we have that α∗µα
is nonzero if and only if there exists ν ∈ E∗ \ E0 such that µα = αν, which
is impossible since this would imply that f = αj for some 1 ≤ j ≤ α − 1.
Thus α∗µα = 0.
(cid:3)
Lemma 6.2. Let E be a graph satisfying Condition (L), and let R be a
commutative ring with unit. If x ∈ LR(E) is a polynomial in only real edges
and x 6= 0, then there exist paths α, β ∈ E∗ such that α∗xβ = rv for some
v ∈ E0 and some r ∈ R \ {0}.
Proof. We will prove by induction on N that if x ∈ LR(E) is a nonzero
polynomial in only real edges with deg x = N , then there exist paths α, β ∈
E∗ such that α∗xβ = rv for some v ∈ E0 and some r ∈ R \ {0}. In the
i=1 rivi for vi ∈ E0 and nonzero
base case we have deg x = 0, so that x =PM
ri ∈ R with vi 6= vj for i 6= j. If we let α = β = v1, then α∗xβ = r1v1.
In the inductive step, we assume that our claim holds for all nonzero
polynomials in real edges with degree N − 1 or less. Suppose x ∈ LR(E) is
a nonzero polynomial in real edges with deg x = N . If x has no terms of
degree 0, then we may write
M
Xi=1
with each xi a nonzero polynomial in real edges of degree N − 1 or less, and
ei ∈ E1 with ei 6= ej for i 6= j. Then e∗
1x = x1 is a nonzero polynomial of
degree N − 1 or less, so by the inductive hypothesis there exists α′, β ∈ E∗
x =
eixi
LEAVITT PATH ALGEBRAS WITH COEFFICIENTS IN A COMMUTATIVE RING 13
such that (α′)∗x1β = rv for some v ∈ E0 and r ∈ R\{0}. If we let α := e1α′,
1xβ = (α′)∗x1β = rv and the claim holds. On the other
then α∗xβ = (α′)∗e∗
hand, if x does have a term of degree 0, then we may write
M
K
x =
riαi +
sjvj
Xi=1
Xj=1
where the αi's are paths of length 1 or greater, each ri, sj ∈ R \ {0}, and
the vj's are vertices with vj 6= vj ′ for j 6= j′. Let F := {αi : 1 ≤ i ≤ M }.
By Lemma 6.1 there exists α ∈ E∗ such that s(α) = v1 and for every αi we
have α∗αiα = 0. If we let β = α, then we have
M
K
α∗xβ =
riα∗αiα +
sjα∗vjα = s1α∗v1α = s1r(α).
Xi=1
Xj=1
By the Principle of Mathematical Induction, we may conclude that the claim
holds for all N , and hence the lemma holds for all nonzero polynomials in
only real edges.
(cid:3)
Lemma 6.3. Let E be a graph and let R be a commutative ring with unit.
Let x ∈ LR(E) and suppose that x is a polynomial in real edges with x 6= 0.
If there exists v ∈ E0 with xv = x, then for any e ∈ E1 with s(e) = v we
have xe 6= 0.
Proof. Since LR(E) is graded with LR(E) = Lk∈Z LR(E)k, it suffices to
case we may write x = PM
prove the claim when x is homogeneous of degree k for some k ≥ 0. In this
i=1 riαi with each ri ∈ R \ {0} and each αi ∈ Ek
with αi 6= αi′ for i 6= i′. Since xv = x, we may also assume that r(αi) = v
for all i. For any e ∈ E1 with s(e) = v we see that αie ∈ Ek+1. If xe = 0,
then
r1r(e) = e∗α∗
1(r1α1e) = e∗α∗
1 M
Xi=1
riαi! e = e∗α∗
1(xe) = 0,
which contradicts Proposition 3.4. Hence xe 6= 0.
(cid:3)
Lemma 6.4. Let E be a graph and let R be a commutative ring with unit.
If x ∈ LR(E) and x 6= 0 then there exists γ ∈ E∗ such that xγ 6= 0 and xγ
is a polynomial in only real edges.
Proof. Define
riαiβ∗
i : ri ∈ R, αi, βi ∈ E∗, and βi ≤ N for 1 ≤ i ≤ M) .
AN :=( M
Xi=1
Then LR(E) =S∞
in only real edges. In the base case we have N = 0, and x = PM
N =0 AN . We will prove by induction on N that if x ∈ AN
and x 6= 0, then there exists γ ∈ E∗ such that xγ 6= 0 and xγ is a polynomial
i=1 riαi so
that x is a polynomial in real edges. Choose v ∈ E0 such that xv 6= 0. Then
xv is a polynomial in only real edges, and the claim holds.
14
MARK TOMFORDE
In the inductive step, we assume that for all nonzero x′ ∈ AN −1, there
exists γ ∈ E∗ such that x′γ 6= 0 and x′γ is a polynomial in only real edges.
i ∈ AN , we may choose v ∈ E0 such that
i=1 riαiβ∗
Given an element x =PM
xv 6= 0. By regrouping terms, we may write
xv =
P
Xj=1
xje∗
j + y
where the xj's are polynomials in which each term has N − 1 ghost edges
or fewer (so that xj ∈ AN −1), each ej ∈ E1 with s(ej) = v and ej 6= ej ′
for j 6= j′, and y is a polynomial in only real edges with yv = y. If y = 0,
then xve1 = x1 6= 0 and by the inductive hypothesis there exists γ′ such
that x1γ′ is a nonzero polynomial in only real edges.
If γ := e1γ′, then
xγ = xve1γ′ = x1γ′ is a nonzero polynomial in only real edges.
If y 6= 0, then we consider three possibilities for v. If v is a regular vertex,
then v =Ps(e)=v ee∗ and xv =PP
j=1 xje∗
j +Ps(e)=v yee∗ and by regrouping
we are as in the situation described in the previous paragraph, so we may
argue as done there. If v is a sink, then there are no edges whose source
is v, so xv = y and we may choose γ := v, and the claim holds. If v is an
infinite emitter, then we may choose e ∈ E1 with s(e) = v and e 6= ej for all
j e + ye = ye.
Since y is a nonzero polynomial in only real edges with yv = y, it follows
from Lemma 6.3 that ye is a nonzero polynomial in only real edges. By the
Principle of Mathematical Induction, we may conclude that the claim holds
for all N , and hence the lemma holds for all nonzero x ∈ LR(E).
(cid:3)
1 ≤ j ≤ P . If we let γ := e, then xγ = xe = xve = PP
j=1 xje∗
Theorem 6.5 (Cuntz-Krieger Uniqueness Theorem). Let E be a graph sat-
isfying Condition (L), and let R be a commutative ring with unit. If S is
a ring and φ : LR(E) → S is a ring homomorphism with the property that
φ(rv) 6= 0 for all v ∈ E0 and for all r ∈ R \ {0}, then φ is injective.
Proof. Suppose x ∈ ker φ and x 6= 0. By Lemma 6.4 there exists γ ∈ E∗ such
that xγ is a nonzero polynomial in all real edges. Consequently, Lemma 6.2
implies that there exists α, β ∈ E∗ such that α∗xγβ = rv for some v ∈ E0
and some r ∈ R \ {0}. Then φ(rv) = φ(α∗)φ(x)φ(γβ) = 0, which is a
contradiction. Hence ker φ = {0}, and φ is injective.
(cid:3)
Corollary 6.6. Let E be a graph satisfying Condition (L), and let K be a
field. If S is a ring and φ : LK(E) → S is a ring homomorphism with the
property that φ(v) 6= 0 for all v ∈ E0, then φ is injective.
Remark 6.7. The Cuntz-Krieger Uniqueness Theorem has a long history
in the C ∗-algebra setting, and the Cuntz-Krieger Uniqueness Theorem for
graph C ∗-algebras may be viewed as a vast generalization of Coburn's The-
orem [22, Theorem 3.5.18].
LEAVITT PATH ALGEBRAS WITH COEFFICIENTS IN A COMMUTATIVE RING 15
The first Cuntz-Krieger Uniqueness Theorem was proven by Cuntz and
Krieger in [8, Theorem 2.13], where they showed that if A is a finite {0, 1}-
matrix satisfying Condition (L) then any two Cuntz-Krieger A-families com-
posed of nonzero partial isometries generate isomorphic C ∗-algebras. This
was generalized to C ∗-algebras of locally finite graphs in [19, Theorem 3.7]
using groupoid techniques, and [19] is also where Condition (L) was first
introduced. In [5, Theorem 3.1] a Cuntz-Krieger Theorem was proven for
C ∗-algebras of row-finite graphs, and the proof avoided groupoid methods
in favor of a direct analysis of the AF-core. A Cuntz-Krieger Uniqueness
Theorem for C ∗-algebras of non-row-finite graphs was obtained in [24, The-
orem 1.5] by realizing the graph C ∗-algebra as an increasing union of C ∗-
algebras of finite graphs. When Cuntz and Krieger's original theorem [8,
Theorem 2.13] is translated into a theorem about graphs, one obtains the
Cuntz-Krieger Uniqueness Theorem for C ∗-algebras of finite graphs with no
sinks and, moreover, Condition (I) is equivalent to Conidition (L) for these
graphs. Additionally, the Cuntz-Krieger Uniqueness Theorem was extended
to Cuntz-Krieger algebras of infinite matrices [10, Theorem 13.1].
Although there is a Cuntz-Krieger Uniqueness Theorem for Leavitt path
algebras, this result is independent of the graph C ∗-algebra result -- neither
the Cuntz-Krieger Uniqueness Theorem for Leavitt path algebras nor the
Cuntz-Krieger Uniqueness Theorem for graph C ∗-algebras may be used to
obtain the other. In [1, Corollary 3.3], Abrams and Aranda-Pino first proved
a weak version of the Cuntz-Krieger Uniqueness Theorem for LK(E), where
K is a field and E is a row-finite graph. Later, the author proved a lemma
(see [26, Lemma 6.5]) that, with [1, Corollary 3.3], gives a full Cuntz-Krieger
Uniqueness Theorem for LK(E) when E is a row-finite graph. A proof of
the Cuntz-Krieger Uniqueness Theorem for LK (E), where K is a field and
E is an arbitrary graph, was given by the author in [26, Theorem 6.8]. The
proof in [26] uses the process of desingularization [26, Lemma 6.7] to show
that the Cuntz-Krieger Uniqueness Theorem in the row-finite case implies
the Cuntz-Krieger Uniqueness Theorem for arbitrary graphs. The proof in
Theorem 6.5 uses different techniques than [1] or [26], and does not require
one to consider the row-finite case first.
7. Ideals in Leavitt path algebras
In this section we analyze ideals in LR(E). We will see that for ideals of
LR(E) we will not only be concerned with which vertices are in the ideal,
but also which multiples of the vertices are in the ideal. To motivate the
results in this section, we start with an example.
Example 7.1. Let E be the graph with two vertices and no edges, and let
R = Z. Then LZ(E) ∼= Z ⊕ Z.
If we consider the ideals of LZ(E), we
see that they are of the form nZ ⊕ mZ for n, m ∈ {0, 1, 2, . . . , ∞}. We
would like to consider the ideals that are reflected in the structure of the
graph -- in particular, those ideals that are generated by vertices of the
16
MARK TOMFORDE
graph. However, if we list the vertices of E as E0 = {v, w}, then there are
four subsets of vertices, ∅, {v}, {w}, {v, w}, and the ideals generated by these
sets are 0, Z⊕0, 0⊕Z, Z⊕Z. These are the only ideals generated by subsets
of vertices, and each of them has the property that if a nonzero multiple of
a vertex in in the ideal, then that vertex is in the ideal. Consequently, it is
only these kind of ideals that will be determined by subsets of vertices in
the graph. This motivates the following definition.
Definition 7.2. Let R be a commutative ring with unit, and let E be a graph.
If I is an ideal in LR(E), we say that I is basic if whenever r ∈ R \ {0} and
v ∈ E0, we have rv ∈ I implies v ∈ I.
Remark 7.3. Observe that if K is a field, then every ideal in LK(E) is basic.
In this section we show that saturated hereditary subsets of vertices cor-
respond to graded basic ideals. Throughout this section we restrict our
attention to the case of row-finite graphs in order to avoid many of the
complications that arise in the non-row-finite case. Our hope is that this
will make our investigations easier for the reader to follow. Despite this,
most of the results in this section do generalize to the non-row-finite setting,
provided one uses admissible pairs in place of saturated hereditary subsets.
Definition 7.4. Let E be a graph. A subset H ⊆ E0 is hereditary if for
all e ∈ E0 and s(e) ∈ H imply that r(e) ∈ H. A hereditary subset H is
saturated if whenever v ∈ E0
reg then r(s−1(v)) ⊆ H implies that v ∈ H.
For any hereditary set X, we define the saturation X to be the smallest
saturated hereditary subset of E0 containing X.
Observe that intersections of saturated hereditary subsets are saturated
hereditary. Also, unions of saturated hereditary subsets are hereditary, but
not necessarily saturated.
In any R-algebra A, the ideals of A are partially ordered by inclusion and
form a lattice under the operations I ∧ J := I ∩ J and I ∨ J := I + J.
(Note that I + J is the smallest ideal containing I ∪ J.) This lattice has a
maximum element A and a minimum element {0}.
Likewise, for any graph E = (E0, E1, r, s), the saturated hereditary sub-
sets of E0 are partially ordered by inclusion and form a lattice under the
operations H1 ∧ H2 := H1 ∩ H2 and H1 ∨ H2 := H1 ∪ H2. This lattice has
a maximum element E0 and a minimum element ∅.
Definition 7.5. Let E = (E0, E1, r, s) be a graph and H ⊆ E0 be a saturated
hereditary subset. We define (E\H) to be the graph with (E\H)0 := E0\H,
(E \ H)1 := E1 \ r−1(H), and r(E\H) and s(E\H) are obtained by restricting
r and s to (E \ H)1. We also define EH to be the graph with E0
H := H,
E1
H := s−1(H), and rEH and sEH are obtained by restricting r and s to E1
H.
Lemma 7.6. Let E be a graph, and let R be a commutative ring with unit.
If I is an ideal of LR(E), then the set HI := {v : v ∈ I} is a saturated
hereditary subset.
LEAVITT PATH ALGEBRAS WITH COEFFICIENTS IN A COMMUTATIVE RING 17
Proof. If e ∈ E1 and s(e) ∈ H, then s(e) ∈ I so r(e) = e∗e = e∗s(e)e ∈ I
and r(e) ∈ H. Thus H is hereditary.
If v ∈ E0
reg and r(s−1(v)) ⊆ H, then for each e ∈ s−1(v) we have r(e) ∈ H
and r(e) ∈ I so ee∗ = er(e)e∗ ∈ I. Thus v = Ps(e)=v ee∗ ∈ I, and v ∈ H.
Hence H is saturated.
(cid:3)
Proposition 7.7. Let E be a graph, and let R be a commutative ring with
unit. If H is a saturated hereditary subset of E0, and IH is the two-sided
ideal in LR(E) generated by {v : v ∈ H}, then
IH = spanR{αβ∗ : α, β ∈ E∗ and r(α) = r(β) ∈ H},
IH is a graded basic ideal, and {v ∈ E0 : v ∈ IH} = H. Moreover, IH is a
selfadjoint ideal that is also an idempotent ring.
Proof. We first observe that the multiplication rules of (2.1) imply that
spanR{αβ∗ : α, β ∈ E∗ and r(α) = r(β) ∈ H} is a two-sided ideal containing
H. It follows that IH ⊆ spanR{αβ∗ : α, β ∈ E∗ and r(α) = r(β) ∈ H}.
Furthermore, if v ∈ H, then for any α, β ∈ E∗ with r(α) = r(β) = v, the
element αvβ∗ = αβ∗ is in any ideal containing v. Hence IH = spanR{αβ∗ :
α, β ∈ E∗ and r(α) = r(β) ∈ H}.
To see that IH is graded it suffices to notice that αβ∗ is homogeneous of
degree α − β. In addition, we see IH is selfadjoint because (αβ∗) = βα∗.
Next we show that IH is a basic ideal. Let v ∈ E0 and suppose that rv ∈ IH
for some r ∈ R \ {0}. Let E \ H be the graph of Definition 7.5. Then the
vertices, edges, and ghost edges of E \ H, which generate LR(E \ H), may
be extended to a Leavitt E-family by simply defining elements to be zero
if v ∈ H or r(e) ∈ H. By the universal property of LR(E), we obtain an
R-algebra homomorphism φ : LR(E) → LR(E \ H) with
φ(v) =(v
0
if v ∈ E0 \ H
if v ∈ H
φ(e) =(e
if r(e) ∈ E0 \ H
0 if r(e) ∈ H
and
φ(e∗) =(e∗
0
if r(e) ∈ E0 \ H
if r(e) ∈ H.
Thus ker φ is a two-sided ideal of LR(E) containing H, and it follows that
IH ⊆ ker φ. Hence rφ(v) = φ(rv) = 0, and since v is a vertex in E0, either
φ(v) = v or φ(v) = 0. But Proposition 3.4 implies that in LR(E \ H) we
have rv 6= 0 for all v ∈ (E \ H)0 and all r ∈ R \ {0}. Thus φ(v) = 0 and
v ∈ H. Hence v ∈ IH, and IH is a basic ideal.
We next show that the set {v ∈ E0 : v ∈ IH} is precisely H. To begin,
we trivially have H ⊆ {v ∈ E0 : v ∈ IH}. For the reverse inclusion we use
the fact that IH ⊆ ker φ to conclude that v /∈ H implies that φ(v) 6= 0 so
that v /∈ ker φ and v /∈ IH . Hence {v ∈ E0 : v ∈ IH} = H.
Finally we show that IH is an idempotent ring. Any x ∈ IH has the
i with r(αi) = r(βi) ∈ H. For each i, define vi :=
i=1 riαiβ∗
form x = PN
18
MARK TOMFORDE
r(αi) = r(βi). Then riαiβ∗
viβ∗
for a1, . . . , aN , b1, . . . , bN ∈ IH. Thus IH is an idempotent ring.
i ), and since riαivi ∈ IH and
i ∈ IH, we see that any x ∈ IH may be written as x = a1b1 + . . . + aN bN
(cid:3)
i = (riαivi)(viβ∗
Lemma 7.8. Let E be a graph, and let R be a commutative ring with unit.
If X is a hereditary subset of E0, and IX is the two-sided ideal in LR(E)
generated by {v : v ∈ X}, then
IX = IX.
In particular, IX is a graded basic ideal that is also an idempotent ring.
Proof. Since X ⊆ X, we have IX ⊆ IX . Conversely, if we let H := {v ∈ E0 :
v ∈ IX}, then it follows from Lemma 7.6 that H is a saturated hereditary
subset containing X. Thus X ⊆ H, and v ∈ X implies v ∈ IX . Hence
IX ⊆ IX.
(cid:3)
Theorem 7.9. Let E be a graph, and let R be a commutative ring with
unit. Using the notation of Definition 7.5 and Proposition 7.7, we have the
following:
(1) The map H 7→ IH is a lattice isomorphism from the lattice of satu-
rated hereditary subsets of E0 onto the lattice of graded basic ideals
of LR(E). In particular, the graded basic ideals of LR(E) form a
lattice with
IH1 ∧ IH2 = IH1∩H2
and
IH1 ∨ IH2 = IH1∪H2
for any saturated hereditary subsets H1 and H2.
(2) For any saturated hereditary subset H we have that LR(E)/IH is
canonically isomorphic to LR(E \ H).
(3) For any hereditary subset X the ideal IX and the Leavitt path algebra
LR(EX ) are Morita equivalent as rings.
Proof. We shall first prove (2), then (1), and then (3).
Proof of (2): We shall show that LR(E)/IH ∼= LR(E \ H). Let {v : v ∈
E0}∪{e, e∗ :∈ E1} be the generators for LR(E). Then {v +IH : v ∈ E \H}∪
e + IH, e∗ + IH : r(e) /∈ H} is a collection of elements satisfying the Leav-
itt path algebra relations for EH and generating LR(E)/IH . Hence there
exists a surjective R-algebra homomorphism φ : LR(EH ) → LR(E)/IH .
Proposition 7.7 shows that IH is a graded ideal, and hence φ is a graded
homomorphism. Furthermore, if v ∈ E0
H, then v /∈ H and the previous
paragraph implies that v /∈ IH. Since Proposition 7.7 shows that IH is a
basic ideal, for all v ∈ E0
H and all r ∈ R \ {0}, we have φ(rv) = rv + IH 6= 0.
It follows from the Graded Uniqueness Theorem 5.3 that φ is injective. Thus
φ is an isomorphism and LR(E)/IH
Proof of (1): We shall show that H 7→ IH is a lattice isomorphism. To
see that this map is surjective, let I be a graded basic ideal in LR(E), and
set H := {v ∈ E0 : v ∈ I}. Since IH ⊆ I, we see that IH and I contain the
same v's. Therefore, just as in the proof of Part (2), we see that LR(E)/IH
∼= LR(E \ H).
LEAVITT PATH ALGEBRAS WITH COEFFICIENTS IN A COMMUTATIVE RING 19
and LR(E)/I are generated by nonzero elements satisfying the Leavitt path
algebra relations for E \ H. Since both IH and I are graded, both quotients
are graded, and the quotient map π : LR(E)/IH → LR(E)/I is a graded
homomorphism. Furthermore, since I and IH contain the same v's, and
since I is a basic ideal, it follows that if v ∈ E0 \ H, then v /∈ IH and rv /∈ I
for all r ∈ R \ {0}. Thus the Graded Uniqueness Theorem implies that the
quotient map π : LR(E \ H) ∼= LR(E)/IH → LR(E)/I is injective. Hence
I = IH.
The fact that H 7→ IH is injective follows immediately from the fact that
{v ∈ E0 : v ∈ IH} is precisely H, which was obtained in Proposition 7.7.
Thus the correspondence H 7→ IH is bijective. Since H 7→ IH is a bijection
that preserves inclusions, the map H 7→ IH is a poset isomorphism and
hence automatically a lattice isomorphism
Proof of (3):
To see that IX is Morita equivalent to LR(EX ), list the elements of X =
{v1, v2, . . .}, let
Λ :=({1, 2, . . . , X}
{1, 2, . . .}
if X is finite
if X is infinite,
and let en :=Pn
i=1 vi for n ∈ Λ.
If we consider the elements {v : v ∈ H} and {e, e∗ : e ∈ E1 and s(e) ∈ H}
in LR(E), we see that they are a Leavitt EX -family and thus there exists
a homomorphism π : LR(EX ) → LR(E) taking the generators of LR(EX )
to these elements. Since this homomorphism is graded, Theorem 5.3 shows
that π is injective. Hence we may identify LR(EX ) with the subalgebra
spanR{αβ∗ : α, β ∈ E∗
X and r(α) = r(β) ∈ X}
n=1 enLR(E)en.
enLR(E), ψ, φ!
In addition,
of LR(E). With this identification, we see that LR(EX ) =P∞
Moreover, Lemma 7.7 shows that IX =P∞
Xn∈Λ
LR(E)enLR(E),Xn∈Λ
enLR(E)en,Xn∈Λ
LR(E)en,Xn∈Λ
n=1 LR(E)enLR(E).
with ψ(m ⊗ n) = mn and φ(n ⊗ m) = nm is a (surjective) Morita con-
text for the idempotent rings LR(EX) and IX .
It then follows from [13,
Proposition 2.5] and [13, Proposition 2.7] that LR(EX ) and IX are Morita
equivalent.
Corollary 7.10. Let E be a graph, and let R be a commutative ring with
unit. Then every graded basic ideal of LR(E) is selfadjoint.
Using the Cuntz-Krieger Uniqueness Theorem we can characterize those
graphs whose associated Leavitt path algebras have the property that every
basic ideal is a graded ideal.
(cid:3)
20
MARK TOMFORDE
Definition 7.11. We say that a closed path α = e1 . . . en ∈ En is simple if
s(ei) 6= s(e1) for i = 2, 3, . . . , n.
Definition 7.12. A graph E satisfies Condition (K) if every vertex in E0 is
either the base of no closed path or the base of at least two simple closed
paths.
The following proposition is well known. It has been proven in [25, Propo-
sition 1.17] and [4, Theorem 4.5(2),(3)].
Proposition 7.13. If E is a row-finite graph, then E satisfies Condition (K)
if and only if for every saturated hereditary subset H, the graph E \ H of
Definition 7.5 satisfies Condition (L).
Lemma 7.14. If E is the graph consisting of a single simple closed path of
length n; i.e.,
E0 = {v1, . . . , vn} E1 = {e1, . . . en}
s(ei) = vi
for 1 ≤ i ≤ n
r(ei) = vi+1
for 1 ≤ i < n
and
r(en) = v1,
an d R is a commutative ring with unit, then LR(E) ∼= Mn(R[x, x−1]).
The proof of Lemma 7.14 is the same as the proof of [26, Lemma 6.12].
Lemma 7.15. Let R be a commutative ring with unit, let E be a row-finite
graph, and let H be a saturated hereditary subset of E. Then the ideal IH
in LR(E) is a ring with a set of local units.
The proof of Lemma 7.15 is the same as the proof of [26, Lemma 6.14].
Lemma 7.16. Let R be a commutative ring with unit, and let E be a row-
finite graph that contains a simple closed path with no exit. Then LR(E)
contains an ideal that is basic but not graded.
Proof. Let α := e1 . . . en be a simple closed path with no exits in E. If we
let X := {s(ei)}n
i=1, then since α has no exits, X is a hereditary subset of
E0. By Theorem 7.9(3) LR(EX ) is Morita equivalent to the ideal IX in
LR(E). However, EX is the graph which consists of a single closed path,
and thus LR(EX ) ∼= Mn(R[x, x−1]) by Lemma 7.14. Theorem 7.9(1) implies
that LR(E) ∼= Mn(R[x, x−1]) has no proper nontrivial graded ideals. Let
I := hx + 1i be the ideal in R[x, x−1] generated by x + 1. Then any element
of I has the form p(x)(x + 1) for some p(x) ∈ R[x, x−1] and hence has −1
as a root. It follows that for every r ∈ R \ {0} we have that r1 /∈ I. Since
v = 1 in R[x, x−1], it follows that rv /∈ I for all r ∈ R \{0}. Thus I is a basic
ideal. It follows that Mn(I) is a proper nontrivial ideal of Mn(R[x, x−1]),
which is basic but not graded. Because the Morita context described in the
proof of Theorem 7.9(3) gives a lattice isomorphism from ideals of LR(EX )
to ideals of IX that preserves the grading, we may conclude that IX contains
an ideal that is basic but not graded. Since IX has a set of local units by
Lemma 7.15, it follows from Lemma 4.14 that ideals of IX are ideals of
LR(E). Hence LR(E) contains an ideal that is basic but not graded.
(cid:3)
LEAVITT PATH ALGEBRAS WITH COEFFICIENTS IN A COMMUTATIVE RING 21
These results together with the Cuntz-Krieger Uniqueness Theorem give
us the following theorem.
Theorem 7.17. Let R be a commutative ring with unit. If E is a row-
finite graph, then E satisfies Condition (K) if and only if every basic ideal
in LR(E) is graded.
Proof. Suppose that E satisfies Condition (K). If I is a basic ideal of LR(E),
let H := {v : v ∈ I}. Then IH ⊆ I, and we have a canonical surjection
q : LR(E)/IH → LR(E)/I. By Theorem 7.9(2) there exists a canonical
isomorphism φ : LR(E \ H) → LR(E)/IH . Since I is basic, the composition
q ◦ φ : LR(E \ H) → LR(E)/I has the property that (q ◦ φ)(rv) 6= 0 for all
v ∈ E0 and r ∈ R \ {0}. Since E satisfies Condition (K), it follows from
Proposition 7.13 that E \ H satisfies Condition (L). Hence we may apply
Theorem 6.5 to conclude that q ◦ φ is injective. Since φ is an isomorphism,
this implies that q is injective and I = IH. It then follows from Lemma 7.7
that I is graded.
Conversely, suppose that E does not satisfy Condition (K). Then Proposi-
tion 7.13 implies that there exists a saturated hereditary subset H such that
E \ H does not satisfy Condition (L). Thus there exists a closed simple path
with no exit in E\H, and by Lemma 7.16 the algebra LR(E\H) ∼= LR(E)/IH
contains an ideal I that is basic and not graded. If we let q : LR(E) →
LR(E)/IH be the quotient map, then q is graded and q−1(I) is an ideal of
LR(E) that is basic but not graded.
(cid:3)
Corollary 7.18. If E is a row-finite graph that satisfies Condition (K),
then the map H 7→ IH is a lattice isomorphism from the lattice of saturated
hereditary subsets of E onto the lattice of basic ideals of LR(E).
Definition 7.19. The Leavitt path algebra LR(E) is basically simple if the
only basic ideals of LR(E) are {0} and LR(E). (Note that if R = K is a
field, then LK(E) is basically simple if and only if LK(E) is simple.)
Theorem 7.20. Let R be a commutative ring with unit, and let E be a
row-finite graph. The Leavitt path algebra LR(E) is basically simple if and
only if E satisfies both of the following conditions:
(i) The only saturated hereditary subsets of E are ∅ and E0, and
(ii) The graph E satisfies Condition (L).
Proof. Suppose that LR(E) is basically simple. Then the only basic ideals of
LR(E) are {0} and LR(E), both of which are graded. By Theorem 7.17 we
have that E satisfies Condition (K). It then follows from Theorem 7.9(1) and
the fact that LR(E) is basically simple, that the only saturated hereditary
subsets of E are ∅ and E0. Hence (i) holds. In addition, since Condition (K)
implies Condition (L) (cf. Proposition 7.13) we have that (ii) holds.
Conversely, suppose that (i) and (ii) hold. We shall show that E satisfies
Condition (K). Let v be a vertex and let α = e1 . . . en be a closed simple
path based at v. By (ii) we know that α has an exit f ; i.e., there exists
22
MARK TOMFORDE
f ∈ E1 with s(f ) = s(ei) and f 6= ei for some i. If we let H be the set of
vertices in E0 such that there is no path from that vertex to v, then H is
saturated hereditary. By (i) we must have either H = ∅ or H = E0. Since
v /∈ H, we have H = ∅. Hence for every vertex in E0, there is a path from
that vertex to v. Choose a path β ∈ E∗ from r(f ) to v of minimal length.
Then e1 . . . ei−1f β is a simple closed path based at v that is distinct from α.
Hence E satisfies Condition (K). It then follows from Theorem 7.9(1) and
(i) that LR(E) is basically simple.
(cid:3)
Condition (i) and (ii) in the above theorem can be reformulated in a
number of equivalent ways. The equivalence of the statements (2) -- (5) in
Proposition 7.22 are elementary facts about directed graphs (cf. [25, Theo-
rem 1.23] and [2, Proposition 3.2]).
Definition 7.21. A graph E is cofinal if whenever e1e2e3 . . . is an infinite
path in E and v ∈ E0, then there exists a finite path from v to s(ei) for
some i ∈ N.
Proposition 7.22. Let E be a row-finite graph, let R be a commutative
ring with unit, and let LR(E) be the associated Leavitt path algebra. Then
the following are equivalent.
(1) LR(E) is basically simple.
(2) E satisfies Condition (L), and the only saturated hereditary subsets
of E0 are ∅ and E0.
(3) E satisfies Condition (K), and the only saturated hereditary subsets
of E0 are ∅ and E0.
(4) E satisfies Condition (L), E is cofinal, and whenever v is a sink in
E and w ∈ E0 there is a path from w to v.
(5) E satisfies Condition (K), E is cofinal, and whenever v is a sink in
E and w ∈ E0 there is a path from w to v.
Remark 7.23. The techniques used in this section are similar to those used
to analyze ideals of graph C ∗-algebras, which were inspired by the work of
Cuntz and Krieger. In [8] Cuntz and Krieger showed that the Cuntz-Krieger
algebra of an irreducible matrix satisfying Condition (I) is simple.
In [7,
Theorem 2.5] Cuntz showed that for the Cuntz-Krieger algebra of a matrix
satisfying Condition (II) there is a bijective correspondence between the
ideals of the Cuntz-Krieger algebra and the hereditary subsets of a certain
finite partially ordered set associated with the matrix. Subsequently, it was
shown in [15, Theorem 3.5] that the hereditary subsets of this partially
ordered set correspond to the gauge-invariant ideals in any universal Cuntz-
Krieger algebra of a finite {0, 1}-matrix.
In [20, Theorem 6.6], the authors introduced Condition (K) for graphs,
and showed that for a locally finite graph satisfying Condition (K) there
is a bijective correspondence between ideals in the graph C ∗-algebra and
saturated hereditary subsets of the graph. Their proof used groupoid tech-
niques and relied on realizing the graph C ∗-algebra as the C ∗-algebra of a
LEAVITT PATH ALGEBRAS WITH COEFFICIENTS IN A COMMUTATIVE RING 23
groupoid. In [5, Theorem 4.1] it was shown that for C ∗-algebras of row-finite
graphs there is a bijective correspondence between gauge-invariant ideals in
the graph C ∗-algebra and saturated hereditary subsets of the graph, and in
[5, Theorem 4.4] it is proven that when a graph satisfies Condition (K) all
the ideals of the associated C ∗-algebra are gauge invariant. The techniques
used in [5] avoided the use of groupoids, and instead used methods similar
to those used by Cuntz in [7]. In [6, Theorem 3.6] and [9, Theorem 3.5] the
analysis of ideals was extended to non-row-finite graphs, where new phenom-
ena had to be accounted for, and it was shown that gauge-invariant ideals of
the graph C ∗-algebra are in bijective correspondence with admissible pairs;
i.e., pairs consisting of a saturated hereditary set and a subset of break-
ing vertices for this saturated hereditary subset. Furthermore, these results
have been generalized to Cuntz-Pimsner algebras, and it has been shown
that the gauge-invariant ideals in a Cuntz-Pimsner algebra correspond to
admissible pairs of ideals in the coefficient algebra of the Hilbert bimodule
[18, Theorem 8.6].
In the past five years, methods similar to those in [8], [7], [15], [5], [6], and
[9] have been used to analyze the ideal structure of Leavitt path algebras over
fields. It has been shown in [1, Theorem 3.11] that LK(E) is simple if and
only if E satisfies Condition (L) and the only saturated hereditary subsets
of E are ∅ and E0. In [3, Theorem 5.3] it was shown that if E is a row-finite
graph, then the graded ideals of LK(E) are in bijective correspondence with
the saturated hereditary subsets of E. Furthermore, in [26, Theorem 5.7]
it was shown that for a non-row-finite graph E the graded ideals of LK(E)
are in bijective correspondence with the admissible pairs of E. Moreover, it
was proven in [26, Theorem 6.16] that a graph E satisfies Condition (K) if
and only if every ideal in the Leavitt path algebra LK(E) is graded.
8. Tensor products and changing coefficients
If R is a commutative ring with unit and if A and B are R-algebras,
then the tensor product A ⊗R B is an R-module that may also be given
the structure of an R-algebra with a multiplication satisfying (a1 ⊗ b1)(a2 ⊗
b2) = a1a2 ⊗ b1b2. (See [16, Ch.IV, Theorem 7.4] for details on how this
multiplication is obtained.) Furthermore, if R is a commutative ring with
unit that contains a unital subring S, then we may view R as an S-algebra.
If, in addition, A is an S-algebra, then R ⊗S A is an R-algebra with r1(r2 ⊗
a) = r1r2 ⊗ a.
Theorem 8.1. Let R be an algebra over the commutative unital ring S, and
let E be a graph. Then
LR(E) ∼= R ⊗S LS(E)
as R-algebras.
Proof. One can verify that
{1 ⊗ v : v ∈ E0} ∪ {1 ⊗ e, 1 ⊗ e∗ : e ∈ E1}
24
MARK TOMFORDE
is a Leavitt E-family in the R-algebra R ⊗S LS(E), and hence there exists
an R-algebra homomorphism φ : LR(E) → R ⊗S LS(E) with φ(v) = 1 ⊗ v,
φ(e) = 1 ⊗ e, and φ(e∗) = 1 ⊗ e∗. Furthermore, LR(E) is an S-algebra that
contains a Leavitt E-family {v : v ∈ E0} ∪ {e, e∗ : e ∈ E1}. Thus there
exists an S-algebra homomorphism φ : LS(E) → LR(E) with φ(v) = v,
φ(e) = e, and φ(e∗) = e∗. Furthermore, using the universal property of
the tensor product, one can verify that there exists an R-module morphism
ψ : R ⊗S LS(E) → LR(E) with ψ(r ⊗ x) = rφ(x). Finally, one can verify
that ψ is an inverse for φ (simply check on generators), and hence φ is an
R-algebra isomorphism.
(cid:3)
Corollary 8.2. Let E be a graph, and let K be a field. If we view K as a
Z-module, then
LK(E) ∼= K ⊗Z LZ(E).
Furthermore, if K has characteristic p > 0, then we may view K as a Zp-
module and
LK(E) ∼= K ⊗Zp LZp(E).
(Here LZ(E) denotes the Leavitt path ring of characteristic 0 associated to
E, and LZp(E) denotes the Leavitt path ring of characteristic p associated
to E, as described in Definition 3.2.)
Let R be a commutative ring with unit that contains a unital subring S,
and let E be a row-finite graph. For a saturated hereditary subset H of
E, let I S
H denote the ideal in LS(E) generated by {v : v ∈ H} and let I R
H
denote the ideal in LR(E) generated by {v : v ∈ H}. Theorem 7.9 shows
that any graded basic ideal of LS(E) has the form I S
H , and any graded
basic ideal of LR(E) has the form I R
H 7→ I R
H. Thus the map I S
H is a lattice
isomorphism from the lattice of graded basic ideals of LS(E) onto the lattice
of graded basic ideals of LR(E). If we use Theorem 8.1 to identify LR(E)
with R ⊗S LR(E) via the isomorphism described in the proof, then I R
H =
R ⊗ I S
H , and we see that I 7→ R ⊗ I is a map from ideals of LS(E) onto
ideals of LR(E) that restricts to an isomorphism from graded basic ideals
of LS(E) onto graded basic ideal of LR(E). In the special case that S = Z
and R = K is a field (respectively, a field of characteristic p), we see that
all ideals of LK(E) are basic, and hence the map I 7→ K ⊗ I is a lattice
isomorphism from the lattice of graded basic ideals of LZ(E) (respectively,
LZp(E)) onto the lattice of graded ideals of LK (E). This suggests that
properties of graded ideals of LK(E) may derived from properties of graded
basic ideals of LZ(E) and LZn(E).
In the study of Leavitt path algebras over fields, it has frequently been
found that properties of LK(E) depend only on properties of the graph E
and are independent of the particular field K that is chosen. The fact that
LK(E) ∼= K ⊗Z LZ(E) (and LK(E) ∼= K ⊗Zp LZp(E) if char K = p), suggests
that properties of LK(E) may consequences of properties of the Leavitt path
rings LZ(E) and LZp(E). One may speculate that this is the reason many
properties of LK(E) are independent of K.
LEAVITT PATH ALGEBRAS WITH COEFFICIENTS IN A COMMUTATIVE RING 25
References
[1] G. Abrams and G. Aranda-Pino, The Leavitt path algebra of a graph, J. Algebra 293
(2005), 319 -- 334.
[2] G. Abrams and G. Aranda-Pino, The Leavitt path algebras of arbitrary graphs, Hous-
ton J. Math 34 (2008), 423 -- 442.
[3] P. Ara, M. A. Moreno, and E. Pardo, Nonstable K-theory for graph algebras, Al-
gebr. Represent. Theory 10 (2007), 157 -- 178.
[4] G. Aranda-Pino, E. Pardo, and M. Siles-Molina, Exchange Leavitt path algebras and
stable rank, J. Algebra 305 (2006), 912 -- 936.
[5] T. Bates, D. Pask, I. Raeburn and W. Szyma´nski, The C ∗-algebras of row-finite
graphs, New York J. Math. 6 (2000), 307 -- 324.
[6] T. Bates, J. H. Hong, I. Raeburn, and W. Szyma´nski, The ideal structure of C ∗-
algebras of infinite graphs, Illinois J. Math 46 (2002), 1159 -- 1176.
[7] J. Cuntz, A class of C ∗-algebras and topological Markov chains II: reducible chains
and the Ext-functor for C ∗-algebras, Invent. Math. 63 (1981), 25 -- 40.
[8] J. Cuntz and W. Krieger, A class of C ∗-algebras and topological Markov chains,
Invent. Math. 56 (1980), 251 -- 268.
[9] D. Drinen and M. Tomforde, The C ∗-algebras of arbitrary graphs, Rocky Mountain
J. Math. 35 (2005), 105 -- 135.
[10] R. Exel and M. Laca, Cuntz-Krieger algebras for infinite matrices, J. reine angew.
Math. 512 (1999), 119 -- 172.
[11] N. Fowler, P. Muhly, and I. Raeburn, Representations of Cuntz-Pimsner algebras,
Indiana Univ. Math. J. 52 (2003), 569 -- 605.
[12] K. Fuller, On rings whose left modules are direct sums of finitely generated modules,
Proc. Amer. Math. Soc. 54 (1976), 39 -- 44.
[13] J. L. Garc´ıa and J J. Sim´on, Morita equivalence
for
idempotent
rings,
J. Pure Appl. Algebra 76 (1991), 39 -- 56.
[14] K. R. Goodearl, Leavitt path algebras and direct limits, preprint.
[15] A. an Huef and I. Raeburn, The ideal structure of Cuntz-Krieger algebras, Ergodic
Theory Dynam. Systems 17 (1997), 611 -- 624.
[16] T. W. Hungerford, Algebra. Reprint of the 1974 original. Graduate Texts in Mathe-
matics 73. Springer-Verlag, New York-Berlin, 1980. xxiii+502 pp.
[17] T. Katsura, On C ∗-algebras associated with C ∗-correspondences, J. Funct. Anal. 217
(2004), 366 -- 401.
[18] T. Katsura, Ideal structure of C ∗-algebras associated with C ∗-correspondences, Pacific
J. Math. 230 (2007), 107 -- 145.
[19] A. Kumjian, D. Pask, and I. Raeburn, Cuntz-Krieger algebras of directed graphs,
Pacific J. Math. 184 (1998), 161 -- 174.
[20] A. Kumjian, D. Pask, I. Raeburn, and J. Renault, Graphs, groupoids, and Cuntz-
Krieger algebras, J. Funct. Anal. 144 (1997), 505 -- 541.
[21] W. G. Leavitt, Modules without invariant basis number, Proc. Amer. Math. Soc. 8
(1957), 322 -- 328.
[22] G. J. Murphy, C ∗-algebras and Operator Theory. Academic Press, Inc., Boston, MA,
1990. x+286 pp.
[23] I. Raeburn, Graph algebras. CBMS Regional Conference Series in Mathematics, 103,
Published for the Conference Board of the Mathematical Sciences, Washington, DC;
by the American Mathematical Society, Providence, RI, 2005. vi+113 pp.
[24] I. Raeburn and W. Szyma´nski, Cuntz-Krieger algebras of infinite graphs and matrices,
Trans. Amer. Math. Soc. 356 (2004), 39 -- 59.
[25] M. Tomforde, Structure of graph C*-algebras and their generalizations, Chapter in the
book "Graph Algebras: Bridging the gap between analysis and algebra", Eds. Gonzalo
26
MARK TOMFORDE
Aranda Pino, Francesc Perera Dom`enech, and Mercedes Siles Molina, Servicio de
Publicaciones de la Universidad de M´alaga, M´alaga, Spain, 2006.
[26] M. Tomforde, Uniqueness theorems and ideal structure for Leavitt path algebras, J. Al-
gebra 318 (2007), 270 -- 299.
Department of Mathematics, University of Houston, Houston, TX 77204-
3008, USA
E-mail address: [email protected]
|
1605.01202 | 1 | 1605 | 2016-05-04T10:02:12 | A dynamical characterization of diagonal preserving $*$-isomorphisms of graph $C^*$-algebras | [
"math.OA"
] | We characterize when there exists a diagonal preserving $*$-isomorphism between two graph $C^*$-algebras in terms of the dynamics of the boundary path spaces. In particular, we refine the notion of "orbit equivalence" between the boundary path spaces of the directed graphs $E$ and $F$ and show that this is a necessary and sufficient condition for the existence of a diagonal preserving $*$-isomorphism between the graph $C^*$-algebras $C^*(E)$ and $C^*(F)$. | math.OA | math |
A DYNAMICAL CHARACTERIZATION OF DIAGONAL
PRESERVING ∗-ISOMORPHISMS OF GRAPH C∗-ALGEBRAS
SARA E. ARKLINT, SØREN EILERS, AND EFREN RUIZ
Abstract. We characterize when there exists a diagonal preserving ∗-iso-
morphism between two graph C ∗-algebras in terms of the dynamics of the
boundary path spaces. In particular, we refine the notion of "orbit equivalence"
between the boundary path spaces of the directed graphs E and F and show
that this is a necessary and sufficient condition for the existence of a diagonal
preserving ∗-isomorphism between the graph C ∗-algebras C ∗(E) and C ∗(F ).
1. Introduction
The notion of continuous orbit equivalence, pioneered by Matsumoto ([Mat13]),
has proven to be an extremely important vehicle for understanding the relationship
between dynamical systems and the C∗-algebras that they define.
Indeed, this
concept was a key ingredient which allowed Matsumoto and Matui to prove that the
stabilized Cuntz-Krieger algebras become complete invariants for flow equivalence
of irreducible shifts of finite type when considered not just as C∗-algebras, but as
C∗-algebras containing a canonical commutative subalgebra, the diagonal. The key
result in [MM14] thus gave an extremely elegant answer to the question that has
been left open since Rørdam as a key step in the proof of his seminal classification
result [Rør95] showed that two such shift spaces can give the same Cuntz-Krieger
algebra without being flow equivalent: The diagonal is precisely the extra structure
which is needed for the C∗-algebra to remember its underlying shift space.
The success of the approach of Matsumoto and Matui begs the question of
whether or not similar results hold true for more general C∗-algebras such as non-
simple Cuntz-Krieger algebras and (simple or non-simple) graph C∗-algebras, ob-
jects which are currently (see [ERRS16]) giving way to classification in a way paral-
lelling Rørdam's results. In a sweeping generalization of Matsumoto's fundamental
result, Brownlowe, Carlsen and Whittaker in [BCW14] showed that continuous or-
bit equivalence exactly translates to diagonal-preserving isomorphism of the graph
C∗-algebras for any graph with the so-called Condition (L), and proved by example
that this condition is necessary.
In the paper at hand, we will study continuous orbit equivalences preserving even-
tually periodic points and prove that they exactly correspond to diagonal-preserving
isomorphism of graph C∗-algebras. This small adjustment of the notions studied
in [MM14] and [BCW14] thus allow a complete understanding also when Condition
(L) fails. In particular, we prove that the original notion of orbit equivalence cor-
responds to diagonal-preserving isomorphism when the graphs are finite and have
no sinks, a case prominently containing the classical Cuntz-Krieger case.
Our method of proof involves reducing the general case to the Condition (L)
case and hence to the main result of [BCW14] by an elaboration the concept of
"plugging" and "unplugging" graphs introduced in [ERRS16].
Date: June 28, 2021.
2010 Mathematics Subject Classification. Primary: 46L55; Secondary: 46L35, 37B10.
Key words and phrases. Orbit equivalence, diagonal preserving ∗-isomorphism.
1
2
SARA E. ARKLINT, SØREN EILERS, AND EFREN RUIZ
After having circulated an early version of this paper, we were made aware that
our main result had been simultaneously obtained by Carlsen and Winger ([CW16])
by completely different methods.
2. Preliminaries
In this section, we provide the definitions of the objects considered in this paper.
We start with some background on directed graphs, graph C∗-algebras and their
diagonal subalgebras. The definitions of the boundary path space and the graph
groupoid of a directed graph are also provided.
2.1. Graph C∗-algebras and the diagonal subalgebra. A directed graph E =
(E0, E1, r, s) consists of sets E0 and E1 and functions, r, s : E1 → E0 called the
range and source maps, respectively. The elements of E0 are called the vertices of
E and the elements of E1 are called the edges of E.
Assumption 2.1. Throughout the paper, unless stated otherwise, when we say a
graph we mean a directed graph. Moreover, we will only consider graphs such that
the set of vertices and the set of edges are countable sets.
Let E be a graph. A path of length n in E is a finite sequence µ = e1e2 . . . en
with ei ∈ E1 and r(ei) = s(ei+1) for all i = 1, . . . , n − 1. We will regard the vertices
E0 of E as paths of length zero. Denote the set of paths of length n in E by En. Set
E∗ =Sn∈N0
En and set E≥k =Sn≥k En. We extend the range and source maps to
E∗ by r(v) = s(v) = v for v ∈ E0, and s(e1 · · · en) = s(e1) and r(e1 · · · en) = r(en).
A loop in E is an edge e in E such that s(e) = r(e). A cycle in E is a path
µ ∈ E≥1 such that s(µ) = r(µ). A cycle e1e2 · · · en in E is said to have an exit if
there exists an f ∈ E1 such that s(ek) = s(f ) for some k = 1, 2, . . . , n with f 6= ek.
A vertex-simple cycle in E is a cycle µ = e1e2 · · · en such that r(ei) 6= r(ej ) for all
i 6= j. A return path in E is a cycle µ = e1e2 · · · en such that r(ei) 6= r(µ) for all
i = 1, 2, . . . , n − 1.
An infinite path in E is an infinite sequence (en)∞
n=1, denoted e1e2 · · · , such that
ei ∈ E1 and r(ei) = s(ei+1) for all i. The set of infinite paths in E will be denoted
by E∞. If e is a loop and n ∈ N, then en will denote the cycle of length n with
edges equal to e and e∞ will denote the infinite path with edges equal to e. If µ is
a cycle in E, then µ∞ denotes the infinite path µµµ · · · .
Definition 2.2. A graph E is said to have Condition (L) if every cycle in E has
an exit.
Let V, W be subsets of E0, S a subset of E∗ ∪ E∞, and n ∈ N0. Define subsets
of E∗ ∪ E∞, labelled V S, SW , and V SW , by V S = {µ ∈ S s(µ) ∈ V }, SW =
{µ ∈ S r(µ) ∈ W }, and V SW = V S ∩ SW . Note that SW and V SW are subsets
of E∗ since the range map is only defined on E∗. We will write vS if V = {v}.
Similarly for SW and V SW . A vertex v ∈ E0 is called regular if 0 < vE1 < ∞.
Denote the set of regular vertices in E by E0
reg. A singular vertex is a vertex v in
E that is not regular. We denote the set of singular vertices by E0
sing. A vertex v
in E is called a sink if vE1 = 0 and is called an infinite emitter if vE1 = ∞.
Denote the set of sinks by E0
inf . Hence,
E0
sing = E0
We call an infinite path e1e2 · · · in E a tail if s(ei)E1 = {ei} = E1s(ei+1), and
non-wandering if s(ei)E1 = {ei}. If µ ∈ E∗ is a cycle with no exits, then µ∞ is
non-wandering.
Definition 2.3. Let E be a graph. A Cuntz-Krieger E-family in a C∗-algebra A
sink and the set of infinite emitters by E0
consists of a set of mutually orthogonal projections (cid:8)Pv(cid:12)(cid:12) v ∈ E0(cid:9) ⊆ A and a set
of partial isometries(cid:8)Se(cid:12)(cid:12) e ∈ E1(cid:9) ⊆ A satisfying
sink ∪ E0
inf .
The topology on ∂E is given as follows: For µ ∈ E∗, the cylinder set of µ is the set
where µx is the concatenation of paths. For µ ∈ E∗ and a finite subset F of r(µ)E1,
set
∂E = E∞ ⊔(cid:8)µ ∈ E∗(cid:12)(cid:12) r(µ) ∈ E0
sing(cid:9) .
ZE(µ) = {µx ∈ ∂E x ∈ r(µ)∂E} ,
ZE(µ\F ) = ZE(µ)\ [e∈F
ZE(µe)! .
DIAGONAL PRESERVING ∗-ISOMORPHISMS OF GRAPH C ∗-ALGEBRAS
3
(CK1) S∗
(CK2) S∗
(CK3) SeS∗
e Sf = 0 for all e, f ∈ E1 with e 6= f ;
e Se = Pr(e) for all e ∈ E1;
e ≤ Ps(e) for all e ∈ E1; and
e for all v ∈ E0
(CK4) Pv =Pe∈vE1 SeS∗
sv = pv. Then the C∗-subalgebra span(cid:8)sµs∗
reg.
The graph C∗-algebra C∗(E) is the universal C∗-algebra generated by a Cuntz-
Krieger E-family.
If µ = e1e2 · · · en ∈ E≥2, we set sµ = se1 se2 · · · sen and for v ∈ E0, we set
diagonal subalgebra of C∗(E) and is denoted by D(E).
Definition 2.4. Let E and F be graphs. A ∗-isomorphism Φ : C∗(E) → C∗(F ) is
a diagonal preserving ∗-isomorphism if Φ(D(E)) = D(F ).
µ(cid:12)(cid:12) µ ∈ E∗(cid:9) of C∗(E) is called the
2.2. Boundary path space and the graph groupoid of a graph. The defini-
tions given in this section follows that of [BCW14, Section 2.2 and 2.3].
Definition 2.5. Let E be a graph. The boundary path space of E is the space
(When there is no cause for confusion, we will at times omit the subscripts.) The
topology of ∂E is the topology generated by
BE =(cid:8)ZE(µ\F )(cid:12)(cid:12) µ ∈ E∗, F a finite subset of r(µ)E1(cid:9) .
The boundary path space ∂E is a locally compact Hausdorff space with basis
BE and every U ∈ BE is compact and open (see [Web14, Theorem 2.1 and Theo-
rem 2.2]).
The key relationship between ∂E and D(E) is the following theorem.
Theorem 2.6 ([Web14, Theorem 3.7]). There exists a unique homeomorphism hE
from ∂E to the spectrum of D(E) given by
hE(x)(sµs∗
µ) =(1
0
if x ∈ Z(µ)
otherwise.
Lemma 2.7. Let E be a graph and let S be a subset of E0
(i.e., closed and open) subset of ∂E.
Proof. First note that for all v ∈ E0
sink and for all µ ∈ E∗ and F finite subset
of r(µ)E1, v ∈ ZE(µ\F ) if and only if v = µ and F = ∅. In particular, for all
v ∈ E0
again that the only cylinder set that contains a sink v is {v}, we conclude that if
x ∈ ∂E\S, then any cylinder set ZE(µ) containing x will satisfy ZE(µ) ⊆ ∂E\S.
So, ∂E\S is open which implies that S is closed.
(cid:3)
sink, ZE(v) = {v}. Therefore, S = Sv∈S ZE(v), and hence open. Using
sink. Then S is a clopen
If x ∈ ∂E, then set
x =(∞ if x ∈ E∞
∂E since ∂E≥n is equal toSµ∈En Z(µ).
n
For n ∈ N0, set ∂E≥n = {x ∈ ∂E x ≥ n}. Note that ∂E≥n is an open subset of
if x ∈ En for some n ∈ N0.
4
SARA E. ARKLINT, SØREN EILERS, AND EFREN RUIZ
Definition 2.8. Let E be a graph. Define the shift map σE : ∂E≥1 → ∂E on E by
σE(x) =(e2e3 · · ·
r(x)
if x = e1e2 · · · ∈ ∂E≥2
if x ∈ ∂E ∩ E1.
E : ∂E≥n → ∂E will be the n-fold composition of σE with itself and
For n ≥ 1, σn
σ0
E : ∂E → ∂E will be the identity map.
One can check that for all n ∈ N0, σn
E is continuous and moreover, σn
E is a local
homeomorphism.
We now define the graph groupoid of a graph E.
Definition 2.9. Let E be a graph. The graph groupoid GE is defined as follows:
As a set,
GE = {(x, m − n, y) x, y ∈ ∂E with x ≥ m, y ≥ n, and σm
E (x) = σn
E(y)} .
The product is defined by (x, k, y)(w, l, z) = (x, k + l, z) if y = w and undefined
otherwise, and the inverse of (x, k, y) is (y, −k, x). The set of units G(0)
E of GE is
{(x, 0, x) x ∈ ∂E}.
Let m, n ∈ N0, U be an open subset of ∂E≥m such that σm
EV is injective. Suppose σm
E U is injective, and V
E(V ).
E (U ) = σn
be an open subset of ∂E≥n such that σn
Set
ZE(U, m, n, V ) = {(x, m − n, y) ∈ GE x ∈ U, y ∈ V, σm
E (x) = σn
E(y)} .
(When there is no cause for confusion, we will at times omit the subscripts.) Then
GE is a locally compact, Hausdorff, étale topological groupoid with the topology
generated by the basis consisting of sets ZE(U, m, n, V ).
One checks that the map µ ∈ ∂E 7→ (µ, 0, µ) ∈ G(0)
E . We will freely identify G(0)
E is a homeomorphism from ∂E
to G(0)
E with ∂E using this map throughout the paper
without further mention. Thus, we have range and source maps r, s : GE → ∂E
defined by r((x, k, y)) = x and s((x, k, y)) = y.
By [ADR00, Proposition 3.3.5 and 6.1.8], the reduced and universal C∗-algebra
of GE are equal since GE is topologically amenable (see [Yee07, Proposition 6.2]).
We denote this C∗-algebra by C∗(GE).
Theorem 2.10 ([BCW14, Proposition 2.2]). If E is a graph, then there exists
a unique ∗-isomorphism ΦE : C∗(E) → C∗(GE) such that ΦE(pv) = 1Z(v,v) for all
v ∈ E0 and ΦE(se) = 1Z(e,r(e)) for all e ∈ E1, and such that ΦE(D(E)) = C0(G(0)
E ).
3. Orbit equivalence preserving periodic points and pseudogroups
We now define orbit equivalence between graphs E and F preserving periodic
points. When E and F are graphs with finitely many vertices and no sinks, or when
E and F are graphs satisfying Condition (L), then this notion of orbit equivalence
coincides with the notion of orbit equivalence defined in [BCW14, Definiton 3.1].
3.1. Orbit equivalence preserving periodic points. Let E be a graph, and let
∂Eiso denote the set of isolated points in ∂E.
Definition 3.1. Let E be a graph. Then x ∈ ∂E is eventually periodic if x = µν∞
where µ ∈ E∗, ν is a cycle in E, and r(µ) = s(ν).
Note that an eventually periodic point x ∈ ∂E is an isolated point if and only if
x = µν∞ where µ ∈ E∗ and ν ∈ E∗ is a cycle with no exits satisfying r(µ) = s(ν).
sink, and call x even-
tually non-wandering if x = µy where µ ∈ E∗ and y ∈ r(µ)E∞ is non-wandering.
We call x ∈ ∂E eventually a sink if x ∈ E∗ with r(x) ∈ E0
DIAGONAL PRESERVING ∗-ISOMORPHISMS OF GRAPH C ∗-ALGEBRAS
5
The isolated points in ∂E are exactly the points that are eventually sinks or eventu-
ally non-wandering. Clearly, the eventually periodic isolated points are exactly the
eventually periodic eventually non-wandering points. We will refer to the eventually
non-wandering points that are are not eventually periodic as eventually non-periodic
non-wandering.
Definition 3.2. Let E and F be graphs and let κ : ∂E → ∂F be a homeomor-
phism. We say that κ is an orbit equivalence if there exist continuous functions
l, m : ∂E≥1 → N0 and l′, m′ : ∂F ≥1 → N0 such that
(1) σm(x)
(2) σm′(y)
F
(κ(σE(x))) = σl(x)
(κ−1(σF (y))) = σl′(y)
F (κ(x)) for all x ∈ ∂E≥1; and
E (κ−1(y)) for all y ∈ ∂F ≥1.
E
If such a κ exists, we say that E and F are orbit equivalent or there exists an orbit
equivalence between E and F .
If, in addition, κ satisfies
(3) for all x ∈ ∂Eiso, x is eventually periodic if and only if κ(x) is eventually
periodic,
then we say that κ is an orbit equivalence preserving periodic points. If such a κ
exists, then we say that there exists an orbit equivalence between E and F preserving
periodic points.
All eventually non-wandering points in ∂E are eventually periodic if the graph E
has finitely many vertices. Hence if E and F are graphs with finitely many vertices
and no sinks, an orbit equivalence between E and F will automatically preserve
periodic points. Likewise, if E and F are graphs satisfying Condition (L), any orbit
equivalence between E and F will preserve periodic points, as E and F contain no
eventually periodic isolated points.
In general, if E and F are orbit equivalent graphs, there may not exist an or-
bit equivalence between E and F preserving periodic points, as all three types of
isolated points may be interchanged by an orbit equivalence. Consider the graphs
E :
•
and
F :
•
.
By [BCW14, Example 5.2], E and F are orbit equivalent but there is no orbit
equivalence between E and F preserving periodic points, as the isolated point in
∂E is a sink while the isolated point in ∂F is periodic. In Example 3.3 we provide
examples of orbit equivalences that interchange eventually periodic points with
eventually non-periodic non-wandering points, and eventually sinks with eventually
non-periodic non-wondering points.
Example 3.3. Consider the graphs E, F and G:
E :
e2
e1
...
•
•
f
F :
e1
e2
•
•
...
G :
...
•
v
e2
e1
D
D
6
SARA E. ARKLINT, SØREN EILERS, AND EFREN RUIZ
Then all points in ∂E are eventually periodic isolated points, all points in ∂F are
eventually non-periodic non-wandering, and all points in ∂G are eventually sinks.
We now show that E, F , and G are orbit equivalent.
Define κ1 : ∂E → ∂F and κ2 : ∂F → ∂G by κ1(ei · · · e1f ∞) = ei+1ei+2 · · ·
for i ≥ 1 and κ1(f ∞) = e1e2 · · · , and κ2(eiei+1 · · · ) = ei−1 · · · e1 for i ≥ 2 and
κ2(e1e2 · · · ) = v. Clearly, κ1 and κ2 are bijective, and they are continuous and
open since ∂E, ∂F , and ∂G carry the discrete topologies. One readily checks that
m1, l1 : ∂E≥1 → N0 defined by m1(f ∞) = 0, m1(ei · · · e1f ∞) = 1, and l1(x) = 0 for
all x ∈ ∂E≥1 satisfies
σm1(x)
F
for all x ∈ ∂E≥1, and that m′
satisfies
(κ1(σE(x))) = σl1(x)
1 : ∂F ≥1 → N0 defined by m′
(κ1(x))
F
1, l′
1(x)
σm′
E
(κ−1
1 (σF (x))) = σl′
1(x)
E
(κ−1
1 (x))
1(x) = 1 and l′
1(x) = 0
for all x ∈ ∂F ≥1. So κ1 is an orbit equivalence, since the maps m1, l1, m′
1 are
automatically continuous. Similarly, m2, l2 : ∂F ≥1 → N0 defined by m2(x) = 1 and
l2(x) = 0, and m′
2(x) = 0, lets us
conclude that κ2 is an orbit equivalence.
2 : ∂G≥1 → N0 defined by m′
2(x) = 1 and l′
1, l′
2, l′
3.2. The pseudogroup PE and the groupoid of germs of PE. We now recall
the groupoid of germs defined in [Ren08, Section 3].
Let X be a topological space. A homeomorphism h : U → V where U and V
are open subsets of X is called a partial homeomorphism. Under composition and
inverse, the collection of partial homeomorphisms on X is an inverse semigroup.
A pseudogroup on X is a family of partial homeomorphisms of X stable under
composition and inverse.
Let P be a pseudogroup on X. A partial homeomorphism h : U → V is said
to locally belong to P if for all x ∈ U , there exists an open neighborhood W of x
and there exists g ∈ P such that hW = gW . The pseudogroup P is ample if each
partial homeomorphism h : U → V that locally belongs to P must also be element
in P.
Definition 3.4. Let P be a pseudogroup on the topological space X. The groupoid
of germs of P is
GP = {[x, h, y] h ∈ P, y ∈ dom(h), x = h(y)}
where [x, h, y] = [x, g, y] if and only if there exists a neighborhood V of y in X such
that hV = gV .
The range and source maps are given by
r([x, h, y]) = x and s([x, h, y]) = y.
The partially defined product is [x, h, y][y, g, z] = [x, h ◦ g, z], undefined otherwise
and the inverse [x, h, y]−1 = [y, h−1, x]. The groupoid GP is given the topology
given by basic open sets
U(U, h, V ) = {[x, g, y] ∈ GP x ∈ U, y ∈ V }
where U and V are open subsets of X and h ∈ P.
We recall how to construct a pseudogroup from an étale groupoid G. A subset A
of a groupoid G is called a bisection if rA and sA are injective functions. Then the
set of all open bisections S(G) forms an inverse semigroup with composition law
and
AB =nγγ′(cid:12)(cid:12)(cid:12) (γ, γ′) ∈ (A × B) ∩ G(2)o
A−1 =(cid:8)γ−1(cid:12)(cid:12) γ ∈ A(cid:9) .
DIAGONAL PRESERVING ∗-ISOMORPHISMS OF GRAPH C ∗-ALGEBRAS
7
Let A be an open bisection. Then define αA : s(A) → r(A) by αA(s(γ)) = r(γ)
for all γ ∈ A. One checks that αA is a homeomorphism. Then the pseudogroup on
G(0) is
P(G) = {αA A is an open bisection} .
Assumption 3.5. Isomorphisms between topological groupoids are isomorphisms
between groupoids that are also homeomorphisms.
The following proposition follows from [Ren08, Proposition 3.6] and the proofs
of [Ren08, Proposition 3.2 and Corollary 3.3].
Proposition 3.6. Let G be an étale groupoid. Define ϕG : G → GP(G) by
ϕG(γ) = [r(γ), αA, s(γ)]
where A is an open bisection containing γ. Then ϕG is a well-defined surjective ho-
momorphism of groupoids. Moreover, if G is Hausdorff and topologically principal,
then ϕG is an isomorphism.
As an immediate consequence, we get the following corollary.
Corollary 3.7. Let G and H be étale groupoids. Suppose there exists a homeomor-
phism κ : G(0) → H(0) such that
Then there exists an isomorphism ψκ : GP(G) → GP(H) defined by
κ ◦ P(G) ◦ κ−1 :=(cid:8)κ ◦ g ◦ κ−1(cid:12)(cid:12) g ∈ P(G)(cid:9) = P(H).
ψκ([x, g, y]) = [κ(x), κ ◦ g ◦ κ−1, κ(y)].
Consequently, if G and H are Hausdorff and topological principal, then ψκ induces
an isomorphism eψκ : G → H such that eψκG(0) = κ.
Of particular interest to us is the pseudogroup of the étale groupoid GE for a
graph E. So, for a graph E, we denote P(GE) by PE and we call PE the pseudogroup
of E.
In [BCW14, Proposition 3.4], the authors prove that E and F are orbit equivalent
if and only if the pseudogroups of E and F are isomorphic, i.e., there exists a
homeomorphism κ : ∂E → ∂F such that
κ ◦ PE ◦ κ−1
E =(cid:8)κ ◦ g ◦ κ−1(cid:12)(cid:12) g ∈ PE(cid:9) = PF .
They actually prove a stronger statement in the sense that the orbit equivalence
between E and F induces the isomorphism between the pseudogroups of E and F
and vice versa. We record this in the following proposition.
Proposition 3.8 ([BCW14, Proposition 3.4]). Let E and F be graphs and κ from
∂E to ∂F be a homeomorphism. Then κ is an orbit equivalence (preserving periodic
points) if and only if κ ◦ PE ◦ κ−1 = PF (and for all x ∈ ∂Eiso, x is eventually
periodic if and only if κ(x) is eventually periodic).
Proposition 3.9. For all graphs E1, E2, and E3, if κ1 from ∂E1 to ∂E2 and κ2
from ∂E2 to ∂E3 are orbit equivalences (preserving periodic points), then κ2 ◦ κ1
from ∂E1 to ∂E3 is an orbit equivalence (preserving periodic points).
Proof. Let E1, E2, and E3 be graphs. Suppose κ1 : ∂E1 → ∂E2 and κ2 : ∂E2 → ∂E3
are orbit equivalences. By Proposition 3.8, we have κ1 ◦ PE1 ◦ κ−1
1 = PE2 and
κ2 ◦ PE2 ◦ κ−1
2 = PE3. It follows that (κ2 ◦ κ1) ◦ PE1 ◦ (κ2 ◦ κ1)−1 = PE3. Thus, by
Proposition 3.8, κ2 ◦ κ1 : ∂E1 → ∂E3 is an orbit equivalence.
Suppose κ1 and κ2 are orbit equivalences preserving periodic points. Then,
as above (κ2 ◦ κ1) ◦ PE1 ◦ (κ2 ◦ κ1)−1 = PE3. Moreover, since κ1 and κ2 are
orbit equivalences preserving periodic points, for all isolated points x in ∂E1, x is
8
SARA E. ARKLINT, SØREN EILERS, AND EFREN RUIZ
eventually periodic if and only if κ1(x) is eventually periodic if and only if (κ2◦κ1)(x)
is eventually periodic.
(cid:3)
We are now able to prove a stronger version of [BCW14, Theorem 5.1]. This
version will be important for us in the proof of Theorem 5.3.
Theorem 3.10 (cf. [BCW14, Theorem 5.1]). Let E and F be graphs satisfying
Condition (L). Suppose κ : ∂E → ∂F is an orbit equivalence from E to F . Then
there exists an isomorphism ϕ : GE → GF such that ϕ∂E = κ.
Proof. By Proposition 3.8, κ ◦ PE ◦ κ−1 = PF . Since E and F satisfy Condi-
tion (L), by [BCW14, Proposition 2.3], GE and GF are topologically principal. By
Corollary 3.7, there exists an isomorphism ϕ : GE → GF such that ϕ∂E = κ.
(cid:3)
3.3. The unplugged graph and orbit equivalence. For a graph E, let E0
cycle
be the set of vertices of E that is on a vertex-simple cycle with no exits. Suppose
E satisfies the property that if ν is a vertex-simple cycle with no exits, then ν is a
loop. This entails that every vertex v ∈ E0
cycle supports a unique loop ev. Note that
cycle and e ∈ E1 such that s(e) = v, then e = ev. Denote the set of all loops
if v ∈ E0
based at a vertex in E0
cycle. Also
note that if e, f ∈ E1
cycle with s(e) = s(f ) (equivalently, r(e) = r(f )), then e = f .
cycle. Note that s(E1
cycle) = r(E1
cycle) = E0
cycle by E1
Let E be a graph such that all vertex-simple cycles with no exits are loops. Let
the unplugged graph Eg of E be the graph defined by
E0
g = E0
and E1
g = E1\E1
cycle
with the range and source maps of Eg the restrictions of the range and source maps
of E respectively.
Proposition 3.11. Let E be a graph such that each vertex-simple cycle with no
exits is a loop. Define κE : ∂Eg → ∂E by
κE(x) =(xe∞
x,
r(x),
g with r(x) ∈ E0
cycle
if x ∈ E∗
otherwise.
Then κE is an orbit equivalence such that for each isolated point x ∈ ∂Eg, rEg(x) ∈
E0
cycle if and only if κE(x) is an isolated point in ∂E that is eventually periodic.
Proof. A computation shows that κE is a bijection with κ−1
µe∞
all other x. Let µ = e1e2 · · · en ∈ E∗. Suppose µ ∈ E∗
r(µ) for some µ = e1 · · · en with r(µ) ∈ E0
g. Then
cycle and en 6= er(µ), and κ−1
E (x) = µ when x =
E (x) = x for
κ−1
E (ZE(µ)) = {µx ∈ ∂Eg x ∈ r(µ)∂Eg} = ZEg (µ)
which is open. Suppose µ /∈ E∗
and ej /∈ E1
ZE(µ) = ZE(e1 · · · ei0−1) = {e1 · · · ei0−1e∞
Hence,
cycle
cycle for all j < i0. Then ei0 = ek for all i0 ≤ k ≤ n. Therefore,
i0 }, where e1 · · · ei0−1 = s(µ) if i0 = 1.
g. Let i0 ∈ N with 1 ≤ i0 ≤ n such that ei0 ∈ E1
κ−1
E (ZE(µ)) = κ−1
E ({e1 · · · ei0−1e∞
i0 }) = {e1 · · · ei0−1} = ZEg(e1 · · · ei0−1)
which is open. Let F be a finite nonempty subset of r(µ)E1.
then F = {er(µ)} so ZE(µ\F ) = ∅, hence κ−1
If r(µ) ∈ E0
E (ZE(µ\F )) is trivially open.
cycle
If
DIAGONAL PRESERVING ∗-ISOMORPHISMS OF GRAPH C ∗-ALGEBRAS
9
r(µ) /∈ E0
cycle then µ ∈ E∗
g and F ⊆ r(µ)E1
g. Hence
κ−1
E (ZE(µ\F )) = κ−1
E (ZE(µ))\ [e∈F
= ZEg(µ)\ [e∈F
E (ZE(µe))!
κ−1
ZEg(µe)! = ZEg(µ\F )
which is open. We have just shown that κE is continuous.
Let µ ∈ E∗
g. Then κE(ZEg(µ)) = ZE(µ). Let F be a finite subset of r(µ)E1
g.
Then κE(ZEg(µe)) = ZE(µe) for all e ∈ F , hence
κE(ZEg (µ\F )) = ZE(µ)\ [e∈F
ZE(µe)! = ZE(µ\F ),
which is open. Hence, κE is an open map. Therefore, κE is a homeomorphism.
Define m : ∂E≥1 → N0 by
m(y) =(0,
if y ∈ κE(E0
cycle)
1, otherwise.
cycle ⊆ (Eg)0
Since E0
homeomorphism, κE(E0
sink, by Lemma 2.7, E0
cycle is clopen in ∂Eg. Since κE is a
cycle) is clopen in ∂E, so m is continuous.
A computation shows that
σE(κE(x)) = κE(σEg(x))
and σm(y)
Eg
(κ−1
E (y)) = κ−1
E (σE(y)).
Therefore, κE is an orbit equivalence. The last part of the proposition follows
immediately from the construction of κE.
(cid:3)
The next lemma shows that we can adjust an orbit equivalence so that sinks
are sent to sinks. This will be used in the proof of Theorem 5.3 to construct a
diagonal preserving ∗-isomorphism. Note that if µ ∈ E∗ such that r(µ) is a sink,
then µ ∈ ∂Eiso.
Lemma 3.12. Let E be a graph and let F be a subset of E≥1 such that
(1) for all µ ∈ F , r(µ) ∈ E0
(2) for all µ, ν ∈ F , r(µ) = r(ν) if and only if µ = ν, and
(3) F is a closed subset of ∂E.
sink,
Define κ : ∂E → ∂E by
κ(x) =
x
r(x)
µ
if x /∈ F ∪ r(F )
if x ∈ F
if x = r(µ) for some µ ∈ F .
Then κ is an orbit equivalence.
Proof. We must show that κ is a homeomorphism and there exist continuous func-
tions l, m : ∂E≥1 → N0 and l′, m′ : ∂E≥1 → N0 such that
σm(x)
E
(κ(σE(x))) = σl(x)
E (κ(x))
and σm′(x)
E
(κ−1(σE(x))) = σl′(x)
E (κ−1(x))
for all x ∈ ∂E≥1. A computation shows that κ ◦ κ = id. Hence, it is enough to
show that κ is continuous and there exist l, m : ∂E≥1 → N0 such that
σm(x)
E
(κ(σE(x))) = σl(x)
E (κ(x))
for all x ∈ ∂E≥1.
Since r(F ) ⊆ E0
sink, by Lemma 2.7, r(F ) is a clopen subset of ∂E. Since
F ⊆ ∂Eiso, F is open in ∂E. By Assumption (3), F is closed in ∂E. As F ,
10
SARA E. ARKLINT, SØREN EILERS, AND EFREN RUIZ
r(F ) and thereby also ∂E\ (F ∪ r(F )) are clopen, it suffices to check for continuity
on the three sets individually. Since F and r(F ) consist of isolated points and
thereby carry the discrete subspace topology, κ is automatically continuous on F
and r(F ). Since κ restricts to the identity on ∂E\ (F ∪ r(F )), κ is also continuous
on ∂E\ (F ∪ r(F )). So κ is continuous and hence a homeomorphism.
We now produce continuous functions l, m : ∂E≥1 → N0 such that
σm(x)
E
(κ(σE(x))) = σl(x)
E (κ(x))
for all x ∈ ∂E≥1. Note that F ⊆ E∗ (paths of finite length). For each v ∈ r(F ), we
will denote the unique element in F with range v by µv. Note that if x ∈ σ−1
E (F ),
then x ≥ 2 since σE(x) ∈ F and F ⊆ ∂E≥1.
Define l, m : ∂E≥1 → N0 by
and
E (F ),
0
x
1
if x ∈ F ,
if x ∈ σ−1
otherwise,
l(x) =
m(x) =
E (F )(cid:1), l is also continuous on ∂E≥1\(cid:0)F ∪ σ−1
µr(x)
x − 1 if x ∈ F ∩ ∂E≥2,
0
if x ∈ E1r(F ),
otherwise.
We first show that l is continuous. Since F is clopen in ∂E≥1 and σE is continu-
E (F ) consist
E (F ). As l is constant
ous, we have that σ−1
of isolated points, l is automatically continuous on F and σ−1
E (F ) is a clopen subset of ∂E≥1. Since F and σ−1
on ∂E≥1\(cid:0)F ∪ σ−1
on ∂E≥1.
E (F )(cid:1) and thereby
For continuity of m, we first note that F ∩ ∂E≥2 and E1r(F ) are both clopen in
∂E≥1 as F and ∂E≥2 are, and as E1r(F ) = σ−1
E (F ) ∩ E1 with ∂E≥1 ∩ E1 clopen
in ∂E≥1. Since E1r(F ) and F ∩ ∂E≥2 consist of isolated points, they carry the
discrete subspace topology, so m is continuous on E1r(F ) and F ∩ ∂E≥2. As m is
constant on the complement ∂E≥1\(E1r(F ) ∪ (F ∩ ∂E≥2)) we conclude that m is
continuous.
Let x ∈ ∂E≥1. Suppose x ∈ E1r(F ). Then
σm(x)
E
(κ(σE(x))) = σ
µr(x)
E
(κ(r(x))) = σ
µr(x)
E
(µr(x)) = r(x)
and
E (κ(x)) =(κ(x)
σl(x)
σE(κ(x))
if x ∈ F
if x ∈ E1r(F )\F
= r(x).
Suppose x ∈ F ∩ ∂E≥2. Then σE(x) /∈ F since r(σE(x)) = r(x) with x ∈ F , and
σE(x) /∈ r(F ) as σE(x) ∈ ∂E≥1, hence κ(σE(x)) = σE(x). Thus
and
σm(x)
E
(κ(σE(x))) = σx−1
E
(σE(x)) = r(x)
σl(x)
E (κ(x)) = σ0
E(r(x)) = r(x).
DIAGONAL PRESERVING ∗-ISOMORPHISMS OF GRAPH C ∗-ALGEBRAS
11
Suppose x /∈ E1r(F ) and x /∈ F ∩ ∂E≥2. Then x /∈ F and σE(x) /∈ r(F ), so
and κ(x) = x, hence
if x ∈ σ−1
if x /∈ σ−1
E (F )
E (F )
if x ∈ σ−1
if x /∈ σ−1
E (F )
E (F ),
if x ∈ σ−1
if x /∈ σ−1
E (F )
E (F )
if x ∈ σ−1
if x /∈ σ−1
E (F )
E (F ).
σm(x)
E
(κ(σE(x))) = κ(σE(x))
σE(x)
σE(x)
=(r(σE (x))
=(r(x)
E (κ(x))) =(σx
=(r(x)
E (x)
σE(x)
σE(x)
σl(x)
We have just shown that l, m : ∂E≥1 → N0 are continuous functions and
σm(x)
E
(κ(σE(x))) = σl(x)
E (κ(x))
for all x ∈ ∂E≥1. We conclude that κ is an orbit equivalence.
(cid:3)
4. The extended Weyl groupoid
In [BCW14], the authors prove that a diagonal preserving ∗-isomorphism be-
tween C∗(E) and C∗(F ) implies that E and F are orbit equivalent. In this section,
we point out that their arguments even prove that the existence of a diagonal pre-
serving ∗-isomorphism between C∗(E) and C∗(F ) implies the existence of an orbit
equivalence between E and F preserving periodic points. The arguments are actu-
ally contained in [BCW14, Section 4]. For the convenience of the reader, we provide
the arguments here.
First we need to recall the extended Weyl groupoid of (C∗(E), D(E)) as defined
in [BCW14, Section 4].
Definition 4.1. Let E be a graph. The normalizer of D(E) is defined to be the
set
N (D(E)) = {n ∈ C∗(E) ndn∗, n∗dn ∈ D(E) for all d ∈ D(E)} .
By [Ren08, Lemma 4.6], for all n ∈ N (D(E)), nn∗ and n∗n are elements in
D(E). Therefore, we may define for n ∈ N (D(E), the sets
and
dom(n) = {x ∈ ∂E hE(x)(n∗n) > 0}
ran(n) = {x ∈ ∂E hE(x)(nn∗) > 0} .
By [Ren08, Proposition 4.7], for each n ∈ N (D(E)), there exists a unique homeo-
morphism αn : dom(n) → ran(n) such that for all d ∈ D(E),
hE(x)(n∗dn) = hE(αn(x))(d)hE (x)(n∗n).
For each x ∈ ∂Eiso, we let px denote the unique element in D(E) satisfying
hE(y)(px) = 1 if y = x and zero otherwise, i.e., px is the unique element in D(E)
corresponding to the characteristic function χ{x} ∈ C0(∂E) under the canonical
∗-isomorphisms D(E) ∼= C0(Spec(D(E))) ∼= C0(∂E). By [BCW14, Lemma 4.3],
if x ∈ ∂Eiso, then pxD(E)px is either isomorphic to C (when x is not eventually
periodic) or C(T) (when x is eventually periodic).
12
SARA E. ARKLINT, SØREN EILERS, AND EFREN RUIZ
By [BCW14, Lemma 4.4], for each x ∈ ∂Eiso, n1, n2 ∈ N (D(E)) such that
x ∈ dom(n1) ∩ dom(n2) and αn1 (x) = αn2(x),
U(x,n1,n2) = (hE(x)(n∗
1n1n∗
2n2))−1/2 pxn∗
1n2px
is a unitary in pxC∗(E)px.
We define an equivalence relation ∼ on {(n, x) n ∈ N (D(E)), x ∈ dom(n)} by
(n1, x1) ∼ (n2, x2) if either
(1) x1 = x2 ∈ ∂Eiso, αn1 (x1) = αn2 (x2), and [U(x1,n1,n2)] = 0 in K1(px1C∗(E)px1 )
(2) x1 = x2 /∈ ∂Eiso and there is an open set V such that x1 ∈ V ⊆ dom(n1) ∩
dom(n2) and αn1 (y) = αn2 (y) for all y ∈ V .
It is shown in [BCW14, Proposition 4.6] that this relation is in fact an equivalence
relation.
Let G(C ∗(E),D(E)) be the collection of equivalence classes. Define a partially
defined product by
[(n1, x1)] · [(n2, x2)] = [(n1n2, x2)]
if αn2 (x2) = x1
and undefined otherwise, define an inverse map by
[(n, x)]−1 = [(n∗, αn(x))].
By [BCW14, Proposition 4.7 and Proposition 4.8], G(C ∗(E),D(E)) is a groupoid and
is a topological groupoid with the topology generated by
{{[(n, x)] x ∈ dom(n)} n ∈ N (D(E))} .
Moreover, by [BCW14, Proposition 4.8 and its proof], the map ϕE from GE to
G(C ∗(E),D(E)) defined by
ϕE((x, k, y)) = [(sµs∗
ν, y)]
morphism.
where x = µz, y = νz, k = µ − ν for some µ, ν ∈ E∗ and z ∈ ∂E, and
(cid:8)pv, se(cid:12)(cid:12) v ∈ E0, e ∈ E1(cid:9) be a Cuntz-Krieger E-family generating C∗(E), is an iso-
Proposition 4.2 (cf. [BCW14, Proposition 4.11]). Let E and F be graphs. Suppose
there exists a diagonal preserving ∗-isomorphism Ψ : C∗(E) → C∗(F ). Then there
exists an isomorphism β : GE → GF and a homeomorphism κ : ∂E → ∂F such that
β((µ, 0, µ)) = (κ(µ), 0, κ(µ)) and for all x ∈ ∂Eiso, x is eventually periodic if and
only if κ(x) is eventually periodic.
Proof. Let(cid:8)pv, se(cid:12)(cid:12) v ∈ E0, e ∈ E1(cid:9) be a Cuntz-Krieger E-family generating C∗(E).
Since Ψ : C∗(E) → C∗(F ) is a ∗-isomorphism such that Ψ(D(E)) = D(F ), there
exists a homeomorphism κ : ∂E → ∂F such that hE(x)(f ) = hF (κ(x))Ψ(f ) for all
f ∈ D(E) and the map λ : G(C ∗(E),D(E)) → G(C ∗(F ),D(E)) given by
λ([(n, x)]) = [(Ψ(x), κ(x))]
is an isomorphism. Now, β = ϕ−1
F ◦ λ ◦ ϕE : GE → GF is an isomorphism.
We claim that β((µ, 0, µ)) = ((κ(µ), 0, κ(µ)) for all µ ∈ ∂E. Let µ ∈ ∂E. Then
ϕF (β((µ, 0, µ))) = λ([(sµs∗
µ, µ)]) = [(Ψ(sµs∗
µ), κ(µ))].
µ), κ(µ))]. Since ϕF ((ν, 0, ν)) = [(sν s∗
Since ϕF is an isomorphism, there exists (ν, 0, ν) ∈ G(0)
[(Ψ(sµs∗
ϕ−1
F ([(Ψ(sµs∗
We will now show that for all x ∈ ∂Eiso, x is eventually periodic if and only if
κ(x) is eventually periodic. To do this, we first show that for all x ∈ ∂Eiso, then
F such that ϕF ((ν, 0, ν)) =
ν, ν)], we have that ν = κ(µ). Hence,
µ), κ(µ))]) = (κ(µ), 0, κ(µ)), thus proving the claim.
DIAGONAL PRESERVING ∗-ISOMORPHISMS OF GRAPH C ∗-ALGEBRAS
13
Ψ(px) = pκ(x). Let x ∈ ∂Eiso. Suppose y ∈ ∂F . Then
hF (y)(Ψ(px)) = hE(κ−1(y))(px)
0
=(1
=(1
0
if κ−1(y) = x
otherwise
if κ(x) = y
otherwise
= hF (y)(pκ(x)).
Therefore, by the uniqueness of pκ(x), we have that ψ(px) = pκ(x). Hence, proving
the claim.
Let x ∈ ∂Eiso. Then
pκ(x)D(F )pκ(x) = Ψ(pxD(E)px),
and hence pκ(x)D(F )pκ(x)
only if pxD(E)px ∼= C(T) if and only if pκ(x)D(F )pκ(x)
is eventually periodic.
∼= pxD(E)px. Therefore, x is eventually periodic if and
∼= C(T) if and only if κ(x)
(cid:3)
Theorem 4.3. Let E and F be graphs. Suppose there exists a diagonal preserving
∗-isomorphism Ψ : C∗(E) → C∗(F ). Then there exists an orbit equivalence between
E and F preserving periodic points.
Proof. By Proposition 4.2, there exist an isomorphism β : GE → GF and a homeo-
morphism κ : ∂E → ∂F such that β((µ, 0, µ)) = (κ(µ), 0, κ(µ)) and for all x ∈ ∂Eiso,
x is eventually periodic if and only if κ(x) is eventually periodic. One can check
that κ ◦ PE ◦ κ−1 = PF . By Proposition 3.8, κ is an orbit equivalence between E
and F preserving periodic points.
(cid:3)
5. Main result
Let E be a graph and let S be a subset of E0
sink. Define Ef,S to be the graph
with vertices E0
f,S = E0 and edges
E1
f,S = E1 ⊔ {e(v) v ∈ S}
where the range and source maps extends the range and source maps of E respec-
tively, and rEf,S (e(v)) = sEf,S (e(v)) = v for all v ∈ S.
Proposition 5.1. Let E and F be graphs, S1 be a nonempty subset of E0
sink, S2 be a
nonempty subset of F 0
sink. Suppose there exist a bijection w : S1 → S2 and a diagonal
preserving ∗-isomorphism Φ : C∗(E) → C∗(F ) such that Φ(Pv) = Qw(v) for all
Then there exists a diagonal preserving ∗-isomorphism Ψ : C∗(Ef,S1 ) → C∗(Ff,S2 ).
v ∈ S1, where (cid:8)Pv, Se(cid:12)(cid:12) v ∈ E0, e ∈ E1(cid:9) is a Cuntz-Krieger E-family generating
C∗(E) and(cid:8)Qv, Te(cid:12)(cid:12) v ∈ F 0, e ∈ F(cid:9) is a Cuntz-Krieger F -family generating C∗(F ).
Proof. Let (cid:8)pv, se(cid:12)(cid:12) v ∈ E0
f,S1(cid:9) be a Cuntz-Krieger Ef,S1 -family gen-
erating C∗(Ef,S1 ) and (cid:8)qv, te(cid:12)(cid:12) v ∈ F 0
f,S2(cid:9) be a Cuntz-Krieger Ff,S2 -
family generating C∗(Ff,S2 ). Clearly,
f,S2 , e ∈ F 0
f,S1, e ∈ E1
(cid:8)pv, se(cid:12)(cid:12) v ∈ E0, e ∈ E1(cid:9)
is a Cuntz-Krieger E-family in C∗(Ef,S1 ). Therefore, there exists a ∗-homo-
morphism Φ1 : C∗(E) → C∗(Ef,S1 ) such that Φ1(Pv) = pv and Φ1(Se) = se for
all v ∈ E0 and e ∈ E1. For all µ ∈ E∗, Φ1(Sµ) = sµ. Hence, for µ a vertex-
simple cycle in E with no exits, µ is a vertex-simple cycle in Ef,S1 with no exits,
and so Φ1(Sµ) = sµ is a partial unitary with spectrum equal to T. And since
Φ1(Pv) 6= 0 for all v ∈ E0, by [Szy02, Theorem 1.2], we have that Φ1 is injective.
14
SARA E. ARKLINT, SØREN EILERS, AND EFREN RUIZ
Similarly, there exists an injective ∗-homomorphism Φ2 : C∗(F ) → C∗(Ff,S2 ) such
that Φ2(Qv) = qv and Φ2(Te) = te for all v ∈ F 0 and e ∈ F 1.
which we denote by D(A) and
Set A = Φ1(C∗(E)) and B = Φ2(C∗(F )). Note that A is the C∗-subalgebra of
C∗(Ef,S1 ) generated by (cid:8)pv, se(cid:12)(cid:12) v ∈ E0, e ∈ E1(cid:9) and B is the C∗-subalgebra of
C∗(Ff,S1 ) generated by(cid:8)qv, te(cid:12)(cid:12) v ∈ F 0, e ∈ F 1(cid:9). Moreover,
µ(cid:12)(cid:12) µ ∈ E∗(cid:9)
µ(cid:12)(cid:12) µ ∈ F ∗(cid:9)
which we denote by D(B). Therefore, Φ induces a ∗-isomorphism eΦ : A → B such
that eΦ(D(A)) = D(B) and eΦ(pv) = qw(v) for all v ∈ S1.
Φ1(D(E)) = span(cid:8)sµs∗
Φ2(D(F )) = span(cid:8)tµt∗
f,S1 , set pv =eΦ(pv) and for e ∈ E1
For v ∈ E0
f,S1, set
if e ∈ E1
if e = e(v) for some v ∈ S1.
se =(eΦ(se)
One can check that(cid:8)pv, se(cid:12)(cid:12) v ∈ E0
te(w(v))
f,S1 , e ∈ E1
f,S1(cid:9) is a Cuntz-Krieger Ef,S1-family
in C∗(Ff,S2 ). Hence, there exists a ∗-homomorphism Ψ : C∗(Ef,S1 ) → C∗(Ff,S2)
such that Ψ(pv) = pv and Ψ(se) = se for all v ∈ E0
f,S1 . We claim
that Ψ is a ∗-isomorphism.
f,S1 and e ∈ E1
First we show that Ψ is surjective. Let w ∈ F 0
f,S2 . Since F 0
f,S2 = F 0, we have
that w ∈ F 0. Thus, qw is in the image of eΦ and hence qw is in the image of Ψ.
f,S2. Suppose e ∈ F 1. Then te is in the image of eΦ which implies that
Let e ∈ F 1
te is in the image of Ψ. Suppose e = e(z) for some z ∈ S2. Since w : S1 → S2 is a
bijection, z = w(v) for some v ∈ S1. Hence, Ψ(se(v)) = te(w(v)) = te(z) = te. Thus,
Ψ is surjective.
To show that Ψ is injective, we will first show that for every vertex-simple cycle
µ in Ef,S1 with no exits, Ψ(sµ) is a partial unitary with spectrum equal to T. Let
µ be a vertex-simple cycle in Ef,S1 with no exits. Note from the construction of
Ef,S1, µ is either a vertex-simple cycle in E with no exits or µ = e(v) for some
v ∈ S1. Suppose µ is a vertex-simple cycle in E with no exits. Then Ψ(sµ) =eΦ(sµ),
and since eΦ is a ∗-isomorphism, Ψ(sµ) =eΦ(sµ) is a partial unitary with spectrum
equal to T. Suppose µ = e(v) for some v ∈ S1. Then Ψ(sµ) = te(w(v)) which is a
partial unitary with spectrum equal to T. Since Ψ(pv) 6= 0 for all v ∈ E0
f,S1, by
[Szy02, Theorem 1.2] Ψ is injective.
We have just shown that Ψ is a ∗-isomorphism. We are left with showing that
Ψ(D(Ef,S1 )) = D(Ff,S2 ). Noting that sµe(v)n s∗
µ for all v ∈ S1 and µ ∈
E∗ with r(µ) = v, we have that D(Ef,S1 ) = D(A). Similarly, D(Ff,S2 ) = D(B).
(cid:3)
It is now clear that Ψ(D(Ef,S1 )) = D(Ff,S2 ) since eΦ(D(A)) = D(B).
To use the above result to prove that an orbit equivalence preserving periodic
points implies diagonal preserving isomorphism, we must show that we may reduce
the problem to graphs satisfying the property that all vertex-simple cycles with no
exits are loops. These are the loops that we will unplug, and then plug again.
µe(v)n = sµs∗
Proposition 5.2. Let E be a graph. Then there exists a graph F such that each
vertex-simple cycle in F with no exits is a loop and there exists a diagonal preserving
∗-isomorphism Ψ : C∗(F ) → C∗(E).
Proof. For each cycle µ, let Vµ be the vertices that support the cycle µ.
If µ
and ν are vertex-simple cycles in E with no exits, then Vµ = Vν if and only if
Vµ ∩ Vν 6= ∅. Define a relation ≈ on the vertex-simple cycles in E with no exits by
DIAGONAL PRESERVING ∗-ISOMORPHISMS OF GRAPH C ∗-ALGEBRAS
15
µ ≈ ν if Vµ = Vν . Clearly, ≈ is an equivalence relation, and we may pick a set F
of representatives of the equivalence classes. Then F ⊆ E∗ such that
(a) For each µ ∈ F , µ is a vertex-simple cycle with no exits;
(b) For each µ, ν ∈ F , Vµ ∩ Vν 6= ∅ if and only if µ = ν; and
(c) For each vertex-simple cycle µ in E with no exits, there exists ν ∈ F such
that Vµ = Vν .
sF (ν) = sE(ν) for all ν ∈ F .
It is clear from the construction of F , that each
vertex-simple cycle in F with no exits is a loop. In fact, the vertex-simple cycles
with no exits are {ν ν ∈ F}.
Set S = (cid:8)e ∈ E1(cid:12)(cid:12) s(e) = s(ν) for some ν ∈ F(cid:9). Define F by F 0 = E0, F 1 =
(cid:0)E1\S(cid:1) ⊔ {ν ν ∈ F}, and rF E1\S = rEE1\S, sF E1\S = sEE1\S, rF (ν) =
Let(cid:8)pv, se(cid:12)(cid:12) v ∈ E0, e ∈ E1(cid:9) be a generating Cuntz-Krieger E-family in C∗(E) ,
and let(cid:8)qv, te(cid:12)(cid:12) v ∈ F 0, e ∈ F 1(cid:9) be a generating Cuntz-Krieger F -family in C∗(F ).
shows that(cid:8)Qv, Te(cid:12)(cid:12) v ∈ F 0, e ∈ F 1(cid:9) is a Cuntz-Krieger F -family in C∗(E). Hence,
We now define a Cuntz-Krieger F -family in C∗(E). For each v ∈ F 0, set Qv = pv
and for each e ∈ E1\S, set Te = se. For ν ∈ F , set Tν = sν. A computation
there exists a ∗-homomorphism Ψ : C∗(F ) → C∗(E) such that Ψ(qv) = Qv and
Ψ(te) = Te for all v ∈ F 0, e ∈ F 1.
Since the only vertex-simple cycles of F with no exits are the ν's and Ψ(tν ) = sν
where ν is a vertex-simple cycle with no exits, Ψ(tν) is a partial unitary with
spectrum equal to T. Since Ψ(qv) = pv 6= 0 for all v ∈ F 0, by [Szy02, Theorem 1.2],
Ψ is injective.
We now show that Ψ is surjective. Clearly, pv, se ∈ im Ψ, for all v ∈ E0 and
e ∈ E1\S. For each ν ∈ F , s−1
E (sE(ν)) = {eν} for some eν ∈ E1 since ν is a
vertex-simple cycle with no exits. So S = {eν ν ∈ F}. Let ν ∈ F . If ν = 1 then
ν = eν so seν = Tν ∈ im Ψ. Assume ν ≥ 2. Then ν = eνµ where µ = e1 · · · eν−1
with each ei not an element of S. Hence, sei = Tei ∈ im Ψ for each i. Since ν is a
vertex-simple cycle with no exits, we have that seis∗
ei = psE (ei). We now have
seν = seν psE (e1) = seν se1 s∗
e1
e1 = seν se1 se2 s∗
e2 s∗
e1
= seν se1 psE (e2)s∗
...
= seν sµs∗
µ
= sνs∗
µ
= Ψ(tνtµ) ∈ ψ(C∗(F )).
Therefore, Ψ is surjective, and hence Ψ is a ∗-isomorphism.
It is clear that Ψ(D(F )) ⊆ D(E). Let µ ∈ E∗. Note that sµs∗
ν for a path
ν ∈ E∗ where all edges of ν are not in S. Indeed, if µ = νe1 · · · en with e1 ∈ S,
e2, . . . , en ∈ E1, and ν ∈ E∗ where all edges of ν are not in S, then s(ei)E1 = {ei}
for all i since e1 ∈ S. So pr(ν) = se1···en s∗
ν. Therefore,
Ψ(tνt∗
µ. Hence, Ψ(D(F )) ⊇ D(E), so Ψ is diagonal preserving. (cid:3)
e1···en , hence sµs∗
ν) = sν s∗
ν = sµs∗
µ = sνs∗
µ = sν s∗
We provide an example to illustrate the construction in the proof of Proposi-
tion 5.2. Let E denote the left-most graph below. Then E contains four vertex-
simple cycles with no exits. Let F = {ν} with ν = e3e4e1e2. Then S = {e3} and
16
SARA E. ARKLINT, SØREN EILERS, AND EFREN RUIZ
F is the right-most graph below.
•
•
•
e4
?
_❃❃❃❃❃❃❃❃
e3
E :
•
•
F :
•
•
•
e4
?
ν
•
e1
❃❃❃❃❃❃❃❃
e2
e1
❃❃❃❃❃❃❃❃
e2
•
.
We are now ready to prove our main result.
Theorem 5.3. Let E and E′ be graphs. Then the following are equivalent.
(1) There exists a diagonal preserving ∗-isomorphism Ψ : C∗(E) → C∗(E′).
(2) GE and GE ′ are isomorphic.
(3) There exists a homeomorphism κ : ∂E → ∂E′ such that κ ◦ PE ◦ κ−1 = PE ′
and for each x ∈ ∂Eiso, x is eventually periodic if and only if κ(x) is
eventually periodic.
(4) There exists an orbit equivalence from E to E′ preserving periodic points.
Proof. The equivalence of (1) and (2) is [BCW14, Theorem 5.1 (1) ⇐⇒ (2)]. The
equivalence of (3) and (4) follows from Proposition 3.8. We are left to showing that
(1) and (4) are equivalent.
1 such that all vertex-simple cycles in E1 and E′
By Theorem 4.3, we get that (1) implies (4). We now prove (4) implies (1).
We will first show that we may assume that all vertex-simple cycles in E and
E′ with no exits are loops. Indeed, by Proposition 5.2, there are graphs E1 and
E′
1 with no exits are loops, and
diagonal preserving ∗-isomorphisms from C∗(E) to C∗(E1) and from C∗(E′) to
C∗(E′
1). By Theorem 4.3, there exists an orbit equivalence from E to E1 preserving
periodic points and there exists an orbit equivalence from E′ to E′
1 preserving
periodic points. By Proposition 3.9, there exists an orbit equivalence from E1 to
E′
1 preserving periodic points if and only if there is an orbit equivalence from E
to E′ preserving periodic points. And clearly, there is a diagonal preserving ∗-iso-
1) if and only if there is one from C∗(E) to C∗(E′).
morphism from C∗(E1) to C∗(E′
This establishes the claim.
Assume all vertex-simple cycles in E and E′ with no exits are loops, and that
there exists an orbit equivalence β from E to E′ preserving periodic points. Let
g → ∂E′ be the orbit equivalences provided in Propo-
κE : ∂Eg → ∂E and κE ′ : ∂E′
sition 3.11. Then by Propostion 3.9, λ = κ−1
g is an orbit equiv-
E ′ ◦β◦κE : ∂Eg → ∂E′
alence. Let V =nv ∈ E0
cycle(cid:12)(cid:12)(cid:12) λ(v) ∈ (∂E′
g)≥1o. Set F = λ(V ). By Lemma 2.7, V
g.
Let v ∈ E0
is closed in ∂Eg. Since λ is a homeomorphism, we have that F is closed in ∂E′
Hence, F satisfies (3) in Lemma 3.12.
cycle. Then, κE(v) = e∞
v . Since β is an orbit equivalence preserving
cycle and µ ∈ (E′)∗ with no edges equal to
periodic points, there exist w ∈ (E′)0
g)∗ with
ew and rE ′ (µ) = w such that β(e∞
rE ′ (µ) ∈ (E′)0
w ) = µ. Hence,
g(F ) ⊆ (E′
sink,
rE ′
so F satisfies (1) in Lemma 3.12. A similar argument using λ−1 shows that
rEg(λ−1((E′)0
cycle. Therefore, λ(v) = κ−1
w . So, in particular, µ ∈ (E′
cycle. In particular, F ⊆ (E′
E ′ (β(κE(v))) = κ−1
E ′ (µe∞
cycle)) ⊆ (E′)0
g)≥1 with rE ′
v ) = µe∞
g(λ(E0
cycle)) ⊆ E0
cycle.
g)0
?
_
?
$
$
DIAGONAL PRESERVING ∗-ISOMORPHISMS OF GRAPH C ∗-ALGEBRAS
17
Let v ∈ (E′)0
cycle. Then λ−1(v) = µ ∈ E∗
g with rE(µ) = w ∈ E0
v . By [BCW14, Lemma 3.5], there exist n, m ∈ N0 such that
β(µe∞
w ) = e∞
cycle and
E ′ (β(µe∞
As σn
edges in ν equal to ev. Hence λ(w) = ν with rE ′
w )) = e∞
v , β(e∞
g (ν) = v, so
E ′ (β(σµ
σn
E (µe∞
w ) = νe∞
v
E ′ (β(µe∞
w ))) = σn
for some ν ∈ (E′)∗ with rE ′ (ν) = v and no
w )).
rE ′
g(λ(rEg (λ−1(v)))) = rE ′
g (λ(rEg (µ))) = rE ′
g(λ(w)) = v.
Applying this to rEg (λ−1((E′)0
and conclude that (E′)0
cycle = rE ′
g (λ(E0
cycle)).
cycle)) ⊆ E0
cycle we see that (E′)0
cycle ⊆ rE ′
g(λ(E0
cycle))
g(λ(v1)) = rE ′
w1 and β(e∞
We now show that F satisfies (2) in Lemma 3.12. Let v1, v2 ∈ E0
g(λ(v2)). We will show that λ(v1) = λ(v2). Then β(e∞
cycle and assume
that rE ′
v1 ) =
cycle, µi ∈ (E′)∗, no edges in µi are
µ1e∞
g(λ(vi)) = wi.
equal to ewi and rE ′ (µi) = wi. So, λ(vi) = µi, which implies that rE ′
So w1 = w2. Since β is an orbit equivalence, by [BCW14, Lemma 3.5], there exist
n1, m1, n2, m2 ∈ N0 such that
w2 where wi ∈ (E′)0
v2 ) = µ2e∞
E (β−1(σµi
σni
F (µie∞
w1))) = σmi
E (β−1(µie∞
w1))
for all i ∈ {1, 2}. Thus,
v1 = σn2
e∞
E (e∞
v1 ) = σn1+n2
E
(β−1(e∞
w1 )) = σn1
E (e∞
v2 ) = e∞
v2 .
g(λ(E0
g → ∂E′
Let κ : ∂E′
g (λ(v)) /∈ rE ′
g. Let v ∈ E0
cycle)) = (E′)0
cycle, we have that γ(E0
cycle. Either way, γ(v) = κ(λ(v)) = rE ′
cycle. If λ(v) ∈ F then κ(λ(v)) = rE ′
g (F ) as we saw above that rE ′
This implies that v1 = v2. We have just shown that F satisfies (2) in Lemma 3.12.
g be the orbit equivalence given in Lemma 3.12 for the
graph E′
g and the set F . Then γ = κ ◦ λ is an orbit equivalence from ∂Eg to
∂E′
g(λ(v)). If λ(v) /∈ F then
g (λ(w)) im-
g(λ(v)) = rE ′
λ(v) = rE ′
plies v = w for all w ∈ E0
g (λ(v)). Since
sink. Since
rE ′
Eg and E′
g are graphs satisfying Condition (L), by Theorem 3.10, there exists
an isomorphism ϕ : GEg → GE ′
g such that ϕ∂Eg = γ. By [BCW14, Proposi-
tion 2.2], this isomorphism of groupoids induces a ∗-isomorphism Φ : C∗(Eg) →
C∗(E′
cycle, where
(cid:8)pv, se(cid:12)(cid:12) v ∈ E0
(cid:8)qv, te(cid:12)(cid:12) v ∈ (E′
g(cid:9) is a Cuntz-Krieger Eg-family generating C∗(Eg) and
g)1(cid:9) is a Cuntz-Krieger E′
g).
∼= E′, Proposition 5.1 implies that there
Since (Eg)f,E0
exists a diagonal preserving ∗-isomorphism Ψ : C∗(E) → C∗(E′). Hence (4) implies
(1).
(cid:3)
g) and Φ(pv) = qγ(v) for all v ∈ E0
g) such that Φ(D(Eg)) = D(E′
g-family generating C∗(E′
g, e ∈ E1
g)0, e ∈ (E′
cycle) = (E′)0
∼= E and (E′
cycle ⊆ (E′
g)f,(E ′)0
g)0
cycle
cycle
We will denote the C∗-algebra of compact operators on ℓ2(N) by K and the
maximal abelian subalgebra of K consisting of diagonal operators by C. For a
commutative ring R with identity, we write M∞(R) for the ring of finitely supported,
countably infinite square matrices over R and D∞(R) for the abelian subring of
M∞(R) consisting of diagonal matrices.
We write (C∗(E), D(E)) ∼= (C∗(F ), D(F )) if there exists a diagonal preserv-
ing ∗-isomorphism Ψ from C∗(E) to C∗(F ), and write (C∗(E) ⊗ K, D(E) ⊗ C) ∼=
(C∗(F ) ⊗ K, D(F ) ⊗ C) if there exists a ∗-isomorphism from Ψ from C∗(E) ⊗ K to
C∗(F ) ⊗ K such that Ψ(D(E) ⊗ C) = D(F ) ⊗ C.
Let R be a commutative ring with identity. If there is a ring ∗-isomorphism Ψ
from LR(E) to LR(F ) such that Ψ(DR(E)) = DR(F ), then we write
(LR(E), DR(E)) ∼= (LR(F ), DR(F )).
18
SARA E. ARKLINT, SØREN EILERS, AND EFREN RUIZ
Similarly, if there is a ring ∗-isomorphism Ψ from LR(E) ⊗ M∞(R) to LR(F ) ⊗
M∞(R) such that Ψ(DR(E) ⊗ D∞(R)) = DR(F ) ⊗ D∞(R), then we write
(LR(E) ⊗ M∞(R), DR(E) ⊗ D∞(R)) ∼= (LR(F ) ⊗ M∞(R), DR(F ) ⊗ D∞(R)).
Let R be the full equivalence relation on N × N. We can regard R as a discrete
principal groupoid with unit space N.
Corollary 5.4. Let E and F be graphs, and let R be a commutative integral domain
with 1. The following are equivalent:
(1) (C∗(E) ⊗ K, D(E) ⊗ C) ∼= (C∗(F ) ⊗ K, D(F ) ⊗ C);
(2) (LR(E)⊗ M∞(R), DR(E)⊗ D∞(R)) ∼= (LR(F )⊗ M∞(R), DR(F )⊗ D∞(R));
(3) (C∗(SE), D(SE)) ∼= (C∗(SF ), D(SF ));
(4) (LR(SE), DR(SE)) ∼= (LR(SF ), DR(SF ));
(5) GE × R ∼= GF × R;
(6) GSE ∼= GSF ;
(7) There exists an orbit equivalence from SE to SF preserving periodic points.
Proof. By [CRS16, Theorem 4.2], (1) through (6) are equivalent. (6) ⇐⇒ (7)
follows from Theorem 5.3 for the graphs SE and SF .
(cid:3)
We end by noting that the results above combine with [ERRS16] to completely re-
solve the relationship between orbit and flow equivalence for countable shift spaces.
Corollary 5.5. Let E and F be finite graphs with no sinks and sources such that
the edge shift spaces XE and XF are countable sets.
(1) If E and F are orbit equivalent, then XE and XF are flow equivalent.
(2) SE and SF are orbit equivalent if and only if XE and XF are flow equiva-
lent.
Proof. Suppose E and F are orbit equivalent. Since E and F have no sinks, there
exists an orbit equivalence from E to F preserving periodic points. Hence, by
Theorem 5.3, C∗(E) ∼= C∗(F ). One easily sees by contradiction that the asserted
countability translates to the condition that every vertex of E and F either supports
exactly one return path or does not support a return path. Hence, [ERRS16,
Theorem 7.1 (3) =⇒ (4)] applies, and thus the shift spaces XE and XF are flow
equivalent, proving (1).
For (2), we note that if SE and SF are orbit equivalent then by Corollary 5.4,
C∗(E)⊗ K ∼= C∗(F )⊗ K, and the forward implication follows as above. In the other
direction, suppose XE and XF are flow equivalent. By [ERRS16, Lemma 5.1], then
E and F are Move equivalent. By [CRS16, Corollary 4.8], there exists a diagonal
preserving isomorphism from C∗(E) ⊗ K to C∗(F ) ⊗ K. Hence, by Corollary 5.4,
SE and SF are orbit equivalent.
(cid:3)
6. Acknowledgements
This work was partially supported by the Danish National Research Founda-
tion through the Centre for Symmetry and Deformation (DNRF92), by VILLUM
FONDEN through the network for Experimental Mathematics in Number Theory,
Operator Algebras, and Topology, and by a grant from the Simons Foundation (#
279369 to Efren Ruiz).
This work was completed while all three authors were attending the research
program Classification of operator algebras: complexity, rigidity, and dynamics at
the Mittag-Leffler Institute, January -- April 2016. We thank the institute and its
staff for the excellent work conditions provided.
The authors thank Aidan Sims for many helpful discussions.
DIAGONAL PRESERVING ∗-ISOMORPHISMS OF GRAPH C ∗-ALGEBRAS
19
References
[ADR00] C. Anantharaman-Delaroche and J. Renault, Amenable groupoids, Monographies
de L'Enseignement Mathématique [Monographs of L'Enseignement Mathématique],
vol. 36, L'Enseignement Mathématique, Geneva, 2000, With a foreword by Georges
Skandalis and Appendix B by E. Germain.
[BCW14] Nathan Brownlowe, Toke Meier Carlsen, and Michael F. Whittaker, Graph al-
gebras and orbit equivalence, to appear in Ergod. Theory and Dynam. Systems,
doi:10.1017/etds.2015.52 .
[CRS16] Toke Meier Carlsen, Efren Ruiz, and Aidan Sims, Equivalence and stable isomorphism
of groupoids, and diagonal-preserving stable isomorphisms of graph C ∗-algebras and
Leavitt path algebras, eprint arXiv:1602.02602 (2016).
Toke M. Carsen and Marius L. Winger, Orbit equivalence of graphs and isomorphism
of graph groupoids, private communication, 2016.
[CW16]
[ERRS16] Søren Eilers, Gunnar Restorff, Efren Ruiz, and Adam P. W. Sørensen, Geometric
classification of graph C ∗-algebras over finite graphs, eprint arXiv:1604.05439, 2016.
[Mat13] Kengo Matsumoto, Classification of Cuntz-Krieger algebras by orbit equivalence of
topological Markov shifts, Proc. Amer. Math. Soc. 141 (2013), 2329 -- 2342.
[MM14] Kengo Matsumoto and Hiroki Matui, Continuous orbit equivalence of topological
Markov shifts and Cuntz-Krieger algebras, Kyoto J. Math. 54 (2014), no. 4, 863 -- 877,
doi:10.1215/21562261-2801849.
Jean Renault, Cartan subalgebras in C ∗-algebras, Irish Math. Soc. Bull. (2008), no. 61,
29 -- 63.
[Ren08]
[Rør95] Mikael Rørdam, Classification of Cuntz-Krieger algebras, K-Theory 9 (1995), no. 1,
31 -- 58, doi:10.1007/BF00965458.
[Szy02] Wojciech Szymański, General Cuntz-Krieger uniqueness theorem, Internat. J. Math.
[Web14]
[Yee07]
13 (2002), no. 5, 549 -- 555, doi:10.1142/S0129167X0200137X.
Samuel B. G. Webster, The path space of a directed graph, Proc. Amer. Math. Soc.
142 (2014), no. 1, 213 -- 225, doi:10.1090/S0002-9939-2013-11755-7.
Trent Yeend, Groupoid models for the C ∗-algebras of topological higher-rank graphs,
J. Operator Theory 57 (2007), no. 1, 95 -- 120.
Department of Mathematical Sciences, University of Copenhagen, Universitets-
parken 5, DK-2100 Copenhagen, Denmark
E-mail address: [email protected]
Department of Mathematical Sciences, University of Copenhagen, Universitets-
parken 5, DK-2100 Copenhagen, Denmark
E-mail address: [email protected]
Department of Mathematics, University of Hawaii, Hilo, 200 W. Kawili St., Hilo,
Hawaii, 96720-4091 USA
E-mail address: [email protected]
|
1106.3143 | 1 | 1106 | 2011-06-16T05:25:41 | Ideals in Operator Space Projective Tensor Product of $C^*$-algebras | [
"math.OA"
] | For $C^*$-algebras $A$ and $B$, we prove the slice map conjecture for ideals in the operator space projective tensor product $A \hat\otimes B$. As an application, a characterization of prime ideals in the Banach $\ast$-algebra $A\hat\otimes B$ is obtained. Further, we study the primitive ideals, modular ideals and the maximal modular ideals of $A\hat\otimes B$. It is also shown that the Banach $\ast$-algebra $A\hat\otimes B$ possesses Wiener property; and that, for a subhomogenous $C^*$-algebra $A$, $A\hat\otimes B$ is symmetric. | math.OA | math |
IDEALS IN OPERATOR SPACE PROJECTIVE TENSOR
PRODUCT OF C ∗-ALGEBRAS
RANJANA JAIN AND AJAY KUMAR
Abstract. For C ∗-algebras A and B, we prove the slice map conjecture
for ideals in the operator space projective tensor product A b⊗B. As an
application, a characterization of prime ideals in the Banach ∗-algebra A b⊗B
is obtained. Further, we study the primitive ideals, modular ideals and
the maximal modular ideals of A b⊗B. It is also shown that the Banach ∗-
algebra A b⊗B possesses Wiener property; and that, for a subhomogeneous
C ∗-algebra A, A b⊗B is symmetric.
1.
Introduction
A systematic study of tensor products of subspaces and subalgebras of C ∗-
algebras was initiated by Blecher and Paulsen [7], and Effros and Ruan [9, 10].
Analogous constructions to those of Banach spaces; for example, quotients,
duals and tensor products were defined and studied. For a Hilbert space H,
let B(H) denote the bounded operators on H. An operator space X on H
is just a closed subspace of B(H). If E and F are operator spaces, then the
operator space projective tensor product, denoted by Eb⊗F , is the completion
of the algebraic tensor product E ⊗ F under the norm
kuk∧ = inf{kαkkvkkwkkβk : u = α(v ⊗ w)β},
where the infimum runs over arbitrary decompositions with v ∈ Mp(E), w ∈
Mq(F ), α ∈ M1, pq, β ∈ Mpq, 1 with p, q ∈ N arbitrary; Mk, l being the space
If E and F are C ∗-algebras, then Eb⊗F admits a
of k × l matrices over C.
Banach algebra with canonical isometric involution [16]. The main objective of
this paper is to study the closed ∗-ideals of this Banach ∗-algebra.
In Section 2, we study the slice map problem for ideals of Ab⊗B. Tomiyama
[25] studied the slice maps on the tensor product of C ∗-algebras with respect
to the 'min'-norm. Later, Wassermann [26] discussed the slice map problem
in greater detail, which was then studied and used in different contexts - see,
for instance, [2, 27]. It is interesting to know that the slice map property is
not true for the 'min' norm for all C ∗-algebras.
In fact, for the 'min' norm
the slice map problem for ideals is equivalent to the problem of whether every
tensor product A ⊗min B has Property F of Tomiyama [26, Remark 24].
In
1991, Smith [23] studied the slice map property for the Haagerup norm and
proved that the slice map conjecture is true for all subspaces of B(H). We give
2000 Mathematics Subject Classification. 46L06,46L07,47L25.
Key words and phrases. C ∗-algebras, Operator space projective tensor norm, Haagerup
tensor product.
1
2
R. JAIN AND A. KUMAR
an affirmative answer to the slice map conjecture for ideals with respect to the
operator space projective tensor norm.
The ideal structure for the Haagerup tensor product and the 'min' norm
has been studied extensively in [1], [3] and [24]. In [16] and [14], the authors
investigated some properties of the closed ideals of the projective tensor product
In Section 3, we discuss a characterization of prime ideals, primitive ideals, and
Ab⊗B, for example, sum of the product ideals, minimal and the maximal ideals.
maximal modular ideals of the Banach ∗-algebra Ab⊗B. Finally, in Section 4,
certain ∗-algebraic properties of Ab⊗B, namely, Wiener property and symmetry
are studied. Throughout the paper, A and B will denote C ∗-algebras unless
otherwise specified.
Recall that the Haagerup norm of an element u in the algebraic tensor prod-
uct A ⊗ B of two C ∗-algebras A and B is defined by
kukh = inf{kΣi aia∗
i k1/2 kΣi b∗
i bik1/2 : u = Σn
i=1ai ⊗ bi}.
The Haagerup tensor product A ⊗h B is defined to be the completion of A ⊗ B
in the norm k · kh. Also, the Banach space projective norm of u ∈ A ⊗ B is
given by
kukγ = inf{Σi kaikkbik : u = Σn
i=1 ai ⊗ bi}.
The norms k · kh, k · k∧ and k · kγ on the tensor product A ⊗ B of two C ∗-algebras
A and B satisfy
k · kh ≤ k · k∧ ≤ k · kγ.
Necessary and sufficient conditions on A and B for the equivalence of these
norms can be seen in [17].
2. Slice Map Property for Ideals
For each φ ∈ A∗, define a linear map Rφ : A ⊗ B → B by
Rφ(Σn
i=1 ai ⊗ bi) = Σn
i=1 φ(ai)bi.
Then, it can be easily seen that Rφ is well defined. Also, it is continuous with
respect to the 'min'-norm [26] and hence for the larger operator space projective
tensor norm with kRφk ≤ kφk; so, it can be extended to Ab⊗B as a bounded
linear map and is known as the right slice map associated to φ. Similarly, one
can define the left slice map Lψ for each ψ ∈ B∗. For a closed ideal J of B,
Ab⊗J is a closed ideal of Ab⊗B [16] and clearly Rφ(x) ∈ J for all x ∈ Ab⊗J. We
prove the converse of this statement which is known as the slice map problem
for ideals.
Lemma 2.1. The set {Rφ : φ ∈ A∗} is total on Ab⊗B, that is, if x ∈ Ab⊗B and
Rφ(x) = 0 for all φ ∈ A∗, then x = 0.
Proof. For φ ∈ A∗ and ψ ∈ B∗, consider φ ⊗ ψ : A ⊗ B → C given by
(φ ⊗ ψ)(Σi ai ⊗ bi) = Σi φ(ai)ψ(bi).
Note that, by the definition of the Banach space injective norm λ [24, page
188], we have Σi φ(ai)ψ(bi) ≤ kφkkψkkΣiai ⊗ bikλ. Thus φ ⊗ ψ is continuous
with respect to larger norms, in particular, 'min'-norm and '∧'-norm; so, φ ⊗ ψ
can be extended to continuous linear functionals on A ⊗min B and Ab⊗B. Let
IDEALS IN O.S. PROJECTIVE TENSOR PRODUCT
3
us denote its extensions by φ ⊗min ψ and φb⊗ψ respectively. We claim that the
set {φb⊗ψ : φ ∈ A∗, ψ ∈ B∗} is total on Ab⊗B. For this, consider an element
x ∈ Ab⊗B such that
(φb⊗ψ)(x) = 0, ∀φ ∈ A∗, ψ ∈ B∗.
Observe that for the canonical map i : Ab⊗B → A ⊗min B, the maps φb⊗ψ
and (φ ⊗min ψ) ◦ i both are continuous on Ab⊗B and agree on A ⊗ B, giving
(φ ⊗min ψ)(i(x)) = 0 for all φ ∈ A∗, ψ ∈ B∗. Now, for faithful representations
{πA, H} and {πB, K} of A and B respectively, for ξi ∈ H, ηi ∈ K, i = 1, 2;
φ := hπA(·)ξ1, ξ2i ∈ A∗, ψ := hπB(·)η1, η2i ∈ B∗; so
0 = (φ ⊗min ψ)(i(x)) = h(πA ⊗ πB)(i(x))ξ1 ⊗ η1, ξ2 ⊗ η2i.
This holds for all ξi ∈ H, ηi ∈ K; i = 1, 2 giving (πA ⊗ πB)(i(x)) = 0. Using
the facts that πA ⊗ πB is faithful [24, Theorem IV.4.9], and that i is injective
[13, Corollary 1] we obtain the claim. Finally, the relation
hx, φb⊗ψi = hRφ(x), ψi = hLψ(x), φi, ∀x ∈ Ab⊗B,
gives the required result.
(cid:3)
Recall that, for Banach spaces X and Y , a mapping θ : X → Y is said to be
a quotient map if it maps the open unit ball of X onto that of Y [9]. Clearly,
a quotient map is surjective, and for Banach space X and a closed subspace
Y of X, the canonical quotient map π : X → X/Y is a quotient map in the
above sense. Like in the case of Haagerup tensor product [1], the operator space
projective tensor product of quotient maps behaves nicely. Although straight
forward, we include a proof of the following for the sake of convenience:
Lemma 2.2. Let I and J be closed ideals of the C ∗-algebras A and B, and
π : A → A/I and ρ : B → B/J be the quotient maps. Then,
(1) πb⊗ρ : Ab⊗B → (A/I)b⊗(B/J) is a quotient map with
ker(πb⊗ρ) = Ab⊗J + Ib⊗B.
(2) for a closed ideal K of Ab⊗B containing ker(πb⊗ρ), (πb⊗ρ)(K) is a closed
ideal of (A/I)b⊗(B/J) with
(πb⊗ρ)−1((πb⊗ρ)(K)) = K.
Proof. (1) This follows directly from [14, Proposition 3.5].
(2) Consider an element (πb⊗ρ)(x) ∈ (A/I)b⊗(B/J) such that (πb⊗ρ)(x) ∈
cl((πb⊗ρ)(K)), where x ∈ Ab⊗B. Given an arbitrary ǫ > 0, there exists k ∈ K
such that
k(πb⊗ρ)(k − x)k(A/I) b⊗(B/J) < ǫ.
Using part (1) above, there is an isomorphism between (Ab⊗B)/Z and (A/I)b⊗
(B/J), where Z = ker(πb⊗ρ). Therefore,
k(k − x) + Zk(A b⊗B)/Z < cǫ,
for some constant c. So, there exists some z ∈ Z ⊆ K with k(k + z) −
xk(A b⊗B)/Z ≤ cǫ. Since K is closed and k + z ∈ K, we must have x ∈ K,
4
R. JAIN AND A. KUMAR
which proves the claim. Finally, the equation in the statement is a routine
verification.
(cid:3)
We are now prepared to present a proof of the slice map problem for ideals.
Theorem 2.3. Let J be a closed ideal of B. Then
Ab⊗J = {x ∈ Ab⊗B : Rφ(x) ∈ J for all φ ∈ A∗}.
Lemma 2.2, corresponding to the quotient map π : B → B/J, we have another
Proof. Consider an element x ∈ Ab⊗B such that Rφ(x) ∈ J for all φ ∈ A∗. From
quotient map ib⊗π : Ab⊗B → Ab⊗(B/J) with ker(ib⊗π) = Ab⊗J, where 'i' is the
identity map on A. Also observe that, by continuity and agreement on A ⊗ B,
π ◦ Rφ = rφ ◦ (ib⊗π),
where rφ : Ab⊗(B/J) → B/J is the right slice map. Using the fact that Rφ(x) ∈
J for all φ ∈ A∗, we obtain rφ(ib⊗π(x)) = 0 for all φ ∈ A∗. Thus, by Lemma 2.1,
ib⊗π(x) = 0; so that x ∈ ker(ib⊗π) = Ab⊗J. The other containment is easy. (cid:3)
We next give an application of Theorem 2.3 which will be used later to
characterize the prime ideals. For the Haagerup norm such a result was proved
for subspaces of B(H) in [23, Corollary 4.6].
Proposition 2.4. Let A1, A2 and B1, B2 be closed ideals of A and B, respec-
tively. Then,
(A1b⊗B1) ∩ (A2b⊗B2) = (A1 ∩ A2)b⊗(B1 ∩ B2).
Proof. Since Aib⊗Bi, i = 1, 2 are closed ideals of Ab⊗B [16], it is easy to see that
(A1 ∩ A2)b⊗(B1 ∩ B2) ⊆ (A1b⊗B1) ∩ (A2b⊗B2).
For the other containment, consider an element v ∈ (A1b⊗B1)∩(A2b⊗B2). Then,
Rφ(v) ∈ B1 ∩ B2 for all φ ∈ A∗; so, by Theorem 2.3, v ∈ Ab⊗(B1 ∩ B2).
Next, consider any ψ ∈ (B1 ∩ B2)∗ and let ψ be an extension on B∗. Again,
L ψ(v) ∈ (A1 ∩ A2) and Lψ(v) = L ψ(v); so that Lψ(v) ∈ (A1 ∩ A2). This is true
for every ψ ∈ (B1 ∩ B2)∗; so, applying the slice map property once again for the
left slice map, we obtain v ∈ (A1 ∩ A2)b⊗(B1 ∩ B2), which proves the claim. (cid:3)
Using the slice map property for the right and the left slice maps, and the
technique of extending linear functionals as done in Proposition 2.4, we can
easily deduce the following:
Corollary 2.5. For closed ideals I and J of A and B respectively, we have
Ib⊗J = {x ∈ Ab⊗B : Rφ(x) ∈ J, Lψ(x) ∈ I; ∀ φ ∈ A∗, ∀ ψ ∈ B∗}.
3. Ideal Structure for Ab⊗B
This section deals with the structure of prime ideals, primitive ideals and
modular ideals of Ab⊗B which play an important role in determining the struc-
ture of a Banach ∗-algebra.
In a Banach algebra a proper closed ideal K is
said to be prime if for any pair of closed ideals I and J satisfying IJ ⊆ K,
either I ⊆ K or J ⊆ K. It is well known that a proper closed ideal K of a C ∗-
algebra A is prime if and only if for any pair of closed ideals I and J satisfying
IDEALS IN O.S. PROJECTIVE TENSOR PRODUCT
5
I ∩ J ⊆ K, either I ⊆ K or J ⊆ K. This property is also true for Ab⊗B as
can be explicitly observed from the following result. The proof of the following
result is largely inspired by [1].
Now consider the quotient maps π : A → A/E and ρ : B → B/F . Since
Proof. Let K be a closed prime ideal. We can choose closed ideals E and F in
Theorem 3.1. A closed ideal K in Ab⊗B is prime if and only if K = Ab⊗F +
Eb⊗B for some prime ideals E and F in A and B respectively.
A and B which are maximal with respect to the property Ab⊗F + Eb⊗B ⊆ K.
ker(π ⊗ ρ) ⊆ K, by Lemma 2.2, (π ⊗ ρ)(K) is a closed ideal of A/Eb⊗B/F . We
claim that (π ⊗ ρ)(K) = 0; this would imply K = Ab⊗F + Eb⊗B. If possible, let
the ideal (π ⊗ ρ)(K) be non-zero. Then, it must contain a non-zero elementary
tensor, say, π(a) ⊗ ρ(b), where a ⊗ b ∈ K [14, Proposition 3.7]. Let E0 and F0
be the closed ideals generated by a and b respectively. Then, the product ideal
E0b⊗F0 is contained in K. Now, consider the product ideals M = Ab⊗(F + F0)
and N = (E + E0)b⊗B. Using Proposition 2.4 and [14, Proposition 3.6], we have
M N ⊆ M ∩ N = Eb⊗F + Eb⊗F0 + E0b⊗F + E0b⊗F0.
It is clear that M N ⊆ K, so that either M ⊆ K or N ⊆ K. Using the
maximality property of E and F , we have either E0 ⊆ E or F0 ⊆ F . Thus,
either π(a) = 0 or ρ(b) = 0 contradicting the fact that (π ⊗ ρ)(a ⊗ b) 6= 0.
Next we prove that E and F are prime ideals. Note that E and F both are
proper ideals, K being proper. Let I ∩J ⊆ E for some closed ideals I and J of A.
Then, (Ib⊗B)(J b⊗B) ⊆ (Ib⊗B)∩(J b⊗B) ⊆ K; so, either Ib⊗B ⊆ K or J b⊗B ⊆ K.
Without loss of generality, let Ib⊗B ⊆ K. Consider any φ ∈ E⊥ ⊆ A∗ and
0 6= ψ ∈ F ⊥. Then, (φ ⊗ ψ)(K) = 0 which further gives (φ ⊗ ψ)(Ib⊗B) = 0.
Since this is true for any φ ∈ E⊥, we must have I ⊆ E. Thus, E is prime and
by a similar argument F is also prime.
For the converse, let us assume that K = Ab⊗F + Eb⊗B for some prime ideals
E and F in A and B respectively. Let IJ ⊆ K for some closed ideals I and J
of Ab⊗B. Define the closed ideals M and N as
M = cl(I + K) and N = cl(J + K).
Then K ⊆ M, K ⊆ N and M N ⊆ K. We claim that either M = K or N = K,
which further implies that either I ⊆ K or J ⊆ K. Suppose, on the contrary,
that both the containments K ⊆ M and K ⊆ N are strict. We now claim
that M contains a product ideal M1b⊗N1 which is not contained in K. As done
previously, since K ( M, (π ⊗ ρ)(M ) is a non-zero closed ideal of A/Eb⊗B/F
with (π⊗ρ)−1((π⊗ρ)(M )) = M . So, (π⊗ρ)(M ) contains a non-zero elementary
tensor say π(a) ⊗ ρ(b). Define M1 and N1 to be the closed ideals generated by
a and b. Then M1b⊗N1 is contained in M but not in K. Similarly, N contains
a product ideal M2b⊗N2 which is not contained in K. By routine calculations,
it is easily seen that
M1M2b⊗N1N2 = cl((M1b⊗N1)(M2b⊗N2)) ⊆ cl(M N ) ⊆ K,
which further gives
π(M1M2) ⊗ ρ(N1N2) ⊆ (π ⊗ ρ)(M1M2b⊗N1N2) = {0}.
6
R. JAIN AND A. KUMAR
So either M1M2 ⊆ ker π = E or N1N2 ⊆ ker ρ = F . Now, both E and F are
prime, so at least one of the following containments must hold:
M1 ⊆ E, M2 ⊆ E, N1 ⊆ F, N2 ⊆ F.
In all these cases, either M1b⊗N1 or M2b⊗N2 is contained in K, which is a
contradiction. Thus, K is prime.
(cid:3)
A closed ideal I of a Banach ∗-algebra E is said to be primitive if it is
the kernel of an irreducible ∗-representation of E on some Hilbert space. The
following gives a characterization of the primitive ideals of Ab⊗B.
Theorem 3.2. For C ∗-algebras A and B, we have the following:
Eb⊗B is also a primitive ideal of Ab⊗B.
(1) If E and F are primitive ideals of A and B respectively, then Ab⊗F +
(2) If K is a primitive ideal of Ab⊗B, then K = Ab⊗F + Eb⊗B for some
(3) If A and B are separable, then K is primitive if and only if K = Ab⊗F +
Eb⊗B for some primitive ideals E and F of A and B, respectively.
prime ideals E and F of A and B, respectively.
Proof. (1) Since E and F are primitive ideals, there exist irreducible ∗- repre-
sentations π1 : A → B(H1) and π2 : B → B(H2) such that E = ker π1 and
F = ker π2. Define π : A ⊗ B → B(H1 ⊗ H2) by
π(a ⊗ b) = π1(a) ⊗ π2(b).
Then, by the definition of min-norm [24], π is bounded with respect to the
min-norm and hence the '∧' norm; so, π can be extended to Ab⊗B as a bounded
∗-representation. We first claim that π is irreducible, equivalently, π(Ab⊗B)′ =
CI. Since π(Ab⊗B) ⊃ π1(A)⊗π2(B), we have π(Ab⊗B)′ ⊆ (π1(A)⊗π2(B))′,
where ⊗ denotes the weak closure. Further, π1 and π2 being irreducible, π1(A)
and π2(B) are non-degenerate ∗-subalgebras of B(H1) and B(H2), respectively;
so that, by Double Commutant Theorem, π1(A) and π2(B) are weakly dense
In particular, π1(A)⊗π2(B) = π1(A)′′ ⊗π2(B)′′; and,
in π1(A)′′ and π2(B)′′.
an appeal to Tomita's Commutation Theorem then yields (π1(A)⊗π2(B))′ =
π1(A)′ ⊗π2(B)′ ⊆ CI, which shows that π is irreducible.
are both contained in ker π; so that K ⊆ ker π. For the other containment,
Next we claim that ker π = Ab⊗F + Eb⊗B = K(say). Clearly, Ab⊗F and Eb⊗B
consider the quotient map θ : Ab⊗B → A/Eb⊗B/F with ker θ = K. Since,
ker π contains ker θ, by Lemma 2.2, θ(ker π) is a closed ideal of A/Eb⊗B/F
with θ−1(θ(ker π)) = ker π. If θ(ker π) 6= 0, then it must contain a non-zero
elementary tensor say (a + E) ⊗ (b + F ) [14, Proposition 3.7]. Now a ⊗ b ∈ ker π
implies π1(a) ⊗ π2(b) = 0, which further implies that either a ∈ E or b ∈ F , so
that (a + E) ⊗ (b + F ) = 0, which is a contradiction. Thus, ker π ⊆ ker θ = K.
(2) Let K = ker π for some irreducible ∗-representation π of Ab⊗B on H. By
[24, Lemma IV.4.1], there exist commuting ∗-representations π1 : A → B(H)
and π2 : B → B(H) such that
π(a ⊗ b) = π1(a)π2(b), ∀ a ∈ A, b ∈ B.
IDEALS IN O.S. PROJECTIVE TENSOR PRODUCT
7
Now, π(A ⊗ B) = π1(A)π2(B), so π(Ab⊗B) ⊆ cl(π1(A)π2(B)). Thus, we obtain
′
′
′
(π1(A)π2(B))
= cl(π1(A)π2(B))
⊆ π(Ab⊗B)
= CI.
Also, note that π1 and π2 are both factor representations as for P = π1(A)
and Q = π2(B)
, we have
′′
′′
′
P ∩ P
′′
′
′
′
∩ π1(A)
= π1(A)
= (π1(A)
∪ π1(A))
⊆ (π2(B) ∪ π1(A))
⊆ {π1(A)π2(B)}
= CI.
′
′
(as π1(A) and π2(B) commute)
Now, let E = ker π1 and F = ker π2. Then E and F , being kernels of factor
representations, are both prime ideals [4, II.6.1.11]. Also, by the definition of
′′
π, Ab⊗F + Eb⊗B ⊆ K. For the reverse containment, consider a ⊗ b ∈ K. Then,
we have π1(a)π2(b) = 0. Since π1(A)
, using
[24, Proposition IV.4.20], we see that either π1(a) = 0 or π2(b) = 0, i.e., a ⊗ b
belongs to either Ab⊗F or Eb⊗B. In both cases, a ⊗ b ∈ Ab⊗F + Eb⊗B. Finally,
exactly on the lines of (1), we conclude that K ⊆ Ab⊗F + Eb⊗B.
(3) If A and B are separable, then every prime ideal is a primitive ideal. So,
(cid:3)
is a factor and π2(B)
the result follows from parts (1) and (2).
⊆ π1(A)
′′
′
In particular, among all the five proper closed ideals of B(H)b⊗B(H) - see
[14, Theorem 3.12]- namely, {0}, B(H)b⊗K(H), K(H)b⊗B(H), B(H)b⊗K(H) +
K(H)b⊗B(H) and K(H)b⊗K(H), the first four are prime as well primitive.
We now discuss the modular ideals of Ab⊗B.
In a Banach algebra A, an
ideal I is said to be modular (or regular) if there exists an e ∈ A such that
xe − x, ex − x ∈ I for all x ∈ A, or equivalently, if A/I is unital. It is clear that
every proper ideal in a unital Banach algebra is modular. Also, {0} is modular
if and only if A is unital.
If I is a closed modular ideal of A, then the product ideal Ib⊗A need not be
modular in Ab⊗A. This can be seen by considering A = C0(X), where X is
a locally compact Hausdorff space (non-compact). A closed modular ideal of
C0(X) is of the form I(E) = {f ∈ A : f (E) = 0}, where E is a compact subset
of X [15]. So let us consider a closed modular ideal I = I(E) of A. Now note
that
Ib⊗A ⊆ Ab⊗A ⊆ A ⊗λ A = C0(X × X),
we have the following result which characterizes the modular product ideals.
where 'λ' is the Banach space injective tensor product. This shows that Ib⊗A ⊆
I(E × X). Thus, Ib⊗A is not modular, I(E × X) not being modular. In fact,
Theorem 3.3. For closed modular ideals I and J of A and B respectively, Ib⊗J
is modular in Ab⊗B if and only if both A and B are unital.
Proof. If A and B are both unital, then so is Ab⊗B; so that every ideal is
modular. Conversely, let Ib⊗J be a modular ideal. Since Ab⊗J and Ib⊗B both
contain Ib⊗J, both are modular ideals of Ab⊗B. Using Lemma 2.2, we have
an isomorphism between (Ab⊗B)/(Ab⊗J) and Ab⊗(B/J), and similarly between
8
R. JAIN AND A. KUMAR
(cid:3)
which further show that A and B are both unital [20, Theorem 1].
(Ab⊗B)/(Ib⊗B) and (A/I)b⊗B. Therefore, Ab⊗(B/J) and (A/I)b⊗B are unital
In particular, K(H)b⊗K(H) is a closed modular ideal of B(H)b⊗B(H), but it
is not modular in B(H)b⊗K(H). However, the maximal modular ideals behave
well in Ab⊗B as can be seen in the following result:
Theorem 3.4. A closed ideal K of Ab⊗B is maximal modular if and only if
K = Ab⊗N + M b⊗B for some maximal modular ideals M and N of A and B,
Proof. Let K be a maximal modular ideal of Ab⊗B. Since every maximal mod-
ular ideal is also a maximal ideal, K is of the form K = Ab⊗N + M b⊗B for
Now (Ab⊗B)/K is unital and is isomorphic to A/M b⊗B/N , by Lemma 2.2 ;
some maximal ideals M and N of A and B respectively [14, Theorem 3.11].
respectively.
therefore, the latter space is unital. But this implies that A/M and B/N are
both unital [20, Theorem 1]. Thus, M and N are also modular ideals of A and
B respectively.
modular ideals of A and B respectively. Then, M and N being maximal, by
For the converse, let K = Ab⊗N + M b⊗B, where M and N are maximal
[14, Theorem 3.11], K is also a maximal ideal. Also, the facts that (Ab⊗B)/K
and A/M b⊗B/N are isomorphic, and A/M and that B/N are both unital,
together imply that Ab⊗B/K is unital, so that K is modular.
(cid:3)
4. Wiener Property and Symmetry
A Banach ∗-algebra is said to have Wiener property if every proper closed
two-sided ideal is annihilated by an irreducible ∗-representation [22]. Wiener
property for group algebras and the weighted group algebras has been studied in
[12, 21] and others. It is well known that every C ∗-algebra has Wiener property.
Theorem 4.1. The Banach ∗-algebra Ab⊗B has Wiener property.
Proof. Consider a proper closed two-sided ideal J of Ab⊗B. Let Jmin denote
the closure of i(J) in A ⊗min B, where i : Ab⊗B → A ⊗min B is the canonical
homomorphism. By [16, Theorem 6], Jmin is also a proper closed two-sided
ideal of the C ∗-algebra A ⊗min B, and so it is annihilated by an irreducible
∗-representation π : A ⊗min B → B(H). Note that the isometry of involution
gives i is ∗-preserving, so that we have a ∗-representation π := π ◦ i of Ab⊗B
on H. Using injectivity of i [13], we have π(J) = {0}. Also, the relation
π(A ⊗ B) = π(A ⊗ B) gives
π(Ab⊗B)′ ⊆ π(A ⊗ B)′ = π(A ⊗min B)′ = CI,
where the equality between the middle expressions follows from the norm den-
sity of π(A ⊗ B) in π(A ⊗min B). This further implies that π is irreducible;
hence, Ab⊗B has Wiener property.
(cid:3)
A Banach ∗-algebra is said to be symmetric if every element of the form
x∗x has positive spectrum, or equivalently, every self adjoint element has a
real spectrum [22, Theorem 10.4.17]. Symmetry in group algebras has been
IDEALS IN O.S. PROJECTIVE TENSOR PRODUCT
9
investigated by various authors, see, for instance, [21, 19]. One can easily verify
that a Banach ∗-algebra A is symmetric if and only if for every left modular
ideal I of A with modular unit α, the set SI of Hermitian sesquilinear forms
given by
SI = {B : A × A → C Bα = B, B(I, A) = {0}, B(u, u) ≥ 0,
B(uw, vw) = B(v∗uw, w), ∀u, v, w ∈ A}
is non-trivial, where Bα(v, w) := B(vα, wα), ∀v, w ∈ A [21]. It is well known
that every C ∗-algebra is symmetric [22]. For C ∗-algebras A and B, we do not
know whether the Banach ∗-algebra Ab⊗B is symmetric or not, but if one of
them is subhomogeneous, then we have an affirmative answer. Recall that a
C ∗-algebra A is subhomogeneous if there exists a positive integer n such that
each irreducible representation of A has dimension less than or equal to n.
We first modify a result from [16] in terms of operator algebras. We say that
a Banach algebra A is an operator algebra if there exists a Hilbert space H and
a bicontinuous homomorphism of A into B(H).
Proposition 4.2. If A and B are operator algebras, then Ab⊗B is a Banach
algebra. If A and B both have isometric involutions then Ab⊗B is a Banach
∗-algebra.
Proof. It is known that if A is an operator algebra then the multiplication
operator m : A ⊗h A → A given by m(a ⊗ b) = ab is completely bounded [6,
Theorem 1.3]. Using this result, we get the completely bounded operators
mA : A ⊗h A → A and mB : B ⊗h B → B.
Now consider the canonical map i : Ab⊗A → A ⊗h A, which is a completely
A : Ab⊗A → A,
contractive homomorphism. Then, the multiplication operator m′
which can be regarded as m′
A = mA ◦ i, is completely bounded. Similarly,
B : Bb⊗B → B is also completely bounded. In
the multiplication operator m′
particular, the operator
B : (Ab⊗A)b⊗(Bb⊗B) → Ab⊗B
A ⊗ m′
m′
is bounded. Using the commutativity of '∧', the operator
m′
A ⊗ m′
B : (Ab⊗B)b⊗(Ab⊗B) → Ab⊗B
follows as in [16].
is also bounded. Hence, Ab⊗B is a Banach algebra. The proof for involution
Lemma 4.3. Let A and B be C ∗-algebras with either A or B finite-dimensional.
Then Ab⊗B is a symmetric operator algebra.
Proof. If A or B is finite dimensional, then clearly, Ab⊗B is ∗-isomorphic to
A ⊗min B, which gives the required result.
(cid:3)
Lemma 4.4. If A is a commutative unital C ∗-algebra and B is a symmetric
unital operator algebra with isometric involution, then Ab⊗B is symmetric.
(cid:3)
10
R. JAIN AND A. KUMAR
denote the set of maximal ideals of A, then it is in one-one correspondence
with the space of non-zero ∗-homomorphisms of A. For M ∈ Φ(A), define
Proof. Note that Ab⊗B is a Banach ∗-algebra by Proposition 4.2. Let Φ(A)
hM : A ⊗ B → B by hM (P ai ⊗ bi) = P ai(M )bi. It is bounded with respect
to '∧'-norm, so can be extended to Ab⊗B as a ∗-homomorphism. Then, by [18,
Corollary 2], an element x of Ab⊗B is invertible if and only if hM (x) is invertible
for each maximal ideal M of A. Thus,
σ(x) = [
M ∈Φ(A)
σ(hM (x)),
where σ(x) denotes the spectrum of x in Ab⊗B. Now consider a self-adjoint
element u in Ab⊗B. For any M ∈ Φ(A), hM being ∗-preserving, hM (u) is
self-adjoint in B. But B is symmetric, so
σ(u) = [
σ(hM (u)) ⊆ R.
M ∈Φ(A)
(cid:3)
Hence, Ab⊗B is symmetric.
Remark 4.5. Note that one can also prove the above lemma using an argument
similar to that in [8, Corollary 3.3].
Theorem 4.6. If A is a subhomogeneous C ∗-algebra, then for any C ∗-algebra
B, Ab⊗B is symmetric.
Proof. Since Ab⊗B can be isometrically embedded in A∗∗b⊗B∗∗ as a closed ∗-
subalgebra, it is sufficient to show that A∗∗b⊗B∗∗ is symmetric. Let A be n-
subhomogeneous, then A∗∗ is a direct sum of type Im von Neumann algebras
for m ≤ n [4, Theorem IV.1.4.6]. Also each type Im von Neumann algebra is
isomorphic to Mm⊗C, where Mm is the set of m × m complex matrices and
C is a commutative von Neumann algebra [4, III.1.5.12]. Thus, A∗∗b⊗B∗∗ is ∗-
isomorphic (not necessarily isometrically) to a direct sum of some Mm(C)b⊗B∗∗.
For each m, Mm(C) is isomorphic to Mmb⊗C; so, using the commutativity and
associativity of the operator space projective norm, we get Mm(C)b⊗B∗∗ is ∗-
isomorphic to Cb⊗(Mmb⊗B∗∗). Note that, Lemma 4.3 gives Mmb⊗B∗∗ is an
4.4, Mm(C)b⊗B∗∗ is symmetric. Hence, A∗∗b⊗B∗∗ is symmetric being the direct
Remark 4.7. If A is commutative and B is any C ∗-algebra, then, by [8, Corol-
lary 3.3], A ⊗γ B is symmetric. However, the symmetry of A ⊗γ B when A is
subhomogeneous and B is any C ∗-algebra follows as in Theorem 4.6.
operator algebra with an isometric involution and is symmetric; so, by Lemma
sum of symmetric Banach ∗-algebras [22, Theorem 11.4.2]
(cid:3)
References
[1] S. D. Allen, A.M. Sinclair and R. R. Smith, The ideal structure of the Haagerup tensor
product of C ∗-algebras, J. Reine Angew. Math. 442(1993), 111-148.
[2] R. J. Archbold, A counter example for commutation in tensor products of C ∗-algebras,
Proc. Amer. Math Soc. 81(4)(1981), 562-564.
[3] R. J. Archbold, E. Kaniuth, G. Schlichting and D. W. B. Somerset, Ideal space of the
Haagerup tensor product of C ∗-algebras, Internat. J. Math. 8(1997), 1-29.
IDEALS IN O.S. PROJECTIVE TENSOR PRODUCT
11
[4] B. Blackadar, Operator algebras- Theory of C ∗-algebras and von-Neumann algebras,
Springer-Verlag, 2006.
[5] D. P. Blecher, Tensor products which do not preserve operator algebras, Math. Proc.
Camb. Phil. Soc. 108(1990), 395-403.
[6] D. P. Blecher and C. LeMerdy, On quotients of function algebras and operator algebra
structures on l
p, J. Operator Theory 34(1995), 315-346.
[7] D. P. Blecher and V. I. Paulsen, Tensor product of operator spaces, J. Funct. Anal.
99(1991), 262-292.
[8] R. A. Bonic, Symmetry in group algebras of discrete groups, Pacific J. Math 11(1)(1961),
73-94.
[9] E. G. Effros and Z. J. Ruan, On matricially normed spaces, Pacific J. Math. 132(2)(1988),
243-264.
[10] E. G. Effros and Z. J. Ruan, A new approach to operator space, Canad. Math. Bull.
34(3)(1991), 329-337.
[11] U. Haagerup, The Grothendieck inequality for bilinear forms on C ∗-algebras, Advances
in Math. 56(1985), 93-116.
[12] W. Hauenschild, E. Kaniuth and A. Kumar, Ideal structure of Beurling algebras on [F C]−
groups, J. Funct. Anal. 51(1983), 213-228.
[13] R. Jain and A. Kumar, Operator space tensor products of C ∗-algebras, Math. Zeit.
260(2008), 805-811.
[14] R. Jain and A. Kumar, Operator space projective tensor product: Embedding into second
dual and ideal structure, arXiv:1106.2644v1 [math.OA] 14 Jun 2011.
[15] E. Kaniuth, A course in commutative Banach algebras, Springer-Verlag, 2009.
[16] A. Kumar, Operator space projective tensor product of C ∗-algebras, Math. Zeit.
237(2001), 211-217.
[17] A. Kumar and A. M. Sinclair, Equivalence of norms on operator space tensor products
of C ∗-algebras, Trans. Amer. Math. Soc. 350(1998), 2033-2048.
[18] A. Lebow, Maximal ideals in tensor products of Banach algebras, Bull. Amer. Math. Soc.
74(11)(1968), 1020-1022.
[19] H. Leptin and D. Poguntke, Symmetry and non symmetry for locally compact groups, J.
Funct. Anal. 33(1979), 119-134.
[20] R. J. Loy, Identities in tensor product of Banach algebras, Bull. Aust. Math. Soc. 2(1970),
253-260.
[21] J. Ludwig, A class of symmetric and a class of Wiener group algebras, J. Funct. Anal.
31(1979), 187-194.
[22] T. W. Palmer, Banach algebras and the general theory of ∗-algebras II, Cambridge Uni-
versity Press, 2001.
[23] R. R. Smith, Completely bounded module maps and the Haagerup tensor product, J.
Funct. Anal. 102(1991), 156-175.
[24] M. Takesaki, Theory of operator algebras I, Springer-Verlag, 2000.
[25] J. Tomiyama, Applications of Fubini type theorem to the tensor products of C ∗-algebras,
Tohoku Math J. 19(2)(1967), 213-226.
[26] S. Wassermann, The slice map problem for C ∗-algebras, Proc. London Math. Soc.
32(3)1976, 537-559.
[27] S. Wassermann, On tensor products of certain group C ∗-algebras, J. Funct. Anal.
23(1976), 239-254.
Department of Mathematics, Lady Shri Ram College for Women, New Delhi-
110024, India.
E-mail address: ranjanaj [email protected]
Department of Mathematics, University of Delhi, Delhi-110007, India.
E-mail address: [email protected]
|
1811.00456 | 1 | 1811 | 2018-11-01T15:52:03 | Higher variations for free L\'evy processes | [
"math.OA",
"math.PR"
] | For a general free L\'evy process, we prove the existence of its higher variation processes as limits in distribution, and identify the limits in terms of the L\'evy-It\^o representation of the original process. For a general free compound Poisson process, this convergence holds almost uniformly, This implies joint convergence in distribution to a $k$-tuple of higher variation processes, and so the existence of $k$-fold stochastic integrals as almost uniform limits. If the existence of moments of all orders is assumed, the result holds for free additive (not necessarily stationary) processes and more general approximants. In the appendix we note relevant properties of symmetric polynomials in non-commuting variables. | math.OA | math | HIGHER VARIATIONS FOR FREE L ´EVY PROCESSES
MICHAEL ANSHELEVICH, ZHICHAO WANG
ABSTRACT. For a general free L´evy process, we prove the existence of its higher variation processes
as limits in distribution, and identify the limits in terms of the L´evy-Ito representation of the original
process. For a general free compound Poisson process, this convergence holds almost uniformly,
This implies joint convergence in distribution to a k-tuple of higher variation processes, and so the
existence of k-fold stochastic integrals as almost uniform limits. If the existence of moments of all
orders is assumed, the result holds for free additive (not necessarily stationary) processes and more
general approximants.
In the appendix we note relevant properties of symmetric polynomials in
non-commuting variables.
8
1
0
2
v
o
N
1
]
.
A
O
h
t
a
m
[
1
v
6
5
4
0
0
.
1
1
8
1
:
v
i
X
r
a
1. INTRODUCTION
A free (additive) L´evy process (in law; we will typically omit this qualifier) is a family of self-
adjoint random variables {X(t) : t > 0} affiliated to a non-commutative probability space (A, τ )
which starts at zero, has free, stationary increments, and is stochastically continuous:
free,
(a) X(0) = 0,
(b) For all n ∈ N and t0 < t1 < . . . < tn, {X(t0), X(t1) − X(t0), . . . , X(tn) − X(tn−1)} are
(c) The distribution of the increment X(t + h) − X(t) depends only on h (and will be denoted
(d) For all ε > 0, limh→0 µh(x > ε) = 0.
µh),
The distributions of increments of a free L´evy process form a semigroup with respect to the additive
free convolution ⊞, and so are ⊞-infinitely divisible. This implies that the Voiculescu transform of
the distribution µt of X(t) has the form
(1)
ϕµt(z) = tη + t
a
z
+ tZR(cid:20) z2
z − x − z − x1[−1,1](x)(cid:21) dρ(x),
where η ∈ R, a ∈ R+, and ρ is a L´evy measure. Barndorff-Nielsen and Thorbjørnsen proved that a
free L´evy process has a free L´evy-Ito decomposition.
Theorem 1 (Theorems 6.4, 6.5 in [BNT05]). Let {X(t) : t > 0} be a free L´evy process, with the
generating triple (η, a, ρ) as above. Then, X(t) is equal in distribution to a sum of three freely
Date: November 2, 2018.
2010 Mathematics Subject Classification. Primary 46L54; Secondary 60F05, 60G51.
This work was supported in part by a Simons Foundation Collaboration Grant.
Parts of this article form part of the second author's 2018 Master's thesis at Texas A&M University.
1
2
MICHAEL ANSHELEVICH, ZHICHAO WANG
independent parts. In general,
(2) X(t) d= ηt1A0 + √aS(t)
+ lim
ǫց0(cid:16)Z(0,t]×{x>ǫ}
In particular, whenR[−1,1] xρ(dx) is finite and η := η −R 1
X(t) d= ηt1A0 + √aS(t) +Z(0,t]×R
xdM(t, x) −Z(0,t]×{ǫ<x61}
(3)
−1 xρ(dx), then
xdM(t, x).
x(Leb ⊗ ρ)(dt, dx)1A0(cid:17).
Here, S(t) is the free Brownian motion (in some W ∗-probability space (A0, τ 0)) and M is a free
Poisson random measure on the measure space (R+ × R,B(R+ × R), Leb ⊗ ρ) with values in
(A0, τ 0). The limit is taken in probability.
In the representation in the theorem above, define the k'th variation of the process by
(4)
X (k)(t) = atδk,21A +Z(0,t]×R
xkdM(t, x).
We will show that these objects are well defined, and again form a free L´evy process. Later in the
article we will define the corresponding object when xk is replaced by a more general function p(x).
Our first main result concerns convergence in distribution to a higher variation process.
Theorem 2. For each N ∈ N, let {Xi,N : i ∈ N} be free, identically distributed, self-adjoint ran-
dom variables affiliated to (A, τ ). Suppose that for t > 0,
Then for each k,
the limits being taken in distribution.
lim
N→∞
lim
N→∞
Xi,N
d= X(t).
X k
i,N
d= X (k)(t),
[N t]Xi=1
[N t]Xi=1
We next discuss joint convergence in distribution. In the non-commutative case, there is at this
point no universally accepted definition of this notion. Recall the following.
Definition 1. A family of self-adjoint operators (a1,N , . . . , ak,N ) affiliated to a non-commutative
probability space (A, τ ) converges to (a1, . . . , ak) jointly in moments if for any non-commutative
self-adjoint polynomial P (x1, . . . , xk), τ [P (a1,N , . . . , ak,N )] is well-defined and
τ [P (a1,N , . . . , ak,N )] → τ [P (a1, . . . , ak)]
The family converges jointly in distribution if for any P as above,
P (a1,N , . . . , ak,N ) → P (a1, . . . , ak)
in distribution (see [MS13] for a related notion).
HIGHER VARIATIONS FOR FREE L ´EVY PROCESSES
3
Recall that convergence in distribution and convergence in moments coincide for bounded opera-
tors, but in general neither implies the other.
The next result applies to free additive processes whose increments are not necessarily stationary.
Theorem 3. For each N ∈ N, let {Xi,N : i ∈ N} be free self-adjoint random variables affiliated to
(A, τ ) all of whose moments are finite. Suppose that for t > 0,
converges in moments to X(t) as N → ∞. Suppose in addition that
Xi,N
[N t]Xi=1
(5)
NXi=1
τ [X k
i,N ]2 → 0
[N t]Xi=1
Xi,N ,
[N t]Xi=1
X 2
i,N , . . . ,
[N t]Xi=1
as N → ∞, for all k. Then there exist free additive processes(cid:8)X (j)(t)(cid:9) such that we have joint
convergence in moments
X k
i,N →(cid:0)X(t), X (2)(t), . . . , X (k)(t)(cid:1)
i,N ] → 0 andPN
i=1 τ [X 2
as N → ∞.
Remark 1. For triangular arrays of centered random variables with finite variance, the standard
condition for convergence is max16i6N τ [X 2
i,N ] 6 c < ∞, see for example
Section 22 in [Lo`e77]. The assumption (5) is clearly significantly stronger. On the other hand, it
is significantly weaker that assuming that all Xi,N are identically distributed. In the latter case, the
result follows from the limit theorem 13.1 in [NS06], itself based on a result of Speicher [Spe90].
The second case where we can prove joint convergence is when individual convergence holds in
probability. In the following theorem, we actually have almost uniform convergence. Recall that in
the commutative case, by Egorov's theorem this mode of convergence corresponds to the conver-
gence almost surely.
Theorem 4. Let ρ be a finite probability measure, and
X(t) =Z(0,t]×R
x dM(t, x)
the corresponding free compound Poisson process. Then for Xi,N = X( i
N ) − X( i−1
N ), we have
the limit being taken almost uniformly.
lim
N→∞
X k
i,N = X (k)(t),
[N t]Xi=1
We expect similar convergence, in probability, for general free L´evy processes. At this point we
have the following partial result.
4
MICHAEL ANSHELEVICH, ZHICHAO WANG
Theorem 5. Let {X(t) : t > 0} be a free L´evy process whose increments have symmetric distribu-
tions, and Xi,N = X( i
N ) − X( i−1
N ). Then
lim
N→∞
[N t]Xi=1
X 2
i,N = X (2)(t),
the limit being taken in probability.
Corollary 6. For a free compound Poisson process {X(t) : t > 0} and increments Xi,N as above,
we have joint convergence in distribution
[N t]Xi=1
Xi,N ,
[N t]Xi=1
X 2
i,N , . . . ,
[N t]Xi=1
X k
i,N →(cid:0)X(t), X (2)(t), . . . , X (k)(t)(cid:1)
as N → ∞.
Corollary 7. Let {Xi,N : 1 6 i 6 N, N ∈ N} be as in either Theorem 3 or Corollary 6. Then for
t > 0,
(6)
lim
N→∞
X16i(1),i(2),...,i(k)6[N t]
i(1)6=i(2),i(2)6=i(3),...,i(k−1)6=i(k)
Xi(1),N Xi(2),N . . . Xi(k),N
=
kXj=1
(−1)k−j Xm1,...,mj >1
m1+...+mj =k
X (m1)(t) . . . X (mj )(t).
Here under the assumptions of Theorem 3 the limit is in moments, while under the assumptions of
Corollary 6 the limit is almost uniformly, and so also in distribution.
It was shown in Proposition 1 of [Ans00] that for free L´evy processes with bounded, centered
increments, the limits (in norm) of the left-hand side of (6) and of
(7)
X16i(1),i(2),...,i(k)6[N t]
{i(1),i(2),...,i(k)}=k
Xi(1),N Xi(2),N . . . Xi(k),N .
coincide. These limits should be interpreted as the free stochastic integral
Z[0,t]k
dX(s1) . . . dX(sk).
See the end of the introduction, and the appendix, for the explanation of why the expression (6) is
more appropriate in the free case.
Prior results. The initial motivation for our analysis was the article [AT86] by Avram and Taqqu.
We briefly compare some of their results with ours; the reader should consult their article for more
details. Let {X(t)} be a L´evy process, and define its higher variations pathwise using jumps. Note
that such a definition is unavailable in the non-commutative case. On the other hand, while the
HIGHER VARIATIONS FOR FREE L ´EVY PROCESSES
5
classical version of the representation (4) is surely known, we have not found it in the literature. Let
{Xi,N : 1 6 i 6 N, N ∈ N} be a triangular array with i.i.d. rows, such that
in distribution as N → ∞. Then a multivariate limit theorem implies that
NXi=1
Xi,N → X(t)
(8)
[N t]Xi=1(cid:0)Xi,N , X 2
i,N , . . . , X k
i,N(cid:1) →(cid:0)X(t), X (2)(t), . . . , X (k)(t)(cid:1)
jointly in distribution. At this point, in the non-commutative case such a theorem is only available
for convergence in moments. On the other hand, we actually prove Theorem 2 not just for powers
but for polynomials, that is, linear combinations of powers. For commuting variables, convergence
in distribution of linear combinations is equivalent to joint convergence in distribution (an easy
exercise left to the reader). So the appropriate commutative analog of Theorem 2 also implies the
joint convergence in (8).
Next, recall that the elementary symmetric polynomial
ek(x1, . . . , xN ) =
X16i(1)<i(2)<...<i(k)6N
xi(1)xi(2) . . . xi(k)
is a polynomial Pk(p1, . . . , pk) in the power sum symmetric polynomials
(the polynomial Pk can be written down explicitly). Consequently,
X
16i(1)<i(2)<...<i(k)6[N t]
converges in distribution as N → ∞. Its limit is naturally identified with the multiple integral
xj
i
pj(x1, . . . , xN ) =
NXi=1
Xi(1),N Xi(2),N . . . Xi(k),N = Pk
[N t]Xi=1
Z06s1<s2<...<sk6t
dX(s1) dX(s2) . . . dX(sk).
Xi,N ,
[N t]Xi=1
X 2
i,N , . . . ,
[N t]Xi=1
X k
i,N
Note that as explained in the appendix, if the variables {xi} do not commute, ek is not a polynomial
in the pj's. Its natural replacement in the non-commutative setting is
ek(x1, . . . , xN ) =
X16i(1),i(2),...,i(k)6N
i(1)6=i(2),i(2)6=i(3),...,i(k−1)6=i(k)
xi(1)xi(2) . . . xi(k)
used in equation (6).
Motivated by [RW97], the first author studied related objects in [Ans00], but only for the case of
free L´evy processes with compactly supported distributions. We are not aware of other sources
where these specific topics are studied in the free probability setting. See however the study of
homogeneous sums in [DN14, Sim15].
6
MICHAEL ANSHELEVICH, ZHICHAO WANG
The article is organized as follows. After the introduction and background in Section 2, Section 3
treats, for general free L´evy processes, convergence in distribution to the higher variation pro-
cesses, and their generalization from powers to more general continuous functions. The key result
is Theorem 17. Section 4 treats joint convergence in moments for more general additive processes.
Section 5 contains results about almost uniform convergence and convergence in probability, as
well as an alternative definition of joint convergence in distribution for non-commuting variables.
Finally, in the appendix we explain which symmetric polynomials in non-commuting variables can
be expressed in terms of the basic power sum symmetric polynomials.
Acknowledgements. The authors are grateful to Matthieu Josuat-Verg`es for the references in the
Appendix.
2. BACKGROUND AND THE FREE POISSON RANDOM MEASURE
2.1. Unbounded Operators and Affiliated Operators. A W ∗-probability space is a pair (A, τ ),
where A is a von Neumann algebra acting on a Hilbert space and τ is a faithful normal tracial state
on A. Throughout most of the paper, we will work with possibly unbounded operators affiliated to
A. A self-adjoint operator a is affiliated to A if all of its spectral projections are in A. Equivalently,
for any bounded Borel function, f (a) ∈ A. We denote the collection of all self-adjoint operators
affiliated to A by Asa. A general closed, densely defined operator a is affiliated to A if in its polar
decomposition a = u a, we have u ∈ A and a ∈ Asa. The collection of all such operators is
denoted by A. Murray and von Neumann [MVN36] proved that A is an algebra, that is, if a, b ∈ A,
then a + b and ab are densely defined and closable, and their closures are in A.
For a ∈ Asa, its distribution is the unique probability measure µa on R such that for any bounded
Borel function,
(9)
τ [f (a)] =ZR
f (x) dµa(x).
Definition 2. ([BNT02]) Let (A, τ ) be a W ∗-probability space and (an)n∈N be a sequence of op-
erators affiliated with A. We say that an → a in probability if an − a → 0 in distribution as
n → ∞.
Here, a := √a∗a, which is self-adjoint. When an and a are self-adjoint operators affiliated with
A, an → a in probability if and only if an − a converges to zero in distribution, i.e. the distribution
of an − a as a probability measure on R converges weakly to probability measure δ0.
We list the following proposition for completeness. Compare with Proposition 2.18 in [BNT02]
Proposition 8. The following are equivalent.
(a) an → a in probability.
(b) ∀ε > 0, the traces of the spectral projections τ [1(ε,∞)(an − a)] → 0.
(c) Denote
N (ε, δ) =nb ∈ A : ∃ projection p ∈ A such that τ [1 − p] < δ, bp ∈ A,kbpk < εo .
Then ∀ε, δ > 0, for sufficiently large n, an − a ∈ N (ε, δ).
This mode of convergence is also called convergence in measure.
HIGHER VARIATIONS FOR FREE L ´EVY PROCESSES
7
2.2. Freely infinitely divisible distributions and limit theorems. As mentioned in the introduc-
tion, a probability measure µ on R is ⊞-infinitely divisible if and only if its Voiculescu transform
has a representation
(10)
ϕµ(z) = η +
where η ∈ R, a ∈ R+, and ρ is a L´evy measure, that is,
z − x − z − x1[−1,1](x)(cid:21) dρ(x),
a
z
+ZR(cid:20) z2
ρ({0}) = 0 and ZR
ϕµ(z) = γ +ZR
min(1, x2) dρ(x) < ∞.
1 + xz
z − x
dσ(x).
ϕµ also has an alternative representation
(11)
For future reference, we record the relation between the generating triple (a, η, ρ) and the generating
pair (γ, σ) for the same measure µ:
(12)
and, conversely,
(13)
x2
1 + x2 ρ(dx)
1
σ(dx) = aδ0(dx) +
γ = η −ZR
x(cid:20)1[−1,1](x) −
1 + x2(cid:21) dρ(x)
a = σ({0})
η = γ +ZR\{0}
ρ(dx) =
1 + x2
x2
1 + x2
x
(cid:20)1[−1,1](x) −
1
1 + x2(cid:21) dσ(x)
1R\{0}(x)σ(dx).
The following fundamental limit theorem was proved by Bercovici and Pata in [BP99].
Theorem 9. For a sequence of probability measures {µn} and a strictly increasing sequence of
positive integers (kn), the following assertions are equivalent:
(a) the sequence of kn-fold free convolutions µ⊞kn
n
converges weakly to a probability measure
µ;
(b) there exist a finite positive Borel measure σ on R and a real number γ such that
(14)
(15)
and
kn
x2
x2 + 1
dµn(x) w.→ dσ(x)
lim
n→∞
knZR
x
1 + x2 dµn(x) = γ.
The pair of parameters (γ, σ) comes from the Voiculescu transform (11) of µ. This also implies the
⊞-infinite divisibility of µ.
8
MICHAEL ANSHELEVICH, ZHICHAO WANG
2.3. Free Poisson Random Measures.
Definition 3 (Free Poisson Random Measures). Let (Θ,E, ν) be a measure space and put E0 =
{E ∈ E : ν(E) < ∞}. Let further (A, τ ) be a W ∗−probability space and let A+ denote the cone
of positive operators in A. A free Poisson random measure on (Θ,E, ν) with values in (A, τ ) is a
mapping M : E0 → A+ with the following properties:
(a) the distribution of M(E) is a free Poisson distribution Poiss⊞(ν(E));
(b) for mutually disjoint sets A1, ..., An in E0, the random variables M(A1), M(A2), ..., M(An)
are freely independent and M(∪n
j=1 M(Aj).
Here, the free Poisson distribution Poiss⊞(λ) is obtained by the limit in distribution of
j=1Aj) =Pn
(cid:18)(1 −
)δ0 +
λ
N
λ
N
δ1(cid:19)⊞N
,
as N → ∞ (see Lecture 12 in [NS06]). The existence of free Poisson random measures is proved
by Barndorff-Nielsen and Thorbjørnsen in [BNT05]. For an alternative approach, see Remark 3
below.
We next discuss integration with respect to a free Poisson random measure.
Definition 4. Let s be a real-valued simple function in L1(Θ,E, ν) of the form s = Pr
j=1 aj 1Ej ,
where aj ∈ R \ {0} and Ej are disjoint sets from E0. Then, we define the integral of s with respect
to M as
ZΘ
rXj=1
sdM =
ajM(Ej) ∈ A.
ple function in L1(Θ,E, ν). Next, we can extend this integration to general functions in L1(Θ,E, ν).
Lemma 10. [BNT05, Proposition 4.3] Let f be a real-valued function in L1(Θ,E, ν). Choose
a sequence of real-valued simple functions (sn) in L1(Θ,E, ν) which satisfies the assumptions of
converges in probability to a self-adjoint (possibly unbounded) operator affiliated with A. This
operator is independent of the choice of approximating sequence (sn). We denote this operator by
Because M(Ej) are positive in A, the elementRΘ sdM is self-adjoint in A, for any real-valued sim-
the Dominated Convergence Theorem, such that sn(θ) → f (θ), for all θ ∈ Θ. Then, RΘ sndM
RΘ f dM .
The proof of the following lemma follows by the same techniques as Proposition 4.3 and Corol-
lary 4.5 in [BNT05].
Lemma 11. Let f be a real-valued function in L1(Θ,E, ν). Choose a sequence of real-valued func-
tions (fn) in L1(Θ,E, ν) which satisfies the assumptions of the Dominated Convergence Theorem,
such that fn(θ) → f (θ), for all θ ∈ Θ. Then,RΘ fndM converges in probability toRΘ f dM .
In fact, we only use a special measure space with a concrete intensity measure in our situation. Let
D = R+ × R and B(D) be the set of all Borel subsets of D. In our case,
(Θ,E, ν) = (D,B(D), Leb ⊗ ρ),
HIGHER VARIATIONS FOR FREE L ´EVY PROCESSES
9
where ρ is a L´evy measure. The free Poisson random measure M that we will use is defined on
(D,B(D), Leb ⊗ ρ) with values in a W ∗−probability space (A, τ ). Besides, the integration with
respect to this free Poisson measure M we will use is also a special case.
Lemma 12. Let ρ be a L´evy measure on the real line, and let M be a free Poisson random measure
on (D,B(D), Leb ⊗ ρ) with values in the W ∗−probability space (A, τ ). Suppose that p(x) is any
continuous function on R.
(a) For any ǫ > 0 and 0 6 s < t < ∞, the integral
Z(s,t]×{ǫ<x6n}
p(x)M(dt, dx)
converges in probability, as n → ∞, to some self-adjoint operator affiliated with A, which
is denoted by
p(x)M(dt, dx).
Z(s,t]×{ǫ<x<∞}
Z(s,t]×{x6n}
(b) IfR[−1,1] p(x)ρ(dx) < ∞, then for any ǫ > 0 and 0 6 s < t < ∞, the integral
p(x)M(dt, dx)
converges in probability to some self-adjoint operator affiliated with A, as n → ∞. We
denote it by
Z(s,t]×R
p(x)M(dt, dx).
The statement of Lemma 12 is quite similar with Lemma 6.3 of [BNT05]. In the paper [BNT05],
the authors only proved the situation when p(x) = x but their methods in Lemma 6.1 and Lemma
6.2 of [BNT05] still work well for Lemma 12. According to Lemma 6.3 of [BNT05], there are only
two things for us to check. Since ρ is a L´evy measure, we have that
Z(s,t]×{ǫ<x6n} p(x)Leb ⊗ ρ(du, dx) = (t − s)Z{ǫ<x6n} p(x)ρ(dx) < ∞.
IfR[−1,1] p(x)ρ(dx) < ∞, we have that
Z(s,t]×{x6n} p(x)Leb⊗ ρ(du, dx) = (t− s)(cid:20)Z{x61} p(x)ρ(dx) +Z{1<x6n} p(x)ρ(dx)(cid:21) < ∞.
Thus, integralsR(s,t]×{ǫ<x6n} p(x)M(dt, dx) andR(s,t]×{x6n} p(x)M(dt, dx) are well-defined by
Proposition 4.3 of [BNT05]. Then, we can copy the proof of Lemma 6.3 of [BNT05] and replace
the function f (x) = x by arbitrary continuous function p(x) directly to prove Lemma 12. The idea
for proving Lemma 6.3 is employing the Bercovici-Pata bijection to transform the statement into
classical sense and then using Lebesgue's dominated convergence theorem.
10
MICHAEL ANSHELEVICH, ZHICHAO WANG
3. THE HIGHER VARIATIONS OF FREE L ´EVY PROCESSES
Proposition 13. If there exist a finite Borel measure σ and a constant γ such that
N
x2
x2 + 1
dµN (x) w.→ dσ(x)
lim
N→∞
NZR
x
1 + x2 dµN (x) = γ,
(16)
and
(17)
and
w.→ µt, for any
then there exists a family {µt}t>0 of probability measures on R such that µ⊞[N t]
t ∈ [0,∞). Each µt is ⊞-infinitely divisible and its Voiculescu transform is ϕµt(z) = tγ +
tRR
Moreover, there exists a free L´evy process {X(t)}t>0 such that the distribution of each X(t) is µt,
for all t > 0.
1+xz
z−x dσ(x) = tϕµ(z), where µ := µ1 is the distribution of X(1).
N
Proof. By Theorem 9, we know that if there exist a finite Borel measure σ and a constant γ such
that (16) and (17) hold, then µ⊞N
N
w.→ µ1. For any t ∈ [0,∞), we have that
x2
dµN (x) w.→ tdσ(x) =: dσt(x)
x2 + 1
[Nt]
x
[Nt]ZR
lim
N→∞
1 + x2 dµN (x) = t lim
1 + x2 dµN (x) = tγ =: γt.
w.→ µt. Accord-
Therefore, for any t ∈ [0,∞), there exists a probability measure µt such that µ⊞[N t]
ing to Theorem 9, for any t ∈ [0,∞), µt is ⊞-infinitely divisible since the Voiculescu transform of
µt is
N→∞
N
x
NZR
ϕµt(z) = γt +ZR
1 + xz
z − x
dσt(x) = tϕµ(z),
where µ := µ1. Therefore, ϕµt = ϕµt−s + ϕµs, when t > s > 0. In other words, µt = µt−s ⊞ µs.
w.→ δ0, as t → 0. Then, by Remark 6.7 in
Meanwhile, ϕµt → 0 when t → 0, which means µt
[BNT05], we can conclude that there exists a free L´evy process {X(t)}t>0, which is a family of
self-adjoint operators affiliated with some W ∗-probability space (A0, τ 0), such that the distribution
of each X(t) is µt, for all t > 0.
Lemma 14. Let (A, τ ) be a W ∗-probability space. Let a ∈ Asa with distribution µ, and p(x)
be a continuous real-valued function. Then the distribution µ(p) of operator p(a) (obtained via
continuous functional calculus) can be obtained by the following formula:
(cid:3)
ZR
f (p(x))dµ(x) =ZR
f (x)dµ(p)(x),
for any bounded Borel function f : R → R.
Proof. By definition of the distribution, for any bounded Borel function f : R → R,
τ (f (a)) =ZR
f (x)dµ(x).
HIGHER VARIATIONS FOR FREE L ´EVY PROCESSES
Then, f ◦ p(x) is still a bounded Borel function. Thus,
ZR
f (p(x))dµ(x) = τ (f (p(a))) =ZR
f (x)dµ(p)(x).
11
(cid:3)
Generally, Lemma 14 shows how to change variables between different probability measures.
Note the difference between the notation µ(p) in the preceding lemma and ρp in the following one.
Lemma 15. Let p(x) be any real-valued continuous function such that p(0) = 0 and p′(0) exists.
Suppose that M is a free Poisson random measure determined by a L´evy measure ρ on the Borel
measure space (D,B(D), Leb ⊗ ρ) with values in some W ∗-probability space A. If ρp is another
measure defined by
(18)
f (p(x))1R\{0}(p(x))dρ(x),
ZR
f (x)dρp(x) =ZR
for any bounded Borel function f (x) on R, then ρp is a L´evy measure. The free Poisson random
measure M (p) defined by ρp on (D,B(D), Leb ⊗ ρp) has the following relation with M :
(19)
for any t, ǫ > 0, and
(20)
Z(0,t]×{ǫ<x}
Z(0,t]×R
xdM (p)(t, x) d=Z(0,t]×{ǫ<p(x)}
xdM (p)(t, x) d=Z(0,t]×R
provided thatR[−1,1] xdρp(x) < ∞.
p(x)dM(t, x),
p(x)dM(t, x),
∀t > 0,
Proof. Since p(0) = 0, there exists an ε > 0 such that p(x) 6 1 when x 6 ε. Since p′(0) exists,
the function
h(x) :=( p(x)
x , x 6= 0
p′(0), x = 0,
RR
ZR
is continuous on R. First, we show that ρp is a L´evy measure. If f (x) = 1{0}(x), then ρp({0}) =
1{0}(x)dρp(x) is zero by the definition (18). Next, if f (x) = min{1, x2}, then we can get the
following conclusion:
1R\[−1,1](x)dρp(x)
min{1, x2}dρp(x) =ZR
=ZR
6Z{x∈R:ε<p(x)61}
6ZR\[−ε,ε]
1[−1,1](x)x2dρp(x) +ZR
1[−1,1]\{0}(p(x))(p(x))2dρ(x) +ZR
p(x)2dρ(x) +Z ε
−ε6x6εh(x)2Z 1
1dρ(x) + max
−ε
1R\[−1,1](p(x))dρ(x)
h(x)2x2dρ(x) +ZR\[−ε,ε]
x2dρ(x) +ZR\[−ε,ε]
−1
1dρ(x)
1dρ(x) < ∞.
Therefore, ρp is a L´evy measure.
12
MICHAEL ANSHELEVICH, ZHICHAO WANG
since
Second, we show that the relation (20) holds. IfR[−1,1] xdρp(x) is finite, then so isR 1
−1 p(x)dρ(x) < ∞.
1[−1,1]\{0}(p(x)) · p(x)dρ(x) =Z 1
Z 1
−1 xdρp(x) =ZR
−1 p(x)dρ(x),
Thus, the right-hand side and left-hand side of (20) make sense by Lemma 12. According to Lemma
12, we only need to show that
for all t > 0 and n ∈ N. For any N ∈ N, consider mutually disjoint intervals
p(x)dM(t, x),
Z(0,t]×{x:−n6x<n}
EN
N
2nm
,−n +
2n(m − 1)
xdM (p)(t, x) d=Z(0,t]×{x:−n6p(x)<n}
m =h−n +
N (cid:17),
(cid:19) 1
NXm=1(cid:18)−n +
sN (x)dM (p)(t, x) →Z(0,t]×{x:−n6x<n}
N
m (x)
EN
where 1 6 m 6 N and m ∈ N. Then, the simple functions
2n(m − 1)
sN (x) =
converge to f (x) = x, for any x ∈ [−n, n) as N → ∞. Thus,
Z(0,t]×{x:−n6x<n}
in probability. Let
xdM (p)(t, x)
Then, ∪N
m=1J N
m = {x : −n 6 p(x) < n} and {J N
m} are mutually disjoint. The simple functions
J N
m = {x : p(x) ∈ EN
m}, (1 6 m 6 N, m ∈ N).
for any x ∈ {x : −n 6 p(x) < n}, as N → ∞. Therefore, when N → ∞,
in probability. We conclude that it suffices to show the equality in distribution
m (x) → p(x)
J N
N
(cid:19) 1
2n(m − 1)
NXm=1(cid:18)−n +
gN (x)dM(t, x) →Z(0,t]×{x:−n6p(x)<n}
gN (x)dM(t, x) d=Z(0,t]×{−n6x<n}
m . By Definition 4, we know that
p(x)dM(t, x)
sN (x)dM (p)(t, x).
gN (x) =
Z(0,t]×{x:−n6p(x)<n}
Z(0,t]×{x:−n6p(x)<n}
Z(0,t]×{x:−n6x<n}
Z(0,t]×{x:−n6p(x)<n}
m and K N
Let F N
m = (0, t] × EN
m = (0, t] × J N
sN (x)dM (p)(t, x) =
and
gN (x)dM(t, x) =
NXm=1(cid:18)−n +
NXm=1(cid:18)−n +
2n(m − 1)
N
2n(m − 1)
N
m ),
(cid:19) M (p)(F N
(cid:19) M(K N
m ).
HIGHER VARIATIONS FOR FREE L ´EVY PROCESSES
13
By Definition 3, the distribution of M(K N
Poiss⊞(tρp(EN
m )). According to (18), we conclude that, when m 6= N
2 + 1, 0 /∈ EN
m , so
m ) is Poiss⊞(tρ(J N
m )) and the distribution of M (p)(F N
m ) is
ρp(EN
m ) =ZR
=ZR
=ZR
1
m (x)dρp(x)
EN
1
m \{0}(p(x))dρ(x)
EN
1
{x:p(x)∈EN
m}(x)dρ(x) = ρ(J N
m ).
m )) = Poiss⊞(tρp(EN
So, Poiss⊞(tρ(J N
M(K N
and n ∈ N, we can apply the same method and show that
m )), m 6= N
) and M (p)(F N
1+ N
2
1+ N
2
2 + 1. Notice that the coefficients in front of
) are zero. Then, we get the final result (20). In general, for any t, ǫ > 0
Z(0,t]×{ǫ<x<n}
xdM (p)(t, x) d=Z(0,t]×{ǫ<p(x)<n}
p(x)dM(t, x),
to prove equation (19).
(cid:3)
Lemma 16. Let p(x) be any real-valued continuous function such that p(0) = p′(0) = 0 and
p′′(0) = 2c exists. Then whether or notR 1
Proof. Denote
−1 xdρ(x) is finite,R 1
q(x) :=( p(x)
x2 , x 6= 0
c, x = 0.
−1 xdρp(x) is finite.
Then q(x) is a continuous function. So we can check that
1[−1,1](x)xdρp(x)
Z 1
−1 xdρp(x) =ZR
=ZR
=Z 1
6 kqkC([−1,1])Z 1
6 CZR
−1
−1
1[−1,1]\{0}(p(x))p(x)dρ(x)
1[−1,1]\{0}(p(x))p(x)dρ(x) +ZR\[−1,1]
x2dρ(x) +ZR\[−1,1]
min{1, x2}dρ(x) < ∞.
1[−1,1]\{0}(p(x))p(x)dρ(x)
1{0<p(x)61}(x)p(x)dρ(x)
(cid:3)
Theorem 2 follows from the following more general result by taking p(x) = xk.
Theorem 17. For each N ∈ N, let {Xi,N : i ∈ N} be free, identically distributed, self-adjoint
random variables affiliated to (A, τ ). Suppose that for t > 0,
[N t]Xr=1
XN,r
d.→ X(t),
14
MICHAEL ANSHELEVICH, ZHICHAO WANG
where X(t) is a free L´evy process. Let p(x) be any real-valued continuous function such that
p(0) = 0, p′(0) = b, and p′′(0) = 2c. Then, there exists a L´evy process X p(t) such that
(21)
[N t]Xr=1
p (XN,r) d.→ X p(t).
In addition, if X(t) has the L´evy-Ito decomposition (2) with the generating triple (η, a, ρ), then
X p(t) has a representation in the form:
(22)
X p(t) d= bX(t) + act1A0 +Z(0,t]×R
(p(x) − bx)dM(t, x),
where M is a free Poisson random measure coming from the L´evy-Ito decomposition of X(t). This
is the case whether or notR 1
−1 xdρ(x) is finite. In particular, if p′(0) = 0,
X p(t) d= act1A0 +Z(0,t]×R
p(x)dM(t, x).
Proof. Let X(1) be generated by the pair (γ, σ). Let µN and µp
p(XN,r) respectively. Recall that by Lemma 14, RR f (x)dµp
real-valued and bounded Borel function f (x). Let
N be the distributions of XN,r and
N (x) = RR f (p(x))dµN (x), for any
q(x) :=( p(x)−bx
x2
c, x = 0.
, x 6= 0
Then q(x) is a continuous function. Therefore,
lim
N→∞
= t lim
N→∞
[Nt]ZR
NhZR
x
bx
N (x) = t lim
N→∞
NZR
1 + x2 dµp
1 + x2 dµN (x) +ZR(cid:18) p(x)
NZR
1 + x2 dµN (x),
gp(x)
x2
= tbγ + t lim
N→∞
p(x)
1 + p(x)2 dµN (x)
1 + p(x)2 −
bx
1 + x2(cid:19) dµN (x)i
where gp(x) = p(x)+q(x)−b(b+xq(x))p(x)
1+p(x)2
∈ Cb(R) and gp(0) = c. So γp is defined by
1 + x2 dµp
where γ and σ are defined by (14) and (15). Define
γp := lim
N→∞
N (x) = bγ +ZR
gp(x)dσ(x),
x
NZR
h(x) :=( p(x)
x , x 6= 0
b, x = 0.
HIGHER VARIATIONS FOR FREE L ´EVY PROCESSES
15
Then, h(x) is a continuous function and xh(x) = p(x). For any f (x) ∈ Cb(R),
[Nt]ZR
f (x)
x2
x2 + 1
dµp
N (x) = [Nt]ZR
= [Nt]ZR
N→∞−→ tZR
f (p(x))
p(x)2
p(x)2 + 1
dµN (x)
f (p(x))
p(x)2
x2
x2 + 1
p(x)2 + 1
p(x)2 + h(x)2
p(x)2 + 1
f (p(x))
dσ(x).
x2 + 1
x2 dµN (x)
Let hp(x) := p(x)2+h(x)2
p(x)2+1
, which is a positive bounded Borel function on R. We denote by deσ(x)
the measure hp(x)dσ(x). The measure dσp(x) is defined byRR f (x)dσp(x) =RR f (p(x))deσ(x), for
any bounded Borel function f (x). Then,
N
x2
x2 + 1
dµp
N (x) w.→ dσp(x),
Borel measure. Thus, the conclusion (21) follows immediately from Theorem 9. By Proposition
as N → ∞. Since σ is a finite positive Borel measure on R, we know that σp is also a finite positive
13, we know that {X p(t)}t>0 can be a free L´evy process affiliated with some W ∗-probability space.
Denote the free generating triplet of X p(1) by (ap, ηp, ρp).
Next, based on Theorem 1, L´evy process X p(t) has a decomposition in the form of (2) with free
generating triplet (ap, ηp, ρp). Hence, to prove the representation (17) of X p(t), it is necessary
to compute the free generating triplet (ap, ηp, ρp) in terms of free generating pair (γ, σ) or free
1{0}(x)dσp(x) = σ({0})h(0)2 =
generating triplet (a, η, ρ) of X(t). Firstly, ap = σp({0}) = RR
ab2. Secondly, for L´evy measure ρp and any bounded Borel function f (x), we have that
ZR
f (x)dρp(x) =ZR
=ZR
=ZR
=ZR
1 + x2
f (x)
1R\{0}(x)dσp(x)
x2
1 + (p(x))2
p(x)2
f (p(x))
1R\{0}(p(x))
p(x)2 + h(x)2
1 + p(x)2
dσ(x)
f (p(x))1R\{0}(p(x))
1 + x2
x2 dσ(x)
f (p(x))1R\{0}(p(x))dρ(x).
Therefore, ρp is precisely the measure from Lemma 15, and in particular a L´evy measure. Thirdly,
1+x2 )dσp(x), and the corresponding relation
1+x2
x (1[−1,1](x) − 1
by the relation ηp = γp +RR\{0}
16
MICHAEL ANSHELEVICH, ZHICHAO WANG
between η and γ, using also a = σ({0}) and relation (12), we can deduce that
ηp = bγ +ZR
+ZR
= bγ −ZR\{0}
1{p(x)6=0}(x)
p
1
1 + (p(x))2(cid:19) dσ(x)
b
x
h2 + p2
1{p(x)=0}(x)
1{p(x)=0}(x)gp(x)dσ(x)
1{p(x)6=0}(x)gp(x)dσ(x) +ZR
(cid:18)1{−16p(x)61}(x) −
dσ(x) +ZR
1{−16p61}(x)(cid:19) dσ −Z{p(x)>ǫ}
dσ(x) −Z{x>ǫ}
1{0<p(x)61}(x)p(x) − 1{0<x61}(x)bx(cid:19) dρ(x).
1{x=0}(x)cdσ(x)
1{ǫ<p(x)61}(x)
h2 + p2
b
x
b
x
p
+ lim
ǫց0(cid:20)Z{ǫ<p(x)}(cid:18)h2 + p2
p
= bγ + ac + lim
ǫց0(cid:20)ZR
= bη + ac +(cid:18)ZR
dσ(x)(cid:21)
dσ(x)(cid:21)
Note that for some ε > 0, p(x) 6 1 for x 6 ε. So
ZR(cid:12)(cid:12)1{0<p(x)61}(x)p(x) − 1{0<x61}(x)bx(cid:12)(cid:12) dρ(x)
=Z ε
−ε q(x) x2 dρ(x) +ZR(cid:12)(cid:12)1{0<p(x)61,x>ε}(x)p(x) − 1{ε<x61}(x)bx(cid:12)(cid:12) dρ(x)
−ε6x6εq(x)Z ε
x2 dρ(x) + 2Z{x>ε}
dρ(x) < ∞
6 sup
−ε
since ρ is a L´evy measure, and so the expression above makes sense.
Combine three results we got above and recall the general free L´evy-Ito decomposition of X p(t)
with the free generating triplet (ap, ηp, ρp). Let M (p) be the free Poisson random measure on
(D,B(D), Leb ⊗ ρp). Then, we can simplify the last part of L´evy-Ito decomposition of X p(t)
with respect to the free Poisson random measure M (p):
lim
ǫց0(cid:20)Z(0,t]×{x>ǫ}
ǫց0(cid:20)Z(0,t]×{p(x)>ǫ}
ǫց0(cid:20)Z(0,t]×{p(x)>ǫ}
xdM (p)(t, x) −Z(0,t]×{ǫ<x61}
p(x)dM(t, x) − tZR
p(x)dM(t, x) − tZR
xLeb ⊗ ρp(dt, dx)1A0(cid:21)
x1{ǫ<x61}(x)dρp(x)1A0(cid:21)
p(x)1{ǫ<x61}(p(x))dρ(x)1A0(cid:21)
= lim
= lim
.
HIGHER VARIATIONS FOR FREE L ´EVY PROCESSES
17
Here, we employ Lemma 15, integration by substitution with respect to free Poisson random mea-
sures and relation (13). Thus finally,
+ lim
ǫց0(cid:18)ZR
X p(t) d= t(cid:18)bη + ac + lim
ǫց0(cid:20)Z(0,t]×{p(x)>ǫ}
= bhηt1A0 + √aS(t)
ǫց0(cid:16)Z(0,t]×{x>ǫ}
+ lim
+ act + lim
1{ε<p(x)61}(x)p(x) − 1{ε<x61}(x)bx(cid:19) dρ(x)(cid:19) 1A0 + √abS(t)
p(x)dM(t, x) − tZR
xdM(t, x) − tZ{ǫ<x61}
1{ǫ<p(x)61}(x)p(x)dρ(x)1A0(cid:21)
x dρ(x)1A0(cid:17)i
p(x)dM(t, x) − bZ(0,t]×{x>ǫ}
xdM(t, x)(cid:17)
ǫց0(cid:16)Z(0,t]×{p(x)>ǫ}
ǫց0Z(0,t]×{x>ǫ}
= bX(t) + act + lim
(p(x) − bx)dM(t, x)
p(x)(1{p(x)>ǫ} − 1{x>ǫ})dM(t, x).
Here we used the fact that the distribution of S(t) is symmetric. Since
(p(x) − bx)dM(t, x) =Z(0,t]×R
xdM (p(x)−bx)(t, x)
exists by Lemmas 16 and 12, and the functions (p(x) − bx)1x6ε have a uniform integrable bound
and converge to zero pointwise as ε → 0, by Lemma 11 we have
+ lim
ǫց0Z(0,t]×R
Z(0,t]×R
lim
ǫց0Z(0,t]×{x>ǫ}
=Z(0,t]×R
=Z(0,t]×R
(p(x) − bx)dM(t, x)
(p(x) − bx)dM(t, x) − lim
(p(x) − bx)dM(t, x)
ǫց0Z(0,t]×{x≤ǫ}
(p(x) − bx)dM(t, x).
Finally, the functions
p(x)(1{p(x)>ǫ} − 1{x>ǫ}) = −p(x)(1{p(x)≤ǫ} − 1{x≤ǫ})
also have a uniform integrable bound and converge to zero pointwise as ε → 0. Therefore by
Lemma 11,
lim
ǫց0Z(0,t]×R
p(x)(1{p(x)>ǫ} − 1{x>ǫ})dM(t, x) = 0.
(cid:3)
Remark 2. It is natural to consider, more generally, free additive (not necessarily) stationary pro-
cesses approximated by free, non-identically distributed triangular arrays which are infinitesimal,
that is, their distributions µi,N satisfy
lim
N→∞
max
16i6kN
µi,N ({x > ε}) = 0
18
MICHAEL ANSHELEVICH, ZHICHAO WANG
for every ε > 0. The following very simple example shows how without additional assumptions,
the results immediately break down. Let
Xi,N =
1
N
+ (−1)i 1
N α , i = 1, . . . , 2N.
Then clearly the array {Xi,N} is infinitesimal, and limN→∞P[2N t]
i=1 Xi,N = 2t. But
X 2
i,N ∼
2t
N 2α−1
[2N t]Xi=1
diverges for α < 1
do not converge to it. Compare with the remarks on page 494 of [AT86].
2 . So while the quadratic variation of a non-random process is zero, these sums
4. CONVERGENCE IN MOMENTS
For a non-crossing partition π ∈ NC(n), denote
Recall that the free cumulant functional is defined by
Mob(π)τπ[a1, . . . , an],
τ"Yi∈V
ai# .
τπ [a1, . . . , an] = YV ∈π
R[a1, . . . , an] = Xπ∈NC(n)
where Mob is the Mobius function on the lattice of non-crossing partitions. The key property of the
free cumulant functional is that if a1, . . . , ak are free, then
unless u(1) = . . . = u(n).
R[au(1), . . . , au(n)] = 0
Proof of Theorem 3. Note first that by freeness and the free moment-cumulant formula,
X u(1)
i,N , . . . ,
R
[N t]Xi=1
[N t]Xi=1
[N t]Xi=1(cid:16)R(cid:16)X u(1)
[N t]Xi=1 Xπ∈NC(k)
=
=
π6=1k
i,N , . . . , X u(k)
Mob(π)τπhX u(1)
X u(k)
X u(1)+...+u(k)
i,N
[N t]Xi=1
i,N − τ
i,N (cid:17) − τhX u(1)+...+u(k)
i,N i .
i,N , . . . , X u(k)
i,N
i(cid:17)
HIGHER VARIATIONS FOR FREE L ´EVY PROCESSES
19
The absolute value of this expression is bounded by
i,N
τhXPj∈V2
i,N
u(j)
1/2
i2
π6=1k
π6=1k
u(j)
i,N
max
i,N
1/2
[N t]Xi=1
i(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
i2
Mob(π)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
τhXPj∈V u(j)
[N t]Xi=1 YV ∈π
Xπ∈NC(k)
Mob(π)
τhXPj∈V1
[N t]Xi=1
6 Xπ∈NC(k)
(cid:12)(cid:12)(cid:12) ,
16i6[N t](cid:12)(cid:12)(cid:12)XPj∈V u(j)
× YV ∈π\{V1,V2}
τ
i,N
[N t]Xi=1
τ
i,N − Rk
Xi,N → 0
[N t]Xi=1
[N t]Xi=1
Rk
Xi,N → Rk(X(t)).
[N t]Xi=1
lim
N→∞
X k
X k
as N → ∞. Finally, by assumption
which goes to zero as N → ∞, by assumption. So to prove joint convergence in moments, it
suffices to show that the limit
exists for each k. Indeed, applying the derivation above to the case u(1) = . . . = u(k) = 1,
The statement about processes follows as in Proposition 13.
(cid:3)
5. CONVERGENCE IN PROBABILITY
Definition 5. an → a almost uniformly (a.u.) if for any δ > 0, there exists a projection p such that
τ [1 − p] < δ and k(an − a)pk → 0.
We now quote a result from [BV93].
Lemma 18 (Lemma 4.4). Let (A, τ ) be a W ∗-probability space, T1, T2, T ′
orthogonal projections. Suppose T ′
such that
2 ∈ A, and p1, p2 ∈ A
j = Tjpj, for j = 1, 2. Then there exist projections p, q ∈ A
1, T ′
1T ′
(a) (T1T2)p = (T ′
(b) (T1 + T2)q = (T ′
2)q, and
(c) τ [p], τ [q] > τ [p1] + τ [p2] − 1.
2)p,
1 + T ′
We do not have a reference for the following result (compare with [Pet84, Sau91]), and so provide
a short proof.
Lemma 19. Let an → a and bn → b a.u. Then an + bn → a + b and anbn → ab a.u.
20
MICHAEL ANSHELEVICH, ZHICHAO WANG
Proof. The first statement is obvious. For the second, let ε > 0, and choose projections p1, p2, p3, p4
so that
k(an − a)p1k → 0,
k(bn − b)p2k → 0,
p3 6 p1,
kap3k < ∞,
kbp4k < ∞
and τ [p1], τ [p2], τ [p3], τ [p4] > 1 − ε. By Lemma 18, we may choose projections p′, p′′ with
τ [p′], τ [p′′] > 1 − 2ε so that
(an − a)bp′ = (an − a)p1bp4p′,
an(bn − b)p′′ = anp3(bn − b)p2p′′.
Since p3 6 p1, also supn kanp3k < ∞. Finally taking p = p′ ∧ p′′, with τ [p] > 1 − 4ε, we get
k(anbn − ab)pk 6 k(an − a)bpk + kan(bn − b)pk
6 k(an − a)p1kkbp4k + kanp3kk(bn − b)p2k → 0
as n → ∞.
Remark 3. Let ρ be a probability measure on R. In the tracial non-commutative probability space
C = L∞((0, 1] × R, Leb ⊗ ρ), consider the projections P (B) = χB for every Borel set B. Let s be
a semicircular element free from C. Then according to [NS96], the family of operators M : B →
sP (B)s satisfies all the properties of a free Poisson random measure in Definition 3. Next, let
(cid:3)
e(t) =ZR
x P ((0, t] × dx),
meaning that the spectral projections of et are {P ((0, t] × (−∞, x))}. Then {e(t) : t ∈ (0, 1]} is a
process with orthogonal increments, and {se(t)s : t ∈ (0, 1]} is a free compound Poisson process.
Note that
se(t)s =ZR
xs P ((0, t] × dx)s =Z(0,t]×R
x dM(t, x).
Proposition 20. Let Z1, . . . , Zk be bounded and centered, free from a stationary process {e(t)}
with orthogonal increments. Then
NXi=1
em0
i,N Z1em1
i,N Z2 . . . emk−1
i,N Zkemk
i,N → 0
a.u. as N → ∞. Here we denote as usual ei,N = e( i
Proof. Without loss of generality, assume that {e(t)} has the form in Remark 3. For arbitrary ε > 0,
choose T so that for q = χ[0,1]×(−T,T ), we have τ (q) = ρ((−T, T )) > 1 − ε. Then {e′(t) = e(t)q}
is a bounded process with orthogonal increments, and ke′(1)k = T . According to Lemma 18, there
N ) − e( i−1
N ).
j=0 mjε such that
i,N Zkemk
i,N p = (e′
i,N )m0Z1(e′
i,N )m1Z2 . . . (e′
i,N )mk−1Zk(e′
i,N )mkp
On the other hand, according to Theorem 3 from [Ans00],
em0
i,N Z1em1
i,N Z2 . . . emk−1
is a projection p with τ (p) > 1 −Pk
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
NXi=1
The result follows.
i,N )m0Z1(e′
i,N )m1Z2 . . . (e′
(e′
i,N )mk−1Zk(e′
i,N )mk(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
6 42k(maxkZjk)kT Pk
j=0 mj N −k/2.
(cid:3)
in probability as N → ∞. Next, write X(t) = se(t)s as before. By the same reasoning as in
Remark 3,
Z(cid:0)0, [Nt]
N (cid:3)×R
Z(0,t]×R
xk dM(t, x) →Z(0,t]×R
xk dM(t, x)
xk dM(t, x) = se(t)ks.
HIGHER VARIATIONS FOR FREE L ´EVY PROCESSES
21
Proof of Theorem 4. By using the addition part of Lemma 19, we may assume that t ∈ (0, 1]. Note
first that by Lemma 11,
Therefore
[N t]Xi=1
X k
i,N −Z(cid:0)0, [Nt]
N (cid:3)×R
k−1Xj=1 Xm0,m1,...,mj >1
m0+m1+...+mj=k
=
s
[N t]Xi=1
xk dM(t, x) =
[N t]Xi=1
(sei,N s)k −
sek
i,N s
[N t]Xi=1
em0
i,N (s2 − 1)em1
i,N (s2 − 1) . . . emj−1
i,N (s2 − 1)emj
i,N s.
Now note that τ (s2 − 1) = 0 and apply Proposition 20.
(cid:3)
Proof of Corollary 6. By Lemma 19, addition and multiplication are continuous with respect to the
i,No converge to the
topology of a.u. convergence. Thus by Theorem 4, polynomials innP[N t]
corresponding polynomials in(cid:8)X (j)(t)(cid:9) a.u. Finally, convergence a.u. clearly implies convergence
in probability, and by Proposition 2.19 in [BNT02] (see also Proposition 2.1 in [LP97]), conver-
gence in probability implies convergence in distribution.
(cid:3)
i=1 X j
Proof of Corollary 7. According to Corollary 25,
X16i(1),i(2),...,i(k)6[N t]
i(1)6=i(2),i(2)6=i(3),...,i(k−1)6=i(k)
Xi(1),N Xi(2),N . . . Xi(k),N
=
kXj=1
(−1)k−j Xm1,...,mj >1
m1+...+mj =k
[N t]Xi(1)=1
X m1
i(1),N . . .
[N t]Xi(j)=1
X mj
i(j),N .
Now apply either Theorem 3 or Corollary 6.
(cid:3)
See the second author's thesis for a direct proof.
Remark 4. In the case of a process which is not necessarily centered, normalizing it so that
τ [X(t)] = t, a more natural definition of an n-fold stochastic integral ψn, according to Theorem 4
of [Ans00], is
ψn = Xψn−1 +
nXj=2
(−1)j−1
n−jXk=0(cid:18)k + j − 2
j − 2 (cid:19)X (j)ψn−j−k.
22
MICHAEL ANSHELEVICH, ZHICHAO WANG
The recursion
Pn = NXi=1
xi! Pn−1 +
nXj=2
(−1)j−1
n−jXk=0(cid:18)k + j − 2
j − 2 (cid:19) NXi=1
i! Pn−j−k.
xj
for polynomials Pn(x1, . . . , xN , t) can be solved explicitly, but we find the resulting formula com-
plicated and not particularly illuminating, and omit it from the article.
We can similarly upgrade various results proven in [Ans00] for bounded free L´evy processes and
uniform limits to general free compound Poisson processes and almost uniform limits. This applies
to Theorem 1 (stochastic measures corresponding to crossing partitions are zero), Proposition 1 (for
a centered process, stochastic measures corresponding to partitions with inner singletons are zero)
and its corollary on the equality of expressions (6) and (7),
Remark 5. Let µ, ν be probability measures on R, such that µ = µa, ν = µb for free a, b ∈ ( Asa, τ ).
The additive free convolution µ ⊞ ν is the distribution of a + b. If µ is supported on R+ (so that
a is positive), the multiplicative free convolution µ ⊠ ν is the distribution of a1/2ba1/2, which we
identify (since τ is a trace) with the distribution of ab.
According to Proposition 3.5 in [BN08], we have the relation
(23)
(µ⊞t) ⊠ (ν ⊞t) = (µ ⊠ ν)⊞t ◦ D1/t,
where D1/t is the dilation operator corresponding to multiplying the operator by t. Note that in the
proposition, the relation is stated for t > 1, but the same argument shows that it holds whenever all
the convolution powers on the left-hand side are defined and at least one of them is supported on
R+.
Proposition 21. LetnX (1)
i,N : 1 6 i 6 N, N ∈ No∪nX (2)
i,N : 1 6 i 6 N, N ∈ No ⊂(cid:16)( A, τ )sa(cid:17) be
two triangular arrays with free, identically distributed rows, free from each other, the first of which
consists of positive operators. Denote
and suppose that
X (j)
i,N = X (j)
N ,
j = 1, 2
NXi=1
lim
N→∞
X (j)
N = X (j),
j = 1, 2
in distribution, for some(cid:8)X (1), X (2)(cid:9). Then as N → ∞,
i,N(cid:17)1/2
i,N(cid:16)X (1)
i,N(cid:17)1/2
NXi=1(cid:16)X (1)
X (2)
in distribution, and so also in probability.
→ 0
Proof. Using the identity from the preceding remark,
µ
(cid:16)X (1)
i,N(cid:17)1/2
and so
i,N(cid:16)X (1)
X (2)
i,N(cid:17)1/2 = (µ⊞(1/N )
X (1)
N
) ⊠ (µ⊞(1/N )
X (2)
N
) = (µX (1)
N
⊠ µX (2)
N
)⊞(1/N ) ◦ DN
µ
i=1(cid:16)X (1)
i,N(cid:17)1/2
PN
i,N(cid:16)X (1)
X (2)
i,N(cid:17)1/2(z) = (µX (1)
N
⊠ µX (2)
N
) ◦ DN .
HIGHER VARIATIONS FOR FREE L ´EVY PROCESSES
23
N
⊠ µX (2)
As N → ∞, µX (1)
weakly.
Remark 6. Denote Cµ(z) = zϕµ(1/z) the free cumulant transform. A measure σ is free regular if
N → µX (1) ⊠ µX (2) weakly, and so the distribution above converges to δ0
(cid:3)
Cσ(z) = η′z +ZR(cid:18) 1
1 − zx − 1(cid:19) ν(dx)
for some η′ > 0 and ν((−∞, 0]) = 0. By Proposition 6.2 in [AHS13], if µ is ⊞-infinitely divisible
and symmetric, then
Here µ2 = µ(x2) in our earlier notation, m is the standard free Poisson distribution, and σ is a free
regular measure. Moreover by Theorem 11 from [PAS12], this is equivalent to
µ2 = m ⊠ σ.
Cµ(z) = Cσ(z2).
(cid:0)(µ(cid:3)ν)⊞1/2(cid:1)2
= µ2 ⊠ ν2.
Next, let µ, ν be probability measures on R, such that µ = µa, ν = µb for free a, b ∈ ( Asa, τ ).
Denote by µ(cid:3)ν the distribution of the anti-commutator ab + ba. If µ, ν are both symmetric, it
coincides with the distribution of the commutator i(ab − ba), and satisfies
See [NS98], Lectures 15 and 19 in [NS06], and Corollary 6.5 in [AHS13].
We also note that if in Remark 5, µ is free regular, then by Theorem 4.2 in [AHS13], µ⊞t is the
distribution of a positive operator for all t > 0. So if in addition ν is ⊞-infinitely divisible, the
identity (23) holds for all such t.
Proposition 22. LetnX (1)
i,N : 1 6 i 6 N, N ∈ No ∪nX (2)
i,N : 1 6 i 6 N, N ∈ No ⊂ (cid:16)( A, τ )sa(cid:17)
be two triangular arrays with free, identically distributed rows, free from each other, all of whose
distributions are symmetric. Denote
X (j)
i,N = X (j)
N ,
j = 1, 2
N is ⊞-infinitely divisible and
N = X (j),
j = 1, 2
X (j)
and suppose that the distribution of each X (j)
in distribution, for some(cid:8)X (1), X (2)(cid:9). Then as N → ∞,
NXi=1
lim
N→∞
NXi=1(cid:16)X (1)
NXi=1
i,N X (2)
i,N + X (2)
i,N X (1)
i,N(cid:17) → 0
X (1)
i,N X (2)
i,N → 0
in distribution, and so also in probability, and
(24)
in probability.
24
MICHAEL ANSHELEVICH, ZHICHAO WANG
Proof. Denote by µj,N the distribution of X (j)
N . Using the preceding remark, we may write
where σj,N is a free regular measure, such that
µ2
j,N = m ⊠ σj,N ,
Cµj,N (z) = Cσj,N (z2).
Note that
Thus
Cµ
⊞(1/N)
j,N
(z) =
1
N Cµj,N (z) =
1
N Cσj,N (z2) = Cσ
⊞(1/N)
j,N
(z2).
Next,
(cid:18)(cid:16)µ⊞(1/N )
Therefore
1,N (cid:3)µ⊞(1/N )
2,N
(cid:17)⊞(1/2)(cid:19)2
= m ⊠ σ⊞(1/N )
.
j,N
j,N
(cid:16)µ⊞(1/N )
(cid:17)2
=(cid:16)µ⊞(1/N )
(cid:17)2
1,N
⊠(cid:16)µ⊞(1/N )
2,N
(cid:17)2
= m ⊠ σ⊞(1/N )
1,N
⊠ m ⊠ σ⊞(1/N )
2,N
.
C(cid:16)µ
⊞(1/N)
1,N
(cid:3)µ
⊞(1/N)
2,N
(cid:17)⊞(N/2)(z) = NCσ
⊞(1/N)
1,N
⊠m⊠σ
⊞(1/N)
2,N
(z2).
Applying the relation (23) twice and distributing the dilation, we get
(cid:16)σ⊞(1/N )
1,N
⊠ m ⊠ σ⊞(1/N )
2,N
(cid:17)⊞N
=(cid:0)σ1,N ⊠ m⊞N ⊠ σ2,N(cid:1) ◦ DN 2
=(cid:0)m⊞N ◦ DN(cid:1) ⊠ ((σ1,N ⊠ σ2,N ) ◦ DN ) .
Using the (noncommutative) law of large numbers, or by a direct calculation, m⊞N ◦ DN → δ1, so
these measures converge to δ0 weakly. Therefore their free cumulant transforms converge to zero
pointwise, which implies that
Since the same convergence in probability holds for the commutators
(cid:16)µ⊞(1/N )
i(cid:16)X (1)
NXj=1
1,N (cid:3)µ⊞(1/N )
2,N
(cid:17)⊞(N/2)
→ δ0.
j,N X (2)
j,N − X (2)
j,N X (1)
j,N(cid:17) ,
it holds for their linear combination (24).
Proof of Theorem 5. Let
X(t) = ηt1A0 +√aS(t) + lim
Fix α ∈ (0, 1). Denote
X ′(t) =(cid:18)η −Z{α6x61}
xdM(t, x)−Z(0,t]×{ǫ<x61}
ǫց0(cid:16)Z(0,t]×{x>ǫ}
x ρ(dx)(cid:19) t1A0 + √aS(t)
(cid:3)
x((Leb⊗ ρ)(dt, dx)1A0)(cid:17).
+ lim
ǫց0(cid:16)Z(0,t]×{ε<x<α}
xdM(t, x) −Z(0,t]×{ǫ<x<α}
x((Leb ⊗ ρ)(dt, dx)1A0)(cid:17).
HIGHER VARIATIONS FOR FREE L ´EVY PROCESSES
25
and
X ′′(t) =Z(0,t]×{x>α}
xdM(t, x).
Note that {X ′′(t)} is an (unbounded) free compound Poisson process, X(t) = X ′(t) + X ′′(t),
{X ′(t)} and {X ′′(t)} are free from each other, and all of their distributions are ⊞-infinitely divisible
and symmetric. Then
X 2
i,N =
[N t]Xi=1
[N t]Xi=1(cid:0)X ′
i,N(cid:1)2 +
[N t]Xi=1(cid:0)X ′′
i,N(cid:1)2 +
[N t]Xi=1(cid:0)X ′
i,N X ′′
i,N + X ′′
i,N X ′
i,N(cid:1) .
By Theorem 4, the second term converges almost uniformly to (X ′′)(2) (t). By Proposition 22, the
third term converges to zero in probability. By Theorem 2, for fixed α, the first term converges in
distribution to
(X ′)(2) (t) = at1A +Z(0,t]×(−α,α)
x2dM(t, x).
and
Finally, as α → 0, (X ′)(2) (t) → at1A in probability. Thus, given ε, δ > 0, we may choose α small
i=1(cid:0)X ′
i,N(cid:1)2−at1A ∈ N (ε, δ)
so that (X ′)(2) (t)−at1A ∈ N (ε, δ). Then for sufficiently large N ,P[N t]
i,N − (X ′′)(2) (t) − at1A ∈ N (ε, δ).
X 2
[N t]Xi=1
It remains to note that also
X (2)(t) − (X ′′)(2) (t) − at1A = (X ′)(2) (t) − at1A ∈ N (ε, δ).
(cid:3)
We finish this section with another possible definition of joint convergence in distribution. As al-
ready noted, for commuting variables, convergence in distribution of linear combinations is equiv-
alent to joint convergence in distribution. As pointed out by ´Eduard Maurel-Segala and Maxime
Fevrier, this is not the case for non-commuting variables. However the following matricial version
is its natural replacement. By the well-known linearization trick [HT05] (see also Chapter 10 of
[MS17]), it implied the definition in the introduction; we do not know if they are in general equiv-
alent. We show that convergence in probability implies joint convergence in this possibly stronger
sense as well.
Definition 6. Let
{xi,N : 1 6 i 6 k, N ∈ N} ∪ {xi : 1 6 i 6 k} ⊂ ( Asa, τ ).
We say that (x1,N , . . . , xk,N ) → (x1, . . . xk) jointly in distribution if for any d and any Hermitian
matrices A1, . . . , Ak ∈ Md(C), and any B ∈ Md(C) with ℑB > εI for some ε > 0, the Cauchy
transforms
(I ⊗ τ ) B ⊗ 1 −
Ai ⊗ xi,N!−1
→ (I ⊗ τ ) B ⊗ 1 −
kXi=1
Ai ⊗ xi!−1
kXi=1
in norm in MN (C).
Proposition 23. If for each i, xi,N → xi in probability, then (x1,N , . . . , xk,N ) → (x1, . . . xk) in the
sense of Definition 6.
By assumption and a short argument, for any ε, δ > 0 there is an n such that for N > n, there is a
projection pN with τ [pN ] > 1 − δ and
Ai ⊗ xi,N! B ⊗ 1 −
Ai ⊗ xi!−1
.
kXi=1
and in particular this operator is bounded. By the resolvent identity,
reader's convenience. Note first that for X ∈ Md( Asa),
kXi=1
(cid:13)(cid:13)(B ⊗ 1 − X)−1(cid:13)(cid:13) 6(cid:13)(cid:13)(ℑB)−1(cid:13)(cid:13) ,
Ai ⊗ xi,N!−1
B ⊗ 1 −
− B ⊗ 1 −
Ai ⊗ xi!−1
kXi=1
= B ⊗ 1 −
Ai ⊗ xi,N!−1 kXi=1
kXi=1
kXi=1
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
kXi=1
Ai ⊗ xi,N!−1
kXi=1
Ai ⊗ xi,N! (I ⊗ pN )(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
kXi=1
− B ⊗ 1 −
Ai ⊗ xi −
Ai ⊗ xi −
kXi=1
Thus for some projection qN with the same property,
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
B ⊗ 1 −
< ε
kAik .
kXi=1
Ai ⊗ xi!−1 (I ⊗ qN )(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
kXi=1
6 ε(cid:13)(cid:13)(ℑB)−1(cid:13)(cid:13)2
Ai ⊗ xi!−1 → 0.
kXi=1
kAik .
(cid:3)
26
MICHAEL ANSHELEVICH, ZHICHAO WANG
Proof. The argument in Proposition 2.19 in [BNT02] largely goes through; we outline it for the
In particular, the same estimate holds on each matrix entry on the left-hand side. Applying the rest
of the argument from Proposition 2.19 in [BNT02] entry-wise, it follows that
(I ⊗ τ ) B ⊗ 1 −
Ai ⊗ xi,N!−1
kXi=1
− B ⊗ 1 −
APPENDIX A. SYMMETRIC POLYNOMIALS IN NON-COMMUTING VARIABLES
Symmetric functions in non-commuting variables (not to be confused with non-commutative sym-
metric functions) have been considered in [RS06, BRRZ08] and subsequent work. We need the
following observation, whose explicit statement we could not find in the literature.
Proposition 24. Let pk =PN
i be the basic power sum symmetric polynomials. In the algebra
of non-commutative polynomials Chx1, . . . , xNi, the subalgebra generated by {pk : k > 1} is the
linear span of polynomials
i=1 xk
Pu(x) =
NXi(1),i(2),...=1
neighbors distinct
xu(1)
i(1) xu(2)
i(2) . . .
HIGHER VARIATIONS FOR FREE L ´EVY PROCESSES
27
for all choices of u with coordinates u(i) > 1. Note that these polynomials are obviously linearly
independent. In particular, the elementary symmetric functions
ek = Xi(1)6=i(2)6=...6=i(r)
i(1)+i(2)+...+i(r)=k
xi(1)xi(2) . . . xi(r)
are not in this subalgebra for k > 1.
Proof. Clearly the algebra generated by all pk is the span of all
Qu(x) = pu(1)(x)pu(2)(x) . . . =
NXi(1),i(2),...=1
xu(1)
i(1) xu(2)
i(2) . . . ,
where the i(j) are not necessarily distinct. Denote by Int(n) the interval partitions of [n]. Then we
may re-index these polynomials as
Qπ(x) =
NXi(1),i(2),...,i(r)=1
rYj=1
xVj
i(j) =
pVj(x)
rYj=1
for π = {V1, . . . , Vr} ∈ Int(n) for some n. For u ∈ [N]r, denote ker(u) ∈ P(n) the partition such
that u(i) = u(j) if and only if i, j lie in the same block of ker(u). Note that for V ∈ ker(u), the
notation u(V ) is unambiguous. Also, for π ∈ P(n), let I(π) be the largest interval partition such
that I(π) 6 π. Note that I(π) = τ if π > τ and if V, V ′ are neighboring blocks of τ , they lie in
different blocks of π. Finally, for π = {V1, . . . , Vr} ∈ Int(n), denote
Pπ(x) =
neighbors distinct
xVj
i(j).
rYj=1
NXi(1),i(2),...,i(r)=1
I(π)=τ Xi:ker i=πYV ∈π
τ >σ Xπ∈P(n)
i(V ) = Xτ ∈Int(n)
i(V ) = Xτ ∈Int (n)
Pτ (x).
xV
xV
τ >σ
xV
i(V )
Then for σ ∈ Int(n),
π>σ Xi:ker i=πYV ∈π
Qσ(x) = Xπ∈P(n)
= Xτ ∈Int(n)
τ >σ
NXi(1),i(2),...,i(τ )=1
neighbors distinct YV ∈τ
Then by Mobius inversion on the lattice Int(n), the spans of {Qπ} and of {Pπ} are the same. (cid:3)
Corollary 25. In the notation of the preceding proof,
Pσ = Xπ∈Int(n)
π>σ
(−1)σ−πYV ∈π
pV (x).
In particular,
NXi(1),i(2),...,i(n)=1
neighbors distinct
nYj=1
xi(j) = Xπ∈Int(n)
(−1)n−πYV ∈π
pV (x).
28
MICHAEL ANSHELEVICH, ZHICHAO WANG
Proof. The first statement follows by Mobius inversion, since the Mobius function on the lattice
Int(n) is Mob(σ, π) = (−1)σ−π. The second statement follows from the fact that the left-hand
side is P0n(x).
(cid:3)
REFERENCES
[]Ans00 Michael Anshelevich, Free stochastic measures via noncrossing partitions, Adv. Math. 155 (2000), no. 1,
154 -- 179. MR 1789851 (2001k:46102)
[]AHS13 Octavio Arizmendi, Takahiro Hasebe, and Noriyoshi Sakuma, On the law of free subordinators, ALEA Lat.
[]AT86
Am. J. Probab. Math. Stat. 10 (2013), no. 1, 271 -- 291. MR 3083927
Florin Avram and Murad S. Taqqu, Symmetric polynomials of random variables attracted to an infinitely
divisible law, Probab. Theory Relat. Fields 71 (1986), no. 4, 491 -- 500. MR 833266
[]BNT02 Ole E. Barndorff-Nielsen and Steen Thorbjørnsen, Self-decomposability and L´evy processes in free proba-
bility, Bernoulli 8 (2002), no. 3, 323 -- 366. MR 1913111
[]BNT05
, The L´evy-Ito decomposition in free probability, Probab. Theory Related Fields 131 (2005), no. 2,
[]BN08
[]BP99
[]BV93
197 -- 228. MR 2117952
Serban T. Belinschi and Alexandru Nica, On a remarkable semigroup of homomorphisms with respect
to free multiplicative convolution, Indiana Univ. Math. J. 57 (2008), no. 4, 1679 -- 1713. MR 2440877
(2009f:46087)
Hari Bercovici and Vittorino Pata, Stable laws and domains of attraction in free probability theory, Ann. of
Math. (2) 149 (1999), no. 3, 1023 -- 1060, With an appendix by Philippe Biane. MR 2000i:46061
Hari Bercovici and Dan Voiculescu, Free convolution of measures with unbounded support, Indiana Univ.
Math. J. 42 (1993), no. 3, 733 -- 773. MR 1254116 (95c:46109)
[]DN14
[]BRRZ08 Nantel Bergeron, Christophe Reutenauer, Mercedes Rosas, and Mike Zabrocki, Invariants and coinvari-
ants of the symmetric groups in noncommuting variables, Canad. J. Math. 60 (2008), no. 2, 266 -- 296.
MR 2398749
Aur´elien Deya and Ivan Nourdin, Invariance principles for homogeneous sums of free random variables,
Bernoulli 20 (2014), no. 2, 586 -- 603. MR 3178510
Uffe Haagerup and Steen Thorbjørnsen, A new application of random matrices: Ext(C ∗
group, Ann. of Math. (2) 162 (2005), no. 2, 711 -- 775. MR 2183281 (2009k:46121)
J. Martin Lindsay and Vittorino Pata, Some weak laws of large numbers in noncommutative probability,
Math. Z. 226 (1997), no. 4, 533 -- 543. MR 1484709
red(F2)) is not a
[]HT05
[]LP97
[]Lo`e77 Michel Lo`eve, Probability theory. I, fourth ed., Springer-Verlag, New York-Heidelberg, 1977, Graduate
[]MS13
[]MS17
Texts in Mathematics, Vol. 45. MR 0651017
Tobias Mai and Roland Speicher, Operator-valued and multivariate free Berry-Esseen theorems, Limit the-
orems in probability, statistics and number theory, Springer Proc. Math. Stat., vol. 42, Springer, Heidelberg,
2013, pp. 113 -- 140. MR 3079141
James A. Mingo and Roland Speicher, Free probability and random matrices, Fields Institute Monographs,
vol. 35, Springer, New York; Fields Institute for Research in Mathematical Sciences, Toronto, ON, 2017.
MR 3585560
[]MVN36 F. J. Murray and J. Von Neumann, On rings of operators, Ann. of Math. (2) 37 (1936), no. 1, 116 -- 229.
[]Nel74
[]NS96
[]NS98
[]NS06
MR 1503275
Edward Nelson, Notes on non-commutative integration, J. Functional Analysis 15 (1974), 103 -- 116.
MR 0355628
Alexandru Nica and Roland Speicher, On the multiplication of free N -tuples of noncommutative random
variables, Amer. J. Math. 118 (1996), no. 4, 799 -- 837. MR 98i:46069
, Commutators of free random variables, Duke Math. J. 92 (1998), no. 3, 553 -- 592. MR 1620518
(99d:46084)
, Lectures on the combinatorics of free probability, London Mathematical Society Lecture Note
Series, vol. 335, Cambridge University Press, Cambridge, 2006. MR 2266879 (2008k:46198)
[]PAS12 Victor P´erez-Abreu and Noriyoshi Sakuma, Free infinite divisibility of free multiplicative mixtures of the
Wigner distribution, J. Theoret. Probab. 25 (2012), no. 1, 100 -- 121. MR 2886381
HIGHER VARIATIONS FOR FREE L ´EVY PROCESSES
29
[]Pet84
[]RS06
[]RW97
[]Sau91
[]Sim15
[]Spe90
D´enes Petz, Quantum ergodic theorems, Quantum probability and applications to the quantum theory of
irreversible processes (Villa Mondragone, 1982), Lecture Notes in Math., vol. 1055, Springer, Berlin, 1984,
pp. 289 -- 300. MR 782910
Mercedes H. Rosas and Bruce E. Sagan, Symmetric functions in noncommuting variables, Trans. Amer.
Math. Soc. 358 (2006), no. 1, 215 -- 232. MR 2171230
Gian-Carlo Rota and Timothy C. Wallstrom, Stochastic integrals: a combinatorial approach, Ann. Probab.
25 (1997), no. 3, 1257 -- 1283. MR 98m:60081
Jean-Luc Sauvageot, A note on almost uniform convergence in von Neumann algebras, Quantum probability
& related topics, QP-PQ, VI, World Sci. Publ., River Edge, NJ, 1991, pp. 385 -- 390. MR 1149839
Rosaria Simone, Universality of free homogeneous sums in every dimension, ALEA Lat. Am. J. Probab.
Math. Stat. 12 (2015), no. 1, 213 -- 244. MR 3343483
Roland Speicher, A new example of "independence" and "white noise", Probab. Theory Related Fields 84
(1990), no. 2, 141 -- 159. MR 1030725 (90m:46116)
DEPARTMENT OF MATHEMATICS, TEXAS A&M UNIVERSITY, COLLEGE STATION, TX 77843-3368
E-mail address: [email protected], [email protected]
|
1406.6160 | 2 | 1406 | 2015-04-26T15:03:40 | Asymptotic structure of free Araki-Woods factors | [
"math.OA"
] | The purpose of this paper is to investigate the structure of Shlyakhtenko's free Araki-Woods factors using the framework of ultraproduct von Neumann algebras. We first prove that all the free Araki-Woods factors $\Gamma(H_{\mathbb R}, U_t)^{\prime \prime}$ are $\omega$-solid in the following sense: for every von Neumann subalgebra $Q \subset \Gamma(H_{\mathbb R}, U_t)^{\prime \prime}$ that is the range of a faithful normal conditional expectation and such that the relative commutant $Q' \cap M^\omega$ is diffuse, we have that $Q$ is amenable. Next, we prove that the continuous cores of the free Araki-Woods factors $\Gamma(H_{\mathbb R}, U_t)^{\prime \prime}$ associated with mixing orthogonal representations $U : \mathbb R \to \mathcal O(H_{\mathbb R})$ are $\omega$-solid type ${\rm II_\infty}$ factors. Finally, when the orthogonal representation $U : \mathbb R \to \mathcal O(H_{\mathbb R})$ is weakly mixing, we prove a dichotomy result for all the von Neumann subalgebras $Q \subset \Gamma(H_{\mathbb R}, U_t)^{\prime \prime}$ that are globally invariant under the modular automorphism group $(\sigma_t^{\varphi_U})$ of the free quasi-free state $\varphi_U$. | math.OA | math |
Asymptotic structure of free Araki-Woods factors
by Cyril Houdayer1 and Sven Raum2
Abstract
The purpose of this paper is to investigate the structure of Shlyakhtenko's free Araki-Woods
factors using the framework of ultraproduct von Neumann algebras. We first prove that all the
free Araki-Woods factors Γ(HR, Ut)′′ are ω-solid in the following sense: for every von Neumann
subalgebra Q ⊂ Γ(HR, Ut)′′ that is the range of a faithful normal conditional expectation and
such that the relative commutant Q′ ∩ M ω is diffuse, we have that Q is amenable. Next, we
prove that the continuous cores of the free Araki-Woods factors Γ(HR, Ut)′′ associated with
mixing orthogonal representations U ∶ R → O(HR) are ω-solid type II∞ factors. Finally, when
the orthogonal representation U ∶ R → O(HR) is weakly mixing, we prove a dichotomy result for
all the von Neumann subalgebras Q ⊂ Γ(HR, Ut)′′ that are globally invariant under the modular
automorphism group (σϕU
t ) of the free quasi-free state ϕU .
1
Introduction and statement of the main results
Free Araki-Woods factors were introduced by Shlyakhtenko in [Shl97]. In the context of Voiculescu's
free probability theory, these factors can be regarded as the analogues of the hyperfinite factors
coming from the canonical anticommutation relations (CAR) functor. Alternatively, they can also
be regarded as the analogues of the free group factors in the setting of type III factors.
Following [Shl97], to any orthogonal representation U ∶ R → O(HR) on a separable real Hilbert
space, one associates a von Neumann algebra denoted by Γ(HR , Ut)′′, called the free Araki-Woods
von Neumann algebra. The von Neumann algebra Γ(HR , Ut)′′ comes equipped with a unique free
quasi-free state ϕU that is always normal and faithful (see Subsection 2.2 for a detailed construction).
dim(HR)) when U ∶ R → O(HR) is the trivial representation and Γ(HR , Ut)′′
We have Γ(HR , id)′′ ≅ L(F
is a full type III factor when U ∶ R → O(HR) is not the trivial representation.
Free Araki-Woods factors were first studied using the framework of Voiculescu's free probability
theory. A complete description of their type classification as well as fullness and computation
of their Connes's τ and Sd invariants was obtained in [Shl97, Shl98, Shl99] (see also the survey
[Vae06]). More recently, free Araki-Woods factors were studied using the framework of Popa's
Deformation/Rigidity theory [Pop06a]. This new approach allowed to obtain various indecom-
posability results in [Hou08] and complete metric approximation property and absence of Cartan
subalgebra in [HR10]. Because of their rich structure, free Araki-Woods factors form one of the
most prominent classes of type III factors.
The purpose of this paper is to investigate the asymptotic structure of free Araki-Woods factors
using the framework of ultraproduct von Neumann algebras. Before stating our main results, we
first introduce some terminology.
1Research supported by the ANR Grant NEUMANN and JSPS Invitation Fellowship Program for Research in
Japan FY2014
2Research supported by the ANR Grant NEUMANN
1
We will say that a von Neumann subalgebra Q ⊂ M is with expectation if there exists a faithful
normal conditional expectation EQ ∶ M → Q. We will say that a diffuse von Neumann algebra M
is solid if for every von Neumann subalgebra Q ⊂ M with expectation whose relative commutant
Q′ ∩ M is diffuse, we have that Q is amenable [Oza04]. The first class of solid von Neumann
algebras was discovered by Ozawa in [Oza04]. He showed that every Gromov-word hyperbolic
group G gives rise to a solid von Neumann algebra L(G). More conceptually, Ozawa showed that
every finite diffuse von Neumann algebra satisfying the Akemann-Ostrand property (abbreviated
property (AO) hereafter, see Subsection 2.4) is solid. It was observed in [VV05] that in fact every
diffuse von Neumann algebra satisfying property (AO) is solid.
Let now ω ∈ β(N)∖N be a non-principal ultrafilter. We refer to Subsection 2.3 for the construction of
the ultraproduct von Neumann algebra M ω. We will say that a von Neumann algebra M is ω-solid
if for every von Neumann subalgebra Q ⊂ M with expectation whose relative commutant Q′ ∩ M ω
is diffuse, we have that Q is amenable. Since M sits in M ω as a von Neumann subalgebra with
expectation, any ω-solid von Neumann is obviously solid. As of today, the converse implication is
an open problem3. Ozawa proved in [Oza10] that any finite diffuse von Neumann algebra satisfying
property (AO) is ω-solid. Our first result generalises Ozawa's result [Oza10] to arbitrary diffuse
von Neumann algebras with separable predual satisfying property (AO).
Theorem A. Any von Neumann algebra with separable predual satisfying property (AO) is ω-solid.
In particular, any free Araki-Woods factor Γ(HR , Ut)′′ associated with an orthogonal representation
U ∶ R → O(HR) on a separable real Hilbert space is ω-solid.
The proof of Theorem A combines Ozawa's original argument [Oza04] together with several tech-
niques from [AH12] on the structure of ultraproduct von Neumann algebras. The proof of Theo-
rem A is carried out in Section 3. The fact that all free Araki-Woods factors satisfy property (AO)
was proven in [Hou07, Chapter 4].
It follows from Theorem A that any von Neumann subal-
gebra with expectation and with property Gamma of any free Araki-Woods factor is necessarily
amenable. We also show in Proposition 3.3 that for every ω-solid von Neumann algebra M and
every von Neumann subalgebra Q ⊂ M with expectation and with no amenable direct summand,
the relative commutant Q′ ∩M ω is necessarily discrete and hence equal to Q′ ∩M (see Theorem 2.3).
An interesting motivation for studying ω-solidity in the setting of type III factors is the fact that
Connes's τ-invariant [Con74] is computable for all the ω-solid type III1 factors that possess faithful
normal states with non-amenable centralizer. More precisely, we show in Proposition 3.10 that for
every ω-solid factor M and for every faithful normal state ϕ ∈ M∗ such that the centralizer M ϕ is
a non-amenable II1 factor, Connes's invariant τ (M ) is the weakest topology on R that makes the
map R → Aut(M ) ∶ t ↦ σϕ
Extending [Hou08, Theorem 1.2], we next show that the continuous cores of the free Araki-Woods
factors associated with mixing orthogonal representations U ∶ R → O(HR) are ω-solid and so are
their finite corners.
t continuous.
Theorem B. Let U ∶ R → O(HR) be any orthogonal representation on a separable real Hilbert space
that is the direct sum of a mixing representation and a representation of dimension less than or
3The proof of [Oza04, Proposition 7] requires N0 = M and only shows that any finite diffuse von Neumann algebra
that is solid and that has property Gamma is amenable.
2
equal to 1. Let M = Γ(HR , Ut)′′ be the associated free Araki-Woods factor. Then its continuous core
c(M ) is an ω-solid type II∞ factor.
For the proof of Theorem B, we can no longer rely on property (AO). Instead, we work within
the framework of Popa's Deformation/Rigidity theory [Pop06a] and we apply Popa's spectral gap
rigidity [Pop06b] to the free malleable deformation of the free Araki-Woods factors arising from
second quantisation (see Subsection 2.2 for details). The proof of Theorem B is carried out in
Section 3.
When dealing with weakly mixing orthogonal representations U ∶ R → O(HR), we obtain a dichotomy
result for all von Neumann subalgebras of the free Araki-Woods factors Γ(HR , Ut)′′ that are globally
invariant under the modular automorphism group of the free quasi-free state. This result constitutes
a new feature in the structure theory of type III factors.
Theorem C. Let U ∶ R → O(HR ) be any weakly mixing orthogonal representation on a separable
real Hilbert space and (M, ϕ) = (Γ(HR , Ut)′′, ϕU ) the associated free Araki-Woods factor. Let Q ⊂ M
be any von Neumann subalgebra that is globally invariant under the modular automorphism group
t ) of the free quasi-free state ϕ. Then either Q = C1 or Q is a full non-amenable type III1 factor
(σϕ
such that Q′ ∩ M ω = C1.
Theorem C shows in particular that any amenable von Neumann subalgebra of M that is globally
invariant under the modular automorphism group (σϕ
t ) is necessarily trivial. Note that if the
orthogonal representation U ∶ R → O(HR) is not weakly mixing then the centralizer M ϕ is not
trivial. This shows that the assumption of U ∶ R → O(HR) being weakly mixing is necessary
in Theorem C. The proof of Theorem C is based on the recent work of the first named author
[Hou12a, Hou12b, Hou14] and uses in a novel fashion Popa's asymptotic orthogonality property
[Pop83] in the framework of ultraproduct von Neumann algebras. The proof of Theorem C is
carried out in Section 4.
Acknowledgments
This paper was completed when the first named author was visiting the Research Institute for
Mathematical Sciences (RIMS) in Kyoto during Summer 2014. He warmly thanks Narutaka Ozawa
and the RIMS for their kind hospitality. The authors also thank Stefaan Vaes for useful remarks
regarding a first draft of this manuscript. Finally, the authors thank the anonymous referees for
carefully reading the paper and providing valuable comments.
2 Preliminaries
For a von Neumann algebra M, we will denote by Z(M ) the centre of M, by U(M ) the group of
unitaries in M and by Ball(M ) the unit ball of M with respect to the uniform norm ⋅ ∞.
Let now M be any σ-finite von Neumann algebra and ϕ ∈ M∗ any faithful normal state. We
denote by L2(M, ϕ) (or simply L2(M ) when no confusion is possible) the GNS L2-completion of
3
M with respect to the inner product defined by ⟨x, y⟩ϕ = ϕ(y∗x) for all x, y ∈ M. We denote by
Λϕ ∶ M → L2(M)∶ x ↦ Λϕ(x) the canonical embedding and by Jϕ ∶ L2(M) → L2(M) the canonical
conjugation. We have xΛϕ(y) = Λϕ(xy) for all x, y ∈ M.
ϕ = ϕ(x∗x+ xx∗)1~2 for all x ∈ M. Recall that on Ball(M),
We will write xϕ = ϕ(x∗x)1~2 and x#
ϕ ) coincides with the strong (resp. ∗-strong) topology. When
the topology given by ⋅ϕ (resp. ⋅#
ϕ = τ is a faithful normal tracial state, we will simply write x2 = τ(x∗x)1~2 for all x ∈ M. We will
say that a von Neumann algebra M is tracial if it is endowed with a faithful normal tracial state τ.
2.1 The continuous core of a σ-finite von Neumann algebra
Let (M, ϕ) be any σ-finite von Neumann algebra endowed with a faithful normal state. We denote
by (σϕ
t ) the modular automorphism group with respect to the state ϕ. The centraliser M ϕ of
the state ϕ is by definition the fixed point algebra of (M,(σϕ
t )). The continuous core of M with
respect to ϕ, denoted by cϕ(M), is the crossed product von Neumann algebra M⋊σϕ R. The natural
inclusion πϕ ∶ M → cϕ(M) and the unitary representation λϕ ∶ R → cϕ(M) satisfy the covariance
relation
λϕ(s)πϕ(x)λϕ(s)∗ = πϕ(σϕ
s(x))
for all x ∈ M and all s ∈ R .
Because of Connes's Radon-Nikodym cocycle theorem [Con73, Théorème 1.2.1] (see also [Tak03,
There is a unique faithful normal conditional expectation Eϕ ∶ cϕ(M) → Lϕ(R) satisfying
Eϕ(xλϕ(s)) = ϕ(x)λϕ(s). The semifinite faithful normal trace f ↦ ∫R exp(−t)f(t) on L∞(R) gives
rise to a semifinite faithful normal trace Trϕ on Lϕ(R) via the Fourier transform. The formula
Trϕ = Trϕ○Eϕ extends it to a semifinite faithful normal trace on cϕ(M).
Theorem VIII.3.3]), the semifinite von Neumann algebra cϕ(M) together with its trace Trϕ does
M, there is a canonical surjective ∗-isomorphism Πψ,ϕ ∶ cϕ(M) → cψ(M) such that Πψ,ϕ○ πϕ = πψ
and Trψ○Πψ,ϕ = Trϕ. Note however that Πψ,ϕ does not map the subalgebra Lϕ(R) ⊂ cϕ(M) onto
the subalgebra Lψ(R) ⊂ cψ(M).
not depend on the choice of ϕ in the following precise sense. If ψ is another faithful normal state on
2.2 Free Araki-Woods factors
Let U ∶ R → O(HR) be any orthogonal representation on a separable real Hilbert space. Denote by
H = HR⊗R C the complexified Hilbert space of HR and by U ∶ R → U(H) the corresponding unitary
representation. Let A be the positive selfadjoint closed operator defined on H satisfying Ait = Ut
for all t ∈ R. Then there is an isometric embedding of HR into H given by
HR → H ∶ ξ ↦
ξ,
2
1+ A−11~2
whose image we denote by KR. One can check that KR∩iKR ={0} and KR+iKR is dense in H. We
denote by J the canonical conjugation on H = HR ⊕ iHR and by I = J A−1~2. Then I is an invertible
anti-linear closed operator on H satisfying I = I−1. Oserve that KR = {ξ ∈ dom(T) ∶ Iξ = ξ}. From
now on, we will simply write I ∶ ξ + iη ↦ ξ + iη = ξ − iη for all ξ, η ∈ KR.
4
The full Fock space of H is given by
F(H) = CΩ ⊕ ∞ࣷ
n=1
H⊗n.
We call the vector Ω ∈ F(H) the vacuum vector. For all ξ ∈ H, the left creation operator ℓ(ξ) ∈
B(F(H)) is given by the formulae
ℓ(ξ)Ω = ξ
and
ℓ(ξ)(ξ1 ⊗ ⋯ ⊗ ξn) = ξ ⊗ ξ1 ⊗ ⋯ ⊗ ξn.
Note that ℓ(ξ) = ξ and ℓ(ξ) is an isometry if ξ = 1. Put W(ξ) = ℓ(ξ) + ℓ(ξ)∗ for all ξ ∈ KR.
Following [Shl97], we define the free Araki-Woods factor associated with U ∶ R → O(HR) by
Γ(HR , Ut)′′ ={W(ξ) ξ ∈ KR}′′.
t
. In particular, we have σϕU
The vector state ϕU(x) =⟨xΩ, Ω⟩ on Γ(HR , Ut)′′ is called the free quasi-free state. It is faithful and
= Ad(F(Ut)) for all
one can show that the modular automorphism group of ϕU is given by σϕU
t ∈ R, where F(Ut) = 1 ⊕ ࣷn≥1 U⊗n
The GNS-representation of Γ(HR , Ut)′′ with respect to ϕU is isomorphic with its representation on
F(H) with cyclic vector Ω. It is easy to check that for all n ≥ 1 and all ξ1, . . . , ξn ∈ KR + iKR there
is a unique element W(ξ1 ⊗ ⋯ ⊗ ξn) ∈ Γ(HR , Ut)′′ such that W(ξ1 ⊗ ⋯ ⊗ ξn)Ω = ξ1 ⊗ ⋯ ⊗ ξn. We have
W(ξ) = ℓ(ξ) + ℓ(ξ)∗ for all ξ ∈ KR + iKR. The following proposition describes a Wick-type formula
Proposition 2.1 ([Hou12a, Hou12b, HR10]). Let ξj, ηk ∈ KR + iKR, for j, k ≥ 1. The following
statements are true:
t (W(ξ)) = W(Utξ) for all ξ ∈ KR.
for such elements.
t
(i) The Wick formula W(ξ1 ⊗ ⋯ ⊗ ξn) =
(ii) The product W(ξ1 ⊗ ⋯ ⊗ ξr)W(η1 ⊗ ⋯ ⊗ ηs) equals
n
Q
k=0
ℓ(ξ1)⋯ℓ(ξk)ℓ(ξk+1)∗⋯ℓ(ξn)∗ holds.
W(ξ1 ⊗ ⋯ ⊗ ξr ⊗ η1 ⊗ ⋯ ⊗ ηs) +⟨ξr, η1⟩W(ξ1 ⊗ ⋯ ⊗ ξr−1)W(η2 ⊗ ⋯ ⊗ ηs) .
(iii) We have W(ξ1 ⊗ ⋯ ⊗ ξn)∗ = W(ξn ⊗ ⋯ ⊗ ξ1).
(iv) The linear span of {1, W(ξ1 ⊗ ⋯ ⊗ ξn) ∶ n ≥ 1, ξi ∈ KR + iKR} forms a unital σ-strongly dense
∗-subalgebra of Γ(HR , Ut)′′.
Proof. The proof of (i) is borrowed from [HR10, Lemma 3.2]. We prove the formula by induction
on n. For n ∈{0, 1}, we have W(Ω) = 1 and we already observed that W(ξi) = ℓ(ξi) + ℓ(ξi)∗.
Next, for ξ0 ∈ KR + iKR, we have
W(ξ0)W(ξ1 ⊗ ⋯ ⊗ ξn)Ω = W(ξ0)(ξ1 ⊗ ⋯ ⊗ ξn)
=(ℓ(ξ0) + ℓ(ξ0)∗)ξ1 ⊗ ⋯ ⊗ ξn
= ξ0 ⊗ ξ1 ⊗ ⋯ ⊗ ξn +⟨ξ0, ξ1⟩ ξ2 ⊗ ⋯ ⊗ ξn.
5
Q
0≤j≤r−1
ℓ(ξ1)⋯ℓ(ξj)ℓ(ξj+1)∗⋯ℓ(ξr)∗ℓ(η1)∗⋯ℓ(ηs)∗
ℓ(ξ1)⋯ℓ(ξr)ℓ(η1)⋯ℓ(ηk)ℓ(ηk+1)∗⋯ℓ(ηs)∗
+ Q
0≤k≤s
+⟨ξr, η1⟩
Q
0≤j≤r−1,1≤k≤s
Therefore W(ξ1 ⊗ ⋯ ⊗ ξr)W(η1 ⊗ ⋯ ⊗ ηs) is equal to
ℓ(ξ1)⋯ℓ(ξj)ℓ(ξj+1)∗⋯ℓ(ξr−1)∗ℓ(η2)⋯ℓ(ηk)ℓ(ηk+1)∗⋯ℓ(ηs)∗ .
So, we obtain
W(ξ0 ⊗ ⋯ ⊗ ξn) = W(ξ0)W(ξ1 ⊗ ⋯ ⊗ ξn) −⟨ξ0, ξ1⟩W(ξ2 ⊗ ⋯ ⊗ ξn)
= ℓ(ξ0)∗W(ξ1 ⊗ ⋯ ⊗ ξn) −⟨ξ0, ξ1⟩W(ξ2 ⊗ ⋯ ⊗ ξn) + ℓ(ξ0)W(ξ1 ⊗ ⋯ ⊗ ξn).
Using the assumption for n and n − 1 and the relation ℓ(ξ0)∗ℓ(ξ1) =⟨ξ0, ξ1⟩, we obtain
ℓ(ξ0)∗W(ξ1 ⊗ ⋯ ⊗ ξn) =⟨ξ0, ξ1⟩W(ξ2 ⊗ ⋯ ⊗ ξn) + ℓ(ξ0)∗ℓ(ξ1)∗⋯ℓ(ξn)∗ .
Since ℓ(ξ0)W(ξ1 ⊗ ⋯ ⊗ ξn) gives the last n + 1 terms in the Wick formula at order n + 1 and
ℓ(ξ0)∗ℓ(ξ1)∗⋯ℓ(ξn)∗ gives the first term, we are done.
We now prove (ii). By the Wick formula, we have that W(ξ1 ⊗ ⋯ ⊗ ξr)W(η1 ⊗ ⋯ ⊗ ηs) is equal to
Q
0≤j≤r,0≤k≤s
ℓ(ξ1)⋯ℓ(ξj)ℓ(ξj+1)∗⋯ℓ(ξr)∗ℓ(η1)⋯ℓ(ηk)ℓ(ηk+1)∗⋯ℓ(ηs)∗ .
Recall that we have ℓ(ξr)∗ℓ(η1) =⟨ξr, η1⟩. Therefore the above sum equals
W(ξ1 ⊗ ⋯ ⊗ ξr ⊗ η1 ⊗ ⋯ ⊗ ηs) +⟨ξr, η1⟩W(ξ1 ⊗ ⋯ ⊗ ξr−1)W(η2 ⊗ ⋯ ⊗ ηs).
It is now clear that (i) ⇒ (iii). Moreover, (iv) follows from (iii) using an induction procedure.
2.3 Ultraproduct von Neumann algebras
Fix a non-principal ultrafilter ω ∈ β(N) ∖ N. Let M be any σ-finite von Neumann algebra. Define
Iω(M) ={(xn)n ∈ ℓ∞(N, M) xn → 0 ∗ -strongly as n → ω}
Mω(M) ={(xn)n ∈ ℓ∞(N, M) (xn)n Iω(M) ⊂ Iω(M) and Iω(M)(xn)n ⊂ Iω(M)} .
ideal. Following [Ocn85, Chapter 5], we define the ultraproduct von Neumann algebra M ω by
The multiplier algebra Mω(M) is a C∗-algebra and Iω(M) ⊂ Mω(M) is a norm closed two-sided
M ω = Mω(M)~Iω(M). We denote the image of (xn)n ∈ Mω(M) by (xn)ω ∈ M ω.
For all x ∈ M, the constant sequence (x)n lies in the multiplier algebra Mω(M). We will identify
M with (M + Iω(M))~Iω(M) and regard M ⊂ M ω as a von Neumann subalgebra. The map
Eω ∶ M ω → M ∶ (xn)ω ↦ σ-weak limn→ω xn is a faithful normal conditional expectation. For every
6
well-known proposition.
faithful normal state ϕ ∈ M∗, the formula ϕω = ϕ ○ Eω defines a faithful normal state on M ω.
Observe that ϕω((xn)ω) = limn→ω ϕ(xn) for all (xn)ω ∈ M ω.
Put H = L2(M, ϕ). The ultraproduct Hilbert space Hω is defined to be the quotient of ℓ∞(N, H)
by the subspace consisting in sequences (ξn)n satisfying limn→ωξnH = 0. We denote the image
of (ξn)n ∈ ℓ∞(N, H) by (ξn)ω ∈ Hω. The inner product space structure on the Hilbert space Hω
is defined by ⟨(ξn)ω,(ηn)ω⟩Hω = limn→ω⟨ξn, ηn⟩H. The GNS Hilbert space L2(M ω, ϕω) can be
embedded into Hω as a closed subspace by Λϕω((xn)ω) ↦(Λϕ(xn))ω.
Put xϕ = ϕ(⋅ x) and ϕx = ϕ(x ⋅) for all x ∈ M and all ϕ ∈ M∗. We will be using the following
Proposition 2.2. Let (M, ϕ) be any σ-finite von Neumann algebra endowed with a faithful normal
(i) For every λ > 0 and every (xn)n ∈ ℓ∞(N, M) satisfying limn→ωxnϕ − λϕxn = 0, we have
(xn)n ∈ Mω(M) and (xn)ωϕω = λϕω(xn)ω.
(ii) For every projection e ∈ M ω, there exists a sequence of projections (en)n ∈ Mω(M) such
that e = (en)ω. If M is moreover diffuse, the projections en ∈ M may be chosen such that
ϕ(en) = ϕω(e) for all n ∈ N.
Proof. (i) Let (xn)n ∈ ℓ∞(N, M) such that limn→ωxnϕ − λϕxn = 0. Let (bn)n ∈ Iω(M). We may
assume that max{xn∞,bn∞ ∶ n ∈ N} ≤ 1. Using the Cauchy-Schwarz inequality, for all n ∈ N, we
state.
have
(xnbn#
ϕ)2 = ϕ(b∗n x∗nxnbn) + ϕ(xn bnb∗nx∗n)
≤bnϕx∗nxnbnϕ + λ−1 (xnϕ − λϕxn)(bnb∗nx∗n) + λ−1 ϕ(bn b∗nx∗nxn)
≤bnϕ + λ−1xnϕ − λϕxnbnb∗nx∗n∞ + λ−1b∗nϕb∗nx∗nxnϕ
≤bnϕ + λ−1xnϕ − λϕxn + λ−1b∗nϕ.
Therefore, we obtain limn→ωxnbn#
ϕ = 0 and so (xnbn)n ∈ Iω(M). Likewise, for all n ∈ N, we have
(bnxn#
ϕ)2 = ϕ(x∗n b∗nbnxn) + ϕ(bn xnx∗nb∗n)
≤ (λx∗nϕ − ϕx∗n)(b∗nbnxn) + λ ϕ(b∗n bnxnx∗n) +b∗nϕxnx∗nb∗nϕ
≤λx∗nϕ − ϕx∗nb∗nbnxn∞ + λbnϕbnxnx∗nϕ +b∗nϕ
≤xnϕ − λϕxn + λbnϕ +b∗nϕ.
For the first part of the proof of (ii), see the proof of [Hou14, Proposition 2.4 (3)]. It remains to prove
ϕ = 0 and so (bnxn)n ∈ Iω(M). This shows that (xn)n ∈
Therefore, we obtain limn→ωbnxn#
Mω(M). Moreover, (xn)ωϕω = λϕω(xn)ω by [AH12, Lemma 4.36].
the moreover part of (ii) when M is diffuse. Let p ∈ M ω be any projection and (pn)n ∈ Mω(M)
a sequence of projections such that p = (pn)ω. Let n ≥ 1. Assume that ϕ(pn) ≥ ϕω(p). Since
pnM pn is diffuse, we may choose a projection rn ∈ pnM pn such that ϕ(rn) = ϕω(p). Assume that
7
ϕ(pn) ≤ ϕω(p). Since (1−pn)M(1−pn) is diffuse, we may choose a projection sn ∈(1−pn)M(1−pn)
such that ϕ(sn) = ϕω(p) − ϕ(pn). Put rn = pn + sn.
ϕ = limn→ω ϕ(pn − rn) = 0 and hence (pn − rn)n ∈ Iω(M). Thus, we have
We obtain limn→ωpn − rn2
p =(rn)ω and ϕ(rn) = ϕω(p) for all n ∈ N.
The next theorem will be very useful to prove Theorem A. It is a generalization of [Ioa12, Lemma 2.7]
to arbitrary von Neumann algebras.
Theorem 2.3. Let Q ⊂ M be an inclusion of von Neumann algebras with faithful normal conditional
expectation EQ ∶ M → Q. Assume that Q has separable predual. Denote by z ∈ Z(Q′ ∩ M ω) the
unique maximal central projection such that (Q′ ∩ M ω)z is discrete. Then
• z ∈ Z(Q′ ∩ M ω) ∩ Z(Q′ ∩ M),
• (Q′ ∩(M ω)ϕω)(1 − z) is diffuse for all faithful normal states ϕ ∈ M∗ such that ϕ ○ EQ = ϕ and
• (Q′ ∩ M ω)z =(Q′ ∩ M)z.
We start by proving the following two lemmas.
Lemma 2.4. Let Q ⊂ M be an inclusion of von Neumann algebras with faithful normal conditional
expectation EQ ∶ M → Q. Assume that Q has separable predual. Let ϕ ∈ M∗ be any faithful normal
state such that ϕ ○ EQ = ϕ. Denote by e ∈ Z(Q′ ∩(M ω)ϕω) the unique maximal central projection
such that (Q′ ∩(M ω)ϕω)e is discrete. Then
• e ∈ Z(Q′ ∩(M ω)ϕω) ∩ Z(Q′ ∩ M ϕ) and
• (Q′ ∩(M ω)ϕω)e =(Q′ ∩ M ϕ)e.
tion 2.5]). Put Q = Q′ ∩(M ω)ϕω and denote by e ∈ Z(Q) the unique maximal central projection in
Q such that Qe is discrete. Choose a sequence of projections (en)n ∈ Mω(M) such that e =(en)ω.
Let a = σ-weak limn→ω en ∈ Z(Q′ ∩ M ϕ).
Next, we construct by induction a sequence of projections (fm)m≥1 in Q such that
ϕω(efi) = ϕ(a2), ϕω(efia) = ϕ(a3) and ϕω(efifj) = ϕω(efia), ∀1 ≤ i < j.
Proof. The proof is a generalisation of [Ioa12, Lemma 2.7] (see also the proof of [Hou14, Proposi-
(2.1)
Indeed, assume that f1, . . . , fm ∈ Q have been constructed. For every 1 ≤ j ≤ m, choose a sequence
of projections (fj,n)n ∈ Mω(M) such that fj = (fj,n)ω. Let (xi)i∈N be a ⋅#
ϕ -dense sequence in
Ball(Q). Since e = (en)ω ∈(M ω)ϕω , limn→ωenxi − xien#
ϕ = 0 for all i ∈ N and en → a σ-weakly as
n → ω, we can find an increasing sequence (kn)n in N such that for every n ≥ 1, we have
(P1) ekn ϕ − ϕekn ≤ 1
n,
(P2) ekn xi − xiekn#
ϕ ≤ 1
n for all 1 ≤ i ≤ n,
8
n for all 1 ≤ j ≤ m.
(P3) ϕ(enekn) − ϕ(ena) ≤ 1
n,
(P4) ϕ(enekna) − ϕ(ena2) ≤ 1
n and
(P5) ϕ(enfj,nekn) − ϕ(enfj,na) ≤ 1
Properties (P1) and (P2) together with Proposition 2.2 imply that (ekn)n ∈ Mω(M) and f =
(ekn)ω ∈ Q′ ∩(M ω)ϕω. Property (P3) implies that ϕω(ef) = ϕω(ea) = ϕ(a2), Property (P4) implies
that ϕω(ef a) = ϕω(ea2) = ϕ(a3) and Property (P5) implies that ϕω(efjf) = ϕω(efja) for all
1 ≤ j ≤ m. We can now put fm+1 = f. This finishes the proof of the induction.
Define pm = fme which is a projection in Qe. We have ϕω(pj) = ϕ(a2) and ϕω(pjpm) = ϕ(a3)
countable direct sum of finite dimensional factors, Ball(Qe) is ⋅ϕω -compact. Thus, we may choose
a subsequence (pmk)k≥1 that is ⋅ϕω-convergent in Ball(Qe). By the Cauchy-Schwarz inequality,
for all 1 ≤ j < m. Observe that since Qe is a discrete tracial von Neumann algebra and hence a
for all 1 ≤ j < k, we have
ϕω(pmj pmk) − ϕω(pmj) = ϕω(pmj(pmk − pmj)) ≤pmj − pmkϕω .
Put ε = 1
2f −EeM e(f)ϕω
Taking the limit as (j, k) → ∞ and using (2.1), we obtain ϕ(a2) = ϕ(a3) and so 0 ≤ a ≤ 1 is a
ϕ = ϕ(en) + ϕ(a) − 2ϕ(ena) → 0 as n → ω and so
projection in Q′ ∩ M ϕ. Thus we have en − a2
e =(en)ω = a ∈ Z(Q′ ∩(M ω)ϕω) ∩ Z(Q′ ∩ M ϕ).
It remains to prove that (Q′ ∩ (M ω)ϕω)e = (Q′ ∩ M ϕ)e. Assume by contradiction that this is
not the case and choose a nonzero projection f ∈ (Q′ ∩(M ω)ϕω)e such that f ∉ (Q′ ∩ M ϕ)e. Let
ϕe be the faithful normal state on eM e defined by ϕe = ϕ(e⋅e)ϕ(e)
by EeM e ∶ (eM e)ω → eM e the canonical faithful normal conditional expectation. Recall that
e > 0. Moreover, for all y ∈ Ball(eM e),
ϕe○EeM e = ϕω
we have
e > 0 .
e . Since f ∉(Q′∩M ϕ)e, we have f −EeM e(f)ϕω
e ≥f − EeM e(f)ϕω
. Recall that e ∈ M ϕ. Denote
f − yϕω
n ≥ 1, we have
e , limk→ωf1,kxi − xif1,k#
ϕe = 0 for all i ∈ N and limk→ωf1,k − fj,nϕe =f − fj,nϕω
e and f1 = f ∈ Qe. Next, we construct by induction a sequence of projections
fm ∈ Qe such that fp − fqϕω
e ≥ ε for all p, q ≥ 1 such that p ≠ q. Assume that f1, . . . , fm ∈ Qe
have been constructed. For every 1 ≤ j ≤ m, choose a sequence of projections (fj,n)n ∈ Mω(eM e)
such that fj = (fj,n)ω. Let (xi)i∈N be a ⋅#
ϕe-dense sequence in Ball(Qe). Since f = f1 = (f1,n)ω ∈
((eM e)ω)ϕω
e ≥ 2ε
for all 1 ≤ j ≤ m and all n ∈ N, we can find an increasing sequence (kn)n in N such that for every
(P1) f1,knϕe − ϕef1,kn ≤ 1
n,
(P2) f1,knxi − xif1,kn#
ϕe ≤ 1
n for all 1 ≤ i ≤ n and
(P3) f1,kn − fj,nϕe ≥ ε for all 1 ≤ j ≤ m.
(f1,kn)n ∈ Mω(eM e) and (f1,kn)ω ∈ (Qe)′ ∩((eM e)ω)ϕω
By the same reasoning as before, Properties (P1) and (P2) together with Proposition 2.2 imply that
e = Qe. Moreover, Property (P3) implies
9
of the induction.
that (f1,kn)ω − fjϕω
So, we have constructed a sequence of projections fm ∈ Qe such that fp − fqϕω
such that p ≠ q. This however contradicts the fact that Ball(Qe) is ⋅ϕω
e ≥ ε for all 1 ≤ j ≤ m. We can now put fm+1 =(f1,kn)ω. This finishes the proof
e ≥ ε for all p, q ≥ 1
e -compact and finishes the
proof Lemma 2.4.
Lemma 2.5. Let Q ⊂ M be an inclusion of von Neumann algebras with faithful normal conditional
expectation EQ ∶ M → Q. Assume that Q has separable predual. Let ϕ ∈ M∗ be any faithful normal
state such that ϕ ○ EQ = ϕ. If Q′ ∩(M ω)ϕω = C then Q′ ∩ M ω = C.
Proof. The proof is a straightforward generalisation of [AH12, Theorem 5.2] and so we will only
sketch it.
Assume that Q′ ∩(M ω)ϕω = C. Since (Q′ ∩ M ω)ϕω = Q′ ∩(M ω)ϕω = C, [AH12, Lemma 5.3] shows
that Q′ ∩ M ω = C or Q′ ∩ M ω is a type III1 factor. By contradiction, assume that Q′ ∩ M ω is a
type III1 factor. Choose (ai)i∈N a ⋅#
ϕ -dense sequence in Ball(Q). Proceeding as in the proof of
[AH12, Theorem 5.2], for all n ∈ N and all i, j ∈ {1, 2}, we find elements f(n)ij
conditions of [AH12, Theorem 5.2, Claim 1] with respect to the sequence (ai)i∈N in Ball(Q). As in
[AH12, Theorem 5.2, Claim 2], we obtain that (f(n)ij )n ∈ Mω(M) for all i, j ∈ {1, 2}. Finally, we
obtain a projection g11 ∈ Q′ ∩(M ω)ϕω such that ϕω(g11) ≠ 0, 1. This is a contradiction and finishes
∈ M that satisfy the
the proof of Lemma 2.5.
Proof of Theorem 2.3. Let Q ⊂ M be an inclusion of von Neumann algebras and fix a faithful
normal state ϕ ∈ M∗ such that ϕ ○ EQ = ϕ. By Lemma 2.4, the unique maximal central projection
e ∈ Z(Q′ ∩(M ω)ϕω) such that (Q′ ∩(M ω)ϕω)e is discrete satisfies e ∈ Z(Q′ ∩(M ω)ϕω)∩Z(Q′ ∩ M ϕ)
and (Q′ ∩(M ω)ϕω)e =(Q′ ∩ M ϕ)e. Observe that the projection e may a priori depend on the state
ϕ. However, we will prove that this is not the case and show that the projection e satisfies the
conclusion of Theorem 2.3.
Since (Q′ ∩(M ω)ϕω)e is discrete, choose a family (pi)i∈I of pairwise orthogonal minimal projections
in (Q′ ∩(M ω)ϕω)e such that ∑i∈I pi = e. We have pi ∈(Q′ ∩ M ϕ)e and (Qpi)′ ∩((piM pi)ω)ϕω
pi = Cpi
where ϕpi = ϕ(pi⋅pi)
ϕ(pi)
. Lemma 2.5 applied to the inclusion Qpi ⊂ piM pi implies that
pi(Q′ ∩ M ω)pi =(Qpi)′ ∩(piM pi)ω = Cpi
and hence pi is a minimal projection in e(Q′ ∩ M ω)e. Since ∑i∈I pi = e, we have that e(Q′ ∩ M ω)e
is discrete. Denote by z(e) the central support of the projection e in Q′ ∩ M ω. We obtain that
(Q′ ∩ M ω)z(e) is still discrete. Since z(e) ∈ Z(Q′ ∩ M ω) and Q′ ∩ M ω is globally invariant under
the modular automorphism group (σϕω
t ), we have z(e) ∈ Z(Q′ ∩(M ω)ϕω) and (Q′ ∩(M ω)ϕω)z(e)
is discrete. By Lemma 2.4 and since e ≤ z(e), we obtain e = z(e) ∈ Z(Q′ ∩ M ω) ∩ Z(Q′ ∩ M) and
(Q′ ∩ M ω)e is discrete. Observe that (Q′ ∩ M ω)(1 − e) is diffuse since (Q′ ∩(M ω)ϕω)(1 − e) is a
diffuse subalgebra with expectation. Therefore, e ∈ Z(Q′∩M ω)∩Z(Q′∩M) is the unique projection
such that (Q′ ∩ M ω)e is discrete and (Q′ ∩ M ω)(1 − e) is diffuse and hence e does not depend on
the choice of the faithful normal state ϕ ∈ M∗ satisfying ϕ = ϕ ○ EQ. Thus, the above proof shows
that (Q′ ∩(M ω)ψω)(1 − e) is diffuse for all faithful normal states ψ ∈ M∗ satisfying ψ = ψ ○ EQ.
10
It remains to prove that (Q′ ∩M ω)e =(Q′ ∩M)e. Recall that there exists a family (pi)i∈I of pairwise
orthogonal minimal projections in (Q′ ∩ M ω)e such that ∑i∈I pi = e and pi ∈(Q′ ∩ M ϕ)e for all i ∈ I.
In order to show that(Q′∩M ω)e =(Q′∩M)e, it suffices to prove that eF(Q′∩M ω)eF = eF(Q′∩M)eF
for all finite subsets F ⊂ I, with eF = ∑i∈F pi ∈(Q′ ∩ M ϕ)e.
Assume by contradiction that (Q′ ∩M ω)e ≠(Q′ ∩M)e. Hence there exists a finite subset F ⊂ I such
that eF(Q′ ∩ M ω)eF ≠ eF(Q′ ∩ M)eF . For notational convenience, put q = eF and Q = Q′ ∩ M ω.
Observe that qQq is discrete and finite. Let ϕq be the faithful normal state on qM q defined by
. Recall that q ∈ Q′∩M ϕ. Denote by EqM q ∶(qM q)ω → qM q the canonical faithful normal
ϕq = ϕ(q⋅q)ϕ(q)
q . Since qQq =(Qq)′ ∩(qM q)ω is discrete, finite
conditional expectation. Recall that ϕq ○ EqM q = ϕω
q restricted to qQq is diagonalizable. For every
and hence of type I, the faithful normal state ϕω
eigenvalue λ > 0, we will denote by
Eλ =(xn)ω ∈ qQq (xn)ωϕω
q = λϕω
q(xn)ω
the spectral subspace of (qQq, ϕω
q) corresponding to the eigenvalue λ. Since qQq ≠ q(Q′ ∩ M)q, we
may choose an eigenvalue λ > 0 and a nonzero element f ∈ Ball(Eλ) such that f ∉ q(Q′ ∩ M)q. Since
f ∉ q(Q′ ∩ M)q, we have f − EqM q(f)ϕω
f − yϕω
q > 0. Moreover, for all y ∈ Ball(qM q), we have
q ≥f − EqM q(f)ϕω
q > 0 .
Put ε = 1
2f − EqM q(f)ϕω
ϕq -dense sequence in Ball(Qq). Since f = f1 = (f1,n)ω ∈ Eλ, limk→ωf1,kxi − xif1,k#
q and f1 = f ∈ Ball(Eλ). Next, we construct by induction a sequence
of elements fm ∈ Ball(Eλ) such that fm − fpϕω
q ≥ ε for all m, p ≥ 1 such that m ≠ p. As-
sume that f1, . . . , fm ∈ Ball(Eλ) have been constructed. For every 1 ≤ j ≤ m, choose a sequence
(fj,n)n ∈ Mω(qM q) such that fj,n ∈ Ball(qM q) for all n ∈ N and fj = (fj,n)ω. Let (xi)i∈N be a
⋅#
ϕq = 0 for all
i ∈ N and limk→ωf1,k − fj,nϕq = f − fj,nϕω
q ≥ 2ε for all 1 ≤ j ≤ m and all n ∈ N, we can find an
increasing sequence (kn)n in N such that for every n ≥ 1, we have
(P1) f1,knϕq − λϕqf1,kn ≤ 1
n,
(P2) f1,knxi − xif1,kn#
n for all 1 ≤ i ≤ n and
ϕq ≤ 1
(P3) f1,kn − fj,nϕq ≥ ε for all 1 ≤ j ≤ m.
(f1,kn)n ∈ Mω(qM q) and (f1,kn)ω ∈ Ball(Eλ). Moreover, Property (P3) implies that (f1,kn)ω −
e ≥ ε for all 1 ≤ j ≤ m. We can now put fm+1 =(f1,kn)ω. This finishes the proof of the induction.
fjϕω
So, we have constructed a sequence of elements fm ∈ Ball(Eλ) such thatfm−fpϕω
q ≥ ε for all m, p ≥ 1
such that m ≠ p. However, since qQq is discrete and finite and hence a countable direct sum of
finite dimensional factors, Ball(Eλ) is ⋅ϕω
q -compact and hence we have obtained a contradiction.
By the same reasoning as before, Properties (P1) and (P2) together with Proposition 2.2 imply that
This finishes the proof of Theorem 2.3.
Following [Con74], we define the asymptotic centraliser Mω of the von Neumann algebra M by
Mω =(xn)n ∈ ℓ∞(N, M) ∀ψ ∈ M∗, lim
n→ωxnψ − ψxn = 0~Iω(M) .
By [Con74, Proposition 2.8], we have Mω =(M ′ ∩ M ω)ϕω for every faithful normal state ϕ ∈ M∗.
11
Corollary 2.6. Let M be any factor with separable predual such that M ′ ∩ M ω ≠ C1. Then Mω is
diffuse.
Proof. Let ϕ ∈ M∗ be any faithful normal state. We have Mω =(M ′ ∩ M ω)ϕω = M ′ ∩(M ω)ϕω. Since
Mω ≠ C1 and since M ′ ∩ M = C1, the projection z obtained in Theorem 2.3 satisfies z = 0 and hence
Mω is diffuse.
2.4 The Akemann-Ostrand property (AO)
from the work of
The Akemann-Ostrand property for von Neumann algebras arises
Akemann-Ostrand [AO75] and was introduced by Ozawa in [Oza04]. A von Neumann algebra
M ⊂ B(H) has property (AO) if there are unital σ-weakly dense C∗-subalgebras B ⊂ M and C ⊂ M ′
such that B is locally reflexive and such that the map
B ⊗alg C → B(H)~K(H) ∶ x ⊗ y ↦ xy
is continuous with respect to the minimal tensor C∗-norm. We recall the following well-known
result. For a proof, we refer the reader to [Hou07, Chapter 4].
Proposition 2.7. Any free Araki-Woods factor satisfies property (AO).
2.5 Intertwining-by-bimodules techniques
Popa introduced his powerful intertwining-by-bimodule techniques in [Pop02, Pop03, Pop04]. We
first recall the intertwining-by-bimodule criterion in the case of finite von Neumann algebras. Let
(M, τ) be any tracial von Neumann algebra together with von Neumann subalgebras A ⊂ 1AM1A
and B ⊂ 1BM1B. Denote by EB ∶ 1BM1B → B the unique trace preserving faithful normal
conditional expectation. Then the following statements are equivalent (see [Pop03, Lemma 2.1 and
Corollary 2.3]):
• There is n ≥ 1, a nonzero partial isometry v ∈ M1,n(1AM1B) and a possibly non-unital normal
• There is no net of unitaries (wi)i in U(A) such that EB(x∗wiy) → 0 ∗-strongly as i → ∞ for
∗-homomorphism π ∶ A → Mn(B) such that av = vπ(a) for all a ∈ A.
all x, y ∈ 1AM1B.
We will say that A embeds into B inside M and write A ⪯M B if one of the above equivalent
conditions is satisfied.
Let (M, Tr) be any semifinite von Neumann algebra endowed with a semifinite faithful normal
trace. Let B ⊂ M be any von Neumann subalgebra such that Tr B is semifinite. Denote by
EB ∶ M → B the unique trace preserving faithful normal conditional expectation. Let p ∈ M be
any nonzero finite trace projection and A ⊂ pMp any von Neumann subalgebra. Let q ∈ B be any
nonzero finite trace projection. Observe that p ∨ q is a nonzero finite trace projection in M. We
will say that A embeds into qBq inside M and write A ⪯M qBq if A ⪯
(p∨q)M(p∨q) qBq in the usual
sense for finite von Neumann algebras.
12
We will need the following useful intertwining-by-bimodule criterion for semifinite von Neumann
algebras (see [HR10, Lemma 2.2] or [BHR14, Lemma 2.3]).
Lemma 2.8. Let (M, Tr) be any semifinite von Neumann algebra endowed with a semifinite faithful
normal trace. Let B ⊂ M be any von Neumann subalgebra such that Tr B is semifinite. Denote by
EB ∶ M → B the unique trace preserving faithful normal conditional expectation.
Let p ∈ M be any nonzero finite trace projection and A ⊂ pMp any von Neumann subalgebra. The
following conditions are equivalent:
(i) For every nonzero finite trace projection q ∈ B, we have A M qBq.
(ii) There exists an increasing sequence of nonzero finite trace projections qn ∈ B such that qn → 1
strongly and A M qnBqn for all n ∈ N.
(iii) There exists a net of unitaries (wi) in U(A) such that limkEB(x∗wiy)2,Tr = 0 for all x, y ∈
pM.
2.6 Deformation/Rigidity theory for free Araki-Woods factors
We introduce the s-malleable deformation of free Araki-Woods factors. It is an analogue of the
malleable deformations considered in [Pop01, IPP05].
Let U ∶ R → O(HR) be any orthogonal representation on a separable real Hilbert space. Denote by
(M, ϕ) = (Γ(HR , Ut)′′, ϕU) the associated free Araki-Woods factor together with its free quasi-free
state ϕ. Put M = cϕ(M) and simply denote by Tr the canonical semifinite faithful normal trace on
M. Furthermore, we write (M ,ϕ) = (Γ(HR ⊕ HR , Ut ⊕ Ut)′′, ϕU ⊕U) and M = cϕ(M). By [Shl97],
there are ∗-isomorphisms (M ,ϕ) ≅ (M, ϕ) ∗(M, ϕ) and M ≅ M ∗Lϕ(R) M. We will identify M
with its first copy in M and M with its first copy in M.
The orthogonal representation V ∶ R → O(HR ⊕ HR) given by
2 s) − sin( π
2 s)
Vs =cos( π
2 s)
sin( π
cos( π
2 s)
commutes with the orthogonal representation U ⊕ U ∶ R → O(HR ⊕ HR). Hence the associated
transformation Γ(Vs) on the free Fock space of HR ⊕ HR induces a ∗-automorphism αs of M . It
satisfies α1(x ∗ 1) = 1 ∗ x for all x ∈ M.
Since Γ(Vs) fixes the vacuum vector, it preserves the free quasi-free state ϕ on M. Hence it induces
a trace preserving ∗-automorphism of M that we still denote by αs. Likewise, the orthogonal
transformation
induces a trace preserving ∗-automorphism β of M which moreover satisfies β2 = id M, β M = idM
and βαs = α−sβ for all s ∈ R. Therefore, the deformation (αs, β)s∈R is s-malleable in the sense of
Popa [Pop03] and satisfies the following transversality property.
Proposition 2.9 (See [HR10, Proposition 4.2] and [Pop06b, Lemma 2.1]).
x − α2s(x)2 ≤√2αs(x) −(EM ○ αs)(x)2
for all x ∈ L2(M, Tr) and all s ∈ R.
0
1
0 −1
13
3 Proofs of Theorems A and B
3.1 Preliminaries on ω-solidity
Definition 3.1. Let ω ∈ β(N) ∖ N be a non-principal ultrafilter. We will say that a diffuse von
Neumann algebra M is ω-solid if for every von Neumann subalgebra Q ⊂ M with expectation whose
relative commutant Q′ ∩ M ω is diffuse, we have that Q is amenable.
We first show that ω-solidity is stable under amplifications.
is amenable and so is Q.
Proof. If M is a diffuse ω-solid von Neumann algebra, then pM p is ω-solid for every nonzero
Proposition 3.2. Let M be any diffuse ω-solid von Neumann algebra. Then p(M ⊗ B(ℓ2))p is
ω-solid for every nonzero projection p ∈ M ⊗ B(ℓ2).
projection p ∈ M. Indeed, let Q ⊂ pM p be any von Neumann subalgebra such that Q′ ∩(pM p)ω is
diffuse. Put Q = Q ⊕ C(1 − p). Then Q′ ∩ M ω ⊃ Q′ ∩(pM p)ω ⊕(1 − p)M ω(1 − p) is diffuse. Thus Q
It remains to prove that if M is a diffuse ω-solid von Neumann algebra, then M ⊗ B(ℓ2) is ω-solid.
We may assume that M is not amenable. Observe that if M is properly infinite, then M ⊗B(ℓ2) ≅ M.
Since any von Neumann algebra is the direct sum of a finite von Neumann algebra and a properly
infinite von Neumann algebra, after cutting down by a central projection, we may assume that M
is a diffuse ω-solid finite von Neumann algebra. Since M is the direct sum of a diffuse amenable von
Neumann algebra and at most countably many non-amenable ω-solid II1 factors, after cutting down
by a central projection, we may further assume without loss of generality that M is a non-amenable
ω-solid II1 factor.
We first prove that M t is ω-solid for all t > 0. Using the first paragraph of the proof, it suffices to
corner of M, we obtain that M is not ω-solid. This is a contradiction.
prove that M ⊗ Mn(C) is ω-solid for all n ≥ 1. Let Q ⊂ M ⊗ Mn(C) be any von Neumann subalgebra
such that Q′ ∩(M ⊗ Mn(C))ω is diffuse. Assume by contradiction that Q is not amenable. We
may choose a projection q ∈ Q such that qQq is not amenable and (τ ⊗ Trn)(q) ≤ 1~n. Note that
(qQq)′ ∩(q(M ⊗ Mn(C))q)ω = (Q′ ∩(M ⊗ Mn(C))ω)q is diffuse. Regarding q(M ⊗ Mn(C))q as a
We now prove that M = M ⊗ B(ℓ2) is ω-solid. Let Q ⊂ M be any von Neumann subalgebra with
faithful normal conditional expectation EQ ∶ M → Q and such that Q′ ∩ Mω is diffuse. Denote by
Eω ∶ Mω → M the canonical faithful normal conditional expectation. Let ϕ ∈ M∗ be any faithful
normal state such that ϕ○EQ = ϕ. By Theorem 2.3, the relative commutant Q′ ∩(Mω)ϕω is diffuse
as well.
Fix a tracial faithful normal semifinite weight Tr on M. By [AH12, Lemma 4.26], Trω = Tr ○Eω is
a tracial faithful normal semifinite weight on Mω. Denote by T ∈ L1(M, Tr)+ the unique positive
selfadjoint operator affiliated with M satisfying ϕ = Tr(T ⋅). By [Con73, Lemme 1.2.3 (b) and
Lemme 1.4.4], we have ϕω = Trω(T ⋅).
Q = Q ∨ B. Since ϕω = Trω(T ⋅), we have Mϕω = B ′ ∩ Mω and hence
Q′ ∩ Mω = Q′ ∩ B ′ ∩ Mω = Q′ ∩(Mω)ϕω
Denote by B ⊂ M the von Neumann subalgebra generated by all the spectral projections of T . Put
14
is diffuse. Observe that Q is globally invariant under the modular automorphism group (σϕ
t ).
Since T ∈ L1(M, Tr), we may choose a sequence of finite trace projections pk ∈ B ⊂ Q such that
(pkQpk)′ ∩(pkMpk)ω =(Q′ ∩Mω)pk =(Q′ ∩(Mω)ϕω)pk is diffuse, we have that pkQpk is amenable.
pk → 1 strongly. Since pkMpk is an ω-solid II1 factor by the first part of the proof and since
Since amenability is stable under direct limits, we finally obtain that Q is amenable. Since Q ⊂ Q
is a von Neumann subalgebra with expectation, Q is also amenable. This finishes the proof of
Proposition 3.2.
Next, we prove a useful characterisation of ω-solidity.
Proposition 3.3. Let M be any von Neumann algebra with separable predual that has no amenable
direct summand. The following conditions are equivalent.
(i) For every von Neumann subalgebra Q ⊂ M with expectation, if Q′ ∩ M ω is diffuse then Q is
amenable.
(ii) For every von Neumann subalgebra Q ⊂ M with expectation that has no amenable direct
summand, the relative commutant Q′ ∩ M ω is discrete.
Proof. (i) ⇒ (ii). Let Q ⊂ M be any von Neumann subalgebra with expectation that has no
amenable direct summand. By Theorem 2.3, there is a unique central projection z ∈ Z(Q′ ∩ M ω) ∩
Z(Q′ ∩ M) such that (Q′ ∩ M ω)z is diffuse and (Q′ ∩ M ω)(1 − z) is discrete. Put Q = Qz ⊕ C(1 − z).
Since Q′ ∩ M ω ⊃(Q′ ∩ M ω)z ⊕(1 − z)M ω(1 − z) is diffuse, we have that Q is amenable. Thus z = 0
and Q′ ∩ M ω is discrete.
(ii) ⇒ (i). Let Q ⊂ M be any von Neumann subalgebra with expectation such that Q′ ∩ M ω is
diffuse. Denote by z ∈ Z(Q) the unique central projection such that Qz has no amenable direct
summand and Q(1 − z) is amenable. Since (1 − z)M(1 − z) has no amenable direct summand,
Q = Qz ⊕(1 − z)M(1 − z) has no amenable direct summand either. Then Q′ ∩ M ω is discrete and
so is (Q′ ∩ M ω)z =(Q′ ∩ M ω)z. Thus z = 0 and Q is amenable.
3.2 Proof of Theorem A
Proof of Theorem A. Let M be any von Neumann algebra with separable predual that satisfies
property (AO). Denote by Eω ∶ M ω → M the canonical faithful normal conditional expectation. Let
Q ⊂ M be any von Neumann subalgebra with faithful normal conditional expectation EQ ∶ M → Q
and such that Q′ ∩ M ω is diffuse. Fix a faithful normal state ϕ ∈ M∗ satisfying ϕ ○ EQ = ϕ. By
Theorem 2.3, Q′∩(M ω)ϕω is diffuse and hence there is a sequence of unitaries Uk in U(Q′∩(M ω)ϕω)
such that Uk → 0 weakly as k → ∞. Choose a sequence (uk
m ∈ Ball(M)
ϕ -dense sequence in Ball(Q) and (ψj)j≥1 be a
for all m ∈ N and Uk = (uk
⋅-dense sequence in M∗.
There exists an increasing sequence(kn)n in N such that for every n ∈ N, we have limm→ω ψj(ukn
m) =
(ψj ○ Eω)(Ukn) < 1
n for all 1 ≤ j ≤ n. Therefore, there exists an increasing sequence (mn)n in N
such that for every n ∈ N, the element un = ukn
m)m ∈ Mω(M) such that uk
m)ω. Let (xi)i≥1 be a ⋅#
mn ∈ Ball(M) satisfies
15
n for all 1 ≤ j ≤ n.
(P1) unϕ − ϕun ≤ 1
n,
nun#
n#
n and 1 − unu∗
(P2) 1 − u∗
ϕ ≤ 1
ϕ ≤ 1
n
(P3) unxi − xiun#
ϕ ≤ 1
n for all 1 ≤ i ≤ n and
(P4) ψj(un) ≤ 1
Property (P1) and Proposition 2.2 show that (un)n ∈ Mω(M) and together with Properties (P2)
and (P3) they show that U = (un)ω ∈ U(Q′ ∩(M ω)ϕω). Finally, Property (P4) shows that un → 0
Regard M ⊂ B(H) where the Hilbert space H is given by property (AO). Define the unital
completely positive map Θ ∶ B(H) → B(H) by Θ(T) = σ-weak limn→ω unT u∗
n. Observe that
Θ(x) = Eω(U xU ∗) ∈ M for all x ∈ M and hence Θ M is normal. Next, define the unital com-
weakly as n → ∞.
pletely positive maps
Ψk = 1
k
k
Q
j=1
Θ○j ∶ B(H) → B(H)
and let Ψ be the pointwise σ-weak limit of (Ψk)k, as k → ω. Since Θ(M) ⊂ M, we have Ψk(M) ⊂ M
for all k ≥ 1 and hence Ψ(M) ⊂ M. Note that since Θ M is normal, we have Ψ M = Θ ○ Ψ M . Since
U ∈ U((M ω)ϕω), we also have ϕ ○ Θ M = ϕ. This implies that ϕ ○ Ψk M = ϕ for all k ≥ 1 and hence
ϕ ○ Ψ M = ϕ.
Put Q ={U, U ∗}′ ∩ M. Observe that Q ⊂ Q ⊂ M and that Q is globally invariant under the modular
t ). Let x ∈ M and put y = Ψ(x) ∈ M. We have y = Ψ(x) = Θ(Ψ(x)) =
automorphism group (σϕ
Θ(y) = Eω(U yU ∗). Since U ∈ U((M ω)ϕω) and since yϕω =yϕ, we obtain U yU ∗ϕω =yϕ and
hence
y − U yU ∗2
ϕω − 2 Re ϕω(U y∗U ∗y)
ϕω =y2
= 2y2
= 2y2
= 2y2
ϕω +U yU ∗2
ϕ − 2 Re ϕ(Eω(U y∗U ∗y))
ϕ − 2 Re ϕ(Eω(U y∗U ∗)y)
ϕ − 2 Re ϕ(y∗y) = 0 .
Therefore, we have y = U yU ∗ and hence Ψ(x) = y ∈ Q. Combining this with the fact that Θ(x) = x
for all x ∈ Q, we see that Ψ M is a norm one projection onto Q. We already saw that Ψ M is
ϕ-preserving and hence we infer that Ψ M = EQ ∶ M → Q is the unique ϕ-preserving conditional
j=1 EQ(bi) ci. By definition of Ψ, we have Ψ(c) = c
Define ΦQ ∶ M ⊗alg M ′ → B(H) ∶ ∑n
for all c ∈ M ′. Therefore [Cho74, Theorem 3.1] implies that for all n ≥ 1, all bi ∈ M and all ci ∈ M ′,
we have
expectation from M onto Q.
j=1 bi ⊗ ci ↦ ∑n
Ψ( n
Q
j=1
bi ci) =
n
Q
j=1
Ψ(bi) ci =
n
Q
j=1
EQ(bi) ci = ΦQ( n
bi ⊗ ci) .
Q
j=1
The fact that un → 0 σ-weakly as n → ω implies that Θ(T) = σ-weak limn→ω unT u∗
n = 0 for all
T ∈ K(H). This shows that Ψ(T) = 0 for all T ∈ K(H). Hence K(H) ⊂ ker Ψ. Denote by
16
π ∶ B(H) → B(H)~K(H) the canonical quotient map. Then there exists a unital completely positive
map Ψ ∶ B(H)~K(H) → B(H) such that Ψ = Ψ ○ π.
By property (AO) of M, there is a unital σ-weakly dense locally reflexive C∗-subalgebra B ⊂ M
and a unital σ-weakly dense C∗-subalgebra C ⊂ M ′ together with a ∗-homomorphism
ν ∶ B ⊗alg C → B(H)~K(H) ∶
n
Q
j=1
that is continuous with respect to the minimal tensor norm on B ⊗alg C. Therefore ΦQ = Ψ ○ ν is
continuous with respect to the minimal tensor norm on B ⊗alg C. Applying [Oza04, Lemma 5], we
obtain that Q is amenable and so is Q.
bi ⊗ ci ↦ π( n
bi ci)
Q
j=1
3.3 Proof of Theorem B
The next theorem is a generalisation of [Hou08, Theorem 3.4] regarding the position of the rel-
ative commutant of large subalgebras of the continuous core of free Araki-Woods factors in the
ultraproduct framework.
state ϕ.
Theorem 3.4. Let U ∶ R → O(HR) be any orthogonal representation on a separable real Hilbert
space. Denote by (M, ϕ) =(Γ(HR , Ut)′′, ϕU) the corresponding free Araki-Woods factor together with
its free quasi-free state and by M = cϕ(M) the continuous core associated with the free quasi-free
Then for every nonzero finite trace projection p ∈ Lϕ(R) ⊂ M and every von Neumann subalgebra
q ∈ Lϕ(R) such that
Q ⊂ pMp that has no amenable direct summand, there exists a nonzero finite trace projection
Q′ ∩ pMωp ⪯Mω Lϕ(R)ωq .
Proof of Theorem 3.4. The proof is very much inspired by [Pet09, Theorems 4.3 and 4.5] (see
also [Hou12b, Theorem D]). Let αt ∶ M → M be the trace preserving s-malleable deformation
introduced in Subsection 2.6. Write M = M ∗Lϕ(R) α1(M). Observe that if (xn)n ∈ Iω(M), then
also (αt(xn))n ∈ Iω(M) for all t ∈ R. It follows that (αt) extends to a one-parameter family of
trace preserving ∗-automorphisms of the ultraproduct von Neumann algebra Mω that we denote
t ). We emphasise however that t ↦ αt(x) need not be continuous when x ∈ Mω.
by (αω
Step 1: Uniform convergence in ⋅2 of (αω
t ) on Ball(Q′ ∩ pMωp). Assume by contradiction
that (αω
t ) does not converge uniformly in ⋅2 on Ball(Q′ ∩ pMωp). Thus there exist c > 0, a
sequence (tk)k of positive reals such that limk tk = 0 and a sequence (Xk)k in Ball(Q′ ∩ pMωp)
such that Xk − αω
2tk(Xk)2 ≥ 2c for all k ∈ N. Write Xk =(xk,n)ω with xk,n ∈ Ball(pMp) satisfying
limn→ωyxk,n − xk,ny2 = 0 and Xk − αω
2tk(Xk)2 = limn→ωxk,n − α2tk(xk,n)2 for all k ∈ N and all
Denote by I the directed set of all pairs (F , ε) with ε > 0 and F ⊂ Ball(Q) finite subset. Let
i = (F , ε) ∈ I. Choose k ∈ N large enough so that a − αtk(a)2 ≤ ε~3 for all a ∈ F. Then choose
n ∈ N large enough so that xk,n − α2tk(xk,n)2 ≥ c and axk,n − xk,na2 ≤ ε~3 for all a ∈ F.
y ∈ Q.
17
Put ξi = αtk(xk,n) − EpMp(αtk(xk,n)) ∈ L2(pMp) ⊖ L2(pMp). By Proposition 2.9, we have
ξi2 ≥ 1√2xk,n − α2tk(xk,n)2 ≥ c√2
.
For all x ∈ pMp, we have
xξi2 =(1 − EpMp)(xαtk(xk,n))2 ≤xαtk(xk,n)2 ≤x2 .
By Popa's spectral gap argument [Pop06b], for all a ∈ F, we have
aξi − ξia2 =(1 − EpMp)(aαtk(xk,n) − αtk(xk,n)a)2 ≤aαtk(xk,n) − αtk(xk,n)a2
Lemma 5.1]), it follows that Q has an amenable direct summand by Connes' characterisation of
≤ 2a − αtk(a)2 +axk,n − xk,na2 ≤ ε.
Hence ξi ∈ L2(pMp) ⊖ L2(pMp) is a net of vectors satisfying lim supixξi2 ≤x2 for all x ∈ pMp,
and limiaξi − ξia2 = 0 for all a ∈ Q. Since the pMp-pMp-bimodule L2(pMp)⊖
lim inf iξi2 ≥ c√2
L2(pMp) is weakly contained in the coarse pMp-pMp-bimodule L2(pMp) ⊗ L2(pMp) (see [HR10,
amenability [Con76]. This is a contradiction and hence (αω
t ) does converge uniformly in ⋅2 on
Ball(Q′ ∩ pMωp).
We now proceed by contradiction and assume that Q′∩pMωp Mω Lϕ(R)ωq for every nonzero finite
trace projection q ∈ Lϕ(R). By Lemma 2.8, there exists a net (Uk)k of unitaries in U(Q′ ∩ pMωp)
such that limkELϕ(R)ω(X ∗UkY)2 = 0 for all X, Y ∈ pMω.
Step 2: Uniform convergence in ⋅2 of (αt) on Ball(Q). Take ε > 0. Since (αω
t ) converges
uniformly in ⋅2 on Ball(Q′ ∩ pMωp), there is some t0 > 0 such that for all t ∈ [0, t0], we have
t (X)−X2 < ε2~4 for all X ∈ Ball(Q′∩pMωp). We show that for all t ∈ [0, t0] and all x ∈ Ball(Q),
αω
we have αt(x) − x2 < ε.
Take t ∈ [0, t0] and x ∈ Ball(Q). Let (yi)i be a ⋅ 2-dense sequence in Ball(pM). There is an
increasing sequence (kn)n such that for every n ≥ 1, the unitary Ukn ∈ U(Q′ ∩ pMωp) satisfies
m ∈ Ball(pMp) for
ELϕ(R)ω(y∗
all m ∈ N. There exists an increasing sequence (mn)n in N such that for every n ≥ 1, the element
vn = ukn
i Ukn yj)2 < 1~n for all i, j ∈ {1, . . . , n}. Write Ukn = (ukn
m )ω with ukn
mn ∈ Ball(pMp) satisfies
n − x2 ≤ 1~n,
• vnxv∗
• ELϕ(R)(y∗
• αt(vn) − vn2 ≤ ε2~4.
i vnyj)2 ≤ 1~n for all i, j ∈ {1, . . . , n}, and
Since (yi)i is ⋅ 2-dense in Ball(pM), the second condition implies that ELϕ(R)(a∗vnb)2 → 0 for
all a, b ∈ pM. Writing now δt(x) = αt(x) − EpMp(αt(x)) ∈ p Mp ⊖ pMp, we obtain
δt(x)2
2 = ⟨δt(x), δt(x)⟩
≤ ⟨δt(vnxv∗
≤ ⟨vnδt(x)v∗
≤ ⟨vnδt(x)v∗
n), δt(x)⟩ + vnxv∗
n, δt(x)⟩ + vnxv∗
n, δt(x)⟩ + 1~n + ε2~2 .
n − x2
n − x2 + 2vn − αt(vn)2
18
Observe moreover that by Cauchy-Schwarz inequality, we have
n, δt(x)⟩ = Tr(δt(x)∗vnδt(x)v∗
n)
⟨vnδt(x)v∗
= Tr(EpMp(δt(x)∗vnδt(x))v∗
n)
≤ EpMp(δt(x)∗vnδt(x))2.
Since δt(x) ∈ p( M⊖M) and since limn ELϕ(R)(a∗vnb)2 = 0 for all a, b ∈ pM, by [BHR14, Theorem
2.5, Claim], it follows that limn EpMp(δt(x)∗vnδt(x))2 = 0 and hence limn ⟨vnδt(x)v∗
n, δt(x)⟩ = 0.
Hence, the transversality property of Proposition 2.9 now yields x − α2t(x)2 ≤ √2δt(x)2 ≤ ε.
Thus, (αt) converges uniformly in ⋅ 2 on Ball(Q).
Step 3: Deducing a contradiction. Since (αt) converges uniformly in ⋅ 2 on Ball(Q),
[HR10, Theorem 4.3] implies that there exists a nonzero finite trace projection r ∈ Lϕ(R) such that
Q ⪯M Lϕ(R)r. Since Lϕ(R)r is amenable, it follows that Q has an amenable direct summand, con-
tradicting our assumption that it does not. It follows that the assumption Q′ ∩pMωp Mω Lϕ(R)ωq
for every nonzero finite trace projection q ∈ Lϕ(R) of the previous step is wrong. This finishes the
proof of the theorem.
Before we can proceed to the proof of Theorem B, we need a few basic results regarding mixing
inclusions in semifinite amalgamated free products. Recall that an inclusion of tracial von Neumann
algebras B ⊂ (M, τ) is mixing if for every uniformly bounded net (wk)k of elements in B that goes
to 0 weakly, we have
k EB(xwky)2 = 0, ∀x, y ∈ M ⊖ B .
lim
Let now (M, Tr) be any semifinite von Neumann algebra endowed with a semifinite faithful normal
trace. Let B ⊂ M be any von Neumann subalgebra such that Tr B is semifinite. Denote by
EB ∶ M → B the unique trace preserving faithful normal conditional expectation.
Definition 3.5. Keep the same notation. We will say that the inclusion B ⊂ M is mixing if for
every nonzero finite trace projection q ∈ B and for every uniformly bounded net (wk)k in qBq that
goes to 0 weakly, we have
k EB(x∗wky)2 = 0, ∀x, y ∈ q(M ⊖ B) .
lim
We prove a useful characterisation of mixing inclusions of semifinite von Neumann algebras.
Lemma 3.6. Keep the same notation. The following conditions are equivalent.
(i) The inclusion B ⊂ M is mixing.
(ii) For every nonzero finite trace projection q ∈ B, the inclusion of tracial von Neumann algebras
qBq ⊂ qMq is mixing.
(iii) There exists an increasing sequence of nonzero finite trace projections qn ∈ B such that the
inclusion of tracial von Neumann algebras qnBqn ⊂ qnMqn is mixing for all n ∈ N.
19
Proof. (i) ⇒ (ii) ⇒ (iii) are obvious. For (iii) ⇒ (i), let q ∈ B be a nonzero finite trace projection,
(wk)k a net of elements in Ball(qBq) that goes to 0 weakly and x, y ∈ Ball(M) ∩ q(M ⊖ B).
Take ε > 0. Since Tr(q) < ∞ and qn → 1 strongly, there exists n ∈ N such that
q − qnq2 + q − qqn2 + x∗q − qnx∗q2 + qy − qyqn2 ≤ ε
(3.1)
.
4
This implies in particular that for all k, we have
(3.2)
wk − qnwkqn2 ≤ wk − qnwk2 + qn(wk − wkqn)2 ≤ ε
4
.
Since the inclusion of tracial von Neumann algebras qnBqn ⊂ qnMqn is mixing,
since
qnqyqn, qnx∗qqn ∈ qnMqn ⊖ qnBqn and since qnwkqn → 0 weakly as k → ∞, there exists k0 such
that for all k ≥ k0, we have
(3.3)
EB(qnx∗qqn qnwkqn qnqyqn)2 = Tr(qn)1~2 EqnBqn(qnx∗qqn qnwkqn qnqyqn)2,τqnMqn
≤ ε
2
.
Combining (3.1), (3.2) and (3.3), we obtain
EB(x∗wky)2 ≤ EB(qnx∗q wk qyqn)2 +
ε
4
≤ EB(qnx∗q qnwkqn qyqn)2 +
≤ ε.
ε
2
An interesting class of mixing inclusions of semifinite von Neumann algebras arises from modular
automorphism groups.
Proposition 3.7. Let (M, ϕ) be any von Neumann algebra together with a faithful normal state
t ) is mixing, that is, for all x, y ∈ M, we have
such that the modular automorphism group (σϕ
t (x)y) = ϕ(x)ϕ(y). Denote by cϕ(M) the continuous core associated with ϕ. Then the
lim t →∞ ϕ(σϕ
inclusion Lϕ(R) ⊂ cϕ(M) is mixing.
Proof. By Fourier transform, identify Lϕ(R) with L∞(R). The proof of [Hou08, Theorem 3.7]
shows that the inclusion of tracial von Neumann algebras Lϕ(R)q ⊂ qcϕ(M)q is mixing for all
nonzero projections q corresponding to the bounded intervals of the form [−T, T] with T > 0. Then
Lemma 3.6 shows that the inclusion Lϕ(R) ⊂ cϕ(M) is mixing.
let B ⊂ Mi be an inclusion of von Neumann algebras with faithful normal
For all i ∈ {1, 2},
conditional expectation Ei ∶ Mi → B. Assume that B is semifinite with faithful normal semifinite
tracial weight Tr. Assume moreover that Tr ○Ei is still a semifinite trace on Mi. Consider the
amalgamated free product von Neumann algebra (M, E) = (M1, E1) ∗B (M2, E2) and observe that
Tr ○E is still a faithful normal semifinite trace on M (see [BHR14, Section 2.2]). We say in that
case that M = M1 ∗B M2 is a semifinite amalgamated free product von Neumann algebra.
We prove the analogue of [Hou12a, Proposition 4.7] in the setting of semifinite amalgamated free
product von Neumann algebras.
20
Proposition 3.8. Let M = M1 ∗B M2 be a semifinite amalgamated free product von Neumann
algebra. Assume that the inclusion B ⊂ M2 is mixing. Then the inclusion M1 ⊂ M is mixing.
Proof. Denote by EM1 ∶ M → M1 the unique trace preserving faithful normal conditional ex-
pectation. To prove that the inclusion M1 ⊂ M is mixing, using Kaplansky's density theo-
rem and Lemma 3.6, it suffices to show that for all nonzero finite trace projections q ∈ B, all
nets (wk)k of elements in Ball(qM1q) that go to 0 weakly and all elements x, y ∈ q(M ⊖ M1)
of the form x = qx1⋯x2m+1 and y = qy1⋯y2n+1 with m, n ≥ 1, x1, x2m+1, y1, y2n+1 ∈ Ball(M1),
x2, . . . , x2m, y2, . . . , y2n ∈ Ball(M2)∩(M2⊖B) and x3, . . . , x2m−1, y3, . . . , y2n−1 ∈ Ball(M1)∩(M1⊖B),
we have
k EM1(x∗wky)2 = 0 .
lim
Using the property of freeness with amalgamation over B, we have
EM1(x∗wky) = EM1(x∗
= EM1(x∗
= EM1(x∗
2m+1⋯x∗
2m+1⋯x∗
2m+1⋯x∗
2 x∗
2 EB(x∗
3 EB(x∗
1qwkqy1 y2⋯y2n+1)
1 wky1) y2⋯y2n+1)
2 EB(x∗
1 wky1) y2) y3⋯y2n+1) .
Take ε > 0. Since Tr(q) < +∞, we may choose a large enough finite trace projection p ∈ B such that
qy1 − qy1p2 + x∗
1 q − px∗
1q2 ≤ ε .
We infer that EB(x∗
1wky1) − EB(px∗
EM1(x∗wky) − EM1(x∗
1q wk qy1p)2 ≤ ε for all k and hence
2m+1⋯x∗
k
lim sup
3 EB(x∗
Since the inclusion B ⊂ M2 is mixing, since (EB(px∗
weakly and since px2, py2 ∈ p(M2 ⊖ B), it follows that limk EB(x∗
and hence
2 EB(px∗
1 q wk qy1p) y2) y3⋯y2n+1)2 ≤ ε .
1 q wk qy1p))k is a net in Ball(pBp) that goes to 0
1q wk qy1p) py2)2 = 0
2 p EB(px∗
2m+1⋯x∗
3 EB(x∗
lim
k EM1(x∗
1 q wk qy1p) y2) y3⋯y2n+1)2 = 0 .
Since ε > 0 is arbitrary, we deduce that
2 EB(px∗
This implies that lim supk EM1(x∗wky)2 ≤ ε.
limk EM1(x∗wky)2 = 0.
Proof of Theorem B. By Proposition 3.2, it suffices to prove that finite corners of continuous cores
of free Araki-Woods factors are ω-solid.
Let U ∶ R → O(HR) be any orthogonal representation on a separable real Hilbert space that is the
direct sum of a mixing representation and a representation of dimension less than or equal to 1.
Denote by (M, ϕ) = (Γ(HR , Ut)′′, ϕU) the associated free Araki-Woods factor together with its free
quasi-free state and M = cϕ(M) its continuous core with respect to the free quasi-free state ϕ.
Observe that M is a type III1 factor and hence M is a type II∞ factor. Let p ∈ Lϕ(R) be any
nonzero finite trace projection and Q ⊂ pMp any von Neumann subalgebra such that Q′ ∩(pMp)ω
is diffuse.
21
Assume by contradiction that Q is not amenable. Let z ∈ Z(Q) be a nonzero central projection
such that Qz has no amenable direct summand. Since pMp is a II1 factor and since Lϕ(R)p is
diffuse, there exists u ∈ U(pMp) and q ∈ Lϕ(R)p such that uzu∗ = q. So up to conjugating by a
unitary and taking a smaller projection in Lϕ(R)p, we may assume without loss of generality that
Q ⊂ pMp has no amenable direct summand and that Q′ ∩ (pMp)ω is diffuse.
By Theorem 3.4, we know that there exists a nonzero finite trace projection q ∈ Lϕ(R) such that
Q′ ∩ (pMp)ω ⪯Mω (Lϕ(R)q)ω. Up to replacing q by p ∨ q ∈ Lϕ(R), we may assume that p ≤ q.
If (Ut) is mixing, then [Hou08, Proposition 2.4] and Proposition 3.7 show that the inclusion Lϕ(R) ⊂
M is mixing. Applying [Ioa12, Lemma 9.5], we obtain that Q ⪯qMq Lϕ(R)q.
If (Ut) is the direct sum of a mixing orthogonal representation with an orthogonal representation
of dimension one, then M = N ∗ L(Z), where N is the Araki-Woods factor associated with the
mixing part of (Ut). Writing N = cϕ N (N), we obtain M ≅ N ∗Lϕ(R) (L(Z) ⊗ Lϕ(R)). Hence
Proposition 3.8 shows that the inclusion L(Z) ⊗ Lϕ(R) ⊂ M is mixing. Moreover, we know that
Q′ ∩ (pMp)ω ⪯
(qMq)ω (Lϕ(R)q)ω and hence Q′ ∩ (pMp)ω ⪯
(qMq)ω ((L(Z) ⊗ Lϕ(R))q)ω. Applying
[Ioa12, Lemma 9.5], we obtain that Q ⪯qMq (L(Z) ⊗ Lϕ(R))q.
However, in both cases, this contradicts the fact that Q has no amenable direct summand.
3.4 Computation of Connes's τ -invariant for ω-solid factors
Let M be any von Neumann algebra with separable predual. We endow Aut(M) with the topology
of pointwise convergence in M∗, that is,
αn → id in Aut(M) as n → ∞ if and only if
n→∞ϕ ○ αn − ϕ = 0 for all ϕ ∈ M∗ .
lim
Endowed with this topology, Aut(M) becomes a Polish group.
Recall from [Con74] that when M is a factor, we have that M is full if and only if the subgroup
Inn(M) of inner automorphisms is closed in Aut(M). Equivalently, we have Mω = C1 for some (or
any) ω ∈ β(N) ∖ N. In that case, the quotient group Out(M) = Aut(M)~Inn(M) endowed with
the quotient topology is a Polish group. We will denote by π ∶ Aut(M) → Out(M) the quotient
homomorphism.
By Connes's Radon-Nikodym cocycle theorem [Con73, Théorème 1.2.1] (see also [Tak03, Theorem
VIII.3.3]), the homomorphism δ ∶ R → Out(M) ∶ t ↦ π(σϕ
t ) is well-defined and does not depend on
the choice of a particular state on M.
Definition 3.9 ([Con74]). Let M be a full factor of type III1 with separable predual. We define
τ(M) to be the weakest topology that makes the map δ ∶ R → Out(M) continuous.
It is typically difficult to calculate Connes's τ-invariant for arbitrary type III1 factors.
In the
case of the free Araki-Woods factors M = Γ(HR , Ut)′′, using a 14ε-type argument, it is proven in
[Shl97, Vae06] that τ(M) is the weakest topology that makes the map R → O(HR) ∶ t ↦ Ut strongly
continuous.
In the next proposition, we show that Connes's τ-invariant is computable for a fairly large class
of ω-solid type III1 factors. Our proof no longer relies on a 14ε-type argument and works in great
generality.
22
t continuous.
Proposition 3.10. Let M be any ω-solid factor of type III1 with separable predual and ϕ ∈ M∗ any
faithful normal state whose centralizer is a non-amenable II1 factor. Then M is a full factor and
τ(M) is the weakest topology on R that makes the map R → Aut(M) ∶ t ↦ σϕ
Proof. Let M be any ω-solid factor of type III1 with separable predual and ϕ ∈ M∗ a faithful normal
state whose centralizer is a non-amenable II1 factor. Since M is a non-amenable ω-solid factor,
M ′ ∩ M ω is discrete by Proposition 3.3 and hence M ′ ∩ M ω = C1 by Corollary 2.6. This implies
that M is a full factor. We next have to show that if (tn)n is a sequence in R that converges to 0
with respect to τ(M), then σϕ
By Theorem A and Proposition 3.3, the relative commutant (M ϕ)′ ∩ M ω is discrete. Applying
Theorem 2.3, we have that (M ϕ)′ ∩ M ω = (M ϕ)′ ∩ M. Since ((M ϕ)′ ∩ M)ϕ = (M ϕ)′ ∩ M ϕ = C1,
[AH12, Lemma 5.3] implies that (M ϕ)′ ∩ M = C1 or (M ϕ)′ ∩ M is a factor of type III1. Since
(M ϕ)′ ∩ M ω = (M ϕ)′ ∩ M is discrete, we obtain that (M ϕ)′ ∩ M ω = C1. Observe that this implies
that (M ϕ)′ ∩ M ω = C1 for all non-principal ultrafilter ω ∈ β(N) ∖ N.
Now take a sequence (tn)n in R that converges to 0 with respect to τ(M). Then there is a
tn → id in Aut(M). Fix ω ∈ β(N) ∖ N a
sequence of unitaries (un)n in M such that (Ad un) ○ σϕ
non-principal ultrafilter. As in the proof of [Ued11, Proposition 3.1], we have that (un)n ∈ Mω(M)
and (un)ω ∈ (M ϕ)′ ∩ M ω. Indeed, for all n ∈ N, we have
n = ϕ ○ (Ad un) − ϕ = ϕ ○ (Ad un) ○ σϕ
tn = ϕ ○ (Ad un) ○ σϕ
tn → id in Aut(M).
nϕ − ϕu∗
u∗
tn − ϕ ○ σϕ
tn − ϕ.
nϕ − ϕu∗
tn − ϕ = 0 and hence
tn − ϕ = 0, we have limn→ω ϕ ○ (Ad un) ○ σϕ
Since limn→∞ ϕ ○ (Ad un) ○ σϕ
n = 0. Therefore (un)n ∈ Mω(M) and (un)ω ∈ (M ω)ϕω by Proposition 2.2.
limn→ω u∗
tn(x) → x strongly as n → ∞ for all x ∈ M ϕ. This implies that
We moreover have (Ad un) ○ σϕ
n − xϕ = 0 for all x ∈ M ϕ. Since (un)n ∈ Mω(M) and (un)ω ∈ (M ω)ϕω, we finally
limn→ω unxu∗
obtain (un)ω ∈ (M ϕ)′ ∩ M ω.
Since (M ϕ)′ ∩ M ω = C1, we have limn→ω un − ϕ(un)1ϕ = (un)ω − ϕω((un)ω)ϕω = 0. Since this is
true for every ω ∈ β(N) ∖ N, we obtain limn→∞ un − ϕ(un)1ϕ = 0.
Proceeding now exactly as in the proof of [Con74, Theorem 5.2], we conclude that σϕ
tn
Aut(M).
→ id in
4 Proof of Theorem C
We first recall a basic fact on ε-orthogonality.
Definition 4.1. Let H be a complex Hilbert space and ε ≥ 0. Two (not necessarily closed) subspaces
K, L ⊂ H are called ε-orthogonal if ⟨ξ, η⟩ ≤ εξη for all ξ ∈ K and all η ∈ L. In that case, we will
denote K ⊥ε L.
Proposition 4.2 ([Hou12a, Proposition 2.3]). There is a continuous function δ ∶ [0, 1~2) → R≥0
satisfying δ(0) = 0 and the following property. If k ≥ 1 and 0 ≤ ε < 1~2 are such that δ○(k−1)(ε) < 1~2,
23
then for all projections pi ∈ B(H), i ∈ {1, . . . , 2k}, satisfying piH ⊥ε pjH for all i, j ∈ {1, . . . , 2k},
i ≠ j, we have
2k
Q
i=1piξ2 ≤
k−1
M
j=0(1 + δ○j(ε))2P ξ2 ,
where P = ⋁2k
i=1 pi is the projection onto the closed linear span span ⋃2k
i=1 piH.
The main result of this section is the following asymptotic orthogonality result in the framework of
ultraproducts of free Araki-Woods factors and is inspired by [Pop83, Lemma 2.1].
Theorem 4.3. Let U ∶ R → O(HR) be any weakly mixing orthogonal representation on a separable
real Hilbert space and (M, ϕ) = (Γ(HR , Ut)′′, ϕU) the associated free Araki-Woods factor. Then for
all x, y ∈ (M ω)ϕω ⊖ C1 and all a, b ∈ M ⊖ C1, we have ϕω(b∗y∗ax) = 0.
Proof. Let H = HR ⊕ iHR and denote by H = F(H) the full Fock space. We view KR + iKR ⊂ H as
a dense subspace of H. Put κt = id ⊕ ࣷn≥1 U ⊗n
∈ U(H). For every x ∈ M, we have
t
t (x)Ω = κt(xΩ) .
σϕ
Since the linear span of 1 and of all the reduced words W(ξ1 ⊗⋯⊗ ξm) with m ≥ 1 and ξj ∈ KR +iKR
is a unital σ-strongly dense ∗-subalgebra of M, it suffices to prove the result when a = W(ξ1⊗⋯⊗ξk)
and b = W(η1 ⊗ ⋯ ⊗ ηℓ) are reduced words with ξ1, . . . , ξk, η1, . . . , ηℓ ∈ KR + iKR. Approximating
ηj ∈ KR + iKR by 1[λ−1,λ](A)(ηj) ∈ KR + iKR for all 1 ≤ j ≤ ℓ and for λ > 1 sufficiently large, we may
further assume that ηj = 1[λ−1,λ](A)(ηj) for all 1 ≤ j ≤ ℓ. It follows that the map R → KR + iKR ∶
t ↦ Utηj can be extended to an entire analytic function which takes values in KR + iKR for all
1 ≤ j ≤ ℓ. This implies that the map R → M ∶ t ↦ W(Utηℓ ⊗ ⋯ ⊗ Utη1) can be extended to an
t (W(ηℓ ⊗ ⋯ ⊗ η1)) = W(Utηℓ ⊗ ⋯ ⊗ Utη1) for all t ∈ R,
M-valued entire analytic function. Since σϕ
we obtain that W(ηℓ ⊗ ⋯ ⊗ η1) is analytic for the modular automorphism group (σϕ
t ) and we have
σϕ
z (W(ηℓ ⊗ ⋯ ⊗ η1)) = W(Aizηℓ ⊗ ⋯ ⊗ Aizη1) for all z ∈ C.
From now on and for the rest of the proof, define L = span(ξk, ξk, η1, η1) ⊂ KR + iKR. We will use
the following notation:
• X1 ⊂ H is the closed subspace generated by the linear span of all the reduced words e1 ⊗⋯⊗en
with n ≥ 1 and such that e1 ∈ L.
• X2 ⊂ H is the closed subspace generated by the linear span of all the reduced words e1 ⊗⋯⊗en
with n ≥ 1 and such that en ∈ L.
• Y ⊂ H is the closed subspace generated by the linear span of all the reduced words e1 ⊗ ⋯⊗ en
with n ≥ 1 and such that e1, en ∈ (KR + iKR) ∩ L⊥.
Observe that we have
CΩ ⊕ (X1 + X2)
⋅ϕ ⊕ Y = H .
Claim 1. Let ε ≥ 0 and t ∈ R be such that Ut(L) ⊥ε~ dim(L) L. Then for all i ∈ {1, 2}, we have
κt(Xi) ⊥ε Xi .
24
Choose an orthonormal basis (ζ1, . . . , ζdim(L)) of L. We first prove the claim for X1. We will identify
X1 with L ⊗ H using the following unitary defined by
V1 ∶ H ⊗ H ∋ ζ ⊗ (e1 ⊗ ⋯ ⊗ en) ↦ ζ ⊗ e1 ⊗ ⋯ ⊗ en ∈ H ,
for all n ≥ 1 and all ζ, e1, . . . , en ∈ H. Observe that κtV1 = V1(Ut ⊗ κt) for all t ∈ R. Let ξ, η ∈ X1
be such that ξ = ∑dim(L)
ζj ⊗ νj with µi, νj ∈ H. Further observe that
ξ2 = ∑dim(L)
νj2. We have κtξ = ∑dim(L)
µi2 and η2 = ∑dim(L)
ζi ⊗ µi and η = ∑dim(L)
Utζi ⊗ κtµi and hence
j=1
i=1
j=1
i=1
i=1
⟨κtξ, η⟩ ≤
dim(L)
Q
i,j=1 ⟨Utζi, ζj⟩ µiνj .
Since ⟨Utζi, ζj⟩ ≤ ε~ dim(L), we obtain ⟨κtξ, η⟩ ≤ εξη by the Cauchy-Schwarz inequality.
Next, we prove the claim for X2. We identify X2 with H ⊗ L using the unitary defined by
V2 ∶ H ⊗ H ∋ (e1 ⊗ ⋯ ⊗ en) ⊗ ζ ↦ e1 ⊗ ⋯ ⊗ en ⊗ ζ ∈ H ,
for all n ≥ 1 and all ζ, e1, . . . , en ∈ H. Observe that κtV2 = V2(κt ⊗ Ut) for all t ∈ R. Let ξ, η ∈ X2
be such that ξ = ∑dim(L)
νj ⊗ ζj with µi, νj ∈ H. Further observe that
ξ2 = ∑dim(L)
νj2. We have κtξ = ∑dim(L)
µi2 and η2 = ∑dim(L)
µi ⊗ ζi and η = ∑dim(L)
κtµi ⊗ Utζi and hence
j=1
i=1
j=1
i=1
i=1
⟨κtξ, η⟩ ≤
dim(L)
Q
i,j=1 ⟨Utζi, ζj⟩ µiνj .
Since ⟨Utζi, ζj⟩ ≤ ε~ dim(L), we obtain ⟨κtξ, η⟩ ≤ εξη by the Cauchy-Schwarz inequality. This
finishes the proof of the claim.
Claim 2. For every x = (xn)ω ∈ (M ω)ϕω, we have
and
n→ωPX1(xnΩ)ϕ = 0
lim
n→ωPX2(xnΩ)ϕ = 0 .
lim
Let x ∈ (M ω)ϕω. We may assume that x ∈ Ball((M ω)ϕω
) and then choose a sequence (xn)n ∈
Mω(M) such that xn ∈ Ball(M) for all n ∈ N and x = (xn)ω. For all i ∈ {1, 2}, all t ∈ R and all
n ∈ N, we have
PXi(xnΩ)2
ϕ
ϕ = κtPXi(xnΩ)2
≤ 2κtPXi(xnΩ) − Pκt(Xi)(xnΩ)2
= 2Pκt(Xi)(κt(xnΩ) − xnΩ)2
≤ 2σϕ
t (xn) − xn2
ϕ + 2Pκt(Xi)(xnΩ)2
ϕ.
ϕ + 2Pκt(Xi)(xnΩ)2
ϕ
ϕ + 2Pκt(Xi)(xnΩ)2
ϕ
Furthermore, [AH12, Theorem 4.1] says that for all t ∈ R
(xn)ω = x = σϕω
t (x) = (σϕ
t (xn))ω
25
holds. This implies that limn→ω xn − σϕ
Fix p ≥ 1. Choose ε > 0 very small according to Proposition 4.2 so that ∏p−1
j=0(1 + δ○j(ε))2 ≤ 2. Since
U ∶ R → O(HR) is weakly mixing and since L is finite dimensional, with ε′ = ε~ dim(L), we can
choose inductively t1, . . . , t2p ∈ R such that
ϕ = 0 for all t ∈ R.
t (xn)#
Using Claim 1, this implies that
Utj(L) ⊥ε′ Uti(L), ∀1 ≤ i < j ≤ 2p .
κtj(X1) ⊥ε κti(X1) and κtj(X2) ⊥ε κti(X2), ∀1 ≤ i < j ≤ 2p .
Thus, using the above inequalities and Proposition 4.2, we obtain
ϕ = lim
n→ω
2p
Q
ϕ
lim
n→ω
≤ lim
n→ω
2pPXi(xnΩ)2
j=1κtj PXi(xnΩ)2
2p
2σϕ
tj(xn) − xn2
Q
j=1
4xn2
ϕ .
ϕ ≤ 22−p for all p ≥ 1. Thus, we have limn→ω PXi(xnΩ)ϕ = 0.
We conclude that limn→ω PXi(xnΩ)2
This finishes the proof of the claim.
Claim 3. The subspaces W(ξ1 ⊗ ⋯ ⊗ ξk) Y and Jϕσϕ
−i~2(W(ηℓ ⊗ ⋯ ⊗ η1))Jϕ Y are orthogonal in H.
Let m, n ≥ 1 and e1, . . . , em, f1 . . . , fn ∈ KR + iKR. Assume moreover that e1, em, f1, f n ∈ L⊥ so that
e1 ⊗ ⋯ ⊗ em ∈ Y and f1 ⊗ ⋯ ⊗ fn ∈ Y. Then by Proposition 2.1 (ii) and since ξk, η1 ∈ L, we have
2Pκtj (Xi)(xnΩ)2
ϕ + lim
n→ω
≤ lim
n→ω
2p
Q
j=1
ϕ
W(ξ1 ⊗ ⋯ ⊗ ξk)(e1 ⊗ ⋯ ⊗ em) = W(ξ1 ⊗ ⋯ ⊗ ξk)W(e1 ⊗ ⋯ ⊗ em)Ω
= W(ξ1 ⊗ ⋯ ⊗ ξk ⊗ e1 ⊗ ⋯ ⊗ em)Ω
= ξ1 ⊗ ⋯ ⊗ ξk ⊗ e1 ⊗ ⋯ ⊗ em ,
Jϕσϕ
−i~2(W(ηℓ ⊗ ⋯ ⊗ η1))Jϕ (f1 ⊗ ⋯ ⊗ fn) = W(f1 ⊗ ⋯ ⊗ fn)W(η1 ⊗ ⋯ ⊗ ηℓ)Ω
= W(f1 ⊗ ⋯ ⊗ fn ⊗ η1 ⊗ ⋯ ⊗ ηℓ)Ω
= f1 ⊗ ⋯ ⊗ fn ⊗ η1 ⊗ ⋯ ⊗ ηℓ.
Since ⟨ξ1, f1⟩ = 0, we see that the vectors
W(ξ1 ⊗ ⋯ ⊗ ξk)(e1 ⊗ ⋯ ⊗ em)
and
Jϕσϕ
−i~2(W(ηℓ ⊗ ⋯ ⊗ η1))Jϕ (f1 ⊗ ⋯ ⊗ fn)
are orthogonal in H. Finally, using the density of the linear span of the words e1 ⊗ ⋯ ⊗ em and
f1 ⊗ ⋯ ⊗ fn in Y finishes the proof of the claim.
We are now ready to finish the proof of Theorem 4.3. Let x, y ∈ (M ω)ϕω ⊖ C1. Using Claim 2 and
the fact that limn→ω PC Ω(xnΩ)ϕ = 0, we have
Λϕω(ax) = (W(ξ1 ⊗ ⋯ ⊗ ξk) xnΩ)ω
= (W(ξ1 ⊗ ⋯ ⊗ ξk) PY(xnΩ))ω ,
Λϕω(yb) = (Jϕσϕ
= (Jϕσϕ
−i~2(W(ηℓ ⊗ ⋯ ⊗ η1))Jϕ ynΩ)ω
−i~2(W(ηℓ ⊗ ⋯ ⊗ η1))Jϕ PY(ynΩ))ω .
26
By Claim 3, we know that
W(ξ1 ⊗ ⋯ ⊗ ξk) PY(xnΩ) ⊥ Jϕσϕ
−i~2(W(ηℓ ⊗ ⋯ ⊗ η1))Jϕ PY(ynΩ) ,
for all n ∈ N. Hence Λϕω(ax) ⊥ Λϕω(yb) in Hω, which implies that ϕω(b∗y∗ax) = 0.
Theorem 4.4. Let U ∶ R → O(HR) be any weakly mixing orthogonal representation on a separable
real Hilbert space and (M, ϕ) = (Γ(HR , Ut)′′, ϕU) the associated free Araki-Woods factor. Let Q ⊂ M
be any von Neumann subalgebra such that Q′ ∩ (M ω)ϕω ≠ C1. Then Q = C1.
Proof. Assume that Q′ ∩ (M ω)ϕω ≠ C1. Choose a projection e ∈ Q′ ∩ (M ω)ϕω such that e ∉ {0, 1}.
Then choose a sequence of projections (en)n ∈ Mω(M) such that e = (en)ω and limn→ω σϕ
t (en) −
ϕ = 0 for all t ∈ R. Put a = σ-weak limn→ω en ∈ Q′ ∩ M ϕ. Since M ϕ = C1, we obtain a = ϕ(a)1.
en#
Since e ∉ {0, 1}, we have ϕ(a) ∉ {0, 1}.
Let y ∈ Q ⊖ C1. By Theorem 4.3, we have
(e − ϕ(a)1)y2
ϕω = ϕω(y∗(e − ϕ(a)1)∗(e − ϕ(a)1)y) = ϕω(y∗(e − ϕ(a)1)∗y(e − ϕ(a)1)) = 0
Moreover,
ϕω = lim
n→ω(en−ϕ(a)1)y2
ϕ = lim
n→ω⟨(en−2ϕ(a)en+ϕ(a)21)yΩ, yΩ⟩ϕ = (ϕ(a)−ϕ(a)2)y2
ϕ .
(e−ϕ(a)1)y2
Since ϕ(a) ∉ {0, 1}, it follows that y = 0 and hence Q = C1.
Proof of Theorem C. Let Q ⊂ M be any von Neumann subalgebra that is globally invariant under
the modular automorphism group (σϕ
t ). There is a unique ϕ-preserving faithful normal conditional
expectation EQ ∶ M → Q. Assume that Q′ ∩ M ω ≠ C1. Then we have Q′ ∩ (M ω)ϕω ≠ C1 by Lemma
2.5. Therefore, we obtain Q = C1 by Theorem 4.4.
References
[AH12]
[AO75]
H. Ando and U. Haagerup. Ultraproducts of von Neumann algebras. J. Funct. Anal. 266,
6842 -- 6913, 2014.
C. A. Akemann and P. A. Ostrand. On a tensor product C∗ -algebra associated with the free
group on two generators. J. Math. Soc. Japan 27, 589 -- 599, 1975.
[BHR14] R. Boutonnet, C. Houdayer, and S. Raum. Amalgamated free product type III factors with at
most one Cartan subalgebra. Compos. Math. 150 (1), 143 -- 174, 2014.
[Cho74] M. D. Choi. A Schwarz inequality for positive linear maps on C∗-algebras. Illinois J. Math. 18,
565 -- 574, 1974.
[Con73] A. Connes. Une classification des facteurs de type III. Ann. Sci. Éc. Norm. Supér. (4) 6, 133 -- 252,
1973.
[Con74] A. Connes. Almost periodic states and factors of type III1. J. Funct. Anal. 16, 415 -- 445, 1974.
[Con76] A. Connes. Classification of injective factors. Cases II1, II∞, IIIλ, λ ≠ 1. Ann. Math. (2) 74,
73 -- 115, 1976.
27
[Hou07]
[Hou08]
C. Houdayer. Sur la classification de certaines algèbres de von Neumann. PhD thesis, Université
de Paris VII, 2007.
C. Houdayer. Structural results for free Araki-Woods factors and their continuous cores. J. Inst.
Math. Jussieu 9 (4), 741 -- 767, 2010.
[Hou12a] C. Houdayer. A class of II1 factors with an exotic abelian maximal amenable subalgebra. Trans.
Am. Math. Soc. 366, 3693 -- 3707, 2014.
[Hou12b] C. Houdayer. Structure of II1 factors arising from free Bogoljubov actions of arbitrary groups.
[Hou14]
[HR10]
[Ioa12]
[IPP05]
Adv. Math. 260, 414 -- 457, 2014.
C. Houdayer. Gamma stability in free product von Neumann algebras. arXiv:1403.4098, to appear
in Commun. Math. Phys.
C. Houdayer and É. Ricard. Approximation properties and absence of Cartan subalgebra for free
Araki-Woods factors. Adv. Math. 228 (2), 764 -- 802, 2011.
A. Ioana. Cartan subalgebras of amalgamated free product II1 factors. arXiv:1207.0054, to appear
in Ann. Sci. Éc. Norm. Supér.
A. Ioana, J. Peterson, and S. Popa. Amalgamated free products of weakly rigid factors and
calculation of their symmetry groups. Acta Math. 200 (1), 85 -- 153, 2008.
[Ocn85] A. Ocneanu. Actions of discrete amenable groups on von Neumann algebras, volume 1138 of
[Oza04]
[Oza10]
[Pet09]
[Pop83]
[Pop01]
[Pop02]
[Pop03]
[Pop04]
Lecture Notes in Mathematics. Berlin-Heidelberg-New York: Springer-Verlag, 1985.
N. Ozawa. Solid von Neumann algebras. Acta Math. 192 (1), 111 -- 117, 2004.
N. Ozawa. A comment on free group factors. In Noncommutative harmonic analysis with applica-
tions to probability II, volume 89 of Banach Center Publ., pages 241 -- 245. Polish Acad. Sci. Inst.
Math., Warsaw, 2010.
J. Peterson. L2-rigidity in von Neumann algebras. Invent. Math. 175 (2), 417 -- 433, 2009.
S. Popa. Maximal injective subalgebras in factors associated with free groups. Adv. Math. 50,
27 -- 48, 1983.
S. Popa. Some rigidity results for non-commutative Bernoulli shifts. J. Funct. Anal. 230 (2),
273 -- 328, 2006.
S. Popa. On a class of type II1 factors with Betti numbers invariants. Ann. Math. (2) 163 (3),
809 -- 899, 2006.
S. Popa. Strong rigidity of II1 factors arising from malleable actions of w-rigid groups. I. Invent.
Math. 165 (2), 369 -- 408, 2006.
S. Popa. Strong rigidity of II1 factors arising from malleable actions of w-rigid groups. II. Invent.
Math. 165 (2), 409 -- 451, 2006.
[Pop06a] S. Popa. Deformation and rigidity for group actions and von Neumann algebras. In M. Sanz-Solé
et al., editors, Proceedings of the international congress of mathematicians, Madrid, Spain, Au-
gust 22 -- 30, 2006, volume I: Plenary lectures and ceremonies, pages 445 -- 477. Zürich: European
Mathematical Society, 2007.
[Pop06b] S. Popa. On the superrigidity of malleable actions with spectral gap. J. Am. Math. Soc. 21 (4),
[Shl97]
[Shl98]
981 -- 1000, 2008.
D. Shlyakhtenko. Free quasi-free states. Pac. J. Math. 177 (2), 329 -- 368, 1997.
D. Shlyakhtenko. Some applications of freeness with amalgamation. J. Reine Angew. Math. 500,
191 -- 212, 1998.
D. Shlyakhtenko. A-valued semicircular systems. J. Funct. Anal. 166 (1), 1 -- 47, 1999.
[Shl99]
[Tak03] M. Takesaki. Theory of operator algebras III. Berlin-Heidelberg-New York: Springer-Verlag, 2003.
[Ued11]
[Vae06]
Y. Ueda. On type III1 factors arising as free products. Math. Res. Lett. 18 (5), 909 -- 920, 2011.
S. Vaes. États quasi-libres libres et facteurs de type III.
In Séminaire Bourbaki. 2003/2004,
volume 937 of Astérisque. Paris: Société Mathématique de France, 2007.
28
[VV05]
S. Vaes and R. Vergnioux. The boundary of universal discrete quantum groups, exactness, and
factoriality. Duke Math. J. 140 (1), 35 -- 84, 2007.
Cyril Houdayer
CNRS - Université Paris-Est - Marne-la-Vallée
LAMA UMR 8050
77454 Marne-la-Vallée cedex 2
France
[email protected]
Sven Raum
RIMS
Kitashirakawa-oiwakecho
606-8502 Sakyo-ku, Kyoto
Japan
[email protected]
29
|
1410.2051 | 5 | 1410 | 2017-01-09T19:30:56 | Inverse semigroup actions on groupoids | [
"math.OA"
] | We define inverse semigroup actions on topological groupoids by partial equivalences. From such actions, we construct saturated Fell bundles over inverse semigroups and non-Hausdorff \'etale groupoids. We interpret these as actions on C*-algebras by Hilbert bimodules and describe the section algebras of these Fell bundles.
Our constructions give saturated Fell bundles over non-Hausdorff \'etale groupoids that model actions on locally Hausdorff spaces. We show that these Fell bundles are usually not Morita equivalent to an action by automorphisms. That is, the Packer-Raeburn Stabilisation Trick does not generalise to non-Hausdorff groupoids. | math.OA | math |
INVERSE SEMIGROUP ACTIONS ON GROUPOIDS
ALCIDES BUSS AND RALF MEYER
Abstract. We define inverse semigroup actions on topological groupoids by
partial equivalences. From such actions, we construct saturated Fell bundles
over inverse semigroups and non-Hausdorff étale groupoids. We interpret these
as actions on C ∗-algebras by Hilbert bimodules and describe the section alge-
bras of these Fell bundles.
Our constructions give saturated Fell bundles over non-Hausdorff étale
groupoids that model actions on locally Hausdorff spaces. We show that these
Fell bundles are usually not Morita equivalent to an action by automorphisms.
That is, the Packer -- Raeburn Stabilisation Trick does not generalise to non-
Hausdorff groupoids.
Contents
Introduction
Inverse semigroup actions on groupoids
Inverse semigroup actions on C∗-algebras
1.
2. Partial equivalences
3.
3.1. Compatibility with order and involution
3.2. Transformation groupoids
3.3. Examples: group actions and actions on spaces
3.4. Morita invariance of actions by partial equivalences
3.5. Local centralisers
3.6. Decomposing proper Lie groupoids
4.
5. Fell bundles from actions of inverse semigroups
5.1. A Haar system on the transformation groupoid
5.2. Construction of the Fell bundle
5.3. Another construction of the Fell bundle
6. Actions of inverse semigroups and groupoids
6.1. The motivating example
6.2.
7. Actions by automorphisms are not enough
8. A simple explicit example
Appendix A. Preliminaries on topological groupoids
A.1. Topological groupoids, principal bundles, and equivalences
A.2. Basic actions versus free and proper actions
A.3. Covering groupoids and equivalence
Appendix B. Fields of Banach spaces over locally Hausdorff spaces
B.1. Proof of Theorem 5.5
References
Inverse semigroup models for étale groupoids
2
4
11
14
18
20
23
25
26
27
31
32
32
35
36
38
40
43
44
46
47
49
50
53
56
57
2010 Mathematics Subject Classification. 46L55, 20M18, 22A22.
Key words and phrases. Inverse semigroups, groupoids, actions, partial equivalences, Fell bun-
dles, stabilisation trick.
Supported by CNPq/CsF (Brazil) and the German Research Foundation (Deutsche Forschungs-
gemeinschaft (DFG)) through the grant "Actions of 2-groupoids on C*-algebras."
1
2
ALCIDES BUSS AND RALF MEYER
1. Introduction
Two of the most obvious actions of a groupoid G are those by left and right
If G is Hausdorff, they induce continuous
translations on its arrow space G1.
actions of G on the C∗-algebra C0(G1). What happens if G is non-Hausdorff?
Let G be a non-Hausdorff, étale groupoid with Hausdorff, locally compact object
space G0. Then G1 is locally Hausdorff, that is, it has an open covering U = (Ui)i∈I
by Hausdorff subsets: we may choose Ui so that the range and source maps restrict
to homeomorphisms from Ui onto open subsets of the Hausdorff space G0.
The covering U yields an étale, locally compact, Hausdorff groupoid H with
object space H 0 := Fi∈I Ui, arrow space H 1 := Fi,j∈I Ui ∩ Uj, range and source
maps r(i, j, x) := (i, x) and s(i, j, x) := (j, x), and multiplication (i, j, x) · (j, k, x) =
(i, k, x). The groupoid H is known as the Čech groupoid for the covering U. In
noncommutative geometry, we view the groupoid C∗-algebra C∗(H) as the algebra
of functions on the non-Hausdorff space G1.
Is there some kind of action of G
on C∗(H) that corresponds to the translation action of G on G1?
There is no action of G on C∗(H) in the usual sense because there is no action
of G on H by automorphisms. The problem is that arrows g ∈ G1 have many
liftings (i, g) ∈ H 0. To let g ∈ G act on H, we must choose k ∈ I with gh ∈ Uk for
h ∈ Uj with r(h) = s(g). It may, however, be impossible to choose k continuously
when h varies in Uj. This article introduces actions by partial equivalences in order
to make sense of the actions of G on H and C∗(H).
At first, we replace G by its inverse semigroup of bisections S = Bis(G). This
inverse semigroup cannot act on H by partial groupoid isomorphisms for the same
reasons as above. It does, however, act on H by partial equivalences because the
equivalence class of H is independent of the covering (see also [16, Lemma 4.1]); thus
partial homeomorphisms on G1 lift to partial equivalences of H in a canonical way.
We will see that an S-action by partial equivalences on a Čech groupoid for a locally
Hausdorff space Z is equivalent to an S-action on Z by partial homeomorphisms.
Let S act on a groupoid H by partial equivalences. Then we build a transforma-
tion groupoid H ⋊ S. Special cases of this construction are the groupoid of germs
for an action of S on a space by partial homeomorphisms, the semidirect product
for a group(oid) action on another group(oid) by automorphisms, and the linking
groupoid of a single Morita -- Rieffel equivalence. The original action is encoded in
the transformation groupoid L := H ⋊ S and open subsets Lt ⊆ L with
Lt · Lu = Ltu, L−1
t = Lt∗ , Lt ∩ Lu = [v≤t,u
Lv, L1 = [t∈S
Lt
and H = L1. We call such a family of subsets an S-grading on L with unit fibre H.
Any S-graded groupoid is a transformation groupoid for an essentially unique action
of S by partial equivalences on its unit fibre. This is a very convenient characteri-
sation of actions by partial equivalences.
An action of an inverse semigroup S on H by partial equivalences cannot induce,
in general, an action of S on C∗(H) by partial automorphisms in the usual sense
(as defined by Sieben [33]). But we do get an action by partial Morita -- Rieffel
equivalences, that is, by Hilbert bimodules. We show that actions of S by Hilbert
bimodules are equivalent to (saturated) Fell bundles over S. Along the way, we also
drastically simplify the definition of Fell bundles over inverse semigroups in [13].
Our approach clarifies in what sense a Fell bundle over an inverse semigroup is an
"action" of the inverse semigroup on a C∗-algebra.
In the end, we want an action of the groupoid G itself, not of the inverse semi-
group Bis(G). For actions by automorphisms, Sieben and Quigg [28] characterise
which actions of Bis(G) come from actions of G. We extend this characterisation to
INVERSE SEMIGROUP ACTIONS ON GROUPOIDS
3
Fell bundles: a Fell bundle over Bis(G) comes from a Fell bundle over G if and only
if the restriction of the action to idempotents in Bis(G) commutes with suprema of
arbitrarily large subsets. This criterion only works for Bis(G) itself. In practice, we
may want to "model" G by a smaller inverse semigroup S such that G0 ⋊ S ∼= G.
We characterise which Fell bundles over such S come from Fell bundles over G.
In particular, our action of Bis(G) on C∗(H) for the Čech groupoid associated
to G1 does come from an action of G, so we get Fell bundles over G that describe
the left and the right translation actions on G1. For these Fell bundles, we show
that the section C∗-algebras are Morita equivalent to C0(G0). More generally, for
any principal G-bundle X → Z, the section algebra of the Fell bundle over G that
describes the action of G on a Čech groupoid for X is Morita -- Rieffel equivalent
to C0(Z), just as in the more classical Hausdorff case (see Proposition 6.5).
For any action of an inverse semigroup S on a locally compact groupoid H by
partial equivalences, we identify the section C∗-algebra of the resulting Fell bundle
over S with the groupoid C∗-algebra of the transformation groupoid.
In brief
notation,
This generalises the well-known isomorphism
C∗(H) ⋊ S ∼= C∗(H ⋊ S).
C0(X) ⋊ S ∼= C∗(X ⋊ S)
for inverse semigroup actions on Hausdorff locally compact spaces by partial home-
omorphisms.
For a Hausdorff locally compact groupoid, any Fell bundle is equivalent to an
ordinary action on a stabilisation (Packer -- Raeburn Stabilisation Trick, see also
[6, Proposition 5.2]).
In contrast, our Theorem 7.1 shows that a non-Hausdorff
groupoid has no action by automorphisms that describes its translation action
on G1. Thus we really need Fell bundles to treat these actions of a non-Hausdorff
groupoid.
Now we explain the results of the individual sections of the paper.
In Section 2, we study partial equivalences between topological groupoids. We
show, in particular, that the involution that exchanges the left and right actions on
a partial equivalence behaves like the involution in an inverse semigroup.
Section 3 introduces inverse semigroup actions by partial equivalences. We show
that the rather simple-minded definition implies further structure, which is needed
to construct the transformation groupoid. Once we know that actions by partial
equivalences are essentially the same as S-graded groupoids, we treat many exam-
ples. This includes actions on spaces and Čech groupoids; in particular, an S-action
on a space by partial homeomorphisms induces an action by partial equivalences on
any Čech groupoid for a covering of the space. We describe a group action by (par-
tial) equivalences as a kind of extension by the group. We show that any (locally)
proper Lie groupoid is a transformation groupoid for an inverse semigroup action
on a very simple kind of groupoid: a disjoint union of transformation groupoids
of the form V ⋊ K, where V is a vector space, K a compact Lie group, and the
K-action on V is by an R-linear representation. This is meant as an example for
gluing together groupoids along partial equivalences.
In Section 4 we define inverse semigroup actions on C∗-algebras by Hilbert bimod-
ules. The theory is parallel to that for actions on groupoids by partial equivalences
because both cases have the same crucial algebraic features. We show that actions
by Hilbert bimodules are equivalent to saturated Fell bundles. This simplifies the
original definition of Fell bundles over inverse semigroups in [13].
In Section 5, we turn inverse semigroup actions on groupoids by partial equiva-
lences into actions on groupoid C∗-algebras by Hilbert bimodules. We do this in
4
ALCIDES BUSS AND RALF MEYER
two different (but equivalent) ways, by using transformation groupoids and abstract
functorial properties of our constructions. The approach using transformation
groupoids suggests that the section C∗-algebra of the resulting Fell bundle is simply
the groupoid C∗-algebra of the transformation groupoid: C∗(H) ⋊ S ∼= C∗(H ⋊ S).
We prove this, and a more general result for Fell bundles over H ⋊ S.
In Section 6 we relate inverse semigroup actions to actions of corresponding
étale groupoids.
In particular, we characterise when an action of Bis(G) comes
from an action of G. Finally, we can then treat our motivating example and turn a
groupoid action on a locally Hausdorff space Z into a Fell bundle over the groupoid.
We may also describe the section C∗-algebra in this case, which plays the role of the
crossed product. If the action is free and proper, then the result is Morita -- Rieffel
equivalent to C0(Z/G). We also define "proper actions" of inverse semigroups on
groupoids. We show that a free and proper action can only occur on a groupoid
that is equivalent to a locally Hausdorff and locally quasi-compact space.
Section 7 shows that the translation action of a non-Hausdorff étale groupoid
on its arrow space cannot be described by a groupoid action by automorphisms in
the usual sense. Our previous theory shows, however, that we may describe such
actions by groupoid Fell bundles. Thus the no-go theorem in Section 7 shows that
the Packer -- Raeburn Stabilisation Trick fails for non-Hausdorff groupoids, so Fell
bundles are really more general than ordinary actions in that case.
In Section 8, we examine a very simple explicit example to illustrate the no-go
theorem and to see how our main results avoid it.
Appendix A deals with topological groupoids, their actions on spaces and equiv-
alences between them. The main point is to define principal bundles and (Morita)
equivalence for non-Hausdorff groupoids in such a way that the theory works just
as well as in the Hausdorff case. Among others, we show that a non-Hausdorff
space is equivalent to a Hausdorff, locally compact groupoid if and only if it is
locally Hausdorff and locally quasi-compact, answering a question in [8].
Appendix B contains a general technical result about upper semicontinuous fields
of Banach spaces over locally Hausdorff spaces and uses it to prove C∗(H) ⋊ S ∼=
C∗(H ⋊S) for inverse semigroup actions on groupoids and a more general statement
involving Fell bundles over H ⋊ S.
2. Partial equivalences
In this section and the next one, we work in the category of topological spaces
and continuous maps, without assuming spaces to be Hausdorff or locally com-
pact. Appendix A shows how topological groupoids, their actions, principal bun-
dles, and equivalences between them should be defined so that the theory goes
through smoothly without extra assumptions on the underlying topological spaces.
Our main applications deal with groupoids that have a Hausdorff, locally com-
pact object space and a locally Hausdorff, locally quasi-compact arrow space. We
care about actions of such groupoids G on locally Hausdorff spaces Z. It is very
convenient to encode such an action by the transformation groupoid G ⋉ Z. Its
object space Z is only locally Hausdorff. When we allow such topological groupoids,
the usual definition of equivalence for topological groupoids breaks down because
orbit spaces of proper actions are always Hausdorff, so the actions on an equivalence
bispace cannot be proper unless the object spaces of the two groupoids are Haus-
dorff. Jean-Louis Tu's definition in [36] works -- it is equivalent to what we do. But
the theory becomes more elegant if we also drop the local compactness assumption
and thus no longer use proper maps in our basic definitions. The replacement for
free and proper actions are "basic" actions, which are characterised by the map
G ×s,G0,r X → X × X,
(g, x) 7→ (gx, x),
INVERSE SEMIGROUP ACTIONS ON GROUPOIDS
5
being a homeomorphism onto its image with the subspace topology from X × X.
Readers already familiar with the usual theory of locally compact groupoids
may read on and only turn to Appendix A in cases of doubt; they should note that
range and source maps of groupoids are assumed to be open, whereas anchor maps
of groupoid actions are not assumed open. Less experienced readers should read
Appendix A first.
Definition 2.1. Let G and H be topological groupoids. A partial equivalence
from H to G is a topological space with anchor maps r : X → G0 and s: X → H 0
and multiplication maps G1 ×s,G0,r X → X and X ×s,H0,r H 1 → X, which we write
multiplicatively, that satisfy the following conditions:
(P1) s(g · x) = s(x), r(g · x) = r(g) for all g ∈ G1, x ∈ X with s(g) = r(x), and
s(x · h) = s(h), r(x · h) = r(x) for all x ∈ X, h ∈ H 1 with s(x) = r(h);
(P2) associativity: g1 · (g2 · x) = (g1 · g2) · x, g2 · (x · h1) = (g2 · x) · h1, x · (h1 · h2) =
(x · h1) · h2 for all g1, g2 ∈ G1, x ∈ X, h1, h2 ∈ H 1 with s(g1) = r(g2),
s(g2) = r(x), s(x) = r(h1), s(h1) = r(h2);
(P3) the following two maps are homeomorphisms:
G1 ×s,G0,r X → X ×s,H0,s X,
X ×s,H0,r H 1 → X ×r,G0,r X,
(g, x) 7→ (x, g · x),
(x, h) 7→ (x, x · h);
(P4) s and r are open.
The first two conditions say that X is a G, H-bispace. The only difference be-
tween a partial and a global equivalence is whether the anchor maps are assumed
surjective or not: conditions (P1) -- (P4) are the same as conditions (E1) -- (E4) in
Proposition A.5.
We view a partial equivalence X from H to G as a generalised map from H
to G. Indeed, there is a bicategory with partial equivalences as arrows H → G
(Theorem 2.15).
Definition 2.2. Let G be a groupoid. A subset U ⊆ G0 is G-invariant if r−1(U ) =
s−1(U ). In this case, U and r−1(U ) = s−1(U ) are the object and arrow spaces of a
subgroupoid of G, which we denote by GU .
The canonical projection p: G0 → G0/G induces a bijection between G-invariant
subsets U ⊆ G0 and subsets p(U ) ⊆ G0/G. We are mainly interested in open
invariant subsets. Since p is open and continuous, open G-invariant subsets of G0
correspond to open subsets of G0/G.
Lemma 2.3. Let G and H be topological groupoids. A partial equivalence X from H
to G is the same as an equivalence from HV to GU for open, invariant subsets
U ⊆ G0, V ⊆ H 0. Here U = r(X), V = s(X).
Proof. Let U ⊆ G0 be G-invariant. A left GU -action is the same as a left G-action
U = U and G1 ×s,G0,r X ∼=
for which the anchor map takes values in U because G0
G1
U ×s,G0,r X if r(X) ⊆ U . Thus the commuting actions of GU and HV for an
equivalence from HV to GU may also be viewed as commuting actions of G and H,
respectively. This gives a partial equivalence (see Definition 2.1). Conversely, let X
be a partial equivalence. Let U := r(X) ⊆ G0 and V := s(X) ⊆ H 0. These are
open subsets because r and s are open, and they are invariant by (P1). The actions
of G and H are equivalent to actions of GU and HV , respectively. After replacing
G and H by GU and HV , respectively, all conditions (E1) -- (E5) in Proposition A.5
hold; thus X is an equivalence from HV to GU .
(cid:3)
6
ALCIDES BUSS AND RALF MEYER
Lemma 2.4. Let X be a partial equivalence from H to G and let U ⊆ G0 and
V ⊆ H 0 be invariant open subsets. Then
U XV := {x ∈ X r(x) ∈ U, s(x) ∈ V }
is again a partial equivalence from H to G.
We also write U X and XV for U XH0 and G0XV , respectively.
Proof. The subset U XV is open because r and s are continuous and U and V are
open, and it is invariant under the actions of G and H because U and V are invariant
and the two anchor maps are either invariant or equivariant with respect to the two
actions. Hence we may restrict the actions of G and H to U XV . Conditions
(P1) -- (P2) and (P4) in Definition 2.1 are inherited by an open invariant subspace.
The inverse to the first homeomorphism in (P3) maps U XV ×s,H0,s U XV into
G1 ×s,G0,r U XV , and the inverse to the second one maps U XV ×r,G0,r U XV
into U XV ×s,H0,r H 1. Thus U XV also inherits (P3) and is a partial equivalence
from H to G.
(cid:3)
Equivalences are partial equivalences, of course. In particular, the identity equiv-
alence G1 with G acting by left and right multiplication is also a partial equivalence.
Let X and Y be partial equivalences from H to G and from K to H, respectively.
Their composite is defined as for global equivalences, and still denoted by ×H:
X ×H Y := X ×s,H0,r Y / (x · h, y) ∼ (x, h · y),
equipped with the quotient topology and the induced actions of G and K by left
and right multiplication. The canonical map X ×s,H0,r Y → X ×H Y is a principal
H-bundle for the H-action defined by (x, y) · h := (x · h, h−1 · y); this follows from
the general theory in [21].
Example 2.5. We associate an equivalence Hf from G to H to a groupoid isomor-
phism f : G → H. The functor f consists of homeomorphisms f i : Gi → H i for
i = 0, 1. We take X = H 1 with the usual left H-action and the right G-action by
h · g := h · f 1(g) for all h ∈ H 1, g ∈ G1 with s(h) = r(f 1(g)) = f 0(r(g)); so the
right anchor map is (f 0)−1 ◦ s = s ◦ (f 1)−1.
We claim that an equivalence is of this form if and only if it is isomorphic to H 1
as a left H-space. Since H\H 1 ∼= H 0, the right anchor map gives a homeomorphism
H 0 → G0 in this case; let f 0 : G0 → H 0 be its inverse. The right action of g ∈ G1
on h ∈ H 1 with s(h) = f 0(r(g)) must be of the form h · g = h · f 1(g) for a unique
f 1(g) ∈ H 1 with r(f 1(g)) = f 0(r(g)) and s(f 1(g)) = s(h · g) = f 0(s(g)).
It is
routine to check that f 0 and f 1 give a topological groupoid isomorphism.
∼−→ H 1
When do two isomorphisms f, ϕ: G → H give isomorphic equivalences? Let
ϕ be an isomorphism. Define a continuous map σ : G0 → H 1 by
u : H 1
f
σ(x) := u(1f 0(x)) for all x ∈ G0. This satisfies r(σ(x)) = f 0(x) and s(σ(x)) = ϕ0(x)
for all x ∈ G0 because u is compatible with anchor maps. Since u is left H-invariant,
u(h) = u(h · 1s(h)) = h · (σ ◦ (f 0)−1 ◦ s)(h) for all h ∈ H 1, so σ determines u. The
right G-invariance of u translates to σ(r(g)) · ϕ1(g) = f 1(g) · σ(s(g)) for all g ∈ G.
Thus
(1)
ϕ0(x) = s(σ(x)),
ϕ1(g) = σ(r(g))−1 · f 1(g) · σ(s(g)).
Roughly speaking, f and ϕ differ by an inner automorphism.
Let an equivalence f : G → H and a continuous map σ : G0 → H 1 with r(σ(x)) =
f 0(x) for all x ∈ H 0 be given. Assume that H 0 → G0, x 7→ s(σ(x)),
is a
homeomorphism. Then (1) defines an isomorphism ϕ: G → H such that h 7→
h · σ((f 0)−1(s(h))) is an isomorphism between the equivalences Hf and Hϕ.
INVERSE SEMIGROUP ACTIONS ON GROUPOIDS
7
Example 2.6. If G and H are minimal groupoids in the sense that G0 and H 0 have
no proper open invariant subsets, then any partial equivalence is either empty or a
full equivalence G ∼−→ H. This holds, in particular, if G and H are groups.
Example 2.7. Any non-empty (partial) equivalence between two groups is isomor-
phic to one coming from a group isomorphism G ∼= H. Indeed, since X/H ∼= G0
and G\X ∼= H 0 are a single point, both actions on X are free and transitive. Fix
x0 ∈ X. Since the actions are free and transitive and part of principal bundles,
the maps G → X, g 7→ g · x0, and H → X, h 7→ x0 · h, are homeomorphisms.
The composite map G ∼−→ X ∼−→ H is an isomorphism of topological groups. This
isomorphism depends on the choice of x0. The isomorphisms G ∼−→ H for different
choices of x0 differ by an inner automorphism.
Lemma 2.8. The composition ×H is associative and unital with the identity equiv-
alence as unit, up to the usual canonical bibundle isomorphisms
(X ×H Y ) ×K Z ∼= X ×H (Y ×K Z),
G1 ×G X ∼= X ∼= X ×H H 1.
Proof. For global equivalences with arbitrary topological spaces, this is contained
in [21, Proposition 7.10]. The proofs in [21] can be extended to the partial case
as well. Alternatively, we may reduce the partial to the global case by restricting
our partial equivalences to global equivalences between open subgroupoids as in
Lemma 2.4. This works because
U (X ×H Y )V ∼= (U X) ×H (Y V )
for U ⊆ G0, V ⊆ K 0 open and invariant and partial equivalences X from H to G
and Y from K to H. Details are left to the reader.
(cid:3)
Proposition 2.9. Let G and H be topological groupoids. Let X1 and X2 be partial
equivalences from H to G. There is no bibundle map X1 → X2 unless r(X1) ⊆
r(X2) and s(X1) ⊆ s(X2). Any G, H-bibundle map ϕ: X1 → X2 is an isomorphism
onto the open sub-bibundle r(X1)X2 = X2s(X1). The map ϕ is invertible if r(X2) ⊆
r(X1) or s(X2) ⊆ s(X1). In this case, r(X2) = r(X1) and s(X2) = s(X1).
Proof. Since rX2 ◦ ϕ = rX1 and sX2 ◦ ϕ = sX1, we must have r(X1) ⊆ r(X2) and
s(X1) ⊆ s(X2) if there is a bibundle map ϕ: X1 → X2. Assume this from now
on. The image of a bibundle map is contained in r(X1)X2 and in X2s(X1). Since
r(X1) ⊆ r(X2) and s(X1) ⊆ s(X2), we have r(r(X1)X2) = r(X1) and s(X2s(X1)) =
s(X1). All remaining assertions now follow once we prove that a bibundle map
ϕ: X1 → X2 is invertible if r(X1) = r(X2) or s(X1) = s(X2). We treat the case
r(X2) = r(X1); the other one is proved in the same way, exchanging left and right.
Since Xi is a partial equivalence, it is a principal H-bundle over Xi/H ∼= r(Xi).
The map ϕ induces a homeomorphism on the base spaces because r(X2) = r(X1)
both carry the subspace topology from G0. Hence ϕ is a homeomorphism by [21,
Proposition 5.9].
(cid:3)
In particular, the restricted multiplication maps G1
U ×H X ⊆ G1 ×G X → X and
V ⊆ X ×H H 1 → X are bibundle maps. Proposition 2.9 shows that they
X ×H H 1
induce bibundle isomorphisms
(2)
G1
U ×G X ∼= U X,
X ×H H 1
V
∼= XV .
Partial equivalences carry extra structure similar to an inverse semigroup. The
adjoint operation is the following:
Definition 2.10. Given a partial equivalence X from H to G, we define the dual
partial equivalence X ∗ by exchanging the left and right actions on X. More precisely,
X ∗ is X as a space, the anchor maps r∗ : X ∗ → H 0 and s∗ : X ∗ → G0 are r∗ = sX
8
ALCIDES BUSS AND RALF MEYER
and s∗ = rX , and the left H- and right G-actions are defined by h ·∗ x = x · h−1
and x ·∗ g := g−1 · x, respectively.
If X gives an equivalence from HV to GU for open invariant subsets U ⊆ G0,
V ⊆ H 0, then X ∗ gives the "inverse" equivalence from GU to HV .
The following properties of duals are trivial:
• naturality: a bibundle map X → Y induces a bibundle map X ∗ → Y ∗;
• (X ∗)∗ = X;
• there is a natural isomorphism σ : (X ×H Y )∗ ∼= Y ∗ ×H X ∗, (x, y) 7→ (y, x),
with σ2 = Id.
Let Map(Y1, Y2) be the space of bibundle maps between two partial equivalences
Y1, Y2 from H to G.
Proposition 2.11. Let X be a partial equivalence from H to G. Then there are
natural isomorphisms
X ∗ ×G X ∼= H 1
that make the following diagrams of isomorphisms commute:
X ×H X ∗ ∼= G1
r(X),
s(X)
(3)
X ×H X ∗ ×G X
X ×H H 1
s(X)
G1
r(X) ×G X
X,
X ∗ ×G X ×H X ∗
X ∗ ×G G1
r(X)
H 1
s(X) ×G X ∗
X ∗.
If K is another groupoid and Y and Z are partial equivalences from K to G and
from K to H, respectively, with r(Y ) ⊆ r(X) and r(Z) ⊆ s(X), then there are
natural isomorphisms
Map(X ×H Z, Y ) ∼= Map(Z, X ∗ ×G Y ),
Map(Y, X ×H Z) ∼= Map(X ∗ ×G Y, Z).
Both map the subsets of bibundle isomorphisms onto each other.
r(X) and X ∗ ×G X ∼= H 1
Proof. Lemma 2.3 shows that X is an equivalence from Hs(X) to Gr(X). Hence
the usual theory of groupoid equivalence gives canonical isomorphisms X ×H X ∗ ∼=
G1
s(X). The first one maps the class of (x1, x2) with s(x1) =
s(x2) to the unique g ∈ G1 with x1 = g · x2. In particular, it maps [x, x] 7→ 1r(x).
The second one maps the class of (x1, x2) with r(x1) = r(x2) to the unique h ∈ H 1
with x2 = x1 · h.
7→ 1s(x). Then the composite
isomorphisms X ×H X ∗ ×G X → X and X ∗ ×G X ×H X ∗ → X ∗ map [x, x, x] 7→ x,
respectively. Since any element in X ×H X ∗ ×G X or X ∗ ×G X ×H X ∗ has a
representative of the form (x, x, x), we get the two commuting diagrams in (3).
In particular, it maps [x, x]
The assumption r(Y ) ⊆ r(X) implies s(X ∗ ×G Y ) = s(Y ) because for any
y ∈ Y there is x ∈ X ∗ with (x, y) ∈ X ∗ ×G Y . Similarly, r(Z) ⊆ s(X) implies
s(X ×H Z) = s(Z). By Proposition 2.9, a bibundle map X ×H Z → Y exists only
if s(Z) ⊆ s(Y ), and then it is an isomorphism onto Y s(Z); and a bibundle map
Z → X ∗ ×G Y exists only if s(Z) ⊆ s(Y ), and then it is an isomorphism onto
X ∗ ×G Y s(Z). Thus we may as well replace Y by Y s(Z) to achieve s(Y ) = s(Z);
then all bibundle maps X ×H Z → Y or Z → X ∗ ×G Y are bibundle isomorphisms.
INVERSE SEMIGROUP ACTIONS ON GROUPOIDS
9
The second isomorphism reduces in a similar way to the case where also s(Y ) = s(Z)
and where we are dealing only with bibundle isomorphisms.
A bibundle map ϕ: X ×H Z → Y induces IdX ∗ ×G ϕ: X ∗ ×G X ×H Z → X ∗ ×G Z;
we compose this with the natural isomorphism
X ∗ ×G X ×H Z ∼= H 1
s(X) ×H Z ∼= s(X)Z = Z
to get a bibundle map Z → X ∗ ×G Y ; here we used s(X) ⊇ r(Z). We claim
that this construction gives the desired bijection between Map(X ×H Z, Y ) and
Map(Z, X ∗ ×G Y ). Since composing with an isomorphism is certainly a bijection,
it remains to show that
Map(X ×H Z, Y ) → Map(X ∗ ×G X ×H Z, X ∗ ×G Y ), ϕ 7→ IdX ∗ ×G ϕ,
is bijective. Since X ×H X ∗ ∼= G1
r(X) and r(X) ⊇ r(Y ), we have natural isomor-
phisms X ×H X ∗ ×G Y ∼= Y and X ×H X ∗ ×G X ×H Z ∼= X ×H Z. Naturality
means that they intertwine ϕ 7→ IdX×H X ∗ ×G ϕ and ϕ. Since IdX×H X ∗ ×G ϕ =
IdX ×H IdX ∗ ×G ϕ, we see that ϕ 7→ IdX ∗ ×G ϕ is injective and has ψ 7→ IdX ×H ψ
for ψ : Z → X ∗ ×G Y as a one-sided inverse. The same argument also shows that
ψ 7→ IdX ×H ψ is injective, so both constructions are bijective.
(cid:3)
Applying duality, we also get bijections Map(Z ∗×H X ∗, Y ∗) ∼= Map(Z ∗, Y ∗×GX)
and Map(Y ∗, Z ∗ ×H X ∗) ∼= Map(Y ∗ ×G X, Z ∗) under the same hypotheses.
The canonical isomorphisms
(4)
X ×H X ∗ ×G X ∼= X,
X ∗ ×G X ×H X ∗ ∼= X ∗
from Proposition 2.11 characterise X ∗ uniquely in the following sense:
Proposition 2.12. Let X and Y be partial equivalences from H to G and from G
to H, respectively. If there are bibundle isomorphisms
X ×H Y ×G X ∼= X,
Y ×H X ×G Y ∼= Y,
then there is a unique bibundle isomorphism X ∗ ∼= Y such that the composite map
(5)
X ∼= X ×H X ∗ ×G X ∼= X ×H Y ×G X ∼= X
is the identity map.
Proof. When we multiply the inverse of the isomorphism X ×H Y ×G X ∼= X on
both sides with X ∗ and use (2), we get an isomorphism
X ∗ ∼= X ∗ ×G X ×H X ∗ ∼= X ∗ ×G X ×H Y ×G X ×H X ∗
∼= H 1
s(X) ×H Y ×G G1
r(X)
∼= s(X)Y r(X).
This implies s(X) = r(X ∗) ⊆ r(Y ) and r(X) = s(X ∗) ⊆ s(Y ) by Proposition 2.9.
Exchanging X and Y , the isomorphism Y ×H X ×G Y ∼= Y gives s(Y ) ⊆ r(X) and
r(Y ) ⊆ s(X). Hence r(Y ) = s(X) and s(Y ) = r(X), so s(X)Y r(X) = Y . This
gives an isomorphism α : X ∗ ∼−→ Y .
A diagram chase using the commuting diagrams in (3) shows that the composite
of the map X ×H X ∗ ×G X → X ×H Y ×G X induced by the isomorphism α and
the given isomorphism X ×H Y ×G X → Y (which we used to construct α) is the
canonical map X ×H X ∗ ×G X → X as in (4). Hence the composite in (5) is the
identity map for the isomorphism α.
The isomorphisms in Proposition 2.11 give a canonical bijection
Map(X ∗, Y ) ∼= Map(X ∗ ×G X ×H X ∗, Y ) ∼= Map(X ×H X ∗, X ×H Y )
∼= Map(X, X ×H Y ×G X) ∼= Map(X, X).
10
ALCIDES BUSS AND RALF MEYER
Inspection shows that it maps an isomorphism X ∗ ∼−→ Y to the composite map
in (5). Hence there is only one isomorphism X ∗ ∼−→ Y for which the composite map
in (5) is the identity map.
(cid:3)
Proposition 2.13. Let X be a partial equivalence from G to itself and let µ: X ×G
X → X be a bibundle isomorphism. Then there is a unique isomorphism ϕ: X ∼−→
U for an open G-invariant subset U ⊆ G0 such that the following diagram com-
G1
mutes:
(6)
X ×G X
ϕ ×G ϕ
G1
U ×G G1
U
µ
µ0
X
ϕ
G1
U ,
µ0(g1, g2) = g1 · g2.
Hence r(X) = s(X) and µ is associative.
Proof. The isomorphism µ induces an isomorphism
X ×G X ×G X
µ×GIdX
−−−−−→ X ×G X
µ
−→ X
Hence Y = X satisfies the two conditions in Proposition 2.11 that ensure X = Y ∼=
X ∗. This gives an isomorphism ϕ: X ∼= X ×G X ∼= X ×G X ∗ ∼= G1
r(X). Since ϕ is
a bibundle map, the diagram (6) commutes if and only if µ is the composite map
X ×G X
ϕ×GIdX
−−−−−→ G1
r(X) ×G X ∼= X,
where the map G1
Sending an isomorphism ϕ: X → G1
of the bijections in Proposition 2.11, namely, the first one for X = Y = Z:
r(X) ×G X ∼= X is the left multiplication map, [g, x] 7→ g · x.
∼= X ×G X ∗ to this composite map is one
r(X)
Map(X, G1
s(X)) ∼= Map(X, X ∗ ×G X) ∼= Map(X ×G X, X).
Hence there is exactly one isomorphism ϕ that corresponds under this bijection
to µ.
(cid:3)
Proposition 2.11 implies that isomorphism classes of partial equivalences from G
to itself form an inverse semigroup fpeq(G). The idempotents in this inverse semi-
group are in bijection with G-invariant open subsets of G0 by Proposition 2.13.
These are, in turn, in bijection with open subsets of the orbit space G0/G by the
definition of the quotient topology on G0/G. These also correspond to the idempo-
tents of the inverse semigroup pHomeo(G0/G) of partial homeomorphisms of the
topological space G0/G.
A partial equivalence X from H to G induces a partial homeomorphism
X∗ : H 0/H ⊆ s(X) → r(X) ⊆ G0/G
by X∗([h]) = [g] if there is x ∈ X with s(x) ∈ [h], r(x) ∈ [g]. If Y is another partial
equivalence from K to H, then (X ×H Y )∗ = X∗ ◦ Y∗ by definition. This gives a
canonical homomorphism of inverse semigroups
fpeq(G) → pHomeo(G0/G).
Remark 2.14. The homomorphism fpeq(G) → pHomeo(G0/G) is neither injective
nor surjective in general, although it is always an isomorphism on the semilattice
of idempotents. Consider, for instance, the disjoint union G = Z/3 ⊔ {pt}. This
groupoid is a group bundle, and G0/G has two points. The partial homeomorphism
that maps one point to the other does not lift to a partial equivalence because the
stabilisers are not the same and equivalences must preserve the stabiliser groups.
The group Z/3 has non-inner automorphisms, so there are non-isomorphic partial
INVERSE SEMIGROUP ACTIONS ON GROUPOIDS
11
equivalences of G defined on Z/3 that induce the same partial homeomorphism
on G0/G.
In our definition of an inverse semigroup action (see Sections 3 and 4 below),
certain isomorphisms of partial equivalences are a crucial part of the data. We
could not construct transformation groupoids and Fell bundles without them. If we
identify isomorphic partial equivalences as above, then we can no longer talk about
two isomorphisms of partial equivalences being equal. The correct way to take into
account isomorphisms of partial equivalences is through a bicategory (see [1, 6, 19]).
The following remarks are intended for readers familiar with bicategories.
Our bicategory has topological groupoids as objects and partial equivalences
as arrows. Let G and H be topological groupoids and let X1 and X2 be partial
equivalences from H to G. As 2-arrows X1 ⇒ X2, we take all G, H-bibundle
isomorphisms X1 → X2, so all 2-arrows are invertible. The vertical product of
2-arrows is the composition of bibundle maps. Unit 2-arrows are identity maps
on partial equivalences. The composition of arrows is ×H. The unit arrow on a
topological groupoid G is G1 with the standard bibundle structure. Lemma 2.8
provides invertible 2-arrows
(X ×H Y ) ×K Z ⇒ X ×H (Y ×K Z),
G1 ×G X ⇒ X ⇐ X ×H H 1,
which we take as associator and left and right unit transformations. Let X1, X2
be partial equivalences from H to G and let Y1, Y2 be partial equivalences from K
to H. The horizontal product of two bibundle maps f : X1 → X2 and g : Y1 → Y2
is f ×H g : X1 ×H Y1 → X2 ×H Y2.
Theorem 2.15. The data above defines a bicategory peq.
Proof. It is routine to check that partial equivalences from H to G with bibundle
maps between them form a category C(G, H) for the vertical product of bibundle
maps, and that the composition of partial equivalences with the horizontal prod-
uct of bibundle maps is a functor C(G, H) × C(H, K) → C(G, K). The associator
and both unit transformations are natural isomorphisms of functors; the associa-
tor is clearly compatible with unit transformations and makes the usual pentagon
commute, see [19, p. 2].
(cid:3)
Remark 2.16. We still get a bicategory if we allow all bibundle maps as 2-arrows.
We restrict to invertible 2-arrows to get the correct notion of inverse semigroup
actions below.
An arrow f : x → y in a bicategory is called an equivalence if there are an arrow
g : y → x and invertible 2-arrows g ◦ f ⇒ Idx and f ◦ g ⇒ Idy. The equivalences
in peq are precisely the global bibundle equivalences.
The duality X 7→ X ∗ with the canonical flip maps (X ×H Y )∗ ∼−→ Y ∗ ×H X ∗
gives a functor I : peq → peqop with I 2 = Idpeq. It seems useful to formalise the
properties of this functor and look for examples in more general bicategories. But
we shall not go into this question here.
3. Inverse semigroup actions on groupoids
We give two equivalent definitions for actions of inverse semigroups on topological
groupoids by partial equivalences. The first is exactly what it promises to be.
The second, more elementary, definition does not mention groupoids or partial
equivalences.
Let S be an inverse semigroup with unit 1. Let G be a topological groupoid.
Definition 3.1. An action of S on G by partial equivalences consists of
• partial equivalences Xt from G to G for t ∈ S;
12
ALCIDES BUSS AND RALF MEYER
• bibundle isomorphisms µt,u : Xt ×G Xu
∼−→ Xtu for t, u ∈ S;
satisfying
(A1) X1 is the identity equivalence G1 on G;
(A2) µt,1 : Xt ×G G1 ∼−→ Xt and µ1,u : G1 ×G Xu
∼−→ Xu are the canonical iso-
morphisms, that is, the left and right G-actions, for all t, u ∈ S;
(A3) associativity: for all t, u, v ∈ S, the following diagram commutes:
(Xt ×G Xu) ×G Xv
ass
Xt ×G (Xu ×G Xv)
µt,u ×G IdXv
IdXt ×G µu,v
Xtu ×G Xv
Xt ×G Xuv
µtu,v
Xtuv
µt,uv
If S has a zero object 0, then we may also ask X0 = ∅.
Remark 3.2. Let S be an inverse semigroup possibly without 1. We may add a unit
1 formally and extend the multiplication by 1 · s = s = s · 1 for all s ∈ S ∪ {1}.
If partial equivalences (Xt)t∈S and bibundle isomorphisms (µt,u)t,u∈S are given
satisfying associativity for all t, u, v ∈ S, then we may extend this uniquely to an
action of S ∪ {1}: we put X1 := G1 and let µt,1 and µ1,u be the right and left
G-action, respectively. The associativity condition is trivial if one of t, u, v is 1, so
associativity holds for all t, u, v ∈ S ∪{1}. As a result, an action of S ∪{1} by partial
equivalences is the same as (Xt)t∈S and (µt,u)t,u∈S satisfying only Condition (A3).
Similarly, we may add a zero 0 to S and extend the multiplication by 0 · s = 0 =
s · 0 for all s ∈ S ∪ {0}. We extend an S-action by X0 := ∅, so that X0 ×G Xt = ∅ =
Xt ×G X0, leaving no choice for the maps µt,0, µ0,u : ∅ → ∅. This gives an action of
S ∪ {0} with X0 = ∅.
If 0, 1 ∈ S and we ask no conditions on X0 and X1, then r(Xt), s(Xt) ⊆ r(X1) =
s(X1) for all t ∈ S, and Xt restricted to r(X0) = s(X0) is the trivial action where
all Xt act by the identity equivalence. Hence all the action is on the locally closed,
invariant subset r(X1) \ r(X0) ⊆ G0. The conditions on X0 and X1 merely rule out
such degeneracies.
Remark 3.3. An inverse semigroup may be viewed as a special kind of category
with only one object, which is also a very special kind of bicategory. An inverse
semigroup action by partial equivalences is exactly the same as a functor from this
category to the bicategory peq of partial equivalences (see [19]).
Lemma 3.4. For an inverse semigroup action (Xt, µt,u), we have r(Xt) = r(Xtt∗ ) =
s(Xtt∗) = s(Xt∗) and s(Xt) = s(Xt∗t) = r(Xt∗t) = r(Xt∗ ) for each t ∈ S.
Proof. If e ∈ S idempotent, then Proposition 2.13 applied to the isomorphism
µe,e : Xe ×G Xe ∼= Xe gives r(Xe) = s(Xe). The existence of an isomorphism
µt,t∗ : Xt ×G Xt∗ ∼= Xtt∗ implies r(Xt) ⊇ r(Xtt∗ ) and s(Xt∗) ⊇ s(Xtt∗). Similarly,
the isomorphism µtt∗,t gives r(Xtt∗ ) ⊇ r(Xt), and µt,t∗t gives s(Xt∗t) ⊇ s(Xt). Now
everything follows.
(cid:3)
Definition 3.5. Let S be an inverse semigroup with unit. A simplified action of S
on a topological groupoid consists of
• a topological space G0;
• topological spaces Xt for t ∈ S;
• continuous maps s, r : Xt → G0;
• continuous maps
µt,u : Xt ×s,G0,r Xu → Xtu,
(x, y) 7→ x · y,
for t, u ∈ S;
INVERSE SEMIGROUP ACTIONS ON GROUPOIDS
13
satisfying
(S1) s(x · y) = s(y), r(x · y) = r(x) for all t, u ∈ S, x ∈ Xt, y ∈ Xu with
s(x) = r(y);
(S2) r : Xt → G0 and s: Xt → G0 are open for all t ∈ S;
(S3) the maps r, s: X1 → G0 are surjective;
(S4) µt,u is surjective for each t, u ∈ S;
(S5) the map
Xt ×s,G0,r Xu → Xu ×s,G0,s Xtu,
(x, y) 7→ (y, x · y),
is a homeomorphism if t = 1 and u ∈ S;
(S6) the map
Xt ×s,G0,r Xu → Xt ×r,G0,r Xtu,
(x, y) 7→ (x, x · y),
is a homeomorphism if t ∈ S and u = 1;
(S7) for all t, u, v ∈ S, the following diagram commutes:
(Xt ×s,G0,r Xu) ×s,G0,r Xv
µt,u ×s,G0,r IdXv
Xtu ×s,G0,r Xv
(7)
ass
µtu,v
µt,uv
Xtuv
Xt ×s,G0,r (Xu ×s,G0,r Xv)
IdXt ×s,G0,r µu,v
Xt ×s,G0,r Xuv
If S has a zero element, we may also ask X0 = ∅.
This definition is more elementary because it does not mention groupoids or
partial equivalences. It seems less elegant than Definition 3.1, but is simpler because
much of the complexity of Definition 3.1 is hidden in the conditions (P1) -- (P4)
defining partial equivalences of topological groupoids.
It is clear that an inverse semigroup action by partial equivalences gives a sim-
plified action: forget the multiplication on G1 and the left and right actions of G
on the spaces Xt. The isomorphisms in (S5) for t = 1 and in (S6) for u = 1 are
those in (P3), and all other conditions in Definition 3.5 are evident. The converse
is more remarkable:
Proposition 3.6. Any simplified inverse semigroup action on groupoids comes
from a unique action by partial equivalences. Thus actions and simplified actions
of inverse semigroups by partial equivalences are equivalent. Furthermore, the maps
in (S5) and (S6) are isomorphisms and the maps µt,u are open for all t, u ∈ S.
Proof. The spaces G0 and G1 := X1 with range and source maps r and s and mul-
tiplication µ1,1 satisfy the conditions (G1) -- (G4) in Proposition A.1 because these
are special cases of our conditions (S1) -- (S7). Hence this data defines a topological
groupoid. Similarly, the anchor maps r : Xt → G0 and s: Xt → G0 and the multi-
plication maps µ1,t and µt,1 satisfy conditions (P1) -- (P4) in Definition 2.1 and thus
turn Xt into a partial equivalence from G to itself.
Let t, u ∈ S. The associativity of the maps µ for t, 1, u, 1, t, u and t, u, 1 implies
that µt,u descends to a G, G-bibundle map ¯µt,u : Xt ×G Xu → Xtu. Since µt,u is
surjective by (S4), so is ¯µt,u. Hence it is a bibundle isomorphism by Proposition 2.9.
The groupoid structure on X1 and the left and right actions on Xt are defined so
that X1 is the identity equivalence on G and the maps ¯µ1,u and ¯µt,1 are the canon-
ical isomorphisms. The associativity condition for the bibundle isomorphisms ¯µt,u
follows from the corresponding property of the maps µt,u. Thus we have got an
action by partial equivalences. This is the only action that simplifies to the given
data because of the assumptions about X1, µ1,u, and µt,1 in Definition 3.1.
14
ALCIDES BUSS AND RALF MEYER
By definition, Xt ×G Xu is the orbit space of the G-action on Xt ×s,G0,r Xu by
(x1, x2) · g := (x1 · g, g−1 · x2). The canonical projection Xt ×s,G0,r Xu → Xt ×G Xu
is open by Proposition A.3. The map µt,u is the composite of this projection with
the homeomorphism ¯µt,u : Xt ×G Xu → Xtu, hence it is also open.
Finally, we check that the maps in (S5) are isomorphisms for all t, u ∈ S; ex-
changing left and right gives the same for the maps in (S6). The map in (S5) is
G-equivariant if we let G act on Xt ×s,G0,r Xu by g · (x, y) := (xg−1, gy) and on
Xu ×s,G0,s Xtu by g · (y, x) := (gy, x). Both actions are part of principal bundles:
the bundle projection on Xt ×s,G0,r Xu is the canonical map to Xt ×G Xu, and the
bundle projection on Xu×s,G0,sXtu is s×G0,sIdXtu to Xtur(Xu). Our G-equivariant
map induces the map µt,u on the base spaces, which is a homeomorphism. Hence
so is the map on the total spaces by [21, Proposition 5.9].
(cid:3)
3.1. Compatibility with order and involution. Let S be an inverse semigroup
with unit. Define a partial order on S by t ≤ u if t = tt∗u or, equivalently, t = ut∗t.
The multiplication and involution preserve this order: t1t2 ≤ u1u2 and t∗
1 if
t1 ≤ u1 and t2 ≤ u2 (see [18]).
1 ≤ u∗
Let (Xt)t∈S, (µt,u)t,u∈S be an action of S on G. We are going to prove that the
action is compatible with this partial order and the involution on S. To prepare for
the proofs of analogous statements for inverse semigroup actions on C∗-algebras,
we give rather abstract proofs, which carry over literally to the C∗-algebraic case.
Proposition 3.7. There are unique bibundle maps ju,t : Xt → Xu for t, u ∈ S
with t ≤ u such that the following diagrams commute for all t1, t2, u1, u2 ∈ S with
t1 ≤ u1, t2 ≤ u2:
(8)
Xt1 ×G Xt2
ju1,t1 ×G ju2,t2
Xu1 ×G Xu2
µt1,t2
µu1,u2
Xt1t2
ju1u2,t1t2
Xu1u2
The map ju,t is a bibundle isomorphism onto Xus(Xt) = r(Xt)Xu. We have jt,t =
IdXt for all t ∈ S and jv,u ◦ ju,t = jv,t for t ≤ u ≤ v in S.
Proof. Let E(S) ⊆ S be the subset of idempotents and let e ∈ E(S). Proposi-
tion 2.13 gives a unique isomorphism Xe ∼= G1
Ue intertwining µe,e : Xe ×G Xe → Xe
Ue; here Ue := r(Xe) = s(Xe) is an open G-invariant
and the multiplication in G1
subset of G0. The diagram (8) for (e, e) ≤ (1, 1) shows that j1,e has to be this
particular isomorphism Xe ∼= G1
Ue ⊆ G1. To simplify notation, we now identify Xe
with G1
Ue for all e ∈ E(S) using these unique isomorphisms, and we transfer the
multiplication maps µs,t for idempotent s, t or st accordingly. This gives an iso-
morphic action of S by partial equivalences. So we may assume that Xe = G1
Ue and
that µe,e : Xe ×G Xe → Xe is the usual multiplication map on G1
Ue for all e ∈ E(S).
Let e ∈ E(S) and let t, u ∈ S satisfy t∗t ≤ e and uu∗ ≤ e. Thus te = t, eu = u
∼−→
and teu = tu. We show that µt,e : Xt ×G G1
Xu are the obvious maps µ0
e,u from the left and right G-actions in this
case. Associativity of the multiplication maps gives us a commuting diagram of
isomorphisms
Ue → Xt and µe,u : G1
t,e or µ0
Ue ×G Xu
Xt ×G Xe ×G Xu
µt,e ×G IdXu
Xt ×G Xu
IdXt ×G µe,u
µt,u
Xt ×G Xu
µt,u
Xtu
INVERSE SEMIGROUP ACTIONS ON GROUPOIDS
15
We may cancel the isomorphism µt,u to get IdXt ×G µe,u = µt,e ×G IdXu. Now we
consider two cases: t = e or e = u. If t = e, then µt,e = µ0
t,e is the multiplication
map on G1
Ue. Hence so is µt,e ×G IdXu. Thus µe,u and µ0
e,u induce the same map
G1
Ue ×G G1
Ue ×G Xu → G1
Ue because
s(Xe) = Ue ⊇ r(Xu) ⊇ r(G1
e,u if e is idempotent and
e ≥ uu∗. A similar argument in the other case e = u gives µt,e = µ0
Ue ×G Xu. We may use (2) to cancel the factor G1
Ue ×G Xu). Thus µe,u = µ0
t,e if t∗t ≤ e.
Now let t ≤ u, that is, t = tt∗u = ut∗t. Then we get two candidates for the
bibundle map ju,t : Xt → Xu:
(9)
Xt
Xt
µtt∗ ,u←−−−−
∼=
µu,t∗ t←−−−−
∼=
Xtt∗ ×G Xu = G1
Utt∗ ×G Xu
Xu ×G Xt∗t = Xu ×G G1
Ut∗ t
µ0
tt∗ ,u−−−−→
∼=
µ0
u,t∗ t−−−−→
∼=
Utt∗ Xu ⊆ Xu,
XuUt∗ t ⊆ Xu.
We claim that both maps Xt → Xu are equal, so we get only one map ju,t : Xt → Xu.
Let e = tt∗ and f = t∗t. Then there is a commuting diagram of isomorphisms
Xe ×G Xu ×G Xf
IdXe ×G µu,f
µ0 ×G IdXf
IdXe ×G µ0
Ue Xu ×G Xf
(10)
µe,u ×G IdXf
Xe ×G Xt
µe,t = µ0
e,t
Xt ×G Xf
Xe ×G XuUf
µu,f
µe,u
Xt
µt,f = µ0
t,f
The large rectangle commutes by associativity. The argument above gives µe,t =
e,t and µt,f = µ0
µ0
t,f . The lower left and upper right triangles commute because µe,u
and µu,f are bibundle maps, so they are compatible with µ0. Hence the interior
quadrilateral commutes. Thus the two definitions of ju,t in (9) are equal.
The first construction of ju,t in (9) gives the unique map for which the diagram (8)
commutes for (e, t) ≤ (1, u) and the inclusion map j1,e. Since we already saw
that j1,e is unique, the diagrams (8) characterise the bibundle maps ju,t uniquely
for all t ≤ u in S. The map jt,t is the identity on Xt because µtt∗,t = µ0
Now let t ≤ u ≤ v, define e = tt∗ and f = uu∗ and identify Xe and Xf with
subsets of G1. In the following diagram, we abbreviate ×G to ∗, and µ0 denotes
the left and right actions for subsets of G1:
tt∗,t.
ju,t
µe,u
Xt
Xe ∗ Xu
µ0
Ue Xu
µe,v
Id ∗ µf,v
Ue µf,v
jv,t
Xe ∗ Xv
µe,f ∗ Id
Xe ∗ Xf ∗ Xv
µ0 ∗ Id
Ue Xf ∗ Xv
Ue jv,u
µ0
Ue Xv
µ0
Id ∗ µ0
Xe ∗ (Uf Xv)
µ0
µ0
Ue Xv
16
ALCIDES BUSS AND RALF MEYER
The top left square commutes because the multiplication maps are associative, the
top right square because they are bibundle maps. The bottom left square commutes
because µe,f = µ0, and the bottom right square commutes for trivial reasons. The
bent composite arrows are the maps j by construction. Thus the whole diagram
commutes, and this means that jv,u ◦ ju,t = jv,t.
If t1 ≤ u1 and t2 ≤ u2 in S, then there is a commuting diagram of isomorphisms
Xt1 ∗ Xt2
µt1 ,t2
Xt1t2
µt1 t∗
1
,u1
∗µu2 ,t∗
2
t2
µt1 t∗
1
,u1u2 ,t∗
2
t2
(11)
Xt1t∗
1
∗ Xu1 ∗ Xu2 ∗ Xt∗
2 t2
Id∗µu1 ,u2 ∗Id
Xt1t∗
1
∗ Xu1u2 ∗ Xt∗
2 t2
µ0
µ0
Ut1 t∗
1
Xu1 ∗ Xu2Ut∗
2
t2
µu1 ,u2
Ut1 t∗
1
Xu1u2 Ut∗
2
t2
t ,u1,u2,t∗
Here we abbreviate ×G to ∗, µ0 denotes the left and right actions for subsets of G1,
2t2 denotes the appropriate combination of two multiplication maps,
and µttt∗
which is well-defined by associativity. The upper square commutes by associativity.
The lower square commutes because µu1,u2 is a bibundle map. The left vertical
is ju1,t1 ∗ ju2,t2 because the two
isomorphism from Xt1 ∗ Xt2 to Ut1 t∗
constructions in (9) coincide. It remains to see that the right vertical isomorphism
from Xt1t2 to Ut1 t∗
Xu1 ∗ Xu2Ut∗
is ju1u2,t1t2.
Xu1u2 Ut∗
t2
t2
1
2
The proof of this is similar to the proof that the two maps in (9) coincide.
1. Since r(Xt1t2) = Ue and (11) is a diagram of
2t2. Furthermore,
∗Xu1u2 ∗Xt∗
Let e = (t1t2)(t1t2)∗, so e ≤ t1t∗
isomorphisms, we have Xe∗Xt1t∗
the isomorphism
∗Xu1u2 ∗Xt∗
∼= Xt1t∗
2t2
1
1
1
2
µe,t1t∗
1
∗ Id : Xe ∗ Xt1t∗
1
∗ Xu1u2 ∗ Xt∗
2 t2 → Xe ∗ Xu1u2 ∗ Xt∗
2 t2
is equal to the standard multiplication map µ0
1. This fact
and associativity show that the right vertical isomorphism in (11) is equal to the
composite map
∗ Id because e ≤ ttt∗
e,t1t∗
1
Xt1t2
µe,u1 u2 ,t2 t∗
2
←−−−−−−−−
∼=
Xe ∗ Xu1u2 ∗ Xt∗
2 t2
µ0
−→
∼=
Ue Xu1u2 Ut∗
2
t2
= Ue Xu1u2.
Similarly, we get the same composite map if we replace t∗
2t2 on the right by the
smaller idempotent f = (t1t2)∗(t1t2). Now the diagram (10) shows that the map
we get is ju1u2,t1t2 as desired. Hence (8) commutes.
(cid:3)
Remark 3.8. Let E be a semilattice with unit 1, viewed as an inverse semigroup.
An E-action on a topological groupoid G is the same as a unital semilattice map
from E to the lattice of open G-invariant subsets of G0, that is, a map e 7→ Ue
satisfying U1 = G0 and Ue ∩ Uf = Uef for all e, f ∈ E. The corresponding action by
partial equivalences is defined by Xe := G1
→ G1
.
Proposition 3.7 implies that every action of E is isomorphic to one of this form.
Ue and µe,f = µ0 : G1
Ue ×G G1
Uf
Uef
Proposition 3.9. There are unique bibundle isomorphisms Jt : X ∗
which the following composite map is the identity:
t → Xt∗ for
(12)
Xt
∼= Xt ×G X ∗
t ×G Xt
IdXt ×GJt×GIdXt
−−−−−−−−−−−−→ Xt ×G Xt∗ ×G Xt
µt,t∗ ,t−−−−→ Xt.
INVERSE SEMIGROUP ACTIONS ON GROUPOIDS
17
These involutions also make the following diagrams commute:
Xt ×G X ∗
t
G1
Utt∗
(13)
IdXt ×G Jt
t ×G Xt
X ∗
Jt ×G IdXt
G1
Ut∗ t
Xt ×G Xt∗
Xtt∗
µt,t∗
Xt∗ ×G Xt
Xt∗t
µt∗,t
Here the unlabelled arrows are the canonical isomorphisms from Propositions 2.11
and 2.13. Furthermore, (Jt∗)∗ ◦ Jt : X ∗
is the identity map for all
t ∈ S and the following diagrams commute for all t, u, v ∈ S with t ≤ u:
t → Xt∗ → X ∗
t
(14)
u ×G X ∗
v
X ∗
Ju ×G Jv
Xu∗ ×G Xv∗
µ∗
v,u
µu∗,v∗
X ∗
vu
Jvu
Xu∗v∗
X ∗
t
Jt
Xt∗
j∗
u,t
X ∗
uUtt∗
JuUtt∗
Xu∗Utt∗
ju∗,t∗
Write x∗ := Jt(x) for x ∈ Xt and µt,u(x, y) = x · y for x ∈ Xt, y ∈ Xu with
s(x) = r(y). The above diagrams and equations of maps mean that the involution
is characterised by x · x∗ · x = x for all x ∈ Xt and has the properties x · x∗ = 1r(x),
x∗ · x = 1s(x), (x∗)∗ = x, (x · y)∗ = y∗ · x∗, and ju∗,t∗(x∗) = ju,t(x)∗.
Proof. The two isomorphisms µt,t∗,t : Xt ×G Xt∗ ×G Xt → Xt and µt∗,t,t∗ : Xt∗ ×G
Xt ×G Xt∗ → Xt∗ that we may build from µ are equal by associativity. Propo-
∼= Xt∗ for
sition 2.12 for these isomorphisms gives a unique isomorphism Jt : X ∗
t
which (12) becomes the identity map.
We claim that (12) is the identity if and only if either of the diagrams in (13)
commutes. The proofs for both cases differ only by exchanging left and right, so
we only write down one of them. Assume that the first diagram in (13) commutes.
Applying the functor ␣×GIdXt to it, we get that the isomorphism (12) is the identity
map because the multiplication map µtt∗,t : Xtt∗ ×G Xt → Xt is just the left action
if we identify Xtt∗ ∼= G1
Utt∗ as usual. Conversely, assume that the isomorphism
in (12) is the identity map. Take a further product with Xt∗ and then identify
Xt ×G Xt∗ ∼= Xtt∗ via µt,t∗. Using again that the multiplication with Xtt∗ is just
the G-action, this gives the first diagram in (13).
Next we show that Jt∗ = (J −1
t
commuting diagram
)∗, which implies J ∗
t∗ ◦ Jt = IdXt . We use the
X ∗
t∗ ×G Xt∗
(J −1
t
)∗ ×G J −1
t
Xt ×G X ∗
t
IdXt ×G Jt
Xt ×G Xt∗
G1
Utt∗
G1
Utt∗
µt,t∗
Xtt∗
t
The top rectangle commutes because the pairing X ×G X ∗ → G1
r(X) is natural. The
bottom diagram is the first one in (13). The large rectangle is the second diagram
in (13) for t∗ with (J −1
)∗ instead of Jt∗. Since this diagram characterises Jt∗, we
get Jt∗ = (J −1
)∗ as asserted.
t
Since the involution Jvu is uniquely characterised by a diagram like the first one
in (13), we may prove the first diagram in (14) by showing that the composite map
µu∗,v∗ ◦ (Ju ×G Jv) ◦ (µ∗
vu → Xu∗v∗ also makes the diagram in (13) for
t = vu commute. This is a routine computation using the same diagrams for Ju
and Jv and that the multiplication maps involving Xe for idempotent e ∈ S are
always given by the left or right action because of the compatibility with j1,e. This
v,u)−1 : X ∗
18
ALCIDES BUSS AND RALF MEYER
proof is a variant of the usual proof that (xy)−1 = y−1x−1 in a group because
y−1x−1 · (xy) = 1.
Similarly, we get the second diagram in (14) by showing that the composite
t → Xt∗ satisfies the defining condition for Jt because ju,t
(cid:3)
map j−1
and ju∗,t∗ are compatible with the multiplication maps.
u∗,t∗ ◦ Ju ◦ j∗
u,t : X ∗
3.2. Transformation groupoids. Let (Xt, µt,u)t,u∈S be an action of a unital in-
verse semigroup S on a topological groupoid G by partial equivalences. Define the
embeddings ju,t : Xt → Xu for t ≤ u in S and the involutions X ∗
t → Xt∗ as in
Propositions 3.7 and 3.9.
Let X := Ft∈S Xt and define a relation ∼ on X by (t, x) ∼ (u, y) for x ∈ Xt,
y ∈ Xu if there are v ∈ S with v ≤ t, u and z ∈ Xv with jt,v(z) = x and ju,v(z) = y.
Lemma 3.10. The relation ∼ is an equivalence relation. Equip X∼ := X/∼ with
the quotient topology. The quotient map π : X → X∼ is a local homeomorphism. It
restricts to a homeomorphism from Xt onto an open subset of X∼ for each t ∈ S.
Thus X∼ is locally quasi-compact or locally Hausdorff if and only if all Xt are.
Proof. It is clear that ∼ is reflexive and symmetric. For transitivity, take (t1, x1) ∼
(t2, x2) ∼ (t3, x3). Then there are t12 ≤ t1, t2, t23 ≤ t2, t3, x12 ∈ Xt12, and x23 ∈
Xt23 with jti,t12(x12) = xi for i = 1, 2 and jti,t23(x23) = xi for i = 2, 3. Thus
s(x12) = s(x2) = s(x23) ∈ s(t23) = s(t∗
23t23, so that t ≤
∼= Xt.
t12 and t ≤ t2t∗
Let x be the image of x12 under this isomorphism. Then jt12,t(x) = x12. Hence
jti,t(x) = jti,t12(jt12,t(x)) = xi for i = 1, 2. Since jt2,t23(x23) = x2 = jt2,t23(jt23,t(x))
and jt2,t23 is injective by Proposition 2.9, we get jt23,t(x) = x23 and hence also
jt3,t(x) = x3. Thus x1 ∼ x3 as desired.
2t23 = t23. We have x12 ∈ Xt12 Ut∗
23t23). Let t := t12t∗
∼= Xt12 ×G Xt∗
23t23
t23
23
We prove that π is open. Any open subset of X is a disjoint union of open
subsets of the spaces Xt; so π is open if and only if all the maps Xt → X∼ are
open. Let U ⊆ Xt be open, then we must check that π−1(π(U )) is open. This set
is a union over the set of triples t, v, w ∈ S with w ≤ t, v, where the set for t, v, w
is contained in Xv and consists of all jv,w(x) with x ∈ j−1
t,w(U ). The map jv,w is
open by Proposition 2.9, and jt,w is continuous, so jv,w(j−1
t,w(U )) is open. Hence
π−1(π(U )) is open as a union of open subsets of X, showing that π is open.
If (t, x) ∼ (t, y), then there are u ≤ t and z ∈ Xu with x = jt,u(z) = y; so
the map from Xt to X∼ is injective. Since π is open and continuous, it restricts
to a homeomorphism from Xt onto an open subset of X∼. Thus π is a local
homeomorphism. Since being locally Hausdorff or locally quasi-compact are local
properties and π is a local homeomorphism, X∼ has one of these two properties if
and only if X has, if and only if each Xt has.
(cid:3)
The space X∼ need not be Hausdorff, just as for étale groupoids constructed
from inverse semigroup actions on spaces, where X∼ will be the groupoid of germs
of the action (by Theorem 3.18).
From now on, we identify Xt with its image in X∼, using that πXt : Xt → X∼
is a homeomorphism onto an open subset by Lemma 3.10.
We are going to turn X∼ into a topological groupoid with the same object
space G0 as G. Since ju,t is a bibundle map, it is compatible with range and
source maps. So the maps r, s: Xt ⇒ G0 induce well-defined maps r, s: X∼ ⇒ G0.
The multiplication maps µt,u give a continuous map X ×s,G0,r X → X, by map-
ping the t, u-component of X ×s,G0,r X to the tu-component of X by µt,u. Equa-
tion (8) shows that this descends to a well-defined continuous map µ: X∼ ×s,G0,r
X∼ → X∼.
INVERSE SEMIGROUP ACTIONS ON GROUPOIDS
19
Lemma 3.11. The maps r, s: X∼ ⇒ G0 and µ: X∼ ×s,G0,r X∼ → X∼ define a
topological groupoid X∼. It contains G as an open subgroupoid. Hence X∼ is étale
if and only if G is.
Proof. The multiplication is associative already on X by (A3) and the associativity
of S. The maps r and s are open on X∼ because they are so on each Xt. The
maps r, s, µ restricted to G1 = X1 reproduce the groupoid structure on G by (A1).
Even more, (A2) implies that multiplication in X∼ with elements of X1 is the same
as the G-action. In particular, unit elements in G1 act identically, so they remain
If x ∈ Xt, then x∗ ∈ Xt∗ satisfies µt,t∗(x, x∗) = 1r(x) by
unit elements in X∼.
Proposition 3.9. Hence
π(x, t) · π(x∗, t∗) := π(µt,t∗ (x, x∗), tt∗) = π(1r(x), tt∗).
This is equivalent to the unit element (1r(x), 1) in X∼ because j1,tt∗ is the usual
inclusion map (more precisely, the computation above assumes that we identify
Xtt∗ ∼= G1
Utt∗ ⊆ G1 using j1,tt∗ ). Similarly, π(x, t) · π(x∗, t∗) ∼ (1s(x), 1) is a
unit element. Thus π(x∗, t∗) is inverse to π(x, t). The map π(x, t) 7→ π(x∗, t∗)
is continuous. Thus we have a topological groupoid. We have seen above that it
contains G as an open subgroupoid. Therefore, X∼ is étale if and only if G is. (cid:3)
Definition 3.12. The groupoid X∼ is called the transformation groupoid of the
S-action (Xt, µt,u) on G and denoted by G ⋊ S, or by G ⋊Xt,µt,u S if the action
must be specified.
Our proof shows that G ⋊ S with the family of open subsets (Xt)t∈S encodes
all the algebraic structure of our action by partial equivalences. The next defi-
nition characterises when a groupoid H with a family of subsets (Ht)t∈S is the
transformation groupoid of an inverse semigroup action.
Definition 3.13. Let S be an inverse semigroup. A (saturated) S-grading on a
topological groupoid H is a family of open subsets (Ht)t∈S of H 1 such that
(Gr1) Ht · Hu = Htu for all t, u ∈ S;
(Gr2) H −1
t = Ht∗ for all t ∈ S;
(Gr3) Ht ∩ Hu =Sv≤t,u Hv for all t, u ∈ S;
(Gr4) H 1 =St∈S Ht.
If S has a zero element 0, we may also require H0 = ∅.
The conditions (Gr1) and (Gr2) imply that H1 is a subgroupoid of H, called the
unit fibre of the grading. (Gr4) and (Gr1) imply that s(H1) = r(H1) = H 0. (Gr3)
implies Hv ⊆ Hu for v ≤ u.
A non-saturated S-grading would be defined by weakening (Gr1) to Ht·Hu ⊆ Htu
for all t, u ∈ S. We only use saturated gradings and drop the adjective.
Theorem 3.14. Let S be an inverse semigroup with unit. The transformation
groupoid G⋊S of an S-action on a groupoid G by partial equivalences is an S-graded
groupoid. Any S-graded groupoid (H, (Ht)t∈S) is isomorphic to one of this form,
where G0 = H 0 and G1 = H1 ⊆ H. Two actions by partial equivalences are
isomorphic if and only if their transformation groupoids are isomorphic in a grading-
preserving way.
Here an isomorphism between actions (Xt)t∈S and (Yt)t∈S by partial equiva-
lences on two groupoids G and H means the obvious thing: a family of homeo-
morphisms Xt ∼= Yt compatible with the range, source, and multiplication maps in
Definition 3.5.
20
ALCIDES BUSS AND RALF MEYER
Proof. It follows directly from our construction that the subspaces Xt ⊆ G ⋊ S for
an S-action by partial equivalences satisfy (Gr1) -- (Gr4). It is also clear that the
transformation groupoid construction is natural for isomorphisms of S-actions.
Let H with the subspaces Ht for t ∈ S be an S-graded topological groupoid.
Then G1 := H1 with G0 = H 0 is an open subgroupoid of H. Let Xt = Ht
with the restriction of the range and source map of H, and with the G-action and
maps µt,u : Xt ×s,G0,r Xu → Xtu from the multiplication map in H. This satisfies
(S3) by definition, (S4) because Ht · Hu = Htu, (S1) and (S7) because H is a
groupoid, and (S2) because Xt is open in H and the range and source maps of H
are open. If (y, z) ∈ Xu ×s,G0,s Xtu, then zy−1 ∈ XtuXu∗ = Xtuu∗ ⊆ Xt because
tuu∗ ≤ t. Hence (y, z) 7→ (zy−1, y) gives a continuous inverse for the map in (S5),
so that the latter is a homeomorphism. A similar argument shows that the map
in (S6) is a homeomorphism. Thus we get an S-action by partial equivalences.
This construction is natural in the sense that isomorphic S-graded groupoids give
isomorphic actions by partial equivalences actions.
If we start with an action by partial equivalences, turn it into a graded group-
oid, and then back into an action by partial equivalences, then we get an isomor-
phic action by construction. When we start with a graded groupoid, go to an
action by partial equivalences and back to a graded groupoid, then we also get
back our original S-graded groupoid. The only non-trivial point is that the map
π : Ft∈S Ht →(cid:0)Ft∈S Ht(cid:1)∼ identifies x ∈ Ht and y ∈ Hu for t, u ∈ S if and only if
x = y in H; this is exactly the meaning of (Gr3).
(cid:3)
3.3. Examples: group actions and actions on spaces. The equivalence be-
tween actions by partial equivalences and graded groupoids makes it easy to de-
scribe all actions of groups on groupoids and all actions of inverse semigroups on
spaces.
Theorem 3.15. Let G be a topological groupoid and let S be a group, viewed as an
inverse semigroup. Then an S-action on G by (partial) equivalences is equivalent
to a groupoid H containing G as an open subgroupoid with H 0 = G0, and with a
continuous groupoid homomorphism π : H ։ S such that π−1(1) = G and, for each
x ∈ H 0 and t ∈ S there is h ∈ H 1 with s(h) = x and π(h) = t. In this situation,
H is the transformation groupoid G ⋊ S. If G is also a group, this is the same as
a group extension G H ։ S.
Proof. Since tt∗ = 1 for any t ∈ S, any action of S by partial equivalences will be
an action by global equivalences. By Theorem 3.14, we may replace an S-action by
partial equivalences by an S-graded groupoid (H, (Ht)t∈S). We have H 0 = G0 by
construction. Since S is a group, (Gr3) says that Ht ∩ Hu = ∅ for t 6= u. Thus we
get a well-defined map π : H 1 → S with π−1(t) = Ht; in particular, G = π−1(1).
The map π is continuous because the subsets Ht are open. The condition on the
existence of h for given x, t says that the map s: Ht → H 0 is onto, that is, Ht is
a global equivalence. Thus an S-action on G gives π : H → S with the asserted
properties.
For the converse, let π : H → S be a groupoid homomorphism as in the statement.
Define Ht := π−1(t) ⊆ H 1. These are open subsets because π is continuous. If
t, u ∈ S, then HtHu ⊆ Htu is trivial. If h ∈ Htu, then our technical assumption
gives h2 ∈ Hu with s(h2) = s(h). Then h1 := hh−1
2 ∈ Ht, so h ∈ HtHu. Thus (Gr1)
holds. The remaining conditions for an S-grading are trivial in this case, and H is
the transformation groupoid G ⋊ S by construction.
If G is also a group, then so is H because G0 = H 0, and then the condition on π
simply says that it is a surjection π : H ։ S with kernel G. This is the same as a
group extension.
(cid:3)
INVERSE SEMIGROUP ACTIONS ON GROUPOIDS
21
The obvious definition of a group action by automorphisms on another group
only covers split group extensions. We need some kind of twisted action by auto-
morphisms to allow for non-trivial group extensions as well. Our notion of action
by equivalences achieves this very naturally.
For groupoid extensions, one usually requires the kernel to be a group bundle;
this need not be the case here. There are many examples of groupoid homomor-
phisms (or 1-cocycles) with the properties required in Theorem 3.15. We mention
one typical case:
Example 3.16. Let H be the groupoid associated to a self-covering σ : X → X of
a compact space X as in [9]. The canonical Z-valued cocycle π : H → Z on it
clearly has the properties needed to define a Z-grading on H. The subgroupoid
G := π−1(0) is the groupoid that describes the equivalence relation generated by
x ∼ y if σk(x) = σk(y) for some k ∈ N. The action of σ on X preserves this
equivalence relation and hence gives an endomorphism of G; this endomorphism
is an equivalence, and our Z-action on G by equivalences is generated by this self-
equivalence of G. But unless σ is a homeomorphism, σ is not invertible on G, so it
gives no action of Z by automorphisms.
Now we turn to actions of inverse semigroups on topological spaces. Let S be an
inverse semigroup with unit and let Z be a topological space. First we recall Exel's
construction of the groupoid of germs for an inverse semigroup action by partial
homeomorphisms [12].
Let pHomeo(Z) be the inverse semigroup of partial homeomorphisms of Z. An
action of S on Z by partial homeomorphisms is a monoid homomorphism θ : S →
pHomeo(Z). This gives partial homeomorphisms θt : Dt∗t → Dtt∗ for t ∈ S with
open subsets De ⊆ Z for e ∈ E(S). The groupoid of germs has object space Z,
and its arrows are the "germs" [t, z] for t ∈ S, z ∈ Dt∗t; by definition, [t, z] = [u, z′]
if and only if there is e ∈ E(S) with z = z′ ∈ De and te = ue. The groupoid
structure is defined by s[t, z] = z, r[t, z] = θt(z), [t, z] · [u, z′] = [tu, z′] if z = θu(z′),
and [t, z]−1 = [t∗, θt(z)]. The subsets {[t, z] z ∈ U } for t ∈ S and an open subset
U ⊆ Dt∗t form a basis for the topology on the arrow space.
Remark 3.17. Many authors use another germ relation that only requires an open
subset V of Z with z ∈ V and θtV = θuV . This may give a different groupoid, of
course. Exel's germ groupoids need not be essentially principal (see [32]).
Theorem 3.18. Let Z be a topological space viewed as a topological groupoid, and
let S be an inverse semigroup with unit. Isomorphism classes of actions of S on Z
by partial equivalences are in natural bijection with actions of S on Z by partial
homeomorphisms. The transformation groupoid Z ⋊ S for an action by partial
equivalences is the groupoid of germs defined by Exel [12].
Proof. Let θ : S → pHomeo(Z) be an action of S by partial homeomorphisms.
Exel's groupoid of germs carries an obvious S-grading by the open subsets Xt :=
{[t, z] z ∈ Dt∗t} with X1 = Z. The conditions in Definition 3.13 are trivial to
check. Hence Exel's groupoid is the transformation groupoid Z ⋊ S for an action
of S on Z by partial equivalences by Theorem 3.14. Conversely, an S-action on Z is
equivalent to the S-graded groupoid Z ⋊S. This groupoid is étale. The assumptions
of an S-grading imply that the subsets Xt ⊆ Z ⋊ S form an inverse semigroup of
bisections that satisfies the assumptions in [12, Proposition 5.4], which ensures that
the groupoid of germs is Z ⋊ S.
(cid:3)
Corollary 3.19. Let G be an étale groupoid, let S be an inverse semigroup, and
let f : S → Bis(G) be a semigroup homomorphism. This induces an isomorphism
22
ALCIDES BUSS AND RALF MEYER
G0 ⋊ S ∼= G if and only if St∈S f (t) = G and f (t) ∩ f (u) =Sv∈S,v≤t,u f (v) for all
t, u ∈ S.
Here G0 ⋊ S uses the action of S on G0 induced by f and the usual action
of Bis(G).
Proof. Add a unit to S and map it to the unit bisection G0 ⊆ G, so that we may
apply Theorem 3.14. For t ∈ S, let Gt := f (t) ⊆ G1; these are open subsets
because each f (t) is a bisection. Since f is a semigroup homomorphism, (Gr1)
and (Gr2) hold. The other two conditions are exactly the technical assumptions
of the corollary. Thus these two assumptions are equivalent to (Gt)t∈S being an
S-grading on G. If they hold, then Theorem 3.14 says that G ∼= G1 ⋊ S = Z ⋊ S.
Conversely, the transformation groupoid Z ⋊ S is S-graded by Theorem 3.14, so if
G ∼= Z ⋊ S, then it satisfies the two technical assumptions.
(cid:3)
A subsemigroup S ⊆ Bis(G) with the properties required in Corollary 3.19 is
called wide. Corollary 3.19 explains why they appear so frequently (see, for instance,
[3, 12, 28]). [12, Proposition 5.4] already shows that Z ⋊ S = G if S is wide, but we
have not seen the converse statement yet.
Since the proof of Theorem 3.18 is not explicit, we give another pedestrian proof.
Let θt : Dt∗t → Dtt∗ for t ∈ S give an action of S on Z by partial homeomor-
phisms. This is a groupoid isomorphism from Dt∗t to Dtt∗, which we turn into a
partial equivalence from Z to itself as in Example 2.5. Here this means that we take
X ′
(z). Since all arrows in Z are
units, the range and source maps determine the partial equivalence. The homeo-
t to Xt = Dt∗t with r(z) := θt(z)
morphism θt gives a bibundle isomorphism from X ′
and s(z) := z. The comparison with Exel's groupoid is more obvious for the second
choice, which we take from now on.
t := Dtt∗ with anchor maps r′(z) := z, s′(z) := θ−1
t
There is an obvious homeomorphism
Xt ×Z Xu
∼−→ {z ∈ Dtt∗ θt(z) ∈ Duu∗} = D(tu)∗(tu),
such that the range and source maps are θtu and the inclusion map, respectively.
We choose this isomorphism for µt,u to define our action by partial equivalences.
Actually, this is no choice at all because the range and source maps are injective here,
so there is at most one bibundle map between any two partial equivalences. (We
will see more groupoids with this property in Section 3.5.) Hence the associativity
condition in the definition of an inverse semigroup action holds automatically. Thus
we have turned an action by partial homeomorphisms on Z into an action by partial
equivalences on Z, viewed as a topological groupoid.
Our construction of the transformation groupoid above is exactly the construc-
tion of the groupoid of germs in this special case, so the isomorphism between Z ⋊S
and the groupoid of germs from [12] is trivial.
Next we check that every partial equivalence X of Z is isomorphic to one coming
from a partial homeomorphism. Since all arrows in Z are units, we have X/Z =
X = Z\X. Hence the anchor maps G0 ← X → H 0 are continuous, open and
injective by condition (P3) in Proposition A.5. The map θ := r ◦ s−1 : s(X) → r(X)
is a partial homeomorphism from G0 to H 0, and r : X → r(X) is an isomorphism
of partial equivalences from X to Xθ.
Since there is always only one isomorphism between partial equivalences coming
from the same partial homeomorphism, an inverse semigroup action on Z is deter-
mined uniquely by the isomorphism classes of the Xt, which are in bijection with
partial homeomorphisms of Z. This proves the first statement in Theorem 3.18.
INVERSE SEMIGROUP ACTIONS ON GROUPOIDS
23
3.4. Morita invariance of actions by partial equivalences.
Proposition 3.20. Let Y be an equivalence from H to G, and let (Xt, µt,u) be an
t := Y ×G Xt ×G Y ∗ and let
action of an inverse semigroup with unit on G. Let X ′
µ′
t,u : X ′
tu be the composite isomorphism
t ×H X ′
u → X ′
Y ×G Xt ×G Y ∗ ×H Y ×G Xu ×G Y ∗ ∼−→ Y ×G Xt ×G G1 ×G Xu ×G Y ∗
∼−→ Y ×G Xt ×G Xu ×G Y ∗ µt,u−−→
∼=
Y ×G Xtu ×G Y ∗,
where the first two isomorphisms are the canonical ones from Proposition 2.11 and
Lemma 2.8. Then µ′
t,u is an action of S on H by partial equivalences. Its transfor-
mation groupoid H ⋊ S is equivalent to G ⋊ S.
When we translate the action on Y back to X using the inverse equivalence Y ∗,
we get an action on G that is isomorphic to the original one.
1 := H 1 for t = 1, and let µ′
1 as defined above is only isomorphic to H 1 in a very
Proof. More precisely, X ′
obvious way. We should only use the above definition of X ′
t for t 6= 1 and let
X ′
t,1 be the canonical isomorphisms. We
should also put in associators for the composition of partial equivalences, which
only cause notational complications, however. Up to these technicalities, it is clear
that the maps µ′ inherit associativity from the maps µ. The action that we get by
translating µ′ back to G with Y ∗ is canonically isomorphic to the original action
because Y ∗ ×H Y ∼= G1.
1,t and µ′
It remains to prove the equivalence of the transformation groupoids G ⋊ S and
H ⋊ S. Here we use the linking groupoid L of the equivalence; its object space is
L0 = G0 ⊔ H 0, its arrow space is G1 ⊔ Y ⊔ Y ∗ ⊔ H 1, its range and source maps
are r and s on each component, and its multiplication consists of the multiplica-
tions in G and H, the G, H-bibundle structure on Y , the H, G-bibundle structure
on Y ∗, and the canonical isomorphisms Y ×G Y ∗ ∼−→ H 1 and Y ∗ ×H Y ∼−→ G1 from
Proposition 2.11. This gives a topological groupoid L. There is a canonical right
action of L on G1 ⊔ Y = r−1(G1) ⊆ L1 that provides an equivalence from L to G
when combined with the left actions of G on G1 and Y ; there is a similar canonical
equivalence H 1 ⊔ Y ∗ from L to H.
We may transport the S-action on G to L because it is equivalent to G. When we
transport this action on L further to H, we get the action described above because
the composite equivalence (G1 ⊔ Y ) ×L (H 1 ⊔ Y ∗)∗ from H to G is isomorphic to Y .
The action on L is given by bibundles
(G1 ⊔ Y )∗ ×G Xt ×G (G1 ⊔ Y )
∼= Xt ⊔ (Xt ×G Y ) ⊔ (Y ∗ ×G Xt) ⊔ (Y ∗ ×G Xt ×G Y ),
where we cancelled factors of G1 using Lemma 2.8. When we restrict the trans-
formation groupoid L ⋊ S to G0 ⊆ L0 or to H 0 ⊆ L0, then we only pick the
components Xt and Y ∗ ×G Xt ×G Y in the above decomposition, so we get the
transformation groupoids G ⋊ S and H ⋊ S, respectively. Routine computations
show that the other two parts r−1(G0) ∩ s−1(H 0) and r−1(H 0) ∩ s−1(G0) of L ⋊ S
give an equivalence from H ⋊ S to G ⋊ S, such that L ⋊ S is the resulting linking
groupoid.
It can be shown with less routine computations that the embedding G ⋊ S ֒→
L ⋊ S is fully faithful and essentially surjective. We checked "fully faithful" above.
Being "essentially surjective" means that the map G0 ×⊂,L0,r L1 → L0, (x, l) 7→ s(l),
is open and surjective. It is open because r : L1 → L0 is open and G0 ⊂ L0 is open,
and surjective because already G0 ×L0 Y ⊂ G0 ×L0 L1 surjects onto H 0. Since both
G ⋊ S ֒→ L ⋊ S and H ⋊ S ֒→ L ⋊ S are fully faithful and essentially surjective,
24
ALCIDES BUSS AND RALF MEYER
they induce equivalence bibundles by [21, Proposition 6.8], which we may compose
to an equivalence from H ⋊ S to G ⋊ S. Of course, this gives the same equivalence
as the argument above.
(cid:3)
Corollary 3.21. Let S be an inverse semigroup with unit. Let f : X → Z be
an open continuous surjection and let G(f ) be its covering groupoid, see Defini-
tion A.8. Then S-actions by partial equivalences on G(f ) are canonically equivalent
to S-actions on Z by partial homeomorphisms, such that G(f ) ⋊ S is equivalent to
Z ⋊ S.
Here "equivalent" means an equivalence of categories, where the arrows are iso-
morphisms of S-actions that fix the underlying groupoid.
Proof. G(f ) is canonically equivalent to Z viewed as a groupoid, so the assertion
follows from Theorem 3.18 and Proposition 3.20.
(cid:3)
In particular, Corollary 3.21 applies to the Čech groupoid GU of an open cov-
ering U of a locally Hausdorff space Z by Hausdorff open subsets. Thus we may
replace an S-action by partial homeomorphisms on a locally Hausdorff space Z by
an "equivalent" action by partial equivalences on a Hausdorff groupoid GU, and the
resulting transformation groupoids Z ⋊ S and GU ⋊ S are equivalent.
The quickest way to describe the resulting S-action on GU explicitly is by de-
scribing GU ⋊ S and an S-grading on it. Let X := FU∈U U and let p: X → Z be
the canonical map, which is an open surjection. The pull-back p∗(Z ⋊ S) of Z ⋊ S
along p is a groupoid with object space X, arrow space X ×p,Z,r (Z ⋊ S)1 ×s,Z,p X,
r(x1, g, x2) = x1, s(x1, g, x2) = x2, and (x1, g, x2) · (x2, h, x3) = (x1, g · h, x3) (see
[21, Example 3.13]). Let
Xt := {(x1, g, x2) ∈ X ×p,Z,r (Z ⋊ S)1 ×s,Z,p X g ∈ t}.
Proposition 3.22. The subspaces Xt ⊆ p∗(Z ⋊S)1 form an S-grading on p∗(Z ⋊S).
The resulting S-graded groupoid is the transformation groupoid for the S-action
on GU that we get by translating the S-action on Z along the equivalence to GU.
Proof. The subspaces Xt form an S-grading because the bisections t ∈ S give an
S-grading on Z ⋊ S and p is surjective. Hence they describe an S-action on GU.
The equivalence from GU to Z is given by the canonical action of GU on G0
U = X
and the projection p: X → Z. Hence Xt is exactly what we get when we translate
t ⊆ Z ⋊ S along the equivalence.
(cid:3)
For instance, let H be an étale groupoid with locally Hausdorff arrow space and
let S be some inverse semigroup of bisections with H ∼= H 0 ⋊ S; we could take
S = Bis(H). Let Z = H 1 with the action of H by left multiplication. This induces
an action of S on Z. Its transformation groupoid H 1 ⋊S is H 1 ⋊H with the obvious
S-grading by H 1 ×H0 t for t ∈ S.
The left multiplication action of H on H 1 with the bundle projection s: H 1 →
H 0 is a trivial principal bundle. In particular, the transformation groupoid H 1⋊S ∼=
H 1 ⋊ H is isomorphic to the covering groupoid of the cover s: H 1 ։ H 0. Hence
it is equivalent to the space H 0, viewed as a groupoid with only unit arrows. The
S-grading on H 1 ⋊ S does, however, not carry over to H 0.
If we replace the S-action on H 1 by an equivalent S-action on GU for a Hausdorff
open cover of H 1, then the transformation groupoid GU ⋊ S is equivalent to H 1 ⋊ S
and hence also equivalent to the space H 0. In particular, the groupoid GU ⋊ S is
basic (see Section A.2). If a groupoid is equivalent to a space, then this space has
to be its orbit space. So if H 0 is Hausdorff, then the groupoid GU ⋊ S is a free and
proper, Hausdorff groupoid by Proposition A.7.
INVERSE SEMIGROUP ACTIONS ON GROUPOIDS
25
3.5. Local centralisers. We are going to show that for many groupoids G the
bibundles Xt already determine the multiplication maps µt,u and thus the inverse
semigroup action. This happens, among others, for essentially principal topological
groupoids (meaning that the isotropy group bundle has no interior; see [32]) and
for groups with trivial centre.
Definition 3.23. A local centraliser of G is a map γ : U → G1 defined on an
open G-invariant subset U of G0 with s(γ(x)) = r(γ(x)) = x for all x ∈ U and
γ(r(g)) · g = g · γ(s(g)) for all g ∈ G. We say that G has no local centralisers if all
local centralisers are given by γ(x) = 1x for x ∈ U and some U as above.
Local centralisers defined on the same subset U form an Abelian group under
pointwise multiplication. All local centralisers form an Abelian inverse semigroup.
It is the centre of Bis(G) if G is étale.
Lemma 3.24. Let X be a partial equivalence from H to G. Then Map(X, X) is
isomorphic to the group of local centralisers of G defined on r(X), and to the group
of local centralisers of H defined on s(X).
If G has no local centralisers and X and Y are partial equivalences from G or
to G, then there is at most one bibundle map X → Y , so bibundle isomorphisms
are unique if they exist.
Proof. The two descriptions of Map(X, X) are equivalent by taking X ∗, so we only
prove one. Every bibundle map X → X is invertible by Proposition 2.9. The
canonical group homomorphisms
Map(X, X) ␣×H X ∗
−−−−−→ Map(X ×H X ∗, X ×H X ∗) ∼= Map(G1
r(X))
r(X), G1
␣×GX−−−−→ Map(X, X)
U , G1
are inverse to each other by the proof of Proposition 2.11. Thus it remains to iden-
U ) for an open G-invariant subset U of G0 with the group
tify the set Map(G1
of local centralisers defined on U . We may view G1
U to
itself associated to the identity functor on G1
U . We described all bibundle isomor-
phisms between such equivalences in Example 2.5. Specialising Example 2.5 to the
automorphisms of the identity functor gives exactly the local centralisers defined
on U . A quick computation shows that the composition of bibundle isomorphisms
corresponds to the pointwise multiplication of local centralisers.
U as the equivalence from G1
Let f1, f2 : X → Y be bibundle maps. Then both are bibundle isomorphisms
X → Y s(X), and we may form a composite bibundle isomorphism f −1
2 ◦f1 : X → X.
Since there are no local centralisers, the first part of the lemma shows that this is
the identity map, so f1 = f2. In particular, if two partial equivalences G → H or
H → G are isomorphic, then the isomorphism is unique.
(cid:3)
tial equivalences on G.
More precisely, isomorphism classes of S-actions on G by partial equivalences are
Theorem 3.25. Let G be a topological groupoid without local centralisers. An ac-
Recall that fpeq(G) denotes the inverse semigroup of isomorphism classes of par-
tion of an inverse semigroup S on G is equivalent to a homomorphism S → fpeq(G).
in canonical bijection with homomorphisms S → fpeq(G).
Proof. A homomorphism f : S → fpeq(G) gives us bibundles Xt with Xt ×G Xu
∼=
Xtu and X1 ∼= G1; we may as well assume X1 = G1. By Lemma 3.24, the isomor-
∼−→ Xtu above are unique, so there is no need to specify them.
phisms µt,u : Xt×G Xu
The conditions (A2) and (A3) hold because any two parallel bibundle isomorphisms
are equal. Thus f determines an S-action by partial equivalences. Conversely, an
26
ALCIDES BUSS AND RALF MEYER
action by partial equivalences determines such a homomorphism by taking the iso-
morphism classes of the Xt and forgetting the µt,u. Since isomorphisms of partial
equivalences are unique if they exist, this forgetful functor is actually not forget-
ting anything here, so we get a bijection between isomorphism classes of actions by
(cid:3)
partial equivalences and homomorphisms S → fpeq(G).
The results in this section are inspired by the notion of a "quasi-graphoid" used
by Debord in [10]. Debord already treats partial equivalences of groupoids as
arrows between groupoids and uses them to glue together groupoids constructed
locally. She restricts, however, to a situation where bibundle isomorphisms are
uniquely determined. Even more, she wants the range and source maps to already
determine a partial equivalence uniquely. For this, she assumes that a smooth map
γ : U → G1 defined on an open subset U of G0 has to be the unit section already
if it only satisfies s(γ(x)) = r(γ(x)) = x for all x ∈ U . This condition holds for
holonomy groupoids of foliations -- even for the mildly singular foliations that she
is considering.
3.6. Decomposing proper Lie groupoids. A manifold may be constructed by
taking a disjoint union of local charts and gluing them together along the coordi-
nate change maps, which are partial homeomorphisms, or diffeomorphisms in the
smooth case. When constructing groupoids locally, it is more likely that the coordi-
nate change maps are no longer partial isomorphisms but only partial equivalences.
Actually, it may well be that the local pieces are, to begin with, only local group-
oids and not groupoids (see [10]); this is not covered by our theory. Therefore, we
know no good examples where groupoids have been constructed by gluing together
smaller groupoids along partial equivalences.
Instead, we take a groupoid as given and analyse it using local information. The
local information should say that the groupoid locally is equivalent to one of a par-
ticularly simple form. Then the groupoid is globally equivalent to a transformation
groupoid for an inverse semigroup action by partial equivalences on a disjoint union
of groupoids having the desired simple form.
We now get more concrete and consider a proper Lie groupoid H. To formulate
stronger results, we shall work with (partial) equivalences of Lie groupoids in this
section; that is, spaces are replaced by smooth manifolds, continuous maps by
smooth maps, and open maps by submersions. This does not change the theory
significantly, see [21].
First we formulate the local linearisability of proper Lie groupoids. This was
conjectured by Weinstein [37] and proved by Zung [38]. Both authors try to de-
scribe the local structure of proper Lie groupoids up to isomorphism. Following
Trentinaglia [35], we only aim for a description up to Morita equivalence:
Theorem 3.26. Let H be a proper Lie groupoid. For every x ∈ H 0 there are
an open H-invariant neighbourhood Ux of x in H 0, a linear representation of the
stabiliser group Hx on a finite-dimensional vector space Wx, and a Lie groupoid
equivalence from the transformation groupoid Wx ⋊ Hx to HUx.
The vector space Wx is the normal bundle to the H-orbit Hx of x, with its
canonical representation of Hx.
Weinstein and Zung impose extra assumptions on H to describe HUx up to
isomorphism. The argument in [35, Section 4] shows how to deduce Theorem 3.26
quickly from [38, Theorem 2.3] without extra assumptions.
Actually, we do not need H to be proper. Since we only need local structure, it
is enough for H to be locally proper, that is, each x ∈ H 0 has an H-invariant open
neighbourhood U such that HU is proper; this allows the orbit space H 0/H to be
a locally Hausdorff but non-Hausdorff manifold.
INVERSE SEMIGROUP ACTIONS ON GROUPOIDS
27
Assume now that H is a locally proper groupoid. By Theorem 3.26, there is
a covering U of H 0 by open, H-invariant subsets and, for each U ∈ U, a Lie
groupoid equivalence XU from a transformation groupoid WU ⋊ GU for a compact
Lie group GU and a linear representation WU of GU to the restriction HU . Now let
G := GU∈U
WU ⋊ GU .
Let K be the covering groupoid of H 0 for the covering U. Since HU U∩V =
This disjoint union is a groupoid with object space F WU .
HU∩V = HV U∩V , the inverse semigroup S := Bis(K) acts on FU∈U HU : each
The disjoint union X :=FU∈U XU gives an equivalence from G toFU∈U HU , so we
element of Bis(K) acts by the identity equivalence between the appropriate restric-
tions of HU and HV , and all the multiplication maps are the canonical isomorphisms.
may transfer this S-action to G.
We make the action on G more concrete. Any bisection of K is a disjoint union
of bisections of the form
(U1, D, U2) := {(U1, x, U2) x ∈ D}
for U1, U2 ∈ U and an open subset D ⊆ U1 ∩ U2. The product (U1, D1, U2) ·
(U ′
2, D2, U3) is empty if U2 6= U ′
The partial equivalence XU1,D,U2 on G associated to (U1, D, U2) is the composite
2, and is equal to (U1, D1 ∩ D2, U3) if U2 = U ′
2.
partial equivalence
G ⊇ WU1
⋊ GU1
D X ∗
U1−−−−→ HD
XU2 D−−−−→ WU2
⋊ GU2 ⊆ G.
The composite of XU1,D1,U2 and XU ′
be. If U2 = U ′
2, then there is a canonical isomorphism of partial equivalences
2,D2,U3 is clearly empty for U2 6= U ′
2, as it should
µ(U1,D1,U2),(U2,D2,U3) : XU1,D1,U2 ×G XU2,D2,U3 → XU1,D1∩D2,U3,
using the restriction of the canonical pairing XU2 ×G X ∗
U2 → HU2 to remove the
extra two factors in the middle. This is exactly what happens if we translate the
"trivial" action of S on F HU described above to G along the equivalence F XU .
Theorem 3.27. The locally proper Lie groupoid H is equivalent to the transfor-
mation groupoid G ⋊ S for the action of S on G described above.
Proof. Since we constructed the action of S on G by translating the action on
FU∈U HU , Proposition 3.20 shows that G⋊S is equivalent toFU∈U HU ⋊S. Since S
acts "trivially" on F HU , this transformation groupoid is easy to understand: it is
the pull-back p∗(H) of H for the canonical map p: FU∈U U → H 0. Since p is a
surjective submersion, p∗(H) is equivalent to H.
(cid:3)
As a result, any locally proper Lie groupoid is equivalent to a transformation
groupoid for an inverse semigroup action on a disjoint union of linear actions of
compact groups. Such transformation groupoids need not be locally proper, how-
ever, so we do not have a characterisation of locally proper Lie groupoids. The
groupoid G ⋊ S is étale if and only if G is, if and only if the stabilisers Hx are finite.
This means that H is an orbifold (see [22]).
4. Inverse semigroup actions on C∗-algebras
We now define inverse semigroup actions on C∗-algebras by Hilbert bimodules,
in parallel to actions on groupoids by partial equivalences.
Definition 4.1. A Hilbert A, B-bimodule H is a left Hilbert A-module and a
right Hilbert B-module such that the left and right multiplications commute, and
28
ALCIDES BUSS AND RALF MEYER
hhxyiiA ·z = x·hyziB for all x, y, z ∈ H. A Hilbert A, B-bimodule map is a bimodule
map that also intertwines both inner products.
Let H be a Hilbert A, B-bimodule. Let I ⊳ A and J ⊳ B be the closed linear
spans of the elements hhxyiiA and hxyiB with x, y ∈ H, respectively. These are
closed ideals in A and B, and H is an I, J-imprimitivity bimodule by restricting
the left multiplications to I and J. Ideals in a C∗-algebra are in bijection with open
subsets of its primitive ideal space, so ideals are the right analogues of open invariant
subsets of groupoids. Hence we denote the ideals I and J above as I := r(H) and
J := s(H), and we think of Hilbert A, B-bimodules as partial Morita equivalences
from B to A.
Given an ideal K ⊳ A, we define the restriction of a Hilbert bimodule H to K
as KH := K · H ⊆ H, which is canonically isomorphic to K ⊗A H. We restrict to
ideals in B in a similar way.
The left action of A on a Hilbert bimodule is by a nondegenerate ∗-homomor-
phism A → B(H) into the adjointable operators on H. Thus a Hilbert A, B-
bimodule becomes a correspondence by forgetting the left inner product.
Lemma 4.2. A correspondence H carries a Hilbert bimodule structure if and only
if there is an ideal I ⊳ A such that the left action ϕ: A → B(H) restricts to an
isomorphism from I onto K(H). This ideal and the left inner product are uniquely
determined by the correspondence.
Proof. First let H be a Hilbert bimodule. Then H is an imprimitivity bimodule
from s(H) to r(H), so ϕr(H) is an isomorphism from r(H) onto K(H). If I ⊳ A is
another ideal with ϕ(I) = K(H), then ϕ(r(H) · I) = K(H) as well. Thus r(H) is
the minimal ideal that ϕ maps onto K(H), and the only one on which this happens
isomorphically. Thus r(H) is already determined by the underlying correspondence.
Let H′ be another Hilbert A, B-bimodule with the same underlying correspon-
dence as H and with left A-valued inner product hhxyii′
A. Then
ϕ(hhxyii′
A)z = x · hyziB = ϕ(hhxyiiA)z
for all x, y, z ∈ H. Since r(H) = r(H′) depends only on the correspondence and
the restriction of ϕ to r(H) is faithful, we get H = H′ as Hilbert bimodules.
Now let H be a correspondence and let I ⊳ A be an ideal that is mapped
isomorphically onto K(H). Transfer the usual K(H)-valued left inner product on H
through this isomorphism to one with values in A ⊇ I. This turns H into a Hilbert
A, B-bimodule.
(cid:3)
Proposition 4.3. Let H and H′ be Hilbert A, B-bimodules. If there is a Hilbert
bimodule map f : H → H′, then s(H) ⊆ s(H′) and r(H) ⊆ r(H′). Such a Hilbert
bimodule map is an isomorphism from H onto the submodule H′ · s(H) = r(H) · H′
in H′. So it is an isomorphism onto H′ if and only if s(H′) ⊆ s(H), if and
only if r(H′) ⊆ r(H), if and only if the map K(H) → K(H′) induced by f is an
isomorphism.
Proof. Since the norm on a Hilbert bimodule is generated by the inner products,
Hilbert bimodule maps are norm isometries and thus injective. Moreover,
f (H) = f (r(H) · H) = r(H) · f (H) ⊆ r(H) · H′.
Thus r(H′) ⊆ r(H) is necessary for f to be an isomorphism. Conversely, if r(H′) ⊆
r(H), then even r(H′) = r(H) because a bimodule map preserves the left inner
product. Then the map from r(H) ∼= K(H) to r(H′) ∼= K(H′) that sends ξihη to
f (ξ)ihf (η) for ξ, η ∈ H is an isomorphism K(H) ∼= r(H′) ∼= K(H′). Since K(H′) ·
H′ = H′, the linear span of elements of the form f (ξ)ihf (η)ζ′ = f (ξ) · hf (η)ζ′i
for ξ, η ∈ H, ζ′ ∈ H′ is dense in H′. Since f (H) is a right B-module, this implies
INVERSE SEMIGROUP ACTIONS ON GROUPOIDS
29
that f is surjective. Hence it is an isomorphism of Hilbert bimodules. A similar
argument for the right inner product instead of the left one shows that all the listed
conditions for f are indeed equivalent to f being an isomorphism.
If r(H) 6= r(H′), then we may restrict f to a Hilbert bimodule map H → r(H)·H′.
Since r(H) · r(H) · H′ = r(H) · H′, this is an isomorphism by the first statement. A
similar argument on the other side shows that f is an isomorphism onto H′ · s(H),
so H′ · s(H) = r(H) · H′.
(cid:3)
A Hilbert A, B-bimodule H has a dual Hilbert B, A-bimodule H∗, where we
exchange left and right structures using adjoints: b · x∗ · a := (a∗ · x · b∗)∗ for a ∈ A,
b ∈ B, x ∈ H, and hx∗y∗iA = hhyxiiA, hhx∗y∗iiB = hyxiB. We will see that this
construction has the same formal properties as the dual for partial equivalences
of groupoids. To begin with, a Hilbert bimodule map X → Y remains a Hilbert
bimodule map X ∗ → Y ∗, and (X ∗)∗ = X. Furthermore, (ξ ⊗ η)∗ 7→ η∗ ⊗ ξ∗
defines a Hilbert bimodule map σ : (X ⊗B Y )∗ → Y ∗ ⊗B X ∗ with dense range,
hence an isomorphism. Applying σ twice gives the identity map. (More precisely,
σY ∗,X ∗ ◦ σX,Y = Id(X⊗B Y )∗.)
Proposition 4.4. Let H be a Hilbert A, B-bimodule. The inner products on H
give Hilbert bimodule isomorphisms H ⊗B H∗ ∼= r(H) and H∗ ⊗A H ∼= s(H), and
the restrictions of the left and right actions give Hilbert bimodule isomorphisms
r(H) ⊗A H ∼= H ∼= H ⊗B s(H),
s(H) ⊗B H∗ ∼= H∗ ∼= H∗ ⊗A r(H).
that make the following diagrams of isomorphisms commute:
H ⊗B H∗ ⊗A H
H ⊗B s(H)
H∗ ⊗A H ⊗B H∗
H∗ ⊗A r(H)
(15)
r(H) ⊗A H
H,
s(H) ⊗B H∗
H∗.
Let D be another C∗-algebra, let K be a Hilbert A, D-bimodule and let L be a
Hilbert B, D-bimodule with r(K) ⊆ r(H) and r(L) ⊆ s(H). Then Hilbert A, D-
bimodule maps H ⊗B L → K are naturally in bijection with Hilbert B, D-bimodule
maps L → H∗ ⊗A K, and this bijection maps isomorphisms again to isomorphisms.
Similarly, Hilbert A, D-bimodule maps H ⊗B L ← K are naturally in bijection with
Hilbert B, D-bimodule maps L ← H∗ ⊗A K.
Proof. The Hilbert bimodule isomorphisms H ⊗B H∗ ∼= r(H), H∗ ⊗A H ∼= s(H),
r(H) ⊗A H ∼= H ∼= H ⊗B s(H) and s(H) ⊗B H∗ ∼= H∗ ∼= H∗ ⊗A r(H) are routine
to check using that H is full as a Hilbert r(H), s(H)-bimodule. The diagrams
in (15) are equivalent to the requirement hhxyiiA · z = x · hyziB in the definition
of a Hilbert bimodule. The claim about Hilbert bimodule maps is proved like the
analogous one about partial equivalences of groupoids in Proposition 2.12; now
we use the canonical isomorphisms just established and Proposition 4.3 instead of
Proposition 2.9.
(cid:3)
Proposition 4.5. Up to isomorphism, H∗ is the unique Hilbert B, A-bimodule K
for which there are isomorphisms
H ⊗B K ⊗A H ∼= H,
K ⊗A H ⊗B K ∼= K.
More precisely, if there are such isomorphisms then there is a unique Hilbert bimod-
ule isomorphism H∗ ∼−→ K such that the following map is the identity map:
H ∼−→ H ⊗B K ⊗A H ∼−→ H ⊗B H∗ ⊗A H ∼−→ r(H) ⊗A H ∼−→ H.
Proof. Repeat the proof of Proposition 2.12, replacing ×G by ⊗A.
(cid:3)
30
ALCIDES BUSS AND RALF MEYER
Proposition 4.6. Let H be a Hilbert A, A-bimodule and let µ: H ⊗A H → H be
a bimodule isomorphism. Then there is a unique isomorphism from H onto an
ideal I ⊳ A that intertwines µ and the multiplication map I ⊗A I ∼−→ I. We have
I = r(H) = s(H), and the multiplication µ is associative.
Proof. This is proved exactly like Proposition 2.13.
Definition 4.7. Let S be an inverse semigroup with unit and let A be a C∗-algebra.
An S-action on A by Hilbert bimodules consists of
(cid:3)
• Hilbert A, A-bimodules Ht for t ∈ S;
• bimodule isomorphisms µt,u : Ht ⊗A Hu
∼−→ Htu for t, u ∈ S;
satisfying
(AH1) H1 is the identity Hilbert A, A-bimodule A;
(AH2) µt,1 : Ht ⊗A A ∼−→ Ht and µ1,u : A ⊗A Hu
∼−→ Hu are the canonical isomor-
phisms for all t, u ∈ S;
(AH3) associativity: for all t, u, v ∈ S, the following diagram commutes:
(Ht ⊗A Hu) ⊗A Hv
ass
Ht ⊗A (Hu ⊗A Hv)
µt,u ⊗A IdHv
IdHt ⊗A µu,v
Htu ⊗A Hv
Ht ⊗A Huv
µtu,v
Htuv
µt,uv
If S has a zero element 0, we may also require H0 = {0}.
Theorem 4.8. Let S be an inverse semigroup with unit, let A be a C∗-algebra.
Then actions of S on A by Hilbert bimodules are equivalent to saturated Fell bundles
over S (as defined in [13]) with unit fibre A.
More precisely, let (Ht)t∈S and (µt,u)t,u∈S be an S-action by Hilbert bimodules
on A. Then there are unique Hilbert bimodule maps ju,t : Ht → Hu for t ≤ u
that make the following diagrams commute for all t1, t2, u1, u2 ∈ S with t1 ≤ u1,
t2 ≤ u2:
(16)
Ht1 ⊗A Ht2
µt1,t2
Ht1t2
ju1,t1 ⊗A ju2,t2
Hu1 ⊗A Hu2
µu1,u2
ju1u2,t1t2
Hu1u2
The map ju,t is a Hilbert bimodule isomorphism onto Hu · s(Ht) = r(Ht) · Hu. We
have jt,t = IdHt for all t ∈ S and jv,u ◦ ju,t = jv,t for t ≤ u ≤ v in S. And
∼−→ Ht∗, x 7→ x∗, such
there are unique Hilbert bimodule isomorphisms Jt : H∗
t
that µt,t∗,t(x, x∗, x) = x · hxxiA = hhxxiiA · x for all x ∈ Ht. These also satisfy
µt,t∗(x ⊗ x∗) = hhxxiiA, µt∗,t(x∗, x) = hxxiA and (x∗)∗ = x for all x ∈ Ht;
µt,u(x, y)∗ = µu∗,t∗(y∗, x∗) for all x ∈ Ht, y ∈ Hu, t, u ∈ S; and ju,t(x)∗ =
ju∗,t∗(x∗) for all t ≤ u in S, x ∈ Ht.
Conversely, a saturated Fell bundle (At)t∈S over S with A = A1 becomes an
S-action by Hilbert bimodules by taking Ht = At with the multiplication maps µt,u
and the A-bimodule structure induced by the Fell bundle multiplication, and the left
and right inner products hhxyiiA := x · y∗, hxyiA := x∗ · y for x, y ∈ Ht.
Proof. We construct the inclusion maps jt,u and the involutions Jt and show their
properties exactly as in the proofs of Propositions 3.7 and 3.9.
(cid:3)
With Theorem 4.8, it becomes easier to construct saturated Fell bundles over in-
verse semigroups because Definition 4.7 needs far less data and has correspondingly
fewer conditions to check.
INVERSE SEMIGROUP ACTIONS ON GROUPOIDS
31
Remark 4.9. The correspondence bicategory introduced in [6] is not suitable for
our purposes by the following observation: Let I ֒→ A ։ A/I be a split extension
of C∗-algebras. Then p: A → A/I → A is an idempotent endomorphism. It re-
mains an idempotent arrow in the correspondence bicategory. More generally, if A
is Morita equivalent to an ideal in a C∗-algebra B, then we can translate p to a
correspondence H from B to itself that is idempotent in the sense that H⊗B H ∼= H
with an associative isomorphism. Thus there are more idempotent endomorphisms
in the correspondence bicategory than usual for inverse semigroup actions. Fur-
thermore, the idempotent arrows no longer commute up to isomorphism; thus a
very basic assumption for inverse semigroups fails in this case. This is why we only
allowed Hilbert bimodules above.
Proposition 4.10. There is a bicategory with C∗-algebras as objects, Hilbert bimod-
ules as arrows, Hilbert bimodule isomorphisms as 2-arrows, and ⊗B as composition
of arrows.
Proof. The correspondence bicategory is constructed already in [6]. Lemma 4.2
allows to identify Hilbert bimodules with a subset of correspondences. It is well-
known that composites of Hilbert bimodules are again Hilbert bimodules. Hence
the Hilbert bimodules form a sub-bicategory in the opposite of the correspondence
bicategory.
(cid:3)
5. Fell bundles from actions of inverse semigroups
All groupoids in this section are assumed to be locally quasi-compact, locally
Hausdorff and with (locally compact) Hausdorff object space and a Haar system,
so that they have groupoid C∗-algebras. Let G be such a groupoid and let S be
a unital inverse semigroup acting on G by partial equivalences. We want to turn
this into an action of S on C∗(G) by Hilbert bimodules; equivalently, we want a
Fell bundle over S with unit fibre C∗(G). There are two closely related ways to
construct this. We are going to explain one approach in detail and only sketch the
other one briefly in Section 5.3.
We give details for the construction of the Fell bundle using the transforma-
tion groupoid L = G ⋊ S because this also suggests how to describe the section
C∗-algebra of the resulting Fell bundle. The transformation groupoid L comes with
an S-grading (Lt)t∈S. Roughly speaking, our Fell bundle over S will involve the
subspaces of C∗(L) of elements supported on the open subsets Lt. Since G1 = L1,
the unit fibre of the Fell bundle will be C∗(G). This also suggests that the section
C∗-algebra of the Fell bundle over S is C∗(L). This is indeed the case, but the
technical details need some care.
First, we need a Haar system on L. We show in Proposition 5.1 that the Haar
system on G extends uniquely to a Haar system on L. Secondly, it is non-trivial
that C∗(G) is contained in C∗(L): this means that the maximal C∗-norm that
defines C∗(G) extends to a C∗-norm on C∗(L). A related issue is to show that
an element of C∗(L) supported in G actually belongs to C∗(G). These problems
become clearer if we construct a pre-Fell bundle using the dense ∗-algebra that
defines C∗(L) and then complete it.
In the non-Hausdorff case, continuous functions with compact support are re-
placed by finite linear combinations of certain functions that are not continuous.
The identification of C∗(L) with the section C∗-algebra of the Fell bundle requires
a technical result about these functions. We prove it in Appendix B in the more
general setting of sections of upper semicontinuous Banach bundles because this is
not more difficult and allows us to generalise our main results to Fell bundles over
groupoids.
32
ALCIDES BUSS AND RALF MEYER
We write S(X) for the space of linear combinations of compactly supported
functions on Hausdorff open subsets of a locally Hausdorff, locally quasi-compact
space X. This is the space of compactly supported continuous functions on X if
and only if X is Hausdorff, and it is often denoted by Cc(X). We find this notation
misleading, however, because its elements are not continuous functions.
5.1. A Haar system on the transformation groupoid. Before we enter the
construction of Haar systems, we mention an important trivial case: if G is étale,
then so is L. Therefore, L certainly has a canonical Haar system if G is étale. This
already covers many examples, and the reader only interested in étale groupoids
may skip the construction of the Haar system on L.
We define Haar systems as in [31, Section 1]. Thus our Haar system (λx
on G is left invariant, so supp λx
for all g ∈ G. The continuity requirement for (λx
G)x∈G0
G = λr(g)
G)x∈G0 is that the function λG(f )
G(g) is continuous on G0 for all f ∈ S(G).
By the definition of S(G) (see Definition B.1), it suffices to check continuity if f is
a continuous function with compact support on a Hausdorff open subset U of G.
on G0 defined by λG(f )(x) :=RG f (g) dλx
G = Gx = {g ∈ G1 r(g) = x} and g∗λs(g)
G
Proposition 5.1. The Haar system on G extends uniquely to a Haar system on
the transformation groupoid L.
Proof. Fix x ∈ G0 = L0. We are going to describe the measure λx
Haar system. Since L = St∈S Lt is an open cover, the measure λx
L on Lx in the
L is determined
by its restrictions to Lt for all t ∈ S. If x /∈ r(Lt), then there is nothing to do, so
t := Lt ∩ Lx
consider t ∈ S with x ∈ r(Lt), and fix g ∈ Lt with r(g) = x. If A ⊆ Lx
is measurable, then A = g · (g−1 · A) with g−1 · A ⊆ L−1
· Lt = L1 = G. Since
we want (λx
L(A) =
λs(g)
G (g−1 · A) if g ∈ Lt satisfies r(g) = x and A ⊆ Lx
t is measurable. Hence there is
at most one Haar measure on L extending the given Haar measure on G.
L) to be left invariant and to extend (λx
G), we must have λx
t
· g2 ∈ L−1
1
G) with respect to G implies that λs(g)
If g1, g2 ∈ Lt satisfy r(g1) = r(g2) = x, then g−1
t Lt = L1 = G;
the left invariance of (λx
G (g−1 · A) does not
depend on the choice of g. If ∅ 6= A ⊆ Lx
u, then we may pick the same element
g ∈ A to define the measure of A as a subset of Lx
u. Thus the definitions
t for t ∈ S are compatible. Thus there is a unique measure λx
of λx
L
on Lx with λx
t is measurable and g ∈ Lt
satisfies r(g) = x. If l ∈ L has s(l) = x, then l∗(λx) is a measure on Lr(l) with the
same properties that characterise λr(l) uniquely; so we get the left invariance of our
family of measures: l∗(λs(l)) = λr(l) for all l ∈ L.
G (g−1 · A) whenever A ⊆ Lx
L on the sets Lx
L(A) = λs(g)
t and of Lx
t ∩ Lx
Checking continuity by hand is unpleasant, so we use a different description of
the same Haar system for this purpose. Recall that Lt is an equivalence between
restrictions of G to open invariant subsets of G0. The proof that equivalent group-
oids have Morita -- Rieffel equivalent groupoid C∗-algebras uses a family of measures
on the equivalence bibundle in order to define the right inner product; this measure
on Lt is exactly the one described above (see the proof of [31, Corollaire 5.4]), and
its continuity is known, even in the non-Hausdorff case. Thus our family of mea-
sures restricts to a continuous family on each Lt. Since the map L S(Lt) → S(L)
in Proposition B.2 is surjective, the family of measures (λx
L) is continuous.
(cid:3)
5.2. Construction of the Fell bundle. We know now that L has a Haar system.
So we get a ∗-algebra structure on S(L) as in [25,31]. Since the Haar measure on L
extends the one on G, the map S(G) → S(L) induced by the open embedding
G → L is a ∗-algebra isomorphism onto its image. The groupoid C∗-algebras of L
and G are the completions of S(L) and S(G) for suitable C∗-norms.
INVERSE SEMIGROUP ACTIONS ON GROUPOIDS
33
Lemma 5.2. The involution on S(L) maps S(Lt) onto S(L−1
convolution product maps S(Lt) × S(Lu) to S(Ltu).
t ) = S(Lt∗). The
Proof. The claim for the involution is trivial. The claim for the convolution product
follows, of course, from Lt · Lu ⊆ Ltu, but requires some care in the non-Hausdorff
case because the convolution product is not defined directly, see the proof of [25,
Proposition 4.4]. If f1 ∈ S(U ), f2 ∈ S(V ) for Hausdorff open subsets U ⊆ Lt and
V ⊆ Lu, and if U · V is also Hausdorff, then we directly get f1 ∗ f2 ∈ S(U · V )
with U · V ⊆ Ltu. If U · V is non-Hausdorff, a partition of unity is used to write
f1 and f2 as finite sums of functions on smaller Hausdorff open subsets U ′ ⊆ U ,
V ′ ⊆ V for which U ′ · V ′ is Hausdorff. Since U ′ · V ′ ⊆ U · V ⊆ Ltu, we get
S(Lt) ∗ S(Lu) ⊆ S(Ltu) as desired.
(cid:3)
1 ∗ f2 ∈ S(G) and f1 ∗ f ∗
Lemma 5.2 gives S(G) ∗ S(Lt) ⊆ S(Lt) and S(Lt) ∗ S(G) ⊆ S(Lt), so S(Lt) is
a S(G)-bimodule; it also implies f ∗
2 ∈ S(G) for all f1, f2 ∈
S(Lt), which gives S(G)-valued left and right inner products on S(Lt). We also
have f1∗f2 ∈ S(Ltu) for f1 ∈ S(Lt) and f2 ∈ S(Lu), and these multiplication maps
are associative and "isometric" with respect to the S(G)-valued inner products. We
put "isometric" in quotation marks because we have not yet talked about norms.
Lemma 5.3. f ∗ ∗ f ∈ S(G) is positive in C∗(G) for each t ∈ S, f ∈ S(Lt), and
the closed linear span of f ∗
1 ∗ f2 for f1, f2 ∈ S(Lt) is dense in C∗(Gr(Lt)).
Proof. We have already used in the proof of Proposition 5.1 that the Haar measure
on L restricts to the usual family of measures on the partial equivalence space Lt.
In that context, the positivity of such inner products is already proved in [25, 31]
in order to show that S(Lt) may be completed to a Hilbert C∗(G)-bimodule. The
proof that an equivalence induces a Morita -- Rieffel equivalence also shows that the
inner product defined above is full, that is, the closed linear span of f ∗
1 ∗ f2 for
f1, f2 ∈ S(Lt) is dense in C∗(Gr(Lt)).
(cid:3)
Hence we may complete S(Lt) to a Hilbert bimodule C∗(Lt) over C∗(G). The
densely defined convolution map S(Lt) × S(Lu) → S(Ltu) extends to a Hilbert
bimodule map
µt,u : C∗(Lt) ⊗C∗(G) C∗(Lu) → C∗(Ltu)
because it is isometric for the S(G)-valued inner products. Since C∗(Lt) is full as
a Hilbert bimodule over C∗(Gr(Lt)) and C∗(Gs(Lt)), it follows that the maps µt,u
above are surjective.
The associativity of the multiplication on the dense subspaces S(Lt) extends
to C∗(Lt). Thus we have constructed an action of S by Hilbert bimodules on C∗(G).
By Theorem 4.8, this is equivalent to a saturated Fell bundle C∗(Lt)t∈S over S.
Theorem 5.4. The section C∗-algebra C∗(S, C∗(Lt)t∈S) is naturally isomorphic
to the groupoid C∗-algebra C∗(L).
This theorem looks almost trivial from our construction; but the proof requires
a technical result about S(L) to be proved in Appendix B. Before we turn to that,
we first add coefficients in a Fell bundle over L.
The above construction still works in almost literally the same way if we replace
S(Lt) by S(Lt, B) everywhere, where B is a Fell bundle over the groupoid L.
Unfortunately, we could not find a reference for the generalisation of Lemma 5.3 to
this context. The references on groupoid crossed products we could find consider
either Fell bundles over Hausdorff groupoids (such as [24]) or a more restrictive class
of actions for non-Hausdorff groupoids (such as [25,31]), but not both. In particular,
the positivity of the inner product on S(Lt, B) for a partial equivalence Lt is
only proved in some cases: for arbitrary upper semicontinuous Fell bundles over
34
ALCIDES BUSS AND RALF MEYER
Hausdorff groupoids in [24]; for Green twisted actions of non-Hausdorff groupoids
on continuous fields of C∗-algebras over G0 in [31]; and for untwisted actions by
automorphisms of non-Hausdorff groupoids on C0(G0)-algebras in [25]. This is
probably only a technical issue that will be resolved eventually, but not in this
paper. So we add an assumption about it in our next theorem.
Theorem 5.5. Let B be a Fell bundle over L. Assume that f ∗ ∗ f ∈ S(G, B) is
positive in C∗(G, B) for all f ∈ S(Lt, B), t ∈ S, and that the linear span of these
inner products is dense in C∗(Gs(Lt), B). Then there is a Fell bundle C∗(Lt, B)t∈S
over S that has the section C∗-algebra of the restriction C∗(G, BG) as unit fibre.
The section C∗-algebra C∗(S, C∗(Lt, B)t∈S) is naturally isomorphic to the section
C∗-algebra of the groupoid Fell bundle C∗(L, B).
Theorem 5.4 is a special case of Theorem 5.5 for the constant Fell bundle C.
It remains to prove Theorem 5.5. This will be done in Appendix B.1, after some
preliminary results about Banach bundles in Appendix B.
Corollary 5.6. Let L be an étale topological groupoid with Hausdorff locally com-
pact object space and with a Haar system. Let S be a wide inverse subsemigroup
= t1 ∩ t2 for all t1, t2 ∈ S. Then
of Bis(L), that is, St∈S t = L and St∈S,t⊆t1∩t2
the groupoid C∗-algebra of L is isomorphic to the crossed product C0(L0) ⋊ S.
More generally, if B is a Fell bundle over L, then the section C∗-algebra C∗(L, B)
is isomorphic to the section C∗-algebra of the associated Fell bundle over S.
Proof. The assumptions on S ensure that L is an S-graded groupoid by Lt := t
with unit fibre G = L0. So Theorem 5.4 gives the first assertion, and Theorem 5.5
gives the second one. In this case, positivity is not an issue because we are dealing
with a space G, so positivity in C∗(G, B) is equivalent to pointwise positivity in all
x ∈ L0 = G0. The value (f ∗ ∗ f )(x) for f ∈ S(Lt, B) is either zero or f (l)∗f (l) for
the unique l ∈ Lt with s(l) = x. This is assumed to be positive in the definition of
a Fell bundle over a groupoid.
(cid:3)
The isomorphism C∗(L) ∼= C0(L0) ⋊ S is already proved in [12, Theorem 9.8]
(if L0 is second countable and S is countable). The more general result for (sepa-
rable) Fell bundles over (second countable) étale groupoids is proved in [4, Theo-
rem 2.13].
Another special case worth mentioning are group extensions. Let G H ։ S
be an extension of locally compact groups with discrete S. This gives an action
of S, viewed as an inverse semigroup, on G by Theorem 3.15. We get a Fell bundle
over S with unit fibre C∗(G) and section C∗-algebra C∗(H). More generally, we get
a similar result for a Fell bundle over H (compare with [5, Example 3.9]). Our Fell
bundle also comes from a Green twisted action of (H, G), and in this formulation,
our theorem is well-known in this case (see [7, 14]).
Corollary 5.7. In the situation of Theorem 5.5, the canonical map from C∗(G, B)
to C∗(L, B) is injective.
Proof. The unit fibre of the Fell bundle in Theorem 5.5 is C∗(G, B) and the section
C∗-algebra is C∗(L, B). The unit fibre always embeds into the section C∗-algebras
of a Fell bundle over an inverse semigroup, see [13, Corollary 8.10].
(cid:3)
Next, we note a useful variant of Theorem 5.4 for group-valued cocycles.
Let L be a locally quasi-compact, locally Hausdorff groupoid, let S be a group,
and let c: L → S be a 1-cocycle. Let Lt := c−1(t) ⊆ L for t ∈ S, and let G =
L1 = c−1(1). Since we do not assume anything about c, this need not be an
S-grading (compare Theorem 3.15). Nevertheless, we may complete S(Lt) to a
INVERSE SEMIGROUP ACTIONS ON GROUPOIDS
35
Hilbert bimodule over C∗(G) and thus get a Fell bundle over S. The difference to
the situation above is that this Fell bundle need not be saturated any more.
Theorem 5.8. The section C∗-algebra of the Fell bundle over S with unit fi-
bre C∗(G) just described is isomorphic to C∗(L). Hence the canonical map C∗(G) →
C∗(L) is faithful.
Proof. The proofs of Theorem 5.4 and Corollary 5.7 still work for non-saturated
Fell bundles (even over inverse semigroups). Alternatively, we may replace our non-
saturated Fell bundle over G by a saturated Fell bundle over an inverse semigroup
associated to G, just as for partial actions (see [11]). This does not change the
section C∗-algebra, and afterwards Theorem 5.4 applies literally.
(cid:3)
5.3. Another construction of the Fell bundle. The construction of the Fell
bundle over S in Section 5.2 used the transformation groupoid. Now we construct
this Fell bundle using the abstract functorial properties of actions on groupoids
and their corresponding actions on C∗-algebras. Actually, some aspects of this
have been used to prove Lemma 5.3 above.
It is well-known that two equivalent groupoids have Morita -- Rieffel equivalent
C∗-algebras (see [23]), even in the non-Hausdorff case (see [31]). The proof is
constructive: given an equivalence X from H to G, the space S(X) is completed
to a C∗(G)-C∗(H)-imprimitivity bimodule, using certain natural formulas for a
S(G)-S(H)-bimodule structure and S(G)- and S(H)-valued inner products. An
important ingredient here is that the Haar measures on G and H give canonical
families of measures on the fibres of the range and source maps of X, which may
be used to integrate functions on X.
Even if X is only a partial equivalence, the same formulas still work and give a
Hilbert bimodule C∗(X) from C∗(H) to C∗(G) by completing S(X). If f : X →
X ′ is an isomorphism between two partial equivalences, then f∗ : S(X) → S(X ′)
defined by f∗(h) = h ◦ f −1 is an isomorphism that preserves all structure, so it
extends to an isomorphism C∗(X) ∼−→ C∗(X ′).
Theorem 5.9. The maps G 7→ C∗(G) from groupoids to C∗-algebras, X 7→ C∗(X)
from partial equivalences to Hilbert bimodules, and f 7→ f∗ from bibundle isomor-
phisms to Hilbert bimodule isomorphisms are part of a functor from the bicategory of
partial groupoid equivalences to the bicategory of C∗-algebras and Hilbert bimodules.
Proof. The above map is strictly compatible with unit arrows: the unit arrow G1
on G is sent to C∗(G1) = C∗(G), and the unit transformations in both bicategories
are also preserved. To complete the above data to a functor of bicategories, it
remains to give natural isomorphisms C∗(X) ⊗C∗(H) C∗(Y ) ∼= C∗(X ×H Y ) and
check that they satisfy the expected associativity condition for three composable
partial equivalences. They are constructed by writing down the "convolution map"
S(X) ⊙ S(Y ) → S(X ×H Y ) given by the formula
(17)
ξ(x · h)η(h−1 · y) dλu(h),
(ξ · η)(x, y) :=ZH1
for all ξ ∈ S(X), η ∈ S(Y ) and (x, y) ∈ X ×H Y , where u = s(x) = r(y). It is
routine to check that the map (17) has dense range and is a bimodule map and an
isometry for both inner products; thus it extends to an isomorphism between the
completions: C∗(X) ⊗C∗(H) C∗(Y ) ∼= C∗(X ×H Y ).
One way to construct the convolution maps and check their properties is like our
construction above using the transformation groupoid: build an appropriate linking
groupoid containing all the data. For two composable equivalences Y and X from K
to H and from H to H, this linking groupoid has object space G0⊔H 0⊔K 0; its arrow
36
ALCIDES BUSS AND RALF MEYER
space is a disjoint union of G1, H 1, K 1, X, Y , X ∗, Y ∗, X ×H Y , and Y ∗ ×H X ∗,
the source and range maps are the obvious ones, and the multiplication is defined
using the left and right actions of G, H and K and canonical maps. This is indeed
a topological groupoid, and it inherits a canonical Haar system if G, H and K have
Haar systems. The convolution map is the restriction of the convolution in this
larger groupoid to X ×H Y . Given three composable partial equivalences, there is
a similar linking groupoid combining all the relevant data, and the associativity of
its convolution product on X ×H Y ×K Z gives the associativity coherence of the
isomorphisms C∗(X) ⊗C∗(H) C∗(Y ) ∼= C∗(X ×H Y ).
(cid:3)
Remark 5.10. The above theorem is extended in the thesis of Rohit Holkar [15],
where a similar functor from a bicategory of groupoid correspondences to the bi-
category of C∗-correspondences is constructed. This construction is more difficult
because the family of measures needed to write down the right inner product is
no longer canonical and becomes part of the data. Hence the behaviour of the
measures under composition has to be studied as well.
An inverse semigroup action by partial equivalences may be defined as a functor
(of bicategories) from the inverse semigroup to the bicategory of groupoids and
partial equivalences. Composing it with the functor in the theorem gives a functor
from the inverse semigroup to the bicategory of Hilbert bimodules, which is the
same thing as an action by Hilbert bimodules. This is the same as a saturated Fell
bundle over the inverse semigroup by Theorem 4.8. This is the second construction
of the Fell bundle over S.
It gives an isomorphic Fell bundle because the Haar
measure on L used above is the same as the combination of the measure families
on the partial equivalences Lt that are used to define the convolution maps in
Theorem 5.9.
More concretely, an action (Xt, µt,u) of S on G yields the action on C∗(G) given
by the Hilbert bimodules C∗(Xt) with the multiplication maps
C∗(Xt) ⊗A C∗(Xu) ∼−→ C∗(Xt ×G Xu) C∗(µt,u)
−−−−−→
∼=
C∗(Xtu),
which involve the convolution isomorphisms C∗(Xt) ⊗A C∗(Xu) ∼−→ C∗(Xt ×G Xu).
This is associative by the associativity coherence of these convolution isomorphisms.
6. Actions of inverse semigroups and groupoids
Let H be an étale groupoid with locally compact Hausdorff object space. So far,
we have constructed actions of the inverse semigroup Bis(H) on certain C∗-algebras.
Instead, we would like to construct actions of H itself. In this section, we are going
to see that both kinds of actions are very closely related. Here an action of Bis(H)
is as above: an action by Hilbert bimodules or, equivalently, a saturated Fell bundle
over Bis(H). The corresponding "actions" of H are saturated Fell bundles over H.
First we explain how to turn a Fell bundle over H into one over Bis(H). So let
B = (Bh)h∈H be a Fell bundle over H (see [4, 17]). Let A := C0(H 0, B) be the
C∗-algebra of C0-sections of B over H 0; this is a C0(H 0)-C∗-algebra by construction.
If t ∈ Bis(H), then the Fell bundle operations turn Ht := C0(t, B) into a Hilbert
C0(r(t), B)-C0(s(t), B)-bimodule. The multiplication in the Fell bundle induces
multiplication maps µt,u : Ht ⊗A Hu → Htu. This gives an action of Bis(H) on A
by Hilbert bimodules.
Not every action of Bis(H) by Hilbert bimodules is of this form. The obstruction
lies in how idempotents in Bis(H) act. Idempotents in Bis(H) are the same as open
subsets of H 0. We identify the idempotent semilattice E(Bis(H)) with the complete
lattice O(H 0) of open subsets of H 0. So the action of idempotents in Bis(H)
becomes a map from O(H 0) to the complete lattice I(A) of ideals in A.
INVERSE SEMIGROUP ACTIONS ON GROUPOIDS
37
Theorem 6.1. An action (Ht, µt,u)t∈Bis(H) of the inverse semigroup Bis(H) on
a C∗-algebra A by Hilbert A-bimodules comes from a Fell bundle over H if and
only if the map from E(Bis(H)) ∼= O(H 0) to I(A) commutes with suprema. This
Fell bundle over H is unique up to isomorphism, and the Fell bundles over Bis(H)
and H have the same section C∗-algebras.
Proof. A map O(H 0) → I(A) comes from a continuous map Prim(A) → H 0 if and
only if it commutes with finite infima and arbitrary suprema by [20, Lemma 2.25];
here we need H 0 to be a sober space, a very mild condition that certainly allows
all locally Hausdorff spaces. Compatibility with finite infima says that it is a mor-
phism of semilattices, which we assume anyway; compatibility with suprema is an
extra condition. A continuous map Prim(A) → H 0 is equivalent to an isomor-
phism between A and the C∗-algebra of C0-sections of an upper semicontinuous
field (Ax)x∈H0 of C∗-algebras over H 0 (see [26]). Thus the criterion in the theorem
is necessary and sufficient for A to come from such an upper semicontinuous field.
This gives a Fell bundle over H 0 ⊆ H. It remains to extend this to all of H.
Let t ∈ Bis(H). Then Ht is a Hilbert A-bimodule. For h ∈ t ⊆ H 1, we define
Hh,t := Ht ⊗A As(h); this is a Hilbert As(h)-module. If ξ ∈ Ht, then kξk2 = khξ, ξik,
and for hξ, ξi ∈ A, the norm is the supremum of the norms of its images in Ax for
all x ∈ H 0. Therefore, the canonical map from Ht to Qh∈t Hh,t is isometric. Thus
we view Ht as a space of sections of the bundle of Banach spaces Hh,t over t. This
is an upper semicontinuous bundle on t because (Ax)x∈H0 is and the norm on Ht
is given by kξk2 = khξ, ξik with hξ, ξi ∈ A.
If t, u ∈ Bis(H) and h ∈ t ∩ u, then both Hh,t and Hh,u are candidates for
the fibre Hh of our Fell bundle at h. These are isomorphic through the canonical
isomorphisms jt,t∩u : Ht∩u → Hts(t∩u) and ju,t∩u : Ht∩u → Hus(t∩u) from Theo-
rem 4.8.
For each h ∈ H 1, choose some th ∈ Bis(H) with h ∈ th and define Hh := Hh,th.
If t ∈ Bis(H), then there are canonical isomorphisms Hh ∼= Hh,t for all h ∈ t. We
use them to transport the topology on the bundle (Hh,t)h∈t to the bundle (Hh)h∈t.
These topologies are compatible on t ∩ u for all t, u ∈ Bis(H). Since the subsets t ∈
Bis(H) form an open cover of H 1, there is a topology on the whole bundle (Hh)h∈H1
that coincides with the topology on (Hh)h∈t described above for each t ∈ Bis(H).
In particular, the space of C0-sections of (Hh)h∈H1 on t coincides naturally with Ht.
Let A(U ) for U ∈ O(G) be the ideal of C0-sections of (Ax) vanishing outside U .
Then A(U ) = HU if we view U ∈ E(Bis(G)). We have
(18)
Ht ⊗A A(U ) = Ht·U = Ht(U)·t = A(t(U )) ⊗A Ht
for all t ∈ Bis(H), U ∈ O(H 0) with U ⊆ s(t). Here we view each t ∈ Bis(H) as
a partial homeomorphism s(t) → r(t) and write t(U ) for the image of U under
this map. This is exactly how Bis(H) acts on H 0. Equation (18) implies that
∼= Ar(h) ⊗A Ht. Thus Hh is a Hilbert Ar(h)-As(h)-bimodule. The isomorphism
Hh,t
Ht ⊗A Hu → Htu is A-linear and hence C0(H 0)-linear. Thus it restricts to an
isomorphism on the fibres, Hg,t ⊗A Hh,u → Hgh,tu for all g ∈ t, h ∈ u with
s(g) = r(h). The compatibility of the multiplication with the inclusion maps from
Theorem 4.8 shows that these maps on the fibres do not depend on the choice of t
and u with h ∈ t and h ∈ u. Thus we get well-defined isomorphisms Hg ⊗As(g) Hh →
Hgh for all g, h ∈ H 1 with s(g) = r(h). Since they can be put together to maps
Ht ⊗A Hu → Htu for all t, u ∈ Bis(H) and since Bis(H) covers H 1, they are locally
∼= Ht∗ must come
continuous, hence continuous. Similarly, the isomorphisms H∗
t
from well-defined, continuous maps H∗
h → Hh−1 for h ∈ H 1 by restricting them
to fibres. The remaining algebraic conditions needed for a Fell bundle over the
groupoid H 1 all follow easily because (Ht, µt,u) gives a Fell bundle over Bis(H).
38
ALCIDES BUSS AND RALF MEYER
If we turn the Fell bundle over H constructed above into a Fell bundle over Bis(H)
again, we clearly get back the original Fell bundle over Bis(H) because Ht is the
space of C0-sections of (Hh)h∈t. Conversely, if we start with a Fell bundle over H,
turn it into a Fell bundle over Bis(H), and then use the above construction to go
back, we get an isomorphic Fell bundle over H. Hence we get a bijection between
isomorphism classes of the two types of Fell bundles. Theorem 5.5 shows that the
passage from Fell bundles over H to Fell bundles over Bis(H) does not change the
section C∗-algebras.
(cid:3)
We assumed G0 to be Hausdorff and locally compact so far because Fell bundles
over groupoids have not yet been defined in greater generality. We suggest to use
the necessary and sufficient criterion in Theorem 6.1 as a definition:
Definition 6.2. Let G be an étale topological groupoid for which G0 (and hence G1)
is sober. An action of G on a C∗-algebra A is an action of Bis(G) by Hilbert bimod-
ules for which the resulting map O(G0) → I(A) commutes with arbitrary suprema.
Sobriety of G0 is needed to turn a map O(G0) → I(A) that commutes with
suprema into a continuous map Prim(A) → G0 (see [20, Lemma 2.25]).
Let G be a sober space G0 viewed as a groupoid. Then an action of G is the
same as a continuous map Prim(A) → G0. In the notation of [20], this turns A into
a C∗-algebra over G0. It is unclear what the "fibres" of such a C∗-algebra over G0
should be if G0 is badly non-Hausdorff. Therefore, it is not clear how to describe
actions of étale sober groupoids in the sense of Definition 6.2 as Fell bundles over G.
If G0 is locally Hausdorff and locally quasi-compact, then Definition 6.2 seems to
work quite well; we plan to discuss this in greater detail elsewhere.
The criterion in Theorem 6.1 also suggests how to define actions of étale group-
oids on other groupoids:
Definition 6.3. Let G be an étale topological groupoid for which G0 (and hence G1)
is sober, and let H be an arbitrary topological groupoid. An action of G on H is an
action of Bis(G) on H by partial equivalences for which the map O(G0) → O(H 0/H)
that describes the action of E(Bis(G)) commutes with arbitrary suprema.
The extra assumption in Definition 6.3 and [20, Lemma 2.25] ensure that the
map O(G0) → O(H 0/H) for an action of G on H comes from a continuous map
H 0/H → G0 or, equivalently, an H-invariant continuous map H 0 → G0.
Proposition 6.4. Let H be a locally quasi-compact, locally Hausdorff groupoid
with Hausdorff object space and with a Haar system. An action of G on H induces
an action of G on C∗(H) as well.
Proof. In Section 5.2, we turn an action of Bis(G) on H into an action of Bis(G)
on C∗(H). For any open H-invariant subset U of H 0, the closure of S(HU ) in C∗(H)
is an ideal C∗(HU ) in C∗(H). The map O(H 0/H) → I(C∗(H)), U 7→ C∗(HU ), com-
mutes with suprema. Hence Theorem 6.1 applies to the action of Bis(G) on C∗(H)
if the action of Bis(G) satisfies the condition in Definition 6.3.
(cid:3)
6.1. The motivating example. Now we consider our motivating example: an
action of a locally Hausdorff, locally quasi-compact, étale groupoid H on a locally
Hausdorff, locally quasi-compact space Z. Let U be a Hausdorff open covering of Z
and let GU be the associated covering groupoid, which is étale, locally compact
and Hausdorff. Its C∗-algebra C∗(GU) is our noncommutative model for the non-
Hausdorff space Z. We want to construct an "action" of H on it that models the
given action of H on Z.
INVERSE SEMIGROUP ACTIONS ON GROUPOIDS
39
To construct it, we use the inverse semigroup S := Bis(H) of bisections of H.
First we turn the action of H on Z into an action of S on Z by partial home-
omorphisms in the usual way: a bisection t ∈ S acts by the homeomorphism
r−1(s(t)) → r−1(r(t)), z 7→ gr(z)·z, where gx is the unique arrow in t with s(gx) = x.
We have seen in Corollary 3.21 that the S-action on Z induces an S-action
on GU by partial equivalences. The transformation groupoid GU ⋊ S for this action
is Hausdorff, étale and locally compact. It is equivalent to Z ⋊ S by Corollary 3.21.
Let p: X := FU∈U U → Z be the canonical map. Then GU = p∗(Z). An
idempotent U ∈ O(H 0) in Bis(H) acts on GU by the identity map on the open
invariant subgroupoid GU(r◦p)−1(U). That is, O(H 0) acts on GU through the map
U/GU), U 7→ (r ◦ p)−1(U ); this commutes with suprema and infima.
O(H 0) → O(G0
Thus our action of Bis(H) on GU is also an action of H in the sense of Definition 6.3.
We may identify Z ⋊ S ∼= Z ⋊ H using the obvious S-grading on Z ⋊ H and
Theorem 3.14, so GU ⋊ S is equivalent to Z ⋊ H.
The S-action on GU induces a Fell bundle over S with unit fibre C∗(GU), which
we view as an action of S on C∗(GU). Theorem 5.4 gives an isomorphism between
its section C∗-algebra C∗(GU) ⋊ S and the groupoid C∗-algebra C∗(GU ⋊ S). We
may turn our Fell bundle over Bis(H) into a Fell bundle over the groupoid H by
Proposition 6.4.
Theorem 6.1 also says that the section C∗-algebra of the Fell bundle over H is
isomorphic to C∗(GU) ⋊S ∼= C∗(GU ⋊S). The restriction to the unit fibre is C∗(GU)
by construction. We are going to describe this Fell bundle over H.
We have GU ⋊ S ∼= p∗(Z ⋊ H), that is, the object space of GU ⋊ S is X and the
arrow space is homeomorphic to the space of triples (x1, h, x2), x1, x2 ∈ X, h ∈ H 1
with r(p(x1)) = r(h) and r(p(x2)) = s(h) in H 0. Here (x1, h, x2) is an arrow from x2
to x1, and the multiplication is (x1, h1, x2)·(x2, h2, x3) = (x1, h1h2, x3). For h ∈ H 1,
let Kh be the subspace of triples (x1, h, x2) for x1, x2 ∈ X, r(p(x1)) = r(h) and
r(p(x2)) = s(h). Since p and H are étale, this is a discrete set. The fibre at h of
our Fell bundle over H is the completion of the space Cc(Kh) of finitely supported
functions on Kh to a Hilbert bimodule over C∗(K1r(h) ) and C∗(K1s(h)).
Proposition 6.5. Let Z be a basic action of H with Hausdorff quotient space H\Z,
for instance, Z = H 1 with the action by left or right multiplication and quotient
space H 0. Then the groupoid GU ⋊ S is equivalent to H\Z and C∗(GU) ⋊ S is
Morita equivalent to C0(H\Z).
Proof. The groupoid GU ⋊ S is equivalent to Z ⋊ S. This is the same as Z ⋊ H
by Theorem 3.14, using the evident S-grading on Z ⋊ H. Since the H-action on Z
is basic, Z ⋊ H is equivalent to H\Z. This space is assumed to be Hausdorff, and
GU ⋊ S is also a groupoid with Hausdorff object space. So the equivalence between
them is of the usual type, involving free and proper actions, by Proposition A.7.
Hence it induces a Morita -- Rieffel equivalence from C0(H\Z) to C∗(GU) ⋊ S. (cid:3)
In the situation of Proposition 6.5, GU ⋊ S has Hausdorff arrow space because it
must be isomorphic to the covering groupoid of the open surjection G0
U → (GU ⋊
∼= H\Z between two Hausdorff spaces. In this case, it is also easy to see that
S)\G0
U
any Fell bundle over the groupoid GU ⋊ S is a pull-back of a Fell bundle over H\Z,
which is the same as a C0(H\Z)-C∗-algebra B. The section C∗-algebra of the Fell
bundle over GU ⋊ S is Morita -- Rieffel equivalent to this C0(H\Z)-C∗-algebra B. By
Theorem 5.5, this is also the section C∗-algebra of the Fell bundle over S associated
to B.
Many properties like properness, amenability, essential principality are shared by
an action of a groupoid on a space and its transformation groupoid. This suggests
how to extend these notions to inverse semigroup actions on groupoids. We take
40
ALCIDES BUSS AND RALF MEYER
this as a definition for proper actions of inverse semigroups on locally compact
groupoids:
Definition 6.6. An action of an inverse semigroup S on a topological groupoid G
is proper if the groupoid G ⋊ S is proper, that is, the following map is proper (that
is, stably closed):
(s, r): (G ⋊ S)1 → G0 × G0,
g 7→ (s(g), r(g)).
The action is called free if this map is injective.
Let L be a proper groupoid such that L0 is a locally compact Hausdorff space.
Then the image of L1 in L0 × L0 is locally compact and Hausdorff because it is a
closed subspace of a locally compact Hausdorff space. Since this subspace is closed
and the orbit space projection L0 → L\L0 is open, it also follows that L\L0 is
locally compact Hausdorff (see Proposition A.3). The groupoid L itself need not
be Hausdorff: the non-Hausdorff group bundle in Section 8 is proper in this sense
because it is quasi-compact and the image of (s, r) is closed. If L acts freely and
properly on a Hausdorff space L0, however, then L1 must be Hausdorff. In this case,
we also get information about any open subgroupoid, which leads to the following
proposition:
Proposition 6.7. Let S act properly and freely on a locally Hausdorff, locally
quasi-compact groupoid G. Then G is a basic groupoid, so that G is equivalent to
the locally Hausdorff, locally quasi-compact space G\G0.
Proof. The map in Definition 6.6 is a homeomorphism onto its image because it
is continuous, injective, and closed. Hence its restriction to the open subspace
G1 ⊆ (G ⋊ S)1 is still a homeomorphism onto its image. This means that G is a
basic groupoid, so G is equivalent to G0/G. This is locally Hausdorff and locally
quasi-compact by Proposition A.14.
(cid:3)
Thus the free and proper actions of S all come from actions on locally Hausdorff
spaces that are desingularised by replacing the space by a Hausdorff groupoid G.
6.2. Inverse semigroup models for étale groupoids. Let G be an étale group-
oid. So far, we have described actions of G through actions of the inverse semi-
group Bis(G). Since Bis(G) is usually quite big, even uncountable, we now replace
it by smaller inverse semigroups. The following definition describes which inverse
semigroups we allow as "models" for G:
Definition 6.8. An inverse semigroup model for an étale groupoid G consists of
an inverse semigroup S, an S-action on the space G0 by partial homeomorphisms,
and an isomorphism G0 ⋊ S ∼= G of étale groupoids that is the identity on objects.
In particular, if S ⊆ Bis(G) is a wide inverse subsemigroup, then S with its usual
action on G0 and the canonical isomorphism G0 ⋊ S ∼= G from Corollary 3.19 is a
model for G.
Lemma 6.9. An inverse semigroup model for G is equivalent to an inverse semi-
group S with a homomorphism ϕ: S → Bis(G) that induces an isomorphism G0 ⋊
S → G0 ⋊ Bis(G) ∼= G, where we use the canonical action of Bis(G) on G0 and ϕ
to let S act on G0.
Proof. Let S act on G0. There is a canonical homomorphism S → Bis(G0 ⋊ S),
see [12]. Combined with an isomorphism G0 ⋊ S ∼= G, we get a homomorphism
ϕ: S → Bis(G). Conversely, such a homomorphism induces an action of S on G0
and then a continuous groupoid homomorphism G0 ⋊ S → G0 ⋊ Bis(G) ∼= G.
Routine computations show that these two constructions are inverse to each other.
(cid:3)
INVERSE SEMIGROUP ACTIONS ON GROUPOIDS
41
The following lemma characterises inverse semigroup models more concretely
when we take S = Bis(G).
Lemma 6.10. Let S and S be inverse semigroups, let ϕ: S → S be a homo-
morphism, and let S act on Z by partial homeomorphisms. The induced groupoid
homomorphism ϕ : Z ⋊ S → Z ⋊ S is an isomorphism if and only if
(1) for all t1, t2 ∈ S and every z ∈ Z with z ∈ Dt∗
2 t2 and every f ∈ E( S)
with z ∈ Df and ϕ(t1)f = ϕ(t2)f , there is e ∈ E(S) with z ∈ De and
t1e = t2e;
(2) for every u ∈ S and every z ∈ Z with z ∈ Du∗u, there is t ∈ S with z ∈ Dt∗t
1 t1 ∩ Dt∗
and there is f ∈ E( S) with z ∈ Df and uf = ϕ(t)f .
In this case, we call ϕ a Z-isomorphism.
Proof. The groupoid homomorphism ϕ is the identity on objects and always contin-
uous and open on arrows, so the only issue is whether ϕ is bijective on arrows. It is
routine to check that (1) is equivalent to injectivity and (2) to surjectivity of ϕ. (cid:3)
Let S and ϕ: S → Bis(G) be an inverse semigroup model for an étale topolog-
ical groupoid G. Which actions of S on groupoids by partial equivalences or on
C∗-algebras by Hilbert bimodules come from actions of G?
First we consider a trivial special case to see why we need more data. Let G be
just a topological space, viewed as a groupoid. In this case, the trivial inverse semi-
group {1} is an inverse semigroup model. An action of S contains no information.
An action of G on a topological groupoid H or a C∗-algebra is simply a continuous
map ψ : H 0/H → G0 or ψ : Prim(A) → G0, respectively.
Theorem 6.11. Let G be a sober étale topological groupoid and let S and ϕ: S →
Bis(G) be an inverse semigroup model for G. Let H be a topological groupoid.
An action of G on H by partial equivalences is equivalent to a pair consisting of an
action of S on H by partial equivalences and an S-equivariant map ψ : H 0/H → G0.
The transformation groupoid for an action of G (that is, Bis(G)) and its restriction
to S are the same.
The S-equivariance of ψ refers to the actions of S on H 0/H and G0 by partial
homeomorphisms induced by the action on H and by ϕ.
Proof. First let G act on H; more precisely, Bis(G) acts on H and the resulting
map O(G0) = E(Bis(G)) → O(H 0/H) commutes with suprema (Definition 6.3).
[20, Lemma 2.25] shows that it comes from a continuous map ψ : H 0/H → G0.
This map is Bis(G)-equivariant and hence S-equivariant.
Now let S act on H and let ψ : H 0/H → G0 be an S-equivariant map. Let
L := H ⋊ S with its canonical S-grading (Lt)t∈S. We claim that there is a unique
Bis(G)-grading ( ¯Lt)t∈Bis(G) on L with ¯Lϕ(t) = Lt for all t ∈ S, and ¯LU = H 1
ψ−1(U)
for U ∈ O(G0). These two conditions on the Bis(G)-grading say exactly that it
corresponds to the given S-action and map ψ. So the proof of the claim will finish
the proof of the theorem.
For t ∈ Bis(G) and u ∈ S, we may form t ∩ ϕ(u) ∈ Bis(G). We have
t ∩ ϕ(u) = t · Vt,u = ϕ(u) · Vt,u
for Vt,u = s(t ∩ ϕ(u)) ∈ O(G0);
here we also view Vt,u as an idempotent element of Bis(G). Since S models G,
we have t = Su∈S t ∩ ϕ(u) and hence s(t) = Su∈S Vt,u. Any Bis(G)-grading with
ψ−1(V ) for all V ∈ O(G0) satisfies
¯LV = H 1
¯Ltψ−1(Vt,u) = ¯Lt · ¯LVt,u = ¯Lt∩ϕ(u) = ¯Lϕ(u)ψ−1(Vt,u)
42
ALCIDES BUSS AND RALF MEYER
for all t ∈ Bis(G), u ∈ S. Since s(t) =Su∈S Vt,u and ψ is S-equivariant, this shows
that there is at most one Bis(G)-grading with the required properties, namely,
¯Lt = [u∈S
Luψ−1(Vt,u).
More explicitly, l ∈ ¯Lt if and only if l ∈ Lu for some u ∈ S for which t and ϕ(u)
have the same germ at ψ(s(l)). We must prove that ( ¯Lt)t∈Bis(G) is a grading with
all desired properties.
First we check ¯Lϕ(u) = Lu for u ∈ S. The inclusion ⊇ is trivial. If l ∈ ¯Lϕ(u),
then l ∈ Lu′ for some u′ ∈ S for which ϕ(u) and ϕ(u′) have the same germ at
ψ(s(l)) ∈ G0. Hence there is an idempotent element e ∈ S with ψ(s(l)) ∈ ϕ(e) and
ue = u′e. Since Le = H 1
ψ−1(e), we get l ∈ Lu′Le = Lu′e = Lue = LuLe ⊆ Lu. This
finishes the proof that ¯Lϕ(u) = Lu for all u ∈ S.
Next we check ¯LW = H 1
ψ−1(W ) for W ∈ O(G0). The inclusion ⊇ holds because
VW,1 = W . Conversely, let l ∈ ¯LW . Then l ∈ Lu for some u ∈ S for which ϕ(u)
and IdW have the same germ at ψ(s(l)). Since G0 ⋊ S ∼= G, there is an idempotent
e ∈ S with ψ(s(l)) ∈ ϕ(e) and ue = e. An argument as in the previous paragraph
shows that l ∈ LuLe = Le ⊆ H 1. Thus ¯LW = H 1
ψ−1(W ) for all W ∈ O(G0).
If t ∈ Bis(G), u ∈ S, then (ϕ(u) ∩ t)∗ = ϕ(u∗) ∩ t∗. Hence Vt∗,u∗ = t(Vt,u) =
ϕ(u)(Vt,u). This implies Lt∗ = L−1
t
for all t ∈ Bis(G).
Let t1, t2 ∈ Bis(G). We claim that ¯Lt1 · ¯Lt1 = ¯Lt1t2. The inclusion ⊆ follows
because (ϕ(u1) ∩ t1) · (ϕ(u2) ∩ t2) ⊆ ϕ(u1u2) ∩ t1t2. For the converse inclusion, take
l ∈ ¯Lt1t2. Then t ∈ Lu for some u ∈ S for which t1t2 and ϕ(u) have the same germ
at ψ(s(l)). Factor this germ as g1g2 with gj ∈ tj for j = 1, 2. There are uj ∈ S
with gj ∈ ϕ(uj) for j = 1, 2 because G ∼= G0 ⋊ S. Thus ϕ(u1)ϕ(u2) = ϕ(u1u2)
and t1t2 have the same germ g1g2 at ψ(s(l)). Then u1u2 and u also have the same
germ there, and an argument as above shows that l ∈ Lu1u2 as well. Using (Gr1)
for the S-grading, we get lj ∈ Luj for j = 1, 2 with l = l1l2. Then s(l2) = s(l) and
, so that l1 ∈ ¯Lt1. Hence
r(l1) = r(l). This allows to prove l2 ∈ ¯Lt2 and l−1
the Bis(G)-grading satisfies (Gr1).
It is clear that ¯Lt1 ⊆ ¯Lt2 if t1 ≤ t2 in Bis(G), so ¯Lt1 ∩ ¯Lt2 ⊇Sv≤t1,t2
¯Lv = ¯Lt1∩t2
for all t1, t2 ∈ Bis(G). For the converse inclusion, take l ∈ ¯Lt1 ∩ ¯Lt2. Then there are
u1, u2 ∈ S with l ∈ Lu1 ∩ Lu2, such that tj and ϕ(uj) have the same germ at ψ(s(l))
for j = 1, 2. (Gr3) for the S-grading gives v ∈ S with v ≤ u1, u2 and l ∈ Lv.
Since ψ is S-equivariant, ψ(s(l)) belongs to the domain of ϕ(v), so the germs of
ϕ(v) and ϕ(ui) at ψ(s(l)) are equal. Then the germs of t1 and t2 at ψ(s(l)) are
equal as well, that is, t1 ∩ t2 is defined at ψ(s(l)) and has the same germ there
as ϕ(v). This means that l ∈ ¯Lt1∩t2. This verifies (Gr3) for the Bis(G)-grading.
Since ¯Lϕ(u) = Lu for all u ∈ S andSu∈S Lu = L1, we also getSt∈Bis(G)
¯Lt = L1,
(cid:3)
which is (Gr4).
1 ∈ ¯Lt∗
1
The following lemma is needed to formulate a similar result for actions on
C∗-algebras:
Lemma 6.12. An action of S on a C∗-algebra A by Hilbert bimodules induces an
action of S on Prim(A) by partial homeomorphisms.
Proof. The Rieffel Correspondence (see [29, Corollary 3.33]) says that an imprim-
itivity bimodule H from B to A induces a homeomorphism Prim(B) ∼−→ Prim(A).
The corresponding lattice isomorphism
I(B) = O(Prim(B)) ∼−→ O(Prim(A)) = I(A)
INVERSE SEMIGROUP ACTIONS ON GROUPOIDS
43
sends an ideal J ⊆ B to the unique ideal I ⊆ A with I · H = H · J. A Hilbert
A, B-bimodule induces a partial homeomorphism Prim(B) → Prim(A) because it is
an imprimitivity bimodule between certain ideals in A and B, which correspond to
open subsets of the primitive ideal spaces. Isomorphic Hilbert bimodules induce the
same partial homeomorphism, of course. The partial homeomorphism associated
to a tensor product bimodule H1 ⊗B H2 is the composite of the partial homeo-
morphisms associated to H1 and H2. Thus the map from S to pHomeo(Prim(A))
induced by an action on A by Hilbert bimodules is a homomorphism.
(cid:3)
Theorem 6.13. Let G be a sober étale topological groupoid and let S and ϕ: S →
Bis(G) be an inverse semigroup model for G. Let A be a C∗-algebra. An action of G
on A by Hilbert bimodules is equivalent to a pair consisting of an action of S on A
by Hilbert bimodules and an S-equivariant map ψ : Prim(A) → G0. The section
C∗-algebras of the corresponding Fell bundles over Bis(G) and S are the same.
The S-equivariance of ψ refers to the action of S on Prim(A) from Lemma 6.12.
Proof. Assume first that G0 is locally compact Hausdorff. In that case, an action
of G is the same as a Fell bundle over G by Theorem 6.1. This determines an
action of Bis(G), which we may compose with ϕ to get an action of S; we also
get an S-equivariant map ψ. Conversely, let an action of S and a continuous
S-equivariant map ψ : Prim(A) → G0 be given. Since G ∼= G0 ⋊ S, we may carry
over the proof of Theorem 6.1. The S-equivariance of ψ gives the compatibility
condition (18). Hence literally the same argument still works.
If G0 is only a sober topological space, we need a different proof because we
cannot describe G-actions fibrewise. We first construct the section C∗-algebra B of
the Fell bundle over S corresponding to the action by Theorem 4.8. This C∗-algebra
is S-graded by construction: it is the Hausdorff completion of the ∗-algebraLt∈S Ht
in the maximal C∗-seminorm that vanishes on ju,t(ξ)δu − ξδt for all t, u ∈ S with
t ≤ u and all ξ ∈ Ht, and we let Bt ⊆ B be the image of Ht in B. In particular, we
may identify A = B1. Now we must construct a Bis(G)-grading ( ¯Bt)t∈Bis(G) on B
with ¯Bϕ(t) = Bt for all t ∈ S and ¯BU = A(U ) for all U ∈ O(G0), where A(U ) denotes
the ideal in A corresponding to ψ−1(U ) ∈ O(Prim(A)). This is done similarly to the
proof of Theorem 6.11. Since this is rather technical and we already have another
proof in the locally compact Hausdorff case, we leave it to the determined reader
to spell out the details of this argument.
(cid:3)
7. Actions by automorphisms are not enough
The following theorem shows that the multiplication action of a non-Hausdorff
groupoid on its own arrow space cannot be described by a continuous groupoid
action by automorphisms.
Theorem 7.1. Let G be a locally quasi-compact, locally Hausdorff, étale groupoid
with Hausdorff G0, such that G1 is not Hausdorff. Let A be a C∗-algebra with
Prim(A) ∼= G1. There is no continuous (twisted) action of G on A by automor-
phisms that induces the left multiplication action on Prim(A) ∼= G1.
Proof. Since Prim(A) ∼= G1, the lattice of ideals in A is order-isomorphic to the
lattice of open subsets in G1. Let A(U ) ⊳ A for an open subset U ⊆ G1 be the
corresponding ideal in A. Then Prim(A(U )) ∼= U .
Part of a continuous action of G on A is a continuous map Prim(A) → G0.
(This is equivalent to a C0(G0)-algebra structure.) Since we want to have the
left multiplication action of G1 on Prim(A), we assume that this map becomes
the range map G1 → G0 when we identify Prim(A) ∼= G1. The fibre at x ∈ G0
is the restriction of A to the closed subset Gx = {g ∈ G1 r(g) = x}, which
44
ALCIDES BUSS AND RALF MEYER
we denote by AGx; we have Prim(AGx) = Gx. A G-action on A must provide
isomorphisms αg : AGs(g) → AGr(g) for g ∈ G1. We assume that αg induces the
map Gs(g) → Gr(g), h 7→ gh, on the primitive ideal space.
What does continuity of g 7→ αg mean? Let U, V ⊆ G1 be bisections, then
U · V is also a bisection. If g ∈ U , h ∈ V satisfy s(g) = r(h), then αg restricts to
an isomorphism αg,h : Ah → Agh. Any element of U · V is of the form g · h for
unique g ∈ U , h ∈ V . Continuity of (αg) means that for all bisections U, V and all
a = (ah)h∈V in A(V ), the section (g · h) 7→ αg,h(ah) for g ∈ U , h ∈ V is continuous
on U · V , that is, it belongs to A(U · V ) (see also [27, Definition 2.3]). Thus we
get isomorphisms αU : A(V ) → A(U · V ).
In brief, Bis(G) acts on A by partial
isomorphisms.
Since G1 is non-Hausdorff, there are g1, g2 ∈ G1 that cannot be separated by
open subsets. Then r(g1) = r(g2) and s(g1) = s(g2). Let U1 and U2 be bisections
of G containing g1 and g2, respectively. Shrinking them, we may achieve that
s(U1) = s(U2). Let
V := U ∗
1 U1 = {1x x ∈ s(U1)} = U ∗
2 U2;
then U1V = U1 and U2V = U2. Since g1 and g2 cannot be separated, there is a
net (hn) in U1 ∩ U2 that converges both to g1 and to g2.
Let f ∈ A(V ) with f (1s(g1)) 6= 0. Then αU1(f ) ∈ A(U1V ) and αU2(f ) ∈ A(U2V )
by our continuity assumption. Thus
ψ := αU1(f ) · αU2(f )∗ ∈ A(U1V ) ∩ A(U2V ) = A(U1 ∩ U2),
so ψ vanishes at g1 and g2. At hn ∈ U1 ∩ U2, we have
αU1(f )(hn) = αU1∩U2(f )(hn) = αU2(f )(hn) = αhn(f (1s(hn))).
Since each αhn is an isomorphism, we get
kψ(hn)k = kαhn(f (1s(hh))f (1s(hn))∗)k = kf (1s(hh))k2.
If U ⊆ G1 is Hausdorff and a ∈ A(U ), then U ∋ x 7→ kakx is continuous by [26,
Corollary 2.2] because the map Prim A(U ) → U is open and U is Hausdorff and
locally compact. Therefore, kψ(hn)k converges towards kψ(g1)k = 0. At the same
time, kψ(hn)k converges towards kf (1s(g1))k2 6= 0 because s(hn) → s(g1) inside
the Hausdorff open subset V . This contradiction shows that there is no continuous
action of G on A that lifts the multiplication action on Prim(A) ∼= G1.
(cid:3)
Remark 7.2. More generally, if we only assume an open continuous surjection
p: Prim(A) → G1, then there is no continuous action of G on A such that p is
G-equivariant for the induced action of G on Prim(A) and the left multiplication
action on G1; the proof is exactly the same.
The proof of Theorem 7.1 does not care about the multiplicativity of the action,
so allowing "twisted" actions of G does not help. There are only two ways around
this. First, we may allow Fell bundles over G. Secondly, we may allow actions
of the inverse semigroup Bis(G). After stabilisation every Fell bundle becomes a
twisted action by partial automorphisms (see [3]). We cannot remove the twist,
however, because an untwisted action of Bis(G) by automorphisms would give an
action of G by automorphisms as well, which cannot exist by Theorem 7.1.
8. A simple explicit example
Let G be the group bundle over G0 = [0, 1] with trivial isotropy groups G(x)
for x 6= 0 and with G(0) ∼= Z/2 = {1, −1}. So, as a set, G is (0, 1] ∪ {0+, 0−}
with 0+ corresponding to +1 ∈ Z/2 and 0− to −1 ∈ Z/2. The topology on G
is the quotient topology from [0, 1] × Z/2, where we divide by the equivalence
INVERSE SEMIGROUP ACTIONS ON GROUPOIDS
45
relation generated by (x, 1) ∼ (x, −1) for x 6= 0. With this topology, G is an étale,
quasi-compact, second countable, locally Hausdorff, non-Hausdorff groupoid (even
a group bundle). The points 0+ and 0− cannot be topologically separated: any net
in (0, 1] converging to 0+ also converges to 0−, and vice versa.
Let H be the groupoid of the equivalence relation ∼ on [0, 1] × Z/2 just defined.
Its C∗-algebra C∗(H) ∼= C∗
r (H) is
A := {f ∈ C([0, 1], M2) : f (0) is diagonal}.
(This can be proved using the same idea as in [8, Example 7.1].) This is a C∗-algebra
∼= C2, and it has A ∼= Prim(A) ∼= G1
over [0, 1] with fibres Ax
(this is a special case of [8, Corollary 5.4]). Theorem 7.1 shows that there is no action
of G on A by automorphisms that would model the left multiplication action of G
on G1.
∼= M2 at x 6= 0 and A0
Since A is the groupoid C∗-algebra of the Čech groupoid for the covering [0+, 1]∪
[0−, 1] = H 1, our main results give an action of G on A by Hilbert bimodules. We
first describe it as an inverse semigroup action for a very small inverse semigroup S
that models G. We consider three special bisections of G:
1 = [0+, 1] = G1 \ {0−},
g = [0−, 1] = G1 \ {0+},
e = (0, 1] = g ∩ 1.
The bisection 1 is the unit bisection of G, so 1x = x = x1 for all x ∈ {1, g, e}.
Moreover, g2 = 1, e2 = e, and eg = ge = e. Thus S := {1, e, g} is an inverse
semigroup with x∗ = x for all x ∈ {1, e, g}. A bisection t of G cannot contain both
0+ and 0−. Hence either 0+ ∈ t ⊆ 1, 0− ∈ t ⊆ g, or t ⊆ e = 1 ∩ g.
The groupoid G is the étale groupoid associated to the trivial action of S on G0;
here the trivial action has 1 and g acting by the identity on G0 and e acting by the
identity on (0, 1] ⊆ G0. An action of G on a groupoid or a C∗-algebra is equivalent
to an action of S together with a compatible action of G0 = [0, 1] (Theorem 6.13).
The transformation groupoid L of the S-action on H may be identified with the
groupoid of the equivalence relation on [0, 1] ⊔ [0, 1] that identifies the two copies
of (0, 1], so that
L1 = [0, 1] × {(+, +), (+, −), (−, +), (−, −)}
⊆ ([0, 1] × {(+, +), (+, −), (−, +), (−, −)})2.
The S-grading on L has
L1 = (0, 1] × {(+, +), (+, −), (−, +), (−, −)} ⊔ {0} × {(+, +), (−, −)},
Lg = (0, 1] × {(+, +), (+, −), (−, +), (−, −)} ⊔ {0} × {(+, −), (−, +)},
Le = (0, 1] × {(+, +), (+, −), (−, +), (−, −)} = L1 ∩ Lg.
So L1 ∼= H is open but not closed. The C∗-algebra of L is B := C([0, 1], M2).
To let S act on the C∗-algebra A of H, we use the transformation groupoid
C∗-algebra B and the involution u := (cid:0) 0 1
1 0(cid:1) ∈ B. We have u = u∗ and u2 = 1,
u · A(0, 1] = A(0, 1] = A(0, 1] · u and uA = Au as subsets of B. Let A1 := A,
Ae := A(0, 1] ⊆ A1, and Ag := uA = Au. These subspaces Ax for x ∈ S satisfy
A∗
x = Ax = Ax∗ for all x ∈ S and Ax · Ay = Axy for all x, y ∈ S; in particular, Ag is
a full Hilbert bimodule over A1 with inner products given by the usual formulas
a∗
1 · a2 and a1 · a∗
2. Furthermore, A1 ∩ Ag = Ae and A1 + Ag = B because elements
of Ag are precisely those f ∈ B with off-diagonal f (0). Hence the map g 7→ Ag
defines an action of S on A by Hilbert bimodules. Since A1 +Ag is already complete
in the C∗-norm of B, there is only one C∗-norm on A1 + Ag that extends the given
C∗-norm on A1. Thus the sectional C∗-algebra for the resulting Fell bundle over S
is B, which is Morita -- Rieffel equivalent to C[0, 1].
46
ALCIDES BUSS AND RALF MEYER
The S-action on A extends to all bisections of G because they are all contained
in 1 or g: if t ⊆ G1 is a bisection, then let At = A1s(t) if t ⊆ 1 and At = Ags(t) if
t ⊆ g; this is consistent for t ⊆ 1 ∩ g = e because Ae = A1 ∩ Ag.
Next we describe a twisted S-action by partial automorphisms of A that induces
the S-action by Hilbert bimodules described above. (This is possible by [3, Corol-
lary 4.16] because our saturated Fell bundle is regular in the notation of [3].)
A twisted S-action by partial automorphisms is given by ideals A1 = A and Ae
with isomorphisms αx : Axx∗ → Ax∗x and unitary multipliers (the twists) ω(x, y) in
M(Axyy∗x∗) for x, y ∈ S. For the idempotent elements x = e, 1, the isomorphism αx
is the identity; for x = g, it is the order-2-automorphism αg : A → A, a 7→ uau,
because a1 · ua2 = u · (ua1u · a2) for all a1 ∈ A1, ua2 ∈ Ag. The automorphism αg is
not inner on A1 because u ∈ B does not belong to M(A). The restriction of αg to
the ideal Ae becomes inner, however, because u ∈ M(A(0, 1]). This unitary u enters
in the twisting unitaries ω(x, y) for x, y ∈ S; they are 1 if x = 1 or y = 1, or if (x, y)
is (e, e) or (g, g) (α2
g = IdA = α1). The remaining cases are ω(e, g) = ω(g, e) = uAe,
that is, u viewed as a multiplier of the ideal Ae = A(0, 1]. It is routine to check
that this data gives a twisted action of S on A in the sense of [3, Definition 4.1] and
that the resulting saturated Fell bundle over S is isomorphic to the one described
above. Incidentally, this is not a twisted action in the sense of Sieben [34] because
ω(e, g) and ω(g, e) are non-trivial although e is idempotent.
This twisted S-action cannot be turned into a groupoid action of G by partial
automorphisms because for x ∈ 1∩g, the restrictions of αg and α1 to As(x) differ by
a non-trivial inner automorphism. This impossibility is in accord with Theorem 7.1.
Remark 8.1. The Packer -- Raeburn Stabilisation Trick replaces a twisted group ac-
tion by an untwisted action on a suitable C∗-stabilisation. We claim that this cannot
be done for the above inverse semigroup twisted action. Let D be a C∗-algebra with
an untwisted action of S by automorphisms. Then 1 and e act by the identity on D
and by some ideal De ⊳ D, respectively, and g acts by some automorphism αg on D.
If there is no twist, then αgDe = α1De is the identity on De because eg = e = ge.
Suppose that D is also a C∗-algebra over [0, 1], with D((0, 1]) = De. Then this
allows to define an action of the groupoid G on D by letting elements of g or 1
act by the fibre restrictions of αg and IdD, respectively. This gives a well-defined,
untwisted action of G on D. Theorem 7.1 implies that Prim(D) 6∼= G1, so that
A and D cannot be Morita -- Rieffel equivalent. This example therefore shows that
the Packer -- Raeburn Stabilisation Trick cannot be extended from groups to inverse
semigroups or non-Hausdorff groupoids.
Appendix A. Preliminaries on topological groupoids
This appendix defines topological groupoids and equivalences between them, fol-
lowing [21]. The point is that all this works smoothly without assuming topological
spaces to be Hausdorff or locally (quasi)compact, if we choose appropriate defini-
tions. The theory of possibly non-Hausdorff topological groupoids becomes very
natural if one treats topological groupoids, Lie groupoids, infinite-dimensional Lie
groupoids (modelled on Banach or Fréchet manifolds), and other types of groupoids
simultaneously as in [21]. Here we recall the results and definitions from [21] that
are relevant for us.
The theory of topological groupoids and their principal bundles and equivalences
depends on a choice of "covers" in the category of topological spaces (see [21]). We
choose the open surjections as covers. This means that we require the range and
source maps in a topological groupoid, the bundle projection in a principal bundle,
and the anchor maps in a (bibundle) equivalence to be open surjections.
INVERSE SEMIGROUP ACTIONS ON GROUPOIDS
47
Following Bourbaki, we require compact and locally compact spaces to be Haus-
dorff. Since many authors allow non-Hausdorff locally compact spaces, we usually
speak of "Hausdorff locally compact" spaces to avoid confusion. A topological space
is locally quasi-compact if every point has a neighbourhood basis consisting of quasi-
compact neighbourhoods. This is strictly more than having a single quasi-compact
neighbourhood, but both notions coincide in the locally Hausdorff case, which is the
case we are interested in. Recall that a topological space is locally Hausdorff if every
point has a Hausdorff neighbourhood (and thus a neighbourhood basis consisting of
Hausdorff neighbourhoods). A space is locally Hausdorff, locally quasi-compact if
and only if every point has a compact (hence Hausdorff) neighbourhood. It would
make sense to call such spaces "locally compact," if it were not for the conflict with
other established notation.
A.1. Topological groupoids, principal bundles, and equivalences. We now
specialise the general definitions of groupoids, groupoid actions, principal bundles,
basic groupoid actions and bibundle equivalences in [21] to the category of (all)
topological spaces with open surjections as covers.
Proposition A.1. A topological groupoid consists of topological spaces G0 and G1
and continuous maps r, s: G1 ⇒ G0 and m: G1 ×s,G0,r G1 → G1, (g1, g2) 7→ g1 · g2,
such that
(G1) s(g1 · g2) = s(g2) and r(g1 · g2) = r(g1) for all g1, g2 ∈ G1;
(G2) m is associative: (g1 · g2) · g3 = g1 · (g2 · g3) for all g1, g2, g3 ∈ G1 with
s(g1) = r(g2), s(g2) = r(g3);
(G3) the following two maps are homeomorphisms:
G1 ×s,G0,r G1 → G1 ×s,G0,s G1,
G1 ×s,G0,r G1 → G1 ×r,G0,r G1,
(g1, g2) 7→ (g1 · g2, g2),
(g1, g2) 7→ (g1, g1 · g2),
(G4) r and s are open surjections.
Then m is open and surjective and there are continuous maps G0 → G1 and G1 →
G1 with the usual properties of unit and inversion. Conversely, the maps in (G3)
are homeomorphisms if G has continuous unit and inversion maps.
Proof. Our definition of a groupoid is exactly [21, Definition 3.4]. It implies that m
is open and surjective and is equivalent to the usual one with unit and inverse by
[21, Proposition 3.6].
(cid:3)
Let G be a topological groupoid as above.
Proposition A.2. A (right) G-action is a space X with continuous maps s: X →
G0 and m: X ×s,G0,r G1 → X, (x, g) 7→ x · g, such that
(A1) s(x · g) = s(g) for all x ∈ X, g ∈ G1 with s(x) = r(g);
(A2) m is associative: (x · g1) · g2 = x · (g1 · g2) for all x ∈ X, g1, g2 ∈ G1 with
s(x) = r(g1) and s(g1) = r(g2);
(A3) m is surjective.
Condition (A3) holds if and only if x · 1s(x) = x for all x ∈ X, if and only if m is
an open surjection, if and only if the following map is a homeomorphism:
X ×s,G0,r G1 → X ×s,G0,s G1,
(x, g) 7→ (x · g, g).
Proof. This is contained in [21, Definition and Lemma 4.1].
(cid:3)
Left actions are defined similarly and are equivalent to right actions by g · x =
x · g−1. The transformation groupoid X ⋊ G of a groupoid action is a topological
groupoid by [21, Definition and Lemma 4.11]. Any groupoid acts on G0 by r(g) ·
48
ALCIDES BUSS AND RALF MEYER
g := s(g) for all g ∈ G1, and on G1 both on the left and right by left and right
multiplication.
Proposition A.3. For any G-action on a topological space X, the orbit space
projection X → X/G is an open surjection, and X/G is Hausdorff if and only if
X ×X/G X = {(x1, x2) ∈ X
there is g ∈ G1 with s(x1) = r(g) and x1 · g = x2}
is a closed subset of X × X.
Proof. The orbit space projection is open by [21, Proposition 9.31] because the
range and source maps of G are open. By [21, Proposition 9.18], X/G is Hausdorff
if and only if X ×X/G X is closed in X × X (open surjections are clearly biquotient
maps, see the discussion in [21, Section 9.6]).
(cid:3)
We now specialise the general concepts of basic actions and principal bundles
from [21] to our context.
Proposition A.4. A right G-action is basic if the map
(19)
X ×s,G0,r G1 → X × X,
(x, g) 7→ (x, x · g),
is a homeomorphism onto its image with the subspace topology.
A principal right G-bundle is a space X with continuous maps s: X → G0,
p: X → Z, and m: X ×s,G0,r G1 → X, (x, g) 7→ x · g, such that
(Pr1) s(x · g) = s(g) and p(x · g) = p(x) for all x ∈ X, g ∈ G1 with s(x) = r(g);
(Pr2) m is associative: (x · g1) · g2 = x · (g1 · g2) for all x ∈ X, g1, g2 ∈ G1 with
s(x) = r(g1) and s(g1) = r(g2);
(Pr3) the map
X ×s,G0,r G1 → X ×p,Z,p X,
(x, g) 7→ (x, x · g),
is a homeomorphism;
(Pr4) the map p is open and surjective.
Then x · 1s(x) = x for all x ∈ X, and there is a unique homeomorphism Z ∼= X/G
intertwining p and the canonical projection X → X/G. Thus a principal G-bundle
is equivalent to a basic G-action with a homeomorphism X/G ∼= Z.
Proof. A principal bundle in the sense above also satisfies x · 1s(x) = x for all x ∈ X
because of (Pr3) (see [21, Lemma 5.3]). Hence s and m give a right G-action, and
all conditions for a principal bundle in [21] are met. [21, Lemma 5.3] also gives the
unique homeomorphism X/G ∼= Z intertwining p and the canonical map X → X/G.
A groupoid action is called basic in [21] if it becomes a principal bundle with
X → X/G as bundle projection. The canonical map X → X/G is automatically
G-invariant, and it is an open surjection by [21, Proposition 9.31]. Thus the second
half of (Pr1) and (Pr4) hold for any G-action with this choice of p. The first half
of (Pr1) and (Pr2) are part of the definition of a groupoid action. The image of
the map in (19) is X ×X/G X by the definition of X/G, so that (Pr3) is equivalent
to (19) being a homeomorphism onto its image.
(cid:3)
Next, we consider the notion of equivalence between groupoids as defined in [21].
We will relate it to notions of equivalence by other authors in Appendix A.2.
Proposition A.5. Let G and H be topological groupoids. A bibundle equivalence
from H to G consists of a topological space X, continuous maps r : X → G0,
s: X → H 0 (anchor maps), G1 ×s,G0,r X → X and X ×s,H0,r H 1 → X (multi-
plications), satisfying the following conditions:
INVERSE SEMIGROUP ACTIONS ON GROUPOIDS
49
(E1) s(g · x) = s(x), r(g · x) = r(g) for all g ∈ G1, x ∈ X with s(g) = r(x), and
s(x · h) = s(h), r(x · h) = r(x) for all x ∈ X, h ∈ H 1 with s(x) = r(h);
(E2) associativity: g1 · (g2 · x) = (g1 · g2) · x, g2 · (x · h1) = (g2 · x) · h1, x · (h1 · h2) =
(x · h1) · h2 for all g1, g2 ∈ G1, x ∈ X, h1, h2 ∈ H 1 with s(g1) = r(g2),
s(g2) = r(x), s(x) = r(h1), s(h1) = r(h2);
(E3) the following two maps are homeomorphisms:
G1 ×s,G0,r X → X ×s,H0,s X,
X ×s,H0,r H 1 → X ×r,G0,r X,
(g, x) 7→ (x, g · x),
(x, h) 7→ (x, x · h);
(E4) s and r are open;
(E5) s and r are surjective.
Then 1r(x) · x = x = x · 1s(x) for all x ∈ X, and the anchor maps descend to
homeomorphisms G\X ∼= H 0 and X/H ∼= G0.
Proof. Condition (E1) and (E3) are equivalent to (Pr1) and (Pr3) for both the
left G-action with p = s and the right H-action with p = r, respectively. Condi-
tion (E2) means that the left G-action and the right H-action satisfy (Pr2) and
commute. Conditions (E4) and (E5) together are equivalent to (Pr4) for both
actions. Thus the conditions (E1) -- (E5) characterise bibundle equivalences in the
notation of [21]. The last sentence follows from the general properties of principal
bundles, see Proposition A.4.
(cid:3)
In the following, we abbreviate "bibundle equivalence" to "equivalence" because
we do not use any other equivalences between groupoids.
We have switched the direction of a bibundle equivalence compared to [21] be-
cause this is convenient here. Going from right to left is also consistent with our
notation s and r for the right and left anchor maps.
A.2. Basic actions versus free and proper actions. We now compare our basic
actions with free and proper actions. A continuous map f : X → Y is closed if it
maps closed subsets of X to closed subsets of Y , and proper if IdZ × f : Z × X →
Z ×Y is closed for all topological spaces Z or, equivalently, f is closed and f −1(y) is
quasi-compact for all y ∈ Y (see [2, Theorem 1 in I.10.2]). A map from a Hausdorff
space X to a Hausdorff locally compact space Y is proper if and only if preimages
of compact subsets are compact.
In this case, X is necessarily locally compact
([2, Proposition 7 in I.10.3]).
Definition A.6. A right action of a topological groupoid G on a topological
space X is proper if the map in (19) is proper. The action is free if the map (19) is
injective.
Groupoids for which the action on its unit space is free (that is, for which the
map s × r : G1 → G0 × G0 is injective) are often called principal (see [30]). This
terminology conflicts, however, with the usual notion of a principal bundle, which
requires extra topological conditions besides freeness of the action.
We call a groupoid basic if its canonical action on the object space is basic, that
is, the map s × r : G1 → G0 × G0 is a homeomorphism onto its image.
Proposition A.7. A groupoid action is free and proper if and only if it is basic
and has Hausdorff orbit space.
If G and H are topological groupoids with Hausdorff object spaces, then an equiva-
lence from H to G in our sense is the same as a topological space X with commuting
free and proper actions of G and H, such that the anchor maps induce homeomor-
phisms G\X ∼= H 0 and X/H ∼= G0.
50
ALCIDES BUSS AND RALF MEYER
Proof. The characterisation of free and proper actions is [21, Corollary 9.32]; the
main point of the proof is that the orbit space is Hausdorff if and only if the orbit
equivalence relation is closed in X ×X (Proposition A.3). The left and right actions
on an equivalence are basic with X/H ∼= G0 and G\X ∼= H 0; hence they are free
and proper if and only if G0 and H 0 are Hausdorff, respectively. Conversely, if the
actions of G and H on X are free and proper, then both actions are basic, and both
anchor maps are open because they are equivalent to orbit space projections; thus
we have an equivalence in our sense.
(cid:3)
For a general action of a groupoid G on a space X, the image of the map (19)
is the orbit equivalence relation X ×X/G X ⊆ X × X. Thus the map (19) is a
homeomorphism (the action is basic) if and only if the action is free and the map
that sends (x1, x2) ∈ X ×X/G X to the unique g ∈ G1 with s(x1) = r(g) and
x1 · g = x2 is continuous.
If G, H and X are locally compact Hausdorff, then an equivalence in our sense
is the same as a (G, H)-equivalence in the notation of [23]; the main result of [23]
is that such an equivalence induces a Morita equivalence between the groupoid
C∗-algebras of G and H (for any Haar systems).
For non-Hausdorff groupoids, Jean-Louis Tu defined a notion of equivalence
in [36], using a technical variant of proper actions: he calls a groupoid G ρ-proper
with respect to a G-invariant continuous map ρ: G0 → T if the map
(r, s): G1 → G0 ×ρ,T,ρ G0,
g 7→ (r(g), s(g)),
is proper. If T is non-Hausdorff, then G0 ×ρ,T,ρ G0 need not be closed in G0 ×G0, so
that this is weaker than properness. In the definition of equivalence, he takes ρ to
be the anchor map on the other side, so he requires the maps in (E3) to be proper.
These maps are continuous bijections because the actions are free. A continuous,
proper bijection, being closed, must be a homeomorphism. Thus Tu's notion of
equivalence is equivalent to ours.
A.3. Covering groupoids and equivalence.
Definition A.8. Let f : X → Z be a continuous, open surjection. The covering
groupoid G(f ) has object space X, arrow space X ×f,Z,f X, range and source maps
r(x1, x2) := x1, s(x1, x2) := x2, and multiplication (x1, x2) · (x2, x3) := (x1, x3) for
all x1, x2, x3 ∈ X with f (x1) = f (x2) = f (x3).
The assumption on f implies that it is a quotient map, that is, we may identify Z
with the quotient space X/∼ by the following equivalence relation: x ∼ y if and
only if f (x) = f (y); and f becomes the quotient map X → X/∼. The covering
groupoid G(f ) is the groupoid associated to this equivalence relation. In particular,
Z can be identified with the orbit space X/G(f ) for the canonical action of G(f )
on its unit space X.
Every covering groupoid is basic, that is, its action on the unit space is basic.
Conversely, if G is a basic groupoid, then it is isomorphic to a covering groupoid.
The map r×s: G1 → G0×G0 gives a homeomorphism from G1 onto G0×f,G0/G,f G0,
where f : G0 → G0/G denotes the quotient map. This yields an isomorphism of
topological groupoids G ∼= G(f ).
Example A.9 (Čech groupoids). Let Z be a topological space and let U be an open
covering of Z. Let X := FU∈U U and let f : X → Z be the canonical map: f is
the inclusion map on each U ∈ U. This map is an open surjection. It is even étale,
that is, a local homeomorphism. We denote the covering groupoid of f by GU and
call it the Čech groupoid of the covering.
INVERSE SEMIGROUP ACTIONS ON GROUPOIDS
51
Assume that Z is locally Hausdorff and choose the open covering U to consist
of Hausdorff open subsets U ⊂ Z. Then the Čech groupoid GU is a Hausdorff,
étale topological groupoid (see also [8, Lemma 4.2]). If, in addition, Z is locally
quasi-compact, then GU is a (Hausdorff) locally compact, étale groupoid. This is
the situation we are mainly interested in.
Proposition A.10. Let fi : Xi → Z for i = 1, 2 be two continuous, open surjec-
tions. Then X1×f1,Z,f2 X2 with the obvious left and right actions of G(f1) and G(f2)
gives an equivalence from G(f2) to G(f1).
Proof. This is [21, Example 6.4].
(cid:3)
If G(f1) and G(f2) are Hausdorff locally compact, then so is the equivalence
X1 ×f1,Z,f2 X2 between them. If the maps f1 and f2 are both étale -- for instance,
if they come from open coverings of Z -- then the groupoids G(f1) and G(f2) are
étale, and the anchor maps X1 ← X1 ×f1,Z,f2 X2 → X2, x1 ← (x1, x2) → x2, are
étale as well.
Proposition A.11. The covering groupoid G(f ) of a continuous open surjection
f : X → Z is always equivalent (as a topological groupoid) to the space Z viewed as
a groupoid with only identity arrows. In particular, the Čech groupoid of a covering
of Z is equivalent to Z.
Conversely, if X is an equivalence from a space Z to a topological groupoid G,
then G is isomorphic to the covering groupoid of the anchor map s: X → Z.
Hence covering groupoids are exactly the groupoids that are equivalent to spaces.
Proof. The first part is a consequence of Proposition A.10 applied to f1 = f and
f2 = IdZ (see also [21, Example 6.3]). For the second part, observe that the action
of Z on X is simply the anchor map s: X → Z, which must be an open surjection.
The anchor map r : X → G0 must be a homeomorphism (because it must be the
projection map X → Z\X = X), so we may as well assume X = G0. Then
G1 ×s,G0,r X ∼= G1, and the first isomorphism in (E3) identifies G1 with X ×s,Z,s X.
This yields an isomorphism from G to the covering groupoid G(s) of s: X → Z. (cid:3)
Let Z be a space, view Z as a groupoid with only identity arrows. When is Z
equivalent to a locally compact, Hausdorff groupoid? If Z is equivalent to a topo-
logical groupoid G, then G is necessarily the covering groupoid G(f ) of a cover
f : X → Z by Proposition A.11.
Given a space Z, we thus seek a locally compact, Hausdorff space X and an
open, continuous surjection f : X → Z such that X ×f,Z,f X is locally compact.
The question when X ×f,Z,f X is locally compact is also asked in [8] at the end of
Section 4. We answer this question in Proposition A.14 below: X ×f,Z,f X is locally
compact if and only if Z is locally Hausdorff. Proposition A.16 says that the only
topological spaces Z that are equivalent to locally compact Hausdorff groupoids
are the locally Hausdorff, locally quasi-compact ones; for them, Example A.9 gives
such an equivalence, where the groupoid is even étale. We need some preparation
in order to prove Proposition A.14.
Definition A.12 ([2, I.3.3, Définition 2, Proposition 5]). A subset S of a topolog-
ical space X is locally closed if it satisfies the following equivalent conditions:
(1) any x ∈ S has a neighbourhood U such that S ∩ U is relatively closed in U ;
(2) S is open in its closure;
(3) S is an intersection of an open and a closed subset of X.
The following proposition generalises [2, I.9.7, Propositions 12 and 13] to the
locally Hausdorff case.
52
ALCIDES BUSS AND RALF MEYER
Proposition A.13. A subset S of a locally Hausdorff, locally quasi-compact space X
is locally quasi-compact in the subspace topology if and only if it is locally closed.
Proof. First let S be locally closed. Write S = A∩U with A closed and U open in X.
Let x ∈ S. Since X is locally quasi-compact, the quasi-compact neighbourhoods
of x in X form a neighbourhood basis of X. Since x ∈ U , those quasi-compact
neighbourhoods of x that are contained in U form a neighbourhood basis in U .
Their intersections with A remain quasi-compact because A is closed in X. They
form a neighbourhood basis of x in S, proving that S is locally quasi-compact.
Conversely, assume that S is locally quasi-compact in the subspace topology.
Let x ∈ S. Let U be a Hausdorff open neighbourhood of x in X. Then S ∩ U is
a neighbourhood of x in S and hence contains a quasi-compact neighbourhood K
of x in S because S is locally quasi-compact. We have K = S ∩ V for some
neighbourhood V of x in X, and we may assume V ⊆ U because K ⊆ U . The
subset S ∩ V is relatively closed in V because U ⊇ V is Hausdorff and S ∩ V is
quasi-compact. Thus S is locally closed.
(cid:3)
Proposition A.14. Let f : X → Z be a continuous, open surjection. The equiva-
lence relation Xf,Z,f X ⊆ X × X defined by f is locally closed if and only if Z is
locally Hausdorff. In particular, if X is locally quasi-compact and locally Hausdorff,
then Xf,Z,f X is locally quasi-compact if and only if Z is locally Hausdorff.
Proof. Assume Z to be locally Hausdorff first. Let (x1, x2) ∈ X ×f,Z,f X and let
U ⊆ Z be a Hausdorff open neighbourhood of f (x1) = f (x2). Then f −1(U ) ⊆ X is
an open subset such that f : f −1(U ) → U is an open map onto a Hausdorff space.
Hence
f −1(U ) ×f,U,f f −1(U ) =(cid:0)X ×f,Z,f X(cid:1) ∩(cid:0)f −1(U ) × f −1(U )(cid:1)
is relatively closed in f −1(U ) × f −1(U ) by [21, Proposition 9.15]. Thus X ×f,Z,f X
is locally closed in X × X.
Conversely, assume X ×f,Z,f X to be locally closed in X × X. Let x ∈ X. Then
(x, x) has a neighbourhood in X × X so that X ×f,Z,f X restricted to it is relatively
closed. Shrinking this neighbourhood, we may assume that it is of the form U × U
for an open neighbourhood of x, by the definition of the product topology on X ×X.
The map f U : U → f (U ) is open, and (X ×f,Z,f X) ∩ (U × U ) = U ×f U ,f (U),f U U .
Since this is relatively closed by assumption, [21, Proposition 9.15] shows that f (U )
is Hausdorff. Since x was arbitrary, this means that Z is locally Hausdorff.
The last sentence follows from the first one and Proposition A.13.
(cid:3)
Corollary A.15. A topological space X is locally Hausdorff if and only if the
diagonal {(x, x) x ∈ X} is a locally closed subset in X × X.
Proof. Apply Proposition A.14 to the identity map.
(cid:3)
Proposition A.16. Let G be a locally quasi-compact, locally Hausdorff groupoid
and let X be a basic right G-action. Then X/G is locally quasi-compact and locally
Hausdorff.
If X is an equivalence from a space Z to G, then Z ∼= X/G is locally quasi-
compact and locally Hausdorff.
Proof. Since G and X are locally quasi-compact and locally Hausdorff, so is their
product X × G1. Since G0 is locally Hausdorff, the diagonal in G0 is locally closed
by Corollary A.15. The fibre product X ×s,G0,r G1 is the preimage of the diagonal
in G0 × G0 under the continuous map r × s: X × G1 → G0 × G0; hence X ×s,G0,r G1
is locally closed in X × G1. Thus X ×s,G0,r G1 is locally quasi-compact and locally
Hausdorff by Proposition A.13.
INVERSE SEMIGROUP ACTIONS ON GROUPOIDS
53
Since the G-action on X is basic, X ×s,G0,r G1 is homeomorphic to the subset
X ×X/G X ⊆ X × X. Now Proposition A.13 shows that X ×X/G X is locally closed
in X × X. Then X/G is locally Hausdorff by Proposition A.14. Since continuous
images of quasi-compact subsets are again quasi-compact, X/G is also locally quasi-
compact.
An equivalence from a space Z to G is the same as a basic G-action with a
homeomorphism X/G ∼= Z. If this exists, then Z must be locally Hausdorff and
locally quasi-compact by the above argument.
(cid:3)
Appendix B. Fields of Banach spaces over locally Hausdorff spaces
Let X be a locally quasi-compact, locally Hausdorff space. Thus any Hausdorff
open subset of X is locally compact.
Definition B.1 (see [25] and the references there). An upper semicontinuous field
of Banach spaces on X is a family of Banach spaces (Bx)x∈X with a topology
on B = Fx∈X Bx such that, for each Hausdorff open subset U of X, BU is an
upper semicontinuous field of Banach spaces on U . In particular, the norm of any
continuous section of BU is an upper semicontinuous scalar-valued function on U .
Let S(U, B) denote the vector space of continuous, compactly supported sections
of BU . This is the union (hence inductive limit) of the subspaces S0(K, B) of
continuous sections on K vanishing on ∂K, where K runs through the directed set
of compact subsets of U and ∂K = K ∩ U \ K is the boundary of K in U . Each
S0(K, B) is a Banach space for the supremum norm
kf k∞ := sup{kf (x)k x ∈ K}.
We call a subset of S(U, B) bounded if it is the image of a norm-bounded subset of
S0(K, B) for some K.
If f ∈ S(U, B) for a Hausdorff open subset U of X, then we always extend f to
a section of B on all of X by taking f (x) := 0 for x /∈ U . Let S(X, B) be the vector
i=1 fi
for fi ∈ S(Ui, B) and Hausdorff open subset Ui of X. We call such sections of B
quasi-continuous.
space of all sections of B that may be written as finite linear combinationsPm
A subset A of S(X, B) is bounded if there are Hausdorff open subsets U1, . . . , Um
of X and bounded subsets Ai ⊆ S(Ui, B) for i = 1, . . . , m such that every element
To simplify our proofs, we use bornological language, that is, we speak of bounded
instead of open subsets. For a Hausdorff locally compact space X, S(X, B) with its
usual topology is an inductive limit of Banach spaces. The inductive limit topology
is determined by its continuous seminorms. A seminorm is continuous if and only
if it is bounded in the sense that its supremum over each bounded subset is finite;
this is so because a seminorm on a Banach space is continuous if and only if it is
bounded. For locally Hausdorff X, the bounded seminorms are those that restrict
to bounded seminorms on all the subspaces S(U, B) for U ⊆ X open and Hausdorff;
this is the same as the quotient topology from the map LU S(U, B) → S(X, B),
where U runs through the Hausdorff open subsets of X. Thus the usual topology on
S(X, B) -- which is the quotient topology induced by the inductive limit topologies
on the direct sums of the spaces S(U, B) -- is the topology generated by all bounded
seminorms.
Let U be a family of open subsets of X with the following two properties:
(1) X =SU∈U U , that is, for each x ∈ X there is U ∈ U with x ∈ U ;
(2) U1 ∩U2 =S{U ∈ U U ⊆ U1 ∩U2} for all U1, U2 ∈ U; that is, if x ∈ U1 ∩U2,
then there is U ∈ U with U ⊆ U1 ∩ U2 and x ∈ U .
of A may be written as a sum Pm
i=1 fi with fi ∈ Ai for i = 1, . . . , m.
54
ALCIDES BUSS AND RALF MEYER
In our main application, the open subsets in U will not be Hausdorff. Thus S(U, B)
for U ∈ U is defined in the same way as S(X, B), by taking finite linear combina-
tions of continuous compactly supported sections on Hausdorff open subsets of U .
We view S(U, B) as a subspace in S(X, B) by extending functions on U by 0 out-
side U . This gives an injective, bounded linear map S(U, B) → S(X, B). Being
bounded means that it maps bounded subsets to bounded subsets.
Let ιU : S(U, B) → LU∈U S(U, B) for U ∈ U denote the inclusion map of the
U -summand. We call a subset A of LU∈U S(U, B) bounded if there are finitely
of A may be written as Pm
many U1, . . . , Um ∈ U and bounded subsets Ai of S(Ui, B) such that any element
i=1 ιUi(fi) with fi ∈ Ai.
Proposition B.2. The map
E : MU∈U
S(U, B) → S(X, B)
is bounded linear and a bornological quotient map in the sense that any bounded
subset of S(X, B) is the image of a bounded subset ofLU∈U S(U, B); in particular,
it is surjective.
The kernel of E is the closed linear span of the set of elements of the form
ιU (f ) − ιV (f ) for f ∈ S(U, B), U, V ∈ U with U ⊆ V .
The "closure" in the description of the kernel is the bornological one, defined
using Mackey's notion of convergence in a bornological vector space. For any el-
ement g ∈ ker E, we will find a bounded subset A ⊆ LU∈U S(U, B), and linear
combinations gn of ιU (f ) − ιV (f ) for f ∈ S(U, B), U, V ∈ U with U ⊆ V such that
g − gn ∈ 2−n · A. This implies convergence in any bounded seminorm.
Remark B.3. Proposition B.2 implies that E is a quotient map with respect to the
canonical topologies on the spaces involved. That is, a seminorm p on S(X, B) is
continuous if and only if p ◦ E is a continuous seminorm on LU∈U S(U, B). The
proof uses that continuity and boundedness are equivalent for seminorms on both
spaces and that E is a bornological quotient map. It seems inconvenient, however,
to prove this directly without bornological language.
Proof. In the proof, we abbreviate S(U ) := S(U, B) because the Banach space
bundle is fixed throughout. We first show that E is a bornological quotient map.
Let A ⊆ S(X) be bounded. By definition, there are finitely many Hausdorff
open subsets V1, . . . , Vm ⊆ X, compact subsets Ki ⊆ Vi and scalars Ci > 0 such
i=1 fi with fi ∈ S0(Ki) having kfik∞ ≤ Ci.
Since the subsets U ∈ U cover X, they cover the compact subset Ki. Since
compact spaces are paracompact, there is a finite subordinate partition of unity
(ψi,U )U∈U, that is, ψi,U : Ki → [0, 1] is continuous and has compact support Li,U
that any f ∈ A may be written as Pm
If fi ∈ S0(Ki), then fi ·ψi,U ∈ S0(Ki ∩Li,U ) ⊆ S(Vi ∩U ) and kfi ·ψi,U k∞ ≤ kfik∞.
i=1 fi with fi ∈ S0(Ki) having kfik∞ ≤ Ci, and then
contained in U ∩ Ki, only finitely many ψi,U are non-zero, and PU∈U ψi,U (x) = 1.
Now write f ∈ A first asPm
i=1PU∈U fi · ψi,U . This sum is still finite because only finitely many ψi,U are
asPn
non-zero for each i, and each summand fi · ψi,U runs through a bounded subset
of S(Vi ∩ U ) and hence of S(U ) because we have uniform control on the supports
supp fiψi,U ⊆ Ki ∩ Li,U and norms kfi · ψi,U k∞ ≤ Ci of the summands. Hence A
is contained in the E-image of a bounded subset in L S(U ).
Now we describe the kernel of E. Let N be the linear span of elements of the form
ιU (f ) − ιV (f ) for all f ∈ S(U ), U, V ∈ U with U ⊆ V . Since E(ιU (f ) − ιV (f )) = 0,
we have N ⊆ ker E. If U1, U2, V ∈ U satisfy V ⊆ U1 ∩ U2 and f ∈ S(V ), then
ιU1 (f ) − ιU2(f ) = −(ιV (f ) − ιU1(f )) + (ιV (f ) − ιU2(f )) ∈ N.
INVERSE SEMIGROUP ACTIONS ON GROUPOIDS
55
i
i
let K ◦
1, . . . , m be such that f = Pm
Pm
i
We are going to modify a given element of ker E by adding elements of N so that the
norms of its constituents become arbitrarily small, without enlarging their supports.
A generic element f ∈LU∈U S(U ) is of the form f =P ιU (fU ) with fU ∈ S(U )
and fU = 0 for all but finitely many U . Each non-zero fU is a sum fU =PkU
j=1 fU,j
with fU,j ∈ S(VU,j) for finitely many Hausdorff open subsets VU,1, . . . , VU,kU ⊆
U . We renumber the finitely many Hausdorff open subsets VU,j consecutively as
V1, . . . , Vm and relabel our sections fi ∈ S(Vi) accordingly. Let Ui ∈ U for i =
i=1 ιUi(fi); so Vi ⊆ Ui. Let Ki := supp fi ⊆ Vi and
i be the interior of Ki inside Vi; thus x ∈ K ◦
i for all x ∈ X with fi(x) 6= 0.
) with f (0) = f , f (j+1) − f (j) ∈ N , and kf (m)
Now assume f ∈ ker(E) and let ǫ > 0. We will construct a finite sequence f (j) =
i=1 ιUi(f (j)
k < ǫ for all i = 1, . . . , m.
Furthermore, our construction ensures that the support of f (j)
is contained in Ki
for all i, j. Letting ǫ run through a sequence going to 0, the differences f −f (m) in N
will converge to f in the sense explained above because each constituent fi − f (m)
converges to fi in the normed space S0(Ki). Our construction will be such that
i = f (i)
f (j)
for j ≥ i, that is, in the jth step we keep f1, . . . , fj−1 fixed. To make
the following steps possible, we aim for stronger norm estimates kf (j)
i k < 2j−mǫ.
) with f − f (j) ∈ N
i k < 2j−mǫ for i = 1, . . . , j; for j = 0, this is satisfied for f (0) = f . We
) with f (j) − f (j+1) ∈ N and hence
Assume that we have already constructed f (j) =Pm
are going to construct f (j+1) =Pm
i=1 ιUi(f (j+1)
for i = 1, 2, . . . , j, and kf (j+1)
f − f (j+1) ∈ N , with f (j+1)
i=1 ιUi(f (j)
and kf (j)
= f (j)
j+1(x)k ≥ 2j+1−mǫ}. This is a closed subset of K ◦
j+1 k < 2j+1−mǫ.
j+1
because the norm function is upper semicontinuous. Since Kj+1 is compact, Aj+1
(x) = 0 for all
Let Aj+1 = {x ∈ Vj+1 kf (j)
x ∈ X. If x ∈ Aj+1, then this gives
is compact. Since E(f ) = 0 and E(f − f (j)) = 0, we have Pm
jXi=1
j+1(x)k −
j+1Xi=1
f (j)
i
f (j)
i
i=1 f (j)
kf (j)
i k∞ > 0.
i
i
i
i
i
(cid:13)(cid:13)(cid:13)(cid:13)
mXi=j+2
(x)(cid:13)(cid:13)(cid:13)(cid:13) =(cid:13)(cid:13)(cid:13)(cid:13)
(x)(cid:13)(cid:13)(cid:13)(cid:13) ≥ kf (j)
i
i
i
Hence there must be i > j + 1 with f (j)
i . Thus the open
subsets K ◦
i , then x ∈ Ui ∩ Uj+1. By our
assumption on U, there is U ∈ U with x ∈ U and U ⊆ Ui ∩ Uj+1. Thus the open
subsets K ◦
i ∩ U for i > j + 1 and U ∈ U with U ⊆ Ui ∩ Uj+1 cover Aj+1.
i for i > j + 1 cover Aj+1. If x ∈ Aj+1 ∩ K ◦
(x) 6= 0, so that x ∈ K ◦
Since Aj+1 is compact and contained in the Hausdorff locally compact space
Vj+1, there is a subordinate finite partition of unity (ψi,U ). That is, all but finitely
many ψi,U are non-zero, ψi,U : Aj+1 → [0, 1] is a continuous function with compact
support contained in K ◦
non-zero ψi,U from Aj+1 to a continuous function ¯ψi,U : Kj+1 → [0, 1] vanishing in a
neighbourhood of ∂Kj+1 and on Kj+1 \(K ◦
i ∩U ) because these two compact subsets
of Aj+1 are disjoint from the compact support of ψi,U in K ◦
i ∩ U . If necessary, we
i ∩U , andP ψi,U (x) = 1 for x ∈ Aj+1. We may extend each
multiply all ¯ψi,U with a suitable cut-off function so that P ¯ψi,U (x) ≤ 1 for all
x ∈ Kj+1.
Now we let
f (j+1) = f (j) +Xi,U
ιUi (f (j)
j+1
¯ψi,U ) − ιUj+1 (f (j)
j+1
¯ψi,U ).
By construction, f (j)
j+1
K ◦
f (j+1) − f (j) ∈ N as desired. Since only i > j + 1 appear in the sum, f (j+1)
¯ψi,U is continuous and supported in a compact subset of
¯ψi,U ) ∈ N , so
= f (j)
i ∩ U with U ⊆ Ui ∩ Uj+1, U ∈ U. Hence ιUi(f (j)
¯ψi,U ) − ιUj+1 (f (j)
j+1
j+1
i
i
56
ALCIDES BUSS AND RALF MEYER
for i < j + 1. We get
¯ψi,U (x)(cid:19).
This has supremum norm less than 2j+1−mǫ because 1−Pi,U
j+1(x) ·(cid:18)1 −Xi,U
j+1 (x) = f (j)
f (j+1)
¯ψi,U (x) vanishes where
is
j+1(x)k ≥ 2j+1−mǫ and is at most 1 everywhere else. The support of f (j+1)
kf (j)
still contained in Kj+1 by construction.
j+1
For i > j + 1, we get
f (j+1)
i
= f (j)
i +XU
j+1 · ¯ψi,U .
f (j)
This still has support Ki because ¯ψi,U is supported there. This completes the
induction step and thus the proof.
(cid:3)
Remark B.4. If X is Hausdorff, then a partition-of-unity argument as in the proof
of [4, Theorem 2.13] shows that ker(E) is the linear span without closure of ιU (f ) −
ιV (f ) with U, V ∈ U. Hence this linear span is already closed for the natural
topology onLU∈U S(U, B). Convergent infinite series are needed to generate ker E
from ιU (f ) − ιV (f ) with U, V ∈ U. This happens in simple examples, such as the
space X = [0, 1] ⊔(0,1] [0, 1] discussed in Section 8 with the trivial bundle C and the
standard open cover by two Hausdorff open subsets with their intersection (0, 1].
B.1. Proof of Theorem 5.5. We apply Proposition B.2 to X = L, the cover
(Lt)t∈S, and the given Fell bundle B as in the statement of Theorem 5.5. The
subsets B∗ and B1∗B2 for bounded subsets B, B1, B2 ⊆ S(L, B) are again bounded;
this is routine to check. Thus S(L, B) is a bornological ∗-algebra. (The continuity
of the operations for the "inductive limit topology" is also known but somewhat
more difficult.)
We are going to cite some results of [31] below, which follow from the Disintegra-
tion Theorem and the Morita Equivalence Theorem. We assume that they hold for
the Fell bundle B in question and its restriction to G; this is not yet proved in the
literature, see the discussion before Theorem 5.5. Remark B.6 sketches a slightly
more complicated proof that uses only the Morita Equivalence Theorem, that is,
the assumptions in Theorem 5.5.
Lemma B.5. The C∗-algebra C∗(L, B) is the completion of S(L, B) in the maxi-
mal bounded C∗-seminorm.
Proof. Usually, C∗(L, B) is defined as the completion of S(L, B) in the maxi-
mal C∗-seminorm that is bounded with respect to the I-norm, a certain norm
on S(L, B). [31, Corollaire 4.8] shows that a representation of S(L, B) that is con-
tinuous with respect to the "inductive limit topology" is bounded for the I-norm.
Hence a C∗-seminorm on S(L, B) is continuous with respect to the "inductive limit
topology" if and only if it is bounded with respect to the I-norm. The topology
on S(L, B) called "inductive limit topology" in [31] is really the quotient topology
induced by the inductive limit topology on LU∈U S(U, B), where U is the set of
all Hausdorff open subsets of L andLU∈U S(U, B) is viewed as the inductive limit
of the Banach subspaces LU∈F S0(KU , B) where F is a finite subset of U and
KU ⊆ U for U ∈ F are compact subsets. As we remarked above, a seminorm
is continuous in this sense if and only if it is bounded in the canonical bornology
on S(L, B) introduced in Appendix B.
(cid:3)
Let D := Lt∈S S(Lt, B). This carries a canonical direct sum bornology as in
Appendix B. The Fell bundle operations turn it into a ∗-algebra. The multiplication
INVERSE SEMIGROUP ACTIONS ON GROUPOIDS
57
and involution are bounded, so we even have a bornological ∗-algebra. The map
E : D → S(L, B) from Proposition B.2 is a bounded ∗-homomorphism.
Since E is a bornological quotient map by Proposition B.2, a C∗-seminorm p
on S(L, B) is bounded if and only if p ◦ E is a bounded C∗-seminorm on D. A
bounded C∗-seminorm on D is of the form p ◦ E for a C∗-seminorm p on S(L, B) if
and only if it vanishes on the kernel of E. By Proposition B.2, a bounded seminorm
on D vanishes on ker E if and only if it vanishes on ιt(f )−ιu(f ) for all f ∈ S(Lt, B),
t, u ∈ S, t ≤ u. Thus C∗(L, B) is isomorphic to the completion of D in the maximal
C∗-seminorm q on D that is bounded and vanishes on ιt(f ) − ιu(f ) for all f, t, u as
above.
The restriction of this C∗-seminorm q to S(G, B) ⊆ D is a bounded C∗-seminorm.
Since C∗(G, B) is defined as the completion of S(G, B) with respect to the maxi-
mal bounded C∗-seminorm on S(G, B), q extends to a C∗-seminorm on C∗(G, B).
Since q(f )2 = q(f ∗ ∗ f ) for f ∈ S(Lt, B), the restriction of q to S(Lt, B) is domi-
nated by the Hilbert module norm from C∗(Lt, B). Thus q automatically extends
f ∈ C∗(Lt, B), t, u ∈ S, t ≤ u because S(Lt, B) is dense in C∗(Lt, B). Conversely, a
t, u ∈ S, t ≤ u restricts to a C∗-seminorm q on D that annihilates ιt(f ) − ιu(f )
to the sumLt∈S C∗(Lt, B)t∈S. Furthermore, q still annihilates ιt(f ) − ιu(f ) for all
C∗-seminorm onLt∈S C∗(Lt, B) that annihilates ιt(f )−ιu(f ) for all f ∈ C∗(Lt, B),
for all f ∈ S(Lt, B), t, u ∈ S, t ≤ u. Since D is dense in Lt∈S C∗(Lt, B), this
says that C∗(L, B) is isomorphic to the completion of Lt∈S C∗(Lt, B) in the max-
imal C∗-seminorm that annihilates ιt(f ) − ιu(f ) for all f ∈ C∗(Lt, B), t, u ∈ S,
t ≤ u. This is exactly the definition of the section C∗-algebra of the Fell bundle
C∗(Lt, B)t∈S over S. This finishes the proof of Theorem 5.5.
Remark B.6. We may also prove Theorem 5.5 without Lemma B.5, using the usual
definition of C∗(L, B) involving the I-norm on D. This variant of the proof has
the advantage that it does not require the Disintegration Theorem. We still need
the Morita equivalence theorem for our Fell bundles, however, so that our inner
products are positive and generate the expected ideals.
We only explain the new points in this alternative proof. The I-norm on S(L, B)
restricts to the I-norm on S(G, B). Consider a C∗-seminorm q on D that annihi-
lates ιt(f ) − ιu(f ) for all f ∈ S(Lt, B), t, u ∈ S, t ≤ u and satisfies q(f ) ≤ kf kI
for all f ∈ S(G, B). Then q(f ) = q(f ∗ ∗ f )1/2 ≤ kf ∗ ∗ f k1/2
I ≤ kf kI for all
f ∈ S(Lt, B), t ∈ S. Thus q is bounded with respect to our bornology as well, so
it factors as q ◦ E for a bounded seminorm q on S(L, B) by Proposition B.2. This
seminorm satisfies q(f ) ≤ kf kI for all f ∈ S(Lt, B), t ∈ S. But then q(f ) ≤ kf kI
follows for all f ∈ S(L, B), t ∈ S.
References
[1] Jean Bénabou, Introduction to bicategories, Reports of the Midwest Category Seminar,
Springer, Berlin, 1967, pp. 1 -- 77, doi: 10.1007/BFb0074299. MR 0220789
[2] Nicolas Bourbaki, Topologie générale. Chapitres 1 à 4, Éléments de mathématique, Hermann,
Paris, 1971. MR 0358652
[3] Alcides Buss and Ruy Exel, Twisted actions and regular Fell bundles over inverse semigroups,
Proc. Lond. Math. Soc. (3) 103 (2011), no. 2, 235 -- 270, doi: 10.1112/plms/pdr006.MR 2821242
group-
at
oids,
http://www.theta.ro/jot/archive/2012-067-001/2012-067-001-007.pdf. MR 2881538
semigroups
no.
twisted
153 -- 205,
étale
available
over
Theory
inverse
67
Fell
bundles
and
1,
(2012),
[4]
,
J.
Operator
[5] Alcides Buss and Ralf Meyer, Crossed products for actions of crossed modules on C∗-algebras,
J. Noncommut. Geom. (2016), accepted. arXiv: 1304.6540.
[6] Alcides Buss, Ralf Meyer, and Chenchang Zhu, A higher category approach to twisted
(2) 56 (2013), no. 2, 387 -- 426,
actions on C∗-algebras, Proc. Edinb. Math. Soc.
doi: 10.1017/S0013091512000259. MR 3056650
58
ALCIDES BUSS AND RALF MEYER
[7] Jérôme Chabert and Siegfried Echterhoff, Twisted equivariant KK-theory and the
Baum -- Connes conjecture for group extensions, K-Theory 23 (2001), no. 2, 157 -- 200,
doi: 10.1023/A:1017916521415. MR 1857079
[8] Lisa Orloff Clark, Astrid an Huef, and Iain Raeburn, The equivalence relations of local
homeomorphisms and Fell algebras, New York J. Math. 19 (2013), 367 -- 394, available at
http://nyjm.albany.edu/j/2013/19_367.html. MR 3084709
[9] Valentin Deaconu, Groupoids associated with endomorphisms, Trans. Amer. Math. Soc. 347
(1995), no. 5, 1779 -- 1786, doi: 10.2307/2154972. MR 1233967
[10] Claire Debord, Holonomy groupoids of singular foliations, J. Differential Geom. 58 (2001),
no. 3, 467 -- 500, available at http://projecteuclid.org/euclid.jdg/1090348356.MR 1906783
[11] Ruy Exel, Partial actions of groups and actions of inverse semigroups, Proc. Amer. Math.
Soc. 126 (1998), no. 12, 3481 -- 3494, doi: 10.1090/S0002-9939-98-04575-4. MR 1469405
[12]
[13]
, Inverse semigroups and combinatorial C ∗-algebras, Bull. Braz. Math. Soc. (N.S.) 39
(2008), no. 2, 191 -- 313, doi: 10.1007/s00574-008-0080-7. MR 2419901
, Noncommutative Cartan subalgebras of C ∗-algebras, New York J. Math. 17 (2011),
331 -- 382, available at http://nyjm.albany.edu/j/2011/17-17.html. MR 2811068
[14] Philip Green, The local structure of twisted covariance algebras, Acta Math. 140 (1978),
no. 3-4, 191 -- 250, doi: 10.1007/BF02392308. MR 0493349
[15] Rohit Dilip Holkar, Topological construction of C∗-correspondences for groupoid C∗-algebras,
Ph.D. Thesis, Georg-August-Universität Göttingen, 2014.
[16] Gennadi G. Kasparov and Georges Skandalis, Groups acting on buildings, operator K-theory,
and Novikov's conjecture, K-Theory 4 (1991), no. 4, 303 -- 337, doi: 10.1007/BF00533989.
MR 1115824
[17] Alex Kumjian, Fell bundles over groupoids, Proc. Amer. Math. Soc. 126 (1998), no. 4, 1115 --
1125, doi: 10.1090/S0002-9939-98-04240-3. MR 1443836
[18] Mark V. Lawson, Inverse semigroups: the theory of partial symmetries, World Scientific
Publishing Co., River Edge, NJ, 1998.
[19] Tom Leinster, Basic Bicategories (1998), eprint. arXiv: math/9810017.
[20] Ralf Meyer
bootstrap
215 -- 252,
the
http://nbn-resolving.de/urn:nbn:de:hbz:6-10569452982. MR 2545613
and
Ryszard
class, Münster
Nest,
J. Math.
C ∗-Algebras
(2009),
over
2
topological
available
spaces:
at
[21] Ralf Meyer
and
topology,
http://www.tac.mta.ca/tac/volumes/30/55/30-55abs.html.
Theory
Categ.
30
Chenchang
Appl.
Zhu,
Groupoids
(2015),
categories
in
1906 -- 1998,
with
available
pre-
at
[22] Ieke Moerdijk, Orbifolds as groupoids: an introduction, Orbifolds in mathematics and physics
(Madison, WI, 2001), Contemp. Math., vol. 310, Amer. Math. Soc., Providence, RI, 2002,
pp. 205 -- 222, doi: 10.1090/conm/310/5405. MR 1950948
[23] Paul S. Muhly, Jean N. Renault, and Dana P. Williams, Equivalence and isomor-
phism for groupoid C ∗-algebras, J. Operator Theory 17 (1987), no. 1, 3 -- 22, available at
http://www.theta.ro/jot/archive/1987-017-001/1987-017-001-001.pdf. MR 873460
[24] Paul S. Muhly and Dana P. Williams, Equivalence and disintegration theorems for Fell
bundles and their C ∗-algebras, Dissertationes Math. (Rozprawy Mat.) 456 (2008), 1 -- 57,
doi: 10.4064/dm456-0-1. MR 2446021
[25]
, Renault's equivalence theorem for groupoid crossed products, NYJM Monographs,
vol. 3, State University of New York University at Albany, Albany, NY, 2008. available at
http://nyjm.albany.edu/m/2008/3.htm MR 2547343
[26] May Nilsen, C ∗-bundles and C0(X)-algebras, Indiana Univ. Math. J. 45 (1996), no. 2, 463 --
477, doi: 10.1512/iumj.1996.45.1086. MR 1414338
[27] Radu Popescu, Equivariant E-theory for groupoids acting on C ∗-algebras, J. Funct. Anal.
209 (2004), no. 2, 247 -- 292, doi: 10.1016/j.jfa.2003.04.001. MR 2044224
[28] John Quigg and Nándor Sieben, C ∗-actions of r-discrete groupoids and inverse semigroups,
J. Austral. Math. Soc. Ser. A 66 (1999), no. 2, 143 -- 167, doi: 10.1017/S1446788700039288.
MR 1671944
[29] Iain Raeburn and Dana P. Williams, Morita equivalence and continuous-trace C ∗-algebras,
Mathematical Surveys and Monographs, vol. 60, American Mathematical Society, Providence,
RI, 1998. MR 1634408
[30] Jean Renault, A groupoid approach to C∗-algebras, Lecture Notes in Mathematics, vol. 793,
Springer, Berlin, 1980. doi: 10.1007/BFb0091072 MR 584266
[31]
[32]
,
Operator
Représentation
Theory
des
18
produits
(1987),
croisés
no.
J.
http://www.theta.ro/jot/archive/1987-018-001/1987-018-001-005.pdf. MR 912813
available
1,
d'algèbres
67 -- 97,
de
groupoïdes,
at
, Cartan subalgebras in C ∗-algebras, Irish Math. Soc. Bull. 61 (2008), 29 -- 63, available
at http://www.maths.tcd.ie/pub/ims/bull61/S6101.pdf. MR 2460017
INVERSE SEMIGROUP ACTIONS ON GROUPOIDS
59
[34]
,
[33] Nándor Sieben, C ∗-crossed products by partial actions and actions of inverse semigroups, J.
Austral. Math. Soc. Ser. A 63 (1997), no. 1, 32 -- 46, doi: 10.1017/S1446788700000306.MR 1456588
ac-
at
tions,
http://www.theta.ro/jot/archive/1998-039-002/1998-039-002-009.pdf. MR 1620499
C ∗-crossed
J. Operator
by
(1998),
twisted
no.
available
semigroup
2,
361 -- 393,
inverse
products
Theory
39
[35] Giorgio Trentinaglia, Tannaka duality for proper Lie groupoids, Ph.D. Thesis, Utrecht Uni-
versity, 2008. arXiv: 0809.3394.
[36] Jean-Louis
K-theory,
http://www.mathematik.uni-bielefeld.de/documenta/vol-09/26.html. MR 2117427
Non-Hausdorff
Math.
9
groupoids,
(2004),
proper
565 -- 597,
actions
available
Tu,
Doc.
and
at
[37] Alan Weinstein, Linearization of regular proper groupoids, J. Inst. Math. Jussieu 1 (2002),
no. 3, 493 -- 511, doi: 10.1017/S1474748002000130. MR 1956059
linearization, affin-
(4) 39 (2006), no. 5, 841 -- 869,
[38] Nguyen Tien Zung, Proper
groupoids and momentum maps:
ity, and convexity, Ann. Sci. École Norm. Sup.
doi: 10.1016/j.ansens.2006.09.002. MR 2292634
E-mail address: [email protected]
Departamento de Matemática, Universidade Federal de Santa Catarina, 88.040-900
Florianópolis-SC, Brazil
E-mail address: [email protected]
Mathematisches Institut, Georg-August-Universität Göttingen, Bunsenstrasse 3 -- 5,
37073 Göttingen, Germany
|
1706.00554 | 1 | 1706 | 2017-06-02T04:53:01 | Quantum groups, property (T), and weak mixing | [
"math.OA",
"math.FA",
"math.QA"
] | For second countable discrete quantum groups, and more generally second countable locally compact quantum groups with trivial scaling group, we show that property (T) is equivalent to every weakly mixing unitary representation not having almost invariant vectors. This is a generalization of a theorem of Bekka and Valette from the group setting and was previously established in the case of low dual by Daws, Skalsi, and Viselter. Our approach uses spectral techniques and is completely different from those of Bekka--Valette and Daws--Skalski--Viselter. By a separate argument we furthermore extend the result to second countable nonunimodular locally compact quantum groups, which are shown in particular not to have property (T), generalizing a theorem of Fima from the discrete setting. We also obtain quantum group versions of characterizations of property (T) of Kerr and Pichot in terms of the Baire category theory of weak mixing representations and of Connes and Weiss in term of the prevalence of strongly ergodic actions. | math.OA | math |
QUANTUM GROUPS, PROPERTY (T), AND WEAK MIXING
MICHAEL BRANNAN AND DAVID KERR
Abstract. For second countable discrete quantum groups, and more generally sec-
ond countable locally compact quantum groups with trivial scaling group, we show
that property (T) is equivalent to every weakly mixing unitary representation not
having almost invariant vectors. This is a generalization of a theorem of Bekka and
Valette from the group setting and was previously established in the case of low
dual by Daws, Skalsi, and Viselter. Our approach uses spectral techniques and is
completely different from those of Bekka -- Valette and Daws -- Skalski -- Viselter. By a
separate argument we furthermore extend the result to second countable nonuni-
modular locally compact quantum groups, which are shown in particular not to have
property (T), generalizing a theorem of Fima from the discrete setting. We also ob-
tain quantum group versions of characterizations of property (T) of Kerr and Pichot
in terms of the Baire category theory of weak mixing representations and of Connes
and Weiss in term of the prevalence of strongly ergodic actions.
1. Introduction
Introduced by Kazhdan in the 1960s for the purpose of showing that many lattices
are finitely generated, Property (T) has come to play a foundational role in the study
of rigidity in Lie groups, ergodic theory, and von Neumann algebras through work of
Margulis, Zimmer, Connes, Popa, and others [34, 5, 28]. Over the last twenty-five
years it has been extended in stages to the realm of quantum groups, first via Kac
algebras [15], then in the algebraic [3] and discrete [13, 24] settings, and finally in the
general framework of locally compact quantum groups as defined by Kusterman and
Vaes [10]. In one notable recent application, Arano showed in [1, 2] that the Drinfeld
double of a q-deformation of compact simple lie group has property (T) and that this
implies that the duals of these q-deformations have a central version of property (T),
a fact which has inspired progress in the theory of C∗-tensor categories and underpins
Popa and Vaes's construction of subfactors with property (T) standard invariant that
do not come from groups [29, 26].
By definition, a locally compact group G does not have property (T) if it admits
a unitary representation which does not have a nonzero invariant vector (ergodic-
ity) but does have a net of unit vectors which is asymptotically invariant on each
group element (having almost invariant vectors). Because ergodicity has poor per-
manence properties, it can be hard to leverage this definition so as to obtain global
information about the representation theory of a group without property (T), and
in particular to determine to what extent the kind of flexible behaviour exhibited by
1
2
MICHAEL BRANNAN AND DAVID KERR
amenable groups persists in this more general setting. Bekka and Valette provided
a remedy for this in the separable case by showing that one can equivalently replace
ergodicity above with weak mixing, which is characterized by the absence of nonzero
finite-dimensional subrepresentions, or alternatively by the ergodicity of the tensor
product of the representation with its conjugate [4]. This leads for example to a short
proof of a theorem of Wang that characterizes property (T) in terms of the isolation
of finite-dimensional representations in the spectrum ([4], Section 4) and a stream-
lined proof of the Connes -- Weiss characterization of property (T) in terms of strongly
ergodic probability-measure-preserving actions ([5], Section 6.3).
Using the fact that weak mixing is preserved under tensor products with arbitrary
representations, Kerr and Pichot applied the Bekka -- Valette theorem to show that if
a second countable locally compact group does not have property (T) then within the
set of all unitary representations of the group on a fixed separable infinite-dimensional
Hilbert space the weakly mixing ones form a dense Gδ in the weak topology [20],
strengthening a result of Glasner and Weiss that gave the same conclusion for ergodic
representations [14]. The idea is that any representation will approximately absorb a
representation with almost invariant vectors under tensoring (since locally it is as if
we were tensoring with the trival representation) and so such a tensor product will
be "close" to the original representation while also inheriting any properties of the
second one that are preserved under tensoring, such as weak mixing. By a similar
principle requiring a more subtle implementation, Kerr and Pichot also established
an analogous conclusion for the measure-preserving actions of the group on a fixed
standard atomless probability space.
Using the theory of positive-definite functions as in [16, 27], Daws, Skalsi, and Visel-
ter demonstrated in [11] that the conclusion of the Bekka -- Valette theorem also holds
for second countable discrete unimodular quantum groups with low dual, and as an ap-
plication they derive analogues of the Connes -- Weiss theorem and the representation-
theoretic Kerr -- Pichot theorem. Low dual is the rather restrictive assumption that
there is a bound on the dimensions of the irreducible representations of the quantum
group, and the authors of [11] wonder, somewhat pessimistically, whether it can be
removed. In the present paper we show that the Bekka -- Valette and Kerr -- Pichot the-
orems actually hold for all second countable discrete quantum groups, and even more
generally for all second countable locally compact quantum groups with trivial scal-
ing group (Theorems 4.8 and 4.9) as well as for all second countable nonunimodular
locally compact quantum groups (Theorem 6.3). The methods of Daws, Skalsi, and
Viselter can then also be applied to extend their version of the Connes -- Weiss theorem
to all second countable locally compact quantum groups with trivial scaling group
(Theorem 5.1).
Our approach is completely different from those of Daws -- Skalsi -- Viselter and Bekka --
Valette and consists in applying the quantum group version of Wang's characterization
of property (T) mentioned above in order to reduce the problem to a purely spectral
question concerning C∗-algebras. In Theorems 3.7 and 3.8 we prove that the following
QUANTUM GROUPS, PROPERTY (T), AND WEAK MIXING
3
hold for a separable unital C∗-algebra A and a fixed separable infinite-dimensional
Hilbert space H :
(i) if the spectrum of A contains no isolated finite-dimensional representations
then the set of weakly mixing unital representations of A on H is a dense Gδ,
and
(ii) if the set of finite-dimensional representations in the spectrum of A is nonempty
and contains only isolated points then the set of weakly mixing unital repre-
sentations of A on H is closed and nowhere dense.
A version of the argument establishing (i) for unitary representations of countable
discrete groups has also been included in the book [19] by Li and the second author.
Theorems 4.8 and 4.9 then follow from (i) and (ii) whenever Wang's characterization
of property (T) holds in the quantum group context, and this is known to be the case
when the scaling group is trivial (see Section 4). By a completely different argument
we also prove in Theorem 6.3 that the conclusions Theorems 4.8 and 4.9 are valid for
second countable nonunimodular locally compact quantum groups, which we show in
particular not to have property (T), generalizing a result of Fima from the discrete
case [13].
We begin in Section 2 by reviewing some of the basic theory of locally compact
quantum groups and their unitary representations as developed by Kustermans and
Vaes [22, 23, 30, 21]. In Section 3 we study weak mixing for C∗-algebra representa-
tions and establish the two key spectral results (i) and (ii) concerning separable unital
C∗-algebras.
In Section 4 we discuss weak mixing and property (T) for quantum
groups, record the quantum group incarnation of Wang's theorem, and then establish
our versions of the Bekka -- Valette and Kerr -- Pichot theorems. Section 5 contains the
Connes -- Weiss-type dynamical characterization of property (T). Finally, the nonuni-
modular case is treated in Section 6.
Acknowledgements. M.B. was partially supported by NSF grant DMS-1700267. D.K.
was partially supported by NSF grant DMS-1500593.
2. Preliminaries
For a C∗-algebra A we write M(A) for its multiplier algebra. A representation of
A is understood to mean a ∗-homomorphism from A into the C∗-algebra of bounded
linear operators on some Hilbert space. When working with tensor products of Hilbert
spaces H and K , we denote by Σ the tensor flip map from H ⊗ K to K ⊗ H . For
linear operators on multiple tensor products, we use leg notation. For example, if U
is a unitary operator on a Hilbert space tensor product H ⊗ K we write U13 for the
unitary operator on a Hilbert space tensor product of the form H ⊗ J ⊗ K which
is given by V (U ⊗ id)V −1 where V is the shuffle map H ⊗ K ⊗ J → H ⊗ J ⊗ K
defined on elementary tensors by ξ ⊗ ζ ⊗ κ 7→ ξ ⊗ κ ⊗ ζ, i.e., V = idH ⊗ Σ.
4
MICHAEL BRANNAN AND DAVID KERR
2.1. Locally compact quantum groups. Our main references for generalities on
locally compact quantum groups are [22, 23, 30]. Formally speaking, a (von Neumann
algebraic) locally compact quantum group is a von Neumann algebra with coassociative
coproduct and left and right Haar weights, but as usual we use the simple notation
G so that we can conveniently and suggestively refer to the various objects that are
canonically attached to it just as one does for locally compact groups, although there
is no longer anything like an underlying group. The von Neumann algebra itself is thus
written L∞(G), and the coproduct is a unital normal ∗-homomorphism ∆ : L∞(G) →
L∞(G)⊗L∞(G) satisfying the coassociativity condition
(∆ ⊗ id)∆ = (id ⊗ ∆)∆.
The left and right Haar weights are normal semifinite weights ϕ and ψ on L∞(G) such
that for every ω ∈ L∞(G)+
∗ one has
for all a ∈ L∞(G)+ with ϕ(a) < ∞ and
ϕ((ω ⊗ id)∆(a)) = ϕ(a)ω(1)
ψ((id ⊗ ω)∆(a)) = ψ(a)ω(1)
for all a ∈ L∞(G)+ with ψ(a) < ∞. The predual of L∞(G) is written as L1(G),
and becomes a completely contractive Banach algebra with respect to the convolution
product
ω1 ⋆ ω2 = (ω1 ⊗ ω2) ◦ ∆,
ω1, ω2 ∈ L1(G).
Associated to G is a canonical weakly dense sub-C∗-algebra of L∞(G), written
C0(G), which plays the role of the C∗-algebra of continuous functions vanishing at
infinity in the case of ordinary groups. We say that G is second countable if C0(G)
is separable. The coproduct restricts to a unital ∗-homomorphism ∆ : C0(G) →
M(C0(G) ⊗ C0(G)). The algebras C0(G) and L∞(G) are standardly represented on
the GNS Hilbert space L2(G) associated to the left Haar weight.
In the case of a
locally compact group, the notations L∞(G), L1(G), C0(G), and L2(G) have their
ordinary meaning.
k·k
There is a (left) fundamental unitary operator W on L2(G) ⊗ L2(G) which satisfies
the pentagonal relation W12W13W23 = W23W12 and unitarily implements the coproduct
∆ on L∞(G) via the formula ∆(x) = W ∗(1 ⊗ x)W . Using W one has C0(G) =
{(id ⊗ ω)W : ω ∈ B(L2(G))∗}
, and one can define the antipode of G as the (generally
only densely defined) linear operator S on C0(G) (or L∞(G)) satisfying the identity
(S⊗id)W = W ∗. The antipode admits a polar decomposition S = R◦τ−i/2 where R is
an antiautomorphism of L∞(G) (the unitary antipode) and {τt}t∈R is a one-parameter
group of automorphisms (the scaling group). In the case of a locally compact group, the
scaling group is trivial and the antipode is the antiautomorphism sending a function
f ∈ C0(G) to the function s 7→ f (s−1). Using the antipode S one can endow the
convolution algebra L1(G) with a densely defined involution by considering the norm-
dense subalgebra L1
♯ (G) of L1(G) consisting of all ω ∈ L1(G) for which there exists an
QUANTUM GROUPS, PROPERTY (T), AND WEAK MIXING
5
k·k
equivalent to C0(G) being a direct sum of matrix algebras.
cally identified with the original quantum group G. One says that a locally compact
ω♯ ∈ L1(G) with hω♯, xi = hω, S(x)∗i for each x ∈ D(S). It is known from [21] and
Section 2 of [23] that L1
♯ (G) is an involutive Banach algebra with involution ω 7→ ω♯
and norm kωk♯ = max{kωk,kω♯k}.
Associated to any locally compact quantum group G is its dual locally compact
quantum group bG, whose associated algebras, coproduct, and fundamental unitary are
⊆ B(L2(G)), L∞(bG) = C0(bG)′′,
given by C0(bG) = {(ω ⊗ id)W : ω ∈ B(L2(G))∗}
∆(x) = W ∗(1 ⊗ x) W , and W = ΣW ∗Σ. Then in fact W ∈ M(C0(G) ⊗ C0(bG)), and
the Pontryagin duality theorem asserts that the bidual quantum group bbG is canoni-
quantum group G is compact if C0(G) is unital, and discrete if bG is compact, which is
For a locally compact quantum group G, we can always assume that the left and right
Haar weights are related by ψ = ϕ◦ R, where R is the unitary antipode. If the left and
right Haar weights ϕ and ψ of G coincide then we say that G is unimodular. In general,
the failure of ψ to be left-invariant is measured by the modular element, which is a
strictly positive element δ affiliated with L∞(G) satisfying the identities ∆(δ) = δ ⊗ δ
and ψ(·) = ϕ(δ1/2 · δ1/2). Compact quantum groups are always unimodular, and the
corresponding Haar weight can always be chosen to be a state. Although discrete
groups are always unimodular, discrete quantum groups need not be. We recall that
a discrete quantum group G is said to be of Kac type (or a Kac algebra) if it is
unimodular, which is equivalent to the Haar state on bG being a trace.
2.2. Unitary representations.
Definition 2.1. A unitary representation of a locally compact quantum group G on
a Hilbert space H is a unitary U ∈ M(C0(G)⊗ K (H )) ⊆ B(L2(G)⊗ H ) such that
(∆ ⊗ id)(U) = U13U23.
In the above definition one can replace M(C0(G)⊗ K (H )) with the larger algebra
L∞(G)⊗B(H ), for if U is a unitary in the latter which satisfies (∆⊗ id)(U) = U13U23
then U automatically belongs to the former (see for example Theorem 4.12 of [6]).
Associated to a unitary representation U ∈ M(C0(G) ⊗ K (H )) is an adjointable
operator on the Hilbert module C0(G)⊗ H which we write using the boldface version
U of the symbol in question. The relation between U and U is given by
hU(a ⊗ ξ), b ⊗ ζi = b∗(id ⊗ ωξ,ζ)(U)a
for all a, b ∈ C0(G) and ξ, ζ ∈ H , where ωξ,ζ is the vector functional x 7→ hxξ, ζi.
Associated to G are two distinguished unitary representations, the one-dimensional
trivial representation 1G ∈ M(C0(G)) given by the unit of L∞(G), and the left regular
representation given by the fundamental unitary W ∈ M(C0(G) ⊗ C0(bG)).
Two unitary representations U ∈ M(C0(G)⊗K (H1)) and V ∈ M(C0(G)⊗K (H2))
of G are (unitarily) equivalent if there is a unitary isomorphism u : H1 → H2 such that
6
MICHAEL BRANNAN AND DAVID KERR
V = (id ⊗ Ad u)(U). A subrepresentation of a unitary representation U ∈ M(C0(G) ⊗
K (H )) is a unitary representation of the form Q = (1⊗P )U(1⊗P ) ∈ L∞(G)⊗B(H0)
where H0 is a closed subspace of H , P is the orthogonal projection of H onto H0,
and 1 ⊗ P commutes with U. In this case, we write Q ≤ U. The subrepresentation is
said to be finite-dimensional if H0 is finite-dimensional.
Let G be a unimodular locally compact quantum group. Let U ∈ M(C0(G) ⊗
K (H )) be a unitary representation of G. Write H for the conjugate of H , i.e.,
the Hilbert space which is the same as H as an additive group but with the scalar
multiplication (c, ξ) 7→ ¯cξ for c ∈ C and inner product hξ, ζiH = hζ, ξiH . Letting
T : B(H ) → B(H ) be the transpose map T (a)(ξ) = a∗(ξ), we define the conjugate
of U, written U, to be the unitary representation
(R ⊗ T )(U) ∈ M(C0(G) ⊗ K (H ))).
The tensor product of two unitary representations U ∈ M(C0(G) ⊗ K (H )) and
V ∈ M(C0(G) ⊗ K (K )) is the unitary representation
U ⊙ V := U12V13 ∈ M(C0(G) ⊗ K (H ⊗ K )) ⊆ L∞(G)⊗B(H )⊗B(K ).
There is a bijective correspondence between unitary representations U ∈ M(C0(G)⊗
♯ (G) → B(H) ([21], Corol-
K (H )) and nondegenerate ∗-representations πU : L1
lary 2.13). This correspondence is given by
πU (ω) = (ω ⊗ id)U ∈ B(H),
ω ∈ L1
♯ (G).
k·k
Let C u
♯ (G). This is a univer-
k·k
♯ (G))
and C0(G) = λ(L1
C0(bG) = λ(L1
At the level of ∗-representations of L1
♯ (G), the trivial representation 1G corresponds
to the ∗-character ω 7→ ω(1), and the left regular representation is written as ω 7→
λ(ω) = (ω ⊗ id)W ∈ C0(bG) ⊆ B(L2(G)). As expected, we have the dual relations
where λ is the left regular representation of bG.
sal version of C0(bG) which encodes the unitary representation theory of G (since
♯ (bG))
0 (bG) denote the universal enveloping C∗-algebra of L1
its nondegenerate representations are in bijective correspondence with the nonde-
generate ∗-representations of L1
In par-
ticular, the left regular representation W gives rise to a surjective representation
λ : C u
(dual) counit εu : C u
∗-representations of L1
0 (bG) → C0(bG) ⊆ B(L2(G)), and the trivial representation gives rise to the
0 (bG) → C. We will generally use the same symbols to denote
0 (bG))
0 (bG), ∆u) into a universal C∗-algebraic locally compact
which can be used to turn (C u
quantum group. For our purposes, we only need the fact that ∆u allows one to express
the tensor product U ⊙ V of two unitary representations U ∈ M(C0(G)⊗ K (H )) and
0 (bG) admits a coproduct ∆u : C u
0 (bG).
0 (bG) → M(C u
♯ (G) as bounded Hilbert space operators).
♯ (G) and their unique extensions to C u
0 (bG)⊗ C u
As was shown in [21], C u
QUANTUM GROUPS, PROPERTY (T), AND WEAK MIXING
7
V ∈ M(C0(G) ⊗ K (K )) in terms of the representation (πU ⊗ πV ) ◦ σ ∆u : C u
0 (bG) →
0 (bG) ⊗ C u
B(H ⊗ K ), where σ denotes the tensor flip map on C u
0 (bG) (or of ∗-representations
Finally, note that at the level of representations of C u
of L1
♯ (G)) the notions of subrepresentation and unitary equivalence of unitary repre-
sentations reduce to their standard meanings. Indeed, given a unitary representation
U ∈ M(C0(G)⊗K (H )) and a projection P in B(H ), the projection 1⊗P commutes
0 (bG)), in which case the representation
with U if and only if P commutes with πU (C u
0 (bG) → B(P H ) associated to Q = (1⊗ P )U(1⊗ P ) is given by a 7→ P πU (a)P .
πQ : C u
Similarly, if V ∈ M(C0(G)⊗K (K )) is another unitary representation, then a unitary
isomorphism u : H → K implements an equivalence between U and V if and only if it
implements a unitary equivalence between πU and πV in the sense that πV = Ad u◦πU .
0 (bG).
3. Weak mixing and representations of C∗-algebras
This section is purely C∗-algebraic and aims to establish two results concerning
the prevalence of weak mixing among unital representations of a separable unital C∗-
algebra on a fixed Hilbert space (Theorems 3.7 and 3.8).
Throughout this section A will denote a separable unital C∗-algebra. For a fixed
Hilbert space H , the set of all unital representations of A on H will be written
Rep(A, H ). We equip Rep(A, H ) with the point-strong operator topology, which is
equivalent to the point-weak operator topology, and also to the point-∗-strong operator
topology since the strong and ∗-strong operator topologies agree on the unitary group
of B(H ) and A is linearly spanned by its unitaries.
representations, will be thought of as actual representations via their representatives,
Points in the spectrum bA, while formally defined as equivalence classes of irreducible
following convention. The set of finite-dimensional representations in bA will be written
bAfin.
We begin with a discussion of weak mixing for unital representations of unital C∗-
algebras.
Definition 3.1. We say that a unital representation of A on a Hilbert space is weakly
mixing if it has no nonzero finite-dimensional subrepresentations.
Recall that weak mixing for a unitary representation π : G → U (H ) of a group
can be expressed in either of the following equivalent ways (see Theorem 2.23 in [19],
and note that the countability assumption there is not needed):
(i) for every finite set Ω ⊆ H and ε > 0, there exists an s ∈ G such that
(ii) π has no nonzero finite-dimensional subrepresentations.
hπ(s)ξ, ζi < ε for all ξ, ζ ∈ Ω,
Since a unital C∗-algebra is linearly spanned by its unitaries, from (ii) we immediately
obtain the following, justifying the terminology of Definition 3.1.
8
MICHAEL BRANNAN AND DAVID KERR
Proposition 3.2. A unital representation of A is weakly mixing if and only if its
restriction to the unitary group of A is weakly mixing.
Next we consider a C∗-algebra version of Zimmer's notion [34] of weak containment
for unitary representations of groups.
Definition 3.3. Let π : A → B(H ) and ρ : A → B(K ) be unital representations.
We write π ≺ ρ if for every finite set Ω ⊆ A, orthonormal set {ξ1, . . . , ξn} ⊆ H ,
and ε > 0 there is an orthonormal set {ζ1, . . . , ζn} ⊆ K such that hπ(a)ξi, ξii −
hρ(a)ζi, ζii < ε for all a ∈ Ω and i = 1, . . . , n.
This is the same as the usual notion of weak containment when the representation
π is irreducible, but is different in general. In fact π is weakly contained in ρ if and
only if π ≺ ρ⊕N.
The following is a straightforward consequence of Definition 3.3. The version for
unitary group representations was noted in the remark after Proposition H.2 in [18].
Lemma 3.4. Let π : A → B(H ) and ρ : A → B(K ) be unital representations
on separable infinite-dimensional Hilbert spaces. Then π ≺ ρ if and only if π ∈
{κ ∈ Rep(A, H ) : κ ∼= ρ}.
Denote by WM(A, H ) ⊆ Rep(A, H ) the subcollection of all weakly mixing repre-
sentations.
Lemma 3.5. Let H be a separable Hilbert space. Then WM(A, H ) is a Gδ in
Rep(A, H ).
Proof. Write G for the unitary group of A. Take an increasing sequence Ω1 ⊆ Ω2 ⊆ . . .
of finite subsets of H with dense union in H . For every n ∈ N define Γn to be the
set of all ϕ ∈ Rep(A, H ) such that there exists a u ∈ G satisfying hϕ(u)ξ, ζi < 1/n
for all ξ, ζ ∈ Ωn. Then Γn is open, and so the set Γ = T∞
n=1 Γn is a Gδ. By the
characterization of weak mixing for unitary group representations described before
Proposition 3.2, Γ is precisely the set of all representations whose restriction to G is
weakly mixing, which is equal to WM(A, H ) by Proposition 3.2.
(cid:3)
Lemma 3.6. Suppose that each point in bAfin is isolated in bA. Let ρ : A → B(H ) be
a representation in bAfin. Then there exists a weakly mixing representation θ of A on a
separable Hilbert space such that ρ ≺ θ.
Proof. We may assume that ρ is not the limit of a sequence {πn} of infinite-dimensional
representations in bA, for in that case the representation π =L∞
n=1 πn is weakly mixing
and ρ ≺ π. Since bA is second countable ([12], Proposition 3.3.4), we can then find
a countable neighbourhood basis {Un}n∈N for ρ in bA such that no Un contains an
Let n ∈ N. We construct a weakly mixing representation θn in bA as follows. First we
argue that Un is uncountable. Suppose that this is not the case. As bA is a Baire space
infinite-dimensional representation.
QUANTUM GROUPS, PROPERTY (T), AND WEAK MIXING
9
([12], Theorem 3.4.13) and Un is open, Un is itself a Baire space. For every ω ∈ Un
the singleton {ω} is closed by finite-dimensionality (see Section 3.6 of [12]), and so
by the Baire property there exists an ω0 ∈ Un such that {ω0} is open, which means
that ω0 is isolated, contradicting our hypothesis. Thus Un is uncountable. We can
consequently find a dn ∈ N such that Un contains uncountably many dn-dimensional
representations.
Fix a Hilbert space Hn of dimension dn. Denote by Irr(A, Hn) the set of irreducible
representations in Rep(A, Hn). We observe the following:
are equivalent to some element of U is open.
(i) Irr(A, Hn) is open in Rep(A, Hn),
(ii) every equivalence class in Irr(A, Hn) is closed in Irr(A, Hn),
(iii) for every open set U ⊆ Irr(A, Hn) the set of all elements in Irr(A, Hn) which
Assertion (iii) is clear. To verify (i), let {πk} be a convergent sequence in Rep(A, Hn)
whose terms are not irreducible and let us show that its limit π is not irreducible.
For every k choose a nonzero projection Pk ∈ πk(A)′ of rank less than dn. In view
of the finite-dimensionality of Hn, we may assume by passing to a subsequence that
the sequence {Pk} converges in B(Hn), in which case its limit P is a nonzero pro-
jection which commutes with π(A). This means that π is not irreducible, yielding
(i). Finally, to verify (ii) we let {πk} be a convergent sequence in Irr(A, Hn) such
that for every k there exists a unitary operator Zk which conjugates πk to π1. By the
finite-dimensionality of Hn, there is a subsequence {Zkj}j that converges to a unitary
operator Z, which must then conjugate the limit of {πk} to π1, yielding (ii).
Since open subsets of Polish spaces are themselves Polish spaces ([17], Theorem 3.11),
we infer from (i) that Irr(A, Hn) is a Polish space. Assertions (ii) and (iii) then per-
mit us to apply a standard selection theorem ([17], Theorem 12.16) which provides a
Borel set Bn ⊆ Irr(G, Hn) of representatives for the relation of unitary equivalence.
Write Wn for the set of all π ∈ Irr(A, Hn) which, as elements in bA, belong to Un.
This is clearly an open set in Irr(A, Hn), and it is uncountable by our choice of dn.
Thus Bn ∩ Wn is an uncountable Borel set, which means that it is isomorphic as a
Borel space to the unit interval with Lebesgue measure and hence admits an atomless
Borel probability measure of full support. Let µn be the push forward of this measure
under the inclusion Bn∩ Wn ֒→ Irr(A, Hn), and note that µn(C) = 0 for every unitary
equivalence class C in Irr(A, Hn),
Setting Yn = Irr(A, Hn), we next consider the Hilbert space L2(Yn, Hn) of (classes
of) Hn-valued functions on Yn with inner product
hf, gi =ZYnhf (π), g(π)i dµn(π).
Finally, we define the representation θn : A → B(L2(Yn, Hn)) by setting
(θn(a)ζ)(π) = π(a)ζ(π)
for a ∈ A, ζ ∈ L2(Yn, Hn), and π ∈ Yn.
10
MICHAEL BRANNAN AND DAVID KERR
We next verify that θn is weakly mixing. Suppose that this is not the case. Then
there is a finite-dimensional irreducible representation π : A → B(H ) and an isomet-
ric operator Z : H → L2(Yn, Hn) such that Zπ(a) = θn(a)Z for all a ∈ A. Then for
a.e. ρ ∈ Yn and every a ∈ A and η ∈ H we have
(Zπ(a)η)(ρ) = (θi(a)Zη)(ρ) = ρ(a)((Zη)(ρ)).
so that the operator Zρ : H → Hn given by Zρη = (Zη)(ρ) satisfies Zρπ(a) = ρ(a)Z
for all a ∈ A. As Z is isometric, the operator Zρ must be nonzero for all ρ in a nonull
subset of Yn. But each such ρ is equivalent to π by irreducibility, contradicting the
fact that the measure of every unitary equivalence class is zero. Therefore θn is weakly
mixing.
Now set θ = L∞
n=1 θn. Then θ is weakly mixing since each summand is weakly
It remains to show that ρ ≺ θ. Let Ω be a finite subset of A and ε >
mixing.
0. Then we can find an n ∈ N such that for every π ∈ Un there is an isometry
V : H → Hn such that kV ρ(a) − π(a)V k < ε/2 for all a ∈ Ω. Since bounded
sets in B(Hn) are precompact and representations are contractive, we can find an
open set U ⊆ Yn with µ(U) > 0 such that for all a ∈ Ω and π, π′ ∈ U one has
kπ(a) − π′(a)k < ε/2. Choose a π0 ∈ U and an isometry V : H → Hn such that
for all a ∈ Ω we have kV ρ(a) − π0(a)V k < ε/2 and hence kV ρ(a) − π(a)V k < ε for
every π ∈ U. Writing 1U for the indicator function of U, we set f = µn(U)−1/21U ,
which is a unit vector in L2(Yn, Hn). Define an isometry V : H → L2(Yn, Hn) by
V ξ = f ⊗ V ξ ∈ L2(Yn, µn) ⊗ Hn ∼= L2(Yn, Hn). Then for all a ∈ Ω and norm-one
vectors ξ ∈ H we have
k( V ρ(a) − θn(a) V )ξk2 = kf ⊗ V ρ(a)ξ − θn(a)(f ⊗ V ξ)k2
1
1
µn(U)ZU k(V ρ(a) − π(a)V )ξk2 dµn(π)
µn(U)ZU
ε2 dµn(π)
=
≤
= ε2,
so that k V ρ(a) − θn(a) V k < ε. We conclude that ρ ≺ θ, as desired.
We now come to the main theorems of this section.
(cid:3)
Theorem 3.7. Suppose that no point in bAfin is isolated in bA. Let H be a separable
Hilbert space. Then WM(A, H ) is a dense Gδ in Rep(A, H ).
Proof. By Lemma 3.5 it suffices to show the density of WM(A, H ). Let π ∈ Rep(A, H ).
By a maximality argument involving the collection of direct sums of finite-dimensional
subrepresentations of π, we can write π = π0 ⊕Li∈I πi where π0 is weakly mixing and
πi is finite-dimensional for every i ∈ I. By decomposing further we may assume that πi
is irreducible for each i ∈ I. By Lemma 3.6, for each i ∈ I we can find a weak mixing
QUANTUM GROUPS, PROPERTY (T), AND WEAK MIXING
11
(cid:3)
the desired density.
H be a separable infinite-dimensional Hilbert space. Then WM(A, H ) is closed and
nowhere dense in Gδ in Rep(A, H ).
representation θi on a separable Hilbert space such that πi ≺ θi. Set ρ = π0 ⊕Li∈I θi.
Then π ≺ ρ, and ρ acts on a separable Hilbert space since I is countable by the sep-
arability of H . It follows by Lemma 3.4 that π belongs to the closure of the set of
all κ ∈ Rep(G, H ) such that κ ∼= ρ, and hence to the closure of WM(A, H ), yielding
Theorem 3.8. Suppose that bAfin 6= ∅ and each point in bAfin is isolated in bA. Let
Proof. By assumption there exists a ρ ∈ bAfin. Let π ∈ Rep(A, H ). Then clearly
π ≺ π ⊕ ρ and so by Lemma 3.4 the representation π belongs to the closure of the
set of all κ ∈ Rep(G, H ) such that κ ∼= π ⊕ ρ, showing that the complement of
WM(A, H ) is dense in Rep(A, H ).
Now let π be a representation in Rep(A, H ) which is not weakly mixing. Then
we can write π = π0 ⊕ π1 where π1 is finite-dimensional, and we may assume that
π1 is irreducible. Now suppose that {ρn} is a sequence in Rep(A, H ) converging
to ρ and set ρ = L∞
n=1 ρn. Then π ≺ ρ and hence π1 ≺ ρ, which implies that
π1 is a subrepresentation of ρ since π1 is isolated in bAfin ([33], Theorem 1.7). Since
π1 is irreducible, there must exist an n ∈ N such that π1 is a subrepresentation of
ρn. We deduce from this that π has a neighbourhood in Rep(A, H ) which does not
intersect WM(A, H ). This shows that the complement of WM(A, H ) is open and
hence completes the proof.
(cid:3)
4. Weak mixing and Property (T) for quantum groups
We now return to the context of quantum groups and discuss the notions of weak
mixing and property (T). Let G be a locally compact quantum group.
Definition 4.1. A unitary representation U ∈ M(C0(G) ⊗ K (H )) of G is weakly
mixing if it contains no nonzero finite-dimensional subrepresentation.
Since finite-dimensionality is preserved under the canonical correspondence between
0 (bG), by Propo-
unitary representations of G and nondegenerate representations of C u
sition 3.2 we obtain the following.
Proposition 4.2. Let U ∈ M(C0(G) ⊗ K (H )) be a unitary representation, let πU
be the corresponding representation of C u
U be the canonical extension
0 (bG)+. Then U is weakly mixing
of πU to a unital representation of the unitization C u
0 (bG)+ is weakly mixing.
if and only if the restriction of π+
Let H be a fixed Hilbert space. Write Rep(G, H ) for the collection of all unitary
representations of G on H , and equip it with the point-strict topology it inherits as
a subset of M(C0(G) ⊗ K (H )). Let WM(G, H ) ⊆ Rep(G, H ) be the subcollection
0 (bG), and let π+
U to the unitary group of C u
12
MICHAEL BRANNAN AND DAVID KERR
0 (bG), H ) for the collection of non-
of all weakly mixing representations. Write Rep(C u
0 (bG) on H (which is consistent with our notation in
degenerate representations of C u
Section 3 for unital C∗-algebras). In Proposition 5.1 of [10] it is shown that the topol-
0 (bG), H ) under the canonical bijection between Rep(G, H )
ogy T induced on Rep(C u
0 (bG), H ) is the point-strict topology. Since the strict topology and the
and Rep(C u
∗-strong operator topology coincide on bounded subsets of B(H ), this is the same as
0 (bG), H ). Since the ∗-strong operator
the point-∗-strong operator topology on Rep(C u
topology and the strong operator topology agree on the unitary group of B(H ), and a
unital C∗-algebra is spanned by its unitaries, we therefore have πn → π in the topology
n → π+ in the point-strong operator topology
T on Rep(C u
n and π+ are the canonical unital extensions of πn and π
on Rep(C u
0 (bG)+. Combining these observations with Lemma 3.5, we obtain:
to the unitization C u
0 (bG), H ) if and only if π+
0 (bG)+, H ), where π+
Proposition 4.3. Let H be a separable infinite-dimensional Hilbert space. Then
WM(G, H ) is a Gδ in Rep(G, H ).
Next we recall the definition of property (T).
Definition 4.4. Let G be a locally compact quantum group and let U ∈ M(C0(G) ⊗
K (H )) be a unitary representation of G. A vector ξ ∈ H is said to be invariant for
U if U(η ⊗ ξ) = η ⊗ ξ for all η ∈ L2(G). We say that U has almost invariant vectors
if there is a net {ξi}i of unit vectors in H such that
for all η ∈ L2(G), which by Proposition 3.7 of [10] is equivalent to
kU(η ⊗ ξi) − η ⊗ ξik → 0
kπU (a)ξi − εu(a)ξik → 0
for all a ∈ C u
Definition 4.5. A locally compact quantum group G has property (T) if every unitary
representation G having almost invariant vectors has a nonzero invariant vector.
0 (bG).
In order to establish the two main results of this section, Theorems 4.8 and 4.9, we
require a characterization of property (T) in terms of the isolation of finite-dimensional
representations in the spectrum. Specifically, we will need the equivalence of (i) and
(iv) in Theorem 4.7 below. For locally compact groups the equivalence between (i),
(iii), and (iv) in Theorem 4.7 is due to Wang [33], and is known more generally for
locally compact quantum groups with trivial scaling group, although it does not seem
to be explicitly stated in this generality in the literature (see Remark 5.3 of [24] and
Section 3 of [7]). For quantum groups the idea is to adapt the argument of Bekka,
de la Harpe, and Valette for groups in Section 1.2 of [5]. This requires Lemma 4.6,
which generalizes a well known fact for locally compact groups. The discrete case
of Lemma 4.6 appears in Section 2.5 of [24], although the proof there works more
QUANTUM GROUPS, PROPERTY (T), AND WEAK MIXING
13
generally, as observed in Proposition 7.2 of [9]. See also Section 3 of [7] and the first
paragraph of the proof of Proposition 3.5 in [32].
Lemma 4.6. Let U and V be finite-dimensional unitary representations of a second
countable locally compact quantum group G with trivial scaling group. Then 1G ≤ U⊙ ¯V
if and only if U and V contain a common nonzero subrepresentation.
Armed with Lemma 4.6, one can now establish the following result by repeating
mutatis mutandis the argument in Section 1.2 of [5], as was done in Section 3 of [7] in
the discrete unimodular case.
Theorem 4.7. For a second countable locally compact quantum group G with trivial
scaling group the following are equivalent:
(i) G has property (T),
(ii) εu is isolated in the spectrum of C u
(iii) every finite-dimensional representation in the spectrum of C u
(iv) there exists a finite-dimensional representation in the spectrum of C u
0 (bG),
0 (bG) is isolated,
0 (bG) which
is isolated.
Theorem 4.8. A second countable locally compact quantum group G with trivial scal-
ing group has property (T) if and only if every weakly mixing unitary representation
of G fails to have almost invariant vectors.
Proof. For the nontrivial direction, if G does not have property (T) then by Theo-
rem 4.7 and Lemma 3.6 there is a weakly mixing unitary representation of G with
almost invariant vectors.
(cid:3)
Theorem 4.9. Let G be a second countable locally compact quantum group with trivial
scaling group. Let H be a separable infinite-dimensional Hilbert space. If G does not
have property (T) then WM(G, H ) is a dense Gδ in Rep(G, H ), while if G has
property (T) then WM(G, H ) is closed and nowhere dense in Rep(G, H ).
Proof. Apply Theorem 4.7 in conjunction with Theorems 3.7 and 3.8.
(cid:3)
5. Property (T) and strongly ergodic actions
We establish in Theorem 5.1 a quantum group version of a result of Connes and
Weiss [8] for countable discrete groups. It was verified by Daws, Skalski, and Viselter
under the additional hypothesis that the quantum group is discrete and has low dual
([11], Theorem 9.3). In fact to obtain the conclusion we can simply apply the argument
of Daws, Skalsi, and Viselter by replacing their Theorem 7.3 with our Theorem 4.8,
or rather a slight strengthening of the latter in line with Remark 7.4 of [11], as we
explain below.
An n.s.p. (normal-state-preserving) action G yα (N, σ) is a normal injective unital
∗-homomorphism α : N → L∞(G)⊗N, where N is a von Neumann algebra with a
14
MICHAEL BRANNAN AND DAVID KERR
faithful normal state σ, such that (id⊗ α)α = (∆⊗ id)α and (id⊗ σ)α(x) = σ(x)1 for
all x ∈ N. We drop the symbol α if we don't need to refer to it explicitly.
Let G yα (N, σ) be an n.s.p. action. A bounded net {xi} ⊂ N in said to be
asymptotically invariant if for every normal state ω ∈ L1(G) we have (ω⊗ id)(α(xi))−
xi → 0 strongly, and trivial if xi − σ(xi)1 → 0 strongly. The action is said to be
strongly ergodic if every asymptotically invariant net is trivial.
Recall that R denotes the unitary antipode of G, which acts on B(L2(G)) as R(x) =
JRx∗JR where JR is the modular conjugation associated to the left Haar weight on
L∞(G). We say that a unitary representation U ∈ M(C0(G) ⊗ K (H )) of G is self-
conjugate (referred to as condition R in [11]) if there exists an anti-unitary operator
J : H → H such that the anti-isomorphism j : B(H ) → B(H ) given by x 7→ Jx∗J ∗
satisfies (R ⊗ j)(U) = U.
Let U be a unitary representation of G on a Hilbert space H . Recall that the
conjugate representation U on the conjugate Hilbert space H is defined as (R⊗T )(U)
where T : B(H ) → B(H ) is the map given by T (b)ξ = b∗ξ. Let V = U ⊙ U be the
tensor product of U and U . Write J for the anti-unitary operator on H ⊗ H given
on elementary tensors by J(ξ ⊗ ζ) = ζ ⊗ ξ. and let j : B(H ) → B(H ) be the anti-
isomorphism given by x 7→ Jx∗J ∗. Let y =Pi∈I ai ⊗ bi be a finite sum of elementary
tensors in L∞(G)⊗ B(H ). Since R(aiR(aj)) = ajR(ai) and j(bi ⊗ T (bj)) = bj ⊗ T (bi)
for all i, j ∈ I, it follows that the element
x = y12[(R ⊗ T )y]13 = Xi,j∈I
aiR(aj) ⊗ bi ⊗ T (bj) ∈ L∞(G) ⊗ B(H ) ⊗ B(H )
satisfies
(R ⊗ j)(x) = x.
Since the maps R⊗ j and R⊗ T are ∗-strongly continuous, multiplication is ∗-strongly
continuous on the unit ball of B(L2(G) ⊗ H ⊗ H ), and U is a ∗-strong limit of
operators of norm at most one in L∞(G)⊗ B(H ) by Kaplansky density, we conclude
that (R ⊗ j)(V ) = V , so that V is self-conjugate.
Now if U has almost invariant vectors then so does V , as is easily seen, and if U
is weakly mixing and G has trivial scaling group then V is weakly mixing by Theo-
rem 3.11 of [32]. It thus follows from Theorem 4.8 that if G has trivial scaling group
and does not have property (T) then it admits a weakly mixing self-conjugate unitary
representation with almost invariant vectors.
This extra self-conjugacy condition is needed in the argument of Daws -- Skalski --
Viselter, who apply it in Lemma 9.2 of [11] so as to permit the use of Vaes's construction
of actions on the free Araki -- Wood factors from [31]. Now that we have also have it
in our more general setting, we can apply the argument in Section 9 of [11] to deduce
the following theorem. Note that the assumption of trivial scaling group is not merely
required for the application of Theorem 3.11 of [32] in the previous paragraph, but is
also a hypothesis in Lemma 9.2 of [11]. Here (L F∞, τ ) is the von Neumann algebra of
QUANTUM GROUPS, PROPERTY (T), AND WEAK MIXING
15
the free group on a countably infinite set of generators along with its canonical tracial
state.
Theorem 5.1. For a second countable locally compact quantum group G with trivial
scaling group the following are equivalent:
(i) G has property (T),
(ii) every weakly mixing n.s.p. action G y (N, σ) is strongly ergodic,
(iii) every ergodic n.s.p. action G y (N, σ) is strongly ergodic.
One can also replace (N, σ) with the fixed pair (L F∞, τ ) in (ii) and (iii).
6. The general nonunimodular case
In this final section, our aim is to prove Theorem 6.3, which extends Theorems 4.8
and 4.9 so as to cover the general nonunimodular case. It shows in particular that a
nonunimodular second countable locally compact quantum group cannot have property
(T), which in the discrete situation was established in [13].
Let G be a second countable locally compact quantum group. Recall that the mod-
ular element of G is strictly positive unbounded operator δ on L2(G) affiliated with
the von Neumann algebra L∞(G). The map t 7→ δit from R to L∞(G) ⊆ B(L2(G)) is
continuous in the strong operator topology by Stone's theorem and so it defines a uni-
tary operator U on L2(R, L2(G)) ∼= L2(R)⊗ L2(G) which, when viewing L2(R, L2(G))
as the Hilbert space direct integral of copies of L2(G) over R with respect to Lebesgue
measure, is a decomposable operator and as such is expressed by the direct integral
R ⊕
R δit dt. Thus for all η1, η2 ∈ L2(G) and ξ1, ξ2 ∈ L2(R) we have
hU(η1 ⊗ ξ1), η2 ⊗ ξ2i =ZRhξ1(t)δitη1, ξ2(t)η2i dt
(1)
ξ1(t)ξ2(t)hδitη1, η2i dt.
=ZR
Since ∆(δit) = δit ⊗ δit (see for example the proof of Proposition 1.9.11 in [30]), we see
that (∆⊗id)U and U13U12 are both decomposable operators on L2(R, L2(G)⊗L2(G)) ∼=
L2(G) ⊗ L2(G) ⊗ L2(R) which can be expressed as the direct integral R ⊕
R δit ⊗ δit dt.
Thus U is a unitary representation of G on L2(R). (In fact, U ∈ L∞(G)⊗L∞(R) can
be regarded as a unitary representation of both G and R simultaneously.)
Lemma 6.1. The unitary representation U of G has almost invariant vectors.
Proof. For every n ∈ N, writing 1[0, 1
√n1[0, 1
formula (1), for every n we have
n ] we set ξn =
n ], which is a unit vector in L2(R). Let η ∈ L2(G) be a unit vector. Using the
n ] for the indicator function of [0, 1
hU(η ⊗ ξn), η ⊗ ξni − 1 = nZ[0, 1
n ]h(δit − 1)η, ηi dt
16
MICHAEL BRANNAN AND DAVID KERR
and the expression on the right converges to zero as n → ∞ by the strong operator
continuity of the map t 7→ δit. Thus U has almost invariant vectors.
(cid:3)
Let V be any unitary representation of G on a separable Hilbert space H , and con-
sider the tensor product representation Z = V ⊙ U = V12U13. Viewing L2(R, L2(G) ⊗
H ) ∼= L2(G)⊗ H ⊗ L2(R) as the Hilbert space direct integral of copies of L2(G)⊗ H
over R with respect to Lebesgue measure, the operator V12 is decomposable and can
expressed by the direct integral R ⊕
R V dt, so that for η ∈ L2(G) and ζ ∈ L2(R, H ) ∼=
H ⊗ L2(R) the vector V ∗
12(η ⊗ ζ), viewed as an element of L2(R, L2(G)⊗ H ), is equal
to t 7→ V ∗(η ⊗ ζ(t)). Thus for η1, η2 ∈ L2(G) and ζ1, ζ2 ∈ L2(R, H ) ∼= H ⊗ L2(R) we
have the formula
(2)
hZ(η1 ⊗ ζ1), η2 ⊗ ζ2i = hU13(η1 ⊗ ζ1i), V ∗
12(η2 ⊗ ζ2i)
=ZRhδitη1 ⊗ ζ1(t), V ∗(η2 ⊗ ζ2(t))i dt.
Lemma 6.2. Suppose that G is not unimodular. Let V be any unitary representation of
G on a separable Hilbert space H . Then the tensor product representation Z = V ⊙ U
is weakly mixing. Moreover, if H is infinite-dimensional then the closure of the set
of unitary conjugates of Z in Rep(G, H ) contains V .
Proof. Since G is not unimodular, there is a real number c 6= 0 such that ec belongs
to the spectrum of the modular element δ. The equation ∆(δ) = δ ⊗ δ then implies,
via elementary spectral theory, that enc belongs to the spectrum of δ for every n ∈ N.
Let E be a nonzero finite-dimensional subspace of L2(R)⊗ H . Choose an orthonor-
mal basis {ζ1, . . . , ζK} for E . Let ε > 0. Then there are a b > 0 and an N ∈ N
k ∈ L2([−b, b], H ) ∼= L2([−b, b]) ⊗ H
such that for each k = 1, . . . , K we can find a ζ ′
which is an H -valued step function on [−b, b] taking at most N different values and
satisfying kζ ′
k(t)k. By the proof of
the Riemann -- Lebesgue lemma for step functions, there is an n ∈ N depending only on
N and M such that, setting s = nc, every step function f : [−b, b] → C which takes
at most N 2 values and is bounded in modulus by M 2 satisfies
k − ζkk < ε/3. Set M = maxk=1,...,K maxt∈[−b,b] kζ ′
(3)
f (t)eist dt(cid:12)(cid:12)(cid:12)(cid:12) <
ε
6
Z[−b,b]
(cid:12)(cid:12)(cid:12)(cid:12)
Write K for the range of the spectral projection of δ corresponding to [0, 2es]. Since
es belongs to the spectrum of δ, we can find a norm-one vector η ∈ K such that
kδη − esηk is small enough so that by the continuous functional calculus, applied to δ
acting (boundedly) on K , we have kδitη−eistηk < ε/(6M 2) for all t ∈ [−b, b]. For every
norm-one vector θ ∈ L2(G) and k = 1, . . . , n the function t 7→ hη⊗ ζ ′
k(t))i
on [−b, b] is a step function which takes at most N 2 values and is bounded in modulus
1(t), V ∗(θ⊗ ζ ′
QUANTUM GROUPS, PROPERTY (T), AND WEAK MIXING
17
by M 2, and so using the formula (2) and applying (3) we obtain
hZ(η ⊗ ζ ′
1), θ ⊗ ζ ′
ki =(cid:12)(cid:12)(cid:12)(cid:12)
≤(cid:12)(cid:12)(cid:12)(cid:12)
≤(cid:12)(cid:12)(cid:12)(cid:12)
1(t), V ∗(θ ⊗ ζ ′
Z[−b,b]hδitη ⊗ ζ ′
Z[−b,b]heistη ⊗ ζ ′
+(cid:12)(cid:12)(cid:12)(cid:12)
Z[−b,b]hη ⊗ ζ ′
t∈[−b,b](cid:0)kδitη − eistηkkζ ′
ε
6M 2 · M 2
1(t), V ∗(θ ⊗ ζ ′
Z[−b,b]hδitη − eistη ⊗ ζ ′
1(t), V ∗(θ ⊗ ζ ′
+ sup
k(t))i dt(cid:12)(cid:12)(cid:12)(cid:12)
k(t))i dt(cid:12)(cid:12)(cid:12)(cid:12)
k(t))ieist dt(cid:12)(cid:12)(cid:12)(cid:12)
k(t)k(cid:1)
1(t), V ∗(θ ⊗ ζ ′
1(t)kkζ ′
+
k(t))i dt(cid:12)(cid:12)(cid:12)(cid:12)
<
=
and hence
ε
6
ε
3
(4)
1), θ ⊗ ζ ′
hZ(η ⊗ ζ1), θ ⊗ ζki ≤ hZ(η ⊗ ζ ′
ε
= ε.
+
3
ε
3
ε
3
+
<
ki + kζ1 − ζ ′
1k + kζk − ζ ′
kk
Now given a norm-one vector κ ∈ L2(G) ⊗ E we can write it asPK
PK
k=1 kθkk2 = 1, and so if we take ε = 1/K then from (4) we get
k=1 θk ⊗ ζk where
hZ(η ⊗ ζ1), κi ≤
hZ(η ⊗ ζ1), θk ⊗ ζki <
KXk=1
KXk=1
1
nkθkk ≤ Kε = 1.
Since the vector Z(η⊗ζ1) has norm one by the unitarity of Z, it follows that Z(η⊗ζ1) /∈
E , which shows that L2(G)⊗E is not U-invariant. We conclude that U has no nonzero
finite-dimensional subrepresentations.
Suppose now that H is infinite-dimensional and let us show that the closure of the
set of unitary conjugates of Z in Rep(G, H ) contains V . Fix an orthonormal basis
{ζk}∞
k=1 of H . Let Ω be a finite set of norm-one elements in C0(G). Let K ∈ N and
ε > 0. Working in the Hilbert module C0(G) ⊗ H , for each k = 1, . . . , K there exist
xk,1, . . . , xk,Lk, yk,1, . . . , yk,Lk ∈ H such that kV(a ⊗ ζk) −PLk
l=1 xk,l ⊗ ζlk < ε/3 and
kV∗(a ⊗ ζk) −PLk
l=1 yk,l ⊗ ζlk < ε/3 for all a ∈ Ω. Set L = max{k, L1, . . . , LK}. Using
the characterization of having almost invariant vectors given in Proposition 3.7(v) of
[10], Lemma 6.1 yields a unit vector ξ ∈ H such that kU(a ⊗ ξ) − a ⊗ ξk < ε/3
for all a ∈ Ω. Since H is separable and infinite-dimensional we can find a unitary
isomorphism u : H ⊗ L2(R) → H which sends ζk ⊗ ξ to ζk for every k = 1, . . . , L.
18
MICHAEL BRANNAN AND DAVID KERR
Then for every a ∈ Ω and k = 1, . . . , K we have
k((id ⊗ u)Z(id ⊗ u)−1 − V)(a ⊗ ζk)k
≤ k(id ⊗ u)V12(U13(a ⊗ ζk ⊗ ξ) − a ⊗ ζk ⊗ ξ)k
+(cid:13)(cid:13)(cid:13)(cid:13)(id ⊗ u)(cid:18)V12(a ⊗ ζk ⊗ ξ) −
xk,l ⊗ ζl − V(a ⊗ ζk)(cid:13)(cid:13)(cid:13)(cid:13)
+(cid:13)(cid:13)(cid:13)(cid:13)
LkXl=1
= ε
LkXl=1
xk,l ⊗ ζl ⊗ ξ(cid:19)(cid:13)(cid:13)(cid:13)(cid:13)
<
ε
3
+
+
ε
3
ε
3
and similarly k((id ⊗ u)Z∗(id ⊗ u)−1 − V∗)(a ⊗ ζk)k < ε. We conclude that V belongs
to the closure of the set of unitary conjugates of Z in Rep(G, H ).
(cid:3)
We can now conclude with the main result of the section.
Theorem 6.3. Let G be a second countable nonunimodular locally compact quantum
group. Then there exists a a weakly mixing unitary representation of G which has
almost invariant vectors. Moreover, if H is a separable infinite-dimensional Hilbert
space then WM(G, H ) is a dense Gδ in Rep(G, H ).
Proof. The representation U has almost invariant vectors by Lemma 6.1, and it is
weakly mixing by Lemma 6.2, as we can take V there to be the trivial representation.
Finally, if H is a separable infinite-dimensional Hilbert space then WM(G, H ) is
(cid:3)
a dense Gδ in Rep(G, H ) by Lemmas 3.5 and 6.2.
References
[1] Y. Arano. Unitary spherical representations of Drinfeld doubles. To appear in J. Reine Angew.
Math.
[2] Y. Arano. Comparison of unitary duals of Drinfeld doubles and complex semisimple Lie groups.
Comm. Math. Phys. 351 (2017), 1137 -- 1147.
[3] E. B´edos, R. Conti, and L. Tuset. On amenability and co-amenability of algebraic quantum
groups and their corepresentations. Canad. J. Math. 57 (2005), 17 -- 60.
[4] M. E. B. Bekka and A. Valette. Kazhdan's property (T) and amenable representations. Math.
Z. 212 (1993), 293 -- 299.
[5] B. Bekka, P. de la Harpe, and A. Valette. Kazhdan's Property (T). New Mathematical Mono-
graphs, 11. Cambridge University Press, Cambridge, 2008.
[6] M. Brannan, M. Daws, and E. Samei. Completely bounded representations of convolution
algebras of locally compact quantum groups. Munster J. Math. 6 (2013), 445 -- 482.
[7] X. Chen and C.-K. Ng. Property T for locally compact groups. Intl. J. Math. 26 (2015), 1550024,
13 pp.
[8] A. Connes and B. Weiss. Property T and asymptotically invariant sequences. Israel J. Math.
37 (1980), 209 -- 210.
[9] B. Das and M. Daws. Quantum Eberlein compactifications and invariant means. Indiana Univ.
Math. J. 65 (2016), 307 -- 352.
QUANTUM GROUPS, PROPERTY (T), AND WEAK MIXING
19
[10] M. Daws, P. Fima, A. Skalski, S. White. The Haagerup property for locally compact quantum
groups. J. Reine Angew. Math. 711 (2016), 189 -- 229.
[11] M. Daws, A. Skalsi, and A. Viselter. Around property (T)
for quantum groups.
arXiv:1605.02800v1.
[12] J. Dixmier. C∗-Algebras. Translated from the French by Francis Jellett. North-Holland Mathe-
matical Library, Vol. 15. North-Holland Publishing Co., Amsterdam-New York-Oxford, 1977.
[13] P. Fima. Kazhdan's property T for discrete quantum groups. Internat. J. Math. 21 (2010),
47 -- 65.
[14] E. Glasner and B. Weiss. Kazhdan's property T and the geometry of the collection of invariant
measures. Geom. Funct. Anal. 7 (1997), 917 -- 935.
[15] M. Joit¸a and S. Petrescu. Property (T) for Kac algebras. Rev. Roumaine Math. Pures Appl. 37
(1992), 163 -- 178.
[16] P. Jolissaint. On property (T) for pairs of topological groups. Enseign. Math. (2) 51 (2005),
31 -- 45.
[17] A. S. Kechris. Classical Descriptive Set Theory. Graduate Texts in Mathematics, 156. Springer-
Verlag, New York, 1995.
[18] A. S. Kechris. Global Aspects of Ergodic Group Actions. Mathematical Surveys and Monographs,
160. American Mathematical Society, Providence, RI, 2010.
[19] D. Kerr and H. Li. Ergodic Theory: Independence and Dichotomies. Springer, 2016.
[20] D. Kerr and M. Pichot. Asymptotic Abelianness, weak mixing, and property T. J. Reine Angew.
Math. 623 (2008), 213 -- 235.
[21] J. Kustermans. Locally compact quantum groups in the universal setting. Internat. J. Math.
12 (2001), 289 -- 338.
[22] J. Kustermans and S. Vaes. Locally compact quantum groups. Ann. Sci. ´Ecole Norm. Sup. (4)
33 (2000), 837 -- 934.
[23] J. Kustermans and S. Vaes. Locally compact quantum groups in the von Neumann algebraic
setting. Math. Scand. 92 (2003), 68 -- 92.
[24] D. Kyed and P. Soltan. Property (T) and exotic quantum group norms. J. Noncommut. Geom.
6 (2012), 773 -- 800.
[25] E. C. Lance. Hilbert C∗-Modules. A Toolkit for Operator Algebraists. London Mathematical
Society Lecture Note Series, 210. Cambridge University Press, Cambridge, 1995.
[26] S. Neshveyev and M. Yamashita. Drinfeld center and representation theory for monoidal cate-
gories. Comm. Math. Phys. 345 (2016), 385 -- 434.
[27] J. Peterson and S. Popa. On the notion of relative property (T) for inclusions of von Neumann
algebras. J. Funct. Anal. 219 (2005), 469 -- 483.
[28] S. Popa. Deformation and rigidity for group actions and von Neumann algebras. In: Interna-
tional Congress of Mathematicians. Vol. I., 445 -- 477. Eur. Math. Soc., Zurich, 2007.
[29] S. Popa, and S. Vaes. Representation theory for subfactors, λ-lattices and C∗-tensor categories.
Comm. Math. Phys. 340 (2015), 1239 -- 1280.
[30] S. Vaes. Locally compact quantum groups. Thesis, Catholic University of Leuven, 2001.
[31] S. Vaes. Strictly outer actions of groups and quantum groups. J. Reine Angew. Math. 578
(2005), 147 -- 184.
[32] A. Viselter. Weak mixing for locally compact quantum groups. To appear in Ergodic Theory
Dynam. Systems.
[33] P. S. Wang. On isolated points in the dual spaces of locally compact groups. Math. Ann. 218
(1975), 19 -- 34.
[34] R. J. Zimmer. Ergodic Theory and Semisimple Groups. Monographs in Mathematics, 81.
Birkhauser Verlag, Basel, 1984.
20
MICHAEL BRANNAN AND DAVID KERR
Michael Brannan, Department of Mathematics, Mailstop 3368, Texas A&M Univer-
sity, College Station, TX 77843-3368, USA
E-mail address: [email protected]
David Kerr, Department of Mathematics, Mailstop 3368, Texas A&M University,
College Station, TX 77843-3368, USA
E-mail address: [email protected]
|
1611.01556 | 2 | 1611 | 2016-11-08T14:59:31 | Dirac type operators on the quantum solid torus with global boundary conditions | [
"math.OA"
] | We define a noncommutative space we call the quantum solid torus. It is an example of a noncommutative manifold with a noncommutative boundary. We study quantum Dirac type operators subject to Atiyah-Patodi-Singer like boundary conditions on the quantum solid torus. We show that such operators have compact inverse, which means that the corresponding boundary value problem is elliptic. | math.OA | math |
DIRAC TYPE OPERATORS ON THE QUANTUM SOLID TORUS WITH
GLOBAL BOUNDARY CONDITIONS
SLAWOMIR KLIMEK AND MATT MCBRIDE
Abstract. We define a noncommutative space we call the quantum solid torus.
It is
an example of a noncommutative manifold with a noncommutative boundary. We study
quantum Dirac type operators subject to Atiyah-Patodi-Singer like boundary conditions on
the quantum solid torus. We show that such operators have compact inverse, which means
that the corresponding boundary value problem is elliptic.
1. Introduction
In this paper we are continuing our study of quantum analogs of Dirac type operators on
manifolds with boundary and global boundary conditions of Atiyah, Patodi, and Singer type
[2]. Despite similarities with our previous papers on the subject, the problem studied here
is much more complex. We consider a higher dimensional space, namely three dimensional
solid torus, and its quantization. The resulting space is noncommutative and its boundary
is also noncommutative. Dirac type operators are matrix valued operators and the Atiyah-
Patodi-Singer condition cannot be generalized straightforwardly.
One of the main objectives of our previous work was to find appropriate non-commutative
analogs of Dirac type operators on planar domains with boundary and to investigate their
functional analytic properties. Some examples of Dirac type operators in simple domains,
such as the disk, annulus, and punctured disk were described in [5] and [6]. These papers
exhibited strong similarities in the setup and the results between the commutative and
quantum cases. Also in those papers the global boundary condition imposed on Dirac-like
operators was essentially the classical APS boundary condition. In reference [8] we discussed
Dirac type operators on the solid torus, in the commutative sense, with a different nonlocal
boundary condition, that was inspired by [9]. This is because the natural metric on the disk
does not have the product structure near boundary, required for APS theory. This was not
a problem for two dimensional domains but becomes an issue in dimension three. In this
paper we follow up the analysis [8] of the operators on the commutative solid torus, with
same type of investigation for their quantum analogs.
Instead of constructing a quantum analog of the non-local boundary condition that was
used in [8], we consider a large class of boundary conditions that yield desired analytical
properties of the parametrices of the quantum Dirac operators. The boundary condition of
[8] required extending functions in the domain beyond the boundary of the solid torus, which
makes sense geometrically in the classical case. However in the quantum analog, there are
obvious obstructions, for example what does we mean by the "outside" of the boundary of
the quantum solid torus. This will be addressed in our future work.
Date: January 16, 2018.
1
2
SLAWOMIR KLIMEK AND MATT MCBRIDE
The solid torus is the product of a disk and a circle. In quantum case we take a noncom-
mutative disk and then a twisted product of it with the unit circle to produce the quantum
solid torus. The operators we consider respect this decomposition. Unlike our previous work
on d-bar operators on the quantum disk [5], we consider here a bigger variety of related Dirac
like operators on the disk. They yield a wide class of operators on the solid torus. The crux
of the analysis presented here is the proof that the parametrices of our Dirac type operators
are compact operators, like it was shown in [8] that the parametrix of the classical Dirac
type operator was also compact.
The paper is organized as follows.
In section 2 the quantum solid torus is introduced
and discussed. This is followed by section 3 on partial Fourier series on the quantum solid
torus. The relevant Hilbert spaces are also defined there. Section 4 introduces our Dirac
type operators, while section 5 contains a discussion of some properties of their coefficients.
The boundary conditions are defined in section 6, based on properties of special solutions
introduced there. The computation of the parametrix of the Dirac type operator is described
in section 7. Finally, section 8 contains analysis of a parametrix culminating in the proof of
the main theorem on compactness.
2. Non-commutative solid Torus
In this section we define our version of the quantum solid torus. From the topological
point of view such a quantum torus is a noncommutative C ∗−algebra which has a structure
similar in some ways to the ordinary solid torus. The idea here is simple: the solid torus
can be thought of as the product of the disk and the circle. To obtain a quantum solid torus
we take the quantum disk of [4] and take its twisted product with the circle. The result can
be described either as a suitable crossed product algebra or as a universal C ∗−algebra with
generators and relations. We proceed to describe them now.
Definition: Let Aθ be the universal C ∗−algebra with generators U and V such that U is
an isometry i.e. U ∗U = I, V is a unitary i.e. V ∗V = V V ∗ = I and such that they satisfy
the commutation relation
V U = e2πiθUV.
Let {ek} be the canonical basis for ℓ2(Z≥0) and let W be the unilateral shift, i.e. W ek =
ek+1. Let T be the C ∗−algebra generated by W . The algebra T is called the Toeplitz algebra
and by Coburn's theorem it is the universal C ∗−algebra with generator W satisfying the
relation W ∗W = I, i.e. an isometry. Reference [4] shows that this algebra can be thought
of as a quantum unit disk. It's structure is described by a short exact sequence, namely:
(2.1)
where K is the ideal of compact operators in ℓ2(Z≥0). In fact K is the commutator ideal of
the algebra T .
0 −→ K −→ T −→ C(S1) −→ 0.
∼= T ⋊θ Z
Proposition 2.1. For any θ we have the following isomorphism of algebras: Aθ
where the crossed product on the right hand side is defined, for n ∈ Z, by the automorphisms
Θn(U) = e2πinθU for some θ ∈ [0, 2π).
DIRAC TYPE OPERATORS ON THE QUANTUM SOLID TORUS
3
Note that because of the universality of the Toeplitz algebra, we only need to define an
automorphism on its generator and check that it satisfies the defining relation.
Proof. By definition, a representation π of the universal C ∗−algebra Aθ, consists of a Hilbert
space, H and bounded operators π(U) and π(V ) on H, that satisfy the defining relations of
Aθ. Then Aθ is the completion of the algebra of polynomials a in U and V and their adjoints
with respect to the maximal norm, i.e.
kakmax = sup
π
kπ(a)k.
On the other hand, the covariant representation of the dynamical system (Z, Θn, T ) consists
of a Hilbert space X, a bounded representation ρ of T on X, and a unitary operator ρ(V )
on X, such that the following commutation relation holds:
ρ(Θ1(b)) = ρ(V )ρ(b)ρ(V )∗,
for any b ∈ T . By the universality of the Toeplitz algebra, its representation ρ is completely
determined by an isometry ρ(U) i.e. ρ(U)ρ(U)∗ = I. Moreover, the above commutation
relation with ρ(V ) becomes the following:
e2πiθρ(U) = ρ(V )ρ(U)ρ(V )∗.
Consequently the crossed product T ⋊θ Z is obtained by completion of the algebra of poly-
nomials a in U and V and their adjoints with respect to the maximal norm, i.e.
kakmax = sup
ρ
kρ(a)k.
So we see that the two concepts are identical, establishing an isomorphism between the
algebras, and completing the proof.
(cid:3)
By the amenability of Z, the crossed product of the Toeplitz algebra with Z is equal to
the reduced crossed product of the Toeplitz algebra with Z. Consequently, the norm of an
element of the crossed product can computed from a single representation by combining
a faithful representation of the Toeplitz algebra with the left regular representation of Z.
Explicitly, choosing the defining representation of the Toeplitz algebra, we can view the
reduced crossed product as the C ∗−algebra generated by the operators described below
acting in the Hilbert space ℓ2(Z, ℓ2(Z≥0)) = ℓ2(Z) ⊗ ℓ2(Z≥0) = ℓ2(Z≥0 × Z), where the
equality of Hilbert spaces follows from [10].
Corollary 2.2. Let {ek,l} be the canonical basis in ℓ2(Z≥0 × Z). Defining two operators:
Uek,l = e−2πilθek+1,l and V ek,l = ek,l+1, we have Aθ = C ∗(U, V ).
For θ ∈ [0, 2π) let T 2
θ is the
universal C ∗−algebra with two unitary generators u and v such that vu = e2πiθuv, see for
example [3]. The following proposition describes the structure of Aθ.
θ be the two dimensional quantum torus. In other words, T 2
Proposition 2.3. We have the following short exact sequence:
0 −→ K ⊗ C(S1) −→ Aθ −→ T 2
θ −→ 0.
4
SLAWOMIR KLIMEK AND MATT MCBRIDE
Proof. Consider the crossed product of the short exact sequence (2.1) for the Toeplitz algebra
with Z. We get:
0 −→ K ⋊θ Z −→ T ⋊θ Z −→ C(S1) ⋊θ Z −→ 0.
∼= C(S1) ⋊θ Z, see [3]. Moreover, it is easy to see that K ⋊θ Z ∼=
It is well known that T 2
θ
K ⊗ C(S1), see [12] for the details. Explicitly, the cross product K ⋊θ Z can be "untwisted"
by noticing that, in the tensor product of Hilbert spaces ℓ2(Z) ⊗ ℓ2(Z≥0), if a is compact in
ℓ2(Z≥0) then:
(I ⊗ a)(V ′)n = (I ⊗ e−2πiKθa) · (V ⊗ I)n.
The operators above are Kek,l = kek,l and V ′ek,l = e−2πikθek,l+1.
In addition define the
operator U ′ek,l = ek+1,l, then we have Aθ = C ∗(U ′, V ′) by universality, while e−2πiKθa is still
compact. Combination of all those facts above yields the above short exact sequence.
(cid:3)
In view of the results about Aθ contained in this section it natural to call this algebra the
quantum solid torus.
3. Fourier series
The purpose of this section is to introduce a partial Fourier transform on the quantum
torus. As stated in Corollary 2.2, the algebra Aθ is generated by the following two operators:
Uek,l = e−2πilθek+1,l and V ek,l = ek,l+1, satisfying the relations: V U = e2πiθUV , U ∗U = I
and V ∗V = V V ∗ = I. We also reuse the following diagonal label operator Kek,l = kek,l, so,
for a bounded function f : Z≥0 → C, we have f (K)ek,l = f (k)ek,l. We have the following
two useful commutation relations for a diagonal operator f (K):
f (K) U = Uf (K + 1) and f (K) V = V f (K).
(3.1)
Let P ol(U, V ) be the set of all polynomials in U, U ∗, V and V ∗. We call a function f :
Z≥0 → C eventually constant, if there exists a natural number k0 such that f (k) is constant
for k ≥ k0. The set of all such functions will be denoted by EC. Using the above notation
we consider the following operators in ℓ2(Z≥0 × Z) expressed as finite sums:
N
N
a =
Xn=0,m≤M
Xn=1,m≤M
V mU nf +
m,n(K) +
f −
m,n(K)V m(U ∗)n
for some M, N ≥ 0 and f ±
following proposition:
m,n(k) ∈ EC. Let A be the set of all such operators. We have the
Proposition 3.1. A = P ol(U, V ).
Proof. Using the commutation relations (3.1) it is easy to see that a product of two elements
of A and the adjoint of an element of A are still in A. It follows that A is a ∗−subalgebra
of Aθ. Since U and V are in A, it follows that P ol(U, V ) ⊂ A. To prove the other way
inclusion it suffices to show that for any f ∈ EC the operator f (K) is in P ol(U, V ), as the
remaining parts of the sum are already polynomials in U, V , U ∗ and V ∗. To show that
f (K) ∈ P ol(U, V ), we decompose any f (K) ∈ EC in the following way:
DIRAC TYPE OPERATORS ON THE QUANTUM SOLID TORUS
5
k0−1
f (K) =
f (k)Pk + f∞P≥k0,
Xk=0
where f∞ = limk→∞ f (k), Pk is the orthogonal projection onto span{ek,l}l∈Z and P≥k0 is
the orthogonal projection onto span{ek,l}k≥k0,l∈Z. A straightforward calculation shows that
U k(U ∗)k = P≥k and Pk = P≥k − P≥k+1. This completes the proof.
(cid:3)
The above considerations tell us that a ∈ A are completely determined by the coefficients
f ±
m,n(K) ∈ EC and so it motivates the following definition of a partial Fourier series. For
f ∈ Aθ, define the formal series:
fseries = Xn≥0,m∈Z
V mU nf +
m,n(K) + Xn≥1,m∈Z
m,n(K)V m(U ∗)n,
f −
where
f +
m,n(k) = hek,0, (U ∗)nV −mf ek,0i
and f −
m,n(k) = hek,0, f U nV −mek,0i.
Similarly to the usual theory of Fourier series, fseries determines f even though in general
the series is not norm convergent. Other types of convergence results can be obtained along
the lines of the usual Fourier analysis.
We now proceed to describing the Hilbert spaces hosting our Dirac type operators. They
are analogs of the L2 spaces on the classical solid torus. Let {an(k)} be a sequence of positive
numbers, which we call weights, such that the sum
exists and such that s(n) goes to zero as n → ∞. For any formal series define a norm by
s(n) :=
1
an(k)
∞
Xk=0
kfseriesk2 =
∞
Xk=0 Xn≥0,m∈Z
1
an(k)
f +
m,n(k)2 +
∞
Xk=0 Xn≥1,m∈Z
1
an(k)
f −
m,n(k)2.
Let H0 be the Hilbert space whose elements are the above formal series fseries such that
kfseriesk is finite.
Let Q0 be the orthogonal projection in ℓ2(Z≥0 ×Z) onto span{ek,0}k∈Z≥0. Define the linear
functional τ on B1(ℓ2(Z≥0 × Z)), the space of trace class operators, by τ (a) = tr(aQ0).
Lemma 3.2. For a bounded operator a, and b ∈ B1(ℓ2(Z≥0 × Z)) we have the following
inequality:
Proof. We estimate as follows:
τ (ab) ≤ kak(τ (b∗b))1/2 := kakkbkτ .
τ (ab) = tr(abQ0) ≤ kak(tr((bQ0)∗bQ0))1/2 = kak(tr(Q∗
0b∗bQ0))1/2
= kak(tr(b∗bQ0))1/2 = kak(τ (b∗b))1/2,
where the inequality follows from the usual estimate for the trace.
(cid:3)
6
SLAWOMIR KLIMEK AND MATT MCBRIDE
We use this lemma to show that Aθ is dense in H0.
Proposition 3.3. If f ∈ Aθ, then fseries converges in H0. Moreover the map Aθ ∋ f 7→
fseries ∈ H0 is continuous, one-to-one, and the image is dense in H0.
Proof. We follow here the similar argument from [7]. For f ∈ Aθ we need to estimate the
norm of fseries. Notice first that if f ∈ A, then fseries = f . Since such f 's are dense in Aθ, it
suffices to estimate the norm of the finite sums. Without loss of generality, we assume that
f has only the U ∗ and V terms as the same proof works for the remaining terms. For such
f we have:
N
kfseriesk2
H0 =
Xk=0
= τ
Xn=1,m≤M
= τ
Xn=1,m≤M
N
Next we estimate:
Consequently, using the inequality from Lemma 3.2, we get
∞
N
Xn=1,m≤M
1
an(k)
f −
N
m,n(k)2 = τ
Xn=1,m≤M
m,n(K)
1
an(K)
m,n(K)V −m (U ∗)n V mU nf −
f −
1
an(K)
f −
m,n(K)2
kfseriesk2
kf k.
1
an(K)
f −
f ∗
m,n(K)V m(U ∗)n
.
m,n(K)V m(U ∗)n(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)τ
an(K)
f −
N
2
1
H0 ≤(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xn=1,m≤M
m,n(K)V m(U ∗)n(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
f −
1
=
τ
f −
m,n(K)V m(U ∗)n
an(K)
1
an(k)an(k)
N
Xl=1,j≤M
V −jU lf −
j,l(K)
1
an(K)
n,k (cid:18) 1
an(k)(cid:19) .
f −
m,n(k)2 ≤ kfseriesk2
H0 sup
an(K)
N
N
1
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xn=1,m≤M
= τ
Xn=1,m≤M
Xn=1,m≤M
Xk=0
=
∞
N
Using the the summability conditions on the weights we obtain:
sup
n,k (cid:18) 1
an(k)(cid:19) ≤ sup
n ∞
Xk=0
1
an(k)! = sup
n
(s(n)) ≤ const,
and hence we get kfseriesk2
Aθ ∋ f 7→ fseries ∈ H0, and consequently fseries converges in H0 for every f ∈ Aθ.
H0 ≤ constkfserieskH0kf k. This shows the continuity of the map
Next we show that the map Aθ ∋ f 7→ fseries ∈ H0 is one-to-one. Let f and g belong to Aθ
and suppose that fseries = gseries. This means that the Fourier coefficients are equal, that is:
f +
m,n(k) = g+
m,n(k) for all n ≥ 1 and all
m,n(k) for all n ≥ 0 and all m and k, and f −
m,n(k) = g−
DIRAC TYPE OPERATORS ON THE QUANTUM SOLID TORUS
7
m and k. From f +
m,n(k) = hek,0, f U nV −mek,0i it follows
that all matrix coefficients of f and g are the same so we must have f = g. Thus the map
f 7→ fseries ∈ H0 is one-to-one.
m,n(k) = hek,0, (U ∗)nV −mf ek,0i and f −
To prove density we define the following indicator function
δl(k) =(cid:26) 1 l = k
0 l 6= k
.
It is clear that V mU nδl(K) and δl(K)V m(U ∗)n are in Aθ, and moreover they form an orthog-
onal basis for H0. Finally finite linear combinations of V mU nδl(K) and δl(K)V m(U ∗)n form
a dense subspace of Aθ as they are polynomials in U and V and hence are in A, making it
a dense subspace of H0. Thus the proof is complete.
(cid:3)
4. Dirac type operators
The purpose of this section is to introduce the main object of study of this paper: Dirac
type operators on the quantum solid torus. We start with reviewing the definition of such
operators, contained in [8], on the classical solid torus.
In that paper we considered the
following formally self-adjoint Dirac operator Dcl defined in L2(ST 2, C2) ∼= L2(ST 2) ⊗ C2 by
where ST 2 = D × S1 is the solid torus, and D = {z ∈ C :
Dcl =(cid:18) 1
∂
i
∂θ
−2 ∂
2 ∂
∂z
∂z − 1
∂
i
∂θ (cid:19) ,
z ≤ 1} is the unit disk, while
S1 =(cid:8)eiθ ∈ C : 0 ≤ θ ≤ 2π(cid:9) is the unit circle. Using Fourier decomposition:
gm,n(r) (cid:19) einϕ+imθ,
F = Xm,n∈Z(cid:18) fm,n(r)
eiϕ(cid:0) ∂
m
∂r + n
e−iϕ(cid:0)− ∂
r(cid:1)
the operator D becomes:
DclF = Xm,n∈Z(cid:18)
∂r − n
−m
r(cid:1)
(cid:19)(cid:18) fm,n(r)
gm,n(r) (cid:19) einϕ+imθ.
The boundary condition we studied in [8], geometrically very much the same as the APS
condition, considered the solid torus as a subset of the bigger noncompact space of the plane
cross the unit circle. We defined the domain of the Dirac operator Dcl as consisting of those
sufficiently regular (first Sobolev class) vectors F which extend to square integrable solutions
of DclF = 0 on the complement of the solid torus.
Denote by In, Kn the modified Bessel functions of the first and second kind respectively.
Without boundary conditions, the operator Dcl on the solid torus has an infinite dimen-
sional kernel consisting of linear combinations of special solutions In for which, in Fourier
decomposition, the only nonzero components for m 6= 0 are:
fm,n+1(r) = −
m
m
In+1(mr), gm,n(r) = In(mr).
Outside of the solid torus in the plane cross the unit circle the square integrable functions
in kernel of Dcl are linear combinations of solutions Kn for which the only non vanishing
8
SLAWOMIR KLIMEK AND MATT MCBRIDE
components when m 6= 0 are:
fm,n+1(r) =
m
m
Kn+1(mr), gm,n(r) = Kn(mr).
Consequently, see [8] for details, using the Fourier decomposition of F , the boundary condi-
tion can be rephrased as
mKn+1(m)gm,n(1) − mKn(m)fm,n+1(1) = 0,
if m 6= 0 and f0,n(1) = 0 for n ≤ 0. Additionally, for m = 0, we have g0,n(1) = 0 for n ≥ 0.
We now proceed to define our Dirac type operators for the noncommutative solid torus,
in a way that is analogous to the commutative case. Let c1,n(k) ≤ 1 and c2,n(k) ≤ 1 be
sequences of positive numbers such that Qk c1,n(k), Qk c2,n(k) exist and are not zero, and
such that there exists a constant, κ, such that for all k or n we have 1/κ ≤ 1/ci,n(k) ≤ 1,
i = 1, 2.
Let ℓ2
an(Z≥0) be the following Hilbert space of sequences:
ℓ2
an(Z≥0) =({h(k)}k≥0 : Xk
1
an(k)
h(k)2 < ∞) .
We introduce the Jacobi type difference operators by formulas:
Bnh(k) = an(k)(h(k) − c2,n(k − 1)h(k − 1)) : ℓ2
Bnh(k) = an+1(k)(h(k) − c1,n(k)h(k + 1)) : ℓ2
an+1(Z≥0) → ℓ2
an(Z≥0)
an(Z≥0) → ℓ2
an+1(Z≥0) : kBhkℓ2
an+1(Z≥0)
an (Z≥0) < ∞} and the
(4.1)
defined on maximal domains, i.e. dom(B) = {h ∈ ℓ2
domain of B is defined in the same way.
Using the above general one-step difference operators we define operators δ0 and δ2 which
are noncommutative analogs of ∂
∂z and ∂
V mU n+1Bnf +
∂z as:
δ0(f ) = − Xn≥0,m∈Z
δ2(f ) = − Xn≥1,m∈Z
m,n(K) + Xn≥1,m∈Z
m,n(K) + Xn≥0,m∈Z
Bn−1f −
m,n(K)V m (U ∗)n−1
V mU n−1Bn−1f +
Bnf −
m,n(K)V m (U ∗)n+1 .
Those operators are more general versions of the d-bar like operators considered in [5].
Next define another diagonal label operator L on ℓ2(Z≥0 × Z) by Lek,l = lek,l. We use it
for the definition of δ1 = [L, · ], the analog of 1
i
∂
∂θ . It is easy to see that
δ1(f ) = Xn≥0,m∈Z
mV mU nf +
m,n(K) + Xn≥1,m∈Z
again considered on its maximal domain.
mf −
m,n(K)V m (U ∗)n ,
Let H = H0 ⊗ C2. Any F ∈ H has the following Fourier decomposition:
F =(cid:18) Pn≥1,m∈Z V mU nf +
Pn≥0,m∈Z V mU ng+
m,n(K) +Pn≥0,m∈Z f −
m,n(K) +Pn≥1,m∈Z g−
We define the quantum Dirac type operator D to be
m,n(K)V m(U ∗)n
m,n(K)V m(U ∗)n (cid:19) .
(4.2)
DIRAC TYPE OPERATORS ON THE QUANTUM SOLID TORUS
9
Initially we set the domain of D to be the maximal domain:
D =(cid:18) δ1
δ2 −δ1 (cid:19) .
δ0
dom(D) = {F ∈ H : kDF k < ∞}.
(4.3)
(4.4)
Using Fourier series (4.2) we relate D to finite difference operators with matrix coefficients.
For brevity throughout the rest of the paper we only study the "positive" part of the Fourier
transform of the elements of the Hilbert space, as the "negative" part can be analyzed in
an analogical fashion, but writing down all the repetative details would have significantly
increase the length of the paper (and decrease its readability). Because of this, the "+"
superscript is dropped for simplicity.
Define the following Jacobi type difference operator with matrix valued coefficients
Am,n(cid:18) f
g (cid:19) (k) = Am,n(k + 1)(cid:20)(cid:18) f (k + 1)
g(k + 1) (cid:19) − Cm,n(k)(cid:18) f (k)
g(k) (cid:19)(cid:21) ,
Am,n(k + 1) =(cid:18) an+1(k)c1,n(k)
m
0
an(k + 1) (cid:19) ,
where
and
Cm,n(k) =
1
c1,n(k)
−m
an(Z≥0) ⊗ ℓ2
an(k+1)c1,n(k)
an+1(Z≥0) → ℓ2
−m
an+1(k)c1,n(k)
m2
an(k+1)an+1(k)c1,n(k) ! .
c2,n(k) +
(4.5)
(4.6)
(4.7)
Notice that Am,n : ℓ2
domain for Am,n to be
an+1(Z≥0) ⊗ ℓ2
an(Z≥0). Initially we define the
The domain will be modified later to include a boundary condition.
dom Am,n =(cid:8)h ∈ ℓ2
an(Z≥0) ⊗ ℓ2
an+1(Z≥0) : kAm,nhkan+1⊗an < ∞(cid:9) .
(4.8)
Proposition 4.1. For F, G ∈ H where F = (f, g)t, G = (p, q)t, and f, g, p, q ∈ H0, the
"positive" part of the equation DF = G is equivalent to the following equations:
Am,n(cid:18) gm,n
fm,n+1 (cid:19) (k) =(cid:18) pm,n+1(k)
−qm,n(k + 1) (cid:19) ,
for n ≥ 0, m ∈ Z, k ≥ 0, and
an(0)fm,n+1(0) + mgm,n(0) = qm,n(0).
(4.9)
(4.10)
The last equation (4.10) will be referred to below as the initial regularity condition as it
is both a starting point of recurrence proofs and it is analogous to the regularity condition
at origin for solutions of Dirac operator for the classical solid torus.
Proof. In the Fourier series for F and G there is an ambiguity whether n = 0 term should go
with positive or with negative terms. Because we change summation indices in the calculation
below, the ambiguity is resolved differently for different components of F and G.
10
SLAWOMIR KLIMEK AND MATT MCBRIDE
Using the definition of D and shifting n in the first sum we get, ignoring the "negative"
part,
DF = Xn≥0,m∈Z
V m(cid:18) U n+1(mfm,n+1(k) − Bngm,n(k))
U n(−Bnfm,n+1(k) − mgm,n(k)) (cid:19) = G.
Shifting k in the second equation for k ≥ 1 we get the following equivalent system of equations
mfm,n+1(k) − an+1(k)(gm,n(k) − c1,n(k)gm,n(k + 1)) = pm,n+1(k)
an(k + 1)(fm,n+1(k + 1) − c2,n(k)fm,n+1(k)) + mgm,n(k + 1) = −qm,n(k + 1),
(cid:26)
while the case k = 0 in the second equation leads to equation (4.10).
The system above can be rewritten in matrix form:
m
(cid:18) an+1(k)c1,n(k)
−(cid:18) an+1(k)
0
0
an(k + 1) (cid:19)(cid:18) gm,n(k + 1)
fm,n+1(k + 1) (cid:19)
an(k + 1)c2,n(k) (cid:19)(cid:18) gm,n(k)
−m
fm,n+1(k) (cid:19) =(cid:18) pm,n+1(k)
−qm,n(k + 1) (cid:19) .
The first matrix above is Am,n(k +1), therefore factoring it out of the left side of the equation
gives the desired result and completes the proof.
(cid:3)
5. Properties of the coefficients
It follows from Proposition 4.1 that to analyze D we need to study the properties of the
operators Am,n subject to the initial regularity condition. This section discusses important
properties of the matrix coefficients of those operators.
Unless specified differently, in all formulas in the section, n ≥ 0, m ∈ Z, k ≥ 0. First
notice that det Am,n(k + 1) = an+1(k)an(k + 1)c1,n(k) 6= 0 for any k and n which means that
the inverse of the matrices Am,n(k + 1) exists for any k and n. Notice also that det Cm,n(k) =
c2,n(k)/c1,n(k). Additionally we have:
Proposition 5.1. The infinite product
exists and is invertible.
Cm,n :=
Cm,n(k)
∞
Yk=0
Here and everywhere below a product of matrices is understood from right to left, that is:
Cm,n(i) = Cm,n(k)Cm,n(k − 1) · · · Cm,n(0).
k
Yi=0
Proof. We consider Cm,n(k) − I, since by [11], if the series Pk kCm,n(k) − Ik converges
and det Cm,n(k) 6= 0, then the infinite product of Cm,n(k) converges. We already have
det Cm,n(k) 6= 0 for all k, m, n, so consider
DIRAC TYPE OPERATORS ON THE QUANTUM SOLID TORUS
11
Cm,n(k) − I =
−
1
c1,n(k)−1
m
an(k+1)c1,n(k)
−
an+1(k)c1,n(k)
m
an(k+1)an+1(k)c1,n(k) ! .
m2
c2,n(k) − 1 +
Estimating the matrix norm by the sum of absolute values of the coefficients, we have
kCm,n(k) − Ik ≤(cid:12)(cid:12)(cid:12)(cid:12)
1
c1,n(k) − 1(cid:12)(cid:12)(cid:12)(cid:12)
+
m
+
m
an+1(k)c1,n(k)
an(k + 1)c1,n(k)
+ c2,n(k) − 1 +
m2
an(k + 1)an+1(k)c1,n(k)
.
By assumption, the infinite products of c1,n(k) and c2,n(k) exist. Using this and the summa-
bility criteria on 1/an(k), we see thatPk kCm,n(k) − Ik converges. Therefore we can deduce
that the infinite product Qk Cm,n(k) exists, thus completing proof.
Below we use simplified notation for the products:
(cid:3)
J1(n) =
∞
Yk=0
c1,n(k)
and J2(n) =
c2,n(k).
∞
Yk=0
We have
k
k
det Cm,n = lim
k→∞
det Cm,n(i) = lim
k→∞
Yi=0
c2,n(i)
c1,n(i)
Yi=0
= Q∞
Q∞
i=0 c2,n(i)
i=0 c1,n(i)
=
J2(n)
J1(n)
.
The next proposition describes the structure of the infinite product of the Cm,n(k) matrices.
Proposition 5.2. The infinite product Cm,n has the following structure:
Cm,n =(cid:18) Q∞
i=0
1
c1,n(i) + F0(m2)
−mF2(m2)
−mF1(m2)
i=0 c2,n(i) + F3(m2) (cid:19) ,
Q∞
where the Fj(m2) are power series in m2 with positive coefficients for j = 0, 1, 2, 3, and
additionally
∞
1
form
c1,n(i)
F3(m2) = ∞
Yi=0
+ F0(m2)!−1 m2F1(m2)F2(m2) − F0(m2)
Yi=0
Proof. It is easy to verify by induction that for each k the product Qk
Yi=0
Cm,n(i) = Qk
i=0 u2(n, k)(m2)i
i=0 u0(n, k)(m2)i
−mPk
i=0 c2,n(i) +Pk
Qk
c1,n(i) +Pk
−mPk
i=0
k
1
where the above sums are polynomials in m2 with positive coefficients. Moreover as k →
∞, the polynomials converge to a power series, by the Weierstrass Analytic Convergence
Theorem, see [1]. Thus the result follows.
(cid:3)
c2,n(i)! .
i=0 Cm,n(i) is of the
i=0 u1(n, k)(m2)i
i=0 u3(n, k)(m2)i ! ,
12
SLAWOMIR KLIMEK AND MATT MCBRIDE
6. The Class of Boundary Conditions
As defined above, the operator D is not a Fredholm operator on its maximal domain: it
has an infinite dimensional kernel, as is typical for differential operators on manifolds with
boundary. In this section we single out two types of solutions in the kernel of Am,n that play
the role of similar solutions in [8] made of modified Bessel functions of first and second kind.
Like in the commutative case, they are then used to define a class boundary conditions for
D that turns D into an invertible operator. First we describe the kernel of D through the
kernel of Am,n.
Proposition 6.1. Let Am,n be the operator given by equation (4.5), then
for some vector v.
Ker Am,n =( k−1
Yi=0
Cm,n(i)! v)
Proof. By formula (4.5) we need to solve the equation:
Am,n(k + 1)(cid:20)(cid:18) f (k + 1)
g(k + 1) (cid:19) − Cm,n(k)(cid:18) f (k)
g(k) (cid:19)(cid:21) =(cid:18) 0
0 (cid:19) .
Solving recursively we see that
for some arbitrary vector v.
g(k + 1) (cid:19) = k
(cid:18) f (k + 1)
Yi=0
Cm,n(i)! v
(cid:3)
If Fm,n(k) ∈ Ker Am,n, then by Proposition 5.1 limk→∞ Fm,n(k) = Fm,n(∞) exists and is
finite.
We define Im,n(k) to be a special element of Ker Am,n, so:
Im,n(k) = k−1
Yi=0
Cm,n(i)! Im,n(0) = k−1
Yi=0
Cm,n(i)! I (1)
m,n(0) ! ,
m,n(0)
I (2)
(6.1)
which additionally satisfies the initial regularity equation (4.10) with zero right-hand side.
We normalize Im,n(k) in the following way: we set
I (1)
m,n(0) = −1
and I (2)
m,n(0) =
m
an(0)
.
We also want to define a class of complementary solutions, denoted Km,n(k), for which
Km,n(k) = k−1
Yi=0
Cm,n(i)! Km,n(0) = k−1
Yi=0
Cm,n(i)! K (1)
m,n(0) ! ,
m,n(0)
K (2)
(6.2)
and such that for every k, n ≥ 0 and m ∈ Z, the set {Im,n(k), Km,n(k)} is linearly inde-
pendent. This can be achieved by requesting the following first three sign conditions at
infinity:
DIRAC TYPE OPERATORS ON THE QUANTUM SOLID TORUS
13
1.) m > 0 : K (1)
2.) m < 0 : K (1)
3.) m = 0 : K (1)
K (1)
K (2)
4.) m 6= 0 :
m,n(∞) > 0 and K (2)
m,n(∞) < 0 and K (2)
0,n(∞) = 0 and K (2)
m,n(∞)
m,n(∞)
m,n(∞) > 0
m,n(∞) > 0
0,n(∞) 6= 0
→ 0 as
m → ∞ uniformly in n.
(6.3)
Additionally, the sign conditions, and the normalization of Im,n, lead to monotonicity prop-
erties of Im,n and Km,n that are crucially used in estimates in the last section. The key
asymptotic fourth condition above is also required to prove compactness of the resolvent.
Any collection Km,n(k) satisfying (6.2) and (6.3) will be referred to, abusing the terminology,
as a quantum K function.
We have:
Proposition 6.2. Let Km,n(k) be any quantum K function. For every k, n ≥ 0 and m ∈ Z
the vectors Im,n(k) and Km,n(k) are linearly independent in C2. Moreover for all n ≥ 0 and
m ∈ Z, Im,n(∞) and Km,n(∞) are also linearly independent.
Proof. Consider the case m > 0. Recall that we have
Im,n(k) = k−1
Yi=0
Cm,n(i)!(cid:18) −1
an(0) (cid:19) and Km,n(k) = k−1
Yi=0
m
Cm,n(i)! K (1)
m,n(0) ! .
m,n(0)
K (2)
Using the formulas for the product of Cm,n(k) in the proof of Proposition 5.2 we write out
the components to get
u0(n, k)(m2)i +
m
an(0) k
Yi=0
i=0 u1(n, k)(m2)i
an(0)
! ,
u3(n, k)(m2)i! .
m2Pk
Xi=0
k
c2,n(i) +
I (1)
m,n(k) = − k
Yi=0
Xi=0
I (2)
m,n(k) = m
k
1
c1,n(i)
+
k
Xi=0
u2(n, k)(m2)i +
Since all coefficients are positive and m > 0, we see that I (1)
positive.
m,n(k) is negative and I (2)
m,n(k) is
To analyze Km,n(k) we use an alternative, equivalent, way to write it as:
Km,n(k) = ∞
Yi=k
Cm,n(i)!−1
Km,n(∞).
Since the components of Km,n(∞) are both positive, and the matrix Cm,n(i)−1 has all positive
entries for m > 0, it follows that K (1)
m,n(k) are positive. Therefore since one of
the components of Im,n(k) is negative and both components of Km,n(k) are positive, it is
impossible for them to be linearly dependent.
m,n(k) and K (2)
In the case m < 0 both components of Im,n(k) are negative and one of the components
of Km,n(k) is positive again showing they can not be linearly dependent. When m = 0
14
SLAWOMIR KLIMEK AND MATT MCBRIDE
then all matrices Cm,n(k) are diagonal and so K (1)
m,n(k) = 0, which implies
independence. Finally the same arguments are valid when k = ∞, thus the proof is complete.
(cid:3)
m,n(k) = 0 and I (2)
Now we can define the class of boundary conditions we consider for D. Given an F ∈ H
such that DF ∈ H, the Fourier coefficients fm,n+1(k) and gm,n(k) given by equation (4.2),
have limits as k → ∞ for all n ≥ 0 and m ∈ Z. This follows from the calculation of the
resolvent in the next section, see equation (7.3) for details. These limits will be denoted by
fm,n+1(∞) and gm,n(∞).
Any quantum K function defines a boundary condition in the following way.
Definition: Given Km,n(k) satisfying (6.2) and (6.3) we set the domain dom(D) of D to be
the space of all F ∈ H such that DF ∈ H and such that for all m ∈ Z, n ≥ 0 there exist
βm,n ∈ C such that
gm,n(∞) = βm,nK (1)
fm,n+1(∞) = βm,nK (2)
m,n(∞)
m,n(∞).
(6.4)
The above conditions at infinity can be restated to mimic the boundary condition in the
classical case:
gm,n(∞)K (2)
m,n(∞) − fm,n+1(∞)K (1)
m,n(∞) = 0.
The following theorem is the main result of this paper and is proved at the end of this
paper.
Theorem 6.3. The quantum Dirac operator D, defined by (4.3) subject to the boundary
condition defined in equation (6.4), is an invertible operator whose inverse Q, is a compact
operator.
We end this section with listing of useful recurrence relations for Im,n(k) and Km,n(k)
that are easy consequences of their definitions. Let Hm,n(k) be Im,n(k) or Km,n(k) with
components Hm,n(k) = (H (1)
m,n(k))t, then we have
m,n(k), H (2)
Hm,n(k + 1) = Cm,n(k)Hm,n(k).
(6.5)
This gives the following recurrence relation:
H (1)
m,n(k + 1) −
1
c1,n(k)
H (1)
m,n(k) = −
H (2)
m,n(k + 1) − c2,n(k)H (2)
m,n(k) = −
H (1)
m,n(k)
(6.6)
H (2)
m,n(k)
m
an+1(k)c2,n(k)
m
an(k + 1)c1,n(k)
+
m2
an+1(k)an(k + 1)c1,n(k)
H (2)
m,n(k).
Similarly, using the relation
we can produce the following two (equivalent) equations:
Hm,n(k) = (Cm,n(k))−1Hm,n(k + 1),
DIRAC TYPE OPERATORS ON THE QUANTUM SOLID TORUS
15
H (2)
m,n(k) −
1
c2,n(k)
H (2)
m,n(k + 1) =
m
an(k + 1)c2,n(k)
H (1)
m,n(k + 1)
H (1)
m,n(k) − c1,n(k)H (1)
m,n(k + 1) =
m
an+1(k)c2,n(k)
H (2)
m,n(k + 1)+
(6.7)
+
m2
an+1(k)an(k + 1)c2,n(k)
H (1)
m,n(k + 1).
The recurrence relations above are extensively used in the next two section in the analysis
of the parametrix of D.
7. Parametrix of the quantum Dirac type operator
In the previous section we introduced a new domain of D. To account for this, we redefine
an(Z≥0) ⊗
the domain of Am,n in the following way: dom(Am,n) is the space of all h ∈ ℓ2
ℓ2
an+1(Z≥0) such that kAm,nhkan+1⊗an < ∞, and such that there exists β ∈ C such that
x(∞) = βK (1)
y(∞) = βK (2)
m,n(∞)
m,n(∞),
(7.1)
where h = (x, y)t. The existence of such limits x(∞) and y(∞) follows from the calculation
of the resolvent in this section, see equation (7.3).
Proposition 7.1. Given h ∈ dom Am,n, if h ∈ Ker Am,n and h = (x, y)t satisfies
an(0)y(0) + mx(0) = 0,
then h(k) = 0 for every k ≥ 0. In other words, the kernel of Am,n subject to initial regularity
condition (4.10) is trivial.
Proof. First notice that we can write any element h(k) ∈ Ker Am,n as
h(k) = c1Im,n(k) + c2Km,n(k),
since by Proposition 6.2, the nonzero two-vectors Im,n(k) and Km,n(k) are linearly indepen-
dent. However if k = 0 then both h(0) and Im,n(0) satisfy the equation an(0)y(0)+mx(0) = 0
so there exists some constant c3 such that:
c3Im,n(0) = h(0) = c1Im,n(0) + c2Km,n(0).
This would imply that Km,n(0) is a scalar multiple of Im,n(0), which is impossible by the
linear independence of the solutions, hence c2 = 0. If k = ∞ then, since h ∈ dom Am,n,
there is some constant c4, such that:
c4Km,n(∞) = h(∞) = c1Im,n(∞),
which implies that Km,n(∞) is a scalar multiple of Im,n(∞), but again this is impossible by
Proposition 6.2, hence c1 = 0. Therefore h(k) = 0 for every k. This completes the proof. (cid:3)
16
SLAWOMIR KLIMEK AND MATT MCBRIDE
Next we discuss the non-homogeneuous equation DF = G, the solution of which leads to
the parametrix (in our case the inverse) of the quantum Dirac type operator D, subject to
boundary conditions (6.4).
We use the following notation: for a vector v = (v1, v2)t we write v⊥ := (v2, −v1)t.
Proposition 7.2. Let Am,n be the one-step matrix difference operator defined by equation
(4.5), then Am,n, subject to the boundary conditions given by the equation (7.1), and subject
to the initial regularity condition (4.10) is an invertible operator with inverse Q(m,n) given
by (7.7) below.
Proof. The goal is to solve the equation
which becomes the following difference equation with matrix coefficients
Am,n(cid:18) x
y (cid:19) (k) =(cid:18)
p(k)
−q(k + 1) (cid:19) ,
Am,n(k + 1)(cid:20)(cid:18) x(k + 1)
y(k + 1) (cid:19) − Cm,n(k)(cid:18) x(k)
y(k) (cid:19)(cid:21) =(cid:18)
p(k)
−q(k + 1) (cid:19)
with Am,n(k + 1) and Cm,n(k) defined in formulas (4.6) and (4.7) respectively, while the
initial regularity condition is
an(0)y(0) + mx(0) = q(0).
Relabeling h(k) = (x(k), y(k))t and r(k + 1) = (p(k), −q(k + 1))t, the system becomes
Am,n(k + 1)(h(k + 1) − Cm,n(k)h(k)) = r(k + 1).
Before applying the boundary condition, the solution h(k) is not unique, with non-uniqueness
due to one-dimensional kernel Im,n(k). With this in mind we choose h(0) to be the following
particular solution of the initial regularity condition:
Solving the difference equation by using variation of constants method we get
h(0) =(cid:18)0,
q(0)
an(0)(cid:19)t
.
Q(m,n)r(k) := h(k) =
Cm,n(i)
Cm,n(j)!−1
(Am,n(i))−1 r(i) + αIm,n(k)
(7.2)
k−1
Yi=0
k
Xi=0 i−1
Yj=0
for some parameter α. Here, for convenience, we set the Qk−1
an(0)(cid:17)t
and also we define (Am,n(0))−1 r(0) := h(0) =(cid:16)0, q(0)
j=k Cm,n(j) = 1 for any m and n,
from above. To apply the boundary
conditions in equation (6.4), we need to know that Q(m,n)r(∞) is well defined; this follows
from Proposition 5.1 and the summability of Am,n(k)−1. Therefore the boundary condition,
equation (6.4), is well defined, and gives:
Q(m,n)r(∞) = Cm,n
∞
Xi=0 i−1
Yj=0
= βKm,n(∞)
Cm,n(j)!−1
(Am,n(i))−1 r(i) + αIm,n(∞) =
(7.3)
DIRAC TYPE OPERATORS ON THE QUANTUM SOLID TORUS
17
for some constant β. The goal is to solve for α. Below we use the following notation:
(7.4)
Multiplying by (Cm,n)−1 and taking the inner product of both sides of equation (7.3) with
τm,n := hKm,n(0), Im,n(0)⊥i.
Km,n(0)⊥, we get
* ∞
Xi=0 i−1
Yj=0
Cm,n(j)!−1
(Am,n(i))−1 r(i), Km,n(0)⊥+ + ατm,n = 0,
which can now be solved for α:
α =
−1
τm,n * ∞
Xi=0 i−1
Yj=0
Cm,n(j)!−1
(Am,n(i))−1 r(i), Km,n(0)⊥+ .
There is a partial cancellation between the two terms in equation (7.2). To see this,
we notice that, since Im,n(0) and Km,n(0) are linearly independent, we can decompose any
two-vector v as
where
v = v1Im,n(0) + v2Km,n(0),
v1 = −(cid:10)v, Km,n(0)⊥(cid:11)
τm,n
Applying this decomposition to vector
and
v2 = (cid:10)v, Im,n(0)⊥(cid:11)
τm,n
,
v =
k
Xi=0 i−1
Yj=0
Cm,n(j)!−1
(Am,n(i))−1 r(i),
and using the formula for α, and the formulas for Im,n(k) and Km,n(k), equations (6.1) and
(6.2) respectively, we get
k−1
1
Cm,n(i)
Q(m,n)r(k) =
Yi=0
Xi=k+1 i−1
τm,n * ∞
Yj=0
Cm,n(j)!−1
Xi=0 i−1
Yj=0
×
−
τm,n * k
+
1
Cm,n(j)!−1
(Am,n(i))−1 r(i), Km,n(0)⊥+ Im,n(0)
(Am,n(i))−1 r(i), Im,n(0)⊥+ Km,n(0)
.
(Am,n(i))−1 r(i), Km,n(0)⊥+
Cm,n(j)!−1
(7.5)
The coefficients in the formula above will be denoted by:
e(1)
m,n(k) :=
and
−1
τm,n * ∞
Xi=k+1 i−1
Yj=0
18
SLAWOMIR KLIMEK AND MATT MCBRIDE
e(2)
m,n(k) :=
1
τm,n * k
Xi=0 i−1
Yj=0
Cm,n(j)!−1
(Am,n(i))−1 r(i), Im,n(0)⊥+ .
With this notation we get the formula for Q(m,n):
Q(m,n)r(k) = e(1)
m,n(k)Im,n(k) + e(2)
k
Q(m,n)r(k) =
The proof is complete.
Xi=0 (cid:18) 0
0
1
an(i)Qk−1
m,n(k)Km,n(k) m 6= 0
0
j=i c2,n(j) (cid:19) r(i) m = 0.
(7.6)
(7.7)
(cid:3)
As computed above, the operator Q(m,n) is not easy to analyze, mostly because it contains
products of non-commuting matrices Cm,n(k). Our strategy is to re-write the formulas for
Q(m,n) in terms of quantum I and K functions, similarly to the commutative case, and then
estimate the Hilbert-Schmidt norm of the parametrics using the recurrence relations (6.6)
and (6.7).
We need the following observations. For vectors v = (v1, v2)t and u = (u1, u2)t we have
hv⊥, ui = −hv, u⊥i, where h·, ·i is the standard Euclidean inner product on R2. If R is a 2 × 2
matrix and v = (v1, v2)t, then we have (Rv)⊥ = (det R) (Rt)−1 v⊥.
As a consequence of the above and using the formula for the determinant of the Cm,n(k)
matrix, we get:
Im,n(k)⊥ =
Km,n(k)⊥ =
k−1
Yi=0
Yi=0
k−1
c2,n(i)
c1,n(i)
c1,n(i)
c2,n(i)
k−1
Yi=0
k−1
Yi=0
t
t
Cm,n(i)!−1
Cm,n(i)!−1
Im,n(0)⊥
Km,n(0)⊥.
(7.8)
Consequently, inserting equations (7.8) into formulas for the coefficients e(1)
(7.6), gives:
m,n, e(2)
m,n, (7.5) and
e(1)
m,n(k) =
1
τm,n
Similarly we also have:
∞
Xi=k+1
i−1
Yj=0
c1,n(j)
c2,n(j)
(K (2)
m,n(i), −K (1)
m,n(i)) (Am,n(i))−1 r(i).
e(2)
m,n(k) =
1
τm,n
c1,n(j)
c2,n(j)
(I (2)
m,n(i), −I (1)
m,n(i)) (Am,n(i))−1 r(i).
k
i−1
Xi=0
Yj=0
These are to be understood as a row of a matrix times a column vector.
It is convenient to rewrite Q(m,n) in the matrix notation:
DIRAC TYPE OPERATORS ON THE QUANTUM SOLID TORUS
19
Q(m,n)rn(k) = e(1)
m,n(k)Im,n(k) + e(2)
m,n(k)Km,n(k)
= (Im,n(k), Km,n(k)) e(1)
m,n(k) ! = I (1)
m,n(k)
e(2)
m,n(k) K (1)
I (2)
m,n(k) K (2)
m,n(k)
m,n(k) ! e(1)
m,n(k) ! .
m,n(k)
e(2)
Inserting the above formulas for the coefficients e(1)
m,n, e(2)
m,n we get:
Q(m,n)rn(k) =
j=0
i=k+1Qi−1
P∞
i=0Qi−1
Pk
j=0
1
c1,n(j)
m,n(k)
τm,n I (1)
m,n(k) K (1)
I (2)
m,n(k) K (2)
c2,n(j)K (2)
c2,n(j)I (2)
m,n(i)
m,n(k) ! ×
i=k+1Qi−1
m,n(i) −P∞
i=0Qi−1
−Pk
c1,n(j)
j=0
j=0
c1,n(j)
m,n(i)
c2,n(j)K (1)
c2,n(j)I (1)
m,n(i) ! (Am,n(i))−1 r(i).
c1,n(j)
Next we focus on multiplying the second matrix with (Am,n(i))−1. We use the recurrence
relations, (6.6) and (6.7), to get:
Q(m,n)rn(k) =
1
τm,n I (1)
m,n(k) K (1)
I (2)
m,n(k) K (2)
an+1(i−1) K (2)
an+1(i−1) I (2)
m,n(k)
m,n(k) ! ×
i=k+1Qi−1
m,n(i − 1) P∞
i=0Qi−1
Pk
m,n(i − 1)
j=0
1
1
j=0
j=0
i=k+1Qi−2
P∞
i=0Qi−2
Pk
j=0
c1,n(j)
c2,n(j)
c1,n(j)
c2,n(j)
c1,n(j)
c2,n(j)
c1,n(j)
c2,n(j)
1
m,n(i)
an(i) K (1)
an(i) I (1)
m,n(i) ! r(i).
1
(7.9)
This formula finally expresses Q(m,n) through the quantum I and K functions. Its primary
benefit is that it shows that the entries the matrix of Q(m,n) are integral operators in weighted
ℓ2 spaces. Below we explicitly write down those operators.
For α, β = 1, 2, m 6= 0 and n ≥ 0, we define the following integral operators X αβ
ℓ2
an−1+β (Z≥0) → ℓ2
an−1+α(Z≥0) by:
m,n, Y αβ
m,n :
X αβ
m,nr(k) = I (α)
m,n(k)
Y αβ
m,nr(k) = K (α)
m,n(k)
m,n(i − β + 1)
an−1+β(i − β + 1)
r(i)
(7.10)
k
∞
c1,n(j)
c1,n(j)
Xi=k+1 i−β
c2,n(j)! K (β)
Yj=0
Xi=0 i−β
c2,n(j)! I (β)
Yj=0
Xi=0 k−1
Yj=i
c2,n(j)! r(i)
an(Z≥0) → ℓ2
k
an(i)
m,n(i − β + 1)
an−1+β(i − β + 1)
r(i).
Z0,nr(k) =
.
(7.11)
Also, for n ≥ 0, define integral operator Z0,n : ℓ2
an+1(Z≥0) by
Lemma 7.3. The parametrix, Q(m,n) for the operator Am,n from above for m 6= 0 and n ≥ 0
is given by the following equivalent formula:
Q(m,n)(cid:18) r(1)(k)
r(2)(k) (cid:19) =
1
τm,n p(1)
m,n(k)
p(2)
m,n(k) ! ,
20
where
SLAWOMIR KLIMEK AND MATT MCBRIDE
p(1)
m,n(k) = X 12
p(2)
m,n(k) = X 22
m,nr(1)(k) + Y 12
m,nr(1)(k) + Y 22
m,nr(1)(k) + X 11
m,nr(1)(k) + X 21
m,nr(2)(k) + Y 11
m,nr(2)(k) + Y 21
m,nr(2)(k)
m,nr(2)(k).
Moreover when m = 0 and n ≥ 0 the parametrix is given by the following formula:
Q(0,n)(cid:18) r(1)(k)
r(2)(k) (cid:19) =(cid:18)
0
Z0,nr(2)(i) (cid:19) .
Proof. Multiplying out the two matrices in (7.9) and applying it to the vector r(i) =
(cid:0)r(1)(i), r(2)(i)(cid:1), and using the definitions of the integral operators, equation (7.10), gives
the desired result. The case m = 0 immediately follows from equations (7.7) and (7.11).
Thus the proof is complete.
(cid:3)
8. Analysis of the Parametrix
This section contains analysis of the parametrix Q(m,n), culminating in the proof of the
main theorem 6.3.
The case m = 0 is the easiest: all matrices are diagonal and the analysis is not any different
than the analysis for the quantum disk [5] and [7]. Below we only consider the case m > 0
since for m < 0 the computations are virtually identical. We start by gathering the basic
information about the quantum I and K functions.
Lemma 8.1. For m > 0 the components −I (1)
positive for all n, k ≥ 0.
m,n(k), and K (2)
m,n(k), K (1)
m,n(k), I (2)
m,n(k) are all
Proof. This immediately follows from the definitions and the proof of Proposition 6.2.
(cid:3)
Lemma 8.2. We have the following inequalities for all n, k ≥ 0 and m > 0:
− I (1)
m,n(k) < −I (1)
m,n(k + 1),
I (2)
m,n(k) <
1
c2,n(k)
I (2)
m,n(k + 1),
K (1)
m,n(k + 1) <
1
c1,n(k)
K (1)
m,n(k), K (2)
m,n(k + 1) < K (2)
m,n(k).
Proof. Using recurrence relation (6.6), (6.7), the assumption that m > 0 and the previous
lemma we have
0 <
m
an(k + 1)
K (1)
m,n(k + 1) = c2,n(k)K (2)
m,n(k) − K (2)
m,n(k + 1).
Consequently, using c2,n(k) ≤ 1 , we have
m,n(k + 1) < c2,n(k)K (2)
proving one of the inequalities. Similarly, we have
K (2)
m,n(k) ≤ K (2)
m,n(k),
0 <
m
an+1(k)c(k)
K (2)
m,n(k) =
1
c1,n(k)
K (1)
m,n(k) − K (1)
m,n(k + 1),
implying the other inequality involving Km,n. The proofs for the Im,n(k) are identical. (cid:3)
DIRAC TYPE OPERATORS ON THE QUANTUM SOLID TORUS
21
The next lemma is the crux of the compactness argument.
It establishes estimates on
the components of quantum I and K functions in a similar fashion to the estimates of the
modified Bessel functions in [8]. Consider the quantity:
The series above is convergent by summability assumptions on an(k).
ε(m, n) =
an+1(k)
m2 + an(k)an+1(k)
.
∞
Xk=0
Lemma 8.3. For m 6= 0 and n, k ≥ 0 the following inequalities hold:
I (2)
m,n(k) ≤ −mε(m, n)I (1)
m,n(k+1) and K (1)
m,n(k+1) ≤ m ε(m, n) +
K (1)
K (2)
m,n(∞)
m,n(∞)! K (2)
m,n(k).
Moreover ε(m, n) → 0 as m, n → ∞.
Proof. Assuming m > 0 and using the recurrence relations (6.6), (6.7), we have
I (2)
m,n(k)
−I (1)
m,n(k + 1)
=
m
an+1(k) +
c1,n(k)
1
m
an(k) +c2,n(k−1)
(2)
m,n (k−1)
I
−I
(1)
m,n(k)
≤ c1,n(k)
1 +
m
an(k)
m2
an(k)an+1(k)
+ c2,n(k − 1)
I (2)
m,n(k − 1)
−I (1)
m,n(k) !
≤
1 +
m
an(k)
m2
an(k)an+1(k)
+
I (2)
m,n(k − 1)
−I (1)
m,n(k)
,
since c1,n(k) ≤ 1 and c2,n(k) ≤ 1 for all k ≥ 0, assuming I (2)
rearranging the terms, that
m,n(−1) = 0.
It follows, by
m,n(k)
1
m I (2)
−I (1)
m,n(k + 1)
−
I (2)
m,n(k − 1)
−I (1)
m,n(k) ! ≤
1 +
1
an(k)
m2
.
an(k)an+1(k)
Summing both sides and telescoping the left side we get
1
m,n(k)
m I (2)
m,n(k + 1)! ≤
−I (1)
1
an(k)
m2
an(k)an+1(k)
1 +
= ε(m, n).
∞
Xk=0
This establishes the first inequality. To obtain the second inequality we proceed similarly,
using the recurrence relations (6.6), (6.7), but this time telescoping sum goes to infinity.
The next step is to show ε(m, n) goes to zero as m and n go to infinity.
It follows
immediately from the definition that:
and the sum on the right goes to zero as n → ∞ from the condition on an(k), and hence
ε(m, n) → 0 as n → ∞. Now for any η > 0 pick N > 0 such that we have
ε(m, n) ≤
1
an(k)
,
∞
Xk=0
22
SLAWOMIR KLIMEK AND MATT MCBRIDE
and pick M > 0 such that
1
an(k)
≤
η
2
,
Xk>N
1
an(k)
m2
an(k)an+1(k)
≤
η
2
1 +
Xk≤N
for m > M. It now follows that ε(m, n) → 0 as n, m → ∞. The case m < 0 is analogous,
and thus the desired result follows.
(cid:3)
The next technical lemma deals with sums appearing in the definitions of the integral
operators which comprise the parametrix.
Lemma 8.4. The following summation estimates are true for m 6= 0 and n, k ≥ 0:
∞
Xi=k+1 i−2
Yj=0
c1,n(j)! K (2)
m,n(i − 1)
an+1(i − 1)
≤
1
m
k−1
Yj=0
c1,n(j)K (1)
m,n(k) and
K (1)
m,n(i)
an(i)
≤
1
m
K (2)
m,n(k).
∞
Xi=k+1
Proof. As usual we consider only m > 0. Using equation (6.6), multiplying both sides of the
equation by a product of c1,n(j)'s, and summing we get:
∞
Xi=k+1 i−2
Yj=0
c1,n(j)! K (2)
m,n(i − 1)
an+1(i − 1)
=
≤
1
m
1
m
c1,n(j)K (1)
m,n(i − 1) −
c1,n(j)K (1)
m,n(i)!
i−1
Yj=0
∞
Xi=k+1 i−2
Yj=0
Yj=0
k−1
c1,n(j)K (1)
m,n(k),
where the last inequality is true because the difference above is a telescoping sum. The
second inequality follows from the recurrence relation (6.7) and the same telescoping sum
trick. This completes the proof.
(cid:3)
The next and final lemma in preparation for the proof of the main result deals with
pointwise estimates of the products of different components of quantum I and K functions
through the quantity τm,n = hKm,n(0), Im,n(0)⊥i.
Lemma 8.5. If m 6= 0 then, for all n ≥ 0,
K (1)
m,n(k)I (2)
m,n(k) ≤ τm,n
−I (1)
m,n(k + 1)K (2)
m,n(k) ≤
τm,n
c1,n(k)
c2,n(i)
c1,n(i)
c2,n(i)
c1,n(i)
k−1
k−1
Yi=0
Yi=0
and − K (2)
m,n(k)I (1)
m,n(k) ≤ τm,n
k−1
Yi=0
and I (2)
m,n(k)K (1)
m,n(k + 1) ≤
τm,n
c1,n(k)
c2,n(i)
c1,n(i)
c2,n(i)
c1,n(i)
.
k−1
Yi=0
DIRAC TYPE OPERATORS ON THE QUANTUM SOLID TORUS
23
Proof. Assume as usual that m > 0. Using the matrix valued recurrence relation (6.5) and
properties of the inner product we get
hKm,n(k + 1), Im,n(k + 1)⊥i =
=
c2,n(i)
c1,n(i)* k
Yi=0
Cm,n(i)Km,n(0),
c2,n(i)
c1,n(i)
hKm,n(0), Im,n(0)⊥i =
k
k
Yi=0
Yi=0
Writing out the above inner product gives
k
Yi=0(cid:0)Cm,n(i)−1(cid:1)t Im,n(0)⊥+
Yi=0
c2,n(i)
c1,n(i)
τm,n.
k
(8.1)
hKm,n(k + 1), Im,n(k + 1)⊥i = K (1)
m,n(k + 1)I (2)
m,n(k + 1) + K (2)
m,n(k + 1)[−I (1)
m,n(k + 1)].
Consequently, the above equality (8.1) implies that
K (1)
m,n(k)I (2)
m,n(k) ≤ τm,n
c2,n(i)
c1,n(i)
and − K (2)
m,n(k)I (1)
m,n(k) ≤ τm,n
c2,n(i)
c1,n(i)
.
k−1
Yi=0
k−1
Yi=0
To get the second set of inequalities we use the above equation (8.1) and recurrence
relations (6.6), (6.7) to get
τm,n
c2,n(i)
c1,n(i)
k−1
Yi=0
= −I (1)
m,n(k)K (2)
m,n(k) + I (2)
m,n(k)K (1)
m,n(k)
m,n(k + 1) −
m
an+1(k)
m,n(k + 1) +
+ I (2)
=(cid:18)−c1,n(k)I (1)
m,n(k)(cid:18)c1,n(k)K (1)
= c1,n(k)(cid:0)−I (1)
I (2)
m
K (2)
m,n(k)
m,n(k)(cid:19) K (2)
m,n(k)(cid:19)
m,n(k + 1)(cid:1) .
m,n(k)K (1)
an+1(k)
m,n(k + 1)K (2)
m,n(k) + I (2)
From the last equality it follows that:
I (1)
m,n(k + 1)K (2)
m,n(k) ≤
τm,n
c1,n(k)
and the lemma is proved.
c2,n(i)
c1,n(i)
k−1
Yi=0
and I (2)
m,n(k)K (1)
m,n(k + 1) ≤
τm,n
c1,n(k)
k−1
Yi=0
c2,n(i)
c1,n(i)
,
(cid:3)
We have gathered now enough information to analyze compactness of the inverse of D. The
Fourier transform decomposes that inverse into a direct sum of parametrices Qm,n, which in
turn consist of the integral operators X αβ
m,n, Z0,n. Those integral operators are compact
operators, in fact even Hilbert-Schmidt operators as stated in the following theorem.
m,n, Y αβ
Theorem 8.6. Let n ≥ 0. If m 6= 0, then the integral operators X αβ
m,n defined in
equation (7.10) are Hilbert-Schmidt operators for α, β = 1, 2, and, if m = 0, the integral
m,n and Y αβ
24
SLAWOMIR KLIMEK AND MATT MCBRIDE
operator Z0,n defined in equation (7.11) is a Hilbert-Schmidt operator. Moreover the Hilbert-
Schmidt norms of X αβ
m,n and Y αβ
m,n go to zero as m, n → ∞, and the Hilbert-Schmidt norm
of Z (0,n) goes to zero as n → ∞.
Proof. We start with the case m 6= 0. Using the definition of X αβ
that for α, β = 1, 2 we have
m,n and Y αβ
m,n, it is easy to see
kX αβ
m,nk2
HS =
kY αβ
m,nk2
HS =
an−1+α(k)
m,n(k)(cid:17)2
(cid:16)I (α)
m,n(k)(cid:17)2
(cid:16)K (α)
an−1+α(k)
∞
Xk=0
Xk=0
∞
∞
c1,n(j)
c2,n(j)!2
Xi=k+1 i−β
Yj=0
c2,n(j)!2
Xi=0 i−β
Yj=0
c1,n(j)
k
an−1+β(i − β + 1)
m,n(i − β + 1)(cid:17)2
· (cid:16)K (β)
m,n(i − β + 1)(cid:17)2
· (cid:16)I (β)
an−1+β(i − β + 1)
.
(8.2)
There are eight norms to estimate; however that number can be reduced to four, since by
Fubini's Theorem we have:
kX 11
kX 12
m,nk2
m,nk2
HS = kY 11
HS = kY 21
m,nk2
m,nk2
HS,
HS,
kX 22
kX 21
m,nk2
m,nk2
HS = kY 22
HS = kY 12
m,nk2
m,nk2
HS
HS.
As usual consider the case m > 0. Using Lemmas 8.2, 8.4, 8.5 in the formula for the
Hilbert-Schmidt norm of X 11
m,n, we obtain:
kX 11
m,nk2
HS =
≤
≤
∞
∞
Xk=0
J1(n)2
J2(n)2 κ
an(k)
m,n(k)(cid:17)2
(cid:16)I (1)
Xi=k+1 i−1
Yj=0
m,n(k)(cid:17)2
(cid:16)I (1)
h−I (1)
Xk=0
J1(n)
J2(n)
τm,nκ
m
∞
∞
·
an(k)
Xk=0
c1,n(j)
c2,n(j)!2
· (cid:16)K (1)
m,n(i)(cid:17)2
an(i)
K (1)
m,n(k + 1)
K (1)
m,n(i)
an(i)
∞
Xi=k+1
m,n(k)i K (1)
an(k)
m,n(k + 1)
.
We estimate the last term in the following way:
τm,nκJ1(n)
mJ2(n)
h−I (1)
∞
Xk=0
≤
J2(n)
J1(n)
τ 2
m,nκJ1(n)
J2(n)
= τ 2
m,nκ ε(m, n) +
m,n(k + 1)
∞
an(k)
m,n(k)i K (1)
Xk=0 ε(m, n) +
m,n(∞)! s(n),
K (1)
K (2)
m,n(∞)
·
K (2)
K (2)
m,n(k)
m,n(k)
K (1)
K (2)
m,n(∞)
m,n(∞)! 1
an(k)
by using Lemmas 8.3, 8.5, and the inequality 1/c1,n(k) ≤ κ. Therefore we have
DIRAC TYPE OPERATORS ON THE QUANTUM SOLID TORUS
25
kX 11
m,nk2
HS ≤ τ 2
m,nκ ε(m, n) +
K (1)
K (2)
m,n(∞)
m,n(∞)! s(n).
Similarly, using Lemmas 8.2, 8.4, and 8.5, we can estimate the Hilbert-Schmidt norm of
m,n by:
X 22
kX 22
m,nk2
HS =
∞
Xk=0
∞
an+1(k)
m,n(k)(cid:17)2
(cid:16)I (2)
Xi=k+1 i−2
Yj=0
m,n(k)(cid:17)2
(cid:16)I (2)
an+1(k)
∞
J1(n)
J2(n)2
Xk=0
an+1(i − 1)
c1,n(j)
c2,n(j)!2 (cid:16)K (2)
Xi=k+1 i−2
Yj=0
m,n(i − 1)(cid:17)2
c1,n(j)! K (2)
∞
K (2)
m,n(k)
m,n(i − 1)
an+1(i − 1)
≤
≤
J1(n)2
J2(n)2
τm,n
m
·
J2(n)
J1(n)
I (2)
m,n(k)K (2)
an+1(k)
m,n(k)
.
∞
Xk=0
The above is then equal to
τm,n
m
·
J1(n)
J2(n)
∞
Xk=0
m,n(k)K (2)
I (2)
an+1(k)
m,n(k)
·
I (1)
m,n(k + 1)
I (1)
m,n(k + 1)
≤ τ 2
m,nκ
ε(m, n)
an+1(k)
= τ 2
m,nκε(m, n)s(n + 1),
∞
Xk=0
again by using Lemmas 8.3 and 8.5 and the fact that 1/c1,n(k) is less than κ. Thus we have
Very similar arguments, using the same steps as above, show that
kX 22
m,nk2
HS ≤ τ 2
m,nκε(m, n)s(n + 1).
kX 12
m,nk2
HS ≤ τ 2
m,nκ ε(m, n) +
K (1)
K (2)
m,n(∞)
m,n(∞)! s(n)
and kX 21
m,nk2
HS ≤ τ 2
m,nκε(m, n)s(n + 1).
Therefore for m 6= 0 the Hilbert-Schmidt norms are finite for all of the operators.
When m = 0 we have:
kZ0,nk2
HS =
since again c2,n(k) ≤ 1.
1
an+1(k)
∞
Xk=0
k
Xi=0 k−1
Yj=i
c2,n(j)!2
1
an(i)
≤ s(n)s(n + 1),
By assumption s(n) goes to zero as n → ∞, Lemma 8.3 implies that ε(m, n) → 0 as
m, n → ∞, and the boundary condition given in equation (6.4) requires that
K (1)
K (2)
m,n(∞)
m,n(∞)
→ 0
as
m → ∞.
uniformly in n. Consequently, the Hilbert-Schmidt norms of X αβ
m,n go to zero as
m, n → ∞, and the Hilbert-Schmidt norm of Z (0,n) goes to zero as n → ∞, and the proof
is finished.
(cid:3)
m,n and Y αβ
26
SLAWOMIR KLIMEK AND MATT MCBRIDE
The following statement is an immediate consequence of the previous theorem, since the
parametrix Q(m,n) is comprised of the integral operators estimated in it.
Corollary 8.7. The parametrix Q(m,n) for m 6= 0 and n ≥ 0 is a Hilbert-Schmidt operator
and the Hilbert-Schmidt norm of Q(m,n) goes to zero as m, n → ∞. Moreover for m = 0 the
parametrix Q(0,n) is a Hilbert-Schmidt operator and its Hilbert-Schmidt norm goes to zero as
n → ∞.
We can now close out this section by proving the main theorem of this paper, that the Dirac
operator defined by equation (4.3) subject to the boundary condition (6.4) has a compact
inverse.
Proof. (Proof of Theorem 6.3) It follows from Proposition 4.1 that inverse Q of the Dirac
operator D is (essentially) a direct sum of Q(m,n) and its analogs for "negative" terms in the
Fourier decomposition. Corollary 8.7 shows that Q(m,n) is a Hilbert-Schmidt operator for all
m ∈ Z and n ≥ 0. Moreover the same corollary stipulates that the Hilbert-Schmidt norms
of Q(m,n) go to zero as m, n → ∞. This means that since Q is a direct sum of compact
operators with norms going to zero, which implies that Q is a compact operator. Thus the
proof is complete.
(cid:3)
References
[1] Ahlfors, L., Complex Analysis, McGraw-Hill, 1979.
[2] Atiyah, M. F., Patodi, V. K. and Singer I. M., Spectral asymmetry and Riemannian geometry I, II, III,
Math. Proc. Camb. Phil. Soc. 77, 43-69, 78, 43-432, 79, 71-99, 1975-1976.
[3] Davidson, K., C ∗−algebras by Example, American Mathematical Society, 1991.
[4] Klimek, S. and Lesniewski, A., Quantum Riemann surfaces, I. The unit disk, Comm. Math. Phys., 146,
103-122, 1992.
[5] Klimek, S. and McBride M., D-bar Operators on Quantum Domains. Math. Phys. Anal. Geom., 13, 357
- 390, 2010.
[6] Klimek, S. and McBride M., A note on Dirac Operators on the Quantum Punctured Disk. SIGMA, 6,
56 - 68, 2010.
[7] Klimek, S. and McBride M., A Note on Gluing Dirac Type Operators on a Mirror Quantum Two-Sphere.
SIGMA, 10, 36 - 51, 2014.
[8] Klimek, S. and McBride M., Global boundary conditions for a Dirac operator on the solid torus. Jour.
Math. Phys., 52, 1 - 14, 2011.
[9] Mishchenko, A. V. and Sitenko, Yu, Spectral Boundary Conditions and Index Theorem for Two-
Dimensional Compact Manifold with Boundary. Annals of Physics, 218, 199 - 232, 1992.
[10] Reed, M. and Simon, B., Functional Analysis, Academic Press, 1980.
[11] Trgo, A., Monodromy Matrix for Linear Difference Operators with Almost Constant Coefficients. Jour.
Math. Anal., 194, 697 - 719, 1995.
[12] Williams, D., Crossed Products of C ∗−algebras, American Mathematical Society, 2007.
Department of Mathematical Sciences, Indiana University-Purdue University Indianapo-
lis, 402 N. Blackford St., Indianapolis, IN 46202, U.S.A.
E-mail address: [email protected]
Department of Mathematics and Statistics, Mississippi State University, 175 President's
Cir., Mississippi State, MS 39762, U.S.A.
E-mail address: [email protected]
|
1602.04807 | 1 | 1602 | 2016-02-15T09:53:32 | Woronowicz's Tannaka-Krein duality and free orthogonal quantum groups | [
"math.OA",
"math.CT",
"math.QA"
] | Given a finite dimensional Hilbert space H and a collection of operators between its tensor powers satisfying certain properties, we give a category-free proof of the existence of a compact quantum group G with a fundamental representation U on H such that the intertwiners between the tensor powers of U coincide with the given collection of operators. We then explain how the general version of Woronowicz's Tannaka-Krein duality can be deduced from this. | math.OA | math |
WORONOWICZ'S TANNAKA-KREIN DUALITY AND FREE ORTHOGONAL
QUANTUM GROUPS
SARA MALACARNE
Abstract. Given a finite dimensional Hilbert space H and a collection of operators between its tensor
powers satisfying certain properties, we give a category-free proof of the existence of a compact quantum
group G with a fundamental representation U on H such that the intertwiners between the tensor powers
of U coincide with the given collection of operators. We then explain how the general version of Woronowicz's
Tannaka-Krein duality can be deduced from this.
Introduction
The aim of the paper is to give a category-free proof of Woronowicz's Tannaka-Krein duality Theorem
[5]. We consider a finite dimensional Hilbert space H and a collection of operators between its tensor powers
satisfying certain properties. Categorically speaking, we deal with a C∗-tensor category with conjugates that
is a subcategory of the category of finite dimensional Hilbert spaces, Hilbf , and assume that such category is
generated by one self-conjugate Hilbert space. We prove the existence of a compact quantum group G, such
that its representation category Rep G is our given category. The proof consists of an explicit reconstruction
of the Hopf ∗-algebra C[G], sometimes denoted by Pol G, generated by the coefficients of all finite dimensional
representations of G. The relations defining such Hopf ∗-algebra are directly obtained through morphisms
in the category, or equivalently, through the collection of operators between tensor powers of H. The version
of Woronowicz's Tannaka-Krein Theorem that we prove is essentially formulated in the paper by T. Banica
and R. Speicher [1], where the duality is used for the construction of new examples of free quantum groups,
via subcategories of the so-called categories of noncrossing partitions. Even though the proof here presented
is, in many respects, similar to the proofs of Woronowicz's Tannaka-Krein duality appearing in [5] and [2],
we wish to point out that this version is more algebraic, mostly category-free and the key part is based on
simple duality statements for finite dimensional vector spaces. Related to this reconstruction process, it is
also important to mention P. Schauenburg's paper [3], in which a proof of Tannaka-Krein duality is given
in a more general setting: monoidal categories that are not semisimple are considered and, correspondingly,
arbitrary Hopf algebras are recovered.
The paper is structured in the following way:
in Section 1 we define a bialgebra, which we will later
prove to be the Hopf ∗-algebra C[G]. The relations defining this bialgebra are obtained from the collection
of operators between tensor powers of H, denoted by C. In Section 2 we prove that the bialgebra defined in
Section 1 can be equipped with a Hopf ∗-algebra structure. For this we first consider a smaller collection
of operators, CF , and show that they define the free orthogonal quantum group O+
F , [4]. We want to stress
that nothing, apart from the fact that C[O+
F ] is a well-defined Hopf ∗-algebra, is used. In Section 3 we prove
the equivalence between the C∗-tensor category generated by one selfdual Hilbert space H and Rep G. In
Section 4 we show how the particular case analysed in Section 3 can be extended to the general case of a
not necessarily finitely generated C∗-tensor category.
Acknowledgments. I would like to thank my supervisor Sergey Neshveyev for his help and precious
advice throughout this work. I am grateful to Teodor Banica for useful suggestions. Thanks also to Marco
Matassa for fruitful discussions.
Date: February 15, 2016.
Supported by the European Research Council under the European Union's Seventh Framework Programme (FP/2007-2013)/
ERC Grant Agreement no. 307663 (PI: S. Neshveyev).
1
1. Singly generated categories of Hilbert spaces
Our goal is to prove the following version of Woronowicz's Tannaka-Krein duality.
Theorem 1.1. Let H be a finite dimensional Hilbert space. Suppose we are given a collection C of spaces
C(k, l) of operators H ⊗k → H ⊗l for all k, l ≥ 0 satisfying the following properties:
(1) if T, S ∈ C, then T ⊗ S ∈ C;
(2) if T, S ∈ C are composable, then T S ∈ C;
(3) T ∈ C implies T ∗ ∈ C;
(4) C(k, k) contains the identity operator for all k ≥ 0;
(5) C(0, 2) contains an operator R such that (R∗ ⊗ ι)(ι ⊗ R) = ±ι on H.
Then there exists a unique up to isomomorphism compact quantum group G with a self-conjugate funda-
mental representation U on H such that HomG(H ⊗k, H ⊗l) = C(k, l) for all k, l ≥ 0.
In the last section we will discuss how the general form of Woronowicz's Tannaka-Krein duality can be
easily deduced from this.
Denote by A the tensor algebra of the space of linear functionals on B (H), i.e.,
A := T (cid:0)B (H)∗(cid:1) =
∞
Mk=0
B(cid:0)H ⊗k(cid:1)∗
.
Let U ∈ B (H) ⊗ B(H)∗ ⊂ B (H) ⊗ A be the "fundamental matrix" of A, so U is characterized by the
property
In other words,
(ι ⊗ T )(U ) = T ∀T ∈ B(H).
(1.1)
where the eij 's are matrix units in B (H) and {uij}i,j is the dual basis of B (H)∗ such that uij (ekl) = δikδjl.
The tensor algebra A is a bialgebra with comultiplication ∆ defined by duality from the multiplication on
B(H), so that
eij ⊗ uij,
U :=Xi,j
or equivalently, using the leg-numbering notation, (ι ⊗ ∆)(U ) = U12U13.
∆ (uij) =Xk
uik ⊗ ukj,
Next, denote by An ⊂ A the subspace given by
An :=
and denote by Bn the commutant
Bn :=
n
Mk,l=0
Finally, let
B(cid:0)H ⊗k(cid:1)∗
n
Mk=0
C(k, l)
′
⊆
n
Mk=0
= n
Mk=0
B(cid:0)H ⊗k(cid:1)!∗
B(H ⊗k) ⊂ B n
Mk=0
,
H ⊗k! .
and denote by I the union I :=S∞
Lemma 1.2. I is a bi-ideal in the bialgebra A.
In :=(cid:8)a ∈ An : aBn
= 0(cid:9) ,
n=0 In. Note that In+1 ∩ An = In, so I is a subspace of A.
Proof. We will first prove that I is an ideal. Assume a ∈ In and b ∈ B(H ⊗m)∗; we have to check that a ⊗ b
vanishes on (cid:16)⊕n+m
k,l=mC(k, l)(cid:17)′
. Since (cid:16)⊕n+m
k,l=mC(k, l)(cid:17)′
simply follows from the assumption that a vanishes on (⊕n
⊆ (cid:16)⊕n
k,l=0C(k, l)(cid:17)′
k=0C(k, l))′.
⊗(cid:16)⊕m
k,l=0C(k, l)(cid:17)′
, the statement
To prove that I is a coideal we have to show that ∆(I) ⊆ I ⊗ A + A ⊗ I. For this purpose we use an
equivalent definition of I, that is, we consider the space spanned by the slices
(ω ⊗ ι)(cid:0)(T ⊗ 1)U ⊗k − U ⊗l(T ⊗ 1)(cid:1)
2
(1.2)
for all ω ∈ B(H ⊗k, H ⊗l)∗, T ∈ C(k, l) and k, l ≥ 0. This space indeed coincides with I, since using (1.1)
k=0B(H ⊗k) vanishes on the elements (1.2) for all k, l ≤ n if and only it lies
we see that an operator S ∈ ⊕n
in Bn. We choose an orthonormal basis of H ⊗k, {ξj}j, and of H ⊗l, {ηi}i, and assume ω is of the form
ωij = h · ξj , ηii. We set V := U ⊗k and W := U ⊗l. Then, using the leg-numbering notation,
∆(ωij ⊗ ι) ((T ⊗ 1)V − W (T ⊗ 1))
is equal to
The expression in the parentheses can be written as
(ωij ⊗ ι ⊗ ι) ((T ⊗ 1 ⊗ 1)V12V13 − W12W13(T ⊗ 1 ⊗ 1)) .
((T ⊗ 1 ⊗ 1)V12 − W12(T ⊗ 1 ⊗ 1)) V13 + W12 ((T ⊗ 1 ⊗ 1)V13 − W13(T ⊗ 1 ⊗ 1)) .
Now, if we just consider the first part of the sum
(ωij ⊗ ι ⊗ ι) (((T ⊗ 1 ⊗ 1)V12 − W12(T ⊗ 1 ⊗ 1)) V13) ,
it can be expressed as
(ωik ⊗ ι ⊗ ι) ((T ⊗ 1 ⊗ 1)V12 − W12(T ⊗ 1 ⊗ 1)) (ωkj ⊗ ι ⊗ ι)(V13),
Xk
which belongs to I ⊗ A. Similarly the other part lies in A ⊗ I, as wanted.
By the previous Lemma, A/I is a bialgebra. What we wish to prove is that A/I ∼= C[G], for a compact
quantum group G, and to do so we need for A/I to be a Hopf ∗-algebra and for U to be unitary (see Theorem
1.6.7 of [2]), and this is not obvious written in this manner. In fact it is not even clear whether A/I has
a ∗-structure. We shall proceed with an intermediate step. The idea is the following: we will introduce
another bi-ideal in A, IF , and show that A/IF ∼= C[O+
F is the free orthogonal quantum group.
Thus, A/IF will automatically inherit a Hopf ∗-algebra structure. Finally we will show that I/IF is a Hopf
∗-ideal in A/IF and then conclude that there exists a compact quantum group G such that
F ], where O+
(cid:3)
again by Theorem 1.6.7 of [2].
A/I ∼= A/IF.I/IF ∼= C[G],
2. Representation category of a free orthogonal quantum group
Following the strategy described above, we now consider the case when C is the smallest collection of
spaces as in Theorem 1.1 containing a fixed operator R : C → H ⊗2 such that (R∗ ⊗ ι)(ι ⊗ R) = ±ι. It is
known, and not difficult to see, that if we fix an orthonormal basis e1, . . . , en in H, then R has the form
(ι ⊗ F )r, where r : C → H ⊗ H is given by r(1) =Pi ei ⊗ ei and F ∈ GLn(C) is such that F ¯F = ±1, where
¯F is the matrix obtained from F by taking the complex conjugate of every entry. We will use the subindex
F for the constructions of the previous section related to this smallest collection, so we write CF , BF,n, IF,n,
etc.
Consider the universal unital algebra C[O+
F ] generated by entries of a matrix U = (uij )i,j satisfying the
relations
U F tU t(F −1)t = 1, F tU t(F −1)tU = 1.
for an invertible n by n matrix F . It is again known and easy to see that this is a Hopf ∗-algebra with
comultiplication ∆(uij) =Pk uik ⊗ ukj and involution given by U ∗ = F tU t(F −1)t. The compact quantum
group O+
F thus defined is known in literature as the free orthogonal quantum group [4], but we do not need
to know any properties of this quantum group apart from the fact that it is well-defined. The following
Lemma is a simple consequence of our definitions.
Lemma 2.1. We have A/IF ∼= C[O+
F ].
3
Proof. By definition, the bialgebra C[O+
elements
F ] can be written as A/L where L is the ideal generated by the
with U ∈ B(H) ⊗ A being the fundamental matrix of A. In order to prove the Lemma we need to show
that IF = L. To show that L ⊆ IF , consider the linear functionals ω1,ij := h · 1, ei ⊗ eji ∈ B(C, H ⊗2)∗ and
ωij,1 := h · ei ⊗ ej, 1i ∈ B(H ⊗2, C)∗. Then
(cid:0)F t − (U F tU t)(cid:1)ij , (cid:0)(F −1)t − U t(F −1)tU(cid:1)ij
∀i, j
and
(cid:0)F t − (U F tU t)(cid:1)ij = (ω1,ij ⊗ ι)(cid:0)(R ⊗ ι) − U ⊗2(R ⊗ ι)(cid:1)
where we recall that R = (ι ⊗ F )r and in the second equality we use that F ∗ = ±(F −1)t. Hence L ⊆ IF .
(cid:0)(F −1)t − U t(F −1)tU(cid:1)ij = ±(ωij,1 ⊗ ι)(cid:0)(R∗ ⊗ ι) − (R∗ ⊗ ι)U ⊗2(cid:1) ,
Conversely, let us show that IF ⊆ L. As follows from the above identities, R and R∗ are morphisms in
F . But this implies IF ⊆ L
F . It follows that any operator in CF is a morphism in Rep O+
the category Rep O+
because any relation defined by elements of CF has to be satisfied in C[O+
F ].
Therefore in order to prove Theorem 1.1 for C = CF it remains to establish the following Lemma. We
remark that the Lemma itself is not needed for the general case, but its proof will be reused.
Lemma 2.2. We have HomO+
F
(H ⊗k, H ⊗l) = CF (k, l) for all k, l ≥ 0.
(cid:3)
Proof. We have to show that for every n
n
As both sides are (finite dimensional) von Neumann algebras, the equality is equivalent to the equality of
their commutants:
CF (k, l) = EndO+
Mk,l=0
F n
Mk=0
H ⊗k! .
BF,n = EndO+
F n
Mk=0
H ⊗k!′
.
Recall now that for any finite dimensional representation V ∈ B(HV ) ⊗ C[G] of a compact quantum
group G we have a representation πV of the algebra C[G]∗ on HV defined by πV (φ) = (ι ⊗ φ)(V ), and
then πV (C[G]∗) = EndG(HV )′. Therefore we have to show that πLn
immediately follows from the previous Lemma, as C[O+
F ]∗(cid:1) = BF,n. But this
k=0 U ⊗k(cid:0)C[O+
F ] = A/IF ⊃ An/IF,n = B∗
F,n.
(cid:3)
3. Proof of the Theorem
We now turn to a general C as in Theorem 1.1. Let R ∈ C(0, 2) be an operator such that (R∗ ⊗ ι)(ι ⊗ R) =
±ι. As in the previous section, we fix an orthonormal basis e1, . . . , en in H and write R as (ι ⊗ F )r. Denote
by J the bi-ideal I/IF in C[O+
F ] = A/IF . Note that J can still be described as the space spanned by the
elements (ω ⊗ ι)(cid:0)(T ⊗ 1)U ⊗k − U ⊗l(T ⊗ 1)(cid:1), for T ∈ C(k, l) and ω ∈ B(H ⊗k, H ⊗l)∗, where we slightly abuse
the notation and denote by the same symbol U the fundamental matrix of A and its image in C[O+
F ].
Lemma 3.1. J is a Hopf ∗-ideal in C[O+
F ].
Proof. We denote by S the antipode of C[O+
under taking the adjoints and is invariant under S. Let a∗ = (ω ⊗ ι)(cid:0)(T ⊗ 1)U ⊗k − U ⊗l(T ⊗ 1)(cid:1)∗
We have to show that it lies in J for any k, l ≥ 0. We first note that C(1, 1) is closed under the operation
.
F ]. Since J is a bi-ideal, we only need to check that J is closed
∨ defined by
(ι ⊗ T )R = (T ∨ ⊗ ι)R,
since T ∨ = ±(ι ⊗ R∗)(ι ⊗ T ⊗ ι)(R ⊗ ι).
As (ι ⊗ T )r = (T t ⊗ ι)r, we have T ∨ = (F −1T F )t, and the inverse operation, still preserving C(1, 1),
is T 7→ F T tF −1. We recall from the previous section that we also have U ∗ = F tU t(F −1)t. Analogous
formulas hold for T ∈ C(k, l). In fact, if we denote by Fk = F ⊗k and Uk = U ⊗k,
4
then T ∨ = (F −1
l T Fk)t ∈ C(l, k) and U ∗
k = F t
kU t
k(F −1
k )t. Therefore, choosing an orthonormal basis of
H ⊗k, {ξj}j, and of H ⊗l, {ηi}i, and assuming ω is of the form ωij = h · ξj , ηii, we have
a∗ =(ωij ⊗ ι) ((T ⊗ 1)Uk − Ul(T ⊗ 1))∗ =
(ωji ⊗ ι) (U ∗
k (T ∗ ⊗ 1) − (T ∗ ⊗ 1)U ∗
l ) =
k )t − (U ∗
k )t − (U ∗
k UkFk) − (F −1
l )t((T ∗)t ⊗ 1)(cid:1)t(cid:17) =
l )t((T ∗)t ⊗ 1)(cid:1) =
l UlFl)((T ∗)t ⊗ 1)(cid:1) =
k ⊗ 1)Uk − Ul(Fl(T ∗)tF −1
(ωji ⊗ ι)(cid:16)(cid:0)((T ∗)t ⊗ 1)(U ∗
(ωij ⊗ ι)(cid:0)((T ∗)t ⊗ 1)(U ∗
(ωij ⊗ ι)(cid:0)((T ∗)t ⊗ 1)(F −1
(ωij ⊗ ι)(cid:0)(F −1
Xm,n
(F −1
l
l ⊗ 1)(cid:0)(Fl(T ∗)tF −1
)im(Fk)nj (ωmn ⊗ ι)(cid:16)( T ⊗ ι)Uk − Ul( T ⊗ ι)(cid:17) ,
k ⊗ 1)(cid:1) (Fk ⊗ 1)(cid:1) =
where T = Fl(T ∗)tF −1
operation of ∨. Hence, a∗ ∈ J .
k ∈ C(k, l), since T 7→ FlT tF −1
k
is a map from C(l, k) to C(k, l), being the inverse
The invariance of J under the antipode immediately follows from its invariance under involution.
If
a = (ωij ⊗ ι) ((T ⊗ 1)Uk − Ul(T ⊗ 1)) then
S(a) =(ωij ⊗ ι) ((T ⊗ 1)(ι ⊗ S)(Uk) − (ι ⊗ S)(Ul)(T ⊗ 1))
=(ωij ⊗ ι) ((T ⊗ 1)U ∗
=(ωji ⊗ ι) (Uk(T ∗ ⊗ 1) − (T ∗ ⊗ 1)Ul)∗ ∈ J .
l (T ⊗ 1))
k − U ∗
Given the above Lemma, we conclude that there exists a compact quantum group G such that A/I ∼= C[G].
It remains to show that HomG(H ⊗k, H ⊗l) = C(k, l). But this is done in exactly the same way as in
Lemma 2.2.
(cid:3)
To finish the proof of Theorem 1.1 we have to show that the compact quantum group G is unique up to
isomorphism. Let G′ be another compact quantum group satisfying the assumptions of Theorem 1.1, that
is, having a fundamental representation V = (vij )ij on H such that HomG′(H ⊗k, H ⊗l) = C(k, l). We can
identify C[G′] with A/I ′, for a bi-ideal I ′ ⊂ A. The only thing to check is that the bi-ideal I ′ is completely
determined by the operator spaces C(k, l). Since C(k, l) and HomG′(H ⊗k, H ⊗l) coincide, from the proof of
Lemma 2.2 we see that this implies that An/I ′
n. Hence, the
spaces I ′
n are completely determined by the spaces C(k, l). Thus Theorem 1.1 is proved.
n = I ′ ∩ An and I ′ = Sn≥0 I ′
n, where I ′
n = B∗
4. General version of the Tannaka-Krein duality
In this section we want to explain, without too many details, how using Theorem 1.1 one can recover the
following result.
Theorem 4.1 (Woronowicz's Tannaka-Krein duality). Let C be an essentially small C∗-tensor category with
conjugates, τ : C → Hilbf be a unitary fiber functor. Then there exists a compact quantum group G and
a unitary monoidal equivalence θ : C → Rep G such that τ is naturally unitarily monoidally isomorphic to
the composition of the canonical fiber functor π : Rep G → Hilbf with θ. Furthermore, the Hopf ∗-algebra
(C[G], ∆) for such a G is uniquely determined up to isomorphism.
We remark that for C∗-tensor categories we follow the conventions of [2], in particular, we assume that
they are closed under finite direct sums and subobjects.
We concentrate only on the existence of G. We may assume that C is a subcategory of Hilbf and τ is the
embedding functor. If there exist an object H in C such that any other object is isomorphic to a subobject
of H ⊗n for some n ≥ 0, and a morphism R : C → H ⊗ H such that (R∗ ⊗ ι)(ι ⊗ R) = ±ι , then the result
follows from Theorem 1.1. For general C let us distinguish between three cases:
(i) C is generated, as a C∗-tensor category with conjugates, by one object;
5
(ii) C is generated by a finite number of objects;
(iii) C is infinitely generated.
(i) Assume C is generated by one object K, so every object of C is isomorphic to a subobject of a tensor
product of copies of K and an object ¯K conjugate to K. Let (R′, ¯R′) be a solution of the conjugate equations
for K and ¯K. Then letting H = K ⊕ ¯K and R = R′ ⊕ ¯R′, considered as a morphism C → H ⊗ H, we have
(R∗ ⊗ ι)(ι ⊗ R) = ι, so we are back to the case covered by Theorem 1.1.
(ii) The case when C is generated by a finite number of objects H1, . . . , Hn is not much different from (i),
as then C is generated by H1 ⊕ · · · ⊕ Hn.
(iii) For general C, choose a generating set F in C and let E be the family of finite subsets of F ordered
by inclusion. For each E ∈ E let CE be the full rigid C∗-tensor subcategory of C generated by the finite
set of objects in E. By the previous case, for each subcategory CE we get a compact quantum group GE
with representation category CE. Moreover, if E ⊂ E′, then, since CE ⊂ CE′, by the uniqueness part of
Theorem 1.1, the quantum group GE is a quotient of GE′ , that is, we have an embedding C[GE] ֒→ C[GE′ ]
of Hopf ∗-algebras. Then C[G] is defined as the inductive limit of the Hopf ∗-algebras C[GE].
In the following example we can see how to recover the free unitary quantum group following the procedure
explained in point (i) of the above.
unitary quantum group.
Example 4.2 (Free Unitary quantum group). We denote by C[U +
Q ] the universal unital ∗-algebra generated
by the entries of matrices V = (vij )i,j and ¯V = (¯vij )i,j such that V and ¯V are unitary with involution
defined by V ∗ = Qt ¯V t(Q−1)t and ¯V ∗ = (Q−1)∗V tQ∗, for an invertible n by n matrix Q. The algebra C[U +
Q ]
Q is known in literature as the free
is a Hopf ∗-algebra with comultiplication ∆(vij ) = Pk vik ⊗ vkj and U +
We wish to prove the equivalence between the representation category of the free unitary quantum group
and a concrete C∗-tensor category having certain properties. More specifically, consider the Hilbert space
K = Cn and its complex conjugate ¯K. Let CQ be the smallest collection of operators between tensor powers
of H := K ⊕ ¯K, as in Theorem 1.1, containing the operator R : C → H ⊗2 such that (R∗ ⊗ ι)(ι ⊗ R) = ι,
and the projection p : K ⊕ ¯K → K. The operator R is equal to (ι ⊗ F )r for F ∈ GL2(Mn(C)) with entries
F11 = F22 = 0, F12 = ¯Q−1 and F21 = Q, where ¯Q is the matrix whose coefficients are the complex conjugates
of the entries of Q.
We claim that HomU +
Q
(H ⊗k, H ⊗l) = CQ(k, l) for all k, l ≥ 0. We show that C[U +
Q ] ∼= A/I, where I is the
ideal generated by slice maps (ω ⊗ ι)(cid:0)(T ⊗ 1)U ⊗k − U ⊗l(T ⊗ 1)(cid:1), for T ∈ CQ(k, l) and ω ∈ B(H ⊗k, H ⊗l)∗.
The claim will then follow from Theorem 1.1. By definition, C[U +
ideal generated by the relations
Q ] can be written as A/L where L is the
U F tU t(F −1)t = 1, F tU t(F −1)tU = 1, U12 = 0, U21 = 0.
We already know that the ideal I contains slices of the first two relations, since we showed in Lemma 2.1
that they correspond to the slice maps
(ω1,ij ⊗ ι)(cid:0)(R ⊗ ι) − U ⊗2(R ⊗ ι)(cid:1)i,j ,
(ωij,1 ⊗ ι)(cid:0)(R∗ ⊗ ι)U ⊗2 − (R∗ ⊗ ι)(cid:1)i,j .
The other two relations correspond to (ωij ⊗ι) ((p ⊗ ι)U − U (p ⊗ ι))i,j. Hence L ⊆ I. The opposite inclusion
follows analogously to the second part of the proof of Lemma 2.1.
References
[1] T. Banica and R. Speicher. Liberation of Orthogonal Lie groups. Advances in Mathematics, 222: 1461-1501, 2009.
[2] S. Neshveyev and L. Tuset. Compact quantum groups and their representation categories, volume 20 of Cours Sp´ecialis´es.
Societe Math´ematique de France, Paris, 2013.
[3] P. Schauenburg. Tannaka duality for Arbitrary Hopf Algebras. Algebra Berichte 66, Fischer, Munich, 1991.
[4] A. Van Daele and S. Wang. Universal Quantum Groups. Int. J. Math., 7(2): 255-263, 1996.
[5] S. L. Woronowicz. Tannaka-Krein duality for compact matrix pseudogroups. Twisted SU(N) groups. Invent. Math., 93(1):
35-76, 1988.
6
E-mail address: [email protected]
Department of Mathematics, University of Oslo, P.O. Box 1053 Blindern, NO-0316 Oslo, Norway
7
|
1203.5063 | 3 | 1203 | 2012-09-06T20:26:21 | Closed quantum subgroups of locally compact quantum groups | [
"math.OA",
"math.FA"
] | We investigate the fundamental concept of a closed quantum subgroup of a locally compact quantum group. Two definitions - one due to S.Vaes and one due to S.L.Woronowicz - are analyzed and relations between them discussed. Among many reformulations we prove that the former definition can be phrased in terms of quasi-equivalence of representations of quantum groups while the latter can be related to an old definition of Podle\'s from the theory of compact quantum groups. The cases of classical groups, duals of classical groups, compact and discrete quantum groups are singled out and equivalence of the two definitions is proved in the relevant context. A deep relationship with the quantum group generalization of Herz restriction theorem from classical harmonic analysis is also established, in particular, in the course of our analysis we give a new proof of Herz restriction theorem. | math.OA | math |
CLOSED QUANTUM SUBGROUPS OF LOCALLY COMPACT QUANTUM
GROUPS
MATTHEW DAWS, PAWE L KASPRZAK, ADAM SKALSKI, AND PIOTR M. SO LTAN
Dedicated to Leonid Vainerman on the occasion of his 65th birthday
Abstract. We investigate the fundamental concept of a closed quantum subgroup of a locally
compact quantum group. Two definitions -- one due to S. Vaes and one due to S.L. Woronowicz
-- are analyzed and relations between them discussed. Among many reformulations we prove
that the former definition can be phrased in terms of quasi-equivalence of representations of
quantum groups while the latter can be related to an old definition of Podle´s from the theory of
compact quantum groups. The cases of classical groups, duals of classical groups, compact and
discrete quantum groups are singled out and equivalence of the two definitions is proved in the
relevant context. A deep relationship with the quantum group generalization of Herz restriction
theorem from classical harmonic analysis is also established, in particular, in the course of our
analysis we give a new proof of Herz restriction theorem.
1. Introduction
In this paper we study the notion of a closed quantum subgroup of a locally compact quan-
tum group. The theory of quantum groups phrased in operator algebra language is already well
established as a rapidly developing field on the border between noncommutative geometry and
abstract harmonic analysis. Nevertheless, the fundamental notion of a closed (quantum) subgroup
has not received enough attention so far. There have been several "working definitions" of such an
object, but most efforts were directed toward developing other aspects of the theory. The first to
look at quantum subgroups of (compact) quantum groups was P. Podle´s ([31, 32], see also a later
discussion in [30]). His view was motivated by the straightforward noncommutative generalization
of the inclusion homomorphism from the subgroup to the group and required the existence of a
surjective ∗-homomorphism between the algebras of continuous functions on respective quantum
groups. This point of view, however, has many disadvantages and drastically limits the number
of subgroups (e.g. many quantum groups do not have the trivial subgroup in this sense). Soon
it was realized that in the context of compact quantum groups one should rather require the ex-
istence of a surjective ∗-homomorphism between the universal versions of algebras of continuous
functions on respective quantum groups. This approach, adopted for example in [3] and [4], avoids
the problems mentioned above and also enables a purely algebraic reformulation in terms of the
underlying Hopf ∗-algebras. It was not clear, however, whether it would lead to a satisfactory
notion for arbitrary locally compact quantum groups.
In 2005 in [43] S. Vaes proposed another definition of a closed quantum subgroup of a locally
compact quantum group, phrased in the language of von Neumann algebras. This definition was
used in the same paper to develop the full force of the theory of induced representations and
homogeneous spaces for quantum groups. Earlier another definition of a closed subgroup of a
locally compact quantum group was proposed in [44, Definition 2.9] by Vaes and Vainerman. We
show that the definition of Vaes and that given by Vaes and Vainerman are equivalent. It should
be stressed that the argument needed to show that the definitions of a closed quantum subgroup
proposed in [43, 44] give the standard notion of a closed subgroup in classical case is quite subtle.
It can be formulated as saying that an inclusion of a closed subgroup H into a locally compact
Date: November 6, 2018.
2010 Mathematics Subject Classification. Primary: 20G42, 22D25, Secondary: 43A30, 46L89.
Key words and phrases. Quantum group, quantum subgroup, representation, quasi-equivalence, Herz restriction
theorem.
1
2
MATTHEW DAWS, PAWE L KASPRZAK, ADAM SKALSKI, AND PIOTR M. SO LTAN
group G induces a normal inclusion of respective group von Neumann algebras vN(H) ֒→ vN(G)
and is equivalent to the fact that the restriction to H of regular representation of G is quasi-
equivalent to the regular representation of H. This is, in turn, equivalent to the conclusion of the
Herz restriction theorem which says that the map of Fourier algebras associated to H ֒→ G is a
surjective contraction ([15], cf. also Section 4). All this has been known to the experts for a long
time (cf. [16, 42, 10, 11, 44]); a detailed proof can be found in the 2008 thesis of C. Zwarich ([52]).
The definition given in [43] is very well adapted to the problems studied in that paper, but it was
not clear whether it is optimal in other contexts and how it relates to the notion studied earlier for
compact quantum groups. As mentioned above it is also relatively difficult to see that it actually
generalizes the classical notion of a closed subgroup. Yet another possible definition, related to the
recently introduced notion of morphisms between quantum groups ([27]), was suggested to us by
S.L. Woronowicz. Woronowicz's definition is phrased entirely in the language of C∗-algebras and
it is notably easier to see that it generalizes the ordinary notion of a closed subgroup of a locally
compact group (see Section 4). The main focus of this paper is on understanding the relations
between the definitions of a closed quantum subgroup of a locally compact quantum group given
by Vaes and Woronowicz and providing their equivalent reformulations.
The definition of Woronowicz is deeply connected with the notion of a C∗-algebra generated by a
quantum family of multipliers (which we analyze in Subsection 1.1) and turns out to be equivalent
to the reformulation of the original idea of Podle´s, i.e. corresponds to the existence of a surjective ∗-
homomorphism between the universal versions of the algebras of continuous, vanishing at infinity,
functions on respective locally compact quantum groups. On the other hand the definition of Vaes
can be rephrased in a simplified way (still in the von Neumann algebraic language) and turns out
to be intimately connected with the notion of quasi-equivalence of representations of quantum
groups ([40], cf. Theorem 3.4). Moreover, we show that this definition of a closed subgroup of a
quantum group is strongly tied to the generalization to quantum groups of the Herz restriction
theorem (cf. Remark 3.8).
We show that the definition of Vaes is stronger than the definition of Woronowicz, in the sense
that if H and G are locally compact quantum groups and H is a closed quantum subgroup of G
in the sense of Vaes, then it is also a closed quantum subgroup of G in the sense of Woronowicz.
Further we prove that they are equivalent in all special cases one usually considers: classical groups
(both definitions describe the standard notion of a closed subgroup), duals of classical groups (both
definitions describe a group epimorphism in the opposite direction), compact quantum groups
and discrete quantum groups (Sections 4, 5 and 6). In particular this opens the way to finding
all compact quantum subgroups of a given locally compact quantum group G via the theory of
idempotent states (as studied for example in [36]) since each compact quantum subgroup of G gives
rise to a state on the algebra of functions on G, which is idempotent with respect to the convolution
product, and such states can be sometimes computed directly using Fourier transform methods.
In the context of compact quantum groups this strategy was employed in [14] to re-establish the
list of all quantum subgroups of SUq(2), originally found by Podle´s in [32].
In the course of our investigation we make crucial use of the quantum group versions of the
Fourier and Fourier-Stieltjes algebras (cf. [9, Section 8]). It is worth noting that our work produces
a new proof of the classical Herz restriction theorem (see Section 4). In the group-dual case we use
the results of M. Ilie and R. Stokke on weak∗-continuous maps of Fourier-Stieltjes algebras ([18])
which we are also able to generalize (to some extent) to the quantum group setting (Proposition
5.3). This exemplifies the connections of our article with recent extensions of noncommutative
harmonic analysis to the context of locally compact quantum groups (see for example [17] and
references therein).
Finally let us note that the differences between the definitions of a closed quantum sub-
group according to Vaes and Woronowicz bear a striking similarity to the interplay between the
Kustermans-Vaes definition of a locally compact quantum group (formulated in [23]) and the def-
inition of a quantum group used in [40, 27] and based on the theory of manageable and modular
multiplicative unitaries ([50, 39]). Again, the former definition is stronger and in all examples one
finds that the two approaches are equivalent. Moreover, in special cases of classical groups, duals
CLOSED QUANTUM SUBGROUPS OF LOCALLY COMPACT QUANTUM GROUPS
3
of classical groups, compact and discrete quantum groups we have results on existence of Haar
measures, so the Kustermans-Vaes approach is equivalent with the one used by So ltan-Woronowicz.
At the present stage of research in the theory of quantum groups it is very difficult to predict
whether the definitions of a closed quantum subgroup given by Vaes and Woronowicz are equiv-
alent. We conjecture that in the full generality they are different. However, it seems very likely
that in large classes of well-behaved locally compact quantum groups, e.g. the regular or even
semi-regular ones, the two definitions will turn out to be equivalent.
Let us give now a brief description of the paper. In the remainder of this section we collect
necessary preliminaries from the theory of C∗-algebras (Subsection 1.1), locally compact quantum
groups (Subsection 1.2) and homomorphisms of quantum groups as defined in [27] (Subsection
1.3). Section 2 focuses on the theory of representations of quantum groups and the notion of quasi-
equivalence of such representations. We also relate this notion to the problem of generation of C∗-
algebras by quantum families of multipliers, which later turns out to be crucial for understanding
the interplay between the definitions of closed quantum subgroups given by Vaes and Woronowicz.
These are introduced in Section 3 with the relations between them unraveled. We provide several
equivalent reformulations of either definition and show that the former implies the latter (in the
sense described above). We also give sufficient conditions for the two definitions to be equivalent.
Section 4 is devoted to the study of both definitions of a closed quantum subgroup in the special
case of classical groups. We prove there in detail that both are equivalent to the standard definition
of a closed subgroup and discuss the direct connection between the definition of Vaes and the Herz
restriction theorem. Then in Section 5 we conduct a similar investigation for the case of duals of
classical groups. In this case also the definitions of Vaes and Woronowicz agree. Finally in Section
6 we show that the two definitions are equivalent for compact and discrete quantum groups (more
precisely a compact quantum group H is a closed subgroup of a locally compact quantum group
G in the sense of Vaes if and only if it is a closed subgroup of G in the sense of Woronowicz, and
a similar result holds for subgroups of discrete quantum groups).
1.1. C∗-algebras and morphisms. Throughout the paper we will use the language of the theory
of C∗-algebras as introduced in [47, 48, 49, 24]. In particular for C∗-algebras A and B a morphism
from A to B is a ∗-homomorphism Φ from A into the multiplier algebra M(B) of B which is
non-degenerate, i.e. the set Φ(A)B of linear combinations of products of the form Φ(a)b (a ∈ A,
b ∈ B) is dense in B (by the Cohen factorization theorem this is equivalent to the condition
that Φ(A)B = B). The set of all morphisms from A to B will be denoted by Mor(A, B). The
non-degeneracy of morphisms ensures that each Φ ∈ Mor(A, B) extends uniquely to a unital
∗-homomorphism M(A) → M(B) which we will sometimes denote by ¯Φ. This also defines the
operation of composition of morphisms (see [47, 49, 24]). For a Hilbert space H the C∗-algebra
of compact operators on H will be denoted by K(H). Any C∗-algebra A acting on H (written
A ⊂ B(H)) will act non-degenerately, so that the identity map idA : A → A is a morphism from A
to K(H). More generally a representation of A on H is by definition an element of Mor(cid:0)A, K(H)(cid:1).
The notion of a morphism of C∗-algebras generalizes that of a continuous map between locally
compact Hausdorff spaces. We have the following well known result:
Theorem 1.1. Let X and Y be locally compact Hausdorff spaces and let B = C0(X) and A =
C0(Y ). Then
(1) any continuous φ : X → Y defines a morphism Φ ∈ Mor(A, B) via
Φ(f ) = f ◦φ,
(f ∈ A);
(1.1)
(2) for any Φ ∈ Mor(A, B) there exists a continuous φ : X → Y such that (1.1) holds.
Fixing Φ and φ so that (1.1) holds we moreover have
(3) the range of Φ is contained in B = C0(X) if and only if φ is a proper map,
(4) φ has dense image if and only if Φ is injective,
(5) φ is injective if and only if Φ has strictly dense range.
4
MATTHEW DAWS, PAWE L KASPRZAK, ADAM SKALSKI, AND PIOTR M. SO LTAN
The strict topology on a multiplier algebra mentioned in Theorem 1.1 is described e.g. in [49,
Section 2] or [24, Chapter 1]. The proof of the above theorem is a simple exercise in elementary
topology and we leave it to the reader (see e.g. [46, Exercises to Chapter 2]).
Let A be a C∗-algebra. The dual space A∗ is naturally a module over A and we will denote the
natural left action of a ∈ A on ϕ ∈ A∗ by a · ϕ, so that (a · ϕ)(b) = ϕ(ba) for all b ∈ A. Note that
if C ⊂ B(H) then C acts in a natural way on the functionals in B(H)∗ = K(H)∗ and we have
C · K(H)∗ = C K(H) · K(H)∗ = K(H) · K(H)∗ = K(H)∗.
(1.2)
(all sets above are automatically closed by the Cohen factorization theorem).
For C∗-algebras A and B their minimal tensor product will be denoted by A ⊗ B. For von
Neumann algebras M and N the von Neumann algebra tensor product of M and N will be denoted
by M ¯⊗N. The tensor flip a ⊗ b 7→ b ⊗ a will be denoted by σ regardless of which C∗-algebras
are being considered. We will also use the same symbol "⊗" to denote tensor product of Hilbert
spaces.
In [49] S.L. Woronowicz introduced a very important notion of a C∗-algebra generated by
elements which do not necessarily belong to it. We will use a crucial part of his theory dealing
with C∗-algebras "generated by a quantum family of multipliers". Let A and C be C∗-algebras and
let T ∈ M(C ⊗ A). By analogy with the classical situation (when C is commutative) the element
T is referred to as a quantum family of elements of M(A) labeled by the spectrum of C (cf. [49,
Formula (2.5)]).
Definition 1.2 ([49, Definition 4.1]). Let A and C be C∗-algebras. We say that A is generated by
T ∈ M(C ⊗ A) if for any Hilbert space H, any representation ρ of A on H and any C∗-algebra
B ⊂ B(H) the condition that (id ⊗ ρ)(T ) ∈ M(C ⊗ B) implies that ρ ∈ Mor(A, B).
Examples of the situation described in Definition 1.2 are plentiful. For the simplest case consider
a C∗-algebra A generated by a finite set of elements a1, . . . , an ∈ A (in the usual sense, i.e. the
closure of the set of algebraic combinations of the elements a1, . . . , an and their adjoints coincides
with A). Then A is generated by T ∈ M(Cn ⊗ A) with
T =
nXi=1
ei ⊗ ai,
where {e1, . . . , en} is the standard basis of Cn. More complicated examples of C∗-algebras gener-
ated by quantum families of multipliers are given in [49, Section 4]. In this paper we will be mostly
interested in examples of this situation arising from representations of locally compact quantum
groups to be studied in Subsection 1.2, Section 2 and Section 3.
Remark 1.3. Let H and K be Hilbert spaces and consider C∗-algebras A1, A2 ⊂ B(H) and C ⊂ B(K).
Suppose that T ∈ B(K ⊗ H) is such that T ∈ M(C ⊗ A1) ∩ M(C ⊗ A2) and T generates both A1
and A2. Then A1 = A2, as the identity representation of A1 is a morphism in Mor(A1, A2), and
similarly the identity representation of A2 is a morphism in Mor(A2, A1). This argument appeared
already in [49].
Usually it is difficult to check that a given T ∈ M(C⊗A) generates A. For the needs of this paper
it will be very useful to apply the following criterion. Note that if T ∈ M(C ⊗ A) and C ⊂ B(H),
then each functional ω ∈ B(H)∗ defines an element of C∗, so that, in particular, (ω ⊗id)(T ) ∈ M(A)
([24, Proposition 8.3], [25, Lemma A.3]).
Lemma 1.4. Let A and C be C∗-algebras with C ⊂ B(H) for a Hilbert space H. Let T ∈ M(C ⊗ A)
be unitary and define
If S ⊂ A and S generates A (as a subset of the C∗-algebra A) then T ∈ M(C ⊗ A) generates A.
S =(cid:8)(ω ⊗ id)(T ) ω ∈ B(H)∗(cid:9) ⊂ M(A).
Proof. Let K be a Hilbert space and let ρ be a representation of A on K such that (id ⊗ ρ)(T ) ∈
M(C ⊗ B) for a certain C∗-algebra B ⊂ B(K). It is easily seen that ρ(S) ⊂ M(B), which implies
CLOSED QUANTUM SUBGROUPS OF LOCALLY COMPACT QUANTUM GROUPS
5
that ρ(A) ⊂ M(B) because S generates A. Furthermore
(cid:0)ρ(S)B(cid:1) -- k·k =(cid:8)ρ(cid:0)(c · ω ⊗ id)(T )(cid:1)b c ∈ C, b ∈ B, ω ∈ B(H)∗(cid:9) -- k·k
=(cid:8)(ω ⊗ id)(cid:0)(id ⊗ ρ)(T ) (c ⊗ b)(cid:1) c ∈ C, b ∈ B, ω ∈ B(H)∗(cid:9) -- k·k
=(cid:8)(ω ⊗ id)(c ⊗ b) c ∈ C, b ∈ B, : ω ∈ B(H)∗(cid:9) -- k·k = B,
where in the first equality we used the formula (1.2) and in the last one we used the fact that
(id ⊗ ρ)(T ) is unitary in M(C ⊗ B). This shows that ρ ∈ Mor(A, B) and ends the proof.
(cid:3)
Remark 1.5. Sometimes it is important to use the notion of a C∗-algebra generated by a quantum
family of multipliers in a different version. More precisely let A and C be C∗-algebras and let
T ∈ M(A ⊗ C) (note the different order of tensor factors from the one in Definition 1.2). We
will say that T ∈ M(A ⊗ C) generates A if σ(T ) ∈ M(C ⊗ A) generates A in the sense described
in Definition 1.2. It can happen that a given T ∈ M(C ⊗ A) generates A and at the same time
T ∈ M(C ⊗ A) generates C. Coming back to the analogy with classical situation we would say that
in the first statement T is a quantum family of multipliers of A labeled by the spectrum of C and
in the second statement T is a quantum family of multipliers of C labeled by the spectrum of A.
Throughout the paper we will use the so-called leg-numbering notation. This is explained in a
number of texts on quantum groups, e.g. [33, 2].
1.2. Locally compact quantum groups and their universal versions. For the theory of
locally compact quantum groups we refer the reader to [23] and to [25] for an equivalent approach
with different initial axioms. Most results of this paper are true in a potentially more general
setting of quantum groups defined by modular multiplicative unitaries ([39, 40, 27]), but we will
stay within the theory of Kustermans and Vaes. For a locally compact quantum group G the
corresponding C∗-algebra of "continuous functions on G vanishing at infinity" will be denoted by
C0(G). This C∗-algebra is equipped with a comultiplication ∆G ∈ Mor(cid:0)C0(G), C0(G) ⊗ C0(G)(cid:1).
There is also the reduced bicharacter WG ∈ M(cid:0)C0(bG) ⊗ C0(G)(cid:1) (see [17, Page 53]), where bG
denotes the dual of G. The Haar weights provide a realization of both C0(G) and C0(bG) on the
Hilbert space L2(G). Then WG ∈ B(cid:0)L2(G) ⊗ L2(G)(cid:1) is a multiplicative unitary ([2]) called the
Kac-Takesaki operator of G ([25]). The comultiplication is then implemented by WG:
∆G(f ) = WG(f ⊗ 1)(WG)∗
for all f ∈ C0(G) (note that we are using the conventions of [2, 50, 25, 40, 27] favoring right Haar
weights over left ones). The embedding of C0(G) into B(cid:0)L2(G)(cid:1) defines also the von Neumann
algebra L∞(G) as C0(G)′′. Moreover we have
C0(G) =(cid:8)(ω ⊗ id)(WG) ω ∈ B(cid:0)L2(G)(cid:1)∗(cid:9) -- k·k.
In fact C0(G) is generated by the quantum family WG ∈ M(cid:0)C0(bG)⊗ C0(G)(cid:1) in the sense described
in Definition 1.2 ([50]). Moreover the C∗-algebra C0(bG) is generated by quantum family WG ∈
M(cid:0)C0(bG) ⊗ C0(G)(cid:1) (note the difference, cf. Remark 1.5).
The dense subspace
(1.3)
(1.4)
AG =(cid:8)(ω ⊗ id)(WG) ω ∈ B(cid:0)L2(G)(cid:1)∗(cid:9) ⊂ C0(G)
(no closure) is called the Fourier algebra of G ([9, Section 8]). Note that the vector space AG is
by the functionals which vanish on C0(G) with L∞(G)∗. It is clear that one can use this space of
indeed a subalgebra of C0(G) ([2, Proposition 1.4]). We will identify the quotient of B(cid:0)L2(G)(cid:1)∗
functionals instead of B(cid:0)L2(G)(cid:1)∗ in all formulas of the form (1.4) or (1.3).
M(cid:0)C0(bG)(cid:1) is non-zero.
Lemma 1.6. Let G be a quantum group and let η ∈ C0(G)∗ be non-zero. Then (id ⊗ η)(WG) ∈
6
MATTHEW DAWS, PAWE L KASPRZAK, ADAM SKALSKI, AND PIOTR M. SO LTAN
Proof. If η 6= 0 then it must be non-zero on the norm dense set AG. Therefore there is a normal
functional ω on B(cid:0)L2(G)(cid:1) such that η(cid:0)(ω ⊗ id)(WG)(cid:1) 6= 0. Consequently
ω(cid:0)(id ⊗ η)(WG)(cid:1) 6= 0
which clearly implies that (id ⊗ η)(WG) 6= 0.
(cid:3)
of view of classical harmonic analysis ([12]). We will view the Fourier algebra both as a Banach
space and a subspace of C0(G).
The first consequence of Lemma 1.6 is that AG is isomorphic as a vector space to L∞(bG)∗; in
particular it is a Banach space with the norm transported from L∞(bG)∗. Indeed this is the point
0(G)(cid:1). This object was introduced and analyzed
a comultiplication ∆u
in [22]. In the more general setting of quantum groups defined by modular multiplicative unitaries
the universal C∗-algebra corresponding to G is studied in [40, Section 5]. The reduced bicharacter
lifts to the universal level, i.e. we have the universal bicharacter
The universal object related to G is a C∗-algebra which we will denote by Cu
G ∈ Mor(cid:0)Cu
0(G), endowed with
0(G) ⊗ Cu
0(G), Cu
([22, Proposition 3.8] and [27, Proposition 4.8]). Following the conventions of [40] the reducing mor-
0(bG) ⊗ Cu
phisms for G and bG will be denoted by ΛG ∈ Mor(cid:0)Cu
respectively (see [40, Definition 35]). We have
VVG ∈ M(cid:0)Cu
0(G)(cid:1)
0(G), C0(G)(cid:1) and Λ bG ∈ Mor(cid:0)Cu
0(bG), C0(bG)(cid:1)
(Λ bG ⊗ ΛG)(VVG) = WG.
The elements (id ⊗ ΛG)(VV
G) and (Λ bG ⊗ id)(VV
G) will be denoted by
WG ∈ M(cid:0)Cu
and W G ∈ M(cid:0)C0(bG) ⊗ Cu
0(bG) ⊗ C0(G)(cid:1)
0(G) =(cid:8)(ω ⊗ id)(W G) ω ∈ B(cid:0)L2(G)(cid:1)∗(cid:9) -- k·k
0(G)(cid:1) (by Lemma 1.4, cf. [40] and Proposition 2.1).
Cu
0(G)(cid:1)
respectively. We have
W G ∈ M(cid:0)C0(bG) ⊗ Cu
The universal dual is determined by the quantum group G only up to isomorphism, so when ΛG
0(G) = C0(G) and ΛG = id.
is an isomorphism (i.e. G is coamenable) then we can declare that Cu
Then
([40, Formula (5.14)]) and consequently the C∗-algebra Cu
0(G) is generated by the quantum family
(1.5)
(1.6)
(1.7)
VVG = W G and W
bG = W
bG.
Similarly, when bG is coamenable then
VV
bG = W
bG and W G = WG.
(1.8)
Note that quantum groups which are classical (i.e. quantum groups G for which C0(G) is com-
mutative) are always coamenable.
Proposition 1.7. Let G be a locally compact quantum group. Then the reducing map ΛG is
injective on the subspace
The proof is obvious:
(cid:16)(id ⊗ η)(W
(cid:8)(id ⊗ η)(W
bG) η ∈ B(cid:0)L2(bG)(cid:1)∗(cid:9).
bG) 6= 0(cid:17) =⇒ (cid:0)η 6= 0(cid:1) =⇒ (cid:16)(id ⊗ η)(W
(1.9)
bG) 6= 0(cid:17)
very useful [51, Proposition 3.2]. The image of ΛG on the subspace (1.9) is exactly the Fourier
by Lemma 1.6 applied to bG. Note that Proposition 1.7 can be viewed as a generalization of the
algebra AG. It follows that ΛG(cid:0)Cu
0(G)(cid:1) = C0(G).
The Fourier-Stieltjes algebra of G is the space
BG =(cid:8)(η ⊗ id)(VVG) η ∈ Cu
0(bG)∗(cid:9) ⊂ M(cid:0)Cu
0(G)(cid:1)
(see [9, Section 8], note that in that paper BG was embedded into M(cid:0)C0(G)(cid:1) and not into
0(G)(cid:1)). A reasoning analogous to that in the proof of Lemma 1.6 shows that BG is isomorphic
M(cid:0)Cu
CLOSED QUANTUM SUBGROUPS OF LOCALLY COMPACT QUANTUM GROUPS
7
as a vector space to Cu
In what follows we shall utilize both pictures of AG and BG -- as Banach spaces of functionals
0(bG)∗. Indeed, as (cid:8)(cid:0)id ⊗ [ω ◦ ΛG](cid:1)(WG) ω ∈ L∞(G)∗(cid:9) is dense in Cu
0(bG)
([40, Section 5]), a non-zero η must be non-zero on some element of the form(cid:0)id ⊗ [ω◦ΛG](cid:1)(WG),
so (ω ◦ΛG)(cid:0)(η ⊗ id)(VVG)(cid:1) 6= 0. In particular (η ⊗ id)(VVG) 6= 0.
and at the same time as (non-closed) subspaces of C0(G) and M(cid:0)Cu
instead of C0(G). Dually, G is discrete if bG is compact. In this case C0(G) is a c0-direct sum
of matrix algebras and we write c0(G) instead of C0(G). We also write in this case ℓ∞(G) for
L∞(G). Discrete quantum groups are always coamenable ([33]). We refer to [51] for the complete
account of the theory of compact quantum groups and to [33, Section 3] for a thorough treatment
of discrete quantum groups.
A quantum group G is compact if the C∗-algebra C0(G) is unital. In this case we write C(G)
0(G)(cid:1) respectively.
Finally let us mention that on the level of bicharacters the duality between G and bG is imple-
mented by the tensor flip and the adjoint operation:
VV
bG = σ(VVG)∗.
(1.10)
It follows that
W
bG = σ(WG)∗
and W
bG = σ(WG)∗.
1.3. Homomorphisms of locally compact quantum groups. Let G and H be locally compact
quantum groups. In [27] it is shown that the following three classes of objects are in a one-to-one
correspondence:
(1) strong quantum homomorphisms: morphisms
(2) bicharacters (from H to G): unitaries
π ∈ Mor(cid:0)Cu
0(G), Cu
0(H)(cid:1)
(π ⊗ π)◦∆u
G = ∆u
H ◦π;
V ∈ M(cid:0)C0(bG) ⊗ C0(H)(cid:1)
such that
such that
such that
(∆bG ⊗ idC0(H))(V ) = V23V13,
(idC0( bG) ⊗ ∆H)(V ) = V12V13.
(1.11)
(3) right quantum homomorphisms: morphisms
ρ ∈ Mor(cid:0)C0(G), C0(G) ⊗ C0(H)(cid:1)
(∆G ⊗ id)◦ρ = (id ⊗ ρ)◦∆G,
(id ⊗ ∆H)◦ρ = (ρ ⊗ id)◦ρ.
All these should be thought of as alternative descriptions of a fixed homomorphism from H to
G. Note that the reduced bicharacter WG of G introduced in Subsection 1.2 is a bicharacter from
G to G in the above sense and describes the identity homomorphism. Sometimes, to simplify the
language, we will refer to a strong quantum homomorphism π as above as a homomorphism from
H to G. A strong quantum homomorphism π is related to the bicharacter V via the formula
V =(cid:0)Λ bG ⊗ [ΛH◦π](cid:1)(VV
G),
(1.12)
while the right quantum homomorphism ρ is given by
for any x ∈ C0(G) ⊂ B(cid:0)L2(G)(cid:1).
ρ(x) = V (x ⊗ 1C0(H))V ∗
8
MATTHEW DAWS, PAWE L KASPRZAK, ADAM SKALSKI, AND PIOTR M. SO LTAN
One can also check (see [27, Lemma 3.4]) that for a unitary V ∈ M(cid:0)C0(bG) ⊗ C0(H)(cid:1) the
conditions (1.11) are equivalent to the following "twisted" pentagonal equations:
12 = WG
V23WG
WH
12V13V23,
23V12 = V12V13WH
23,
in M(cid:0)C0(bG) ⊗ K(cid:0)L2(G)(cid:1) ⊗ C0(H)(cid:1),
in M(cid:0)C0(bG) ⊗ K(cid:0)L2(H)(cid:1) ⊗ C0(H)(cid:1).
(1.13a)
(1.13b)
cπ2 ◦cπ1 = \π1 ◦π2,
The next result, namely [27, Proposition 3.14], describes in the simplest way the construction
of the dual homomorphisms (cf. (1.11)).
Proposition 1.8. If V is a bicharacter from H to G, the unitary bV = σ(V ∗) ∈ M(cid:0)C0(H)⊗C0(bG)(cid:1)
is a bicharacter from bG to bH.
Proposition 1.8 makes possible the following definition:
Definition 1.9. Let π be a morphism from H to G with corresponding bicharacter V . Then the
strong quantum homomorphism defined by bV is called the dual of π and will be denoted by bπ, so
that bπ ∈ Mor(cid:0)Cu
Let us note the most fundamental equality relating π to bπ (and determining bπ uniquely) con-
tained in [27, Theorem 4.15]:
0(bH), Cu
0(bG)(cid:1).
By applying (Λ bG ⊗ ΛH) to both sides and using (1.12) we obtain
(id ⊗ π)(VVG) = (bπ ⊗ id)(VVH).
Moreover if π1 and π2 are strong quantum homomorphisms associated with homomorphisms from
G1 to G2 and from G2 to G3 respectively then
(cid:0)Λ bG ⊗ [ΛH ◦π](cid:1)(VVG) = V =(cid:0)[Λ bG ◦bπ] ⊗ ΛH(cid:1)(VVH).
(1.14)
(1.15)
(1.16)
since (1.14) characterizes the dual strong quantum homomorphism. Thus if π is an isomorphism
of C∗-algebras then so is bπ.
from H to G such that the corresponding π ∈ Mor(cid:0)Cu
0(G) onto Cu
a one-to-one map from Cu
C0(H) such that πr ◦ΛG = ΛH ◦π.
Theorem 1.10. Let G and H be locally compact quantum groups. Consider a homomorphism
0(H). Then there exists an isomorphism πr of C0(G) onto
0(G), Cu
0(H)(cid:1) is an isomorphism, i.e. π is
Theorem 1.10 says that isomorphisms in the category of locally compact quantum groups con-
sidered in [27] drop down to C∗-algebraic isomorphisms of the reduced level. In what follows we
will refer to this situation by simply saying that G and H are isomorphic. A proof of this result
may be given along the lines of [22, Proposition 8.7] (cf. also [22, Proposition 7.1]). In Section 3
we will give a short proof of Theorem 1.10 using representation theory of locally compact quantum
groups and techniques developed in this paper. Let us note that these techniques make no use of
the existence of Haar weights and are equally applicable to quantum groups arising from modular
multiplicative unitaries.
2. Representations of locally compact quantum groups
In this section we recall some basic notions of the representation theory of locally compact quan-
tum groups ([2], [40, Section 3]) and establish alternative characterizations of quasi-equivalence of
two representations of a given quantum group (Theorem 2.2).
Let G be a locally compact quantum group and let H be a Hilbert space. A strongly continuous
unitary representation of G on H is a unitary element U ∈ M(cid:0)K(H) ⊗ C0(G)(cid:1) such that
usually decorated by the subscript U , so that U ∈ M(cid:0)K(HU ) ⊗ C0(G)(cid:1).
For such a representation U of G the subspace
We will usually write simply of "representations of G". Moreover the Hilbert space H will be
(id ⊗ ∆G)(U ) = U12U13.
AU =(cid:8)(id ⊗ ω)(U ) ω ∈ L∞(G)∗(cid:9) -- k·k
CLOSED QUANTUM SUBGROUPS OF LOCALLY COMPACT QUANTUM GROUPS
9
is a non-degenerate C∗-subalgebra of B(H) (it was denoted "BU " in [40]). In fact U is a multiplier
of AU ⊗ C0(G) and the quantum family U ∈ M(cid:0)AU ⊗ C0(G)(cid:1) generates AU ([50, 40]).
We will also use at some point the notation
AU =(cid:8)(id ⊗ ω)(U ) ω ∈ L∞(G)∗(cid:9).
It is easy to check that AU is an algebra -- this is a quantum group analogue of the "Fourier
space of a representation" defined in [1, Definition (2.1)]; note for example that AWG is the Fourier
algebra AbG of bG. Observe further that a bicharacter from H to G is a representation of H on L2(G)
and it follows from Proposition 1.8 that bV is a representation of bG on L2(H).
The generating property for representations can be reformulated in terms of their slices. In the
following proposition note the use of the notion of a C∗-algebra generated by a quantum family
of multipliers in the version described in Remark 1.5.
Proposition 2.1. Let U be a representation of G on a Hilbert space H and let A be a non-
degenerate C∗-subalgebra of B(H). Assume that U ∈ M(cid:0)A ⊗ C0(G)(cid:1). Then the following are
equivalent:
(1) U ∈ M(cid:0)A ⊗ C0(G)(cid:1) generates A;
(2) A = AU .
Proof. A direct consequence of the fact that U generates AU and Remark 1.3 (cf. [40, Subsection
3.5]).
(cid:3)
The standard notions of representation theory were all collected in [40, Section 3].
• Two representations U and V of G are equivalent if there exists a unitary operator T ∈
B(HU , HV ) such that
(T ⊗ 1)U = V (T ⊗ 1).
• If H is a Hilbert space then the trivial representation of G on H is
• The tensor product of two representations U and V is the representation
IH = 1B(H) ⊗ 1C0(G) ∈ M(cid:0)K(H) ⊗ C0(G)(cid:1).
U
⊤ V ∈ M(cid:0)K(HU ⊗ HV ) ⊗ C0(G)(cid:1)
defined by
• Representation U and V are quasi-equivalent if there exists a Hilbert space H such that
U
⊤ V = U13V23.
IH
⊤ U and IH
⊤ V are equivalent ([40, Proposition 13]).
The following theorem will be crucial in the next section, when we analyze a definition of a
closed quantum subgroup proposed in [43]. The implication (1)⇒(2) is [40, Corollary 15].
Theorem 2.2. Let U and V be representations of G on HU and HV respectively. The following
three conditions are equivalent:
(1) U is quasi-equivalent to V ;
(2) there exists a (necessarily unique) normal ∗-isomorphism γ : A′′
U → A′′
V such that
(γ ⊗ id)(U ) = V ;
(3) we have
(cid:8)(η ⊗ idB(L2(G)))(U ) η ∈ B(HU )∗(cid:9) =(cid:8)(µ ⊗ idB(L2(G)))(V ) µ ∈ B(HV )∗(cid:9).
Proof. (1)⇒(2). Suppose that U and V are quasi-equivalent. Let K be a Hilbert space and let
T : K ⊗ HU → K ⊗ HV be a unitary such that
Take ω ∈ B(cid:0)L2(G)(cid:1)∗ and put x = (id ⊗ ω)(U ) and y = (id ⊗ ω)(V ). Equation (2.2) shows that
T (1K ⊗ x)T ∗ = 1K ⊗ y. This implies that T (1K ⊗ A′′
in the converse direction we observe that actually T (1K ⊗ A′′
V . Applying a similar argument
V so there exists a
U )T ∗ = 1K ¯⊗A′′
U )T ∗ ⊂ 1K ¯⊗A′′
T12U23T ∗
12 = V23.
(2.1)
(2.2)
10
MATTHEW DAWS, PAWE L KASPRZAK, ADAM SKALSKI, AND PIOTR M. SO LTAN
normal ∗-isomorphism γ : A′′
equation (2.2) we see that (γ ⊗ id)(U ) = V .
U → A′′
V such that 1K ⊗ γ(x) = T (1K ⊗ x)T ∗ for all x ∈ A′′
U . Using
U → A′′
(2)⇒(1). Let γ : A′′
V be an normal ∗-isomorphism such that (γ ⊗ id)(U ) = V . It is a
well known fact (see e.g. [7, Theorem III.2.2.8]) that γ is of the form 1K ⊗ γ(x) = T (1K ⊗ x)T ∗
for some Hilbert space K and a unitary operator T : K ⊗ HU → K ⊗ HV . It is then easy to check
that T12U23T ∗
12 = V23, which proves the quasi-equivalence of U and V .
(2)⇒(3). Since for any µ ∈ B(HV )∗ the composition µ◦γ is a normal functional on A′′
V , there
exists η ∈ B(HV )∗ such that µ◦γ = η(cid:12)(cid:12)A′′
V
(cid:8)(η ⊗ idB(L2(G)))(U ) ω ∈ B(HU )∗(cid:9) ⊃(cid:8)(µ ⊗ idB(L2(G)))(V ) µ ∈ B(HV )∗(cid:9).
Exchanging the roles of U and V we get the opposite inclusion; hence (3) follows.
. This shows that
(3)⇒(2). Let κ be the extension of the antipode κ of G to an unbounded operator acting
on M(cid:0)C0(G)(cid:1) (see [50, Theorem 1.6]). Recall that κ is a densely defined operator acting on its
domain D(κ) ⊂ M(cid:0)C0(G)(cid:1) such that for any representation U ∈ M(cid:0)K(HU ) ⊗ C0(G)(cid:1) of G and any
η ∈ B(HU )∗ we have (η ⊗ id)(U ) ∈ D(κ) and
(2.3)
Consider the set X ⊂ L∞(G)∗ defined so that ω ∈ X if and only if ω∗ ◦ κ extends to a bounded
normal functional on L∞(G). Define further AX
the fact that U is a representation ensures that AX
U = (cid:8)(id ⊗ ω)(U ) ω ∈ X(cid:9). Equation (2.3) and
U is a weakly dense ∗-subalgebra of A′′
U .
κ(cid:0)(η ⊗ id)(U )(cid:1) = (η ⊗ id)(U ∗).
Let us define a map γ0 : AX
U → A′′
V by the following formula:
γ0(cid:0)(id ⊗ ω)(U )(cid:1) = (id ⊗ ω)(V ),
ω ∈ X.
Fix ω ∈ X. Since V ∈ M(cid:0)AV ⊗ C0(G)(cid:1), the expression (id ⊗ ω)(V ) makes sense. Moreover if
(id ⊗ ω)(U ) = 0 then for any η ∈ B(HU )∗ we have ω(cid:0)(η ⊗ id)(U )(cid:1) = 0 and by our assumption
ω(cid:0)(µ ⊗ id)(V )(cid:1) = 0 for any µ ∈ B(HV )∗. The last property means that (id ⊗ ω)(V ) = 0 and
shows that γ0 is well-defined (cf. Lemma 1.6). It can be checked that γ0 is a ∗-homomorphism,
for example
γ0(cid:0)(id ⊗ ω)(U )(cid:1)∗
=(cid:0)(id ⊗ ω)(V )(cid:1)∗
=(cid:0)id ⊗ [ω∗◦ κ](cid:1)(V ) = γ0(cid:0)(cid:0)(id ⊗ ω)(U )(cid:1)∗(cid:1).
to a certain element y ∈ A′′
In the next step we shall show that γ0 may be extended to a normal ∗-isomorphism γ : A′′
U → A′′
V .
Take x ∈ AU . Using Kaplansky's density theorem, we may find a bounded net (xi)i∈I of elements
in AX
xi = x. Let M ∈ R+ be the
U , say xi = (id ⊗ ωi)(U ) with ωi ∈ X such that w-lim
i∈I
corresponding bound: kxik ≤ M . In what follows we shall prove that(cid:0)γ0(xi)(cid:1)i∈I weakly converges
V )∗. For each i ∈ I we have µ(cid:0)γ0(xi)(cid:1) =
ωi(cid:0)(µ ⊗ id)(V )(cid:1). For η ∈ (A′′
U )∗ such that (µ ⊗ id)(V ) = (η ⊗ id)(U ) we obtain µ(cid:0)γ0(xi)(cid:1) = η(xi).
In particular(cid:12)(cid:12)µ(cid:0)γ0(xi)(cid:1)(cid:12)(cid:12) ≤ M kηk and lim
we conclude that the family (cid:0)γ0(xi)(cid:1)i∈I of functionals on (A′′
Banach-Steinhaus theorem it is norm bounded. Let N ∈ R+ be a bound: (cid:13)(cid:13)γ0(xi)(cid:13)(cid:13) ≤ N for all
µ(cid:0)γ0(xi)(cid:1) ∈ C is a bounded functional with the
i ∈ I. Noting that the map (A′′
norm not greater than N we conclude the existence of y ∈ A′′
V )∗
V )∗ is pointwise bounded. By the
µ(cid:0)γ0(xi)(cid:1) = η(x). Interpreting A′′
V . Take now µ ∈ (A′′
V as the dual of (A′′
V )∗ ∋ µ 7→ lim
i∈I
i∈I
V , such that y = w-lim
i∈I
0 then for each
γ0(xi). This
The equality γ(x∗) = γ(x)∗ for any x ∈ AX
continuous imply that γ(x∗) = γ(x)∗ for any x ∈ A′′
that for any x, x′ ∈ A′′
U and the fact that the star operation is weakly
U . We will now show using once again (2.1)
U we have γ(xx′) = γ(x)γ(x′). Note that, in the notation of the previous
U )∗. Passing to the limit
paragraph, for any i ∈ I we have µ(cid:0)γ0(xi)(cid:1) = η(xi) for a certain η ∈ (A′′
we get µ(cid:0)γ(x)(cid:1) = η(x), for any x ∈ A′′
i)(cid:1) =(cid:0)µ · γ(x)(cid:1)(cid:0)γ(x′
i)(cid:1)
i) = µ(cid:0)γ(xx′
U . Note also that
i) = η(xx′
(η · x)(x′
enables us to define the aforementioned extension by putting γ(x) = y. If xi −−→
i∈I
η(xi) = 0, so that y = 0. This implies that γ is well defined.
µ as above lim
i∈I
µ(cid:0)γ0(xi)(cid:1) = lim
i∈I
CLOSED QUANTUM SUBGROUPS OF LOCALLY COMPACT QUANTUM GROUPS
11
for any x, x′
i ∈ AX
U . Again, passing to the limit, we get
(η · x)(x′) =(cid:0)µ · γ(x)(cid:1)(cid:0)γ(x′)(cid:1)
for any x ∈ AX
elements of AX
Finally we compute:
U and x′ ∈ A′′
U . Replacing x ∈ AX
U and passing to the limit yields (η · x)(x′) = (cid:0)µ · γ(x)(cid:1)(cid:0)γ(x′)(cid:1) for any x, x′ ∈ A′′
µ(cid:0)γ(xx′)(cid:1) = η(xx′) = (η · x)(x′) =(cid:0)µ · γ(x)(cid:1)(cid:0)γ(x′)(cid:1) = µ(cid:0)γ(x)γ(x′)(cid:1)
U with a bounded, weakly convergent net (xi) of
U .
which shows that γ(xx′) = γ(x)γ(x′) for any x, x′ ∈ A′′
U .
Exchanging the roles of U and V leads to the inverse ∗-homomorphism γ−1 : A′′
U . This
shows that γ is normal, since isomorphisms of von Neumann algebras are automatically normal
([41, Corollary 3.10, page 135]).
(cid:3)
V → A′′
It was shown in [40] that a unitary representation U of G is quasi-equivalent to WG if it is
right absorbing, i.e. for any other representation V of G the tensor product V
⊤ U is equivalent to
⊤ U (this can be viewed as a version of the Fell absorption principle). We finish the section
IHV
with a proposition which describes relation between quasi-equivalence of a given representation U
of G with WG and the fact that U ∈ M(cid:0)K(HU ) ⊗ C0(G)(cid:1) generates C0(G).
element U ∈ M(cid:0)AU ⊗ C0(G)(cid:1) generates C0(G). On the other hand a representation U which
Proposition 2.3. Let U be a representation of G quasi-equivalent to WG. Then the unitary
generates C0(G) need not be quasi-equivalent to WG (even when G is a locally compact group).
Proof. From Theorem 2.2(3) it follows that
(cid:8)(ω ⊗ idB(L2(G)))(U ) ω ∈ B(HU )∗(cid:9) =(cid:8)(µ ⊗ idB(L2(G)))(WG) µ ∈ B(cid:0)L2(G)(cid:1)∗(cid:9).
For the second part it suffices to observe the following fact: let U1 and U2 be representations
of a locally compact quantum group G and let U be their direct sum ([40, Subsection 3.3.1]). If
Since for Y = (cid:8)(µ ⊗ idB(L2(G)))(WG) µ ∈ B(cid:0)L2(G)(cid:1)∗(cid:9) we have Y = C0(G), we see that Y
generates C0(G) as a C∗-algebra. Lemma 1.4 implies that U ∈ M(cid:0)AU ⊗ C0(G)(cid:1) generates C0(G).
U1 ∈ M(cid:0)K(HU1 ) ⊗ C0(G)(cid:1) generates C0(G) then so does U ∈ M(cid:0)K(HU1 ⊕ HU2 ) ⊗ C0(G)(cid:1).
U1 ⊕ U2 ∈ M(cid:0)K(ℓ2(Z) ⊕ C) ⊗ c0(Z)(cid:1) generates c0(Z). It cannot be quasi-equivalent to WZ, as
Let G = Z and let U1 and U2 be the regular and trivial representation of G. Then U =
then, according to Theorem 2.2, we would have a (normal) ∗-isomorphism between von Neumann
algebras A′′
WZ = L∞(T). However, the latter algebra is non-atomic, so we
would have a contradiction.
U = L∞(T) ⊕ C and A′′
(cid:3)
3. Closed quantum subgroups of locally compact quantum groups
This section is central to our paper. We begin by introducing two possible definitions of a closed
quantum subgroup of a given quantum group, the first of which appears in [43] and the second
was suggested to us by S.L. Woronowicz. Then we provide alternative, simplified descriptions for
both of them (Theorems 3.4 and 3.6) and analyze their mutual relations (Theorems 3.5 and 3.7).
We also present here a proof of Theorem 1.10.
The aforementioned definitions are as follows:
Definition 3.1 ([43, Definition 2.5]). Let G, H be locally compact quantum groups. Then H is
said to be a closed quantum subgroup of G in the sense of Vaes if there exists a morphism π from
H to G and a normal injective ∗-homomorphism γ : L∞(bH) → L∞(bG) such that
(3.1)
Definition 3.2 (Woronowicz). Let G, H be locally compact quantum groups. Then H is said to be
a closed quantum subgroup of G in the sense of Woronowicz if there exists a morphism π from H
to G such that the associated bicharacter V ∈ M(cid:0)C0(bG) ⊗ C0(H)(cid:1) generates C0(H).
γ(cid:12)(cid:12)C0( bH)◦Λ bH = Λ bG ◦bπ.
12
MATTHEW DAWS, PAWE L KASPRZAK, ADAM SKALSKI, AND PIOTR M. SO LTAN
The conditions above take as a starting point a morphism π from H to G. We will sometimes
say that H is a closed quantum subgroup of G in the sense of Vaes (respectively, in the sense of
Woronowicz) via the morphism π. In Section 4 we will explain why when both H and G are locally
compact groups both definitions are equivalent to the classical notion of H being (homeomorphic
to) a closed subgroup of G.
We will see later that the various examples of quantum subgroups considered in the literature
are all closed quantum subgroups in the sense of both Vaes and Woronowicz. The case of compact
and discrete subgroups is treated in Section 6. The non-compact examples of quantum subgroups
in [38, Sections 3 and 4] and those coming from Rieffel deformation presented in [20] are all closed
subgroups in the sense of Vaes and Woronowicz ([20, Section 6], cf. also Theorem 3.5). Another
class of examples is provided by the bicrossed product construction (see e.g. [44, 45]). If (G1, G2)
is a matched pair of locally compact quantum groups in the sense of [45, Definition 2.1] then bG1
is a closed quantum subgroup of the bicrossed product of G1 and G2 both in the sense of Vaes
and Woronowicz.
In the next theorem we note that the definition of Vaes can be reformulated in various simplified
ways (note especially condition (2), which does not assume a priori the existence of a homomor-
phism between H and G). In particular the definition of Vaes-Vainerman ([44, Definition 2.9]) is
equivalent to Definition 3.1.
Theorem 3.3. Let G, H be locally compact quantum groups. Then the following conditions are
equivalent:
(1) H is a closed quantum subgroup of G in the sense of Vaes;
(γ ⊗ γ)◦∆bH = ∆bG ◦γ;
(2) there exists a normal injective ∗-homomorphism γ : L∞(bH) → L∞(bG) such that
(3) there exists a normal injective ∗-homomorphism γ : L∞(bH) → L∞(bG) such that the unitary
(γ ⊗ id)(WH) ∈ L∞(bG) ¯⊗L∞(H) is a bicharacter from H to G -- in particular it belongs
to M(cid:0)C0(bG) ⊗ C0(H)(cid:1).
It will become clear from the proof of Theorem 3.3 that the map γ mentioned in point (2) is
the same as the one in (3) and still the same as the map γ from Definition 3.1. Moreover we show
in the proof that γ restricted to C0(bH) is an element of Mor(cid:0)C0(bH), C0(bG)(cid:1).
(2)⇒(3). The map γ′ = γ(cid:12)(cid:12)C0( bH) is naturally a representation of the C∗-algebra C0(bH) on L2(G).
Consider the unitary V = (γ′ ⊗ id)(WH) ∈ M(cid:0)K(cid:0)L2(G)(cid:1) ⊗ C0(H)(cid:1). Applying γ ⊗ γ ⊗ id to both
Proof of Theorem 3.3. (1)⇒(2) -- trivial.
sides of the equality
(3.2)
(∆bH ⊗ id)(WH) = WH
23WH
13
see that
(3.3)
12V ∗
23.
V13 = (WG
12)∗V23WG
precisely (1.13a). The application of γ ⊗ id ⊗ id to the pentagonal equation for WH implies that
(1.13b) holds and V is a bicharacter.
(viewed as an equality of operators in L∞(bH) ¯⊗ L∞(bH) ¯⊗ L∞(H)) and using the equation (3.2) we
The right side of the above expression belongs to M(cid:0)C0(bG) ⊗ K(cid:0)L2(G)(cid:1) ⊗ C0(H)(cid:1) and V13 has
legs only in the first and third tensor factor. Thus V ∈ M(cid:0)C0(bG) ⊗ C0(H)(cid:1). Note that (3.3) is
(3)⇒(1). Note first that as V ∈ M(cid:0)C0(bG) ⊗ C0(H)(cid:1), it follows that
because WH ∈ M(cid:0)C0(bH) ⊗ C0(H)(cid:1) generates C0(bH).
Let π ∈ Mor(cid:0)Cu
acter V and let bπ ∈ Mor(cid:0)Cu
0(H)(cid:1) be the strong quantum homomorphism associated with the bichar-
0(bH), C0(bG)(cid:1) be the dual quantum homomorphism. Then on one hand
γ′ = γ(cid:12)(cid:12)C0( bH) ∈ Mor(cid:0)C0(bH), C0(bG)(cid:1)
0(G), Cu
CLOSED QUANTUM SUBGROUPS OF LOCALLY COMPACT QUANTUM GROUPS
13
we have (recall the dependencies between V , bV and bπ listed in Subsection 1.3)
(id ⊗ γ′)(W
and on the other hand
bH) = bV
bV =(cid:0)ΛH ⊗ [Λ bG◦bπ](cid:1)(VV
W
bH = (ΛH ⊗ Λ bH)(VV
bH)
bH).
(this is (1.15) combined with (1.10)). Comparing the above and using the fact that
This can be rewritten as
(i.e. (1.5) for the quantum group bH) we obtain
(cid:0)ΛH ⊗ [γ′◦Λ bH](cid:1)(VV
(cid:0)id ⊗ [γ′ ◦Λ bH](cid:1)(W
(γ′◦Λ bH)(cid:0)(ω ⊗ id)(W
and upon application of (ω ⊗ id) with ω ∈ B(cid:0)L2(H)(cid:1)∗ yields
bH) =(cid:0)ΛH ⊗ [Λ bG ◦bπ](cid:1)(VV
bH) =(cid:0)id ⊗ [Λ bG ◦bπ](cid:1)(W
bH)(cid:1) = (Λ bG ◦bπ)(cid:0)(ω ⊗ id)(W
for any such ω. By (1.7) this implies that (3.1) holds. This ends the proof.
bH)(cid:1)
bH).
bH).
(cid:3)
Theorem 3.3 and a straightforward application of Theorem 2.2 yields the following result.
Theorem 3.4. Let G, H be locally compact quantum groups and suppose that V ∈ M(cid:0)C0(bG) ⊗
C0(H)(cid:1) is a bicharacter describing a morphism π from H to G. Then the following conditions are
equivalent:
(1) H is a closed quantum subgroup of G in the sense of Vaes via the morphism π;
(2) the bicharacter V is quasi-equivalent to WH (as a representation of H);
(3) A bV = AW bH .
Proof. The equivalence of (2) and (3) follows immediately from Theorem 2.2, as the sets appearing
in (3) here coincide with the analogous sets in condition (3) of that theorem (recall that bV = σ(V ∗),
bH = σ(WH)∗). The equivalence of (1) and (2) follows again from Theorem 2.2 and (the proof
(cid:3)
W
of) Theorem 3.3.
It now follows from Theorem 3.4 and Proposition 2.3 that there is a natural relation between
Definitions 3.1 and 3.2.
Theorem 3.5. If H is a closed quantum subgroup of G in the sense of Vaes, it is also a closed
quantum subgroup of G in the sense of Woronowicz.
Proof. Immediate consequence of Theorem 3.4 and Proposition 2.3.
(cid:3)
It is not clear if Definitions 3.1 and 3.2 are equivalent; in other words, whether Theorem 3.5
admits the converse. This would follow if we could show that a bi-character U ∈ M(cid:0)C0(bG)⊗C0(H)(cid:1)
describing a homomorphism from H to G which generates C0(H) must be quasi-equivalent to WH
(the example in Proposition 2.3 showed it need not be the case if we only assume that U is
a representation of H).
In the following sections we will show that in fact the equivalence of
Definitions 3.1 and 3.2 holds in many natural cases.
Now we show that Definition 3.2 also admits several natural equivalent reformulations. We
collect them in the next theorem.
0(G), Cu
Theorem 3.6. Let G, H be locally compact quantum groups and consider a homomorphism from
H to G described by a bicharacter V ∈ M(cid:0)C0(bG) ⊗ C0(H)(cid:1), a strong quantum homomorphism
π ∈ Mor(cid:0)Cu
0(H)(cid:1) and a right quantum homomorphism ρ ∈ Mor(cid:0)C0(G), C0(G) ⊗ C0(H)(cid:1).
Then the following conditions are equivalent (recall that bV := σ(V )∗ is a representation of bG):
(1) V ∈ M(cid:0)C0(bG) ⊗ C0(H)(cid:1) generates C0(H) (in other words H is a closed quantum subgroup
of G in the sense of Woronowicz);
14
MATTHEW DAWS, PAWE L KASPRZAK, ADAM SKALSKI, AND PIOTR M. SO LTAN
(3.4)
(2) A bV = C0(H);
(3) the right quantum homomorphism ρ is strongly non-degenerate:
ρ(cid:0)C0(G)(cid:1)(cid:0)C0(G) ⊗ 1C0(H)(cid:1) = C0(G) ⊗ C0(H)
0(H);
(2)⇔(3). We compute:
0(G)(cid:1) = Cu
(in particular the left hand side of (3.4) is contained in the right hand side);
0(G)(cid:1) = C0(H).
(4) π(cid:0)Cu
(5) (ΛH ◦π)(cid:0)Cu
Proof. (1)⇔(2). This follows from Proposition 2.1 and an obvious fact that V ∈ M(cid:0)C0(bG)⊗C0(H)(cid:1)
generates C0(H) if and only if bV ∈ M(cid:0)C0(H) ⊗ C0(bG)(cid:1) generates C0(H) (cf. Remark 1.5).
ρ(cid:0)C0(G)(cid:1)(cid:0)C0(G) ⊗ 1(cid:1)
=(cid:8)V(cid:0)(x · ω ⊗ id)(WG)∗ ⊗ 1(cid:1)V ∗(y ⊗ 1) x ∈ C0(bG), y ∈ C0(G), ω ∈ B(cid:0)L2(G)(cid:1)∗(cid:9) -- k·k
23(x ⊗ y ⊗ 1)(cid:1) x ∈ C0(bG), y ∈ C0(G), ω ∈ B(cid:0)L2(G)(cid:1)∗(cid:9) -- k·k
=(cid:8)(ω ⊗ id ⊗ id)(cid:0)V23(WG
=(cid:8)(ω ⊗ id ⊗ id)(cid:0)V ∗
12)∗(x ⊗ y ⊗ 1)(cid:1) x ∈ C0(bG), y ∈ C0(G), ω ∈ B(cid:0)L2(G)(cid:1)∗(cid:9) -- k·k
=(cid:8)(ω ⊗ id ⊗ id)(cid:0)V ∗
= C0(G) ⊗(cid:8)(ω ⊗ id)(V ) ω ∈ B(cid:0)L2(G)(cid:1)∗(cid:9) -- k·k = C0(G) ⊗ A bV
13(x ⊗ y ⊗ 1)(cid:1) x ∈ C0(bG), y ∈ C0(G), ω ∈ B(cid:0)L2(G)(cid:1)∗(cid:9) -- k·k
In the third equality we used the bicharacter property of V (Eq. (1.13a)) and in the fourth equality
we used the fact that WG ∈ M(cid:0)C0(bG) ⊗ C0(G)(cid:1) is unitary. The above computation shows that
A bV = C0(H) if and only if ρ(cid:0)C0(G)(cid:1)(cid:0)C0(G) ⊗ 1(cid:1) = C0(G) ⊗ C0(H).
(2)⇔(5). Taking into account (1.7), (1.6) and (1.12) we find that V =(cid:0)id ⊗ [ΛH◦π](cid:1)(WG) and
(ΛH ◦π)(cid:0)(ω ⊗ id)(WG)(cid:1) = (ω ⊗ id)(V ).
12)∗V ∗
13(WG
(3.5)
Thus
(ΛH ◦π)(cid:0)Cu
0(G)(cid:1) = (ΛH ◦π)(cid:0)(cid:8)(ω ⊗ id)(WG) ω ∈ B(cid:0)L2(G)(cid:1)∗(cid:9) -- k·k(cid:1)
=(cid:8)(ΛH ◦π)(cid:0)(ω ⊗ id)(WG)(cid:1) ω ∈ B(cid:0)L2(G)(cid:1)∗(cid:9) -- k·k = A bV .
(4)⇔(5). Since (4)⇒(5) is clear, it remains to show the converse implication. Consider the
universal lift of V u ∈ M(cid:0)C0(bG) ⊗ Cu
0(H)(cid:1) defined as V u = (id ⊗ π)(W G), cf. [27, Section 4]. To
show the desired implication it suffices to establish the following equality:
(cf. the proof of (2)⇔(5)). Noting that1
(cid:8)(ω ⊗ id)(V u) ω ∈ B(cid:0)L2(G)(cid:1)∗(cid:9) -- k·k = Cu
0(H)
13 = V ∗
V u
12
W H
23V12(W H
23)∗
(3.6)
we compute:
(cid:8)(ω ⊗ id)(V u) ω ∈ B(cid:0)L2(G)(cid:1)∗(cid:9) -- k·k
=(cid:8)(ω ⊗ µ ⊗ id)(cid:0)V ∗
=(cid:8)(ω ⊗ µ ⊗ id)(cid:0)WH
=(cid:8)(η ⊗ µ ⊗ id)(cid:0)W H
=(cid:8)(η ⊗ µ ⊗ id)(WH
=(cid:8)(η · x ⊗ id)(W H
12
W H
23V12(W H
12
23V12(WH
23WH
12(W H
W H
23)∗(cid:1) ω ∈ B(cid:0)L2(G)(cid:1)∗, µ ∈ B(cid:0)L2(H)(cid:1)∗(cid:9) -- k·k
23)∗(cid:1) ω ∈ B(cid:0)L2(G)(cid:1)∗, µ ∈ B(cid:0)L2(H)(cid:1)∗(cid:9) -- k·k
23)∗(cid:1) η ∈ B(cid:0)L2(H)(cid:1)∗, µ ∈ B(cid:0)L2(H)(cid:1)∗(cid:9) -- k·k
13) η ∈ B(cid:0)L2(H)(cid:1)∗, µ ∈ B(cid:0)L2(H)(cid:1)∗(cid:9) -- k·k
13) η ∈ B(cid:0)L2(H)(cid:1)∗, x ∈ C0(bH)(cid:9) -- k·k = Cu
0(H).
1To prove (3.6) we first note that we have (∆bH
0(H)(cid:1) described in [27, Proposition 4.14]. This can be rewritten as WH
in M(cid:0)C0(bH) ⊗ Cu
Slicing with ω ∈ B(cid:0)L2(H)(cid:1)∗ on the left leg we obtain the formula (ΛH ⊗ id)∆u
x ∈ Cu
H)(V u) = V u
reads W H
0(H). Now we apply id ⊗ ΛH ⊗ id to both sides of (id ⊗ ∆u
13 which is (3.6).
23)∗ = V12 V u
V12(W H
V u
23 as WH is the unique lift of WH to a bicharacter
23)∗.
H(x) = WH(cid:0)ΛH(x) ⊗ 1(cid:1)(WH)∗ for all
13. By the previous formula this
13 = W H
12(W H
23WH
W H
⊗id)WH = WH
12
WH
13
12
23
CLOSED QUANTUM SUBGROUPS OF LOCALLY COMPACT QUANTUM GROUPS
15
W H
W H
12(W H
23)∗ = WH
12
13 (see derivation of formula (3.6)).
The second equality follows from the unitarity of V , in the third one we used the fact that
V ∈ M(cid:0)C0(bG) ⊗ C0(H)(cid:1) generates C0(H) and in the fourth equality we used the equation
23WH
Condition (3) in Theorem 3.6 classically corresponds to properness and freeness of the natural
action of H on G induced by the homomorphism from H to G, see Section 4. It was introduced
in the context of quantum groups by Podle´s in his thesis [31, Definicja 2.2], see also [32] and
[38, Proposition 2.3] for a complete discussion. Condition (4) is a natural reflection of a general
principle that injectivity on the level of point transformations is equivalent to surjectivity on the
level of induced transformations on algebras of functions (but cf. Theorem 1.1(5)). One can ask
the following question: are the conditions (1)-(3) above equivalent in general to the surjectivity
of the extension of π to multiplier algebras ¯π : M(cid:0)Cu
0(G)(cid:1) → M(cid:0)Cu
0(H)(cid:1)? The Pedersen-Tietze
theorem ([46, Theorem 2.3.9]) implies that ¯π is surjective if π is, provided that Cu
Cu
0(H)) is σ-unital.
With the results obtained in this section in hand we are ready to give a proof of Theorem 1.10.
0(G) (and hence
(cid:3)
0(G), Cu
Proof of Theorem 1.10. We are assuming that
Mor(cid:0)Cu
bπ ∈ Mor(cid:0)Cu
with a Vaes-closed subgroup of bH. Let bV be the bicharacter associated to bπ:
0(H)(cid:1) is an isomorphism. Consider the dual strong quantum homomorphism
0(bG)(cid:1). By (1.16) bπ is an isomorphism. We will show that bπ identifies bG
0(bH), Cu
the strong quantum homomorphism π ∈
bH).
bV =(cid:0)ΛH ⊗ [Λ bG◦bπ](cid:1)(VV
As
for all x ∈ Cu
W
bH(cid:0)Λ bH(x) ⊗ 1(cid:1)(W
0(bH) (see Footnote 1), it follows that
bV(cid:0)Λ bH(x) ⊗ 1(cid:1)(bV )∗ =(cid:0)Λ bH ⊗ [Λ bG ◦bπ](cid:1)∆u
bH)∗ = (Λ bH ⊗ id)∆u
bH(x)
bH(x).
so
(id ⊗ ∆u
bH). Then
bH) = Y . Set X = (ϕ ⊗bπ−1)(VV
bH)(X) = X12X13 because bπ−1 intertwines the coproducts. Further
Let Y ∈ M(cid:0)K(HY ) ⊗ C0(H)(cid:1) be a representation of H. By the results of [22, 27], there exists a
unique ϕ ∈ Mor(cid:0)Cu
0(H), K(H)(cid:1) such that (ϕ ⊗ Λ bH)(VV
(cid:0)id ⊗ Λ bH ⊗ [Λ bG ◦π](cid:1)(X12X13) = bV23(cid:0)(id ⊗ Λ bH)(X)(cid:1)12(bV23)∗,
12bV23(cid:0)(id ⊗ Λ bH)(X)(cid:1)12
(cid:0)id ⊗ [Λ bG◦π](cid:1)(X)
⊤bV =(cid:0)(id ⊗ Λ bH)(X)(cid:1)∗
⊤bV . However(cid:0)id⊗[Λ bG◦π](cid:1)(X) = Y , so
which means that(cid:0)id⊗[Λ bG◦π](cid:1)(X)
⊤bV is equivalent to IHY
that, as Y was arbitrary, bV is right-absorbing. It follows from [40] (see remark before Proposition
2.3) that bV is quasi-equivalent to W
bH. By Theorem 3.4 bG is a closed quantum subgroup of bH in
By Theorem 3.3 and the comment after it, there exists a morphism bγ1 ∈ Mor(cid:0)C0(G), C0(H)(cid:1)
bγ1◦ΛG = ΛH ◦π.
Applying identical reasoning to bπ−1 we obtain the existence ofbγ2 ∈ Mor(cid:0)C0(H), C0(G)(cid:1) such that
bγ2◦ΛH = ΛG ◦π−1
(note that we use once again the fact that bπ−1 = dπ−1). Since ΛG and ΛH are surjections we see
that bγ1 and bγ2 are mutually inverse and we can set πr =bγ1.
We recall from Subsection 1.2 that the Fourier algebra and the Fourier-Stieltjes algebra of a
the sense of Vaes.
such that
(cid:3)
locally compact quantum group are the Banach spaces
AG = L∞(bG)∗
and
BG = Cu
0(bG)∗
16
MATTHEW DAWS, PAWE L KASPRZAK, ADAM SKALSKI, AND PIOTR M. SO LTAN
which we embedded into C0(G) and M(cid:0)Cu
0(G)(cid:1) respectively with the maps
AG ∋ ω 7−→ (ω ⊗ id)(WG),
G).
BG ∋ η 7−→ (η ⊗ id)(VV
We also note that AG embeds in BG via Λ∗
bG ◦ı∗, where ı is the embedding C0(G) ֒→ L∞(G) (we
will also use the symbol "ı" to denote the analogous embedding for other quantum groups). It
is easy to check that this embedding is isometric. Moreover the induced embedding of AG into
M(cid:0)Cu
0(G)(cid:1) actually embeds the Fourier algebra into Cu
0(G).
Theorem 3.7. Let H be a closed subgroup of G in the sense of Woronowicz via the morphism
π : Cu
0(G) → Cu
(1) π restricts to a map T : AG → AH which has dense range (for the AH norm);
0(H). Then the following are equivalent:
(2) bπ∗ : BG → BH restricts to a map S : AG → AH which has dense range;
(3) H is a closed subgroup of G in the sense of Vaes.
Moreover we can replace "dense range" by "surjection" in (1) and (2).
If these conditions
hold, then S and T are the same map, which is nothing but the pre-adjoint of the implicit map
bG(cid:0)ı∗(ω)(cid:1) ∈ Cu
γ : L∞(bH) → L∞(bG) appearing in (3).
Proof. Let ω ∈ L∞(bG)∗ = AG, set µ = Λ∗
0(G)(cid:1), so a = (µ ⊗ id)(VVG). Then
BG ⊂ M(cid:0)Cu
so that π(a) is (the image of) bπ∗(µ) in BH. It is now clear that (1) and (2) are equivalent.
isometries, S must be bounded. Set γ = S∗ : L∞(bH) → L∞(bG), so as S has dense range, γ is
π(a) = (µ ⊗ π)(VVG) =(cid:0)bπ∗(µ) ⊗ id(cid:1)(VVH),
0(bG)∗, and let a be the image of ω in
If (2) holds then the map S satisfies Λ∗
injective. Then we have that
bG ◦ ı∗, and as Λ∗
bG ◦ ı∗ and Λ∗
bH ◦ ı∗ are
bH ◦ ı∗ ◦ S = bπ∗ ◦ Λ∗
bH (cid:12)(cid:12)Cu
0 ( bH) = ı∗∗ ◦Λ∗∗
bG ◦bπ∗∗(cid:12)(cid:12)Cu
0 ( bH) = ı◦Λ bG◦bπ.
γ ◦Λ bH = S∗ ◦Λ bH = S∗◦ı∗∗◦Λ∗∗
As γ is weak∗-continuous, it now follows that γ is a ∗-homomorphism, and so (3) holds.
Finally, if (3) holds, then we have a normal injective ∗-homomorphism γ : L∞(bH) → L∞(bG)
with γ ◦ı◦Λ bH = ı◦Λ bG◦bπ. Thus, for ω ∈ L∞(bG)∗, we have that
and so bπ∗ restricts to a map S : AG → AH. As γ is injective and hence an isometry, γ∗ is a
surjection and so S, which agrees with γ∗ once appropriate identifications are made, is also a
surjection. This shows (2), and also demonstrates the claim about replacing "dense range" by
"surjection".
(cid:3)
bG ◦ı∗)(ω)(cid:1) = (Λ∗
bπ∗(cid:0)(Λ∗
bH ◦ı∗◦γ∗)(ω),
Remark 3.8. From Theorem 3.7 we immediately see that H is a closed subgroup of G in the
0(H), if and only if π restricts to a surjection
sense of Vaes, via the morphism π : Cu
AG → AH. In the classical case, the Herz restriction theorem ([15, 1]) says exactly that if H is a
closed subgroup of G, then the restriction map (which is nothing but π : C0(G) → C0(H)) gives a
surjection AG → AH . In other words the definition of a Vaes-closed subgroup is tailored exactly
so that the quantum version of the Herz restriction theorem holds.
0(G) → Cu
4. Commutative case
Let now G and H be locally compact groups, so in particular Cu
0(H) =
C0(H). Any homomorphism from H to G (in the sense of quantum groups -- as defined in
Subsection 1.3) is then described by a π ∈ Mor(cid:0)C0(G), C0(H)(cid:1). Moreover π is necessarily of the
form π(f ) = f ◦θ, where θ : H → G is a continuous homomorphism (cf. Theorem 1.14).
0(G) = C0(G) and Cu
Given a situation as above, consider the natural right action of H on the topological space G
given by
G × H ∋ (g, h) 7−→ g · h = g θ(h) ∈ G.
(4.1)
CLOSED QUANTUM SUBGROUPS OF LOCALLY COMPACT QUANTUM GROUPS
17
Let us also introduce the so called canonical map γ : G × H → G × G for this action
γ(g, h) = (g, g · h) =(cid:0)g, g θ(h)(cid:1).
(4.2)
([37]). Let ρ ∈ Mor(cid:0)C0(G), C0(G) ⊗ C0(H)(cid:1) and Γ ∈ Mor(cid:0)C0(G) ⊗ C0(G), C0(G) ⊗ C0(H)(cid:1) be
the morphisms of C∗-algebras corresponding to (4.1) and (4.2):
f ∈ C0(G),
ρ(f )(g, h) = f (g · h),
Γ(F )(g, h) = F(cid:0)γ(g, h)(cid:1), F ∈ C0(G) ⊗ C0(G),
g ∈ G, h ∈ H,
g ∈ G, h ∈ H.
Lemma 4.1. Let θ : H → G be a continuous homomorphism with corresponding action of H on
G as in (4.1) and canonical map γ : G × H → G × G. Then the following are equivalent:
(1) θ is a homeomorphism onto its closed image;
(2) the action of H on G is free and proper i.e. γ is injective and proper;
(3) (cid:0)ρ(cid:0)C0(G)(cid:1)(cid:0)C0(G) ⊗ 1(cid:1) = C0(G) ⊗ C0(H).
Proof. (1)⇒(2). If θ is a homeomorphism, then γ is a homeomorphism onto its range, as for (g, g′)
in the range of γ (i.e. g′ = g · h for some h ∈ H) we have γ−1(g, g′) =(cid:0)g, θ−1(g−1g′)(cid:1). Hence γ is
in particular injective and proper.
(2)⇒(1). Assume that γ is injective and proper. Clearly θ is then injective. Similarly, if θ
were not proper, then there would be a compact set K ⊂ G with θ−1(K) non-compact. But then
γ−1(cid:0){e} × K(cid:1) = {e} × θ−1(K) would not be compact either.
Hence θ is injective and proper. Proper continuous maps between locally compact spaces
are automatically closed ([8, Chapter 1, §10]). Hence θ has a closed image, and as a bijective
continuous closed map is in fact a homeomorphism.
(2)⇔(3). Note first that
(cid:0)ρ(cid:0)C0(G)(cid:1)(cid:0)C0(G) ⊗ 1(cid:1) = Γ(cid:0)C0(G) ⊗ C0(G)(cid:1).
Hence (3) is equivalent to the fact that γ is injective and proper by Theorem 1.1.
(cid:3)
Theorem 4.2. Suppose that G and H are locally compact groups. Then the following conditions
are equivalent:
(1) H is a closed quantum subgroup of G in the sense of Vaes;
(2) H is a closed quantum subgroup of G in the sense of Woronowicz;
(3) H is homeomorphic to a closed subgroup of G.
Proof. Condition (1) implies (2) by Theorem 3.5. Conditions (2) and (3) are equivalent by Lemma
4.1 and Theorem 3.6. It remains to note that (3) implies (1). By theorem 3.3 this is precisely [52,
Corollary 4.2.6] (which is a consequence of [1, Theorem (3.23)]).
(cid:3)
Remark 4.3.
(1) Let G be a locally compact group and let H be a locally compact quantum group. If H is a
closed subgroup of G in the sense of Woronowicz then by Theorem 3.6 there is a surjection
from C0(G) onto C0(H), so that H is in fact a classical group. By Theorem 4.2, H is then
also a closed subgroup of G in the usual sense.
(2) Let H be a locally compact group and let G be a locally compact quantum group.
If
H is a closed subgroup of G in the sense of Woronowicz then the associated morphism
0(G) → C0(H) factors through the algebra C0( G), where G is the intrinsic group of
π : Cu
G as defined by Kalantar and Neufang (a locally compact group associated to G, see [19]).
It follows from Theorem 3.6 that H is a closed subgroup of G (again in the usual sense).
Theorem 3.7 shows that the existence of an injective normal ∗-homomorphism from vN(H)
to vN(G) is naturally very closely related to the Herz restriction theorem (cf. Remark 3.8). To
analyze the situation closer assume that H is a closed subgroup of a locally compact group G and
consider the following statements:
(1) the restriction map from C0(G) to C0(H) yields a surjective map from AG to AH ;
18
MATTHEW DAWS, PAWE L KASPRZAK, ADAM SKALSKI, AND PIOTR M. SO LTAN
(2) the prescription λh 7→ λ(G)
L2(H) and λ(G)
injective ∗-homomorphism from vN(H) to vN(G);
h
h , h ∈ H (where λh denotes the (unitary) left shift by h on
the corresponding (unitary) left shift by h on L2(G)) extends to a normal
(3) the restriction of the left regular representation of G to H is quasi-equivalent to the left
regular representation of H.
It is very easy to see that they are all logically equivalent ((2) is essentially the definition of
quasi-equivalence in (3), and the equivalence of (1) and (2) follows from a basic functional analytic
argument, appearing already in [16, Section 0]). The first condition is the Herz restriction theorem.
The third one can be viewed as a statement related to the theory of induced representations, and
the induction-restriction procedure, as [26, Theorem 4.2] states that the left regular representation
of G is the induction of the left regular representation of H. Interestingly, we could not locate an
explicit statement of the condition (3) in literature. In the remainder of this section we will give
an alternative proof of the implication (3)⇒(1) in Theorem 4.2. In particular this gives a new
proof of Herz restriction theorem (cf. Remark 3.8). Our reasoning is based on existence of locally
Baire cross-sections for the canonical projection G → G/H ([21]).
Following the notation of [21, Section 4] we let q : G/H → G be a locally bounded Baire
cross-section to the canonical quotient map G → G/H. We denote by Φ the bijection
Let µ and β be Haar measures on G and H respectively and let λ be a quasi-invariant measure
on G/H with associated ρ-function ρ ([13, Section 2.6]). In [21, Section 4] E.T. Kehlet shows that
(G/H) × H ∋(cid:0)[g], h(cid:1) 7−→ q([g])h ∈ G.
is a unitary map L2(G, µ) → L2(cid:0)(G/H) × H, λ × β(cid:1). We note that
and the function ρ satisfies
ψ 7−→ ρ− 1
2 · ψ◦Φ
Φ(cid:0)[g], h(cid:1)h′ = Φ(cid:0)[g], hh′(cid:1),
ρ(gh′) = ∆H (h′)
∆G(h′) ρ(g)
(4.3)
(4.4)
(4.5)
([13]).
We identify L2(G/H, λ) ⊗ L2(H) with L2(cid:0)(G/H) × H, λ × β(cid:1) in the usual way (the respective
measures are regular) and define a unitary T : L2(G/H, λ) ⊗ L2(H) → L2(G) as the inverse of
(4.3), i.e.
(T ψ)(g) = ρ(g)
1
2 ψ(cid:0)Φ−1(g)(cid:1).
1
1
1
1
2 ∆H (h′)
2(cid:16)(cid:0)(1 ⊗ Rh′ )T ∗(cid:1)ψ(cid:17)(cid:0)Φ−1(g)(cid:1)
2 (T ∗ψ)(cid:0)[g0], h0h′(cid:1)
2 ρ(cid:0)Φ(cid:0)[g0], h0h′(cid:1)(cid:1)− 1
2 (cid:16) ∆H (h′)
∆G(h′)(cid:17)− 1
2 ∆H (h′)
2 ∆H (h′)
2 ρ(gh′)− 1
2 ∆H (h′)
2 ψ(gh′)
2
1
1
1
1
1
= ρ(g)
= ρ(g)
= ρ(g)
= ρ(g)
= ∆G(h′)
1
2 ψ(gh′) = (RG
h′ ψ)(g).
2 ψ(cid:0)Φ(cid:0)[g0], h0h′(cid:1)(cid:1)
ρ(g)− 1
2 ψ(gh′)
For h′ ∈ H let Rh′ be the unitarized operator of right translation by h′ on L2(H) and let RG
h′
denote the operator of right translation by h′ on L2(G). Fix g ∈ G and let Φ−1(g) = (cid:0)[g0], h0(cid:1).
Taking into account (4.4) and (4.5) we compute
(cid:16)(cid:0)T (1 ⊗ Rh′ )T ∗(cid:1)ψ(cid:17)(g) = ρ(g)
Thus T (1 ⊗ Rh′ )T ∗ = RG
h′ . This means that the (right) group von Neumann algebra vN(H)
is isomorphic to the von Neumann subalgebra of vN(G) generated by the right shifts on G by
elements from the subgroup H. This embedding is the map γ from Definition 3.1. In particular
H is a closed subgroup of G in the sense of Vaes.
CLOSED QUANTUM SUBGROUPS OF LOCALLY COMPACT QUANTUM GROUPS
19
5. Cocommutative case
Let again G and H be locally compact groups. Recall that in this case the dual locally compact
r(G),
quantum groups bG and bH of G and H are respectively defined by putting C0(bG) = C∗
C0(bH) = C∗
Theorem 5.1. Let π be a morphism from bH to bG and let, as usual, bπ denote the dual morphism
r(H). We have the following result.
from G to H, so that
bπ : C0(H) ∋ f 7−→ f ◦θ ∈ M(cid:0)C0(G)(cid:1),
(f ∈ C0(H))
for some continuous homomorphism θ : G → H. Then the following conditions are equivalent:
(1) bH is a closed quantum subgroup of bG in the sense of Vaes (via the morphism π);
(2) bH is a closed quantum subgroup of bG in the sense of Woronowicz (via the morphism π);
(3) θ maps G onto H and the induced map θ : G/ker θ → H is a homeomorphism.
Proof. That (1)⇒(2) is Theorem 3.5.
Suppose that (2) holds, so that the morphism π : C∗(G) → C∗(H) maps to C∗(H) and is
surjective. Since G and H are classical groups, the algebras AH , BH , AG and BG (as defined in
Subsection 1.2) are the classical Fourier and Fourier-Stieltjes algebras of H and G respectively.
In particular we have that BH = C∗(H)∗, and similarly BG = C∗(G)∗. Thus π∗ : BH → BG is
weak∗-weak∗-continuous. Let G0 be the closure of the image of θ in H, and let θ0 : G → G0 be
the corestriction of θ. By [18, Lemma 4.2] it follows that θ0 is an open surjection. We claim that
G0 = H, from which (3) will follow. Indeed, if G0 6= H then as AH (and hence also BH ) is a
regular algebra of functions on H (see [12, Lemme 3.2] or [52, Proposition 4.1.8]) we can find a
non-zero b ∈ BH with b(s) = 0 for all s ∈ G0. As a map between function algebras, π∗ is simply
π∗(b) = b◦θ, and so π∗(b) = 0. However, as π is surjection, π∗ is an isometry, and so π∗(b) 6= 0, a
contradiction. Thus G0 = H as required.
If (3) holds then as both K = ker θ and G/K are locally compact groups in their own right, they
carry Haar measures, which we may normalize so that the Weyl formula holds: for f ∈ C00(G),
Z
f (s) ds = Z
Z
G
G/K
K
It is not hard to see that the map
IK : L1(G) ∋ f 7−→Z
K
f (st) dt d(sK).
f (st) dt ∈ L1(G/K)
is an algebra homomorphism and a metric surjection; see [28, Section 1.9.12] for example. We
notice that then I ∗
K : L∞(G/K) → L∞(G) is an injective normal ∗-homomorphism which inter-
twines the coproducts. As G/K is homeomorphic to H, the Haar measures on H and on G/K
are proportional, and so the map
γ0 : L∞(H) ∋ F 7−→ F ◦ θ ∈ L∞(G/K)
is well-defined, and is hence a normal ∗-isomorphism which intertwines the coproduct. Then set
γ = I ∗
K ◦γ0 : L∞(H) → L∞(G). So γ is an injective normal ∗-homomorphism which intertwines
(cid:3)
the coproducts, and a simple check shows that γ(cid:12)(cid:12)C0(H) = bπ, so (1) holds.
H is a closed subgroup of bG in the sense of Woronowicz then π : C∗(G) → Cu
Remark 5.2. Let G be a locally compact group and let H be a locally compact quantum group. If
0(H) is a surjection,
0(H) is cocommutative, hence of the form C∗(H) for some H, and it then follows that H
and so Cu
is a quotient of G.
Let us give some indications of how the proof of [18, Lemma 4.2] proceeds. Firstly, arguing
as in the proof of (3)⇒(1) above, it is not hard to reduce the problem to the case when θ is an
injection. The key result is then [6, Theorem 1.3] which tells us that π∗(BH ) contains AG (as it
is a weak∗-closed, conjugate closed, C∗(G)-module which, as a space of functions on G, separates
the points of G; this final claim uses the assumption that θ is injective).
20
MATTHEW DAWS, PAWE L KASPRZAK, ADAM SKALSKI, AND PIOTR M. SO LTAN
These ideas can be readily generalized to the setting of locally compact quantum groups.
Proposition 5.3. Let G and H be locally compact quantum groups and let π : Cu
0(H) be a
strong quantum homomorphism identifying H as a Woronowicz-closed subgroup of G. Furthermore,
suppose that π∗(BH) contains the image of AG under the map Λ∗
G. Then G and H are isomorphic.
0(G) → Cu
Proof. As π is onto, π∗ is an isometry onto its range, and so there is an isometric map φ : L∞(G)∗ →
Cu
G; clearly φ is a Banach algebra homomorphism. Let
0(H)∗ with π∗ ◦ φ = Λ∗
ψ = φ∗(cid:12)(cid:12)Cu
G (cid:12)(cid:12)Cu
Cu
0(G)
π /
/ Cu
0(H)
Then ψ◦π = φ∗ ◦π∗∗(cid:12)(cid:12)Cu
0 (H) : Cu
0(H) −→ L∞(G).
0 (G) = Λ∗∗
0 (G) = ΛG, so we have the diagram
✉
✉
✉
✉
ψ
✉
✉
✉
✉
ΛG
z✉
C0(G)
0(H) → Cu
As π is onto, it follows that ψ is a ∗-homomorphism. Therefore it follows easily that ψ intertwines
the coproducts. From the results of [27, Section 4] there is a strong quantum homomorphism
0(G) with ΛG ◦ ψ0 = ψ. Thus ΛG ◦ ψ0 ◦ π = ΛG. By passing to bicharacters and
ψ0 : Cu
0(G). In particular, π must be
applying [27, Lemma 4.13] it follows that ψ0◦π is the identity on Cu
injective, and so an isomorphism. Thus the quantum groups G and H are isomorphic by Theorem
1.10.
(cid:3)
6. Compact and discrete cases
In this section we establish the equivalence of Definitions 3.1 and 3.2 when a potential quantum
subgroup is compact (Theorem 6.1) and when the "larger" quantum group is discrete (Theorem
6.2). The first of these results shows in particular that if both quantum groups in question are
compact, the definitions studied in this paper coincide with the one currently adopted in literature
(see [3], [4], etc.); the second can be thought of as the generalization of the Herz restriction theorem
to the context of discrete quantum groups.
6.1. Compact subgroups. Let G and H be locally compact quantum groups and assume further
that H is compact. We will show in Theorem 6.1 that H is a closed subgroup of G in the sense
of Vaes if and only if it is a closed subgroup of G in the sense of Woronowicz. However, before
proceeding with this theorem let us make the following observation: consider a homomorphism
Theorem 1.1(5) one could define injectivity of the homomorphism from H to G as the property
from H to G described by π ∈ Mor(cid:0)Cu
0(G), Cu(H)(cid:1) (remember that H is compact). Based on
that the range of π is strictly dense in M(cid:0)Cu(H)(cid:1). But Cu(H) is unital, so strict density of the
range of π is equivalent to its norm-density. Moreover, since the image of a C∗-algebra under
a ∗-homomorphisms is closed, π must be a surjection. By Theorem 3.6 this means that H is a
closed subgroup of G in the sense of Woronowicz. In other words, the above argument shows that
a compact quantum group H with an injective homomorphism into G is automatically a closed
subgroup of G in the sense of Woronowicz (thus by Theorem 6.1 it is also closed in the sense of
Vaes). In particular the notion of a quantum subgroup used e.g. in [38, Sections 4 and 5], [36], [4]
is identical to those given in Definitions 3.1 and 3.2.
Before proceeding let us also quickly note that a Woronowicz-closed subgroup of a compact
quantum group is automatically compact (so, by Theorem 6.1 it is also Vaes-closed). The reason
for this is that a quotient of a unital C∗-algebra is obviously unital (cf. Theorem 3.6(4)).
Theorem 6.1. Let H be a closed subgroup of G in the sense of Woronowicz and assume that H
is compact. Then H is a closed subgroup of G in the sense of Vaes.
Proof. The subgroup H is compact, so we can write Cu
0(bH) = c0(bH), as the quantum group bH
is discrete and hence coamenable. Moreover the C∗-algebra c0(bH) is a c0-direct sum of matrix
algebras. It is not difficult to see the following
/
/
z
CLOSED QUANTUM SUBGROUPS OF LOCALLY COMPACT QUANTUM GROUPS
21
• the multiplier algebra M(cid:0)c0(bH)(cid:1) is canonically isomorphic to the double dual c0(bH)∗∗,
• for any C∗-algebra C of operators and any Φ ∈ Mor(cid:0)c0(bH), C(cid:1) the extension of Φ to a
mapping M(cid:0)c0(bH)(cid:1) → M(C) ⊂ C′′ is σ-weakly continuous; in fact the extension of Φ to
multipliers coincides with its normal extension ([29, Theorem 3.7.7]).
Now, just as in the proof of Theorem 3.6 (Eq. (3.5)) we have
Since V ∈ M(cid:0)C0(G) ⊗ C0(H)(cid:1) generates C0(H), we have by Proposition 2.1 that
(ΛH ◦π)(cid:0)(ω ⊗ id)(WG)(cid:1) = (ω ⊗ id)(V ).
(6.1)
is dense in C(H).
follows that
Therefore if ω ∈ L∞(H)∗ is non-zero then it must be non-zero on some element (η ⊗ id)(V ). It
(cid:8)(η ⊗ id)(V ) η ∈ L∞(bG)∗(cid:9)
η(cid:0)(id ⊗ ω)(V )(cid:1) 6= 0,
so (id ⊗ ω)(V ) 6= 0. In view of (6.1) this means that Λ bG◦bπ is injective on the subspace
which coincides with the Fourier algebra
(cid:8)(id ⊗ ω)(WH) ω ∈ L∞(H)∗(cid:9) ⊂ c0(bH)
AbH =(cid:8)(id ⊗ ω)(WH) ω ∈ L∞(H)∗(cid:9),
as bH is coamenable (cf. (1.8)). This last subspace contains the Pedersen ideal of c0(bH) (cf. (6.3)).
By [7, Proposition II.8.2.4] this implies injectivity of Λ bG◦bπ on all of c0(bH). Finally Λ bG◦bπ remains
injective after extension to ℓ∞(bH) = M(cid:0)c0(bH)(cid:1) because this extension coincides with the extension
to the multiplier algebra and such extensions always preserve injectivity ([24, Proposition 2.1]). (cid:3)
The arguments similar to these above appeared earlier in [35], an article which studies the
relations between compact quantum subgroups of a coamenable locally compact quantum group
G and left invariant C∗-subalgebras of C0(G).
6.2. Subgroups of discrete quantum groups. The main result of this subsection is the fol-
lowing:
Theorem 6.2. Let H be a closed subgroup of G in the sense of Woronowicz and assume that G
is discrete. Then H is discrete and H is a closed subgroup of G in the sense of Vaes.
We will prove Theorem 6.2 by generalizing to the setting of discrete quantum groups the theorem
of Herz [15], [1, Proposition 3.23] and using Theorem 3.7 (cf. Remark 3.8).
Since G is a discrete quantum group, the C∗-algebra C0(G) = Cu
0(G) = c0(G) is a c0-direct
sum:
Mnα
c0(G) = Mα∈R
0(G)(cid:1) is in this case
and the embedding of BG into M(cid:0)Cu
(and AG is then mapped to the space of slices of WG = WG with normal functionals on L∞(bG)).
BG = Cu(bG)∗ ∋ η 7−→ (η ⊗ id)(VVG) ∈ M(cid:0)c0(G)(cid:1) = ℓ∞(G)
In particular one can use the functionals dual to the canonical basis of the Hopf ∗-algebra sitting
inside C(G) ([51, Theorem 2.2], [5, Theorem 5.1]). These are normal and we easily see that their
image in the mapping (6.2) spans the Pedersen ideal c00(G) of c0(G). The ideal c00(G) is the
algebraic direct sum of the same family of matrix algebras. On the other hand these functionals
(6.2)
are linearly dense in L∞(bG)∗ (they correspond to density matrices on L2(bG) which are of finite
rank). Therefore
c00(G) ⊂ AG
(6.3)
with AG viewed as a subspace of c0(G).
Theorem 6.3. The space AG is the closure in BG of c00(G).
22
MATTHEW DAWS, PAWE L KASPRZAK, ADAM SKALSKI, AND PIOTR M. SO LTAN
Proof. As we mentioned before stating Theorem 6.3 the space c00(G) viewed inside BG is the space
of functionals which are normal on L∞(bG) and whose density matrix is a finite rank operator.
The closure of this space of functionals inside the space of all functionals on Cu(bG) is the space of
all functionals which are normal on L∞(bG), i.e. the space AG.
Proof of Theorem 6.2. The C∗-algebra C0(G) = c0(G) is a c0-direct sum of matrix algebras:
(cid:3)
c0(G) = Mα∈R
Mnα.
By [7, Proposition II.8.2.4] any ideal in c0(G) is of the form
Mα∈R0
Mnα.
for some R0 ⊂ R (the direct sum is still in c0-sense). Now if π : c0(G) → Cu
0(H) is the epimorphism
corresponding to the embedding of H into G and R0 corresponds to the kernel of π, we see that
Cu
0(H) is the c0-direct sum
Cu
0(G) = Mα∈R\R0
Mnα.
For the same reason the algebra C0(H) which is a (potentially proper) quotient of Cu
0(H) is also
a c0-direct sum of matrix algebras. In particular C0(H) is an ideal in C0(G)∗∗. By [34, Theorem
4.4] H is a discrete quantum group. In particular Cu
0(H) = C0(H) = c0(H).
Consider the adjoint of the map bπ : Cu(bH) → Cu(bG), i.e.
bπ∗ : BG −→ BH.
We now note that bπ∗ maps c00(G) into c00(H). Indeed, bπ∗ is the operation of pre-composing a
functional with bπ. In particular, on the level of ℓ∞(G), where BG is embedded, we have
bπ∗(cid:0)(η ⊗ id)(VVH)(cid:1) =(cid:0)[η◦bπ] ⊗ id(cid:1)(VVH)
= (η ⊗ id)(cid:0)(bπ ⊗ id)(VVH)(cid:1)
= (η ⊗ id)(cid:0)(id ⊗ π)(VVG)(cid:1)
G)(cid:1),
= ¯π(cid:0)(η ⊗ id)(VV
where ¯π is the canonical extension of π to M(cid:0)c0(G)(cid:1) = ℓ∞(G). Also bπ∗ is a contraction for the
norms on BG and BH (as an adjoint map of a contraction bπ : Cu(bH) → Cu(bG)). It follows from
Theorem 6.3 that bπ∗ restricts to a contraction
T : AG −→ AH
with dense range; this completes the proof by applying Theorem 3.7.
(cid:3)
Acknowledgments
The authors wish to thank Nico Spronk for helpful suggestions and pointing out several im-
portant references. The second and third authors were supported by the National Science Centre
(NCN) grant no. UMO-2011/01/B/ST1/06474. The fourth author was supported by National Sci-
ence Centre (NCN) grant no. 2011/01/B/ST1/05011. The third author's visit to the University
of Leeds was supported by EPSRC grant EP/I002316/1.
References
[1] G. Arsac: Sur l'espace de Banach engendr´e par les coefficients d'une repr´esentation unitare. Pub. D`ep.
Math. Lyon 13 (1976), 1 -- 101.
[2] S. Baaj & G. Skandalis: Unitaires multiplicatifs et dualit´e pour les produits crois´es de C∗-alg`ebres.
Ann. Scient. ´Ec. Norm. Sup., 4e s´erie, t. 26 (1993), 425 -- 488.
[3] T. Banica & J. Bichon: Quantum groups acting on 4 points. J. Reine Angew. Math. 626 (2009), 75 -- 114.
[4] T. Banica, A. Skalski & P.M. So ltan: Noncommutative homogeneous spaces:
the matrix case.
J. Geom. Phys. 62 (2012), 1451 -- 1466.
CLOSED QUANTUM SUBGROUPS OF LOCALLY COMPACT QUANTUM GROUPS
23
[5] E. Bedos, G.J. Murphy & L. Tuset: Co-amenability for compact quantum groups. J. Geom. Phys. 40
(2001), 130 -- 153.
[6] M.E.B. Bekka, A.T. Lau & G. Schlichting: On invariant subalgebras of the Fourier-Stieltjes algebra of a
locally compact group. Math. Ann. 294 (1992), 513 -- 522.
[7] B. Blackadar: Operator algebras. Theory of C∗-algebras and von Neumann algebras. Encyclopedia of
Mathematical Sciences, Vol. 122, Springer-Verlag 2006.
[8] N. Bourbaki: Elements of Mathematics. General Topology. Chapters 1-4. Springer-Verlag 1989.
[9] M. Daws: Multipliers, Self-Induced and Dual Banach Algebras. Dissertationes Mathematicae (Rozprawy
Matematyczne) 470 (2010).
[10] J. Delaporte & A. Derighetti: On Herz' extension theorem. Boll. Un. Mat. Ital. A 6 (1992), 245-247.
[11] A. Derighetti: Relations entre les convoluteurs d'un groupe localement compact et ceux d'un sous-groupe
ferm´e. Bull. Sci. Math. 106 (1982), 69-84.
[12] P. Eymard: L'alg`ebre de Fourier d'un groupe localement compact. Bull. Soc. Math. France 92 (1964),
181 -- 286.
[13] G.B. Folland: A course in abstract harmonic analysis. CRC Press 1995.
[14] U. Franz, A. Skalski & R. Tomatsu: Idempotent states on the compact quantum groups and their classi-
q(3). To appear in Journal of Noncommutative Geometry, available at
q(2) and SO
fication on U
arXiv:0903.2363v2 [math.OA].
q(2), SU
[15] C. Herz: Le rapport entre l'alg`ebre A
p d'un groupe et d'un sous-groupe. C. R. Acad. Sci. Paris S´er. A-B
271 (1970), 244 -- 246.
[16] C. Herz: Harmonic synthesis for subgroups. Ann. Inst. Fourier (Grenoble) 23 (1973), 91 -- 123.
[17] Z. Hu, M. Neufang & Z.-J. Ruan: Completely bounded multipliers over locally compact quantum groups.
Proc. Lond. Math. Soc. (3) 103 (2011), 1 -- 39.
[18] M. Ilie & R. Stokke: Weak∗-continuous homomorphisms of Fourier-Stieltjes algebras, Math. Proc. Cam-
bridge Philos. Soc. 145 (2008), 107 -- 120.
[19] M. Kalantar & M. Neufang: From Quantum Groups to Groups. Available at arXiv:1110.5129v1
[math.OA].
[20] P. Kasprzak: Rieffel deformation of homogeneous spaces. J. Funct. Anal. 260 (2011), 146 -- 163.
[21] E.T. Kehlet: Cross sections for quotient maps of locally compact groups.Math. Scand. 55 (1984), 152 -- 160.
[22] J. Kustermans: Locally compact quantum groups in the universal setting. Int. J. Math. 12 (2001) 289 -- 338.
[23] J. Kustermans & S. Vaes: Locally compact quantum groups. Ann. Scient. ´Ec. Norm. Sup. 4e s´erie, t. 33
(2000), 837 -- 934.
[24] C.E. Lance: Hilbert C∗-modules: a toolkit for operator algebraists. London Mathematical Society Lecture
Note Series 210, Cambridge University Press 1995.
[25] T. Masuda, Y. Nakagami & S.L. Woronowicz: A C∗-algebraic framework for the quantum groups.
Int. J. Math. 14 (2003), 903 -- 1001.
[26] G. Mackey: Induced representations of locally compact groups I. Ann. Math. 55 (1952), 101 -- 139.
[27] R. Meyer, S. Roy & S.L. Woronowicz: Homomorphisms of quantum groups, Munster J. Math. 4 (2011),
101 -- 124.
[28] T.W. Palmer: Banach algebras and the general theory of ∗-algebras, Volume 1, Cambridge University
Press 1994.
[29] G.K. Pedersen: C∗-algebras and their automorphism groups. Academic Press 1979.
[30] C. Pinzari: Embedding ergodic actions of compact quantum groups on C∗-algebras into quotient spaces.
Int. J. Math. 18 (2007), 137 -- 164.
[31] P. Podle´s: Przestrzenie kwantowe i ich grupy symetrii (Quantum spaces and their symmetry groups).
Ph.D. Thesis, Department of Mathematical Methods in Physics, Faculty of Physics, University of Warsaw
(1989) (in Polish).
[32] P. Podle´s: Symmetries of quantum spaces. Subgroups and quotient spaces of quantum SU(2) and SO(3)
groups. Commun. Math. Phys. 170 (1995), 1 -- 20.
[33] P. Podle´s & S.L. Woronowicz: Quantum deformation of Lorentz group. Comm. Math. Phys. 130 (1990),
381 -- 431.
[34] V. Runde: Characterizations of compact and discrete quantum groups through second duals. J. Op. Th.
60 (2008), 415 -- 428.
[35] P. Salmi: Compact quantum subgroups and left invariant C∗-subalgebras of locally compact quantum
groups. J. Funct. Anal. 261 (2011), 1 -- 24.
[36] P. Salmi & A. Skalski: Idempotent states on locally compact quantum groups. To appear in Quarterly
Journal of Mathematics, available at arXiv:1102.2051v2 [math.OA].
[37] H.-J. Schneider: Principal homogeneous spaces for arbitrary Hopf algebras. Isr. J. Math. 72 (1990),
167 -- 195.
[38] P.M. So ltan: Examples of non-compact quantum group actions. J. Math. Anal. Appl. 372 (2010), 224 -- 236.
[39] P.M. So ltan & S.L. Woronowicz: A remark on manageable multiplicative unitaries. Lett. Math. Phys. 57
(2001), 239 -- 252.
[40] P.M. So ltan & S.L. Woronowicz: From multiplicative unitaries to quantum groups II. J. Funct. Anal. 252
(2007), 42 -- 67.
24
MATTHEW DAWS, PAWE L KASPRZAK, ADAM SKALSKI, AND PIOTR M. SO LTAN
[41] M. Takesaki: Theory of Operator Algebras I. Springer-Verlag 1979.
[42] M. Takesaki & N. Tatsuuma: Duality and subgroups. Ann. Math. 93 (1971), 344 -- 364.
[43] S. Vaes: A new approach to quantum and imprimitivity results. J. Funct. Anal. 229 (2005), 317 -- 374.
[44] S. Vaes & L. Vainerman: On low-dimensional locally compact quantum groups. In Locally compact quan-
tum groups and groupoids, IRMA Lect. Math. Theor. Phys. 2, de Gruyter, Berlin, (2003), pp. 127 -- 187.
[45] S. Vaes & L. Vainerman: Extensions of locally compact quantum groups and the bicrossed product
construction. Adv. Math. 175 (2003), 1 -- 101.
[46] N.E. Wegge-Olsen: K-theory and C∗-algebras: a friendly approach. Oxford University Press 1993.
[47] S.L. Woronowicz: Pseudogroups, pseudospaces and Pontryagin duality. Proceedings of the International
Conference on Mathematical Physics, Lausanne 1979 Lecture Notes in Physics, 116, pp. 407 -- 412.
[48] S.L. Woronowicz: Unbounded elements affiliated with C∗-algebras and non-compact quantum groups.
Commun. Math. Phys. 136 (1991), 399 -- 432.
[49] S.L. Woronowicz: C∗-algebras generated by unbounded elements. Rev. Math. Phys. 7, (1995), 481 -- 521.
[50] S.L. Woronowicz: From multiplicative unitaries to quantum groups. Int. J. Math. 7, (1996), 127 -- 149.
[51] S.L. Woronowicz: Compact quantum groups. In: Sym´etries quantiques, les Houches, Session LXIV 1995,
Elsevier 1998, pp. 845 -- 884.
[52] C. Zwarich: Von Neumann algebras for abstract harmonic analysis. Thesis at University of Waterloo,
available at http://hdl.handle.net/10012/3920.
School of Mathematics, University of Leeds, Leeds LS2 9JT, United Kingdom
E-mail address: [email protected]
Department of Mathematical Methods in Physics, Faculty of Physics, University of Warsaw, Poland
and Institute of Mathematics of the Polish Academy of Sciences, ul. ´Sniadeckich 8, 00 -- 956 Warszawa,
Poland
E-mail address: [email protected]
Institute of Mathematics of the Polish Academy of Sciences, ul. ´Sniadeckich 8, 00 -- 956 Warszawa,
Poland
E-mail address: [email protected]
Department of Mathematical Methods in Physics, Faculty of Physics, University of Warsaw, Poland
E-mail address: [email protected]
|
1806.03189 | 1 | 1806 | 2018-06-07T14:11:51 | Local Lie derivations on von Neumann algebras and algebras of locally measurable operators | [
"math.OA"
] | Let $\mathcal{A}$ be a unital associative algebra and $\mathcal{M}$ be an $\mathcal{A}$-bimodule. A linear mapping $\varphi$ from $\mathcal{A}$ into an $\mathcal{A}$-bimodule $\mathcal{M}$ is called a Lie derivation if $\varphi[A,B]=[\varphi(A),B]+[A,\varphi(B)]$ for each $A,B$ in $\mathcal{A}$, and $\varphi$ is called a \emph{local Lie derivation} if for every $A$ in $\mathcal{A}$, there exists a Lie derivation $\varphi_{A}$ (depending on $A$) from $\mathcal{A}$ into $\mathcal{M}$ such that $\varphi(A)=\varphi_{A}(A)$. In this paper, we prove that every local Lie derivation on von Neumann algebras is a Lie derivation; and we show that if $\mathcal M$ is a type I von Neumann algebra with atomic lattice of projections, then every local Lie derivation on $LS(\mathcal M)$ is a Lie derivation. | math.OA | math | Local Lie derivations on von Neumann algebras and
algebras of locally measurable operators
Jun He1 and Guangyu An2∗
1Department of Mathematics, Anhui Polytechnic University
Wuhu 241000, China
2Department of Mathematics, Shaanxi University of Science and Technology
8
1
0
2
n
u
J
7
]
.
A
O
h
t
a
m
[
1
v
9
8
1
3
0
.
6
0
8
1
:
v
i
X
r
a
Xi'an 710021, China
Abstract
Let A be a unital associative algebra and M be an A-bimodule. A linear mapping
ϕ from A into an A-bimodule M is called a Lie derivation if ϕ[A, B] = [ϕ(A), B] +
[A, ϕ(B)] for each A, B in A, and ϕ is called a local Lie derivation if for every A
in A, there exists a Lie derivation ϕA (depending on A) from A into M such that
ϕ(A) = ϕA(A). In this paper, we prove that every local Lie derivation on von Neumann
algebras is a Lie derivation; and we show that if M is a type I von Neumann algebra
with atomic lattice of projections, then every local Lie derivation on LS(M) is a Lie
derivation.
Keywords: Lie derivation, local Lie derivation, von Neumann algebra, locally
measurable operator.
Mathematics Subject Classification(2010): 46L57; 47L35; 46L50
1
Introduction
Let A be a unital associative algebra over the complex field C and M be an A-
bimodule. An linear mapping δ from A into M is called a derivation if δ(AB) =
δ(A)B + Aδ(B) for each A and B in A.
In particular, a derivation δM defined by
∗Corresponding author. E-mail address: [email protected]
1
δM (A) = M A − AM for every A in A is called an inner derivation, where M is a fixed
element in M.
In [31], S. Sakai proves that every derivation on von Neumann algebras is an inner
derivation. In [13], E. Christensen shows that every derivation on nest algebras on a
Hilbert space H is an inner derivation. For more information on derivations and inner
derivations, we refer to [14, 15, 19].
In [23, 25], R. Kadison and D. Larson introduce the concept of local derivations. A
linear mapping δ from A into M is called a local derivation if for every A in A, there
exists a derivation δA (depending on A) from A into M such that δ(A) = δA(A).
In [23], R. Kadison proves that every continuous local derivation from a von Neu-
mann algebra into its dual Banach module is a derivation. In [25], D. Larson and A.
Sourour prove that if X is a Banach space, then every local derivation on B(X) is a
derivation. In [21], B. Jonson shows that every local derivation from a C ∗-algebra into
its Banach bimodule is a derivation. In [17, 18], D. Hadwin and J. Li characterize lo-
cal derivations on non self-adjoint operator algebras such as nest algebras and CDCSL
algebras.
A linear mapping ϕ from A into an A-bimodule M is called a Lie derivation if
ϕ[A, B] = [ϕ(A), B] + [A, ϕ(B)] for each A and B in A, where [A, B] = AB − BA is the
usual Lie product. A Lie derivation ϕ is said to be standard if it can be decomposed
as ϕ = δ + τ , where δ is a derivation from A into M and τ is a linear mapping from A
into Z(A, M) such that τ [A, B] = 0 for each A and B in A, where Z(A, M) = {M ∈
M : AM = M A for every A in A}.
In [22], B. Johnson proves that every continuous Lie derivation from a C ∗-algebra
into its Banach bimodule is standard. In [28], M. Mathieu and A. Villena prove that
every Lie derivation on a C ∗-algebra is standard. In [12], W. Cheung characterizes Lie
derivations on triangular algebras. In [27], F. Lu proves that every Lie derivation on
a completely distributed commutative subspace lattice algebra is standard. In [4], D.
Benkovic proves that every Lie derivation on matrix algebra Mn(A) is standard, where
n ≥ 2 and A is a 2-torsion free unital algebra.
Similar to local derivations, In [10], L. Chen, F. Lu and T. Wang introduce the
concept of local Lie derivations. A linear mapping ϕ from A into M is called a local
Lie derivation if for every A in A, there exists a Lie derivation ϕA (depending on A)
from A into M such that ϕ(A) = ϕA(A).
In [10], L. Chen, F. Lu and T. Wang prove that every local Lie derivation on B(X) is
a Lie derivation, where X is a Banach space of dimension exceeding 2. In [11], L. Chen
and F. Lu prove that every local Lie derivation on nest algebras is a Lie derivation. In
[26], D. Liu and J. Zhang prove that under certain conditions, every local Lie derivation
on triangular algebras is a Lie derivation. In [20], J. He, J. Li, G. An and W. Huang
prove that every local Lie derivation on some algebras such as finite von Neumann
2
algebras, nest algebras, Jiang-Su algebra and UHF algebras is a Lie derivation.
Compare with the characterizations of derivations on Banach algebras, investigation
of derivations on unbounded operator algebras begin much later.
In [32], I. Segal studies the theory of noncommutative integration, and introduces
various classes of non-trivial ∗-algebras of unbounded operators.
In this paper, we
mainly consider the ∗-algebra S(M) of all measurable operators and the ∗-algebra
LS(M) of all locally measurable operators affiliated with a von Neumann algebra M.
In [32], I. Segal shows that the algebraic and topological properties of the measurable
operators algebra S(M) are similar to the von Neumann algebra M. If M is a com-
mutative von Neumann algebra, then M is ∗-isomorphic to the algebra L∞(Ω, Σ, µ)
of all essentially bounded measurable complex functions on a measure space (Ω, Σ, µ);
and S(M) is ∗-isomorphic to the algebra L0(Ω, Σ, µ) of all measurable almost every-
where finite complex-valued functions on (Ω, Σ, µ).
In [5], A. Ber, V. Chilin and F.
Sukochev show that there exists a derivation on L0(0, 1) is not an inner derivation, and
the derivation is discontinuous in the measure topology. This result means that the
properties of derivations on S(M) are different from the derivations on M.
In [1, 2], Albeverio, Ayupov and Kudaybergenov study the properties of derivations
on various classes of measurable algebras. If M is a type I von Neumann algebra, in [1],
the authors prove that every derivation on LS(M) is an inner derivation if and only if it
is Z(M) linear; in [2], the authors give the decomposition form of derivations on S(M)
and LS(M); they also prove that if M is a type I∞ von Neumann algebra, then every
derivation on S(M) or LS(M) is an inner derivation. If M is a properly infinite von
Neumann algebra, in [6], A. Ber, V. Chilin and F. Sukochev prove that every derivation
on LS(M) is continuous with respect to the local measure topology t(M); and in [7],
the authors show that every derivation on LS(M) is an inner derivation.
In [3], S.
Albeverio and S. Ayupov give a characterization of local derivations on S(M), where
M is an abelian von Neumann algebra. In [16], D. Hadwin and J. Li prove that if M is
a von Neumann algebra without abelian direct summands, then every local derivation
on LS(M) or S(M) is a derivation. In [9], V. Chilin and I. Juraev show that every Lie
derivation on LS(M) or S(M) is standard.
This paper is organized as follows. In Section 2, we recall the definitions of algebras
of measurable operators and local measurable operators.
In Section 3, we generalize the Corollary 3.2 in [20] and prove that every local Lie
derivation on von Neumann algebras is a Lie derivation.
In Section 4, we prove that if M is a type I von Neumann algebra with an atomic
lattice of projections, then every local Lie derivation on LS(M) is a Lie derivation.
3
2 Preliminaries
Let H be a complex Hilbert space and B(H) be the algebra of all bounded linear
operators on H. Suppose that M is a von Neumann algebra on H and Z(M) = M∩M′
is the center of M, where
M′ = {a ∈ B(H) : ab = ba for every b in M}.
Denote by P(M) = {p ∈ M : p = p∗ = p2} the lattice of all projections in M and by
Pf in(M) the set of all finite projections in M. For each p and q in P(M), if we define
the inclusion relation p ⊂ q by p ≤ q, then P(M) is a complete lattice. Suppose that
{pl}l∈λ is a family of projections in M, we denote
pl = [
sup
l∈λ
l∈λ
plH and inf
l∈λ
pl = \
plH.
l∈λ
If {pl}l∈λ is an orthogonal family of projections in M, then we have that
pl = X
pl.
sup
l∈λ
l∈λ
Let x be a closed densely defined linear operator on H with the domain D(x), where
D(x) is a linear subspace of H. x is said to be affiliated with M, denote by xηM, if
u∗xu = x for every unitary element u in M′.
A linear operator affiliated with M is said to be measurable with respect to M,
if there exists a sequence {pn}∞
n=1 ⊂ P(M) such that pn ↑ 1, pn(H) ⊂ D(x) and
p⊥
n = 1 − pn ∈ Pf in(M) for every n ∈ N, where N is the set of all natural numbers.
Denote by S(M) the set of all measurable operators affiliated with the von Neumann
algebra M.
A linear operator affiliated with M is said to be locally measurable with respect to
M, if there exists a sequence {zn}∞
n=1 ⊂ P(Z(M)) such that zn ↑ 1 and znx ∈ S(M)
for every n ∈ N. Denote by LS(M) the set of all locally measurable operators affiliated
with the von Neumann algebra M.
In [29], Muratov and Chilin prove that S(M) and LS(M) are both unital ∗-algebras
and M ⊂ S(M) ⊂ LS(M); the authors also show that if M is a finite von Neumann
algebra or dim(Z(M)) < ∞, then S(M) = LS(M); if M is a type III von Neumann
algebra and dim(Z(M)) = ∞, then S(M) = M and LS(M) 6= M.
3 Local Lie derivations on von Neumann alge-
bras
4
In this section, we consider local Lie derivations on von Neumann algebras. To
prove our main theorem, we need the following lemma.
Lemma 3.1. Let A1 and A2 be two unital algebras and A = A1 L A2. If the following
five conditions hold:
(1) each Lie derivation on A is standard;
(2) each derivation on A is inner;
(3) each local derivation on A is a derivation;
(4) Z(A1) ∩ [A1, A1] = {0};
(5) A2 = [A2, A2],
then every local Lie derivation on A is a Lie derivation.
Proof. Denote the units of A, A1 and A2 by I, P and Q, respectively. For each A in
A, we have that A = P A + QA = A1 + A2, where Ai ∈ Ai, i = 1, 2.
In the following we suppose that ϕ is a local Lie derivation on A.
By the definition of local Lie derivation, we know that for every A1 in A1, there
exists a Lie derivation ϕA1 on A such that ϕ(A1) = ϕA1(A1). Since ϕA1 is standard
and each derivation on A is inner, we can obtain that
ϕ(A1) = ϕA1(A1) = δA1(A1) + τA1(A1) = [A1, TA1 ] + P τA1(A1) + QτA1(A1),
where δA1 is a derivation on A, TA1 is an element in A, and τA1 is a linear mapping
from A into Z(A) such that τA1([A, A]) = 0.
It means that ϕ has a decomposition at A1. Next we show that the decomposition
at A1 is unique. Assume there is another decomposition at A1, that is
ϕ(A1) = ϕ
A1(A1) = δ
A1(A1) + τ
A1(A1) = [A1, T
A1 ] + P τ
′
′
′
′
′
A1(A1) + Qτ
′
A1(A1),
′
A1 is a derivation on A, T
where δ
from A into Z(A) such that τ
′
A1([A, A]) = 0.
′
A1 is an element in A and τ
′
A1 is a linear mapping
Then we have that
[A1, TA1] + P τA1(A1) + QτA1(A1) = [A1, T
′
A1] + P τ
′
A1(A1) + Qτ
′
A1(A1).
Thus
[A1, TA1] − [A1, T
A1] = P τ
′
′
A1(A1) − P τA1(A1) + Qτ
′
A1(A1) − QτA1(A1).
Since [A1, TA1] − [A1, T
A1(A1)−QτA1(A1) belongs to A2, we have that Qτ
A1(A1) − P τA1(A1) ∈ A1 belong to A1, and
A1(A1)−QτA1(A1) = 0. Moreover,
A1] and P τ
′
′
′
′
Qτ
we can obtain that
[A1, TA1] − [A1, T
A1] = [A1, P TA1] − [A1, P T
A1] ∈ [A1, A1],
′
′
5
and
P τ
′
A1(A1) − P τA1(A1) ∈ Z(A1).
By condition (4), it follows that [A1, TA1] − [A1, T
It implies that δA1(A1) = δ
unique.
′
A1(A1) and τA1(A1) = τ
′
′
A1] = P τ
A1(A1) − P τA1(A1) = 0.
A1(A1). Hence the decomposition is
′
Now we have ϕA1 = δ1 + τ1, where δ1 is a mapping from A1 into A1 such that
δ1(A1) = [A1, SA1] for some element SA1 in A1, and τ1 is a mapping from A1 into Z(A)
such that τ1([A1, A1]) = 0.
Next we prove that δ1 and τ1 are linear mappings. For each A1 and B1 in A1, we
have that
and
ϕ(A1) = δ1(A1) + τ1(A1) = [A1, SA1] + τ1(A1),
ϕ(B1) = δ1(B1) + τ1(B1) = [B1, SB1 ] + τ1(B1),
ϕ(A1 + B1) = δ1(A1 + B1) + τ1(A1 + B1) = [A1 + B1, SA1+B1] + τ1(A1 + B1).
Since ϕ is additive, through a discussion similar to that before, it implies that
[A1 + B1, SA1+B1] = [A1, SA1] + [B1, SB1]
and
τ1(A1 + B1) = τ1(A1) + τ1(B1).
It means that δ1 and τ1 are additive mappings. Using the same technique, we can prove
that δ1 and τ1 are homogeneous. Hence δ1 and τ1 are linear mappings.
For every A2 in A2, we have that
ϕ(A2) = ϕA2(A2) = δA2(A2) + τA2(A2) = [A2, TA2] + τA2(A2),
where δA2 is a derivation on A, TA2 is an element in A and τA2 is a linear mapping from
A into Z(A) such that τA2([A, A]) = 0. By condition (5), we have that τA2(A2) = 0.
Thus ϕ(A2) = [A2, TA2] = [A2, QTA2].
Let ϕA2 = δ2. Then we have δ2(A2) = [A2, SA2] for some element SA2 in A2. And
obviously, δ2 is linear.
Define two linear mappings as follows:
δ(A) = δ1(A1) + δ2(A2), τ (A) = τ1(A1),
for all A = A1 + A2 ∈ A. By the previous discussion, τ is a linear mapping from A into
Z(A) such that τ ([A, A]) = 0. In addition,
δ(A) = δ1(A1) + δ2(A2) = [A1, SA1] + [A2, SA2] = [A1 + A2, SA1 + SA2] = [A, SA1 + SA2].
6
It means that δ is a local derivation. By condition (3), δ is a derivation. Notice that
ϕ(A) = ϕ(A1) + ϕ(A2) = δ1(A1) + τ1(A1) + δ2(A2) = δ(A) + τ (A).
Hence ϕ is a standard Lie derivation.
By Lemma 3.1, we have the following result.
Theorem 3.2. Every local Lie derivation on a von Neumann algebra is a Lie deriva-
tion.
Proof. Let A be a von Neumann algebra. It is well known that A = A1 L A2, where
A1 is a finite von Neumann algebra, and A2 is a proper infinite von Neumann algebra.
By [28, Theorem 1.1], we know that every Lie derivation on A is standard, by [31,
Theorem 1], we know that every derivation on A is inner, and by [21, Theorem 5.3],
we know that every local derivation on A is a derivation. Since A2 is a proper infinite
von Neumann algebra, we known that A2 = [A2, A2](see in [33]).
Hence it is sufficient to prove that Z(A1) ∩ [A1, A1] = {0}. Since A1 is finite and
by [24, Theorem 8.2.8], it follows that there is a center-valued trace τ on A1 such that
τ (Z) = Z for every Z in Z(A1) and τ ([A, B]) = 0 for each A and B in A1. Suppose
that A ∈ Z(A1) ∩ [A1, A1], then we have that τ (A) = A and τ (A) = 0. it implies that
A = 0.
By Lemma 3.1, we know that every local Lie derivation on a von Neumann algebra
is a Lie derivation.
4 Local Lie derivations on algebras of locally
measurable operators
In this section, we mainly consider local Lie derivations on algebras of all locally
measurable operators affiliated with a type I von Neumann algebra. To prove the main
result, we need the following lemmas.
Lemma 4.1. Suppose that A is a commutative unital algebra and J = Mn(A). Then
Z(J ) ∩ [J , J ] = {0}
Proof. Let {ei,j}n
A in J , we have that A = Pn
i,j=1 aij eij, where aij ∈ A.
i,j=1 be the system of matrix units in Mn(A). Then for every element
Define a linear mapping τ from J into A by τ (A) = Pn
i=1 aii for every A =
i,j=1 aijeij ∈ J . Since A is commutative, it is not difficult to verify that τ ([A, B]) = 0
Pn
for each A and B in J .
7
It should be noticed that Z(J ) = {A : A = Pn
i=1 aeii, a ∈ A}. Suppose that
i=1 aeii is an element in Z(J ) ∩ [J , J ], then by the definition of τ , we have that
A = Pn
τ (A) = na and τ (A) = 0. It implies that A = 0.
Lemma 4.2. Suppose that A = Qi∈Λ Ai. If Z(Ai) ∩ [Ai, Ai] = {0} for every i ∈ Λ,
then we have that Z(A) ∩ [A, A] = {0}.
Proof. Let A = {ai}i∈Λ be an element in Z(A) ∩ [A, A]. Then for every i ∈ Λ, we have
that ai ∈ Z(Ai) ∩ [Ai, Ai]. By assumption, it follows that ai = 0. Hence A = 0.
Lemma 4.3. Suppose that M is a type I∞ von Neumann algebra. Then LS(M) =
[LS(M), LS(M)].
Proof. By [30], we know that for every x in LS(M), there exists a sequence {zn}
of mutually orthogonal central projections in M with P∞
n=1 zn = I, such that x =
P∞
n=1 znx, and znx ∈ M for every n ∈ N. Since M is a proper infinite von Neumann
algebra, it is well known that M = [M, M]. Thus we have that znx = Pk
i , bn
i ],
where an
i ∈ M for each n and i.
i=1[an
i , bn
Set si = P∞
n=1 znan
i and ti = P∞
n=1 znbn
i . By the definition of locally measurable
operators, it is easy to show that si and ti are two elements in LS(M).
Since that {zn} are mutually orthogonal central projections, we can obtain that
∞
[si, ti] = [
X
znan
i ,
n=1
∞
X
n=1
znbn
i ] =
∞
X
n=1
zn[an
i , bn
i ],
moreover, we have that
k
X
[si, ti] =
i=1
k
∞
X
X
i=1
n=1
zn[an
i , bn
i ] =
∞
X
n=1
k
zn(
X
i=1
[an
i , bn
i ]) =
∞
X
n=1
znx = x.
It follows that x ∈ [LS(M), LS(M)].
In the following we show the main result of this section.
Theorem 4.4. Suppose that M is a type I von Neumann algebra with an atomic
lattice of projections. Then every local Lie derivation from LS(M) into itself is a Lie
derivation.
Proof. By [24, Theorem 6.5.2], we know that M = M1 L M2, where M1 is a type
If inite von Neumann algebra and M2 is a type I∞ von Neumann algebra. Hence by [2,
Proposition 1.1], we have that LS(M) ∼= LS(M1) L LS(M2).
In the following we will verify the conditions (1) to (5) in Lemma 3.1 one by one.
By [9, Theorem 1], we know that every Lie derivation on LS(M) is standard; by [2,
Corollary 5,12], we know that every derivation on LS(M) is inner for a von Neumann
algebra with atomic lattice of projections.
8
It is proved in [17] that every local derivation on LS(M) is a derivation for a von
Neumann algebra without abelian direct summands. While for an abelian von Neumann
algebra with atomic lattice of projections, by [3, Theorem 3.8] we know that every local
derivation on LS(M) is a derivation. Associated the two results, we can obtain each
local derivation on LS(M) is a derivation for a von Neumann algebra with atomic
lattice of projections.
Since M1 is a type If inite von Neumann algebra, we know that M1 = L∞
n=1 An,
where each An is a homogenous type In von Neumann algebra. Hence LS(M1) ∼=
Q∞
n=1 LS(An). Since An is a homogenous type In von Neumann algebra, by [2] we
know that LS(An) ∼= Mn(Z(LS(An))). By Lemmas 4.1 and 4.2, we know that the
condition (4) in Lemma 3.1 holds. And by Lemma 4.3, the condition (5) in Lemma 3.1
holds.
References
[1] S. Albeverio, S. Ayupov, K. Kudaybergenov. Derivations on the algebra of measur-
able operators affiliated with a type I von Neumann algebra. Siberian Adv. Math.,
2008, 18: 86-94.
[2] S. Albeverio, S. Ayupov, K. Kudaybergenov. Structure of derivations on various
algebras of measurable operators for type I von Neumann algebras. J. Func. Anal.,
2009, 256: 2917-2943.
[3] S. Albeverio, S. Ayupov, K. Kudaybergenov, B. Nurjanov. Local derivations on
algebras of measurable operators. Comm. In Contem. Math., 2011, 13: 643 -- 657.
[4] D. Benkovic. Lie triple derivations of unital algebras with idempotents. Linear
Multilinear Algebra, 2015, 63: 141 -- 165.
[5] A. Ber, V. Chilin, F. Sukochev. Non-trivial derivation on commutative regular
algebras. Extracta Math., 2006, 21: 107 -- 147.
[6] A. Ber, V. Chilin, F. Sukochev. Continuity of derivations of algebras of locally
measurable operators. Integr. Equ. Oper. Theory, 2013, 75: 527 -- 557.
[7] A. Ber, V. Chilin, F. Sukochev. Continuous derivations on algebras of locally
measurable operators are inner. Proc. London Math. Soc., 2014, 109: 65 -- 89.
[8] M. Bresar, E. Kissin, S. Shulman, Lie ideals: from pure algebra to C*-algebras, J.
reine angew. Math., 2008, 623: 73 -- 121.
[9] V. Chilin, I. Juraev. Lie derivations on the algebras of locally measurable operators.
2016, arXiv: 1608. 03996v1.
[10] L. Chen, F. Lu, T. Wang. Local and 2-local Lie derivations of operator algebras
on Banach spaces. Integr. Equ. Oper. Theory, 2013, 77: 109 -- 121.
9
[11] L. Chen, F. Lu. Local Lie derivations of nest algebras. Linear Algebra Appl., 2015,
475: 62 -- 72.
[12] W. Cheung. Lie derivations of triangular algebra. Linear Multilinear Algebra, 2003,
512: 299 -- 310.
[13] E. Christensen. Derivations of nest algebras. Math. Ann., 1977, 229: 155 -- 161.
[14] H. Du, J. Zhang. Derivations on nest-subalgebras of von Neumann algebras. Chin.
Ann. Math., 1996, A 17: 467 -- 474.
[15] H. Du, J. Zhang. Derivations on nest-subalgebras of von Neumann algebras II.
Acta. Math., 1997, A 40: 357 -- 362.
[16] D. Hadwin, J. Li, Q. Li, X. Ma. Local derivations on rings containing a von Neu-
mann algebra and a question of Kadison. 2013, arXiv:1311.0030v1.
[17] D. Hadwin, J. Li. Local derivations and local automorphisms on some algebras. J.
Oper. Theory, 2008, 60: 29 -- 44.
[18] D. Hadwin, J. Li. Local derivations and local automorphisms. J. Math. Anal. Appl.,
2004, 290: 702 -- 714.
[19] J. He, J. Li, D. Zhao. Derivations, Local and 2-Local derivations on some algebras
of operators on Hilbert C*-bodules. Mediterr. J. Math., 2017, 14: article 230.
[20] J. He, J. Li, G. An, W. Huang. Characterizations of 2-local derivations and local
Lie derivations on some algebras. Sib. Math. J., 2017, accepted.
[21] B. Johnson. Local derivations on C∗-algebras are derivations. Trans. Amer. Math.
Soc., 2001, 353: 313 -- 325.
[22] B. Johnson. Symmetric amenability and the nonexistence of Lie and Jordan deriva-
tions. Math. Proc. Cambd. Philos. Soc., 1996, 120: 455 -- 473.
[23] R. Kadison. Local derivations. J. Algebra, 1990, 130: 494 -- 509.
[24] R. Kadison, J. Ringrose, Fundamentals of the Theory of Operator Algebras, I,
Pure Appl. Math. 100, Academic Press, New York, 1983.
[25] D. Larson, A. Sourour. Local derivations and local automorphisms. Proc. Sympos.
Pure Math., 1990, 51: 187 -- 194.
[26] D. Liu, J. Zhang. Local Lie derivations on certain operator algebras. Ann. Funct.
Anal., 2017: 270 -- 280.
[27] F. Lu. Lie derivation of certain CSL algebras. Israel J. Math., 2006, 155: 149 -- 156.
[28] M. Mathieu, A. Villena. The structure of Lie derivations on C ∗-algebras. J. Funct.
Anal., 2003, 202: 504 -- 525.
[29] M. Muratov, V. Chilin. Algebras of measurable and locally measurable operators.
Kyiv, Pratse In-ty matematiki NAN ukraini., 2007, 69: 390 pp, (Russian).
10
[30] M. Muratov, V. Chilin. Central extensions of *-algebras of measurable operators.
Reports of the National Academy of Science of Ukraine, 2009, 7: 24 -- 28. (Russian).
[31] S. Sakai. Derivations of W ∗-algebras. Ann. Math., 1966, 83: 273 -- 279.
[32] I. Segal. A non-commutative extension of abstract integration. Ann, Math., 1953,
57: 401 -- 457.
[33] H. Sunouchi. Infinite Lie rings. Tohoku Math. J., 1956, 8: 291 -- 307.
11
|
1711.09466 | 1 | 1711 | 2017-11-26T21:42:46 | Measures of noncompactness on the standard Hilbert $C^*$-module | [
"math.OA"
] | We define a measure of noncompactness $\lambda$ on the standard Hilbert $C^*$-module $l^2(\mathcal A)$ over a unital $C^*$-algebra, such that $\lambda(E)=0$ if and only if $E$ is $\mathcal A$-precompact (i.e.\ it is $\varepsilon$-close to a finitely generated projective submodule for any $\varepsilon>0$) and derive its properties. Further, we consider the known, Kuratowski, Hausdorff and Istr\u{a}\c{t}escu measure of noncomapctnes on $l^2(\mathcal A)$ regarded as a locally convex space with respect to a suitable topology, and obtain their properties as well as some relationship between them and introduced measure of noncompactness $\lambda$. | math.OA | math |
MEASURES OF NONCOMPACTNESS ON THE STANDARD
HILBERT C∗-MODULE
DRAGOLJUB J. KEČKIĆ AND ZLATKO LAZOVIĆ
Abstract. We define a measure of noncompactness λ on the standard Hilbert
C ∗-module l2(A) over a unital C ∗-algebra, such that λ(E) = 0 if and only if E
is A-precompact (i.e. it is ε-close to a finitely generated projective submodule
for any ε > 0) and derive its properties. Further, we consider the known, Kura-
towski, Hausdorff and Istrăţescu measure of noncomapctnes on l2(A) regarded
as a locally convex space with respect to a suitable topology, and obtain their
properties as well as some relationship between them and introduced measure
of noncompactness λ.
1. Introduction
Measures of noncompactness (MNCs in further) have been studied almost for a
century. Roughly speaking, a MNC is a function which assigns a real number to
any bounded set in a given metric space, and this real number can be regarded as
a characteristic of the extent to which A is not totally bounded, (that is relatively
compact when completeness is supposed). There are many different MNCs on
metric spaces, among them the most cited are: Kuratowski, Hausdorf and Istrăţescu
MNC. Their definitions is given by:
Definition 1.1. Let (M, d) be a metric space, and let A ⊆ M be a bounded set.
a) The Kuratowski measure of noncompactness of A, denoted by α(A), is the
infimum of all d > 0 such that A admits a partitioning into finitely many subsets
whose diameters are less than d ([8], see also [9]).
b) The Hausdorff measure of noncompactness of A, denoted by χ(A), is the
infimum of all ε > 0 such that A has a finite ε-mash in M . (If we require that
this ε-mash belongs to A, we refer to inner Hausdorff measure of noncompactness,
denoted by χi(A).)
c) The Istrăţescu measure of noncompactness, denoted by I(A) , is the infimum
of all d > 0 such that there is an ε-discrete sequence in A, that is, a sequence
xn ∈ A, such that d(xm, xn) ≥ ε for all m 6= n ([6]).
Although different in general, these three functions share some common proper-
ties, e.g. the following:
Proposition 1.1. Let (M, d) be a metric space, and let µ be any of functions α,
χ, I defined above. Then µ has the following properties:
(a) regularity: µ(A) = 0 iff A is relatively compact;
2010 Mathematics Subject Classification. Primary: 46L08, Secondary: 47H08, 54E15.
Key words and phrases. Hilbert module, measures of noncompactness, uniform spaces.
The authors was supported in part by the Ministry of education and science, Republic of
Serbia, Grant #174034.
1
2
DRAGOLJUB J. KEČKIĆ AND ZLATKO LAZOVIĆ
(b) non-singularity: µ is equal to zero on any single-element set;
(c) monotonicity: µ(A) ≤ µ(B), whenever A ⊆ B;
(d) subadditivity: µ(A ∪ B) = max{µ(A), µ(B)}, for all A and B;
If, in addition, M is a normed space, i.e., M is a linear space and the metric d
is defined via d(x, y) = x − y then µ has additional properties:
(e) algebraic semi-additivity: µ(A + B) ≤ µ(A) + µ(B), for all A, B;
(f) semi-homogeneity: µ(tA) = tµ(A), for t ∈ C and all A ⊆ M ;
(g) invariance under translations: µ(x+A) = µ(A), for all A ⊆ M and x ∈ M ;
(h) Lipschitz continuity: µ(A) − µ(B) ≤ LdH (A, B), where L is an absolute
constant (Lα = 2, Lχ = 1, LI = 2) and d(A, B) is the so called Hausdorff
distance, that is dH (A, B) = max{supx∈A d(x, B), supy∈B d(y, A)};
(i) invariance to the transition to the closure and to the convex hull: µ(A) =
µ(A) = µ(co A);
(j) The functions α, χ and I are equivalent to each other, that is,
(1.1)
χ(A) ≤ I(A) ≤ α(A) ≤ 2χ(A)
for all bounded A.
These properties are well known and their proofs can be found throughout liter-
ature, for instance in [16]. (Though, it should be mentioned that inner Hausdorff
MNC does not satisfy the equality χi(co A) = χi(A), see [1, page 9])
Some of the mentioned properties were singled out in order to establish the
axiomatic definition of the abstract notion of the MNC, and this was done in various
different manners.
For more detailed exposition on MNCs on metric or normed spaces, the reader
is referred to [1], [4] or [17].
There is some extension of the MNC theory to the framework of uniform spaces
which will be described in Section 3. Among all uniform spaces, we are specially
interested in the standard Hilbert module l2(A) over a unital C∗-algebra A. It is
defined by
l2(A) = nx = (ξ1, ξ2, . . . ) ξj ∈ A,
+∞
X
j=1
j ξj converges in the norm topologyo,
ξ∗
and it is equipped with the A-valued inner product
l2(A) × l2(A) ∋ (x, y) 7→
+∞
X
j=1
ξ∗
j ηj ∈ A,
x = (ξ1, ξ2, . . . ),
y = (η1, η2, . . . ).
Since an arbitrary A-linear bounded operator on l2(A) does not need to have an
adjoint, the natural algebra of operators is Ba(l2(A)) - the algebra of all A-linear
bounded operators on l2(A) having an adjoint. It is known that Ba(l2(A)) is a
C∗-algebra, as well.
Among all operators in Ba(l2(A)), those that belong to the linear span of the
operators of the form x 7→ Θy,z(x) = z hy, xi (y, z ∈ l2(A)) are called finite rank
operators. The norm closure of finite rank operators is known as the algebra of
all "compact" operators. The quotation marks are usually written in order to
emphasize the fact that "compact" operators does not maps bounded sets into
relatively compact sets, as it is the case in the framework of Hilbert (and also
Banach) spaces, though they share many properties of proper compact operators
on a Hilbert space, [12], [13].
MEASURES OF NONCOMPACTNESS
3
(For general literature concerning Hilbert modules over C∗ algebras, including
the standard Hilbert module, the reader is referred to [10] or [14].)
In our earlier work [7], we construct a locally convex topology τ on l2(A) such
that T ∈ Ba(l2(A)) is "compact" implies that its image of the unit ball is totally
bounded with respect to τ . The converse is obtained in the special case A = B(H).
The aim of this note is to introduce a MNC on the standard Hilbert module l2(A)
and to derive its properties. In Section 2, we introduce the MNC λ such that for any
bounded set E ⊆ l2(A) the equality λ(E) = 0 holds if and only if E is precompact
in the sense of [14, Proposition 2.6], which implies that the corresponding MNC of
a given operator T is equal to 0 if and only if T is "compact".
In Section 3 we list some known results on MNCs on uniform spaces and also
derive some simple generalizations.
In Section 4, we discuss generalizations of Kuratowski, Hausdorff and Istrăţescu
measures of noncomapctness on l2(A) and obtain their relationship with MNC λ.
We use the following basic and also simple facts on Hilbert modules, that can
be found throughout the literature. To make proofs more easy we list and prove
them.
(F1) Let z1 ⊥ z2. Then z1 + z2 ≥ z1.
Indeed, we have
hz1 + z2, z1 + z2i = hz1, z1i + hz2, z2i ≥ hz1, z1i ,
in the order defined by the positive cone in A. Therefore
z1 + z22 = hz1 + z2, z1 + z2i ≥ hz1, z1i = z12.
(F2) Let M be a projective finitely generated submodule of l2(A), and let x ∈
l2(A) be arbitrary. Then
d(x, M ) = x − PM x,
where PM is orthogonal projection onto M with null-space M ⊥.
Indeed, by [14, Theorem 1.4.5] M is orthogonally complemented, l2(A) = M ⊕
M = PM , the range of PM
M ⊥, and there is PM : l2(A) → l2(A) such that P 2
is M and the kernel of PM is M ⊥. Let y ∈ M be arbitrary. Then
M = P ∗
x − y = (x − PM x) + (PM x − y) ≥ x − PM x,
by (F1), because x − PM x ∈ M ⊥ and PM x − y ∈ M .
2. Measure of noncompactness λ
Throughout this section, A will always denote a unital C∗-algebra, and its unit
will be denoted by 1. Also, l2(A) will denote the standard Hilbert C∗-module over
A defined in the introduction.
In [14, Proposition 2.6] A-precompact sets were defined as those bounded sets E
such that for all ε > 0 there is a free finitely generated module M ∼= An such that
d(E, M ) := sup
x∈E
d(x, M ) = sup
x∈E
inf
y∈M
d(x, y) < ε.
We generalize this notion in the following way.
Definition 2.1. Let E ⊆ l2(A) be a bounded set. The measure of noncompactness
of E, denoted by λ(E) is the greatest lower bound of all η > 0 for which there is a
4
DRAGOLJUB J. KEČKIĆ AND ZLATKO LAZOVIĆ
free finitely generated module M ≤ l2(A) such that
d(E, M ) := sup
x∈E
inf
y∈M
d(x, y) < η.
Proposition 2.1. The measure of noncompactness λ(E) can be computed as:
(1) λ(E) = inf M∈F supx∈E d(x, M ), where F is the set of all free finitely gen-
erated modules;
(2) λ(E) = limn→+∞ supx∈E x − Pnx = inf n≥1 supx∈E x − Pnx, where
Pn : l2(A) → l2(A) is given by Pn(x1, x2, . . . ) = (x1, x2, . . . , xn, 0, 0, . . . ).
Proof. The equation (1) is obvious. The sequence I − Pn is decreasing. Hence
limn→+∞ supx∈E x − Pnx = inf n≥1 supx∈E x − Pnx. Since Pnl2(A) is a free
finitely generated module, we have immediately
λ(E) ≤ inf
n≥1
sup
x∈E
x − Pnx.
To get the opposite inequality, let ε > 0. Then, there is a free finitely generated
module M such that d(x, M ) < λ(E) + ε. Denote the projection on M by Q. By
[14, Proposition 2.2.1.], Q − PnQ → 0 as n → +∞. Since PnQx ∈ Pnl2(A), we
have by (F2)
x − Pnx =d(x, Pnl2(A)) ≤ x − PnQx ≤
≤x − Qx + Qx − PnQx ≤ λ(E) + ε + KQ − PnQ,
where x ≤ K for all x ∈ E (E is bounded). Thus
x − Pnx ≤ λ(E) + ε + KQ − PnQ → λ(E) + ε,
as n → +∞,
sup
x∈E
which finishes the proof.
(cid:3)
Proposition 2.2. The measure of noncompactness λ has the following properties
(1) if E ⊆ F then λ(E) ≤ λ(F );
(2) λ(E ∩ F ) ≤ min{λ(E), λ(F )};
(3) λ(E ∪ F ) ≤ max{λ(E), λ(F )};
(4) λ(E + F ) ≤ λ(E) + λ(F ).
(5) λ(Ea) ≤ λ(E)a, where a ∈ A. If, in addition, a is invertible, then also
In particular, when a is unitary then λ(Ea) =
a−1−1λ(E) ≤ λ(Ea).
λ(E).
(6) λ(co E) = λ(E), where co E = {Pn
is the convex hull of E.
i=1 tixi 0 ≤ ti ∈ R,Pn
i=1 ti = 1, xi ∈ E}
Proof. (1) It is obvious.
(2) This follows from (1).
(3) Let d = max{λ(E), λ(F )}. Then for all x ∈ E, as well as for all x ∈ F , we
have x − Pnx ≤ d + ε, for all ε > 0 and n large enough. Hence the result.
(4) Let z ∈ E + F . Then z = x + y for some x ∈ E, y ∈ F . We have
z − Pnz ≤ x − Pnx + y − Pny ≤ λ(E) + ε + λ(F ) + ε,
for n large enough.
(5) Any z ∈ Ea is of the form z = ya for some y ∈ E. Therefore z − Pnz ≤
y − Pny a and the inequality follows by taking a limit. If a has the inverse
a−1, then E = (Ea)a−1 and by previous λ(E) ≤ λ(Ea)a−1.
MEASURES OF NONCOMPACTNESS
5
i=1 tixi for some xi ∈ E, and positive ti such that
Pn
(6) Let x ∈ co E. Then x = Pn
i=1 ti = 1. We have
n
x − Pnx = (cid:13)(cid:13)(cid:13)
ti(xi − Pnxi)(cid:13)(cid:13)(cid:13)
X
i=1
≤
n
X
i=1
tixi − Pnxi ≤ sup
x∈E
x − Pnx.
Thus, supx∈co E x−Pnx ≤ supx∈E x−Pnx. The opposite inequality is obvious.
The required follows from Proposition 2.1-(2).
(cid:3)
Proposition 2.3. Let B denote the unit ball in l2(A). Then λ(B) = 1.
Proof. Any submodule contains the origin. Hence λ(B) ≤ 1. Let 0 < δ < 1, and
let M be some free finitely generated submodule of l2(A). Then, there is nontrivial
y ∈ M ⊥ and δy−1y ∈ B ∩ M ⊥. We have
d(B, M ) ≥ d(δy−1y, M ).
However, δy−1y ⊥ M which implies that for all x ∈ M we have
δy−1y − x2 = δ2 + x2 ≥ δ2.
Hence d(B, M ) ≥ δ. Thus λ(B) ≥ δ.
Corollary 2.4. If E ⊆ F + δB then λ(E) ≤ λ(F ) + δ.
Proof. Immediately follows from Proposition 2.2-(4) and Proposition 2.3.
(cid:3)
(cid:3)
Proposition 2.5. The measure of noncompactness λ has the following continuity
properties:
(1) λ(E) − λ(F ) ≤ dH (E, F ) = max{d(E, F ), d(F, E)} (dH stands for the so
called Hausdorf distance.)
(2) λ(E) = λ(E) (E stands for the norm closure of E);
(3) λ(E) = 0 iff E is A-precompact;
(4) λ(E) ≤ supx∈E x.
Proof. (1) Let d = dH (E, F ). Then E ⊆ F + dB and by Corollary 2.4
λ(E) ≤ λ(F ) + d,
i.e.
λ(E) − λ(F ) ≤ d.
Similarly, F ⊆ E + dB implying λ(F ) − λ(E) ≤ d.
(2) As it is easy to see dH (E, E) = 0 we can apply the previous item.
(3) Follows directly from the definition.
(4) Follows from E ⊆ (supx∈E x) · B.
(cid:3)
Example 2.1. In Proposition 2.2-(5) the strict inequalities might hold.
Indeed, let the algebra A contain a nontrivial projection, say p, and let A(1 − p)
is isomorphic to A. (For instance A = L∞(0, 1) and p = χ[0,1/2].) Let
E = n(a1p + b1(1 − p), a2p + b2(1 − p), . . . )
∞
X
n=1
a∗
nan ≤ 1,
∞
X
n=1
nbn ≤ 4o
b∗
and let a = p. Then p = 1 and λ(E) ≥ 2, since E contains a copy of a ball of
radius 2 in l2(A) (when aj = 0). On the other hand λ(Ep) ≤ 1. Indeed, Ep is
contained in the unit ball of l2(A).
Finally, we want to define MNC of an operator T ∈ Ba(l2(A)). As it is expected,
it will be the MNC of its image of the unit ball.
6
DRAGOLJUB J. KEČKIĆ AND ZLATKO LAZOVIĆ
Definition 2.2. Let T ∈ Ba(l2(A)) be an adjointable operator. We set
where B1 is the unit ball in l2(A).
λ0(T ) = λ(T (B1)),
Proposition 2.6. The function λ0 has the following properties:
(a) λ0 is subadditive, i.e.
λ0(T1 + T2) ≤ λ0(T1) + λ0(T2);
(b) λ0 is positively homogeneous, i.e.
λ0(cT ) = cλ0(T ),
for all c > 0 and all T ∈ Ba(l2(A));
(c) λ0(T ) ≤ T , for all T ∈ Ba(l2(A));
Proof. Direct verification.
(cid:3)
We will be able to say more on the MNC λ0 in section 4.
3. Measure of noncompactness on uniform spaces - known results
Let us recall some basic definitions and facts concerning uniform spaces. For
more details see [18] or [5].
Uniform spaces are those topological spaces in which one can deal with notions
such as Cauchy sequence, Cauchy net or uniform continuity. Although it is usual
to define them as spaces endowed with a family of sets in X × X given as some
kind of neighborhoods of the diagonal, so called entourages, for our purpose it is
more convenient to give an equivalent definition, via a family of semi-metrics.
Definition 3.1. A nonempty set endowed with a family of semi-metrics, functions
dα : X × X → [0, +∞) satisfying (i) dα(x, y) ≥ 0; (ii) dα(x, y) = dα(y, x); (iii)
dα(x, z) ≤ dα(x, y) + dα(y, z) is called a uniform space.
Remark 3.1. There is some ambiguity in literature; sometimes, functions dα from
the previous definition are called "pseudo-metrics", whereas the term "semi-metric"
is reserved for a different notion.
All dα are metrics except they do not distinguish points, i.e. there might be
dα(x, y) = 0 for some x = y. However it is provided that for all x = y there is an α
such that dα(x, y) > 0. The family of sets Bdα(x; ε) = {y ∈ Xdα(x, y) < ε} makes
a basis for some topology. It is well known that a topological space X is a uniform
space if and only if it is completely regular.
Any locally convex topological vector space is a uniform space. Indeed, there
is a family of semi-norms generating its topology. This family can be obtained by
Minkowski functionals of basic neighborhoods of zero. And an arbitrary semi-norm
define a semi-metric in a natural way. Conversely, any family of semi-norms that
distinguishes points leads to a locally convex Hausdorff topological vector space.
We point out two generalizations of the notion of MNC to the framework of
uniform spaces, i.e, those topological spaces that arise from the family of semi-
metrics.
Sadovskii [18] considered a uniform space X and a family P of semi-metrics that
are uniformly continuous on X × X. Starting with Kuratowski and Hausdorff MNC
on a semi-metric space (which is defined exactly as on a metric space), Sadovskii
MEASURES OF NONCOMPACTNESS
7
defined corresponding MNCs α, χ : Υ → G, where Υ denotes the family of all
subsets that are bounded with respect to any semi-metric p ∈ P on the given
uniform space, and G denotes the set of functions g : P → [0, +∞) with uniformity
generated by pointwise convergence and the natural partial ordering: g1 ≤ g2 ⇔
(∀p ∈ P ) g1(p) ≤ g2(p). Their definitions are
[α(E)](p) = inf nd > 0 : E =
m
[
j=1
Ej, for some Ej, diam(Ej) < do,
[χ(E)](p) = inf nε > 0 : E ⊆
m
[
j=1
Bp(xj ; ε) for some xj ∈ Xo.
For such defined α and χ, in [18, §1.2.3. and §1.2.5.], the following is proved,
provided that the family P generates the topology on X.
Theorem 3.1. The Kuratowski and Hausdorff measures of noncompactness (µ = α
or µ = χ) have the following properties:
(a) µ is non-singular, that is, they are zero on any single-element set;
(b) µ is continuous, that is, for all E ∈ Υ, p ∈ P and ε > 0 there is an
entourage V in X such that for all E1 that is V -close to E there holds
µ(E1)(p) − µ(E)(p) < ε;
(c) µ is semi-additive, that is, for all E1, E2 we have
µ(E1 ∪ E2) = max{µ(E1), µ(E2)};
(d) The function is algebraically semi-additive, that is,
µ(E1 + E2)(p) ≤ µ(E1)(p) + µ(E2)(p)
for all E1, E2 ∈ Υ;
(e) µ is invariant under shifts, that is,
µ(x + E) = µ(E)
for all E ∈ Υ, x ∈ X;
(f) µ is invariant under the transition to its closure and to the convex hull of
the set, that is,
µ(co E) = µ(E) = µ(E)
for all E ∈ Υ;
(g) µ is uniformly continuous, that is, for all p ∈ P and ε > 0 there is
an entourage V in X such that for all V -close E1 and E2 there holds
µ(E1)(p) − µ(E2)(p) < ε;
(h) The functions α, χ and I are equivalent to each other, that is,
χ(Ω) ≤ I(Ω) ≤ α(Ω) ≤ 2χ(Ω)
for all Ω ∈ Υ.
Although it was not done in [18], one can also define the Istrăţescu MNC in a
similar way, i.e.
[I(E)](p) = inf{ε > 0 : E ∋ xn, p(xn, xm) ≥ ε for all m 6= n}.
By (1.1) one can easily derive
(3.1)
[χ(E)](p) ≤ [I(E)](p) ≤ [α(E)](p) ≤ 2[χ(E)](p)
for all bounded E.
Also, Theorem 3.1 hold for µ = I. Part (a) is obvious, parts (d) and (e) follow
from [3, Proposition 1], part (c) follows from [6], parts (b) and (g) from (3.1) and
the corresponding properties of χ, and finally (f) can be derived from [1, Theorem
1.3.4].
Arandjelović [2] dealt with an arbitrary uniform space, and gave the following:
8
DRAGOLJUB J. KEČKIĆ AND ZLATKO LAZOVIĆ
Definition 3.2. Let X be a uniform space, metric space or semi-metric space. Any
function Φ defined on the partitive set of X, which satisfies the following:
(1) Φ(E) = +∞ if and only if E is unbounded;
(2) Φ(E) = Φ(E);
(3) from Φ(E) = 0 follows that E is totally bounded set;
(4) from E ⊆ F it follows Φ(E) ≤ Φ(F );
(5) if X is complete, and if {En}n∈N is a sequence of closed subsets of X
such that En+1 ⊆ En for each n ∈ N and limn→∞ Φ(En) = 0, then K =
Tn∈N En is a nonempty compact set.
is called a measure of noncompactness on X.
Remark 3.2. Note that the only nontrivial requirement in (5) is that K is nonempty.
Moreover, condition (5) can be replaced by a weaker one - Φ(A ∪ {x}) = Φ(A). It
was shown in [15], see also [].
Theorem 3.2. [2, Theorem 3] Let X be a uniform space and let {dii ∈ I} be a
family of semi-metrics which defines topology on X. Denote by µi arbitrary MNC
on the semi-metric space (X, di) for each i ∈ I. Then the function µ∗ : X → [0, +∞]
defined by
for each E ∈ X, is a measure of noncompactness on X.
µ∗(E) = sup
i∈I
µi(E)
Uniform spaces make a proper subclass of all topological spaces, but still wide
enough. For instance all topological vector spaces are uniform spaces. Hence, we
can apply results of Arandjelović to topological vector spaces.
4. Measures of "noncompactness" over standard Hilbert W ∗-modul
l2(B(H))
In this section we shall discuss standard Hilbert modules over a W ∗-algebra A,
a narrower class then that considered in section 2. As it is well known, A always
has a unit.
In our earlier work [7], we construct a locally convex topology τ on l2(A) such
that T ∈ Ba(l2(A)) is "compact" implies its image of the unit ball is totally bounded
with respect to τ . This topology is defined via the family of semi-norms pϕ,y
(4.1)
∞
pϕ,y(x) = vuut
X
j=1
ϕ(η∗
j ξj)2,
x = (ξ1, ξ2, . . . ) ∈ l2(A),
where ϕ ∈ A∗ is a normal state and y = (η1, η2, ...) is a sequence of elements in A
such that
(4.2)
ϕ(η∗
j ηj) = 1.
sup
j≥1
Also, in special case, where A = B(H) is the full algebra of all bounded operators
on a Hilbert space H, the converse is also proved, i.e. that any T ∈ Ba(l2(A)) whose
image of the unit ball is totally bounded with respect to τ must be "compact".
Construction described in the previous section endows the space l2(A) by the
corresponding Kuratowski, Hausdorff and Istrăţescu measure of noncompactness,
MEASURES OF NONCOMPACTNESS
9
α, χ, I : Υ → G,
[α(E)](pϕ,y) = inf nε > 0 : E =
n
[
i=1
Si, pϕ,y(x′ − x′′) < ε, ∀x′, x′′ ∈ Sio,
[χ(E)](pϕ,y) = inf nε > 0 : E ⊂
[
i=1
where Bpϕ,y (xi, ε) = {ypϕ,y(y − xi) < ε}, and
[I(E)](pϕ,y) = sup{ε > 0 : there is S ⊂ E such that pϕ,y(x′−x′′) ≥ ε, ∀x′, x′′ ∈ S}.
Bpϕ,y (xi, ε), xi ∈ l2(A)o,
n
The function α, χ and I can be regarded as functions depending on two variables,
on the bounded set Ω and on the semi-norm pϕ,y. If we want to obtain a MNC
that not depends on a particular semi-norm, we can use the functions χ∗, α∗,
I ∗ : Υ → [0, +∞) defined by
(4.3)
χ∗(E) = sup
pϕ,y ∈P
α∗(E) = sup
pϕ,y ∈P
I ∗(E) = sup
pϕ,y∈P
[χ(E)](pϕ,y),
[α(E)](pϕ,y),
[I(E)](pϕ,y)
for each E ∈ Υ, where P is the set of all semi-norms of the form (4.1). Since α, χ
and I annihilates all singletons (Theorem 3.1-(a)), they satisfy condition in Remark
3.2, and hence they are measures of noncompactness in the sense of Definition 3.2.
By Theorem 3.2 α∗, χ∗ and I ∗ are measures of noncompactness on (l2(A), τ ) in the
sense of Definition 3.2, as well. Also, properties (c), (d), (e) and (f) in Theorem 3.1
are easily transferred to α∗, χ∗ and I ∗, by taking a supremum. Finally, by (3.1),
we have
(4.4)
χ∗(E) ≤ I ∗(E) ≤ α∗(E) ≤ 2χ∗(E).
Remark 4.1. Note that l2(A) is rarely complete, due to [7, Proposition 3.3]. There-
fore, the condition (5) in Definition 3.2 is vague, unless l2(A)′ ∼= l2(A) which is
equivalent to the condition that A is finite dimensional.
We want to place the MNC λ, discussed in section 2, somewhere in the preceding
chain of inequalities.
Proposition 4.1. For any bounded set E ⊆ l2(A), we have
χ∗(E) ≤ λ(E).
Proof. Let E be a bounded set, and let Pn denote the projection to the first n
coordinates in l2(A), i.e. Pn(ξ1, ξ2, . . . ) = (ξ1, ξ2, . . . , ξn, 0, . . . ). Since E ⊆ PnE +
(I − Pn)E, and since χ∗ is subadditive, we have
χ∗(E) ≤ χ∗(PnE) + χ∗((I − Pn)E).
However, by ([7, Proposition 3.4.]), the set PnE is totally bounded, and we have
χ∗(PnE) = 0. Hence
χ∗(E) ≤ χ∗((I − Pn)E) ≤ sup
x∈E
(I − Pn)x,
for all n ∈ N. Therefore, by Proposition 2.1-(2), χ∗(E) ≤ λ(E).
(cid:3)
10
DRAGOLJUB J. KEČKIĆ AND ZLATKO LAZOVIĆ
The preceding Proposition establishes a lower bound of λ. Before we obtained
an upper bound for λ, in a special case, we introduce balanced sets.
Definition 4.1. Let E ⊂ l2(A) be a bounded set.
(a) We say that E is A-balanced if x · u ∈ E whenever x ∈ E and u ∈ A is
unitary. (This definition is motivated by the notion of balanced sets on
topological vector spaces over the field C, where u unitary is reduced to
u = 1.)
(b) By A-balanced hull of E we assume the minimal balanced set containing
E, that is S Eu, where the union is taken over all unitaries u ∈ A.
In Proposition 2.2-(5) we proved λ(Eu) = λ(E). We give two extensions of this
statement, the first of them concerning the balanced hull.
Proposition 4.2. Let E ⊆ l2(A) be a bounded set and let F be its A-balanced hull.
Then λ(F ) = λ(E).
Proof. For all x ∈ E and all unitaries u, we have u = 1, and hence xu−Pnxu ≤
x − Pnxu. Therefore, by Proposition 2.1-(2), we have
λ(F ) = lim
n→+∞
sup
x∈E,u−unitary
xu − Pnxu ≤ lim
n→+∞
sup
x∈E
x − Pnx = λ(E).
The opposite inequality λ(E) ≤ λ(F ) follows from E ⊆ F .
(cid:3)
Proposition 4.3. Let E ⊆ l2(A) be a bounded set, let u ∈ A be a unitary and let
µ stands for any of Kuratowski, Hausdorff or Istrăţescu MNC. Then
µ∗(Eu) = µ∗(E).
Proof. First, observe that given a normal state ϕ on A and a unitary u ∈ A, the
mapping ϕu, ϕu(x) = ϕ(u∗xu) is also a normal state. Obviously, ϕu(1) = 1 and
ϕu(x) ≥ 0 whenever x ≥ 0. Hence, it suffices to prove that ϕu is normal. Let xα be
an increasing net with the least upper bound x. Then uxαu∗ is also an increasing
net, bounded by uxu∗. Thus, its least upper bound is less then uxu∗. Moreover, it
is equal to uxu∗ by interchanging roles. Therefore
sup ϕu(xα) = sup ϕ(uxαu∗) = ϕ(uxu∗) = ϕu(x).
Let ε > 0 be arbitrary. By (4.3) there is a semi-norm pϕ,y ∈ P such that
[µ(Eu)](pϕ,y) > µ∗(Eu) − ε. Next, we have pϕ,y(xu) = pϕu,yu, where yu∗ =
(η1u∗, η2u∗, . . . ). Indeed
+∞
+∞
pϕ,y(xu)2 =
ϕ(η∗
j ξju)2 =
X
j=1
X
j=1
ϕ(u∗(ηj u∗)∗ξju)2 = pϕu,yu∗(x)2.
(Note that the pair (ϕu, yu∗) trivially satisfies (4.2).) Therefore, [µ(E)](pϕu,yu∗) =
[µ(Eu)](pϕ,y) > µ∗(Eu) − ε and hence µ∗(E) ≥ µ∗(Eu). The opposite inequality
follows by E = (Eu)u−1.
(cid:3)
Remark 4.2. We don't know whether Kuratowski, Hausdorff and Istrăţescu MNCs
are stable with respect to balanced hull.
Now, we are going to derive an upper bound for the MNC λ. Namely, in a special
case, where A = B(H), we can obtain the exact position of λ in the inequality chain
(4.4) for balanced sets.
MEASURES OF NONCOMPACTNESS
11
Theorem 4.4. Let A = B(H). Let E ⊆ l2(A) be an A-balanced set. Then
where E = supx∈E x.
λ(E) ≤ pEI ∗(E),
Proof. Let Pk denote the projection to the first k coordinates, i.e. Pk(ξ1, ξ2, . . . )
= (ξ1, . . . , ξk, 0, 0, . . . ). It is well known that all Pk are "compact". If inf k≥1 (I −
Pk)E = 0, then from Proposition 4.1 and (4.4), it follows λ∗(E) = χ∗(E) =
I ∗(E) = 0. So, let
δ = inf
k≥1
(I − Pk)E > 0.
Then immediately, δ ≤ E. Choose ε > 0 such that ε < δ2. Define the sequence
of projections Qn ∈ {P1, P2, . . . } and the sequences of vectors xn and zn ∈ l2(A)
in the following way. Let Q0 = 0.
If Qn−1 is already defined, there is xn ∈ E
2 (cid:16)δ + pδ2 − ε/2(cid:17). Since
such that xn ≥ (I − Qn−1)xn > C1, where C1 = 1
limk→+∞ (I − Pk)(I − Qn−1)xn = 0, there is a positive integer kn such that
2 (cid:16)δ − pδ2 − ε/2(cid:17)o.
(I − Pkn )(I − Qn−1)xn < C2, where C2 = minn ε
Define Qn = Pkn and
2E , 1
(4.5)
zn = Qn(I − Qn−1)xn.
The sequences xn and zn have the following properties:
Firstly, by definition, there hold the inequalities
(4.6)
(4.7)
(4.8)
(I − Qn)(I − Qn−1)xn < C2,
zn ≤ xn ≤ E,
zn ≥ (I − Qn−1)xn − (I − Qn)(I − Qn−1)xn > C1 − C2.
Secondly,
(4.9)
hzn, xni = hzn, zni .
Indeed, since zn = Qn(I − Qn−1)xn, we have
hzn, xni = hQn(I − Qn−1)xn, xni =
= hQn(I − Qn−1)xn, (I − Qn−1)Qnxni = hzn, zni .
Thirdly, for m > n we have
(4.10)
hzm, xni < C2E.
Indeed, for such m and n we have Qn−1 ≤ Qn ≤ Qm−1, i.e. I − Qm−1 ≤ I − Qn ≤
I − Qn−1, implying I − Qm−1 = (I − Qm−1)(I − Qn)(I − Qn−1), and thus
hzm, xni = h(I − Qm−1)zm, xni =
= hzm, (I − Qm−1)(I − Qn)(I − Qn−1)xni =
= hzm, (I − Qn)(I − Qn−1)xni .
Therefore, by (4.6) and (4.8)
hzm, xni ≤ zm · (I − Qn)(I − Qn−1)xn ≤ C2E.
12
DRAGOLJUB J. KEČKIĆ AND ZLATKO LAZOVIĆ
Let us construct a semi-norm p, continuous in τ , and a totally discrete sequence
from E. Since by (4.8) zn2 = hzn, zni > (C1 − C2)2, we can choose a normal
state ϕ and υj, νj ∈ A according to [7, Lemma 4.6.]), such that
(4.11)
ϕ(υ∗
n hzn, zni νn) > (C1 − C2)2.
Consider the semi-norm p given by
+∞
p(x) = vuut
X
j=1
ϕ(hzjυj, xi)2.
By (4.5) there is a sequence ζj ∈ A such that
Define ωj = ζjυn/ϕ(υ∗
Also, for x = (ξ1, ξ2, . . . ) we have
zn = (0, . . . , 0, ζkn−1+1, . . . , ζkn , 0, . . . ).
nζ∗
j ζjυn)1/2, for kn−1 + 1 ≤ j ≤ kn. Obviously ϕ(ω∗
j ωj) = 1.
ϕ(hznυn, xi)2 = (cid:12)(cid:12)(cid:12)
kn
X
j=kn−1+1
ϕ(υ∗
nζ∗
j ζjυn)1/2ϕ(ω∗
kn
≤
X
j=kn−1+1
ϕ(υ∗
nζ∗
j ζjυn)
kn
X
j=kn−1+1
≤
2
j ξj)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)ϕ(ω∗
2
=
j ξj)(cid:12)(cid:12)
= ϕ(υ∗
n hzn, zni υn)
kn
X
j=kn−1+1
2
.
(cid:12)(cid:12)ϕ(ω∗
j ξj)(cid:12)(cid:12)
Including (4.7) we obtain ϕ(υ∗
hence
n hzn, zni υn) ≤ υ∗
n hzn, zni υn = zn2 ≤ E2 and
p(x)2 =
+∞
X
n=1
ϕ(hznυn, xi)2 ≤ E2
+∞
X
j=1
ϕ(ω∗
j ξj)2 = E2pϕ,ω1,...,ωn,...(x)2.
Thus, we conclude that p is well defined and also that it is continuous with respect
to τ .
Also, E is A-balanced, so xnνn ∈ E. Finally we shall prove that xnνn is a totally
discrete sequence. Indeed, for m > n we have
p(xmνm − xnνn) ≥ ϕ(hzmυm, xmνm − xnνni) ≥
m hzm, xmi νm) − ϕ(υ∗
≥ ϕ(υ∗
m hzm, xni νn) .
However, by (4.9) and (4.11),
ϕ(υ∗
m hzm, zmi νm) > (C1 − C2)2
and, by (4.10)
Therefore
and
ϕ(υ∗
m hzm, xni νn) ≤ hzm, xni < C2E.
p(xmνm − xnνn) > (C1 − C2)2 − C2E ≥ δ2 − ε
pϕ,ω1,...,ωn,...(xmνm − xnνn) >
δ2 − ε
E
.
For ε ∈ (0, δ2), we have I ∗(E) ≥ δ2−ε
λ(E) ≤ pEI ∗(E).
E and hence I ∗(E) ≥ δ2
E = λ(E)2
E . Thus,
(cid:3)
MEASURES OF NONCOMPACTNESS
13
Remark 4.3. The preceding proof is adapted proof of our earlier result [7, Theorem
4.10].
Corollary 4.5. On the standard Hilbert module l2(B(H)) over the algebra B(H)
of all bounded operators on a Hilbert space H, there holds
χ∗(E) ≤ λ(E) ≤ pEI ∗(E) ≤ pEα∗(E) ≤ p2Eχ∗(E),
for any balanced set, i.e. a set for which x ∈ E, u-unitary implies xu ∈ E.
Finally, we discuss some relationship between the MNC of an arbitrary ad-
jointable operator λ0 introduced in Definition 2.2, and corresponding MNCs derived
from α∗, χ∗ and I ∗.
Definition 4.2. Let A is an arbitrary W ∗-algebra and let T ∈ Ba(l2(A)) be an
adjointable operator. The functions α∗
0 : Ba(l2(A)) → [0, +∞) defined by
0, χ∗
0, I ∗
α∗
0(T ) = α∗(T (B1)), χ∗
0(T ) = χ∗(T (B1)),
I ∗
0 (T ) = I ∗(T (B1))
are called, respectively, Kuratowski, Hausdorff and Istrăţescu measure of noncom-
pactness of the operator T .
Proposition 4.6. Let A be an arbitrary W ∗-algebra, let T , S ∈ Ba(l2(A)) and let
µ stands for any of MNCs α, χ, I. Then
0, χ∗
(a) All α∗
0 and I ∗
µ∗
0(T + S) ≤ µ∗
(b) The functions α∗
0(S), µ∗
0(cT ) = cµ∗(T ),
0 are subadditive and positivel homogeneous, i.e. there holds
0(T ) + µ∗
0, χ∗
0 and I ∗
χ∗
0(T ) ≤ I ∗
0(T ) ≤ λ0(T ).
0 are equivalent to each other, that is,
0 (T ) ≤ α∗
0(T ) ≤ 2χ∗
for all c > 0.
0(T ).
Also, there holds χ∗
0(T ), λ0(T ) ≤ T and α∗
0(T ), I ∗
0 (T ) ≤ 2T .
(c) χ∗
(d) If T is "compact", i.e. T belongs to the closed linear space generated by
x 7→ z hy, xi, then λ0(T ) = χ0(T ) = α0(T ) = I(T ) = 0. In general, the
converse might not hold.
0(T + K) = µ∗
operators K.
0(T ), as well as λ0(T + K) = λ0(T ) for all "compact"
(e) µ∗
Proof. Part (a) follows easily from Theorem 3.1-(d), whereas part (b) follows from
(4.4) and Proposition 4.1.
Since T (B1) ⊆ B(0; T ) = T B1, it follows λ0(T ) ≤ T according to Propo-
sition 2.3. Other inequalities in part (c) follows from part (a).
If T is "compact", then T is norm limit of finite rank operators. Hence T (B1)
is A-precompact. Therefore λ0(T ) = 0. This, together with (b) proves (d). The
converse does not always hold due to [7, Example 5.1.].
Finally, (e) follows from (d).
(cid:3)
In the case A = B(H), we can obtain more.
Proposition 4.7. Let A = B(H) and let T ∈ Ba(l2(B(H))). Then
(a) There holds
0(T ) ≤ λ0(T ) ≤ qT I ∗
χ∗
0(T ) = 0, µ ∈ {α, χ, I} iff λ0(T ) = 0 iff T is a "compact" operator;
0(T ) ≤ q2T χ∗
0 (T ) ≤ qT α∗
0(T ).
(b) µ∗
14
DRAGOLJUB J. KEČKIĆ AND ZLATKO LAZOVIĆ
Proof. Part (a) follows from Proposition 4.6-(b) and Theorem 4.4, since T (B1) is
a balanced set (y = T x ∈ T (B1) implies yu = T (xu) ∈ T (B1)).
Part (b) follows from part (a) and Proposition 4.6-(d). Indeed, if any of four
0(T ) = 0, and hence, T (B1) is totally
(cid:3)
MNCs annihilate T , then, by part (a), α∗
bounded in the topology τ . By [7, Theorem 4.10.], T is "compact".
5. Three questions
Question 5.1. It is easy to obtain the following: Let E ⊆ l2(A) be bounded. For
any ε > 0 there is an A-precompact set Cε such that E ⊆ Cε + (λ(E) + ε)B. (For
instance Cε = E ∩ M for a suitable free finitely generated M .)
Is it possible to get something stronger: There is an A-precompact set C such
that E ⊆ C + λ(E) · B?
Question 5.2. Among all properties of MNCs on a Banach space, it turns out that
the most important is µ(co E) = µ(E). This was proved in this note for λ if we co E
1 cjxj P cj = 1, cj ∈ R, xj ∈ E}.
regard as a real field convex hull, i.e. co E = {Pn
However, there is a notion of A-convex hull (see for instance [11])
coA E = n
n
X
j=1
j xjaj xj ∈ E, aj ∈ A,X a∗
a∗
j aj = 1o.
Is it possible to obtain λ(coA E) = λ(E)?
Question 5.3. As it was mentioned in Remark 4.2, we ask for the following: Is it
true µ∗(E) = µ∗(F ), where F denotes the balanced hull of E, i.e. F = Su Eu (the
union runs through all unitaries u), and µ denotes any of Kuratowski, Hausdorff or
Istrăţescu MNC?
Acknowledgement. The authors was supported in part by the Ministry of edu-
cation and science, Republic of Serbia, Grant #174034.
References
[1] R. R. Akhmerov, M. I. Kamenskii, A. S. Potapov, A. E. Rodkina, and B. N. Sadovskii.
Measures of Noncompactness and Condensing Operators. Birkhäuser Verlag, Basel, Boston,
Berlin, 1992. Translated from Russian.
[2] Ivan D. Arandjelović. Measure of noncompactness on uniform spaces. Math. Morav., 2:1–8,
1998.
[3] Ivan D. Arandjelović. Some properties of Istratescu's measure of noncompactness. Filomat,
13:99–104, 1999.
[4] J. M. Ayerbe Toledano, T. Dominguez Benavides, and G. Lopez Acedo. Measures of Non-
compactness in Metric Fixed Point Theory. Springer Basel AG, 1997.
[5] N. Bourbaki. General Topology, chapters 1–4. Springer, 1987.
[6] Vasile Ion Istrăţescu. On a measure of noncompactness. Bull. Math. Soc. Sci. Math. R.S.
Roumanie (N.S), 16:195–197, 1972.
[7] Dragoljub J. Kečkić and Zlatko Lazović. Compact and "compact" operators on the standard
Hilbert module over a W ∗ algebra. Ann. Funct. Anal., to appear:13 pages.
[8] C. Kuratowsky. Sur les espaces complets. Fund. Math., 15:301–309, 1930.
[9] C. Kuratowsky. Topologie I. Monografje Matematyczne, Warszaw, 4th edition, 1958.
[10] K. L. Lance. Hilbert C ∗-Modules: A toolkit for operator algebraists. Cambridge University
Press, 1995.
[11] Bojan Magajna. C ∗-convex sets and completely positive maps. Integral Eq. Op. Th., 85:37–
62, 2016.
MEASURES OF NONCOMPACTNESS
15
[12] V. M. Manuilov. Diagonalization of compact operators in Hilbert modules over finite W ∗-
algebras. Ann. Global Anal. Geom., 13(3):207–226, 1995.
[13] V. M. Manuilov. Diagonalization of compact operators on Hilbert modules over C ∗-algebras
of real rank zero. Math. Notes, 62(6):726–730, 1997.
[14] V. M. Manuilov and E. V. Troitsky. Hilbert C ∗-modules. Translations of mathematical mono-
graphs Vol. 226. AMS, Providence, Rhode Island, 2005.
[15] Marina M. Milovanović-Arandjelović. Measures of noncompactness on uniform spaces–the
axiomatic approach. Filomat, 15:221–225, 2001.
[16] Mohammad Mursaleen, Syed M. H. Rizvi, and Bessem Samet. Measures of noncompactness
and their applications. In J. Banas, M. Jleli, M. Mursaleen, B. Samet, and Vetro C., editors,
Advances in Nonlinear Analysis via the Concept of Measure of Noncompactness, chapter 2,
pages XIII, 487. Springer, 2017.
[17] Vladimir Rakočević. Measures of noncompactness and some applications. Filomat, 12.
[18] A. I. Sadovskii. Limit-compact and condensing operators. Russian Math. Surveys, 27(1):85–
155, 1972.
University of Belgrade, Faculty of Mathematics, Studentski trg 16-18, 11000
Beograd, Serbia
E-mail address: [email protected]
University of Belgrade, Faculty of Mathematics, Studentski trg 16-18, 11000
Beograd, Serbia
E-mail address: [email protected]
|
1806.10110 | 1 | 1806 | 2018-06-26T17:08:20 | Infinite characters on $GL_n(\mathbf{Q})$, on $SL_n(\mathbf{Z}),$ and on groups acting on trees | [
"math.OA",
"math.GR"
] | Answering a question of J. Rosenberg, we construct the first examples of infinite characters on $GL_n(\mathbf{K})$ for a global field $\mathbf{K}$ and $n\geq 2.$ The case $n=2$ is deduced from the following more general result. Let $G$ a non amenable countable subgroup acting on locally finite tree $X$. Assume either that the stabilizer in $G$ of every vertex of $X$ is finite or that the closure of the image of $G$ in ${\rm Aut}(X)$ is not amenable. We show that $G$ has uncountably many infinite dimensional irreducible unitary representations $(\pi, \mathcal{H})$ of $G$ which are traceable, that is, such that the $C^*$-subalgebra of $\mathcal{B}(\mathcal{H})$ generated by $\pi(G)$ contains the algebra of the compact operators on $\mathcal{H}.$ In the case $n\geq 3,$ we prove the existence of infinitely many characters for $G=SL_n(R)$, where $n\geq 3$ and $R$ is an integral domain such that $G$ is not amenable. In particular, the group $SL_n(\mathbf{Z})$ has infinitely many such characters for $n\geq 2.$ | math.OA | math |
INFINITE CHARACTERS ON GLn(Q), ON SLn(Z), AND
ON GROUPS ACTING ON TREES
BACHIR BEKKA
Abstract. Answering a question of J. Rosenberg from [Ros -- 89],
we construct the first examples of infinite characters on GLn(K)
for a global field K and n ≥ 2. The case n = 2 is deduced from
the following more general result. Let G a non amenable countable
subgroup acting on locally finite tree X. Assume either that the
stabilizer in G of every vertex of X is finite or that the closure
of the image of G in Aut(X) is not amenable. We show that
G has uncountably many infinite dimensional irreducible unitary
representations (π, H) of G which are traceable, that is, such that
the C ∗-subalgebra of B(H) generated by π(G) contains the algebra
of the compact operators on H. In the case n ≥ 3, we prove the
existence of infinitely many characters for G = GLn(R), where
n ≥ 3 and R is an integral domain such that G is not amenable. In
particular, the group SLn(Z) has infinitely many such characters
for n ≥ 2.
1. Introduction
is, the set of equivalence classes of irreducible unitary representations
standard Borel space exactly when G is virtually abelian, by results
of Glimm and Thoma (see [Gli -- 61] and [Tho -- 68]). So, unless G is
virtually abelian (in which case the representation theory of G is well
Let G be a countable discrete group and bG the unitary dual of G, that
of G. The space bG, equipped with a natural Borel structure, is a
understood), a description of bG is hopeless or useless. There are at
than bG:
least two other dual objects of G, which seem to be more accessible
• Thoma's dual space E(G), that is, the set of indecomposable
positive definite central functions on G;
• the space Char(G) of characters of G, that is, the space of
lower semi-continuous semi-finite (not necessarily finite) traces
t on the maximal C ∗-algebra C ∗(G) of G (see Subsection 2.1)
The author acknowledges the support by the ANR (French Agence Nationale de
la Recherche) through the projects Labex Lebesgue (ANR-11-LABX-0020-01) and
GAMME (ANR-14-CE25-0004).
1
2
BACHIR BEKKA
which satisfies the following extremality condition: every lower
semi-continuous semi-finite trace on C ∗(G) dominated by t on
the ideal of definition of t is proportional to t.
The space Char(G) parametrizes the quasi-equivalence classes of fac-
torial representations of C ∗(G) which are traceable ; recall that a
unitary representation π is factorial if the von Neumann algebra M
generated by π(G) is a factor and that a factorial representation π is
traceable if there exists a faithful normal (not necessarily finite) trace τ
on M and a positive element x ∈ C ∗(G) such that 0 < τ (π(x)) < +∞.
If this is the case, then t = τ ◦ π belongs to Char(G). Conversely, every
element of Char(G) is obtained in this way. Traceable representations
are also called normal representations.
Two traceable factorial representations π1 and π2 are quasi-equivalent
if there exists an isomorphism Φ : M1 → M2 such that Φ(π1(g)) =
π2(g) for all g ∈ G, where Mi is the factor generated by πi(G).
Observe that an irreducible unitary representation (π, H) of G is
traceable if and only if π(C ∗(G)) contains the algebra of compact op-
erators on H. The character associated to such a representation is given
by the usual trace on B(H) and so does not belong to E(G) whenever
H is infinite dimensional; in this case, the character is said to be of type
I∞, in accordance with the type classification of von Neumann algebras.
Observe also that two irreducible traceable representations of a group
G are quasi-equivalent if and only if they are unitarily equivalent.
Thoma's dual space E(G) is a subspace of Char(G) and classifies the
quasi-equivalence classes of the factorial representations π of C ∗(G) for
which the factor M generated by π(G) is finite, that is, such that the
trace τ on M takes only finite values (for more detail on all of this, see
Chapters 6 and 17 in [Dix -- 77]).
Thoma's dual space E(G) was determined for several examples of
countable groups G, among them G = GLn(K) or G = SLn(K)
for an infinite field K and n ≥ 2 ([Kiri -- 65]; see also [PeT -- 16]), and
G = SLn(Z) for n ≥ 3 ([Bek -- 07]); a procedure is given in [How -- 77,
Proposition 3] to compute E(G) when G is a nilpotent finitely gener-
ated group.
The space Char(G) has been described for some amenable groups G:
• when G is nilpotent, we have E(G) = Char(G) (see [CaM -- 84,
Theorem 2.1]);
• the space Char(G) is determined in [Guic -- 63] for the Baumslag-
Solitar group BS(1, 2) and in [VeK -- 91] for the infinite symmet-
ric group;
INFINITE CHARACTERS
3
• for G = GLn(K) and n ≥ 2, it is shown in [Ros -- 89] that
E(G) = Char(G) in the case where K an algebraic extension of
a finite field. (Observe that GLn(K) is amenable if and only if
K an algebraic extension of a finite field; see Proposition 9 in
[HoR -- 89] or Proposition 11 below.)
J. Rosenberg asked in [Ros -- 89, Remark after Th´eor`eme 1] whether
there exists an infinite character on G = GLn(K), that is, whether
Char(G) 6= E(G), for a field K which is not an algebraic extension
of a finite field. We will show below that the answer to this question
is positive, by exhibiting as far we know the first examples of such
characters. The case where n = 2 and K is a global field (see below)
will be deduced from a general result concerning groups acting on trees,
which we now state.
Recall that a graph X is locally finite if every vertex on X has only
finitely many neighbours. In this case, the group Aut(X) of automor-
phisms of X, equipped with the topology of pointwise convergence, is
a locally compact group for which the vertex stabilizers are compact.
Concerning the notion of weakly equivalent representations, see Chap-
ters 3 and 18 in [Dix -- 77] (see also Section 2.1).
Theorem 1. Let X be a tree and G a countable subgroup acting on X.
Assume that
(a) either G is not amenable and the stabilizer in G of every vertex
of X is finite, or
(b) X is locally finite and the closure of the image of G in Aut(X)
is not amenable.
There exists an uncountable family (πt)t of irreducible unitary represen-
tations of G with the following properties: πt is infinite dimensional, is
traceable and is not weakly equivalent to πt′ for t′ 6= t.
Recall that a global field is a finite extension of either the field Q of
rational numbers or of the field Fp(T ) of rational functions in T over
the finite field Fp (see Chapter III in [Wei -- 67]).
Corollary 2. Let G be either
(i) GL2(K) or SL2(K) for a global field K, or
(ii) SL2(Z), or
(iii) Fn, the free non abelian group over n ∈ {2, . . . , +∞} genera-
tors.
There exists an uncountable family (πt)t of unitary representations of
G with the properties from Theorem 1; moreover, in case G = Fn, the
representations πt are all faithful.
4
BACHIR BEKKA
Turning to the case n ≥ 3, we prove a result for G = GLn(R) or
G = SLn(R), valid for every integral domain R such that G is not
amenable.
Theorem 3. Let R be a countable unital commutative ring which is
an integral domain; in case the characteristic of R is positive, assume
that the field of fractions of R is not an algebraic extension of its prime
field. For n ≥ 3, let G = GLn(R) or G = SLn(R). There exists an
infinite dimensional irreducible unitary representation of G which is
traceable.
In the case where R is a field or the ring of integers, we can even pro-
duce infinitely many non equivalent representations as in Theorem 3.
Corollary 4.
(i) For n ≥ 3, let G = GLn(K) for a countable
field K which is not an algebraic extension of a finite field.
There exists an uncountable family (πt)t of pairwise non equiv-
alent infinite dimensional irreducible unitary representations of
G which are traceable. Moreover, the representations πt all have
a trivial central character, that is, the πt's are representations
of PGLn(K).
(ii) Let G = SLn(Z) for n ≥ 3. There exists an infinite family of
pairwise non equivalent infinite dimensional irreducible unitary
representations of G which are traceable.
The methods of proofs of Theorem 1 and Theorem 3 are quite dif-
ferent in nature:
• the proof of Theorem 1 is based on properties of a remark-
able family of unitary representations of groups acting on trees
constructed in [JuV -- 84] and used to show their K-theoretic
amenability, a notion which originated from [Cun -- 83]
in the
case of free groups;
• the traceable representations we construct in Theorem 3 are
induced representations from suitable subgroups. The case n ≥
4 uses the existence of appropriate subgroups of GLn(R) with
Kazhdan's Property (T).
Remark 5. For a group G as in Theorem 1 or Theorem 1, our results
show that the set Char(G) contains characters of type I∞.
For, say, G = GLn(Q), we do not know whether Char(G) contains
characters of type II∞, that is, characters for which the corresponding
factorial representation generates a factor of type II∞.
INFINITE CHARACTERS
5
This paper is organized as follows. In Section 2, we establish some
preliminary facts which are necessary to the proofs of our results. Sec-
tion 3 is devoted to the proofs of Theorem 1 and Corollary 2; Theorem 3
and Corollary 4 are proved in Section 4.
2. Some preliminary results
.
2.1. C ∗-algebras. Let G be a countable group. Recall that a unitary
representation of G is a homomorphism π : G → U(H) from G to the
unitary group of a complex separable Hilbert space H. From now on,
we will simply write representation of G instead of "unitary repre-
sentation of G".
Every representation (π, H) of G extends naturally to a ∗-representation,
denoted again by π, of the group algebra C[G] by bounded operators
on H.
Recall that the maximal C ∗-algebra C ∗(G) of G is the completion of
C[G] of G with respect to the norm
f 7→ sup
kπ(f )k,
π∈Rep(G)
where Rep(G) denotes the set of representations (π, H) of G in a sep-
arable Hilbert space H.
We can view G as subset of C[G] and hence as a subset of C ∗(G).
The C ∗-algebra C ∗(G) has the following universal property: every rep-
resentation (π, H) of G extends to a unique representation (that is, ∗-
homomorphism) π : C ∗(G) → B(H). The correspondence G → C ∗(G)
is functorial: every homomorphism ϕ : G1 → G2 between two count-
able groups G1 and G2 extends to a unique morphism
ϕ∗ : C ∗(G1) → C ∗(G2)
of C ∗-algebras. In particular, given a subgroup H of a group G, the
injection map i : H → G extends to a morphism i∗ : C ∗(H) → C ∗(G);
the map i∗ is injective and so C ∗(H) can be viewed naturally as a
subalgebra of C ∗(G):
indeed, this follows from the fact that every
representation σ of H occurs as subrepresentation of the restriction
to H of some representation π of G (one may take as π the induced
representation IndG
H σ, as shown below in Proposition 9).
The following simple lemma will be one of our tools in order to show
that π(C ∗(G)) contains a non-zero compact operator for a representa-
tion π of G.
6
BACHIR BEKKA
Let A be a C ∗-algebra. Recall that a representation π : A → B(H)
weakly contains another representation ρ : A → B(K) if
kρ(a)k ≤ kπ(a)k
for all a ∈ A,
or, equivalently, ker π ⊂ ker ρ (see Chapter 3 in [Dix -- 77]). Two repre-
sentations π and ρ are weakly equivalent if π weakly contains ρ and ρ
weakly contains π, that is, if ker π = ker ρ.
Lemma 6. Let A be a C ∗-algebra and π : A → B(H) a representation
of A. Assume that H contains a non-zero finite dimensional π(A)-
invariant subspace K and that the restriction π1 of π to K is not weakly
contained in the restriction π0 of π to the orthogonal complement K⊥.
Then π(A) contains a non-zero compact operator.
Proof. The ideal ker π0 is not contained in ker π1, since π1 is not weakly
contained in π0. Hence, there exists a ∈ A with π0(a) = 0 and π1(a) 6=
0. Then π(a) = π1(a) has a finite dimensional range and is non-zero. (cid:3)
Knowing that a representation of A contains in its image a non-
zero compact operator, the following lemma enables us to construct an
irreducible representation of A with the same property.
Lemma 7. Let A be a C ∗-algebra and π : A → B(H) a representation
of A in a separable Hilbert space H Let a ∈ A be such that π(a) is a
non-zero compact operator. Then there exists an irreducible subrepre-
sentation σ of π such that σ(a) is a compact operator and such that
kσ(a)k = kπ(a)k.
Ω πωdµ(ω) of irre-
ducible representations πω; thus, we can find a probabilility measure µ
on a standard Borel space Ω, a measurable field ω → πω of irreducible
representations of A in a measurable field ω → Hω of separable Hilbert
Ω Hωdµ(ω)
Proof. We can decompose π as a direct integral R ⊕
spaces on Ω, and a Hilbert space isomorphism U : H →R ⊕
such that
Uπ(x)U −1 =Z ⊕
Ω
πω(x)dµ(ω).
for all x ∈ A
R ⊕
(see [Dix -- 77, §8.5]). Without loss of generality, we will identify π with
Ω πωdµ(ω).
Let a ∈ A be such that π(a) is a non-zero compact operator. Since
kσ(a∗a)k = kσ(a)k2 for every representation σ of A, upon replacing a
by a∗a, we can assume that a is a positive element of A. So π(a) is a
positive selfadjoint compact operator on H with π(a) 6= 0.
INFINITE CHARACTERS
7
There exists an orthonormal basis (Fn)n≥1 of H =R ⊕
Ω Hωdµ(ω) con-
sisting of eigenvectors of π(a), with corresponding eigenvalues (λn)n≥1,
counted with multiplicities. For every ω ∈ Ω and every n ≥ 1, we have
(∗)
πω(a)(Fn(ω)) = λnFn(ω).
Let n0 ≥ 1 be such that λn0 = max{λn n ≥ 1}. Then kπ(a)k = λn0.
Set
Ω0 = {ω ∈ Ω Fn0(ω) 6= 0.}.
Since Fn0 6= 0, we have µ(Ω0) > 0. We claim that Ω0 is a finite subset
of Ω consisting of atoms of µ. Indeed, assume by contradiction that is
not the case. Then there exists an infinite sequence (Ak)k of pairwise
disjoint Borel subsets of Ω0 with µ(Ak) > 0. Observe that 1AkFn0 is a
non-zero vector in H and that h1AkFn0, 1AlFn0i = 0 for every k 6= l.
Moreover, we have
π(a)(1AkFn0) =Z ⊕
= λn0Z ⊕
Ak
Ak
= λn01AkFn0.
πω(a)(Fn0(ω))dµ(ω)
Fn0(ω)dµ(ω)
Since π(a) is a compact operator and λn0 6= 0, this is a contradiction.
Let ω0 ∈ Ω0 be such that µ({ω0}) > 0. We claim that the linear span
of {Fn(ω0) n ≥ 1} is dense in Hω0. Indeed, let v ∈ Hω0 be such that
hv Fn(ω0)i = 0
for all n ≥ 1.
Let F = 1ω0 ⊗ v ∈ H be defined by F (ω0) = v and F (ω) = 0 for
ω 6= ω0. Then hF Fni = 0 for all n ≥ 1. Hence, F = 0, that is, v = 0,
since (Fn)n≥1 is a basis of H.
By (∗), Fn(ω0) is an eigenvector of πω0(a) with eigenvalue λn for
every n ≥ 1 such that Fn(ω0) 6= 0. Since {Fn(ω0) n ≥ 1} is a total
subset of Hω0, it follows that there exists a basis of Hω0 consisting of
eigenvectors of πω0(a). As
lim
n→∞
λn = 0
(in case the sequence (λn)n≥1 is infinite), it follows that πω0(a) is a
compact operator on Hω0. Moreover, we have
kπω0(a)k = max{λn n ≥ 1} = λn0 = kπ(a)k.
Finally, an equivalence between πω0 and a subrepresentation of π is
provided by the unitary linear map
Hω0 → H, v 7→ 1ω0 ⊗ v.
8
BACHIR BEKKA
(cid:3)
2.2. Induced representations of groups. In the sequel, we will of-
ten consider group representations which are induced representations.
Let G be a countable group, H a subgroup of G and (σ, K) a represen-
tation of H. Recall that the induced representation IndG
H σ of G may
be realized as follows. Let H be the Hilbert space of maps f : G → K
with the following properties
(i) f (hx) = σ(h)f (x) for all x ∈ G, h ∈ H;
(ii) Px∈H\G kf (x)k2 < ∞. (Observe that kf (x)k only depends on
the coset of x in H\G.)
The induced representation π = IndG
H σ is given on H by right trans-
lation:
(π(g)f )(x) = f (xg)
for all g ∈ G, f ∈ H and x ∈ G.
Recall that the commensurator of H in G is the subgroup, denoted
by CommG(H), of the elements g ∈ G such that gHg−1 ∩ H is of finite
index in both H and g−1Hg.
The following result appeared in [Mac -- 51] in the case where σ is
of dimension 1 and was extended to its present form in [Kle -- 61] and
[Cor -- 75].
Theorem 8. Let G be a countable group and H a subgroup of G such
that CommG(H) = H.
(i) For every finite dimensional irreducible representation σ of H,
the induced representation IndG
H σ is irreducible.
(ii) Let σ1 and σ2 be non equivalent finite dimensional irreducible
H σ2
representations of H. The representations IndG
are non equivalent.
H σ1 and IndG
We will need to decompose the restriction to a subgroup of an in-
duced representation IndG
H σ as in Theorem 8. For g ∈ G, we denote
by σg the representation of g−1Hg defined by σg(x) = σ(gxg−1) for
x ∈ g−1Hg.
For the convenience of the reader, we give a short and elemen-
tary proof of the following special case of the far more general result
[Mac -- 52, Theorem 12.1].
Proposition 9. Let G be a countable group, H, L subgroups of G and
(σ, K) a representation of H. Let S be a system of representatives for
the double coset space H\G/L. The restriction πL to L of the induced
INFINITE CHARACTERS
9
representation π = IndG
H σ is equivalent to the direct sum
IndL
s−1Hs∩L(σss−1Hs∩L)
Ms∈S
Proof. Let H be the Hilbert space of π, as described above. For every
s ∈ S, let Hs be the space of maps f ∈ H such that f = 0 outside the
double coset HsL. We have an orthogonal L-invariant decomposition
H =Ms∈S
Hs.
Fix s ∈ S. The Hilbert space H′
s of IndL
s−1Hs∩L(σss−1Hs∩L) consists
• f (tx) = σ(sts−1)f (x) for all t ∈ s−1Hs ∩ L, x ∈ L;
of the maps f : L → K such that
• Px∈s−1Hs∩L\L kf (x)k2 < ∞.
Define a linear map U : Hs → H′
Uf (x) = f (sx)
f ∈ Hs, x ∈ L.
s by
for all
Observe that, for t ∈ s−1Hs ∩ L, x ∈ L and f ∈ Hs, we have
Uf (tx) = f (stx) = f ((sts−1)sx) = σ(sts−1)f (sx) = σs(t)Uf (x)
and that
Xx∈s−1Hs∩L\L
kUf (x)k2 = Xx∈s−1Hs∩L\L
kf (sx)k2 = Xy∈H\G
kf (y)k2 < ∞,
so that Uf ∈ H′
map U is invertible, with inverse given by
s and U is an isometry. It is easy to check that the
U −1f (y) =(σ(h)f (x)
0
if y = hsx ∈ HsL
otherwise ,
for f ∈ H′
IndL
s. Moreover, U intertwines the restriction of πL to Hs and
s−1Hs∩L(σss−1Hs∩L) : for g, x ∈ L and f ∈ H′
s, we have
(cid:0)Uπ(g)U −1f(cid:1) (x) =(cid:0)π(g)U −1f(cid:1) (sx)
= (U −1f (sxg)
= f (xg)
=(cid:0)IndL
s−1Hs∩L(σss−1Hs∩L)(g)f(cid:1) (x)
(cid:3)
We will need the following elementary lemma about induced repre-
sentations containing a finite dimensional representation. Recall that
a representation π of a group G contains another representation σ of
G if σ is equivalent to a subrepresentation of π. Recall also that, if π is
10
BACHIR BEKKA
finite dimensional representation of a group G, then π ⊗ ¯π contains the
trivial representation 1G, where ¯π is the conjugate representation of π
and π ⊗ ρ denotes the (inner) tensor product of the representations π
and ρ (see [BHV -- 08, Proposition A. 1.12]).
Proposition 10. Let G be a countable group, H a subgroup of G, and
σ a representation of H. Assume that the induced representation IndG
H σ
contains a finite dimensional representation of G. Then H has finite
index in G.
Proof. By assumption, π := IndG
resentation σ. Hence, π ⊗ ¯π contains 1G. On the other hand,
H σ contains a finite dimensional rep-
π ⊗ ¯π = (IndG
H σ) ⊗ ¯π
is equivalent to IndG
H(ρ), where ρ = σ⊗(¯πH); see [BHV -- 08, Proposition
E. 2.5]. So, there exists a non-zero map f : G → K in the Hilbert space
of IndG
H(ρ) which is G-invariant, that is, such that f (xg) = f (x) for all
g, x ∈ G. This implies that the L2-function x 7→ kf (x)k2 is constant
on H\G. This is only possible if H\G is finite.
(cid:3)
2.3. Amenability. Let G be a topological group and UCB(G) the
Banach space of the left uniformly continuous bounded functions on G,
equipped with the uniform norm. Recall that G is amenable if there
exists a G-invariant mean on UCB(G) (see Appendix G in [BHV -- 08]).
The following proposition characterizes the integral domains R for
which GLn(R) or SLn(R) is amenable; the proof is an easy extension
of the proof given in Proposition 9 in [HoR -- 89] for the case where R is
a field.
Proposition 11. Let R be a countable unital commutative ring which
is an integral domain. Let K be the field of fractions of R and G =
GLn(R) or G = SLn(R) for an integer n ≥ 2. The following properties
are equivalent:
(i) G is not amenable.
(ii) K is not an algebraic extension of a finite field.
(iii) R contains Z if the characteristic of K is 0 or the polynomial
ring Fp[T ] if the characteristic of K is p > 0.
Proof. Assume that K is an algebraic extension of a finite field Fq.
Then K = Sm Km for an increasing family of finite extensions Km
of Fq; hence, GLn(K) = Sm GLn(Km) is the inductive limit of the
finite and hence amenable groups GLn(Km); it follows that GLn(K) is
amenable and therefore GLn(R) and SLn(R) are amenable. This shows
that (i) implies (ii).
INFINITE CHARACTERS
11
Assume that (ii) holds. If the characteristic of K is 0, then K con-
tains Q and hence R contains Z. So, we can assume that the charac-
teristic of K is p > 0. We claim that R contains an element which is
not algebraic over the prime field Fp. Indeed, otherwise, every element
in R is algebraic over Fp. As the set of elements in K which are alge-
braic over Fp is a field, it would follow that the field fraction field K is
algebraic over Fp. This contradiction shows that (ii) implies (iii).
Assume that (iii) holds. Then SLn(R) contains a copy of SL2(Z) or a
copy of SL2(Fp[T ]). It is well-known that both SL2(Z) and SL2(Fp[T ])
contain a subgroup which is isomorphic to the free group on two gen-
erators. Therefore, G is not amenable and so (iii) implies (i).
(cid:3)
Let G be a locally compact group, with Haar measure m. Recall
that the amenability of G is characterized by the Hulanicki-Reiter the-
orem (see [BHV -- 08, Theorem G.3.2]): G is amenable if and only if
the regular representation (λG, L2(G, m)) weakly contains the trivial
representation 1G, where m is Haar measure on G; when this is the
case, λG weakly contains every representation of G
The following result shows the amenability of G can be detected by
the restriction of λG to a dense subgroup; for a more general result,
see [Guiv -- 80, Proposition 1] or [Bek -- 16, Theorem 5.5].
Proposition 12. Let G be a locally compact group and G a countable
dense subgroup of G. Assume that the restriction to G of the regular
representation λG of G weakly contains the trivial representation 1G.
Then G is amenable.
Proof. By assumption, there exists a sequence (fn)n in L2(G, m) with
kfnk = 1 such that
lim
n
kλG(g)fn − fnk = 0
for all
g ∈ G.
Then, since fn(g−1x) − fn(x) ≤ fn(g−1x) − fn(x) for g, x ∈ G,
we have
(∗)
kλG(g)fn − fnk = 0
for all
g ∈ G.
lim
n
Set ϕn :=pfn. Then ϕn ≥ 0 andRG ϕndm = 1. Every ϕn defines a
mean Mn : f 7→RG f ϕndm on UCB(G). Let M be a limit of (Mn)n
for the weak-*-topology on the dual space of UBC(G). It follows from
(∗) that M is invariant under G. Since, for every f ∈ UCB(G), the
map
G → UCB(G),
g 7→g f
is continuous (where gf denotes left translation by g ∈ G), it follows
that M is invariant under G. Hence, G is amenable.
12
BACHIR BEKKA
2.4. Special linear groups over a subring of a field. We will
use the following elementary lemma about subgroups of SLn(K) which
stabilize a line in Kn.
Lemma 13. For an infinite field K and n ≥ 2, let L be a subgroup of
SLn(K) which stabilizes a line ℓ in Kn. Then L ∩ SLn(R) has infinite
index in SLn(R) for every infinite unital subring R of K.
Proof. Let {v1, . . . , vn} be a basis of Kn with ℓ = Kv1. Fix i, j ∈
{1, . . . , n} with i 6= j and, for λ ∈ K, let Eij(λ) be the corresponding
elementary matrix in SLn(K), that is,
Eij(λ) = In + λ∆ij,
where ∆ij denotes the matrix with 1 at the position (i, j) and 0 other-
wise.
For every l = 1, . . . , n, let ϕl : K → K be defined by
Eij(λ)(v1) =
nXi=1
ϕl(λ)vi
for λ ∈ K.
Every ϕl is a polynomial function (in fact, an affine function) on K
and, for l = 2, . . . , n, we have ϕl(λ) = 0 for every λ ∈ K such that
Eij(λ) ∈ L.
Assume, by contradiction, that L∩SLn(R) has finite index in SLn(R)
for an infinite subring R of K. Then the subgroup
Li,j(R) := L ∩ {Eij(λ) λ ∈ R}
has finite index in the subgroup {Eij(λ) λ ∈ R} of SLn(R). In partic-
ular, Li,j(R) is infinite. It follows that ϕl has infinitely many roots in
K and hence that ϕl = 0, for every l = 2, . . . , n. Therefore, every ele-
mentary matrix Eij(λ) fixes the line ℓ, for i, j ∈ {1, . . . , n} and λ ∈ K.
Since SLn(K) is generated by elementary matrices, it follows that every
matrix in SLn(K) fixes the line ℓ; this of course is impossible.
(cid:3)
3. Proofs of Theorem 1 and Corollary 2
3.1. Proof of Theorem 1. Let X be a tree, with X 0 the set of vertices
and X 1 the set of edges of X. Let G be a locally compact group acting
on X.
Julg and Valette constructed in [JuV -- 84] (see also [Szw -- 91] and
[Jul -- 15]) a remarkable family of representations (πt)t∈[0,1] of G, all de-
fined on ℓ2(X 0), with the following properties:
INFINITE CHARACTERS
13
(i) π0 is the natural representation of G on ℓ2(X 0) and π1 is equiv-
alent to 1G ⊕ ρ1, where ρ1 is the natural representation of G on
ℓ2(X 1);
(ii) for every t ∈ [0, 1], there exists a bounded operator Tt on ℓ2(X 0)
with inverse T −1
defined on the subspace of functions of X 0
with finite support such that πt(g) := T −1
t π0(g)Tt extends to
a unitary operator on ℓ2(X 0) for every g ∈ G; so, a unitary
representation g 7→ πt(g) of G is defined on ℓ2(X 0);
t
(iii) πt(g) − π0(g) is a finite-rank operator on ℓ2(X 0), for every t ∈
[0, 1] and g ∈ G;
(iv) we have
hπt(g)T −1
t δx T −1
t δyi = td(gx,y),
for every t ∈ (0, 1), g ∈ G and x, y ∈ X 0, where d denotes the
natural distance on X 0;
(v) the map
[0, 1] → R+, t 7→ kπt(g) − π0(g)k
is continuous for every g ∈ G.
(Our representation πt is g 7→ Uλρλ(g)U −1
representation ρλ and the operator Uλ appearing in §2 of [JuV -- 84].)
t with λ = − log t, for the
Let G be a countable group acting on X. Assume that
(a) either G is not amenable and the stabilizer in G of every vertex
of X is finite or
(b) X is locally finite and the closure of the image of G in Aut(X)
is not amenable.
Set G = G in case (a) and let G be the closure of G in Aut(X) in case
(b). Let (πt)t∈[0,1] be the family of representations of G as above.
• First step. For every a ∈ C ∗(G) and every t ∈ [0, 1], the operator
πt(a) − π0(a) is compact and the map
[0, 1] → R+, t 7→ kπt(a) − π0(a)k
is continuous.
Indeed, this follows from Properties (iii) and (v) of the family (πt)t
and from the fact that C[G] is dense in C ∗(G).
• Second step. The restriction π0G of π0 to G does not weakly
contain the trivial representation 1G.
Indeed, the representation π0 of G is equivalent to the direct sum
⊕s∈T λG/Gs, where S is a system of representatives for the G-orbits
in X 0 and Gs is the stabilizer in G of s ∈ S. Since Gs is compact
(and even finite in case (a)) and hence amenable, λG/Gs = IndG
Gs 1Gs
14
BACHIR BEKKA
is weakly contained in the regular representation λG of G and so π0 is
weakly contained in λG. Hence, π0 does not weakly contain the triv-
ial representation 1G in case (a). In case (b), the claim follows from
Proposition 12, since G is not amenable and G is dense in G.
• Third step. There exists an element a ∈ C ∗(G) and 0 ≤ t0 < 1
with the following properties: πt0(a) = 0, πt(a) is a non zero compact
operator for every t ∈ (t0, 1], and the map
[t0, 1] → R+, t 7→ kπt(a)k
is continuous.
Indeed, by the second step, there exists a ∈ C ∗(G) such that π0(a) =
0 and 1G(a) 6= 0. Therefore, π1(a) 6= 0 and πt(a) = πt(a) − π0(a) for
every t ∈ [0, 1] and so the claim follows from the first step.
• Fourth step. Let a ∈ C ∗(G) and 0 ≤ t0 < 1 be as in the third step.
There exists an irreducible infinite dimensional subrepresentation σt
of πt such that σt(a) is a compact operator and such that kσt(a)k =
kπt(a)k for every t ∈ (t0, 1).
Indeed, it follows from the third step and Lemma 7 that πtG con-
tains an irreducible subrepresentation σt such that σt(a) is a compact
operator with kσt(a)k = kπt(a)k. It remains to show that σt is infinite
dimensional for every t ∈ (t0, 1).
Assume, by contradiction, σt is finite dimensional for some t ∈ (t0, 1).
Since G is dense in G, the closed subspace Kt of ℓ2(X 0) defining σt is
invariant under G and so σt is the restriction to G of a dimensional
subrepresentation of πt, again denoted by σt. On the one hand, G acts
properly on X 0, since the stabilizers of vertices are compact (and even
finite in case (a)). So, we have
lim
g→+∞:g∈G
d(gx, x) = 0
for all x ∈ X 0.
It follows from Property (iv) of the family (πt)t that πt (and hence σt)
is a C0-representation, that is,
lim
hπt(g)v wi = 0
g→+∞:g∈G
for every v, w ∈ ℓ2(X 0). On the other hand, since σt is finite dimen-
sional, σt ⊗σt contains 1G. As G is not compact, this is a contradiction
to the fact that σt is a C0-representation.
• Fifth step. There exists uncountably many real numbers t ∈ (t0, 1)
such that the subrepresentations σt of πtG as in the fourth step are
pairwise non weakly equivalent.
INFINITE CHARACTERS
15
Indeed, by the third step, the function f : t 7→ kπt(a)k is continuous
on [t0, 1], with f (t0) = 0 and f (1) > 0. So, the range of f contains a
whole interval. Let t, s ∈ (t0, 1) be such that f (t) 6= f (s). Then
kσt(a)k = kπt(a)k = f (t) 6= f (s) = kπs(a)k = kσs(a)k,
and so σt and σs are not weakly equivalent.
3.2. Proof of Corollary 2. The following remarks show how Corol-
lary 2 follows from Theorem 1.
(i) Let K be global field K. Choose a non trivial discrete valuation
v : K∗ → Z. The completion of K at v is a non archimedean local
field Kv. The tree Xv associated to v (see Chapter II in [Ser -- 80]) is a
locally finite regular graph. The group G = GL2(K) acts as a group of
automorphisms of Xv, with vertex stabilizers conjugate to GL2(Ov∩K),
where Ov is the compact subring of the integers in Kv. The closure of
the image of G in Aut(Xv) coincides with PGL2(Kv) and is therefore
non amenable. A similar remark applies to G = SL2(K).
(ii) As is well-known, the group G = SL2(Z) is an amalgamated product
Z/4Z∗Z/2ZZ/6Z. It follows that G acts on a tree with vertices of valence
2 or 3 with vertex stabilizers of order 4 or 6 (see Chapter I, Examples
4.2. in [Ser -- 80])
(iii) The free non abelian group F2 acts freely on its Cayley graph X,
which is a 4-regular tree. It follows that Fn acts freely on X for every
n ∈ {2, . . . , +∞}. Observe that, in this case, the representations πt
and σt as in the proof of Theorem 1 are faithful for t 6= 1 (since there
are even C0-representations).
(cid:3)
4. Proofs of Theorem 3 and Corollary 4
4.1. Proof of Theorem 3. Let R be a countable unital commutative
ring which is an integral domain and K its field of fractions. In case
the characteristic of K is positive, assume that K is not an algebraic
extension of its prime field.
Let n ≥ 3 and G = GLn(R). We consider the natural action of G on
the projective space P(Kn). Let ℓ0 = Ke1 ∈ P(Kn) be the line defined
by the first unit vector e1 in Kn. The stabilizer of ℓ0 in G is
H =(cid:18)R×
0 GLn−1(R)(cid:19) .
Rn−1
16
BACHIR BEKKA
Let σ be a finite dimensional representation of H and π := IndG
H σ. We
claim that π is irreducible and that π(C ∗(G)) contains a non zero com-
pact operator. For the proof of this claim, we have to treat separately
the cases n = 3 and n ≥ 4.
4.1.1. Case n = 3. • First step. We claim that gHg−1∩H is amenable,
for every g ∈ G \ H.
Indeed, let g ∈ G \ H. Then ℓ0 and gℓ0 are distinct lines in Kn and
are both stabilized by gHg−1 ∩ H. Hence, gHg−1 ∩ H is isomorphic to
a subgroup of the solvable group
K∗
0 K
0 K∗ K
0
0 K∗
and is therefore amenable.
• Second step. We claim that the representation π is irreducible.
Indeed, in view of Theorem 8, we have to show that CommG(H) = H.
Let g ∈ G \ H. On the one hand, gHg−1 ∩ H is amenable, by the first
step. On the other hand, H is non amenable, by Proposition 11. This
implies that gHg−1 ∩ H is not of finite index in H and so g is not in
the commensurator of H in G.
• Third step. We claim that the C ∗-algebra π(C ∗(G)) contains a
non-zero compact operator.
Indeed, let S be a system of representatives for the double cosets
space H\G/H with e ∈ S. By Proposition 9, the restriction πH of π
to H is equivalent to the direct sum
Ms∈S
IndH
s−1Hs∩H(σss−1Hs∩H) = σ ⊕ Ms∈S\{e}
IndH
s−1Hs∩H(σss−1Hs∩H)
Let s ∈ S \ {e}. By the first step, s−1Hs ∩ H is amenable and hence
σss−1Hs∩H is weakly contained in the regular representation λs−1Hs∩H of
s−1Hs ∩ H, by the Hulanicki-Reiter theorem. By continuity of induc-
tion (see [BHV -- 08, Theorem F.3.5]), it follows that IndH
is weakly contained in the regular representation λH of H. Therefore,
s−1Hs∩H(σss−1Hs∩H)
π0 := Ms∈S\{e}
IndH
s−1Hs∩H(σss−1Hs∩H)
is weakly contained in λH. It follows that π0 does not weakly contain σ;
indeed, assume by contradiction that σ is weakly contained in π0. Then
λH ⊗ λH , which is a multiple of λH , weakly contains σ ⊗ σ. However,
since σ is finite dimensional, σ ⊗ σ contains 1H. Hence, 1H is weakly
INFINITE CHARACTERS
17
contained in λH and this is a contradiction to the non amenability of
H.
It follows from Lemma 6 that π(C ∗(H)) contains a non-zero compact
operator. Since C ∗(H) can be viewed a subalgebra of C ∗(G), the claim
is proved for G = GL3(R).
4.1.2. Case n ≥ 4. For every unital subring R′ of R, set
L(R′) :=(cid:18)1
0 SLn−1(R′)(cid:19) ,
0
which is a subgroup of H isomorphic to SLn−1(R′).
• First step. Let g0 ∈ G \ H and R′ an infinite unital subring of R.
We claim that g0Hg−1
0 ∩ L(R′) has infinite index in L(R′).
Indeed, the group L := g0Hg−1
0 ∩ L(R′) stabilizes the two lines ℓ0
and g0ℓ0. Let V be the linear span of the n − 1 unit vectors e2, . . . , en.
Denote by ℓ the projection on V of the line g0ℓ0, parallel to ℓ0. As
gℓ0 6= ℓ0, we have ℓ 6= {0}. Moreover, L stabilizes ℓ, since L stabilizes
ℓ0 and V . So, identifying L(R′) with the group SLn−1(R′), we see can
view L as a subgroup of SLn−1(K) which stabilizes a line in Kn−1.
Lemma 13 shows that L has infinite index in SLn−1(R′), as claimed.
• Second step. We claim that the representation π is irreducible. In
view of Theorem 8, it suffices to show that CommG(H) = H.
Let g0 ∈ G \ H. By the first step, g0Hg−1
0 ∩ L(R) has infinite index
0 ∩ H has infinite index in H, since L(R) is a
in L(R); hence, g0Hg−1
subgroup of H.
• Third step. We claim that π(C ∗(G)) contains a non-zero compact
operator.
Indeed, since K is not an algebraic extension over its prime field, R
contains a subring R′ which is a copy Z or a copy of the polynomial
ring Fp[T ], by Proposition 11. The corresponding subgroup
L := L(R′)
of G is isomorphic to SLn−1(Z) or SLn−1(Fp[T ]). Observe that L is a
lattice in the locally group G = SLn−1(R) or G = SLn−1(Fp((T −1))),
where Fp((T −1)) is the local field of Laurent series over Fp. Since
n − 1 ≥ 3, the group G and hence L has Kazhdan's Property (T); see
[BHV -- 08, §. 1.4, 1.7].
Let S be a system of representatives for the double cosets space
H\G/H with e ∈ S. By Proposition 9, the restriction πL to L of π is
18
BACHIR BEKKA
equivalent to the direct sum σL ⊕ π0, where
π0 := Ms∈S\{e}
IndL
s−1Hs∩L(σss−1Hs∩L).
We claim that π0 does not weakly contain σL. Indeed, assume by con-
tradiction that π0 weakly contains σL. Since σ is finite dimensional
and L has Property (T), it follows that π0 contains σL (see [BHV -- 08,
Theorem 1.2.5]). Therefore, IndL
s−1Hs∩L(σss−1Hs∩L) contains a subrep-
resentation of σL for some s ∈ S \ {e}. Hence, s−1Hs ∩ L has finite
index in L, by Proposition 10. Since L = L(R′) for an infinite unital
subring R′ of R, this is a contradiction to the first step.
As in the proof for the case n = 3, we conclude that π(C ∗(G))
contains a non-zero compact operator.
This proves Theorem 3 for G = GLn(R) when n ≥ 3. The case
G = SLn(R) is proved in exactly the same way.
4.2. Proof of Corollary 4. For n ≥ 3, let G = GLn(R) for a ring R
as above. The irreducible traceable representations of G constructed
in the proof of Theorem 3 are of the form π = IndG
H σ for a finite
dimensional representation of the subgroup H = (cid:18)R×
0 GLn−1(R)(cid:19).
Rn−1
Observe that, π = IndG
contains Z.
H σ is trivial on the center Z of G, since H
By Theorem 8, there are infinitely (respectively, uncountably) many
non equivalent such representations π, provided there exists infinitely
(respectively, uncountably) non equivalent finite dimensional irreducible
representations of H. This will be the case if GLn−1(R) ⋉ Rn−1, which
is a quotient of H, has infinitely (respectively, uncountably) many non
equivalent finite dimensional irreducible representations.
(i) Assume that R = K is an infinite field. It is easy to show that the
finite dimensional irreducible representations of GLn−1(K) ⋉ Kn−1 are
all of the form
(cid:18)∗ ∗
0 A(cid:19) 7→ χ(det A),
A ∈ GLn−1(K)
for some χ in the unitary dual cK∗ of K∗; as K∗ is infinite, cK∗ is a
compact infinite group and is therefore uncountable.
(ii) Assume that R = Z.
• Case n = 3. The free group F2 is a subgroup of finite index in
GL2(Z). There exists uncountably many unitary characters (that is
one-dimensional unitary representations) of F2. For every such unitary
INFINITE CHARACTERS
19
i
i
F2
character χ, the representation IndGL2(Z)
χ is finite dimensional and so
has a decomposition ⊕iσ(χ)
as a direct sum of finite dimensional irre-
ducible representations σ(χ)
of GL2(Z). One can choose uncountably
many pairwise non equivalent representations among the σ(χ)
's and we
obtain in this way uncountably many non equivalent finite dimensional
irreducible representations of GL2(Z) and hence of GL2(Z) ⋉ Z2.
• Case n ≥ 4. The group GLn−1(Z) ⋉Zn−1 has Kazhdan's property (T)
and so has at most countably many non equivalent finite dimensional
representations (see [Wan -- 75, Theorem 2.1]). There are indeed infin-
itely many such representations:
for every integer N ≥ 1, the finite
group
i
GN = GLn−1(Z/N Z)) ⋉ (Z/N Z)n−1
is a quotient of GLn−1(Z)⋉Zn−1; infinitely many representations among
the irreducible representations of the GN 's are pairwise non equivalent
when viewed as representations of GLn−1(Z) ⋉ Zn−1.
References
[Bek -- 07] B. Bekka. Operator-algebraic superrigidity for SLn(Z), n ≥ 3. Invent.
Math. 169 (2007), 401 -- 425.
[Bek -- 16] B. Bekka. Spectral rigidity of group actions on homogeneous spaces.
Prepint 2017, to appear in "Handbook of group actions, Volume III"
(Editors: L. Ji, A. Papadopoulos, S-T Yau); ArXiv 1602.02892.
[BHV -- 08] B. Bekka, P. de la Harpe, A. Valette. Kazhdan's Property (T). Cambridge
University Press 2008.
[CaM -- 84] A. Carey and W. Moran. Characters of nilpotent groups. Math. Proc.
Camb. Phil. Soc. 96 (1984), no. 1, 123 -- 137.
[Cor -- 75] L. Corwin. Induced representations of discrete groups. Proc. Amer.
Math. Soc. 47 (1975), 279 -- 287.
[Cun -- 83] J. Cuntz. K-theoretic amenability for discrete groups. J. Reine Angew.
Math. 344 (1983), 180 -- 195
[Dix -- 77] J. Dixmier. C ∗-algebras. North-Holland 1977.
[Gli -- 61]
[Guic -- 63] A. Guichardet. Caract`eres des alg`ebres de Banach involutives. Ann. Inst.
J. Glimm. Type I C*-algebras. Ann. of Math. (2) 73 (1961), 572 -- 612.
Fourier 13 (1963), 1 -- 81.
[Guiv -- 80] Y. Guivarc'h. Quelques propri´et´es asymptotiques des produits de matri-
ces al´eatoires. Lecture Notes Math. 774, 17-250, Springer, 1980.
[How -- 77] R. Howe. On representations of discrete, finitely generated, torsion-free,
nilpotent groups. Pacific J. Math. 73 (1977), no. 2, 281 -- 305.
[HoR -- 89] R. Howe, J. Rosenberg. The unitary representation theory of GL(n) of
an infinite discrete field. Israel J. Math. 67 (1989), 67 -- 81.
[Jul -- 15] P. Julg. A new look at the proof of K-theoretic amenability for groups
acting on trees. Bull. Belg. Math. Soc. Simon Stevin 22 (2015), 263 -- 269.
[JuV -- 84] P. Julg, A. Valette. K-theoretic amenability for SL2(Qp), and the action
on the associated tree. J. Funct. Anal. 58 (1984),194 -- 215.
20
BACHIR BEKKA
[Kiri -- 65] A.A. Kirillov. Positive definite functions on a group of matrices with
elements from a discrete field. Soviet. Math. Dokl. 6 (1965), 707 -- 709.
[Kle -- 61] A. Kleppner. On the intertwining number theorem. Proc. Amer. Math.
Soc. 12 (1961), 731 -- 733.
[Mac -- 51] G.W. Mackey. On induced representations of groups. Amer. J. Math. 73
(1951), 576 -- 592.
[Mac -- 52] G.W. Mackey. Induced representations of locally compact groups. I. Ann.
of Math. (2) 55 (1952), 101 -- 139.
[PeT -- 16] J. Peterson, A. Thom. Character rigidity for special linear groups. J.
reine angew. Math. 716 (2016), 207 -- 228.
[Ros -- 89] J. Rosenberg. Un compl´ement `a un th´eor`eme de Kirillov sur les caract`eres
de GL(n) d'un corps infini discret. C.R. Acad. Sci. Paris 309 (1989), S´erie
I, 581 -- 586.
J-P. Serre. Trees. Springer-Verlag, Berlin-New York, 1980.
[Ser -- 80]
[Szw -- 91] R. Szwarc. Groups acting on trees and approximation properties of the
Fourier algebra. J. Funct. Anal. 95 (1991), 320 -- 343.
[VeK -- 91] A.M. Vershik, S.V. Kerov. Asymptotic theory of the characters of a sym-
metric group. Functional Anal. Appl. 15 (1982), 246 -- 255.
[Tho -- 68] E. Thoma. Eine Charakterisierung diskreter Gruppen vom Typ I. Invent.
Math. 6 (1968), 190 -- 196.
[Wan -- 75] S.P. Wang. On isolated points in the dual spaces of locally compact
groups. Math. Ann. 218 (1975), 19 -- 34.
[Wei -- 67] A. Weil. Basic number theory. Die Grundlehren der mathematischen
Wissenschaften 144, Springer 1967.
Bachir Bekka, Univ Rennes, CNRS, IRMAR -- UMR 6625, Campus Beaulieu,
F-35042 Rennes Cedex, France
E-mail address: [email protected]
|
1602.04760 | 1 | 1602 | 2016-02-15T18:44:21 | Piecewise conjugacy for multivariable dynamics over the Jacobson spectrum of a C*-algebra | [
"math.OA"
] | We show that if (A,a) and (B,b) are automorphic multivariable C*-dynamical systems with isometrically isomorphic tensor algebras (or semi crossed products), then the systems are piecewise conjugate over their Jacobson spectrum. This answers a question of Kakariadis and the author. | math.OA | math |
PIECEWISE CONJUGACY FOR MULTIVARIABLE DYNAMICS
OVER THE JACOBSON SPECTRUM OF A C∗-ALGEBRA
ELIAS G. KATSOULIS
Abstract. We show that if (A, α) and (B, β) are automorphic multivariable
C∗-dynamical systems with isometrically isomorphic tensor algebras (or semi
crossed products), then the systems are piecewise conjugate over their Jacob-
son spectrum. This answers a question of Kakariadis and the author.
1. Introduction
The concept of a non-selfadjoint operator algebra associated with a multivari-
able dynamical system is new and yet fruitfull. Such algebras appeared for the
first time in the work of Power [8] and Donsig, Katavolos and Manoussos [5] but
their systematic study only started recently with the Memoirs of Davidson and the
author [2]. In that paper, given a multivariable dynamical system over a locally
compact Hausdorff space, we isolated two associated operator algebras, the tensor
algebra and the semi crossed product, and we made the case that various properties
of the dynamical system are encoded in these algebras. Rather surprisingly, the
classification of such algebras in [2] has had an impact beyond operator algebras,
as witnessed in the work of Cornelissen and Marcolli [1] on class field theory.
Inspired by [1, 2], Kakariadis and the author [7] extended the study of mul-
tivariable dynamics beyond commutative C∗-algebras. It turns out that the non-
commutative context allows for questions that do not materialize in the commuta-
tive setting. One such question revolves around the various notions of a spectrum
for a C∗-algebra.
In [7] we showed that if (A, α) and (B, β) are multivariable
dynamical systems with isometrically isomorphic tensor algebras (or semi crossed
products), then the systems are piecewise conjugate over the spectrum, as described
by Ernest in [6]. However the conjugacy over the Jacobson spectrum, i.e., the prim-
itive ideal space with the hull-kernel topology, was left open and it was asked as a
question at the end of the paper [7, Question 3]. In this note we observe that the
presence of a continuous open map between the spectra, combined with the results
of [7] settles this question in the affirmative.
2. The main result
We adhere to the notation of [7] and use as references for the properties of the
various spectra of C∗-algebras the book of Dixmier [4], and Ernest's paper [6].
If A is a C∗-algebra, then Prim(A) will denote its Jacobson spectrum. Let A be
the collection of all (equivalence classes of non-trivial) irreducible representations
2000 Mathematics Subject Classification. 47L55, 47L40, 46L05, 37B20.
Key words and phrases: C∗-algebra, multivariable dynamical system, piecewise conjugacy,
spectrum.
1
2
of A and
(1)
E.G.KATSOULIS
θ : A → Prim(A); A ∋ π 7→ ker π.
The space A is equipped with the smallest topology that makes θ continuous. This
forces θ to be an open mapping as well.
In [7] we worked exclusively with J. Ernest's picture for the spectrum for a C∗-
algebra. Let R(A) be the collection of all railway representations of A; these are
representations equivalent to appropriately large ampliations of irreducible repre-
sentations of A, all acting on the same Hilbert space. The space R(A) is equipped
with the topology of pointwise-SOT convergence. If ϕ : R(A) → A is the map
that associates a railway representation to the equivalence class of its associated
irreducible representation, then [6] shows that ϕ is continuous and open.
Lemma 2.1. Let X, Y, Z be topological spaces and f, g, h maps so that the following
diagram
f
X
❅
❅
❅
❅
g
❅
❅
❅
❅
Y
h
Z
commutes. Assume that f is continuous and open, Z is equipped with the quotient
topology for g and h is a bijective surjection. Then h is a homeomorphism.
Proof. Assume that U ⊆ Z is open. Since g is continuous, g−1(U ) = f −1(h−1(U ))
is also open. Since f is open, we obtain
f (g−1(U )) = f (cid:0)f −1(h−1(U ))(cid:1) = h−1(U )
is open and so h is continuous.
Now let U ⊆ Y be open. Then,
g−1(h(U )) = (cid:0)g−1(h−1)−1(cid:1) (U )
= (h−1g)−1(U ) = f −1(U )
is open. Since Z is equipped with the quotient topology for g, h(U ) is open, i.e., h
is open.
If we take X = R(A), Y = A, Z = R(A)/ ∼, i.e., unitary equivalence classes
of railway representations, f = ϕ, g the quotient map and h the map that assigns
an equivalence class of irreducible representations to the equivalence class of the
corresponding railway representation, then Lemma 2.1 shows that the map h is a
canonical homeomorphism between A and R(A)/ ∼. In the sequel, we will not be
distinguishing between these two spectra.
Let X and Y be topological spaces and let σ = (σ1, σ2, . . . , σn) and τ =
(τ1, τ2, . . . , τn) be multivariable dynamical systems consisting of selfmaps of X and
Y respectively. Davidson and Katsoulis [2, Definition 3.16] define (X, σ) and (Y, τ )
to be piecewise conjugate if there exists a homeomorphism h : X → Y and an open
cover {Ug g ∈ Sn} of Y so that
τi(y) = h σg(i) h−1(y), for each y ∈ Ug, g ∈ Sn and 1 ≤ i ≤ n.
/
/
JACOBSON SPECTRUM
3
An (automorphic) multivariable C∗-dynamical system (A, α) consists of a C∗-
algebra A and ∗-automorphisms α = (α1, α2, . . . , α) acting on A. Any automor-
phism (resp. multivariable system) α of A induces a homeomorphism (resp. multi-
variable dynamical system) α on A, that maps the equivalence class [ρ] of a railway
representation to [ρα]. Similarly, there is a map α on Prim(A) that maps ker π to
ker πα.
Below we answer affirmatively [7, Question 3]. We will not be explaining the
operator algebras appearing in the Theorem below as we do not require any of their
defining properties. Instead we direct the reader to [7, Section 2].
Theorem 2.2. Let (A, α) and (B, β) be multivariable dynamical systems consisting
of ∗-automorphisms.
If the associated operator algebras alg(A, α) and alg(B, β)
are isometrically isomorphic as operator algebras, then the multivariable dynamical
systems (Prim(A), α) and (Prim(B), β) are piecewise conjugate.
Proof. We begin by noticing that for any choice of C∗-algebras C, D and a
∗-automorphism δ : C → D we have a commutative diagram
(2)
C
δ
y
D
θ−−−−→ Prim(C)
δ
y
θ−−−−→ Prim(D)
where θ is defined in (1).
In [7, Theorem 4.9] we proved that if alg(A, α) and alg(B, β) are isometrically
isomorphic as operator algebras, then the dynamical systems ( A, α) and ( B, β) are
piecewise conjugate. Furthermore, it follows from the proof of [7, Theorem 4.9]
that the homeomorphism h implementing the piecewise conjugacy comes from a
∗-automorphism γ : A → B, i.e., h = γ.
Let {Ug g ∈ Sn} be an open cover of B so that
βi(y) = γ αg(i) γ−1(y), for each y ∈ Ug, g ∈ Sn and 1 ≤ i ≤ n.
Since θ is open, {θ(Ug) g ∈ Sn} be an open cover of Prim(B). Furthermore,
repeated use of (2) for the appropriate choices of C, D and δ implies that
βiθ(y) = θ βi(y) = θγ αg(i) γ−1(y)
= γθ αg(i)γ−1(y) = · · · = γ αg(i) γ−1θ(y).
Hence, βi θ(Ug)= γ αg(i) γ−1 θ(Ug ) and we are done.
References
[1] G. Cornelissen and M. Marcolli, Quantum Statistical Mechanics, L-series and Anabelian
Geometry, manuscript.
[2] K. Davidson, E. Katsoulis, Operator algebras for multivariable dynamics, Mem. Amer. Math.
Soc. 209 (2011), no. 982, viii+53 pp.
[3] K. Davidson, E. Katsoulis, Isomorphisms between topological conjugacy algebras, J. reine
angew. Math. 621 (2008), 29-51.
[4] J. Dixmier, C∗-algebras, North-Holland Mathematical Library; v. 15, 1977.
[5] A. Donsig, A. Katavolos and A. Manoussos, The Jacobson radical for analytic crossed prod-
ucts J. Funct. Anal. 187 (2001), 129 -- 145.
[6] J. Ernest, On the topology of the spectrum of a C∗-algebra, Math. Ann. 216 (1975), 149 -- 153.
4
E.G.KATSOULIS
[7] E.T.A. Kakariadis and E. Katsoulis, Isomorphism invariants for multivariable C∗-dynamics.
J. Noncommut. Geom. 8 (2014), 771 -- 787.
[8] S.C. Power, Classification of analytic crossed product algebras, Bull. London Math. Soc. 24
(1992), 368 -- 372.
Dept. Math., East Carolina University, Greenville, NC 27858, USA
E-mail address: [email protected]
|
1601.01259 | 2 | 1601 | 2016-01-13T17:14:16 | A "quantum" Ramsey theorem for operator systems | [
"math.OA",
"math.CO",
"math.FA",
"math.RA",
"quant-ph"
] | Let V be a linear subspace of M_n(C) which contains the identity matrix and is stable under the formation of Hermitian adjoints. We prove that if n is sufficiently large then there exists a rank k orthogonal projection P such that dim(PVP) = 1 or k^2. | math.OA | math |
A "QUANTUM" RAMSEY THEOREM FOR OPERATOR
SYSTEMS
NIK WEAVER
Abstract. Let V be a linear subspace of Mn(C) which contains the identity
matrix and is stable under the formation of Hermitian adjoints. We prove that
if n is sufficiently large then there exists a rank k orthogonal projection P such
that dim(P V P ) = 1 or k2.
An operator system in finite dimensions is a linear subspace V of Mn(C) with
1. Background
the properties
• In ∈ V
• A ∈ V ⇒ A∗ ∈ V
where In is the n × n identity matrix and A∗ is the Hermitian adjoint of A. In this
paper the scalar field will be complex and we will write Mn = Mn(C).
Operator systems play a role in the theory of quantum error correction. In clas-
sical information theory, the "confusability graph" is a bookkeeping device which
keeps track of possible ambiguity that can result when a message is transmitted
through a noisy channel. It is defined by taking as vertices all possible source mes-
sages, and placing an edge between two messages if they are sufficiently similar that
data corruption could lead to them being indistinguishable on reception. Once the
confusability graph is known, one is able to overcome the problem of information
loss by using an independent subset of the confusability graph, which is known as
a "code". If it is agreed that only code messages will be sent, then we can be sure
that the intended message is recoverable.
When information is stored in quantum mechanical systems, the problem of error
correction changes radically. The basic theory of quantum error correction was laid
down in [3]. In [2] it was suggested that in this setting the role of the confusability
graph is played by an operator system, and it was shown that for every operator
system a "quantum Lov´asz number" could be defined, in analogy to the classical
Lov´asz number of a graph. This is an important parameter in classical information
theory. See also [5] for much more along these lines.
The interpretation of operator systems as "quantum graphs" was also proposed in
[8], based on the more general idea of regarding linear subspaces of Mn as "quantum
relations", and taking the conditions In ∈ V and A ∈ V ⇒ A∗ ∈ V to respectively
express reflexivity and symmetry conditions. The idea is that the edge structure
of a classical graph can be encoded in an obvious way as a reflexive, symmetric
relation on a set. This point of view was explicitly connected to the quantum error
correction literature in [9].
Date: Jan. 13, 2016.
1
2
NIK WEAVER
Ramsey's theorem states that for any k there exists n such that every graph with
at least n vertices contains either a k-clique or a k-anticlique, i.e., a set of k vertices
among which either all edges are present or no edges are present. Simone Severini
asked the author whether there is a "quantum" version of this theorem for operator
systems. The natural notions of k-clique and k-anticlique are the following.
Definition 1.1. Let V ⊆ Mn be an operator system. A quantum k-clique of V
is an orthogonal projection P ∈ Mn (i.e., a matrix satisfying P = P 2 = P ∗)
whose rank is k, such that P VP = {P AP : A ∈ V} is maximal; that is, such
that P VP = P MnP ∼= Mk, or equivalently, dim(P VP ) = k2. A quantum k-
anticlique of V is a rank k projection P such that P VP is minimal; that is, such
that P VP = C · P ∼= M1, or equivalently, dim(P VP ) = 1.
The definition of quantum k-anticlique is supported by the fact that in quantum
error correction a code is taken to be the range of a projection satisfying just this
condition, P VP = C · P [3]. As mentioned earlier, classical codes are taken to
be independent sets, which is to say, anticliques. See also Section 4 of [9], where
intuition for why P VP is correctly thought of as a "restriction" of V is given.
The main result of this paper is a quantum Ramsey theorem which states that for
every k there exists n such that every operator system in Mn has either a quantum
k-clique or a quantum k-anticlique. This answers Severini's question positively. The
quantum Ramsey theorem is not merely analogous to the classical Ramsey theorem;
using the bimodule formalism of [8], we can formulate a common generalization of
the two results. This will be done in the final section of the paper.
I especially thank Michael Jury for stimulating discussions, and in particular for
conjecturing Proposition 2.3 and improving Lemma 4.2.
Part of this work was done at a workshop on Zero-error information, Operators,
and Graphs at the Universitat Aut`onoma de Barcelona.
2. Examples
If G = (V, E) is any finite simple graph, without loss of generality suppose
V = {1, . . . , n} and define VG to be the operator system
VG = span{Eij : i = j or {i, j} ∈ E} ⊆ Mn.
Here we use the notation Eij for the n × n matrix with a 1 in the (i, j) entry and
0's elsewhere. Also, let (ei) be the standard basis of Cn, so that Eij = eie∗
j .
The inclusion of the diagonal Eii matrices in VG corresponds to including a loop
at each vertex in G. In the error correction setting this is natural: we place an
edge between any two messages that might be indistinguishable on reception, and
this is certainly true of any message and itself. Once we adopt the convention that
every graph has a loop at each vertex, an anticlique should no longer be a subset
S ⊆ V which contains no edges, it should be a subset which contains no edges
except loops. Such a set corresponds to the projection PS onto span{ei : i ∈ S},
which has the property that PSVGPS = span{Eii : i ∈ S}. Or course this is very
different from a quantum anticlique where P VP is one-dimensional.
To illustrate the dissimilarity between classical and quantum cliques and anti-
cliques, consider the diagonal operator system Dn ⊆ Mn consisting of the diagonal
n × n complex matrices. In the notation used above, this is just the operator sys-
tem VG corresponding to the empty graph on n vertices. It might at first appear
to falsify the desired quantum Ramsey theorem, because of the following fact.
A QUANTUM RAMSEY THEOREM
3
Proposition 2.1. Dn has no quantum k-anticlique for k ≥ 2.
Proof. Let P ∈ Mn be a projection of rank k ≥ 2. Since rank(Eii) = 1 for all
i, it follows that rank(P EiiP ) = 0 or 1 for each i.
If P EiiP = 0 for all i then
i=1 P EiiP = 0, contradiction. Thus we must have rank(P EiiP ) = 1 for
some i, but then P EiiP cannot belong to C · P = {aP : a ∈ C}, since every matrix
in this set has rank 0 or k. So P DnP 6= C · P .
(cid:3)
P = Pn
Since every operator system of the form VG contains the diagonal matrices, none
of these operator systems has nontrivial quantum anticliques. The surprising thing
is that for n sufficiently large, they all have quantum k-cliques. This follows from
the next result.
Proposition 2.2. If n ≥ k2 + k − 1 then Dn has a quantum k-clique.
Proof. Without loss of generality let n = k2 + k − 1. Start by considering Mk acting
on Ck. Find k2 vectors v1, . . . , vk2 in Ck such that the rank 1 matrices viv∗
i are
linearly independent. (For example, we could take the k standard basis vectors ei
plus the k2−k
vectors ei + iej for i 6= j.
The corresponding rank 1 matrices span Mk and thus they must be independent
since dim(Mk) = k2.) Making the identification Cn ∼= Ck ⊕ Ck2−1, we can extend
the vi to orthogonal vectors wi ∈ Cn as follows: take w1 = v1 ⊕ (1, 0, . . . , 0),
w2 = v2 ⊕ (a1, 1, 0, . . . , 0), w3 = v3 ⊕ (b1, b2, 1, 0, . . . , 0), etc., with a1, b1, b2, . . .
successively chosen so that hwi, wj i = 0 for i 6= j. We need k2 − 1 extra dimensions
to accomplish this. Now let P be the rank k projection of Cn onto Ck and let Dn be
the diagonal operator system relative to any orthonormal basis of Cn that contains
the vectors wi
i for all i, so
dim(P DnP ) = k2.
(cid:3)
kwik for 1 ≤ i ≤ k2. Then P DnP contains P wiw∗
vectors ei + ej for i 6= j plus the k2−k
i P = viv∗
2
2
A stronger version of this result will be proven in Lemma 4.3. The value n =
k2 + k − 1 may not be optimal, but note that in order for Dn to have a quantum
k-clique n must be at least k2, since dim(Dn) = n and we need dim(P DnP ) = k2.
Next, we show that operator systems of arbitrarily large dimension may lack
quantum 3-cliques.
Proposition 2.3. Let Vn = span{In, E11, E12, . . . , E1n, E21, . . . , En1} ⊆ Mn.
Then Vn has no quantum 3-cliques.
Proof. Let P ∈ Mn be any projection. If P e1 = 0 then P E1iP = P Ei1P = 0 for
all i, so P is a quantum anticlique. Otherwise let k = rank(P ) and let f1, . . . , fk be
an orthonormal basis of ran(P ) with f1 = P e1
i P = f1v∗
i
where vi = kP e1kP ei. The span of these matrices f1v∗
i },
i
since the projections of the ei span ran(P ). Similarly, the span of the matrices
1 }. So P VnP is just Vk ⊆ Mk ∼= P MnP , relative to
P Ei1P is precisely span{fif ∗
If k ≥ 3 then dim(Vk) = 2k < k2, so P cannot be a quantum
the (fi) basis.
clique.
(cid:3)
kP e1k . Then P E1iP = P e1e∗
is precisely span{f1f ∗
3. Quantum 2-cliques
In contrast to Proposition 2.3, we will show in this section that any operator
system whose dimension is at least four must have a quantum 2-clique. This result
is clearly sharp. It is somewhat analogous to the trivial classical fact that any graph
that contains at least one edge must have a 2-clique.
4
NIK WEAVER
Define the Hilbert-Schmidt inner product of A, B ∈ Mn to be Tr(AB∗). Denote
the set of Hermitian n × n matrices by M h
n . Observe that any operator system is
spanned by its Hermitian part since any matrix A satisfies A = Re(A) + iIm(A)
where Re(A) = 1
2 (A + A∗) and Im(A) = 1
2i (A − A∗).
Lemma 3.1. Let V ⊆ Mn be an operator system and suppose dim(V) ≤ 3. Then
its Hilbert-Schmidt orthocomplement is spanned by rank 2 Hermitian matrices.
Proof. Work in M h
matrices in V ⊥
complex spans then yields the desired result.
n . Let V0 = V ∩M h
0 whose rank is 2. We will show that W0 = V ⊥
n and let W0 be the real span of the Hermitian
n ); taking
0 (in M h
Suppose to the contrary that there exists a nonzero Hermitian matrix B ∈ V ⊥
0
which is orthogonal to W0. Say V0 = span{In, A1, A2}, where A1 and A2 are
not necessarily distinct from In. Since B ∈ V ⊥
0 , we have Tr(InB) = Tr(A1B) =
Tr(A2B) = 0, but Tr(B2) 6= 0. We will show that there is a rank 2 Hermitian
matrix C whose inner products against In, A1, A2, and B are the same as their
inner products against B. This will be a matrix in W0 which is not orthogonal to
B, a contradiction.
Since B is Hermitian, we can choose an orthonormal basis (fi) of Cn with
respect to which it is diagonal, say B = diag(b1, . . . , bn). We may assume
b1, . . . , bj ≥ 0 and bj+1, . . . , bn < 0. Let B+ = diag(b1, . . . , bj, 0, . . . , 0) and
B− = diag(0, . . . , 0, −bj+1, . . . , −bn) be the positive and negative parts of B,
so that B = B+ − B−. Let α = Tr(B+) = Tr(B−) (they are equal since
Tr(B) = Tr(InB) = 0). Then 1
α B+ is a convex combination of the rank 1 ma-
α Tr(AB+) is a convex
trices f1f ∗
combination of the linear functionals A 7→ hAfi, fii for 1 ≤ i ≤ j. By the convexity
of the joint numerical range of three Hermitian matrices [1], there exists a unit
vector v ∈ Cn such that 1
α Tr(AB+) = hAv, vi for A = A1, A2, and B. Similarly,
there exists a unit vector w such that 1
α Tr(AB−) = hAw, wi for A = A1, A2, and
B. Then C = α(vv∗ − ww∗) is a rank 2 Hermitian matrix whose inner products
against In, A1, A2, and B are the same as their inner products against B. So C
has the desired properties.
(cid:3)
j ; that is, the linear functional A 7→ 1
1 , . . ., fjf ∗
Lemma 3.2. Let V ⊆ M3 be an operator system and suppose dim(V) = 4. Then
V has a quantum 2-clique.
Proof. The proof is computational. Say V = span{A0, A1, A2, A3} where A0 = I3
and the other Ai are Hermitian. It will suffice to find two vectors v, w ∈ C3 such
that the four vectors [hAiv, vi, hAiv, wi, hAiw, vi, hAiw, wi] ∈ C4 for 0 ≤ i ≤ 3 are
independent. That is, we need the 4×4 matrix whose rows are these vectors to have
nonzero determinant. Then letting P be the orthogonal projection onto span{v, w}
will verify the lemma.
We can simplify by putting the Ai in a special form. First, by choosing a basis
of eigenvectors, we can assume A1 is diagonal. By subtracting a suitable multiple
of A0 from A1, multiplying by a nonzero scalar, and possibly reordering the basis
vectors, we can arrange that A1 has the form diag(0, 1, a). (Note that dim(V) = 4
implies that A1 cannot be a scalar multiple of A0.) These operations do not affect
span{A0, A1, A2, A3}. Then, by subtracting suitable linear combinations of A0 and
A QUANTUM RAMSEY THEOREM
5
b12
0
¯b23
b13
b23
b33
.
1
0
β
,
0
¯a12
¯a13
a12
0
¯a23
A1, we can arrange that A2 and A3 have the forms
0
¯b12
¯b13
w =
v =
a13
a23
a33
1
α
0
and
and
Let
then evaluate the determinant of the 4 × 4 matrix described above and expand it
as a polynomial in α, β, ¯α, and ¯β. We just need this determinant to be nonzero
for some values of α and β; if this fails, then the polynomial coefficients must all
be zero, and direct computation shows that this forces one of A2 and A3 to be a
scalar multiple of the other. We omit the tedious but straightforward details. (cid:3)
Theorem 3.3. Let V ⊆ Mn be an operator system and suppose dim(V) ≥ 4. Then
V has a quantum 2-clique.
Proof. Without loss of generality we can suppose that dim(V) = 4. Say V =
span{In, A1, A2, A3} where the Ai are Hermitian and linearly independent.
We first claim that there is a projection P of rank at most 3 such that P InP ,
P A1P , and P A2P are linearly independent. If A1 and A2 are jointly diagonalizable
then we can find three common eigenvectors v1, v2, and v3 such that the vectors
(1, 1, 1), (λ1, λ2, λ3), (µ1, µ2, µ3) ∈ C3 are linearly independent, where λi and µi are
the eigenvalues belonging to vi for A1 and A2, respectively. Then the projection
onto span{v1, v2, v3} verifies the claim. If A1 and A2 are not jointly diagonalizable,
then we can find two eigenvectors v1 and v2 of A1 such that hA2v1, v2i 6= 0. Let-
ting v3 be a third eigenvector of A1 with the property that the eigenvalues of A1
belonging to v1, v2, and v3 are not all equal, we can again use the projection onto
span{v1, v2, v3}. This establishes the claim.
Now let P be as in the claim and find B ∈ Mn such that P InP , P A1P , P A2P ,
and P BP are linearly independent. By Lemma 3.2 we can then find a rank 2 pro-
jection Q ≤ P such that QInQ, QA1Q, QA2Q, and QBQ are linearly independent.
If QInQ, QA1Q, QA2Q, and QA3Q are linearly independent then we are done.
Otherwise, let α, β, and γ be the unique scalars such that QA3Q = αQInQ +
βQA1Q + γQA2Q. By Lemma 3.1 we can find a rank 2 Hermitian matrix C such
that Tr(InC) = Tr(A1C) = Tr(A2C) = 0 but Tr(A3C) 6= 0. Then C = vv∗ − ww∗
for some orthogonal vectors v and w. Thus, hAv, vi = hAw, wi for A = In, A1, and
A2, but not for A = A3. It follows that the two conditions
hA3v, vi = αhInv, vi + βhA1v, vi + γhA2v, vi
and
hA3w, wi = αhInw, wi + βhA1w, wi + γhA2w, wi
cannot both hold. Without loss of generality suppose the first fails. Then letting Q′
be the projection onto span(ran(Q) ∪ {v}), we cannot have Q′A3Q′ = αQ′InQ′ +
βQ′A1Q′ + γQ′A2Q′. Thus rank(Q′) = 3 and dim(Q′VQ′) = 4. The theorem now
follows by applying Lemma 3.2 to Q′VQ′.
(cid:3)
Theorem 3.3 does not generalize to arbitrary four-dimensional subspaces of Mn.
For instance, let V = span{E11, E12, E13, E14} ⊂ M4; by reasoning similar to that in
the proof of Proposition 2.3, if P is any rank 2 projection in M4 then dim(P VP ) ≤ 2.
6
NIK WEAVER
4. The main theorem
The proof of our main theorem proceeds through a series of lemmas.
If dim(V) ≥
Lemma 4.1. Suppose the operator system V is contained in Dn.
k2 + k − 1 then V has a quantum k-clique. If dim(V) ≤ n−k
k−1 then V has a quantum
k-anticlique. If n ≥ k3 − k + 1 then V has either a quantum k-clique or a quantum
k-anticlique.
Proof. If dim(V) ≥ k2 + k − 1 = m then we can find a set of indices S ⊆ {1, . . . , n}
of cardinality m such that dim(P VP ) = m where P is the orthogonal projection
onto span{ei : i ∈ S}. Then P VP ∼= Dm ⊆ Mm ∼= P MnP and Proposition 2.2
yields that P VP , and hence also V, has a quantum k-clique. If dim(V) ≤ n−k
k−1 then
a result of Tverberg [6, 7] can be used to extract a quantum k-anticlique; this is
essentially Theorem 4 of [4]. Thus if k2 + k − 1 ≤ n−k
k−1 then one of the two cases
must obtain, i.e., V must have either a quantum k-clique or a quantum k-anticlique.
A little algebra shows that this inequality is equivalent to n ≥ k3 − k + 1.
(cid:3)
Lemma 4.2. Let v1, . . . , vr be vectors in Cs. Then there are vectors w1, . . . , wr ∈
Cr−1 such that the vectors vi ⊕ wi ∈ Cs+r−1 are pairwise orthogonal and all have
the same norm.
Proof. Let G be the Gramian matrix of the vectors vi and let kGk be its operator
norm. Then rank(kGkIr − G) ≤ r − 1, so we can find vectors wi ∈ Cr−1 whose
Gramian matrix is kGkIr − G. The Gramian matrix of the vectors vi ⊕ wi is then
kGkIr, as desired.
(cid:3)
Then next lemma improves Proposition 2.2.
Lemma 4.3. Let n = k2 + k − 1 and suppose A1, . . . , Ak2 are Hermitian matrices
in Mn such that for each i we have hAiei, eii = 1, and also hAier, esi = 0 whenever
max{r, s} > i. Then V = span{I, A1, . . . , Ak2 } has a quantum k-clique.
Proof. Let Ai have matrix entries (ai
such that the matrices
rs). The goal is to find vectors v1, . . . , vk2 ∈ Ck
A′
i = X1≤r,s≤k2
ai
rsvrv∗
s ∈ Mk
are linearly independent. Once we have done this, find vectors wi ∈ Ck2−1 as in
N (vi ⊕ wi) ∈ Cn ∼= Ck ⊕ Ck2−1 where N is the common
Lemma 4.2 and let fi = 1
norm of the vi ⊕ wi. Then the fi form an orthonormal set in Cn, so they can be
extended to an orthonormal basis, and the operators whose matrices for this basis
i on the initial Ck, which are linearly
are the Ai compress to the matrices
independent. So P VP contains k2 linearly independent matrices, where P is the
orthogonal projection onto Ck, showing that V has a quantum k-clique.
N 2 A′
1
The vectors vi are constructed inductively. Once v1, . . . , vi are chosen so that
A′
1, . . . , A′
i are independent, future choices of the v's cannot change this since
A1, . . . , Ai all live on the initial i × i block. We can let v1 be any nonzero vec-
tor in Ck, since A1 = e1e∗
1 and this only has to be nonzero.
Now suppose v1, . . . , vi−1 have been chosen and we need to select vi so that A′
i is
rsvrv∗
independent of A′
s .
i−1. After choosing vi we will have A′
1, so that A′
i = P1≤r,s≤i ai
1 = v1v∗
1, . . . , A′
A QUANTUM RAMSEY THEOREM
7
Let B be this sum restricted to 1 ≤ r, s ≤ i − 1. That part is already determined
since vi does not appear. Also let
u = ai
1iv1 + · · · + ai
(i−1)ivi−1;
then we will have
A′
(using the assumption that ai
i = B + uv∗
ii = 1). That is,
i + viu∗ + viv∗
i
A′
i = (B − uu∗) + (u + vi)(u + vi)∗ = B′ + uu∗
where u = u + vi is arbitrary, and the question is whether u can be chosen to make
this matrix independent of A′
i span Mk
- there is no matrix which is Hilbert-Schmidt orthogonal to B′ + uu∗ for all u
- so there must be a choice of u which makes A′
i−1, as
desired.
(cid:3)
i−1. But the possible choices of A′
i independent of A′
1, . . . , A′
1, . . . , A′
Next we prove a technical variation on Lemma 4.3.
Lemma 4.4. Let n = k4 + k3 + k − 1 and let V be an operator system contained
in Mn. Suppose V contains matrices A1, . . . , Ak4+k3 such that for each i we have
hAiei, ei+1i 6= 0, and also hAier, esi = 0 whenever max{r, s} > i + 1 and r 6= s.
Then V has a quantum k-clique.
rr, . . . , ai
Proof. Let Ai have matrix entries (ai
rs). Observe that for each i the compression
of Ai to span{ei+2, . . . , en} is diagonal. For each r > i + 1 let the r-tail of Ai
be the vector (ai
nn). Suppose there exist indices i1, . . . , ik2+k−1 such that
the r-tails of the Aij , 1 ≤ j ≤ k2 + k − 1, are linearly independent, where r =
maxj{ij + 2}. Then the compression of V to span{er, . . . , en} contains k2 + k − 1
linearly independent diagonal matrices, so it has a quantum k-clique by the first
assertion of Lemma 4.1. Thus, we may assume that for any k2 + k − 1 distinct
indices ij the matrices Aij have linearly dependent r-tails.
We construct an orthonormal sequence of vectors vi and a sequence of Her-
mitian matrices Bi ∈ V, 1 ≤ i ≤ k2, such that the compressions of the Bi to
span{v1, . . . , vk2 , ek4+k3+1, . . . , ek4+k3+k−1} satisfy the hypotheses of Lemma 4.3.
This will ensure the existence of a quantum k-clique.
1 = Pk2+k−1
The first k2 + k − 1 matrices A1, . . . , Ak2+k−1 have linearly dependent r-tails for
r = k2 + k + 1. Thus there is a nontrivial linear combination B′
αiAi
whose r-tail is the zero vector. Letting j be the largest index such that αj is
nonzero, we have hB′
1ej, ej+1i 6= 0 because hAj ej, ej+1i 6= 0 but hAiej, ej+1i = 0
for i < j. Thus the compression of B′
1 to span{e1, . . . , ek2+k} is nonzero, so there
exists a unit vector v1 in this span such that hB′
1v1, v1i 6= 0. Then let B1 be a scalar
multiple of either the real or imaginary part of B′
1 which satisfies hB1v1, v1i = 1.
Note that hB1er, esi = 0 for any r, s with max{r, s} > k2 + k. Apply the same
reasoning to the next block of k2 + k − 1 matrices Ak2+k+1, . . . , A2k2+2k−1 to find
v2 and B2, and proceed inductively. After k2 steps, k2(k2 + k) = k4 + k3 indices
will have been used up and k − 1 (namely, ek4+k3+1, . . . , ek4+k3+k−1) will remain,
as needed.
(cid:3)
i=1
Theorem 4.5. Every operator system in M8k11 has either a quantum k-clique or
a quantum k-anticlique.
8
NIK WEAVER
Proof. Set n = 8k11 and let V be an operator system in Mn. Find a unit vector
v1 ∈ Cn, if one exists, such that the dimension of Vv1 = {Av1 : A ∈ V} is less than
8k8. Then find a unit vector v2 ∈ (Vv1)⊥, if one exists, such that the dimension of
(Vv1)⊥ ∩ (Vv2) is less than 8k8. Proceed in this fashion, at the rth step trying to
find a unit vector vr in
(Vv1)⊥ ∩ · · · ∩ (Vvr−1)⊥
such that the dimension of
(Vv1)⊥ ∩ · · · ∩ (Vvr−1)⊥ ∩ (Vvr)
is less than 8k8. If this construction lasts for k3 steps then the compression of V to
span{v1, . . . , vk3 } ∼= Mk3 is contained in Dk3 , so this compression, and hence also
V, has either a quantum k-clique or a quantum k-anticlique by Lemma 4.1.
Otherwise, the construction fails at some stage d. This means that the compres-
sion V ′ of V to F = (Vv1)⊥ ∩ · · · ∩ (Vvd)⊥ has the property that the dimension of
V ′v is at least 8k8, for every unit vector v ∈ F .
Work in F . Choose any nonzero vector w1 ∈ F and find A1 ∈ V ′ such that w2 =
A1w1 is nonzero and orthogonal to w1. Then find A2 ∈ V ′ such that w3 = A2w2
is nonzero and orthogonal to span{w1, w2, A1w1, A∗
1w2}. Continue in
this way, at the rth step finding Ar ∈ V ′ such that wr+1 = Arwr is nonzero and
orthogonal to span{wj, Aiwj, A∗
i wj : i < r and j ≤ r}. The dimension of this span
is at most 2r2 − r, so as long as r ≤ 2k4 its dimension is less than 8k8 and a vector
wr+1 can be found. Compressing to the span of the wi then puts us in the situation
of Lemma 4.4 with n = 2k4, which is more than enough. So there exists a quantum
k-clique by that lemma.
(cid:3)
1w1, A1w2, A∗
The constants in the proof could easily be improved, but only marginally. Very
likely the problem of determining optimal bounds on quantum Ramsey numbers is
open-ended, just as in the classical case.
5. A generalization
In this section we will present a result which simultaneously generalizes the
classical and quantum Ramsey theorems. This is less interesting than it sounds
because the proof involves little more than a reduction to these two special cases.
Perhaps the statement of the theorem is more significant than its proof.
At the beginning of Section 2 we showed how any simple graph G on the vertex
set {1, . . . , n} gives rise to an operator system VG ⊆ Mn. This operator system has
the special property that it is a bimodule over Dn, i.e., it is stable under left and
right multiplication by diagonal matrices. Conversely, it is not hard to see that any
operator system in Mn which is also a Dn-Dn-bimodule must have the form VG
for some G ([8], Propositions 2.2 and 2.5). The general definition therefore goes as
follows:
Definition 5.1. ([8], Definition 2.6 (d)) Let M be a unital ∗-subalgebra of Mn. A
quantum graph on M is an operator system V ⊆ Mn which satisfies M′VM′ = V.
Here M′ = {A ∈ Mn : AB = BA for all B ∈ M} is the commutant of M. This
if M and N are ∗-isomorphic
definition is actually representation-independent:
unital ∗-subalgebras of two matrix algebras (possibly of different sizes), then the
quantum graphs on M naturally correspond to the quantum graphs on N ([8],
A QUANTUM RAMSEY THEOREM
9
Theorem 2.7). More properly, one could say that the pair (M, V) is the quantum
graph, just as a classical graph is a pair (V, E).
If M = Mn then its commutant is C·In and the bimodule condition in Definition
5.1 is vacuous: any operator system in Mn is a quantum graph on Mn. On the other
hand, the commutant of M = Dn is itself, so that by the comment made above,
the quantum graphs on Dn - the operator systems which are Dn-Dn-bimodules -
correspond to simple graphs on the vertex set {1, . . . , n}. In this correspondence,
subsets of {1, . . . , n} give rise to orthogonal projections P ∈ Dn, and the k-cliques
and k-anticliques of the graph are realized in the matrix picture as rank k orthogonal
projections P ∈ Dn which satisfy P VGP = P MnP or P DnP , respectively. This
suggests the following definition.
Definition 5.2. Let M be a unital ∗-subalgebra of Mn and let V ⊆ Mn be a
quantum graph on M. A rank k projection P ∈ M is a quantum k-clique if it
satisfies P VP = P MnP and a quantum k-anticlique if it satisfies P VP = P M′P .
Since every operator system contains the identity matrix, if V is a quantum
graph on M then M′ ⊆ V. So P VP = P M′P is the minimal possibility, as
P VP = P MnP is the maximal possibility. Note the crucial requirement that P
must belong to M.
If M = Mn then M′ = C · In and the preceding definition duplicates the notions
of quantum k-clique and quantum k-anticlique used earlier in the paper, whereas if
M = Dn it effectively reproduces the classical notions of k-clique and k-anticlique
in a finite simple graph. In the classical setting fewer operator systems count as
graphs, but one also has less freedom in the choice of P when seeking cliques or
anticliques.
We require only the following simple lemma.
Lemma 5.3. Let V ⊆ Mnd ∼= Mn ⊗ Md be a quantum graph on Mn ⊗ Id.
If
nd ≥ 8k11 then there is a projection in Mn ⊗ Id whose rank is at least k, and which
is either a quantum clique or a quantum anticlique of V.
Proof. Since V is a bimodule over (Mn⊗Id)′ = In⊗Md, it has the form V = W ⊗Md
for some operator system W ⊆ Mn. If d = 1 then the desired result was proven in
Theorem 4.5, and if d ≥ k then any projection of the form P ⊗ Md, where P is a
rank 1 projection in Mn, will have rank at least k and be both a quantum clique
and a quantum anticlique. So assume 2 ≤ d < k.
Now if d ≥ 3 then d10/11 > 2, so d < k(d10/11 − 1). Thus 1 < k
d (d10/11 − 1), i.e.,
k
d + 1 < k
d · d10/11 = k
d1/11 , which implies ( k
d . So finally
n ≥
8k11
d
> 8(cid:18) k
d
d + 1)11 < k11
d(cid:25)11
+ 1(cid:19)11
> 8(cid:24) k
.
If d = 2 then d < 3(d10/11 − 1), so the same reasoning leads to the same inequality
n ≥ 8⌈ k
d ⌉11 provided k ≥ 3, and the inequality is immediate when k = d = 2. So
we conclude that in all cases n ≥ 8⌈ k
d ⌉-
clique or a quantum ⌈ k
d ⌉-anticlique Q ∈ Mn. Then Q ⊗ Id is correspondingly either
a quantum ⌈ k
d ⌉ · d-anticlique of V, which is enough. (cid:3)
d ⌉11. By Theorem 4.5, W has a quantum ⌈ k
d ⌉ · d-clique or a quantum ⌈ k
Note that we cannot promise a quantum k-clique or -anticlique, only a ≥ k-clique
or -anticlique, since the rank of any projection in Mn ⊗ Id is a multiple of d.
10
NIK WEAVER
Theorem 5.4. For every k there exists n such that if M is a unital ∗-subalgebra
of Mn and V ⊆ Mn is an operator system satisfying M′VM′ = V, then there is a
projection P ∈ M whose rank is at least k and such that P VP = P MnP or P M′P .
Proof. Let R(k, k) be the classical Ramsey number and set n = 8k11 · R(k, k).
Now M has the form (Mn1 ⊗ Id1) ⊕ · · · ⊕ (Mnr ⊗ Idr ) for some pair of sequences
(n1, . . . , nr) and (d1, . . . , dr) such that n1d1 + · · · + nrdr = n. Thus if r ≤ R(k, k)
then for some i we must have nidi ≥ 8k11, and compressing to that block then
yields the desired conclusion by appealing to the lemma. Otherwise, if r > R(k, k),
then choose a sequence of rank 1 projections Qi ∈ Mni and work in QMnQ where
Q = (Q1 ⊗ Id1 ) ⊕ · · · ⊕ (Qr ⊗ Idr ). Then QMnQ ∼= Md1+···+dr , QMQ ∼= C · Id1 ⊕
∼= Dr, and QVQ is a bimodule over the commutant of QMQ in QMnQ,
· · · ⊕ C · Idr
i.e., the ∗-algebra Md1 ⊕ · · · ⊕ Mdr . It follows that there is a graph G = (V, E) on
the vertex set V = {1, . . . , r} such that QVQ has the form
QVQ = X Eij ⊗ Mdidj ,
taking the sum over the set of pairs {(i, j) : i = j or {i, j} ∈ E} ([8], Theorem 2.7).
Since r > R(k, k), there exists either a k-clique or a k-anticlique in G, and this
gives rise to a diagonal projection in QMQ whose rank is at least k and which is
either a quantum k-clique or a quantum k-anticlique of V.
(cid:3)
Again, when M = Mn Theorem 5.4 recovers the quantum Ramsey theorem and
when M = Dn it recovers the classical Ramsey theorem (though in both cases with
worse constants).
Theorem 5.4 could also be proven by mimicking the proof of Theorem 4.5. How-
ever, in order to accomodate the requirement that P belong to M we need to modify
the last part of the proof so as to be sure that each wr belongs to Ww1∩· · ·∩Wwr−1.
This means that instead of needing Wv to have sufficiently large dimension for each
v, we need it to have sufficiently small codimension. Ensuring that this must be
the case if the construction in the first part of the proof fails then requires that
construction to take place in a space whose dimension is exponential in k. This
explains the dramatic difference between classical and quantum Ramsey numbers
(the first grows exponentially, the second polynomially).
References
[1] Y. H. Au-Yeung and Y. T. Poon, A remark on the convexity and positive definiteness
concerning Hermitian matrices, Southeast Asian Bull. Math. 3 (1979), 85-92.
[2] R. Duan, S. Severini, and A. Winter, Zero-error communication via quantum channels,
noncommutative graphs, and a quantum Lov´asz number, IEEE Trans. Inform. Theory
59 (2013), 1164-1174.
[3] E. Knill and R. Laflamme, Theory of quantum error-correcting codes, Phys. Rev. A
55 (1997), 900-911.
[4] E. Knill, R. Laflamme, and L. Viola, Theory of quantum error correction for general
noise, Phys. Rev. Lett. 84 (2000), 2525-2528.
[5] D. Stahlke, Quantum source-channel coding and non-commutative graph theory,
arXiv:1405.5254.
[6] H. Tverberg, A generalization of Radon's theorem, J. London Math. Soc. 41 (1966),
123-128.
[7] ---, A generalization of Radon's theorem, II, Bull. Austral. Math. Soc. 24 (1981),
321-325.
[8] N. Weaver, Quantum relations, Mem. Amer. Math. Soc. 215 (2012), v-vi, 81-140.
[9] ---, Quantum graphs as quantum relations, arXiv:1506.03892.
A QUANTUM RAMSEY THEOREM
11
Department of Mathematics, Washington University, Saint Louis, MO 63130
E-mail address: [email protected]
|
1211.0239 | 2 | 1211 | 2013-03-25T19:13:51 | KMS states for the generalized gauge action on graph algebras | [
"math.OA"
] | Given a positive function on the set of edges of an arbitrary directed graph $E=(E^0,E^1)$, we define a one-parameter group of automorphisms on the C*-algebra of the graph $C^*(E)$, and study the problem of finding KMS states for this action. We prove that there are bijective correspondences between KMS states on $C^*(E)$, a certain class of states on its core, and a certain class of tracial states on $C_0(E^0)$. We also find the ground states for this action and give some examples. | math.OA | math |
KMS STATES FOR THE GENERALIZED GAUGE
ACTION ON GRAPH ALGEBRAS
Gilles G. de Castro and Fernando de L. Mortari
Abstract. Given a positive function on the set of edges of an arbitrary di-
rected graph E = (E 0, E 1), we define a one-parameter group of automorphisms
on the C*-algebra of the graph C ∗(E), and study the problem of finding KMS
states for this action. We prove that there are bijective correspondences be-
tween KMS states on C ∗(E), a certain class of states on its core, and a certain
class of tracial states on C0(E 0). We also find the ground states for this action
and give some examples.
1. Introduction
Given a directed graph E = (E0, E1), we can associate to it a C*-algebra
C∗(E), and an interesting problem that arises is to find relations between the
algebraic properties of the algebra and the combinatorial properties of the graph
[?]. One such problem is to determine the set of KMS states for a certain action
on the algebra.
Graph algebras are a generalization of Cuntz algebras and Cuntz-Krieger alge-
bras. For the Cuntz algebra, there is a very natural action of the circle, the gauge
action, which can be extended to an action of the real line. The KMS states for
this action are studied in [?] and later generalized to a more general action of the
line, that can be thought of as a generalized gauge action [?]. The same is done for
the Cuntz-Krieger algebras [?], [?].
Recently there were similar results proven for the C*-algebra associated to a
finite graph. This is done in [?] for an arbitrary finite graph, in [?] for a certain class
of finite graphs via groupoid C*-algebras and in [?] for the Toeplitz C*-algebra of
the graph.
Our goal is to generalize these results to the case of an arbitrary graph. First
we analyze which conditions the restrictions of a KMS state to the core of C∗(E)
and to C0(E0) must satisfy. By using a description of the core as an inductive limit,
we can build a KMS state on C∗(E) from a tracial state on C0(E0) satisfying the
conditions found.
In section 2 we review some of the basic definitions and results about graph
algebras as well as the description of the core as an inductive limit. In section 3
Partially supported by Funpesquisa/UFSC
The first named author was also supported by project mathamsud U11-MATH05 and
Capes/Math-AmSud 013/10.
1
2
GILLES G. DE CASTRO AND FERNANDO DE L. MORTARI
we establish the results concerning KMS states, followed by a discussion on ground
states in section 4. In section 5, some examples are given.
2. Graph algebras
Definition 2.1. A (directed) graph E = (E0, E1, r, s) consists of nonempty
sets E0, E1 and functions r, s : E1 → E0; an element of E0 is called a vertex of the
graph, and an element of E1 is called an edge. For an edge e, we say that r(e) is
the range of e and s(e) is the source of e.
Definition 2.2. A vertex v in a graph E is called a source if r−1(v) = ∅, and
is said to be singular if it is either a source, or r−1(v) is infinite.
Definition 2.3. A path of length n in a graph E is a sequence µ = µ1µ2 . . . µn
such that r(µi + 1) = s(µi) for all i = 1, . . . , n − 1. We write µ = n for the length
of µ and regard vertices as paths of length 0. We denote by En the set of all paths
of length n and E∗ = ∪n≥0En. We extend the range and source maps to E∗ by
defining s(µ) = s(µn) and r(µ) = r(µ1) if n ≥ 2 and s(v) = v = r(v) for n = 0.
Definition 2.4. Given a graph E, we define the C*-algebra of E as the uni-
versal C*-algebra C∗(E) generated by mutually orthogonal projections {pv}v∈E0
and partial isometries {se}e∈E1 with mutually orthogonal ranges such that
ese = ps(e);
(1) s∗
(2) ses∗
e ≤ pr(e) for every e ∈ E1;
(3) pv =Pe∈r−1(v) ses∗
e for every v ∈ E0 such that 0 < r−1(v) < ∞.
For a path µ = µ1 . . . µn, we denote the composition sµ1 . . . sµn by sµ, and for
v ∈ E0 we define sv to be the projection pv.
Propositions 2.5, 2.6, 2.8 and 2.9 below are found in [?] (as Corollary 1.15,
Proposition 2.1, Proposition 3.2 and Corollary 3.3, respectively) in the context of
row-finite graphs, but their proofs hold just the same for general graphs as above.
Proposition 2.5. For α, β, µ, ν ∈ E∗ we have
(sµs∗
ν)(sαs∗
β) =
sµα′ s∗
β
sµs∗
βν′
0
if α = να′
if ν = αν′
otherwise
and C∗(E) = span{sµs∗
ν : µ, ν ∈ E∗, s(µ) = s(ν)}.
Proposition 2.6. Let E be a graph. Then there is an action γ of T on C∗(E),
called a gauge action, such that γz(se) = zse for every e ∈ E1 and γz(pv) = pv for
every v ∈ E0.
Definition 2.7. The core of the algebra C∗(E) is the fixed-point subalgebra
for the gauge action, denoted by C∗(E)γ .
Proposition 2.8. C∗(E)γ = span{sµs∗
ν : µ, ν ∈ E∗, s(µ) = s(ν), µ = ν}.
Proposition 2.9. There is a conditional expectation Φ : C∗(E) → C∗(E)γ
such that Φ(sµs∗
ν) = [µ = ν]sµs∗
ν.
It is useful to describe the core as an inductive limit of subalgebras, as was
done in an appendix in [?]. The idea is as follows. For k ≥ 0 define the sets
Fk = span{sµs∗
ν : µ, ν ∈ Ek, s(µ) = s(ν)},
KMS STATES FOR THE GENERALIZED GAUGE ACTION ON GRAPH ALGEBRAS
3
Ek = span{sµs∗
ν : µ, ν ∈ Ek and s(µ) = s(ν) is singular},
Also, for a given vertex v we define
Ck = F0 + · · · + Fk.
Fk(v) = span{sµs∗
ν : µ, ν ∈ Ek, s(µ) = s(ν) = v}
so that
(2.1)
Fk = Mv∈E0
Fk(v)
as a direct sum of C*-algebras.
Lemma 2.10. Let Λ be the set of all finite subsets of Ek and for λ ∈ Λ define
sµs∗
µ.
uλ = Xµ∈λ
Then {uλ}λ∈Λ is an approximate unit of Fk consisting of projections.
Proof. This is a direct consequence of Proposition 2.5.
(cid:3)
The following result is a combination of Proposition A.1 and Lemma A.2 in [?].
Proposition 2.11. With the notation as above for a graph E, the following
hold for k ≥ 0:
(a) Ck is a C*-subalgebra of C∗(E)γ , Fk+1 is an ideal in Ck, Ck ⊆ Ck+1 and
C∗(E)γ = lim−→ Ck.
(b) Fk ∩ Fk+1 =L{Fk(v) : 0 < r−1(v) < ∞}. (C*-algebraic direct sum)
(c) Ck = E0 ⊕ E1 ⊕ · · · ⊕ Ek−1 ⊕ Fk (vector space direct sum)
With the above, we can now prove the following.
Proposition 2.12. For each k ≥ 0, Ck ∩ Fk+1 = Fk ∩ Fk+1.
Proof. Obviously one has Fk ∩ Fk+1 ⊆ Ck ∩ Fk+1. On the other hand, given
x ∈ Ck ∩ Fk+1, one can decompose x as sums in Ck and Ck+1 with Proposition
2.11(c), use the fact that Fk = Ek ⊕ Fk ∩ Fk+1 and the uniqueness of the direct sum
decompositions of x to conclude that x ∈ Fk.
(cid:3)
3. KMS states for the generalized gauge action
In this section we define an action on C∗(E) from a function c : E1 → R∗
+
similar to what is done in [?] for the Cuntz algebras and in [?] for the Cuntz-
Krieger algebras. We will always suppose that there is a constant k > 0 such that
c(e) > k for all e ∈ E1 and that β > 0. Observe that in this case c−β is bounded.
+ by defining c(v) = 1
We extend a function as above to a function c : E∗ → R∗
if v ∈ E0 and c(µ) = c(µ1) . . . c(µn) if µ = µ1 · · · µn ∈ En.
Proposition 3.1. Given a function c : E1 → R∗
tinuous action σc : R → Aut(C∗(E)) given by σc
σc
t (se) = c(e)itse for all e ∈ E1.
+, there is a strongly con-
t (pv) = pv for all v ∈ E0 and
4
GILLES G. DE CASTRO AND FERNANDO DE L. MORTARI
e = ses∗
Proof. Let Te = c(e)itse and note that Te is a partial isometry with T ∗
e Te =
ese and TeT ∗
s∗
e. It follows that the sets {pv}v∈E0 and {Te}e∈E1 satisfy the
same relations as {pv}v∈E0 and {se}e∈E1. By the universal property, there is a
homomorphism σc
t (pv) = pv for all v ∈ E0 and
σc
t (se) = Te = c(e)itse for all e ∈ E1.
t : C∗(E) → C∗(E) such that σc
It is easy to see that σc
t1 ◦ σc
t2 = σc
t1+t2 and σc
0 = Id. Hence σc
t is an automor-
phism with inverse σc
−t.
x = Pµ,ν∈E∗ λµ,νsµs∗
λµ,ν 6= 0, there is δµ,ν such that
To prove continuity, let a ∈ C∗(E), t ∈ R and ε > 0. Take x to be a finite sum
ν such that ka − xk < ε/3. For each pair of paths µ, ν with
c(µ)itc(ν)−it − c(µ)iuc(ν)−iu <
ε
3Pµ,ν∈E∗ kλµ,νsµs∗
ν k
for all u ∈ R with t − u < δµ,ν. If we take δ to be the minimum of all such δµ,ν,
then for all u ∈ R with t − u < δ we have
kσc
t (x) − σc
(c(µ)itc(ν)−it − c(µ)iuc(ν)−iu)λµ,ν sµs∗
<
u(x)k =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xµ,ν∈E∗
3Pµ,ν∈E∗ kλµ,νsµs∗
ε
kλµ,νsµs∗
νk =
ε
3
νk Xµ,ν∈E∗
ν(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
and hence
kσc
≤ kσc
t (a) − σc
u(a)k = kσc
t (a) − σc
t (a − x)k + kσc
t (x) − σc
t (x) + σc
u(x)k + kσc
u(x) + σc
t (x) − σc
u(x − a)k ≤ ε/3 + ε/3 + ε/3 = ε.
u(x) − σc
u(a)k ≤
From now on, we will write simply σ instead of σc. The next result shows that
KMS states on C∗(E) are determined by their values at the core algebra.
(cid:3)
Proposition 3.2. Suppose c : E1 → R∗
+ is such that c(µ) 6= 1 for all µ ∈
If two (σ, β)-KMS states ϕ1, ϕ2 on C∗(E) coincide at the core algebra
E∗\E0.
C∗(E)γ, then ϕ1 = ϕ2.
sµs∗
Proof. Taking an arbitrary sµs∗
ν ∈ C∗(E)γ and thus ϕ1(sµs∗
Suppose then that µ 6= ν, and denote the functional ϕ2 − ϕ1 by ϕ. Using
ν such that s(µ) = s(ν), if µ = ν then
ν ) = ϕ2(sµs∗
ν).
the KMS condition, one obtains
ϕ(sµs∗
ν ) = ϕ(s∗
νc(µ)−β sµ) =
c(µ)−βϕ(s∗
c(µ)−β ϕ(s∗
ν′ )
µ′ )
0
if ν = µν′
if µ = νµ′
otherwise
.
It is therefore sufficient to show that ϕ(sµ) = ϕ(s∗
notice that if C∗(E) has a unit, then
µ) = 0 if µ ≥ 1. To see this,
ϕ(sµ) = ϕ(sµ1) = ϕ(1c(µ)−βsµ) = c(µ)−β ϕ(sµ),
whence ϕ(sµ) = 0 since c(µ) 6= 1 by hypothesis; the non-unital case is established
analogously with the use of an approximate unit.
(cid:3)
KMS STATES FOR THE GENERALIZED GAUGE ACTION ON GRAPH ALGEBRAS
5
Theorem 3.3. Suppose c : E1 → R∗
+ is such that c(µ) 6= 1 for all µ ∈ E∗\E0.
If ϕ is a (σ, β)-KMS state on C∗(E) then its restriction ω = ϕC ∗(E)γ to C∗(E)γ
satisfies
ω(sµs∗
(3.1)
conversely, if ω is a state on C∗(E)γ satisfying (3.1) then ϕ = ω◦Φ is a (σ, β)-KMS
state on C∗(E), where Φ is the conditional expectation from proposition 2.9. The
correspondence thus obtained is bijective and preserves convex combinations.
ν) = [µ = ν]c(µ)−βω(ps(µ));
Proof. Let ϕ be a (σ, β)-KMS state on C∗(E) and ω its restriction to C∗(E)γ.
If µ, ν are paths such that µ = ν and s(µ) = s(ν) then
ν σiβ(sµ)) = ϕ(s∗
ν) = ϕ(sµs∗
ν) = ϕ(s∗
ω(sµs∗
ν c(µ)−β sµ) =
= [µ = ν]c(µ)−βϕ(ps(µ)) = [µ = ν]c(µ)−β ω(ps(µ)).
Conversely, let ω be a state on C∗(E)γ satisfying (3.1) and ϕ = ω ◦ Φ; we
have to show that ϕ satisfies the KMS condition. By continuity and linearity, it is
η where µ, ν, ζ, η ∈ E∗
sufficient to verify this for elements x = sµs∗
are paths such that s(µ) = s(ν) and s(ζ) = s(η).
ν and y = sζs∗
We need to check that ϕ(xy) = ϕ(yσiβ(x)). First note that
xy = (sµs∗
ν)(sζ s∗
and
yσiβ(x) = c(µ)−βc(ν)β (sζs∗
η)(sµs∗
if ζ = νζ′
if ν = ζν′
otherwise
(1)
(2)
(3)
sµζ′ s∗
η
sµs∗
ην′
0
η) =
ν ) = c(µ)−βc(ν)β
sζµ′ s∗
ν
sζs∗
νη′
0
if µ = ηµ′
if η = µη′
otherwise
(a)
(b)
(c)
.
There are nine cases to consider. In each case it must be checked whether the
resulting paths have the same size, for they will be otherwise sent to 0 by Φ.
Case 1-a. In this case ζ = νζ′ and µ = ηµ′ so that ζ = ν + ζ′ and µ =
η + µ′. We claim that µζ′ = µ + ζ′ = η if and only if ζµ′ = ζ + µ′ = ν,
and in this case µ = η and ν = ζ. In fact,
µ + ζ′ = η ⇔ η + µ′ + ζ′ = η ⇔ µ′ + ζ′ = 0 ⇔
⇔ ν + ζ′ + µ′ = ν ⇔ ζ + µ′ = ν.
Observe that, in this case, we have µ′ + ζ′ = 0 so that µ′ = ζ′ = 0, and hence
µ = η, ν = ζ.
It follows that, if µζ′ 6= η, then
ϕ(xy) = ω ◦ Φ(xy) = ω(0) = ω ◦ Φ(yσiβ(x)) = ϕ(yσiβ (x))
and, if µζ′ = η, we get
ϕ(xy) = ϕ(sµs∗
µ) = ω(sµs∗
µ) = c(µ)−βω(ps(µ))
and on the other hand
ϕ(yσiβ (x)) = c(µ)−βc(ν)β ϕ(sν s∗
ν) = c(µ)−βc(ν)β ω(sνs∗
ν) =
= c(µ)−β c(ν)βc(ν)−β ω(ps(µ)) = c(µ)−βω(ps(µ)).
Case 1-b. Now, we have that ζ = νζ′ and η = µη′ so that ζ = νζ′ = ν + ζ′
and η = µη′ = µ + η′; as before, we can check that µ + ζ′ = η if and
6
GILLES G. DE CASTRO AND FERNANDO DE L. MORTARI
only if ζ = ν + η′. If that is not the case then ϕ(xy) = 0 = ϕ(yσiβ (x)). If the
equivalent conditions are true then
ϕ(xy) = ϕ(sµζ′ s∗
η) = ω(sµζ′ s∗
η) = [µζ′ = η]c(η)−β ω(ps(η))
and
ϕ(yσiβ (x)) = c(µ)−βc(ν)β ϕ(sζs∗
νη′ ) = c(µ)−βc(ν)β [ζ = νη′]c(ζ)−β ω(ps(ζ)).
Since ζ = νζ′ and η = µη′, we have that µζ′ = η if and only if ζ = νη′ and if both
are true, then ζ′ = η′ and
c(µ)βc(ν)−β c(ζ)−β = c(µ)−βc(ν)β c(ν)−βc(η′)−β = c(µ)−βc(η′)−β =
= c(µ)−βc(ζ′)−β = c(η)−β.
From our original hypothesis, we have that s(η) = s(ζ) so we conclude that ϕ(xy) =
ϕ(yσiβ (x)).
Case 1-c. In this case ϕ(yσiβ (x)) = 0, so we need to check that ϕ(xy) = 0. As
with the previous case, we have that ϕ(xy) = [µζ′ = η]c(η)−β ω(ps(η)); however, in
case (c) µζ′ 6= η for all ζ′ and therefore ϕ(xy) = 0.
The other cases are analogous to these three, except for case 3-c, where ϕ(xy) =
0 = ϕ(yσiβ (x)) since xy = 0 = yσiβ (x).
That the correspondence obtained is bijective follows from Proposition 3.2 and
(cid:3)
that it preserves convex combinations is immediate.
Next, we want to show that there is also a bijective correspondence between
(σ, β)-KMS states on C∗(E) and a certain class of tracial states on C0(E0). We
build this correspondence by first describing a correspondence between this class of
tracial states on C0(E0) and states ω on C∗(E)γ satisfying (3.1).
The conditions found for the states on C0(E0) are similar to those in [?], al-
though as discussed in [?], their results cannot be used directly for an arbitrary
graph; nevertheless, the results of Theorem 1.1 of [?] still apply in the general set-
ting, and we use them to build a certain kind of transfer operator on the dual of
C0(E0).
Let us first recall how to construct C∗(E) as C*-algebra associated to a C*-
correspondence [?]. If we let A = C0(E0), then Cc(E1) has a pre-Hilbert A-module
structure given by
hξ, ηi (v) = Xe∈s−1(v)
ξ(e)η(e)
for
v ∈ E0,
(ξa)(e) = ξ(e)a(s(e))
for
e ∈ E1,
where ξ, η ∈ Cc(E1) and a ∈ A; it follows that the completion X of Cc(E1) with
respect to the norm given by kξk = k hξ, ξi k1/2 is a Hilbert A-module. A represen-
tation iX : A → L(X) is then defined by by
iX (a)(ξ)(e) = a(r(e))ξ(e)
for
v ∈ E0,
where L(X) is the C*-algebra of adjointable operators on X.
Let K(X) be the C*-subalgebra of L(X) generated by the operators θξ,η given
by θξ,η(ζ) = ξ hη, ζi. For each e ∈ E1, let χe ∈ Cc(E1) be the characteristic function
of {e} and observe that
(tλ =Xe∈λ
θχe,χe)λ∈Λ
,
KMS STATES FOR THE GENERALIZED GAUGE ACTION ON GRAPH ALGEBRAS
7
where Λ is the set of all finite subsets of E1, is an approximate unit of K(X) . It
is essentially the same approximate unit given by Lemma 2.10.
If τ is a tracial state on C0(E0), as in Theorem 1.1 of [?] we define a trace Trτ
on L(X) by
where T ∈ L(X).
Trτ (T ) = lim
λ→∞Xe∈λ
τ (hχe, T χei)
For a function c : E1 → R∗
+ as in the beginning of the section and β > 0,
we have that c−β ∈ Cb(E1) and so it defines an operator on L(X) by pointwise
multiplication.
Definition 3.4. Given c and β as above and τ a tracial state on C0(E0), we
define a trace Fc,β(τ ) on C0(E0) by
Fc,β(τ )(a) = Trτ (iX (a)c−β).
Now, observe that C0(E0) ∼= span{pv}v∈E0; regarding this as an equality, for
a given tracial state τ on C0(E0) we will write τ (pv) = τv. For v ∈ E0, it can be
verified that
Fc,β(τ )(pv) = lim
c(e)−βτs(e),
D→r−1(v)Xe∈D
where the limit is taken on finite subsets D of r−1(v), and Fc,β(τ )(pv) = 0 if
r−1(v) = ∅.
Remark 3.5. By Theorem 1.1 of [?], if Fc,β(τ )(a) < ∞ for all a ∈ C0(E0), then
Fc,β(τ ) is actually a positive linear functional; also, if V ⊆ E0 and Fc,β(τ )(pv) < ∞
for all v ∈ V then Fc,β(τ ) is a positive linear functional on span{pv : v ∈ V }.
Definition 3.6. For a vertex v ∈ E0 and a positive integer n, we define
r−n(v) = {µ ∈ En : r(µ) = v}.
Lemma 3.7. If Fc,β(τ )(pv) ≤ τv for all v ∈ E0 then
lim
D→r−n(v) Xµ∈D
for all v ∈ E0 and for all n ∈ N∗.
c(µ)−β τs(µ) ≤ τv
Proof. This is proved by induction. The case n = 1 is the hypothesis. Now
suppose it is true for n, then
lim
D→r−(n+1)(v) Xµ∈D
c(µ)−βτs(µ) =
lim
D→r−n(v) Xν∈r≤n
c(ν)−β Xe∈r−1(s(ν))
c(e)−βτs(e)[νe ∈ D] ≤
≤
lim
D→r−n(v)Xν∈D
c(ν)−β τs(ν) ≤ τv
where the first inequality is true due to the fact that since c is a positive function
then the net Pe∈D c(e)−βτs(e) for finite subsets D of r−1(s(ν)) is nondecreasing
and less than or equal to τs(ν) by hypothesis. The last inequality is the induction
hypothesis.
(cid:3)
The next lemma is found in [?] for unital algebras, but their proof carries out
the same in the non-unital case by using an approximate unit instead of a unit.
8
GILLES G. DE CASTRO AND FERNANDO DE L. MORTARI
Lemma 3.8 (Exel-Laca). Let B be a C*-algebra, A be a C*-subalgebra such that
an approximate unit of A is also an approximate unit of B and I a closed bilateral
ideal of B such that B = A + I. Let ϕ be a state on A and ψ a linear positive
functional on I such that ϕ(x) = ψ(x) ∀x ∈ A ∩ I and ψ(x) ≤ ϕ(x) ∀x ∈ A+, where
ψ(x) = limλ ψ(buλ) for an approximate unit {uλ}λ∈Λ of I. Then there is a unique
state Φ on B such that ΦA = ϕ and ΦI = ψ.
We want to use this lemma for A = Cn, I = Fn+1 and B = Cn+1, defined
in section 2. For that, we first note that Fn+1 is indeed an ideal of Cn+1 by
Proposition 2.11 and that the approximate unit for F0 given by Lemma 2.10 is also
an approximate unit of Cn for all n. We also need to know what the intersection
A ∩ I is, and for that we need a preliminary result.
Lemma 3.9. Suppose c, β and τ are such that Fc,β(τ )(a) ≤ τ (a) for all a ∈
C0(E0)+, then for each k ≥ 1 there is a unique positive linear functional ψk on Fk
defined by
(3.2)
ψk(sµs∗
ν) = [µ = ν]c(µ)−β τs(µ).
Proof. Since {sµs∗
tion 3.2 defines a unique linear functional on span{sµs∗
To extend to the closure, it is sufficient to prove that ψk is continuous.
ν : µ, ν ∈ Ek, s(µ) = s(ν)} is linearly independent, equa-
ν : µ, ν ∈ Ek, s(µ) = s(ν)}.
If x ∈ span{sµs∗
ν : µ, ν ∈ Ek, s(µ) = s(ν)} then
x = Xv∈V X(µ,ν)∈Gv
µ,νsµs∗
av
ν
where V is a finite subset of E0 and Gv is a finite subset of {(µ, ν) ∈ En × En :
s(µ) = s(ν) = v}. Using the decomposition given by equation 2.1 and observing
that {sµs∗
ν : (µ, ν) ∈ Gv} can be completed to generators of a matrix algebra, we
have that
k(av
µ,ν)µ,ν k
where the last norm is the matrix norm.
If Tr is the usual matrix trace we have
= max
v∈V
kxk = max
av
µ,νsµs∗
av
µ,νsµs∗
ν(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
v∈V (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
X(µ,ν)∈Gv
ψk(x) =(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
ψk
ν
Xv∈V X(µ,ν)∈Gv
=(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
µ,ν [µ = ν]c(µ)−β τs(µ)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xv∈V X(µ,ν)∈Gv
µ,ν )µ,ν diag(c(µ)−βτs(µ))(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
=(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xv∈V
≤ Xv∈V (cid:12)(cid:12)Tr((av
µ,ν )µ,ν diag(c(µ)−βτs(µ))(cid:12)(cid:12) ≤
≤ Xv∈V
µ,ν)µ,ν k Xµ:(µ,µ)∈Gv
Tr((av
k(av
av
=
=
≤
c(µ)−βτs(µ) ≤lemma 3.7
KMS STATES FOR THE GENERALIZED GAUGE ACTION ON GRAPH ALGEBRAS
9
≤ Xv∈V
k(av
µ,ν)µ,ν kτv ≤ max
v∈V
(k(av
µ,ν )µ,νk)Xv∈V
τv =
= kxkXv∈V
τv ≤ kxk
where the last inequality comes from the fact that τ comes from a probability
measure on a discrete space.
(cid:3)
Theorem 3.10. If ω is a state on C∗(E)γ satisfying (3.1) then its restriction
τ to C0(E0) satisfies:
(K1) Fc,β(τ )(a) = τ (a) for all a ∈ span{pv : 0 < r−1(v) < ∞},
(K2) Fc,β(τ )(a) ≤ τ (a) for all a ∈ C0(E0)+.
Conversely, if τ is a tracial state on C0(E0) satisfying (K1) and (K2) then there is
unique state ω on C∗(E)γ satisfying (3.1). This correspondence preserves convex
combinations.
Proof. Let ω be a state on C∗(E)γ satisfying (3.1) and τ its restriction to
C0(E0). By Remark 3.5, to establish (K1) it is sufficient to consider a = pv where
v ∈ E0 is such that 0 < r−1(v) < ∞, and in this case
τ (pv) = ω(pv) = ω
= Xe∈r−1(v)
Xe∈r−1(v)
ses∗
e
= Xe∈r−1(v)
c(e)−βτs(e) = Fc,β(τ )(pv).
c(e)−βω(ps(e)) =
For (K2), let a ∈ C0(E0)+ and write a =Pv∈E0 avpv; again, by remark 3.5 it
is sufficient to show the result for a = pv where v ∈ E0. If 0 < r−1(v) < ∞, then
we have an equality as shown above. If r−1(v) = 0, then Fc,β(τ )(pv) = 0 ≤ τ (pv).
If r−1(v) = ∞, then
Fc,β(τ )(pv) = lim
D→r−1(v)Xe∈D
D→r−1(v)Xe∈D
= lim
c(e)−βτs(e) = lim
D→r−1(v)Xe∈D
ω(ses∗
e) =
ω(pvses∗
e) ≤ ω(pv) = τ (pv).
To see the inequality above, we observe that ses∗
e are mutually orthogonal
projections that commute with pv so that
pvses∗
pv −Xe∈D
e = pv 1 −Xe∈D
ses∗
e! = 1 −Xe∈D
ses∗
e! pv 1 −Xe∈D
ses∗
e! ≥ 0.
Now, let τ be a tracial state on C0(E0) satisfying (K1) and (K2). We will use
Lemma 3.8 and the discussion after it. Observe that F0 = C0(E0) and let ψ0 = τ .
For n ≥ 1, by Lemma 3.9 there exists a positive linear functional ψn on Fn defined
by
ψn(sµs∗
ν) = [µ = ν]c(µ)−β τs(µ).
Let us show by induction that there is a unique state ϕn on Cn such that the
restriction to Fn is ψn. For n = 1, we use Lemma 3.8 with A = C0(E0), I = F1,
10
GILLES G. DE CASTRO AND FERNANDO DE L. MORTARI
B = C1, ϕ = τ and ψ = ψ1. By Proposition 2.12, in this case A ∩ I = span{pv :
v ∈ E0, 0 < r−1(v) < ∞} and if pv ∈ A ∩ I then
ψ(pv) = ψ1(pv) = ψ1(svs∗
v) = τv = τ (pv).
Using the approximate unit given by Lemma 2.10, for any v ∈ E0 we have
ψ(pv) = ψ1(pv) = lim
λ→∞
ψ1(pvuλ) = lim
D→r−1(v)Xe∈D
ψ1(ses∗
e) =
= lim
D→r−1(v)Xe∈D
c(e)−βτs(e) = Fc,β(τ )(pv) ≤ τ (pv),
where the last inequality is exactly (K2).
Now suppose that there is a unique state ϕn on Cn such that the restriction to
Fn is ψn and let us show that this is also true for n + 1. We set A = Cn, I = Fn+1,
B = Cn+1, ϕ = ϕn and ψ = ψn+1 on Lemma 3.8. By Proposition 2.12, we have
that A ∩ I = span{sµs∗
ν : µ, ν ∈ En, s(µ) = s(ν), µ = ν, 0 < r−1(s(µ)) < ∞}.
Let sµs∗
ν ∈ A ∩ I. Since 0 < r−1(s(µ)) < ∞ we have that
ψ(sµs∗
ν ) = ψn+1(sµs∗
ν) = Xe∈r−1(s(µ))
ψn+1(sµes∗
νe) = Xe∈r−1(s(µ))
[µe = νe]c(µe)−βτs(µe) =
= Xe∈r−1(s(µ))
[µ = ν]c(µ)−β c(e)−βτs(e) = [µ = ν]c(µ)−β Xe∈r−1(s(µ))
c(e)−βτs(e) =
= [µ = ν]c(µ)−βFc,β(τ )(ps(µ)) = [µ = ν]c(µ)−β τ (ps(µ)) = ψn(sµs∗
ν ) = ϕn(sµs∗
ν).
Again, using the approximate unit given by Lemma 2.10, if sµs∗
ν ∈ Cn, then
ψ(sµs∗
ν ) = ψn+1(sµs∗
ν) = lim
λ→∞
ψn+1(sµs∗
ν uλ) =
lim
D→r≤n+1−ν(s(µ))Xζ∈D
ψn+1(sµζ s∗
νζ) =
[µζ = νζ]c(νζ)−β τs(νζ) =
=
lim
D→r≤n+1−ν(s(ν))Xζ∈D
D→r≤n+1−ν(s(ν))Xζ∈D
lim
=
[µ = ν]c(ν)−β c(ζ)−βτs(ζ) =
= [µ = ν]c(ν)−β
lim
D→r≤n+1−ν(s(ν))Xζ∈D
c(ζ)−β τs(ζ) ≤
≤ [µ = ν]c(ν)−β τs(ν) = ϕn(sµs∗
ν),
where the inequality is given by Lemma 3.7, which is a consequence of (K2).
By the description of the core C∗(E)γ as an inductive limit of the Cn, we can
define a state ω as the inductive limit of ϕn. By construction, ω satisfies (3.1) and,
since each ϕn is uniquely defined by (3.1), so is ω.
Finally, it is easily seen that the correspondence built preserves convex combi-
nations by construction.
(cid:3)
KMS STATES FOR THE GENERALIZED GAUGE ACTION ON GRAPH ALGEBRAS
11
4. Ground states
In this section, we let a function c : E1 → R∗
+ be given and define a one-
parameter group of automorphisms σ as in the last section.
The following definition of a ground state will be used [?].
Definition 4.1. We say that φ is a σ-ground state if for all a, b ∈ C∗(E)a,
the entire analytic function ζ 7→ φ(aσζ (b)) is uniformly bounded in the region
{ζ ∈ C : Im(ζ) ≥ 0}, where C∗(E)a is the set of analytic elements for σ.
Proposition 4.2. If τ is a tracial state on C0(E0) such that supp(τ ) ⊆ {v ∈
E0 : v is singular} then there is a unique state φ on C∗(E) such that
(i) φ(pv) = τ (pv) for all v ∈ E0;
(ii) φ(sµs∗
ν ) = 0 if µ > 0 or ν > 0.
Proof. First, observe that a state φ satisfying (ii) is uniquely determined by
its values on C∗(E)γ because (ii) implies that φ = φC ∗(E)γ ◦ Φ, where Φ is the
conditional expectation given by Proposition 2.9.
Given τ as in the statement of the proposition, a state ω on C∗(E)γ can be
built in the same way as in the proof of Theorem 3.10. For each n, use Lemma 3.8
with A = Cn, B = Cn+1, I = Fn+1, ψn ≡ 0 and ϕn is given by the previous step,
where for the first step we have ϕ0 = τ . For ω = lim−→ ϕn, we have that φ = ω ◦ Φ
satisfies (i) and (ii) and is unique by construction.
(cid:3)
Proposition 4.3. If c is such that c(e) > 1 for all e ∈ E1, then a state φ on
ν ) = 0 whenever µ > 0 or
C∗(E) is a σ-ground state for σ if and only if φ(sµs∗
ν > 0.
Proof. If φ is a ground state then for each pair µ, ν ∈ E∗ the function ζ 7→
ν )) is bounded on the upper half of the complex plane. If ζ = x + iy then
φ(sµσζ(s∗
φ(sµσζ(s∗
ν )) = φ(sµc(ν)−iζ s∗
ν) = c(ν)y−ixφ(sµs∗
ν) = c(ν)yφ(sµs∗
ν).
If ν > 0, we have that c(ν) > 1 and so the only possibility for the above function
to be bounded is if φ(sµs∗
It is shown analogously that if µ > 0 then
φ(sµs∗
ν) = 0.
ν) = 0.
For the converse, observe that if µ = ν = 0 then φ(sµσζ (s∗
1. It can be now readily verified that if φ(sµs∗
then φ is a ground state.
ν ≤
ν) = 0 whenever µ > 0 or ν > 0
(cid:3)
ν)) = φ(sµs∗
Theorem 4.4. If c is such that c(e) > 1 for all e ∈ E1 then there is a bijective
correspondence, given by restriction, between σ-ground states φ and tracial states τ
on C0(E0) such that supp(τ ) ⊆ {v ∈ E0 : v is singular}.
Proof. This is an immediate consequence of Propositions 4.2 and 4.3. Just
note that if φ is a σ-ground state and v ∈ E0 is not singular then
φ(pv) = φ
Xe∈r−1(v)
ses∗
e
= 0.
(cid:3)
12
GILLES G. DE CASTRO AND FERNANDO DE L. MORTARI
5. Examples
In this section we give two examples with infinite graphs and study the KMS
states on the C*-algebras associated to these graphs.
Example 5.1 (The Cuntz algebra O∞). Let E0 = {v} be any unitary set and
E1 = {en}n∈N any countably infinite set with r(en) = s(en) = v ∀n ∈ N, then
C∗(E) ∼= O∞.
If c(en) = e (Euler's number) then we have the usual gauge action. In this
case, Fc,β(τ )(pv) = ∞ so that condition (K2) from Theorem 3.10 is not satisfied
and we have no KMS states for finite β. Since we have only one state on C0(E0)
and v is a singular vertex, by Theorem 4.4 there exists a unique ground state.
n=0 a−β
P∞
that Fc,β(τ )(pv) = P∞
n converges, then there exists β0 > 0 such thatP∞
Now if c(en) = an where an ∈ (1, ∞) is such that there is β > 0 for which
n = 1. Observing
and using again the fact that there exists only one
state on C0(E0), we conclude from Theorems 3.3 and 3.10 that there is no KMS
state for β < β0, there exists a unique KMS state for each β ≥ β0 and, as with the
gauge action, there is a unique ground state.
n=0 a−β
n=0 a−β
n
Example 5.2 (A graph with infinitely many sources). Let E0 = {vn}n∈N and
E1 = {en}n∈N\{0} be countably infinite sets and define r(en) = v0 and s(en) = vn
for all n ∈ N \ {0}.
Again, let an ∈ (1, ∞), n ∈ N \ {0}, be such that P∞
converges for
some β > 0. For n 6= 0 we have that Fc,β(τ )(pvn ) = 0 and for n = 0 we have
n τvn . Condition (K1) of Theorem 3.10 is trivially satisfied,
n=1 a−β
n
n=1 a−β
Fc,β(τ )(pv0 ) =P∞
and for condition (K2) we need P∞
n=1 a−β
n τvn ≤ τv0 .
If τv0 > 0, since 0 ≤ τvn ≤ 1 for all n there exists β0 > 0 such thatP∞
τv0 so that (K2) is verified for all β ≥ β0 and so there are infinitely many KMS
states. And for β < β0 (K2) is not verified so that there are no KMS states.
n=1 a−β0
n
τvn =
For ground states, since all vertices are singular, we have no restriction on τv0 ;
every state τ on C0(E0) gives a ground state on C∗(E).
Departamento de Matem´atica, Universidade Federal de Santa Catarina, 88040-970
Florian´opolis SC, Brazil.
E-mail address: [email protected], [email protected]
|
1604.00713 | 1 | 1604 | 2016-04-04T01:20:08 | Notes on ergodic theorems in non-commutative symmetric spaces | [
"math.OA"
] | In this paper we establish individual ergodic theorem for positive kernels (or so called Danford Shwartz (DS+) operators acting on non commutative symmetric spaces. | math.OA | math |
NOTES ON ERGODIC THEOREMS IN NON-COMMUTATIVE
SYMMETRIC SPACES
GENADY YA. GRABARNIK
Abstract. In this paper we establish individual ergodic theorem for positive
kernels (or so called Danford Shwartz (DS+) operators acting on non commu-
tative symmetric spaces.
1. Introduction
The goal of the paper is to see that one of the results by Veksler [20] or Muratov,
Pashkova and Rubshtein [16] remains valid for the non commutative case. 1 Let
M, τ be a semifinite von Neumann algebra with semifinite normal faithful trace τ .
In addition we assume that M, τ satisfy homogeneity property, it may be pre-
sented in the form of the resonant property on trace, see for example [1].
The space of all measurable operators affiliated with M, τ in the Sigal [19] sense
is denoted by L0, see for details [19, 5, 6].
Notions of L1(M, τ ) and L∞(M, τ ) was naturally introduced in the same paper.
Since we fix algebra M and trace τ , we omit them from the notations from now
on.
Let F ∈ M be a set of finite linear combinations of orthogonal projections with
finite trace.
Space R0 = (L1 + L∞)0 is the closure of F in the norm kxk = inf {kx1k1 +
kx2k∞, x = x1 + x2, x1 ∈ L1, x2 ∈ L∞}.
Remark 1. The space R0 is not necessary separable, it is sufficient to consider
M = B(H) of all bounded operators in the Hilbert space H with not separable H
with natural trace τ .
Definition 1. Non-commutative re-arrangement invariant (or symmetric) space L
for the fully symmetric case were introduced by Yeadon in [24]. For the definition
of the symmetric spaces we refer to the recent book of Lord, Sukochev, Zanin on
singular traces, with original proofs due to Kalton and Sukochev.
Definition 2. Re-arrangement invariant space L over (M, τ ) is called minimal if
F is dense in L by norm of L.
Date: March 1, 2016.
1991 Mathematics Subject Classification. Primary 05C38, 15A15; Secondary 05A15, 15A18.
Key words and phrases. Ergodic theory, Operator algebras, re-arrangement invariant spaces.
1 It become known to author that Litvinov and Chilin also saw this result at the same time and
wrote it at the same time. I suggested to them to combine results and names on paper. Waiting
for the answer.
1
2
GENADY YA. GRABARNIK
2. Embedding theorem
The following refinement of the embedding theorem [12] take place.
Theorem 1. Let L be a re-arrangement invariant space over (M, τ ). Then
(1) If L is minimal, then
(2) If L is not minimal L, then
L1 ∩ L∞ ⊆ L ⊆ R0
L∞ ⊆ L ⊆ L1 + L∞
Proof. Proof is given in the forthcoming paper of author and some co-authors. (cid:3)
3. Individual ergodic theorem for minimal symmetric spaces
3.1. The Largest Minimal Symmetric Space. The largest minimal symmetric
space is R0 = (L1 + L∞)0 is a set of all measurable integrable with trace operators
plus bounded with not increasing re-arrangement functions decreasing to 0.
The space R0 is minimal symmetric space. Any minimal symmetric space is a
subset of R0.
3.2. Positive double contraction on R0. Any positive kernel (T ∈ DS+) leaves
R0 invariant.
3.3. Mean Ergodic Theorem. Von Neumann Mean ergodic Theorem on L2 fol-
lows from the general von Neumann ergodic theorem for the contractions on the
general Hilbert spaces.
Mean convergence on L1 ∩L∞ follows from the the fact that both L1 and L∞ are
invariant under positive kernels. The space L1 ∩ L∞ itself is invariant under action
of positive kernel, and positive kernel is a contraction of the space L1 ∩ L∞. The
von Neumann Ergodic theorem and the closedness of L1 ∩ L∞ in its norm implies
mean ergodic theorem on L1 ∩ L∞.
3.4. Mean convergence on R0.
Proposition 1. The Cesaro averages Sn(T )x converge to some x in norm of R0.
Proof. Follows from von Neumann ergodic theorem and the fact that L2 is dense
R0. Details. We show that sl(T )x, l = 1, 2, ... is fundamental sequence in R0.
Indeed, each x ∈ R0 may be presented as x = x1,n + x2,n with x1,n ∈ L1 and
x2,n ∈ L∞ with kx2,nk∞ < 2−n. Then we can apply von Neumann or Yeadon's
Mean Ergodic theorem 4.2 [24] for L1 to x1,n and find l(n) such that for l, m ≥ l(n)
holds
ksl(T )x1,n − sm(T )x1,nkL1 < 2−n.
Then for l, m ≥ l(n)
ksl(T )x − sm(T )xkR0 ≤ ksl(T )x1,n − sm(T )x1,nkL1+
+ksl(T )x2,n − sm(T )x2,nkL∞ ≤ 4 ∗ 2−n,
and, hence, the sequence sl(T )x is fundamental. Completeness of R0 implies exis-
tence of x liml→∞ sl(T )x.
(cid:3)
NOTES ON NON-COMMUTATIVE ERGODIC THEOREMS
3
Remark 2. Mean ergodic theorem for fully symmetric spaces is due to Yeadon [24],
Theorem 4.2. Note that we do not require the space L to be fully symmetric here.
Condition ii) of the theorem 4.2 [24] means that the space L is minimal. The space
R0 does not satisfy condition iii) in theorem 4.2 [24].
3.5. Individual Ergodic Theorem in L1. Individual Ergodic theorem for L1
was established by Yeadon [22], among other authors.
3.6. Individual ergodic theorem for R0. The goal of the section is to show
double side almost everywhere convergence for the operators from the R0.
Definition 3. The sequence xn from L0 is called converging double side almost
everywhere to x0 ∈ L0 if for every ǫ > 0 there exist orthogonal projection E ∈ M
such that τ (1 − E) < ǫ, E(xn − x0)E ∈ M and E(xn − x0)E → 0.
Theorem 2. Let M, τ, R0 are as above and T is positive kernel on M . For x ∈ R0,
Cesaro averages Sn(T )x converge d.s.a.e. in R0.
Proof. The proof follows the line of the proof of Proposition 1 and uses Yeadon's
individual ergodic theorem for L1 [22], see also Chilin, Litvinov [?], [?]
We show that sl(T )x, l = 1, 2, ... is fundamental d.s.a.e. sequence in R0. Indeed,
each x ∈ R0 may be presented as x = x1,n + x2,n with x1,n ∈ L1 and x2,n ∈ L∞
with kx2,nk∞ < 2−4∗n. In turn, x1,n = x1,1,n + x1,2,n, with x1,1,n ∈ L∞, x1,2,n ∈ L1
and kx1,2,nkL1 < 2−8∗n, n = 1, 2, ....
Then we can apply Yeadon's Individual Ergodic theorem 1, [22] for L1 to x1,1,n
and find l(n) and projector E(n) ∈ M such that τ (1 − E(n)) < 2−4∗n and for
l, m ≥ l(n) holds
kE(n)(sl(T )x1,1,n − sm(T )x1,1,n)E(n)kL∞ } → 0.
We can represent x1,2,n = P∞
k=1 x2,n,k, with x2,n,k ∈ L∞ and kx2,n,kkL1 <
2−8∗(n+k).
Then, we can find E(1, n) = ∧kE(1, n, k), with τ (1 − E(1, n)) ≤ 2−4∗n and
E(1, n)sl(x1,2,n)E(1, n) ∈ L∞ and kE(1, n)sl(x2,n)E(1, n)kL∞ < 2−4∗n, where
projections E(1, n, k) are obtain by Theorem 1 from [22] applied to x2,n,k and
ǫ = 2−4(n+k).
By choosing E(2, n) = ∧∞
k=1E(1, n + k) ∧ ∧∞
k=1E(n + k), we have τ (1 − E(2, n)) <
2−4∗n.
Moreover, for l, m ≥ l(n)
kE(2, n)(sl(T )x − sm(T )x)E(2, n)kL∞
≤ kE(2, n)(sl(T )x1,n−sm(T )x1,n)E(2, n)kL∞+kE(2, n)(sl(T )x2,n−sm(T )x2,n)E(2, n)kL∞
≤ kE(2, n)(sl(T )x1,1,n − sm(T )x1,1,n)E(2, n)kL∞ + kE(2, n)(sl(T )x1,2,nE(2, n)kL∞
+kE(2, n)sm(T )x1,2,nE(2, n)kL∞ + 2 ∗ 2−4∗n ≤ 8 ∗ 2−n,
and, hence, the sequence sl(T )x is fundamental d.s.a.e. The sequence sl(T )x also
converges in norm in R0 to x, hence it converges in measure. This implies conver-
gence of sl(T )x d.s.a.e. to x.
(cid:3)
Remark 3. In the case when space L is not minimal, it contains M . Then it
is possible to show [16], that even in the commutative case, there exists ergodic
automorphism of the space with measure such that ergodic averages do not converge
almost everywhere.
4
GENADY YA. GRABARNIK
Corollary 1. Let L be a minimal non-commutative symmetric space. Let T be
a positive kernel such that T leaves L invariant and T acts as contraction on L.
Then Cesaro averages sn(T )x converge in norm and d.s.a.e. for any x ∈ L.
Proof. Since the space L is minimal, the set of L1 ∩ L∞ is dense in L. Since
L1 ∩ L∞ ⊆ L, hence
kxkL1∩L∞ ≥ C ∗ kxkL
for any x ∈ L1 ∩ L∞, which in turn implies convergence of Cesaro averages of
sn(T )x in norm of L for x ∈ L1 ∩ L∞. Fix real ǫ > 0. Find xk ∈ L1 ∩ L∞ with
kx − xkkL < ǫ/2. Then Cesaro averages are sn(T )x are within ǫ of the xk, where
xk = limn→∞ sn(T )xk. This implies norm convergence of sn(T )x.
The d.s.a.e. convergence follows from the embedding theorem 1, since a minimal
re-arrangement invariant non-commutative function space L is a subspace of R0.
(cid:3)
Corollary 2. Let L be a minimal fully symmetric space. Let T be a positive kernel
on (M, τ ). Then T leaves L invariant and act on L as contraction. Moreover, the
Cesaro averages Sn(T )x converge d.a.e. for any x ∈ L.
Corollary 3. (see Chilin Litvinov [3]). Let LΨ be a non-commutative Orlicz space
with function Ψ satisfying conditions δ2 and ∆2. Let T be a positive kernel. Then
the Cesaro averages Sn(T )x converge d.s.a.e. for any x ∈ L.
Proof. The Orlicz space LΨ with function Ψ satisfying conditions δ2 and ∆2 is
minimal [1, 12, 15]. Since the space LΨ is fully symmetric, it is interpolation space
[1, 12] and hence T leaves LΨ invariant and acts on LΨ as a contraction. Then we
are in the assumptions of the Corollary 2, and hence Sn(T )x converges d.a.e. . (cid:3)
References
[1] C. Bennett, R. Sharpley, Interpolation of Operators, Academic Press Inc. (London) LTD,
1988.
[2] V. Chilin, S. Litvinov, On pointwise ergodic theorems for infinite measure, arhiv, 2015
[3] V. Chilin, S. Litvinov, Individual ergodic theorems in noncommutative Orlicz spaces, arhiv,
2016
[4] V. I. Chilin, F. A. Sukochev, Weak convergence in non-commutative symmetric spaces, J.
Operator Theory, 31 (1994), 35-65.
[5] P. G. Dodds, T. K. Dodds, and B. Pagter, Fully symmetric operator spaces, Integr. Equat.
Oper. Theory, 15 (1992), 942-972.
[6] P. G. Dodds, T. K. Dodds, and B. Pagter, Noncommutative Kothe duality, Trans. Amer.
Math. Soc., 339(2) (1993), 717-750.
[7] P. G. Dodds, T. K. Dodds, F. A. Sukochev, and O. Ye. Tikhonov, A Non-commutative
Yoshida-Hewitt theorem and convex sets of measurable operators closed locally in measure,
Positivity, 9 (2005), 457-484.
[8] P. G. Dodds, B. Pagter and F. A. Sukochev, Sets of uniformly absolutely continuous norm
in symmetric spaces of measurable operators, Trans. Amer. Math. Soc., (2015).
[9] N. Dunford and J. T. Schwartz, Linear Operators, Part I: General Theory, John Willey
and Sons, 1988.
[10] T. Fack, H. Kosaki, Generalized s-numbers of τ -mesaurable operators, Pacific J. Math.,
123(1986), 269-300.
[11] R. V. Kadison, A generalized Schwarz inequality and algebraic invariants for operator alge-
bras, Ann. of Math. (2), 56(1952), 494-503.
[12] S. G. Krein, Ju. I. Petunin, and E. M. Semenov, Interpolation of Linear Operators,
Translations of Mathematical Monographs, Amer. Math. Soc., 54, 1982.
[13] J. Lindenstraus, L. Tsafriri, Classical Banach spaces I-II, Springer-Verlag, Berlin Heidel-
berg New York. 1977.
NOTES ON NON-COMMUTATIVE ERGODIC THEOREMS
5
[14] E. Nelson, Notes on non-commutative integration, J. Funct. Anal., 15 (1974), 103-116.
[15] B.Z. Rubshtein, G. Ya. Grabarnik, M. A. Muratov, Pashkova, SSMF, in press, 2017
[16] M. A. Muratov, J. Pashkova, B-Z. Rubshtein Order Convergence Ergodic Theorems in Re-
arrangement Invariant Spaces, Operator Methods in Mathematical Physics: Conference on
Operator Theory, Analysis and Mathematical Physics, 2013
[17] B.Z. Rubshtein, M. A. Muratov, Pashkova, Embedding theorem, oral communication, 2017
[18] S. Sakai, C ∗
−algebras, Springer-Verlag, Berlin Heidelberg New York,
−algebras and W ∗
1971.
[19] I. E. Segal, A non-commutative extension of abstract integration, Ann. of Math., 57 (1953),
401-457.
[20] A. Veksler, An ergodic theorem in symmetric spaces, Subirsk. Mat. Zh, 24 (1985), 189-191
(in Russian).
[21] H. Yanhou, T. N. Bekjan, The dual on noncommutative Lorentz spaces, Acta Math. Sci., 31
B(5)(2011), 2067-2080.
[22] F. J. Yeadon, Non-commutative Lp
−spaces, Math. Proc. Camb. Phil. Soc., 77(1975), 91-102.
[23] F. J. Yeadon, Ergodic theorems for semifinite von Neumann algebras-I, J. London Math.
Soc., 16 (2)(1977), 326-332.
[24] F. J. Yeadon, Ergodic theorems for semifinite von Neumann algebras: II, Math. Proc. Camb.
Phil. Soc., 88 (1980), 135-147.
(GYaG) St Johns University, Queens, NY, USA
E-mail address, GYaG: [email protected]
|
1611.01632 | 1 | 1611 | 2016-11-05T11:17:59 | Characterizations of 2-local derivations and local Lie derivations on some algebras | [
"math.OA"
] | We prove that every 2-local derivation from the algebra $M_n(\mathcal{A})(n>2)$ into its bimodule $M_n(\mathcal{M})$ is a derivation, where $\mathcal{A}$ is a unital Banach algebra and $\mathcal{M}$ is a unital $\mathcal{A}$-bimodule such that each Jordan derivation from $\mathcal{A}$ into $\mathcal{M}$ is an inner derivation, and that every 2-local derivation on a C*-algebra with a faithful traceable representation is a derivation. We also characterize local and 2-local Lie derivations on some algebras such as von Neumann algebras, nest algebras, Jiang-Su algebra and UHF algebras. | math.OA | math |
Characterizations of 2-local derivations and local Lie
derivations on some algebras
Jun He∗, Jiankui Li, Guangyu An, and Wenbo Huang
Department of Mathematics, East China University of Science and Technology
Shanghai 200237, China
Abstract
We prove that every 2-local derivation from the algebra Mn(A)(n > 2) into its
bimodule Mn(M) is a derivation, where A is a unital Banach algebra and M is a unital
A-bimodule such that each Jordan derivation from A into M is an inner derivation, and
that every 2-local derivation on a C*-algebra with a faithful traceable representation
is a derivation. We also characterize local and 2-local Lie derivations on some algebras
such as von Neumann algebras, nest algebras, Jiang-Su algebra and UHF algebras.
Keywords: 2-local derivation, local Lie derivation, 2-local Lie derivation, matrix
algebra, von Neumann algebra
Mathematics Subject Classification(2010): 46L57; 47B47; 47C15
1
Introduction
Let A be a Banach algebra and M an A-bimodule. We recall that a linear map D :
A → M is called a derivation if D(ab) = D(a)b + aD(b) for all a, b ∈ A, and a linear
map D : A → M is called a Jordan derivation if D(a2) = D(a)a + aD(a) for all a ∈ A.
A derivation Da defined by Da(x) = ax − xa for all x ∈ A is called an inner derivation,
where a is a fixed element in M.
In [17], R. Kadison introduces the concept of local derivation in the following sense:
a linear mapping T from A into M such that for every a ∈ A, there exists a derivation
Da : A → X , depending on a, satisfying T (a) = Da(a). Also in [17], the author proves
∗Corresponding author. E-mail address: [email protected]
1
that each continuous local derivation from a von Neumann algebra into its dual Banach
module is a derivation. B. Jonson [16] extends the above result by proving that every
local derivation from a C*-algebra into its Banach bimodule is a derivation. Based
on these results, many authors have studied local derivations on operator algebras, for
example, see in [12, 14, 19, 27].
In [28] P. Semrl introduces the concept of 2-local derivations. Recall that a map
∆ : A → M (not necessarily linear) is called a 2-local derivation if for each a, b ∈ A,
there exists a derivation Da,b : A → M such that ∆(a) = Da,b(a) and ∆(b) = Da,b(b).
Moreover, the author proves that every 2-local derivation on B(H) is a derivation.
In [18] S. Kim and J. Kim give a short proof of that every 2-local derivation on the
algebra Mn(C) is a derivation. Later J. Zhang and H. Li [30] extend the above result
for arbitrary symmetric digraph matrix algebras and construct an example of 2-local
derivation which is not a derivation on the algebra of all upper triangular complex 2 × 2
matrices.
In [2], S. Ayupov and K. Kudaybergenov suggest a new technique and prove that
every 2-local derivation on B(H) is a derivation for arbitrary Hilbert space H. Then
they consider the cases for several kinds of von Neumann algebras in succession in
[3, 4, 5], and finally prove that any 2-local derivation on arbitrary von Neumann algebra
is a derivation. Quite recently, in [6] the authors study the case for matrix algebras
over unital semiprime Banach algebras, and prove that every 2-local derivation on the
algebra M2n(A), n ≥ 2, is a derivation, where A is a unital semiprime Banach algebra
with the inner derivation property.
In Section 2, we improve the above result([6, Theorem 2.1]) for arbitrary n > 2.
More specifically, we prove that if A is a unital Banach algebra and M is a unital A-
bimodule such that each Jordan derivation from A into M is an inner derivation, then
each 2-local derivation ∆ from Mn(A)(n > 2) into Mn(M) is a derivation. Moreover,
if we only consider the case M = A, then the assumption of innerness can be relaxed
to spatial innerness. That is, if A is a unital Banach algebra such that each Jordan
derivation on A is a derivation and for each derivation D on A, there exists an element
a in B such that D(x) = [a, x] for all x ∈ A, where B is an algebra containing A, then
each 2-local derivation on Mn(A)(n > 2) is a derivation. Based on these, we obtain
some applications. We also prove that every 2-local derivation on a C*-algebra with a
faithful traceable representation is a derivation.
Recall that a linear map ϕ : A → A is called a Lie derivation if ϕ[a, b] = [ϕ(a), b] +
[a, ϕ(b)], for all a, b ∈ A, where [a, b] = ab−ba is the usual Lie product. A Lie derivation
ϕ is said to be standard if it can be decomposed as ϕ = D + τ , where D is a derivation
on A and τ is a linear map from A into the center of A such that τ [a, b] = 0 for all
a, b ∈ A.
It is natural to ask under which conditions each Lie derivation is standard. This
2
problem has been studied by many authors. M. Mathieu and A. Villena [26] prove
that each Lie derivation on a C*-algebra is standard. W. Cheung [10] characterizes
Lie derivations on triangular algebras. F. Lu [24] proves that each Lie derivation on a
CDCSL(completely distributed commutative subspace lattice) algebra is standard.
Obviously, one can define local and 2-local Lie derivations in a similar way as the
local and 2-local derivations. In [9], L. Chen, F. Lu and T. Wang prove that every local
Lie derivation on B(X) is a Lie derivation, where X is a Banach space of dimension
exceeding 2. Also in this paper [9], the authors characterize 2-local Lie derivations on
B(X). Later, L. Chen, and F. Lu [8] prove that every local Lie derivation from a nest
algebra algN into B(H) is a Lie derivation, where N is a nest on the Hilbert space H.
Quite recently, L. Liu [21] characterizes 2-local Lie derivations on a semi-finite factor
von Neumann algebra with dimension greater than 4.
In Section 3, we study local and 2-local Lie derivations on some algebras by using
a new technique. On the algebras including factor von Neumann algebras, finite von
Neumann algebras, nest algebras, UHF(uniformly hyperfinite) algebras, and the Jiang-
Su algebra, we prove that every local Lie derivation is a Lie derivation. On the algebras
including factor von Neumann algebras, UHF algebras, and the Jiang-Su algebra, we
prove that every 2-local Lie derivation is a Lie derivation. Besides, for a finite von
Neumann algebra A which is not a factor, we construct an example of 2-local Lie
derivation but not a Lie derivation on A.
2
2-Local derivations
Through this paper, we denote by Mn(A) the set of all matrices (xi,j)n×n, where
xi,j ∈ A. Clearly, if A is a Banach algebra, then so is Mn(A), and if M is an A-
bimodule, then Mn(M) is also an Mn(A)-bimodule.
Theorem 2.1. Let A be a unital Banach algebra and M be a unital A-bimodule. If each
Jordan derivation from A into M is an inner derivation, then each 2-local derivation
∆ from Mn(A)(n > 2) into Mn(M) is a derivation.
Let {ei,j}n
the linear span of the set {ei,j}n
matrices in Mn(A) by Dn(A).
i,j=1 be the system of matrix units in Mn(A) and denote by span{ei,j}n
i,j=1
i,j=1. Besides, we denote the subalgebra of all diagonal
To prove Theorem 2.1, we need several lemmas. Firstly, we give the following two
lemmas. Since the proofs are completely similar as the proofs of Lemma 2.2 and Lemma
2.9 in [6], we omit it.
Lemma 2.2. Let A be a unital Banach algebra and M be a unital A-bimodule. If each
derivation from A into M is an inner derivation, then each derivation from Mn(A)
3
into Mn(M) is also an inner derivation.
Lemma 2.3. Let A be a unital Banach algebra and M be a unital A-bimodule.
If
each Jordan derivation from A into M is an inner derivation, then for each 2-local
derivation ∆ from Mn(A)(n > 2) into Mn(M), there exists an element a ∈ Mn(M)
such that ∆Dn(A) = DaDn(A) and ∆span{ei,j}n
= Daspan{ei,j}n
.
i,j=1
i,j=1
In following Lemmas 2.4-2.7, the conditions of Theorem 2.1 hold. We consider a 2-
=
local derivation δ from Mn(A) into Mn(M) such that δDn(A) = 0 and δspan{ei,j }n
0.
i,j=1
In addition, from the definition, it is easy to see that δ is homogeneous, i.e. δ(λx) =
λδ(x), for each x ∈ Mn(A) and λ ∈ C. Thus, if necessary, we can assume that kxi,jk < 1,
and so e + xi,j is invertible in A, where 1 ≤ i, j ≤ n, xi,j is the (i, j)− entry of x and e
is the unit of A.
For each x = (xi,j)n×n ∈ Mn(A) , we denote by dxi,j the matrix in Mn(A) such that
the (i, j)− entry is xi,j and the others are zero, i.e.
dxi,j = ei,ixej,j =
0 . . .
0
...
...
0 . . . xi,j
...
...
0
0 . . .
. . .
. . .
. . .
0
...
0
...
0
Lemma 2.4. For all x ∈ Mn(A), [δ(x)]i,i = 0, i = 1, 2, · · ·, n.
Proof. Let y = ei,i + cxi,i. By lemma 2.2, there exists an element a in Mn(M) such that
δ(x) = [a, x], δ(y) = [a, y]. Since y ∈ Dn(A), we have δ(y) = 0. For any k 6= i,
0 = [δ(y)]i,k = [a, y]i,k =
nX
j=1
(ai,jyj,k − yi,jaj,k) = −yi,iai,k = −(e + xi,i)ai,k.
Since e + xi,i is invertible, it follows that ai,k = 0. Similarly, ak,i = 0, for any k 6= i.
Now we can obtain
[δ(x)]i,i = [a, x]i,i =
nX
k=1
(ai,kxk,i − xi,kak,i) = [ai,i, xi,i] = [a, y]i,i = 0.
The proof is complete.
Lemma 2.5. For each x ∈ Mn(A), let
0
...
xi,1
...
0
y =
. . .
. . . x1,j
...
xi,j
...
. . . xn,j
4
. . .
0
...
. . . xi,n
...
0
. . .
i.e. y = Pn
k=1(dxi,k + dxk,j) −dxi,j. Then [δ(x)]i,j = [δ(y)]i,j .
Proof. Take an element a from Mn(M) such that δ(x) = [a, x], and δ(y) = [a, y]. It is
easy to verify the above result by direct calculation.
Lemma 2.6. For all x ∈ Mn(A), [δ(x)]i,j = [δ(dxi,j )]i,j, whenever i 6= j.
Proof. Let
By Lemma 2.5, [δ(x)]i,j = [δ(y)]i,j . Let s 6= i, and z = y − dxs,j + es,i, i.e.
y =
nX
k=1
(dxi,k + dxk,j) −dxi,j.
0
...
xi,1
...
0
...
0
z =
. . .
0
...
. . . xi,i
...
e
...
0
. . .
. . .
. . .
. . . x1,j
...
xi,j
...
0
...
. . . xn,j
. . .
. . .
0
...
. . . xi,n
...
0
...
0
. . .
. . .
Take an element a in Mn(M) such that δ(y) = [a, y], δ(z) = [a, z]. Then
[δ(y)]i,i − [δ(z)]i,i = [a, y − z]i,i = [a,dxs,j − es,i]i,i = −ai,s.
By Lemma 2.4, [δ(y)]i,i = [δ(z)]i,i = 0, and so ai,s = 0. Thus
[δ(y)]i,j − [δ(z)]i,j = [a, y − z]i,j = [a,dxs,j − es,i]i,j = ai,sxs,j = 0,
i.e. [δ(y)]i,j = [δ(z)]i,j . Let
w = z − es,i =
0
...
xi,1
...
0
...
0
. . .
0
...
. . . xi,i
...
0
...
0
. . .
. . .
. . .
. . . x1,j
...
xi,j
...
0
...
. . . xn,j
. . .
. . .
0
...
. . . xi,n
...
0
...
0
. . .
. . .
again by Lemma 2.5, we have [δ(z)]i,j = [δ(w)]i,j . Now we have [δ(y)]i,j = [δ(y−dxs,j)]i,j.
Repeating the above steps, we can obtain that [δ(y)]i,j = [δ(u)]i,j , where
u =
nX
k=1dxi,k =
0
...
xi,1
...
0
. . .
0
...
. . . xi,n
...
0
. . .
. . .
0
...
. . . xi,j
...
0
. . .
5
Similarly, we can show that
[δ(u)]i,j = [δ(u −dxi,s + ej,s)]i,j = [δ(u −dxi,s)]i,j , f ors 6= j.
It follows that [δ(u)]i,j = [δ(dxi,j)]i,j by taking other indexes successively. Hence
[δ(x)]i,j = [δ(y)]i,j = [δ(u)]i,j = [δ(dxi,j )]i,j .
The proof is complete.
Lemma 2.7. For all x ∈ Mn(A), [δ(x)]i,j = 0, whenever i 6= j.
Proof. Assume that i < j. Let
v =
n−1X
k=1
ek+1,k =
0
e 0
0 e
...
0
0
. . .
. . .
. . .
e
0
It is not difficult to check that every matrix a in Mn(M) commuting with v satisfies
the following properties:
ak,k = as,s, ak,s = 0, whenever k < s.
Let y be a matrix in Mn(A) such that yi,i = e+ xi,j, yi,j = xi,j and the other entries of y
are all zero. Take an element a from Mn(M) such that δ(y) = [a, y], δ(v) = [a, v]. Since
v ∈ span{ei,j}n
i,j=1, δ(v) = [a, v] = 0. Thus ai,i = aj,j, and ai,j = 0. Since yj,i = 0, by
Lemma 2.6, [δ(y)]j,i = 0. Then
0 = [δ(y)]j,i = [a, y]j,i = aj,iyi,i = aj,i(e + xi,j).
Since e + xi,j is invertible, it implies that aj,i = 0. Thus
[δ(y)]i,i = [a, y]i,i = ai,iyi,i − yi,iai,i − yi,jaj,i = ai,ixi,j − xi,jai,i.
Therefore
[δ(y)]i,j = [a, y]i,j = ai,iyi,j − yi,iai,j − yi,jaj,j = ai,ixi,j − xi,jai,i = [δ(y)]i,i = 0.
Again by Lemma 2.6,
[δ(x)]i,j = [δ(y)]i,j = 0.
Similarly we can show that [δ(x)]i,j = 0 if i > j. The proof is complete.
Now we are in position to prove Theorem 2.1.
6
Proof of Theorem 2.1. According to Lemma 2.3, there exists a derivation Da such
. Let δ = ∆ − Da.
that ∆Dn(A) = DaDn(A) and ∆span{ei,j}n
= Daspan{ei,j}n
Then by Lemmas 2.4 and 2.7, we have δ = 0, i.e. ∆ = Da.
It follows that ∆ is a
derivation.
i,j=1
i,j=1
The following corollaries are some specific examples for applying Theoerm 2.1.
Corollary 2.8. Let A be a unital commutative C*-algebra and M be a unital A-
bimodule. Then each 2-local derivation from Mn(A)(n > 2) into Mn(M∗) is a deriva-
tion.
Proof. Since M is a unital A-bimodule, so is M∗. We know A is amenable, that is
each derivation from A into M∗ is an inner derivation [29]. Besides, there is a classic
result that each Jordan derivation from a C*-algebra into its bimodule is a derivation
(see [15]). By Theorem 2.1, the proof is complete.
Corollary 2.9. Let A be a C*-algebra. Then each 2-local derivation from Mn(A)(n >
2) into Mn(A∗) is a derivation.
Proof. Every C*-algebra A is weakly amenable, that is, each derivation from A into
A∗ is an inner derivation [29]. By Theorem 2.1, the proof is complete.
Corollary 2.10. Each 2-local derivation on the matrix algebra Mn(A)(n > 2) is a
derivation, if the algebra A satisfies one of the following conditions:
(1) A = algL, where L is a subspace lattice on a Hilbert space H with the property that
0+ 6= 0 or H− 6= H;
(2) A = alg(N1N N2N · · ·N Nn), where each Ni is a nest, i = 1, 2, · · ·, n.
Proof. (1) According to [22, Theorem 2.1], each Jordan derivation on A is a derivation.
And see in [20] , each derivation on A is an inner derivation.
(2) It is known that
N1N N2N · · ·N Nn is still a CSL, so Jordan derivations on A are derivations [23]. In
[13], F. Gilfeather, A. Hopenwasser and D. Larson prove that the cohomology vanishes
for CSL algebras whose lattices are generated by finite independent nests. Applying
this result, we immediately obtain that derivations on A are inner derivations. By
Theorem 2.1, the proof is complete.
Corollary 2.11. Let A be a nest algebra of infinite multiplicity. Then each 2-local
derivation on A is a derivation.
Proof. Since A is of infinite multiplicity, it is isomorphic to Mn(B) for some nest algebra
B and some integer n > 2. It is known that each derivation on nest algebras is an inner
derivation [11]. By Theorem 2.1, the proof is complete.
7
Lemma 2.12. Let A = L∞
i=1 Ai be a Banach algebra with the inner derivation prop-
erty, i.e. all derivations on A are inner derivations. If each 2-local derivation on Ai is
a derivation for any i ∈ N, then each 2-local derivation on A is also a derivation.
Proof. We only consider the case A = A1L A2. For a 2-local derivation δ on A, denote
by δi the restriction of δ in Ai, i = 1, 2. One can easily verify that δ(ai) = δi(ai) ∈ Ai
for each ai ∈ Ai, moreover, δi is a 2-local derivation on Ai, and thus a derivation.
For each a ∈ A, we write a = a1 + a2, where ai ∈ Ai, i = 1, 2. By the definition of 2-
local derivation, there exists an element x ∈ A such that δ(a) = [a, x] = [a1, x1]+[a2, x2]
and δ(a1) = [a1, x] = [a1, x1]. Similarly, there exists an element y ∈ A such that
δ(a) = [a, y] = [a1, y1] + [a2, y2] and δ(a2) = [a2, y] = [a2, y2]. So we can obtain
δ(a) = δ(a1) + [a2, x2] = [a1, y1] + δ(a2).
It follows that δ(a1) = [a1, y1], δ(a2) = [a2, x2] and thus δ(a) = δ(a1) + δ(a2). Hence δ
is linear. Moreover, we can obtain
δ(ab) = δ(a1b1 + a2b2)
= δ(a1b1) + δ(a2b2)
= δ(a1)b1 + a1δ(b1) + δ(a2)b2 + a2δ(b2)
= (δ(a1) + δ(a2))(b1 + b2) + (a1 + a2)(δ(b1) + δ(b2))
= δ(a)b + aδ(b)
Hence δ is a derivation on A. The proof is complete.
Corollary 2.13. Let A be a von Neumann algebra without direct summand of type I1
and I2. Then each 2-local derivation on A is a derivation.
Proof. It is well known that A = L∞
i=1 Ai, where each Ai is isomorphic to a matrix
algebra Mn(B) for some von Neumann algebra B and some integer n > 2. By Theorem
2.1 and Lemma 2.12, the result follows.
Remark 2.14. In [4] and [5], the authors show that each 2-local derivation on a von
Neumann algebra is a derivation in other ways. By comparison, our proof is more
simple. However, we can not handle the case for type I2. In addition, for the case of
type I1, the result is trival, since each derivation on an abelian von Neumann algebra
is zero.
Corollary 2.15. Let A be a unital algebra with the inner derivation property and n ≥ 6
be a positive integer but not a prime number. Then each 2-local derivation on the matrix
algebra Mn(A) is a derivation.
8
Proof. Suppose n = rt, where r > 2 and t > 1. Then Mn(A) is isomorphic to
Mr(Mt(A)). In [1], the author proves that each Jordan derivation on Mt(A)(t > 1) is
a derivation. By Theorem 2.1, the proof is complete.
Through the same technique with the proof of Theorem 2.1, we can show the fol-
lowing theorem.
Theorem 2.16. Let A be a unital Banach algebra such that:
(1) each Jordan derivation on A is a derivation;
(2) for each derivation D on A, there exists an element a in B such that D(x) = [a, x]
for all x ∈ A, where B is an algebra containing A.
Then each 2-local derivation on Mn(A)(n > 2) is a derivation.
Corollary 2.17. Let A be a C*-algebra. Then each 2-local derivation on Mn(A)(n > 2)
is a derivation.
Let A be a C*-algebra, π a representation of A, and M the von Neumann algebra
generated by π(A). Then π is said to be a traceable representation if there exists a
faithful normal trace τ on M+ such that π(A)T Mτ is weakly dense in M, where Mτ
denotes the span of the set {M ∈ M+ : τ (M ) < ∞}. Especially, a finite representation
the von Neumann algebra generated by π(A) is finite) can be regarded as a
π(i.e.
specific traceable representation.
Theorem 2.18. Let A be a C*-algebra with a faithful traceable representation π. Then
each 2-local derivation on A is a derivation.
Proof. Let M be the von Neumann algebra generated by π(A), and τ a faithful normal
trace on M+ such that Aτ is weakly dense in M, where Aτ denotes π(A)T Mτ . It is
known that Mτ is a two-side ideal of M, thus Aτ is also a two-side ideal of π(A).
For a 2-local derivation ∆ on A, define δ = π ◦∆◦π−1. Then δ is a 2-local derivation
on π(A). For each x ∈ π(A) and y ∈ Aτ , there exists an element m ∈ M such that
δ(x) = [m, x] and δ(y) = [m, y]. Hence
τ (δ(x)y) = τ ((mx − xm)y) = τ (mxy) − τ (xmy)
= τ (xym) − τ (xmy) = τ (x(ym − my))
= −τ (xδ(y)).
For each a, b ∈ π(A) and x ∈ Aτ , we have
τ (δ(a + b)x) = −τ ((a + b)δ(x)) = −τ (aδ(x)) − τ (bδ(x))
= τ (δ(a)x) + τ (δ(b)x)
= τ ((δ(a) + δ(b))x).
9
It means that τ (ux) = 0 for any x ∈ Aτ , where u denotes δ(a + b) − δ(a) − δ(b). For
any y ∈ Aτ , by taking x = yy∗u∗, we obtain τ (uyy∗u∗) = 0. Since τ is faithful and Aτ
is weakly dense, we have uy = 0 and thus u = 0.
Hence δ is additive, and thus a local derivation, moreover, a derivation. It follows
that ∆ = π−1 ◦ δ ◦ π is a derivation on A. The proof is complete.
As direct applications of the above theorem, we have the following corollaries.
Corollary 2.19. Each 2-local derivation on a UHF C*-algebra is a derivation.
Corollary 2.20. Each 2-local derivation on the Jiang-Su algebra is a derivation.
3 Local and 2-local Lie derivations
In this section, we study local and 2-local Lie derivations on some algebras. The
main results are as follows.
Theorem 3.1. Let B be a Banach algebra over C satisfying the following conditions:
(1) each Lie derivation on B is standard;
(2) each derivation on B is inner;
(3) each local derivation on B is a derivation;
(4) every nonzero element in Z(B)(the center of B) can not be written in the form
Pn
i=1[Ai, Bi], where Ai, Bi ∈ B.
Then each local Lie derivation ϕ on B is a Lie derivation.
Proof. By the definition of local Lie derivation, for any A ∈ B, there exists a Lie
derivation ϕA such that ϕA(A) = ϕ(A). Since ϕA is standard, ϕA = δA + τA, where δA
is a derivation and τA is a linear mapping from B into Z(B) such that τA[X, Y ] = 0,
for all X, Y ∈ B. Thus
ϕ(A) = δA(A) + τA(A) = [A, TA] + τA(A)
for some TA ∈ B, which means that ϕ has a decomposition at each point.
following we prove that the decomposition is unique. Suppose
In the
ϕ(A) = δ
A(A) + τ
A(A) = [A, T
A] + τ
′
′
′
′
A(A).
Then
Thus
[A, TA] + τA(A) = [A, T
′
A] + τ
′
A(A).
[A, TA] − [A, T
′
A] = τ
′
A(A) − τA(A).
10
Since every nonzero element in Z(B) can not be written in the form Pn
i=1[Ai, Bi], we
obtain
[A, TA] − [A, T
′
A] = τ
′
A(A) − τA(A) = 0,
and so
Now we can define
δA(A) = δ
A(A), τA(A) = τ
′
′
A(A).
δ(A) = δA(A) = [A, TA] and τ (A) = τA(A),
for every A ∈ B. We shall prove δ and τ are linear. For all A, B ∈ B, since
ϕ(A + B) = ϕ(A) + ϕ(B),
we obtain that
[A, TA] + [B, TB] − [A + B, TA+B] = τ (A + B) − τ (A) − τ (B).
Hence
[A + B, TA+B] = [A, TA] + [B, TB], τ (A + B) = τ (A) + τ (B).
It means that δ and τ are additive. Similarly, we can prove that δ and τ are homoge-
neous, and so linear. Hence δ is a local derivation, by assumption, a derivation, and τ
is a linear mapping from B into Z(B) such that τ [X, Y ] = 0, for all X, Y ∈ B. It follows
that ϕ is a Lie derivation. The proof is complete.
Corollary 3.2. Each local Lie derivation on a factor von Neumann algebra M is a
Lie derivation.
Proof. It is known that every element in M can be written in the form λI+Pn
Pn
every element in M has the form Pn
i=1[Ai, Bi],
where λ ∈ C and Ai, Bi ∈ M(see, [7]). If the unit I can not be written in the form
i=1[Ai, Bi], where Ai, Bi ∈ M, then by Theorem 3.1, the result follows. Otherwise,
i=1[Ai, Bi]. Hence each Lie derivation is a deriva-
tions, so every local Lie derivation is a Lie derivation.
The following result is proved in [8]. We show it by Theorem 3.1 as an alternative
method.
Corollary 3.3. Each local Lie derivation on a nest algebra A = algN is a Lie deriva-
tion.
Proof. If A is of infinite multiplicity, then every A ∈ A has the form Pn
i=1[Ai, Bi],
where Ai, Bi ∈ A([25, Theorem 4.7]). In this case, all Lie derivations are derivations,
and so the result is trival.
11
Otherwise, suppose A is not of infinite multiplicity. Then N has a finite-dimensional
atom P . It is known that A satisfies the conditions (1) to (3) of Theorem 3.1(see in
[24, 11, 14]), so it is sufficient to check the condition (4). We know Z(A) = CI. Assume
i=1[Ai, Bi]. It is known that the linear
map σ : A → P AP defined by σ(A) = P AP for all A ∈ A is a homomorphism. So we
have
that the unit I can be written in the form Pn
P = P IP =
nX
i=1
P [Ai, Bi]P =
[σ(Ai), σ(Bi)].
nX
i=1
It follows that there is no trace on P AP , which contradicts that P AP has finite di-
i=1[Ai, Bi]. The proof is
mension. Therefore I can not be written in the form Pn
complete.
Theorem 3.4. Let B be an algebra over C satisfying the following conditions:
(1) each Lie derivation on B is standard;
(2) each derivation on B is inner;
(3) each local derivation on B is a derivation;
(4) there exists a center-valued trace on B.
Then each local Lie derivation ϕ on B is a Lie derivation.
Proof. Since there exists a center-valued trace on B, every nonzero element in Z(B)
i=1[Ai, Bi], where Ai, Bi ∈ B. By Theorem 3.1, the
can not be written in the form Pn
result follows.
Corollary 3.5. Each local Lie derivation on a finite von Neumann algebra M is a Lie
derivation.
Proof. Since M is finite, there exists a central valued trace on M. By Theorem 3.4,
the result follows.
Remark 3.6. For a properly infinite von Neumann algebra M, since every element
i=1[Ai, Bi], where Ai, Bi ∈ M(see, [7]), all Lie derivations are
in M has the form Pn
derivations, so every local Lie derivation is a Lie derivation.
Also as applications of Theorem 3.4, we can immediately obtain the following two
corollaries.
Corollary 3.7. Each local Lie derivation on a UHF C*-algebra A is a Lie derivation.
Corollary 3.8. Each local Lie derivation on the Jiang-Su algebra Z is a Lie derivation.
Theorem 3.9. Let B be an algebra over C satisfying the following conditions:
(1) each Lie derivation on B is standard;
(2) each derivation on B is inner;
(3) each 2-local derivation on B is a derivation;
12
(4) Z(B) = CI, where Z(B) is the center of B;
i=1[Ai, Bi], where λ ∈ C and
(5) every element in B can be written in the form λI +Pn
Ai, Bi ∈ B.
Then each 2-local Lie derivation ϕ on B is a Lie derivation.
Proof. Similarly with the foregoing discussion, if I can be written in the formPn
then the result is trival. We assume that I can not be written in the form Pn
i=1[Ai, Bi].
In this case, we have proved that ϕ has a unique decomposition at each point A in B,
i.e.
i=1[Ai, Bi],
ϕ(A) = ϕA(A) = δA(A) + τA(A),
where ϕA is a Lie derivation, δA is a derivation and τA is a linear mapping from B into
Z(B) such that τA[X, Y ] = 0, for all X, Y ∈ B. Similarly we can define
δ(A) = δA(A) and τ (A) = τA(A),
for all A ∈ B. Now it is sufficient to show that δ is a derivation and τ is linear. Given
A and B in B, there exists a Lie derivation ϕA,B such that
and
ϕ(A) = ϕA,B(A) = δA,B(A) + τA,B(A),
ϕ(B) = ϕA,B(B) = δA,B(B) + τA,B(B),
where δA,B + τA,B is the standard decomposition of ϕA,B. By the uniqueness of the
decomposition, δ(A) = δA,B(A) and δ(B) = δA,B(B). Hence δ is a 2-local derivation
and thus a derivation. Now Our task is to show that τ is linear. It is easy to see that
i=1[Ai, Bi], then the
i=1[Ai, Bi],
ϕ is homogeneous, and so is τ . If both A and B have the form Pn
equation τ (A + B) = τ (A) + τ (B) is trival. Otherwise, suppose A = λI +Pn
where λ 6= 0. Take a Lie derivation ϕA,A+B such that
ϕ(A) = ϕA,A+B(A) = δA,A+B(A) + τA,A+B(A),
and
ϕ(A + B) = ϕA,A+B(A + B) = δA,A+B(A + B) + τA,A+B(A + B),
where δA,A+B + τA,A+B is the standard decomposition of ϕA,A+B. From
τ (A) = τA,B(A) = τA,A+B(A),
we have τA,B(I) = τA,A+B(I), and thus
τA,B(B) = τA,A+B(B).
Now we can obtain that
τ (A + B) = τA,A+B(A + B) = τA,A+B(A) + τA,A+B(B)
= τ (A) + τA,B(B) = τ (A) + τ (B).
Hence τ is additive, and so linear. The proof is complete.
13
Directly applying the above theorem, we can obtain the following corollaries.
Corollary 3.10. Each 2-local Lie derivation on the Jiang-Su algebra is a Lie deriva-
tion.
Corollary 3.11. Each 2-local Lie derivation on a UHF C*-algebra is a Lie derivation.
Corollary 3.12. Each 2-local Lie derivation on a factor von Neumann algebra is a Lie
derivation.
Example 3.13. If M is a finite von Neumann algebra but not a factor, then there
exists a 2-local Lie derivation on M but not a Lie derivation.
Proof. There exist a center-valued trace τ on M and a non-trival central projection P
in M. Suppose the Hilbert space that M acts on is H. Take two unital vectors x and
y in P H and (I − P )H, respectively. For each element M ∈ M, denote < τ (M )x, x >
and < τ (M )y, y > by xM and yM , respectively.
Define a functional f on M by follows: f (M ) equals to x2
M
yM
if yM 6= 0, else 0,
and define a mapping δ from M into Z(M) by δ(M ) = f (M )P . Obviously, δ is not
linear and thus not a Lie derivation. Now we only need to check that δ is a 2-local Lie
for each A, B in M, there exists a Lie derivation φ on M such that
derivation, i.e.
φ(A) = δ(A) and φ(B) = δ(B).
case1 xAyB 6= xByA.
Take
φ(M ) =
(f (A)yB − f (B)yA)xM + (f (B)xA − f (A)xB)yM
P.
xAyB − xByA
It is not difficult to check that φ is a linear mapping from M into Z(M) and φ([X, Y ]) =
0 for all X, Y in M. Hence φ is a Lie derivation. Moreover, φ(A) = δ(A) and φ(B) =
δ(B).
case2 xAyB = xByA.
If both yA and yB are zero, then it is sufficient to take φ ≡ 0. Otherwise, without
P . One can easily
= 0 and
loss of generality, we may assume yA 6= 0, then take φ(M ) = xAxM
yA
check that φ is a Lie derivation and φ(A) = δ(A). If yB = 0, then xB = xAyB
yA
φ(B) = 0 = δ(B). If yB 6= 0, then φ(B) = xAxB
yA
P = x2
B
yB
P = δ(B).
The proof is complete.
Acknowledgements. This paper was partially supported by National Natural Science
Foundation of China(Grant No. 11371136).
References
[1] R. Alizadeh, Jordan derivations of full matrix algebras, Linear Algebra Appl. 430
(2009), 574-578
14
[2] S. Ayupov, K. Kudaybergenov, 2-Loacl derivations and automorphisms on
B(H), J. Math. Anal. Appl. 395(1)(2012), 15-18.
[3] S. Ayupov, K. Kudaybergenov, B. Nurjanov, A. Alauatdinov, Local
and 2-local derivations on noncommutative Arens algebras, Mathematica Slovaca
64(2014), 423-432.
[4] S. Ayupov, F. Arzikulov, 2-Local derivations on semi-finite von Neumann al-
gebras, Glasgow Math. J. 56(2014), 9-12.
[5] S. Ayupov, K. Kudaybergenov, 2-Local derivations on von Neumann algebras,
Positivity 19(3)(2014), 445-455.
[6] S. Ayupov, K. Kudaybergenov, 2-Local derivations on matrix algebras over
semi-prime Banach algebras and on AW*-algebras, Journal of Physics Conference
Series 697(2016).
[7] M. Bresar, E. Kissin, S. Shulman, Lie ideals:
from pure algebra to C*-
algebras, J. reine angew. Math. 623 (2008), 73121.
[8] L. Chen, F. Lu, Local Lie derivations of nest algebras, Linear Algebra Appl.
475(2015), 62-72.
[9] L. Chen, F. Lu, T. Wang, Local and 2-local Lie derivations of operator algebras
on Banach spaces, Integr. Equ. Oper. Theory 77(2013), 109-121.
[10] W. Cheung, Lie derivations of triangular algebra, Linear Multilinear Algebra
51(2003), 299-310.
[11] E. Christensen, Derivations of nest algebras, Math. Ann 229(1977), 155-161.
[12] R. Crist, Local derivations on operator algebras, J. Funct. Anal. 135(1996), 72-
92.
[13] F. Gilfeather, A. Hopenwasser, D. Larson, Reflexive algebras with finite
width lattices: tensor products, cohomology, compact perturbations, J. Funct.
Anal. 55(2)(1984), 176-199.
[14] D. Hadwin, J. Li, Local derivations and local automorphisms on some algebras,
J. Operator Theory 60(1)(2008), 29-44.
[15] N. Jacobson, C. Rickart, Jordan homomorphisms of rings, Trans. Amer. Math.
Soc. 69(3)(1950), 479-502.
[16] B. Johnson, Local derivations on C*-algebras are derivations, Trans. Amer. Math.
Soc. 353(2001), 313-325.
[17] R. Kadison, Local derivations, J. Algebra 130(1990), 494-509.
[18] S. Kim, J. Kim, Local automorphisms and derivations on Mn(C), Proc. Amer.
Math. Soc. 132(5)(2004), 1389-1392.
15
[19] D. Larson, A. Sourour, Local derivations and local automorphisms, Proc. Sym-
pos. Pure Math. 51(1990), 187-194.
[20] J. Li, H. Pendharkar, Derivations of certain algebras, Internat. J. Math.
24(5)(2000), 345-349.
[21] L. Liu, 2-Local Lie derivations on semi-finite factor von Neumann algebras, Linear
Multilinear Algebra 64(2016), 1679-1686.
[22] F. Lu, Jordan derivations of reflexive algebras, Integr. Equ. Oper. Theory
67(2010), 51-56.
[23] F. Lu, The Jordan structure of CSL algebras, Studia Math. 190(2009), 283-299.
[24] F. Lu, Lie derivation of certain CSL algebras, Israel J. Math. 155(2006), 149-156.
[25] L. Marcoux, A. Sourour, Conjugation-invariant subspaces and Lie ideals in
non-selfadjoint operator algebras, J. London Math. Soc. 65(2)(2002), 493-512.
[26] M. Mathieu, A. Villena, The structure of Lie derivations on C*-algebras, J.
Funct. Anal. 202(2003), 504-525.
[27] Y. Pang, W. Yang, Derivations and local derivations on strongly double triangle
subspace lattice algebras, Linear Multilinear Algebra 58(2010), 855-862.
[28] P. Semrl, Local automorphisms and derivations on B(H), Proc. Amer. Math.
Soc. 125(1997), 2677-2680.
[29] V. Runde, Lectures on Amenability, I. Springer Verlag, 2002.
[30] J. Zhang, H. Li, 2-Loacl derivations on digraph algebras, Acta Math. Sinica,
Chinese series 49(2006), 1401-1406.
16
|
1811.04043 | 3 | 1811 | 2019-05-03T16:19:05 | The universal C*-algebra of a contraction | [
"math.OA",
"math.FA"
] | We say that a contractive Hilbert space operator is universal if there is a natural surjection from its generated C*-algebra to the C*-algebra generated by any other contraction. A universal contraction may be irreducible or a direct sum of (even nilpotent) matrices; we sharpen the latter fact and its proof in several ways, including von Neumann-type inequalities for noncommutative *-polynomials. We also record properties of the unique C*-algebra generated by a universal contraction, and we show that it can be used similarly to C*(F_2) in various Kirchberg-like reformulations of Connes' Embedding Problem (some known, some new). Finally we prove some analogous results for universal C*-algebras of noncommuting row contractions and universal Pythagorean C*-algebras. | math.OA | math |
THE UNIVERSAL C∗-ALGEBRA OF A CONTRACTION
KRISTIN COURTNEY AND DAVID SHERMAN
Abstract. We say that a contractive Hilbert space operator is universal if
there is a natural surjection from its generated C ∗-algebra to the C ∗-algebra
generated by any other contraction. A universal contraction may be irreducible
or a direct sum of (even nilpotent) matrices; we sharpen the latter fact and its
proof in several ways, including von Neumann-type inequalities for noncom-
mutative *-polynomials. We also record properties of the unique C ∗-algebra
generated by a universal contraction, and we show that it can be used simi-
larly to C ∗(F2) in various Kirchberg-like reformulations of Connes' Embedding
Problem (some known, some new). Finally we prove some analogous results
for universal C ∗-algebras of row contractions and universal Pythagorean C ∗-
algebras.
Contents
Introduction
1.
2. Universal contraction operators
3. Von Neumann-type inequalities for noncommutative *-polynomials
4. Universal C∗-algebras and projectivity
5. Finite-dimensional and nilpotent representations of A
6. A and Connes' Embedding Problem
6.1. The WEP, LLP, and some results of Kirchberg
6.2. A versus C∗(F2)
7. C∗-algebras generated by multiple universal contractions
7.1. Universal row contractions
7.2. Universal Pythagorean C∗-algebras
References
1
3
7
10
12
17
18
20
22
22
25
26
1. Introduction
For T a contractive linear operator on a Hilbert space, let C∗(T, I) denote the
unital C∗-algebra generated by T and the identity I. We say that T is a universal
contraction if for any other contractive Hilbert space operator S there is a unital
*-homomorphism from C∗(T, I) to C∗(S, I) taking T to S. This is equivalent to
requiring that for any noncommutative *-polynomial q, the norm of q(T ) is as large
Date: May 6, 2019.
2010 Mathematics Subject Classification. 46L05.
Key words and phrases. universal contraction, universal C ∗-algebra, von Neumann's inequal-
ity, residual finite dimensionality, primitivity, nilpotent operators, Connes Embedding Problem.
The research of the first-named author was partially supported by the Eric Nordgren Research
Fellowship Fund at the University of New Hampshire.
1
2
KRISTIN COURTNEY AND DAVID SHERMAN
as it can be for a contraction. (It should not be confused with "universal (model)
operators" as introduced by Rota [47, 48], although we explain a connection in
Remark 2.8.)
One goal of this paper is to describe universal contractions themselves. On a
given separable infinite-dimensional Hilbert space, universal contractions comprise
a single approximate unitary equivalence class that is strong* dense in the set of
all contractions (Propositions 2.6 and 2.7). But there are qualitative differences
among them: a single universal contraction can be irreducible (Theorem 2.5) or a
direct sum of nilpotent matrices (Theorem 5.1). The latter result is known, but the
existing proof is quite intricate. In Section 5 we give two alternative proofs, and in
Section 7.1 we give a generalization.
If T1 and T2 are universal contractions, then C∗(T1, I) is *-isomorphic to C∗(T2, I)
via the map sending T1 to T2. Thus we may call this C∗-algebra (equipped with its
distinguished generator) the universal unital C∗-algebra of a contraction. We will
denote it here by A, and we will denote the non-unital C∗-algebra C∗(T1) ≃ C∗(T2)
by A0. (It is indeed non-unital; see Remark 4.3.) The familiar reader will notice
that A can be identified with the universal unital C∗-algebra associated to the op-
erator space C, as in [43], and the maximal C∗-dilation of the disk algebra, as in
[6]. In the usual generator-relation notation for universal C∗-algebras, we write
A0 = C∗hx : kxk ≤ 1i.
Adding a unit to the generating set and relations above gives us the unitization of
A0, which is exactly A. The two are essentially interchangeable in this paper, but
we often make use of the unit and so choose A for the main discussion.
A C∗-algebra is called residually finite dimensional (RFD) if it has a separating
family of finite-dimensional representations. The fact that a universal contraction
can be a direct sum of matrices says that A is RFD and gives us a *-polynomial
analogue of von Neumann's inequality: the maximal norm of any noncommutative
*-polynomial whose entry ranges over contractive Hilbert space operators can be
determined by considering only contractive (even nilpotent) matrices. In Section
3 we sharpen this by showing that the maximal norm is actually achieved at a
contractive matrix whose dimension depends only on the degree of the *-polynomial
(Theorem 3.1). But not all elements of A attain their norm in a finite-dimensional
representation (Proposition 3.3). In Section 4 we discuss the projectivity of A as a
unital C∗-algebra, which is stronger than RFD [35, Lemma 8.1.4] and implies that
A has trivial topological invariants such as K-theory and shape theory. Note that
any separable RFD algebra has a faithful tracial state, so A is a finite C∗-algebra.
On the other hand, the algebra A is the mother of all singly-generated unital
C∗-algebras, so it also has various "largeness" properties. Because it surjects onto
a non-exact algebra, it is not exact and a fortiori not nuclear ([8, Corollary 1] or
[6, Section 6.1], although "K ⊗ B" in the latter should be unitized). Because it
surjects onto a non-type I algebra (e.g., a UHF algebra, singly generated by [50]),
it is not type I. This leads to comparisons with other large C∗-algebras, like the
universal C∗-algebra of a partial isometry, to which A is Morita equivalent [8], and
full C∗-algebras of free groups. In fact several of the results for A in Section 2 are
patterned on corresponding results for C∗(F2) obtained in the beautiful 1980 paper
of Choi [12]. Note that C∗(F2) is not singly generated as a unital C∗-algebra, so it
cannot be a quotient of A. However, A is a quotient of C∗(F2), and they embed
into each other relatively weakly injectively (Theorem 6.10). Using Theorem 6.10
THE UNIVERSAL C ∗-ALGEBRA OF A CONTRACTION
3
and results from Kirchberg's seminal work [31], we give in Theorem 6.11 several
characterizations of Connes' Embedding Problem in terms of A (some of these were
deduced differently in [43, 6]).
In Section 7 we adapt some of our arguments and results for A to other C∗-
algebras generated by multiple universal contractions. Theorem 7.4 generalizes
Theorem 5.1 to row contractions. This leads to a Popescu-von Neumann inequality
for noncommutative *-polynomials on row contractions (Corollary 7.6). Section 7.2
establishes that the universal Pythagorean C∗-algebras from [9] are RFD.
Our notation throughout the paper is fairly standard. For an operator T ∈ B(H),
we denote its spectrum by σ(T ) and its essential spectrum by σe(T ). We denote
the open unit disk and unit circle of C by D and T, respectively. We write Mn for
the complex n × n matrices and freely associate it with B(Cn).
After this article was completed, the authors became aware of unpublished work
from 1989 by Froelich and Salas along similar themes. Upon request Prof. Salas
graciously shared their manuscript; most of the overlap concerns results in Section
2. We decided to make no changes to the present article except mention of the
reference [17].
The authors are grateful to David Blecher, Scott Atkinson, and an anonymous
referee for useful comments on a draft of this article, and to Don Hadwin for many
valuable discussions and perspectives. Some of this material is taken from the
first-named author's 2018 PhD dissertation at the University of Virginia [15].
2. Universal contraction operators
We start with an easy observation.
Proposition 2.1. The following are equivalent for a contractive operator T on a
Hilbert space.
(1) The operator T is a universal contraction; i.e., for any other contractive
Hilbert space operator S, the assignment T 7→ S induces a *-homomorphism
from C∗(T, I) to C∗(S, I).
(2) For any noncommutative *-polynomial q,
(2.1)
kq(T )k = sup
kq(S)k,
S
where the supremum is taken over all contractive Hilbert space operators.
Proof. We have that (1) implies (2) because *-homomorphisms are contractive, and
(2) implies (1) because the assignment q(T ) 7→ q(S) densely defines a continuous
*-homomorphism between the C∗-algebras.
(cid:3)
How can we produce universal contractions? For the reader versed in universal
C∗-algebras defined by generators and relations, it is clear that A is a separable
nonzero algebra (see for instance [35, Section 3.1]). So it has a faithful represen-
tation on ℓ2, and the image of its distinguished generator is a separably-acting
universal contraction.
The next proposition exhibits a universal contraction that is a little more con-
crete. It is "known to the experts" and seems to have been first mentioned as a
tool in the proof of [25, Theorem 5.1], but we lack a reference that provides the
details. Since it plays an important role in this paper, for the benefit of the reader
we give a short argument using only basic operator theory. (It also can be proved
by appealing to the projectivity of A; see the text after Remark 4.3.)
4
KRISTIN COURTNEY AND DAVID SHERMAN
Proposition 2.2.
(1) The supremum in (2.1) is unchanged if S ranges only over matrix contrac-
tions.
(2) Let T be any direct sum of contractive matrices such that the n × n sum-
mands are dense in (Mn)≤1 for each n. Then T is a universal contraction.
(3) The algebra A is RFD.
Proof. All of these statements effectively say the same thing, so it suffices to prove
(1). (To see that A is RFD, note that restriction to the matrix summands of the
operator in (2) gives a separating family of representations.)
Let S be a contractive operator on the Hilbert space H. For F a finite-dimensional
subspace of H, let PF be the projection onto F . Then the net {PF SPF }, ordered
by inclusion of the subspaces F , converges strong* to S. For any noncommuta-
tive *-polynomial q, we also have q(PF SPF ) → q(S) strong*. By strong* lower
semicontinuity of the norm, we have
kq(S)k = ks∗ − lim q(PF SPF )k ≤ lim inf kq(PF SPF )k.
Now PF SPF is the direct sum of a matrix and some multiple of the 1 × 1 zero
operator. It follows that kq(S)k is not more than the maximal norm of q evaluated
on contractive matrices. (In fact it is a maximum; see Theorem 3.1 below.)
(cid:3)
In Section 5 we give proofs and variations for the much harder fact, basically due
to Herrero, that a universal contraction can be built as a direct sum of nilpotent
matrices.
So must a universal contraction be a direct sum of some sort? No, we show
below in Theorem 2.5 that there are irreducible universal contractions. This is
equivalent to proving that A is primitive, meaning that it has a faithful irreducible
representation. Our argument mimics Choi's proof of primitivity for C∗(F2) [12,
Theorem 6], with additional reliance on the fact that A is RFD.
Proposition 2.3. The algebra A contains no nontrivial projections.
faithful representation of A, then π(A) contains no nonzero compact operators.
If π is a
Proof. This can be proved in the same way as [12, Theorem 1 and Corollary 2]. (cid:3)
Lemma 2.4. Let T be a universal contraction and K be a compact operator. If
T + K is a contraction, then it is a universal contraction.
Proof. Let T and K be as stated, let q be a noncommutative *-polynomial, and
let π : B(H) → B(H)/K(H) be the Calkin map.
It follows from Proposition
2.3 that the restriction of π to C∗(T, I) is isometric. Since π(q(T )) = q(π(T )) =
q(π(T + K)) = π(q(T + K)), we have
kq(T + K)k ≥ kπ(q(T + K))k = kπ(q(T ))k = kq(T )k
and are done by Proposition 2.1.
(cid:3)
Theorem 2.5. The algebra A is primitive. Equivalently, there exist irreducible
universal contractions.
Proof. Choi used the following lemma, based on techniques of Radjavi-Rosenthal
[46, Theorem 7.10 and Theorem 8.30]. Let A and B be operators on the same
Hilbert space, expressed as matrices with respect to a basis. If A is diagonal with
distinct entries, and B has no zero entries in its first column, then A and B have
THE UNIVERSAL C ∗-ALGEBRA OF A CONTRACTION
5
no common nontrivial reducing subspace. We will find a universal contraction
whose real and imaginary parts can be taken as A and B above. Since a projection
commutes with an operator if and only if it commutes with the real and imaginary
parts, the contraction will have no nontrivial reducing subspaces.
As in Proposition 2.2, construct a universal contraction T = ⊕∞
j=1Tj on H =
j=1Cnj as a direct sum of matrices Tj ∈ Mnj . We may choose the matrices Tj
⊕∞
to be strict contractions, and we may choose the ordering so that kTjk < 1 − 1
j .
By taking unitary conjugates we may also assume that the real part of each Tj is
diagonal.
For each j ≥ 1, perturb the diagonal entries of Re Tj by less than 1
2j so that
the entries are distinct from each other and all diagonal entries of Ti for i < j.
This perturbs T to T ′ so that Re T ′ is diagonal with distinct entries, and the cor-
responding summands have kT ′
2j . Because this is a compact perturbation,
T ′ is still a universal contraction by Lemma 2.4.
jk < 1 − 1
Now perturb T ′ to T ′′ as follows. We only change the first column and row so
that these are everywhere nonzero in Im T ′′. We go one block at a time, taking
advantage of the little bit of norm wiggle room. In the first block T ′
1, change all
entries in the first column and row by some small identical pure imaginary amount
λ1, making them all nonzero and keeping kT ′′
3 . Then change the part of
the first column and row of T ′ corresponding to the second block - these are all zero
in T ′ - by a small identical pure imaginary amount λ2, making them all nonzero
and preserving that the submatrix of T ′′ corresponding to the first two blocks has
norm < 1 − 1
j. Visually, we are perturbing
T ′ by the operator
6 . Continue in this way for all blocks T ′
1 k < 1 − 1
· · · λ1 λ2
· · · λn
· · ·
0
R =
λ1 λ1
λ1
...
λ1
λ2
...
λn
...
where each λj ∈ iR is repeated nj times.
The submatrix of T ′′ = T ′ + R corresponding to the initial string of n blocks has
norm < 1 − 1
3n . The operator T ′′ is a strong limit of these submatrices (considered
as infinite-dimensional operators by filling in the rest of the matrix with zeroes),
so T ′′ is also a contraction. And because we have only changed the first row and
column to go from T ′ to T ′′, it is a compact perturbation, and T ′′ is still a universal
contraction by Lemma 2.4.
Notice that the matrix for Re T ′′ = Re T ′ is diagonal with distinct entries,
and the matrix for Im T ′′ has nonzero entries in the first column. By the result
mentioned at the beginning of the proof, we are done.
(cid:3)
It is apparent from the preceding results that separably-acting universal con-
tractions need not be unitarily equivalent, but the truth is not so far from that.
Recall that two operators are said to be approximately unitarily equivalent (a.u.e.),
6
KRISTIN COURTNEY AND DAVID SHERMAN
denoted here ∼a, if one is the norm limit of unitary conjugates of the other. Two
representations of the same C∗-algebra are a.u.e. if one representation is the point-
norm limit of unitary conjugates of the other.
Parts of the next two propositions are noted or implied in Hadwin's work on
Voiculescu's noncommutative Weyl-von Neumann theorem [24, 25, 26] and the sub-
sequent discussion in [17, Section 7]. We include some citations but make all the
reasoning explicit here.
Proposition 2.6. Separably-acting universal contractions form a single a.u.e. class.
Proof. Let T1 and T2 be separably-acting universal contractions. By sending the
distinguished generator of A to Tj, we obtain two faithful representations of A whose
ranges contain no nontrivial compact operators by Proposition 2.3. It then follows
from Voiculescu's noncommutative Weyl-von Neumann theorem [51, Corollary 1.4]
that the representations are a.u.e., which means that T1 ∼a T2.
If S1 ∼a S2, then for any noncommutative *-polynomial q we have q(S1) ∼a q(S2)
and kq(S1)k = kq(S2)k. So any operator a.u.e. to a universal contraction must also
be one by Proposition 2.1.
(cid:3)
Proposition 2.7. For T ∈ B(ℓ2)≤1, the following are equivalent.
(1) T is a universal contraction.
(2) T ∼a (T ⊕ S) for every S ∈ B(ℓ2)≤1.
(3) The strong* closure of the unitary orbit of T is B(ℓ2)≤1.
Proof. (1) ⇐⇒ (2): We first show that the operators on ℓ2 satisfying the condition
in (2) form a single a.u.e. class. If T1, T2 satisfy the condition, then
And if T ∼a T1 and S ∈ B(ℓ2)≤1, then
T1 ∼a T1 ⊕ T2 ∼a T2.
T ⊕ S ∼a T1 ⊕ S ∼a T1 ∼a T.
The conclusion then follows from Proposition 2.6 and Hadwin's observation [26,
Example 7.3(1)] that a universal contraction constructed in Proposition 2.2(2) sat-
isfies the condition in (2).
(2) ⇒ (3): Hadwin shows in [24, Corollary 3.4 and Theorem 4.3] that for any
T ∈ B(ℓ2), the strong* closure of the unitary orbit of T is the set of operators that
are summands of operators a.u.e. to T .
(3) ⇒ (1): Assume (3), and let S be any contraction on ℓ2. Let {Uj} be unitaries
with U ∗
j T Uj → S strong*, and let q be a noncommutative *-polynomial. Then
kq(S)k = kq(s∗ − lim(U ∗
j T Uj))k = ks∗ − lim q(U ∗
j T Uj)k
= ks∗ − lim U ∗
j q(T )Ujk ≤ lim inf kU ∗
j q(T )Ujk = kq(T )k.
Thus T is universal by Proposition 2.1.
(cid:3)
Remark 2.8. "Universality" has various meanings for operators, some connected to
our situation and some not. In the context of operator ideals it is a factorization
property (e.g., [41]) -- this is quite different from the present paper.
In the definition perhaps most familiar to an operator theorist, a Hilbert space
operator T is said to be universal, or be a universal model, if every separably-
acting operator can be scaled to become similar to the restriction of T to some
invariant subspace. See [33, Chapter 6] for a discussion of the area. The terminology
THE UNIVERSAL C ∗-ALGEBRA OF A CONTRACTION
7
originates with Rota [47], who proved in [48] that B, the backward shift with infinite
multiplicity, is a universal model in a slightly more precise sense: any operator
with spectral radius < 1 is similar to the restriction of B to an invariant subspace.
Proposition 2.7(2) characterizes the universality of T (as defined in this paper)
analogously: any operator with norm ≤ 1 is unitarily equivalent to the restriction
of an operator a.u.e. to T to a reducing subspace.
Operator theorists may be tempted to look for universal contractions among
the weighted shifts, but that cannot succeed because weighted shifts are centered
[39]; they satisfy relations saying that the family {T mT ∗m, T ∗nT n : m, n ≥ 0}
is commutative.
In fact the authors know of no explicit "naturally-occurring"
universal contractions. Our constructions all rely in some way on the direct sum of
a dense set.
3. Von Neumann-type inequalities for noncommutative *-polynomials
Let Mz denote multiplication by z on L2(T, m), and let p be a polynomial. From
compactness, the maximum modulus principle, and spectral theory, we have
(3.1)
[sup or max]¯Dp(λ) = [sup or max]Tp(λ) = kp(Mz)k.
The celebrated von Neumann inequality [40] is the fact that for any contraction S,
the norm of p(S) is dominated by the quantity in (3.1).
We may then say that Mz is a universal contraction for polynomials. (In fact
a contraction has this property if and only if its spectrum contains T.)1 Equiva-
lently, Mz generates the universal unital operator algebra2 of a contraction, which
is isometrically isomorphic as a Banach algebra to the disk algebra A(D) via the
map densely defined by p ↔ p(Mz) [7, Example 2.2]. All of the foregoing is still
true if we replace p with any f ∈ A(D), and so von Neumann's inequality opens the
door for an analytic functional calculus for all contractions. (For any contraction
S, f (S) is well-defined by von Neumann's inequality as the limit of pj(S), where
{pj} is any sequence of polynomials converging to f in A(D).)
In summary, von Neumann's inequality says that there are sufficiently many one-
dimensional representations of the universal unital operator algebra of a contraction
to determine the norm of a polynomial in the generator (and thus we can identify
the algebra with A(D) and the one-dimensional representations with ¯D). In fact
the norm is achieved. The smaller set T suffices to determine the norm, which is
still achieved. For any element of A(D) the norm is achieved at a one-dimensional
representation.
We pursue analogues of every clause in the preceding paragraph, in the context
of noncommutative *-polynomials and C∗-algebras, and call these analogues (when
valid) "von Neumann-type inequalities."
Let x be the canonical generator of A, S any contraction, and q a noncommu-
tative *-polynomial. By definition we have that kq(S)k ≤ kq(x)k. We typically
cannot determine kq(x)k with the one-dimensional representations of A (i.e., re-
placing x with scalars): consider q(z) = z∗z − zz∗. But the finite-dimensional
1The backward implication holds because the operator norm is not smaller than the spectral
radius. For the forward implication, let λ ∈ T and T be a contraction with this property. The
spectral radius of (T + λ) is lim k(T + λ)nk1/n = lim k(z + λ)nk1/n
C(T) = 2, which forces λ ∈ sp(T ).
2Here an operator algebra is a (not necessarily self-adjoint) norm-closed subalgebra of some
B(H), or any matrix-normed Banach algebra that can be represented as such.
8
KRISTIN COURTNEY AND DAVID SHERMAN
representations of A do suffice (i.e., replacing x with contractive matrices). We
already proved this in Proposition 2.2(1), which may thus be considered a von
Neumann-type inequality:
(3.2)
kq(S)k ≤ kq(x)k =
sup
kq(M )k.
M contractive matrix
Here are our questions. Note that there is no compactness available.
(1) Is the supremum in (3.2) achieved?
(2) Can we replace the contractive matrices with a smaller natural set? And
will the supremum still be achieved?
(3) Since noncommutative *-polynomials in x form a dense set in A, it follows
from the above that for any a ∈ A, kak = supπ kπ(a)k, as π ranges over
the finite-dimensional representations of A. Is this supremum achieved?
Here are the answers, the first and third of which we proceed to show in the
remainder of this section.
(1) Yes (Theorem 3.1).
(2) Contractive nilpotent matrices suffice, but the supremum is not achieved
in general (Theorem 5.1 and Remark 5.2).
(3) No, it is not achieved for all elements of A, but it is for some elements that
are not of the form q(x) (Proposition 3.3).
In [16, Theorem 3.2] it was shown that a C∗-algebra is RFD exactly when it
has a dense subset of elements that attain their norm under a finite-dimensional
representation of the algebra. In some cases, this dense subset contains all non-
commutative *-polynomials in the standard generators. For instance, Fritz, Netzer,
and Thom prove in [22, Lemma 2.7] that every element in CFn attains its norm
under some finite-dimensional representation of C∗(Fn), where Fn is a free group
on n ≤ ∞ generators. Our proof of Theorem 3.1 below is a simplified version of
the proof of [22, Lemma 2.7], which itself is an adaptation of Choi's argument in
[12, Theorem 7] that C∗(Fn) is RFD.
The degree of a noncommutative *-polynomial is the length of its longest mono-
mial.
Theorem 3.1. Let q be a noncommutative *-polynomial of degree d, and let x be
the canonical generator of A. Then
(3.3)
kq(x)k = max{kq(M )k : M ∈ M2d+1, kM k ≤ 1}.
Proof. Let ϕ be a state on A with ϕ(q(x)∗q(x)) = kq(x)k2, and let (π, H, ξ) be the
associated GNS representation. Then kq(x)k = kπ(q(x))ξk. Define
H0 = span{π(g(x))ξ : g is a *-monomial of length ℓ(g) ≤ d}.
Note that dim(H0) ≤ 2d+1. Let V be the inclusion of H0 in H, so that V V ∗ is the
projection in B(H) onto H0. By construction
(3.4)
V q(V ∗π(x)V )V ∗ξ = q(π(x))ξ,
since V V ∗ acts as the identity everywhere in the expansion of the left-hand side.
We can think of V ∗π(x)V ∈ B(H0) as a contractive matrix of size ≤ 2d+1, so
the assignment x 7→ V ∗π(x)V induces a unital *-homomorphism π0 : A → B(H0).
THE UNIVERSAL C ∗-ALGEBRA OF A CONTRACTION
9
Compute
kq(π0(x))(V ∗ξ)k ≥ kV q(π0(x))V ∗ξk
= kV q(V ∗π(x)V )V ∗ξk = kq(π(x))ξk = kπ(q(x))ξk = kq(x)k.
Thus the norm of q(π0(x)) must be kq(x)k (it cannot be bigger). Finally note that
B(H0) can be (not necessarily unitally) included in M2d+1 .
(cid:3)
The reader may wonder if A has a separating family of representations of bounded
(finite) dimension, so that the maximum in (3.3) can be taken in some fixed MN
for all q. (For polynomials N = 1 works!) This is not so: for a C∗-algebra, the
existence of a separating family of representations of dimension ≤ N implies that
all irreducible representations have dimension ≤ N (see [18, Proposition 3.6.3(i)]).
Actually, it follows from [16, Theorem 5.1] that there are elements of A whose
maximal norm in n-dimensional representations grows according to any prescribed
finite pattern in n. But the methods in [16] are nonconstructive, so that one cannot
exhibit such elements, and they are unlikely to be noncommutative *-polynomials
in the generator. As a complement to these results we display here explicit non-
commutative *-polynomials that achieve their maximal norm at any prescribed
dimension and no lower.
Example 3.2. For any n ≥ 1, consider the noncommutative *-polynomial
qn(z) = (z∗z + z∗2z2 + ... + z∗nzn) + (1 − zz∗).
We claim that the maximal norm of qn(M ), where M ranges over all contractions,
is attained in Mn+1 but no smaller matrix algebra.
We have that qn(M ) is a sum of n + 1 positive contractions, so its norm cannot
be more than n + 1. Taking Sn to be the forward shift in Mn+1, qn(Sn) sends the
first basis element e1 to (n + 1)e1, so it has the maximal norm n + 1.
Now suppose that M is a matrix such that the positive matrix qn(M ) has norm
n + 1. Let v be a unit eigenvector for qn(M ) for the value n + 1. It follows that
v is fixed by all of M ∗M, . . . , M ∗nM n and annihilated by M M ∗; this means that
M ∗v = 0 and M jv is a unit vector for j = 1, . . . n. We have M v ∈ ran(M ) ⊥
ker(M ∗) ∋ v. We also have M 2v ∈ ran(M 2) ⊥ ker(M ∗2) ∋ v, M v. (For the last
membership, note M ∗2(M v) = M ∗(M ∗M v) = M ∗v = 0.) Continuing, we get that
v, M v, M 2v, . . . , M nv form an orthonormal set of (n + 1) vectors on which M acts
as the forward shift. Thus M is a matrix of size at least (n + 1) × (n + 1).
Let AF denote the set of elements in A that achieve their norm under some
finite-dimensional representation, i.e., a ∈ AF if there exists a finite-dimensional
representation π of A such that kak = kπ(a)k. Then Theorem 3.1 says that the
set P∗ of noncommutative *-polynomials in x is contained in AF .
It is natural
to ask whether every element in A attains its norm under some finite-dimensional
representation, i.e., whether or not AF is the entire space. The first-named author
and T. Shulman explored this question for general C∗-algebras in [16]. As the
next proposition shows, not every element of A achieves its norm under a finite-
dimensional representation, but there are some elements in A\P∗ that do.
Proposition 3.3. Let AF be defined as above. Then
P∗ ( AF ( A.
10
KRISTIN COURTNEY AND DAVID SHERMAN
Proof. By [16, Theorem 4.4], we know that AF ( A if and only if A has a simple,
infinite-dimensional AF subquotient. The CAR algebra M2∞ is a simple, unital,
infinite-dimensional AF C∗-algebra, which is singly generated by [50]. Hence, it is
isomorphic to a quotient of A, and so
AF ( A.
It also follows from [16, Theorem 5.5] that AF is not closed under addition (or
multiplication) and hence cannot equal P∗, i.e.,
P∗ ( AF .
(cid:3)
The proof of Proposition 3.3 is again nonconstructive. A specific element of
AF \ P∗ is ex∗x, which attains its norm of e in the one-dimensional representation
of A sending x to 1. More interestingly, we can use Example 3.2 to exhibit an
element of A \ AF .
Example 3.4. For any contractive element y of a C∗-algebra, consider the norm
convergent series
q(y) =
2−n qn(y)
n + 1
,
∞
Xn=1
where the qn are as in Example 3.2. Since kqn(x)k = n + 1, we have kq(x)k ≤ 1.
Taking S to be the unilateral shift on ℓ2 and e1 the first standard basis vector, we
also have kq(S)k ≥ kq(S)e1k = ke1k = 1. Thus kq(x)k = 1. But it follows from
Example 3.2 that kq(M )k < 1 for any finite-dimensional contraction M .
4. Universal C∗-algebras and projectivity
In this section, we consider properties of A as a universal C∗-algebra. Because
other universal C∗-algebras will soon make appearances, now is a good time to
introduce them formally. We will forgo constructions and refer the reader to those
given in [35, Chapter 3] or [5, Section 1]. Though they can be defined in more
generality, we will restrict ourselves to separable universal C∗-algebras defined by
imposing norm constraints and noncommutative *-polynomial relations on the gen-
erators.
Let G = {x1, ..., xn} be a finite set and R a finite set of relations of the form
• r(x1, ..., xn) ≥ 0, or
• ks(x1, ..., xn)k ≤ C for some C ≥ 0
where r and s are noncommutative *-polynomials. To fend off existence issues, we
require that the relations in R enforce norm bounds on the elements of G and that
there exists some tuple (A1, ..., An) of operators on some Hilbert space that satisfy
R. The universal C∗-algebra with generators G subject to relations R is denoted by
C∗hG : Ri and has the following defining universal property: given Hilbert space
operators T1, ..., Tn satisfying R, the assignments xi 7→ Ti, 1 ≤ i ≤ n, induce a
surjective *-homomorphism C∗hG : Ri → C∗(T1, ..., Tn).
If C∗hG : Ri has a unit, then the *-homomorphism is assumed to be unital.
If it is not unital, then we denote its unitization by C∗
uhG : Ri; we also adopt
the convention of identifying a unital C∗-algebra with its unitization. Note that
unitizing a universal C∗-algebra is the same as adding a unit to the generating set,
i.e.,
C∗h{y1, ..., yn} ∪ {I} : R ∪ {I = I ∗ = I 2, yiI = yi = Iyi, 1 ≤ i ≤ n}i.
THE UNIVERSAL C ∗-ALGEBRA OF A CONTRACTION
11
To see this, we faithfully represent
C∗h{y1, ..., yn} ∪ {I} : R ∪ {I = I ∗ = I 2, yiI = yi = Iyi, 1 ≤ i ≤ n}i
as C∗(IH, Y1, ..., Yn) on some Hilbert space H. By universality, we have a surjective
*-homomorphism C∗hG : Ri → C∗(Y1, ..., Yn) induced by sending xi 7→ Yi. This
extends to a surjective unital *-homomophism C∗
uhG : Ri → C∗(IH, Y1, ..., Yn),
which is injective by the universality of the latter algebra. In particular, this means
adding a unit to the generators of A0 and unitizing A0 give the same algebra, namely
A. (We shall see in Remark 4.3 why A0 cannot have a unit.) As a convention, we
say the unitization of the zero C∗-algebra is C.
Now we are ready to change gears and introduce one of the most important
properties of A. We call a C∗-algebra A projective if given any C∗-algebra B with
closed two-sided ideal I and quotient map π : B → B/I, any *-homomorphism
φ : A → B/I lifts to a *-homomorphsim ψ : A → B so that φ = π ◦ ψ. In other
words, a projective C∗-algebra is a projective object in the category of C∗-algebras
with *-homomorphisms. By [5, Proposition 2.5], a C∗-algebra is projective if and
only if its unitization is projective in the category of unital C∗-algebras with unital
*-homomorphisms. For simplicity, if we call a unital C∗-algebra projective, we
mean in the category of unital C∗-algebras with unital *-homomorphisms.
Proposition 4.1. [35, Lemma 8.1.4] Both A0 and A are projective.
The projectivity of A follows from that of A0, which is a consequence of the fact
that any contraction in a C∗-quotient lifts to a contraction.
Remark 4.2. We mentioned in the introduction that C∗(F2) is not projective. This
means exactly that unitaries do not in general lift to unitaries, e.g., unitaries in the
Calkin algebra with nonzero Fredholm index.
Remark 4.3. By [5, Proposition 2.3], a projective C∗-algebra is contractible, which
implies that it is non-unital and has trivial shape in the sense of [5] and trivial K-
theory. (We say a unital C∗-algebra is contractible when it is homotopy equivalent
to C.) That A is contractible and has trivial K1 is also proved in [8, Theorem 3.1].
Projectivity is a powerful (and rare) property because many other nice properties
can be rephrased in terms of lifting problems. For instance, we can quickly establish
Propositions 2.2 and 2.3. Proposition 2.2 is a special case of the following.
Proposition 4.4. Any separable projective C∗-algebra is RFD.
See [35, Theorem 11.2.1] or [27] for a full proof. The main ideas are that any
separable C∗-algebra is the quotient of an RFD C∗-algebra, and that residual fi-
nite dimensionality passes to subalgebras. For projective objects, being a quotient
implies being a subobject.
Meanwhile Proposition 2.3 follows from the next proposition, the proof of which
uses only the statement of [12, Proposition 1] as opposed to its proof.
Proposition 4.5. Any separable projective C∗-algebra B has no nontrivial projec-
tions, and if (π, H) is a faithful representation of B, then π is essential (i.e., π(B)
contains no nonzero compact operators).
Proof. It will suffice to prove the unital case. To that end, let B be a separable
unital C∗-algebra that is projective as a unital C∗-algebra. Then B is isomorphic
to a quotient of C∗(F∞). By projectivity, this isomorphism lifts to an embedding
12
KRISTIN COURTNEY AND DAVID SHERMAN
of B into C∗(F∞). Since C∗(F∞) has no nontrivial projections [12, Theorem 1],
neither does B.3 It then follows, by the same argument as for [12, Corollary 2], that
any faithful representation of B on a Hilbert space H trivially intersects K(H). (cid:3)
Contractibility makes it difficult to distinguish projective C∗-algebras, such as A
i , i = 1, 2i.
and the free product C[−1, 1] ∗C C[−1, 1] ≃ C∗
(That this particular C∗-algebra is projective follows from [19, Theorem 3.2].)
uhx1, x2 : kxik ≤ 1, xi = x∗
Question 4.6. Are A and C[−1, 1] ∗C C[−1, 1] isomorphic?
We point out why the most obvious guess at an isomorphism does not answer
Question 4.6. Let x be the distinguished generator of A, and b, c be the canonical
generators of the C[−1, 1] factors in the free product. Since Re x and Im x are
self-adjoint contractions with spectrum [−1, 1], one might propose an isomorphism
via Re x 7→ b, Im x 7→ c. But this fails, because x = Re x + iIm x is a contraction,
while b+ic has norm 2. (One can deduce this from the representation of C[−1, 1]∗C
i
0(cid:19), and b + ic to (cid:18)0
2
0
0(cid:19).)
C[−1, 1] sending b to (cid:18)0
1
1
0(cid:19), c to (cid:18) 0
−i
Consider the "abelianized" version of this question. The universal unital C∗-
algebra of a normal contraction is C(D), and the tensor product C[−1, 1] ⊗ C[−1, 1]
is C([−1, 1]2). These two C∗-algebras are isomorphic, because the closed unit disk
is homeomorphic to the closed (solid) 2 × 2 square. But the homeomorphism is
more complicated than simply rescaling coordinates. Question 4.6 is asking for
a noncommutative version of this homeomorphism. With poetic license: are the
noncommutative disk and noncommutative solid square homeomorphic?
5. Finite-dimensional and nilpotent representations of A
In this section we discuss the rather surprising fact that a universal contrac-
tion can be a countable direct sum of nilpotent matrices. This was originally
proved by Herrero via intricate computations; here we give two alternative argu-
ments. The first, "`a la Choi" and fairly short, proceeds by dilating slightly rescaled
finite-dimensional compressions of a universal contraction to contractive nilpotent
matrices. The second, which we later generalize in the proof of Theorem 7.4, asymp-
totically factorizes a faithful representation of A through the universal C∗-algebras
of nilpotents of increasing order, relying on the classical nilpotent approximation
theorem of Apostol-Foias-Voiculescu and a result of Shulman on lifting nilpotent
contractions.
Theorem 5.1. There exists a separating family of finite-dimensional representa-
tions of A that map the generator to contractive nilpotent matrices. In other words,
for any noncommutative *-polynomial q and contractive Hilbert space operator S,
kq(S)k ≤ sup{kq(M )k : M ∈ Mn, kM k ≤ 1, M n = 0, n ≥ 1}.
Theorem 5.1 gives us a different refinement (from Theorem 3.1) of the von
Neumann-type inequality we got from the residual finite-dimensionality of A in
Section 3.
Since the representations guaranteed by Theorem 5.1 are separating, by [16,
Theorem 3.2], A has a dense subset of elements that attain their norm under one
3This fact also follows from [28, Proposition 3].
THE UNIVERSAL C ∗-ALGEBRA OF A CONTRACTION
13
of these representations. However, this dense subset need not contain the canonical
dense subset P∗ of noncommutative *-polynomials in the generator, as the next
example shows.
Example 5.2. Let x be the canonical generator of A. Then x+x∗ cannot attain its
norm under a finite-dimensional representation mapping x to a nilpotent operator.
First note that kx + x∗k = 2, which can be seen by mapping x to 1. Now,
suppose M ∈ Mn is nilpotent with kM k = 1. Since M + M ∗ is self-adjoint, if
kM + M ∗k = 2, then there exists a unit vector v such that either
(M + M ∗)v = 2v or (M + M ∗)v = −2v.
Because the unit vector v is an extreme point of the unit ball, it follows that M v = v
or M v = −v, either way violating the fact that σ(M ) = 0.
Now, let us commence with the proofs of Theorem 5.1. The first proof is truly
due to Herrero, though he does not draw this explicit conclusion. Let T be a
universal contraction as constructed in Proposition 2.2(2). Herrero shows in [30,
Corollary 4.8] that T is the norm limit of contractive nilpotent operators that
are block diagonal. Hence by Proposition 2.6, the same holds for any universal
contraction on B(ℓ2), and we have the following extension of [30, Corollary 4.8].
Proposition 5.3. Every universal contraction in B(ℓ2) is the norm limit of block-
diagonal nilpotent contractive operators with finite-dimensional blocks.
Theorem 5.1 follows by taking the family of representations induced by mapping
the generator x to each of the blocks in the sequence.
Remark 5.4. Note that no universal contraction T ∈ B(H) can be the norm limit
of nilpotent contractions that lie in C∗(T, I). Indeed, if N ∈ C∗(T, I) is nilpotent
and π : C∗(T, I) → C maps T 7→ 1, then
kT − N k ≥ kπ(T ) − π(N )k = k1 − 0k = 1.
Thanks to Don Hadwin for originally pointing out this fact.
Our second argument proves Theorem 5.1 directly by an "asymptotic dilation"
argument. This is close in spirit to Choi's proof [12, Theorem 7] that C∗(F2) is
RFD, but the details are somewhat different.4
Proof. For each n > 1, let Jn ∈ Mn denote the n×n Jordan block with eigenvalue 0.
By [23, Proposition 1], there is a unit vector ξn ∈ Cn so that w(Jn) = hJnξn, ξni =
cos π
n+1 , where w(Jn) is the numerical radius of Jn. Let Un ∈ Mn be a unitary so
that Une1 = ξn (where e1 is from the standard basis of Cn). For each n > 1, define
Mn := U ∗
nJnUn = (a(n)
ij )1≤i,j≤n.
Then Mn is a nilpotent contraction with a(n)
11 = cos π
n+1 .
Let H = ℓ2 with the standard basis {e1, e2, ...}; let Pn be the orthogonal projec-
tion of H onto span{e1, ..., en}; let T be a universal contraction operator on H; and
let Tn := PnT Pn for each n > 1. Then Tn ⊗ Mn ∈ PnB(H)Pn ⊗ Mn is a nilpotent
contraction, and, if we view PnB(H)Pn as Mn, then the assignments T 7→ Tn ⊗ Mn
4Arveson has characterized in [3, Theorem 1.3.1] the contractive Hilbert space operators that
can be power dilated to contractive nilpotent operators -- for powers less than the order of the
nilpotent. But we need all *-monomials, at least asymptotically.
14
KRISTIN COURTNEY AND DAVID SHERMAN
induce a family of finite-dimensional representations πn : C∗(T, I) → Mn2 where
πn(T ) is nilpotent. To see that this family is separating, it will suffice to show that
⊕nπn is isometric, which will follow from showing that kq(T )k = supn kq(Tn ⊗Mn)k
for any noncommutative *-polynomial q. To that end, let q be any nonzero non-
commutative *-polynomial, and let ε > 0.
For each n > 1, define Cn = (c(n)
ij ) ∈ Mn−1 by c(n)
ij
:= a(n)
i+1,j+1 for each 1 ≤
i, j ≤ n − 1. Since a(n)
11 = cos π
n+1 → 1 as n → ∞, we have
k[a(n)∗
1j
]2≤j≤nkℓ2
n
as n → ∞, and so
→ 0 and k[a(n)
i1 ]2≤i≤nkℓ2
n
→ 0
a(n)
12
0
...
0
0
a(n)
21
...
a(n)
1n
≤ k[a(n)∗
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
1j
... a(n)
1n
0
...
...
0
...
...
+ k[a(n)
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
kMn −(cid:16)cos( π
n+1 ) ⊕ Cn(cid:17) k =
]2≤j≤nkℓ2
n
i1 ]2≤i≤nkℓ2
n
→ 0
as n → ∞, which implies that
kTn ⊗ Mn − Tn ⊗ (cos( π
n+1 ) ⊕ Cn)k → 0
as n → ∞. Now, q is Lipschitz on the unit ball of Mn2 ≃ PnB(H)Pn ⊗ Mn with
Lipschitz constant independent of n. From this we see that
(5.1)
kq(Tn ⊗ Mn) − q((Tn ⊗ (cos( π
n+1 ) ⊕ Cn))k −−−−→
n→∞
0.
Assume for simplicity that kq(T )k = 1. Since Tn
and we have
S ∗OT−−−−→ T , so does cos( π
n+1 )Tn,
(5.2)
kq(cos( π
n+1 )Tn)k > 1 −
ε
2
for n sufficiently large. Combining (5.1) and (5.2), we have, for n sufficiently large,
kq(Tn ⊗ Mn)k > kq(Tn ⊗ (cos( π
n+1 ) ⊕ Cn))k −
ε
2
= kq((Tn ⊗ cos( π
n+1 )) ⊕ (Tn ⊗ Cn))k −
= kq((Tn ⊗ cos( π
n+1 )) ⊕ q(Tn ⊗ Cn)k −
≥ kq(cos( π
n+1 )Tn)k −
ε
2
= 1 − ε.
> 1 −
ε
2
−
ε
2
ε
2
ε
2
Since ε was arbitrary, this shows that ⊕nπn is an isomorphism.
(cid:3)
Our final proof of Theorem 5.1 will require a few preliminary results. Though
it is a little more work on the outset, this approach is what we will generalize to
prove Theorem 7.4, at which point we will appreciate the preliminaries established
here. We begin by recalling a few definitions.
THE UNIVERSAL C ∗-ALGEBRA OF A CONTRACTION
15
Definition 5.5. A separable C∗-algebra A is quasidiagonal (QD) if there exists a
sequence of completely positive contractive maps φn : A → Mkn that are asymp-
totically multiplicative and asymptotically isometric.
It is easy to see that a separable RFD C∗-algebra is QD since it embeds faithfully
into some product Qn Mkn .
Definition 5.6. A set of operators Ω ⊆ B(H) is called quasidiagonal (QD) if there
exists an increasing sequence of finite-rank projections P1 ≤ P2 ≤ ... converging
strongly to I such that kPnT − T Pnk → 0 for every T ∈ Ω.
A representation π of a C∗-algebra A on H is called quasidiagonal (QD) if π(A)
is a QD set of operators on H.
Note that π being a QD representation of A is not equivalent to π(A) being a
QD C∗-algebra (see [10, Remark 7.5.3]). The issue arises from a Fredholm index
obstruction, which can be avoided by taking essential representations, i.e., those in
which the image contains no nonzero compact operators. In fact, Voiculescu has
shown in [52] that a separable C∗-algebra A is QD if and only if every faithful
essential representation of A (on a separable Hilbert space) is quasidiagonal.
Definition 5.7. An operator T on H is called quasitriangular if there exists an
increasing sequence of finite-rank projections P1 ≤ P2 ≤ ... converging strongly to
I such that kT Pn − PnT Pnk → 0; it is called bi-quasitriangular if T and T ∗ are
quasitriangular.
A QD operator is automatically bi-quasitriangular, and so if π is a QD rep-
resentation of a C∗-algebra A, then every π(a) ∈ π(A) is bi-quasitriangular. In
particular, if π is an essential representation of a separable RFD C∗-algebra, then
every π(a) ∈ π(A) is bi-quasitriangular.
Our first step is an observation which utilizes Apostol, Foias, and Voiculescu's
characterization of the norm closure of nilpotents in B(H) [2, Theorem 2.7] (which
is also a key ingredient in the proof of Herrero's result [30, Corollary 4.8]).
Proposition 5.8. If π is a faithful representation of a projective C∗-algebra A on
a separable Hilbert space H, then for any a ∈ A, π(a) is the norm limit of nilpotent
operators in B(H) if and only if its spectrum is connected and contains 0.
In particular, if T ∈ B(H) is a universal contraction, then there exists a sequence
(Tn)n≥1 of nilpotent operators in B(H) that converge to T in norm. Moreover, these
can be chosen to be contractions satisfying T n
n = 0 for each n ≥ 1.
Proof. It will suffice to prove the unital case. Let π be a faithful representation
of A, which is also essential by Proposition 4.5. By [2, Theorem 2.7], we know
that an operator T ∈ B(H) is the norm limit of nilpotent operators if and only
if T is bi-quasitriangular, σ(T ) and σe(T ) are connected, and 0 ∈ σe(T ). Because
A is projective, it is RFD. Since π is essential, every element π(a) ∈ π(A) is
automatically bi-quasitriangular, and σ(π(a)) = σe(π(a)). Hence we have that
π(a) is the norm limit of nilpotent operators in B(H) if and only if σ(π(a)) is
connected and contains zero.
In the case of A, note that the spectrum of the generator must be D, and any
sequence of nilpotent operators converging in norm to T can be easily adjusted to
satisfy the extra criteria.
(cid:3)
16
KRISTIN COURTNEY AND DAVID SHERMAN
Of course, Proposition 5.3 says we can actually arrange it so that each Tn is also
block-diagonal. However, this will not be of any help when we generalize Theorem
5.1 in Section 7.1, and Proposition 5.8 is really all we need.
The following theorem gives an "exactness"-type result, which says that any
faithful representation of A on a separable Hilbert space "asymptotically factorizes"
through the family of universal C∗-algebras generated by nilpotent contractions of
increasing finite orders.
Theorem 5.9. Let π : A → B(H) be a faithful representation of A on a separable
Hilbert space H, and let
φn : A → An := C∗
uhxn : kxnk ≤ 1, xn
n = 0i
be the unital *-homomorphisms induced by mapping x 7→ xn for each n ≥ 1. Then
there exists a sequence of *-homomorphisms
such that ψn ◦ φn converges to π pointwise in norm.
ψn : An → B(H)
Proof. By Proposition 5.8 there exists a sequence (Tn) ⊆ B(H) that converges in
norm to π(x) and satisfies kTnk ≤ 1 and T n
n = 0 for each n ≥ 1. For each n ≥ 1,
let
ψn : An → C∗(Tn, I) ⊆ B(H)
be the canonical surjective unital *-homomorphsim induced by mapping xn 7→ Tn.
Then, for any a ∈ A,
kψn ◦ φn(a) − ak → 0. (cid:3)
Remark 5.10. By Remark 5.4, the identity map on A does not factor through the
family {An}n≥1, which means we cannot hope to extend this to a "nuclearity"
result.
With Theorem 5.9, we can give our final proof of Theorem 5.1.
Proof. The family {φn} from Theorem 5.9 separates the elements of A. By [49,
Theorem 5], for each n ≥ 1 the C∗-algebra An is projective and hence also RFD
by Proposition 4.4. Let {ρn,k}k∈N be a separating family of finite-dimensional
representations of An. Then the compositions {ρn,k ◦ φn}n,k∈N form a separating
family of finite-dimensional representations of A, and, moreover, ρn,k ◦ φn(x) is
nilpotent for each n, k ∈ N where x denotes the generator of A.
(cid:3)
We conclude this section with a slight generalization of Theorem 5.1.
Theorem 5.11. For each λ ∈ D, A has a separating family of finite-dimensional
representations that send the generator to contractive matrices with spectrum {λ}.
The proof due to Herrero can be adapted to show this by replacing [30, Corollary
4.8] with the remark in [30] just after it. In the remainder of this section we sketch
how our second proof can also be adapted to prove Theorem 5.11.
As it turns out, by [38, Theorem 10], the universal C∗-algebra
Aλ,n = C∗
uhxn : kxnk ≤ 1, (xn − λ)n = 0i
is RFD for each n ≥ 1. So all we need is another "exactness" result, like Theorem
5.9, which says that every faithful representation (π, H) of A asymptotically fac-
torizes through the Aλ,n. To define the maps ψn as in Theorem 5.9, it was crucial
THE UNIVERSAL C ∗-ALGEBRA OF A CONTRACTION
17
that a universal contraction T is the norm limit of nilpotent contractions Tn, each
with appropriate nilpotency order. For this more general setting, we need T to be
the norm limit of contractions Tn satisfying (Tn − λI)n = 0 for each n ≥ 1, which
we show in Lemma 5.12 below. The rest of the argument for Theorem 5.11 will
then run like the third proof of Theorem 5.1.
Lemma 5.12. Let H be a separable Hilbert space, T ∈ B(H) be a universal con-
traction operator, and λ ∈ D. Then there exists a sequence of contractive operators
(Tn) ⊆ B(H) that converges in norm to T and satisfies (Tn − λI)n = 0 for each
n ≥ 1.
Proof. Since σ(T − λI) = ¯D − λ is connected and contains 0, by Proposition 5.8,
there is a sequence (Nn) ⊆ B(H) of nilpotents so that Nn → T − λI in norm and
N n
n = 0 for each n ≥ 1. Then Nn + λI → T in norm; however, each Nn + λI may
not be a contraction. To remedy this, we define Tn := cnNn + λI for each n ≥ 1
where
cn =( 1
: kNn + λIk ≤ 1
: kNn + λIk > 1
.
1−λ
kNn+λIk−λ
Then (Tn − λI)n = (cnNn)n = 0 for each n ≥ 1. Since kNn + λIk → kT k = 1, it
follows that cn → 1 and hence that Tn → T in norm. Moreover, if kNn + λIk ≤ 1,
then Tn = cnNn + λI = Nn + λI is a contraction. If kNn + λIk > 1, then 1 > cn,
and we have that
kTnk = kcnNn + λIk = kcn(Nn + λI) + (1 − cn)λIk ≤ cnkNn + λIk + λ1 − cn
= cnkNn + λIk + λ − λcn = cn(kNn + λIk − λ) + λ = (1 − λ) + λ = 1.
Thus (Tn) is the desired sequence.
(cid:3)
Remark 5.13. We cannot replace λ ∈ D from Theorem 5.11 with λ ∈ C\D in either
of the preceding results. Of course, if λ > 1, then no contraction Y ∈ B(H) can
satisfy (Y − λI)n = 0. And if Y ∈ B(H) is any contraction with (Y − λI)n = 0
for some n ≥ 1 and some λ ∈ T, then Y = λI. Indeed, by [38, Lemma 4] we know
that Y is unitarily equivalent to an upper triangular array (yij)i,j≥1 with yii = λ
for i ≥ 1. But, since kY k ≤ 1, it follows that yij = 0 for i 6= j, and Y = λI. In
other words, for any λ ∈ T and n ≥ 1,
C∗
uhx : kxk ≤ 1, (x − λ)n = 0i ≃ C
via the map x 7→ λ.
This argument shows that kN +λIk > 1 for any λ ∈ T and any nonzero nilpotent
operator N . (Just take y = N + λI, and run the previous argument as a contradic-
tion.) But this fails if we assume only that N is quasinilpotent, i.e., σ(N ) = {0}.
For example, let V be the Volterra integration operator, and N = (I + V )−1 − I.
We know from [29, Problem 190] that kN + Ik = k(I + V )−1k = 1, and σ(N ) = {0}.
However, N 6= 0.
6. A and Connes' Embedding Problem
As we mentioned in the introduction, A often behaves much like C∗(F2), in
particular when it comes to the role of C∗(F2) in Kirchberg's characterizations
of Connes' Embedding Problem.
In this section, we elucidate the relation of A
to Connes' Embedding Problem and consider some related questions. Before we
begin, we must first establish some relevant background.
18
KRISTIN COURTNEY AND DAVID SHERMAN
6.1. The WEP, LLP, and some results of Kirchberg. For two C∗-algebras
A and B, A ⊗max B and A ⊗min B refer to the completion of the algebraic tensor
product A ⊙ B of A and B with respect to the maximal and minimal C∗-norms,
respectively. The embedding A ⊙ B → A ⊗min B extends uniquely to a surjective
*-homomorphism A ⊗max B → A ⊗min B. See [10, Chapter 3] for definitions and
other relevant properties, and see [5, Section 2] for a short discussion on universal
C∗-algebras and tensor products.
Though it is readily verifiable that, for any C∗-algebras A, B and C with A ⊆ B,
A ⊗min C ⊆ B ⊗min C,
this containment can easily fail for the maximal tensor product because subalgebras
tend to have more representations. In [34], Lance characterized embeddings of C∗-
algebras that are always preserved under the maximal tensor product in terms of
completely positive extensions (see [10, Proposition 3.6.6] for a proof). We say a
linear map is positive if it sends positive elements to positive elements and completely
positive if this remains true even after amplification with matrix algebras.
If a
completely positive map is contractive, we call it cpc, and if it is moreover unital,
we call it ucp. We give Lance's result as a proposition/definition hybrid.
Proposition 6.1. Let A and B be unital C∗-algebras and 1B ∈ A ⊂ B an inclu-
sion. The inclusion is called relatively weakly injective if the following equivalent
conditions hold.
(1) There exists a ucp map φ : B → πu(A)′′ so that φ(a) = πu(a) for all
a ∈ A, where (πu, Hu) is the universal representation of A from the Gelfand-
Naimark Theorem.
(2) For any C∗-algebra C, there is a natural inclusion
A ⊗max C ⊆ B ⊗max C.
The map φ : B → πu(A)′′ is called a weak expectation. If a C∗-algebra A embeds
relatively weakly injectively into B(H) for any Hilbert space H, then it is said to
have Lance's Weak Expectation Property (WEP).
Proposition 6.2. If A is a projective C∗-algebra and B surjects onto A, then A
embeds relatively weakly injectively into B.
Indeed, the identity map idA : A → A lifts to a *-homomorphism ι : A → B,
which must be injective since idA is. The weak expectation B → πu(A) is just the
induced surjective *-homomorphism B → πu(A).
Let A and B be C∗-algebras and J a closed two-sided ideal in B with quotient
map π : B → B/J. A cpc map φ : A → B/J is liftable if there is a cpc map
φ : A → B such that π ◦ φ = φ. We say a cpc map φ : A → B/J is locally liftable
if for any finite dimensional operator system S ⊆ A there is a cpc map φ : S → B
such that π ◦ φ = φS.
A unital C∗ algebra A has the (local) lifting property or (L)LP if any ucp5 map
from A into a quotient C∗-algebra is locally liftable. A non-unital C∗-algebra has
the (L)LP if and only if its unitization does.
5Yes, we did unceremoniously shift from cpc maps to ucp maps, but it turns out to be sufficient
to restrict ourselves to ucp maps. Using the fact that the set of liftable cpc maps from a separable
operator system into a C ∗-quotient is closed ([4, Theorem 6]), one can argue (as in [10, Lemma
13.1.2]) that any ucp map with a (local) cpc lift has a (local) ucp lift.
THE UNIVERSAL C ∗-ALGEBRA OF A CONTRACTION
19
We can view the LP as projectivity in the category of unital C∗-algebras with
(unital) cpc maps. Then the LLP can be thought of as a "localized ucp projectivity"
and the WEP as "weak injectivity."
Proposition 6.3. [31, Corollaries 2.6 and 3.3] The LLP and WEP pass to relatively
weakly injective subalgebras.
A key class of examples comes from Kirchberg [32, Lemma 3.3].
Lemma 6.4. For any free group F, C∗(F) has the LLP. Moreover, C∗(F) has the
LP if F has countably many generators.
With this, we can readily prove the following proposition, which greatly simplifies
checking whether or not a given C∗-algebra has the LLP. This is a mild rephrasing of
[10, Corollary 13.1.4], and we include a brief argument for the reader's convenience.
Proposition 6.5. A unital C∗-algebra A has the LLP if and only if the identity
on A is locally liftable.
To see why it suffices to lift the identity of A, choose a free group F such that we
have a surjection π : C∗(F) → A. Let B be a C∗-algebra with two-sided closed ideal
J and φ : A → B/J a ucp map. Suppose S ⊆ A is a finite dimensional operator
system, and let ρ : S → C∗(F) be the lift of idAS. Then φ ◦ π : C∗(F) → B/J is a
ucp map, which lifts to a ucp map ψ : C∗(F) → B. Our desired lift of φS is ψ ◦ ρ.
B
ψ
φ
φ◦π
B/J
C∗(F)
ρ
S
⊆
π
A
Given the above proposition, the following is surely known to the experts. (The
special case for A follows immediately from [42, Proposition 2.2].)
Proposition 6.6. Any projective C∗-algebra has the LLP. In particular, A has the
LLP.
Next we give Kirchberg's deep and elegant characterization of Connes' Embed-
ding Problem. Below is an augmented and abridged version of [31, Proposition 8.1]
(omitting some of the equivalent conditions and adding one that quickly follows).
Theorem 6.7 (Kirchberg). The following conjectures are equivalent.
(1) Every finite von Neumann algebra with separable predual is embeddable into
an ultrapower of the hyperfinite II1-factor.
(2) C∗(F) ⊗max C∗(F) ≃ C∗(F) ⊗min C∗(F) canonically (where F is any non-
abelian free group).
(3) C∗(F) ⊗max C∗(F) is RFD.
(4) The LLP implies the WEP.
(5) C∗(F) has the WEP.
20
KRISTIN COURTNEY AND DAVID SHERMAN
Item (1) is traditionally referred to as Connes' Embedding Problem (CEP). Item
(3) is not included in [31]; however, it is equivalent to (2) by virtue of the following
proposition, which has often been alluded to in the literature. We briefly sketch an
argument here; for a complete proof, see [15, Proposition A.0.1].
Proposition 6.8. Given two RFD C∗-algebras A and B, A ⊗max B is RFD if and
only if A ⊗max B = A ⊗min B canonically.
By taking tensor products of finite dimensional representations, one can readily
verify that the minimal tensor product of two RFD C∗-algebras is again RFD. For
the other implication, it will suffice to show that any finite dimensional represen-
tation of the maximal tensor product factors through the minimal tensor product.
To that end, let A and B be C∗-algebras and π : A ⊗max B → Mn be a finite di-
mensional representation with restrictions πA : A → Mn and πn : B → Mn, i.e., πA
and πB have commuting ranges and πA⊙B = πA × πB. Then we have the induced
*-homomorphism πA ⊗ πB : A ⊗min B → πA(A) ⊗min πB(B) ≃ πA(A) ⊗max πB(B).
On the other hand, since πA(A) and πB(B) commute, the natural embeddings of
πA(A) and πB(B) into Mn induce a representation of πA(A) ⊗max πB(B) such that
πA(a)⊗πB(b) 7→ πA(a)πB(b) = π(a⊗b) for each a ∈ A and b ∈ B. The composition
of this representation with πA ⊗ πB gives a representation of A ⊗min B that agrees
with π on A ⊙ B.
6.2. A versus C∗(F2). Let us first show that A is "smaller" than C∗(F2).
Lemma 6.9. We have that A is a quotient of C∗(F2), but not vice versa.
Proof. Let x be the canonical generator of A, with x1 + ix2 its decomposition into
real and imaginary parts. Then A = C∗(x, 1) = C∗(x1, x2, 1) = C∗(eix1 , eix2), since
the spectrum of eixj is contained in a proper arc of the circle, and a continuous
branch of −i log(·) takes eixj back to xj . Any C∗-algebra generated by two unitaries
is a quotient of C∗(F2).
To establish the other statement, it suffices to recall that C∗(F2) is not singly
generated (as a unital C∗-algebra). For C∗(F2) surjects onto C(T2), since the latter
is generated by two unitaries, and C(T2) is not singly generated because T2 is not
homeomorphic to a compact subset of the plane.
(cid:3)
Although C∗(F2) and A are not isomorphic, they play the same universal role
for CEP. This has been observed before, specifically by Pisier in [43, Proposition
16.13], where he shows that CEP has an affirmative answer if and only if A has
the WEP (compare with Theorem 6.7(5)). It turns out that this can be linked to
a strong structural relationship between the two algebras.
Theorem 6.10. A and C∗(F2) embed relatively weakly injectively into one another.
Proof. We will use Proposition 6.1(1); in this case it will turn out that the required
ucp maps can be taken to be *-homomorphisms.
From Lemma 6.9, there exists a surjection π : C∗(F2) → A. So, A embeds
relatively weakly injectively into C∗(F2) by Proposition 6.2.
For the other relatively weakly injective embedding, let πu : C∗(F2) → B(Hu)
be the universal representation of C∗(F2), and let U1, U2 ∈ B(Hu) be the images
of the generators under πu. Let A1, A2 ∈ B(Hu) be self-adjoint elements such that
Uj = eiAj and
THE UNIVERSAL C ∗-ALGEBRA OF A CONTRACTION
21
C∗(U1, U2) ⊆ C∗(A1, A2) = C∗( A1
α + i A2
α ) ⊆ C∗(U1, U2)′′,
where α = kA1 + iA2k. Let x = x1 + ix2 be the canonical generator of A as before.
Then eiαxj are unitaries in A and
C∗(eiαx1 , eiαx2) ⊆ C∗(x1, x2) = A.
By universality, there exist surjective unital *-homomorphisms,
φ : C∗(U1, U2) → C∗(eiαx1 , eiαx2) and ψ : A → C∗( A1
α + i A2
α )
with φ(Uj) = eiαxj and ψ(xj ) = Aj
α . Then
ψ ◦ φ(Uj) = ψ(eiαxj ) = eiα( 1
α Aj) = eiAj = Uj.
It follows that ψ ◦ φ = idC ∗(U1,U2), and this completes the proof.
(cid:3)
We now leverage relative weak injectivity to get the following characterizations
of CEP. Items (1) and (3) were deduced differently by Pisier [43, Proposition 16.13]
and Blecher-Duncan [6, Section 6.1], respectively.
Theorem 6.11. The following are equivalent to CEP.
(1) A has the WEP.
(2) Projectivity implies the WEP.
(3) A ⊗max A ≃ A ⊗min A canonically.
(4) A ⊗max A is RFD.
(5) A⊗max A has a separating family of finite-dimensional representations that
map the two generators x ⊗ 1 and 1 ⊗ x to nilpotent matrices.
Proof. (2) ⇒ (1) ⇐⇒ CEP ⇒ (2): Condition (2) above implies (1) above by the
fact that A is projective (Proposition 4.1). By Theorem 6.10 and Proposition 6.3,
A has the WEP if and only if C∗(F2) does, which is item (5) in Theorem 6.7. Item
(4) in Theorem 6.7 implies (2) above because projective algebras have the LLP
(Proposition 6.6).
(3) ⇐⇒ (4) ⇐⇒ CEP: The equivalence of (3) and (4) follows from Proposition
6.8. With Theorem 6.10 and a few applications of Lance's characterization of
relatively weakly injective embeddings (Proposition 6.1), we have the embeddings
and
A ⊗max A ⊂ C∗(F2) ⊗max C∗(F2)
C∗(F2) ⊗max C∗(F2) ⊂ A ⊗max A.
Since residual finite dimensionality passes to subalgebras, we have that (4) is equiv-
alent to C∗(F2) ⊗max C∗(F2) being RFD (item (3) of Theorem 6.7).
(4) ⇐⇒ (5): The reverse direction is obvious, so we assume (4) and show
(5). Since (3) and (4) have already been shown to be equivalent, it suffices to
prove the claim in (5) for A ⊗min A, and this essentially amounts to the argument
that the minimal tensor product of two RFD C∗-algebras is RFD. By Theorem
5.1, for i = 1, 2, there exist separating families of finite-dimensional representations
{σ(i)
n maps the generator to a nilpotent matrix. For
n, m ≥ 1, let πn,m = σ(1)
m denote the *-homomorphism on A ⊗min A that
maps a ⊗ b to σ(1)
m (b). Then {πn,m}m,n∈N form the desired family of
representations.
(cid:3)
n }n∈N of A where each σ(i)
n ⊗ σ(2)
n (a) ⊗ σ(2)
22
KRISTIN COURTNEY AND DAVID SHERMAN
Remark 6.12. To highlight the difficulty in proving part (3) of Theorem 6.11, con-
sider the following. Because residual finite dimensionality is preserved by exten-
sions, A ⊗max A is RFD if and only if A0 ⊗max A0 is. Moreover, by Proposition 6.8
A0 ⊗max A0 is RFD if and only if it is canonically isomorphic to A0 ⊗min A0. How-
ever, recall that A0 is contractible, and contractibility is preserved under taking
tensor products (just take the homotopy between the zero map and the identity on
the contractible C∗-algebra and tensor it with the identity on the other C∗-algebra).
In particular, this means that both A0 ⊗min A0 and A0 ⊗max A0 are contractible,
which makes them very difficult to distinguish with topological invariants. The
authors would like to thank Hannes Thiel for clarifying some points on this.
Remark 6.13. Using their universality, one can identify A⊗max A with the universal
unital C∗-algebra generated by a pair of doubly commuting contractions (i.e., the
generators commute and commute with each other's adjoints). In [13] it was asked
whether the universal C∗-algebra generated by a pair of commuting contractions,
which can be identified with the maximal C∗-algebra for the bidisk algebra [7,
Example 2.3], is RFD. An affirmative answer would amount to a noncommutative
*-polynomial analogue (in the spirit of (3.2)) to Ando's von Neumann Inequality
from [1]. It does not seem to be known whether this question is easier than the
question for A ⊗max A.
The universal (unital) C∗-algebra generated by a pair of unrelated contractions
is a full (unital) free product of RFD C∗-algebras, which is RFD by [19, Theorem
3.2].
Remark 6.14. Though the question of whether or not A ⊗max A is RFD proves to
be quite challenging, the proof of [10, Proposition 7.4.5] can be readily adapted to
show that A ⊗max A is quasidiagonal.
7. C∗-algebras generated by multiple universal contractions
We conclude with a few remarks on C∗-algebras generated by two or more uni-
versal contractions.
B(H)m an m-row contraction if kP SiS∗
7.1. Universal row contractions. We call an m-tuple of operators (S1, ..., Sm) ∈
i k ≤ 1. (Note that the Si are not assumed
to be commuting.) Such tuples have received considerable attention in multivari-
ate operator theory, especially with regard to generalizing the commutant lifting
theorem and von Neumann's inequality (e.g., [11], [20], [21], [44], and [45]).
In this section, we generalize Theorems 3.1 and 5.1 to universal C∗-algebras
generated by m-row contractions6:
Rm := C∗
m
u*y1, ..., ym :(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xi=1
≤ 1+ .
yiy∗
i(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Some of the techniques we have developed for A = R1 can be generalized to prove
the same results for Rm. In particular, where before we had von Neumann-type
inequalities for noncommutative *-polynomials in one variable, in this section we
obtain a noncommutative *-polynomial analogue to Popescu's von Neumann in-
equality for row contractions [45, Theorem 2.1].
6Note that the row contraction relation automatically enforces a norm condition on the gen-
erators. As we mentioned in Section 4, this will guarantee that these universal C ∗-algebras exist
(and are separably acting).
THE UNIVERSAL C ∗-ALGEBRA OF A CONTRACTION
23
Proposition 7.1. Fix 1 < m < ∞ and let {y1, ..., ym} denote the generators of
Rm. Then for any 1 ≤ k < m, any subset {yi1, ..., yik } ⊆ {y1, ..., ym} forms a
jik = 1, and, moreover,
universal k-row contraction.
each yi is a universal contraction.
j=1 yji y∗
In particular, kPk
Proof. Since
k
m
0 ≤
Xj=1
yjiy∗
ji ≤
yjy∗
j ,
Xj=1
it follows that the {yi1, ..., yik } form a k-row contraction. For any k-row contraction
(Z1, ..., Zk), the restriction to C∗(yi1 , ..., yik ) of the map Rm → C∗(Z1, ..., Zk, 1)
sending yij → Zj for 1 ≤ j ≤ k and the rest of the generators to 0 is a surjection. (cid:3)
By [36, Theorem 5.1], Rm is projective for 1 ≤ m < ∞, just like A. From this it
follows that Rm is RFD (see [13, Example 3] for a more direct proof). It turns out
we can say moreover that the noncommutative *-polynomials in the generators of
Rm form a dense subset of elements that actually attain their norms under some
finite-dimensional representation (just as with A).
Theorem 7.2. Fix 1 ≤ m < ∞, and let q be a noncommutative *-polynomial in
m variables of degree d. Then there exists a representation π of Rm on a (2m)d+1-
dimensional Hilbert space such that kπ(q(y1, ..., ym))k = kq(y1, ..., ym)k.
Proof. To prove this, we will need to change little from the proof of Theorem 3.1.
The key there was that compressions of π0(x) are still contractions, which means
they induce representations of A. Hence, the argument can be adapted for any
universal C∗-algebra C∗hGRi with a finite set of generators and with relations
that are preserved under compressions. All we have to prove is that Rm is such an
algebra. To that end, suppose Y1, ..., Ym ∈ B(H) form a row contraction, and let
P ∈ B(H) be a projection. Since P ≤ IH, we have that YiP Y ∗
for each
1 ≤ i ≤ m. Then
i ≤ YiY ∗
i
m
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xi=1
P YiP Y ∗
YiP Y ∗
YiY ∗
≤ 1. (cid:3)
i P(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
m
≤(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xi=1
i (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
m
≤(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xi=1
i (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Now we turn our attention back to nilpotents.
In Section 5, we gave three
arguments showing that the universal contraction can be faithfully represented
as a countable direct sum of nilpotent matrices. The third can be generalized
to show the same for Rm, for 1 ≤ m < ∞. The main idea of the proof is to
establish an "exactness"-type result like Theorem 5.9, which says that any faithful
representation of Rm "asymptotically factorizes" though the family of universal
C∗-algebras generated by order n nilpotents that form row contractions:
Rm,n = C∗
u(cid:28)y1,n, ..., ym,n :
kyi,nk ≤ 1, yn
i,n = 0, i = 1, ..., m,
i=1 yi,ny∗
i,nk ≤ 1
(cid:29)
Just as in the proof of Theorem 5.9, we will first need to know that any faithful
representation of a universal row contraction is the norm limit of row contractions
whose entries are each nilpotent. To that end, we have the following lemma.,
kPm
Lemma 7.3. Let 1 ≤ m < ∞ and π : Rm → B(H) a faithful representation
on a separable Hilbert space. Then, for each 1 ≤ j ≤ m, there exists a sequence
24
KRISTIN COURTNEY AND DAVID SHERMAN
(Tj,n)n≥1 in B(H) that converges in norm to π(yj ), and for each n ≥ 1, T n
and (T1,n, ..., Tm,n) forms a row contraction.
j,n = 0
Proof. By Propositions 7.1 and 5.8, each π(yj ) is the norm limit of a sequence
of nilpotent contractions (Nj,n)n≥1, which we can assume satisfy N n
j,n = 0 for all
j,nk−1/2} (where we
n ≥ 1. Now for each n ≥ 1, we let αn = min{1, kPm
assume αn = 1 if Nj,n = 0 for all 1 ≤ j ≤ m), and define
j=1 Nj,nN ∗
Tj,n := αnNj,n
for each 1 ≤ j ≤ m. We immediately have T n
j,n = 0 for all n ≥ 1
and 1 ≤ j ≤ m. Moreover, since Nj,n → π(yj ) for each 1 ≤ j ≤ m, it follows
from Proposition 7.1 that αn → 1, and hence Tj,n → π(yj) for each 1 ≤ j ≤ m.
Finally we check that (T1,n, ..., Tm,n) forms a row contraction for each n ≥ 1. If
j,nk > 1, then
j,n = αnN n
j=1 Nj,nN ∗
m
m
(N1,n, ..., Nm,n) is a row contraction, then αn = 1. If kPm
−1(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
j,n(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xj=1
=(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xj=1
=(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xj=1
j,n(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
j,n(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
α2
nNj,nN ∗
Nj,nN ∗
Tj,nT ∗
m
m
Xj=1
Nj,nN ∗
j,n(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
= 1.
(cid:3)
Theorem 7.4. For 1 ≤ m < ∞, Rm has a separating family of finite-dimensional
representations that map each generator to a nilpotent. In particular, there exist
nilpotent matrices Nk1, ..., Nkm ∈ Mjk , k ∈ N, such that (⊕kNk1 , ..., ⊕kNkm ) is a
universal m-row contraction.
Proof. Let π be a faithful representation of Rm on a separable Hilbert space H,
and let (Ti,n)n≥1 be the sequences of nilpotents guaranteed by Lemma 7.3 for each
1 ≤ i ≤ m.
By universality, for each n ≥ 1, we can define a surjective *-homomorphsim ψn
from
kPm
Rm,n = C∗
u(cid:28)y1,n, ..., ym,n :
kyi,nk ≤ 1, yn
i,n = 0, i = 1, ..., m,
i=1 yi,ny∗
i,nk ≤ 1
(cid:29)
to C∗(1, T1,n, ..., Tm,n) by sending yi,n 7→ Ti,n for each i = 1, ..., m. Let φn : Rm →
Rm,n denote the surjection induced by sending generators to generators. Then, for
each i = 1, ..., m,
kψn ◦ φn(yi) − π(yi)k → 0,
which means that ψn ◦ φn converge pointwise in norm to π.
In particular, the
{φn} separate the points of Rm. By [37, Theorem 2.3], each Rm,n is projective
and hence RFD by Proposition 4.4. So, composing the {φn} with the separating
families of finite-dimensional representations of the Rm,n's gives us the desired
family of representations of Rm.
The matrices in the statement of the theorem come from taking the direct sum
(cid:3)
of this family of representations.
Remark 7.5. Since the generators of Rm are all universal contractions, Example
5.2 also shows that Theorem 7.2 will not hold if we also want the generators to be
mapped to nilpotents.
We conclude this section by rephrasing the results in terms of a Popescu-von
Neumann inequality for noncommutative *-polynomials on row contractions.
THE UNIVERSAL C ∗-ALGEBRA OF A CONTRACTION
25
Corollary 7.6. Let q be a noncommutative *-polynomial in m variables of degree
d and (T1, ..., Tm) a row contraction. Then
kq(T1, ..., Tm)k ≤ sup{kq(A1, ..., Am)k : (A1, ..., Am) ∈ (Mm
n )≤1, n ≥ 1}
= max{kq(A1, ..., Am)k : (A1, ..., Am) ∈ (Mm
= sup{kq(N1, ..., Nm)k : (N1, ..., Nm) ∈ (Mm
(2m)d+1 )≤1}
n )≤1, N n
i = 0, 1 ≤ i ≤ m, n ≥ 1}.
7.2. Universal Pythagorean C∗-algebras. An m-tuple of operators (S1, ..., Sm)
in B(H)m satisfying the identity
m
S∗
i Si = 1
Xi=1
is often called a column isometry. Such operators have recently been utilized in
[9] to construct several interesting unitary representations of Thompson's groups
F and T (and their n-ary versions Fn and Tn). We adopt the language of [9] in
calling
Pm = C∗
u*a1, ..., am :
universal Pythagorean C∗-algebras.
a∗
i ai = 1+ ,
m
Xi=1
Below, we adapt Choi's technique from [12, Theorem 7] to show that each Pm is
RFD for m > 1. Note that this is not true for m = 1. The universal C∗-algebra of
a single isometry is not RFD because every finite-dimensional isometry is unitary.
Theorem 7.7. For each m > 1, Pm is RFD.
Proof. First, we faithfully and nondegenerately represent Pm on a separable Hilbert
space H, denoting the generators by A1, ..., Am. Let (Pn) be an increasing sequence
of projections, each of rank n, such that Pn converges strongly to IH. For each
n ≥ 1 and 1 ≤ i ≤ m, let B(n)
, and note that
Cn ≤ Pn for each n ≥ 1.
Define operators A(n)
i=1 B(n)∗
i B(n)
1 , ..., A(n)
i
m ∈ B(PnH ⊕ PnH) by
i = PnAiPn and Cn =Pm
2 =(cid:18)
(Pn − Cn)1/2
B(n)
0
2
A(n)
1 =(cid:18)B(n)
1
0
0
Pn(cid:19) , A(n)
for each 2 < i ≤ m. For each n, we compute
0(cid:19) , and A(n)
i =(cid:18)B(n)
i
0
0
0(cid:19)
0
0(cid:19)
i A(n)
A(n)∗
i
m
Xi=1
1
0
0
1 B(n)
=(cid:18)B(n)∗
=(cid:18)Pn
Pn(cid:19) +(cid:18)B(n)∗
Pn(cid:19) = IPnH⊕PnH.
0
0
2 B(n)
2 + Pn − Cn
0
0
0(cid:19) +(cid:18)Pm
i=3 B(n)∗
i B(n)
0
i
1 , ..., A(n)
So, (A(n)
M2n. For each n ≥ 1, let πn : C∗(IH, A1, ..., Am) → C∗(Pn ⊕ Pn, A(n)
B(PnH ⊕ PnH) be the representation induced by Ai 7→ A(n)
It remains to show that π = ⊕nπn is an isometry.
m ) satisfies the column isometry identity in B(PnH ⊕ PnH) ≃
m ) ⊂
for each 1 ≤ i ≤ m.
1 , ..., A(n)
i
We now consider PnH ⊕ PnH as a subspace of H ⊕ H in the obvious way, thereby
as an operator on H ⊕ H (acting as zero on (PnH ⊕ PnH)⊥).
viewing each A(n)
i
26
KRISTIN COURTNEY AND DAVID SHERMAN
Since the Pn converge *-strongly to IH, we have that A(n)
to A1 ⊕ 1H and A(n)
noncommutative *-polynomial q in two variables,
converges *-strongly
converges *-strongly to Ai ⊕ 0 for 2 ≤ i ≤ m. So, for any
1
i
q(A(n)
1 , ..., A(n)
m ) S ∗OT−−−−→ q(A1 ⊕ IH, A2 ⊕ 0, ..., Am ⊕ 0).
Note that (A1 ⊕ IH, A2 ⊕ 0, ..., Am ⊕ 0) is also a universal m-column isometry.
Indeed, it clearly satisfies the column isometry identity, and, since
kq(A1 ⊕ IH, A2 ⊕ 0, ..., Am ⊕ 0)k = kq(A1, ..., Am)⊕ q(IH, 0, ..., 0)k ≥ kq(A1, ..., Am)k
for any noncommutative *-polynomial q in two variables, the natural surjection
induced by mapping A1 7→ A(n)
i ⊕ 0 for 2 ≤ i ≤ m is isometric.
Now, let q be a noncommutative *-polynomial in m variables and ǫ > 0. Assume
1 ⊕ 1H and Ai 7→ A(n)
kq(A1, ..., Am)k = 1. Then kq(A1 ⊕ IH, A2 ⊕ 0, ..., Am ⊕ 0)k = 1 and
for all sufficiently large n. Hence,
kq(A(n)
1 , ..., A(n)
m )k ≥ 1 − ǫ,
kπ(q(A1, ..., Am))k ≥ kπn(q(A1, ..., Am))k = kq(A(n)
1 , ..., A(n)
m )k ≥ 1 − ǫ
for all sufficiently large n. Since ǫ was arbitrary, π is isometric on the set of
noncommutative *-polynomials on A1, ..., Am, which, by continuity, means it is
isometric on C∗(IH, A1, ..., Am) ≃ Pm.
(cid:3)
References
[1] T. Ando, On a pair of commutative contractions, Acta Sci. Math. (Szeged) 24 (1963), 88 -- 90.
[2] C. Apostol, C. Foias, and D. Voiculescu, On the norm-closure of nilpotents, II, Rev. Roum.
Math. Pures et Appl. 19 (1974), 549 -- 577.
[3] W. Arveson, Subalgebras of C ∗-algebras, II, Acta Math. 128 (1972), 271 -- 308.
[4] W. Arveson, Notes on extensions of C ∗-algebras, Duke Math. J. 44 (1977), 329 -- 355.
[5] B. Blackadar, Shape theory for C ∗-algebras, Math. Scand. 56 (1985), 249 -- 275.
[6] D. P. Blecher and B. Duncan, Nuclearity-related properties for nonselfadjoint operator alge-
bras, J. Operator Theory 65 (2011), 47 -- 70.
[7] D. P. Blecher and V. I. Paulsen, Explicit construction of universal operator algebras and
applications to polynomial factorization, Proc. Amer. Math. Soc. 112 (1991), 839 -- 850.
[8] B. Brenken and Z. Niu, The C ∗-algebra of a partial isometry, Proc. Amer. Math. Soc. 140
(2012), 199 -- 206.
[9] A. Brothier and V. Jones, Pythagorean representations of Thompson's groups, preprint, 2018
arXiv: 1807.06215.
[10] N. Brown and N. Ozawa, C ∗-Algebras and Finite-Dimensional Approximations, Graduate
Studies in Mathematics 88, American Mathematical Society, Providence, 2008.
[11] J. Bunce, Models for n-tuples of non-commuting operators, J. Funct. Anal. 57 (1984), 21 -- 30.
[12] M. Choi, The full C ∗-algebra of the free group on two generators, Pacific J. Math. 87 (1980),
41 -- 48.
[13] R. Clouatre and C. Ramsey, Residually finite-dimensional operator algebras, preprint, 2018
arXiv:1806.00038.
[14] A. Connes, Classification of injective factors. Cases II1, II∞, IIIλ, λ 6= 1, Ann. Math. (2)
104 (1976), 73 -- 115.
[15] K. Courtney, C ∗-algebras and their finite-dimensional representations, PhD thesis (2018),
University of Virginia, Charlottesville, United States.
[16] K. Courtney and T. Shulman, Elements of C ∗-algebras attaining their norm in a finite-
dimensional representation, Canad. J. Math. to appear. arXiv:1707.01949.
[17] R. E. Curto and D. A. Herrero, On closures of joint similarity orbits, Integral Equations
Operator Theory 8 (1985), 489 -- 556.
[18] J. Dixmier, C ∗-algebras, North Holland Publ. Co., Amsterdam, 1977.
THE UNIVERSAL C ∗-ALGEBRA OF A CONTRACTION
27
[19] R. Exel and T. Loring, Finite-dimensional representations of free product C ∗-algebras, Int.
J. Math, 3 (1992), 469 -- 476.
[20] A. Frazho, Models for non-commuting operators, J. Funct. Anal. 48 (1982), 1 -- 11.
[21] A. Frazho, Complements to models for noncommuting operators, J. Funct. Anal. 59 (1984),
445 -- 461.
[22] T. Fritz, T. Netzer, and A. Thom, Can you compute the operator norm? Proc. Amer. Math.
Soc. 142 (2014), 4265 -- 4276.
[23] U. Haagerup and P. de la Harpe, The numerical radius of a nilpotent operator on a Hilbert
space, Proc. Amer. Math. Soc. 115 (1992), 371 -- 379.
[24] D. Hadwin, An operator-valued spectrum, Indiana University Math. J. 26 (1977), 329 -- 340.
[25] D. Hadwin, An asymptotic double commutant theorem for C ∗-algebras, Trans. Amer. Math.
Soc. 244 (1978), 273 -- 297.
[26] D. Hadwin, Nonseparable approximate equivalence, Trans. Amer. Math. Soc. 266 (1981),
203 -- 231.
[27] D. Hadwin, A lifting characterization of RFD C ∗-algebras, Math. Scand. 115 (2014), 85 -- 95.
[28] D. Hadwin and T. Shulman, Variations of projectivity for C ∗-algebras, preprint,
arXiv:1709.01379.
[29] P. R. Halmos, A Hilbert Space Problem Book, Second Edition, Springer-Verlag, New York,
1982.
[30] D. A. Herrero, Quasidiagonality, similarity and approximation by nilpotent operators, Indiana
Univ. Math. J. 30 (1981), 199 -- 232.
[31] E. Kirchberg, On non-semisplit extensions, tensor products and exactness of group C ∗-
algebras, Invent. Math. 112 (1993), 449 -- 489.
[32] E. Kirchberg, Commutants of unitaries in UHF algebras and functorial properties of exact-
ness, J. Reine Angew. Math. 452 (1994), 39 -- 78.
[33] C. S. Kubrusly, An Introduction to Models and Decompositions in Operator Theory,
Birkhauser Boston, Inc., Boston, MA, 1997.
[34] E. C. Lance, On nuclear C ∗-algebras, J. Funct. Anal. 12 (1973), 157 -- 176.
[35] T. Loring, Lifting Solutions to Perturbing Problems in C ∗-Algebras, Fields Institutue Mono-
graphs 8, American Mathematical Society, Providence, 1997.
[36] T. Loring and T. Shulman, Non-commutative semialgebraic sets and associated lifting prob-
lems, Trans. Amer. Math. Soc. 364 (2012), 721 -- 744.
[37] T. Loring and T. Shulman, Non-commutative semialgebraic sets in nilpotent variables, New
York J. Math. 14 (2012), 361 -- 372.
[38] T. Loring and T. Shulman, Lifting algebraic contractions in C ∗-algebras, Oper. Theory Adv.
Appl. 233 (2014), 85-92.
[39] B. B. Morrel and P. S. Muhly, Centered operators, Studia Math. 51 (1974), 251 -- 263.
[40] J. von Neumann, Eine Spektraltheorie fur allgemeine Operatoren eines unitaren Raumes,
Math. Nachr. 4 (1951), 258 -- 281.
[41] T. Oikhberg, A note on universal operators, in: Ordered Structures and Applications, pp.
339-347, Trends Math., Birkhauser/Springer, Cham, 2016.
[42] N. Ozawa, On the lifting property for universal C ∗-algebras of operator spaces, J. Operator
Theory, 46 (2001), 579-591.
[43] G. Pisier, Introduction to Operator Space Theory, London Mathematical Society Lecture
Note Series 294, Cambridge University Press, 2003.
[44] G. Popescu, Isometric dilations for infinite sequences of noncommuting operators, Trans.
Amer. Math. Soc. 316 (1989), 523 -- 536.
[45] G. Popescu, Von Neumann inequality for (B(H)n )1, Math. Scand. 68 (1991), 292 -- 304.
[46] H. Radjavi and P. Rosenthal, Invariant Subspaces, Springer-Verlag, New York, 1973.
[47] G.-C. Rota, Note on the invariant subspaces of linear operators, Rend. Circ. Mat. Palermo
(2) 8 (1959), 182 -- 184.
[48] G.-C. Rota, On models for linear operators, Comm. Pure Appl. Math. 13 (1960), 469 -- 472.
[49] T. Shulman, Lifting of nilpotent contractions, Bull. London Math. Soc. 40 (2008), 1002 -- 1006.
[50] D. Topping, UHF algebras are singly generated, Math. Scand. 22 (1968), 224 -- 226.
[51] D. Voiculescu, A non-commutative Weyl-von Neumann theorem, Rev. Roum. Math. Pures et
Appl. 21 (1976), 97 -- 113.
[52] D. Voiculescu, A note on quasi-diagonal C ∗-algebras and homotopy, Duke Math. J. 62 (1991),
267 -- 271.
28
KRISTIN COURTNEY AND DAVID SHERMAN
Department of Mathematics, University of Virginia, 141 Cabell Drive, Kerchof
Hall, P.O. Box 400137, Charlottesville, VA 22904-4137 USA
Current address: Mathematical Institute, WWU Munster, Einsteinstr. 62, 48149 Munster,
Germany
E-mail address: [email protected]
Department of Mathematics, University of Virginia, 141 Cabell Drive, Kerchof
Hall, P.O. Box 400137, Charlottesville, VA 22904-4137 USA
E-mail address: [email protected]
|
1708.03258 | 1 | 1708 | 2017-08-10T15:33:54 | E-theory Spectra for graded C*-algebras | [
"math.OA",
"math.AT"
] | This paper brings together C*-algebras and algebraic topology in terms of viewing a C*-algebraic invariant in terms of a topological spectrum. E-theory, E(A,B), is a bivariant functor in the sense that is a cohomology functor in the first variable and a homology functor in the second variable but underlying goes from the category of separable C*-algebras and *-homomorphisms to the category of abelian groups and group homomorphisms. Here we create a generalisation of a orthogonal spectrum to quasi-topological spaces for E-theory. This includes a rich product structure in the context of graded separable C*-algebras. | math.OA | math |
E-theory Spectra for graded C ∗-algebras
Sarah L. Browne
May 14, 2018
Abstract
This paper brings together C ∗-algebras and algebraic topology in terms of viewing
a C ∗-algebraic invariant in terms of a topological spectrum. E-theory, E(A, B), is a
bivariant functor in the sense that is a cohomology functor in the first variable and
a homology functor in the second variable but underlying goes from the category of
separable C ∗-algebras and ∗-homomorphisms to the category of abelian groups and
group homomorphisms. Here we create a generalisation of a orthogonal spectrum to
quasi-topological spaces for E-theory . This includes a rich product structure in the
context of graded separable C ∗-algebras.
Introduction
Separable C ∗-algebras are analytical objects used in non-commutative geometry. E-theory,
E(A, B), is a invariant of C ∗-algebras and is a bivariant functor from the category of separable
C ∗-algebras and ∗-homomorphisms to the category of abelian groups and group homomor-
phisms. Here we will only consider real graded C ∗-algebras since then we have a useful tool
for application in differential geometry. E-theory is a cohomology theory in its first variable
and a homology theory in its second variable. It was first defined by Higson [8] in a categor-
ical manner and Connes and Higson [4] gave a concrete description using homotopy classes
of certain morphisms of C ∗-algebras. This will form section 1.
Spectrum is a topological notion used to represent stable homotopy theories. Here we re-
quire the notion of a quasi-spectrum, which is considered as a spectrum over quasi-topological
spaces. We incorporate an action of the orthogonal group and create an orthogonal quasi-
spectrum. This has a rich product structure as we will see.
The main construction encodes the Bott periodicity of E-theory proving in the complex
graded case by Guentner-Higson [6] and in the real case by Browne [2], as an Ω-quasi-
spectrum coming from weak equivalences. Additionally we also encode the product structure
of E-theory. This then will give us a stable homotopy theory. It is worth noting that the
spectrum will be dependent on C ∗-algebras just like the K-theory spectrum for C ∗-categories
is dependent on C ∗-categories defined by Paul D. Mitchener [11]. The author likes to think
of this as a "local" spectrum and it is suited for possible applications to positive scalar
curvature by generalising work of Weiss-Williams [14]
1
Further we include a way of connecting K-theory spectrum and K-homology spectrum
using the relations of E-theory with both K-theory and K-homology in the final section of
the paper.
Acknowledgements
The author would like to thank their PhD supervisor, Paul Mitchener for all the valuable
meetings and discussions during her PhD. The author would also like to thank the EPSRC
for funding her PhD which made this possible. Lastly the author wishes to thank many
mathematicians for conversations and particularly Jamie Gabe for his insight during a visit
at the University of Southampton.
1 E-theory
This section details the analytic necessities which we need to inlude in our definition of
the quasi-spectrum for E-theory. A complex C ∗-algebra is a complex Banach space with
involution ∗ satisfying for all a ∈ A, the C ∗-identity a∗a = a2. A real C ∗-algebra is
a real Banach space with an involution, satisfying th C ∗-identity and additonally that the
element 1 + a∗a is invertible in A for all a ∈ A. A C ∗-algebra is called separable if its
underlying topological space has a countable dense subset and all of our C ∗-algebras will
be separable. A map f : A → B of real C ∗-algebras is called a ∗-homomorphism and is an
algebra homomorphism and satisfies f (a∗) = (f (a))∗ for all a ∈ A. A grading on a C ∗-
algebra A is an automorphism δA : A → A such that δ2
A = 1. We can also think of a grading
by allocating an even and odd notion on A. That is we have A = Aeven ⊕ Aodd, where
Aeven = {a ∈ A δA(a) = a} and Aodd = {a ∈ A δA(a) = −a}.
Then additionally we define the degree of an element a ∈ A by:
deg(a) =(0,
1,
if a ∈ Aeven
if a ∈ Aodd.
An important example of a graded real C ∗-algebra is the algebra of continuous real-valued
functions vanishing at infinity, S = C0(R), under the supremum norm and grading given
by δ(f )(x) = f (−x) for all x ∈ R. Denote this algebra by S. Also we can consider the
suspension of a C ∗-algebra
ΣA = {f : [0, 1] → A f (0) = f (1) = 0}.
If A is graded then ΣA has grading coming from A. Also we can consider the n-fold suspen-
sion, ΣnA = Σn−1ΣA.
Now we should note that a graded ∗-homomorphism is a ∗-homomorphism that preserves
the grading. That is, if an element is even then its image is even and if an element is odd
then its image is odd. Furthermore we can define a different type of morphism between
C ∗-algebras:-
2
Definition 1.1. Let A and B be real graded C ∗-algebras with gradings δA and δB respec-
tively. A graded asymptotic morphism ϕ : A 99K B is a family of functions {ϕt}t∈[1,∞) such
that:
1. the map t 7→ ϕt(a), from [1, ∞) to B is continuous and bounded for each a ∈ A,
2. limt→∞ ϕt(ab) − ϕt(a)ϕt(b) = 0, for each a, b ∈ A,
3. limt→∞ ϕt(a + λb) − ϕt(a) − λϕt(b) = 0, for each a, b ∈ A, λ ∈ R,
4. limt→∞ ϕt(a∗) − ϕt(a)∗ = 0, for each a ∈ A,
5. limt→∞ ϕt(δA(a)) − δB(ϕt(a)) = 0 for each a ∈ A.
Denote the set of these by Asyg(A, B). It will be useful to consider the graded tensor
product:
Definition 1.2. Let A and B be graded C ∗-algebras with gradings δA and δB respectively.
Then define Ab⊗B to be the completion of the algebraic tensor product of A and B in the
ϕ(ai)ψ(bi)
norm
ai ⊗ bi = sup
ϕ,ψ
where ϕ : A 99K C, ψ : A 99K C are graded asymptotic morphisms to a common graded
C ∗-algebra C. We equip Ab⊗B with involution, multiplication and grading defined by:
1. (ab⊗b)∗ = (−1)deg(a)deg(b)a∗ ⊗ b∗
2. (ab⊗b)(cb⊗d) = (−1)deg(b)deg(c)(ac ⊗ bd)
3. γ(ab⊗b) = α(a) ⊗ β(b)
Extending by linearity gives Ab⊗B.
We have the following result, Lemma 4.5 in [7], for asymptotic morphisms:
Lemma 1.3. Let ϕ : A1 99K A2 and ψ : B1 99K B2 be (graded) asymptotic morphisms, then
the compositions
Xi
Xi
and
A1b⊗B1
A1b⊗B1
ϕ b⊗1
−−→ A2b⊗B1
1 b⊗ψ−−→ A1b⊗B2
1 b⊗ψ
−−→ A2b⊗B2,
ϕ b⊗1−−→ A2b⊗B2,
are equal. The 1 symbolises the relevant identity morphism.
Now E-theory is built out of homotopy classes so we need the notion of a homotopy of
graded asymptotic morphism:-
Definition 1.4. A homotopy of graded asymptotic morphisms ϕt, ψt : A 99K B is a graded
asymptotic homomorphism θt : A 99K C([0, 1], B) such that
θt(a)(0) = ϕt(a) and θt(a)(1) = ψt(a).
3
Denote the set of homootopy classes of asymptotic morphisms from A to B by JA, BK.
We require the next remark to impose a particular grading on compact operators.
Remark 1.5. Let H be a Hilbert space equipped with the orthogonal decomposition
H = H0 ⊕ H1,
where H0 denotes the even elements and H1 denotes the odd elements. Then the C ∗-algebra
K(H) of compact operators on such a Hilbert space is graded. For this grading, we consider
2 × 2 matrices of operators where the diagonal matrices are even and the off diagonal ones
are odd. That is we have a grading
defined by
β : K(H) → K(H)
β(T ) =(T
−T
if T is even
if T is odd.
Then the graded E-theory groups are defined by:
En
g (A, B) = JSb⊗Ab⊗K(H), ΣnBb⊗K(H)K,
where ΣnB for n ≥ 0 is the n-fold suspension of a C ∗-algebra and it's grading comes from
B, and where K(H) denotes the compact operators on some real graded Hilbert space with
grading as detailed in Remark 1.5. If we have a graded ∗-homomorphisms ϕ, ψ : A → B, we
can also form homotopy classes of these, by setting ϕ = ϕt, ψ = ψt for all t ∈ [1, ∞). Denote
the set of homotopy classes of ∗-homomorphisms from A to B by [A, B].
It will become important to consider the graded ∗-homomorphism
defined in [6] and also detailed in [2]. The idea is to restrict to the set of continuous functions
on the interval [−R, R], denoted by SR and use functional calculus and define
∆ : S → Sb⊗S,
f (XRb⊗1 + 1b⊗XR) ∈ SRb⊗SR,
and thereafter obtain a graded ∗-homomorphism.
Now in E-theory we have a Bott periodicity result that can be found in [2]. Let F = R
or C. Before we state the result we note that Fn,0 is the Clifford algebra on n generators
e1, e2, . . . en such that e2
i = 1 and eiej = −ejei for all i, j. Then we have the following:
Theorem 1.6. There is a ∗-homomorphim b : S → Σb⊗F1,0 inducing an isomorphism
Also for Section 4, we have the following definitions coming from the E-theory definition.
E(A, B) ∼= E(A, ΣBb⊗F1,0).
The K-theory groups are given by:
and the K-homology is given by
Kn(A) = [S, ΣnAb⊗K(H)] ∼= En(R, ΣnA),
Khom(A) = En(A, R).
4
2 Quasi-topological spaces
In this section we generalise the notion of orthogonal spectra to quasi-topological spaces.
In order to do this, we have to define the notion of a quasi-continuous group action and
further prove we have a symmetric monoidal structure on the category of quasi-orthogonal
sequences which we also define.
Firstly we recall some notions of quasi-topological spaces by Spanier [12] which we will
need as it is not known if we can put a standard topology on the set of asymptotic morphisms.
Definition 2.1. A quasi-topology on a set X, is a collection of sets of maps from C to X for
each compact Hausdorff space C, written Q(C, X), called quasi-continuous and satisfying:
• any constant map C → X belongs to Q(C, X),
• if f : C1 → C2 is a map of compact Hausdorff spaces and g ∈ Q(C2, X) then gf ∈
Q(C1, X),
• for a disjoint union C = C1 ∐ C2 of closed compact Hausdorff spaces, a map g : C → X
is contained in Q(C, X) if and only if gCi ∈ Q(Ci, X) for i = 1, 2,
• for every f : C1 → C2 surjective map of compact Hausdorff spaces, then a map h : C2 →
X is quasi-continuous if h ◦ f is quasi-continuous.
A quasi-topological space is a set X endowed with a quasi-topology as described above.
If X is a topological space we can obtain a quasi-topology on X by considering Q(C, X)
as the set of continuous maps from C to X in the topology of X.
A map of quasi-topological spaces f : X → Y is called quasi-continuous if g ∈ Q(C, X)
implies that the composite f g ∈ Q(C, Y ). Also by the definition of quasi-continuous maps,
a composite of quasi-continuous maps is also quasi-continuous. A quasi-homeomorphism
f : X → Y between quasi-topological spaces is a quasi-continuous bijection with a quasi-
continuous inverse g : Y → X.
Definition 2.2. Let X, Y be quasi-topological spaces and f, g : X → Y be quasi-continuous
maps. Then a homotopy is a quasi-continuous map
such that for all x ∈ X, H(x)(0) = f (x) and H(x)(1) = g(x).
H : X → C([0, 1], Y ),
The suspension and loop space of a quasi-topological space X are defined similarly to
the case of topological spaces by
ΣtopX = S1 ∧ X,
and
ΩX = {µ : S1 → X µ is quasi-continuous and basepoint preserving}
and we consider the circle S1 with the quasi-topology that comes from the standard topology
on R2. That is, our quasi-continuous maps are the continuous maps from every compact
Hausdorff space in to S1 in the topology from R2. Now we check that Σtop and Ω are
5
adjoints in the category where objects are quasi-topological spaces and arrows are quasi-
continuous maps.
In order to do this, we consider an abstract result and obtain it as a
corollary.
Proposition 2.3. Let X, Y and Z be quasi-topological spaces. Then
F (X ∧ Y, Z) and F (X, F (Y, Z)),
are quasi-homeomorphic.
Proof. We define α : F (X ∧ Y, Z) → F (X, F (Y, Z)) by
((α(f ))(x))(y) = f (x ∧ y),
where f ∈ F (X ∧ Y, Z), x ∈ X and y ∈ Y . Then α is quasi-continuous since f is quasi-
continuous.
Define β : F (X, F (Y, Z)) → F (X ∧ Y, Z) by
(β(g))(x ∧ y) = (g(x))(y),
where g ∈ F (X, F (Y, Z)), x ∈ X and y ∈ Y . So β is quasi-continous since g is quasi-
continuous.
Finally α and β are inverses, so we obtain a quasi-homeomorphism.
Corollary 2.4. Σtop and Ω are adjoints in the category of quasi-topological spaces. That is
F (ΣtopX, Y ) and F (X, ΩY )
are quasi-homeomorphic.
Proof. This follows from Proposition 2.3 since
and
F (ΣtopX, Y ) = F (X ∧ S1, Y ),
F (X, ΩY ) = F (X, F (S1, Y )).
Let A, B be C ∗-algebras. Further by work of Dardalat-Meyer [5] we can define a quasi
topology on Asy(A, B), the set of asymptotic morphisms from A to B. We define the
set of quasi-continuous maps from a compact Hausdorff space Y to Asy(A, B) to be the
Asy(A, C(Y, B)), mentioned in [5]. That is, more precisely we have
Definition 2.5. For a compact Hausdorff space Y , a map h : Y → Asy(A, B) is quasi-
continuous when for each t ∈ [1, ∞) the map eht(a) : Y → B defined by
is continuous.
eht(a)(y) = h(y)t(a),
6
Now we check that this is a quasi-topology.
Proposition 2.6. The set of asymptotic morphisms from A to B, Asy(A, B), is a quasi-
topological space when equipped with the above quasi-topology.
Proof. We must check the axioms. Let c : Y → Asy(A, B) be constant. Then for y ∈ Y ,
c(y) = ft : A 99K B for a fixed f . Then we need to show that c is quasi-continuous. That is
a ∈ A, y ∈ Y by
to show for each map c, the map ect(a) : Y → B is continuous. Then we define this map for
ect(a)(y) = c(y)t(a) = ft(a),
which is continuous as a function of Y and hence c is continuous.
Now let f : Y1 → Y2 be a map of compact Hausdorff spaces and let g : Y2 → Asy(A, B)
be quasi-continuous. Then we want to show that gf : Y1 → Asy(A, B) is quasi-continuous.
That is, we need to show thatfgf t(a) : Y1 → B is continuous. Now f is continuous and since
g is quasi-continuous, we have eachegt(a) : Y2 → B is continuous and
Now,
egt(a)(y) = g(y)t(a).
fgf t(a)(y) =egt(a)f (y),
Let Y = Y1 ∐ Y2 of compact Hausdorff spaces. Then we need to show that g : Y →
Asy(A, B) is quasi-continuous if and only if gYi is quasi-continuous for i = 1, 2. Suppose
which is continuous in Y so is fgf t(a) is continuous, yielding that gf is quasi-continuous.
g : Y → Asy(A, B) is quasi-continuous. Then map egt(a) : Y → B defined by
continuous function is continuous. Suppose that gYi is quasi-continuous, then (egYi)t(a) : Yi →
B is continuous. Then by properties of continuous functions we know that if the restrictions
are continuous then the map of the disjoint uniion will be continuous.
is continuous. Now by properties of continuous functions we know that the restriction of a
egt(a)(y) = g(y)t(a),
Finally, we need to check for every surjective map f : Y1 → Y2 of compact Hausdorff spaces
that g : Y2 → Asy(A, B) is quasi-continuous if gf : Y1 → Asy(A, B) is quasi-continuous. Then
for a particular map f and by the above argument
is continuous. Let f : Y1 → Y2 be the identity, then
fgf t(a)(y) =egt(a)f (y),
fgf t(a)(y) =egt(a)f (y) =egt(a),
on Asy(A, B) defined as above.
and henceegt(a) is continuous and the result follows. So we do indeed have a quasi-topology
For a quasi-topological space X, let X+ denote the space with a basepoint.
Proposition 2.7. The quasi-topological spaces ΩAsy(A, B) and Asy(A, ΣB) are quasi-homeomorphic.
7
Proof. By the definition of a quasi-topology, we know that
ΩAsy(A, B) : = Q(S1, Asy(A, B))+.
Then by the definition 2.5, we have that
Q(Y, Asy(A, B))+ = Asy(A, C(Y, B)+),
and then that
ΩAsy(A, B) = Q(S1, Asy(A, B))+
= Asy(A, C(S1, B)+)
= Asy(A, ΣB).
The above results hold in the case of graded asymptotic morphisms.
2.1 Group actions
The following definition makes sense since a topological group can be viewed as a quasi-
topological group.
Definition 2.8. Let G be a topological group acting on a quasi-topological space X. Then
the group action is called quasi-continuous if the map G × X → X is quasi-continuous. If
this is the case we say that the set X is a quasi G-space.
Definition 2.9. A map f : X → Y of quasi G-spaces is a quasi G-map if it is G-equivariant.
That is, for all g ∈ G, we have
Now we consider basepoint preserving group actions.
f (gx) = g(f (x)).
Proposition 2.10. Let G and H be groups. Let X be a quasi G-space, Y a quasi H-space
where the group actions preserve the basepoints of both X and Y . Then there are basepoint
preserving actions of G × H on X × Y , X ∨ Y and X ∧ Y . These actions are defined in the
obvious way.
We now need the notion of a balanced smash product.
Definition 2.11. Let X be a right quasi G-space and Y a left quasi G-space, then we can
from the balanced smash product X ∧G Y , which is the quotient space X ∧ Y / ∼G where
(xg ∧ y) ∼G (x ∧ g−1y) ⇔ (x ∧ y) ∼G (xg ∧ g−1y),
for all g ∈ G.
8
Let the equivalence class of x ∧ y be denoted by x ∧G y. Now using these we can construct
a left quasi G-space.
Let G be a topological group and H a subgroup. Then let X be a based left quasi H-
space where G acts by preserving the basepoint. Let G+ = G ∐ {∗}, then we can construct
the right quasi G-space denoted G+ ∧H X using the above equivalence classes. Additionally
we can actually define a left quasi G-action on this space by the following map:
(f, g ∧H x) 7→ f g ∧H x,
for all f ∈ G.
To prove this is well-defined action it is a formality of using the fact that H is a subgroup
of G.
Let X and Y be based quasi G-spaces. Then let QG(X, Y ) denote the set of basepoint
preserving quasi G-maps. Then we have the following result:
Proposition 2.12. Let H be a subgroup of a group G. Let X be a left quasi H-space and
Y a left quasi G-space. There there is a natural bijection
QH(X, Y ) ←→ QG(G+ ∧H X, Y ).
Proof. We define α : QH(X, Y ) → QG(G+ ∧H X, Y ). Let f ∈ QH (X, Y ) and g ∧H x ∈
G+ ∧H X, then we define
α(f )(g ∧H x) = gf (x),
and we then need to check that α is well-defined and also G-equivariant. Since (g ∧H x) ∼
(gh ∧H h−1x), then
α(f )(gh ∧H h−1x) = ghf (h−1x) = ghh−1f (x)
since f is a quasi H-map
= gf (x) = α(f )(g ∧H x).
Then α(f ) is G-equivariant since
α(f )(g′g ∧H x) = g′gf (x) = g′(gf (x)) = g′α(f )(g ∧H x),
for all g′ ∈ G.
Now define β : QG(G+ ∧H X, Y ) → QH(X, Y ). Let k ∈ QG(G+ ∧H X, Y ), x ∈ X, then
where e denotes the identity in G. Then β is H-equivariant since
β(k)(x) = k(e ∧H x),
β(k)(hx) = k(e ∧H hx)
= k(h−1h ∧H hx)
= k(h ∧H x) by equivalence relations
= hk(e ∧H x) as k is a H-map
= hβ(k)x.
Then is is clear that α and β are inverse maps, and both are natural in X and Y , so the
result follows.
9
2.2 Quasi-Orthogonal sequences
Let O be the category of finite dimensional real Euclidean inner product spaces and linear
isometric isomorphisms where we have objects to be the set
and morphisms are
obj(O) = {Rn n = 0, 1, . . .}
O(A, B) =(O(n),
∅,
if A = B = Rn
otherwise.
It should be noted that this is a small category since the collection of objects is a set.
Let T denote the category of quasi-topological spaces with basepoints and quasi-continuous
maps. So obj(T ) is the collection of quasi-topological spaces with basepoints and the mor-
phisms T (X, Y ) are the set of basepoint preserving quasi-continuous maps from X to Y .
Then we can obtain the product category T × T where obj(T × T ) are pairs (X, Y )
of quasi-topological spaces with basepoints, and morphisms are
(T × T )((X, Y )(Z, W )) = {(f, g) f ∈ T (X, Z), g ∈ T (Y, W )}.
Proposition 2.13. The smash product ∧ : T × T → T of quasi-topological spaces is a
functor.
The following definition of a quasi-orthogonal sequence is going to form part of the
definition of a orthogonal quasi-spectrum.
Definition 2.14. Let O and T be the categories defined above. Then we define the category
of quasi orthogonal sequences formed as the functor category T O with objects
obj(T O) = {functors X : O → T Xn := X(Rn)},
together with a left quasi-continuous basepoint preserving action of O(n) on each Xn for all
n ≥ 0, and morphisms
T O (X, Y ) = {ϕ : X → Y ϕ is a natural transformation},
and such that a natural transformation is formed of sets of quasi-continuous basepoint pre-
serving maps ϕn : Xn → Yn that are O(n)-equivariant for n ≥ 0, or equivalently that the
map ϕn commutes with the group action of O(n) on Xn and Yn.
A useful example of such a functor category will be the unit sequence coming from the
orthogonal sequence defined below. Consider a based topological space K, then define the
orthogonal sequence with n-space:
(GpK)n =(O(n)+ ∧ K,
{∗},
if n = p
otherwise
10
Then the unit sequence is when we just have the topological space S0, given by the sequence
G0S0 = {S0, ∗, ∗, . . .}.
We also consider quasi-biorthogonal sequences since they will help us in defining our
smash product structure.
The category of quasi-biorthogonal sequences is defined to be the category with objects
obj(T O×O ) = {X : O × O → T X is a functor},
together with a quasi-continuous basepoint preserving left-action of O(m) × O(n), and
T O×O(X, Y ) = {ψ : X → Y ψ is a natural transformation},
formed of sets of quasi-continuous basepoint preserving maps ψm,n : Xm,n → Ym,n, where
Xm,n := X(Rm, Rn), that are O(m) × O(n)-equivariant for all m, n ≥ 0.
Using this we can define the external smash product of two quasi-orthogonal sequence X
and Y .
Definition 2.15. Define the external smash product X∧Y to be the quasi-biorthogonal
sequence given by the composition
O × O X×Y−−−→ T × T
∧−→ T ,
defined by
(X∧Y )m,n = (X∧Y )(Rm, Rn) = X(Rm) ∧ Y (Rn) = Xm ∧ Yn.
Then by Proposition 2.10, the quasi-topological space Xn ∧ Ym has a quasi-O(n) × O(m)-
action.
For a general quasi-orthogonal sequence X we can define a quasi-biorthogonal sequence
X ◦ ⊕ by:
(X ◦ ⊕)m,n = (X ◦ ⊕)(Rm, Rn) = X(Rm+n) = Xm+n.
Now we can construct the tensor product of quasi-orthogonal sequences since the category
T is complete and cocomplete.
Definition 2.16. For quasi orthogonal sequence X and Y we define the tensor product of
X and Y to be the quasi-orthogonal sequence
(X ⊗ Y )n = _p+q=n
O(n)+ ∧O(p)×O(q) Xp ∧ Yq,
where we define the O(n)-action on (X ⊗ Y )n by acting on each wedge summand.
Then we can combine the external smash product and tensor product of quasi-orthogonal
sequences as a natural bijection:
Proposition 2.17. For quasi-orthogonal sequences X, Y and Z, there is a natural bijection
T O×O (X∧Y, Z ◦ ⊕) ←→ T O(X ⊗ Y, Z).
11
Proof. Let f : X∧Y → Z◦⊕ be a natural transformation in the category of quasi-biorthogonal
sequences. Then fp,q : Xp ∧ Yq → Z ◦ ⊕ is quasi O(p) × O(q)-equivariant and then by propo-
sition 2.12, this corresponds to a quasi O(n)-equivariant map, with n = p + q
f p,q : O(n)+ ∧O(p)×O(q) Xp ∧ Yq → Zn.
Now fixing n and letting p and q vary, this allows us to obtain a quasi O(n)-equivariant map
f n = _p+q=n
O(n)+ ∧O(p)×O(q) Xp ∧ Yq → Zn,
which is a quasi-continuous basepoint preserving O(n)-equivariant map in T O from X ⊗ Y
to Z.
Now we construct a map the other way. Let g ∈ T O (X ⊗ Y, Z). Then g is a wedge
summand of basepoint preserving quasi-continuous O(n)-equivariant maps
gn : _p+q=n
O(n)+ ∧O(p)×O(q) Xp ∧ Yq → Zn,
for all n ≥ 0. Also, we can write that gn =Wp+q=n gp,q, where
gp,q : O(n)+ ∧O(p)×O(q) Xp ∧ Yq → Zn,
and by proposition 2.12, we obtain a basepoint preserving quasi-continuous O(p) × O(q)-
equivariant map as required.
Let G0S0 be the unit quasi-orthogonal sequence
(G0S0)n = (S0, ∗, ∗, . . .).
For details of the following see Chapter 4 of [3].
Proposition 2.18. The category of quasi-orthogonal sequences forms a symmetric monoidal
category (T O , ⊗, G0S0).
Let S = (S0, S1, S2 . . .) be the quasi-orthogonal sequence defined in terms of quasi-
topological spaces. For a proof of the subsequent result see Proposition 4.3.13 in [3].
Proposition 2.19. The orthogonal sequence of quasi-topological spaces S = (S0, S1, S2 . . .)
is a commutative monoid in the symmetric monoidal category (T O, ⊗, G0S0).
3 E-theory orthogonal quasi-spectra
This section brings together ideas from the previous section, since we will define the notion
of an orthogonal quasi-spectrum which is a quasi-orthogonal sequence with added structure.
We will show that we have an orthogonal quasi-spectrum representing the graded E-theory
groups and thereafter show we have a smash product.
12
3.1 Quasi-Spectra
We begin by defining concepts we have seen before in terms of quasi-topological spaces.
A quasi-spectrum is a sequence of based quasi-topological spaces X0, X1, . . . with structure
maps ǫ : Xm → ΩXm+1 that are quasi-continuous. An Ω-quasi-spectrum is a quasi-spectrum
where for all natural numbers m the structure maps ǫ : Xm → ΩXm+1 are weak equivalences.
Then we can define an orthogonal quasi-spectrum:
Definition 3.1. An orthogonal quasi-spectrum is
• a sequence of based quasi-topological spaces X0, X1, . . .
• a basepoint preserving quasi-continuous left action of O(m) on each Xm for all m, and
• a collection of based structure maps σ = σm : Xm ∧ S1 → Xm+1 that are quasi-
continuous,
such that for each m, n ≥ 0, the iterated map
m : Xm ∧ Sn → Xm+1 ∧ Sn−1 → . . . → Xm+n,
σn
is quasi-continuous and O(m) × O(n)-equivariant.
In the same manner, we have that a morphism of orthogonal quasi-spectrum f : X → Y
is a collection of quasi-O(m)-equivariant maps fm : Xm → Ym for all m, which satisfy the
following commutative diagram:
Xm ∧ S1
f ∧idS1
/ Ym ∧ S1
σm
σm
Xm+1
fm+1
/ Ym+1,
or alternatively that the following diagram commutes:
Xm
ǫm
fm
/ Ym
ǫm
ΩXm+1
Ωfm+1
/ ΩYm+1.
It is easily seen that any orthogonal spectrum is an orthogonal quasi-spectrum. By Corol-
lary 2.4, the structure maps in the definition of quasi-spectrum can be defined in terms of
loop spaces. Notice that an orthogonal quasi-spectrum is a quasi-orthogonal sequence with
more structure.
Proposition 3.2. The category of right S-modules, mod-S is naturally equivalent to the
category of orthogonal quasi-spectrum.
13
/
/
/
/
Proof. Consider the multiplication map ν : X ⊗ S → S for a right S-module X. Then by
Proposition 2.17 we have a set of O(m) × O(n)-equivariant maps
νn
m : Xm ∧ Sn → Xm+n,
for m, n ≥ 0 with unit quasi-homeomorphism ν0
that the structure maps are then defined by νm.
m. Now this action is associative so it follows
Conversely, consider the set of structure maps
σn
p : Xp ∧ Sp → Xn+p,
for a spectrum X and p, n ≥ 0, with unit quasi-homeomorphism σ0
p. Then we have a
multiplicative map ν : X ⊗ S → X defining a right S-module. Since these constructions are
inverses, we have a natural equivalence of these two categories.
Hence we can obtain a tensor product of orthogonal quasi-spectrum since we have a
tensor product in the category of right S-modules.
Definition 3.3. Let X be an orthogonal quasi-spectrum with spaces Xn. For each integer
k ∈ Z we define the k-th stable homotopy group πk(X) to be the direct limit
πk(X) = lim−→
n
πk+nXn,
under the maps ǫ∗ : πk+nXn → πk+n+1Xn+1 induced from the structure maps ǫ : Ωk+nXn →
Ωk+n+1Xn+1.
3.2 Graded E-theory Spectra
Let Asyg(A, B) denote the set of graded asymptotic morphisms from A to B with the quasi-
topology as defined in Definition 2.5.
Proposition 3.4. The map of quasi-topological spaces
defined by
is quasi-continuous for all xt ∈ Asyg(A, B).
f : Asyg(A, B) → Asyg(Db⊗A, Db⊗B)
f (xt) = idDb⊗xt,
Proof. Since gradings follow immediately, we consider ungraded asymptotic morphisms through-
out the proof. To prove a map of quasi-topological spaces is quasi-continuous, wwe need to
check that for a quasi-continuous map g : Y → Asy(A, B) where Y is a compact Haus-
g : Y → Asy(A, B) where Y is a compact Hausdorff space is quasi-continuous. Then by defi-
dorff space, that the composition f g : Y → Asy(Db⊗A, Db⊗B) is quasi-continuous. Suppose
nition 2.5 we know that g is quasi-continuous when for each t ∈ [1, ∞) the mapegt(a) : Y → B
defined by
egt(a)(y) = g(y)t(a)
14
is continuous. Then we define f g : Y → Asy(Db⊗A, Db⊗B) by for each t ∈ [1, ∞)
f (g(y)t)(a) = idDb⊗g(y)t(a) = idDb⊗egt(a)(y)
but since g is quasi-continuous and that
it follows from the definition of quasi-topology on the set of asymptotic morphisms that f g
is quasi-continuous.
f (g(y)t)(a) =ff gt(a)(y),
Definition 3.5. Let K = K(H). Define X(A, B) to be the sequence of based quasi-
topological spaces
where m ≥ 0. Define maps ǫm : Xm → ΩXm+1:
Xm = Asyg(Sb⊗Ab⊗K, Bb⊗Fm,0b⊗K)
Asyg(Sb⊗Ab⊗K, Bb⊗Fm,0b⊗K)
by:
/ ΩAsyg(Sb⊗Ab⊗K, Bb⊗Fm+1,0b⊗K)
Asyg(Sb⊗Ab⊗K, Σ(Bb⊗Fm+1,0)b⊗K)
∼=
ǫ(xt) = (bb⊗idB b⊗Fm,0
b⊗K) ◦ (idSb⊗xt) ◦ (∆b⊗idA b⊗K),
rem 1.6. Alternatively, we also have a map σm : Xm ∧ S1 → Xm+1 defined by
for all xt ∈ Asyg(Sb⊗Ab⊗K, Bb⊗Fm,0b⊗K) and the Bott map b ∈ Homg(S, ΣF1,0) from Theo-
with xt ∈ Asyg(Sb⊗Ab⊗K(H), Bb⊗Fm,0b⊗K(H)) and s ∈ S1.
Definition 3.6. We define a quasi-continuous action of the group O(m) on the space Xm as
follows. First we consider the alternative definition of the Clifford algebra Fm,0 as Cliff(V ).
Recall that for V an m-dimensional Euclidean vector space, Cliff(V ) = G(V )/ ∼ where
G(V ) is the algebra generated by V subject to the equivalence relation ∼ defined by
σm(xt, s) = ǫm(xt)(s),
v2 = v2 · 1
for all v ∈ V . We write ab for the product of two elements a, b ∈ Cliff(V ).
If V = Rm, then we have a natural group action (H, v) 7→ Hv where H ∈ O(m), v ∈ V .
Then we can define a group action of O(m) on G(V ) by
H(v1 . . . vk) 7→ H(v1) . . . H(vk) and H(1) = 1
for all H ∈ O(m). Then this gives a group action of O(m) on Cliff(V ) since
H(v2) = H(v)H(v) = (H(v))2
= H(v)2 · 1 = v · 1 since H is orthogonal.
15
/
So then we get a group action
λ : O(m) × Fm,0 → Fm,0,
by
λ(H, (e1, e2, . . . em)) = H(e1)H(e2) . . . H(em),
where H ∈ O(m), e1, e2, . . . em are the generators of the algebra Fm,0. Then we define
by
λ∗ : O(m) × Bb⊗Fm,0b⊗K(H) → Bb⊗Fm,0b⊗K(H)
λ∗(H, bb⊗xb⊗p) = bb⊗λ(H, x)b⊗p
with H ∈ O(m), b ∈ B, x ∈ Fm,0 and p ∈ K(H). Then we finally define a group action of
O(m) on Xm
by
λ∗∗ : O(m) × Asyg(Sb⊗Ab⊗K(H), Bb⊗Fm,0b⊗K(H))
−→ Asyg(Sb⊗Ab⊗K(H), Bb⊗Fm,0b⊗K(H)),
λ∗∗(H, αt)(x) = λ∗(H, αt(x)),
where we have αt ∈ Asyg(Sb⊗Ab⊗K(H), Bb⊗Fm,0b⊗K(H)), x ∈ Sb⊗Ab⊗K(H) and H ∈ O(m).
Then it follows that this action is O(m)-equivariant.
The following is true by Proposition 3.4.
Proposition 3.7. The action in the previous definition is a basepoint preserving quasi-
continuous action of O(m) on Xm.
Proposition 3.8. The map ǫm : Xm → ΩXm+1 is quasi-continuous and hence the map
σm : Xm ∧ S1 → Xm+1 is quasi-continuous.
Proof. Since both
b⊗K(H)
bb⊗idB b⊗Fm,0
and ∆b⊗idA b⊗K(H)
are ∗-homomorphisms, these two maps are continuous. So it suffices to check the map
maps yields a quasi-continuous map. By proposition 3.4, with D = S it follows that the
(idSb⊗xt) is quasi-continuous since the a composition of continuous and quasi-continuous
map (idSb⊗xt) is quasi-continuous. Hence the map ǫm is quasi-continuous. Since σm is defined
in terms of ǫm it is also quasi-continuous.
We define the iterated map σn
m : Xm ∧ Sn → Xm+n by
σn
m(xt, s1, s2, . . . sn) = ǫn(xt)(s1)(s2) . . . (sn),
. . . ∧ S1.
where xt ∈ Asyg(Sb⊗Ab⊗K(H), Bb⊗Fm,0b⊗K(H)) and s1, s2, . . . sn is contained in S1 ∧ S1 ∧
16
Proposition 3.9. The iterated map σn
and O(m) × O(n)-equivariant.
m : Xm ∧ Sn → Xm+n defined above quasi-continuous
Proof. By other results it suffices to check that the map is O(m) × O(n)-equivariant.
Firstly it is clear that Xm ∧ Sn and Xm+n are quasi O(m) × O(n)-spaces. Let i : O(m) ×
O(n) → O(m + n) be the inclusion map.
defined earlier.
J(xt) = (idBb⊗Jb⊗idK(H)) ◦ xt
Now O(m + n) acts on Asyg(Sb⊗Ab⊗K(H), Bb⊗Fm+n,0b⊗K(H)) by
for all J ∈ O(m + n) and xt : Sb⊗Ab⊗K(H) → Bb⊗Fm+n,0b⊗K(H). Here J acts of Fm+n,0 as
Then O(m) × O(n) acts on Asyg(Sb⊗Ab⊗K(H), Bb⊗Fm,0b⊗K(H)) ∧ Sn by,
for all H ∈ O(m), K ∈ O(n), xt ∈ Asyg(Sb⊗Ab⊗K(H), Bb⊗Fm,0b⊗K(H)) and s ∈ Sn. Then
(H, K)(xt, s) = ((idBb⊗Hb⊗idK(H)) ◦ xt, Ks),
we need to show that for σ = σn
m : Xm ∧ Sn → Xm+n, that
σ((H, K)(xt, s)) = i(H, K)σ(xt, s),
that is,
That is to show, by definition of σ that,
σ((Hxt, Ks) = i(H, K)σ(xt, s).
ǫ(Hxt)(Ks) = i(H, K)ǫ(xt)(s).
Then
Let bn = bb⊗ . . .b⊗b be the n-fold graded tensor product of the Bott map, b : S → ΣF1,0.
we have bb⊗ . . .b⊗b(λ)(s) ∈ Fn,0 for λ ∈ S n and s ∈ Sn. Then for K ∈ O(n),
bn = bb⊗ . . .b⊗b : S n → Σnb⊗Fn,0,
(bb⊗ . . .b⊗b)(λ)(Ks) = K(bb⊗ . . .b⊗b)(λ)(s).
Then by permuting copies of Σ and extending by linearity we have an action of the
orthogonal group.
Reconsidering
the left hand side yields
ǫ(Hxt)(Ks) = i(H, K)ǫ(xt)(s),
and the right hand side yields
ǫ(Hxt)(Ks) = ((bnb⊗idB b⊗Fm,0
i(H, K)ǫ(xt)(s) = i(H, K)(bb⊗idB b⊗Fm,0
b⊗K(H)) ◦ (idSb⊗Hxt) ◦ (∆b⊗idA b⊗K(H))(Ks)),
b⊗K(H)) ◦ (idSb⊗xt) ◦ (∆b⊗idA b⊗K(H)).
Then
17
ǫ(Hxt)(Ks)
= ((bnb⊗idB b⊗Fm,0
= i(H, 1)((bnb⊗idB b⊗Fm,0
= i(H, 1)i(1, K)(bnb⊗idB b⊗Fm,0
= i(H, K)(bnb⊗idB b⊗Fm,0
b⊗K(H)) ◦ (idSb⊗Hxt) ◦ (∆b⊗idA b⊗K(H)))(Ks)
b⊗K(H)) ◦ (idSb⊗xt) ◦ (∆b⊗idA b⊗K(H))(Ks))
b⊗K(H)) ◦ (idSb⊗xt) ◦ (∆b⊗idA b⊗K(H))(s)
b⊗K(H)) ◦ (idSb⊗xt) ◦ (∆b⊗idA b⊗K(H))(s)
= i(H, K)ǫ(xt)(s)
Then the result follows.
The proof of the following result follows from the above propositions, namely Proposi-
tion 3.7, Proposition 3.8 and Proposition 3.9.
Proposition 3.10. The spectrum X(A, B) is an orthogonal quasi-spectrum.
Proposition 3.11. If G0, G1, G2, . . . is a sequence of groups with isomorphisms θn : Gn →
Gn+1 for n ≥ 0, then
Proof. We first need to construct a commutative diagram.
Gn = G0.
lim−→
n
Gn
θn
.
G0
δ
"❊❊❊❊❊❊❊❊
<②②②②②②②②
ψ
Gn+1
As θn is an isomorphism for all n ≥ 0, we have inverses, so δ = (θ0)−1 . . . (θn−1)−1 and
ψ = (θ0)−1 . . . (θn−1)−1(θn)−1 and hence the diagram commutes. Now we check that G0 is
unique. Suppose we have a group H such that we have a group homomorphism f : G0 → H
which fits into the following diagram
Gn
θn
µ1
.
G0
f
/ H
µ2
δ
#❋❋❋❋❋❋❋❋❋
;①①①①①①①①
ψ
Gn+1
Then define f = µ1δ−1 so our diagram commutes. Suppose that we have another group
homomorphism g : G0 → H fitting into the diagram. Then by commutativity we have
gδ = µ1, so g = µ1δ−1 = f so f is unique.
18
"
<
#
$
$
/
;
:
:
Proposition 3.12. The direct limit lim−→n
Proof. This result follows from Proposition 3.11 where
Eg(A, Σk+nBb⊗Fn,0) is Eg(A, ΣkB).
and using Proposition 1.6.
Gn = Eg(A, Σk+nBb⊗Fn,0)
Proposition 3.13. For all positive integers k,
πkX(A, B) = Eg(A, ΣkB).
Proof. Since X is an orthogonal quasi-spectrum we have that
πkX(A, B) = lim−→
n
πk+nEn.
Then
lim−→
n
πk+nXn = lim−→
n
n
= lim−→
π0Ωk+nAsyg(Sb⊗Ab⊗K(H), Bb⊗Fn,0b⊗K(H))
π0Asyg(Sb⊗Ab⊗K(H), Σk+nBb⊗Fn,0b⊗K(H))
JSb⊗Ab⊗K(H), Σk+nBb⊗Fn,0b⊗K(H)K
Eg(A, Σk+nBb⊗Fn,0)
= lim−→
= Eg(A, ΣkB) by Proposition 3.12.
= lim−→
n
n
Proposition 3.14. The orthogonal quasi-spectrum X(A, B) is an Ω-quasi-spectrum.
Proof. We just need to check that the structure map ǫ : En → ΩEn+1 is a weak equivalence.
That is the map πkEn → πkΩEn+1 is an isomorphism for all k. Now this gives us the map:
which is an isomorphism for all k by Theorem 1.6.
Eg(A, Σk(Bb⊗Fn,0)) → Eg(A, Σk+1(Bb⊗Fn+1,0)),
Theorem 3.15. Let A, B and C be graded C ∗-algebras. Then there is a natural map of
orthogonal quasi-spectra
µm,n : X(A, B) ∧ X(B, C) → X(A, C),
defined by
(α ∧ β)t 7→ (βr(t)b⊗idFm,0) ◦ (idSb⊗αt) ◦ (∆b⊗idA b⊗K),
where α ∈ Asyg(Sb⊗Ab⊗K, Bb⊗Fm,0b⊗K) and β ∈ Asyg(Sb⊗Bb⊗K, Cb⊗Fn,0b⊗K). In addition
the product is associative up to homotopy.
19
Proof. The product gives a natural O(m) × O(n)-equivariant map:
Asyg(Sb⊗Ab⊗K, Bb⊗Fm,0b⊗K) ∧ Asyg(Sb⊗Bb⊗K, Cb⊗Fn,0b⊗K)
−→ Asyg(Sb⊗Ab⊗K, Cb⊗Fm+n,0b⊗K),
given by permuting the m, n and m + n copies of F1,0.
Now compatibility with the structure maps follows from the naturality of the structure
maps and also since we have the following two diagrams:
and
Xm(A, B) ∧ Xn(B, C)
µm,n
/ Xm+n(A, C)
ǫ∧id
ǫ
ΩXm+1(A, B) ∧ Xn(B, C)
µm+1,n
/ ΩXm+n+1(A, C),
Xm(A, B) ∧ Xn(B, C)
µm,n
/ Xm+n(A, C)
id∧ǫ
ǫ
Xm(A, B) ∧ ΩXn+1(B, C)
µm,n+1
/ ΩXm+n+1(A, C),
where id denote the obvious identities, and the ǫ's denote the required structure maps. These
diagrams commute since,
µm+1,n(ǫ ∧ id)(α ∧ β)t = µm+1,n(ǫ(α) ∧ β)t
b⊗K) ◦ (idSb⊗αt) ◦ (∆b⊗idA b⊗K)i) ◦ (∆b⊗idA b⊗K)
b⊗K) ◦ (idS b⊗Sb⊗αt) ◦ (∆b⊗idS b⊗A b⊗K) ◦ (∆b⊗idA b⊗K)
= (βr(t)b⊗idΣFm+1,0) ◦ (idSb⊗ǫ(αt)) ◦ (∆b⊗idA b⊗K)
= (βr(t)b⊗idΣFm+1,0) ◦ (idSb⊗h(bb⊗idB b⊗Fm,0
= (idΣF1,0b⊗βr(t)b⊗idFm,0) ◦ (bb⊗idS b⊗B b⊗Fm,0
b⊗K) ◦ (idSb⊗βr(t)b⊗idFm,0) ◦ (idS b⊗Sb⊗αt)
= (bb⊗idC b⊗Fm+n,0
◦ (∆b⊗idS b⊗A b⊗K) ◦ (∆b⊗idA b⊗K) by Lemma 1.3,
= (bb⊗idC b⊗Fm+n,0
b⊗K) ◦ (idSb⊗
(cid:2)(βr(t)b⊗idFm,0) ◦ (idSb⊗αt) ◦ (∆b⊗idA b⊗K)(cid:3)) ◦ (∆b⊗idA b⊗K)
= ǫ((βr(t)b⊗idFm,0) ◦ (idSb⊗αt) ◦ (∆b⊗idA b⊗K))
= ǫ(µm,n)(α ∧ β)t,
20
/
/
/
/
and
µm,n+1(id ∧g ǫ)(α ∧g β)t = µm,n+1(α ∧ ǫ(β))t
= (ǫ(β)r(t)b⊗idFm,0) ◦ (idSb⊗αt) ◦ (∆b⊗idA b⊗K)
= h(bb⊗idC b⊗Fn,0
= (bb⊗idC b⊗Fm+n,0
= (bb⊗idC b⊗Fm+n,0
= (bb⊗idC b⊗Fm+n,0
= ǫ((βr(t)b⊗idFm,0) ◦ (idSb⊗αt) ◦ (∆b⊗idA b⊗K))
b⊗K) ◦ (idSb⊗βr(t) ◦ (∆b⊗idB b⊗K)ib⊗idFm,0) ◦ (idSb⊗αt) ◦ (∆b⊗idA b⊗K)
b⊗K) ◦ (idSb⊗αt) ◦ (∆b⊗idA b⊗K)
b⊗K) ◦ (idSb⊗βr(t)b⊗idFm,0) ◦ (∆b⊗idB b⊗Fm,0
b⊗K) ◦ (idSb⊗βr(t)b⊗idFm,0) ◦ (idS b⊗Sb⊗αt)
◦ (∆b⊗idS b⊗A b⊗K) ◦ (∆b⊗idA b⊗K) by Lemma 1.3,
b⊗K) ◦ (idSb⊗(cid:2)(βr(t)b⊗idFm,0) ◦ (idSb⊗αt) ◦ (∆b⊗idA b⊗K)(cid:3)) ◦ (∆b⊗idA b⊗K)
= ǫ(µm,n)(α ∧ β)t,
for all α ∧ β ∈ Em(A, B) ∧ En(B, C).
Now we check that our product is associative up to homotopy.
Let α ∈ Asyg(Sb⊗Ab⊗K, Bb⊗Fm,0b⊗K) and β ∈ Asyg(Sb⊗Bb⊗K, Cb⊗Fn,0b⊗K) and γ ∈
Asyg(Sb⊗Cb⊗K, Db⊗Fp,0b⊗K). Then take the homotopy classes of these elements and we ob-
tain E-theory groups, and we know that the E-theory product is associative.
4 Connecting graded K and E-theory spectra
This section connects together graded K-theory and E-theory spectra. In particular we form
a smash product in terms of these two spectra and consequently combine K-theory and K-
homology in a smash product. Here we obtain a connection between K-theory, K-homology
and E-theory spectra. Let Hom(A, B) denote the set of graded ∗-homomorphisms from A
to B. Then recall that
and the K-homology is given by
Kn(A) = [S, ΣnAb⊗K(H)] ∼= En
g (F, A),
Khom(A) = En(A, F).
4.1 A topology on graded ∗-homomorphisms
Let Homg(A, B) denote the set of graded ∗-homomorphisms from A to B. We equip
Homg(A, B) with the compact open topology as detailed below.
Definition 4.1. A basis for a topology on a set A is a collection of subsets A of A such
that A is a union of sets from A and such that if A1, A2 are in A then their intersection is a
union of sets from A.
A subbasis is a collection of subsets B of A where the set A of all finite intersections of
sets in B is a basis.
21
Definition 4.2. Let A and B be graded C ∗-algebras. The compact open topology on the
set of graded ∗-homomorphisms from A to B, Homg(A, B), is generated by subsets of the
following form,
B(K, U) = {f ∈ Homg(A, B) f (K) ⊂ U},
where K is compact in A and U is open in B. Here generated means that the sets defined
form a subbasis for the open sets when we think of a topology. This then generates a basis
for the topology.
For simplicity of notation let Homg(A, B) denote the space of graded ∗-homomorphisms
from A to B equipped with the compact open topology. Denote the loop space of this
space by ΩHomg(A, B). Note that the basepoints for both of these spaces is just the zero
∗-homomorphism, which will we denote by 0.
The compact open topology is the choice for our topology since it gives us the correct
path components for our loop space and it also allows us to have continuity of particular
maps as we will see soon.
Now let us consider the generators of the compact open topology on the spaces ΩHomg(A, B)
and Homg(A, ΣB). Now we have a basis for Homg(A, B), so we just extend this for ΩHomg(A, B),
and it is not to hard to see that a basis for the loop space is the set generated by B(K ′, V )
such that K ′ ⊆ [0, 1] compact and V ⊆ Homg(A, B) open.
Combining these, we obtain the following definition.
Definition 4.3. The compact open topology on ΩHomg(A, B) is generated by sets of the
form B(K ′, B(K, U)), where K ⊆ A compact, K ′ ⊆ [0, 1] compact and U ⊆ B open. The
compact open topology on Homg(A, ΣB) is generated by sets of the form B(K, B(K ′, U))
where K ⊆ A compact, K ′ ⊆ [0, 1] compact and U ⊆ B open.
Before we check we have the continuity of maps in the following proof, it is worth noting
that it is sufficient to check that a map is continuous under a topology by considering a
subbasis. That is, to check a map of topological spaces is continuous we just need to check
that continuity holds at the level of generating sets for a basis of a topology. For details,
see [13], Application 3.2.5.
Proposition 4.4. The spaces ΩHomg(A, B) and Homg(A, ΣB) are homeomorphic.
Proof. We consider ungraded ∗-homomorphisms since the grading property is immediate.
Define f : ΩHom(A, B) → Hom(A, ΣB) as follows. Let µ ∈ ΩHom(A, B) based at 0, then
define
f (µ)(a)(s) = µ(s)(a),
for all a ∈ A and s ∈ [0, 1].
Now define g : Hom(A, ΣB) → ΩHom(A, B) as follows. Let τ ∈ Hom(A, ΣB), define
g(τ )(s)(a) = τ (a)(s).
Both f and g are well defined since µ and τ are ∗-homomorphisms.
We need to show that f ◦ g = id, and g ◦ f = id where id stands for the natural identities.
22
Let ϕ ∈ Hom(A, ΣB), then for all a ∈ A, s ∈ [0, 1].
f g(ϕ)(a)(s) = g(ϕ)(s)(a) = ϕ(a)(s).
Similarly, let ψ ∈ ΩHom(A, B), then for all a ∈ A, s ∈ [0, 1],
gf (ψ)(s)(a) = f (ψ)(a)(s) = ψ(s)(a).
Then f ◦ g = id and g ◦ f = id as required.
Now we check that f and g are continuous. By the above discussion, it suffices to check
that:
and
f −1[B(K ′, B(K, U))] = B(K, B(K ′, U))
g−1[B(K, B(K ′, U))] = B(K ′, B(K, U)),
for all K ⊆ A compact, K ′ ⊆ [0, 1] compact and U ⊆ B open.
Let y ∈ B(K, B(K ′, U)), then f −1(y) = {x f (x) = y}. Now let x ∈ f −1(y), then we
know
f (x)(s)(a) = x(a)(s) = y(s)(a),
so x must be contained in B(K ′, B(K, U)), and similarly we can check the converse, so f is
continuous.
Similarly we can prove that g is continuous.
4.2 K-theory spectra
Now we can define the K-theory spectrum.
Definition 4.5. Let K = K(H). Define K(A) to be the sequence of based topological spaces
where m ≥ 0. Define maps ǫn : Kn → ΩKn+1:
Kn = Homg(Sb⊗Fb⊗K, Ab⊗Fn,0b⊗K)
Homg(Sb⊗Fb⊗K, Ab⊗Fn,0b⊗K)
∼=
/ ΩHomg(Sb⊗Fb⊗K, Ab⊗Fn+1,0b⊗K)
Homg(Sb⊗Ab⊗K, Σ(Ab⊗Fn+1,0)b⊗K)
for all xt ∈ Asyg(Sb⊗Ab⊗K, Bb⊗Fn,0b⊗K) and the Bott map b ∈ Homg(S, ΣF1,0).
b⊗K) ◦ (idSb⊗xt) ◦ (∆b⊗idF b⊗K),
ǫ(xt) = (bb⊗idA b⊗Fn,0
by:
We now give an alternative definition for the spectrum of graded K-theory in terms of
asymptotic morphisms.
23
/
Definition 4.6. Let K = K(H). Define K′(A) to be the orthogonal quasi-spectrum with
the sequence of based quasi-topological spaces
where n ≥ 0. The structure maps ǫ : K ′
K ′
n = Asyg(Sb⊗Fb⊗K, Ab⊗Fn,0b⊗K)
n → ΩK ′
n+1:
Asyg(Sb⊗Fb⊗K, Ab⊗Fn,0b⊗K)
/ ΩAsyg(Sb⊗Fb⊗K, Ab⊗Fn+1,0b⊗K)
Asyg(Sb⊗Fb⊗K, Σ(Ab⊗Fn+1,0)b⊗K)
for all xt ∈ Asyg(Sb⊗Ab⊗K, Bb⊗Fn,0b⊗K) and the Bott map b ∈ Homg(S, ΣF1,0).
b⊗K) ◦ (idSb⊗xt) ◦ (∆b⊗idF b⊗K),
ǫ(xt) = (bb⊗idA b⊗Fn,0
are defined by:
∼=
We now notice that Definition 4.5 and Definition 4.6 are orthogonal and orthogonal quasi-
spectra for the same reason that 3.5 forms one and consequently the following result comes
from the stable homotopy groups coming from these spectra.
Proposition 4.7. The map of spectrum f : K(A) → K′(A), defined by f (ϕ) = ϕ for all
ϕ ∈ K(A) is a weak equivalence.
Proof. Consider the map f ′ : Homg(S, A) → Asyg(S, A) then this induces the map f ′
JS, AK. Now the map,
∗ : [S, A] →
induces an isomorphism at the level of π0, and therefore the map
Homg(S, Ab⊗Fn+1,0) → Asyg(S, Ab⊗Fn+1,0),
Homg(S, Ab⊗F1,0) → Asyg(S, Ab⊗F1,0),
induces an isomorphism at the level of πn. Therefore we have a weak equivalence. Then
we can also consider the map f above and the same applies, since we obtain this map by
tensoring with the suspension and the complex numbers.
Corollary 4.8. The map of spectrum f : K(A) → K′(A) has a natural inverse g : K′ → K
at the level of stable homotopy groups.
Theorem 4.9. Let A and B be C ∗-algebras. Then there is a natural map of orthogonal
quasi-spectra
ν ′
m,n : K(A) ∧ E(A, B) → K′(B),
defined by
(α ∧ βt)t 7→ (βtb⊗idFm,0) ◦ (idSb⊗α) ◦ (∆b⊗idF b⊗K(H)),
where α ∈ Homg(Sb⊗Ab⊗K(H), Bb⊗Fm,0b⊗K(H)) and β ∈ Asyg(Sb⊗Bb⊗K(H), Cb⊗Fn,0b⊗K(H)).
24
/
Proof. Since the composition of a ∗-homomorphism and an asymptotic morphism is an
asymptotic morphism it is clear that α ∧ β is an asymptotic morphism and lies in the
required spectra.
By the above theorem and Corollary 4.8, we obtain
Corollary 4.10. There is a natural map of spectra
νm,n : K(A) ∧ E(A, B) → K(B),
with the above criteria.
Now we finalise this section by combining the graded K-theory spectrum and K-homology
spectrum noting that Khom(A) = E(A, F).
Theorem 4.11. There is a canonical map
of orthogonal quasi-spectra. The map S is natural in the variable B in the obvious sense and
natural in the variable A, in the sense that if we have a ∗-homomorphism f : A → A′ then
we have the following commutative diagram
S : K(Ab⊗B) ∧ Khom(A) → K′(B)
S
/ K′(B)
K′(B)
f∗∧id
id∧f ∗
K(Ab⊗B) ∧ Khom(A)
K(Ab⊗B) ∧ Khom(A′)
K(A′b⊗B) ∧ Khom(A′)
f∗(α)(λ) = (fb⊗idB b⊗Fm,0
f ∗(βt)(a) = βt(fb⊗idS b⊗K(H))(a),
S /
/ K′(B)
b⊗K(H))(α(λ)),
where f∗ and f ∗ are defined by:
and
Proof. Writing Khom(A) = E(A, F), we can extend the definition of S to a composition of
maps, in order to obtain the following diagram:
with α ∈ K(Ab⊗B), β ∈ Khom(A′), a ∈ Sb⊗Ab⊗K and λ ∈ Sb⊗Fb⊗K.
/ K(Ab⊗B) ∧ E(Ab⊗B, B)
id∧ b⊗B /
ν ′
m,n
/ K′(B)
K′(B)
ν ′
m,n
/ K′(B)
id∧f ∗
K(Ab⊗B) ∧ E(A, F)
K(Ab⊗B) ∧ E(A′, F)
K(A′b⊗B) ∧ E(A′, F)
f∗∧id
id∧ b⊗B /
/ K(A′b⊗B) ∧ E(A′ ⊗ B, B)
25
/
O
O
/
O
O
/
Then we have
K(Ab⊗B) ∧ E(Ab⊗B, B) → K′(B).
ν ′
= ν ′
m,n(id ∧b⊗B)(id ∧ f ∗)(α ∧ βt) = ν ′
m,n(α ∧ f ∗(βt)b⊗idB)
m,n(id ∧b⊗B)(α ∧ f ∗(βt))
= (f ∗(βt)b⊗idB b⊗Fm,0) ◦ (idSb⊗α) ◦ (∆b⊗idF b⊗K(H))
= ([βt ◦ (fb⊗idS b⊗K(H))]b⊗idB b⊗Fm,0) ◦ (idSb⊗α) ◦ (∆b⊗idF b⊗K(H))
b⊗K(H)) ◦ (idSb⊗α) ◦ (∆b⊗idF b⊗K(H))
= (βtb⊗idB b⊗Fm,0) ◦ (fb⊗idS b⊗B b⊗Fm,0
b⊗K(H)) ◦ α] ◦ (∆b⊗idF b⊗K(H))
= (βtb⊗idB b⊗Fm,0) ◦ (idSb⊗[(fb⊗idB b⊗Fm,0
= (βtb⊗idB b⊗Fm,0) ◦ (idSb⊗f∗(α)) ◦ (∆b⊗idF b⊗K(H))
m,n(f∗(α) ∧ (βtb⊗idB))
m,n(id ∧b⊗B)(f∗ ∧ id)(α ∧ βt).
m,n(id ∧b⊗B)(f∗(α) ∧ βt) = v′
= v′
= v′
Corollary 4.12. There is a canonical map
S : K(A ⊗ B) ∧ Khom(A) → K(B)
of orthogonal quasi-spectra. The map S is natural in the variable B in the obvious sense and
natural in the variable A, in the sense that if we have a ∗-homomorphism f : A → A′ then
we have the following commutative diagram
K(A ⊗ B) ∧ Khom(A)
S
/ K(B)
id∧f ∗
K(A ⊗ B) ∧ Khom(A′)
K(B)
f∗∧id
K(A′ ⊗ B) ∧ Khom(A′)
S /
/ K(B).
References
[1] G. Bredon, Topology and Geometry, Springer GTM 139, 1993.
[2] S. L. Browne, A Bott periodicity proof for real graded C ∗-algebras, arXiv:1611.09887.
[3] S. L. Browne, E-theory spectra, PhD thesis, University of Sheffield, available at
http://etheses.whiterose.ac.uk/id/eprint/17812.
26
/
O
O
[4] A. Connes and N. Higson, D´eformations, morphismes asymptotiques et K-th´eorie bi-
variante, C. R. Acad. Sci. Paris S´er. I Math., 311, 1990, 101-106.
[5] M. Dadarlat and R. Meyer, E-theory for C ∗-algebras over topological spaces, J. Funct.
Anal., 263(1):216-247, 2012.
[6] E. Guentner and N. Higson, Group C ∗-algebras and K-theory, Noncommutative geom-
etry, Lecture Notes in Math., 1831, 137–251, Springer, Berlin, 2004.
[7] E. Guentner, N. Higson and J. Trout, Equivariant E-Theory for C ∗-algebras, Memoirs
of the American Mathematical Society, 2000.
[8] N. Higson, Categories of fractions and excision in KK-theory, J. Pure Appl. Algebra,
65, 1990, 119-138.
[9] N. Higson and G. Kasparov, Operator K-theory for groups which act properly and iso-
metrically on Hilbert space, Electronic Research Announcements of the American Math-
ematical Society, 3 (1997), pages 131-142.
[10] M. Hovey, B. Shipley and J. Smith, Symmetric spectra, J. Amer. Math. Soc., volume
13, pages 149–208, 2000.
[11] P. D. Mitchener, Symmetric K-theory spectra of C ∗-categories, K-theory, 24 (2001),
pages 157-201.
[12] E. Spanier, Quasi-topologies, Duke Math. J., 20:1–14, 1963.
[13] W. A. Sutherland, Introduction to Metric and Topological Spaces, Oxford Science Pub-
lications, 1975.
[14] M. Weiss and B. Williams. Assembly. In Novikov conjectures, index theorems and rigid-
ity, Vol. 2 (Oberwolfach, 1993), volume 227 of London Math. Soc. Lecture Note Ser.,
pages 332-352. Cambridge Univ. Press, Cambridge, 1995.
27
|
1811.01922 | 3 | 1811 | 2018-11-28T18:58:53 | Quantum extensions of ordinary maps | [
"math.OA",
"math-ph",
"math.GN",
"math-ph"
] | We define a loop to be quantum nullhomotopic if and only if it admits a nonempty quantum set of extensions to the unit disk. We show that the canonical loop in the unit circle is not quantum nullhomotopic, but that every loop in the real projective plane is quantum nullhomotopic. Furthermore, we apply Kuiper's theorem to show that the canonical loop admits a continuous family of extensions to the unit disk that is indexed by an infinite quantum space. We obtain these results using a purely topological condition that we show to be equivalent to the existence of a quantum family of extensions of a given map. | math.OA | math |
QUANTUM EXTENSIONS OF ORDINARY MAPS
ANDRE KORNELL
Department of Mathematics
University of California, Davis
Abstract. We define a loop to be quantum nullhomotopic if and only if it admits a
nonempty quantum set of extensions to the unit disk. We show that the canonical loop
in the unit circle is not quantum nullhomotopic, but that every loop in the real projective
plane is quantum nullhomotopic. Furthermore, we apply Kuiper's theorem to show that the
canonical loop admits a continuous family of extensions to the unit disk that is indexed by
an infinite quantum space. We obtain these results using a purely topological condition that
we show to be equivalent to the existence of a quantum family of extensions of a given map.
Quantum families of extensions. Noncommutative mathematics draws an analogy be-
tween unital C*-algebras and compact Hausdorff spaces. Commutative unital C*-algebras
correspond bijectively to compact Hausdorff spaces, up to isomorphism of the former and
up to homeomorphism of the latter, so noncommutative unital C*-algebras are viewed as
"quantum compact Hausdorff spaces". Specifically, each compact Hausdorff space X corre-
sponds to the commutative C*-algebra C(X) of continuous complex-valued functions on X,
and conversely, each noncommutative unital C*-algebra is imagined to have the same form,
but for some fictitious space X, whose local topology is quantum in the tautological sense
that some real-valued functions on this space fail to commute, in the manner of incompatible
observables.
This correspondence between compact Hausdorff spaces and commutative unital C*-
algebras extends to a correspondence between continuous maps and unital ∗-homomorphisms.
Each map f from a compact Hausdorff space X to a compact Hausdorff space Y induces
a unital ∗-homomorphism from C(Y ) to C(X), by precompostion. We thus have a con-
travariant equivalence of categories. Each unital ∗-homomorphism between noncommuta-
tive unital C*-algebras is thus viewed as a continuous function between the corresponding
quantum spaces, but in the opposite direction. This "noncommutative metaphor" can be
extended further and further [6]. For the purposes of the this paper, we only recall that
C(X ⊔ Y ) ∼= C(X) ⊕ C(Y ) and C(X × Y ) ∼= C(X) ⊗ C(Y ), for compact Hausdorff spaces X
and Y , so we take the direct sum and the minimal tensor product of C*-algebras to be the
appropriate generalizations of disjoint union and Cartesian product of compact Hausdorff
spaces. The minimal tensor product is preferred over other notions because it distributes
over the direct sum.
Let S and T be compact Hausdorff spaces, and f an ordinary map from S to T . If S is a
subspace of a larger compact Hausdorff space S, then an extension of f to S is of course a
E-mail address: [email protected].
1
map f : S → T making the following diagram commute:
S
j
S
f
f
T
We write j : S ֒→ S for the inclusion map. A family of extensions indexed by a compact
Hausdorff space Y is instead a continuous function f of two variables making making the
following diagram commute:
S × Y
j×id
f
S × Y
f ×id
T × Y
T
proj1
The existence of a nonempty family of extensions indexed by an ordinary compact Haus-
dorff space is of course equivalent to the existence of a single extension. We will show that
there is no such equivalence when we consider nonempty families of extensions indexed by a
quantum compact Hausdorff space. To make this claim precise, we first apply the functor C
to the diagram above, to obtain a diagram in the category of unital C*-algebras and unital
∗-homomorphisms:
C( S) ⊗ C(Y )
j ⋆⊗id
C(S) ⊗ C(Y )
f ⋆⊗id
C(T ) ⊗ C(Y )
C(T )
proj⋆
1
Allowing the index space Y to be quantum, we replace C(Y ), with an arbitrary unital C*-
algebra B. We say that the quantum space is nonempty just in case the C*-algebra B is
nonzero. Thus, we ask whether there exists a nonzero unital C*-algebra B, and a unital
normal ∗-homomorphism φ : C(T ) → C( S) ⊗ B making the following diagram commute:
C( S) ⊗ B
j ⋆⊗id
φ
C(S) ⊗ B
f ⋆⊗id
C(T ) ⊗ B
C(T )
a⊗1←[a
The core result of the paper provides a purely topological condition equivalent to the
existence of a family of extensions indexed by a quantum compact Hausdorff space.
Theorem 1.6. Let B be a unital C*-algebra. Let ı : T → Hom(C(T ), B) be the map taking
each point of T to evaluation at that point. There exists a unital ∗-homomorphism C(T ) →
C( S) ⊗ B making the above diagram commute if and only if the ordinary map ı ◦ f extends
to S:
S
f
j
S
T
f
ı
2
Hom(C(T ), B)
Thus, the functor Hom(C( − ), B) is analogous to the quantum monad of Abramsky, Barbosa,
de Silva, and Zapata [1].
Quantum nullhomotopic loops. Recall that a loop f : S1 → T is nullhomotopic if and
only if f extends to a map on the unit disk D2. As an application of the above theorem, we
investigate the corresponding quantum notion. Setting S = S1 and S = D2, we ask whether
a map f : S1 → T admits a family of extensions indexed by a nonempty quantum compact
Hausdorff space.
This notion of quantum nullhomotopy degenerates if we allow arbitrary quantum compact
Hausdorff spaces as index spaces. As a consequence of Kuiper's theorem [8] we show that
the canonical loop S1 → S1 admits a quantum family of extensions to the unit disk:
Theorem 2.6. Let H be an infinite-dimensional Hilbert space, and let L(H) be the unital
C*-algebra of bounded operators on H. There is a unital ∗-homomorphism φ such that the
following diagram commutes:
C(D2) ⊗ L(H)
j ⋆⊗id
φ
C(S1) ⊗ L(H)
a⊗1←[a
C(S1)
It follows that every loop in every compact Hausdorff space admits a quantum family of
extensions to the unit disk indexed by the quantum compact Hausdorff space corresponding
to L(H). Therefore, we make the following definition:
Definition 2.1. A loop f : S1 → T is quantum nullhomotopic just in case there is a nonzero
finite-dimensional Hilbert space H and unital ∗-homomorphism φ such that the following
diagram commutes.
C(D2) ⊗ L(H)
j ⋆⊗id
φ
C(S1) ⊗ L(H)
f ⋆⊗id
C(T ) ⊗ L(H)
C(T )
a⊗1←[a
Equivalently, we may replace the arbitrary nonzero matrix algebra L(H) with an arbitrary
nonzero finite-dimensional C*-alegebra B in this definition.
This notion of quantum nullhomotopy does not degenerate:
Proposition 2.4. The identity function S1 → S1 is not quantum nullhomotopic as a loop
in S1.
Furthermore, it is distinct from the ordinary notion of nullhomotopy of loops:
Corollary 3.3. Every loop in RP2 is quantum nullhomotopic.
Corollary 3.5. There is a loop in S1 ∨ S1 that is quantum nullhomotopic, but not nullho-
motopic in the ordinary sense.
3
Quantum sets. If X is a locally compact Hausdorff space, then
C0(X) = {f ∈ (X, C)
lim
x→∞
f (x) = 0}
is a C*-algebra, which is unital iff X is compact. Thus, C*-algebras are commonly viewed to
be the quantum generalization of locally compact Hausdorff spaces. In particular, quantum
sets should correspond to a class of C*-algebras.
The noncommutative dictionary does not yet include a widely accepted quantum general-
ization of sets. If the C*-algebra of compact operators is taken to correspond to a quantum
set, then in fact there does exist a nonempty quantum set of extensions of the canonical
loop to the unit disk. The unital ∗-homomorphism φ of theorem 2.6 can be viewed as a
morphism, in the sense of Woronowicz [6], from C(S1) to C(D2) ⊗ L0(H), and thus, as a
quantum family of extensions that is indexed by the quantum set corresponding the compact
operator algebra L0(H).
The notion of quantum set preferred by the author excludes those quantum locally com-
pact Hausdorff spaces that correspond to infinite-dimensional compact operator C*-algebras.
Instead, quantum sets are taken to be those quantum locally compact Hausdorff spaces that
correspond to c0-direct sums of finite matrix C*-algebras. This notion of discreteness first
appeared in the context of quantum Gelfand duality [13] [5] [15] [4]. It was later promoted
by the author [7].
Decomposing the C*-algebra of a quantum set into matrix algebras, we find that a loop
is quantum nullhomotopic if and only if it admits a quantum set of extensions to the unit
disk, that is, a quantum family of extensions indexed by a discrete quantum space.
Quantum pseudotelepathy. Quantum families of functions can be interpreted as quantum
strategies for games in which two players, traditionally named Alice and Bob, cooperate
against a Referee without communicating with one another. In some instances, Alice and
Bob have a winning strategy that utilizes quantum entanglement, despite having no winning
strategy classically. The availability of a quantum strategy is equivalent to the existence of
a quantum set of functions of one or another kind [2, proposition 1] [9] [1] [11] [7, 1.2].
In this context, the domain and codomain spaces are taken to be finite. However, it is also
possible to interpret the quantum extension problem considered in this paper as an idealized
game of the same kind. In the example of theorem 2.6, the Referee sends Alice and Bob
elements of D2, and they respond with elements of S1. Alice and Bob lose if their responses
disagree, provided that the Referee sent them both the same element of D2, or if either Alice
or Bob fails to mirror the Referee's move, provided that the Referee sent them an element
of S1.
I speculate that this idealized game is not so far removed from the physical world as one
might imagine. The players could perhaps exchange continuous values as the momenta of
particles, and the Referee could be certain of sending the same value to both Alice and Bob
by using a mechanism such as pair creation. The strategies implemented by Alice and Bob
would necessarily be continuous, as a physical limitation.
Preliminary computations suggest that it is possible to formulate some of the results of the
present paper in terms of discrete games. However, these games retain some infinitary ele-
ment, e. g., the number of possible moves is infinite, or the length of play is unbounded. Fur-
thermore, the quantum advantage becomes probabilistic, rather than deterministic. Thus,
the significance of such reformulations is uncertain.
4
Acknowledgements. I thank Neil Ross and Peter Selinger for organizing the Quantum
Physics and Logic conference in Halifax, an enriching experience that led me to consider
this quantum extension problem. I thank Greg Kuperberg for advising me to emphasize the
topological aspect of the problem, in favor of the quantum communication aspect. I thank
Rui Soares Barbosa and Sam Staton for giving me the opportunity to present this research
at the Oxford Advanced Seminar on Informatic Structures.
1. quantum extensions
Let S and T be compact Hausdorff spaces, and let H be a nonzero finite-dimensional
Hilbert space.
In this section, we demonstrate a one-to-one correspondence between the
unital ∗-homomorphisms from C(T ) to C(S) ⊗ L(H), and the continuous functions from S
to Hom(C(T ), L(H)). There are several reasonable ways to gloss the term C(S) ⊗ L(H) that
turn out to be equivalent, and likewise for Hom(C(T ), L(H)). The C*-algebras C(S) and
L(H) are both nuclear, so there is a unique cross-norm on their algebraic tensor product.
In fact, their algebraic tensor product is already isomorphic to the C*-algebra of dim(H) ×
dim(H) matrices over C(S), so the term C(S)⊗L(H) can refer equivalently to the maximum
tensor product, the minimum tensor product, or the algebraic tensor product of the two C*-
algebras.
The space Hom(C(T ), L(H)) consists of unital ∗-homomorphisms from C(T ) to L(H).
This set is typically equipped with the point-norm topology, which is characterized by the
condition that a net (φα) converges to φ∞ if and only if for each function b in C(T ),
the net (φα(b)) converges to φ∞(b) in the operator norm topology. We instead equip
Hom(C(T ), L(H)) with the kaonization of the compact-open topology, which we will later
show to be equivalent to the point-norm topology. We do so to reason in Steenrod's conve-
nient category of compactly generated spaces [14]. We summarize a few of its basic aspects.
Steenrod's convenient category is a subcategory of the category of topological spaces and
continuous functions that includes all locally compact spaces and all metrizable spaces. It is
convenient primarily in the sense that it is Cartesian closed: for all objects X, Y , and Z of
the category, the space of continuous functions C(Y, Z) is itself an object of the category, and
there is a natural one-to-one correspondence between the continuous function from X × Y to
Z and the continuous functions from X to C(Y, Z). However, the category is inconvenient in
the sense that the topology of the category-theoretic product X × Y is sometimes different
from the usual product topology.
The objects of Steenrod's category are the so-called compactly generated Hausdorff spaces.
These spaces are characterized by the criterion that any set whose intersection with every
compact set is closed must itself be closed. Kaonization is the functor from Hausdorff spaces
to compactly generated Hausdorff spaces that adds closed sets according to this criterion;
it is typically denoted by the lower-case letter k. The category-theoretic product of two
compactly generated Hausdorff spaces X and Y is the kaonization of the usual topological
product. The morphisms of Steenrod's category are ordinary continuous functions, and
the set of continuous functions C(Y, Z) is canonically equipped with the kaonization of the
If X is locally compact, then the product topology
compact-open topology on this set.
on X × Y is already compactly generated.
If Y is compact, and Z is metrizable, then
the compact-open topology on C(Y, Z) is also already compactly generated; it is simply the
uniform topology. In particular, in Steenrod's category, C(S) = C(S, C) and C(T ) = C(T, C)
are canonically equipped with their usual norm topologies.
5
With this overview of the compactly generated spaces, we are ready to formally state and
prove the claimed one-to-one correspondence.
Definition 1.1. Let A and B be unital C*-algebras. We write Hom(A, B) for the set of all
unital ∗-homomorphisms from A to B, equipped with the kaonization of the compact-open
topology.
The space Hom(A, B) is a subspace of the space C(A, B), itself equipped with the kaoniza-
tion of its compact open topology. It easy to see that for closed subsets, the kaonization of
the subspace topology coincides with the subpace topology of the kaonization.
Lemma 1.2. We have a homeomorphism Hom(A, C(S) ⊗ B) ∼= C(S, Hom(A, B)), natural
in the compact Hausdorff space S, and the unital C*-algebras A and B.
Implicitly, the morphisms of compact Hausdorff spaces are continuous functions, and the
morphisms of unital C*-algebras are unital ∗-homomorphisms.
Proof. The compact-open topology on C(S, B) is the topology of uniform convergence. It
is a metrizable topology, so it is compactly generated; thus, the topology on C(S, B) in
Steenrod's convenient category is just the usual one. There is a well known isomorphism
between the C*-algebras C(S) ⊗ B and C(S, B) [10, theorem 6.4.17], which is natural in S
and B, so it is sufficient to establish that Hom(A, C(S, B)) ∼= C(S, Hom(A, B)).
We obtain this natural homeomorphism as a restriction of the following composition:
C(A, C(S, B)) ∼= C(A × S, B) ∼= C(S, C(A, B))
Both natural homeomorphisms are instances of the Cartesian closedness of Steenrod's conve-
nient category. Each continuous function π from A to C(S, B) corresponds to a continuous
function π′ from S to C(A, B), which is defined by π′(s)(a) = π(a)(s) for all a ∈ A, for
all s ∈ S. Thus, if π is a unital ∗-homomorphism, then for each s ∈ S, π′(s) is a uni-
tal ∗-homomorphism, and vica versa, because the algebraic structure of C(S, B) is defined
pointwise. We conclude that the natural homeomorphism C(A, C(S, B)) ∼= C(S, C(A, B))
restricts to a natural homeomorphism Hom(A, C(S, B)) ∼= C(S, Hom(A, B)).
(cid:3)
For reference, we record that each unital ∗-homomorphism φ : A → C(S) ⊗ B corresponds
to the continuous function s 7→ (evals ⊗ idB)(φ( · )).
Lemma 1.3. Let V and W be Banach spaces. Write L(V, W )1 for the set of linear operators
from V to W of norm at most 1. The point-norm topology on L(V, W )1 is equal to the
compact-open topology.
The point-norm topology is generated by subbasis elements of the form {t kt(v)−wk < ǫ}
for v ∈ V , w ∈ W , and ǫ > 0. The compact-open topology is of course generated by subbasis
elements of the form {t t(K) ⊆ U} for K ⊆ V compact, and U ⊆ W open. Both Banach
spaces are equipped with their norm topologies.
Proof. It is well known that convergence in the compact-open topology is equivalent to
uniform convergence on compact subsets of the domain, whenever the codomain is a metric
space. The point-norm topology is just the topology of pointwise convergence. Convergence
of the former kind clearly implies convergence of the latter kind, so it remains only to show
the converse.
6
Let (tα) be a net in L(V, W )1 converging pointwise to a linear operator t∞, itself au-
tomatically of norm at most 1. Let K be a compact subset of V , and let ǫ > 0. Being
compact, the set V is totally bounded, so we can cover K by a finitely many open balls
Bǫ(v1), Bǫ(v2), . . . , Bǫ(vn) of radius ǫ, with centers v1, v2, . . . , vn ∈ V . Consider α sufficiently
large so that ktα(v1) − t(v1)k < ǫ, ktα(v2) − tα(v2)k < ǫ, . . ., and ktα(vn) − tα(vn)k < ǫ. Every
element v of K is within distance ǫ of some vi, so we calculate:
ktα(v) − t(v)k ≤ ktα(v) − tα(vi)k + ktα(vi) − t∞(vi)k + kt∞(vi) − t∞(v)k
≤ ktαk · kv − vik + ǫ + kt∞k · kv − vik ≤ 3ǫ
Thus, (tα) converges uniformly to t∞ on K. Therefore, pointwise convergence implies uniform
convergence on compact subsets.
(cid:3)
Proposition 1.4. We have a bijection, natural in the compact Hausdorff space S, and the
unital C*-algebras A and B,
Hom(A, C(S) ⊗ B) ∼= C(S, Hom(A, B)),
where Hom(A, B) denotes the set of all unital ∗-homomorphisms from A to B, equipped with
the point-norm topology. It takes each each unital ∗-homomorphism φ : A → C(S) ⊗ B to
the function s 7→ (evals ⊗ id)(φ( · )).
Proof. This proposition is just lemma 1.2, with a different topology on Hom(A, B). The
point-norm topology on Hom(A, B) is equal to the compact-open topology by lemma 1.3.
Thus, the topology on Hom(A, B) in lemma 1.2 is equivalently the kaonization of the point-
norm topology. Since S is a compact Hausdorff space, the set of continuous functions from S
to Hom(A, B) is the same if we equip Hom(A, B) with just the point-norm topology, without
kaonizing it. Thus, the proposition follows.
(cid:3)
Lemma 1.5. Let T be a compact Hausdorff space. The bijection of proposition 1.4 takes
the canonical unital ∗-homomorphism π : C(T ) → C(T ) ⊗ B, defined by π(a) = a ⊗ 1 to the
continuous function ı : T → Hom(C(T ), B), defined by ı(t)(a) = a(t) · 1.
Proof. Let p be the function from T to Hom(C(T ), B) that corresponds to π under the
bijection of proposition 1.4. For each t ∈ T , and all a ∈ C(T ), we compute:
p(t)(a) = (evalt ⊗ id)(π(a)) = (evalt ⊗ id)(a ⊗ 1) = a(t) ⊗ 1 = a(t) · 1.
(cid:3)
Theorem 1.6. Let S and T be compact Hausdorff spaces, and let B be a unital C ∗-algebra.
Let S be a closed subset of S, with inclusion function j : S ֒→ S. Let f be a continuous
function from S to T . Use the notation of proposition 1.4 and lemma 1.5. The following are
equivalent:
(1) The function ı ◦ f : S → Hom(C(T ), B) extends to a continuous function f : S →
Hom(C(T ), B).
S
j
S
f
T
f
ı
7
Hom(C(T ), B)
(2) There is a unital ∗-homomorphism φ : C(T ) → C( S) ⊗ B such that
(j⋆ ⊗ id) ◦ φ = (f ⋆ ⊗ id) ◦ π.
C( S) ⊗ B
j ⋆⊗id
φ
C(S) ⊗ B
f ⋆⊗id
C(T ) ⊗ B
C(T )
π
Proof. Apply proposition 1.4, with A = C(T ) to obtain the following commutative diagram:
Hom(C(T ), C( S) ⊗ B))
(j ⋆⊗id)◦
Hom(C(T ), C(S) ⊗ B))
(f ⋆⊗id)◦
Hom(C(T ), C(T ) ⊗ B))
∼=
∼=
∼=
C( S, Hom(C(T ), B))
◦j
C(S, Hom(C(T ), B))
◦f
C(T, Hom(C(T ), B))
The homomorphism π is an element of Hom(C(T ), C(T ) ⊗ B)), and the map ı is an element
of C(T, Hom(C(T ), B)). By lemma 1.5, the two elements correspond to each other under
the indicated bijection. Condition (1) is equivalently that there exists an element f of
C( S, Hom(C(T ), B)) whose image in C(S, Hom(C(T ), B)) is equal to the image of ı in that
set. Condition (2) is equivalently that there exists an element φ of Hom(C(T ), C( S) ⊗
B)) whose image in Hom(C(T ), C(S) ⊗ B)) us equal to the image of π in that set. The
commutative diagram demonstrates that the two conditions are equivalent, establishing the
theorem.
(cid:3)
We close this section by observing that the space Hom(C(T ), Mn(C)) is necessarily com-
pact for every n.
Lemma 1.7. Let A be a unital C*-algebra, and let n be a positive integer. The space
Hom(A, Mn(C)) of unital ∗-homomorphisms from A to Mn(C), equipped with the compact-
open topology, or equivalently, with the point-norm topology, is compact. If A is separable,
then Hom(A, Mn(C)) has a countable basis for the point-norm topology.
Proof. The two topologies are equivalent as a special case of lemma 1.3. To observe com-
pactness, we put the set Hom(A, Mn(C)) through the following isomorphisms in Steenrod's
convenient category:
C(A, Mn(C)) ∼= C(A, C × · · · × C) ∼= C(A, C) × · · · × C(A, C)
The set Hom(A, Mn(C)) is closed in C(A, Mn(C)). A unital ∗-homomorphism has norm
at most 1, so Hom(A, Mn(C)) corresponds to a closed subset of L(A, C)1 × · · · × L(A, C)1.
By lemma 1.3, the compact-open topology on L(A, C)1 is just the point-norm topology, or
equivalently, the weak* topology. By Alaoglu's theorem, the unit ball L(A, C)1 is compact.
Thus, the topology on L(A, C)1 as a subspace of C(A, C) in Steenrod's convenient category is
just the weak* topology. Therefore the space Hom(A, Mn(C)) is compact. Furthermore, if A
is separable, then the unit ball L(A, C)1 is metrizable [12, exercise 2.5.3], implying the same
8
for Hom(A, Mn(C)). Of course, for compact Hausdorff spaces metrizability is equivalent to
the existence of a countable basis.
(cid:3)
2. quantum loops
We write S1 for the unit circle, D2 for the unit disk, and j : S1 ֒→ D2 for the inclusion
map.
Definition 2.1. Let T be a compact Hausdorff space. A loop in T is a function f : S1 → T .
A loop f is quantum nullhomotopic just in case there is a nonzero finite-dimensional Hilbert
space H and a unital ∗-homomorphism φ : C(T ) → C(D2) ⊗ L(H) such that (j⋆ ⊗ id)φ(a) =
f ⋆(a) ⊗ 1H for all a ∈ C(T ).
C(D2) ⊗ L(H)
j ⋆⊗id
φ
C(S1) ⊗ L(H)
f ⋆⊗id
C(T ) ⊗ L(H)
C(T )
a⊗1←[a
Equivalently, f is nullhomotopic if and only if there exists an positive integer n, and a unital
∗-homomorphism φ : C(T ) → Mn(C(D2)) such that applying j⋆ to every element of the
matrix φ(a) yields the matrix f ⋆(a) · In for all a ∈ C(T ).
Corollary 2.2. Let T be a compact Hausdorff space. A loop f in T is quantum nullhomotopic
if and only if there is a positive integer n such that the loop ı⋆(f ) = ı◦f in Hom(C(T ), Mn(C))
is nullhomotopic in the ordinary sense, where ı is defined by ı(t)(a) = a(t) · In.
Proof. Apply theorem 1.6, with S = S1 and S = D2.
(cid:3)
Proposition 2.3. Let g be a continuous function from a compact Hausdorff space T1 to a
compact Hausdorff space T2. If a loop f : S1 → T1 is quantum nullhomotopic in T , then the
loop g ◦ f : S1 → T2 is quantum nullhomotopic in T2.
Proof. Apply corollary 2.2. Let n be a positive integer such that the loop ı⋆(f ) : S1 →
Hom(C(T1), Mn(C)) is nullhomotopic in the ordinary sense. We show that the loop ı⋆(g ◦
f ) : S1 → Hom(C(T2), Mn(C)) is a continuous image of the loop ı⋆(f ).
ı⋆(f )
S1
Hom(C(T1), Mn(C))
ı⋆(g◦f )
◦g⋆
Hom(C(T2), Mn(C))
For each point s on the circle S1, and all a2 ∈ C(T2), we calculate that
ı⋆(g ◦ f )(s)(a2) = a2(g(f (s))) · In = g⋆(a2)(f (s)) · In = ı⋆(f )(s)(g⋆(a2)) = (ı⋆(f )(s) ◦ g⋆)(a2).
Thus, the homomorphism ı⋆(g◦f )(s) is obtained by composing ı⋆(f )(s) with g⋆. This is a con-
tinuous function from Hom(C(T1), Mn(C)) to Hom(C(T2), Mn(C)), since Hom(C( − ), Mn(C))
is a functor from the category of compact Hausdorff spaces and continuous functions to
Steenrod's convenient category. Being a continuous image of a nullhomotopic loop, the loop
ı⋆(g ◦ f ) is also nullhomotopic. Appealing a second time to corollary 2.2, we conclude that
g ◦ f is quantum nullhomotopic.
(cid:3)
9
Proposition 2.4. The identity function f : S1 → S1 is not quantum nullhomotopic as a
loop in S1.
Proof. Let f : S1 → S1 be the identity function, and suppose that f is quantum nullho-
motopic as a loop in S1. By corollary 2.2, there is a positive integer n such that the loop
ı⋆(f ) : S1 → Hom(C(S1), Mn(C)) is nullhomotopic. We will compose ı⋆(f ) with two other
continuous functions to obtain a loop which we know is not nullhomotopic, arriving at a
contradiction.
First, let z be the inclusion of S1 into C; it is a unitary operator in C(S1). Evaluation at
z is a function from Hom(C(S1), Mn(C)) to the space U(n) of unitary n × n matrices. This
function evalz is continuous because all evaluation functions are continuous in Steenrod's
convenient category.
Second, the determinant is a continuous function from U(n) to S1. Thus, we obtain a
nullhomotopic loop det ◦ evalz ◦ f in S1.
ı⋆(f )
S1
Hom(C(S1, Mn(C))
evalz
U(n)
det
S1
For each point s on the circle S1 we calculate that
(det ◦ evalz ◦ ı⋆(f ))(s) = det(z(f (s)) · In) = det(s · In) = sn.
It is a basic fact that this loop is not nullhomotopic. We have reached a contradiction.
Therefore, f is not quantum nullhomotopic.
(cid:3)
Lemma 2.5. Let B a be a unital C*-algebra, and let z : S1 ֒→ C be the inclusion map.
Evaluation at z is a homeomorphism
Hom(C(S1), B) → U(B),
where U(B) denotes the set of unitary operators in B, equipped with the norm topology.
Proof. Recall that the C*-algebra C(S1) is isomorphic to the group C*-algebra C∗(Z) [3,
proposition VII.1.1]. The inclusion function z : S1 ֒→ C is mapped to the generator u1
of C∗(Z). The C*-algebra C∗(Z) is evidently the universal C*-algebra for a single unitary
operator:
for any C*-algebra B, there is a bijective correspondence between the unital
∗-homomorphisms from C∗(Z) to B, and the unitary operators in B. Each such unital ∗-
homomorphism sends the generator u1 to the corresponding unitary in B. Therefore, we
have a bijection evalz from Hom(C(S1), B) to U(B), defined by ρ 7→ ρ(z).
The bijection evalz is continuous, because all evaluation functions are continuous in Steen-
rod's convenient category. To show that it is a homeomorphism,
let (ρλ) be a net in
Hom(C(S1), B) whose image (evalz(ρλ)) converges to evalz(ρ) for some homomorphism ρ
in Hom(C(S1), B). In other words, (ρλ(z)) converges to ρ(z). The operations that make
up the ∗-algebra structure of B are all continuous in the norm topology, so in fact, (ρλ(a))
converges to ρ(a) for all a in the unital ∗-subalgebra A0 of C(S1) that is generated by z.
Let a be any element of C(S1). The algebra A0 is dense in C(S1) by the Stone-Weierstrass
theorem, so we can choose a sequence (an) in A0 such that kan − ak ≤ 1/n for each positive
10
integer n. We now estimate that for each positive integer n,
kρ(a) − ρλ(a)k ≤ kρ(a) − ρ(an)k + kρ(an) − ρλ(an)k + kρλ(an) − ρλ(a)k
≤ kρk · kan − ak + kρ(an) − ρλ(an)k + kρλk · kan − ak
≤
=
1
n
2
n
+ kρ(an) − ρλ(an)k +
1
n
+ kρ(an) − ρλ(an)k
The term kρ(an) − ρλ(an)k converges to 0 as λ goes to infinity, so lim supλ kρ(a) − ρλ(a)k ≤
2/n. We have this estimate for each positive integer n, so limλ kρ(a) − ρλ(a)k = 0.
In
other words, ρλ converges to ρ in the point-norm topology. Therefore, eval−1
: U(B) →
z
Hom(C(S1), B) is norm-(point-norm) continuous, and thus, by lemma 1.3,
it is norm-
(compact-open) continuous.
It is an elementary fact about compactly generated spaces that kaonization does not
introduce or eliminate compact subset. In this case, if a set K ⊆ Hom(C(S1), B) if compact
in the kaonized compact-open topology, then it is compact in the compact-open topology
itself, and vice versa. Thus, we have shown that evalz induces a one-to-one correspondence
between the compact subsets of Hom(C(S1), B) and the compact subsets of U(B). Since
both spaces are compactly generated, evalz must be a homeomorphism.
(cid:3)
Theorem 2.6. Let H be a separable infinite-dimensional Hilbert space. There is a unital
∗-homomorphism φ : C(S1) → C(D2) ⊗ L(H) such that (j⋆ ⊗ id)(φ(a)) = a ⊗ 1H for all
a ∈ C(S1), where j : S1 ֒→ D2 is the inclusion map.
C(D2) ⊗ L(H)
j ⋆⊗id
φ
C(S1) ⊗ L(H)
a⊗1←[a
C(S1)
The proof of this theorem is a more elaborate variation on the proof of proposition 2.4.
We write U(H) for the set of unitary operators in L(H), equipped with the norm topology.
Proof. We apply theorem 1.6, with S = S1, T = S1, S = D2, B = L(H), and f = idS1.
Specifically, we show that the function ı : S1 → Hom(C(S1), L(H)), defined by ı(s)(a) =
a(s) · 1H for all a ∈ C(S1) and s ∈ S1, extends to D2.
Composing ı with the homeomorphism evalz of lemma 2.5, we obtain the loop evalz ◦ ı
in U(H). By Kuiper's theorem [8, theorem (3)], U(H) is contractible, so the loop evalz ◦ ı
is nullhomotopic. Since evalz is a homeomorphism by lemma 2.5, the loop ı must also be
nullhomotopic. Applying, theorem 1.6, we reach the desired conclusion.
(cid:3)
3. every loop in RP2 is quantum nullhomotopic
We define the circle S1 to be a subspace of C, and the sphere S2 to be a subspace of C × R.
Explicitly, the circle S1 consists of complex numbers α such that α2 = 1, and the sphere
S2 consists of pairs (α, t) in C × R such that α2 + t2 = 1. The sphere S2 is homeomorphic
to CP1; we explicitly describe one such homeomorphism that is convenient in this context.
11
Write R1(M2(C)) for the space of reflection matrices of negative determinant, i. e., for
the space of 2 × 2 complex matrices b such that b∗ = b, b2 = 1, and det(b) = −1. There is a
nice homeomorphism h : S2 → R1(M2(C)) given by the formula
h(α, t) = (cid:18) t α
α −t(cid:19) .
It satisfies h(−x) = −h(x) for all points x ∈ S2.
The elements of R1(M2(C)) are exactly the matrices with spectrum equal to {1, −1}. Thus,
for any matrix b in R1(M2(C)), the matrices (1 + b)/2 and (1 − b)/2 are rank-one projections,
orthogonal to each other. We obtain a pair of homeomorphisms x 7→ (1 + h(x))/2, and
x 7→ (1 − h(x))/2, from the space S2 to the space P1(M2(C)) of rank-one projections, with
the property that their values at every point x of the sphere S2 are orthogonal.
Lemma 3.1. Let T be a compact Hausdorff space. The space Hom(C(T ), M2(C)) is a
quotient of S2 × T × T :
S2 × T × T
q1
(S2 × T × T )/Z2
q2
Hom(C(T ), M2(C))
All three spaces are compact Hausdorff spaces. The space (S2 × T × T )/Z2 is the quotient of
S2 × T × T by the involution (x, t1, t2) 7→ (−x, t2, t1), and q1 is the corresponding quotient
map. The quotient map q2 takes each orbit {(x, t1, t2), (−x, t2, t1)} to the homomorphism
(∗)
evalt1( · )
I2 + h(x)
2
+ evalt2( · )
I2 − h(x)
2
.
The quotient map q2 identifies distinct orbits {(x, t1, t2), (−x, t2, t1)} and {(x′, t′
if and only if t1 = t2 = t′
1 = t′
2.
1, t′
2), (−x′, t′
2, t′
1)}
Proof. The first space S2 × T × T is a compact Hausdorff space because it is a product of
compact Hausdorff spaces. The second space (S2 × T × T )/Z2 is a compact Hausdorff space
because it is the quotient of a compact Hausdorff space by the action of a finite group. The
third space Hom(C(T ), M2(C)) is a compact Hausdorff space by lemma 1.7.
The expression (∗) clearly names a morphism in Steenrod's convenient category, i. e., a
continuous function from S2 × T × T to C(C(T, C), M2(C)). It is immediately apparent from
the expression that this continuous function is invariant under the action of Z2. Therefore,
it factors through the quotient map q1, via a continuous function q2.
For all points x in S2, the matrices (1 + h(x))/2 and (1 − h(x))/2 are rank-one projections
that sum to the identity matrix. Furthermore, for all points t1 and t2 in T the evaluation
functions evalt1 and evalt2 are unital ∗-homomorphisms. It follows that for all points (x, t1, t2)
in the product space S2 × T × T , the expression (∗) names a unital ∗-homomorphism.
Therefore, the range of q2 is a subset of Hom(C(T ), M2(C)).
We collect a couple of basic facts. First, the range of any element ρ in Hom(C(T ), M2(C))
is a commutative unital C*-subalgebra of M2(C). In particular, it is isomorphic to either C
or C2. Second, if ρ = (q2 ◦ q1)(x, t1, t2) for some (x, t1, t2) ∈ S2 ×T ×T , then the dimension of
ρ(C(T )) depends on whether the points t1 and t2 are equal or distinct. If they are equal, then
ρ = evalt1(·)I2, so ρ(C(T )) ∼= C. If t1 and t2 are distinct, then the distinct characters evalt1
and evalt2 both factor through ρ : C(T ) → ρ(C(T )), so ρ(C(T )) cannot be one-dimensional,
and thus, ρ(C(T )) ∼= C2.
12
1, t′
1 = t′
1( · )I2, so t1 = t′
Let ρ be any element of Hom(C(T ), M2(C)). First, assume that ρ(C(T )) ∼= C. It follows
that ρ is the composition of a unital ∗-homomorphism γ : C(T ) → C with the canonical
inclusion C ֒→ M2(C). By Gelfand duality, any such homomorphism γ is equal to evalt for
some point t. Thus, ρ = evalt( · )I2 = q2(q1((1, 0), t, t)).
Now assume that ρ(C(T )) ∼= C2. It follows that we can write ρ as γ1( · )b1 + γ2( · )b2, for
pairwise orthogonal projections b1 and b2 in ρ(C(T )), and unital ∗-homomorphisms γ1 and
γ2 from C(T ) to C. By Gelfand duality, there exist points t1 and t2 such that γ1 = evalt1
and γ2 = evalt2. There also exists a point x in the sphere S2 such that b1 = (1 + h(x))/2, and
consequently, such that b2 = (1 − h(x))/2. Thus, ρ = evalt1( · )((1 + h(x))/2) + evalt2( · )((1 −
h(x))/2) = q2(q1(x, t1, t2)). Therefore, q2 is surjective. As any continuous surjective function
between compact Hausdorff spaces it must be a quotient map.
2), (−x′, t′
2, then q(w) = evalt1( · )I2 = evalt′
1)} be orbits in (S2 × T ×
2, t′
1( · )I2 = q2(w′). Conversely,
T )/Z2. If t1 = t2 = t′
assume that q2(w) = q2(w′), and write ρ = q2(w) = q2(w′). The range ρ(C(T )) is isomorphic
to either C or C2. If it is isomorphic to C, then t1 = t2, t′
2, and evalt1( · )I2 = q2(w) =
q2(w′) = evalt′
Let w = {(x, t1, t2), (−x, t2, t1)} and w′ = {(x′, t′
1. Thus, t1 = t2 = t′
1 = t′
2.
1 = t′
Therefore, assume that the range ρ(C(T )) is isomorphic to C2. The two minimal projec-
tions in q2(w)(C(T )) are (1 + h(x))/2 or (1 − h(x))/2, and similarly the two minimal projec-
tions in q2(w′)(C(T )) are (1 + h(x′))/2 or (1 − h(x′))/2. Since the C*-algebras q2(w)(C(T ))
and q2(w′)(C(T )) are equal, the projection (1 + h(x))/2 must be equal to either (1 + h(x′))/2
or to (1 − h(x′))/2. In other words, h(x) must be equal to either h(x′) or h(−x′), so w = w′.
Therefore, q2(w) = q2(w′) if and only if w = w′ or t1 = t2 = t′
(cid:3)
1 = t′
2.
If ψ is a path in S2, and φ1 and φ2 are paths in T , we write [ψ, φ1, φ2] for the homotopy
class of the path (ψ, φ1, φ2) defined by τ 7→ (ψ(τ ), φ1(τ ), φ2(τ )).
Theorem 3.2. Let T be a compact Hausdorff space with distinguished point t0. Write q for
the composition of the quotient maps q1 and q2 in lemma 3.1, and write
q∗ : π1(S2 × T × T, ((1, 0), t0, t0)) → π1(Hom(C(T ), M2(C)), evalt0)
for the induced homomorphism of fundamental groups. For every path ϕ : [0, 1] → T , begin-
ning and ending at t0, we have that q∗[cnst(1,0), ϕ, cnstt0] = q∗[cnst(1,0), cnstt0, ϕ].
Proof. Let σ be the path in S2, beginning at (1, 0) and ending at (−1, 0), defined by σ(τ ) =
(exp(πiτ ), 0). The path q ◦ (σ, cnstt0, cnstt0) is the constant path at evalt0, by lemma 3.1. It
follows that q∗[σ, cnstt0, cnstt0] is the identity element of π1(Hom(C(T ), M2(C)), evalt0).
Let ϕ : [0, 1] → T be any path in T beginning and ending at t0. Appealing to lemma 3.1
in the first step, we calculate that
q∗[cnst(1,0), ϕ, cnstt0] = q∗[cnst(−1,0), cnstt0, ϕ]
= q∗[cnst(−1,0), cnstt0, ϕ] · q∗[σ, cnstt0, cnstt0]
= q∗([cnst(−1,0), cnstt0, ϕ] · [σ, cnstt0, cnstt0])
= q∗[σ, cnstt0, ϕ]
= q∗([σ, cnstt0, cnstt0] · [cnst(1,0), cnstt0, ϕ])
= q∗[σ, cnstt0, cnstt0] · q∗[cnst(1,0), cnstt0, ϕ]
= q∗[cnst(1,0), cnstt0, ϕ]
13
The embedding ı : T Hom(C(T ), M2(C)) clearly factors through q:
(cid:3)
S2 × T × T
cnst(1,0)×∆
q
T
ı
Hom(C(T ), M2(C))
The map diagonal map ∆ : T → T × T is defined by ∆(t) = (t, t).
Corollary 3.3. Every loop in RP2 is quantum nullhomotopic.
Proof. Let t0 be any point of T = RP2, and let ϕ : [0, 1] → T be any path that begins and
ends at t0.
ı∗[ϕ] = q∗((cnst(1,0) × ∆)∗[ϕ]) = q∗[cnst(1,0), ϕ, ϕ]
= q∗([cnst(1,0), ϕ, cnstt0] · [cnst(1,0), cnstt0, ϕ])
= q∗[cnst(1,0), ϕ, cnstt0] · q∗[cnst(1,0), cnstt0, ϕ]
= q∗[cnst(1,0), ϕ, cnstt0] · q∗[cnst(1,0), ϕ, cnstt0]
= q∗([cnst(1,0), ϕ, cnstt0] · [cnst(1,0), ϕ, cnstt0])
(cid:3)
Thus, ı∗[φ] is the identity element of π1(Hom(C(T ), M2(C)), evalt0). Therefore, for every
loop f : S1 → T , the loop ι ◦ f is nullhomotopic. By corollary 2.2, we conclude that every
loop in T = RP2 is quantum nullhomotopic.
Proposition 3.4. Let T be a compact Hausdorff space with distinguished point t0. The group
ı∗(π1(T, t0)) is commutative.
Proof. Let ı∗[ϕ1] and ı∗[ϕ2] be elements of ı∗(π1(T, t0)). Appealing to theorem 3.2, we cal-
culate that
q∗[cnst(1,0), ϕ1, cnstt0] · q∗[cnst(1,0), ϕ2, cnstt0] = q∗[cnst(1,0), ϕ1, cnstt0] · q∗[cnst(1,0), cnstt0, ϕ2]
= q∗[cnst(1,0), ϕ1, ϕ2]
= q∗[cnst(1,0), cnstt0, ϕ2] · q∗[cnst(1,0), ϕ1, cnstt0]
= q∗[cnst(1,0), ϕ2, cnstt0] · q∗[cnst(1,0), ϕ1, cnstt0]
Similarly q∗[cnst(1,0), cnstt0, ϕ1] and q∗[cnst(1,0), cnstt0, ϕ2] commute. Since we can write
ı∗[ϕ1] = q∗[cnst(1,0), cnstt0, ϕ1] · q∗[cnst(1,0), ϕ1, cnstt0],
and likewise for ı∗[ϕ2], it follows that ı∗[ϕ1] and ı∗[ϕ2] commute. Therefore, the group
ı∗(π1(T, t0)) is commutative.
(cid:3)
Corollary 3.5. There is a loop in (S1, 1) ∨ (S1, 1) that is quantum nullhomotopic, but not
nullhomotopic in the ordinary sense.
Proof. The fundamental group of T = S1 ∨S1 is the free group on two generators. Therefore,
the commutator of its two generators yields a loop f in S1 ∨ S1 that is not nullhomotopic.
However, the image of this loop f in Hom(C(T ), M2(C)) must be nullhomotopic by propo-
sition 3.4. It follows that f is quantum nullhomotopic by corollary 2.2.
(cid:3)
14
References
[1] S. Abramsky, R. S. Barbosa, N. de Silva, and O. Zapata, The Quantum Monad on Relational Structures,
Proc. MFCS 2017 (2017).
[2] P. J. Cameron, A. Montanaro, M. W. Newman, S. Severini, and A. Winter, On the quantum chromatic
number of a graph, Electron. J. Combin. 14 (2007), no. 1.
[3] K. R. Davidson, C*-algebras by Example, Fields Institute Monographs 6 (1996).
[4] K. De Commer, P. Kasprzak, A. Skalski, and P. So ltan, Quantum actions on discrete quantum spaces
and a generalization of Clifford's theory of representations (2016), available at arXiv:1611.10341.
[5] E. Effros and Z.-J. Ruan, Discrete Quantum Groups I. The Haar Measure, Int. J. Math 5 (1994).
[6] J. M. Garcia-Bondia, J. C. Varilly, and H. Figueroa, Elements of Noncommutative Geometry, Birkauser,
2000.
[7] A. Kornell, Quantum sets, conditionally accepted for publication in J. Math. Phys.
[8] N. H. Kuiper, The homotopy type of the unitary group of Hilbert space, Topology 3 (1965).
[9] L. Mancinska and D. E. Roberson, Quantum homomorphisms, J. Combin. Theory, Series B 118 (2016).
[10] G. J. Murphy, C*-algebras and Operator Theory, Academic Press, 1990.
[11] B. Musto, D. J. Reutter, and D. Verdon, A compositional approach to quantum functions, to appear in
J. Math. Phys. (2017), available at arXiv:1711.07945.
[12] G. K. Pedersen, Analysis NOW, Springer Science+Business Media, 1995.
[13] P. Podle´s and S. L. Woronowicz, Quantum deformation of Lorentz Group, Comm. Math. Phys. 130
(1990), no. 2.
[14] N. E. Steenrod, A convenient category of topological spaces, Michigan Math J. 14 (1967), no. 2.
[15] A. Van Daele, Discrete quantum groups, J. Algebra 180 (1996).
15
|
1508.07904 | 3 | 1508 | 2017-07-19T20:08:35 | Part II, Free Actions of Compact Groups on C*-Algebras | [
"math.OA",
"math.QA"
] | We study a simple subclass of free actions of non-Abelian groups on unital C*-algebras, namely cleft actions. These are characterized by the fact that the associated noncommutative vector bundles are trivial. In particular, we provide a complete classification theory for these actions and describe its relations to classical principal bundles. | math.OA | math |
Part II, Free Actions of Compact
Groups on C∗-Algebras
Kay Schwieger ∗
Stefan Wagner †
Abstract
We study a simple class of free actions of non-Abelian groups on unital C∗-alge-
bras, namely cleft actions. These are characterized by the fact that the associated
noncommutative vector bundles are trivial. In particular, we provide a complete
classification theory for these actions and describe its relations to classical principal
bundles.
Keyword: Weakly cleft action, C∗-algebra, factor system, cocycle action
MSC2010: 46L85, 37B05 (primary), 55R10, 16D70 (secondary).
1 Introduction
In this presentation we investigate a special class of group actions on unital C∗-algebras.
The experience with group actions on topological spaces shows that free actions are
easier to understand and to classify.
In this context a free action of a group G on
a space P is typically regarded as a topological principal G-bundle over the quotient
P/G. For instance, locally trivial bundles are, up to equivalence, characterized by the
Čech cohomology H 1(X, G).
It is therefore reasonable to expect that free actions on
noncommutative spaces are easier to understand and classify, too.
In the first part of this series [18] we investigated free actions of compact Abelian groups
on unital C∗-algebras. This classification relies on the fact that the corresponding iso-
typic components are Morita self-equivalence over the fixed point algebra. For non-
Abelian compact groups the bimodule structure is more subtle. For this reason the
current article concentrates on a simple class of free actions of non-Abelian groups,
∗University of Helsinki, [email protected]
†Blekinge Tekniska Högskola, [email protected]
1
namely cleft actions. Regarded as noncommutative principal bundles, these actions are
characterized by the fact that all associated noncommutative vector bundles are trivial.
In particular, for such actions all Chern classes (cf. [2]) vanish. Although this prop-
erty looks very limiting, in fact, many noncommutative phenomena already show up
here. Therefore, cleft actions may be used as toy models of noncommutative principal
bundles, e. g. for Chern-Simmons actions (see [23, 12]).
The earliest classification result for free actions of compact, but not necessarily Abelian
groups goes back to Wassermann [21]. He showed that free and ergodic actions, i. e.,
actions with full multiplicity and trivial fixed point space, are, up to equivalence, char-
acterized by unitary 2-cocycles on the dual of the group. For finite groups Davydov [3]
presented an alternative classification using classical group cohomology. Although free
and ergodic actions only correspond to principal bundles over the singleton base space,
these result yet show that noncommutative geometry admits more interesting examples
than the classical theory.
The prototypes of free and ergodic actions are the quantum 2-tori Aθ, θ ∈ T, equipped
with the gauge action of T2. Varying θ ∈ T, these are in fact all non-equivalent free
ergodic actions of the 2-torus T2. Non-ergodic examples can then be obtained by taking
continuous bundles of ergodic actions. A prominent example of this type is the gauge
action of the 2-torus T2 on the Heisenberg group C∗-algebra A (see [7]). The algebra A
can be written as C∗ -- bundle over T, where for each θ ∈ T the fiber is the noncommutative
2-tori Aθ. The action of T2 on the algebra A is the fiberwise gauge action. Echterhoff,
Nest, and Oyono-Oyono [7] proposed such bundles as noncommutative principal torus
bundles. Concerning non-Abelian groups, the classification results of Wassermann can
easily be extended to bundles of free ergodic actions. Such bundle actions are up to
equivalence determined by the continuous family of unitary 2-cocycles corresponding to
the ergodic actions in each fiber.
Beyond the concept of bundles of ergodic actions there are multiple directions to in-
troduce further noncommutativity. Any C∗-algebra obtained by forming a bundle of
ergodic actions over a compact base space X always contains C(X) in its center.
In
order to explore noncommutative principal bundles over a noncommutative base space,
this requirement must be abolished. Without this restriction new examples are immedi-
ately available even for a classical base space. For instance, given a coaction γ of G on
X, the crossed product C(X) ⋊γ G with the dual action of G is cleft (cf. Example 4.9)
but C(X) is central only if the coaction is inner, i. e., if it is trivial up to a 1-cocycle (cf.
Example 5.15).
A second direction to explore is to not only consider actions of classical groups but of
quantum groups. An algebraic approach to cleft actions of quantum groups (alias Hopf
algebras) is already established in the theory of Hopf-Galois extensions (see e. g. [6, 17]).
There cleft actions are free actions with a convolution invertible cleaving map. Doi [5]
has shown that cleft actions can be written as a twisted crossed product and provided
2
a classification for these crossed products. Also the work of Vaes and Vainerman [20]
should be mentioned, who studied cleft actions of locally compact Hopf von Neumann
algebras. An essential part of our presentation will lift the algebraic constructions to
the C∗-algebraic framework. In this article we restricts ourselves to classical compact
groups for sake of a simple presentation and because most essential problems are already
present in this restricted context.
The article is structured as follows. After this introduction and some preliminaries,
we recall the basic decomposition of a C∗-dynamical system into (generalized) isotypic
components in Section 3. In Section 4 we introduce cleft C∗-dynamical systems and a
weaker notion, called weakly cleft, and discuss their relations. Section 5 presents the
characterization of weakly cleft systems in terms of factor systems, alias cocycle actions.
Moreover, we discuss some relations between the type of the dynamical system and
the form of its factor system. In particular, we classify all cleft topological (classical)
principal bundles. Finally, in Section 6 we show that factor systems and weakly cleft
C∗-dynamical systems are, up to equivalence, indeed in 1-1-correspondence by explicitly
constructing the dynamical system.
2 Preliminaries and Notations
Once and for the rest of the paper we fix a compact group G. All integrals over G are
taken with respect to the Haar probability measure. By a representation of G we always
mean a finite-dimensional unitary representation. For a representation (π, V ) we write
dπ for its dimension, for the dual representation we write (¯π, ¯V ). The set of equivalence
classes of irreducible representations will be denoted by G. All our constructions behave
naturally with respect to intertwiners and hence, for sake of brevity, we do not distinguish
between a representation of G and its equivalence class.
Let A be a unital C∗-algebra. For the unit of A we write 1A or simply 1. For an
element u ∈ A we denote by Ad[u] : A → A the map x 7→ uxu∗. All tensor products of
C∗-algebras are taken with respect to the minimal tensor product. We will frequently
deal with multiple tensor products of unital C∗-algebras A, B, and C.
If there is no
ambiguity, we regard A, B, and C as subalgebras of A ⊗ B ⊗ C and extend maps on
A, B, or C canonically by tensoring with the identity map. For sake of clarity we may
occasionally use the leg numbering notation, e. g., for x ∈ A ⊗ C we write x13 to denote
the corresponding element in A ⊗ B ⊗ C.
Our main focus in this paper will be on C∗-dynamical systems, by which we mean triples
(A, G, α) consisting of a unital C∗-algebra A together with a group of ∗-automorphisms
αg : A → A, g ∈ G, such that for each x ∈ A the map g 7→ αg(x) is continuous. We
typically write B = AG for the fixed point algebra of the dynamical system and we
3
denote by P0 the associated the conditional expectation P0(x) := RG αg(x) dg, x ∈ A.
More general, for an irreducible representation π ∈ G we denote by Pπ : A → A the
G-equivariant projection onto the isotypic component A(π) := Pπ(A), which is given
by
Pπ(x) := dπZG
Tr(π∗
g) αg(x) dg,
x ∈ A.
Two C∗-dynamical system (A, G, α) and (A′, G, α′) are called equivalent if there is an
isomorphism ϕ : A → A′ with ϕ ◦ αg = α′
g ◦ ϕ for all g ∈ G.
For the necessary background on modules of C∗-algebras we recommend [1], here we only
briefly recall some relevant definition. For a C∗-algebra B a right pre-Hilbert B-module is
a right B-module H equipped with a sesquilinear map h·, ·iB : H×H → B that satisfies the
usual axioms of a definite inner product with B-linearity in the second component.1 We
may define a norm on H by putting kxkH = khx, xiBk1/2. If H is complete with respect
to this norm then H is called a right Hilbert B-module. A linear operator T : H → H
on a right Hilbert B-module H is called adjointable if there is an operator T ⋆ : H → H
satisfying hT x, yiB = hx, T ⋆yiB for all x, y ∈ H. Adjointable operators are automatically
bounded but the converse does not hold. The set L(H) of all adjointable operators on a
right Hilbert B-module is a C∗-algebra. A correspondence over B, or a right Hilbert B-
bimodule, is a B-bimodule H equipped with a B-valued inner product h·, ·iB which turns
it into a right Hilbert B-module such that the left action of B on H is via adjointable
operators. For two correspondences H, K over the same algebra B we denote by H ⊗B K
their tensor product, which is again a correspondence over B. The elementary tensors
x ⊗ y (x ∈ H, y ∈ K) are total in H ⊗B K. The inner product on H ⊗B K is given by
hx1 ⊗ y1, x2 ⊗ y2iB = hy1, hx1, x2iB . y2iB
for all x1, x2 ∈ H1 and y1, y2 ∈ K.
3 Decomposition of C∗-Dynamical Systems
As a background for later discussions we first would like to recall the general decom-
position of C∗-dynamical systems (A, G, α).
In analogy to the GNS-construction the
conditional expectation P0 onto the fixed point space B := AG allows to equip A with
the definite B-valued inner product
hx, yiB := P0(x∗y) = ZG
αg(x∗y) dg
for x, y ∈ A. We write L2(A) for the right Hilbert B-module obtained by taking the
completion of A with respect to the corresponding norm. The C∗-algebra A admits a
1 In the literature the notion is usually relaxed even further to pre-C∗-algebras and non-definite inner
products. But we do not need this more general framework.
4
faithful representation on L2(A) given by
λ : A → L(cid:0)L2(A)(cid:1),
λ(x)y := x · y.
This allows to identify A with the subalgebra λ(A) ⊆ L(cid:0)L2(A)(cid:1) and we implicitly do
so unless confusion arise. For each g ∈ G we have a unitary operator on L2(A) given
by Ugx := αg(x) for x ∈ A ⊆ L2(A). The map g 7→ Ug is a strongly continuous
representation of G that implements the automorphisms αg, g ∈ G, in the sense that
λ(cid:0)αg(x)(cid:1) = Ug λ(x) U ⋆
g ,
x ∈ A.
As every representation of G, the algebra A can be decomposed into its isotypic compo-
nents with respect to the actions. More precisely, the sum of the isotypic components is
A(π) is a dense ∗-subalgebra of A. Moreover, the isotypic components
are mutually orthogonal, closed, linear subspaces of L2(A) with
direct and Palg
π∈ G
L2(A) = Mπ∈ G
A(π).
One aspect of this presentation will be the multiplication of A. It is worth noting that
the multiplication is determined by the action of A(π) ⊆ A on A(ρ) ⊆ L2(A) for all
pairs of irreducible representations π and ρ.
Instead of dealing with isotypic components, it will be more convenient to consider, for
a representation π, the generalized isotypic component
A2(π) := (cid:8)x ∈ A ⊗ L(V )(cid:12)(cid:12) πg · αg(x) = x ∀g ∈ G(cid:9).
Obviously, A2(π) is a B-bimodule for the usual left and right multiplication. In addition,
we may equip A2(π) with the B-valued inner product
hx, yiB := 1
dπ
(idA ⊗ Tr)(x∗y)
for x, y ∈ A2(π). Then the space A2(π) is a correspondence over B (see e. g. [4] for
completeness of the norm).
If π is irreducible, the map x 7→ (idA ⊗ Tr)(x) gives an
isomorphism of correspondences from A2(π) to the dual isotypic component A(¯π) with
inverse given by y 7→ dπRG αg(y) ⊗ πg dg. In the following we will frequently use this
identification.
The multiplication between isotypic components is well captured by family of maps
mπ,ρ : A2(π) ⊗B A2(ρ) −→ A2(π ⊗ ρ) ⊆ A ⊗ L(V ) ⊗ L(W )
mπ,ρ(x ⊗ y) := x12 · y13
for pairs of representations (π, V ) and (ρ, W ) of G. For an irreducible representation
σ ∈ G and a representation π of G let us denote by Pσ,π : L(Vπ) → L(Vσ) the map
5
Pσ,π(x) := Pk=1 v∗
kxvk with an orthonormal basis of intertwiners v1, . . . , vm : Vσ → Vπ.
The map Pσ,π does not depend on the choice of intertwiners. Then for irreducible
representations π, ρ ∈ G and elements x ∈ A2(π) and y ∈ A2(ρ) the operator λ(x) takes
the form
λ(x)y = Mσ∈ G
Pσ,π⊗ρ(cid:0)mπ,ρ(x ⊗ y)(cid:1).
(1)
Since the sum of the isotypic components is norm dense in A, the C∗-algebra λ(A) is
generated by the set of these operators λ(x) with x ∈ A2(π), π ∈ G. In particular, the
multiplication of A can be recovered from the maps mπ,ρ for π, ρ ∈ G in this way.
4 Cleft and Weakly Cleft C∗-Dynamical Systems
In principle, C∗-dynamical systems (A, G, α) with a given fixed point algebra B can be
classified in a functorial way in terms of the module structure of the generalized isotypic
components and their multiplicative relation (see [11]). In this presentation we will focus
on the class of cleft actions.
Definition 4.1. A C∗-dynamical system (A, G, α) is called cleft if for every irreducible
representation (π, V ) of G the set A2(π) ⊆ A ⊗ L(V ) contains a unitary element.
Cleft C∗-dynamical systems are precisely the so called semidual actions discussed in
[22]. For such a C∗-dynamical system (A, G, α) it follows along the same lines as in [22,
Thm. 10] that the crossed product A ⋊ G is isomorphic to AG ⊗ K, generalizing Green's
Theorem (cf. [8, Cor. 15] and [7]). In the algebraic theory of actions of Hopf algebras
these actions are up to equivalence given by twisted crossed products (cf. [6, 5], see also
[20]). Most arguments in our discussion rely on the following weaker hypothesis only
and establishing the results in a slightly wider framework has some technical advantages
later on.
Definition 4.2. A C∗-dynamical system (A, G, α) is called weakly cleft if for every
irreducible representation (π, V ) of G the set A2(π) ⊆ A ⊗ L(V ) contains an element s
such that
s∗s = 1
and
ss∗x = x
for all x ∈ A2(π). We call such an element s a non-degenerate isometry.
Unfortunately, we cannot present an examples of a weakly cleft but not cleft action.
In fact, in simple examples our results show that weakly cleft dynamical systems are
automatically cleft (see Lemma 4.6 and 5.9). We do not yet know whether this holds
in general. Most proofs of this article only rely on the weakly cleft assumption but can
6
be simplified for cleft systems. The advantage of dealing with weakly cleft actions is
that this property can be characterized by the right Hilbert module structure of the
(generalized) isotypic components. More precisely, the following lemma shows that the
C∗-system (A, G, α) is weakly cleft if and only if each isotypic component A(π), π ∈ G,
is a free right Hilbert B-module of rank d2
π. In the ergodic case, B = C1, this is the same
as saying that the Hilbert space A(π) has its maximal dimension d2
π, i. e., A(π) has full
multiplicity.
Lemma 4.3. For an element s ∈ A2(π) the following statements are equivalent:
(a) s is a non-degenerated isometry.
(b) The map ϕ : B ⊗ L(Vπ) → A2(π), ϕ(x) := sx is an isomorphism of right Hilbert
B-modules.
Proof. For one implication notice that for an isometry s the map ϕ(x) = sx is an isometry
for the right inner product. If s is non-degenerate, ϕ admits the inverse ϕ−1(y) = s∗y.
Together this proves the implication from (a) to (b). For the converse implication,
suppose that ϕ is an isomorphism. Then the selfadjoint element p := s∗s ∈ B ⊗ L(V ) is
a projection, since it satisfies
hp, xiB = hϕ(p), ϕ(x)iB = 1
dπ
(id ⊗ Tr)(s∗ss∗sx) = hp2, xiB
for all x ∈ B ⊗ L(V ). Injectivity of ϕ then implies that s is in fact an isometry because
ϕ(p) = ss∗s = s = ϕ(1). It follows that the inverse of ϕ is given by ϕ−1(x) = s∗x and
hence we have x = ϕ(cid:0)ϕ−1(x)(cid:1) = ss∗x for all x ∈ A2(π).
To distinguish cleft and weakly cleft, let us introduce a third property, which is of great
independent interest. Let (A, G, α) be a C∗-dynamical system. For a representation
(π, V ) of G we write A2(π)A2(π)∗ for the linear subspace generated by products xy∗ of
elements x, y ∈ A2(π), and we put
C(π) := {x ∈ A ⊗ L(V ) (cid:0)αg ⊗ Ad[πg](cid:1)(x) = x ∀g ∈ G}.
Definition 4.4. A C∗-dynamical system (A, G, α) is called free if for all representations
π of G we have A2(π)A2(π)∗ = C(π).
Free actions have attained special interest in the literature, see, e. g., [13, 16, 15, 4].
For the definition given here and the relation to noncommutative principal bundles, we
refer to [18, Section 3]. It should be noted that A2(π)A2(π)∗ is always a ∗-ideal in C(π).
Therefore, the dynamical system is free if and only if the closure of A2(π)A2(π)∗ contains
the unit 1A⊗L(V ).
7
Lemma 4.5. A C∗-dynamical system is cleft if and only if it is weakly cleft and free.
Proof. If (A, G, α) is cleft, it is obviously weakly cleft. Since for every representation
(π, V ) of G we find a unitary u ∈ A2(π), the set A2(π)A2(π)∗ contains uu∗ = 1A⊗L(V )
and hence the dynamical system is free. To show the converse, suppose that (A, G, α)
is weakly cleft and free and let (π, V ) be a representation of G. Then we find a non-
degenerated isometry s ∈ A2(π) and obtain
A2(π)A2(π)∗ = (cid:0)s · B ⊗ L(V )(cid:1) ·(cid:0)s ⊗ B ⊗ L(V )(cid:1)∗ = s · B ⊗ L(V ) · s∗.
This set is closed and, since the dynamical system is free, it contains the unit. It follows
that s must have full range, i. e., s is a unitary element and hence the system is cleft.
Lemma 4.6. For a compact Abelian group G, every weakly cleft C∗-dynamical system
is cleft.
Proof. For an Abelian groups generalized isotypic components and the isotypic compo-
nent of the dual representation literally coincide. If the C∗-dynamical system is weakly
cleft, then for each π ∈ G we find an isometry s ∈ A(¯π). Hence for the dual representa-
tion ¯π the set A(π)A(π)∗ = A(¯π)∗A(¯π) contain the element s∗s = 1.
Example 4.7. We start with the most basic example as a prototype. Let B be a
unital C∗-algebra and denote by C(G) the C∗-algebra of continuous functions on G. We
consider the C∗-dynamical system(cid:0)B⊗C(G), G, id ⊗r(cid:1) where the action on C(G) is given
by the right translation (rgf )(h) := f (hg) for f ∈ C(G) and h ∈ G. Clearly, the fixed
point space of this system is B = B ⊗ 1G. We may identify elements in a tensor product
with C(G) with functions on G in the usual way. Then, for an irreducible representation
(π, V ) of G, the correspondence A2(π) over B is given by
A2(π) = {f : G → B ⊗ L(V ) continuous πgf (g) = f (e) ∀g ∈ G}.
The function u(g) := π∗
the system is cleft.
g then obviously is a unitary element in A2(π), which shows that
If B is commutative, i. e., B = C(X) for some compact Hausdorff space X, then the
dynamical system of Example 4.7 can be be understood as a trivial principal G-bundle
over the space X. The next example shows that also non-trivial principal bundles
may give rise to cleft actions. Moreover, in the later Corollary 5.10 we will provide
a characterization of cleft topological principal bundles.
Example 4.8. Fix n ∈ N and denote by Cn := {ζ ∈ C ζ n = 1} the group of n-th roots
of unity. We consider the C∗-algebra A := C(T) of continuous function on the circle
with the action of Cn given by rotations, i. e., for ζ ∈ Cn and f ∈ C(T) put
(αζf )(z) := f (ζ · z),
z ∈ T.
8
For an irreducible representation of Cn, i. e., an element k ∈ Z/nZ, we have
A2(−k) = A(k) = {f ∈ C(T) f (ζ · z) = ζ k · f (z) ∀z ∈ T}.
The action is cleft, because an invertible element in A2(−k) is, for instance, given by the
function f (z) := zk, z ∈ T. This dynamical system can be understood as the non-trivial
principal Cn-bundle corresponding to the n-fold covering p : T → T, p(z) := zn. In fact,
since all vector bundles over T are trivial (cf. [19, Section 18]), every principal bundles
over T with arbitrary compact structure group G gives rise to a cleft C∗-dynamical
system.
Example 4.9. Consider a compact group G and a closed normal subgroup N and
suppose that the action of N on C(G) by right translations is cleft (e. g. suppose that
G/N is finite). Furthermore, let δ : A → M(cid:0)A ⊗ C ∗(G)) be a coaction of G on a
unital C∗-algebra A. We recall that a twist over G/N is a unitary corepresentation
W ∈ M(A ⊗ C ∗(G/N)) of G/N such that (id ⊗q) ◦ δ = Ad[W ] and (δ ⊗ id)(W ) = W13,
where q denotes the natural ∗-homomorphism q : C ∗(G) → C ∗(G/N). Such a twist
gives rise to an ideal IW of the crossed product A ⋊δ G, called twisting ideal, which
is invariant under the canonical action of N on A ⋊δ G (see [14]). The twisted crossed
product A ⋊δ,W G := (A ⋊δ G)/IW then carries an action of N, which we denote by α,
that is, we obtain a C∗-dynamical system
(A ⋊δ,W G, N, α).
Since the natural embedding of C(G) into A ⋊δ G is G-equivariant, the factorized ho-
momorphism kG : C(G) → A ⋊δ,W G is N-equivariant. Finally, a few moments thought
shows that being cleft is preserved under equivariant ∗-homomorphisms and, therefore,
it follows that the above C∗-dynamical system is cleft.
5 Factor Systems
Let (A, G, α) be a C∗-dynamical system with fixed point algebra B. Suppose for the
moment that the system is weakly cleft, that is, for each irreducible representation
π ∈ G we find a non-degenerate isometry sπ ∈ A2(π) ⊆ A ⊗ L(Vπ). For the trivial
representation, denoted by 1, we pick s1 := 1B. We may extend this family of isome-
tries to non-irreducible representations by decomposing each representation (π, V ) into
a direct sum of irreducible representations (σk, Vk) with intertwiners vk : Vk → V for
k. It is easily checked that this provides a
non-degenerate isometry in A2(π) and that the construction does not depend on the
choice of intertwiners.
each 1 ≤ k ≤ m and put sπ := Pm
k=1 vksσkv∗
By Lemma 4.3, for each π ∈ G the space A2(π) is isomorphic to B ⊗ L(Vπ) as a right
Hilbert B-module, but in general not as a left B-module. In order to describe the left
9
action of B consider the map
γπ : B → B ⊗ L(Vπ),
γπ(b) := s∗
π(b ⊗ 1π)sπ.
(2)
The correspondence A2(π) is then isomorphic to the vector space B ⊗ L(Vπ) equipped
with the usual right multiplication by B, the usual right B-valued inner product, and
the left multiplication given by
b . x := γπ(b) x,
for all x ∈ B ⊗ L(Vπ) and b ∈ B.
Lemma 5.1. The map γπ is a unital ∗-homomorphism.
Proof. Obviously, γπ(b∗) = γπ(b)∗ for all b ∈ B and γπ is unital, since sπ is an isometry.
For each b ∈ B the element (b ⊗ 1)sπ lies in A2(π). The non-degeneracy of sπ then
implies for all b1, b2 ∈ B that
γπ(b1)γπ(b2) = s∗
π(b1 ⊗ 1)(b2 ⊗ 1)sπ = γπ(b1b2).
π(b1 ⊗ 1)sπs∗
π(b2 ⊗ 1)sπ = s∗
Remark 5.2. We would like to point out that, since sπ is non-degenerated, the map
γπ : B → B ⊗ L(Vπ) is uniquely determined by the property (b ⊗ 1)sπ = sπγπ(b) for all
b ∈ B.
In the following we will frequently deal with tensor products in which precisely one factor
is B. In this cases we will allow more flexibility for the position of the factor B, that is,
we reshuffle the tensor factor is such a way that B is at a convenient position, usually
the first factor, and the other factors are kept in order.
Coming back to the dynamical system, the multiplicative structure among the sets A2(π)
for different representations can be phrased in terms of the elements sπ, too. For two
irreducible representations π, ρ ∈ G consider the multiplication map
mπ,ρ : A2(π) ⊗B A2(ρ) −→ A2(π ⊗ ρ) ⊆ A ⊗ L(Vπ) ⊗ L(Vρ),
mπ,ρ(x ⊗ y) := x12 y13.
(3)
This is a module map for the right action of B ⊗ L(Vπ) ⊗ L(Vρ) on domain and codomain.
Therefore, it is uniquely determined by the element mπ,ρ(sπ, sρ). For this element there
is a unique element ω(π, ρ) ∈ B ⊗ L(Vπ ⊗ Vρ) with mπ,ρ(sπ, sρ) = sπ⊗ρ · ω(π, ρ). In fact,
ω(π, ρ) is the isometry given by
With this element the multiplication map can be written as
ω(π, ρ) = s∗
π⊗ρ (sπ)12 (sρ)13.
mπ,ρ(sπx ⊗ sρy) = sπ⊗ρ · ω(π, ρ) · (idπ ⊗γρ)(x) · (1π ⊗ y)
(4)
(5)
10
for all x ∈ B ⊗ L(Vπ), y ∈ B ⊗ L(Vρ).
We want to classify and characterize weakly cleft C∗-dynamical systems in terms of the
∗-homomorphisms γπ and the isometries ω(π, ρ). For this purpose, we fix a group G and
a C∗-algebra B and we consider pairs (γ, ω) consisting of two families γ = (γπ)π∈ G and
ω = (cid:0)ω(π, ρ)(cid:1)π,ρ∈ G where
1. for each π ∈ G, we have a unital ∗-homomorphism γπ : B → B ⊗ L(Vπ) and
2. for each π, ρ ∈ G, we have an isometry ω(π, ρ) ∈ B ⊗ L(Vπ ⊗ Vρ).
Definition 5.3.
1. The pair (γ, ω) is called a factor system for (G, B) if ω(1, 1) = 1B
and the family satisfies
ω(π, ρ) · (idπ ⊗γρ)(cid:0)γπ(b)(cid:1) = γπ⊗ρ(b) · ω(π, ρ),
(cid:0)1π ⊗ ω(ρ, σ)(cid:1) · (idπ⊗ρ ⊗γσ)(cid:0)ω(π, ρ)(cid:1)∗ = ω(π, ρ ⊗ σ)∗ · ω(π ⊗ ρ, σ)
for all π, ρ, σ ∈ G and b ∈ B.
(6)
(7)
2. Two factor systems (γ, ω) and (γ′, ω′) for (G, B) are called conjugated if there is a
family v = (vπ)π∈ G of unitaries vπ ∈ B ⊗ L(Vπ) such that
γ′
π = Ad[v∗
π] ◦ γπ,
vπ⊗ρ · ω′(π, ρ) = ω(π, ρ) · (idπ ⊗γρ)(vπ) · vρ.
Equations (6) and (7) are twisted versions of the equations for a coaction and for a
2-cocycle. For this reason, we refer to condition (6) as the coaction condition and
to condition (7) as the cocycle condition.
Immediate examples of factor systems are
accordingly given a by coaction with a trivial cocycle or by a 2-cocycle with a trivial
coaction (see Remark 5.4).
Remark 5.4.
1. The normalization condition ω(1, 1) = 1B can always be achieved
by passing to a normalized, conjugated system (in the straightforwardly generalized
sense). Together with the coaction and cocycle condition the normalization implies
γ1 = idB and ω(π, 1) = 1 = ω(1, π) for all π ∈ G.
2. It is worth mention that we may rephrase things in terms of the group C∗-algebra
C ∗(G), more precisely its multiplier algebra M C ∗(G). The family (γπ)π∈ G may
be equivalently written as a single unital ∗-homomorphism γ : B → B ⊗ M C ∗(G)
and the family (ω(π, ρ))π,ρ∈ G as an isometry ω ∈ B ⊗ M C ∗(G × G). Then Equa-
tions (6) and (7) can be casted in the form
ω ·(cid:0)(id ⊗γ) ◦ γ(cid:1)(b) = (δ ◦ γ)(b) · ω,
(1 ⊗ ω) · (id ⊗γ)(ω∗) = (id ⊗δ)(ω∗) · (δ ⊗ id)(ω),
11
where δ : M C ∗(G) → M C ∗(G × G) denotes the usual comultiplication. This
formulation is used for extending the theory to compact quantum groups. If ω is
unitary, the pair (γ, ω) is sometimes called a cocycle action (see e. g. [20]).
The following statement summarizes that the construction from the beginning of this
section indeed provides examples of factor systems.
Lemma 5.5. Let (A, G, α) be a weakly cleft C∗-dynamical system and let B denote
its fixed point algebra. Furthermore, let s = (sπ)π∈ G be a family of non-degenerate
isometries sπ ∈ A2(π) with s1 = 1B. If γ = (γπ)π∈ G and ω = (cid:0)ω(π, ρ)(cid:1)π,ρ∈ G are the
associated families of homomorphisms and unitaries given by Equations (2) and (4),
respectively, then the pair (γ, ω) is a factor system.
Proof. In order to show the coaction condition (6), we recall that we have sπγπ(b) =
(b ⊗ 1)sπ for every b ∈ B (see Remark 5.2). Successively applying this relation and its
starred version then yields Equation (6), i. e., in B ⊗ L(Vπ) ⊗ L(Vρ) we obtain
ω(π, ρ) · γρ(cid:0)γπ(b)(cid:1) = s∗
π⊗ρsπsργρ(cid:0)γπ(b)(cid:1) = s∗
= γπ⊗ρ(b)s∗
π⊗ρsπsρ = γπ⊗ρ(b) · ω(π, ρ)
π⊗ρbsπsρ
for all π, ρ ∈ G and b ∈ B. In order to verify the cocycle condition (7) let us first consider
its left hand side
ω(ρ, σ) · γσ(cid:0)ω(π, ρ)(cid:1)∗ = s∗
ρ⊗σsρsσ · s∗
σ(s∗
π⊗ρsπsρ)∗sσ = s∗
ρ⊗σsρsσs∗
σs∗
ρs∗
πsπ⊗ρsσ
The rightmost product s∗
degeneracy of sσ allows us to cancel the factor sσs∗
in L(Vπ) ⊗ A2(ρ), which allows us to cancel even further to obtain
πsπ⊗ρsσ lies in L(Vπ ⊗ Vρ) ⊗ A2(σ). Therefore, the non-
πsπ⊗ρ lies
σ. Similarly, the element s∗
ρs∗
ω(ρ, σ) · γσ(cid:0)ω(π, ρ)(cid:1)∗ = s∗
ρ⊗σs∗
πsπ⊗ρsσ.
For the right hand side of the cocycle condition, the non-degeneracy of sπ⊗ρ⊗σ likewise
implies
ω(π, ρ ⊗ σ)∗ · ω(π ⊗ ρ, σ) = s∗
ρ⊗σs∗
πsπ⊗ρ⊗σs∗
π⊗ρ⊗σsπ⊗ρsσ = s∗
ρ⊗σs∗
πsπ⊗ρsσ.
Comparing with the simplification of the left side then yields the cocycle condition for
all π, ρ, σ ∈ G.
The next result states that the weakly cleft C∗-dynamical systems are uniquely deter-
mined by their factor systems up to equivalence.
12
Theorem 5.6. Let (A, G, α) and (A′, G, α′) be weakly cleft C∗-dynamical systems with
the same fixed point algebra B and let (γ, ω) and (γ′, ω′) be associated factor systems,
respectively. Then the following statements are equivalent:
(a) The dynamical systems (A, G, α) and (A′, G, α′) are equivalent.
(b) The factor systems (γ, ω) and (γ′, ω′) are conjugated.
Proof. As a distinction we add a prime to all notions referring to (A′, G, α′).
1. To prove that (a) implies (b) it suffice to show that for the same dynamical sys-
tem (A, G, α) different choices of non-degenerate isometries sπ ∈ A2(π), π ∈ G,
π, π ∈ G, be
lead to conjugated factor systems. For this purpose let sπ and s′
two such choices and let us denote by (γ, ω) and (γ′, ω′) the associated factor sys-
tems, respectively. Consider first a fixed representation π ∈ G. By Lemma 4.3
there are unique elements vπ, v′
πv′
π.
Uniqueness implies vπv′
πvπ and, since sπ is a isometry, we also have
π = v∗
πsπvπ = (s′
πvπ = v∗
v∗
π.
For the ∗-homomorphisms of the factor systems we therefore obtain
π) = 1. Hence vπ and v′
π = 1 = v′
π)∗(s′
π ∈ B ⊗ L(Vπ) with s′
π = sπvπ and sπ = s′
π are unitaries with v′
πs∗
γ′
π(b) = (s′
π)∗(b ⊗ 1)s′
π = v∗
πs∗
π(b ⊗ 1)sπvπ = v∗
πγπ(b)vπ,
for every π ∈ G and b ∈ B. For the isometries of the factor systems we may use
the non-degeneracy of sρ and the fact that vπsρvρ ∈ L(Vπ) ⊗ A2(ρ) to conclude for
all π, ρ ∈ G:
vπ⊗ρ ω′(π, ρ) = vπ⊗ρ(sπ⊗ρvπ⊗ρ)∗sπvπsρvρ = s∗
π⊗ρsπ(sρs∗
ρ)vπsρvρ
= s∗
π⊗ρsπsργρ(vπ)vρ = ω(π, ρ) γρ(vπ) vρ.
2. For the converse implication, (b) ⇒ (a), let sπ ∈ A2(π) and s′
2(π), π ∈ G,
by non-degenerate isometries with associated factor systems (γ, ω) and (γ′, ω′),
respectively. Furthermore, let vπ, π ∈ G, be a family of unitaries realizing the
conjugation of the factor systems as in Definition 5.3. For every representation π
of G the map
π ∈ A′
ϕπ : A′
2(π) → A2(π),
s′
πx 7→ sπvπx
for all x ∈ B ⊗ L(Vπ) is well-defined by Lemma 4.3. Moreover, it is straightforward
to check that ϕπ is a unitary map between the two right Hilbert B-modules. Since
L2(A) = Lπ∈ GA2(π) and likewise for A′, taking direct sums yields a unitary map
Furthermore, the maps ϕπ, π ∈ G, intertwine with the multiplication maps, that
is, for all π, ρ ∈ G we have
V : L2(A′) → L2(A), V := Mπ∈ G
ϕπ.
mπ,ρ(cid:0)ϕπ(sπx) ⊗ ϕρ(sρy)(cid:1) = ϕπ⊗ρ(cid:0)m′
π,ρ(sπx ⊗ sπy)(cid:1)
13
for all x ∈ B ⊗ L(Vπ) and y ∈ B ⊗ L(Vρ). Together with Equation (1) this shows
that the homomorphism
Φ(x) := V xV ⋆
maps A′ ⊆ L(cid:0)L2(A′)(cid:1) into A ⊆ L(cid:0)L2(A)(cid:1) and hence may be restricted to an
injective ∗-homomorphism Φ : A′ → A. Exchanging the role of sπ and s′
π shows
that Φ is in fact a ∗-isomorphism. Obviously, we have ϕπ(xπ∗
g) = ϕπ(x)π∗
g for all
x ∈ A2(π) and g ∈ G. It follows that V intertwines the G-action on L2(A) and
L2(A′), that is, we have V Ug = U ′
gV for all g ∈ G. Consequently, Φ intertwines
αg = Ad[Ug] on A and α′
g = Ad[U ′
g] on A′.
The following lemma rephrases freeness in terms of the multiplication maps defined in
(3). As a consequence we find that cleft dynamical systems are characterized factor
systems where the isometries ω(π, ρ) are in fact unitaries.
Lemma 5.7. A C∗-dynamical system (A, G, α) is free if and only if for all π ∈ G the
multiplication map
mπ,¯π : A2(π) ⊗B A2(¯π) → A2(π ⊗ ¯π), mπ,¯π(x ⊗ ¯y) := x12 · ¯y13
has dense range or, equivalently, is surjective.
Proof. First we note that mπ,¯π is an isometry of correspondences over B and hence it is
surjective if and only if it has dense range. Let us fix a finite-dimensional representation
(π, V ) of G and denote by d its dimension. For sake of a convenient notation we fix a basis
of V and write elements x ∈ A ⊗ L(V ) as matrices x = (xi,j)1≤i,j≤d with entries in A.
Likewise we write elements of A ⊗ L( ¯V ) and A ⊗ L(V ⊗ ¯V ) as matrices with respect to
the dual basis on ¯V and the product basis, respectively. A straightforward computation
shows that the transpose map A2(¯π) → A2(π)∗, x = (xi,j)i,j 7→ xt := (xj,i)i,j is a linear
bijection. Moreover, similar computations show that the map
ϕ : A2(π ⊗ ¯π) → C(π) ⊗ L(V ), ϕ(x)(i,j),(k,ℓ) := x(i,k),(j,ℓ)
is a linear bijection. Now consider the composition ψ := ϕ ◦ mπ,¯π, which takes the
concrete form
ψ(x ⊗ ¯y)(i,j),(k,ℓ) = xi,j ¯yk,ℓ
for x ∈ A2(π) and ¯y ∈ A2(¯π). Since A2(π) and A2(¯π) are right L(V )- and L( ¯V )-modules,
respectively, the range of ψ is a bimodule for 1C(π) ⊗ L(V ). It follows that the range
of ψ is of the form J ⊗ L(V ) for some subspace J ⊆ C(π). Furthermore, for elements
x, y ∈ A2(π) we may put ¯y := (y∗)t and find
(idC(π) ⊗ Tr)(cid:0)ψ(x ⊗ ¯y)(cid:1) = (cid:18) d
Xℓ=1
xi,ℓ y∗
ℓ,j(cid:19)i,j
= xy∗,
which shows that J = A2(π)A2(π)∗. We conclude that the multiplication map mπ,¯π is
surjective if and only if ψ has full range if and only if C(π) = J = A2(π)A2(π)∗.
14
Theorem 5.8. For a weakly cleft C∗-dynamical system (A, G, α) the following state-
ments are equivalent:
(a) The dynamical system is cleft or, equivalently, free.
(b) For some factor system (γ, ω) -- and hence for all factor systems -- all elements
ω(π, ρ) for π, ρ ∈ G are unitary.
Proof. By Lemma 4.5, for a cleft system we may choose unitary elements sπ in A2(π).
π⊗ρsπsρ (π, ρ ∈ G) of the corresponding
Then it follows that the elements ω(π, ρ) = s∗
factor system are unitary, too, which proves one implication. For the converse implica-
tion we take advantage of Lemma 5.7. Indeed, suppose that we have non-degenerated
isometries sπ ∈ A2(π) for each π ∈ G such that the elements ω(π, ρ) = s∗
π⊗ρsπsρ of the
corresponding factor system are unitaries for all π, ρ ∈ G. Then A2(π ⊗ ρ) is given by
A2(π ⊗ ρ) = sπ⊗ρ · B ⊗ L(Vπ ⊗ Vρ)
= sπ⊗ρ ω(π, ρ) · B ⊗ L(Vπ ⊗ Vρ) = sπsρ · B ⊗ L(Vπ ⊗ Vρ)
= sπ ·(cid:2)1B ⊗ L(Vπ) ⊗ 1ρ(cid:3) · sρ ·(cid:2)B ⊗ 1π ⊗ L(Vρ)(cid:3)
Since sπ ·(cid:0)1B ⊗ L(Vπ)(cid:1)⊆ A2(π) and sρ ·(cid:0)B ⊗ L(Vρ)(cid:1) = A2(ρ), we conclude that mπ,ρ is
surjective and hence the dynamical system is free.
Corollary 5.9. Every weakly cleft C∗-dynamical system (A, G, α) with commutative or
finite-dimensional fixed point algebra B is cleft.
Proof. Under the hypothesis on B an isometry ω ∈ B ⊗ L(V ) for finite-dimensional V is
automatically unitary.
In Section 6 we will discuss whether the converse of Theorem 5.6 holds, that is, whether
every factor system gives rise to a C∗-dynamical system. We postpone this problem
for the moment and continue to investigate the relation between the form of the fac-
tor system and the type of the dynamical system. First, let us characterize classical
systems.
Corollary 5.10. Let (A, G, α) be a cleft C∗-dynamical system with a commutative fixed
point algebra B and let (γ, ω) be an arbitrary associated factor system. Then A is com-
mutative if and only if
γπ(b) = b ⊗ 1π,
ω(ρ, π) = σ(cid:0)ω(π, ρ)(cid:1)
for all π, ρ ∈ G and b ∈ B, where σ denotes the tensor flip of L(Vπ) ⊗ L(Vρ).
and
15
Proof. First suppose that A is commutative. Let sπ ∈ A2(π), π ∈ G, be a family of
non-degenerated isometries and denote by (γ, u) the associated factor system. Since A
is commutative, every element b ⊗ 1 commutes with sπ in A ⊗ L(Vπ). It follows that
γπ(b) = s∗
π(b ⊗ 1)sπ = b ⊗ 1 for all b ∈ B. Likewise the elements (sπ)12 and (sρ)13
commute in A ⊗ L(Vπ) ⊗ L(Vρ). It follows that
σ(cid:0)ω(ρ, π)(cid:1) = σ(cid:0)s∗
ρ⊗πsρsπ(cid:1) = s∗
π⊗ρ sπ sρ = ω(π, ρ).
Conversely, let sπ ∈ A2(π), π ∈ G, be a family of non-degenerate isometries and suppose
the asserted condition on the corresponding factor system holds. Consider the family of
multiplication maps
mπ,ρ : A2(π) ⊗B A2(ρ) → A2(π ⊗ ρ), mπ,ρ(x ⊗ y) = x12 y13.
Looking at Equation (5), we have for all x ∈ B ⊗ L(Vπ) and y ∈ B ⊗ L(Vρ):
σ(cid:0)mρ,π(sρy ⊗ sπx)(cid:1) = σ(cid:0)sρ⊗π ω(ρ, π) y12 x13(cid:1) = sπ⊗ρ ω(π, ρ) x12 y13
= mπ,ρ(sπx ⊗ sρy),
because x12 and y13 commute in B ⊗ L(Vπ) ⊗ L(Vρ). From the general construction of
Section 3 we may then deduce that A is commutative.
Remark 5.11.
1. Corollary 5.10 provides a classification of cleft topological princi-
pal bundles.
2. Similar arguments as in the proof of Corollary 5.10 show the following more general
statement for a weakly cleft system: The center of B is contained in the center of
A if and only if for every factor system (γ, ω) the homomorphisms act trivially on
the center of B, i. e., γπ(z) = z ⊗ 1 for every central element z ∈ B and π ∈ G.
For a given group G and a given fixed point algebra B there clearly seems to be a trivial
dynamical system, namely (B ⊗C(G), G, id ⊗r). From our discussions in Example 4.7 we
immediately derive from Theorem 5.6 that for a weakly cleft dynamical system (A, G, α)
the following statements are equivalent:
(a) The dynamical system (A, G, α) is equivalent to (cid:0)B ⊗ C(G), G, id ⊗r(cid:1).
(b) Every factor system of (A, G, α) is conjugated to the factor system (γ, ω) given by
γπ(b) := b ⊗ 1
ω(π, ρ) := 1,
π, ρ ∈ G, b ∈ B
A next simple class of C∗-dynamical systems are those which are essentially ergodic
actions, that is, the dynamical system (A, G, α) arises from an ergodic dynamical system
(A0, G, α0) by tensoring with B.
16
Corollary 5.12. For a C∗-dynamical system (A, G, α) the following statements are
equivalent:
(a) (A, G, α) is isomorphic to (B ⊗ A0, G, idB ⊗α0) with an ergodic cleft C∗-dynamical
system (A0, G, α0).
(b) (A, G, α) is (weakly) cleft and admits a factor system of the form
γπ(b) = b ⊗ 1,
ω(π, ρ) ∈ 1B ⊗ L(Vπ ⊗ Vρ),
π, ρ ∈ G, b ∈ B.
Proof. For the system (B ⊗ A0, G, idB ⊗α0) the generalized isotypic component of π ∈ G
is of the form
A2(π) = B ⊗ A(0)
2 (π),
where A(0)
Since α0 is a cleft action, for each π ∈ G we find a unitary element s(0)
obtain a unitary sπ ∈ A2(π) by putting sπ := 1B ⊗ s(0)
system defined by Equations (2) and (4) is of the asserted form.
2 (π) ⊆ A0 ⊗ L(Vπ) denotes the generalized isotypic component of (A0, G, α0).
2 (π) and
π . Then the corresponding factor
π ∈ A(0)
Conversely, suppose that a factor system (γ, ω) for (A, G, α) has the asserted form.
Then ω(π, ρ) is a unitary for each π, ρ ∈ G and the cocycle condition (7) states that
ω = (cid:0)ω(π, ρ)(cid:1)π,ρ forms a unitary 2-cocycle for G in the usual sense. By [22] this 2-
cocycle provides an ergodic cleft C∗-dynamical system (A0, G, α0) whose factor system
is given by ω together with the trivial homomorphisms. (This may also be derived from
the construction presented in Section 6.) By the arguments in the first part of the proof
the dynamical system (B ⊗ A0, G, idB ⊗α0) then has the factor system (γ, ω) and, by
Theorem 5.6, it is isomorphic to (A, G, α).
Example 5.13. Let B = L(H) for some Hilbert space H. We claim that up to iso-
morphism every cleft dynamical system with fixed point algebra L(H) is of the form
(L(H) ⊗ A0, G, id ⊗α0) with an ergodic cleft action (A0, G, α0). This can be proved
directly by looking at minimal projections p ∈ L(H) and the corresponding restricted dy-
namical system on the algebra pAp. Alternatively, we may consider an arbitrary factor
system (γ, ω) for (A, G, α). Then for each representation π ∈ G the ∗-homomorphism
γπ : L(H) → L(H) ⊗ L(Vπ) is necessarily of the form γπ(b) = v∗
π(b ⊗ 1)vπ, b ∈ B, for
some unitary vπ ∈ L(H) ⊗ L(Vπ). Passing from (γ, ω) to a conjugated factor system,
we may without loss of generality assume γπ(b) = b ⊗ 1 for all b ∈ B. Then the coaction
condition states that for all π, ρ ∈ G the element ω(π, ρ) ∈ L(H ⊗ Vπ ⊗ Vρ) commutes
with L(H) ⊗ 1π⊗ρ and hence lies in 1 ⊗ L(Vπ ⊗ Vρ). Finally, Corollary 5.12 proves the
claim.
17
Corollary 5.14. For a C∗-dynamical system (A, G, α) and a unital C∗-subalgebra B0 ⊆
B the following statements are equivalent:
(a) There is an α-invariant C∗-subalgebra B0 ⊆ A0 ⊆ A such that the restricted system
(A0, G, αA0 ) is cleft with fixed point algebra B0.
(b) There is a factor system (γ, ω) of A with unitary elements ω(π, ρ) ∈ B0 ⊗L(Vπ ⊗Vρ)
and γπ(B0) ⊆ B0 ⊗ L(Vπ) for all π, ρ ∈ G.
In this case A0 can be chosen to have the factor system (cid:0)γπB0, ω(π, ρ)(cid:1)π,ρ∈ G and A is
generated by A0 and B.
To have a concise statement we claim equivalence here but at this point we prove only
one implication. The converse implication will be deduced later as Corollary 6.10.
Proof of (a) ⇒ (b). For an α-invariant subalgebra A0 ⊆ A its generalized isotypic com-
ponent of π ∈ G, denoted by A(0)
2 (π), is contained in the generalized isotypic com-
If the action on A0 is cleft, then for each π ∈ G we find a
ponent A2(π) of A.
2 (π). The elements sπ, π ∈ G, give rise to a factor system
unitary element sπ ∈ A(0)
for A0 and to a factor system for A. Both factor systems share the same unitaries
π⊗ρsπsρ ∈ B0 ⊗ L(Vπ ⊗ Vρ) for all π, ρ ∈ G. The ∗-homomorphisms of the
ω(π, ρ) = s∗
π(b ⊗ 1)sπ for π ∈ G, that is, they only
two factor systems are both given by γπ(b) = s∗
differ by their domains B0 and B, respectively, and the corresponding codomains.
Example 5.15. We would like to present a simple example that is not a tensor product
with a free ergodic action. For this purpose consider the group G = SU2 and the
commutative C∗-algebra C(X) for an arbitrary compact space X on which we fix a
non-trivial continuous reflection h : X → X, h ◦ h = idX. The group SU2 has up
to equivalence for every dimension precisely one irreducible representation. So we may
identify G with the natural numbers starting with V0 denoting the trivial representation
and V1 = C2 as the standard representation of SU2. As a factor system (γ, ω) we choose
the coaction γn : C(X) → C(X) ⊗ L(Vn), n ∈ N, given by
γn(b) := (b ⊗ 1
(b ◦ h) ⊗ 1 if n is odd
if n is even,
accompanied by the trivial 2-cocycle ω(n, m) := 1 for n, m ∈ N. The coaction condition
for γ = (γn)n∈N is easily verified, e. g., using the Clebsch-Gordan formula for SU2. Sup-
pose for a moment that there exists a C∗-dynamical system (A, G, α) with the factor
system (γ, ω). (This will be proven in Section 6.) It follows from Corollary 5.14 that
there is an α-invariant unital subalgebra A0 ⊆ A such that the restricted system is equiv-
alent to the trivial system(cid:0)C(SU2), G, r(cid:1) and the algebra A is generated by A0 = C(SU2)
and B = C(X). However, these two algebras satisfy non-trivial commutation relations.
18
Namely, we may collect all odd and even isotypic components and split A into its odd
and even part, that is, we call x ∈ A even if α−1(x) = x and odd if α−1(x) = (−x).
Recall that for each (π, V ) ∈ G the isotypic component A(π), as correspondence over B,
is isomorphic to B ⊗ L(V ) with the usual right right multiplication and the left mul-
tiplication b . x = γπ(b)x for x ∈ A(π) and b ∈ B. Since γπ maps into the center of
B ⊗ L(V ), we obtain for π odd that b . x = ((b ◦ h) ⊗ 1)x = x . (b ◦ h) and for π even
likewise b . x = x . b. Hence, for every x ∈ A0 = C(SU2) and b ∈ B = C(X) we have
bx = xb
bx = x(b ◦ h)
if x is even,
if x is odd.
Theses relations determine the C∗-algebra A and the action of G on A uniquely, that
is, A is the unique C∗-algebra generated by C(SU2) and C(X) subject to the above
relations. The algebra A is a particular case of the a twisted tensor product (see [9, 10]).
Remark 5.16. It is a well-known fact from homotopy theory that there exists, up to
isomorphy, exactly one principal SU2-bundle over the 3-sphere S3, namely the trivial
principal SU2-bundle S3 × SU2. It is used as a toy model for Chern-Simons theory devel-
oped by Witten [23] to derive a 3-dimensional quantum field theory in order to give an
intrinsic definition of the Jones polynomial and its generalizations dealing with knots in
three dimensional space. Example 5.15 shows that the noncommutative setting provides
more possibilities of constructing "noncommutative principal SU2-bundles" over S3.
6 Construction of Cleft Systems
In the previous section we have seen that factor systems provide an invariant of weakly
cleft C∗-dynamical systems. In the following we will see that they actually provide a full
classification. That is, we will show that for every factor system (γ, ω) for (B, G) there
actually is a weakly cleft C∗-dynamical system (A, G, α) with factor system (γ, ω). We
would like to mention that the construction principle is not limited to weakly cleft actions
and can be carried out more abstractly for general actions of compact quantum groups
on C∗-algebras (see [11]). However, our restricted setting allows some simplifications.
We will split the construction into three steps. In the first step we will show how a factor
system gives rise to a multiplication and hence an algebra. The second step will concern
the construction of an involution by exploiting the Hilbert module structure. In the last
step we proceed along the lines sketched in Section 3 to finally construct the C∗-algebra
and the dynamical system.
19
6.1 Associativity and Factor Systems
Once and for all let us fix a compact group G and a unital C∗-algebra B, and let (γ, ω) be
a pair consisting of a family γ = (γπ)π∈ G of unital ∗-homomorphisms γπ : B → B ⊗ L(Vπ)
and a family of isometries ω(π, ρ) ∈ B ⊗ L(Vπ ⊗ Vρ) for all π, ρ ∈ G. For the moment
we do not assume that (γ, ω) is a factor system. We extend both families naturally
to arbitrary representations of G, that is, we decompose given representations (π, V ),
(ρ, W ) into direct sums of irreducible components and define the corresponding maps γπ
and elements ω(π, ρ) componentwise. Equivalently, γπ : B → B ⊗ L(V ) and ω(π, ρ) ∈
B ⊗ L(V ⊗ W ) are the unique ∗-homomorphisms and unitary elements, respectively, such
that
γπ(b)v = vγσ(b)
and
ω(π, ρ)(v ⊗ w) = (v ⊗ w)ω(σ, σ′)
(8)
for all isometric intertwiners v : Vσ → V and w : Vσ′ → W with irreducible representa-
tions σ, σ′ ∈ G and all b ∈ B.
For each representation (π, V ) of G we consider the correspondence B⊗L(V ) over B given
by the usual right multiplication, the B-valued inner product hx, yiB := 1
(id ⊗ Tr)(x∗y),
dπ
and the left multiplication
b . x := γπ(b)x
for all b ∈ B and x, y ∈ B ⊗ L(V ). We write kxk2 := hx, xi1/2
for the corresponding
norm of x ∈ B ⊗ L(V ) and for an irreducible representation σ ∈ G, we write Pσ,π
for the map Pσ,π : B ⊗ L(V ) → B ⊗ L(Vσ) given by Pσ,π(x) := Pm
kxvk, where
v1, . . . , vm : Vσ → V is an orthonormal basis of intertwiners. The map does not depend
on the choice of the orthonormal basis. Moreover, the map Pσ,π is adjointable with
adjoint given by P ⋆
k for all z ∈ B ⊗ L(Vσ).
k=1 v∗
B
σ,π(z) = dπ
k=1 vkyv∗
dσ Pm
For each pair of representations (π, V ) and (ρ, W ) of G we define a linear map by
mπ,ρ : [B ⊗ L(V )] ⊗B [B ⊗ L(W )] −→ B ⊗ L(V ⊗ W )
mπ,ρ(x ⊗ y) := ω(π, ρ) · (idπ ⊗γρ)(x) · (1π ⊗ y).
We refer to mπ,ρ as multiplication map. For sake of a concise notation we will frequently
drop the subindex and simply write m if the respective domains are not substantial and
clearly determined by the context. Moreover, we amplify in the canonical way. With
this simplifications the map mπ,ρ takes the form
mπ,ρ(x ⊗ y) = ω(π, ρ) γρ(x)y
for all x ∈ B ⊗ L(V ) and y ∈ B ⊗ L(W ).
20
Lemma 6.1. The map mπ,ρ is an adjointable isometry.
Proof. Recall that for any C∗-algebra A, equipped with the natural right inner product
hx, yiA := x∗y (x, y ∈ A), and for any isometry s ∈ A the map A → A, a 7→ sa is an
adjointable isometry. In particular, it follows that the map z 7→ ω(π, ρ) z on B⊗L(V ⊗W )
is an adjointable isometry. Therefore, the assertion follows from the fact that the tensor
product [B⊗L(V )]⊗B [B⊗L(W )] is canonically isometrically isomorphic to B⊗L(V ⊗W )
via the isomorphism x ⊗ y 7→ γρ(x)y for all x ∈ B ⊗ L(V ) and y ∈ B ⊗ L(W ).
Lemma 6.2. Suppose (γ, ω) is a factor system. Then, for all representations π, ρ of G
the map mπ,ρ is a B-bimodule map. Moreover, we have
(idπ ⊗mρ,σ) ◦ (m⋆
π,ρ ⊗ idσ) = m⋆
π,ρ⊗σ ◦ mπ⊗ρ,σ.
(9)
for all representations π, ρ, σ of G. In particular, the family of maps mπ,ρ is associative,
i. e.,
mπ,ρ⊗σ ◦ (idπ ⊗mρ,σ) = mπ⊗ρ,σ ◦ (mπ,ρ ⊗ idσ).
Proof. The right module property of mπ,ρ is obviously satisfied even without assuming
that (γ, ω) is a factor system. Furthermore, by the coaction condition of the factor
system, Equation (6), we obtain
m(cid:0)b . (x ⊗ y)(cid:1) = m(cid:0)γπ(b)x ⊗ y(cid:1) = ω(π, ρ) γρ(cid:0)γπ(b)x(cid:1) y
(6)= γπ⊗ρ(b) ω(π, ρ) γρ(x)y = b . m(x ⊗ y)
for all b ∈ B, x ∈ B ⊗ L(V ), and y ∈ B ⊗ L(W ). This shows that mπ,ρ is indeed
a B-bimodule map. In particular, the left and right hand side of (9) are well-defined
maps on the correspondence [B ⊗ L(Vπ ⊗ Vρ)] ⊗B [B ⊗ L(Vσ)].
In order to verify (9)
for representations π, ρ, σ of G, we canonically identify the tensor product of the three
correspondences B ⊗ L(Vπ), B ⊗ L(Vρ), and B ⊗ L(Vσ) over B with the correspondence
B ⊗ L(Vπ ⊗ Vρ ⊗ Vσ) via the isometric isomorphism x ⊗ y ⊗ z 7→ γσ(cid:0)γρ(x)y(cid:1)z. Then the
cocycle condition, Equation (7), implies that
(idπ ⊗mρ,σ)(cid:0)(m⋆
π,ρ ⊗ idσ)(x)(cid:1) = ω(ρ, σ) γσ(cid:0)ω(π, ρ)(cid:1)∗
(7)= ω(π, ρ ⊗ σ)∗ ω(π ⊗ ρ, σ) x = m⋆
x,
π,ρ⊗σ(cid:0)mπ⊗ρ,σ(x)(cid:1)
for all x ∈ B ⊗ L(Vπ ⊗ Vρ ⊗ Vσ), which proves Equation (9). Multiplying (9) with
(mπ,ρ ⊗ idσ) from the right and mπ,ρ⊗σ from the left yields
mπ,ρ⊗σ(idπ ⊗mρ,σ) = mπ,ρ⊗σ ◦ m⋆
π,ρ⊗σ ◦ mπ⊗ρ,σ ◦ (mπ,ρ ⊗ idσ).
Since the left hand side of this equation and the term S := mπ⊗ρ,σ ◦ (mπ,ρ ⊗ idσ) on
the right hand side are both isometries, the orthogonal projection mπ,ρ⊗σ ◦ m⋆
π,ρ⊗σ acts
trivially on the rage of S. Therefore, we may cancel mπ,ρ⊗σ ◦ m⋆
π,ρ⊗σ, which shows
associativity.
21
Remark 6.3. The normalization condition ω(1, 1) = 1B of Definition 5.3 is equivalent
to the fact that for the trivial representation the multiplication map m recovers the
B-bimodule structure, that is, for a representation (π, V ) of G we have
m1,π(b ⊗ x) = b . x = γπ(b)x,
mπ,1(x ⊗ b) = x . b = x(b ⊗ 1π)
for all elements x ∈ B ⊗ L(V ) and b ∈ B. In particular, in this case m1,1 coincides with
the usual multiplication of B.
In order to define an algebra, we consider the algebraic direct sum of the correspondences
B ⊗ L(Vπ) for π ∈ G:
B ⊗ L(Vπ).
A := Mπ∈ G
The left and right action of B are given componentwise and the B-valued inner product
is hx, yiB = Pσ∈ Ghxσ, yσiB, where xσ and yσ denote the components of x and y, respec-
tively. On A we define the product of x ∈ B ⊗ L(Vπ) and y ∈ B ⊗ L(Vρ) with π, ρ ∈ G
by
x • y := Xσ∈ G
Pσ,π⊗ρ(cid:0)m(x ⊗ y)(cid:1).
Bilinear extension then yields a bilinear map (x, y) 7→ x • y on A. For x ∈ B or y ∈ B
this product coincides with the left or right action of B, respectively.
Due to the intertwining relations (8) the family of maps mπ,ρ behaves nicely with respect
to intertwiners. It is straightforward to check that for all representations π, ρ of G and
every intertwiner v : Vσ → Vπ we have
m(vxv∗ ⊗ y) = (v ⊗ 1ρ) m(x ⊗ y) (v ⊗ 1ρ)∗,
m(v∗zv ⊗ y) = (v ⊗ 1ρ)∗m(z ⊗ y)(v ⊗ 1ρ)
(10)
for all x ∈ B ⊗ L(Vπ) and y ∈ B ⊗ L(Vρ), and similar equations hold for adjoining inter-
twiners in the right tensor factor. Using this relations, the associativity of Lemma 6.2
and some algebra straightforwardly yields the following:
Lemma 6.4. Suppose (γ, ω) is a factor system. Then the product • on A is associative,
that is, (A, •) is an algebra.
6.2 Constructing the Involution
Throughout the remainder of the section let us assume that (γ, ω) is a factor system.
In order to define an involution on the algebra A, let consider a fixed π ∈ G. We put
pπ := P ∗
π-th multiple of the orthogonal projection onto the
1,π⊗¯π(1B), that is, pπ is the d2
22
fixed point space of the representation π ⊗ ¯π. For every x ∈ B ⊗ L(Vπ) we define an
element in B ⊗ L( ¯Vπ) by
J(x) := (L⋆
x ◦ m⋆
π,¯π)(pπ),
where we denote by Lx : B ⊗ L( ¯Vπ) → [B ⊗ L(Vπ)] ⊗B [B ⊗ L( ¯Vπ)] the adjointable
map given by Lxy := x ⊗ y. Extending this antilinearly to all summands provides an
antilinear map J : A → A. On the subalgebra B ⊆ A (i. e., for π = 1) we immediately
find J(b) = b∗ for all b ∈ B. For the other summands, however, J does not coincide with
the usual involution on B ⊗ L(Vπ) in general.
Theorem 6.5. For all x, y, z ∈ A we have
hJ(x) • y, ziB = hy, x • ziB.
Proof. Let x ∈ B ⊗ L(Vπ), y ∈ B ⊗ L(Vρ), and z ∈ B ⊗ L(Vσ) with π, ρ, σ ∈ G. Since
(γ, ω) is a factor system, the multiplication maps satisfy (9) of Lemma 6.2. Then, writing
z := P ⋆
σ,¯π⊗ρ(z) for short, we obtain
hJ(x) • y, ziB = hm¯π,ρ(cid:0)J(x) ⊗ y(cid:1), ziB = hm¯π,ρ(cid:0)L⋆
π,¯π ⊗ idρ)(pπ ⊗ y), ziB
π,¯π(pπ) ⊗ y(cid:1), ziB
xm⋆
π,¯π⊗ρmπ⊗¯π,ρ(pπ ⊗ y), ziB
x(idπ ⊗m¯π,ρ)(m⋆
xm⋆
= hL⋆
(9)= hL⋆
= hmπ⊗¯π,ρ(pπ ⊗ y), mπ,¯π⊗ρ(x ⊗ z)iB.
(8)= h(P ⋆
id,π⊗¯π ⊗ idρ)(y), (idπ ⊗P ⋆
σ,¯π⊗ρ)m(x ⊗ z)iB.
We fix an isometric intertwiner w : C → Vπ⊗¯π and we choose an arbitrary orthonormal
bases of intertwiners v1, . . . , vm : Vσ → ¯Vπ ⊗ Vρ. Then uk := q dπdρ
(1π ⊗ vk)∗(w ⊗ 1ρ) for
1 ≤ k ≤ m form an orthonormal basis of intertwiners from ρ to π ⊗ σ. Orthonormality
follows from the fact that for every T ∈ L( ¯Vπ) we have w∗(1π ⊗ T )w = 1
Tr(T ). Com-
dπ
pleteness holds because the dimension of the intertwiner space from ρ to π ⊗ σ coincides
with the dimension of the intertwiner space from σ to ρ ⊗ ¯π. With this intertwiners
we find Pρ,π⊗σ = (Pid,π⊗¯π ⊗ idρ)(idπ ⊗P ⋆
σ,¯π⊗ρ). Continuing the above computation, we
obtain
dσ
hJ(x) • y, ziB = hy, Pρ,π⊗σm(x ⊗ z)iB = hy, x • ziB.
The assertion for arbitrary x, y, z ∈ A then follows by extending (anti-)linearly.
Corollary 6.6. Let P0 : A → B denote the orthogonal projection onto the direct sum-
mand B ⊆ A corresponding to the trivial representation. Then for all x, y ∈ A we have
hx, yiB = P0(cid:0)J(x) • y(cid:1).
Proof. The element 1B is a unit for the multiplication of A and fixed by the orthogonal
projection P0. Hence by Theorem 6.5 we obtain
hx, yiB = h1B, J(x) • yiB = (cid:10)1B, P0(cid:0)J(x) • y(cid:1)(cid:11)B = P0(cid:0)J(x) • y(cid:1).
23
Remark 6.7. We would like to mention that by Corollary 6.6 the element J(x) can be
equivalently be characterized as the unique element of A satisfying hJ(x), yiB = P0(x • y)
for all y ∈ A.
Corollary 6.8. A is an involutive algebra, i. e., for all x, y ∈ A we have
J(cid:0)J(x)(cid:1) = x,
J(x • y) = J(y) • J(x).
Proof. By applying Theorem 6.5 twice we get
and
(cid:10)J(cid:0)J(x)(cid:1), z(cid:11)B = h1B, J(x) • ziB = hx, ziB
hJ(x • y), ziB = h1B, x • y • ziB = hJ(x), y • ziB = hJ(y) • J(x), ziB
for all z ∈ A. Since the inner product separates points, this yields the assertion.
6.3 Construction of Free Actions
Having the algebra A and the involution on A in hands, the construction of the C∗-alge-
bra and the cleft action follows the outline presented in Section 3. We consider A as a
right pre-Hilbert B-module. The inner product is positive definite but, unless G is finite,
A is not closed with respect to the induced norm
kxk2 := khx, xiBk1/2
x ∈ A.
op = kP0(cid:0)J(x) • x(cid:1)k1/2
op ,
We denote by ¯A the completion with respect to this norm. Equivalently, the correspon-
dence ¯A over B is the direct sum of the previously discussed correspondences B ⊗ L(Vπ)
with π ∈ G.
Theorem 6.5 allows us to extend each left multiplication by x ∈ A to an adjointable map
on the completion ¯A, which we again denote by λ[x] : ¯A → ¯A. Then the map
λ : A → L( ¯A),
x 7→ λ[x]
is an representation of the ∗-algebra A by adjointable operators on ¯A. The vector 1B ∈ ¯A
is clearly cyclic and separating for the subalgebra λ(A). In particular, λ is faithful and
hence isometric on B ⊆ A. We denote by A the C∗-algebra generated by the range of λ.
To simplify the notation we identify the algebra (A, •) with the subalgebra λ(A) ⊆ A.
For sake of clarity, we extend the notation for the ∗-algebra A to the C∗-algebra L( ¯A),
that is, for elements x, y ∈ L( ¯A) we write x • y for their product and for x ∈ A we write
J(x) for its adjoint. Since for x ∈ A we have kλ[x]k2 ≥ kλ[x]1Bk2
2, we may
regard the C∗-algebra A as a subset of ¯A, so that A ⊆ A ⊆ ¯A.
2 = kxk2
24
On each direct summand B ⊗ L(Vρ) ⊆ ¯A, π ∈ G, we have a continuous action of G by
the usual right multiplication
Ug(x) := x (1B ⊗ π∗
g)
for g ∈ G and x ∈ B ⊗ L(Vπ). For each g ∈ G the map Ug is unitary on B ⊗ L(Vρ) with
respect to the B-valued inner product. Taking direct sums and continuous extension, we
obtain a strongly continuous unitary representation g 7→ Ug ∈ L( ¯A) of the group G on
the correspondence ¯A. We denote by α = (αg)g∈G the associated automorphism group
on L( ¯A), i. e., we put
αg(x) := Ug • x • U ⋆
g ,
x ∈ L( ¯A).
For each element x ∈ A ⊆ A we have αg(x) = Ug(x), that is, on the algebra A ⊆ ¯A
the actions α = (αg)g∈G and U = (Ug)g∈G coincide. Since A is dense in A, it follows
that g 7→ αg(x) is continuous for every x ∈ A. Summarizing, we have constructed a
C∗-dynamical system (A, G, α).
Theorem 6.9. The C∗-dynamical system (A, G, α) is weakly cleft with fixed point algebra
B and factor system (γ, ω).
Proof. In the first part of the proof we will show that for the action U = (Ug)g∈G on
¯A the isotypic component of π ∈ G is given by the direct summand B ⊗ L( ¯Vπ). This
verifies in particular that B is the fixed point space. By Theorem 6.5 it follows that the
canonical B-valued inner product on A coincides with the inner product on the larger
space ¯A, that is,
ZG
αg(cid:0)J(x) • y(cid:1) dg = P0(cid:0)J(x) • y)
6.5
= hx, yiB.
for all x, y ∈ A. Using Lemma 4.3 we conclude that the dynamical system (A, G, α) is
weakly cleft. In the second part of the proof we confirm that (γ, ω) is indeed a factor
system of (A, G, α).
1. We consider the action U = (Ug)g∈G on the correspondence ¯A over B. Obviously,
all elements of B ⊆ ¯A are fixed by the action. For an element x ∈ B ⊗ L(Vπ) in a di-
g dg vanishes
because π is irreducible. Taking linear combinations and continuous extensions
rect summand of ¯A with id 6= π ∈ G the integral RG Ug(x) dg = RG xπ∗
then shows that RG Ug(x) dg = P0(x) for all x ∈ ¯A. Adapting the arguments, we
obtain for every π ∈ G and x ∈ ¯A
dπZG
Tr(π∗
g)Ug(x) dg = P¯π(x),
where P¯π : ¯A → B ⊗ L( ¯Vπ) denotes the orthogonal projection onto the direct
summand B ⊗ L( ¯Vπ) ⊆ ¯A. We conclude that B ⊗ L( ¯Vπ) is the isotypic component
of π in ¯A and hence also in A ⊆ ¯A.
25
2. It remains to verify that (γ, ω) is a factor system of (A, G, α). According to the
first part of the proof the isotypic component of ¯π ∈ G in A is given by B ⊗ L(Vπ).
As a right B-module B ⊗ L(Vπ) is generated by the element 1B ⊗ 1π of norm 1.
The natural isomorphism of the isotypic and the generalized isotypic component
together with Lemma 4.3 therefore provides us with the non-degenerate isometry
sπ ∈ A2(π) ⊆ A ⊗ L(Vπ) given by
sπ = dπZG
1B ⊗ π∗
g ⊗ πg dg = (1B ⊗ F ),
where F ∈ L(V ) ⊗ L(V ) denotes the tensor flip. This family s = (sπ)π∈ G of
isometries then gives rise to a factor system (γ, ω) by Equations (2) and (4). For
each π ∈ G the homomorphism γπ is uniquely determined by the left action of B
on sπ. For convenience we also write • for the multiplication on all A ⊗ L(Vπ),
π ∈ G. Then the left action of B on sπ is given by
(b ⊗ 1π) • sπ = dπZG(cid:0)b • (1B ⊗ π∗
g)(cid:1) ⊗ πg
= (cid:0)γπ(b) ⊗ 1π(cid:1)(1B ⊗ F ) = sπ(1B ⊗ F )(cid:0)γπ(b) ⊗ 1π(cid:1)(1B ⊗ F )
g)(cid:1) ⊗ πg = dπZG(cid:0)γπ(b)(1B ⊗ π∗
= sπ • γπ(b)
for every b ∈ B. Consequently, γπ = γπ. With some algebra involving the multipli-
cation •, which we leave to the reader, it is then straightforward to show that we
also have
sπ • sρ = sπ⊗ρ • ω(π, ρ)
for all π, ρ ∈ G. This proves ω(π, ρ) = s∗
π⊗ρ • sπ • sρ = ω(π, ρ). Summarizing, we
have shown that (γ, ω) is indeed the factor system associated with the isometries
sπ, π ∈ G, which finishes the proof.
This concrete representation theorem of weakly cleft actions finally allows us as a corol-
lary to complete the proof of Corollary 5.14, which we asserted in Section 5. For conve-
nience we repeat the relevant statement as a reminder in a slightly more general form.
Corollary 6.10. Let (A, G, α) be a C∗-dynamical system and B0 ⊆ B a unital C∗-sub-
algebra of fixed points. Furthermore, let (γ, ω) be a factor system such that
γπ(B0) ⊆ B0 ⊗ L(Vπ),
ω(π, ρ) ∈ B0 ⊗ L(Vπ ⊗ Vρ),
for all π, ρ ∈ G. Then there is an α-invariant subalgebra A0 ⊆ A such that the restricted
dynamical system (A0, G, αA0 ) is weakly cleft and A is generated by A0 and B.
Proof. By Theorem 6.9 and Theorem 5.6 we may assume that A admits an α-invariant
dense subalgebra of the form A = Lπ∈ G B ⊗ L(Vπ) with the multiplication
x • y = Xσ∈ G
Pσ(cid:0)ω(π, ρ) γρ(x) y(cid:1)
26
and the action αg(x) = x(1B ⊗ π∗
π, ρ ∈ G. By the hypotheses on (γ, ω) the subset
g) for all x ∈ B ⊗ L(Vπ), y ∈ B ⊗ L(Vρ), g ∈ G, and
A0 := Mπ∈ G
B0 ⊗ L(Vπ)
is an α-invariant ∗-subalgebra of A (cf. also Remark 6.7). Let A0 ⊆ A be the closure of
A0 with respect to the operator norm. By Theorem 6.9, the dynamical system restricted
to A0 is weakly cleft with the restriction of (γ, ω) to B0 as factor system. The space A
is linearly generated by all products x • b = (b ⊗ x) with x ∈ 1B ⊗ L(Vπ), π ∈ G, and
b ∈ B. Since A0 contains all elements x ∈ 1B ⊗ L(Vπ), π ∈ G, the algebra A is generated
by A0 and B. Taking closures, A is generated by A0 and B.
Acknowledgment
We would like to acknowledge the Center of Excellence in Analysis and Dynamics Re-
search (Academy of Finland, decision no. 271983 and no. 1138810) for supporting this
research.
References
[1] B. Blackadar. Operator Algebras, volume 112, Operator Algebras and Non-
Communtative Geometry of Encyclopaedia Math. Sci. Springer, 2006.
[2] T. Brzeziński and P. M. Hajac. The Chern-Galois character. C. R. Math. Acad.
Sci. Paris, 338(2):113 -- 116, 2004.
[3] A. A. Davydov. Galois algebras and monoidal functors between categories of repre-
sentations of finite groups. Journal of Algebra, 244(1):273 -- 301, 2001.
[4] K. De Commer and M. Yamashita. A construction of finite index C∗-algebra in-
clusions from free actions of compact quantum groups. Publ. Res. Inst. Math. Sci.,
49(4):709 -- 735, 2013.
[5] Y. Doi.
Equivalent crossed products for a Hopf algebra. Comm. Algebra,
17(12):3053 -- 3085, 1989.
[6] Y. Doi and M. Takeuchi. Cleft comodule algebras for a bialgebra. Comm. Algebra,
14(5):801 -- 817, 1986.
27
[7] S. Echterhoff, R. Nest, and H. Oyono-Oyono. Principal non-commutative torus
bundles. Proc. Lond. Math. Soc., 99(1):1 -- 31, 2009.
[8] P. Green. C∗-algebras of transformation groups with smooth orbit space. Pacific J.
Math., 72(1):71 -- 97, 1977.
[9] G. G. Kasparov. The operator K-functor and extensions of C∗-algebras. Mathemat-
ics of the USSR-Izvestiya, 16(3), 1981.
[10] R. Meyer, S. Roy, and S. L. Woronowicz. Quantum group-twisted tensor products
of C∗-algebras. Internat. J. Math., 25(02):1450019, 2014.
[11] S. Neshveyev. Duality theory for nonergodic actions. Münster J. Math., 2013. to
appear.
[12] O. Pfante. A Chern-Simons action for noncommutative spaces in general with the
example SUq(2). J. Noncommut. Geom., 8(3):611 -- 654, 2014.
[13] C. N. Phillips. Equivariant K-Theory and Freeness of Group Actions on C∗-Algebras,
volume 1274 of Lecture Notes in Math. Springer, 1987.
[14] J. Phillips and I. Raeburn. Twisted crossed products by coactions. J. Aust. Math.
Soc., 56(3):320 -- 344, June 1994.
[15] N. C. Phillips. Freeness of actions of finite groups on C∗-algebras.
In Operator
structures and dynamical systems, volume 503 of Contemp. Math., pages 217 -- 257.
Amer. Math. Soc., 2009.
[16] M. A. Rieffel. Proper actions of groups on C∗-algebras.
In H. Araki and R. V.
Kadison, editors, Mappings of Operator Algebras, volume 84 of Progr. Math, pages
141 -- 182. Birkhäuser, 1991.
[17] P. Schauenburg. Hopf-Galois and bi-Galois extensions. In G. Janelidze, B. Pareigis,
and W. Tholen, editors, Galois Theory, Hopf Algebras, and Semiabelian Categories,
volume 43 of Fields Inst. Commun., pages 469 -- 515. Amer. Math. Soc., 2004.
[18] K. Schwieger and S. Wagner. Part I, Free actions of compact Abelian groups on
C∗-algebras. Adv. Math., 317:224 -- 266, 2017.
[19] N. Steenrod. The Topology of Fibre Bundles. Princeton Univ. Press, 1st edition,
1999.
[20] S. Vaes and L. Vainerman. Extensions of locally compact quantum groups and the
bicrossed product construction. Adv. Math., 175(1):1 -- 101, 2003.
28
[21] A. Wassermann. Ergodic actions of compact groups an operator algebras. II. Clas-
sification of full multiplicity ergodic actions. Canad. J. Math., 40(6):1482 -- 1527,
1988.
[22] A. Wassermann. Ergodic actions of compact groups on operator algebras. I. General
theory. Ann. of Math. (2), 130(2):273 -- 319, 1989.
[23] E. Witten. Quantum field theory and the Jones polynomial. Comm. Math. Phys,
121(3):351 -- 399, 1989.
29
|
1210.2964 | 1 | 1210 | 2012-10-10T16:03:26 | Tensorial Function Theory: From Berezin transforms to Taylor's Taylor series and back | [
"math.OA",
"math.CV",
"math.FA",
"math.RA"
] | Let $H^{\infty}(E)$ be the Hardy algebra of a $W^{*}$-correspondence $E$ over a $W^{*}$-algebra $M$. Then the ultraweakly continuous completely contractive representations of $H^{\infty}(E)$ are parametrized by certain sets $\mathcal{AC}(\sigma)$ indexed by $NRep(M)$ - the normal *-representations $\sigma$ of $M$. Each set $\mathcal{AC}(\sigma)$ has analytic structure, and each element $F\in H^{\infty}(E)$ gives rise to an analytic operator-valued function $\hat{F}_{\sigma}$ on $\mathcal{AC}(\sigma)$ that we call the $\sigma$-Berezin transform of $F$. The sets ${\mathcal{AC}(\sigma)}_{\sigma\in\Sigma}$ and the family of functions ${\hat{F}_{\sigma}}_{\sigma\in\Sigma}$ exhibit "matricial structure" that was introduced by Joeseph Taylor in his work on noncommutative spectral theory in the early 1970s. Such structure has been exploited more recently in other areas of free analysis and in the theory of linear matrix inequalities. Our objective here is to determine the extent to which the matricial structure characterizes the Berezin transforms. | math.OA | math |
TENSORIAL FUNCTION THEORY:
FROM BEREZIN TRANSFORMS TO TAYLOR'S TAYLOR SERIES AND
BACK
PAUL S. MUHLY AND BARUCH SOLEL
Abstract. Let H ∞(E) be the Hardy algebra of a W ∗-correspond-
ence E over a W ∗-algebra M . Then the ultraweakly continuous
completely contractive representations of H ∞(E) are parametrized
by certain sets AC(σ) indexed by N Rep(M ) - the normal ∗-repres-
entations σ of M . Each set AC(σ) has analytic structure, and
each element F ∈ H ∞(E) gives rise to an analytic operator-valued
function bFσ on AC(σ) that we call the σ-Berezin transform of F .
The sets {AC(σ)}σ∈Σ and the family of functions {bFσ}σ∈Σ exhibit
"matricial structure" that was introduced by Joeseph Taylor in
his work on noncommutative spectral theory in the early 1970s.
Such structure has been exploited more recently in other areas of
free analysis and in the theory of linear matrix inequalities. Our
objective here is to determine the extent to which the matricial
structure characterizes the Berezin transforms.
1. Introduction
Our purpose in this paper is to explore relations among three subjects:
(1) the theory of Berezin transforms that arise from the representation
theory of tensor algebras and Hardy algebras of W ∗-correspondences,
(2) infinite dimensional holomorphy, and (3) the theory of free holo-
morphic functions initiated by Joseph Taylor in [Tay72, Tay73]. In this
introduction we indicate the connections we have in mind and provide
a bit of context. Details and a fuller account, including relevant defini-
tions of terms left undefined here, will be given in subsequent sections.
Suppose first that M is a W ∗-algebra and that E is a W ∗-correspond-
ence over M.1 With these ingredients one may build two operator
algebras, T+(E) and H ∞(E), that are generated by a copy of M and
The research of both authors was supported in part by a US-Israel Binational
Science Foundation grant. The second author was also supported by the Technion
V.P.R. Fund.
1We shall assume throughout that M has a separable predual and that E is
countably generated.
1
TENSORIAL FUNCTION THEORY
2
the creation operators {Tξ ξ ∈ E} acting on the full Fock space
F (E) = M ⊕ E ⊕ E⊗2 ⊕ E⊗3 ⊕ · · · : T+(E), the tensor algebra of E,
is the norm closed algebra generated by these objects and H ∞(E), the
Hardy algebra of E, is its ultraweak closure. One may think of T+(E) as
a generalization of the disc algebra A(D), and H ∞(E) may be viewed
as a generalization of the classical Hardy space, H ∞(D), consisting of
the bounded analytic functions on D.
Indeed, when M = C = E,
then T+(E) is naturally completely isometrically isomorphic to A(D)
and H ∞(E) is naturally completely isometrically isomorphic and weak-
∗ homeomorphic to H ∞(D). Another important example to keep in
mind is that which arises when M = C and E = Cd, d ≥ 2.
In
this case, T+(E) is naturally completely isometrically isomorphic to
Gelu Popescu's noncommutative disc algebra Ad [Pop96] and H ∞(E)
has come to be called the algebra of noncommutative analytic Toeplitz
operators. The terminology is due to Davidson and Pitts [DP98], and
much of the initial theory of these algebras is due to them. Other
interesting finite dimensional settings can be constructed from graphs
or quivers. (See [Muh97, MS99, KK06] for examples.)
It is worthwhile to emphasize that even if one were interested only in
these finite dimensional examples, it is useful to work with the gen-
eral theory. One reason is that the general theory is invariant under
Morita equivalence [MS00, MS11a] and, among many things, Morita
theory allows one to study every T+(E) and H ∞(E) in terms of an-
alytic crossed products - generalizations of twisted polynomial rings -
that have played such a prominent role in the theory of non-self-adjoint
operator algebras. The point is that although T+(E) and H ∞(E) look
very much like algebras of functions of several (noncommutative) vari-
ables, as we shall see, they behave much more like algebras of functions
of one variable than one might expect.
In [MS98, Theorem 2.9], we showed that every completely contractive
representation of T+(E) is given by a pair, (σ, z), where σ is a normal
representation of M on a Hilbert space Hσ and z : E ⊗σ Hσ → Hσ is
an operator of norm at most 1 that intertwines σE ◦ ϕ and σ, where
σE is the representation of L(E) that is induced by σ in the sense of
Rieffel [Rie74a] and where ϕ denotes the left action of M on E. We
denote the representation associated to (σ, z) by σ × z.
(In general, if A is a not-necessarily-self-adjoint algebra and if π and ρ
are two representations of A by bounded operators on Hilbert spaces
Hπ and Hρ, respectively, then we shall write I(π, σ) for the collection of
all operators C from Hπ to Hρ such that Cπ(a) = ρ(a)C for all a ∈ A
TENSORIAL FUNCTION THEORY
3
and we call I(π, ρ) the intertwiner space or the space of intertwiners
from π to ρ.)
For each normal representation σ of M, we endow I(σE ◦ ϕ, σ) with
the operator norm and we write D(0, 1, σ) for the open unit ball in
I(σE ◦ ϕ, σ). Then, when σ is fixed, each F ∈ T+(E) gives rise to a
B(Hσ)-valued function bFσ defined on the closed unit ball D(0, 1, σ) by
the formula
(1.1)
bFσ(z) := σ × z(F ).
all normal representations of M on separable Hilbert space, NRep(M),
the σ-Berezin transform of F .2 The collection of all σ-Berezin trans-
a continuous function on D(0, 1, σ) with values in B(Hσ), where both
Because of formulas that we derived in [MS09, Theorem 13], we call bFσ
forms, {bFσ}σ∈N Rep(M ), obtained by letting σ range over the collection of
is called the Berezin transform of F and will be written simply as bF .
It is easy to see that for F ∈ T+(E), each σ-Berezin transform bFσ is
spaces are given the norm operator topology. Further, bFσ is holomor-
phic in the sense of Frechet [HP74, 112 and 778] when restricted to
D(0, 1, σ).
A similar sort of representation exists for elements F ∈ H ∞(E). How-
ever, for these F , bFσ makes good sense only on the set of zs in D(0, 1, σ)
such that σ × z extends from T+(E) to an ultraweakly continuous rep-
resentation of H ∞(E) in B(Hσ). We denote the set of such points by
AC(σ) and for reasons spelled out in [MS11b] we call them the abso-
lutely continuous points in D(0, 1, σ).
It turns out that D(0, 1, σ) ⊆
AC(σ) [MS04, Corollary 2.14]. Thus, for F ∈ H ∞(E), bFσ makes sense
It is, in fact, bounded and holomorphic
The Frechet power series of bFσ is easy to calculate and has a remarkably
simple expression in terms of the tensorial "Fourier series" in which F
may be expressed using the gauge automorphism group built from the
number operator on the full Fock space F (E). We call power series
with this special form tensorial power series (Definition 2.9).
It is
natural to inquire about the structure of such power series, in general,
and one soon sees that much of standard elementary theory of complex
as a function on D(0, 1, σ).
with respect to the norm topologies on D(0, 1, σ) and B(Hσ).
2We are indebted to Lew Coburn for calling our attention to the connection be-
tween our formulas and the classical Berezin transform associated with the Hardy
space on the open unit disc in the complex plane. We note, too, that our termi-
nology agrees with that of Gelu Popescu [Pop08] in those settings where his theory
and ours overlap.
TENSORIAL FUNCTION THEORY
4
analysis on the open unit disc can be recapitulated in the more general
setting we are describing. When M = C and E = Cd, this has been
done already by Popescu in [Pop06].
is an algebra under pointwise multiplication that may be identified
naturally with a quotient of H ∞(E) by an ultraweakly closed ideal
For a given σ, we write B(σ) for {bFσ F ∈ H ∞(E)}. Then B(σ)
and the map F → bFσ is a complete quotient map. If σ is a faithful
normal representation of M of infinite multiplicity, then the map is
a completely isometric isomorphism from H ∞(E) onto B(σ) [MS08,
Lemma 3.8]. In general, however, B(σ) is a proper quotient of H ∞(E).
Indeed, when M = C, and E = Cd, with d ≥ 2, then each normal
representation of M is, of course, (unitarily equivalent to) a multiple
of the representation σ1 of C on the one-dimensional Hilbert space, C.
The σ1-Berezin transforms {bFσ1 F ∈ H ∞(E)} are the multipliers of
the Drury-Arveson space and so form a commutative algebra. On the
other hand, B(∞σ1) is isomorphic to the algebra of noncommutative
analytic Toeplitz operators in the fashion just described. For 2 ≤
n < ∞, B(nσ1) is a completion of the algebra of d generic n × n
matrices. So, as soon as n ≥ 2, it is noncommutative. But also, by
virtue of the polynomial identities these algebras satisfy, it is easy to see
that when n 6= m, B(nσ1) ≇ B(mσ1), and that no finite dimensional
representation of C yields a faithful representation of either H ∞(E) or
of T+(E) in terms of Berezin transforms. Thus there arise very natural
questions: How much is lost when forming the σ-Berezin transform
of an element in H ∞(E)? How might one reconstruct an F from its
finite dimensional σ-Berezin transforms? What extra information is
required? While we are still far from giving definitive answers to these
questions, we believe that what we accomplish here is a helpful start.
There is an important feature of the discs D(0, 1, σ) that plays a central
roll in our theory: For any two normal representations of M, σ and τ ,
(1.2)
D(0, 1, σ) ⊕ D(0, 1, τ ) ⊆ D(0, 1, σ ⊕ τ ).
z12
The meaning of this inclusion is easy to understand when one realizes
that I((σ ⊕ τ )E ◦ ϕ, σ ⊕ τ ) may be viewed as a set of operator matrices
(cid:20)z11
z22(cid:21) acting as operators from HσE ◦ϕ ⊕ Hτ E◦ϕ to Hσ ⊕ Hτ , where
z21
z11 ∈ I(σE ◦ ϕ, σ), z12 ∈ I(τ E ◦ ϕ, σ), z21 ∈ I(σE ◦ ϕ, τ ), and z22 ∈
I(τ E ◦ ϕ, τ ). So D(0, 1, σ) ⊕ D(0, 1, τ ) is just the collection of those
matrices (cid:20)z11
z22(cid:21) ∈ D(0, 1, σ ⊕ τ ) in which the off-diagonal entries
z12
z21
vanish.
TENSORIAL FUNCTION THEORY
5
Definition 1.1. A family of sets {U(σ)}σ∈N Rep(M ), with U(σ) ⊆ I(σE◦
ϕ, σ), satisfying U(σ)⊕ U(τ ) ⊆ U(σ ⊕ τ ) is called a matricial family of
sets.
Matricial families, in particular, families of discs {D(0, 1, σ)}σ∈N Rep(M ),
enjoy properties that are very similar to the properties of the domains
that Taylor first considered in [Tay72, Section 6], when he introduced
a notion of localization for free algebras. They are very closely related
to the fully matricial sets of Dan Voiculescu [Voi04, Voi10], which are
matricial sets in our terminology but also satisfy additional conditions
which we do not use here. In contexts when E is finite dimensional,
matricial sets are essentially the noncommutative sets that Bill Helton,
Igor Klep, Scott McCullough and others study in the setting of lin-
ear matrix inequalities [HKM11b, HKM11a, HKMS09]. They are also
closely connected to the noncommutative sets in the work of Dmitry
Kaliuzhnyi-Verbovetskyi and Victor Vinnikov devoted to noncommu-
tative function theory [KVV09, KV10, KVV].
There is another property that the discs also enjoy, viz. for any con-
traction t ∈ I(τ, σ), the inclusion,
(1.3)
tD(0, 1, τ )(IE ⊗ t∗) ⊆ D(0, 1, σ),
holds. This shows that the discs are matricially convex in the sense
of operator space theory (see [ER00]). While matricial convexity does
not play a role in our immediate considerations, it already has proved
useful elsewhere.
satisfies the equation
The Berezin transform, bF = {bFσ}σ∈N Rep(M ), of an element F ∈ H ∞(E)
(1.4) bFσ⊕τ (z⊕ w) = bFσ(z)⊕bFτ (w),
z⊕ w ∈ D(0, 1, σ)⊕ D(0, 1, τ ).
This, too, is a critical feature of the functions in Taylor's theory and
in the other places just cited. In fact, the Berezin transforms have an
additional property that we will use repeatedly:
Definition 1.2. Suppose {U(σ)}σ∈N Rep(M ) is a matricial family of sets
and suppose that for each σ ∈ NRep(M), fσ : U(σ) → B(Hσ) is a
function. We say that f := {fσ}σ∈N Rep(M ) is a matricial family of
functions in case
(1.5)
for every z ∈ U(σ), every w ∈ U(τ ) and every C ∈ I(σ, τ ) such that
(1.6)
Cfσ(z) = fτ (w)C
Cz = w(IE ⊗ C).
TENSORIAL FUNCTION THEORY
6
When the family {U(σ)}σ∈N Rep(M ) is {D(0, 1, σ)}σ∈N Rep(M ), and f =
{fσ}σ∈N Rep(M ) is a Berezin transform, then it is easy to see that the
assumptions on an operator C : Hσ → Hτ that C ∈ I(σ, τ ) and satisfies
equation (1.6) express the fact that C lies in I(σ × z, τ × w). But
then, equation (1.5) is immediate. It is simply a manifestation of the
structure of the commutant of the representation (σ ⊕ τ ) × (z⊕ w). In
this setting also, the defining hypothesis for a matricial family can be
written simply as
I(σ × z, τ × w) ⊆ I(fσ(z), fτ (w)),
(1.7)
for all σ, τ ∈ NRep(M), z ∈ AC(σ), and w ∈ AC(τ ). Consequently, we
sometimes say that a matricial family respects intertwiners. Observe
that if a family respects intertwiners, then it automatically satisfies
equations like (1.4).
What is surprising is the following converse - a nonlinear double com-
mutant theorem of sorts - which extends the double commutant the-
orem for induced representations of Hardy algebras [MS04, Corollary
3.10]. We shall prove a more refined statement in Theorem 4.4.
Theorem 1.3. If f = {fσ}σ∈N Rep(M ) is a matricial family of functions,
with fσ defined on AC(σ) and mapping to B(Hσ), then there is an
F ∈ H ∞(E) such that f is the Berezin transform of F , i.e., fσ = bFσ
for every σ.
Three features of Theorem 1.3 deserve comment: (1) the domain of
each fσ is AC(σ), (2) all the representations in NRep(M) are used,
and (3) no hypotheses on the nature of the functions fσ are made.
They are not assumed to be analytic or continuous; they are not even
assumed to be bounded.
If we relax (1) and assume only that the
functions are defined on D(0, 1, σ), then the theorem breaks down: In
Example 4.5 we will exhibit an unbounded family f = {fσ}σ∈N Rep(M ),
where fσ is defined only on D(0, 1, σ), that satisfies (1.7). However,
we shall also prove that if the fσ : D(0, 1, σ) → B(Hσ) are bounded
uniformly in σ ∈ NRep(M) and if they respect intertwiners, then f is
a Berezin transform (Theorem 4.8).
With regard to (2), it is helpful to reflect on the fact that NRep(M) is
really a W ∗-category in the sense of [GLR85]. In fact, it is essentially
Riefel's category Normod − M [Rie74b]. We do not need much of the
theory of such categories to achieve our goals here, but that theory has
guided our thinking. The objects of NRep(M) are the normal repre-
sentations of M on separable Hilbert space. (We have abused notation
a bit and have simply written σ ∈ NRep(M), when σ is a normal rep-
resentation of M.) The set of morphisms from σ to τ is the intertwiner
TENSORIAL FUNCTION THEORY
7
space I(σ, τ ). We want to consider subcategories Σ of NRep(M) with
the property that if σ and τ are in Σ, then so is σ ⊕ τ . Such a category
is an additive subcategory of NRep(M). In the literature, particularly
that dealing with the setting where M = C and E = Cd, it is im-
portant to extract as much information as is possible from the finite
dimensional representations of M, and, of course, the finite dimensional
representations of M determine an additive subcategory of NRep(M).
Although, we do not know how to modify the hypotheses in Theorem
1.3 so that we can restrict to every additive subcategory, we can prove
that if Σ is a full additive subcategory of NRep(M) that consists of
faithful representations of M, if f = {fσ}σ∈Σ is a uniformly bounded
matricial family of functions, with fσ : D(0, 1, σ) → B(Hσ), then each
of the fσs is Frechet holomorphic in D(0, 1, σ) and the Frechet-Taylor
expansion of fσ can be expressed explicitly in terms of tensors from the
tensor powers of E in the same way that the Taylor series for Berezin
transforms can be expressed, i.e., in terms of our tensorial power series.
But to do this, we must first use ideas from Kaliuzhnyi-Verbovetskyi
and Vinnikov [KVV09, KV10, KVV] to show that the fσs have Taylor
series in the sense of Joseph Taylor [Tay73] and then use a variant of
our duality theorem [MS04, Theorem 3.6] to show that Taylor's Taylor
coefficients are tensors of the desired type. Thus, we will show that
there are strong interconnections among three notions of holomorphy
and their associated power series: Frechet holomorphy for maps be-
tween Banach spaces, Taylor's notion of free holomorphy, and tensorial
holomorphy. We were led to these by studying Berezin transforms of
elements of H ∞(E). Although Theorem 4.8 shows that a matricial
family of functions f = {fσ}σ∈N Rep(M ), with fσ : D(0, 1, σ) → B(Hσ),
is a Berezin transform if and only if f is uniformly bounded in σ,
the proof very much depends upon properties of NRep(M) that are
not shared by all subcategories. The best we can say at this stage
is that if Σ is a full additive subcategory of NRep(M) consisting of
faithful representations, if f = {fσ}σ∈Σ is a family of functions with
fσ : D(0, 1, σ) → B(Hσ), and if supσ∈Σ supz∈D(0,1,σ) kfσ(z)k < ∞, then
f is a family of tensorial power series if and only if f is a matricial
family (Theorem 5.1). Additional information appears to be needed to
conclude that such an f is a Berezin transform.
Section 2 is dedicated to recapping a number of facts we need from the
theory of tensor algebras and Hardy algebras. Tensorial power series
are also introduced and some of their properties developed. Section 3 is
devoted to developing general properties of matricial sets and functions.
Section 4 is devoted to proving a refined version of Theorem 1.3. The
TENSORIAL FUNCTION THEORY
8
connection between Taylor's Taylor series and our Tensorial Taylor
series is made in Section 5. Finally in Section 6 we extend our work in
Section 5 to cover maps between correspondence duals.
Acknowledgement. We are pleased to thank Victor Vinnikov for en-
lightening conversations we had with him over matricial function the-
ory. His work in this area, especially that with Dmitry Kaliuzhnyi-
Verbovetskyi and Mihai Popa, has been a source of inspiration to us.
2. Preliminaries
Throughout, M will be a fixed W ∗-algebra with separable predual.
However, we shall not fix a representation in advance, but instead we
shall want to access the entire category of normal representations of M
on separable Hilbert space, NRep(M). We shall always assume that
such representations are unital. Strictly speaking, we shall not want to
identify unitarily equivalent representations. But when M = C, there
is no harm in doing so. Also, throughout this paper, Σ will denote an
additive subcategory of NRep(M). This means that whenever σ and τ
are in Σ so is σ⊕τ , where Hσ⊕τ := Hσ⊕Hτ and σ⊕τ (a) := σ(a)⊕τ (a).
In particular, we shall be interested in additive subcategory generated
by a single representation σ. It consists of all the finite multiples of σ,
which we write nσ. When we want to consider an infinite multiple of
σ, we will be explicit about this and denote it by ∞σ. Only at certain
points will we need to assume that Σ is a full subcategory of NRep(M),
meaning that when σ and τ are in Σ, then all the intertwiners between
σ and τ are also in Σ. We will be explicit about where this assumption
is used. We will, however, always assume that if σ lies in Σ and if
τ = mσ, then the natural injection ιk of Hσ into Hτ , that identifies Hσ
with the kth summand of Hτ , is a morphism in Σ, and so is its adjoint,
ι∗
k, which is the projection of Hτ onto Hσ.
We shall fix a W ∗-correspondence E over M. Thus, E is a self-dual
right Hilbert module over M that is endowed with a left action given by
a faithful normal representation ϕ of M in the algebra of all continuous
linear right module maps on E, L(E). (Recall that because E is a self-
dual right Hilbert module over M, every bounded right module map on
E is automatically adjointable.) As with Hilbert space representations,
we shall assume normal homomorphisms between W ∗-algebras are uni-
tal and, in particular, we shall assume that ϕ is unital. Like M, E is a
dual space. We shall refer to the weak-∗ topology on a W ∗-algebra or
on a W ∗-correspondence as the ultraweak topology [MS11b, Paragraph
2.2].
TENSORIAL FUNCTION THEORY
9
The k-fold tensor powers of E, balanced over M (via ϕ), will be denoted
E⊗k. Recall that E⊗k is the self-dual completion of the C ∗-tensor power
and so is a W ∗-correspondence over M. The left action of M on E⊗k
will be denoted by ϕk, i.e., ϕk(a)(ξ1 ⊗ ξ2 ⊗ · · · ⊗ ξk) := (ϕ(a)ξ1) ⊗ ξ2 ⊗
· · · ⊗ ξk. In particular, ϕ0 is just left multiplication of M on M and
ϕ1 = ϕ [MS11b, Paragraph 2.7].
There are a number of spaces and algebras that we will want to build
from the tensor powers E⊗k. First, there is the Fock space F (E) :=
M ⊕ E ⊕ E⊗2 ⊕ E⊗3 ⊕ · · · Here we interpret the sum as the self-dual
completion of the Hilbert C ∗-module direct sum. It is a natural W ∗-
correspondence over M, with the left action ϕ∞ := ϕ0 ⊕ ϕ1 ⊕ ϕ2 ⊕ · · ·
[MS11b, Paragraph 2.7]. For ξ ∈ E, we write Tξ for the creation
operator determined by ξ. By definition, Tξη := ξ ⊗ η, η ∈ F (E).
It is easy to check that Tξ is bounded, with norm dominated by kξk.
Since we are assuming ϕ is injective and unital, it follows that kTξk =
kξk. The norm-closed subalgebra of L(F (E)) generated by ϕ∞(M)
and {Tξ}ξ∈E will be denoted T+(E) and called the tensor algebra of
E. The Hardy algebra of E is defined to be the ultraweak closure of
T+(E) in L(F (E)). The (nonclosed) subalgebra of T+(E) generated by
ϕ∞(M) and{Tξ}ξ∈E will be denoted by T0(E) and called the algebraic
tensor algebra of E. In the purely algebraic setting it would simply be
called the tensor algebra of E.3
Before getting too far along with the theory, it will be helpful to have
an example that we can follow as the theory develops.
Example 2.1. (The Basic Example) In this example we let M =
C and we let E = Cd. The left and right actions coincide and the
inner product is the usual one, except we choose it to be conjugate
linear in the right hand variable. We interpret Cd as ℓ2(N), when
d = ∞. The representations of M on Hilbert space, in this case, are
all multiples of the identity representation σ1 which represents C on
C via multiplication, i.e., we may view the objects in NRep(M) as
{nσ1 n ∈ N ∪ {∞}}. For m, n ∈ N ∪ {∞}, the intertwiner space
I(mσ1, nσ1) may be identified with the n × m matrices, where when
either m or n is infinite, we interpret the n × m matrices as bounded
linear operators. To keep the notation simple, when it is convenient we
shall write Mn(C) for B(ℓ2(N)) when n = ∞. An additive subcategory
of NRep(M) is determined by any additive subsemigroup of N ∪ {∞}.
3Strictly speaking, because our coefficient algebra M may be noncommutative,
in the algebra literature T0(E) would be called the tensor M -ring determined by
the bimodule E. (See, e.g., [Coh06, P. 134].)
TENSORIAL FUNCTION THEORY
10
It is customary to identify F (Cd) with ℓ2(F+
The Fock space F (E) is usually called the full Fock space of the Hilbert
d ), where F+
space Cd.
d
denotes the free semigroup on d generators. The identification is carried
out explicitly by letting {ei 1 ≤ i ≤ d} denote the standard basis for
Cd and then sending the decomposable tensor ei1 ⊗ ei2 ⊗ · · · ⊗ eik ∈
(Cd)⊗k ⊆ F (Cd) to δw, where δw is the point mass at the word w =
i1i2 · · · ik. The vacuum vector Ω in F (Cd) is sent to the empty word
∅ in F+
d . When this identification is made, then the creation operator
Tei is identified with the Si, where Si is the isometry defined by the
equation
Siδw = δiw,
w ∈ F+
d .
Also under this identification, the algebraic tensor algebra T0(Cd) is
identified with the free algebra on d generators, ChX1, X2,· · · , Xdi,
realized as convolution operators on ℓ2(F+
d ) through the correspondence
Xi ↔ Si. Of course, T+(Cd) may be viewed as the norm closed, unital
algebra generated by the Si, and, as we mentioned in the introduction,
in this guise it is known as Popescu's noncommutative disc algebra
Its ultraweak closure, H ∞(E), which in this case coincides
[Pop96].
with its weak operator closure, is the noncommutative analytic Toeplitz
algebra.
Definition 2.2. The full Cartesian product Πk≥0E⊗k will be called
the formal tensor series algebra determined by M and E and will be
denoted by T+((E)). Elements in T+((E)) will be denoted by formal
sums, θ ∼Pk≥0 θk, θk ∈ E⊗k.
It is easy to see that T+((E)) is, indeed, an algebra under coordinate-
wise addition and scalar multiplication, with the product given by the
formula
where θ ∼ Pk≥0 θk, η ∼ Pk≥0 ηk, and ζ ∼ Pk ζk are related by the
equation
θ ∗ η = ζ,
ζk := Xk=l+m
θl ⊗ ηm,
k, l, m ≥ 0.
Further, we may view T0(E), T+(E) and H ∞(E) as subsets of T+((E))
using the "Fourier coefficient operators" calculated with respect to the
gauge automorphism group acting on L(F (E)) [MS11b, Paragraph
2.9]. That is, if Pn is the projection of F (E) onto E⊗n and if
Wt :=
eintPn,
∞Xn=0
TENSORIAL FUNCTION THEORY
11
then {Wt}t∈R is a one-parameter (2π-periodic) unitary group in L(F (E))
such that if {γt}t∈R is defined by the formula γt = Ad(Wt), then {γt}t∈R
is an ultraweakly continuous group of ∗-automorphisms of L(F (E)),
called the gauge automorphism group of L(F (E)). This group leaves
each of the algebras T0(E), T+(E), and H ∞(E) invariant and so do
each of the Fourier coefficient operators {Φj}j∈Z defined on L(F (E))
by the formula
(2.1)
Φj(F ) :=
1
2π 2π
0
e−intγt(F ) dt,
F ∈ L(F (E)),
where the integral converges in the ultraweak topology. Further, if
F ∈ H ∞(E), then Φj(F ) is of the form Tθj , where θj ∈ E⊗j. This
is clear if F ∈ T0(E). For a general element F ∈ H ∞(E), the result
follows from the facts that T0(E) is ultraweakly dense in H ∞(E) and
Φj is ultraweakly continuous. In fact, if Σk is defined by the formula
(2.2)
Σk(F ) := Xj<k
(1 − j
k
)Φj(F ),
F ∈ L(F (E)), then for F ∈ L(F (E)), limk→∞ Σk(F ) = F , in the ultra-
weak topology. It follows that for F ∈ H ∞(E), F determines and is
uniquely determined by its sequence of Fourier coefficients θ = {θj}j≥0,
where Tθj := Φj(F ).
The correspondence E determines an endo-functor ΦE from NRep(M)
into NRep(M). For an object σ in NRep(M), ΦE(σ) = σE ◦ ϕ, where
σE is the representation of L(E) that is induced from σ in the sense of
Rieffel [Rie74a]: For X ∈ L(E), σE(X) = X ⊗ IHσ acting on E ⊗σ Hσ.
It is [Rie74a, Theorem 5.3] that ensures that for T ∈ I(σ, τ ), ΦE(T )
should be IE ⊗ T ∈ I(ΦE(σ), ΦE(τ )).
Following the notation in [MS11b, Paragraph 2.6] and elsewhere in our
work, we set Eσ∗ = I(σE ◦ ϕ, σ) for each σ ∈ NRep(M). Also, for
z0 ∈ Eσ∗ and 0 ≤ R ≤ ∞, we write D(z0, R, σ, E) for the open disc or
ball in Eσ∗ of radius R centered at z0. Usually, E will be dropped from
the notation, since it will be understood from context.
The space Eσ∗ is a bimodule over σ(M)′, where for a, b ∈ σ(M)′ and
ξ ∈ Eσ∗, a · ξ · b := aξ(IE ⊗ b). In fact, it is a left Hilbert module over
σ(M)′, where the inner product is given by the formula hξ, ηi = ξη∗
and as such, Eσ∗ becomes a correspondence over σ(M)′. Although
the correspondence properties of Eσ∗ have figured heavily elsewhere in
our work, they do not play much of a role here. Only the bimodule
(In Section 6 we will
properties are of consequence in this section.
TENSORIAL FUNCTION THEORY
12
use properties of the correspondence Eσ, which is a right Hilbert W ∗-
module over σ(M)′.)
Once σ is fixed, each z ∈ Eσ∗ determines a homomorphism, denoted
σ×z, from T0(E) to B(Hσ), which is defined on the generators of T0(E)
via the formulas
σ × z(ϕ∞(a)) = σ(a),
a ∈ M,
σ × z(Tξ)(h) = z(ξ ⊗ h),
ξ ∈ E, h ∈ Hσ.
The fact that these equations define a representation of T0(E) on
B(Hσ) follows from the universal properties of tensor algebras once
it is checked that the map ξ → σ × z(Tξ) is a bimodule map in the
sense that
σ × z(Ta·ξ·b) = σ(a)(σ × z(Tξ))σ(b),
a, b ∈ M, ξ ∈ E.
And conversely, every representation ρ of T0(E) on Hilbert space such
that the restriction of ρ to ϕ∞(M) in T0(E) is a normal representation
of M, say σ, and the restriction of ρ to the subspace {Tξ}ξ∈E yields an
ultraweakly continuous bimodule map of E, must be of the form σ × z,
where z is given by the formula z(ξ ⊗ h) = ρ(Tξ)h.
As we indicated in the introduction, it is of fundamental importance for
our theory, that σ × z extends from T0(E) to a completely contractive
representation of T+(E) on B(Hσ) if and only if z lies in the norm closed
disc D(0, 1, σ) [MS11b, Theorem 2.9]. Moreover, for kzk < 1, σ × z
extends further to be an ultraweakly continuous, completely contractive
representation of H ∞(E) on Hσ [MS04, Corollary 2.14]. We write
AC(σ, E), or simply AC(σ), when E is understood, for the collection of
all z ∈ D(0, 1, σ) such that σ × z extends to an ultraweakly continuous,
completely contractive representation of H ∞(E) on Hσ. We then call
z an absolutely continuous point in D(0, 1, σ) and we say that σ × z is
an absolutely continuous representation. The terminology derives from
the fact that one can define an H ∞-functional calculus in the sense of
Sz.-Nagy and Foiaş for a contraction operator if and only if its minimal
unitary dilation is absolutely continuous.
(See [MS11b], where this
perspective is developed at length.)
Remark 2.3. There is now a bit of ambiguity that needs to be clarified.
When we write I(σ × z, τ × w) we shall always mean the set of all
operators C : Hσ → Hτ such that C ∈ I(σ, τ ) and Cz = w(IE ⊗
C). It is straightforward to see that this happens if and only if C ∈
I(σ, τ ) and Cσ × z(F ) = τ × w(F )C for all F ∈ T0(E). Since T0(E) is
norm dense in T+(E) and since σ × z and τ × w extend to continuous
TENSORIAL FUNCTION THEORY
13
representations of T+(E) when kzk,kwk ≤ 1, it follows that I(σ×z, τ ×
w) = {C : Hσ → Hτ Cσ × z(F ) = τ × w(F )C for all F ∈ T+(E)},
when kzk,kwk ≤ 1. Likewise, since AC(σ) is the set of all z ∈ Eσ∗ such
that σ×z extends from T+(E) to an ultraweakly continuous completely
contractive representation of H ∞(E), we find that I(σ × z, τ × w) =
{C : Hσ → Hτ Cσ × z(F ) = τ × w(F )C for all F ∈ H ∞(E)} when
z ∈ AC(σ) and w ∈ AC(τ ).
With the notation we have established, it follows that T0(E) may be
represented as a space polynomial-like, B(Hσ)-valued functions on Eσ∗
via the formula:
(2.3)
bFσ(z) := σ × z(F ),
F ∈ T0(E), z ∈ Eσ∗.
Since our primary objective is to understand the nature of the functions
bFσ, it will be helpful to work carefully through some examples that
illuminate their definition. Consider first the case when F = Tξ1Tξ2 =
Tξ1⊗ξ2. Then for h ∈ Hσ, we have
(2.4)
(σ × z)(Tξ1Tξ2)(h) = (σ × z)(Tξ1)(σ × z)(Tξ2)(h)
= z(ξ1 ⊗ z(ξ2 ⊗ h))
= z(IE ⊗ z)(ξ1 ⊗ ξ2 ⊗ h).
More generally, we find that
(σ× z)(Tξ1⊗ξ2⊗···⊗ξk)(h) = z(IE ⊗ z)· · · (IE⊗k−1 ⊗ z)(ξ1⊗ξ2⊗· · ·⊗ξk⊗h).
Observe that z(IE ⊗ z)· · · (IE⊗k−1 ⊗ z) is a bounded linear operator
from E⊗k ⊗ Hσ to Hσ, with norm dominated by kzkk. Further, z(IE ⊗
z)· · · (IE⊗k−1⊗z) intertwines σE⊗k◦ϕk and σ. Thus z(IE⊗z)· · · (IE⊗k−1⊗
z) is an element of I(σE⊗k ◦ ϕk, σ) = (E⊗k)σ∗. In [MS04, Page 363]
z(IE ⊗ z)· · · (IE⊗k−1 ⊗ z) is called the kth generalized power of z. We
shall denote it here by z(k) and we shall denote the function z → z(k),
from Eσ∗ to (E⊗k)σ∗, by Zk, i.e., Zk(z) := z(k) = z(IE⊗z)· · · (IE⊗k−1⊗z).
By convention, we set Z0(z) ≡ IHσ .
Example 2.4. In the context of our basic example, Example 2.1, E⊗σ
Hσ is Cd ⊗ Hσ, which we shall view simply as columns of length d
of vectors from Hσ. We may do this, since there is no consequential
balancing going on. After all, σ is a representation of C. It follows that
Eσ∗ := I(σE ◦ ϕ, σ) may be identified with all d-tuples of operators in
B(Hσ) arranged as a row. That is, we may and shall write Eσ∗ = {z =
(Z1, Z2,· · · , Zd) Zi ∈ B(Hσ)}. So, if F ∈ T0(E) and if we identify
T0(E) with the free algebra on d generators, as we did in our basic
TENSORIAL FUNCTION THEORY
14
becomes
example, we may write F = Pw∈F+
bFσ(z) = Xw∈F+
d
d
awZw,
awXw, and then equation (2.3)
where Zw = Zi1Zi2 · · · Zik, w = i1i2 · · · ik. In particular, if σ happens
to be an n-dimensional representation with n < ∞, then Eσ∗ is the
collection of all d-tuples of n × n matrices and equation (2.3) yields
the representation of the free algebra on d generators on the algebra
of d generic n × n matrices. The function Zk in this setting assigns
to z = (Z1, Z2,· · · , Zd) the tuple (Zw)w=k of length dk that is ordered
lexicographically by the words of length k.
In the following lemma and, indeed, throughout this paper, we identify
E⊗(k+l) with E⊗k ⊗ E⊗l; E⊗0 := M.
Lemma 2.5. The functions Zk, from Eσ∗ to (E⊗k)σ∗, satisfy the equa-
tion
Zk+l(z) = Zk(z)(IE⊗k ⊗ Zl(z)),
z ∈ Eσ∗.
Further, each Zk is Frechet differentiable and the Frechet derivative of
Zk, denoted DZk, is given by the formula
(2.5)
(DZk)(z)[ζ] =
k−1Xl=0
Zl(z)(IE⊗l ⊗ ζ)(IE⊗(l+1) ⊗ Zk−l−1(z))
for all z, ζ ∈ Eσ∗.
The proof is a straightforward calculation based on the definition of
Zk and will be omitted. Observe that for each z ∈ Eσ∗, DZk(z) is a
bounded linear operator from Eσ∗ to (E⊗k)σ∗ whose norm is dominated
by kkzkk−1.
If θ ∈ E⊗k, we shall write Lθ for the linear operator from Hσ to E⊗k⊗σ
Hσ that maps h to θ⊗ h. Evidently, kLθk ≤ kθk, with equality holding
if σ is faithful. The calculations that led from equation (2.4) through
Lemma 2.5 immediately yield
Theorem 2.6. Suppose F =Pn
k=0 Tθk ∈ T0(E), with θk ∈ E⊗k. Then
for σ ∈ NRep(M) and z ∈ Eσ∗, the operator bFσ(z) in B(Hσ) is given
by the formula
(2.6)
bFσ(z) =
nXk=0
Zk(z)Lθk.
TENSORIAL FUNCTION THEORY
15
Further, as a function of z ∈ Eσ∗, bFσ is Frechet differentiable and its
Frechet derivative, mapping Eσ∗ to B(Hσ), is given by the formula
(2.7)
DbFσ(z)[ζ] =
nXk=0
DZk(z)[ζ]Lθk.
Of course, we would like to extend our function theory beyond the realm
of the algebraic tensor algebra. For this purpose, we follow Popescu
[Pop06] (who followed Hadamard, who followed Cauchy) and introduce
the following definition.
Definition 2.7. For θ ∼Pk≥0 θk in T+((E)), we define R(θ) to be
1
R(θ) = (lim sup
k→∞ kθkk
k )−1,
and we call R(θ) the radius of convergence of the formal series θ.
Of course, R(θ) is a non-negative real number or +∞.
Proposition 2.8. Suppose θ ∼ Pk≥0 θk in T+((E)) and R = R(θ) is
not zero. Then for each σ ∈ NRep(M), for each z0 ∈ Eσ∗, and for
each z ∈ D(z0, R, σ), the series
∞Xk=0
(2.8)
Zk(z − z0)Lθk
converges in the norm of B(Hσ). The convergence is uniform on any
subdisk D(z0, ρ, σ), ρ < R; and it is uniform in σ, as well. The resulting
function, fσ, is analytic as a map from D(z0, R, σ) to B(Hσ) and the
Frechet derivative of fσ is given by the formula
Dfσ(z)[ζ] =
∞Xk=1
DZk(z − z0)[ζ]Lθk,
z ∈ D(z0, R, σ), ζ ∈ Eσ∗.
1
Proof. The uniform convergence of (2.8) follows a standard argument:
R(θ) = lim supk kθkk1/k, there
Fix ρ, ρ′ such that 0 < ρ < ρ′ < R. Since
is an m such that, for all k ≥ m, kθkk1/k < 1/ρ′. For such k, for all
σ and for all z ∈ D(z0, ρ, σ), kZk(z − z0)Lθkk ≤ kz − z0kkkθkk ≤ ( ρ
ρ′ )k.
Since ρ < ρ′, we see that (2.8) converges uniformly on D(z0, ρ, σ). The
remaining assertions can be proved by similar elementary estimates.
Alternatively, one can appeal to [HP74, Theorems 3.17.1 and 3.18.1].
(cid:3)
Definition 2.9. The series (2.8) is called the tensorial power series
determined by the formal tensor series θ, the point z0, and the repre-
sentation σ.
TENSORIAL FUNCTION THEORY
16
Remark 2.10. Following the line of thought developed in elementary
texts on complex function theory, one might expect that if z ∈ Eσ∗ has
norm larger than R(θ), then the series (2.8) diverges. However, the
situation is quite a bit more complicated than in one complex variable.
For example, if M = E = C and if dim Hσ ≥ 2, then we may identify
Eσ∗ with B(Hσ), and if z is any nilpotent element in Eσ∗, the series
(2.8) converges no matter what R(θ) is. The situation is even more
complicated: There are series θ with finite R(θ) such that for "most"
σ the series (2.8) converges on all of Eσ∗, i.e., almost all fσ are entire.
A particularly instructive example, due to Luminet [Lum86, Exam-
ple 2.9], is constructed as follows. Let Sk(X1, X2,· · · , Xk) denote the
standard identity in k noncommuting variables:
Sk(X1, X2,· · · , Xk) = Xs∈Sk
sign(s)Xs(1)Xs(2) · · · Xs(k),
where Sk denotes the permutation group on k letters and where sign(s)
is 1 if s is even and −1 if s is odd. Then the set of these identities
determines an element θ in T+((C2)) by identifying T+((C2)) with the
free formal series in two noncommuting variables X1 and X2 and setting
θ(X1, X2) =Xk≥2
Sk(X1, X1X2,· · · , X1X k−1
2
).
When this series is written as a series of words in X1 and X2,
θ(X1, X2) = Xw∈F+
2
λwXw,
one sees that for each positive integer d there is a word w of length at
least d so that λw = 1. It follows that R(θ) < ∞. However, if σ is any
finite dimensional representation of C = M, the series (2.8) converges
throughout Eσ∗. Indeed, for any given z there are only finitely many
nonzero terms in the series. On the other hand, we will be able to show
later that given R′ > R(θ), there is a σ ∈ NRep(M) and an element
z ∈ Eσ∗ with kzk = R′ such that the series (2.8) diverges.
The next proposition establishes a bridge between the theory developed
in [MS04, MS05, MS08, MS09] and the focus of this paper. First, recall
that if F ∈ H ∞(E), then the Fourier coefficients Φj(F ) are of the form
Tθj for θj ∈ E⊗j. Further, the series Pj≥0 Tθj is Cesaro summable to
F (2.2). On the other hand, each Fourier coefficient operator Φj is
contractive and so the norm of each θj is dominated by kFk. Thus the
radius of convergence of the series θ ∼ Pj≥0 θj is at least one. Since
TENSORIAL FUNCTION THEORY
17
σ × z(Tθ) = Zj(z)Lθ for θ ∈ E⊗j and z ∈ D(0, 1, σ), the validity of the
following proposition is immediate.
Proposition 2.11. For F ∈ H ∞(E) and σ ∈ Σ, the σ-Berezin trans-
form bFσ, defined on AC(σ) by the formula, bFσ(z) := σ × z(F ), admits
the tensorial power series expansion,
(2.9)
∞Xj=0
bFσ(z) =
Zj(z)Lθj ,
where Tθj = Φj(F ). The series converges in norm for z ∈ D(0, 1, σ),
uniformly in σ and on any sub-disk, D(0, ρ, σ), ρ < 1. Moreover, the
function bFσ(z) is bounded by kFk throughout AC(σ).
Remark 2.12. In view of Proposition 2.11, one might wonder if every
bounded analytic function on D(0, 1, σ), with tensorial power series
P∞
j=0 Zj(z)Lθj , comes from a function in H ∞(E). Thanks to a very
detailed study by Arveson, such is not the case. He shows that in one
of the simplest settings, when M = C, E = Cd (d ≥ 2), and σ is one-
dimensional, there are bounded analytic functions on D(0, 1, σ) (which
is the unit ball in Cd) which are not of the form bFσ(z) for any element
F in H ∞(Cd) (see [Arv98, Theorem 3.3]).
3. Matricial Families and Functions
0
0
z21
0
z12
Suppose σ and τ are two normal representations of our W ∗-algebra
M on Hilbert spaces Hσ and Hτ , respectively. We have noted that
if one writes σ ⊕ τ matricially as (σ ⊕ τ )(·) = (cid:20)σ(·)
τ (·)(cid:21), then
E(σ⊕τ )∗ = I((σ ⊕ τ )E ◦ ϕ, (σ ⊕ τ )) may be written as matrices of oper-
ators(cid:20)z11
z22(cid:21), viewed as operators from HσE ◦ϕ ⊕ Hτ E ◦ϕ to Hσ ⊕ Hτ ,
where z11 ∈ I(σE◦ϕ, σ), z12 ∈ I(τ E◦ϕ, σ), z21 ∈ I(σE◦ϕ, τ ), and where
z22 ∈ I(τ E ◦ ϕ, τ ). In particular, note that all the matrices of the form
(cid:20)z11
z22(cid:21), with z11 ∈ I(σE ◦ ϕ, σ) and z22 ∈ I(τ E ◦ ϕ, τ ) are contained
0
E(σ⊕τ )∗. We abbreviate this fact by writing Eσ∗ ⊕ Eτ ∗ ⊆ E(σ⊕τ )∗. We
will have occasion later to think about higher order matrices. That
is, if σ1, σ2,· · · , σn are n (not-necessarily distinct) representations in
NRep(M), then E(σ1⊕σ2⊕···⊕σn)∗ can be viewed as n× n matrices whose
i, j-entries lie in I(σE
j ◦ ϕ, σi). When this is done, we will be able to
write P⊕
1≤i≤n Eσi∗ ⊆ E(σ1⊕σ2⊕···⊕σn)∗.
TENSORIAL FUNCTION THEORY
18
Remark 3.1. In Definition 1.1, we defined the notion of a matricial
family of sets indexed by all of NRep(M). Evidently, that notion
can be "relativized" to additive subcategories. So if Σ is an additive
subcategory of NRep(M), then a family of sets U = {U(σ)}σ∈Σ that
satisfies the following two properties will be called a matricial E, Σ-
family of sets. If Σ or E are clear from context, we shall simply call U
a matricial family of sets.
sums, i.e.,
(1) Each U(σ) is contained in Eσ∗.
(2) The family {U(σ)}σ∈Σ is closed with respect to taking direct
U(σ) ⊕ U(τ ) :=(cid:20)U(σ)
U(τ )(cid:21) ⊆ U(σ ⊕ τ ).
0
0
If each of the members of U is described by a common property, such
as being open or a domain, then we shall adjust the terminology ap-
propriately, e.g., by saying that the family is a matricial E, Σ-family of
open sets or domains. We shall say that U unitarily invariant, if each
U(σ) is unitarily invariant in the sense that for each unitary operator
u ∈ σ(M)′ and each z ∈ U(σ), u−1·z·u := u−1z(IE⊗u) lies in U(σ). Re-
latedly, U is called matricially convex, if for any σ and τ in Σ and for all
v ∈ I(σ, τ ) such that vv∗ = IHτ , v ·U(σ)· v∗ = vU(σ)(IE ⊗ v∗) ⊆ U(τ ).
Example 3.2. We already have noted that {D(0, 1, σ)}σ∈N Rep(M ) is a
matricially convex, matricial family of sets. However, the best we can
say, in general, is that {AC(σ)}σ∈N Rep(M ) is unitarily invariant. For a
more general class of discs, suppose that {ζσ}σ∈Σ is a family of vectors
with ζσ ∈ Eσ∗ such that ζσ⊕τ = ζσ ⊕ ζτ for all σ, τ ∈ Σ. We shall call
ζ := {ζσ}σ∈Σ an additive field of vectors over Σ. Given such a field ζ
and an R, 0 ≤ R ≤ ∞, we shall call D(ζ, R) := {D(ζσ, R, σ)}σ∈Σ the
matricial disc determined by the field ζ. If ζσ = 0 for all σ ∈ Σ, we
shall simply write D(0, R), calling it the matricial disc {D(0, R, σ)}σ∈Σ
we already have defined. It is clear that D(ζ, R) is a matricial set when
ζ is an additive field of vectors over Σ, but in general D(ζ, R) won't
be matricially convex. However, it will be matricially convex when
ζ = {ζσ}σ∈Σ is a central additive field in the sense that ζτ (IE ⊗ C) =
Cζσ for all C ∈ I(σ, τ ), σ, τ ∈ Σ, as is evident from the definition.
Note that for ζ to be a central additive field requires more than each
ζσ being central in the sense of [MS08, Definition 4.11], which simply
means that ζσ(IE ⊗ C) = Cζσ for all C ∈ I(σ, σ) = σ(M)′. However,
if each ζσ is central in that sense, then D(ζ, R) is a unitarily invariant
matricial set. Note, too, that if Σ is the subcategory consisting of
all multiples of a single representation σ, together with all the spaces
TENSORIAL FUNCTION THEORY
19
I(nσ, mσ), n, m ∈ N, and if ζσ is a central vector in Eσ∗, then {ζmσ}m∈N
is a central additive field on Σ, where ζmσ is the m-fold direct sum of
copies of ζσ.
To see these examples in the simplest, most concrete setting, consider
the following addition to our basic example, Example 2.1.
Example 3.3. Let M = C and let E = Cd. Then since every repre-
sentation σ of C is simply a multiple of the basic, 1-dimensional rep-
resentation σ1, i.e., σ = nσ1 for a suitable positive integer n or ∞, it
follows that Eσ∗ = {z = (Z1, Z2,· · · Zd) Zi ∈ Mn(C)}, where we inter-
pret M∞(C) as B(ℓ2(N)). The disc D(0, R, σ) is {z = (Z1, Z2,· · · Zd)
kP ZiZ ∗
i k 1
2 < R}, and a z = (Z1, Z2,· · · , Zd) ∈ D(0, 1, σ) is in AC(σ),
when σ is a finite multiple of σ1, if and only if z is completely non-
coisometric [MS11b, Corollary 5.7] in the sense that kz(k)∗hk → 0 for
every h ∈ Cn [MS04, Remark 7.2]. (No such simple characterization of
AC(∞σ1) is known.) For a z ∈ Eσ∗ and unitary u ∈ σ(M)′ = Mn(C),
u−1 · z · u = u−1z(IE ⊗ u) = (u−1Z1u, u−1Z2u,· · · , u−1Zdu). Thus z is
central if and only if each Zi is a scalar multiple of the identity. Fur-
ther, as we have seen, an additive subcategory Σ of NRep(M) is simply
determined by an additive subsemigroup of N∪ {∞} and a central ad-
ditive field {ζσ}σ∈Σ is a family of central elements such that for each i,
1 ≤ i ≤ d, the scalar that forms the scalar multiple of the identity in
the ith element of ζσ is independent of σ.
Remark 3.4. The notion of a matricial family of functions indexed by
NRep(M), defined in Definition 1.2, can be "relativized" to additive
subcategories, too. So, if {U(σ)}σ∈Σ is a matricial E, Σ-family, where
Σ is an additive subcategory of NRep(M), then a family of functions
f = {fσ}σ∈Σ, with
fσ : U(σ) → B(Hσ),
σ ∈ Σ,
is an E, Σ-matricial family of functions, or simply a matricial family of
functions, in case f respects intertwiners in the sense that for every z ∈
U(σ), every w ∈ U(τ ), I(σ×z, τ×w) ⊆ I(fσ(z), fτ (w)). (Thanks to our
convention established in Remark 2.3, I(σ×z, τ ×w) ⊆ I(fσ(z), fτ (w))
if and only if every C ∈ I(σ, τ ) that satisfies the equation Cz = w(IE ⊗
C) also satisfies the equation Cfσ(z) = fτ (w)C.) We shall say that f
is continuous, or holomorphic, etc., in case each fσ has the indicated
property.
Proposition 3.5. Let Σ be an additive subcategory of NRep(M) and
let D(ζ, R) be the matricial disc determined by a central additive field
ζ on Σ, where 0 < R ≤ ∞. If θ ∈ T+((E)) has radius of convergence
TENSORIAL FUNCTION THEORY
20
at least R, then the collection {fσ}σ∈Σ of tensorial power series deter-
mined by θ, fσ(z) := Pk≥0 Zk(z − ζσ)Lθk, forms a matricial family of
holomorphic functions on D(ζ, R).
Proof. We saw in Proposition 2.8 that each fσ is a holomorphic B(Hσ)-
valued function defined throughout D(ζσ, R, σ). We therefore need only
check that {fσ}σ∈Σ preserves intertwiners. For this, it suffices to check
that for each k, the kth terms of {fσ}σ∈Σ preserve intertwiners.
If
C ∈ I(σ × z, τ × w), then by definition C ∈ I(σ, τ ), so C ∈ I(Z0(z −
ζσ)Lθ0,Z0(w−ζτ )Lθ0) because Z0(z−ζσ) is identically IHσ in z, similarly
for Z0(w − ζτ ), and because when we view Lθ0 as a map from Hσ
to M ⊗σ(M ) Hσ we identify θ0 with σ(θ0), when we identify Hσ with
M ⊗σ(M ) Hσ, as is customary; similarly for Lθ0 and τ (θ0). To handle
I(Z1(z− ζσ)Lθ1,Z1(w− ζτ )Lθ1), observe first that Lθ1C = (IE ⊗ C)Lθ1
whether Lθ1 is viewed as a map from Hσ to E ⊗σ Hσ or from Hτ
to E ⊗τ Hτ . Further, Cζσ = ζτ (IE ⊗ C) by definition of a central
family and Cz = w(IE ⊗ C) by the hypothesis that C ∈ I(σ × z, τ ×
w). Thus C ∈ I(Z1(z − ζσ)Lθ1,Z1(w − ζτ )Lθ1). The general case
I(Zk(z− ζσ)Lθk,Zk(w− ζτ )Lθk) is handled by noting that it suffices to
check the intertwining condition when θk is a decomposable tensor, say
θk = ξ1⊗ξ2⊗· · ·⊗ξk and noting that in this case, Zk(z−ζσ)Lθk = Z1(z−
ζσ)Lξ1Z1(z − ζσ)Lξ2 · · ·Z1(z − ζσ)Lξk and similarly Zk(w − ζτ )Lθk =
Z1(w − ζτ )Lξ1Z1(w − ζτ )Lξ2 · · ·Z1(w − ζτ )Lξk . For these expressions,
it is obvious that C intertwines, by virtue of the fact that C ∈ I(Z1(z−
ζσ)Lθ1,Z1(w − ζτ )Lθ1).
A concept that is closely related to the notion of a matricial family of
functions is given in
(cid:3)
Definition 3.6. Suppose E and F are two W ∗-correspondences over
the same W ∗-algebra, M and suppose Σ is an additive subcategory
of NRep(M). If {U(σ)}σ∈Σ is an E, Σ-matricial family of sets and if
{V(σ)}σ∈Σ is an F, Σ-matricial family of sets, then we call a family
of maps {fσ}σ∈Σ, where fσ maps U(σ) to V(σ), an E, F, Σ-matricial
family of maps, or for short, a matricial family of maps, in case
Cfσ(z) = fτ (w)(IF ⊗ C)
for every C : Hσ → Hτ in I(σ, τ ) such that Cz = w(IE ⊗ C), for all
z ∈ U(σ) and all w ∈ V(τ ). To say the same thing more succinctly,
{fσ}σ∈Σ is a matricial family of maps in case
(3.1)
I(σ × z, τ × w) ⊆ I(σ × fσ(z), τ × fτ (w)),
z ∈ U(σ), w ∈ U(τ ).
TENSORIAL FUNCTION THEORY
21
To help forestall confusion that may develop, we emphasize that we
will be consistent in our distinction between functions and maps: ma-
tricial families of functions map matricial families of sets to algebras
of operators; matricial families of maps are families of maps between
matricial families of sets. As we shall see later, matricial families of
maps are connected to homomorphisms between Hardy algebras.
4. A special Generator
In ring theory, a module G is a generator for the category of left mod-
ules over a ring R, RM, in case every M ∈ RM is the image of a
homomorphism from the algebraic direct sum of a suitable number of
copies of G [AF92, Page 193]. In [Rie74b], Rieffel defines a generator
for NRep(M) in a similar fashion, but allows infinite Hilbert space
direct sums. In [Rie74b, Proposition 1.3], he proves that a representa-
tion σ is a generator for NRep(M) in this extended sense if and only
if σ is faithful. Here we shall develop a useful analogue of the notion
of a generator for the category of ultraweakly continuous, completely
contractive representations of H ∞(E).
Definition 4.1. We shall say that a generator π for NRep(M) is an
infinite generator in case it is an infinite multiple of a generator for
NRep(M), i.e., an infinite multiple of a faithful normal representation
of M. We shall say that σ0 is a special generator for NRep(M) if
σ0 = πF (E) ◦ ϕ∞ for an infinite generator for NRep(M).
Remark 4.2. Of course σ0 and π are equivalent in NRep(M) if π is an
infinite generator. However, we want to consider additive subcategories
of NRep(M) that are not necessarily closed under forming infinite di-
rect sums. Consequently, it is important for our considerations to make
a distinction between σ0 and π.
If σ0 = πF (E) ◦ ϕ∞, acting on Hσ0 = F (E)⊗π Hπ, is a special generator
for NRep(M), and if s0 is defined by the formula
s0(ξ ⊗ h) = Tξh,
ξ ∈ E, h ∈ F (E) ⊗π Hπ,
then σ0 × s0 is an induced representation of H ∞(E) in the sense of
[MS99]. In [MS11b, Proposition 2.3] we show that σ0 × s0 is unique up
to unitary equivalence in the sense that if π′ has the same properties as
0 is constructed from π′ in a similar fashion to σ0×s0, then
π and if σ′
σ′
0⊗ s′
0 is unitarily equivalent to σ0× s0. Further, if σ× z is any induced
representation of H ∞(E), then there is a subspace K of Hπ that reduces
π such that σ× z is unitarily equivalent to σ0 × s0F (E)⊗π K. Observe
that by construction σ0×s0 is absolutely continuous, so s0 ∈ AC(σ0). In
0⊗s′
TENSORIAL FUNCTION THEORY
22
fact, σ0×s0 is a generator for the category of all ultraweakly continuous
completely contractive representations of H ∞(E). This assertion is the
content of the following theorem, which is a summary of Theorems 4.7
and 4.11 of [MS11b].
Theorem 4.3. Suppose σ ∈ NRep(M) and z ∈ D(0, 1, σ). Then z lies
in AC(σ) if and only if Hσ is the closed linear span of the ranges of
the operators in I(σ0 × s0, σ × z). In this event, the stronger equation
holds:
Hσ =[{Ran(C) C ∈ I(σ0 × s0, σ × z)}.
With this theorem at our disposal, we are able to prove the following
theorem that has Theorem 1.3 as an immediate corollary.
Theorem 4.4. Suppose Σ is a full additive subcategory of NRep(M)
that contains a special generator σ0 for NRep(M). If f = {fσ}σ∈Σ is
a family of functions such that each fσ maps AC(σ) to B(Hσ), then f
is a Berezin transform (restricted to {AC(σ)}σ∈Σ) if and only if f is
an E, Σ-matricial family.
Proof. We already have noted Berezin transforms are matricial fam-
ilies of functions on {AC(σ)}σ∈N Rep(M ). So certainly, their restric-
tions to {AC(σ)}σ∈Σ are matricial E, Σ-families. For the converse,
suppose that {fσ}σ∈Σ is a matricial family of functions defined on
{AC(σ)}σ∈Σ. Since Σ is assumed to be full, for every σ and τ in Σ,
IΣ(σ, τ ) = IN Rep(M )(σ, τ ) where the subscripts indicate the category
under consideration. It follows that for every z ∈ AC(σ) and for ev-
ery w ∈ AC(τ ), IΣ(σ × z, τ × w) = IN Rep(M )(σ × z, τ × w). So our
hypotheses guarantee that for every C ∈ IN Rep(M )(σ0 × s0, σ0 × s0),
Cfσ0(s0) = fσ0(s0)C. That is, fσ0(s0) lies in the double commutant
of σ0 × s0(H ∞(E)). However, σ0 × s0 is the restriction of πF (E) to
H ∞(E), where σ0 = πF (E) ◦ ϕ∞, and πF (E)(H ∞(E)) is its own double
commutant by [MS04, Corollary 3.10]. Thus there is an F ∈ H ∞(E)
so that
fσ0(s0) = bFσ0(s0).
(4.1)
If σ is an arbitrary representation in Σ and if z ∈ AC(σ), then for
every C ∈ IN Rep(M )(σ0 × s0, σ × z), fσ(z)C = Cfσ0(s0) because Σ is full
and {fσ}σ∈Σ preserves intertwiners by hypothesis. However, by (4.1),
Cfσ0(s0) = CbFσ0(s0). Hence we have
fσ(z)C = CbFσ0(s0) = bFσ(z)C,
TENSORIAL FUNCTION THEORY
23
(cid:3)
where the second equality is justified by Remark 2.3. Since the ranges
of the C in IN Rep(M )(σ0 × s0, σ × z) cover all of Hσ, by Theorem 4.3,
we conclude that fσ(z) = bFσ(z).
We digress momentarily to provide an example promised in the intro-
duction that shows that Theorems 4.4 and 1.3 can fail if the hypothesis
that the matricial function f = {fσ}σ∈Σ in question is defined only on
{D(0, 1, σ)}σ∈Σ and not on the collection of larger sets, {AC(σ)}σ∈Σ.
Example 4.5. We let M = C = E. Then NRep(M) may be identified
with {nσ1}0<n≤∞, where, recall, σ1 is the one-dimensional representa-
tion of C on C. The disc D(0, 1, σ) is just the collection of all operators
of norm less than 1 in B(Hσ). We set fσ(z) = (IHσ − z)−1, i.e., fσ is
just the resolvent operator restricted to D(0, 1, σ). Then it is immedi-
ate that f = {fσ}σ∈N Rep(M ) is a matricial family of functions defined
on {D(0, 1, σ)}σ∈N Rep(M ). However, since none of the fσs is bounded,
f is not a Berezin transform of an element in H ∞(E) ≃ H ∞(T).
This example should be compared with Theorem 4.8.
We would like to use Theorem 4.4 to obtain information about which
matricial families of functions come from tensorial power series that
have a given radius of convergence. First, however, we take up an issue
that was left hanging after Proposition 2.8.
Proposition 4.6. Suppose θ ∼P θk ∈ T+((E)) has a finite radius of
convergence R = R(θ). If R′ > R, then there a σ ∈ NRep(M) and z ∈
Eσ∗, with kzk = R′, such that the tensorial power series Pk Zk(·)Lθk
diverges at z; indeed, Pk kZk(z)Lθkk = ∞.
Proof. Choose ρ, with R < ρ < R′, let σ be σ0 and set z = R′s0. Since
1
ρ < 1
R, there are infinitely many ks for which kθkk1/k > 1
ρ. On the
other hand, Zk(z)Lθk = Zk(R′s0)Lθk = R′k(Tθk ⊗π IK0). Consequently,
kZk(z)Lθkk = R′kkθkk. So for each k satisfying kθkk1/k > 1
ρ, we have
kZk(z)Lθkk > ( R′
ρ )k > 1. Since there are infinitely many such ks, the
series Pk kZk(z)Lθkk diverges to ∞.
Theorem 4.7. Suppose Σ is a full additive subcategory of NRep(M)
containing a special generator for NRep(M). If f = {fσ}σ∈Σ is a fam-
ily of functions, with fσ mapping D(0, R, σ) to B(Hσ), then there is a
formal tensor series θ with R(θ) ≥ R such that f is the family of tenso-
rial power series determined by θ, {Pk≥0 Zk(z)Lθk z ∈ D(0, R, σ)}σ∈Σ,
if and only if f is an E, Σ-matricial family of functions.
(cid:3)
TENSORIAL FUNCTION THEORY
24
Proof. Arguments we have used before show that if the fσs admit ten-
sorial power series expansions of the indicated kind, then the family
preserves intertwiners. Consequently, we shall concentrate on the con-
verse assertion. So assume that {fσ}σ∈Σ preserves intertwiners. Since
Σ is assumed to be full, we may and will drop the subscripts Σ and
NRep(M) on all intertwining spaces. Also, we will let σ0 = πF (E)◦ ϕ∞,
for a suitable infinite generator π. The key to our analysis is to focus on
fσ0 in order to bring properties of H ∞(E) into play and then to use the
intertwining property of the family {fσ}σ∈Σ to propagate to the other
discs D(0, R, σ). Fix 0 < r < R and consider fσ0(rs0). Recall that
σ0 × s0(H ∞(E)) = πF (E)(H ∞(E)). Consequently, every C in the com-
mutant of πF (E)(H ∞(E)) lies in I(σ0×s0, σ0×s0) = I(σ0×rs0, σ0×rs0)
and, thus commutes with fσ0(rs0). Since πF (E)(H ∞(E)) equals its
own double commutant, there is an element Fr ∈ H ∞(E) such that
fσ0(rs0) = πF (E)(Fr) = cFrσ0(s0). Now take a z ∈ D(0, r, σ). Then
k 1
r zk < 1 and so σ × 1
r z is absolutely continuous. We conclude, by
Theorem 4.3, that
(4.2)
_{Ran(C) C ∈ I(σ0 × s0, σ ×
1
r
z)} = Hσ.
Also, for every C ∈ I(σ0 × s0, σ × 1
r z) = I(σ0 × rs0, σ × z), Cfσ0(rs0) =
fσ(z)C by hypothesis. Since fσ0(rs0) =cFrσ0(s0), we see that fσ(z)C =
CcFrσ0(s0). But by Remark 2.3, CcFr σ0(s0) =cFrσ( 1
kzk < r < R.
r z)C, so we conclude
from (4.2) that ,
(4.3)
1
r
z),
fσ(z) =cFr σ(
We need to remove the dependence of Fr on r. So if 0 < r < r1 <
z).
r2 < R and if kzk ≤ r, we obtain the equation cFr1 σ( 1
z) = cFr2 σ( 1
In particular,
r1
r2
We now would like to apply the "Fourier coefficient maps" Φk to equa-
tion (4.4). To give this its proper meaning, note that, whenever X ∈
H ∞(E) and 0 < t < 1,
(4.5)
\Φk(X)σ0
(ts0) = tkΦk(X) ⊗ IHπ.
This is easy to verify by first taking X = Tξ for ξ ∈ E⊗m, and then using
linearity and ultra-weak continuity. Thus, applying Φk to equation
(4.4)
r
r1
cFr1 σ0(
s0) = cFr2 σ0(
r
r2
s0).
TENSORIAL FUNCTION THEORY
25
(4.4), we obtain
(4.6)
rk
rk
1
Φk(Fr1) ⊗ IHπ =
rk
Φk(Fr2) ⊗ IHπ.
rk
2
for all k ≥ 0, which implies that r 7→ 1
rk Φk(Fr) is constant in r, 0 < r <
R, for every k. Consequently, since the image of Φk is {Tθk θk ∈ E⊗k},
there is a θk ∈ E⊗k, independent of r, so that
Tθk ⊗ IHπ =
1
rk Φk(Fr) ⊗ IHπ,
0 < r < R.
Φk(Fr) = rkTθk = Trkθk,
Canceling "⊗IHπ ", as we may, we conclude that
(4.7)
Now fix 0 < r < R and z ∈ Eσ∗ with kzk < r. For 0 ≤ k, let
ξk = rkθk ∈ E⊗k. Then with Fr in place of F and 1
r z in place of z, we
find that
0 < r < R.
1
r
z) =Xk≥0
1
rkZk(z)Lξk
cFrσ(
1
andP 1
rkkZk(z)Lξkk < ∞. By (4.3), we conclude that fσ(z) =cFr σ( 1
rkZk(z)Lξk = Pk≥0 Zk(z)Lθk. Thus PkZk(z)Lθkk < ∞ and
Pk≥0
fσ(z) =Pk≥0 Zk(z)Lθk. By Theorem 4.6, R(θ) ≥ R.
It remains to show that the series θ ∼P θk is uniquely determined by
{fσ}σ∈Σ. In fact, it is uniquely determined by fσ0. Suppose θ′ ∼P θ′
is another series with R(θ′) ≥ R and suppose
Zk(z)Lθk = fσ0(z) =Xk≥0
Zk(z)Lθ′
Xk≥0
k
k
r z) =
Zk(ts0)Lθk =Xk≥0
for all z ∈ D(0, R, σ0). However, as we have seen from our analysis
that yielded (4.5), Zk(ts0)Lθk = tkTθk ⊗ IHπ . Thus we conclude that
for 0 ≤ t < min{1, R(θ)},
tkTθk ⊗ IHπ =Xk≥0
Xk≥0
where all the series converge in the operator norm on B(F (E)⊗π Hπ).
Since the map X 7→ πF (E)(X), X ∈ L(F (E)), is a faithful normal
representation of L(F (E)),
Xk≥0
tkTθk =Xk≥0
=Xk≥0
Zk(ts0)Lθ′
k ⊗ IHπ ,
tkTθ′
tkTθ′
k
k
as norm-convergent series in H ∞(E). So, if we apply Φk to both sides,
we conclude that tkTθk = tkTθ′
k for every k
and θ = θ′.
(cid:3)
for every k. Hence θk = θ′
k
TENSORIAL FUNCTION THEORY
26
represents a bounded holomorphic function fσ on D(0, 1, σ) for a par-
As we noted in Remark 2.12, it can happen that a seriesPk≥0 Zk(z)Lθk
ticular σ and R(θ) ≥ 1, θ ∼P θk, but fσ is not a σ-Berezin transform.
However, the following proposition shows that if fσ is a member of a
matricial family of functions f = {fσ}σ∈Σ that are uniformly bounded
in σ, then f is a Berezin transform.
Theorem 4.8. Suppose Σ is an additive full subcategory of NRep(M)
that contains a special generator for NRep(M) and suppose f = {fσ}σ∈Σ
is an E, Σ-matricial family of functions, with fσ defined on D(0, 1, σ)
and mapping to B(Hσ). Then f is a Berezin transform restricted to
{D(0, 1, σ)}σ∈Σ if and only if
(4.8)
sup{kfσ(z)k σ ∈ Σ, z ∈ D(0, 1, σ)} < ∞.
Proof. If there is an F ∈ H ∞(E) such that fσ = bFσ for all σ, then
certainly kfσ(z)k ≤ kFk for all σ and z. So we shall attend to the
converse and suppose sup{kfσ(z)k σ ∈ NRepO(M), z ∈ D(0, 1, σ)} =
A < ∞. If 0 < r < 1, then as we saw in the proof of Theorem 4.7, there
is an Fr ∈ H ∞(E) such that fσ0(rs0) = πF (E)(Fr) and fσ(z) =cFr σ( 1
r z)
for all kzk < r and all σ (see (4.3)).
Also, it follows from Theorem 4.7 that f is a tensorial power series
{Pk≥0 Zk(·)Lθk
z ∈ D(0, 1, σ)} where the series θ ∼ Pk≥0 θk in
T+((E)) has R(θ) ≥ 1. For 0 < t < r < 1, we thus have
s0) = fσ0(ts0) =Xk≥0
cFrσ0(
Zk(ts0)Lθk =Xk≥0
tkTθk ⊗ IHπ
t
r
and it follows that for every integer m ≥ 0,
for every 0 < r < 1.
Note that, for every 0 < r < 1, Fr = πF (E)(Fr) = fσ0(rs0) ≤ A.
Thus {Fr} is a bounded set.
If rn ր 1 and if F is an ultraweak limit point of {Frn}, say F =
limα Frnα for an appropriate subnet of {rn}, then for every m ≥ 0
we have Φm(πF (E)(F )) = lim Φm(πF (E)(Frnα )) = lim Φm(dFrnα σ0
(s0)) =
lim rm
nαTξm⊗π IHπ = Tξm⊗π IHπ. It follows that Φm(F ) = Tθm and, using
Therefore
t
r
s0)) = tmTθk ⊗ IHπ.
Φm(cFrσ0(
Φm(cFrσ0(s0)) = rmTθk ⊗ IHπ
TENSORIAL FUNCTION THEORY
27
Proposition 2.11, we have, for every σ ∈ Σ and every z ∈ D(0, 1, σ),
fσ(z) =Xk≥0
Zk(z)Lθk = bFσ(z).
(cid:3)
Thus fσ = bFσ.
In the following theorem we want to consider two W ∗-correspondences
over M, E and F , and relate maps between their families of absolutely
continuous representations to homomorphisms between their Hardy al-
gebras.
Theorem 4.9. Suppose Σ is a full additive subcategory of NRep(M)
that contains a special generator for NRep(M). Suppose also that E
and F are two W ∗-correspondences over M and that f = {fσ}σ∈Σ is
a family of maps, with fσ : AC(σ, E) → AC(σ, F ). Then f is an
E, F, Σ-matricial family of maps if and only if there is an ultraweakly
continuous homomorphism α : H ∞(F ) → H ∞(E) such that for every
z ∈ AC(σ, E) and every Y ∈ H ∞(F ),
(4.9)
Proof. Suppose that f = {fσ}σ∈Σ is a matricial family of maps and let
Y be an element of H ∞(F ). For each σ ∈ Σ, define gσ : AC(σ, E) →
B(Hσ) by gσ(z) = bY (fσ(z)). If z ∈ AC(σ, E), if w ∈ AC(τ, E), and if
C ∈ I(σ× z, τ × w), then since f is assumed to be a matricial family of
maps, C ∈ I(σ× fσ(z), τ × fτ (w)). By Theorem 4.4, Cgσ(z) = gτ (w)C.
Thus, by Theorem 4.4 again, there is an operator α(Y ) in H ∞(E)
such that [α(Y )(z) = gσ(z) for all z ∈ AC(σ, E). Thus equation 4.9
is satisfied. Note that α(Y ) is uniquely determined by virtue of the
uniqueness assertion in Theorem 4.7. Since
[α(Y )(z) = bY (fσ(z)).
(4.10)
[α(Y )(z) = gσ(z) = bY (fσ(z)),
it is clear that α is a homomorphism. It remains to prove that α is
ultraweakly continuous. For that purpose, let σ be the special generator
σ0 and let z = s0 in (4.10), to conclude that
α(Y ) ⊗ IHπ = bY (fσ0(s0)),
from which it follows immediately that α is ultraweakly continuous.
For the converse, suppose the family f implements a homomorphism
α via equation (4.10). To show that f is an E, F, Σ-matricial family of
maps, we must show that the family preserves intertwiners in the sense
of equation (3.1). So let C ∈ I(σ × z, τ × w) and apply Theorem 4.4
to conclude that CbY (fσ(z)) = C [α(Y )(z) = [α(Y )(w)C = bY (fτ (w))
TENSORIAL FUNCTION THEORY
28
ξ = fτ (w)Lτ
(cid:3)
ξ h := ξ ⊗ h and if we define Lτ
ξ C = fτ (w)(IE ⊗ C)Lσ
for all Y ∈ H ∞(F ).
In particular, this holds for Y = Tξ for every
ξ ∈ F . So, if we write Lσ
ξ for the map from Hσ → F ⊗σ Hσ defined
by the equation Lσ
ξ similarly, then for
ξ and bY (fτ (w)) = fτ (w)Lτ
Y = Tξ we conclude that bY (fσ(z)) = fσ(z)Lσ
ξ .
Thus Cfσ(z)Lσ
ξ . Since this equation
holds for all ξ ∈ F , we conclude that Cfσ(z) = fτ (w)(IE ⊗ C). Thus
C ∈ I(fσ(z), fτ (w)).
Remark 4.10. Recall that an element z ∈ Eσ∗ lies in the center of Eσ∗
in case b · z = z · b for all b ∈ σ(M)′ [MS08, Definition 4.11], in which
case we write z ∈ Z(Eσ∗). Thus z ∈ Z(Eσ∗) if and only if σ(M)′ ⊆
I(σ × z, σ × z). It follows that if {fσ}σ∈N Rep(M ) is a matricial family of
maps, with fσ : AC(σ, E) → AC(σ, F ), then fσ(AC(σ, E) ∩ Z(Eσ∗)) ⊆
AC(σ, F ) ∩ Z(F σ∗).
5. Function Theory without a generator
In this section we shall focus on additive subcategories Σ of NRep(M)
that do not necessarily contain a special generator and address the
problem of deciding which matricial families of functions {fσ}σ∈Σ have
tensorial power series representations as in Proposition 2.8. In partic-
ular, we shall prove the following theorem which complements and is
something of a converse to Proposition 3.5.
Theorem 5.1. Suppose Σ is an additive subcategory of NRep(M) and
that D(ζ, R) is a matricial disc determined by an additive field ζ =
{ζσ}σ∈Σ on Σ and R, 0 < R ≤ ∞. Suppose, also, that f = {fσ}σ∈Σ
is a matricial family of functions defined on D(ζ, R) that is locally
uniformly bounded on D(ζ, R) in the sense that for each r < R,
sup
σ∈Σ
z∈D(ζσ,r)kfσ(z)k < ∞.
sup
Then:
(1) Each fσ is Frechet analytic on D(ζσ, R).
(2) If the subcategory is full, if the additive field ζ is also central,
and if each σ ∈ Σ is faithful, then f is a family of tensorial
power series {Pk≥0 Z k(z − ζσ)Lθk z ∈ D(ζk, R, σ)}σ∈Σ, where
θ ∼Pk≥0 θk has R(θ) ≥ R.
To achieve this goal, we use the matrix analysis initiated by Tay-
lor in [Tay72], and developed further in the work of Voiculescu in
[Voi04, Voi10], in the work of Helton and his collaborators [HKMS09,
TENSORIAL FUNCTION THEORY
29
HKM11a, HKM11b] and especially in the investigations of Kaliuzhnyi-
Verbovetskyi and Vinnikov [KVV09, KV10, KVV06]. Indeed, the first
half of Theorem 5.1 is proved by showing that Taylor's matrix analysis
leads to one of Taylor's Taylor series for each fσ. The existence of such
a series implies that fσ is Frechet analytic. The second half of the the-
orem involves showing that each of Taylor's Taylor coefficients comes
from an element of a suitable tensor power of E⊗k.
Throughout this section, Σ will denote a fixed additive subcategory
of NRep(M). No other assumptions will be placed on Σ, except as
explicitly stated in added hypotheses in the statements of results. In
particular, we do not assume that Σ is full, nor do we assume that Σ
contains a special generator. The matricial sets we will consider will
primarily be matricial discs D(ζ, R) = {D(ζσ, R, σ)}σ∈Σ, where ζ is an
additive field of vectors on Σ (see Example 3.2). Note, in particular,
that the assumption that ζ is an additive field on Σ guarantees that
ζmσ = ζσ ⊕ ζσ ⊕ · · · ⊕ ζσ (m-summands).
The following result can be found in [KV] for the case when M = C.
Lemma 5.2. Let f = {fσ}σ∈Σ be a matricial family of functions de-
fined on a matricial E, Σ-family {U(σ)}σ∈Σ where Σ is an additive
subcategory of NRep(M). Suppose σ, τ ∈ Σ, z ∈ U(σ), w ∈ U(σ) and
u ∈ I(τ E ◦ ϕ, σ) are such that (cid:18) z u
0 w (cid:19) ∈ U(σ ⊕ τ ). Then there is an
operator ∆fσ,τ (z, w)(u) ∈ B(Hτ , Hσ) such that
(1)
fσ⊕τ ((cid:18) z u
0 w (cid:19)) =(cid:18) fσ(z) ∆fσ,τ (z, w)(u)
fτ (w)
0
(cid:19) .
(2) If b ∈ σ(M)′ is such that z(IE⊗b) = bz and such that(cid:18) z
0 w (cid:19) ∈
U(σ⊕τ ), then fσ(z)b = bfσ(z) and ∆fσ,τ (z, w)(bu) = b∆fσ,τ (z, w)(u).
In particular ∆fσ,τ (z, w)(tu) = t∆fσ,τ (z, w)(u) for t ∈ C such
that
bu
(cid:18) z
0 w (cid:19) ∈ U(σ ⊕ τ ).
tu
(3) If b ∈ τ (M)′ is such that w(IE ⊗ b) = bw and such that
(cid:18) z u(IE ⊗ b)
w
0
(cid:19) ∈ U(σ ⊕ τ ),
then fτ (w)b = bfτ (w) and ∆fσ,τ (z, w)(u(IE⊗b)) = ∆fσ,τ (z, w)(u)b.
TENSORIAL FUNCTION THEORY
30
(4) If σ = τ , z ∈ U(σ) and b ∈ σ(M)′ such that (cid:18) z
0 z + bu (cid:19) ∈
bu
U(σ ⊕ σ), then
fσ(z + bu) − fσ(z) = ∆fσ,σ(z, z + bu)(bu)
and, if t ∈ C,
fσ(z + tu) − fσ(z) = t∆fσ,σ(z, z + tu)(u).
Remark 5.3. It should be emphasized that in (1), ∆fσ,τ (z, w)(u) repre-
sents an operator in B(Hτ , Hσ), whose full dependence on z, w and u
has still to be determined. Among other things, part (2) of the lemma
proves that ∆fσ,τ (z, w)(u) extends to be homogeneous in u of degree
one, but the additivity in u will be proved later in Lemma 5.9. It should
also be emphasized that to prove that ∆fσ,τ (z, w)(u) extends to be ho-
mogeneous in u the full force of the assumption that f is matricial is
not used: We may assume that σ = τ and that the intertwiners in-
volved are operator matrices whose entries are scalar multiples of the
identity (on Hσ). These are present in any of the categories we use, as
was stated at the outset of Section 2.
Proof. Write Iσ for the identity operator on Hσ. Then
(5.1)
and
(5.2)
(5.3)
and
(5.4)
(cid:19) =(cid:18) Iσ
0
(cid:18) z u
0 w (cid:19)(cid:18) IE ⊗ Iσ
0 (cid:19) z
(cid:0) 0 Iτ (cid:1)(cid:18) z u
0 w (cid:19) = w(cid:0) 0 IE ⊗ Iτ (cid:1) .
fσ⊕τ ((cid:18) z u
0 w (cid:19))(cid:18) Iσ
0 (cid:19) fσ(z)
(cid:0) 0 Iτ (cid:1) fσ⊕τ ((cid:18) z u
0 w (cid:19)) = fτ (w)(cid:0) 0 Iτ (cid:1) .
a21 a22(cid:21) , then (5.3) shows that a11 =
0 w(cid:21)) as(cid:20)a11 a12
0 (cid:19) =(cid:18) Iσ
Since f preserves intertwiners, we may write
So, if we write fσ⊕τ ((cid:20)z u
fσ(z) and a21 = 0, while (5.4) also shows that a21 = 0 as well as
a22 = fτ (w). The remaining entry, a12, is taken as the definition of
∆fσ,τ (z, w)(u). This proves (1).
TENSORIAL FUNCTION THEORY
31
For (2) note that the equality
(cid:18) b
0 Iτ (cid:19)(cid:18) z u
0 w (cid:19) =(cid:18) z
0 w (cid:19)(cid:18) IE ⊗ b
bu
0
0
0
IE ⊗ Iτ (cid:19)
implies
(cid:18) b
0 Iτ (cid:19)(cid:18) fσ(z) ∆fσ,τ (z, w)(u)
fτ (w)
0
0
(cid:19)
=(cid:18) fσ(z) ∆fσ,τ (z, w)(bu)
fτ (w)
0
(cid:19)(cid:18) b
0 Iτ (cid:19) ,
0
proving (2).
The proof of (3) is similar using the equality
(cid:18) Iσ 0
b (cid:19)(cid:18) z u(IE ⊗ b)
w
0
0
(cid:19) =(cid:18) z u
0 w (cid:19)(cid:18) IE ⊗ Iσ
0
0
IE ⊗ b (cid:19) .
For (4), simply note that
(cid:0) −Iσ
Iτ (cid:1)(cid:18) z
0 z + bu (cid:19) = z(cid:0) −IE ⊗ Iσ
bu
IE ⊗ Iτ (cid:1)
(cid:3)
and use the properties of f .
Even though the linearity in u of the operator ∆fσ,τ (z, w)(u) has still
to be shown, ∆fσ,τ (z, w) can be viewed as a noncommutative difference
operator with values in B(Hτ , Hσ).
Definition 5.4. With the hypotheses and notation as in Lemma 5.2,
we call ∆fσ,τ (z, w) the Taylor difference operator determined by f and
the points z and w. If σ = τ and z = w, we call ∆fσ,σ(z, z) the Taylor
derivative of f at z and denote it ∆fσ(z, z) or ∆fσ(z).
Even though its linearity in u has still to be shown, we can also define
non commutative difference operators of higher order. We will need
these first. So to this end, note that by applying part (1) of Lemma 5.2
repeatedly one finds that for every σ0, . . . , σn in Σ, for every zi ∈ U(σi)
(0 ≤ i ≤ n) and for every uj ∈ I(σE
j ◦ ϕ, σj−1) (1 ≤ j ≤ n), the matrix
u1
z0
(5.5)
fσ0⊕σ1⊕···⊕σn(
z1
. . .
0
...
...
0 · · ·
0
...
0
· · ·
. . .
. . .
zn−1 un
zn
0
)
0
. . .
. . .
. . .
· · ·
TENSORIAL FUNCTION THEORY
32
has a block upper triangular form with fσ0(z0), . . . , fσn(zn) on the main
diagonal, assuming of course that the argument is in U(σ0 ⊕ σ1 ⊕ · · ·⊕
σn), i.e.,
(5.6) fσ0⊕σ1⊕···⊕σn(
=
z0
u1
z1
. . .
0
...
...
0 · · ·
fσ0(z0)
0
...
...
0
0
. . .
. . .
. . .
· · ·
0
...
0
· · ·
. . .
. . .
zn−1 un
zn
a02
fσ1(z1) a12
. . .
a01
0
)
· · ·
. . .
. . .
a0n
a1n
...
fσn−1(zn−1)
an−1n
fσn(zn)
· · ·
· · ·
Definition 5.5. The function of u1, u2, · · · , un defined by the 0, n
entry of the right hand side of Equation 5.6 will be called the nth-
order Taylor difference operator determined by z0, z1,· · · ,zn, and will
be denoted ∆nfσ0,σ1,··· ,σn(z0, . . . , zn). If z0 = z1 = · · · = zn = z, we call
∆nfσ,σ,··· ,σ(z, z,· · · , z) := ∆nfσ(z) the nth-order Taylor derivative of fσ
at z.
The arguments for the proofs of parts (2) and (3) of Lemma 5.2 can be
adapted easily to prove the following lemma. We omit the details.
Lemma 5.6. Suppose {zi} and {uj} are as above. Fix 1 ≤ k ≤ n + 1
and choose b ∈ σk−1(M)′ so that bzk−1 = zk−1(IE ⊗ b) (in particular, b
can be in C), then:
(1) If 1 < k ≤ n and if both
∆nfσ0,σ1,··· ,σn(z0, . . . , zn)(u1, . . . , uk−1, buk, . . . , un)
and ∆nfσ0,σ1,··· ,σn(z0, . . . , zn)(u1, . . . , uk−1(IE⊗b), uk, . . . , un) are
well defined, in the sense that the argument matrices in the
expression 5.5 lie in U(σ0 ⊕ σ1 ⊕ · · · ⊕ σn), then
∆nfσ0,σ1,··· ,σn(z0, . . . , zn)(u1, . . . , uk−1, buk, . . . , un)
= ∆nfσ0,σ1,··· ,σn(z0, . . . , zn)(u1, . . . , uk−1(IE ⊗ b), uk, . . . , un).
TENSORIAL FUNCTION THEORY
33
(2) If k = 1 and if ∆nfσ0,σ1,··· ,σnf (z0, . . . , zn)(bu1, . . . , un) is well
defined, then
∆nfσ0,σ1,··· ,σn(z0, . . . , zn)(bu1, . . . , un)
= b∆nfσ0,σ1,··· ,σn(z0, . . . , zn)(u1, . . . , un)
.
(3) If k = n + 1 and if ∆nfσ0,σ1,··· ,σn(z0, . . . , zn)(u1, . . . , un(IE ⊗ b))
is well defined, then
∆nfσ0,σ1,··· ,σn(z0, . . . , zn)(u1, . . . , un(IE ⊗ b))
= ∆nfσ0,σ1,··· ,σn(z0, . . . , zn)(u1, . . . , un)b.
Remark 5.7. Again we want to note that the argument for Lemma 5.6
shows that if we take the σis to be one and the same σ, so that the
matrix
z0
0
...
...
0
u1
z1
. . .
· · ·
0
. . .
. . .
. . .
· · ·
0
...
0
· · ·
. . .
. . .
zn−1 un
zn
0
acts from E ⊗σ Hmσ to Hmσ, then the only matrices necessary to show
that ∆nfσ0,σ1,··· ,σn(z0, . . . , zn)(u1, . . . , un) extends to be homogeneous in
each of the uis are block matrices whose entries are scalar multiples of
the identity on Hσ.
Lemma 5.8. Given σ0, . . . , σn in Σ, zi ∈ U(σi) (0 ≤ i ≤ n) and
uj ∈ I(σE
j ◦ ϕ, σj−1) (1 ≤ j ≤ n), then for 0 ≤ i < j ≤ n, the i, j entry,
aij of the right hand side of Equation (5.6) is
ai,j = ∆jfσ0,σ1,··· ,σj (z0, . . . , zj)(u1, . . . , uj).
Proof. The proof proceeds by induction. For n = 1, the assertion is
Lemma 5.2(1). So assume it holds for n and write σ = ⊕n+1
i=0 σi. We
apply Lemma 5.2(1) repeatedly. If we partition σ as σ = (⊕n
i=0σi) ⊕
σn+1, Lemma 5.2 and the induction hypothesis prove the lemma for all
j ≤ n and for i = j = n + 1. To obtain the formula for ai,n+1, simply
write σ as σ = (⊕i−2
k=i−1σk), then apply Lemma 5.2 and the
induction assumption.
(cid:3)
k=0σk) ⊕ (⊕n+1
The argument used in the proof of the following lemma was shown
to us by Victor Vinnikov. It will appear in his joint work with D. S.
Kaliuzhnyi-Verbovetskyi [KVV].
TENSORIAL FUNCTION THEORY
34
Lemma 5.9. Let f = {fσ}σ∈Σ be a matricial family of functions de-
fined on a matricial E, Σ-family {U(σ)}σ∈Σ where Σ is an additive sub-
category of NRep(M). Suppose σ, τ ∈ Σ, z ∈ U(σ), w ∈ U(σ) and ui ∈
w (cid:19) ∈ U(σ⊕ τ ),
I(τ E ◦ ϕ, σ) for i = 1, 2 are such that A :=(cid:18) z u1 + u2
B :=
∈ U(2σ⊕τ ).
∈ U(2σ⊕τ ) and C :=
0
z 0 u2
0 z u1
0 0 w
z 0 u1
0 z u2
0 0 w
Then
∆fσ,τ (z, w)(u1 + u2) = ∆fσ,τ (z, w)(u1) + ∆f (z, w)(u2).
Proof. Considering f2σ⊕τ (B) and f2σ⊕τ (C) and using Lemma 5.2(1)
with 2σ ⊕ τ split as σ ⊕ (σ ⊕ τ ) we find that these matrices have the
following from:
and
f2σ⊕τ (B) =
f2σ⊕τ (C) =
fσ(z)
x
y
0
0
fσ(z)
0
0
fσ(z) ∆fσ,τ (z, w)(u2)
0
u
fτ (w)
v
fσ(z) ∆fσ,τ (z, w)(u1)
0
fτ (w)
.
for some x, y, u, and v. Writing S for the 3 × 3 permutation matrix
associated with the transposition (1, 2), we see that C = SBS−1 =
SBS. Thus f (C) = Sf (B)S and, therefore, y = ∆fσ,τ (z, w)(u1) and
f2σ⊕τ (B) =
f (z)
0
0
x ∆fσ,τ (z, w)(u1)
f (z) ∆fσ,τ (z, w)(u2)
0
fτ (w)
Now write D =(cid:18) I
fσ⊕τ (A)D. Since fσ⊕τ (A) = (cid:18) fσ(z) ∆fσ,τ (z, w)(u1 + u2)
0 0 I (cid:19). Then DB = AD and, thus Df2σ⊕τ (B) =
(cid:19), we may
fτ (w)
I 0
0
equate the (1, 3) entries to obtain the result.
(cid:3)
Remark 5.10. Again, in Lemma 5.9, if we set τ = σ, we see that
the only matrices used in the proof are matrices that are block scalar
matrices. Combining Lemma 5.9 with Lemma 5.2(2), we conclude that
∆fσ,τ (z, w)(u) extends to be linear in u.
We can now deduce the multilinearity of ∆nfσ0,σ1,··· ,σn(z0, . . . , zn)(·, . . . ,·)
(with z0, . . . , zn fixed) from the linearity of ∆fσ,τ (z, w)(u) in u. In order
TENSORIAL FUNCTION THEORY
35
to avoid specifying the conditions on the variables as in the statement
of the lemma above, we restrict ourselves to matricial discs.
Corollary 5.11. Let f = {fσ}σ∈Σ be a matricial family of functions,
defined in a matricial disc D(ζ, r). Then for every σ0, . . . , σn in Σ and
every zi ∈ D(ζσi, r, σi), (0 ≤ i ≤ n), the function
∆nfσ0,σ1,··· ,σn(z0, . . . , zn)(u1, . . . , un),
n ◦ ϕ, σn−1).
j ◦ ϕ, σj−1) (1 ≤ j ≤ n) with norm sufficiently
1 ◦ ϕ, σ0) × · · · ×
defined for uj ∈ I(σE
small can be extended to a multilinear map on I(σE
I(σE
Proof. For n = 1, this is shown in Lemma 5.9. For the general case, fix
1 ≤ j ≤ n and write
z0
u1
0
...
...
0
z1
. . .
· · ·
0
. . .
. . .
. . .
· · ·
0
...
0
· · ·
. . .
. . .
zn−1 un
zn
0
=(cid:18) X Z
0 Y (cid:19)
where X is a j×j block and Z is a j×(n−j) block with uj in the bottom
left corner and all other entries 0. Applying fσ0⊕σ1⊕···⊕σn to this matrix
we get, on one hand, a matrix with ∆nfσ0,σ1,··· ,σn(z0, . . . , zn)(u1, . . . , un)
in the upper right corner and, on the other hand, a 2 × 2 block matrix
that we shall simply write as ∆f (X, Y )(Z) in the (1, 2) corner. Using
Lemma 5.9, we know that ∆f (X, Y )(Z) is linear in Z and, thus, linear
in uj. Therefore ∆nfσ0,σ1,··· ,σn(z0, . . . , zn)(u1, . . . , un), being a corner, is
linear in uj. This completes the proof.
(cid:3)
The following theorem is an analogue of Taylor's Taylor Theorem with
remainder [Tay73, Proposition 4.2].
Theorem 5.12. Let f = {fσ}σ∈Σ be a matricial family of functions
defined on a matricial disc D(ζ, r). Fix σ ∈ Σ and choose z and w in
D(ζσ, r, σ) so that the matrix
(5.7)
z
. . .
z w 0
. . .
. . .
. . .
· · ·
0
...
...
0 · · ·
· · ·
. . .
. . .
z
0
0
...
0
w
z + w
TENSORIAL FUNCTION THEORY
36
lies in D(ζ(n+1)σ, r, (n + 1)σ). Then
fσ(z+w) =
n−1Xk=0
∆kfσ(z)(w, . . . , w)+∆nfσ,σ,··· ,σ(z, . . . , z, z+w)(w, . . . , w).
Proof. Denote the matrix (5.7) by A and note that,
A
IE ⊗ Iσ
...
...
...
IE ⊗ Iσ
=
Iσ
...
...
...
Iσ
(z + w).
Applying the intertwining property of f(n+1)σ and Lemma 5.8 completes
the proof.
(cid:3)
We would like to pass to the limit as n → ∞, in Theorem 5.12, but to
do this, we must impose an additional hypothesis on f , viz. f must be
locally uniformly bounded in the sense of Theorem 5.1. The following
theorem should be compared with Theorem 4.8. There, we assumed
that the additive subcategory Σ is full and contains a special generator
for NRep(M). Here we make no assumptions on Σ other than it is an
additive subcategory of NRep(M).
Theorem 5.13. Let f = {fσ}σ∈Σ be a matricial family of functions de-
fined on a matricial disc D(ζ, r) and suppose that f is locally uniformly
bounded in the sense of Theorem 5.1. Then:
(1) Each fσ is Frechet differentiable in z, z ∈ D(ζσ, r, σ), and
f ′
σ(z)(w) = ∆f (z)(w).
(2) Each fσ may be expanded on D(ζσ, r, σ) as
(5.8)
fσ(ζσ + z) =
∞Xk=0
∆kfσ(ζσ)(z, . . . , z),
where the series converges absolutely and uniformly on every
disc D(0, r0, σ) with r0 < r.
Proof. To prove (1), let r0 < r and let M be greater than
sup
σ∈Σ
sup
z∈D(ζσ,r0)kfσ(z)k,
TENSORIAL FUNCTION THEORY
37
which is finite by assumption. Fix z ∈ D(ζσ, r0, σ) and note that for w
with norm sufficiently small (say, w ≤ s), the matrix
(5.9)
lies in D(ζ3σ, r0, 3σ) = D(ζσ ⊕ ζσ ⊕ ζσ, r0, 3σ). But this implies in
particular, that
z w
0
z
0 0 z + w
0
w
(cid:20)z w
z(cid:21)
0
lies in D(ζ2σ, r0, 2σ). Our boundedness assumption, then implies that
for kwk ≤ s,
k∆fσ(z)(w)k ≤ k(cid:20)fσ(z) ∆fσ(z)(w)
fσ(z)
0
(cid:21) k = kf2σ((cid:20)z w
z(cid:21))k ≤ M,
0
i.e., ∆fσ(z)(·) extends to be a bounded linear operator. Further, the
same sort of argument shows that when we apply f3σ to the matrix (5.9)
for kwk ≤ s, we have ∆2fσ,σ,σ(z, z, z+w)(w, w) ≤ M, by definition of
∆2f·,·,·(·,·,·)(·,·). But by Theorem 5.12, fσ(z+w)−fσ(z)−∆fσ(z)(w) =
∆2fσ,σ,σ(z, z, z+w)(w, w). So, if we write w0 for
w w, then fσ(z+w)−
s )2∆2fσ,σ,σ(z, z, z + w)(w0, w0) ≤ w2M
fσ(z) − ∆fσ(z)(w) = ( w
,
which proves that
s2
s
1
wfσ(z + w) − fσ(z) − ∆fσ(z)(w) −→ 0
as w → 0. Thus (1) is proved.
To prove (2), fix 0 < r0 < r1 < r, write q := r1/r0 and let r2 :=
r1√1+q2
< r0. Then for every n and every z ∈ D(ζσ, r2, σ),
q(z − ζσ)
ζσ
w :=
0
...
...
0
ζσ
. . .
· · ·
· · ·
. . .
. . .
ζσ
0
0
. . .
. . .
. . .
· · ·
0
...
0
q(z − ζσ)
z
lies in D(ζ(n+1)σ, r1, (n + 1)σ) ⊆ D(ζ(n+1)σ, r, (n + 1)σ). Thus
∆nfσ,σ,··· ,σ(ζσ, . . . , ζσ, z)(q(z − ζσ), . . . , q(z − ζσ)) ≤ f (w) ≤ M.
From Lemma 5.8, we conclude that
∆nfσ,σ,··· ,σ(ζσ, . . . , ζσ, z)(z − ζσ, . . . , z − ζσ) ≤ q−nM −→n→∞
0.
So Theorem 5.12 proves (5.8) for z ∈ D(ζσ, r2, σ).
TENSORIAL FUNCTION THEORY
38
To pass to z in the larger disc D(ζσ, r0, σ), observe that for such z
u :=
ζσ
0
...
...
0
q(z − ζσ)
ζσ
. . .
· · ·
· · ·
. . .
. . .
ζσ
0
0
. . .
. . .
. . .
· · ·
0
...
0
q(z − ζσ)
ζσ
lies in the disc D(ζ(n+1)σ, r1, (n + 1)σ) ⊆ D(ζ(n+1)σ, r, (n + 1)σ). Conse-
quently
∆nfσ,σ,··· ,σ(ζσ, . . . , ζσ)(z − ζσ, . . . , z − ζσ) =
q−n∆nfσ,σ,··· ,σ(ζσ, . . . , ζσ)(q(z − ζσ), . . . , q(z − ζσ)) ≤ q−nM.
It follows that the series in Equation (5.8) converges absolutely and
uniformly on D(0, r0).
To see that the sum is fσ(ζσ + z), fix z ∈ D(0, r0, σ) and consider the
function h(λ) := fσ(ζσ +λz). By part (1) h is analytic in a disc centered
at the origin in the complex plane having radius bigger than 1. On the
other hand we may also form
g(λ) :=
∞Xk=0
∆kfσ,σ,··· ,σ(ζσ, . . . , ζσ)(λz, . . . , λz)
=
∞Xk=0
λk∆kfσ,σ,··· ,σ(ζσ, . . . , ζσ)(z, . . . , z),
which is also analytic in a disc centered at the origin of radius bigger
than 1. By what we have just shown, these two Banach space-valued
functions agree on a neighborhood of the origin in the complex plane.
Therefore, they agree on the intersection of their domains of definition
([HP74, Theorem 3.11.5]), which includes 1, i.e., Equation (5.8) is valid
throughout the disc D(0, r0, σ).
(cid:3)
Remark 5.14. As a special case of Theorem 5.13, we obtain a formula
that was inspired by [KV]. If σ is a normal representation of M, then
the subcategory Σ generated by σ is just the collection of finite mul-
tiples of σ. The collection of morphisms from mσ to nσ is just the
n× m matrices over σ(M)′. A ζ0 ∈ Eσ∗ generates an additive field over
Σ simply by setting ζkσ = ζ0 ⊕ ζ0 ⊕ · · · ⊕ ζ0 (k summands). Then, if
f = {fmσ}m∈N is a locally bounded, matricial function on the matricial
disc D(ζ, r), ζ = {ζmσ}m∈N, then Theorem 5.8 implies that for every
TENSORIAL FUNCTION THEORY
39
m ∈ N and every z in D(ζmσ, r, mσ),
∞Xk=0
fmσ(z) =
∆kfmσ(ζmσ)(z − ζmσ, . . . , z − ζmσ).
Suppose f = {fσ}σ∈Σ is a matricial function defined on a matricial disc
D(ζ, r) and suppose that f is locally uniformly bounded as in Theorems
(5.1) and (5.8). Then we have just seen that each fσ is Frechet differ-
entiable throughout D(ζσ, r, σ). By [HP74, Theorem 26.3.10] fσ can be
expanded in a unique power series about each point z ∈ D(ζσ, r, σ), and
the series converges at least in the largest open ball centered at z con-
tained in D(ζσ, r, σ). The terms of the power series are built from the
higher order Frechet derivatives of fσ. Recall from the general theory of
differentiable functions on Banach spaces (applied to our setting) that
the nth order Frechet derivative of fσ is a B(Hσ)-valued function, de-
noted Dnfσ, that is defined on D(ζσ, r, σ)×Eσ∗×Eσ∗×· · ·×Eσ∗ and has
the following properties: For each z, Dnfσ(z)(u1, . . . , un) is a bounded,
symmetric, multilinear function of (u1, . . . , un) ∈ Eσ∗×Eσ∗×· · ·×Eσ∗,
where the norm kDnfσ(z)(·, . . . ,·)k is locally bounded in z ∈ D(0, r, σ)
([HP74, Theorem 26.3.5]). If we write Dnfσ(z)(u) for Dnfσ(z)(u, . . . , u),
then Dnfσ(z)(u) is homogeneous of degree n in u and for each a ∈
D(ζσ, r, σ) there is an r′, depending on a, such that
(5.10)
fσ(z + u) =
∞Xn=0
1
n!
Dnfσ(z)(u),
with the convergence uniform for z − a < r′ and u < r′ ([HP74,
Theorem 3.17.1]).
When z = ζσ we find that
(5.11)
fσ(ζσ + u) =
∞Xn=0
1
n!
Dnfσ(ζσ)(u)
Since each of the summands in equations (5.8) and (5.11) is homoge-
neous, we conclude that
∆kfσ(ζσ)(z) =
1
k!
Dkfσ(ζσ)(z)
for all k and z for which the left hand side is well defined.
We may therefore summarize our analysis as follows.
Corollary 5.15. Let Σ be an additive subcategory of NRep(M), let
f = {fσ}σ∈Σ be a matricial family of functions defined on a matricial
disc D(ζ, r), and suppose that f is locally uniformly bounded on D(ζ, r).
TENSORIAL FUNCTION THEORY
40
Then, for every σ ∈ Σ, every k ≥ 0, and every z ∈ D(0, r, σ), Taylor's
Taylor derivatives and the Frechet derivatives of fσ are related by the
equation
(5.12)
∆kfσ(ζσ)(z, . . . , z) =
1
k!
Dkf (ζσ)(z),
and so ∆kfσ(ζσ)(z) is the restriction to the diagonal of D(ζσ, r, σ)k of
a bounded, symmetric, k-linear map on Eσ∗ × Eσ∗ · · · × Eσ∗. Conse-
quently, we may write
fσ(ζσ + z) =
=
∞Xk=0
∞Xk=0
∆kfσ(ζσ)(z)
1
k!
Dkf (ζσ)(z)
Remark 5.16. At this point we pause to take stock of what we have
proved, and with what hypotheses. First of all, we have proved the
first assertion of Theorem 5.1. Further, we identified the higher Frechet
derivatives of f with Taylor's Taylor derivatives (Equation (5.12)). We
did this with the minimal hypotheses that we have placed on our addi-
tive categories, vis., that they contain the natural injections and pro-
jections for finite direct sums.
To obtain the second assertion of Theorem 5.1, we will need the as-
sumptions stated there. Even though in some of the results to follow
we can we can get by with slightly weaker assumptions, for the remain-
der of this section, we shall assume:
Σ is full; every σ in Σ is faithful; and ζ is central.
The assumption that ζ is central allows us to assume that ζσ = 0 for
each σ. That is, we can translate the disc D(ζ, R) to D(0, R) whenever
ζ is central.
Lemma 5.17. Suppose Σ is a full additive subcategory of NRep(M),
and suppose that ζ = {ζσ}σ∈Σ is a central additive field on Σ. Suppose,
also, that f = {fσ}σ∈Σ is a matricial function define on the matricial
disc D(ζ, R), 0 < R ≤ ∞, and define g = {gσ}σ∈Σ by setting gσ(z) :=
fσ(z+ζσ). Then g is a matricial function on D(0, R) and for all positive
integers k,
∆kfσ(ζσ)(z − ζσ) = ∆kgσ(0)(z − ζσ).
Proof. The fact that g is matricial is immediate from the observation
that when ζ is central, I(σ × (z + ζσ), τ × (w + ζτ )) = I(σ × z, τ × w).
TENSORIAL FUNCTION THEORY
41
The equation for Taylor's derivatives is immediate from their definition,
Definition 5.5.
(cid:3)
We want to show that this symmetric multilinear map is the restriction
of a completely bounded map defined on the tensor product Eσ∗ ⊗
Eσ∗ · · · ⊗ Eσ∗, balanced over σ(M)′. To accomplish this, we need the
following technical lemma.
Lemma 5.18. Suppose f = {fσ}σ∈Σ is a matricial function defined on
a matricial disc centered at the origin D(0, r) = {D(0, r, σ}σ∈Σ. Suppose
that p and k are positive integers and that for 1 ≤ j ≤ k, integers l(j)
and m(j) are defined and satisfy
1 ≤ l(1) < m(1), l(2) < m(2), l(3) < · · · < m(k) ≤ p.
Suppose also that u1, . . . , uk are in Eσ∗ and let U be the p × p matrix
over Eσ∗ (viewed as an element of Epσ∗) whose (l(j), m(j)) entry is
uj, for all j, 1 ≤ j ≤ k, and whose other entries are all zero. Then
A := fpσ(U), viewed as a p × p matrix over B(Hσ), will have non zero
entries only in positions (l(j), m(j + s)) where m(j) = l(j + 1), m(j +
1) = l(j + 2), . . . , m(j + s − 1) = l(j + s). In these positions we will
have
Al(j),m(j+s) = ∆s+1fσ(0)(uj, . . . , uj+s).
Proof. Suppose first that, for every j, m(j) = l(j) + 1. Then the non
zero entries of U are all on the first diagonal above the main one and the
result follows easily from Lemma 5.8 (noting that ∆rf (0)(z1, . . . , zr) =
0 if one of the zi is 0). For the general case, let θ be a permutation
on {1, . . . , p} such that θ(l(1)) = 1, θ(m(j)) = θ(l(j)) + 1 for all j
and θ(l(j + 1)) = θ(m(j)) + m(j) − l(j + 1) for all 1 ≤ j ≤ p − 1.
Such θ always exists (but, in general, is not unique). Write S for
the permutation matrix associated with θ and consider U′ = SU(I ⊗
S−1). This matrix will have non zero entries only on the first diagonal
above the main diagonal and, thus, f (U′) = SAS−1 will be of the form
described above and it will follow that A satisfies the assertion of the
lemma.
(cid:3)
Recall that if V1, V2,· · · , Vk, and X are operator spaces and if ϕ : V1 ×
V2 × · · · × Vk → X is a multilinear map, then one writes ϕ(n) for the
multilinear map
ϕ(n) : Mn(V1) × · · · × Mn(Vk) → Mn(X)
defined by
ϕ(n)((α1 ⊗ v1),· · · , (αk ⊗ vk)) = α1α2 · · · αk ⊗ ϕ(v1, v2, . . . , vk)
TENSORIAL FUNCTION THEORY
42
for αi ∈ Mn(C) and vi ∈ Vi [ER00, Section 9.1].
Lemma 5.19. Let f = {fσ}σ∈Σ be a matricial family of functions
defined on a matricial E, Σ-family of discs {D(0, r, σ)}σ∈Σ, where Σ
is a full additive subcategory of NRep(M). For σ ∈ Σ, write ϕ :
Eσ∗ × Eσ∗ · · · × Eσ∗ → B(Hσ) for the map
ϕ(u1, . . . , uk) = ∆kfσ(0)(u1, . . . , uk),
ui ∈ Eσ∗, i = 1, 2,· · · , k.
Then, when Mn(Eσ∗) is identified with Enσ∗, we have
ϕ(n)(U1, . . . , Uk) = ∆kfσ(0)(U1, . . . , Uk)
(5.13)
for Ui ∈ Enσ∗.
Proof. Since the functions on both sides of (5.13) are k-linear, it suf-
fices to prove the lemma for matrices Ui that have only one non zero
entry. If U is such a matrix and the only non zero entry is u, which
lies in the (i, j) position, we write U = εi,j ⊗ u. So, we write Uj =
εr(j),s(j)⊗uj and, using the definition of ϕ(n), we have ϕ(n)(U1, . . . , Uk) =
εr(1),s(1) · · · εr(k),s(k) ⊗ ∆kfσ(0)(u1, . . . , uk). Thus
ϕ(n)(U1, . . . , Uk) = εr(1),s(k) ⊗ ∆kfσ(0)(u1, . . . , uk)
provided s(j) = r(j + 1) for all 1 ≤ j ≤ k−1, and it equals 0 otherwise.
In order to compute the right hand side of (5.13), we form the matrix
B :=
0 U1
0
. . .
0
...
...
0 · · ·
0
. . .
. . .
. . .
· · ·
0
...
0
· · ·
. . .
. . .
0 Uk
0
0
,
write A for the (k + 1) × (k + 1) matrix f(k+1)nσ(B) and note that
∆kfnσ(0)(U1, . . . , Uk) is the n × n block of A in the (1, k + 1) position.
If we now view A as a matrix of size n(k+1)×n(k+1) (over B(Hσ)), we
see that A satisfies the assumptions of Lemma 5.18, with p = n(k + 1),
l(1) = r(1), m(1) = n + s(1), l(2) = n + r(2) etc.
It follows from
that lemma that the only non zero entry in the upper-right n× n block
can be in the (l(1), m(k)) position and this will be non zero only if
m(1) = l(2), m(2) = l(3), . . . , m(k − 1) = l(k). Using our notation
here, this entry will be non zero if and only if s(j) = r(j + 1) for all
1 ≤ j ≤ k − 1. If this is the case, then by Lemma 5.18, this entry will
be ∆kfσ(0)(u1, . . . , uk). This completes the proof.
(cid:3)
TENSORIAL FUNCTION THEORY
43
Proposition 5.20. Let f = {fσ}σ∈Σ be a matricial family of functions
defined on a matricial disc D(0, r) = {D(0, r, σ)}σ∈Σ where Σ is a full,
additive subcategory of NRep(M). Suppose f is uniformly bounded in
norm by M so that, for every σ ∈ Σ, and for every z ∈ D(0, r, σ),
fσ(z) ≤ M. Then, for every k, the map
(5.14)
is a k-linear map, balanced over σ(M)′, and is completely bounded, with
∆kfσ(0)(·,· · · ,·) : Eσ∗ × Eσ∗ · · · × Eσ∗ → B(Hσ)
∆kfσ(0)(·,· · · ,·)cb ≤
M
rk .
Proof. With the notation preceding Lemma 5.19, we have
∆kfσ(0)(·,· · · ,·)cb = sup{∆kfσ(0)(n) : n ≥ 1}.
Note that in [ER00] this norm is denoted · mb but we follow the
notation in [BLM04] and in other places in the literature. Using
Lemma 5.19, it suffices to show that ∆kfσ(0)(·,· · · ,·) ≤ M
rk . For
this, consider ui ∈ D(0, 1, σ) and write u′
0
) ≤ M.
i := rui ∈ D(0, r, σ). Then
u′
1
0
. . .
k) ≤ f(k+1)σ(
∆kfσ(0)(u′
1, . . . , u′
0
...
...
0 · · ·
· · ·
. . .
. . .
0
0
0
. . .
. . .
. . .
· · ·
0
...
0
u′
k
0
By the k-linearity of the map, we get
∆kfσ(0)(u1, . . . , uk) ≤
M
rk ,
proving the complete norm estimate. The only thing left to prove is the
fact that the map is balanced, but this follows from Lemma 5.6.
(cid:3)
Recall that, in the discussion preceding Lemma 2.5, Zk is defined by
the formula, Zk(z) := z(k) = z(IE ⊗ z)· · · (IE⊗k−1 ⊗ z), for z ∈ Eσ∗.
For the statement of the following theorem it will be convenient to
use the natural extension of Zk to Eσ∗ × Eσ∗ × · · · × Eσ∗ and write
Zk(u1, . . . , uk) := u1(IE ⊗ u2)· · · (IE⊗k−1 ⊗ uk), for u1, . . . , uk in Eσ∗.
Theorem 5.21. Let f = {fσ}σ∈Σ be a matricial family of functions
defined on a matricial disc D(0, r) over a full, additive subcategory Σ
of faithful representations in NRep(M). Suppose that f is defined and
bounded uniformly M on D(0, r). Then, for every k, there is a unique
θk ∈ E⊗k, with θk ≤ M
(5.15)
∆kfσ(0)(u1, . . . , uk)h = Zk(u1, . . . , uk)(θk ⊗ h)
rk , such that
TENSORIAL FUNCTION THEORY
44
for every u1, . . . , uk in Eσ∗ and every h ∈ Hσ.
Proof. Consider the map ψ : Eσ × Eσ × · · · × Eσ → B(H) defined by
ψ(η1, . . . , ηk) = ∆kfσ(0)(η∗
k, . . . , η∗
1)∗.
k, . . . , η∗
k, . . . , η∗
1(I ⊗ b∗))∗ = (c∗∆kf (0)(η∗
For simplicity, we write here N for the von Neumann algebra σ(M)′.
Using Proposition 5.20, we see that this is a k-linear map, balanced over
σ(M)′, with norm not exceeding M
rk . Applying [BMP00, Theorem 2.3],
we find that it induces a linear, completely bounded map Ψ : Eσ ⊗hN
Eσ ⊗hN · · · ⊗hN Eσ → B(Hσ), where ⊗hN is the module Haagerup
tensor product, and Ψ ≤ M
rk . But Eσ ⊗hN Eσ ⊗hN · · · ⊗hN Eσ =
Eσ ⊗C ∗ Eσ ⊗C ∗ · · · ⊗C ∗ Eσ where ⊗C ∗ is the internal tensor product
of C ∗-correspondences (see [Ble97, Theorem 4.3] ). For b, c ∈ N and
η1, . . . , ηk ∈ Eσ, we have Ψ(b·η1⊗η2⊗· · ·⊗ηk·c) = Ψ((IEσ ⊗b)η1⊗η2⊗
· · ·⊗ ηkc) = ∆kfσ(0)(c∗η∗
1)(I ⊗
b∗))∗ = bΨ(η1⊗η2⊗· · ·⊗ηk)c (using Lemma 5.6). Thus, Ψ is a bimodule
map. Write F for the C ∗-correspondence Eσ ⊗C ∗ Eσ ⊗C ∗ · · · ⊗C ∗ Eσ
(over N = σ(M)′). Using the terminology of [MS98], we say that
(Ψ, ι) is a completely bounded covariant representation of F where ι
is the identity representation of σ(M)′ on Hσ. It follows from [MS98,
Lemma 3.5] that there is a bounded map Ψ : F ⊗ι Hσ → Hσ such that
Ψ(ϕF (·) ⊗ IH) = ι(·) Ψ and Ψ ≤ M
rk . Now note that (Eσ)⊗k is the
self-dual completion of F and, using Remark 1.8 in [Vis11], we have
F ⊗ι Hσ = (Eσ)⊗k⊗ι Hσ. Thus we can view Ψ as a map from (Eσ)⊗k⊗ι
Hσ into Hσ satisfying Ψ(ϕ(·)⊗IH) = ι(·) Ψ. Applying [MS04, Theorem
3.6 and Lemma 3.7], there is an element θk ∈ E⊗k that corresponds
to Ψ via the isomorphism ((Eσ)⊗k)ι ∼= E⊗k. More precisely, we have,
using Equation (3.1) in [MS04], for every η1, η2, . . . , ηk in Eσ and every
h ∈ Hσ,
(5.16) L∗
θk(IE⊗(k−1)⊗ηk)· · · (IE⊗η2)η1h = Ψ(ηk⊗ηk−1⊗· · ·⊗η1⊗h) =
= ∆kfσ(0)(η∗
Taking adjoints and writing ui for η∗
k)∗h.
1, . . . , η∗
i , we obtain the desired result. (cid:3)
The following corollary is now immediate, by Theorem 5.13 and The-
orem 5.21.
Corollary 5.22. Let f = {fσ}σ∈Σ be a uniformly bounded matri-
cial family of functions defined on a matricial disc D(0, r), where Σ
is a full additive subcategory of faithful representations in NRep(M).
Then there is a uniquely determined series θ ∼ Pk≥0 θk in T+((E))
TENSORIAL FUNCTION THEORY
45
with R(θ) ≥ r such that f is the family of tensorial power series
{Pk≥0 Zk(z)Lθk z ∈ D(0, r, σ)}σ∈Σ .
Remark 5.23. Corollary 5.22 and Lemma 5.17 immediately yield the
second assertion of Theorem 5.1
6. Series for matricial families of maps
In the previous section we studied matricial families of functions in
contexts where special generators are not present in the category under
consideration. In this section we focus on matricial families of maps
(as in Theorem 4.9). Many of the results proved for functions extend to
the setting of such families with only minor changes necessary. Indeed,
the formulas that go into defining ∆kfσ when f is a map are minor
variants of the formulas that enter into the definitions of ∆kfσ when f
is a function. One has only to replace equations like Cfσ(z) = fτ (w)C
with Cfσ(z) = fτ (w)(IF ⊗ C). To illustrate, recall that to say f is a
matricial family of maps means that for each pair, σ and τ in Σ, for
each z ∈ D(0,r,σ) , for each w ∈ D(0, r, τ ), and for each C ∈ I(σ, τ )
such that Cz = w(IE ⊗ C) we have
So, if (cid:20)z u
Cfσ(z) = fτ (w)(IF ⊗ C).
0 w(cid:21) lies in D(0, r, σ ⊕ τ ) and if fσ⊕τ ((cid:20)z u
0 w(cid:21)) = (cid:20)a11 a21
a21 a22(cid:21),
0 w (cid:19)(cid:18) IE ⊗ Iσ
(cid:19) ,
a21 a22(cid:19)(cid:18) IF ⊗ Iσ
(cid:19) ,
(cid:18) Iσ
0 (cid:19) z =(cid:18) z u
(cid:18)Iz
0(cid:19) fσ(z) =(cid:18)a11 a12
0
0
then because
we must have
(6.1)
which implies that a11 = fσ(z) and a21 = 0. Formula (6.1) is essen-
tially formula (5.3), and the other formulas in the analysis of matricial
functions have similar modifications for matricial maps.
In particu-
lar, one can proceed to define ∆fσ,τ (z, w)(u) as a12, and prove that
∆fσ,τ (z, w)(u) is linear in u. With matricial maps, however, ∆fσ,τ (z, w)(·)
is a map from Eσ∗ to F σ∗.
Once the distinction between matricial maps and functions is recog-
nized, the entire body of results that begins with part (1) of Definition
5.2, ends with Remark 5.16, and does not inolve the bimodule properties
of Eσ∗ (as a bimodule over σ(M)′), goes through mutatis mutandis for
matricial maps. In particular, the series expansions of functions found
in Corollary 5.15 make sense and remain valid for maps. However, the
TENSORIAL FUNCTION THEORY
46
notion of a "tensorial power series of maps" has not been defined and
the series expansion found in assertion (2) of Theorem 5.1 does not
make sense in the setting of maps. So our principal goal, Theorem 6.1,
is to exhibit the appropriate replacement and to give a definition of
tensorial power series of maps.
We shall assume our category Σ ⊆ NRep(M) is additive, full, and
has the property that every representation σ ∈ Σ is faithful. Although
matricial maps can be defined on arbitrary matricial sets, for simplicity
we shall restrict ourselves to matricial discs centered at the origin.
Thus we will consider matricial maps f = {fσ}σ∈Σ defined on an E, Σ-
matricial disc D(0, r), for some r. We shall assume that f maps into an
F, Σ-matricial disc. This is tantamount to assuming that f is uniformly
bounded in σ.
We will write MLM (E, F ) for the maps in L(E, F ) that are bimod-
ule maps. That is, T ∈ L(E, F ) lies in MLM (E, F ) if and only if
T (ϕE(a)ξb) = ϕF (a)T (ξ)b, for all a, b ∈ M. Our goal is to prove the
following theorem that complements part (2) of Theorem 5.1.
Theorem 6.1. Let E and F be two W ∗-correspondences over the
same W ∗-algebra, M, and suppose Σ is a full additive subcategory
of NRep(M) whose objects are all faithful representations of M.
If
f = {fσ}σ∈Σ is a matricial family of maps, mapping an E, Σ-disc
D(0, r, E) to an F, Σ-disc D(0, R, F ), then there is a uniquely defined
sequence of bimodule maps {Dkf}∞
k=0, where for each k, Dkf lies in
MLM (F, E⊗k), such that for every z ∈ D(0, r, σ),
fσ(z) = fσ(0) +Xk≥1
(6.2)
Zk(z)(Dkf ⊗ IHσ ).
Theorem 6.1 immediately suggests the following definition for a tenso-
rial power series of maps between two correspondence duals.
Definition 6.2. A map f defined from an open disc D(ζ0, r, σ) in Eσ∗
to F σ∗ is said to have a tensorial power series expansion on D(ζ0, r, σ)
in case there is a sequence {Θk}k≥0, with Θk in MLM (F, E⊗k), such
that
f (z) =Xk≥0
Zk(z − ζ0)(Θk ⊗ IHσ )
for all z ∈ D(ζ0, r, σ).
Under our standing hypotheses on Σ, the derivatives ∆kfσ(0) are really
k-linear bimodule maps on (Eσ∗)k mapping to F σ∗, when we view Eσ∗
as a bimodule over σ(M)′, and they are balanced over σ(M)′. The
principal obstacle to proving Theorem 6.1 turns out to be isolating the
TENSORIAL FUNCTION THEORY
47
dependence of ∆kfσ(0) on σ. This, in turn requires a careful study
of the dualities involved. Our analysis therefore rests on the following
two lemmas. Note that the first is couched in terms of maps from Eσto
F σ instead of maps from Eσ∗ to F σ∗. The reason is that as we noted
on page 12, Eσ and F σ are right correspondences over σ(M)′ and this
allows us to apply our Duality Theorem, [MS04, Theorem 3.6], to iden-
tify E and F with the second duals, Eσ,ι and F σ,ι, respectively, where
ι is the identity representation of σ(M)′ on Hσ. This identification will
prove central to what follows.
Lemma 6.3. If E and F are two W ∗ − correspondences over M and
if σ ∈ NRep(M) is a faithful normal representation, then:
(1) Every τ ∈σ(M )′ Lσ(M )′(Eσ, F σ) induces is a unique bounded map
τ∗ : F → E such that for every η ∈ Eσ and every θ ∈ F ,
(6.3)
η∗Lτ∗(θ) = τ (η)∗Lθ,
where Lθ : Hσ → F ⊗σ Hσ is given by Lθh = θ ⊗ h.
(2) The map τ∗ lies in MLM (F, E) and the correspondence τ 7→
τ∗ is contravariant and surjective, i.e., (τ1τ2)∗ = τ2∗τ1∗ and
(σ(M )′Lσ(M )′(Eσ, F σ))∗ =M LM (F, E).
Proof. Recall that Eσ := I(σ, σE ◦ ϕ) is a (right) W ∗-correspondence
over σ(M)′. Recall, too, that if ι denotes the identity representation of
σ(M)′ on Hσ, then the map WE : E → Eσ,ι such that WE(ξ)∗(η⊗ h) =
L∗
ξ(ηh), where ξ ∈ E, η ∈ Eσ and h ∈ Hσ, and where Lξ : Hσ → E⊗Hσ
is given by Lξh := ξ ⊗ h, is a correspondence isomorphism [MS04,
Theorem 3.6]. Similarly, one has a correspondence isomorphism WF :
F → F σ,ι. Also, note that, given g ∈ F σ,ι (so that g : Hσ → F σ ⊗ι Hσ
and g(bh) = (ϕι(b) ⊗ IHσ)gh, where b ∈ σ(M) and ϕι(·) is the left
action of σ(M) on F σ,ι) and τ ∈σ(M )′ Lσ(M )′(Eσ, F σ), we have that
(τ ∗ ⊗ IHσ )g ∈ Eσ,ι. Now define
τ∗(θ) = W −1
E ((τ ∗ ⊗ IHσ)WF (θ)).
It follows easily that τ∗ ∈M LM (F, E). Also, (τ ∗⊗I)WF (θ) = WE(τ∗(θ))
and, for every η ∈ Eσ and h ∈ Hσ, WE(τ∗(θ))∗(η⊗ h) = WF (θ)(τ ∗(η)⊗
h). Using the definitions of WE and WF , we find that L∗
τ∗(θ)(ηh) =
L∗
θ(τ (η)h). By taking adjoints, we obtain Equation (6.3). For the
uniqueness statement in (1), suppose that ξ ∈ E satisfies η∗Lξ =
τ (η)∗Lθ for all η ∈ Eσ. Then, for every η ∈ Eσ and h ∈ Hσ,
ξ(ηh) = L∗
τ∗(θ)(ηh) and, since the images of all η ∈ Eσ span E⊗Hσ, we
L∗
find that L∗
τ∗(θ) = L∗
ξ on E ⊗ Hσ, which in turn implies that τ∗(θ) = ξ.
TENSORIAL FUNCTION THEORY
48
The fact that the map τ 7→ τ∗ is contravariant follows easily from the
definition. The fact that it is surjective follows from duality.
(cid:3)
Lemma 6.4. Suppose E and F are two W ∗-correspondences over the
same W ∗-algebra M. For i = 1, 2, let σ1 and σ2 be two faithful represen-
tations in NRep(M) and let τi be a map in σi(M )′Lσi(M )′(Eσi, F σi). If
for every c, d ∈ I(σ1, σ2) we have τ1((I ⊗ c∗)ηd) = (I ⊗ c∗)τ2(η)d for ev-
ery η ∈ Eσ2, then τ1 = τ2. Consequently, under our standing hypothe-
ses that Σ is additive, full, and composed of faithful representations, if
{τσ}σ∈Σ is a family of maps, with τσ ∈σ(M )′ Lσ(M )′(Eσ, F σ), and if for
every σ1, σ2 ∈ Σ and every c, d ∈ I(σ1, σ2) we have τσ1((I ⊗ c∗)ηd) =
(I ⊗ c∗)τσ2(η)d for every η ∈ Eσ2, then the maps τσ∗, σ ∈ Σ, obtained
from Lemma 6.3 are independent of σ.
Proof. First note that (IE ⊗ c∗)ηd lies in Eσ1 for every c, d ∈ I(σ1, σ2)
and every η ∈ Eσ2.
If we write ηc,d for (IE ⊗ c∗)ηd, then by the
assumption, τ1(ηc,d) = (I ⊗ c∗)τ2(η)d. So, for every θ ∈ F , we have
η∗
c,dLτ1∗(θ) = τ1(ηc,d)∗Lθ = ((I ⊗ c∗)τ2(η)d)∗Lθ = d∗τ2(η)∗Lθ(I ⊗ c) =
d∗η∗Lτ2∗(θ)(I ⊗ c) = d∗η∗(I ⊗ c)Lτ2∗(θ) = η∗
(6.4)
for every η ∈ Eσ2 and c, d ∈ I(σ1, σ2). Since both σ1 and σ2 are normal
faithful representations of M, W{d(Hσ1) : d ∈ I(σ1, σ2)} = Hσ2 and
W{c∗(Hσ2) : c ∈ I(σ1, σ2)} = Hσ1. It follows from [MS04, Lemma 3.5]
that W{η(Hσ2) : η ∈ Eσ2} = E ⊗σ2 Hσ2 and, therefore, that
_{ηc,d(Hσ1) : η ∈ Eσ2, c, d ∈ I(σ1, σ2)} = E ⊗σ1 Hσ1.
c,dLτ2∗(θ). Thus
L∗
τ1∗(θ)ηc,d = L∗
τ2∗(θ)ηc,d
(6.5)
Combining these equations above we conclude that τ1∗ = τ2∗.
(cid:3)
Corollary 6.5. Under the hypotheses of Theorem 6.1, the matricial
map f has a well-defined Taylor derivative ∆fσ(0)(·) that is a map
from Eσ∗ to F σ∗ in σ(M )′Lσ(M )′(Eσ∗, F σ∗).
Its transposed map η 7→
(∆fσ(0)(η∗))∗ lies in σ(M )′Lσ(M )′(Eσ, F σ) and has the following depen-
dence on σ: There is a map Df ∈M LM (F, E) such that for every
σ ∈ Σ and every z ∈ Eσ∗,
∆fσ(0)(z) = z ◦ (Df ⊗ IHσ).
Proof. As we indicated at the beginning of this section, the existence
and linearity of ∆fσ(0)(·) is proved using the arguments of Lemma 5.2(1)
and Lemma 5.9. The fact that ∆fσ(0)(·) is a bimodule map uses the
arguments of Lemma 5.2(2) and (3).
(Note that we are entitled to
apply these because we are assuming Σ is full.) Thus the transposed
TENSORIAL FUNCTION THEORY
49
map η 7→ (∆fσ(0)(η)∗)∗ lies in σ(M )′Lσ(M )′(Eσ, F σ), also, and it fol-
lows from Lemma 6.3 and Lemma 6.4 that there is an element Df of
MLM (F, E) such that, for every σ ∈ Σ, every η ∈ Eσ and every θ ∈ F ,
η∗LDf (θ) = ∆fσ(0)(η∗)Lθ. Writing z in place of η∗ and applying the
two sides of this equality to h ∈ Hσ, we obtain the equation
∆fσ(0)(z)(θ ⊗ h) = z(Df (θ) ⊗ h) = z(Df ⊗ IH)(θ ⊗ h)
and the result follows.
(cid:3)
A similar analysis allows us to identify the dependence of ∆kfσ(0) on
σ, k > 1.
Lemma 6.6. Under the hypotheses of Theorem 6.1, we conclude that
for all k ∈ N and σ ∈ Σ, the Taylor derivative ∆kfσ(0)(·, . . . ,·) is
well defined and is a completely bounded k-linear, bimodule map from
Eσ∗ ×· · ·× Eσ∗ to F σ∗ that is balanced over σ(M)′. Moreover, there is
a uniquely determined map Dkf in MLM (F, E⊗k) such that for every
σ ∈ Σ and every z ∈ Eσ∗,
∆kfσ(0)(z) = Zk(z) ◦ (Dkf ⊗ IHσ).
Proof. The proof of the existance of ∆kfσ(0)(·, . . . ,·) uses the same
arguments as the analogous result for matricial families of functions,
as we already have mentioned. To get the bimodule properties of
∆kfσ(0)(·, . . . ,·), note that as in the proof of Theorem 5.21, ∆kfσ(0)(·, . . . ,·)
induces a bimodule map Ψ : Eσ ⊗C ∗ Eσ ⊗C ∗ · · · ⊗C ∗ Eσ → F σ. Using
Ψ in place of τ in Lemma 6.3, we obtain a map Ψ∗ ∈M LM (F, E⊗k).
Note that we do not know that Ψ induces a bimodule map on (Eσ)⊗k,
but all we needed in the proof of Lemma 6.3 is to know that the
map Ψ ⊗ IHσ : (Eσ)⊗k ⊗ι Hσ → F ⊗ι Hσ is well defined and inter-
twines the actions of σ(M)′. This holds here since (Eσ)⊗k ⊗ι H =
Eσ ⊗C ∗ Eσ ⊗C ∗ · · · ⊗C ∗ Eσ ⊗ι H, thanks to an observation of Viselter
[Vis11, Remark 1.8]. So we do indeed obtain a map Ψ∗ ∈M LM (F, E⊗k)
such that, for every θ ∈ F and every z ∈ Eσ∗,
Zk(z)LΨ∗(θ) = ∆kfσ(0)(z)Lθ.
The map Dkf that we want is Ψ∗.
(cid:3)
The proof of Theorem 6.1 is essentially complete. All that is necessary
is to observe that an analogue of the expansion (5.8) holds also for ma-
tricial families of maps. Since we have identified the Taylor derivatives
with the Dkf , the proof is complete.
We conclude by showing how the Schur class automorphisms we con-
sidered in [MS08] fit into theory we have developed here.
TENSORIAL FUNCTION THEORY
50
Example 6.7. For a central element γ ∈ D(Z(Eσ)), the map gγ :
D(Eσ∗) → D(Eσ∗) is defined by
gγ(z) = ∆γ(I − zγ)−1(γ∗ − z)∆−1
γ∗
where ∆γ = (IH − γ∗γ)1/2 ∈ Z(σ(M)′) = σ(Z(M)) and ∆γ∗ = (IE⊗H −
γγ∗)1/2 ∈ B(E ⊗σ Hσ). In [MS08, Lemma 4.20] it is shown that there
is a completely isometric automorphism αγ of H ∞(E) such that
\αγ(X)(z) = bX(gγ(z))
for z ∈ D(0,1,σ). Using Theorem 4.9 we see that gγ is a matricial
family of maps (with E = F ) and, thus, Corollary 6.1 applies and we
can write
(6.6)
gγ(z) = gγ(0) +Xk≥1
Zk(z)(Dkgγ ⊗ IHσ ).
On the other hand, using the expansion (I − zγ)−1 =Pk≥0(zγ)k, one
can write gγ as
(6.7)
gγ(z) = ∆γγ∗∆−1
γ∗ +Xk≥1
∆γzγzγ · · · γz∆γ∗
where, in the k-th term, z appears k times.
At first glance, it might not be evident that the terms in the expansion
(6.7) can be written in the form of the terms in (6.6). To deal with
the zeroth term, simply note that ∆γγ∗ = γ∗∆γ∗ so that ∆γγ∗∆−1
γ∗ =
γ∗ = I − γγ∗ ∈ (ϕE(M)⊗ I)′
γ∗ = gγ(0). For the first term, note that ∆2
(since γ ∈ Eσ) but also, for a ∈ σ(M)′, γ∗(I ⊗a) = aγ (since γ is in the
center of Eσ) and, thus, ∆2
γ∗ ∈ (IE ⊗ σ(M)′)′ = L(E) ⊗ IH. It follows
that one can write ∆γ∗ = X⊗IH for an X ∈ L(E)∩ϕE(M)′. Note also
that ∆γ is in Z(σ(M)) = σ(Z(M)) and (identifying ∆γ with σ−1(∆γ)),
the first term in (6.7) can be written ∆γz∆γ∗ = z(ϕE(∆γ) ⊗ IHσ )(X ⊗
IHσ) = z((ϕE(∆γ)X) ⊗ IHσ) so that, writing Dgγ = ϕE(∆γ)X ∈
L(E) ∩ ϕE(M)′ =M LM (E), we have the first term of (6.7) written in
the form of (6.6). (Note that ϕE(∆γ) ∈ ϕE(M)′ since ∆γ ∈ Z(M)).
For k ≥ 2 a similar computation is possible. It is a little less straightfor-
ward than the computation of Dgγ, but the case when k = 2 illustrates
amply what to do. With γ as above, define a map Y : E⊗H → E⊗2⊗H
by Y ζh = (IE ⊗ ζ)γh for every ζ ∈ Eσ and h ∈ Hσ. To see that
this map is well defined and bounded, compute Pi(I ⊗ ζi)γhi2 =
Pi,jh(I ⊗ ζi)γhi, (I ⊗ ζj)γhji = Pi,jhγ∗(I ⊗ ζ ∗
j ζi)γhi, hji. Since γ is
in the center of Eσ, the last expression equals Pi,jhγ∗γζ ∗
j ζihi, hji. As
ζj ∈ Eσ, this is equal toPi,jhζ ∗
j (ϕE(γ∗γ)⊗ IH )ζihi, hji = (ϕE(γ∗γ)⊗
TENSORIAL FUNCTION THEORY
51
IH)1/2Pi ζihi2. Thus Y ≤ γ. It is also easy to check that Y in-
tertwines IE ⊗ b and IE⊗2 ⊗ b for every b ∈ σ(M)′. Thus, there is some
Y0 ∈M LM (E, E⊗2) such that Y = Y0⊗IH . For z ∈ Eσ∗ and ζ ∈ Eσ, we
have zζ ∈ σ(M)′ and, since γ is central, we conclude that for h ∈ Hσ,
γzζh = (I ⊗ zζ)γh = (I ⊗ z)(Y0 ⊗ I)ζh. Thus γz = (IE ⊗ z)(Y0 ⊗ IHσ).
We now compute:
∆γzγz∆γ∗ = zγz(ϕE(∆γ) ⊗ IH )(X ⊗ IH) = Z2(z)(Y0ϕE(∆γ)X ⊗ IH),
which shows that D2gγ = Y0ϕE(∆γ)X.
References
[AF92]
Frank W. Anderson and Kent R. Fuller, Rings and categories of mod-
ules, second ed., Graduate Texts in Mathematics, vol. 13, Springer-
Verlag, New York, 1992. MR 1245487 (94i:16001)
[Arv98] William Arveson, Subalgebras of C ∗-algebras. III. Multivariable oper-
ator theory, Acta Math. 181 (1998), no. 2, 159 -- 228. MR 1668582
(2000e:47013)
David P. Blecher, A new approach to Hilbert C ∗-modules, Math. Ann.
307 (1997), no. 2, 253 -- 290. MR 1428873 (98d:46063)
[Ble97]
[BLM04] David P. Blecher and Christian Le Merdy, Operator algebras and their
modules -- an operator space approach, London Mathematical Society
Monographs. New Series, vol. 30, The Clarendon Press Oxford Univer-
sity Press, Oxford, 2004, Oxford Science Publications. MR 2111973
(2006a:46070)
[Coh06]
[BMP00] David P. Blecher, Paul S. Muhly, and Vern I. Paulsen, Categories of
operator modules (Morita equivalence and projective modules), Mem.
Amer. Math. Soc. 143 (2000), no. 681, viii+94. MR 1645699
(2000j:46132)
P. M. Cohn, Free ideal rings and localization in general rings, New
Mathematical Monographs, vol. 3, Cambridge University Press, Cam-
bridge, 2006. MR 2246388 (2007k:16020)
Kenneth R. Davidson and David R. Pitts, The algebraic structure of
non-commutative analytic Toeplitz algebras, Math. Ann. 311 (1998),
no. 2, 275 -- 303. MR 1625750 (2001c:47082)
Edward G. Effros and Zhong-Jin Ruan, Operator spaces, London Math-
ematical Society Monographs. New Series, vol. 23, The Clarendon Press
Oxford University Press, New York, 2000. MR 1793753 (2002a:46082)
[GLR85] P. Ghez, R. Lima, and J. E. Roberts, W ∗-categories, Pacific J. Math.
[DP98]
[ER00]
120 (1985), no. 1, 79 -- 109. MR 808930 (87g:46091)
[HKM11a] J. William Helton, Igor Klep, and Scott McCullough, Analytic mappings
between noncommutative pencil balls, J. Math. Anal. Appl. 376 (2011),
no. 2, 407 -- 428. MR 2747767 (2012b:46135)
[HKM11b]
, Proper analytic free maps, J. Funct. Anal. 260 (2011), no. 5,
1476 -- 1490. MR 2749435
TENSORIAL FUNCTION THEORY
52
[HP74]
[KK06]
[HKMS09] J. William Helton, Igor Klep, Scott McCullough, and Nick Slinglend,
Noncommutative ball maps, J. Funct. Anal. 257 (2009), no. 1, 47 -- 87.
MR 2523335 (2011b:47037)
Einar Hille and Ralph S. Phillips, Functional analysis and semi-groups,
American Mathematical Society, Providence, R. I., 1974, Third printing
of the revised edition of 1957, American Mathematical Society Collo-
quium Publications, Vol. XXXI. MR 0423094 (54 #11077)
Elias Katsoulis and David W. Kribs, The C ∗-envelope of the tensor
algebra of a directed graph, Integral Equations Operator Theory 56
(2006), no. 3, 401 -- 414. MR 2270844 (2007m:46088)
D. S. Kaliuzhnyi-Verbovetskyi, Noncommutative functions and fixed
point theorems, Slides from 2012 joint Mathematics meeting.
D. S. Kaliuzhnyi-Verbovetskyi and V. Vinnikov, Noncommutative ra-
tional functions, their difference-differential calculus and realizations,
ArXiv e-prints (2010).
D. S. Kaliuzhnyi-Verbovetskyi and V. Vinnikov, Foundations of non-
commutative function theory, In Preparation.
[KV]
[KV10]
[KVV]
[KVV06] Dmitry S. Kalyuzhnyi-Verbovetskiı and Victor Vinnikov, Non-
commutative positive kernels and their matrix evaluations, Proc. Amer.
Math. Soc. 134 (2006), no. 3, 805 -- 816 (electronic). MR 2180898
(2006f:47019)
[KVV09] Dmitry S. Kaliuzhnyi-Verbovetskyi and Victor Vinnikov, Singularities
of rational functions and minimal factorizations: the noncommutative
and the commutative setting, Linear Algebra Appl. 430 (2009), no. 4,
869 -- 889. MR 2489365 (2010a:47032)
[Lum86] Denis Luminet, A functional calculus for Banach PI-algebras, Pacific J.
[MS98]
[MS99]
[MS00]
[MS04]
[MS05]
[MS08]
[MS09]
Math. 125 (1986), no. 1, 127 -- 160. MR 860755 (88c:46060)
Paul S. Muhly and Baruch Solel, Tensor algebras over C ∗-
correspondences: representations, dilations, and C ∗-envelopes, J. Funct.
Anal. 158 (1998), no. 2, 389 -- 457. MR 1648483 (99j:46066)
, Tensor algebras, induced representations, and the Wold de-
composition, Canad. J. Math. 51 (1999), no. 4, 850 -- 880. MR 1701345
(2000i:46052)
, On the Morita equivalence of tensor algebras, Proc. London
Math. Soc. (3) 81 (2000), no. 1, 113 -- 168. MR 1757049 (2001g:46128)
, Hardy algebras, W ∗-correspondences and interpolation theory,
Math. Ann. 330 (2004), no. 2, 353 -- 415. MR 2089431 (2006a:46073)
, Hardy algebras associated with W ∗-correspondences (point eval-
uation and Schur class functions), Operator theory, systems theory and
scattering theory: multidimensional generalizations, Oper. Theory Adv.
Appl., vol. 157, Birkhäuser, Basel, 2005, pp. 221 -- 241. MR 2129649
(2006d:46071)
, Schur class operator functions and automorphisms of Hardy
algebras, Doc. Math. 13 (2008), 365 -- 411. MR 2520475 (2010g:46098)
, The Poisson kernel for Hardy algebras, Complex Anal. Oper.
Theory 3 (2009), no. 1, 221 -- 242. MR 2481905 (2011a:47166)
[MS11a]
, Morita transforms of tensor algebras, New York J. Math. 17A
(2011), 87 -- 100. MR 2782729
TENSORIAL FUNCTION THEORY
53
[MS11b]
[Muh97]
[Pop96]
[Pop06]
[Pop08]
, Representations of Hardy algebras: absolute continuity, in-
tertwiners, and superharmonic operators, Integral Equations Operator
Theory 70 (2011), no. 2, 151 -- 203. MR 2794388 (2012g:47218)
Paul S. Muhly, A finite-dimensional introduction to operator algebra,
Operator algebras and applications (Samos, 1996), NATO Adv. Sci.
Inst. Ser. C Math. Phys. Sci., vol. 495, Kluwer Acad. Publ., Dordrecht,
1997, pp. 313 -- 354. MR 1462686 (98h:46062)
Gelu Popescu, Non-commutative disc algebras and their representations,
Proc. Amer. Math. Soc. 124 (1996), no. 7, 2137 -- 2148. MR 1343719
(96k:47077)
Anal. 241 (2006), no. 1, 268 -- 333. MR 2264252 (2007h:47015)
, Free holomorphic functions on the unit ball of B(H)n, J. Funct.
, Noncommutative Berezin transforms and multivariable oper-
ator model theory, J. Funct. Anal. 254 (2008), no. 4, 1003 -- 1057.
MR 2381202 (2009j:47015)
[Rie74a] Marc A. Rieffel, Induced representations of C ∗-algebras, Advances in
Math. 13 (1974), 176 -- 257. MR 0353003 (50 #5489)
[Rie74b]
, Morita equivalence for C ∗-algebras and W ∗-algebras, J. Pure
[Tay72]
[Tay73]
[Vis11]
[Voi04]
[Voi10]
Appl. Algebra 5 (1974), 51 -- 96. MR 0367670 (51 #3912)
Joseph L. Taylor, A general framework for a multi-operator functional
calculus, Advances in Math. 9 (1972), 183 -- 252. MR 0328625 (48 #6967)
, Functions of several noncommuting variables, Bull. Amer.
Math. Soc. 79 (1973), 1 -- 34. MR 0315446 (47 #3995)
Ami Viselter, Covariant representations of subproduct systems, Proc.
Lond. Math. Soc. (3) 102 (2011), no. 4, 767 -- 800. MR 2793449
(2012d:46170)
Dan Voiculescu, Free analysis questions. I. Duality transform for the
coalgebra of ∂X : B, Int. Math. Res. Not. (2004), no. 16, 793 -- 822.
MR 2036956 (2005a:46140)
Dan-Virgil Voiculescu, Free analysis questions II: the Grassmannian
completion and the series expansions at the origin, J. Reine Angew.
Math. 645 (2010), 155 -- 236. MR 2673426 (2012b:46144)
Department of Mathematics, University of Iowa, Iowa City, IA 52242
E-mail address: [email protected]
Department of Mathematics, Technion, 32000 Haifa, Israel
E-mail address: [email protected]
|
1210.4670 | 1 | 1210 | 2012-10-17T08:55:44 | Completeness of $n$--tuple of projections in $C^*$--algebras | [
"math.OA",
"math.FA"
] | Let $(P_1,...,P_n)$ be an $n$--tuple of projections in a unital $C^*$--algebra $\aa$. We say $\pn$ is complete in $\aa$ if $\aa$ is the linear direct sum of the closed subspaces $P_1\aa,...,P_n\aa$. In this paper, we give some necessary and sufficient conditions for the completeness of $\pn$ and discuss the perturbation problem and topology of the set of all complete $n$--tuple of projections in $\aa$. Some interesting and significant results are obtained in this paper. | math.OA | math |
Completeness of n -- tuple of projections in C∗ -- algebras
Shanwen Hu∗ and Yifeng Xue†
Department of mathematics and Research Center for Operator Algebras
East China Normal University, Shanghai 200241, P.R. China
Abstract
Let (P1,··· , Pn) be an n -- tuple of projections in a unital C ∗ -- algebra A. We say
(P1,··· , Pn) is complete in A if A is the linear direct sum of the closed subspaces
P1A,··· , PnA. In this paper, we give some necessary and sufficient conditions
for the completeness of (P1,··· , Pn) and discuss the perturbation problem and
topology of the set of all complete n -- tuple of projections in A. Some interesting
and significant results are obtained in this paper.
2010 Mathematics Subject Classification: 46L05, 47B65
Key words: projection, idempotent, complete n -- tuple of projections
0 Introduction
Throughout the paper, we always assume that H is a complex Hilbert space with
inner product < ·,· >, B(H) is the C ∗ -- algebra of all bounded linear operators on H
and A is a C ∗ -- algebra with the unit 1. Let A+ denote the set of all positive elements
in A. It is well -- known that A has a faithful representation (ψ, Hψ) with ψ(1) = I
(cf. [8, Theorem 1.6.17] or [17, Theorem 1.5.36]). For T ∈ B(H), let Ran(T ) (resp.
Ker(T )) denote the range (resp. kernel) of T .
Let V1, V2 be closed subspaces in H such that H = V1 ∔ V2 = V ⊥
1 ∔ V2, that is, V1
and V2 is in generic position (cf. [6]). Let Pi be projection of H onto Vi, i = 1, 2. Then
H = Ran(P1) ∔ Ran(P2) = Ran(I − P1) ∔ Ran(P2). In this case, Halmos gave very
useful matrix representations of P1 and P2 in [6]. Following Halmos' work, Sunder
investigated in [14] the n -- tuple closed subspaces (V1,··· , Vn) in H which satisfying
the condition H = V1 ∔··· ∔ Vn. If let Pi be the projection of H onto Vi, i = 1,··· , n,
then the condition H = V1 ∔ ··· ∔ Vn is equivalent to H = Ran(P1) ∔ ··· ∔ Ran(Pn).
Now the question yields: when does the relation H = Ran(P1) ∔ ··· ∔ Ran(Pn)
hold for an n -- tuple of projections (P1,··· , Pn)? When n = 2, Buckholdtz proved
in [3] that Ran(P1) ∔ Ran(P2) = H iff P1 − P2 is invertible in B(H) iff I − P1P2 is
invertible in B(H) and iff P1+P2−P1P2 is invertible in B(H). More information about
two projections can be found in [2]. Koliha and Rakocevi´c generalized Buckholdtz's
∗E-mail: [email protected]
†E-mail: [email protected]
1
work to the set of C ∗ -- algebras and rings. They gave some equivalent conditions for
decomposition R = P R ∔ Q R or R = R P ∔ R Q in [9] and [10] for idempotent
elements P and Q in a unital ring R. They also characterized the Fredhomness of
the difference of projections on H in [11]. For n ≥ 3, the question remains unknown
so far. But there are some works concerning with this problem. For example, the
estimation of the spectrum of the finite sum of projections on H is given in [1] and
the C ∗ -- algebra generated by certain projections is investigated in [13] and [15], etc..
Let Pn(A) denote the set of n -- tuple (n ≥ 2) of non -- trivial projections in A and
put
PCn(A) = {(P1,··· , Pn) ∈ Pn(A) P1A ∔ ··· ∔ PnA = A}.
It is worth to note that if A = B(H) and (P1,··· , Pn) ∈ Pn(B(H)), then (P1,··· , Pn)
∈ PCn(B(H)) if and only if Ran(P1) ∔ ··· ∔ Ran(Pn) = H (see Theorem 1.2 below).
In this paper, we will investigate the set PCn(A) for n ≥ 3. The paper consists of
four sections. In Section 1, we give some necessary and sufficient conditions that make
(P1,··· , Pn) ∈ Pn(A) be in PCn(A). In Section 2, using some equivalent conditions
for (P1,··· , Pn) ∈ PCn(A) obtained in §1, we obtain an explicit expression of Pi1 ∨
··· ∨ Pik for {i1,··· , ik} ⊂ {1,··· , n}. We discuss the perturbation problems for
(P1,··· , Pn) ∈ PCn(A) in Section 3. We find an interesting result: if (P1,··· , Pn) ∈
Pi invertible in A, then kPiA−1Pjk < (cid:2)(n − 1)kA−1kkAk2(cid:3)−1,
Pn(A) with A =
i 6= j implies PiA−1Pj = 0, i 6= j, i, j = 1,··· , n in this section. We show in this
section that for given ǫ ∈ (0, 1), if (P1,··· , Pn) ∈ Pn(A) satisfies condition kPiPjk < ǫ,
then there exists an n -- tuple of mutually orthogonal projections (P ′
n) ∈ Pn(A)
such that kPi−P ′
ik < 2(n−1)ǫ, i = 1,··· , n, which improves a conventional estimate:
ik < (12)n−1n!ǫ, i = 1,··· , n (cf. [8]). In the final section, we will study the
kPi − P ′
topological properties and equivalent relations on PCn(A).
1,··· , P ′
n
Pi=1
1 Equivalent conditions for complete n -- tuples of projec-
tions in C∗ -- algebras
Let GL(A) (resp. U (A)) denote the group of all invertible (resp. unitary) elements
in A. Let Mk(A) denote matrix algebra of all k × k matrices over A. For any a ∈ A,
we set aA = {ax x ∈ A} ⊂ A.
Definition 1.1. An n -- tuple of projections (P1,··· , Pn) in A is called complete in A,
if (P1,··· , Pn) ∈ PCn(A).
Theorem 1.2. Let (P1,··· , Pn) ∈ Pn(A). Then the following statements are equiv-
alent:
(1) (P1,··· , Pn) is complete in A.
(2) Hψ = Ran(ψ(P1)) ∔ ··· ∔ Ran(ψ(Pn)) for any faithful representation (ψ, Hψ)
of A with ψ(1) = I.
2
(3) Hψ = Ran(ψ(P1)) ∔ ··· ∔ Ran(ψ(Pn)) for some faithful representation (ψ, Hψ)
of A with ψ(1) = I.
(4) Pj6=i
Pj + λPi ∈ GL(A), i = 1, 2,··· , n and ∀ λ ∈ [1 − n, 0).
Pj(cid:1) + Pi ∈ GL(A) for 1 ≤ i ≤ n and all λ ∈ C\{0}.
(5) λ(cid:0)Pj6=i
Pi=1
Pi ∈ GL(A) and PiA−1Pi = Pi, i = 1,··· , n.
Pi=1
Pi ∈ GL(A) and PiA−1Pj = 0, i 6= j, i, j = 1,··· , n.
(7) A =
(6) A =
n
n
(8) there is an n-tuple of idempotent operators (E1,··· , En) in A such that EiPi =
n
Ei, PiEi = Pi, i = 1,··· , n and EiEj = 0, i 6= j, i, j = 1,··· , n,
Ei = 1.
Pi=1
In order to show Theorem 1.2, we need following lemmas.
Lemma 1.3. Let B, C ∈ A+\{0} and suppose that λB + C is invertible in A for
every λ ∈ R\{0}. Then there is a non -- trivial orthogonal projection P ∈ A such that
B = (B + C)1/2P (B + C)1/2, C = (B + C)1/2(1 − P )(B + C)1/2.
Proof. Put D = B + C and Dλ = λB + C, ∀ λ ∈ R\{0}. Then D ≥ 0, D and Dλ are
all invertible in A, ∀ λ ∈ R\{0}.
Put B1 = D−1/2BD−1/2, C1 = D−1/2CD−1/2. Then B1 + C1 = 1 and
D−1/2DλD−1/2 = λB1 + C1 = λ + (1 − λ)C1 = (1 − λ)(λ(1 − λ)−1 + C1)
λ
1 − λ
is invertible in A for any λ ∈ R\{0, 1}. Since λ 7→
is a homeomorphism from
R\{0, 1} onto R\{−1, 0}, it follows that σ(C1) ⊂ {0, 1}. Note that B1 and C1 are all
non -- zero. So σ(C1) = {0, 1} = σ(B1) and hence P = B1 is a non -- zero projection in
A and B = D1/2P D1/2, C = D1/2(1 − P )D1/2.
Lemma 1.4. Let B, C ∈ A+\{0}. Then the following statements are equivalent:
(1) for any non -- zero real number λ, λB + C is invertible in A.
(2) B + C is invertible in A and B(B + C)−1B = B.
(3) B + C is invertible in A and B(B + C)−1C = 0.
(4) B + C is invertible in A and for any B′, C ′ ∈ A+ with B′ ≤ B and C ′ ≤ C,
B′(B + C)−1C ′ = 0.
3
Proof. (1)⇒(2) By Lemma 1.3, there is a non -- zero projection P in A such that
B = D1/2P D1/2, C = D1/2(1 − P )D1/2, where D = B + C ∈ GL(A). So
B(B + C)−1B = D1/2P D1/2D−1D1/2P D1/2 = B.
The assertion (2) ⇔ (3) follows from
B(B + C)−1B = B(B + C)−1(B + C − C) = B − B(B + C)−1C.
(3) ⇒ (4) For any C ′ with 0 ≤ C ′ ≤ C,
0 ≤ B(B + C)−1C ′(B + C)−1B ≤ B(B + C)−1C(B + C)−1B = 0,
we have B(B +C)−1C ′ = B(B +C)−1C ′1/2C ′1/2 = 0. This implies C ′(B +C)−1B = 0.
In the same way, we get that for any B′ with 0 ≤ B′ ≤ B, C ′(B + C)−1B′ = 0.
(4)⇒(3) is obvious.
(2)⇒(1) Set X = (B + C)−1/2B and Y = (B + C)−1/2C. Then X, Y ∈ A and
X ∗X = B, X + Y = (B + C)1/2. Thus, for any λ ∈ R\{0},
X + λY = (B + C)−1/2(B + λC)
(X + λY )∗(X + λY ) = ((1 − λ)X + λ(B + C)1/2)∗((1 − λ)X + λ(B + C)1/2)
= (1 − λ)2B + 2λ(1 − λ)B + λ2(B + C)
= B + λ2C
and consequently, (X + λY )∗(X + λY ) ≥ B + C if λ > 1 and (X + λY )∗(X + λY ) ≥
λ2(B + C) when λ < 1. This indicates that (X + λY )∗(X + λY ) is invertible in A.
Noting that B + C ≥ k(B + C)−1k−1 · 1, we have, for any λ ∈ R\{0},
(B + λC)2 = (X + λY )∗(B + C)(X + λY )
≥ k(B + C)−1k−1(X + λY )∗(X + λY ).
Therefore, B + λC is invertible in A, ∀ λ ∈ R\{0}.
Pibi.
n
Pi=1
...
. Then
Put I =
. . .
, X =
1
0
0
P1
0
...
0
Now we begin to prove Theorem 1.2.
(1)⇒(6) Statement (1) implies that there are b1,··· , bn ∈ A such that 1 =
and Y =
b1 0 ···
...
. . .
bn 0 ···
Pi=1
I = XY = XY Y ∗X ∗ ≤ kY k2XX ∗ = kY k2
··· Pn
0
···
...
. . .
0
···
0
. . .
0
...
0
0
n
Pi
4
and so that A =
n
Pi=1
Pi is invertible in A. Therefore, from A = P1A ∔ ··· ∔ PnA and
Pi = P1A−1Pi + ··· + PiA−1Pi + ··· + PnA−1Pi = 0 + ··· + 0
}
{z
i−1
,
+Pi + 0 + ··· + 0
}
{z
n−i
i = 1,··· , n, we get that Pi = PiA−1Pi, i = 1,··· , n.
(2)⇒(3) is obvious.
(3)⇒(4) Set Qi = ψ(Pi), i = 1,··· , n. From Hψ = Ran(Q1) ∔ ··· ∔ Ran(Qn),
Pi=1
we obtain idempotent operators F1,··· , Fn in B(Hψ) such that
Fi = I, FiFj = 0,
i 6= j and FiHψ = QiHψ, i, j = 1,··· , n. So FiQi = Qi, QiFi = Fi and FjQi = 0,
i 6= j, 1 ≤ i, j ≤ n. Using these relations, it is easy to check that
n
n
Xi=1
Xi=1
n
(cid:0)
(cid:0)
n
λiQi(cid:1)(cid:0)
λ−1
i F ∗
Xi=1
i Fi(cid:1)(cid:0)
n
Xi=1
λ−1
i F ∗
i Fi(cid:1) =
λiQi(cid:1) =
n
Xi=1
Xi=1
n
Fi = I,
F ∗
i = I,
for any non -- zero complex number λi, i = 1,··· , n. Particularly, for any λ ∈ [1−n, 0),
Qj(cid:1) + Qi(cid:1)−1 = λ−1Xj6=i
(cid:0)λ(cid:0)Xj6=i
Qj(cid:1) + Qi is invertible ψ(A), 1 ≤ i ≤ n by [17, Corollary 1.5.8]
F ∗
j Fj + F ∗
i Fi
in B(Hψ). Thus, λ(cid:0)Pj6=i
and so that λ(cid:0)Pj6=i
(4)⇒(5) Put Ai(λ) = Pj6=i
(Ai(λ))2 ≤ 2(cid:0)Xj6=i
Pj(cid:1) + Pi ∈ GL(A) since ψ is faithful and ψ(1) = I.
Pj + λPi, i = 1,··· , n, λ ∈ R\{0}, then
Pj(cid:1)2 + 2λ2Pi ≤ 2(n − 1)Xj6=i
≤ 2 max{n − 1, λ2}(P1 + ··· + Pn).
Pj + 2λ2Pi
So Ai(λ) is invertible in A, ∀ λ ∈ [1− n, 0) means that A = P1 +··· + Pn is invertible
in A. Consequently, Ai(λ) is invertible in A when λ > 0, ∀ 1 ≤ i ≤ n.
Now we show that Ai(λ) is invertible in A for i = 1,··· , n and λ < 1 − n. Put
A1i = PiAPi, A2i = PiA(1 − Pi), A4i = (1 − Pi)A(1 − Pi), i = 1,··· , n.
A4i(cid:21), i = 1,··· , n. Noting that
Express Ai(λ) as the form Ai(λ) = (cid:20)A1i + (λ − 1)Pi A2i
A4i is invertible in (1 − Pi)A(1 − Pi) (A ≥ kA−1k−1 · 1, A4i ≥ kA−1k−1(1 − Pi)) and
A∗
2i
Ai(λ)(cid:20)
Pi
−A−1
4i A∗
0
2i 1 − Pi(cid:21) = (cid:20)A1i − A2iA−1
4i A∗
0
A4i(cid:21) ,
2i + (λ − 1)Pi A2i
5
we get that Ai(λ) is invertible iff A1i − A2iA−1
i = 1,··· , n. Since A1i ≤ nPi, it follows that
4i A∗
2i + (λ − 1)Pi is invertible in PiAPi,
−A1i + A2iA−1
4i A∗
2i − (λ − 1)Pi ≥ (1 − n − λ)Pi + A2iA−1
4i A∗
2i ≥ (1 − n − λ)Pi
n
n
Pi=1
when λ < 1 − n, i = 1,··· , n. Therefore, Ai(λ) is invertible in A for λ < 1 − n and
i = 1,··· , n.
Applying Lemma 1.4 to Pj6=i
(5)⇒(6) and (6)⇒ (7) easily.
(7)⇒(8) Set Ei = PiA−1, i = 1··· , n. Then Ei is an idempotent elements in A
Ei = 1 and PiEi = Ei,
Pj and Pi, 1 ≤ i ≤ n, we can get the implications
(8)⇒(1) Let E1,··· , En be idempotent elements in A such that EiEj = δijEi,
Ei = 1 and EiPi = Pi, PiEi = Ei, i, j = 1,··· , n. Then EiA = PiA, i = 1,··· , n
and EiEj = 0, i 6= j, i, j = 1,··· , n. It is obvious that
EiPi = Pi, i = 1,··· , n.
Pi=1
and A = E1A ∔ ··· EnA = P1A ∔ ··· ∔ PnA.
Pi=1
resentation of A with ψ(1) = I. Put E′
E′
iE′
i = I and Ran(E′
Hψ = Ran(Q1) ∔ ··· ∔ Ran(Qn).
Remark 1.5. (1) Statement (3) in Theorem 1.2 can not be replaced by "for any
(8)⇒(2) Let E1,··· , En be idempotent elements in A such that EiEj = δijEi,
Ei = 1 and EiPi = Pi, PiEi = Ei, i, j = 1,··· , n. Let (ψ, Hψ) be any faithful rep-
i = ψ(Ei) and Qi = ψ(Pi), i = 1,··· , n. Then
i) = Ran(Qi), i, j = 1,··· , n. Consequently,
j = δijE′
i,
Pi=1
E′
n
n
4
Li=1
H and put A = B(H (4)),
, P2 =
, P3 =
0
0
I
I
I
0
0
.
I
1 ≤ i ≤ n, Pi − Pj6=i
Pj is invertible".
For example, let H (4) =
I
I
0
0
P1 =
Clearly, Pi − Pj6=i
is, (P1, P2, P3) is not complete in A.
Pi ∈ GL(A), then Pi6=j
∈ Pn(A), if
Pi=1
Pj is invertible, 1 ≤ i ≤ 3, but P2 + P3 − 2P1 is not invertible, that
(2) According to the proof of (3)⇒(4) of Theorem 1.2, we see that for (P1,··· , Pn)
Pi − λPi ∈ GL(A), ∀ 1 ≤ i ≤ n and λ > n − 1.
n
Corollary 1.6 ([3, Theorem 1]). Let P1, P2 be non -- trivial projections in B(H). Then
H = Ran(P1) ∔ Ran(P2) iff P1 − P2 is invertible in B(H).
Proof. By Theorem 1.2, H = Ran(P1) ∔ Ran(P2) implies that P1 − P2 ∈ GL(B(H)).
6
Conversely, if P1 − P2 ∈ GL(B(H)), then from 2(P1 + P2) ≥ (P1 − P2)2, we get
that P1 + P2 ∈ GL(B(H)) and so that P1 − λP2, P2 − λP1 ∈ GL(B(H)), ∀ λ > 1 by
Remark 1.5 (2). Thus, for any λ ∈ (0, 1], P1 − λP2 and P2 − λP1 are all invertible in
B(H). Consequently, H = Ran(P1) ∔ Ran(P2) by Theorem 1.2.
2 Some representations concerning the complete n -- tuple
of projections
We first statement two lemmas which are frequently used in this section and the later
sections.
Lemma 2.1. Let B ∈ A+ such that 0 ∈ σ(B) is an isolated point. Then there is a
unique element B† ∈ A+ such that
BB†B = B, B†BB† = B†, BB† = B†B.
Proof. The assertion follows from Proposition 3.5.8, Proposition 3.5.3 and Lemma
3.5.1 of [17].
Remark 2.2. The element B† in Lemma 2.1 is called the Moore -- Penrose inverse of
B. When 0 6∈ σ(B), B† , B−1. The detailed information can be found in [17].
The following lemma comes from [17, Lemma 3.5.5] and [4, Lemma 1]:
Lemma 2.3. Let P ∈ A be an idempotent element. Then
(1) P + P ∗ − 1 ∈ GL(A).
(2) R = P (P + P ∗ − 1)−1 is a projection in A satisfying P R = R and RP = P .
Moreover, if R′ ∈ A is a projection such that P R′ = R′ and R′P = P , then R′ = R.
Pi. By Theorem 1.2, A ∈ GL(A)
Let (P1,··· , Pn) ∈ PCn(A) and put A =
n
and Ei = PiA−1, 1 ≤ i ≤ n are idempotent elements satisfying conditions
Pi=1
EiEj = 0, i 6= j, EiPi = Pi, PiEi = Ei, i = 1,··· , n, and
n
Xi=1
Ei = 1.
i + Ei− 1)−1, 1 ≤ i ≤ n. So the C ∗ -- algebra C ∗(P1,··· , Pn)
By Lemma 2.3, Pi = Ei(E∗
generated by P1,··· , Pn is equal to the C ∗ -- algebra C ∗(E1,··· , En) generated by
E1,··· , En.
Put Qi = A−1/2PiA−1/2, i = 1,··· , n. Then QiQj = δijQi by Theorem 1.2,
i, j = 1,··· , n and
Qi = 1. Thus,
n
Pi=1
Pi = A1/2QiA1/2 and Ei = PiA−1 = A1/2QiA−1/2, i = 1,··· , n.
(2.1)
7
Proposition 2.4. Let (P1,··· , Pn) ∈ PCn(A) with A =
i = 1,··· , n, (cid:0) n
Pi=1
λiPi(cid:1)−1 = A−1(cid:0) n
Pi=1
λ−1
Proof. Keeping the symbols as above. We have
i Pi(cid:1)A−1.
Pi=1
n
n
Xi=1
(cid:0)
λiPi(cid:1)−1 = A−1/2(cid:0)
n
Xi=1
λ−1
i Qi(cid:1)A−1/2 = A−1(cid:0)
n
Pi=1
Pi. Then for any λi 6= 0,
λiQi(cid:1)A1/2. Thus,
n
λiPi = A1/2(cid:0) n
Pi=1
i Pi(cid:1)A−1.
Pr=1
Xi=1
λ−1
k
k
Now for i1, i2,··· , ik ∈ {1, 2,··· , n} with i1 < i2 < ··· < ik, put A0 =
Pr=1
Qir . Then A0, Q0 ∈ A and Q0 is a projection. From (2.1), A0 = A1/2Q0A1/2.
Q0 =
Thus, σ(A0)\{0} = σ(Q0AQ0)\{0} (cf.
[17, Proposition 1.4.14]). Since Q0AQ0 is
invertible in Q0AQ0, it follows that 0 ∈ σ(Q0AQ0) is an isolated point and so that
0 ∈ σ(A0) is also an isolated point. So we can define Pi1 ∨···∨Pik to be the projection
A†
0A0 ∈ A by Lemma 2.1. This definition is reasonable:
if P ∈ A is a projection such that P ≥ Pir , r = 1,··· , k, then P A0 = A0 and
Pir and
hence P A0A†
0 = A0A†
0 = (1 − A†
0, i.e., P ≥ Pi1 ∨ ··· ∨ Pik ; Since A0 ≥ Pir , we have
0A0)
0A0) ≥ (1 − A†
0A0)Pir (1 − A†
0A0)A0(1 − A†
and consequently, Pir (1 − A†
Proposition 2.5. Let (P1,··· , Pn) ∈ PCn(A) with A =
above and {j1,··· , jl} = {1,··· , n}\{i1,··· , ik} with j1 < ··· < jl. Then
0A0) = 0, that is, Pir ≤ Pi1 ∨ ··· ∨ Pik , i = 1,··· , k.
Pi. Let i1,··· , ik be as
Pi=1
n
k
k
Pi1 ∨ ··· ∨ Pik = A1/2(cid:2)(cid:0)
Xr=1
= (cid:0)
Xr=1
Pir(cid:1)(cid:2)(cid:0)
Qir(cid:1)A(cid:0)
Xr=1
Xr=1
Pir(cid:1)2 +
Qir(cid:1)(cid:3)−1A1/2
Xt=1
Pjt(cid:3)−1(cid:0)
k
k
l
(2.2)
(2.3)
k
Xr=1
Pir(cid:1).
Proof. Using the symbols Pi, Qi, Ei as above. According to (2.1),
k
k
k
k
Xr=1
Xr=1
Pir = A1/2(cid:0)
Thus (cid:0) k
Eir(cid:1)(cid:0) k
Pir(cid:1) =
Pr=1
Pr=1
Pr=1
Xr=1
Eir(cid:1)Pi1 ∨ ··· ∨ Pik = Pi1 ∨ ··· ∨ Pik , Pi1 ∨ ··· ∨ Pik(cid:0)
(cid:0)
Xr=1
Eir = A1/2(cid:0)
Eir = (cid:0) k
Pr=1
Qir(cid:1)A1/2,
Pr=1
Xr=1
Pir and
k
k
k
Pir(cid:1)A−1. Then we have
Xr=1
Eir(cid:1) =
Xr=1
k
k
Qir(cid:1)A−1/2.
Eir ,
according to the definition of Pi1 ∨ ··· ∨ Pik .
8
Since
k
Pr=1
Eir is an idempotent element in A, it follows from Lemma 2.3 that
k
k
Xr=1
Xr=1
Pi1 ∨ ··· ∨ Pik = (cid:0)
Eir(cid:1)(cid:2)
Noting that (cid:0) k
Qir(cid:1)A(cid:0) k
Qir(cid:1) ∈ GL(cid:0)(cid:0) k
Pr=1
Pr=1
Pr=1
is invertible in (cid:0) k
Qjt(cid:1)A(cid:0) k
Qjt(cid:1) and
Pt=1
Pt=1
Xr=1
Xr=1
ir + Eir ) − 1 = A−1/2(cid:2)(cid:0)
Xr=1
= A−1/2(cid:2)(cid:0)
Qir(cid:1)A + A(cid:0)
Xr=1
Qir(cid:1)A(cid:0)
(E∗
k
k
k
k
(E∗
ir + Eir ) − 1(cid:3)−1 ∈ A.
Qir(cid:1)A(cid:0) k
Qir(cid:1)(cid:1); (cid:0) l
Pr=1
Pt=1
(2.4)
Qjt(cid:1)
Qjt(cid:1)A(cid:0) k
Pt=1
k
Xr=1
Qir(cid:1)−(cid:0)
Qir(cid:1) − A(cid:3)A−1/2
Xt=1
Qjt(cid:1)A(cid:0)
Xt=1
l
l
Qjt(cid:1)(cid:3)A−1/2,
we obtain that
k
Xr=1
(cid:2)
(E∗
ir + Eir ) − 1(cid:3)−1
Xr=1
= A1/2(cid:2)(cid:2)(cid:0)
k
Qir(cid:1)A(cid:0)
k
Xr=1
Qir(cid:1)(cid:3)−1−(cid:2)(cid:0)
l
Xt=1
Qjt(cid:1)A(cid:0)
l
Xt=1
Qjt(cid:1)(cid:3)−1(cid:3)A1/2.
Pjt = A1/2(cid:0) l
Pt=1
Qjt(cid:1)A1/2 and(cid:0) k
Pr=1
Pir(cid:1)2
k
Note that
Combining this with (2.4), we can get (2.2).
l
Pir = A1/2(cid:0) k
Qir(cid:1)A1/2,
Pr=1
Pr=1
Pt=1
= A1/2(cid:0) k
Qir(cid:1)A(cid:0) k
Qir(cid:1)A1/2. Therefore,
Pr=1
Pr=1
Xt=1
Xr=1
Xr=1
Xr=1
Pjt(cid:3)−1(cid:0)
(cid:0)
Pir(cid:1)
Pir(cid:1)(cid:2)(cid:0)
Pir(cid:1)2 +
Xr=1
Xr=1
Xr=1
= A1/2(cid:0)
Qir(cid:1)(cid:0)(cid:2)(cid:0)
Qir(cid:1)A(cid:0)
Qir(cid:1)(cid:3)−1 +
= Pi1 ∨ ··· ∨ Pik
k
k
k
k
k
k
l
l
Xt=1
Qjt(cid:1)(cid:0)
k
Xr=1
Qir(cid:1)A1/2
by (2.2).
3 Perturbations of a complete n -- tuple of projections
Recall that for any non -- zero operator C ∈ B(H), the reduced minimum modulus
γ(C) is given by γ(C) = {kCxk x ∈ (Ker(C))⊥, kxk = 1} (cf. [17, Remark 1.2.10]).
We list some properties of the reduced minimum modulus as our lemma as follows.
Lemma 3.1 (cf. [17]). Let C be in B(H)\{0}, Then
9
(1) kCxk ≥ γ(C)kxk, ∀ x ∈ (Ker(C))⊥.
(2) γ(C) = inf{λ λ ∈ σ(C)\{0}}, where C = (C ∗C)1/2.
(3) γ(C) > 0 iff Ran(C) is closed iff 0 is an isolated point of σ(C) if 0 ∈ σ(C).
(4) γ(C) = kC −1k−1 when C is invertible.
(5) γ(C) ≥ kBk−1 when CBC = C for B ∈ B(H)\{0}.
For a ∈ A+, put β(a) = inf{λ λ ∈ σ(a)\{0}}. Combining Lemma 3.1 with the
n
H , H and H1 =
n
Li=1
T 2
T1T2
1
T 2
T2T1
2
···
···
TnT1 T2T2
faithful representation of A, we can obtain
Corollary 3.2. Let a ∈ A+. Then
(1) β(a) > 0 if and only if 0 ∈ σ(a) is isolated when a 6∈ GL(A).
(2) β(c) ≥ kck−1 when aca = a for some c ∈ A+\{0}.
Let E be a C ∗ -- subalgebra of B(H) with the unit I. Let (T1,··· , Tn) be an n -- tuple
Ran(Ti) ⊂
Ker(Ti) ⊂ H. Since H = Ran(Ti) ⊕ Ker(Ti), i = 1,··· , n, it
of positive operators in E with Ran(Ti) closed, i = 1,··· , n. Put H0 =
Li=1
follows that H0 ⊕ H1 = H. Put Tij = TiTj(cid:12)(cid:12)Ran(Tj ), i, j = 1,··· , n and set
T =
∈ Mn(E), T =
··· T1n
··· T2n
∈ B(H0),
···
···
··· Tnn
(3.1)
Clearly, H1 ⊂ Ker(T ) and it is easy to check that Ker(T ) = H1 when Ker( T ) = {0}.
Thus, in this case, T can be expressed as T = (cid:20) T 0
0(cid:21) with respect to the orthogonal
decomposition H = H0 ⊕ H1 and consequently, σ(T ) = σ( T ) ∪ {0}.
Lemma 3.3. Let (T1,··· , Tn) be an n -- tuple of positive operators in E with Ran(Ti)
closed, i = 1,··· , n. Let H0, H1, H be as above and T, T be given in (3.1). Suppose
that T is invertible in B(H0). Then
(1) σ( T ) = σ(cid:0) n
Pi=1
(2) 0 is an isolated point in σ(cid:0) n
Pi=1
(3) {T1a1,··· , Tnan} is linearly independent for any a1,··· , an ∈ E with Tiai 6= 0,
··· T1Tn
··· T2Tn
···
···
T 2
···
n
Ti(cid:1) if 0 ∈ σ(cid:0) n
Pi=1
T11 T12
T21 T22
···
···
Tn1 Tn2
T 2
i (cid:1)\{0}.
0
Ti(cid:1).
n
Li=1
i = 1,··· , n.
10
∈ Mn(E). Then ZZ ∗ =
T1
0
...
0
Proof. (1) Put Z =
Z ∗Z = T . Thus, σ(cid:0) n
Pi=1
··· Tn
···
0
...
. . .
···
0
i (cid:1)\{0} = σ(T )\{0} = σ( T ).
(2) According to (1), 0 is an isolated point of σ(cid:0) n
Pi=1
in E. So by Lemma 2.1, there is G ∈ E+ such that
T 2
T 2
i
n
Pi=1
0
. . .
0
and
T 2
i (cid:1) if
n
Pi=1
T 2
i
is not invertible
n
T 2
n
n
n
n
T 2
n
T 2
(cid:0)
T 2
T 2
T 2
T 2
Xi=1
Xi=1
Xi=1
Xi=1
Xi=1
i , G(cid:0)
i (cid:1)G = G, (cid:0)
Xi=1
i (cid:1).
i (cid:1)G(cid:0)
i (cid:1) =
i (cid:1)G ∈ E. Then P0 is a projection with Ran(P0) = Ker(cid:0) n
Put P0 = I −(cid:0) n
Pi=1
Pi=1
i (cid:1) = Ker(cid:0) n
Noting that Ker(cid:0) n
Ti(cid:1) =
Pi=1
Pi=1
Ti=1
Pi=1
T 2
1≤i≤n kTik)
i ≤ ( max
i (cid:1).
T 2
i ∈ GL((I −P0)E(I −P0))
Ti is invertible in
i (cid:1)G = G(cid:0)
with the inverse G and
Ti(cid:1) when 0 ∈ σ(cid:0) n
(I − P0)E(I − P0). Thus, 0 is an isolated point of σ(cid:0) n
Pi=1
Pi=1
i ∈ E+ such that TiT †
(3) By Lemma 3.1 (3) and Lemma 2.1, there is T †
Ti, we get that
Ker(Ti),
Pi=1
n
Pi=1
n
Pi=1
T 2
T 2
n
n
n
T †
i TiT †
i = T †
i , T †
i Ti = TiT †
i , i = 1,··· , n. Thus, Ran(Ti) = Ran(TiT †
n
Let a1,··· , an ∈ E with Tiai 6= 0, i = 1,··· , n such that
Pi=1
Ti(cid:1).
i Ti = Ti,
i ), i = 1,··· , n.
λiTiai = 0 for some
λ1,··· , λn ∈ C. For any ξ ∈ H, put x =
since T is invertible. Thus, λiTiT †
n
Li=1
λiTiT †
i aiξ ∈ H0. Then T x = 0 and x = 0
i aiξ = 0, ∀ ξ ∈ H and hence λi = 0, i = 1,··· , n.
The following result duo to Levy and Dedplanques is very useful in Matrix Theory:
Lemma 3.4 (cf.
[7]). Suppose complex n × n self -- adjoint matrix C = [cij ]n×n is
strictly diagonally dominant, that is, Pj6=icij < cii, i = 1,··· , n. Then C is invertible
and positive.
Proposition 3.5. Let T1,··· , Tn ∈ A+. Assume that
(1) γ = min{β(T1),··· , β(Tn)} > 0 and
(2) there exists ρ ∈ (0, γ] such that η = max{kTiTjk i 6= j, i, j = 1,··· , n} <
(n − 1)−1ρ2.
Then for any δ ∈ [η, (n − 1)−1ρ2), we have
(1) σ(cid:0) n
Pi=1
i (cid:1)\{0} ⊂ [ρ2 − (n − 1)δ, ρ2 + (n − 1)δ].
T 2
11
Li=1
··· S1n
··· S2n
···
···
··· Snn
Ti(cid:1).
Ti(cid:1) if 0 ∈ σ(cid:0) n
Pi=1
(2) 0 is an isolated point of σ(cid:0) n
Pi=1
(3) (cid:0) n
Ti(cid:1)A = T1A ∔ ··· ∔ TnA.
Pi=1
Proof. (1) Let (ψ, Hψ) be a faithful representation of A with ψ(1) = I. We may
assume that H = Hψ and E = ψ(A). Put Si = ψ(Ti), Sij = SiSj(cid:12)(cid:12)Ran(Sj ), i, j =
1,··· , n. Then max{kSiSjk 1 ≤ i 6= j ≤ n} = η and γ(Si) = β(Ti) by Lemma 3.1,
1 ≤ i ≤ n. Set H0 =
Ran(Si) and
n
S =
S11 S12
S21 S22
···
···
Sn1 Sn2
Then for any λ < ρ2 − (n− 1)δ, we have Pj6=ikSijk ≤ (n− 1)η < ρ2 − λ. It follows from
ρ2 − λ −kS12k ··· −kS1nk
··· −kS2nk
−kS21k
···
···
···
ρ2 − λ
−kSn1k −kSn2k ···
∈ B(H0), S0 =
Lemma 3.4 that S0 is positive and invertible. Therefore the quadratic form
ρ2 − λ
···
,
f (x1, x2,··· , xn) =
n
Xi=1
(ρ2 − λ)x2
i − 2 X1≤i<j≤n
kSijkxixj
is positive definite and hence there exists α > 0 such that for any (x1,··· , xn) ∈ Rn,
f (x1,··· , xn) ≥ α(x2
1 + ··· + x2
n).
So for any ξ =
i = 1,··· , n, by Lemma 3.1 and
n
Li=1
ξi ∈ H0, kSiξik ≥ γ(Si)kξik ≥ ρkξik, ξi ∈ Ran(Si) = (Ker(Si))⊥,
< ( S − λI)ξ, ξ > =
≥
n
Xi=1
Xi=1
n
n
λkξik2 + X1≤i<j≤n
Xi
kSiξik2 −
(ρ2 − λ)kξik2 − 2 X1≤i<j≤n
Xi=1
k
= f (kξ1k,··· ,kξkk) ≥ α
kξik2.
kSijkkξikkξjk
(< Sijξj, ξi > + < S∗
ijξi, ξj >)
Therefore, S − λI is invertible.
Similarly, for any λ > ρ2 + (n − 1)δ, we can obtain that λI − S is invertible.
So σ( S) ⊂ [ρ2 − (n − 1)δ, ρ2 + (n − 1)δ] ⊂ (0, ρ2 + (n − 1)δ] and consequently,
i (cid:1)\{0} = σ(cid:0) n
σ(cid:0) n
Pi=1
Pi=1
Ti(cid:1) = σ(cid:0) n
(2) Since σ(cid:0) n
Pi=1
Pi=1
i(cid:1)\{0} ⊂ [ρ2 − (n − 1)δ, ρ2 + (n − 1)δ] by Lemma 3.3.
Si(cid:1), the assertion follows from Lemma 3.3 (2).
T 2
S2
12
(3) By (2) and Lemma 2.1, (cid:0) n
Pi=1
Ti(cid:1)† ∈ A exists. Set E = (cid:0) n
Pi=1
Ti(cid:1)(cid:0) n
Pi=1
Ti(cid:1)†.
Ti(cid:1)A ⊂ T1A + ··· + TnA for E(cid:0) n
Pi=1
Obviously, EA = (cid:0) n
Ti(cid:1) =
Pi=1
Pi=1
Ti, we get that (1− E)Ti(1− E) ≤ (1− E)(cid:0) n
Ti(cid:1)(1− E) = 0, i.e.,
Pi=1
Ti = ETi, i = 1,··· , n. So TiA ⊂ EA, i = 1,··· , n and hence
From Ti ≤
Pi=1
Ti.
n
n
T1A + ··· + TnA ⊂ EA = (cid:0)
n
Xi=1
Ti(cid:1)A ⊂ T1A + ··· + TnA.
Since for any a1,··· , an ∈ A with Tiai 6= 0, i = 1,··· , n, {S1ψ(ai),··· , Snψ(an)}
is linearly independent in E, we have {T1a1,··· , Tnan} is linearly independent in A.
Therefore, (cid:0) n
Pi=1
Ti(cid:1)A = EA = T1A ∔ ··· ∔ TnA.
Let P1, P2 be projections on H. Buckholtz shows in [3] that Ran(P1) ∔ Ran(P2) =
H iff kP1 + P2 − Ik < 1. For (P1,··· , Pn) ∈ Pn(A), we have
Corollary 3.6. Let (P1,··· , Pn) ∈ Pn(A) with (cid:13)(cid:13)
(P1,··· , Pn) is complete in A.
Proof. For any i 6= j,
Pi=1
n
Pi − 1(cid:13)(cid:13) < (n − 1)−2. Then
kPiPjk2 = kPiPjPik ≤ (cid:13)(cid:13)Pi(cid:0)Xk6=i
Pk − 1(cid:1)Pi(cid:13)(cid:13) ≤ (cid:13)(cid:13)
= (cid:13)(cid:13)Pi(cid:0)
Pk(cid:1)Pi(cid:13)(cid:13)
Xk=1
Xk=1
n
n
Pk − 1(cid:13)(cid:13) <
1
(n − 1)2 .
Thus kPiPjk < (n − 1)−1. Noting that
ρ = min{β(P1),··· , β(Pn)} = 1, η = max{kPiPjk 1 ≤ i < j ≤ n} <
1
n − 1
,
we have (cid:0) n
Pi(cid:1)A = P1A ∔ ··· ∔ PnA by Proposition 3.5.
Pi=1
Pi − 1(cid:13)(cid:13) < (n − 1)−2, we have
From (cid:13)(cid:13)
Pi=1
A = P1A ∔ ··· ∔ PnA. Thus, (P1,··· , Pn) is complete in A.
Pi=1
n
n
Pi is invertible in A and so that
Combing Corollary 3.6 with Theorem 1.2 (3), we have
Corollary 3.7. Let P1,··· , Pn be non -- trivial projections in B(H) with (cid:13)(cid:13)
Pi−I(cid:13)(cid:13) <
(n − 1)−2. Then H = Ran(P1) ∔ ··· ∔ Ran(Pn).
Let (P1,··· , Pn) ∈ Pn(A). A well -- known statement says: "for any ǫ > 0, there
is δ > 0 such that if kPiPjk < δ, i 6= j, i, j = 1,··· , n, then there are mutually
orthogonal projections P ′
ik < ǫ, i = 1,··· , n". It may
n ∈ A with kPi − P ′
1,··· , P ′
Pi=1
n
13
be the first time appeared in Glimm's paper [5]. By using the induction on n, he
gave its proof. But how δ depends on ǫ is not given. Lemma 2.5.6 of [8] states this
statement and the author gives a slightly different proof. We can find from the proof
of [8, Lemma 2.5.6] that the relation between δ and ǫ is δ ≤
(12)(n−1)n!
ǫ
.
The next corollary will give a new proof of this statement with the relation δ =
ǫ
for ǫ ∈ (0, 1).
n
ǫ
2(n − 1)
Corollary 3.8. Let (P1,··· , Pn) ∈ Pn(A). Given ǫ ∈ (0, 1). If P1,··· , Pn satisfy
condition: kPiPjk < δ =
, 1 ≤ i < j ≤ n, then there are mutually orthogonal
projections P ′
2(n − 1)
1,··· , P ′
Pi=1
n ∈ A such that kPi − P ′
Pi. Noting that γ = min{β(P1),··· , β(Pn)} = 1, kPiPjk <
ik < ǫ, i = 1,··· , n.
Proof. Set A =
,
n − 1
1 ≤ i < j ≤ n and taking ρ = 1, we have σ(A)\{0} ⊂ [1 − (n − 1)δ, 1 + (n − 1)δ] by
Proposition 3.5 (1). So the positive element A† exists by Lemma 2.1. Set P = A†A =
AA† ∈ A. From AA†A = A and A†AA† = A†, we get that Pi ≤ P , i = 1,··· , n and
AP = P A = A, A†P = P A† = A†. So A ∈ GL(PAP ) with the inverse A† ∈ PAP .
Let σP AP (A†) stand for the spectrum of A† in PAP . Then
1
σP AP (A†) = σ(A†)\{0} = {λ−1 λ ∈ σ(A)\{0}}
⊂ [(1 + (n − 1)δ)−1, (1 − (n − 1)δ)−1],
(3.2)
Now by Proposition 3.5, PA = AA = P1A ∔ ··· ∔ PnA. Thus, by using Pi ≤ P ,
i = 1,··· , n, we have PAP = P1(PAP ) ∔ ··· ∔ Pn(PAP ) and then PiA†Pj = δijPi,
i = (A†)1/2Pi(A†)1/2 ∈ A, i = 1,··· , n. Then
i, j = 1,··· , n by Theorem 1.2. Put P ′
1,··· , P ′
P ′
n are mutually orthogonal projections and moreover, for 1 ≤ i ≤ n,
kP ′
i − Pik ≤ k(A†)1/2Pi(A†)1/2 − Pi(A†)1/2k + kPi(A†)1/2 − Pik
(3.3)
Note that 0 < (n − 1)δ < 1/2. Applying Spectrum Mapping Theorem to (3.2),
≤ (k(A†)1/2k + 1)k(A†)1/2 − Pk.
we get that k(A†)1/2k) ≤ (1 − (n − 1)δ)−1/2 < √2 and
kP − (A†)1/2k ≤ (1 − (n − 1)δ)−1/2 − 1 <
2
1 + √2
(n − 1)δ.
Thus kP ′
i − Pik < 2(n − 1)δ = ǫ by (3.3).
Applying Theorem 1.2 and Corollary 3.8 to an n -- tuple of linear independent unit
vectors, we have:
Corollary 3.9. Let (α1,··· , αn) be an n -- tuple of linear independent unit vectors in
Hilbert space H.
(1) There is an invertible, positive operator K in B(H) and an n -- tuple of mutually
orthogonal unit vectors (γ1,··· , γn) in H such that γi = Kαi, i = 1,··· , n.
14
(2) Given ǫ ∈ (0, 1).
If < αi, αj > <
, 1 ≤ i < j ≤ n, then there
exists an n -- tuple of mutually orthogonal unit vectors (β1,··· , βn) in H such
that kαi − βjk < 2ǫ, i = 1,··· , n.
2(n − 1)
ǫ
n
Proof. Set H1 = span{α1,··· , αn} and Piξ =< ξ, αi > αi, ∀ ξ ∈ H1, i = 1,··· , n.
Then (P1,··· , Pn) ∈ Pn(B(H1)) and Ran(P1) ∔ ··· ∔ Ran(Pn) = H1.
Pi is invertible in B(H1) and PiA−1
By Theorem 1.2, A0 =
1,··· , n. Put K = A−1/2
αi, i = 1,··· , n, where P0 is the
0
projection of H onto H ⊥
1 . It is easy to check that K is invertible and positive in B(H)
with γi = Kαi, i = 1,··· , n and (γ1,··· , γn) is an n -- tuple of mutually orthogonal
unit vectors. This proves (1).
Pi=1
+ P0 and γi = A−1/2
0 Pj = δijPi, i, j =
0
(2) Note that kPiPjk = < αi, αj > <
ik < ǫ, i = 1,··· , n. Put β′
Corollary 3.8, there are mutually orthogonal projections P ′
kPi − P ′
mutually orthogonal and kαi − β′
Then < βi, βj >= δijβi, i, j = 1,··· , n and
kαi − βik ≤ kαi − β′
ik < ǫ, i = 1,··· , n. Set βi = kβ′
ik + 1 − kβ′
ik < 2ǫ,
i = P ′
ǫ
2(n − 1)
, 1 ≤ i < j ≤ n. Thus, by
n ∈ A such that
1,··· , β′
n are
i, i = 1,··· , n.
i αi, i = 1,··· , n. Then β′
ik−1β′
1,··· , P ′
for i = 1,··· , n.
Now we give a simple characterization of the completeness of a given n -- tuple of
n
projections in C ∗ -- algebra A as follows.
Theorem 3.10. Let P1,··· , Pn be projections in A. Then (P1,··· , Pn) is complete
Pi is invertible in A and kPiA−1Pjk < (cid:2)(n − 1)kA−1kkAk2(cid:3)−1,
Pi=1
if and only if A =
∀ i 6= j, i, j = 1,··· , n.
Proof. If (P1,··· , Pn) is complete, then by Theorem 1.2, A is invertible in A and
PiA−1Pj = 0, ∀ i 6= j, i, j = 1,··· , n.
Now we prove the converse.
Put Ti = A−1/2PiA−1/2, 1 ≤ i ≤ n. Then
Ti = 1. Since Ti(A1/2PiA1/2)Ti = Ti,
we have β(Ti) ≥ kA1/2PiA1/2k−1 ≥ kAk−1, i = 1,··· , n by Corollary 3.2. Put
ρ = kAk−1. Then
Pi=1
n
kTiTjk ≤ kA−1kkPiA−1Pjk < (cid:2)(n − 1)kAk2(cid:3)−1 =
ρ2
n − 1
, i 6= j, i, j = 1,··· , n.
Thus by Proposition 3.5 (3), A = T1A ∔ ··· ∔ TnA. Note that TiA = A−1/2(PiA),
i = 1,··· , n. So P1A ∔ ··· ∔ PnA = A1/2A = A, i.e., (P1,··· , Pn) ∈ PCn(A).
Corollary 3.11. Let (P1,··· , Pn) ∈ PCn(A) and let (P ′
that kPi − P ′
then (P ′
ik < (cid:2)4n2(n− 1)kA−1k2(nkA−1k + 1)(cid:3)−1, i = 1,··· , n, where A =
n) ∈ Pn(A). Assume
Pi,
1,··· , P ′
Pi=1
n
1,··· , P ′
n) ∈ PCn(A).
15
Proof. Set B =
n
Pi=1
i . Since nkA−1k ≥ kAkkA−1k ≥ 1, it follows that kA − Bk <
P ′
1
2kA−1k
. Thus B is invertible in A with
kB−1k ≤
kA−1k
1 − kA−1kkA − Bk
< 2kA−1k, kB−1 − A−1k < 2kA−1k2kA − Bk.
Note that PiA−1Pj = 0, i 6= j, i, j = 1,··· , n, we have
kP ′
i B−1P ′
jk ≤ kP ′
i (B−1 − A−1)P ′
jk + k(P ′
i − Pi)A−1P ′
jk + kPiA−1(Pj − P ′
j)k
≤ 2kA−1k2kA − Bk + kA−1kkPi − P ′
<
<
1
1
ik + kA−1kkPj − P ′
jk
2n2(n − 1)kA−1k
(n − 1)kB−1kkBk2 .
So (P ′
1,··· , P ′
n) is complete in A by Theorem 3.10.
4 Some equivalent relations and topological properties
on PCn(A)
n
Let A be a C ∗ -- algebra with the unit 1 and let GL0(A) (resp. U0(A)) be the connected
component of 1 in GL(A) (resp. in U (A)). Set
Xi=1
PIn(A) = (cid:8)(P1,··· , Pn) ∈ Pn(A)
Xi=1
POn(A) = (cid:8)(P1,··· , Pn) ∈ Pn(A)
Definition 4.1. Let (P1,··· , Pn) and (P ′
(1) We say (P1,··· , Pn) is equivalent to (P ′
Pi ∈ GL(A)(cid:9)
Pi = 1, PiPj = 0, i 6= j, i, j = 1,··· , n(cid:9).
1,··· , P ′
n), denoted by (P1,··· , Pn) ∼
n) be in PCn(A).
1,··· , P ′
n
(P ′
1,··· , P ′
n), if there are U1,··· , Un ∈ A such that Pi = U ∗
i Ui, P ′
i = UiU ∗
i .
(2) (P1,··· , Pn) and (P ′
(P1,··· , Pn) ∼u (P ′
i = 1,··· , n.
1,··· , P ′
1,··· , P ′
n) are called to be unitarily equivalent, denoted by
n), if there is U ∈ U (A) such that U PiU ∗ = P ′
i ,
(3) (P1,··· , Pn) and (P ′
(P1,··· , Pn) ∼h (P ′
PCn(A) such that F (0) = (P1,··· , Pn) and F (1) = (P ′
n) are called homotopically equivalent, denoted by
n), if there exists a continuous mapping F : [0, 1] →
1,··· , P ′
1,··· , P ′
1,··· , P ′
n).
It is well -- know that
1,··· , P ′
n) ⇒ (P1,··· , Pn) ∼ (P ′
1,··· , P ′
n)
(P1,··· , Pn) ∼h (P ′
and if U (A) is path -- connected,
(P1,··· , Pn) ∼u (P ′
1,··· , P ′
n) ⇒ (P1,··· , Pn) ∼h (P ′
1,··· , P ′
n).
16
n
Lemma 4.2. Let (P1,··· , Pn) be in PCn(A) and C be a positive and invertible
element in A with PiC 2Pi = Pi, i = 1,··· , n. Then (CP1C,··· , CPnC) ∈ PCn(A)
and (P1,··· , Pn) ∼h (CP1C,··· , CPnC) in PCn(A).
Proof. From (CPiC)2 = CPiC 2PiC = CPiC, 1 ≤ i ≤ n, we have (CP1C,··· , CPnC)
Pi=1
Pi ∈ GL(A) and PiA−1Pi =
∈ Pn(A). (P1,··· , Pn) ∈ PCn(A) implies that A =
Pi, 1 ≤ i ≤ n by Theorem 1.2. So (CPiC)(cid:16) n
(CPiC)(cid:17)−1
Pi=1
(CPiC) = CPiA−1PiC and
hence (CP1C,··· , CPnC) ∈ PCn(A) by Theorem 1.2.
Put Ai(t) = C tPiC t and Bi(t) = C −tPiC −t, ∀ t ∈ [0, 1], i = 1,··· , n. Then
Qi(t) , Ai(t)Bi(t) = C tPiC −t is idempotent and Ai(t)Bi(t)Ai(t) = Ai(t), ∀ t ∈ [0, 1],
i = 1,··· , n. Thus Ai(t)A = Qi(t)A, ∀ t ∈ [0, 1], i = 1,··· , n.
By Lemma 2.3, Pi(t) , Qi(t)(Qi(t)+(Qi(t))∗−1)−1 is a projection in A satisfying
Qi(t)Pi(t) = Pi(t) and Pi(t)Qi(t) = Qi(t), ∀ t ∈ [0, 1], i = 1,··· , n. Clearly, Ai(t)A =
Qi(t)A = Pi(t)A, ∀ t ∈ [0, 1] and t 7→ Pi(t) is a continuous mapping from [0, 1] into
A, i = 1,··· , n. Thus, from
(C tP1C t)A ∔ ··· ∔ (C tPnC t)A = A,
∀ t ∈ [0, 1],
we get that F (t) = (P1(t),··· , Pn(t)) ∈ PCn(A), ∀ t ∈ [0, 1]. Note that F : [0, 1] →
PCn(A) is continuous with F (0) = (P1,··· , Pn). Note that Ai(1) = CPiC is a
projection with Ai(1)Qi(1) = CPiCCPiC −1 = Qi(1) and Qi(1)Ai(1) = Ai(1), i =
1,··· , n. So Pi(1) = Ai(1), i = 1,··· , n and F (1) = (CP1C,··· , CPnC). The
assertion follows.
For (P1,··· , Pn) ∈ PCn(A), A =
Pi ∈ GL(A) and Qi = A−1/2PiA−1/2 is
a projection with QiQj = 0, i 6= j, i, j = 1,··· , n (see Theorem 1.2), that is,
(Q1,··· , Qn) ∈ POn(A). Since C = A−1/2 satisfies the condition given in Lemma
4.2, we have the following:
n
Pi=1
Corollary 4.3. Let (P1,··· , Pn) ∈ PCn(A) and let (Q1,··· , Qn) be as above. Then
(P1,··· , Pn) ∼h (Q1,··· , Qn) in PCn(A).
Theorem 4.4. Let (P1,··· , Pn) and (P ′
statements are equivalent:
n) ∈ PCn(A). Then the following
1,··· , P ′
1,··· , P ′
n).
(1) (P1,··· , Pn) ∼ (P ′
(2) there is D ∈ GL(A) such that for 1 ≤ i ≤ n, PiDD∗Pi = Pi and P ′
(3) there is (S1,··· , Sn) ∈ PCn(A) such that
i = D∗PiD.
(P1,··· , Pn) ∼u (S1,··· , Sn) ∼h (P ′
1,··· , P ′
n).
17
n
n
Pi, A′ =
i = 1,··· , n. Put A =
Proof. The implication (3)⇒(1) is obvious. We now prove the implications (1)⇒(2)
and (2)⇒(3) as follows.
Pi=1
Xi=1
W ∗W = A′−1/2(cid:0)
Xi=1
= A′−1/2(cid:0)
i Ui = Pi, UiU ∗
(1)⇒ (2) Let Ui ∈ A be partial isometries such that U ∗
i and W = A−1/2(cid:0) n
Pi=1
PiU ∗
Xi=1
i(cid:1)A′−1/2
i UiPi(cid:1)A−1(cid:0)
i P ′
Xi=1
i(cid:1)A′−1/2 = A′−1/2(cid:0)
i P ′
P ′
i UiPiU ∗
i = P ′
i ,
i(cid:1)A′−1/2. Then
i P ′
i(cid:1)A′−1/2 = 1.
Pi=1
PiU ∗
P ′
P ′
P ′
n
n
n
n
Similarly, W W ∗ = 1. Thus, W ∈ U (A). Set D = A−1/2W A′1/2 ∈ GL(A). Then, for
1 ≤ i ≤ n,
D∗PiD = (cid:0)
n
Xi=1
P ′
i UiPi(cid:1)A−1PiA−1(cid:0)
n
Xi=1
PiU ∗
i(cid:1) = P ′
i P ′
i UiPiU ∗
i P ′
i = P ′
i
and PiDD∗Pi = Pi follows from (D∗PiD)2 = D∗PiD.
(2)⇒(3) Put U = (DD∗)−1/2D. Then U ∈ U (A). Set C = U ∗(DD∗)1/2U
and Si = U ∗PiU , 1 ≤ i ≤ n. Then (S1,··· , Sn) ∈ PCn(A) with (S1,··· , Sn)∼u
(P1,··· , Pn) and (P ′
n) = (CS1C,··· , CSnC).
Since SiC 2Si = U ∗PiDD∗PiU = Si, i = 1,··· , n, it follows from Lemma 4.2 that
1,··· , P ′
n) ∼h (S1,··· , Sn) in PCn(A).
1,··· , P ′
(P ′
Proposition 4.5. For Pn(A), PCn(A), PIn(A) and POn(A), we have
(1) PIn(A) is open in Pn(A).
(2) PCn(A) is a clopen subset of PIn(A).
(3) POn(A) is a strong deformation retract of PCn(A).
(4) PCn(A) is locally connected. Thus every connected component of PCn(A) is
path -- connected.
(5) (P1,··· , Pn), (P ′
1,··· , P ′
n) ∈ PCn(A) are in the same connected component iff
there is D ∈ GL0(A) such that Pi = D∗P ′
i D, i = 1,··· , n.
Proof. (1) Since h(P1,··· , Pn) =
and GL(A) is open in A, it follows that PIn(A) = h−1(GL(A)) is open in Pn(A).
Pi is a continuous mapping from Pn(A) into A
n
Pi=1
(2) Define F : PIn(A) → R by
F (P1,··· , Pn) = X1≤i<j≤n
(n − 1)(cid:13)(cid:13)
n
Xi=1
2(cid:13)(cid:13)(cid:0)
Pi(cid:13)(cid:13)
n
Xi=1
Pi(cid:1)−1(cid:13)(cid:13)(cid:13)(cid:13)Pi(cid:0)
n
Xi=1
Pi(cid:1)−1Pjk.
18
Clearly, F is continuous on PIn(A). By means of Theorem 3.10, we get that PCn(A) =
F −1((−1, 1)) is open in PIn(A) and PCn(A) = F −1({0}) is closed in PIn(A).
(3) Define the continuous mapping r : PCn(A) → POn(A) by
r(P1,··· , Pn) = (cid:0)A−1/2P1A−1/2,··· , A−1/2PnA−1/2(cid:1), A =
n
Xi=1
Pi.
by Theorem 1.2. Clearly, r(P1,··· , Pn) = (P1,··· , Pn) when (P1,··· , Pn) ∈ POn(A).
This means that POn(A) is a retract of PCn(A).
For any t ∈ [0, 1] and i = 1,··· , n, put
Hi(P1,··· , Pn, t) = A−t/2PiAt/2(A−t/2PiAt/2 + At/2PiA−t/2 − 1)−1.
Similar to the proof of Lemma 4.2, we have
H(P1,··· , Pn, t) = (H1(P1,··· , Pn, t),··· , Hn(P1,··· , Pn, t))
is a continuous mapping from PCn(A) × [0, 1] to PCn(A) with H(P1,··· , Pn, 0) =
(P1,··· , Pn) and H(P1,··· , Pn, 1) = r(P1,··· , Pn). Furthermore, when (P1,··· , Pn)
∈ POn(A), A = 1. In this case, H(P1,··· , Pn, t) = (P1,··· , Pn), ∀ t ∈ [0, 1]. There-
fore, POn(A) is a strong deformation retract of PCn(A).
(4) Let (P1,··· , Pn) ∈ PCn(A). Then by Corollary 3.11, there is δ ∈ (0, 1/2)
such that for any (R1,··· , Rn) ∈ Pn(A) with kPi − Rik < δ, 1 ≤ i ≤ n, we have
(R1,··· , Rn) ∈ PCn(A).
Let (R1,··· , Rn) ∈ PCn(A) with kPj − Rjk < δ, i = 1,··· , n. put Pi(t) = Pi,
Ri(t) = Ri and ai(t) = (1 − t)Pi + tRi, ∀ t ∈ [0, 1], i = 1,··· , n. Then Pi, Ri, ai
are self -- adjoint elements in C([0, 1],A) = B and kPi − aik = max
t∈[0,1]kPi − ai(t)k < δ,
i = 1,··· , n.
It follows from [17, Lemm 6.5.9 (1)] that there exists a projection
fi ∈ C ∗(ai) (the C ∗ -- subalgebra of B generated by ai) such that kPi−fik ≤ kPi−aik <
δ, i = 1,··· , n. Thus, kPi − fi(t)k < δ, i = 1,··· , n and consequently, F (t) =
(f1(t),··· , fn(t)) is a continuous mapping of [0, 1] into PCn(A). Since ai(0) = Pi,
ai(1) = Ri and fi(t) ∈ C ∗(ai(t)), ∀ t ∈ [0, 1], we have f (0) = (P1,··· , Pn) and
f (1) = (R1,··· , Rn). This means that PCn(A) is locally path -- connected.
(5) There is a continuous path P (t) = (P1(t),··· , Pn(t)) ∈ PCn(A), ∀ t ∈ [0, 1]
such that P (0) = (P1,··· , Pn) and P (1) = (P ′
n). By [12, Corollary 5.2.9.],
there is a continuous mapping t 7→ Ui(t) of [0, 1] into U (A) with Ui(0) = 1 such that
Pi(t) = Ui(t)P1U ∗
1,··· , P ′
i (t), ∀ t ∈ [0, 1] and i = 1,··· , n. Set
i (t)Pi(t)(cid:17)(cid:16) n
Xi=1
i (t)(cid:17)1/2
Pi(cid:17)−1/2(cid:16) n
Xi=1
W (t)(cid:16) n
Pi=1
W (t) = (cid:16) n
Xi=1
Pi(cid:17)−1/2
PiU ∗
and D(t) = (cid:16) n
Pi=1
, ∀ t ∈ [0, 1]. Then W (t) ∈ U (A)
with W (0) = 1, D(t) ∈ GL(A) with D(0) = 1 and W (t), D(t) are all continuous on
Ui(t)PiU ∗
Ui(t)PiU ∗
i (t)(cid:17)−1/2
19
[0, 1] with D∗(t)PiD(t) = Pi(t) (see the proof of (1)⇒(2) in Theorem 4.4), ∀ t ∈ [0, 1]
and i = 1,··· , n. Put D = D(1). Then D ∈ GL0(A) and D∗PiD = P ′
i , i = 1,··· , n.
i , i = 1,··· , n. Then
i U ∗ = (DD∗)1/2Pi(DD∗)1/2,
Conversely, if there is D ∈ GL0(A) such that D∗PiD = P ′
U = (DD∗)−1/2D ∈ U0(A) and PiDD∗Pi = Pi, U P ′
1U ∗,··· , U P ′
i = 1,··· , n. Thus, (P ′
n) ∼h (U P ′
1,··· , P ′
nU ∗) and
((DD∗)1/2P1(DD∗)1/2,··· , (DD∗)1/2Pn(DD∗)1/2) ∼h (P1,··· , Pn)
by Lemma 4.2. Consequently, (P ′
n) ∼h (P1,··· , Pn).
As ending of this section, we consider following examples:
1,··· , P ′
n
n−1
Pi=1
Pi=1
Example 4.6. Let A = Mk(C), k ≥ 2. Define a mapping ρ : PCn(A) → Nn−1 by
ρ(P1,··· , Pn) = (Tr(P1),··· , Tr(Pn−1)), where 2 ≤ n ≤ k and Tr(·) is the canonical
trace on A.
By Theorem 1.2, (P1,··· , Pn) ∈ PCn(A) means that A =
Pi ∈ GL(A)
and (A−1/2P1A−1/2,··· , A−1/2PnA−1/2) ∈ POn(A). Put Qi = A−1/2PiA−1/2, i =
1,··· , n. Since (P1,··· , Pn) ∼h (Q1,··· , Qn) by Corollary 4.3, it follows that Tr(Pi) =
Tr(Qi), i = 1,··· , n and Tr(A) = k. Thus Tr(Pn) = k −
Pi.
Note that U (A) is path -- connected. So, for (P1,··· , Pn), (P ′
n) ∈ PCn(A),
1,··· , P ′
n) are in the same connected component if and only if
1,··· , P ′
n).
(P1,··· , Pn) and (P ′
ρ(P1,··· , Pn) = ρ(P ′
The above shows that PCk(A) is connected and PCn(A) is not connected when
k ≥ 3 and 2 ≤ n ≤ k − 1.
Example 4.7. Let H be a separable complex Hilbert space and K(H) be the C ∗ --
algebra of all compact operators in B(H). Let A = B(H)/K(H) be the Calkin algebra
and π : B(H) → A be the quotient mapping. Then PCn(A) is path -- connected.
In fact, if (P1,··· , Pn), (P ′
1,··· , Q′
1,··· , Q′
n) ∈ PCn(A), then we can find (Q1,··· , Qn),
n) ∈ POn(A) such that (P1,··· , Pn) ∼h (Q1,··· , Qn) and (P ′
(Q′
n) ∼h
(Q′
n) by Corollary 4.3. Since B(H) is of real rank zero, it follows from
[17, Corollary B.2.2] or [16, Lemma 3.2] that there are projections R1,··· , Rn and
R′
1,··· , R′
n in B(H) such that π(Ri) = Qi, π(R′
1,··· , P ′
1,··· , P ′
1,··· , P ′
i) = Q′
RiRj = δijRi, R′
iR′
j = δijR′
i, i, j = 1,··· , n,
n
i, i = 1,··· , n and
Xi=1
Xi=1
Ri =
R′
n
i = I.
n
Pi=1
1,··· , R′
Vi. Then V ∈ U (B(H)) and V RiV ∗ = R′
Note that R1,··· , Rn, R′
tors V1,··· , Vn in B(H) such that V ∗
V =
U (A). Then (U Q1U ∗,··· , U QnU ∗) = (Q′
is path -- connected, we have (Q1,··· , Qn) ∼h (Q′
(P1,··· , Pn) ∼h (P ′
n 6∈ K(H). So there are partial isometry opera-
i, i = 1,··· , n. Put
i, i = 1,··· , n. Put U = π(V ) ∈
1,··· , Qn) in POn(A). Since U (B(H))
n) in PCn(A). Finally,
n). This means that PCn(A) is path -- connected.
i Vi = Ri, ViV ∗
1,··· , Q′
1,··· , P ′
i = R′
20
References
[1] P.E. Bjørstad and J. Mandel, On the spectra of sums of orthogonal projections
with applications to parallel computing, BIT, 31 (1991), 76 -- 88.
[2] A. Bottcher and I.M. Spitkovsky, A gentle guide to the basics of two projections
theory, Linear Algebra Appl., 432 (2010), 1412 -- 1459.
[3] D. Buckholtz, Hilber space idempotents and involutions, Proc. Amer. Math.
Soc., 128 (2000), 1415 -- 1418.
[4] G. Chen and Y. Xue, The expression of generalized inverse of the perturbed
operators under type I perturbation in Hilbert spaces, Linear Algebra Appl.,
285 (1998), 1 -- 6.
[5] J. Glimm, On a certain class of operator algebras, Trans. Amer. Math. Soc., 95
(1960), 318 -- 340.
[6] P. Halmos, Two subspace, Tran. Amer. Math. Soc., 144 (1969), 381 -- 389.
[7] R.A. Horn and R. Johnson, Matrix Analysis, Cambridge University Press, 1986.
[8] H. Lin, An Intruduction to the Classification of Amenable C ∗-Algebras, World
Scintific, 2001.
[9] J.J. Koliha and V. Rakocevi´c, Invertibility of the sum of idempotents, Linear
and Multilinear Algebra, 50 (2002), 285 -- 292.
[10] J.J. Koliha and V. Rakocevi´c, Invertibility of the difference of idempotents, Lin-
ear and Multilinear Algebra 51 (2003), 97 -- 110.
[11] J.J. Koliha and V. Rakocevi´c, Fredholm properties of the difference of orthogonal
projections in a Hilbert space, Integr. Equ. Oper. Theory, 52 (2005), 125 -- 134.
[12] N.E. Wegge -- Olsen, K -- Theory and C*-Algebras, A Friendly Approach, OUP,
1993.
[13] T. Shulman, On universal C ∗-algebras generated by n projections with scalar
sum, Amer. Math. Soc. 137(1) (2009), 115 -- 122.
[14] V. S. Sunder, N subspaces, Can. J. Math., XL (1) (1988), 38 -- 54.
[15] N.L. Vasilevski, C ∗ -- algebras generated by orthogonal projections and their ap-
plications, Integr. Equ. Oper. Theory, 31 (1998), 113 -- 132.
[16] Y. Xue, The reduced minimum modulus in C ∗ -- algebras, Integr. Equ. Oper.
Theory, 59 (2007), 269 -- 280.
[17] Y. Xue, Stable Perturbations of Operators and Related Topics, World Scientific,
2012.
21
|
1302.5922 | 1 | 1302 | 2013-02-24T16:33:25 | Factors from trees | [
"math.OA"
] | We construct factors of type $\tn$ for $n\in\NN, n\geq 2$ from group actions on homogeneous trees and their boundaries. Our result is a discrete analogue of a result of R.J Spatzier, where the hyperfinite factor of type $\tone$ is constructed from a group action on the boundary of the universal cover of a manifold. | math.OA | math |
FACTORS FROM TREES
JACQUI RAMAGGE AND GUYAN ROBERTSON
Abstract. We construct factors of type III1/n for n ∈ N, n ≥ 2
from group actions on homogeneous trees and their boundaries.
Our result is a discrete analogue of a result of R.J Spatzier, namely [S,
Proposition 1], where the hyperfinite factor of type III1 is con-
structed from a group action on the boundary of the universal
cover of a manifold.
1. Introduction
Let Γ be a group acting simply transitively on the vertices of a ho-
mogeneous tree T of degree n + 1 < ∞. Then, by [FTN, Ch. I, Theo-
rem 6.3],
Γ ∼= Z2 ∗ · · · ∗ Z2 ∗ Z ∗ · · · ∗ Z
where there are s factors of Z2, t factors of Z, and s + 2t = n + 1. Thus
Γ has a presentation
Γ = (cid:10)a1, . . . , as+t : a2
i = 1 for i ∈ {1, . . . , s}(cid:11) ,
we can identify the Cayley graph of Γ constructed via right multipli-
cation with T and the action of Γ on T is equivalent to the natural
action of Γ on its Cayley graph via left multiplication.
We can associate a natural boundary to T , namely the set Ω of semi-
infinite reduced words in the generators of Γ. The action of Γ on T
induces an action of Γ on Ω.
For each x ∈ Γ, let
Ωx = {ω ∈ Ω : ω = x · · · }
be the set of semi-infinite reduced words beginning with x. The set
{Ωx}x∈Γ is a set of basic open sets for a compact Hasudorff topology
on Ω. Denote by x the length of a reduced expression for x. Let
V m = {x ∈ Γ : x = m} and define Nm = V m. Then Ω is the disjoint
union of the Nm sets Ωx for x ∈ V m.
We can also endow Ω with the structure of a measure space. Ω has
a unique distinguished Borel probability measure ν such that
ν (Ωx) =
1
n + 1 (cid:18) 1
n(cid:19)x−1
1
2
JACQUI RAMAGGE AND GUYAN ROBERTSON
for every nontrivial x ∈ Γ. The sets Ωx, x ∈ Γ generate the Borel
σ-algebra.
This measure ν on Ω is quasi-invariant under the action of Γ, so that
Γ acts on the measure space (Ω, ν) and enabling us to extend the action
of Γ to an action on L∞(Ω, ν) via
g · f (ω) = f (g−1 · ω)
for all g ∈ Γ, f ∈ L∞(Ω, ν), and ω ∈ Ω. We may therefore consider the
von Neumann algebra L∞(Ω, ν) ⋊ Γ which we shall write as L∞(Ω) ⋊ Γ
for brevity.
2. The Factors
We note that the action of Γ on Ω is free since if gω = ω for some
g ∈ Γ and ω ∈ Ω then we must have either ω = ggg · · · or ω =
g−1g−1g−1 · · · and
ν(cid:0)ggg · · · , g−1g−1g−1 · · ·(cid:1) = 0.
The action of Γ on Ω is also ergodic by the proof of [PS, Propo-
sition 3.9], so that L∞(Ω) ⋊ Γ is a factor. Establishing the type of
the factor is not quite as straightforward. We begin by recalling some
classical definitions.
Definition 2.1. Given a group Γ acting on a measure space Ω, we
define the full group, [Γ], of Γ by
[Γ] = {T ∈ Aut(Ω) : T ω ∈ Γω for almost every ω ∈ Ω} .
The set [Γ]0 of measure preserving maps in [Γ] is then given by
[Γ]0 = {T ∈ [Γ] : T ◦ν = ν}
Definition 2.2. Let G be a countable group of automorphisms of the
measure space (Ω, ν). Following W. Krieger, define the ratio set r(G)
to be the subset of [0, ∞) such that if λ ≥ 0 then λ ∈ r(G) if and only
if for every ǫ > 0 and Borel set E with ν(E) > 0, there exists a g ∈ G
and a Borel set F such that ν(F ) > 0, F ∪ gF ⊆ E and
for all ω ∈ F .
dν◦g
dν
(cid:12)(cid:12)(cid:12)(cid:12)
(ω) − λ(cid:12)(cid:12)(cid:12)(cid:12)
< ǫ
Remark 2.3. The ratio set r(G) depends only on the quasi-equivalence
class of the measure ν, see [HO, §I-3, Lemma 14]. It also depends only
on the full group in the sense that
[H] = [G] ⇒ r(H) = r(G).
FACTORS FROM TREES
3
The following result will be applied in the special case where G = Γ.
However, since the simple transitivity of the action doesn't play a role
in the proof, we can state it in greater generality.
Proposition 2.4. Let G be a countable subgroup of Aut(T ) ≤ Aut(Ω).
Suppose there exist an element g ∈ G such that d(ge, e) = 1 and a
subgroup K of [G]0 whose action on Ω is ergodic. Then
r(G) = (cid:8)nk : k ∈ Z} ∪ {0(cid:9) .
Proof. By Remark 2.3, it is sufficient to prove the statement for some
group H such that [H] = [G]. In particular, since [G] = [hG, Ki] for
any subgroup K of [G]0, we may assume without loss of generality that
K ≤ G.
By [FTN, Chapter 2, part 1)], for each g ∈ G and ω ∈ Ω we have
dν◦g
dν
(ω) ∈ (cid:8)nk : k ∈ Z} ∪ {0(cid:9) .
Since G acts ergodically on Ω, r(G) \ {0} is a group. It is therefore
enough to show that n ∈ r(G). Write x = ge and note that νx = ν◦g−1.
By [FTN, Chapter 2, part1)] we have
(1)
dνx
dν
(ω) = n, for all ω ∈ Ωx
e .
Let E ⊆ Ω be a Borel set with ν(E) > 0. By the ergodicity of K, there
exist k1, k2 ∈ K such that the set
F = {ω ∈ E : k1ω ∈ Ωx
e and k2g−1k1ω ∈ E}
has positive measure.
Finally, let t = k2g−1k1 ∈ G. By construction, F ∪ tF ⊆ E. More-
over, since K is measure-preserving,
dν◦t
dν
(ω) =
dν◦g−1
dν
(k1ω) =
dνx
dν
(k1ω) = n for all ω ∈ F
by (1), since k1 ∈ Ωx
e . This proves n ∈ r(G), as required.
(cid:3)
Corollary 2.5. If, in addition to the hypotheses for Proposition 2.4,
the action of G is free, then L∞(Ω) ⋊ G is a factor of type III1/n.
Proof. Having determined the ratio set, this is immediate from [C1,
Corollaire 3.3.4].
(cid:3)
Thus, if we can find a countable subgroup K ≤ [Γ]0 whose action
on Ω is ergodic we will have shown that L∞(Ω) ⋊ Γ is a factor of type
III1/n. To this end, we prove the following sufficiency condition for
ergodicity.
4
JACQUI RAMAGGE AND GUYAN ROBERTSON
Lemma 2.6. Let K be group which acts on Ω. If K acts transitively
on the collection of sets {Ωx : x ∈ Γ, x = m} for each natural number
m, then K acts ergodically on Ω.
Proof. Suppose that X0 ⊆ Ω is a Borel set which is invariant under
K and such that ν(X0) > 0. We show that this necessarily implies
ν(Ω \ X0) = 0, thus establishing the ergodicity of the action.
Define a new measure µ on Ω by µ(X) = ν(X ∩ X0) for each Borel
set X ⊆ Ω. Now, for each g ∈ K,
µ(gX) = ν(gX ∩ X0) = ν(X ∩ g−1X0)
≤ ν(X ∩ X0) + ν(X ∩ (g−1X0 \ X0))
= ν(X ∩ X0)
= µ(X),
and therefore µ is K-invariant. Since K acts transitively on the basic
open sets Ωx associated to words x of length m this implies that
µ(Ωx) = µ(Ωy)
whenever x = y. Since Ω is the union of Nm disjoint sets Ωx, x ∈ V m,
each of which has equal measure with respect to µ, we deduce that
µ(Ωx) =
c
Nm
for each x ∈ V m, where c = µ(X0) = ν(X0) > 0. Thus µ(Ωx) = cν(Ωx)
for every x ∈ Γ.
Since the sets Ωx, x ∈ Γ generate the Borel σ-algebra, we deduce
that µ(X) = cν(X) for each Borel set X. Therefore
ν(Ω \ X0) = c−1µ(Ω \ X0)
= c−1ν((Ω \ X0) ∩ X0) = 0,
thus proving ergodicity.
(cid:3)
In the last of our technical results, we give a constructive proof of
the existence of a countable ergodic subgroup of [Γ]0.
Lemma 2.7. There is a countable ergodic group K ≤ Aut(Ω) such
that K ≤ [Γ]0.
Proof. Let x, y ∈ V m. We construct a measure preserving automor-
phism kx,y of Ω such that
(1) kx,y is almost everywhere a bijection from Ωx onto Ωy,
(2) kx,y is the identity on Ω \ (Ωx ∪ Ωy).
FACTORS FROM TREES
5
It then follows from Lemma 2.6 that the group
K = hkx,y : {x, y} ⊆ V m, m ∈ Ni
acts ergodically on Ω and the construction will show explicitly that
K ≤ [Γ]0.
Fix x, y ∈ V m and suppose that we have reduced expressions x =
x1 . . . xm, and y = y1 . . . ym.
Define kx,y to be left multiplication by yx−1 on each of the sets Ωxz
where z = 1 and z /∈ {x−1
m }. Then kx,y is a measure preserving
bijection from each such set onto Ωyz. If ym = xm then kx,y is now well
defined everywhere on Ωx.
m , y−1
Suppose now that ym 6= xm. Then kx,y is defined on the set Ωx \Ωxy−1
, which it maps bijectively onto Ωy \ Ωyx−1
m . Now define kx,y to be left
m z where z = 1
multiplication by yx−1
and z /∈ {xm, ym}. Then kx,y is a measure preserving bijection of each
such Ωxy−1
m ymx−1 on each of the sets Ωxy−1
m z onto Ωyx−1
m z.
m
on the set Ωx \ Ωxy−1
Thus we have extended the domain of kx,y so that it is now defined
m ym.
m ymx−1 on the
m xm, which it maps bijectively onto Ωy \ Ωyx−1
Next define kx,y to be left muliplication by yx−1
m ymx−1
sets Ωxy−1
m xmz where z = 1 and z /∈ {x−1
m , y−1
m }.
Continue in this way. At the jth step kx,y is a measure preserving
bijection from Ωx \ Xj onto Ωy \ Yj where ν(Xj) → 0 as j → ∞ so that
eventually kx,y is defined almost everywhere on Ω. Finally, define
kx,y(xy−1
m xmy−1
m xmy−1
m xm . . .) = yx−1
m ymx−1
m ymx−1
m ym . . .
thus defining kx,y everywhere on Ω in such a way that its action is
pointwise approximable by Γ almost everywhere. Hence
K = hkx,y : {x, y} ⊆ V m, m ∈ Ni
is a countable group with an ergodic measure-preserving action on Ω
and K ≤ [Γ]0.
(cid:3)
We are now in a position to prove our main result.
Theorem 2.8. The von Neumann algebra L∞(Ω) ⋊ Γ is the hyperfinite
factor of type III1/n.
Proof. By applying Corollary 2.5 with G = Γ, g ∈ Γ any generator of
Γ, and K as in Lemma 2.7 we conclude that L∞(Ω) ⋊ Γ is a factor of
type III1/n.
To see that the factor is hyperfinite simply note that the action
of Γ is amenable as a result of [A, Theorem 5.1]. We refer to [C2,
6
JACQUI RAMAGGE AND GUYAN ROBERTSON
Theorem 4.4.1] for the uniqueness of the hyperfinite factor of type
III1/n.
(cid:3)
Remark 2.9. Taking different measures on Ω should yield hyperfinite
factors of type IIIλ for any 0 < λ < 1. We have concentrated on the
geometrically interesting case.
Remark 2.10. In [Sp1], Spielberg constructs IIIλ factor states on the
algebra O2. The reduced C ∗-algebra C(Ω) ⋊r Γ is a Cuntz-Krieger
algebra OA by [Sp2]. What we have done is construct a type III1/n
factor state on some of these algebras OA.
Remark 2.11. From [C2, p. 476], we know that if Γ = Q ⋊ Q∗ acts
naturally on Qp, then the crossed product L∞(Qp)⋊Γ is the hyperfinite
factor of type III1/p. This may be proved geometrically as above by
regarding the the boundary of the homogeneous tree of degree p + 1 as
the one point compactification of Qp as in [CKW].
References
[A]
S. Adams. Boundary amenability for word hyperbolic groups and an ap-
plication to smooth dynamics of simple groups. Preprint 1993.
[C1]
[CKW] D. I. Cartwright, V. Kaimanovich, and W. Woess. Random walks on
the affine group of local fields and homogeneous trees. M.S.R.I. Preprint
No. 022-94, 1993.
A. Connes. Une classification des facteurs de type III. Ann. Scient. Ec.
Nrom. Sup., 6 (1973), pp. 133 -- 252.
A. Connes. On the classification of von Neumann algebras and their au-
tomorphisms. Symposia Mathematica XX, pp. 435 -- 478, Academic Press
1978.
[C2]
[HO]
[FTN] A. Fig`a-Talamanca and C. Nebbia, Harmonic Analysis and Representa-
tion Theory for Groups Acting on Homogeneous Trees, Cambridge Uni-
versity Press, London 1991.
T. Hamachi and M. Osikawa, Ergodic Groups Acting of Automorphisms
and Krieger's Theorems, Seminar on Mathematical Sciences No. 3, Keio
University, Japan, 1981.
C. Pensavalle and T. Steger, Tensor products and anisotropic principal
series representations for free groups. Preprint, University of Georgia,
June 1993.
R. J. Spatzier, An example of an amenable action from geometry, Ergod.
Th. & Dynam. Sys. (1987), 7, 289 -- 293.
J. Spielberg, Diagonal states on O2, Pac. J. Maths. 144 (1990), 351 -- 382.
J. Spielberg, Free product groups, Cuntz-Krieger algebras and covariant
maps, Int. J. Maths. (1991), 2, 457 -- 476.
[Sp1]
[Sp2]
[PS]
[S]
FACTORS FROM TREES
7
Mathematics Department, University of Newcastle, Callaghan, NSW
2308, Australia
E-mail address: [email protected]
Mathematics Department, University of Newcastle, Callaghan, NSW
2308, Australia
E-mail address: [email protected]
|
1201.0792 | 1 | 1201 | 2012-01-04T00:36:49 | Amenability for Fell bundles over groupoids | [
"math.OA"
] | We establish conditions under which the universal and reduced norms coincide for a Fell bundle over a groupoid. Specifically, we prove that the full and reduced C*-algebras of any Fell bundle over a measurewise amenable groupoid coincide, and also that for a groupoid G whose orbit space is T_0, the full and reduced algebras of a Fell bundle over G coincide if the full and reduced algebras of the restriction of the bundle to each isotropy group coincide. | math.OA | math |
AMENABILITY FOR FELL BUNDLES OVER GROUPOIDS
AIDAN SIMS AND DANA P. WILLIAMS
Abstract. We establish conditions under which the universal and reduced
norms coincide for a Fell bundle over a groupoid. Specifically, we prove that the
full and reduced C ∗-algebras of any Fell bundle over a measurewise amenable
groupoid coincide, and also that for a groupoid G whose orbit space is T0, the
full and reduced algebras of a Fell bundle over G coincide if the full and reduced
algebras of the restriction of the bundle to each isotropy group coincide.
Contents
Introduction
1. Fell bundles
2. Amenable groupoids
3. The disintegation theorem revisited
4. Fell bundles over amenable groupoids
5. Measurewise amenable groupoids
6. Fibrewise-amenable Fell bundles
Appendix A. Nondegenerate Borel ∗-functors
References
1
2
4
4
5
8
9
11
12
Introduction
If G is an amenable group, then the reduced crossed product and full crossed
product for any action of G on a C∗-algebra coincide. This result was proved for
discrete groups by Zeller-Meyer in [20] and in general by Takai in [18]. Since the C∗-
algebra of a Fell bundle over a groupoid G is a very general sort of crossed product
by G, it is reasonable to expect the universal norm and reduced norm to coincide on
Γc(G; B) when G is suitably amenable. Immediately the situation is complicated
because amenability for groupoids is not as clear cut as it is for groups. There
are three reasonable notions of amenability for a second countable locally compact
Hausdorff groupoid: (topological) amenability, measurewise amenability and, for
lack of a better term, "metric amenability" by which we simply mean that the
reduced norm and universal norm on Cc(G) coincide.
Amenability implies measurewise amenability which in turn implies metric amenabil-
ity. While there are situations where the converses hold, it is unknown if they hold
in general. Our main result here, Theorem 1, is that if G is measurewise amenable
as defined in [1], then the reduced norm and universal norm on Γc(G; B) coincide
Date: 22 December 2011.
2000 Mathematics Subject Classification. 46L55.
Key words and phrases. Groupoid; Fell bundle; amenable; reduced C ∗-algebra.
1
2
AIDAN SIMS AND DANA P. WILLIAMS
for any Fell bundle B over G. This result subsumes the usual result for group
dynamical systems and the result for groupoid dynamical systems; for a discus-
sion of this, see [16, Examples 10 and 11]. The result for groupoid systems is also
asserted in [1, Proposition 6.1.10] where they cite [15, Theorem 3.6]. Since it is
usually hard to determine if a groupoid found in the wild is amenable in any given
one the three flavors mentioned above, we also prove in Theorem 4 that groupoids
which act nicely on their unit spaces in the sense that that G\G(0) is T0 and whose
stability groups are all amenable are themselves measurewise amenable. This result
may be known to experts, but seems worth advertising. We also show that if G is
a groupoid whose orbit space is T0 and if B is a Fell bundle over G such that the
full and reduced C∗-algebras of the restriction of B to each isotropy group in G
coincide, then the full and reduced C∗-algebras of the whole bundle coincide. This
is a formally stronger result than the combination of Theorem 4 and Theorem 1:
there are many examples of Fell bundles over non-amenable groups whose full and
reduced C∗-algebras coincide (see, for example, [3]).
We start with very short sections on Fell bundles and amenable groupoids to
clarify our definitions and point to the relevant literature. In Section 3 we point out
a simple strengthening of the disintegration theorem for Fell bundles (from [10])
which is needed here. For readability, the details are shifted to Appendix A. In
Section 4 we prove our main theorem. In Section 6 we show that groupoids with T0
orbit space and amenable stability groups are measurewise amenable. In Section 6
we prove that bundles over groupoids with T0 orbit space whose restrictions to
isotropy groups are metrically amenable are themselves metrically amenable.
Since we appeal to the disintegration theorem for Fell bundles, we require sep-
arability for our results. In particular, all the groupoids and spaces that appear
will be assumed to be second countable, locally compact and Hausdorff. Except
when it is clearly not the case, for example B(H) and other multiplier algebras, all
the algebras and Banach spaces that appear are separable. We also assume that
our Fell bundles are always saturated. The underlying Banach bundles are only
required to be upper semicontinuous.
1. Fell bundles
We will refer to [10, §1] for details of the definition of a Fell bundle p : B → G over
a groupoid as well as of the construction of the associated C∗-algebra C∗(G, B).
(The examples in [10, §2] would be very helpful supplementary reading.) Roughly
speaking, a Fell bundle p : B → G is an upper-semicontinuous Banach bundle
endowed with a partial multiplication compatible with p such that the fibres A(u)
over units u are C∗-algebras and such that each fibre B(x) is an A(r(x)) -- A(s(x))-
imprimitivity bimodule with respect to the inner products and actions induced by
the multiplication on B. In particular, when x and y are composable, multiplication
in B implements isomorphisms B(x) ⊗A(s(x)) B(y) ∼= B(xy). The space Γc(G; B)
of continuous sections of B then carries a natural convolution and involution. The
C∗-algebra C∗(G, B) is the completion of Γc(G; B) with respect to the universal
norm for representations which are continuous with respect to the inductive-limit
topology on G.
Regarding our notation: as above, we use a roman letter, B(x), for the fibre
over x together with its Banach space structure, but we will use both A(u) and
B(u) for the fibre over a unit u so as to distinguish its dual roles. The Fell bundle
AMENABILITY FOR FELL BUNDLES OVER GROUPOIDS
3
axioms imply that A := Γ0(G(0); B) is a C∗-algebra which is called the C∗-algebra
of B over G(0); in particular it is a C0(G(0))-algebra. So for u ∈ G(0) we write
A(u) for the fibre over u when we are thinking of it as a C∗-algebra, and we write
B(u) when we are thinking of it instead as an A(u) -- A(u)-imprimitivity bimod-
ule. We assume that our Fell bundles are separable, so in addition to G being
second countable, we assume that the Banach space Γ0(G; B) is separable. By
axiom, our Fell bundles are saturated in that B(x)B(y) = B(xy), where B(x)B(y)
denotes span{ bxby : bx ∈ B(x), by ∈ B(y) }. If F is a locally closed subset1 of G(0),
then we abuse notation slightly and write Γc(F ; B) in place of Γc(F ; BF ) (as we
have already done for A = Γ0(G(0); B) above). If we let G(F ) := GF = { x ∈
G : s(x) ∈ F and r(x) ∈ F } be the restriction of G to F , then G(F ) is a locally
compact groupoid with Haar system {λu}u∈F . As above, we write C∗(G(F ), B) in
place of C∗(G(F ), BG(F )).
Recall the definition of the reduced norm on Γc(G; B) from [16].
If π is a
representation of A = Γ0(G(0); B), then using [19, Example F.25] and the discussion
preceding [11, Definition 7.9], we can assume that there is a Borel Hilbert bumdle
G(0) ∗ H , a finite Radon measure µ on G(0) and representations πu of A on H(u),
factoring through Au, such that
π =Z ⊕
G(0)
πu dµ(u).
For u ∈ G(0), we frequently regard the πu as representations of A(u). Even if π is
nondegenerate, we can only assume that µ-almost all of the πu are nondegenerate.
Indeed, we could have πu = 0 for a null set of u. The formula (Ind π)(f )(g ⊗ h) =
(f ∗ g) ⊗ h for f, g ∈ Γc(G; B) and h ∈ L2(G(0) ∗ H , µ) determines a representation
Ind π of Γc(G; B) on the completion of Γc(G; B) ⊙ L2(G(0) ∗ H , µ) with respect to
the inner product
(f ⊗ h g ⊗ k) =(cid:0)π(g∗ ∗ f )h k(cid:1)
=ZG(0)ZG(cid:0)πu(cid:0)g(x−1)∗f (x−1)(cid:1)h(u) k(u)(cid:1) dλu(x) dµ(u)
=ZG(0)ZG(cid:0)πu(cid:0)g(x)∗f (x)(cid:1)h(u) k(u)(cid:1) dλu(x) dµ(u).
(1)
The reduced norm on Γc(G; B) is given by
kf kr := sup{ k(Ind π)(f )k : π is a representation of A }.
Since ker(Ind π) depends only on ker π (by, for example, [12, Corollary 2.73]) this
definition of k · kr agrees with other definitions in the literature -- for example Exel's
in [3] and Moutuou and Tu's in [8]. So C∗
r (G, B) is the quotient of C∗(G, B) by
r (G,B) of Ind π for any faithful representation π of the C∗-algebra
the kernel IC ∗
A = Γ0(G(0); B) of B over G(0).
1Recall that a subset of a locally compact Hausdorff space is locally compact if and only if it
is locally closed and is locally closed if and only if it is the intersection of a closed set and an open
set [19, Lemmas 1.25 and 1.26].
4
AIDAN SIMS AND DANA P. WILLIAMS
2. Amenable groupoids
Let G be a second-countable locally compact groupoid with Haar system λ. Re-
nault [14, p. 92] originally defined G to be topologically amenable, or just amenable,
if there is a net {fi} ⊂ Cc(G) such that
(a) the functions u 7→ fi ∗ f ∗
(b) fi ∗ f ∗
i → 1 uniformly on compact subsets of G.
i (u) are uniformly bounded on C0(G(0)), and
Later, in the extensive treatment by Anatharaman-Delaroche and Renault, an a
priori different definition was given:
[1, Definition 2.2.8]; however, [1, Proposi-
tion 2.2.13(iv)] and its proof show that the two notions of amenability are equiva-
lent. It is not hard to see, using standard criteria such as [19, Proposition A.17],
that a group is amenable as a groupoid if and only if it is amenable as a group.
measure on G (that is, ν(·) = RG(0) λu(·) dµ(u)).
Let µ be a quasi-invariant measure on G(0), and let ν := µ ◦ λ be the induced
In [1, Definition 3.2.8], µ is
called amenable if there exists a suitably invariant mean on L∞(G, ν). The pair
(G, λ) is measurewise amenable if every quasi-invariant measure µ is amenable
[1, Definition 3.2.8]. Since L∞(G, ν) depends only on the equivalence class of ν,
if µ′ is equivalent to µ and µ is amenable, then so is µ′. Since [1] considers only
σ-finite measures, to demonstrate that (G, λ) is measurewise amenable, it suffices
to show that every finite quasi-invariant measure µ is amenable.
It follows from [1, Theorem 2.2.17] and [1, Theoerem 3.2.16] that amenability
and measurewise amenability, respectively, are preserved under groupoid equiva-
lence. Theorem 17 of [17] implies that metric amenability is preserved as well. In
particular, none of the three flavors of amenability of G depend on the choice of
Haar system λ.
In this note, we will use the characterization of amenability of (G, λ, µ) given
in [1, Proposition 3.2.14(v)]. If G is amenable then it is measurewise amenable by
[1, Proposition 3.3.5]. If G is measurewise amenable then it is metrically amenable
by [1, Proposition 6.1.8].
3. The disintegation theorem revisited
Our main tool here is the disintegration theorem from [10]. Fix a nondegenerate
representation L of C∗(G, B). Then [10, Theorem 4.13] implies that there are a
quasi-invariant measure µ on G(0), a Borel Hilbert bundle G(0) ∗ H , and a Borel
∗-functor b 7→(cid:0)r(b), π(b), s(b)(cid:1) (see [10, Definition 4.5]) from B into End(G(0) ∗ H )
such that L is equivalent to the integrated form of the associated strict representa-
tion (µ, G(0) ∗ H , π) of B. For h, k ∈ L2(G(0) ∗ H , µ) and f ∈ Γc(G; B), we then
have
(cid:0)L(f )h k(cid:1) =ZG(cid:16)π(f (x))h(s(x)) (cid:12)(cid:12) k(r(x)(cid:17)∆(x)− 1
2 dν(x).
Regrettably, the authors of [10] neglected to point out that the Borel ∗-functor
associated to L constructed in [10, Theorem 4.13] is nondegenerate in the sense
that for all x ∈ G,
(2)
π(B(x))H(s(x)) = span{ π(b)v : b ∈ B(x) and v ∈ H(s(x)) } = H(r(x)).
We outline why this is true in Appendix A, and at the same time, we tidy up
some details of the proof of the disintegration theorem itself.
AMENABILITY FOR FELL BUNDLES OVER GROUPOIDS
5
4. Fell bundles over amenable groupoids
Our first main theorem says that every Fell bundle over a measurewise amenable
groupoid is metrically amenable.
Theorem 1. Let G be a second-countable locally compact Hausdorff groupoid with
Haar system {λu}u∈G(0) . Suppose that p : B → G is a separable Fell bundle over
G. If G is measurewise amenable, then the reduced norm on Γc(G; B) is equal to
the universal norm, so C∗
r (G, B) = C∗(G, B).
Our proof follows the lines of Renault's proof of the corresponding result for
groupoid C∗-algebras, suitably modified for the bundle context. Before getting
into the proof, we need to do a little set-up. We will continue with the following
notation for the remainder of the section.
Fix a ∈ IC ∗
r (G,B) and let L be a nondegenerate representation of C∗(G, B). As
in Section 3, we may assume that L is the integrated form of a strict representation
(cid:0)µ, G(0) ∗ H , π(cid:1) of B which is nondegenerate in the sense that (2) holds for all x.
Fix a unit vector h in L2(G(0) ∗ H , µ), and let ωh be the associated vector state.
To prove Theorem 1, it suffices to see that ωh(a) = 0.
Let G ∗ Hr be the pullback of G(0) ∗ H over the range map. We may describe it
as follows. Let (hj)∞
j=1 be a special orthorgonal fundamental sequence for G(0) ∗ H
as in [19, Proposition F.6]. For each j, let hj(x) = hj(r(x)) ∈ H(r(x)). Then
j=1.
fundamental sequence (hj)∞
Let ν = µ ◦ λ be the measure on G induced by µ, and recall that ν−1 denotes
G ∗ Hr is isomorphic to the Borel Hilbert bundle built from `x∈G H(r(x)) with
the measure ν−1(f ) = RG f (x−1)dν(x). Since µ is quasi-invariant, ν and ν−1 are
equivalent measures. By passing to an equivalent measure, we may assume that
the Radon-Nikodym derivative ∆ = dν/dν−1 is multiplicative from G to (0, ∞) --
there is a nice proof of this in [9, Theorem 3.15].2
Lemma 2. Define U : Γc(G; B) ⊙ L2(G(0) ∗ H , µ) → L2(G ∗ Hr, ν−1) by U (f ⊗
h)(x) = π(cid:0)f (x)(cid:1)h(cid:0)s(x)(cid:1). Then U is isometric and extends to a unitary, also de-
noted by U , from HInd πµ onto L2(G ∗ Hr, ν−1). Furthermore, U intertwines the
regular representation Ind πµ with the representation Mπ of C∗(G, B) on L2(G ∗
H , ν−1) given on f ∈ Γc(G; B) by
(3)
(cid:0)Mπ(f )ξ η(cid:1) =ZGZG(cid:16)π(f (xy))ξ(y−1)(cid:12)(cid:12) η(x)(cid:17) dλs(x)(y) dν−1(x).
Proof. That π is a Borel ∗-functor, f is a continuous section and
(cid:0)U (f ⊗ h)(x) hj(x)(cid:1) =(cid:0)π(f (x))h(s(x)) hj(r(x)(cid:1)
∞
=
Xk=1(cid:0)h(s(x)) hk(s(x))(cid:1)(cid:0)π(f (x))hk(s(x)) hj(r(x))(cid:1),
imply that x 7→ (cid:0)U (f ⊗ h)(x) hj(x)(cid:1) is Borel. Thus U (g ⊗ h) ∈ B(G ∗ Hr).
The representation πµ comes from a Borel ∗-functor defined on all of B, so the
2The proof in [9] unfortunatly remains unpublished, but it is based on Hahn's [4, Corol-
lary 3.14].
6
AIDAN SIMS AND DANA P. WILLIAMS
formula (1) for the inner product on HInd πµ becomes
(cid:0)f ⊗ h g ⊗ k(cid:1) =ZG(0)ZG(cid:0)π(cid:0)g(x)∗f (x)(cid:1)h(u) k(u)(cid:1) dλu(x) dµ(u)
=ZG(cid:0)π(f (x))h(s(x)) π(g(x))k(s(x))(cid:1) dν−1(x)
=ZG(cid:0)U (f ⊗ h)(x) U (g ⊗ k)(x)(cid:1) dν−1(x).
Hence U (f ⊗ h) ∈ L2(G ∗ H , ν−1) and U is an isometry. Since π is nondegenerate,
an argument like that of [19, Lemma F.17], shows that the range of U is dense.
Hence, U is a unitary as claimed.
For the last assertion, recall that Ind πµ(f ) acts by convolution. Thus
(cid:0)Mπ(f )ξ η(cid:1) =ZG(cid:0)Mπ(f )ξ(x) η(x)(cid:1) dν−1(x)
=ZGZG(cid:0)π(f (y))ξ(y−1x) η(x)(cid:1) dλr(x)(y) dν−1(x)
=ZGZG(cid:0)π(f (xy))ξ(y−1) η(x)(cid:1) dλs(x)(y) dν−1(x).
(cid:3)
To prove Theorem 1, we invoke measurewise amenability in the form of [1, Propo-
sition 3.2.14(v)]. So we fix a sequence (fn)∞
n=1 of Borel functions on G such that
(a) u 7→RG fn(x)2 dλu(x) is bounded on G(0),
n ∗ fn(u) ≤ 1 for all u ∈ G(0) and
n ∗ fn → 1 in the weak-∗ topology on L∞(G, ν).
(b) f ∗
(c) f ∗
To keep notation compact, we denote f ∗
n ∗ fn by en, so that for y ∈ G,
en(y) =ZG
fn(x−1)fn(x−1y) dλr(y)(x),
Proof of Theorem 1. Recall the notation fixed at the beginning of the section: in
particular, L is the integrated form of a nondegenerate strict representation of
C∗(G, B) on a Hilbert bundle G(0) ∗ G, h is a unit vector in L2(G(0) ∗ H , µ) and
ωh is the associated vector state. We claim that ωh(g) ≤ k(Ind πµ)(g)k for all
g ∈ Γc(G; B).
Fix g ∈ Γc(G; B). Then
Define a sequence (αn)∞
n=1 of complex numbers by
ωh(g) =(cid:0)L(g)h h(cid:1) =ZG(cid:0)π(g(y))h(s(y)) h(r(y))(cid:1)∆(y)− 1
αn :=ZG
=ZG(0)ZGZG
en(y)(cid:0)π(g(y))h(s(y)) h(r(y))(cid:1)∆(y)− 1
fn(x−1)fn(x−1y)(cid:0)π(g(y))h(s(y)) h(r(y))(cid:1)∆(y)− 1
2 dν(y)
2
2 dν(y).
(4)
dλu(x) dλu(y) dµ(u).
(It is tempting to write ωh(eng) for αn, but the en are assumed only to be Borel,
so the pointwise products eng may not belong to Γc(G; B).) By assumption on the
en, the αn converge to ωh(g). So it suffices to show that
(5)
αn ≤ k(Ind πµ)(g)k
for all n ∈ N.
AMENABILITY FOR FELL BUNDLES OVER GROUPOIDS
7
Fix n ∈ N. Define hn : G → H by
hn(x) = ∆(x)
Then for each j, the function
x 7→(cid:0)hn(x) hj(x)(cid:1) = ∆(x)
1
2 fn(x−1)h(cid:0)r(x)(cid:1).
2 fn(x−1)(cid:0)h(r(x)) hj(r(x))(cid:1)
1
is Borel, so hn ∈ L2(G ∗ H , ν−1). Starting from (4), we apply Fubini's theorem,
substitute xy for x, and then use first that ν = ∆ν−1 and then that ∆ is multi-
plicative to calculate:
2 dλs(x)(y) dν(x)
αn =ZGZG
fn(x−1)fn(y)(cid:0)π(g(xy))h(s(y)) h(r(x))(cid:1)∆(xy)− 1
=ZGZG
fn(x−1)fn(y)(cid:0)π(g(xy))h(s(y)) h(r(x))(cid:1)
=ZGZG(cid:0)π(g(xy))hn(y−1) hn(x)(cid:1) dλs(x)(y) dν−1(x)
=(cid:0)Mπ(g)hn hn(cid:1).
khnk2 =ZG
khn(x)k2 dν−1(x) =ZG
2 ∆(x) dλs(x)(y) dν−1(x)
∆(xy)− 1
We have
and since ν = ∆ν−1 and en(u) ≤ 1 for all u, it follows that
fn(x−1)2kh(r(x))k2∆(x) dν−1(x),
khnk2 =ZG(0)
en(u)kh(u)k2 dµ(u) ≤ khk2 = 1.
Hence the Cauchy-Schwarz inequality gives (5). Thus ωh(g) ≤ k(Ind πµ)(g)k
for all g ∈ Γc(G; B) as claimed. Since Γc(G; B) is dense in C∗(G, B), it follows
that
ωh(a) ≤ k(Ind πµ)(a)k
for all a ∈ C∗(G, B).
In particular, if a ∈ IC ∗
required.
r (G,B), then (Ind πµ)(a) = 0, and hence ωh(a) = 0 as
(cid:3)
Example 3. Recall from [2] that given a row-finite k-graph Λ with no sources, a Λ-
system of C∗-correspondences consists of an assignment v 7→ Av of C∗-algebras to
vertices and an assignment λ 7→ Xλ of an Ar(λ) -- As(λ) correspondence to each path
λ, together with isomorphisms χµ,ν : Xµ ⊗As(µ) Xν → Xµν for each composable
pair µ, ν ∈ Λ, all subject to an appropriate associativity condition on the χµ,ν (see
[2, Definitions 3.1.1 and 3.1.2] for details). Suppose that X is such a system, and
suppose that each Xλ is nondegenerate as a left Ar(λ)-module, and full as a right
Hilbert As(λ)-module, and that the left action of Ar(λ) is by compact operators.
By [2, Theorem 4.3.1], the construction of Sections 4.1 and 4.2 of the same paper
associates to X a saturated Fell bundle EX over the k-graph groupoid GΛ of [7].
Moreover, [2, Theorem 4.3.6] says that the C∗-algebra C∗(A, X, χ) of the Λ-system
is isomorphic to the reduced C∗-algebra C∗
r (GΛ, EX ) of the Fell bundle.
Theorem 5.5 of [7] says that GΛ is amenable, and hence also measurewise
r (GΛ, EX ) = C∗(GΛ, EX ); in par-
amenable. Hence our Theorem 1 implies that C∗
ticular C∗(A, X, χ) ∼= C∗(GΛ, EX ).
8
AIDAN SIMS AND DANA P. WILLIAMS
Since 1-graphs are precisely the path-categories E∗ of countable directed graphs
E, and since an E∗-system of correspondences can be constructed from any assign-
ment of C∗-algebras Av to vertices v, and Ar(e) -- As(e) C∗-correspondences Xe to
edges e (see [2, Remark 3.1.5]), Example 3 provides a substantial library of examples
of our result
5. Measurewise amenable groupoids
Our initial motivation for proving Theorem 1 was to show that if G has T0
orbit space and amenable stability groups then the full and reduced C∗-algebras of
any Fell bundle over G coincide: roughly, since C∗(G, B) is a C0(G\G(0))-algebra,
representations will factor through restrictions to orbit groupoids G([u]), each of
which is amenable because is is equivalent to the amenable stability group G(u) :=
{x ∈ G : r(x) = u = s(x)} (see section 6 for details). However, the following
argument shows that the result follows directly from Theorem 1. We thank Jean
Renault for pointing us in the direction of [1, Proposition 5.3.4].
Theorem 4. Suppose that G is a second countable locally compact Hausdorff
groupoid with Haar system {λu}u∈G(0) . Suppose that the orbit space G\G(0) is T0
and that each stability group G(u) is amenable. Then G is measurewise amenable.
Our proof requires some straightforward observations as well as some nontrivial
results from [1].
Lemma 5. Suppose that µ is a quasi-invariant finite measure on G(0) and that
F ⊂ G(0) is a locally compact G-invariant subset such that µ(G(0) \ F ) = 0. Then
(G, λ, µ) is amenable if and only if (G(F ), λF , µF ) is amenable.
Proof. Recall that (G, λ, µ) is amenable if there is an invariant mean on L∞(G, ν)
where ν = µ◦λ. Since µF ◦λF = νG(F ), we have L∞(G, ν) ∼= L∞(G(F ), µF ◦λF ).
In particular, an invariant mean on L∞(G) gives an invariant mean on L∞(G(F ))
and vice versa.
(cid:3)
Lemma 6. Suppose that G\G(0) is T0. Then, as a Borel space, G\G(0) is countably
separated and each orbit [u] is locally closed in G and hence locally compact.
Proof. Since subsets of a locally compact Hausdorff space are locally compact if and
only if they are locally closed (see [19, Lemma 1.26]), the lemma is an immediate
consequence of the Mackey-Glimm-Ramsay dichotomy [13, Theorem 2.1].
(cid:3)
Proof of Theorem 4. Suppose that µ is a finite quasi-invariant measure on G(0). It
suffices to show that (G, λ, µ) is amenable. Let p : G → G\G(0) be the orbit map,
and let µ be the forward image µ(f ) = µ(f ◦ p) of µ under p. By Lemma 6, G\G(0)
is countably separated as a Borel space. Hence we can disintegrate µ -- as, for
example, in [19, Theorem I.5] -- so that for each orbit [u] there is a probability
measure ρ[u] on G(0) supported on [u] such that
µ =ZG\G(0)
ρ[u] dµ([u]).
It follows from [1, Proposition 5.3.4] that ρ[u] is quasi-invariant for almost all [u] and
that (G, λ, µ) is amenable if each (G, λ, ρ[u]) is. Since ρ[u](G(0) \ [u]) = 0, Lemma 5
implies that it is enough to see that each (G([u]), λ[u], µ[u]) is amenable. Since [u] is
locally compact, G([u]) is a locally compact transitive groupoid equivalent to G(u),
AMENABILITY FOR FELL BUNDLES OVER GROUPOIDS
9
which is assumed to be amenable. Hence [1, Theorem 2.2.13] implies that G([u]) is
amenable, and therefore also measurewise amenable by [1, Proposition 3.3.5]. (cid:3)
6. Fibrewise-amenable Fell bundles
In the preceding section, we showed that if G\G(0) is T0 and each stability group
is amenable, then G is measurewise amenable. In particular, if p : B → G is a
bundle over such a groupoid, then its full and reduced algebras coincide. In this
section, we show that it suffices that G\G(0) is T0 and that for each u ∈ G(0),
the full and reduced algebras of the restriction of B to the isotropy group G(u)
coincide. To see that this is a strictly stronger theorem, and also that the hypothesis
is genuinely checkable, we refer the reader to the results, for example, of [3].
Theorem 7. Let G be a second-countable locally compact Hausdorff groupoid with
Haar system {λu}u∈G(0) , and let p : B → G be a separable Fell bundle over G.
Suppose that the orbit space G\G(0) is T0 and that for each unit u, the full and
reduced cross-sectional algebras C∗(G(u), B) and C∗
r (G(u), B) coincide. Then the
full and reduced norms on Γc(G; B) are equal and hence C∗
r (G, B) = C∗(G, B).
To prove the theorem, we first use the equivalence theorem of [16] to see that
the full and reduced C∗-algebras of a Fell bundle over transitive groupoid coincide
whenever the full and reduced algebras of its restriction to any isotropy group
coincide.
Lemma 8. Let G be a second-countable locally compact Hausdorff groupoid with
Haar system {λu}u∈G(0) , and let p : B → G be a separable Fell bundle over G.
Suppose that G is transitive. Then the following are equivalent.
(a) For some unit u, the full and reduced cross-section algebras C∗(G(u), B)
and C∗
r (G(u), B) coincide.
(b) For every unit u, the full and reduced cross-section algebras C∗(G(u), B)
(c) The full and reduced norms on Γc(G; B) are equal and hence C∗
r (G, B) =
and C∗
r (G(u), B) coincide.
C∗(G, B).
Proof. Fix u ∈ G(0). Then Gu := s−1(u) is a (G, G(u))-equivalence, and as in [5,
Theorem 1], E := p−1(Gu) implements an equivalence between B and p−1(G(u)).
Consequently, [16, Theorem 14] implies that the natural surjection of C∗(G, B)
onto C∗
r (G, B) is an isomorphism if and only if the kernel Ir of the natural map of
C∗(G(u), B) onto C∗
r (G(u), B) is trivial. Since u ∈ G(0) was arbitrary, the result
follows.
(cid:3)
To finish off our proof of Theorem 7, we need the following special case of [6,
Theorem 3.7]. As above, let p : B → G be a separable Fell bundle over G with
associated C∗-algebra A = Γ0(G(0); B). Recall from [6, Proposition 2.2] that G
acts on Prim A which we identify with { (u, P ) : u ∈ G(0) and P ∈ Prim A(u) }.
Let U be an open G-invariant subset of G(0) with complement F . Then { (u, P ) ∈
Prim A : u ∈ F } is a closed invariant subset of Prim(A), and corresponds to the
G-invariant ideal { a ∈ A : a(u) = 0 for all u ∈ F } of A. By [6, Proposition 3.3],
the corresponding bundle BI is the one with fibres
BI (x) =(B(x)
{0}
if x ∈ G(U )
if u ∈ G(F ),
10
AIDAN SIMS AND DANA P. WILLIAMS
so we can identify it with the bundle BG(U) over G(U ). Moreover, BI is the
complementary bundle
BI (x) =({0}
B(x)
if u ∈ G(U )
if u ∈ G(F ),
which we may identify with the bundle BG(F ) over G(F ). Thus, as a special case
of Theorem 3.7 of [6], we obtain the following result.
Lemma 9. Let G be a second-countable locally compact Hausdorff groupoid with
Haar system {λu}u∈G(0) , and let p : B → G be a separable Fell bundle over G.
Suppose that U is a G-invariant open subset of G(0) with complement F . There is
a short exact sequence of C∗-algebras
0
/ C∗(G(U ), B)
ι
/ C∗(G, B)
q
/ C∗(G(F ), B)
/ 0,
where ι is induced by inclusion and q by restriction on sections.
As an application of Lemma 9, recall3 that there is a nondegenerate map M :
C0(G(0)) → M (C∗(G, B)) given on sections by
M (φ)f (x) = φ(cid:0)r(x)(cid:1)f (x).
Suppose that the orbit space G\G(0) is Hausdorff. Then we may identify C0(G\G(0))
with the subalgebra of Cb(G(0)) consisting of functions which are constant on orbits
and vanish at infinity on the orbit space. We extend M to Cb(G(0)) and restrict
to C0(G\G(0)) to obtain a nondegenerate map of C0(G\G(0)) into the center of
M (C∗(G, B)), making C∗(G, B) into a C0(G\G(0))-algebra. As usual, if u ∈ G(0),
we let [u] be the corresponding orbit in G\G(0).
Corollary 10. Let G be a second-countable locally compact Hausdorff groupoid
with Haar system {λu}u∈G(0), and let p : B → G be a separable Fell bundle
over G. If G\G(0) is Hausdorff, then C∗(G, B) is a C0(G\G(0))-algebra with fibres
C∗(G, B)([u]) ∼= C∗(G([u]), B).
Proof. Recall that C∗(G, B)([u]) is the quotient of C∗(G, B) by the ideal J[u] =
span{ φ · a : φ ∈ C0(G\G(0)), φ([u]) = 0 and a ∈ C∗(G, B) }. Using Lemma 9, we
can identify J[u] with C∗(G(U ), B), where U = G(0) \ [u], and C∗(G, B)/J[u] with
C∗(G([u]), B) as claimed.
(cid:3)
Proof of Theorem 7. Fix in irreducible representation π of C∗(G; B) and an ele-
ment f ∈ Γc(G; B). It suffices to show that kπ(f )k ≤ kf kC ∗
r (G;B).
By [13, Theorem 2.1], the orbit space G\G(0) is locally Hausdorff and every orbit
[u] is locally closed in G(0). Since G\G(0) is second countable, [19, Lemma 6.3] im-
plies that there is a countable ordinal γ and a nested open cover { Un : 0 ≤ n ≤ γ }
of G\G(0) such that U0 = ∅, Uγ = G\G(0) and Un+1 \ Un is Hausdorff (and dense)
in (G\G(0)) \ Un. For n ≤ γ, let Vn := { u ∈ G(0) : [u] ∈ Un }. Then each Vn is
an open invariant subset of G(0). Using Lemma 9, we can identify C∗(G(Vn), B)
with an ideal in C∗(G, B).
In fact, { C∗(G(Vn), B) }n≤γ is a composition series
of ideals in C∗(G, B). By [19, Lemma 8.13], there exists 0 < n ≤ γ such that π
lives on the subquotient C∗(G(Vn), B)/C∗(G(Vn−1), B); that is, π is the canon-
ical lift ¯ρ of an irreducible representation ρ of the ideal C∗(G(Vn), B) such that
3A proof can be constructed along the lines of [10, Proposition 4.2].
/
/
/
/
AMENABILITY FOR FELL BUNDLES OVER GROUPOIDS
11
ker ρ ⊃ C∗(G(Vn−1), B). Lemma 9 implies that C∗(G(Vn), B)/C∗(G(Vn−1), B) ∼=
C∗(G(Vn \ Vn−1), B). By construction, G(Vn \ Vn−1) has Hausdorff orbit space
Un \ Un−1. Hence C∗(G(Vn \ Vn−1), B) is a C0(Un \ Un−1)-algebra and ρ fac-
tors through a fibre C∗(G(Vn \ Vn−1), B)([u]) ∼= C∗(G([u]), B) for some u ∈ Vn
by [19, Proposition C.5]. Since, by assumption, C∗(G[u]), B) = C∗
r (G([u]), B),
we have ker(Ind π[u]) ⊂ ker ρ where π[u] factors through a faithful representation
of the quotient AVn ([u]) of the C∗-algebra A(Vn) of C∗(G(Vn), B) corresponding
to the closed set [u] ⊂ Vn. (Note that A(Vn) is the ideal of A corresponding to
Vn ⊂ G(0).) The kernel of π = ¯ρ is then contained in the kernel of the canoni-
cal lift of Ind π[u] to C∗(G, B). It is not hard to check that the canonical lift of
Ind π[u] to C∗(G, B) is Ind ¯π[u] where ¯π[u] is the canonical lift of π[u] to A. Hence
kπ(f )k = k ¯ρ(f )k ≤ k Ind ¯π[u](f )k ≤ kf kC ∗
(cid:3)
r (G,B).
Appendix A. Nondegenerate Borel ∗-functors
Let p : B → G be a Fell bundle over a second-countable locally compact Haus-
dorff groupoid, and let G(0) ∗ H be a Borel Hilbert bundle. A Borel ∗-functor π
from B to End(G(0) ∗ H ) is a map
π : b 7→(cid:0)r(b), π(b), s(b)(cid:1)
such that π(b) ∈ B(cid:0)H(s(b)), H(r(b))(cid:1) for all b and such that π respects adjoints
and the partial linear and multiplicative structure of B (see [10, Definition 4.5]).
Following [10, §4], a strict representation of B is a triple (µ, G(0) ∗ H , π) consisting
of a quasi-invariant measure µ on G(0), a Borel Hilbert bundle G(0) ∗ H and a Borel
∗-functor π. It is common practice to use π and π interchangeably, and we will drop
the caret henceforth. A strict representation determines a bounded representation
via integration (see [10, Proposition 4.10]); indeed, a Borel ∗-functor defined on
p−1(GF ) for any µ-conull set F ⊂ G(0) is sufficient. Nevertheless, it is convenient
to have π defined everywhere.
The purpose of this section is to point out that the disintegration theorem [10,
Theorem 4.13] for Fell bundles can be strengthened to assert that π can be taken
to be nondegenerate as defined in §3. At the same time, we correct an error in the
construction of π in [10].
In the proof of [10, Theorem 4.13], starting from a pre-representation L of B
on a dense subspace H0 of a Hilbert space H, the authors showed that for any
orthonormal basis {ζi : i ∈ N} for span{L(f )ξ : f ∈ Γc(G; B), ξ ∈ H0}, setting
H′
00 := span{ζi : i ∈ N}, there is a saturated Borel µ-conull set F ⊂ G(0) and a
Borel Hilbert bundle F ∗ H whose fibres H(u) are Hilbert-space completions of
Γc(Gu; B) ⊙ H′
00 (see [10, Lemma 5.18] and [10, Lemma 5.20]). For f ∈ Γc(G; B)
and h ∈ H′
00, the class of f ⊗ h in H(u) is denoted by f ⊗u h. The space H(u)
may be trivial for some u. For each z ∈ GF , b ∈ B(z) and f ∈ Γc(G; B), let π(b)f
denote a section satisfying
π(b)f (x) = ∆(z)
1
2 bf (z−1x)
for x ∈ Gr(b).
The Borel ∗-functor in the disintegration of L constructed in [10] is defined by
π(b)(f ⊗s(b) ζi) = π(b)f ⊗r(b) ζi.
Since F is saturated, G is the disjoint union of GF and GG(0)\F . Since the latter
is ν-null, we can extend π to all of G by defining it as we please on p−1(GG(0)\F ),
and this will not affect the integrated representation. To ensure that π is still a
12
AIDAN SIMS AND DANA P. WILLIAMS
genuine Borel ∗-functor, one sets H(u) := {0} for each u /∈ F and π(b) := 0 for
b /∈ p−1(GF ). (In [10], the authors mistakenly let (G(0) \ F ) ∗ H be a constant
field and let π(b) be the identity operator, but such a π is not a ∗-functor since as it
doesn't preserve the partial linear structure.) We claim that π is nondegenerate in
the sense that (2) holds for all z ∈ G. It holds trivially for z 6∈ GF , so fix z ∈ GF ,
and let u := r(z). We start with two observations.
(A) If fi → f in the inductive limit topology on Γc(Gu; B) then fi ⊗u ζk →
f ⊗u ζk in H(u). To see this, observe that equation (5.19) of [10] is bounded
by Kkf k∞kgk∞ where K is constant depending only on supp f and supp g.
(B) If { ei } is an approximate identity in A(u), and, for each i, eig represents any
section in Γc(G; B) such that (eig)(x) = eig(x) for x ∈ Gu, then eig → g in
the inductive limit topology on Γc(Gu; B). This follows from a compactness
argument using that A(u) acts nondegenerately on B(x).
By (B), to establish (2) for z, it suffices to see that each eig ⊗r(z) ζk belongs to
π(B(z)H(s(z)). Fix b1, . . . bn ∈ B(z) such that
bjb∗
j ∼ ei.
Xj
Then by (A), we have
and this suffices.
π(bj)(π(b∗
j )g ⊗ ζk) ∼ eig ⊗ ζk,
Xj
Remark 11. Just as ∗-functors are automatically bounded (see [10, Remark 4.6]),
there is a sense in which the Borel ∗-functor appearing in any strict representation
(µ, G(0) ∗ H , π) is essentially nondegenerate. We claim that
(6)
for all x ∈ G.
π(cid:0)B(x)(cid:1)H(cid:0)(s(x)(cid:1) = π(cid:0)A(r(x)(cid:1)H(cid:0)(r(x)(cid:1)
The right-hand side of (6) is the essential space of the representation πr(x) of
A(cid:0)r(x)(cid:1) determined by π, so (2) holds whenever πr(x) is nondegenerate. So if
the representation πµ of A = Γ0(G(0); B) determined by π is nondegenerate, then
πu is nondegenerate for µ-almost all u, so (2) holds on a ν-conull subset of G (where,
as usual, ν = µ ◦ λ).
To establish (6), we use that B is saturated: one the one hand,
while on the other hand,
π(cid:0)B(x)(cid:1)H(cid:0)s(x)(cid:1) = π(cid:0)A(r(x)(cid:1)(cid:1)π(cid:0)B(x)(cid:1)H(cid:0)s(x)(cid:1) ⊂ π(cid:0)A(r(x)(cid:1)(cid:1)H(cid:0)r(x)(cid:1),
π(cid:0)B(x)(cid:1)H(cid:0)s(x)(cid:1) ⊃ π(cid:0)B(x)(cid:1)π(cid:0)B(x∗)(cid:1)H(cid:0)r(x)(cid:1) = π(cid:0)A(cid:0)r(x)(cid:1)(cid:1)H(cid:0)r(x)(cid:1).
References
[1] Claire Anantharaman-Delaroche and Jean Renault, Amenable groupoids, Monographies de
L'Enseignement Math´ematique [Monographs of L'Enseignement Math´ematique], vol. 36,
L'Enseignement Math´ematique, Geneva, 2000. With a foreword by Georges Skandalis and
Appendix B by E. Germain.
[2] Valentin Deaconu, Alex Kumjian, David Pask, and Aidan Sims, Graphs of C ∗-
correspondences and Fell bundles, Indiana U. Math. J. 59 (2011), 1687 -- 1735.
[3] Ruy Exel, Amenability for Fell bundles, J. reine angew. Math. 492 (1997), 41 -- 73.
[4] Peter Hahn, Haar measure for measure groupoids, Trans. Amer. Math. Soc. 242 (1978), 1 -- 33.
AMENABILITY FOR FELL BUNDLES OVER GROUPOIDS
13
[5] Marius Ionescu and Dana P. Williams, A classic Morita equivalence result for Fell bundle
C ∗-algebras, Math. Scand. 108 (2011), 251 -- 263.
[6]
, Remarks on the ideal structure of Fell bundle C ∗-algebras, Houston J. Math. (2012),
in press. (arXiv:math.OA.0912.1124).
[7] Alex Kumjian and David Pask, Higher rank graph C ∗-algebras, New York J. Math. 6 (2000),
1 -- 20.
[8] El-Kaıoum M. Moutuou and Jean-Louis Tu, Equivalence of fell systems and their reduced
C ∗-algebras, preprint, 2011. (arXiv:math.OA.1101.1235v1).
[9] Paul S. Muhly, Coordinates in operator algebra, CMBS Conference Lecture Notes (Texas
Christian University 1990), 1999. In continuous preparation.
[10] Paul S. Muhly and Dana P. Williams, Equivalence and disintegration theorems for Fell bun-
dles and their C ∗-algebras, Dissertationes Math. (Rozprawy Mat.) 456 (2008), 1 -- 57.
[11]
, Renault's equivalence theorem for groupoid crossed products, NYJM Monographs,
vol. 3, State University of New York University at Albany, Albany, NY, 2008. Available at
http://nyjm.albany.edu:8000/m/2008/3.htm.
[12] Iain Raeburn and Dana P. Williams, Morita equivalence and continuous-trace C ∗-algebras,
Mathematical Surveys and Monographs, vol. 60, American Mathematical Society, Providence,
RI, 1998.
[13] Arlan Ramsay, The Mackey-Glimm dichotomy for foliations and other Polish groupoids, J.
Funct. Anal. 94 (1990), 358 -- 374.
[14] Jean Renault, A groupoid approach to C ∗-algebras, Lecture Notes in Mathematics, vol. 793,
Springer-Verlag, New York, 1980.
[15]
, The ideal structure of groupoid crossed product C ∗-algebras, J. Operator Theory 25
(1991), 3 -- 36.
[16] Aidan Sims and Dana P. Williams, An equivalence theorem for reduced Fell bundle C ∗-
algebras, preprint, 2011. (arXiv:math.OA.1111.5753v1).
[17]
, Renault's equivalence theorem for reduced groupoid C ∗-algebras, J. Operator Theory
(2012), in press. (arXiv:math.OA.1002.3093).
[18] Hiroshi Takai, On a duality for crossed products of C ∗-algebras, J. Funct. Anal. 19 (1975),
25 -- 39.
[19] Dana P. Williams, Crossed products of C ∗-algebras, Mathematical Surveys and Monographs,
vol. 134, American Mathematical Society, Providence, RI, 2007.
[20] Georges Zeller-Meier, Produits crois´es d'une C ∗-alg`ebre par un groupe d'automorphismes, J.
Math. Pures Appl. (9) 47 (1968), 101 -- 239.
School of Mathematics and Applied Statistics, University of Wollongong, NSW
2522, Australia
E-mail address: [email protected]
Department of Mathematics, Dartmouth College, Hanover, NH 03755-3551
E-mail address: [email protected]
|
1311.1193 | 1 | 1311 | 2013-11-05T20:53:26 | Noncommutative solenoids and their projective modules | [
"math.OA"
] | Let p be prime. A noncommutative p-solenoid is the C*-algebra of Z[1/p] x Z[1/p] twisted by a multiplier of that group, where Z[1/p] is the additive subgroup of the field Q of rational numbers whose denominators are powers of p. In this paper, we survey our classification of these C*-algebras up to *-isomorphism in terms of the multipliers on Z[1/p], using techniques from noncommutative topology. Our work relies in part on writing these C*-algebras as direct limits of rotation algebras, i.e. twisted group C*-algebras of the group Z^2 thereby providing a mean for computing the K-theory of the noncommutative solenoids, as well as the range of the trace on the K_0 groups. We also establish a necessary and sufficient condition for the simplicity of the noncommutative solenoids. Then, using the computation of the trace on K_0, we discuss two different ways of constructing projective modules over the noncommutative solenoids. | math.OA | math | NONCOMMUTATIVE SOLENOIDS AND THEIR PROJECTIVE
MODULES
FR´ED´ERIC LATR´EMOLI`ERE AND JUDITH A. PACKER
pi × Zh 1
pi twisted by a multiplier of that group, where Zh 1
Abstract. Let p be prime. A noncommutative p-solenoid is the C ∗-algebra of
Z h 1
pi is the additive
subgroup of the field Q of rational numbers whose denominators are powers
of p. In this paper, we survey our classification of these C ∗-algebras up to
pi, using techniques
*-isomorphism in terms of the multipliers on Zh 1
from noncommutative topology. Our work relies in part on writing these C ∗-
algebras as direct limits of rotation algebras, i.e. twisted group C*-algebras
of the group Z2, thereby providing a mean for computing the K-theory of the
noncommutative solenoids, as well as the range of the trace on the K0 groups.
We also establish a necessary and sufficient condition for the simplicity of
the noncommutative solenoids. Then, using the computation of the trace on
K0, we discuss two different ways of constructing projective modules over the
noncommutative solenoids.
pi × Zh 1
3
1
0
2
v
o
N
5
]
.
A
O
h
t
a
m
[
1
v
3
9
1
1
.
1
1
3
1
:
v
i
X
r
a
1. Introduction
Twisted group algebras and transformation group C∗-algebras have been studied
since the early 1960's [8] and provide a rich source of examples and problems in C*-
algebra theory. Much progress has been made in studying such C∗-algebras when
the groups involved are finitely generated (or compactly generated, in the case of Lie
groups). Even when G = Zn, these C∗-algebras give a rich class of examples which
have driven much development in C*-algebra theory, including the foundation of
noncommutative geometry by Connes [2], the extensive study of the geometry of
quantum tori by Rieffel [14, 16, 17, 18], the expansion of the classification problem
from AF to AT algebras by G. Elliott and D. Evans [4], and many more (L. Baggett
and A. Kleppner [1], and S. Echterhoff and J. Rosenberg [3]).
In this paper, we present our work on twisted group C∗-algebras of the Cartesian
square of the discrete group Zh 1
pi of p-adic rationals, i.e. the additive subgroup
of Q whose elements have denominators given by powers of a fixed p ∈ N, p ≥ 1.
The Pontryagin duals of these groups are the p-solenoid, thereby motivating our
terminology in calling these C∗-algebras noncommutative solenoids. We review our
computation of the K-groups of these C∗-algebras, derived in their full technicality
in [10], and which in and of itself involves an intriguing problem in the theory of
Abelian group extensions. We were also able to compute the range of the trace
Date: March 31, 2013.
1991 Mathematics Subject Classification. Primary 46L40, 46L80; Secondary 46L08, 19K14.
Key words and phrases. C*-algebras; solenoids; projective modules; p-adic analysis.
1
2
FR ´ED ´ERIC LATR ´EMOLI `ERE AND JUDITH A. PACKER
on the K0-groups, and use this knowledge to classify these C∗-algebras up to ∗-
isomorphism, in [10], and these facts are summarized in a brief survey of [10] in the
first two sections of this paper.
This paper is concerns with the open problem of classifying noncommutative
solenoids up to Morita equivalence. We demonstrate a method of constructing an
equivalence bimodule between two noncommutative solenoids using methods due
to M. Rieffel [16], and will note how this method has relationships to the theory of
wavelet frames. These new matters occupy the last three sections of this paper.
Acknowledgments. The authors gratefully acknowledge helpful conversations with
Jerry Kaminker and Jack Spielberg.
2. Noncommutative Solenoids
This section and the next provide a survey of the main results proven in [10]
concerning the computation of the K-theory of noncommutative solenoids and its
application to their classification up to *-isomorphism. An interesting connection
between the K-theory of noncommutative solenoids and the p-adic integers is un-
earthed, and in particular, we prove that the range of the K0 functor on the class
of all noncommutative solenoids is fully described by all Abelian extensions of the
group of p-adic rationals by Z. These interesting matters are the subject of the
next section, whereas we start in this section with the basic objects of our study.
We shall fix, for this section and the next, an arbitrary p ∈ N with p > 1. Our
story starts with the following groups:
Definition 2.1. Let p ∈ N, p > 1. The group Zh 1
inductive limit of the sequence of groups:
pi of p-adic rationals is the
Z
z7→pz
−−−−→ Z
z7→pz
−−−−→ Z
z7→pz
−−−−→ Z
z7→pz
−−−−→ · · ·
which is explicitly given as the group:
(2.1)
endowed with the discrete topology.
p(cid:21) =(cid:26) z
pk ∈ Q : z ∈ Z, k ∈ N(cid:27)
Z(cid:20) 1
From the description of Zh 1
pi as an injective limit, we obtain the following
Proposition 2.2. Let p ∈ N, p > 1. The Pontryagin dual of the group Zh 1
pi is the
result by functoriality of the Pontryagin duality. We denote by T the unit circle
{z ∈ C : z = 1} in the field C of complex numbers.
p-solenoid group, given by:
endowed with the induced topology from the injection Sp ֒→ TN. The dual pairing
between QN and SN is given by:
n+1 = zn(cid:9) ,
Sp =(cid:8)(zn)n∈N ∈ TN : ∀n ∈ N zp
pk , (zn)n∈N(cid:29) = zq
(cid:28) q
pi and (zn)n∈N ∈ Sp.
k,
where q
pk ∈ Zh 1
We study in [10] the following C*-algebras.
NONCOMMUTATIVE SOLENOIDS AND THEIR PROJECTIVE MODULES
3
Definition 2.3. A noncommutative solenoid is a C*-algebra of the form
C∗(cid:18)Z(cid:20) 1
p(cid:21) × Z(cid:20) 1
p(cid:21), σ(cid:19) ,
where p is a natural number greater or equal to 2 and σ is a multiplier of the group
The first matter to attend in the study of these C*-algebras is to describe all
are T-valued unless otherwise specified, with T the unit circle in C. Note that the
pi × Zh 1
pi.
Zh 1
the multipliers of the group Zh 1
group Zh 1
Using [8], we compute in [10] the group H 2(cid:16)Zh 1
pi up to equivalence, as follows:
tipliers of Zh 1
pi up to equivalence, where our multipliers
pi has no nontrivial multiplier, so our noncommutative solenoids are the
pi, T(cid:17) of T-valued mul-
pi × Zh 1
pi × Zh 1
pi × Zh 1
natural object to consider.
Theorem 2.4. [10, Theorem 2.3] Let p ∈ N, p > 1. Let:
Ξp = {(αn) : α0 ∈ [0, 1) ∧ (∀n ∈ N ∃k ∈ {0, . . . , N − 1} pαn+1 = αn + k)}
which is a group for the pointwise addition modulo one. There exists a group iso-
and α = ρ(σ), and if f is a multiplier of class σ, then f is cohomologous to:
morphism ρ : H 2(cid:16)Zh 1
Ψα :(cid:18)(cid:18) q1
pk1
,
pi × Zh 1
pk2(cid:19) ,(cid:18) q3
pi, T(cid:17) → Ξp such that if σ ∈ H 2(cid:16)Zh 1
pk4(cid:19)(cid:19) 7−→ exp(cid:0)2iπα(k1+k4)q1q4(cid:1) .
pk3
q4
q2
,
pi × Zh 1
pi, T(cid:17)
For any p ∈ N, p > 1, the groups Ξp and Sp are obviously isomorphic as topo-
logical groups; yet it is easier to perform our computations in the additive group Ξp
pi, T(cid:17) is isomorphic
in what follows. Thus, as a topological group, H 2(cid:16)Zh 1
pi × Zh 1
to Sp. Moreover, we observe that a corollary of Theorem (2.4) is that Ψα and Ψβ
are cohomologous if and only if α = β ∈ Ξp. The proof of Theorem (2.4) involves
the standard calculations for cohomology classes of multipliers on discrete Abelian
groups, due to A. Kleppner, generalizing results of Backhouse and Bradley.
With this understanding of the multipliers of Zh 1
pi, we thus propose
pi, σ(cid:17). Let us start by
to classify the noncommutative solenoids C∗(cid:16)Zh 1
is the C∗-completion of the involutive Banach algebra (cid:0)ℓ1 (Γ) , ∗σ, ·∗(cid:1), where the
recalling [20] that for any multiplier σ of a discrete group Γ, the C*-algebra C∗ (Γ, σ)
pi × Zh 1
pi × Zh 1
twisted convolution ∗σ is given for any f1, f2 ∈ ℓ1 (Γ) by
f1(γ1)f2(γ − γ1)σ(γ1, γ − γ1),
f1 ∗σ f2 : γ ∈ Γ 7−→ Xγ1∈Γ
while the adjoint operation is given by:
f ∗
1 : γ ∈ Γ 7−→ σ(γ, −γ)f1(−γ).
The C*-algebra C∗ (Γ, σ) is then shown to be the universal C*-algebra generated by
a family (Wγ)γ∈Γ of unitaries such that WγWδ = σ(γ, δ)Wγδ for any γ, δ ∈ Γ [20].
We shall henceforth refer to these generating unitaries as the canonical unitaries of
C∗ (Γ, σ).
4
FR ´ED ´ERIC LATR ´EMOLI `ERE AND JUDITH A. PACKER
One checks easily that if σ and η are two cohomologous multipliers of the discrete
group Γ, then C∗ (Γ, σ) and C∗ (Γ, η) are *-isomorphic [20]. Thus, by Theorem
Noncommutative solenoids, defined in Definition (2.3) as twisted group algebras
Notation 2.5. For any p ∈ N, p > 1 and for any α ∈ Ξp, the C*-algebra
the form Ψα with α ∈ Ξp. With this in mind, we introduce the following notation:
pi × Zh 1
pi of
(2.4), we shall henceforth restrict our attention to multipliers of Zh 1
C∗(cid:16)Zh 1
pi, Ψα(cid:17), with Ψα defined in Theorem (2.4), is denoted by A S
pi × Zh 1
pi, also have a presentation as transformation group C∗-algebras, in
pi × Zh 1
of Zh 1
action of Zh 1
pi on Sp defined for all q
The C*-crossed-product C(Sp) ⋊θα Zh 1
pk ∈ Zh 1
pi is *-isomorphic to A S
((zn)n∈N) =(cid:16)e(2iπα(k+n)q)zn(cid:17)n∈N
pi and for all (zn)n∈N ∈ Sp by:
Proposition 2.6. [10, Proposition 3.3] Let p ∈ N, p > 1 and α ∈ Ξp. Let θα be the
a manner similar to the situation with rotation C*-algebras:
Whichever way one decides to study them, there are longstanding methods in
place to determine whether or not these C∗ algebras are simple (see for instance []).
For now, we concentrate on methods from the theory of twisted group C∗-algebras.
α .
θα
q
pk
.
α .
Theorem-Definition 2.6.1. [13, Theorem 1.5] The symmetrizer group of a mul-
tiplier σ : Γ × Γ → T of a discrete group Γ is given by
Sσ =(cid:8)γ ∈ Γ : ∀g ∈ Γ σ(γ, g)σ(g, γ)−1 = 1(cid:9) .
The C*-algebra C∗(Γ, σ) is simple if, and only if the symmetrizer group Sσ is
reduced to the identity of Γ.
In [10], we thus characterize when the symmetrizer group of the multipliers of
Zh 1
pi × Zh 1
pi given by Theorem (2.4) is non-trivial:
Theorem 2.7. [10, Theorem 2.12] Let p ∈ N, p > 1. Let α ∈ Ξp. Denote by Ψα
the multiplier defined in Theorem (2.4). The following are equivalent:
(1) the symmetrizer group SΨα is non-trivial,
(2) the sequence α has finite range, i.e. the set {αj : j ∈ N} is finite,
(3) there exists k ∈ N such that (pk − 1)α0 ∈ Z,
(4) the sequence α is periodic,
(5) there exists a positive integer b ∈ N such that:
SΨα = bZ(cid:20) 1
p(cid:21) × Z(cid:20) 1
p(cid:21) =(cid:26)(br1, br2), (r1, r2) ∈ Z(cid:20) 1
p(cid:21) × Z(cid:20) 1
p(cid:21)(cid:27) .
Theorem (2.8), when applied to noncommutative solenoids via Theorem (2.7),
allows us to conclude:
Theorem 2.8. [10, Theorem 3.5] Let p ∈ N, p > 1 and α ∈ Ξp. Then the following
are equivalent:
(1) the noncommutative solenoid A S
(2) the set {αj : j ∈ N} is infinite,
α is simple,
NONCOMMUTATIVE SOLENOIDS AND THEIR PROJECTIVE MODULES
5
(3) for every k ∈ N, we have (pk − 1)α0 6∈ Z.
In particular, if α ∈ Ξp is chosen with at least one irrational entry, then by
definition of Ξp, all entries of α are irrational, and by Theorem (2.8), the noncom-
mutative solenoid A S
α is simple. The reader may observe that, even if α ∈ Ξp only
has rational entries, the noncommutative solenoid may yet be simple -- as long as
α has infinite range. We called this situation the aperiodic rational case in [10].
Example 2.9 (Aperiodic rational case). Let p = 7, and consider α ∈ Ξ7 given by
α =(cid:18) 2
7
,
2
49
,
2
343
,
2
2401
, · · ·(cid:19) =(cid:18) 2
7n(cid:19)n∈N
.
Note that αj ∈ Q for all j ∈ N, yet Theorem (2.8) tells us that the noncommutative
solenoid A S
α is simple!
The following is an example where the symmetrizer subgroup is non-trivial, so
that the corresponding C∗-algebra is not simple.
Example 2.10 (Periodic rational case). Let p = 5, and consider α ∈ Ξ5 given by
α =(cid:18) 1
62
,
25
62
,
5
62
,
1
62
, · · ·(cid:19) .
Theorem (2.7) shows that the symmetrizer group of the multiplier Ψα of(cid:0)Z(cid:2) 1
5(cid:3)(cid:1)2
given by Theorem (2.4) is:
Sα =(cid:26)(cid:18) 62j1
5k ,
62j2
5k (cid:19) ∈ Q : j1, j2 ∈ Z, k ∈ N(cid:27) .
Hence the noncommutative solenoid A S
α is not simple by Theorem (2.8).
We conclude this section with the following result about the existence of traces
on noncommutative solenoids, which follows from [7], since the Pontryagin dual
α for any α ∈ Ξp via the dual
Sp × Sp of Zh 1
action:
pi × Zh 1
pi acts ergodically on A S
Theorem 2.11. [10, Theorem 3.8] Let p ∈ N, p > 1 and α ∈ Ξp. There exists at
least one tracial state on the noncommutative solenoid A S
α . Moreover, this tracial
state is unique if, and only if α is not periodic.
Moreover, since noncommutative solenoids carry an ergodic action of the com-
pact groups Sp, if one chooses any continuous length function on Sp, then one may
employ the results found in [18] to equip noncommutative solenoids with quantum
compact metric spaces structures and, for instance, use [19] and [9] to obtain vari-
ous results on continuity for the quantum Gromov-Hausdorff distance of the family
of noncommutative solenoids as the multiplier and the length functions are left to
vary. In this paper, we shall focus our attention on the noncommutative topology
of our noncommutative solenoids, rather than their metric properties.
In [10, Theorem 3.17], we provide a full description of noncommutative solenoids
as bundles of matrix algebras over the space S 2
p , while in contrast, in [10, Propo-
sition 3.16], we note that for α with at least (and thus all) irrational entry, the
noncommutative solenoid A S
α is an inductive limit of circle algebras (i.e. AT), with
real rank zero. Both these results follow from writing noncommutative solenoids as
inductive limits of quantum tori, which is the starting point for our next section.
6
FR ´ED ´ERIC LATR ´EMOLI `ERE AND JUDITH A. PACKER
3. Classification of the noncommutative Solenoids
Noncommutative solenoids are classified by their K-theory; more precisely by
their K0 groups and the range of the traces on K0. The main content of our
paper [10] is the computation of the K-theory of noncommutative solenoids and its
application to their classification up to *-isomorphism.
The starting point of this computation is the identification of noncommutative
solenoids as inductive limits of sequences of noncommutative tori. A noncommuta-
tive torus is a twisted group C*-algebra for Zd, with d ∈ N, d > 1 [14]. In particular,
for d = 2, we have the following description of noncommutative tori. Any multiplier
of Z2 is cohomologous to one of the form:
σθ :(cid:18)(cid:18)z1
z2(cid:19) ,(cid:18)y1
y2(cid:19)(cid:19) 7−→ exp(2iπθz1y2)
for some θ ∈ [0, 1). Consequently, for a given θ ∈ [0, 1), the C*-algebra C∗(cid:0)Z2, σθ(cid:1)
is the universal C*-algebra generated by two unitaries U, V such that:
U V = e2iπθV U .
We will employ the following notation throughout this paper:
Notation 3.1. The noncommutative torus C∗(cid:0)Z2, σθ(cid:1), for θ ∈ [0, 1), is denoted by
Aθ. Moreover, the two canonical generators of Aθ (i.e. the unitaries corresponding
to (1, 0), (0, 1) ∈ Z2), are denoted by Uθ and Vθ, so that UθVθ = e2iπθVθUθ.
For any θ ∈ [0, 1), the noncommutative torus Aθ is *-isomorphic to the crossed-
product C*-algebra for the action of Z on the circle T generated by the rotation of
angle 2iπθ, and thus Aθ is also known as the rotation algebra for the rotation of
angle θ -- a name by which it was originally known.
pi × Zh 1
Zh 1
The following question naturally arises: since Aθ is a twisted Z2 algebra, and
pi can be realized as a direct limit group built from embeddings of Z2
into itself, is it possible to build our noncommutative solenoids A S
α as a direct
limits of rotation algebras? The answer is positive, and this observation provides
much structural information regarding noncommutative solenoids.
Theorem 3.2. [10, Theorem 3.7] Let p ∈ N, p > 1 and α ∈ Ξp. For all n ∈ N, let
ϕn be the unique *-morphism from Aα2n into Aα2n+2 given by:
(Uα2n 7−→ U p
Vα2n 7−→ V p
α2n+2
α2n+2
Then:
Aα0
ϕ0−→ Aα2
ϕ1−→ Aα4
ϕ2−→ · · ·
converges to the noncommutative solenoid A S
is the family of canonical unitary generators of A S
tion algebra Aα2n embeds in A S
α . Moreover, if (Wr1,r2)(r1,r2)∈Z[ 1
p ]×Z[ 1
p ]
α , then, for all n ∈ N, the rota-
α via the unique extension of the map:
Uα2n 7−→ W( 1
Vα2n 7−→ W(0, 1
pn ,0)
pn ).
to a *-morphism, given by the universal property of rotation algebras; one checks
that this embeddings, indeed, commute with the maps ϕn.
NONCOMMUTATIVE SOLENOIDS AND THEIR PROJECTIVE MODULES
7
Our choice of terminology for noncommutative solenoids is inspired, in part, by
Theorem (3.2), and the well established terminology of noncommutative torus for
rotation algebras. Moreover, as we shall now see, our study of noncommutative
solenoids is firmly set within the framework of noncommutative topology.
The main result from our paper [10] under survey in this section and the pre-
vious one is the computation of the K-theory of noncommutative solenoid and its
application to their classification. An interesting connection between our work on
which in turn are classified by means of the group of p-adic integers, emerges as a
consequence of our computation. We shall present this result now, starting with
noncommutative solenoid and classifications of Abelian extensions of Zh 1
some reminders about the p-adic integers and Abelian extensions of Zh 1
pi by Z,
pi, and refer
to [10] for the involved proof leading to it.
Theorem-Definition 3.2.1. Let p ∈ N, p > 1. The set:
Zp =(cid:8)(Jk)k∈N : J0 = 0 and ∀k ∈ N Jk+1 ≡ Jk mod pk(cid:9)
is a group for the operation defined as:
(Jk)k∈N + (Kk)K∈N = ((Jk + Kk) mod pk)k∈N
for any (Jk)k∈N, (Kk)k∈N ∈ Zp. This group is the group of p-adic integers.
following manner:
One may define the group of p-adic integer simply as the set of sequences valued
in {0, . . . , p − 1} with the appropriate operation, but our choice of definition will
make our exposition clearer. We note that we have a natural embedding of Z as
a subgroup of Zp by sending z ∈ Z to the sequence (z mod pk)k∈N. We shall
henceforth identify Z with its image in Zp when no confusion may arise.
Theorem 3.3. [10] Let p ∈ N, p > 1 and let J = (Jk)k∈N ∈ Zp. Define the map
pi, i.e. a map
We can associate, to any p-adic integer, a Schur multiplier of Zh 1
pi × Zh 1
pi → Z which satisfies the (additive) 2-cocycle identity, in the
ξj : Zh 1
ξJ : Zh 1
pi × Zh 1
pi → Z by setting, for any q1
pk2(cid:19) =
ξJ(cid:18) q1
where all fractions are written in their reduced form, i.e. such that the exponent
of p at the denominator is minimal (this form is unique). Then:
if k2 > k1,
if k1 > k2,
if k1 = k2, with q
− q1
− q2
q
pr (Jk1 − Jr)
pi:
pk2 ∈ Zh 1
pk1 (Jk2 − Jk1 )
pk2 (Jk1 − Jk2 )
pk1 , q2
pr = q1
pk1 + q2
pk2 ,
• For any J, K ∈ Zp, the Schur multipliers ξJ and ξK are cohomologous if,
• ξJ is a Schur multiplier of Zh 1
• Any Schur multiplier of Zh 1
pi [10, Lemma 3.11].
pi is cohomologous to ξJ for some J ∈ Zp [10,
and only if J − K ∈ Z [10, Theorem 3.14].
Theorem 3.16].
q2
,
pk1
In particular, Ext(cid:16)Zh 1
pi, Z(cid:17) is isomorphic to Zp /Z .
8
FR ´ED ´ERIC LATR ´EMOLI `ERE AND JUDITH A. PACKER
Schur multipliers provide us with a mean to describe and classify Abelian ex-
tensions of Zh 1
pi by Zp. Our interest in Theorem (3.3) lies in the remarkable
observation that the K0 groups of noncommutative solenoids are exactly given by
these extensions:
be the Schur multiplier of Zh 1
with underlying set Z × Zh 1
for all (z1, r1), (z2, r2) ∈ Z × Zh 1
of Zh 1
Then:
Theorem 3.4. [10, Theorem 3.12] Let p ∈ N, p > 1 and let α = (αk)k∈N ∈ Ξp. For
any k ∈ N, define Jk = pkαk − α0, and note that by construction, J ∈ Zp. Let ξJ
pi defined in Theorem (3.3), and let QJ be the group
pi and operation:
(z1, r1) ⊞ (z2, r2) = (z1 + z2 + ξJ (r1, r2) , r1 + r2)
1
Furthermore, we have:
α (cid:1), char-
and, moreover, all tracial states of A S
acterized by:
pi. By construction, QJ is an Abelian extension
pi by Z given by the Schur multiplier ξJ .
K0(cid:0)A S
α (cid:1) = QJ
α lift to a single trace τ on K0(cid:0)A S
τ : (1, 0) 7→ 1 and (cid:18)0,
α (cid:1) = Z(cid:20) 1
K1(cid:0)A S
pk(cid:19) 7→ αk.
p(cid:21) × Z(cid:20) 1
p(cid:21).
We observe, in particular, that given any Abelian extension of Zh 1
pi by Z, one
pi of the form ξJ for some
can find, by Theorem (3.3), a Schur multiplier of Zh 1
α =(cid:16) α0+Jk
only Abelian extensions of Zh 1
pi by Z are given as K0 groups of noncommutative
J ∈ Zp, and, up to an arbitrary choice of α0 ∈ [0, 1), one may form the sequence
, and check that α ∈ Ξp; thus all possible Abelian extensions, and
solenoids. With this observation, the K0 groups of noncommutative solenoids are
uniquely described by a p-adic integer modulo an integer, and the information
contained in the pair (K0(A S
α ), τ ) of the K0 group of a noncommutative solenoid
and its trace, is contained in the pair (J, α0) with J ∈ Zp /Z as defined in Theorem
(3.4).
pk (cid:17)k∈N
Remark 3.5. For any p ∈ N, p > 1 and α ∈ Ξp, the range of the unique trace
τ on K0(A S
α ), as described by Theorem (3.4), is the subgroup Z ⊕ ⊕k∈NαnZ.
Let γ = z + z1αn1 + . . . zkαnk be an arbitrary element of this set, where, to fix
notations, we assume n1 < . . . < nk. Then, since αn+1 ≡ pαn mod 1 for any
n ∈ N, we conclude that we can rewrite γ simply as z′ + yαnk , for some z′, y ∈ Z.
Thus the range of our trace on K0(A S
α ) is given by:
τ(cid:0)K0(cid:0)A S
α (cid:1)(cid:1) = {z + yαk : z, y ∈ Z, k ∈ N} .
We thus have a complete characterization of the K-theory of noncommutative
solenoids. This noncommutative topological invariant, in turn, contains enough
information to fully classify noncommutative solenoids in term of their defining
NONCOMMUTATIVE SOLENOIDS AND THEIR PROJECTIVE MODULES
9
multipliers. We refer to [10, Theorem 4.2] for the complete statement of this clas-
sification; to keep our notations at a minimum, we shall state the corollary of [10,
Theorem 4.2] when working with p prime:
Theorem 3.6. [10, Corollary 4.3] Let p, q be two prime numbers and let α ∈ Ξp
and β ∈ Ξq. Then the following are equivalent:
(1) The noncommutative solenoids A S
(2) p = q and a truncated subsequence of α is a truncated subsequence of β or
β are *-isomorphic,
α and A S
(1 − βk)k∈N.
Theorem (3.6) is given in greater generality in [10, Theorem 4.2], where p, q are
not assumed prime; the second assertion of the Theorem must however be phrased
in a more convoluted manner: essentially, p and q must have the same set of prime
factors, and there is an embedding of Ξp and Ξq in a larger group Ξ, whose elements
are still sequences in [0, 1), such that the images of α and β for these embeddings
are sub-sequences of a single element of Ξ.
We conclude this section with an element of the computation of the K0 groups
exists a Rieffel-Powers projection in Aα2k whose image in A S
α for the embedding
given by Theorem (3.2) has K0 class the element γ, whose trace is thus naturally
given by Theorem (3.4). Much work is needed, however, to identify the range of
pk(cid:17) ∈ K0(cid:0)A S
in Theorem (3.4). Given γ = (cid:16)0, 1
K0 as the set of all Abelian extensions of Zh 1
turn, by Zp /Z , as we have shown in this section.
α (cid:1), if α0 is irrational, then there
pi by Z, and parametrize these, in
We now turn to the question of the structure of the category of modules over
noncommutative solenoids. In the next two sections, we show how to apply some
constructions of equivalence bimodules to the case of noncommutative solenoids
as a first step toward solving the still open problem of Morita equivalence for
noncommutative solenoids.
4. Forming projective modules over noncommutative solenoids from
the inside out
Projective modules for rotation algebras and higher dimensional noncommuta-
tive tori were studied by M. Rieffel ([16]). F. Luef has extended this work to build
modules with a dense subspace of functions coming from modulation spaces (e.g.,
Feichtinger's algebra) with nice properties ([11], [12]). One approach to building
projective modules over noncommutative solenoids is to build the projective mod-
ules from the "inside out".
We first make some straightforward observations in this direction. We recall that,
by Notation (2.5), for any p ∈ N, p > 1, and for any α ∈ Ξp, where Ξp is defined in
Theorem (2.4), the C*-algebra C∗(cid:16)Zh 1
pi × Zh 1
pi, Ψα(cid:17), where the multiplier Ψα
was defined in Theorem (2.4), is denoted by A S
α . In this section, we will work with
p a prime number. Last, we also recall that by Notation (3.1), the rotation algebra
for the rotation of angle θ ∈ [0, 1) is denoted by Aθ, while its canonical unitary
generators are denoted by Uθ and Vθ, so that UθVθ = e2iπθVθUθ.
Theorem (3.4) describes the K0 groups of noncommutative solenoids, and, among
other conclusions, state that there always exists a unique trace on the K0 of any
10
FR ´ED ´ERIC LATR ´EMOLI `ERE AND JUDITH A. PACKER
noncommutative solenoid, lifted from any tracial state on the C*-algebra itself.
With this in mind, we state:
Proposition 4.1. Let p be a prime number, and fix α ∈ Ξp, with α0 6∈ Q. Let
γ = z + qαN for some z, q ∈ Z and N ∈ N, with γ > 0. Then there is a left
projective module over A S
α whose K0 class has trace γ, or equivalently, whose K0
class is given by(cid:16)z, q
pi.
pk(cid:17) ∈ Z × Zh 1
Proof. By Remark (3.5), γ is the image of some class in K0(A S
α ) for the trace on
this group. Now, since αN +1 = pαN + j for some j ∈ Z by definition of Ξα, we may
as well assume N is even. As K0(A S
α ) is the inductive limit of K0(Aαk )k∈2N by
Theorem (3.2), γ is the trace of an element of K0(AαN )), where AαN is identified
as a subalgebra of A S
α (again using Theorem (3.2). By [14], there is a projection
Pγ in AαN whose K0 class has trace γ, and it is then easy to check that the left
projective module P A S
(cid:3)
α fulfills our proposition.
α over A S
So, for example, with the notations of the proof of Proposition (4.1), if Pγ is a
α with trace γ ∈ (0, 1), one can construct the equivalence
projection in Aαn ⊂ A S
bimodule
A S
α
− A S
α Pγ − Pγ A S
α Pγ.
From this realization, not much about the structure of Pγ A S
α Pγ can be seen,
although it is possible to write this C∗-algebra as a direct limit of rotation algebras.
Let us now discuss this matter.
Suppose we have two directed sequences of C∗-algebras:
ϕ2−−−−→ · · ·
ϕ0−−−−→ A1
ϕ1−−−−→ A2
A0
and
B0
ψ0−−−−→ B1
ψ1−−−−→ B2
ψ2−−−−→ · · ·
Suppose further that for each n ∈ N there is an equivalence bimodule Xn between
An and Bn
An − Xn − Bn,
and that the (Xn)n∈N form a directed system, in the following sense: there exists
a direct system of module monomorphisms
X0
i0−−−−→ X1
i1−−−−→ X2
i2−−−−→ · · ·
satisfying, for all f, g ∈ Xn and b ∈ Bn:
hin(f ), in(g)iBn+1 = ψn(hf, giBn )
and
in(f · b) = in(f ) · ψn(b),
with analogous but symmetric equalities holding for the Xn viewed as left-An mod-
ules.
Now let A be the direct limit of (An)n∈N, B be the direct limit of (Bn)n∈N and X
be the direct limit of (Xn)n∈N (completed in the natural C∗-module norm). Then
X is an A − B bimodule. If one further assumes that the algebra of adjointable
operators on X viewed as a A − B bimodule, L(X ), can be obtained via an appro-
priate limiting process from the sequence of adjointable operators {L(Xn)}∞
n=1 (
NONCOMMUTATIVE SOLENOIDS AND THEIR PROJECTIVE MODULES
11
where each Xn is a An − Bn bimodule), then in addition one has that X is a strong
Morita equivalence bimodule between A and B.
So suppose that γ ∈ (0, 1) is as in the statement of Proposition (4.1), for some
α ∈ Ξp not equal to zero, and suppose that we know that there is a positive integer
N and a projection Pγ in AαN whose K0 class has trace γ. Again, without loss of
generality, we assume that N is even. Then setting
An = AαN +2n , Xn = AαN +2n Pγ, and Bn = PγAαN +2n Pγ,
all of the conditions in the above paragraphs hold a priori, since A S
α is a direct limit
of the AαN +2n, so that certainly B = Pγ A S
α Pγ is a direct limit of the PγAαN +2nPγ,
and X = A S
α Pγ can be expressed as a direct limit of the Xn = AαN +2nPγ, again
by construction, with the desired conditions on the adjointable operators satisfied
by construction.
It would be interesting to see how far this set-up could be extended to more
general directed systems of Morita equivalence bimodules over directed systems of
C∗-algebras, but we leave this project to a future endeavor.
We discuss very simple examples, to show how the directed system of bimodules
is constructed.
Example 4.2. Fix an irrational α0 ∈ [0, 1), let p = 2, and consider α ∈ Ξ2 given
by
α = (α0, α1 =
, α2 =
, · · · , αn =
αn
2n , · · · , ),
α0
2
α0
4
Consider Pα0 ∈ Aα0 ⊂ Aα1 a projection of trace α0 = 2α1. The bimodule
is equivalent to Rieffel's bimodule
Aα0 − Aα0 · Pα0 − Pα0 Aα0 Pα0
Aα0 − Cc(R) − A 1
α0
= B0.
Let β0 = 1
α0
a bimodule
. Rieffel's theory, specifically Theorem 1.1 of [15], again shows there is
is the same as
Aα2 − Aα2 · Pα0 − Pα0 Aα2 Pα0
Aα2 − Aα2 · P4α2 − P4α2 Aα2 P4α2
which is equivalent to Rieffel's bimodule
Aα2 − Cc(R × F4) − C(T × F4) ⋊τ1 Z = B1,
where F4 = Z/4Z, and the action of Z on T × F4 is given by multiples of ( β2
for β2 = 1
α2
, [1]F4]), i.e. multiples of (β0, [1]F4).
, i.e. multiples of ( 1
α0
4 , [1]F4),
At the nth stage, using Theorem 1.1 of [15] again, we see that
is the same as
Aα2n − Aα2n · Pα0 − Pα0 Aα2n Pα0
Aα2n − Aα2n · P2nαn − P2nαn Aαn P2nαn
which is equivalent to
Aαn − Cc(R × F4n ) − C(T × F4n ) ⋊τn Z = Bn,
12
FR ´ED ´ERIC LATR ´EMOLI `ERE AND JUDITH A. PACKER
where the action of Z on T × F4n is given by multiples of ( β2n
1
4n , [1]F4n ), for β2n =
, [1]F4n ), i.e. multiples of (β0, [1]F4n ), for F4n =
, i.e. multiples of ( 1
α0
= 4n
α0
α2n
Z/4nZ.
From calculating the embeddings, we see that for α = (α0, α0
2 , · · · , α
2n , · · · ) ∈ Ξ2,
we have that
AS
α
is strongly Morita equivalent to a direct limit B of the Bn. The structure of B is
not clear in this description, although each Bn is seen to be a variant of a rotation
algebra. As expected, one calculates
tr(K0(AS
α)) = α0 · tr(K0(B)).
5. Forming projective modules over noncommutative solenoids using
p-adic fields
Under certain conditions, one can construct equivalence bimodules for A S
α (α ∈
Ξp,p prime) by using a construction of M. Rieffel [16]. The idea is to first embed
pi as a co-compact 'lattice' in a larger group M , and the quotient
group M /Γ will be exactly the solenoid Sp. We thank Jerry Kaminker and Jack
Spielberg for telling us about this trick.
pi × Zh 1
Γ = Zh 1
We start with a brief description of the field of p-adic numbers, with p prime.
Algebraically, the field Qp is the field of fraction of the ring of p-adic integers Zp --
we introduce Zp as a group, though there is a natural multiplication on Zp turning
it into a ring. A more analytic approach is to consider Qp as the completion of the
field Q for the p-adic metric dp, defined by dp(r, r′) = r − r′p for any r, r′ ∈ Q,
where · p is the p-adic norm defined by:
r =(p−n if r 6= 0 and where r = pn a
0 if r = 0.
b with a, b are both relatively prime with p,
If we endow Q with the metric dp, then series of the form:
ajpj
∞Xj=k
will converge, for any k ∈ Z and aj ∈ {0, . . . , p − 1} for all j = k, . . .. This is
the p-adic expansion of a p-adic number. One may easily check that addition and
multiplication on Q are uniformly continuous for dp and thus extend uniquely to Qp
to give it the structure of a field. Moreover, one may check that the group Zp of p-
adic integer defined in Section 3 embeds in Qp as the group of p-adic numbers of the
j=0 ajpj with aj ∈ {0, . . . , p − 1} for all j ∈ N. Now, with this embedding,
one could also check that Zp is indeed a subring of Qp whose field of fractions is Qp
(i.e. Qp is the smallest field containing Zp as a subring) and thus, both constructions
described in this section agree. Last, the quotient of the (additive) group Qp by its
subgroup Zp is the Prufer p-group Z(p∞) = {z ∈ T : ∃n ∈ N z(pn) = 1}.
formP∞
5.1. Embedding Z( 1
completion of Q and Zh 1
p ) as a lattice in a self-dual group. Since Qp is a metric
pi is a subgroup of Q, we shall identify, in this section,
NONCOMMUTATIVE SOLENOIDS AND THEIR PROJECTIVE MODULES
13
Zh 1
pi as a subgroup of Qp with no further mention. We now define a few group
homomorphisms to construct a short exact sequence at the core of our construction.
Let ω : R → Sp be the standard "winding line" defined for any t ∈ R by:
ω(t) =(cid:16)e2πit, e2πi t
p , e2πi t
p2 , · · · , e2πi t
pn , · · ·(cid:17) .
in {0, . . . , p − 1}. We define the sequence ζ(γ) by setting for all j ∈ N:
j=k ajpj for a (unique) family (aj)j=k,... of elements
Let γ ∈ Qp and write γ =P∞
with the convention thatPk
We thus may define the map
ζj(γ) = e
2πi(cid:16)Pj
m=k
am
pj−m+k (cid:17)
j · · · is zero if k < j.
Π :(Qp × R −→ Sp
γ
7−→ Π(γ, t) = ζj (γ) · ω(t).
then one checks that the following is an exact sequence:
ι−−−−→ Qp × R
Π−−−−→ Sp −−−−→ 1
pi −→ Qp × R
7−→ ι(r) = (r, −r),
If we set
r
ι :(Zh 1
pi
1 −−−−→ Zh 1
1 −−−−→ Zh 1
pi × Zh 1
different terms in Zh 1
family of different embeddings of Zh 1
It follows that there is an exact sequence
Indeed, we will show later that it is possible to perturb the embeddings of the
pi −−−−→ [Qp × R] × [Qp × R] −−−−→ Sp × Sp −−−−→ 1.
pi by elements of Qp \ {0} and R \ {0} to obtain a
pi × Zh 1
pi × Zh 1
pi into [Qp × R]2.
We now observe that M = Qp × R is self-dual. We shall use the following
standard notation:
Notation 5.1. The Pontryagin dual of a locally compact group G is denoted by
bG. The dual pairing between a group and its dual is denoted by h·, ·i : G × bG → T.
Let us show that M ∼= cM . To every x ∈ Qp, we can associate the character
χx : q ∈ Qp 7→ e2iπi{x·q}
where {x · q}p is the fractional part of the product x · q in Qp, i.e. it is the sum of
the terms involving the negative powers of p in the p-adic expansion of x · q. The
map x ∈ Qp 7→ χx ∈ cQp is an isomorphism of topological group. Similarly, every
character of R is of the form χr : t ∈ R 7→ e2iπrt for some r ∈ R. Therefore every
character of M is given by
χ(x,r) : (q, t) ∈ Qp × R 7−→ χx(q)χr(t)
for some (x, r) ∈ Qp × R (see [6]) for further details on characters of specific locally
compact abelian groups). It is possible to check that the map (x, r) 7→ χ(x,r) is a
group isomorphism between M and cM , so that M = Qp × R is indeed self-dual.
14
FR ´ED ´ERIC LATR ´EMOLI `ERE AND JUDITH A. PACKER
5.2. The Heisenberg representation and the Heisenberg equivalence bi-
prime number, and now let M = [Qp × R]. We have shown in the previous section
that M is self-dual, since both Qp and R are self-dual. Now suppose there is an
pi where p is some
module of Rieffel. In this section, we write Γ = Zh 1
embedding ι : Γ → M ×cM . Let the image ι(Γ) be denoted by D. In the case we are
considering, D is a discrete co-compact subgroup of M ×cM . Following the method
of M. Rieffel [16], the Heisenberg multiplier η : (M × cM ) × (M × cM ) → T is
pi × Zh 1
η((m, s), (n, t)) = hm, ti, (m, s), (n, t) ∈ M ×cM .
(We note we use the Greek letter 'η' rather than Rieffel's 'β', because we have used
'β' elsewhere. Following Rieffel, the symmetrized version of η is denoted by the
letter ρ, and is the multiplier defined by:
defined by:
L2(M ), defined as π, where
ρ((m, s), (n, t)) = η((m, s), (n, t))η((n, t), (m, s)), (m, s), (n, t) ∈ M ×cM .
M. Rieffel [16] has shown that CC (M ) can be given the structure of a left
C∗(D, η) module, as follows. One first constructs an η-representation of M ×cM on
π(m,s)(f )(n) = hn, si f (n + m), (m, s) ∈ M ×cM , n ∈ M .
When the representation π is restricted to D, we still have a projective η-representation
of D, on L2(M ), and its integrated form gives CC (M ) the structure of a left C∗(D, η)
module, i.e. for Φ ∈ CC (D, η), f ∈ CC (M ),
Φ((d, χ))π(d,χ)(f )(n)
Φ((d, χ)) hn, χi f (n + d).
π(Φ) · f (n) = X(d,χ)∈D
= X(d,χ)∈D
f (n)π(d,χ)(g)(n)dn = ZM
There is also a CC (D, η) valued inner product defined on CC (M ) given by:
Moreover, Rieffel has shown that setting
hf, giCC (D,η) =ZM
D⊥ = {(n, t) ∈ M ×cM : ∀(m, s) ∈ D ρ((m, s), (n, t)) = 1},
CC (M ) has the structure of a right C∗(D⊥, η) module. Here the right module
structure is given for all f ∈ Cc(M ), Ω ∈ Cc(D⊥) and n ∈ M by:
f (n)hn, χi g(n + d)dn.
f · Ω(n) = X(c,ξ)∈D⊥
π∗
(c,ξ)(f )(n)Ω(c, ξ),
and the CC (D⊥, η)-valued inner product is given by
hf, giCC (D⊥,η)(c, ξ) =ZM
f (n)π(c,ξ)(g)(n)dn =ZM
where f, g ∈ CC (M ), Ω ∈ CC (D⊥, η), and (c, ξ) ∈ D⊥.
f (n) hn, ξi g(n + c)dn,
Moreover, Rieffel shows in [16, Theorem 2.12] that C∗(D, η) and C∗(D⊥, η) are
strongly Morita equivalent, with the equivalence bimodule being the completion of
CC (M ) in the norm defined by the above inner products.
NONCOMMUTATIVE SOLENOIDS AND THEIR PROJECTIVE MODULES
15
In order to construct explicit bimodules, we first define the multiplier η more
precisely, and then discuss different embeddings of Zh 1
pi × Zh 1
In the case examined here, the Heisenberg multiplier η : [Qp×R]2×[Qp×R]2 → T
pi into M ×cM .
is given by:
Definition 5.2. The Heisenberg multiplier η : [Qp × R]2 × [Qp × R]2 → T is defined
by
η[((q1, r1), (q2, r2)), ((q3, r3), (q4, r4))] = e2πir1r4 e2πi{q1q4}p ,
where {q1q4}p is the fractional part of the product q1 · q4, i.e. the sum of the terms
involving the negative powers of p in the p-adic expansion of q1q4.
pi × Zh 1
The following embeddings of Zh 1
Definition 5.3. For θ ∈ R, θ 6= 0, we define ιθ : Zh 1
different embeddings of Zh 1
pi:
We start by observing that for r1, r2, r3, r4 ∈ Zh 1
pi in [Qp × R]2 will prove interesting:
pi → [Qp × R]2 by
pi into [Qp × R]2 and their influence on the
We examine the structure of the multiplier η more precisely and then discuss
different equivalence bimodules they allow us to construct.
pi × Zh 1
pi × Zh 1
ιθ(r1, r2) = [(r1, θ · r1), (r2, r2)].
η(ιθ(r1, r2)), ιθ(r3, r4)) = e2πi{r1r4}p e2πiθr1r4
= e2πir1r4e2πiθr1r4 = e2πi(θ+1)r1r4.
(Here we used the fact that for ri, rj ∈ Z( 1
One checks that setting Dθ = ιθ(cid:16)Zh 1
exactly *-isomorphic to the noncommutative solenoid A S
pi(cid:17), the C∗-algebra C∗(Dθ, η) is
α for
p ), {rirj }p ≡ rirj modulo Z.)
θ + 1
pi × Zh 1
pn , · · ·(cid:19) =(cid:18) θ + 1
pi × Zh 1
pn (cid:19)n∈N
.
pi as the discrete subgroup D inside
θ =(cid:26)(cid:16)r1, −
θ(cid:17) , (r2, −r2)i ,h(cid:16)r3, −
θ(cid:17) , (r2, −r2) : r1, r2 ∈ Z(cid:20) 1
p(cid:21)(cid:27) .
θ(cid:17) , (r4, −r4)i(cid:17) = e−2πi( 1
θ , η) is also a non-commutative solenoid A S
r3
r1
θ +1)r1r4.
β where β =
α =(cid:18)θ + 1,
θ + 1
p
, · · · ,
For this particular embedding of Zh 1
M ×cM , we calculate that
D⊥
r1
Moreover,
η(cid:16)h(cid:16)r1, −
pnθ(cid:17)n∈N
.
Note that for
(cid:16)1 − θ+1
It is evident that C∗(D⊥
and
α =(cid:18)θ + 1,
θ + 1
p
, · · · ,
θ + 1
pn , · · ·(cid:19) ,
β =(cid:18)1 −
θ + 1
pnθ (cid:19)n∈N
,
16
FR ´ED ´ERIC LATR ´EMOLI `ERE AND JUDITH A. PACKER
we have
θ · τ(cid:0)K0(cid:0)A S
β (cid:1)(cid:1)
α (cid:1)(cid:1) = τ(cid:0)K0(cid:0)A S
with the notations of Theorem (3.4). Thus in this case we do see the desired
relationship mentioned in Section 4: the range of the trace on the K0 groups of the
two C∗-algebras are related via multiplication by a positive constant.
We can now generalize our construction above as follows.
Definition 5.4. For any x ∈ Qp \ {0}, and any θ ∈ R \ {0}, there is an embedding
defined for all r1, r2 ∈ Zh 1
ιx,θ : Z(cid:20) 1
pi by
p(cid:21) × Z(cid:20) 1
p(cid:21) → [Qp × R]2
ιx,θ(r1, r2) = [(x · r1, θ · r1), (r2, r2)].
Then, we shall prove that for all α ∈ Ξp there exists x ∈ Qp \ {0} and θ ∈ R \ {0}
such that, by setting
Dx,θ = ιx,θ(cid:18)Z(cid:20) 1
p(cid:21) × Z(cid:20) 1
p(cid:21)(cid:19)
the twisted group C*-algebra C∗(D, η) is *-isomorphic to A S
α .
As a first step, we prove:
Lemma 5.5. Let p be prime, and let M = Qp × R. Let (x, θ) ∈ [Qp \ {0}]× [R\ {0}],
and define ιx,θ : Zh 1
pi × Zh 1
pi → [Qp × R]2 ∼= M ×cM by:
Then
ιx,θ(r1, r2) = [(x · r1, θ · r1), (r2, r2)] for all r1, r2 ∈ Z(cid:20) 1
p(cid:21).
Let η denote the Heisenberg cocycle defined on [M ×cM ]2 and let
p(cid:21) × Z(cid:20) 1
p(cid:21)(cid:19) .
θ(cid:19)(cid:21) : t1, t2 ∈ Z(cid:20) 1
p(cid:21)(cid:27) .
D = ιx,θ(cid:18)Z(cid:20) 1
x,θ =(cid:26)(cid:20)(t1, −t1),(cid:18)x−1t2, −
x,θ =(cid:26)[(q1, s1), (q2, s2)] : ∀r1, r2 ∈ Z(cid:20) 1
=(cid:26)[(q1, s1), (q2, s2)] : ∀r1, r2 ∈ Z(cid:20) 1
=(cid:26)[(q1, s1), (q2, s2)] : ∀r1, r2 ∈ Z(cid:20) 1
Proof. By definition,
D⊥
D⊥
t2
Now if r2 = 0, and r1 = pn, for any n ∈ Z, this implies
p(cid:21) ρ([ιx,θ(r1, r2)], [(q1, s1), (q2, s2)]) = 1(cid:27)
p(cid:21) ρ([(x · r1, θ · r1), (r2, r2)], [(q1, s1), (q2, s2)]) = 1(cid:27)
p(cid:21) e2πiθr1s2e2πi{x·r1q2}p e2πis1r2e2πi{q1r2}p = 1(cid:27) .
∀n ∈ Z e2πiθpns2 e2πi{x·pnq2}p = 1,
so that if we choose s2 = − t2
Likewise, if we take r1 = 0, and r2 = pn, for any n ∈ Z, we need (q1, s1) such that
pi ⊆ R, we need q2 = x−1t2.
θ for some t2 ∈ Zh 1
∀n ∈ Z e2πis1pn
e2πi{q1pn}p = 1.
NONCOMMUTATIVE SOLENOIDS AND THEIR PROJECTIVE MODULES
17
Again fixing q1 = t1 ∈ Zh 1
pi, this forces s1 = −t1. Thus
D⊥
x,θ =(cid:26)(cid:20)(t1, −t1),(cid:18)x−1t2, −
t2
θ(cid:19)(cid:21) : t1, t2 ∈ Z(cid:20) 1
p(cid:21)(cid:27) ,
as we desired to show.
(cid:3)
One thus sees that the two C∗-algebras C∗(Dx,θ, η) and C∗(D⊥
x,θ, η) are strongly
Morita equivalent (but not isomorphic, in general), and also the proof of this lemma
shows that C∗(D⊥
x,θ, η) is a noncommutative solenoid.
We can use Lemma (5.5) to prove the following Theorem:
Theorem 5.6. Let p be prime, and let α = (αi)i∈N ∈ Ξp, with α0 ∈ (0, 1). Then
there exists (x, θ) ∈ [Qp \{0}]×[R\{0}] with C∗(Dx,θ, η) isomorphic to the noncom-
mutative solenoid A S
of Rieffel produces an equivalence bimodule between A S
algebra B, and B is itself isomorphic to a noncommutative solenoid.
α , where Dx,θ = ιx,θ(cid:16)Zh 1
pi(cid:17). Moreover, the method
α and another unital C∗-
pi × Zh 1
Proof. By definition of Ξp, for all j ∈ N there exists bj ∈ {0, . . . , p − 1} such that
j=0 bjpj ∈
Zp ⊂ Qp. Let θ = α0, and now consider for this specific x and this specific θ the
C∗-algebra C∗(Dx,θ, η). By Definition (5.4), ιx,θ(r1, r2) = [(x · r1, θ · r1), (r2, r2)],
pαj+1 = αj + bj. We construct an element of the p-adic integers, x =P∞
for r1, r2 ∈ Zh 1
η (ιx,θ(r1, r2), ιx,θ(r3, r4)) = η ([(x · r1, θ · r1) , (r2, r2)] , [(x · r3, θ · r3) , (r4, r4)])
pi. Then
= e2πiθr1r4 e2πi{xr1r4}p , r1, r2, r3, r4 ∈ Z(cid:20) 1
p(cid:21),
and, setting ri = ji
pki , 1 ≤ i ≤ 4, and setting θ = α0, we obtain
η(cid:18)ιx,α0(cid:18) j1
pk1
,
j2
pk2(cid:19) , ιx,α0(cid:18) j3
for all
j1 j4
pk1 +k4 e
2πi{x j1 j4
pk1+k4
}p
We now note that the relation pαj+1 = αj + bj, bj ∈ {0, 1, · · · , p − 1} allows us to
prove inductively that
∀n ≥ 1 αn =
j1
pk1
,
,
,
,
j4
pk3
2πiα0
j4
pk4
j3
pk3
j2
pk2
pk4(cid:19)(cid:19) = e
p(cid:21).
∈ Z(cid:20) 1
α0 +Pn−1
pi × Zh 1
pk4(cid:19)(cid:19) = e2πi(α(k1 +k4)j1j4)
pk4 ∈ Zh 1
j=0 bjpj
pn
pk1 and j4
2πi α0j1 j4
= e
j4
.
By Theorem (2.4), the multiplier Ψα on Zh 1
Ψα(cid:18)(cid:18) j1
pk1
,
j2
pk2(cid:19) ,(cid:18) j3
pk3
,
pi is defined by:
A p-adic calculation now shows that for j1
Zp, we have {x j1j4
bjpj) ·
pk1 +k4 }p = (Pk1+k4−1
2πi{x j1 j4
j=0
pk1 +k4
e
}p = e2πi(Pk1 +k4 −1
j=0
bj pj j1j4)/pk1 +k2 .
pk1 +k4 e2πi(Pk1 +k4 −1
j=0
bj pj j1j4)/pk1 +k2 .
pi and x =P∞
j=0 bjpj ∈
j1j4
pk1 +k2 modulo Z, so that
18
FR ´ED ´ERIC LATR ´EMOLI `ERE AND JUDITH A. PACKER
We thus obtain
η(ιx,θ(r1, r2), ιx,θ(r3, r4)) = Ψα((r1, r2), (r3, r4))
for all r1, r2, r3, r4 ∈ Zh 1
pi,as desired.
To prove the final statement of the Theorem, we use Lemma 5.5. We have
α is isomorphic to C∗(Dx,θ, η), and the discussion prior to the statement of
x,θ, η) = B.
shown A S
Lemma 5.5 shows that C∗(Dx,θ, η) is strongly Morita equivalent to C∗(D⊥
But the proof of Lemma 5.5 gives that D⊥
C∗(D⊥
x,θ is isomorphic to Zh 1
x,θ, η) = B is a noncommutative solenoid, as we desired to show.
pi × Zh 1
pi, so that
(cid:3)
Remark 5.7. It remains an open question to give necessary and sufficient conditions
under which two noncommutative solenoids A S
β would be strongly Morita
equivalent, although it is evident that a necessary that the range of the trace on K0
of one of the C∗-algebras should be a constant multiple of the range of the trace
on the K0 group of the other. By changing the value of θ to be α0 + j, j ∈ Z,
and adjusting the value of x ∈ Qp accordingly, one can use the method of Theorem
α and A S
5.6 to construct a variety of embeddings ιx,θ of Zh 1
pi × Zh 1
pi into [Qp × R]2 that
provide lattices Dx,θ such that C∗(Dx,θ, η) and A S
α are ∗-isomorphic, but such
that the strongly Morita equivalent solenoids C∗(D⊥
x,θ, η) vary in structure. This
might lend some insight into classifying the noncommutative solenoids up to strong
Morita equivalence, as might a study between the relationship between the two
different methods of building equivalence modules described in Sections 4 and 5.
has used the Heisenberg equivalence bimodule construction of Rieffel to construct
different families of Gabor frames in modulation spaces of L2(Rn) for modulation
and translation by Zn ([11], [12]). It is of interest to see how far this analogy can
Remark 5.8. In the case where the lattice Z2n embeds into Rn × cRn, F. Luef
pi acting on
be taken when studying modulation and translation operators of Zh 1
L2(Qp × R), and we are working on this problem at present.
References
[1] L. Baggett and A. Kleppner, Multiplier representations of abelian groups, J. Functional Anal-
ysis 14 (1973), 299-324.
[2] A. Connes, C* -- alg`ebres et g´eom´etrie differentielle, C. R. de l'academie des Sciences de Paris
(1980), no. series A-B, 290.
[3] S. Echterhoff and J. Rosenberg, Fine structure of the Mackey machine for actions of abelian
groups with constant Mackey obstruction, Pacific J. Math. 170 (1995), 17-52.
[4] G. Elliott and D. Evans, Structure of the irrational rotation C ∗-algebras, Annals of Mathe-
matics 138 (1993), 477 -- 501.
[5] L. Fuchs, Infinite Abelian Groups, Volume I, Academic Press, New York and London, 1970.
[6] E. Hewitt and K. Ross, Abstract Harmonic Analysis, Volume II, Springer-Verlag Berlin,
1970.
[7] R. Hoegh-Krohn, M. B. Landstad, and E. Stormer, Compact ergodic groups of automor-
phisms, Annals of Mathematics 114 (1981), 75 -- 86.
[8] A. Kleppner, Multipliers on Abelian groups, Mathematishen Annalen 158 (1965), 11 -- 34.
[9] F. Latr´emoli`ere, Approximation of the quantum tori by finite quantum tori for the quantum
gromov-hausdorff distance, Journal of Funct. Anal. 223 (2005), 365 -- 395, math.OA/0310214.
[10] F. Latr´emoli`ere and J. Packer, Noncommutative solenoids, Submitted (2011), 30 pages,
ArXiv: 1110.6227.
NONCOMMUTATIVE SOLENOIDS AND THEIR PROJECTIVE MODULES
19
[11] F. Luef, Projective modules over noncommutative tori and multi-window Gabor frames for
modulation spaces, J. Funct. Anal. 257 (2009), 1921 -- 1946.
[12]
, Projections in noncommutative tori and Gabor frames, Proc. Amer. Math. Soc. 139
(2011), 571 -- 582.
[13] J. Packer and I. Raeburn, On the structure of twisted group C ∗-algebras, Trans. Amer. Math.
Soc. 334 (1992), no. 2, 685 -- 718.
[14] M. A. Rieffel, C*-algebras associated with irrational rotations, Pacific Journal of Mathematics
93 (1981), 415 -- 429.
[15]
[16]
[17]
[18]
[19]
, The cancellation theorem for the projective modules over irrational rotation C ∗-
algebras, Proc. London Math. Soc. 47 (1983), 285 -- 302.
, Projective modules over higher-dimensional non-commutative tori, Can. J. Math.
XL (1988), no. 2, 257 -- 338.
, Non-commutative tori -- a case study of non-commutative differentiable manifolds,
Contemporary Math 105 (1990), 191 -- 211.
, Metrics on states from actions of compact groups, Documenta Mathematica 3 (1998),
215 -- 229, math.OA/9807084.
, Gromov-Hausdorff distance for quantum metric spaces, Mem. Amer. Math. Soc. 168
(March 2004), no. 796, math.OA/0011063.
[20] G. Zeller-Meier, Produits crois´es d'une C*-alg`ebre par un groupe d' Automorphismes, J.
Math. pures et appl. 47 (1968), no. 2, 101 -- 239.
Department of Mathematics, University of Denver, 80208
E-mail address: [email protected]
Department of Mathematics, Campus Box 395, University of Colorado, Boulder, CO,
80309-0395
E-mail address: [email protected]
|
1511.01193 | 3 | 1511 | 2016-05-07T01:49:16 | A short note on Cuntz splice from a viewpoint of continuous orbit equivalence of topological Markov shifts | [
"math.OA"
] | Let $A$ be an $N\times N$ irreducible matrix with entries in $\{0,1\}$. We present an easy way to find an $(N+3)\times (N+3)$ irreducible matrix $\bar{A}$ with entries in $\{0,1\}$ such that their Cuntz--Krieger algebras ${\mathcal{O}}_A$ and ${\mathcal{O}}_{\bar{A}}$ are isomorphic and $ \det(1 -A) = - \det(1-\bar{A}). $ As a consequence, we know that two Cuntz--Krieger algebras ${\mathcal{O}}_A$ and ${\mathcal{O}}_B$ are isomorphic if and only if the one-sided topological Markov shift $(X_A, \sigma_A)$ is continuously orbit equivalent to $(X_B, \sigma_B)$ or $(X_{\bar{B}}, \sigma_{\bar{B}}).$ | math.OA | math | A short note on Cuntz splice from a viewpoint of continuous
orbit equivalence of topological Markov shifts
6
1
0
2
y
a
M
7
]
Kengo Matsumoto
Department of Mathematics
Joetsu University of Education
Joetsu, 943-8512, JAPAN
Abstract
Let A be an N ×N irreducible matrix with entries in {0, 1}. We present an easy way
to find an (N + 3) × (N + 3) irreducible matrix ¯A with entries in {0, 1} such that their
Cuntz -- Krieger algebras OA and O ¯A are isomorphic and det(1 − A) = −det(1 − ¯A). As
a consequence, we know that two Cuntz -- Krieger algebras OA and OB are isomorphic
if and only if the one-sided topological Markov shift (XA, σA) is continuously orbit
equivalent to either (XB, σB) or (X ¯B, σ ¯B).
For an N × N irreducible matrix A with entries in {0, 1}, let us denote by G(A) the
abelian group ZN /(1 − At)ZN and by uA the position of the class [(1, . . . , 1)] of the vector
(1, . . . , 1) in the group G(A). Throughout this short note, matrices are all assumed to be
irreducible and not any permutation matrices. J. Cuntz in [3] has shown that the pair
(K0(OA), [1]) of the K0-group K0(OA) of the Cuntz -- Krieger algebra OA and the class [1]
of the unit in K0(OA) is isomorphic to (G(A), uA). In [12], M. Rørdam has shown that
(G(A), uA) is a complete invariant of the isomorphism class of OA (see [6] for N ≤ 3). For
an N × N irreducible matrix A = [A(i, j)]N
i,j=1 with entries in {0, 1}, the (N + 2) × (N + 2)
irreducible matrix A− defined by
.
A
O
h
t
a
m
[
3
v
3
9
1
1
0
.
1
1
5
1
:
v
i
X
r
a
A− =
A(N − 1, 1)
A(1, 1)
...
A(N, 1)
0
0
. . .
A(1, N )
A(1, N − 1)
0 0
...
...
. . . A(N − 1, N − 1) A(N − 1, N ) 0 0
1 0
. . .
. . .
1 1
1 1
. . .
A(N, N − 1)
A(N, N )
...
0
0
...
1
0
is called the Cuntz splice for A, which has been first introduced in [4] by J. Cuntz, related
to classification problem for Cuntz -- Krieger algebras.
In [4], he had used the notation
A− instead of the above A−. The crucial property of the Cuntz splice is that G(A−) is
isomorphic to G(A) and det(1 − A−) = −det(1 − A). The Cuntz splice
1 1 0 0
1 1 1 0
0 1 1 1
0 0 1 1
1
1 1 ] is denoted by 2−.
for the matrix [ 1 1
In the proof of the above Rørdam's result [12,
Theorem 6.5], J. Cuntz's theorem [12, Theorem 7.2] is used which says that O2 ∼= O2−
implies OA ⊗ K ∼= OA− ⊗ K for all irreducible non-permutation matrices A. Since Rørdam
has proved O2 ∼= O2− ([12, Lemma 6.4]), the result OA ⊗ K ∼= OA− ⊗ K holds for all
irreducible non-permutation matrices A. By using this result, Rørdam has also obtained
that the group G(A) is a complete invariant of the stable isomorphism class of OA.
Let us denote by BF(A) the abelian group G(At) = ZN /(1 − A)ZN , which is called the
Bowen -- Franks group for N × N matrix A ([1]). Although BF(A) is isomorphic to G(A) as
a group, there is no canonical isomorphism between them. Related to classification theory
of symbolic dynamical systems, J. Franks has shown that the pair (BF(A), sgn(det(1−A)))
is a complete invariant of the flow equivalence class of the two-sided topological Markov
shift ( ¯XA, ¯σA) by using Bown -- Franks's result [1] for the group BF(A) and Parry -- Sullivan's
result [11] for the determinant det(1 − A). Combining this with the Rørdam's result for
the stable isomorphism classes of the Cuntz -- Krieger algebras, OA is stably isomorphic to
OB if and only if ( ¯XA, ¯σA) is flow equivalent to either ( ¯XB, ¯σB) or ( ¯XB−, ¯σB−).
In [9], the author has introduced a notion of continuous orbit equivalence in one-
sided topological Markov shifts to classify Cuntz -- Krieger algebras from a view point of
topological dynamical system. In [10], H. Matui and the author have shown that the triple
(G(A), uA, sgn(det(1 − A))) is a complete invariant of the continuous orbit equivalence
class of the right one-sided topological Markov shift (XA, σA). This result is rephrased
by using the above mentioned Rørdam's result for isomorphism classes of the Cuntz --
Krieger algebras such that the pair (OA, sgn(det(1 − A))) is a complete invariant of the
continuous orbit equivalence class of the one-sided topological Markov shift (XA, σA). The
C ∗-algebra OA− is not necessarily isomorphic to OA, whereas they are stably isomorphic,
because the position uA− in G(A−) generally is different from the position uA in G(A). We
note that the group G(A) determines the absolute value det(1 − A). If G(A) is infinite,
Ker(1 − A) is not trivial so that det(1 − A) = 0. If G(A) is finite, it forms a finite direct
sum Z/m1Z ⊕ · · · ⊕ Z/mrZ for some m1, . . . , mr ∈ N so that det(1 − A) = m1 · · · mr (cf.
[4], [5], [12]).
By [10, Lemma 3.7], we know that there is a matrix A′ with entries in {0, 1} such that
the triple (G(A), uA, sgn(det(1 − A))) is isomorphic to (G(A′), uA′, −sgn(det(1 − A′))),
which means that there exists an isomorphism Φ : G(A) → G(A′) such that Φ(uA) = uA′
and sgn(det(1 − A)) = −sgn(det(1 − A′)). Following the given proof of [10, Lemma 3.7],
the construction of the matrix A′ seems to be slightly complicated and the matrix size of
A′ becomes much bigger than that of A. It is not an easy task to present the matrix A′
for the given matrix A in a concrete way.
In this short note, we directly present an (N + 3) × (N + 3) matrix ¯A with entries in
{0, 1} such that (G(A), uA, sgn(det(1−A))) is isomorphic to (G( ¯A), u ¯A, −sgn(det(1− ¯A))).
The matrix ¯A is constructed such that if A is an irreducible non-permutation matrix, so
is ¯A.
2
A(1, 1)
...
A(N − 1, 1)
0
A(N, 1)
A(1, 1)
...
A(N − 1, 1)
0
A(N, 1)
0
0
...
. . .
A(1, N )
A(1, N − 1)
0
...
. . . A(N − 1, N − 1) A(N − 1, N ) 0
1
. . .
. . .
0
A(N, N − 1)
A(N, N )
0
0
...
...
...
. . .
A(1, N )
A(1, N − 1)
0 0 0
...
...
. . . A(N − 1, N − 1) A(N − 1, N ) 0 0 0
1 0 0
. . .
. . .
0 1 0
1 1 1
. . .
. . .
0 1 1
A(N, N − 1)
A(N, N )
0
0
0
0
0
0
...
.
(1)
We define
A◦ =
and
¯A = (A◦)− =
The operation A → A◦ is nothing but an expansion defined by Parry -- Sullivan in [11], and
preserves their determinant: det(1 − A) = det(1 − A◦). The following figure is a graphical
expression of the matrix ¯A from A.
vN
vN
vN +1
Figure 1:
vN +2
vN +3
We provide two lemmas. The first one is seen in [1]. The second one is seen in [4] and
[12] in a different form.
Lemma 1 ([1, Theorem 1.3]). The map
ηA : (x1, . . . , xN −1, xN , xN +1) ∈ ZN +1 → (x1, . . . , xN −1, xN + xN +1) ∈ ZN
induces an isomorphism ¯ηA from G(A◦) to G(A) such that ¯ηA([(1, . . . , 1, 0)]) = uA.
Lemma 2 (cf. [4, Proposition 2], [12, Proposition 7.1]). The map
ξA : (x1, . . . , xN ) ∈ ZN → (x1, . . . , xN , 0, 0) ∈ ZN +2
induces an isomorphism ¯ξA from G(A) to G(A−) such that ¯ξA([(1, . . . , 1, 0)]) = uA−.
3
Proof. For y = (y1, . . . , yN ) ∈ ZN , put
We then have
ξA(z) =
= (1 − At
−)
z1
...
zN
z =
z1
...
zN
0
0
x1
...
xN
0
0
= (1 − At)
y1
...
yN
.
.
y1
...
yN
0
−yN
= (1 − At
−)
z1
...
zN
zN +1
zN +2
−)ZN +2 so that ξA : ZN → ZN +2 induces a
Hence we have ξA((1 − At)ZN ) ⊂ (1 − At
homomorphism from G(A) to G(A−) denoted by ¯ξA. Suppose that [ξ(x1, . . . , xN )] = 0 in
G(A−) so that
for some (z1, . . . , zN +2) ∈ ZN +2. It then follows that zN +1 = 0, zN +2 = −zN so that
x1
...
xN
= (1 − At)
z1
...
zN
.
This implies [(x1, . . . , xN )] = 0 in G(A) and hence ¯ξA is injective.
For (x1, . . . , xN , xN +1, xN +2) ∈ ZN +2, we have
x1
...
xN
xN +1
xN +2
=
x1
...
xN −1
xN − xN +2
0
0
0
...
0
xN +2
xN +1
xN +2
x1
...
xN −1
0
0
+
=
xN − xN +2
+ (1 − At
−)
0
...
0
−zN +2
−zN +1
.
This implies that [(x1, . . . , xN , xN +1, xN +2)] = ¯ξA([(x1, . . . , xN −1, xN − xN −2)]) in G(A−).
Therefore ¯ξA : G(A) → G(A−) is surjective and hence an isomorphism. In particular, we
see that [(1, . . . , 1, 1, 1)] = ¯ξA([(1, . . . , 1, 0)]) in G(A−).
We have the following theorem by the preceding two lemmas.
Theorem 3. For an N × N matrix A with entries in {0, 1}, let ¯A be the (N + 3) × (N + 3)
matrix with entries in {0, 1} defined in (1). Then there exists an isomorphism Φ : G(A) →
G( ¯A) such that Φ(uA) = u ¯A and the matrices A, ¯A satisfy det(1 − A) = −det(1 − ¯A). If
A is an irreducible non-permutation matrix, so is ¯A.
4
Proof. Define Φ : G(A) → G( ¯A) by Φ = ¯ξA◦ ◦¯η−1
Since det(1 − ¯A) = −det(1 − A◦) = −det(1 − A), we see the desired assertion.
A so that Φ(uA) = ¯ξA◦([(1, . . . , 1, 0)]) = u ¯A.
Let P be an N × N permutation matrix coming from a permutation of the set
{1, 2, . . . , N }. Since there exists a natural isomorphism ΦP : G(A) −→ G(P AP −1) such
that ΦP (uA) = uP AP −1 and det(1−A) = det(1−P AP −1), the triplet (G(A), uA, det(1−A))
does not depend on the choice of the vertex vN in the directed graph of the matrix A.
We have some corollaries.
Corollary 4. Let A be an irreducible non-permutation matrix with entries in {0, 1}. Then
OA is isomorphic to O ¯A and det(1 − A) = −det(1 − ¯A).
Let ¯1 denote the matrix
0 1 0 0
1 0 1 0
0 1 1 1
0 0 1 1
which is the matrix ¯A for the 1 × 1 matrix A = [1]. By the above theorem, we have
Corollary 5. (K0(O¯1), u¯1) = (Z, 1).
Hence the simple purely infinite C ∗-algebra O¯1 has the same K-theory as the C ∗-
algebra O1 = C(S1) of the continuous functions on the unit circle S1 with the positions
of their units, whereas (K0(O1−), u1− ) = (Z, 0) for the matrix 1− =h 1 1 0
0 1 1i by [6] (cf. [4,
p. 150]).
1 1 1
The following corollary has been shown in [10]. Its proof is now easy by using [12].
Corollary 6 ([10, Lemma 3.7]). Let F be a finitely generated abelian group and u an
element of F . Let s = 0 when F is infinite and s = −1 or 1 when F is finite. Then there
exists an irreducible non-permutation matrix A such that
(F, u, s) = (G(A), uA, sgn(det(1 − A)).
Proof. By [12, Proposition 6.7 (i)], we know that there exists an irreducible non-permutation
matrix A such that (F, u) = (G(A), uA). If s = sgn(det(1−A)), the matrix A is the desired
one, otherwise ¯A is the desired one.
Let A and B be two irreducible non-permutation matrices with entries in {0, 1}. The
one-sided topological Markov shifts (XA, σA) and (XB, σB) are said to be flip continuously
orbit equivalent if (XA, σA) is continuously orbit equivalent to either (XB, σB) or (X ¯B, σ ¯B).
Similarly two-sided topological Markov shifts ( ¯XA, ¯σA) and ( ¯XB, ¯σB) are said to be flip
flow equivalent if ( ¯XA, ¯σA) is flow equivalent to either ( ¯XB, ¯σB), or ( ¯X ¯B, ¯σ ¯B). We thus
have the following corollaries.
Corollary 7. Let A, B be irreducible and not any permutation matrices with entries in
{0, 1}.
(i) OA is isomorphic to OB if and only if the one-sided topological Markov shifts (XA, σA)
and (XB, σB) are flip continuously orbit equivalent.
5
(ii) OA is stably isomorphic to OB if and only if the two-sided topological Markov shifts
( ¯XA, ¯σA) and ( ¯XB, ¯σB) are flip flow equivalent.
Let us denote by [OA] the isomorphism class of the Cuntz -- Krieger algebra OA as a
C ∗-algebra. Since (G(A), uA) is isomorphic to (G( ¯A), u ¯A), we have [OA] = [O ¯A]. We
regard the sign sgn(det(1 − A)) of det(1 − A) as the orientation of the class [OA]. Then
we can say that the pair ([OA], sgn(det(1 − A))) is a complete invariant of the continuous
orbit equivalence class of the one-sided topological Markov shift (XA, σA).
In the rest of this short note, we present another square matrix A of size N + 3
i,j=1 of size N such that OA is isomorphic to O A and
from a square matrix A = [A(i, j)]N
det(1 − A) = −det(1 − A). Define (N + 3) × (N + 3) matrix A by setting
A(N − 1, 1)
A =
A(1, 1)
...
0
A(N, 1)
0
0
...
. . .
A(1, N )
A(1, N − 1)
0 0 0
...
...
. . . A(N − 1, N − 1) A(N − 1, N ) 0 0 0
1 0 0
. . .
0 1 0
. . .
. . .
1 0 1
0 1 1
. . .
A(N, N − 1)
A(N, N )
0
0
0
...
0
0
0
...
.
(2)
The difference between the previous matrix ¯A in (1) and the above matrix A is the only
Its graphical expression of the matrix A from A is the
((N + 2), (N + 2))-component.
following figure.
vN
vN
vN +1
Figure 2:
vN +2
vN +3
By virtue of [6], we know the following proposition.
Proposition 8. The Cuntz -- Krieger algebras O ¯A and O A are isomorphic, and det(1− ¯A) =
det(1 − A).
Proof. Let us denote by ¯Ai the ith row vector of the matrix ¯A of size N + 3. We put Ei
i
the row vector of size N + 3 such that Ei = (0, . . . , 0,
1, 0, . . . , 0) where the ith component
is one, and the other components are zero. Then we have ¯AN +2 = EN +1 + ¯AN +3. Since
6
the (N + 2)th row AN +2 of A is AN +2 = EN +1 + EN +3, and the other rows of A are the
same as those of ¯A, the matrix A is obtained from ¯A by the primitive transfer
¯A
=⇒
EN +1+ ¯AN +3→ AN +2
A
in the sense of [6, Definition 3.5]. We obtain that O ¯A is isomorphic to O A by [6, Theorem
3.7], and det(1 − ¯A) = det(1 − A) by [6, Theorem 8.4].
Before ending this short note, we refer to differences among the three matrices A−, ¯A, A
from a view point of dynamical system. As (G(A−), det(1 − A−)) = (G( ¯A), det(1 − ¯A)) =
(G( A), det(1 − A)), there is a possibility that their two sided topological Markov shifts
( ¯XA−, ¯σA−), ( ¯X ¯A, ¯σ ¯A), ( ¯X A, ¯σ A) are topologically conjugate. We however know that they
are not topologically conjugate to each other in general by the following example. Denote
by pn(¯σA) the cardinal number of the n-periodic points {x ∈ ¯XA ¯σn
A(x) = x} of the
topological Markov shift ( ¯XA, ¯σA). The zeta function ζA(z) for ( ¯XA, ¯σA) is defined by
ζA(z) = exp ∞
Xn=1
pn(¯σA)
n
zn!
(cf.[8]).
It is well-known that the formula ζA(z) =
the matrices A−, ¯A, A for [ 1 1
1 1 ] respectively. It is direct to see that
1
det(1−zA) holds ([2]). Let us denote by 2−, ¯2, 2
ζ2−(z) =
1
1 − 4z + 3z2 + 2z3 − z4 ,
ζ¯2(z) =
1
1 − 3z + 4z3 − z4 ,
ζ2(z) =
1
1 − 3z + z2 + z3 + z4 .
The zeta function is invariant under topological conjugacy so that ( ¯X2−, ¯σ2−), ( ¯X¯2, ¯σ¯2), ( ¯X2, ¯σ2)
are not topologically conjugate to each other.
This paper is a revised version of the paper entitled "Continuous orbit equivalence of
topological Markov shifts and Cuntz splice" arXiv:1511.01193v2 [math.OA].
Acknowledgment. This work was supported by JSPS KAKENHI Grant Number 15K04896.
References
[1] R. Bowen and J. Franks, Homology for zero-dimensional nonwandering sets, Ann.
Math. 106(1977), pp. 73 -- 92.
[2] R. Bowen and O. E. Lanford III, Zeta functions of the shift transformation,
Trans. Amer. Math. Soc. 112(1964), pp. 55 -- 66.
[3] J. Cuntz, A class of C ∗-algebras and topological Markov chains II: reducible chains
and the Ext- functor for C ∗-algebras, Invent. Math. 63(1980), pp. 25 -- 40.
[4] J. Cuntz, The classification problem for the C ∗-algebra OA, Geometric methods
in operator algebras, Pitman Research Notes in Mathematics Series 123(1986), pp.
145 -- 151.
7
[5] J. Cuntz and W. Krieger, A class of C ∗-algebras and topological Markov chains,
Invent. Math. 56(1980), pp. 251 -- 268.
[6] M. Enomoto, M. Fujii and Y. Watatani, K0-groups and classifications of Cuntz --
Krieger algebras, Math. Japon. 26(1981), pp. 443 -- 460.
[7] J. Franks, Flow equivalence of subshifts of finite type, Ergodic Theory Dynam.
Systems 4(1984), pp. 53 -- 66.
[8] D. Lind and B. Marcus, An introduction to symbolic dynamics and coding,
Cambridge University Press, Cambridge, 1995.
[9] K. Matsumoto, Orbit equivalence of topological Markov shifts and Cuntz -- Krieger
algebras, Pacific J. Math. 246(2010), pp. 199 -- 225.
[10] K. Matsumoto and H. Matui, Continuous orbit equivalence of topological Markov
shifts and Cuntz -- Krieger algebras, Kyoto J. Math. 54(2014), pp. 863 -- 878.
[11] W. Parry and D. Sullivan, A topological invariant for flows on one-dimensional
spaces, Topology 14(1975), pp. 297 -- 299.
[12] M. Rørdam, Classification of Cuntz -- Krieger algebras, K-theory 9(1995), pp. 31 -- 58.
8
|
1210.4533 | 2 | 1210 | 2013-03-04T17:21:05 | The Cuntz semigroup and stability of close C*-algebras | [
"math.OA"
] | We prove that separable C*-algebras which are completely close in a natural uniform sense have isomorphic Cuntz semigroups, continuing a line of research developed by Kadison - Kastler, Christensen, and Khoshkam. This result has several applications: we are able to prove that the property of stability is preserved by close C*-algebras provided that one algebra has stable rank one; close C*-algebras must have affinely homeomorphic spaces of lower-semicontinuous quasitraces; strict comparison is preserved by sufficient closeness of C*-algebras. We also examine C*-algebras which have a positive answer to Kadison's Similarity Problem, as these algebras are completely close whenever they are close. A sample consequence is that sufficiently close C*-algebras have isomorphic Cuntz semigroups when one algebra absorbs the Jiang-Su algebra tensorially. | math.OA | math |
THE CUNTZ SEMIGROUP AND STABILITY OF CLOSE
C ∗-ALGEBRAS
FRANCESC PERERA, ANDREW TOMS, STUART WHITE,
AND WILHELM WINTER
Abstract. We prove that separable C ∗-algebras which are completely
close in a natural uniform sense have isomorphic Cuntz semigroups,
continuing a line of research developed by Kadison-Kastler, Christensen,
and Khoshkam. This result has several applications: we are able to
prove that the property of stability is preserved by close C ∗-algebras
provided that one algebra has stable rank one; close C ∗-algebras must
have affinely homeomorphic spaces of lower-semicontinuous quasitraces;
strict comparison is preserved by sufficient closeness of C ∗-algebras. We
also examine C ∗-algebras which have a positive answer to Kadison's
Similarity Problem, as these algebras are completely close whenever they
are close. A sample consequence is that sufficiently close C ∗-algebras
have isomorphic Cuntz semigroups when one algebra absorbs the Jiang-
Su algebra tensorially.
1. introduction
In 1972 Kadison and Kastler introduced a metric d on the C ∗-subalgebras
of a given C ∗-algebra by equipping the unit balls of the subalgebras with
the Hausdorff metric (in norm) ([27]). They conjectured that sufficiently
close C ∗-subalgebras of B(H) should be isomorphic, and this conjecture was
recently established by Christensen, Sinclair, Smith and the last two named
authors ([17]) when one C ∗-algebra is separable and nuclear. The one-sided
version of this result -- that a sufficiently close near inclusion of a nuclear
separable C ∗-algebra into another C ∗-algebra gives rise to a true inclusion --
was later proved by Hirshberg, Kirchberg, and the third named author ([23]).
These results and others (see [16], [10]) have given new momentum to the
perturbation theory of operator algebras.
The foundational paper [27] was concerned with structural properties of
close algebras, showing that the type decomposition of a von Neumann al-
gebra transfers to nearby algebras. We continue this theme here asking
Date: October 9, 2018.
Research partially supported by EPSRC (grants No. EP/G014019/1 and No.
EP/I019227/1), by the DFG (SFB 878), by NSF (DMS-0969246), by the DGI MICIIN
(grant No. MTM2011-28992-C02-01), and by the Comissionat per Universitats i Recerca
de la Generalitat de Catalunya. Andrew Toms is partially supported by the 2011 AMS
Centennial Fellowship.
1
2
F. PERERA, A. TOMS, S. WHITE, AND W. WINTER
"Which properties or invariants of C ∗-algebras are preserved by small per-
turbations?" With the proof of the Kadison-Kastler conjecture the answer
for nuclear separable C ∗-algebras is, "All of them." Here we consider gen-
eral separable C ∗-algebras where already, there are some results. Sufficiently
close C ∗-algebras have isomorphic lattices of ideals ([34]) and algebras whose
stabilizations are sufficiently close have isomorphic K-theories ([29]). This
was extended to the Elliott invariant consisting of K-theory, traces, and
their natural pairing, in [16]. A natural next step is to consider the Cuntz
semigroup of (equivalence classes of) positive elements (in the stabilisation)
of a C ∗-algebra, due both to its exceptional sensitivity in determining non-
isomorphism ([42]), classification results using the semigroup ([39]) and the
host of C ∗-algebraic properties that can be formulated as order-theoretic
properties of the semigroup: for example there is strong evidence to suggest
that the behaviour of the Cuntz semigroup characterises important algebraic
regularity properties of simple separable nuclear C ∗-algebras ([31, 44, 45]).
We prove that algebras whose stabilizations are sufficiently close do indeed
have isomorphic Cuntz semigroups, a surprising fact given the sensitivity of
a Cuntz class to perturbations of its representing positive element. This is
in stark contrast with the case of Murray-von Neumann equivalence classes
of projections, where classes are stable under perturbations of the represent-
ing projection of size strictly less than one. The bridge between these two
situations is that we can arrange for the representing positive element of a
Cuntz class to be almost a projection in trace. We exploit this fact through
the introduction of what we call very rapidly increasing sequences of positive
contractions, increasing sequences where each element almost acts as a unit
on its predecessor.
The Kadison-Kastler metric d is equivalent to a complete version dcb
(given by applying d to the stabilisations) if and only if Kadison's Similar-
ity Problem has a positive solution [16, 9]; the latter is known to hold in
considerable generality, for instance in the case of Z-stable algebras ([25]).
We show how this result, and a number of other similarity results for C ∗-
algebras, can be put in a common framework using Christensen's property
Dk ([12]), and, building on [16], make a more careful study of automatic
complete closeness and its relation to property Dk. We prove that if an
algebra A has Dk for some k, then d(A ⊗ K, B ⊗ K) ≤ C(k)d(A, B), where
C(k) is a constant independent of A and B; as a consequence sufficiently
close C ∗-algebras have isomorphic Cuntz semigroups provided one algebra
is Z-stable.
Stability is perhaps the most basic property one could study in pertur-
bation theory, yet proving its permanence under small perturbations has
seen very little progress. We take a significant step here by proving that
stability is indeed preserved provided that one of the algebras considered
has stable rank one. The proof is an application of our permanence re-
sult for the Cuntz semigroup. Another application is our proof that stably
THE CUNTZ SEMIGROUP AND STABILITY OF CLOSE C ∗-ALGEBRAS
3
close C ∗-algebras have affinely homeomorphic spaces of lower semicontin-
uous 2-quasitraces. This extends and improves an earlier results from [16]
showing that the affine isomorphism between the trace spaces of stably close
C ∗-algebras obtained in [16] is weak∗-weak∗ continuous.
The paper is organized as follows: Section 2 contains the preliminaries on
the Cuntz semigroup and the Kadison-Kastler metric; Section 3 establishes
the permanence of the Cuntz semigroup under complete closeness; Section 4
discusses property Dk and proves our permanence result for stability; Section
5 proves permanence for quasitraces.
2. Preliminaries
Throughout the paper we write A+ for the positive elements of a C ∗-
1 for the positive contractions in
algebra A, A1 for the unit ball of A and A+
A.
In the next two subsections we review the definition and basic properties
of the Cuntz semigroup. A complete account can be found in the survey [4].
Sn Mn(A)+.
2.1. The Cuntz semigroup. Let A be a C ∗-algebra. Let us consider on
(A ⊗ K)+ the relation a - b if vnbv∗
n → a for some sequence (vn) in A ⊗ K.
Let us write a ∼ b if a - b and b - a.
In this case we say that a is
Cuntz equivalent to b. Let Cu(A) denote the set (A ⊗ K)+/ ∼ of Cuntz
equivalence classes. We use hai to denote the class of a in Cu(A). It is clear
that hai ≤ hbi ⇔ a - b defines an order on Cu(A). We also endow Cu(A)
with an addition operation by setting hai + hbi := ha′ + b′i, where a′ and b′
are orthogonal and Cuntz equivalent to a and b respectively (the choice of a′
and b′ does not affect the Cuntz class of their sum). The semigroup W (A)
is then the subsemigroup of Cu(A) of Cuntz classes with a representative in
Alternatively, Cu(A) can be defined to consist of equivalence classes of
countably generated Hilbert modules over A [18]. The equivalence relation
boils down to isomorphism in the case that A has stable rank one, but is
rather more complicated in general and as we do not require the precise
definition of this relation in the sequel, we omit it. We note, however, that
the identification of these two approaches to Cu(A) is achieved by associating
the element hai to the class of the Hilbert module aℓ2(A).
2.2. The category Cu. The semigroup Cu(A) is an object in a category of
ordered Abelian monoids denoted by Cu introduced in [18] with additional
properties. Before stating them, we require the notion of order-theoretic
compact containment. Let T be a pre-ordered set with x, y ∈ T . We say
that x is compactly contained in y -- denoted by x ≪ y -- if for any increasing
sequence (yn) in T with supremum y, we have x ≤ yn0 for some n0 ∈ N. An
object S of Cu enjoys the following properties (see [18, 4]), which we use
repeatedly in the sequel. In particular the existence of suprema in property
4
F. PERERA, A. TOMS, S. WHITE, AND W. WINTER
P3 is a crucial in our construction of a map between the Cuntz semigroups
of stably close C ∗-algebras.
P1 S contains a zero element;
P2 the order on S is compatible with addition: x1 + x2 ≤ y1 + y2
whenever xi ≤ yi, i ∈ {1, 2};
P3 every countable upward directed set in S has a supremum;
P4 for each x ∈ S, the set x≪ = {y ∈ S y ≪ x} is upward directed
with respect to both ≤ and ≪, and contains a sequence (xn) such
that xn ≪ xn+1 for every n ∈ N and supn xn = x;
P5 the operation of passing to the supremum of a countable upward
directed set and the relation ≪ are compatible with addition: if S1
and S2 are countable upward directed sets in S, then S1 + S2 is
upward directed and sup(S1 + S2) = sup S1 + sup S2, and if xi ≪ yi
for i ∈ {1, 2}, then x1 + x2 ≪ y1 + y2 .
We say that a sequence (xn) in S ∈ Cu is rapidly increasing if xn ≪ xn+1
for all n. We take the scale Σ(Cu(A)) to be the subset of Cu(A) obtained
as supremums of increasing sequences from A+.
For objects S and T from Cu, the map φ : S → T is a morphism in the
category Cu if
M1 φ is order preserving;
M2 φ is additive and maps 0 to 0;
M3 φ preserves the suprema of increasing sequences;
M4 φ preserves the relation ≪.
2.3. The Kadison-Kastler metric. Let us recall the definition of the met-
ric on the collection of all C ∗-subalgebras of a C ∗-algebra introduced in [27].
Definition 2.1. Let A, B be C ∗-subalgebras of a C ∗-algebra C. Define a
metric d on all such pairs as follows: d(A, B) < γ if and only if for each x in
the unit ball of A or B, there is y in the unit ball of the other algebra such
that kx − yk < γ.
In this definition, we typically take C = B(H) for a Hilbert space H.
The complete, or stabilised version, of the Kadison-Kastler metric is defined
by dcb(A, B) = d(A ⊗ K, B ⊗ K) inside C ⊗ K (here K is the compact
operators on ℓ2(N)); the notion dcb is used for this metric as dcb(A, B) ≤ γ
is equivalent to the condition that d(Mn(A), Mn(B)) ≤ γ for every n.
We repeatedly use the standard fact that if d(A, B) < γ, then given a
positive contraction a ∈ A+
1 , there exists a positive contraction b ∈ B+
1 with
ka − bk < 2α. One way of seeing this is to use the hypothesis d(A, B) < γ
to approximate a1/2 by some c ∈ B1 with ka1/2 − ck < γ. Then take b = cc∗
so that
ka − bk ≤ ka1/2(a1/2 − c)k + k(a1/2 − c∗)ck < 2γ.
THE CUNTZ SEMIGROUP AND STABILITY OF CLOSE C ∗-ALGEBRAS
5
There is also a one-sided version of closeness introduced by Christensen
in [13], which is referred to as a γ-near inclusion:
Definition 2.2. Let A, B be C ∗-subalgebras of a C ∗-algebra C and let
γ > 0. Write A ⊆γ B if for every x in the unit ball of B, there is y ∈ B
such that kx− yk ≤ γ (note that y need not be in the unit ball of B). Write
A ⊂γ B if there exists γ′ < γ with A ⊆γ ′ B. As with the Kadison-Kastler
metric, we also use complete, or stabilised, near inclusions: write A ⊆cb,γ B
when A ⊗ Mn ⊆γ B ⊗ Mn for all n, and A ⊂cb,γ B when there exists γ′ < γ
with A ⊆cb,γ B.
3. Very rapidly increasing sequences and the Cuntz semigroup
We start by noting that, for close C ∗-algebras of real rank zero, an isomor-
phism between their Cuntz semigroups can be deduced from existing results
in the literature. For a C ∗-algebra A, let V (A) be the Murray and von
Neumann semigroup of equivalence classes of projections in S∞
n=1 A ⊗ Mn
and write Σ(V (A)) = {[p] ∈ V (A) p = p2 = p∗ ∈ A}. This is a local
semigroup in the sense that if p, q, p′ and q′ are projections in A with
p′q′ = 0 and p ∼ p′, q ∼ q′, then [p] + [q] = [p′ + q′] ∈ Σ(V (A)). Recall
that, if A has real rank zero, then the work of Zhang [46] shows that V (A)
has the Riesz refinement property. By definition, this means that when-
ever x1, . . . , xn, y1, . . . , ym ∈ V (A) satisfy Pi xi = Pj yj, then there exist
zi,j ∈ V (A) with Pj zi,j = xi and Pi zi,j = yj for each i, j. The case
m = n = 2 of this can be found as [3, Lemma 2.3], and the same proof
works in general.
The Cuntz semigroup of a C ∗-algebra of real rank zero is completely
determined by its semigroup of projections (see [33] when A additionally
has stable rank one and [1] for the general case). We briefly recall how this
is done. An interval in V (A) is a non-empty, order hereditary and upward
directed subset I of V (A), which is said to be countably generated provided
there is an increasing sequence (xn) in V (A) such that I = {x ∈ V (A)
x ≤ xn for some n}. The set of countably generated intervals is denoted
by Λσ(V (A)), and it has a natural semigroup structure. Namely, if I and
J have generating sequences (xn) and (yn) respectively, then I + J is the
interval generated by (xn + yn). Given a positive element a in A ⊗ K in a
σ-unital C ∗-algebra of real rank zero A, put I(a) = {[p] ∈ V (A) p - a}.
The correspondence [a] 7→ I(a) defines an ordered semigroup isomorphism
Cu(A) ∼= Λσ(V (A)).
Theorem 3.1. Let A and B be σ-unital C ∗-subalgebras of a C ∗-algebra C,
with d(A, B) < 1/8. If A has real rank zero, then B also has real rank zero
and Cu(A) ∼= Cu(B).
Proof. That B has real rank zero follows from [16, Theorem 6.3]. We know
from [35, Theorem 2.6] that there is an isomorphism of local semigroups
Φ1 : Σ(V (A)) → Σ(V (B)) (with inverse, say, Ψ1). This is defined as Φ1[p] =
6
F. PERERA, A. TOMS, S. WHITE, AND W. WINTER
[q], where q is a projection in B such that kp− qk < 1/8. Given p ∈ Mn(A),
by [46, Theorem 3.2] we can find projections {pi}i=1,...,n in A such that
[p] = Pi[pi]. Now extend Φ1 to Φ : V (A) → V (B) by Φ([p]) = Pi Φ1([pi]).
Let us check that Φ is well defined. If [p] = Pi[pi] = Pj[qj] for projections
pi and qj in A, then use refinement to find elements aij ∈ V (A) such that
[pi] = Pj aij and [qj] = Pi aij. We may also clearly choose projections
zij, z′
ij], and such that zij ⊥ zik if j 6= k, and
ij ⊥ z′
z′
ij ∈ A such that aij = [zij] = [z′
lj if i 6= l. Then:
X Φ1([pi]) = X
= X
X
X
Φ1([zij])
Φ1([z′
j
i
i
j
ij]) = X
j
X
i
Φ1([z′
ij]) = X
j
Φ1([qj]) .
It is clear that Φ is additive and that ΦΣ(V (A)) = Φ1. Using Ψ1, we construct
an additive map Ψ : V (B) → V (A), with ΨΣ(V (B)) = Ψ1. Since Ψ1 ◦ Φ1 =
idΣ(V (A)), it follows that Ψ ◦ Φ = idV (A). Similarly Φ ◦ Ψ = idV (B).
Now, since Cu(A) ∼= Λσ(V (A)) and Cu(B) ∼= Λσ(V (B)), it follows that
Cu(A) is isomorphic to Cu(B).
(cid:3)
We turn now to very rapidly increasing sequences. These provide the key
tool we use to transfer information between close algebras at the level of the
Cuntz semigroup.
n=1 in A+
Definition 3.2. Let A be a C ∗-algebra. We say that a rapidly increasing
sequence (an)∞
1 is very rapidly increasing if given ε > 0 and n ∈ N,
there exists m0 ∈ N such that for m ≥ m0, there exists v ∈ A1 with
k(vamv∗)an − ank < ε. Say that a very rapidly increasing sequence (an)∞
in (A ⊗ K)1
+ represents x ∈ Cu(A) if supnhani = x.
n=1
(3.1)
The following two functions are used in the sequel to manipulate very
rapidly increasing sequences. Given a ∈ A+ and ε > 0, write (a − ε)+
for hε(a), where hε is the continuous function hε(t) = max(0, t − ε). For
0 ≤ β < γ, let gβ,γ be the piecewise linear function on R given by
0,
t−β
γ−β , β < t < γ;
1,
With this notation, the standard example of a very rapidly increasing se-
n=1 for a ∈ A+
quence is given by (g2−(n+1),2−n(a))∞
1 . This sequence represents
hai. In this way every element of the Cuntz semigroup of A is represented by
a very rapidly increasing sequence from (A ⊗ K)+
1 . In the next few lemmas
we develop properties of very rapidly increasing sequences, starting with a
technical observation.
t ≤ β;
t ≥ γ.
gβ,γ(t) =
THE CUNTZ SEMIGROUP AND STABILITY OF CLOSE C ∗-ALGEBRAS
7
Lemma 3.3. Let A be a C ∗-algebra and let a, b ∈ A+
1 and v ∈ A1 satisfy
kv∗bva − ak ≤ δ for some δ > 0. Suppose that 0 < β < 1 and γ ≥ 0 satisfy
γ + δβ−1 < 1, then h(a − β)+i ≤ h(b − γ)+i in Cu(A).
Proof. Let p ∈ A∗∗ denote the spectral projection of a for the interval [β, 1].
When p = 0, then (a − β)+ = 0 and the result is trivial, so we may assume
that p 6= 0. Then ap is invertible in pA∗∗p with inverse x satisfying kxk ≤
β−1. Compressing (v∗bva− a) by p and multiplying by x, we have kpv∗bvp−
pk ≤ δβ−1. Thus
kpv∗(b − γ)+vp − pk ≤ k(b − γ)+ − bk + kpv∗bvp − pk ≤ γ + δβ−1,
and so
pv∗(b − γ)+vp ≥ (cid:0)1 − (γ + δβ−1)(cid:1) p.
As p acts as a unit on (a − β)+, we have
(a − β)+ = (a − β)1/2
+
+ p(a − β)1/2
≤ (cid:0)1 − (γ + δβ−1)(cid:1)−1
= (cid:0)1 − (γ + δβ−1)(cid:1)−1
Thus (a − β)+ - (b − γ)+.
(a − β)1/2
(a − β)1/2
+ pv∗(b − γ)+vp(a − β)1/2
+ v∗(b − γ)+v(a − β)1/2
+ .
+
(cid:3)
The next lemma encapsulates the fact that the element of the Cuntz
semigroup represented by a very rapidly increasing sequence (an)∞
n=1 of con-
tractions depends only on the behaviour of parts of the an with spectrum
near 1.
n=1 be a very rapidly increasing sequence in A+
Lemma 3.4. Let (an)∞
1 . Then
for each λ < 1, the sequence (h(an − λ)+i)∞
n=1 has the property that for each
n ∈ N, there is m0 ∈ N such that for m ≥ m0, we have h(an − λ)+i ≪
h(am − λ)+i. Furthermore
sup
(3.2)
n h(an − λ)+i = sup
n hani.
Proof. Fix n ∈ N and 0 < ε < λ and take 0 < δ small enough that λ+ε−1δ <
1. As (an)∞
n=1 is very rapidly increasing, there exists m0 such that for
m ≥ m0, there exists v ∈ A1 with k(v∗amv)an − ank < δ. Lemma 3.3 gives
h(an − ε)+i ≤ h(am − λ)+i,
so that h(an− λ)+i ≪ ham− λ)+i as ε < λ. This shows that (h(ar − λ)+i)∞
is upward directed and that
r=1
h(an − δ)+i ≤ sup
r h(ar − λ)+i,
for all n and all ε > 0, from which (3.2) follows.
(cid:3)
We can modify elements sufficiently far down a very rapidly increasing
sequences with contractions so that they almost act as units for positive
contractions dominated in the Cuntz semigroup by the sequence.
8
F. PERERA, A. TOMS, S. WHITE, AND W. WINTER
Lemma 3.5. Let A be a C ∗-algebra.
(1) Suppose that a, b ∈ A+
(2) Let (an)∞
1 satisfy a - b. Then for all ε > 0, there exists
n=1 be a very rapidly increasing sequence in A+
v ∈ A with kv∗bva − ak ≤ ε and kv∗bvk ≤ 1.
1 and suppose a ∈
A+
1 satisfies hai ≪ suphani. Then, for every ε > 0, there exists m0 ∈ N
such that for m ≥ m0, there exists v ∈ A1 with k(v∗amv)a − ak < ε.
Proof. (1). Fix ε > 0 and find r > 0 so that ka1+r − ak ≤ ε/2. Now ar - b,
so there exists w ∈ A with kar − w∗bwk ≤ ε/4. Thus kw∗bwk ≤ 1 + ε/4, and
so, writing v = (1+ε/4)−1/2w, we have kv∗bvk ≤ 1 and kw∗bw−v∗bvk ≤ ε/4.
As such kar − v∗bvk ≤ ε/2 and so
kv∗bva − ak ≤ kv∗bv − arkkak + ka1+r − ak ≤ ε/2 + ε/2 = ε,
as claimed.
(2). As hai ≪ suphani, there exists some m1 ∈ N with a - am1 ∼ a2
m1.
Fix ε > 0 and by part (1), find w ∈ A with k(w∗a2
m1w)a − ak < ε/2 and
kw∗a2
n=1 is very rapidly
increasing, find some m0 > m1 such that for m ≥ m0 there exists t ∈ A1
with k(t∗amt)am1 − am1k ≤ ε′. Given such m and t, we have
m1wk ≤ 1. Now set ε′ = ε/(2kwk) and, as (an)∞
k(w∗am1t∗amtam1w)a − ak ≤ kw∗am1kk(t∗amt)am1 − am1kkwkkak
+ k(w∗a2
≤ kwkε′ + ε/2 = ε,
m1w)a − ak
as kw∗am1k ≤ 1. As such we can take v = tam1w ∈ A1.
(cid:3)
It follows immediately from part (2) above, that two very rapidly increas-
ing sequences representing the same element of the Cuntz semigroup can be
intertwined to a single very rapidly increasing sequence.
n)∞
n=1, (a′
Proposition 3.6. Let (an)∞
n=1 be very rapidly increasing sequences
in a C ∗-algebra A representing the same element x ∈ Cu(A). Then these
sequences can be intertwined after telescoping to form a very rapidly increas-
ing sequence which also represents x, i.e. there exists m1 < m2 < ··· and
n1 < n2 < ··· such that (am1 , a′
n2,··· ) is a very rapidly increasing
sequence.
n1, am2, a′
Given a rapidly increasing sequence in A+
1 , we can use the functions gβ,γ
from (3.1) to push the spectrum of the elements of the sequence out to 1
and extract a very rapidly rapidly increasing sequence representing the same
element of the Cuntz semigroup.
Lemma 3.7. Let A be a C ∗-algebra and (an)∞
sequence in A+
sequence (g2−(mn +1),2−mn (an))∞
1 . Then there exists a sequence (mn)∞
n=1 is very rapidly increasing and
n=1 be a rapidly increasing
n=1 in N such that the
n hg2−(mn +1),2−mn (an)i = sup
sup
n hani.
THE CUNTZ SEMIGROUP AND STABILITY OF CLOSE C ∗-ALGEBRAS
9
In particular, every element of the scale Σ(Cu(A)) can be expressed as a
very rapidly increasing sequence of elements from A+
1 .
Proof. We will construct the mn so that an−1 - g2−(mn +1),2−mn (an) and for
each 1 ≤ r < n, there exists v ∈ A1 with
k(v∗g2−(mn +1),2−mn (an)v)g2−(mr +1),2−mr (ar) − g2−(mr +1),2−mr (ar)k < 2−n.
Fix n ∈ N and suppose m1,··· , mn−1 have been constructed with these
properties. As (g2−(m+1),2−m(an))∞
m=1 is a very rapidly increasing sequence
representing hani and han−1i ≪ hani, there exists fmn such that han−1i ≪
h(g2−(m+1),2−m(an))i for m ≥ fmn. Further, for 1 ≤ r < n,
hg2−(mr +1),2−mr (ar)i ≪ hari ≪ sup
m h(g2−(m+1),2−m(an)i
and so the required mn can be found using Part (2) of Lemma 3.5.
The resulting sequence (g2−(mn +1),2−mn (an))∞
n=1 is very rapidly increasing
by construction. As an−1 - g2−(mn +1),2−mn (an) - an for all n, we have
supnhg2−(mn +1),2−mn (an)i = supnhani.
(cid:3)
We now consider the situation where we have two close C ∗-algebras acting
on the same Hilbert space. The following lemma ensures that we can produce
a well defined map between the Cuntz semigroups.
Lemma 3.8. Let A, B be C ∗-algebras acting on the same Hilbert space and
suppose that a ∈ A+
1 satisfy ka − bk < 2α for some α < 1/27.
Suppose that (an)∞
1 with hai ≪
suphani. Then, there exists n0 ∈ N with the property that for n ≥ n0 and
bn ∈ B+
1 and b ∈ B+
n=1 is a very rapidly increasing sequence in A+
1 with kbn − ank < 2α, we have
h(b − 18α)+i ≪ h(bn − γ)+i ≪ h(bn − 18α)+i
in Cu(B), for all γ with 18α < γ < 2/3.
Proof. Fix γ with 2/3 > γ > 18α (which is possible as α < 1/27). By taking
ε = 2α−ka− bk in Lemma 3.5 (2), there exists n0 ∈ N such that for n ≥ n0,
there exists v ∈ A1 with
k(v∗anv)a − ak < 2α − ka − bk.
Fix such an n ≥ n0 and v ∈ A1, and take bn ∈ B+
choose some w ∈ B1 with kw − vk < α. We have
1 with kan − bnk < 2α and
kw∗bnw − v∗anvk ≤ 2kw − vk + kbn − ank < 4α
so that
k(w∗bnw)b − bk ≤ k((w∗bnw) − 1)(b − a)k
+ k(w∗bnw − v∗anv)ak
+ k(v∗anv)a − ak
≤ kb − ak + 4α + k(v∗anv)a − ak
≤ 6α.
10
F. PERERA, A. TOMS, S. WHITE, AND W. WINTER
Taking δ = 6α, β = 18α and 2/3 > γ′ > γ > 18α so that γ′ + δβ−1 < 1,
Lemma 3.3 gives
h(b − 18α)+i ≤ h(bn − γ′)+i ≪ h(bn − γ)+i ≪ h(bn − 18α)+i,
as claimed.
(cid:3)
Proposition 3.9. Let A, B be C ∗-algebras acting on the same Hilbert space
with the property that there exists α < 1/27 such that for each a ∈ A1 there
exists b ∈ B1 with ka−bk < α. Then there is a well defined, order preserving
map Φ : Σ(Cu(A)) → Σ(Cu(B)) given by
Φ(suphani) = suph(bn − 18α)+i,
n=1 is a very rapidly increasing sequence in A+
1 and bn ∈ B+
whenever (an)∞
have kan − bnk < 2α for all n ∈ N. Moreover, if d(A, B) < α for α < 1/42,
then Φ is a bijection with inverse Ψ : Σ(Cu(B)) → Σ(Cu(A)) obtained from
interchanging the roles of A and B in the definition of Φ.
1
n=1
n=1 and (b′
n)∞
Proof. Suppose first that α < 1/27. To see that Φ is well defined, we apply
Lemma 3.8 repeatedly. Firstly, given a very rapidly increasing sequence
n=1 in A+
(an)∞
1 representing an element x ∈ Σ(Cu(A)) and a sequence (bn)∞
in B+
1 with kan − bnk < 2α for all n, Lemma 3.8 shows that the sequence
(h(bn − 18α)+i)∞
n=1 is upward directed. Indeed, for each m, take a = am
and b = bm in Lemma 3.8 so that h(bm − 18α)+i ≪ h(bn − 18α)+i for all
sufficiently large n. As such supnh(bn − 18α)+i exists in Σ(Cu(B)).
Secondly, this supremum does not depend on the choice of (bn)∞
n=1. Con-
sider two sequences (bn)∞
n=1 satisfying kbn − ank < 2α and
kb′
n − ank < 2α for all n. For each n, Lemma 3.8 shows that there exists m0
such that for m ≥ m0, we have
h(bn − 18α)+i ≪ h(b′
m − 18α)+i,
Thus supnh(bn − 18α)+i = supnh(b′
Thirdly, given two very rapidly increasing sequences (a′
n)∞
n=1
in A+
n=1 and (bn)∞
n)∞
n=1 in
B+
nk,kbn − ank < 2α for all n, Lemma 3.8 gives supnh(b′
1 with kb′
n −
n=1 and (an)∞
18α)+i ≤ supnh(bn − 18α)+i.
n=1
represent the same element of Σ(Cu(A)), this shows that the map Φ given
in the proposition is well defined.
In general, this third observation also
shows that Φ is order preserving.
ni ≤ supnhani, and sequences (b′
n − 18α)+i ≪ h(bm − 18α)+i.
and h(b′
n − 18α)+i.
1 with supnha′
In particular, when (a′
n=1 and (an)∞
n − a′
n)∞
Now suppose that d(A, B) < α < 1/42 and let Ψ : Σ(Cu(B)) → Σ(Cu(A))
be the order preserving map obtained by interchanging the roles of A and
B above. Take x ∈ Σ(Cu(A)) and fix a very rapidly increasing sequence
(an)∞
1 with kan−bnk <
2α for all n. For each n, Lemma 3.8 gives m > n with
1 representing x. Fix a sequence (bn)∞
n=1 in B+
n=1 in A+
h(bn − 18α)+i ≪ h(bm − γ)+i ≪ h(bm − 18α)+i,
for any γ with 18α < γ < 2/3. Passing to a subsequence if necessary, we
can assume this holds for m = n + 1 and hence ((bn − 18α)+)∞
n=1 is a rapidly
THE CUNTZ SEMIGROUP AND STABILITY OF CLOSE C ∗-ALGEBRAS
11
n=1 in B+
ni = sup
n=1 in A+
increasing sequence. By Lemma 3.7, there exists a sequence (mn)∞
so that, defining b′
increasing sequence (b′
n=1 in N
n = g2−(mn+1),2−mn ((bn − 18α)+), we have a very rapidly
1 with
n h(bn − 18α)+i = Φ(x).
1 with kcn − b′
n)∞
n hb′
sup
Choose a sequence (cn)∞
nk < 2α for each n so that
the definition of Ψ gives Ψ(Φ(x)) = suph(cn − 18α)+i. We now show that
x ≤ Ψ(Φ(x)) ≤ x.
Fix 0 < β < 1 with α(18 + 24β−1) < 1. This choice can be made as
α < 1/42. Fix n ∈ N. As h(bn − 18α)+i ≪ suprhb′
ri, Lemma 3.5 (2) provides
m0 ∈ N such that for m ≥ m0, there exists w ∈ B1 with
(3.3)
Take v ∈ A1 with kv − wk < α. Then
(3.4)
Combining the estimates (3.3), (3.4) and noting that kw∗b′
w is a contraction, gives
mw)(bn − 18α)+ − (bn − 18α)+k < 2α − kan − bnk.
mk + 2kv − wk < 4α.
k(v∗cmv − w∗b′
mw)k ≤ kcm − b′
k(w∗b′
mw − 1k ≤ 1 as
k(v∗cmv)an − ank ≤ k((v∗cmv) − 1)(an − bn)k
mw)bnk
+ k(v∗cmv − w∗b′
+ k(w∗b′
+ k(w∗b′
< kan − bnk + 4α + 18α + (2α − kan − bnk) = 24α.
mw − 1)(bn − (bn − 18α)+)k
mw)(bn − 18α)+ − (bn − 18α)+k
Taking γ = 18α, δ = 24α, Lemma 3.3 gives
h(an − β)+i ≤ h(cm − 18α)+i ≤ Ψ(Φ(x)).
k playing the role of a, (b′
As n was arbitrary, supnh(an − β)+i ≤ Ψ(Φ(x)). As β < 1, Lemma 3.4 gives
supnh(an − β)+i = supnhani = x so that x ≤ Ψ(Φ(x)).
For the reverse inequality, fix k ∈ N and apply Lemma 3.8 (with the
roles of A and B reversed, b′
n=1 the role of
(an)) and γ = 1/2, so 18α < γ < 2/3) to find some n ∈ N such that
h(ck − 18α)+i ≤ h(cn − 1/2)+i. Now, just as in the proof of Lemma 3.8,
there is z ∈ B1 with k(z∗bn+1z)bn − bnk ≤ 6α. Let p ∈ B∗∗ be the spectral
projection of bn for [18α, 1], so that, just as in the proof of Lemma 3.3,
kz∗bn+1zp − pk ≤ 1/3. Fix y ∈ A1 with ky − zk ≤ α. Since p is a unit for
(bn − 18α)+, it is a unit for b′
n = g2−(mn+1),2−mn ((bn − 18α)+), giving the
estimate
n)∞
ky∗an+1ycn − cnk ≤ ky∗an+1ycn − z∗bn+1zcnk
+ k(z∗bn+1z − 1)(cn − b′
n)k
+ k(z∗bn+1z)b′
≤ 4α + 2α + 1/3 = 6α + 1/3.
n − b′
nk
12
F. PERERA, A. TOMS, S. WHITE, AND W. WINTER
Take δ = 6α + 1/3, β = 1/2 and γ = 0, so that γ + β−1δ = 2/3 + 12α < 1.
Thus Lemma 3.3, gives
and hence
h(cn − 1/2)+i ≤ han+1i,
h(ck − 18α)+i ≤ han+1i ≤ x.
(cid:3)
n=1 is a sequence in (B ⊗ K)+
n=1 is a very rapidly increasing sequence in (A ⊗ K)+
1 with kan − bnk < 2α for all n ∈ N.
Taking the supremum over k gives Ψ(Φ(x)) ≤ x.
Theorem 3.10. Let A and B be C ∗-algebras acting on the same Hilbert
space with dcb(A, B) < α < 1/42. Then (Cu(A), Σ(Cu(A))) is isomor-
phic to (Cu(B), Σ(Cu(B))). Moreover, an order preserving isomorphism
Φ : Cu(A) → Cu(B) can be defined by Φ(suphani) = suph(bn − 18α)+i,
whenever (an)∞
1 and
(bn)∞
Proof. We have d(A⊗ K, B⊗ K) < α < 1/42. By definition Σ(Cu(A⊗ K)) =
Cu(A) and Σ(Cu(A ⊗ K)) = Cu(B). By applying Proposition 3.9 to A ⊗ K
and B⊗K, we obtain mutually inverse order preserving bijections Φ : Cu(A⊗
K) → Cu(B ⊗ K) and Ψ : Cu(B ⊗ K) → Cu(A⊗ K), given by Φ(suphani) =
suph(bn − 18α)+i, whenever (an)∞
n=1 is a very rapidly increasing sequence in
(A⊗ K)+
1 with kan − bnk < 2α for all
n=1 in A+
n ∈ N. Given a very rapidly increasing sequence (an)∞
1 representing
an element x ∈ Σ(Cu(A)), we can find a sequence (bn)∞
1 with
kan − bnk < 2α, so that Φ(x) = suph(bn − 18α)+i ∈ Σ(Cu(B)). Since Φ
and Φ−1 are order preserving bijections, they also preserve the relation ≪
of compact containment and suprema of countable upward directed sets,
as these notions are determined by the order relation ≤. Further, taking
an = bn = 0 for all n, shows that Φ(0Cu(A)) = 0Cu(B). Finally, note that
Φ preserves addition: given very rapidly increasing sequences (an)∞
n=1 and
n=1 in (A ⊗ K)+
1 representing x and y in Cu(A), the sequence (an ⊕ a′
n)∞
(a′
n)
is very rapidly increasing in M2(A ⊗ K) ∼= A ⊗ K. If (bn)∞
n=1 have
kan − bnk,ka′
n − b′
nk < 2α for all n, then
n=1 is a sequence in (B ⊗ K)+
1 and (bn)∞
n=1 in B+
n=1, (b′
n)∞
k(an ⊕ a′
n) − (bn ⊕ b′
n)k < 2α,
and has
((bn ⊕ b′
n) − 18α)+ = (bn − 18α)+ ⊕ (b′
n − 18α)+.
In this way we see that Φ(x + y) = Φ(x) + Φ(y).
(cid:3)
In particular properties of a C ∗-algebra which are determined by its Cuntz
semigroup transfer to completely close C ∗-algebras. One of the most notable
of these properties is that of strict comparison.
Corollary 3.11. Let A and B be C ∗-algebras acting on the same Hilbert
space with dcb(A, B) < 1/42 and suppose that A has strict comparison. Then
so too does B.
THE CUNTZ SEMIGROUP AND STABILITY OF CLOSE C ∗-ALGEBRAS
13
4. Z-stability and automatic complete closeness
Given a C ∗-algebra A ⊂ B(H), [9] shows that the metrics d(A,·) and
dcb(A,·) are equivalent if and only if A has a positive answer to Kadison's
similarity problem from [26]. The most useful reformulation of the similar-
ity property for working with close C ∗-algebras is due to Christensen and
Kirchberg. Combining [29] and [14, Theorem 3.1], it follows that a C ∗-
algebra A has a positive answer to the similarity problem if and only if A
has Christensen's property Dk from [13] for some k.
Definition 4.1. Given an operator T ∈ B(H), we write ad(T ) for the
derivation ad(T )(x) = xT − T x. A C ∗-algebra A has property Dk for some
k > 0 if, for every non-degenerate representation π : A → B(H), the in-
equality
d(T, π(A)′) ≤ kkad(T )π(A)k
(4.1)
holds for all T ∈ B(H). A von Neumann algebra A is said to have the
property D∗
k if the inequality (4.1) holds for all unital normal representations
π on H and all T ∈ B(H).
By taking weak∗-limit points, it follows that if A is a weak∗-dense C ∗-
subalgebra of a von Neumann algebra M and A has property Dk, then M
has property D∗
k.
That property Dk converts near containments to completely bounded
near containments originates in [13, Theorem 3.1]. The version we give
below improves on the bounds γ′ = 6kγ from [13] and γ′ = (1 + γ)2k − 1
from [16, Corollary 2.12].
Proposition 4.2. Suppose that A has property Dk for some k > 0. Then for
γ > 0, every near inclusion A ⊆γ B (or A ⊂γ B) with A and B acting non-
degenerately on the same Hilbert space, gives rise to a completely bounded
near inclusion A ⊆cb,γ ′ B (or A ⊂cb,γ ′), where γ′ = 2kγ.
Proof. Suppose A ⊆γ B is a near inclusion of C ∗-algebras acting non-
degenerately on H and fix n ∈ N. Let C = C ∗(A, B) and let π : C → B(K)
be the universal representation of C. Then π(A)′′ has property D∗
k so that
π(A)′′ ⊆cb,2kγ π(B)′′ by [8, Proposition 2.2.4]. By definition, for n ∈ N we
have π(A)′′ ⊗ Mn ⊆2kγ π(B)′′ ⊗ Mn. As π is the universal representation of
C the Hahn-Banach argument used to deduce [13, equation (3)] from [13,
equation (2)] gives A ⊗ Mn ⊆2kγ B ⊗ Mn, as required. The result when we
work with strict near inclusions A ⊂γ B follows immediately.
(cid:3)
C ∗-algebras with no bounded traces (such as stable algebras) where shown
to have the similarity property in [20]. Using the property Dk version of this
fact, the previous proposition gives automatic complete closeness when one
algebra has no bounded traces. The argument below which transfers the
absence of bounded traces to a nearby C ∗-algebra essentially goes back to
[27, Lemma 9]. We use more recent results in order to get better estimates.
14
F. PERERA, A. TOMS, S. WHITE, AND W. WINTER
Corollary 4.3. Suppose that A and B are C ∗-algebras which act non-
degenerately on the same Hilbert space and satisfy d(A, B) < γ for γ <
(2 + 6√2)−1. Suppose that A has no bounded traces (for example if A is sta-
ble). Then B has no bounded traces, and therefore A ⊂cb,3γ B, B ⊂cb,3γ A
and dcb(A, B) < 6γ.
Proof. Suppose d(A, B) < (2 + 6√2)−1 and τ : B → C is a bounded trace.
Let π : B → B(H) be the GNS-representation of B corresponding to τ . Then
there is a larger Hilbert space H and a representation π : C ∗(A, B) → B( H)
such that π is a direct summand of πB. That is, the projection p from
H onto H is central in π(B) and π(b) = pπ(b)p for all b ∈ B. Then, by
[27, Lemma 5], we have d(π(A)′′, π(B)′′) ≤ d(A, B), and hence there is a
projection q ∈ π(A)′′ with kp − qk ≤ γ/√2 by [28, Lemma 1.10(ii)]. If q is
an infinite projection in π(A)′′, then as d(A, B) < (2+6√2)−1, one can follow
the argument of [16, Lemma 6.1] (using the estimate kp−qk < γ/√2 in place
of kp − qk < 2γ) to see that p is infinite in π(B)′′, giving a contradiction. If
q is finite, then qπ(A)′′q has a finite trace ρ and ρ ◦ πA defines a bounded
trace on A, and again we have a contradiction. Thus B has no bounded
traces.
Theorem 2.4 of [12] shows that a properly infinite von Neumann algebra
has property D∗
3/2. As such, every C ∗-algebra with no bounded traces has
property D3/2. Since A and B both have property D3/2, Proposition 4.2
gives A ⊂cb,3γ B and B ⊂cb,3γ A, whence dcb(A, B) < 6γ.
(cid:3)
Corollary 4.4. Suppose that A and B are C ∗-algebras which act non-
degenerately on the same Hilbert space and satisfy d(A, B) < 1/252 and
suppose that A has no bounded traces (for example when A is stable). Then
(Cu(A), Σ(Cu(A)) ∼= (Cu(B), Σ(Cu(B)).
Proof. Combine Corollary 4.3 with Theorem 3.10 (noting that 6d(A, B) <
1/42).
(cid:3)
We can use the Cuntz semigroup to show that stability transfers to close
C ∗-algebras provided one algebra has stable rank one. To detect stability
for a σ-unital C ∗-algebra we use the following criterion from [32, Lemma
5.4] which reformulates the earlier characterisation from [24].
Lemma 4.5. Let A be a σ-unital C ∗-algebra and let c ∈ A be a strictly
positive element. Then, A is stable if and only if for every ǫ > 0, there is
b ∈ A+ such that (c − ǫ)+ ⊥ b and (c − ǫ)+ - b.
Following [40], we say that a C ∗-algebra A has weak cancellation provided
Cu(A) satisfies the property that x + z ≪ y + z implies x ≤ y. It was proved
in [41, Theorem 4.3] that if A has stable rank one, then W (A) has the
property defining weak cancelation. When A has stable rank one, so too
does A ⊗ K [38, Theorem 3.6], and so A has weak cancelation.
THE CUNTZ SEMIGROUP AND STABILITY OF CLOSE C ∗-ALGEBRAS
15
Lemma 4.6. Let A be a σ-unital C ∗-algebra with weak cancelation. Then
A is stable if and only if Cu(A) = Σ(Cu(A)).
Proof. If A is stable, then Σ(Cu(A)) = Cu(A). Indeed, given n ∈ N, choose
an automorphism θn : K ⊗ K → K ⊗ K with θn(K ⊗ e11) = K ⊗ Mn and let
ψ : A → A ⊗ K be an isomorphism. Then (ψ−1 ⊗ idK)(idA ⊗ θn)(ψ ⊗ idK)
is an automorphism of A ⊗ K which maps A ⊗ e11 onto A ⊗ Mn. In this
way the class of a positive element in A ⊗ Mn lies in the scale Σ(Cu(A)).
For x ∈ (A ⊗ K)+ and ε > 0, we have (x − ε)+ ∈ S∞
n=1(A ⊗ Mn), and
hence h(x − ε)+i ∈ Σ(Cu(A)). Since the scale is defined to be closed under
suprema, it follows that Cu(A) = Σ(Cu(A)).
Conversely, let c ∈ A be a strictly positive element so that Σ(Cu(A)) =
{x ∈ Cu(A) : x ≤ hci} and let ǫ > 0 be given. The hypothesis ensures that
2hci ≤ hci, and so we can find δ > 0 such that 2h(c − ǫ
4 )+i ≪ h(c − δ)+i.
Now write
(c − δ)+ = (c − δ)+gǫ/2,ǫ(c) + (c − δ)+(1M (A) − gǫ/2,ǫ(c)),
and observe that
h(c − δ)+gǫ/2,ǫ(c)i ≤ hgǫ/2,ǫ(c)i = h(c − ǫ
2 )+i ≪ h(c − ǫ
4 )+i.
We now have that
2h(c − ǫ
4 )+i ≪ h(c − δ)+i ≤ h(c − ǫ
2 )+i + h(c − δ)+(1M (A) − gǫ/2,ǫ(c))i
and so weak cancellation enables us to conclude that
h(c − ǫ
4 )+i ≤ h(c − δ)+(1 − gǫ/2,ǫ(c))i.
Let b = (c − δ)+(1 − gǫ/2,ǫ(c)).
(c − ǫ)+ ≤ (c − ǫ
A is stable.
It is clear that b ⊥ (c − ǫ)+ and that
4 )+ - b. Thus we may invoke Lemma 4.5 to conclude that
(cid:3)
Theorem 4.7. Let A and B be σ-unital C ∗-algebras with A stable and
d(A, B) < 1/252 and suppose either A or B has stable rank one. Then B is
stable.
Proof. By Corollary 4.4, we have an isomorphism
(Cu(A), Σ(Cu(A)) ∼= (Cu(B), Σ(Cu(B)).
Since A is stable Cu(A) = Σ(Cu(A)). Our isomorphism condition now tells
us that Cu(B) = Σ(Cu(B)).
If B has stable rank one, then it has weak
cancelation, whereas if A has stable rank one, A has weak cancelation and,
as weak cancelation is a property of the Cuntz semigroup, so too does B.
The result now follows from Lemma 4.6.
(cid:3)
We now turn to the situation in which one C ∗-algebra is Z-stable. In [12],
Christensen shows that McDuff II1 factors have property D5/2, and hence
via the estimates of [36], have similarity length at most 5. (In fact McDuff
factors, and more generally II1 factors with Murray and von Neumann's
property Γ have length 3 [15], but at present we do not know how to use
16
F. PERERA, A. TOMS, S. WHITE, AND W. WINTER
this fact to obtain better estimates for automatic complete closeness of close
factors with property Γ.)
In [37, 30, 25] analogous results have been established in a C ∗-setting:
in particular Z-stable C ∗-algebras ([25]) and C ∗-algebras of the form A ⊗
B, where B is nuclear and has arbitrarily large unital matrix subalgebras
([37]) have similarity degree (and hence length) at most 5. Here we show
how to use the original von Neumann techniques from [12] to show that
a class of algebras generalising both these examples have property D5/2
(recapturing the upper bound 5 on the length). A similar result has been
obtained independently by Hadwin and Li [22, Corollary 1] working in terms
of the similarity degree as opposed to property Dk. Once we have this Dk
estimate, Proposition 4.2 applies. In particular we obtain uniform estimates
on the cb-distance dcb(A, B) in terms of d(A, B) when A is Z-stable.
Given a von Neumann algebra M ⊂ B(H) and x ∈ B(H), write cow
M(x)
for the weak∗ closed convex hull of {uxu∗ : u ∈ U(M)}. If M is injective,
M(x) ∩ M′ is non-empty for all x ∈ B(H).
then by Schwartz's property P, cow
Note that for a non-degenerately represented C ∗-algebra A ⊂ B(H), we have
kad(T )Ak = kad(T )A′′k. We say that an inclusion A ⊂ C of C ∗-algebras is
non-degenerate if the inclusion map is non-degenerate.
Proposition 4.8. Let C be a C ∗-algebra and A, B ⊂ C be commuting non-
degenerate C ∗-subalgebras which generate C. Suppose B is nuclear and has
no non-zero finite dimensional representations. Then C has property D5/2,
and hence similarity length at most 5.
Proof. Suppose C is non-degenerately represented on H and fix x ∈ B(H).
The non-degeneracy assumption ensures that A and B are non-degenerately
represented on H. Note that C ′′ has no finite type I part as B has no non-
zero finite dimensional representations. Let p be the central projection in
C ′′ so that C ′′p is type II1 and C ′′(1 − p) is properly infinite. Fix a unital
type I∞ subalgebra M0 ⊂ (1 − p)C ′′(1 − p) and let M = (M0 ∪ pB)′′ which
M(x)∩ (M∪{p})′.
is injective. By Schwartz's property P, there exists y ∈ cow
As in [12, Theorems 2.3, 2.4], ky − xk ≤ kad(x)C ′′k and kad(y)C ′′k ≤
kad(x)C ′′k. Write y1 = yp and y2 = y(1 − p). If p 6= 1, then the properly
2}′ and so by [12, Corollary 2.2],
infinite algebra M0 lies in C ′′(1− p)∩{y2, y∗
kad(y2)C ′′(1−p)k = 2d(y2, C ′(1 − p)). Take x2 ∈ C ′′(1 − p) with kx2 − y2k =
kad(y2)C ′′(1−p)k/2 ≤ kad(x)C ′′k/2.
If p 6= 0, then we argue exactly as in the proof of [12, Proposition 2.8]
to produce first z1 ∈ A′p with ky1 − z1k ≤ kad(y1)C ′′pk/2 ≤ kad(x)C ′′k/2.
Continuing with the proof of [12, Proposition 2.8], as B′′p and A′′p commute,
B ′′p(z1) ∩ B′p
B ′′p(z1) is contained in A′p and hence there exists x1 ∈ cow
cow
with
kx1 − z1k ≤ kad(z1)B ′′pk ≤ kad(z1 − y1)B ′′pk ≤ 2kz1 − y1k ≤ kad(x)C ′′k.
Then
ky1 − x1k ≤ ky1 − z1k + kz1 − x1k ≤ 3kad(x)C ′′k/2.
THE CUNTZ SEMIGROUP AND STABILITY OF CLOSE C ∗-ALGEBRAS
17
If p = 0, take x1 = 0 and the same inequality holds. The element x1+x2 ∈ C ′
has
kx − (x1 + x2)k
≤kx − yk + k(y1 − x1) + (y2 − x2)k
≤kad(x)C ′′k + max(ky1 − x2k,ky2 − x2k) ≤ 5kad(x)C ′′k/2.
Therefore C has property D5/2, and so by [36, Remark 4.7] has length at
most 5.
(cid:3)
Corollary 4.9. Let A be a Z-stable C ∗-algebra. Then A has property D5/2
and length at most 5.
The main result of [16] is that the similarity property transfers to close
C ∗-algebras. This work is carried out with estimates depending on the
length and length constant of A, but it is equally possible to carry out
this work entirely in terms of property Dk so it can be applied to Z-stable
algebras. Our objective is to obtain a version of [16, Corollary 4.6] replacing
the hypothesis that A has length at most ℓ and length constant at most K
with the formally weaker hypothesis that A has property Dk (if A has the
specified length and length constants, then it has property Dk for k = Kℓ/2,
conversely if A has property Dk, then it has length at most ⌊2k⌋, but a
length constant estimate is not known in this case, see [36, Remark 4.7]).
This enables us to use Corollary 4.9 obtain an isomorphism between the
Cuntz semigroups of sufficiently close C ∗-algebras when one algebra is Z-
stable. To achieve a Dk version of [16, Section 4], we adjust the hypotheses
in Lemma 4.1, Theorem 4.2 and Theorem 4.4 of [16] in turn, starting with
Lemma 4.1. We begin by isolating a technical observation.
Lemma 4.10. Let M be a finite von Neumann algebra with a faithful tracial
state acting in standard form on H and let J be the conjugate linear modular
conjugation operator inducing an isometric antisomorphism x 7→ JxJ of M
onto M′ ∼= Mop. Suppose that S is another von Neumann algebra acting
k, then M′ ⊂cb,2kγ
nondegenerately on H with M′ ⊂γ S. If M has property D∗
S.
Proof. As J is isometric, M ⊂γ J SJ, so that M ⊂cb,2kγ J SJ by Proposition
4.2. Now, for each n ∈ N, let Jn denote the isometric conjugate linear
operator of component wise complex conjugation on Cn so that J ⊗ Jn is a
conjugate linear isometry on H ⊗ Cn. We can conjugate the near inclusion
M ⊗ Mn ⊂2kγ JSJ ⊗ Mn by J ⊗ Jn to obtain M′ ⊗ Mn ⊂2kγ S ⊗ Mn, as
required.
(cid:3)
The next lemma is the modification of [16, Lemma 4.1]. The expression
for β below is a slight improvement over that of the original.
Lemma 4.11. Let M and N be von Neumann algebras of type II1 faithfully
and non-degnerately represented on H with common centre Z which admits
18
F. PERERA, A. TOMS, S. WHITE, AND W. WINTER
a faithful state. Suppose d(M, N) = α and M has property D∗
k. If α satisfies
24(12√2k + 4k + 1)α < 1/200,
then d(M′, N′) < 2β + 1200kα(1 + β) where β = 96kα(600k + 1).
Proof. This amounts to showing that the hypothesis in [16, Lemma 4.1] that
M contains an weak∗ dense C ∗-algebra A of length at most ℓ and length
constant at most K can be replaced by the statement that M has property
D∗
k (and that the specified expressions on β are valid). The hypothesis that
M has such a weak∗ dense C ∗-algebra is initially used to see that M has
property Dk at the beginning of the lemma and then applied to a unital
normal representation to obtain [16, equation (4.5)]. As such property D∗
k
suffices for this estimate.
The other use of this hypothesis comes on page 385 in the last paragraph
of the lemma, to obtain [16, equation (4.28)]. Using the notation of this
paragraph, the von Neumann algebra TM is a cutdown of M acting as M⊗IG
on H ⊗ G by the projection ei0,i0 from the commutant of M on this space.
Since ei0,i0 is unitarily equivalent in this commutant to a projection of the
form e ⊗ g0, where e is a projection from the commutant of M on H of full
central support and g0 is a minimal projection in B(G), it follows that ei0,i0
has full central support in the commutant of M on H ⊗ G. As such TM is
isomorphic to M, so has property D∗
k. Thus Lemma 4.10 can be applied to
M ⊂48(600kα+α) T ′
the near inclusion T ′
N2
from [16, equation (4.25)] giving
M ⊂cb,96k(600kα+α)) T ′
T ′
N2 .
It then follows that
TM ⊗ B(ℓ2(Λ)) ⊂96k(600kα+α) TN ⊗ B(ℓ2(Λ)),
which is precisely [16, equation (4.28)] with our new estimate for β replacing
that of the original. We then deduce that d(M′, N′) ≤ 2β + 1200kα(1 + β) in
just the same way that [16, equation (4.30)] is obtained from [16, equation
(4.28)].
(cid:3)
Now we adjust Theorem 4.2 of [16]. The resulting constant β is obtained
by taking α = 11γ in the previous lemma. Note that there is an unfortunate
omission in the value of β in Theorem 4.2 of [16] which should be given by
taking α = 11γ in Lemma 4.1 of [16], so should be K((1 + 316800kγ +
528γ)ℓ − 1): this has no knock on consequences to Theorem 4.4 of [16]
where the correct value of β is used.
Lemma 4.12. Let A and B be C ∗-algebras acting on a Hilbert space and
suppose that d(A, B) = γ. Suppose A has property Dk and 24(12√2k + 4k +
1)γ < 1/2200. Then
d(A′, B′) ≤ 10γ + 2β + 13200kγ(1 + β),
where β = 1056k(600kγ + γ).
THE CUNTZ SEMIGROUP AND STABILITY OF CLOSE C ∗-ALGEBRAS
19
Proof. This amounts to replacing the hypothesis that A has length at most
ℓ and length constant at most K with the condition that A has property
Dk in Theorem 4.2 of [16]. The length hypothesis on A is used to show
that certain II1 von Neumann closures of A satisfy [16, Lemma 4.1], but
since the weak∗-closure of a C ∗-algebra with property Dk has property D∗
k,
Lemma 4.11 can be used in place of [16, Lemma 4.1]. Note that in the
proof of [16, Theorem 4.2] the reference to injective von Neumann algebras
n=1 Mn
has the similarity property). The correct statement is that these algebras
have property D∗
(cid:3)
having property D1 is incorrect (it is an open question whether Q∞
1 which is all that is used.
Finally we can convert Theorem 4.4 of [16]. Note the typo in the statement
1−2η−kγ .
of this theorem, the definition of k should be
The same change should be made in Corollary 4.6 of [16].
k
k
1−2η−2kγ rather than
Proposition 4.13. Let A and B be C ∗-subalgebras of some C ∗-algebra
C with d(A, B) < γ and suppose that A has property Dk. Write β =
1056(600kγ + γ) and η = 10γ + 2β + 13200kγ(1 + β) and suppose that
(4.2)
24(12√2k + 4k + 1)γ <
,
2η + 2kγ < 1.
Then dcb(A, B) ≤ 4kγ, where
k =
1
2200
k
1 − 2η − 2kγ
.
.
Proof. We check that B has property Dk. This amounts to weakening the
hypothesis of [16, Theorem 4.4] in just the same way as the preceeding
lemmas. Applying Lemma 4.12 in place of Theorem 4.2 of [16] in the proof
of Theorem 4.4 of [16], shows that under the hypotheses of this proposition
B has property Dk, where
k =
k
1 − 2η − 2kγ
This is valid as property Dk descends to quotients so, following the proof of
[16, Theorem 4.4], the algebra ρ(A) inherits property Dk allowing the use
of Lemma 4.12 above in place of [16, Theorem 4.2]. Note that one should
take care with issues of degeneracy here. In particular, the representation π
of B in the proof of Theorem 4.4 of [16] should be assumed non-degenerate.
Proposition 4.2 now shows that B ⊂cb,2k A and A ⊂cb,2k B. Therefore
(cid:3)
dcb(A, B) ≤ 2 max(2kγ, 2kγ) = 4kγ
Corollary 4.14. Let A be a C ∗-algebra generated by two commuting non-
degenerate C ∗-subalgebras one of which is nuclear and has no finite dimen-
sional irreducible representations. Suppose that A ⊂ B(H) and B is an-
other C ∗-subalgebra of B(H) with d(A, B) < γ for γ < 1/6422957. Then
dcb(A, B) < 1/42 and (Cu(A), Σ(Cu(A)) ∼= (Cu(B), Σ(Cu(B)).
20
F. PERERA, A. TOMS, S. WHITE, AND W. WINTER
Proof. By Proposition 4.8, A has property Dk for k = 5/2 so in Proposition
4.13, β = 1585056γ and η = 3203122γ + 52306848000γ2 so that 2η + 2kγ <
1011γ < 1 for γ < 10−11. The bound on γ ensures that (4.2) holds so that
Proposition 4.13 applies. Further this bound gives
4kγ
1 − 2η − 2kγ
<
1
42
,
and so the result follows from Proposition 4.13 and Theorem 3.10.
(cid:3)
In particular, C ∗-algebras sufficiently close to Z-stable algebras are au-
tomatically completely close and have the Cuntz semigroup of a Z-stable
algebra. The question of whether the property of Z-stability transfers to
sufficiently close subalgebras raised in [17] remains open.
Corollary 4.15. Let A be a Z-stable C ∗-algebra and suppose that B is
another C ∗-algebra acting on the same Hilbert space as A with d(A, B) <
1/6422957. Then dcb(A, B) < 1/42, (Cu(A), Σ(Cu(A)) ∼= (Cu(B), Σ(Cu(B)).
In particular B has the Cuntz semigroup of a Z-stable algebra.
5. Quasitraces
In this section we use our isomorphism between the Cuntz semigroups
of completely close C ∗-algebras to give an affine homeomorphism between
the lower semicontinuous quasitraces on such algebras. This isomorphism
is compatible with the affine isomorphism of the trace spaces of close C ∗-
algebras constructed in [16, Section 5].
Given a C ∗-algebra A, write T (A) for the cone of lower semicontinuous
traces on A and QT2(A) for the cone of lower semicontinuous 2-quasitraces
on A. Precisely, a trace τ on A is a linear function τ : A+ → [0,∞] van-
ishing at 0 and satisfying the trace identity τ (xx∗) = τ (x∗x) for all x ∈ A.
A 2-quasitrace is a function τ : A+ → [0,∞] vanishing at 0 which satis-
fies the trace identity and which is linear on commuting elements of A+.
Write Ts(A) for the simplex of tracial states on A and QT2,s(A) for the
bounded 2-quasitraces on A of norm one. Lower semicontinuous traces and
2-quasitraces on A extend uniquely to lower semicontinuous traces and 2-
quasitraces respectively on A ⊗ K (see [7, Remark 2.27(viii)]).
In [19, Section 4], Elliott, Robert and Santiago extend earlier work of
Blackadar and Handelman, setting out how functionals on Cu(A) arise from
elements of QT2(A). Precisely a functional on Cu(A) is a map f : Cu(A) →
[0,∞] which is additive, order preserving, has f (0) = 0 and preserves
the suprema of increasing sequences. Given τ ∈ QT2(A), the expression
dτ (hai) = limn→∞ τ (a1/n) gives a well defined functional on Cu(A), where
we abuse notation by using τ to denote the extension of the original lower
semicontinuous 2-quasitrace to A ⊗ K. Alternatively, one can define dτ by
dτ (hai) = limn→∞ τ (an), where (an)∞
n=1 is any very rapidly increasing se-
quence from (A⊗ K)+
1 representing hai. Conversely, given a functional f on
Cu(A), a lower semicontinuous 2-quasitrace on A ⊗ K (and hence on A) is
THE CUNTZ SEMIGROUP AND STABILITY OF CLOSE C ∗-ALGEBRAS
21
0 f (h(a − t)+i)dt. With this notation, the assignments
The topology on QT2(A) is specified by saying that a net (τi) in QT2(A)
given by τf (a) = R ∞
τ 7→ dτ and f 7→ τf are mutually inverse (see [19, Proposition 4.2]).
converges to τ ∈ QT2(A) if and only if
lim sup
i
τi((a − ε)+) ≤ τ (a) ≤ lim inf
i
τi(a)
for all a ∈ A+ and ε > 0. With this topology QT2(A) is a compact Hausdorff
space [19, Theorem 4.4] and T (A) is compact in the induced topology [19,
Theorem 3.7].
In a similar fashion, the cone of functionals on Cu(A) is
topologised by defining λi → λ if and only if
lim sup
i
λi(h(a − ε)+i) ≤ λ(hai) ≤ lim inf
i
λi(hai)
for all a ∈ (A ⊗ K)+ and ε > 0. Theorem 4.4 of [19] shows that the affine
map τ 7→ dτ is a homeomorphism between the cone QT2(A) and the cone of
functionals on the Cuntz semigroup.
Theorem 5.1.
(1) Let A, B be C ∗-algebras acting non-degenerately on
a Hilbert space with dcb(A, B) < 1/42. Then the isomorphism
Φ : (Cu(A), Σ(Cu(A))) → (Cu(B), Σ(Cu(B)))
given by Theorem 3.10 induces an affine homeomorphism
satisfying
(5.1)
bΦ : QT2(B) → QT2(A)
d Φ(τ )(x) = dτ (Φ(x))
for all x ∈ Cu(A) and τ ∈ QT2(B).
(2) Suppose additionally that A and B are unital and dcb(A, B) < γ <
1/2200. Then the map Φ from (1) is compatible with the map Ψ :
Ts(B) → Ts(A) given in Lemma 5.4 of [16]. Precisely, for τ ∈ Ts(B),
we have bΦ(τ ) ∈ Ts(A) ⊂ QT2(A) and bΦ(τ ) = Ψ(τ ).
Proof. The first part of the theorem is a consequence of Theorem 3.10 and
[19, Proposition 4.2]: given τ ∈ QT2(B), define bΦ(τ ) to be the lower semi-
continuous 2-quasitrace induced by the functional dτ ◦ Φ on Cu(A). It is
immediate from the construction that the map bΦ is affine, bijective and the
identity (1) holds.
To show that bΦ is continuous, we use the homeomorphism between the
cone of lower semicontinuous quasi-traces and functionals on the Cuntz semi-
group in [19, Theorem 4.4]. Consider a net (τi) in QT2(B) with τi → τ . Fix
a ∈ A+, then,
dτ (Φ(hai)) ≤ lim inf
i
dτi(Φ(hai)),
as dτi → dτ . Now take ε > 0 and fix a contraction b ∈ (B ⊗ K)+ with
Φ(hai) = hbi. As (h(b−1/n)+i)∞
n=1 is very rapidly increasing with supremum
22
F. PERERA, A. TOMS, S. WHITE, AND W. WINTER
hbi, there exists n ∈ N with Φ(h(a − ε)+i) ≤ h(b − 1/n)+i. As
lim sup
dτi(hb − 1/n)+i) ≤ dτ (hbi),
it follows that
i
lim sup
i
d bΦ(τi)(h(a − ε)+i) ≤ d bΦ(τ )(hai) ≤ lim inf
i
d bΦ(τ )(hai).
Thus d bΦ(τi) → d bΦ(τ ) and so, using the homeomorphism between QT2(A) and
functionals on Cu(A), we have bΦ(τi) → bΦ(τ ). Therefore bΦ is continuous,
and hence a homeomorphism between QT2(B) and QT2(A).
For the second part we first need to review the construction of the map
Ψ from [16]. Suppose dcb(A, B) < γ < 1/2200. Write C = C ∗(A, B) and
let C ⊂ B(H) be the universal representation of C so that M = A′′ and
N = B′′ are isometrically isomorphic to A∗∗ and B∗∗ respectively. Note
that the Kaplansky density argument of [27, Lemma 5] gives dcb(M, N) ≤
dcb(A, B). Following the proof of [16, Lemma 5.4] we can find a unitary
u ∈ (Z(M) ∪ Z(N))′′ such that Z(uMu∗) = Z(N) and ku − 1Ck ≤ 5γ.
We write A1 = uAu∗ and M1 = uMu∗. There is now a projection zfin ∈
Z(M1) = Z(N) which simultaneously decomposes M1 = M1zfin ⊕ M1(1 −
zfin) and N = Nzfin ⊕ N(1 − zfin) into the finite and properly infinite parts
respectively ([16, Lemma 3.5] or [27]). Given a tracial state τ on B, there
is a unique extension τ ′′ to N, which then factors uniquely through the
centre valued trace TrNzfin on Nzfin. That is, τ ′′(x) = (φτ ◦ TrNzfin )(xzfin)
for some state φτ on Nzfin. The map Ψ in [16] is then given by defining
Ψ(τ )(y) = (φτ ◦ TrM1zfin)(uyu∗zfin) for y ∈ A.
Now fix τ ∈ Ts(B). For m ∈ N and a ∈ (A⊗Mm)+
1 , consider the standard
very rapidly increasing sequence (g2−(n+1),2−n(a))∞
n=1 which represents hai.
Let pn ∈ M ⊗ Mm be the spectral projection for a for [2−(n+1), 1], so that
the alternating sequence
g2−2,2−1(a), p1, g2−3,2−2(a), p2, g2−4,2−3(a), p3, . . .
is very rapidly increasing. Then
(5.2)
dΨ(τ )(hai) = sup
Choose bn ∈ (B ⊗ Mm)+
n
(Ψ(τ ))(g2−n ,2−(n+1)(a)) = sup
n
Ψ(τ )′′(pn).
1 with kg2−(n+1),2−n(a)− bnk ≤ 2γ and projections
qn ∈ N ⊗ Mm with kpn − qnk ≤ 2γ (by a standard functional calculus
argument, see [11, Lemma 2.1]). Note that dcb(M1, N) ≤ 11γ and the
algebras (M1 ⊗ Mm)(zfin ⊗ 1m) and (N1 ⊗ Mm)(zfin ⊗ 1m) have the same
centre. Since k(u ⊗ 1m)p(u ⊗ 1m)∗(zfin ⊗ 1m) − q(zfin ⊗ 1m)k < 1/2, Lemma
3.6 of [16] applies to show that
(TrM1zfin⊗trm)((u⊗1m)pn(u⊗1m)∗(zfin⊗1m)) = (TrNzfin⊗trm)(q(zfin⊗1m)).
This ensures that Ψ(τ )′′(pn) = τ ′′(qn) for all n.
As each (qn − 18γ)+ = qn, the sequence
(b1 − 18γ)+, q1, (b2 − 18γ)+, q2, (b3 − 18γ)+, q3, . . .
THE CUNTZ SEMIGROUP AND STABILITY OF CLOSE C ∗-ALGEBRAS
23
is upwards directed by Lemma 3.8 and the supremum of this sequence defines
Φ(hai). We then have
(5.3)
τ ′′(qn).
dτ (Φ(hai)) = sup
n
Indeed, dτ (Φ(hai)) is given by sup τ (cn), where (cn)∞
n=1 is any very rapidly
increasing sequence in (B⊗K)+ representing Φ(hai). But, working in Cu(N),
Proposition 3.6 shows that any such very rapidly increasing sequence (cn)∞
n=1
can be intertwined with the very rapidly increasing sequence (qn)∞
n=1 after
telescoping, and this establishes (5.3). Combining (5.2) and (5.3), we have
(5.4)
dΨ(τ )(hai) = d bΦ(τ )(hai)
for all m ∈ N and a ∈ (A ⊗ Mm)+. As functionals on the Cuntz-semigroup
preserve suprema, (5.4) holds for all a ∈ (A⊗K)+, whence Ψ(τ ) = bΦ(τ ). (cid:3)
The homeomorphism between the lower semicontinuous quasi-traces can
be used to establish the weak∗-continuity of the map between the tracial
state spaces of close unital C ∗-algebras from [16, Section 5] resolving a point
left open there. In particular this shows that the map defined in [16] provides
an isomorphism between the Elliott invariants of completely close algebras,
as a priori the
For any closed two-sided ideal I ✂A, the subcone TI (A) of T (A) consists of
those τ ∈ T (A) such that the closed two-sided ideal generated by {x ∈ A+ :
τ (x) < ∞} is I. Proposition 3.11 of [19] shows that the relative topology
on TI (A) is the topology of pointwise convergence on the positive elements
of the Pedersen ideal of I. In particular, Ts(A) ⊂ TA(A). In particular, the
induced topology on Ts(A) is just the weak∗-topology.
Corollary 5.2. Suppose that A and B are unital C ∗-algebras acting non-
degenerately on a Hilbert space with dcb(A, B) < 1/42 and d(A, B) < 1/2200.
Then the affine isomorphism Ψ : Ts(B) → Ts(A) between tracial state spaces
in [16, Section 5] is a homeomorphism with respect to the weak∗-topologies.
We end with two further corollaries of Theorem 5.1.
Corollary 5.3. Let A and B be unital C ∗-algebras acting non-degenerately
on the same Hilbert space with dcb(A, B) < 1/2200. Suppose every bounded
2-quasitrace on A is a trace, then the same property holds for B.
Proof. Given τ ∈ QT2,s(B), its image bΦ(τ ) lies in QT2,s(A) = Ts(A). By
Theorem 5.1 (2) (applied with A and B interchanged)
as claimed.
τ = bΦ−1(bΦ(τ )) = Ψ−1(bΦ(τ )) ∈ Ts(B),
(cid:3)
The question of whether exactness transfers to (completely) close C ∗-
algebras raised in [16] remains open, but we do at least obtain the following
corollary.
24
F. PERERA, A. TOMS, S. WHITE, AND W. WINTER
Corollary 5.4. Let A and B be unital C ∗-algebras acting non-degenerately
on the same Hilbert space with dcb(A, B) < 1/2200 and suppose A is exact.
Then every bounded 2-quasitrace on B is a trace.
Proof. This is immediate from Haagerup's result that bounded 2-quasitraces
on exact C ∗-algebras are traces ([21]) and the previous corollary.
(cid:3)
We end by noting that the isomorphism between the Cuntz semigroups
of completely close algebras in Theorem 3.10 can also be used to directly
recapture an isomorphism between the Elliott invariants in significant cases.
Let CuT be the functor A 7→ Cu(A ⊗ C(T)) mapping the category of C ∗-
algebras into the category Cu introduced in [18] and let Ell be the Elliott
invariant functor taking values in the category Inv whose objects are the
4-tuples arising from the Elliott invariant. Let C be the subcategory of
separable, unital, simple finite and Z-stable algebras A with QT2(A) = T (A)
(for example if A is exact). Then, building on work from [6, 5], Theorem 4.2
of [2] provides functors F : Inv → Cu and G : Cu → Inv such that there
are natural equivalences of functors F ◦ EllC ∼= CuTC and G◦ CuTC ∼= EllC
(a similar result for simple unital ASH algebras which are not type I and
have slow dimension growth can be found in [43]). Note that in Theorem 4.2
there is an implicit nuclearity hypothesis, which is only actually used in order
to see QT2(A) = T (A) -- the result holds in the generality stated. Thus if
A and B are Z-stable C ∗-algebras with dcb(A, B) sufficiently small, and A
is simple, separable, unital finite and has QT2(A) = T (A), then B enjoys
all these properties. Further, since tensoring by an abelian algebra does not
increase the complete distance between A and B (see [13, Theorem 3.2] for
this result in the context of near inclusions -- the same proof works for the
metric dcb), CuT(A) ∼= CuT(B) by Theorem 3.10. Thus Ell(A) ∼= Ell(B).
Acknowledgements. This work was initiated at the Centre de Recerca
Matem`atica (Bellaterra) during the Programme "The Cuntz Semigroup and
the Classification of C ∗-algebras" in 2011. The authors would like to thank
the CRM for the financial support for this programme and the conducive
research environment.
References
[1] R. Antoine, J. Bosa, and F. Perera. Completions of monoids with applications to the
Cuntz semigroup. Internat. J. Math., 22(6):837 -- 861, 2011.
[2] R. Antoine, M. Dadarlat, F. Perera and L. Santiago. Recovering the Elliott invariant
from the Cuntz semigroup. Trans. Amer. Math. Soc., in press, arXiv:1109.5803v1,
[math.OA], 2011.
[3] P. Ara and E. Pardo. Refinement monoids with weak comparability and applications
to regular rings and C ∗-algebras. Proc. Amer. Math. Soc., 124(3):715 -- 720, 1996.
[4] P. Ara, F. Perera, and A. S. Toms. K-theory for operator algebras. Classification of
C ∗-algebras. In Aspects of operator algebras and applications, volume 534 of Contemp.
Math., pages 1 -- 71. Amer. Math. Soc., Providence, RI, 2011.
[5] N. P. Brown, A. S. Toms. Three applications of the Cuntz semigroup. Int. Math. Res.
Not. IMRN, no. 19, Art. ID rnm068, 14 pp, 2007.
THE CUNTZ SEMIGROUP AND STABILITY OF CLOSE C ∗-ALGEBRAS
25
[6] N. P. Brown, F. Perera, A. S. Toms. The Cuntz semigroup, the Elliott conjecture,
and dimension functions on C∗-algebras. J. Reine Angew. Math. 621, 191-211, 2008.
[7] E. Blanchard and E. Kirchberg. Non-simple purely infinite C ∗-algebras: the Hausdorff
case. J. Funct. Anal., 207:461 -- 513, 2004.
[8] J. Cameron, E. Christensen, A. Sinclair, R. R. Smith, S. White, and A. Wiggins.
Kadison-Kastler stable factors. arXiv:1209.4116 [math.OA], 2012.
[9] J. Cameron, E. Christensen, A. Sinclair, R. R. Smith, S. White, and A. Wiggins. A
remark on the similarity and perturbation problems. C. R. Math. Rep. Acad. Sci.
Canada, in press, arXiv:1206.5405 [math.OA], 2012.
[10] J. Cameron, E. Christensen, A. Sinclair, R. R. Smith, S. White, and A. Wiggins.
Type II1 factors satisfying the spatial isomorphism conjecture. Proc. Natl. Acad. Sci.
USA.,109(5) 20338-20343, 2012.
[11] E. Christensen. Perturbations of type I von Neumann algebras. J. London Math. Soc.
(2), 9:395 -- 405, 1975.
[12] E. Christensen. Perturbations of operator algebras. II. Indiana Univ. Math. J.,
26(5):891 -- 904, 1977.
[13] E. Christensen. Near inclusions of C ∗-algebras. Acta Math., 144(3-4):249 -- 265, 1980.
[14] E. Christensen. Extensions of derivations. II. Math. Scand., 50(1):111 -- 122, 1982.
[15] E. Christensen. Finite von Neumann algebra factors with property Γ. J. Funct. Anal.,
186:366-380, 2001.
[16] E. Christensen, A. Sinclair, R. R. Smith, and S. White. Perturbations of C ∗-algebraic
invariants. Geom. Funct. Anal., 20(2):368 -- 397, 2010.
[17] E. Christensen, A. Sinclair, R. R. Smith, S. White, and W. Winter. Perturbations of
nuclear C ∗-algebras. Acta Math., 208:93 -- 150, 2012.
[18] K. T. Coward, G. A. Elliott, and C. Ivanescu. The Cuntz semigroup as an invariant
for C ∗-algebras. J. Reine Angew. Math., 623:161 -- 193, 2008.
[19] G. Elliott, L. Robert, and L. Santiago. The cone of lower semicontinuous traces on a
C ∗-algebra. Amer. J. Math., 133(4):969 -- 1005, 2011.
[20] U. Haagerup. Solution of the similarity problem for cyclic representations of C ∗-
algebras. Ann. of Math. (2), 118(2):215 -- 240, 1983.
[21] U. Haagerup. Quasi-traces on exact C ∗-algebras are traces. Preprint, 1991.
[22] D. Hadwin and W. Li The similarity degree of some C ∗-algebras. arXiv:1210.6015.
[math.OA], 2012.
[23] I. Hirshberg, E. Kirchberg, and S. White. Decomposable approximations of nuclear
C ∗-algebras. Adv. Math., 230:1029 -- 1039, 2012.
[24] J. Hjelmborg and M. Rørdam. On stability of C ∗-algebras. J. Funct. Anal.,
155(1):153 -- 170, 1998.
[25] M. Johanesova and W. Winter. The similarity problem for -stable C ∗-algebras. Bull.
London Math. Soc., 44(6):1215 -- 1220, 2012.
[26] R. V. Kadison. On the orthogonalization of operator representations. Amer. J. Math.,
77:600 -- 620, 1955.
[27] R. V. Kadison and D. Kastler. Perturbations of von Neumann algebras. I. Stability
of type. Amer. J. Math., 94:38 -- 54, 1972.
[28] M. Khoshkam. Perturbations of C ∗-algebras and K-theory. J. Operator Theory,
12(1):89 -- 99, 1984.
[29] E. Kirchberg. The derivation problem and the similarity problem are equivalent. J.
Operator Theory, 36(1):59 -- 62, 1996.
[30] W. Li and J. Shen. A note on approximately divisible C ∗-algebras. arXiv:0804.0465
[math.OA], 2008.
[31] H. Matui and Y. Sato. Strict comparison and Z-absorption of nuclear C ∗-algebras.
Acta. Math., in press. arXiv:1111.1637v1 [math.OA], 2011.
26
F. PERERA, A. TOMS, S. WHITE, AND W. WINTER
[32] E. Ortega, F. Perera, and M. Rørdam. The Corona factorization property, stability,
and the Cuntz semigroup of a C ∗-algebra. Int. Math. Res. Not. IMRN, (1):34 -- 66,
2012.
[33] F. Perera. The structure of positive elements for C ∗-algebras with real rank zero.
Internat. J. Math., 8(3):383 -- 405, 1997.
[34] J. Phillips. Perturbations of C ∗-algebras. Indiana Univ. Math. J., 23:1167 -- 1176,
1973/74.
[35] J. Phillips and I. Raeburn. Perturbations of AF-algebras. Canad. J. Math.,
31(5):1012 -- 1016, 1979.
[36] G. Pisier. The similarity degree of an operator algebra. St. Petersburg Math. J.,
10(1):103 -- 146, 1999.
[37] F. Pop. The similarity problem for tensor products of certain C ∗-algebras. Bull. Aus-
tral. Math. Soc., 70(3):385 -- 389, 2004.
[38] M. A. Rieffel. Dimension and stable rank in the K-theory of C ∗-algebras. Proc. Lon-
don Math. Soc.,, 46:301 -- 333, 1983.
[39] L. Robert Classification of inductive limits of 1-dimensional NCCW complexes. Adv.
Math., 231(5):2802 -- 2836, 2012.
[40] L. Robert and L. Santiago. Classification of C ∗-homomorphisms from C0(0, 1] to a
C ∗-algebra. J. Funct. Anal., 258(3):869 -- 892, 2010.
[41] M. Rørdam and W. Winter. The Jiang-Su algebra revisited. J. Reine Angew. Math.,
642:129 -- 155, 2010.
[42] A. Toms. On the classification problem for nuclear C ∗-algebras. Ann. of Math. (2),
167(3):1029 -- 1044, 2008.
[43] A. Tikuisis. The Cuntz semigroup of continuous functions into certain simple C ∗-
algebras. Int. J. Math. 22, 1-37, 2011.
[44] W. Winter. Decomposition rank and Z-stability. Invent. Math., 179(2):229 -- 301, 2010.
[45] W. Winter. Nuclear dimension and Z-stability of pure C ∗-algebras. Invent. Math.,
187(2):259 -- 342, 2012.
[46] S. Zhang. Diagonalizing projections in multiplier algebras and in matrices over a
C ∗-algebra. Pacific J. Math., 145(1):181 -- 200, 1990.
Francesc Perera, Department of Mathematics, 08193 Bellaterra, Barcelona,
Spain.
E-mail address: [email protected]
Andrew Toms, Department of Mathematics, Purdue University, 150 North
University Street, West Lafayette, IN 47907, USA.
E-mail address: [email protected]
Stuart White, School of Mathematics and Statistics, University of Glasgow,
University Gardens, Glasgow Q12 8QW, Scotland.
E-mail address: [email protected]
Wilhelm Winter, Mathematisches Institut der WWU Munster, Einsteinstrasse
62, 48149, Munster, Germany.
E-mail address: [email protected]
|
1203.1806 | 2 | 1203 | 2012-05-15T12:19:03 | Some analysis on amalgamated free products of von Neumann algebras in non-tracial setting | [
"math.OA"
] | Several techniques together with some partial answers are given to the questions of factoriality, type classification and fullness for amalgamated free product von Neumann algebras. | math.OA | math |
SOME ANALYSIS ON AMALGAMATED FREE PRODUCTS OF
VON NEUMANN ALGEBRAS IN NON-TRACIAL SETTING
YOSHIMICHI UEDA
Abstract. Several techniques together with some partial answers are given to the questions
of factoriality, type classification and fullness for amalgamated free product von Neumann
algebras.
1. Introduction
It was quite recent that the complete answers were given in [27, 28] to the questions of
factoriality, type classification, fullness and Sd- and τ -invariants for arbitrary free product von
Neumann algebras.
It is natural as a next project to consider the same questions for more
general amalgamated free product von Neumann algebras. Such attempts were already made
by us [21, 24, 23] almost 10 years ago for amalgamated free products over Cartan subalgebras.
However the results there are far from satisfactory as compared to those on plain free prod-
uct von Neumann algebras. The aim of this paper is to take a still very first step towards
'satisfactory' answers to those questions for amalgamated free product von Neumann algebras.
As simple consequences we will give partial answers at least when amalgamated free products
are taken over type I von Neumann algebras, which are improvements of our previous works
[21, 24, 23, 25, 26].
The proofs in [27, 28] are divided into analytical and combinatorial parts in essence. Combi-
natorial parts are completed by some 'induction arguments', whose essential idea originates in
several works due to Dykema, especially [6]. On the other hand, analytical parts are devoted
to proving several inequalities involving the Hilbert space norms arising from some states of
particular form (instead of so-called free product states themselves), whose essential ideas ap-
parently go back to the ICC argument for factoriality of group von Neumann algebras and the
so-called 14 ε-argument both due to Murray and von Neumann. However our problems are of
the nature of type III von Neumann algebras, and thus the lack of trace causes main difficulties.
Hence the key is to overcome such difficulties. Here we will take up such analytical aspects
in the general amalgamated free product setup, and indeed improve the analytical results in
[27, 28] with new techniques from the recent amazing development on type II1 factors opened
by several breakthroughs due to Popa. We hope that the technical facts provided in this paper
are sufficient as analytical parts in future 'best-possible' answers to the questions mentioned
above at least in the case where amalgamated free products are taken over type I von Neumann
subalgebras.
The organization of this paper is as follows. Section 2 is preliminaries on amalgamated
free product von Neumann algebras. In section 3 we provide a non-tracial version of one of
the results in Ioana -- Peterson -- Popa's article [9, Theorem 1.1]. In relation to it we provide a
non-tracial adaptation of the so-called intertwining-by-bimidule criterion due to Popa, which
may be of independent interest as future reference.
In the same section we also generalize
our previous results of controlling central sequences [27, Proposition 3.5],[28, Proposition 3.1]
to the amalgamated free product setting.
In section 4, we give several partial answers to
1
2
Y. UEDA
the questions mentioned above by utilizing technologies developed in §3. Those include an
answer to the factoriality and non-amenability questions of a given amalgamated free product
(M, E) = (M1, E1) ⋆N (M2, E2) when M1 is 'diffuse relative to N ', M2 'non-trivial relative to
N ', and N of type I.
Standard notation rule here follows our previous papers [27, 28]; for example, the center,
the unitary group and the set of projections of a given von Neumann algebra M are denoted
by Z(M ), M u and M p, respectively, and also the central support of e ∈ M p in M by cM
e .
Notations and facts concerning amalgamated free products of von Neumann algebras will be
summarized in next section 2.
2. Amalgamated Free Product von Neumann Algebras
Let M1 ⊇ N ⊆ M2 be σ-finite von Neumann algebras, and faithful normal conditional
expectations E1 : M1 → N , E2 : M2 → N be given. Their amalgamated free product
(M, E) = (M1, E1) ⋆N (M2, E2) is a pair of von Neumann algebra M containing M1 ⊇ N ⊆ M2
and faithful normal conditional expectation E : M → N satisfying (i) M = M1 ∨ M2, (ii)
2 )≡ 0, where Λ◦(M ◦
2 ) denotes the set of all
E ↾Mk = Ek (k = 1, 2) and (iii) E ↾Λ◦(M ◦
alternating words in M ◦
2 := Ker(E2). The construction of such a pair is
a bit complicated, but this simple formulation perfectly serves as a working definition. The
construction was introduced in the tracial setting in [15] based on the C∗-algebraic one [31].
Its modular theoretical treatment was given in [21], and will be reviewed below.
1 ,M ◦
1 := Ker(E1) and M ◦
1 , M ◦
Let χ be a faithful normal semifinite weight on N . Then the modular automorphism σχ◦E
,
t
t ∈ R, is simply computed as
σχ◦E
t
↾Mk = σχ◦Ek
t
(k = 1, 2),
(2.1)
see [21, Theorem 2.6]. This formula together with famous Takesaki's criterion shows that
for each k = 1, 2 there is a unique faithful normal conditional expectation EMk : M → Mk
characterized by
(2.2)
EMk ↾Λ◦(M ◦
1 ,M ◦
2 )\M ◦
k≡ 0.
This fact is easily confirmed in the exactly same way as in [27, Lemma 2.1]. It is clear that
E ◦ EMk = E holds. Consider the natural inclusion of the so-called continuous cores:
fM := M ⋊σχ◦E R ⊇ fMk := Mk ⋊σχ◦Ek R (k = 1, 2) ⊇ eN := N ⋊σχ R,
which is independent of the choice of χ thanks to Connes's Radon-Nikodym cocycle theorem.
(2.3)
σχ◦E R,
eEk := Ek ¯⊗IdB(L2(R)) ↾Mk⋊
The canonical liftings (still being faithful normal conditional expectations) eE : fM → eN , eEk :
fMk → eN (k = 1, 2) are constructed by
eE := E ¯⊗IdB(L2(R)) ↾M ⋊
Remark that the original E and Ek are recovered as the restrictions of eE and eEk to M and
Mk via the canonical embeddings M ֒→ fM and Mk ֒→ fMk, respectively. Here is a simple but
important fact [21, Theorem 5.1] that fM1 and fM2 are freely independent with amalgamation
over eN with respect to eE, and moreover fM = fM1 ∨ fM2. Consequently the following natural
(k = 1, 2) and Tr eN on fM , fMk and
= Tr eN ◦ eEk
eN , respectively, (see [20, Theorem XII.1.1]) must satisfy TrfM = Tr eN ◦ eE and Tr fMk
(see e.g. [14, §4]).
(fM , eE) = (fM1, eE1) ⋆ eN (fM2, eE2).
The canonical faithful normal semifinite traces Tr fM , TrfMk
formula holds:
(2.4)
R .
σχ◦Ek
(2.5)
NON-TRACIAL AMALGAMATED FREE PRODUCTS
3
Let M ω ⊇ M ω
k (k = 1, 2) ⊇ N ω be the ultraproducts of M ⊇ Mk (k = 1, 2) ⊇ N . Here
the inclusion relation is guaranteed by the existence of conditional expectations, and E and
Ek (k = 1, 2) can be lifted up to Eω : M ω → N ω and Eω
k → N ω, respectively. All
the necessary facts on ultraproducts of von Neumann algebras are summarized in [27, §§2.2].
Remark that M ω
2 are freely independent with amalgamation over N ω with respect to
Eω, see [23, Proposition 4]. However it is hopeless due to [18, Lemma 2.2] that M ω = M ω
1 ∨M ω
2
holds.
1 and M ω
k : M ω
3. Technical Results
3.1. A non-tracial adaptation of Popa's intertwining-by-bimodule criterion. Let M
be an arbitrary σ-finite (possibly type III) von Neumann algebra, and A, B be its (possibly non-
unital) von Neumann subalgebras with units 1A, 1B, respectively. Suppose that B is semifinite
with a faithful normal semifinite trace TrB and furthermore that there is a faithful normal
conditional expectation EB : 1BM 1B → B.
Proposition 3.1. The following are equivalent:
(i) There is no net uλ of unitaries in A which satisfies EB(y∗uλx) −→ 0 σ-strongly for
any x, y ∈S(cid:8)1AM p p ∈ Bp; TrN (p) < +∞}.
(ii) There are a normal (possibly non-unital) ∗-homomorphism ρ : A → Mn(C) ¯⊗B with
finite n ∈ N and a non-zero partial isometry w ∈ Mn(C) ¯⊗M such that
-- (Trn ¯⊗TrB)(ρ(1A)) < +∞,
-- ww∗ ≤ e11 ⊗ 1A and w∗w ≤ ρ(1A), and
-- (e11 ⊗ a)w = wρ(a) for all a ∈ A.
(iii) There are non-zero projections e ∈ A, f ∈ B, a normal unital ∗-isomorphism θ : eAe →
f Bf and a non-zero partial isometry v ∈ M such that
e is finite in A and TrB(f ) < +∞,
-- the central support cA
-- vv∗ ≤ e and v∗v ≤ f , and
-- xv = vθ(x) for all x ∈ eAe.
Suppose further that M has an almost periodic weight ψ such that both A and B sit inside the
centralizer Mψ, ψ ↾B is still semifinite, and the EB is the unique ψ ↾1BM1B -preserving one.
Then the w in (ii) and the v in (iii) can be chosen in such a way that there is a common
eigenvalue λ of ∆ψ so that (idn ¯⊗ σψ
t (v) = λitv for all t ∈ R.
t )(w) = λitw and σψ
As usual let us write A (cid:22)M B (with EB and TrB) if the above equivalent conditions (i) -- (iii)
hold. Remark that no assumption on A is necessary. The proof is of course modeled after Popa's
original one for finite von Neumann algebras, but some cares are necessary. Indeed we observed
this fact with B finite several years ago, through our attempt to get better understanding of the
fundamental articles [16, 17] due to Popa. Houdayer and Vaes informed us that they have also
observed it with B finite independently (see [8, Theorem 2.3]), and moreover Vaes corrected our
misunderstanding on some argument in [4, §2]. The proof below is just a combination and/or
a reformulation of several existing proofs of Popa's criterion [16, Appendix],[17, §2] (also see
[2, Appendix F],[29, Appendix C] for its exposition) and its variants [1, §3],[4, §2],[7, §4], etc.
The same idea as in e.g. the proof of (1) ⇒ (4) in [29, Proposition C.1] perfectly works for (ii)
⇒ (i). (Note that the proof of (4) ⇒ (1) in [2, Theorem F.12] does not work at this point due
to the lack of finite trace. Thus we could not prove (iii) ⇒ (i) directly.) Hence the main parts
below are (ii) ⇔ (iii) and (i) ⇒ (ii).
Proof of (ii) ⇒ (i): We may assume that ρ(1A) = Pn
to [11, Corollary 3.20]. Since (Trn ¯⊗TrB)(ρ(1A)) < +∞, one has w = Pn
k=1 ekk ⊗ pk with pk ∈ Bp thanks
k=1 e1k ⊗ wk with
4
Y. UEDA
One can find a net uλ in Au in such a way that EB(w∗
i,j=1 eij ⊗ EB(w∗
wk = wkpk ∈ S(cid:8)1AM p p ∈ Bp; TrN (p) < +∞}. On contrary, suppose that (i) is not true.
i uλwj ) −→ 0 σ-strongly for all i, j,
and hence ρ(uλ)(id ¯⊗EB)(w∗w) = Pn
i uλwj) −→ 0 σ-strongly. Therefore,
k(idn ¯⊗EB)(w∗w)kTrn ¯⊗TrB = kρ(uλ)(idn ¯⊗EB)(w∗w)ρ(1A)kTrn ¯⊗TrB −→ 0, a contradiction to
w 6= 0.
(cid:3)
k=1 ek = cA
e1 = cA
k=1 e1k ⊗ cA
kvk = e1 and vkv∗
i,j=1 eij ⊗ θ(v∗
i=1 viv∗
i,k=1 e1ieik ⊗ viv∗
i = cA
Proof of (iii) ⇒ (ii): Since v∗v ∈ θ(eAe)′, one can find a non-zero z ∈ Z(eAe)p = (Z(A)e)p
in such a way that the normal ∗-homomorphism x ∈ zAz = (eAe)z 7→ θ(x)v∗v is injective.
Since cA
e is finite in A, by [12, Proposition 8.2.1] one can find non-zero, mutually orthogonal
and equivalent (in A) e1, . . . , en ∈ Ap in such a way that e1 ≤ z and Pn
e1 . We
have e1vv∗ 6= 0, since θ(e1)v∗v = v∗(e1vv∗)v by the choice of z and e1 ≤ z. Then one gets
partial isometries v1 := e1, v2, . . . , vn ∈ A so that v∗
k = ek (k = 2, . . . , n).
Since e1Ae1 ⊆ eAe, we can construct a normal ∗-homomorphism ρ : A → Mn(C) ¯⊗B by
ρ(a) := Pn
i avj ), a ∈ A. Set w := Pn
k=1 e1k ⊗ vkv with v in (iii), which defines
i vj v = δijv∗e1v and v∗e1v = θ(e1)v∗v 6= 0 as remarked
a non-zero partial isometry, since v∗v∗
ek for all k = 2, . . . , n, we have wρ(a) = Pn
before. Since Pn
i,j,k=1 e1iejk ⊗
i avkv =Pn
j avk) =Pn
vivθ(v∗
ek avkv = (e11 ⊗ a)w for all a ∈ A. Since
ρ(1A) ≤ 1n ⊗ f , one has (Trn ¯⊗TrB)(ρ(1A)) < +∞.
(cid:3)
Proof of (ii) ⇒ (iii): As in (ii) ⇒ (i) we may and do assume that ρ(1A) = Pn
k=1 ekk ⊗ pk
with TrB-finite pk ∈ Bp. Note that any union of finite number of TrB-finite projections is
again TrB-finite thanks to the Kaplansky formula [12, Theorem 6.1.7]. Thus p = Wn
k=1 pk is
TrB-finite, and replacing B by pBp (if necessary) we may and do assume that TrB(1B) < +∞.
Notice that A must be of the form A = A0 ⊕ Ker(ρ(−)w∗w) with A0 finite, since ρ(A) is finite.
Note here that w∗w ∈ ρ(A)′, and thus ρ(−)w∗w is a normal ∗-homomorphism.
Let us first assume that A0 has a type II1 direct summand. By [12, Lemma 6.5.6] one can find
nonzero, mutually orthogonal and equivalent (in A0) e1, . . . , en ∈ Ap
0 whose sum is the unit of
the type II1 direct summand. With the center-valued trace τ : Mn(C) ¯⊗B → C1 ¯⊗Z(B) we have
nτ (ρ(e1)) ≤ τ (1⊗ 1B) = nτ (e11 ⊗ 1B), implying that there is a partial isometry v1 ∈ Mn(C) ¯⊗B
such that v∗
1 ≤ e11 ⊗ 1B, we can
construct a normal unital ∗-isomorphism θ : eAe → f Bf with e := e1, f := θ(e) in such a way
1 for x ∈ eAe. Since w∗w ∈ ρ(A)′ ∩ ρ(1A)(cid:0)Mn(C) ¯⊗M(cid:1)ρ(1A) and
that e11 ⊗ θ(x) = v1ρ(x)v∗
ww∗ ∈(cid:0)Ce11 ¯⊗A(cid:1)′
1 is a non-zero
partial isometry whose left and right support projections are less than e11 ⊗ e and e11 ⊗ f ,
respectively, and hence wv∗
1 = e11 ⊗ v for some non-zero partial isometry v ∈ eM f . Then one
has e11 ⊗ xv = (e11 ⊗ x)wv∗
1 = wρ(x)v∗
We next consider the case that A0 is of type I, that is, there is an abelian (in A) e ∈ Ap
e = 1A0. With a MASA A between ρ(eAe) ⊕ Cρ(e)⊥ ⊆ Mn(C) ¯⊗B one can choose, by
with cA
[11, Theorem 3.18], mutually orthogonal and equivalent (in Mn(C) ¯⊗B) projections q1, . . . , qn
from A with Pn
k=1 qk = 1n ⊗ 1B. Then one immediately observes (by looking at their center-
valued traces) that every qk is equivalent to e11 ⊗ 1B in Mn(C) ¯⊗B. Since ρ(e)w∗w 6= 0, some
q := qk must satisfy qρ(e)w∗w 6= 0. In this way, we can choose a non-zero partial isometry
v1 ∈ Mn(C) ¯⊗B in such a way v∗
1 ≤ e11 ⊗ 1B, and thus v∗
1v1 ∈ ρ(eAe)′
1 v1 = qρ(e)(≤ ρ(e)), v1v∗
1 6= 0 (since qρ(e)w∗w 6= 0). Then we can construct a unital normal ∗-homomorphism
and wv∗
θ : eAe → f Bf with f := θ(e) by e11 ⊗ θ(x) = v1ρ(x)v∗
1 for x ∈ eAe and a non-zero y ∈ eM f
by e11 ⊗ y = wv∗
1 ) =
e11 ⊗ yθ(x) for x ∈ eAe, since v∗
1v1 ∈ ρ(eAe)′. Hence xy = yθ(x) for x ∈ eAe. Replacing e by
suitable z ∈ Z(eAe)p (if necessary) we can make θ injective with keeping both θ(e) = f and
∩ (e11 ⊗ 1A)(cid:0)Mn(C) ¯⊗M(cid:1)(e11 ⊗ 1A), it is easy to see that wv∗
1 = e11 ⊗ vθ(x) for x ∈ eAe.
1 . Moreover we have e11 ⊗ xy = (e11 ⊗ x)wv∗
1 ≤ e11 ⊗ 1B. Since v1ρ(e1)v∗
1 = wρ(x)v∗
1 = wv∗
1 (v1ρ(x)v∗
1v1 = ρ(e1) and v1v∗
1 = v1v∗
1 = wv∗
1 v1ρ(x)v∗
0
NON-TRACIAL AMALGAMATED FREE PRODUCTS
5
y = eyf . With the polar decomposition y = vy we get vv∗ ≤ e, v∗v ≤ f and xv = vθ(x) for
x ∈ eAe.
(cid:3)
We have two ways for completing the final part of the proof of (i) ⇒ (ii) below; one is the
use of Haagerup's Lp-space technologies and the other that of standard forms due to Araki,
Connes and Haagerup. Here we use the latter as easy way. In what follows (M y H, JM , P♮
M )
denotes a standard form of M , see [20, Definition IX.1.13].
Proof of (i) ⇒ (ii): Note that EB(y∗uλx) −→ 0 σ-strongly if and only if kEB(y∗uλx)kTrB −→
0 for any x, y ∈S{1AM p p ∈ Bp; TrB(p) < +∞}. Thus there are ε > 0 and F ⋐S{1AM p p ∈
Bp; TrB(p) < +∞} so that
Xx,y∈F kEB(y∗ux)kTrB ≥ ε
for all u ∈ Au.
(3.1)
Bx1⊥
B)1⊥
BM 1⊥
B, and set B := B ⊕ C1⊥
Each x ∈ F has a TrB-finite px ∈ Bp with x = xpx, and p := Wx∈F px must be TrB-finite as
remarked in (ii) ⇒ (iii). Thus, replacing B by pBp (if necessary) we may and do assume that
TrB is a finite trace, that is, TrB(1B) < +∞.
Choose a faithful normal state ϕ0 on 1⊥
B and E B : x ∈ M 7→
EB(1Bx1B) + ϕ0(1⊥
B giving a faithful normal conditional expectation from the whole
M onto B. Clearly B is still finite (since we have assumed that TrB is a finite trace), and the
B ∈ B 7→ TrB(b) + α ∈ C defines a faithful normal trace (not weight !) Tr B on
mapping b + α1⊥
B. Set ϕ := Tr B ◦ E B, a faithful normal positive linear functional on M , and let ξ0 ∈ P♮
M be its
unique representing vector. It is standard, by a usual exhaustion argument like e.g. the proof
of [19, Theorem IV.5.5], to see that there is a family of vectors {ξi}i∈I in H so that ξ0 is in the
family (thus 0 is regarded as a distinguished element in I) and moreover H =P⊕
i∈I [JM BJM ξi].
Therefore, one can construct an isometry U : H → ℓ2(I)⊗ L2( B) satisfying U ξ0 = δ0 ⊗ ΛTr B (1)
and U (JM x∗JM ) = (1⊗ J Bx∗J B)U for x ∈ B, where L2( B) is the usual standard Hilbert space
constructed out of Tr B, ΛTr B the canonical embedding of B to L2( B) and J B the canonical
unitary conjugation on L2( B). By the construction we observe that P := U U ∗ ∈ B(ℓ2(I)) ¯⊗ B
and moreover that the pair P(cid:0)B(ℓ2(I)) ¯⊗ B(cid:1)P and PCδ0 ⊗ 1 with the rank 1 projection PCδ0
onto Cδ0 is nothing but a concrete realization, modulo the unitary equivalence by U , of the
basic extension hM, Bi and the Jones projection e B associated with E B. Then
TrhM, Bi(−) := (TrB(ℓ2(I)) ¯⊗Tr B)(U (−)U ∗)
(3.2)
with the usual trace TrB(ℓ2(I)) on B(ℓ2(I)) gives a faithful normal semifinite trace on the
basic extension hM, Bi. For x ∈ B one has U xe BU ∗ = PCδ0 ⊗ x and hence TrhM, Bi(xe B) =
(TrB(ℓ2(I)) ¯⊗Tr B)(U xe BU ∗) = (TrB(ℓ2(I)) ¯⊗Tr B)(PCδ0 ⊗ x) = Tr B(x). Therefore, we get
TrhM, Bi(xe By) = TrhM, Bi(e Byxe B) = TrhM, Bi( EB(yx)e B) = ϕ(yx),
x, y ∈ M.
(3.3)
Let d :=Py∈F ye By∗ ∈ hM, Bi+, and then TrhM, Bi(d) = Py∈F ϕ(y∗y) < +∞ by (3.3). In
the exactly same way as in the proof of (1) ⇒ (2) of [2, Theorem F.12] we see, by using (3.1),
that the σ-weakly closed convex hull C of {u∗du u ∈ Au} does not contain 0. Moreover, it
is plain to see that JM 1BJM d = d. Since 1B ∈ Z( B) and hence JM 1BJM ∈ Z(hM, Bi), we
conclude that C sits in (1AJM 1BJM )hM, Bi(1AJM 1BJM ). Since d ≥ 0 and TrhM, Bi(d) < +∞,
C is embedded, as a closed convex set, into L2(hM, Bi, TrhM, Bi), the usual GNS Hilbert space
associated with TrhM, Bi. Hence one can choose a unique minimal point d0 ∈ C with respect to
, which in turn falls in (1AJM 1BJM )hM, Bi(1AJM 1BJM )∩A′
the Hilbert space norm k−kTrhM, Bi
6
Y. UEDA
aξj = Xn
i=1
and satisfies TrhM, Bi(d0) < +∞. Choosing a suitable spectral projection of d0 we get a nonzero
projection e ∈ hM, Bi ∩ A′ such that e ≤ 1AJM 1BJM and TrhM, Bi(e) < +∞.
The projection e apparently gives an A -- B bimodule K := eH with left and right (unital)
actions a · ξ · b := aJM b∗JM ξ for a ∈ A, ξ ∈ K, b ∈ B. The GNS representation of B
associated with TrB is simply given by the restriction B y L2(B) := 1BL2( B) with the
canonical embedding ΛTrB := ΛTr B
↾B, and moreover the canonical unitary conjugation JB is
also just the restriction of J B to L2(B). Thus we get the right B-module embedding U0 :=
U ↾K: K ֒→ ℓ2(I) ¯⊗L2(B)B (⊆ (1 ⊗ 1B)(ℓ2(I) ¯⊗L2( B)), and U0U ∗
0 ∈ B(ℓ2(I)) ¯⊗B satisfies
(TrB(ℓ2(I)) ¯⊗TrB)(U0U ∗
0 ) = TrhM, Bi(e) < +∞ by (3.2). By the same reason as in the beginning
of the proof of [2, Proposition F.10] or by [29, Lemma A.1] there are n ∈ N and a nonzero
z ∈ Z(B)p such that K0 := JM zJMK is still a non-trivial A -- B bimodule and (U0 ↾JM zJM K
0 - Pn ⊗ z in B(ℓ2(I)) ¯⊗B =
)(U0 ↾JM zJM K)∗ = (1 ⊗ JBzJB)U0U ∗
(C1 ¯⊗JBBJB)′, where Pn is a rank n projection in B(ℓ2(I)). Choose a partial isometry v ∈
(C1 ¯⊗JBBJB)′ with v∗v = (U0 ↾JM zJM K)(U0 ↾JM zJM K)∗ and vv∗ ≤ Pn ⊗ z, and then we can
define a right B-module embedding V : K0 ֒→ Cn ¯⊗L2(B) by V := v(U0 ↾JM zJM K) with a
fixed identification Pnℓ2(I) = Cn. The embedding V gives the normal (possibly non-unital)
∗-homomorphism ρ : a ∈ A 7→ V aV ∗ ∈ Mn(C) ¯⊗B.
Let δi (1 ≤ i ≤ n) be a standard basis of Cn, and set ξi
(1 ≤ i ≤ n). For a ∈ A, write ρ(a) = Pn
:= V ∗(δi ⊗ ΛTrB (1B)) ∈ K0
i,j=1 eij ⊗ ρ(a)ij with the matrix units eij associated
JM ρ(a)∗
0 (1 ⊗ JBzJB) = (1 ⊗ z)U0U ∗
with the δi, and then
ij JM ξi,
1 ≤ j ≤ n.
M. Set ξ := Pn
(3.4)
Consider M := Mn+1(C) ¯⊗M y L2(M) := Mn+1(C) ¯⊗H (by left matrix-multiplication) with
the canonical unitary conjugation JM defined by JM(eij ⊗ ξ) := eji ⊗ (JM ξ) for eij ⊗ ξ ∈ L2(M).
The natural cone determined by (M y L2(M), JM) is denoted by P♮
k=1 e0k ⊗
ξk ∈ L2(M), and define a normal (possibly non-unital) ∗-homomorphism ρ : A ֒→ M by
ρ(a) := e00 ⊗ a + Pn
i,j=1 eij ⊗ ρ(a)ij for a ∈ A. Here a standard matrix unit system eij
in Mn+1(C) is indexed by 0, 1, . . . , n. By (3.4) one has ρ(a) ξ = JM ρ(a)∗JM ξ for a ∈ A. A
standard fact on polar decomposition in standard forms (c.f. [20, Exercise IX.1.2],[1, Lemma
3.1]) guarantees the existence of a vector ξ ∈ P♮
M and a partial isometry w ∈ M satisfying that
w ξ = ξ, w∗ w = [M′ ξ], w w∗ = [M′ ξ] and ρ(a) w = w ρ(a) for a ∈ A. Since (e00 ⊗ 1A) ξ = ξ,
one has (e00 ⊗ 1A)[M′ ξ] = [M′ ξ], and thus w w∗ ≤ e00 ⊗ 1A. Here (ρ(A) ⊆) Mn(C) ¯⊗M is
naturally regarded as a corner of M by the numbering of the matrix units eij's. Then one has,
by (3.4) again, JMρ(1A)JM ξ = ξ, and hence JMρ(1A)JM ξ = ξ. By JM ξ = ξ ∈ P♮
M we get
ρ(1A)[M′ ξ] = [M′ ξ] so that w∗ w ≤ ρ(1A) ≤Pn
k=1 ekk ⊗ 1B. Therefore, w =Pn
k=1 e0k ⊗ wk
with wk ∈ 1AM 1B. Letting w := (e10 ⊗ 1A) w = Pn
k=1 e1k ⊗ wk ∈ Mn(C) ¯⊗M we have
w∗w ≤ ρ(1A), ww∗ ≤ e11 ⊗ 1A and (e11 ⊗ a)w = (e10 ⊗ 1A)ρ(a) w = (e10 ⊗ 1A) w ρ(a) = wρ(a)
for a ∈ A. We have assumed (by cutting by a projection in B) that TrB(1B) < +∞, and hence
(Trn ¯⊗TrB)(ρ(1A)) < +∞ is now trivial. Hence we are done.
(cid:3)
Proof of the second part of the assertion: Only the proof of (i) ⇒ (ii) needs small modification
to prove this. Let us explain this in what follows. The standard form (M y H, JM , P♮
M ) is
constructed from ψ so that JM ∆ψJM = ∆−1
ψ . The TrB is given by ψ ↾B. We need an extra
argument in relation to the d0 ∈ (1AJM 1BJM )hM, Bi(1AJM 1BJM ) ∩ A′. By the assumption
here the modular operator ∆ψ has a diagonalization ∆ψ = Pλ>0 λ eψ
ψ ∈
hM, Bi ∩ A′ for all t ∈ R. Hence all the eψ
0 with some
λ defines a non-zero element in hM, Bi ∩ A′. Since eψ
λ commutes with 1AJM 1BJM and since
λ 's fall in hM, Bi ∩ A′. Thus eψ
λ and satisfies ∆it
λ d1/2
NON-TRACIAL AMALGAMATED FREE PRODUCTS
7
λ ) = TrhM, Bi(d1/2
0 eψ
λ d1/2
0
λ d0 eψ
ψ ) ξ still falls in P♮
λ . Hence the A -- B bimodule K0 can be chosen as a subspace of eψ
TrhM, Bi(eψ
) ≤ TrhM, Bi(d0) < +∞, we may and do assume d0 =
λ H. Therefore, the ξ ∈
eψ
λ d0 eψ
ψ ) ξ = λit ξ for all t ∈ R. Since IMn+1(C) ¯⊗ ∆ψ
L2(M) = Mn+1(C) ¯⊗H satisfies that (IMn+1(C) ¯⊗ ∆it
is the modular operator of Trn+1 ¯⊗ ψ on M, (IMn+1(C) ¯⊗ ∆it
M, see [20, Lemma
IX.1.4]. Remark here that JM there is nothing but the one constructed from Trn+1 ¯⊗ ψ. Hence,
ψ ) ξ = ξ hold
by the uniqueness of polar decomposition (id ¯⊗σψ
for every t ∈ R. These modifications are enough to complete the proof.
(cid:3)
Remark 3.2. Let E B and ϕ = Tr B ◦ E B be as in the proof of (i) ⇒ (ii) above. Let cE B :
hM, Bi → M be the dual operator-valued weight associated with E B in the sense of [13, §§1.2].
It is known that the modular operator ∆ϕ and Connes's spacial derivative (d(ϕ◦ cE B ))/(d(Tr B ◦
AdJM ((−)∗))) must coincide, see e.g. the proof of [10, Proposition 2.2]. Moreover ∆ϕ is affiliated
with hM, Bi, since ϕ = Tr B ◦ E B. With these two facts one can prove that the modular operator
∆ϕ is the Radon -- Nikodym derivative of ϕ ◦ cE B , i.e., ϕ ◦ cE B = TrhM, Bi,∆ϕ
in the sense of [20,
Lemma VIII.2.8]. This explains, in full generality, the relationship that was pointed out in [17,
Eq.(1.3.1)] in the almost periodic case.
t )( w) = λit w and (IMn+1(C) ¯⊗ ∆it
3.2. A non-tracial version of Ioana -- Peterson -- Popa's theorem. Let us investigate an
amalgamated free product (M, E) = (M1, E1) ⋆N (M2, E2).
Proposition 3.3. Let A be a (unital) von Neumann subalgebra of the centralizer (M1)ϕ of a
certain faithful normal state ϕ, and M1 be a (possibly non-unital) dense (in any von Neumann
algebra topology) ∗-subalgebra of M1 with E1(M1) ⊆ M1. Suppose that there is a net vλ of
unitaries in A such that E1(y∗vλx) −→ 0 σ-strongly for all x, y ∈ M1. Then any unitary u ∈ M
with uAu∗ ⊆ M1 must fall in M1. In particular, NM (A) = NM1 (A) and A′ ∩ M = A′ ∩ M1.
Here NP (Q) denotes the set of unitaries u ∈ P with uQu∗ = Q for a given unital inclusion
P ⊇ Q of von Neumann algebras.
This is nothing but a non-tracial version of [9, Theorem 1.1] due to Ioana, Peterson and Popa.
Although the proof below is modeled after their proof, we need to overcome some difficulties
due to the lack of trace by utilizing modular theoretic technologies.
Proof. Let (M y H, JM , P♮
M be the unique repre-
senting vector of ϕ ◦ EM1 . Let eM1 be the so-called Jones projection associated with EM1 , i.e.,
eM1xξ0 = EM1 (x)ξ0 for x ∈ M , and the basic extension hM, M1i is defined to be M ∨{eM1}′′ =
1JM y H. Consider the projection p := [AJM M1JM u∗ξ0] ∈ A′ ∩ (JM M1JM )′ =
JM M ′
A′ ∩ hM, M1i. Notice that aJM x∗JM u∗ξ0 = JM x∗JM u∗(uau∗)ξ0 for a ∈ A and x ∈ M1,
and moreover that uau∗ ∈ M1 can be approximated in any von Neumann algebra topology, by
analytic elements, say yλ, in M1 with respect to the modular action σϕ. Those altogether show
that
M ) be a standard form of M , and ξ0 ∈ P♮
aJM x∗JM u∗ξ0 = lim
λ
JM x∗JM u∗yλξ0 = lim
λ
JM x∗σϕ
i/2(yλ)∗JM u∗ξ0 ∈ [JM M1JM u∗ξ0]
↾M1 = σϕ
t
ϕ◦EM1
thanks to σ
t
(t ∈ R) and [20, Lemma VIII.3.18 (ii)]. Consequently we get
p ≤ [JM M1JM u∗ξ0] = u∗eM1u, which and dEM1 (eM1) = 1 imply kdEM1 (p)k∞ < +∞, where
dEM1 : hM, M1i → M denotes the dual operator-valued weight of EM1 . See [13, §§1.2, Lemma
In fact, if this is the case, then p ≤ eM1 so
3.1]. We will prove (1 − eM1)p(1 − eM1) = 0.
that u∗ξ0 = eM1u∗ξ0 = EM1 (u∗)ξ0, implying u = EM1 (u) ∈ M1 since ξ0 is separating for
M y H. Since kdEM1(p)k∞ < +∞ and dEM1 (eM1 ) = 1 as before, any spectral projection f of
(1−eM1)p(1−eM1) corresponding to [δ, 1] with arbitrary δ > 0 still satisfies kdEM1 (f )k∞ < +∞.
8
Y. UEDA
Therefore, it suffices to prove that any projection f ∈ A′ ∩hM, M1i satisfying both f ≤ 1− eM1
and kdEM1 (f )k∞ < +∞ must be 0.
In what follows we denote by A the ∗-subalgebra of M consisting of all analytic elements
with respect to σϕ◦EM1 , which is well-known to be dense in any von Neumann algebra topol-
ogy. Set ψ := ϕ ◦ EM1 ◦ dEM1 , a faithful normal semifinite weight on hM, M1i, and let
hM, M1i y L2(hM, M1i, ψ) be the GNS representation with canonical embedding Λψ : nψ :=
{x ∈ hM, M1i ψ(x∗x) < +∞} → L2(hM, M1i, ψ) and norm k − kψ associated with the weight
t ◦ EM1 for all t ∈ R) and thus
ψ. Remark that EM1 (A) ⊆ A (thanks to EM1 ◦ σ
span(AeM1A) becomes a dense (in any von Neumann algebra topology) ∗-subalgebra of n∗
ψ∩nψ,
and hence Λψ(span(A eM1A)) is dense in L2(hM, M1i, ψ) by [10, Lemma 2.1]. Thus one can
choose a sequence Tn ∈ span(A eM1A) in such a way that kΛψ(Tn − f )kψ −→ 0 as n → ∞,
where note that f clearly falls in nψ. Since f ≤ 1 − eM1 and σψ
t (eM1 ) = eM1 (t ∈ R) [13,
Lemma 5.1], we also have kΛψ((1 − eM1 )Tn(1 − eM1) − f )kψ −→ 0 as n → ∞ so that may and
do assume that Tn = (1 − eM1)Tn(1 − eM1) for all n.
On contrary, suppose f 6= 0, that is, γ := kΛψ(f )kψ (cid:9) 0. Then one can choose T := Tn0 ∈
span(AeM1A) with some n0 in such a way that
ϕ◦EM1
t
= σϕ
kΛψ(T )kψ ≤ 3γ/2,
kΛψ(T − f )kψ ≤ γ/5.
(3.5)
For any v ∈ Au we compute
γ2 − ψ(T ∗vT v∗) ≤ ψ(f vf v∗) − ψ(T ∗vT v∗)
≤ ψ((f − T )∗vf v∗) + ψ(T ∗v(f − T )v∗)
≤ kΛψ(f − T )kψkΛψ(vf v∗)kψ + kΛψ(T )kψkΛψ(v(f − T )v∗)kψ
≤ kΛψ(f − T )kψkΛψ(f )kψ + kΛψ(T )kψkΛψ(f − T )kψ
≤ γ2/2,
where the first, the third, the fourth and the fifth inequalities follow from f ∈ A′ ∩ hM, M1i,
the Cauchy -- Schwarz inequality, v ∈ (M1)ϕ ⊂ hM, M1iψ, and (3.5), respectively. Therefore,
γ2 ≤ 2ψ(T ∗vT v∗) holds for all v ∈ Au. Since T = (1 − eM1)T (1 − eM1 ), we can write
T =Pm
k=1 xkeM1 yk with xk, yk ∈ A ∩ Ker(EM1 ). Thus, for every v ∈ Au we have
keM1x∗
γ2 ≤ 2Xm
= 2Xm
= 2Xm
= 2Xm
≤ 2 max
k,l=1 ψ(y∗
kEM1 (x∗
k,l=1 ψ(y∗
k,l=1 ϕ ◦ EM1 (y∗
k,l=1 ϕ ◦ EM1 (σ
1≤k≤mkykk∞ max
1≤l≤mkσ
kvxleM1ylv∗)
kvxl)eM1ylv∗)
kEM1 (x∗
kvxl)ylv∗)
(yl)v∗y∗
kEM1 (x∗
ϕ◦EM1
i
ϕ◦EM1
i
(yl)k∞Xm
kvxl))
k,l=1 kEM1 (x∗
kvxl)kϕ.
Here the third equality is due to dEM1 (eM1) = 1, the fourth one follows from v ∈ (M1)ϕ ⊆
and yl ∈ A with the so-called modular condition, and finally the last inequality is due
Mϕ◦EM1
to the Cauchy -- Schwarz inequality. Consequently we have chosen x1, . . . , xm ∈ A ∩ Ker(EM1 )
and a universal constant C > 0 so that
γ2 ≤ CXm
k,l=1 kEM1(x∗
kvxl)kϕ
for all v ∈ Au.
(3.6)
Set M◦
1 := M1 ∩ M ◦
1 . By the assumption on M1 and by the Kaplansky density theorem any
1 can be approximated in any von Neumann algebra topology by a bounded net of
element x ∈ M ◦
NON-TRACIAL AMALGAMATED FREE PRODUCTS
9
λ = xλ − E1(xλ) ∈ M◦
2 )\ M◦
1 with xλ ∈ M1, xλ −→ x. Thus M1 + span(Λ◦(M◦
elements x◦
1)
is also dense in M in any von Neumann algebra topology so that the Kaplansky density theorem
enables us to approximate each xk (= xk − EM1 (xk)) by a net xk,λ in span(Λ◦(M◦
2 )\ M◦
1);
namely kxk,λk∞ ≤ 2kxkk∞ and xk,λ −→ xk σ-∗-strongly. Then we have, for every v ∈ Au,
1, M ◦
1, M ◦
kEM1 (x∗
≤ kEM1(x∗
≤ kEM1(x∗
kvxl)kϕ ≤ kEM1(x∗
k,λvxl,λ)kϕ + kEM1 (x∗
k,λvxl,λ)kϕ + k(xk − xk,λ)∗vxl)kϕ◦EM1
k − x∗
k,λvxl,λ)kϕ + kσ
(xl)k∞kx∗
ϕ◦EM1
i/2
k,λvxl,λ)kϕ
kvxl − x∗
+ kxk,λv(xl − xl,λ)kϕ◦EM1
k,λkϕ◦EM1
+ 2kxkk∞kxl − xl,λkϕ◦EM1
,
where we used, in the last line, that xl ∈ A with [20, Lemma VIII.3.18 (ii)] and v ∈ (M1)ϕ.
Let ε > 0 be arbitrary chosen. Then some λ (being independent of v's) satisfies that γ2 ≤
ε + CPm
1 is
written as azb with a, b ∈ {1} ∪ M◦
2 whose leftmost and
rightmost letters are chosen from M ◦
j , i = 1, 2,
j = 1, . . . , m′, and positive constants Cj > 0, j = 1, . . . , m′, so that
k,λvxl,λ)kϕ for all v ∈ Au. Since any element in Λ◦(M◦
1, z an alternating word in M◦
j z(i)
2 , there are finitely many such words a(i)
k,l=1 kEM1(x∗
2 ) \ M◦
j b(i)
1, M ◦
1, M ◦
γ2 ≤ ε + Xm′
= ε + Xm′
j=1
j=1
CjkEM1 (a(1)
Cjka(1)
j z(1)
j b(1)
j va(2)
j z(2)
j b(2)
j )kϕ
j EM1 (z(1)
j E1(b(1)
j va(2)
j )z(2)
j
)b(2)
j kϕ
for all v ∈ Au, where the equality comes from the free independence of M1, M2 and (2.2).
Applying the above estimate of γ2 to the net v = vλ in our hypothesis we get γ2 ≤ ε (at the
limit in λ), a contradiction to γ (cid:9) 0, since ε is arbitrary.
(cid:3)
Remark 3.4. It is worth while to note that the inequality (3.6) is a general fact. Let P ⊇ Q
be σ-finite von Neumann algebras with a faithful normal conditional expectation EQ : P → Q
and A be a von Neumann subalgebra of the centralizer Qϕ with some faithful normal state ϕ.
The middle part of discussion above shows that for each projection f ∈ A′ ∩ hP, Qi satisfying
both f ≤ 1 − eQ and kcEQ(f )k∞ < +∞ there are analytic (with respect to σϕ◦EQ) elements
x1, . . . , xm ∈ P and a universal constant C > 0 such that
k,l=1 kEQ(x∗
for all v ∈ Au.
kΛϕ◦EQ◦ EQ
ϕ◦EQ◦ EQ ≤ CXm
kvxl)kϕ
(f )k2
3.3. A result for controlling central sequences in amalgamated free products. Let us
investigate central sequences in an amalgamated free product (M, E) = (M1, E1) ⋆N (M2, E2).
The next result is an adaptation and/or an improvement of the methods of [27, Proposition
3.5] and [28, Proposition 3.1] to amalgamated free product von Neumann algebras.
In this
subsection we use the notations and facts summarized in [27, §§2.2].
Proposition 3.5. Suppose that there is a faithful normal state ϕ on M1 satisfying the following
conditions:
(a) σϕ
(b) For every n ∈ N with n ≥ 2 there are unitaries uk = u(n)
t (N ) = N for all t ∈ R.
k ∈ (M1)ϕ, 0 ≤
k ≤ n − 1, such that E1(u∗
k1 vk2 ) = 0 for all 0 ≤ k1 6= k2 ≤ n − 1, where
Eϕ
N denotes the unique ϕ-preserving conditional expectation from M1 onto N , whose
existence follows from (a) and Takesaki's criterion.
k , vk = v(n)
k1 uk2) = Eϕ
N (v∗
ϕ ∩ M ω, any y ∈ M ◦
2 and any sequence (tm)m of real numbers we have
Then, for any x ∈ (M1)′
with z :=(cid:2)(σ
ϕ◦EM1
tm
kE2(y∗y)1/2(x − (EM1 )ω(x))k(ϕ◦EM1 )ω ≤ kyx − xzk(ϕ◦EM1 )ω ,
(y))m(cid:3) ∈ M ω.
10
Y. UEDA
Remark here that any bounded sequence (σ
(x(m)))m with arbitrary (x(m))m giving
an element in M ω gives again an element in M ω, as shown in the proof of [28, Proposition 3.1].
A key fact behind this is that any modular action σψ satisfies ψ ◦ σψ
t = ψ for all t ∈ R. In
particular, the element z in the statement above makes sense.
ϕ◦EM1
tm
2
M ◦
1 , M ◦
1 or Eϕ
1 , M ◦
1 , that span(Λ◦(M ◦
Proof. Write M ▽
N + M ◦
linear span of the following sets of words:
1 := Ker(Eϕ
N (x) + (x − Eϕ
1 ··· M ◦
M ▽
}
{z
2 ··· M ◦
N ). One can easily see, by using x ∈ M1 7→ E1(x) + (x − E1(x)) ∈
N (x)) ∈ N + M ▽
1 ) coincides with the
1 ··· M ◦
}
{z
Define four closed subspaces X1 := (cid:2)Λϕ◦EM1
X3 := (cid:2)Λϕ◦EM1
Denote by Pi, i = 1, 2, 3, 4, the projection from H onto Xi. Remark that
2 ) \ M ◦
2 ··· M ◦
}
{z
2 )(cid:3),
1 )(cid:3), X2 := (cid:2)Λϕ◦EM1
1 ··· M ◦
2 )(cid:3) in H := L2(M, ϕ ◦ EM1 ), and
2 ··· M ◦
}
{z
1 ··· M ◦
2 M ▽
1 )(cid:3), X4 := (cid:2)Λϕ◦EM1
(M ◦
2 ··· M ◦
H = Λϕ◦EM1
(M1) ⊕ X1 ⊕ X2 ⊕ X3 ⊕ X4.
1 , M ◦
, M ◦
clearly
2 M ▽
alternating
alternating
alternating
alternating
(M ◦
(M ◦
(M ◦
M ▽
2
2
2
.
(cid:18)IH − X4
Pi(cid:19) Λϕ◦EM1
i=1
x ∈ M.
Let n ∈ N with n ≥ 2 be fixed. Define unitary operators Sk = S(n)
(vkxv∗
0, . . . , n − 1) on H by
(x) = Λϕ◦EM1
(EM1 (x)),
(ukxu∗
(x) := Λϕ◦EM1
SkΛϕ◦EM1
k),
(3.7)
(k =
k , Tk = T (n)
k
TkΛϕ◦EM1
(x) := Λϕ◦EM1
x ∈ M,
in our hypothesis. Here are simple claims.
k),
with uk = u(n)
k , vk = v(n)
k ∈ (M1)ϕ ⊆ Mϕ◦EM1
(A) {SkXi}n−1
(B) {TkX2}n−1
k=0 is an orthogonal family of closed subspaces, i = 3, 4.
k=0 is an orthogonal family of closed subspaces.
The proofs of those are essentially same, but (A) is easier than (B). Thus we prove only (B)
here and leave (A) to the reader. By using x 7→ Ei(x) + (x − Ei(x)) ∈ N + M ◦
i (i = 1, 2) again
and again we have
(vk2 (M ◦
k2 )∗(vk1 (M ◦
1 ··· M ◦
= vk2 (M ◦
2 )v∗
2 ··· (M ◦
1 ··· M ◦
2 )v∗
k1 )
1 )··· M ◦
2 )v∗
1 v∗
k2 vk1 M ◦
k1 ⊆ vk2 N v∗
k1 + vk2 Ker(EM1 )v∗
k1 .
The desired assertion immediately follows from that vk ∈ (M1)ϕ; in fact, if k1 6= k2, then
ϕ ◦ EM1 (vk2 N v∗
ϕ ◦ EM1 (vk2 Ker(EM1 )v∗
k1 ) = ϕ(N v∗
k1 vk2 ) = ϕ(N Eϕ
N (v∗
k1 vk2 )) = {0},
k1 ) = ϕ(vk2 EM1 (Ker(EM1 ))v∗
k1 ) = {0}.
Let us choose arbitrary x ∈ (M1)′
ϕ ∩ M ω with representative (x(m))m. For each ε > 0 and
each n ∈ N with n ≥ 2 one can choose a neighborhood W = Wε,n in β(N) at ω so that
(x(m) − vkx(m)v∗
k , vk = v(n)
k
k)kH < ε
are as above. For each
kΛϕ◦EM1
(x(m) − ukx(m)u∗
k)kH < ε,
kΛϕ◦EM1
for all 0 ≤ k ≤ n − 1 and m ∈ W ∩ N, where the uk = u(n)
i = 3, 4 and every m ∈ W ∩ N we have, with the above Sk = S(n)
k ,
kPiΛϕ◦EM1
nXn−1
=
nXn−1
k=0 kSkPiΛϕ◦EM1
k=0nkSkPiΛϕ◦EM1
(x(m))k2
(x(m)) − SkPiS∗
(x(m))k2
(x(m))k2
kΛϕ◦EM1
≤
1
2
H
H
H + kSkPiS∗
kΛϕ◦EM1
Ho
(x(m))k2
NON-TRACIAL AMALGAMATED FREE PRODUCTS
11
(ukx(m)u∗
k − x(m))k2
H +
2
nXn−1
k=0 kSkPiS∗
kΛϕ◦EM1
(x(m))k2
H
=
2
nXn−1
< 2ε2 +
kΛϕ◦EM1
k=1 kSkPiS∗
nXn−1
2
2
nkΛϕ◦EM1
k=0 kSkPiS∗
(x(m))k2
∞/n.
≤ 2ε2 +
≤ 2ε2 + 2k((x(m))mk2
Similarly, using the claim (B) with T (n)
k
kΛϕ◦EM1
(x(m))k2
H
H (by the claim (A))
kP2Λϕ◦EM1
(x(m))k2
k we have
instead of S(n)
H < 2ε2 + 2k((x(m))mk2
∞/n
for every m ∈ W ∩ N. Since n and ε are arbitrary, for each δ > 0 one can find a neighborhood
Wδ in β(N) at ω so that
k(P2 + P3 + P4)Λϕ◦EM1
(x(m))kH < δ
(3.8)
for all m ∈ Wδ ∩ N.
In the standard embedding L2(M ω, (ϕ ◦ EM1 )ω) ֒→ Hω we have, by (3.7) and (3.8),
[Dϕ ◦ EM1 : Dχ ◦ E]t σχ◦E
t
(y) [Dϕ ◦ EM1 : Dχ ◦ E]∗
t
e−t2/ℓ dt
√ℓπ
with a fixed faithful normal state χ on N . Clearly yℓ falls in the σ-weak (or σ-strong) closure of
span(M1M ◦
2 M1), since [Dϕ ◦ EM1 : Dχ ◦ E]t = [Dϕ : Dχ ◦ E1]t ∈ M1 by [20, Corollary IX.4.22
(yℓ))m(cid:3) ∈ M ω, which is well-defined as
(ii)] and σχ◦E
remarked just before the proof. Note that σ
each ℓ we have, by (3.7), (3.8) as before and by [20, Lemma VIII.3.18 (ii)],
2 by (2.1). Set zℓ := (cid:2)(σ
ϕ◦EM1
tm
ϕ◦EM1
−i/2
(y) ∈ M ◦
(yℓ)). For
(yℓ)) = σ
ϕ◦EM1
−i/2
ϕ◦EM1
tm
ϕ◦EM1
tm
(σ
(σ
t
= lim
(cid:13)(cid:13)(cid:13)Λ(ϕ◦EM1 )ω ((x − (EM1 )ω(x))zℓ) −(cid:2)(Jσ
m→ω(cid:13)(cid:13)Jσ
(yℓ))∗J(cid:0)Λϕ◦EM1
m∈Wδ ∩N(cid:13)(cid:13)Jσ
≤ sup
ϕ◦EM1
−i/2
ϕ◦EM1
−i/2
ϕ◦EM1
tm
ϕ◦EM1
tm
(σ
(σ
ϕ◦EM1
−i/2
(σ
ϕ◦EM1
tm
(yℓ))∗JP1Λϕ◦EM1
(x(m) − EM1 (x(m))) − P1Λϕ◦EM1
(x(m)))m(cid:3)(cid:13)(cid:13)(cid:13)Hω
(x(m))(cid:1)(cid:13)(cid:13)H
ϕ◦EM1
−i/2
(yℓ)k∞δ
(x(m))(cid:13)(cid:13)H < kσ
(yℓ))∗J(P2 + P3 + P4)Λϕ◦EM1
(x(m)))m(cid:3)(cid:13)(cid:13)(cid:13)Hω
(x(m))(cid:13)(cid:13)H
= lim
= lim
(y(x(m) − EM1 (x(m)))) − yP1Λϕ◦EM1
(cid:13)(cid:13)(cid:13)Λ(ϕ◦EM1 )ω (y(x − (EM1 )ω(x))) −(cid:2)(yP1Λϕ◦EM1
m→ω(cid:13)(cid:13)Λϕ◦EM1
(x(m))(cid:13)(cid:13)H
m→ω(cid:13)(cid:13)y(P2 + P3 + P4)Λϕ◦EM1
(x(m))(cid:13)(cid:13)H < kyk∞δ,
m∈Wδ ∩N(cid:13)(cid:13)y(P2 + P3 + P4)Λϕ◦EM1
Λ(ϕ◦EM1 )ω (y(x − (EM1 )ω(x))) =(cid:2)(yP1Λϕ◦EM1
≤ sup
(x(m)))m(cid:3)
(yEM1 (x(m)) − EM1 (x(m))σ
ϕ◦EM1
tm
(3.9)
(3.10)
(y)))m(cid:3).
and hence
in Hω, since δ is arbitrary. Trivially, in Hω,
Λ(ϕ◦EM1 )ω (y(EM1 )ω(x) − (EM1 )ω(x)z)
=(cid:2)(Λϕ◦EM1
Set
ϕ◦EM1
σ
t
(y)
e−t2/ℓ dt
√ℓπ
yℓ :=Z +∞
=Z +∞
−∞
−∞
12
Y. UEDA
with the modular conjugation J of M y H = L2(M, ϕ ◦ EM1 ). Hence, for each ℓ,
Λ(ϕ◦EM1 )ω ((x − (EM1 )ω(x))zℓ)
=(cid:2)(Jσ
in Hω, since δ is arbitrary. Note that
ϕ◦EM1
−i/2
(σ
ϕ◦EM1
tm
(yℓ))∗JP1Λϕ◦EM1
(x(m)))m(cid:3)
(3.11)
On the other hand,
yP1Λϕ◦EM1
(xm)) ∈ spanΛϕ◦EM1
(M ◦
2 M ◦
1 ··· M ◦
2 M ▽
1 ).
Λϕ◦EM1
and
(yEM1 (x(m)) − EM1 (x(m))σ
∈ spanΛϕ◦EM1
(y))
ϕ◦EM1
tm
(M ◦
⊕ spanΛϕ◦EM1
2 ) ⊕ spanΛϕ◦EM1
(M ◦
(M ◦
2 ) ⊕ spanΛϕ◦EM1
1 M ◦
2 M ▽
1 )
(M ◦
1 M ◦
2 M ▽
1 )
Jσ
ϕ◦EM1
−i/2
(σ
ϕ◦EM1
tm
(x(m))
(yℓ))∗JP1Λϕ◦EM1
1 ··· M ◦
(M ◦
∈ spanΛϕ◦EM1
⊆ Λϕ◦EM1
2 M ▽
1 σ
(M1) ⊕ spanΛϕ◦EM1
ϕ◦EM1
tm
(M ◦
1 M ◦
(yℓ))
2 ··· ).
(yℓ) falls in the σ-
Here the last fact follows from [20, Lemma VIII.3.18 (ii)] and that σ
strong closure of span(M1M ◦
2 M1). Therefore, we see, by (3.9) -- (3.11), that Λ(ϕ◦EM1 )ω (y(x −
(EM1 )ω(x))) is orthogonal to both Λ(ϕ◦EM1 )ω (y(EM1 )ω(x) − (EM1 )ω(x)z) and Λ(ϕ◦EM1 )ω ((x −
−tm (y))m(cid:3), both of which
(EM1 )ω(x))zℓ). Finally, letting x := (cid:2)(σ
−tm (x(m)))m(cid:3), y :=(cid:2)(σ
fall in M ω as remarked just before the proof, we have
ϕ◦EM1
ϕ◦EM1
ϕ◦EM1
tm
(cid:0)Λ(ϕ◦EM1 )ω ((x − (EM1 )ω(x))z)Λ(ϕ◦EM1 )ω (y(x − (EM1 )ω(x)))(cid:1)(ϕ◦EM1 )ω
= (ϕ ◦ EM1 )ω((x − (EM1 )ω(x))∗y∗(x − (EM1 )ω(x))z)
= (ϕ ◦ EM1 )ω((x − (EM1 )ω(x))∗ y∗(x − (EM1 )ω(x))y)
= lim
ℓ→∞
= lim
ℓ→∞
= lim
(ϕ ◦ EM1 )ω((x − (EM1 )ω(x))∗ y∗(x − (EM1 )ω(x))yℓ)
(ϕ ◦ EM1 )ω((x − (EM1 )ω(x))∗y∗(x − (EM1 )ω(x))zℓ)
ℓ→∞(cid:0)Λ(ϕ◦EM1 )ω ((x − (EM1 )ω(x))zℓΛ(ϕ◦EM1 )ω (y(x − (EM1 )ω(x)))(cid:1)(ϕ◦EM1 )ω = 0.
Consequently we get ky(x − (EM1 )ω(x))k(ϕ◦EM1 )ω ≤ kyx − xzk(ϕ◦EM1 )ω . We have, by (3.9),
ky(x − (EM1 )ω(x))k2
(ϕ◦EM1 )ω
(x(m)))m(cid:3)(cid:13)(cid:13)(cid:13)
=(cid:13)(cid:13)(cid:13)(cid:2)(yP1Λϕ◦EM1
m→ωn(E2(y∗y)P1Λϕ◦EM1
(yP1Λϕ◦EM1
= lim
m→ω
= lim
2
Hω
(x(m))yP1Λϕ◦EM1
(x(m)))ϕ◦EM1
(x(m)))ϕ◦EM1
+ ((y∗y − E2(y∗y))P1Λϕ◦EM1
in X3 orthogonal to X1
(x(m))
P1Λϕ◦EM1
{z
in X1
}
(x(m)))ϕ◦EM1
(x(m))
)ϕ◦EM1o
}
(x(m))P1Λϕ◦EM1
{z
(x(m)))m(cid:3)(cid:13)(cid:13)(cid:13)
(x(m))P1Λϕ◦EM1
.
Hω
2
= lim
m→ω
(E2(y∗y)P1Λϕ◦EM1
=(cid:13)(cid:13)(cid:13)(cid:2)(E2(y∗y)1/2P1Λϕ◦EM1
NON-TRACIAL AMALGAMATED FREE PRODUCTS
As in showing (3.9) one has
(cid:2)(E2(y∗y)1/2P1Λϕ◦EM1
and the proof is completed.
(x(m)))m(cid:3) = iΛ(ϕ◦EM1 )ω (E2(y∗y)1/2(x − (EM1 )ω(x))),
13
(cid:3)
4. Some Consequences
We first formulate that P is 'non-trivial relative to Q' for a given inclusion of von Neumann
algebras P ⊇ Q, and then provide some technical facts.
Definition 4.1. A (unital) inclusion P ⊇ Q of von Neumann algebras is said to be entirely
non-trivial, if no non-zero direct summand of Q is a direct summand of P .
Let P ⊇ Q be an inclusion of von Neumann algebras with a faithful normal conditional
expectation EQ. If zP = Qz (as set) for some non-zero z ∈ Z(Q)p, then P z = Qz too by
taking adjoints, and thus for x ∈ P one has zx = EQ(zx) = zEQ(x) = EQ(x)z = EQ(xz) = xz,
implying z ∈ Z(P ). Hence Qz is a direct summand of P . Therefore, P ⊇ Q is entirely non-
trivial if and only if P z 6= Qz or equivalently zP 6= Qz for any non-zero projection z ∈ Z(Q),
where P z and zP denote the one-sided ideals of all xz and zx, respectively, with x ∈ P .
The next simple lemma, especially (3) there, will frequently be used later.
again entirely non-trivial.
i ≤ e, and then P ei = P v∗
e =Pi∈I xei =Pi∈I EQ(x)ei = EQ(x)cQ
e , and therefore P cQ
e = QcQ
e .
i yi)) = 1, where s(x) denotes the support projection of x = x∗.
Lemma 4.2. Let P ⊇ Q be an inclusion of von Neumann algebras with a faithful normal
conditional expectation EQ : P → Q.
(1) The following are equivalent:
(i) P ⊇ Q is entirely non-trivial.
(ii) P e 6= Qe or equivalently eP 6= eQ for any non-zero projection e ∈ Q.
(2) If P ⊇ Q is entirely non-trivial and f ∈ Q a projection with cQ
f = 1, then f P f ⊇ f Qf is
(3) If P ⊇ Q is entirely non-trivial, then there is a family {yi}i∈I of elements in Ker(EQ)
so that Pi∈I s(EQ(y∗
Proof. (1) By the discussion above (i) is equivalent to P z 6= Qz or equivalently zP 6= Qz for
any non-zero z ∈ Z(Q)p. Thus (ii) ⇒ (i) is trivial, and it suffices to show (i) ⇒ (ii). Suppose
that P e = Qe for some non-zero e ∈ Qp. By a standard exhaustion argument based on the
comparison theorem we can choose an orthogonal family {ei}i∈I of projections in Q such that
ei - e in Q for all i ∈ I and cQ
i vi = ei
and viv∗
i vi ⊆ P evi = Qevi ⊆ Qei, implying P ei = Qei ⊆ Q. For
x ∈ P one has xcQ
(2) By (1) it suffices to prove that eP f 6= eQf for any non-zero e ∈ Qp with e ≤ f . As in
(1) one can find an orthogonal family {fi}i∈I of projections in Q such that fi - f in Q for all
i ∈ I and Pi∈I fi = cQ
f = 1. On contrary, suppose that eP f = eQf for some non-zero e ∈ Qp
with e ≤ f . Then one has eP fi = eQfi in the same way as in (1). Hence, as in the above (1)
one can justify, by using EQ, the following computation: eP =Pi∈I eP fi =Pi∈I eQfi = eQ,
a contradiction to the entire non-triviality of P ⊇ Q thanks to (1).
(3) Choose a maximal (with respect to set-inclusion) family {yi}i∈I of elements in Ker(EQ) so
that {s(EQ(y∗
i yi)) 6=
1. Set e := 1 −Pi∈I s(EQ(y∗
i yi)) ∈ Qp \ {0}. Since P ⊇ Q is entirely non-trivial, one has
P e 6= Qe by (1), and hence can choose x ∈ P with xe 6∈ Q. Hence xe − EQ(xe) 6= 0 and
set y := xe − EQ(xe) ∈ Ker(EQ). Clearly, ye = y, and thus EQ(y∗y) = eEQ(y∗y)e, implying
s(EQ(y∗y)) ≤ e = 1 −Pi∈I s(EQ(y∗
(cid:3)
i yi))}i∈I is an orthogonal family of projections in Q. SupposePi∈I s(EQ(y∗
e = Pi∈I ei. Choose a partial isometry vi ∈ Q with v∗
i yi)), a contradiction to the maximality of {yi}i∈I .
14
Y. UEDA
Let (M, E) = (M1, E1) ⋆N (M2, E2) be an amalgamated free product throughout the rest of
this section.
Theorem 4.3. Assume that there is a faithful normal state ϕ on M1 such that one can find
a (possibly non-unital) dense (in any von Neumann algebra topology) ∗-subalgebra M1 of M1
with E1(M1) ⊆ M1 and a net vλ of unitaries in the centralizer (M1)ϕ in such a way that
E1(y∗vλx) −→ 0 σ-strongly for all x, y ∈ M1. Assume also that M2 ⊇ N is entirely non-trivial.
Then we have:
t
t
(0) ((M1)ϕ)′ ∩ M = ((M1)ϕ)′ ∩ M1.
(1) Z(M ) = Z(M1) ∩ Z(M2) ∩ Z(N ).
(2) Let χ be an arbitrary faithful normal semifinite weight on N . Then, if a unitary u
= Adu for some t ∈ R, then u must fall in N . In particular,
= Adu = σχ◦E2
(3) M is semifinite if and only if there is a faithful normal semifinite trace TrN such that
in M satisfies σχ◦E
T (M ) = {t ∈ R σχ◦E1
both TrN ◦ E1 and TrN ◦ E2 are traces.
for some u ∈ N u}.
t
ϕ◦EM1
(2) One has σ
t
Proof. (0) is nothing but what Proposition 3.3 says.
(4) Z(fM ) = Z(fM1) ∩ Z(fM2) ∩ Z(eN ).
(1) Let x ∈ Z(M ) be arbitrary, and then x must be in M1 by (0). For any y ∈ M ◦
2 one
has y(x − E1(x)) + yE1(x) = yx = xy = E1(x)y + (x − E1(x))y, and thus {yE1(x), E1(x)y},
y(x − E1(x)) and (x − E1(x))y are orthogonal with respect to E due to the free independence
between M1 and M2. Thus y(x − E1(x)) = 0 so that (by looking at the E-value of the
product of its adjoint and itself) we get (x − E1(x))∗E2(y∗y)(x − E1(x)) = 0. Therefore,
E2(y∗y)(x − E1(x)) = 0 for all y ∈ M ◦
2 . By taking its adjoint one can easily see that (x −
E1(x))∗ ↾ran(E2(y∗y))≡ 0 so that (x − E1(x))∗s(E2(y∗y)) = 0 for all y ∈ M ◦
2 . By Lemma 4.2 (3)
2 so that Pi∈I s(E2(y∗
one can find a family {yi}i∈I of elements in M ◦
i yi)) = 1, which implies
x = E1(x) ∈ N . The desired assertion is now immediate.
= Ad([Dϕ : Dχ◦ E1]t u) by Connes's Radon -- Nikodym cocycle theorem
, we have [Dϕ : Dχ ◦ E1]t u ∈ M1 by
1 . For y ∈ M ◦
(y)(u −
(y)u = uy = E1(u)y + (u − E1(u))y, and as in (1) we get
2 by (2.1). The same argument as in (1) again
and [20, Corollary IX.4.20]. Since (M1)ϕ ⊆ Mϕ◦EM1
(0). In particular, u ∈ M1, since [Dϕ : Dχ ◦ E1]t ∈ M u
E1(u)) + σχ◦E
(u− E1(u))y = 0, since σχ◦E
shows u = E1(u) ∈ N . The T-set computation is straightforward.
(3) M is semifinite if and only if there is a 1-parameter unitary group u(t) in M so that
σχ◦E
= Adu(t), t ∈ R, for a fixed faithful normal state χ on N . See [20, Theorem VIII.3.14].
t
Then u(t) ∈ N by (2). By Stone's theorem u(t) = H it with some positive non-singular, self-
adjoint H affiliated with N . Since σχ
(u(t)) = u(t), H must indeed be affiliated
with the centralizer Nχ. Hence, by [20, Lemma VIII.2.8] we can construct a faithful normal
semifinite weight χH−1 on N , and by the construction we observe that χH−1 ◦ E = (χ◦ E)H−1 .
χH−1 ◦E
(−)H it = id. Hence the
Moreover, by [20, Lemma VIII.2.11] we have σ
t
χH−1 is a desired faithful normal semifinite trace on N .
(4) By (0) together with the same argument as in [22, Corollary 4] we observe that ((M1)ϕ)′∩
(M ⋊σϕ◦EM1 R) = ((M1)ϕ)′∩(M1 ⋊σϕ R), where (M1)ϕ ⊂ M1 ⊆ M ֒→ M ⋊σϕ◦EM1 R canonically
as in §2. It follows that (fM1)′ ∩fM = Z(fM1), where we need Connes's Radon -- Nikodym cocycle
theorem together with [20, Theorem X.1.7]. Choose an arbitrary x ∈ Z(fM ). Then x must fall in
Z(fM1) ⊆ fM1. For y ∈ M ◦
2 one has y(x−eE(x))+yeE(x) = yx = xy = eE(x)y+(x−eE(x))y,
and thus y(x − eE(x)) = 0 since fM1, fM2 are freely independent with respect to eE as remarked
2 ⊂ fM ◦
(y)E1(u) = σχ◦E
t
(y) = σχ◦E2
t
2 we have σχ◦E
t
t (u(t)) = σχ◦E
t
= H −itσχ◦E
t
t
t
(y) ∈ M ◦
NON-TRACIAL AMALGAMATED FREE PRODUCTS
15
2 as in (1). Therefore, using
(cid:3)
in §2. In particular, we get E2(y∗y)(x − eE(x)) = 0 for all y ∈ M ◦
Lemma 4.2 (3) as in (1) once again we can prove x = eE(x) ∈ eN . Hence we are done.
Let us illustrate how the above theorem is useful by giving next two corollaries. The first
corollary shows that Proposition 3.1 is useful to confirm the necessary hypothesis of the theorem.
The second one does that the theorem is still applicable beyond the case where N is semifinite.
Remark that the first one can be viewed as a simultaneous generalization of both [27, Theorem
3.4] and [21, §4].
Corollary 4.4. Assume that M1 is diffuse, N of type I and M2 ⊇ N entirely non-trivial. Let
z ∈ Z(N ) be the unique projection so that N z is diffuse and N z⊥ atomic, and assume further
that M1cM1
has no type I direct summand when z 6= 0 (i.e., this last assumption is fulfilled if
M1 has no type I direct summand). Then all the assertions of Theorem 4.3 holds with a certain
faithful normal state ϕ on M1.
z
t
z
Proof. Let us fix a faithful normal semifinite trace TrN on N . Write c := cM1
for simplicity.
Clearly σTrN ◦E1
(c) = c for all t ∈ R, and thus Takesaki's criterion shows that there is a TrN ◦E1-
preserving unique conditional expectation EL : M1 → L := N ∨ {c}′′ = N c ⊕ N c⊥ (⊇ N ). In
particular, one observes that E1 ◦ EL = E1 holds. As in the proof of [27, Theorem 3.4] one can
choose a faithful normal state ϕ on M1 such that (M1c)ϕ↾M1c has no type I direct summand
and (M1c⊥)ϕ↾M1c⊥ is just only diffuse. Then it is clear that (M1c)ϕ↾M1c 6(cid:22)M1c N c with EL ↾M1c
and TrN ◦ E1 ↾N c and that (M1c⊥)ϕ↾
M1c⊥ 6(cid:22)M1c⊥ N c⊥ with EL ↾M1c and TrN ◦ E1 ↾N c⊥, since
N c⊥ = (N z⊥)c⊥ is a reduced von Neumann algebra of the atomic part N z⊥. Therefore, by
the equivalent condition (i) in Proposition 3.1 there are two nets v(1)
of unitaries in
λ
2v(2)
(M1c)ϕ↾M1c and (M1c⊥)ϕ↾M1c⊥ , respectively, so that EL(y∗
λ x2) −→ 0
σ-strongly for all x1, y1 ∈S{M1p p ∈ (N c)p; TrN ◦ E1(p) < +∞} and all x2, y2 ∈S{M1p p ∈
(N c⊥)p; TrN ◦ E1(p) < +∞}. Remark that EL = (EL ↾M1c)⊕ (EL ↾M1c⊥) in M1 = M1c⊕ M1c⊥
and that TrN (p) < +∞ implies both TrN ◦ E1(pc) < +∞ and TrN ◦ E1(pc⊥) < +∞ for
p ∈ N p. Thus, letting vλ := v(1)
M1c⊥ = (M1)ϕ one has, for all
x, y ∈ S{M1p p ∈ N p; TrN (p) < +∞}, EL(y∗vλx) −→ 0 σ-strongly and hence E1(y∗vλx) =
E1(EL(y∗vλx)) −→ 0 σ-strongly. Hence we can apply Theorem 4.3 with the above ϕ and
M1 := S{pM1p p ∈ N p; TrN (p) < +∞}. Note here that M1 is indeed a ∗-algebra thanks to
the Kaplansky formula [12, Theorem 6.1.7] and dense in any von Neumann algebra topology
due to the semifiniteness of TrN .
(cid:3)
and v(2)
λ
λ x1) −→ 0 and EL(y∗
λ ∈ (M1c)ϕ↾M1c ⊕ (M1c⊥)ϕ↾
λ ⊕ v(2)
1v(1)
Corollary 4.5. Assume that (M1, E1) is one of the following: (i) M1 = N ⋊α G and E1 is
the canonical conditional expectation from M1 = N ⋊α G onto N , where α : G y N is an
infinite discrete group action preserving a faithful normal state ψ on N . (ii) M1 = Q ¯⊗N and
E1 = ψ ¯⊗idN , where Q is a diffuse von Neumann algebra with a faithful normal state ψ. Assume
also that M2 ⊇ N is entirely non-trivial. Then all the assertions of Theorem 4.3 holds with
ϕ = ψ ◦ E1 in (i) and with ϕ = ϕ0 ¯⊗χ in (ii), where Qϕ0 is diffuse (such a state ϕ0 certainly
exists) and χ arbitrary.
Proof. Case (i): Since ψ is invariant under the action α, the restriction (ψ ¯⊗idB(ℓ2(G))) ↾N ⋊αG
gives a faithful normal conditional expectation from Eψ : M1 = N ⋊α G → L(G) = C1 ⋊ G, and
it is plain to see that ψ ◦ E1 := τG ◦ Eψ with the canonical tracial state τG on L(G). Clearly
L(G) = C1 ⋊ G sits inside (N ⋊α G)ψ◦E1 and is diffuse (see e.g. [6, Proposition 5.1]). With
ϕ := ψ ◦ E1 = τG ◦ Eψ and M1 := span{xλg x ∈ N, g ∈ G} one can choose a net vλ from
L(G) = C ⋊ G as in Theorem 4.3, since L(G) is diffuse and E1 ↾L(G)=C1⋊G= τG(−)1.
16
Y. UEDA
Case (ii): As in the proof of [27, Theorem 2.4] one can choose a faithful normal state ϕ0 on
Q in such a way that the centralizer Qϕ0 is diffuse. Set ϕ := ϕ0 ¯⊗χ with a faithful normal state
χ on N and M1 := Q ⊙ N = span{x ⊗ y x ∈ Q, y ∈ N}. Then one can choose a net vλ from
Qϕ0
(cid:3)
¯⊗C1 as in Theorem 4.3, since Qϕ0 is diffuse.
The next lemma seems well-known, but we do give it for the reader's convenience as a
reference for the discussions below.
Lemma 4.6. Let (P, F ) = (P1, F1) ⋆Q (P2, F2) be an amalgamated free product. If a projec-
tion f ∈ Q has cQ
f = 1, then (f P f, F ↾f P f ) = (f P1f, F1 ↾f P1f ) ⋆f Qf (f P2f, F2 ↾f P2f ) holds
canonically.
viv∗
1 , P ◦
i1 )(vi1 x2v∗
Proof. Clearly f P1f and f P2f are freely independent with respect to F ↾f P f , and hence
it suffices to see that those generate f P f as von Neumann algebra. As in the proof of
i vi = cQ
f = 1 and
2 ) one has f xf =
i2 )··· (vin−1 xnf ) σ-strongly, which falls in the σ-strong closure of
1 , P ◦
2 ),
(cid:3)
Lemma 4.2 one can find partial isometries {vi}i∈I in Q such that Pi∈I v∗
i ≤ f for all i ∈ I. For any alternating word x = x1 ··· xn ∈ Λ◦(P ◦
Pi1,...,in−1∈I(f x1v∗
the linear span of Λ◦((f P1f )◦, (f P2f )◦)). Since P is the σ-strong closure of Q+spanΛ◦(P ◦
the assertion is immediate.
Lemma 4.7. Let P ⊇ Q be an inclusion of σ-finite von Neumann algebras with a faithful
normal conditional expectation EQ : P → Q, and assume that Q is commutative.
(1) If P has no type I direct summand and a faithful normal semifinite trace TrP on P with
TrP ◦ EQ = TrP , then there is a faithful normal state χ on Q so that for each n ∈ N with n ≥ 2
one can find a unitary un ∈ Pχ◦EQ in such a way that EQ(uk
n) = 0 for all 1 ≤ k ≤ n − 1, i.e.,
EQ(uk1
n
(2) If P is diffuse and Q is atomic, then there is a faithful normal state ϕ on P such that
(a) the centralizer Pϕ contains Q,
(b) there are two unitaries u, v ∈ Pϕ so that EQ(uk) = Eϕ
Q(vk) = 0 as long as k 6= 0,
i.e., EQ(uk1 ∗ uk2 ) = Eϕ
Q denotes the unique
ϕ-preserving conditional expectation from P onto Q whose existence follows from (a)
and Takesaki's criterion.
Q(vk1 ∗ vk2 ) = 0 for all k1 6= k2. Here Eϕ
n ) = 0 for all 0 ≤ k1 6= k2 ≤ n − 1.
∗ uk2
(3) Let z ∈ Z(P ) be the central support projection of the type I direct summand of P . Assume
that P is diffuse and Qz atomic. Then there is a faithful normal state ϕ on the continuous core
eP of P such that
(a) the centralizer (eP )ϕ contains eQ, where eQ = Q ⋊σχ R ֒→ eP = P ⋊σχ◦EQ R with a faithful
(b) for each n ∈ N with n ≥ 2 one can find a unitary un ∈ (eP )ϕ in such a way that
normal state or semifinite weight χ on Q,
∗ uk2
∗ vk2
n
(vk1
n
n ) = Eϕ
n ) = 0 for
eQ
denotes the unique
(uk
n) = Eϕ
eQ
n) = 0 for all 1 ≤ k ≤ n − 1, i.e., eEQ(uk1
eEQ(uk
all 0 ≤ k1 6= k2 ≤ n − 1. Here eEQ = (EQ ¯⊗idB(L2(R))) ↾ eP , and Eϕ
ϕ-preserving conditional expectation from eP onto eQ as in (2).
eQ
The same assertion also holds for P ⊇ Q with EQ themselves, if it is further assumed that P
is semifinite and EQ preserves a faithful normal semifinite trace TrP on P .
Proof. (1) By assumption TrP ↾Q is semifinite, and thus one can choose an orthogonal sequence
{qm}m of projections in Q with TrP (qm) < +∞ and Pm∈N qm = 1. Consider the faithful
normal state χ := Pm∈N
m∈N Qqm = Q(cid:1) so that Pχ◦EQ
Clearly the centralizer Pχ◦EQ contains P⊕
2mTrP (qm) TrP ↾Qqm on Q. (Remark here that Q is commutative.)
m∈N qmP qm (cid:0) ⊇ P⊕
1
NON-TRACIAL AMALGAMATED FREE PRODUCTS
17
n)) = 0.
n) = EQ(EA(uk
n) = 0 so that EQ(uk
m∈N Cqm. Clearly EQ factors as P
Since P is diffuse, so are all qmP qm; hence by the proof of [27, Theorem 3.4] there are faith-
must be of type II1. Choose a MASA A in Pχ◦EQ that contains Q. By [11, Corollary 3.16], for
each n ∈ N with n ≥ 2 there are n orthogonal e0, . . . , en−1 ∈ Ap, all of which are equivalent
in Pχ◦EQ , and Pn−1
i=0 ei = 1. Then one can construct a unitary un ∈ Pχ◦EQ such that une0 =
e1un, une1 = e2un, . . . , unen−1 = e0un. Let EA : P → A be the χ ◦ EQ-preserving conditional
expectation (whose existence follows from Takesaki's criterion), and clearly EQ ◦ EA = EQ.
Then, for every 1 ≤ k ≤ n − 1 one has EA(uk
(2) Write Q = P⊕
EQ′ ∩P−→ Q′ ∩ P Ψ−→ Q, where Q′ ∩
m∈N qmP qm and EQ′∩P (x) = Pm∈N qmxqm for x ∈ P . Moreover Ψ is of the form
P = P⊕
Ψ(Pm∈N xm) =Pm∈N ψm(xm)qm for xm ∈ qmP qm with faithful normal states ψm on qmP qm.
ful normal states ϕm on qmP qm with (qmP qm)ϕm diffuse for all m. Define Φ(Pm∈N xm) =
Pm∈N ϕm(xm)qm for xm ∈ qmP qm, giving a faithful normal conditional expectation from Q′∩P
onto Q. Set ϕ := χ◦ Φ◦ EQ′∩P , a faithful normal state on P , with a faithful normal state χ on
Q. Then Q′ ∩ Pϕ =P⊕
m∈N(qmP qm)ϕm, a direct sum of diffuse von Neumann algebras. One can
choose, for each m, unitaries um, vm ∈ (qmP qm)ϕm so that ϕm(uk
m) = 0 as long as
k 6= 0. (See the proof of [27, Theorem 3.7].) Then u :=Pm∈N um, v :=Pm∈N vm are unitaries
in Q′ ∩ Pϕ, and moreover EQ(uk) = Ψ(uk) = 0 and Eϕ
(3) Consider P = P z ⊕ P z⊥ ⊇ R := Q ∨ {z}′′ = Qz ⊕ Qz⊥ ⊇ Q. Let χ be an arbitrary
faithful normal state on Q. As in the proof of Corollary 4.4 one can show that there is a unique
faithful normal conditional expectation ER : P → R with EQ ◦ ER = EQ. Then we have
Q(vk) = Φ(vk) = 0 as long as k 6= 0.
m) = ψm(vk
eER
^EQ↾R
⊇ eQ = Q ⋊σχ R,
where eER = (ER ¯⊗idB(L2(R))) ↾ eP and ^EQ ↾R = ((EQ ↾R) ¯⊗idB(L2(R))) ↾ eR= eEQ ↾ eR. Since ER =
(ER ↾P z) ⊕ (ER ↾P z⊥) in P = P z ⊕ P z⊥, we have, by [20, Theorem X.1.7 (ii)],
⊇ eR = R ⋊σχ◦(EQ↾R ) R
eP = P ⋊σχ◦EQ R
(cid:16)eP
eER
⊇ eR(cid:17) ∼=(cid:16)fP z
^ER↾P z
⊇ fQz(cid:17) ⊕(cid:16)gP z⊥
^ER↾P z⊥
⊇ gQz⊥(cid:17),
where the continuous cores and the conditional expectations in the right-hand side are defined
z⊥,n) = ( ^ER ↾P z⊥)(uk
m∈N Cem, and ER ↾P z factors as P z
similarly as above. Since gP z⊥ has no type I direct summand by the assumption here and [20,
Theorem XII.1.1] and since ^ER ↾P z⊥ preserves the canonical trace on gP z⊥ see e.g. [14, §4], we
can apply (1) to the second (cid:16)gP z⊥ ⊇ gQz⊥(cid:17) with ^ER ↾P z⊥ directly, and get a faithful normal
state ϕz⊥ on gP z⊥ with ϕz⊥ ◦ ( ^ER ↾P z⊥) = ϕz⊥ such that for each n ∈ N with n ≥ 2 one can
find a unitary uz⊥,n ∈ (fP z)ϕz⊥ in such a way that eER(uk
z⊥,n) = 0 for all
1 ≤ k ≤ n− 1. Write Qz =P⊕
E(Qz)′∩P z−→ (Qz)′ ∩ P z Ψ−→ Qz,
where (Qz)′ ∩ P z = P⊕
m∈N em(P z)em and E(Qz)′∩P z(x) = Pm∈N emxem for x ∈ P z. More-
over, Ψ is of the form Ψ(Pm∈N xm) = Pm∈N ψm(xm)em for xm ∈ em(P z)em with faithful
normal states ψm on em(P z)em. By the assumption here P z is diffuse and of type I, and
thus so are the em(P z)em; hence the centers of those must be diffuse, and so are all the
(em(P z)em)ψm . In the same way as in (2), one can find a unitary uz ∈ ((Qz)′ ∩ P z)χz◦Ψ with
'any' faithful normal state χz on Qz in such a way that Ψ(uk
z ) = 0 for all k 6= 0. Denote
by λ(t) the generators of C ⋊ R in fP z = (P z) ⋊σχz ◦(ER↾P z ) R (←֓ (Qz) ⋊σχz R = fQz canoni-
cally), and set ϕz := τ ◦ ( ^ER ↾P z), a faithful normal state on fP z, with a fixed faithful normal
tracial state τ := χz ¯⊗τ0 on fQz = Qz ¯⊗λ(R)′′. Note that λ(t)uz = σχz ◦(ER↾P z)
uzλ(t) for all t ∈ R. Thus, for any finite sum x = Pk xkλ(tk) ∈ fP z with xk ∈ P z we
(uz)λ(t) =
t
18
Y. UEDA
ϕ(x) := 1
z ) = ER(uk
z ) = ( ^ER ↾P z)(uk
have ϕz(uzx) = Pk τ (Ψ(uzE(Qz)′∩P z(xk))λ(tk)) = Pk χz(Ψ(uzE(Qz)′∩P z(xk)))τ0(λ(tk)) =
Pk χz(Ψ(E(Qz)′∩P z(xk)uz))τ0(λ(tk)) = Pk ϕz(xkuzλ(tk)) = ϕz(xuz). It follows that uz falls
in (fP z)ϕz . Clearly eER(uk
z ) = 0 for all k 6= 0. Set
2 (ϕz(xz) + ϕz⊥(xz⊥)) for x ∈ eP , and then ϕ becomes a faithful normal state on eP
and satisfies ϕ ◦ eER = ϕ, implying the desired condition (a), since eR is commutative. For
each n ∈ N with n ≥ 2 we define the unitary un := uz ⊕ uz⊥,n ∈ fP z ⊕ gP z⊥ = eP , and thus
eER(uk
z⊥,n) = 0 for all 1 ≤ k ≤ n − 1. Hence the desired
condition (b) is immediate as in (1) from the fact that eEQ = eEQ ◦ eER and Eϕ
eQ ◦ eER (the
latter follows from ϕ ◦ eER = ϕ). The final assertion is shown in the exactly same way (but
z ) ⊕ ( ^ER ↾P z⊥)(uk
n) = ( ^ER ↾P z)(uk
easier) as above.
z ) = Ψ(uk
(cid:3)
= Eϕ
eQ
We will give two applications of Proposition 3.5. The latter is a straightforward generalization
of both [27, Theorem 3.7] and [28, Proposition 3.1]. Remark that the former reproves the
assertions (1), (4) in Corollary 4.4 without any use of the technologies provided in §§3.1 -- 3.2.
Theorem 4.8. Assume that M1 diffuse, N of type I and M2 ⊇ N entirely non-trivial. Let
z ∈ Z(N ) be the unique projection such that N z is diffuse and N z⊥ atomic, and assume further
that (M1)cM1
has no type I direct summand when z 6= 0 (i.e., this last assumption is fulfilled if
M1 has no type I direct summand). Then (fM )ω =(cid:0)fM(cid:1)′
∩Z(eN )ω. In particular,
fM and hence M itself are non-amenable. If M is additionally assumed to be semifinite, then
Mω = M ′ ∩ M ω = M ′ ∩ Z(N )ω also holds.
=(cid:0)fM(cid:1)′
∩(cid:0)fM(cid:1)ω
z
After the completion of the main part of the present work we learned that Houdayer and
Vaes have also independently been obtained a similar (but not same) result as above under
different assumptions with different (and simpler) methods. See [8, Theorem 5.8]. More on this
will be discussed at the end of this section.
eN
f = 1. Let us first prove:
Proof. Note that (eN ⊇ N ) ∼= (N ¯⊗λ(R)′′ ⊇ N ¯⊗C1). Since N is of type I, one can choose an
abelian f ∈ N p (⊂ eN p) with c
For each x ∈ eN ω with representative (x(m))m one has f xf = [(f x(m)f )m], and for every
m there is a unique z(m) ∈ Z(eN ) with f x(m)f = z(m)f . By c
f = 1 the mapping x′ ∈
eN ′ 7→ x′f ∈ eN ′f gives a bijective normal ∗-homomorphism (thus k−k∞-preserving), and hence
(z(m))m defines z ∈ Z(eN )ω. Consequently we get f xf = zf ∈ Z(eN )ωf .
By Lemma 4.6 together with (2.5) we have the identification
f = Z(eN )ωf.
f(cid:0)eN(cid:1)ω
(4.1)
eN
(cid:0) ]f M f, ^E ↾f Mf(cid:1) =(cid:0)^f M1f , ^E1 ↾f M1f(cid:1) ⋆]f N f (cid:0)^f M2f , ^E2 ↾f M2f(cid:1).
Let c ∈ Z(M1) be the central support projection of the type I direct summand of M1. Then
e = cf is that of f M1f too, and f N f e = Z(N )f e must be atomic (or 0 if e = 0) by the
e ∼= Z(N )e = Z(N )f e is not
assumption here. In fact, if this was not the case, then Z(N )cN
atomic, and hence zcN
z c ≥ zc ≥ ze 6= 0, a contradiction to that
M1cM1
has no type I direct summand. Therefore, by Lemma 4.7 (3) we can apply Proposition
3.5 to (4.2) and thus any x ∈(cid:0) ]f M f(cid:1)′
and any y ∈(cid:0)^f M2f(cid:1)◦
e 6= 0, i.e., ze 6= 0, implying cM1
∩(cid:0) ]f M f(cid:1)ω
must satisfy that
z
( ^E2 ↾f M2f )(y∗y)(x − ( ^Ef M1f )ω(x)) = 0,
(4.2)
(4.3)
NON-TRACIAL AMALGAMATED FREE PRODUCTS
19
and (4.1) imply that
, that
= π−1
= (cid:0) ]f M f(cid:1)′
∩(cid:0) ]f M f(cid:1)ω
∩(cid:0)]f N f(cid:1)ω
.
(x) ∈ (cid:0)^f M1f(cid:1)ω
(cid:0) ]f M f(cid:1)′
. Consequently (cid:0) ]f M f(cid:1)′
and (cid:0)^f M2f(cid:1)ω
∩(cid:0) ]f M f(cid:1)ω
=(cid:0) ]f M f(cid:1)′
where Ef M1f is the unique conditional expectation from f M f onto f M1f determined as (2.2).
Note that ^f M2f ⊇ ]f N f with ^E2 ↾f M2f contains f M2f ⊇ f N f with E2 ↾f M2f canonically.
Hence, by Lemma 4.2 (2), (3) one can find a family {yi}i∈I in (f M2f )◦ in such a way that
Pi∈I s(E2(y∗
i yi)) = f (= 1f N f ). Therefore, it follows from (4.3) as in the proof of Theorem 4.3
that x = (cid:0) ^Ef M1f(cid:1)ω
∩(cid:0)^f M1f(cid:1)ω
.
In the same way as in the proof of Theorem 4.3, we see, by using the above {yi}i∈I again and
the free independence between (cid:0)^f M1f(cid:1)ω
Choose a faithful normal semifinite trace TrN on N , and fM ⊇ fMk (k = 1, 2) ⊇ eN are realized
as fM = M ⋊σTrN ◦E R ⊇ fMk = Mk ⋊σTrN ◦Ek R ⊇ eN = N ⋊σTrN R. Since σTrN ◦E
t ∈ R, ffM f ⊇ ffMkf ⊇ f eN f are naturally identified with ]f M f ⊇ ^f Mkf ⊇ ]f N f . Hence (4.4)
f =(cid:0)fM(cid:1)′
f ∩ f eN ωf =(cid:0)fM(cid:1)′
f ∩ Z(eN )ωf.
Let πf be the normal surjective ∗-homomorphism x ∈(cid:0)fM(cid:1)′
7→ xf ∈(cid:0)fM ′∩(cid:0)fM(cid:1)ω(cid:1)f =
∩(cid:0)fM(cid:1)ω
(cid:0)fM(cid:1)′
f ∩ (f(cid:0)fM(cid:1)ω
∩(cid:0)fM(cid:1)ω
f = 1 too). By (4.5) we have (cid:0)fM(cid:1)′
f (cid:0)(cid:0)fM(cid:1)′
f ∩ Z(eN )ωf(cid:1). As in the proof
of Lemma 4.2 one can choose partial isometries {vi}i∈I in eN so that Pi∈I v∗
i ≤ f for all i ∈ I. Then, if x = zf ∈ (cid:0)fM(cid:1)′
f ∩ Z(eN )ωf with z ∈ Z(eN )ω, then
i2 )xvi2 = Pi1,i2∈I v∗
i2vi2 = Pi1,i2∈I v∗
we have yz = Pi1,i2∈I v∗
∩Z(eN )ω. Hence(cid:0)fM(cid:1)′
i2 vi2 = zy for y ∈ fM , implying z ∈(cid:0)fM(cid:1)′
Pi1,i2∈I v∗
f ∩Z(eN )ωf =
∩(cid:0)fM(cid:1)ω
∩ Z(eN )ω(cid:1)f . Consequently (cid:0)fM(cid:1)′
(cid:0)(cid:0)fM(cid:1)′
f ∩ Z(eN )ωf(cid:1) = π−1
f (cid:0)(cid:0)fM(cid:1)′
f (cid:0)(cid:0)(cid:0)fM(cid:1)′
∩
∩(cid:0)fM(cid:1)ω
∩ Z(eN )ω. Since (cid:0)fM(cid:1)′
= (cid:0)fM(cid:1)′
Z(eN )ω(cid:1)f(cid:1) = (cid:0)fM(cid:1)′
∩ Z(eN )ω is commutative, it must
equal (fM )ω as observed in [23, (8) in page 360].
The final assertion is also shown in the exactly same way as above by using the final assertion
in Lemma 4.7 (3), since there is a faithful normal semifinite trace TrN on N so that TrN ◦ Ek
(k = 1, 2) are traces again thanks to Corollary 4.4 (3).
(cid:3)
f ) (c.f. [30, Lemma 4.1 (i)]), which is also injective due to c
(cid:0)fM(cid:1)′
f ∩ f(cid:0)fM(cid:1)ω
eN
f = 1 (and hence
t
(f ) = f for all
i1 (vi1 yv∗
= π−1
fM
c
and viv∗
i vi = c
eN
f = 1
i1 x(vi1 yv∗
i2 )vi2 =
(4.4)
(4.5)
i1 vi1 yzv∗
i1 vi1 zyv∗
Remark 4.9. The same type argument as in Theorem 4.3 (3) works for constructing a faithful
normal state χ on N with σχ◦E
T = Id with T = −2π/ log λ, 0 < λ < 1, when M is known to be
a factor of type IIIλ under the same set of assumptions as in Theorem 4.8. Hence the discrete
core of such M can also be written as an amalgamated free product von Neumann algebra of
the same form as the continuous core, and an analogous formula for its asymptotic centralizer
holds. In particular, the discrete core of such a factor of type IIIλ is an ∞-amplification of a
non-strongly stable type II1 factor. Further and more detailed discussions related to this aspect
will be given elsewhere.
Theorem 4.10. If M1 is diffuse, N of atomic type I and M2 ⊇ N entirely non-trivial, then
the following hold true:
(1) Mω = M ′ ∩ M ω = M ′ ∩Z(N ) (= Z(M )). Hence M does never have no type III0 direct
summand (see [5, Theorem 2.12]), and becomes full in the sense of Connes [5] under
the separability of preduals.
(2) The Connes τ -invariant τ (M ) (see [5]) is determined under the separability of preduals
as follows. Let χ be a faithful normal state on N . Then tm −→ 0 in τ (M ) as m → ∞
if and only if there is a unitary w ∈ N so that σχ◦E
tm −→ Adw in Aut(M ) as m → ∞.
20
Y. UEDA
fi
= 1 and σχ
Proof. (1) This is proved along the same line as in the proof of Theorem 4.8 by using only
Lemma 4.7 (2) instead together with a well-known fact Z(N ) = Z(N )ω due to the assumption
that it is atomic.
(2) We can write N = P⊕
i∈I B(Hi). Looking at this structure with the given χ we can
choose a collection {ei}i∈I of abelian projections in N with Pi∈I ei = 1 such that for each
t (fi) = fi (t ∈ R).
i ∈ I there is a larger abelian fi ∈ N p so that ei ≤ fi, cN
Assume that tm −→ 0 in τ (M ) as m → ∞. Then there is a sequence (um)m of unitaries
in M such that Adum ◦ σχ◦E
tm −→ id in Aut(M ) as m → ∞. As observed in the proof [28,
Proposition 3.1] the (um)m defines a unitary u ∈ M ω, and clearly ufi = fiu for all i ∈ I.
Hence fiu defines a unitary in fiM ωfi = (fiM fi)ω, and we denote it by ui for simplicity. Since
fiM1fi is still diffuse, looking at fiM1fi ⊇ fiN fi = Z(N )fi one can choose a faithful normal
state ϕ on fiM1fi as in Lemma 4.7 (2). Set ϕ(x) := ϕ(fixfi) + χ ◦ E1(f ⊥
i ), x ∈ M1,
which becomes a faithful normal positive linear functional on M1. Clearly fi ∈ (M1) ϕ and
thus fi[Dχ ◦ E1 : D ϕ]t = [Dχ ◦ E1 ↾fiM1fi : Dϕ]t for all t ∈ R by the uniqueness part of
Connes's Radon-Nikodym cocycle theorem. As observed in the proof of [28, Proposition 3.1]
again the sequence vm := [Dχ ◦ E1 : D ϕ]tm defines a unitary v ∈ M ω
1 and also the sequence
1 fi = (fiM1fi)ω. Since ϕ ◦ EM1 ↾fiMfi = ϕ ◦ (EM1 ↾fiMfi
fivm = vmfi does a unitary vi ∈ fiM ω
(y))m(cid:3) = uvz = uiviz for
), we have yuivi = yuv = (cid:2)(yumvm)m(cid:3) = (cid:2)(umvmσ
(y))m(cid:3) ∈ (fiM fi)ω = fiM ωfi in the identification
y ∈ (fiM2fi)◦ with z = (cid:2)(σ
(fiM fi, E ↾fiMfi ) = (fiM1fi, E1 ↾fiM1fi) ⋆fiN fi (fiM2fi, E2 ↾fiM2fi ) provided by Lemma 4.6.
By Proposition 3.5 we get (E2 ↾fiM2fi )(y∗y)(uivi − (EM1 ↾fiMfi )ω(uivi)) = 0 for y ∈ (fiM2fi)◦.
By using Lemma 4.2 (2), (3) twice as in the proof of Theorem 4.8 we can prove firstly that
uivi ∈ (fiM1fi)ω = fiM ω
1 fi), and finally that
ui ∈ fiN ωfi = Z(N )ωfi = Z(N )fi. Therefore, u = Pi∈I eiu = Pi∈I eifiu = Pi∈I eiui ∈ N .
Letting w := u∗ ∈ N u we have Adw∗ ◦ σχ◦E
1 fi (since vi ∈ fiM ω
tm −→ id in Aut(M ) as m → ∞.
1 fi, secondly that ui ∈ fiM ω
ϕ◦(EM1↾fiM fi )
tm
ϕ◦(EM1↾fiM fi )
tm
i xf ⊥
(cid:3)
The next proposition shows that Proposition 3.5 is still useful beyond the case where N is
of type I or even semifinite. The proof goes along the same line as that of Theorem 4.8 but is
easier than it. Hence the proof is left to the reader.
Proposition 4.11. Assume that there is a faithful normal state ϕ on M1 satisfying the following
conditions:
(a) σϕ
(b) For every n ∈ N with n ≥ 2 there are unitaries uk = u(n)
t (N ) = N for all t ∈ R.
k ∈ (M1)ϕ, 0 ≤
k ≤ n − 1, such that E1(u∗
vk2 ) = 0 for all 0 ≤ k1 6= k2 ≤ n − 1, where
Eϕ
N denotes the unique ϕ-preserving conditional expectation from M1 onto N , whose
existence follows from (a) and Takesaki's criterion.
k , vk = v(n)
uk2) = Eϕ
N (v∗
k1
k1
Assume also that M2 ⊇ N is entirely non-trivial. Then M ′ ∩ M ω = M ′ ∩ N ω holds. Moreover,
if it is further assumed that N is finite, then Mω = M ′ ∩ M ω = M ′ ∩ Nω.
It is easy to confirm that the (M1, E1) in Corollary 4.5 satisfies the assumption of Proposition
4.11. Thus M ′ ∩ M ω = M ′ ∩ N ω holds under the set of assumptions in Corollary 4.5.
Assume that M1 is a von Neumann algebra with separable predual and that N is a Cartan
subalgebra in M1.
It was proved in [21, Lemma 4.2] that if M1 is further assumed to be a
non-type I factor, then there are a faithful normal state ϕ on M1 with ϕ◦ E1 = ϕ and a unitary
u ∈ (M1)ϕ such that E1(uk) = 0 as long as k 6= 0. The same assertion can indeed be proved
even when M1 is further assumed only to have no type I direct summand (i.e., without being
a factor). The proof is similar to [21, Lemma 4.2] but tedious based on disintegration. Hence
such (M1, E1) satisfies the assumption of Proposition 4.11.
NON-TRACIAL AMALGAMATED FREE PRODUCTS
21
Remark 4.12. Almost all the results obtained above have appropriate 'HNN variants' thanks
to tricks given in [26]. Here it should be emphasized that our results so far essentially need
assumptions for only one free component. The notion of HNN extensions of von Neumann alge-
bras as well as their basic properties including their modular theoretic aspects were established
in [25].
In closing of this section we discuss one of Houdayer and Vaes's results [8, Theorem 5.8].
This part of the present paper is added after receiving a draft of [8] in order to point out only
one consequence obtained from this and that papers without any new idea. Therefore, some
facts provided in [8] are necessary below. The original aim of the present work is to provide
amalgamated free product counterparts of the results in [27, §3]. One issue to do so is how to
formulate a suitable assumption saying that M1 is 'diffuse relative to N ' which corresponds to
that M1 is diffuse when N = C1. The requirement for M1 ⊇ N in Theorem 4.3 seems to be
one strong form of them without any restriction on N , but it seems not so easy to check it in
general. Thus we propose the requirement for M1 ⊇ N in Corollary 4.4 and Theorem 4.8 as
such a candidate in the special case when N is of type I. However a more sophisticated one
in the special case seems to be that M1 ⊇ N has no trivial corner, which is proposed in [8,
§5] by a different motivation. In fact, Houdayer and Vaes [8, Theorem 5.8] give a factoriality
and non-amenability result under the set of assumptions that both Mk ⊇ N , k = 1, 2, have no
trivial corner and that N is of type I, and establish their primeness result under the same set
of assumptions. Here an inclusion P ⊇ Q of von Neumann algebras is said to have no trivial
corner if pP p 6= Qp for any non-zero projection p ∈ Q′ ∩ P . Any exact general relationship
between theirs and ours is not immediately clear. However the proof of Theorem 4.8 and general
properties on inclusions without trivial corner provided in [8, §§5.1] altogether immediately give
an improvement of [8, Theorem 5.8], though it is not immediately clear whether the primeness
result in [8, Theorem E] holds or not under the new set of assumptions.
Theorem 4.13. If M1 ⊇ N has no trivial corner, N is of type I and M2 ⊇ N entirely non-
trivial, then the following hold true:
(1) Z(M ) = Z(M1) ∩ Z(M2) ∩ Z(N ).
(2) Z(fM ) = Z(fM1) ∩ Z(fM2) ∩ Z(eN ).
(3) (fM )ω =(cid:0)fM(cid:1)′
=(cid:0)fM(cid:1)′
∩(cid:0)fM(cid:1)ω
In particular, (3) explains that M does never become amenable.
∩ Z(eN )ω.
Proof. It is trivial that (3) ⇒ (2) ⇒ (1), see e.g. the proof of [25, Theorem 5.2] for (3) ⇒ (2)
and [20, Theorem X.II.1.1] for (2) ⇒ (1). Thus it suffices to prove only (3). The line of the
proof below is exactly identical to that of Theorem 4.8, and thus we keep the notations there.
In fact, only one modification is sufficient. By [8, Lemma 5.2, Proposition 5.5] the inclusion
^f M1f ⊇ ]f N f also has no trivial corner. Then it suffices to prove the exactly same assertion as
in Lemma 4.7 (1) with replacing the assumption that P has no type I direct summand by that
P ⊇ Q has no trivial corner. In fact, by using this new assertion instead of Lemma 4.7 (3) one
gets the same equation (4.3) and the rest of the proof there works well.
Let P ⊇ Q be an inclusion of von Neumann algebras without trivial corner. Assume that
Q is commutative, P has a faithful normal semifinite trace TrP and there is a faithful normal
conditional expectation EQ : P → Q satisfying TrP ◦ EQ = TrP . As in the proof of Lemma 4.7
(1) we choose the qm's and χ. Then we apply [8, Lemma 5.4 (3)] (note that it holds without
assuming the separability of preduals, see Lemma 4.14 below) with q = p := qm and get a
m) = 0 as long as k 6= 0. Letting u :=Pm∈N um we
unitary um ∈ qmP qm satisfying that EQ(uk
have u ∈ Pχ◦EQ and EQ(uk) = 0 as long as k 6= 0. Hence we are done.
(cid:3)
22
Y. UEDA
As remarked in [8, Lemma 5.3] the next lemma immediately follows from Rohlin's general
theorem on Lebesgue spaces under the separability of preduals. Thus only the advantage of the
proof below is no use of disintegration; hence the separability of preduals is not necessary in [8,
Lemma 5.4]. Although it is a rather minor point, we do give it for the sake of completeness.
Lemma 4.14. Let B ⊇ A be (unital) inclusion of commutative σ-finite von Neumann algebras
with a faithful normal conditional expectation EA : B → A.
If Bf 6= Af for any nonzero
projection f ∈ B, then there is a unitary u ∈ B such that EA(uk) = 0 as long as k 6= 0.
Proof. Choose non-zero f ∈ Bp. Since Bf 6= Af , there is x ∈ B such that x 6∈ Af and
0 ≤ x ≤ f . Since EA(x) ≤ EA(f ), one can choose a positive contraction c ∈ A so that
cEA(f ) = EA(x) (since A is commutative). Letting y := x − cf ∈ Bf we have y = y∗ 6= 0 (due
to x 6∈ Af ) and EA(y) = 0. Therefore, an idea given in the proof of [3, Lemma 2.1] enables
us to construct projections e(ε1,...,εn) ∈ B, n ∈ N, εk ∈ {1, 2}, in such a way that e(ε1,...,εn) =
e(ε1,...,εn,1) + e(ε1,...,εn,2) and EA(e(ε1,...,εn)) = 1
2n 1. The proof is done by induction. Assume
that we have chosen up to n-th stage. Set Λe := {x = x∗ ∈ Bekxk∞ ≤ 1, EA(x) = 0} with
e := e(ε1,...,εn). It is a σ-weakly compact convex subset, and thus has sufficiently many extremal
points due to the Krein -- Milman theorem. Let a ∈ Λe be an extremal point. Then it suffices to
prove a = 2e0 − e for some e0 ∈ Bp with e0 ≤ e, since it clearly implies that EA(e0) = 1
2 EA(e).
On contrary, suppose that it is not the case. By the spectral decomposition of a one can find
δ > 0 and non-zero f ∈ Bp in such a way that f ≤ e and −(1 − δ)f ≤ af ≤ (1 − δ)f . By
what we have shown above, there is a non-zero y = y∗ ∈ Bf such that −δf ≤ y ≤ δf and
EA(y) = 0, and hence a + y, a − y ∈ Λe and a = 1
2 (a − y), a contradiction. Thus
e(ε1,...,εn,1) := e0 and e(ε1,...,εn,2) := e − e0 become desired ones in (n + 1)-th stage. Hence we
have proved the claim. Let (C, ω) be the von Neumann algebraic infinite tensor product of
C ⊕ C with equal weights {1/2, 1/2}. Once passing GNS representations one can construct an
injective normal ∗-homomorphism from C ¯⊗A into B which intertwines ω ¯⊗idA and EA. Hence
the desired assertion follows, since (C, ω) ∼= (L(Z), τZ) thanks to [19, Theorem III.1.22].
(cid:3)
2 (a + y) + 1
The entire non-triviality of an inclusion P ⊇ Q of von Neumann algebras is nothing but
just the non-triviality of P when Q = C1, and hence Theorem 4.13 is no longer true under
assuming only that M1 ⊇ N is entirely non-trivial instead.
In fact, the plain free product
of two 2-dimensional algebras with suitable states provides a counter example, see [27] for
suitable references therein. Finally we conjecture that Corollary 4.4, especially a strong kind
of irreducibility ((M1)ϕ)′ ∩ M ⊆ M1 for some faithful normal state ϕ, should also hold under
the same set of assumptions of Theorem 4.13. This is rather technical, but such a property
may have some potential in further analysis. We will consider it in future work beyond the case
where Z(M ) = Z(M1) ∩ Z(M2) ∩ Z(N ) need not hold.
Acknowledgment
I thank Professors Cyril Houdayer and Stefaan Vaes for several fruitful conversations in
Dec. 2011 and in Jan. 2012 and also for sending us a draft of [8] prior to putting it on the
ArXiv.
References
[1] J. Asher, A Kurosh-type theorem for type III factors, Proc. Amer. Math. Soc., 137 (2009), 4109 -- 4116.
[2] N.P. Brown and N. Ozawa, C∗-algebras and finite-dimensional approximations. Graduate Studies in Math-
ematics, 88, American Mathematical Society, Providence, RI, 2008.
[3] J. Cameron, J. Fang and K. Mukherjee, Mixing subalgebras of finite von Neumann algebras, Preprint,
arXiv:1001.0169.
NON-TRACIAL AMALGAMATED FREE PRODUCTS
23
[4] I. Chifan and C. Houdayer, Bass -- Serre rigidity results in von Neumann algebras, Duke Math. J., 153
(2010), 23 -- 54.
[5] A. Connes, Almost periodic states and factors of type III1, J. Funct. Anal., 16 (1974), 415 -- 445.
[6] K. Dykema, Free products of hyperfinite von Neumann algebras and free dimension, Duke Math. J., 69
(1993), 97 -- 119.
[7] C. Houdayer, Construction of type II1 factors with prescribed countable fundamental group, J. reine
angew. Math. 634 (2009), 169 -- 207.
[8] C. Houdayer and S. Vaes, Type III factors with unique Cartan decomposition, Preprint, 2012.
[9] A. Ioana, J. Peterson and S. Popa, Amalgamated free products of weakly rigid factors and calculation of
their symmetry groups, Acta Math., 200 (2008), 85 -- 153.
[10] M. Izumi, R. Longo and S. Popa, A Galois correspondence for compact groups of automorphisms of von
Neumann algebras with a generalization to Kac algebras, J. Funct. Anal., 155 (1998), 25 -- 63.
[11] R.V. Kadison, Diagonalizing matrices, Amer. J. Math., 106 (1984), 1451 -- 1468.
[12] R.V. Kadison and J. Ringrose, Fundamentals of the theory of operator algebras, Vol. II, Advanced theory,
Graduate Studies in Mathematics, 16, American Mathematical Society, Providence, RI, 1997.
[13] H. Kosaki, Extension of Jones' theory on index to arbitrary factors, J. Funct. Anal., 66 (1986), 123 -- 140.
[14] R. Longo, Index of subfactors and statistics of quantum fields, I, Commun. Math. Phys., 126 (1989),
217 -- 247.
[15] S. Popa, Markov traces on universal Jones algebras and subfactors of finite index, Invent. math., 111
(1993), 375 -- 405.
[16] S. Popa, On a class of type II1 factors with Betti numbers invariants, Ann. of Math. (2), 163 (2006),
809 -- 899.
[17] S. Popa, Strong rigidity of II1 factors arising from malleable actions of w-rigid groups, I, Invent. math.,
165 (2006), 369 -- 408.
[18] S. Popa, On Ozawa's property for free group factors, Int. Math. Res. Not., IMRN 2007, Art. ID rnm036.
[19] M. Takesaki, Theory of Operator Algebras, I, Encyclopedia of Mathematical Sciences, 124, Operator Alge-
bras and Non-commutative Geometry, 5, Springer, Berlin, 2002.
[20] M. Takesaki, Theory of Operator Algebras, II, Encyclopedia of Mathematical Sciences, 125, Operator
Algebras and Non-commutative Geometry, 6, Springer, Berlin, 2003.
[21] Y. Ueda, Amalgamated free product over Cartan subalgebra, Pacific J. Math., 191 (1999), 359 -- 392.
[22] Y. Ueda, Remarks on free products with respect to non-tracial states, Math. Scand., 88 (2001), 111 -- 125.
[23] Y. Ueda, Fullness, Connes' χ-groups, and ultra-products of amalgamated free products over Cartan subal-
gebras, Trans. Amer. Math. Soc., 355 (2003), 349 -- 371.
[24] Y. Ueda, Amalgamated free product over Cartan subalgebra, II: Supplementary Results & Examples,
Advanced Studies in Pure Mathematics, 38 (2004), 239-265.
[25] Y. Ueda, HNN extensions of von Neumann algebras, J. Funct. Anal., 225 (2005), 383 -- 426.
[26] Y. Ueda, Remarks on HNN extensions in operator algebras, Illinois J. Math., 52 (2008), 705 -- 725.
[27] Y. Ueda, Factoriality, type classification and fullness for free product von Neumann algebras, Adv. Math.,
228 (2011), 2647-2671.
[28] Y. Ueda, On type III1 factors arising as free products, Math. Res. Lett., 18 (2011), 909 -- 920.
[29] S. Vaes, Rigidity results for Bernoulli actions and their von Neumann algebras (after Sorin Popa), S´eminaire
Bourbaki, Vol. 2005/2006, Ast´erisque, 311 (2007), Exp. No. 961, viii, 237 -- 294.
[30] B. J. Voeden, Normalcy in von Neumann algebras, Proc. London Math. Soc. (3), 27 (1973), 88 -- 100.
[31] D. Voiculescu, Symmetries of some reduced free product C ∗-algebras, in Operator algberas and their con-
nections with topology and ergodic theory, Lect. Notes in Math., 1132 (1985), 566 -- 588.
Graduate School of Mathematics, Kyushu University, Fukuoka, 810-8560, Japan
E-mail address: [email protected]
|
1701.01338 | 1 | 1701 | 2017-01-05T15:01:59 | Functors induced by Cauchy extension of C*-algebras | [
"math.OA",
"math.FA"
] | In this paper we give three functors $\mathfrak{P}$, $[\cdot]_K$ and $\mathfrak{F}$ on the category of C$^\ast$-algebras. The functor $\mathfrak{P}$ assigns to each C$^\ast$-algebra $\mathcal{A}$ a pre-C$^\ast$-algebra $\mathfrak{P}(\mathcal{A})$ with completion $[\mathcal{A}]_K$. The functor $[\cdot]_K$ assigns to each C$^\ast$-algebra $\mathcal{A}$ the Cauchy extension $[\mathcal{A}]_K$ of $\mathcal{A}$ by a non-unital C$^\ast$-algebra $\mathfrak{F}(\mathcal{A})$. Some properties of these functors are also given. In particular, we show that the functors $[\cdot]_K$ and $\mathfrak{F}$ are exact and the functor $\mathfrak{P}$ is normal exact. | math.OA | math |
Functors induced by Cauchy extension of
C∗-algebras
Kourosh Nourouzi1 ∗, Ali Reza2
1,2 Faculty of Mathematics, K. N. Toosi University of
Technology,
P.O. Box 16315-1618, Tehran, Iran.
October 14, 2018
Abstract
In this paper we give three functors P, [·]K and F on the category
of C∗- algebras. The functor P assigns to each C∗-algebra A a pre-C∗-
algebra P(A) with completion [A]K . The functor [·]K assigns to each
C∗-algebra A the Cauchy extension [A]K of A by a non-unital C∗-algebra
F(A). Some properties of these functors are also given. In particular, we
show that the functors [·]K and F are exact and the functor P is normal
exact.
Keywords: Pre-C∗- algebras; Extensions of C∗- algebras; Exact functors;
Cauchy extension.
1 Introduction
Given a complex C∗-algebra A, the algebra A[[Z]] consists of all sequences
(an)∞
n=0 in A with pointwise linear operations and Cauchy product
((an)∞
n=0)((bn)∞
n=0) = (cn)∞
n=0,
where each cn = Pn
the formal power series in one variable of the form P∞
k=0 akbn−k. It is natural to think of elements of A[[Z]] as
n=0 anZ n with product
(
∞Xn=0
∞Xn=0
anZ n)(
bnZ n) =
∞Xn=0
cnZ n,
∗Corresponding author, e-mail: [email protected]
2010 Mathematics Subject Classification. 46L05, 46M15
1
where cn's are as above. One may consider the complex subalgebra
A[Z] = {
∞Xn=0
anZ n :
∞Xn=0
kank < ∞},
of A[[Z]]. It is of interest to find a C∗-algebra via A[Z] to be an extension of
A. Recall that an extension B of C by A is a short exact sequence
0 −→ A
f
−→ B
g
−→ C −→ 0.
(1)
of C∗-algebras (see, e.g., [1, 5, 6, 10]). For any subset Kof [−1, 1] such that 0
is a limit point of K, we will define a pre-C∗-norm on A[Z]. The completion of
A[Z], denoted by [A]K, is an extension of A (Proposition 7 (iii)) which will be
called the Cauchy extension of A.
The outline of this work is as follows.
In Section 2 we introduce pre-C∗-
In Proposition 5, it is shown that A[Z] is not a C∗-algebra.
algebra A[Z].
Proposition 7 shows that the completion [A]K of pre-C∗-algebra A[Z] is an
extension of A. We also introduce the functors P, [·]K and F on the category
of C∗-algebras. The functor P assigns to each C∗-algebra A a pre-C∗-algebra
P(A) = A[Z]. The functor [·]K assigns to each C∗-algebra A an extension
[A]K of A by a non-unital C∗-algebra F(A), where the C∗-algebra F(A) is the
completion of the ideal
A1 = {
∞Xn=0
anZ n ∈ A[Z] : a0 = 0}
of A[Z]. Some properties of functors P, [·]K and F are listed in Proposition 8.
In Section 3 we show that the functors [·]K and F are exact. In Section 4, using
the notion of normal exact sequence of the normed spaces introduced by Yang
[16], we prove that the functor P is normal exact. More precisely, for any short
exact sequence of C∗-algebra (1) the corresponding short exact sequence
0 −→ A[Z]
f
−→ B[Z]
g
−→ C[Z] −→ 0
is a normal exact sequence of pre-C∗-algebras. That is, B(Z)/ ker g −→ C[Z] is
an isometry. Among other results we also show that for any closed ideal I of
a C∗-algebra A, the pre-C∗-algebra I[Z] is a closed ideal of A[Z] (Proposition
8 (iii)) and the quotient A[Z]/I[Z] is a pre-C∗-algebra (Theorem 3) which is
isometric ∗-isomorphic to (A/I)[Z] (Theorem 4). Finally in Section 5, we show
that the Cauchy extension [A]K of a C∗-algebra A can be considered as a C∗-
subalgebra of Cb(K, A), the C∗-algebra of all bounded continuous functions
from K to A (Theorem 5 (i)).
In particular, if K is compact, then [A]K is
∗-isomorphic to C(K, A). We also give some other results in Theorem 5. A
minimax type result is given in Corollary 5.
2
2 Cauchy extension of C∗- algebras
Let A be a complex Banach algebra and A[[Z]] be the complex algebra consisting
of all formal power series in A. If A has a unit, then an element F = F (Z) =
n=0 anZ n ∈ A[[Z]] is invertible if and only if a0 is an invertible element in A.
In particular, 1 + Z 2 is invertible in A[[Z]] and we have
P∞
(1 + Z 2)(
∞Xn=0
The subalgebra
(−1)nZ 2n) = (
(−1)nZ 2n)(1 + Z 2) = 1.
∞Xn=0
A[Z] = {
∞Xn=0
anZ n ∈ A[[Z]] :
∞Xn=0
kank < ∞}
can be equipped with a norm as
kF k =
∞Xn=0
kank,
(2)
(3)
for all F (Z) =P∞
n=0 anZ n ∈ A[Z].
Proposition 1. Let A be a Banach algebra. Then A[Z] with the norm given
in (3) is a Banach algebra.
n=0 aknZ n) be a
Proof. To show that A[Z] is a Banach algebra, let (Fk) = (P∞
sequence in A[Z] such that P∞
∞Xn=0
k=0 kFkk < ∞. Then
kaknk < ∞.
kaknk =
∞Xk=0
∞Xn=0
k=0 akn and F =P∞
kPN
∞Xk=0
Let cn =P∞
There exists a positive integer N such thatP∞
n=0(P∞
n=0 kP∞
n=0P∞
k=N +1P∞
k=0 Fk − F k = kP∞
= P∞
≤ P∞
= P∞
k=N +1 akn)Z nk
k=N +1 aknk
k=N +1 kaknk
n=0 kaknk
< ε.
n=0 cnZ n. Then F ∈ A[Z]. Let ε > 0 be given.
n=0 kaknk < ε. We have
k=N +1P∞
This completes the proof.
Proposition 2. Let A be a Banach algebra. If F (Z) = P∞
then PN
n=0 anZ n → F (Z) as N → ∞,
n=0 anZ n ∈ A[Z],
3
Proof. Since
kF (Z) −
NXn=0
anZ nk = k
∞Xn=N +1
anZ nk =
∞Xn=N +1
kank,
we get the desired limit.
gent series in A[Z].
Now one can consider any element F (Z) =P∞
P∞
lution and the norm given in (3) is a ∗-Banach algebra.
If A is a C∗-algebra, we can define an involution ∗ in A[Z] by F ∗(Z) =
n=0 a∗
nZ n for any F (Z) ∈ A[Z]. In this case, A[Z] equipped with this invo-
n=0 anZ n ∈ A[Z] as a conver-
Proposition 3. Let A be a C∗-algebra. There is no norm on involutive algebra
(A[Z], ∗) which makes it a C∗-algebra. In particular, (A[Z], ∗) equipped with the
norm given in (3) is not a C∗-algebra.
Proof. We suppose on the contrary that there exists a norm k · k such that
(A[Z], ∗, k · k) is a C∗-algebra. Suppose that A is unital. By (2) the element
1 + Z 2 is not invertible in A[Z]. This implies that −1 ∈ σ(Z 2) which is a
contradiction. Now let A be non-unital and a ∈ A be self-adjoint with kak > 1.
Applying (2) for aZ we get that 1 + a2Z 2 is not invertible in (A ⊕ C)(Z). That
is, −1 ∈ σ(a2Z 2) which is again a contradiction.
then
For a C∗-algebra (A, k · k) if F (Z) = P∞
∞Xn=0
kantnk ≤
∞Xn=0
kank < ∞.
n=0 anZ n ∈ A[Z] and −1 ≤ t ≤ 1
Hence F (t) =P∞
n=0 antn is norm-convergent in A.
For any F (Z), G(Z) ∈ A[Z] and λ ∈ C, t ∈ [−1, 1] we have
(λF (Z))(t) = λF (t),
(F (Z) + G(Z))(t) = F (t) + G(t),
(F (Z)G(Z))(t) = F (t)G(t).
(4)
(5)
(6)
Note that the equalities (4) and (5) are clear and the proof of (6) is similar to
that of complex case (see [15, p. 74]).
The following proposition has a straightforward proof which is omitted here.
Proposition 4. Suppose that K is a subset of [−1, 1] such that 0 is a limit
point of K and (an)∞
n=0 is a sequence in C∗-algebra A. If
(i) F (Z) =P∞
n=0 anZ n such that P∞
(ii) F (t) = 0 for any t ∈ K,
n=0 kank < ∞;
then an = 0 for all n.
4
Hereafter, throughout the paper K will denote a subset of [−1, 1] such that
0 is a limit point of it.
Proposition 5. The following statements hold:
(i) The functional k · kK defined by
kF kK = sup
t∈K
k
∞Xn=0
antnk,
for all F = F (Z) =P∞
n=0 anZ n ∈ A[Z], is a norm;
(ii) (A[Z], ∗, k · kK) is a pre-C∗-algebra but not a C∗-algebra;
(iii) kF kK ≤ kF k for all F ∈ A[Z];
(iv) If F (Z) =P∞
n=0 anZ n, then PN
n=0 anZ n → F (Z) as N → ∞ in k · kK.
Proof. (i) From (4), (5), (6) and Proposition 4 it is easily seen that k · kK is a
norm. (ii) By the definition of k · kK we have the identity kF ∗F kK = kF k2
K.
Therefore (A[Z], ∗, k · kK) is a pre-C∗-algebra which by Proposition 3 is not a
C∗-algebra. (iii) By the definition of k · kK is clear. (iv) The proof follows from
Proposition 2 and Part (ii).
We will call the completion [A]K of pre-C∗-algebra (A, ∗, k · kK) the K -
It is clear that [A]K is a C∗-
Cauchy or simply the Cauchy extension of A.
algebra.
Proposition 6. Let A be a C∗-algebra. The following hold:
(i) If I is an ideal of A[Z], then the completion I of (I, k · kK) is a closed ideal
of [A]K;
(ii) If I is a closed ideal of A, then [I]K is a closed ideal of [A]K.
Proof. (i) Let I be an ideal of A[Z]. Then the completion I of (I, k · kK) is
a closed ideal of [A]K. Choose any element F ∈ I and G ∈ [A]K. Let (Fn)
and (Gk) be two sequences in I and A[Z] respectively converging to F ∈ I
and G ∈ [A]K . For any k, n ≥ 1 we have FnGk, GkFn ∈ I. This implies that
F GK, GkF ∈ I, for all k ≥ 1 and so F G, GF ∈ I. That is I is a closed ideal of
[A]K.
(ii) Consider F ∈ I[Z] and G ∈ A[Z]. It is clear that F G, GF ∈ I[Z], i.e., I[Z]
is an ideal of A[Z]. Now Part (i) implies that \(I[Z]) = [I]K is a closed ideal of
[A]K.
5
For a C∗-algebra A define
A0 = {F (Z) =
∞Xn=0
anZ n ∈ A[Z] : an = 0 f or n > 0},
A1 = {F (Z) =
∞Xn=0
anZ n ∈ A[Z] : a0 = 0}.
Denote the completion of A1 by A1.
It is clear that A1 is an ideal of A[Z]
and by Proposition 6, A1 is a closed ideal of [A]K . Hence if A 6= 0, then [A]K
has a proper closed ideal A1. Consequently no simple C∗-algebra is a Cauchy
extension of some C∗-algebra. It is worth mentioning that there is no ideal I of
A such that I[Z] = A1. Since A0 is naturally ∗-isomorphic to A we always use
A instead of A0 as a subalgebra of A[Z].
Suppose that A, B, E are C∗-algebras such that B is an ideal of E. It is said
to be E an extension of A by B if there is a short exact sequence
0 −→ B i−→ E
p
−→ A −→ 0,
where i(B) = ker p and i, p are injective and surjective ∗-homomorphisms re-
spectively (see, e.g.,[1]).
Proposition 7. Let A be a C∗-algebra. The following statements hold:
(i) Every element F of [A]K has a unique representation F = a + G, where
a ∈ A and G ∈ A1;
(ii) kakK = kak ≤ ka + GkK, for all a ∈ A and G ∈ A1;
(iii) [A]K is an extension of A by A1;
(iv) A1 is not unital as a C∗-subalgebra of [A]K .
Proof. (i) Let (Fk) be a Cauchy sequence in (A[Z], k · kK), where
Fk =
∞Xn=0
aknZ n ∈ A[Z].
Let ε > 0 be given. Then kFk − Fk′ kK < ε for sufficiently large k, k′ . Suppose
that (tm) is a sequence in K such that tm −→ 0 as m −→ ∞. By the definition
of k · kK we have
kak0 − ak′0k = lim
m→∞
k
∞Xn=0
(akn − ak′n)tn
mk ≤ sup
t∈K
k
∞Xn=0
(akn − ak′n)tnk
= kFk − Fk′ kK < ε,
6
for sufficiently large k, k′. Furthermore
k
sup
t∈K
∞Xn=1
(akn − ak′n)tnk < 2ε.
n=1 aknZ n) are Cauchy in A and A1,
n=0 aknZ n ∈
n=1 aknZ n −→ G ∈ A1 as
k → ∞. Since A1 ∩ A = 0, then this representation is unique. Hence [A]K is
the internal direct sum of subspaces A and A1, i.e., [A]K = A ⊕ A1.
Therefore the sequences (ak0) and (P∞
respectively. For F ∈ [A]K , let F = limk→∞ Fk, where Fk = P∞
A[Z]. Then F = a + G, where ak0 −→ a ∈ A and P∞
(ii) Note that if a ∈ A and G = limk→∞P∞
n=1 aknZ n ∈ A1, then
ka +
∞Xn=1
aknZ nkK = sup
t∈K
ka +
akntnk,
∞Xn=1
ka +P∞
for all k ≥ 1. A similar method to that used in Part (i) implies that kak ≤
n=1 aknZ nk, for all k ≥ 1. Therefore kak ≤ ka + Gk, for all a ∈ A and
G ∈ A1.
(iii) Define pA : [A]K −→ A by pA(a + G) = a, for all a ∈ A and G ∈ A1. It is
easily seen that pA is a surjective ∗-homomorphism and ker pA = A1. Therefore
we have the short exact sequence
0 → A1
i
֒→ [A]K
pA→ A → 0.
(7)
This shows that [A]K is an extension of A by A1.
(iv) Suppose on the contrary that A1 is unital with unit U (Z). Since aZU (Z) =
aZ for all a ∈ A, we have taU (t) = ta for any t ∈ K and a ∈ A. This implies that
aU (t) = a for all t 6= 0 and therefore limt→0 U (t) 6= 0, which is a contradiction.
Remark 1. Each ∗-homomorphism f : A −→ B of C∗-algebras induces a ∗-
homomorphism f : A[Z] −→ B[Z] between pre-C∗-algebras A[Z] and B[Z] by
f (
∞Xn=0
anZ n) =
∞Xn=0
f (an)Z n,
(8)
where P∞
n=0 anZ n ∈ A[Z].
Remark 2. If we define P(A) = A[Z] for any C∗-algebra A and P(f ) = f ,
for any ∗-homomorphism f : A −→ B of C∗-algebras, then P is a functor
from the category of C∗-algebras to the category of pre-C∗-algebras. Each ∗-
homomorphism f : A[Z] −→ B[Z] defined by (8) induces a ∗-homomorphism
f : [A]K −→ [B]K. It is easy to see that [·]K is a functor from the category
of C∗-algebras to itself as [f ]K = f . Now defining F(A) = A1 and F(A
−→
B) = f A1
: A1 −→ B1, for C∗-algebras A, B and ∗-homomorphism f , we get a
functor on the category of C∗-algebras which assigns, by Proposition 7 (iv), to
any C∗-algebra a non-unital C∗-algebra.
f
7
By A ∼= B we mean that the C∗-algebras A and B are ∗-isomorphic.
Proposition 8. Let f : A −→ B be a ∗-homomorphism of C∗-algebras. Then
(i) f is a contraction;
(ii) f and f are isometries provided that f is an isometry;
(iii) f is surjective provided that f is surjective;
(iv) If f is a ∗-isomorphism, then both f and f are ∗-isomorphisms;
(v) ker f = (ker f )[Z];
(vi) Im f = (Im f )[Z];
(vii) If I is a closed ideal of A, then I[Z] is a closed ideal of (A[Z], k · kK). In
particular,
and
0 −→ I[Z] ֒→ A[Z]
p′
−→ A[Z]/I[Z] −→ 0
0 −→ I[Z] ֒→ A[Z]
p
−→ (A/I)[Z] −→ 0
are short exact sequences;
(viii) [A ⊕ B]K ∼= [A]K ⊕ [B]K;
(ix) \(A ⊕ B)1
∼= A1 ⊕ B1.
Proof. The proof of (iv) follows from (ii) and (iii). The proofs of (v) and (vi)
are straightforward and the proof of (ix) is similar to Part (viii). We prove the
others.
(i) suppose that F (Z) =P∞
n=0 anZ n ∈ A[Z]. Then
k f (F )kK = kP∞
n=0 f (an)Z nkK
= supt∈K kP∞
= supt∈K kf (P∞
≤ supt∈K kP∞
= kF kK.
n=0 f (an)tnk
n=0 antn)k
n=0 antnk
(ii) If f is an isometry, then the proof of (i) shows that k f (F )kK = kF kK, for
all F ∈ A[Z]. That is f and consequently f is an isometry.
8
(iii) Let f be surjective and G = P∞
n=0 bnZ n ∈ B[Z]. For any integer n ≥ 0,
there exists an ∈ A such that bn = f (an). For any integer n ≥ 0 there
exists a′
n ∈ ker f such that
kan + a′
nk ≤ kan + ker f k + 2−n.
Since A/ ker f ∼= B, we have
kan + ker f k = kf (an)k = kbnk.
(9)
(10)
Define a′′
F (Z) =P∞
f (F ) = G.
n = an + a′
n=0 a′′
n, for all n ≥ 0. Now we see from (9) and (10) that
n) = bn, for each n ≥ 0, and therefore
nZ n ∈ A[Z] and f (a′′
(vii) Exactness of first diagram is clear. Part (iii) shows that A[Z]
p
→ (A/I)[Z]
−→ A/I is surjective. By (v) ker p = I[Z] is
p
induces by the projection A
a closed ideal of A[Z]. This completes the proof.
(viii) It is easily seen that
T : A[Z] ⊕ B[Z] −→ (A ⊕ B)[Z]
defined by
T (
anZ n,
∞Xn=0
∞Xn=0
n=0 anZ n ∈ A[Z] and P∞
for all P∞
bnZ n) =
∞Xn=0
(an, bn)Z n,
n=0 bnZ n ∈ B[Z], is a ∗-isomorphism.
3 Exactness of the functor [·]K
In this section we show that [·]K is an exact functor. We first recall some
definitions of the category theory [11].
f
g
Recall that a map X
−→ Z and Y
−→ Y in a category C is called an epimorphism if for
h−→ Z in C with g ◦ f = h ◦ f , we have g = h. In
all maps Y
the category of C∗-algebras, a ∗-homomorphism f : A → B is an epimorphism
if and only if it is surjective [13].
f
Suppose that X
j
−→ X is a kernel of f if f ◦ j = 0 and for any map Z ′
−→ Y is a map in a category C with zero object. A map
Z
−→ X in C such that
f ◦ g = 0, there exists a unique map Z ′ h−→ Z such that j ◦ h = g. For example,
f
−→ B is a ∗-homomorphism of C∗-algebras, then the inclusion ker f ֒→ A
if A
is a kernel of f .
g
Theorem 1. The functor [·]K is exact.
9
Proof. Suppose that
0 −→ A
f
−→ B
g
−→ C −→ 0
is a short exact sequence of C∗-algebras. We must show that
0 −→ [A]K
f
−→ [B]K
g
−→ [C]K −→ 0
(11)
is a short exact sequence of C∗-algebras. We first show that if f : A → B is
a surjective ∗-homomorphism of C∗-algebras, then f : [A]K → [B]K is also a
h→ C
surjective ∗-homomorphism of C∗-algebras. To do this suppose that [B]K
g
→ C are ∗-homomorphism of C∗-algebras such that h ◦ f = g ◦ f .
and [B]K
From Proposition 8 (iii), we have f (A[Z]) = B[Z]. So for any G(Z) ∈ B[Z]
there exists an element F (Z) ∈ A[Z] such that
h(G(Z)) = h( f (F (Z)))
= (g ◦ f )(F (Z))
= g(G(Z)).
This implies that hB[Z] = gB[Z] and therefore g = h. Hence f is an epimorphism
and consequently is surjective by [13].
Now we show that if A
g
ker f = [ker f ]K. To prove this, suppose that C
of C∗-algebras such that f ◦ g = 0. If gC = pA ◦ g and C
then gC ◦ iC = g. Since ker f
∗-homomorphism C h−→ ker f such that the diagram
f
−→ B is a ∗-homomorphism of C∗-algebras, then
−→ [A]K is a ∗-homomorphism
iC֒→ [C]K is the injection,
j
֒→ A is a kernel of f , there exists a unique
ker f
h
f
B
A
gC
C
is commutative. Since [·]K is a functor we get the commutative diagram
[ker f ]K
j
h
f
[B]K
[A]K
gC
[C]K
10
Putting h′ = h ◦ iC we get j ◦ h′ = g, since j ◦ h = gC. Now we show that
k−→ [ker f ]K such
h′ is unique. Suppose that there is a ∗-homomorphism C
that j ◦ k = g = j ◦ h′. Since j is an injection, then k = h′, which proves the
uniqueness of h′. It is clear that ker f = [ker f ]K. Now the Parts (ii), (v) and
(vi) of Proposition 8 imply that (11) is a short exact sequence of C∗-algebras,
or equivalently [·]K is an exact functor. The diagram
[ker f ]K
j
[A]K
f
[B]K
h
gC
k
[C]K
g
iC
C
shows the detials above.
Corollary 1. If I is a closed ideal of a C∗-algebra A, then [A/I]K ∼= [A]K/[I]K.
Proof. By Theorem 1, the short exact sequence
0 −→ I ֒→ A −→ A/I −→ 0
induces the short exact sequence
0 −→ [I]K ֒→ [A]K −→ [A/I]K −→ 0
which implies that [A]K/[I]K ∼= [A/I]K.
In the following corollary we use 3 × 3 lemma in homological algebra for the
C∗-algebras as complex vector spaces (see, e.g., [14]).
Corollary 2. If
0 −→ A
f
−→ B
g
−→ C −→ 0
is a short exact sequence of C∗-algebras, then
0 −→ A1
f A1−→ B1
g B1−→ C1 −→ 0
is also a short exact sequence of C∗-algebras, i.e., F is an exact functor (see
Remark 2). Furthermore, if I is a closed ideal of A, then (dA/I)1 ∼= A1/ I1
Proof. In the commutative diagram
11
0
0
0
f A1
f
f
0
A1
[A]K
pA
A
0
0
B1
[B]K
pB
B
0
g B1
g
g
0
C1
[C]K
pC
C
0
0
0
0
the middle row is exact by Theorem 1 and all columns are exact by (7). Now
3 × 3 Lemma [14] shows that the top row is also exact. By a similar argument
as in Corollary, 1 we get
(dA/I)1 ∼= A1/ I1.
Recall that an ideal I of a C∗-algebra A is called modular if there is an
element u ∈ A such that ua − a, au − a ∈ A, for all element a ∈ A. Note that
I is modular if and only if A/I is unital [12].
Corollary 3. Let I be a closed ideal of a C∗-algebra A. Then I is a modular
ideal of A if and only if [I]K is a modular ideal of [A]K.
Proof. We first show that a C∗-algebra B is unital if and only if [B]K is unital.
It can be easily seen that if B is unital, then [B]K is also unital. Now, by
Proposition 7 (i) suppose that [B]K is unital with unit a + G for some a ∈ B and
G ∈ B1. Consider an arbitrary element b + F ∈ [B]K with b ∈ B and F ∈ B1.
Then (b + F )(a + G) = b + F or equivalently ba + F G + F a + bG = b + F . It
follows that ba − b = H, for some H ∈ B1. Since B ∩ B1 = 0, then ba = b.
Similarly ab = b. This shows that a is the unit of B. Now let I be a closed ideal
of A. Then by Corollary 1, I is modular if and only if [A/I]K ∼= [A]K/[I]K is
unital. Hence I is modular if and only if [I]K is modular.
12
4 Normal exactness of the functor P
Suppose that A is a C∗-algebra and I is a closed ideal of A. It follows from
Proposition 8 (vii) that A[Z]/I[Z] is a normed algebra with the usual quotient
norm. In this section, we show that A[Z]/I[Z] is a pre-C∗-algebra. Also using
Five Lemma and Theorem 2 below, we will show that A[Z]/I[Z] is isometric
∗-isomorphic to (A/I)[Z]. This implies that the functor P is, in fact, normal
exact.
We remind that the Five Lemma in homological algebra (see, e.g., [14]) says
that in the commutative diagram
A1
t1
B1
A2
t2
B2
A3
t3
B3
A4
t4
B4
A5
t5
B5
of commutative R-modules with exact rows if t1, t2, t4 and t5 are isomorphisms,
so is t3.
Definition 1. [16] The exact sequence
· · · −→ An
fn−→ An+1
fn+1−→ An+2 −→ · · ·
of normed spaces with contraction fn (kfnk ≤ 1 for any n) is called normal exact
if the induced map An/ ker fn −→ fn(An) defined by x + ker fn 7−→ fn(x), is an
isometry. Note that any short exact sequence of C∗-algebras is normal exact.
The following theorem is the main one in [16].
Theorem 2. Suppose that
0 −→ Y
i−→ X
p
−→ Z −→ 0
is a normal exact sequence of normed spaces. Then
0 −→ Y
i
−→ X
p
−→ Z −→ 0
is a normal exact sequence of corresponding completion Banach spaces.
Theorem 3. Let I be a closed ideal of a C∗-algebra A. Then A[Z]/I[Z] is a
pre-C∗-algebra.
Proof. We first show that
(i) If (uλ)λ∈Λ is an approximate unit for A, then (uλ)λ∈Λ is also an approximate
unit for A[Z];
13
(ii) If (uλ)λ∈Λ is an approximate unit for I, then for any F (Z) ∈ A[Z] we have
kF (Z) + I[Z]k = limλ kF (Z) − uλF (Z)kK
= limλ kF (Z) − F (Z)uλkK.
n=0 anZ n ∈ A[Z] and ε > 0 be given. Since
n=N +1 2kank < ε.
Now for any λ ∈ Λ we have
To prove (i), let F (Z) = P∞
P∞
n=0 kank < ∞, there is a positive integer N such that P∞
kF (Z) − uλF (Z)kK = kP∞
≤ P∞
= PN
< PN
n=0 kan − uλank +P∞
n=0(an − uλan)Z nkK
n=0 kan − uλank
n=0 kan − uλank + ε.
n=N +1 kan − uλank
Therefore
lim
λ
sup kF (Z) − uλF (Z)kK ≤ ε.
Since ε > 0 was arbitrary, we have
lim
λ
lim
λ
kF (Z) − uλF (Z)kK = 0.
kF (Z) − F (Z)uλkK = 0.
Similarly we get
To prove (ii) let
α = kF (Z) + I[Z]k = inf{kF (Z) + H(Z)kK : H(Z) ∈ I[Z]}.
Let ε > 0 be given. There exists an element G(Z) ∈ I[Z] such that kF (Z) −
G(Z)kK < α + ε. We have
α ≤ kF (Z) − F (Z)uλkK
≤ k(F (Z) − G(Z)) − (F (Z) − G(Z))uλkK + kG(Z) − G(Z)uλkK
= k(F (Z) − G(Z))(1 − uλ)kK + kG(Z) − G(Z)uλkK
≤ kF (Z) − G(Z)kK + kG(Z) − G(Z)uλkK
< α + ε + kG(Z) − G(Z)uλkK.
Now by Part (i) we have
α ≤ limλ inf kF (Z) − F (Z)uλkK
≤ α + ε,
α ≤ limλ sup kF (Z) − F (Z)uλkK ≤ α + ε.
14
Since ε > 0 was arbitrary, we have α = limλ kF (Z) − F (Z)uλkK. Similarly,
α = limλ kF (Z) − uλF (Z)kK.
To prove the theorem let (uλ)λ∈Λ be an approximate unit for I. If F (Z) ∈
A[Z] and G(Z) ∈ I[Z], by Parts (i), (ii) and Proposition 5 (i) we have
kF (Z) + I[Z]k2 = limλ kF (Z) − F (Z)uλk2
K
= limλ k(1 − uλ)F ∗(Z)F (Z)(1 − uλ)kK
≤ limλ k(1 − uλ)(F ∗(Z)F (Z) + G(Z))(1 − uλ)kK
+ limλ k(1 − uλ)G(Z)(1 − uλ)kK
≤ kF ∗(Z)F (Z) + G(Z)kK.
Therefore
kF (Z) + I[Z]k2 ≤ kF ∗(Z)F (Z) + I[Z]k
and consequently we get the equality
kF (Z) + I[Z]k2 = kF ∗(Z)F (Z) + I[Z]k,
which completes the proof.
Now we are ready to show that the functor P is normal exact.
Theorem 4. The functor P is normal exact.
Proof. Let I be a closed ideal of a C∗-algebra A. First we show that there
exists an isometric ∗-isomorphism between A[Z]/I[Z] and (A/I)[Z]. Define
T : A[Z]/I[Z] −→ (A/I)[Z] by
T (
∞Xn=0
anZ n + I[Z]) =
∞Xn=0
(an + I)Z n,
n=0 anZ n ∈ A[Z]. It is clear that T is well defined, linear and preserves
the involution. We are going to show that (a) T is injective, (b) T is surjective,
(c) T is a contraction, and (d) T is an isometry. We proceed as follows:
n=0(an + I)Z n = I,
i.e., an ∈ I for n = 0, 1, 2, · · · . Therefore F (Z) ∈ I[Z] and so T is
injective.
n=0 anZ n ∈ A[Z] with T (F ) = I, thenP∞
for allP∞
(a) If F (Z) =P∞
(b) Let G =P∞
n=0(an + I)Z n ∈ (A/I)[Z]. For each n = 0, 1, 2, · · · there is an
n=0 cnZ n,
n=0 kan + Ik < ∞
element bn ∈ I such that kan + bnk < kan + Ik + 2−n. Let P∞
where cn = an + bn for each n = 0, 1, 2, · · · . Since P∞
we have F (Z) ∈ A[Z]. Therefore
T (F (Z) + I[Z]) = P∞
= P∞
= G,
n=0(cn + I)Z n
n=0(an + I)Z n
15
that is T is surjective.
(c) Let F (Z) =P∞
n=0 anZ n ∈ A[Z]. Then
kT (F (Z) + I[Z])k = kP∞
n=0(an + I)Z nkK
n=0(an + I)tnk
n=0 antn + Ik
= supt∈K kP∞
= supt∈K kP∞
= supt∈K inf b∈I kP∞
≤ inf b∈I supt∈K kP∞
= inf G(Z)∈I[Z] supt∈K kP∞
= inf G kF (Z) + G(Z)kK
n=0 antn + bk
n=0 antn + bk
n=0 antn + G(t)k
= kF (Z) + I[Z]k,
that is T is a contraction. (Note that sup inf f ≤ inf sup f for every real
valued function f in two variables)
(d) Suppose that \(A[Z]/I[Z]) is the completion of A[Z]/I[Z] with respect to
the quotient norm and
T : ( \A[Z]/I[Z]) −→ [A/I]K,
is the extension of T . By Theorem 3, T is a ∗-homomorphism of C∗-
algebras. Now we show that T is a ∗-isomorphism. The diagram
n=0 anZ n
P∞
n=0 anZ n
P∞
p′
p
n=0 anZ n + I[Z]
P∞
T
n=0(an + I)Z n
P∞
shows that the diagram
0
0
I[Z]
A[Z]
I[Z]
A[Z]
p′
p
A[Z]/I[Z]
T
(A/I)[Z]
0
0
of pre-C∗-algebras is commutative, where p′ is the quotient map and p is
the map induced by the projection A
−→ A/I (see Definition 1). The ex-
actness of two rows follow from Proposition 8 (vii). Now the commutative
diagram
p
16
0
0
[I]K
[A]K
[I]K
[A]K
p′
p
( \A[Z]/I[Z])
T
[A/I]K
0
0
of C∗-algebras have exact rows.
In fact, the exactness of first row is a
consequence of Theorem 2 and the second one follows from Theorem 1.
Applying Five Lemma for commutative diagram
[I]K
[A]K
t1
t2
[I]K
[A]K
p′
p
( \A[Z]/I[Z])
t3 = T
[A/I]K
t4
0
0
t5
0
0
with exact rows shows that T is a ∗-isomorphism. This implies, partic-
ularly, that T is an isometry. Now consider the short exact sequence of
C∗-algebras
0 −→ I
i
֒→ A
g
−→ B −→ 0.
Applying functor P we get a short exact sequence of pre-C∗-algebras
0 −→ I[Z]
i
֒→ A[Z]
g
−→ B[Z] −→ 0.
(12)
Note that we have the ∗-isomorphism g1 : A/I −→ B, induced by g. By
Part (d) we have the composition of isometric ∗-isomorphism of pre-C∗-
algebras
A[Z]/I[Z] T−→ (A/I)(Z)
g1−→ B[Z]
such that
∞Xn=0
anZ n + I[Z] 7→
∞Xn=0
(an + I)Z n 7→
∞Xn=0
g(an)Z n.
That, is the induced map A[Z]/I[Z] −→ B[Z] by g is an isometry. There-
fore (12) is a normal exact sequence of pre-C∗-algebras.
From (c) and (d) of Theorem 4 we have the following.
Corollary 4. Suppose that I is a closed ideal of a C∗-algebra A and a0, a1,
a2, · · · , is a sequence in A such that P∞
n=0 kank < ∞. Then
inf
b∈I
sup
t∈K
k
∞Xn=0
antn + bk = sup
t∈K
k
inf
b∈I
17
∞Xn=0
antn + bk.
5 Cauchy extension [A]K as C∗-subalgebra of
Cb(K, A)
In this section, we characterize the Cauchy extensions of C∗-algebras as C∗-
valued function spaces. Using the obtained characterization, we give some re-
sults on the Cauchy extensions of C∗-algebras.
Recall that for a C∗-algebra A and a topological space X, Cb(X, A) is the set
of all bounded continuous functions from X to A. The addition, scalar multipli-
cation and the product on Cb(X, A) are defined pointwise. The involution can
be defined as α∗(x) = (α(x))∗, for all α ∈ Cb(X, A) and x ∈ X. Furthermore,
defining kαk∞ = supx∈X kα(x)k for all α ∈ Cb(X, A), the algebra Cb(X, A) be-
comes a C∗-algebra. If X is a locally compact Hausdroff space, then C0(X, A)
consisting of all continuous functions f ∈ Cb(X, A) vanishing at infinity is a
C∗-subalgebra of Cb(X, A) (see [12, p.37] ). If X is a compact Hausdorff space,
then Cb(X, A) = C0(X, A) = C(X, A).
It is easy to see that for C∗-algebras A1, A2, · · · , An, we have
Cb(X, A1 ⊕ · · · ⊕ An) ∼= Cb(X, A1) ⊕ · · · ⊕ Cb(X, An).
(13)
In particular, if A = C, we use C(X), Cb(X) and C0(X) for C(X, C), Cb(X, C)
and C0(X, C), respectively. Recall that a C∗-algebra A is called nuclear if for
each C∗-algebra B, there is a unique C∗-norm on tensor product A ⊗ B. An
ideal I of a C∗-algebra A is called essential if aI = 0 implies that a = 0.
Theorem 5. Suppose that A and B are two C∗-algebras and K ⊆ J = [−1, 1]
such that 0 is a limit point of K. Then
(i) [A]K is ∗-isomorphic to a C∗-subalgebra of Cb(K, A);
(ii) If K is a compact interval, then [A]K ∼= C(K, A);
(iii) [A]K ∼= {f K : f ∈ C([−1, 1], A)};
(iv) If K is compact, then [A]K ∼= C(K, A). Furthermore, [A ⊗ B]K ∼= [A]K ⊗
B ∼= A ⊗ [B]K;
(v) There is a closed ideal IK of [A]J such that [A]J /IK ∼= [A]K;
(vi) A is nuclear if and only if [A]K is nuclear;
(vii) I is an essential ideal of A if and only if [I]K is an essential ideal of
[A]K ;
(viii) If 0 /∈ K and K is a locally compact subspace of J such that K ′ =
K ∪ {0} is compact, then [A]K ∼= C(K ′, A). If A is finite dimensional,
then M ( A1) ∼= Cb(K, A), where M ( A1) is the multiplier algebra of A1;
(ix) A ∼= B if and only if [A]K ∼= [B]K for any compact set K.
18
Proof.
n=0 antn) = P∞
(i) It is clear that for any sequence (an) in A with P∞
tion f (t) = P∞
T : A(K) −→ A[Z] defined by T (P∞
n=0 kank < ∞ the summa-
n=0 antn, where t ∈ K defines a function from K to A. Denote
the set of all such functions by A(K). It is clear that f is a bounded contin-
uous function on K and A(K) is a ∗-subalgebra of Cb(K, A). Now the map
n=0 anZ n is an isometric
∗-isomorphism. That is [A]K is ∗-isomorphic to a C∗-subalgebra of Cb(K, A).
(ii) For the case that A = C, since C(K) is a self-adjoint subalgebra of C(K)
which separate points of K and contains the constant functions one can see, by
Stone-Weierstrass Theorem (see [15, p.165]), that [C]K ∼= C(K). Now for any
C∗-algebra A and any compact interval K one can use approximate Berstein
Theorem (see [2, p.182] ), as follows: We may assume that K = [0, 1]. Let
f ∈ C(K, A). Because f is uniformly continuous (see [8, p.60] ), define the
Bernstein Polynomials
βn(t) =
f (m/n)(cid:18) n
m(cid:19)tm(1 − t)n−m,
nXm=0
for any t ∈ K and integer n > 0. Note that βn ∈ A(K) for any n = 1, 2, 3, · · · .
By a similar argument as in the proof of the Berstein Theorem, we see that βn
is convergent uniformly to f . This shows that [A]K ∼= C(K, A).
(iii) Define T : A(J) −→ A(K) by T (f ) = f K, for each f ∈ A(J). It is clear
that T is a bijective bounded linear operator. We claim that the extension
T : [A]J −→ [A]K is surjective. Note that Parts (i) and (ii) imply that T is of
the form T (g) = gK for all g ∈ [A]J . Suppose that H, G : [A]K −→ B are two
∗-homomorphisms such that G ◦ T = H ◦ T . This implies that H ◦ T A(J) =
G ◦ T A(J) or HA(K) = GA(K). Since A(K) ∼= A[Z] is dense in [A]K , then
H = G. Hence T is surjective (see [13]). By (ii) we have [A]J ∼= C(J, A) and
therefore [A]K ∼= {f K : f ∈ C(J, A)}.
(iv) By Tietze's Theorem ([9, Theorem 4.1]), any continuous function f : K −→
A has a continuous extension f1 : J −→ A. This fact together with Part (iii)
show that [A]K ∼= C(K, A). From ([3, II.6.4.4] ) we have C(K, A) ∼= C(K) ⊗ A
and therefore
[A ⊗ B]K ∼= C(K) ⊗ (A ⊗ B) ∼= [A]K ⊗ B ∼= A ⊗ [B]K.
(v) Let T : [A]J −→ [A]K be the given surjective ∗-homomorphism in Part (iii).
If IK = ker T , then [A]J /IK ∼= [A]K .
In fact, [A]J is an extension of any
Cauchy extension [A]K .
(vi) Let A be nuclear. By Part (ii) we have [A]J ∼= C(J, A). Since C(J) is
nuclear (see [12, Theorem 6.4.15]) and C(J, A) ∼= C(J) ⊗ A (see [3, II.6.4.4])
we imply that [A]J is nuclear (see [3, IV.3.1.1]). Since every closed ideal of a
nuclear C∗-algebra is nuclear (see [3, II.9.6.3]), then A1 is nuclear. In particular,
the closed ideal IK (given in part (v)) is nuclear. Since [A]J /IK is nuclear (see
19
[3, IV 3.1.13]), Part (v) implies that [A]K is also nuclear. Conversely, if [A]K
is nuclear, then the ideal A1 is nuclear. By (7) we have A ∼= [A]K / A1 which
shows that A is nuclear too.
(vii) Let I be an essential ideal of A. By Part (i) we can consider [A]K as a C∗-
subalgebra of Cb(K, A). Choose G : K −→ A in [A]K such that f G = Gf = 0
for any f : K −→ I ∈ [I]K . For any t ∈ K we have f (t)G(t) = G(t)f (t) = 0.
Let b be an arbitrary element in I and let fb : K −→ I be a constant function
with value fb(t) = b. Now for any t ∈ K we have
or
fb(t)G(t) = G(t)fb(t) = 0
bG(t) = G(t)b = 0.
This implies that G(t) = 0 for all t ∈ K. Therefore [I]K is an essential ideal of
[A]K. The converse statement can be proved similarly.
(viii) Suppose that C1(K) = {f ∈ C(K) : f (0) = 0}, where C(K) is as given in
Part (i). For f ∈ C1(K) and ε > 0, suppose that X = {t ∈ K : f (t) ≥ ε}
and x is a limit point of X. Then x 6= 0 and x is a limit point of K ′, and
therefore x ∈ K. This implies that X is compact. That is f vanishes at infinity,
so C1(K) ⊆ C0(K). Now suppose that 0 6= a ∈ C and g(x) = xa for all x ∈ K.
Then g ∈ C1(K) and for any t ∈ K we have g(t) 6= 0. In addition, for any t1 6= t2
in K, g(t1) 6= g(t2), that is, C1(K) strongly separates points of K. It is clear
that C1(K) is self-adjoint. By the Stone-Weierstrass Theorem (see [7, p.151])
we have C1 ∼= C0(K) and therefore [C]K ∼= C ⊕ C0(K) ∼= C(K ′) (see [3, p.53]).
Parts (iii) and (iv) and the fact that kf kK = kf kK ′ for any f ∈ C(J, A) imply
that the map f K 7→ f K ′ is a ∗-isomorphism between [A]K and C(K ′, A). Now
suppose that A is a finite dimensional C∗-algebra. By ([12, p.194]) we have
A ∼= Mn1(C) ⊕ Mn2(C) ⊕ · · · ⊕ Mnm(C).
(14)
We first show that for any positive integer n, \(Mn(C))1 ∼= Mn( C1). To see this,
note that the completion of C1(K) is ∗-isomorphic to C0(K). Now the map
G : (Mn(C))1(K) −→ Mn(C1(K)) defined by G(F ) = (Fij ), where
F (t) =
∞Xm=1
Bmtm = (Fij (t))
and Fij ∈ C1(K), for any i, j = 1, 2, · · · , n is an isometric ∗-isomorphism with
norm k(Fij)k = supt∈K kFij (t)k = kF k. Suppose that F = (Fij ) ∈ Mn(C0(K)),
then Fij ∈ C0(K) for i, j = 1, 2, · · · , n. There exist sequences (Fmij ) in C1(K)
for i, j = 1, 2, · · · , n such that Fmij −→ Fij as m → ∞ in norm k · kK. If F :
K −→ Mn(C) is a continuous function such that for any t ∈ K, F (t) = (Fij (t)),
then
k(Fmij ) − (Fij )k = supt∈K k(Fmij(t)) − Fij (t)k
≤ supt∈KPi,j kFmij (t) − Fij (t)k
≤ Pi,j supt∈K kFmij (t) − Fij (t)k.
20
This implies that (Fmij ) −→ (Fij ) as m −→ ∞. Now by completion we see that
\(Mn(C))1
∼= Mn(C0(K)) ∼= Mn( C1).
Also we have clearly the ∗-isomorphism
Mn(Cb(K)) ∼= Cb(K, Mn(C)).
(15)
(16)
Suppose that B, A1, A2, · · · , An are C∗-algebras. We have the following for the
multipliers algebras (see [4, p.84])
M (Mn(B)) ∼= Mn(M (B))
(17)
M (A1 ⊕ A2 ⊕ · · · ⊕ An) ∼= M (A1) ⊕ M (A2) ⊕ · · · ⊕ M (An),
(18)
(see [3, p.155]). We also have M (C0(K)) ∼= Cb(K) (see [12, p.83]). Now
from (13) − (18), and Proposition 8 (ix), we have
A1 ∼= Mn1(C0(K)) ⊕ Mn2(C0(K)) ⊕ · · · ⊕ Mnm(C0(K)).
M ( A1) ∼= M (Mn1(C0(K))) ⊕ · · · ⊕ M (Mnm(C0(K)))
∼= Cb(K, Mn1(C)) ⊕ · · · ⊕ Cb(K, Mnm(C))
∼= Cb(K, Mn1(C) ⊕ · · · ⊕ Mnm(C))
∼= Cb(K, A).
(ix) If A ∼= B, then [A]K ∼= [B]K by Proposition 8 (iv). Let ϕn : [A]Kn −→ [B]Kn
be a ∗-isomorphism between [A]Kn and [B]Kn, where Kn = [−1/n, 1/n] for
n=1 Kn = {0}. Now
n = 1, 2, 3, · · · .
([A]Kn , pn)∞
It is clear that (Kn) is nested with T∞
n=1 is a direct sequence of C∗-algebras, where each map
pn : [A]Kn −→ [A]Kn+1
defined by f Kn 7→ f Kn+1, for all f ∈ [A]K is a ∗-homomorphism. Part (iv)
and [3, II.6.4.4] show that
[A]Kn
∼= C(Kn, A) ∼= C(Kn) ⊗ A,
for all n. Furthermore by [3, II.9.6.5] we have the direct limit
[A]Kn
lim
−→
∼= lim
−→
(C(Kn) ⊗ A) ∼= (lim
−→
C(Kn)) ⊗ A ∼= C({0}) ⊗ A ∼= C ⊗ A ∼= A.
From the commutative diagram
[A]Kn
pn
ϕn
[B]Kn
qn
[A]Kn+1
ϕn+1
[B]Kn+1
21
where ([B]Kn, qn)∞
for any f ∈ [A]Kn , we conclude that
n=1 is the direct sequence defined by qn(ϕn(f )) = ϕn+1(f Kn+1),
A ∼= lim
−→
[A]Kn
∼= lim
−→
[B]Kn
∼= B,
as desired.
Any C∗-algebra of the form
B = Mn1(C[a1, b1]) ⊕ · · · ⊕ Mnm(C[an, bn])
where ai < bi for i = 1, 2, · · · , n are real numbers, is a Cauchy extension of some
C∗-algebra. In fact
B ∼= Mn1(C[−1, 1]) ⊕ Mn2 (C[−1, 1]) ⊕ · · · ⊕ Mnm(C[−1, 1]).
Therefore B ∼= [A]J , where A is the C∗-algebra defined in (14).
Corollary 5. Suppose that A is a C∗-algebra and I is a closed ideal of A. If
K = [0, 1] and F ∈ C(K, A), then
inf
b∈I
sup
t∈K
kF (t) + bk = sup
t∈K
inf
b∈I
kF (t) + bk.
Proof. Let ε > 0 be given. By Theorem 5 (ii) there exists an element Fn ∈ A(K)
such that supt∈K kF (t) − Fn(t)k < ε. For any t ∈ K we have
kF (t) + bk ≤ kF (t) − Fn(t)k + kFn(t) + bk < ε + kFn(t) + bk.
On the other hand
kFn(t) + bk ≤ kFn(t) − F (t)k + kF (t) + bk < ε + kF (t) + bk,
for any t ∈ K. By Corollary 4 we have
inf b∈I supt∈K kF (t) + bk ≤ ε + inf b∈I supt∈K kFn(t) + bk
= ε + supt∈K inf b∈I kFn(t) + bk
≤ 2ε + supt∈K inf b∈I kF (t) + bk.
Since ε > 0 was arbitrary, then
inf
b∈I
sup
t∈K
kF (t) + bk ≤ sup
t∈K
inf
b∈I
kF (t) + bk.
This completes the proof.
22
References
[1] Arveson, W. Notes on extensions of C
∗
-algebras. Duke Math. J. 44 (1977),
no. 2, 329 -- 355.
[2] Bartle, R.G., The elements of real analysis. Second edition. John Wiley &
Sons, New York-London-Sydney, 1976.
[3] Blackadar, B., Operator algebras. Theory of C∗ -algebras and Von Neu-
mann algebras. Encyclopaedia of Mathematical Sciences, 122. Operator Al-
gebras and Non-commutative Geometry, III. Springer-Verlag, Berlin, 2006.
[4] Blecher, D.P.; Le Merdy, C. Operator algebras and their modules an op-
erator space approach. London Mathematical Society Monographs. New
Series, 30. Oxford Science Publications. The Clarendon Press, Oxford Uni-
versity Press, Oxford, 2004.
[5] Brown, L.G.; Douglas, R. G.; Fillmore, P. A. Extensions of C∗-algebras
and K-homology. Ann. of Math. (2) 105 (1977), no. 2, 265 -- 324.
[6] Busby, R.C., Double centralizers and extensions of C∗ -algebras. Trans.
Amer. Math. Soc. 132 (1968) 79 -- 99.
[7] Conway, J.B., A course in functional analysis. Graduate Texts in Mathe-
matics, 96. Springer-Verlag, New York, 1985.
[8] Dieudonne, J., Foundations of modern analysis. Enlarged and corrected
printing. Pure and Applied Mathematics, Vol. 10-I. Academic Press, New
York-London, 1969.
[9] Dugundji, J., An extension of Tietze's theorem. Pacific J. Math. 1, (1951).
353 -- 367.
[10] Kasparov, G.G., The operator K -functor and extensions of C∗ -algebras.
(Russian) Izv. Akad. Nauk SSSR Ser. Mat. 44 (1980), no. 3, 571 -- 636, 719.
[11] MacLane, S., Categories for the working mathematician. Graduate Texts
in Mathematics, Vol. 5. Springer-Verlag, New York-Berlin, 1971.
[12] Murphy, G.J., C∗ -algebras and operator theory. Academic Press, Inc.,
Boston, MA, 1990.
[13] Reid, G. A., Epimorphisms and surjectivity. Invent. Math. 9 (1969/1970),
295 -- 307.
[14] Rotman, J.J., An introduction to homological algebra. Second edition. Uni-
versitext. Springer, New York, 2009.
[15] Rudin, W., Principles of mathematical analysis. Third edition. Interna-
tional Series in Pure and Applied Mathematics. McGraw-Hill Book Co.,
New York-Auckland-DCsseldorf, 1976.
23
[16] Yang, K.W., Completion of normed linear spaces. Proc. Amer. Math. Soc.
19 (1968), 801 -- 806.
24
|
1201.3879 | 1 | 1201 | 2012-01-18T19:01:49 | The generator problem for Z-stable C*-algebras | [
"math.OA"
] | The generator problem was posed by Kadison in 1967, and it remains open until today. We provide a solution for the class of C*-algebras absorbing the Jiang-Su algebra Z tensorially. More precisely, we show that every unital, separable, Z-stable C*-algebras A is singly generated, which means that there exists an element x in A that is not contained in any proper sub-C*-algebra of A.
To give applications of our result, we observe that Z can be embedded into the reduced group C*-algebra of a discrete group that contains a non-cyclic, free subgroup. It follows that certain tensor products with reduced group C*-algebras are singly generated. In particular, the tensor product of two reduced free group C*-algebras is singly generated. | math.OA | math | THE GENERATOR PROBLEM FOR Z-STABLE C ∗-ALGEBRAS
HANNES THIEL AND WILHELM WINTER
ABSTRACT. The generator problem was posed by Kadison in 1967, and it remains
open until today. We provide a solution for the class of C ∗-algebras absorbing the
Jiang-Su algebra Z tensorially. More precisely, we show that every unital, separa-
ble, Z-stable C ∗-algebra A is singly generated, which means that there exists an
element x ∈ A that is not contained in any proper sub-C ∗-algebra of A.
To give applications of our result, we observe that Z can be embedded into
the reduced group C ∗-algebra of a discrete group that contains a non-cyclic, free
subgroup. It follows that certain tensor products with reduced group C ∗-algebras
are singly generated. In particular, C ∗
r (F∞) is singly generated.
r (F∞) ⊗ C ∗
.
A
O
h
t
a
m
[
1
v
9
7
8
3
.
1
0
2
1
:
v
i
X
r
a
1. INTRODUCTION
By an operator algebra we mean a ∗-subalgebra of B(H) that is either closed in
the norm topology (a concrete C ∗-algebra) or the weak operator topology (a von
Neumann algebra). One way of realizing an operator algebra is to take a subset of
B(H) and consider the smallest operator algebra containing it.
In a trivial way, every operator algebra can be obtained this way. The situation
becomes interesting if one imposes restrictions on the generating set, and one nat-
ural possibility is to require that it consists of only one element, i.e., to consider
operator algebras that are generated by a single operator. It is an old problem to
determine which operator algebras arise this way.
More generally, one tries to compute the minimal number of elements that gen-
erate a given operator algebra, see 2.1.
It is often convenient to consider self-
adjoint generators. Note that two self-adjoint elements a, b generate the same op-
erator algebra as the element a + ib. Thus, if we ask whether an operator algebra is
singly generated, it is equivalent to ask whether it is generated by two self-adjoint
elements.
In the case of von Neumann algebras, the generator problem was included in
Kadison's famous 'Problems on von Neumann algebras', [Kad67]. This problem
list has turned out to be very influential, yet its original form remains unpublished.
It is indirectly available in an article by Ge, [Ge03], where a brief summary of the
developments around Kadison's famous problems is given.
Date: November 12, 2018.
2000 Mathematics Subject Classification. Primary 46L05, 46L85; Secondary 46L35.
Key words and phrases. C ∗-algebras, generator problem, single generation, Z-stability.
This research was partially supported by the Centre de Recerca Matem`atica, Barcelona. The first
named author was partially supported by the Danish National Research Foundation through the Cen-
tre for Symmetry and Deformation, Copenhagen. The second named author was partially supported
by EPSRC Grants EP/G014019/1 and EP/I019227/1.
1
2
HANNES THIEL AND WILHELM WINTER
Question 1.1 (Kadison, [Kad67, Problem 14], see also [Ge03]). Is every separably-
acting1 von Neumann algebra singly generated?
We just mention an incomplete list of results.
As noted in [She09], there exist singly generated von Neumann algebras that
are not separably-acting. However, the separably-acting von Neumann algebras
are the natural class for which one might expect single generation. The answer to
Question 1.1 is still open in general, but many authors have contributed to show
that large classes of separably-acting von Neumann algebras are singly generated.
It starts with von Neumann,
[vN31], who showed that the abelian operator algebras named after him are gener-
ated by a single self-adjoint element, thus implicitly raising the generator problem.
Some thirty years later, this was extended by Pearcy, [Pea62], who showed that
all von Neumann algebras of type I are singly generated. Then Wogen, [Wog69,
Theorem 2], proved that all properly infinite von Neumann algebras are singly
generated, thus reducing the generator problem to the type II1 case.
Later, this was further reduced to the case of a II1-factor by Willig, [Wil74], and
then to the case of a finitely-generated II1-factor by Sherman, [She09, Theorem
3.8]. This means that Question 1.1 has a positive answer if every separably-acting,
finitely generated II1-factor is singly generated.
There are many properties known to imply that a II1-factor is singly generated.
We just mention that Ge and Popa, [GP98, Theorem 6.2], show that every tensori-
ally non-prime2 II1-factor is singly generated. Our main result Theorem 3.5 can be
considered as a partial C ∗-algebraic analog of this result.
Let us also mention that the free group factors W ∗(Fk) are the outstanding
examples of separably-acting von Neumann algebra for which it is not known
whether they are singly generated.
In the case of C ∗-algebras, the generator problem is more subtle. There is al-
ready no obvious class of C ∗-algebras for which one conjectures that they are
singly generated. Every singly generated C ∗-algebra is separable3. However, the
converse is false, and counterexamples can be found among the commutative C ∗-
algebras.
In fact, the C ∗-algebra C0(X) is generated by n self-adjoint elements if and only
if X can be embedded into Rn. Thus, C0(X) is singly generated if and only if X is
planar, i.e., can be embedded into the plane R2.
It is easy to see that a C ∗-algebra A is generated by n self-adjoint elements if and
only if its minimal unitization eA is generated by n self-adjoint elements. Therefore,
we will mostly consider the generator problem for separable, unital C ∗-algebra. In
that case, taking the tensor product with a matrix algebra has the effect of reduc-
ing the necessary number of generators. If A is generated by n2 + 1 self-adjoint
elements, then A ⊗ Mn is singly generated, see e.g. [Nag04, Theorem 3].
One derives the principle that a C ∗-algebra needs less generators if it is 'more
non-commutative'. Consequently, one might expect a (separable) C ∗-algebra to be
singly generated if it is 'maximally non-commutative'. As a non-unital instance
1A von Neumann algebra is called 'separably-acting', or just 'separable', if it is a subalgebra of
B(l2N), or equivalently if it has a separable predual.
2A II1-factor M is called tensorially non-prime if it is isomorphic to a tensor product, M1 ¯⊗M2, of
two II1-factors M1, M2.
3A C ∗-algebra is called 'separable' if it contains a countable, norm-dense subset
THE GENERATOR PROBLEM FOR Z-STABLE C ∗-ALGEBRAS
3
of this principle, we note that the stabilization, A ⊗ K, of a separable unital C ∗-
algebra A is singly generated, [OZ76, Theorem 8]. In the unital case, there are
at least three natural cases when one considers a C ∗-algebra A to be 'maximally
non-commutative', which are the following:
(1) A contains a simple, unital, nonelementary sub-C ∗-algebra,
(2) A contains a sequence of pairwise orthogonal, full elements,
(3) A has no finite-dimensional irreducible representations.
In general, the implications (1) ⇒ (2) ⇒ (3) hold; it is not known if the con-
verses are true.
Conditions (2) and (3) can also be considered for possibly non-unital C ∗-alge-
bras, and we let (2∗) be the weaker statement that A contains two orthogonal, full
elements. The implication '(3) ⇒ (2)' holds exactly if the implication '(3) ⇒ (2∗)'
holds.
The Global Glimm halving problem asks the following: Given a (possibly non-
unital) C ∗-algebra A that satisfies condition (3), does there exist a full map from
the cone over M2 to A? It is not known whether the Global Glimm halving prob-
lem has a positive answer, but if it does then it shows that implication '(3) ⇒ (2)'
holds, since the cone over M2 contains two orthogonal, full elements.
Let us remark that the analogs of conditions (1)−(3) for von Neumann algebras
are all equivalent. In fact, if a von Neumann algebra M has no finite-dimensional
representations, then the hyperfinite II1-factor R unitally embeds into M .
Historically, the generator problem for C ∗-algebras is mostly asked for C ∗-al-
gebras that are simple ore more generally have no finite-dimensional representa-
tions:
Question 1.2. Is every simple, separable, unital C ∗-algebra singly generated?
Question 1.3. Is a separable, unital C ∗-algebra singly generated provided it has
no finite-dimensional irreducible representations?
The answers to both questions are open. A positive answer to Question 1.3
implies a positive answer to Question 1.2, of course. The converse is not clear.
Let us mention some results that solve the generator problem for particular
classes of separable C ∗-algebras.
It was shown by Topping, [Top68], that ev-
ery UHF-algebra is singly generated. This was generalized by Olsen and Zame,
[OZ76, Theorem 9], who showed that the tensor product, A ⊗ B, of any separable,
unital C ∗-algebra A with a UHF-algebra B is singly generated.
Later, it was shown by Li and Shen, [LS10, Theorem 3.1], that every unital, ap-
proximately divisible4 C ∗-algebra is singly generated. This generalizes the result
of Olsen and Zame, since the tensor product with a UHF-algebra is always ap-
proximately divisible.
In this article we prove that every separable, unital, Z-stable C ∗-algebra is
singly generated, see Theorem 3.7. This generalizes the result of Li and Shen,
since every approximately divisible C ∗-algebra is Z-stable, see [TW08, Theorem
2.3]. The notion of Z-stability has proven to be very important in the classifica-
tion program of nuclear C ∗-algebras, see e.g. [Win07] or [ET08], and it is has been
4A unital C ∗-algebra A is 'approximately divisible' if for every ε > 0 and finite subset F ⊂ A
there exists a finite-dimensional, unital sub-C ∗-algebra B ⊂ A such that B has no characters and
kxb − bxk ≤ εkbk for all x ∈ F, b ∈ B.
4
HANNES THIEL AND WILHELM WINTER
shown that many nuclear, simple C ∗-algebras are Z-stable, see e.g. [Win10]. Z-
stability is also relevant in the non-nuclear context; for example, unital Z-stable
C ∗-algebras satisfy Kadison's similarity property, see [JW11].
This paper proceeds as follows:
In Section 2 we set up our notation and give some basic facts about the genera-
tor rank, see 2.1, and C0(X)-algebras, see 2.4.
Section 3 contains the proof of our main result, which states that the tensor
product A ⊗max B of two separable, unital C ∗-algebras is singly generated, if A
satisfies condition (2) from above (e.g. A is simple and non-elementary) and B
admits a unital embedding of the Jiang-Su algebra Z, see Theorem 3.5.
We derive that every separable, unital, Z-stable C ∗-algebra is singly generated,
see Theorem 3.7. Our main result can be considered as a (partial) C ∗-algebraic
analog of a theorem of Ge and Popa, [GP98, Theorem 6.2], which shows that a
tensor product, M ¯⊗N , of two II1-factors M, N is singly generated. In fact, we can
reprove their theorem with our methods, see Corollary 3.11.
In Section 4 we give further applications of our main theorem to tensor prod-
ucts with reduced group C ∗-algebras. We first observe that Z embeds unitally into
r (F∞), the reduced group C ∗-algebra of the free group on infinitely many gen-
C ∗
erators, see Lemma 4.1. Consequently, if a discrete group Γ contains a non-cyclic
free subgroup, then Z embeds unitally into C ∗
r (Γ), see Proposition 4.2.
We deduce that tensor products of the form A ⊗max C ∗
r (Γ) are singly generated
if A is a separable, unital C ∗-algebra satisfying condition (2) from above, and Γ
is a group containing a non-cyclic free subgroup, see Corollary 4.4. For example,
r (F∞) is singly generated, although this C ∗-algebra is not Z-stable,
C ∗
see Example 4.5.
r (F∞) ⊗ C ∗
2. PRELIMINARIES
By a morphism between C ∗-algebras we mean a ∗-homomorphism, and by an
ideal of a C ∗-algebra we understand a closed, two-sided ideal. If A is a C ∗-al-
gebra, then we denote by eA its minimal unitization. Often, we write Mk for the
C ∗-algebra of k-by-k matrices Mk(C).
2.1. Let A be a C ∗-algebra, and Asa ⊂ A the subset of self-adjoint elements. We
say that a set S ⊂ Asa generates A, denoted A = C ∗(S), if the smallest sub-C ∗-
algebra of A containing S is A itself. We denote by gen(A) the smallest number
n ∈ {1, 2, 3, . . . , ∞} such that A contains a generating subset S ⊂ Asa of cardinality
n, and we call gen(A) the generating rank of A.
We stress that for the definition of gen(A), the generators are assumed to be
self-adjoint. Two self-adjoint elements a, b generate the same C ∗-algebra as the
(non-self-adjoint) element a + ib. Therefore, a C ∗-algebra A is said to be singly
generated if gen(A) ≤ 2.
For more details on the generating rank we refer the reader to Nagisa, [Nag04],
where also the following simple facts are noted for C ∗-algebras A and B:
(1) gen(eA) = gen(A),
THE GENERATOR PROBLEM FOR Z-STABLE C ∗-ALGEBRAS
5
(2) gen(C ∗(A, B)) ≤ gen(A) + gen(B), if A, B are sub-C ∗-algebras of a common
C ∗-algebra, and where C ∗(A, B) denotes the sub-C ∗-algebra they generate
together,
(3) gen(A ⊕ B) = max{gen(A), gen(B)} if at least one of the algebras is unital.
Let I ✁ A be an ideal in a C ∗-algebra A.
It is easy to see that the generat-
ing rank of the quotient A/I is not bigger than the generating rank of A, i.e.,
gen(A/I) ≤ gen(A), and the generating rank of A can be estimated as gen(A) ≤
gen(I) + gen(A/I). The following result gives an estimate for gen(I), and it is
probably well-known to experts; since we could not locate it in the literature, we
include a short proof.
Proposition 2.2. Let A be a C ∗-algebra, and I ✁ A an ideal. Then gen(I) ≤ gen(A) + 1.
Proof. We may assume gen(A) is finite. So let a1, . . . , ak be a set of self-adjoint
generators for A. Then A and I are separable, and so I contains a strictly positive
element h. It follows that C ∗(h) contains a quasi-central approximate unit, see
[AP77, Corollary 3.3] and [Arv77]. It is straightforward to show that I is generated
by the k + 1 elements h, ha1h, . . . , hakh.
(cid:3)
The following result is attributed to Kirchberg in [Nag04].
Theorem 2.3 (Kirchberg). Every separable, unital, properly infinite C ∗-algebra is singly
generated.
Proof. We sketch a proof based on the proof of [OZ76, Theorem 9]. Let A be a sep-
arable, unital, properly infinite C ∗-algebra. Then there exist isometries s1, s2, . . . ∈
A with pairwise orthogonal ranges (i.e., A contains a unital copy of the Cuntz
algebra O∞).
Let a1, a2, . . . ∈ A be a sequence of (positive) generators for A such that their
spectra satisfy σ(ak) ⊂ [1/2 · 1/4k, 1/4k]. A generator for A is given by:
x :=Xk≥1
(skaks∗
k + 1/2ksk).
As in in the proof of [OZ76, Theorem 9], one can show that σ(x) ⊂ {0} ∪
Sk≥1[1/2 · 1/4k, 1/4k]. Let B := C ∗(x) ⊂ A. Proceeding inductively, one shows
that ak, sk ∈ B. We only sketch this for k = 1. Set p := s1s∗
1. Let fn be a sequence
of polynomials converging uniformly to 1 on [1/8, 1/4] and to 0 on [0, 1/16]. Then
fn(x) converges to an element y ∈ B of the form y = p + pb(1 − p) for some
b ∈ A. We compute yy∗ = p(1A + b(1 − p)b∗)p. Then for a continuous function
f : R → R with f (0) = 0 and f (t) = 1 for t ≥ 1, we get f (yy∗) = p ∈ B. Then
s1a1s∗
(cid:3)
1 = pxp ∈ B and s1 = 2 · px(1 − p) ∈ B, and then also a1 ∈ B.
2.4. Let X be a locally compact σ-compact Hausdorff space. A C0(X)-algebra is a
C ∗-algebra A together with a morphism η : C0(X) → Z(M (A)), from the commu-
tative C ∗-algebra C0(X) to the center of the multiplier algebra of A, such that for
any approximate unit (uλ)Λ of C0(X), η(uλ)a → a for any a ∈ A, or equivalently,
the closure of η(C0(X))A is all of A. Thus, if X is compact, then η is necessarily
unital. We will usually suppress reference to the structure map, and simply write
f a or f · a instead of η(f )a for the product of a function f ∈ C0(X) and an element
a ∈ A.
6
HANNES THIEL AND WILHELM WINTER
Let Y ⊂ X be a closed subset, and U := X \ Y its complement (an open subset).
Then C0(U ) · A is an ideal of A, denoted by A(U ). The quotient A/A(U ) is denoted
by A(Y ).
Given a point x ∈ X, we write A(x) for A({x}), and we call this C ∗-algebra
the fiber of A at x. For an element a ∈ A, we denote by a(x) the image of a in
the fiber A(x). For each a ∈ A, we may consider the map a : x 7→ ka(x)k. This
is a real-valued, upper-semicontinuous function on X, vanishing at infinity. The
C0(X)-algebra A is called continuous if a is a continuous function for each a ∈ A.
For more information on C0(X)-algebras we refer the reader to [Kas88, §1] or
the more recent [Dad09, §2].
2.5. The Jiang-Su algebra Z was constructed in [JS99]; it may be regarded as a C ∗-
algebraic analog of the hyperfinite II1-factor. It can be obtained as an inductive
limit of prime dimension drop algebras Zp,q := {f : [0, 1] → Mp ⊗ Mq f (0) ∈
1p ⊗ Mq, f (1) ∈ Mp ⊗ 1q}.
For more details, we refer the reader to [Win11], where Z is characterized in an
entirely abstract manner, and to [Rør04] and [RW10], where it is shown that the
generalized dimension drop algebra Z2∞,3∞ := {f : [0, 1] → M2∞ ⊗ M3∞ f (0) ∈
1 ⊗ M3∞, f (1) ∈ M2∞ ⊗ 1} embeds unitally into Z; in fact, Z can be written as a
stationary inductive limit of Z2∞,3∞ .
3. RESULTS
Lemma 3.1. Let A be a separable, unital C ∗-algebra. Then gen(A ⊗ Z2∞,3∞) ≤ 5.
Proof. Consider the ideal I := C0(0, 1) ⊗ M6∞ in B := A ⊗ Z2∞,3∞ . The quotient
B/I is isomorphic to (A⊗M2∞ )⊕(A⊗M3∞). Thus, we have a short exact sequence:
A ⊗ C0(0, 1) ⊗ M6∞
/ A ⊗ Z2∞,3∞
/ (A ⊗ M2∞) ⊕ (A ⊗ M3∞ )
It follows from [OZ76] that the tensor product of a unital, separable C ∗-alge-
bra with a UHF-algebra is singly generated. In particular, gen(A ⊗ M2∞), gen(A ⊗
M3∞) ≤ 2. Thus, the quotient satisfies gen(B/I) = max{gen(A ⊗ M2∞), gen(A ⊗
M3∞)} ≤ 2, see 2.1.
Note that I is an ideal in the C ∗-algebra C := A ⊗ C(S1) ⊗ M2∞ . We have
gen(C) ≤ 2, and then gen(I) ≤ gen(C) + 1 ≤ 3, by Proposition 2.2. Then, the
extension is generated by at most 2 + 3 = 5 self-adjoint elements.
(cid:3)
The following is a Stone-Weierstrass type result. We prove it using the factorial
Stone-Weierstrass conjecture, which states that a sub-C ∗-algebra B ⊂ A exhausts
A if it separates the factorial states of A. The factorial Stone-Weierstrass conjec-
ture was proved for separable C ∗-algebras independently by Longo, [Lon84], and
Popa, [Pop84].
See 2.4 for a short introduction to C0(X)-algebras.
Lemma 3.2. Let A be a separable, continuous C0(X)-algebra, and B ⊂ A a sub-C ∗-al-
gebra such that the following two conditions are satisfied:
(i) For each x ∈ X, B exhausts the fiber A(x),
(ii) B separates the points of X by full elements, i.e., for each distinct pair of points
x0, x1 ∈ X there exists some b ∈ B such that b(x1) is full in B(x1) = A(x1) and
b(x0) = 0.
/
/
THE GENERATOR PROBLEM FOR Z-STABLE C ∗-ALGEBRAS
7
Then A = B.
Condition (ii) is for instance satisfied if B contains the image of the structure map
η : C0(X) → Z(M (A)).
Proof. Set Y := Prim(Z(M (A))), and identify Z(M (A)) with C(Y ). Let π : A →
B(H) be a non-degenerate factor representation. Then π extends to a represen-
It is straightforward to show π(A)′′ = π(M (A))′′,
tation π : M (A) → B(H).
so that π is a factor representation of M (A). For any c ∈ Z(M (A)), we have
c ∈ π(A)′ ∩ π(M (A))′′ = C · 1H. Thus, there exists a point y ∈ Y such that
π(c) = c(y) · 1H for all c ∈ Z(M (A)). Since η(C0(X)) contains an approximate unit
for A, we have that π ◦ η is non-zero. Thus, there exists a point x ∈ X such that
π ◦ η(f ) = f (x) · 1H for all f ∈ C0(X). This means that π ◦ η vanishes on the ideal
A(X \ {x}), so that π factors through the fiber A(x).
Let us show that B ⊂ A separates the factors states of A. So let ϕ1, ϕ2 be two
different, non-degenerate factors states of A. We have shown above that there
are two points x1, x2 ∈ X such that ϕi factors through A(xi), and we denote by
¯ϕi : A(xi) → C the induced factor state on A(xi), for i = 1, 2. We distinguish two
cases:
Case 1: x1 = x2. In this case, since ϕ1 6= ϕ2, there exists an element a ∈ A such
that ϕ1(a) 6= ϕ2(a). By condition (i), there exists some element b ∈ B such that
b(x1) = a(x1). Note that ϕi(b) = ¯ϕi(b(x1)) = ¯ϕi(a(x1)) = ϕi(a), for i = 1, 2. Thus,
b separates the two states.
Case 2: x1 6= x2. In this case, by condition (ii), there exists an element b ∈ B
such that b(x2) is full in A(x2) and b(x1) = 0. Since ϕ2 6= 0, there exists an element
a ∈ A such that ϕ2(a) = ¯ϕ2(a(x2)) ≥ 1.
Since b(x2) is full, there exist finitely many elements gi, hi ∈ A(x2) such that
ka(x2) −Pi cib(x2)dik < 1. By condition (i), there exist elements gi, hi ∈ B such
that gi(x2) = gi and hi(x2) = hi. Set b′ := Pi cib di. Then ϕ2(b′) = ¯ϕ2(b′(x2)) >
0, while b′(x1) = 0. This shows that b′ separates the two states.
It follows that B separates the factor states of A, and therefore B = A by the
factorial Stone-Weierstrass conjecture, proved independently by Longo, [Lon84],
and Popa, [Pop84].
(cid:3)
Lemma 3.3. Let A be a unital C ∗-algebra with gen(A) ≤ 3. Then there exist a positive
element x ∈ A ⊗ Z2,3 and two positive, full elements y′, z′ ∈ Z2,3 such that A ⊗ Z2,3 is
generated by x and 1 ⊗ y′, and further y′ and z′ are orthogonal.
Proof. We consider Z2,3 as the C ∗-algebra of continuous functions from [0, 1] to M6
with the boundary conditions
Y
f (0) =
Y
f (1) =(cid:18)Z
Y
QZQ∗(cid:19) ,
where Y ∈ M2 and Z ∈ M3 are arbitrary matrices, and Q ∈ M3 is the following
fixed permutation matrix:
Q =
1
1
.
1
8
HANNES THIEL AND WILHELM WINTER
This means that f (0), f (1) ∈ M6 have the following form:
f (0) =
f (1) =
µ11 µ12
µ21 µ22
µ11 µ12
µ21 µ22
µ11 µ12
µ21 µ22
λ11 λ12 λ13
λ21 λ22 λ23
λ31 λ22 λ33
,
λ33 λ31 λ32
λ13 λ11 λ12
λ23 λ21 λ22
for numbers µi,j, λi,j ∈ C.
Note that Z2,3 is naturally a continuous C([0, 1])-algebra, with fibers Z2,3(0) ∼=
M2, Z2,3(1) ∼= M3, and Z2,3(t) ∼= M6 for points t ∈ (0, 1) ⊂ [0, 1].
Let a, b, c ∈ A be a set of invertible, positive generators for A. Denote by ei,j the
matrix units in M6. To shorten notation, for indices i, j set fi,j := ei,j + ej,i. For
t ∈ [0, 1] we define the following element of A ⊗ M6:
xt :=a ⊗ (e1,1 + (1 − t) · e3,3 + e5,5)
+b ⊗ (f1,2 + (1 − t) · f3,4 + f5,6)
+c ⊗ (e2,2 + (1 − t) · e4.4 + e6,6)
+1A ⊗ (t · f2,3 + t · f4,5 + δ(t) · f1,3)
where δ : [0, 1] → [0, 1] is a continuous function on [0, 1] that takes the value 0 at
the endpoints 0 and 1, and is strictly positive at each point t ∈ (0, 1), e.g., δ could
be given by δ(t) = 1/4 − (t − 1/2)2. We also define for t ∈ [0, 1] two elements of
M6:
y′
t :=e1,1 + (1 − t) · e3,3 + e5,5
z′
t :=e2,2 + (1 − t) · e4,4 + e6,6
It is easy to check that the assignment x : t 7→ xt defines an element x ∈ A⊗Z2.3.
t. In matrix
Similarly, we get two elements y′, z′ ∈ Z2.3 defined via t 7→ y′
form, these elements look as follows:
t and t 7→ z′
xt :=
y′
t :=
a
b
δ(t)
b
c
t
δ(t)
t
(1 − t)a (1 − t)b
(1 − t)c
(1 − t)b
1
(1 − t)
1
t
t
a b
b
c
z′
t :=
1
(1 − t)
1
Set y := 1 ⊗ y′, and let D := C ∗(x + 1, y) be the sub-C ∗-algebra of E := A ⊗ Z2,3
generated by the two self-adjoint elements x + 1 and y. Since x ≥ 0, we get that
both 1 and x lie in C ∗(x + 1). It follows that D = C ∗(1, x, y), and we will show that
D = E. Note that E has a natural continuous C([0, 1])-algebra structure (induced
THE GENERATOR PROBLEM FOR Z-STABLE C ∗-ALGEBRAS
9
by the one of Z2,3), with fibers E(0) ∼= A ⊗ M2, E(1) ∼= A ⊗ M3, and E(t) ∼= A ⊗ M6
for points t ∈ (0, 1) ⊂ [0, 1].
Let J := E((0, 1)) ✁ E be the natural ideal corresponding to the open set (0, 1) ⊂
[0, 1]. Note that J ∼= A⊗ C0((0, 1))⊗ M6, and J is naturally a continuous C0((0, 1))-
algebra. We will show in two steps that D exhausts the ideal J (i.e., D ∩ J = J)
and the quotient E/J (i.e., D/(D ∩ J) = E/J).
Step 1: We want to apply Lemma 3.2 to the C((0, 1))-algebra J with sub-C ∗-
algebra D ∩ J. To verify condition (ii), note that the C ∗-algebra generated by y′
contains C0((0, 1)) ⊗ e3,3. Therefore, D ∩ J contains 1A ⊗ C0((0, 1)) ⊗ e3,3, which
separates the points of (0, 1). Since 1A ⊗ e3,3 ∈ E(t) ∼= A ⊗ M6 is full, condition (ii)
of Lemma 3.2 holds and it remains to verify condition (i).
We need to show that D ∩ J exhausts all fibers of J. Fix some t ∈ (0, 1), and
set Dt := C ∗(1, xt, yt) ⊂ A ⊗ M6. To simplify notation, we write ¯ei,j for the matrix
units 1A ⊗ ei,j ∈ A ⊗ M6. We need to show that Dt is all of A ⊗ M6. This will follow
if Dt contains all ¯ei,j, and for this it is enough to show that the off-diagonal matrix
units ¯ei,i+1 are in Dt, for i = 1, . . . , 5.
The spectrum of yt is {0, 1 − t, 1}. Applying functional calculus to yt we obtain
that the following three elements lie in Dt:
u := ¯e1,1 + ¯e5,5
v := ¯e3,3
w := 1 − v − u = ¯e2,2 + ¯e4,4 + ¯e6,6
Then, we proceed as follows:
1. ¯e1,3 = δ(t)−1uxtv ∈ Dt and so ¯e1,1, ¯e5,5 ∈ Dt.
2. g := b ⊗ e1,2 = ¯e1,1xtw ∈ Dt. It follows b ⊗ e1,1 = (gg∗)1/2 ∈ Dt, cf. [OZ76].
Then b−1 ⊗ e1,1 ∈ C ∗(b ⊗ e1,1) ⊂ Dt and so ¯e1,2 = (b−1 ⊗ e1,1) · g ∈ Dt and
¯e2,2 ∈ Dt.
3. b ⊗ e3,4 = (1 − t)−1¯e3,3xt(w − ¯e2,2) ∈ Dt. Arguing as above, it follows that
¯e3,4 ∈ Dt, and then ¯e4,4, ¯e6,6 ∈ Dt.
4. ¯e2,3 = t−1¯e2,2xt¯e3,3 ∈ Dt.
5. ¯e4,5 = t−1¯e4,4xt¯e5,5 ∈ Dt.
6. b ⊗ e5,6 = ¯e5,5xt¯e6,6 ∈ Dt and so ¯e5,6 ∈ Dt.
This shows that D ∩ J exhausts the fibers of J. We may apply Lemma 3.2 and
deduce D ∩ J = J, which finishes step 1.
Let us denote the matrix units in M2 by e(0)
M3 by e(1)
1A ⊗ e(k)
Step 2: We want to show that D/J exhausts E/J = E({0, 1}) ∼= A ⊗ (M2 ⊕ M3).
i,j , i = 1, 2, and the matrix units in
i,j for the matrix units
i,j ∈ A ⊗ (M2 ⊕ M3). Let us denote the image of x and y in D/J by v and w:
2,3 + ¯e(1)
i,j , i = 1, 2, 3. To simplify notation, we write ¯e(k)
2,1) + c ⊗ (e(0)
v = a ⊗ (e(0)
2,2) + ¯e(1)
2,2 + e(1)
1,2 + e(0)
2,1 + e(1)
1,2 + e(1)
3,2
0
.
0
=(cid:18)a b
b
w = ¯e(0)
1,1 + ¯e(1)
1
1,1 + e(1)
c(cid:19) ⊕
1,1 =(cid:18)1 0
1,1) + b ⊗ (e(0)
a b
c
b
1
0 0(cid:19) ⊕
1
10
HANNES THIEL AND WILHELM WINTER
As in step 1, it is enough to show that D/J contains the off-diagonal matrix units
¯e(0)
1,2, ¯e(1)
1. g := wv(1 − w) = b ⊗ (e(0)
2,3. We argue as follows:
1,2 and ¯e(1)
1,2) ∈ D/J. As in step 1, it follows that b ⊗ (e(0)
1,2 + ¯e(1)
1,1 +
1,2 =
1,2 + e(1)
1,1) = (gg∗)1/2 ∈ D/J. Then b−1 ⊗ (e(0)
e(1)
1,1 + e(1)
(b−1 ⊗ (e(0)
3,3 = 1 − w − (¯e(0)
2,3 = v¯e(1)
1,1 + e(1)
1,1)) · g ∈ D/J. It follows that ¯e(0)
2,2 + ¯e(1)
3,3 ∈ D/J, and so ¯e(1)
2,2) ∈ D/J.
2,2 ∈ D/J.
1,1) ∈ D/J, and so ¯e(0)
2,2 + ¯e(1)
2,2 ∈ D/J.
2,2 ∈ D/J. Again, this implies ¯e(1)
1,2 ∈ D/J and so ¯e(1)
1,1 ∈ D/J.
1,2 = wv¯e(1)
2. ¯e(1)
3. ¯e(1)
4. b ⊗ e(1)
5. ¯e(0)
6. ¯e(0)
7. b ⊗ e(0)
This finishes step 2.
1,1 = w − ¯e(1)
2,2 = 1 − w − ¯e(1)
1,1v¯e(0)
1,2 = ¯e(0)
1,1 ∈ D/J.
3,3 ∈ D/J.
2,2 − ¯e(1)
2,2 ∈ D/J. Again, this implies ¯e(0)
1,2 ∈ D/J.
We have seen that A ⊗ Z2,3 is generated by x + 1 and y. Moreover, z′ is full,
(cid:3)
positive and orthogonal to y′.
Lemma 3.4. Let A be a separable, unital C ∗-algebra. Then there exist a positive element
x ∈ A ⊗ Z2∞,3∞ and two positive, full elements y′, z′ ∈ Z2∞,3∞ such that A ⊗ Z2∞,3∞
is generated by x and y := 1 ⊗ y′, and further y′ and z′ are orthogonal.
Proof. Let B := A ⊗ Z2∞,3∞. Note that Z2∞,3∞ ⊗ Z2,3 is naturally a C([0, 1] × [0, 1])-
algebra. Then, the quotient corresponding to the diagonal {(t, t) t ∈ [0, 1]} ⊂
[0, 1] × [0, 1] is isomorphic to Z2∞,3∞ , and we denote the resulting surjective mor-
phism by π : Z2∞,3∞ ⊗ Z2,3 → Z2∞,3∞. We proceed in two steps.
Step 1: We show that gen(B) ≤ k + 1 implies gen(B) ≤ k for k ≥ 2. So assume
B is generated by the self-adjoint, invertible elements a1, . . . , ak+1. The sub-C ∗-al-
gebra C := C ∗(ak−1, ak, ak+1) ⊂ B is unital and satisfies gen(C) ≤ 3. Consider the
C ∗-algebra B ⊗ Z2,3. By Lemma 3.3, the sub-C ∗-algebra C ⊗ Z2,3 is generated by
two self-adjoint elements, say b, c.
One readily checks that B ⊗ Z2,3 is generated by the k self-adjoint elements
a1 ⊗ 1, . . . , ak−2 ⊗ 1, b, c. Since B = A ⊗ Z2∞,3∞ is isomorphic to a quotient of
B ⊗ Z2,3 = A ⊗ Z2∞,3∞ ⊗ Z2,3, we obtain gen(B) ≤ gen(B ⊗ Z2,3) ≤ k.
Step 2: By Lemma 3.1, we have gen(B) ≤ 5. Applying Step 1 several times, we
obtain gen(B) ≤ 3.
It follows from Lemma 3.3 that there exists a positive element x ∈ B ⊗ Z2,3 and
two positive, full elements y′, z′ ∈ Z2,3 such that B ⊗ Z2,3 is generated by x and
1 ⊗ y′, and further y′ and z′ are orthogonal.
Consider the surjective morphism id ⊗π : A ⊗ Z2∞,3∞ ⊗ Z2,3 → A ⊗ Z2∞,3∞ .
One checks that the elements x := (id ⊗π)(x) ∈ A ⊗ Z2∞,3∞ , and y′ := π(y′), z′ :=
π(z′) ∈ Z2∞,3∞ have the desired properties.
(cid:3)
Theorem 3.5. Let A, B be two separable, unital C ∗-algebras. Assume the following:
(1) A contains a sequence a1, a2, . . . of full, positive elements that are pairwise or-
thogonal,
(2) B admits a unital embedding of the Jiang-Su algebra Z.
Then A ⊗max B is singly generated. Every other tensor product A ⊗λ B is a quotient of
A ⊗max B, and therefore is also singly generated.
THE GENERATOR PROBLEM FOR Z-STABLE C ∗-ALGEBRAS
11
Proof. There exists a unital embedding of Z2∞,3∞ in Z, so we may assume that
there is a unital embedding of Z2∞,3∞ in B. We may assume that the elements
a1, a2, . . . ∈ A are contractive.
Choose a sequence b1, b2, . . . ∈ B of contractive, positive elements that is dense
in the set of all contractive, positive elements of B.
Consider the sub-C ∗-algebra A⊗ Z2∞,3∞ ⊂ A⊗max B. By Lemma 3.4, there exist
a positive element x ∈ A ⊗ Z2∞,3∞ and two full, positive elements y′, z′ ∈ Z2∞,3∞
such that A ⊗ Z2∞,3∞ is generated by x and y := 1 ⊗ y′, and further y′ and z′ are
orthogonal.
Define the following two elements of A ⊗max B:
v := x,
w := 1 ⊗ y′ −Xk≥1
1/2k · ak ⊗ (z′bkz′).
Let D := C ∗(v, w) be the sub-C ∗-algebra of A ⊗max B generated by v and w. We
claim that D = A ⊗ B.
Step 1: We show A ⊗ Z2∞,3∞ ⊂ D. Note that the two elements 1 ⊗ y′ and
positive part of w, and therefore 1⊗y′ ∈ D. Therefore, C ∗(v, 1⊗y′) = A⊗Z2∞,3∞ ⊂
D.
Pk≥1 1/2k · ak ⊗ (z′bkz′) are positive and orthogonal. It follows that 1 ⊗ y′ is the
Step 2: We show 1 ⊗ B ⊂ D. We have g := Pk≥1 1/2k · ak ⊗ (z′bkz′) ∈ D. It
k is full, there exist finitely many elements ci, di ∈ A such that 1A =Pi cia2
Step 1, we have ci ⊗ 1, di ⊗ 1 ∈ D. Then 1 ⊗ (z′bkz′) =Pi(ci ⊗ 1)(a2
k ⊗ (z′bkz′) = 2k · (ak ⊗ 1)g ∈ D. Since
kdi. By
k ⊗ (z′bkz′))(di ⊗
follows from Step 1 that ak ⊗ 1 ∈ D, and so a2
a2
1) ∈ D, for each k.
Let b ∈ B be a contractive, positive element. Then b = limj bk(j) for certain in-
dices k(j). Then 1 ⊗ (z′bz′) = limj 1 ⊗ (z′bk(j)z′) ∈ D. It follows that the hereditary
sub-C ∗-algebra 1 ⊗ z′Bz′ is contained in D. Since z′ is full in Z2∞,3∞ , there exist
finitely many elements ci, di ∈ Z2∞,3∞ such that 1B = Pi ciz′di. We have seen
that 1 ⊗ z′bz′ ∈ D for any b ∈ B. Then 1 ⊗ bz′ = Pi(1 ⊗ ci)(1 ⊗ z′dibz′) ∈ D for
any b ∈ B. Similarly 1 ⊗ b =Pi(1 ⊗ bciz′)(1 ⊗ di) ∈ D for any b ∈ B, as desired.
It follows from Steps 1 and 2 that for each a ∈ A and b ∈ B the simple tensor
a ⊗ b is contained in D. The conclusion follows since A ⊗max B is the closure of the
linear span of simple tensors.
(cid:3)
Corollary 3.6. Let A, B be two separable, unital C ∗-algebras that both admit a unital
embedding of the Jiang-Su algebra Z. Then A ⊗max B is singly generated.
Proof. It is easy to verify that condition (i) of Theorem 3.5 is fulfilled if A admits a
unital embedding of Z.
(cid:3)
Theorem 3.7. Let A be a unital, separable C ∗-algebra. Then A ⊗ Z is singly generated.
Proof. Note that A ⊗ Z ∼= (A ⊗ Z) ⊗ Z. It is clear that both A ⊗ Z and Z admit
unital embeddings of Z. Then apply the above Corollary 3.6.
(cid:3)
Corollary 3.8. Let A be a separable C ∗-algebra. Then gen(A ⊗ Z) ≤ 3.
Proof. Let eA be the minimal unitization of A.
gen(eA ⊗ Z) ≤ 2. Since A ⊗ Z is an ideal in eA ⊗ Z, we get gen(A ⊗ Z) ≤ gen(eA ⊗
Z) + 1 ≤ 3 from Proposition 2.2, as desired.
It follows from Theorem 3.7 that
(cid:3)
12
HANNES THIEL AND WILHELM WINTER
Our results allow us to give new proofs for results about single generation of cer-
tain von Neumann algebras.
Proposition 3.9. Assume M, N are separably-acting von Neumann algebras that both
admit a unital embedding of the hyperfinite II1-factor. Then M ¯⊗N is singly generated.
Proof. Consider the GNS-representation π : Z → B(H) of the Jiang-Su algebra
with respect to its tracial state. The weak closure, π(Z)′′, is isomorphic to the
hyperfinite II1-factor R. Thus, there exists a weakly dense, unital copy of Z inside
R.
Choose weakly dense, separable, unital C ∗-algebras A0 ⊂ M , and similarly
B0 ⊂ N . Consider Z ⊂ R ⊂ M and set A := C ∗(A0, Z) ⊂ M . Similarly set
B := C ∗(B0, Z) ⊂ N .
Then A and B are separable, unital C ∗-algebras that both contain unital copies
of the Jiang-Su algebra. By Corollary 3.6, A ⊗max B is singly generated.
Consider the sub-C ∗-algebra C := C ∗(A ¯⊗1, 1 ¯⊗B) ⊂ M ¯⊗N . Then C is a quo-
tient of A ⊗max B, and therefore singly generated. Since C is weakly dense in
M ¯⊗N , we obtain that M ¯⊗N is singly generated, as desired.
(cid:3)
Remark 3.10. We note that a von Neumann algebra M admits a unital embedding
of R if and only if M has no (non-zero) finite-dimensional representations.
The analogous statement for C ∗-algebras would be that a C ∗-algebra A admits
a unital embedding of Z if and only if A has no (non-zero) finite-dimensional
representations. It was shown by Elliott and Rørdam, [ER06], that this is true for
C ∗-algebras of real rank zero. However, in [DHTW09] a simple, separable, unital,
non-elementary AH-algebra is constructed into which Z does not embed.
As a particular case of Proposition 3.9 we obtain the following result of Ge and
Popa.
Corollary 3.11 (Ge, Popa, [GP98, Theorem 6.2]). Assume M, N are separably-acting
II1-factors. Then M ¯⊗N is singly generated.
4. APPLICATIONS
In this section we show that the Jiang-Su algebra Z embeds unitally into the re-
duced group C ∗-algebras, C ∗
r (Γ), of groups Γ that contain a non-cyclic free sub-
group, see Proposition 4.2. We only consider discrete groups, and we let Fk denote
the free group with k generators (k ∈ {2, 3 . . . , ∞}).
We can apply Theorem 3.5 to show that certain tensor products of the form
r (F∞) ⊗
r (Γ) are singly generated, see Corollary 4.4.
In particular, C ∗
A ⊗max C ∗
C ∗
r (F∞) is singly generated, although it is not Z-stable, see Example 4.5.
r (F∞). A key observation is that C ∗
4.1. It was shown by Robert, [Rob10], that the Jiang-Su algebra Z embeds unitally
into C ∗
r (F∞) has strict comparison of positive
elements. This follows from the work of Dykema and Rørdam on reduced free
product C ∗-algebras, see [DR98] and [DR00].
Dykema and Rørdam study the comparison of projections, but this can be gen-
eralized to obtain results about the comparison of positive elements, as noted by
Robert, [Rob10]. In particular, [DR98, Lemma 5.3] and [DR00, Theorem 2.1] can be
generalized, and it follows that C ∗
r (F∞) has strict comparison of positive elements.
THE GENERATOR PROBLEM FOR Z-STABLE C ∗-ALGEBRAS
13
Proposition 4.2. If Γ is a discrete group that contains F∞ as a subgroup, then Z embeds
unitally into C ∗
r (Γ).
Proof. In general, for any subgroup Γ1 of a discrete group Γ, we have a unital
embedding C ∗
r (Γ) contains
a unital copy of C ∗
(cid:3)
r (Γ). Hence, if F∞ is a subgroup of Γ, then C ∗
r (F∞), which in turn contains a unital copy of Z.
r (Γ1) ⊂ C ∗
Remark 4.3. Every non-cyclic free group Fk (k ≥ 2) contains F∞ as a subgroup. In
general, by the Nielsen-Schreier theorem, every subgroup of a free group is again
free. Thus, if a, b are free elements, then the the elements akbk generate a subgroup
Γ = hakbk, k ≥ 1i that is free, and since none of the elements akbk is contained in
the subgroup generated by the other elements, we have Γ ∼= F∞.
Thus, when we ask which discrete groups contain F∞ as a subgroup, we are
equivalently asking which groups Γ contain a non-cyclic free subgroup.
It is a
necessary condition that Γ is non-amenable. The converse implication is known
as the von Neumann conjecture, but this was disproved in 1980 by Ol'shanskij.
A counterexample are the so-called Tarski monster groups, in which every non-
trivial proper subgroup is cyclic of some fixed prime order. Clearly, such a group
cannot contain F∞ as a subgroup, and it is Ol'shanskij's contribution to show that
Tarski monster groups exist and are non-amenable.
On the other hand, every group with the weak Powers property, as defined in
[BN88], has a non-cyclic free subgroup. A proof can be found in [dlH07], which
also lists classes of groups that have the (weak) Powers property. We just mention
that all free products Γ1 ∗ Γ2 with Γ1 ≥ 2, Γ2 ≥ 3 have the Powers property, and
therefore Proposition 4.2 applies.
We may derive the following from Theorem 3.5 and Proposition 4.2:
Corollary 4.4. Let A be a separable, unital C ∗-algebra that contains a countable sequence
of pairwise orthogonal, full elements (e.g., A is simple and nonelementary), and let Γ be a
group that contains a non-cyclic free subgroup. Then A ⊗max C ∗
r (Γ) is singly generated.
r (Γ1 × Γ2) ∼= C ∗
Example 4.5. Let Γ1, Γ2 be two groups that contain non-cyclic free subgroups.
Then C ∗
r (Γ2) is singly generated. For example, for
any k, l ∈ {2, 3, . . . , ∞}, the C ∗-algebra C ∗
r (Fl) is singly generated. In
particular, C ∗
r (F∞) is singly generated.
r (F∞) ⊗max C ∗
r (Γ1) ⊗max C ∗
r (Fk) ⊗max C ∗
It was pointed out to the authors by S. Wassermann that C ∗
r (Fl) is not
r (Fl) ∼= A ⊗ B ⊗ C, then
Z-stable, for any k, l ∈ {2, 3, . . . , ∞}. In fact, if C ∗
one of the three algebras A, B or C is isomorphic to C. This is a generalization of
the fact that C ∗
r (Fk) is tensorially prime, and it can be proved similarly.
r (Fk) ⊗ C ∗
r (Fk) ⊗ C ∗
We note that it is a difficult open problem whether C ∗
r (Fk) is singly generated
itself.
Question 4.6. Given a non-amenable (discrete) group Γ. Does C ∗
tal embedding of Z?
r (Γ) admit a uni-
For each group Γ, the trivial group-morphism Γ → {1} induces a surjective mor-
phism C ∗(Γ) → C. Thus, the Jiang-Su algebra can never unitally embed into a full
r (Γ) ∼= C ∗(Γ), and consequently there
group C ∗-algebra. If Γ is amenable, then C ∗
is no unital embedding of Z into the reduced group C ∗-algebra of an amenable
group.
14
HANNES THIEL AND WILHELM WINTER
On the other hand, if Γ contains a non-cyclic free subgroup, then Proposition 4.2
gives a positive answer to Question 4.6. As noted in Remark 4.3, not every non-
amenable group contains a non-cyclic free subgroup. However, it is known that
the reduced group C ∗-algebra of a non-amenable group has no finite-dimensional
representations, which is a necessary condition for the Jiang-Su algebra to embed.
ACKNOWLEDGMENTS
The first named author thanks Mikael Rørdam for valuable comments, especially
on the applications in Section 4.
REFERENCES
[AP77]
C.A. Akemann and G.K. Pedersen, Ideal perturbations of elements in C ∗-algebras, Math.
Scand. 41 (1977), 117 -- 139.
[Arv77] W. Arveson, Notes on extensions of C ∗-algebras, Duke Math. J. 44 (1977), 329 -- 355.
[BN88]
F. Boca and V. Nit¸ica, Combinatorial properties of groups and simple C ∗-algebras with a unique
trace, J. Oper. Theory 20 (1988), no. 1, 183 -- 196.
[Dad09] M. Dadarlat, Continuous fields of C ∗-algebras over finite dimensional spaces, Adv. Math. 222
(2009), no. 5, 1850 -- 1881.
[DHTW09] M. Dadarlat, I. Hirshberg, A.S. Toms, and W. Winter, The Jiang-Su algebra does not always
[dlH07]
[DR98]
[DR00]
[ER06]
[ET08]
[Ge03]
[GP98]
[JS99]
[JW11]
[Kad67]
[Kas88]
[Lon84]
[LS10]
embed, Math. Res. Lett. 16 (2009), no. 1, 23 -- 26.
P. de la Harpe, On simplicity of reduced C ∗-algebras of groups, Bull. Lond. Math. Soc. 39 (2007),
no. 1, 1 -- 26.
K.J. Dykema and M. Rørdam, Projections in free product C ∗-algebras, Geom. Funct. Anal. 8
(1998), no. 1, 1 -- 16.
, Projections in free product C ∗-algebras. II, Math. Z. 234 (2000), no. 1, 103 -- 113.
G.A. Elliott and M. Rørdam, Perturbation of Hausdorff moment sequences, and an application to
the theory of C ∗-algebras of real rank zero, Bratteli, Ola (ed.) et al., Operator algebras. The Abel
symposium 2004. Proceedings of the first Abel symposium, Oslo, Norway, September 3 -- 5,
2004. Berlin: Springer. Abel Symposia 1, 97-115, 2006.
G.A. Elliott and A.S. Toms, Regularity properties in the classification program for separable
amenable C ∗-algebras, Bull. Am. Math. Soc., New Ser. 45 (2008), no. 2, 229 -- 245.
L.M. Ge, On 'Problems on von Neumann algebras by R. Kadison, 1967', Acta Math. Sin., Engl.
Ser. 19 (2003), no. 3, 619 -- 624.
L. Ge and S. Popa, On some decomposition properties for factors of type II1, Duke Math. J. 94
(1998), no. 1, 79 -- 101.
X. Jiang and H. Su, On a simple unital projectionless C ∗-algebra, Am. J. Math. 121 (1999), no. 2,
359 -- 413.
M. Johanesova and W. Winter, The similarity problem for Z-stable C ∗-algebras, preprint,
arXiv:1104.2067, 2011.
R. Kadison, Problems on von Neumann algebras, unpublished manuscript, presented at Con-
ference on Operator Algebras and Their Applications, Louisiana State Univ., Baton Rouge,
La., 1967.
G.G. Kasparov, Equivariant KK-theory and the Novikov conjecture, Invent. Math. 91 (1988),
no. 1, 147 -- 201.
R. Longo, Solution of the factorial Stone-Weierstrass conjecture. An application of the theory of
standard split W ∗-inclusions, Invent. Math. 76 (1984), 145 -- 155.
W. Li and J. Shen, A note on approximately divisible C ∗-algebras, preprint, arXiv:0804.0465,
2010.
[Nag04] M. Nagisa, Single generation and rank of C ∗-algebras, Kosaki, Hideki (ed.), Operator alge-
bras and applications. Proceedings of the US-Japan seminar held at Kyushu University,
Fukuoka, Japan, June 7 -- 11, 1999. Tokyo: Mathematical Society of Japan. Advanced Stud-
ies in Pure Mathematics 38, 135-143, 2004.
C.L. Olsen and W.R. Zame, Some C ∗-algebras with a single generator, Trans. Am. Math. Soc.
215 (1976), 205 -- 217.
[OZ76]
THE GENERATOR PROBLEM FOR Z-STABLE C ∗-ALGEBRAS
15
[Pea62]
[Pop84]
[Rob10]
[Rør04]
[RW10]
[She09]
[Top68]
[TW08]
[vN31]
[Wil74]
C. Pearcy, W ∗-algebras with a single generator, Proc. Am. Math. Soc. 13 (1962), 831 -- 832.
S. Popa, Semiregular maximal Abelian ∗-subalgebras and the solution to the factor state Stone-
Weierstrass problem, Invent. Math. 76 (1984), 157 -- 161.
L. Robert, Classification of inductive limits of 1-dimensional NCCW complexes, preprint,
arXiv:1007.1964, 2010.
M. Rørdam, The stable and the real rank of Z-absorbing C ∗-algebras, Int. J. Math. 15 (2004),
no. 10, 1065 -- 1084.
M. Rørdam and W. Winter, The Jiang-Su algebra revisited, J. Reine Angew. Math. 642 (2010),
129 -- 155.
D. Sherman, On cardinal invariants and generators for von Neumann algebras, preprint,
arXiv:0908.4565, 2009.
D.M. Topping, UHF algebras are singly generated, Math. Scand. 22 (1968), 224 -- 226.
A.S. Toms and W. Winter, Z-stable ASH algebras, Can. J. Math. 60 (2008), no. 3, 703 -- 720.
J. von Neumann,
no. 2, 191 -- 226.
P. Willig, Generators and direct integral decompositions of W∗-algebras, Tohoku Math. J., II. Ser.
26 (1974), 35 -- 37.
Uber Funktionen von Funktionaloperatoren, Ann. of Math. (2) 32 (1931),
[Win07] W. Winter, Localizing the Elliott conjecture at strongly self-absorbing C ∗-algebras, preprint,
arXiv:0708.0283, 2007.
[Win10]
[Win11]
, Nuclear dimension and Z-stability of pure C ∗-algebras, preprint, arXiv:1006.2731, to
appear in Inventiones, 2010.
, Strongly self-absorbing C ∗-algebras are Z-stable, J. Noncommut. Geom. 5 (2011),
no. 2, 253 -- 264.
[Wog69] W.R. Wogen, On generators for von Neumann algebras, Bull. Am. Math. Soc. 75 (1969), 95 -- 99.
DEPARTMENT OF MATHEMATICAL SCIENCES, UNIVERSITY OF COPENHAGEN, UNIVERSITETSPARKEN
5, DK-2100, COPENHAGEN Ø, DENMARK
E-mail address: [email protected]
MATHEMATISCHES INSTITUT DER UNIVERSIT AT M UNSTER, EINSTEINSTR. 62, 48149 M UNSTER,
GERMANY
E-mail address: [email protected]
|
1301.7347 | 1 | 1301 | 2013-01-30T20:12:10 | C*-algebras associated with topological group quivers II: K-groups | [
"math.OA"
] | Topological quivers generalize the notion of directed graphs in which the sets of vertices and edges are locally compact (second countable) Hausdorff spaces. Associated to a topological quiver $Q$ is a $C^*$-correspondence, and in turn, a Cuntz-Pimsner algebra $C^*(Q).$ Given $\Gamma$ a locally compact group and $\alpha$ and $\beta$ endomorphisms on $\Gamma,$ one may construct a topological quiver $Q_{\alpha,\beta}(\Gamma)$ with vertex set $\Gamma,$ and edge set $\Omega_{\alpha,\beta}(\Gamma)= \{(x,y)\in\Gamma\times\Gamma\st \alpha(y)=\beta(x)\}.$ In \cite{Mc1}, the author examined the Cuntz-Pimsner algebra $\cO_{\alpha,\beta}(\Gamma):=C^*(Q_{\alpha,\beta}(\Gamma))$ and found generators (and their relations) of $\cO_{\alpha,\beta}(\Gamma).$ In this paper, the author uses this information to create a six term exact sequence in order to calculate the $K$-groups of $\cO_{\alpha,\beta}(\Gamma).$ | math.OA | math |
C ∗-ALGEBRAS ASSOCIATED WITH TOPOLOGICAL GROUP
QUIVERS II:
K-GROUPS
SHAWN J. McCANN
Abstract. Topological quivers generalize the notion of directed graphs in which
the sets of vertices and edges are locally compact (second countable) Hausdorff
spaces. Associated to a topological quiver Q is a C ∗-correspondence, and in turn,
a Cuntz-Pimsner algebra C ∗(Q). Given Γ a locally compact group and α and
β endomorphisms on Γ, one may construct a topological quiver Qα,β(Γ) with
vertex set Γ, and edge set Ωα,β(Γ) = {(x, y) ∈ Γ × Γ(cid:12)(cid:12) α(y) = β(x)}. In [52], the
author examined the Cuntz-Pimsner algebra Oα,β(Γ) := C ∗(Qα,β(Γ)) and found
generators (and their relations) of Oα,β(Γ). In this paper, the author uses this
information to create a six term exact sequence in order to calculate the K-groups
of Oα,β(Γ).
1. Introduction and Notation
1.1. Background. Given a quintuple Q = (X, E, r, s, λ), where X and E are locally
compact (second countable) Hausdorff spaces, r and s are continuous maps from
X to E with r open, and λ = {λx}x∈E is a system of Radon measures, one can
create a corresponding Cuntz-Pimsner C ∗-algebra C ∗(Q). In [24], Exel, an Huef
and Raeburn define C ∗-algebras associated with a system (B, α, L) where α is an
endomorphism of a unital C ∗-algebra B and L is a positive linear map L : B → B
such that L(α(a)b) = aL(b) for all a, b ∈ B called a transfer operator. In fact, the
C ∗-algebra they generate is a Cuntz-Pimsner algebra and under certain restrictions,
a C ∗-algebra associated with a topological quiver; in particular, when B = C(Td)
the continuous function on the d-torus, F ∈ Md(Z) and α is the endomorphism
α(f )(e2πit) = f (e2πiF t)
for f ∈ C(Td) and t ∈ Rd. Exel, an Huef and Raeburn then determine a six term
exact sequence in which to use to calculate the K-groups of these C ∗-algebras. In
[52], the author considers a certain class of topological quivers (which extend the
notions of Exel, an Huef and Raeburn) Q = (Γ, Ωα,β(Γ), r, s, λ) where Γ is a locally
compact group, α and β are endormorphism of Γ,
Ωα,β(Γ) = {(x, y) ∈ Γ × Γ(cid:12)(cid:12) α(y) = β(x)}
and λ is an appropriate family of Radon measures. The resulting Cuntz-Pimsner C ∗-
algebra, denoted Oα,β(Γ), was then examined and certain generators and relations
where found. We now proceed to generalize the six term exact sequence considered
in [24] to C ∗-algebras of the form Oα,β(Γ) where Γ is a compact group.
1
2
SHAWN J. McCANN
1.2. Notation. The sets of natural numbers, integers, rationals numbers, real num-
bers and complex numbers will be denoted by N, Z, Q, R, and C, respectively.
Convention: N does not contain zero. Z+
0 will denote the set N ∪ {0}, R+ denotes
0 = R+ ∪ {0}. Finally, Zp denotes the abelian group
the set {r ∈ R(cid:12)(cid:12) r > 0} and R+
Z/pZ = {0, 1, ..., p − 1 mod p} and T denotes the torus {z ∈ C(cid:12)(cid:12) z = 1}. Whenever
For a topological space Y , the closure of Y is denoted Y . Given a locally compact
convenient, view Zp ⊂ T by Zp
∼= {z ∈ T(cid:12)(cid:12) zp = 1}.
Hausdorff space X, let
(1) C(X) be the continuous complex functions on X;
(2) Cb(X) be the continuous and bounded complex functions on X;
(3) C0(X) be the continuous complex functions on X vanishing at infinity;
(4) Cc(X) be the continuous complex functions on X with compact support.
The supremum norm is denoted · ∞ and defined by
f ∞ = sup
x∈X
{f (x)}
for each continuous map f : X → C. For a continuous function f ∈ Cc(X), denote
supp f = osuppf .
the open support of f by osupp f = {x ∈ X(cid:12)(cid:12) f (x) 6= 0} and the support of f by
For C ∗-algebras A and B, A is isomorphic to B will be written A ∼= B; for
example, we use C(Td) ⊗ MN (C) ∼= MN (C(Td)). Moreover, A⊕n denotes the n-fold
direct sum A ⊕ · · · ⊕ A. Given a group Γ and a ring R, a normal subgroup, N, of Γ is
denoted N ✁ Γ and an ideal, I, of R is denoted I ✁ R. Note if R is a C ∗-algebra then
the term ideal denotes a closed two-sided ideal. Furthermore, End(Γ) (End(R)) and
Aut(Γ) (Aut(R)) denotes the set of endomorphisms of Γ (R) and automorphisms of
Γ (R), respectively. For a map γ : Γ → Aut(A), the fixed point set is denoted Aγ
and defined by
Let α ∈ C(X) then α# ∈ End(C(X)) denotes the endomorphism of C(X) defined
Aγ = {a ∈ A(cid:12)(cid:12) γ(g)(a) = a for each g ∈ Γ}.
by
Let S be a set and define the Kronecker delta function δ : S × S → {0, 1} by
α#(f ) = f ◦ α
for each f ∈ C(X).
δr
s := δ(s, r) =(0 if s 6= r
1 if s = r
.
The set of n by n matrices with coefficients in a set R will be denoted Mn(R) and
for any F ∈ Mn(R), the transpose of F is denoted F T . Given a function σ : R → S,
we may create an augmented function σn : Mn(R) → Mn(S) via
σn((ri,j)n
i,j=1) = (σ(ri,j))n
i,j=1
for each (ri,j)n
i,j=1 ∈ Mn(R). Given vectors v = (v1, ..., vn) of length n and w =
(w1, ..., wm) of length m, denote (v, w) to be the vector (v, w) = (v1, ..., vn, w1, ..., wm)
of length n + m.
C ∗-ALGEBRAS ASSOCIATED WITH TOPOLOGICAL GROUP QUIVERS II
3
2. Preliminairies
2.1. Hilbert C ∗-modules. We begin by defining Hilbert C ∗-modules. Further de-
tails and references can be found in [48, 63].
Definition 2.1. [48] If A is a C ∗-algebra, then a (right) Hilbert A-module is a
Banach space EA together with a right action of A on EA and an A-valued inner
product h·, ·iA satisfying
(1) hξ, ηaiA = hξ, ηiAa
(2) hξ, ηiA = hη, ξi∗
A
(3) hξ, ξi ≥ 0 and ξ = hξ, ξi1/2
A A
for all ξ, η ∈ EA and a ∈ A (if the context is clear, we denote EA simply by E).
For Hilbert A-modules E and F , call a function T : E → F adjointable if there
is a function T ∗ : F → E such that hT (ξ), ηiA = hξ, T ∗(η)iA for all ξ ∈ E and
η ∈ F . Let L(E, F ) denote the set of adjointable (A-linear) operators from E to F .
If E = F , then L(E) := L(E, E) is a C ∗-algebra (see [48].) Let K(E, F ) denote the
closed two-sided ideal of compact operators given by
where
θE,F
ξ,η (ζ) = ξhη, ζiA
K(E, F ) := span{θE,F
ξ,η (cid:12)(cid:12) ξ ∈ E, η ∈ F }
for each ζ ∈ E.
Similarly, K(E) := K(E, E) and θE
ξ,η (or θξ,η if understood) denotes θE,E
ξ,η . For Hilbert
A-module E, the linear span of {hξ, ηi(cid:12)(cid:12) ξ, η ∈ E}, denoted hE, Ei, once closed is a
two-sided ideal of A. Note that EhE, Ei is dense in E ([48]). The Hilbert module E
is called full if hE, Ei is dense in A. The Hilbert module AA refers to the Hilbert
module A over itself, where ha, bi = a∗b for all a, b ∈ A.
An algebraic generating set for E is a subset {ui}i∈I ⊂ E for some indexing set I
such that E equals the linear span of {ui · a(cid:12)(cid:12) i ∈ I, a ∈ A}.
Definition 2.2. [37] A subset {ui}i∈I ⊂ E is called a basis provided the following
reconstruction formula holds for all ξ ∈ E :
ui · hui, ξi
(in E, · .)
ξ =Xi∈I
If hui, uji = δj
i as well, call {ui}i∈I an orthonormal basis of E.
Remark 2.3. The preceding definition is in accordance with the finite version in
[37], but many other versions exist such as in [24] where {ui}n
i=1 is called a finite
Parseval frame, or in [68] where this is taken as the definition for finitely generated.
There has been substantial work done on similar frames (see [32]).
The following notions of C ∗-correspondence and morphism may be found in [56,
9, 10, 11, 24, 25, 26, 39]
Definition 2.4. [10, 11] If A and B are C ∗-algebras, then an A−B C ∗-correspondence
E is a right Hilbert B-module EB together with a left action of A on E given by a
∗-homomorphism φA : A → L(E), a · ξ = φA(a)ξ for a ∈ A and ξ ∈ E. We may
4
SHAWN J. McCANN
occasionally write, AEB to denote an A − B C ∗-correspondence and φ instead of
φA. Furthermore, if A1EB1 and A2FB2 are C ∗-correspondences, then a morphism
(π1, T, π2) : E → F consists of ∗-homomorphisms πi : Ai → Bi and a linear map
T : E → F satisfying
(i) π2(hξ, ηiA2) = hT (ξ), T (η)iB2
(ii) T (φA1(a1)ξ) = φB1(π1(a1))T (ξ)
(iii) T (ξ)π2(a2) = T (ξa2)
for all ξ, η ∈ E and ai ∈ Ai.
Notation 2.5. When A = B, we refer to AEA as a C ∗-correspondence over A. For
E a C ∗-correspondence over A and F a C ∗-correspondence over B, a morphism
(π, T, π) : E → F will be denoted by (T, π).
Definition 2.6. [56] If F is the Hilbert module CCC where C is a C ∗-algebra with
the inner product hx, yiB = x∗y then call a morphism (T, π) : AEB → C of Hilbert
modules a representation of AEB into C.
Remark 2.7. Note that a representation of AEB need only satisfying (i) and (ii)
of definition 2.4 as it was unnecessary to require (iii). For a proof, see [52, Remark
2.7].
A morphism of Hilbert modules (T, π) : E → F yields a ∗-homomorphism ΨT :
K(E) → K(F ) by
ΨT (θE
ξ,η) = θF
T (ξ),T (η)
for ξ, η ∈ E and if (S, σ) : D → E, and (T, π) : E → F are morphisms of Hilbert
modules then ΨT ◦ ΨS = ΨT ◦S. In the case where F = B a C ∗-algebra, we may
first identify K(B) as B, and a representation (T, π) of E in a C ∗-algebra B yields
a ∗-homomorphism ΨT : K(E) → B given by
ΨT (θξ,η) = T (ξ)T (η)∗.
Definition 2.8. [56] For a C ∗-correspondence E over A, denote the ideal φ−1(K(E))
of A by J(E), and let JE = J(E) ∩ (ker φ)⊥ where (ker φ)⊥ is the ideal {a ∈ A(cid:12)(cid:12) ab =
If AEA and BFB are C ∗-correspondences over A and B re-
0 for all b ∈ ker φ} .
spectively and K ✁ J(E), a morphism (T, π) : E → F is called coisometric on K
if
for all a ∈ K, or just coisometric, if K = J(E).
ΨT (φA(a)) = φB(π(a))
Notation 2.9. We denote C ∗(T, π) to be the C ∗-algebra generated by T (E) and
π(A) where (T, π) : E → B is a representation of AEA in a C ∗-algebra B. Further-
more, if ρ : B → C is a ∗-homomorphism of C ∗-algebras, then ρ ◦ (T, π) denotes the
representation (ρ ◦ T, ρ ◦ π) of E.
Definition 2.10. [56] A morphism (TE , πE) coisometric on an ideal K is said to be
universal if whenever (T, π) : E → B is a representation coisometric on K, there
exists a ∗-homomorphism ρ : C ∗(TE , πE) → B with (T, π) = ρ ◦ (TE, πE). The
universal C ∗-algebra C ∗(TE, πE ) is called the relative Cuntz-Pimsner algebra of E
C ∗-ALGEBRAS ASSOCIATED WITH TOPOLOGICAL GROUP QUIVERS II
5
determined by the ideal K and denoted by O(K, E). If K = 0, then O(K, E) is
denoted by T (E) and called the universal Toeplitz C ∗-algebra for E. We denote
O(JE, E) by OE .
2.2. Topological Quivers.
Definition 2.11. [56] A topological quiver (or topological directed graph) Q =
(X, E, Y, r, s, λ) is a diagram
s
..
......................................................................................
.........
...
..r
..........
.....................................................................................
...
Y
E
X
where X, E, and Y are second countable locally compact Hausdorff spaces, r and
s are continuous maps with r open, along with a family λ = {λyy ∈ Y } of Radon
measures on E satisfying
(1) supp λy = r−1(y) for all y ∈ Y , and
(2) y 7→ λy(f ) =RE f (α)dλy(α) ∈ Cc(Y ) for f ∈ Cc(E).
Remark 2.12. If X = Y then write Q = (X, E, r, s, λ) in lieu of (X, E, X, r, s, λ).
Remark 2.13. The author provides a broad history and a series of examples of
topological quivers in [51, 52].
Given a topological quiver Q = (X, E, Y, r, s, λ), one may associate a correspon-
dence EQ of the C ∗-algebra C0(X) to the C ∗-algebra C0(Y ). Define left and right
actions
(a · ξ · b)(e) = a(s(e))ξ(e)b(r(e))
by C0(X) and C0(Y ) respectively on Cc(E). Furthermore, define the Cc(Y )-valued
inner product
hξ, ηi(y) =Zr−1(y)
ξ(α)η(α)dλy(α)
for ξ, η ∈ Cc(E), y ∈ Y, and let EQ be the completion of Cc(E) with respect to the
norm
ξ = hξ, ξi1/2∞ = λy(ξ2)1/2
∞ .
Definition 2.14. Given topological quiver Q over a space X, define the C ∗-algebra,
C ∗(Q) associated with Q to be the Cuntz-Pimnser C ∗-algebra OEQ of the correspon-
dence EQ over A = C0(X).
2.3. Topological Group Quivers.
Definition 2.15. Let Γ be a (second countable) locally compact group and let
α, β ∈ End(Γ) be continuous. Define the closed subgroup, Ωα,β(Γ), of Γ × Γ,
and let Qα,β(Γ) = (Γ, Ωα,β(Γ), r, s, λ) where r and s are the group homomorphisms
defined by
Ωα,β(Γ) = {(x, y) ∈ Γ × Γ(cid:12)(cid:12) α(y) = β(x)}
for each (x, y) ∈ Ωα,β(Γ) and λx for x ∈ Γ is the measure on
r(x, y) = x
and
s(x, y) = y
r−1(x) = {x} × α−1(β(x))
6
SHAWN J. McCANN
defined by
λx(B) = µ(y−1s(B ∩ r−1(x)) ∩ ker α)
(for any y ∈ α−1(β(x)))
for each measurable B ⊆ Ωα,β(Γ) where µ is a left Haar measure (normalized if
possible) on r−1(1Γ) = {1} × ker α (a closed normal subgroup of Γ × Γ; hence, a
locally compact group). Note if r−1(x) = ∅ then α−1(β(x)) = ∅ and so λx = 0. This
measure is well-defined,
supp λx = {x} × y ker α = {x} × α−1(β(x)) = r−1(x)
and y 7→ λy(f ) is a continuous compactly supported function (cf.
3.1].
[52, Definition
Call Qα,β(Γ) a topological group relation. Define Eα,β(Γ) to be the C0(Γ)-correspondence
EQα,β(Γ) and form the Cuntz-Pimsner algebra
Oα,β(Γ) := C ∗(Qα,β(Γ)) = O(JEα,β (Γ), Eα,β(Γ))
and the Toeplitz-Pimsner algebra
Tα,β(Γ) := T (Qα,β(Γ)).
Remark 2.16. It will be implicitly assumed that Γ is second countable. Further-
more, since Γ is locally compact Hausdorff, r−1(x) is closed and locally compact.
Moreover, whenever r is a local homeomorphism, r−1(x) is discrete and hence, λx is
counting measure (normalized when ker α < ∞.)
Example 2.17 ([52]). For the compact abelian group Td, note End(Td) ∼= Md(Z)
([67]); that is, an element σ ∈ End(Td) is of the form σF for some F ∈ Md(Z) where
σF (e2πit) = e2πiF t
for each t ∈ Zd. To simplify notation, use F and G in place of σF and σG whenever
convenient. For instance,
and the C ∗-correspondence
QF,G(Td) := QσF ,σG(Td)
EF,G(Td) := EσF ,σG(Td)
where F, G ∈ Md(Z). We will consider the cases when these maps are surjective;
that is, det F and det G are non-zero.
Let F, G ∈ Md(Z) where det F, det G 6= 0. Then ker σF = det F and so, the
C(Td)-valued inner product becomes
hξ, ηi(x) =
1
det F XσF (y)=σG(x)
ξ(x, y)η(x, y)
for ξ, η ∈ EF,G(Td) and x ∈ Td. This is a finite sum since the number of solutions,
y, to σF (y) = σG(x) given any x ∈ Td is det F < ∞.
C ∗-ALGEBRAS ASSOCIATED WITH TOPOLOGICAL GROUP QUIVERS II
7
Remark 2.18. The left action, φ, is defined by
φ(a)ξ(x, y) = a(y)ξ(x, y)
for a ∈ C(Td), ξ ∈ C(ΩF,G(Td)) and (x, y) ∈ ΩF,G(Td). Note: φ is injective. To see
this claim, let a ∈ C(Td) and assume φ(a)ξ = 0 for each ξ ∈ C(ΩF,G(Td)). Then
a(y)ξ(x, y) = 0 for each (x, y) ∈ ΩF,G(Td) and ξ ∈ C(ΩF,G(Td)). Since s(ΩF,G(Td)) =
{y ∈ Td(cid:12)(cid:12) (x, y) ∈ ΩF,G(Td)} = Td by the surjectivity of σF , a = 0.
Remark 2.19. It was shown in [52] that one may assume the matrix F is positive
diagonal.
Let F = Diag(a1, ..., ad) ∈ Md(Z), G = (bjk)d
j = 1, ..., d, det G 6= 0 and let Gj denote the j-th row of G, (bjk)d
j,k=1 ∈ Md(Z) where aj > 0 for each
k=1. Further, let
N = det F =Qd
j=1 aj > 0 and let
I(F ) = {ν = (νj)d
The C(Td)-valued inner product becomes
j=1 ∈ Zd(cid:12)(cid:12) 0 ≤ νj ≤ aj − 1}.
N XσF (y)=σG(x)
ξ(x, y)η(x, y)
1
hξ, ηi(x) =
for all ξ, η ∈ C(ΩF,G(Td)) and x ∈ Td.
Given ν ∈ I(F ), define uν ∈ C(ΩF,G(Td)) by
uν(x, y) = yν =
yνj
d
Yj=1
for (x, y) ∈ ΩF,G(Td). It was shown in [52] that {uν}ν∈I(F ) is a basis for EF,G(Td)
and also the following:
Theorem 2.20. [52, Theorem 3.23] Let F = Diag(a1, ..., ad), G ∈ Md(Z) where
det F, det G 6= 0 and let Gj be the j-th row vector of G. Further, let I(F ) denote
the set {ν = (νj)d
algebra generated by isometries {Sν}ν∈I(F ) and (full spectrum) commuting unitaries
{Uj}d
j=1 ∈ Zd(cid:12)(cid:12) 0 ≤ νj ≤ aj − 1}. Then OF,G(Td) is the universal C ∗-
j=1 that satisfy the relations
(1) S∗
(2) U νS = Sν for all ν ∈ I(F ),
(3) U aj
ν Sν ′ = huν, uν ′i = δν ′
ν ,
j S = SU Gj , for all j = 1, ..., d and
(4) 1 =Pν∈I(F ) SνS∗
where U ν denotesQd
j=1 U νj
ated by isometries {Sν}ν∈I(F ) and commuting unitaries {Uj}d
(1)-(3)
j . Furthermore, Tα,β(Γ) is the universal C ∗-algebra gener-
j=1 that satisfy relations
ν =Pν∈I(F ) U νSS∗U −ν
8
SHAWN J. McCANN
3. Six Term Exact Sequence for Oα,β(Γ)
In this section, we follow and extend the approach of [24] to create a six term
exact sequence. Let Γ be a compact group with α, β ∈ End(Γ). Suppose the left
action for the correspondence, φ, is injective where φ is defined by
φ(a)ξ(x, y) = a(y)ξ(x, y)
for a ∈ C(Γ), ξ ∈ C(Ωα,β(Γ)) and (x, y) ∈ Ωα,β(Γ). Furthermore, we shall assume
the existence of an orthonormal basis (see Defintion 2.2) {ui}N −1
i=0
for Eα,β(Γ).
In order to construct our exact sequence for K∗(Oα,β(Γ)), note the short exact
sequence
0
..........
...........................................................................
...
..
ker q
..ι
..........
........................................................................
...
Tα,β(Γ)
..q
..........
.............................................................................
...
Oα,β(Γ)
..........
................................................................................
...
..
0,
where q : Tα,β(Γ) → Oα,β(Γ) is the canonical quotient map and ι : ker q → Tα,β(Γ) is
the inclusion homomorphism, induces the six-term exact sequence of K-groups (see
[65])
(3.1)
K0(ker q)
δ1
..
.
.
.
.
.
.
.
.
.
.
..
.
..
.
.
.
.
.
.
.
...
...
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
ι∗
.........
...............................................................................................................
...
..
K1(Oα,β(Γ))
.............................................................................................
..........
..
.. q∗
K0(Tα,β(Γ))
K1(Tα,β(Γ))
q∗
...
.........
.............................................................................................
..
K0(Oα,β(Γ))
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
...
...
.
.
.
.
.
..
..
.
.
.
.
.
.
.
.
.
.
.
.
.
.
..
.δ0
K1(ker q)
...............................................................................................................
..........
..
..
ι∗
Let (T, π) denote the universal Toeplitz representation on Eα,β(Γ); that is, π = q◦π
is the morphism C(Γ) → Oα,β(Γ). As shown in [60, Theorem 4.4], the homomor-
phism π : C(Γ) → Tα,β(Γ) induces an isomorphism of Ki(C(Γ)) onto Ki(Tα,β(Γ)).
Thus we may replace Ki(Tα,β(Γ)) with Ki(C(Γ)) provided we can identify the re-
sulting maps. We intend to show that (3.1) induces the six-term exact sequence
(3.2)
K0(C(Γ))
..
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
...
...
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
1 − Ω∗
...
.........
..............................................................................................................................
..
K1(Oα,β(Γ))
............................................................................................................
..........
..
.. π∗
K0(C(Γ))
K1(C(Γ))
π∗
...
.........
............................................................................................................
..
K0(Oα,β(Γ))
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
..
..
.
.
.
.
..
..
.
.
.
.
.
.
.
.
.
.
.
.
.
.
..
..
..
K1(C(Γ))
..
..............................................................................................................................
..........
.. 1 − Ω∗
for π = q ◦ π : C(Γ) → Oα,β(Γ) and an appropriately chosen homomorphism Ω :
C(Γ) → MN (C(Γ)).
Lemma 3.1. Define Ω : C(Γ) → MN (C(Γ)) by Ω(a) = (hui, a · uji)N −1
i,j=0. Then
Ω is a unital homomorphism and Ω(α#(a)) is the diagonal matrix β#(a)1N for all
a ∈ C(Γ).
C ∗-ALGEBRAS ASSOCIATED WITH TOPOLOGICAL GROUP QUIVERS II
9
Proof. Let a, b ∈ C(Γ). Then the (i, j)-entry of Ω(a)Ω(b) is
N −1
(Ω(a)Ω(b))i,j =
hui, a · ukihuk, b · uji
Xk=0
= hui, a · (Xk
= hui, a · (b · uj)i
= Ω(ab)i,j.
uk · huk, b · uji)i
Furthermore, for a∗ denoting the map a∗(x) = a(x) for x ∈ Γ,
Ω(a∗) = (hui, a∗ · uji)i,j = (ha · ui, uji)i,j = (huj, a · uii∗)i,j = Ω(a)∗
and
Finally, let x ∈ Γ. Then
Ω(1) = (hui, uji)i,j = (δj
i )i,j = 1N .
Ω(α#(a))i,j(x) = hui, α#(a) · uji(x)
ui(e)a(α(s(e)))uj(e) dλx(e)
ui(e)a(β(x))uj(e) dλx(e)
ui(e)uj(e) dλx(e)
=Zr−1(x)
=Zr−1(x)
= a(β(x))Zr−1(x)
= a(β(x))hui, uji(x)
= δj
i β#(a)(x);
hence, Ω(α#(a)) = β#(a)1N .
(cid:3)
In order to describe ker q, use the notation E ⊗k := Eα,β(Γ)⊗k for the k-fold internal
tensor product of C ∗-correspondences ([48]) Eα,β(Γ) ⊗ · · · ⊗ Eα,β(Γ), which is itself a
C ∗-correspondence over A = C(Γ). For the universal covariant representation (T, π) :
Eα,β(Γ) → Tα,β(Γ) (that is, q(Tj) is the isometry Sj with T (uj) = Tj,) there is, in
i=1 T (ξi)
for all elementary tensors ξ = ξ1 ⊗ · · · ⊗ ξk where ξi ∈ Eα,β(Γ) (see [27, Proposition
1.8] where the term "Hilbert bimodule" is used instead of C ∗-correspondence.) Note
Eα,β(Γ)⊗0 := C(Γ) and T ⊗0 := π. By [27, Lemma 2.4],
fact, a Toeplitz representation (T ⊗k, π) of Tα,β(Γ) such that T ⊗k(ξ) = Qk
Tα,β(Γ) = span{T ⊗k(ξ)T ⊗k′
j . The proceeding lemmas and propositions are essentially
those found in [24, Lemma 3.2, Lemma 3.3 & Proposition 3.4] with some changes.
j=0 TjT ∗
Next, let p = PN −1
(η)∗(cid:12)(cid:12) k, k′ ≥ 0, ξ ∈ E ⊗k, η ∈ E ⊗k′
}.
Lemma 3.2. With the preceding notation:
(1) p is a projection which commutes with π(a) for all a ∈ C(Γ)
10
SHAWN J. McCANN
(2) 1 − p is a full projection in ker q
(3) (1 − p)T ⊗k(ξ) = 0 for all ξ ∈ E ⊗k and k ≥ 1
(4) ker q = span{T ⊗k(ξ)(1 − p)T ⊗k′(η)∗(cid:12)(cid:12) k, k′ ≥ 0, ξ ∈ E ⊗k, η ∈ E ⊗k′}
i Tj = π(hui, uji) = δj
Proof. (1) Recall that T ∗
more,
i . Thus, p2 = p and p∗ = p. Further-
pπ(a)p =
=
=
and so,
N −1
N −1
Xj,k=1
Xj,k=0
Xk=0
N −1
TjT ∗
j π(a)TkT ∗
k =
N −1
Xj,k=0
T (ujhuj, a · uki)T ∗
k =
Tj π(huj, a · uki)T ∗
k
N −1
Xk=0
T (
N −1
Xj=0
ujhuj, a · uki)T ∗
k
T (a · uk)T ∗
k = π(a)p
pπ(a) = (π(a)∗p)∗ = (pπ(a)∗p)∗ = pπ(a)p = π(a)p.
(2) Recall φ(a) =PN −1
j=0 θa·uj ,uj , so
ΨT (φ(a)) =
T (a · uj)T (uj)∗ = π(a)p
N −1
Xj=0
and
q(1 − p) = q(π(1) − π(1)p) = q(π(1) − ΨT (φ(1))) = 0.
Hence, 1 − p = π(1) − π(1)p ∈ ker q and since ker q is the ideal in Tα,β(Γ) generated
by {π(a) − ΨT (φ(a))(cid:12)(cid:12) a ∈ C(Γ)} and 1 − p ∈ ker q, ker q is the ideal generated by
{π(a)(1 − p)(cid:12)(cid:12) a ∈ C(Γ)}. Hence 1 − p is full.
(3) Let ξ ∈ Eα,β(Γ) then
N −1
N −1
pT (ξ) =
Xj=0
T (uj)T (uj)∗T (ξ) =
Xj=0
T (ujhuj, ξi) = T (ξ).
Thus, (1 − p)T (ξ) = 0. Now for k > 1, let ξ = ξ1 ⊗ ... ⊗ ξk. Then
(1 − p)T ⊗k(ξ) = (1 − p)
T (ξj) = 0
k
Yj=0
and hence, by linearity and continuity, (3) has been proven.
(4) Since ker q = Tα,β(Γ)(1−p)Tα,β(Γ), the description of Tα,β(Γ) preceding Lemma
3.2 paired with (3) gives the desired result.
(cid:3)
Lemma 3.3. There exists a homomorphism ρ : C(Γ) → ker q ⊂ Tα,β(Γ) such that
ρ(a) = π(a)(1 − p) and ρ is an isomorphism of C(Γ) onto the full corner C ∗-algebra
C ∗-ALGEBRAS ASSOCIATED WITH TOPOLOGICAL GROUP QUIVERS II
11
(1 − p)ker q(1 − p).
Proof. By the previous lemma,
(1 − p)π(a)(1 − p) = π(a)(1 − p) ∈ ker q.
Thus, ρ(a) = π(a)(1 − p) defines a homomorphism ρ : C(Γ) → (1 − p) ker q(1 − p) ⊂
ker q. Using the previous lemma,
(1 − p) ker q(1 − p) = span{(1 − p)T ⊗k(ξ)(1 − p)T ⊗k′
(η)∗(1 − p)(cid:12)(cid:12) k, k′ ≥ 0, ξ ∈ E ⊗k, η ∈ E ⊗k′
}
= span{(1 − p)π(a)(1 − p)π(b)∗(1 − p)(cid:12)(cid:12) a, b ∈ C(Γ)}
= span{π(a)(1 − p)(cid:12)(cid:12) a ∈ C(Γ)} = ran ρ.
Hence, ρ is surjective.
In order to show the injectivity of ρ, choose a faithful representation π0 : C(Γ) →
B(H) and consider the Fock representation (TF , πF ) of Eα,β(Γ) induced from π0 as
described in [27, Example 1.4]. The underlying space of this Fock representation
is F (Eα,β(Γ)) ⊗A H := ⊕k≥0(E ⊗k ⊗A H) where A = C(Γ) acts diagonally on the
left and Eα,β(Γ) acts by creation operators. Then TF (ξ)∗ is an annihilation operator
vanishing on the subspace A ⊗A H of F (Eα,β(Γ)) ⊗A H. Now, for a ∈ A,
0 = (TF × πF )(ρ(a)) = (TF × πF )(π(a)(1 − p) = πF (a)(1 −
TF (uj)TF (uj)∗).
N −1
Xj=0
Since TF (uj)∗ vanishes on A ⊗A H, we have that ρ(a) = 0 implies
πF (a)(1 −
TF (uj)TF (uj)∗)(1 ⊗A h) = 0
N −1
Xj=0
for all h ∈ H and so, πF (a)(1 ⊗A h) = 0 for all h ∈ H. Thus, a ⊗A h = 0 for all
h ∈ H and hence, π0(a)h = 0 for all h ∈ H which implies a = 0 since π0 is faithful.
Hence, ρ is injective.
Lemma 3.4. [24, Lemma 3.5] Suppose that A is a C ∗-algebra, r ≥ 1 and N ≥ 2
are integers, and
(cid:3)
is a subset of A. For m, n satisfying 0 ≤ m, n < rN − 1, define
{bj,s;k,t(cid:12)(cid:12) 0 ≤ j, k < N and 0 ≤ s, t < r}
cm,n = bj,s;k,t where m = sN + j and n = lN + k, and
dm,n = bj,s;k,t where m = jr + s and n = kr + t.
Then there is a scalar unitary permutation matrix U such that the matrices C :=
(cm,n)m,n and Dm,n := (dm,n)m,n are related by C = UDU ∗.
The following is standard (and also appears in [24]):
Lemma 3.5. Suppose that S is an isometry in a unital C ∗-algebra A. Then
is a unitary element of M2(A) and its class in K1(A) is the identity.
S∗ (cid:19)
U :=(cid:18)S 1 − SS∗
0
12
SHAWN J. McCANN
Proposition 3.6. Let (T, π) denote the universal Toeplitz representation on Eα,β(Γ)
and let {uj}N −1
j=0 TjT ∗
j
where Tj = T (uj). Then, with the maps Ω : C(Γ) → MN (C(Γ)) and ρ : C(Γ) →
ker q ⊂ Tα,β(Γ) defined by
j=0 be an orthonormal basis of Eα,β(Γ). Further, let p = PN −1
and
Ω(a) = (hui, a · uji)N −1
i,j=0
ρ(a) = π(a)(1 − p)
as in Lemmas 3.1 and 3.3, the following two diagrams (i = 0, 1) commute:
(3.3)
Ki(C(Γ))
ρ∗
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
...
...
.
.
.
.
.
..
..
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
..
Ki(ker q)
1 − Ω∗
...........
...............................................................................................................................
.
.
ι∗
........
....................................................................................................................
.
...
..
Ki(C(Γ))
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
...
...
.
.
.
.
.
..
..
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
..
π∗
Ki(Tα,β(Γ))
Proof. First, let i = 0. Let z = (zs,t) ∈ Mr(C(Γ)) be a projection and let πr denote
the augmentation map, π ⊗ idr, of π on Mr(C(Γ)). Then
ρ∗([z]) = [(ρ(zs,t))s,t] = [(π(zs,t)(1 − p))s,t] = [πr(z)] − [πr(z)(p1r)]
and
Hence, it suffices to show that
π∗ ◦ (1 − Ω∗)([z]) = [πr(z)] − π∗ ◦ Ω∗([z]).
π∗ ◦ Ω∗([z]) = [πr(z)(p1r)].
Note that
so
Ω∗([z]) = [(Ω(zs,t))s,t] = [((huj, zs,t · uki)j,k)s,t],
π∗ ◦ Ω∗([z]) = [((π(huj, zs,t · uki))j,k)s,t] = [πrN ◦ Ωr(z)].
Set bj,s;k,t = π(huj, zs,t · uki) and C = (cm,n)m,n = πrN (Ωr(z)) as in Lemma 3.4.
Let
T01r T11r
0
...
0
0
...
0
T =
...
...
... TN −11r
...
0
...
0
∈ MN (Mr(Oα,β(Γ))).
Then T T ∗ = p1r ⊕ 0r(N1) and since πr(z) is a projection which commutes with p1r,
(πr(z) ⊕ 0r(N −1))T
is a partial isometry which implements a Murray-von Neumann equivalence between
T ∗(πr(z) ⊕ 0r(N −1))T
and
thus,
(πr(z) ⊕ 0r(N −1))T T ∗(πr(z) ⊕ 0r(N −1)) = πr(z)(p1r) ⊕ 0r(N −1);
[πr(z)(p1r)] = [T ∗(πr(z) ⊕ 0r(N −1))T ].
C ∗-ALGEBRAS ASSOCIATED WITH TOPOLOGICAL GROUP QUIVERS II
13
Furthermore,
T ∗(πr(z) ⊕ 0r(N −1))T = T ∗
πr(z)T0
... πr(z)TN −1
...
0
...
...
0
...
= (T ∗
j πr(z)Tk)j,k
so the (j, k) entry is (π(huj, zs,t · uki))s,t. Recall bj,s;k,t = π(huj, zs,t · uki) and so
T ∗(πr(z) ⊕ 0r(N −1))T = D = (dm,n)m,n as in Lemma 3.4. Thus, by Lemma 3.4, there
exists a unitary U such that C = U ∗DU which gives us
[πr(z)(p1r)] = [D] = [C] = [πrN ◦ ΩR(z)]
as desired.
For the case i = 1, let u ∈ Mr(C(Γ)) be a unitary. Note ρ∗ : K1(C(Γ)) →
K1(ker q) is the composition of a unital isomorphism of C(Γ) onto (1 −p) ker q(1 −p)
with the inclusion of (1−p) ker q(1−p) as a full corner in the non-unital algebra ker q;
that is, [u] 7→ [ρr(u)] = [πr(u)((1−p)1r)] 7→ [πr(u)((1−p)1r)+p1r] ∈ K1((ker q)+) =
K1(ker q). Furthermore,
π∗ ◦ Ω∗([u]) = [πr(u)] − [πrN ◦ Ωr(u)]
and hence, we need only show
[(πr(u)((1 − p)1r) + p1r) ⊕ 1r(N −1)] = [πr(u) ⊕ 1r(N −1)] − [πrN ◦ Ωr(u)]
in K1(Tα,β(G)).
We take a brief moment to make an aside: If C ∈ M2rN (Tα,β(Γ)) is invertible
with K1-class the identity 1 then the K1-class is unchanged by pre- and post-
multiplication by C. In particular, when C is equal to:
(1) (Lemma 3.5) a unitary of the form
S∗ (cid:19)
(cid:18)S 1 − SS∗
0
where S is an isometry
(2) an upper- or lower-triangular matrix of the form
(cid:18)1 A
0 1(cid:19)
or
(which are connected to 1 via t 7→(cid:18)1 tA
0
(3) any constant invertible matrix in GL2rN (C) (because GL2rN (C) is path con-
nected); this implies that row and column operations may be used without
changing the K1-class.
(cid:18) 1 0
A 1(cid:19)
1 (cid:19) and likewise for the transpose)
Recall: T =
T01r T11r
0
...
0
0
...
0
...
...
... TN −11r
...
0
...
0
.
14
SHAWN J. McCANN
With this in mind, calculate
[(πr(u)((1 − p)1r) + p1r) ⊕ 1r(N −1)]
0rN
=h(cid:18)(πr(u)((1 − p)1r) + p1r) ⊕ 1r(N −1) 0rN
=h(cid:18)(πr(u)((1 − p)1r) + p1r) ⊕ 1r(N −1) 0rN
=h(cid:18)((πr(u)(1 − p)1r) + p1r) ⊕ 1r(N −1))T π(u)((1 − p)1r) ⊕ 1r(N −1)
1rN(cid:19)ih(cid:18) T
1rN(cid:19)ih(cid:18) T
1rN − T T ∗
0rN
0rN
0rN
0rN
T ∗
T ∗
T ∗
(cid:19)i
(cid:19)i
(1 − p)1r ⊕ 1r(N −1)
and recall (1 − p)Ti = 0 by Lemma 3.2(3), hence (1 − p)1rT = 0 and
(cid:19)i
[(πr(u)((1 − p)1r) + p1r) ⊕ 1r(N −1)]
0rN
=h(cid:18) T
=h(cid:18) T
=h(cid:18) T
0rN
0rN
πr(u)((1 − p)1r) ⊕ 1r(N −1)
T ∗
πr(u)((1 − p)1r) ⊕ 1r(N −1)
T ∗
(cid:19)i
(cid:19)ih(cid:18)1rN T ∗(πr(u) ⊕ 1r(N −1))
1rN
0rN
(cid:19)i
πr(u) ⊕ 1r(N −1)
T ∗
(cid:19)i
since T T ∗ = p1r ⊕0r(N −1) and (p1r)πr(u) = πr(u)(p1r). Using elementary operations,
compute
[(πr(u)((1 − p)1r) + p1r) ⊕ 1r(N −1)]
T
T
0rN(cid:19)i
0rN(cid:19)ih(cid:18)1rN −(πr(u∗) ⊕ 1r(N −1))T
−T ∗(πr(u∗) ⊕ 1r(N −1))T(cid:19)i
1rN
0rN
0rN
(cid:19)i
T ∗
T ∗
=h(cid:18)πr(u) ⊕ 1r(N −1)
=h(cid:18)πr(u) ⊕ 1r(N −1)
=h(cid:18)πr(u) ⊕ 1r(N −1)
=h(cid:18)
=h(cid:18)πr(u) ⊕ 1r(N −1)
0rN
1rN
T ∗
0rN
−T ∗(πr(u∗) ⊕ 1r(N −1)) 1rN(cid:19)ih(cid:18)πr(u) ⊕ 1r(N −1)
−T ∗(πr(u∗) ⊕ 1r(N −1))T(cid:19)ih(cid:18)1rN
−T ∗(πr(u∗) ⊕ 1r(N −1))T(cid:19)i
0rN −1rN(cid:19)i
0rN
0rN
0rN
T ∗
= [πr(u) ⊕ 1r(N −1)] + [T ∗(πr(u∗) ⊕ 1r(N −1))T ].
Furthermore,
[T ∗(πr(u∗) ⊕ 1r(N −1))T ] = [πrN (Ωr(u−1))] = −[πrN (Ωr(u))].
Hence,
[(πr(u)((1 − p)1r) + p1r) ⊕ 1r(N −1)] = [πr(u) ⊕ 1r(N −1)] − [πrN ◦ Ωr(u)]
as desired.
(cid:3)
C ∗-ALGEBRAS ASSOCIATED WITH TOPOLOGICAL GROUP QUIVERS II
15
Theorem 3.7. Let (S, π) = q ◦ (T, π) be the universal Cuntz-Pimsner covariant
representation of Eα,β(Γ) in Oα,β(Γ). Then the following diagram is exact:
(3.4)
..
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
...
...
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
ρ−1
∗ ◦ δ0
K1(Oα,β(Γ))
K0(C(Γ))
1 − Ω∗
.
..
.........
..............................................................................................................................
..
K0(C(Γ))
π∗
.
..
.........
............................................................................................................
..
K0(Oα,β(Γ))
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
..
..
.
.
.
.
.
..
..
.
.
.
.
.
.
.
.
.
.
.
.
.
.
..
..
..ρ−1
∗ ◦ δ1
...
...........................................................................................................
..........
.. π∗
K1(C(Γ))
...
.............................................................................................................................
..........
.. 1 − Ω∗
K1(C(Γ))
Proof. Note ρ : C(Γ) → ker q is an isomorphism onto a full corner, implying ρ∗ is
an isomorphism. Further note π∗ : Ki(C(Γ)) → Ki(Tα,β(Γ)) is an isomorphism (see
comments prior to Lemma 3.1). Then (3.1) and the previous proposition give the
stated result.
(cid:3)
4. K-groups of OF,G(Γ)
In this section, the approach of [24] is made easier and extended. For this section,
let α, β ∈ End(C(Td)) defined by α = σ#
G where F = Diag(a1, ..., ad) ∈
Md(Z)+ and G ∈ Md(Z) such that det F > 0 and det G 6= 0. We know there exists
F and β = σ#
an orthonormal basis for EF,G(Td), {uj(cid:12)(cid:12) 0 ≤ j ≤ N − 1}; this is the basis {uν}ν∈I(F ),
described in Section 3.3, reindexed by 0, 1, ..., N − 1. Let Uj be the unitary defined
by Uj(x) = xj for x = (xi)d
i=1 ∈ Td for j ∈ {1, ..., d}. Further, let
and J ′ = {1, ..., d} \ J in increasing order. Define
{[1]0}
Ik = {J ⊂ {1, ..., d}(cid:12)(cid:12) J = k, J = {j1 < ... < jk}}
{[UJ ]0 = [Uj1]0 ∧ ... ∧ [Ujk]0(cid:12)(cid:12) J ∈ Ik} if k > 0 is even
{[UJ ]1 = [Uj1]1 ∧ ... ∧ [Ujk]1(cid:12)(cid:12) J ∈ Ik} if k > 0 is odd
if k = 0
Ek =
If it is understood, the notation [·] will be used in lieu of [·]i.
It is well known (see [34] and [33, Example 3.11 and 3.15]) that
with basis {Ek}k even and
K0(C(Td)) ∼= ^evens
K1(C(Td)) ∼= ^odds
Zd = Z2d−1
Zd = Z2d−1
with basis {Ek}k odd. For subsets J and I of the same size, define FJ,I to be the
square submatrix of F whose entries belong to the rows in J and the columns in I.
j=1 →
j=1 → span{[Uj]}d
With these identifications, the (K1-group) induced map α∗V1 Zd : span{[Uj]}d
j=1 is multiplication by F T = F, and β∗V1 Zd : span{[Uj]}d
j=1
span{[Uj]}d
is multiplication by GT , the transpose of G. We have
U bjk
k
β∗([Uj]) = [β(Uj)] = [U Gj ] = [Yk
bjk[Uk].
] =Xk
16
SHAWN J. McCANN
Do likewise to prove α∗V1 Zd is multiplication by F. One can also check that α∗ and
α∗[1] = [α(1)] = 1 = [β(1)] = β∗[1]
since α and β are group homomorphisms.
β∗ act on V0 Zd by
Lemma 4.1. For 1 ≤ k ≤ d, the matrix Ak representating α∗ : Vk Zd → Vk Zd
with respect to the basis Ek is the diagonal matrix Diag(aI)I∈Ik (aI =Qi∈I ai) and
matrix Bk representating β∗ :Vk Zd →Vk Zd is (det GJI)I,J∈Ik.
Proof. Begin by noting, Ak = Vk F T and Bk = Vk GT . Let [UI] ∈ Ek with I =
{i1 < ... < ik}. Then
k
k
β∗([UI]) = (
= GT [Ui1] ∧ ... ∧ GT [Uik ]
^ GT )([Ui1] ∧ ... ∧ [Uik ])
d
=
=
^ GT )[UI] = (
Xm1,...,mk=1
X[UJ ]∈Ek,J={m1,...,mk}
= X[UJ ]∈Ek Xσ∈Sk
= X[UJ ]∈Ek Xσ∈Sk
= X[UJ ]∈Ek
det GI,J[UJ ]
bi1,m1...bik,mk ([Um1] ∧ ... ∧ [Umk ])
bi1,m1...bik,mk[UJ ]
bi1,σ(j1)...bik,σ(jk)([Uσ(j1)] ∧ ... ∧ [Uσ(jk)])
(−1)deg σbi1,σ(j1)...bik,σ(jk)([Uj1] ∧ ... ∧ [Ujk])
where Sk denotes the symmetric group on k elements. The result for Ak follows by
specializing to the diagonal case.
(cid:3)
Let Ak and Bk (for k = 0, ..., d) denote the matrices described in Lemma 4.1; that
is, A0 = B0 = 1, Ak = Diag(aI)I∈Ik and Bk = (det GJI)I,J∈Ik for k ∈ {1, ..., d}.
From Lemma 3.1, the map Ω(σ#
F ) : C(Td) → MN (C(Td)) is the diagonal matrix
Ω(σ#
G and d = (ds,t) ∈ Mr(C(Td)),
G (a)1N and so, for α = σ#
F (a)) = σ#
F , β = σ#
Ω∗ ◦ α∗([d]) = [(Ω ◦ α)r(d)]
= [(Ω(α(ds,t))s,t]
= [(β(ds,t)1N )s,t]
= N[βr(d)]
= Nβ∗[d]
where Ωr, αr, βr denote the appropriate augmented maps on Mr(C(Td)). Using the
previous lemma and the above equation, the matrix Ck representing Ω∗ on Vk Zd
C ∗-ALGEBRAS ASSOCIATED WITH TOPOLOGICAL GROUP QUIVERS II
17
satisfies
CkAk = NBk
where it is expected that Ck is a matrix with integer entries for each k = 0, ..., d.
Recall Ak = Diag(aI)I∈Ik. Let I ′ = {1, ..., d} \ I ordered so that I = {i1 < ... <
ik}, I ′ = {ik+1 < ... < id}. Then set C0 = N = det F ∈ Z and
Ck = BkDiag(aI ′)I∈Ik = (aJ ′ det GJI)I,J∈Ik
for k = 1, ..., d. Note that A0 = B0 = 1 by the calculations before Lemma 4.1.
Hence,
C0A0 = NB0
and
CkAk = BkDiag(aI ′)I∈IkDiag(aI)I∈Ik
= BkDiag(aI∪I ′)I∈Ik
= BkDiag(a1,...,d)I∈Ik
= BkDiag(N)I∈Ik = NBk
for k = 1, ..., d.
In order to calculate the K-theory of OF,G(Td), one needs only to calculate ker(1−
Ck) and coker(1 − Ck) for each k = 0, ..., d as the next theorem will demonstrate.
Theorem 4.2. Let F = Diag(a1, ..., ad), G ∈ Md(Z) such that det G 6= 0 and aj ∈ N
for each j = 1, ..., d. With Ck defined as above and OF,G(Td) defined as in Example
2.17,
(1) K0(OF,G(Td)) =(cid:0) M0≤k≤d, even
(2) K1(OF,G(Td)) =(cid:0) M0≤k≤d, odd
coker(1 − Ck)(cid:1) ⊕(cid:0) M0≤k≤d, odd
coker(1 − Ck)(cid:1) ⊕(cid:0) M0≤k≤d, even
Proof. By 3.4 in Theorem 3.7,
ker(1 − Ck)(cid:1), and
ker(1 − Ck)(cid:1).
1 − Ck
M0≤k≤d, even
.
..
.........
.................................................................................................................................................................................................................................................
..
K0(C(Td)
...
..
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
..
..
.
.
...
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
K0(C(Td))
.
..
.........
........................................................................................................................................................................................................................
..
K0(OF,G(Td))
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
..
..
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
..
..
..
..
..
K1(OF,G(Td))
...
.......................................................................................................................................................................................................................
..........
..
K1(C(Td))
is exact, so there are two exact sequences
K1(C(Td))
...
...........................................................................................................................................................................................................................................
..........
..
1 − Ck
M0≤k≤d, odd
0
..........
................................................................................
..
..
.
and
0
..........
..................................................................................
...
..
M0≤k≤d, even
M0≤k≤d, odd
coker(1 − Ck)
..........
............................................................
..
..
.
K0(OF,G(Td))
..........
.............................................................................
..
..
.
coker(1 − Ck)
..........
..............................................................
...
..
K1(OF,G(Td))
..........
...........................................................................
...
..
M0≤k≤d, odd
M0≤k≤d, even
ker(1 − Ck)
..........
...............................................
..
..
.
ker(1 − Ck)
..
..........
.............................................
...
0
0
18
SHAWN J. McCANN
which split since Vk Zd and hence, ker(1 − Ck) is free for each k. Thus,
K0(OF,G(Td)) =(cid:0) M0≤k≤d, even
K1(OF,G(Td)) =(cid:0) M0≤k≤d, odd
coker(1 − Ck)(cid:1) ⊕(cid:0) M0≤k≤d, odd
coker(1 − Ck)(cid:1) ⊕(cid:0) M0≤k≤d, even
ker(1 − Ck)(cid:1)
ker(1 − Ck)(cid:1).
Definition 4.3. A matrix Z ∈ Md(Z) is called an integer dilation matrix provided
each eigenvalue λ of Z satisfies λ > 1.
Remark 4.4. The case where F is an integer dilation and G = 1d was computed
in [24, Theorem 4.9] where it was found that
and
and
K0(OF,1(Td)) =(cid:0) M0≤k≤d, even
K1(OF,1(Td)) =(cid:0) M0≤k≤d, odd
coker(1 − Qk)(cid:1) ⊕(cid:0) M0≤k≤d, odd
coker(1 − Qk)(cid:1) ⊕(cid:0) M0≤k≤d, even
ker(1 − Qk)(cid:1)
ker(1 − Qk)(cid:1)
for the matrix Qk satisfying the relation
k).
Qk(det FJI)I,J∈Ik = N1(d
Recall the Smith normal form of F, F = UDV where U, V ∈ Md(Z) are unimodular
matrices and D = Diag(aj)d
j=1 ∈ Md(Z) is a positive diagonal matrix. Then by
properties of matrix minors,
(det FJI)I,J∈Ik = UkDkVk
where Uk = (det UJI)I,J∈Ik, Vk = (det VJI)I,J∈Ik and Dk = Diag(aI)I∈Ik. Also note
for (U −1)k = (det(U −1)JI)I,J∈Ik,
Uk(U −1)k = (UU −1)k = 1(d
k);
that is, Uk is unimodular. Hence, for G = U −1V −1, QkUkDkVk = N1(d
k Diag(aI ′)I∈Ik = BkDiag(aI ′)I∈Ik = Ck.
U −1
k QkUk = U −1
k V −1
k) implies
Therefore,
and likewise,
ker(1 − Ck) = ker(U −1
k (1 − Qk)Uk) ∼= ker(1 − Qk)
coker(1 − Ck) ∼= coker(1 − Qk);
that is, Theorem 4.2 extends Theorem 4.9 of [24].
Here we first consider the case where F = n1d and G ∈ Md(Z). To this end, note
that Ck has a simple form. First of all, note Ak is nk1(d
k) and hence,
Ck = nd−kBk.
(see the calculations preceding Theorem 4.2.)
C ∗-ALGEBRAS ASSOCIATED WITH TOPOLOGICAL GROUP QUIVERS II
19
Theorem 4.5. For d ∈ N, let G ∈ Md(Z) (det G 6= 0) and n ∈ N. Then with
Ck = nd−kBk where Bk = (det GJ,I)I,J∈Ik ∈ M(d
k)(Z) as above,
(1) If d > 1 and n > 1, then
coker(1 − Ck)
coker(1 − Ck)(cid:1)
coker(1 − Ck)
if det G 6= 1
if det G = 1
if det G 6= 1
if det G = 1
(a) K0(On,G(Td)) =
M0≤k≤d, even
Z ⊕(cid:0) M0≤k≤d−1, even
(b) K1(On,G(Td)) =
M0≤k≤d, odd
Z ⊕(cid:0) M0≤k≤d−1, odd
(a) K0(On,G(Td)) = Z ⊕(cid:0) M1≤k≤d, even
(b) K1(On,G(Td)) = Z ⊕(cid:0) M1≤k≤d, odd
coker(1 − Ck)(cid:1)
coker(1 − Ck)(cid:1)
coker(1 − Ck)(cid:1)
(2) If n = 1 and G is an integer dilation matrix, then
(3) If d = 1, then
(a) K0(On,m(T)) = Zn−1 and K1(On,m(T)) = Zm−1 for n > 1 and m 6= 0, 1
(b) K0(O1,m(T)) = Z and K1(O1,m(T)) = Z ⊕ Zm−1 for m 6= 0, 1
(c) K0(On,1(T)) = Z ⊕ Zn−1 and K1(On,1(T)) = Z for n > 1
Proof. For (1), let d, n > 1 and det G 6= 1. Then C0 = nd 6= 1 and Cd = det G 6= 1;
that is, 1 − C0 and 1 − Cd are injective. Furthermore, for 1 ≤ k ≤ d − 1,
Ck = nd−k(det GJ,I)J,I∈Ik.
Since the characteristic polynomial of a matrix is monic, it follows from Gauss'
Lemma that any rational eigenvalue of a matrix in Md(Z) must actually be an
1
nd−k /∈ σ((det GJ,I)I,J) if and only if 1 /∈ σ(Ck). Thus, 1 − Ck is
integer. That is,
injective for each k = 1, ..., d − 1 and so, ker(1 − Ck) = 0 for each k = 0, ..., d. By
Theorem 4.2,
and
K0(On,G(Td)) = M0≤k≤d, even
K1(On,G(Td)) = M0≤k≤d, odd
coker(1 − Ck),
coker(1 − Ck).
Now assume det G = 1, then, as above, ker(1 − Ck) = 0 for k = 0, ..., d − 1. But
ker(1 − Cd) = Z and coker(1 − Cd) = Z whether d is even or odd. Theorem 4.2 gives
and
K0(On,G(Td)) = Z ⊕(cid:0) M0≤k≤d−1, even
K1(On,G(Td)) = Z ⊕(cid:0) M0≤k≤d−1, odd
coker(1 − Ck)(cid:1)
coker(1 − Ck)(cid:1).
20
SHAWN J. McCANN
For (2), let n = 1 and λ > 1 for all eigenvalues λ of G. Then C0 = 1 and
Cd = det G 6= 1. We now wish to show that det(1 − Ck) 6= 0 for k = 1, ..., d. To
this end, choose a basis of Cd such that G becomes upper triangular (not necessarily
with integer entries); that is,
a11 a12
0
a22
...
. . .
0
0
G =
... a1d
... a2d
. . .
... add
.
Then Ck is lower triangular with diagonal entries det GII = Qi∈I aii
6= 1 (since
λ > 1 for all λ ∈ σ(G)) and so det(1 − Ck) 6= 0. Hence, 1 − Ck is injective for
k = 1, ..., d and ker(1 − C0) = coker(1 − C0) = Z. Theorem 4.2 now gives
and
K0(On,G(Td)) = Z ⊕(cid:0) M1≤k≤d, even
K1(On,G(Td)) = Z ⊕(cid:0) M1≤k≤d, odd
coker(1 − Ck)(cid:1),
coker(1 − Ck)(cid:1).
Finally, let us prove (3).
If n > 1 and m 6= 0, 1 then 1 − n and 1 − m are
injective and coker(1 − n) = Zn−1 := Z/(n − 1)Z; likewise, coker(1 − m) = Zm−1.
Thus, K0(On,m(T)) = Zn−1 and K1(On,m(T)) = Zm−1. If n = 1 and m 6= 0, 1, then
ker(1 − n) = coker(1 − n) = Z, coker(1 − m) = Zm−1 and ker(1 − m) = 0 thus,
K0(O1,m(T)) = Z and K1(O1,m(T)) = Z ⊕ Zm−1. If n > 1 and m = 1, then similarly,
K0(On,1(T)) = Z ⊕ Zn−1 and K1(On,1(T)) = Z.
Corollary 4.6. Let d = 2, n = 1, and 1 /∈ σ(G), the spectrum of G. Then
(cid:3)
(1) if det G = 1, then
(a) K0(O1,G(T2)) = Z2
(b) K1(O1,G(T2)) = Z2 ⊕ coker(1 − GT )
(2) if det G 6= 1, then
(a) K0(O1,G(T2)) = Z ⊕ Z1−det G
(b) K1(O1,G(T2)) = Z ⊕ coker(1 − GT )
Proof. Begin with calculating C0 = n2 = 1, C1 = nGT = GT and C2 = det G. With
1 not an eigenvalue of G, it is guaranteed that det(1 − GT ) 6= 0 and hence, 1 − C1
is injective. This leaves us to calculate
(1) ker(1 − C0) = Z
(2) coker(1 − C0) = Z
(3) ker(1 − C1) = 0
(4) coker(1 − C1) = coker(1 − GT )
(5) ker(1 − C2) =(Z if det G = 1
if det G 6= 1
0
, and
C ∗-ALGEBRAS ASSOCIATED WITH TOPOLOGICAL GROUP QUIVERS II
21
(6) coker(1 − C2) =(Z
if det G = 1
Z1−det G if det G 6= 1
We now calculate the K-theory when F, G ∈ Md(Z) are both diagonal matrices.
Let F = Diag(a1, ..., ad) and G = Diag(b1, ..., bd) be diagonal integral matrices of
non-zero determinant and such that 1 ≤ a1 ≤ ... ≤ ad. Let f denote the number of
(cid:3)
1's in F ; that is, 1 = a1 = ... = af < af +1 ≤ ... ≤ ad. Let aI = Qi∈I ai for I ∈ Ik.
Then Ak = Diag(aI)I∈Ik, Bk = Diag(bI)I∈Ik and
Ck = (det F )BkA−1
k = Diag(bI aI ′)I∈Ik.
So ker(1 − Ck) = Zdk where dk is the number of 1's in Ck; that is, the number of I
making bIaI ′ = 1. Furthermote, bI aI ′ = 1 implies bI = aI ′ = 1 and so {f + 1, ..., d} ⊂
I. So let v be the number of negative ones in {bf +1, ..., bd} and let p be the number
of ones in the same set. Then
dk =Xr∈I (cid:18) v
2r(cid:19)(cid:18)
p
(k − d + f ) − 2r(cid:19)
where I = {r ∈ N(cid:12)(cid:12) 0 ≤ 2r ≤ v, p − (k − d + f ) ≤ 2r ≤ k − d + f }.
Lemma 4.7. With the context of the preceding paragraph, let p > 0 be the number
of ones while v is the number of negative ones. Let pk(p, v) be the number of
combinations of choosing k digits of 1′s and −1′s that multiply to 1 and let vk(p, v)
be similar, but multiplying to −1; that is,
and
pk(p, v) = X0,k−p≤2r≤v,k(cid:18) v
vk(p, v) = X0,k−p≤2r+1≤v,k(cid:18) v
2r(cid:19)(cid:18) p
2r + 1(cid:19)(cid:18)
k − 2r(cid:19)
k − 2r − 1(cid:19)
p
pk(p, v) =(2p
2p+v−1
if v = 0 and p > 0
if v 6= 0
.
X0≤k≤p
for 0 < k ≤ p, p0(p, v) = 1 if p > 0, and v0(p, v) = vk(p, v) = pk(p, v) = p0(0, v) = 0
for k > p. Then
Proof. First realize that pk(p, v) + vk(p, v) =(cid:0)p+v
nations of choosing k digits to multiply to either 1 or −1. Thus,
k (cid:1) since it is the number of combi-
Xk
pk(p, v) +Xk
So if v = 0, then the proof is done.
vk(p, v) =Xk (cid:18)p + v
k (cid:19) = 2p+v.
Assume v 6= 0 and let v be odd. Then claim vk(p, v) = pp−k(p, v). This can
be easily seen by realizing that, for each choice of k digits to multiply to −1, the
22
SHAWN J. McCANN
remaining digits multiply to 1. Thus, Pk vk(p, v) = Pk pp−k(p, v) = Pk pk(p, v)
and so, Pk pk(p, v) = 1
Now let v be even. Then, for even k 6= 0
22p+v = 2p+v−1.
p
p
k − 2r − 1(cid:19)
2r + 1(cid:19)(cid:18)
vk(p, v) − vk(p, v − 1) = X0,k−p≤2r+1≤v,k(cid:18) v − 1
k − 2r − 1(cid:19)
2r + 1(cid:19)(cid:18)
− X0,k−p+1≤2r+1≤v,k(cid:18) v − 1
2r + 1(cid:19)](cid:18)
2r + 1(cid:19) −(cid:18) v − 1
[(cid:18) v
= X0,k−p≤2r+1≤v,k
k − 2r − 1(cid:19)
2r (cid:19)(cid:18)
= X0,k−p≤2r+1≤v,k(cid:18)v − 1
(k − 1) − 2r(cid:19)
2r(cid:19)(cid:18)
X0,(k−1)−p≤2r≤v−1,k−1(cid:18) v
p − 1
=
p
= pk−1(p, v − 1).
p
k − 2r − 1(cid:19)
Since v − 1 is odd and v0(p, v) = 0, we get
Xk
vk(p, v) =Xk>0
and consequently,
vk(p, v − 1) +Xk>0
pk−1(p, v − 1) = 2p+v−2 + 2p+v−2 = 2p+v−1
pk(p, v) = 2p+v−1.
Xk
(cid:3)
Theorem 4.8. Let F = Diag(a1, ..., ad) and G = Diag(b1, ..., bd) be diagonal integral
matrices of non-zero determinant and such that 1 ≤ a1 ≤ ... ≤ ad. Let f > 0 denote
the number of 1's in F , let v be the number of negative ones in {bf +1, ..., bd} and let
p be the number of ones in that same set. Then
(1) if p = 0 and v = 0, then
(2) if p = 0 and v > 0, then
Z1−bI aI ′ ]
Z1−bI aI ′ ]
[MI∈Ik
[MI∈Ik
(a) K0(OF,G(Td)) = M0≤k≤d, even
(b) K1(OF,G(Td)) = M0≤k≤d, odd
(a) K0(OF,G(Td)) = Z2v−1−1 ⊕(cid:0) M0≤k≤d, odd
(b) K1(OF,G(Td)) = Z2v−1−1 ⊕(cid:0) M0≤k≤d, odd
(a) K0(OF,G(Td)) = Z2p ⊕(cid:0) M0≤k≤d, even
(3) if v = 0, p > 0, we have
[ MI∈Ik,bI aI ′ 6=1
[ MI∈Ik,bI aI ′ 6=1
[ MI∈Ik,bI aI ′ 6=1
Z1−bI aI ′ ](cid:3)
Z1−bI aI ′ ](cid:1)
Z1−bI aI ′ ](cid:1)
C ∗-ALGEBRAS ASSOCIATED WITH TOPOLOGICAL GROUP QUIVERS II
23
(4) if v 6= 0, p > 0, then
(b) K1(OF,G(Td)) = Z2p ⊕(cid:0) M0≤k≤d, odd
(a) K0(OF,G(Td)) = Z2p+v−1 ⊕(cid:0) M0≤k≤d, even
(b) K1(OF,G(Td)) = Z2p+v−1 ⊕(cid:0) M0≤k≤d, odd
Proof. Begin by calculating
[ MI∈Ik,bI aI ′ 6=1
Z1−bI aI ′ ](cid:1)
[ MI∈Ik,bI aI ′ 6=1
[ MI∈Ik,bI aI ′ 6=1
Z1−bI aI ′ ](cid:1)
Z1−bI aI ′ ](cid:1)
Zdk
ker(1 − Ck) = Mk, even(odd)
coker(1 − Ck) = Mk, even(odd)(cid:0) MbI aI ′ 6=1,I∈Ik
(1) Mk, even(odd)
(2) Mk, even(odd)
K0(OF,G(Td)) =(cid:0) M0≤k≤d, even
coker(1 − Ck)(cid:1) ⊕(cid:0) M0≤k≤d, odd
Thus,
Zdk(cid:1).
Z1−bI aI ′(cid:1) ⊕(cid:0) Mk, even(odd)
ker(1 − Ck)(cid:1)
Z1−bI aI ′(cid:1) ⊕(cid:0) Mk, even
Zdk(cid:1)
Z1−bI aI ′(cid:1)
Z1−bI aI ′(cid:1).
Zdk(cid:1) ⊕(cid:0) Mk, even(cid:0) MbI aI ′ 6=1,I∈Ik
=(cid:0) Mk, odd
=(cid:0)Mk
Zdk(cid:1) ⊕(cid:0) Mk, even(cid:0) MbI aI ′ 6=1,I∈Ik
= ZPk dk ⊕(cid:0) Mk, even(cid:0) MbI aI ′ 6=1,I∈Ik
Similarly for K1;
K1(OF,G(Td) = ZPk dk ⊕(cid:0) Mk, odd(cid:0) MbI aI ′ 6=1,I∈Ik
Z1−bI aI ′(cid:1).
If p = 0 and v = 0, then there is no way to multiply to get 1. Hence, 1 − Ck is
injective for all k and the result follows readily. If p = 0 and v > 0, then dk = 0 for
all odd k or k > v and dk =(cid:0)v
dk = X2≤2k≤v(cid:18) v
k(cid:1) for all even k ≤ v. Thus,
2k(cid:19) = ( X0≤2k≤v(cid:18) v
Xk
2k(cid:19)) − 1 = 2v−1 − 1
and the rest follows.
If p > 0, then
dk =Xr∈I (cid:18) v
2r(cid:19)(cid:18)
p
(k − d + f ) − 2r(cid:19)
where I = {r ∈ N(cid:12)(cid:12) 0 ≤ 2r ≤ v, p − (k − d + f ) ≤ 2r ≤ k − d + f }. Let k′ = k − d + f
and then
dk = pk′(p, v)
24
SHAWN J. McCANN
where pk′(p, v) was defined in the above lemma and so,
Xk
dk =Xk
and once again the rest follows.
pk(p, v) =(2p
2p+v−1
if v = 0
if v 6= 0
Remark 4.9. If f, the number of ones in F, is 0 then Ck is injective for k =
0, ..., d − 1. It is then very simple to calculate the K-groups whether det G = 1 or
det G 6= 1.
Corollary 4.10. If F = n1d, G = m1d ∈ Md(Z) where n ∈ N and m ∈ Z are
non-zero, then
(1) if either n > 1 or m 6= 1, then
(cid:3)
(a) K0(On,m(Td)) = M0≤k≤d, even
(b) K1(On,m(Td)) = M0≤k≤d, odd
k)
(d
1−nd−kmk
Z
k)
(d
1−nd−kmk
Z
(2) if n = m = 1, then K0(O1,1(Td)) = K1(O1,1(Td)) = Z2d
(3) if n = 1, m = −1, then K0(O1,−1(Td)) = K1(O1,−1(Td)) = Z2d−1 ⊕ Z2d−1
2
5. Acknowledgements
This paper is the second product of my doctoral dissertation at the University
of Calgary. It would not have been possible without Dr. Berndt Brenken for his
insight on corrections and enhancements of the material discussed here.
I am also indebted to NSERC, the Department of Mathematics and Statistics at
the University of Calgary and the Department of Mathematics and Statistics at the
University of Regina in providing funding to finance my mathematical studies.
References
[1] an Huef, A., and Raeburn, I., The ideal structure of Cuntz-Krieger algebras, Ergodic Theory
Dyn. Sys. 17 (1997) 611-624.
[2] Bates, T., Hong, J. H., Raeburn, I., and Szymanski, W., The ideal structure of C∗-algebras
of infinite graphs, Illinois J. Math 46 (2002), 1159-1176.
[3] Bates, T., Pask, D., Raeburn, I., and Szymanski, W., The C ∗-algebras of row-finite graphs,
New York J. Math. 6 (2000), 307-324.
[4] Blackadar, B., K-theory for Operator Algebras, M. S. R. I. Monographs, vol. 5, Springer-
Verlag, Berlin and New York, 1986.
[5] Blackadar, B., Operator Algebras: Theory of C*-Algebras and von Neumann Algebras, En-
cyclopaedia of Mathematical Sciences, vol. 122, Springer-Verlag, Berlin, 2006.
[6] Brenken, B., The local product structure of expansive automorphisms of solenoids and their
associated C ∗-algebras, Can. J. Math., 48, (1996), 692-709.
[7] Brenken, B., C ∗-algebras associated with topological relations, J. Ramanujan Math. Soc. 19,
No. 1 (2004), 1-21.
[8] Brenken, B., Endomorphisms of type I von Neumann algebras with discrete center, J. Operator
Theory 51 (2004), no. 1, 19-34.
C ∗-ALGEBRAS ASSOCIATED WITH TOPOLOGICAL GROUP QUIVERS II
25
[9] Brenken, B., The isolated ideal of a correspondence associated with a topological quiver. New
York J. Math., 12, (2006), 1-16.
[10] Brenken, B., A Dynamical Core for Topological Directed Graphs, Munster J. of Math. 3 (2010),
111-144.
[11] Brenken, B., Topological Quivers as Multiplicity Free Relations, Math. Scand., 106, (2010),
217-242.
[12] Brown, N.P., and Ozawa, N., C ∗-Algebras and Finite-Dimensional Approximations, Graduate
Studies in Mathematics, 88. Amer. Math. Soc. Providence, Rhode Island, 2008.
[13] Conway, J.B., A Course in Functional Analysis, Second Edition, Graduate Texts in Mathe-
matics, 96. Springer, New York, 1990.
[14] Cuntz, J., Simple C ∗-algebras generated by isometries. Comm. Math. Phys. 57 (1977), no. 2,
173-185.
[15] Cuntz, J., and Krieger, W., A class of C ∗-algebras and topological Markov chains, Inventiones
Math., 56 (1980), 251-268.
[16] Davidson, K., C ∗-Algebras by Example, Fields Institute Monograph, Amer. Math. Soc. Prov-
idence, Rhode Island, 1996.
[17] Deaconu, V., Groupoids associated with endomorphisms, Trans. Amer. Math. Soc. 347 (1995),
1779-1786.
[18] Deaconu, V., A path model for circle algebras, J. Operator Theory 34 (1995), 57-89.
[19] Deaconu, V., Generalized Cuntz-Krieger algebras, Proc. Amer. Math. Soc. 124 (1996), 3427-
3435.
[20] Deaconu, V., Continuous graphs and C ∗-algebras, in Operator theoretical methods (Timisoara,
1998), 137-149, Theta Found., Bucharest, 2000.
[21] Deaconu, V., and Muhly, P., C ∗-algebras associated with branched coverings, Proc. Amer.
Math. Soc. 129 (2001), 1077-1086.
[22] Dummit, D.S., and Foote, R.M., Abstract Algebra, 3rd edition, Wiley and Sons, 2004.
[23] Evans, D.E., The C*-algebras of topological Markov chains, Lecture Notes, Tokyo Metropolitan
University (1983).
[24] Exel, R., an Huef, A., and Raeburn, I., Purely Infinite Simple C ∗-algebras associated to Integer
Dilation Matrices, Indiana Univ. Math. J. 60 (2011), no. 3, 1033-1058.
[25] Fowler, N.J., Laca, M., and Raeburn, I., The C ∗-algebras of infinite graphs, Proc. Amer.
Math. Soc. 8 (2000), 2319-2327.
[26] Fowler, N.J., Muhly, P.S., and Raeburn, I., Representations of Cuntz-Pimsner Algerbas, Indi-
ana Univ. Math. J., 52(3) (2003), 569-605.
[27] Fowler, N.J., Raeburn, I., The Toeplitz algebra of a Hilbert bimodule, Indiana Univ. Math. J.
48 (1999), 155-181.
[28] Gabriel, P., Unzerlegbare Darstellungen I (Oberwolfach 1970), Manuscr. Math. 6 (1972), 71-
103.
[29] Hajac, P. M., Matthes, R., and Szyma´nski, W., Graph C ∗-algebras and Z2-quotients of quan-
tum spheres. Proceedings of the XXXIV Symposium on Mathematical Physics (Toru, 2002).
Rep. Math. Phys. 51 (2003), no. 2-3, 215-224.
[30] Hajac, P.M., Matthes, R., and Szyma´nski, W., Quantum real projective space, disc and
spheres, Algebr. Represent. Theory 6 (2003), 169-192.
[31] Hong, J.H., and Szyma nski, W., Quantum spheres and projective spaces as graph algebras,
Comm. Math. Phys. 232 (2002), 157-188.
[32] Han, D., Jing, W., Larson, D., and Mohapatra, R., Riesz bases and their dual modular frames
in Hilbert C ∗-modules. J. Math. Anal. Appl. 343 (2008), no. 1, 246-256.
[33] Hatcher, A., Algebraic Topology, Cambridge University Press, New York, 2001.
[34] Ji, R., On Crossed Product C ∗-Algebras Associated with Furstenberg Transformations on Tori,
PhD Thesis, State University of New York, Stony Brook, 1986.
26
SHAWN J. McCANN
[35] Kadison, R.V., and Ringrose, J.R., Fundamentals of the theory of operator algebras. Vol. I.
Elementary theory, Graduate Studies in Mathematics, 15. Amer. Math. Soc., Providence, RI,
1997.
[36] Kadison, R.V., and Ringrose, J.R., Fundamentals of the theory of operator algebras. Vol. II.
Advanced theory, Graduate Studies in Mathematics, 16. Amer. Math. Soc., Providence, RI,
1997.
[37] Kajiwara, T., and Watatani, Y., Hilbert C ∗-bimodules and continuous Cuntz-Krieger algebras.
J. Math. Soc. Japan 54 (2002), no. 1, 35-59.
[38] Katsura, T., On C ∗-algebras associated with C ∗-correspondences. J. of Functional Analysis
217 (2004), 366-401.
[39] Katsura, T., A class of C ∗-algebras generalizing both graph algebras and homeomorphism C ∗-
algebras. I. Fundamental results. Trans. Amer. Math. Soc. 356 (2004), no. 11, 4287-4322.
[40] Katsura, T., A class of C ∗-algebras generalizing both graph algebras and homeomorphism C ∗-
algebras. II. Examples. Internat. J. Math. 17 (2006), no. 7, 791-833.
[41] Katsura, T., A class of C ∗-algebras generalizing both graph algebras and homeomorphism C ∗-
algebras. III. Ideal structures. Ergodic Theory Dynam. Systems 26 (2006), no. 6, 1805-1854.
[42] Katsura, T., A class of C ∗-algebras generalizing both graph algebras and homeomorphism C ∗-
algebras. IV. Pure infiniteness. J. Funct. Anal. 254 (2008), no. 5, 1161-1187.
[43] Kumjian, A.,Notes on C*-algebras of graphs, Contemporary Math. 228 (1998) 189-200.
[44] Kumjian, A., and Pask, D., Higher rank graph C ∗-algebras, New York J. Math. 6 (2000), 1-20.
[45] Kumjian, A., and Pask, D., Actions of Zk associated to higher rank graphs, Ergodic Theory
& Dynamical Systems, 23 (2003), 1153-1172.
[46] Kumjian, A., Pask, D., and Raeburn, I., Cuntz-Krieger algebras of directed graphs, Pacific J.
Math. 184 (1998), 161-174.
[47] Kumjian, A., Pask, D., Raeburn, I., and Renault, J., Graphs, groupoids, and Cuntz-Krieger
algebras, J. Funct. Anal. 144 (1997), 505-541.
[48] Lance, E.C., Hilbert C ∗-modules: A toolkit for operator algebraists, London Mathematical
Society Lecture Note Series, vol. 210, Cambridge University Press, 1995.
[49] Mann, M.H., Raeburn, I., and Sutherland, C.E., Representations of finite groups and Cuntz-
Krieger algebras, Bull. Austral. Math. Soc. 46 (1992), 225243.
[50] Mann, M.H., Raeburn, I., and Sutherland, C.E., Representations of compact groups, Cuntz-
Krieger algebras, and groupoid C ∗-algebras in Miniconference on probability and analysis
(Sydney, 1991), 135144, Proc. Centre Math. Appl. Austral. Nat. Univ., 29, Austral. Nat.
Univ., Canberra, 1992.
[51] McCann, S.J., C ∗-algebras associated with topological group quivers, PhD Thesis, University
of Calgary, Calgary, Alberta, Canada, 2012.
[52] McCann, S.J., C ∗-algebras associated with topological group quivers I: generators, relations
and spatial structure, preprint
[53] Muhly, P., and Solel, B., Tensor algebras over C ∗-correspondences (representations, dilations,
and C ∗-envelopes), J. Funct. Anal. 158 (1998), 389-457.
[54] Muhly, P., and Solel, B., On the Morita Equivalence of Tensor algebras, Proc. London Math.
Soc. 81 (2000), 113-168.
[55] Muhly, P., and Tomforde, M., Adding tails to C ∗-correspondences, Doc. Math. 9 (2004), 79-
106.
[56] Muhly, P., and Tomforde, M., Topological quivers. Internat. J. Math. 16 (2005), no. 7, 693-755.
[57] Munkres, J. R., Topology, 2nd edition, Prentice-Hall, 2000.
[58] Pask, D., and Sutherland, C.E., Filtered inclusions of path algebras; a combinatorial approach
to Doplicher-Roberts duality, J. Operator Theory 31 (1994), 99-121.
[59] Paulsen, V.I., Completely bounded maps and operator algebras, Cambridge Studies in Ad-
vanced Math., 78, Cambridge University Press, Cambridge, 2002.
C ∗-ALGEBRAS ASSOCIATED WITH TOPOLOGICAL GROUP QUIVERS II
27
[60] Pimsner, M., A class of C ∗-algebras generating both Cuntz-Krieger algebras and crossed prod-
ucts by Z, in Free Probability Theory, fields inst. Commun., vol. 12, Amer. Math. Soc.,
Providence, 1997, pages 189-212.
[61] Schweizer, J., Crossed Product by C ∗-correspondences and Cuntz-Pimsner Algebras, 'in C ∗-
algebras, Proceedings of the SFB-Workshop on C ∗-algebras, Muenster, 1999,' (Eds.) J. Cuntz,
S. Echterhoff, Springer Verlag, Berlin, 2000.
[62] Stacey, P.J., Crossed products of C ∗-algebras by endomorphisms, J. Austral. Math. Soc. (Series
A) 54 (1993), 204-212.
[63] Raeburn, I., Williams, D.P., Morita Equivalence and Continuous-Trace C ∗-Algebras, Math.
Surveys & Monographs, vol. 60, Amer. Math. Soc., Providence, 1998.
[64] Rieffel, M., C ∗-algebras associated with irrational rotations, Pacific J. Math. 93(2) (1981),
415-429.
[65] Rφrdam, M., Larsen, F., Lausten, N.J., An Introduction to K-theory for C*-algebras. 256
pp. London Mathematical Society, Student Text 49, Cambridge University Press, Cambridge,
2000.
[66] Rφrdam, M., Stφrmer, E., Operator Algebras and Non-Commutative Geometry, Vol VII:
Classification of Nuclear C*-Algebras. Entropy in Operator Algebras. Encyclopaedia of Math-
ematical Sciences 126. Springer Verlag, Heidelberg, 2001.
[67] Walters, P., An introduction to ergodic theory, Springer, New York, 1982.
[68] Yamashita, S., Circle Correspondence C ∗-algebras, Houston J. Math. 37 (2011), no. 4, 1181-
1202.
|
1906.09723 | 1 | 1906 | 2019-06-24T04:54:34 | Absolutely compatible pair of elements in a von Neumann algebra-II | [
"math.OA"
] | Let $A$ be a unital C$^*$-algebra with unity $1_A$. A pair of elements $0 \le a, b \le 1_A$ in $A$ is said to be \emph{absolutely compatible} if, $\vert a - b \vert + \vert 1_A - a - b \vert = 1_A.$ In this paper we provide a complete description of absolutely compatible pair of strict elements in a von Neumann algebra. The end form of such a pair has a striking resemblance with that of a `generic pair' of projections on a complex Hilbert space introduced by Halmos. | math.OA | math |
ABSOLUTELY COMPATIBLE PAIR OF ELEMENTS IN A VON
NEUMANN ALGEBRA-II
ANIL KUMAR KARN
Abstract. Let A be a unital C∗-algebra with unity 1A. A pair of elements
0 ≤ a, b ≤ 1A in A is said to be absolutely compatible if, a − b + 1A − a − b =
1A. In this paper we provide a complete description of absolutely compatible
pair of strict elements in a von Neumann algebra. The end form of such a
pair has a striking resemblance with that of a 'generic pair' of projections on
a complex Hilbert space introduced by Halmos.
1. Introduction
Let A be a C∗-algebra. A pair of elements a, b ∈ A is said to be orthogonal, if
ab = 0 = ba = a∗b = ab∗.
Orthogonal pair of positive elements play an important role in the theory of C∗-
algebras. For example, it follows from the functional calculus that every self-adjoint
element a ∈ Asa has a unique decomposition: a = a+ − a− in A+, where a+
is algebraically orthogonal to a−. Recently, the author proved an order theoretic
characterization of algebraic orthogonality among positive elements of a C∗-algebra
[7]. (Also see [5, 6].)
Orthogonal pairs of positive elements of norm ≤ 1 exhibit an interesting property.
Let A be a unital C∗-algebra with unity 1A. For a pair of elements 0 ≤ a, b ≤ 1A
in A, we have ab = 0 (a is algebraically orthogonal to b) if and only if a + b ≤ 1A
and a − b + 1A − a − b = 1A. We isolate the later part and propose the following
definition: A pair of elements 0 ≤ a, b ≤ 1A in A is said to be absolutely compatible,
([7]), if
a − b + 1A − a − b = 1A.
It was proved in [7] that a projection p in a C∗-algebra A is absolutely compatible
with a positive element a of A with kak ≤ 1 if, and only if, ap = pa. However,
the two notions are distinct in general. The notion of absolute compatibility was
introduced as an instrument to prove a spectral decomposition theorem in the
context of "absolute order unit spaces" [7]. As an absolute order unit space is
an order theoretic generalization of unital C∗-algebras [7], absolute compatibility
appears to be a significant property.
Keeping this point of view, the author, along with Jana and Peralta, initiated
a study of absolute compatibility in operator algebras [3, 4]. Let M be a von
Neumann algebra and let 0 ≤ a ≤ 1M . We write
s(a) := sup{p ∈ P(M ) : p ≤ a}
2010 Mathematics Subject Classification. Primary 46L10; Secondary 46B40.
Key words and phrases. Absolutely ordered space, absolute oder unit space, order isometry,
absolute value preserving maps, absolute matrix order unit space.
1
2
and
ANIL KUMAR KARN
n(a) := sup{p ∈ P(M ) : pa = 0}.
For 0 ≤ a ≤ 1M , we say that a is strict in M , if s(a) = 0 and n(a) = 0.
In
[3], it was proved that an absolutely compatible pair of positive elements in a von
Neumann algebra has a (matricial) decomposition as a direct sum of commuting
and 'strict' elements. Let 0 ≤ a, b ≤ 1M such that a is absolutely compatible with
b. Then there exist mutually orthogonal projections p1, p2, s, n1, n2 ∈ P(M ) with
p1 + p2 + s + n1 + n2 = 1M such that
and
a = p1 ⊕ a1 ⊕ a2 ⊕ 0 ⊕ a3
b = b1 ⊕ p2 ⊕ b2 ⊕ b3 ⊕ 0
with respect to {p1, p2, s, n1, n2} with a2 and b2 are strict and absolutely compatible
in sM s [3, Theorem 2.10]. Thus the study of absolutely compatible pair of elements
reduces to such strict pairs. It was further proved that
Theorem 1.1. [3, Theorem 2.12] Let M be a von Neumann algebra, and let a, b
be elements in [0, 1]M . Then a is absolutely compatible with b if and only if there
exists a projection p1 in M so that a and b have matrix representations, say a =
(cid:18) a11 a12
12 a22 (cid:19) , and b = (cid:18) b11
b∗
12
a∗
b12
b22 (cid:19) with respect to the set {p1, 1 − p1 = p2} (i.e.,
aij = piapj and bij = pibpj) satisfying:
(i) a12 + b12 = 0;
(ii) a12a∗
(iii) a∗
(iv) a12 = a11a12 + a12a22 = b11a12 + a12b22.
12 = (p1 − a11)(p1 − b11);
12a12 = a22b22 = b22a22;
Using these decompositions, a complete description of absolutely compatible pair
of strict elements was given for the finite dimensional algebras Mn in the same paper
[3, Theorem 3.9].
In the present paper we continue the investigation and provide a complete de-
scription of absolutely compatible pair of strict elements in a von Neumann algebra.
However, as expected, the finite dimensional (matricial) techniques used in [3] fail
to work. At this juncture, the author would like to thank Professor Kalyan B.
Sinha for his suggestion to use polar decomposition. This idea works just perfectly.
The main results of the paper are Theorems 1.2 and 1.3.
Theorem 1.2. Let M be a von Neumann algebra with the underlying Hilbert space
H. Assume that a, b ∈ [0, 1]M be strict and absolutely compatible pair. Put p =
1 − r(a ◦ b).
(1) Then pH is isometrically isomorphic to (1−p)H. In particular, H ≡ K⊕K,
where K = pH.
(2) There exist strict elements a1, b1 ∈ [0, p] ∩ M with a1b1 = b1a1, a1 + b1 ≤ p
together with p − (a1 + b1) strict in pM p; and a unitary U : H → K ⊕ K
such that
a = U ∗(cid:20)
a1
(a1b1)
1
2
1
2
(a1b1)
p − a1 (cid:21) U and b = U ∗(cid:20)
b1
−(a1b1)
1
2
1
2
−(a1b1)
p − b1 (cid:21) U.
ABSOLUTE COMPATIBILITY
3
Theorem 1.3. Let M be a von Neumann algebra with the underlying Hilbert space
H. Assume that a, b ∈ [0, 1]M be strict and commuting pair such that a2 + b2 ≤ 1
with 1 − (a2 + b2) strict. Put a1 = (cid:20)a2
ab
ab
1 − a2(cid:21) and b1 = (cid:20) b2
−ab
−ab
1 − b2(cid:21). Then
a1, b1 ∈ [0, 1]M2(M) and a1 is absolutely compatible with b1.
It is interesting and surprising to note that the end form of an absolutely com-
patible pair of strict elements bears a striking resemblance with that of a 'generic
pair' of projections on a complex Hilbert space (studied by Halmos [2]).
A pair of closed subspaces M and N of a Hilbert space H is said to be in generic
position, if each of the following four subspaces of H: M ∩ N , M ∩ N ⊥, M ⊥ ∩ N
and M ⊥ ∩ N ⊥ are trivial. Let P and Q be the projections of H on M and N
respectively. In [2], Halmos proved that if M and N are in generic position, then
there exists a Hilbert space K, commuting, positive and invertible contractions C
and S in B(K) and a unitary operator U : H → K ⊕ K such that
P = U ∗(cid:20) C2 CS
CS
S2(cid:21) U and Q = U ∗(cid:20) C2 −CS
S2 (cid:21) U.
−CS
This observation further signifies the importance of a pair of strict absolutely
compatible elements in an operator algebra. This paper is in the sequel of [3].
We obtain some basic results to prove Theorems 1.2 and 1.3.
2. The main results
Lemma 2.1. Let M be a von Neumann algebra with the underlying Hilbert space
H and let a ∈ [0, 1]M be strict. If p ∈ P(M ), then, for the matrix representation
a = (cid:20)a11 a12
12 a22(cid:21) with respect to {p, 1 − p}, we have a11 ∈ [0, p] and a22 ∈ [0, 1 − p]
a∗
are also strict in the von Neumann algebras pM p and (1 − p)M (1 − p) respectively.
Proof. Without any loss of generality, we may assume that p 6= 0, p 6= 1. First, we
show that r(a11) = p. Put p − r(a11) = p0. Then p0a11 = 0 so that
0 ≤ (cid:20)p0
0
0
1 − p(cid:21)(cid:20)a11 a12
12 a22(cid:21)(cid:20)p0
a∗
0
Thus p0a12 = 0 and consequently,
0
1 − p(cid:21) = (cid:20) 0
a∗
12p0
p0a12
a22 (cid:21) .
(cid:20)p − p0
0
0
1 − p(cid:21) a(cid:20)p − p0
0
0
1 − p(cid:21) = a.
Now, it follows that 1 ≤ r(a) ≤ 1 − p0 so that p0 = 0. Thus r(a11) = p. Similarly,
r(a22) = 1 − p. Since n(x) = 1 − r(x) for any x ∈ [0, 1]M , we have n(a11) = 0 and
n(a22) = 0.
Next, as a is strict, so is 1 − a = (cid:20)p − a11
−a∗
12
−a12
1 − p − a22(cid:21). Thus, as above,
r(p − a11) = p and r(1 − p − a22) = 1 − p. Therefore, s(a11) = 0 and s(a22) = 0.
Hence a11 ∈ [0, p] and a22 ∈ [0, 1 − p] are strict in the von Neumann algebras pM p
and (1 − p)M (1 − p) respectively.
(cid:3)
Lemma 2.2. Let M be a von Neumann algebra with the underlying Hilbert space H
and let a ∈ [0, 1]M be strict. If ax = 0 for some x ∈ M , then x = 0. In particular,
if aξ = 0 for some ξ ∈ H, then ξ = 0.
4
ANIL KUMAR KARN
Proof. Let ax = 0. Then ax∗2 = axx∗ = 0 so that ax∗ = 0 and consequently,
ar(x∗) = 0. Thus r(x∗) ≤ n(a) = 0 so that r(x∗) = 0. Now, it follows that
x = 0.
Next, if aξ = 0, then ap = 0 where p is the projection of H on the span of ξ.
(cid:3)
Now, by the first step, p = 0 so that ξ = 0.
Lemma 2.3. Let M be a von Neumann algebra with the underlying Hilbert space
H and let a, b ∈ [0, 1]M be strict.
1
(1) Then a
(2) If ab = ba, then ab is also strict in M .
2 is also strict in M .
Proof.
(1) Let p = s(a
1
2 ). Then p = pa
1
2 = a
1
2 p. Squaring, we get p = pa = ap
2 ) = 0. Also,
1
so that p ≤ s(a) = 0. Thus p = 0 and consequently, s(a
n(a
2 ) = n(a) = 0. Therefore, a
2 is strict.
1
1
(2) Let ab = ba. Then ab ∈ [0, 1]M with ab ≤ a. Thus s(ab) ≤ ab ≤ a so that
s(ab) ≤ s(a) = 0. Therefore, s(ab) = 0. Next, we have abn(ab) = 0. As a
and b are strict, a repeated use of Lemma 2.2 yields that n(ab) = 0. Thus
ab is also strict.
(cid:3)
Remark 2.4. Let a ∈ [0, 1]M . Then a is strict if and only if a2 is strict,
The following results are a compilation of operator algebra folklore which can
easily be found in the literature. (See, for example, [1, Section I.5.2].)
Folklore 2.5. Let M be a von Neumann algebra with the underlying Hilbert space
H and let x ∈ M . If x = ux is the polar decomposition of x, then
(1) u : (1 − r(x))H → r(x)H is an unitary. In particular, u is a partial
isometry in M .
(2) x = u∗x.
(3) r(x) = u∗u.
(4) x∗k = uxku∗ for all k ∈ N.
(5) xk = u∗x∗ku for all k ∈ N.
Now we prove the main results of the paper.
Proof of Theorem 1.2.
(1) Put r(a ◦ b) = p1 so that p = 1 − p1. Now, by
Theorem 1.1, a and b have matrix representations
a = (cid:20)a11 a12
b22 (cid:21)
12 a22(cid:21) and b = (cid:20) b11 −a12
−a∗
12
a∗
12a12 = a22b22; and
12 = (p1 − a11)(p1 − b11);
with respect to {p1, p} such that
(a) a12a∗
(b) a∗
(c) a12 = a11a12 + a12a22 = b11a12 + a12b22.
Since a and b are strict, we have p1 6= 0, p1 6= 1. Also, by Lemma 2.1, a11
and b11 are strict in p1M p1 and a22 and b22 are strict in pM p. Consider
the polar decomposition
a12 = ua12
ABSOLUTE COMPATIBILITY
5
so that U : u∗uH → uu∗H is a unitary. By (a) and Folklore 2.5(4), we have
(p1 − a11)(p1 − b11) = a∗
122 = ua122
= u(a22b22)u∗ ≤ uu∗ ≤ p1.
Now, as a11 and b11 are strict elements in [0, p1] with a11b11 = b11a11, by
Lemma 2.3(2), we may conclude that (p1 − a11)(p1 − b11) is also a strict
element in [0, p1]. Thus r((p1 −a11)(p1 −b11)) = p1. Thus uu∗ = p1 = 1−p.
In a similar way, by using (b) and Folklore 2.5(5), we may prove that
u∗u = p. Thus u : pH → (1 − p)H is a unitary.
(2) By (b), we have a22b22 = a122 so that a22, b22 and a12 commute with
each other. Thus by (c), we get
ua12 = a12 = a11a12 + a12a22
= a11ua12 + ua12a22
= a11ua12 + ua22a12.
In other words,
a12(a11u + ua22 − u)∗(a11u + ua22 − u) = 0.
By Lemma 2.3, a12 is strict in pM p so that, by Lemma 2.2, a11u+ua22 = u.
Therefore,
p1 = uu∗ = a11uu∗ + ua22u∗ = a11 + ua22u∗
whence ua22u∗ = p1 − a11. In the same way we can show that ub22u∗ =
p1 − b11. Now, put a22 = a1, b22 = b1 and U = (cid:20) 0
u∗
p
0(cid:21). Then a1 and b1
are strict elements in [0, p] and U : H → K ⊕ K is a unitary. Also, the
matrix multiplications yield that
U ∗(cid:20)
a1
(a1b1)
1
2
1
2
(a1b1)
p − a1 (cid:21) U = a and U ∗(cid:20)
b1
−(a1b1)
1
2
1
2
−(a1b1)
p − b1 (cid:21) U = b.
Finally, we show that p − a1 − b1 is a strict element of [0, p]. Note that
a ◦ b = (cid:20)a11 + b11 − p1
0
0
0(cid:21). Thus r(a11 + b11 − p1) = r(a ◦ b) = p1. Next,
a11 + b11 − p1 = p1 − ua1u∗ + p1 − ub1u∗ − p1
= u(p − a1 − b1)u∗
for upu∗ = uu∗uu∗ = p2
1 = p1. Also, then
u∗(a11 + b11 − p1)ku = (p − a1 − b1)k
for all k ∈ N. Since for any x ∈ M +, r(x) is limit of the increasing sequence
{( 1
n + x)−1x} in the strong operator topology, we may conclude that
r(p − a1 − b1) = u∗r(a1 + b11 − p1)u = u∗p1u = p.
Thus n(p − a1 − b1) = 0. Also
s(p − a1 − b1) ≤ p − a1 − b1 ≤ p − a1
so that
s(p − a1 − b1) ≤ s(p − a1) = 0
for p − a1 is strict. Therefore, p − a1 − b1 is a strict element in [0, p].
(cid:3)
6
ANIL KUMAR KARN
Proof of Theorem 1.3. Let ξ, η ∈ H. Then
habη, ξi2 = hbη, aξi2 ≤ hb2η, ηiha2ξ, ξi
≤ h(1 − a2)η, ηiha2ξ, ξi.
Thus we obtain that
(cid:28)(cid:20)a2
ab
ab
1 − a2(cid:21)(cid:20)ξ
η(cid:21)(cid:29) = ha2ξ, ξi + habη, ξi + habξ, ηi + h(1 − a2)η, ηi
Therefore, a1 ≥ 0. Similarly, we can also show that b1 ≥ 0. Again,
≥ (cid:16)ha2ξ, ξi
1
2 − h(1 − a2)η, ηi
≥ 0.
1
2(cid:17)2
1M2(M) − a1 = (cid:20)1 − a2 −ab
a2 (cid:21) and 1M2(M) − b1 = (cid:20)1 − b2 ab
b2(cid:21)
−ab
ab
so that a1, b1 ∈ [0, 1]M2(M).
Now, we show that a1 is strict. Let p be a projection in M2(M ) such that p ≤ a1.
Then p = a1p. For (cid:20)ξ
η(cid:21) ∈ p(H ⊕ H), we have a1(cid:20)ξ
η(cid:21) = (cid:20)ξ
η(cid:21) so that
ξ = a2ξ + abη
and η = abξ + (1 − a2)η.
By the second condition, we get a2η = abξ so that a(aη − bξ) = 0. Since a is strict,
by Lemma 2.2, we may conclude that aη = bξ. Using this in the first relation, we
get ξ = a2ξ + b2ξ so that (1 − a2 − b2)ξ = 0. Since 1 − a2 − b2 is also strict, invoking
Lemma 2.2 again we conclude that ξ = 0. But then, aη = 0 whence η = 0 as a is
also strict. Now, it follows that p = 0 so that s(a1) = 0.
Next, let q be a projection in M2(M ) such that a1q = 0. If (cid:20)ξ
have a1(cid:20)ξ
η(cid:21) = 0. Thus
η(cid:21) ∈ q(H ⊕ H), we
a2ξ + abη = 0 and abξ + (1 − a2)η = 0
and, as above we again conclude that q = 0. Therefore, n(a1) = 0 too whence a1 is
strict. In a similar manner, we can conclude that b1 is also strict. Finally, we show
that a1 is absolutely compatible with b1. We have
= (cid:20)(a2 + b2)2
0
0
(a2 + b2)2(cid:21)
so that a1 − b1 = (cid:20)a2 + b2
0
2ab
2ab
(a1 − b1)2 = (cid:20)a2 − b2
b2 − a2(cid:21)2
a2 + b2(cid:21). Also
1M2(M) − a1 − b1 = (cid:12)(cid:12)(cid:12)(cid:12)
0
(cid:20)a − a2 − b2
= (cid:20)1 − a2 − b2
0
0
0
a2 + b2 − 1(cid:21)(cid:12)(cid:12)(cid:12)(cid:12)
1 − a2 − b2(cid:21) .
0
It follows that a1 − b1 + 1M2(M) − a1 − b1 = 1M2(M) so that a1 is absolutely
compatible with b1.
(cid:3)
Acknowledgements: The author is thankful to Antonio M. Peralta for intro-
ducing to him the notion of 'strict' elements.
ABSOLUTE COMPATIBILITY
7
References
[1] B. Blackadar, Operator algebras, Springer-Verlag Heidelberg Berling, 2006.
[2] P. R. Halmos, Two subspaces, Trans. Amer. math. Soc., 144(1969), 181-189.
[3] N. K. Jana, A. K. Karn and A. M. Peralta, Absolutely compatible pairs in a von Neumann
algebra, (https://arxiv.org/abs/1801.01216), (Communicated for publication).
[4] N. K. Jana, A. K. Karn and A. M. Peralta, Contractive linear preservers of absolutely
compatible pairs between C∗-algebras, Revista de la Real Academia de Ciencias Exactas,
Fsicas y Naturales. Serie A. Matemticas (RCSM) 113(2019) no. 3, 2731-2741.
[5] A. K. Karn, Orthogonality in lp-spaces and its bearing on ordered Banach spaces, Positivity,
18(02) (2014), 223-234.
[6] A. K. Karn, Orthogonality in C∗-algebras, Positivity, 20(03) (2016), 607-620.
[7] A. K. Karn, Algebraic orthogonality and commuting projections in operator algebras, Acta
Sci. Math. (Szeged), 84(2018), 323-353.
School of Mathematical Sciences, National Institute of Science Education and Re-
search, HBNI, Bhubaneswar, P.O. - Jatni, District - Khurda, Odisha - 752050, India.
E-mail address: [email protected]
|
1904.06431 | 5 | 1904 | 2019-07-13T15:03:16 | Invariant subspaces of generalized Hardy algebras associated with compact abelian group actions on W*-algebras | [
"math.OA"
] | We consider an action of a compact group whose dual is archimedean linearly ordered or a direct product (or sum) of such groups on a von Neumann algebra, M. We define the generalized Hardy subspace of the Hilbert space of a standard representation the algebra, and the Hardy subalgebra of analytic elements of M with respect to the action. We find conditions in order that the Hardy algebra is a hereditarily reflexive algebra of operators. In particular if every non zero spectral subspace, contains a unitary operator, the condition is satisfied and therefore the Hardy algebra is hereditarily reflexive. This is the case if the action is the dual action on a crossed product, or an ergodic action, or, if, in some situations, the fixed point algebra is a factor. | math.OA | math |
Invariant subspaces of generalized Hardy alge-
bras associated with compact abelian group ac-
tions on W*-algebras
Costel Peligrad
Department of Mathematical Sciences, University of Cincinnati, PO Box
210025, Cincinnati, OH 45221-0025, USA. E-mail address: [email protected]
Key words and phrases. W*-dynamical system, invariant subspaces, analytic
elements, generalized Hardy algebra, reflexive algebra.
2013 Mathematics Subject Classification. Primary 46L10, 46L40, 47L75;
Secondary 30H10, 47B35.
ABSTRACT. We consider an action of a compact abelian group whose dual
is any subgroup of the additive group of real numbers (so, an archimedean
linearly ordered group) or a direct product (or sum) of such groups on a W*-
algebra, M . We define the generalized Hardy subspace of the Hilbert space of
a standard representation the algebra, and the Hardy subalgebra of analytic
elements of M with respect to the action. We find conditions in order that the
Hardy algebra is a hereditarily reflexive algebra of operators. In particular if
every non zero spectral subspace, contains a unitary operator, the condition is
satisfied and therefore the Hardy algebra is hereditarily reflexive. This is the
case if the action is the dual action on a crossed product, or an ergodic action,
or, if, in some situations, the fixed point algebra is a factor.
1 Introduction
This paper is concerned with the study of invariant subspaces and reflexivity
of operator algebras associated with compact group actions on W*-algebras.
Recall first the definition of a reflexive operator algebra.
Let A ⊂ B(X) be a weakly closed algebra of operators on a Banach space
X. Denote by Lat(A) the lattice of closed subspaces of X that are invariant for
all operators a ∈ A. Let
algLat(A) = {b ∈ B(X) : bK ⊂ K for all K ∈ Lat(A)} .
The algebra A is called reflexive if A = algLat(A). Hence, a reflexive operator
algebra is completely determined by the lattice of its invariant subspaces. An
algebra A ⊂ B(X) is called hereditarily reflexive if every unital weakly closed
subalgebra of A is reflexive. Sarason [19], proved two results: (1) every com-
mutative von Neumann algebra is hereditarily reflexive and (2) the algebra of
1
analytic Toeplitz operators on the Hardy space H 2(T ) where T is the the unit
circle T = {z ∈ C : z = 1} , is hereditarily reflexive. In [14] we extended this
result in two directions: (1) to the case of H p(T ), 1 < p < ∞ and (2) to the
not necessarily commutative case of non selfadjoint crossed products of finite
von Neumann algebras by the semigroup Z+. Later, in [9], Kakariadis has con-
sidered the more general case of reduced w*-semicrossed products and, among
other results, he has extended the particular case of our reflexivity result in [14,
Proposition 4.5,] for p = 2 to the semicrossed product setting [9, 2.10.H.]. This
result was considered later by Helmer [5] in the context of W*-correspondences
[12]. Further, in [15], we studied a related problem in a more general setting
than the crossed product or the reduced w∗semicrossed product considered in
[14] and [9, 2.9] for the case of von Neumann algebras. We considered a W*-
dynamical system (M, T,α) where T = {z ∈ C : z = 1} is the circle group and
M is a σ−finite W*-algebra. We constructed a standard covariant representa-
tion of the system on a certain Hilbert space, H, a generalized Hardy space,
H+ and the corresponding Hardy algebra M+ ⊂ B(H+). We have shown that
if the spectral subspace corresponding to the smallest positive element of the
spectrum contains a unitary element, then, the algebra M+ is reflexive. Actu-
ally, [15, Theorem 3.5.] shows that if M ⊂ B(H) is a σ-finite von Neumann
algebra in its standard representation such that each spectral subspace contains
a unitary operator (as is, in particular, the algebra of analytic Toeplitz opera-
tors considered by Sarason), then, M+ ⊂ B(H+) is a reflexive operator algebra.
Recently Bickerton and Kakariadis [2] have obtained results about reflexivity of
algebras associated with actions of Zd
+ (the direct product of d copies of Z+).
In this paper we make two significant steps towards solving the reflexiv-
ity problem of Hardy algebras associated to one-parameter dynamical systems
(M, R,α): 1. We consider a W*-dynamical system (M, G, α) where M is a von
Neumann algebra in standard form, and 2. G is a compact abelian group whose
dual is an arbitrary subgroup of R (possibly R itself with the discrete topology),
so, a discrete group with a linear archimedean order. We also consider actions of
compact abelian groups, G, whose duals, Γ, are direct products or direct sums
of discrete groups with linearly archimedean order and consider the lattice order
on Γ (see for instance [3]). In Section 2.1. we define a standard covariant repre-
sentation of the system that will be the framework for the rest of the paper. In
the Corollary to Proposition 2.4. we show that every von Neumann algebra in
standard form (in particular every maximal abelian von Neumann algebra) is
hereditarily reflexive, thus extending the first result of Sarason mentioned above
to every von Neumann algebra in its standard representation. In Section 3 we
consider the case when the dual Γ of G has an archimedean linear order, or is
a direct product of such groups, we define a generalized Hardy space H+ ⊂ H
and a corresponding Hardy algebra M+ ⊂ B(H+), where H is.the Hilbert space
of the standard covariant representation of the system (M, G, α) and we prove
that, in some conditions, including the conditions in [15], M+ ⊂ B(H+) is
hereditarily reflexive (in [15] we proved only reflexivity for the particular case
when Γ = Z).We do not assume as in [15] that M is σ-finite. Also, if Γ is an
arbitrary archimedean linearly ordered discrete group, it can be any subgroup
2
of R with the discrete topology, not only Z as in [15], [9], [14]. Examples include
the Hardy algebra of analytic Toeplitz operators, H ∞(T), the results in [15],
w∗-crossed products by abelian archimedean ordered discrete groups or a direct
product of such groups, some reduced w∗-semicrossed products considered in
[9], [2] and other situations as stated in the Corollaries 3.14., 3.15., 3.16. and
3.17.
2 Preliminary results and notations
2.1 Standard representations of W*-algebras
In this section we review some concepts and results related to the standard
representation of a von Neumann algebra. Some of these results are certainly
known, but we did not find an exact reference for them. We provide proofs
of these results for the convenience of the reader. In Proposition 2.4. and its
Corollary we prove that every von Neumann algebra in its standard representa-
tion is hereditarily reflexive. In particular, every abelian von Neumann algebra
is hereditarily reflexive ([19, Theorem 1]).
Let M be a W*-algebra and let ρ be a weight on the positive part, M +, of
M, that is a mapping ρ : M + → [0, ∞) ∪ {∞} such that
ρ(m + n) = ρ(m) + ρ(n), m, n ∈ M +
and
ρ(λm) = λρ(m), m ∈ M +, λ ∈ R, λ > 0
with the convention 0 · ∞ = 0. As it is customary ([8], [20]), denote
Nρ = {m ∈ M : ρ(m∗m) < ∞} .
Nρ = {m ∈ M : ρ(m∗m) = 0} .
Fρ =(cid:8)m ∈ M + : ρ(m) < ∞(cid:9) .
Mρ = linear span of Fρ.
It is immediate that Nρ is a left ideal of M . The weight ρ is called faithful if
Nρ = {0} , normal if it is the sum of a family {ϕι} of positive normal linear
functionals and semifinite if Mρ or, equivalently [20, 2.1.], Nρ is w*- dense in
M.
Now let M be a W*-algebra, M0 ⊂ M a W*-subalgebra, and P0 : M → M0 a
w*-continuous projection of norm 1 of M onto M0 which is, in addition, faithful
on the set of positive elements of M. Let ρ0 be a faithful normal semifinite
weight on M . It is known that such a weight exists. Indeed, consider a family
{ϕι} of positive normal linear functionals of M0 such that their supports {pι}
3
form a maximal family of mutually orthogonal projections of M0, in particular
P pι = I. Then ρ0 =P ϕι is a faithful normal semifinite weight of M0.
The following fact is stated in [19, Corollary 10.5] as a consequence of a
theorem of Takesaki [20, Theorem 10.1.]. We present a short proof of this fact
in our setting for the convenience of the reader.
Lemma 2.1. ρ = ρ0 ◦ P0 is a faithful normal semifinite weight on M.
Proof. Since ρ0 and P0 are faithful, it follows that ρ is faithful. Since ρ0
is normal and P0 is w*-continuous, it follows that ρ is normal. To prove that ρ
is semifinite, notice that from the definition of Nρ we have that M Nρ0 ⊂ Nρ,
where M Nρ0 denotes the linear span of {mn : m ∈ M, n ∈ Nρ0 } . Since Nρ0 is
w*-dense in M0 it follows that M Nρ0 and therefore Nρ is dense in M so ρ is
semifinite.
By [8, Theorem 7.5.3.], there exists a faithful normal representation πρ of
M on the completion Hρ of Nρ ⊂ M with respect to the inner product
hm, ni = ρ(n∗m).
This representation is uniquely determined up to unitary equivalence and is, in
that sense, independent of the choice of the weight ρ0. We will use the version
of Tomita-Takesaki Theorem from [8, Theorem 9.2.37.]. If S is the conjugate
ρ by S(n) = n∗, then S is a preclosed densely
linear operator defined on Nρ ∩ N ∗
2 , in
defined operator on Hρ and its closure has the polar decomposition J∆
which ∆ is an invertible positive operator and J is a conjugate linear isometry
acting on Hρ such that J 2 = I and J πρ(M )J = πρ(M )
is the
commutant of πρ(M ) in B(Hρ). We have Nρ0 ⊂ Nρ, where, as above
where πρ(M )
1
′
′
and
Nρ0 = {m ∈ M0 : ρ0(m∗m) < ∞} .
Nρ = {m ∈ M : ρ(m∗m) < ∞} .
In the rest of the paper if ρ is a faithful, normal semifinite weight on M +
we will identify πρ(M ) with M and will write m instead of πρ(m), m ∈ M. We
will call this representation the standard representation of M and we will refer
to the inclusion M ⊂ B(Hρ) as the standard form of M. Also, we will denote
Hρ by H and the closure of Nρ0 in H by H0.
Lemma 2.2. The restriction of P0 to Nρ extends to the orthogonal projec-
tion of H onto H0.
4
Proof. Clearly, P0(Nρ0 ) = Nρ0 . We will prove next that P0(Nρ) ⊂ Nρ0 .
Indeed, let n ∈ Nρ. Then n = P0(n) + (I − P0)(n) = n0 + n1. Since n ∈ Nρ, we
have ρ(n∗n) < ∞, so
ρ(n∗
0n0 + n∗
0n1 + n∗
1n0 + n∗
1n1) = ρ0(P0(n∗
0n0 + n∗
0n1 + n∗
1n0 + n∗
1n1)) =
ρ0(n∗
0n0 + n∗
0P0(n1) + P0(n∗
1)n0 + P0(n∗
1n1)) =
ρ0(n∗
0n0 + P0(n∗
1n1)) < ∞.
Therefore, ρ0(n∗
0n0) < ∞ and we are done. On the other hand,
hP0(n), mi = hn, P0(m)i .
since both of the above terms equal ρ0(P0(m∗)P0(n)), so P0 is.self adjoint.
Lemma 2.3. i) M H0 is dense in H and H0 is a separating set for M that
is, if m ∈ M is such that mξ0 = 0 for all ξ0 ∈ H0, then m = 0. Here, M H0
denotes the linear span of {mξ : m ∈ M, ξ ∈ H0} .
ii) M ′H0 is dense in H and H0 is a separating set for M ′ that is, if m′ ∈ M ′
is such that m′ξ0 = 0 for all ξ0 ∈ H0, then m′ = 0.
Proof.
i) We will prove that M H0 is dense in Nρ ⊂ H and, since Nρ is
dense in H, the first part of i) will follow. Let x ∈ Nρ. Therefore, ρ(x∗x) =
ρ0(P0(x∗x)) < ∞. Since ρ0 is a faithful normal semifinite weight on M0 we can
assume that ρ0 is the sum of a family of normal positive linear functionals {ϕι}
on M0 such that their suports {pι} form a maximal family of mutually ortog-
onal projections in M0 and P pι = I. Since ρ0(P0(x∗x)) = P ϕι(P0(x∗x)) <
P∞
i=1 ϕi(P0(x∗x)) < ∞. Let qn = Pi=n
∞, it follows that the summable family of positive numbers {ϕι(P0(x∗x))}ι
is at most countable, say {ϕi(P0(x∗x))}i∈N with ρ(x∗x) = ρ0(P0(x∗x)) =
i=1 pi, where, for each i ∈ N, pi is the
supportof ϕi. Then, qn ∈ Nρ0 ⊂ H0 and xqn ∈ M Nρ0 ⊂ M H0 for every n ∈ N.
Clearly, since pi is the support of ϕi we have ϕi(y) = ϕi(piy) = ϕi(ypi) for
all i ∈ N, y ∈ M0 and ϕi(qny) = ϕi(y) for every i, n ∈ N with i 6 n, y ∈ M0
and ϕi(qnx) = 0 if i > n We will show that limn→∞ xqn = x in H.Let ǫ > 0.
i=N +1 ϕi(P0(x∗x)) < ǫ2. Therefore,
Then, there exists N = Nǫ > 0 such thatP∞
if n > N
hqnx − x, qnx − xi = ρ((qnx − x)∗(qnx − x)) =
ϕi(P0((qnx − x)∗(qnx − x)))+
nXi=1
+
∞Xi=n+1
ϕi(P0((qnx − x)∗(qnx − x))
Using the preceding observations, if i 6 n, we get
ϕi(P0((qnx − x)∗(qnx − x))) =
5
ϕi(qnP0(x∗x)qn − qnP0(x∗x) − P0(x∗x)qn + P0(x∗x)) = 0
and, if i > n > N
ϕi(P0((qnx − x)∗(qnx − x) = ϕi(P0(x∗x))
Hence hqnx − x, qnx − xi < ǫ2 and the claim is proven. To prove the second
part, let m ∈ M such that mξ0 = 0 for every ξ0 ∈ H0, in particular, mn = 0 for
every n ∈ Nρ0 ⊂ H0. Since Nρ0 is w*-dense in M0 ⊂ M and I ∈ M0, it follows
that m = 0 and part i) is proven.
ii). Denote by K the closure of M ′H0. Then K is a closed subspace of H
which is invariant for every m′ ∈ M ′, so, the orthogonal projection, p, of H
on K comutes with M ′, and therefore p ∈ M. Since H0 ⊂ K, it follows that
(1 − p)H0 = {0} . Since by i) H0 is a separating set for M, we have 1 − p = 0
and thus K = H.To prove the second part of ii), let m′ ∈ M ′ be such that
m′H0 = {0} . It follows that M m′H0 = {0} , so m′M H0 = {0} . Since, by i)
M H0 is dense in H, it follows that m′ = 0.
Some of the statements in the next Proposition are probably known, but we
did not find a reference for any of them.
Proposition 2.4. i) Let M ⊂ B(H) be a von Neumann algebra in standard
form. Then, every normal linear functional, ϕ, on M is a vector functional,
that is, there exist ξ, η ∈ H such that ϕ(m) = hmξ, ηi , m ∈ M.
ii) If N ⊂ B(H) is an abelian von Neumann algebra, not necessarily in
standard form, then every normal linear functional on N is a vector functional.
iii) If M0 ⊂ B(H0) is a maximal abelian von Neumann algebra, then it is
spatially isomorphic with its standard form.
Proof. i) Let p ∈ M be a countably decomposable projection. According
to [8, 9.6.18.], the hypotheses of [8, 9.6.20.] are satisfied, so there exists ξ0 ∈ H
such that J ξ0 = ξ0 and M ′ξ0 = pH. Therefore, every countably decomposable
projection p ∈ M is a cyclic projection. Now, let ϕ be a normal linear functional
on M. By the polar decomposition of normal linear functionals [8, Theorem
7.3.2.], it is enough to prove the statement for normal positive linear functionals.
Let ψ be a normal positive functional on M and p its support (that is, p is the
complement of the supremum of all projections q ∈ M for which ψ(q) = 0).
Then, p is countably decomposable, so by the previous arguments, p is a cyclic
projection. Applying [8, Proposition 7.2.7.] it follows that ψ is a vector normal
positive functional.
ii) Let ϕ be a normal linear functional on M. As argued in i), using the polar
decomposition of normal linear functionals it is enough to prove the statement in
ii) for normal positive functionals. Let ψ be a normal positive functional on M
and p its support which is a countably decomposable projection. Without loss of
generality we can assume that p = I. We will show that there exists a separating
6
vector, ξ0 ∈ H for N and therefore, cyclic for N ′. Let {ξi ∈ H : i ∈ A} be a maxi-
mal family of orthogonal unit vectors such that, the projections {pi ∈ N : i ∈ A}
onto(cid:8)Hi = N ′ξi ⊂ H : i ∈ A(cid:9) are mutually orthogonal, so H =P⊕ Hi. Since
Suppose A ⊆ N. Let ξ0 =Pi∈A
p = I is countably decomposable, it follows that the set A is at most countable.
2i ξi. and let m ∈ N + be such that mξ0 = 0.
Hence hmξ0, ξii = 0, i ∈ A. Then, since N is abelian, so N ⊆ N ′, it follows that
hmξi, ξii = 0 for every i. Therefore, since m ∈ N +, it follows that mξi = 0,
so mHi = {0} for every i, and thus m = 0. Hence ξ0 is separating for N. The
statement ii) follows from [8, 7.2.7.].
1
iii) Let {pι} be a maximal family of mutually orthogonal countably decom-
posable projections of M0 and {ϕι} a family of positive linear functionals such
that the support of ϕι is pι. ClearlyP pι = I. Let ρ0 =P ϕι. Then ρ0 is a faith-
ful normal semifinite weight on M +
0 . By ii) for every ι there exists ξι ∈ pιH0 such
that ϕι(m) = hmξι, ξιi , m ∈ M0. Obviously, ξι is a cyclic and separating vector
of M pιpιH0 for every ι.It is also clear that hm1ξι, m2ξιi = ϕι(m∗
2m1), m1, m2 ∈
M0, for every ι. If Nρ0 is as above,
Nρ0 = {m ∈ M0 : ρ0(m∗m) < ∞} ,
then the mapping mpι → mξι extends to a unitary operator from Hρ0 to H0
and we are done.
Corollary i) Every von Neumann algebra in standard form is hereditarily
reflexive.
ii) [19, Theorem 2] Every abelian von Neumann algebra, not necessarily in
standard form is hereditarily reflexive.
Proof. i) Follows from Proposition 2.4. i) and [10, Theorem 3.5.].
ii) Follows from Proposition 2.4. ii) and [10 Theorem 3.5.].
2.2 W*-dynamical systems with compact abelian groups
Let (M, G,α) be a W*-dynamical system, where M is a W ∗−algebra, G is a
compact abelian group with dual Γ, and α a faithful w∗−continuous action of
G on M, .that is αg 6= id if g 6= 0, where id is the identity automorphism of M
and the mapping g → ϕ(αg(m)) for every m ∈ M and every ϕ ∈ M∗, where M∗
denotes the predual of M. For each γ ∈ Γ, denote by
Mγ =(cid:26)Z hg, γiαg(m)dg : m ∈ M(cid:27) .
where the integral is taken in the w∗−topology. In particular, if γ = 0, M0 is
the fixed point algebra of the system. It can immediately be checked that
Mγ = {m ∈ M : αg(m) = hg, γi m}
7
It is clear that the mapping Pγ : M → Mγ defined by Pγ(m) =Z hg, γiαg(m)dg
is a w*-continuous projection of M onto the closed subspace Mγ ⊂ M. In
particular, P0 is a w*-continuous projection of M onto M0 which is clearly
faithful (on M +).
It is well known that M is the w∗-closed linear span of
{Mγ : γ ∈ Γ} . The Arveson spectrum of the action α is, by definition ([1], [13])
sp(α) = {γ ∈ Γ : Mγ 6= {0}} .
Lemma 2.5. i) M−γ = M ∗
ii) Mγ1Mγ2 ⊂ Mγ1+γ2 where Mγ1Mγ2 is the linear span of {xy : x ∈ Mγ1 , y ∈ Mγ2} .
iii) If m ∈ Mγ has polar decomposition m = u x , then u ∈ Mγ and
γ = {m∗ : m ∈ Mγ} M ∗
γ = {m∗ : m ∈ Mγ}.
γ ,where M ∗
x ∈ M0.
Proof. i) and ii) are obvious.
iii) is a straightforward consequence of the
uniqueness of the polar decomposition of m.
Let (M, G, α) be as above, ρ0 a faithful normal semifinite weight on M0 and
ρ = ρ0 ◦ P0. Consider the corresponding normal faithful representation πρ on
Hρ and the Tomita-Takesaki operators S, J as in 2.1. above. As in 2.1. we will
write H instead of Hρ and M instead of πρ(M ). In the case when Γ is a partially
ordered group, this representation will allow us to construct a generalized Hardy
space on which, in certain situations, the subalgebra of analytic elements of the
system (M, G,α) is hereditarily reflexive.
For every g ∈ G define the unitary operator Ug ∈ B(H) as the unique
extension of Ug(n) = αg(n), n ∈ Nρ to H. Then, since clearly, Nρ is an α-
invariant left ideal of M, it is straightforward to check that the group of unitary
operators {Ug : g ∈ G} implements the action α. Also, from the definition of S it
follows that SUg = UgS and S ∗Ug = UgS ∗ for all g ∈ G. Therefore, J Ug = UgJ,
g ∈ G. It follows that the group {Ug : g ∈ G} implements an action α′ of G on
M
, namely
′
α′
g(J mJ) = UgJ mJ U ∗
g = J αg(m)J.
Similarly with the projections Pγ of M onto Mγ, γ ∈ Γ one can define the
projections P ′
γ of M ′ onto M ′
γ =(cid:8)x ∈ M ′ : α′
γ(x) =Z hg, γiα′
P ′
g(x) = hg, γi x(cid:9)
g(x)dg.
The proof of the following lemma is a straghtforward application of the
definitions.
Lemma 2.6. With the notations above, we have the following:
i) α′
g is an action of G on M ′, where M ′ is the commutant
g(m′) = Ugm′U ∗
of M in B(H).
8
ii) If g ∈ G, then Ug commutes with J, and J αg(m)J = α′
g(J mJ), m ∈
M, g ∈ G.
iii) ( M ′)γ = JM−γJ, γ ∈ Γ.
iv) Ug(m′ξ) = α′
v) sp(α) = sp(α′).
g(m′)ξ, ξ ∈ H0, m′ ∈ M ′.
We will need also the following
Remark 2.7.
If M is a finite W*-algebra, then M ′ is a finite W*-
algebra.This fact is immediate from the definition of the standard representa-
tions.
Let Hγ =(cid:26)Z hg, γiUg(ξ)dg : ξ ∈ H(cid:27) = {ξ ∈ H : Ugξ = hg, γi ξ} . Then, the
map P H
γ from H to Hγ defined as follows
P H
γ (ξ) =Z hg, γiUg(ξ)dg : γ ∈ Γ, ξ ∈ H.
is an orthogonal projection of H onto the closed supspace Hγ . Applying Lemma
2.2., we see that if γ = 0, the Hilbert subspace H0 ⊂ H coincides with the
Hilbert subspace H0 considered in Section 2.1.
Lemma 2.8. i) If γ1 6= γ2, then Hγ1 and Hγ2 are orthogonal.
ii) For every γ ∈ sp(α),we have MγH0 = Hγ , where
MγH0 = {mξ0 : m ∈ Mγ, ξ0 ∈ H0} .
iii) The direct sum of Hilbert spaces XHγ equals H.
iv) For every γ ∈ sp(α) we have (M ′ )γ H0 = Hγ , where
( M
v) For all γ, γ ′ ∈ sp(α) we have MγHγ ′ ⊂ Hγ+γ ′ and M ′
)γ , ξ0 ∈ H0}.
)γ H0 = {m
ξ0 : m
′
′
′
∈ (M
′
γHγ ′ ⊂ Hγ+γ ′.
Proof. i) Let ξ ∈ Hγ1 , η ∈ Hγ2 . Then, by definition, Ug(ξ) = hg, γ1i ξ and
Ug(η) = hg, γ2i η, for all g ∈ G. Since the operators Ug are unitary, we have
hξ, ηi = hUgξ, Ugηi = hg, γ1 − γ2i hξ, ηi , g ∈ G.
Hence, if γ1 6= γ2 it follows that hξ, ηi = 0.
ii) Since, by Lemma 2.3. i), H0 is a cyclic set for M, the subspace M H0 =
γ and Pγ are the above pro-
{mξ0 : m ∈ M, ξ0 ∈ H0} is dense in H. Then, if P H
jections, we have
MγH0 = Pγ(M )H0 =(cid:26)Z hg, γiαg(m)ξ0dg : m ∈ M, ξ0 ∈ H0(cid:27) =
=(cid:26)Z hg, γiUg(ξ)dg : ξ = mξ0, m ∈ M, ξ0 ∈ H0(cid:27) =
.
9
=(cid:8)P H
γ (ξ) : ξ = mξ0, m ∈ M, ξ0 ∈ H0(cid:9) .
Since by Lemma 2.3. i) the subspace M H0 is dense in H and P H
projection, the result stated in ii) follows.
γ is an orthogonal
iii) Let η ∈ H be such that η ⊥ Hγ for all γ ∈ Γ. Since M is the w∗-closed
linear span of {Mγ : γ ∈ Γ} , it follows from ii) that η ⊥ M H0, so, since by
Lemma 2.3. i) H0 is cyclic for M, it follows that η = 0.
iv) The proof is similar with that of ii) taking into account that, according
to Lemma 2.3. ii), H0 is cyclic for M
′
as well.
v) Immediate from definitions.
3 Hereditary reflexivity of generalized Hardy al-
gebras
In this section we will construct the generalized Hardy space and the generalized
Hardy algebra and prove the main results of this paper, Theorem 3.8. and
Theorem 3.9.
Let (M, G, α) be a W*-dynamical system with G compact abelian. Through-
out this section we will assyme that M ⊂ B(H) where H is the Hilbert space
constructed in Section 2.1. Suppose, in addition, that Γ is an archimedean lin-
early ordered discrete group, or Γ ⊆ Πι∈I Γι is the direct product or the direct
sum of archimedean linearly ordered (discrete) groups Γι. If (Γι)+ is the semi-
group of non negative elements of Γι, denote by Γ+ = Πι∈I (Γι)+ . Then, Γ+ is
a sub semigroup of Γ such that
and
Γ+ ∩ (−Γ+) = {0}
Γ+ − Γ+ = Γ
so Γ+ defines a partial order on Γ, namely γ1 6 γ2 if γ2 − γ1 ∈ Γ+.
Lemma 3.1. Let γ1, γ2 ∈ Γ+, γ1 6= γ2. Then, either
i) γ1, γ2 are not comparable under the above order relation, or,
ii) γ1 < γ2, or,
iii) γ1 > γ2 and, in this case, there exists p ∈ N such that either
iii a) γ1 > pγ2 and γ1 < (p + 1)γ2, or,
iii b) γ1 > pγ2 and γ1 and (p + 1)γ2 are not comparable.
Proof. Suppose that γ1, γ2 are comparable. Thus, either γ1 < γ2, or
γ1 > γ2. Suppose that γ1 > γ2. Since γ1, γ2 ∈ Πι∈I (Γι)+ , we can write
1 > γι0
γ1 = (γι
2
for some ι0 ∈ I. Since for every ι ∈ I , Γι has an archimedean order, there exists
a largest pι ∈ N such that γι
2. If L ⊂ N is the set of non repeating p′
ιs
2 ∈ (Γι)+ , so γι
2, ι ∈ I and γι0
1)ι∈I , γ2 = (γι
2)ι∈I with γι
1, γι
1 > γι
1 > pιγι
10
then, since N is well ordered, there exists p = min L. Therefore, γ1 > pγ2. By
the definition of p ∈ N, γ1 (cid:11) (p + 1)γ2, so either iii a) or iii b) must hold.
The following consequence of the above Lemma will be used
Corollary 3.2. Let γ0 ∈ Γ+ r {0} and γ ∈ Γ+.Then, there exists p ∈ Z+
such that either pγ0 6 γ < (p + 1)γ0 or pγ0 6 γ and γ is not comparable with
(p + 1)γ0.
Proof. If γ < γ0 or γ and γ0 are not comparable, then p = 0 satisfies the
conclusion. If γ > γ0, the statement follows from Lemma 3.1. iii).
Lemma 3.3.
If Γ ⊆ Πι∈IΓι is the direct product or the direct sum of
linearly ordered discrete groups Γι, then Γ is lattice ordered (see [3]), i.e.
if
A = {γ1, γ2, ...γn} is a finite subset of Γ then there exists inf A and sup A in Γ.
Proof.
sup A = (νι)ι∈I .
j(cid:1) , 1 6 j 6 n, let µι = min(cid:8)γι
j : j = 1, 2, ...n(cid:9) and
If γj = (cid:0)γι
j : j = 1, 2, ...n(cid:9) for each ι ∈ I. Then clearly inf A = (µι)ι∈I and
νι = max(cid:8)γι
If (M, G, α), M ⊂ B(H), H0, bG = Γ and Γ+ are as above, define
H+ = Xγ∈Γ+
Hγ
Let p+ be the orthogonal projection of H onto H+. By Lemma 2.8. v), the
(closed) subspace H+ ⊂ H is invariant for ∨γ>0Mγ, and for ∨γ>0(M ′)γ .We will
denote by M+ the weak operator closure
in B(H+) and similarly
M+ = p+(∨γ∈Γ+Mγ)p+
wo
(M ′)+ = p+(∨γ>0(M ′)γ)p+
wo
where ∨γ>0Mγ , is the algebra generated by {Mγ : γ > 0}) and ∨γ>0(M ′)γ is the
algebra generated by {(M ′)γ : γ > 0}). Then, we will call H+ the generalized
Hardy space and M+ the generalized Hardy algebra of analytic elements of the
dynamical system (M, G, α).
To prove hereditary reflexivity, we also need the following
Lemma 3.4. If ψ is a weakly continuous functional on M+ ⊂ B(H+), then
ψ is a vector functional (that is, there exist ξ, η ∈ H+ such that ψ(m) = hmξ, ηi
for all m ∈ M+).
Proof. Since ψ is weakly continuous, there exist n ∈ N and ξi, ηi ∈ H+, 1 6
i 6 n such that ψ(m+) =Pi hm+ξi, ηii , m+ ∈ M+. Now let eψ be the functional
11
Proposition 2.4. to this restriction, it follows that there exist ξ, η ∈ H such
on defined by eψ(b) =Pi hbξi, ηii , b ∈ B(H). Since ξi, ηi ∈ H+, it follows that
eψ(b) = eψ(p+bp+), b ∈ B(H), where, as above, p+ is the projection of H onto
H+. The restriction of eψ to M is a normal linear functional of M. Applying
that eψ(m) = hmξ, ηi , m ∈ M. Since, as noticed before, eψ(m) = eψ(p+mp+), we
can take ξ, η ∈ H+. Therefore, in particular, ψ(mγ) = eψ(mγ) = eψ(p+mp+) =
ψ(mγ) = hmγ ξ, ηi for every mγ ∈ Mγ, γ ∈ Γ+. The definition of M+ implies
that ψ(m+) = hm+ξ, ηi , m+ ∈ M+ and the proof is completed.
Loginov and Sul'man [10, Theorem 2.3.] have shown, in particular, that if
a reflexive algebra satisfies the hypothesis of Lemma 3.4. then it is hereditarily
reflexive, that is, all unital weakly closed subalgebras are reflexive. Under the
name of super reflexivity this fact has been also considered in [4, Proposition
2.5. (1)].
In Theorem 3.8. below we will assume that (M, G, α) satisfies the following
condition:
(C) For every γ ∈ sp(α) \ {0} there exists an element uγ ∈ Mγ such that
γ = eγ where eγ is a central projection of M and Mγ = M0uγ. For
u∗
γuγ = uγu∗
γ = 0 we will take u0 = I.
Examples of dynamical systems (M, G, α) satisfying this condition include
the following:
a) If M is a finite W*-algebra and the center, Z(M0), of M0 is contained in
the center, Z(M ), of M [18]. This is the case, in particular, when M is a finite
W*-algebra and M0 is a factor (this case will be discussed in a more general
context in part b)). The conditions M finite and Z(M0) ⊂ Z(M ) also hold if
M = ⊕Mi and (Mi, G, α) is a finite W*-algebra and the fixed point algebra is
a factor. Corollary 3.14. below will refer to these example.
b) If M is a semifinite injective von Neumann algebra such that M0 is a
factor, except when M is type III and M0 is a type II1 factor [21]. Thomsen
has proved that in these cases, the action α has full unitary spectrum, that is
every nonzero spectral subspace contains unitary operators. This is the case, in
particular, when M is a finite W*-algebra and M0 is a factor. In particular, this
latter situation occurs if α is a prime action of the compact abelian group G on
the hyperfinite type II1 factor [6], [7], in particular if α is egodic. Recall that an
action is called prime if the fixed point algebra is a factor. In particular if the
action α is faithful, then the all the examples in this part b) satisfy sp(α) = Γ.
Corollary 3.15. below will refer to these examples. Also, the Condition (C)
is satisfied if M is the crossed product of a von Neumann algebra M0. by an
abelian discrete group Γ. Corollaries 3.16. and 3.17. will consider this case.
Lemma 3.5. Suppose that condition (C).is satisfied. Then
12
i) Mγ = uγM0
ii) There exists an element wγ ∈ M ′ = JM J such that wγ w∗
γ = w∗
γ wγ = eγ,
and (M ′)γ = (M ′)0wγ = wγ(M ′)0.
then x = meγuγ = eγmuγ =
uγ(u∗
i) Clearly, if x = muγ for some
Proof.
γmuγ) ∈ uγM0 and conversely.
ii) Obviously, wγ = J u∗
γJ satisfies the equality wγ w∗
γ = w∗
γ wγ = J eγJ.Since
eγ is a central projection of M, we can apply [8, 9.6.18.] to get J eγJ = eγ.
Lemma 3.6. Let (M, G, α), M ⊂ B(H) be a W*-dynamical system with
6= 0, we have
G compact abelian as above. Then, if γ, γ ′ ∈ sp(α) and eγeγ ′
γ ′ − γ ∈ sp(α) (therefore, γ − γ ′ ∈ sp(α)) and eγeγ ′ 6 eγ−γ ′.
Proof. Since eγeγ ′
6= 0. Hence,
applying Lemma 2.5. ii), it follows that Mγ ′−γ 6= 0, so γ ′ − γ ∈ sp(α). To prove
the last statement of the lemma, notice that by Lemma 2.5. and Lemma 3.5.
6= 0, we have uγu∗
6= 0. so u∗
γuγ ′u∗
γ ′
γuγ ′
eγ eγ ′ = uγu∗
γuγ ′u∗
γ ′ = uγuγ ′−γmu∗
γ ′ = uγ uγ ′−γ meγ ′u∗
γ ′ = uγ uγ ′−γeγ′ mu∗
γ ′
for some m ∈ M0. Further, using repeatedly Lemma 3.5. we get
eγeγ ′ = uγuγ ′−γeγ′ mu∗
γ ′ = uγuγ ′−γ u∗
γ ′uγ ′mu∗
γ ′ =
= uγu∗
γ ′(uγ ′uγ ′−γ u∗
γ ′)uγ ′mu∗
γ ′ =
uγ−γ ′m1uγ ′−γ m2uγ ′mu∗
γ ′ = uγ−γ ′uγ ′−γ m3m2eγ−γ ′m1uγ ′mu∗
γ ′ = eγ−γ ′m4
for some m1, m2, m3, m4 ∈ M0\ {0} . Therefore,
eγeγ ′ = eγ−γ ′m4m∗
4eγ−γ ′ 6 km4m∗
4k eγ−γ ′
So eγeγ ′ 6 eγ−γ ′.
Lemma 3.7. Suppose that Condition (C) is satisfied. Then
i) M+ is the w*-closed subalgebra of B ( H+) generated by M0 and
{uγ : γ ∈ sp(α), γ > 0} .
ii) (M ′)+ is the w*-closed subalgebra of B ( H+) generated by (M ′)0 and
{wγ : γ ∈ sp(α′) = sp(α), γ > 0} .
Proof. Follows from Lemma 3.5.
We will prove next our results about reflexivity.
In Theorem 3.8. we assume that Γ is archimedean linearly ordered and that
Condition (C) is satisfied.
Theorem 3.8. Let (M, G, α) be such that bG = Γ is archimedean linearly
ordered and Condition (C) is satisfied. Then M+ ⊂ B(H+) is hereditarily
reflexive.
13
In Theorem 3.9. we assume that Γ is a direct product (or a direct sum) of
archimedean linearly ordered discrete groups, but we assume a stronger condi-
tion than Condition (C).
Theorem 3.9. Let (M, G, α) be such that bG = Γ ⊆ Πι∈I Γι is the direct
product, or the direct sum, of archimedean linearly ordered discrete groups, Γι.
Suppose that sp(α) = Γ and that for every γ ∈ sp(α) = Γ, there exists a unitary
operator uγ ∈ Mγ. Then M+ ⊂ B(H+) is hereditarily reflexive.
The proofs of these theorems will be given after some auxiliary results.
Lemma 3.10. Suppose that Γ is archimedean linearly ordered and that
Condition (C) is satisfied. Then
i) (M+)′ = (M ′)+, where ( M+)′ denotes the commutant of M+ in B(H+)
and
ii) ((M
′
′
)+)
= M+
Proof.
i) Let x ∈ (M ′)γ1 and m ∈ Mγ2 , γ1, γ2 ∈ sp(α) ∩ Γ+ . Then,
p+xp+mp+ = p+xmp+ = p+mxp+ = p+mp+xp+
so, (M ′)+ ⊂ (M+)′. To prove the converse inclusion, let x ∈ (M+)′ ⊂ B(H+).
Consider the following dense subspace of H+
H ′ =Xγ∈F
Hγ : F ⊂ sp(α) ∩ Γ+ a finite subset.
Then, clearly, the subspace
.
H ′′ = lin {uγξ : ξ ∈ H ′, γ ∈ sp(α)} .
Pn
where uγ is the partial isometry in Condition (C), is dense in H. Let η =
i=1 uγi ξi ∈ H". Without loss of generality, we will assume in the rest of this
proof that γ1 6 γ2 6 ... 6 γn. If uγi u∗
uγi = eγi , 1 6 i 6 n are the
central projections from Condition (C) above, then, a standard calculation in
the commutative W*-algebra Z(M ), shows that
γi = u∗
γi
eγ1 ∨ eγ2 ∨ ... ∨ eγn = eγ1 +
nXi=2
(1 − eγ1)...(1 − eγi−1)eγi .
(1)
where eγ1 ∨ eγ2 ∨ ... ∨ eγn = sup {eγi : 1 6 i 6 n} . Clearly, the terms of the
above sum are mutually orthogonal central projections in Z(M ). So if η =
i=1 uγi ξi ∈ H" it follows that
Pn
η = eγ1 η +
nXi=2
(1 − eγ1)...(1 − eγi−1)eγi η.
(2)
14
or,
η =
nXi=1
piη
(3)
where p1 = eγ1 and pi = (1 − eγ1)...(1 − eγi−1 )eγi , 2 6 i 6 n. Define the operator
bx on H ′′ as follows
We prove first that bx is well defined. Indeed, suppose thatPn
will show thatPn
nXi=1
i=1 uγi xξi = 0.Now, if η =Pn
nXi=1
bx(
piη = 0, 1 6 i 6 n.
uγi ξi) =
uγi xξi.
i=1 uγi ξi = 0. We
i=1 uγi ξi = 0, it follows that
We must show that
pi
uγj xξj = 0, 1 6 i 6 n.
nXj=1
These equalities imply thatPn
i=1 uγi xξi = 0, sobx is well defined. Since piη = 0,
nXj=1
piuγj piξj = 0.
nXj=i
piuγj ξj =
we have
(4)
Thus, factoring out uγi
uγi
nXj=1
piu∗
γi
uγj piξj = 0
By multiplying the above equality by u∗
γi
we get
, and taking into account that pi 6 eγi
where u∗
γi
x ∈ (M+)′, it follows that
uγj ∈ Mγj −γi . Since γi 6 γj, if i 6 j 6 n so, Mγj −γi ⊂ M+ and
x(pieγi ξi +
nXi=2
piu∗
γi
uγj ξj) = pieγi xξi +
nXj=i+1
piu∗
γiuγj xξj = 0.
By multiplying the above equality by uγi , we get
piuγixξi +
nXj=i+1
piuγj xξj = pi
uγj xξj = pi
nXj=i
uγj xξj = 0.
nXj=i
15
so
piu∗
γi
uγj ξj = 0.
nXj=i
pieγi ξi +
piu∗
γi uγj ξj = 0.
nXj=i+1
Further, since γi 6 γj, and pjuγi = 0 when i 6 j and x ∈ (M+)′ we have
and this proves (4). Therefore, bx is well defined. From the definition of bx it
follows that bxm = mbx on H" for all m ∈ M andbx(η) = x(η) for every η ∈ H ′,
so if, as we will prove,bx is bounded, it follows thatbx ∈ M ′. Next we prove that
the operatorbx is bounded. Indeed, if, as above, η =P uγi eγi ξi, then, using the
equality (3) and the fact thatbx commutes with pi, 1 6 i 6 n, we have
nXi=1
kpibx (piη)k2.
piuγj xξj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
nXj=i
γi uγj ξj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Therefore, kbx(η)k 6 kxk kηk so bx is bounded. As noticed above, bx ∈ M ′ and,
since obviously, p+bxp+ = x it follows that x ∈ (M ′)+ and we are done.
=vuut
kbx(η)k =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)bx nXi=1
piη!(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
=(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
pibx (piη)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
nXi=1
=(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
piuγj ξj)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
kpibx (piη)k2 =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
nXj=i
nXj=1
pibx(
6(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
uγj xξj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
=(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
piuγj xξj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
nXj=i
nXj=i
6 kxk2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
γi uγj xξj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
=(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
nXj=1
nXj=i
γiη(cid:13)(cid:13)2
= kxk2(cid:13)(cid:13)piu∗
uγj xξj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
ii) follows from i) by replacing M with M
6(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
6 kxk2 kpiηk2 .
pix
piu∗
pi
piu∗
γi
pi
piu∗
2
=
2
=
2
6
pi
piuγi
piu∗
γi
2
pi
2
2
2
′
.
The following version of Lemma 3.10. will be used in the proof of Theorem
3.9. Here and in Theorem 3.9 we assume that Γ is a direct product or a direct
sum of archimedean linearly ordered discrete groups, but we also assume a
stronger version of Condition (C), namely that for every γ ∈ sp(α), Mγ contains
a unitary operator and that the action is faithful (i.e. αg = id implies g = 0).
These conditions imply that sp(α) = Γ (see for instance [21 Lemma 2.2.]).
Lemma 3.11. Let (M, G, α) be such that bG = Γ is a direct product, Γ =
Πι∈IΓι, (or a direct sum) of archimedean linearly ordered discrete groups Γι.
Suppose that α is faithful and for every γ ∈ sp(α), there exists a unitary operator
uγ ∈ Mγ (as noticed above, these conditions imply that sp(α) = Γ). Then
i) (M+)′ = (M ′)+, where ( M+)′ denotes the commutant of M+ in B(H+)
and
ii) ((M
′
′
)+)
= M+.
Proof.
it follows that
(M ′)+ ⊂ (M+)′. To prove the opposite inclusion, let x ∈ (M+)′. Further, let us
i) As in the proof of the previous Lemma 3.10.
16
denote
and
H ′ =Xγ∈F
Hγ : F ⊂ Γ+ a finite subset.
H ′′ = lin {uγξ : ξ ∈ H ′, γ ∈ Γ} .
on H" as follows
Clearly H ′ is dense in H+ and H" is dense in H. Define the linear operator bx
uγi ξi) =
uγi xξi.
We will prove first that bx is well defined. Suppose that
uγiξi = 0.
(5)
nXi=1
bx(
nXi=1
nXi=1
Since, by Lemma 3.3., Γ is lattice ordered, let ν = inf {γi : 1 6 i 6 n} ∈ Γ.
Since, by hypothesis sp(α) = Γ, we have ν ∈ sp(α). Let uν ∈ Mν be a unitary
operator as in the hypothesis. By multiplying (5) by u∗
ν we get
uγi−ν ξi = 0.
nXi=1
Since γi − ν ∈ Γ+, i = 1, 2, ..., n and x ∈ (M+)′ it follows that
x
uν
so
u∗
uγi ξi) = 0
u∗
ν uγi ξi =
u∗
ν uγi xξi =
nXi=1
ν uγi xξi = 0.
nXi=1
nXi=1
nXi=1
uγi xξi =bx(
sobx is well defined. We will prove next that bx is continuous. Indeed
ν uγi xξi(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
uγi xξi(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
=(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
=(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
nXi=1
nXi=1
ν uγi ξi(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
6 kxk(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
ν uγiξi(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
=(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
nXi=1
nXi=1
uγi ξi(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
= kxk(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
uγi ξi(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
nXi=1
nXi=1
sobx is continuous. As in the proof of the previous Lemma 3.10. we can see that
bx ∈ M ′ and p+bxp+ = x, so x ∈ (M ′)+.
nXi=1
uγiξi)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)bx(
nXi=1
=(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
ν uγi xξi(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
nXi=1
kxk(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
ii) follows from i) by replacing M with M
x
u∗
u∗
ν
uν
u∗
u∗
u∗
=
=
.
′
.
17
It is worth mentioning that the previous two Lemmas imply, in particular,
that (M+) " = M+ but we will not use this fact.
In the next two lemmas we will assume that (M, G, α) is a W*-dynamical
system that satisfies Condition (C) and M is in standard form. We also assume
that Γ is a direct product or sum of archimedean linearly ordered abelian discrete
groups. If γ0 ∈ sp(α) ∩ Γ+ r {0}and γ ∈ sp(α) ∩ Γ+ according to Corollary 3.2.
there exists p ∈ Z+ such that either pγ0 6 γ < (p + 1)γ0 or pγ0 6 γ and γ is not
comparable with (p + 1)γ0. If, in addition, eγeγ0 6= 0 we will denote for every
λ ∈ C, λ < 1
x(λ, γ0, γ, ξ) =Xn>0
λnun
γ0
y(λ, γ0, γ, ξ) =Xn>0
λnwn
and
Kγ0,γ,λ =
Lγ0,γ,λ =
uγ−pγ0ξ : ξ ∈ eγ0 H0
γ0 wγ−pγ0 ξ : ξ ∈ eγ0 H0
⊂ eγ0 H+.
⊂ eγ0 H
Lemma 3.12. Let (M, G, α) be a W*-dynamical system that satisfies Con-
dition (C). Γ and γ0, γ ∈ sp(α) ∩ Γ+ be as above. Then,
Kγ0 = lin {Kγ0,γ,λ : γ ∈ sp(α) ∩ Γ+, λ ∈ C, λ < 1}
is dense in eγ0 H+.
Proof. Notice first that if γ0, γ are as in the hypothesis of the lemma, then,
by the definition of eγ0 in the Condition (C),we have that pγ0 ∈ sp(α) for every
p ∈ Z+ and, by Lemma 3.6., γ − pγ0 ∈ sp(α), so uγ−pγ0 exists, and thus the
definition of Kγ0,γ,λ in the hypothesis of the Lemma is consistent. Now, taking
λ = 0 in Kγ0,γ,λ, it follows that eγ0 uγ−pγ0 H0 ⊂ Kγ0,γ ⊂ Kγ0 . In particular, for
p = 0 (so, when either 0 6 γ < γ0 or 0 6 γ and γ is not comparable with γ0),
we have eγ0 uγH0 = eγ0 Hγ ⊂ Kγ0. We will prove that eγ0 Hγ = uγeγ0 H0 ⊂ Kγ0
eγ0 Hγ ⊂ Kγ0 and
eγ0 Hγ ⊂ K γ0 , so Kγ0 is dense in eγ0 H+ as claimed.
uγ−pγ0 H0 ⊂ Kγ0,γ . The case p = 0 was proved
for every γ ∈ sp(α) ∩ Γ+. This fact will imply that Pγ∈Γ+
therefore, eγ0 H+ =Pγ∈Γ+
We will prove first that eγ0 up
γ0
above. Suppose that p > 0. We will prove by induction on k that
eγ0 uk
γ0 uγ−pγ0 H0 ⊂ Kγ0
for every k, in particular for k = p. If k = 0, the above inclusion follows, as
noticed at the beginning of this proof from the definition of Kγ0,γ for λ = 0.
Suppose by induction that ul
γ0uγ−pγ0 ξ ∈ Kγ0 for l = 0, 1, ...k − 1.Then
λnun
γ0 uγ−pγ0 ξ ∈ Kγ0
Xn>k
18
for every ξ ∈ eγ0 H0, λ ∈ C, λ < 1. Thus
λkuk
γ0 uγ−pγ0ξ + Xn>k+1
λnun
γ0 uγ−pγ0 ξ ∈ Kγ0
By dividing the above relation by λk, λ 6= 0 and then taking the limit as λ → 0,
we get that uk
γ0 uγ−pγ0 ξ ∈ Kγ0 for every
ξ ∈ H0. Since, obviously, up
ii),
M−pγ0Mγ ⊂ Mγ−pγ0, it follows that
γ0)p = eγ0 for p > 0 and, by Lemma 2.5.
γ0 uγ−pγ0ξ ∈ Kγ0 , so, in particular, up
γ0(u∗
eγ0 Hγ = eγ0 uγH0 = up
γ0(u∗
γ0)puγH0 ⊂ up
γ0
uγ−pγ0H0 ⊂ Kγ0
and we are done.
The following lemma can be proven similarly with the previous Lemma 3.12.
Lemma 3.13. Let (M, G, α) be a W*-dynamical system that satisfies Con-
dition (C), Γ and γ0, γ ∈ sp(α) ∩ Γ+ be as above. Then,
Lγ0 = lin {Lγ0,γ,λ : pγ0 6 γ < (p + 1)γ0 for some p ∈ Z+, λ ∈ C, λ < 1}
is dense in eγ0 H+.
′
′
′
)0p+ ⊂ B(H+). Therefore, for every projection f ∈ (M
Proof. of Theorem 3.8. We will prove first that M+ ⊂ B(H+) is reflexive
and then apply Lemma 3.4. and the subsequent discussion to infer that M+ is
hereditarily reflexive. Let γ0 ∈ sp(α)∩Γ+. Since uγ0 ∈ M, it follows that uγ0 f =
f uγ0 for every projection f ∈ (M ′), in particular for every projection f ∈ (M ′)0.
By Lemma 2.8. iv), since (M ′)γ H0 = Hγ , γ ∈ sp(α), we have, in particular that
)0, f H+belongs to
(M
Lat(M+). Let x ∈ algLat(M+) ⊂ B(H+). We will prove that x ∈ ((M
and,
then, applying Lemma 3.11. ii) it will follow that x ∈ M+, so M+ is a reflexive
operator algebra. The way to prove this fact is to use Lemma 3.14. to show
that x∗ ∈ ((M
= M+.As
noticed above, f H+ ∈ LatM + for every projection f ∈ (M
)0,so xf = f x and
therefore x∗f = f x∗ for every projection f ∈ (M
)0. It follows that x commutes
with every element of (M
)0. We will prove next that xwγ = wγ x for every
γ ∈ sp(α′) ∩ Γ+ = sp(α) ∩ Γ+ and then apply Lemma 3.10. ii) to infer that x ∈
= M+. To this end, let γ0 ∈ sp(α) ∩ Γ+. If γ0 = 0, then as convened,
((M
uγ0 = wγ0 = I, so nothing to prove. Let γ0 ∈ sp(α) = sp(α′), γ0 > 0. Denote
) the operator uγ0 ∈ M+ (respectively wγ0 ∈ (M ′)+)
by Tuγ0
defined on H+. Then the adjoints of Tuγ0
and then, clearly, it will follow that x ∈ ((M
(respectively Twγ0
on H+ are
, Twγ0
)+)
′
′
)∗
+)
)+)
′
′
)+)
′
′
′
′
′
′
T ∗
uγ0
ξγ = 0 if 0 6 γ < γ0 and T ∗
uγ0
ξγ = u∗
γ0
ξγ if γ > γ0.
and similarly
T ∗
wγ0
ξγ = 0 if 0 6 γ < γ0 and T ∗
wγ0
ξγ = w∗
γ0
ξγ if γ > γ0.
19
Since uγ0 , wγ0 (so Tuγ0
well. Notice that if, for λ ∈ C, λ < 1, we denote
, Twγ0
) commute, it follows that T ∗
uγ0
, T ∗
wγ0
commute as
eLγ0,λ =nξ ∈ H+ : T ∗
wγ0
ξ = λξo .
wγ0
then, since T ∗
x ∈ algLat(M+) it follows that.Lγ0,γ,λ ∈ Lat(x∗) for every λ ∈ C, λ < 1. Since
for the eigenvalue λ, then, it follows that
commutes with (M+)∗, we haveeLγ0,γ,λ ∈ Lat(M+)∗ and, since
x∗ξ for every ξ ∈eLγ0,λ . On the other hand, it is clear that
eLγ0,γ,λ consists of eigenvectors of T ∗
ξ = T ∗
x∗T ∗
wγ0
wγ0
wγ0
Lγ0,γ,λ ⊂eLγ0,λ
wγ0
where Lγ0,γ,λ is as in Lemma 3.13., so x∗T ∗
By Lemma 3.13.,
ξ = T ∗
wγ0
x∗ξ for every ξ ∈ Lγ0,γ,λ .
Lγ0 = lin {Lγ0,γ,λ : pγ0 6 γ < (p + 1)γ0 for some p ∈ Z+, λ ∈ C, λ < 1}
x∗ξ for every ξ ∈ eγ0 H+, so
is dense in eγ0 H+, and therefore x∗T ∗
+ . Since x∗
x∗ commutes with T ∗
wγ0
commutes with (M ′)∗
+ , it follows that x commutes with (M ′)+, so by Lemma
3.10. ii) x ∈ M+ so M+ is reflexive. Finally, by applying Lemma 3.4. and the
discussion following it, we see that M+ is hereditarily reflexive and we are done.
, γ0 ∈ sp(α) ∩ Γ+ and therefore with (M ′)∗
ξ = T ∗
wγ0
wγ0
Corollary 3.14. below refers to the Example a) to Condition (C).
Corollary 3.14. Let (M, G, α) be a W *-dynamical system with G compact
abelian and M a finite W*-algebra in standard form such that Z(M0) ⊂ Z(M ).
Suppose that the dual Γ of G has an archimedean linear order. Then M+ ⊂
B(H+) is reflexive.
Proof. According to [18, Theorem 2.3.], if M is finite and Z(M0) ⊂ Z(M ),
then the Condition (C) is satisfied and therefore the result follows from Theorem
3.8.
Proof. of Theorem 3.9. The proof is very similar with the proof of Theorem
3.8. The only modification is using Lemma 3.11 instead of Lemma 3.10.
The next Corollary refers to Examples b) to Condition (C).
Corollary 3.15. Let (M, G, α) be a W *-dynamical system with M an
injective von Neumann algebra in standard form and G a compact abelian group
such that the dual Γ of G is a direct product (or a direct sum) of archimedean
linearly ordered discrete groups. Suppose that α is prime and faithful and it
is not the case that M is of type III and M0 is of type II 1. Then M+ is
hereditarily reflexive.
20
Proof. Since α is faithful, we have sp(α) = Γ (see for instance [21, Lemma
2.2.]). By [21, Theorem 2.3.] each spectral subspace Mγ contains a unitary
operator. The conclusion of the Corollary follows from Theorem 3.9.
The concept of nonselfadjoint crossed product, or more generally that of
w*-semicrossed product were defined in [3], [11], [16].
Corollary 3.16. Let (M0, Γ, β) be a W*-dynamical system such that M0 ⊂
B(H0) is in standard form and Γ is a discrete abelian group. Suppose that
Γ ⊆ ΠΓι is a direct product (or a direct sum) of archimedean linearly ordered
groups. Let M = M0 ×α Γ ⊂ B(l2(Γ, H0)) be the corresponding crossed product .
Then, the non selfadjoint crossed product M+ = M0 ×β Γ+ (i.e. the algebra of
elements of M with non negative spectrum) is a hereditarily reflexive operator
algebra in B(H+) where H+ = l2(Γ+, H0).
Proof. If G denotes the (compact) dual of Γ, and α = bβ is the dual action
of β on M, then, consider the canonical conditional expectation P0 : M → M0.
Since M0 ⊂ B(H0) is in standard form, from Lemma 2.1. and the subsequent
discussion it follows that M ⊂ B(l2(Γ, H0)) is in standard form and by the
definition of the crossed product, Mγ contains a unitary operator uγ for every
γ ∈ Γ and thus the W*-dynamical system (M, G,bβ) satisfies the conditions of
Corollary 3.15.
Corollary 3.17. Let (M0, Γ, β) be a W*-dynamical system such that M0 ⊂
B(H0) is a maximal abelian von Neumann algebra and Γ is a discrete abelian
group. Suppose that Γ ⊆ ΠΓι is a direct product (or a direct sum) of archimedean
linearly ordered groups. Let M = M0 ×α Γ ⊂ B(l2(Γ, H0)) be the corresponding
crossed product . Then, the non selfadjoint crossed product M+ = M0 ×β Γ+ is
a hereditarily reflexive operator algebra in B(H+) where H+ = l2(Γ+, H0).
Proof. Since M0 ⊂ B(H0) is a maximal abelian von Neumann algebra, by
Proposition 2.4. it is spatially isomorphic with its standard form, so the result
folows from the previous Corollary 3.16.
In [2, Corollary 5.14.] it is stated that if M0 ⊂ B(H0) is a maximal abelian
von Neumann algebra and Γ = Zd, d ∈ N, then M0 ×β Γ+ is reflexive, so the
Corollary 3.18. above extends that result by showing also hereditary reflexivity
in the special case Γ = Zd, d ∈ N.
REFERENCES
1. W. B. Arveson, The harmonic analysis of automorphism groups, operator
algebras and applications, Proc. Sympos. Pure Math., vol. 38, Amer. Math.
Soc., Providence, R. I., 1982.
2. R. T. Bickerton, and E. T. A. Kakariadis, Free multivariate w*-semicrossed
products: reflexivity and the bicommutant property, Canad. J. Math. 70(2018),
1201 -- 1235.
21
3. K, R. Davidson, A. H. Fuller and E. T. A. Kakariadis, Semicrossed
Products of Operator Algebras by Semigroups, Memoirs of the AMS, Vol. 247,
2017.
4. D. W. Hadwin and E.A. Nordgren, Subalgebras of reflexive algebras. J.
Operator Theory 7 (1982), 3 -- 23.
5. L. Helmer, Reflexivity of non-commutative Hardy algebras, J. Funct.
Anal. 272(2017), 2752 -- 2794.
6. V. F. R. Jones, Prime actions of compact abelian groups on the hyperfinite
type II1 factor, J. Operator Theory, 9(1983), 181-186.
7. V. F. R. Jones and M. Takesaki, Actions of compact abelian groups on
semifinite injective factors, Acta Math. 153 (1984), 213 -- 258.
8. R. V. Kadison and J. R. Ringrose, Fundamentals of the theory of operator
algebras, Vol. II Advanced Theory,Academic Press 1986.
9. E. T. A. Kakariadis, Semicrossed products and reflexivity, J. Operator
Theory, 67(2012), 379-395.
10. A I Loginov and V S Sul'man, Hereditary and intermediate reflexivity
of W*-algebras, Mathematics of the USSR-Izvestiya 9(1975), 1189-1202.
11. M. McAsey, P. S. Muhly, and K.-S. Saito, Nonselfadjoint crossed prod-
ucts (invariant subspaces and maximality), Trans. Amer. Math. Soc., 248(1979),
381 -- 409.
12. P. S. Muhly and B. Solel, Hardy algebras, W*-correspondences and
interpolation theory, Math. Ann. 330 (2004), 353 -- 415.
13. G. K. Pedersen, C*-algebras and their automorphism groups, Academic
Press 1979.
14. C. Peligrad, Reflexive operator algebras on noncommutative Hardy
spaces, Math. Annalen, 253(1980), 165-175.
15. C. Peligrad, Invariant subspaces of algebras of analytic elements associ-
ated with periodic flows on von neumann algebras, Houston J. Math., 42(2016),
1331-1445.
16. J. R. Peters, Semicrossed products of C*-algebras. J. Funct. Anal.,
59(1984), 498 -- 534.
17. H. Radjavi and P. Rosenthal, Invariant subspaces, 2nd edition, Dover
Publications, Mineola, New, York, 2003.
18. K.-S. Saito, Nonselfadjoint subalgebras associated with compact abelian
group actions on finite von Neumann algebras, Tohoku Math. Journ. 34(1982),
485-494.
19. D. Sarason, Invariant subspaces and unstarred operator algebras, Pacific
J. Math., 17(1966), 511-517.
20. S. Stratila, Modular theory in operator algebras, Bucharest; Abacus
Press, Tunbridge Wells, 1981
21. K. Thomsen, Compact abelian prime actions on von Neumann algebras,
Trans. Amer. Math. Soc., 315(1989), 255-273.
22
|
1003.2676 | 1 | 1003 | 2010-03-13T05:22:46 | Automorphisms of the bipartite graph planar algebra | [
"math.OA"
] | For any abstract subfactor planar algebra $P$, there exists a finite index extremal subfactor $M_0 \subset M_1$ with $P$ as its standard invariant. In this paper, we classify the automorphism group of a bipartite graph planar algebra, and obtain subfactor planar subalgebras by taking fixed points under groups of automorphisms. This construction provides both new examples of subfactors and new descriptions of the planar algebras of previously known examples. | math.OA | math | Automorphisms of the bipartite graph planar algebra
R. D. Bursteina,∗
aDepartment of Mathematics
1326 Stevenson Center
Vanderbilt University
Nashville, TN 37240
1. Introduction
0
1
0
2
r
a
M
3
1
]
.
A
O
h
t
a
m
[
1
v
6
7
6
2
.
3
0
0
1
:
v
i
X
r
a
Planar algebras were introduced by Jones in [13]. They are a powerful tool
for studying subfactors, providing a graphical calculus on the standard invariant
of an finite index extremal subfactor of type II1 ([13],[22]). The planar operad
is the set of planar tangles with zero or more internal disks and a checkerboard
shading, with a distinguished region of the boundary and each internal disk,
all taken up to isotopy. The operation is gluing: a tangle may be pasted into
an internal disk of another tangle (matching up the distinguished boundary
regions) if the number of strands and the shading are compatible. A planar
algebra is a graded vector space P = (V ±
n ), n ≥ 0, along with an associative
action of the planar operad.
If M0 ⊂ M1 is a finite index extremal II1 subfactor with Jones tower M0 ⊂
M1 ⊂ M2 ⊂ ..., then we may take V +
1 ∩ Mn+1. There is
an operad action defined in [13] making this into a planar algebra. Conversely,
Jones describes a certain list of additional properties which make P into a Popa
system, implying that there exists a subfactor M0 ⊂ M1 of which P is the
standard invariant [22]. A planar algebra with this list of properties is called a
subfactor planar algebra (SPA). Other methods of constructing a subfactor from
an SPA have since been obtained ([11],[15],[16]), which use more diagrammatic
notation than Popa's original method.
0 ∩ Mn, V −
n = M ′
n = M ′
Constructing an SPA abstractly therefore implies the existence of a corre-
sponding subfactor. This method may be used to find new subfactors (as in [6]
or [1]), or provide new proofs of the existence of subfactors with specified proper-
ties. Abstract constructions of SPAs of previously known subfactors can provide
new insight into the structure of their standard invariants (e.g.
[3],[4],[19]).
A planar algebra may be constructed from any finite bipartite graph [14],
and some infinite graphs as well [11]. These bipartite graph planar algebras
(BGPAs) are almost never of subfactor type, because their vector spaces are
too large. However, they possess several of the necessary properties required for
∗Corresponding author. Tel. 1-615-322-6672 Fax 1-615-343-0215
Email address: [email protected] (R. D. Burstein)
Preprint submitted to Elsevier
October 30, 2018
SPAs, which are inherited by planar subalgebras. We may therefore try to find
SPAs by looking at small planar subalgebras of BGPAs.
It is generally difficult to show that a graded subspace of a BGPA is closed
under the action of the planar operad, although progress has been made in the
single generator case (e.g.
[7],[8],[19],[1]). In this paper, we describe a method
for doing so by taking fixed points of a BGPA under a group of automorphisms.
In section 2 we describe BGPAs, with particular attention to infinite graphs.
We introduce a slightly different notation from [11] and [14]; this simplifies the
computations of section 3, where we compute the automorphism group of an
arbitrary BGPA PΓ. In section 4 we find conditions for a planar subalgebra of a
BGPA to be an SPA, and we conclude in section 5 by presenting some examples
of SPAs obtained by this planar fixed point construction.
2. The bipartite graph planar algebra
The bipartite graph planar algebra (BGPA) is described by Jones in [14].
The data required are a finite bipartite graph Γ and a spin vector: i.e. a function
from the vertices of the graph to the positive real numbers. Given such data,
there is a planar algebra PΓ = (V ±
n has a basis labelled
by loops of length 2n in the graph, starting at even or odd vertices depending
on sign.
n ), n ≥ 0, where each V ±
We will consider BGPAs on infinite graphs as well. In this case, the vector
spaces will be infinite dimensional, so there are some new topological consider-
ations. To simplify later computations, we will describes the vector spaces of
the planar algebra as operators on a certain Hilbert space.
In our diagrams, we will use the convention of [15] that a thick line represents
as many parallel strands as necessary. Also we will omit shading in a diagram
when both shadings can occur. When not otherwise specified, the distinguished
boundary region of a tangle is on the left.
The discussion that follows is taken from [14], adjusted where necessary to
allow for infinite graphs.
Let Γ be a locally finite bipartite graph, i.e. each vertex of Γ is an endpoint
of at most A edges. Let µ be a function from the vertices of Γ to the positive
real numbers obeying the following local boundedness condition: there is some
M > 0 such that for any two adjacent vertices v and w, we have µ(v)/µ(w) < M .
Let l±
n be the set of loops of length 2n on Γ, starting at an even (+) or odd (−)
vertex. Then the vector space V ±
n of the bipartite graph planar algebra PΓ is
the set of bounded functions from l±
n to C.
The boundary type of a tangle is the ordered pair (n,±), where 2n strands
intersect the tangle boundary transversely and the distinguished region of the
tangle is shaded (+) or unshaded (−). The same definition is used for the
boundary type of each internal disk of a tangle.
Let T be a planar tangle with k internal disks. Let the boundary type of
T be (n,±), and let the internal disks of T have boundary types respectively
(n1,±), (n2,±), ..., (nk,±). Then to describe the planar operad, for each set of
2
inputs (x1, ..., xk) (with xk ∈ V ±
n . This
assignment must agree with gluing, be multilinear in the inputs, and be isotopy
invariant. We define Z(T ) following [14].
nk ) we must assign the output Z(T ) ∈ V ±
A state of the tangle is a function σ which maps the strands of T to edges
of Γ, and the regions of T to vertices of Γ. Shaded regions are mapped to
positive vertices, and unshaded regions to negative ones. A state must obey a
compatibility condition: if a strand S is adjacent to a region R, then σ(R) must
be one of the endpoints of σ(S).
A state σ is compatible with a given loop if treading the output of σ coun-
terclockwise around the boundary of T , starting from the distinguished region,
produces that loop.
A singularity of T is a local maximum or local minimum of a strand. Fix
a state σ; let v be the vertex associated to the concave side of the singularity
by σ, and w the vertex associated to the convex side. Then the value of the
singularity is µ(v)/µ(w).
A state σ associates a loop Li to each internal disk Di of T , obtained by
reading the output of σ counterclockwise starting from the distinguished region
of the disk. The value of the state on Di is then the value of the ith input xi
at this loop Li.
Now we can define the value of Z(T ) on a specified loop L ∈ l±
n . This is
Xσ compatible with L
σ(Di)
Y
σ(s)
YDi ∈internal disks
s∈singularities of T
where σ is evaluated on disks and internal singularities as above.
By local finiteness of Γ, only finitely many states are compatible with L.
Specifically, let m be the largest number of strands that must be crossed to get
from the distinguished boundary region of T to any other region (boundary or
internal). Then if v is the first vertex of L, every compatible state must assign
all regions of T to vertices within a distance of m from v on the graph. A state
is determined by its value on strands. Let each vertex of Γ contact at most emax
edges. Then each strand must be chosen from at most em
max possibilities, and if
T contains a distinct strands, each loop is compatible with at most eam
max states.
The local boundedness condition on µ (µ(v)/µ(w) < M < ∞ for all adjacent
vertices v and w) means that each singularity has value at most M , for any state.
Let T have b singularities. Elements of V ±
ni are bounded; let N be the largest
bound of any input xi, 1 ≤ i ≤ k.
on each loop L as a finite sum. Moreover, this sum is bounded by eam
so the output is a bounded function on l±
element of V ±
n .
Putting this together, we find that the tangle output Z(T ) may be evaluated
maxM bN k,
n and this evaluation rule produces an
The proofs of [14] that this map is multilinear, isotopy invariant and respects
gluing may be used without alteration, since local finiteness of Γ implies that
all necessary sums are finite. Therefore the above definition of Z(T ) produces
a planar algebra.
3
There is a natural antilinear involution, which we refer to as ∗, of each vector
space V ±
n , the reversed loop L′ consists of the
n (see [14]). For each loop L ∈ l±
same list of vertices and edges, taken in the opposite order. Then the involution
is defined by A∗(L) = A(L′), for all A ∈ V ±
Let ρT be the map corresponding to some tangle T with k internal disks,
i.e. ρT (x1 ⊗ x2 ⊗ ... ⊗ xk) is equal to Z(T ) when the inputs are (x1, x2, ..., xk).
k) = ρT ′ (x1 ⊗ x2 ⊗ ... ⊗ xk)∗, where
Then as in [14] we have ρT (x∗
the tangle T ′ is the mirror image of T . From [13], this means that the BGPA
PΓ is a planar ∗-algebra, with involution as above.
of bounded linear operators on a certain Hilbert space.
For ease of later computation, it will be convenient to describe V ±
n and L ∈ l±
n .
2 ⊗ ... ⊗ x∗
n as a set
1 ⊗ x∗
Let H ±
n be the Hilbert space with basis {xp} labelled by the paths of length
n on Γ whose initial vertex is even (+) or odd (−). These vector spaces are all
infinite dimensional when Γ is infinite. We also define the total path Hilbert
space H as the direct sum of the H ±
n 's: H =Pn,± H ±
n .
n . A loop L ∈ l±
Each element A of Vn± naturally defines a linear map on H ±
n
may be described as a pair of paths (π, ǫ) of length n whose endpoints are the
same. Then π and ǫ correspond to basis vectors xπ, xǫ in H ±
n , and we take
hA(xǫ), xπi = A(L).
is equal to (ǫ, π).
hA∗(xǫ), xπi = hA(xπ), xǫi. In other words, the involution acts on V ±
adjoint operation for B(H ±
n is equal to the pair of paths (π, ǫ), then the reversed loop L′
It follows from the above definition of the involution that
n as the
If L ∈ l±
n ).
Every path p has a starting vertex s(p) and a terminal vertex t(p). For each
vertex v of Γ we may define a projection sv ∈ B(H ±
n ) which fixes the closed
linear span of {xps(p) = v}, and likewise tv which fixes the closed span of
{xpt(p) = v}. The sv's and tw's form an abelian algebra; in fact the projections
{svtwv, w ∈ vertices of Γ} are a partition of unity. Since π and ǫ above always
have the same endpoints, we have each A ∈ V ±
n commuting with sv and tv for
all vertices v.
This means that A ∈ V ±
n may be thought of as a (potentially infinite) sum
of operators Avw each acting on the subspace svtw(H ±
n ). Each svtw is of finite
rank, and this rank is universally bounded by local finiteness of Γ, so these
subspaces have bounded dimension. Since the components of A are bounded by
definition, it follows that A is bounded in norm as a linear operator on H.
We then have V ±
n ⊂ B(Hn)±, and in fact V ±
n ⊂ {sv, twv, w ∈ vertices of Γ}′.
Each svtwB(H ±
n ) is a finite dimensional matrix algebra, with matrix units given
by partial isometries from xp to xq where p and q are paths from of length n from
v to w. All such partial isometries are contained in V ±
n ) ⊂ V ±
n for
all v, w. {sv, tw}′ consists of bounded formal sums of elements of svtwB(H ±
n ),
so from the definition of V ±
n = {sv, tw}′ as operators on
H ±
n . This is a von Neumann algebra by the bicommutant theorem.
n 's; these are just the scalars
We will freely refer to scalar elements of the V ±
n it then follows that V ±
n , so svtwB(H ±
as operators on the appropriate Hilbert space.
To assist with computations using the above notation, we now define a con-
4
catenation operation c on the H ±
n 's:
m → H ±
Definition 2.1. Let p and q be two paths on the graph Γ. Then c(xp, xq) is
zero if t(p) 6= s(q), and otherwise is xr where r is the path obtained by first
following p and then q. This operation extends linearly and continuously to maps
H ±
n ⊗ H ±
From the definition of the bases of the V ±
n 's, every concatenation map is
surjective. Furthermore, these maps are associative: c(c(x, y), z) = c(x, c(y, z)).
This means we may freely apply c to multiple inputs via the inductive definition
c(x1, x2, ..., xn) = c(c(x1, x2, ..., xn−1), xn).
k , where k = n + m and the signs are chosen appropriately.
We also define an antilinear path reversal operator rev:
Definition 2.2. Let p be a path on the graph Γ. Let q be the reverse path of p,
i.e. the same edges and vertices taken in the opposite order. Then rev(xp) = xq.
This operation extends antilinearly and continuously to maps from H ±
n to H ±
n
or H ∓
n , depending on the value of n.
Both c and rev are bounded, so they extend to the total Hilbert space H.
rev is an antilinear involution.
Later, we will be interested in demonstrating that certain maps on the V ±
n 's
commute with the action of the planar operad. To do this it suffices to show that
such maps commute with particular tangles that generate the planar operad (see
e.g.
[4]). We now describe the action of one set of such generating tangles in
terms of the above notation. All of these actions may be readily verified directly
from the operad definition.
The multiplication tangle:
B
A
This tangle corresponds to operator multiplication. The output is AB.
The left embedding tangle l(A):
A
For A ∈ V ±
where x ∈ H ±
n , we have l(A) ∈ V ∓
n and v ∈ H ∓
n+1 defined by l(A)(c(v, x)) = c(v, A(x)),
1 . This operation is bounded in norm by emax.
5
The right embedding tangle r(A) :
A
For A ∈ V ±
1 or H −
where x ∈ H ±
is norm bounded by emax as well.
n , we have r(A) ∈ V ±
n and v ∈ H +
n+1 defined by r(A)(c(x, v)) = c(A(x), v),
1 depending on the value of n. This operation
Both the left and right embedding operators are strongly continuous; this
follows directly from the definition.
Temperley-Lieb generators (T L+ and T L−):
Using the above operad definition, the T L+ element acts on the path basis
and
as follows:
Let a and b be paths of length 1 on Γ which start at positive vertices,
c and d paths of length 1 which start at negative vertices, with xa, xb, xc,
xd the corresponding basis vectors in V +
1 . Then we should have
1
hT L(c(xa, xc)), c(xb, xd)i = 0 unless a is the reverse of c and b is the reverse
of d, with additionally s(a) = s(b).
If these conditions hold, then the inner
product should be µ(t(a))µ(t(c))/µ(s(a))2.
In other words, for each vertex v ∈ P +
and V −
0 , let
yv = Xes(e)=v
µ(t(e))c(xe, rev(xe))
where the sum is taken over all paths of length 1 on Γ.
Then
T L(yv) = Xe1,e2s(e1,2)=v
= Pes(e)=v µ(t(e))2
µ(v)2
!
Xes(e)=v
= Pes(e)=v µ(t(e))2
µ(v)2
yv
µ(te1 )µ(te2 )2/µ(v)2c(xe2 , rev(xe2 ))
µ(t(e))c(xe, rev(xe))
For each positive vertex v, yv is an eigenvector for T L with eigenvalue δv =
Pes(e)=v µ(t(e))2/µ(v)2, and T L is zero off the closed linear span of these yv's.
We also have a T L− element. This is the same diagram as above but with
reversed shading, and is defined by reversing all signs in the above definition.
yv is defined as above, but now for v being a negative vertex. Then T L− has
zero off the closed linear span of these yv's.
each such yv as an eigenvector with eigenvalue Pes(e)=v µ(t(e))2/µ(v)2, and is
We can now compute left and right capping operators, respectively LC(A) :
n → V ∓
V ±
n−1 and and RC(A) : V ±
n → V ±
n−1.
6
A
A
These operators may be described in terms of embedding and the TL gen-
erators:
rn−1(T L)l(A)rn−1(T L) = rn−2(T L)l2(LC(A))
ln−1(T L)r(A)ln−1(T L) = ln−2(T L)r2(RC(A))
This uniquely defines the capping operation: from the definitio of left and
right embedding rn−2(T L)l2(LC(A)) and ln−2(T L)l2(RC(A)) are each zero
only if LC(A), RC(A) respectively are zero. Note that a graded bounded linear
map on the V ±
n 's which fixes the Temperley-Lieb algebra and commutes with
multiplication and embedding also commutes with capping; for such a map ω
we have
ln−2(T L)r2(RC(ω(A))) =
ω(ln−2(T l)r2(RC(A))) =
ln−2(T L)r2(ω(RC(A)))
by the above definition, and the same holds for LC.
0 -valued sesquilinear form on V ±
We now describe the interaction of the involution on V ±
n (defined above)
with these tangles. Since conjugation by the involution corresponds to tangle
reflection (see [13]), it follows that (AB)∗ = B∗A∗, T L± are self-adjoint, and
the involution commutes with left and right embedding. This is consistant with
the above description of this involution as the adjoint operation on each B(H ±
n ).
One important property of the BGPA in [14] was the existence of a positive
definite V ±
n , namely hx, yi = RCn(y∗x) in the
above notation. We would like this form to be positive definite here as well.
Let p, q, r, s be path basis elements in V ±
n , and A, B rank one partial isometries
from (respectively) p to q and r to s. Then it follows directly from the operad
definition that hA, Bi is a positive scalar multiple of a rank one projection in
V +
0 if p = q and r = s, and is zero otherwise. It follows that the form is positive
definite on bounded formal sums of such elements, which constitute all of V ±
n .
In order for a BGPA to be useful in a subfactor context, we should have both
shaded and unshaded circles being equal to some scalar δ. This condition on a
planar algebra is called modulus δ [13]. By capping off the single vertical strand
l(1), we see this is true when T L2 = δT L with both shadings. This occurs when
is independent of v, or
δv = Xes(e)=v
µ(t(e))2/µ(v)2
µ(t(e))2 = δµ(v)2
Xes(e)=v
7
for all v. Another way of describing this situation is that the vector with com-
ponents µ(v)2 is an eigenvector for the connection matrix of the graph Γ, with
eigenvalue δ. Under these circumstances we will say that the spin vector itself
(or the BGPA) has modulus δ.
A modulus δ spin vector is necessarily locally bounded and locally finite,
√δ and emax ≤ δ2. Note that on a finite graph, there is only one
with M ≤
modulus δ spin vector (up to normalization). Its entries are the square roots of
the Perron-Frobenius eigenvector entries for the inclusion matrix, and δ is the
Perron-Frobenius eigenvalue. On an infinite locally finite graph, there may be
many modulus δ spin vectors, with different values of δ.
3. Automorphisms of BGPAs
An automorphism of a planar algebra is a graded linear map on the V ±
n 's
which commutes with the entire planar operad. If the planar algebra has an
involution, we will require the map to commute with the involution as well. In
this section we describe the automorphism group of an arbitrary BGPA.
These automorphisms of planar algebras may be viewed as a generalization
of the automorphisms of the standard invariant of a subfactor (see [18]; c.f.
[23],[12] for specific examples).
In the BGPA case, these automorphisms are
similar to those of a Jones tower of finite-dimensional C∗-algebras, computed in
[10].
Lemma 3.1. Let Γ be a bipartite graph, with path Hilbert space H as in section
2. Let U be a unitary operator on H which respects the grading and commutes
with the concatenation operator c. Then the action of U on H ±
0 is described by
a graph automorphism. Moreover, AdU leaves the BGPA PΓ invariant.
Proof. Let x = xv be a standard basis element of H +
0 , corresponding to a
vertex v. We have c(x, x) = x, so c(U (x), U (x)) = U (x) as well by the proper-
ties of U . The only elements of H +
0 which have this property are of the form
Pw∈S xw, where S is some subset of the even vertices of Γ. Unitarity of U
implies that in fact U (xv) = xw for some w. So U acts by permutation on the
even vertices of Γ, and likewise on the odd vertices by an identical argument.
Let this permutation be σ.
As in section 2, let sv be the projection onto the basis elements corresponding
to paths starting at v, and tv the projection onto paths terminating at v. The
dimension of tvsw(H +
1 ) is the number of edges between v and w, or zero if they
are not adjacent. Let n(v, w) be this number of edges.
For any path p from v to w (with corresponding basis element xp), we
have c(xv, xp, xw) = xp, implying that c(xσ(v), U (xp), xσ(w)) = U (xp). In other
words, U tvsw(H +
1 ). Moreover U ∗ also commutes with con-
catenation and respects grading, so applying the above argument to U ∗ gives
us equality of the above subspaces. This means that n(σ(v), σ(w)) = n(v, w),
and σ is a graph automorphism.
1 ) ⊂ tσ(v)sσ(w)(H +
8
The above also implies that AdU leaves the algebra generated by the s's and
t's invariant, since U svtwU ∗ = sσ(v)tσ(w). Therefore AdU leaves the commutant
of this algebra invariant as well. This commutant is precisely V ±
n , implying that
AdU acts on the planar algebra as desired.
We recall that the path reversal operator rev defined in section 2 is an
involution sending H +
1 to H −
1 . From the definition, this operator obeys
hrev(x), rev(y)i = hy, xi
1 or H −
1 .
where h·,·i is the inner product on H +
Theorem 3.1. Let Γ be a locally finite bipartite graph with locally bounded spin
vector. Let H be the path Hilbert space on Γ as above. Let U be a unitary element
of B(H) which agrees with the grading of H and commutes with the concatena-
tion operator. Assume further that the restriction of U to H ±
1 commutes with
the path reversal operator, and that the vertex permutation induced by U scales
the measure on H0.
Then AdU commutes with the entire planar operad (and adjoint) on V ±
n .
Proof. We know from lemma 3.1 that AdU acts on the planar algebra.
n , v ∈ H ±
Recall l is the left embedding operator. Let x be in V ±
n . Then since U commutes with concatentation,
1 and
w ∈ H ∓
U ∗l(x)U c(v, w) = U ∗l(x)c(U (v), U (w))
= U ∗c(U (v), xU (w)) = c(v, U ∗xU (w))
l(U ∗xU )c(v, w) = c(v, U ∗xU (w))
while
which is the same. So AdU commutes with left embedding, and with right
embedding as well by a similar argument.
Since U is unitary, AdU commutes with operator multiplication and taking
adjoint.
It remains only to show that AdU commutes with the Temperley-
Lieb diagrams, ie with the elements of the operad representing tangles with no
internal disks. This algebra is generated by the generators T L+ and T L− along
with multiplication and embedding, so we need to show that AdU fixes these
generators, i.e. that U commutes with them.
Let v and w be two adjacent vertices of Γ. Let H vw
1 be the subspace of H1
the subspace of H2 spanned by paths
1 . Take
spannned by paths from v to w, and H vw
from v to w and back to v. Let {ai} be any orthonormal basis for H vw
yvw = Pi c(ai, rev(ai)), using the reversal operator defined above. Because of
the interaction of rev with inner product described above, the inner product
of yvw with any element of the form c(v, rev(v)) is equal to the squared norm
of v regardless of which basis is chosen. Such elements span H vw
2 , so yvw is
independent of the choice of specific basis. This is true if v is odd or even.
2
9
Let σ be the permutation action of U on vertices of Γ. As in lemma 3.1, we
, and so U maps an orthonormal basis for the first
have U (H vw
vector space to one for the second. This means that U (yvw) = yσ(v)σ(w).
1 ) = H σ(v)σ(w)
1
We recall from section 2 that the T L generators in B(H ±
2 ) each leave the
closed linear span of certain vectors {xv} invariant, where v is taken from the
set of positive or negative vertices depending on the sign of the generator, and
are zero off the span of these vectors. We defined
yv = Xes(e)=v
µ(t(e))c(xe, rev(xe))
for each vertex v but we can also write
yv = Xwn(v,w)6=0
µ(w)yvw
using the notation above.
From lemma 3.1 σ is a graph automorphism of Γ. By assumption, σ scales
the spin vector by a fixed constant λ. Therefore
U (µ(w)yvw) = λ−1(µ(σ(w))yσ(v)σ(w)
Since σ is a graph automorphism, it maps the set of vertices adjacent to v to
the set of vertices adjacent to σ(v), and we have as well
λ−1µ(w)yσ(v)w
U
Xwn(w,v)6=0
µ(w)yvw
= Xwn(w,σ(v))6=0
So U maps one standard basis vector of the subspace acted on by either T L
generator to λ−1 times another such vector.
properties of σ,
We have T L(yv) = δvyv, where δv = Pes(e)=v µ(t(e))2/µ(v)2. From the
δσ(v) = Xes(e)=σ(v)
= Xes(e)=v
µ(t(e))2/µ(σ(v))2 = Xes(e)=v
λ2µ(t(e))2/λ2µ(σ(v))2 = Xes(e)=v
µ(σ(t(e)))2/µ(σ(v))2
µ(t(e))2/µ(v)2 = δv
So U leaves each eigenspace of both T L generators invariant.
This means that U commutes with the T L generators. So U commutes with
a set of generating tangles for the planar operad, and hence with the entire
operad.
Now we describe two classes of linear maps on a BGPA which meet all the
above conditions.
10
Definition 3.1. Let Γ be a bipartite graph with spin vector µ. Let κ be a
permutation of the vertices of Γ, preserving connection numbers and parity and
preserving or scaling spin. Label each n-fold multiple edge by {1, ..., n}. We may
then extend κ to the edges of Γ by asserting that it preserves this numbering.
Then κ gives rise to a permutation of the paths on Γ, and therefore a map U
on the path Hilbert space H. Then AdU is the graph automorphism operator
associated with κ.
Lemma 3.2. Graph automorphism operators are automorphisms of the BGPA.
Proof. Let AdU be a graph automorphism operator as above. It follows di-
rectly from the definition that AdU commutes with path reversal and concate-
nation, and agrees with the grading. We have also assumed that the underlying
graph automorphism is trace scaling. So all of the conditions of theorem 3.1 are
satisfied, and AdU is a planar algebra automorphism.
Definition 3.2. Let O be an element of V +
acting on H +
path basis element of H +
1 . Let O′ = rev ◦ O ◦ rev, acting on H −
1 (for i even) or H −
1 which is unitary as an operator
1 . For 1 ≤ i ≤ n, let pi be a
1 (for i odd). Then let
U (c(p1, p2, ..., pn)) = c(O(p1), O′(p2), O(p3)...)
and U extends to a unique bounded linear operator on H +
on H −
n as the extension of the map
n . Define U similarly
U (c(p1, p2, ..., pn)) = c(O′(p1), O(p2), O′(p3)...)
where the pi's are again in H +
1 or H −
1 as appropriate.
Then AdU is the multiplication operator associated with O.
If O acts nontrivially on only one of the subspaces svtw(H +
1 ), while leaving
all others fixed, then we will call it a basic multiplication operator associated to
the vertex pair {v, w}. Every multiplication operator is a product of basic mul-
tiplication operators, and basic multiplication operators associated to different
vertex pairs commute with each other. A multiplication operator is scalar if
the restriction to B(svtw(H +
1 )) for each v, w is a scalar multiple of the identity.
The scalar multiplication operators are the center of the multiplication operator
group. If the graph has no multiple edges then every multiplication operator is
scalar.
Lemma 3.3. Multiplication operators are automorphisms of the BGPA.
Proof. Let AdU be a multiplication operator as above. From the definition,
it commutes with concatenation and path reversal and respects grading. Since
O comes from an element of V +
1 , it commutes with sv and tv for vertices v.
Therefore the associated graph automorphism σ is trivial, and preserves the
trace. So all the conditions of theorem 3.1 are satisfied, and AdU is a planar
algebra isomorphism.
11
Direct computation shows that conjugating a basic multiplication operator
by a graph automorphism operator produces a basic multiplication operator
associated to different vertex pair. So the group of automorphisms generated
by these two types of operators has a crossed product structure: the subgroup
of multiplication operators is normal.
Now we will show that the two types of operators described above in fact
generate the entire automorphism group of a BGPA.
Lemma 3.4. Any automorphism of a BGPA is strongly continuous.
Proof. Let α be an automorphism of a BGPA.
n ⊂ B(H ±
n ), where V ±
From the definition in section 3.2, we have V ±
n is a
(potentially infinite) direct sum of type I factors. Every automorphism of such
a von Neumann algebra can be written AdU , where U is a unitary element of
B(H ±
commutes with multiplication and involution, this restric-
tion of α is a von Neumann algebra isomorphism, and may be written AdU ±
n
for unitary U ∈ B(H ±
n s provides a unitary operator on
the total Hilbert space H whose adjoint action on the graded vector space V ±
n
agrees with α.
n ). Summing all the U ±
n ). Since αV ±
n
Multiplication is strongly continuous.
This means that two automorphisms of a BGPA are equal if and only if they
agree on loops, since the loops span a strongly dense set in each V ±
n .
Lemma 3.5. Let α be an automorphism of a BGPA. Then there is a graph
automorphism operator β such that α and β agree on V ±
0 .
0 , V −
Proof. Let pv, qw be the atomic projections in V +
0 associated with even
and odd vertices v and w. Since α is a BGPA automorphism, it must send pv
to some other atomic projection pσ(v) in V +
0 , and likewise for qw. Therefore α
induces a permutation σ on the vertices of Γ.
To see that this permutation is a graph automorphism, note that l(pv)r(qw)
is a minimal central projection of V +
1 (or zero), and the dimension of l(pv)r(qw)V +
1
is the square of the number of edges between v and w. This dimension is pre-
served by α, i.e.
1 . There-
fore (v, w) and (σ(v), σ(w)) have the same number of edges between them, i.e.
n(v, w) = n(σ(v), σ(w)).
it agrees with the dimension of l(pσ(v))r(qσ(w))V +
Next we must show that σ preserves or scales the trace. For this we note
that pv with a circle around it evaluates to
qw
µ(v)
µ(w)
Xwn(v,w)6=0
Since α commutes with the tangle, we must have
µ(v)/µ(w) = µ(σ(v))µ(σ(w))
for all adjacent v, w, implying the desired result.
12
From the above description of graph automorphism operators, there is a
such an operator β whose induced permutation action on the vertices of Γ is
the same as α's. These two operators then agree on V ±
0 .
Lemma 3.6. Let α be an automorphism of a BGPA which acts trivially on V ±
0 .
Then there is a multiplication operator β such that β and α agree on V ±
1 .
Proof. Since α acts trivially on V ±
0 , it fixes all elements of the form r(pv)l(qw)
in V +
1 . These elements are the center of V +
taken as a von Neumann algebra,
1
so α acts as an inner automorphism on V +
1 . There is a multiplication operator
β whose action on V +
is any desired inner automorphism. Then α and β agree
1
on V +
1 . Both of these automorphisms commute with the half rotation
A
which is a bijective map from V +
1
to V −
1 , so they agree on V −
1 as well.
Lemma 3.7. Let α be an automorphism of a BGPA which acts trivially on V ±
1 .
Then α is a scalar multiplication operator.
Proof. From the proof of lemma 3.4, we can write the action of α on each V ±
n
as AdU ±
n is a unitary in B(H ±
n , where U ±
n ).
First note that for any path p on the graph Γ, with corresponding basis
n . This
n , there is a rank one projection onto xp contained in V ±
vector xp ∈ H ±
projection may be written as the product
rn(pe1 )rn−1l(pe2 )...rln−1(pen−1)ln(pen )
1 or V −
where ei is the ith edge of p and pei is the rank one projection in V +
1
onto the vector xei corresponding to the path ei. Since α acts trivially on V ±
1 ,
it fixes this projection, and U ±
n must therefore map each xp to a scalar multiple
of itself.
Now let bl ∈ V ±
n be a rank one partial isometry corresponding to a loop
n . We have bl = xpblxq for certain rank one projections xp, xq as above,
l ∈ l±
implying that
α(bl) = α(xpblxq) = xpα(bl)xq
The only way this can be true is if α sends bl to a scalar multiple of itself as
well.
In other words, every loop is an eigenvector of α, and α induces a map ρ
from the set of loops to the complex scalars of modulus 1.
Since α commutes with the half rotation, ρ(l) is independent of the basepoint
1 , ρ(l) = 1 for any loop l of length 2. Finally, since α
of l. Because α fixes V ±
commutes with this diagram
13
bl1
bl2
if l3 is the concatenation of the loops l1 and l2 then we have ρ(l3) = ρ(l1)ρ(l2).
Putting this together we find that ρ is necessarily a 1-dimensional repre-
sentation of the fundamental group of Γ. But any such representation may
be obtained from a scalar multiplication operator. We may always find a set
of free generators {l1, l2, ...} for the fundamental group with the property that
each generator li contains some edge ei which does not appear in any other
loop. Then a basic scalar multiplication operator with value λ associated to the
endpoints of some ek corresponds to the representation of π1(Γ) sending lk to
λ (or possible λ, depending on the direction of lk) and all other generators to
1. All other representations of π1 may be obtained similarly from basic scalar
multiplication operators associated to various ei's.
This implies that there is a scalar multiplication operator which agrees with
α on loops. Since the loops span a strongly dense subset of each V ±
n , and
automorphisms of a BGPA are strongly continuous by lemma 3.4, it follows
that α itself is a scalar multiplication operator.
Since scalar multiplication operators themselves act trivially on V ±
1 , this
lemma in fact shows that the scalar multiplication operators of a BGPA are
isomorphic to the 1-dimensional representations of the fundamental group of
the graph.
Theorem 3.2. Let PΓ be a BGPA, with multiplication operators E and graph
automorphism operators A. Let α be an automorphism of PΓ. Then α = ae for
some a ∈ A, e ∈ E.
Proof. This follows from the proceeding lemmas. There is β1 ∈ A such that
β−1
1 α acts trivially on V ±
1 α acts trivially on
V ±
1 . So β−1
1 α is a scalar multiplication operator β3, and α = β1β2β3 with
β1 ∈ A and β2β3 ∈ E.
0 . There is β2 ∈ E such that β−1
2 β−1
2 β−1
Since conjugation by graph automorphisms leaves the multiplication oper-
ator group invariant, we may in fact write AutPΓ = E ⋊ A with notation as
above.
4. Planar fixed point subfactors
A subfactor planar algebra is a planar algebra with the following additional
properties [13]:
• dimV ±
• dimV ±
0 = 1
n < ∞ ∀n
14
• Spherical: Since V +
the scalars. A planar algebra is spherical if for any A ∈ V +
right caps LC(A) and RC(A) agree.
0 = C, we may equate these vector spaces with
1 , the left and
0 = V −
• Involution: There is an antilinear isometry on each V ±
with tangles as reflection.
n which interacts
• Positive definiteness: Involution gives us a scalar sesquilinear form, namely
hx, yi = RCn(y∗x) where the multiplication and right capping tangles are
as in section 2. This form should be positive definite.
described as a subfactor planar algebra ([13],[22]). We take V +
V −
n = M ′
The standard invariant of any finite index extremal II1 subfactor may be
0 ∩ Mn,
Conversely, if P is a subfactor planar algebra, then there exists a finite index
extremal II1 subfactor such that the standard invariant of this subfactor is P
([22], c.f. [13],[11], [16],[15]).
1 ∩ Mn+1, and the operad definition is given in [13].
n = M ′
For finite graphs with the correct spin vector, any sufficiently small planar
subalgebra is of subfactor type.
Lemma 4.1. Let PΓ be a bipartite graph planar algebra with spin vector µ. Let
x be in V +
0 obtained by capping off to the
left, and xr ∈ V +
from capping off to the right. Suppose both xl and xr are
scalars. Then there is some constant α, independent of x, such that xl = αxr.
1 ; let xl represent the element of V −
0
Proof. From [14], there is a partition function defined on V +
0 and V0− as the
linear extension of xv → µ4(v), and this function has the same value on xl and
xr.
If xl and xr are scalars, this means xl = λ1Pv xv, xr = λ2Pv xv.
The partition function on x1 is λ1Pv µ4(xv), and on xr is λ2Pv µ4(xv).
Since these are the same we must have λ1 = λ2
α = Pv µ4(xv )
Pv µ4(xv ) and xl = αxr for all x such that xl and xr are scalars.
Pv µ4(xv)
Pv µ4(xv) .
In other words
0 = C. Then X is a subfactor planar algebra.
Theorem 4.1. Let PΓ be a finite bipartite graph planar algebra whose spin
vector µ has modulus δ. Let X be a planar ∗-subalgebra of PΓ such that X∩V +
0 =
X ∩ V −
Proof. The BGPA has an involution, giving rise to a positive definite V +
0 -
valued sesquilinear form (see section 2). The involution and form are inherited
by X, and the restriction of the form to X is scalar valued since X ∩ V ±
0 = C.
Since Γ is finite, each V ±
n is finite dimensional, and so this is true of their
intersections with X as well.
To show that X is of subfactor type, it remains only to demonstrate spher-
icality. Let x ∈ X be an element of V +
1 . Capping off to left or right produces
scalars, since X ∩ V ±
is scalar. Therefore the conditions of the lemma above
are satisfied. Because µ has modulus δ, shaded and unshaded circles represent
the same scalar. These diagrams are the left and right caps of a single vertical
0
15
strand, so the constant α in the lemma is equal to 1. It follows that xl and xr
are equal as scalars, and X is spherical.
A small subalgebra of an infinite BGPA still corresponds to a subfactor, but
the subfactor need not be extremal.
0 = C. Then X ∩ V ±
Theorem 4.2. Let PΓ be a locally finite bipartite graph planar algebra. Let
X ⊂ PΓ be a spherical planar ∗-subalgebra with X ∩ V ±
n is
finite dimensional.
Proof. Let pv ∈ V +
0 be a minimal projection corresponding to some even
n , and let RCn be the diagram con-
vertex v, and q = rn(pv). Take x ∈ X ∩ V +
sisting of capping off all strands to the right (this is the form from section 2).
Then RCn(x) is a scalar from the properties of X, and RCn(qx) = qRCn(x) by
isotopy invariance. q commutes with x and x∗ with respect to the usual multipli-
cation tangle. If qx = 0, then qx∗x = 0, giving RCn(qx∗x) = 0 = qRCn(x∗x).
Since RCn(x∗x) is a scalar, this means that px = 0 implies RCn(x∗x) = 0. But
RCn gives a positive definite form on V +
n , so x itself is zero in this case.
This means that the map x → qx is injective on X ∩ V +
n has basis
labelled by loops of length 2n which start and end at v. By local finiteness of
Γ, this set is finite. Therefore pV +
n is as well.
n is finite dimensional, and X ∩ V +
n . But qV +
n is finite dimensional.
The same argument shows that X ∩ V −
Burns described a class of rigid planar C∗-algebras in [9], generalizing Jones'
definition of a SPA. Every rigid planar C∗-algebra is the standard invariant of
a finite index II1 subfactor, but this subfactor need not be extremal. From
Burns' definition, a planar algebra having all the characteristics of a SPA except
sphericality is a rigid planar C∗-algebra.
Let PΓ be a modulus δ BGPA coming from an infinite finite graph, and
X ⊂ PΓ a planar subalgebra with dim(X ∩ V ±
0 ) = 1. The above lemma tells us
that X∩V ±
n is finite dimensional. The BGPA has a positive definite sesquilinear
form and an involution with the right properties, which are inherited by X. This
means that X is a rigid planar C∗-algebra. If X is additionally spherical, then
it is an SPA. This is automatic when X is irreducible, but is not true in general.
This means that there is a finite index II1 subfactor whose standard invariant
is X. The subfactor will be extremal if and only if X is spherical.
5. Examples
5.1. Introduction
In order for the fixed points P G
We can construct a wide range of subfactor planar algebras under this fixed
Γ to be an SPA, we need it
point technique.
to have 1-dimensional intersection with V +
0 , which is equivalent to
having the graph automorphism part of G act transitively on both positive and
negative vertices. We should also check sphericality for non-irreducible infinite
graph examples. Some SPAS thus are the standard invariants of previously
known subfactors, while others seem to be previously unclassified. A few such
examples are described below.
0 and V −
16
Definition 5.1. If PΓ is a BGPA, and G ⊂ AutPΓ with P G
corresponding subfactor is a planar fixed point subfactor.
Γ an SPA, then the
5.2. Group-subgroup subfactors
A specific example of the planar fixed point construction is found in [12]. In
this paper, Gupta starts with the BGPA on the graph with n odd vertices all
connected to one even vertex. Then G is some group acting by permutation on
the set of odd vertices, and H is the subgroup which fixes some specified vertex.
Gupta shows that the fixed points of the BGPA by this action constitute a
subfactor planar algebra, and that this is in fact the standard invariant of the
group-subgroup subfactor (see e.g. [17]) corresponding to the inclusion H ⊂ G,
namely M G ⊂ M H for some outer action of some finite group G on a II1 factor
M .
5.3. Wassermann subfactors
Here we let Γ be the graph with two vertices connected by an n-fold multiple
edge. Let G be any compact subgroup of the unitaries Mn(C). Then G may
be embedded in the multiplication operators on this graph. The fixed points of
this G-action are of subfactor type. They are identical to the standard invariant
of the Wasserman subfactor
(1 ⊗ Mn ⊗ Mn ⊗ ...
st
)G ⊂ (Mn ⊗ Mn ⊗ Mn ⊗ ...
st
)G
where G acts pointwise on the tensor products (see [24]).
5.4. Diagonal subfactors
Let G be a finitely generated group of outer automorphisms of a II1 factor
M . Let G have distinct generators {g1, ..., gn}, and take Γ to be the graph which
has one odd and one even vertex for each element of G. Two vertices v+
x and
v−
y are connected by a single edge if y = xh, for h ∈ {1, g1, ..., gn}. The spin
vector of this graph is 1 at every vertex; it has modulus n + 1.
g ) = vxg. Associating
each such graph automorphism with the corresponding graph automorphism
operator on PΓ gives an action of G on PΓ.
Then G acts on the graph by left translation: αx(v±
sphericality may be directly verified: V +
projection has left and right trace 1/(n + 1). Therefore P G
This action is transitive on even and odd vertices. When the graph is infinite,
1 has dimension n+ 1, and each minimal
Γ is of subfactor type.
Since vertices are labelled by group elements, loops of length 2n in this graph
may be written as a list of vertices
x − xa1 − xa1a−1
2 xa1a−1
2 a3... − xa1a−1
2 ...a−1
2n = x
where each ai comes from the set {1, g1, ..., gn}. So we may think of this loop as
a starting point x along with a list of generators (a1, a2, ...) whose alternating
product is the identity in G.
17
The group action moves the base point (via left translation) while keeping
the generator list invariant. It follows that these generator lists label a basis for
the intersection of P G
Γ with V ±
n .
This basis is precisely that described in [3] for the planar algebra of the
diagonal subfactor (see [21],[2]). It may be verified that the operations of left
and right embedding, involution, and multiplication on P G
Γ agree with the planar
algebra of [3], and the Temperley-Lieb algebra embeds in the same way, so these
planar algebras are isomorphic. It follows that the planar algebra constructed
in this way is that of a diagonal subfactor without cocycle.
5.5. Bisch-Haagerup subfactors
Let G be a group of outer automorphisms of a II1 factor M , generated by
finite subgroups H and K. For simplicity we require H ∩ K = {1}.
Let Γ be the graph which has one even vertex for each H right coset in G,
and one odd vertex for each K right coset. Then the edges of Γ are labelled by
group elements. The endpoints of each edge eg are the even vertex vgH and the
odd vertex vgK . The spin vector has value H on each even vertex and value K
We may write a loop of length 2n as a list of edges:
on each odd vertex, and it has modulus pHK.
ex1 − ex2... − ex2n − e1
Assume first that the loop begins at a positive vertex. For the path to be
connected, ex1 and ex2 must share a vertex, so x1 and x2 are in the same K-
coset and x2 = x1k1. Likewise x2 and x3 are in the same H-coset. So we can
view this loop as a starting point x along with a list of alternating elements of
k and h, with the restriction that k1h1k2h2...knhn is equal to the identity in G.
Again G acts on the graph by left translation, and this gives a G-action
on PΓ. This action shifts the basepoint of loops while leaving the list of ki's
and hi's the same. So we may identify a basis for P G
n , namely n-tuples of
elements of K and H obeying k1h1k2h2...knhn = 1; the basis for V −
n is obtained
by reversing the roles of H and K. This labelling set is the same as that of the
planar algebra of [4] (c.f.
[5]); again it may be shown that the two planar
algebras are isomorphic. It follows that this fixed point planar algebra is the
standard invariant of the Bisch-Haagerup subfactor M H ⊂ M ⋊ K.
5.6. The cube graph
Γ ∩ V +
Let Γ be the graph of a cube, and PΓ the corresponding BGPA. We obtain
several subfactor planar algebras by taking fixed points under various group
actions.
We note that the automorphism group of the bipartite graph is S4, since each
such automorphism may be described uniquely as a certain permutation of the
even vertices. The only modulus δ spin vector assigns weight 3 to every vertex;
δ = 3, so every subfactor planar subalgebra produces an index 9 subfactor. We
can write down a biprojection which is invariant under every automorphism, so
there is always an index 3 intermediate subfactor.
18
generated by (12)(34) and (13)(24), then P G
subfactor; the group generators are each order 2, and the group is Z 2
A4, then P G
order-3 automorphisms of M together generate the group A4.
We mention some of the possibilities described above. If G is the subgroup
Γ is the planar algebra of the diagonal
2 . If G is
Γ is the Bisch-Haagerup subfactor M Z3 ⊂ M ⋊ Z3, where the two
Its
Γ is some other subfactor planar algebra.
If we take G = S4, then P G
principal graph may be directly computed from the group action:
∗
The dual principal graph is the same.
It does not appear to be of any
previously categorized type, although we have tentatively identified it as a com-
position of two group-subgroup subfactors.
Finally, we may take G = AutPΓ. Since the multiplication operator group
of PΓ is (S1)5, this group is infinite, and P G
Γ is infinite depth.
5.7. The degree (3,2) tree graph
Let Γ be the graph which branches twice at each even vertex and three
times at each odd vertex. The operator norm of this graph is 4√2. There are
no multiplication operators on PΓ, so all automorphisms will come from graph
automorphisms.
First assign weight 2 to every even vertex and 3 to every odd vertex. This
spin vector has modulus √6. Every automorphism, of Γ here gives rise to
an automorphism of PΓ. Taking G = Aut(PΓ), we obtain a subfactor planar
algebra P G
Γ . This subfactor is irreducible, hence automatically spherical, and
is non-amenable; we conjecture it is obtained as a composition of two group-
subgroup subfactors, where the groups involved are variations of the Grigorchuk
lamplighter group.
We may also obtain a transitive subgroup as follows: color the edges of
the graph with three colors so that no vertex contacts two edges of the same
color, and then consider all automorphisms which leave the coloring invariant
or permute the colors. This group is Z3 ∗ Z2, and the resulting subfactor is the
index 6 Bisch-Haagerup subfactor described in [5] corresponding to this group.
Now we describe a non-irreducible example. We color the edges of the graph
red and blue so that each even vertex contacts a red and blue edge, and each
odd vertex contacts two reds and a blue. We now consider G to be the group
of color preserving automorphisms. It may be seen that this group is transitive
on odd and even vertices, but the fixed points have 2-dimensional intersection
with V +
space, corresponding to the 2 G-orbits of edges. With the above spin
1
vector, sphericality fails: the 'blue edge' element has left and right traces of 1/3
and 1/2.
To find a subfactor planar subalgebra of this BGPA, we will need a different
spin vector. We take µ(v0) = 1 for some arbitary base vertex v0, and then define
19
b
b
b
b
b
c
b
c
b
c
b
c
µ elsewhere so that for adjacent vertices x even and y odd, we have µ(y) = µ(x)
if the x − y edge is red, but µ(y) = 21/4µ(x) if the edge is blue. This trace has
modulus 1 + √2. It may be seen that any color preserving automorphism of Γ
will multiply µ by a constant factor, hence providing an automorphism of PΓ.
With this choice of spin vector and G the color preserving graph automorphisms,
the left and right traces of the 'blue' element are both √2 − 1, and sphericality
holds. So from section 5.4, P G
√2)2 = 3 + 2√2, and it does not have any intermediate subfactors.
Γ is a subfactor planar algebra.
The resulting subfactor is infinite depth non-amenable.
Its index is (1 +
From [20], this is the minimum index for an extremal non-irreducible sub-
factor. The above construction is a new way of getting such a subfactor. Any
group which is transitive on even and odd vertices but the same two orbits of
edges as above will also produce a subfactor with this list of properties, so we
actually have many such examples. For example, we might partition the red
edges into 'red' and 'white' so that each odd vertex contacts one vertex of each
color; G′ might then be the group of graph automorphisms which either preserve
color or swap the colors red and white.
There are also spin vectors of any modulus greater than 3 + 2√2 on this
graph which are preserved or scaled by every element of G. Taking fixed points
then gives us a non-extremal subfactor, as in [9]. So we obtain a continuous
family of non-extremal subfactors with the same principal graphs, but different
indices.
Other non-irreducible subfactors with index (1 + √3)2 = 4 + 2√3 may be
constructed similarly from a graph which branches twice at each even vertex
and four times at each odd vertex. This is the third on the list from [20].
Acknowledgement. I would like to thank Professor Dietmar Bisch for his useful
suggestions and corrections.
References
[1] Stephen Bigelow, Scott Morrison, Emily Peters and Noah Snyder. Con-
structing the extended Haagerup planar algebra. arxiv.org, arXiv 0909.4099
[math.OA]:45 pp., 2009.
[2] Dietmar Bisch. Entropy of groups and subfactors J. Funct. Anal. 103
(1992), no. 1, 190 -- 208.
[3] Dietmar Bisch, Paramita Das and Shamindra K. Ghosh. The planar algebra
of diagonal subfactors. Clay Mathematics Proceedings, Vol. 10 (2008), 25
pp.
[4] Dietmar Bisch, Paramita Das and Shamindra K. Ghosh. The planar algebra
of group-type subfactors. J. Funct. Anal. 257 (2009), no. 1, 20 -- 46.
[5] Dietmar Bisch and Uffe Haagerup. Composition of subfactors: new exam-
ples of infinite depth subfactors. Ann. Sci. ´Ecole Norm. Sup. (4), 29(3):329 --
383, 1996.
20
[6] Dietmar Bisch and Vaughan F. R Jones. Algebras associated to interme-
diate subfactors Invent. Math. 128 (1997), no. 1, 89 -- 157.
[7] Dietmar Bisch and Vaughan F. R. Jones. Singly generated planar algebras
of small dimension. Duke Math. J. 101 (2000), no. 1, 41 -- 75.
[8] Dietmar Bisch and Vaughan F. R Jones. Singly generated planar algebras
of small dimension, II. Adv. Math. 175 (2003), no. 2, 297 -- 318.
[9] Michael Burns. Subfactors, planar algebras and rotations. PhD dissertation,
University of California, Berkeley, Department of Mathematics, 2003.
[10] Richard D. Burstein. Group-type subfactors and Hadamard matrices. To
appear, Trans. Amer. Math. Soc.
[11] Alice Guionnet, Vaughan F. R. Jones and Dimitri Shlyakhtenko. Ran-
dom matrices, free probability, planar algebras and subfactors. arXiv.org,
arXiv:0712.2904:31 pp., 2008.
[12] Ved P. Gupta. Planar algebra of the subgroup-subfactor. Proc. Indian
Acad. Sci. Math. Sci. 118 (2008), no. 4, 583 -- 612.
[13] Vaughan F. R. Jones. Planar algebras, I.
arXiv.org, arXiv:9909027
[math.OA]:122 pp, 1999. To appear, New Zealand J. Math.
[14] Vaughan F. R. Jones. The planar algebra of a bipartite graph. in Knots in
Hellas '98 (Delphi), 94 -- 117, World Sci. Publ., River Edge, NJ.
[15] Vaughan F. R. Jones, Dimitri Shlyakhtenko, Kevin Walker. An orthogonal
approach to the subfactor of a planar algebra. arXiv.org, arXiv:0806.4146
[math.OA]: 12pp., 2008.
[16] Vijay Kodiyalam and V. S. Sunder. From subfactor planar algebras to
subfactors Internat. J. Math. 20 (2009), no. 10, 1207 -- 1231.
[17] Hideki Kosaki and Shigeru Yamagami.
Irreducible bimodules associated
with crossed product algebras. Internat. J. Math. 3 (1992), no. 5, 661 -- 676.
MR1189679
[18] Phan H. Loi. On automorphisms of subfactors. J. Funct. Anal. 141 (1996),
no. 2, 275 -- 293.
[19] Emily Peters. A planar algebra construction of the Haagerup subfactor.
arXiv.org, arXiv:0808.0764 [math.OA]: 57pp., 2008. To appear, Int. J. of
Math.
[20] Mihai Pimsner and Sorin Popa. Entropy and index for subfactors. Ann.
Sci. ´Ecole Norm. Sup. (4) 19 (1986), no. 1, 57 -- 106.
[21] Sorin Popa. Sousfacteurs, actions des groupes et cohomologie. C. R. Acad.
Sci. Paris S´er. I Math. 309 (1989), no. 12, 771 -- 776.
21
[22] Sorin Popa. An axiomatization of the lattice of higher relative commutants
of a subfactor, Invent. Math. 120 (1995), no. 3, 427 -- 445.
[23] Anna L. Svendsen. Automorphisms of subfactors from commuting squares.
Trans. Amer. Math. Soc. 356 (2004), no. 6, 2515 -- 2543 (electronic).
[24] Antony Wassermann. Coactions and Yang-Baxter equations for ergodic
actions and subfactors. in Operator algebras and applications, Vol. 2, 203 --
236, Cambridge Univ. Press, Cambridge.
22
|
0906.1401 | 2 | 0906 | 2010-02-22T01:04:17 | Nonseparable UHF algebras I: Dixmier's problem | [
"math.OA",
"math.LO"
] | There are three natural ways to define UHF (uniformly hyperfinite) C*-algebras, and all three definitions are equivalent for separable algebras. In 1967 Dixmier asked whether the three definitions remain equivalent for not necessarily separable algebras. We give a complete answer to this question. More precisely, we show that in small cardinality two definitions remain equivalent, and give counterexamples in other cases. Our results do not use any additional set-theoretic axioms beyond the usual axioms, namely ZFC. | math.OA | math |
NONSEPARABLE UHF ALGEBRAS I: DIXMIER'S
PROBLEM
ILIJAS FARAH AND TAKESHI KATSURA
Abstract. There are three natural ways to define UHF (uniformly hy-
perfinite) C*-algebras, and all three definitions are equivalent for separa-
ble algebras. In 1967 Dixmier asked whether the three definitions remain
equivalent for not necessarily separable algebras. We give a complete
answer to this question. More precisely, we show that in small cardi-
nality two definitions remain equivalent, and give counterexamples in
other cases. Our results do not use any additional set-theoretic axioms
beyond the usual axioms, namely ZFC.
1. Introduction
Let A be a C*-algebra and let ε be a positive number. For an element x
of A and a subset F of A, we write x ∈ε F if there exists y ∈ F such that
kx − yk < ε. For two subsets F,G of A, we write F ⊆ε G if x ∈ε G for all
x ∈ F. For each n ∈ N, we denote by Mn(C) the unital C*-algebra of all
n × n matrices with complex entries. A C*-algebra which is isomorphic to
Mn(C) for some n ∈ N is called a full matrix algebra.
Definition 1.1. A C*-algebra A is said to be
of full matrix algebras.
matrix subalgebras with dense union.
• uniformly hyperfinite (or UHF ) if A is isomorphic to a tensor product
• approximately matricial (or AM ) if it has a directed family of full
• locally matricial (or LM ) if for any finite subset F of A and any
ε > 0, there exists a full matrix subalgebra M of A with F ⊆ε M ,
For a definition of tensor products, see Definition 2.16. The property
LM was called matroid in [6, Definition 1.1]. A UHF algebra is unital
by definition, and it is easy to see that UHF implies AM and that AM
implies LM. In [11, Theorem 1.13], Glimm shows that a unital separable
LM algebra is UHF (see also [6, Remark 1.3 and Theorem 1.6]). Thus for
separable C*-algebras, the three conditions UHF, unital AM and unital LM
coincide. Dixmier asked whether these three conditions coincide for general
C*-algebras in [6, Problem 8.1]. We show that this is not the case. To state
our results precisely, we need the following notion.
Date: November 21, 2018.
1991 Mathematics Subject Classification. 46L05, 03E75.
1
2
ILIJAS FARAH AND TAKESHI KATSURA
Definition 1.2. The character density χ(A) of a C*-algebra A is the small-
est cardinality of a dense subset of A.
Hence A is separable if and only if its character density χ(A) is the first
infinite cardinal ℵ0. Note that χ(A) is equal to the smallest cardinality of
an infinite generating subset of A.
The following are our main results which completely answer [6, Prob-
lem 8.1]. Note that ℵ1 is the smallest uncountable cardinal.
Theorem 1.3.
AM and LM are equivalent.
(1) For a C*-algebra with character density at most ℵ1,
(2) For every cardinal κ > ℵ1, there exists a unital LM algebra with
(3) For every cardinal κ ≥ ℵ1, there exists a unital AM algebra with
character density κ which is not AM.
character density κ which is not UHF.
Proof.
(1) Follows from Proposition 5.2 and Proposition 5.6.
(2) Follows from Proposition 6.10 and Proposition 6.12.
(3) Follows from Proposition 4.5.
(cid:3)
In (3), we can also control the representation density (defined in Defi-
nition 7.1) of the example (Theorem 7.17).
In particular, we distinguish
between AM algebras and UHF algebras faithfully represented on a separa-
ble Hilbert space.
Results similar to (1) and (2) hold for approximately finite-dimensional
(AF) algebras.
Definition 1.4. A C*-algebra A is said to be
of finite-dimensional subalgebras with dense union.
• approximately finite-dimensional (or AF ) if it has a directed family
• locally finite-dimensional (or LF ) if for any finite subset F of A and
any ε > 0, there exists a finite-dimensional subalgebra D of A with
F ⊆ε D.
It is easy to see that AF implies LF. In [3, Theorem 2.2] Bratteli proved
that for a separable C*-algebra, AF and LF are equivalent. We get the
following.
Theorem 1.5.
AF and LF are equivalent.
(1)' For a C*-algebra with character density at most ℵ1,
(2)' For every cardinal κ > ℵ1, there exists an LF algebra with character
density κ which is not AF.
Proof.
(1)' Follows from Proposition 5.6.
(2)' Follows from Proposition 6.10 and Proposition 6.12.
(cid:3)
A C*-algebra is AM (resp. AF) if and only if it is obtained as a direct
limit of full matrix algebras (resp. finite-dimensional algebras) over a general
directed set (not necessarily a sequence). On the other hand, it is not hard
NONSEPARABLE UHF ALGEBRAS I: DIXMIER'S PROBLEM
3
to see that a C*-algebra is LM (resp. LF) if and only if it is obtained as
a direct limit of (separable) AM (resp. AF) algebras (Lemma 2.13). Hence
the two theorems above imply the following.
Corollary 1.6. The classes of AM algebras and AF algebras are not closed
under taking direct limits.
Some of the results of the present paper were announced in [14]. By
extending our methods the first author constructed an AM algebra that has
faithful irreducible representations both on a separable Hilbert space and on
a nonseparable Hilbert space ([8]). In the sequel to this paper [10] we show
that the classification problems for UHF and AM algebras are significantly
different.
Organization of the paper. In §2 we set up the toolbox used in the paper.
In §3 we use the Jiang -- Su algebra to distinguish LM algebras from UHF
algebras. σ-complete directed systems are used in §4 to distinguish between
AM and UHF algebras. The relation between AM and LM algebras as well
as the one between AF and LF algebras are explained in §5 and §6. In §7 we
introduce the representation density, and using it distinguish between AM
algebras and UHF algebras faithfully represented on a given Hilbert space.
2. Preliminary
In the present section we fix the terminology and prove some standard
facts from set theory, σ-complete directed systems and tensor products (re-
spectively).
2.1. Set theory. By X ∐ Y we denote the disjoint union of sets X and Y .
If f : X → Y and Z ⊆ X then we write f [Z] = {f (z) : z ∈ Z} instead of
the notation f (Z) commonly accepted outside of set theory. Let us denote
the cardinality of a set X by X. The countable infinite cardinal and the
smallest uncountable cardinal are denoted by ℵ0 and ℵ1, respectively. The
smallest uncountable ordinal is denoted by ω1.
Lemma 2.1. Let X be a set. For each x ∈ X, choose a countable subset
Yx ⊆ X with x ∈ Yx. If X > ℵ1 then one can find two elements x, y ∈ X
such that x /∈ Yy and y /∈ Yx.
Proof. Take Z ⊆ X with Z = ℵ1. Choose x ∈ X \Sz∈Z Yz and y ∈ Z \ Yx.
Then x and y are as required.
(cid:3)
Remark 2.2. The conclusion may be false if X ≤ ℵ1. To see this consider
X = ω1 and Yx = {y ∈ ω1 : y ≤ x} for x ∈ ω1.
Definition 2.3. A directed set Λ is said to be σ-complete if every countable
directed Z ⊆ Λ has the supremum sup Z ∈ Λ.
The ordered set ω1 is σ-complete. The following is another σ-complete
directed set considered in this paper.
4
ILIJAS FARAH AND TAKESHI KATSURA
Definition 2.4. For an infinite set X, we denote by [X]ℵ0 the set of all
countable infinite subsets of X, considered as a directed set with respect to
the inclusion.
Definition 2.5. Let Λ be a σ-complete directed set. A subset Λ0 of Λ is
said to be closed if for every countable directed Z ⊆ Λ0 we have sup Z ∈ Λ0,
and cofinal if for every λ ∈ Λ there exists λ0 ∈ Λ0 such that λ (cid:22) λ0.
A closed and cofinal subset is called a club.
A club is an abbreviation of a closed and unbounded set. The condition
'unbounded' (meaning 'not having an upper bound') is equivalent to 'cofinal'
for totally ordered sets such as ω1, but is strictly weaker than 'cofinal' for
general directed sets. A widely accepted custom among set theorists is
calling closed and cofinal subsets of [X]ℵ0 closed and unbounded sets (or
clubs). Reluctantly, we continue this unfortunate abuse of terminology in
our paper. This can be justified by the fact that ω1 and [X]ℵ0 are the only
σ-complete directed sets that we will consider from the next section on.
0 be clubs of
0 be an order isomorphism. Then there exists a club Λ00
Lemma 2.6. Let Λ be a σ-complete directed set. Let Λ0 and Λ′
Λ and φ : Λ0 → Λ′
of Λ such that Λ00 ⊆ Λ0 ∩ Λ′
Proof. Set Λ00 := {λ ∈ Λ0 ∩ Λ′
0 : φ(λ) = λ}. It is easy to see that Λ00 is
closed. We will see that it is cofinal. Take λ ∈ Λ. Since Λ0 is cofinal, there
exists λ1 ∈ Λ0 with λ (cid:22) λ1. Since Λ′
0 with
1 and φ(λ1) (cid:22) λ′
λ1 (cid:22) λ′
0 for
n = 1, 2, . . . such that
0 and φ ↾Λ00= id.
0 is cofinal, there exists λ′
1. Recursively, we can find λn ∈ Λ0 and λ′
1 ∈ Λ′
n ∈ Λ′
λn (cid:22) λ′
n, φ(λn) (cid:22) λ′
n,
λ′
n (cid:22) λn+1, φ−1(λ′
n) (cid:22) λn+1.
Then
λ00 := sup{λn}∞
n=1 = sup{λ′
n}∞
n=1 ∈ Λ0 ∩ Λ′
0
satisfies φ(λ00) = λ00. Thus we have found λ00 ∈ Λ00 with λ (cid:22) λ00.
Lemma 2.7. Let X and Y be infinite sets. For a club C in [X ∐ Y ]ℵ0, there
exists a club C0 in [X]ℵ0 such that for every µ0 ∈ C0 there exists µ ∈ C with
µ0 = µ ∩ X.
Proof. This is a well-known and very useful fact. We provide a proof for the
reader's convenience.
(cid:3)
Let [X]<ℵ0 denote the set of all finite subsets of X. Since C is cofinal,
we can find an increasing map f : [X]<ℵ0 → C satisfying s ⊆ f (s) for all
s ∈ [X]<ℵ0 by induction on s. We define g : [X]ℵ0 → [X ∐ Y ]ℵ0 by g(µ0) :=
Ss⊆µ0
f (s) for µ0 ∈ [X]ℵ0 . For every µ0 ∈ [X]ℵ0 , we have µ0 ⊆ g(µ0) and
g(µ0) ∈ C because f is increasing and C is closed. We set
C0 := {µ0 ∈ [X]ℵ0 : µ0 = g(µ0) ∩ X}.
NONSEPARABLE UHF ALGEBRAS I: DIXMIER'S PROBLEM
5
Then C0 is closed because for a countable directed Z ⊆ [X]ℵ0 , we have
[µ0∈Z
g(µ0) = g(cid:16) [µ0∈Z
µ0(cid:17).
It remains to show that C0 is cofinal in [X]ℵ0 . Take λ0 ∈ [X]ℵ0 arbitrarily.
We define λ1, λ2, . . . ∈ [X]ℵ0 by λn+1 := g(λn) ∩ X for n = 0, 1, . . .. Then
{λn}∞
n=0 λn is in C0.
Thus C0 is cofinal. Therefore we get a club C0 in [X]ℵ0 as required.
(cid:3)
n=0 is an increasing sequence in [X]ℵ0 and µ0 := S∞
We note that by a well-known result of Kueker for every club C in [X]ℵ0
there exists h : [X]<ℵ0 → X such that every µ ∈ [X]ℵ0 closed under h belongs
to C.
2.2. σ-complete directed families of subalgebras. By a subalgebra of a
C*-algebra we mean a C*-subalgebra, and by a unital subalgebra of a unital
C*-algebra we mean a C*-subalgebra containing the unit of the original C*-
algebra. By a directed family {Aλ}λ∈Λ of subalgebras of a C*-algebra A, we
mean that Λ is a directed set, and λ (cid:22) µ if and only if Aλ ⊆ Aµ. Thus by
definition Λ ∋ λ 7→ Aλ is injective.
Definition 2.8. A directed family {Aλ}λ∈Λ of subalgebras of a C*-algebra
A is said to be σ-complete if Λ is σ-complete and for every countable directed
Z ⊆ Λ, Asup Z is the closure of the union of {Aλ}λ∈Z .
In other words, a directed family {Aλ}λ∈Λ is σ-complete if Sλ∈Z Aλ is in
the family for every countable directed Z ⊆ Λ.
Lemma 2.9. Let A be a C*-algebra, and let {Aλ}λ∈Λ be a σ-complete di-
rected family of subalgebras of A with dense union. Then for a club Λ0 ⊆ Λ,
the restriction {Aλ}λ∈Λ0 is also a σ-complete directed family with dense
union.
Proof. The restriction {Aλ}λ∈Λ0 is σ-complete because Λ0 is closed, and its
union is dense because Λ0 is cofinal.
(cid:3)
Lemma 2.10. Every C*-algebra A has a σ-complete directed family of sep-
arable subalgebras with dense union.
Proof. We can take the family of all separable subalgebras of A ordered by
the inclusion.
(cid:3)
Lemma 2.11. Let A be a C*-algebra, and let {Aλ}λ∈Λ be a σ-complete
directed family of subalgebras of A with dense union. For every separable
subalgebra A0 of A there exists λ ∈ Λ such that A0 ⊆ Aλ.
Proof. Let {a1, a2, . . .} be a dense sequence of A0. For each n ∈ N, one
can inductively find λn ∈ Λ such that ai ∈1/n Aλn for i = 1, 2, . . . , n and
λn−1 (cid:22) λn because the family {Aλ}λ∈Λ is directed and its union is dense in
A. Then λ := sup{λn : n ∈ N} ∈ Λ satisfies A0 ⊆ Aλ.
(cid:3)
6
ILIJAS FARAH AND TAKESHI KATSURA
By the lemma above, we can see that the union of a σ-complete directed
0 ⊆ Λ′ and an order isomorphism φ : Λ0 → Λ′
family is automatically closed.
Proposition 2.12. Let A and B be C*-algebras, and {Aλ}λ∈Λ and {Bλ′}λ′∈Λ′
be σ-complete directed families of separable subalgebras of A and B with
dense union. Let Φ : A → B be an isomorphism. Then there exist clubs
Λ0 ⊆ Λ and Λ′
0 such that
Φ[Aλ] = Bφ(λ) for all λ ∈ Λ0. If Λ = Λ′, then one can take Λ0 = Λ′
0 and
φ = id.
Proof. Let Λ0 be the set of all λ ∈ Λ such that there exists λ′ ∈ Λ′ with
0 ⊆ Λ′ as the set of all λ′ ∈ Λ′ such that
Φ[Aλ] = Bλ′. Similarly we define Λ′
there is λ ∈ Λ with Φ−1[Bλ′] = Aλ. Then there exists an order isomorphism
φ : Λ0 → Λ′
0 such that Φ[Aλ] = Bφ(λ) for all λ ∈ Λ0. We are going to show
that Λ0 ⊆ Λ is a club. It is clear that Λ0 is closed. Take λ ∈ Λ. Since Aλ is
separable, there exists λ′
by Lemma 2.11. By
the same reason, there exists λ1 ∈ Λ such that Φ−1[Bλ′
] ⊆ Aλ1. Then we
have Aλ ⊆ Aλ1. In this way, we can find sequences
Aλ ⊆ Aλ1 ⊆ Aλ2 ⊆ Aλ3 ⊆ ···
1 ∈ Λ′ such that Φ[Aλ] ⊆ Bλ′
1
1
Bλ′
1 ⊆ Bλ′
such that Bλ′
and λ′
2 ⊆ Bλ′
n ⊆ Φ[Aλn] and Φ[Aλn] ⊆ Bλ′
0 ∈ Λ′ be the supremums of {λn}∞
for n = 1, 2, . . .. Let λ0 ∈ Λ
n=1. Then we have
, we get λ0 ∈ Λ0.
0 ⊆ Λ′ is
This shows that Λ0 is cofinal, and hence it is a club. Similarly Λ′
a club. This shows the former assertion. The latter assertion follows from
Lemma 2.6.
(cid:3)
n}∞
n. Since Φ[Aλ0] = Bλ′
Aλ0 =S∞
n=1 and {λ′
=S∞
n=1 Aλn and Bλ′
0
n=1 Bλ′
3 ⊆ ···
n+1
0
Lemma 2.13. A C*-algebra A is LF if and only if it has a σ-complete
directed family of separable AF algebras with dense union.
Proof. We only need to prove the direct implication. We see that A has
a σ-complete directed family of separable subalgebras {Aλ}λ∈Λ with dense
union by Lemma 2.10. Since by [3, Theorem 2.2] every separable LF algebra
is AF, it suffices to show that the set Λ0 of all λ ∈ Λ such that Aλ is LF is a
club. Clearly Λ0 is closed. To show that Λ0 is cofinal, it suffices to see that
for any separable subalgebra A0 of A, there exists a separable subalgebra
A′
0 containing A0 such that for any finite subset F of A0 and any ε > 0,
there exists a finite-dimensional subalgebra M of A′
0 with F ⊆ε M . This is
easy to see.
(cid:3)
In the same way, one can show that a C*-algebra A is LM if and only if
it has a σ-complete directed family of separable AM subalgebras with dense
union.
Remark 2.14. Lemma 2.13 is just a special case of the downward Lowen-
heim -- Skolem theorem for logic of metric structures ([1], or [9] for a version
NONSEPARABLE UHF ALGEBRAS I: DIXMIER'S PROBLEM
7
suitable for study of C*-algebras and II1 factors). Similar arguments have
been used by C*-algebraists to reflect properties of nonseparable algebras to
separable subalgebras (see [2, II.8.5]) such as for example simplicity or the
existence of the unique trace.
2.3. Tensor products. In this subsection, we give a definition and some
properties of tensor products of C*-algebras. We try to avoid using results
on nuclear C*-algebras as much as possible.
In fact, we use the nuclear-
ity only in Proposition 2.24 (and Lemma 2.22) which is used in the proof
of Proposition 4.5 (3). We are interested in tensor products of possibly
uncountably many unital C*-algebras, and for this purpose the maximal
tensor products are easier to treat than the minimal ones. We remark that
we mainly deal with nuclear C*-algebras for which there is no distinction
between the minimal tensor products and the maximal ones.
Definition 2.15. A family {Ax}x∈X of subalgebras of a C*-algebra A is said
to mutually commute if for distinct x, y ∈ X, every element of Ax commutes
with every element of Ay.
Definition 2.16. For a family {Ax}x∈X of unital C*-algebras, its (maximal)
tensor product Nx∈X Ax is the C*-algebra having (an isomorphic copy of)
Ax as unital subalgebras for x ∈ X satisfying the following two properties:
(1) the family {Ax}x∈X of subalgebras ofNx∈X Ax mutually commutes,
and its union Sx∈X Ax generates Nx∈X Ax.
(2) for a unital C*-algebra B and a family {ϕx}x∈X of unital ∗-homomor-
phisms ϕx : Ax → B such that {ϕx[Ax]}x∈X is a mutually commuting
family of unital subalgebras of B, there exists a unital ∗-homomor-
phism ϕ : Nx∈X Ax → B such that ϕ ↾Ax= ϕx for all x ∈ X.
When Ax = A for all x ∈ X, we simply write NX A for Nx∈X Ax.
It is not difficult to see that the tensor product exists and is unique. A
nice exposition of tensor products of C*-algebras can be found e.g., in [4].
The condition (2) is called the universal property of the tensor product. A
nice exposition of universal C*-algebras can be found e.g., [2, II.8.3].
Let A and B be unital C*-algebras. Since we consider A and B as unital
subalgebras of A⊗ B, each a ∈ A and each b ∈ B are considered as elements
of A ⊗ B. Thus the product ab ∈ A ⊗ B makes sense whereas this element
is usually denoted by a ⊗ b ∈ A ⊗ B. Similarly, for a family {Ax}x∈X of
unital C*-algebras and a finite family {ax}x∈Y of elements with ax ∈ Ax for
x ∈ Y ⊆ X, we denote by Qx∈Y ax ∈ Nx∈X Ax the product of {ax}x∈Y .
Note that this product does not depend on the order of multiplications
because the family {ax}x∈Y in Nx∈X Ax mutually commutes.
The referee pointed out that the version of the next lemma when A is as-
sumed to be nuclear and simple instead of LM is true (cf. [4, Corollary 9.4.6]).
Since one can prove that LM algebras are nuclear and simple, this gives a
proof of this lemma. We give an elementary proof for the reader's conve-
nience.
8
ILIJAS FARAH AND TAKESHI KATSURA
Lemma 2.17. Let A and B be unital subalgebras of a unital C*-algebra D
commuting with each other. If A is LM, then the natural map from A⊗ B to
the C*-subalgebra C ∗(A∪B) of D generated by A∪B ⊆ D is an isomorphism.
Proof. We first show the statement in the case that A is a full matrix algebra
i,j=1 be a matrix unit of A ∼= Mn(C). Then every element
Mn(C). Let {ei,j}n
of A ⊗ B can be written as Pn
i,j=1 ei,jbi,j for bi,j ∈ B. In C ∗(A ∪ B) ⊆ D,
we have
ek,i′(cid:16) nXi,j=1
nXk=1
ei,jbi,j(cid:17)ej′,k
for i′, j′ = 1, 2, . . . , n. Hence if an element Pn
i,j=1 ei,jbi,j ∈ A ⊗ B is sent
to 0 ∈ D by the natural map A ⊗ B → D, then bi,j = 0 for all i, j which
implies Pn
i,j=1 ei,jbi,j = 0 in A ⊗ B. Thus when A is a full matrix algebra,
the natural map A⊗ B → C ∗(A∪ B) is injective, and hence an isomorphism.
Now suppose that A is LM. Let π : A⊗B → C ∗(A∪B) be the natural map.
Take x ∈ A ⊗ B. Take ε > 0 arbitrarily. Then there exist a1, a2, . . . , an ∈ A
and b1, b2, . . . , bn ∈ B such that
bi′,j′ =
(cid:13)(cid:13)(cid:13)x −
nXi=1
aibi(cid:13)(cid:13)(cid:13) < ε.
Since A is LM, we may assume (by perturbing ai's slightly if necessarily)
that a1, a2, . . . , an ∈ M for some unital full matrix subalgebra M of A. Then
by the first part of the proof, we have
(cid:13)(cid:13)(cid:13)
Hence we get
aibi(cid:17)(cid:13)(cid:13)(cid:13).
aibi(cid:13)(cid:13)(cid:13) =(cid:13)(cid:13)(cid:13)π(cid:16) nXi=1
nXi=1
(cid:12)(cid:12)kxk − kπ(x)k(cid:12)(cid:12) < 2ε.
Since ε was arbitrary, we have kxk = kπ(x)k. This shows that the natural
map π : A ⊗ B → C ∗(A ∪ B) is injective, and hence an isomorphism.
(cid:3)
We take advantage of Lemma 2.17 and use the notation A ⊗ B whenever
it is justified by this lemma. Note that this lemma is false if we replace
LM by LF. To see this, just consider A = B = D = C ⊕ C. For a family
{Ax}x∈X of unital C*-algebras, and unital subalgebras Dx ⊆ Ax, we some-
times denote by Nx∈X Dx the subalgebra of Nx∈X Ax generated by the
mutually commuting family {Dx}x∈X of unital subalgebras ofNx∈X Ax. In
fact, this unital subalgebra is the image of the ∗-homomorphism from the
tensor product Nx∈X Dx to Nx∈X Ax, but no confusion should arise.
We use the following well-known fact without mentioning. We give its
proof for the reader's convenience.
Lemma 2.18. Let A and B be unital C*-algebras, and A0 ⊆ A and B0 ⊆ B
be unital subalgebras. Then we have (A0 ⊗ B0) ∩ B = B0 in A ⊗ B.
NONSEPARABLE UHF ALGEBRAS I: DIXMIER'S PROBLEM
9
Proof. Take a state ϕ of A. Define a linear map E : A⊗ B → B by E(ab) =
ϕ(a)b for a ∈ A and b ∈ B. Since E(b) = b for b ∈ B and E(A0 ⊗ B0) ⊆ B0,
we get (A0 ⊗ B0) ∩ B ⊆ B0. The inverse inclusion is easy to see.
(cid:3)
For two families {Ax}x∈X1 and {Ax}x∈X2 of unital C*-algebras, the tensor
product Nx∈X1∐X2
Ax is naturally isomorphic to
(cid:16) Ox∈X1
Ax(cid:17) ⊗(cid:16) Ox∈X2
Ax(cid:17).
We identify these two tensor products. In particular, we can and will con-
sider Nx∈Y Ax as a unital subalgebra of Nx∈X Ax for a subset Y of X.
We use the convention that Nx∈Y Ax = C for Y = ∅. We remark that the
subalgebra Nx∈Y Ax coincides with Nx∈X Ax for a subset Y of X if and
only if Ax = C for all x ∈ X \ Y .
The following is easy to see.
Lemma 2.19. Let {Ax}x∈X be an infinite family of unital C*-algebras, and
set A =Nx∈X Ax. Then {Nx∈λ Ax}λ∈[X]ℵ0 is a σ-complete directed system
of subalgebras of A with dense union.
(cid:3)
Lemma 2.20. If A = Nx∈X Ax, X is infinite, and each Ax is separable
and not C, then the character density χ(A) of A is equal to X.
Proof. Fix a countable dense Cx ⊆ Ax for each x. Their union has cardi-
nality X and generates A. This shows χ(A) ≤ X. Take a subset Z ⊆ A
with cardinality less than X. For each z ∈ Z, there exists λz ∈ [X]ℵ0 with
Ax by Lemma 2.19. Since the set Sz∈Z λz ⊆ X has cardinality
z ∈ Nx∈λz
less than X, we can find x ∈ X outside of this set. Since Ax is not C, Z is
not dense in A. Hence χ(A) = X.
(cid:3)
For a unitary u of a unital C*-algebra A, an automorphism Ad u on A
is defined by Ad u(a) = uau∗ for a ∈ A. Let {Ax}x∈X be a family of
unital C*-algebras. By the universality, a family {αx}x∈X of automorphisms
αx on Ax determines the automorphism α on Nx∈X Ax with α ↾Ax= αx
which we denote by Nx∈X αx. For a subset Y ⊆ X and a family {αx}x∈Y
of automorphisms αx on Ax, we denote by Nx∈Y αx the automorphism
Nx∈X αx ofNx∈X Ax where αx = idAx for x ∈ X\Y . For unitaries ux ∈ Ax
for x ∈ Y , we get an automorphism Nx∈Y Ad ux on A =Nx∈X Ax. When
Y is finite, we get Nx∈Y Ad ux = Ad u where u = Qx∈Y ux ∈ A, but in
general, Nx∈Y Ad ux is not in the form Ad u for a unitary u of A.
2.4. Relative commutants. For a subset A of a C*-algebra B, we denote
by ZB(A) the relative commutant (or centralizer ) of A inside B;
ZB(A) := {b ∈ B : ab = ba for all a ∈ A}
which is a subalgebra of B if A is closed under the ∗-operation (for example
if A is a subalgebra). We avoid the common notation A′ ∩ B for ZB(A) in
order to increase the readability of certain formulas. For a subset A of a
10
ILIJAS FARAH AND TAKESHI KATSURA
If A is LM, then
C*-algebra B, we denote by C ∗(A) the subalgebra generated by A. Note
that ZB(C ∗(A)) = ZB(A) for a subset A closed under the ∗-operation. We
also note that ZB(A1 ∪ A2) = ZB(A1) ∩ ZB(A2).
Lemma 2.21. Let A and D be unital C*-algebras.
ZA⊗D(A) = D.
Proof. It is clear from the definition of tensor products that ZA⊗D(A) ⊃ D.
Take x0 ∈ ZA⊗D(A). For any ε > 0, there exist elements a1, . . . , an ∈ A
and d1, . . . , dn ∈ D such that kx0 −Pn
i=1 aidik < ε. Since A is LM, we may
assume that a1, . . . , an are in a full matrix unital subalgebra M of A. Let
E : A⊗D → A⊗D be a contractive linear map defined by E(x) =RU uxu∗du
for x ∈ A⊗D where du is the normalized Haar measure on the unitary group
U of M . Since x0 ∈ ZA⊗D(A), we have E(x0) = x0. For a ∈ M and d ∈ D,
we have E(ad) = tr(a)d where tr : M → C is the normalized trace. Hence
we have kx0 −Pn
i=1 tr(ai)bik < ε. This means that x0 ∈ε D. Since ε was
arbitrary, x0 ∈ D. Thus we get ZA⊗D(A) ⊆ D, and therefore ZA⊗D(A) = D.
We are done.
(cid:3)
By letting D = C in the lemma above, we see that the center ZA(A) of
an LM algebra A is C. Thus one can write the conclusion of Lemma 2.21 as
ZA⊗D(A) = ZA(A)⊗D. The referee pointed out that ZA⊗D(A) = ZA(A)⊗D
holds for minimal tensor products by [12, Theorem 1]. Since one can prove
that LM algebras are nuclear and satisfy ZA(A) = C, this gives an indirect
proof of Lemma 2.21.
To prove Proposition 4.5 (3), we need some facts on nuclear C*-algebras
(Lemma 2.22 and Proposition 2.24). When we apply Proposition 2.24 in the
proof of Proposition 4.5 (3), we use the fact that a UHF algebra is a tensor
product of separable nuclear C*-algebras because full matrix algebras are
nuclear. A nice exposition of nuclearity of C*-algebras can be found e.g., in
[4].
Lemma 2.22. Let A and D be unital C*-algebras, and A0 a unital subal-
gebra of A. Suppose that D is nuclear. Then ZA⊗D(A0) = ZA(A0) ⊗ D.
Proof. Clearly we have ZA(A0) ⊗ D ⊆ ZA⊗D(A0). Let
F := {c ∈ A ⊗ D : (id ⊗ω)(c) ∈ ZA(A0) for all ω ∈ D∗}.
For a ∈ A ⊆ A ⊗ D and c ∈ A ⊗ D, we have
(id⊗ω)(ac) = a(id⊗ω)(c),
(id⊗ω)(ca) = (id⊗ω)(c)a
for all ω ∈ D∗. Hence we get ZA⊗D(A0) ⊆ F . We claim that F =
ZA(A0)⊗ D. This equality is usually referred to as the slice map property of
the triple (D, A, ZA(A0)) (see [4, Definition 12.4.3]). Here we remark that
the (maximal) tensor product considered in this paper coincides with the
minimal one because D is nuclear. By [4, Theorem 12.4.4 (2)] (see [4, Defi-
nition 12.4.1] and note nuclear⇔CPAP⇒SOAP), the triple (D, A, ZA(A0))
NONSEPARABLE UHF ALGEBRAS I: DIXMIER'S PROBLEM
11
has the slice map property because D is nuclear. Thus we have ZA(A0)⊗D =
ZA⊗D(A0).
(cid:3)
Definition 2.23. Let A be a unital C*-algebra, and A0 a unital subalgebra
of A. We say that A0 is complemented in A if C ∗(A0 ∪ ZA(A0)) = A.
In a tensor product A =Nx∈X Ax of unital C*-algebras Ax, a subalgebra
AY =Nx∈Y Ax is complemented for every subset Y of X.
Proposition 2.24. Let A be a unital C*-algebra. Suppose that there exists a
unital C*-algebra D such that A⊗ D is a tensor product of separable nuclear
C*-algebras. Then for a σ-complete directed system {Aλ}λ∈Λ of separable
subalgebras of A with dense union, there exists a club Λ0 ⊆ Λ such that for
each λ ∈ Λ0, Aλ is complemented in A.
Proof. Fix a dense X ⊆ A and a dense Y ⊆ D. Then {C ∗(µ)}µ∈[X∐Y ]ℵ0 is
a σ-complete directed family of separable subalgebras of A ⊗ D with dense
union. Since A⊗ D is a tensor product of separable C*-algebras, A⊗ D has
a σ-complete directed system of separable complemented subalgebras with
dense union by Lemma 2.19. Hence by Proposition 2.12, there exists a club
C ⊆ [X ∐ Y ]ℵ0 such that C ∗(µ) is complemented in A ⊗ D for all µ ∈ C.
By Lemma 2.7 there exists a club C0 ⊆ [X]ℵ0 such that for every µ0 ∈ C0
there exists µ ∈ C with µ0 = µ ∩ X. By Lemma 2.9, {C ∗(µ0)}µ0∈C0 is a
σ-complete directed family of separable subalgebras of A with dense union.
Hence by Proposition 2.12 applied with id : A → A, we get a club Λ0 ⊆ Λ
such that for each λ ∈ Λ0 there exists µ0 ∈ C0 with Aλ = C ∗(µ0).
It remains to prove that Aλ is complemented in A for every λ ∈ Λ0.
Take λ ∈ Λ0. Then by the arguments above, there exists µ ∈ C such that
Aλ = C ∗(µ∩ X). Then C ∗(µ) is complemented in A⊗ D, and we have Aλ ⊆
C ∗(µ) ⊆ Aλ ⊗ D. Since A⊗ D is a tensor product of nuclear C*-algebras, D
is nuclear by [4, Proposition 10.1.7]. Hence we get ZA(Aλ)⊗ D = ZA⊗D(Aλ)
by Lemma 2.22. Therefore we have
A ⊗ D = C ∗(cid:0)C ∗(µ) ∪ ZA⊗D(C ∗(µ))(cid:1)
⊆ C ∗(cid:0)(Aλ ⊗ D) ∪ ZA⊗D(Aλ)(cid:1)
= C ∗(cid:0)(Aλ ⊗ D) ∪ (ZA(Aλ) ⊗ D)(cid:1)
= C ∗(Aλ ∪ ZA(Aλ)) ⊗ D.
This shows that Aλ is complemented in A and finishes the proof.
(cid:3)
3. LM but not UHF
Proposition 3.2 gives examples of unital LM algebras that are not UHF,
answering part of [6, Problem 8.1]. Recall that Z is the Jiang -- Su algebra.
We shall need the following properties of Z proved in [13]:
• Z is a unital, separable C*-algebra which is not UHF.
• Nℵ0 Z ∼= Z.
• A ⊗ Z ∼= A for any infinite-dimensional separable UHF algebra A.
12
ILIJAS FARAH AND TAKESHI KATSURA
M2(C) is called the CAR algebra.
Definition 3.1. The UHF algebra Nℵ0
Proposition 3.2. For two sets X and Y , define AX,Y := NX M2(C) ⊗
NY Z. Suppose that X is infinite. Then we have the following.
(1) AX,Y is a unital LM algebra with χ(AX,Y ) = X + Y .
(2) AX,Y is UHF if and only if X ≥ Y .
(3) AX,Y ⊗ D is UHF for any UHF algebra D with χ(D) ≥ Y .
Proof. Since X is infinite, we can identify AX,Y with NX A ⊗NY Z where
A is the CAR algebra. For each λ ∈ [X]ℵ0 and λ′ ∈ [Y ]ℵ0 we set
Dλ,λ′ :=Oλ
A ⊗Oλ′ Z ⊆ AX,Y
Then Dλ,λ′ is the CAR algebra for all λ and λ′. Since {Dλ,λ′} is a σ-complete
directed system with dense union, we see that AX,Y is LM. By Lemma 2.20,
we have χ(AX,Y ) = X + Y . This shows (1).
By rearranging the factors, we see that AX,Y is UHF if X ≥ Y and
that AX,Y ⊗ D is UHF for a UHF algebra D with χ(D) ≥ Y . It remains
to show that AX,Y is UHF only if X ≥ Y . For the sake of obtaining
a contradiction, assume that X < Y and AX,Y is UHF. Let us denote
by Ax = M2(C) for x ∈ X and Ay = Z for y ∈ Y the unital subalgebra of
AX,Y so that AX,Y =Nx∈X Ax⊗Ny∈Y Ay. Let Φ : AX,Y →Nz∈Z Mz be an
isomorphism where Z is a set and {Mz}z∈Z is a family of full matrix algebras.
For each x ∈ X, there exists a finite Fx ⊆ Z such that Φ[Ax] ⊆Nz∈Fx
Mz.
If we set Z1 = Sx∈X Fx ⊆ Z, then we get Z1 = X and Φ(cid:2)Nx∈X Ax(cid:3) ⊆
Nz∈Z1
Mz. Similarly, for each z ∈ Z1 there exists a finite Gz ⊆ Y such that
Mz ⊆ Φ(cid:2)Nx∈X Ax ⊗Ny∈Gz
Gz ⊆ Y , then we
get Y1 ≤ Z1 = X and
Oz∈Z1
Ay(cid:3). If we set Y1 = Sz∈Z1
Ayi.
Ax ⊗ Oy∈Y1
Mz ⊆ ΦhOx∈X
If we set Z2 = Z1 ∪Sy∈Y1
Ax ⊗ Oy∈Y1
ΦhOx∈X
Mz.
Next for each y ∈ Y1 there exists a countable Cy ⊆ Z such that Φ[Ay] ⊆
Nz∈Cy
Cy ⊆ Z, then we get Z1 ⊆ Z2,
Z2 = X and
Ayi ⊆ Oz∈Z2
Recursively, we can find increasing sequences {Yk}n
k=1 and {Zk}n
sets of Y and Z, respectively, such that X ∐ Yk = Zk = X and
k=1 of sub-
Mz.
Oz∈Zk
Mz ⊆ ΦhOx∈X
for k = 1, 2, . . .. We set Y ′ :=S∞
we have X ∐ Y ′ = Z ′ = X and
Ayi ⊆ Oz∈Zk+1
Ax ⊗ Oy∈Yk
k=1 Yk ⊆ Y and Z ′ :=S∞
Mz
ΦhOx∈X
Ax ⊗ Oy∈Y ′
Ayi = Oz∈Z ′
Mz.
k=1 Zk ⊆ Z. Then
NONSEPARABLE UHF ALGEBRAS I: DIXMIER'S PROBLEM
13
Since Nz∈Z ′ Mz is UHF and hence LM, we have
ZAX,Y(cid:16)Ox∈X
Ax ⊗ Oy∈Y ′
ZNz∈Z Mz(cid:16)Oz∈Z ′
Ay(cid:17) = Oy∈Y \Y ′
Mz(cid:17) = Oz∈Z\Z ′
Ay,
Mz.
by Lemma 2.21. Thus we get Φ(cid:2)Ny∈Y \Y ′ Ay(cid:3) =Nz∈Z\Z ′ Mz. Since Y ′ ≤
X < Y , we see that Y \ Y ′ is infinite. Hence Z \ Z ′ is also infinite. By
Proposition 2.12 and Lemma 2.19, there exist C ∈ [Y \ Y ′]ℵ0 and C ′ ∈
[Z \ Z ′]ℵ0 such that Φ(cid:2)Ny∈C Ay(cid:3) = Nz∈C′ Mz. This is a contradiction
because Ny∈C Ay ∼= Z is not UHF.
We thank the referee who pointed out an error of a proof of Proposi-
tion 3.2 (3) in an earlier draft. By Proposition 3.2, the unital C*-algebra
AX,Y is LM but not UHF if X < Y . When X = ℵ0 and Y = ℵ1, we
see that AX,Y is AM by Theorem 1.3 (1). In the other case, we do not know
whether AX,Y is AM or not.
Problem 3.3. Let X, Y be sets such that ℵ0 ≤ X < Y and ℵ1 < Y . Is
AX,Y =NX M2(C) ⊗NY Z AM?
(cid:3)
4. AM but not UHF
In this section, for each infinite set X we define a unital AM-algebra
BX with χ(BX ) = X, and show that BX, or even BX ⊗ D for a unital
C*-algebra D, is not UHF when X ≥ ℵ1.
Lemma 4.1. A C*-algebra generated by two self-adjoint unitaries v, w with
vw = −wv is always isomorphic to M2(C).
Proof. A C*-algebra A generated by two self-adjoint unitaries v, w with
vw = −wv is spanned (as a vector space) by 4 elements {1, v, w, vw}, and
it is noncommutative. Hence it is isomorphic to M2(C) which is the unique
noncommutative C*-algebra with dimension ≤ 4. A concrete isomorphism
from A to M2(C) is given by sending v and w to the unitaries
0
(cid:18)1
0 −1(cid:19)
and
(cid:18)0 1
1 0(cid:19)
in M2(C).
(cid:3)
Let us take a set X. For each x ∈ X, let Ax be a C*-algebra generated by
two self-adjoint unitaries vx, wx with vxwx = −wxvx. By Lemma 4.1, Ax is
isomorphic to M2(C). We define a UHF algebra AX by AX :=Nx∈X Ax ∼=
NX M2(C). We define an automorphism α on AX by α := Nx∈X Ad vx.
Note that α2 = id. Let {ei,j}2
and define an embedding
i,j=1 be a system of matrix units of M2(C),
ι : AX ∋ a 7→ ae1,1 + α(a)e2,2 ∈ AX ⊗ M2(C).
14
ILIJAS FARAH AND TAKESHI KATSURA
Let u ∈ AX ⊗ M2(C) be a self-adjoint unitary defined by u := e1,2 + e2,1.
Set BX := C ∗(ι(AX ) ∪ {u}). We consider AX as a unital subalgebra of BX
and omit ι. Then we have ua = α(a)u for a ∈ AX and BX = {au + a′ :
a, a′ ∈ AX}.
Remark 4.2. The C*-algebra BX is nothing but the crossed product AX ⋊α
(Z/2Z).
For Y ⊆ X, we denote by AY the subalgebra Nx∈Y Ax ⊆ AX ⊆ BX,
and define BY := C ∗(AY ∪ {u}) ⊆ BX . It is easy to see that AY ⊆ AX is
globally invariant under α, and hence BY = {au + a′ : a, a′ ∈ AY }.
Lemma 4.3. If Y is infinite then ZBX (AY ) = AX\Y .
Proof. Since ZAX (AY ) = AX\Y by Lemma 2.21, it suffices to show that
ZBX (AY ) ⊆ AX. Take au + a′ ∈ ZBX (AY ) with a, a′ ∈ AX, and we will
show a = 0.
For any ε > 0 there is a finite F ⊆ X such that a ∈ε AF . Since Y is
infinite, pick y ∈ Y \ F . The unitary wy ∈ AY satisfies uwy = −wyu. Hence
wy(au + a′) = (au + a′)wy yields (wya + awy)u + (wya′ − a′wy) = 0. Since
bu + b′ = 0 for b, b′ in AX implies b = b′ = 0, we have wya = −awy. Thus
we get
kak = kwyak = kwya + wyak/2 = kwya − awyk/2.
Since a ∈ε AF and wy commutes with AF , we have kak = kwya−awyk/2 < ε.
Since ε was arbitrary, a = 0. We are done.
(cid:3)
Lemma 4.4. If Y ( X and Y is infinite, then BY is not complemented
in BX .
Proof. Since BY = C ∗(AY ∪ {u}), we have
ZBX (BY ) = ZBX (AY ) ∩ ZBX ({u}) = AX\Y ∩ ZBX ({u})
by Lemma 4.3. Hence
C ∗(BY ∪ ZBX (BY )) =(cid:8)au + a′ : a, a′ ∈ AY ⊗ (AX\Y ∩ ZBX ({u}))(cid:9)
which does not contain wy ∈ AX\Y for y ∈ X \ Y .
Proposition 4.5.
(1) If X is infinite, then BX is a unital AM algebra
(cid:3)
with χ(BX) = X.
(2) If X is uncountable, then BX is not UHF.
(3) If X is uncountable, then BX ⊗ D is not UHF for any unital C*-
algebra D.
Proof. (1) Suppose X is infinite.
Let us set
and define
Λ = {(F, y) : F ⊆ X finite, and y ∈ X \ F}
D(F,y) = C ∗(cid:0)BF ∪ {wy}(cid:1) ⊆ BX .
NONSEPARABLE UHF ALGEBRAS I: DIXMIER'S PROBLEM
15
for (F, y) ∈ Λ. Then it is clear that {D(F,y)}(F,y)∈Λ is a directed family
with dense union. It remains to show that D(F,y) is a full matrix algebra for
(F, y) ∈ Λ. Take (F, y) ∈ Λ with F = n ∈ N. Then we have AF ∼= M2n(C),
and the restriction of α to AF coincides with Ad v where
v = Yx∈F
vx ∈ AF .
Then the two self-adjoint unitaries uv and wy in D(F,y) satisfy wy(uv) =
−(uv)wy and commute with AF . By Lemma 4.1, the subalgebra of D(F,y)
generated by uv and wy is isomorphic to M2(C), and commute with AF .
Since D(F,y) is generated by AF and this subalgebra, D(F,y) is isomorphic to
M2n+1(C). We are done.
(2) Suppose X is uncountable. Then {BY }Y ∈[X]ℵ0 is a σ-complete di-
rected family of separable subalgebras of BX with dense union. By Lemma 4.4,
neither one of these subalgebras is complemented. By Lemma 2.19, a UHF
algebra has a σ-complete directed system of separable complemented subal-
gebras with dense union. Hence BX cannot be UHF by Lemma 2.7.
(3) As in (2), BX has a σ-complete directed system of separable subal-
gebras with dense union neither one of which is complemented. By Propo-
sition 2.24, BX ⊗ D cannot be UHF for any unital C*-algebra D because
every UHF algebra is a tensor product of separable nuclear C*-algebras. (cid:3)
Note that an example of a unital LM algebra A that is not UHF given in
Proposition 3.2 has the property that A ⊗ D is UHF for some UHF algebra
D, but the one given in Proposition 4.5 does not have this property.
The following answers [6, Problem 8.3] negatively although it was cer-
tainly known.
Corollary 4.6. There is a proper subalgebra A of the CAR algebra B such
that A is also CAR algebra and ZB(A) = C1. In particular, B 6= A⊗ZB(A).
Proof. Use Proposition 4.5 with X = N. Then AX is the CAR algebra. The
C*-algebra BX is also the CAR algebra because it is a separable unital LM
algebra obtained as a direct limit of algebras of the form M2n(C) by the
proof of Proposition 4.5. By Lemma 4.3, we have ZBX (AX ) = C1.
(cid:3)
5. AM = LM and AF = LF for character density ≤ ℵ1
We first show AM = LM + AF. We use the following well-known result
repeatedly. Recall that a finite-dimensional C*-algebra D is isomorphic to a
direct sum of finitely many full matrix algebras (e.g., [5, Theorem III.1.1]),
and the cardinality F of a system F of matrix units of D as defined after
[5, Theorem III.1.1] coincides with the dimension of D.
Lemma 5.1 ([5, Corollary III.3.3]). Given d ∈ N, there exists δ > 0 so that
if D is a finite-dimensional subalgebra of a C*-algebra A with a system F of
matrix units such that F = d and B is a subalgebra of A such that F ⊆δ B,
16
ILIJAS FARAH AND TAKESHI KATSURA
there exists a unitary u in the unitization of A satisfying uDu∗ ⊆ B and
commuting with D ∩ B.
Moreover, for a previously given ε > 0 in addition to d, there exists δ > 0
such that one can choose u as above so that ku − 1k < ε.
(cid:3)
Proposition 5.2. A C*-algebra is AM if and only if it is LM and AF.
Proof. We only need to prove that if a C*-algebra A is LM and AF, then it
is AM. Take a directed family {Dλ}λ∈Λ of finite-dimensional subalgebras of
A with dense union. To show that A is AM, it suffices to show that for any
λ ∈ Λ there exists a full matrix subalgebra M containing Dλ and contained
in Dλ′ for some λ′ (cid:23) λ. Then the set of such full matrix subalgebras is
directed and has dense union.
Take λ ∈ Λ. Let F be a system of matrix units of Dλ. Let δ > 0 be
as in Lemma 5.1 for d = F ∈ N. Since A is LM, it has a full matrix
subalgebra M0 such that F ⊆δ M0. By Lemma 5.1, there exists a unitary
u in the unitization of A satisfying uDλu∗ ⊆ M0. Let F ′ be a system
of matrix units of u∗M0u. Let δ′ > 0 be as in Lemma 5.1 for d = F ′.
Since {Dλ}λ∈Λ has dense union, there exists λ′ ∈ Λ such that λ′ (cid:23) λ and
F ′ ⊆δ′ Dλ′. By Lemma 5.1, there exists a unitary u′ in the unitization of A
satisfying u′(u∗M0u)u′∗ ⊆ Dλ′ and commuting with (u∗M0u) ∩ Dλ′. Since
Dλ ⊆ (u∗M0u) ∩ Dλ′, the full matrix subalgebra M := u′(u∗M0u)u′∗ of A
satisfies Dλ ⊆ M ⊆ Dλ′. This completes the proof.
(cid:3)
By Proposition 5.2, the statement AM = LM is reduced to AF = LF
because LM implies LF. Thus we only show that AF = LF for character
density at most ℵ1 although the same argument as below works for showing
AM = LM directly by just changing "F" to "M" and "finite-dimensional"
to "full matrix" in all statements and proofs.
Lemma 5.3. Let A be a separable AF algebra. For an increasing sequence
{Dn}n∈N of finite-dimensional subalgebras of A there exists an increasing
sequence {D′
n}n∈N of finite-dimensional subalgebras with dense union such
that Sn∈N Dn ⊆Sn∈N D′
Proof. This is well-known to specialists, and can be shown in a similar way
to [5, Theorem III.3.5]. For the reader's convenience, we give a proof.
n.
Let {Bk}k∈N be an increasing sequence of finite-dimensional subalgebras
of A with dense union. We construct inductively an increasing sequence
{kn}n∈N in N and a sequence {un}n∈N of unitaries in the unitization of A
with kun − 1k < 2−n such that for each n ∈ N the finite-dimensional algebra
is contained in Bkn and commutes with un+1. We first construct k1 and u1.
Choose k1 ∈ N such that F ⊆δ Bk1 where F is a system of matrix units of
D1 and δ > 0 be as in the latter statement of Lemma 5.1 for d = F and
ε = 2−1. By Lemma 5.1, there exists a unitary u1 in the unitization of A
satisfying u1D1u∗
1 ⊆ Bk1 and ku1− 1k < 1/2. Suppose that k1, . . . , kn−1 ∈ N
un ··· u2u1Dnu∗
1u∗
2 ··· u∗
n
NONSEPARABLE UHF ALGEBRAS I: DIXMIER'S PROBLEM
17
and unitaries u1, . . . , un−1 were chosen. Choose kn ∈ N such that kn > kn−1
and F ′ ⊆δ′ Bkn where F ′ is a system of matrix units of
un−1 ··· u2u1Dnu∗
1u∗
2 ··· u∗
n−1
and δ′ > 0 be as in the latter statement of Lemma 5.1 for d = F ′ and
ε = 2−n. Lemma 5.1 gives us a unitary un in the unitization of A satisfying
unun−1 ··· u2u1Dnu∗
1u∗
2 ··· u∗
n−1u∗
n ⊆ Bkn,
commuting with
un−1 ··· u2u1Dn−1u∗
1u∗
2 ··· u∗
n−1 ⊆ un−1 ··· u2u1Dnu∗
1u∗
2 ··· u∗
n−1 ∩ Bkn
and satisfying kun − 1k < 2−n. Thus we get the desired sequences {kn}n∈N
and {un}n∈N.
Since kun − 1k < 2−n for all n ∈ N andPn∈N 2−n = 1 < ∞, the sequence
{un ··· u2u1}n∈N converges to a unitary u in the unitization of A. Since
un ··· u2u1Dnu∗
2 ··· u∗
1u∗
n = un+1un ··· u2u1Dnu∗
1u∗
⊆ un+1un ··· u2u1Dn+1u∗
2 ··· u∗
1u∗
nu∗
2 ··· u∗
n+1
nu∗
n+1,
1u∗
1u∗
1u∗
n−1u∗
2 ··· u∗
2 ··· u∗
λ for each λ ∈ Λ.
2 ··· u∗
n := u∗Bknu for n ∈ N. Then {D′
n.
un ··· u2u1Dnu∗
n commutes with un+2. By repeating this argument,
one can see that un ··· u2u1Dnu∗
n commutes with um for all m >
n. Hence we get uDnu∗ = unun−1 ··· u2u1Dnu∗
n ⊆ Bkn. We
set D′
n}n∈N is an increasing sequence
of finite-dimensional subalgebras with dense union such that Sn∈N Dn ⊆
Sn∈N D′
(cid:3)
In the next lemma, for two families Υ = {Dλ}λ∈Λ and Υ′ = {D′
λ}λ∈Λ′ of
subalgebras, Υ ⊆ Υ′ means that Λ ⊆ Λ′ and Dλ = D′
Lemma 5.4. Let A be a separable AF algebra contained in a separable AF
algebra A′. For a countable directed family Υ of finite-dimensional subalge-
bras of A with dense union, there exists a countable directed family Υ′ of
finite-dimensional subalgebras of A′ with dense union such that Υ ⊆ Υ′.
Proof. Let us write Υ = {Dλ}λ∈Λ. Since Λ is countable, we can choose a
subsequence {λn}n∈N of Λ such thatSλ∈Λ Dλ =Sn∈N Dλn. By Lemma 5.3,
there exists an increasing sequence {D′
n}n∈N of finite-dimensional subalge-
bras of A′ with dense union such thatSn∈N Dλn ⊆Sn∈N D′
n. For each λ ∈ Λ,
there exists n ∈ N such that Dλ ⊆ D′
n because Dλ is finite-dimensional. Let
Λ′ := Λ ∐ N, ordered by requiring that Λ and N have their natural order-
n. Then the family Υ′ := {D′
ings and λ (cid:22) n if Dλ ⊆ D′
λ}λ∈Λ′ defined by
D′
λ := Dλ for λ ∈ Λ satisfies the desired properties.
(cid:3)
Lemma 5.5. Each LF algebra of character density at most ℵ1 has a σ-
complete directed family of separable AF subalgebras with dense union in-
dexed by the ordinal ω1.
18
ILIJAS FARAH AND TAKESHI KATSURA
Proof. Let A be an LF algebra with χ(A) ≤ ℵ1. Fix a dense subset {xγ :
γ ∈ ω1} of A, and define Aλ := C ∗({xγ : γ < λ}) for each λ ∈ ω1. Then
{Aλ}λ∈ω1 is a σ-complete directed family of separable subalgebras of A.
By Lemma 2.13, A also has a σ-complete direct family of separable AF
subalgebras with dense union. By Proposition 2.12 applied with id : A → A,
there is a club Λ ⊆ ω1 such that Aλ is AF for λ ∈ Λ. As ordered sets, Λ is
isomorphic to ω1, and {Aλ}λ∈Λ is the desired family.
(cid:3)
Proposition 5.6. Each LF algebra of character density at most ℵ1 is an
AF algebra.
Proof. Let A be an LF algebra with χ(A) ≤ ℵ1. Let {Aξ}ξ∈ω1 be a σ-
complete directed family of separable AF subalgebras of A with dense union
as in Lemma 5.5. Using transfinite recursion, we are going to construct
an increasing family of countable directed families Υξ of finite-dimensional
subalgebras whose union is dense in Aξ for each ξ ∈ ω1. For ξ = 0, choose
an increasing sequence of finite-dimensional subalgebras of A0 with dense
union, and set it Υ0. If Υξ has been defined, then Υξ+1 is defined using
Lemma 5.4. If η is a limit ordinal and Υξ has been defined for all ξ < η,
required.
let Υη =Sξ<η Υξ. Since Aη is the closure of the union of {Aξ}ξ<η, Υη is as
Finally let Υ = Sξ∈ω1
Υξ. Then this is a directed family of finite-
dimensional subalgebras of A with dense union. Thus A is an AF alge-
bra.
(cid:3)
The example of the following section easily shows that the version of
Lemma 5.4 for nonseparable algebras is false.
6. AM 6= LM and AF 6= LF for character density > ℵ1
In this section, we construct an LM algebra which is not AF. This C*-
algebra shows the difference between the classes of AM and LM algebras as
well as between the classes of AF and LF algebras. To show that a given
C*-algebra is not AF, we use the following criterion.
The converse direction in the following lemma was proved by George
Elliott, following a remark by Tamas Matrai, during the first author's talk
at a set theory seminar in Toronto in April 2009.
Lemma 6.1. A C*-algebra A is AF if and only if there exists a map ρ : A →
A such that ka − ρ(a)k < 1 for every a ∈ A and C ∗({ρ(a)}a∈F ) is finite-
dimensional for every finite subset F of A.
Proof. Assume A is AF and let {Aλ}λ∈Λ be a directed family of finite-
dimensional subalgebras of A with dense union. For each a ∈ A there exists
λa ∈ Λ such that there exists ρ(a) ∈ Aλa with ka − ρ(a)k < 1. For every
finite subset F of A there exists λ ∈ Λ such that λ (cid:23) λa for all a ∈ F . Then
C ∗({ρ(a)}a∈F ) ⊆ Aλ is finite-dimensional.
NONSEPARABLE UHF ALGEBRAS I: DIXMIER'S PROBLEM
19
Now assume that ρ : A → A is as in the statement of the lemma. If Λ is
the family of all finite subsets of A then Aλ = C ∗({ρ(a)}a∈λ) form a directed
family of finite-dimensional subalgebras of A. Fix a ∈ A and ε > 0. Let
λ = {a/ε}. Then ερ(a/ε) ∈ Aλ and ka − ερ(a/ε)k < ε. Since a and ε were
arbitrary, we conclude A is AF.
(cid:3)
We also use the following lemma (for the case when A is the CAR algebra)
in the proof of Proposition 6.12
Lemma 6.2. Let A be a unital LM subalgebra of a unital C*-algebra B.
Take a1, a2, . . . , an ∈ A and b1, b2, . . . , bn ∈ ZB(A).
i=1 is linearly
independent in A and Pn
Proof. Since A is LM, the natural map from A ⊗ ZB(A) to B is injective
by Lemma 2.17. It is well known that the inclusion map from the algebraic
tensor product of A and ZB(A) to the (maximal) tensor product A⊗ ZB(A)
is injective (see [2, II.9.1.3]). The conclusion follows from these lemmas. (cid:3)
i=1 aibi = 0 in B, then we have bi = 0 for all i.
If (ai)n
Definition 6.3. We say that a pair (v1, v2) of self-adjoint unitaries v1, v2 in
a unital C*-algebra is generic if the family
(cid:0)(v1v2)n, (v1v2)nv1(cid:1)n∈Z
= (1, v1, v2, v1v2, v2v1, v1v2v1, v2v1v2, v1v2v1v2, v2v1v2v1, v1v2v1v2v1, . . .)
is linearly independent.
In other words, (v1, v2) is generic if and only if the map sending the natural
generators of the group algebra C((Z/2Z) ∗ (Z/2Z)) to v1, v2 is injective.
Lemma 6.4. Let v1, v2, w1, w2 be the four self-adjoint unitaries in the C*-
algebra C([0, 1], M2(C)) defined by
0
v1(t) =(cid:18)1
0 −1(cid:19) ,
1 0(cid:19) ,
w1(t) =(cid:18)0 1
sin(πt)
v2(t) =(cid:18)cos(πt)
w2(t) =(cid:18)− sin(πt)
sin(πt) − cos(πt)(cid:19) ,
sin(πt)(cid:19)
cos(πt)
cos(πt)
for t ∈ [0, 1]. Then v1, v2, w1, w2 satisfy v1w1 = −w1v1, v2w2 = −w2v2 and
the pair (v1, v2) is generic.
Proof. It is routine to check the two equalities v1w1 = −w1v1 and v2w2 =
−w2v2. That the pair (v1, v2) is generic comes from the fact that {cos(nπt)+
√−1 sin(nπt)}n∈Z is linearly independent in C([0, 1]). We leave the details
to the readers.
(cid:3)
Let X be an infinite set, and [X]2 be the set of all subsets of X with
cardinality 2. For ξ = {x, y} ∈ [X]2 let Aξ be the CAR algebra. We fix
four self-adjoint unitaries vx,y, vy,x, wx,y, wy,x in Aξ such that vx,ywx,y =
−wx,yvx,y, vy,xwy,x = −wy,xvy,x and the pair (vx,y, vy,x) is generic. Such
unitaries exist by Lemma 6.4 because there exists a unital embedding from
C([0, 1], M2(C)) to the CAR algebra.
20
ILIJAS FARAH AND TAKESHI KATSURA
M2(C).
We define a UHF algebra A[X]2 by A[X]2 =Nξ∈[X]2 Aξ ∼=N[X]2×ℵ0
For a subset Y of X, we set A[Y ]2 =Nξ∈[Y ]2 Aξ ⊆ A[X]2.
Definition 6.5. For a set X, we denote by GX the abelian group consisting
of all finite subsets of X where the operation is the symmetric difference ∆.
We often identify an element x of X with a subset {x} of X. Thus
the group GX is generated by the family {x}x∈X of mutually commuting
involutions. Hence GX is isomorphic to the group LX(Z/2Z) of the direct
sum of X copies of Z/2Z.
For g ∈ GX we define an automorphism αg on A[X]2 by
αg = Ox ∈ g and y /∈ g
Ad vx,y.
If x /∈ g define unitaries Vg;x and Vx;g in A[X]2 via
Vg;x =Yy∈g
vy,xvx,y
and
vx,yvy,x.
Vx;g =Yy∈g
Lemma 6.6. If x /∈ g then αg ◦ αx = Ad(Vg;x) ◦ αg∪{x} and αg∪{x} ◦ αx =
Ad(Vx;g) ◦ αg.
Proof. Note that vx,y and vz,t commute unless z = y and x = t. Using x /∈ g
we have
αg ◦ αx =(cid:16) Oy∈g and z /∈g
Ad vy,z(cid:17) ◦(cid:16)Oz6=x
Ad vy,x ◦ Oy∈g and z /∈g∪{x}
Ad vx,z(cid:17)
Ad vy,z(cid:17) ◦(cid:16)Oz∈g
Ad vx,z ◦ Oz /∈g∪{x}
Ad vx,z(cid:17)
=(cid:16)Oy∈g
= Ad(Vg;x) ◦ αg∪{x}
This proves the first equality. Since αx is an involution and Vx;g = V ∗
first equality implies the second equality.
g;x, the
(cid:3)
Let us choose a faithful representation A[X]2 ⊆ B(H) on some Hilbert
space H (see Section 7 for one construction of such a representation). Let
ℓ2(GX , H) be the Hilbert space consisting of functions ξ : GX → H with
Pg∈GX kξ(g)k2 < ∞. We embed A[X]2 into B(ℓ2(GX , H)) by
for a ∈ A[X]2, ξ ∈ ℓ2(GX , H) and g ∈ GX . For each x ∈ X, we define
ux ∈ B(ℓ2(GX , H)) by
(aξ)(g) = αg(a)ξ(g) ∈ H
(uxξ)(g) = Vg;x ξ(g ∪ {x}) ∈ H
(uxξ)(g ∪ {x}) = Vx;g ξ(g) ∈ H
for ξ ∈ ℓ2(GX , H) and g ∈ GX with x /∈ g.
Lemma 6.7. For each x ∈ X, ux is a self-adjoint unitary such that Ad ux
and αx agree on A[X]2 ⊆ B(ℓ2(GX , H)).
NONSEPARABLE UHF ALGEBRAS I: DIXMIER'S PROBLEM
21
Proof. For g ∈ GX such that x /∈ g the subspace ℓ2({g, g ∪ {x}}, H) ⊆
ℓ2(GX , H) is invariant for ux, and ux is represented on it as
0 (cid:19) .
This shows that ux is a self-adjoint unitary. To show that Ad ux and αx
agree on A[X]2 ⊆ B(ℓ2(GX , H)), it suffices to see
ux =(cid:18) 0
Vg;x
Vx;g
Ad(Vx;g) ◦ αg = αg∪{x} ◦ αx
Ad(Vg;x) ◦ αg∪{x} = αg ◦ αx
which is Lemma 6.6.
(cid:3)
By Lemma 6.7 we see that for {x, y} ∈ [X]2 and z ∈ X we have Ad ux ↾A{x,y}=
Ad vx,y, and Ad uz ↾A{x,y}= id if z /∈ {x, y}. In particular, uz commutes with
vx,y unless y = z.
Lemma 6.8. For {x, y} ∈ [X]2 the two self-adjoint unitaries uxvx,y and
uyvy,x commute.
Proof. Take {x, y} ∈ [X]2. First note that for h ∈ GX we have αh(vx,y) =
vy,xvx,yvy,x if y ∈ h and x /∈ h, and αh(vx,y) = vx,y otherwise.
Fix g ∈ GX such that x /∈ g and y /∈ g. The subspace
Hg = ℓ2(cid:0){g, g ∪ {x}, g ∪ {y}, g ∪ {x, y}}, H(cid:1) ⊆ ℓ2(GX , H)
is invariant for each of ux, uy, vx,y and vy,x. Using the observation in the
beginning of this proof, we see that ux, uy, vx,y and vy,x are represented on
Hg by
ux =
uy =
0
Vx;g
0
0
0
0
Vy;g
0
0
0
0
Vg;x
0
0
0 Vx;g∪{y}
Vg;y
0
0
Vg∪{y};x
0
0
0 Vg∪{x};y
0
Vy;g∪{x} 0
0
0
0
0
0
vx,y =
,
, vy,x =
vx,y
0
0
0
vy,x
0
0
0
0
vx,y
0
0
0 vy,xvx,yvy,x
0
0
0
vx,yvy,xvx,y
0
0
0
0
vy,x
0
0
0
0
vx,y
0
0
0
vy,x
,
.
Using the computations such as Vx;g∪{y} = Vx;gvx,yvy,x, we see that uxvx,y
and uyvy,x are represented on Hg by
uxvx,y =
uyvy,x =
0
Vg;xvx,y
Vx;gvx,y
0
0
0
0
Vy;gvy,x
0
0
0
0
0
0
0
Vy;gvx,y
0
0
0
0
0
Vg;xvy,x
Vx;gvy,x
Vg;yvy,x
0
0
0
0
0
Vg;yvx,y
0
0
,
.
22
ILIJAS FARAH AND TAKESHI KATSURA
The unitaries Vg;x, Vx;g, Vg;y, Vy;g, vx,y and vy,x occurring in entries of these
two matrices commute with each others except that vx,y does not commute
with vy,x. Using this fact, one can show that both (uxvx,y)(uyvy,x) and
(uyvy,x)(uxvx,y) are equal to
Therefore uxvx,y and uyvy,x commute.
(cid:3)
0
0
0
0
0
0
0
0
0
Vx;gVy;g
Vg;xVg;y
Vg;xVy;gvy,xvx,y
Vx;gVg;yvx,yvy,x
.
B[X]2 := C ∗(A[X]2 ∪ {ux}x∈X ) ⊆ B(cid:0)ℓ2(GX , H)(cid:1).
0
0
0
B[Y ]2 := C ∗(A[Y ]2 ∪ {ux}x∈Y ) ⊆ B[X]2.
Let
For a subset Y ⊆ X, we define
Remark 6.9. The C*-algebra B[X]2 does not depend on the choices of em-
beddings A[X]2 ⊆ B(H), and is isomorphic to a cocycle crossed product
A[X]2 ⋊(α,c) GX for an appropriate cocycle action (α, c) (see [15] for defini-
tions of cocycle actions and cocycle crossed products). In fact, the proof of
Proposition 6.10 shows that any C*-algebra generated by A[X]2 ∪ {ux}x∈X
with the relations in Lemma 6.7 and Lemma 6.8 is isomorphic to B[X]2.
Proposition 6.10. The C*-algebra B[X]2 is a unital LM algebra with χ(B[X]2) =
X.
Proof. By Lemma 2.20, we have χ(A[X]2) = X. This implies χ(B[X]2) =
X.
We are going to show that B[X]2 is a direct limit of CAR algebras. This
implies that B[X]2 is LM. For a finite subset F ⊆ X and an injective map
ι : F → X \ F , define a subalgebra D(F,ι) ⊆ B[X]2 by
D(F,ι) := C ∗(cid:0)B[F ]2 ∪ {wx,ι(x)}x∈F(cid:1) ⊆ B[X]2.
The family {D(F,ι)}(F,ι) of subalgebras is directed because X is infinite, and
its union is dense in B[X]2. Thus it suffices to show that D(F,ι) is the CAR
algebra for every finite subset F ⊆ X and every injective map ι : F → X\F .
Take a finite subset F ⊆ X and an injective map ι : F → X \ F . For
x ∈ F , we define
which is a self-adjoint unitary. Since Lemma 6.7 shows
u′
x := ux Yy∈F \{x}
vx,y ∈ D(F,ι).
Ad ux ↾A[F ]2 = αx ↾A[F ]2 = Ad(cid:16) Yy∈F \{x}
vx,y(cid:17) ↾A[F ]2 ,
x commutes with the subalgebra A[F ]2. The family {u′
u′
x}x∈F mutually com-
mutes by Lemma 6.8. For each x ∈ F , the self-adjoint unitary wx,ι(x) ∈
NONSEPARABLE UHF ALGEBRAS I: DIXMIER'S PROBLEM
23
D(F,ι) commutes with A[F ]2 and {wy,ι(y), u′
x. Therefore C ∗(u′
−wx,ι(x)u′
Lemma 4.1, and the family
{C ∗(u′
x, wx,ι(x))}x∈F ∪ {A[F ]2}
xwx,ι(x) =
x, wx,ι(x)) is isomorphic to M2(C) for x ∈ F by
y}y∈F \{x}, and satisfies u′
mutually commutes. Since D(F,ι) is generated by these mutually commuting
subalgebras, we get
D(F,ι) = A[F ]2 ⊗Ox∈F
C ∗(u′
x, wx,ι(x)) ∼= O[F ]2×ℵ0+F
M2(C).
We are done.
(cid:3)
Lemma 6.11. Let Y be a nonempty proper subset of X. Take x ∈ Y and
y ∈ X \ Y . Then every element in B[Y ]2 ⊆ B[X]2 can be written as avx,y + a′
for a, a′ ∈ ZB[X]2 (A{x,y}).
Proof. Since vx,y is a self-adjoint unitary in A{x,y}, the set of all elements in
the form avx,y +a′ for a, a′ ∈ ZB[X]2 (A{x,y}) is a subalgebra of B[X]2. Hence it
suffices to show that the generators A[Y ]2 ∪{uz}z∈Y of B[Y ]2 are in this form.
We have A[Y ]2 ⊆ ZB[X]2 (A{x,y}) since y /∈ Y . We have uz ∈ ZB[X]2 (A{x,y})
for z ∈ Y \{x}. Finally, we get ux = (uxvx,y)vx,y and uxvx,y ∈ ZB[X]2 (A{x,y}).
We are done.
(cid:3)
Proposition 6.12. If X > ℵ1 then B[X]2 is not AF.
Proof. For the sake of obtaining a contradiction, assume that B[X]2 is AF.
Then by Lemma 6.1 there exists a family {bx}x∈X in B[X]2 with kux−bxk < 1
for all x ∈ X such that C ∗({bx}x∈F ) ⊆ B[X]2 is finite-dimensional for all
finite subsets F of X.
For each x ∈ X, there exists a countable subset Yx of X with x ∈ Yx
such that bx ∈ B[Yx]2. Since X > ℵ1, we can apply Lemma 2.1 to get
{x, y} ∈ [X]2 such that x /∈ Yy and y /∈ Yx. By Lemma 6.11, there exists
x and by = ayvy,x+a′
ax, a′
y.
Since kux − bxk < 1, we have
x, ay, a′
y ∈ ZB[X]2 (A{x,y}) such that bx = axvx,y +a′
(cid:13)(cid:13)(cid:0)(ux − bx) − wx,y(ux − bx)wx,y(cid:1)/2(cid:13)(cid:13) < 1.
x) − (−axvx,y + a′
(cid:0)bx − wx,ybxwx,y(cid:1)/2 =(cid:0)(axvx,y + a′
x)(cid:1)/2 = axvx,y,
We have
and similarly (ux − wx,yuxwx,y)/2 = ux. Hence we get kux − axvx,yk < 1.
Thus kuxvx,y−axk < 1. Since uxvx,y is a unitary, ax is an invertible element.
Similarly, one can show that ay is also invertible.
By the assumption, C ∗({bx, by}) is finite-dimensional. Therefore {(bxby)n}∞
is linearly dependent. Hence there exist N ∈ N and λ0, λ1, . . . , λN ∈ C with
n=0
24
ILIJAS FARAH AND TAKESHI KATSURA
λN 6= 0 such that PN
λn(bxby)n =
NXn=0
n=0 λn(bxby)n = 0. We can write
NXn=0
λn(cid:0)(axvx,y + a′
x)(ayvy,x + a′
y)(cid:1)n = Xv∈V
fvv
x, ay, a′
where
V := {1, vx,y, vy,x, vx,yvy,x, vy,xvx,y, vx,yvy,xvx,y, vy,xvx,yvy,x, . . . , (vx,yvy,x)N}
and for each v ∈ V , fv ∈ ZB[X]2 (A{x,y}) is a sum of products of λ0, λ1, . . . , λN ∈
C and ax, a′
y ∈ ZB[X]2 (A{x,y}). Since V ⊆ A{x,y} is linearly in-
dependent, we get fv = 0 for all v ∈ V by Lemma 6.2.
In particular,
f(vx,yvy,x)N = λN (axay)N ∈ ZB[X]2 (A{x,y}) is 0. This cannot happen because
λN 6= 0 and both ax and ay are invertible. Thus we get a contradiction. We
are done.
(cid:3)
Remark 6.13. When X = ℵ0, B[X]2 is a UHF algebra (in fact CAR
algebra) by Glimm's theorem [11, Theorem 1.13]. When X = ℵ1, B[X]2
is a unital AM algebra by Proposition 6.10 and Theorem 1.3 (1). In this
case one can show that B[X]2 is not UHF in a similar (but much more
complicated) way to the proof of Proposition 4.5 (2) (see [10]).
Remark 6.14. As we pointed out in Remark 4.2, the examples in Section 4
of unital AM algebras which are not UHF are obtained as crossed products
of UHF algebras by the group Z/2Z. The examples in this section of unital
LM algebras which are not AM are obtained as cocycle crossed products (see
Remark 6.9). However we do not know the following.
Problem 6.15. Find an example of a unital LM algebra which is not AM
such that it is obtained as a crossed product of a unital AM (or UHF) algebra
by a discrete group.
Remark 6.16. We can solve the non-unital version of this problem using the
examples in this section. In fact, by [15, Corollary 3.7] the tensor product
B[X]2 ⊗ K is obtained as an (ordinary) crossed product of A[X]2 ⊗ K by
the group GX where K := K(cid:0)ℓ2(GX )(cid:1) is the non-unital AM algebra of all
compact operators on the Hilbert space ℓ2(GX ). Thus for every cardinal
κ > ℵ1, there exists an example of a non-unital LM algebra with character
density κ which is not AM such that it is obtained as a crossed product of
a non-unital AM algebra by a discrete group. Note that B[X]2 ⊗ K is not
AM if B[X]2 is not AM because every corner of an AM algebra is AM.
The same comments can be applied to LF and AF instead of LM and
AM.
7. Representation density and character density
The purpose of this section is to give an answer to the half of the question
raised by Masamichi Takesaki when the second author gave a talk on this
paper. We could not answer the other half (Problem 7.19). The proof uses
NONSEPARABLE UHF ALGEBRAS I: DIXMIER'S PROBLEM
25
the construction (Proposition 7.12) that was given by Bruce Blackadar when
the first author gave a talk. Both authors would like to thank Masamichi
Takesaki and Bruce Blackadar.
For a Hilbert space H, we also denote by χ(H) the smallest cardinality
of a dense subset of H. Note that for an infinite-dimensional Hilbert space
H and an infinite set X, we get χ(H) = X if and only if H is isomorphic
to ℓ2(X).
Definition 7.1. The representation density χr(A) of a C*-algebra A is
the smallest cardinal χ(H) where H is a Hilbert space on which A can be
faithfully represented.
Note that both the representation density χr and the character density χ
(Definition 1.2) are monotonic in the sense that if A is a subalgebra of B
then the density of B is not smaller than the density of A.
Since these cardinal invariants of C*-algebras were apparently not consid-
ered previously, the reader will hopefully excuse us for starting this section
by listing a few trivial statements.
Lemma 7.2. For every C*-algebra A we have that
χ(A) ≥ sup(cid:8)X : X is a family of commuting projections in A(cid:9)
χr(A) ≥ sup(cid:8)X : X is a family of nonzero orthogonal projections in A(cid:9).
Proof. For the first part note that if p and q are distinct commuting projec-
tions then kp − qk = 1. The second part is obvious.
(cid:3)
Lemma 7.3. For every infinite-dimensional Hilbert space H we have
χ(cid:0)B(H)(cid:1) = B(H) = 2χ(H).
Proof. Let us choose an infinite set X with X = χ(H), and identify H
with ℓ2(X). For a subset Y ⊆ X, let pY ∈ B(H) be the projection onto the
subspace ℓ2(Y ) ⊆ H. Then {pY }Y ⊆X is a family of commuting projections
of size 2X. Thus we have χ(B(H)) ≥ 2X by Lemma 7.2. For x, y ∈ X,
p{x}B(H)p{y} is one dimensional, and the map
B(H) ∋ T 7→(cid:0)p{x}T p{y}(cid:1)x,y∈X ∈ Yx,y∈X(cid:0)p{x}B(H)p{y}(cid:1) ∼= Yx,y∈X
C
is injective. Hence we get χ(B(H)) ≤ B(H) ≤ CX×X = 2X. We are
done.
If K = 22ℵ0 with the product topology then C(K) ∼= N2ℵ0 C2 is an
abelian C*-algebra with character density 2ℵ0 and representation density
ℵ0. The first claim follows by Lemma 2.20. The second claim follows from
the fact that K is, being a product of 2ℵ0 separable spaces, separable by the
Hewitt -- Marczewski -- Pondiczery Theorem (see e.g.,
[7, Corollary 2.3.16]).
See also Corollary 7.7, Theorem 7.17 and Problem 7.19.
Lemma 7.4. For every C*-algebra A we have χr(A) ≤ χ(A) ≤ 2χr(A).
(cid:3)
26
ILIJAS FARAH AND TAKESHI KATSURA
Proof. Choose a subset X ⊆ A with X = χ(A). For each x ∈ X, there
exists a cyclic representation πx : A → B(Hx) with kπx(x)k = kxk (see [2,
Corollary II.6.4.9]). Since Hx has a cyclic vector for πx, we have χ(Hx) ≤
χ(A). Then the representation
π := Mx∈X
Hx(cid:17) = Xx∈X
is faithful, and
χ(cid:16)Mx∈X
πx : A → B(cid:16)Mx∈X
Hx(cid:17)
χ(Hx) ≤ X × χ(A) = χ(A)
Hence χr(A) ≤ χ(A). The second inequality χ(A) ≤ 2χr(A) follows from
Lemma 7.3.
Lemma 7.5. Let X0 ∋ x 7→ ξx ∈ H be a map from a set X0 to a Hilbert
space H such that X0 > χ(H). Then for every ε > 0, there exists X1 ⊆ X0
with X1 > χ(H) such that kξx − ξyk < ε for every x, y ∈ X1.
Proof. Choose a dense subset Y ⊆ H with Y = χ(H). For each x ∈ X0
there exists η(x) ∈ Y such that kξx−η(x)k < ε/2. Since X0 > χ(H) = Y ,
there exists η ∈ Y such that the set X1 := {x ∈ X0 : η(x) = η} ⊆ X0 satisfies
X1 > χ(H). Then for every x, y ∈ X1, we get
(cid:3)
kξx − ξyk ≤ kξx − ηk + kξy − ηk < ε.
(cid:3)
Proposition 7.6. For a family {Ax}x∈X of nonabelian unital C*-algebras,
the representation density of the tensor product A = Nx∈X Ax is at least
X.
Proof. Assume the contrary and fix a faithful representation π : A → B(H)
for a Hilbert space H with X > χ(H). Note that this assumption implies
that X is uncountable. For each x ∈ X, fix ax and bx in the unit ball of Ax
such that axbx 6= bxax. Since π is faithful, we can choose a vector ξx ∈ H
such that
π(axbx − bxax)ξx 6= 0.
Since X is uncountable, there exist δ > 0 and a subset X0 ⊆ X with
X0 > χ(H) such that for all x ∈ X0 we have
kπ(axbx − bxax)ξxk ≥ δ.
In this proof, we write a ≈ε b if ka − bk < ε. Since
Set ε = δ/4 > 0.
X0 > χ(H), we can apply Lemma 7.5 to {ξx}x∈X0 and ε > 0 to get
X1 ⊆ X0 with X1 > χ(H) such that ξx ≈ε ξy for every x, y ∈ X1. By
applying Lemma 7.5 three more times to {π(ax)ξx}x∈X1 and so on, we get
X4 ⊆ X1 with X4 > χ(H) such that
π(ax)ξx ≈ε π(ay)ξy,
π(bx)ξx ≈ε π(by)ξy,
π(bxax)ξx ≈ε π(byay)ξy
NONSEPARABLE UHF ALGEBRAS I: DIXMIER'S PROBLEM
27
for every x, y ∈ X4. Since X4 > χ(H) ≥ ℵ0, we can take two distinct
x, y ∈ X4. Then we have
π(axbx)ξx = π(ax)π(bx)ξx ≈ε π(ax)π(by)ξy = π(axby)ξy ≈ε π(axby)ξx
=
π(bxax)ξx ≈ε π(byay)ξy = π(by)π(ay)ξy ≈ε π(by)π(ax)ξx = π(byax)ξx
because ax ∈ Ax ⊆ A and by ∈ Ay ⊆ A commute. Thus we get
(cid:13)(cid:13)π(axbx − bxax)ξx(cid:13)(cid:13) < 4ε = δ,
which is a contradiction. This completes the proof.
(cid:3)
Corollary 7.7. If A is a UHF algebra then χ(A) = χr(A).
(cid:3)
With the possible exception of the algebras AX,Y as defined in §3, each
example of an AM, or even LM, algebra given so far has a UHF subalge-
bra with the same character density. Since the algebras AX,Y are tensor
products of separable algebras, Proposition 7.6 implies that for each AM or
LM algebra A so far defined in this paper we have χ(A) = χr(A). We are
going to show that χ(A) can be any cardinality between χr(A) and 2χr(A)
for unital AM algebras A.
Let X be an infinite set. As in Section 4, let Ax be a C*-algebra generated
by two self-adjoint unitaries vx, wx with vxwx = −wxvx for each x ∈ X, and
let AX := Nx∈X Ax. By Lemma 4.1, Ax ∼= M2(C) for each x ∈ X and
hence AX ∼=NX M2(C) is a UHF algebra. For each Y ⊆ X, we set
AY :=Ox∈Y
Ax ⊆ AX .
We are going to use the GNS representation of AX associated with the
unique tracial state of AX. For the reader's convenience we explain what
it is. For each finite subset F ⊆ X, there exists a unique linear functional
τF : AF → C satisfying the trace condition τF (ab) = τF (ba) for a, b ∈ AF and
the normalized condition τF (1) = 1. If F = n, then we have τF = 2−nTr
where Tr is the usual trace of AF ∼= M2n(C). It is easy to see that τF is
positive and faithful, that is, τF (a∗a) > 0 for all a ∈ AF \ {0}. Let Afin
X :=
SF ⊆X AF ⊆ AX where F runs all finite subsets of X. By the uniqueness of
the tracial state τF , we get τF ′ ↾AF = τF for two finite subsets F ⊆ F ′ ⊆ X.
Thus we get a linear map τ : Afin
X → C such that τ ↾AF = τF for every finite
subset F ⊆ X. Although we do not need it, we would like to remark that τ
can be extended to the unique tracial state of AX (cf. [5, Lemma I.9.5]). We
define an inner product on Afin
X ∋ (a, b) 7→ τ (ab∗) ∈ C. Then
the completion HX of Afin
X with respect to the norm coming from the inner
product defined as above becomes a Hilbert space. The embedding from Afin
X
to HX is denoted by Afin
X ∋ a 7→ a ∈ HX. The image of this embedding is
dense in HX. For each finite subset F ⊆ X and each a ∈ AF , it is easy to see
that the map b 7→ bab extends to a bounded operator on HX. Thus we get a ∗-
homomorphism πF : AF → B(HX) such that πF (a)(bb) = bab for a ∈ AF and
X × Afin
X by Afin
28
ILIJAS FARAH AND TAKESHI KATSURA
X . We have πF ′ ↾AF = πF for two finite subsets F ⊆ F ′ ⊆ X. Since
b ∈ Afin
the family {π{x}[A{x}]}x∈X mutually commutes, we get a representation
π : AX → B(HX) such that π ↾AF = πF for every finite subset F ⊆ X. This
representation is called the GNS representation associated with τ . Since
π(a)(ba∗) = daa∗ 6= 0 for all F ⊆ X and all a ∈ AF \ {0}, π is injective. In
order to simplify the notation we identify AX with the subalgebra π[AX ] of
B(HX).
Lemma 7.8. We have χ(HX) = X.
Proof. Since the union of finite-dimensional subspaces {a ∈ HX a ∈ AF}
for finite subsets F ⊆ X is dense in HX, we have χ(HX ) ≤ X. For distinct
x, y ∈ X, we have τ (uxuy) = 0 because
τ (uxuy) = τ (wx(wxuxuy)) = τ ((wxuxuy)wx)
= τ (wxux(wxuy)) = τ (wx(−wxux)uy) = −τ (uxuy).
Hence we get
kcux −cuyk2 = τ ((ux − uy)(ux − uy)) = τ (2 − 2uxuy) = 2
for all x, y ∈ X with x 6= y. This shows that χ(HX) ≥ X. Thus we get
χ(HX) = X.
(cid:3)
We can consider the power-set P(X) of a set X as an abelian group with
respect to the symmetric difference. This group is naturally isomorphic
to the direct product of X copies of Z/2Z. For g ∈ P(X) consider an
automorphism of AX defined by
αg =Ox∈g
Ad vx.
X .
Then α defines an action of P(X) on AX . For each g ∈ P(X), the automor-
phism αg preserves the subalgebra AF ⊆ AX and satisfies τF ◦ αg = τF for
every finite subset F ⊆ X. Hence we get an element ug ∈ B(HX) such that
ug(bb) = αgb(b) for b ∈ Afin
Lemma 7.9. The elements {ug}g∈P(X) ⊆ B(HX) are self-adjoint unitaries
satisfying ugaug = αg(a) and uguh = ugh for a ∈ AX ⊆ B(HX) and g, h ∈
P(X).
Proof. Take g ∈ P(X). Since αg preserves τ , the element u∗
isfies u∗
because α−1
αg(aαg(b)) = αg(a)b and αg(αh(b)) = αgh(b) for b ∈ AX .
Definition 7.10. For an infinite set X and a subgroup Γ ⊆ P(X) we define
g ∈ B(HX) sat-
X . Hence ug is a unitary. This is self-adjoint
g = αg. The latter two equalities follow from the equations
(cid:3)
g(bb) = α−1
g b(b) for b ∈ Afin
BX,Γ := C ∗(AX ∪ {ug}g∈Γ) ⊆ B(HX).
NONSEPARABLE UHF ALGEBRAS I: DIXMIER'S PROBLEM
29
Remark 7.11. One can show that BX,Γ is isomorphic to the crossed product
AX ⋊α Γ.
In particular BX in Section 4 is isomorphic to BX,Γ for Γ =
{∅, X} ∼= Z/2Z.
Proposition 7.12. The C*-algebra BX,Γ satisfies χ(BX,Γ) = X + Γ and
χr(BX,Γ) = X.
Proof. We have χ(AX ) = X by Lemma 2.20. On the other hand, we have
χ(C ∗({ug}g∈Γ)) ≥ Γ by Lemma 7.2 because {(ug + 1)/2}g∈Γ is a family of
commuting projections. Since BX,Γ is generated by AX and {ug}g∈Γ, we get
X + Γ = max{X,Γ} ≤ χ(BX,Γ) ≤ X + Γ
This shows χ(BX,Γ) = X +Γ. Since BX,Γ ⊆ B(HX), we have χr(BX,Γ) ≤
χ(HX) = X by Lemma 7.8. We also have χr(BX,Γ) ≥ χr(AX ) = X by
Corollary 7.7. Hence we get χr(BX,Γ) = X.
(cid:3)
Proposition 7.13. The unital C*-algebra BX,Γ is AM if every finite subset
of Γ is included in a subgroup generated by g1, g2, . . . , gn ∈ Γ which are
infinite and mutually disjoint.
Proof. Take mutually disjoint infinite elements g1, g2, . . . , gn ∈ Γ. Take a
finite subset F of X and choose xi ∈ gi \ F for i = 1, 2, . . . , n. Let Λ be the
set of all such data λ = ({gi}n
Dλ := C ∗(cid:0){ugi}n
i=1, F,{xi}n
i=1 ∪ AF ∪ {wxi}n
i=1(cid:1) ⊆ BX,Γ.
By the assumption of Γ, the family {Dλ}λ∈Λ of subalgebras is directed and
its union is dense in B[X]2. We are going to show Dλ ∼= M2m+n (C) for
λ = ({gi}n
i=1) as above where m = F. This implies that B[X]2
is AM, and hence completes the proof. For i ∈ {1, 2, . . . , n} define
i=1, F,{xi}n
i=1), and define
u′
i = ugi Yx∈F ∩gi
vx ∈ Dλ.
Since
Ad ugi ↾AF = Ad(cid:16) Yx∈F ∩gi
vx(cid:17) ↾AF ,
u′
i is a self-adjoint unitary and commutes with the subalgebra AF .
It is
easy to see that the family {u′
i}n
i=1 mutually commutes. Since xi ∈ gi \ F
and gi is disjoint from gj for j 6= i, we have that wxi commutes with AF
and {u′
i anti-commute because so do wxi and
ugi. Therefore C ∗(u′
i, wxi) is isomorphic to M2(C) for i ∈ {1, 2, . . . , n} by
Lemma 4.1, and the family
j, wxj}j6=i. Finally wxi and u′
{C ∗(u′
i, wxi)}n
i=1 ∪ {AF}
mutually commutes. Since Dλ is generated by these mutually commuting
subalgebras, we get
Dλ =(cid:16) nOi=1
C ∗(u′
i, wxi)(cid:17) ⊗ AF ∼= On+F
M2(C) ∼= M2n+m(C),
30
ILIJAS FARAH AND TAKESHI KATSURA
(cid:3)
as required.
Remark 7.14. For finite g ∈ P(X), we have αg = Ad(cid:0)Qx∈g vx(cid:1). From
this fact, one can show that BX,Γ is not AM if Γ contains a finite nonempty
element g (one can also show that BX,Γ is always AF). Thus in order for BX,Γ
to be AM it is necessary that every g ∈ Γ\{∅} is infinite. One can show that
this is also sufficient although its proof becomes significantly complicated
compared with Proposition 7.13. We shall not need such generality for
proving Theorem 7.17.
Remark 7.15. One can show that BX,Γ is not UHF when X ≥ ℵ1 and
Γ 6= {∅} in a similar way to the proof of Proposition 4.5 (2). We omit the
proof because we do not need this (see the proof of Theorem 7.17 for some
special cases). One can also show that ZBX,Γ(AX ) = C1 holds when every
g ∈ Γ\{∅} is infinite (even in the case χ(AX ) < χ(BX,Γ)). This shows that a
generalization of question [6, Problem 8.3] for nonseparable AM algebras has
a very strong negative answer (see Corollary 4.6). The authors would like
to thank Bruce Blackadar for pointing out the phenomenon ZBX,Γ(AX ) =
C1. This strong phenomenon does not occur for UHF algebras because we
can show χ(ZB(A)) = χ(B) for a subalgebra A of a UHF algebra B with
χ(A) < χ(B), and hence in this case ZB(A) is huge.
Lemma 7.16. For every cardinal κ with X ≤ κ ≤ 2X, there exists a
subgroup Γ ⊆ P(X) with Γ = κ such that every finite subset of Γ is included
in a subgroup generated by g1, g2, . . . , gn ∈ Γ which are infinite and mutually
disjoint.
Proof. Take a subset Y ⊆ P(X) with Y = κ. Let Γ0 be the Boolean subal-
gebra of P(X) generated by Y , that is the smallest subset of P(X) containing
Y and closed under taking unions, intersections and complements. Then Γ0
is a subgroup of P(X) with Γ0 = κ. Choose a bijection ι : X × N → X and
define an injective homomorphism
ϕ : P(X) ∋ g 7→ ι[g × N] ∈ P(X).
Let Γ := ϕ[Γ0] ⊆ P(X). Then every finite subset of Γ is included in a
finite Boolean subalgebra of Γ. If g1, g2, . . . , gn ∈ Γ are the atoms of this
subalgebra then they clearly satisfy the requirements.
(cid:3)
Theorem 7.17. For every pair of infinite cardinals κ and ν with κ ≥ ℵ1
and ν ≤ κ ≤ 2ν, there exists a unital AM algebra of character density κ and
representation density ν which is not UHF.
Proof. For κ = ν ≥ ℵ1, the example BX in Proposition 4.5 for X = κ
is a unital AM algebra of character density κ and representation density ν
which is not UHF. Suppose ν < κ ≤ 2ν . Take a set X with X = ν. By
Lemma 7.16, there exists a subgroup Γ ⊆ P(X) with Γ = κ satisfying
the assumption of Proposition 7.13. Then BX,Γ is a unital AM algebra of
character density κ and representation density ν by Proposition 7.12 and
Proposition 7.13. This is not UHF by Corollary 7.7.
(cid:3)
NONSEPARABLE UHF ALGEBRAS I: DIXMIER'S PROBLEM
31
From Theorem 7.17 we have the following.
Corollary 7.18. There is a unital AM algebra faithfully represented on a
separable Hilbert space that is not a UHF algebra.
(cid:3)
This corollary answers a half of the question raised by Masamichi Take-
saki. The following is the other half which we could not answer.
Problem 7.19. Is there an LM algebra faithfully represented on a separable
Hilbert space which is not AM?
Since χ(cid:0)B(ℓ2(N))(cid:1) = 2ℵ0, by Theorem 1.3 (1) there is no such a C*-
algebra if we assume the continuum hypothesis 2ℵ0 = ℵ1. We do not know
what happens if we do not assume the continuum hypothesis.
Acknowledgment. Many of the results presented in this paper were proved
while both authors were visiting the Fields Institute in Fall 2007 and in early
2008. I.F. would like to thank colleagues from his department, in particular
Juris Stepr¯ans and Man Wah Wong, for making his stay at the Fields Insti-
tute possible. Both authors would like to thank George Elliott, Toshihiko
Masuda, Narutaka Ozawa, N. Christopher Phillips, Juris Stepr¯ans, Reiji
Tomatsu and Andrew Toms for illuminating conversations. We would also
like to thank the anonymous referee for several useful remarks.
References
[1] I. Ben Yaacov, A. Berenstein, C.W. Henson, and A. Usvyatsov, Model theory for
metric structures, Model Theory with Applications to Algebra and Analysis, Vol. II
(Z. Chatzidakis et al., eds.), Lecture Notes series of the London Math. Society., no.
350, Cambridge University Press, 2008, pp. 315 -- 427.
[2] B. Blackadar, Operator algebras, Encyclopaedia of Mathematical Sciences, vol. 122,
Springer-Verlag, Berlin, 2006, Theory of C ∗-algebras and von Neumann algebras,
Operator Algebras and Non-commutative Geometry, III.
[3] O. Bratteli, Inductive limits of finite dimensional C ∗-algebras, Trans. Amer. Math.
Soc. 171 (1972), 195 -- 234.
[4] N. Brown and N. Ozawa, C ∗-algebras and finite-dimensional approximations, Grad-
uate Studies in Mathematics, vol. 88, Amer. Math. Soc., Providence, RI, 2008.
[5] K.R. Davidson, C ∗-algebras by example, Fields Institute Monographs, vol. 6, Amer.
Math. Soc., Providence, RI, 1996.
[6] J. Dixmier, On some C ∗-algebras considered by Glimm, J. Functional Analysis 1
(1967), 182 -- 203.
[7] R. Engelking, General topology, Heldermann, Berlin, 1989.
[8] I. Farah, Graphs and CCR algebras, Indiana Univ. Math. Journal (to appear).
[9] I. Farah, B. Hart, and D. Sherman, Model theory of operator algebras II: Model theory,
preprint, 2010.
[10] I. Farah and T. Katsura, Nonseparable UHF algebras II: Classification, in preparation,
2010.
[11] J.G. Glimm, On a certain class of operator algebras, Trans. Amer. Math. Soc. 95
(1960), 318 -- 340.
[12] R. Haydon and S. Wassermann, A commutation result for tensor products of C*-
algebras, Bull. London Math. Soc. 5 (1973), 283 -- 28.
32
ILIJAS FARAH AND TAKESHI KATSURA
[13] X. Jiang and H. Su, On a simple unital projectionless C ∗-algebra, Amer. J. Math.
121 (1999), no. 2, 359 -- 413.
[14] T. Katsura, Non-separable UHF algebras, RIMS proceedings (2009).
[15] J.A. Packer and I. Raeburn, Twisted crossed products of C ∗-algebras, Math. Proc.
Cambridge Philos. Soc. 106 (1989), no. 2, 293 -- 311.
Department of Mathematics and Statistics, York University, 4700 Keele
Street, North York, Ontario, Canada, M3J 1P3, and Matematicki Institut,
Kneza Mihaila 34, Belgrade, Serbia
URL: http://www.math.yorku.ca/∼ifarah
E-mail address: [email protected]
Department of Mathematics, Faculty of Science and Technology, Keio Uni-
versity, 3-14-1 Hiyoshi, Kouhoku-ku, Yokohama, JAPAN, 223-8522
E-mail address: [email protected]
|
1608.03515 | 1 | 1608 | 2016-08-11T16:06:35 | Eta-diagonal distributions and infinite divisibility for R-diagonals | [
"math.OA",
"math.CO",
"math.PR"
] | The class of R-diagonal *-distributions is fairly well understood in free probability. In this class, we consider the concept of infinite divisibility with respect to the operation $\boxplus$ of free additive convolution. We exploit the relation between free probability and the parallel (and simpler) world of Boolean probability. It is natural to introduce the concept of an eta-diagonal distribution that is the Boolean counterpart of an R-diagonal distribution. We establish a number of properties of eta-diagonal distributions, then we examine the canonical bijection relating eta-diagonal distributions to infinitely divisible R-diagonal ones. The overall result is a parametrization of an arbitrary $\boxplus$-infinitely divisible R-diagonal distribution that can arise in a C*-probability space, by a pair of compactly supported Borel probability measures on $[ 0, \infty )$. Among the applications of this parametrization, we prove that the set of $\boxplus$-infinitely divisible R-diagonal distributions is closed under the operation $\boxtimes$ of free multiplicative convolution. | math.OA | math |
ETA-DIAGONAL DISTRIBUTIONS AND INFINITE
DIVISIBILITY FOR R-DIAGONALS
HARI BERCOVICI, ALEXANDRU NICA, MICHAEL NOYES, AND KAMIL SZPOJANKOWSKI
Abstract. The class of R-diagonal ∗-distributions is fairly well understood in free prob-
ability. In this class, we consider the concept of infinite divisibility with respect to the
operation ⊞ of free additive convolution. We exploit the relation between free proba-
bility and the parallel (and simpler) world of Boolean probability.
It is natural to in-
troduce the concept of an η-diagonal distribution that is the Boolean counterpart of an
R-diagonal distribution. We establish a number of properties of η-diagonal distributions,
then we examine the canonical bijection relating η-diagonal distributions to infinitely di-
visible R-diagonal ones. The overall result is a parametrization of an arbitrary ⊞-infinitely
divisible R-diagonal distribution that can arise in a C ∗-probability space, by a pair of com-
pactly supported Borel probability measures on [0, ∞). Among the applications of this
parametrization, we prove that the set of ⊞-infinitely divisible R-diagonal distributions is
closed under the operation ⊠ of free multiplicative convolution.
1. Introduction
Free additive convolution ⊞ is a binary operation on the set P of Borel probability
measures on R, reflecting the addition operation for free selfadjoint elements in a noncom-
mutative probability space. The properties of this operation parallel in many respects the
ones of the usual convolution on P, for instance in the treatment of infinite divisibility.
One way to approach ⊞-infinite divisibility is to use a bijection constructed in [4] which re-
lates free independence to another form of noncommutative independence, namely Boolean
independence. In this paper we focus on probability measures with compact support, so we
view this bijection as a map B : Pc → P (inf−div)
, where Pc is the set of probability measures
with compact support on R, while P (inf−div)
consists of those measures µ ∈ Pc which are
⊞-infinitely divisible, that is, have the property that for every n ∈ N, there exists µn ∈ Pc
satisfying
c
c
(1.1)
µn ⊞ · · · ⊞ µn
= µ.
n
{z
}
The bijection B connects the fundamental transforms of free and Boolean probability, the
R-transform and respectively the η-series. For µ ∈ Pc, both of these transforms Rµ(z) and
ηµ(z) are convergent power series. The bijection B is described by the equation
(1.2)
RB(µ) = ηµ, µ ∈ Pc.
More precisely, for every µ ∈ Pc there exists a uniquely determined measure ν ∈ P (inf−div)
such that Rν = ηµ, and one defines B(µ) := ν.
c
At the level of compactly supported distributions, the bijection B is precisely the para-
metrization of ⊞-infinitely divisible distributions provided in [14]. This was extended in [4]
HB: supported in part by a grant from the National Science Foundation of the USA.
AN: research supported by a Discovery Grant from NSERC, Canada.
1
2
H. BERCOVICI, A. NICA, M. NOYES, AND K. SZPOJANKOWSKI
to the space P of all Borel probability measures on R. In a different direction, the bijection
B was extended in [2] to the space of joint distributions for k-tuples of selfadjoint elements
in a C ∗-probability space. Our goal in this paper is to use a multivariate version of the
bijection B in order to study ⊞-infinitely divisibile R-diagonal distributions, a significant
class of ∗-distributions considered in free probability.
To explain our results, we introduce some notation. We let Dc(1, ∗) stand for the collec-
tion of all ∗-distributions of (generally, not selfadjoint) elements in a (generally not tracial)
C ∗-probability space. There is a natural operation ⊞ on Dc(1, ∗) which corresponds to the
addition a + b of two variables a, b in the same space such that {a, a∗} is free from {b, b∗}.
Infinite divisibility in Dc(1, ∗) is defined as in (1.1), and we denote by Dc(1, ∗)(inf−div) the
collection of ⊞-infinitely divisible elements of Dc(1, ∗). The notions of R-transform and
η-series also have natural extensions to the context of ∗-distributions.
The results of [2], specialized to two selfadjoint variables, can be applied to Dc(1, ∗) after
a simple change of coordinates. There is again a bijection B(1,∗) : Dc(1, ∗) → Dc(1, ∗)(inf−div)
defined by the requirement that
(1.3)
RB(1,∗)(µ) = ηµ, µ ∈ Dc(1, ∗).
This is analogous to the condition (1.2) satisfied by the original bijection B, but proving
the existence of B(1,∗) is more than a trivial extension of the proof for B, and requires a
mixture of combinatorial and analytic methods.
We turn now to R-diagonal ∗-distributions, which can be succinctly described as the
distributions in Dc(1, ∗) that are invariant under multiplication by a free Haar unitary
(see [11, Theorem 15.10, p. 244]). For our purposes, it is more useful to consider the
original definition [10] of R-diagonal distributions which asks that the R-transform of the
distribution be in some sense 'diagonal' [11, Definition 15.3, p. 241]. From this point of
view, it is clear how to define the Boolean counterpart of R-diagonality: we simply say that
a ∗-distribution is η-diagonal if its η-series is diagonal. The map B
(1,∗) defined by (1.3) will
then give a bijection between the set of all η-diagonal distributions in Dc(1, ∗) and the set
of R-diagonal distributions in Dc(1, ∗) which are ⊞-infinitely divisible.
The above discussion shows that there is some interest in studying η-diagonal distribu-
tions. In this paper we point out a few general algebraic and combinatorial properties of
such a distribution µ, which actually hold for µ in a larger, purely algebraic space Dalg(1, ∗).
The property of a distribution µ ∈ Dalg(1, ∗) of being η-diagonal has an elegant description
phrased directly in terms of the ∗-moments of µ. This result (Theorem 2.8) is reminiscent
of (but simpler than) the description [9, Theorem 1.2.1] of R-diagonal variables in terms of
their ∗-moments. The η-diagonal distributions also have other algebraic and combinatorial
properties that are analogous to known properties of R-diagonal distributions. In particu-
lar, if a is an η-diagonal element in a ∗-probability space (A, ϕ) (which means, by definition,
that a has η-diagonal ∗-distribution with respect to ϕ) then it follows that aa∗ and a∗a are
Boolean independent elements of A, and that the coefficients of the η-series of aa∗ and a∗a
are read from the so-called determining sequences for the ∗-distribution of a. For details on
the terms used above and for a discussion of why this is indeed analogous to known facts
about R-diagonals, see Remark 3.4 below.
In the case in which the η-diagonal distribution µ is in Dc(1, ∗), we point out a natural
parametrization for µ, given by a pair of compactly supported Borel probability measures
on [0, ∞). That is, we establish a canonical bijection
(1.4)
{µ ∈ Dc(1, ∗) : µ is η-diagonal} ∋ µ ↔ (σ1, σ2) ∈ P +
c × P +
c ,
ETA-DIAGONALS AND INFINITE DIVISIBILITY FOR R-DIAGONALS
3
:= {σ ∈ Pc : σ(cid:0) [0, ∞)(cid:1) = 1}. Without going into details, we mention that all
where P +
c
the ∗-distributions appearing in this paper are defined as linear functionals on the algebra
ChZ, Z ∗i of complex polynomials in the non-commuting indeterminates Z and Z ∗, and that
the correspondence µ ↔ (σ1, σ2) from (1.4) amounts to the equalities
(1.5)
µ((ZZ ∗)n) =Z ∞
0
tn dσ1(t) and µ((Z ∗Z)n) =Z ∞
0
tn dσ2(t), n ∈ N.
In other words, the probability measures σ1 and σ2 which parametrize µ in (1.4) are simply
the distributions of ZZ ∗ and of Z ∗Z with respect to the functional µ. The relevant point here
is that for any given σ1, σ2 ∈ P +
c there exists a unique η-diagonal distribution µ ∈ Dc(1, ∗)
such that (1.5) holds.
When the bijection B(1,∗) is applied to {µ ∈ Dc(1, ∗) : µ is η-diagonal} in (1.4), we obtain
a bijection
(1.6)
nν ∈ Dc(1, ∗) :
ν is R-diagonal and
⊞-infinitely divisible o ∋ ν ↔ (σ1, σ2) ∈ P +
c × P +
c .
Thus, we have a parametrization of a general ⊞-infinitely divisible R-diagonal distribution
by a pair of probability measures from P +
c . Analogously to (1.5), it is possible to write
explicitly the relation connecting σ1, σ2 to the distributions of the elements ZZ ∗ and Z ∗Z
in the noncommutative probability space (ChZ, Z ∗i, ν). Theorem 6.4 below realizes this
parametrization by providing precise formulas for the R-transforms of ZZ ∗ and of Z ∗Z.
An important subclass of R-diagonal distributions are those which satisfy the KMS con-
dition for some parameter t ∈ (0, ∞). This is a generalization of the trace condition, where
the latter corresponds to the special case t = 1 (see the review in Section 3 below). For an
R-diagonal distribution ν which satisfies KMS with parameter t, one can process further
the result of Theorem 6.4 in order to obtain explicit formulas for the distributions of ZZ ∗
and of Z ∗Z, in terms of the probability measures σ1 and σ2 which parametrize ν. These
formulas invoke some commonly used elements of free harmonic analysis on P +
c , and are
given in Proposition 6.8.
As an application of the parametrization from (1.6), we prove that the set of ⊞-infinitely
divisible R-diagonal ∗-distributions is closed under the operation ⊠ of multiplicative con-
volution.
In addition to the present introduction, the paper contains 6 sections. Section 2 intro-
duces η-diagonal ∗-distributions and discusses some of their algebraic properties. Section 3
is devoted to a review of R-diagonal ∗-distributions, with emphasis on facts that are needed
in the present paper. In Section 4 we verify that the bijection B(1,∗) does indeed work on
Dc(1, ∗) in the way described in (1.3). Section 5 presents the operator model for η-diagonals
that was announced in (1.4). Section 6 contains our results concerning the parametriza-
tion of infinitely divisible R-diagonal distributions, and a discussion of the KMS example.
Finally, Section 7 discusses the application to free multiplicative convolution.
2. η-series and η-diagonal ∗-distributions
Notation 2.1. (1) We denote by W + the setF∞
n=1{1, ∗}n consisting of all non-empty words
over the two-letter alphabet {1, ∗}. This is a semigroup (without unit) under the natural
operation of concatenation. We denote by w the number of letters in a word w ∈ W +.
4
H. BERCOVICI, A. NICA, M. NOYES, AND K. SZPOJANKOWSKI
(2) The algebra of complex polynomials in two non-commuting variables Z and Z ∗ is
denoted, as usual, by ChZ, Z ∗i. For every word w = (ℓ1, . . . , ℓn) ∈ W + we write
Z w = Z ℓ1 · · · Z ℓn ∈ ChZ, Z ∗i.
The set {1} ∪ {Z w : w ∈ W +} is a basis of ChZ, Z ∗i as a complex vector space.
(3) An algebraic ∗-distribution is a linear functional µ : ChZ, Z ∗i → C such that µ(1) = 1.
(At this stage we do not require µ to have any additional properties.) The values of µ on
monomials Z w (with w ∈ W +) will be referred to as ∗-moments of µ.
(4) The collection of all algebraic ∗-distributions from (3) is denoted Dalg(1, ∗).
Notation 2.2. (Series and their coefficients.) (1) The algebra of formal power series in two
non-commuting indeterminates z and z∗ is denoted, as usual, by Chhz, z∗ii. The collection
C0hhz, z∗ii ⊂ Chhz, z∗ii of power series with vanishing constant coefficient is a two-sided
ideal in Chhz, z∗ii. An arbitrary element f ∈ C0hhz, z∗ii is of the form
(2.1)
f (z, z∗) =
∞Xn=1 Xℓ1,...,ℓn∈{1,∗}
α(ℓ1,...,ℓn)zℓ1 · · · zℓn = Xw∈W +
αwzw,
where the coefficients αw are complex numbers, and for w = (ℓ1, . . . , ℓn) ∈ W + we use the
notation zw = zℓ1 · · · zℓn.
(2) Given w ∈ W +, we denote by Cfw : C0hhz, z∗ii → C the linear functional which
extracts the coefficient of zw from a series f ∈ C0hhz, z∗ii. That is, if f is given by (2.1),
we have Cfw(f ) = αw, w ∈ W +.
(3) Given a positive integer n, a word w = (ℓ1, . . . , ℓn) ∈ W +, and a partition π of
{1, . . . , n}, we define a functional (non-linear unless π consists of only one block) Cfw;π :
C0hhz, z∗ii → C, as follows. For every block B = {b1, . . . , bm} of π, where 1 ≤ b1 < · · · <
bm ≤ n, we set
wB = (ℓ1, . . . , ℓn)B := (ℓb1, . . . , ℓbm) ∈ {1, ∗}m.
Then we define
(2.2)
Cfw;π(f ) := YB∈π
CfwB(f ),
f ∈ C0hhz, z∗ii.
[Suppose, for instance, that n = 5, π = {{1, 4, 5}, {2, 3}}, and w = (ℓ1, . . . , ℓ5). Then
Cfw;π(f ) = Cf(ℓ1,ℓ4,ℓ5)(f ) · Cf(ℓ2,ℓ3)(f ), f ∈ C0hhz, z∗ii.]
Definition and Remark 2.3. (Moment series, η-series.) Fix µ ∈ Dalg(1, ∗). (1) The
moment series of µ is defined as Mµ :=Pw∈W + µ(Z w)zw ∈ C0hhz, z∗ii.
(2) The η-series of µ is defined as
(2.3)
ηµ := Mµ(1 + Mµ)−1 = (1 + Mµ)−1Mµ ∈ C0hhz, z∗ii,
where all the algebraic operations are performed in the algebra Chhz, z∗ii.
(3) It is immediate from (2.3) that the series Mµ can be retrieved from ηµ by the formula
(2.4)
Mµ = ηµ(1 − ηµ)−1 = (1 − ηµ)−1ηµ.
(4) The right-hand side of (2.4) can be written as a geometric series P∞
converges in the sense that the series P∞
µ (which
µ) contains only finitely many non-zero
n=1 Cfw(ηn
n=1 ηn
ETA-DIAGONALS AND INFINITE DIVISIBILITY FOR R-DIAGONALS
5
terms for every w ∈ W +). This leads to an explicit formula for the coefficients of Mµ in
terms of those of ηµ, namely
Cfw(Mµ) = Xπ∈Int(n)
Cfw;π(ηµ), w ∈ W + with w = n.
(2.5)
(2.6)
Here Int(n) denotes the set of interval partitions of {1, . . . , n}, that is, partitions which
have the property that every block B of π is of the form {a, a + 1, . . . , b} for some a ≤ b in
{1, . . . , n}. An analogous argument converts (2.3) into the formula
Cfw(ηµ) = Xπ∈Int(n)
(−1)1+πCfw;π(Mµ), w ∈ W + with w = n,
where π denotes the number of blocks of the partition π.
Remark 2.4. It is clear that the map Dalg(1, ∗) ∋ µ 7→ Mµ ∈ C0hhz, z∗ii is bijective.
Equations (2.3) and (2.4) show that the map Dalg(1, ∗) ∋ µ 7→ ηµ ∈ C0hhz, z∗ii is a bijection
as well. In other words, we can define a distribution µ ∈ Dalg(1, ∗) by specifying its η-series.
Definition 2.5. A word w ∈ W + is said to be alternating when it is of the form
) = (∗, 1)m,
) = (1, ∗)m or w = (∗, 1, ∗, 1, . . . , ∗, 1
w = (1, ∗, 1, ∗, . . . , 1, ∗
2m
{z
}
2m
{z
}
for some positive integer m.
In the first case w is said to be of type (1, ∗), and in the
second case w is said to be of type (∗, 1). In these formulas, powers are taken relative to
concatenation. Note in particular that alternating words have positive, even length.
Definition 2.6. (1) A distribution µ ∈ Dalg(1, ∗) is said to be η-diagonal if Cfw(ηµ) = 0
for every word w ∈ W + which is not alternating.
(2) If µ is η-diagonal, its η-series is thus of the form
ηµ(z, z∗) =
αn(zz∗)n +
βn(z∗z)n,
∞Xn=1
∞Xn=1
with αn, βn ∈ C for n ∈ N. The sequences (αn)∞
ing sequences of the η-diagonal distribution µ.
n=1 and (βn)∞
n=1 will be called the determin-
The main goal of the present section is to reveal an equivalent characterization of η-
diagonal distributions, which is phrased directly in terms of ∗-moments. For this purpose,
we require one more concept related to words in W +.
Definition and Remark 2.7. (1) A word w = (ℓ1, . . . , ℓn) ∈ W + is said to be mixed-
alternating if n = 2m is even and if the letters of w are such that ℓ2k−1 6= ℓ2k, for k =
1, 2, . . . , m. Equivalently, w is mixed-alternating when it belongs to the concatenation
subsemigroup of W + generated by the words (1, ∗) and (∗, 1).
(2) By grouping factors, it is easily seen that every mixed-alternating word w can be
written in a unique way as a concatenation of alternating words such that consecutive
words are of different types. Indeed, if we write such a word as
(2.7)
w = w1w2 · · · wp
where p ≥ 1, each wi is alternating, and wi is not of the same type as wi+1, i = 1, 2, . . . , p−1,
then the boundaries between the words w1, . . . , wp can be retrieved at the places where w
has two consecutive identical letters.
[For example, w = (1, ∗, 1, ∗, 1, ∗, ∗, 1, ∗, 1, 1, ∗, ∗, 1, ∗, 1) is mixed-alternating and its canoni-
cal factorization (2.7) is w1w2w3w4 with w1 = (1, ∗)3, w2 = (∗, 1)2, w3 = (1, ∗), w4 = (∗, 1)2.]
6
H. BERCOVICI, A. NICA, M. NOYES, AND K. SZPOJANKOWSKI
Theorem 2.8. For every distribution µ ∈ Dalg(1, ∗), the statements (a) and (b) are equiv-
alent.
(a) µ is η-diagonal.
(b) µ satisfies the following conditions1:
(ηDM1) Whenever w ∈ W + is not mixed-alternating, it follows that µ(Z w) = 0.
(ηDM2) Whenever w = w1 · · · wp ∈ W + is mixed-alternating and factored as in (2.7),
it follows that µ(Z w) = µ(Z w1) · · · µ(Z wp).
The proof of the implication (b) ⇒ (a) in the above theorem requires two auxilliary
results.
Lemma 2.9. Suppose that a distribution µ ∈ Dalg(1, ∗) satisfies the condition (ηDM1).
Then Cfw(ηµ) = 0 for every word w ∈ W + that is not mixed-alternating.
Proof. Suppose that w is not mixed alternating and w = n. We prove that Cf w(ηµ) =
0 by showing that each term in the right-hand side of (2.6) vanishes.
Indeed, let π =
{J1, . . . , Jm} ∈ Int(n) be a partition where the intervals J1, . . . , Jm are listed in increasing
order. Observe that w = (wJ1) · · · (wJm) (concatenation product). Since w is not mixed-
alternating, there must exist an index 1 ≤ k ≤ m such that wJk is not mixed-alternating.
For this k, condition (ηDM1) yields CfwJk(Mµ) = 0. Therefore the term indexed by π in
(2.6) vanishes as well because CfwJk(Mµ) is one of its factors. The lemma follows.
(cid:3)
Lemma 2.10. Let µ, ν ∈ Dalg(1, ∗) be such that
(1) Both µ and ν satisfy conditions (ηDM1) and (ηDM2), and
(2) Cfw(ηµ) = Cfw(ην ) for every alternating word w ∈ W +.
Then µ = ν.
In fact, it suffices to prove this equality when w is alternating.
Proof. If w ∈ W + is not mixed-alternating then µ(Z w) = ν(Z w) = 0 because µ and ν
satisfy (ηDM1). Thus it suffices to verify that µ(Z w) = ν(Z w) for mixed-alternating words
w.
Indeed, suppose for
the moment that the equality has been proved for alternating words and let w ∈ W + be
a mixed-alternating word. Consider the canonical factorization w = w1 · · · wp indicated in
(2.7). We have
µ(Z w) = µ(Z w1) · · · µ(Z wp) (by (ηDM2) for µ)
= ν(Z w1) · · · ν(Z wp) (by assumption on alternating moments)
= ν(Z w) (by (ηDM2) for ν).
We conclude the proof by showing that µ(Z w) = ν(Z w) for every alternating word w.
By symmetry, it suffices to verify that µ((Z ∗Z)m) = ν((Z ∗Z)m) for every m ∈ N Fix m
and write µ((Z ∗Z)m) and ν((Z ∗Z)m) as sums indexed by Int(2m), in the way indicated
in (2.5). We show that for every π ∈ Int(2m), the terms indexed by π in the two sums
(for µ and for ν) are equal to each other. If π has a block B of odd cardinality, then the
terms we are looking at are both equal to 0 because they include the factors CfwB(ηµ) and
respectively CfwB(ην ), and these factors are zero by Lemma 2.9. If all the blocks of π are
even, we write π = {J1, . . . , Jk} where the intervals J1, . . . , Jk are listed in increasing order,
and where J1 = 2d1, . . . , Jk = 2dk for some d1, . . . , dk ∈ N. The terms indexed by π in
1The acronym ηDM is meant to suggest η-Diagonality-in-Moments.
ETA-DIAGONALS AND INFINITE DIVISIBILITY FOR R-DIAGONALS
7
the two sums we consider are then
(2.8)
Cf(∗,1)di (ηµ) and respectively
Cf(∗,1)di (ην)
kYi=1
kYi=1
where we used the fact that (∗, 1)mJi = (∗, 1)di , for 1 ≤ i ≤ k. The two products in (2.8)
are indeed equal by assumption (2) in the statement.
(cid:3)
Proof of Theorem 2.8. Suppose first that µ is η-diagonal. We verify that it satisfies (ηDM1)
and (ηDM2). The hypothesis on µ says in particular that Cfw(ηµ) = 0 for every w ∈ W +
which is not mixed-alternating. To prove (ηDM1), we must show that Cfw(Mµ) = 0 for
every such word. This argument is carried precisely as in the proof of Lemma 2.9, with the
roles of Mµ and ηµ being reversed and with (2.5) in place of (2.6). The reader will have no
difficulty verifying the details.
In order to show that µ also satisfies (ηDM2), fix a mixed-alternating word w ∈ W +,
with canonical factorization w = w1 · · · wp as in (2.7). Set n = w = w1 + · · · + wp, and
let ρ0 be the partition of {1, . . . , n} into intervals J1, . . . , Jp (written in increasing order)
with lengths J1 = w1, . . . , Jp = wp. Given a partition π of {1, . . . , n}, we write π ≤ ρ0
if every block B of π is contained in one of the blocks J1, . . . , Jp of ρ0. (This relation is
usually called the reverse refinement order on partitions.)
Next, we use (2.5) to express the coefficient Cfw(Mµ) = µ(Z w) as a sum indexed by
Int(n). The special structure of the coefficients of ηµ implies that a partition π ∈ Int(n) has
a zero contribution to that sum unless π ≤ ρ0. It is immediate that the partitions π ∈ Int(n)
satisfying π ≤ ρ0 are in natural bijective correspondence to tuples of partitions (π1, . . . , πp)
where π1 ∈ Int(J1), . . . , πp ∈ Int(Jp). This correspondence is such that for π ↔ (π1, . . . , πp)
we have
These observations lead to the formula
Cfw;π(ηµ) = Cfw1;π1(ηµ) · · · Cfwp;πp(ηµ).
µ(Z w) = Xπ1∈Int(J1),...,
πp∈Int(Jp)
Cfw1;π1(ηµ) · · · Cfwp;πp(ηµ) =
pYi=1(cid:16) Xπi∈Int(Ji)
Cfwi;πi(ηµ)(cid:17).
In the latter product, one more application of (2.5) identifies
Xπi∈Int(Ji)
Cfwi;πi(ηµ) = µ(Z wi), 1 ≤ i ≤ p,
i=1 µ(Z wi).
thus implying the desired conclusion that µ(Z w) =Qp
Conversely, assume now that µ satisfies conditions (ηDM1) and (ηDM2). We show that it
is η-diagonal by an indirect argument: we construct an η-diagonal distribution ν ∈ Dalg(1, ∗)
and then prove that µ = ν. The distribution ν is defined by specifying its η-series (see
Remark 2.4), namely Cfw(ην) = Cfw(ηµ) if w is alternating and Cfw(ην) = 0 otherwise. To
prove that µ = ν we show that µ and ν satisfy the hypothesis of Lemma 2.10. Indeed, both
µ and ν satisfy (ηDM1) and (ηDM2): µ does so by hypothesis, while ν does so because it
is η-diagonal and by virtue of the implication (a) ⇒ (b) proved above. On the other hand,
if w ∈ W + is an alternating word, the equality Cfw(ηµ) = Cfw(ην) is true by the definition
of ν. This concludes the proof of the theorem.
(cid:3)
8
H. BERCOVICI, A. NICA, M. NOYES, AND K. SZPOJANKOWSKI
Remark 2.11. Let (A, ϕ) be a noncommutative probability space (that is, A is a unital
algebra over C, ϕ : A → C is a linear functional, and ϕ(1) = 1), and let a1, a2 ∈ A. Recall
[13] that a1, a2 are said to be Boolean independent provided that, given positive integers
n, p1, . . . , pn and indices i1, . . . , in ∈ {1, 2} such that ik 6= ik+1 for i = 1, . . . , n − 1, the
following identity is satisfied:
ϕ(ap1
i1
· · · apn
in ) = ϕ(ap1
i1
) · · · ϕ(apn
in ).
Now consider the noncommutative probability space (ChZ, Z ∗i, µ), where µ ∈ Dalg(1, ∗)
is η-diagonal. Condition (ηDM2) of Theorem 2.8 can be restated as saying that ZZ ∗ and
Z ∗Z are Boolean independent in (ChZ, Z ∗i, µ).
Remark and Notation 2.12. Another relevant fact concerning an η-diagonal ∗-distribu-
tion µ concerns the individual η-series of ZZ ∗ and Z ∗Z in the noncommutative probability
space (ChZ, Z ∗i, µ). If a is an element in a noncommutative probability space (A, ϕ), then
its moment series and η-series Ma, ηa ∈ C[[z]] are defined as in Definition 2.3 but using
moments in place of ∗-moments: first we set Ma(z) =P∞
ηa(z) = Ma(z)/(1 + Ma(z)) ∈ C[[z]].
n=1 ϕ(an)zn, and then define
The coefficients of Ma and ηa are related to each other via summations over interval par-
titions which are analogous to those shown in (2.5), (2.6) (and are derived the same way,
by starting from the algebraic relations satisfied by the series themselves). We explicitly
record here the analogue of (2.6):
(2.9)
Cf n(ηa) = Xρ∈Int(n)
(−1)1+ρYB∈ρ
ϕ(aB), n ∈ N,
where (by analogy with Notation 2.2(2)) we use the notation Cfn : C[[z]] → C for the linear
map that extracts the nth coefficient of a series in C[[z]].
When applied to the elements ZZ ∗ and Z ∗Z from the framework of Theorem 2.8, these
observations yield the following result.
n=1, (βn)∞
Proposition 2.13. Suppose that µ ∈ Dalg(1, ∗) is an η-diagonal distribution, and let
(αn)∞
n=1 be its determining sequences (as introduced in Definition 2.6(2)). Then
in the noncommutative probability space (ChZ, Z ∗i, µ), the elements ZZ ∗ and Z ∗Z have
η-series given by
αnzn and ηZ ∗ Z (z) =
βnzn.
∞Xn=1
ηZZ ∗ (z) =
∞Xn=1
Cfn(ηZZ ∗ ) = Xρ∈Int(n)
Proof. By symmetry, it suffices to prove the first formula. Equation (2.9) yields
(2.10)
(−1)1+ρYB∈ρ
µ((ZZ ∗)B), n ∈ N.
For the remainder of the proof, we fix n ∈ N and verify that the right-hand side of (2.10) is
equal to αn.
For every partition ρ = {J1, . . . , Jk} ∈ Int(n), with intervals J1, . . . , Jk written in increas-
ing order, we define a doubled partition bρ = {bJ1, . . . ,cJk} ∈ Int(2n). This is the interval
partition uniquely determined by the requirement that bJ1, . . . ,cJk come in increasing order
and satisfy bJi = 2Ji for 1 ≤ i ≤ k. With this notation, it is easily seen that the right-hand
ETA-DIAGONALS AND INFINITE DIVISIBILITY FOR R-DIAGONALS
9
side of (2.10) can be written as
This, however, is the same as
Xρ∈Int(n)
Xπ∈Int(2n)
(−1)1+bρCf(1,∗)n;bρ(Mµ).
(−1)1+πCf(1,∗)n;π(Mµ).
Indeed, due to the special structure of the ∗-moments of µ described in Theorem 2.8, all
the terms in the latter sum, corresponding to partitions π ∈ Int(2n) which are not of the
form bρ, are equal to 0. We conclude that
Cfn(ηZZ ∗ ) = Xπ∈Int(2n)
(−1)1+πCf(1,∗)n;π(Mµ) = Cf(1,∗)n (ηµ) = αn,
where (2.6) is used in the second equality.
(cid:3)
3. R-transforms and R-diagonal ∗-distributions
The discussion in Section 2 is better put into perspective when one compares it to the
parallel (more elaborate) free probability framework.
In the free probability framework,
instead of η-series one works with R-transforms, and one has the concept of what it means
for a ∗-distribution µ ∈ Dalg(1, ∗) to be R-diagonal. The class of R-diagonal ∗-distributions
is in fact rather well-studied in the free probability literature. Here we review some of their
basic properties, mostly following [11, Lecture 15], and with emphasis on the facts we need
in the present work.
Remark 3.1. On a combinatorial level, switching to the world of free probability comes
to using non-crossing partitions instead of interval partitions. We recall that a crossing of
a partition π of {1, . . . n} consists of integers 1 ≤ a < b < c < d ≤ n such that the set {a, c}
is contained in a block of π and {b, d} is contained in a different block of π. A partition is
non-crossing if it has no crossings. We denote by N C(n) the collection of all non-crossing
partitions of {1, . . . n}.
Similarly to the lattice Int(n), the set N C(n) is partially ordered by reverse refinement.
The minimal and maximal elements with respect to this partial order are denoted by 0n
(the partition of {1, . . . , n} into n singleton blocks) and respectively 1n (the partition of
{1, . . . , n} into one block).
We record a notation and an elementary observation needed in the final part of this
section. For every n ∈ N, we denote by N CE(2n) the collection of all the partitions
π ∈ N C(2n) with the property that every block of π has even cardinality. Observe that
if π ∈ N CE(2n) and if V = {k1 < k2 < · · · < k2m} is a block of π, then the numbers
k1, k2, . . . , k2m have alternating parities. Indeed, for every i = 1, . . . , 2m − 1, the set {ki +
1, ki +2, . . . , ki+1 −1} is a union of blocks of π, and hence has even cardinality, which implies
that ki+1 is of opposite parity from ki.
For a discussion of other elementary facts concerning N C(n), we refer to [11, Lecture 9].
Remark 3.2. (Review of R-transforms.)
(1) The R-transform of a ∗-distribution µ ∈ Dalg(1, ∗) is the series Rµ ∈ C0hhz, z∗ii, whose
coefficients are uniquely determined by the requirement that they relate to the ∗-moments
10
H. BERCOVICI, A. NICA, M. NOYES, AND K. SZPOJANKOWSKI
of µ by the formula
(3.1)
Cfw(Mµ) = Xπ∈N C(n)
Cfw;π(Rµ), w ∈ W + and n = w.
Equation (3.1) is the free probabilistic counterpart of (2.5). It is often referred to as the
moment-cumulant formula for free cumulants (see [11, Lecture 11] for an explanation of
this terminology).
One can also define the series Rµ by an equation involving the series Mµ and Rµ them-
selves (rather than their coefficients, as in (3.1)). More precisely, Rµ is the unique series in
C0hhz, z∗ii that satisfies the functional equation
(3.2)
Rµ(z(1 + Mµ(z, z∗)), z∗(1 + Mµ(z, z∗)) = Mµ(z, z∗)
(see [11, Corollary 16.16]). This is the free probabilistic analogue of (2.3), but now we only
have an implicit functional equation rather than an explicit formula describing the series
Rµ.
An easy inductive argument in the moment-cumulant formula (3.1) shows that one can
recover Mµ from Rµ and that (as in Remark 2.4) we have a bijection
Dalg(1, ∗) ∋ µ 7→ Rµ ∈ C0hhz, z∗ii.
In other words, one can define a distribution µ ∈ Dalg(1, ∗) by specifying its R-transform.
(2) Consider the framework of Remark 2.11, where we discussed the moment series and
η-series Ma, ηa ∈ C[[z]] associated to an element a in a noncommutative probability space
(A, ϕ). In that framework one also has an R-transform associated with the element a ∈ A.
This is the series Ra ∈ C[[z]] that relates to Ma by
(3.3)
and
(3.4)
Cfn(Ma) = Xπ∈N C(n)
Cfn;π(Ra), n ∈ N,
Ra(z(1 + Ma(z))) = Ma(z).
These formulas are analogous to (3.1) or (3.2), respectively. For a detailed discussion of the
algebraic aspects of R-transforms (covering both the series Rµ in part (1) of this remark
and the series Ra in part (2)), see [11, Lecture 16].
Definition 3.3. (1) A ∗-distribution µ ∈ Dalg(1, ∗) is said to be R-diagonal when Cfw(Rµ) =
0 for every word w ∈ W + that is not alternating.
(2) If µ is R-diagonal, its R-transform is thus of the form
(3.5)
Rµ(z, z∗) =
αn(zz∗)n +
βn(z∗z)n,
∞Xn=1
∞Xn=1
with αn, βn ∈ C for n ∈ N. The sequences (αn)∞
sequences of µ.
n=1 and (βn)∞
n=1 are called the determining
Remark 3.4. The concept of an η-diagonal ∗-distribution from Section 2 obviously parallels
the one of an R-diagonal ∗-distribution, with the η-series in place of the R-transform. The
basic properties of η-diagonal distributions proved in Section 2 are the counterparts of
known facts concerning R-diagonal distributions, as noted below.
(1) Remark 2.11 is the Boolean counterpart of [11, Corollary 15.11, p. 244]:
if µ is
R-diagonal, then Z ∗Z and ZZ ∗ are freely independent in the noncommutative probability
space (ChZ, Z ∗i, µ).
ETA-DIAGONALS AND INFINITE DIVISIBILITY FOR R-DIAGONALS
11
(2) Theorem 2.8 is the counterpart of [9, Theorem 1.2.1] which describes R-diagonal
distributions in terms of their ∗-moments.
(3) Proposition 2.13 is analogous to [11, Proposition 15.6, p. 241] which gives a precise
formula, first found in [7], for the coefficients of the one-variable R-transforms RZZ ∗ and
RZ ∗Z in terms of the determining sequences (αn)∞
n=1 from (3.5). This formula
is more elaborate than the relation found in Proposition 2.13 for η-diagonal elements. It
states that
n=1 and (βn)∞
(3.6)
Cfn(RZZ ∗) =
Xπ={V1,...,Vk}∈N C(n)
αV1βV2 · · · βVk,
where the blocks of π are arranged so 1 ∈ V1. The coefficients of RZ ∗Z are obtained by
interchanging the roles of α and β in these formulas. For example, the first three coefficients
of RZZ ∗ are α1, α2 + α1β1 and α3 + 2α2β1 + α1β2 + α1β2
1 .
Equations (3.6) lead to the following observation: an easy induction on n (where one
singles out the terms indexed by the partition 1n ∈ N C(n) on the right-hand sides of these
equations) shows that the determining sequences (αn)∞
n=1 can be retrieved from
the coefficients of RZZ ∗ and RZ ∗Z . Hence the R-diagonal ∗-distribution µ ∈ Dalg(1, ∗) is
completely determined by these R-transforms.
n=1 and (βn)∞
In Section 6 we require a reformulation of (3.6) in terms of operations with series rather
than individual coefficients. This reformulation is given in the next proposition. The for-
mulas (3.7) bear a striking resemblance to the functional equation (3.4) of the R-transform.
In fact, (3.7) collapse to RZZ ∗(z) = RZ ∗Z(z) = Ma(z) in the special case αn = βn, n ∈ N,
in which case we can take a = b.
Proposition 3.5. Let µ ∈ Dalg(1, ∗) be an R-diagonal ∗-distribution with determining
sequences (αn)∞
n=1. Suppose we are given some elements a and b in a noncom-
mutative probability space (A, ϕ) such that
n=1 and (βn)∞
Ra(z) =
∞Xn=1
αnzn and Rb(z) =
βnzn.
∞Xn=1
Then, in the noncommutative probability space (ChZ, Z ∗i, µ), we have
(3.7)
RZZ ∗(z) = Ra(z(1 + Mb(z))) and RZ ∗Z (z) = Rb(z(1 + Ma(z))).
Proof. The argument is analogous to the proof of (3.2) (see, for instance, the proof of [11,
Theorem 16.15]). For the reader's convenience, we describe the basic idea.
By symmetry, it suffices to prove the first equality in (3.7). We show that the coefficients
of zn in the series RZZ ∗(z) and Ra(z(1 + Mb(z))) are equal to each other for every n ∈ N.
The formal series expansion
Ra(z(1 + Mb(z))) =
∞Xm=1
αm(cid:0)z(1 + Mb(z))(cid:1)m
yields
Cfn(cid:0)Ra(z(1 + Mb(z)))(cid:1) =
nXm=1
αmCfn−m(cid:0)(1 + Mb)m(cid:1), n ∈ N.
12
H. BERCOVICI, A. NICA, M. NOYES, AND K. SZPOJANKOWSKI
Recall that the coefficients of 1 + Mb are moments of b, and expand (1 + Mb)m, to obtain
(3.8)
Cfn[Ra(z(1 + Mb)(z))] =
nXm=1
Xk1,...,km≥0
with k1+···+km=n−m
αmϕ(bk1) · · · ϕ(bkm).
On the other hand, (3.6) yields
Cfn(RZZ ∗) =
nXm=1 XS⊆{1,...,n} with
Xπ={V1,...,Vp}∈N C(n)
S=m and 1∈S
with V1=S
αmβV2 · · · βVp, n ∈ N.
For a fixed set S = {s1, . . . , sm} ⊆ {1, . . . , n} with 1 = s1 < s2 < · · · < sm ≤ n, the
collection of non-crossing partitions {π ∈ N C(n) : S is a block of π} is naturally identified
with the Cartesian product
N C(s2 − s1 − 1) × · · · × N C(sm − sm−1 − 1) × N C(n − sm),
in a way that converts the sum
βV2 · · · βVp
with V1=S
Xπ={V1,...,Vp}∈N C(n)
Xπℓ∈N C(sℓ+1−sℓ−1) YV ∈πℓ
mYℓ=1
βV ,
into the product
(3.9)
that QV ∈πℓ
We conclude that
where we set sm+1 = n + 1. An application of the moment-cumulant formula (3.3) shows
βV = ϕ(bsℓ+1−sℓ−1). (See the proof of [11, Theorem 16.15] for more details.)
(3.10)
Cfn(RZZ ∗) =
nXm=1
XS⊆{1,...,n} with
S=m and 1∈S
αm
mYℓ=1
ϕ(bsℓ+1−sℓ−1), n ∈ N.
Finally, observe that for every fixed m ∈ {1, . . . , n} there is a natural bijection between
tuples (k1, . . . , km) ∈ (N ∪ {0})m with k1 + · · · + km = n − m (on the one hand) and subsets
1 ∈ S ⊆ {1, . . . , n} with S = m (on the other), given by the formula
(k1, . . . , km) 7→ S = {1, k1 + 2, k1 + k2 + 3, . . . , k1 + · · · + km−1 + m}.
The inner sums on the right-hand sides of (3.8) and (3.10) are identified term by term via
this bijection, and this concludes the proof.
(cid:3)
In the remainder of this section we discuss the R-diagonal ∗-distributions that satisfy the
KMS condition. This is a special case of the class of ∗-distribution studied in [12] (see, for
instance [12, Remark 2.10]). The best known example of a KMS R-diagonal distribution
is the one where, in the framework of the next definition, one sets α1 = λ, β1 = 1 and
αn = βn = 0 for n ≥ 2; this is called the λ-circular distribution, and is studied in detail in
[12, Section 4].
ETA-DIAGONALS AND INFINITE DIVISIBILITY FOR R-DIAGONALS
13
Definition 3.6. Let µ ∈ Dalg(1, ∗) be an R-diagonal ∗-distribution with determining se-
quences (αn)∞
n=1, and let t be a positive real number. We say that µ satisfies
the KMS condition with parameter t if
n=1 and (βn)∞
(3.11)
αn = tβn, n ∈ N.
The following result shows that the KMS condition is a generalization of the trace prop-
erty, where the latter property occurs for the value t = 1 of the parameter.
Proposition 3.7. Let µ ∈ Dalg(1, ∗) be an R-diagonal ∗-distribution, let t be a positive real
number, and suppose that µ satisfies the KMS condition with parameter t. Denote by
the unique unital algebra homomorphism such that
Ut : ChZ, Z ∗i → ChZ, Z ∗i
Then
(3.12)
Ut(Z) = tZ and Ut(Z ∗) =
Z ∗.
1
t
µ(P Q) = µ(QUt(P )), P, Q ∈ ChZ, Z ∗i.
Proof. Both sides of (3.12) are bilinear in P and Q, so it suffices to check the equation when
both P and Q are monomials. Using the notation Z ∅ = 1, we must show that
(3.13)
µ(Z vZ w) = µ(Z w Ut(Z v)),
v, w ∈ W + ∪ {∅}.
Equivalently, we must show that the set
S = {v ∈ W + ∪ {∅} : µ(Z vZ w) = µ(Z wUt(Z v)), w ∈ W + ∪ {φ}}
is equal to W + ∪ {∅}. The set S is clearly closed under concatenation and contains ∅.
Therefore, it suffices to show that {1, ∗} ⊂ S. In other words it suffices to prove that
(3.14)
µ(ZZ w) = tµ(Z wZ) and µ(Z ∗Z w) =
µ(Z wZ ∗), w ∈ W + ∪ {∅}.
1
t
We only prove the first equality in (3.14); the verification of the second one is analogous.
The case w = ∅ follows from the fact (incorporated in the definition of an R-diagonal
distribution) that µ(Z) = 0. For the remainder part of the proof we fix a word w =
(ℓ1, . . . , ℓn) ∈ W +, for which we prove that µ(ZZ w) = t µ(Z wZ). Observe that
µ(ZZw) = Cfw1(Mµ), µ(ZwZ) = Cfw2(Mµ),
where w1 := (1, ℓ1, . . . , ℓn) and w2 := (ℓ1, . . . , ℓn, 1). It is convenient to view w1 and w2 as
functions from {1, . . . , n + 1} to {1, ∗} and to record the fact that
(3.15)
w2 = w1 ◦ γn+1,
where γn+1 is the cyclic permutation 1 7→ 2 7→ · · · 7→ n + 1 7→ 1 of {1, . . . , n + 1}.
For every partition π = {V1, . . . , Vk} ∈ N C(n+1) we denote by γ−1
in N C(n+1)) whose blocks are γ−1
tCfw2(Mµ) is obtained from
n+1(V1), . . . , γ−1
n+1(π) the partition (still
n+1(Vk). The desired conclusion Cfw1(Mµ) =
(3.16)
Cfw1;π(Rµ) = tCfw2;γ−1
n+1(π)(Rµ),
π ∈ N C(n + 1),
using the moment-cumulant formula. Indeed, sum both sides of (3.16) over π ∈ N C(n + 1)
and invoke (3.1) applied to the words w1 and w2. The sums thus obtained are precisely
Cfw1(Mµ) on the left side and tCfw2(Mµ) on the right.
14
H. BERCOVICI, A. NICA, M. NOYES, AND K. SZPOJANKOWSKI
Thus, it remains to prove (3.16). Fix a partition π = {V1, . . . , Vk} ∈ N C(n + 1) such that
n+1(Vj), 1 ≤ j ≤ k, and n + 1 ∈ W1.
n+1(π) = {W1, . . . , Wk}, where Wj = γ−1
1 ∈ V1. Then γ−1
It follows from (3.15) that w1 Vj = w2 Wj ∈ W + if 2 ≤ j ≤ k, and thus
(3.17)
Cfw1Vj (Rµ) = Cfw2Wj (Rµ),
2 ≤ j ≤ k
For the remaining block, we show that
(3.18)
Cfw1V1(Rµ) = t · Cfw2W1(Rµ).
Indeed, suppose that V1 = {j1, . . . , jm} with 1 = j1 < j2 < · · · < jm, and therefore
W1 = {j2 − 1, . . . , jm − 1, n + 1}. Both sides of (3.18) are 0 if m is odd or if m is even
but w1V1 is not an alternating word. If m is even and w1V1 is alternating, then we find
that w1V1 = {1, ∗}m/2 and w2V2 = {∗, 1}m/2, which implies that Cfw1V1(Rµ) = αm/2 and
Cfw2W1(Rµ) = βm/2. In this case, (3.18) follows from the KMS hypothesis.
Finally, for the partition π fixed in the preceding paragraph we write:
Cfw1;π(Rµ) = Cfw1V1(Rµ)
Cfw1Vj (Rµ)
= tCfw2W1(Rµ)
Cfw2Wj (Rµ) (by (3.17) and (3.18))
kYj=2
kYj=2
= tCfw2;γ−1
n+1(π)(Rµ),
thus concluding the proof of (3.16).
(cid:3)
Remark 3.8. The converse of Proposition 3.7 is also true. More precisely, every R-diagonal
distribution µ ∈ Dalg(1, ∗) that satisfies (3.12) for some t ∈ (0, +∞) must also satisfy
the KMS condition for the same value of t. To see this, let (αn)∞
n=1 be the
determining sequences of µ. Equation (3.12) yields, in particular, the identity
n=1 and (βn)∞
(3.19)
µ((ZZ ∗)n) = tµ((Z ∗Z)n), n ∈ N.
This identity implies αn = tβn, n ∈ N, by induction on n. For the induction step one
invokes the moment-cumulant formula in order to expand both sides of (3.19) as sums over
N C(2n); then the action of the cyclic permutation γ−1
2n on N C(2n) can be used in the same
way as it was done in the proof of Proposition 3.7.
4. The framework of Dc(1, ∗) and Dc(k), BBP bijections
We now introduce the analytic framework which is of interest for the present paper.
Definition 4.1. (1) Let (A, ϕ) be a C ∗-probability space (which means that A is a unital
C ∗-algebra, ϕ : A → C is a positive linear functional, and ϕ(1) = 1), and let a ∈ A. The
∗-distribution of a is the functional µ ∈ Dalg(1, ∗) determined by the requirement that
µ(Z w) = ϕ(aw), w ∈ W +.
(2) We denote by Dc(1, ∗) the set of all elements of Dalg(1, ∗) that are equal to the
∗-distribution of some element in a C ∗-probability space.
(3) Free additive (respectively, multiplicative) convolution is a binary operation on Dc(1, ∗)
denoted by ⊞ (respectively, ⊠). This operation is uniquely determined by the following
property: given elements a, a′ in a C ∗-probability space (A, ϕ) such that {a, a∗} is free from
{a′, (a′)∗}, the ∗-distribution of a + a′ (respectively, aa′) is the free additive (respectively,
ETA-DIAGONALS AND INFINITE DIVISIBILITY FOR R-DIAGONALS
15
multiplicative) convolution of the ∗-distributions of a and a′. See [11, Lectures 5 and 7] for
more details.
(4) An element µ ∈ Dc(1, ∗) is said to be ⊞-infinitely divisible if for every n ∈ N there
exists µn ∈ Dc(1, ∗) such that
The set of all ⊞-infinitely divisible distributions in Dc(1, ∗) is denoted by D(inf−div)
c
(1, ∗).
µ = µn ⊞ · · · ⊞ µn
.
n times
{z
}
The main result of this section is the following theorem. The series Rν and ηµ appearing
in the statement of the theorem are as defined in Sections 2 and 3.
Theorem 4.2. (BBP bijection on Dc(1, ∗).) There exists a bijection
B
(1,∗) : Dc(1, ∗) → D(inf−div)
c
(1, ∗),
determined by the requirement that
(4.1)
RB(1,∗)(µ) = ηµ, µ ∈ Dc(1, ∗).
More precisely, for every µ ∈ Dc(1, ∗) there exists a unique ∗-distribution ν ∈ D(inf−div)
such that Rν = ηµ, and we define B
(1,∗)(µ) := ν.
c
(1, ∗)
Definition and Remark 4.3. (Framework of Dc(k).) We reduce Theorem 4.2 to an
analogous theorem proved in [2] for the space, denoted by Dc(2), of joint distributions of
pairs of selfadjoint elements in a C ∗-probability space. The passage from Dc(1, ∗) to Dc(2)
is natural, and essentially amounts to the change of variables
(a, a∗) 7→(cid:18) a + a∗
2
,
a − a∗
2i (cid:19) ,
for a in a C ∗-probability space (A, ϕ). In order to clarify this idea, we review briefly the
framework of Dc(k). Fix k ∈ N.
(1) We denote by ChX1, . . . , Xki the algebra of polynomials in the non-commuting inde-
terminates X1, . . . , Xk.
(2) Let (A, ϕ) be a C ∗-probability space and let b1, . . . , bk ∈ A be selfadjoint. The joint
distribution of b1, . . . , bk is the linear functional λ : ChX1, . . . , Xki → C which is determined
by the requirement that λ(1) = 1 and
λ(Xi1 · · · Xin) = ϕ(bi1 · · · bin), n ∈ N, i1, . . . , in ∈ {1, . . . , k}.
(3) We denote by Dc(k) the set of all linear functionals λ : ChX1, . . . , Xki → C that can
arise as joint distributions of k-tuples of selfadjoint elements in some C ∗-probability space.
(4) Free additive convolution is a binary operation on Dc(k) denoted 2 by ⊞. This
operation is uniquely determined by the following property: given selfadjoint elements
a1, . . . , ak and b1, . . . , bk in a C ∗-probability space (A, ϕ) such that {a1, . . . , ak} is free from
{b1, . . . , bk}, the joint distribution of a1 + b1, . . . , ak + bk is the free additive convolution of
the joint distributions of a1, . . . , ak and b1, . . . , bk. The concept of ⊞-infinite divisibility in
Dc(k) is introduced as in Definition 4.1(4). The set of ⊞-infinitely divisible distributions in
Dc(k) is denoted by D(inf−div)
(k).
c
(5) We denote by C0hhx1, . . . , xkii the space of those formal power series with complex
coefficients in k non-commuting indeterminates x1, . . . , xk whose constant term is equal to
2It is customary to always denote free additive convolution by "⊞". The setting in which the symbol ⊞
is used should be clear, in each case, from the context.
16
H. BERCOVICI, A. NICA, M. NOYES, AND K. SZPOJANKOWSKI
0. We denote by Cf(i1,...,in)(f ) the coefficient of xi1 · · · xin in a series f ∈ C0hhx1, . . . , xkii.
Every joint distribution λ ∈ Dc(k) has a moment series Mλ, an R-transform Rλ and an
η-series ηλ. These are elements of C0hhx1, . . . , xkii, and their definitions are analogous to
Definitions 2.3 and 3.2. A detailed description of these power series and of the relations
between their coefficients can be found in [2, pp. 14-17].
The proof of Theorem 4.2 will be reduced to the following result from [2] (see also [4] for
the case k = 1).
Theorem 4.4. (BBP bijection on Dc(k).) Let k be a positive integer. There exists a
bijection Bk : Dc(k) → D(inf−div)
(k), determined by the requirement that
c
(4.2)
RBk(λ) = ηλ,
λ ∈ Dc(k).
More precisely, for every λ ∈ Dc(k) there exists a unique ν ∈ D(inf−div)
ηλ, and we define Bk(λ) := ν.
c
(k) such that Rν =
(cid:3)
Remark 4.5. We only need Theorem 4.4 for k = 1 and k = 2. When k = 1, the space Dc(1)
is naturally identified with the space Pc of compactly supported Borel probability measures
on R. Indeed, given b = b∗ in a C ∗-probability space (A, ϕ), Definition 4.3(2) produces
a linear functional λ : Chx1i → C which becomes, via the Riesz representation theorem,
a Borel probability measure supported on the spectrum of b. The original BBP bijection
from [4] was defined on Pc (and on the larger set P of all Borel probability measures on R).
In Section 6, we will simply talk about B(σ) for σ ∈ Pc. In other words if λ denotes the
functional in Dc(1) corresponding to σ, then B(σ) ∈ Pc denotes the probability measure
corresponding to B1(λ).
The following result creates bijections C and D that we use in conjunction with the case
k = 2 of Theorem 4.4. (The letters C and D are meant to suggest complexification and
decomplexification.) The proof is immediate, and therefore omitted.
Proposition 4.6. There exists a bijection D : Dc(1, ∗) → Dc(2) defined as follows. Given
µ ∈ Dc(1, ∗) that is the ∗-distribution of an element a in a C ∗-probability space (A, ϕ), D(µ)
is the joint distribution of the pair
(cid:18) a + a∗
2
,
a − a∗
2i (cid:19) .
The inverse of D is the bijection C : Dc(2) → Dc(1, ∗) defined as follows. Given λ ∈ Dc(2)
that is the joint distribution of a pair b1, b2 of selfadjoint elements in a C ∗-probability space
(A, ϕ), C(λ) is the ∗-distribution of b1 + ib2.
(cid:3)
Definition and Remark 4.7. In addition to the transformations C and D, the proof of
Theorem 4.2 requires the corresponding change of variables for power series. Denote by
(4.3)
t1,1 = t1,∗ =
1
2
and t2,1 =
1
2i
, t2,∗ = −
1
2i
,
the coefficients of the linear transformation b1 = (a + a∗)/2, b2 = (a − a∗)/2i. This trans-
formation can now be written more compactly as
(4.4)
bi = Xℓ∈{1,∗}
ti,ℓaℓ, for i = 1, 2.
ETA-DIAGONALS AND INFINITE DIVISIBILITY FOR R-DIAGONALS
17
Using the coefficients ti,ℓ we define a map eD : C0hhz, z∗ii → C0hhx1, x2ii as follows: given
a series f ∈ C0hhz, z∗ii, the coefficients of the series g = eD(f ) ∈ C0hhx1, x2ii are given by
(4.5) Cf(i1,...,in)(g) := Xℓ1,...,ℓn∈{1,∗}
The map eD is clearly linear and bijective.
Its inverse eC : C0hhx1, x2ii → C0hhz, z∗ii is
defined by a formula analogous to (4.5), but with [ti,ℓ] replaced by the inverse matrix
ti1,ℓ1 · · · tin,ℓnCf(ℓ1,...,ℓn)(f ), n ∈ N, i1, . . . , in ∈ {1, 2}.
t′
1,1 = 1, t′
1,2 = i and t′
∗,1 = 1, t′
∗,2 = −i.
Lemma 4.8. For every ∗-distribution µ ∈ Dc(1, ∗), we have
(4.6)
MD(µ) = eD(Mµ), RD(µ) = eD(Rµ), and ηD(µ) = eD(ηµ).
Proof. Suppose that µ ∈ Dc(1, ∗) is the ∗-distribution of an element a in a C ∗-probability
space (A, ϕ), and set bi =Pℓ∈{1,∗} ti,ℓaℓ for i = 1, 2. By definition, the joint distribution of
b1, b2 is D(µ) ∈ Dc(2). To verify the first identity in (4.6), fix n ∈ N and i1, . . . , in ∈ {1, 2},
and calculate directly
Cf(i1,...,in)(MD(µ)) = ϕ(bi1 · · · bin)
ti1,ℓ1 · · · tin,ℓn ϕ(aℓ1 · · · aℓn)
= ϕ(cid:16)(ti1,1a1 + ti1,∗a∗) · · · (tin,1a1 + tin,∗a∗)(cid:17)
= Xℓ1,...,ℓn∈{1,∗}
= Xℓ1,...,ℓn∈{1,∗}
= Cf(i1,...,in)eD(Mµ).
ti1,ℓ1 · · · tin,ℓnCf(ℓ1,...,ℓn)(Mµ)
The second equality in (4.6) follows from a similar multilinearity argument, using the fact
that the C ∗-probability space (A, ϕ) carries a family of multilinear functionals (κn : An →
C)∞
n=1, called free cumulant functionals, such that
Cf(i1,...,in)(RD(µ)) = κn(bi1 , . . . , bin), n ∈ N, i1, . . . , in ∈ {1, 2}, and
Cf(ℓ1,...,ℓn)(Rµ) = κn(aℓ1, . . . , aℓn), n ∈ N, ℓ1, . . . , ℓn ∈ {1, ∗}.
(This multilinearity argument is precisely the one used to describe the behavior of the
R-transform under linear transformations [11, Proposition 16.12].)
The third equality (4.6) follows from a similar multilinearity argument, using the Boolean
n=1 (for a discussion of Boolean cumulants see, for
(cid:3)
cumulant functionals (βn : An → C)∞
instance, [8, Section 4.6]).
Lemma 4.9. Let D : Dc(1, ∗) → Dc(2) be the bijection defined in Proposition 4.6. Then:
(1) D(µ ⊞ µ′) = D(µ) ⊞ D(µ′) for every µ, µ′ ∈ Dc(1, ∗).
(2) D(D(inf−div)
(1, ∗)) = D(inf−div)
(2).
c
c
18
H. BERCOVICI, A. NICA, M. NOYES, AND K. SZPOJANKOWSKI
Proof. (1) Since every λ ∈ Dc(2) is uniquely determined by its R-transform, it suffices to
verify that D(µ ⊞ µ′) and D(µ) ⊞ D(µ′) have the same R-transform. Indeed,
RD(µ⊞µ′) = eD(Rµ⊞µ′ ) (by Lemma 4.8)
= eD(Rµ) + eD(Rµ′) (since Rµ⊞µ′ = Rµ + Rµ′ and eD is linear)
= RD(µ) + RD(µ′) (by Lemma 4.8)
= RD(µ)⊞D(µ′).
Part (2) follows immediately from (1) and from the definition of ⊞-infinite divisibility. (cid:3)
Proof of Theorem 4.2. We define the required bijection B
(1,∗) so that the diagram
Dc(1, ∗)
D↓
c
B(1,∗)−→ D(inf−div)
↓D
(1, ∗)
Dc(2)
B2−→ D(inf−div)
c
(2)
is commutative, where D is defined in Proposition 4.6 and B2 is provided by Theorem 4.4
for k = 2. More precisely, let D0 : D(inf−div)
(2) be the restriction of D;
this is a bijection by Lemma 4.9(2)). Then define
(1, ∗) → D(inf−div)
c
c
B(1,∗) := D−1
0 ◦ B2 ◦ D.
Pick an arbitrary µ ∈ Dc(1, ∗), denote B
injective, it suffices to verify that eD(Rν) = eD(ηµ). Indeed, the definition of B
B2(D(µ)) = D(ν), and the definition of B2 yields RD(ν) = ηD(µ). Thus
(1,∗)(µ) = ν. We prove that Rν = ηµ. Since eD is
(1,∗) implies
eD(Rν ) = RD(ν) = ηD(µ) = eD(ηµ),
as required.
(cid:3)
5. Parametrization of η-diagonal distributions in Dc(1, ∗)
We show that an η-diagonal ∗-distribution µ ∈ Dc(1, ∗) is naturally parametrized by the
pair of compactly supported probability measures on [0, ∞) that arise as the distributions
of ZZ ∗ and of Z ∗Z in the noncommutative probability space (ChZ, Z ∗i, µ).
Definition and Remark 5.1. Suppose that µ ∈ Dc(1, ∗) is the ∗-distribution of an element
a in a C ∗-probability space (A, ϕ). Basic considerations on positive elements in a C ∗-
probability space (see, for instance, [11, Propositions 3.13 and 3.6]) show the existence of
compactly supported Borel probability measures σ1, σ2 on [0, ∞) such that
0
tn dσ1(t) and ϕ((a∗a)n) =Z ∞
ϕ((aa∗)n) =Z ∞
Z ∞
Z ∞
0
0
0
tn dσ2(t) = µ((Z ∗Z)n) n ∈ N.
tn dσ1(t) = µ((ZZ ∗)n), n ∈ N
Thus σ1 and σ2 satisfy
(5.1)
and
(5.2)
tn dσ2(t), n ∈ N.
ETA-DIAGONALS AND INFINITE DIVISIBILITY FOR R-DIAGONALS
19
Moreover, σ1 and σ2 are uniquely determined by (5.1) and (5.2) since a compactly supported
probability measure on R is determined by its moments. We refer to σ1 and σ2 as the
distributions of ZZ ∗ and Z ∗Z, respectively, in the ∗-probability space (ChZ, Z ∗i, µ).
The following theorem provides the parametrization announced in the title of the section.
c denote the set of all compactly supported Borel probability measures
Theorem 5.2. Let P +
on [0, ∞). There is bijective map
c × P +
Φ : P +
c → {µ ∈ Dc(1, ∗) : µ is η-diagonal}
described as follows: given σ1, σ2 ∈ P +
c , Φ(σ1, σ2) is the unique η-diagonal ∗-distribution
µ ∈ Dc(1, ∗) such that the distributions of ZZ ∗ and Z ∗Z in (ChZ, Z ∗i, µ) are equal to σ1
and σ2, respectively.
c × P +
The point of Theorem 5.2 is that the map Φ is defined on all of P +
c . In other words,
for every σ1, σ2 ∈ P +
c there exists an η-diagonal ∗-distribution µ ∈ Dc(1, ∗) such that (5.1)
and (5.2) hold. We prove this by producing an operator model for µ: starting from σ1 and σ2
we construct explicitly an operator A on a Hilbert space K such that the ∗-distribution of A
with respect to a suitably chosen functional on B(K) is the required η-diagonal distribution.
The bulk of this section is devoted to the description of the operator model. At the end, we
complete the proof of Theorem 5.2. The construction of the operator model is described in
the next remark.
Remark and Notation 5.3. (Description of the operator model.) Fix σ1, σ2 ∈ P +
c which
we take as the input for our construction of an η-diagonal operator. In the description of
the construction, it is convenient to use the symmetric square roots of σ1 and σ2. These
moments given by the formula
are the symmetric compactly supported Borel probability measures eσ1 and eσ2 on R with
Z ∞
−∞
tn deσj(t) =(0,
R ∞
0
n odd
tn/2 dσj(t), n even,
for j = 1, 2. Our construction of an η-diagonal operator proceeds in three steps.
Step 1. We construct a Hilbert space H, an operator X ∈ B(H), and vectors ξ1, ξ2 ∈ H
with the following properties:
(1a) ξ1 = ξ2 = 1,
(1b) hξ1, ξ2i = 0,
(1c) hX kξ1, ξ2i = hX kξ2, ξ1i = 0 for k ∈ N,
0
tk dσj(t) for k ∈ N and j = 1, 2.
algebra B(H). For the actual construction of X consider, for j = 1, 2, Hilbert spaces Mj,
operators Tj ∈ B(Mj), and unit vectors ηj ∈ Mj, such that the distribution of Tj with
(1d) hX 2k−1ξj, ξji = 0 and hX 2kξj, ξji =R ∞
In other words, property (1d) says that X has distribution eσ1 with respect to the vector
state defined by ξ1, and distributioneσ2 with respect to the one defined by ξ2, on the operator
respect to the vector state ηj is eσj. Then set H = M1 ⊕ M2, X = T1 ⊕ T2, ξ1 = η1 ⊕ 0,
and ξ2 = 0 ⊕ η2. Properties (1a) -- (1d) are then easily verified. (In subsequent steps we only
use the properties (1a) -- (1d). The precise description of H, X, ξ1, and ξ2 is not necessary.)
Step 2. Define a rank-one partial isometry Y ∈ B(H) by setting Y (ζ) = hζ, ξ1iξ2 for ζ ∈ H.
We have Y ξ1 = ξ2, Y ∗ζ = hζ, ξ2iξ1 for ζ ∈ H, and
(5.3)
Thus Y Y ∗ and Y ∗Y are the orthogonal projections onto the 1-dimensional spaces generated
by ξ2 and ξ1, respectively.
Y Y ∗ζ = hζ, ξ2iξ2, Y ∗Y ζ = hζ, ξ1iξ1.
20
H. BERCOVICI, A. NICA, M. NOYES, AND K. SZPOJANKOWSKI
Step 3. Consider the Hilbert space K = H ⊗ H and the unit vector ξ = ξ1 ⊗ ξ2 ∈ K, then
consider the C ∗-probability space (B(K), ϕξ), where ϕξ(T ) = hT ξ, ξi for T ∈ B(K). Let
V ∈ B(H⊗H) be the flip operator determined by the requirement that by V (ζ ⊗ζ ′) = ζ ′ ⊗ζ,
ζ, ζ ′ ∈ H. Note that V is a symmetry (that is, it is self-adjoint and V 2 = I). Finally, define
A = V (Y ⊗ X).
This concludes the construction of the variable A in (B(K), ϕξ).
We now take on the proof that the operator A constructed above has the desired η-
diagonal distribution with respect to the functional ϕξ. We start by recording some easily
verified identities satisfied by A, the proof of which is left to the reader.
Lemma 5.4. Consider the framework of Remark 5.3. We have
AA∗ = X 2 ⊗ Y Y ∗, A∗A = Y ∗Y ⊗ X 2,
(5.4)
and
(5.5)
A2 = XY ⊗ Y X,
(A∗)2 = Y ∗X ⊗ XY ∗.
The following lemma establishes the distributions of AA∗ and A∗A along with a few
non-alternating ∗-moments of A.
(cid:3)
Lemma 5.5. Let A be as above, then for any integer k ≥ 0 we have
(2) ϕξ(A(AA∗)k) = 0,
(3) ϕξ(A∗(AA∗)k) = 0,
(1) ϕξ((AA∗)k) =R ∞
(4) ϕξ((A∗A)k) =R ∞
(5) ϕξ(A(A∗A)k) = 0,
(6) ϕξ(A∗(A∗A)k) = 0.
0
0
tk dσ1(t),
tk dσ2(t),
Proof. We verify only the first three equations. The proof of (4) -- (6) is similar. We have
so
(AA∗)k = X 2k ⊗ (Y Y ∗)k = X 2k ⊗ (Y Y ∗),
ϕξ((AA∗)k) = h(X 2k ⊗ (Y Y ∗))ξ, ξi = hX 2kξ1, ξ1i,
and (1) follows from property (1d) in Step 1 of the construction of A. To prove (2), we
calculate
A(AA∗)k = V (Y ⊗ X)(X 2k ⊗ Y Y ∗),
thus
ϕξ(A(AA∗)k) = h(Y ⊗ X)(X 2k ⊗ Y Y ∗)ξ1 ⊗ ξ2, V ξ1 ⊗ ξ2i
= h(Y ⊗ X)(X 2kξ1 ⊗ ξ2), ξ2 ⊗ ξ1i
= hY X 2kξ1, ξ2ihXξ2, ξ1i = 0,
because hXξ2, ξ1i = 0, thereby concluding the proof of (2). Similarly,
A∗(AA∗)k = (X ⊗ Y ∗)V (X 2k ⊗ Y Y ∗),
so
ϕξ(A∗(AA∗)k) =h((X ⊗ Y ∗)V (X 2k ⊗ Y Y ∗)ξ, ξi = hX ⊗ Y ∗(ξ2 ⊗ X 2kξ1), ξ1 ⊗ ξ2i
=hXξ2, ξ1ihY ∗X 2kξ1, ξ2i = 0
by property (1c) in Step 1 of the construction of A.
(cid:3)
ETA-DIAGONALS AND INFINITE DIVISIBILITY FOR R-DIAGONALS
21
The next lemma gives some properties of the operator A that are useful in verifying its
η -- diagonality.
Lemma 5.6. Let A be as above, then for any integer k ≥ 0 we have
(1) A2(AA∗)kξ = 0,
(2) (A∗)2(AA∗)kξ = 0,
(4) A2(A∗A)kξ = 0,
(5) (A∗)2(A∗A)kξ=0,
(3) (A∗A)(AA∗)kξ =R ∞
(6) (AA∗)(A∗A)kξ =R ∞
0
0
tk dσ1(t) · A∗Aξ,
tk dσ2(t) · AA∗ξ.
Proof. As in the previous proof we only verify (1) -- (3). We have
(A)2(AA∗)k = (XY ⊗ Y X)(X 2k ⊗ Y Y ∗) = XY X 2k ⊗ Y XY Y ∗,
and using the fact that Y Y ∗ξ2 = ξ2 we see that Y XY Y ∗ξ2 = hXξ2, ξ1iξ2 = 0 by property
(1c). Similarly,
(A∗)2(AA∗)k = (Y ∗X ⊗ XY ∗)(X 2k ⊗ Y Y ∗) = Y ∗X 2k+1 ⊗ XY ∗,
and (2) follows because Y ∗X 2k+1ξ1 = hX 2k+1ξ1, ξ2iξ1 = 0 by (1c). Finally, (5.4) yields
(A∗A)(AA∗)k = (Y ∗Y ⊗ X 2)(X 2k ⊗ Y Y ∗) = Y ∗Y X 2k ⊗ X 2Y Y ∗.
Observe that Y ∗Y X 2kξ1 = hX 2kξ1, ξ1iξ1 = R ∞
Therefore
0
(A∗A)(AA∗)k =Z ∞
0
tk dσ1(t) · (ξ1 ⊗ X 2ξ2) =Z ∞
0
tk dσ1(t) · ξ1 by (1d), while Y Y ∗ξ2 = ξ2.
tk dσ1(t) · A∗Aξ,
thus proving (3).
(cid:3)
Corollary 5.7. Let W = W1W2 · · · Wd be a mixed-alternating word in A and A∗, factored
as in (2.7) with d ≥ 2. Then
W ξ = ϕξ(W2)ϕξ(W3) · · · ϕξ(Wd)W1ξ.
Proof. Parts (3) and (6) of the preceding lemma yield the conclusion when d = 2. The
general case follows easily by induction on d.
(cid:3)
Proposition 5.8. Let σ1 and σ2 be probability measures in P +
c , and let the operator A in
(B(K), ϕξ) be constructed as in Remark 5.3. Then the ∗-distribution of A is η-diagonal.
Moreover, the distributions of AA∗ and A∗A are σ1 and σ2, respectively.
Proof. The second assertion follows from parts (1) and (4) of Lemma 5.5. It remains to
prove that the distribution of A is η-diagonal, and to do this we verify the conditions in
Theorem 2.8. Let W = W1W2 · · · Wd be a mixed-alternating word in A and A∗, factored as
in (2.7). Corollary 5.7 yields
ϕξ(W ) = ϕξ(W2)ϕξ(W3) · · · ϕξ(Wd)hW1ξ, ξi =
ϕξ(Wj),
dYj=1
thus verifying condition (ηDM2). Finally we verify condition (ηDM1). Suppose that V is a
word in A and A∗ that is not mixed-alternating, and choose a mixed-alternating word W of
22
H. BERCOVICI, A. NICA, M. NOYES, AND K. SZPOJANKOWSKI
maximum length with the property that V can be written as V = U W for some non-empty
word U . Also, write W = W1W2 · · · Wd as in (2.7). We have
ϕξ(V ) = hV ξ, ξi =
ϕξ(Wj)hU W1ξ, ξi
dYj=2
by Corollary 5.7. If U = 1, the equality ϕξ(V ) = 0 follows from Lemma 5.5. If U ≥ 2,
then U is of the form U ′AA or U ′A∗A∗ for some (possibly empty) word U ′. In this case,
ϕξ(V ) = 0 by Lemma 5.6.
(cid:3)
We conclude the discussion of the parametrization announced at the beginning of the
section.
Proof of Theorem 5.2. We first note that the map Φ is well-defined. Indeed, let σ1, σ2 ∈
P +
c be given. The existence of an η-diagonal ∗-distribution µ ∈ Dc(1, ∗) which fulfils the
conditions (5.2) is ensured by Proposition 5.8. The uniqueness of µ follows from the fact
that an η-diagonal ∗-distribution is completely determined by its alternating ∗-moments,
as we saw in Theorem 2.8.
The surjectivity of Φ is immediate from its definition: every η-diagonal ∗-distribution
c are the distributions of ZZ ∗
µ ∈ Dc(1, ∗) can be written as Φ(σ1, σ2), where σ1, σ2 ∈ P +
and respectively Z ∗Z in (ChZ, Z ∗i, µ), in the sense discussed in Definition 5.1.
Finally, the injectivity of Φ is immediate as well. Indeed, if Φ(σ1, σ2) = µ, then the mo-
ments of σ1 and σ2 can be retrieved as alternating moments of µ, and compactly supported
probability measures on R are determined by their moments.
(cid:3)
6. Parametrization of infinitely divisble R-diagonal distributions
In this section we use the BBP method to characterize ⊞-infinitely divisible R-diagonal
distributions. The parametrization mentioned in the title of the section arises naturally, in
the way indicated in the following remark.
Remark and Notation 6.1. Let R(inf−div)
denote the set of all the R-diagonal distribu-
tions in Dc(1, ∗) that are ⊞-infinitely divisible. It is immediate that the bijection B(1,∗) from
Theorem 4.2 induces a bijection (still denoted B
c
(1,∗))
(6.1)
B
(1,∗) : {µ ∈ Dc(1, ∗) : µ is η-diagonal} → R(inf−div)
c
.
On the other hand, Theorem 5.2 provides a natural bijection
(6.2)
The map
(6.3)
Φ : P +
c × P +
c → {µ ∈ Dc(1, ∗) : µ is η-diagonal}.
Ψ := B(1,∗) ◦ Φ : P +
c × P +
c → R(inf−div)
c
is therefore a bijection as well. We refer to Ψ as the BBP parametrization of R(inf−div)
c yields a distribution ν = Ψ(σ1, σ2) ∈ R(inf−div)
Every choice of parameters σ1, σ2 ∈ P +
and every ν ∈ R(inf−div)
arises from a unique pair σ1, σ2.
c
c
c
.
,
We emphasize that the bijection Ψ works in a really straightforward way -- the coefficients
of the η-series of σ1 and σ2 give the determining sequences of ν = Ψ(σ1, σ2). It is actually
worth recording a direct consequence of this fact, as follows.
Notation 6.2. We denote by E +
that f = ησ for some σ ∈ P +
c the collection of those series f ∈ C[[z]] with the property
c (where σ is, a fortiori, uniquely determined).
ETA-DIAGONALS AND INFINITE DIVISIBILITY FOR R-DIAGONALS
23
Proposition 6.3. Let ν ∈ Dc(1, ∗) be an R-diagonal distribution, and let (αn)∞
(βn)∞
ν is ⊞-infinitely divisible if and only if both f and g belong to E +
c .
n=1 be its determining sequences. Set f (z) =P∞
n=1 αnzn and g(z) =P∞
n=1 and
n=1 βnzn. Then
Proof. If ν ∈ R(inf−div)
g = ησ2, and so f, g ∈ E +
for some σ1, σ2 ∈ P +
c
, then ν = Ψ(σ1, σ2) for some σ1, σ2 ∈ P +
c . Conversely, suppose that f, g ∈ E +
c , hence f = ησ1 and
c , so f = ησ1 and g = ησ2
, and the
c
hence ν ∈ R(inf−div)
definition of Ψ shows that eν has the same determining sequences as ν. This forces ν =eν,
c . Then the distribution eν := Ψ(σ1, σ2) belongs to R(inf−div)
The criterion provided by Proposition 6.3 is useful because one can (following the work
in terms of the associated analytic functions. We will
in [1]) characterize the series from E +
c
follow up on this in the application presented in Section 7.
(cid:3)
.
c
Since we are dealing with free probabilistic structures, it is natural to ask what is the
description of the BBP parametrization Ψ in terms of R-transforms. Recall (Remark 3.4)
that an R-diagonal ∗-distribution ν ∈ Dc(1, ∗) is uniquely determined by the R-transforms
RZZ ∗, RZ ∗Z ∈ C[[z]]. The following result thus provides an alternative characterization of
what is Ψ(σ1, σ2).
Theorem 6.4. Let σ1, σ2 ∈ P +
of Z ∗Z in the ∗-probability space (ChZ, Z ∗i, ν) are described as follows:
c and set ν = Ψ(σ1, σ2). Then the R-transforms of ZZ ∗ and
(6.4)
RZZ ∗(z) = RB(σ1)(cid:0)z(1 + MB(σ2)(z))(cid:1) , RZ ∗Z (z) = RB(σ2)(cid:0)z(1 + MB(σ1)(z))(cid:1) ,
where B(σ1) and B(σ2) indicate the original BBP bijection (as discussed in Remark 4.5).
Proof. We set µ := Φ(σ1, σ2), so ν is R-diagonal, µ is η-diagonal, and B
(1,∗)(µ) = ν. Thus
Rν(z, z∗) = ηµ(z, z∗) =
αn(zz∗)n +
βn(z∗z)n,
∞Xn=1
∞Xn=1
n=1 and (βn)∞
where (αn)∞
n=1 are the (common) determining sequences for µ and for ν. By
the definition of the bijection Φ in Theorem 5.2, σ1 has the same moments as the element
ZZ ∗ in the noncommutative probability space (ChZ, Z ∗i, µ). This implies that ησ1 = ηZZ ∗ ,
n=1 αnzn. In a similar way we
and then Proposition 2.13 gives us the formula ησ1(z) =P∞
find that ησ2(z) =P∞
Consider now the probability measures B(σ1), B(σ2) ∈ Pc. The definition of B implies
n=1 βnzn.
RB(σ1)(z) = ησ1 (z) =
αnzn, RB(σ2)(z) = ησ2(z) =
βnzn.
∞Xn=1
∞Xn=1
But then Proposition 3.5 applies to the R-diagonal ∗-distribution ν and yields (6.4).
(cid:3)
As a consequence of Theorem 6.4, we obtain a natural connection between the notions
of ⊞-infinite divisibility in Dc(1, ∗) and in Pc. This is stated in the next corollary. The
converse of the corollary fails even in the tracial framework (see Remark 6.9 below).
Corollary 6.5. Let ν ∈ Dc(1, ∗) be R-diagonal and let τ1, τ2 ∈ P +
c be the distributions of
ZZ ∗ and Z ∗Z in the ∗-probability space (ChZ, Z ∗i, ν) (as discussed in Definition 5.1). If ν
is ⊞-infinitely divisible in Dc(1, ∗), then τ1 and τ2 are ⊞-infinitely divisible in Pc.
Proof. By symmetry, it suffices to show that τ1 is ⊞-infinitely divisible. According to [14,
Theorem 4.3], a compactly supported Borel probability measure on R is ⊞-infinitely divisible
24
H. BERCOVICI, A. NICA, M. NOYES, AND K. SZPOJANKOWSKI
if and only if its R-transform can be extended to an analytic self-map of the upper half-
plane C+. Suppose that ν = Φ(σ1, σ2), where σ1, σ2 ∈ P +
c . The R-transform Rτ1, which
is the same as RZZ ∗, is given by the first Equation (6.4). By [5, Proposition 6.1], the
moment series of the probability measure B(σ2) can be extended analytically to C+ and
this extension satisfies
z ∈ C+ ⇒ z(1 + MB(σ2)(z)) ∈ C+.
Finally, since B(σ1) is ⊞-infinitely divisible, [14, Theorem 4.3] assures us that RB(σ1) extends
analytically to a self-map of C+. We conclude that for every z ∈ C+, RB(σ1) is defined at
z(1 + MB(σ2)(z)), and that
z 7→ RB(σ1)(cid:0) z(1 + MB(σ2))(z)(cid:1)
is an analytic self-map on C+, as required.
(cid:3)
In the remainder of this section, we discuss the KMS example. In this special case one
can process further the formulas from Theorem 6.4 and arrive at explicit formulas (stated in
Proposition 6.8) for the distributions of ZZ ∗ and of Z ∗Z in terms of the probability measures
σ1, σ2 that parametrize ν. These formulas call on some commonly used operations from the
free harmonic analysis of P +
c , that are reviewed in the following remark.
Remark 6.6. (Some elements of free harmonic analysis on P +
c .) (1) Measures σ ∈ Pc have
free additive convolution powers with real exponent t ∈ [1, ∞). More precisely, for every
σ ∈ Pc and t ∈ [1, ∞), there exists a unique measure τ ∈ Pc such that Rτ = tRσ (see [11,
pp. 228-231]). This measure τ is denoted σ⊞t. When t is an integer, σ⊞t is simply the t-fold
convolution σ ⊞ · · · ⊞ σ. The argument in [11, pp. 228-231] also shows that σ⊞t ∈ P +
for
c
all t ∈ [1, ∞) if σ ∈ P +
c .
The analogous result for Boolean convolution provides for every σ ∈ Pc and t ∈ (0, ∞)
a Boolean convolution power σ⊎t ∈ Pc such that ησ⊎t = tησ (see [13, Theorem 3.6]). As in
the free case, σ⊎t ∈ P +
c (see, for instance, the operator model
c
constructed in [2, Proposition 4.8]).
for every t ∈ (0, ∞) if σ ∈ P +
(2) The original BBP bijection B : Pc → P (inf−div)
c
(Remark 4.5) can be expressed using
convolution powers, by the formula
B(σ) =(cid:0)σ⊞2(cid:1)⊎1/2
,
σ ∈ Pc,
which was proved in [3, Theorem 1.2]). The facts reviewed in (1) above imply that B(σ) ∈
P +
c
for every σ ∈ P +
c .
(3) Free multiplicative convolution ⊠ is another binary operation defined on the set P +
c .
This operation corresponds to the product of free random variables. Quite remarkably,
BP +
c was shown in [3, Remark 3.9] to be a homomorphism for ⊠, that is,
B(σ ⊠ σ′) = B(σ) ⊠ B(σ′),
σ, σ′ ∈ P +
c .
(4) The free counterpart of the standard Poisson distribution is the Marchenko-Pastur
distribution Π1 (also known as the the free Poisson distribution). This distribution is
supported on the interval [0, 4] and it is Lebesgue absolutely continuous with density
Its R-transform is
dΠ1(t)/dt =
1
2πp(4 − t)/t,
0 ≤ t ≤ 4.
RΠ1(z) = z/(1 − z),
ETA-DIAGONALS AND INFINITE DIVISIBILITY FOR R-DIAGONALS
25
and a simple calculation using the definition of B shows that
(6.5)
Π1 = B(cid:18) 1
2
(δ0 + δ2)(cid:19) .
A useful property of Π1 is that it converts moment series into R-transforms via the formula
(6.6)
Rσ⊠Π1 = Mσ,
σ ∈ P +
c .
See, for instance, [11, Propositions 17.2 and 17.4].
Remark 6.7. Let σ ∈ P +
(αn)∞
R(inf−div)
n=1 and (βn)∞
satisfy
c
c and let t > 0 be a real number. The determining sequences
n=1 of the infinitely divisible R-diagonal ∗-distribution ν := Ψ(σ⊎t, σ) ∈
βnzn = ησ(z),
αnzn = ησ⊎t (z) = tησ(z).
∞Xn=1
∞Xn=1
Thus ν satisfies the KMS condition with parameter t: αn = tβn, n ∈ N (Definition 3.6).
Proposition 6.8. With the notation of the preceding remark, let τ1, τ2 ∈ P +
c be the dis-
tributions of ZZ ∗ and of Z ∗Z in the noncommutative probability space (ChZ, Z ∗i, ν) (as
discussed in Definition 5.1). Then
(6.7)
τ1 = (B(σ) ⊠ Π1)⊞t and τ2 =(cid:0)B(σ)⊞t ⊠ Π1(cid:1)⊞1/t
.
Proof. The two formulas in (6.7) have similar proofs. We only verify the first one. Since τ1
is the distribution of ZZ ∗, we have Rτ1 = RZZ ∗, and Theorem 6.4 yields
(6.8)
where σ1 = σ⊎t and σ2 = σ. The relation σ1 = σ⊎t and (1.2) imply B(σ1) = B(σ)⊞t, and
hence RB(σ1) = t RB(σ). The equality (6.8) can be continued as follows:
Rτ1(z) = RB(σ1)(cid:0)z(1 + MB(σ2)(z))(cid:1) ,
Rτ1(z) = t · RB(σ)(cid:0)z(1 + MB(σ)(z))(cid:1)
= t · MB(σ)(z) (by (3.4))
= t · RB(σ)⊠Π1 (z) (by (6.6))
= R(B(σ)⊠Π1)⊞t (z).
Thus the probability measures τ1 and (B(σ) ⊠ Π1)⊞t are equal because they have the same
R-transform.
(cid:3)
Remark 6.9. (Tracial case.) In the special case when t = 1, the preceding proposition
reduces to
(6.9)
τ1 = τ2 = B(σ) ⊠ Π1.
Using (6.5) and invoking the multiplicativity of B (Remark 6.6(3)), we can rewrite (6.9) as
(6.10)
τ1 = τ2 = B(cid:18)σ ⊠
1
2
(δ0 + δ2)(cid:19) .
This confirms the fact (Corollary 6.5) that τ1 and τ2 are ⊞-infinitely divisible in Pc.
c
that cannot be written as σ ⊠ 1
We conclude with an argument showing that the converse of Corollary 6.5 does not hold.
2 (δ0 + δ2) for any σ ∈ P +
c .
3 (δ0 + δ1 + δ2) is such a distribution.) Let ν ∈ Dc(1, ∗) be the tracial
R-diagonal ∗-distribution defined by the requirement that the common distribution of ZZ ∗
Choose a distribution eσ ∈ P +
(For instance, eσ = 1
and Z ∗Z in (ChZ, Z ∗i, ν) is equal to B(eσ ) (see [11, Proposition 15.13] for an argument that
26
H. BERCOVICI, A. NICA, M. NOYES, AND K. SZPOJANKOWSKI
ν exists). The distributions of ZZ ∗ and Z ∗Z are ⊞-infinitely divisible in Pc, by construction.
We show that ν is not ⊞-infinitely divisible in Dc(1, ∗). Suppose, to get a contradiction,
that ν is ⊞-infinitely divisible. Then ν = Ψ(σ, σ) for some σ ∈ P +
2 (δ0 + δ2)(cid:1), and thuseσ = σ ⊠ 1
(6.10) yields B(eσ) = B(cid:0)σ ⊠ 1
contrary to the choice ofeσ.
before Definition 3.6. Indeed, it is immediate that in this case the series P∞
P∞
Example 6.10. (λ-circular distribution.) Let λ > 0 be a parameter.
If in the setting
of Remark 6.7 and Proposition 6.8 we take σ = δ1 (Dirac mass at 1) and t = λ, then
the resulting ∗-distribution ν ∈ R(inf−div)
is the λ-circular distribution mentioned right
n=1 βnzn and
n=1 αnzn from Remark 6.7 are reduced to ηδ1(z) = z and respectively to tηδ1(z) = λz;
hence we have α1 = λ, β1 = 1 and αn = βn = 0 for all n ≥ 2, as required in the definition
of the λ-circular distribution.
2 (δ0 + δ2) because B is injective,
c . Since τ1 = τ2 = B(eσ),
c
In this example, the formulas indicated in Proposition 6.8 for the distributions of ZZ ∗
and of Z ∗Z give free Poisson distributions. In order to make this precise, we need to review
another bit of notation: for any two parameters p, q > 0 one has a free Poisson distribution
of rate p and jump size q, which we will denote as Πp;q, and which appears in the free
analogue of the Poisson limit theorem (see e.g. Proposition 12.11 in [11]). The Marchenko-
Pastur distribution reviewed in Remark 6.6(4) corresponds to p = q = 1 (so "Π1" from
there becomes "Π1;1"). For general p, q > 0, the formula given in Remark 6.6(4) for the
R-transform of Π1 extends to
RΠp;q (z) =
pqz
1 − qz
.
Returning to the example of the λ-circular distribution, an immediate processing of the
formulas (6.4) from Theorem 6.4 gives us that the R-transforms of ZZ ∗ and of Z ∗Z in the
noncommutative probability space (ChZ, Z ∗i, ν) are
RZZ ∗(z) =
λz
1 − z
, RZ ∗Z (z) =
z
1 − λz
.
For our example, this shows that the distributions τ1 and τ2 appearing in (6.7) (Proposition
6.8) are free Poisson distributions:
τ1 = Πλ;1 and τ2 = Π1/λ;λ.
7. Stability of R(inf−div)
c
under free multiplicative convolution
Remark 7.1. In this section we consider the operation ⊠ on Dc(1, ∗), which follows the
multiplication of ∗-free random variables (cf. Definition 4.1(3)). One has the remarkable fact
that whenever µ, µ′ ∈ Dc(1, ∗) and at least one of µ, µ′ is R-diagonal, it follows that µ ⊠ µ′ is
R-diagonal as well (see [11, Proposition 15.8]). If we make the additional assumption that
both µ and µ′ are R-diagonal, then we have explicit formulas for the determining sequences
of µ ⊠ µ′ in terms of the determining sequences of µ and of µ′. To be precise, denote
the determining sequences of µ by (αn)∞
n=1.
n=1, and those of µ′ by (α′
n=1, (βn)∞
n=1, (β′
n)∞
n)∞
ETA-DIAGONALS AND INFINITE DIVISIBILITY FOR R-DIAGONALS
27
Tthen the determining sequences (bαn)∞
Xπ⊔ρ∈N C(2n)
n=1, (bβn)∞
bαn =
π={V1,...,Vp}∈N C(1,3,...,2n−1)
with 1∈V1, and
ρ={W1,...,Wr}∈N C(2,4,...,2n)
(7.1)
n=1 of µ ⊠ µ′ are given by:
αV1βV2 · · · βVpα′
W1 · · · α′
Wr,
bβn =
Xπ⊔ρ∈N C(2n)
π={V1,...,Vp}∈N C(1,3,...,2n−1)
with 1∈V1, and
ρ={W1,...,Wr}∈N C(2,4,...,2n)
β′
V1α′
V2 · · · α′
VpβW1 · · · βWr.
The formulas (7.1) were proved in [7, Proposition 3.9]. They can also be rephrased in
terms of equations for power series, as shown in the next proposition. The formulas (7.2)
in the proposition have appeared before (but only as a conjecture, without proof), in [9,
Section 5.3]. For the reader's convenience, we include the proof of how (7.2) is derived out
of (7.1).
Proposition 7.2. With the notation of Remark 7.1, suppose that we have elements a, b, a′, b′
in a noncommutative probability space (A, ϕ) such that {a, b} is free from {a′, b′} and such
that
Ra(z) =
Ra′(z) =
αnzn, Rb(z) =
α′
nzn, Rb′(z) =
∞Xn=1
∞Xn=1
βnzn,
β′
nzn.
∞Xn=1
∞Xn=1
n=1bαnzn =(cid:16)Ra ◦ Rh−1i
P∞
n=1bβnzn =(cid:16)Rb′ ◦ Rh−1i
P∞
a′
b
◦ Mba′(cid:17) (z),
◦ Ma′b(cid:17) (z).
Assume moreover that β1 6= 0 6= α′
relative to composition. Then:
1, so the series Rb and Ra′ have inverses Rh−1i
b
and Rh−1i
a′
(7.2)
Proof. The second equation in (7.2) follows from the first one if we substitute b′, a′, b for
a, b, a′, respectively. To prove the first equation, we fix an n ∈ N and we suitably structure
the formula for bαn provided in (7.1). Let us also momentarily fix an m ≤ n and a set
V1 = {2i1 − 1, . . . , 2im − 1}, where 1 = i1 < · · · < im ≤ n. Denote nk = ik+1 − ik,
k = 1, . . . , m, where im+1 = n + 1. Note that n1 + · · · + nm = n and that V1 can recovered
from n1, . . . , nm. Use the moment-cumulant formula as in the proof of Proposition 3.5
(using the cumulant functionals and the fact that the mixed cumulants of a′ and b vanish
on account of freeness) to obtain
Xπ⊔ρ∈N C(2n)
π={V1,...,Vp}∈N C(1,3,...,2n−1)
ρ={W1,...,Wr}∈N C(2,4,...,2n)
αV1βV2 · · · βVpα′
W1 · · · α′
Wr = αm
ϕ(a′(ba′)nk−1).
mYk=1
Letting V1 vary, (7.1) yields, for the n ∈ N that we had fixed:
nXm=1
αm Xn1+···+nm=n
mYk=1
bαn =
ϕ(a′(ba′)nk−1).
28
H. BERCOVICI, A. NICA, M. NOYES, AND K. SZPOJANKOWSKI
We now let n vary in N, and get that
(7.3)
where
(7.4)
∞Xn=1
∞Xn=1bαnzn =
∞Xn=1
g(z) =
αm(g(z))m = Ra(g(z)),
ϕ(a′(ba′)n−1)zn ∈ C[[z]].
It remains to show that the series g introduced in (7.4) is equal to Rh−1i
◦ Mba′ or,
equivalently, that one has Rb ◦ g = Mba′. To see this, apply again the moment-cumulant
formula (using, as in [11, Theorem 14.4] the fact that b is free from a′) to obtain
b
(7.5)
ϕ((ba′)n) =
βV1 · · · βVpα′
W1 · · · α′
Wr.
Xπ⊔ρ∈N C(2n)
π={V1,...,Vp}∈N C(1,3,...,2n−1)
ρ={W1,...,Wr}∈N C(2,4,...,2n)
In (7.5) we can list the blocks of π such that 1 ∈ V1. A similar argument to the one used
above to structure the formula for bαn shows now that
mYk=1
βm Xn1+···+nm=n
nXm=1
ϕ((ba′)n) =
and this implies the desired relation Mba′ = Rb ◦ g.
ϕ(a′(ba′)nk−1), n ∈ N,
(cid:3)
Remark 7.3. With the notation of the preceding proposition, suppose that βn = 0 for
every n ∈ N. Then the only non-zero term in the first equality in (7.1) corresponds to
The next corollary presents a reformulation of (7.2) which has the advantage that it
π = 1n and ρ = 0n, and therefore bαn = αn(α′
introduces in discussion two power series F and eF , related with the subordination results
Corollary 7.4. In the framework of Proposition 7.2, we have
of [6].
1)n.
(7.6)
where F = M h−1i
b
(P∞
n=1bαnzn = Ra (F (z) (1 + Mb(F (z))))
n=1bβnzn = Rb′(cid:16)eF (z)(cid:16)1 + Ma′(eF (z))(cid:17)(cid:17) ,
P∞
◦ Mba′ and eF = M h−1i
◦ Ma′b.
a′
Proof. By symmetry, it suffices to prove the first of the two equations. Using (7.2), we see
that we must verify the identity
(Rh−1i
(7.7)
b
◦ Mba′)(z) = F (z)(cid:0)1 + Mb(F (z))).
Recalling the assumption that ϕ(b) 6= 0, the functional equation Mb(z) = Rb(z(1 + Mb(z)))
can be rewritten as
(7.8)
Rh−1i
b
(w) = (1 + w)M h−1i
b
(w)
(see [11, Remark 16.18]). Substitute Mba′ for w in (7.8) to find that
(cid:0)Rh−1i
b
◦ Mba′(cid:1)(z) = (1 + Mba′(z)) ·(cid:0)M h−1i
b
◦ Mba′(cid:1)(z) = (1 + Mba′ (z)) · F (z).
Finally, the definition of F implies that Mba′ (z) = Mb(F (z)), and using this equality in the
right hand side of the preceding equality yields (7.7).
(cid:3)
ETA-DIAGONALS AND INFINITE DIVISIBILITY FOR R-DIAGONALS
29
The following lemma is an immediate consequence of the definition of Ψ (Remark 6.1).
Lemma 7.5. Consider a distribution ν ∈ R(inf−div)
n=1 be its
determining sequences. There exist positive elements a, b in a C ∗-probability space (A, ϕ)
such that
n=1 and (βn)∞
, and let (αn)∞
c
Ra(z) =
αnzn,
and Rb(z) =
βnzn.
∞Xn=1
Moreover, the distributions of a and b are ⊞-infinitely divisible.
Proof. Write ν = Ψ(σ1, σ2), with σ1, σ2 ∈ P +
c . Then
∞Xn=1
∞Xn=1
αnzn = ησ1(z) = RB(σ1)(z),
where B(σ1) is a ⊞-infinitely divisible distribution in P +
c (cf. Remark 6.6(2)). Thus taking
a to be a positive element with distribution B(σ1) in some C ∗-probability space will fulfill
the required conditions. The argument for b is similar.
(cid:3)
In reference to the set of power series E +
c
introduced in Notation 6.2, we record a result
which follows easily from [1, Proposition 2.2].
Proposition 7.6. A series f ∈ C[[z]] belongs to the set E +
c
following three conditions:
if and only if it satisfies the
(i) f has real coefficients;
(ii) f has positive convergence radius;
(iii) f can be extended to an analytic map (still denoted f ) of C+ into C+ such that
f (0) = 0 and Arg(z) ≤ Arg(f (z)) for z ∈ C+.
(cid:3)
is not identically zero.
Corollary 7.7. Suppose that the series f (z) = P∞
Then α1 > 0.
n=1 αnzn ∈ E +
c
Proof. Let n be the smallest integer such that αn 6= 0 and suppose, to get a contradiction,
that either n > 1 or n = 1 and αn < 0. Choose γ ∈ C+ such that γ = 1 and ℑ(γnαn) < 0.
We have limr↓0(f (rγ)/rn)) = γnαn, and therefore ℑf (rγ) < 0 for sufficiently small r,
contrary to Proposition 7.6(iii).
(cid:3)
We are now ready for the main result of this section.
c
.
n)∞
c
n)∞
Theorem 7.8. For every ν, ν′ ∈ R(inf−div)
Proof. Let (αn)∞
n=1) denote the determining
sequences of ν (respectively, ν′). Two applications of Lemma 7.5, combined with a free
product construction, allow us to construct a C ∗-probability space (A, ϕ) and positive
elements a, b, a′, b′ ∈ A such that
we have ν ⊠ ν′ ∈ R(inf−div)
n=1 (respectively (α′
n=1 and (βn)∞
n=1 and (β′
n=1 βnzn,
n=1 β′
(c) {a, b} is free from {a′, b′}.
(a) Ra(z) =P∞
(b) Ra′(z) =P∞
We know from Remark 7.1 that ν ⊠ ν′ is an R-diagonal distribution in Dc(1, ∗). Let (bαn)∞
and (bβn)∞
n=1 αnzn and Rb(z) =P∞
nzn and Rb′(z) =P∞
∞Xn=1bαnzn,
n=1 denote the determining sequences of ν ⊠ ν′, and set
∞Xn=1bβnzn.
bf (z) :=
bg(z) :=
nzn, and
n=1 α′
n=1
30
H. BERCOVICI, A. NICA, M. NOYES, AND K. SZPOJANKOWSKI
because α′
yields α′
c . By symmetry, it suffices to show that
n=1 αn(α′
n=1 αnzn belongs to E +
We dispose first of the simple case in which β1 = 0. Corollary 7.7 yields βn = 0 for all
1z)n. The desired conclusion follows
c . Similarly, if α′
1 = 0, Corollary 7.7
By Proposition 6.3, we have to prove that bf ,bg ∈ E +
c , and this is done by verifying that bf satisfies conditions (i) -- (iii) of Proposition 7.6.
bf ∈ E +
n ∈ N, and Remark 7.3 implies that bf (z) =P∞
1 ≥ 0 and the seriesP∞
n = 0 for all n ∈ N, and then (7.1) implies that bf = 0.
It remains to show that bf ∈ E +
that bf = Ra (F (z) (1 + Mb(F (z)))), where F = M h−1i
Therefore bf ∈ E +
composition of the three power series Ra(z), z(1 + Mb(z)), and F (z). We know that Ra ∈
E +
c by
c .
[5, Proposition 6.1]. Proposition 7.6 shows that the set E +
is closed under composition.
c
(cid:3)
c , thus concluding the proof.
Corollary 7.9. Suppose that ν ∈ R(inf−div)
C ∗-probability space. Then the ∗-distribution of an belongs to R(inf−div)
In other words, bf is the
c . The series z(1 + Mb(z)) also belongs to E +
is the ∗-distribution of an element a in some
1. In this case, Corollary 7.4 shows
It was proved in [6] that F ∈ E +
c when β1 6= 0 6= α′
for every n ∈ N.
c
◦ Mba′.
b
c
Proof. This follows from Theorem 7.8 and the known fact [7, Proposition 3.11] that the
distribution of an is equal to ν ⊠n.
(cid:3)
We conclude the section by looking again at the KMS example, and by describing explic-
itly the BBP parametrization for the powers of a λ-circular element.
Remark 7.10. (1) Suppose that ν, ν′ ∈ Dc(1, ∗) are R-diagonal and satisfy the KMS
condition with parameters t, t′ ∈ (0, ∞), respectively. Consider the ∗-distribution ν ⊠ ν′,
which is R-diagonal as well (see Remark 7.1). We claim that ν ⊠ ν′ also satisfies the KMS
condition, with parameter tt′. Using the same notations for determining sequences as in
Remark 7.1, this claim amounts to the fact that bαn = (tt′)bβn for every n ∈ N. In order to
prove this, we replace αV1 and α′
(7.1) to obtain
W1, respectively, in the first formula
W1 by tβV1 and tβ′
β′
W1α′
W2 · · · α′
WlβV1βV2 · · · βVk.
bαn = (tt′)
Xπ⊔ρ∈N C(2n)
π={V1,...,Vk}∈N C(1,3,...,2n−1)
ρ={W1,...,Wl}∈N C(2,4,...,2n)
1∈V1,2∈W1
To see that the last sum equals bβn, we observe that pairs (π, ρ) as above are in a bijective
correspondence with pairs (eπ,eρ) such that eπ ⊔ eρ ∈ N C(2n) and eπ = {fW1, . . . ,fWl} ∈
N C(1, 3, . . . , 2n − 1) andeρ = {eV1, . . . ,eVk} ∈ N C(2, 4, . . . , 2n). Indeed,eρ andeπ are obtained
aseπ ⊔eρ = γ−1
2n (π ⊔ ρ), where we use the permutation γ2n from the proof of Proposition 3.7.
Thus, the sum above is equal to
β′
fW1
α′
fW2
· · · α′
fWl
β eV1 · · · β eVk,
Xπ′⊔ρ′∈N C(2n)
eπ={fW1,...,fWl}∈N C(1,3,...,2n−1)
eρ={ eV1,..., eVk}∈N C(2,4,...,2n)
1∈fW1,2∈ eV1
and this equals bβn by the second formula (7.1).
(2) Now fix a real number λ > 0 and consider the λ-circular distribution ν = Ψ(δλ, δ1), as
in Example 6.10. If a is an element in some ∗-probability space such that the ∗-distribution
ETA-DIAGONALS AND INFINITE DIVISIBILITY FOR R-DIAGONALS
31
of a is equal to ν, then we will say that a is a λ-circular element. Such elements do of course
exist, for instance we can just take a = Z in the ∗-probability space (ChZ, Z ∗i, ν). If a is
a λ-circular element, then Theorem 7.8 and Corollary 7.9 tell us that every power ak has
∗-distribution ν ⊠k ∈ R(inf−div)
. Moreover, part (1) of the present remark assures us that
ν ⊠k satisfies the KMS condition with parameter λk. Hence for every k ∈ N we have a BBP
parametrization of the form
c
ν ⊠k = Ψ(σ⊎λk
c . For k = 1, we know from Example 6.10 that σ1 is
, σk),
for some probability measure σk ∈ P +
the Dirac mass δ1. The next proposition gives a way of describing σk for k ≥ 2.
Proposition 7.11. Let λ and (σk)∞
probability measures with finite support (τk)∞
k=1 defined by
k=1 be as above, and consider on the other hand the
k
Then one has
(7.9)
τk :=
λk
1 + λk δ0 +
1
1 + λk δ1+λk , k ∈ N.
σk = τ1 ⊠ · · · ⊠ τk−1, k ≥ 2.
Proof. As in Remark 7.10(2), we use the notation ν for the λ-circular distribution. We fix
a k ∈ N and invoke Proposition 7.2 in the special case in which the ∗-distributions µ, µ′
considered there are ν ⊠k and ν, respectively. The power series
(7.10)
βnzn,
β′
nzn,
∞Xn=1
∞Xn=1
∞Xn=1bβnzn
σ1 and σk+1, respectively. (For instance the equalityP∞
σ1 = δ1, for the second power series in (7.10) we actually haveP∞
from Proposition 7.2 are equal in this case to the η-series of the probability measures σk,
n=1 βnzn = ησk (z) follows from the
, σk).) Note that, since
The notation of Proposition 7.2 also include some non-commutating random variables
comments at the end of Remark 6.1 and the fact that ν ⊠k = Ψ(σ⊎λk
k
n=1 β′
nzn = z.
a, b, a′, b′, where b is such that
(7.11)
Rb(z) =
βnzn = ησk (z) = RB(σk)(z).
∞Xn=1
From (7.11) we infer that the distribution of b is B(σk). Similar reasoning, based on the
formulas Rb′(z) = z and Ra′(z) = λz, leads to the fact that a′ and b′ have distributions
δλ and δ1, respectively. As a consequence, we may assume without loss of generality that
a′ = λ and b′ = 1 in their noncommutative probability space.
We are interested in the second relation (7.2) from Proposition 7.2. Due to the very
simple form of Ra′ and Rb′, this equation simplifies to
(7.12)
∞Xn=1bβnzn =
1
λ
Mλb(z).
The same argument as used in (7.11) shows that the left-hand side of (7.12) is equal
to RB(σk+1)(z). On the right-hand side of (7.12) we perform the obvious transformation
Mλb(z) = Mb(λz) = MB(σk )(λz), and this leads us to a direct connection between σk and
σk+1:
(7.13)
RB(σk+1)(z) =
1
λ
MB(σk)(λz).
32
H. BERCOVICI, A. NICA, M. NOYES, AND K. SZPOJANKOWSKI
In order to make use of (7.13), it is convenient to resort to another well-known transform
of free probability, the S-transform. For a probability measure σ ∈ Pc with non-vanishing
mean, one defines the S-transform of σ as the power series
Sσ(z) =
1
z
Rh−1i
σ
(z) =
z + 1
z
M h−1i
σ
(z)
(see, for instance, [11, Definition 18.15 and Remark 18.16 on p. 294]). Some straightforward
processing of Equation (7.13) (multiply both sides by λ, take inverses under composition,
and write the resulting series in terms of the suitable S-transforms) then leads to the formula
(7.14)
SB(σk+1)(z) =
1
1 + λz
SB(σk)(λz).
The formula (7.14) was obtained for a fixed (but arbitrary) k ∈ N. We now unfix k and
use a straightforward induction argument, with base case SB(σ1)(z) = Sδ1(z) = 1, in order
to infer that
(7.15)
SB(σk)(z) =
1
1 + λjz
k−1Yj=1
, ∀ k ∈ N.
It remains to make the connection to the τk indicated in the statement of the proposition.
For every j ∈ N, an elementary calculation shows that B(τj) is the free Poisson distribution
Π1/λj ;λj , where the notation "Πp;q" is as in Example 6.10. Another elementary calculation
shows that the S-transform of Π1/λj ;λj is 1/(1 + λjz). Thus the right-hand side of (7.15)
can be written as SB(τ1)(z) SB(τ2)(z) · · · SB(τk−1)(z).
Now, the S-transform is multiplicative with respect to the operation ⊠ ([11, Corollary
18.17]). Since B is multiplicative as well (Remark 6.6(3)), the observations made in the
preceding paragraph lead to the formula
SB(σk) = SB(τ1 ⊠···⊠τk−1), k ≥ 2.
The required Equation (7.9) follows from here, since B is injective and since a probability
measure with non-vanishing mean is uniquely determined by its S-transform.
(cid:3)
References
[1] S.T. Belinschi, H. Bercovici. Partially defined semigroups relative to free multiplicative convolution,
International Mathematics Research Notices (2005), 65-101.
[2] S.T. Belinschi, A. Nica. η-series and a Boolean Bercovici-Pata bijection for bounded k-tuples, Advances
in Mathematics 217 (2008), 1-41.
[3] S.T. Belinschi, A. Nica. On a remarkable semigroup of homomorphisms with respect to free multiplica-
tive convolution, Indiana University Mathematics Journal 57 (2008), 1679-1713.
[4] H. Bercovici, V. Pata. Stable laws and domains of attraction in free probability theory. With an appendix
by P. Biane, The density of free stable distributions, Annals of Mathematics 149 (1999), 1023-1060.
[5] H. Bercovici, D. Voiculescu. Free convolution of measures with unbounded support, Indiana University
Mathematics Journal 42 (1993), 733-773.
[6] P. Biane. Processes with free increments, Mathematische Zeitschrift 227 (1998), 143-174.
[7] B. Krawczyk, R. Speicher. Combinatorics of free cumulants, Journal of Combinatorial Theory, Series
A 90 (2000), 267-292.
[8] F. Lehner. Cumulants in noncommutative probability I. Noncommutative exchangeable systems. Math-
ematische Zeitschrift 248 (2004), 67-100.
[9] A. Nica, D. Shlyakhtenko, R. Speicher. R-diagonal elements and freeness with amalgamation, Canadian
Journal of Mathematics 53 (2001), 355-381.
[10] A. Nica, R. Speicher. R-diagonal pairs -- a common approach to Haar unitaries and circular elements,
Fields Institute Communications 12 (1997), 149-188.
ETA-DIAGONALS AND INFINITE DIVISIBILITY FOR R-DIAGONALS
33
[11] A. Nica, R. Speicher. Lectures on the combinatorics of free probability, London Mathematical Society
Lecture Note Series 335, Cambridge University Press, 2006.
[12] D. Shlyakhtenko. Free quasi-free states, Pacific Journal of Mathematics, 177 (1997), 329-368.
[13] R. Speicher, R. Woroudi. Boolean convolution, Fields Institute Communications 12 (1997), 267-279.
[14] D. Voiculescu. Addition of certain noncommuting random variables, Journal of Functional Analysis 66
(1986), 323 -- 346.
Hari Bercovici: Department of Mathematics, Indiana University, Bloomington, Indiana,
USA.
E-mail address: [email protected]
Alexandru Nica: Department of Pure Mathematics, University of Waterloo, Ontario,
Canada.
E-mail address: [email protected]
Michael Noyes: Department of Mathematics, Bard High School Early College, New York,
New York, USA.
E-mail address: [email protected]
Kamil Szpojankowski: Department of Pure Mathematics, University of Waterloo, Ontario,
Canada, and Faculty of Mathematics and Information Science, Warsaw University of Tech-
nology, Poland.
E-mail address: [email protected], [email protected]
|
1110.4476 | 2 | 1110 | 2013-01-10T11:49:52 | The Weyl group of the Cuntz algebra | [
"math.OA",
"math.DS",
"math.FA",
"math.GR"
] | The Weyl group of the Cuntz algebra O_n, with n finite, is investigated. This is (isomorphic to) the group of polynomial automorphisms of O_n, namely those induced by unitaries that can be written as finite sums of words in the canonical generating isometries and their adjoints. A necessary and sufficient algorithmic combinatorial condition is found for deciding when a polynomial endomorphism restricts to an automorphism of the canonical diagonal MASA. Some steps towards a general criterion for invertibility of such endomorphisms on the whole of O_n are also taken. A condition for verifying invertibility of a certain subclass of polynomial endomorphisms is given. First examples of polynomial automorphisms of O_n not inner related to permutative ones are exhibited, for every n. In particular, the image of the Weyl group in the outer automorphism group of O_n is strictly larger than the image of the reduced Weyl group analyzed in previous papers. Results about the action of the Weyl group on the spectrum of the diagonal are also included. | math.OA | math |
The Weyl Group of the Cuntz Algebra
Roberto Conti∗, Jeong Hee Hong†and Wojciech Szyma´nski*‡
Abstract
The Weyl group of the Cuntz algebra On is investigated. This is (isomorphic
to) the group of polynomial automorphisms λu of On, namely those induced by
unitaries u that can be written as finite sums of words in the canonical generating
isometries Si and their adjoints. A necessary and sufficient algorithmic combinato-
rial condition is found for deciding when a polynomial endomorphism λu restricts
to an automorphism of the canonical diagonal MASA. Some steps towards a gen-
eral criterion for invertibility of λu on the whole of On are also taken. A condition
for verifying invertibility of a certain subclass of polynomial endomorphisms is
given. First examples of polynomial automorphisms of On not inner related to
those induced by unitaries from the core UHF subalgebra are exhibited, for ev-
ery n ≥ 2. In particular, the image of the Weyl group in the outer automorphism
group of On is strictly larger than the image of the restricted Weyl group analyzed
in previous papers. Results about the action of the Weyl group on the spectrum
of the diagonal are also included.
MSC 2010: Primary 46L40, 46L55, Secondary 37B10
Keywords: Cuntz algebra, MASA, automorphism, endomorphism, Cantor set
∗This research was supported through the programme "Research in Pairs" by the Mathematisches
Forschungsinstitut Oberwolfach in 2011.
†This research was supported by Basic Science Research Program through the National Research
Foundation of Korea (NRF) funded by the Ministry of Education, Science and Technology (Grant No.
2012R1A1A2039991).
‡Partially supported by the FNU Rammebevilling 'Operator algebras and applications', the Nord-
Forsk Research Network 'Operator algebra and dynamics' (grant #11580), and the FNU Forskningspro-
jekt 'Structure and Symmetry'.
1
1
Introduction
Consider a finite alphabet {1, 2, . . . , n} with n ≥ 2 letters, and let W be the set of finite
words on this alphabet. We say that two words are orthogonal if one is not the initial
subword of the other. Let Σ be the collection of finite subsets of W consisting of mutually
orthogonal words. We consider the set of n words {α1, α2, . . . , αn} (beginning with the
same subword α and ending with all the distinct letters of the alphabet) equivalent to the
single word α, and this extends to an equivalence relation on Σ. The set of equivalence
{β1, . . . , βr}, with the property that both {α1, . . . , αr} and {β1, . . . , βr} are equivalent
to the empty word. Such a U determines recursively a sequence of transformations
classes is denoted eΣ. We fix U ∈ Σ, comprised of two ordered subsets: {α1, . . . , αr} and
Tk : eΣ → eΣ such that:
if γ = αjµ for some j then T1(γ) = βjµ, and
if Tk−1(γ) = ναjµ for some j and a word ν of length k − 1 then Tk(γ) = νβjµ.
Thus each transformation Tk is determined by a certain Turing machine,
[10], and
hence it is computable for any finite set of inputs. We are interested in the following
stabilization problem of the recursive process Tk ◦ Tk−1 ◦ . . . ◦ T1 : eΣ → eΣ. For
what U it holds that for each A ∈ eΣ there exists an m such that for all k ≥ m we
have Tk ◦ . . . ◦ Tm ◦ . . . ◦ T1 = Tm ◦ . . . ◦ T1? We provide a surprisingly simple complete
solution to this (suitably reformulated in more algebraic terms) stabilization problem in
Theorem 3.7, below.
We can reformulate the above described combinatorial setup in topological terms, as
follows. Let Xn be the space of all (one-sided) infinite words. Then Xn is a Cantor set
with the product topology and elements of eΣ are in bijective correspondence with its
clopen subsets. Our stabilization problem is then equivalent to injectivity of a certain
continuous map ψU : Xn → Xn determined naturally by U. Then, by the Gelfand
duality, this problem is equivalent to surjectivity of a unital, injective ∗-homomorphism
ψU : C(Xn) → C(Xn), dual to ψU . That is, we ask if ψU is a homeomorphism of Xn or,
equivalently, if ψU is an automorphism of C(Xn).
identity. In our setting, the element U gives rise to a unitary u = Pr
Somewhat paradoxically, it is most natural to view this problem in the context
of much larger and noncommutative Cuntz algebras On, [8]. These are C ∗-algebras
generated by n isometries S1, . . . , Sn of a Hilbert space with ranges adding up to the
βj in On,
which in turn leads to a necessarily injective, unital ∗-endomorphism λu of On such that
λu(Sj) = uSj for all j = 1, . . . , n. The C ∗-subalgebra Dn of On generated by ranges
of all finite products of S1, . . . , Sn is maximal abelian in On and naturally isomorphic
to C(Xn). The restriction of endomorphism λu to Dn coincides with ψU . Thus, our
combinatorial stabilization problem is equivalent to the problem of surjectivity of
λuDn. The question of surjectivity of λu itself is very interesting as well and closely
related to the so called Weyl group of the Cuntz algebra. This last problem appears
very difficult and an algorithm for deciding surjectivity of an arbitrary λu has not been
found yet, although we make some headway towards its solution, below.
j=1 Sαj S∗
2
The present paper is a continuation of our investigations of the subgroup Aut(On, Dn)
of automorphisms of On which globally preserve the canonical diagonal MASA Dn, and
of related endomorphisms of On, [6, 4, 5, 11, 1, 2]. As shown in [9], the quotient of
Aut(On, Dn) by its normal subgroup AutDn(On), consisting of those automorphisms
which fix Dn point-wise, is discrete. Since AutDn(On) is a maximal abelian subgroup
of Aut(On), [9], it is natural to call this quotient the Weyl group of On. The Weyl
group contains a natural interesting subgroup corresponding to those automorphisms
which also globally preserve the core UHF-subalgebra Fn of On, called the restricted
Weyl group of On.
It was shown in [2] that the image of the restricted Weyl group
in the outer automorphism group of On can be embedded into the quotient of the
automorphism group of the full two-sided n-shift by its center, and this embedding is
surjective whenever n is prime. In the present article, we focus our attention on the
(full) Weyl group. It was shown in [6] that the Weyl group is isomorphic with the group
of those automorphisms λu ∈ Aut(On) whose corresponding unitaries u may be written
as a sum of words in {Si, S∗
j }. (The collection of all such unitaries in On is denoted
Sn.) The structure of the Weyl group is highly complicated. For example, it contains
the Thompson F group in its intersection with Inn(On), [14]. Our main objective here
is investigation of the structure of the Weyl group of On, its action on the diagonal
MASA, and determining which unitaries u ∈ Sn give rise to automorphisms.
The present paper is organized as follows.
In section 2, we set up notation and
review some basic facts on Cuntz algebras and their endomorphisms.
In section 3,
we study the restriction of an endomorphism λu, u ∈ Sn, to the diagonal Dn. We
give an algorithmic criterion for λuDn to be an automorphism of Dn, Theorem 3.7.
Its proof is combinatorial and involves equivalence of surjectivity of λuDn with the
stabilization problem mentioned above. In section 4, we investigate the problem when
λu is an automorphism of the entire On. In Proposition 4.3, we present a combinatorial
procedure for deciding this question for a certain large class of unitaries u ∈ Sn. In
section 5, we exhibit endomorphisms λu, u ∈ Sn, which are not inner related to the ones
of the form λw with w a unitary in the core UHF-subalgebra Fn. In particular, we show
with concrete examples that the image in Out(On) of the Weyl group is strictly larger
then the image of the restricted Weyl group, Theorem 5.2 and Corollary 5.3. Finally,
in section 6, we look at the action induced by λu on the space Xn, the spectrum of the
diagonal Dn. We characterize homeomorphisms of Xn corresponding to automorphisms
Ad(u), u ∈ Sn, and describe the fixed points in Xn for some exotic automorphisms λu.
2 Notation and preliminaries
infinite C ∗-algebra generated by n isometries S1, . . . , Sn satisfying Pn
If n is an integer greater than 1, then the Cuntz algebra On is a unital, simple, purely
i = 1, [8].
We denote by W k
n the set of k-tuples µ = (µ1, . . . , µk) with µm ∈ {1, . . . , n}, and by Wn
the union ∪∞
n , where W 0
n = {0}. We call elements of Wn multi-indices. If µ ∈ W k
k=0W k
n
then µ = k is the length of µ. For µ, ν ∈ Wn we write µ ≺ ν if µ is an initial subword
n and µ ≺ ν, then we denote by ν − µ the word in W m−k
of ν. If µ ∈ W k
i=1 SiS∗
n , ν ∈ W m
n
3
n
obtained from ν by removing its initial segment µ. Also, if µ ∈ W k
n then we denote by
s(µ) its first letter, and by µ the word in W k−1
obtained from µ by removing s(µ). We
denote by µ ∧ ν the collection of all non-empty words η such that both η ≺ µ and η ≺ ν.
If µ = (µ1, . . . , µk) ∈ Wn then Sµ = Sµ1 . . . Sµk (S0 = 1 by convention) is an isometry
with range projection Pµ = SµS∗
i i = 1, . . . , n} can be uniquely
expressed as SµS∗
µ. Every word in {Si, S∗
ν , for µ, ν ∈ Wn [8, Lemma 1.3].
k=0F k
We denote by F k
n the C ∗-subalgebra of On spanned by all words of the form SµS∗
ν ,
µ, ν ∈ W k
n , which is isomorphic to the matrix algebra Mnk (C). The norm closure Fn of
∪∞
n is the UHF-algebra of type n∞, called the core UHF-subalgebra of On, [8]. We
denote by τ the unique normalized trace on Fn. The core UHF-subalgebra Fn is the
fixed-point algebra for the gauge action γ : U(1) → Aut(On), such that γz(Sj) = zSj for
z ∈ U(1) and j = 1, . . . , n. We denote by E the faithful conditional expectation from
On onto Fn given by averaging with respect to the normalized Haar measure:
E(x) =Zz∈U (1)
γz(x)dz.
For an integer m ∈ Z we denote O(m)
subspace for γ. Then O(0)
have O(m)
n = FnSα and O(−m)
= S∗
n
αFn.
n = Fn and for each positive integer m and each α ∈ W m
:= {x ∈ On : γz(x) = zmx, ∀z ∈ U(1)}, a spectral
n we
n
The C ∗-subalgebra of On generated by projections Pµ, µ ∈ Wn, is a MASA (maximal
abelian subalgebra) in On. We call it the diagonal and denote Dn. Every projection
in Dn of the form Pα for some α ∈ Wn will be called standard. The spectrum of Dn is
naturally identified with Xn -- the full one-sided n-shift space. For d ∈ Dn we denote
by Md a map Md : Dn → Dn such that Md(x) = dx.
As shown by Cuntz in [9], there exists the following bijective correspondence be-
tween unitaries in On (whose collection is denoted U(On)) and unital ∗-endomorphisms
of On (whose collection we denote End(On)). A unitary u ∈ U(On) determines an
endomorphism λu by
λu(Si) = uSi,
i = 1, . . . , n.
Conversely, if ρ : On → On is an endomorphism, then Pn
i = u gives a unitary
u ∈ On such that ρ = λu. Composition of endomorphisms corresponds to a 'convolution'
multiplication of unitaries as follows:
i=1 ρ(Si)S∗
λu ◦ λw = λλu(w)u.
(1)
If A is either a unital C ∗-subalgebra of On or a subset of U(On), then we denote λ(A) =
{λu ∈ End(On) : u unitary in A} and λ(A)−1 = {λu ∈ Aut(On) : u unitary in A}.
We denote by ϕ the canonical shift:
ϕ(x) =
nXi=1
SixS∗
i , x ∈ On.
If we take u =Pn
j then ϕ = λu. For all u ∈ U(On) we have Ad(u) = λuϕ(u∗).
It is well-known that ϕ leaves Dn globally invariant. We denote by φ the standard left
i,j=1 SiSjS∗
i S∗
4
inverse of ϕ, defined as
φ(x) =
1
n
nXi=1
S∗
i xSi, x ∈ On.
If u ∈ U(On) then for each positive integer k we denote
uk = uϕ(u) · · · ϕk−1(u).
(2)
βu∗
β) = ukSαS∗
We often consider elements of On of the form w = P(α,β)∈J cα,βSαS∗
k stands for (uk)∗. If α and β are multi-indices of
Here ϕ0 = id, and we agree that u∗
length k and m, respectively, then λu(SαS∗
m. This is established through a
repeated application of the identity Six = ϕ(x)Si, valid for all i = 1, . . . , n and x ∈ On.
β, where J
is a finite collection of pairs (α, β) of words α, β ∈ Wn and cα,β ∈ C. We denote
J1 = {α : ∃(α, β) ∈ J } and J2 = {β : ∃(α, β) ∈ J }. Of course, such a presentation
(if it exists) is not unique, but once it is chosen then we associate with it two integers:
ℓ = ℓ(J ) = max{α : (α, β) ∈ J } and ℓ′ = ℓ′(J ) = max{α, β : (α, β) ∈ J }. Note
that if w ∈ Fn then w ∈ F ℓ
ν , where ϕ(J ) =
{((i, α), (β, i)) : i ∈ W 1
n , (α, β) ∈ J } and c(i,α),(β,i) = cα,β. Then ℓ(ϕ(J )) = ℓ(J ) + 1
and ℓ′(ϕ(J )) = ℓ′(J ) + 1. In particular, we consider the group Sn of those unitaries in
On which can be written as finite sums of words, i.e.
β.
n. We have ϕ(w) = P(µ,ν)∈ϕ(J ) cµ,νSµS∗
in the form u = P(α,β)∈J SαS∗
Note that such a sum is a unitary if and only if Pα∈J1 Pα = 1 = Pβ∈J2 Pβ. We also
n for the subgroups of Sn consisting of permutative
n = Sn ∩ F k
write Pn = Sn ∩ Fn and P k
unitaries.
For algebras A ⊆ B we denote by NB(A) = {u ∈ U(B) : uAu∗ = A} the normalizer
of A in B and by A′ ∩ B = {b ∈ B : (∀a ∈ A) ab = ba} the relative commutant of A in
B. We also denote by Aut(B, A) the collection of all those automorphisms α of B such
that α(A) = A, and by AutA(B) those automorphisms of B which fix A point-wise.
AutDn(On) is a normal subgroup of Aut(On, Dn), and the corresponding quotient is
called the Weyl group of On. It was shown in [9] that the Weyl group is discrete, and
more recently in [6] that it is isomorphic to λ(Sn)−1. The quotient of Aut(On, Dn) ∩
Aut(On, Fn) by AutDn(On) is called the restricted Weyl group of On. It is isomorphic
to λ(Pn)−1, [6]. The image of λ(Sn)−1 in Out(On) is called the outer Weyl group of
On and such image of λ(Pn)−1 is called the restricted outer Weyl group of On. As
shown in [3, Theorem 3.7], the outer Weyl group is just the quotient of λ(Sn)−1 by
{Ad(u) : u ∈ Sn}. Likewise, the restricted outer Weyl group is the quotient of λ(Pn)−1
by {Ad(w) : w ∈ Pn}.
3 The automorphisms of the diagonal
In this section, we give an algorithmic crierion for deciding if the restriction to Dn of an
endomorphisms λu, u ∈ Sn, gives rise to an automorphism of the diagonal Dn.
Lemma 3.1 Let u ∈ Sn be such that u = P(α,β)∈J SαS∗
n for all k ∈ N.
n) ⊆ Dkℓ
λu(Dk
β, and let ℓ = ℓ(J ). Then
5
Proof. We proceed by induction on k. For k = 1 and i ∈ W 1
n we have
λu(Pi) = uPiu∗ = X(α,β), (α′,β ′)∈J
SαS∗
βPiSβ ′S∗
α′ = X(α,β)∈J , s(β)=i
SαS∗
α
and thus λu(D1
n) ⊆ Dℓ
n. For the inductive step, suppose that λu(Dk
n) ⊆ Dkℓ
n . Then
λu(Dk+1
n
) = λu(D1
nϕ(Dk
n)) = λu(D1
n)(Ad(u)ϕλu)(Dk
n) ⊆ Dℓ
n(Ad(u)ϕ)(Dkℓ
n ) ⊆ Dℓ
nD(k+1)ℓ
n
and thus λu(Dk+1
n
) ⊆ D(k+1)ℓ
n
.
✷
Proposition 3.2 Let u ∈ Sn. Then the following hold.
1. λuDn is an automorphism of Dn if and only if for each α ∈ Wn the sequence
{u∗
kPαuk} eventually stabilizes.
2. λu is an automorphism of On if and only if:
(a) λuDn is an automorphism of Dn, and
(b) there exists a w ∈ Sn such that λwDn = (λuDn)−1.
Proof. Ad 1. This is well-known, [9]. Indeed, the sequence {u∗
kPαuk} eventually stabi-
lizes if and only if Pα belongs to the range of λu (and then λu(lim u∗
kPαuk) = Pα). Thus,
condition 1. is equivalent to λu(Dn) = Dn, i.e. to λuDn being an automorphism of Dn.
Ad 2. If λu is automorphism of On, then λu(Dn) ⊆ Dn since u ∈ NOn(Dn). Thus
λu(Dn) = Dn, since Dn is a MASA in On. Also, there exists w ∈ Sn such that λ−1
u = λw,
[15, 6, 12]. This gives one implication of part 2. For the reversed implication, suppose
that (a) and (b) hold. Then λuλwDn = id. Thus λuλw is an automorphism of On by [1,
Proposition 3.2]. Consequently, λu being surjective is automorphism of On.
✷
1S∗
2 +S2S∗
1 +S1S2S∗
Example 3.3 (a) If u = S1S1S∗
Indeed, a straightforward calculation shows that λu(D2)P2 = CP2.
(b) If u = S2S1S∗
projection P12 does not satisfy (Condition 1) of Proposition 3.2.
(c) If u = u∗ = S1S∗
projection P11 does not satisfy (Condition 1) of Proposition 3.2.
2 + P21 + S2S2S∗
1 + S2S2S∗
2S∗
1S∗
2 + S1S∗
2S∗
2S∗
1 ∈ S2 then λuD2 is not surjective.
Indeed,
Indeed,
2 ∈ S2 then λuD2 is not surjective.
2 ∈ S2 then λuD2 is not surjective.
Our next result shows that in order to verify (Condition 1) in Proposition 3.2 it is
enough to check it only for finitely many projections. Before that, we note the following.
β. Then for each word µ ∈ Wn and for each
Let u ∈ Sn be such that u = P(α,β)∈J SαS∗
(α, β) ∈ J we have
In particular, Ad(Pβ) = Pα.
Ad(u)(Pβµ) = Pαµ.
(3)
6
Lemma 3.4 Let u ∈ Sn be such that u = P(α,β)∈J SαS∗
λuDn is an automorphism of Dn if and only if for each γ ∈ W ℓ′
eventually stabilizes.
β, and let ℓ′ = ℓ′(J ). Then
kPγuk}
n the sequence {u∗
Proof. For short, say a projection Q ∈ Dn is "bad" (relative to u) if the sequence
{u∗
kQuk} does not stabilize, and "good" otherwise. Also, let r be the non-negative
integer uniquely defined by requiring that all projections in Dr
n are good, but there is a
bad projection in Dr+1
is bad as
well. We claim that r + 1 ≤ ℓ′.
. Then at least one of the minimal projections in Dr+1
n
n
n
n
Reasoning by way of contradiction, suppose that ℓ′ < r +1 and let p = Pγ, γ ∈ W r+1
,
. Now, u∗pu can be computed using equation
be such a bad minimal projection in Dr+1
(3), with u replaced by u∗, and hence it is still of the form Pγ1 for some γ1 ∈ Wn. In
this process, by replacing the initial α-segment of γ with the corresponding β, the last
r + 1 − ℓ′ digits will remain unaltered. Now, the assumption that p is bad easily implies
that the projection Pδ := nφ(u∗pu), obtained from u∗pu by deleting the first digit of
γ1, is still bad. By assumption, one must have δ ≥ r + 1, and hence u∗pu /∈ Dr+1
.
In other words, when computing u∗pu we have replaced a word α in γ with a longer
word β. This implies that when in the next step we consider Pγ2 := ϕ(u)∗u∗puϕ(u), the
last r + 1 − ℓ′ digits of γ2 will coincide again with those of γ. Also, (nφ)2(Pγ2) must
be bad, i.e. u∗
. Repeating this argument, one can indeed show that
u∗
for all k = 1, 2, . . ., and moreover the last r + 1 − ℓ′ digits of γk
kpuk = Pγk /∈ Dr+k
coincide with those of γ for any k. All in all, this means that these last digits of γ indeed
play no role in the whole process and defining γ′ simply to be the multi-index obtained
from γ by deleting its last digit, the very same argument would readily show that Pγ ′ is
still bad. But then Pγ ′ ∈ Dr
n, contradicting our assumption.
2pu2 = Pγ2 /∈ Dr+2
n
n
n
By the above, if there are bad projections at all, we can find at least one of them in
n . As a sum of good projections is clearly good, it is also clear that in that case there
✷
Dℓ′
is always such a bad projection of the form Pγ, where γ = ℓ′.
All in all, for u ∈ Sn one has
λu(Dn) = Dn ⇔ Dℓ′
n ⊆ λu(Dn),
(4)
where ℓ′ is as in the statement of Lemma 3.4.
In view of Lemma 3.4, the process of determination if an endomorphism λuDn, u ∈
Sn, is an automorphism of the diagonal can be reduced to verification if a certain finite
collection of projections is contained in its range. This is a very significant reduction but
still it is not clear a priori if this process can be carried out in finately many steps even
for a single projection! This question has a positive answer in the case of a permutative
unitary u ∈ Pn, as shown in [15, 6], but the present case is much more complicated.
Now, we will describe a key construction of the present paper, producing a certain finite
directed graph corresponding to a unitary u ∈ Sn. Non occurence of closed paths on
the graph will turn out to be equivalent to λuDn being automorphism of Dn.
Given u = P(α,β)∈J SαS∗
vertices Γ0
Γu, we proceed by induction.
β in Sn, we define a finite directed graph Γu, whose
u will be identified with certain subsets of J1. In order to construct the graph
7
The initial step. To begin with, we include in Γ0
the empty set ∅. Now, given (α, β) ∈ J , one of the following three cases takes place:
u each singleton subset {α} of J1 and
(i) β = (i) for some i ∈ W 1
n ,
(ii) β = (i, α′, µ) for some i ∈ W 1
n , α′ ∈ J1, and a word µ (possibly empty),
(iii) β = (i, µ) for some i ∈ W 1
elements of J1, namely α′
1, . . . , α′
r.
n and a word µ which is an initial segment of at least two
Depending on the case, we enlarge the graph Γu as follows. In case (i), we add an
edge from vertex {α} to vertex ∅ with label i. In case (ii), we add an edge from vertex
{α} to vertex {α′} with label i. In case (iii), we add a vertex A = {α′
r} and an
edge from {α} to A with label i.
The inductive step. Let A ⊆ J1 be a vertex added to Γ0
u in the preceding step, but
A 6= ∅ and A not a singleton set. For each j ∈ W 1
n we proceed as follows. Let Bk,
k = 1, . . . , m, be the collection of all those already constructed vertices of Γu that there
k=1 Bk = J1 then we add
k=1 Bk
(if such a vertex does not exist already), and we add an edge from A to B with label j.
exists an α ∈ A and an edge from {α} to Bk with label j. If Sm
an edge from A to ∅ with label j. If Sm
k=1 Bk 6= J1 then we add a vertex B =Sm
1, . . . , α′
Continuing inductively in the above described manner, we produce the desired graph
Γu. This is a finite, directed, and labeled graph. Each vertex emits at most n edges,
carrying distinct labels from the set W 1
n . Any finite path on the graph Γu may be
uniquely identified with a pair (A, ν), where A ∈ Γ0
u is the initial vertex of the path and
ν = (ν1, ν2, . . . , νk) is the word such that νj is the label of the jth edge entering this
path. For such a path (A, ν), we denote its terminal vertex by ν(A). We will denote by
Γ1
u the set
of finite paths. Γ∗
u(A), respectively, are the sets of finite paths and paths of
length k which begin at the vertex A.
u the set of paths of length k, and by Γ∗
u the set of edges of the graph, by Γk
u(A) and Γk
Example 3.5 Let u = S1S∗
1. Then the corresponding graph
Γu has five vertices and five edges, and looks as follows. In particular, there is a closed
(directed) path on the graph.
2 + S2S2S∗
2 + S2S1S∗
2 S∗
1S∗
.....................................................
....
..
..
.
..
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
..
.
..
...
. 21
..
..
.
..
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
..
..
..
...
.....................................................
(2)
..................................................................................................................
..................
.............................................................
......................................................
..
..
..
.
.
.
.
..
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
..
.
.
.
..
.
...
. 22
...
..
..
.
..
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
..
.
.
..
..
..
...
.....................................................
(1)
..................................................................................................................
..................
.............................................................
.....................................................
....
..
..
.
..
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
..
.
..
...
. 1
..
..
.
..
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
..
..
..
...
.....................................................
......................................................
..
..
..
.
.
.
.
..
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
..
.
.
.
..
.
...
. ∅
...
..
..
.
..
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
..
.
.
..
..
..
...
.....................................................
(2)
.............................................................................................................................................................................................................................................................................
..................
.................................
..
..
....
.................................
..................
(2)
.............................................................................................................................................................................................................................................................................
.........................................................................................................................................................................................................................................................................................................................
(1)
.................................
..................
.
.
..
.
..
.
..
..
..
..
...
...
...
....
.....
...
....................................
.............................
..
...........................
..
..
..
.
..
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
..
.
.
.
.
..
21, 22
........................
....................................
..
...
...........................
..
.
..
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
..
.
.
..
Example 3.6 Let u = S12S∗
the corresponding graph Γu looks as follows:
21 +S11S∗
221 +S21S∗
222 +S2222S∗
11 +S2221S∗
122 +S221S∗
121. Then
8
............................
.............................
....
..
..
.
..
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
..
.
.
..
..
...
...
..
..
.
..
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
..
.
..
..
....
2222
.........................................................
...........................................................
...
..
..
..
.
.
.
.
.
..
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
..
.
..
.
..
...
11
..
..
..
.
..
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
..
.
..
...
.............................
...........................
....
........................................................
....
...
...
..
..
.
.
..
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
..
.
.
.
..
.
..
..
.
..
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
..
.
.
..
..
..
221
..
....
...........................
...............................
..............................................................................................................................
.............................................................................................................................................................
(1)
..................
..................................
................
............................
............................
...
....
..
11, 12
...
.
..
.
.
..
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
..
.
.
.
..
..
.
..
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
..
.
.
..
..
...
.............................
.............................
................
..
..
...
(2)
............................
....
..
...............................................................
............................
...
..................
..................................
11, 12, 21
..
.
..
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
..
.
.
..
...
.
..
.
.
..
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
..
.
.
.
..
..
...
.............................
...............................................................
..
..
...
...
...
...
.............................
..
...
...
...
...
.
..
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
(2)
..............................
..............................
...
..
.
..
.
.
..
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
..
.
.
.
.
..
..
..
...
12
...
..
.
..
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
..
.
..
..
..
...........................................................
..............................
..............................
...
..
.
..
.
.
..
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
..
.
.
.
.
..
..
..
...
2221
...........................................................
...
..
.
..
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
..
.
..
..
..
(1)
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
..
.
..
...
...
...
..
...
...
...
..
...
...
...
..
...
...
...
(2)
..
...
...
...
...
..
...
...
...
..
...
...
...
..
...
...
.
...
...
..
..
...
...
.
.
.
.
.
.
...
..
...
.
...
...
.
.
..
.
..
.
...
...
..
.
...
..
.....................
..................
.
..
...
..
...
.
...
...
.
..
...
...
.
.
..
..
.
.
.
.
.
.
.
.
.
...
...
...
...
...
...
..
..
...
...
...
...
..
...
...
...
..
...
...
...
..
...
(1)
...
...
..
...
...
...
..
...
..............................
..............................
...
..
.
..
.
.
..
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
..
.
.
.
.
..
..
..
...
∅
...
..
.
..
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
..
.
..
..
..
...........................................................
(2)
.........................................................................................................................................................................................
.....................................
.
.
.
.
.
..
.
..
.
..
...
..
(1)
..................
..................................
..............................................................................................................................
..............................................................................................................................
........................................................
....
...
..
..
.
.
..
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
..
.
.
.
..
.
..
21
...
..
.
..
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
..
.
.
..
..
..
..
....
...........................
...............................
(2)
.............................................................................................................................................................
...............................
..
..
..
.
..
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
..
..
.
...
..............................
..................
..................................
221, 2221, 2222
.............................................................................................................................................................
...............................
...
...
..
...
...
..
..........................
....
...
..
.
..
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
..
.
.
..
..
..
For α ∈ J1, we say that {α} is a splitting vertex if it emits an edge to a vertex
A ⊆ J1 such that A contains at least two elements. This happens when for (α, β) ∈ J
we have that β is an initial subword of more than one α ∈ J1. For example, α = 1 in
Example 3.5 and α1 = 12, α2 = 21, α3 = 2221 and α4 = 2222 in Example 3.6 are all
splitting vertices.
The point of introducing graph Γu is that it conveniently captures the essential
kPαuk, appearing in part 1 of Proposition 3.2.
features of the process of calculating u∗
Indeed, for A ∈ Γ0
u denote PA :=Pα∈A Pα. Then we have
u(A) Xα∈j(A)
Clearly, if Ad(u∗)(Pµ) =Pk Pνk, then for each i ∈ W 1
Ad(ϕ(u∗))(Piµ) =Xk
Ad(u∗)(PA) = X(A,j)∈Γ1
Piνk.
Pjα.
n we have
(5)
(6)
Combining (5) with (6) and proceeding by induction on k, we see that for any A ∈ Γ0
u
and a non-negative integer k we have
Ad(u∗
k)(PA) = X(A,ν)∈Γk
u(A) Xα∈ν(A)
Pνα +
k−1Xm=1 X(A,µ)∈Γm
u (A)
µ(A)=∅
Pµ.
(7)
Now, we are ready to prove a theorem which gives an algorithmic (finite) procedure
for determining if an endomorphism λu, u ∈ Sn, restricts to an automorphism of the
diagonal Dn.
Theorem 3.7 Let u ∈ Sn and let Γu be the directed graph corresponding to u. Then
λuDn is an automorphism of Dn if and only if graph Γu does not contain any closed
(directed) paths.
9
Proof. Firstly, suppose that there is a closed path
A1
(i1)
−→ A2
(i2)
−→ . . .
(ir−1)
−→ Ar
(ir)
−→ A1
in the graph Γu. We denote ν = (i1, i2, . . . , ir) and νk = νν · · · ν (k-fold composition).
With help of formula (7) we see that
Ad(u∗
kr)(PA1) = Pνk ϕkr(PA1) + Xµ∧νk=∅
Pµ.
(8)
Given any k < k′ there exists a non-zero projection q ∈ Dn such that q ≤ Pνk and
qPνk′ = 0. But then formula (8) implies that q Ad(u∗
k′r)(PA1) =
0. Thus the sequence {Ad(u∗
m)(PA1)} never stabilizes and, consequently, projection PA1
does not belong to λu(Dn), Proposition 3.2.
kr)(PA1) 6= 0 while q Ad(u∗
Conversely, suppose that graph Γu does not contain any closed paths. By virtue of
k)(Pµ)} eventually stabilizes for
Lemma 3.4, it suffices to show that the sequence {Ad(u∗
each µ ∈ W ℓ′
n with ℓ′ = ℓ′(J ). To this end, consider the following three cases.
Firstly, we consider the case of Pα, α ∈ J1. Since Γu is a finite graph without closed
paths, there are only finitely many paths and each of them terminates at a sink. By
construction, graph Γu contains exactly one sink, namely vertex ∅. Thus formula (7)
applied to A = {α} shows that for sufficiently large k we have
Ad(u∗
k)(Pα) = X(A,µ)∈Γ∗
u({α})
µ({α})=∅
Pµ,
and thus the sequence {Ad(u∗
k)(Pα)} eventually stabilizes.
Secondly, we consider a word µ such that there exists an α ∈ J1 with µ ≺ α. Then
In this case, the
Pµ = P Pα′, where the sum is over all such α′ ∈ J1 that µ ≺ α′.
k)(Pµ)} stabilizes by the preceding argument.
sequence {Ad(u∗
Thirdly, we must consider the case with µ a word of length at most ℓ′ for which there
exists an α ∈ J1 such that α ≺ µ. Then write µ = αν. Let ({α}, η) be the maximal
path beginning at {α} and such that each vertex on the path is a singleton subset of
J1. Let k be the length of this path. Using formula (7), we see that
Ad(u∗
k)(Pµ) = Pηα′ν
(9)
for some α′ ∈ J1. Now, one of the following two cases happens: either {α′} emits an
edge (with label j) to the sink ∅, or {α′} is a splitting vertex. In the former case, we have
Ad(u∗
k+1)(Pµ) = Pηjν, and the question of stabilization of the sequence corresponding
to the word µ reduces to the same question for the sequence corresponding to the word
ν, which is strictly shorter then µ. In the latter case, let {α′} emit an edge (with label
j=1 Piαj νj , for some αj ∈ J1 and
words νj such that each νj is strictly shorter then ν. Taking into account formula (9), we
obtain Ad(u∗
k)(Pµ)}
stabilizes (with µ = αν) reduces to the same question for all µj = αjνj, where νj < ν.
Consequently, the claim follows for all words µ = αν, α ∈ J1, by induction on ν. ✷
i) to a vertex A. Then we have Ad(u∗)(Pα′) = Pβ ′ =Pm
j=1 Pηiαj νj . Thus, the question if the sequence {Ad(u∗
k+1)(Pµ) = Pm
10
Remark 3.8 We note that for certain special classes of unitaries u ∈ Sn, different
criteria for λuDn ∈ Aut(Dn) were given earlier in [7].
4 The invertibility
In this section, we consider the problem when λu, u ∈ Sn, is an automorphism of On.
Recall that E : On → Fn is the gauge invariant conditional expectation, and for a
β ∈ W k
obtained from β by removing its first
letter.
n the symbol β denotes the word in W k−1
n
Lemma 4.1 If u ∈ Sn is arbitrary then there exists a v ∈ Sn such that E(w) 6= 0 for
w = vuϕ(v∗).
Proof. Let u = P(α,β)∈J SαS∗
P(α,β)∈J vSαS∗
β and suppose that E(u) = 0. If v ∈ Sn then vuϕ(v∗) =
v∗S∗
β1. Thus, it suffices to find a v ∈ Sn such that for certain (α, β) ∈ J
β
v∗ ∈ O(1)
we have vSαS∗
n . Since E(u) = 0, there exists (α, β) ∈ J with α > β. Now,
β
one of the following two cases takes place: either Pα is orthogonal to P β or β ≺ α and
β 6= α. In the former case, put v = S2
β +(other terms). In the latter, we have
α = βµ. Take any ν 6= µ with ν = µ and put v = S1S∗
+ (other terms).
βµ
Then w = vuϕ(v∗) has the required form.
α + S2S∗
2 S∗
βν
+ Sµ
1S∗
✷
Let u = P(α,β)∈J SαS∗
β and ℓ = ℓ(J ). Assume Dℓ
n ⊆ λu(On). Then for each
(α, β) ∈ J and j = 1, . . . , n the element SαS∗
βSj = Pαλu(Sj) belongs to λu(On). Denote
by Zu the collection of all finite products of these elements SαS∗
and their adjoints. The
β
linear span of Zu is dense in λu(On). Also, we denote by hZui the collection of all sums
of elements from Zu.
Lemma 4.2 Let u = P(α,β)∈J SαS∗
β. Denote ℓ = ℓ(J ) and let k ≥ ℓ be any integer
such that there exists a z ∈ Zu, a word of length 2k − 1, with z ∈ O(1)
n . Assume that
λu(Dn) = Dn and E(u) 6= 0. Then for λu to be an automorphism of On it suffices that
F k
n ⊆ λu(On). If this is the case then each SµS∗
n belongs to hZui.
ν with µ, ν ∈ W k
Proof. At first we note that a word z, as in the statement of this lemma, exists since
E(u) 6= 0 by assumption. Then observe that ϕk(Si) belongs to λu(On) for all i =
1, . . . , n. Hence ϕk(Fn) ⊆ λu(On), and consequently λu(On) contains the entire Fn.
Thus Fn and ϕk(Si) are contained in λu(On) and we conclude that λu(On) = On.
We have hZui = λu(h{SµS∗
ν : µ, ν ∈ Wn}i) ⊆ h{SµS∗
invertible then there exists a unitary w ∈ Sn such that λ−1
ν : µ, ν ∈ Wn}i. Now, if λu is
u = λw. Thus we have
λ−1
u (h{SµS∗
ν : µ, ν ∈ Wn}i) ⊆ h{SµS∗
ν : µ, ν ∈ Wn}i
and, consequently, hZui = h{SµS∗
ν : µ, ν ∈ Wn}i.
✷
11
For a while, we restrict our attention to unitaries u = P(α,β)∈J SαS∗
β such that
α − β ∈ {0, ±1} for all (α, β) ∈ J . (We note that endomorphisms corresponding to
such unitaries were studied earlier in [5].)
Given a unitary u as above, we may always find its presentation such that the lengths
of all α coincide. Let k be this common length. Then the collection of all α entering
the presentation of u is equal to W k
n . Now, we define a new directed graph ∆u, as
follows. The set of vertices of ∆u is just W k
n . We put an edge from α1 to α2 whenever
Pα2 ≤ P β1. In view of our assumptions on u, the difference α − β is 0, 1 or 2. We call
this difference the degree of vertex α and of each edge emitted by α. If d is the degree
of α then the vertex α emits exactly nd edges, which end at distinct vertices. With each
edge of degree d > 0 from α1 to α2, we associate a label, which is the terminal subword
of length d of α2. Edges of degree 0 carry empty labels. We extend so defined labels
from edges to finite directed paths on ∆u by concatenation. Also, we define the degree
of a path on ∆u as the sum of the degrees of its edges. We denote the label of a path x
by L(x) and its degree by deg(x).
Now, let ∆∗
u × ∆∗
u be the set of all finite directed paths. In what follows, we consider pairs
(x, y) in ∆∗
u such that x and y end at the same vertex. Let x = x′e and y = y′f ,
where e from α1 to α and f from α2 to α are the last edges of x and y, respectively.
Since e and f end at the same vertex, P β1P β2 6= 0 and thus either β1 ≺ β2 or β2 ≺ β1.
Let µ be the word of length β1 − β2 such that β1 = β2µ or β2 = β1µ. We say that
the pair (x, y) is balanced if the following condition holds: L(x′) = L(y′)µ if β1 = β2µ,
and L(x′)µ = L(y′) if β2 = β1µ. Then we define the total label of (x, y) as L(x′) in
the former case, and L(y′) in the latter. Now, we define a subset Ωu of the Cartesian
product ∆∗
u, as follows. A pair (x, y) belongs to Ωu if and only if:
u × ∆∗
(i) The paths x and y end at the same vertex, but they begin at distinct vertices.
(ii) The paths x and y have identical degrees.
(iii) The pair (x, y) is balanced.
The importance of the set Ωu for our purposes comes from the following Proposition
4.3. Unfortunately, it is not clear to us at the present moment if its hypothesis may be
algorithmically verified in all cases (i.e. for all applicable unitaries u ∈ Sn). However, in
many concrete situations this can be done fairly easily, but preferably with the help of a
computer. Thus, combined with Theorem 3.7, Lemma 4.1 and Lemma 4.2, Proposition
4.3 gives a criterion for deciding invertibility of endomorphism λu.
Proposition 4.3 Let u =P(α,β)∈J SαS∗
the corresponding graph. We assume that λu(Dn) = Dn. Then the following hold.
β be such that α − β ∈ {0, ±1} and let ∆u be
1. Let (x, y) ∈ Ωu have the total label γ and let the paths x, y begin at α and α′,
respectively. Then Zu contains SαPγS∗
α′.
2. If α, α′ ∈ W k
α′ belongs to hZui if and only if there exists a finite collection
(x1, y1), . . . , (xm, ym) in ∆u with the total labels γ1, . . . , γm, respectively, and with
n then SαS∗
12
all xj beginning at α and all yj beginning at α′, such that
1 =
mXj=1
Pγj .
Proof. Ad 1. Let (α, α1, . . . , αm) be the consecutive vertices through which the path x
passes, and likewise let (α′, α′
r) be such vertices for y. Then our definition of Ωu
ensures that
1, . . . , α′
SαPγS∗
α′ = SαS∗
βSα1S∗
β1
· · · SαmS∗
βm
(Sα′S∗
β ′Sα′
1S∗
β ′
1
· · · Sα′
r S∗
β ′
r
)∗,
and thus SαPγS∗
α′ ∈ Zu.
α′ = Pm
Ad 2. Suppose that SαS∗
α′ ∈ hZui, and let SαS∗
νj , with each Sµj S∗
νj
in Zu. Since there are no cancellations among words, each Sµj S∗
νj must be of the form
SαPγj S∗
α′ for some γj ∈ Wn. Now, it is not difficult to verify that an element of Zu has
this form if and only if there exists a pair (xj, yj) in Ωu with the total label γj and such
that xj and yj begin at α and α′, respectively.
j=1 Sµj S∗
The reverse implication is an immediate consequence of part 1 of this proposition.
✷
We end this section with some examples of invertible endomorphisms λu, u ∈ Sn \Pn.
Example 4.4 Let µ, ν be two words such that ν = j1 · · · jr µ with jk ∈ W 1
{µ1, ν1} for all k = 1, . . . , r. Let
n and jk 6∈
u = SνS∗
µ + SµS∗
ν + 1 − Pν − Pµ.
Suppose that λu(Dn) = Dn. We claim that then λu is automatically invertible.
deed, it suffices to check that SνS∗
µSµS∗
Now, SνS∗
µ = Pνλu(Sµ1) and SνS∗
SνS∗
invertible, as claimed.
In-
ν SνS∗
µ.
µ =
ν ∈ λu(On). Consequently, λu is
µ ∈ λu(On). But we have SνS∗
µ = λu(S∗
ν Sj1 · · · Sjr = Pνλu(Sj1) · · · λu(Sjr) and hence SµS∗
ν1)Pµ are both in λu(On). Also, SνS∗
µ = SνS∗
Example 4.5 Let α1, α2, α3 be such that {Pαj } are mutually orthogonal and each αj
begins with the same letter i. Furthermore, suppose that αj = γjµ for some γj which
do not contain the letter i. Let
u = Sα1S∗
α2 + Sα2S∗
α3 + Sα3S∗
α1 + 1 − Pα1 − Pα2 − Pα3.
If λu(Dn) = Dn then automatically λu ∈ Aut(On). Indeed, we have Sk = λu(Sk) for all
k 6= i and thus Sα1S∗
i )Pα2 belongs to λu(On). Similarly, Sα2S∗
α3
and Sα3S∗
α1 are in λu(On) as well. Thus u ∈ λu(On) and λu is invertible. A concrete
example in O2 is obtained by putting
α2 = Pα1λ(Si)Sγ2S∗
γ3λ(S∗
w = S11S∗
121 + S121S∗
1221 + S1221S∗
11 + P1222 + P2,
and then indeed λw is an automorphism of O2.
13
5 The outer Weyl group
In this section, we consider the question whether an endomorphism corresponding to
a unitary in Sn may or may not be equivalent (via an inner automorphism) to one
corresponding to a unitary in the core UHF-subalgebra Fn.
Proposition 5.1 There exist unitaries u ∈ Sn such that λu 6∈ Aut(On)λ(Fn).
Proof. At first we observe that if w ∈ U(Fn) and Q 6= 0 is a projection in On then
the space λw(Dn)Q is infinite dimensional. Indeed, since E(Q) is a non-zero, positive
element of Fn, there is a non-zero projection q ∈ Fn and a scalar t > 0 such that
tq ≤ E(Q). There exists a sequence of indices jk ∈ W 1
n are defined
recursively as α1 = j1, αk+1 = (αk, jk+1) then λw(Pαk)q 6= 0 for all k. The sequence
{qλw(Pαk)q} never stabilizes. Indeed, if qλw(Pαk+m)q = qλw(Pαk)q for all m then
n such that if αk ∈ W k
0 6= τ (qλw(Pαk )q) = τ (qλw(Pαk+m)q) = τ (λw(Pαk+m)qλw(Pαk+m)) ≤ τ (Pαk+m) −→
m→∞
0,
a contradiction. The inequality above holds since w being in U(Fn) the corresponding
endomorphism λw is τ -preserving. Thus, there is a strictly decreasing, infinite sequence
of projections f1 > f2 > . . . in λw(Dn) such that qfkq > qfk+1q for all k. Thus
(fk − fk+1)q 6= 0 for all k, and hence
0 6= (fk − fk+1)tq(fk − fk+1) ≤ (fk − fk+1)E(Q)(fk − fk+1).
Thus (fk−fk+1)Q 6= 0 and, consequently, {(fk−fk+1)Q} is an infinite sequence of linearly
independent elements of λw(Dn)Q, since these are non-zero operators with mutually
orthogonal ranges.
Now, the same conclusion as above holds if λw is replaced by ψλw for some auto-
morphism ψ ∈ Aut(On), since the dimension of (ψλw)(Dn)Q is the same as that of
λw(Dn)ψ−1(Q). Thus, the conclusion of the proposition follows from Example 3.3 (a),
where a unitary u ∈ S2 is exhibited such that λu(D2)P2 is one-dimensional.
✷
Of course, the method of Proposition 5.1 cannot give any information about auto-
morphisms. We treat the automorphism case in Theorem 5.2, below. To the best of
our knowledge, the automorphism entering its proof is the first known example of an
automorphism of On not inner related to an automorphism induced by a unitary from
the core UHF subalgebra Fn.
Theorem 5.2 There exist automorphisms λu, u ∈ Sn, of On such that for all w ∈
U(On) and v ∈ U(Fn) we have λu 6= Ad(w)λv.
Proof. At first we consider the following self-adjoint element of S2 (c.f. Example 4.4):
u = S11S∗
121 + S121S∗
11 + P122 + P2.
(10)
One easily checks that
λu(S1) = S1(S1S∗
21 + S21S∗
1 + P22) and λu(S2) = S2.
14
This yields λ2
u = id. Suppose by contradiction that there exist w ∈ U(O2) and v ∈ U(F2)
such that λu = Ad(w)λv. Then we have D2 = λu(D2) = Ad(w)λv(D2) and λv(D2) ⊆ F2,
thus w∗D2w ⊆ F2. Hence for all d ∈ D2 and z ∈ U(1) we have γz(w∗dw) = w∗dw.
Therefore γz(w)w∗ ∈ D′
2 ∩ O2 = D2. Thus for each z ∈ U(1) there exists a unitary
dz ∈ U(D2) such that γz(w) = dzw. Now we calculate
γz(u) = γz(wvϕ(w∗)) = dzwvϕ(w∗)ϕ(d∗
z) = dzuϕ(d∗
z).
This means that if u =P SαS∗
β then
zα−βSαS∗
βdz = dzSαS∗
β
for all (α, β) and z ∈ U(1). Taking α = β = 2 this yields S2dz = dzS2. By [13], dz is a
scalar, and this in turn implies that γz(u) = u for all z, a contradiction.
Now, if n ≥ 2 is arbitrary, then we consider u = S11S∗
121 + S121S∗
and the same argument as above applies.
11 + P122 + 1 − P1,
✷
As immediate consequences of Theorem 5.2, we obtain the following two corollaries.
Corollary 5.3 The restricted outer Weyl group of On is a proper subgroup of the outer
Weyl group of On.
As shown in [2], the restricted outer Weyl group of On is residually finite and nona-
menable. Thus the outer Weyl group is nonamenable as well, but we do not know if it
is residually finite.
Corollary 5.4 There exist unital subalgebras A of On isomorphic to the UHF algebra
of type {n∞} such that Fn and A are conjugate inside On (by an automorphism of On)
but not inner conjugate.
6 The action on the shift space
Equality (3) easily implies that for all d ∈ Dn and all k > ℓ′(J ) we have
Ad(u)(ϕk(d)) = X(α,β)∈J
ϕk+α−β(d)Pα.
(11)
Consider a map f : Dn → Dn. We say that f eventually preseves standard projec-
tions if there exists an integer m ∈ N such that for each α ∈ Wn, α ≥ m, the image
If u ∈ Sn then Ad(u) eventually preserves standard
f (Pα) is a standard projection.
projections.
Proposition 6.1 If f ∈ Aut(Dn) then there exists a unitary u ∈ Sn such that f =
Ad(u)Dn if and only if;
(i) f eventually preserves standard projections, and
15
(ii) there exist projections Pi, Qi, i = 1, . . . , r, in Dn and non-negative integers ki, mi,
i = 1, . . . , r, such that Pr
i=1 Pi = 1 =Pr
i=1 Qi and
f ◦ MPi ◦ ϕki = MQi ◦ ϕmi,
i = 1, . . . , r.
Proof. Let f ∈ Aut(Dn) satisfy conditions (i) and (ii) of the proposition. For a given
i ∈ {1, . . . , r}, we note that for any subprojection p of Pi we have f ◦ Mp ◦ ϕki =
Mf (p) ◦ ϕmi. Subdividing Pi into a sum of standard projections and using condition (i),
we can assume in condition (ii) that all projections Pi, Qi are standard, say Pi = Pβi
and Qi = Pαi. Define u =Pr
(Ad(u∗) ◦ f ) ◦ MPβi
i=1 SαiS∗
◦ ϕki+αi+h = MPβi
βi, a unitary element of Sn. Then we have
◦ ϕmi+β+h,
i = 1, . . . , r,
for all sufficiently large h ∈ N. We claim that ki + αi = mi + βi for each i. Indeed, fix
an i and suppose that ki + αi ≥ mi + βi (otherwise consider (Ad(u∗) ◦ f )−1 instead).
Then we have
(Ad(u∗) ◦ f )(ϕki+αi−mi−βi(y)Pβi) = yPβi, ∀y ∈ ϕmi+βi+h(Dn),
for all sufficiently large h ∈ N. Fix such an h and let r ≥ mi + βi + h be such that
(Ad(u∗) ◦ f )(ϕki+αi−mi−βi(Dmi+βi+h
n
)Pβi) ⊆ Dr
nPβi
and Pβi ∈ Dr
n. Then we have
(Ad(u∗) ◦ f )(ϕki+αi−mi−βi(Dr
n)Pβi)
= (Ad(u∗) ◦ f )(ϕki+αi−mi−βi(Dmi+βi+h
⊆ Dr
⊆ Dr
nPβiϕmi+βi+h(Dr−mi−βi−h
nPβi.
n
)Pβi
n
)ϕmi+βi+h(Dr−mi−βi−h
n
)Pβi)
Since Ad(u∗) ◦ f is injective and the dimension of ϕki+αi−mi−βi(Dr
than the dimension of Dr
can only happen when ki + αi − mi − βi = 0. Consequently,
n)Pβi is not smaller
nPβi, it follows that these two dimensions are identical, and this
(Ad(u∗) ◦ f ) ◦ MPβi
◦ ϕh = MPβi
◦ ϕh,
i = 1, . . . , r,
for all sufficiently large h ∈ N. Summing over i we get
(Ad(u∗) ◦ f ) ◦ ϕh = ϕh
The opposite direction is clear. Indeed, let u ∈ Sn be such that u = P(α,β)∈J SαS∗
for all sufficiently large h ∈ N. Therefore Ad(u∗) ◦ f = Ad(w)Dn for some w ∈ Pn,
by [2, Lemma 3.2]. Hence f = Ad(uw)Dn and uw ∈ Sn. This proves one direction.
β.
Then condition (i) holds, as noted just above this proposition. One easily checks that
condition (ii) holds with projections Pβ and Pα instead of Pi and Qi, respectively, and
with β and α instead of ki and mi, respectively.
✷
Given u ∈ Sn and considering the homeomorphism Ad(u)∗ of the spectrum Xn of
Dn, we see that the set of fixed points has a very simple structure, as the following
Proposition 6.2 shows.
16
Proposition 6.2 For u ∈ Sn, the set of fixed points in Xn for the homeomorphism
Ad(u)∗ consists of the union of a clopen set and a finite set. Furthermore, each of the
isolated fixed points is either a local attractor or a local repeller.
Proof. Let u =P(α,β)∈J SαS∗
β. It is clear that Ad(u)∗ admits fixed points in Xn if and
only if there exists (α, β) ∈ J such that either α ≺ β or β ≺ α. Thus we arrive at one
of the following three cases. (1) If α = β then the clopen set {x ∈ Xn : β ≺ x} is fixed
by Ad(u)∗. (2) If α = βµ, µ 6= ∅, then x = βµµ . . . is a fixed point and a local attractor.
(3) If β = αµ, µ 6= ∅, then x = αµµ . . . is a fixed point and a local repeller.
✷
In contrast to Proposition 6.2 above, the set of fixed points in Xn corresponding to
an outer automorphism λu, u ∈ Sn, may have a much more complicated structure, as
the following example demonstrates.
Example 6.3 Let u be the unitary in S2 defined by formula (10). It is not difficult to
verify that the corresponding homeomorphism (λu)∗ of X2 fixes an x ∈ X2 if and only
if x does not contain substrings (11) and (121). These fixed points form a compact,
nowhere dense subset K of X2, in which there are no isolated points. Thus K itself is
homeomorphic to the Cantor set and closed under the action of the one-sided shift ϕ∗.
References
[1] R. Conti, Automorphisms of the UHF algebra that do not extend to the Cuntz
algebra, J. Austral. Math. Soc. 89 (2010), 309 -- 315.
[2] R. Conti, J. H. Hong and W. Szyma´nski, The restricted Weyl group of the Cuntz
algebra and shift endomorphisms, J. reine angew. Math. 667 (2012), 177 -- 191.
[3] R. Conti, J. H. Hong and W. Szyma´nski, Endomorphisms of graph algebras, ac-
cepted to J. Funct. Anal., arXiv:1101.4210.
[4] R. Conti, J. Kimberley and W. Szyma´nski, More localized automorphisms of the
Cuntz algebras, Proc. Edinburgh Math. Soc. 53 (2010), 619 -- 631.
[5] R. Conti, M. Rørdam and W. Szyma´nski, Endomorphisms of On which preserve
the canonical UHF-subalgebra, J. Funct. Anal. 259 (2010), 602 -- 617.
[6] R. Conti and W. Szyma´nski, Labeled trees and localized automorphisms of the Cuntz
algebras, Trans. Amer. Math. Soc. 363 (2011), 5847 -- 5870.
[7] R. Conti and W. Szyma´nski, Automorphisms of the Cuntz algebras, accepted
to the Proceedings of the EU-NCG 4th Annual Meeting in Bucharest (2011),
arXiv:1108.0860.
[8] J. Cuntz, Simple C ∗-algebras generated by isometries, Commun. Math. Phys. 57
(1977), 173 -- 185.
17
[9] J. Cuntz, Automorphisms of certain simple C ∗-algebras, in Quantum fields-algebras-
processes, ed. L. Streit, 187 -- 196, Springer, 1980.
[10] S. Eilenberg, Automata, languages and machines, Academic Press, New York, 1974.
[11] J. H. Hong, A. Skalski and W. Szyma´nski, On invariant MASAs for endomorphisms
of the Cuntz algebras, Indiana Univ. Math. J. 59 (2010), 1873 -- 1892.
[12] K. Matsumoto, Orbit equivalence of topological Markov shifts and Cuntz-Krieger
algebras, Pacific J. Math. 246 (2010), 199 -- 225.
[13] K. Matsumoto and J. Tomiyama, Outer automorphisms on Cuntz algebras, Bull.
London Math. Soc. 25 (1993), 64 -- 66.
[14] V. Nekrashevych, Cuntz-Pimsner algebras of group actions, J. Operator Theory 52
(2004), 223 -- 249.
[15] W. Szyma´nski, On localized automorphisms of the Cuntz algebras which preserve
in 'New Development of Operator Algebras', R.I.M.S.
the diagonal subalgebra,
Kokyuroku 1587 (2008), 109 -- 115.
Roberto Conti
Dipartimenti di Scienze
Universit`a di Chieti-Pescara 'G. D'Annunzio'
Viale Pindaro 42, I -- 65127 Pescara, Italy
present address:
Dipartimento di Scienze di Base e Applicate per l'Ingegneria
Sezione di Matematica
Sapienza Universit`a di Roma
Via A. Scarpa 16
00161 Roma, Italy
E-mail: [email protected]
Jeong Hee Hong
Department of Data Information
Korea Maritime University
Busan 606 -- 791, South Korea
E-mail: [email protected]
Wojciech Szyma´nski
Department of Mathematics and Computer Science
The University of Southern Denmark
Campusvej 55, DK-5230 Odense M, Denmark
E-mail: [email protected]
18
|
1201.0609 | 2 | 1201 | 2012-01-04T15:18:36 | Radial multipliers on reduced free products of operator algebras | [
"math.OA"
] | Let A_i be a family of unital C*-algebras, respectively, of von Neumann algebras and phi: N_0 \to C. We show that if a Hankel matrix related to phi is trace-class, then there exists a unique completely bounded map M_phi on the reduced free product of the A_i, which acts as an radial multiplier. Hereby we generalize a result of Wysocza\'nski for Herz-Schur multipliers on reduced group C*-algebras for free products of groups. | math.OA | math |
RADIAL MULTIPLIERS ON REDUCED FREE
PRODUCTS OF OPERATOR ALGEBRAS
UFFE HAAGERUP AND SÖREN MÖLLER
Abstract. Let Ai be a family of unital C*-algebras, respectively,
of von Neumann algebras and φ : N0 → C. We show that if a
Hankel matrix related to φ is trace-class, then there exists a unique
completely bounded map Mφ on the reduced free product of the
Ai, which acts as an radial multiplier. Hereby we generalize a
result of Wysoczański for Herz-Schur multipliers on reduced group
C*-algebras for free products of groups.
1. Introduction
Let C denote the set of functions φ on the non-negative integers N0
for which the matrix
h = (φ(i + j) − φ(i + j + 1))i,j≥0
is of trace class. Let G = ∗i∈IGi be the free product of discrete groups
(Gi)i∈I. In [12] , J. Wysoczański proved that if φ ∈ C and φ : G → C
is defined by φ(e) = φ(0) and φ(g1 . . . gn) = φ(n) for all n > 0 when
gj ∈ Gij \ {e} and i1 6= i2 6= · · · 6= in, then φ is a Herz-Schur multiplier
on G and k φkHS ≤ kφkC , where k · kC is the norm on C defined in
(2.2) below. In particular, there is a unique completely bounded map
Mφ : C ∗
r (G) → C ∗
r (G) such that Mφ(1) = φ(0)1 and
Mφ(λ(g1 . . . gn)) = φ(n)λ(g1 . . . gn)
when gj ∈ Gij \ {e} and i1 6= i2 6= · · · 6= in as above, and kMφkcb ≤
kφkC . Furthermore J. Wysoczański proved that kMφkcb = kφkC in the
cases when I = ∞ and Gi = ∞ for all i ∈ I. In the special case of
φs(n) = sn for n ≥ 0 and s < 1 it follows that
1 − s
1 − s
kMφskcb ≤ kφskC =
.
In this paper we will show that every function φ from C gives rise
to radial multipliers Mφ on reduced free products of C ∗-algebras and
reduced free products of von Neumann algebras (cf. Theorem 2.2),
satisfying kMφkcb ≤ kφkC . Radial multipliers of general reduced free
products of C ∗-algebras were first considered by È. Ricard and Q. Xu
Date: November 15, 2018.
2010 Mathematics Subject Classification. Primary 46L54; Secondary 46L07.
1
RADIAL MULTIPLIERS ON REDUCED FREE PRODUCTS
2
in [10] and the weaker estimate kMφkcb ≤ φ(0) +P∞
be obtained from [10, Corollary 3.3].
n=1 4nφ(n) can
The main result is proved in Section 5. In Section 6 we discuss a
related set of functions C ′ (cf. Definition 6.3).
It was used by T.
Steenstrup, R. Szwarc and the first author in [6] to characterize radial
multipliers on free groups Fn (2 ≤ n ≤ ∞). Moreover, C. Houdayer
and È. Ricard used it in [7] to characterize multipliers on the free
Araki-Woods factor Γ(HR, , Ut)′′ (cf. Section 6.3).
In Section 7 we obtain an integral representation of functions in the
class C which together with N. Ozawa's result in [9], shows that for
every hyperbolic group Γ, and every φ ∈ C , the function
φ(x) = φ(d(x, e))
is a completely bounded Fourier multiplier on Γ (cf. Remark 7.6).
2. The main results
We start by defining the class C , crucial in what follows.
Definition 2.1. Let C denote the set of functions φ : N0 → C for
which the Hankel matrix h = (φ(i + j) − φ(i + j + 1))i,j≥0 is of trace-
class.
If φ ∈ C , then k = (φ(i + j + 1) − φ(i + j + 2))i,j≥0 is of trace-class,
as well. Furthermore, we have
∞
φ(n) − φ(n + 1) ≤ khk1 + kkk1 < ∞,
(2.1)
Xn=0
where kxk1 = T r(x) is the trace-class norm for x ∈ B(l2(N0)). This
implies that c = limn→∞ φ(n) exists. For φ ∈ C set
kφkC = khk1 + kkk1 + c.
(2.2)
The main result of this paper is the following generalization of Wysocza-
ński's result:
Theorem 2.2.
(1) Let A = ∗i∈I(Ai, ωi) be the reduced free product of unital C*-
algebras (Ai)i∈I with respect to states (ωi)i∈I for which the GNS-
representation πωi is faithful, for all i ∈ I.
If φ ∈ C , then there is a unique linear completely bounded
map Mφ : A → A such that Mφ(1) = φ(0)1 and Mφ(a1a2 . . . an) =
φ(n)a1a2 . . . an whenever aj ∈ Aij = ker(ωij ) and i1 6= i2 6=
· · · 6= in. Moreover kMφkcb ≤ kφkC .
(2) Let (M , ω) = ¯∗i∈I(Mi, ωi) be the w*-reduced free product of von
Neumann algebras (Mi)i∈I with respect to normal states (ωi)i∈I
for which the GNS-representation πωi is faithful, for all i ∈ I.
RADIAL MULTIPLIERS ON REDUCED FREE PRODUCTS
3
If φ ∈ C , then there is a unique linear completely bounded
normal map Mφ : M → M such that Mφ(1) = φ(0)1 and
Mφ(a1a2 . . . an) = φ(n)a1a2 . . . an whenever aj ∈ Mij = ker(ωij )
and i1 6= i2 6= · · · 6= in. Moreover kMφkcb ≤ kφkC .
Remark 2.3. By J. Wysoczański's result in [12], the norm estimates
in Theorem 2.2 are best possible, as equality is attained if I = ∞ and
r (Gi), τi) for a family (Gi)i∈I of infinite discrete groups,
(Ai, ωi) = (C ∗
where τi is the canonical trace on C ∗
r (Gi) coming from the left regular
representation. It would be interesting to know for which (Ai, ωi)i∈I,
the equality kMφkcb = kφkC holds for all φ ∈ C .
We start by proving that the operator Mφ is unique, if it exists.
Lemma 2.4 (Uniqueness in Theorem 2.2). The map Mφ is uniquely
determined by the conditions in the hypothesis of Theorem 2.2.
Proof. The algebra C1 +(cid:16)Pi∈I
dense in A and, respectively, C1+(cid:16)Pi∈I
Ai(cid:17) +(cid:16)Pi16=i2
Mi(cid:17)+(cid:16)Pi16=i2
is σ-weakly dense in M . As Mφ is bounded, it is then uniquely defined
on all of A , respectively, on all of M .
(cid:3)
Ai1
Ai2(cid:17) + . . . is norm
Mi2(cid:17)+. . .
Mi1
Now to prove Theorem 2.2 we start by showing that it is enough
to prove the result for the special case of the algebras M = B(Hi, Ωi)
equipped with ωi, the vector state given by Ωi, as this will implie the
result for general C*- and von Neumann-algebras.
Proposition 2.5. If Theorem 2.2 part (2) holds for (Mi, ωi) = (B(Hi), ωΩi)
for Hilbert spaces (Hi, Ωi) and associated vector states ωi then Theorem
2.2 holds in general.
Proof. Assume Theorem 2.2 holds for (B(Hi), ωΩi) for arbitrary Hi and
Ωi. Now let A = ∗i∈I(Ai, ωi), respectively, (M , ω) = ¯∗i∈I(Mi, ωi). Let
(Hi, Ωi) = (Hωi, ξωi) be the Hilbert space and state coming from the
GNS-representation of Ai, respectively, Mi, and let (H, Ω) = ∗i∈I(Hi, Ωi)
be their Hilbert space free product.
Now by [11, Definition 1.5.1], Ai, respectively, Mi can be realized as
subalgebras of B(H) by the action defined as follows. If a ∈ Ai, γ1 ⊗
· · · ⊗ γn ∈ H with γj ∈ Hj := Ω⊥
j then
a(γ1 ⊗ · · · ⊗ γn) = a(Ωi) ⊗ γ1 ⊗ · · · ⊗ γn
if i 6= j, and otherwise
a(γ1 ⊗ · · · ⊗ γn) = (a(γ1) − ha(γ1), ΩiiΩi) ⊗ γ2 ⊗ · · · ⊗ γn
+ ha(γ1), Ωiiγ2 ⊗ · · · ⊗ γn.
Hence MφA and MφM can be obtained by restricting Mφ to the re-
spective subalgebra of B(H) and we then have kMφA kcb ≤ kMφkcb ≤
RADIAL MULTIPLIERS ON REDUCED FREE PRODUCTS
4
kφkC , respectively, kMφM kcb ≤ kMφkcb ≤ kφkC which gives the de-
sired general result.
(cid:3)
We will prove the special case considered in Proposition 2.5 in the
following sections.
3. Preliminaries
We start by introducing some notation.
Let (H, Ω) = ∗i∈I(Hi, Ωi). Also denote Hi = Ω⊥
i , for i ∈ I. Then by
n=0 H(n),
Hi1 ⊗ · · · ⊗ Hin, for n > 0, and H(0) = CΩ.
We will denote the projection from H to H(n) by Pn ∈ B(H), and let
the definition of the Hilbert space free product we have H =L∞
where H(n) := Li16=···6=in
Qn :=P∞
Now choose orthonormal bases Γi for Hi, then Γi = Γi ∪ {Ωi} are
bases for Hi. Put Λ(0) = {Ω} and Λ(n) = {γ1 ⊗ · · · ⊗ γn : γj ∈ Γij , i1 6=
· · · 6= in} for all n ≥ 1. Then Λ(n) is an orthonormal basis for H(n),
n=0 Λ(n) is an orthonormal basis for H. Note
k=n Pk.
for all n ≥ 0 and Λ = S∞
that Λ(1) =Si∈I
Γi considered as a subset of H.
Now we can define the basic operators in B(H). Let γ ∈ Λ(1). Let
Lγ, Rγ ∈ B(H) be the operators for which LγΩ = RγΩ = γ, and for
χ = χ1 ⊗ · · · ⊗ χn ∈ Λ(n) where χj ∈ Γij and γ ∈ Γi we have
respectively,
Lγ(χ) =(cid:26) γ ⊗ χ
0
Rγ(χ) =(cid:26) χ ⊗ γ
0
if i 6= i1
if i = i1
if i 6= in
if i = in.
Note that Lγ and Rγ are well-defined partial isometries in B(H).
Moreover for all γ ∈ Λ(1) and n ≥ 0 we have LγH(n) ⊆ H(n + 1),
respectively, RγH(n) ⊆ H(n + 1). For γ = γ1 ⊗ · · · ⊗ γn ∈ Λ(n) denote
Lγ = Lγ1Lγ2 . . . Lγn, respectively, Rγ = RγnRγn−1 . . . Rγ1, where we set
LΩ = RΩ = 1.
Lemma 3.1. Let B(Hi)= {a ∈ B(Hi) : haΩi, Ωii = 0}. Then the set
spann{Lγ : γ ∈ Γi} ∪ {L∗
in B(Hi)considered as a subset of B(H).
γ : γ ∈ Γi} ∪ {LγL∗
δ : γ, δ ∈ Γi}o is σ-weakly dense
Proof. Let (eγ,δ)γ,δ∈Γi be the matrix units of B(Hi) corresponding to
the basis Γi. Then span{eγ,δ : (γ, δ) 6= (Ωi, Ωi)} is σ-weakly dense in
B(Hi). Moreover, by the natural embedding of B(Hi) in B(H) one
gets for γ, δ ∈ Γi that Lγ = eγ,Ωi, L∗
δ = eγ,δ,
which proves the lemma.
(cid:3)
γ = eΩi,γ, and hence LγL∗
RADIAL MULTIPLIERS ON REDUCED FREE PRODUCTS
5
Definition 3.2. Let a = (ai)i≥0 ∈ l∞(N0). Denote by Da the operator
which is defined by Da(ξ) = anξ for ξ ∈ Λ(n), n ≥ 1, respectively,
Da(Ω) = a0Ω and by linearity is extended to all of H.
Note that Da = P∞
n=0 anPn and that Da is bounded with kDak =
kak∞. Let S denote the standard shift on l∞(N0), i.e., for a = (ai)i≥0 ∈
l∞(N0), let S(a0, a1, a2, . . . ) = (0, a0, a1, a2, . . . ). To ease notation, con-
sider seperately the following two cases, which together contain all pos-
sible situations.
Definition 3.3. For ξ ∈ Λ(k) and η ∈ Λ(l), k, l ≥ 0 we say that we
are in
• Case 1 if ξ = Ω or η = Ω or k, l ≥ 1 and ξ = ξ1 ⊗ · · · ⊗ ξk and
η = η1 ⊗ · · · ⊗ ηl where ξk ∈ Γi, ηl ∈ Γj and i 6= j, i, j ∈ I,
respectively,
• Case 2 if k, l ≥ 1 and ξ = ξ1 ⊗ · · · ⊗ ξk and η = η1 ⊗ · · · ⊗ ηl
where ξk, ηl ∈ Γi for some i ∈ I.
4. Technical lemmas
Definition 4.1. For x, y ∈ l2(N0) and a ∈ B(H) set
Φ(1)
x,y(a) :=
D(S ∗)nxaD∗
(S ∗)ny +
∞
Xn=0
DSnxρn(a)D∗
Sny,
∞
Xn=1
respectively,
Φ(2)
x,y(a) :=
Xn=0
∞
∞
D(S ∗)nxaD∗
(S ∗)ny +
DSnxρn−1(ǫ(a))D∗
Sny
Xn=1
where ρ(a) := Pγ∈Λ(1) RγaR∗
projection onto span{ξ ∈ Λ(n) : n ≥ 1, ξ = γ1 ⊗ · · · ⊗ γn, γn ∈ Γi} for
i ∈ I.
γ and ǫ(a) := Pi∈I qiaqi and qi is the
Lemma 4.2. Let k, l ≥ 0. Then for every ξ ∈ Λ(k) and η ∈ Λ(l) we
have for all n ≥ 0, ρn(LξL∗
η) in
Case 1, and, respectively, ǫ(LξL∗
ηQl+n and ǫ(LξL∗
η) = ρ(LξL∗
η) = LξL∗
η in Case 2.
η) = LξL∗
Proof. For the first statement observe that
ρn(LξL∗
RζLξL∗
ηR∗
ζ
η) = Xζ∈Λ(n)
= Lξ
Xζ∈Λ(n)
= LξQnL∗
η
= LξL∗
ηQl+n.
RζR∗
ζ
L∗
η
RADIAL MULTIPLIERS ON REDUCED FREE PRODUCTS
6
For the second statement, let χ ∈ Λ(m). If m > l then
ǫ(LξL∗
qiLξL∗
ηqi(χ)
η)(χ) =Xi∈I
= LξXi∈I
qiqiL∗
η(χ)
η(χ)
= LξQ(1)L∗
= LξL∗
= LξL∗
ηQ(l + 1)(χ)
η(χ)
While if m = l, χ = η and ηl ∈ Γj for some j ∈ I we have
ǫ(LξL∗
ηqi(η) = qjLξL∗
ηqj(η) = qjLξL∗
qiLξL∗
η(η) = qj(ξ),
η)(η) =Xi∈I
and this is equal to 0 in Case 1 (i.e., ξk /∈ Γj), respectively, equal to ξ
in Case 2 (i.e., ξk ∈ Γj).
Note that both sides vanish if m = l and χ 6= η, or m < l.
To calculate the completely bounded norm of Φ(·)
x,y we use the follow-
(cid:3)
ing result from [3].
Theorem 4.3. [3, Theorem 1.3] If kPi uiu∗
some ui, vi ∈ B(H), then Φ(a) =Pi uiavi defines a normal completely
bounded operator on B(H) and kΦkcb ≤ kPi∈I uiu∗
i vik.
Using this theorem we ontain the following cb-norm estimates.
i kkPi∈I v∗
i k, kPi v∗
i vik < ∞ for
Lemma 4.4. For x, y ∈ l2(N0) we have kΦ(1)
tively, kΦ(2)
x,ykcb ≤ kxk2kyk2.
Proof. Let χ ∈ Λ(m). Then
x,ykcb ≤ kxk2kyk2, respec-
∞
Xn=0
respectively,
∞
D(S ∗)nxD∗
(S ∗)nx(χ) =
x(m + n)D(S ∗)nx(χ)
∞
∞
=
Xn=0
Xn=0
= ∞
Xn=m
x(m + n)2(χ)
x(n)2! (χ),
(DSnxRζ)(DSnxRζ)∗(χ) =
Xn=1 Xζ∈Λ(n)
=
∞
m
Xn=1 Xζ∈Λ(n)
Xn=1
DSnxRζR∗
ζ D∗
Snx(χ)
DSnxD∗
Snx(χ) =
x(n)2(χ).
m−1
Xn=0
RADIAL MULTIPLIERS ON REDUCED FREE PRODUCTS
7
Here the second equality holds since RζR∗
or if n > m.
ζ (χ) = 0 if ζ 6= χm−n+1 . . . χm
On the other hand,
∞
(DSncRζqi)(DSncRζqi)∗(χ)
∞
=
Xn=1 Xζ∈Λ(n−1)Xi∈I
Xn=1 Xζ∈Λ(n−1)Xi∈I
Xn=1 Xζ∈Λ(n−1)
Xn=1
DSncD∗
=
=
m
m
DSncRζqiR∗
ζD∗
Snc(χ)
DSncRζR∗
ζ D∗
Snc(χ)
Snc(χ) =
c(n)2(χ).
m−1
Xn=0
The second equality holds since χm−n+1 ∈ Γi (the rightmost element
ζ (χ)) for a unique i ∈ I if n − 1 < m, and there is no such
of R∗
i ∈ I if n − 1 ≥ m. The third equality holds as RζR∗
ζ(χ) = 0 for
ζ 6= χm−n+2 . . . χm. Hence
D(S ∗)nxD∗
(S ∗)nx +
∞
Xn=0
= kxk2
2 (χ),
∞
Xn=1 Xζ∈Λ(n)
(DSnxRζ)(DSnxRζ)∗
(χ)
respectively,
D(S ∗)nxD∗
(S ∗)nx +
∞
Xn=1
= kxk2
2 (χ).
∞
Xn=1 Xζ∈Λ(n−1)Xi∈I
(DSnxRζqi)(DSnxRζqi)∗
(χ)
Using these calculations for x, y ∈∈ l2(N0) and applying Theorem 4.3
to get the desired result.
(cid:3)
Lemma 4.5. Let k, l ≥ 0. If ξ ∈ Λ(k) and η ∈ Λ(l) then
Φ(1)
x,y(LξL∗
η) = ∞
Xt=0
x(k + t)y(l + t)! LξL∗
η
and, respectively,
Φ(2)
x,y(LξL∗
η) =(cid:26) P∞
P∞
t=0 x(k + t)y(l + t)LξL∗
t=0 x(k + t − 1)y(l + t − 1)LξL∗
η
η
in Case 1
in Case 2.
Proof. We prove this by showing that both sides act similarly on all
simple tensors in H.
RADIAL MULTIPLIERS ON REDUCED FREE PRODUCTS
8
Indeed, let m ≥ 0 and χ ∈ Λ(m) and let n ≥ 0. If χ = η ⊗ ζ, where
ζ ∈ Λ(m − l) for some l ≥ 0 we have for the common type of terms in
Φ(1)
x,y and Φ(2)
x,y that
D(S ∗)nxLξL∗
ηD∗
(S ∗)ny(χ) = y(m + n)D(S ∗)nxLξL∗
η(χ)
(4.1)
= y(m + n)D(S ∗)nx(ξ ⊗ ζ)
= x(k + m − l + n)y(m + n)LξL∗
η(χ).
Otherwise, if there is no ζ such that χ = η ⊗ ζ, then both sides vanish,
wherein we have used the convention that x(p) = 0 for p < 0.
For the other type of terms in Φ(1)
x,y, we get
DSnxρn(LξL∗
η)D∗
Sny(χ) = y(m − n)DSnxρn(LξL∗
η)(χ)
(4.2)
= y(m − n)DSnxLξL∗
ηQl+n(χ)
= y(m − n)DSnxQl+n(ξ ⊗ ζ)
= x(k + m − l − n)y(m − n)LξL∗
ηQl+n(χ)
where in the second equality we use Lemma 4.2 and the fact that both
sides vanish if n > m − l.
We now estimate the other type of terms in Φ(2)
x,y.
In Case 1 we
similarly get
DSnxρn−1(ǫ(LξL∗
η))D∗
Sny(χ) = DSnxρn(LξL∗
Sny(χ)
= x(k + m − l − n)y(m − n)LξL∗
η)D∗
ηQl+n(χ)
with both sides vanishing for n > m − l.
In Case 2 we get by Lemma 4.2
DSnxρn−1(ǫ(LξL∗
η))D∗
Sny(χ) = DSnxρn−1(LξL∗
Sny(χ)
= x(k + m − l − n)y(m − n)LξL∗
η)D∗
ηQl+n−1(χ)
(4.3)
(4.4)
with both sides vanishing for n > m − l + 1.
Combining (4.1) and (4.2) we get
Φ(1)
x,y(LξL∗
η)(χ) =
=
∞
∞
Xn=0
Xn=0
D(S ∗)nxLξL∗
ηD∗
(S ∗)ny +
∞
Xn=1
DSnxρn(LξL∗
η)D∗
Sny
x(k + m − l + n)y(m + n)LξL∗
η(χ)
m−l
+
Xn=1
= ∞
Xn=l−m
x(k + m − l − n)y(m − n)LξL∗
η(χ)
x(k + m − l + n)y(m + n)! LξL∗
η(χ)
RADIAL MULTIPLIERS ON REDUCED FREE PRODUCTS
9
= ∞
Xt=0
x(k + t)y(l + t))! LξL∗
η(χ)
as desired.
Similarly in Case 1, combining (4.1) and (4.3) we get
Φ(2)
x,y(LξL∗
η)(χ) = ∞
Xt=0
x(k + t)y(l + t)! LξL∗
η(χ).
While in Case 2, combining (4.1) and (4.4) we get
Φ(2)
x,y(LξL∗
η)(χ) =
∞
Xn=0
x(k + m − l + n)y(m + n)LξL∗
η(χ)
m−l+1
+
= ∞
Xt=0
This completes the proof.
x(k + m − l − n)y(m − n)LξL∗
Xn=1
x(k + t − 1)y(l + t − 1)! LξL∗
η(χ).
η(χ)
(cid:3)
We now establish some technical results conserning maps φ ∈ C .
Put ψ1(n) = P∞
Lemma 4.6. Let φ ∈ C and let h, k and c be as in Definition 2.1.
i=0(φ(n + 2i) − φ(n + 2i + 1)) and ψ2(n) = ψ1(n + 1),
for n ≥ 0. Then φ(n) = ψ1(n) + ψ2(n) + c for n ≥ 0 and the entries
hi,j and ki,j of h and k are given by hi,j = ψ1(i + j) − ψ1(i + j + 2),
respectively, ki,j = ψ2(i + j) − ψ2(i + j + 2), for i, j ≥ 0.
Proof. By (2.1) we have
∞
lim
n→∞
ψ1(n) ≤ lim
n→∞
φ(n + 2i) − φ(n + 2i + 1) = 0.
Xi=0
A similar statement holds for ψ2 and therefore limn→∞ ψ1(n) = 0 and
limn→∞ ψ2(n) = 0. Next, let n ≥ 0 be fixed. Then simple computations
give ψ1(n) + ψ2(n) = φ(n) − c, and ψ1(n) − ψ1(n + 2) = φ(n) − φ(n + 1),
respectively, ψ2(n) − ψ2(n + 2) = φ(n + 1) − φ(n + 2). Using these
equations, we get the desired formulas for hi,j, respectively ki,j.
(cid:3)
kkk1. Here we use the notation (u ⊙ v)(t) = ht, viu, for u, v, t ∈ l2(N0).
Remark 4.7. Since h, k are trace-class, it is well-known (cf. [6, p. 13])
i=1 xi ⊙ yi and
that there exist xi, yi, zi, wi ∈ l2(N0) such that h = P∞
P kxik2kyik2 = khk1, respectively, k =P∞
in Remark 4.7 we have ψ1(k + l) = P∞
ψ2(k + l) =P∞
i=1 zi⊙wi andP kzik2kwik2 =
i=1P∞
Lemma 4.8. For ψ1 and ψ2 as in Lemma 4.6, and xi, yi, zi, and wi as
t=0 xi(k + t)yi(l + t) and
t=0 zi(k + t)wi(l + t).
i=1P∞
RADIAL MULTIPLIERS ON REDUCED FREE PRODUCTS
10
Proof. Let k, l ≥ 0, then
∞
ψ1(k + l + 2t) − ψ1(k + l + 2t + 2)
hk+t,l+t
∞
xi(k + t)yi(l + t)
xi(k + t)yi(l + t)
ψ1(k + l) =
=
=
=
∞
∞
Xt=0
Xt=0
Xt=0
Xi=1
∞
∞
Xi=1
Xt=0
where the sums are absolutly convergent, and we use Lemma 4.6 for
the first two equalities. A similar reasoning applies to ψ2.
(cid:3)
5. Proof of the main result
As shown in Section 2, it is enough to prove the following lemma in
order to obtain the result of the main theorem.
Proposition 5.1. Let (H, Ω) = ∗i∈I(Hi, Ωi) be the reduced free product
of Hilbert spaces (Hi)i∈I with unit vector Ωi and let ωi(a) = haΩi, Ωii
for a ∈ B(Hi) where we realize B(Hi) as subalgebras of B(H) via the
standard embedding from [11, Definition 1.5.1]. Then for every φ ∈ C ,
there exists a linear completely bounded normal map Mφ on B(H) such
that Mφ(1) = φ(0)1 and Mφ(a1a2 . . . an) = φ(n)a1a2 . . . an whenever
n ≥ 1, i1, . . . in ∈ I with i1 6= i2 6= · · · 6= in and aj ∈ B(Hij )= ker(ωij ).
Moreover, kMφkcb ≤ kφkC .
The proof of Proposition 5.1 will be divided into a series of lemmas.
Lemma 5.2. Let T : B(H) → B(H) be a bounded linear normal map,
and let φ : N0 → C. The following statements are equivalent.
(a) For all n ≥ 1, i1, . . . in ∈ I with i1 6= i2 6= · · · 6= in and aj ∈
B(Hij ) = ker(ωij ), we have T (1) = φ(0)1 and T (a1a2 . . . an) =
φ(n)a1a2 . . . an.
(b) For all k, l ≥ 0 and ξ ∈ Λ(k), η ∈ Λ(l) we have
T (LξL∗
η) =(cid:26) φ(k + l)LξL∗
φ(k + l − 1)LξL∗
η
η
in Case 1
in Case 2.
Proof. Assume (a) and let k, l ≥ 1, ξ ∈ Λ(k), η ∈ Λ(l). Now by the
definition of Lξ we have LξL∗
η = Lξ1 . . . Lξk L∗
ηl . . . L∗
η1.
If we are in Case 1, there exist i, j ∈ I, i 6= j such that ξk ∈ Γi
and ηl ∈ Γj. Hence all adjacent terms above are from different B(Hi),
hence LξL∗
η is of the form a1 . . . an in (a) with n = k + l.
On the other hand, if we are in Case 2, there exists i ∈ I such that
η is of the form
ηl ∈ B(Hi), hence LξL∗
In this case LξkL∗
ξk, ηl ∈ Γi.
RADIAL MULTIPLIERS ON REDUCED FREE PRODUCTS
11
a1 . . . an in (a) with n = k + l − 1. Applying (a) we get the conclusion
of (b) for k, l ≥ 1. If k = 0 or l = 0, e.g., ξ = Ω or η = Ω the result
follows similarly by using LΩ = 1.
Assume (b). Using Kaplansky's density theorem [8, Theorem 5.3.5]
and the fact that the product is jointly σ-strong continuous on bounded
sets, by Lemma 3.1 it is enough to check that T (1) = φ(0)1 and
T (a1 . . . an) = φ(n)a1 . . . an whenever n ≥ 1 and aj ∈ {Lγγ ∈ Γij } ∪
{L∗
δγ, δ ∈ Γij } where i1 6= i2 6= · · · 6= in.
In
γLδ = 0 when γ ∈ Γij and δ ∈ Γij+1, since ij 6= ij+1.
δ1 for some
particular, L∗
Hence a1 . . . an = 0, unless a1 . . . an = Lγ1 . . . Lγk L∗
δl
γj ∈ Γij , δs ∈ Γrs, ij 6= ij+1, rs 6= rs+1 and i1, . . . , ik, r1, . . . , rl ∈ I.
γLδ = 0 when γ, δ ∈ Λ(1), γ 6= δ.
. . . L∗
γγ ∈ Γij } ∪ {LγL∗
It is easy to check that L∗
If we are in Case 1, we have ik 6= rl. Hence neighboring elements
on the right hand side are from different B(Hi)and thus n = k + l.
∈ B(Hik ), thus
If we are in Case 2, we have ik = rl. Hence Lγk L∗
δl
n = k + l − 1. Now (b) gives the result for k ≥ 1 or l ≥ 1. Moreover
the k = l = 0 case of (b) gives T (1) = φ(0)1.
(cid:3)
Next, we explicitly construct such a map T .
Lemma 5.3. Let φ ∈ C . Define maps
T1 =
∞
Xi=1
Φ(1)
xi,yi
and T2 =
Φ(2)
zi,wi
∞
Xi=1
where Φ(·)
x,y are as in Definition 4.1, ψ1, ψ2 as in Lemma 4.6, and
xi, yi, zi, wi as in Remark 4.7. Then for all k, l ≥ 0, ξ ∈ Λ(k), η ∈ Λ(l)
we have T1(LξL∗
η) = ψ1(k + l)LξL∗
η, respectively,
T2(LξL∗
η) =(cid:26) ψ2(k + l)LξL∗
ψ2(k + l − 2)LξL∗
η
η
in Case 1
in Case 2.
Proof. Let k, l ≥ 0 and ξ ∈ Λ(k), η ∈ Λ(l). Now by Lemma 4.5 and
Lemma 4.8 we have
∞
T1(LξL∗
η) =
Φ(1)
xi,yi(LξL∗
η)
Xi=1
= ∞
Xi=1
∞
Xt=0
xi(k + t)yi(l + t)! (LξL∗
η)
= ψ1(k + l)LξL∗
η.
RADIAL MULTIPLIERS ON REDUCED FREE PRODUCTS
12
Furthermore, in Case 1 we have by Lemma 4.5 and Lemma 4.8
∞
T2(LξL∗
η) =
Φ(2)
zi,wi(LξL∗
η)
Xi=1
= ∞
Xi=1
∞
Xt=0
= ψ2(k + l)LξL∗
η,
zi(k + t)wi(l + t)! (LξL∗
η)
respectively, in Case 2
∞
T2(LξL∗
η) =
Φ(2)
zi,wi(LξL∗
η)
Xi=1
= ∞
Xi=1
∞
Xt=0
= ψ2(k + l − 2)LξL∗
η.
zi((k − 1) + t)wi((l − 1) + t)! (LξL∗
η)
This completes the proof.
(cid:3)
Lemma 5.4. Define T = T1 + T2 + c Id where Id denotes the identity
operator on B(H). Then for k, l ≥ 0 and ξ ∈ Λ(k), η ∈ Λ(l) we have
T (LξL∗
η) =(cid:26) φ(k + l)LξL∗
φ(k + l − 1)LξL∗
η
η
in Case 1
in Case 2.
Note that by Lemma 5.2 this implies that T (1) = φ(1)1 and that for
n ≥ 1, Tφ(a1a2 . . . an) = φ(n)a1a2 . . . an.
Proof. Assume we are in Case 1, then
T (LξL∗
η) = T1(LξL∗
η) + T2(LξL∗
η) + cLξL∗
η
= (ψ1(k + l) + ψ2(k + l) + c) LξL∗
η
= φ(k + l)LξL∗
η.
Here we use the definition of T , then Lemma 5.3, and lastly Lemma
4.6. If we are in Case 2, we similarly get
η) + T2(LξL∗
η) = T1(LξL∗
η) + cLξL∗
T (LξL∗
η
= (ψ1(k + l) + ψ2(k + l − 2) + c) LξL∗
η
= (ψ2(k + l − 1) + ψ1(k + l − 1) + c) LξL∗
η
= φ(k + l − 1)LξL∗
η.
Here we furthermore use ψ2(n) = ψ1(n + 1), for n ≥ 0.
(cid:3)
By this result we have proven the existence of Mφ in Proposition 5.1,
and it remains to calculate the cb-norm.
Lemma 5.5. We have kT kcb ≤ kφkC .
RADIAL MULTIPLIERS ON REDUCED FREE PRODUCTS
13
Proof. Let xi, yi ∈ l2(N0) then we have by Lemma 4.4 that kΦ(1)
xi,yikcb ≤
xi,yi we have by Remark
i=1 Φ(1)
4.7
kxik2kyik2. Furthermore, since T1 = P∞
Xi=1
xi,yikcb ≤
kT1kcb ≤
∞
Xi=1
kΦ(1)
respectively,
kxik2kyik2 = khk1,
∞
Xi=1
∞
∞
kT2kcb ≤
kΦ(2)
zi,wikcb ≤
kzik2kwik2 = kkk1.
Xi=1
Hence kT kcb ≤ kT1kcb + kT2kcb + kcIdkcb ≤ khk1 + kkk1 + c = kφkC
as desired.
(cid:3)
Combining Lemmas 5.4 and 5.5 we obtain Porposition 5.1, and there-
fore an application of Proposition 2.5 yields the conclusion of Theorem
2.2.
6. Examples
6.1. The case φs(n) = sn. As a first example we will look at a
simple φ where kφkC can be calculated explicitly.
Corollary 6.1. Let D = {s ∈ Cs < 1} and s ∈ D. Denote by φs
the function φs(n) := sn. Then φs defines a radial multiplier Mφs on
A = ∗i∈I(Ai, ωi), respectively, (M , ω) = ¯∗i∈I(Mi, ωi) as in Theorem
2.2. Moreover, kMφskcb ≤ 1 − s/(1 − s).
Proof. The conclusion follows from Theorem 2.2, once we show that φ
belongs to C and that kφkC = khk1 + kkk1 + c ≤ 1 − s/(1 − s).
Observe first that c = limn→∞ φ(n) = limn→∞ sn = 0 as c < 1.
Furthermore, φ(i + j + 1) − φ(i + j + 2) = s(φ(i + j) − φ(i + j + 1)) so
k = s · h, hence kφkC = (1 + s)khk1. Moreover φ(i + j) − φ(i + j + 1) =
(1 − s)si+j so h = (1 − s)m, where m is the matrix mi,j = si+j. This
gives kφkC = (1+s)khk1 = (1+s)1−skmk1. Now m = a⊙¯a, where
2 = 1/(1 − s2). Combining
a = (sk)k≥0 ∈ l2(N0), hence kmk1 = kak2
these calculations we get kφkC = 1 − s/(1 − s), which proves the
corollary.
(cid:3)
6.2. Wysoczański's theorem. As a second example, we will show
that Wysoczański's result, apart from determining when equality holds,
is a special case of Theorem 2.2.
Theorem 6.2 ([12, Theorem 6.1]). Let G = ∗i∈IGi be the free product
of a family of discrete groups, and let g ∈ G be g = g1g2 . . . gn where
gj ∈ Gij \{e}, j1, . . . , jn ∈ I and j1 6= j2 6= · · · 6= jn. If φ ∈ C then
φ(g) = φ(n) is a Herz-Schur multiplier on G. Moreover k φkHS ≤ kφkC .
RADIAL MULTIPLIERS ON REDUCED FREE PRODUCTS
14
Proof. Let φ ∈ C and g = g1 . . . gn ∈ G as above. Now by [1, p.
301] and [4] we have k φkHS = k φkM0A(G) = k M φkcb where M φ is the
operator M φ(λ(g)) = φ(g)λ(g) for g ∈ G and λ the left regular rep-
resentation. By the definition of φ this is M φ(λ(g)) = φ(n)λ(g) and
by the definition of L(G) we have λ(g) = λ(g1)λ(g2) . . . λ(gn). Hence
M φ(λ(g)) = Mφ(λ(g)), where Mφ is as defined in Theorem 2.2. Ap-
plying the theorem one obtains that φ is a Herz-Schur-multiplier, and
k φkHS ≤ kφkC .
(cid:3)
6.3. Relation to Houdayer and Ricard's results. Recently, C.
Houdayer and É. Ricard in [7] proved results concerning radial mul-
tipliers on free Araki-Woods factors related to functions from a class
C ′ quite similar to C . Although their results apply to different objects
than those considered in this paper, we will discuss in this section the
issue of how their methods could be applied to prove Theorem 2.2.
We start by defining the class of functions C ′, mentioned above.
Definition 6.3. Let C ′ denote the set of functions φ : N0 → C for
which the Hankel matrix h = (φ(i + j) − φ(i + j + 2))i,j≥0 is of trace-
class.
Observe that this implies the existence of c1, c2 ∈ C and a unique
ψ : N0 → C such that φ(n) = c1+(−1)nc2+ψ(n) and limn→∞ ψ(n) = 0.
For φ ∈ C ′ put kφkC ′ = c1 + c2 + khk1.
In [7] the following two results for functions in the class C ′ are proved,
note the resemblance with Theorem 2.2. In what follows T denotes the
Toeplitz algebra, and furthermore, Γ(H, Ut)′′ denotes the free Araki-
Woods factor associated to a real Hilbert space H and a one parameter
group of orthogonal transformations (Ut). (See [7, Sections 2.5 and 3.1]
for more precise definitions).
Theorem 6.4 ([7, Proposition 3.3]). A function φ belongs to C ′ if and
only if the operator γ defined by γ(Si(S∗)j) = φ(i + j) extends to a
bounded map on T . Moreover, kγkT ∗ = kφkC ′, and we say that γ is
the radial functional associated with φ.
Theorem 6.5 ([7, Theorem 3.5]). Let φ : N0 → C. Then φ de-
fines a completely bounded radial multiplier on Γ(H, Ut)′′ if and only
if the radial functional γ on T associated to φ is bounded. Moreover,
kMφkcb = kγkT ∗.
A similar argument as in the proof of these theorems could be used to
prove Theorem 2.2, if one could prove the existence of a ∗-isomorphism
π : C ∗(Lγγ ∈ Λ(1)) → C ∗(Lγγ ∈ Λ(1)) ⊗ C ∗(S2, SS∗)
such that
π(LξL∗
η) =(cid:26) LξL∗
LξL∗
η ⊗ S2k(S∗)2l
η ⊗ S2k−1(S∗)2l−1
in Case 1
in Case 2
(6.1)
RADIAL MULTIPLIERS ON REDUCED FREE PRODUCTS
15
for all k, l ≥ 0 and ξ ∈ Λ(k), η ∈ Λ(l).
Indeed, if this were the case we could choose w ∈ C ∗(S2, SS∗)∗ as
w(Sk(S∗)l) =(cid:26) φ(cid:0) k+l
2 (cid:1)
0
if k + l even
otherwise.
This functional would be bounded if h were trace-class. Moreover
kId ⊗ wkcb = kwk = kφkC . Letting T be defined by T = (Id ⊗ w) ◦ π
we would get
T (LξL∗
η) =(cid:26) φ(k + l)LξL∗
φ(k + l − 1)LξL∗
η
η
in Case 1
in Case 2
with kT kcb ≤ kwkkπkcb ≤ kwk = kφkC . Hence, by Lemma 5.2 T would
be Mφ as defined in Theorem 2.2 and thus be completely bounded.
It is however not possible to construct such an isomorphism. Let for
instance I = 1 and dim(H) = 2 and let eij denote the matrix units
with respect to the basis (Ω, γ) of H. Then we have e01e10 = e00, but
Φ(e01)Φ(e10) = e00 ⊗ S∗S 6= 1 ⊗ 1 − e11 ⊗ SS∗ = Φ(e00).
However, note that it would be sufficient if there existed a unital
completely positive π satisfying (6.1). To find such an operator we can
regard l2(N0) = l2(N0)even ⊕ l2(N0)odd. In this case S2 on l2(N0) can be
realized as S ⊕ S and SS∗ on l2(N0) can be realized as SS∗ ⊕ 1. Then
it would be enough to find unital completely positive operators π1, π2
such that
π1(LξL∗
η) = LξL∗
η ⊗ Sk(S∗)l,
respectively,
π2(LξL∗
η) =(cid:26) LξL∗
LξL∗
η ⊗ Sk(S∗)l
η ⊗ Sk−1(S∗)l−1
(6.2)
(6.3)
in Case 1
in Case 2.
Since π(1) = π(2) = 1 we have kπ1kcb, kπ2kcb ≤ 1 and π = π1 ⊕ π2
is unital completely positive too. Hence T = (Id ⊗ w) ◦ π would be as
desired.
i=0 Pi+n⊗ei0, where eij are the matrix units in B(l2(N0)),
and use the convention Pm = 0 if m < 0. Now it can be shown that
Set Un =P∞
π1(x) =
0
Xn=−∞
0
Un(x ⊗ 1)U ∗
n +
∞
Xn=1
∞
Un(ρn(x) ⊗ 1)U ∗
n,
Un(x ⊗ 1)U ∗
n +
Un(ρn−1(ǫ(x)) ⊗ 1)U ∗
n
respectively,
π2(x) =
Xn=−∞
Xn=1
are unital completely positive and fulfill (6.2) and (6.3). The proof of
this fact can be given by an argument quite similar to that given in
Sections 4 and 5. We leave the details to the reader.
RADIAL MULTIPLIERS ON REDUCED FREE PRODUCTS
16
7. Integral representation of functions from C
In [6] the following integral representation was proved for φ ∈ C ′,
where C ′ is the set of functions φ : N0 → C from Definition 6.3. The
set C ′ is not defined in [6], but the result follows form [6, Theorem 2.12
and Theorem 4.2].
Theorem 7.1. Let ψ : N0 → C be a function. Then the following are
equivalent:
(1) ψ ∈ C ′
(2) There exists a complex Borel measure µ on D and constants
c+, c− ∈ C, such that
ψ(n) = c+ + (−1)nc− +ZD
and
sndµ(s) < ∞
(7.1)
1 − s2
1 − s2 dµ(s) < ∞.
ZD
Moreover, for φ ∈ C ′, the measure µ in (7.1) can be chosen such that
c+ + c− +ZD
1 − s2
1 − s2 dµ(s) ≤
8
π
kψkC ′.
We will prove next a similar characterization of functions in C :
Theorem 7.2. Let φ : N0 → C be a function. Then the following are
equivalent:
(1) φ ∈ C
(2) There exists a constant c ∈ C and a complex Borel measure ν
sndν(s)
(7.2)
Moreover, for φ ∈ C , the measure ν in (7.2) can be chosen such that
c +ZD
1 − s
1 − s
dν(s) ≤
8
π
kφkC .
Proof. (1) implies (2). Let φ ∈ C and put
φ(n) =(cid:26) φ( n
2 )
0
if n is even
if n is odd.
Then by Definition 2.1 and Definition 6.3, φ ∈ C ′ and k φkC ′ = kφkC .
From Theorem 7.1 there exists a complex measure µ on D and constants
on D such that
and
φ(n) = c +ZD
ZD
1 − s
1 − s
dν(s) < ∞.
RADIAL MULTIPLIERS ON REDUCED FREE PRODUCTS
17
c+, c− ∈ C such that
and
sndµ(s) < ∞
ψ(n) = φ(2n) = c+ + c− +ZD
c+ + c− +ZD
1 − s2
1 − s2 dµ(s) ≤
Let ν be the range measure of µ by the map s → s2 of D onto D,
and put c = c+ + c−. Then ν is less or equal to the range measure of
µ by the map s → s2. Hence
8
π
kψkC .
sndν(s)
s2ndµ(s) = c +ZD
φ(n) = c +ZD
dν(s) ≤ c+ + c− +ZD
and
c +ZD
1 − s
1 − s
1 − s2
1 − s2
dµ(s) ≤
8
π
kφkC .
This proves (1) implies (2) and the last statement in Theorem 7.2.
Conversely if (2) holds, the Hankel matrices h, k from Definition 2.1
have the entries
and
hij =ZD
kij =ZD
si+j(1 − s)dν(s)
si+js(1 − s)dν(s).
By the proof of Corollary 6.1,
k(si+j)i,j≥0k1 =
1
1 − s2 ,
s ∈ D.
Hence
khk1 + kkk1 ≤ZD
1 − s + s(1 − s)
1 − s2
dν(s) =ZD
1 − s
1 − s
dν(s) < ∞
which shows that φ ∈ C .
(cid:3)
In [9] N. Ozawa proved that if Γ is a discrete hyperbolic group (in
the sense of M. Gromov [5]), then Γ is weakly amenable. The proof
was obtained by showing that the metric d : Γ × Γ → N0 (w.r.t. the
Cayley graph of Γ) satisfies three properties (1), (2) and (3) listed in
[9, Theorem 1].
As an application of Theorem 7.2, we will show below, that the first
condition (1) from [9] is sufficient to prove that Γ is weakly amenable.
For the definition of weak amenability and of the constant Λ(Γ) for a
weakly amenable group Γ, we refer to [2, Section 12.3].
RADIAL MULTIPLIERS ON REDUCED FREE PRODUCTS
18
Recall that a metric on a discrete metric space (X, d) is called proper
if the ball B(x, r) = {y ∈ X : d(x, y) < r} is finite for all x ∈ X and
all r > 0.
Theorem 7.3. Let Γ be a discrete countable group and let d : Γ × Γ →
N0 be a proper left invariant metric. Put
φs(x) = sd(x,e),
s ∈ D, x ∈ Γ.
Assume that there exists a constant C ≥ 1, such that ψs ∈ M0A(Γ) for
all s ∈ D and
kφskM0A(Γ) ≤ C
1 − s
1 − s
,
s ∈ D.
Then Γ is weakly amenable with constant Λ(Γ) ≤ C.
Remark 7.4. As in [6] we have used the notation M0A(Γ) for the
set of completely bounded Fourier multipliers on Γ. Note that in [2,
Section 12.3] the space M0A(Γ) is denoted B2(Γ).
We first prove
Lemma 7.5.
(1) Put χn(k) = δkn for n, k ≥ 0. Then χn ∈ C and
kχnkC ≤ max{1, 4n},
n ≥ 0.
(2) For r ∈ (0, 1) and l ≥ 0, put
φr(k) = rk
φr,n(k) =(cid:26) rk 0 ≤ k ≤ n
k > n.
0
then φr, φr,n ∈ C , kφrkC = 1 and for fixed r ∈ (0, 1)
lim
n→∞
kφr − φr,nkC = 0.
Proof. From Definition 2.1 we have χn ∈ C , and kχnkC = kHnk1 +
kKnk1 where
Hn(i, j) = χn(i + j) − χn(i + j + 1)
Kn(i, j) = χn(i + j + 1) − χn(i + j + 2).
If H = (hij)∞
i,j=0 is a matrix of complex numbers for whichPi,j hij <
∞, then H is of trace class and kHk1 ≤ P∞
i,j=0 hij. Hence kχnkC ≤
(2n+ 1) + (2n−1) for n ≥ 1 and kχ0kC ≤ 1 which proves (1). It follows
from Corollary 6.1, that kφrkC = 1, 0 < r < 1. By (1),
kφr − φr,nkC = k
which proves (2).
rkχkkC ≤
∞
Xk=n+1
4krk
∞
Xk=n+1
(cid:3)
RADIAL MULTIPLIERS ON REDUCED FREE PRODUCTS
19
Proof of Theorem 7.3. Let φ : N0 → C be a function from C , and put
φ(x) = φ(d(x, e)),
x ∈ Γ.
Then by (7.1) and the integral representation of φ from Theorem 7.2
it follows that φ is a completely bounded Fourier multiplier on Γ and
that
k φkM0A(Γ) ≤
8
π
kφkC .
(7.3)
Let φr and φr,n be as in Lemma 7.5. Then by (7.1) k φrkM0A(Γ) ≤ C.
Moreover by (7.3) and Lemma 7.5
k φr − φr,nkM0A(Γ) = 0
lim
n→∞
for fixed r ∈ (0, 1). Put rk = 1 − 1/k, k ≥ 1 and chose for each k ≥ 2
an nk ≥ k, such that
k φrk − φrk,nkkM0A(Γ) ≤
1
k
.
Then ψk = φrk,nk form a sequence of finitely supported functions
on Γ, such that kψkkM0A(Γ) < C + 1/k, and limk→∞ ψk(x) = 1 for all
x ∈ Γ. Hence Γ is weakly amenable and Λ(Γ) ≤ C.
(cid:3)
Remark 7.6. By [9, Theorem 1] and the proof of Theorem 7.3 it
follows that for every hyperbolic group Γ and every φ ∈ C ′, the function
φ(x) = φ(d(x, e)),
x ∈ Γ
is a completely bounded Fourier multiplier on Γ and k φkM0A(Γ) ≤
π kφkC ′, where C is the constant in [9, Theorem 1 (1)].
8C
References
[1] M. Bożejko and G. Fendler. Herz-Schur multipliers and completely bounded
multipliers of the Fourier algebra of a locally compact group. Boll. Un. Mat.
Ital. A (6), 3(2):297 -- 302, 1984.
[2] N. P. Brown and N. Ozawa. C ∗-algebras and finite-dimensional approxima-
tions, volume 88 of Graduate Studies in Mathematics. American Mathematical
Society, Providence, RI, 2008.
[3] E. Christensen and A. M. Sinclair. A survey of completely bounded operators.
Bull. London Math. Soc., 21(5):417 -- 448, 1989.
[4] J. De Cannière and U. Haagerup. Multipliers of the Fourier algebras of some
simple Lie groups and their discrete subgroups. Amer. J. Math., 107(2):455 --
500, 1985.
[5] M. Gromov. Hyperbolic groups. In Essays in group theory, volume 8 of Math.
Sci. Res. Inst. Publ., pages 75 -- 263. Springer, New York, 1987.
[6] U. Haagerup, T. Steenstrup, and R. Szwarc. Schur multipliers and spherical
functions on homogeneous trees. Internat. J. Math., 21(10):1337 -- 1382, 2010,
0908.4424.
[7] C. Houdayer and É. Ricard. Approximation properties and absence of Cartan
subalgebra for free Araki-Woods factors. Adv. Math., 228(2):764 -- 802, 2011,
1006.3689.
RADIAL MULTIPLIERS ON REDUCED FREE PRODUCTS
20
[8] R. V. Kadison and J. R. Ringrose. Fundamentals of the theory of operator
algebras. Vol. I, volume 15 of Graduate Studies in Mathematics. American
Mathematical Society, Providence, RI, 1997. Elementary theory, Reprint of
the 1983 original.
[9] N. Ozawa. Weak amenability of hyperbolic groups. Groups Geom. Dyn.,
2(2):271 -- 280, 2008, 0704.1635.
[10] É. Ricard and Q. Xu. Khintchine type inequalities
reduced free
products and applications. J. Reine Angew. Math., 599:27 -- 59, 2006,
arXiv:math/0505302.
for
[11] D. V. Voiculescu, K. J. Dykema, and A. Nica. Free random variables, volume 1
of CRM Monograph Series. American Mathematical Society, Providence, RI,
1992. A noncommutative probability approach to free products with applica-
tions to random matrices, operator algebras and harmonic analysis on free
groups.
[12] J. Wysoczański. A characterization of radial Herz-Schur multipliers on free
products of discrete groups. J. Funct. Anal., 129(2):268 -- 292, 1995.
Department of Mathematical Sciences, University of Copenhagen,
Universitetsparken 5, 2100 Copenhagen Ø, Denmark
E-mail address: [email protected]
Department of Mathematics and Computer Science, University of
Southern Denmark, Campusvej 55, 5230 Odense M, Denmark
E-mail address: [email protected]
|
1703.07073 | 3 | 1703 | 2018-03-18T04:32:24 | Heisenberg Modules over Quantum 2-tori are metrized quantum vector bundles | [
"math.OA"
] | The modular Gromov-Hausdorff propinquity is a distance on classes of modules endowed with quantum metric information, in the form of a metric form of a connection and a left Hilbert module structure. This paper proves that the family of Heisenberg modules over quantum two tori, when endowed with their canonical connections, form a family of metrized quantum vector bundles, as a first step in proving that Heisenberg modules form a continuous family for the modular Gromov-Hausdorff propinquity. | math.OA | math | HEISENBERG MODULES OVER QUANTUM 2-TORI ARE
METRIZED QUANTUM VECTOR BUNDLES
FRÉDÉRIC LATRÉMOLIÈRE
Abstract. The modular Gromov-Hausdorff propinquity is a distance on classes
of modules endowed with quantum metric information, in the form of a metric
form of a connection and a left Hilbert module structure. This paper proves
that the family of Heisenberg modules over quantum two tori, when endowed
with their canonical connections, form a family of metrized quantum vector
bundles, as a first step in proving that Heisenberg modules form a continuous
family for the modular Gromov-Hausdorff propinquity.
.
A
O
h
t
a
m
[
3
v
3
7
0
7
0
.
3
0
7
1
:
v
i
X
r
a
1. Introduction
The primary purpose of our research is to bring forth an analytic framework, con-
structed around Gromov-Hausdorff-like hypertopologies on quantum metric spaces,
to bear on problems from mathematical physics and noncommutative geometry
[17, 13, 19, 18, 14, 12, 3, 15]. We constructed an hypertopology on classes of
Hilbert modules over quantum metric spaces in [16] as a far-reaching generaliza-
tion of the Gromov-Hausdorff propinquity. We constructed a distance, up to full
quantum isometry, called the modular Gromov-Hausdorff propinquity, on a class
of objects which generalize Hermitian vector bundles over Riemannian manifolds.
These metrized quantum vector bundles are natural objects for noncommutative
geometry and mathematical physics, as they carry a metric structure and a form of
generalized connection, and we are now able to discuss such questions as continuity
and approximations, not only of quantum compact metric spaces, but also of their
associated modules. As modules are fundamental objects in C*-algebra theory and
their geometry, this new development allows us to further our goal of a geometric
theory of the class of C*-algebras.
This paper brings into our noncommutative metric geometry framework some
very important examples of modules, namely Heisenberg modules over quantum
2-tori. These modules come naturally equipped with a connection induced by the
action of the Heisenberg Lie group. This noncommutative construct played the
central role in the beginning of Connes' noncommutative geometry [5], where the
Heisenberg modules over quantum 2-tori and their connections were first built.
Rieffel [27] then proved that these Heisenberg modules, the finite rank free mod-
ules, and their direct sums, describe all the finitely generated projective modules
Date: September 17, 2018.
2000 Mathematics Subject Classification. Primary: 46L89, 46L30, 58B34.
Key words and phrases. Noncommutative metric geometry, Gromov-Hausdorff convergence,
Monge-Kantorovich distance, Quantum Metric Spaces, Lip-norms, D-norms, Hilbert modules,
noncommutative connections, noncommutative Riemannian geometry, unstable K-theory.
This work is part of the project supported by the grant H2020-MSCA-RISE-2015-691246-
QUANTUM DYNAMICS.
1
2
FRÉDÉRIC LATRÉMOLIÈRE
over quantum tori. Connes and Rieffel [7] proved that the natural connections on
Heisenberg modules solve the noncommutative Yang-Mills problem. We will now
prove that Heisenberg modules are fundamental examples of metrized quantum
vector bundles. Doing so then allows us to discuss in [20] the continuity, for the
modular propinquity, of family of Heisenberg modules as the quantum 2-tori vary
continuously for the propinquity. This will be our first, significant application of the
modular propinquity. Informally, the continuity result in [20] can be understood
as a form of continuity of K-theory. Thus, this paper and [20] are two parts of the
study of the metric geometry of Heisenberg modules.
As a matter of convention throughout this paper, we will use the following no-
tations.
Notation 1.1. By default, the norm of a normed vector space E is denoted by
k·kE. When A is a C*-algebra, the space of self-adjoint elements of A is denoted
by sa (A). The state space of A is denoted by S (A). In this work, all C*-algebras
A will always be unital with unit 1A.
Convention 1.2. If P is some seminorm on a vector subspace D of a vector space
E, then for all x ∈ E \ D we set P (x) = ∞. With this in mind, the domain D of
P is the set {x ∈ E : P (x) < ∞}, with the usual convention that 0∞ = 0 while all
other operations involving ∞ give ∞.
Noncommutative metric geometry [6, 28, 30] studies noncommutative general-
izations of Lipschitz algebras, defined as follows.
Definition 1.3. An ordered pair (A, L) is a Leibniz quantum compact metric space
when A is a unital C*-algebra, and L is a seminorm defined on a dense Jordan-Lie
subalgebra dom (L) of the space of self-adjoint elements sa (A) of A such that:
(1) {a ∈ dom (L) : L(a) = 0} = R1A,
(2) the Monge-Kantorovich metric mkL defined on the state space S (A) of A
by setting, for any two ϕ, ψ ∈ S (A):
mkL(ϕ, ψ) = sup{ϕ(a) − ψ(a) : a ∈ dom (L), L(a) 6 1}
metrizes the weak* topology restricted to S (A),
(3) L is lower semi-continuous,
(4) max(cid:8)L(cid:0) ab+ba
2
(cid:1) , L(cid:0) ab−ba
2i (cid:1)(cid:9) 6 kakAL(b) + kbkAL(a).
Leibniz quantum compact metric spaces, and more generally quasi-Leibniz quan-
tum compact metric spaces (a generalization we will not need in this paper),
form a category with the appropriate notion of Lipschitz morphisms [21], con-
taining such important examples as quantum tori [28], Connes-Landi spheres [22],
group C*-algebras for Hyperbolic groups and nilpotent groups [29, 23], AF alge-
bras [12], Podlès spheres [2], certain C*-crossed-products [1], among others. Any
compact metric space (X, d) give rise to the Leibniz quantum compact metric space
(C(X), Lip) where C(X) is the C*-algebra of C-valued continuous functions over
X, and Lip is the Lipschitz seminorm induced by d.
Rieffel characterized the main property of Leibniz quantum compact metric
spaces as follows:
Theorem 1.4 ([28, Theorem 1.9]). Let (A, L) be a pair with a unital C*-algebra
A and a seminorm L defined on a dense subspace dom (L) of sa (A). The following
assertions are equivalent:
D-NORMS ON HEISENBERG MODULES
3
(1) the Monge-Kantorovich metric mkL defined for any two ϕ, ψ ∈ S (A) by
mkL(ϕ, ψ) = sup{ϕ(a) − ψ(a) : L(a) 6 1}, metrizes the weak* topology on
S (A),
(2) the diameter diam (S (A), mkL) is finite and:
{a ∈ sa (A) : L(a) 6 1 and kakA 6 1}
is norm precompact.
In [16], we extend this idea to noncommutative analogues vector bundles. Our
classical prototype of a metrized quantum vector bundle is given by the module
ΓV of continuous sections of a vector bundle V over a compact Riemannian man-
ifold M with metric g, endowed with a hermitian metric h and some associated
metric connection ∇. For any two ω, η ∈ ΓV , we then set hω, ηiV : x ∈ M 7→
RX hx(ωx, ηx) dVol(x) where Vol is the volume form over M for g, which turns ΓV
into a C(M )-left Hilbert module. We also define, for all ω ∈ M , the norm D(ω)
as the operator norm for the operator ∇ω : X ∈ Γ(T M ) 7→ ∇X ω ∈ ΓV -- noting
that the space of vector fields ΓT M of M has a norm induced by the metric g.
Our general definition for a metrized quantum vector bundle abstracts this picture.
For the present paper, we shall only deal with so-called Leibniz metrized quantum
vector bundles, even though our framework in [16] is more general. This is the main
definition for this paper.
Definition 1.5 ([16, Definition 3.8]). A 5-tuple (M ,h·,·iM , D, A, L) is a metrized
quantum vector bundle when:
(1) a Leibniz quantum compact metric space (A, L) called the base space,
(2) a A-left Hilbert module (M ,h·,·iM ),
(3) a norm D defined on a dense subspace of M such that D(ω) > phω, ωiM
for all ω ∈ M , and such that the set:
is compact in M ,
{ω ∈ M : D(ω) 6 1}
(4) for all a ∈ sa (A) and for all ω ∈ M , we have:
DM (aω) 6 (kakA + LA(a))DM (ω),
which we call the inner Leibniz inequality for DM ,
(5) for all ω, η ∈ M , we have:
max{LA (ℜhω, ηiM ) , LA (ℑhω, ηiM )} 6 2DM (ω)DM (η),
which we call the modular Leibniz inequality for DM .
We refer to [16] for a discussion of these objects, where in particular [16, Ex-
ample 3.10] shows that the prototype of a hermitian vector bundle over a compact
Riemannian manifold, as sketched above, is indeed an example of a metrized quan-
tum vector bundle. We note that Definition (1.5) includes a compactness condition
which mirrors the compactness condition in Theorem (1.4).
Heisenberg modules, equipped with the analogue of a connection as in [5], over
quantum 2-tori, have a similar signature to a metrized quantum vector bundle. The
key difficulty is to prove that the connection can be used to define a D-norm, as in
Definition (1.5), whose unit ball is actually compact in the Hilbert modules norm
of Heisenberg modules. The main result of this paper is to prove that indeed, this
is the case.
4
FRÉDÉRIC LATRÉMOLIÈRE
We begin our work with a presentation of Heisenberg modules, which allow us
to fix our notations for the rest of the paper and [20]. We then prove a series of
lemmas about convergence in the Hilbert modules norm for the Heisenberg modules
-- as these norms are complicated, these lemmas will prove very helpful both in this
paper and in [20]. We prove in the process of this second section that Heisenberg
modules form a continuous field of Banach spaces -- a result which will prove
helpful in [20] and is of independent interest. This result uses the same tools as the
proof that the action of the Heisenberg group on Heisenberg modules is strongly
continuous, which is part of the next section of this paper, where properties of the
Heisenberg group actions which we will need in our work are established. Now,
with all these basic tools in hand, we show how to use Lie group actions to define
D-norm candidates, which have all the desired properties of D-norms except maybe
for the key compactness property of their unit ball. This compactness property for
the Heisenberg modules D-norms is the subject of the last section of this paper,
which conclude our main result.
Importantly, our methods in this paper are designed not only in support of the
main theorem here, but also as key tools for the study of the continuity of the
Heisenberg modules in [20]. For the problem of continuity, we will need not just to
be able to pick finite subsets of the compact unit ball of some D-norm which are
ε-dense for some ε > 0, but also to pick such a finite set which is uniformly ε-dense
across several Heisenberg modules as the D-norms vary. To do so, we will use the
approximation operators introduced in the last section of this paper.
2. Background on Quantum 2-tori and Heisenberg modules
Quantum 2-tori are the twisted convolution C*-algebras of Z2. The projective
finitely generated modules over quantum tori have been extensively studied, and
next to the free modules, the most important class of projective, finitely generated
modules over a quantum torus are the Heisenberg modules. This subsection intro-
duces these modules, as well as the notations we will use throughout this section
regarding quantum tori.
Twisted group C*-algebras are defined by twisting the convolution product over
a locally compact group by a representative of a continuous 2-cocycle of the group.
Notation 2.1. For any θ ∈ R, we define the skew bicharacter of R2:
(2.1)
eθ : ((x1, y1), (x2, y2)) ∈ R2 × R2 7−→ exp (iπθ(x2y1 − x1y2)) .
By [10], any 2-cocycle of Z2 is cohomologous to the restriction of a skew bichar-
acter eθ to Z2 × Z2 for some θ ∈ R. We shall use the same notation for eθ and its
restriction to Z2.
Moreover, for any θ, ϑ ∈ R, the skew bicharacters eθ and eϑ of Z2 are cohomol-
ogous if and only if θ ≡ ϑ mod 1. We note that, as skew bicharacters of R2, they
are cohomologous if and only if θ = ϑ.
We define the twisted convolution products on ℓ1(Z2), where we use the following
notation.
Notation 2.2. For any (nonempty) set E and any p ∈ [1,∞), the set ℓp(E) is the
set of all absolutely p-summable complex valued functions over E, endowed with
D-NORMS ON HEISENBERG MODULES
5
the norm:
for all ξ ∈ ℓp(E).
element of ℓp(E) for all p.
kξkℓp(E) = Xx∈E
1
p
ξ(x)p!
We write δn the function which is 1 at n and 0 otherwise; this function is an
Moreover, if p = 2 then (ℓ2(E),k · kℓ2(E)) is a Hilbert space, where the inner
product hξ, ηiℓ2(E) =Px∈E ξ(x)η(x) for all ξ, η ∈ ℓ2(E).
Definition 2.3. Let θ ∈ R and eθ be defined by Expression (2.1). The twisted
convolution product ∗θ is defined for all f, g ∈ ℓ1(Z2) and for all n ∈ Z2:
We now define:
f∗θg(n) = Xm∈Z2
f (m)g(n − m)eθ(m, n).
The adjoint of any f ∈ ℓ1(Z2) is defined for all n ∈ Z2 by:
f ∗(n) = f (−n).
veloping C*-algebra. To do so, we shall choose a natural faithful *-representation
One checks easily that (cid:0)ℓ1(Z2),∗θ,·∗(cid:1) is a *-algebra. In particular, the adjoint
operation is an isometry of (cid:0)ℓ1(Z2),k · kℓ1(Z2)(cid:1). We now wish to construct its en-
of (cid:0)ℓ1(Z2),∗θ,·∗(cid:1) on ℓ2(Z2). This representation was a key ingredient in the con-
struction of bridges between quantum tori in our work in [11] on convergence of
quantum tori for the quantum propinquity and will play a role in the convergence
of Heisenberg modules.
Notation 2.4. If T : E → F is a continuous linear map between two normed spaces,
we write its norm as TE
Theorem 2.5 ([32]). Let θ ∈ R. We define, for any n ∈ Z2 and ξ ∈ ℓ2(Z2), the
function:
F . When E = F , we simply write TF .
is a unitary eθ-projective representation of Z2, i.e.
θ
θ = eθ(n, m)U n+m
θ ξ : m ∈ Z2 7→ eθ(m, n)ξ(m + n).
U n
The map n ∈ Z2 7→ U n
for all n, m ∈ Z2.
θ U m
U n
If, for all f ∈(cid:0)ℓ1(Z2),∗θ,·∗(cid:1), we define:
πθ(f ) = Xn∈Z2
f (n)U n
θ
θ
which is a bounded operator on ℓ2(Z2) with:
πθ(f )ℓ2(Z2) 6 kfkℓ1(Z2),
then πθ is a faithful *-representation of (ℓ1(Z2),∗θ,∗).
Thus, we may define a C*-norm on ℓ1(Z2) by setting:
kfkAθ = πθ(f )ℓ2(Z2)
for all f ∈ ℓ1(Z2). We thus can define quantum 2-tori.
Definition 2.6. The quantum 2-torus Aθ is the completion of (ℓ1(Z2),∗θ,·∗) for
the norm πθ(·)ℓ2(Z2).
6
FRÉDÉRIC LATRÉMOLIÈRE
As per our general convention, the norm on Aθ is denoted by k·kAθ for all θ ∈ R.
Remark 2.7. Let θ ∈ R. By construction, ℓ1(Z2) is identified with a dense *-
subalgebra of Aθ, and we shall employ this identification all throughout this paper.
With this identification, we also note that for all f ∈ ℓ1(Z2) we have kfkAθ 6
kfkℓ1(Z2), a fact which we will use repeatedly in the next section.
We take one derogation to the convention of using the same symbol for an element
of ℓ1(Z2) and its counter part in a given quantum torus, because the following
notation is at once common and convenient.
Notation 2.8. Let θ ∈ R. The element δ1,0 is denoted by uθ and the element δ0,1 is
denoted by vθ when regarded as elements of Aθ.
The geometry, and in particular the metric geometry [28], of the quantum tori
is obtained by transport of structure using the dual action of the torus given as
follows:
Theorem-Definition 2.9. [32] For all z = (z1, z2) ∈ T2 there exists a unique
*-automorphism βz
θ of Aθ such that, for any f ∈ ℓ1(Z2) and (n, m) ∈ Z2, we have:
βz
θ f (n, m) = zn
1 zm
2 f (n, m).
The map z ∈ T2 7→ βz
action. Moreover, β is ergodic, in the sense that:
θ is a strongly continuous action of T2 on Aθ called the dual
(cid:8)a ∈ Aθ : ∀z ∈ T2 βz(a) = a(cid:9) = C1Aθ .
We now turn to the class of modules to which we shall apply our new modular
propinquity. We construct these modules following [5] using the universal property
of quantum 2-tori, which we now recall.
Proposition 2.10 ([32]). Let θ ∈ R. If U , V are two unitary operators on some
Hilbert space H such that U V = exp(2iπθ)V U for some θ ∈ [0, 1), then there exists
a *-morphism : Aθ → B(H ) such that (uθ) = U and θ(vθ) = V . The range
of is C∗(U, V ).
Another way to state Proposition (2.10) is that, for any θ ∈ R, if ς is some
projective representation of Z2 on some Hilbert space H for some multiplier of
Z2 cohomologous to eθ, then H is a module over Aθ. Indeed, Proposition (2.10)
gives us a *-morphism from Aθ to the C*-algebra B(H ) of all bounded linear
operators on H , with (uθ) = ς 1,0 and (vθ) = ς 0,1. Thus H is a Aθ module.
With this observation in mind, we now turn to the construction of some particular
projective representations of Z2. The idea, found in [5] and explicit in [25], is to
take the tensor product of a projective representation of R2, restricted to Z2, and a
finite dimensional projective representation of Zq for some q ∈ N\{0}. By adjusting
the choice of the multipliers associated with each projective representation, we get
the desired module structure.
Projective representations of R2 are naturally related to the representations of
the Heisenberg group, and we will make important use of this fact in our work. We
thus begin with setting our notations for the Heisenberg group.
Convention 2.11. The vector space Cd is endowed by default with its standard
j=1 zjyj, whose associated norm is
inner product h(z1, . . . , zd), (y1, . . . , yd)iCd = Pd
denoted by k · kCd.
D-NORMS ON HEISENBERG MODULES
7
Notation 2.12. The Heisenberg group is the Lie group given by:
H3 =
1 x u
0
y
0
1
0
: x, y, u ∈ R
1
.
We shall identify H3 with R3 via the natural map (x, y, u) ∈ R3 7→
1
0
which is a Lie group isomorphism once we equip R3 with the multiplication:
1 x u
0
y
0
1
,
(x1, y1, u1)(x2, y2, u2) = (x1 + x2, y1 + y2, u1 + u2 + x1y2)
for all (x1, y1, u1), (x2, y2, u2) ∈ R3.
gleaned by looking at its Lie algebra, which is given by:
The importance of the Heisenberg group for quantum mechanics [8] may be
which is a 2-nilpotent Lie algebra. We easily compute that for all x, y, u ∈ R3:
(2.2)
h =
0 x u
0
y
0
0
0
exp
0
0
0 x u
0
y
0
: x, y, u ∈ R
0
0
=
1
.
1 x u + 1
2 xy
0
y
0
1
0
This expression for the exponential will be important for our construction. Note
that the exponential map is both injective and surjective.
We now set:
0
0
0
P =
1
0
0
0
0
0
, Q =
0
0
0
0 0
0 1
0 0
and T =
0 0
0 0
0 0
1
0
0
.
We easily check that [P, Q] = T = −[Q, P ] while other other commutators between
P , Q and T are null, and spanC{P, Q, T} = h.
We note that in particular, T is central, and thus the relations defining h from the
basis {P, Q, T} are the structural equations of quantum mechanics -- the canonical
commutation relation, as proposed by Heisenberg, in order to express the uncer-
tainty principle between two conjugate observables. We refer to [8] for a detailed
analysis of the Heisenberg group and its connections to the Moyal product, pseudo-
differential calculus, and more fascinating topics.
Thus the study of the irreducible representations of H3 provide the irreducible
representations of the canonical commutation relations. We first note that:
H3.{(0, 0, u) : u ∈ R} = R2
is Abelian, and thus we get a collection of trivial, one-dimensional representations
of H3 by simply lifting the irreducible representations of R2.
If we set, for any ð ∈ R \ {0} and ξ ∈ L2(R):
(2.3)
αx,y,u
ð,1 ξ : s ∈ R 7→ exp(2iπ(ðu + sx))ξ(s + ðy)
then we define a unitary representation of H3, and any nontrivial irreducible unitary
representations of the Heisenberg group is unitarily equivalent to αð,1 for some
8
FRÉDÉRIC LATRÉMOLIÈRE
ð 6= 0 [8]. We note that they all are infinite dimensional (the other, trivial, unitary
representations of H3 are one-dimensional).
Let ð ∈ R \ {0}. For all (x, y) ∈ R2 and for all ξ ∈ L2(R), set:
σx,y
ð,1 ξ = α
(xP +yQ)
ξ
expH3
ð,1
x,y, xy
2
ð,1
= α
ξ : s ∈ R 7→ exp(iπðxy + 2iπsx)ξ(s + ðy).
The map σx,y
ð,1 is a unitary on L2(R) for all (x, y) ∈ R2. Moreover, for all (x1, y1),
(x2, y2) ∈ R2, we note that:
ð,1 σx2,y2
σx1,y1
ð,1 = eð((x1, y1), (x2, y2))σx1+x2,y1+y2
ð,1
,
i.e. σð,1 is a projective representation of R2 on L2(R) for the bicharacter eð, namely
the Schrödinger representation of "Plank constant" ð. Moreover, every nontrivial
irreducible unitary projective representation of R2 is unitarily equivalent to one of
σ1,ð for some ð 6= 0 (by nontrivial, we mean associated with a nontrivial cocycle).
We introduce one more notation which will prove very useful in defining our D-
norm on Heisenberg modules. If d ∈ N with d > 0, we define the following unitarry
operators on L2(R) ⊗ Cd:
αx,y,u
ð,d = αx,y,u
ð,d = σx,y
ð,1 ⊗ id and σx,y
ð,1 ⊗ id
for all x, y, u ∈ R, where id is the identity map on Cd. We trivially check that
αð,d is a unitary representation of H3 on L2(R) ⊗ Cd, while σð,d is a eð-projective
representation of R2 on L2(R) ⊗ Cd. Moreover, we also check immediately that
αx,y,0
ð,d = σx,y
ð,d for all x, y ∈ R.
We now turn to the projective representations of Z2
note that, for any p ∈ Z, the skew bicharacter e p
Z2
q -- which we keep denoting by e p
q
q
q, where q ∈ N\{0}. We first
of Z2 induces a skew bicharacter of
q is cohomologous
. By [10], any multiplier of Z2
to e p
q
for some p ∈ N.
For our purpose, we will thus get, up to unitary equivalence, every possible
q for arbitrary
finite dimensional unitary projective representations of the groups Z2
q ∈ N \ {0} by considering the following family.
Notation 2.13. Let p ∈ Z and q ∈ N \ {0}. Let n ∈ Z 7→ [n] ∈ Zq be the canonical
surjection. Let:
z
z2
. . .
zq−1
0
1
. . .
and vp,q =
. . .
0
. . .
. . .
1
1 0
,
up,q =
1
p,q = vq
p,q = 1, the map:
q (cid:17). Since uq
q 7→ ρz,w
with z = exp(cid:16) 2iπp
ρp,q,1 : (z, w) ∈ Z2
is well-defined. An easy computation shows that ρp,q,1 is a projective representation
of Z2
q.
p,q,1 = exp(cid:18) iπpnm
p,q where [n] = z and [m] = w
q (cid:19) un
p,qvm
D-NORMS ON HEISENBERG MODULES
9
For all d ∈ qN, d > 0, we now set:
ρn,m
p,q,d = ρn,m
p,q,1 ⊗ id d
q
where id d
q
is the identity map on C
d
q .
We remark that ρp,q,d acts on Cd, i.e. we parametrized ρ by the dimension of
the space on which it acts rather than the multiplicity of ρp,q,1, as it will make our
notations much simpler.
If p and q are relatively prime, the representation ρp,q,1 is irreducible, with
range the entire algebra of q × q matrices -- it is in fact, the only irreducible e p
-
projective representation of Z2
q up to unitary equivalence. Thus in general, any
finite dimensional unitary representation of Z2
q is unitarily equivalent to some ρl,r,d
for some l ∈ Z, r ∈ N \ {0}, d ∈ rN \ {0}, with l = 0 and r = 1 or l, r relatively
prime.
q
In order to construct the inner product on the Heisenberg modules, we shall need
to first work on a space of well-behaved functions inside the Hilbert space ℓ2(Z2)
on which quantum tori will act. This space will consist of the Schwarz functions.
Definition 2.14. Let E be a finite dimensional vector space. A function f : R → E
is a E-valued Schwarz function over R when it is infinitely differentiable on R and,
for all j ∈ N and all polynomial p ∈ R[X], we have:
lim
t→±∞(cid:13)(cid:13)(cid:13)
p(t)f (j)(t)(cid:13)(cid:13)(cid:13)E
= 0.
1+sj for all s ∈ R.
The space of all E-valued Schwarz functions over R is denoted by S(E).
We note that if f ∈ S(E) for some finite dimensional space E, then in particular,
f ∈ Lp(R) for all p ∈ [1,∞], since for any j ∈ N, there exists M > 0 such that
kf (s)kE 6 M
We now implement the scheme which we described a few paragraphs above to
construct modules over quantum tori. We refer to the mentioned works of Connes
and Rieffel for the details and justification behind the following construction.
Theorem-Definition 2.15 ( [5], [24], [7] ). Let θ ∈ R and q ∈ N\{0}. Let p ∈ Z,
q ∈ N \ {0} , and let d ∈ qN \ {0}. The Heisenberg module H p,q,d
is the module
over Aθ defined as follows.
of d
the same cocycle. Up to unitary conjugation, we assume that ρp,q,d acts on Cd.
Let ρp,q,d be the projective action of Z2
, consisting of the sum
q copies of the unique, up to unitary equivalence, irreducible representation with
q with cocycle e p
θ
q
Let:
ð = θ −
p
q
.
Let αð,1 be the action of the Heisenberg group H3 on L2(R) given by Expression
(2.3).
and [m], we set:
For (n, m) ∈ Z2, denoting the class of n and m in Z(cid:14)qZ , respectively, by [n]
For all n, m ∈ Z, the map n,m
p,q,ð,d is a unitary of L2(R) ⊗ Cd, and moreover
ð,1 ⊗ ρ[n],[m]
p,q,d .
p,q,ð,d = σn,m
n,m
p,q,ð,d is an eθ-projective representation of Z2.
10
FRÉDÉRIC LATRÉMOLIÈRE
By universality, the Hilbert space L2(R) ⊗ Cd is a module over Aθ, with, in
particular, for all f ∈ ℓ1(Z2) and ξ ∈ L2(R, Cd) = L2(R) ⊗ Cd:
f ξ = Xn,m∈Z
f (n, m)n,m
p,q,ð,dξ.
Let S p,q,d
θ
= S(Cd) ⊆ L2(R) ⊗ Cd. For all ξ, ω ∈ S p,q,d
θ
, define hξ, ωiH p,q,d
θ
as
the function in ℓ1(Z2) given by:
hξ, ωiH p,q,d
θ
: (n, m) ∈ Z2 7−→Dn,m
p,q,ð,dξ, ωEL2(R)⊗E
is the completion of S p,q,d
.
θ
for the norm associ-
The Heisenberg module H p,q,d
θ
ated with the Aθ-inner product h·,·iH p,q,d
θ
.
We note that S p,q,d
θ
the action of the subalgebra:
is not closed under the action of Aθ but it is closed under
{f ∈ ℓ1(Z2) : ∀p ∈ R[X, Y ]
lim
n,m→±∞
p(n, m)f (n, m) = 0}
of (ℓ1(Z2),∗θ,·∗), often referred to as the smooth quantum torus. We will not
use this observation later on, though it is notable that the completion of S p,q,d
is
indeed a Aθ-module.
θ
3. A continuous fields of C∗-Hilbert norms
θ
, as θ varies in R, form a continuous family.
All Heisenberg modules are completions of S(Cd) for some d ∈ N, d > 0. For
a fixed d, it thus becomes possible to ask whether the various C∗-Hilbert norms
k · kH p,q,d
To this end, we establish a succession of lemmas whose primary goal is to provide
us with estimates on the Heisenberg modules' C∗-Hilbert norms in terms of the
norm of ℓ1(Z2). While the Heisenberg modules' C∗-Hilbert norms are in general
delicate to work with as they involve the no-less abstract quantum tori norms,
the ℓ1(Z2) norm, which dominates all of the quantum tori norms, is much more
amenable to computations. For our purpose, we will take full advantage of the
regularity of Schwarz functions, which will enable us to apply various analytic tools
to derive our desired result.
The first step is a lemma which provides a first upper bound to the ℓ1(Z2) norm
of the difference between certain Heisenberg module inner products.
Lemma 3.1. If θ, ϑ ∈ R and p ∈ Z, q ∈ N \ {0}, d ∈ qN \ {0}, and if ω, η and ξ
are C2 functions from R to Cd such that for all f ∈ {ω, η, ξ}:
(1) all of f , f ′ and f ′′ are integrable on R,
(2) limt→±∞ f (t) = limt→±∞ f ′(t) = limt→±∞ f ′′(t) = 0,
then, writing ðθ = θ − p
q and ðϑ = ϑ − p
q , we have:
− hξ, ηiH p,q,d
1
ϑ
(cid:13)(cid:13)(cid:13)ℓ1(Z2)
θ
(cid:13)(cid:13)(cid:13)hω, ηiH p,q,d
6 Xn∈Z
4π2n2 ZR Xm∈Zkω′′(t + ðθm) − ξ′′(t + ðϑm)kCd kη(t)kCd dt+
+ 2ZR Xm∈Zkω′(t + ðθm) − ξ′(t + ðϑm)kCd kη′(t)kCd dt
D-NORMS ON HEISENBERG MODULES
11
+ ZR Xm∈Z
kω(t + ðθm) − ξ(t + ðϑm)kCd kη′′(t)kCd dt! .
Proof. We begin with the observation that for all (n, m) ∈ Z2 we have:
(n, m)
p,q,d ω(t + ðθm), η(t)ECd
p,q,d ξ(t + ðϑm), η(t)ECd
exp(2iπnt) dt
exp(2iπnt) dt
(ω(t + ðθm) − ξ(t + ðϑm)), η(t)ECd
exp(2iπnt) dt.
θ
ϑ
hω, ηiH p,q,d
(n, m) − hξ, ηiH p,q,d
=ZRDρ[n],[m]
−ZRDρ[n],[m]
=ZRDρ[n],[m]
For all n, m ∈ Z, the function:
fn,m : t 7→Dρ[n],[m]
p,q,d
p,q,d ω(t + ðθm) − ξ(t + ðϑm), η(t)ECd
has a first and continuous second derivative which are integrable, and:
lim
t→±∞
fn,m(t) = lim
t→±∞
f ′
n,m(t) = lim
t→±∞
f ′′
n,m(t) = 0.
We consequently may apply integration by part and obtain, for all m, n ∈ Z:
ZRDρ[n],[m]
p,q,d ω(t + ðθm) − ξ(t + ðϑm), η(t)ECd
exp(2iπnt) dt
=ZR
=ZR
fn,m(t) exp(2iπnt) dt = −ZR
f ′′
n,m(t)
exp(2iπnt)
4π2n2
dt.
f ′
n,m(t)
exp(2iπnt)
dt
2iπn
We compute trivially that for all t ∈ R and m, n ∈ Z:
(ω′′(t + ðθm) − ξ′′(t + ðϑm)), η(t)ECd
n,m(t) =Dρ[n],[m]
p,q,d
f ′′
+ 2Dρ[n],[m]
p,q,d
(ω′(t + ðθm) − ξ′(t + ðϑm)), η′(t)ECd
+Dρ[n],[m]
p,q,d
(ω(t + ðθm) − ξ(t + ðϑm)), η′′(t)ECd
.
Thus using Cauchy-Schwarz and since ρ[n],[m]
p,q,d
is a unitary, we conclude:
(cid:13)(cid:13)(cid:13)hω, ηiH p,q,d
θ
− hξ, ηiH p,q,d
ϑ
(cid:13)(cid:13)(cid:13)ℓ1(Z2)
12
FRÉDÉRIC LATRÉMOLIÈRE
dt
1
exp(2iπnt) dt(cid:12)(cid:12)(cid:12)(cid:12)
p,q,d
ZRDρ[n],[m]
n,m(t)(cid:12)(cid:12)
4π2n2
(ω(t + ðθm) − ξ(t + ðϑm)), η(t)ECd
4π2n2 (cid:18)ZR kω′′(t + ðθm) − ξ′′(t + ðϑm)kCdkη(t)kCd dt
= Xn,m∈Z(cid:12)(cid:12)(cid:12)(cid:12)
6 Xm,n∈ZZR (cid:12)(cid:12)f ′′
6 Xm,n∈Z
+ 2ZR kω′(t + ðθm) − ξ′(t + ðϑm)kCdkη′(t)kCd dt
+ZR kω(t + ðθm) − ξ(t + ðϑm)kCdkη′′(t)kCd dt(cid:19)
= Xn∈N
+ 2ZR Xm∈Nkω′(t + ðθm) − ξ′(t + ðϑm)kCd! kη′(t)kCd dt
+ZR Xm∈Nkω(t + ðθm) − ξ(t + ðϑm)kCd! kη′′(t)kCd dt# by Tonelli's theorem.
kω′′(t + ðθm) − ξ′′(t + ðϑm)kCd!kη(t)kCd dt
1
4π2n2 "ZR Xm∈N
This concludes our lemma.
(cid:3)
Our next lemma focuses on the type of estimates given in Lemma (3.1), and
gives a sufficient condition for these upper bounds to converge to 0 when various
parameters are allowed to converge to appropriate values.
Lemma 3.2. Let d ∈ N, d > 0. Let N = N∪{∞} be the one point compactification
of N.
If (ωk)k∈N and (ηk)k∈N are two families of C2-functions from R to Cd and
(ðk)k∈N is a sequence of nonzero real numbers converging to some ð∞ 6= 0 such
that:
(1) (t, k) ∈ R × N 7→ ωk(t) and (t, k) ∈ R × N 7→ ηk(t) are jointly continuous,
(2) there exists M > 0 such that for all k ∈ N and t ∈ R:
M
1 + t2 ,
max {kωk(t)kCd ,kηk(t)kCd} 6
then:
(3.1)
lim
k→∞Xn∈N
1
4π2n2 ZR Xm∈Z
kωk(t + ðkm) − ω∞(t + ð∞m)kCdkηk(t)kCd dt = 0.
Proof. First, we observe that Expression (3.1) is left unchanged if we replace ðk
with −ðk for all k ∈ N, thanks to the summation over m ∈ Z. Consequently, we
may assume without loss of generality that ð∞ > 0 and assume that ðk > 0 for
all k ∈ N (since (ðk)k∈N converges to ð∞ 6= 0, we must have that ðk and ð∞ have
the same sign for k larger than some K ∈ N; we thus can truncate our sequence to
start at K and flip all the signs if necessary to work with positive values).
With this in mind, since (ðk)k∈N is positive and converges to ð∞ > 0, there
exists 0 < ð− < ð+ such that for all k ∈ N, we have ðk ∈ [ð−, ð+].
D-NORMS ON HEISENBERG MODULES
13
We shall employ the Lebesgue dominated convergence theorem. To this end, we
introduce the following function to serve as our upper bound. For all t, m ∈ R we
set:
(3.2)
1+(t+mð−)2 if m > 0 and t > −ð−m, or if m < 0 and t 6 −ð−m,
M otherwise.
M
b(t, m) =(
For a fixed t ∈ R, we note that:
b(t, m) ∼m→±∞
M
−m2 ,
ð2
soPm∈Z b(t, m) < ∞. Moreover, by construction, for all t, m ∈ R and ð ∈ [ð−, ð+],
we have:
M
1 + (t + ðm)2
6 b(t, m).
Therefore, using our hypothesis, for all t ∈ R, m ∈ Z, k ∈ N and ð ∈ [ð−, ð+]:
kωk(t + mð) − ω∞(t + mð∞)kCd 6
1 + (t + mð)2 +
1 + (t + mð∞)2
M
M
Thus for a fixed t ∈ R, we may apply Lebesgue dominated convergence theorem
to conclude:
6 2b(t, m).
(3.3)
lim
k→∞ Xm∈Zkωk(t + mðk) − ω∞(t + mð∞)kCd = 0,
since (t, k) ∈ R × N 7→ ωk(t) is jointly continuous.
We now make another observation. For any fixed ð > 0 and k ∈ N, The function:
t ∈ R 7→ Xm∈Zkωk(t + ðm)kCd
is ð-periodic.
If t ∈ [0, ð+], k ∈ N and ð ∈ [ð−, ð+], then since:
kωk(t + ðm)kCd 6 sup
x∈[0,ð+]
b(x, m)
while, as can easily be checked:
sup
x∈[0,ð+]
b(x, m) ∼m→±∞
M
−m2 ,
ð2
we conclude that the series:
(cid:16)(t, k, ð) ∈ R × N × [ð−, ð+] 7→Xkωk(t + ðm)kCd(cid:17)m∈Z
converges uniformly to its limit on [0, ð+] × N × [ð−, ð+]. In particular:
(t, k, ð) ∈ [0, ð+] × N × [ð−, ð+] 7→ Xm∈Zkωk(t + ðm)kCd
is continuous on a compact domain and so it is bounded. Let C > 0 such that for
all (t, k, ð) ∈ [0, ð+] × N × [ð−, ð+], we have:
Xm∈Zkωk(t + ðm)kCd 6 C.
14
FRÉDÉRIC LATRÉMOLIÈRE
(3.5)
lim
2CM
1 + t2 .
1+t2 is integrable over R. Once again, we apply Lebesgue dominated
We conclude that t 7→ Pm∈Z kωk(t − ðkm)kCd is bounded by C on R, since it
is an ðk-periodic function with ðk 6 ð+, for all k ∈ N.
We thus have that for all t ∈ R and k ∈ N:
(3.4) Xm∈Zkωk(t + mðk) − ω∞(t + mð∞)kCdkηk(t)kCd 6 2Ckηk(t)kCd 6
Now t ∈ R 7→ 2CM
convergence theorem, and we conclude from Expression (3.3) that:
k→∞ZR Xm∈Z
Last, using Inequality (3.4) again, we note that for all k ∈ N:
kωk(t + mðk) − ω∞(t + mð∞)kCdkηk(t)kCd dt 6ZR
kωk(t + mðk) − ω∞(t + mð∞)kCdkηk(t)kCd dt = 0.
ZR Xm∈Z
and thus for all n ∈ Z and k ∈ N:
4π2n2 ZR Xm∈Zkω(t + mðk) − ω(t + mð∞)kCdkη(t)kCd dt 6
with Pn∈Z
k→∞Xn∈Z
2πn2 < ∞; hence we may apply Lebesgue dominated convergence the-
4π2n2 ZR Xm∈Zkωk(t + mðk) − ω∞(t + mð∞)kCdkηk(t)kCd dt = 0.
orem once more to conclude from Expression (3.5):
2CM π
4π2n2 =
CM
2πn2 ,
2CM
1 + t2
CM
1
1
lim
dt = 2CM π
This concludes our lemma.
(cid:3)
Remark 3.3. One may check that Lemma (3.1) and Lemma (3.2) together prove
qo,
that if p, q ∈ N, ξ, ω ∈ S(Cd), for any d ∈ qN with d > 0, and if θ ∈ R \n p
∈ ℓ1(Z2). It is a well-known fact (indeed a basic fact for the very
then hξ, ωiH p,q,d
construction of Heisenberg modules) though maybe not apparent from Theorem-
Definition (2.15) without consulting such sources as [24].
θ
We now bring together Lemma (3.1) and Lemma (3.2) to obtain a first result
of continuity on the Heisenberg module inner products, albeit using the ℓ1(Z2)
norm. This is the core result of this section, and it is phrased at a somewhat higher
level of generality that what is needed for the proof of continuity of the family of
Heisenberg C∗-Hilbert norms. Indeed, this level of generality will prove useful twice
later in this paper: when proving that the Heisenberg group representations αð,d
define strongly continuous actions on Heisenberg modules, and when establishing
that our prospective D-norms on Heisenberg modules will also form a continuous
family of norms in [20].
Lemma 3.4. Let p, q ∈ N with q > 0 and d ∈ qN with d > 0. If (ξk)k∈N is a
family of Cd-valued C2-functions over R such that:
(1) there exists M > 0 such that for all k ∈ N and t ∈ R:
M
1 + t2 ,
max{kξk(t)kCd,kξ′
k(t)kCd ,kξ′′
k (t)kCd} 6
(2) (t, k) ∈ R × N 7→ ξk(t) is continuous,
(cid:13)(cid:13)(cid:13)(cid:13)
6(cid:13)(cid:13)(cid:13)(cid:13)
(cid:13)(cid:13)(cid:13)(cid:13)
(3.7)
observe that:
θ∞ (cid:13)(cid:13)(cid:13)(cid:13)ℓ1(Z2)
θ∞ (cid:13)(cid:13)(cid:13)(cid:13)ℓ1(Z2)
+(cid:13)(cid:13)(cid:13)hξk, ξ∞iH p,q,d
θ∞ (cid:13)(cid:13)(cid:13)(cid:13)ℓ1(Z2)
=(cid:13)(cid:13)(cid:13)(cid:13)
θ∞ (cid:13)(cid:13)(cid:13)(cid:13)ℓ1(Z2)
hξk, ξkiH p,q,d
k (t + ðkm) − ξ′′
θ∞ (cid:13)(cid:13)(cid:13)ℓ1(Z2)
θ∞ (cid:13)(cid:13)(cid:13)(cid:13)ℓ1(Z2)
.
D-NORMS ON HEISENBERG MODULES
15
and if (θk)k∈N is a sequence converging to θ∞ such that θk − p
then we have:
q 6= 0 for all k ∈ N,
lim
k→∞(cid:13)(cid:13)(cid:13)(cid:13)
hξk, ξkiH p,q,d
θk − hξ∞, ξ∞iH p,q,d
= 0.
θ∞ (cid:13)(cid:13)(cid:13)(cid:13)ℓ1(Z2)
Proof. To fix notations, for all k ∈ N, we set ðk = θk − p
sequence of nonzero real numbers converging to ð∞ 6= 0.
We shall prove our result from the following inequality:
q . Note that (ðk)k∈N is a
(3.6)
hξk, ξkiH p,q,d
θk − hξ∞, ξ∞iH p,q,d
θk − hξk, ξ∞iH p,q,d
hξk, ξkiH p,q,d
We begin with the first term of the right hand side of Inequality (3.6). We
θ∞ − hξ∞, ξ∞iH p,q,d
.
θk − hξk, ξ∞iH p,q,d
hξk, ξkiH p,q,d
By Lemma (3.1), we then have for all k ∈ N:
θk − hξ∞, ξkiH p,q,d
θk − hξ∞, ξkiH p,q,d
(cid:13)(cid:13)(cid:13)(cid:13)
hξk, ξkiH p,q,d
6 Xn∈Z
1
4π2n2 ZR Xm∈Zkξ′′
+ 2ZR Xm∈Z
∞(t + ð∞m)kCd kξk(t)kCd dt+
∞(t + ð∞m)kCd kξ′
kξ′
k(t + ðkm) − ξ′
k(t)kCd dt
+ ZR Xm∈Zkξk(t + ðkm) − ξ∞(t + ð∞m)kCd kξ′′
k (t)kCd dt! .
Our assumptions allow us to apply Lemma (3.2) to each term in the right hand
side of Inequality (3.7) to conclude that:
lim
k→∞(cid:13)(cid:13)(cid:13)(cid:13)
hξk, ξkiH p,q,d
θk − hξk, ξ∞iH p,q,d
= 0.
θ∞ (cid:13)(cid:13)(cid:13)(cid:13)ℓ1(Z2)
We handle the second term of Inequality (3.6) in a similar manner.
From Inequality (3.6), our lemma is proven.
(cid:3)
We now conclude this section with the proof that indeed, Heisenberg C∗-Hilbert
norms form continuous families of norms for a fixed projective representation of
some Z2
q.
Proposition 3.5. Let p, q ∈ N and d ∈ qN with d > 0. Let (ξ)k∈N be a family
in S(Cd) such that (k, t) ∈ N × R 7→ ξk(t) is (jointly) continuous and there exists
M > 0 such that kξ(s)
1+t2 for all k ∈ N, t ∈ R and s ∈ {0, 1, 2}.
k (t)kCd 6 M
If (θk)k∈N is a sequence in R converging to θ∞ and such that θk − p
q = 0 for all
k ∈ N, then:
lim
k→∞ kξkkH p,q,d
θk
= kξ∞kH p,q,d
θ∞
.
θk
θk
hξk, ξkiH p,q,d
hξk, ξkiH p,q,d
hξk, ξkiH p,q,d
θk − hξ∞, ξ∞iH p,q,d
=(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
θ∞ (cid:13)(cid:13)(cid:13)Aθ∞(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:13)(cid:13)(cid:13)(cid:13)Aθk
(cid:13)(cid:13)(cid:13)(cid:13)
−(cid:13)(cid:13)(cid:13)hξ∞, ξ∞iH p,q,d
6(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
θ∞ (cid:13)(cid:13)(cid:13)Aθk(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:13)(cid:13)(cid:13)(cid:13)Aθk
(cid:13)(cid:13)(cid:13)(cid:13)
−(cid:13)(cid:13)(cid:13)hξ∞, ξ∞iH p,q,d
+(cid:12)(cid:12)(cid:12)(cid:12)
θ∞ (cid:13)(cid:13)(cid:13)Aθ∞(cid:12)(cid:12)(cid:12)(cid:12)
(cid:13)(cid:13)(cid:13)hξ∞, ξ∞iH p,q,d
θ∞ (cid:13)(cid:13)(cid:13)Aθk −(cid:13)(cid:13)(cid:13)hξ∞, ξ∞iH p,q,d
6(cid:13)(cid:13)(cid:13)(cid:13)
θ∞ (cid:13)(cid:13)(cid:13)(cid:13)Aθk
+(cid:12)(cid:12)(cid:12)(cid:12)(cid:13)(cid:13)(cid:13)hξ∞, ξ∞iH p,q,d
θ∞ (cid:13)(cid:13)(cid:13)Aθ∞(cid:12)(cid:12)(cid:12)(cid:12)
θ∞ (cid:13)(cid:13)(cid:13)Aθk −(cid:13)(cid:13)(cid:13)hξ∞, ξ∞iH p,q,d
6(cid:13)(cid:13)(cid:13)(cid:13)
θ∞ (cid:13)(cid:13)(cid:13)(cid:13)ℓ1(Z2)
+(cid:12)(cid:12)(cid:12)(cid:12)(cid:13)(cid:13)(cid:13)hξ∞, ξ∞iH p,q,d
θ∞ (cid:13)(cid:13)(cid:13)Aθ∞(cid:12)(cid:12)(cid:12)(cid:12)
θ∞ (cid:13)(cid:13)(cid:13)Aθk −(cid:13)(cid:13)(cid:13)hξ∞, ξ∞iH p,q,d
θ∞ (cid:13)(cid:13)(cid:13)(cid:13)ℓ1(Z2)
θ∞ (cid:13)(cid:13)(cid:13)Aθ∞(cid:12)(cid:12)(cid:12)(cid:12)
θ∞ (cid:13)(cid:13)(cid:13)Aθk −(cid:13)(cid:13)(cid:13)hξ∞, ξ∞iH p,q,d
θk − hξ∞, ξ∞iH p,q,d
hξk, ξkiH p,q,d
θk − hξ∞, ξ∞iH p,q,d
= 0.
= 0.
.
We now apply Lemma (3.4) to conclude that:
lim
hξk, ξkiH p,q,d
k→∞(cid:13)(cid:13)(cid:13)(cid:13)
k→∞(cid:12)(cid:12)(cid:12)(cid:12)(cid:13)(cid:13)(cid:13)hξ∞, ξ∞iH p,q,d
lim
16
FRÉDÉRIC LATRÉMOLIÈRE
Proof. For each k ∈ N ∪ {∞}, we set ðk = θk − p
q 6= 0.
We first compute:
(3.8)
(cid:12)(cid:12)(cid:12)(cid:12)
kξkk2
H p,q,d
θk − kξ∞k2
H p,q,d
θ∞ (cid:12)(cid:12)(cid:12)(cid:12)
Now, for any f ∈ ℓ1(Z2), the function θ ∈ R 7→ kfkAθ is continuous by [26,
Corollary 2.7]. Hence, using Remark (3.3):
Thus, we conclude from Inequality (3.8) that:
k→∞ kξkk2
lim
H p,q,d
θk
= kξ∞k2
H p,q,d
θ∞
which, by continuity of the square root, proves our lemma.
Corollary 3.6. Let p, q ∈ N and d ∈ qN with d > 0. Let ξ ∈ S(Cd). If (θk)k∈N is
a sequence in R converging to θ∞ and such that θk − p
q = 0 for all k ∈ N, then:
.
= kξkH p,q,d
k→∞ kξkH p,q,d
lim
(cid:3)
θ∞
θk
Proof. We apply Proposition (3.5) to the family k ∈ N 7→ ξ. We note that since ξ
is a Schwarz function, our assumptions are met.
(cid:3)
4. The action of the Heisenberg group on Heisenberg modules
Our goal in this paper is to prove that Heisenberg modules may be endowed
with a metrized quantum vector bundle structure over quantum 2-tori using a D-
norm built from a Lie group action and inspired by the construction of [28], albeit
involving a projective action of a locally compact group, which will not act via
D-NORMS ON HEISENBERG MODULES
17
isometries of the D-norm. These changes will introduce new difficulties which we
will handle in the next few sections. As a first step, we study the actions of the
Heisenberg group on Heisenberg modules.
One motivation for the results in this section is to establish the properties which
will meet the hypothesis of the main results in our next section, from which our
D-norm will emerge. We also note that the actions αð,d, for all ð ∈ R \ {0} and
d ∈ N\{0}, is a strongly continuous action by isometries of L2(R)⊗Cd, but we need
these results to be proven for the Heisenberg C∗-Hilbert norms, which dominate
the norm of L2(R) ⊗ Cd.
in this section, and thus we group them in the following.
We shall use the same hypotheses for a series of lemmas and our main definition
Hypothesis 4.1. Let p ∈ Z, q ∈ N \ {0}, and let d ∈ qN with d > 0. Let
θ ∈ R \n p
qo. We write ð = θ − p
We shall employ the notations of Theorem-Definition (2.15).
q .
We begin with two lemmas which will prove that H3 acts via isometries of the
norm of the Heisenberg modules on the subspace of Schwarz functions -- where we
have an explicit formula for our inner product -- and thus can indeed be extended
to the entire module.
Lemma 4.2. We assume Hypothesis (4.1). For all (x, y, u) ∈ H3, if z1 = exp (2iπðy)
and z2 = exp (−2iπðx), and if ξ, ω ∈ S p,q,d
ð,d (ω)EH p,q,d
(cid:16)hξ, ωiH p,q,d
Dαx,y,u
ð,d (ξ), αx,y,u
= βz1,z2
(cid:17) .
, then:
θ
θ
θ
θ
Proof. Let n, m ∈ Z. We compute:
Dαx,y,u
ð,d (ω)EH p,q,d
ð,d (ξ), αx,y,u
θ
(n, m)
ð,1 αx,y,u
p,q,ð,dαx,y,u
ð,d ξ, αx,y,u
ð,1 ⊗ ρ[n],[m]
ð,d ωEL2(R)⊗Cd
p,q,d (cid:17) ξ, αx,y,u
ð,1 ⊗ ρ[n],[m]
αx,y,u
ð,1 ⊗ ρ[n],[m]
=Dn,m
=D(cid:16)σn,m
=D(cid:16)α(x,y,u)−1
=Dexp(2iπð(yn − xm))(cid:16)σn,m
p,q,ð,dξ, ωEL2(R)⊗Cd
n,m, nm
2
ð,1
= zn
1 zm
ð,1
α
.
ð,d ωEL2(R)⊗Cd
p,q,d (cid:17) ξ, ωEL2(R)⊗Cd
p,q,d (cid:17) ξ, ωEL2(R)⊗Cd
= βz1,z2
θ
(cid:16)hξ, ωiH p,q,d
θ
(cid:17)
(cid:3)
Therefore, by definition of the dual action β:
2 Dn,m
ð,d (ω)EH p,q,d
θ
ð,d (ξ), αx,y,u
Dαx,y,u
as desired.
To ease our notations in this section, we set:
Notation 4.3. For all (x, y) ∈ R2 and ð > 0, we define:
υð(x, y) = (exp(2iπðy), exp(−2iπðx)) ∈ T2.
18
FRÉDÉRIC LATRÉMOLIÈRE
We now show that the Heisenberg group acts by isometries for the C∗-Hilbert
θ
θ
ð,d
,k · kH p,q,d
norm.
Lemma 4.4. We assume Hypothesis (4.1). For all (x, y, u) ∈ H3, the map αx,y,u
(cid:17).
is an isometry of (cid:16)H p,q,d
Proof. Let (x, y, u) ∈ H3 and ξ ∈ S p,q,d
=(cid:13)(cid:13)(cid:13)(cid:13)Dαx,y,u
=(cid:13)(cid:13)(cid:13)
=(cid:13)(cid:13)(cid:13)hξ, ξiH p,q,d
ð,d ξEH p,q,d
(cid:13)(cid:13)(cid:13)Aθ
hξ, ξiH p,q,d
(cid:13)(cid:13)(cid:13)Aθ
ð,d ξ, αx,y,u
(cid:13)(cid:13)(cid:13)(cid:13)Aθ
by Lemma (4.2),
. We compute:
kαx,y,u
βυr(x,y)
θ
ξk2
= kξk2
H p,q,d
H p,q,d
.
θ
ð
θ
θ
θ
θ
θ
This completes our proof.
(cid:3)
θ
θ
θ
ð,d
,k · kH p,q,d
may thus be extended to H p,q,d
Notation 4.5. We use the notations of Hypothesis (4.1). The action αð,d of H3 on
S p,q,d
for all
(x, y, u) ∈ H3; we shall keep the notation of this extension as αð,d. We note that it
also acts via isometry on (cid:16)H p,q,d
We also use the same notation for σð,d extended to (cid:16)H p,q,d
by extending by continuity αx,y,u
The actions of the Heisenberg group on Heisenberg modules is by morphism
modules, in the sense of [16, Definition 3.5]. This result will play a role in the proof
that our D-norm satisfies the modular version of the Leibniz inequality.
Lemma 4.6. We assume Hypothesis (4.1). For all a ∈ Aθ, ξ ∈ H p,q,d
(x, y, u) ∈ H3, then:
,k · kH p,q,d
(cid:17).
(cid:17).
and
θ
θ
θ
θ
αx,y,u
ð,d
(aξ) = βυð(x,y)
θ
(a)αx,y,u
ð,d (ξ).
Proof. Let n, m ∈ Z and ξ ∈ S p,q,d
θ
and fm,m ∈ ℓ1(Z2) be defined by:
fn,m : (z, w) ∈ Z2 7−→(1 if n = z and m = w,
0 otherwise.
We compute:
αx,y,u
ð,d (fn,mξ) = αx,y,u
ð,d n,m
p,q,ð,dξ
n,m, nm
2
ð,d
ð,d α
p,q,d (cid:17) ξ
=(cid:16)αx,y,u
⊗ ρ[n],[m]
= exp(2iπð(yn − xm))(cid:16)α
= exp(2iπð(yn − xm))n,m
= βυð(x,y)
(fn,m)αx,y,u
ð,d ξ.
θ
n,m, nm
2
ð,d
p,q,ð,d(cid:17) ξ
ð,d ⊗ ρ[n],[m]
αx,y,u
p,q,ð,dαx,y,u
ð,d ξ
Since βθ is an action by *-morphisms, we conclude that for all a ∈ Aθ:
(4.1)
ð,d (aξ) = βυð(x,y)
αx,y,u
θ
(a)αx,y,u
ð,d (ξ)
D-NORMS ON HEISENBERG MODULES
19
as desired. The lemma is concluded by extending Equality (4.1) to H p,q,d
continuity.
θ
by
(cid:3)
An important corollary of Lemma (4.6) is as follows:
Corollary 4.7. We assume Hypothesis (4.1). For all a ∈ Aθ, ξ ∈ H p,q,d
(x, y, u) ∈ H3, we observe that:
αx,y,u
θ
6 kakAθkξkH p,q,d
θ
.
Proof. Let a ∈ Aθ, ξ ∈ H p,q,d
θ
θ
(cid:13)(cid:13)(cid:13)
ð,d (aξ)(cid:13)(cid:13)(cid:13)H p,q,d
ð,d (aξ)(cid:13)(cid:13)(cid:13)H p,q,d
=(cid:13)(cid:13)(cid:13)
6 kβυð(x,y)
θ
θ
αx,y,u
(cid:13)(cid:13)(cid:13)
and (x, y, u) ∈ H3. We compute:
βυð(x,y)
θ
(a)αx,y,u
ð,d ξ(cid:13)(cid:13)(cid:13)H p,q,d
θ
akAθkξkH p,q,d
θ
by Lemma (4.6),
by Lemma (4.4).
This completes our proof.
and
(cid:3)
We have checked that the actions of the Heisenberg group on Heisenberg modules,
which the latter were constructed from, act by isometric module morphisms on the
entire module. Note that we already observed that Heisenberg modules can be
regarded as dense subspaces of L2(R) ⊗ Cd spaces on which the same action of the
Heisenberg group is defined, strongly continuous and isometric; however we needed
to ensure that these actions are well-behaved with respect to the inner product and
norm of the Heisenberg modules.
In order to define our D-norms, we shall require one more important analytic
property: we want our actions to be strongly continuous for the Heisenberg C∗-
Hilbert norms. This is the subject of the next proposition. We actually include
in the next proposition a somewhat more general hypothesis and estimate than
needed for the strong continuity of our actions, as this stronger statement will play
an important role in our study of the continuity properties of our D-norms later on
in [20].
Proposition 4.8. Let p ∈ Z, q ∈ N \ {0} and d ∈ qN with d > 0. Let C > 0 and
M > 0 some constant. Let 0 < ð− < ð+. There exists K > 0 such that for all
ξ ∈ S (Cd) satisfying:
(4.2) max{kξ(s)kCd,ksξ(s)kCd,kξ′(s)kCd ,ksξ′(s)kCd ,
kξ′′(s)kCd,ksξ′′(s)kCd} 6
M
1 + s2 ,
the following holds for all s ∈ R, ð ∈ [ð−, ð+] and (x, y, u) ∈ R3 with x+y+u 6
C:
: n ∈ {0, 1, 2}o 6
K(x + y + u)
1 + s2
.
(4.3)
maxn(cid:13)(cid:13)(cid:13)
αx,y,u
ð,d ξ(n)(s) − ξ(n)(s)(cid:13)(cid:13)(cid:13)Cd
In particular, for all ð 6= 0 and θ = ð + p
q :
αx,y,u
lim
(x,y,u)→0(cid:13)(cid:13)(cid:13)
ð,d ξ − ξ(cid:13)(cid:13)(cid:13)H p,q,d
θ
= 0.
20
FRÉDÉRIC LATRÉMOLIÈRE
Proof. Let ξ ∈ S(Cd) and (x, y, u) ∈ R3. We note that for all s ∈ R, using the
continuity of ξ, we of course have:
αx,y,u
ð,d ξ(s) − ξ(s) = exp(2iπ(u + xs))ξ(s + ðy) − ξ(s)
(x,y,u)→0
−−−−−−→ 0.
However, we wish to apply Lemma (3.4) to obtain convergence in norm, so we
seek a more precise estimate. To this end, let:
fs(t) = αtx,ty,tu
ð,d
ξ(s) = exp(2iπ(ðtu + txs))ξ(s + ðty)
for all t, s ∈ R. We compute for all t, s ∈ R:
f ′
s(t) = exp(2iπ(ðtu + txs)) (2iπ(ðu + xs)ξ(s + ðty) + ðyξ′(s + ðty)) .
Let k(x, y, u)k1 = x +y +u for all (x, y, u) ∈ R2, i.e. k·k1 is the usual 1-norm
on R3. Let us now assume k(x, y, u)k1 6 C -- in particular, y < C. We observe
that for all s ∈ R, using the function b introduced in Expression (3.2) in the proof
of Lemma (3.2):
αx,y,u
(cid:13)(cid:13)(cid:13)
0
f ′
s(t) dt(cid:13)(cid:13)(cid:13)(cid:13)Cd
ð,d ξ(s) − ξ(s)(cid:13)(cid:13)(cid:13)Cd
Z 1
= kfs(1) − fs(0)kCd =(cid:13)(cid:13)(cid:13)(cid:13)
6Z 1
0 kexp(2iπ(ðtu + txs)) (2iπ(ðu + xs)ξ(s + ðty) + ðyξ′(s + ðty))kCd dt
=Z 1
0 k2iπ(ðu + xs)ξ(s + ðty) + ðyξ′(s + ðty)kCd dt
0 k(u, x, y)k1 max
6Z 1
6 2π max{1, ð+}k(x, y, u)k1Z 1
b(s, y)! .
6 2π max{1, ð+}k(x, y, u)k1 sup
k2iπðξ(s + ðty)kCd ,
k2iπsξ(s + ðty)kCd ,
kðξ′(s + ðty)kCd
b(s, ty) dt
y∈[−C,C]
dt
0
Since:
(4.4)
lim
s→±∞
(1 + s2)
sup
b(s, y) = M ,
y∈[−C,C]
we conclude that there exists R > 0 such that for all s ∈ R \ [−R, R], we have:
kαx,y,u
ð,d ξ(s) − ξ(s)kCd 6
M1k(x, y, u)k1
1 + s2
for M1 = 4M π max{1, ð+}. We note that M1 depends only on M , ð+ and C
through Expression (4.4), and not on ξ.
1+s2 is continuous and strictly positive, we may adjust M1 to a
Since s ∈ R 7→ 1
larger value if necessary such that:
min
s∈[−R,R]
M1
1 + s2
> 2πM max{1, ð+}.
D-NORMS ON HEISENBERG MODULES
21
Therefore, we have, for all s ∈ R and (x, y, u) ∈ R3 with k(x, y, u)k1 6 C:
kαx,y,u
ð,d ξ(s) − ξ(s)kCd 6
M1k(x, y, u)k1
1 + s2
6
M1C
1 + s2 .
Now, all the above computations may be applied equally well to ξ′ and ξ′′. We
conclude that indeed, Expression (4.3) holds as stated.
Let now ξ ∈ S ⊗ Cd be chosen. Since ξ is a Schwarz function, there exists M > 0
such that for all s ∈ R, we have:
max {kξ(s)kCd ,ksξ(s)kCd,kξ′(s)kCd ,ksξ′(s)kCd,kξ′′(s)kCd ,ksξ′′(s)kCd}
6
M
1 + s2 .
Thus we can apply our previous work to conclude that Expression (4.3) holds for
some K > 0, having chosen C = 1 for this last part of our proof.
Furthermore, we can apply now Lemma (3.4). For this part, we pick ð > 0; we
need not to worry about the uniformity in ð (we may as well assume ð− = ð+ = ð
here). Thus, if (xn, yn, un)n∈N converges to 0, Lemma (3.4) implies that:
ð,d
ξ − ξkH p,q,d
θ
6rkDαxn,yn,un
ð,d
0 6 kαxn,yn,un
limn→∞rkDαxn,yn,un
To prove our result for a general ð 6= 0, we simply observe that for all (x, y, u) ∈
and thus our proposition is completely proven. (cid:3)
ξ − ξ, αxn,yn,un
kℓ1(Z2) = 0
which concludes the proof of our proposition for ð > 0.
ξ − ξEH p,q,d
ξ − ξ, αxn,yn,un
R3 we have αx,y,u
ð,d = αx,−y,−u
−ð,d
ξ − ξEH p,q,d
θ
ð,d
ð,d
ð,d
θ
kℓ1(Z2)
We wish to use the actions of H3 on Heisenberg modules to define our D-norms.
The next section presents a general source of possible D-norms from actions of Lie
groups satisfying the properties we have established in this section.
5. Seminorms from Lie group actions
Connes introduced a quantized differential calculus on quantum tori in [5] using
the dual action of the tori, using the Lie group structure of the tori. Moreover,
he introduced a noncommutative connection on Heisenberg modules, and these
connections proved to be solutions of the Yang-Mills problem for quantum 2-tori
[7]. These connections were also useful in Rieffel's work on the classification of
modules over quantum tori [24].
Moreover, ergodic actions of metric compact groups on C*-algebras were the first
example of L-seminorms constructed by Rieffel in [28]. In this section, we begin
investigating how to build D-norms from Lie group actions. We will employ as
assumptions the properties which we derived for the action of the Heisenberg group
on Heisenberg modules. Our construction, as we shall see, lies at the intersection
of the purely metric picture of Rieffel and the differential picture of Connes, and is
a noncommutative version of [16, Example 3.10].
Our D-norm will be constructed using the following theorem.
Definition 5.1. Let α be a strongly continuous action of a Lie group G on a
Banach space E . Let w be a nonzero subspace of the Lie algebra of G. An element
22
FRÉDÉRIC LATRÉMOLIÈRE
ξ ∈ E is α-differentiable with respect to w when for all X ∈ w, the limit:
X(ξ) = lim
t→0
αexp(tX)ξ − ξ
t
exists.
In any vector space E, and for any function f : E → R, we denote as usual:
lim sup
x→0
f (x) = inf
δ>0
sup{f (x) : 0 < kxk 6 δ} .
Theorem 5.2. Let α be a strongly continuous action by linear isometries of a Lie
group G on a Banach space E . Let g be the Lie algebra of G and let h ⊆ g be a
nonzero subspace of g.
Let S ⊆ E be the subspace of E consisting of α-differentiable elements of E with
Let k · k be a norm on h. For all ξ ∈ S , the norm of the linear map:
respect to h. We note that S is dense in E .
∇ξ : X ∈ h 7→ ∇X ξ = X(ξ)
is denoted by ∇ξ.
If ξ ∈ S , then, for any δ > 0:
∇ξ = sup((cid:13)(cid:13)αexp(X)ξ − ξ(cid:13)(cid:13)E
= sup((cid:13)(cid:13)αexp(X)ξ − ξ(cid:13)(cid:13)E
X→0 (cid:13)(cid:13)αexp(X)ξ − ξ(cid:13)(cid:13)E
= lim sup
kXk
kXk
kXk
: X ∈ h \ {0})
: X ∈ h \ {0},kXk 6 δ)
.
Proof. A smoothing argument [4] proves that the set:
(cid:26)ξ ∈ E : t > 0 7→
αexp(tX)ξ − ξ
t
has a limit at 0 for all X ∈ g(cid:27)
is dense in E . Therefore, since S contains this set, S is dense in E as well.
Fix ξ ∈ S . Let X ∈ h. We define:
F : t ∈ R 7→ αexp(tX)ξ.
The function F is continuously differentiable, and in particular, F (0) = ξ and
F (1) = αexp(X)ξ.
Moreover, using the fact that t ∈ R 7→ exp(tX) is a continuous group homomor-
phism:
F ′(t) = lim
s→0
αexp((t+s)X)ξ − αexp(tX)ξ
h
= lim
s→0
αexp(tX)(cid:0)αexp(hX)ξ − ξ(cid:1)
h
= αexp(tX)∇X ξ.
Thus:
αexp(X)ξ − ξ =Z 1
0
F ′(t) dt =Z 1
0
αexp(tX) (∇X ξ) dt
D-NORMS ON HEISENBERG MODULES
23
so that:
(cid:13)(cid:13)αexp(X)ξ − ξ(cid:13)(cid:13)E
kXk
This proves that:
αexp(X) (∇X ξ)(cid:13)(cid:13)(cid:13)E
dt
0 ∇ξkXk dt = ∇ξ.
1
1
6
=
= (cid:13)(cid:13)(cid:13)R 1
0 F ′(t) dt(cid:13)(cid:13)(cid:13)E
kXk
kXkZ 1
0 (cid:13)(cid:13)(cid:13)
kXkZ 1
kXkZ 1
sup((cid:13)(cid:13)αexp(X)ξ − ξ(cid:13)(cid:13)E
kXk
6
1
: X ∈ h \ {0}) 6 ∇ξ.
0 k∇X ξkE dt since αexp(tX) is an isometry by hypothesis,
On the other hand, let us now fix some δ > 0. let us now assume that kXk = 1.
We first note that:
∇X ξ = F ′(0)
= lim
t→0
where lim is used for the topology of (E ,k · kE ),
F (t) − F (0)
αexp(tX)ξ − ξ
t
tkXk
= lim
t→0
= lim
t→0
αexp(tX)ξ − ξ
.
ktXk
Thus for all X ∈ h with kXk = 1, since ktXk 6 δ for all t ∈ R with t < δ:
and thus:
We have thus concluded our argument, as the function:
: Y ∈ h \ {0},kY k 6 δ)
: X ∈ h \ {0},kXk 6 δ)
: X ∈ h \ {0}) .
kY k
k∇X ξk 6 sup((cid:13)(cid:13)αexp(Y )ξ − ξ(cid:13)(cid:13)E
∇ξ 6 sup((cid:13)(cid:13)αexp(X)ξ − ξ(cid:13)(cid:13)E
6 sup((cid:13)(cid:13)αexp(X)ξ − ξ(cid:13)(cid:13)E
δ ∈ (0,∞) 7→ sup((cid:13)(cid:13)αexp(X)ξ − ξ(cid:13)(cid:13)E
kXk
kXk
kXk
: X ∈ h \ {0},kXk 6 δ)
(cid:3)
has been shown to be constant.
We note that the seminorms constructed in Theorem (5.2) include Rieffel's L-
seminorms in [28] from actions of compact Lie groups.
Corollary 5.3. Let α be a strongly continuous action by linear isometries of a
compact connected Lie group G on a Banach space E . As a compact Lie group, G
admits an Ad-invariant inner product h·,·ig on g. Let k · k be the norm associated
with h·,·ig. For any g ∈ G, since G is connected and compact, we may define ℓ(g)
as the distance from 1G to g for the Riemannian metric induced by h·,·ig.
24
FRÉDÉRIC LATRÉMOLIÈRE
If ξ ∈ S then:
sup(cid:26)kαgξ − ξkE
ℓ(g)
: g ∈ G \ {1G}(cid:27) = ∇ξ.
Proof. As G is a compact group, it admits a right Haar probability measure µ. Let
h·,·i be any inner product on g. If we set, for all X, Y ∈ g:
hX, Y iG =ZG hAdgX, AdgY i dµ(g)
then one easily verifies that h·,·iG is an Ad-invariant inner product on g.
Now, we endow G with the Riemannian metric induced by left translation of the
inner product h·,·iG. As this metric is induced by an Ad-invariant inner product,
it is in fact right invariant as well.
In particular, G, as a connected compact Riemannian manifold, is geodesically
complete by Hopf-Rinow theorem. As a first application, we let ℓ(g) be the dis-
tance from 1G to g in G for this Riemannian metric, for all g ∈ G. As a second
application, we note that the Riemannian exponential map of G for our metric is
indeed surjective.
It is now possible to check that the exponential map for the Lie group G and
the exponential map for the Riemannian metric coincide. This is done by checking
that the Riemannian exponential map defines a 1-parameter subgroup of G.
With this in mind, we conclude that for all X ∈ g, we have:
ℓ(exp(X)) = inf {kY k : exp(X) = exp(Y )} .
We note that the Lie exponential map is certainly not injective, at least as long as
G is of dimension at least one, though this does not affect our conclusion.
Moreover, since G is a compact connected Lie group, exp is surjective since the
Riemannian exponential is surjective. Thus, our corollary is proven using Theorem
(5.2).
(cid:3)
Now, Rieffel proved in [28] that the obvious necessary condition for a seminorm
of the type given in Corollary (5.3) to be a L-seminorm is, remarkably, sufficient as
well. This fact is highly non-trivial as well, and we record it here as it will be the
source of quantum metrics we put on quantum tori.
Theorem 5.4 ([28, Theorem 1.9]). Let β be a strongly continuous group action
by *-automorphisms of a compact group G on a unital C*-algebra A. Let ℓ be a
continuous length function on G. For all a ∈ A, we define:
L(a) = sup(cid:26)kβg(a) − akA
ℓ(g)
: g ∈ G \ {e}(cid:27) ,
allowing for this quantity to be infinite. Then the following are equivalent:
(1) (A, L) is a quantum compact metric space (which is necessarily Leibniz),
(2) {a ∈ A : ∀g ∈ G βg(a) = a} = C1A.
We note that the proof of Theorem (5.4) involves explicitly the fact that the
spectral subspaces of the action β are finite dimensional under the condition of
ergodicity [9]. This result is not trivial, and worse yet for our purpose, does not carry
to locally compact group. In fact, besides the trivial representation, no irreducible
representation of the Heisenberg group is finite dimensional -- so we are as far as
we can to apply the idea in [28]. In this paper, we shall focus on the Heisenberg
D-NORMS ON HEISENBERG MODULES
25
modules, and we will prove in this case that the seminorms constructed in Theorem
(5.2) have compact unit balls using quite different techniques from Rieffel.
The rest of this section introduces the general scheme to construct D-norms
from Lie group actions which we will employ in this paper, and prove that this
construction meets all our requirements except, maybe, for the compactness of the
unit ball which, in the case of Heisenberg modules, will be the subject of our next
section.
Proposition 5.5. Let β be the action of a compact connected Lie group G on a
unital C*-algebra A via *-automorphisms. Let α be the action by isometric C-linear
isomorphisms of a Lie group H on a Hilbert module (M ,h·,·iM ) over A. We write g
and h the respective Lie algebras of G and H, and expG : g → G and expH : h → H
be the respective Lie exponential maps of G and H.
Let w be a nonzero subspace of h. Let k · k♭ be a norm on g and k · k♯ be a norm
We set for all a ∈ A:
on w ⊆ h.
and for all ξ ∈ E :
: X ∈ g \ {0}) ,
kXk♭
L(a) = sup((cid:13)(cid:13)βexp(X)a − a(cid:13)(cid:13)A
D(ξ) = sup(kξkM ,(cid:13)(cid:13)αexp(X)ξ − ξ(cid:13)(cid:13)M
: X ∈ w \ {0}) .
If there exist two linear maps j : w → g and q : g → w such that:
(1) for all ξ, ω ∈ M and X ∈ w:
βexpG(X)hξ, ωiM =DαexpH (j(X))ξ, αexpH (j(X))ωEE
kXk♯
(5.1)
and:
(5.2)
αexpH (X)(aξ) = βexpG(q(X))(a)αexpH (X)ξ,
(2) j is an isometry from (g,k · k♭) to (w,k · k♯),
(3) q is a surjection of norm at most 1, i.e. kq(X)k♭ 6 kXk♯ for all X ∈ w,
then:
(1) L is a seminorm on a dense subspace of (A,k · kA), and moreover:
L(a) = 0 ⇐⇒ ∀g ∈ G βg(a) = a,
(2) D is a norm on a dense subspace of (M ,h·,·iM ) and D(·) > k · kM ,
(3) L and D are lower semicontinuous,
(4) for all a ∈ A and ξ ∈ M :
D(aξ) 6 kakAD(ξ) + L(a)kξkM ,
(5) for all ξ, ω ∈ M :
L (hξ, ωiM ) 6 kξkM D(ω) + D(ξ)kωkM .
Proof. Let Sg(A) be the subspace of A consisting of all the β-differentiable elements
with respect to g, and Sh(M ) be the subspace of M consisting of all the α-
differentiable elements of M with respect to w.
26
FRÉDÉRIC LATRÉMOLIÈRE
For any a ∈ Sg(A), we define the linear map ∂a : X ∈ g 7→ X(a) whose norm
is denoted by ∂ag
A, where g is endowed with k · k♭. Since g is finite dimensional,
∂a is continuous and thus has finite norm for all a ∈ Sg(A).
For any ξ ∈ Sw(M ), we also define ∇ξ : X ∈ w 7→ X(ξ) whose norm is ∇ξw
where w is endowed by k · k♯ -- since w is finite dimensional, the norm of ∇ξ is
finite as well.
M
By Theorem (5.2), for all a ∈ Sg(A) and for all ξ ∈ Sw(M ), then:
L(a) = ∂ag
A < ∞ and D(ξ) = ∇ξw
M < ∞.
Since Sg(A) and Sw(E ) are dense, we conclude that the domains of L and D are
indeed dense.
Since D(·) > k · kM by construction, D is in particular a norm on its domain.
Moreover if L(a) = 0 for some a ∈ A, we immediately conclude that βga = a for
all g ∈ G since the exponential map of G is surjective.
The function ξ ∈ M 7→ αexp(X)ξ−ξ
is continuous for all X ∈ w \ {0} and thus D
is lower semi-continuous as the pointwise supremum of continuous functions. The
same reasoning and conclusion applies to L.
kXk♯
We are left to prove the two forms of the Leibniz inequalities, which can be easily
checked by direct computation. Let ξ, ω ∈ M . We compute:
L (hξ, ωiE ) = sup((cid:13)(cid:13)βexp(X)hξ, ωiE − hξ, ωiE(cid:13)(cid:13)A
kXk♭
: X ∈ g \ {0})
: X ∈ w \ {0})
: X ∈ w \ {0})
kXk♯
kXk♯
kj(X)k♯
: X ∈ g \ {0})
= sup((cid:13)(cid:13)(cid:10)αexp(j(X))ξ, αexp(j(X))ω(cid:11)E − hξ, ωiE(cid:13)(cid:13)A
6 sup((cid:13)(cid:13)(cid:10)αexp(X)ξ, αexp(X)ω(cid:11)E − hξ, ωiE(cid:13)(cid:13)A
6 sup((cid:13)(cid:13)(cid:10)αexp(X)ξ, αexp(X)ω(cid:11)E −(cid:10)αexp(X)ξ, ω(cid:11)E(cid:13)(cid:13)A
+ sup((cid:13)(cid:13)(cid:10)αexp(X)ξ, ω(cid:11)E − hξ, ωiE(cid:13)(cid:13)A
: X ∈ w \ {0})
: X ∈ w \ {0})
6 sup((cid:13)(cid:13)αexp(X)ξ(cid:13)(cid:13)M (cid:13)(cid:13)αexp(X)ω − ω(cid:13)(cid:13)E
+ sup((cid:13)(cid:13)αexp(X)ξ − ξ(cid:13)(cid:13)E
6 kξkM sup((cid:13)(cid:13)αexp(X)ω − ω(cid:13)(cid:13)E
+ sup((cid:13)(cid:13)αexp(X)ξ − ξ(cid:13)(cid:13)E
: X ∈ w \ {0})kωkM
: X ∈ w \ {0})
: X ∈ w \ {0})kωkM
kXk♯
kXk♯
kXk♯
kXk♯
kXk♯
= kξkM
D(ω) + D(ξ)kωkM .
D-NORMS ON HEISENBERG MODULES
27
Now, let a ∈ A and ξ ∈ M . We compute:
: X ∈ w \ ker q)
kXk♯
kXk♯
: X ∈ w \ {0})
: X ∈ w \ {0})
sup ((cid:13)(cid:13)αexp(X) (aξ) − aξ(cid:13)(cid:13)M
= sup((cid:13)(cid:13)βexp(q(X))(a)αexp(X) (ξ) − aξ(cid:13)(cid:13)M
6 sup((cid:13)(cid:13)βexp(q(X))(a)αexp(X) (ξ) − aαexp(X)ξ(cid:13)(cid:13)M
+ sup((cid:13)(cid:13)aαexp(X) (ξ) − aξ(cid:13)(cid:13)M
: X ∈ w \ {0})
: X ∈ g \ {0})kξkM + kakAD(ξ)
6 sup((cid:13)(cid:13)βexp(q(X))(a) − a(cid:13)(cid:13)M
kq(X)k♭
kXk♯
kXk♭
= L(a)kξkM + kakAD(ξ),
as desired.
(cid:3)
Thus, Proposition (5.5) shows that if we follow the scheme suggested by Theorem
(5.2), then we obtain potential D-norms on modules. The missing property is the
compactness of the closed unit ball for the D-norm candidate.
We conclude our section by connecting our metric framework with the noncom-
mutative differential framework of connections on modules. Let us use the notations
of Proposition (5.5). A direct computation shows that for all X ∈ w, the following
holds:
(5.3)
∇X (aξ) = q(X)a · ξ + a∇X ξ
while for all X ∈ g, we also have:
(5.4)
X(hξ, ωiM ) = hj(X)ξ, ωiM + hξ, j(X)ωiM .
We also denote A ⊗ g∗ by Ω1 and the space of β-differentiable elements of A by
A1. We define ∂ : A1 → Ω1 by setting, for all a ∈ A1:
∂a : X ∈ g 7→ X(a).
We observe trivially that Ω1 is an A-A-bimodule and that ∂ is a derivation, i.e.
∂(ab) = a∂(b) + ∂(a)b for all a, b ∈ A1.
We first note that to get an interesting connection, we want q to be injective, i.e.
g and w to be isomorphic. It is always possible to increase the dimension of g (the
Lie algebra structure is actually not involved in the computations to follow, so this
is always possible), but this would amount to define ∂X = 0 for all vector X not in
g, and this is rather awkward and artificial.
Since, for the differential picture, the norms k · k♭ and k · k♯ do not play a role
in the construction of the connection, we will for now identify g and w and j and
q with the identity map.
With this assumption, Expressions (5.3) translates to the operator ∇ : M →
M ⊗ g∗, defined by:
∇(ξ) : X ∈ g 7→ ∇X ξ
28
FRÉDÉRIC LATRÉMOLIÈRE
for all α-differentiable ξ ∈ M with respect to g, to be a noncommutative connection.
We indeed easily check that for all a ∈ A and ξ ∈ M :
∇(aξ) = a∇(ξ) + ∂(a)ξ.
Expression (5.4) means that the connection ∇ is hermitian, i.e. it is compatible
with the noncommutative equivalent of a metric on the quantum vector bundle
M . It is tempting to call ∇ a Levi-Civita connection, although we do not address
here the computation of the torsion of ∇. Nonetheless, we see that our structure
provides a noncommutative Riemannian geometry. This is the structure which
inspired our definition of metrized quantum vector bundle, and we now can see
how it is implemented through our main example.
In summary, we have constructed a natural D-norm candidate on modules carry-
ing certain Lie group actions. The key difficulty, of course, regards the compactness
of the unit ball of such a D-norm.
6. A D-norm from a connection on Heisenberg modules
We now define our D-norms on Heisenberg modules. Our method employs the
idea of Theorem (5.2) and Proposition (5.5), where the actions of the Heisenberg
group on Heisenberg modules defines a norm which restricts to the operator norm
of a connection constructed via the associated action of the Heisenberg Lie algebra.
As noted at the end of the previous section, we want to only work with a subspace
of the Heisenberg Lie algebra to build our D-norm and its associated connection,
since the central element of the Heisenberg Lie algebra does not act, so to speak,
as a derivation -- it simply acts by multiplication by a scalar. We follow a pattern
which is common in the literature on the Heisenberg group: we only consider the
action of the subspace span{P, Q} in the Lie algebra H.
We thus endow span{P, Q} with a norm. If we were to construct a metric on
the Heisenberg group using this data -- by defining the length of a curve whose
tangent vector at (almost) every point lies in span{P, Q} in the usual manner by
integrating the norm of the tangent vector along the curve, and then defining the
distance between two points as the infimum of the length of all so-called horizontal
curves -- we would actually obtain a sub-Finslerian metric (if our choice of norm
comes from a Hilbert space structure, we would have a sub-Riemannian structure
and our construction would give rise to a Carnot-Carathédory distance on the
Heisenberg group).
However, as discussed, we do not transport the Carnot-Carathédory metric from
the Heisenberg group via its action in this paper. We prefer to carry the norm of the
subspace span{P, Q} of the Heisenberg Lie algebra to our modules. This approach
means that we work with a connection, and seems more natural. In essence, the
Carnot-Caratheodory is the metric obtained on the group while our D-norms are
the quantum metrics obtained on our modules; as the acting group is not compact,
we have no reason to expect them to agree.
With this in mind, we now introduce:
D-NORMS ON HEISENBERG MODULES
29
Definition 6.1. Let p ∈ Z, q ∈ N \ {0} and d ∈ qN with d > 0. Let θ ∈ R \n p
qo.
Let k·k be a norm on R2. We endow the Heisenberg module H p,q,d
: (x, y) ∈ R2 \ {0}
(ξ) = sup
ξ − ξ(cid:13)(cid:13)(cid:13)H p,q,d
2πðk(x, y)k
kξkH p,q,d
with the norm:
,(cid:13)(cid:13)(cid:13)
expH3
ð,d
Dp,q,d
(xP +yQ)
α
θ
θ
θ
θ
where ð = θ − p
q .
We now lighten our notation for the rest of our paper.
θ
simply by Dp,q,d
Convention 6.2. We endow R2 with a fixed norm k·k for the rest of this paper. We
shall denote Dk·k,p,q,d
, as the norm on R2 will not be understood.
We emphasize that k · k is independent of any of the parameters p, q, d and θ.
The norm k · k on R2 provides us with a continuous length function on Aθ for
all θ ∈ R. This length function arises from the invariant Finslerian metric induced
by k · k. A direct computation simply shows that:
θ
ℓ(exp(ix), exp(iy)) = inf{k(x + 2nπ, y + 2mπ)k : n, m ∈ Z2}.
For all θ ∈ R, we denote by Lθ the L-seminorm on Aθ associated with the action
βθ on Aθ and the length function ℓ via [28, Theorem 1.9]. We note that since T2
is compact and Abelian, Corollary (5.3) implies that for all a ∈ Aθ:
: (x, y) ∈ R2 \ {0})
Lθ(a) = sup(kβ
ξ − ξkAθ
expT2 (x,y)
θ
k(x, y)k
and Lθ agrees with the operator norm of derivative for the natural differential
calculus defined by βθ on βθ-differentiable elements. We refer to the previous section
for a discussion of these matters.
We begin by listing various equivalent expressions for our D-norm candidates,
as we shall use whichever may prove useful in this paper.
Remark 6.3. We recall from Notation (2.12) that:
expH3 (xP + yQ) =(cid:18)x, y,
1
2
xy(cid:19)
for all x, y ∈ R.
identities hold:
For all p, q ∈ N, d ∈ qN with d > 0, θ ∈ R\{pq−1} and ξ ∈ H p,q,d
θ
, the following
θ
θ
α
2 xy
x,y, 1
ð,d
Dp,q,d
kξkH p,q,d
2πðk(x, y)k
: (x, y) ∈ R2 \ {0}
: (x, y) ∈ R2 \ {0}
Proposition 6.4. Let p, q ∈ N and d ∈ qN with d > 0. Let θ ∈ R \n p
qo.
ξ − ξ(cid:13)(cid:13)(cid:13)H p,q,d
ð,d ξ − ξ(cid:13)(cid:13)(cid:13)H p,q,d
(ξ) = sup
= sup
,(cid:13)(cid:13)(cid:13)
,(cid:13)(cid:13)(cid:13)
2πðk(x, y)k
kξkH p,q,d
σx,y
.
θ
θ
θ
30
FRÉDÉRIC LATRÉMOLIÈRE
We endow span{P, Q} with the norm 2πðk·k. We also define, for all (x, y) ∈ R2
and ξ ∈ S p,q,d
θ
:
∇ð
x,yξ = lim
t→0
= lim
t→0
(t(xP +yQ))
expH3
ð,d
α
tx,ty, 1
ð,d
α
t
2 t2xy
ξ − ξ
t
ξ − ξ
.
To ease notation, let ·2πð denote the operator norm for linear maps from
,k · kH p,q,d
θ
).
(R2, 2πðk · k) to (H p,q,d
θ
We record:
(1) Dp,q,d
(2) For all ξ ∈ S p,q,d
θ
θ
is a norm on a dense subspace of H p,q,d
,
θ
and for all δ > 0, the following expressions hold:
Dp,q,d
θ
θ
,(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)∇ðξ(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)2πðo
(ξ) = maxnkξkH p,q,d
= sup(kσx,y
(cid:13)(cid:13)(cid:13)
= lim sup
(x,y)→0
ð,dξ − ξkH p,q,d
2πðk(x, y)k
σx,y
ð,d ξ − ξ(cid:13)(cid:13)(cid:13)H p,q,d
2πðk(x, y)k
.
θ
θ
: (x, y) ∈ R2, 0 < k(x, y)k < δ)
(3) If a ∈ Aθ and ξ ∈ H p,q,d
θ
then:
Dp,q,d
θ
(aξ) 6 kakAθ
then:
Dp,q,d
θ
(ξ) + Lθ(a)kξkH p,q,d
θ
.
θ
(4) If ξ, ω ∈ H p,q,d
Lθ(cid:16)hξ, ωiH p,q,d
(ξ)kωkH p,q,d
Proof. The Lie algebra of T2 is R2 with the exponential map given as:
(cid:17) 6 kξkH p,q,d
(ω) + Dp,q,d
Dp,q,d
θ
θ
θ
θ
θ
.
expT2 : (x, y) ∈ R2 7→ (exp(ix), exp(iy)).
Now, the map υð : (x, y) ∈ R2 7→ (2iπðy,−2iπðx) satisfies, according to Lemma
(4.2), the relation:
β
expT2 (υð(x,y))
θ
hξ, ωiH p,q,d
θ
expH3
ð,d
=Dσ
(x,y)
ξ, σ
expH3
ð,d
(x,y,0)
ωEH p,q,d
θ
.
and, according to Lemma (4.6), the relation:
(x,y)
expH3
ð,d
σ
(aξ) = β
expT2 (υð(x,y))
θ
(a)σ
expH3
ð,d
(x,y)
(ξ).
In order to apply Proposition (5.5), since υð is indeed a linear isomorphism, we
endow span{P, Q} with the norm:
kxP + yQk∗ = 2πðk(x, y)k.
We now are in the setting of Proposition (5.5), which allows us to conclude all but
Assertion (2) in our proposition. Assertion (2), in turn, follows from Theorem (5.2),
with our choice of norm.
(cid:3)
D-NORMS ON HEISENBERG MODULES
31
We now turn to the remaining, main issue of the compactness of the closed unit
balls for our D-norm candidates. The strategy we employ relies on a particular
source of finite rank operators naturally associated with the Schödinger represen-
tations of R2 via the Weyl calculus.
Our first step is to introduce the convolution-like operators at the core of our
analysis.
Lemma 6.5. Assume Hypothesis (4.1). If f ∈ L1(R2) and:
σf
ð,d =ZZR2
f (x, y)α
x,y, xy
2
ð,d
dxdy
then σf
ð,d is a well-defined operator on H p,q,d
θ
Proof. Let ξ ∈ H p,q,d
of H p,q,d
θ
θ
for all (x, y, u) ∈ H3, we simply compute:
and (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
σf
ð,d(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)H p,q,d
θ
6 kfkL1(R2).
. Using Lemma (4.4), i.e. the fact that αx,y,u
is an isometry
ð,d
f (x, y)α
x,y, xy
2
ð,d
ZZR2(cid:13)(cid:13)(cid:13)
(ξ)(cid:13)(cid:13)(cid:13)H p,q,d
θ
dxdy =ZZR2 f (x, y)(cid:13)(cid:13)(cid:13)
x,y, xy
2
ð,d
α
(ξ)(cid:13)(cid:13)(cid:13)H p,q,d
dxdy
θ
=ZZR2 f (x, y)kξkH p,q,d
= kfkL1(R2)kξkH p,q,d
.
θ
θ
dxdy
Thus σf
ð,d is well-defined, and moreover:
(cid:13)(cid:13)(cid:13)
σf
ð,d(ξ)(cid:13)(cid:13)(cid:13)H p,q,d
θ
=(cid:13)(cid:13)(cid:13)(cid:13)
ZZR2
f (x, y)σx,y
ð,d(ξ) dxdy(cid:13)(cid:13)(cid:13)(cid:13)H p,q,d
θ
6 kfkL1(R2)kξkH p,q,d
θ
.
(cid:3)
This completes our proof.
We now prove the first of two core lemmas of this section, which provides us with
a mean to approximate elements in Heisenberg modules using our convolution-type
operators, in a manner which is uniform in our prospective D-norms. This lemma
is an adjustment of [30] to our context.
Lemma 6.6. Assume Hypothesis (4.1). Let ε > 0. If f : R2 → [0,∞) is measurable
and satisfies:
then for all ξ ∈ H p,q,d
2πð ,
(1) RR2 f = 1,
(2) RRR2 f (x, y)k(x, y)k dxdy 6 ε
ð,dξ(cid:13)(cid:13)(cid:13)H p,q,d
ξ − σf
(cid:13)(cid:13)(cid:13)
:
θ
θ
6 εDp,q,d
θ
(ξ).
32
FRÉDÉRIC LATRÉMOLIÈRE
Proof. If ξ ∈ H p,q,d
θ
, then:
(cid:13)(cid:13)(cid:13)
ξ − σf
ð,dξ(cid:13)(cid:13)(cid:13)H p,q,d
θ
f (x, y)α
x,y, xy
2
ð,d
ξ dxdy(cid:13)(cid:13)(cid:13)(cid:13)H p,q,d
θ
dxdy
θ
ξkH p,q,d
kξ − α
x,y, xy
2
ð,d
ξkH p,q,d
θ
2πðk(x, y)k
dxdy
θ(ξ) dxdy
=(cid:13)(cid:13)(cid:13)(cid:13)
ZZR2
6ZZR2
6ZZR2
6ZZR2
= Dp,q,d
θ
= εDρ
θ(ξ),
f (x, y)ξ dxdy −ZZR2
f (x, y)kξ − α
x,y, xy
2
ð,d
f (x, y)2πðk(x, y)k
f (x, y)2πðk(x, y)kDρ
(ξ)(cid:18)2πð
2πð(cid:19)
ε
as desired.
(cid:3)
We now ensure that we indeed have an ample source of functions which meet
the hypothesis of Lemma (6.6).
Notation 6.7. If (E, d) is a metric space then the closed ball {x ∈ E : d(x0, x) 6 r}
of center x0 ∈ E and radius r > 0 is denoted by E[x0, r].
The following lemma is valid for any norm on R2; we shall work within our
context with the fixed norm k · k.
Lemma 6.8. For all n ∈ N, let ψn : R2 → [0,∞) be an integrable function
supported on R2h0,
If f : R2 → [0,∞) is integrable on some ball centered at 0 in (R2,k · k), and f
1
continuous at 0, then:
n+1i and with RR2 ψn = 1.
n→∞ZZR2
lim
ψn(x, y)f (x, y) dxdy = f (0).
Proof. Let δ > 0 such that f is integrable on R2[0, δ].
Let ε > 0. Since f is continuous at 0, there exists δc > 0 such that f (x)−f (0) 6
ε for all x ∈ R2[0, δc].
Let N ∈ N be chosen so that
6 min{δ, δc}. For all n > N , we first note
that since ψn is supported on a subset of R2[0, δ], the function ψnf is integrable on
R2. Moreover for all n > N :
N +1
1
ZZR2
(cid:12)(cid:12)(cid:12)(cid:12)
ψn(x, y)f (x, y) dxdy − f (0)(cid:12)(cid:12)(cid:12)(cid:12)
6ZR2 ψn(x, y)(f (x, y) − f (0)) dxdy
=ZZR2[0,n−1] ψn(x, y)f (x, y) − f (0) dxdy
6ZZR2[0,n−1]
ψn(x, y)ε dxdy 6 ε.
Thus we have shown that limn→∞RR2 ψn(x, y)f (x, y) dxdy = f (0).
(cid:3)
D-NORMS ON HEISENBERG MODULES
33
We are now ready to prove the second core lemma of this section. We begin with
an explanation of the ideas and reasons behind this lemma.
operator which maps bounded subsets of E to totally bounded subsets of E.
By a compact operator on a Banach space (E,k · kCd), we mean as usual an
The map f ∈ L1(R2) 7→ σf
ð,d is a *-representation of the twisted convolution
algebra L1(R2) for the convolution product defined for all f, g ∈ L1(R2) and x ∈ R2
by:
and the involution:
f (y)g(x − y)eð(y, x − y) dy
f∗ðg(x) =ZR2
f ∈ L1(R2) 7→ f ∗ = x ∈ R2 7→ f (−x),
σf
operators.
The fact that σf
as can be directly checked, or is established in [8]. It is an important, well-known
fact [8, Theorem 1.30] that this representation is valued in the algebra of compact
operators on L2(R) ⊗ Cd, and is faithful; the completion of (L1(R2),∗ð,∗) for the
is the entire algebra of compact
ð,d is compact as an operator of L2(R)⊗ Cd does not immediately
(cid:17) since in general,
ð,1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)L2(R)
norm f ∈ ℓ1(Z2) 7→ kfkC ∗(R2,eð) = (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
imply that it is compact for the Banach space(cid:16)H p,q,d
. We thus must prove compactness of these
we only know that k·kL2(R) 6 k·kH p,q,d
operators for our C∗-Hilbert norm. However, we can extract the essential tools for
our work from the expansive work on Laguerre expansion of functions and the study
of the Moyal plane. We will prove that, at least when f is a radial function, then
we can approximate σf
ð,d by finite rank operators, in norm. To this end, we need a
supply of finite rank operators, which provide a mean to approximate any σf
ð,d for
f radial. The theory of the quantum harmonic oscillator provides us with a well-
suited family of finite rank projections, obtained as σψ
ð,d for ψ a properly scaled
Laguerre function [8, Ch. 1, sec. 9].
,k · kH p,q,d
θ
θ
θ
To obtain the desired approximation result, however, we need to approximate
our radial functions in the norm of L1(R2) using functions obtained from Laguerre
functions. As Laguerre functions form an orthonormal basis for some L2 space,
we certainly do have a Laguerre expansion which converges in some L2 norm, but
convergence in L1(R2) is highly not trivial.
The work of Sundaram Thangaveru in [31] comes to our rescue, however, by
proving that we may obtain the desired convergence if we replace the Laguerre
expansion series by the sequence of its Césaro averages. We now formalize our
discussion in the next key lemma.
Lemma 6.9. If f : R+ → R is a function such that r ∈ R 7→ rf (r) is Lebesgue
integrable, and if we set:
f ◦ : (x, y) ∈ R2 7→ f(cid:16)px2 + y2(cid:17) ,
then the operator σf ◦
ð,d is a compact operator for the Banach space(cid:16)H p,q,d
,k · kH p,q,d
Proof. Our goal is to write σf ◦
ð,d as a limit, in the operator norm, of finite rank
operators. To this end, let us first assume that ð > 0 and for all n ∈ N, we let ψn
θ
ð
θ
(cid:17).
Note that these functions are given in [31, (6.1.17)] for ð = 1
which will be important for us in later proofs is that ψn
ð = ðψn
obtain all the Laguerre functions we are considering via a simple rescaling.
1 (√ð·), i.e. we can
π . An observation
By slight abuse of notation, we denote by Lp(R+, rdr) the p-Lebesgue space
for the measure defined, for all measurable f : [0,∞] → [0,∞), by R ∞
0 f (r) rdr.
In particular, note that the inner product of L2(R+, rdr) is given for any two
f, g ∈ L2(R, rdr), by:
hf, giL2(R+,rdr) =Z ∞
0
f (r)g(r) rdr.
34
FRÉDÉRIC LATRÉMOLIÈRE
ψn
be the nth Laguerre function defined for all r ∈ [0,∞) by:
2 (cid:19) Ln(cid:0)πðr2(cid:1) ,
(−1)j
j! (cid:18) n
n − j(cid:19)xj.
where Ln is the nth Laguerre polynomials, given for all x ∈ R by:
ð (r) = ð exp(cid:18)−
Ln(x) =
πðr2
n
Xj=1
With all these notations set, we define, for each n ∈ N\{0}, the nth Césaro sum
of the series given by the Laguerre expansion of f :
Cn
ð(f ) =
n
Xj=0
n + 1 − j
n + 1 Df ψj
ð, ψj
ðEL2(R+,rdr)
ψj
ð.
Then by the work of S. Thangavelu in [31, Theorem 6.2.1] -- where our ψj
ð is a
j in [31, Chapter 6] and we use the
rescaled version of the function denoted by ψ0
Césaro sums for "δ = 1" in his notations -- we conclude:
n→∞kCn
lim
ðf − fkL1(R+,rdr) = 0.
Now, a quick computation shows that for all n ∈ N \ {0}:
and therefore:
(cid:13)(cid:13)(cid:13)
(Cj
ð(f ))◦ − f ◦(cid:13)(cid:13)(cid:13)L1(R2)
=(cid:13)(cid:13)(cid:13)
Cj
ð(f ) − f(cid:13)(cid:13)(cid:13)L1(R+,rdr)
,
n→∞k(Cn
lim
ð (f ))◦ − f ◦kL1(R2) = 0
where of course, L1(R2) stands for the 1-Lebesgue space with respect to the usual
Lebesgue measure on R2.
By Lemma (6.5), writing κn = (Cn
ð(f ))◦ for all n ∈ N, we then conclude:
lim
n→∞(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
ð,d − σf ◦
σκn
ð,d(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)H p,q,d
θ
= 0.
By construction, σκn
ð,d is finite rank. Indeed, the operator σκn
bination of the operators σ
turn, projections on CHj
ð)◦
(ψj
ð,d with j ∈ {0, . . . , n}. The operators σ
ð ⊗ Cd ⊆ L2(R) ⊗ Cd, where Hn
ð,d is a linear com-
are, in
ð is the Hermite function:
(ψj
ð,d
ð)◦
Hj
ð : t ∈ R 7→
exp −
t2√2πð
2 ! Hj(cid:16)t√2πð(cid:17)
1
4
(2ð)
pj!2j
D-NORMS ON HEISENBERG MODULES
35
where Hj is the jth Hermite polynomial, given for instance by:
Hj : t ∈ R 7→ (−1)j exp(t2)
dj
dtj exp(−t2).
ð)◦
(ψj
ð,1
Thus the image of the unit ball H p,q,d
are projections on CHj ⊆ L2(R) for all
Indeed, by [8, p. 65], the operators σ
j ∈ N. We note that reassuringly, we will not need the explicit form of the Hermite
polynomials or the Laguerre polynomials in our work.
(cid:17) by σκn
[0, 1] of (cid:16)H p,q,d
(cid:17) for all n ∈ N, as a bounded subset of
totally bounded in (cid:16)H p,q,d
a finite dimensional space (as all norms are equivalent in finite dimension, this
observation does not depend on k · kH p,q,d
,k · kH p,q,d
,k · kH p,q,d
ð,d is compact as the norm limit of compact operators.
Thus σf ◦
We are left to treat the case when ð < 0. We note that for all (x, y, u) ∈ H3, we
ð,d is
).
θ
θ
θ
θ
θ
θ
have:
αx,y,u
ð,d = αx,−y,−u
−ð,d
.
We thus proceed as above with −ð in place of ð, and note that σκn
κn is a radial function. The rest of the proof is left unchanged.
ð,d = −σκn
−ð,d since
(cid:3)
With Lemma (6.9) and Lemma (6.6), we are now able to prove the desired
property for our D-norms:
Lemma 6.10. We assume Hypothesis (4.1). The set:
θ
D1(cid:16)Dp,q,d
,k·kH p,q,d
(cid:17) =nξ ∈ H p,q,d
(cid:17).
θ
θ
: Dp,q,d
θ
(ξ) 6 1o
θ
is compact in (cid:16)H p,q,d
Proof. Let (ψn)n∈N be a sequence of smooth functions from [0,∞) to [0,∞) such
that for all n ∈ N, the function ψn is supported on h− 1
n+1i and:
n+1 ,
1
Thus, using the notations of Lemma (6.9), we note that:
ψn(r) rdr =
1
2π
.
0
Z ∞
2 Z ∞
− π
π
2
0
ZR2
ψ◦
n =Z
ψn(r) rdrdθ =
2π
2π
= 1.
Let ε > 0 be given. By Lemma (6.8), we have:
Thus, there exists N ∈ N such that for all n > N , the following inequality holds:
We may thus apply Lemma (6.6) to conclude that for all ξ ∈ D1(cid:16)Dp,q,d
θ
(cid:17) and
n > N :
lim
n→∞ZZR2
ZZR2
ψ◦
n(x, y)k(x, y)k dxdy = 0.
ψ◦
n(x, y)k(x, y)k dxdy <
ε
4πð
(cid:13)(cid:13)(cid:13)
ξ − σ
ψ◦
n
ð,dξ(cid:13)(cid:13)(cid:13)H p,q,d
θ
6
ε
2
.
36
FRÉDÉRIC LATRÉMOLIÈRE
(cid:17) by Lemma (6.9), and D1(cid:16)Dp,q,d
(cid:17)
(cid:17) by σ
(cid:17) for all n ∈ N. In particular, there exists
ψ◦
n
ð,d
θ
θ
Now, σ
ψ◦
n
ð,d is compact in (cid:16)H p,q,d
θ
,k · kH p,q,d
θ
θ
θ
θ
θ
a ε
ψ◦
N
ð,d
2 -dense subset Bε in σ
by construction. Thus the image of D1(cid:16)Dp,q,d
is bounded for k · kH p,q,d
is totally bounded in (cid:16)H p,q,d
,k · kH p,q,d
(cid:17).
D1(cid:16)Dp,q,d
(cid:17), then there exists η ∈ Bε such that:
Consequently, if ξ ∈ D1(cid:16)Dp,q,d
(cid:13)(cid:13)(cid:13)
ð,d ξ(cid:13)(cid:13)(cid:13)H p,q,d
η − σ
(cid:17) is totally bounded.
Thus kξ − ηkH p,q,d
We thus conclude that D1(cid:16)Dp,q,d
Moreover, for all (x, y) ∈ R2, the map ξ 7→
θ
2πðk(x,y)k
6 ε.
H p,q,d
ξ−ξk
ε
2
ψ◦
N
xy
2
kα
6
x,y,
ð,d
.
θ
θ
θ
θ
is lower semi-continuous with respect to k·kH p,q,d
((−∞, 1]) is closed. Since H p,q,d
θ
θ
is complete and D1(cid:16)Dp,q,d
θ
. Hence D1(cid:16)Dp,q,d
θ
(cid:17) =
(cid:17) is closed
(cid:3)
and totally bounded, it is in fact compact, as desired.
thus Dp,q,d
θ
θ
(cid:16)Dp,q,d
(cid:17)−1
is continuous, and
We summarize the results of this section with the following theorem announcing
that indeed, we have defined D-norms on Heisenberg modules, turning them into
metrized quantum vector bundles over quantum 2-tori.
Theorem 6.11. Let H p,q,d
p ∈ Z, q ∈ N \ {0} and d ∈ qN \ {0}. Let ð = θ − p
a norm on R2. If we set, for all ξ ∈ H p,q,d
be the Heisenberg module over Aθ for some θ ∈ R,
q and assume ð 6= 0. Let k · k be
:
θ
θ
kξkH p,q,d
θ
,(cid:13)(cid:13)(cid:13)
σx,y
ð,dξ − ξ(cid:13)(cid:13)(cid:13)H p,q,d
2πðk(x, y)k
θ
: (x, y) ∈ R2 \ {0}
,
(cid:13)(cid:13)(cid:13)
βexp(ix),exp(iy)
θ
a − a(cid:13)(cid:13)(cid:13)Aθ
: (x, y) ∈ R2 \ {0}
k(x, y)k
,Aθ, Lθ(cid:17) is a Leibniz metrized quantum vector bun-
, Dp,d,q
θ
and for all a ∈ Aθ:
Dp,d,q
θ
(ξ) = sup
Lθ(a) = sup
,h·,·iH p,q,d
θ
then (cid:16)H p,q,d
θ
dle.
Proof. Proposition (6.4) proves that Dp,q,d
is a norm on a dense subspace of H p,q,d
which satisfies the inner and modular quasi-Leibniz inequalities and, by construc-
tion, Dp,q,d
.
θ
θ
θ
> k · kH p,q,d
θ
Lemma (6.10) moreover gives us that D1(cid:16)Dp,q,d
θ
(cid:17) is compact for k · kH p,q,d
θ
. (cid:3)
References
1. S. White J. Zacharias A. Hawkins, A. Skalski, On spectral triples on crossed products arising
from equicontinuous actions, Math. Scand. 113 (2013), 262 -- 291, arXiv:1103.6199.
D-NORMS ON HEISENBERG MODULES
37
2. K. Aguilar and J. Kaad, The Podlès sphere as a spectral metric space, Sub (2018), 23 pages,
ArXiv: 1803.03027.
3. K. Aguilar and F. Latrémolière, Quantum ultrametrics on af algebras and the Gromov --
Hausdorff propinquity, Studia Mathematica 231 (2015), no. 2, 149 -- 194, ArXiv: 1511.07114.
4. O. Bratteli and D. Robinson, Operator algebras and quantum statistical mechanics i, Springer-
Verlag, 1979.
5. A. Connes, C* -- algèbres et géométrie differentielle, C. R. de l'academie des Sciences de Paris
(1980), no. series A-B, 290.
6. A. Connes, Compact metric spaces, Fredholm modules and hyperfiniteness, Ergodic Theory
and Dynamical Systems 9 (1989), no. 2, 207 -- 220.
7. A. Connes and M. A. Rieffel, Yang-mills for noncommutative two-tori, Contemporary Math
62 (1987), no. Operator algebras and mathematical physics (Iowa City, Iowa, 1985), 237 -- 266.
8. G. Folland, Harmonic analysis in phase space, Princteon University Press, 1989.
9. R. Hoegh-Krohn, M. B. Landstad, and E. Stormer, Compact ergodic groups of automorphisms,
Annals of Mathematics 114 (1981), 75 -- 86.
10. A. Kleppner, Multipliers on Abelian groups, Mathematishen Annalen 158 (1965), 11 -- 34.
11. F. Latrémolière, Convergence of fuzzy tori and quantum tori for the quantum Gromov --
Hausdorff Propinquity: an explicit approach., Münster Journal of Mathematics 8 (2015),
no. 1, 57 -- 98, ArXiv: math/1312.0069.
12.
13.
14.
15.
, Curved noncommutative tori as Leibniz compact quantum metric spaces, Journal of
Math. Phys. 56 (2015), no. 12, 123503, 16 pages, ArXiv: 1507.08771.
, The dual Gromov -- Hausdorff Propinquity, Journal de Mathématiques Pures et Ap-
pliquées 103 (2015), no. 2, 303 -- 351, ArXiv: 1311.0104.
, Quantum metric spaces and the Gromov-Hausdorff propinquity, Noncommutative
geometry and optimal transport, Contemp. Math., no. 676, Amer. Math. Soc., 2015, ArXiv:
150604341, pp. 47 -- 133.
, Equivalence of quantum metrics with a common domain, Journal of Mathematical
Analysis and Applications 443 (2016), 1179 -- 1195, ArXiv: 1604.00755.
16. F. Latrémolière, The modular Gromov -- Hausdorff propinquity, Submitted (2016), 67 pages,
ArXiv: 1608.04881.
17. F. Latrémolière, The Quantum Gromov-Hausdorff Propinquity, Trans. Amer. Math.
2015,
electronically published on May 22,
no.
Soc. 368 (2016),
http://dx.doi.org/10.1090/tran/6334, ArXiv: 1302.4058.
365 -- 411,
1,
18.
19.
, A compactness theorem for the dual Gromov-Hausdorff propinquity, Indiana Univer-
sity Journal of Mathematics 66 (2017), no. 5, 1707 -- 1753, ArXiv: 1501.06121.
, The triangle inequality and the dual Gromov-Hausdorff propinquity, Indiana Univer-
sity Journal of Mathematics 66 (2017), no. 1, 297 -- 313, ArXiv: 1404.6633.
20. F. Latrémolière, Convergence of heisenberg modules for the modular Gromov-Hausdorff
propinquity, Submitted (2018), 34 pages.
21. F. Latrémolière and J. Packer, Noncommutative solenoids and the gromov-hausdorff propin-
quity, Proc. Amer 145 (2017), no. 5, 1179 -- 1195, ArXiv: 1601.02707.
22. H. Li, C ∗-algebraic quantum Gromov-Hausdorff distance, (2003), ArXiv: math.OA/0312003.
23. N. Ozawa and M. A. Rieffel, Hyperbolic group C ∗-algebras and free product C ∗-algebras
as compact quantum metric spaces, Canad. J. Math. 57 (2005), 1056 -- 1079, ArXiv:
math/0302310.
24. M. A. Rieffel, The cancellation theorem for the projective modules over irrational rotation
C ∗-algebras, Proc. London Math. Soc. 47 (1983), 285 -- 302.
25.
26.
, Projective modules over higher-dimensional non-commutative tori, Can. J. Math.
XL (1988), no. 2, 257 -- 338.
, Continuous fields of C* -- algebras coming from group cocycles and actions, Math.
Ann. 283 (1989), 631 -- 643.
27. M. A. Rieffel, Deformation-quantization for actions of Rd, Memoirs of the AMS, vol. 106,
American Mathematical Society, 1993.
28. M. A. Rieffel, Metrics on states from actions of compact groups, Documenta Mathematica 3
(1998), 215 -- 229, math.OA/9807084.
29.
, Group C ∗-algebras as compact quantum metric spaces, Documenta Mathematica 7
(2002), 605 -- 651, ArXiv: math/0205195.
38
30.
FRÉDÉRIC LATRÉMOLIÈRE
, Gromov-Hausdorff distance for quantum metric spaces, Mem. Amer. Math. Soc. 168
(March 2004), no. 796, math.OA/0011063.
31. Sundaram Thangavelu, Lectures on hermite and laguerre expansions, Mathematical notes,
vol. 42, Princteon University Press, 1993.
32. G. Zeller-Meier, Produits croisés d'une C*-algèbre par un groupe d' Automorphismes, J. Math.
pures et appl. 47 (1968), no. 2, 101 -- 239.
E-mail address: [email protected]
URL: http://www.math.du.edu/~frederic
Department of Mathematics, University of Denver, Denver CO 80208
|
1907.05832 | 1 | 1907 | 2019-07-12T16:43:17 | An Introduction to Abstract Classification Theory in the Operator Algebraic Setting | [
"math.OA",
"math.LO"
] | In the setting of modern mathematical logic and model theory, classification theory has been one of the landmark achievements of the field. Likewise, the classification of UHF-algebras and AF-algebras were substantial contributions to the field of operator algebra theory. These seemingly disparate topics of study in mathematics, model theory and operator algebras, have in recent years become closely related in many respects. I here attempt to bridge the gap between these two topics by discussing how operator algebraic classifications may be understood in terms of model-theoretic classification theory. This introductory article assumes basic familiarity with model theory and linear operator, but higher-level concepts are introduced when necessary. The focus of this introduction is conceptual and informal, and as such, many results are stated without proof, but relevant sources are cited for completeness. The reader should take this not as a detailed review, but rather as an overview of a general narrative thread connecting these two branches of modern mathematics. | math.OA | math |
An Introduction to Abstract Classification Theory in
the Operator Algebraic Setting
Patrick Fraser
[email protected]
University of Toronto
Abstract
In the setting of modern mathematical logic and model theory, classifica-
tion theory has been one of the landmark achievements of the field. Likewise,
the classification of UHF-algebras and AF-algebras were substantial contri-
butions to the field of operator algebra theory. These seemingly disparate
topics of study in mathematics, model theory and operator algebras, have
in recent years become closely related in many respects. I here attempt to
bridge the gap between these two topics by discussing how operator algebraic
classifications may be understood in terms of model-theoretic classification
theory. This introductory article assumes basic familiarity with model theory
and linear operator, but higher-level concepts are introduced when necessary.
The focus of this introduction is conceptual and informal, and as such, many
results are stated without proof, but relevant sources are cited for complete-
ness. The reader should take this not as a detailed review, but rather as
an overview of a general narrative thread connecting these two branches of
modern mathematics.
1. Introduction
Freeman Dyson famously wrote about a distinction common to the sci-
ences between the diversifiers and the unifiers [1]. In order for a science to
mature, it requires these two groups of people; the diversifiers who explore
the vast jungle of the world, uncovering new, beautiful objects to study, pro-
liferating novel ideas, and the unifiers, who fly above this jungle and piece
together the distinct components, drawing them together into a coherent
picture. Both are essential to understand the world fully. In mathematics,
Based on an essay submitted for MAT437: K-theory and C ∗-algebras
July 15, 2019
too, we see a vast array of diversifiers and unifiers. This essay focuses on the
general goals and methods of the latter.
One of the common ways of unifying ideas in mathematics is through
so-called classification. Initially, the mathematician may notice that there is
a particularly interesting class of objects which have certain properties that
may be used to understand how things work under a new light. Such was the
case with the early formal introduction of what are now standard algebraic
structures, such as groups, rings, and fields. Once such basic objects have
been noticed, one then seeks to explore their behaviour. One might notice,
for instance, that not all rings are commutative, but when they are, certain
features appear which may be exploited to further our understanding. The
initial piece of discovery opens a path, and mathematicians eagerly map their
way along this new path, until they have a fairly detailed understanding of
its cartography. Once the objects under consideration become sufficiently
well understood, it becomes reasonable to try to ask the strongest, most
general questions. Specifically, once mathematicians realize that a certain
class of mathematical objects has interesting behaviour, they might want
an explicit list of all of these objects of a particular form (up to some sort
of equivalence). Often, obtaining such a complete list is very difficult; in
many cases, a classification program first identifies a collection of quantities
which are invariant under isomorphisms, and then claim that the objects in
question are classified if these invariants are enough to tell apart any two
non-isomorphic objects. The invariants themselves may be quite difficult to
compute, and so a complete list is often unattainable (such is the case with
von Neumann algebras, whose classification program is considered complete,
even though it is often quite difficult to say whether or not two von Neumann
algebras are isomorphic) but in principle, these invariants would allow one to
compare any two objects and determine if they are isomorphic or not. This
is the game of classification.
An ideal classification program would consist of a complete list, up to
isomorphism, of all of the objects which behave in a certain way.
In an
elementary group theory course, one might, for instance see the historically
significant classification of finite abelian groups. This classification tells us
that a group G is a finite abelian group if and only if G ∼= Z/n1Z×...×Z/nk Z
for sufficient ni's and k; it gives us a complete list, up to isomorphism, of all
possible finite abelian groups. In more sophisticated settings, classification
theorems provide a powerful correspondence between the abstract and the
concrete, and are usually landmark theorems in their disciplines.
2
Classification theory is a field of study in its own right, generally sub-
sumed as a topic within model theory and mathematical logic. Here, I present
an elementary introduction to classification theory in its model-theoretic set-
ting, and then discuss how it may be concretely applied to the theory of
operator algebras, UHF and AF-algebras in particular.
2. Classification Theory
I begin with the broad notions of classification theory from the model-
theoretic standpoint, based on [2]. In lieu of a basic introduction to model
theory, I point the interested reader towards the very accessible introduc-
tion [3] and the slightly more sophisticated [4]. Assuming familiarity with
the concept of a first-order predicate language,1 I take the following defini-
tions:
Definition: Fix a language L. An L-structure A consists of a set of
elements (the universe), a set of constants of the same cardinality as the
constants of L, and for every relation and function symbol in L, a relation
and function between elements of the universe.
Languages have no meaning, but one may map the symbols of a language
to a structure in which those symbols take on a meaning using an assign-
ment function. Generally, one is concerned with models of a particular set
of formulas; that is, some structure in which all of a certain class of abstract
formulas are true.
Definition: Let L be a language. Given a set T of L-formulas, an L-
structure M is a model of T if M (cid:15) φ for all φ ∈ T .
Structures are objects such as the natural numbers equipped with a binary
addition function and a binary ordering relation, which may be taken as
1Essentially, such a language is a meaningless assemblage of symbols which represent
variables, constants, relations, and functions, with standard logical connectives like ∧, →,
and ∨, as well as a negation symbol ¬ and universal and existential quantification symbols
∀ and ∃ respectively, which have basic rules for constructing grammatical terms, formulas,
and sentences. The language itself is meaningless, but concrete meaning is prescribed to
symbols and formulas when they are interpreted in a structure of that language.
3
models for certain abstract formulas in the language. For example, in a
language with one constant and an order relation L = {0, ≤}, the abstract
formula φ := (∃x)[(∀y)(x ≤ y)] is merely a string of symbols, but when
interpreted in the structure of the natural numbers, this formula is satisfied
because it means that the natural numbers have a smallest element, which
is true. However, this same formula is not satisfied by the L-structure that
is the rational numbers (equipped with the usual ordering), for instance,
because they do not have a lowest element under their usual ordering. As
such, N models φ, but Q does not (and we write N (cid:15) φ and Q 6(cid:15) φ).
When we seek to "classify" a collection of objects, we require first that
they are in fact objects of the same kind; if we are concerned with groups,
we are concerned with groups. If we are concerned with approximately finite
C ∗-algebras, we are concerned with approximately finite C ∗-algebras. But
we are never comparing apples with oranges. Classification takes place al-
ways within the confines of kind. Formally, this amounts to saying that we
are restricting ourselves to L-structures which are all grounded in the same
language L, and further, which all satisfy the same axioms. That is, they
must be models of the same L-theory. With this in mind, I give the fol-
lowing definitions which will help us, in broad strokes, to study the general
behaviour of classes of models for a given theory:
Definition: Let K = (K, ≤K) where K = Mod(T ) the set of all models of
some first-order L-theory T , and ≤K is partial ordering on K. Then K is an
abstract elementary class if: (i) K is closed under isomorphism and ≤K pre-
serves isomorphisms, (ii) if M ≤K N , then M is a substructure of N , (iii)
(downward Lowenheim-Skolem theorem) there exists a cardinal α ≥ ℵ0 + L
such that for every M ∈ K and A ⊂ M, there is a N ∈ K with N ≤K M,
A ⊂ N , and N ≤ A + α, and (iv) (Tarski-Vaught Chains) K is closed
under ≤K-increasing chains of arbitrary regular cardinality. [2, 5]
Essentially, abstract elementary classes are classes of models for a par-
ticular theory which have enough particular constraints to ensure that the
elements in each class all 'look alike' in a precise sense. The models of a
theory, particularly the infinite models, are in general not easy to pin down.
Indeed, the following famous theorem ensures that, if there are any infinite
models of some some first-order theory, there are always many distinct infi-
nite models.
4
Theorem (Lowenheim-Skolem): Suppose T is a countable theory in
a first-order language L which admits an infinite model M. Then for all car-
dinals κ > L, there exists a first-order theory of size κ which is elementary
equivalent to M (see, for instance, [3, 4]).
One may heuristically understand this as saying that an infinite model
of some theory can always be made more or less infinite, for instance, by
extending it to a non-standard model. To understand the behaviour of such
infinite models, which may, in general, be very badly behaved, it is valuable
to understand categoricity:
Definition: An abstract elementary class K is categorical in λ if it has
one and only one model of cardinality λ up to isomorphism). The class of
cardinals for which K is categorical is called the categoricity spectrum of K
and is denoted cat(K).
Categoricity allows us to compare the relative size of elements of some
abstract elementary class, and indeed, there are several nice facts about
categoricity which lend themselves nicely to the demarcation of distinct iso-
morphism classes. Notably, the following theorem is substantial [6]:
Theorem (Morley): Fix some countable first-order L-theory T . Let
λ be a cardinal, and I(λ, K) be the number of models in K of cardinality
λ (up to isomorphism). Then if for some λ > T + ℵ0, I(λ, T ) = 1, then
I(µ, T ) = 1 for all cardinals µ > T + ℵ0.
That is, if the categoricity spectrum of some countable first-order theory
includes a single model of some infinite cardinality larger than that of the
language used, then indeed the categoricity spectrum of that theory extends
to all such cardinals. Categoricity is the gateway to stable model theory
which attempts to understand how well-constrained the infinite models of
some theory are. If one wants to consider more specific properties unique to
a models of a particular theory, the route to formalizing this is definability,
which tells you when the a particular theory admits the specification of par-
ticular subsets based on a precise set of properties which can be formulated
in the language in question.
Definitions: Let A be an L − structure with universe A and let B ⊂ A.
5
A set X ⊂ An is B-definable if there exists some φ(v1, ..., vn, w1, ..., wn) such
that X = {a ∈ Anφ(a, b), b ∈ Bm}.
We can now say that classification theory deals with the problems of un-
derstanding isomorphism classes within abstract elementary classes of mod-
els, and frequently, the understanding of the isomorphism classes relies on
what features are definable within a particular first-order theory and how
the categoricity of the models behaves. The high-level goal of classification
theory is to look at a collection of such classes {K}I and present clear mecha-
nisms for demarcating the "nice" or "well-behaved" classes from those which
are "messy" or "complicated". In this way, one might obtain a strong way to
pin down the well-behaved classes, and prove substantive claims about all of
these classes through more straight-forward case work. The most important
result along these lines has been due to Shelah who determined the precise
conditions for which an abstract elementary class admits a classification (or
in more technical language, has a structure theory) of its isomorphism classes
(the so-called dichotomy theorem) [2, 7, 8]. This theorem states that:
Theorem (Shelah):An abstract elementary class K has a structure the-
ory if there is a class of subsets {A}I of the universe of each model in K
(called cardinal invariants), which are definable in each model, such that, for
any two M, N ∈ K, M ∼= N if and only if AN
for all invariants Ai.
i = AM
i
The low-level goal of classification theory is, given a particular class K
which is known to admit a classification, to determine concretely all of the
elements of K up to isomorphism, or at least to determine a cardinality
bound for how many such distinct elements there may be. The later part of
this project was, again, solved by Shelah [2, 7, 8]. In the operator algebra
setting, once one has established that a particular kind of operator algebra
with a first-order set of axioms in fact does admit a classification (a structure
theory), then the important work is to determine every element of the class of
models of those axioms up to isomorphism. The earliest results along these
lines were the classification of UHF-algebras due to Glimm [9, 10] and the
classification of AF-algebras due to Elliott [9, 11].
Before proceeding to the operator algebra setting, to see a nice case of
classification theory in action in a comfortable setting, consider vector spaces.
Example: Vector Spaces. The class of all vector spaces V is the class
6
of all models of the theory of vector spaces TV S. One can show readily that
TV S is a first-order theory [3]. Therefore, the class of vector spaces forms
an abstract elementary class (ordered under inclusion, where all of the other
more complicated requirements are satisfied more or less trivially). Given a
vector space V ∈ V, there is a definable subset of V which is the set of all
bases of V (the relation under which this set is definable in the language of
Marker [4], is span). It is an elementary fact of linear algebra (and thus a
theorem provable in the theory TV S, and so true of all vector spaces) that the
cardinality of every basis of a particular vector space is the same. Thus, the
cardinality of every element of the set of bases of V is constant. We call this
cardinality the dimension of V , and it is provable using elementary linear
algebra that two finite-dimensional vector spaces are isomorphic if and only
if their dimensions are the same. Therefore, the cardinal invariant for the
structure theory of vector spaces is the dimension, which is the cardinality
of any element of the definable subset of the universe of V which is the set
of bases for V .
3. C ∗-algebras and their K-Theory
In the classification of C ∗-algabras, K-theory is a valuable tool, as it al-
lows one to construct a related structure to the algebra in question, namely
the ordered K0-group (as well as higher K-groups), which encodes lots of the
information about the algebra while allowing questions to be posed in the
context of abelian groups rather than operator algebras. Because of this, K-
theory is helpful for the classification of C ∗-algebras, as definable invariants
for the algebra may be functorially pushed forward to the K-theory setting
where they are more natural and easier to work with. In the special case of
AF-algebras, the classification process is made very clear. After presenting
the basic ideas pertinent to studying C ∗-algebras and their K-theory, I walk
through the examples of UHF-algebras and AF-algebras to understand their
classification in model theoretic terms.
I then discuss more general ideas
presently being pursued for more general classifications. As such, I base the
following on [9].
Definition: A C ∗-algebra is a Banach algebra equipped with a conjugate-
linear homomorphic isometric involution x 7→ x∗ such that x∗x = xx∗ =
x2.
7
Broadly speaking, a C ∗-algebra is an abstract representation of some al-
gebra of bounded, self-adjoint, norm-closed operators on a Hilbert space. In-
deed, the Gelfand-Naimark-Segal (GNS) construction tells us precisely that
every abstract C ∗-algebra admits such a Hilbert space representation. This
notion of concrete representation is made stronger by the existence of the
Gelfand transform, which acts as an isometric ∗-isomorphism between ele-
ments of any given commutative, unital C ∗-algebra and the set of bounded
continuous functions on some particular compact Hausdorff space2. The
study of specific operator algebras over particular Hilbert spaces together
with their associated spectra is a significant area of study in its own right,
but the abstract treatment of general C ∗-algebras is likewise very rich, and
particularly amenable for understanding the behaviour of such systems in
the most unified, general setting.
Even though 'zooming out' to the higher-level theory of abstract operator
algebras allows one to forget about many of the precise details of the systems
they are considering, C ∗-algebras are, in their own right, highly complicated
systems. As such, it is valuable to have tools to simplify problems pertaining
to C ∗-algebras further. One such method for gaining further understanding
comes from K-theory.
When considering any operator algebra, the class of idempotent elements
is of particular importance (one need only look towards von Neumann's for-
malism of quantum mechanics, or more modern POVM-based formalisms to
see how practically powerful projections may be in the wild [12]). Specifi-
cally, projections in an algebra 'pin down' much of the dimension structure
of the algebra3 quite concretely in a precise sense using unitary equivalence
and Murray-von Neumann equivalence:
Definition:Let A be a C ∗-algebra. Let p, q ∈ A. Then p and q are
Murray-von Neumann equivalent (written p ∼ q) if there exists some v ∈ A
such that p = v∗v and q = vv∗.
2In this sense, it may indeed be reasonable to think of the study of C ∗-algebras as an
abstract study of noncommutative topology.
3For a particularly easy example, given the matrix algebras Mn(C) and Mk(C), one
has that Mn(C) ∼= Mk(C) if and only if n = T r([In]0) = T r([Ik]0) = k where [In]0 and [Ik]
are the K-theoretic identities associated with each algebra (here, just the matrix identity
up to Murray-von Neumann equivalence).
8
It can readily be shown that if p, q ∈ A are projections which are homo-
topic or unitarily equivalent, then p ∼ q. There is a canonical matrix algebra
semigroup of projections associated with any C ∗-algebra A given by:
Definition: Let Mn(A) denote the algebra of m×m matrices with entries
in the C ∗-algebra A. Let Pn(A) denote the class of projections in the ma-
trix algebra Mn(A). Then the projection semigroup is P∞(A) := ∪∞
n=1Pn(A).
Addition in this semigroup is given by the usual direct sum ⊕ on matri-
ces. There is an associated notion of Murray-von Neumann equivalence over
P∞(A) given by:
Definition: Let p ∈ Pn(A) and q ∈ Pm(A). Then p ∼0 q if there is
some v ∈ Mm,n(A) such that p = v∗v and q = vv∗.
The triumph of K-theory is that, in light of this equivalence on the pro-
jection semigroup associated with a C ∗-algebra A, one can readily construct
a functor from the category of C ∗-algebras to the category of ordered abelian
groups, which preserves most of the relevant information about the structure
of the algebra through its idemptotent elements. This functor is called the
K0 functor, and it has many beautiful features. The K0 functor (which is
a map which takes a C ∗-algebra A to an ordered abelian group K0(A)) is
defined as follows:
Definition: Let S be a semigroup. The Grothendieck group G(S) is given
by the quotient G(S) = S × S/ ∼G where (x1, y1) ∼G (x2, y2) if and only if
there is some z with x1 + y2 + z = x2 + y1 + z. The K0 group associated with a
C ∗-algebra A is given by the Grothendieck group of the projection semigroup
of A. That is, K0(A) = G(P∞(A)/ ∼0).
To better understand what K0(A) 'looks like', the following results are
quite valuable:
Theorem:
Let [p]0 denote the ∼0 equivalence class of some element
p ∈ P∞(A). Then K0(A) = {[p]0 − [q]0 : p, q ∈ Pn(A), n ∈ N}. Moreover,
[p]0 + [q]0 = [p ⊕ q]0. Finally, if p and q are homotopic, then [p]0 = [q]0.
Furthermore, the functoriality of K0 tells us that it distributes in the
9
natural sense over the composition of ∗-homomorphisms and preserves the
identity map. Additionally, if one has the categorial concept of inductive
limits in place (described below), there is a natural sense in which K0 is
continuous.
With the notion of K0 in place, we may begin to talk about specific
classes of C ∗-algebras in terms of their K-theory. The most elementary class
of C ∗-algebras is that of finite C ∗-algebras (each of which is isomorphic to
the matrix algebra Mn(C) for some n). These algebras are not terribly ex-
citing, but one can form many highly non-trivial classes of C ∗-algebras out
of them using inductive limits. In their general construction, inductive lim-
its act like limits of sequences of objects in some category which are linked
to each other by connecting arrows; they generalize limits to more abstract
categories. Specifically:
Definition: Let C be some category. Let {Oi, φij} be a sequence of
: Oi → Oj (i < j). Then the inductive
objects Oi in C and arrows φij
limit lim→{Oi, φij} is the object O∞, which has a canonical set of maps
φi : Oi → O∞ such that O∞ = ∪∞
i=1φi(Oi) and where φj = φi ◦ φij (i.e.
the below diagram commutes).
φ1
φ23
φ12
O1
O2
...
φ2
O∞
It is important to note that not all categories admit inductive limits (for
instance, the category of sets). When we take the category in question to
be the category of C ∗-algebras, we can form inductive limits (taking the
connecting maps to be ∗-homomorphisms), and it turns out that even the
inductive limits of finite C ∗-algebras can be quite fascinating objects to study.
For our interests here, this setting also provides us with a straightforward
and historically significant instance of classification theory found in the wild.
Specifically, the study of UHF-algebras and AF-algebras (which are inductive
limits of finite C ∗-algebras) provide an interesting setting to connect Shelah's
abstract classification theory to concrete and relevant classification programs.
10
4. The Classification of C ∗-algebras
To proceed, we begin with the following definition of approximately finite
(AF) algebras:
Definition: An AF-algebra is a C ∗-algebra which is isomorphic to the
inductive limit of a sequence of finite C ∗-algebras.
The first example of classification theory as it pertains to C ∗-algebras
which makes clear how these ideas are all inter-connected in a concrete way is
the classification of uniformly hyper-finite (UHF) algebras due to Glimm [10].
An algebra A is a UHF-algebra if it is an AF-algebra where the connecting
∗-homomorphisms are unit preserving. UHF-algebras may be classified using
supernatural numbers which are defined as [9]:
Definition: Let {ni}i∈N be a (finite or countable) set of cardinals where
i where pi
ni ≤ ℵ0, and let n formally be expressed as the product n = Πi∈Npni
is the ith prime number. Any such n is a supernatural number.
It will be shown that {ni}i∈N is the set of cardinal invariants of models of
4, the first-order theory of UHF algebras. Without getting our hands
TU HF
too dirty, stated without proof, we have the following theorems [9]:
Theorem: Let G ⊂ (Q, +) be a subgroup containing 1. Then there is a
supernatural number n associated with G such that G = Q(n), the subgroup
composed of elements of Q where the denominator is a factor of n.
Theorem: Let A be a UHF algebra. Then (K0(A), [1A]0) ∼= (Q(n), 1)
for some supernatural number n.
From this, we see that every UHF-algebra has an associated supernat-
4It should be noted that, in general, C ∗-algebras are not universally first-order ax-
iomatizable, and even when they are, they generally require an uncountable language,
making certain results, such as Lowenheim-Skolem results, more subtle. However, it can
be shown that abelian algebras, and unital finite C ∗-algebras (as well as many other ele-
mentary kinds of algebras) are first-order axiomatizable, and so we do not get into trouble
here [13].
11
ural number n, and so to every UHF-algebra, a set of cardinal invariants
{ni}i∈N may be assigned. This is a step towards classification, however,
we do not yet know that this associated supernatural number determines a
given UHF-algebra uniquely up to isomorphism; a priori, there may exist
two non-isomorphic UHF-algebras (models of TU HF ) with the same associ-
ated supernatural number. However, the major result of Glimm says that
the associated supernatural number of a UHF-algebra is in fact enough to
determine that UHF-algebra up to isomorphism. Specifically, we have:
Theorem: (Glimm): If A and A′ are two UHF algebras with associated
supernatural numbers n and n′ respectively, then A ∼= A′ if and only if n = n′.
In the model-theoretic setting, if n = n′, since prime factorizations are
unique, we see that ni = n′
i}i∈N, and so
we see that A ∼= A′ if and only if ni = n′
i for all i ∈ N and so {ni}i∈N is a
(countable) set of cardinal invariants which constitute the structure theory
for TU HF . Therefore, we see a nice reflection of Shelah's classification theory
in Glimm's theory for UHF-algebras.
i for all elements of {ni}i∈N and {n′
In light of the fact that (K0(A), [1A]0) ∼= (Q(n), 1) from above, we see that
in fact we may restate this classification theorem to be in the language of K-
theory. Specifically, A ∼= A′ if and only if (K0(A), [1A]0) ∼= (K0(A′), [1A′]0),
because the K groups are determined by the associated supernatural num-
bers.
Further connecting the classification of UHF-algebras to classification the-
ory in general, if we recall the low-level goal of determining the cardinality
of the set of isomorphism classes of models for a particular theory (TU HF in
this case), we see that there is a converse theorem which comes in handy. It
is shown in [9] that for every supernatural number n, there is a UHF-algebra
with associated supernatural number n. A supernatural number is exactly
specified by the set {ni}i∈N (noting that we have here ordered the ni's using
the same usual ordering of the corresponding prime numbers, and that prime
factorization is unique). Since there are ℵ0 elements of {ni}i∈N, and since
each ni may take on ℵ0 possible values, there are ℵℵ0
0 many supernatural
numbers. One may show that ℵℵ0
0 = 2ℵ0 = ℵ1 = R, and so there is a (small)
uncountable set of distinct isomorphism classes of UHF algebras. I now look
at unital AF-algebras in general.
Let TAF be the first-order theory of unital AF-algebras (everything to fol-
low may be generalized to the non-unital case [9, 11], but the details involved
12
do not lead to a clearer understanding of the general classification method).
The question which must be answered to classify AF-algebras in the sense of
Shelah is: What set of cardinals {λi}i∈I is definable in any model A (cid:15) TAF
i}i∈I if and only if A ∼= A′? The theorem of Elliott [11]
such that {λi}i∈I = {λ′
of course indicates how this may be answered:
Theorem (Elliott): Two unital AF-algebras A and B are isomorphic if
and only if (K0(A), K0(A)+, [1A]0) ∼= (K0(B), K0(B)+, [1B]0).
Thus, the ordered K0 group with its unit is an algebraic invariant which
is sufficient to classify AF-algebras. But how do we interpret this in terms
of sets of cardinals? The following result, found in [9], brings clarity to the
matter:
Theorem: Let A be an AF algebra. Then (K0(A), K0(A)+) is a dimen-
sion group. Conversely, for every dimension group (G, G+), there exists an
AF-algebra A such that (G, G+) ∼= (K0(A), K0(A)+).
The notion of dimension is valuable, because it allows us to tease out the
cardinals from the discussion. As is explained clearly by Effros et al. [14],
'dimension' may take on many different meanings. In the context of matrix
algebras, dimension refers to the rank of the projections in the algebra. In the
von Neumman algebra setting, dimension becomes a non-negative real value.
In the C ∗-algebra setting however, dimension refers to a value assigned to
Murray-von Neumman equivalence classes of projections in the K0-group of
the algebra; dimension then refers to values in some ordered abelian group.
In this way, we see why the term dimension group is fitting in the AF-algebra
setting; the dimension group is precisely the K0 group in which elements of
the algebra may have their 'dimension' defined.
If we wish to understand how Elliott's classification theorem for AF-
algebras may be expressed in terms of cardinal invariants for the abstract
elementary class of models of the first-order theory for AF-algebras, it is
valuable to follow Bratteli [15] in his diagrammatic representation of AF-
algebras using infinite directed graphs, and read off the cardinal invariants
from this graph in terms of counting the vertices and edges in a particular
way. Bratteli diagram for an AF-algebra A ∼= lim→{Ai, φij} are constructed
thusly: For each Ai, there will be a row of the Bratteli diagram consisting
13
∼= Ls
of a collection of points. Each Ai is isomorphic to some finite factor space
of finite-dimensional complex matrix algebras. That is, Ai
k=1 Mki(C).
Thus, in the ith row of the Bratteli diagram, draw one point for each factor
Mki(C) of Ai, and label it by the integer ki. Given the ith row of a Bratteli
diagram, the i + 1th row is obtained by likewise drawing one point for each
factor space of Ai+1 (in its finite-dimensional complex matrix algebra repre-
sentation, for instance). Then the ith row is connected to the i + 1th row by
drawing edges connecting the points in the ith row to points in the i+1th row
precisely when a given point in the ith row is mapped to the corresponding
point in the i + 1th row under the connecting map φii+1. The number of lines
connecting two points denotes the multiplicity of the map connecting the two
points; if φij : Ai → Aj is a ∗-homomorphism, then the multiplicity of φij
is defined to be T r(φij(p))/T r(p) for any non-zero projection p (it is easy to
show that this quantity is independent of the choice of p). It turns out this
graph representation completely characterizes the inductive limit and thus
completely characterizes up to isomorphism the underlying AF-algebra. Two
examples of Bratteli diagrams (taken from [9]) may be found in figures 1 and
2.
·1
·2
·3
...
·1
·1
·2
...
Figure 1: The Bratteli diagram for the inductive sequence of finite-dimensional C ∗-algebras
constructed by taking An = Mfn (C)⊕Mfn−1(C) where f0, f1, f2, . . . are the usual Fibonacci
numbers, with connecting maps given by: (x, y) 7→ (cid:0)(cid:0) x 0
0 y (cid:1), x(cid:1).
The ordered abelian K0 group (or the dimension group, as one chooses)
associated with an AF-algebra, as Elliott's theorem showed, comes from the
sequence of the K0-groups of the terms in the inductive sequence of finite C ∗-
algebras which induce the AF-algebra in question. As such, to determine the
14
·1
·2
·4
·8
...
Figure 2: The Bratteli diagram for the inductive sequence of finite-dimensional C ∗-algebras
given by An ∼= M2n(C), with the connecting maps given by φii+1(x) = diag(x, x). We
see that the inductive limit of this sequence is isomorphic to M2∞ (C). It may be shown
that this is also the (unique) C ∗-completion of the canonical anti-commutation relation
(CAR) algebra over a separable infinite-dimensional Hilbert space, an algebra which is
very important in mathematical physics, the study of quantum field theories and quan-
tum statistical mechanics in particular (anything which makes use of antisymmetric Fock
space).
algebraically invariant dimension group of an AF-algebra, one may look at
the dimension groups associated with each algebra in the sequence. But this
is effectively the content of a Bratteli diagram! Bratteli diagrams show us,
at each level, the dimension of each finite matrix sub-algebra (using the un-
usual notion of dimension, which may readily be cast as a cardinal) together
with the way in which these finite matrix algebras are 'stitched' together
via their connecting maps, as well as the multiplicity of these connecting
maps (which therefore determines how the rank of projections changes when
pushed forward to the next row of the Bratteli diagram). The 'bottom row'
of a Bratteli diagram thus determines exactly the dimension group of the in-
ductive limit of the sequence, which is therefore isomorphic to the K0 group
of the AF-algebra in question.
The ith row of a Bratteli diagram may be completely specified by the
number associated to each point, together with the edges connecting it to
the previous row. Thus, the ith row may be specified completely by the set of
cardinals {ni, {di,j}ni
} where ni is the number of
j=1, {m(i−1,s)→(i,r)}s=ni−1,r=ni
s,r=1
15
s,r=1
elements of the ith row, di,j is the dimension associated with the point in the
jth position (from left to right) of the ith row, and {m(i−1,s)→(i,r)}s=ni−1,r=ni
denotes the multiplicities of each edge connecting the sth point in the i − 1th
row to the rth point in the ith row (where each m(0,s)→(1,r) = 0 for all s and
r since there are no connecting maps ending in the first row). These values
are all finite integers for any given row. The 'bottom line' (i.e. the ith row
in the limit when i → ℵ0, the countable infinity) may likewise be specified
by this set of cardinals.5 The Bratteli diagram encodes all of the information
about the K0 group of the AF-algebra in question, and so it is an equivalent
description of the invariant necessary to classify AF-algebras, although it is
more conducive to being read as a sequence of cardinals, formally putting us
in a position to claim, in Shelah's language, that models of the theory TAF
are classified up to isomorphism by a set of cardinal invariants.
UHF and AF-algebras are not the only classes of C ∗-algebras to be classi-
fied, nor are they the only classes of C ∗-algebras defined in terms of inductive
limits to be classified. However, noting that, in the UHF-algebra case, the in-
variant required could be equivalently expressed as the double (K0(A), [1A]0),
and in the more general unital AF-algebra case, the invariant needed was only
the slightly more sophisticated triple (K0(A), K0(A)+, [1A]0), one might be
inclined to state a more general thesis, namely that all C ∗-algebras of some
form may be classified by a more general K-theoretic invariant. This is the
present program of classification, and one of the conjectured invariants is the
so-called Elliott invariant [16]:
Ell(A) = (K0(A), K0(A)+, [1A]0, K1(A), T (A), rA)
Here, K1(A) is the standard K1 group (which is isomorphic to the K0
group of SA, the topological suspension of A), T (A) is the tracial space of
A (i.e. the set of all tracial functionals in the dual space of A), and RA is
the so-called coupling map which associates states with traces. The goal of
the classification program was to classify all separable, unital, nuclear simple
C ∗-algebras via this invariant. That is, ideally, one would end up with a
theorem which states that, if A and B are two such algebras, then A ∼= B
if and only if Ell(A) ∼= Ell(B). Unfortunately, there have been counter-
5For the set theorist, this is an terribly dull set of cardinals, as all of them are finite or
countably infinite. However, boring or not, cardinals they are, and so they are satisfactory
for us here.
16
examples preventing such a complete functorial classification, but it has held
for many important cases, such as the aforementioned algebras, as well as
irrational rotation algebras, and AT-algebras in general. There are presently
new candidate invariants being proposed. However, these classification pro-
grams have become extremely complicated and sophisticated, and a complete
functorial classification may be far away; one of the major problems in this
field currently is to try and understand precisely how complex the isomor-
phism maps between C ∗-algebras may in principle be, thereby giving insight
into the complexity of their classification [16].
5. Conclusions
I here presented a general overview of Shelah's classification theory, and
the theory of cardinal invariants. After a cursory glance at the basics of K-
theory for C ∗-algebras, I then recalled several known results, due primarily to
Glimm and Elliott, about the classification of various classes of C ∗-algebras
using particular K-theoretic invariants. I demonstrated how these invariants
may be understood as cardinal invariants in terms of sequences defining su-
pernatural numbers and sequences defining states of Bratteli diagrams. With
this understanding in place, I described the goals of future classification. The
topic of C ∗-algebra classification has, in recent years, become deeply rooted
in model theory and mathematical logic, and so framing the discussion in
terms of Shelah's classification theory reorients the setting of discussion to
be more amenable to such discourse. In this way, we see that the marriage
of model theory and operator algebra theory is, in a sense, preordained by
the sophistication of the field, and will likely serve as an important tool for
future investigation.
References
[1] Freeman Dyson. Infinite in All Directions: Gifford Lectures Given at
Aberdeen, Scotland. Harper Perennial, 1985.
[2] Saharon Shelah. Introduction to: classification theory for abstract ele-
mentary class. arXiv:0903.3428, 2009.
[3] Christopher C. Leary and Lars Kristiansen. A Friendly Introduction to
Mathematical Logic. Milne Library, 2015.
17
[4] David Marker. Model theory: an introduction. Springer, 2002.
[5] Rami Grossberg. Classification theory for abstract elementary classes.
In Yi Zhang, editor, Logic and Algebra. AMS, 2002.
[6] Michael Morley. Categoricity in power. Transactions of the American
Mathematical Society, 114(2), 1965.
[7] Wilfred Hodges. What is a structure theory? Bull. London Math. Soc.,
19:209 -- 237, 1987.
[8] Saharon Shelah. Classification theory. volume 92 of Studies in logic and
the foundations of mathematics. North-Holland, 2 edition, 1990.
[9] Flemming Larsen Mikael Rørdam and Niels J. Laustsen. An Introduction
to K-theory for C*-algebras. Cambridge University Press, 2000.
[10] James G. Glimm. On a certain class of operator algebras. Transactions
of the American Mathematical Society, 95(2):318 -- 340, 1960.
[11] George A. Elliott. On the classification of inductive limits of sequences
of semisimple finite-dimensional algebras. Journal of Algebra, 38:29 -- 44,
1976.
[12] Isaac Chuang Michael Nielsen. Quantum Computation and Quantum
Information: 10th Anniversary Edition. Cambridge University Press,
2010.
[13] I. Farah, B. Hart, M. Lupini, L. Robert, A. Tikuisis, A. Vignati, and
W. Winter. Model theory of c*-algebras. arXiv:1602.08072, 2018.
[14] David E. Handelman Edward G. Effros and Chao-Liang Shen. Dimen-
sion groups and their affine representations. American Journal of Math-
ematics, 102(2):385 -- 407, 1980.
[15] Ola Bratteli. Inductive limits of finite dimensional c*-algebras. Trans-
actions of the American Mathematical Society, 171:195 -- 234, 1972.
[16] Ilijas Farah. Logic for c*-algebras. Draft of a book, April 2019.
18
|
1608.04795 | 1 | 1608 | 2016-08-16T22:24:29 | Interpolation and Commutant Lifting with Weights | [
"math.OA"
] | Our two principle goals are generalizations of the commutant lifting theorem and the Nevanlinna-Pick interpolation theorem to the context of Hardy algebras built from $W^*$-correspondences endowed with a sequence of weights. These theorems generalize theorems of Muhly and Solel from 1998 and 2004, respectively, which were proved in settings without weights. Of special interest is the fact that commutant lifting in our setting is a consequence of Parrott's Lemma; it is inspired by work of Arias. | math.OA | math |
INTERPOLATION AND COMMUTANT LIFTING
WITH WEIGHTS
JENNIFER R GOOD
Abstract. Our two principle goals are generalizations of the com-
mutant lifting theorem and the Nevanlinna-Pick interpolation the-
orem to the context of Hardy algebras built from W ∗-correspon-
dences endowed with a sequence of weights. These theorems gener-
alize theorems of Muhly and Solel from 1998 and 2004, respectively,
which were proved in settings without weights. Of special interest
is the fact that commutant lifting in our setting is a consequence
of Parrott's Lemma; it is inspired by work of Arias.
1. Introduction
This paper concerns weighted interpolation problems in the theory of
operator algebras built from W ∗-correspondences. The classic origins
of this study are summarized in the phrase weighted interpolation. The
original interpolation problem, solved separately by Nevanlinna and
Pick in the early twentieth century, sought conditions under which,
given k ∈ N and two collections, {ωi}k
i=1 ⊆ C, there
exists a function φ ∈ H ∞(D) of norm at most 1 that interpolates this
data; that is,
i=1 ⊆ D and {λi}k
(1.1)
φ(ωi) = λi,
1 ≤ i ≤ k.
Nevanlinna and Pick showed that such a φ exists if and only if
is a positive semidefinite k × k matrix where K is the Szego kernel,
K(w, z) = 1
1−w¯z for w, z ∈ D, reproducing kernel to the Hardy space
H 2(D) [10] [13]. More generally, a weighted Hardy space, as described
in [17], is a reproducing kernel Hilbert space with kernel K : Dr ×Dr →
for a sequence of strictly positive
numbers β = {βi}∞
i=0. Our primary goal is a weighted Nevanlinna-
Pick type interpolation theorem that applies to the so-called weighted
C, defined by K(w, z) =P∞
(wz)i
i=0
β2
i
Key words and phrases. Nevanlinna-Pick interpolation, Commutant Lifting,
Double Commutant, W ∗-correspondence, noncommutative Hardy algebra, Se-
quence of Weights.
1
(1.2)
(cid:2)K(ωi, ωj)(1 − λiλj)(cid:3)k
i,j=1
2
JENNIFER R GOOD
Hardy algebra, a non-self-adjoint operator algebra recently studied in
[9].
While we will expand on the details in a later section, in brief the
setting of [9], inspired by Popescu's work in [14], begins with a W ∗-
algebra M, a W ∗-correspondence over M denoted E, and a sequence
of operator-valued weights Z that has been constructed from a so-
called admissible sequence X. Our algebra of focus is the weighted
Hardy algebra H ∞(E, Z), the algebra of adjointable operators on the
Fock space F(E) that is generated by the weighted creation operators
associated with Z and the left action maps of M on F(E). To gain
insight into H ∞(E, Z) we may view it as an algebra of functions on its
space of representations, as follows. Having fixed σ, a representation of
M on a Hilbert space H, the associated representations of H ∞(E, Z)
are in a sense determined by D(X, σ), a certain collection of operators in
B(E ⊗σ H, H). Every point z ∈ D(X, σ) gives rise to a representation
(σ × z) of H ∞(E, Z) on H, and we may consider an operator Y ∈
H ∞(E, Z) as a B(H)-valued function bY on D(X, σ) defined at z ∈
D(X, σ) by bY (z) = (σ × z)(Y ).
In this setting, let us preview the main theorem. We form a weighted
W ∗-version of the original Szego kernel, a map K from D(X, σ) ×
D(X, σ) to a collection of completely bounded maps on σ(M)′. For
any choice of k ∈ N, {zi}k
i=1 ⊆ D(X, σ), and two collections of opera-
tors {Bi}k
i=1 in B (H), there is an associated map A from
the k × k matrices with entries in σ(M)′ to the k × k matrices with
entries in B (H) that is defined at [Aij]k
i=1 and {Fi}k
i,j=1 to be
j − Fi · K(zi, zj)(Aij) · F ∗
.
i,j=1
j(cid:3)k
The conclusion of our weighted Nevanlinna-Pick theorem is that A is
completely positive if and only if there exists Y ∈ H ∞(E, Z) such that
kY k ≤ 1 and
(1.3)
(1.4)
(cid:2)Bi · K(zi, zj)(Aij) · B∗
Bi(cid:16)bY (zi)(cid:17) = Fi,
1 ≤ i ≤ k.
A comparison of equations (1.1) and (1.4), as well as matrices (1.2) and
(1.3), reveals the similarities with the original interpolation problem.
Indeed, when M = E = C, σ is the one-dimensional representation of
M, Z is the constant sequence at 1, and Bi = 1 for every i, matrix
(1.3) reduces to matrix (1.2) with minor technical adjustments, and we
obtain the original result of Nevanlinna and Pick. Even in the scalar
case, our theorem is more general than that of Nevanlinna and Pick, for
it applies to certain weighted Hardy spaces, precisely those that have
the so-called complete Pick property described in [1]. For example, our
INTERPOLATION AND COMMUTANT LIFTING WITH WEIGHTS
3
results apply to the Hardy and Dirichlet spaces, but not the Bergman
space.
Commutant lifting, pioneered by Sarason [16] and Sz.-Nagy and
Foia¸s [18], serves as the principle tool for the proof of Muhly and
Solel's unweighted Nevanlinna-Pick interpolation result, Theorem 5.3
in [8]. Wishing to follow suit, we must first establish that commutant
lifting can be done in the weighted case. That is, an operator on a
co-invariant subspace for an induced representation of H ∞(E, Z) that
commutes with the compression of the image of the representation may
be lifted to a commuting operator on the full induced space without
increasing the norm. As it turns out, our proof of the weighted com-
mutant lifting theorem is where we differ most substantially from the
unweighted case. Muhly and Solel's commutant lifting result, Theo-
rem 4.4 of [7], is proven using so-called isometric dilations, but in our
setting, the weights create obstacles that make this method difficult or
even impossible. Instead, we adapt the proof for a commutant lifting
theorem given by Arias in [2], an argument that ultimately makes use
of Parrott's lemma [11]. Along with the results themselves, this new
W ∗-approach to commutant lifting utilizing technology that works in
our more general setting is one of the paper's primary attractions. The
W ∗-setting of our results provides for a generality that encompasses,
for instance, the unweighted commutant lifting and Nevanlinna-Pick
theorems of Muhly and Solel in [7] and [8] and the weighted lifting and
interpolation theorems of Popescu in [14].
The paper is organized as follows. In Section 2 we establish defini-
tions and notation; for convenience, we have organized this material
into three subsections. Section 3 is a technical section devoted to the
construction of a family of orthonormal bases in preparation for the
weighted commutant lifting theorem, Theorem 4.1, the principal result
of Section 4. In Section 5 we give a weighted double commutant theo-
rem, Theorem 5.3, that extends its unweighted analogue in [8]. Section
6 contains our main result, the weighted Nevanlinna-Pick interpolation
theorem, Theorem 6.5.
2. Preliminaries
2.1. W ∗-correspondences and the Unweighted Hardy Algebra.
We let N0 = N∪{0}. Hilbert spaces have inner products that are linear
in the second variable and conjugate linear in the first. Throughout
the paper, M will denote a W ∗-algebra, that is a C ∗-algebra that is
4
JENNIFER R GOOD
also a dual space, thought of abstractly, without reference to a partic-
ular representation on Hilbert space. Likewise, E will denote a W ∗-
correspondence over M in the sense of Section 2 in [8]. That is, E is a
self-dual Hilbert C ∗-module over M in the sense of [5] and [12] with a
second-linear inner product that is also a left M-module with respect
to a faithful normal ∗-homomorphism ϕ : M → L(E) where L(E) de-
notes the W ∗-algebra of adjointable operators on E. For simplicity, we
assume that ϕ is unital. At certain points, we will add the assumption
that the right action of E is full in the sense that the ultraweakly closed
linear span of {hξ, ηi ξ, η ∈ E} is all of M. By Proposition 3.8 of [12],
E is a dual space, and we refer to the weak-∗ topology on E as its
ultraweak topology. If E is nonzero, then E has an orthonormal basis,
that is, a family A ⊆ E that is maximal with respect to the following
two properties: for every α ∈ A, hα, αi is a nonzero projection in N,
and if α, β ∈ A and α 6= β, then hα, βi = 0 ([12], proof of Theorem
3.12).
If N is a W ∗-algebra and F is an (M, N) W ∗-correspondence with
left action map σ : M → L(F ), then there is an (M, N) W ∗-correspon-
dence, denoted E ⊗σ F , formed by taking the self-dual completion of
a quotient of the algebraic tensor product of E and F , balanced over
N; the quotient is determined by the semi inner product satisfying
hξ1 ⊗ η1, ξ2 ⊗ η2i = hη1, σhξ1, ξ2i(η2)i for ξ1, ξ2 ∈ E and η1, η2 ∈ F . To
give the left action of M on E⊗σF , first form the induced representation
of L(E), σE : L(E) → L(E ⊗σ F ), defined at S ∈ L(E) by σE(S) =
S ⊗ IF , as in [15]. Then the left action of M on E ⊗σ F is σE ◦ ϕ,
called the induced representation of M, mapping a ∈ M to ϕ(a)⊗IF in
L(E ⊗σ F ). We observe that an (M, C) W ∗-correspondence is simply
a Hilbert space H together with a normal unital ∗-homomorphism σ :
M → B(H); thus, taking F to be H we obtain the Hilbert space
E ⊗σ H, called the induced representation space, and the representation
σE ◦ ϕ : M → B(E ⊗σ H). If, instead, we inductively take F to be
E in the tensor product construction, we form the tensor powers of E,
k=0, a family of W ∗-correspondences over M. We write ϕk for
{E⊗k}∞
the left action of M on E⊗k for each k ∈ N0. To be precise, E⊗0 := M,
E⊗1 := E, and E⊗k := E ⊗ϕk−1 E⊗k−1 for k ≥ 2 where ϕ0 is left
multiplication and ϕ1 = ϕ.
powers of E; that is F(E) := P∞
The Fock space of E, F(E), is the ultraweak direct sum of the tensor
k=0 ⊕E⊗k. The Fock space is a W ∗-
correspondence over M with respect to the left action map ϕ∞ : M →
L(F(E)), defined at a ∈ M by ϕ∞(a) = diag[ϕ0(a), ϕ1(a), ϕ2(a), . . .].
Along with the class of left action maps, a second important class of
Tξ =
0
T (0)
ξ
0
0
0
0
T (1)
ξ
0
0
0
0
T (2)
ξ
. . .
. . .
. . .
,
ξ
INTERPOLATION AND COMMUTANT LIFTING WITH WEIGHTS
5
operators in L(F(E)) is the class of creation operators. The creation
operator Tξ determined by ξ ∈ E is defined at η ∈ F(E) by Tξ (η) =
ξ ⊗ η; matricially, Tξ has a subdiagonal matrix,
where for each j ∈ N0, T (j)
: E⊗j → E⊗(j+1) is defined at η ∈ E⊗j
by T (j)
ξ (η) = ξ ⊗ η. For arbitrary k ∈ N0 and ξ ∈ E⊗k, the creation
operator Tξ is defined in an analogous fashion. The algebraic tensor
algebra, T 0
+ (E), is the subalgebra of L(F(E)) generated by the left
action and creation operators. The principal object of study in the
unweighted setting of [8] is the Hardy algebra, H ∞(E), the closure of
+ (E) in the ultraweak topology of the W ∗-algebra L(F(E)).
T 0
If F is an (M, N) W ∗-correspondence with left action σ and ξ ∈
E, the (left) insertion operator associated with ξ is the map LF
ξ ∈
L(F, E ⊗σ F ), defined at η ∈ F by LF
ξ (η) = ξ ⊗ η. We omit the
superscript F if it is clear from context. We note that a creation
operator is a specific example of an insertion operator. The following
facts are easily verified for ξ, η ∈ E. The operator Lξ is bounded and
kLξk ≤ kξk. For a ∈ M and c ∈ N, La·ξ·c = (ϕ(a) ⊗ IF )Lξσ(c).
If S ∈ L(E), T ∈ L(F ), and T is a left M-module homomorphism,
LSξT = (S ⊗ T )Lξ. For any ζ ∈ F , L∗
ξ(η ⊗ ζ) = hξ, ηi · ζ; hence,
L∗
η = θξ,η ⊗ IF where θξ,η ∈ L(E) denotes
the rank one operator defined at ζ ∈ E by θξ,η(ζ) = ξ · hη, ζi.
ξLη = σhξ, ηi. Finally, LξL∗
2.2. Duality. Let σ : M → B(H) be a normal, unital ∗-homomor-
phism for a Hilbert space H. If ψ : M → B(K) is another such map
for Hilbert space K, I(σ, ψ) denotes the space of intertwiners, i.e. the
collection of operators T ∈ B(H, K) such that T ◦ σ(a) = ψ(a) ◦ T
for every a ∈ M. With this notation, we define Eσ := I(σ, σE ◦ ϕ),
called the σ-dual of E, a subspace of B(H, E ⊗σ H). By Proposition
3.2 of [8], Eσ is a W ∗-correspondence over the W ∗-algebra σ(M)′ with
a · ξ · b := (IE ⊗ a)ξb and hξ, ηi := ξ∗η for a, b ∈ σ(M)′ and ξ, η ∈ Eσ.
We write ϕ′ : σ(M)′ → L(Eσ) for the left action map.
Let us recall several maps that arise in [8], simultaneously providing
notation for future use.
In this subsection, we will suppose that E
is full. For k ∈ N, there is a well-defined σ(M)′ W ∗-correspondence
6
JENNIFER R GOOD
isomorphism Λσ
k : (Eσ)⊗k → (E⊗k)σ with
k
⊗
i=1
ξi ∈ (Eσ)⊗k
Λσ
k(cid:16) k
⊗
i=1
ξi(cid:17) = (Ik−1 ⊗ ξ1) · · · (I2 ⊗ ξk−2)(I1 ⊗ ξk−1)ξk,
k LH
0 : (Eσ)⊗0 → (E⊗0)σ at A ∈ σ(M)′ and h ∈ H by (Λσ
where Ij denotes the identity operator on E⊗j. To include k = 0, define
Λσ
0 (A))(h) = 1M ⊗
Ah. The inclusion map ι : σ(M)′ → B(H) is a faithful, normal, unital
k : (Eσ)⊗k ⊗ι H →
∗-homomorphism, and for every k ∈ N0, the map U σ
E⊗k ⊗σ H defined for ξ ∈ (Eσ)⊗k and h ∈ H by U σ
k (ξ ⊗ h) = Λσ
k(ξ)(h)
is a Hilbert space isomorphism identifying the induced representation
k=0 ⊕U σ
spaces. Equivalently, U σ
k
identifies the spaces F(Eσ) ⊗ι H and F(E) ⊗σ H ([8]; Lemma 3.8). Let
Ad(U σ
∞) : B (F(Eσ) ⊗ι H) → B (F(E) ⊗σ H), denote the isomorphism
that sends an operator T ∈ B (F(Eσ) ⊗ι H) to U σ
∞ . It follows that
∞) ◦ ιF(Eσ) is a
ρσ : L(F(Eσ)) → B(F(E) ⊗σ H), defined by ρσ = Ad(U σ
faithful, normal, unital ∗-homomorphism; ρσ is called "ρ" in Theorem
∞(Y ⊗ IH )U σ∗
3.9 of [8]. Specifically, for Y ∈ L(F(Eσ)), ρσ(Y ) = U σ
∞ .
The map πσ : L(F(E)) → B(F(Eσ) ⊗ι H), defined by πσ = Ad(U σ∗
∞ ) ◦
σF(E) is also a faithful, normal, unital ∗-homomorphism; πσ is called
"ρ" in Section 5 of [8]. For Y ∈ L(F(E)), πσ(Y ) = U σ∗
∞ := P∞
k(ξ). The map U σ
ξ = Λσ
∞T U σ∗
∞ (Y ⊗ IH)U σ
∞.
k, U σ
k , U σ
To summarize, using the W ∗-algebra M, the W ∗-correspondence over
the algebra E, and the representation of the algebra σ, we've defined
the maps Λσ
∞, ρσ, and πσ. We repeat the process with the
W ∗-algebra σ(M)′, the correspondence Eσ, and the representation ι.
The analogue of σ(M)′ is the commutant of the image of ι in B(H),
which is precisely σ(M). The analogue of ι is the inclusion map that
we denote by : σ(M) → B(H). We write Eσι in place of (Eσ)ι, a W ∗-
∞ : σ(M) → L(F(Eσι)).
correspondence over σ(M) with left action ϕ′′
We obtain the analogous collection of maps Λι
∞, ρι, and πι.
Repeating the process once again would utilize the W ∗-algebra σ(M),
the correspondence Eσι, and the representation , but if σ is faithful
this is unnecessary since these constructs may be naturally identified
with M, E, and σ as follows. The identifications of M with σ(M)
and σ with are immediate. After appropriate identifications, the map
ω : E → Eσι defined at ξ ∈ E by ω(ξ) = U σ∗
is an isomorphism
of W ∗-correspondences that is studied in and around Theorem 3.6 of
[8]. Defining ωk := ω ⊗ · · · ⊗ ω identifies E⊗k with (Eσι)⊗k for each k,
k=0 ωk identifies F(E) with F(Eσι). For every k ∈ N0,
k(ωk ⊗ IH) = IE⊗k⊗σH , as demonstrated in and around Corollary
for every ξ ∈ E⊗k. More
U σ
m ⊗ Λι
m for k, m ∈ N0 and
m denotes the identity operator on (Eσ)⊗m. It follows
U σ
3.10 of [8]. Equivalently U σ
generally, U σ
k+m (I ′
ξ ∈ E⊗k, where I ′
and ω∞ := P∞
k(ωkξ)) = L(E⊗m⊗σH)
k(ωkξ) = LH
ξ
k ◦ Λι
1 LH
ξ
k, U ι
k, U ι
k U ι
ξ
πσ(Tξ) =
0
ωξ
0
0
0
I ′
1 ⊗ ωξ
0
0
0
0
0
I ′
2 ⊗ ωξ
. . .
. . .
. . .
.
INTERPOLATION AND COMMUTANT LIFTING WITH WEIGHTS
7
∞U ι
that U σ
how πσ acts on the generators of H ∞(E):
I ′
∞ ⊗ σ(a), and for ξ ∈ E, πσ(Tξ) has the subdiagonal matrix,
∞(ω∞ ⊗IH ) = IF(E)⊗σH . Equations (5.1) and (5.2) of [8] show
for a ∈ M, πσ(ϕ∞(a)) =
2.3. The Weighted Hardy Algebra. We now turn to the weighted
setting of [9]. Throughout, the sequence X = {Xk}∞
k=0 will denote an
admissible sequence; this means that Xk ∈ ϕk(M)c for each k ∈ N0,
Xk ≥ 0 for each k ∈ N0, X0 = 0, X1 is invertible, and finally that
lim supk→∞ kXkk1/k < ∞, where ϕk(M)c denotes the commutant of
ϕk(M) in L(E⊗k) ([9], Definition 4.1). Let R = {Rk}∞
k=0 denote the
sequence
IM ,
Rk =
if k = 0
,
if k > 0
,
i=1 α(i) = ko if 1 ≤ j ≤ k
j=1(cid:16)Pα∈F (k,j)
(cid:16)Pk
Xα(i)(cid:17)(cid:17)1/2
where F (k, j) := nα : {1, . . . , j} → N(cid:12)(cid:12)Pj
j
⊗
i=1
([9], Equation 4.4). Each Rk is a positive, invertible element in ϕk(M)c.
A sequence of operators Z = {Zk}∞
k=0 is called a sequence of weights
associated with X if Zk is invertible and belongs to ϕk(M)c for each k ∈
N0, Z0 = IM , and Z (k)∗Z (k) = R−2
for all k ∈ N0, where Z (k), thought of
k
as the product of the weights {Zk, Zk−1, . . . , Z1, Z0}, is defined Z (k) :=
Zk(I1 ⊗ Zk−1) · · · (Ik−1 ⊗ Z1)(Ik ⊗ Z0) ∈ ϕk(M)c ([9], Definition 4.6).
We define Z (k,j) := Z (k)(Ik−j ⊗ Z (j))−1 when 0 ≤ j ≤ k, equivalently
Z (k,j) = Zk(I1 ⊗ Zk−1) · · · (Ik−j−1 ⊗ Zj+1), which may be thought of
as the product of the weights {Zk, Zk−1, . . . , Zj+1}. One example of a
weight sequence is the canonical sequence Z = {Zk}∞
k=0, with Z0 := IM
and Zk := R−1
k (I1 ⊗ Rk−1), k ≥ 1. We note that the unweighted setting
studied, for instance, in [7] and [8] occurs when Z is the sequence
of identity operators. To work in the setting of [14], we take the W ∗-
correspondence E to be Cd and add the condition that X be a sequence
of diagonal operators.
If Z is a sequence of weights associated with X, then supj∈N0 kZjk <
∞ ([9], proof of Theorem 4.5). It follows that for every k ∈ N0 the
operator DZ
k = Dk := diag[0, . . . , 0, Z (k), Z (k+1,1), Z (k+2,2), . . .] belongs
to L(F(E)). We omit the superscript 'Z' when the weight sequence
8
JENNIFER R GOOD
is understood. For ξ ∈ E⊗k, we define the weighted creation operator
W Z
ξ = Wξ in L(F(E)) by Wξ := Dk · Tξ. The matrix for Wξ is k-
subdiagonal;
Wξ =
Z (k)T (0)
ξ
0
...
0
0
...
0
...
0
Z (k+1,1)T (1)
ξ
0
0
...
0
Z (k+2,2)T (2)
ξ
. . .
. . .
. . .
.
where T (j)
ξ maps η ∈ E⊗j to ξ ⊗ η ∈ E⊗j+k for j ∈ N0. We observe
that while our arrival at the definition of Wξ differs from that in [9], a
comparison of the preceding matrix with equation (3.2) in [9] confirms
that the two definitions agree. When k, l ∈ N0; a, b ∈ M; ξ ∈ E⊗k;
and η ∈ E⊗l; Wa·ξ·b = ϕ∞(a) ◦ Wξ ◦ ϕ∞(b) and Wξ ◦ Wη = Wξ⊗η.
For a ∈ M, Wa = ϕ(a). The Z-algebraic tensor algebra, T 0
+ (E, Z), is
the subalgebra of L(F(E)) generated by the left action and weighted
creation operators. The Z-Hardy algebra, H ∞(E, Z), is the closure of
T 0
+ (E, Z) in the ultraweak topology of L(F(E)).
Let σ : M → B(H) be a normal, unital ∗-homomorphism. We
observe that I(σE ◦ ϕ, σ)∗ = I(σ, σE ◦ ϕ) = Eσ, which was defined
above. For a point z ∈ I(σE ◦ ϕ, σ) and k ∈ N0, we define the kth
(k) := z(I1⊗z) · · · (Ik−1⊗z). The intertwining property
tensorial power, z
of z guarantees that z(k) is a well-defined operator in B(E⊗k ⊗σ H, H);
moreover, z(k) ∈ I(σE⊗k ◦ ϕk, σ). Computation shows that z(k+l) =
z(k)(Ik ⊗ z(l)) for k, l ∈ N0. We define
D(X, σ) :=(z ∈ I(σE ◦ ϕ, σ)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
∞Xk=1
(k)(Xk ⊗ IH )z
z
< 1) ;
(k)∗(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
([9], Definition 4.3). Here and throughout unless stated otherwise, infi-
nite sums indicate convergence with respect to the ultraweak topology
for a W ∗-correspondence or W ∗-algebra and the norm topology for a
Hilbert space. If z ∈ D(X, σ), there is an ultraweakly continuous, com-
pletely contractive representation (σ × z) : H ∞(E, Z) → B(H) such
that (σ × z)(ϕ∞(a)) = σ(a) for every a ∈ M and (σ × z)(Wξ) = z(k)Lξ
for every ξ ∈ E⊗k with k ≥ 1 ([9], Corollary 5.9). For Y ∈ H ∞(E, Z),
we define the noncommutative function bY : D(X, σ) → B(H) at the
point z ∈ D(X, σ) by bY (z) = (σ × z)(Y ).
INTERPOLATION AND COMMUTANT LIFTING WITH WEIGHTS
9
3. An Orthonormal Basis for Tensor Products
The proof of our weighted commutant lifting theorem in Section 4
is modeled after the proof of a commutant lifting theorem by Alvaro
Arias in [2]. Working in the setting of complex n-space, Arias makes
use of the fact that if {ei}n
i=1 is an orthonormal basis for Cn and k ∈ N,
then one orthonormal basis for (Cn)⊗k consists of the simple tensors
k
eα(i) such that α is any function from {1, . . . , k} to {1, . . . , n}. In
⊗
i=1
search of a similarly constructed orthonormal basis for E⊗k or, more
generally, for any correspondence formed by tensor product, let us con-
sider an obstacle that must be overcome. Let F be a nonzero (M, N)
W ∗-correspondence with orthonormal basis A, and let G be a nonzero
(N, P ) W ∗-correspondence with left action σ and orthonormal basis B
for W ∗-algebras M, N, and P . If M = N = P = C, then F and G are
simply Hilbert spaces, the W ∗-tensor product of F and G is the same
as their Hilbert space tensor product, and {α ⊗ β : α ∈ A, β ∈ B},
which is a Hilbert space orthonormal basis for the tensor product, is
also an orthonormal basis in the sense of Hilbert W ∗-modules. For
general M, N, and P , the issue is more complicated. If α ∈ A and
β ∈ B, there is no reason to think that hα ⊗ β, α ⊗ βi is a nonzero
projection in P , in which case α ⊗ β cannot belong to an orthonormal
basis for F ⊗σ G. In fact, it is possible that F ⊗σ G is zero, in which
case F ⊗σ G does not have an orthonormal basis. We begin this sec-
tion by overcoming these obstacles to construct an orthonormal basis
for a nonzero tensor product of two W ∗-correspondences that consists
of simple tensors. Then we obtain, via an inductive process, a family
of orthonormal bases, one basis for each nonzero E⊗k, which we use
in Section 4 to prove the weighted commutant lifting theorem using
Arias' technique.
Along with M, let N and P be W ∗-algebras. Let F be an (M, N)
W ∗-correspondence, and let G be an (N, P ) W ∗-correspondence with
left action σ : N → L(G). We will write a · η in place of σ(a)(η) for
a ∈ N and η ∈ G. If q is a projection in N, then q · G = {q · η η ∈ G}
is the kernel of σ(1N − q). Thus q · G is an ultraweakly closed (right)
P -submodule of G and is therefore a Hilbert W ∗-module over P ([3];
Consequence 1.8). As such, for every projection q ∈ N such that
q · G 6= {0} we may fix an orthonormal basis for q · G. Define Q to be
the, possibly empty, subset of F ,
Q := {ξ ∈ F hξ, ξi is a projection in N and hξ, ξi · G 6= {0}} .
If Q is nonempty and ξ ∈ Q, define B(ξ) to be the orthonormal basis
for hξ, ξi·G chosen above. Fix A, an orthonormal basis for F . If α ∈ A,
10
JENNIFER R GOOD
then hα, αi is a projection in N. The subset of F ⊗σ G,
C := {α ⊗ β α ∈ A ∩ Q and β ∈ B(α)},
is empty if and only if A ∩ Q is empty. For simplicity, when we say that
"α ⊗ β ∈ C", we mean that α ∈ A ∩ Q and β ∈ B(α). The emphasis is
needed since the expression of an element in F ⊗σ G in terms of simple
tensors is not unique. If α ⊗ β ∈ C, then hα, αi · β = β. In our first
theorem, we show that when F ⊗σ G is nonzero, C is an orthonormal
basis.
Theorem 3.1. The correspondence F ⊗σ G is nonzero if and only if
C is nonempty. In this case, C is an orthonormal basis for F ⊗σ G,
and if αi ⊗ βi ∈ C for i = 1, 2, then
if α1 = α2 and β1 = β2
otherwise
.
if A ∩ Q 6= ∅
if A ∩ Q = ∅
.
(3.1)
(3.2)
hα1 ⊗ β1, α2 ⊗ β2i =(hβ2, β2i
ξ ⊗ η =(Pα∈A∩Q α ⊗ hα, ξi · η
0
0
Proof. First, we show that for any ξ ∈ F and η ∈ G,
By Fourier expansion with respect to A, we have ξ =Pα∈A α · hα, ξi
([12], proof of Theorem 3.12). It follows that ξ ⊗ η =Pα∈A α ⊗ hα, ξi ·
η. For every α ∈ A, we have hα · hα, αi − α, α · hα, αi − αi = 0, so
α · hα, αi = α. Thus hα, ξi · η belongs to hα, αi · G, so if α /∈ Q, then
hα, ξi · η = 0. Equation (3.2) follows.
When C is empty, A ∩ Q is also empty, so by equation (3.2) every
simple tensor in F ⊗σ G is zero. Thus F ⊗σ G = {0} since F ⊗σ G is
the ultraweak closure of the linear span of the simple tensors.
Suppose that C is not empty. Using the properties of A and C,
equation (3.1) follows from straightforward computations.
It follows
that for any α ⊗ β ∈ C, hα ⊗ β, α ⊗ βi is a nonzero projection in P ,
so α ⊗ β 6= 0 and F ⊗σ G 6= {0}. It only remains to show that C is an
orthonormal basis for F ⊗σ G. By equation (3.1), C is an orthonormal
set. Towards maximality, suppose that C ′ is an orthonormal subset of
F ⊗σ G such that C ⊆ C ′. For any ξ ∈ F and η ∈ G, if α belongs
to A ∩ Q, then as above, hα, ξi · η is an element in hα, αi · G. Using
the Fourier expansion with respect to the orthonormal basis B(α),
hα, ξi · η =Pβ∈B(α) β · hβ, hα, ξi · ηi =Pβ∈B(α) β · hα ⊗ β, ξ ⊗ ηi. Thus,
by equation (3.2),
(3.3)
ξ ⊗ η = Xα∈A∩Q
Xβ∈B(α)
(α ⊗ β) · hα ⊗ β, ξ ⊗ ηi .
INTERPOLATION AND COMMUTANT LIFTING WITH WEIGHTS
11
Now if C is properly contained in C ′, then there exists ζ ∈ C ′ such
that for every α ⊗ β ∈ C, hζ, α ⊗ βi = 0. Since F ⊗σ G is the ul-
traweak closure of the linear span of the simple tensors, there is a net
converging ultraweakly to ζ in F ⊗σ G. By equa-
tion (3.3),
nPNλ
j=1 ξλ,j ⊗ ηλ,joλ
NλXj=1
Xα∈A∩Q
NλXj=1
hζ, ζi = lim
λ
= lim
λ
hζ, ξλ,j ⊗ ηλ,ji
Xβ∈B(α)
hζ, α ⊗ βihα ⊗ β, ξλ,j ⊗ ηλ,ji = 0,
contradicting the fact that hζ, ζi is a nonzero projection. Therefore, C
is an orthonormal basis for F ⊗σ G.
(cid:3)
Let us inductively construct a family {Ck} of orthonormal bases,
one basis for each nonzero E⊗k. First we establish some notation,
mirroring the discussion prior to Theorem 3.1. Fix an orthonormal
basis for the Hilbert W ∗-module q · E for every projection q ∈ M such
that q · E 6= {0}. Define
Q = {ξ ∈ E hξ, ξi is a projection in M and hξ, ξi · E 6= {0}} .
If ξ ∈ Q, let B(ξ) denote the orthonormal basis for hξ, ξi · E chosen
above. Since E is nonzero, let A be an orthonormal basis for E. Let
C0 = {1M } where 1M is the identity element in M. Let C1 = A. When
k ≥ 2, define Ck to be the possibly empty subset of E⊗k consisting of
ξi = ξ1⊗· · ·⊗ξk such that the following properties
the simple tensors
hold:
k
⊗
i=1
(1) ξ1 ∈ A ∩ Q,
(2) ξi ∈ B(ξi−1) ∩ Q for every i such that 1 < i < k, and
(3) ξk ∈ B(ξk−1).
For simplicity, when we say that
ξi ∈ Ck, we always mean that
k
⊗
i=1
i=1 satisfies properties (1), (2), and (3). Notice that if
{ξi}k
ξi ∈ Ck,
then for every i such that 1 < i ≤ k, we have hξi−1, ξi−1i · ξi = ξi. We
arrive at the following theorem.
k
⊗
i=1
Theorem 3.2. For k ∈ N0, the correspondence E⊗k is nonzero if and
only if Ck is nonempty. In this case, Ck is an orthonormal basis for
12
JENNIFER R GOOD
k
⊗
i=1
E⊗k, and for any ξ =
(3.4) hξ, ηi =(hηk, ηki
0
ξi and η =
k
⊗
i=1
ηi in Ck,
if ξi = ηi for every i such that 1 ≤ i ≤ k
otherwise
.
ξi(cid:17) ⊗ ξk+1(cid:12)(cid:12)(cid:12)(cid:12)
Proof. The cases when k = 0 and k = 1 follow readily. Inductively,
suppose the conclusions of the theorem are satisfied for some k ∈ N.
Towards applying Theorem 3.1 to E⊗k ⊗ϕ E, define
If ξ ∈ Qk, let B(ξ) denote the orthonormal basis for hξ, ξi · E. Define
Qk :=(cid:8)ξ ∈ E⊗k hξ, ξi is a projection in M and hξ, ξi · E 6= {0}(cid:9) .
ξi(cid:17)(cid:27) .
k+1 :=(cid:26)(cid:16) k
ξi ∈ Ck ∩ Qk and ξk+1 ∈ B(cid:16) k
k
⊗
i=1
⊗
i=1
⊗
i=1
C ′
Using the inductive assumption we see that Ck+1 is the image of C ′
k+1
under the natural isomorphism between E⊗k ⊗ϕ E and E⊗k+1. It now
follows from Theorem 3.1 that E⊗k+1 is nonzero if and only if Ck+1
is nonempty and that Ck+1 is an orthonormal basis in that case.
If
ξ =
ηi belong to Ck+1, then (cid:16) k
(cid:16) k
ηi(cid:17) ⊗ ηk+1 belong to C ′
k+1 and(cid:16) k
ξi(cid:17) and(cid:16) k
Since hηk, ηki · ηk+1 = ηk+1 and hξ, ηi = Dξk+1,D k
ξi(cid:17) ⊗ ξk+1 and
ηi(cid:17) belong to Ck.
ηiE · ηk+1E,
equation (3.4) follows from Theorem 3.1 and the inductive hypothesis.
(cid:3)
ξi and η =
k
⊗
i=1
⊗
i=1
k+1
⊗
i=1
k+1
⊗
i=1
⊗
i=1
⊗
i=1
⊗
i=1
⊗
i=1
ξi,
We devote the remainder of the section to establishing notation
and giving technical lemmas that will be useful in proving our main
theorems. Define S := (cid:8)k ∈ N0 E⊗k 6= {0}(cid:9), which is either N0 or
{0, 1, . . . , n} for some n ∈ N. For k ∈ N0, define Ak to be Ck when
k ∈ S and the zero set when k /∈ S. If k ∈ N0 then vk ∈ L(E⊗k, F(E))
denotes the isometry mapping ξ ∈ E⊗k to bξ := (δi=k ξ)∞
i=0 where
δi=k is 1 when i = k and is otherwise zero. A handy fact is that
a diagonal operator S = diag[S0, S1, S2, · · · ] ∈ L(F(E)) may be ex-
pressed as S =P∞
j . The insertion operator Lbξ maps x ∈ H
to bξ ⊗ x ∈ F(E) ⊗σ H. We define Qk in L(F(E)) to be the projection
onto vk(E⊗k), an ultraweakly closed submodule of F(E), noting that
Qk = vkv∗
k. For ξ ∈ E⊗k, θξ ∈ L(E⊗k) denotes the positive rank one
operator that sends η ∈ E⊗k to ξ · hξ, ηi.
j=0 vjSjv∗
Lemma 3.3. For k ∈ N0,
(1) if S ∈ L(E⊗k, F ) for a Hilbert W ∗ M-module F , then SS∗ =
θSξ in L(F );
Pξ∈Ak
INTERPOLATION AND COMMUTANT LIFTING WITH WEIGHTS
13
(2) Qk ⊗ IH =Pξ∈Ak
(3) Pξ∈Ak
LbξL∗
bξ
closed submodule {(ζi)∞
i=0 ζi = 0 when 0 ≤ i < k} of F(E).
ξ is the projection in L(F(E)) onto the ultraweakly
in B(F(E) ⊗σ H); and
TξT ∗
Proof. If k /∈ S, then E⊗k = {0}, and all of the operators in parts (1),
(2), and (3) are zero. Suppose k ∈ S. For η ∈ E⊗k and F a finite subset
of Ak,
θξ! η, η+ =Xξ∈F
0 ≤* Xξ∈F
hη, ξihξ, ηi ≤ Xξ∈Ak
fore,Pξ∈Ak
algebra L(E⊗k) and for any η ∈ E⊗k, D(cid:16)Pξ∈Ak
Thus by polarization,
by a Parseval type identity ([12], proof of Theorem 3.12). There-
θξ converges ultraweakly to a positive operator in the W ∗-
θξ(cid:17) η, ηE = hη, ηi.
hη, ξihξ, ηi = hη, ηi
Ik = Xξ∈Ak
θξ.
(3.5)
(3.6)
(3.6),
P∞
Xξ∈Ak
The map Ad(S) : L(E⊗k) → L(F ) that sends Y ∈ L(E⊗k) to SY S∗ is
linear and ultraweakly continuous, and Ad(S)(θξ) = θSξ for any ξ ∈ Ak.
Therefore, part (1) follows from applying Ad(S) to equation (3.5).
Taking S to be vk in part (1), we obtainPξ∈Ak
Since σF(E) is linear and ultraweakly continuous, Qk ⊗IH =Pξ∈Ak
IH =Pξ∈Ak
θbξ = Qk in L(F(E)).
θbξ ⊗
i−k is
linear and ultraweakly continuous, so it follows from equation (3.5)
that
Let i ∈ N0 such that i ≥ k. The induced representation ϕE⊗k
LbξL∗
, giving part (2).
bξ
Ii = Xξ∈Ak
θξ ⊗ Ii−k .
For any ξ ∈ Ak, we have TξT ∗
i=k vi(θξ ⊗ Ii−k)v∗
ξ = diag[0, . . . , 0, θξ, θξ ⊗ I1, θξ ⊗ I2, . . .] =
i . Summing over all ξ ∈ Ak and using equation
TξT ∗
ξ = Xξ∈Ak ∞Xi=k
vi(θξ ⊗ Ii−k)v∗
i!
vi Xξ∈Ak
=
∞Xi=k
θξ ⊗ Ii−k! v∗
i =
Qi.
∞Xi=k
14
JENNIFER R GOOD
noting that sums may be interchanged as all terms are non-negative.
i=k Qi is the projection in L(F(E)) onto the
(cid:3)
It is readily shown thatP∞
ultraweakly closed submodule {(ζi)∞
i=0 ζi = 0 when 0 ≤ i < k}.
The following technical lemma is similar in flavor to Lemma 5.2 of
[9].
Lemma 3.4. If Z is a sequence of weights associated with X,
WX 1/2
∞Xj=1
ξ = IF(E) − Q0
Proof. First we note that if i ∈ N then Pi
equation (4.7) of [9], so
ξW ∗
X 1/2
j
j
in L(F(E)).
j=1 Xj ⊗ R2
i−j = R2
i by
(3.7)
Z (i)(Xj ⊗ R2
i−j)Z (i)∗ = Z (i)R2
i Z (i)∗ = Ii.
Also when j ∈ N,Pξ∈Aj
(3.8)
θX 1/2
j
ξ = Xj by Lemma 3.3. Thus if i ≥ j ≥ 1,
Z (i)(θX 1/2
ξ ⊗ R2
i−j)Z (i)∗ = Z (i)(Xj ⊗ R2
i−j)Z (i)∗.
j
For j ∈ N and ξ ∈ Aj, computation shows that WX 1/2
a diagonal matrix whose ith diagonal entry is Z (i)(θX 1/2
when i ≥ j and is otherwise zero. Thus by equation (3.8),
j
j
has
ξW ∗
ξ ⊗ R2
j
ξ
X 1/2
i−j)Z (i)∗
Xξ∈Aj
iXj=1
Xξ∈Aj
Xξ∈Aj
WX 1/2
j
ξW ∗
X 1/2
ξ
j
viZ (i)(θX 1/2
ξ ⊗ R2
j
Z (i)(θX 1/2
j
ξ ⊗ R2
i−j)Z (i)∗v∗
i!
i−j)Z (i)∗ v∗
i
viZ (i)(Xj ⊗ R2
i−j)Z (i)∗v∗
i .
=
= Xξ∈Aj ∞Xi=j
viXξ∈Aj
∞Xi=j
∞Xi=j
ξ =
=
∞Xj=1 ∞Xi=j
i−j)Z (i)∗! v∗
Z (i)(Xj ⊗ R2
ξW ∗
X 1/2
j
viZ (i)(Xj ⊗ R2
i−j)Z (i)∗v∗
i =
viv∗
i = IF(E) − Q0,
∞Xi=1
i!
∞Xj=1
j
WX 1/2
Xξ∈Aj
vi iXj=1
∞Xi=1
=
Summing over j and using equation (3.7),
INTERPOLATION AND COMMUTANT LIFTING WITH WEIGHTS
which completes the proof.
15
(cid:3)
To "factor" an element ξ ∈ Ak when k ∈ N we establish the following
notation. If k ∈ S, ξ =
ξj :=
j
⊗
i=1
ξi ∈ E⊗j
k
⊗
i=1
ξi ∈ Ak, and 1 ≤ j ≤ k define
and ξj
◦ :=( k
ξi
i=j+1
hξ, ξi
⊗
if j < k
if j = k
∈ E⊗k−j.
If k /∈ S, let ξj and ξj
◦ to be the zero elements in their respective spaces.
Lemma 3.5.
(1) For k ∈ N0 and ξ ∈ Ak, hξ, ξi is a projection in M that is zero
if and only if k /∈ S. In any case, ξ · hξ, ξi = ξ.
(2) For j ∈ N, l ∈ N0, and ξ ∈ Al+j, we have ξ = ξj ⊗ ξj
◦ and
hξj, ξji · ξj
◦ = ξj
◦. If l + j ∈ S, then ξj ∈ Aj.
(3) If j ∈ N, l ∈ N0, and l + j ∈ S, then as a disjoint union,
(4) For j ∈ N and ξ ∈ Aj,
Al+j = aξ∈Aj
{η ∈ Al+j ηj = ξ}.
ϕ∞hξ, ξi =
∞Xl=0
Xη∈Al+j
ηj =ξ
,
θcηj
◦
where any sum taken over an empty set is assumed to be zero.
vl(ϕlhξ, ξi)v∗
Proof. The proofs of the first three assertions are straightforward. To
show part (4), let l ∈ N0. If η ∈ Al+j, then T (l)∗
◦ which,
by considering separately the cases when l + j /∈ S and l + j ∈ S,
we find to be equal to ηj
◦ when ξ = ηj and zero when ξ 6= ηj. Since
ϕlhξ, ξi = T (l)∗
, it now follows from Lemma 3.3(1) that
(η) = hξ, ηji · ηj
ξ
ξ T (l)
ξ
l = vl Xη∈Al+j
θT (l)∗
ξ
l = vl Xη∈Al+j
η v∗
ηj =ξ
θηj
◦ v∗
l = Xη∈Al+j
ηj =ξ
.
θcηj
◦
Since ϕ∞hξ, ξi = diag[ϕ0hξ, ξi, ϕ1hξ, ξi, ϕ2hξ, ξi, . . .], we obtain part (4)
by summing over l ∈ N0.
(cid:3)
For the final technical lemma of this section, we decompose certain
induced representation spaces in terms of the family {Ak}∞
k=0. Suppose
that H is a Hilbert space and σ : M → B(H) is a faithful, normal,
unital ∗-homomorphism. If p is a projection in M, then pM is a Hilbert
16
JENNIFER R GOOD
W ∗-module over M. Also, σ(p)(H) is a closed subspace of H, and it
is readily shown that pM ⊗σ H and σ(p)(H) are isomorphic Hilbert
spaces under the identification of a ⊗ x with σ(a)(x) for a ∈ pM and
x ∈ H. If k ∈ N0 and ξ ∈ Ak, then hξ, ξi is a projection in M; we let Hξ
⊕Hξ.
denote the Hilbert space σhξ, ξi(H), and we define Hk :=Pξ∈Ak
Note that if k /∈ S, then Hk = {0}.
Lemma 3.6.
(1) For k ∈ N0, there is a Hilbert space isomorphism γk : E⊗k ⊗σ
H → Hk such that for η ∈ E⊗k, h ∈ H, and h = (hξ)ξ∈Ak ∈ Hk,
γk(η ⊗ h) =(cid:16) σhξ, ηi(h)(cid:17)ξ∈Ak
and γ∗
ξ ⊗ hξ.
k(cid:0)h(cid:1) = Xξ∈Ak
(2) Matricially γk(Y ⊗ IH)γ∗
k = [σhξ, Y ηi]ξ,η∈Ak for Y ∈ L(E⊗k),
k ∈ N0.
P∞
(3) There is an isomorphism of Hilbert spaces Γ : F(E) ⊗σ H →
k=0 in F(E), h ∈ H, and
k=0 ⊕Hk such that for η = (ηk)∞
k=0 ∈P∞
k=0
h = ( (hk,ξ)ξ∈Ak )∞
k=0 ⊕Hk,
Proof. When k /∈ S, we define γk to be the zero map. When k ∈ S, Ak
Γ(η ⊗ h) =(cid:16) γk(ηk ⊗ h)(cid:17)∞
is an orthonormal basis for E⊗k, so E⊗k and(cid:16)Pξ∈Ak
and Γ∗(cid:0)h(cid:1) =
∞Xk=0 Xξ∈Akbξ ⊗ hk,ξ! .
⊕ hξ, ξiM(cid:17) are
isomorphic Hilbert W ∗-modules ([12], proof of Theorem 3.12). Using
the discussion preceding the lemma, parts (1) and (3) readily follow.
Part (2) follows from routine computation.
(cid:3)
4. Weighted Commutant Lifting
Commutant lifting is the principle tool used to obtain the Nevan-
linna-Pick interpolation result, Theorem 5.3, in [8]. Towards obtaining
our weighted interpolation theorem in Section 6, the purpose of the
present section is to show that commutant lifting can be done in the
weighted case. Since the weights present obstacles for the method
producing the (unweighted) commutant lifting theorem, Theorem 4.4
in [7], our approach is instead inspired by Arias' proof of Theorem 3.1
in [2], a lifting result in complex n-space that ultimately makes use of
Parrott's lemma, Theorem 1 of [11].
Theorem 4.1 (Weighted Commutant Lifting). Let σ : M → B(H)
be a faithful, normal, unital, ∗-homomorphism for a Hilbert space H,
and let Z be a sequence of weights associated with X. Suppose that
INTERPOLATION AND COMMUTANT LIFTING WITH WEIGHTS
17
J is a closed linear subspace of F(E) ⊗σ H such that for every Y ∈
H ∞(E, Z), (Y ∗ ⊗ IH )(J) ⊆ J. With V the inclusion map of J into
F(E) ⊗σ H, suppose there exists G ∈ B(J) such that for every Y ∈
H ∞(E, Z), G (V ∗(Y ⊗ IH)V ) = (V ∗(Y ⊗ IH)V ) G. Then there exists
eG ∈ B(F(E) ⊗σ H) such that
(1) eG∗(J) ⊆ J,
(2) V ∗eGV = G,
(3) eG(Y ⊗ IH) = (Y ⊗ IH )eG for all Y ∈ H ∞(E, Z), and
(4) keGk = kGk.
j=0 ⊕E⊗j(cid:17) ⊗σ H
P . For n ∈ N0, let Kn be the isomorphic image of(cid:16)Pn
Our construction of eG ∈ B(K) will utilize an ascending sequence of
Proof. First, we consider the case when kGk = 1. Let K = F(E) ⊗σ H,
and let P ∈ B(K) be the projection map onto J. Note that V ∗ is the
range restriction of P to J, V ∗V is the identity map on J, and V V ∗ =
subspaces {Ji} of K and operators {Gi} such that each Gi ∈ B(K, Ji).
We construct these sequences inductively. To begin the process, let
let J1 = J, and define G1 := GV ∗. To organize the
n1 = −1,
numerous properties we wish to maintain at each step of the induc-
If m ∈ N, we say
tive argument, we make the following definition.
({ni}m
i=1) is an "m-triple" if for each i with 1 ≤ i ≤ m
we have that ni ∈ Z, Ji is a closed subspace of K, Gi ∈ B(K, Ji), and
the following conditions hold in all cases: with Vi the inclusion map of
Ji into K,
in K. For technical purposes, let K−1 be the zero subspace of K.
i=1, {Gi}m
i=1, {Ji}m
(Y ∗ ⊗ IH)(Ji) ⊆ Ji,
1 ≤ i ≤ m, Y ∈
i (Y ⊗ IH)Vi)Gi = Gi(Y ⊗ IH ),
1 ≤
1 ≤ i ≤ j ≤ m;
1 ≤ i ≤ m.
n1 < n2 < · · · < nm,
J1 ( J2 ( · · · ( Jm,
(if m ≥ 2);
(if m ≥ 2);
1 ≤ i ≤ m;
Condition (1, m):
Condition (2, m):
Condition (3, m): Kni ⊆ Ji,
Condition (4, m):
H ∞(E, Z);
Condition (5, m):
i ≤ m, Y ∈ H ∞(E, Z);
V ∗
Condition (6, m):
i VjGj = Gi,
kGik = 1,
Condition (7, m):
Our hypotheses guarantee that ({ni}1
(V ∗
i=1, {Ji}1
i=1) is a 1-triple.
i=1)
with m ∈ N such that Jm = K, then to satisfy the conclusion of
Let us show that if there exists an m-triple ({ni}m
i=1, {Gi}m
i=1, {Ji}m
our theorem, we may take eG to be Gm. Here, Gm ∈ B(K), Vm is
the identity map on K, and V1 = V . By Condition (6, m), V ∗Gm =
1 VmGm = G1 = GV ∗. Thus P Gm = V V ∗Gm = V GV ∗ and P GmP =
V ∗
i=1, {Gi}1
18
JENNIFER R GOOD
i=1, {Ji}m
i=1, {Gi}m
i=1, {Ji}m
i=1, {Gi}m
(V GV ∗)P = V GV ∗ = P Gm. It follows that Gm satisfies property (1) in
the conclusion of the theorem. Using again the fact that V ∗Gm = GV ∗,
we have V ∗GmV = GV ∗V = G, so property (2) holds. Condition (5, m)
gives us property (3), and property (4) follows from Condition (7, m).
We are left to consider the case when there is no m ∈ N such
i=1). Proving
if m ∈ N and
i=1) is an m-triple such that Jm ( K, then there
that Jm = K for some m-triple ({ni}m
the following fact will occupy us for quite some time:
({ni}m
exist nm+1, Jm+1, and Gm+1 such that(cid:0){ni}m+1
ofS∞
We begin by defining nm+1 and Jm+1. Since K is the norm-closure
j=0 Kj and Jm 6= K, the set {j ∈ N0 Kj * Jm} is nonempty, so
we define nm+1 to be its least element. Define Jm+1 to be the closed
subspace of K that is generated by Jm ∪Knm+1. It is readily shown that
the first three conditions for an (m + 1)-triple are satisfied. To show
Condition (4, m + 1), we first prove that for any j ∈ N and ξ ∈ E⊗j,
i=1 (cid:1) is
an (m + 1)-triple.
i=1 , {Ji}m+1
i=1 , {Gi}m+1
(4.1)
(4.2)
Pm+1(Wξ ⊗ IH)Pm = Pm+1(Wξ ⊗ IH)
= Pm+1(Wξ ⊗ IH)Pm+1,
ξ ⊗IH )(Knm+1) ⊆ K(nm+1)−1 ⊆ Jm, so (W ∗
where Pm and Pm+1 are the projection maps in B(K) onto Jm and Jm+1,
respectively. First, let j = 1 and ξ ∈ E. By Condition (4, m), (W ∗
ξ ⊗
IH)(Jm) ⊆ Jm. At the same time, from the definition of Wξ it follows
that for any n ∈ N0, (W ∗
ξ ⊗IH )(Kn) ⊆ Kn−1. Therefore, by definition of
nm+1, (W ∗
ξ ⊗IH )(Jm+1) ⊆ Jm.
Thus, equation (4.1) holds when j = 1. Since PmPm+1 = Pm and
E⊗k+1 is generated by elements of the form ξ = η ⊗ ζ where η ∈ E and
ζ ∈ E⊗k, an inductive argument gives equation (4.1) in the general
case. Equation (4.2) follows since PmPm+1 = Pm. For a ∈ M, it is
readily seen that ϕ∞(a∗) ⊗ IH leaves Kn invariant for any n ∈ N and
leaves Jm invariant by Condition (4, m). It follows that for Y = ϕ∞(a),
(4.3)
(Y ∗ ⊗ IH)(Jm+1) ⊆ Jm+1.
By equation (4.2), containment (4.3) is also satisfied for Y = Wξ when
ξ ∈ E⊗j for some j ∈ N. It follows that containment (4.3) holds for
every Y ∈ H ∞(E, Z). This, together with Condition (4, m), implies
Condition (4, m + 1).
Before considering the remaining conditions we must define Gm+1,
which will require some preliminary work. Let us begin by recalling
and establishing notation. Let {Ak}∞
k=0 be the family defined in the
paragraph before Lemma 3.3. For k ∈ N0 and ξ ∈ E⊗k, bξ denotes
i=0 ∈ F(E). For ξ ∈ F(E), Lξ is the insertion operator
vk(ξ) = (δi=k ξ)∞
INTERPOLATION AND COMMUTANT LIFTING WITH WEIGHTS
19
m is the range restriction of Pm, V ∗
∈ B(H, Jm). Note
mapping h ∈ H to ξ ⊗ h. Define g := GmLd1M
that V ∗
mVm is the identity on Jm,
and VmV ∗
m = Pm, with analogous properties for Vm+1. For notational
convenience, define α : H ∞(E, Z) → B(Jm) and β : H ∞(E, Z) →
B(Jm, Jm+1) at Y ∈ H ∞(E, Z) by,
α(Y ) = V ∗
m(Y ⊗ IH)Vm and β(Y ) = V ∗
m+1(Y ⊗ IH)Vm.
The following technical lemma collects facts about α and β that will
be useful in future computations.
Lemma 4.2.
(1) α is contractive, linear, multiplicative, unital, ultraweakly con-
tinuous, and for any Y ∈ H ∞(E, Z),
α(Y )Gm = Gm(Y ⊗ IH).
(2) β is contractive, linear, continuous with respect to the ultra-
weak topology on H ∞(E, Z) and the weak operator topology on
B(Jm, Jm+1), and for any k ∈ N, ξ ∈ E⊗k, and Y ∈ H ∞(E, Z),
β(Wξ)α(Y ) = β(WξY ).
(3) For any a ∈ M, k ∈ N0, and ξ ∈ E⊗k, β(Wξ)∗V ∗
m+1(ϕ∞(a) ⊗
IH)Vm+1 = β(Wa∗·ξ)∗.
(4) For all k ∈ N0 and ξ ∈ E⊗k,
(5) For all a ∈ M, k ∈ N0, and ξ ∈ E⊗k, α(Wξ·a)g = α(Wξ)gσ(a).
(6) For all a ∈ M, k ∈ N0, and ξ ∈ E⊗k,
g∗β(Wξ·a)∗ =
α(W(Z (k))−1ξ)g = GmLbξ,
σ(a∗)g∗β(Wξ)∗.
(7) For any j ∈ N0, η ∈ E⊗j, and T ∈ L(E⊗j),
α(WT η)g = Xξ∈Aj
α(WT ξhξ,ηi)g,
with convergence in the weak operator topology on B(H, Jm).
(8) For any j ∈ N0, η ∈ E⊗j, and T ∈ L(E⊗j),
g∗β(WT η)∗ = Xξ∈Aj
g∗β(WT ξhξ,ηi)∗,
with convergence in the weak operator topology on B(Jm+1, H).
Proof of Lemma 4.2.
(1): The induced representation σF(E) is a normal, unital ∗-homomor-
phism. We form α by composing the restriction of σF(E) to H ∞(E, Z)
with Ad(V ∗
mT Vm. It follows
that α is linear, unital, contractive, and ultraweakly continuous. By
Condition (4, m), for Y ∈ H ∞(E, Z), V ∗
m(Y ⊗ IH)Pm.
m) : B(K) → B(Jm), which sends T to V ∗
m(Y ⊗ IH) = V ∗
20
JENNIFER R GOOD
It follows that α is multiplicative. By Condition (5, m), for any Y ∈
H ∞(E, Z), α(Y )Gm = Gm(Y ⊗ IH).
(2): The linearity, contractivity, and continuity of β follow from
arguments similar to those used above for α. By equation (4.1), if
k ∈ N, ξ ∈ E⊗k, and Y ∈ H ∞(E, Z), then β(Wξ)α(Y ) = V ∗
m+1(Wξ ⊗
IH)Pm(Y ⊗ IH)Vm = V ∗
m+1(Wξ ⊗ IH )(Y ⊗ IH )Vm = β(WξY ).
(3): Since Condition (4, m + 1) holds and ϕ∞(a∗)Wξ = Wa∗·ξ,
m+1(ϕ∞(a∗)⊗IH )Vm+1β(Wξ) = V ∗
V ∗
m+1(ϕ∞(a∗)⊗IH )Pm+1(Wξ⊗IH )Vm
= V ∗
m+1(ϕ∞(a∗) ⊗ IH )(Wξ ⊗ IH)Vm = β(Wa∗·ξ).
The result follows from taking the adjoint.
(4): By part (1), α(W(Z (k))−1ξ)g = Gm(W(Z (k))−1ξ ⊗ IH)Ld1M
(5): Since ϕ(a) = Wa and Z (0) = IM , by part (4)
α(ϕ∞(a))g = α(Wa)g = GmLba = GmLd1M
(4.4)
σ(a) = gσ(a).
= GmLbξ.
Since α is multiplicative, α(Wξ·a)g = α(Wξ)α(ϕ∞(a))g = α(Wξ)gσ(a).
(6): By Condition (4, m) and equation (4.4),
β(Wξ·a)g = V ∗
m+1(Wξ ⊗ IH)(ϕ∞(a) ⊗ IH )Vmg
= V ∗
m+1(Wξ ⊗ IH)Pm(ϕ∞(a) ⊗ IH)Vmg
= β(Wξ)α(ϕ∞(a))g = β(Wξ)gσ(a).
The result follows by taking the adjoint.
(7): We have η = Pξ∈Aj
ξhξ, ηi since Aj is either an orthonormal
basis or the zero set. As an element of L(E⊗j), T is an ultraweakly
continuous, right M-module homomorphism; also the map that sends
ζ ∈ E⊗j to Wζ ∈ L(F(E)) is linear and ultraweakly continuous, as is
α. The result follows.
(8): The result follows along similar lines as part (7).
(cid:4)
One fact we will require in defining Gm+1 is that
(4.5)
G∗
m(x) =
x ∈ Jm.
x! ,
∞Xk=0 Xξ∈Akbξ ⊗ g∗α(cid:0)W(Z (k))−1ξ(cid:1)∗
(cid:16)(cid:0) g∗α(cid:0)W(Z (k))−1ξ(cid:1)∗ x(cid:1)ξ∈Ak (cid:17)∞
k=0
To prove this fact, we will show that
(4.6)
k=0 ⊕Hk, using the notation of Lemma 3.6. Let
k ∈ N0 and ξ ∈ Ak. By Lemma 3.5(1), ξ · hξ, ξi = ξ, so by Lemma
is an element in P∞
INTERPOLATION AND COMMUTANT LIFTING WITH WEIGHTS
21
By Lemma 4.2(4) and Lemma 3.3(2),
4.2(5), we have g∗α(cid:0)W(Z (k))−1ξ(cid:1)∗ x = σhξ, ξig∗α(cid:0)W(Z (k))−1ξ(cid:1)∗ x ∈ Hξ.
(4.7) Xξ∈Ak(cid:13)(cid:13)g∗α(cid:0)W(Z (k))−1ξ(cid:1)∗ x(cid:13)(cid:13)2
= Xξ∈AkDLbξL∗
bξG∗
= Xξ∈Ak(cid:13)(cid:13)(cid:13)L∗
mx(cid:13)(cid:13)(cid:13)
bξG∗
mxE
mx, G∗
2
= h(Qk ⊗ IH)G∗
mx, G∗
mxi ≤ kG∗
mxk2.
2
=
mx, G∗
mxi
= kG∗
mxk2.
h(Qk ⊗ IH)G∗
is otherwise zero. By Lemma 4.2 parts (5), (7), and (4),
operator convergence in B(K), so by computation (4.7) we find,
is a well-
k=0(Qk ⊗ IH) = IK with strong
k=0 ⊕Hk. By Lemma
3.6(3), its image under Γ∗ in K is the element in the right-hand side of
equation (4.5). To show equality with G∗
m(x), let h ∈ H, j ∈ N0, and
We conclude that for every k ∈ N0,(cid:0) g∗α(cid:0)W(Z (k))−1ξ(cid:1)∗ x(cid:1)ξ∈Ak
defined element of Hk. Moreover, P∞
x(cid:1)ξ∈Ak(cid:13)(cid:13)(cid:13)
∞Xk=0(cid:13)(cid:13)(cid:13)(cid:0) g∗α(cid:0)W(Z (k))−1ξ(cid:1)∗
∞Xk=0
Thus, the element in expression (4.6) belongs toP∞
η ∈ E⊗j. For k ∈ N0 and ξ ∈ E⊗k,Dbξ,bηE equals hξ, ηi when k = j and
* ∞Xk=0 Xξ∈Akbξ ⊗ g∗α(cid:0)W(Z (k))−1ξ(cid:1)∗ x! ,bη ⊗ h+
= Xξ∈Aj(cid:10)x, α(cid:0)W(Z (j))−1ξ(cid:1) gσhξ, ηih(cid:11) = Xξ∈AjDx, α(cid:0)W(Z (j))−1ξhξ,ηi(cid:1) g(h)E
mx,bη ⊗ hi.
β(cid:0)W(Z (k))−1η(cid:1) gg∗β(cid:0)W(Z (k))−1η(cid:1)∗! ≤ IJm+1.
=Dx, α(cid:0)W(Z (j))−1η(cid:1) g(h)E = hx, GmLbη(h)i = hG∗
Equation (4.5) follows from routine approximation arguments.
Another fact we need before constructing Gm+1 is that
∞Xk=1 Xη∈Ak
(4.8)
22
JENNIFER R GOOD
For convenience, we break the proof into three main steps, as follows:
(4.9)
IJm+1 ≥
=
(4.10)
(4.11)
=
j
j
Xξ∈Aj
β(cid:16)WX 1/2
∞Xj=1
∞Xk=1 kXj=1 Xη∈Ak
∞Xk=1 Xη∈Ak
ξ(cid:17) GmG∗
β(cid:18)W(cid:16)X 1/2
·g∗β(cid:18)W(cid:16)X 1/2
ξ(cid:17)∗
mβ(cid:16)WX 1/2
j ⊗(Z (k−j))−1(cid:17)η(cid:19) g
j ⊗(Z (k−j))−1(cid:17)η(cid:19)∗(cid:19)(cid:19)
β(cid:0)W(Z (k))−1η(cid:1) gg∗β(cid:0)W(Z (k))−1η(cid:1)∗! .
We begin with inequality (4.9).
ξW ∗
for any j ∈ N, Pξ∈Aj
WX 1/2
j
j
X 1/2
a finite subset of Aj, then because σF(E) is a ∗-homomorphism and
kVmGmk ≤ 1,
It follows from Lemma 3.4 that
If F is
converges in L(F(E)).
ξ
j
ξ(cid:17)∗
mβ(cid:16)WX 1/2
ξ ⊗ IH(cid:17) VmGmG∗
j
j
j
V ∗
≤ V ∗
Xξ∈F
ξ(cid:17) GmG∗
β(cid:16)WX 1/2
m+1(cid:16)WX 1/2
=Xξ∈F
m+1 Xξ∈F
m+1Xξ∈Aj
It follows that Pξ∈Aj
≤ V ∗
weakly in B(Jm+1) and V ∗
j
j
ξ
X 1/2
X 1/2
mV ∗
ξW ∗
WX 1/2
WX 1/2
⊗ IH(cid:19) Vm+1
m(cid:18)W ∗
ξ! ⊗ IH! Vm+1
ξ ⊗ IH Vm+1.
ξ(cid:17)∗
mβ(cid:16)WX 1/2
ξ(cid:17) GmG∗
β(cid:16)WX 1/2
m+1(cid:18)(cid:18)Pξ∈Aj
ξ(cid:19) ⊗ IH(cid:19) Vm+1 is
converges ultra-
WX 1/2
ξW ∗
ξW ∗
X 1/2
X 1/2
j
j
j
j
j
j
an upper bound. This, together with Lemma 3.4, implies that for
INTERPOLATION AND COMMUTANT LIFTING WITH WEIGHTS
23
N ∈ N,
NXj=1
Xξ∈Aj
j
j
ξ(cid:17) GmG∗
mβ(cid:16)WX 1/2
β(cid:16)WX 1/2
m+1
Xξ∈Aj
NXj=1
WX 1/2
j
ξW ∗
≤ V ∗
ξ(cid:17)∗
ξ ⊗ IH Vm+1 ≤ IJm+1.
j
X 1/2
Inequality (4.9) follows. Towards proving equation (4.10), fix j ∈ N
and ξ ∈ Aj. Since ξ · hξ, ξi = ξ, it follows from Lemma 4.2 parts (1)
and (2) that
(4.12)
β(cid:16)WX 1/2
j
ξ(cid:17) Gm = β(cid:16)WX 1/2
j
ξ(cid:17) Gm (ϕ∞hξ, ξi ⊗ IH) .
Also, by Lemma 4.2(4), whenever l ∈ N0 and ζ ∈ E⊗l,
Gm(cid:16)θbζ ⊗ IH(cid:17) G∗
m = GmLbζL∗
Therefore, by Lemma 3.5(4),
(4.13) Gm (ϕ∞hξ, ξi ⊗ IH ) G∗
m = α(cid:0)W(Z (l))−1ζ(cid:1) gg∗α(cid:0)W(Z (l))−1ζ(cid:1)∗ .
bζG∗
m
Xη∈Al+j
Gm(cid:16)θbηj
⊗ IH(cid:17) G∗
∞Xl=0
0(cid:17)∗ .
0(cid:17) gg∗α(cid:16)W(Z (l))−1ηj
α(cid:16)W(Z (l))−1ηj
ηj =ξ
m =
0
=
∞Xl=0
Xη∈Al+j
ηj =ξ
If l ∈ N0 and η ∈ Al+j with ηj = ξ, by Lemma 4.2(2) and Lemma
3.5(2),
(4.14)
β(cid:16)WX 1/2
j
ξ(cid:17) α(cid:16)W(Z (l))−1ηj
0(cid:17) g = β(cid:18)W(cid:16)X 1/2
j ⊗(Z (l))−1(cid:17)η(cid:19) g.
24
JENNIFER R GOOD
Putting together equations (4.12), (4.13), and (4.14),
j
j
j
X 1/2
X 1/2
ξ(cid:19)∗
mβ(cid:18)W
mβ(cid:18)W
ξ(cid:19) Gm (ϕ∞hξ, ξi ⊗ IH) G∗
ξ(cid:19)∗
ξ(cid:19) α(cid:16)W(Z (l))−1ηj
β(cid:18)W
0(cid:17) gg∗α(cid:16)W(Z (l))−1ηj
0(cid:17)∗
j ⊗(Z (l))−1(cid:17)η(cid:19)∗
j ⊗(Z (l))−1(cid:17)η(cid:19) gg∗β(cid:18)W(cid:16)X 1/2
β(cid:18)W(cid:16)X 1/2
j
j
X 1/2
.
ξ(cid:19)∗
j
X 1/2
ξ(cid:19)∗
mβ(cid:18)W
β(cid:18)W(cid:16)X 1/2
j ⊗(Z (l))−1(cid:17)η(cid:19) gg∗β(cid:18)W(cid:16)X 1/2
j ⊗(Z (l))−1(cid:17)η(cid:19) gg∗β(cid:18)W(cid:16)X 1/2
j ⊗(Z (l))−1(cid:17)η(cid:19)∗
j ⊗(Z (l))−1(cid:17)η(cid:19)∗ .
Summing over ξ ∈ Aj, interchanging sums, and using Lemma 3.5(3),
j
j
=
=
ηj =ξ
ηj =ξ
X 1/2
X 1/2
X 1/2
X 1/2
∞Xl=0
β(cid:18)W
ξ(cid:19) GmG∗
= β(cid:18)W
Xη∈Al+j
β(cid:18)W
Xη∈Al+j
∞Xl=0
β(cid:18)W
ξ(cid:19) GmG∗
Xξ∈Aj
Xη∈Al+j
∞Xl=0
= Xξ∈Aj
Xη∈Al+j
β(cid:18)W(cid:16)X 1/2
∞Xl=0
Xξ∈Aj
β(cid:18)W
∞Xj=1
Xη∈Al+j
∞Xj=1
∞Xl=0
Xη∈Ak
kXj=1
∞Xk=1
X 1/2
ηj =ξ
=
=
=
j
Finally, summing over j ∈ N, re-indexing, and interchanging sums,
j
X 1/2
ξ(cid:19)∗
ξ(cid:19) GmG∗
β(cid:18)W(cid:16)X 1/2
β(cid:18)W(cid:16)X 1/2
mβ(cid:18)W
j ⊗(Z (l))−1(cid:17)η(cid:19) gg∗β(cid:18)W(cid:16)X 1/2
j ⊗(Z (k−j))−1(cid:17)η(cid:19) gg∗β(cid:18)W(cid:16)X 1/2
j ⊗(Z (l))−1(cid:17)η(cid:19)∗
j ⊗(Z (k−j))−1(cid:17)η(cid:19)∗
INTERPOLATION AND COMMUTANT LIFTING WITH WEIGHTS
25
as desired. Finally, to prove equation (4.11) it suffices to show that
j ⊗(Z (k−j))−1(cid:17)η(cid:19) gg∗β(cid:18)W(cid:16)X 1/2
β(cid:18)W(cid:16)X 1/2
β(cid:0)W(Z (k))−1η(cid:1) gg∗β(cid:0)W(Z (k))−1η(cid:1)∗
j ⊗(Z (k−j))−1(cid:17)η(cid:19)∗!
(4.15)
kXj=1 Xη∈Ak
= Xη∈Ak
k = Dk(cid:16)Pξ∈Ak
ξ(cid:17) D∗
kg∗β(Wξ)∗xk2 = Xξ∈Ak
Xξ∈Ak
≤ Xξ∈Ak
k(W ∗
for any k ∈ N. Let x ∈ Jm+1. By definition of Dk and Lemma 3.3(3),
DkD∗
ξ . Since kgk ≤ 1,
WξW ∗
TξT ∗
k =Pξ∈Ak
kg∗V ∗
m(W ∗
ξ ⊗ IH)Vm+1xk2
ξ ⊗ IH)xk2 = Xξ∈Ak(cid:10)(WξW ∗
ξ ⊗ IH)x, x(cid:11)
= h(DkD∗
k ⊗ IH)x, xi = k(D∗
k ⊗ IH)xk2 .
Using the notation and results from Lemma 3.6, it follows from Lemma
4.2(6) that for every ξ ∈ Ak, g∗β(Wξ)∗x belongs to Hξ. The preceding
computation shows that (g∗β(Wξ)∗x)ξ∈Ak is an element in Hk; let ex
SS∗⊗IH =Pξ∈Ak
be its image in E⊗k ⊗σ H under the isomorphism γ∗
k. If S is any ele-
ment in L(E⊗k), then using Lemma 3.3(1) and applying σE⊗k, we have
(θSξ ⊗ IH). For each ξ ∈ Ak, the matrix for θSξ ⊗IH,
when viewed as an operator in B(Hk), is [σhµ, SξiσhSξ, νi]µ,ν∈Ak
by
Lemma 3.6(2), so by Lemma 4.2 parts (6) and (8), a matrix computa-
tion reveals,
(4.16)
hex, (SS∗ ⊗ IH)exi
hσhSξ, µig∗β(Wµ)∗x, σhSξ, νi(g∗β(Wν)∗x)i!!
= Xξ∈Ak Xµ∈Ak Xν∈Ak
= Xξ∈Ak Xµ∈Ak Xν∈Ak(cid:10)g∗β(Wµhµ,Sξi)∗x, g∗β(Wνhν,Sξi)∗x(cid:11)!!
hg∗β(WSξ)∗x, g∗β(WSξ)∗xi = Xξ∈Ak
= Xξ∈Ak
k andPk
Using the identities, Z (k)∗Z (k) = R−2
as two applications of equation (4.16), first when S = X 1/2
hβ(WSξ)gg∗β(WSξ)∗x, xi .
j=1 Xj ⊗R2
k, as well
j ⊗(Z (k−j))−1
k−j = R2
26
JENNIFER R GOOD
and second when S = (Z (k))−1, we obtain
* kXj=1
Xη∈Ak
=
j ⊗(Z (k−j))−1(cid:17)η(cid:19)∗ x, x+
β(cid:18)W(cid:16)X 1/2
j ⊗(Z (k−j))−1(cid:17)η(cid:19) gg∗β(cid:18)W(cid:16)X 1/2
kXj=1
k−j) ⊗ IH(cid:1)exi = hex, (R2
hex,(cid:0)(Xj ⊗ R2
k ⊗ IH)exi
= Xη∈AkD(cid:16)β(cid:16)W(Z (k))−1η(cid:17) gg∗β(cid:16)W(Z (k))−1η(cid:17)∗(cid:17) x, xE .
β(cid:0)W(Z (k))−1η(cid:1) gg∗β(cid:0)W(Z (k))−1η(cid:1)∗ converges in B(Jm+1) and
equation (4.15) holds, which completes the proof of inequality (4.8).
Momentarily, we define Gm+1 ∈ B(K, Jm+1) as a 2 × 2 operator-
valued matrix with respect to the decompositions K = (K ⊖ K0) ⊕
K0 and Jm+1 = Jm ⊕ (Jm+1 ⊖ Jm). One of the columns of Gm+1 is
determined by the adjoint of a contraction F : Jm+1 → K ⊖ K0 that
we define as follows at x ∈ Jm+1. By inequality (4.8),
ThusPη∈Ak
(4.17)
∞Xk=1 Xξ∈Ak
=* ∞Xk=1 Xξ∈Ak
xk2!
kg∗β(cid:0)W(Z (k))−1ξ(cid:1)∗
β(cid:0)W(Z (k))−1ξ(cid:1) gg∗β(cid:0)W(Z (k))−1ξ(cid:1)∗! x, x+ ≤ kxk2.
It follows that(cid:16)δk≥1 (cid:0)g∗β(cid:0)W(Z (k))−1ξ(cid:1)∗
ement of P∞
is a well-defined el-
k=0 ⊕Hk where δk≥1 is 1 when k ≥ 1 and 0 when k = 0.
k=0
We define F (x) to be its image under Γ∗ in K, namely
x(cid:1)ξ∈Ak(cid:17)∞
(4.18)
F (x) =
∞Xk=1 Xξ∈Akbξ ⊗ g∗β(cid:0)W(Z (k))−1ξ(cid:1)∗ x! .
M, and h ∈ H,
The operator F is linear, and by inequality (4.17), F is contractive. If
a ∈ M, k ∈ N, and ξ ∈ Ak, thenDbξ,baE = 0. Thus, for x ∈ Jm+1, a ∈
x, σDbξ,baE hE! = 0.
hF (x),ba ⊗ hi =
∞Xk=1 Xξ∈AkDg∗β(cid:0)W(Z (k))−1ξ(cid:1)∗
It follows that F (x) ∈ K ⊖ K0, as desired.
For organizational purposes, we temporarily adopt the following no-
) denote
tation: if L is a Hilbert space with closed subspace L0, let (V L
L0
INTERPOLATION AND COMMUTANT LIFTING WITH WEIGHTS
27
the inclusion isometry in B(L0, L), and let (P L
) denote the projection
L0
)∗ is the range restriction of
in B(L) onto L0.
(P L
) to L0, (V L
) is the identity on
L0
L0
L0. In this notation, let us show that F and Gm satisfy the following
relationship:
It follows that (V L
L0
), and (V L
L0
)∗ = (P L
L0
)∗(V L
L0
)(V L
L0
(4.19)
)∗F ∗ = Gm(V K
If x ∈ Jm then by equations (4.18) and (4.5),
(V Jm+1
Jm
K⊖K0
).
F (V Jm+1
Jm
)x =
=
∞Xk=1 Xξ∈Akbξ ⊗ g∗β(cid:0)W(Z (k))−1ξ(cid:1)∗ (V Jm+1
)x!
∞Xk=1 Xξ∈Akbξ ⊗ g∗α(cid:0)W(Z (k))−1ξ(cid:1)∗ x! = (V K
Jm
)∗G∗
mx.
K⊖K0
Therefore, F (V Jm+1
Jm
) = (V K
K⊖K0
)∗G∗
m, and equation (4.19) follows.
In our 2 × 2 matricial definition of Gm+1, the following maps, R, S,
and T , will comprise three of the four corners. Define
R = (V Jm+1
Jm
S = (V Jm+1
T = Gm(V K
K0
Jm+1⊖Jm
)
)∗F ∗ = Gm(V K
K⊖K0
)
in B(K ⊖ K0, Jm);
)∗F ∗
in B(K ⊖ K0, Jm+1 ⊖ Jm); and
in B(K0, Jm);
noting that R is well-defined by equation (4.19). Matricially,
By Parrott's lemma [11], there exists U ∈ B(K0, Jm+1 ⊖ Jm) such
F ∗ =(cid:20)R
S(cid:21)
and Gm =(cid:2)R T(cid:3) .
(cid:13)(cid:13)(cid:13)(cid:13)(cid:20)R T
S U(cid:21)(cid:13)(cid:13)(cid:13)(cid:13) = max {kF k , kGmk} ,
Gm+1 =(cid:20)R T
S U(cid:21)
(4.20)
that
(4.21)
(4.22)
(4.23)
so at last we define
in B(K, Jm+1). By equation (4.20),
Gm+1(V K
(V Jm+1
Jm
K⊖K0
) = F ∗ and
)∗Gm+1 = Gm.
i=1 , {Ji}m+1
i=1 , {Gi}m+1
Our goal is to show that(cid:0){ni}m+1
triple. We have already shown that the first four (m+1)-level conditions
are satisfied. Condition (7, m + 1) readily follows from equation (4.21).
i=1 (cid:1) is an (m + 1)-
28
JENNIFER R GOOD
Towards proving Condition (5, m + 1), recall that σF(E) ◦ ϕ∞ : M →
B(K) is the normal, unital ∗-homomorphism that maps a ∈ M to
ϕ∞(a) ⊗ IH . Let each of ψ1 : M → B(K ⊖ K0), ψ2 : M → B(K0),
π1 : M → B(Jm), and π2 : M → B(Jm+1 ⊖ Jm) be the compression
of σF(E) ◦ ϕ∞ to the indicated space; for instance, for every a ∈ M,
ψ1(a) = (V K
). The spaces K0, Jm, and Jm+1
reduce σF(E) ◦ ϕ∞, from which it follows that K ⊖ K0 and Jm+1 ⊖ Jm
also reduce σF(E) ◦ ϕ∞. Therefore, ψ1, ψ2, π1, and π2 are normal, unital
∗-homomorphisms. We want to show that
(4.24)
)∗(ϕ∞(a)⊗IH)(V K
K⊖K0
K⊖K0
R ∈ I(ψ1, π1); S ∈ I(ψ1, π2); T ∈ I(ψ2, π1); and U ∈ I(ψ2, π2).
Note that Vm = (V K
Jm
σF(E) ◦ ϕ∞, if a ∈ M, then by Condition (5, m),
) and Vm+1 = (V K
Jm+1
). Since K ⊖ K0 reduces
Rψ1(a) = Gm(P K
K⊖K0
)(ϕ∞(a) ⊗ IH)(V K
)
K⊖K0
= Gm(ϕ∞(a) ⊗ IH)(V K
)
K⊖K0
= V ∗
m(ϕ∞(a) ⊗ IH)VmGm(V K
K⊖K0
) = π1(a)R,
Thus R ∈ I(ψ1, π1); analogously T ∈ I(ψ2, π1). To show that S ∈
I(ψ1, π2), it suffices to show that for every a ∈ M, x ∈ Jm+1 ⊖ Jm,
η ∈ E⊗j with j ∈ N, and h ∈ H,
(4.25)
Thus by Lemma 4.2, parts (6) and (8),
(4.26)
hbη ⊗ h, ψ1(a)S∗xi = Xξ∈Aj(cid:10)h, σhη, a · ξi(cid:0)g∗β(cid:0)W(Z (j))−1ξ(cid:1)∗ x(cid:1)(cid:11)
= Xξ∈Aj(cid:10)h, g∗β(cid:0)W(Z (j))−1ξ·hξ,a∗ηi(cid:1)∗ x(cid:11) =(cid:10)h, g∗β(cid:0)W(Z (j))−1(a∗η)(cid:1)∗ x(cid:11) .
Since K ⊖ K0 reduces σF(E) ◦ ϕ∞, by equation (4.18),
ψ1(a)S∗x = (ϕ∞(a) ⊗ IH)F x
hbη ⊗ h, ψ1(a)S∗xi = hbη ⊗ h, S∗π2(a)xi.
= (ϕ∞(a) ⊗ IH) ∞Xk=1 Xξ∈Akbξ ⊗ g∗β(cid:0)W(Z (k))−1ξ(cid:1)∗ x!!
∞Xk=1 Xξ∈Akda · ξ ⊗ g∗β(cid:0)W(Z (k))−1ξ(cid:1)∗ x! .
=
INTERPOLATION AND COMMUTANT LIFTING WITH WEIGHTS
29
On the other hand, since Jm+1 ⊖ Jm reduces σF(E) ◦ ϕ∞, by Lemma
4.2(3),
S∗π2(a)x = F (V K
Jm+1
)∗(ϕ∞(a) ⊗ IH)(V K
Jm+1
)x
=
=
Jm+1
)∗(ϕ∞(a) ⊗ IH )(V K
Jm+1
∞Xk=1 Xξ∈Akbξ ⊗(cid:16)g∗β(cid:0)W(Z (k))−1ξ(cid:1)∗ (V K
)x(cid:17)!
∞Xk=1 Xξ∈Akbξ ⊗ g∗β(cid:0)Wa∗·(Z (k))−1ξ(cid:1)∗ x! .
hbη ⊗ h, S∗π2(a)xi = Xξ∈Aj(cid:10)h, σhη, ξi(cid:0)g∗β(cid:0)Wa∗·(Z (j))−1ξ(cid:1)∗ x(cid:1)(cid:11)
= Xξ∈Aj(cid:10)h, g∗β(cid:0)Wa∗(Z (j))−1ξhξ,ηi(cid:1)∗
x(cid:11) .
x(cid:11) =(cid:10)h, g∗β(cid:0)W(Z (j))−1(a∗η)(cid:1)∗
Therefore by Lemma 4.2, parts (6) and (8), since (Z (j))−1 ∈ ϕj(M)c,
(4.27)
Combining equations (4.26) and (4.27), we obtain equation (4.25), so
S ∈ I(ψ1, π2). Examining the proof of Parrott's Lemma in [11], we see
that U is given as the weak operator limit of the sequence
(cid:8)−cnS(I − c2
nR∗R)−1R∗T(cid:9)∞
n=0
for some sequence of numbers {cn}∞
n=0. Since R ∈ I(ψ1, π1), S ∈
I(ψ1, π2), and T ∈ I(ψ2, π1), it follows that U ∈ I(ψ2, π2), which gives
(4.24).
To establish Condition (5, m+1), by Conditions (4, m+1) and (5, m),
it suffices to show that
(4.28)
(V ∗
m+1(Y ⊗ IH)Vm+1)Gm+1 = Gm+1(Y ⊗ IH)
in only two cases: when Y = ϕ∞(a) for some a ∈ M and when Y = Wξ
for some ξ ∈ E. Using properties (4.24), we compute
V ∗
0
m+1(ϕ∞(a) ⊗ IH)Vm+1Gm+1 =(cid:20)π1(a)
π2(a)(cid:21)(cid:20)R T
S U(cid:21)
=(cid:20)π1(a)R π1(a)T
π2(a)S π2(a)U(cid:21) =(cid:20)Rψ1(a) T ψ2(a)
Sψ1(a) Uψ2(a)(cid:21)
=(cid:20)R T
S U(cid:21)(cid:20)ψ1(a)
ψ2(a)(cid:21) = Gm+1(ϕ∞(a) ⊗ IH).
0
0
0
This establishes equation (4.28) when Y = ϕ∞(a). For the case when
Y = Wξ, it suffices to show that for every j ∈ N0, η ∈ E⊗j, h ∈ H, and
30
JENNIFER R GOOD
x ∈ Jm+1,
(4.29)
(4.18),
(cid:10)V ∗
m+1(Wξ ⊗ IH )Vm+1Gm+1 (bη ⊗ h) , x(cid:11) = hGm+1(Wξ ⊗ IH) (bη ⊗ h) , xi .
With Wξ(bη) ⊗ h = \Zj+1(ξ ⊗ η) ⊗ h ∈ K ⊖ K0, by equations (4.22) and
hGm+1(Wξ ⊗ IH) (bη ⊗ h) , xi
m+1xE =D \Zj+1(ξ ⊗ η) ⊗ h, F xE
=D \Zj+1(ξ ⊗ η) ⊗ h, (V K
xE! .
∞Xk=1 Xµ∈AkD \Zj+1(ξ ⊗ η) ⊗ h,bµ ⊗ g∗β(cid:0)W(Z (k))−1µ(cid:1)∗
For k ∈ N and µ ∈ Ak, D \Zj+1(ξ ⊗ η),bµE is hZj+1(ξ ⊗ η), µi when
k = j + 1 and is otherwise zero. Recalling that (Z (j+1))−1Zj+1 =
I1 ⊗ (Z (j))−1, we continue the preceding computation using Lemma
4.2, parts (6) and (8),
)∗G∗
K⊖K0
=
(4.30)
hGm+1(Wξ ⊗ IH) (bη ⊗ h) , xi
= Xµ∈Aj+1(cid:10)h, σhZj+1(ξ ⊗ η), µi(cid:0)g∗β(cid:0)W(Z (j+1))−1µ(cid:1)∗ x(cid:1)(cid:11)
= Xµ∈Aj+1Dh, g∗β(cid:16)W(Z (j+1))−1µhµ,Zj+1(ξ⊗η)i(cid:17)∗
xE
=Dh, g∗β(cid:16)W(Z (j+1))−1Zj+1(ξ⊗η)(cid:17)∗
xE =(cid:10)h, g∗β(cid:0)Wξ⊗(Z (j))−1η(cid:1)∗
x(cid:11)
=(cid:10)β(cid:0)Wξ⊗(Z (j))−1η(cid:1) g(h), x(cid:11) .
)∗ and V ∗
On the other hand, we have that PmVm+1 = Vm(V Jm+1
Jm
IH) = V ∗
and Lemma 4.2, parts (4) and (2),
m+1(Wξ⊗
m+1(Wξ ⊗ IH)Pm by equation (4.1). Thus, by equation (4.23)
V ∗
= V ∗
m+1(Wξ ⊗ IH )Vm+1Gm+1 (bη ⊗ h) = V ∗
m+1(Wξ ⊗ IH)Vm(V Jm+1
Jm
m+1(Wξ ⊗ IH )PmVm+1Gm+1Lbηh
= β (Wξ) α(cid:0)W(Z (j))−1η(cid:1) g(h) = β(cid:0)Wξ⊗(Z (j))−1η(cid:1) g(h).
)∗Gm+1Lbηh = β(Wξ)GmLbηh
This fact together with equation (4.30) yields equation (4.29), which
completes the proof of Condition (5, m + 1).
All that remains is Condition (6, m + 1), and we need only consider
the case when j = m + 1.
i VjGj = Gm+1 =
Gi, as desired. Suppose 1 ≤ i ≤ m. Since Ji ⊆ Jm ⊆ Jm+1 ⊆
K, it follows that V ∗
). Taking
If i = m + 1, then V ∗
) and V ∗
j Vi = (V Jm+1
Jm
)(V Jm
Ji
mVi = (V Jm
Ji
INTERPOLATION AND COMMUTANT LIFTING WITH WEIGHTS
31
i=1 , {Ji}m+1
i=1 , {Gi}m+1
i=1 (cid:1) is an (m + 1)-triple.
adjoints, applying equation (4.23), and using Condition (6, m), we have
i VjGj = (V Jm
V ∗
i VmGm = Gi. Thus,
Ji
Condition (6, m + 1) is satisfied. Having shown all seven conditions,
)∗Gm+1 = (V Jm
Ji
)∗Gm = V ∗
)∗(V Jm+1
Jm
we conclude that(cid:0){ni}m+1
i=1, {Gi}m
i=1, {Ji}m
i=1, {Gi}1
Recall that we are considering the case when there is no m ∈ N
such that an m-triple ({ni}m
i=1) exists with Jm = K.
Since we know ({ni}1
i=1, {Ji}1
i=1) is a 1-triple, the result we
have just shown, together with an inductive argument, guarantees the
existence of three sequences {ni}∞
i=1 such that
({ni}m
i=1) is an m-triple for every m ∈ N. As before, for
every i ∈ N, let Vi ∈ B(Ji, K) be the inclusion map, and let Pi ∈ B(K)
be the projection map onto Ji.
i=1, and {Gi}∞
i=1, {Ji}∞
i=1, {Gi}m
i=1, {Ji}m
Towards defining eG, note that {Ji}∞
S∞
Therefore S∞
i=1 is an increasing family of sub-
spaces of K because Condition (2, m) holds for every m, so J∞ :=
k=1 Jk is a linear subspace of K. An inductive argument using Con-
dition (1, m) for m ≥ 2 implies that for all k ∈ N0, k ≤ nk+2. Thus
for each k ∈ N0, Condition (3, k + 2) implies that Kk ⊆ Knk+2 ⊆ Jk+2.
k=0 Kk ⊆ J∞. It follows that K = J∞. Suppose x ∈ J∞
and x ∈ Ji ∩ Jj for some i, j ∈ N. Assuming without loss of general-
ity that i ≤ j, Ji ⊆ Jj by Condition (2, j), and by Condition (8, j),
V ∗
i VjGj = Gi. Therefore, G∗
j x. Using the Condi-
tions (7, m) for m ≥ 1, it follows that there is a well-defined contraction
C ∈ B(K) such that for every k ∈ N and x ∈ Jk, C(x) = G∗
j Vix = G∗
i x = G∗
j V ∗
k(x).
k,
(4.31)
∀k ∈ N.
kx. Thus
1 since V = V1. Since G1 = GV ∗ and P =
stated in the conclusion of the theorem. First, note that for any k ∈ N
Let us define eG := C ∗ and show that eG satisfies the four properties
and x ∈ Jk, eG∗Vkx = Cx = G∗
eG∗Vk = G∗
In particular, eG∗V = G∗
V V ∗ is the projection map onto J in B(K), we have eG∗P = V G∗V ∗.
k (Y ⊗ IH )eG =
k eG(Y ⊗ IH).
k (Y ⊗ IH)PkeG = V ∗
Properties (1) and (2) follow readily. Let Y ∈ H ∞(E, Z) and k ∈ N.
By Condition (4, k), V ∗
k (Y ⊗ IH)Pk; by Condition (5, k)
and two applications of the adjoint of equation (4.31), V ∗
V ∗
k (Y ⊗ IH )VkGk = Gk(Y ⊗ IH) = V ∗
Property (3) follows since K = J∞. Finally, C is a contraction, but
also kCk ≥ 1 since kG1k = 1. This gives property (4), completing
the proof of the theorem in the case when kGk = 1. If G is the zero
operator in B(J), then the zero operator in B(K) satisfies the desired
k (Y ⊗ IH) = V ∗
properties for eG. If G 6= 0, then a straightforward scaling argument
utilizing the case treated above produces the desired result.
(cid:3)
32
JENNIFER R GOOD
The following corollary generalizes Theorem 4.1 to the case where
the W ∗-algebra is represented on two Hilbert spaces. The proof makes
use of the so-called Putnam trick. It is this corollary that we use to
prove the weighted Nevanlinna-Pick interpolation theorem in Section
6.
Corollary 4.3. Let σ1 : M → B(H1) and σ2 : M → B(H2) be faith-
ful, normal, unital ∗-homomorphism for Hilbert spaces H1 and H2, and
let Z be a sequence of weights associated with X. For j = 1, 2, sup-
pose that Jj is a closed linear subspace of F(E) ⊗σj Hj such that for
the inclusion map of Jj into F(E) ⊗σj Hj. Suppose there exists G ∈
B(J1, J2) such that for every Y ∈ H ∞(E, Z), G (V ∗
1 (Y ⊗ IH1)V1) =
(V ∗
H2) such that
every Y ∈ H ∞(E, Z), (cid:0)Y ∗ ⊗ IHj(cid:1) (Jj) ⊆ Jj. For j = 1, 2, let Vj be
2 (Y ⊗ IH2)V2) G. Then there exists eG ∈ B(F(E) ⊗σ1 H1, F(E) ⊗σ2
(1) eG∗(J2) ⊆ J1,
2 eGV1 = G,
(3) eG(Y ⊗ IH1) = (Y ⊗ IH2)eG for all Y ∈ H ∞(E, Z), and
(4) keGk = kGk.
(2) V ∗
Proof. We will use the "Putnam trick" to translate the two-space prob-
lem into a one-space problem where we may apply the original result,
Theorem 4.1. We then return to the two-space setting by picking off
the lower left-hand entries in the 2 × 2 matricial expressions for certain
operators.
Let H = H1 ⊕ H2. Let σ = σ1 ⊕ σ2 : M → B(H). Define the
induced representation spaces K1 = F(E) ⊗σ1 H1, K2 = F(E) ⊗σ2 H2,
and K = F(E) ⊗σ H. We identify K with K1 ⊕ K2 in the usual way
and let J be the image of J1 ⊕ J2 in K. For notational simplicity we
will omit the implied isomorphisms in our computations. Let V be the
inclusion map of J into K. Let P1, P2, and P be the projection maps
in B(K1), B(K2), and B(K) onto J1, J2, and J, respectively. Define
(4.32)
G0 :=(cid:20) 0 0
G 0(cid:21) ∈ B(J).
Towards applying Theorem 4.1, let us show that for all Y ∈ H ∞(E, Z),
(4.33)
(4.34)
(Y ∗ ⊗ IH)(J) ⊆ J and
G0(V ∗(Y ⊗ IH)V ) = (V ∗(Y ⊗ IH)V )G0.
INTERPOLATION AND COMMUTANT LIFTING WITH WEIGHTS
33
By hypothesis, Pj(Y ⊗ IHj ) = Pj(Y ⊗ IHj )Pj for j = 1, 2. Thus,
0
P (Y ⊗ IH) =(cid:20)P1
0 P2(cid:21)(cid:20)Y ⊗ IH1
=(cid:20)P1
0 P2(cid:21)(cid:20)Y ⊗ IH1
Y ⊗ IH2(cid:21)
Y ⊗ IH2(cid:21)(cid:20)P1
0
0
0
0
0
0
0 P2(cid:21) = P (Y ⊗ IH )P.
Containment (4.33) follows. Since V V ∗ = P and V ∗V is the identity
on J,
(4.35) G0V ∗(Y ⊗ IH )V = V ∗V G0V ∗P (Y ⊗ IH)P V
= V ∗(cid:20)
0
V2GV ∗
1
0
0(cid:21)(cid:20)P1
0
0 P2(cid:21)(cid:20)Y ⊗ IH1
= V ∗(cid:20)
0
0
0
Y ⊗ IH2(cid:21)(cid:20)P1
0 P2(cid:21) V
1 (Y ⊗ IH1)P1 0(cid:21) V.
0
0
V2GV ∗
On the other hand,
(4.36) V ∗(Y ⊗ IH)V G0 = V ∗P (Y ⊗ IH)P V G0V ∗V
= V ∗(cid:20)P1
0 P2(cid:21)(cid:20)Y ⊗ IH1
0
0
0
Y ⊗ IH2(cid:21)(cid:20)P1
= V ∗(cid:20)
0
0 P2(cid:21)(cid:20)
0
V2GV ∗
1
0
P2(Y ⊗ IH2)V2GV ∗
1
0
0(cid:21) V
0(cid:21) V.
0
By our hypothesis, V2GV ∗
(4.34) now follows from equations (4.35) and (4.36).
1 (Y ⊗ IH1)P1 = P2(Y ⊗ IH2)V2GV ∗
1 . Equation
Having shown containment (4.33) and equation (4.34), we apply The-
∗
(J) ⊆ J,
orem 4.1 and conclude that there exists fG0 ∈ B(K) such that:
(1′) fG0
(2′) V ∗fG0V = G0,
(3′) fG0(Y ⊗ IH) = (Y ⊗ IH)fG0 for all Y ∈ H ∞(E, Z), and
(4′) kfG0k = kG0k.
With A ∈ B(K1), B ∈ B(K2, K1), C ∈ B(K2), and eG ∈ B(K1, K2)
such that
(4.37)
eG C(cid:21) ,
fG0 =(cid:20)A B
34
JENNIFER R GOOD
0
0
0
0
let us show that eG satisfies properties (1)-(4). By property (1′), PfG0 =
PfG0P , so
eG C(cid:21) = PfG0 = PfG0P
0 P2(cid:21)(cid:20)A B
P2eG P2C(cid:21) =(cid:20)P1
(cid:20)P1A P1B
P2eGP1 P2CP2(cid:21) .
eG C(cid:21)(cid:20)P1
0 P2(cid:21) =(cid:20)P1AP1 P1BP2
=(cid:20)P1
0 P2(cid:21)(cid:20)A B
from property (2′) that PfG0P = V G0V ∗, so by similar computations,
P2eGP1 P2CP2(cid:21) =(cid:20)
0(cid:21) .
(cid:20)P1AP1 P1BP2
Property (2) follows by equating the lower left-hand entries. Property
(3′) implies
Equating the lower left-hand entries, property (1) holds.
It follows
V2GV ∗
1
0
for Y ∈ H ∞(E, Z). Again, equating the lower left-hand entries, we
(Y ⊗ IH2)eG (Y ⊗ IH2)C(cid:21)
eG(Y ⊗ IH1) C(Y ⊗ IH2)(cid:21) =(cid:20)(Y ⊗ IH1)A (Y ⊗ IH1)B
(cid:20)A(Y ⊗ IH1) B(Y ⊗ IH2)
obtain property (3). Finally, it follows from equation (4.37) that keGk ≤
kfG0k. By property (4′) and equation (4.32), kfG0k = kG0k = kGk.
Since kGk ≤ keGk by property (2), we conclude that kGk = keGk. This
gives property (4) and completes the proof.
(cid:3)
5. The Commutant and Double Commutant
In this section we identify the commutant of an induced image of
the weighted Hardy algebra of the dual correspondence and the double
commutant of an induced image of the weighted Hardy algebra of the
original correspondence. For the remainder of the paper, we will assume
that E is full.
Let σ : M → B(H) be a fixed faithful, normal, unital ∗-homomor-
phism for a Hilbert space H, and let Z be a sequence of weights asso-
ciated with X. In Section 7 of [9], Muhly and Solel use X to construct
an admissible sequence for Eσ, X ′ = {X ′
k=0, and Z to construct
a sequence of weights associated with X ′, which we will denote by
Z ′ = {Z ′
k=0. Our proof of the weighted Nevanlinna-Pick theorem
will involve the commutant of the algebra {Y ⊗ IH Y ∈ H ∞(Eσ, Z ′)}
in B(F(Eσ) ⊗ι H); our next goal is to identify it with H ∞(E, Z) in
a certain way. A note of caution is in order, for what we will call
Z ′ is referred to as C in [9]. Let us summarize the construction of
X ′ and Z ′, simultaneously clarifying our notation. For k ∈ N0, we
k}∞
k}∞
k :=(Iσ(M )′
C ′
if k = 0
if k ≥ 1
.
INTERPOLATION AND COMMUTANT LIFTING WITH WEIGHTS
35
k for ϕ′
shall write Ak for ϕk(M)c, the commutant of ϕk(M) in L(E⊗k), and
A′
k(σ(M)′)c, the commutant of the image of σ(M)′ under ϕ′
k in
L((Eσ)⊗k). Using Lemma 7.2 of [9], for each k ∈ N0 we define X ′
k to
be the unique element in A′
k ⊗ IH.
For k ∈ N0, define
k such that U σ∗
k (Xk ⊗ IH)U σ
k = X ′
Ck =(IM
Z (k)(Z (k−1) ⊗ I1)−1
if k = 0
if k ≥ 1
.
k = Z ′
k (Ck ⊗ IH)U σ
k to be unique element in A′
Since Ck ∈ Ak, we define Z ′
k such that
U σ∗
k ⊗ IH ([9], Lemma 7.2). Here is where we have
modified the notation used in [9]; we have interchanged the roles of C
and Z ′ as they were introduced in Lemma 7.4 of [9]. By Theorem 7.6
of [9], X ′ = {X ′
k=0 is a
sequence of weights associated with X ′.
k=0 is an admissible sequence and Z ′ = {Z ′
k}∞
k}∞
We now repeat the construction with Eσ replacing E and Eσι re-
k(σ(M))c, the commutant of
k denote ϕ′′
placing Eσ. For k ∈ N0, let A′′
ϕ′′
k(σ(M)) in L((Eσι)⊗k). For k ∈ N0 define
Z ′(k)(Z ′(k−1) ⊗ I ′
k, there exist unique X ′′
1)−1
k , Z ′′
k = Z ′′
k }∞
k }∞
k (C ′
k (X ′
k ⊗ IH)U ι
k, C ′
k = X ′′
k such that U ι∗
k ∈ A′
k ⊗ IH and U ι∗
k ∈ A′′
k ⊗
k ⊗ IH ([9], Lemma 7.2). It
k=0 is an admissible
k=0 is a sequence of weights associated
Since X ′
IH)U ι
follows from Theorem 7.6 of [9] that X ′′ = {X ′′
sequence for Eσι and Z ′′ = {Z ′′
with X ′′. The following fact will be useful in future computations.
Lemma 5.1. For all k ∈ N0,
Proof. The case when k = 0 follows from the fact that Z0 is the identity
0 is the identity on σ(M)′. Suppose k ≥ 1. By the discus-
on M and C ′
sion preceding Theorem 7.6 in [9], Z ′(k) ⊗IH = U σ∗
k ; more-
over, by Lemma 7.2 of [9], Z ′(k−1)⊗I ′
k .
k (Z (k)⊗IH )U σ
k (Zk ⊗ IH)U σ
1⊗IH = U σ∗
k = C ′
k ⊗ IH.
U σ∗
k (cid:0)I1 ⊗ Z (k−1) ⊗ IH(cid:1) U σ
Therefore, since Z (k)(cid:0)I1 ⊗ Z (k−1)(cid:1)−1
k ⊗ IH = (Z ′(k) ⊗ IH)(Z ′(k−1) ⊗ I ′
C ′
= Zk,
1 ⊗ IH)−1
= U σ∗
k (Z (k) ⊗ IH)(cid:0)I1 ⊗ Z (k−1) ⊗ IH(cid:1)−1
as desired.
U σ
k = U σ∗
k (Zk ⊗ IH)U σ
k ,
(cid:3)
We already know how to naturally identify M with σ(M), E with
Eσι, and σ with . To identify X with X ′′ and Z with Z ′′, that is to
show that
(5.1)
ωkZkω∗
k = Z ′′
k ,
and
ωkXkω∗
k = X ′′
k ,
36
JENNIFER R GOOD
k U ι
note that for all k ∈ N0, U σ
E⊗k ⊗σ H. By Lemma 5.1, U ι∗
k (C ′
ωkZkω∗
ωkXkω∗
k ⊗ IH )U ι
k ⊗ IH , so by the uniqueness of Z ′′
k = X ′′
k .
k(ωk ⊗ IH) is the identity operator on
k =
k . Analogously,
k = U ι∗
k , ωkZkω∗
k (Zk ⊗ IH)U σ
k = Z ′′
k U σ∗
k U ι
Lemma 5.2. Ad(ω∞)(H ∞(E, Z)) = H ∞(Eσι, Z ′′).
(5.2)
Proof. It suffices to show that
∞ = W Z ′′
∞ = ϕ′′
ξ ω∗
ω∞ϕ∞(a)ω∗
ω∞W Z
(5.3)
ωk(ξ),
∞(σ(a)),
∀k ∈ N, ∀ξ ∈ E⊗k,
and
∀a ∈ M.
By straightforward computation with equation (5.1), ωiZ (i,i−k)ω∗
i =
Z ′′(i,i−k) when i ≥ k ≥ 0, so ω∞DZ
∞ = Tωk(ξ),
equation (5.2) follows. Equation (5.3) is simply the case when k =
0.
(cid:3)
k . Since ω∞Tξω∗
∞ = DZ ′′
k ω∗
The following theorem gives two commutant results. The first plays
a role at the crucial step of the weighted Nevanlinna-Pick theorem in
the next section. The second is a weighted double commutant theorem.
Theorem 5.3. In the notation established,
(5.4)
Proof. By Theorem 7.7 in [9],
= πσ(H ∞(E, Z)), and
= σF(E) (H ∞(E, Z)).
(1) (cid:0)ιF(Eσ)(H ∞(Eσ, Z ′))(cid:1)′
(2) (cid:0)σF(E) (H ∞(E, Z))(cid:1)′′
That same theorem, now applied to Eσ, gives(cid:0)ιF(Eσ) (H ∞ (Eσ, Z ′))(cid:1)′
(cid:0)ιF(Eσ) (H ∞ (Eσ, Z ′))(cid:1)′
ρι (H ∞ (Eσι, Z ′′)). Now ρι ◦Ad(ω∞) = πσ since U σ
∞U ι
identity operator on F(E) ⊗σ H, so by Lemma 5.2,
= ρι (Ad(ω∞) (H ∞(E, Z))) = πσ (H ∞(E, Z)) ,
∞) ◦ ιF(Eσ), part (1) now implies
(cid:0)σF(E) (H ∞ (E, Z))(cid:1)′
=
∞ (ω∞ ⊗ IH ) is the
= ρσ (H ∞ (Eσ, Z ′)) .
which gives part (1). Since ρσ = Ad(U σ
that
(cid:0)ρσ (H ∞ (Eσ, Z ′))(cid:1)′ = Ad(U σ
= Ad(U σ
∞)(cid:16)(cid:0)ιF(Eσ) (H ∞ (Eσ, Z ′))(cid:1)′(cid:17)
∞) (πσ (H ∞(E, Z))) .
Since Ad(U σ
∞) ◦ πσ = σF(E) equation (5.4) gives
(cid:0)σF(E) (H ∞(E, Z))(cid:1)′′
=(cid:0)ρσ (H ∞ (Eσ, Z ′))(cid:1)′
= Ad(U σ
∞) (πσ (H ∞(E, Z))) = σF(E) (H ∞(E, Z)) ,
which completes the proof of part (2).
(cid:3)
INTERPOLATION AND COMMUTANT LIFTING WITH WEIGHTS
37
The following corollary to Theorem 5.3(1) will be of use in the next
section. For s, t ∈ N, H (s) denotes the direct sum of s copies of H,
Ms×t(B(H)) denotes the collection of s × t matrices with entries in
B(H), identified with B(H (t), H (s)) in the usual way, and Ms(B(H))
denotes the collection of s×s matrices with entries in B(H). We define
ι(s) : σ(M)′ → Ms(B(H)) to be the direct sum of s copies of ι. Define
πσ
s×t : Ms×t (L(F(E))) → Ms×t (B(F(Eσ) ⊗ι H)) so that
πσ
s×t(cid:16)[Yaf ]s
a=1
t
f =1(cid:17) = [πσ(Yaf )]s
t
f =1 .
a=1
We identify, as usual, F(Eσ) ⊗ι(s) H (s) with (F(Eσ) ⊗ι H)(s).
Corollary 5.4. Let s, t ∈ N. The image of Ms×t (H ∞ (E, Z)) under
πσ
s×t is precisely the collection of elements A ∈ Ms×t (B(F(Eσ) ⊗ι H))
such that for every S ∈ H ∞ (Eσ, Z ′),
A ◦ (ι(t))F(Eσ)(S) = (ι(s))F(Eσ)(S) ◦ A.
6. Weighted Nevanlinna-Pick Interpolation
In this section we give our primary theorem, a weighted W ∗-version
of the classic Nevanlinna-Pick interpolation theorem. While the key
component of its proof, the weighted commutant lifting theorem, is
already at our disposal, we begin this section with a series of technical
lemmas that further facilitate the proof. Let σ : M → B(H) be a fixed
faithful, normal, unital ∗-homomorphism for a Hilbert space H, and
let Z be a sequence of weights associated with X.
Let us establish some notation. If z ∈ D(X, σ), then the map Φz :
k=1 z(k)(Xk ⊗
(k)∗ is completely positive, linear, and ultraweakly continuous by
A)z
Lemma 4.4 of [9]; in addition, kΦzk < 1. Therefore, in the Banach
z converges in
σ(M)′ → σ(M)′, defined for A ∈ σ(M)′ by Φz(A) = P∞
algebra of completely bounded maps on σ(M)′,P∞
norm. In fact,
j=0 Φj
(6.1)
Φj
z (A) =
(k)(R2
k ⊗ A)z
z
(k)∗,
A ∈ σ(M)′
∞Xj=0
∞Xk=0
by Theorem 4.5 of [9], where the summation on the left-hand side of
the equation is with respect to the norm topology on σ(M)′, and the
summation on the right-hand side is with respect to the ultraweak
topology on σ(M)′.
Let z ∈ I(σE ◦ ϕ, σ). For k ∈ N0, (cid:0)(Z (k))−1(cid:1)∗
Thus(cid:0)(cid:0)(Z (k))−1(cid:1)∗ ⊗ IH(cid:1) z(k)∗ belongs to(cid:0)E⊗k(cid:1)σ
cz(k) to be its image in (Eσ)⊗k under the isomorphism Λσ∗
belongs to ϕk(M)c.
z (k) =
k . For w, z ∈
, and we define cZ
38
JENNIFER R GOOD
D(X, σ) and A ∈ σ(M)′, we have hcw(k), A · cz(k)i = w(k)(R2
Define the Z-Cauchy kernel at z to be the tuple,
k ⊗ A)z(k)∗.
z = cz = (cz(k))∞
cZ
k=0 .
Lemma 6.1. If z ∈ D(X, σ), then cz ∈ F(Eσ).
Proof. By equation (6.1),P∞
so for any N ∈ N, PN
P∞
k=0 z(k)(R2
result follows.
k ⊗ IH)z(k)∗ converges in σ(M)′,
k ⊗ IH )z(k)∗ ≤
k ⊗ IH )z(k)∗. Since cz(k) ∈ (Eσ)⊗k for each k ∈ N0, the
(cid:3)
k=0 hcz(k), cz(k)i = PN
k=0 z(k)(R2
k=0 z(k)(R2
Let w, z ∈ D(X, σ). We define K(w, z) : σ(M)′ → B(H) by
K(w, z)(A) = hcw, A · czi =
∞Xk=0
(k)(R2
w
k ⊗ A)z
(k)∗, A ∈ σ(M)′.
For s ∈ N, let K(s)(w, z) : σ(M)′ → Ms(B(H)) be the map defined
at A ∈ σ(M)′ by K(s)(w, z)(A) = ι(s)(K(w, z)(A)) where, as in the
preceding section, ι(s) is the direct sum of s copies of ι.
Remark 6.2. While we will not explore the consequences here, it is read-
ily shown that K : D(X, σ) × D(X, σ) → CB∗(σ(M)′, B(H)) is a "nor-
mal completely positive kernel" in the sense of [4] and [6], independent
of the choice of Z, where CB∗(σ(M)′, B(H)) denotes the completely
bounded ultraweakly continuous maps from σ(M)′ to B(H). As such,
by Proposition 41 of [6], K has an associated "Reproducing Kernel W ∗-
Correspondence", a W ∗-analogue of the classical Reproducing Kernel
Hilbert Space; in fact K, which we may refer to as the (X, σ)-Szego
kernel, is a weighted W ∗-version of the classic Szego kernel.
Recall thatcIH = (δi=0 IH )∞
i=0 belongs to F(Eσ) since IH is the multi-
plicative identity of σ(M)′ = (Eσ)⊗0. If z ∈ D(X, σ), then cz belongs to
F(Eσ) by Lemma 6.1. We write LI and Lz for the insertion operators
LH
cIH
Lemma 6.3. For z ∈ D(X, σ) and Y ∈ H ∞(E, Z),
, respectively.
and LH
cz
(1) L∗
(2) L∗
(3) L∗
z πσ(Y )LIL∗
z = L∗
z πσ(Y ),
z πσ(Y )LI = bY (z), and
z πσ(Y ) = bY (z)L∗
z .
Proof. Let us show that parts (1) and (2) hold when Y = ϕ∞(a) when
a ∈ M. Since πσ(ϕ∞(a)) = I ′
∞ ⊗ σ(a), hcz,cIHi = IH, and σ(a) =
INTERPOLATION AND COMMUTANT LIFTING WITH WEIGHTS
39
(σ × z)(ϕ∞(a)),
L∗
z πσ(ϕ∞(a))LI = L∗
z (I ′
∞ ⊗ σ(a)) LI = L∗
z LIσ(a)
= ιhcz,cIHiσ(a) = σ(a) = (σ × z)(ϕ∞(a)),
which gives part (2). Part (1) now follows from the fact that if η ∈
F(Eσ) and x ∈ H, then L∗
z Lη = hcz, ηi belongs to σ(M)′, so using part
(2),
L∗
z πσ(ϕ∞(a))LIL∗
z (η ⊗ x) = σ(a)L∗
z Lηx = L∗
z Lησ(a)x
= L∗
z (I ′
∞ ⊗ σ(a))Lηx = L∗
z πσ(ϕ∞(a))(η ⊗ x).
Now we show that parts (1) and (2) hold for Y = Wξ when ξ ∈ E.
The ith diagonal entry of the diagonal operator πσ(D1) is zero when
i = 0 and is otherwise U σ∗
i ⊗ IH (Lemma 5.1). It
follows that
i (Zi ⊗ IH)U σ
i = C ′
πσ(Wξ) = πσ(D1)πσ(Tξ)
(6.2)
If Li denotes the insertion operator LH
cially Lz is the column [Li]∞
i=0; hence L∗
By Lemma 5.1, C ′
U σ
1 (cz(1) ⊗ y) = cz(1)(y), when x, y ∈ H,
1 ⊗ IH = U σ∗
1 (Z1 ⊗ IH)U σ
cz(i) for each i ∈ N0, then matri-
z πσ(Wξ)LI = L∗
1 ⊗ IH )(ωξ).
1 ◦ (ωξ) = Lξ and
1 . Since U σ
1(C ′
(cid:10)L∗
z πσ(Wξ)LIx, y(cid:11) = h(C ′
1 ⊗ IH)(ωξ)x, cz(1) ⊗ yi
=(cid:10)(Z1 ⊗ IH )Lξx,(cid:0)(Z −1
1 )∗ ⊗ IH(cid:1) z
∗y(cid:11)
= hzLξx, yi = h(σ × z)(Wξ)x, yi .
We deduce that part (2) holds for Y = Wξ. Let us compare the row
matrices for the operators on either side of the equality in part (1). By
part (2), L∗
j=0. On the other hand, by equa-
tion (6.2), L∗
. Therefore to
obtain part (1), it suffices to show that for every j ∈ N0,
z πσ(Wξ) =(cid:2)L∗
z πσ(Wξ)LIL∗
j+1(C ′
j=0
z = (cid:2)zLξL∗
j(cid:3)∞
j+1 ⊗ IH)(cid:0)I ′
j+1 ⊗ IH)(cid:0)I ′
j+1(C ′
j ⊗ ωξ(cid:1)(cid:3)∞
j ⊗ ωξ(cid:1) .
zLξL∗
j = L∗
(6.3)
0
(C ′
1 ⊗ IH)ωξ
0
0
0
0
(C ′
2 ⊗ IH)(I ′
1 ⊗ ωξ)
0
0
0
0
(C ′
3 ⊗ IH)(I ′
. . .
2 ⊗ ωξ)
. . .
. . .
=
.
40
JENNIFER R GOOD
For any k ∈ N0, U σ
so
k Lk =(cid:0)((Z (k))−1)∗ ⊗ IH(cid:1) z(k)∗ by definition of cz(k),
(6.4)
L∗
kU σ∗
k = z
(k)(cid:0)((Z (k))−1) ⊗ IH(cid:1) ,
k ∈ N0.
Since Z (j+1) = Zj+1(I1 ⊗ Z (j)) and z
of equation (6.4), first when k = j + 1 and then when k = j, yield
(j+1) = z(I1 ⊗ z
(j)), two applications
(6.5) L∗
j+1U σ∗
j+1 (Zj+1 ⊗ IH) = z
= z(cid:16)I1 ⊗ z
(j+1)(cid:0)((Z (j+1))−1) ⊗ IH(cid:1) (Zj+1 ⊗ IH)
(j)(cid:0)((Z (j))−1) ⊗ IH(cid:1)(cid:17) = z(cid:0)I1 ⊗ L∗
j (cid:1)
j ⊗ ωξ(cid:1) = L(E⊗j ⊗σH)
j+1(cid:0)I ′
j ⊗ ωξ(cid:1) = L∗
j ⊗ ωξ(cid:1)
j+1(cid:0)I ′
j+1 (Zj+1 ⊗ IH) U σ
j )L(E⊗j⊗σH)
j = zLξL∗
j ,
= z(I1 ⊗ L∗
j+1U σ∗
j U σ∗
j U σ∗
U σ
ξ
ξ
U σ
j . Thus by
As observed in Section 2.2, U σ
Lemma 5.1 and equation (6.5),
L∗
j+1(C ′
j+1 ⊗ IH)(cid:0)I ′
noting that the final equality follows readily from an elementwise com-
putation. This gives equation (6.3), so part (1) holds for Y = Wξ when
ξ ∈ E.
To show that part (1) holds for arbitrary operators in H ∞(E, Z),
suppose, inductively, that k ∈ N exists such that part (1) holds when-
ever Y = Wξ for ξ ∈ E⊗j with 0 ≤ j ≤ k. For ξ ∈ E⊗k and η ∈ E,
Wξ⊗η = Wξ ⊗ Wη, so
L∗
z πσ(Wξ⊗η)LIL∗
z = L∗
z πσ(Wξ)LIL∗
= L∗
z πσ(Wξ)πσ(Wη)LIL∗
z πσ(Wη)LIL∗
z = L∗
z πσ(Wξ)πσ(Wη) = L∗
= L∗
z πσ(Wξ)LIL∗
z
z πσ(Wη)
z πσ(Wξ⊗η).
Part (1) follows from routine approximation arguments. Towards ob-
taining part (2) in the general case, define the linear and ultraweakly
continuous map τ : H ∞(E, Z) → B(H) at Y ∈ H ∞(E, Z) by τ (Y ) =
L∗
z πσ(Y )LI . By part (1), for Y1, Y2 ∈ H ∞(E, Z),
τ (Y1Y2) = L∗
z πσ(Y2)LI = τ (Y1)τ (Y2).
Thus τ is multiplicative. As both τ and (σ×z) are linear, multiplicative,
and ultraweakly continuous and they agree on elements of the form
Y = ϕ∞(a) when a ∈ M and Y = Wξ when ξ ∈ E, it follows that
τ = (σ × z), which gives part (2). Finally, parts (1) and (2) together
imply part (3).
(cid:3)
z πσ(Y1)πσ(Y2)LI = L∗
z πσ(Y1)LIL∗
Our final technical lemma concerns a weighted creation operator
ξ = W ′
ξ ∈ H ∞(Eσ, Z ′). Recall that if z ∈ D(X, σ), then z∗ is
W Z ′
INTERPOLATION AND COMMUTANT LIFTING WITH WEIGHTS
41
an element in Eσ and for any ξ ∈ Eσ, hξ, z∗i = ξ∗z∗ is an element of
σ(M)′.
Lemma 6.4. For z ∈ D(X, σ), ξ ∈ Eσ, and D ∈ σ(M)′,
(cid:0)W ′
ξ(cid:1)∗ (D · cz) = hD∗ · ξ, z
∗i · cz.
(Z ′
ξ
Λσ
z
k+1(ζ).
(6.6)
(Z ′
ξ
where T (k)
k=0
(Z ′
= z
= z
Thus Λσ
ξ ⊗ η,
k+1(Z ′
k+1ζ ⊗ x)
In particular, when ζ =
∗i · cz = (ξ∗
z
∗ · cz(k))∞
∗ · cz(k), ηE.
= (Ck+1 ⊗ IH )U σ
k+1(ζ ⊗ x) = (Ck+1 ⊗ IH ) Λσ
k+1(ζ)x.
If ζ ∈ (Eσ)⊗k+1 and x ∈ H, then by the definition of Z ′ in terms of C,
Proof. First, let us show that the result holds when D = IH. We
have that (W ′
ξ maps
η ∈ (Eσ)⊗k to ξ ⊗ η ∈ (Eσ)⊗k+1. Also, hξ, z
k=0, so
it suffices to show that for every k ∈ N0 and η ∈ (Eσ)⊗k,
k+1)∗ (cz(k + 1))(cid:17)∞
k+1)∗ (cz(k + 1)) , ηE =Dξ∗
ξ)∗cz = (cid:16)T (k)∗
DT (k)∗
k+1ζ(cid:1) x = U σ
k+1(cid:0)Z ′
k+1(cid:0)Z ′
k+1ζ(cid:1) = (Ck+1 ⊗ IH) Λσ
(6.7) DT (k)∗
k+1)∗ (cz(k + 1)) , ηE =(cid:10)cz(k + 1), Z ′
k+1(ξ ⊗ η)(cid:11)
(k+1)(cid:0)(Z (k+1))−1 ⊗ IH(cid:1) Λσ
k+1(ξ ⊗ η)(cid:1)
k+1(cid:0)Z ′
(k+1)(cid:0)(Z (k+1))−1Ck+1 ⊗ IH(cid:1) Λσ
(k+1)(cid:0)(Z (k))−1 ⊗ I1 ⊗ IH(cid:1) Λσ
(k)(cid:0)(Z (k))−1 ⊗ IH(cid:1) Λσ
(k)(cid:0)(Z (k))−1 ⊗ IH(cid:1) (Ik ⊗ zξ) Λσ
(k+1)(cid:0)(Z (k))−1 ⊗ I1 ⊗ IH(cid:1) Λσ
∞(D) = (W ′
D∗·ξ)∗ · cz = hD∗ · ξ, z∗i · cz, as desired.
From equations (6.7) and (6.8), we obtain equation (6.6). Thus, the
conclusion of the lemma holds when D = IH . For arbitrary D, we
D∗·ξ)∗. Therefore, by our first case,
observe that (W ′
(W ′
(cid:3)
ξ)∗ϕ′
ξ)∗ (D · cz) = (W ′
We are ready to present our main result, a weighted Nevanlinna-Pick
interpolation theorem. We phrase and prove a more-general, matricial
version of the theorem that, aside from increased notational complex-
ity, poses no additional difficulty; the main result, that which gener-
alizes the classic Nevanlinna-Pick interpolation theorem, occurs when
∗ · cz(k), ηE = z
(6.8) Dξ∗
On the other hand,
k (zξ · η)
k (η)
k+1 (ξ ⊗ η) .
= z
= z
= z
k+1(ξ ⊗ η).
k+1(ξ ⊗ η)
ξ
z
42
JENNIFER R GOOD
we take s = t = 1 in the statement below. While we make necessary
adjustments for the weights, our proof mirrors that of the unweighted
Nevanlinna-Pick interpolation result given as Theorem 5.3 in [8].
Theorem 6.5 (Weighted Nevanlinna-Pick Interpolation). Fix s, t ∈ N.
Let k ∈ N. Choose {zi}k
i=1 ⊆
Ms (B (H)) and {Fi}k
i=1 ⊆ Ms×t (B (H)). Define A : Mk(σ(M)′) →
Mk (Ms (B (H))) at [Aij]k
i=1 ⊆ D(X, σ) and two collections, {Bi}k
i,j=1 ∈ Mk(σ(M)′) by
A(cid:0)[Aij]k
i,j=1(cid:1) =
(cid:2)Bi · K(s)(zi, zj)(Aij) · B∗
Then the following conditions are equivalent:
j − Fi · K(t)(zi, zj)(Aij) · F ∗
.
i,j=1
(1) For every n ∈ N and for any {hpi 1 ≤ i ≤ k, 1 ≤ p ≤ n} ⊆
H (s) and {Api 1 ≤ i ≤ k, 1 ≤ p ≤ n} ⊆ σ(M)′,
j(cid:3)k
2
,
i hpi!(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Api · czi ⊗ F ∗
Api · czi ⊗ B∗
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
nXp=1 kXi=1
2
i hpi!(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
≤(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
nXp=1 kXi=1
where on the left-hand side of the inequality, the norm occurs in
the Hilbert space F(Eσ)⊗ι(t) H (t) and on the right in F(Eσ)⊗ι(s)
H (s);
(2) A is completely positive; and
t
(3) there exists Y = [Yaf ]s
a=1
f =1 ∈ Ms×t (H ∞(E, Z)) such that
kY k ≤ 1 and
BihcYaf (zi)is
t
f =1
= Fi
a=1
for every i ∈ N with 1 ≤ i ≤ k.
Remark 6.6. The map A defined in the statement of Theorem 6.5 does
not depend on Z, so if the theorem is satisfied for one sequence of
weights associated with X it holds for every sequence of weights asso-
ciated with X.
Proof. First let us perform a preliminary computation. Let n ∈ N
and choose two collections {hpi 1 ≤ i ≤ k, 1 ≤ p ≤ n} ⊆ H (s) and
p=1 in
{Api 1 ≤ i ≤ k, 1 ≤ p ≤ n} ⊆ σ(M)′. Define eh := (cid:0)(hpi)k
i=1(cid:1)n
INTERPOLATION AND COMMUTANT LIFTING WITH WEIGHTS
43
((H (s))(k))(n). Since K(zi, zj)(A∗
piAqj) = hApi · czi, Aqj · czj i,
j hqj(cid:11)!
Thus, condition (2) implies condition (1).
=
piAqj]k
Api · czi ⊗ F ∗
2
,
piAqj) · B∗
p,q=1
piAqj) · F ∗
p,q=1
where the norm occurs in the Hilbert space F(Eσ) ⊗ι(t) H (t) and the in-
ner product is taken in ((H (s))(k))(n). Since an analogous result follows
with each Fi replaced with Bi we compute,
(eh)(cid:29)
i,j=1in
j(cid:3)k
i hpi, ι(t)(cid:0)K(zi, zj)(A∗
piAqj)(cid:1) F ∗
i hpi!(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
(cid:28)eh,h(cid:2)Fi · K(t)(zi, zj)(A∗
nXp,q=1 kXi,j=1(cid:10)F ∗
=(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
nXp=1 kXi=1
Deh, An(cid:16)(cid:2)[A∗
p,q=1(cid:17) (eh)E
i,j=1(cid:3)n
=(cid:28)eh,h(cid:2)Bi · K(s)(zi, zj)(A∗
(eh)(cid:29)
i,j=1in
j(cid:3)k
−(cid:28)eh,h(cid:2)Fi · K(t)(zi, zj)(A∗
i,j=1in
j(cid:3)k
=(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
−(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
i hpi!(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
i hpi!(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
nXp=1 kXi=1
nXp=1 kXi=1
completely positive. Let A :=(cid:2)δp=1 [δi=1 Aqj]k
i,j=1(cid:3)n
i,j=1(cid:3)n
when i = 1 and is otherwise equal to 0. Then A∗A =(cid:2)[A∗
An(A∗A) is positive in B(cid:0)((H (s))(k))(n)(cid:1). Taking eh := (cid:0)(hpi)k
i=1(cid:1)n
0 ≤Deh, An(A∗A)(eh)E
=(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
i hpi!(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Let us show that condition (2) implies condition (1). Suppose A is
p,q=1 where δi=1 is 1
p,q=1
is positive in Mn(Mk(σ(M)′)). Since An is positive, we have that
in ((H (s))(k))(n) and using equation (6.9) we have that
nXp=1 kXi=1
2
i hpi!(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
−(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
nXp=1 kXi=1
piAqj) · F ∗
p,q=1
Api · czi ⊗ B∗
Api · czi ⊗ F ∗
Api · czi ⊗ B∗
Api · czi ⊗ F ∗
2
.
p=1
2
.
(eh)(cid:29)
piAqj]k
(6.9)
2
44
JENNIFER R GOOD
piAqj]k
Now let us assume condition (1) and show that condition (2) holds.
Let n ∈ N and choose a collection {Api 1 ≤ i ≤ k, 1 ≤ p ≤ n} of ele-
p=1 for some {hpi 1 ≤ i ≤ k, 1 ≤ p ≤ n} ⊆ H (s).
Since we are assuming condition (1), we deduce from equation (6.9)
that
ments in σ(M)′. An arbitrary element of B(cid:0)((H (s))(k))(n)(cid:1) may be writ-
ten eh = (cid:0)(hpi)k
i=1(cid:1)n
Deh, An(cid:16)(cid:2)[A∗
p,q=1(cid:17) (eh)E
i,j=1(cid:3)n
=(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
−(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
i hpi!(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
nXp=1 kXi=1
nXp=1 kXi=1
is nonnegative. Therefore An(cid:16)(cid:2)[A∗
i,j=1(cid:3)n
i,j=1(cid:3)n
a sum of elements of the form(cid:2)[A∗
an arbitrary positive element in Mn(Mk(σ(M)′)) can be expressed as
, it follows that An is
positive. Therefore, A is completely positive; thus condition (1) implies
condition (2).
i hpi!(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
p,q=1(cid:17) is positive. Since
Api · czi ⊗ B∗
Api · czi ⊗ F ∗
piAqj]k
piAqj]k
p,q=1
2
2
i , and Lt
Let us pause for some notational simplifications.
i denote the insertion operators LH
czi
If 1 ≤ i ≤ k,
then Li, Ls
, and
LH (t)
, respectively. Observe that Li is Lzi in the notation of Lemma
czi
, LH (s)
6.3. Also, LI , Ls
,
cIH
and LH (t)
, respectively. Since the identity operators on E⊗s and E⊗t
cIH
do not occur in our computations for the remainder of the proof, we
temporarily write Is in place of IH (s) and It in place of IH (t). Note that
when A ∈ σ(M)′, 1 ≤ i ≤ k, and h ∈ H (s),
I denote the insertion operators LH
cIH
I, and Lt
, LH (s)
czi
(6.10) A · czi ⊗ h = (ϕ′
∞(A) ⊗ Is) (czi ⊗ h) = (ι(s))F(Eσ)(ϕ′
∞(A))Ls
i h.
A similar result holds with t in place of s.
Suppose Y = [Yaf ] satisfies the properties in condition (3), using
a ∈ {1, . . . , s} to indicate indicate rows and f ∈ {1, . . . , t} to indicate
columns for an s×t matrix. We show condition (1). Let i ∈ {1, . . . , k}.
Since Ls
i are diagonal operators with Li in each diagonal entry,
by Lemma 6.3(3),
i and Lt
FiLt∗
i = BihcYaf (zi)i Lt∗
ii
i = BihcYaf (zi)L∗
= Bi [L∗
i πσ(Yaf )] = BiLs∗
i πσ
s×t (Y ) .
INTERPOLATION AND COMMUTANT LIFTING WITH WEIGHTS
45
Taking adjoints, we have Lt
(6.10) and Corollary 5.4, for 1 ≤ p ≤ n,
iF ∗
i B∗
i . Thus by equation
Api · czi ⊗ F ∗
Since kY k ≤ 1 and πσ
nXp=1 kXi=1
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
∞(Api))Lt
iF ∗
i hpi
(ι(s))F(Eσ)(ϕ′
s×t is a linear isometry,
i =(cid:0)πσ
i hpi = (ι(t))F(Eσ)(ϕ′
= (ι(t))F(Eσ)(ϕ′
s×t(Y )(cid:1)∗ Ls
∞(Api))(cid:0)πσ
s×t (Y )(cid:1)∗
=(cid:0)πσ
=(cid:0)πσ
s×t (Y )(cid:1)∗ (Api · czi ⊗ B∗
i hpi!(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
s×t (Y )(cid:13)(cid:13)2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
nXp=1 kXi=1
≤(cid:13)(cid:13)πσ
≤(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
nXp=1 kXi=1
Api · czi ⊗ F ∗
2
Api · czi ⊗ B∗
s×t (Y )(cid:1)∗ Ls
∞(Api))Ls
i hpi) .
i B∗
i B∗
i hpi
i hpi
2
i hpi!(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Api · czi ⊗ B∗
2
.
i hpi!(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Therefore condition (3) implies condition (1).
Finally, we show that condition (1) implies condition (3). Let JB be
the norm-closure of the subspace of F(Eσ) ⊗ι(s) H (s) comprised of the
elements of the form
nXp=1 kXi=1
Api · czi ⊗ B∗
i hpi!
i=1 Api · czi ⊗ F ∗
for some n ∈ N and some collections {hpi 1 ≤ i ≤ k, 1 ≤ p ≤ n} ⊆
H (s) and {Api 1 ≤ i ≤ k, 1 ≤ p ≤ n} ⊆ σ(M)′. Similarly, define JF
to be the norm-closure of the subspace of F(Eσ) ⊗ι(t) H (t) consisting
i hpi(cid:17). As we are
of the elements of the formPn
p=1(cid:16)Pk
assuming condition (1), it follows that there is a well-defined, contrac-
tive linear map R : JB → JF such that for any 1 ≤ i ≤ k, A ∈ σ(M)′,
and h ∈ H (s),
R (A · czi ⊗ B∗
i h) = A · czi ⊗ F ∗
i h.
We aim to apply Corollary 4.3, corollary to the weighted commutant
lifting theorem, with σ(M)′, Eσ, X ′, and Z ′ replacing M, E, X, and Z.
In the statement of that corollary, we take σ1 = ι(t) on H (t), σ2 = ι(s)
on H (s), J1 = JF in F(Eσ) ⊗ι(t) H (t), and J2 = JB in F(Eσ) ⊗ι(s) H (s).
Let VF and VB denote the inclusion maps of JF into F(Eσ) ⊗ι(t) H (t)
46
JENNIFER R GOOD
and JB into F(Eσ) ⊗ι(s) H (s), respectively. Let G = R∗ in B(JF , JB).
Let us show that for any S ∈ H ∞(Eσ, Z ′),
(i) (S∗ ⊗ It) (JF ) ⊆ JF ,
(ii) (S∗ ⊗ Is) (JB) ⊆ JB, and
(iii) G (V ∗
F (S ⊗ It) VF ) = (V ∗
B (S ⊗ Is) VB) G.
It suffices to show that each of these properties holds when S = ϕ′
∞(A)
ξ for ξ ∈ Eσ. When D ∈ σ(M)′,
for A ∈ σ(M)′ and when S = W ′
∞(A)∗ ⊗ It) (D · czi ⊗ F ∗
i h) = A∗D·
1 ≤ i ≤ k, and h ∈ H (s), we have (ϕ′
czi ⊗ F ∗
i h) =
hD∗ · ξ, z∗
i h. Property (i) now follows from the definition of
JF ; property (ii) holds by similar reasoning. To obtain property (iii),
we observe that if D ∈ σ(M)′, 1 ≤ i ≤ k, and h ∈ H (s),
i h. Moreover, by Lemma 6.4, (cid:0)(W ′
ξ)∗ ⊗ It(cid:1) (D · czi ⊗ F ∗
i i · czi ⊗ F ∗
(V ∗
B(ϕ′
∞(A) ⊗ Is)VBG)∗(D · czi ⊗ B∗
i h)
= RV ∗
∞(A∗) ⊗ Is)VB(D · czi ⊗ B∗
i h)
= (A∗D) · czi ⊗ F ∗
B(ϕ′
i h = V ∗
F (ϕ′
= (GV ∗
∞(A) ⊗ It)VF )∗ = (V ∗
F (ϕ′
B(ϕ′
F (ϕ′
Thus (GV ∗
adjoints, we obtain property (iii) for S = ϕ′
6.4,
∞(A∗) ⊗ It)VF R(cid:16)D · czi ⊗ B∗
i h(cid:17)
∞(A) ⊗ It)VF )∗(D · czi ⊗ B∗
∞(A) ⊗ Is)VBG)∗, and by taking
∞(A). Finally, by Lemma
i h)
i h)
i h)
= V ∗
(D · czi ⊗ B∗
(D · czi ⊗ B∗
∗
i i · czi ⊗ F ∗
i h
(cid:0)V ∗
B(cid:0)W ′
= RV ∗
i h) = hD∗ · ξ, z
i h)
adjoints, we obtain property (iii) for S = W ′
Having satisfied the hypothesis of Corollary 4.3, we conclude that
ξ ⊗ Is(cid:1) VBG(cid:1)∗
B(cid:0)(W ′
ξ)∗ ⊗ Is(cid:1) VB (D · czi ⊗ B∗
F(cid:0)(W ′
ξ)∗ ⊗ It(cid:1) VF R (D · czi ⊗ B∗
F(cid:0)W ′
=(cid:0)GV ∗
ξ ⊗ It(cid:1) VF(cid:1)∗
F(cid:0)W ′
Therefore(cid:0)GV ∗
B(cid:0)W ′
ξ ⊗ It(cid:1) VF(cid:1)∗ =(cid:0)V ∗
ξ ⊗ Is(cid:1) VBG(cid:1)∗, and by taking
there exists an operator eG ∈ B(cid:0)F(Eσ) ⊗ι(t) H (t), F(Eσ) ⊗ι(s) H (s)(cid:1) such
(I) eG∗(JB) ⊆ JF ,
BeGVF = R∗,
(III) eG (S ⊗ It) = (S ⊗ Is)eG for all S ∈ H ∞(Eσ, Z ′), and
(IV) keGk = kRk.
Y = [Yaf ] in Ms×t (H ∞ (E, Z)) such that eG = πσ
Property (III) and Corollary 5.4 together imply the existence of some
s×t is
an isometry, property (IV) implies that kY k = kπσ
s×t(Y ). Since πσ
ξ, as desired.
(II) V ∗
that
s×t(Y )k = keGk =
(II), eG∗Ls
iF ∗
Lt
(6.11)
i = eG∗PBLs
i . Taking adjoints, we obtain
F )eG∗(VBV ∗
i eG = FiLt∗
i .
BiLs∗
By Lemma 6.3(2),
(6.12)
i πσ (Yaf ) LI] = Ls∗
hcYaf (zi)i = [L∗
BihcYaf (zi)i = BiLs∗
i eGLt
i Lt
I.
i eGLt
I = FiLt∗
i Lt
I = Fi.
Since czi(0) = IH, we have Lt∗
(6.11),
I = It. Thus, by equations (6.12) and
INTERPOLATION AND COMMUTANT LIFTING WITH WEIGHTS
47
kRk ≤ 1. To complete the proof, let PF ∈ B(F(Eσ) ⊗ι(t) H (t)) denote
the projection map onto JF , and let PB ∈ B(F(Eσ) ⊗ι(s) H (s)) denote
the projection map onto JB. Then VF V ∗
B = PB. For
h ∈ H (s), Ls
i h. It follows that
Ls
i = PBLs
i . Thus by properties (I) and
i B∗
i =
F = PF and VBV ∗
i h = czi ⊗ F ∗
iF ∗
i h = czi ⊗ B∗
i and VF RV ∗
i h and Lt
BLs
iF ∗
i = Lt
i = VF RV ∗
i = (VF V ∗
i B∗
i B∗
B)Ls
i B∗
BLs
i B∗
i B∗
i B∗
i B∗
Thus condition (1) implies condition (3), which completes the proof.
(cid:3)
Remark 6.7. In the scalar case, when M = E = H = C, our kernel
K is simply the reproducing kernel to a weighted Hardy space. Be-
cause one of the hypothesis of the admissible sequence X is that each
Xk is positive, it follows from Theorem 7.33 of [1] that the weighted
Hardy spaces under consideration are those that satisfy the complete
Pick Property that is described in Definition 5.13 in [1]. Thus, for
example, our theorem applies to the Hardy and Dirichlet spaces, but
not the Bergman space, by Corollary 7.37, Corollary 7.41, and Exam-
ple 5.17 of [1]. The original complete Pick property involves certain
matrix-valued multipliers, and while we will not include the discussion
in the present paper, notions such as reproducing kernel Hilbert spaces
and their spaces of multipliers can be generalized to the W ∗-setting,
which produces interesting examples of noncommutative functions. A
promising topic of future work involves the formulation of a W ∗-version
of the complete Pick property and an investigation of its implications
for the representation theory of H(E, Z), as was begun in [9].
Acknowledgements
Thank you to Paul Muhly and Baruch Solel for their support, en-
couragement, and helpful feedback in this endeavor.
48
JENNIFER R GOOD
References
[1] Jim Agler and John E. McCarthy. Pick interpolation and Hilbert function
spaces, volume 44 of Graduate Studies in Mathematics. American Mathemati-
cal Society, Providence, RI, 2002.
[2] Alvaro Arias. Projective modules on Fock spaces. J. Operator Theory,
52(1):139 -- 172, 2004.
[3] Michel Baillet, Yves Denizeau, and Jean-Fran¸cois Havet.
Indice d'une
esp´erance conditionnelle. Compositio Math., 66(2):199 -- 236, 1988.
[4] Stephen D. Barreto, B. V. Rajarama Bhat, Volkmar Liebscher, and Michael
Skeide. Type I product systems of Hilbert modules. J. Funct. Anal.,
212(1):121 -- 181, 2004.
[5] E. C. Lance. Hilbert C ∗-modules, volume 210 of London Math. Soc. Lecture
Note Ser. Cambridge University Press, 1995.
[6] Jonas R. Meyer. Noncommutative Hardy Algebras, Multipliers, and Quotients.
PhD thesis, The University of Iowa, 2010.
[7] Paul S. Muhly and Baruch Solel. Tensor algebras over C ∗-correspondences:
representations, dilations, and C ∗-envelopes. J. Funct. Anal., 158(2):389 -- 457,
1998.
[8] Paul S. Muhly and Baruch Solel. Hardy algebras, W ∗-correspondences and
interpolation theory. Math. Ann., 330(2):353 -- 415, 2004.
[9] Paul S. Muhly and Baruch Solel. Matricial function theory and weighted shifts.
Integral Equations Operator Theory, 84(4):501 -- 553, 2016.
[10] R. Nevanlinna. Uber beschrankte funktionen, die in gegebenen punkten
vorgeschrieben werte annehmen. Ann. Acad. Sci. Fenn. Ser. A, 13(1), 1919.
[11] Stephen Parrott. On a quotient norm and the Sz.-Nagy - Foia¸s lifting theorem.
J. Funct. Anal., 30(3):311 -- 328, 1978.
[12] William L. Paschke. Inner product modules over B ∗-algebras. Trans. Amer.
Math. Soc., 182:443 -- 468, 1973.
[13] Georg Pick. Uber die Beschrankungen analytischer Funktionen, welche
durch vorgegebene Funktionswerte bewirkt werden. Mathematische Annalen,
77(1):7 -- 23, 1915.
[14] Gelu Popescu. Operator theory on noncommutative domains. Mem. Amer.
Math. Soc., 205(964):vi+124, 2010.
[15] Marc A. Rieffel. Induced representations of C ∗-algebras. Advances in Math.,
13:176 -- 257, 1974.
[16] Donald Sarason. Generalized interpolation in H ∞. Trans. Amer. Math. Soc.,
127:179 -- 203, 1967.
[17] Allen L. Shields. Weighted shift operators and analytic function theory. In
Topics in operator theory, pages 49 -- 128. Math. Surveys, No. 13. Amer. Math.
Soc., Providence, R.I., 1974.
[18] B´ela Sz.-Nagy and Ciprian Foia¸s. Dilatation des commutants d'op´erateurs. C.
R. Acad. Sci. Paris S´er. A-B, 266:A493 -- A495, 1968.
Department of Mathematics, University of Wisconsin - Platteville,
Platteville, WI 53818
E-mail address: [email protected]
|
1701.08922 | 1 | 1701 | 2017-01-31T05:32:09 | Hardy-Littlewood inequalities on compact quantum groups of Kac type | [
"math.OA"
] | The Hardy-Littlewood inequality on $\mathbb{T}$ compares the $L^p$-norm of a function with a weighted $\ell^p$-norm of its Fourier coefficients. The approach has recently been studied for compact homogeneous spaces and we study a natural analogue in the framework of compact quantum groups. Especially, in the case of the reduced group $C^*$-algebras and free quantum groups, we establish explicit $L^p-\ell^p$ inequalities through inherent information of underlying quantum group, such as growth rate and rapid decay property. Moreover, we show sharpness of the inequalities in a large class, including $C(G)$ with compact Lie group, $C_r^*(G)$ with polynomially growing discrete group and free quantum groups $O_N^+$, $S_N^+$. | math.OA | math |
HARDY-LITTLEWOOD INEQUALITIES ON COMPACT
QUANTUM GROUPS OF KAC TYPE
SANG-GYUN YOUN
Abstract. The Hardy-Littlewood inequality on T compares the Lp-norm of a
function with a weighted ℓp-norm of its Fourier coefficients. The approach has
recently been studied for compact homogeneous spaces and we study a natu-
ral analogue in the framework of compact quantum groups. Especially, in the
case of the reduced group C ∗-algebras and free quantum groups, we establish
explicit Lp
− ℓp inequalities through inherent information of underlying quan-
tum group, such as growth rate and rapid decay property. Moreover, we show
sharpness of the inequalities in a large class, including C(G) with compact
Lie group, C ∗
r (G) with polynomially growing discrete group and free quantum
N , S+
groups O+
N .
1. Introduction
Hardy and Littlewood [14] showed that, for each 1 < p ≤ 2, there exists a
constant Cp such that
(1.1)
1
1
(Xn∈Z
(1 + n)2−p bf (n)p)
p ≤ Cpf ∼Xn∈Zbf (n)znLp(T) for all f ∈ Lp(T).
This implies that the multiplier Fw : Lp(T) → lp(Z), f 7→ (w(n)bf (n))n∈Z, with
is bounded. Moreover, this is a stronger form of the Sobolev
2−p
p
w(n) :=
1
(1 + n)
embedding theorem
1
p − 1
q
H
p
(T) ⊆ Lq(T), ∀1 < p < q < ∞,
where H s
p(T) :=(cid:8)f ∈ Lp(T) : (1 − ∆) s
2 (f ) ∈ Lp(T)(cid:9) is the Bessel potential space.
"The Hardy-Littlewood inequality (1.1)" had been studied for compact abelian
groups by Hewitt and Ross [15], and was recently extended to compact homogeneous
manifolds by Akylzhanov, Nursultanov and Ruzhansky [[1] and [2]]. For compact
Lie groups G with the real dimension n, the result of [2] is written as follows: For
each 1 < p ≤ 2, there exists a constant Cp > 0 such that
(1.2)
p ≤ CpfLp(G) for all f ∈ Lp(G).
Here, bG denotes a maximal family of mutually inequivalent irreducible unitary
satisfies ∆ : πi,j 7→ −κππi,j for all π = (πi,j )1≤i,j≤nπ ∈ bG and all 1 ≤ i, j ≤ nπ.
The above inequality (1.2) can be reduced to a more familiar form whose left
representations of G, AHS := tr(A∗A)
hand side is a natural weighted ℓp-norm of its Fourier coefficients:
2 and the Laplacian operator ∆ on G
bf (π)p
(Xπ∈ bG
(1 + κπ)
2− p
HS)
n(2−p)
n
1
2
2
π
1
1
(1.3)
(Xπ∈ bG
1
(1 + κπ)
n(2−p)
2
nπbf (π)p
Sp
nπ
)
1
p ≤ CpfLp(G).
Key words and phrases. Hardy-Littlewood inequality, Quantum groups, Fourier analysis.
2010 Mathematics Subject Classification. Primary 20G42, Secondary 46L52
The author is supported by TJ Park Science Fellowship.
1
A notable point is that "the Hardy-Littlewood inequalities on compact Lie groups
(1.2)" are determined by inherent geometric information, namely the real dimension
and the natural length function on bG. Indeed, π 7→ √κπ is equivalent to the natural
length · on bG (See Remark 6.1).
The main purpose of this paper is to establish new Hardy-Littlewood inequal-
ities on compact quantum groups of Kac type through geometric information of
underlying quantum groups. As a part of efforts, we will present some explicit
inequalities in important examples. The reduced group C∗-algebras C∗r (G) with
discrete groups G, the free orthogonal quantum groups O+
N and the free permu-
tation quantum groups S+
N are the main targets. Of course, non-commutative Lp
analysis on quantum groups is widely studied from various perspectives ([7], [9],
[16], [17] and [26]). For details of operator algebraic approach to quantum group
itself, see [19], [20], [24] and [30].
To clarify our intention, let us show the main results of this article on "compact
matrix quantum groups" which admit the natural length function · : Irr(G) →
{0} ∪ N (See Definition 3.3 and Proposition 3.4). The following inequalities are
determined by inherent information of underlying quantum group, namely growth
rate and rapid decay property.
Theorem 1.1. Let G be a compact matrix quantum group of Kac type and denote
by · the natural length function on Irr(G).
(1) Let G have a polynomial growth with Xα∈Irr(G):α≤k
n2
α ≤ (1 + k)γ and γ > 0.
Then, for each 1 < p ≤ 2, there exists a universal constant K = K(p) such
that
( Xα∈Irr(G)
≤( Xα∈Irr(G)
for all f ∼ Xα∈Irr(G)
that
1
1
p
2
α
1
)
1
nα
Sp
HS )
2− p
(1 + α)(2−p)γ nαbf (α)p
bf (α)p
(1 + α)(2−p)γ n
nαtr(bf (α)uα) ∈ Lp(G).
fL∞(G) ≤ C(1 + k)βfL2(G)
p ≤ KfLp(G)
(2) Let G have the rapid decay property with universal constants C, β > 0 such
(1.4)
(1.5)
i,j : α = k, 1 ≤ i, j ≤ nα(cid:9)). Define sk := Xα∈Irr(G):α=k
n2
α.
Then, for each 1 < p ≤ 2, there exists a universal constant K = K(p) such
that
for all f ∈ span((cid:8)uα
(Xk≥0 Xα∈Irr(G):α=k
≤(Xk≥0
for all f ∼ Xα∈Irr(G)
1
1
(2−p)
2
s
k
(1 + k)(2−p)(β+1)
1
p
)
Sp
nα
nαbf (α)p
p
HS )
2 )
1
p ≤ KfLp(G)
(1 + k)(2−p)(β+1) ( Xα∈Irr(G):α=k
nαbf (α)2
nαtr(bf (α)uα) ∈ Lp(G).
In particular, it was known that the rapid decay property of FN can be strength-
ened in general holomorphic setting [18]. The improved result is called the strong
Haagerup inequality. Based on this data, we can improve the Hardy-Littlewood
inequality for C∗r (FN ) by focusing our attention on holomorphic forms. Theorem
2
5.3 justifies the claim and it seems appropriate to call the improved one "strong
Hardy-Littlewood inequality".
A natural view of the Hardy-Littlewood inequalities on compact Lie groups is
multiplier
that a properly chosen weight function w : bG → (0,∞) makes the corresponding
Fw : Lp(G) → ℓp(bG) given by f 7→ (w(π)bf (π))π∈ bG
be bounded for each 1 < p ≤ 2. Indeed, newly derived Hardy-Littlewood inequali-
ties on compact quantum groups will give a decay pair (r, s) whose corresponding
multiplier Fwr,s is bounded, where wr,s(α) :=
rα(1 + α)s . Moreover, in Section
6, we will show that there is no slower decay pair such that Fr,s is bounded when G
is one of the followings: C(G) with compact Lie groups, C∗r (G) with polynomially
growing discrete group and free quantum groups O+
N . See Theorem 6.6.
N , S+
1
This approach is quite natural because it is strongly related to Sobolev embedding
properties. Indeed, for the case G = Td, Fw0,s : Lp(Td) → lp(Zd) is bounded if and
only if
ps
2−p ( 1
q
q − 1
r )
H
(Td) ⊆ Lr(Td) ∀1 < q < r < ∞,
where H s
p(Td) is the Bessel potential space.
Lastly, in Section 7, we present some remarks that are by-products of this re-
search. We show that most of free quantum groups do not admit an infinite (central)
sidon set, and also give a Sobolev embedding theorem type interpretation of our
results for C(G) with compact Lie group and C∗r (G) with polynomially growing
discrete group G. Also, we present an explicit inequality on quantum torus Td
θ.
2. Preliminaries
2.1. Compact quantum groups. A compact quantum group G is given by a
unital C∗-algebra A and a unital ∗-homomorphism ∆ : A → A ⊗min A satisfying
(1) (∆ ⊗ id) ◦ ∆ = (id ⊗ ∆) ◦ ∆;
(2) span{∆(a)(b ⊗ 1A) : a, b ∈ A} , span{∆(a)(1A ⊗ b) : a, b ∈ A} are dense in A.
Every compact quantum group admits the unique Haar state h on A such that
(h ⊗ id)(∆(x)) = h(x)1A = (id ⊗ h)(∆(x)) for all x ∈ A.
A finite dimensional corepresentation of G is given by an element u = (ui,j)1≤i,j≤n ∈
Mn(A) such that ∆(ui,j) =
nXk=1
i,j)1≤i,j≤nα(cid:9)α∈Irr(G)
ui,k ⊗ uk,j for all 1 ≤ i, j ≤ n. We say that the
corepresentation u is unitary if u∗u = uu∗ = Idn ⊗ 1A ∈ Mn(A) and irreducible if
{X ∈ Mn : Xu = uX} = C · Idn where Idn is the identity matrix in Mn.
be a maximal family of mutually inequivalent
finite dimensional unitary irreducible corepresentations of G. It is well known that,
for each α ∈ Irr(G), there is a unique positive invertible matrix Qα ∈ Mnα such
that tr(Qα) = tr(Q−1
Let(cid:8)uα = (uα
h((uβ
s,t)∗uα
i,j) =
α ) and
δα,βδj,t(Q−1
tr(Qα)
α )i,s
, ∀α, β ∈ Irr(G), 1 ≤ i, j ≤ nα, 1 ≤ s, t ≤ nβ,
h(uβ
s,t(uα
i,j)∗) =
δα,βδi,s(Qα)j,t
tr(Qα)
, ∀α, β ∈ Irr(G), 1 ≤ i, j ≤ nα, 1 ≤ s, t ≤ nβ.
case, the Haar state h is tracial.
We say that G is of Kac type if Qα = Idnα ∈ Mnα for all α ∈ Irr(G). In this
Lastly, we define Cr(G) as the image of A in the GNS representation of the Haar
state h and L∞(G) := Cr(G)′′. The Haar state h has a normal faithful extension
to L∞(G).
3
2.2. Non-commutative Lp-spaces. Let M be a von Neumann algebra with a
normal faithful tracial state ϕ. Note that the von Neumann algebra M admits
the unique predual M∗. We define L1(M, ϕ) := M∗ and L∞(M, ϕ) := M, then
consider a contractive injection j : M → M∗, given by [j(x)](y) := h(yx) for all
y ∈ M. The map j has dense range.
Now (M,M∗) is a compatible pair of Banach spaces and we can define non-
commutative Lp-space Lp(M, ϕ) := (M,M∗) 1
is
the complex interpolation space. For any x ∈ L∞(M, ϕ), its Lp-norm is given by
xLp(M,ϕ) = ϕ(xp)
In particular, for all 1 ≤ p ≤ ∞, we denote by Lp(G) the non-commutative
Lp-space associated to the von Neumann algebra L∞(G) of a Kac type compact
quantum group G and the tracial Haar state h. Then the space of polynomials
for all 1 < p < ∞, where (·,·) 1
p for all 1 ≤ p < ∞.
p
p
1
i,j : α ∈ Irr(G), 1 ≤ i, j ≤ nα(cid:9))
P ol(G) := span((cid:8)uα
(Aα)α∈Irr(G) ∈ Yα∈Irr(G)
is dense in Cr(G) and Lp(G) for all 1 ≤ p < ∞.
Under the assumption that G is of Kac type, for 1 ≤ p < ∞,
nαtr(Aαp))
(Aα)α∈Irr(G)ℓ∞(bG) := sup
Mnα : Xα∈Irr(G)
(Aα)α∈Irr(G) ∈ Yα∈Irr(G)
ℓp(bG) :=
(Aα)α∈Irr(G)ℓp(bG) := ( Xα∈Irr(G)
ℓ∞(bG) :=
nαtr(Aαp) < ∞
and the natural ℓp-norm of (Aα)α∈Irr(G) ∈ ℓp(bG) is defined by
p = ( Xα∈Irr(G)
α∈Irr(G)Aα < ∞
and the ℓ∞-norm of (Aα)α∈Irr(G) ∈ ℓ∞(bG) is defined by
It is known that ℓ1(bG) = (ℓ∞(bG))∗ and ℓp(bG) = (ℓ∞(bG), ℓ1(bG)) 1
group G, the Fourier transform F : L1(G) → ℓ∞(bG), ψ 7→ bψ, is defined by
from L2(G) onto l2(bG) ([26], Proposition 3.1 and 3.2). Then, by the interpolation
(bG) for each 1 ≤ p ≤ 2, where p′ is the conjugate of p.
It is also known that F is an injective contractive map and it is an isometry
theorem, we induce the Hausdorff-Young inequality again, i.e. F is a contractive
map from Lp(G) into lp′
∞.
2.3. Fourier analysis on compact quantum groups. For a compact quantum
2.4. The reduced group C∗-algebras. The reduced group C∗-algebra C∗r (G),
can be defined for all locally compact groups, but we only consider discrete groups
in this paper since we want to understand it as a compact quantum group.
Definition 2.1. Let G be a discrete group and define λg ∈ B(l2(G)) for each g ∈ G
by
j,i)∗) for all α ∈ Irr(G), 1 ≤ i, j ≤ nα.
(bψ(α))i,j := ψ((uα
α∈Irr(G)Aα.
for all 1 < p <
p
Also,
1
nαAαp
Sp
nα
1
p .
)
Mnα :
sup
[(λg)(f )](x) := f (g−1x) for all x ∈ G.
Then the reduced group C∗-algebra, C∗r (G), is defined as the norm-closure of the
space span({λg : g ∈ G}) in B(l2(G)). Moreover, if we define a comultiplication
∆ : C∗r (G) → C∗r (G) ⊗min C∗r (G) by λg 7→ λg ⊗ λg for all g ∈ G, then (C∗r (G), ∆)
forms a compact quantum group.
4
Note that, for G = (C∗r (G), ∆) of a discrete group G, L∞(G) is nothing but the
group von Neumann algebra V N (G) and Irr(G) = {λg}g∈G is identified with G.
2.5. Free quantum groups of Kac type.
Definition 2.2. (Free orthogonal quantum group [28])
Let N ≥ 2 and A be the universal unital C∗-algebra which is generated by the N 2
self-adjoint elements ui,j with 1 ≤ i, j ≤ N satisfying the relations:
NXk=1
NXk=1
uk,iuk,j =
ui,kuj,k = δi,j for all 1 ≤ i, j ≤ N.
NXk=1
NXk=1
Also, we define a comultiplication ∆ : A → A ⊗min A by ui,j 7→
ui,k ⊗ uk,j.
Then (A, ∆) forms a compact quantum group called the Free orthogonal quantum
group. We denote it by O+
N .
Definition 2.3. (Free permutation quantum group [29])
Let N ≥ 2 and A be the universal unital C∗-algebra which is generated by the N 2
self-adjoint elements ui,j with 1 ≤ i, j ≤ N satisfying the relations:
u2
i,j = ui,j = u∗i,j and
ui,k =
uk,j = 1A for all 1 ≤ i, j ≤ N.
NXk=1
NXk=1
Also, we define a comultiplication ∆ : A → A ⊗min A by ui,j 7→
ui,k ⊗ uk,j.
Then (A, ∆) forms a compact quantum group called the Free permutation quantum
group. We denote it by S+
N .
These free quantum groups are of Kac type, so that the Haar states are tracial
N +2) can be identified with {0} ∪ N.
states. Also, for all N ≥ 2, Irr(O+
Moreover,
N ) and Irr(S+
nk =
k + 1
2r0−N rk+1
r0
0 + N−r0
2r0−N (N − r0)k+1
if G = O+
if G = O+
2 or S+
4
N or SN +2 with N ≥ 3
where r0 is the largest solution of the equation X 2−N X +1 = 0. Note that nk ≈ rk
if N ≥ 3.
2.6. The noncommutative Marcinkiewicz interpolation theorem. The clas-
sical Marcinkiewicz interpolation theorem has a natural non-commutative analogue
for semi-finite von Neumann algebras.
0
Theorem 2.4. (The non-commutative Marcinkiewicz interpolation theorem [32])
Let M be equipped with a normal semifinite faithful trace φ and 1 ≤ p1 < p <
p2 < ∞. Assume that a sub-linear map A : M → L1(N ) satisfies the following:
There exists C1, C2 > 0 such that for any T1 ∈ Lp1(M ), T2 ∈ Lp2(M ) and for any
y > 0,
(2.1)
φ(1(y,∞)(AT1)) ≤ (
C1
y
C2
y
Then A : Lp(M ) → Lp(N ) is a bounded map.
Proof. The proof of [Theorem 1.22, [32]] is still valid under a natural modification.
)p1T1p1
)p2T2p2
φ(1(y,∞)(AT2)) ≤ (
Lp1 (M),
Lp2 (M).
Also, the direct sum K := M ⊕ N and natural extension eA : K → L1(K) gives
another proof.
(cid:3)
5
If the sub-linear operator A satisfies the inequality (2.1), then we say that A is
of weak type (p1, p1). Also, the boundedness of A : Lp(M ) → lp(N ) implies that A
is of weak type (p, p).
Now denote by c(Irr(G), ν) the space of all functions on the discrete space Irr(G)
with a positive measure ν. Then the above theorem is written as follows:
Corollary 2.5. Let G be a Kac type compact quantum group and let 1 ≤ p1 <
p < p2 < ∞. Assume that A : L∞(G) → c(Irr(G), ν) is sub-linear and satisfies the
following: There exists C1, C2 > 0 such that for any T1 ∈ Lp1(G), T2 ∈ Lp2 (G) and
for any y > 0,
Xα: (AT1)(α)≥y
Xα: (AT2)(α)≥y
ν(α) ≤ (
C1
y
)p1T1p1
Lp1 (G),
ν(α) ≤ (
C2
y
)p2T2p2
Lp2 (G).
Then A : Lp(G) → ℓp(Irr(G), ν) is a bounded map.
3. Paley-type inequalities
3.1. General Approach. In this subsection, we derive a Paley-type inequality for
Kac type compact quantum group G via fundamental techniques such as Hausdorff-
Young inequality, Plancherel theorem and the non-commutative Marcinkiewicz in-
terpolation theorem.
We prove the following theorem by adapting techniques used in [2].
Theorem 3.1. Let G be a Kac type compact quantum group and let w : Irr(G) →
< ∞. Then, for each
(0,∞) be a function such that Cw := sup
1 < p ≤ 2, there exists a universal constant K = K(p) > 0 such that
t · Xα: w(α)≥t
n2
α
t>0
bf (α)p
HS )
1
p ≤ KfLp(G)
2− p
2
α
(3.1)
w(α)2−pn
for all f ∼ Xα∈Irr(G)
( Xα∈Irr(G)
nαtr(bf (α)uα) ∈ Lp(G).
L1(G) → c(Irr(G), ν), f 7→ (bf (α)HS
Proof. Put ν(α) := w(α)2n2
√nαw(α)
from Lp(G) into ℓp(Irr(G), ν) for all 1 < p ≤ 2.
First,
α. We will show that the sub-linear operator A :
)α∈Irr(G), is a well-defined bounded map
Xα∈Irr(G)
(Af )(α)2
HS ν(α) = Xα∈Irr(G)
HS = f2
L2(G).
nαbf (α)2
This implies that A is of (strong) type (2, 2) with C2 = 1.
6
y
√nα
Second, for all y > 0, since bf (α)HS
ν(α) ≤ Xα: w(α)≤ f 1
Xα: Af (α)HS≥y
= Xα: w(α)≤ f 1
=Z ( f 1
= 2Z f 1
≤ 2Cw f1
)2
y
0
0
.
y
y
αdx
0
n2
y Z w(α)2
[ Xx
t( Xα: t≤w(α)≤ f 1
2 ≤w(α)≤ f 1
1
y
y
= (
tr(bf (α)∗bf (α))
nα
w(α)2n2
α
)
1
2 ≤ bf (α) ≤ fL1(G),
n2
α]dx by the Fubini theorem
n2
α)dt by substituting x to t2
This says that A is of weak type (1, 1) with C1 = 2Cw.
Now, by Corollary 2.5,
w(α)2−pn
p( 2
α
p − 1
2 )
[ Xα∈Irr(G)
1
p . fLp(G).
HS ]
bf (α)p
(cid:3)
The left hand side of the inequality (3.1) can be reduced to a more familiar form.
Recall that the natural non-commutative ℓp-norm on ℓ∞(bG) = ℓ∞ − ⊕α∈Irr(G)Mnα
is given by
1
p
(Aα)α∈Irr(G)ℓp(bG) = ( Xα∈Irr(G)
nαAαp
Sp
nα
)
under the condition that G is of Kac type.
Corollary 3.2. Let 1 < p ≤ 2 and w be a function satisfying the condition of
Theorem 3.1. Then we have that
( Xα∈Irr(G)
w(α)2−pnαbf (α)p
nαtr(bf (α)uα) ∈ Lp(G).
for all f ∼ Xα∈Irr(G)
Proof. First,
Sp
nα
))
1
p ≤ KfLp(G)
Put
1
r
=
1
p −
1
2
. Then 2 < r ≤ ∞ and
tr(bf (α)p) ≤ bf (α)p
This completes proof easily.
.
Sp
nα
tr(bf (α)p) = bf (α)p
HSIdnαp
Sr
nα
= n
1− p
2
α
HS .
bf (α)p
(cid:3)
Now we discuss an important subclass of compact quantum groups, namely com-
pact matrix quantum groups which allows the natural length function on Irr(G).
Definition 3.3. A compact matrix quantum group is given by a pair (A, ∆, u) with
a unital C∗-algebra A, a ∗-homomorphism ∆ : A → A ⊗min A and a unitary u =
7
(ui,j)1≤i,j≤n ∈ Mn(A) such that (1) ∆ : ui,j 7→
is invertible in Mn(A) and (3) {ui,j}1≤i,j≤n generates A as a C∗-algebra.
ui,k ⊗ uk,j, (2) u = (u∗i,j)1≤i,j≤n
nXk=1
By definition, free orthogonal quantum groups O+
N and free permutation quantum
groups S+
N are compact matrix quantum groups. Also, in the class of compact
quantum groups, the subclass of compact matrix quantum groups is characterized
by the following proposition. The conjugate α ∈ Irr(G) of α ∈ Irr(G) is determined
by uα := Q
α uαQ− 1
α = ((Qα)
i,j)∗(Qα)− 1
i,i(uα
1
2
1
2
2
2
j,j )1≤i,j≤nα .
Proposition 3.4. ([24])
A compact quantum group is a compact matrix quantum group if and only if
there exists a finite set S := {α1,··· , αn} ⊆ Irr(G) such that any α ∈ Irr(G) is
contained in some iterated tensor product of elements α1, α1,··· , αn, αn and the
trivial corepresentation.
Then there is a natural way to define a length function on Irr(G) ([25]). For
non-trivial α ∈ Irr(G), the natural length α is defined by
min{m ∈ {0} ∪ N : ∃β1,··· , βm such that α ⊆ β1 ⊗ ··· ⊗ βm, βj ∈ {αk, αk}n
The length of the trivial corepresentation is defined by 0.
k=1} .
Then we can extract explicit inequalities from Theorem 3.1 and Corollary 3.2 by
inserting geometric information of underlying quantum group, namely growth rate
that is estimated by the quantities bk := Xα≤k
n2
α [6].
Corollary 3.5. Let a Kac type compact matrix quantum group G satisfy
n2
α ≤ C(1 + k)γ for all k ≥ 0 with C, γ > 0
bk = Xα∈Irr(G):α≤k
with respect to the natural length function. Then, for each 1 < p ≤ 2, there exists a
universal constant K = K(p) such that
HS )
1
p ≤ KfLp(G)
1
p
)
nα
1
1
2
α
Sp
2− p
(3.2)
( Xα∈Irr(G)
≤( Xα∈Irr(G)
for all f ∼ Xα∈Irr(G)
t · Xα:α≤t
(1 + α)(2−p)γ nαbf (α)p
bf (α)p
(1 + α)(2−p)γ n
nαtr(bf (α)uα) ∈ Lp(G).
α
Proof. Consider a weight function w(α) :=
t · Xα:α≤t
= sup
0<t≤1
sup
t>0
γ −1
γ −1
n2
− 1
− 1
1
(1 + α)γ . Then
n2
α
≤ C sup
0<t≤1
t · (t− 1
γ )γ = C.
(cid:3)
Now the conclusion is obtained by Theorem 3.1 and Proposition 3.2.
3.2. A paley-type inequality under the rapid decay property. In this sub-
section, we still assume that G is a compact matrix quantum group of Kac type.
One of the main observations of this paper is that the more detailed geometric infor-
mation actually improves Theorem 3.1 and Corollary 3.2 in various "exponentially
growing" cases. A more refined paley-type inequality can be obtained under the
condition that G has the rapid decay property in the sense of [25].
8
Definition 3.6. ([25])
Let G be a Kac type compact matrix quantum group. Then we say that G has the
rapid decay property with respect to the natural length function on Irr(G) if there
exist C, β > 0 such that
(3.3)
Xα∈Irr(G):α=k
nαXi,j=1
aα
i,juα
i,jL∞(G) ≤ C(1 + k)β Xα∈Irr(G):α=k
nαXi,j=1
aα
i,juα
i,jL2(G)
for any k ≥ 0 and scalars aα
i,j ∈ C.
Notation 1.
(1) When the natural length function on Irr(G) is given, we will
(2) We denote by pk the orthogonal projection from L2(G) to the clousre of
use the notations Sk := {α ∈ Irr(G) : α = k} and sk :=Pα∈Sk
span((cid:8)uα
i,j : α ∈ Sk, 1 ≤ i, j ≤ nα(cid:9)).
Proposition 3.7. Suppose that a Kac type compact matrix quantum group G has
the rapid decay property with respect to the natural length function on Irr(G) and
with inequality (3.3). Then we have that
n2
α.
(3.4)
sup
k≥0
(Pα∈Irr(G):α=k nαbf (α)2
(k + 1)β
HS )
1
2
≤ CfL1(G) for all f ∈ L1(G).
Proof. Since L1(G) is isometrically embedded into the dual space M (G) := Cr(G)∗
and P ol(G) is dense in Cr(G), we have that
fL1(G) =
=
sup
=
≥
(3.5)
sup
x∈P ol(G):xL∞(G)≤1hf, xiL1(G),L∞(G)
x∈P ol(G):xL∞(G)≤1hf, x∗iL1(G),L∞(G)
x∈P ol(G):xL∞(G)≤1 Xα∈Irr(G)
sup
sup
nαtr(bf (α)bx(α)∗)
2 ≤1 Xα∈Irr(G):α=k
HS)
1
1
x∈P ol(G):Pk≥0 C(k+1)β (Pα:α=k nαbx(α)2
≥ sup
k≥0
= sup
k≥0
sup
(Pα:α=k nαbx(α)2
HS)
(Pα∈Irr(G):α=k nαbf (α)2
C(k + 1)β
HS )
1
2
.
2 ≤1 Xα∈Irr(G)
nαtr(bf (α)bx(α)∗)
C(k + 1)β tr(bf (α)bx(α)∗)
nα
(cid:3)
Theorem 3.8. Let a Kac type compact matrix quantum group G have the rapid de-
cay property with respect to the natural length function on Irr(G) and with inequality
(3.3). Also, suppose that a weight function w : {0} ∪ N → (0,∞) satisfies
(3.6)
(Xk≥0
for all f ∼ Xα∈Irr(G)
(k + 1)2β
< ∞.
w(k) ≤ 1
y
Cw := sup
y>0
y · Xk≥0: (k+1)β
w(k)2−p( Xα∈Irr(G):α=k
nαtr(bf (α)uα) ∈ Lp(G).
9
HS )
nαbf (α)2
Then, for each 1 < p ≤ 2, there exists a universal constant K = K(p) > 0 such that
(3.7)
2 )
p
1
p ≤ KfLp(G)
Proof. Put ν(k) := w(k)2. We will show that the sub-linear operator A : L1(G) →
c({0} ∪ N, ν), f 7→ (pk(f )L2(G)
)k≥0, is a well-defined bounded map from Lp(G)
into lp({0} ∪ N, ν) for all 1 < p ≤ 2.
w(k)
Firstly,
(pk(f )L2(G)
w(k)
Xk≥0
)2ν(k) =Xk≥0
pk(f )2
L2(G) = f2
L2(G).
Therefore, A is of (weak) type (2, 2) with C2 = 1.
Secondly, for all y > 0,
Xk≥0:(Af )(k)>y
ν(k) ≤
Xk: w(k)
(k+1)β <
Cf
L1 (G)
y
w(k)2 by Proposition 3.7.
w(k)
(k + 1)β . Then
Now put ew(k) :=
=
Cf
Xk: ew(k)<
≤Z (
= 2Z Cf
0
0
Cf
L1 (G)
y
L1(G)
y
)2
L1 (G)
y
0
Z ew(k)2
Xk:√x≤ ew(k)
t · Xk:t≤ ew(k)
2CwCfL1(G)
.
y
≤
(k + 1)2βdx
(k + 1)2βdx
(k + 1)2βdt by substituting x = t2
Therefore, by Corollary 2.5, we obtain that
(Xk≥0
w(k)2−ppk(f )p
L2(G))
1
p . fLp(G).
(cid:3)
Corollary 3.9. Let a Kac type compact matrix quantum group G have the rapid de-
cay property with respect to the natural length function on Irr(G) and with inequality
(3.3). Then, for each 1 < p ≤ 2, there exists a universal constant K = K(p) > 0
such that
1
(3.8)
(Xk≥0
for all f ∼ Xα∈Irr(G)
(1 + k)(2−p)(β+1) ( Xα∈Irr(G):α=k
nαtr(bf (α)uα) ∈ Lp(G).
10
HS )
nαbf (α)2
p
2 )
1
p ≤ KfLp(G)
Proof. Take w(k) :=
(1 + k)β+1 and E = Irr(G). Then
1
Cw = sup
y
y>0
(1 + k)2β
y · Xk≥0:(1+k)2β+1≤ 1
0<y≤1
t2βdt
y ·Z ( 1
)
0<y≤1(y ·
(2 · ( 1
y )
2β + 1
2β+1 )2β+1
2β+1 +1
y )
1
1
1
≤ sup
≤ sup
2β+1
2β + 1
=
< ∞.
(cid:3)
Corollary 3.10. Let a Kac type compact matrix quantum group G have the rapid
decay property with respect to the natural length function on Irr(G) and with inequal-
ity (3.3). Then, for each 1 < p ≤ 2, there exists a universal constant K = K(p) > 0
such that
(3.9)
(Xk≥0 Xα∈Irr(G):α=k
for all f ∼ Xα∈Irr(G)
1
(1 + α)(2−p)(β+1)(Pβ∈Sk
nαtr(bf (α)uα) ∈ Lp(G).
Proof. Since
=
+
1
p
1
2
2 − p
2p
and n− 1
Xα∈Sk
Sp
nαbf (α)p
nα ≤ Xα∈Sk
2−p
2
n2
β)
nαbf (α)p
Sp
nα
)
1
p ≤ KfLp(G)
nα ≤ n− 1
2
2
2
p
α
1
2
n
HS
2− p
α bf (α)Sp
bf (α)p
αbf (α)HS · n
αbf (α)HS )α∈Skp
· (Xα∈Sk
p −1
α
n2
α)
2−p
1
2
2
= (n
≤ (n
= (Xα∈Sk
α bf (α)HS , we have that
ℓp(Sk)
)α∈Skp
ℓ2(Sk) · (Xα∈Sk
nαbf (α)2
HS )
2−p
2
n2
α)
p
2 .
(cid:3)
Then we obtain the conclusion.
4. Hardy-Littlewood inequalities
This section is devoted to establish explicit Hardy-Littlewood inequalities for the
main targets: the reduced group C∗-algebras C∗r (G) with finitely generated discrete
group G, free orthogonal quantum groups O+
N and free permutation quantum groups
S+
N .
4.1. The reduced group C∗-algebras C∗r (G). In this subsection, we treat finitely
generated discrete groups G. As expected, we found clear evidence that the geo-
metric information of the underlying group is of significant importance for under-
standing non-commutative Lp-spaces Lp(V N (G)).
Definition 4.1. A discrete group with a fixed finite symmetric generating set S is
said to be polynomially growing if there exist C > 0 and k > 0 such that
#{g ∈ G : g ≤ n} ≤ C(1 + n)k for all n ≥ 0.
11
In this case, the polynomial growth rate k0 is defined as the minimum of such k.
Then k0 becomes a natural number and is independent of the choice of generating
set S.
Theorem 4.2.
(1) Let G be a finitely generated discrete group, which has the
polynomial growth rate k0. Then, for each 1 < p ≤ 2, there exists a universal
constant K = K(p) such that
(4.1)
(Xg∈G
1
(1 + g)(2−p)k0 f (g)p)
1
p ≤ Kλ(f )Lp(V N (G))
for all λ(f ) ∼Xg∈G
f (g)λg ∈ Lp(V N (G)).
(2) Let G be a finitely generated discrete group with
bk = #{g ∈ G : gS ≤ k} ≤ Crk for all k ≥ 0,
where · S is the natural length function with respect to a finite symmetric
generating set S. Then, for each 1 < p ≤ 2, there exists a universal constant
K = K(p, S) > 0 such that
(4.2)
(Xg∈G
for all λ(f ) ∼Xg∈G
1
r(2−p)g f (g)p)
1
p ≤ Kλ(f )Lp(V N (G))
f (g)λg ∈ Lp(V N (G)).
Proof. (1) Clear from Corollary 3.5.
(2) Consider w(g) :=
. Then
1
rg
sup
t>0
t · Xg≤logr ( 1
t )
1
= sup
0<t≤1
t · Xg≤logr( 1
t )
1 ≤ C.
(cid:3)
Then the conclusion follows from Theorem 3.1
Remark 4.3.
(1) For every finitely generated discrete group, there exist C, r >
0 such that bk ≤ Crk for all k ≥ 0 by the Fekete's subadditivity lemma.
Therefore, Theorem 4.2 covers all finitely generated discrete group.
(2) In fact, the inequality (4.1) is sharp by Theorem 6.6.
Although we can always find inequality (4.2) for every finitely generated discrete
group, we can get a much better result by adding more detailed geometric infor-
mation of underlying group. Indeed, if we assume hyperbolicity of group, then the
inequality is considerably improved.
Theorem 4.4. Let G be any non-elementary word hyperbolic group with bk ≤ Crk
for all k ≥ 0 with respect to a finite symmetric generating set S. Then, for each
1 < p ≤ 2, there exists a universal constant K = K(S, p) such that
(4.3)
2 )
1
1
p
p
(Xg∈G
(2−p)g
2
r
(1 + g)4−2p f (g)p)
1
p ≤ (Xk≥0
1
(k + 1)4−2p ( Xg∈G:g=k
≤ Kλ(f )Lp(V N (G)).
f (g)2)
for all λ(f ) ∼Xg∈G
f (g)λg ∈ Lp(V N (G)).
Proof. The conclusion follows from Corollary 3.10 and [13].
(cid:3)
12
4.2. Free quantum groups. Let us begin the investigation of 'genuine' quantum
examples: Free orthogonal quantum groups O+
N and free permutation quantum
groups S+
N becomes an equiv-
alence under centrality, positivity, monotonic decrease and a non-oscillation type
condition of Fourier coefficients, as for the result for SU (2) [Theorem 2.10 [2]]. This
is considered as a way to prove sharpness of Hardy-Littlewood inequalities.
N +2. Moreover, the Hardy-Littlewood inequality for O+
Theorem 4.5.
permutation quantum group S+
universal constant K = K(p) such that
(1) Let G be the free orthogonal quantum group O+
2 or the free
4 . Then, for each 1 < p ≤ 2, there exists a
(4.4)
(Xk≥0
1
(1 + k)6−3p nkbf (k)p
for all f ∼Xk≥0
nktr(bf (k)uk) ∈ Lp(G).
Sp
k+1
)
1
p ≤ (Xk≥0
1
(1 + k)6−3p n
2− p
2
k
1
p ≤ KfLp(G)
HS )
bf (k)p
(2) Let G be a free orthogonal quantum group O+
N or a free permutation quantum
N +2 with N ≥ 3. Then, for each 1 < p ≤ 2, there exists a universal
group S+
constant K = K(p) such that
(4.5)
(Xk≥0
1
r(2−p)k
0
(1 + k)4−2p
)
nkbf (k)p
Sp
nk
1
p ≤ (Xk≥0
r(2−p)k
0
≤ KfLp(G)
1
(1 + k)4−2p
n
2− p
2
k
1
p
HS)
bf (k)p
for all f ∼Xk≥0
nktr(bf (k)uk) ∈ Lp(G), where r0 =
N + √N 2 − 4
2
.
Proof. (1) In this case, nk = k + 1 for all k, so that the conclusion follows from
Corollary 3.5.
(2) It is known that free orthogonal quantum groups O+
N and free quantum groups
SN +2 with N ≥ 3 have the rapid decay property with β = 1 ([25] and [5]). Also,
sk = n2
for all k ∈ {0} ∪ N. Therefore, Corollaries 3.9 and 3.10 complete
proof.
(cid:3)
k ≈ r2k
0
Remark 4.6. All results of this paper for S+
N can be extended to quantum auto-
morphism group Gaut(B, ψ) with a δ-trace ψ and dim(B) = N via the same proofs.
An important observation for the free orthogonal quantum groups O+
N is that the
inequalities (4.4) and (4.5) become actually equivalences in a large class. Essentially,
this fact is based on the result for SU(2) [Theorem 2.10, [2]] and the following lemma
moves the result to O+
N .
Lemma 4.7. Let G = O+
for each cases. Then, for f ∼Xn≥0
N or S+
cnχ1
n ∈ Lp(G),
N +2 with N ≥ 2 and consider G = SU (2) or SO(3)
fLp(G) = Φ(f ) ∼Xn≥0
cnχ2
nLp(G) for all 1 ≤ p ≤ ∞.
Here, χ1
representations of G and G respectively.
n = tr(un) and χ2
n = tr(vn) where un and vn are the n-th irreducible unitary
13
Proof. In the above cases, it is known that G and G share the fusion rule. For
k ∈ P ol(G) and m ∈ N,
details, see [Proposition 6.7, [26]]. Now, for any x =Xk≥0
ckχ1
h((x∗x)m) = h((Xk,l≥0
ckclχ1
kχ1
l )m)
= Xk1,l1,··· ,km,lm≥0
= Xk1,l1,··· ,km,lm≥0
=ZG
(Xk,l≥0
=ZG x′2m,
ckclχ2
kχ2
l )m
ck1 ··· ckmcl1 ··· clmh(χ1
ck1 ··· ckmcl1 ··· clmZG
k1 χ1
l1 ··· χ1
km χ1
lm)
k1 χ2
χ2
l1 ··· χ2
kmχ2
lm
where x′ =Xk≥0
proof.
ckχ2
k ∈ P ol(G). Then the Stone-Weistrass theorem completes this
(cid:3)
3
2
< p ≤ 2 and fix D > 0. Also, assume that f ∼
Corollary 4.8. Let N ≥ 2,
N ) satisfies
2 (O+
3
ckχk ∈ L
Xk≥0
(4.6)
Then we have
(4.7)
ck ≥ ck+1 ≥ 0 and Xm≥k
N ) ≈ (Xk≥0
fLp(O+
cm
m + 1 ≤ D · ck for all k ≥ 0.
(1 + k)2p−4cp
k)
1
p .
5. A strong Hardy-Littlewood inequality
The studies of Hardy-Littlewood inequalities in [2], [14] and [15] deal with general
Lp-functions, but a plenty of classical results of harmonic analysis on T shows
that a theorem on a function space can have a stronger form when restricted to
"holomorphic" setting [18].
An evidnce on non-commutative setting is "the strong Haagerup inequality" on
the reduced group C∗-algebras C∗r (FN ). More precisely, it was shown that the rapid
decay property can be strengthen in general holomorphic setting [18].
Let g1,··· , gN be canonical generators of FN and denote by F+
N the the set
of elements of the form gi1gi2 ··· gim with m ∈ {0} ∪ N and 1 ≤ ik ≤ N for all
1 ≤ k ≤ m.
Theorem 5.1. (Strong Haagerup inequality for C∗r (FN ))
Consider a subset E := F+
N and Ek := {g ∈ E : g = k}. Then, for any k ∈
{0} ∪ N, we have that
Xg∈Ek
f (g)λgC ∗
r (FN ) ≤ √e√k + 1(Xg∈Ek
f (g)2)
1
2 .
Based on this information, we can modify the inequality (3.4) as follows.
Proposition 5.2. Let N ≥ 2. Then we have that
(5.1)
(Pg∈Ek f (g)2)
(k + 1) 1
fA(FN ) ≥
1
√e
sup
k≥0
2
1
2
14
for all f ∈ A(FN ).
Proof. We can repeat the proof of Proposition 3.7. The only difference is the im-
provement of (1 + k)β to (1 + k)
2 in inequality (3.5). Then we are able to get
conclusion by restricting support of x ∈ Cc(G) to F+
(cid:3)
Theorem 5.3. Let N ≥ 2. Then, for each 1 < p ≤ 2, there exists a universal
constant K = K(p) > 0 such that
N in the proof.
1
(5.2)
1
3
2
p
1
p
1
2 )
(2−p)g
2 (2−p)N
(1 + g)
f (g)p)
f (g)2)
1
(1 + k) 3
p ≤ (Xk≥0
( Xg∈FN :g=k
2 (2−p)
≤ Kλ(f )Lp(V N (FN ))
f (g)λg ∈ Lp(V N (FN )) with supp(f ) ⊆ F+
N .
(Xg∈FN
for all λ(f ) ∼ Xg∈FN
3.10. The only difference is to replace the operator A with λ(f ) 7→ (f · χEkl2(G))
2 . Also, we choose a weight function w on {0}∪N by w(k) :=
. Then we can derive new inequality for general λ(f ) ∈ Lp(V N (FN )), but
(cid:3)
(1 + k) 3
our consideration is in the case supp(f ) ⊆ F+
N .
Proof. It can be also obtained by repeating the proof of Theorem 3.8 and Corollary
and (1+k)β with (1+k)
w(k)2
1
1
2
)k≥0
The studied Hardy-Littlewood inequalities give a decay pair (r, s) such that the
multiplier
6. Sharpness
is bounded for each cases, where wr,s(α) =
Fwr,s : Lp(G) → ℓp(bG), f 7→ (wr,s(α)bf (α))α∈Irr(G),
1
rα(1 + α)s with respect to the natural
n(2 − p)
) for C(G) with compact Lie
p
length · on Irr(G). Here is the list: (0,
group G, (0,
3(2 − p)
p
) for O+
2 or S+
k0(2 − p)
2−p
,
p
) for C∗r (G) with polynomially growing discrete group G,
p
4 and (r
0
(0,
Remark 6.1. ([21] and [27]) If G is a compact Lie group, then √κπ is equivalent
to the natural length function · S generated by the fundamental generating set S
N or SN +2 with N ≥ 3.
p
β
2(2 − p)
) for O+
of bG. Equivalently, (1 + κπ)
2 ≈ (1 + πS)β.
To assert that the established inequalities are sharp, we will show that there is
no slower decay pair (r, s) such that Fwr,s is bounded for the above cases.
This viewpoint is different from the spirit of [Theorem 2.10 ,[2]] or Theorem 4.8,
which requires finding an equivalence on a subclass. However, our approach is quite
natural since it is strongly related to Sobolev embedding theorem. For example,
Fw0,s : Lp(Td) → lp(Zd) is bounded if and only if H s
(Td) if and only if
H
p is the Bessel potential
space. In this direction, we will provide a Sobolev embedding type interpretation
for results of this section in subsection 7.2.
(Td) ⊆ Lr(Td) for all 1 < q < r < ∞, where H s
p (Td) ⊆ Lp′
2−p ( 1
q
q − 1
r )
ps
In addition, this view has a definite advantage over looking for equivalence be-
cause we can cover much larger class.
Our first strategy is handling an ultracontractivity problem on C(G) with com-
pact Lie groups, C∗r (G) with polynomially growing discrete group. Actually ultra-
contractivity problem is strongly related to Sobolev embedding property [31].
15
Let M be the von Neumann subalgebra generated by {χα}α∈Irr(G) in L∞(G) and
consider Lp(M ) as the non-commutative Lp-space associated to the restriction of
the Haar state on M . Now suppose that l : Irr(G) → (0,∞) is a positive function
and there exist 1 < p < 2 and a universal constant C > 0 such that
(6.1)
J(f ) ∼ Xα∈Irr(G)
1
β
p
l(α)
cαχαLp′ (M) ≤ Cf ∼ Xα∈Irr(G)
cαχαLp(M),
where J is a densely defined positive operator on L2(M ) which maps χα 7→
1
for all α ∈ Irr(G), 1 ≤ i, j ≤ nα. Indeed, J = K∗K where K : χα 7→
χα.
β
2p
l(α)
Now take φ(t) := t
2β
2−p , ψ(z) := z
2β
2−p and L := J− p
2β . Then [Theorem 1.1, [31]]
says that there exists a universal constant C′ > 0 such that
1
β
p
l(α)
χα
(6.2)
e−tL(f ) ∼ Xα∈Irr(G)
for all 0 < t < ∞ and all f ∼ Xα∈Irr(G)
cα
etl(α)
χαL∞(M) ≤ C′ fL1(M)
2−p
2β
t
.
1
2
cαχα ∈ L1(M ).
Our claim is that we get the following observations in the above situation:
(6.3)
sup
0<t<∞
t
2β
2−p Xα∈Irr(G)
n2
α
etl(α)
=: C < ∞.
1
2
if G is given by C(G) with compact Lie group or C∗r (G) with polynomially growing
discrete group.
Lemma 6.2. ([Lemma 4.1, [8]] and [Proposition 5.7, [21]])
n2
π
(1 + κπ) s
2
< ∞
Lemma 6.3.
(1) Let G be a compact Lie group. Then there exist probability
(2) Let G be a finitely generated discrete group with polynomial growth rate k0.
1
if and only if s > n.
(1 + g)s < ∞ if and only if s > k0.
(1) Let G be a compact Lie group with dimension n. Then Xπ∈ bG
Then Xg∈G
measures {νt}t>0 such that bνt(π) =
(2) Let G be a compact Lie group and let f ∼ Xπ∈ bG
that bf (π) ≥ 0 for all π ∈ bG. Then
fL∞(G) = Xπ∈ bG
nπtr(bf (π)).
{νt}t>0 ⊆ L1(G).
1
etκπ
Idnπ for all π ∈ bG. Moreover,
nπtr(bf (π)π) ∈ L∞(G) such
(3) Let G be a discrete group. Then A(G) has a bounded approximate identity
if and only if G is amenable. In this case, the bounded approximate identity
can be chosen as positive and compactly supported functions on G.
(4) If G is an amenable discrete group, we have that
λ(f ) ∼Xg∈G
f (g)λgV N (G) =Xg∈G
f (g)
for any positive function f ∈ l1(G).
16
Proof. (1) Since Xπ∈ bG
n2
π
etκπ
< ∞ by Lemma 6.2, we know that νt ∈ A(G) ⊆ C(G) ⊆
L1(G). The family {νt}t>0 is called the Heat semigroup of measures.
convolution product, we have
(2) Since f 7→ µt ∗ f is a contractive map on L∞(G) for all t > 0 where ∗ is the
fL∞(G) ≥ sup
t>0 ft ∼ Xπ∈ bG
tr(bf (π)) ≤ fL1(G)Xπ
t>0Xπ∈ bG
ft(1) = sup
t>0
nπ
etκπ
Here, sinceXπ
nπ
etκπ
fL∞(G) ≥ sup
The other direction is trivial.
series of ft uniformly converges to ft ∈ C(G). Therefore,
tr(bf (π)) = Xπ∈ bG
nπtr(bf (π)).
nπ
etκπ
n2
π
etκπ
tr(bf (π)π)C(G)
< ∞ by Lemma 6.2, the Fourier
(3) See [Theorem 7.1.3, [23]] and its proof. We also may assume the compact
supportness by considering fǫ := f · χ{g∈G:f (g)>ǫ} for positive f ∈ l1(G).
(4) This is the Kesten's condition that is equivalent to amenability.
(cid:3)
Now we can show that the claim is true.
Proposition 6.4. Let G be C(G) with compact Lie group or C∗r (G) with polyomially
growing discrete group. Also, suppose that the inequality (6.1) holds. Then
(6.4)
Proof. (1) By Lemma 6.3,
=: C < ∞.
t
sup
0<t<∞
Xπ∈ bG
n2
π
etl(π)
1
2
2β
n2
α
etl(α)
2−p Xα∈Irr(G)
r>0Xπ∈ bG
2 erκπ
r>0 e−tL(νr)L∞(G)
C′
= sup
= sup
etl(π)
n2
π
1
1
2
≤
2β
2−p
t
sup
r>0 νrL1(G) ≤
C′
2β
2−p
t
for all 0 < t < ∞.
(2) There exists a bounded approximate identity (ei)i in A(G) that consists of
positive and compactly supported functions since polynomially growing discrete
group is always amenable. Then inequality (6.2) says that
Xg∈G
1
1
2
etl(g)
= sup
i Xg∈G
ei(g)
1
2
etl(g)
= sup
i
Xg∈G
ei(g)
1
2
etl(g)
λgC ∗
r (G) ≤
C′′
2β
2−p
t
since lim
i
ei(g) = 1 for all g ∈ G.
(cid:3)
Proposition 6.4 allows us to extract an interesting quantitive observation.
Proposition 6.5. Let G be C(G) with compact Lie group or C∗r (G) with polyomially
growing discrete group. Also, suppose that the inequality (6.1) holds. Then we have
that
(6.5)
Xα∈Irr(G)
n2
α
l(α) m
2
< ∞ for all natural numbers m >
2β
2 − p
.
17
1
=: C0 < ∞
t
1
2
sup
1
xγ dx
n2
α
etl(α)
n2
α
exl(α)
from (6.4), so that
for all 0 < t ≤ 1.
This implies that
Proof. Choose γ ∈ (maxn 2β
2−p , m − 1o , m). Then we have
0<t≤1
tγ Xα∈Irr(G)
Z 1
t Xα∈Irr(G)
0<t≤1
0<t≤1
tγ−(m−1) Xα∈Irr(G)
Xα∈Irr(G)
2
dx ≤Z 1
2
2
tγ−1 Xα∈Irr(G)
Then there exist D1, D2 > 0 such that
so that we can inductively see that
l(α) m
2 etl(α)
l(α)
2 etl(α)
2 etl(α)
l(α)
n2
α
1
n2
α
m−1
sup
sup
n2
α
1
1
=: C1 < ∞,
2 ≤ D1tm−γ + D2 for all 0 < t ≤ 1
1
=: Cm−1 < ∞.
via the same way.
Lastly, taking the limit t → 0 completes this proof.
(cid:3)
Theorem 6.6. Let 1 < p ≤ 2.
(1) Let G be a compact Lie group with dimension n. Then
(Xπ∈ bG
1
(1 + κπ) s
2
nπbf (π)p
Sp
nπ
)
1
p . fLp(G)
(2) Let G be a finitely generated discrete group with polynomial growth rate k0.
holds if and only if s ≥ n(2 − p).
Then
1
(Xg∈G
(1 + g)sf (g)p)
holds if and only if s ≥ k0(2 − p).
(3) Let G be O+
2 or S+
4 . Then
1
p . λ(f )Lp(V N (G))
(Xk≥0
1
(1 + k)s nkbf (k)p
Sp
k+1
)
1
p . fLp(G)
holds if and only if s ≥ 3(2 − p).
(4) Let G be O+
N or S+
N +2 with N ≥ 3. Then
(Xk≥0
1
r(2−p)k
0
(1 + k)s
nkbf (k)p
Sp
nk
holds if and only if s ≥ 4 − 2p, where r0 =
18
1
)
p . fLp(G)
N + √N 2 − 4
.
2
Proof. "If" parts are obtained from (1.3), (4.1), (4.4) and (4.5). To prove the
2 and (2) 1 + g respectively.
converse direction, firstly, define l(α) by (1) (1 + κπ)
Then the assumed inequality
1
( Xα∈Irr(G)
nα
l(α)s bf (α)p
Sp
nα
)
1
p . fLp(G)
implies the inequality (6.1) for β = s. Then, by Proposition 6.5 and Lemma 6.2,
we get that (1) 2n ≤
, (2) 2k0 ≤
respectively.
In (3) and (4), consider G as SU (2) if G = O+
N +2 for each
cases. Also, denote by χ′k the character corresponding to χk for each cases. Define
l(k) := 1 + k.
N and SO(3) if G = S+
2β
2 − p
2β
2 − p
Firstly, in (3), for each f ∼Xk≥0
ckχk ∈ Lp(G),
Xk≥0
(1 + k)− s
p ckχkLp′ (G) ≤ (Xk≥0
1
1
p
(1 + k)s+p−2 ckp)
. fLp(G) = f′ ∼Xk≥0
ckχ′kLp(G).
by Proposition 4.7 and Hausdorff-Young inequality. On the other hand, in (4), for
each f ∼Xk≥0
Xk≥0
ckχk ∈ Lp(G),
(1 + k)− s+2−p
p
ckχkLp′ (G) ≤ (Xk≥0
1
1
p
(1 + k)sckp)
. fLp(G) = f′ ∼Xk≥0
ckχ′kLp(G)
by the similar way.
Now we can apply Proposition 6.5 and Lemma 6.2 for compact Lie groups again,
(cid:3)
so that (3) s ≥ 6 − 3p and (4) s − p + 2 ≥ 6 − 3p(⇔ s ≥ 4 − 2p) respectively.
7. Some remarks about Sidon sets, Sobolev embedding theorem and
quantum torus
As by-products of this study, we refer to an interesting lacunarity result for com-
pact quantum groups, and present a sobolev embedding theorem type interpretation
for C(G) with compact Lie group and for C∗r (G) with polynomially growing group.
Also, we show an explicit inequality on quantum torus Td
θ.
7.1. Sidon set on compact quantum groups. The study of Lacunarity, espe-
cially on Sidon sets, is one of the major subject in harmonic analysis, and recently
the notion has been extended to the setting of compact quantum groups [26].
Definition 7.1. Let G be a compact quantum group.
(1) A subset E ⊆ Irr(G) is called a Sidon set if there exists K > 0 such that
(2) A subset E ⊆ Irr(G) is called a central Sidon set if there exists K > 0 such
bfℓ1(bG) ≤ KfL∞(G) for all f ∈ P olE(G),
where P olE(G) :=nf ∈ P ol(G) : bf (α) = 0 for all α /∈ Eo.
bfℓ1(bG) ≤ KfL∞(G) for all f = Xα∈Irr(G)
that
19
cαχα ∈ P olE(G).
Let G = (A, ∆) be of Kac type and E ⊆ Irr(G) be a central sidon set. Then
[Proposition 6.4, [26]] implies that there exists µ ∈ M (G) = Cr(G)∗ such that
j,i)∗))1≤i,j≤nα = Idnα for all α ∈ E. Since P ol(G) is dense in Cr(G),
Proposition 3.7 still holds for µ ∈ M (G).
Now, if G satisfies the assumptions of Proposition 3.7 and if E ⊆ Irr(G) is a
central sidon set, then we get
bµ(α) = (µ((uα
1
2
(Pα∈Ek
n2
α)
(1 + k)β
2 minα∈Ek nα
(1 + k)β
Ek
1
,
∞ > sup
k≥0
≥ sup
k≥0
where Ek := {α ∈ E : α = k}.
Irr(G) with r > 1.
Therefore, it can not happen simultaneously that E = ∞ and nα > rα ∀α ∈
Remark 7.2.
(1) The above argument shows that there is no an infinite (cen-
tral) Sidon set in U +
N with N ≥ 3, which are not explained in [26].
(2) Shortly after this research, the author of [26] personally informed me another
2 , the
N cases. Under the identification Irr(U +
N )∼= F+
simple idea to explain U +
fact that
χα4 = (1 + α)
1
4 ∀α ∈ F+
2
implies that there is no an infinite Λ(4) set, so that there is no an infinite
Sidon set on U +
N with N ≥ 2.
7.2. Sobolev embedding properties. The contents of Section 6 can be inter-
preted in terms of Sobolev embeddings properties by [Theorem 1.1, [31]].
For C(G) with compact Lie group whose real dimension is n, the computations
in Section 6 says that
(1 − ∆)− β
2 (f ) ∼ Xπ∈ bG
nπ
(1 + κπ)
β
2
tr(bf (π)π)Lp′ (G) . fLp(G)
if and only if β ≥
n(2 − p)
p
for each 1 < p ≤ 2. Moreover, it is equivalent to that
2 ( 1
p − 1
q )(f )Lq(G) . fLp(G)
if and only if β ≥ n for each 1 < p < q < ∞. If we define the space H s
the above result is interpreted as
(1 − ∆)− β
2 (f ) ∈ Lp(G)(cid:9) as an analogue of Bessel potential space, then
p(G) ⊆ Lq(G) if and only if s ≥ n(
H s
(cid:8)f ∈ Lp(G) : (1 − ∆) s
p(G) :=
1
p −
(7.1)
1
q
)
for each 1 < p < q < ∞.
On the other hand, if G is a finitely generated discrete group with polynomial
growth rate k0, then we define infinitesimal generator L on C∗r (G) by λg 7→ −gλg
for all g ∈ G. Then we can derive the Sobolev embedding property of non-
commutative spaces Lp(V N (G)) as follows.
p − 1
q )(λ(f ))Lq (V N (G)) . λ(f )Lp(V N (G)) if and only if β ≥ k0
(7.2) (1 − L)−β( 1
for each 1 < p < q < ∞.
: λg 7→
−g2λg, but it does not make an essential difference in replacing (1 − L) with
(1 − L′)
The reader may consider another natural infinitesimal generator L′
2 .
1
20
7.3. Hardy-Littlewood inequality on Quantum torus. Quantum torus Td
θ is
a widely studied example of "quantum space". In this case, we can establish Hardy-
θ, which is the same form as for Td. A proof can be given
Littlewood inequality for Td
by repeating the proof of Theorem 3.1.
Remark 7.3. For a quantum torus Td
θ, for each 1 < p ≤ 2, we have that
(7.3)
(Xm∈Zd
1
(1 + m1)d(2−p) bx(m)p)
1
p . xLp(Td
θ ).
Acknowledgement. The author is grateful to Hun Hee Lee for helpful com-
ments on this work. Also, he would like to thank Simeng Wang for discussing the
Sidon set of compact quantum groups.
References
[1] Akylzhanov, R., Nursultanov, E., Ruzhansky, M. (2016). Hardy-Littlewood inequalities and
Fourier multipliers on SU (2). Studia Math., 234 (2016), 1-29.
[2] Akylzhanov, R., Nursultanov, E., Ruzhansky, M. (2015). Hardy-Littlewood, Hausdorff-Young-
Paley inequalities, and Lp-Lq Fourier multipliers on compact homogeneous manifolds. arXiv
preprint arXiv:1504.07043.
[3] Brannan, M. (2012). Quantum symmetries and strong Haagerup inequalities. Communications
in Mathematical Physics, 311(1), 21-53.
[4] Brannan, M. P. (2012). On the Reduced Operator Algebras of Free Quantum Groups, Ph.D
Thesis, Queen's University.
[5] Brannan, M. (2013). Reduced operator algebras of trace-preserving quantum automorphism
groups. Documenta Mathematica, 18, 1349-1402.
[6] Banica, T., Vergnioux, R. (2009). Growth estimates for discrete quantum groups. Infinite
Dimensional Analysis, Quantum Probability and Related Topics, 12(02), 321-340.
[7] Caspers, M. (2013). The Lp-Fourier transform on locally compact quantum groups. Journal
of Operator Theory, 69(1), 161-193.
[8] Dasgupta, A., Ruzhansky, M. (2014). Gevrey functions and ultradistributions on compact Lie
groups and homogeneous spaces. Bulletin des Sciences Mathmatiques, 138(6), 756-782.
[9] Franz, U., Hong, G., Lemeux, F., Ulrich, M. (2015). Hypercontractivity of heat semigroups
on free quantum groups. Journal of Operator Theory, to appear. arXiv:1511.02753.
[10] Folland, G. B. (2013). Real analysis: modern techniques and their applications. John Wiley
& Sons.
[11] Haagerup, U. (1978). An example of a non nuclearC*-algebra, which has the metric approxi-
mation property. Inventiones Mathematicae, 50(3), 279-293.
[12] Haagerup, U. (1979). Lp-spaces associated with an arbitrary von Neumann algebra. In Al-
gebres doprateurs et leurs applications en physique mathmatique (Proc. Colloq., Marseille,
1977) (Vol. 274, pp. 175-184).
[13] de la Harpe, P. (1988). Groupes hyperboliques, algebres doprateurs et un thoreme de Jolis-
saint. CR Acad. Sci. Paris Sr. I Math, 307(14), 771-774.
[14] Hardy, G. H., Littlewood, J. E. (1927). Some new properties of Fourier constants. Mathema-
tische Annalen, 97(1), 159-209.
[15] Hewitt, E., Ross, K. A. (1974). Rearrangements of Lr Fourier series on compact abelian
groups. Proceedings of the London Mathematical Society, 3(2), 317-330.
[16] Junge, M., Mei, T., Parcet, J. (2014). Smooth Fourier multipliers on group von Neumann
algebras. Geometric and Functional Analysis, 24(6), 1913-1980.
[17] Junge, M., Palazuelos, C., Parcet, J., Perrin, M. (2013). Hypercontractivity in group von
Neumann algebras. Memoirs of American Mathematical Society, to appear. arXiv:1304.5789.
[18] Kemp, T., Speicher, R. (2007). Strong Haagerup inequalities for free R-diagonal elements.
Journal of Functional Analysis, 251(1), 141-173.
[19] Kustermans, J., Vaes, S. (2000, November). Locally compact quantum groups. In Annales
Scientifiques de lEcole Normale Suprieure (Vol. 33, No. 6, pp. 837-934). No longer published
by Elsevier.
[20] Kustermans, J., Vaes, S. (2003). Locally compact quantum groups in the von Neumann
algebraic setting. Mathematica scandinavica, 68-92.
[21] Lee, H. H., Youn, S. (2015). New deformations of Convolution algebras and Fourier algebras
on locally compact groups. Canadian Journal of Mathematics, to appear.
[22] Pisier, G., Xu, Q. (2003). Non-commutative Lp-spaces. Handbook of the geometry of Banach
spaces, 2, 1459-1517.
[23] Runde, V. (2002). Lectures on amenability (No. 1774). Springer Science & Business Media.
21
[24] Timmermann, T. (2008). An invitation to quantum groups and duality: From Hopf algebras
to multiplicative unitaries and beyond. European Mathematical Society.
[25] Vergnioux, R. (2007). The property of rapid decay for discrete quantum groups. Journal of
Operator Theory, 57(2), 303-324.
[26] Wang, S. (2016). Lacunary Fourier series for compact quantum groups. Communications in
Mathematical Physics, 1-51.
[27] Wallach.N.R. (1973). Harmonic analysis on homogeneous spaces. Pure and Applied Mathe-
matics, No. 19. Marcel Dekker, Inc., New York.
[28] Wang, S. (1995). Free products of compact quantum groups. Communications in Mathemat-
ical Physics, 167(3), 671-692.
[29] Wang, S. (1998). Quantum symmetry groups of finite spaces. Communications in mathemat-
ical physics, 195(1), 195-211.
[30] Woronowicz, S. L. (1987). Twisted SU (2) group. An example of a non-commutative dif-
ferential calculus. Publications of the Research Institute for Mathematical Sciences, 23(1),
117-181.
[31] Xiong, X. (2016). Noncommutative harmonic analysis on semigroup and ultracontrativity.
arXiv preprint arXiv:1603.04247.
[32] Xu, Q. (2007). Operator spaces and noncommutative Lp. The part on non-commutative
Lp-spaces. Lectures in the Summer School on Banach spaces and Operator spaces, Nankai
University-China.
Sang-Gyun Youn :
E-mail address: [email protected]
22
|
1712.03436 | 1 | 1712 | 2017-12-09T20:19:00 | Derivations on ternary rings of operators | [
"math.OA",
"math.FA"
] | To each projection $p$ in a $C^*$-algebra $A$ we associate a family of derivations on $A$, called $p$-derivations, and relate them to the space of triple derivations on $p A (1-p)$. We then show that every derivation on a ternary ring of operators is spatial and we investigate whether every such derivation on a weakly closed ternary ring of operators is inner. | math.OA | math |
DERIVATIONS ON TERNARY RINGS OF OPERATORS
ROBERT PLUTA AND BERNARD RUSSO
Abstract. To each projection p in a C ∗-algebra A we associate
a family of derivations on A, called p-derivations, and relate them
to the space of triple derivations on pA(1 − p). We then show
that every derivation on a ternary ring of operators is spatial and
we investigate whether every such derivation on a weakly closed
ternary ring of operators is inner.
1. S-derivations on C*-algebras
If A is a C ∗-algebra, we let D(A) denote the Banach Lie algebra of
derivations on A. To be more precise D(A) consists of all operators
δ ∈ B(A) that satisfy δ(xy) = δ(x)y + xδ(x) for every x, y in A. B(A)
denotes the bounded linear operators on A.
A derivation δ ∈ D(A) is called self-adjoint if δ = δ∗, where δ∗ is the
derivation defined by δ∗(x) = δ(x∗)∗ for every x in A. The space of all
self-adjoint derivations on A is a real Banach Lie subalgebra of D(A)
and is denoted D∗(A).
Derivations on C ∗-algebras have suitable counterparts in a more gen-
eral setting of ternary rings of operators, or TROs for short, where they
are sometimes termed triple derivations. However, in this paper we
shall use the term triple derivation to denote a derivation of a Jordan
triple system. For example, if X is a Banach subspace of a C ∗-algebra
and xy∗z +zy∗x ∈ X for every x, y, z in X, then X is called a JC∗-triple
and a triple derivation on X is an operator τ ∈ B(X) satisfying
τ ({xy∗z}) = {τ (x)y∗z} + {xτ (y)∗z} + {xy∗τ (z)}
for every x, y, z in X, where {xyz} = (xy∗z + zy∗x)/2.
We shall use the term TRO-derivation, as follows: If X is a Banach
subspace of a C ∗-algebra and xy∗z ∈ X for every x, y, z in X, then X
is called a TRO and a TRO-derivation on X is an operator τ ∈ B(X)
satisfying
τ (xy∗z) = τ (x)y∗z + xτ (y)∗z + xy∗τ (z)
Date: September 18, 2018.
Key words and phrases. C*-algebra, ternary ring of operators, TRO, derivation,
linking algebra, W*-TRO.
1
2
ROBERT PLUTA AND BERNARD RUSSO
for every x, y, z in X.
It is clear that a TRO (resp. JC∗-triple) can also be defined as a Ba-
nach subspace of B(H, K), the bounded operators from Hilbert space
H to Hilbert space K, which is closed under the triple product xy∗z
(resp. (xy∗z + zy∗x)/2). If a TRO is weakly closed, it is called a W∗-
TRO.
In this section we will introduce the class of S-derivations on a C ∗-
algebra A associated with a subspace S ⊆ A. Of particular interest
will be the case S = pAp for a projection p in A. We will seek to
determine the relationship between the class of pAp derivations (which
we call p-derivations for short) on A and the class of TRO-derivations
on pA(1 − p).
Definition 1.1. Let A be a C ∗-algebra and let S be a subspace of A.
We say that a derivation δ ∈ D(A) is associated with S, or simply that
δ is an S-derivation, if δ leaves S invariant in the sense that δ(S) ⊆ S.
We use DS(A) to denote the set of all S-derivations.
In order to
simplify the notation, we write De(A) for DeAe(A) in case S = eAe,
for some idempotent e ∈ A, and we abuse the terminology slightly by
referring to the elements of De(A) simply as e-derivations.
To repeat, given an arbitrary idempotent e in a C ∗-algebra A, which
in particular may be a projection, by an e-derivation on A we mean a
derivation δ ∈ D(A) satisfying δ(eAe) ⊆ eAe. This condition is easily
seen to be equivalent to the requirement that δ(e) = 0.
Example 1.2. Let A be a C ∗-algebra and let e ∈ A be an idempotent.
Fix a ∈ eAe and b ∈ (1 − e)A(1 − e) = {x − xe − ex + exe : x ∈ A}.
Then δ : A → A defined by δ(x) = (a + b)x − x(a + b) is an e-derivation.
Lemma 1.3. Let A be a C ∗-algebra and let S be a subalgebra with an
identity element 1S (possibly different from the identity element of A
if A is unital). Let δ ∈ D(A) be a derivation. The following state-
ments hold.
(1) If δ(S) ⊆ S then δ(1S) = 0.
(2) If δ(1S) = 0 then δ(S) ⊆ 1SA1S.
Proof. A straightforward consequence of the derivation property. (cid:3)
Lemma 1.4. Let A be a C ∗-algebra and let e ∈ A be an idempotent.
Let δ ∈ D(A) be a derivation. The following statements hold.
(1) If δ(e) = 0, then δ leaves invariant the following subspaces
eAe,
eA(1 − e),
(1 − e)Ae,
(1 − e)A(1 − e).
DERIVATIONS ON TERNARY RINGS OF OPERATORS
3
(2) If δ leaves invariant eAe or (1 − e)A(1 − e), then δ(e) = 0.
Additionally, let δ = δ∗ and e = e∗. Then the following statement holds.
(3) If δ leaves invariant eA(1 − e) or (1 − e)Ae, then δ(e) = 0.
Proof. The assertions (1) and (2) are straightforward consequences of
the derivation property. To prove (3), assume that eA(1−e) is invariant
for δ = δ∗, and e = e∗. Since δ(e) = δ(e)e + eδ(e), we have eδ(e)e = 0
and hence
δ(e) = eδ(e)(1 − e) + (1 − e)δ(e)e.
This shows that both eδ(e) and δ(e)(1 − e) are equal to eδ(e)(1 − e),
and so both eδ(e) and δ(e)(1−e) are elements of the subspace eA(1−e)
which is invariant under δ.
We will show that δ(e) = 0 by showing that δ(e)2 = 0. For this, we
identify A with (cid:16) eAe
(1−e)Ae (1−e)A(1−e)(cid:17) and write δ(e) and δ2(e) as
eA(1−e)
δ(e) = (cid:16)
0
eδ(e)(1−e)
(1−e)δ(e)e
0
δ2(e) = ( a b
c d ) .
(cid:17) ,
Then δ(e)2 = (cid:16) eδ(e)(1−e)δ(e)e
0
0
(1−e)δ(e)eδ(e)(1−e)(cid:17) and since
δ(eδ(e)) = (cid:16) eδ(e)(1−e)δ(e)e+ea
δ(δ(e)(1 − e)) = (cid:16) −eδ(e)(1−e)δ(e)e
0
0
eb
(1−e)δ(e)eδ(e)(1−e)(cid:17) ∈ (cid:0) 0 eA(1−e)
(cid:1) ,
d(1−e)−(1−e)δ(e)eδ(e)(1−e)(cid:17) ∈ (cid:0) 0 eA(1−e)
b(1−e)
0
0
0
0
it follows that (1 − e)δ(e)eδ(e)(1 − e) = 0 = eδ(e)(1 − e)δ(e)e. Thus
δ(e)2 = 0, as desired.
(cid:3)
(cid:1) ,
If A is a C ∗-algebra and p ∈ A is a projection, we let D∗
p(A) denote
the (real) Banach Lie algebra of self-adjoint p-derivations on A. To be
more precise D∗
p(A) consists of all derivations δ ∈ D(A) that satisfy
δ(p) = 0 and δ = δ∗. If X is a TRO, we use DT RO(X) to denote the
(real) Banach Lie algebra of all TRO-derivations on X.
Remark 1.5. Let A be a unital C ∗-algebra and let p ∈ A be a projection.
Then the map
∆ : D∗
p(A) → DT RO(pA(1 − p)),
∆(δ) = δpA(1−p)
is a homomorphism of Banach Lie algebras.
Example 1.6. Let A = M2(C), p = ( 1 0
on A is:
0 0 ) . The set of all p-derivations
Dp(A) = {δ ∈ D(A) : δ(p) = 0}
≃ (cid:8)(cid:0) α 0
0 β(cid:1) : α, β ∈ C(cid:9)
= a complex Banach Lie algebra.
4
ROBERT PLUTA AND BERNARD RUSSO
The set of all self-adjoint p-derivations is:
D∗
p(A) = {δ ∈ Dp(A) : δ = δ∗}
≃ (cid:8)(cid:0) α 0
0 β(cid:1) : α, β ∈ C with ℜ(α) = ℜ(β)(cid:9)
= a real Banach Lie algebra.
The mapping
∆ : D∗
p(A) → DT RO(X),
∆(δ) = δX
defines a linear surjection between the self-adjoint p-derivations on A
and the TRO-derivations on X = pA(1 − p) = ( 0 C
0 0 ) (see Lemma 2.1).
The kernel of ∆ is isomorphic to the center of A, i.e.,
ker ∆ = Z(A) = {( α 0
0 α ) : α ∈ C} .
In other words, the TRO-derivations on X = pA(1−p) = ( 0 C
cisely the self-adjoint p-derivations on the linking algebra ( XX ∗ X
A = M2(C).
0 0 ) are pre-
X ∗ X ∗X ) =
Example 1.7. Let A = M5(C), and let p ∈ A be the projection matrix
with 1 in the (1, 1) and (2, 2) position and zero's elsewhere. The set of
all p-derivations on A is:
Dp(A) = {δ ∈ D(A) : δ(p) = 0}
0 B ) : A ∈ M2(C), B ∈ M3(C)}
≃ {( A 0
= a complex Banach Lie algebra.
p(A) = {δ ∈ Dp(A) : δ = δ∗}
The set of all self-adjoint p-derivations is D∗
and it can be identified with the real Banach Lie algebra consisting of
all matrices of the form ( A 0
0 B ) where A ∈ M2(C), B ∈ M3(C), and
(cid:0) A+A∗
0
0
B+B∗(cid:1) is in the center of A.
2. Derivations on TROs
If A is a unital C*-algebra and e is a projection in A, then X :=
eA(1 − e) is a TRO. Conversely if X ⊂ B(K, H) is a TRO, then with
X ∗ = {x∗ : x ∈ X} ⊂ B(H, K), XX ∗ = span {xy∗ : x, y ∈ X} ⊂
B(H), X ∗X = span {z∗w : z, w ∈ X} ⊂ B(K), Kl(X) = XX ∗n
,
Kr(X) = X ∗X
, we let1
n
AX = (cid:20)Kl(X) + C1H
X ∗
X
Kr(X) + C1K(cid:21) ⊂ B(H ⊕ K)
1If Kl(X) and Kr(X) are unital subalgebras of B(H) and B(K) (resp.), and X is
nondegenerate, that is, XX ∗X is dense in X, then we take AX to beh Kl(X) X
X ∗ Kr (X)i
DERIVATIONS ON TERNARY RINGS OF OPERATORS
5
denote the (unital) linking C*-algebra of X. Then we have a TRO-
isomorphism X ≃ eAX (1 − e), where e = [ 1 0
0 0 ].
Lemma 2.1. Let X be a TRO and let D : X → X be a TRO-derivation
of X. If A0 = ( XX ∗ X
X ∗ X ∗X ), then the map δ0 : A0 → A0 given by
(cid:18)Pi xiy∗
i
y∗ Pj z∗
x
j wj(cid:19) 7→ (cid:18)Pi(xi(Dyi)∗ + (Dxi)y∗
(Dy)∗
i )
Dx
j (Dwj) + (Dzj)∗wj)(cid:19)
Pj(z∗
is well defined and a bounded *-derivation of A0, which extends D
(when X is embedded in AX via x 7→ ( 0 x
0 0 )), and which itself extends
to a *-derivation δ of AX. Thus, the Lie algebra homomorphism ∆ :
δ 7→ δX given in Remark 1.5 is onto.
Proof. If Pi xiy∗
i = 0, then for every z ∈ X,
xiy∗
i z)
0 = D(Xi
= Xi
= (Xi
((Dxi)y∗
i z + xi(Dyi)∗z + xiy∗
i (Dz))
((Dxi)y∗
i + xi(Dyi)∗))z.
i + xi(Dyi)∗) = 0 (see
[5, Lemma 2.3(iv)]) and it follows that δ0 is well defined.
Since this is true for every z, we have Pi((Dxi)y∗
The map δ0 is self-adjoint since if a = (cid:16) Pi xiy∗
δ0(a∗) = δ0(cid:18)Pi yix∗
j zj(cid:19)
= (cid:18)Pi(yi(Dxi)∗ + (Dyi)x∗
x∗ Pj w∗
(Dx)∗
i )
y
i
i
x
y∗ Pj z ∗
j wj(cid:17), then
Dy
j (Dzj) + (Dwj)∗zj)(cid:19)
Pj(w∗
= δ0(a)∗.
It is easy to verify that δ0(a2) = δ0(a)a+aδ0(a) so that δ0 is a Jordan
*-derivation of A0. (We omit that calculation.)
To see that δ0 is bounded, we first note that D is bounded, since it is
a Jordan triple derivation on the JB*-triple X with the Jordan triple
product {xyz} = (xy∗z + zy∗x)/2, and hence bounded by the theorem
i ) by
of Barton and Friedman [1]. Now denoting Pi(xi(Dyi)∗ + (Dxi)y∗
6
ROBERT PLUTA AND BERNARD RUSSO
α, we have (by [5, Lemma 2.3(iv)] again) kPi(xi(Dyi)∗ + (Dxi)y∗
kαk =
sup
kαzk
i )k =
kzk≤1,z∈X
sup
kαz +
kzk≤1,z∈X
sup
kzk≤1,z∈X
sup
kzk≤1,z∈X
kXi
kDXi
=
=
=
z
Xi
D(xiy∗
xiy∗
=0
xiy∗
}
i (Dz) −Xi
i z) −Xi
i z − kDkXi
xiy∗
xiy∗
i (Dz) k
{
i (Dz)k
xiy∗
i
Dz
kDk
k
≤ 2kDkkXi
xiy∗
i k.
Thus δ0 is bounded and therefore extends to a bounded Jordan *-
derivation δ of A0
and hence to AX by setting δ(e) = 0, where e =
(cid:2) 1H 0
0 0(cid:3). By the theorem of Sinclair ([16, Theorem 3.3]), δ is a derivation
of AX.2
(cid:3)
n
For any C*-algebra A ⊂ B(H), the Lie algebra homomorphism A
∋
) is onto (theorem of Kadison and Sakai ([15, 4.1.6]))
w
w
z 7→ ad z ∈ D(A
and so we have the Lie algebra isomorphism
w
) ≃ D(A
w
/Z(A
A
w
).
It follows (cf. [15, 4.1.7]) that
w
{t ∈ A
w
: ad t(A) ⊂ A}/Z(A
) ≃ D(A),
and
w
{t ∈ A
w
: t∗ = −t, ad t(A) ⊂ A}/Z(A
) ≃ D∗(A).
Further, for a projection e in A, we have
w
{t ∈ A
w
: et = te, t∗ = −t, ad t(A) ⊂ A}/Z(A
) ≃ D∗
e(A).
Using these facts in the setting of Lemma 2.1, and noting that, by
[10, page 268], AX
following theorem.
w
= A′′
X = h Kl(X)′′ X
Kr(X)′′i, we can now prove the
X ∗w
w
Theorem 2.2. Every TRO-derivation of a TRO X is spatial in the
sense that there exist α ∈ Kl(X)′′ and β ∈ Kr(X)′′ such that α∗ =
−α, β∗ = −β, and Dx = αx + xβ for every x ∈ X.
2It is also easy to verify directly, by (a more involved) calculation, that δ0 is a
derivation, thereby avoiding the use of Sinclair's theorem
DERIVATIONS ON TERNARY RINGS OF OPERATORS
7
Proof. If D ∈ DT RO(X), choose δ = ad t for some t ∈ AX
t∗ = −t, te = et and
w
with
Moreover
The conditions on t imply that t = (cid:2) α 0
δ(cid:20)0 x
0 0(cid:21) = (cid:20)α 0
0 0(cid:21) −(cid:20)0 x
0 β(cid:21)(cid:20)0 x
(cid:20)0 Dx
0
0 (cid:21) = δ(cid:20)0 x
0 0(cid:21) .
0 β(cid:3) with α∗ = −α and β∗ = −β.
0 β(cid:21) = (cid:20)0 αx + x(−β)
(cid:21) .
0 0(cid:21)(cid:20)α 0
0
0
(cid:3)
j=1(c∗
A TRO derivation D of a TRO X is said to be an inner TRO deriva-
tion if there exist α = −α∗ ∈ XX ∗ and β = −β∗ ∈ X ∗X such that
Dx = αx + xβ for x ∈ X. Note that there exist ai, bi, cj, dj ∈ X,
i ) and β =
i − bia∗
1 ≤ i ≤ n, 1 ≤ j ≤ m such that α = Pn
Pm
j dj − d∗
Corollary 2.3. Every TRO derivation of a C∗-algebra A is of the form
with α∗ = −α, β∗ = −β.
A ∋ x 7→ αx + xβ with elements α, β ∈ A
In particular, every TRO derivation of a von Neumann algebra is an
inner TRO derivation
i=1(aib∗
j cj).
w
Thus, every W∗-TRO which is TRO-isomorphic to a von Neumann
algebra has only inner TRO derivations. For example, this is the case
for the stable W∗-TROs of [13] (see subsection 3.2) and the weak*-
closed right ideals in certain continuous von Neumann algebras acting
on separable Hilbert spaces (see Theorem 3.3).
Theorem 2.2 is an improvement of [18], in which, although proved
for the slightly more general case of derivation pairs, it is assumed
that the TRO (called B*-triple system in [18]) contains the finite rank
operators. For the extension of Zalar's result to unbounded operators,
see [17].
A triple derivation δ of a JC∗-triple X is said to be an inner triple
derivation if there exist finitely many elements ai, bi ∈ X, 1 ≤ i ≤ n,
i=1({aibix} − {biaix}) for x ∈ X, where {xyz} =
(xy∗z + zy∗x)/2. For convenience, we denote the inner triple derivation
x 7→ {abx} − {bax} by δ(a, b). Thus
such that δx = Pn
δ(a, b)(x) = (ab∗x + xb∗a − ba∗x − xa∗b)/2.
Let X be a TRO. As noted in the proof of Lemma 2.1, X is a
JC∗-triple in the triple product (xy∗z + zy∗x)/2, and every TRO-
derivation of X is obviously a triple derivation. On the other hand,
Indeed, if
every inner triple derivation is an inner TRO-derivation.
8
ROBERT PLUTA AND BERNARD RUSSO
δ(x) = {abx} − {bax}, for some a, b ∈ X, then δ(x) = Ax + xB, where
A = ab∗ −ba∗ ∈ XX ∗, B = b∗a−a∗b ∈ X ∗X with A, B skew-hermitian.
Moreover, since by [1, Theorem 4.6], every triple derivation δ on X is
the strong operator limit of a net δα of inner triple derivations, hence
TRO-derivations, we have (i) and (ii) in the following proposition.
Proposition 2.4. Let X be a TRO.
(i): Every TRO-derivation is the strong operator limit of inner
TRO-derivations.
(ii): The triple derivations on X coincide with the TRO-derivations.
(iii): The inner triple derivations on X coincide with the inner
TRO-derivations
(iv): All TRO derivations of X are inner, if and only if, all triple
derivations of X are inner.
Proof. Since (iv) is immediate from (ii) and (iii), we only need to show
part of (iii), that is, that every inner TRO-derivation is an inner triple
derivation. If D is an inner TRO-derivation, then Dx = αx + xβ, with
α∗ = −α ∈ XX ∗ and β∗ = −β ∈ X ∗X. We must show that there
k=1 δ(ak, bk)x where δ(ak, bk) is
i=1 xiy∗
i
j wj, then it suffices to take p = m + n and choose
ai = xi/2, bi = yi for 1 ≤ i ≤ n and an+i = wi, bn+i = zi/2 for
1 ≤ i ≤ m.
(cid:3)
exist elements ak, bk such that Dx = Pp
the inner triple derivation x 7→ {akbkx} − {bkakx}. If α = Pn
and β = Pm
j=1 z∗
3. Derivations on W*-TROs
A von Neumann algebra M is an example of a unital reversible JW ∗-
algebra, and as such, by [7, Theorem 2 and the first sentence in its
proof], every triple derivation on M is an inner triple derivation. Hence
we see that the last statement in Corollary 2.3 follows also from this
and Proposition 2.4(iv). For completeness, we include a proof of the
former result which avoids much of the Jordan theory, starting with
the following lemma, the first part of which is straightforward.
Lemma 3.1. Let A be a unital Banach ∗-algebra equipped with the
ternary product given by {a, b, c} = 1
2 (ab∗c + cb∗a) and the Jordan
product a ◦ b = (ab + ba)/2.
• Let D be an inner derivation, that is, D = ad a : x 7→ ax − xa,
for some a in A. Then D = ad a is a *-derivation whenever
a∗ = −a. Conversely, if D is a *-derivation, then a∗ = −a + z
for some z in the center of A.
• Every triple derivation is the sum of a Jordan *-derivation and
an inner triple derivation.
DERIVATIONS ON TERNARY RINGS OF OPERATORS
9
Proof. To prove the second statement, we modify the proof in [8, Sec-
tion 3] which is in a different context. We note first that for a triple
derivation δ, δ(1)∗ = −δ(1). Next, for a triple derivation δ, the map-
ping δ1(x) = δ(1) ◦ x is equal to the inner triple derivation − 1
4 δ(δ(1), 1)
so that δ0 := δ − δ1 is a triple derivation with δ0(1) = 0. Finally, any
triple derivation which vanishes at 1 is a Jordan *-derivation.
(cid:3)
Theorem 3.2. Every triple derivation on a von Neumann algebra is
an inner triple derivation.
Proof. It suffices, by the second statement in Lemma 3.1, to show that
every self-adjoint Jordan derivation is an inner triple derivation.
If
δ is a self-adjoint Jordan derivation of M, then δ is an associative
derivation (by the theorem of Sinclair, [16, Theorem 3.3])) and hence by
the theorem of Kadison and Sakai ([15, 4.1.6]) and the first statement
in Lemma 3.1, δ(x) = ax − xa where a∗ + a = z is a self adjoint
element of the center of M. Since for every von Neumann algebra, we
have M = Z(M) + [M, M], where Z(M) denotes the center of M (see
[12, Section 3] for a discussion of this fact), we can therefore write
a = z′ +Xj
= z′ +Xj
[bj + icj, b′
j + ic′
j]
([bj, b′
j] − [cj, c′
([cj, b′
j] + [bj, c′
j]),
j]) + iXj
where bj, b′
j, cj, c′
It follows that
j are self adjoint elements of M and z′ ∈ Z(M).
0 = a∗ + a − z = (z′)∗ + z′ − z + 2iXj
([cj, b′
j] + [bj, c′
j])
j] + [bj, c′
j]) belongs to the center of M. We now have
so that Pj([cj, b′
δ = ad a = ad Xj
([bj, b′
j] − [cj, c′
j]).
Pj(cid:0)δ(bj, 2b′
A direct calculation shows that δ is equal to the inner triple derivation
(cid:3)
j) − δ(cj, 2c′
j)(cid:1), completing the proof.
3.1. Weakly closed right ideals in von Neumann algebras. In
this subsection, we shall consider the TRO pM where M is a von
Neumann algebra and p is a projection in M.
A TRO of the form pM, with M a continuous von Neumann algebra,
is classified into four types in [9] as follows.
• II a
• II a
1 if M is of type II1 and p is (necessarily) finite.
∞,1 if M is of type II∞ and p is a finite projection.
10
ROBERT PLUTA AND BERNARD RUSSO
• II a
∞ if M is of type II∞ and p is a properly infinite projection.
• III a if M is of type III and p is a (necessarily) properly infinite
projection.
Similarly, we also define types for pM for M of type I:
1 if M is finite of type I and p is (necessarily) finite.
∞,1 if M is of type I∞ and p is a finite projection.
∞ if M is of type I∞ and p is a properly infinite projection.
• I a
• I a
• I a
The following theorem involves the cases II a
∞, III a and when M is
a factor, the cases I a
1 , I a
1,∞, and I a
∞.
Theorem 3.3. Let X = pM be a TRO, where M is a von Neumann
algebra and p is a projection in M.
(i): If X is of type II a
∞ or III a, and has a separable predual, then
every TRO-derivation of X is an inner TRO-derivation.
(ii): If M is of type III and countably decomposable, then every
TRO-derivation of X = pM is an inner TRO-derivation.
(iii): If M = B(H) is a factor of type I, then
(1) If dim H < ∞, then every TRO-derivation of X = pM is
an inner TRO-derivation.
(2) If dim pH = dim H, then every TRO-derivation of X =
pM is an inner TRO-derivation.
(3) If dim pH < dim H = ∞, then X = pM admits outer
TRO-derivations.
Proof. If M is a continuous von Neumann algebra with a separable
predual and p is a properly infinite projection in M, then it is shown
in [9, Theorem 5.16] that pM is triple isomorphic to a von Neumann
algebra, and hence by Theorem 3.2, every triple derivation is an inner
triple derivation in this case. Consequently, by Proposition 2.4(iv),
every TRO-derivation is an inner TRO-derivation. (Another way to see
this latter fact is to note that by [9, Lemma 5.15], pM is actually TRO-
isomorphic to a von Neumann algebra, and to apply Corollary 2.3.)
This proves (i).
To prove (ii), we note first that if A is a von Neumann algebra with
a projection p ∼ 1, then pA is TRO-isomorphic to A. Indeed, If u is
a partial isometry in A with uu∗ = p and u∗u = 1, then x 7→ u∗x is
a TRO-isomorphism from pA onto A. Now if A is of type III, then
A := c(p)A is of type III, c(p) is the identity of A and pA = p A.
Further, if A is countably decomposable, then by [15, 2.2.14], since in
A, c(p) = 1 A = c(1 A), we have p ∼ 1 A, so A is TRO-isomorphic to
p A = pA.
DERIVATIONS ON TERNARY RINGS OF OPERATORS
11
Finally, let M = B(H). The first statement in (iii) follows from
the fact that every finite dimensional semisimple Jordan triple system
has only inner derivations. This result first appeared in [11, Chapter
11] (for an outline of a proof, see [14, Theorem 2.8,p. 136] and for the
definitions of Jordan triple system and semisimple, see [3, Section 1.2]).
If dim pH = dim H, then pM ≃ B(H) has only inner triple derivations
by Theorem 3.2. On the other hand, if dim pH < dim H = ∞, then
pM ≃ B(H, pH) has outer triple derivations, as shown in [7, Corollary
3]. By Proposition 2.4(iv), this proves (iii)
(cid:3)
Remark 3.4. Although it follows from Theorem 3.3, it is worth pointing
out that the TROs B(C, H) and B(H, C) support outer TRO deriva-
tions if and only if dim H = ∞. According to [9, Lemma 5.15], if B
is a von Neumann algebra of type II∞ or III, and H is a separable
Hilbert space, then B and B⊗B(C, H) are TRO-isomorphic. Corol-
lary 2.3 shows that B⊗B(C, H) has only inner TRO-derivations and
only inner triple derivations, although, as just noted, B(C, H) can have
an outer TRO derivation and an outer triple derivation. This contrasts
the situation of derivations on tensor products of C∗-algebras, as in
[2, Proposition 3.2]).
3.2. W*-TROs of types I,II,III. We begin by recalling some con-
cepts from [13]. If R is a von Neumann algebra and e is a projection
in R, then V := eR(1 − e) is a W*-TRO. Conversely if V ⊂ B(K, H)
is a W*-TRO, then with V ∗ = {x∗ : x ∈ V } ⊂ B(H, K), M(V ) =
XX ∗sot
⊂ B(H), N(V ) = X ∗X
⊂ B(K), we let
sot
RV = (cid:20)M(V )
V ∗
V
N(V )(cid:21) ⊂ B(H ⊕ K)
denote the linking von Neumann algebra of V . Then we have a SOT-
continuous TRO-isomorphism V ≃ eRe⊥, where e = (cid:2) 1H 0
(cid:2) 0 0
0 1K(cid:3).
A W*-TRO V is stable if it is TRO-isomorphic to B(ℓ2)⊗V . A W*-
TRO is of type I,II,or III, by definition, if its linking von Neumann
algebra is of that type as a von Neumann algebra. There is a further
classification of the types I and II depending on the types of M(V ) and
N(V ) leading to the types Im,n, IIα,β where m, n are cardinal numbers
and α, β ∈ {1, ∞}. See [13, Section 4] for detail.
0 0(cid:3) and e⊥ =
In what follows, for ultraweakly closed subspaces A ⊂ M and B ⊂
N, where M and N are von Neumann algebras, A⊗B denotes the
ultraweak closure of the algebraic tensor product A ⊗ B.
We shall use the following results from [13], which we summarize as
a theorem.
12
ROBERT PLUTA AND BERNARD RUSSO
Theorem 3.5 (Ruan [13]). Let V be a W∗-TRO acting on separable
Hilbert spaces.
(i) [13, Theorem 3.2] If V is a stable W*-TRO, then V is TRO-
isomorphic to M(V ) and to N(V ).
(ii) [13, Corollary 4.3] If V is a W*-TRO of one of the types I∞,∞, II∞,∞
or III, then V is a stable W*-TRO, and hence TRO-isomorphic to a
von Neumann algebra.
(iii) [13, Theorem 4.4] If V is a W*-TRO of type II1,∞ (respec-
tively II∞,1), then V is TRO-isomorphic to B(H, C)⊗M (respectively
B(C, H)⊗N), where M (respectively N) is a von Neumann algebra of
type II1.
Because taking a transpose is a triple isomorphism, we have the
following consequence of Theorem 3.5(iii).
Lemma 3.6. A W*-TRO of type II1,∞ is triple isomorphic to a W*-
TRO of type II∞,1. More precisely, B(H, C)⊗M is triple isomorphic
to B(C, H)⊗M t, where xt = Jx∗J, for x ∈ M ⊂ B(H) and J is a
conjugation on H.
Proposition 3.7. Let V be a W*-TRO.
(i): If V acts on a separable Hilbert space and is of one of the
types I∞,∞, II∞,∞ or III, then every triple derivation of V is
an inner triple derivation and every TRO derivation of V is an
inner TRO-derivation.
(ii): If every TRO-derivation of any W∗-TRO of type II1,∞ has
only inner TRO-derivations, then every TRO-derivation of any
W∗-TRO of type II∞,1 has only inner TRO-derivations. The
converse also holds.
Proof. (i) is an immediate consequence of Theorem 3.5(ii), Theorem 3.2
and Proposition 2.4(iv). (ii) is an immediate consequence of Lemma 3.6
and Proposition 2.4(iv).
(cid:3)
It follows from Remark 3.4 that if M is a von Neumann algebra of
type II∞ or III and H is a separable Hilbert space, then B(C, H)⊗M is
triple isomorphic to a von Neumann algebra and hence has only inner
TRO-derivations, giving alternate proofs of parts of Proposition 3.7(i).
By [13, Theorem 4.1], if V is a W*-TRO of type I, then V is TRO-
isomorphic to ⊕αL∞(Ωα)⊗B(Kα, Hα). In the next two results, we con-
sider the related TRO C(Ω, B(H, K)), where Ω is a compact Hausdorff
space.
Lemma 3.8. Let E be a TRO and Ω a compact Hausdorff space.
DERIVATIONS ON TERNARY RINGS OF OPERATORS
13
(i): If every TRO derivation of V := C(Ω, E) is an inner TRO
derivation, then the same holds for E.
(ii): If every triple derivation of V := C(Ω, E) is an inner triple
derivation, then the same holds for E.
Proof. By Proposition 2.4(iv), it is sufficient to prove (i).
If D is a TRO derivation of E, then δf (ω) := D(f (ω)) is a TRO
derivation of V , as is easily checked. Suppose every TRO derivation
of V is an inner TRO derivation. Then δf = αf + f β, where α =
j wj for some
i for some xi, yi ∈ V , and β = −β∗ = Pi z∗
−α∗ = Pi xiy∗
zj, wj ∈ V .
For a ∈ E, let 1 ⊗ a ∈ V be the constant function equal to a. Then
D(a) = D((1 ⊗ a)(ω)) = δ(1 ⊗ a)(ω) = α(ω)a + aβ(ω) for all ω ∈ Ω.
Since α(ω)∗ = −α(ω) ∈ EE∗ and β(ω)∗ = −β(ω) ∈ E∗E, D is an
inner TRO derivation of E.
(cid:3)
Recall from Theorem 3.3(iii) that the TRO B(H, K) supports outer
TRO derivations if and only if it is infinite dimensional and dim H 6=
dim K.
Proposition 3.9. If V = ⊕αC(Ωα, Eα), where Eα = B(Kα, Hα) and
if every triple derivation of V is an inner triple derivation, then for
every α, either dim Eα < ∞ or dim Kα = dim Hα.
Proof. Let δ be a triple derivation of V , and let δα = δC(Ωα,Eα), which is
a triple derivation of the weak*-closed ideal C(Ωα, Eα). Then δ({fα}) =
{δαfα}. Moreover if δ is an inner triple derivation, say δ = Pi δ(ai, bi)
α) is an inner triple
α, bi
for ai = {ai
derivation of C(Ωα, Eα).
α}, bi = {bi
α} ∈ V , then δα = Pi δ(ai
Now suppose that every triple derivation of V is an inner triple
derivation, and that for some α0, Eα0 is infinite dimensional and dim Kα0
does not equal dim Hα0. Then, as noted above, there is an outer triple
derivation D of C(Ωα0, Eα0). Then the triple derivation on V which is
zero on C(Ωα, Eα) for α 6= α0 and equal to D on C(Ωα0, Eα0), cannot
be inner by the preceding paragraph, which is a contradiction.
(cid:3)
4. Some questions left open
Questions 1. It remains to complete the results of Theorem 3.3 to
include the cases where p is a finite projection in a continuous von
Neumann algebra, or when p is arbitrary and M is a general von Neu-
mann algebra of type I. As a possible tool for the first question, we
note that there is an alternate proof of Proposition 2.4 (ii), in the case
X = pM, p finite, using the technique in [6, Section II.B].
14
ROBERT PLUTA AND BERNARD RUSSO
Questions 2. Besides the problem of extending the known cases to
non separable Hilbert spaces, the cases left open in Proposition 3.7 for
arbitrary W*-TROs are those of types II1,1 and II1,∞ (the latter being
equivalent to II∞,1).
Questions 3. Let E be a W*-TRO, and let V = ⊕αL∞(Ωα)⊗B(Kα, Hα)
be a W*-TRO of type I.
• If every derivation of the W*-TRO L∞(Ω)⊗E is inner, does it
follow that every derivation of E is inner?
• If every derivation of V is inner, does it follow that dim B(Kα, Hα) <
∞, for all α; or supα dim B(Kα, Hα) < ∞?
• If supα dim B(Kα, Hα) < ∞, does it follow that every derivation
of V is inner?
Remark 4.1. With respect to Questions 3,
(i) In the first bullet, if E had a separable predual, then a variant of
[15, 1.22.13] would state that L∞(Ω)⊗E = L∞(Ω, E) and the technique
in Lemma 3.8 could be used.
(ii) In the first bullet, suppose that E = pM, with M a von Neumann
algebra in B(H) and p a projection in M, and let D is a derivation
of E. Then δ := id ⊗ D is a derivation of V = L∞(Ω)⊗E. Assuming
that δ is inner, there exist α = −α∗ ∈ V V ∗ = L∞(Ω)⊗(EE∗) (EE∗
denoting the weak closure) and β ∈ V ∗V = L∞(Ω)⊗(E∗E), such that
1 ⊗ Dx = α(1 ⊗ x) + (1 ⊗ x)β,
(x ∈ E).
We have EE∗ = pMp ⊂ B(pH), E∗E ⊂ B(H), and L∞(Ω) ⊂ B(L2(Ω)).
For each ϕ ∈ B(L2(Ω))∗, let Rϕ : B(L2(Ω)⊗pH) → B(pH) be the slice
map of Tomiyama defined by Rϕ(f ⊗ x) = ϕ(f )x ([4, Lemma 7.2.2]).
Since V V ∗ is the ultraweak closure of L∞(Ω) ⊗ EE∗, by the weak*-
continuity of Rϕ, we have
ϕ(1)Dx = Rϕ(α)x + xRϕ(β)
with Rϕ(α) ∈ EE∗ and Rϕ(β) ∈ E∗E. Thus, if dim H = dim pH, or
if E is finite dimensional, then Rϕ(β) ∈ E∗E, so that D is an inner
TRO-derivation, (take ϕ to be a normal state so that Rϕ is self-adjoint
and Rϕ(α)∗ = −Rϕ(α) and Rϕ(β)∗ = −Rϕ(β).) In general, D could
be called a "quasi-inner" TRO-derivation.
(iii) In the second bullet, if each B(Kα, Hα) had a separable predual,
then a variant of [15, 1.22.13] would state that L∞(Ωα)⊗B(Kα, Hα) =
L∞(Ωα, B(Kα, Hα)) and the technique in Proposition 3.9 could be used.
DERIVATIONS ON TERNARY RINGS OF OPERATORS
15
References
[1] T. J. Barton and Y. Friedman, Bounded derivations of JB∗-triples, Quart. J.
Math. Oxford Ser. (2) (1990), 255 -- 268.
[2] C. J. K. Batty, Derivations of tensor products of C*-algebras, J. London Math.
Soc. (2) 17 (1978), 129 -- 140.
[3] Cho-Ho Chu, Jordan structures in geometry and analysis, Cambridge Tracts
in Mathematics, vol. 190, Cambridge University Press, Cambridge, 2012.
[4] E. G. Effros and Z. J. Ruan, Operator Spaces, London Mathematical Society
Monographs, vol. 23, Clarendon Press Oxford, 2000.
[5] M. Hamana, Triple envelopes and Silov boundaries of operator spaces, Math.
J. Toyama Univ. 22 (1999), 77 -- 93.
[6] J. Hamhalter, K. K. Kudaybergenov, A. M. Peralta, and B. Russo, Bounded-
ness of completely additive measures with application to 2-local triple deriva-
tions, J. Math. Physics 57 (2016), no. 2, 22 pp.
[7] T. Ho, J. Mart´ınez-Moreno, A. M. Peralta, and B. Russo, Derivations on real
and complex JB∗-triples, J. London Math. Soc. (2) 65 (2002), no. 1, 85 -- 102.
[8] T. Ho, A. M. Peralta, and B. Russo, Ternary weakly amenable C*-algebras and
JB*-triples, Quarterly J. Math. 64 (2013), 1109 -- 1139.
[9] G. Horn and E. Neher, Classification of continuous JBW∗-triples, Trans. Amer.
Math. Soc. 306 (1988), 553 -- 578.
[10] Manmohan Kaur and Zhong-Jin Ruan, Local properties of ternary rings of
operators and their linking C*-algebras, J. Functional Analysis 195 (2002),
262 -- 305.
[11] Kurt Meyberg, Lectures on algebras and triple systems, The University of Vir-
ginia, Charlottesville, Va., 1972.
[12] Robert Pluta and Bernard Russo, Triple derivations on von Neumann algebras,
Studia Math. 226 (2015), no. 1, 57-73.
[13] Zhong-Jin Ruan, Type decomposition and the rectangular AFD property for
W∗-TROs, Canad. J. Math. 36 (2004), no. 4, 843 -- 870.
[14] Bernard Russo, Derivations and projections on Jordan triples: an introduc-
tion to nonassociative algebra, continuous cohomology, and quantum functional
analysis, In: Advanced courses of mathematical analysis V (2016), 118 -- 227.
World Sci. Publ., Hackensack, NJ.
[15] S. Sakai, C*-algebras and W*-algebras, Ergebnisse der Mathematik und ihrer
Grenzgebiete, vol. 60, Springer-Verlag, New York Heidelberg Berlin, 1971.
[16] A. M. Sinclair, Jordan homomorphisms and derivations on semisimple Banach
algebras, Proc. Amer. Math. Soc. 24 (1970), 209 -- 214.
[17] W. Timmermann, Remarks on automorphism and derivation pairs in ternary
rings of unbounded operators, Arch. Math. (Basel) 74 (2000), no. 5, 379 -- 384.
[18] Borut Zalar, On the structure of automorphism and derivation pairs of B ∗-
triple systems, Topics in operator theory, operator algebras and applications
(Timisoara 1994), Rom. Acad., Bucharest (1995), 265 -- 271.
16
ROBERT PLUTA AND BERNARD RUSSO
Department of Mathematics, University of California, Irvine, CA
92697-3875, USA
E-mail address: [email protected]
Department of Mathematics, University of California, Irvine, CA
92697-3875, USA
E-mail address: [email protected]
|
1804.01930 | 1 | 1804 | 2018-04-05T16:06:28 | Operator-valued Triebel-Lizorkin spaces | [
"math.OA",
"math.FA"
] | This paper is devoted to the study of operator-valued Triebel-Lizorkin spaces. We develop some Fourier multiplier theorems for square functions as our main tool, and then study the operator-valued Triebel-Lizorkin spaces on $\mathbb{R}^d$. As in the classical case, we connect these spaces with operator-valued local Hardy spaces via Bessel potentials. We show the lifting theorem, and get interpolation results for these spaces. We obtain Littlewood-Paley type, as well as the Lusin type square function characterizations in the general way. Finally, we establish smooth atomic decompositions for the operator-valued Triebel-Lizorkin spaces. These atomic decompositions play a key role in our recent study of mapping properties of pseudo-differential operators with operator-valued symbols. | math.OA | math | OPERATOR-VALUED TRIEBEL-LIZORKIN SPACES
RUNLIAN XIA AND XIAO XIONG
8
1
0
2
r
p
A
5
]
Abstract. This paper is devoted to the study of operator-valued Triebel-Lizorkin spaces. We
develop some Fourier multiplier theorems for square functions as our main tool, and then study
the operator-valued Triebel-Lizorkin spaces on Rd. As in the classical case, we connect these
spaces with operator-valued local Hardy spaces via Bessel potentials. We show the lifting the-
orem, and get interpolation results for these spaces. We obtain Littlewood-Paley type, as well
as the Lusin type square function characterizations in the general way. Finally, we establish
smooth atomic decompositions for the operator-valued Triebel-Lizorkin spaces. These atomic
decompositions play a key role in our recent study of mapping properties of pseudo-differential
operators with operator-valued symbols.
Contents
.
A
O
h
t
a
m
[
1
v
0
3
9
1
0
.
4
0
8
1
:
v
i
X
r
a
Introduction and preliminaries
0.
Noncommutative Lp-spaces
Fourier analysis
1. Operator-valued local Hardy spaces
Calder´on-Zygmund theory
Characterizations
Atomic decomposition
2. Multiplier theorems
2.1. Global multipliers
2.2. Hilbert-valued multipliers
2.3. Multipliers on hc
p
3. Operator-valued Triebel-Lizorkin spaces
3.1. Definitions and basic properties
3.2.
3.3. Triebel-Lizorkin spaces with α > 0
4. Characterizations
4.1. General characterizations
4.2. Characterizations via Lusin functions
5. Smooth atomic decomposition
5.1. Smooth atomic decomposition of hc
5.2. Atomic decomposition for F α,c
References
Interpolation
1(Rd,M)
1
(Rd,M)
1
3
3
4
6
6
7
8
8
9
15
18
18
22
23
24
24
31
32
32
36
43
0. Introduction and preliminaries
2 ≤ ξ ≤ 2}, ϕ > 0 on {ξ : 1
Let ϕ be a Schwartz function on Rd such that supp ϕ ⊂ {ξ : 1
ξ < 2}, and Pk∈Z ϕ(2−kξ) = 1 for all ξ 6= 0. For each k ∈ N, let ϕk be the function whose
Fourier transform is equal to ϕ(2−k·), and let ϕ0 be the function whose Fourier transform is equal
to 1 −Pk>0 ϕ(2−k·). Then {ϕk}k≥0 gives a Littlewood-Paley decomposition on Rd. The classical
(inhomogeneous) Triebel-Lizorkin spaces F α
p,q(Rd) for 0 < p < ∞, 0 < q ≤ ∞ and α ∈ R are
2 <
2000 Mathematics Subject Classification: Primary: 46L52, 42B30. Secondary: 46L07, 47L65.
Key words: Noncommutative Lp-spaces, operator-valued Triebel-Lizorkin spaces, operator-valued Hardy spaces,
Fourier multipliers, interpolation, characterizations, atomic decomposition.
1
2
R. Xia and X. Xiong
defined as
with the (quasi-)norm
p,q(Rd) = {f ∈ S′(Rd) : kfkF α
F α
p,q
< ∞}
kfkF α
p,q =(cid:13)(cid:13)(Xj≥0
2qjαϕj ∗ fq)
.
1
q(cid:13)(cid:13)p
We refer the reader to Triebel's books [34] and [35] for more concrete definition and properties
of Triebel-Lizorkin spaces on Rd. This kind of function spaces is closely related to other function
spaces, such as Sobolev and Besov spaces. In particular, Triebel-Lizorkin spaces can be viewed
as generalizations of Hardy spaces, since the Bessel potential J α is known to be an isomorphism
between F α
p,2(Rd) and hp(Rd) (local Hardy spaces introduced in [8]). All these spaces are ba-
sic for many branches of mathematics such as harmonic analysis, PDE, functional analysis and
approximation theory.
This paper is devoted to the study of operator-valued Triebel-Lizorkin spaces. As in the classical
case, it can be viewed as an extension of our recent work [36] on operator-valued local Hardy spaces.
On the other hand, the Triebel-Lizorkin spaces studied here are Euclidean counterparts of those on
usual and quantum tori studied in [40]. Our main motivation is to build a kind of function spaces
where we can carry out the investigation of pseudo-differential operators with operator-valued
symbols.
1
Due to noncommutativity, there are several obstacles on our route, which do not appear in
the classical case. First of all, in the noncommutative integration, the simple replacement of
the usual absolute value by the modulus of operators in the formula (cid:13)(cid:13)(Pj≥0 2qjαϕj ∗ fq)
does not give a norm except for q = 2. Even though one could use Pisier's definition of ℓq-
valued noncommutative Lp-spaces by complex interpolation (see [24]), we will not study that
kind of spaces and will focus only on the case q = 2. The reason for this choice is that, for
q = 2, the Triebel-Lizorkin spaces of operator-valued distributions are isomorphic to the Hardy
spaces developed in [36], as mentioned above. Another difficulty is the lack of pointwise maximal
functions in the noncommutative case. As is well known, the maximal functions play a crucial
role in the classical theory; but they are no longer at our disposal in the noncommutative setting.
In [40], when studying the Triebel-Lizorkin spaces on quantum tori, we use Calder´on-Zygmund
and Fourier multiplier theory as substitution.
In this paper, we will still rely heavily on this
theory. However, we have to consider its local (or inhomogeneous) counterpart, since the theory
used in [40] for quantum tori are nonlocal (or homogeneous) ones. Besides the local nature, we
also develop Hilbert space valued Fourier multiplier theory, which will be used to deduce general
characterizations of operator-valued Triebel-Lizorkin spaces by the Lusin type square function.
q(cid:13)(cid:13)p
Our definition of (column) operator-valued Triebel-Lizorkin spaces is
(Rd,M) = {f ∈ S′(Rd; L1(M) + M) : kfkF α,c
F α,c
p
p < ∞},
where
kfkF α,c
p =(cid:13)(cid:13)(Xj≥0
22jαϕj ∗ f2)
.
1
2(cid:13)(cid:13)p
p
Here the norm k · kp is the norm of the semi-commutative Lp-space Lp(L∞(Rd)⊗M). Different
from the classical case, we have also row and mixture versions; see section 3 for concrete definitions.
We present here two major results of this paper. The first one gives general characterizations
of F α,c
(Rd,M) by any reasonable convolution kernels in place of the Littlewood-Paley decom-
position {ϕj}j≥0. These characterizations can be realized either by the Littlewood Paley type
g-function or by the Lusin type integral function, with the help of the Calder´on-Zygmund theory
and Fourier multiplier theory mentioned above. The second major result is the atomic decom-
position of F α,c
(Rd,M). When α = 0, in [36], we deduce from the h1-bmo duality an atomic
decomposition of hc
1(Rd,M), which does not require any smooth condition on each atom. In this
paper, we refine the smoothness of that atomic decomposition by the Calder´on reproducing identity,
via tent spaces. Using the same trick, we extend that refinement to F α,c
(Rd,M); but compared
with the case of local Hardy spaces, subatoms enter in the game. These smooth atomic decom-
positions will play a crucial role in the study of pseudo-differential operators in the forthcoming
paper [37].
1
1
Inhomogeneous Triebel-Lizorkin spaces
3
In the following, let us recall some notation and background in the interface between harmonic
analysis and operator algebras that we will need throughout the paper, although they are probably
well-known to experts.
Noncommutative Lp-spaces. We start with a brief introduction of noncommutative Lp spaces.
Let M be a von Neumann algebra equipped with a normal semifinite faithful trace τ ; for 1 ≤ p ≤ ∞,
let Lp(M) be the noncommutative Lp-space associated to (M, τ ). The norm of Lp(M) will be often
denoted simply by k · kp. But if different Lp-spaces appear in a same context, we will sometimes
precise the respective Lp-norms in order to avoid possible ambiguity. The reader is referred to
[26] and [41] for more information on noncommutative Lp-spaces. We will also need Hilbert space-
valued noncommutative Lp-spaces (see [14] for more details). Let H be a Hilbert space and v ∈ H
with kvk = 1. Let pv be the orthogonal projection onto the one-dimensional subspace generated
by v. Define
Lp(M; H r) = (pv ⊗ 1M)Lp(B(H)⊗M) and Lp(M; H c) = Lp(B(H)⊗M)(pv ⊗ 1M).
These are the row and column noncommutative Lp-spaces. Like the classical Lp-spaces, noncom-
mutative Lp-spaces form an interpolation scale with respect to the complex interpolation method:
For 1 ≤ p0 < p1 ≤ ∞ and 0 < η < 1, we have
(cid:0)Lp0(M), Lp1 (M)(cid:1)η = Lp(M) with equal norms,
p = 1−η
where 1
for the same indicies, we have
+ η
p1
p0
. Since Lp(M; H c) and Lp(M; H r) are 1-complemented subspaces of Lp(B(H)⊗M),
(cid:0)Lp0 (M; H c), Lp1(M; H c)(cid:1)η = Lp(M; H c) with equal norms.
Fourier analysis. Fourier multipliers will be one of the most important tools of this paper. Let
us give some Fourier multipliers that will be frequently used. They are all very well known in the
classical harmonic theory.
First, we recall the symbols of Littlewood-Paley decomposition on Rd. Fix a Schwartz function
ϕ on Rd satisfying:
(0.1)
2 ≤ ξ ≤ 2}.
2 < ξ < 2},
supp ϕ ⊂ {ξ : 1
ϕ > 0 on {ξ : 1
Pk∈Z ϕ(2−kξ) = 1,∀ ξ 6= 0.
For each k ∈ N, let ϕk be the function whose Fourier transform is equal to ϕ(2−k·), and let ϕ0
be the function whose Fourier transform is equal to 1 −Pk>0 ϕ(2−k·). Then {ϕk}k≥0 gives a
Littlewood-Paley decomposition on Rd such that
suppbϕk ⊂ {ξ ∈ Rd : 2k−1 ≤ ξ ≤ 2k+1},
∀ k ∈ N,
suppbϕ0 ⊂ {ξ ∈ Rd : ξ ≤ 2}
(0.2)
and that
(0.3)
∞Xk=0bϕk(ξ) = 1 ∀ ξ ∈ Rd.
Xk∈Zbϕk(ξ) = 1 ∀ ξ 6= 0.
The homogeneous counterpart of the above decomposition is given by { ϕk}k∈Z. This time, for
(0.4)
every k ∈ Z, these functions are given by bϕk(ξ) = ϕ(2−kξ). We have
The Bessel potential and the Riesz potential are J α = (1− (2π)−2∆)
2 and I α = (−(2π)−2∆)
2 ,
respectively. If α = 1, we will abbreviate J 1 as J and I 1 as I. We denote also Jα(ξ) = (1 + ξ2) α
on Rd and Iα(ξ) = ξα on Rd\{0}. Then Jα(ξ) and Iα(ξ) are the symbols of the Fourier multipliers
J α and I α, respectively.
Given a Banach space X, let S(Rd; X) be the space of X-valued rapidly decreasing functions
on Rd with the standard Fr´echet topology, and S′(Rd; X) be the space of continuous linear maps
from S(Rd) to X. All operations on S(Rd) such as derivations, convolution and Fourier transform
transfer to S′(Rd; X) in the usual way. On the other hand, Lp(Rd; X) naturally embeds into
S′(Rd; X) for 1 ≤ p ≤ ∞, where Lp(Rd; X) stands for the space of strongly p-integrable functions
α
α
2
4
R. Xia and X. Xiong
from Rd to X. By this definition, Fourier multipliers on Rd, in particular the Bessel and Riesz
potentials, extend to vector-valued tempered distributions in a natural way.
We denote by H σ
2 (Rd) the potential Sobolev space, consisting of all tempered distributions f
such that J σ(f ) ∈ L2(Rd). If σ > d
2 , we have
(cid:13)(cid:13)F−1(f )(cid:13)(cid:13)1 =Zs≤1(cid:12)(cid:12)F−1(f )(s)(cid:12)(cid:12)ds +Xk≥0Z2k<s≤2k+1(cid:12)(cid:12)F−1(f )(s)(cid:12)(cid:12)ds
22kσZ2k<s≤2k+1(cid:12)(cid:12)F−1(f )(s)(cid:12)(cid:12)2
2
ds(cid:17) 1
,
ds +Xk≥0
≤ C1(cid:16)Zs≤1(cid:12)(cid:12)F−1(f )(s)(cid:12)(cid:12)2
≤ C2(cid:13)(cid:13)f(cid:13)(cid:13)Hσ
kφ ∗ gkLp(Rd;X) ≤ kφk1kgkLp(Rd;X) ≤ C2kbφkHσ
2
where C1 and C2 are uniform constants. Therefore, ifbφ ∈ H σ
(0.5)
holds for any g ∈ Lp(Rd; X) with 1 ≤ p ≤ ∞. Here X is an arbitrary Banach space. Inequality
(0.5) indicates that functions in H σ
2 (Rd) are the symbols of bounded Fourier multipliers, even in
the vector-valued case.
2 kgkLp(Rd;X)
2 (Rd), the following Young's inequality
In the sequel, we will mainly consider the case X = L1(M) + M, i.e., consider operator-
valued functions or distributions on Rd. We will frequently use the following Cauchy-Schwarz type
inequality for operator-valued square function,
(0.6)
where φ : Rd → C and f : Rd → L1(M) + M are functions such that all integrations of the above
inequality make sense. We also require the operator-valued version of the Plancherel formula. For
sufficiently nice functions f : Rd → L1(M) + M, for example, for f ∈ L2(Rd) ⊗ L2(M), we have
(0.7)
(cid:12)(cid:12)ZRd
≤ZRd φ(s)2dsZRd f (s)2ds,
φ(s)f (s)ds(cid:12)(cid:12)2
ZRd f (s)2ds =ZRd bf (ξ)2dξ.
Throughout, we will use the notation A . B, which is an inequality up to a constant: A ≤ cB for
some constant c > 0. The relevant constants in all such inequalities may depend on the dimension
d, the test function Φ or p, etc, but never on the function f in consideration. The equivalence
A ≈ B will mean A . B and B . A simultaneously.
The layout of this paper is the following. In the next section, we briefly introduce the definition
of local Hardy spaces, and the main results in [36].
In section 2, we develop several Fourier
multiplier theorems: the first one is the inhomogeneous version of the Fourier multiplier theorem
proved in [40], fitted to local Hardy spaces; the second is a Hilbertian Fourier multiplier theorem,
in order to deal with the Lusin area square functions. In section 3, we give the definition of Triebel-
Lizorkin spaces, and some immediate properties. Section 4 is devoted to different characterizations
of Triebel-Lizorkin spaces. The proofs in this section are technical and tedious, based on Calder´on-
Zygmund theory and Fourier multiplier theorems. In the last section, we demonstrate the smooth
atomic decompositions of F α,c
p(Rd,M), and
then extend the result to general α by a similar argument.
(Rd,M): we begin with the space F 0,c
p (Rd,M) = hc
p
1. Operator-valued local Hardy spaces
Let us review the operator-valued local Hardy spaces studied in [36], and collect some of the
main results there that will be useful in this paper. We keep the following notation: (M, τ ) is
trace, and N = L∞(Rd)⊗M is equipped with the tensor
a von Neumann algebra with n.s.f.
trace; letters s, t are used to denote variables of Rd, while letters x, y are reserved for operators in
noncommutative Lp-spaces.
Let P be the Poisson kernel on Rd:
P(s) = cd
1
(s2 + 1)
d+1
2
Inhomogeneous Triebel-Lizorkin spaces
5
with cd the usual normalizing constant and s the Euclidean norm of s. Let
Pε(s) =
) = cd
1
εd P(
s
ε
ε
.
d+1
2
(s2 + ε2)
For any function f on Rd with values in L1(M) + M, its Poisson integral, whenever it exists, will
be denoted by Pε(f ):
The truncated Lusin area square function of f by
Pε(f )(s) =ZRd
sc(f )(s) =(cid:16)ZeΓ(cid:12)(cid:12) ∂
where eΓ is the truncated cone {(t, ε) ∈ Rd+1
p(Rd,M) = {f ∈ L1(M; Rc
hc
L2(Rd,
∂ε
dt
+
where the hc
p(Rd,M)-norm of f is defined by
Pε(s − t)f (t)dt,
(s, ε) ∈ Rd+1
+ .
εd−1(cid:17) 1
Pε(f )(s + t)(cid:12)(cid:12)2 dtdε
, s ∈ Rd,
2
1+td+1 ). For 1 ≤ p < ∞, define the column local Hardy space hc
d) : kfkhc
d) + L∞(M; Rc
p(Rd,M) to be
< ∞},
p
: t < ε < 1}. Denote by Rd the Hilbert space
kfkhc
The row local Hardy space hr
p = kf∗khc
with the norm kfkhr
p(Rd,M) = ksc(f )kLp(N ) + kP ∗ fkLp(N ).
p(Rd,M) is the space of all f such that f∗ ∈ hc
p. Moreover, define the mixture space hp(Rd,M) as follows:
p(Rd,M), equipped
equipped with the sum norm
hp(Rd,M) = hc
p(Rd,M) + hr
p(Rd,M) for 1 ≤ p ≤ 2
kfkhp(Rd,M) = inf{kgkhc
p + khkhr
p : f = g + h, g ∈ hc
p(Rd,M), h ∈ hr
p(Rd,M)},
and
hp(Rd,M) = hc
equipped with the intersection norm
p(Rd,M) ∩ hr
p(Rd,M) for 2 < p < ∞
The local analogue of the Littlewood-Paley g-function of f is defined by
kfkhp = max{kfkhc
p
,kfkhr
p}.
gc(f )(s) =(cid:0)Z 1
∂
∂ε
0
Pε(f )(s)2εdε(cid:1) 1
p ≈ kgc(f )kp + kP ∗ fkp
kfkhc
2 , s ∈ Rd.
It is proved in [36] that
for all 1 ≤ p < ∞.
The dual of hc
any cube Q ⊂ Rd, we denote its volume by Q. Let f ∈ L∞(M; Rc
is denoted by fQ := 1
1(Rd,M) is characterized as a local version of bmo space, defined as follows. For
d). The mean value of f over Q
QRQ f (s)ds. Set
kfkbmoc(Rd,M) = maxn sup
Q<1(cid:13)(cid:13)(
The local version of bmo spaces are defined as
(1.1)
1
QZQ f − fQ2dt)
1
2(cid:13)(cid:13)M
, sup
Q=1(cid:13)(cid:13)(ZQ f2dt)
1
2(cid:13)(cid:13)Mo.
bmoc(Rd,M) = {f ∈ L∞(M; Rc
d) : kfkbmoc < ∞}.
Define bmor(Rd,M) to be the space of all f ∈ L∞(M; Rr
d) such that kf∗kbmoc(Rd,M) is finite, with
the norm kfkbmor = kf∗kbmoc . And bmo(Rd,M) is defined as the intersection of bmoc(Rd,M)
and bmor(Rd,M), equipped with the intersection norm.
The above Hardy and bmo type spaces are local analogues of the spaces studied by Mei [18]. They
turn out to have similar properties with their non-local versions, such as duality and interpolation.
The following two theorems are quoted from [36].
Theorem 1.1. We have hc
is its conjugate index, then hc
1(Rd,M)∗ = bmoc(Rd,M) with equivalent norms. If 1 < p < 2 and q
p(Rd,M)∗ = hc
q(Rd,M) with equivalent norms.
6
R. Xia and X. Xiong
Theorem 1.2. Let 1 < p < ∞. We have
(1) (cid:0)bmoc(Rd,M), hc
(2) (cid:0)X, Y(cid:1) 1
p
1(Rd,M)(cid:1) 1
p
= hc
p(Rd,M).
= Lp(N ), where X = bmo(Rd,M) or L∞(N ), and Y = h1(Rd,M) or L1(N ).
Calder´on-Zygmund theory. The usual Calder´on-Zygmund operators which satisfy the Hormander
condition are not necessarily bounded on local Hardy spaces. In order to guarantee the bound-
p(Rd,M), an extra decay at infinity is imposed on
edness of a Calder´on-Zygmund operator on hc
the kernel in [36]. Let K ∈ S′(Rd; L1(M) + M) coincide on Rd \ {0} with a locally integrable
L1(M) + M-valued function. We define the left singular integral operator K c associated to K by
K(s − t)f (t)dt,
and the right singular integral operator K r associated to K by
K c(f )(s) =ZRd
K r(f )(s) =ZRd
f (t)K(s − t)dt.
Both K c(f ) and K r(f ) are well-defined for sufficiently nice functions f with values in L1(M)∩M,
for instance, for f ∈ S ⊗ (L1(M) ∩ M ).
0(Rd,M) denote the subspace of bmoc(Rd,M) consisting of compactly supported func-
tions. The extra decay of the kernel K given in [36] is condition (2) in the following lemma.
Let bmoc
Lemma 1.3. Assume that
(1) the Fourier transform of K is bounded: supξ∈Rd kbK(ξ)kM < ∞;
(2) K satisfies a size estimate: there exist C1 and ρ > 0 such that
kK(s)kM ≤
, ∀s ≥ 1;
C1
sd+ρ
(3) K has the Lipschitz regularity: there exist a constant C2 and γ > 0 such that
Then K c is bounded on hc
A similar statement also holds for K r and the corresponding row spaces.
kK(s − t) − K(s)kM ≤ C2
p(Rd,M) for 1 ≤ p < ∞ and from bmoc
s − td+γ
, ∀s > 2t.
0(Rd,M) to bmoc(Rd,M).
tγ
Characterizations. Next, we are going to present the characterizations of local Hardy spaces
obtained in [36], which will play an important role when studying the characterizations of Triebel-
Lizorkin spaces in this paper.
The main idea of these characterizations is to replace the Poisson kernel by good enough Schwartz
ε ) for
functions. Let Φ be a Schwartz function on Rd of vanishing mean, and set Φε(s) = ε−dΦ( s
positive ε. Φ is said to be nondegenerate if:
0
= 1,
(1.4)
(1.2)
(1.3)
dε
ε
Then there exists a Schwartz function Ψ of vanishing mean such that
∀ξ ∈ Rd \ {0} ∃ ε > 0 s.t. bΦ(εξ) 6= 0.
Z ∞
bΦ(εξ)bΨ(εξ)
∀ξ ∈ Rd \ {0} .
Furthermore, we can find two functions φ, ψ such that bφ,bψ ∈ H σ
bφ(ξ)bψ(ξ) = 1 −Z 1
0 bΦ(εξ)bΨ(εξ)
Φ(f )(s) =(cid:16)ZZeΓ Φε ∗ f (s + t)2 dtdε
εd+1(cid:17) 1
Φ(f )(s) =(cid:16)Z 1
ε (cid:17) 1
0 Φε ∗ f (s)2 dε
, s ∈ Rd.
For any f ∈ L1(M; Rc
functions of f associated to Φ by
d) + L∞(M; Rc
dε
ε
sc
gc
2
2
, s ∈ Rd,
2 (Rd), bφ(0) > 0,bψ(0) > 0 and
.
d), we define the local versions of the conic and radial square
Fix the four test functions Φ, Ψ, φ, ψ as above. The following theorem is proved in [36].
Inhomogeneous Triebel-Lizorkin spaces
7
p(Rd,M) if and only if sc
Theorem 1.4. Let 1 ≤ p < ∞ and φ, Φ be as above. For any f ∈ L1(M; Rc
Φ(f ) ∈ Lp(N ) and φ ∗ f ∈ Lp(N ) if and only if gc
f ∈ hc
φ ∗ f ∈ Lp(N ). If this is the case, then
(1.5)
Φ(f )kp + kφ ∗ fkp ≈ kgc
with the relevant constants depending only on d, Φ and φ.
Φ(f )kp + kφ ∗ fkp
p ≈ ksc
kfkhc
d) + L∞(M; Rc
d),
Φ(f ) ∈ Lp(N ) and
We have a discrete version of Theorem 1.4. The square functions sc
Φ and gc
Φ can be discretized
as follows:
f =
λj aj,
∞Xj=1
where the aj's are hc
the sense of distribution. We equip hc
1-atoms and λj ∈ C such thatP∞j=1 λj < ∞. The above series converges in
1,at(Rd,M) with the following norm:
kfkhc
1,at = inf{
λj : f =
λj aj; aj's are hc
1 -atoms, λj ∈ C}.
∞Xj=1
∞Xj=1
Similarly, we can define the row and mixture versions. The following theorem is also proved in
[36].
gc,D
Φ (f )(s) =(cid:16)Xj≥1
Φ (f )(s) =,(cid:16)Xj≥1
sc,D
2
,
Φj ∗ f (s)2(cid:17) 1
2djZB(s,2−j ) Φj ∗ f (t)2dt(cid:17) 1
2
.
Here Φj is the inverse Fourier transform of Φ(2−j·). This time, to get a resolvent of the unit on
Rd, we need to assume that Φ satisfies
Then adapting the proof of [32, Lemma V.6] , we can find a Schwartz function Ψ of vanishing mean
such that
∀ ξ ∈ Rd \ {0} ∃ 0 < 2a ≤ b < ∞ s.t. bΦ(εξ) 6= 0, ∀ ε ∈ (a, b].
(1.6)
+∞Xj=−∞bΦ(2−jξ)bΨ(2−jξ) = 1,
Again, there exist two functions φ and ψ such that bϕ,bψ ∈ H σ
∞Xj=1bΦ(2−jξ)bΨ(2−jξ) +bφ(ξ)bψ(ξ) = 1,
(1.7)
∀ξ ∈ Rd \ {0}.
2 (Rd), bφ(0) > 0,bψ(0) > 0 and
∀ξ ∈ Rd.
Now we fix the pairs (Φ, Ψ) and (φ, ψ) satisfying (1.6) and (1.7).
Theorem 1.5. Let φ and Φ be test functions as in (1.7) and 1 ≤ p < ∞. Then for any f ∈
L1(M; Rc
Φ (f ) ∈ Lp(N ) and φ∗ f ∈ Lp(N ) if and
only if gc,D
d) + L∞(M; Rc
Φ (f ) ∈ Lp(N ) and φ ∗ f ∈ Lp(N ). Moreover,
p(Rd,M) if and only if sc,D
d), f ∈ hc
kfkhc
p ≈ ksc,D
Φ (f )kLp(N ) + kφ ∗ fkp ≈ kgc,D
Φ (f )kp + kφ ∗ fkp
with the relevant constants depending only on d, Φ and φ.
Atomic decomposition. Finally, let us include the atomic decomposition of the local Hardy
space hc
1-atom associated with Q
is a function a ∈ L1(M; Lc
1(Rd,M). Let Q be a cube in Rd with Q ≤ 1. If Q = 1, an hc
• supp a ⊂ Q;
2(Rd)) such that
2 ≤ Q− 1
2 .
If Q < 1, we assume additionally:
• τ(cid:0)RQ a(s)2ds(cid:1) 1
• RQ a(s)ds = 0.
Let hc
1,at(Rd,M) be the space of all f admitting a representation of the form
8
R. Xia and X. Xiong
Theorem 1.6. We have hc
1,at(Rd,M) = hc
1(Rd,M) with equivalent norms.
Remark 1.7. In the above definition of atoms, we can replace the support of atoms Q by any
bounded multiple of Q.
2. Multiplier theorems
We are going to develop some Fourier multiplier theorems in this section. They can be viewed as
a special case of Calder´on-Zygmund theory, and will be used to investigate various square funtions
that characterize the Triebel-Lizorkin spaces. Our presentation follows closely the argument in
Section 4.1 of [40].
Recall again that ϕ is a fixed function satisfying (0.1), ϕ0 is the inverse Fourier transform of
1 −Pk>0 ϕ(2−k·), and ϕk is the inverse Fourier transform of ϕ(2−k·) when k > 0. Moreover, we
denote by ϕ(k) the Fourier transform of ϕk for every k ∈ N0 (N0 being the set of nonnegative
integers).
2.1. Global multipliers. Firstly, let us state the following homogeneous version of [40, Theo-
rem 4.1].
Theorem 2.1. Let σ ∈ R with σ > d
functions on Rd\{0} such that
2 . Assume that (φj )j∈Z and (ρj )j∈Z are two sequences of
and
supp φj ρj ⊂ {ξ : 2j−1 ≤ ξ ≤ 2j+1}, j ∈ Z
kφj(2j+k·)ϕkHσ
2 (Rd) < ∞.
sup
j∈Z
−2≤k≤2
Let 1 < p < ∞. Then for any f ∈ S′(Rd; L1(M) + M), we have
kφj(2j+k·)ϕkHσ
22jα φj ∗ ρj ∗ f2)
1
where the constant depends on p, σ, d and ϕ.
(cid:13)(cid:13)(Xj∈Z
(cid:13)(cid:13)(Xj≥K
2(cid:13)(cid:13)p . sup
j∈Z
−2≤k≤2
1
2(cid:13)(cid:13)p . sup
j∈Z
−2≤k≤2
2(cid:13)(cid:13)(Xj∈Z
22jαρj ∗ f2)
,
1
2(cid:13)(cid:13)p
2(cid:13)(cid:13)(Xj≥K
,
1
2(cid:13)(cid:13)p
Proof. Without loss of generality, we may take α = 0. It suffices to show that for any integer K,
(2.1)
φj ∗ ρj ∗ f2)
kφj(2j+k·)ϕkHσ
ρj ∗ f2)
with the relevant constant independent of K ∈ Z. To this end, we set
By easy computation, we have
ψj−K = φj(2K·),
ηj−K = ρj(2K·), and bg = bf (2K·).
supp ψjηj ⊂ {ξ : 2j−1 ≤ ξ ≤ 2j+1}, ∀ j ≥ 0,
φj ∗ ρj ∗ f = 2dK ψj−K ∗ ρj−K ∗ g(2K·).
φj ∗ ρj ∗ f2)
ρj ∗ f2)
(cid:13)(cid:13)(Xj≥K
(cid:13)(cid:13)(Xj≥K
(p−1)dK
p
(p−1)dK
p
1
2(cid:13)(cid:13)p = 2
2(cid:13)(cid:13)p = 2
1
(cid:13)(cid:13)(Xj≥0
(cid:13)(cid:13)(Xj≥0
.
1
2(cid:13)(cid:13)p
ψj ∗ ηj ∗ g2)
ηj ∗ g2)
.
1
2(cid:13)(cid:13)p
and
This ensures
(2.2)
Similarly,
(2.3)
Moreover, since ψj(2j+k·) = φj+K (2j+k+K·), we have
2 = sup
j≥K
−2≤k≤2
≤ sup
j∈Z
−2≤k≤2
kψj(2j+k·)ϕkHσ
−2≤k≤2
sup
j≥0
(2.4)
kφj(2j+k·)ϕkHσ
2
kφj(2j+k·)ϕkHσ
2
.
Inhomogeneous Triebel-Lizorkin spaces
9
Now applying [40, Theorem 4.1] to ψj, ρj and g defined above, we obtain
(cid:13)(cid:13)(Xj≥0
ψj ∗ ηj ∗ g2)
1
2kp . sup
j≥0
−2≤k≤2
kψj(2j+k·)ϕkHσ
ηj ∗ g2)
1
2 kp.
2 ((cid:13)(cid:13)(Xj≥0
Putting (2.2), (2.3) and (2.4) into this inequality, we then get (2.1), which yields Theorem 2.1 by
approximation.
(cid:3)
Theorem 2.1 is developed to deal with the multiplier problem of square functions, and also the
p(Rd,M) by virtue of their characterizations (see [38]). In
multiplier problem of the Hardy spaces Hc
order to deal with the corresponding problems on the inhomogeneous versions of square functions
or Hardy spaces, we need the following global version of Theorem 2.1. The main difference is that
in the inhomogeneous case, we need a careful analysis of the convolution kernel near the origin.
Theorem 2.2. Let 1 < p < ∞, α ∈ R and σ > d
sequences of functions on Rd such that
2 . Assume that (φj)j≥0 and (ρj)j≥0 are two
supp (φj ρj) ⊂ {ξ ∈ Rd : 2j−1 ≤ ξ ≤ 2j+1}, j ∈ N,
supp (φ0ρ0) ⊂ {ξ ∈ Rd : ξ ≤ 2},
and
(2.5)
kφj(2j+k·)ϕkHσ
2 (Rd) < ∞ and kφ0(ϕ(0) + ϕ(1))kHσ
2 (Rd) < ∞.
sup
j≥1
−2≤k≤2
Then for any L1(M) + M-valued distribution f ,
22jα φj ∗ ρj ∗ f2)
(cid:13)(cid:13)(Xj≥0
kφj(2j+k·)ϕkHσ
2
2(cid:9)
,kφ0(ϕ(0) + ϕ(1))kHσ
Proof. This theorem follows easily from its homogeneous version, i.e., Theorem 2.1. Indeed, we
where the constant depends only on p, σ, d and ϕ.
can divide(cid:13)(cid:13)(Pj≥0 22jα φj ∗ ρj ∗ f2)
22jα φj ∗ ρj ∗ f2)
1
(cid:13)(cid:13)(Xj≥0
(cid:13)(cid:13)(Xj≥1
22jα φj ∗ ρj ∗ f2)
,
1
2(cid:13)(cid:13)p
j≥1
−2≤k≤2
22jαρj ∗ f2)
1
1
2(cid:13)(cid:13)p . max(cid:8) sup
·(cid:13)(cid:13)(Xj≥0
2(cid:13)(cid:13)p into two parts
2(cid:13)(cid:13)p ≈(cid:13)(cid:13)(Xj≥1
2(cid:13)(cid:13)p . sup
−2≤k≤2
j≥1
1
1
22jα φj ∗ ρj ∗ f2)
2(cid:13)(cid:13)p + k φ0 ∗ ρ0 ∗ fkp
2(cid:13)(cid:13)(Xj≥1
2(cid:13)(cid:13)p
φ0 ∗ ρ0 ∗ f = F−1(cid:0)φ0(ϕ(0) + ϕ(1))(cid:1) ∗ ρ0 ∗ f.
kφj(2j+k·)ϕkHσ
ρj ∗ f2)
.
1
and treat them separately. Applying Theorem 2.1 to the sequences (φj )j∈Z, (ρj)j∈Z with φj = 0
and ρj = 0 for j ≤ 0, we get the estimate of the first term on the right hand side. The result is
The second term k φ0 ∗ ρ0 ∗ fkp is also easy to handle. By the support assumption on φ0ρ0, we have
Hence,
k φ0 ∗ ρ0 ∗ fkp ≤ kF−1(cid:0)φ0(ϕ(0) + ϕ(1))(cid:1)k1k ρ0 ∗ fkp . kφ0(ϕ(0) + ϕ(1))kHσ
The assertion is proved.
2 k ρ0 ∗ fkp.
(cid:3)
2.2. Hilbert-valued multipliers. In fact, both theorems above deal with Fourier multipliers
acting on Hilbert-valued noncommutative Lp spaces (the Hilbert space being ℓ2). In this subsection
titled "Hilbert-valued multipliers", our target is to extend Theorem 2.2 to the general case where
ℓ2 is replaced with more complicated Hilbert spaces. Assume that we have a sequence of Hilbert
spaces Hj for every j ∈ N0, and denote H = ⊕∞j=0Hj. Then an element f ∈ Lp(N ; H c) has the
form f = (fj)j≥0 with fj ∈ Lp(N ; H c
j ) for every j. In this case, it still makes sense to consider the
action of the Calder´on-Zygmund operator k = ( φj )j≥0.
10
R. Xia and X. Xiong
Since it will be frequently used in the following, we introduce an elementary inequality (see [40,
Lemma 4.2]):
(2.6)
kf gkHσ
2 (Rd;ℓ2) ≤ kfkHσ
2 (Rd;ℓ2)ZRd
(1 + s2)σF−1(g)(s)ds,
where σ > d
2 , and the functions f : Rd → ℓ2 and g : Rd → C satisfy
f ∈ H σ
2 (Rd; ℓ2) and ZRd
(1 + s2)σF−1(g)(s)ds < ∞.
2 (Rd; ℓ2) is the ℓ2-valued Potential Sobolev space of order σ. Note also that ℓ2 could be an
Here H σ
ℓ2-space on an arbitrary index set, depending on the problems in consideration.
The following lemma is an analogue of Lemma 4.3 in [40]. The main difference is that in order
to get a Calder´on-Zygmund operator which is bounded on local Hardy or bmo spaces, we need to
consider the Littlewood-Paley decomposition covering the origin.
Lemma 2.3. Let φ = (φj )j≥0 be a sequence of continuous functions on Rd, viewed as a function
from Rd to ℓ2. For σ > d
2 , we assume that
(2.7)
kφk2,σ
def= max(cid:8) sup
k≥1 kφ(2k·)ϕkHσ
2 (Rd;ℓ2), kφϕ(0)kHσ
2 (Rd;ℓ2)(cid:9) < ∞.
precisely,
Let k = (kj)j≥0 with kj = F−1(φj). Then k is a Calder´on-Zygmund kernel with values in ℓ2, more
(1) kbkkL∞(Rd;ℓ2) . kφk2,σ;
(2) Rs≥ 1
(3) supt∈RdRs>2t kk(s − t) − k(s)kℓ2 ds . kφk2,σ.
The relevant constants depend only on ϕ, σ and d.
2 kk(s)kℓ2 ds . kφk2,σ;
Proof. For any ξ ∈ Rd and k ≥ 1, by the Cauchy-Schwarz inequality, we have
kφ(2kξ)ϕ(ξ)kℓ2 =(cid:13)(cid:13)Z F−1(φ(2k·)ϕ)(s)e−2πis·ξds(cid:13)(cid:13)ℓ2
2 (Rd;ℓ2)(Z (1 + s2)−σds)
≤ kφ(2k·)ϕkHσ
1
2 . kφk2,σ.
In other words, we have kφϕ(2−k·)kL∞(Rd;ℓ2) . kφk2,σ. Likewise, kφϕ(0)kL∞(Rd;ℓ2) . kφk2,σ also
holds. Thus, by (0.2) and (0.3), we easily deduce that kbkkL∞(Rd;ℓ2) . kφk2,σ.
To show the third property of k, we decompose φ into
φ =Xk≥0
φϕ(k).
The convergence of the above series can be proved by a limit procedure of its partial sums, which
is quite formal. By (0.2) and (0.3), we write
φϕ(k) = φ(ϕ(k−1) + ϕ(k) + ϕ(k+1))ϕ(k) def= φ(k)ϕ(k),
k ≥ 0.
Here we make the convention that ϕ(k) = 0 if k < 0. Then for s ∈ Rd,
F−1(φϕ(k))(s) = F−1(φ(k)) ∗ F−1(ϕ(k))(s) = 2kdF−1(φ(k)(2k·)) ∗ F−1(ϕ)(2ks),
k ≥ 0.
By (2.6), we have
(ZRd
(1 + 2ks2)σkF−1(φϕ(k))(s)k2
ℓ2
ds)
1
2 . 2
kd
2 kφ(k)(2k·)kHσ
2 (Rd;ℓ2).
Inhomogeneous Triebel-Lizorkin spaces
11
Notice that if k ≥ 1, we have ϕ(k)(2k·) = ϕ. Thus, if k ≥ 2,
kφ(k)(2k·)kHσ
2 (Rd;ℓ2) ≤
.
=
1Xj=−1
1Xj=−1
1Xj=−1
kφ(2k·)ϕ(k−j)(2k·)kHσ
2 (Rd;ℓ2)
kφ(2k−j·)ϕ(k−j)(2k−j·)kHσ
2 (Rd;ℓ2)
kφ(2k−j·)ϕkHσ
2 (Rd;ℓ2) ≤ 3kφk2,σ.
For k = 0, 1, we treat φ(k)(2k·) in the same way:
kφ(1)(2·)kHσ
kφ(0)kHσ
2 (Rd;ℓ2) . kφϕ(0)kHσ
2 (Rd;ℓ2) . kφϕ(0)kHσ
2 (Rd;ℓ2) + kφ(2·)ϕkHσ
2 (Rd;ℓ2) + kφ(2·)ϕkHσ
2 (Rd;ℓ2) + kφ(4·)ϕkHσ
2 (Rd;ℓ2) ≤ 3kφk2,σ.
2 (Rd;ℓ2) ≤ 3kφk2,σ;
In summary, we obtain
(ZRd
(1 + 2ks2)σkF−1(φϕ(k))(s)k2
ℓ2 ds)
1
2 . 2
kd
2 kφk2,σ.
Thus, by the Cauchy-Schwarz inequality, for any t ∈ Rd \ {0} and k ≥ 0, we have
(1 + 2ks2)−σds)
kF−1(φϕ(k))(s)kℓ2 ds . 2
kd
1
2
(2.8)
Zs>t
2 kφk2,σ(Zs>t
2 −σkφk2,σ.
d
. (2kt)
Consequently,
Zs>2t
kF−1(φϕ(k))(s) − F−1(φϕ(k))(s − t)kℓ2 ds . (2kt)
d
2 −σkφk2,σ.
We notice that d
another estimate
2 − σ < 0, so the estimate above is good only when 2kt ≥ 1. Otherwise, we need
where et(ξ) = e2πiξ·t. Thus,
F−1(φϕ(k))(s) − F−1(φϕ(k))(s − t)
= F−1(φ(k)ϕ(k)(1 − et))(s)
= 2kdF−1(φ(k)(2k·)) ∗ [F−1(ϕ) − F−1(ϕ)(· − 2kt)](2ks),
(ZRd
(1 + 2ks2)σkF−1(φϕ(k))(s) − F−1(φϕ(k))(s − t)k2
2 kφk2,σ2ktZ (1 + s2)σF−1(ϕ)(s − θ2kt)ds
2 kφk2,σ2kt(Z J σ[ϕ(s)e2πis·θ2k t]2ds)
2 kφk2,σ2kt,
. 2
. 2
. 2
ℓ2
kd
kd
kd
1
2
1
2
ds)
where θ ∈ [0, 1]. Then as before, for 2kt < 1, we have
kF−1(φϕ(k))(s) − F−1(φϕ(k))(s − t)kℓ2 ds . 2ktkφk2,σ.
Combining the previous estimates, we obtain
sup
Zs>2t
t∈RdZs>2t
t∈RdXk≥0Zs>2t
t∈RdXk≥0
. kφk2,σ sup
≤ sup
kk(s − t) − k(s)kℓ2 ds
kF−1(φϕ(k))(s) − F−1(φϕ(k))(s − t)kℓ2 ds
min(2kt, (2kt)
d
2 −σ) . kφk2,σ.
12
R. Xia and X. Xiong
Finally, the second estimate of k can be deduced from (2.8) by letting t = 1
2 :
Zs≥ 1
2
kk(s)kℓ2 ds ≤Xk≥0Zs≥ 1
2
(2k−1)
≤Xk≥0
kF−1(φϕ(k))(s)kℓ2 ds
d
2 −σkφk2,σ . kφk2,σ.
The proof is complete.
(cid:3)
We keep the notation H = ⊕∞j=0Hj. By the above lemma, we can apply the (local) Calder´on-
Zygmund theory introduced in section 1, to deduce the following lemma:
Lemma 2.4. Let 1 < p < ∞ and φ = (φj)j≥0 be a sequence of continuous functions on Rd
satisfying (2.7). For any f = (fj)j≥0 ∈ Lp(N ; H c), we have
k( φj ∗ fj)j≥0kLp(N ;Hc) . kφk2,σk(fj)j≥0kLp(N ;Hc),
where the relevant constant depends only on ϕ, σ, p and d.
Proof. Consider k as a diagonal matrix with diagonal entries (kj)j≥0 determined by bkj = φj
and f = (fj)j≥0 as a column matrix. The associated Calder´on-Zygmund operator is defined on
Lp(B(H)⊗N ) by
k(f )(s) =ZRd
k(s − t)f (t)dt.
Now it suffices to show that k is a bounded operator on Lp(N ; H c).
We claim that k is bounded from L∞(N ; H c) into bmo(Rd, B(H)⊗M). Put K(s) = k(s) ⊗
1M ∈ B(H)⊗M, for any s ∈ Rd. Then we have kk(s)kℓ2 ≥ kk(s)kℓ∞
= kK(s)kB(H)⊗M and
kfkL∞(N ;Hc) = kfkB(H)⊗N . Thus, the claim is equivalent to saying that K is bounded from
L∞(N ; H c) into bmo(Rd, B(H)⊗M), if we regard L∞(N ; H c) as a subspace of B(H)⊗N .
First, we show that K is bounded from L∞(N ; H c) into bmoc(Rd, B(H)⊗M). Let Q be a cube
in Rd centered at c. We decompose f as f = g + h with g = f 1 eQ, where eQ = 2Q is the cube which
has the same center as Q and twice the side length of Q. Set
a =ZRd\ eQ
K(c − t)f (t)dt.
Then
where
whence
K(f )(s) − a = K(g)(s) +Z [K(s − t) − K(c − t)]h(t)dt.
Thus, for Q such that Q < 1, we have
1
A =
1
1
B =
QZQ K(f ) − a2ds ≤ 2(A + B),
QZQ K(g)2ds,
QZQ Z [K(s − t) − K(c − t)]h(t)dt2ds.
QA ≤Z bK(ξ)bg(ξ)2dξ =Z bg(ξ)∗bK(ξ)∗bK(ξ)bg(ξ)dξ ≤Z kbK(ξ)k2
.Z kbk(ξ)k2
≤ eQkφk2
2,σZ eQ f (s)2ds
2,σkfk2
ℓ2bg(ξ)2dξ . kφk2
L∞(N ;Hc),
2,σkfk2
B(H)⊗N
= eQkφk2
kAkB(H)⊗M . kφk2
2,σkfk2
L∞(N ;Hc).
The term A is easy to estimate. By Lemma 2.3 and the Plancherel formula (0.7),
B(H)⊗Mbg(ξ)2dξ
Inhomogeneous Triebel-Lizorkin spaces
13
To estimate B, writing h = (hj)j≥0, by Lemma 2.3, we get
(cid:12)(cid:12)Z [K(s − t) − K(c − t)]h(t)dt(cid:12)(cid:12)2
.ZRd\ eQ kK(s − t) − K(c − t)kB(H)⊗MdtZRd\ eQ kK(s − t) − K(c − t)kB(H)⊗Mh(t)2dt
.ZRd\ eQ kk(s − t) − k(c − t)kℓ2 dtZRd\ eQ kk(s − t) − k(c − t)kℓ2h(t)2dt
. kφk2
L∞(N ;Hc).
2,σkfk2
2,σkfk2
. kφk2
B(H)⊗N
Hence,
kBkB(H)⊗M ≤
B(H)⊗M
Combining the previous inequalities, we deduce that, for any Q < 1
ds . kφk2
2,σkfk2
L∞(N ;Hc).
1
1
QZQ(cid:13)(cid:13)Z [K(s − t) − K(c − t)]h(t)dt(cid:13)(cid:13)2
QZQ K(f ) − a2ds)
(cid:13)(cid:13)(cid:13)(
QZQ K(f )2ds ≤ 2
2(cid:13)(cid:13)(cid:13)B(H)⊗M
QZQ K(g)2ds + 2
1
1
1
. kφk2,σkfkL∞(N ;Hc).
1
QZQ K(h)2ds.
Now we consider the case when Q = 1. We have
The first term on the right hand side of the above inequality is equal to the term A, so it remains
2 . Then by
to estimate the second term. When t ∈ Rd\eQ, s ∈ Q and Q = 1, we have s − t ≥ 1
(2) in Lemma 2.3 and the Cauchy-Schwarz inequality (0.6), we easily deduce that
K(h)(s)2 =(cid:12)(cid:12)Z K(s − t)h(t)dt(cid:12)(cid:12)2
≤ZRd\ eQ kK(s − t)kB(H)⊗MdtZRd\ eQ kK(s − t)kB(H)⊗Mh(t)2dt
L∞(N ;Hc)(ZRd\ eQ kk(s − t)kℓ2 dt)2
. kfk2
. kφk2
2,σkfk2
Thus, we have, for any Q = 1,
L∞(N ;Hc).
Next we show that K is bounded from L∞(N ; H c) into bmor(Rd, B(H)⊗M). We still use the
QZQ K(f )2ds(cid:1) 1
(cid:13)(cid:13)(cid:13)(cid:0) 1
2(cid:13)(cid:13)(cid:13)B(H)⊗M
. kφk2,σkfkL∞(N ;Hc).
1
1
A′ =
where
Therefore, K is bounded from L∞(N ; H c) into bmoc(Rd, B(H)⊗M).
same decomposition f = g + h, then we obtain
QZQ [K(f ) − a]∗2ds ≤ 2(A′ + B′),
QZQ K(g)∗2ds,
QZQ(cid:12)(cid:12)Z [(K(s − t) − K(c − t))h(t)]∗dt(cid:12)(cid:12)2
QZQ(cid:13)(cid:13)Z [(K(s − t) − K(c − t))h(t)]∗dt(cid:13)(cid:13)2
QZQ(cid:13)(cid:13)Z [K(s − t) − K(c − t)]h(t)dt(cid:13)(cid:13)2
The estimate of B′ can be reduced to that of B. Indeed,
kB′kB(H)⊗M ≤
B′ =
=
1
1
1
. kφk2
2,σkfk2
L∞(N ;Hc).
ds.
ds
B(H)⊗M
ds
B(H)⊗M
14
R. Xia and X. Xiong
However, for A′, we need a different argument. A′ can be viewed as a bounded operator on
H ⊗ L2(M). So
kA′kB(ℓ2)⊗M = sup
b {
1
QZQ kk(g)(s) bk2
H⊗L2(M)ds},
where the supremum runs over all b in the unit ball of H ⊗ L2(M). By the Plancherel formula
(0.7), we have
ZQ kk(g)(s) bk2
Let diag(fj)j be the diagonal matrix in B(H)⊗N with entries in B(Hj)⊗N . By the Cauchy-
Schwarz inequality, the Plancherel formula (0.7) and Lemma 2.3, we continue the estimate above
as
H⊗L2(M)ds =ZQhk(g)(s) b, k(g)(s) biH⊗L2(M)ds
≤Z hbk(ξ)bg(ξ) b,bk(ξ)bg(ξ) biH⊗L2(M)dξ.
ℓ2Z hbg(ξ) b,bg(ξ) biH⊗L2(M)dξ
ξ kbk(ξ)k2
2,σZ eQ k diag(fj)j(s) bk2
. kφk2
2,σk diag(fj)jk2
. Qkφk2
2,σkfk2
≤ Qkφk2
L∞(N ;Hc),
H⊗L2(M)ds
B(H)⊗Nkbk2
H⊗L2(M)
Z hbk(ξ)bg(ξ) b,bk(ξ)bg(ξ) biH⊗L2(M)dξ ≤ sup
whence,
1
L∞(N ;Hc).
Following the estimate of 1
kA′kB(ℓ2)⊗M . kφk2
QZQ K(f )∗2ds ≤ 2A′ + 2
≤ 2A′ + 2
= 2A′ + 2
2,σkfk2
QRQ K(f )(s)2ds, we get, when Q = 1,
QZQ K(h)∗2ds
QZQ kK(h)∗k2
QZQ kK(h)k2
2,σkfk2
Therefore, K is bounded from L∞(N ; H c) into bmor(Rd, B(H)⊗M).
In summary, we have proved that k is bounded from L∞(N ; H c) into bmo(Rd, B(H)⊗M). It
is also clear that k is bounded from L2(N ; H c) into L2(B(H)⊗N ), then by the interpolation in
Theorem 1.2, k is bounded from Lp(N ; H c) into Lp(B(H)⊗N ) for any 2 ≤ p < ∞. The case
1 < p < 2 is obtained by duality.
(cid:3)
L∞(N ;Hc).
. kφk2
B(H)⊗M
B(H)⊗M
1
1
1
ds
ds
Note that when all Hj degenerate to one dimensional Hilbert space, then H = ℓ2, the above
lemma gives a sufficient condition for (φj )j≥0 being a bounded Fourier multiplier on Lp(N ; ℓc
2). So
we can also use Lemmas 2.3 and 2.4 to prove Theorem 2.2 by an argument similar to the proof of
[40, Theorem 4.1]; details are left to the reader. But here our target is to extend Theorem 2.2 to
a more general setting.
Theorem 2.5. Let p, α, σ, (φj)j≥0 and (ρj )j≥0 be the same as in Theorem 2.2. Then, for any
f ∈ S′(Rd; L1(M) + M),
2j(2α+d)ZB(0,2−j ) φj ∗ ρj ∗ f (· + t)2dt)
kφj(2j+k·)ϕkHσ
2
2(cid:9)
,kφ0(ϕ(0) + ϕ(1))kHσ
−2≤k≤2
2j(2α+d)ZB(0,2−j ) ρj ∗ f (· + t)2dt)
(cid:13)(cid:13)(cid:13)(Xj≥0
. max(cid:8) sup
·(cid:13)(cid:13)(cid:13)(Xj≥0
j≥1
1
2(cid:13)(cid:13)(cid:13)p
2(cid:13)(cid:13)(cid:13)p
1
,
Inhomogeneous Triebel-Lizorkin spaces
15
where the constant depends only on p, σ, d and ϕ.
Proof. Set Hj = L2(cid:0)B(0, 2−j), 2jddt(cid:1) and H = ⊕∞j=0Hj. So we have
2j(2α+d)ZB(0,2−j ) φj ∗ ρj ∗ f (· + t)2dt)
1
(cid:13)(cid:13)(Xj≥0
Let
2(cid:13)(cid:13)p = k(2jα φj ∗ ρj ∗ f (· + ·))jkLp(N ;Hc).
ζj = φj(ϕ(j−1) + ϕ(j) + ϕ(j+1)), j ≥ 2,
ζ1 = φ1(ϕ + ϕ(1) + ϕ(2)),
ζ0 = φ0(ϕ(0) + ϕ) and ζj = 0 if j < 0.
By the support assumption on φjρj, we have that φjρj = ζjρj. So for any f ∈ S′(Rd; L1(M)+M),
φj ∗ ρj ∗ f = ζj ∗ ρj ∗ f, j ∈ N0.
Now we show that ζ = (ζj )j≥0 satisfies (2.7) with ζ instead of φ. Indeed, by the support assumption
of ϕ, the sequence ζ(2k·)ϕ = (cid:0)ζj(2k·)ϕ(cid:1)j≥0 has at most five nonzero terms of indices j with
k − 2 ≤ j ≤ k + 2. Thus for any k ∈ N0,
kζ(2k·)ϕkHσ
2 (Rd;ℓ2) ≤
k+2Xj=k−2
kζj(2k·)ϕkHσ
2
.
Moreover, by (2.6), we have
kζj(2k·)ϕkHσ
Therefore, the condition (2.5) yields
2
. kφj(2k·)ϕkHσ
2
,
k − 2 ≤ j ≤ k + 2.
2 (Rd;ℓ2) . sup
j≥1
−2≤k≤2
kφj(2j+k·)ϕkHσ
2 + kφ0(ϕ(0) + ϕ(1))kHσ
2
< ∞,
where the relevant constant depends only on σ, ϕ and d. In a similar way, we have
kζjϕ(0)kHσ
2
kφj(2j+k·)ϕkHσ
2 + kφ0(ϕ(0) + ϕ(1))kHσ
2
< ∞.
. sup
j≥1
−2≤k≤2
Now applying Lemma 2.4 to fj = 2jα ρj∗f (·+·), and ζj instead of φj , we conclude the theorem. (cid:3)
The above theorem will be useful when we consider the conic square function characterizations
of local Hardy spaces and inhomogeneous Triebel-Lizorkin spaces in section 4.
p. Note that both Theorem 2.2 and Theorem 2.5 do not deal with the case
p with 1 ≤ p ≤ 2 in the
2.3. Multipliers on hc
p = 1. So we include the corresponding Fourier multiplier results for hc
following. When the Hilbert space H degenerates to ℓ2, we have
Lemma 2.6. Let 1 ≤ p ≤ 2 and φ = (φj)j≥0 be a sequence of continuous functions on Rd satisfying
(2.7). For f ∈ hc
p(Rd,M),
sup
k≥1 kζ(2k·)ϕkHσ
2 (Rd;ℓ2) ≤ X0≤j≤2
kζϕ(0)kHσ
The relevant constant depends only on ϕ, σ and d.
Proof. Now we view k = (kj)j≥0 = ( φj )j≥0 as a column matrix and the associated Calder´on-
Zygmund operator k is defined on Lp(N ):
(cid:13)(cid:13)(Xj≥0
φj ∗ f2)
1
2(cid:13)(cid:13)p . kφk2,σkfkhc
.
p
k(f )(s) =ZRd
k(s − t)f (t)dt,
∀s ∈ R.
Thus k maps function with values in Lp(M) to sequence of functions. Then we have to show
that k is bounded from hc
2) for 1 ≤ p ≤ 2. The case p = 2 is trivial, so by
interpolation, it suffices to consider the case p = 1. To prove that k is bounded from hc
1(Rd,M) to
L1(N ; ℓc
2), passing to the dual spaces, it is equal to proving that the adjoint of k is bounded from
p(Rd,M) to Lp(N ; ℓc
16
R. Xia and X. Xiong
L∞(N ; ℓc
sequence f = (fj)j≥0 (viewed as a column matrix), the adjoint of k is defined by
2) to bmoc(Rd,M). We keep all the notation in the proof of Lemma 2.4. For any finite
k∗(f )(s) =ZRdXj ekj(s − t)fj(t)dt,
whereek(s) = k(−s)∗ (so it is a row matrix). Put eK(s) =ek(s)⊗1M. In this case, keK(f )kbmoc(Rd,M) =
keK(f )kbmoc(Rd,B(ℓ2)⊗M). Then we apply the estimates used in Lemma 2.4 by replacing K with
eK. It follows that k∗ is bounded from L∞(N ; ℓc
2) into bmoc(Rd,M), so the desired assertion is
(cid:3)
The next theorem is a complement of Theorem 2.2 for the case p = 1, which relies heavily on
proved.
1(Rd,M) given in Theorem 1.5.
the characterization of hc
Theorem 2.7. We keep the assumption in Theorem 2.2. Assume additionally that for any j ≥ 1,
ρj = ρ(2−j·) for some Schwartz function ρ with supp ρ ⊂ {ξ : 2−1 ≤ ξ ≤ 2} and ρ(ξ) > 0 for
any 2−1 < ξ < 2, and that supp ρ0 ⊂ {ξ : ξ ≤ 2} and ρ0(ξ) > 0 for any ξ < 2. Then for
f ∈ S′(Rd; L1(M) + M), we have
22jα φj ∗ ρj ∗ f2)
(cid:13)(cid:13)(Xj≥0
kφj(2j+k·)ϕkHσ
2
2(cid:9)
,kφ0(ϕ(0) + ϕ(1))kHσ
1
2(cid:13)(cid:13)1 . max(cid:8) sup
·(cid:13)(cid:13)(Xj≥0
j≥1
−2≤k≤2
22jαρj ∗ f2)
.
1
2(cid:13)(cid:13)1
∀ξ ∈ Rd.
Proof. By the assumptions of ρ and ρ0, we can select a Schwartz functioneρ with the same properties
as ρ and a Schwartz function eρ0 satisfying the same conditions as ρ0, such that
Let Ψj = (I−αρ)(2−j·), eΨj = (Iαρ)(2−j·) for j ≥ 1 and Ψ0 = J−αρ0, eΨ0 = Jαρ0. We have
Applying Theorem 1.5 (the equivalence kgc,D
functions in the above identity, we get
kgkhc
∞Xj=1
ρ(2−jξ)eρ(2−jξ) + ρ0(ξ)eρ0(ξ) = 1,
∞Xj=1
Ψj(ξ)eΨj(ξ) + Ψ0(ξ)eΨ0(ξ) = 1,
2(cid:13)(cid:13)1
Φ (f )kp + kφ ∗ fkp ≈ kfkhc
p) to g = J αf with the text
Ψj ∗ g2)
∀ ξ ∈ Rd.
.
1
Now let us show the following equivalence:
1
Ψj ∗ g2)
It is easy to see that Ψ0 ∗ g = ρ0 ∗ f and 2jα ρj ∗ f = Ψj ∗ I αf , so it suffices to prove
(2.9)
.
1
First, let us consider the case α ≥ 0. By [31, Lemma 3.2.2], there exists a finite measure µα on Rd
such that
(cid:13)(cid:13)(Xj≥0
(cid:13)(cid:13)(Xj≥1
(cid:13)(cid:13)(Xj≥1
1
Ψj ∗ I αf2)
22jαρj ∗ f2)
Ψj ∗ J αf2)
1 ≈(cid:13)(cid:13)(Xj≥0
2(cid:13)(cid:13)1 ≈(cid:13)(cid:13)(Xj≥0
2(cid:13)(cid:13)1 ≈(cid:13)(cid:13)(Xj≥1
ξα =bµα(ξ)(1 + ξ2)
Ψj ∗ I αf = µα ∗ Ψj ∗ J αf,
.(cid:13)(cid:13)(Xj≥1
2(cid:13)(cid:13)1
Ψj ∗ I αf2)
2 =bνα(ξ) + ξ−αbλα(ξ).
(1 + ξ2)− α
2 .
∀ j ≥ 1.
Ψj ∗ J αf2)
α
1
1
.
2(cid:13)(cid:13)1
2(cid:13)(cid:13)1
.
1
2(cid:13)(cid:13)1
Thus, we have
This implies that
Then, we move to the case α < 0. Also by [31, Lemma 3.2.2], there exist two finite measures να
and λα on Rd such that
Inhomogeneous Triebel-Lizorkin spaces
17
Let ( ϕk)k∈Z be the homogeneous resolution of the unit defined in (0.4). It follows that
2
ξ−α Xk≥0
(1 + ξ2)− α
ξ−α Xk≥0
ϕk(ξ) = bνα(ξ)
Thus, by the support assumption of bρ, we have
Ψj ∗ I αf = ωα ∗ Ψj ∗ J αf,
ϕk(ξ) +bλα(ξ)Xk≥0
ϕk(ξ).
with
ωα = να ∗Xk≥0
F−1(Iα ϕk) + λα ∗ F−1(Xk≥0
ϕk).
Both F−1(Pk≥0 ϕk) and Pk≥0 F−1(Iα ϕk) are finite measures. Since Pk≥0 ϕk = 1 −Pk<0 ϕk,
andPk<0 ϕk is a Schwartz function, we know that F−1(Pk≥0 ϕk) = δ0−F−1(Pk<0 ϕk) is a finite
measure, where δ0 denotes the Dirac measure at the origin. Moreover, it is known in [40, Lemma
3.4] that kF−1(Iα ϕk)k1 . 2kα. Then we have
kF−1(Xk≥0
2kα < ∞.
Therefore, ωα is a finite measure on Rd. Thus,
1
Iα ϕk)k1 .Xk≥0
2(cid:13)(cid:13)1 .(cid:13)(cid:13)(Xj≥1
2(cid:13)(cid:13)1 .(cid:13)(cid:13)(Xj≥1
1 ≈(cid:13)(cid:13)(cid:0)Xj≥0
1
.
.
1
1
Ψj ∗ I αf2)
Ψj ∗ J αf2)
2(cid:13)(cid:13)1
2(cid:13)(cid:13)1
2(cid:13)(cid:13)1
22jαρj ∗ f2(cid:1) 1
.
Ψj ∗ I αf2)
Ψj ∗ J αf2)
(cid:13)(cid:13)(Xj≥1
(cid:13)(cid:13)(Xj≥1
kgkhc
1 = kJ αfkhc
Similarly, for α ∈ R, we can prove that
In summary, we have proved (2.9), which yields that
Now define a new sequence ζ = (ζj)j≥0 by setting ζj = 2jαI−αφjρj for j ≥ 1 and ζ0 = J−αφ0ρ0.
Then
ζj ∗ g = 2jα φj ∗ ρj ∗ I−αg
and ζ0 ∗ g = φ0 ∗ ρ0 ∗ f.
Repeating the argument for (2.9) with ζ = (ζj )j≥0 instead of Ψ = (Ψj)j≥0, we get
ζj ∗ g2)
22jα φj ∗ ρj ∗ f2)
ζj ∗ I αf2)
1
1
2(cid:13)(cid:13)1 ≈(cid:13)(cid:13)(Xj≥0
.
1
2(cid:13)(cid:13)1
Then, we apply Lemma 2.6 to g with this new ζ instead of φ to get
(cid:13)(cid:13)(Xj≥0
(cid:13)(cid:13)(Xj≥0
2(cid:13)(cid:13)1 =(cid:13)(cid:13)(Xj≥0
2(cid:13)(cid:13)1 . kζk2,σkgkhc
1
ζj ∗ g2)
1 ≈ kζk2,σ(cid:13)(cid:13)(Xj≥0
22jαρj ∗ f2)
.
1
2(cid:13)(cid:13)1
It suffices to estimate the term kζk2,σ. By the definition of ζ = (ζj )j≥, we have
−2≤k≤2
kζj(2j+k·)ϕkHσ
sup
j≥1
kζ0(ϕ(0) + ϕ(1))kHσ
2
2
kφj(2j+k·)ϕkHσ
2
,
. sup
j≥1
−2≤k≤2
. kφ0(ϕ(0) + ϕ(1))kHσ
2
.
So we can use the same argument at the end of the proof of Theorem 2.5, to get
kζk2,σ . max(cid:8) sup
j≥1
−2≤k≤2
kφj(2j+k·)ϕkHσ
2
2(cid:9).
,kφ0(ϕ(0) + ϕ(1))kHσ
Combining the above inequalities, we get the desired assertion.
(cid:3)
In the setting where ℓ2 is replaced by H = ⊕∞j=0Hj with Hj = L2(cid:0)B(0, 2−j), 2jddt(cid:1), the coun-
terpart of Lemma 2.6 is the following:
18
R. Xia and X. Xiong
Lemma 2.8. Let φ = (φj )j≥0 be a sequence of continuous functions on Rd satisfying (2.7). Then
for 1 ≤ p ≤ 2 and f ∈ hc
p(Rd,M),
2djZB(0,2−j ) φj ∗ f (· + t)2dt)
(cid:13)(cid:13)(Xj≥0
1
2(cid:13)(cid:13)p . kφk2,σkfkhc
p
.
The relevant constant depends only on ϕ, σ and d.
Proof. The proof of this lemma is similar to Lemma 2.6; let us point out the necessary change.
Consider the H-valued Calder´on-Zygmund operator k defined on Lp(N ) given by
k(f )j(· + t) = φj ∗ f (· + t).
p(Rd,M) to Lp(N ; H c) for 1 ≤
The lemma is then reduced to showing that k is bounded from hc
p < 2. Since each Hj is a normalized Hilbert space, such that the constant function 1 has Hilbert
norm one, the kernel estimates of our k here are the same as the ones in Lemma 2.4. So we can
repeat the proof in Lemma 2.4 and Lemma 2.6. The desired assertion follows.
(cid:3)
Combining the above lemma with Theorem 1.5 (ksc,D
deduce the analogue of Theorem 2.7 in the setting H = ⊕∞j=0Hj with Hj = L2(cid:0)B(0, 2−j), 2jddt(cid:1).
Its proof is similar to that of Theorem 2.7, and is left to the reader.
Theorem 2.9. Keep the assumption in Theorem 2.5 and assume additionally that for any j ≥ 1,
ρj = ρ(2−j·) for some Schwartz function with supp ρ ⊂ {ξ : 2−1 ≤ ξ ≤ 2} and ρ(ξ) > 0 for any
2−1 < ξ < 2, and that supp ρ0 ⊂ {ξ : ξ ≤ 2} and ρ0(ξ) > 0 for any ξ < 2. Then for any
f ∈ S′(Rd; L1(M) + M),
Φ (f )kLp(N ) + kφ ∗ fkp ≈ kfkhc
p), we can
2j(2α+d)ZB(0,2−j ) φj ∗ ρj ∗ f (· + t)2dt)
kφj(2j+k·)ϕkHσ
2
2(cid:9)
,kφ0(ϕ(0) + ϕ(1))kHσ
−2≤k≤2
2j(2α+d)ZB(0,2−j ) ρj ∗ f (· + t)2dt)
(cid:13)(cid:13)(cid:13)(Xj≥0
. max(cid:8) sup
·(cid:13)(cid:13)(cid:13)(Xj≥0
j≥1
1
2(cid:13)(cid:13)(cid:13)1
2(cid:13)(cid:13)(cid:13)1
1
.
This theorem fills the gap of p = 1 left by Theorem 2.5. Both of them will be useful when we
consider the conic square functions of inhomogeneous Triebel-Lizorkin spaces in section 4.
3. Operator-valued Triebel-Lizorkin spaces
In this section, we give the definition of operator-valued Triebel-Lizorkin spaces, and then prove
some basic properties of them. Among the others, we connect operator-valued Triebel-Lizorkin
spaces with local Hardy spaces introduced in [36] via Bessel potentials. By this connection, we are
able to deduce the duality and the complex interpolation of Triebel-Lizorkin spaces. We also show
that for α > 0 the F α,c
(Rd,M)-norm is the sum of two homogeneous norms.
1
3.1. Definitions and basic properties. Recall that ϕ is a Schwartz function satisfying (0.1).
For each j ∈ N, ϕj is the function whose Fourier transform is equal to ϕ(2−j·), and ϕ0 is the
function whose Fourier transform is equal to 1 −Pj≥1 ϕ(2−j·). Moreover, the Fourier transform
of ϕj is denoted by ϕ(j) for j ∈ N0.
Definition 3.1. Let 1 ≤ p < ∞ and α ∈ R.
(1) The column Triebel-Lizorkin space F α,c
p
(Rd,M) is defined by
F α,c
p
where
(Rd,M) = {f ∈ S′(Rd; L1(M) + M) : kfkF α,c
p < ∞},
kfkF α,c
p =(cid:13)(cid:13)(Xj≥0
22jαϕj ∗ f2)
.
1
2(cid:13)(cid:13)p
(2) The row space F α,r
p
norm kfkF α,r
p = kf∗kF α,c
p
.
(Rd,M) consists of all f such that f∗ ∈ F α,c
p
(Rd,M), equipped with the
Inhomogeneous Triebel-Lizorkin spaces
19
(3) The mixture space F α
p (Rd,M) is defined to be
equipped with
F α
p
F α,c
p (Rd,M) =(F α,c
p =(inf{kgkF α,c
p
kfkF α
(Rd,M) + F α,r
(Rd,M) ∩ F α,r
p
p
(Rd,M)
(Rd,M)
if
if
1 ≤ p ≤ 2
2 < p < ∞,
p + khkF α,r
,kfkF α,r
p }
p
max{kfkF α,c
p
: f = g + h}
if
1 ≤ p ≤ 2
if
2 < p < ∞.
In the sequel, we focus on the study of the column Triebel-Lizorkin spaces. All results obtained in
-norm
the rest of this paper also admit the row versions. The following proposition shows that F α,c
is independent of the choice of the function ϕ satisfying (0.1).
p
Proposition 3.2. Let ψ be another Schwartz function satisfying the same condition (0.1) as ϕ.
For each j ∈ N, let ψj be the function whose Fourier transform is equal to ψ(2−j·), and let ψ0 be
the function whose Fourier transform is equal to 1 −Pj≥1 ψ(2−j·). Then
Proof. For any f ∈ S′(Rd; L1(M) + M), by the support assumption of ψ and ϕ, we have, for any
j ≥ 0,
22jαψj ∗ f2)
2(cid:13)(cid:13)p
kfkF α,c
.
1
p ≈(cid:13)(cid:13)(Xj≥0
1Xk=−1
ψj ∗ ϕk+j ∗ f,
with the convention ϕ−1 = 0. Thus by Theorems 2.2 and 2.7,
ψj ∗ f =
22jαψj ∗ f2)
1
2(cid:13)(cid:13)p
≤
(cid:13)(cid:13)(Xj≥0
1Xk=−1(cid:13)(cid:13)(Xj≥0
. max(cid:8) sup
.(cid:13)(cid:13)(Xj≥0
2
−2≤k≤2 kψ(2k·)ϕkHσ
2(cid:13)(cid:13)p
22jαϕj ∗ f2)
.
1
22jαψj ∗ ϕk+j ∗ f2)
1
2(cid:13)(cid:13)p
,kψ0(ϕ(0) + ϕ(1))kHσ
p
p
(Rd,M) is a Banach space.
(Rd,M) ⊂ F β,c
Changing the role of ϕ and ψ, we get the reverse inequality.
Proposition 3.3. Let 1 ≤ p < ∞ and α ∈ R. Then
(1) F α,c
(2) F α,c
(3) F 0,c
Proof. (1) Let {fi} be a Cauchy sequence in F α,c
(ϕ0 ∗ fi, . . . , 2jαϕj ∗ fi, . . .) is also a Cauchy sequence in Lp(N ; ℓc
function f = (f 0, . . . , f j, . . .) in Lp(N ; ℓc
2(N0)). Formally we take
(3.1)
p(Rd,M) with equivalent norms.
(Rd,M) if α > β.
p (Rd,M) = hc
f j.
p
p
f =Xj≥0
2(cid:9)(cid:13)(cid:13)(Xj≥0
22jαϕj ∗ f2)
1
2(cid:13)(cid:13)p
(cid:3)
(Rd,M). Then, the sequence {ai} with ai =
2(N0)). Thus, ai converges to a
Since for each j ∈ N, suppcf j ⊂ {ξ : 2j−1 ≤ ξ ≤ 2j+1} and suppcf 0 ⊂ {ξ : ξ ≤ 2}, the series
(3.1) converges in S′(Rd; Lp(M)). Let ϕj = 0 if j < 0. By the support assumption of ϕ, when
i → ∞, we get
ϕj ∗ fi =
ϕk ∗ ϕj ∗ fi →
ϕj ∗ f k = ϕj ∗ f,
j+1Xk=j−1
j+1Xk=j−1
which implies that f j = 2jαϕj ∗ f , for any j ≥ 0. Thus, f ∈ F α,c
in F α,c
p
(Rd,M) and {fi} converges to f
p
(Rd,M).
(2) is obvious.
20
R. Xia and X. Xiong
terization of hc
(3) It is easy to see that any ϕ satisfying (0.1) also satisfies (1.6). Then by the discrete charac-
(cid:3)
p(Rd,M) given in Theorem 1.5, we get the desired assertion.
Given a ∈ R+, we define Di,a(ξ) = (2πiξi)a for ξ ∈ Rd, and Da
i to be the Fourier multiplier
(Rd,M). We set Da = D1,a1 ··· Dd,ad and
with symbol Di,a(ξ) on Triebel-Lizorkin spaces F α,c
Da = Da1
i = ∂a
i ,
so there does not exist any conflict of notation. The operator Da can be viewed as a fractional
extension of partial derivatives. The following is the so-called lifting (or reduction) property of
Triebel-Lizorkin spaces.
+. Note that if a is a positive integer, Da
for any a = (a1,··· , ad) ∈ Rd
1 ··· Dad
d
p
Proposition 3.4. Let 1 ≤ p < ∞ and α ∈ R.
(1) For any β ∈ R, J β is an isomorphism between F α,c
(Rd,M) and hc
J α is an isomorphism between F α,c
p
p
(Rd,M) and F α−β,c
p(Rd,M).
p
(2) Let β > 0. Then f ∈ F α,c
p
(Rd,M) if and only if ϕ0 ∗ f ∈ Lp(N ) and Dβ
i f ∈ F α−β,c
p
(Rd,M)
for all i = 1, . . . , d. Moreover, in this case,
(Rd,M). In particular,
kfkF α,c
p ≈ kϕ0 ∗ fkp +
dXi=1
kDβ
i fkF α−β,c
p
.
Proof. (1) Let f ∈ F α,c
p
(Rd,M). Applying Theorems 2.2 and 2.7 with ρ = ϕ, we obtain
2−jβkJβ(2j+k·)ϕkHσ
2
,kJβ(ϕ(0) + ϕ(1))kHσ
2 }
kJ βfkF α−β,c
p
(3.2)
22j(α−β)ϕj ∗ J βf2)
1
2(cid:13)(cid:13)p
=(cid:13)(cid:13)(Xj≥0
·(cid:13)(cid:13)(Xj≥0
. max{ sup
j≥1
−2≤k≤2
22jαϕj ∗ f2)
.
1
2(cid:13)(cid:13)p
It is easy to check that all partial derivatives of 2−jβJβ(2j+k·)ϕ of order less than or equal to
[σ] + 1 are bounded uniformly in j ≥ 1 and −2 ≤ k ≤ 2, and that Jβ(ϕ(0) + ϕ(1)) ∈ H σ
2 (Rd). Thus
(Rd,M), and its inverse
kJ βfkF α−β,c
J−β is also continuous from F α−β,c
. So J β is continuous from F α,c
(Rd,M) to F α−β,c
. kfkF α,c
p
p
p
p
p
(Rd,M) to F α,c
p
(Rd,M).
(2) If we take σ ∈ ( d
2 , β + d
2 ), then we have kDi,βϕ0kHσ
2
< ∞ and kDi,βϕkHσ
2
< ∞. Replacing
J β by Dβ
i
in (3.2), we obtain that, for any i = 1, . . . , d,
kDβ
i fkF α−β,c
p
. kfkF α,c
p
,
which implies immediately that
kϕ0 ∗ fkp +
dXi=1
kDβ
i fkF α−β,c
p
. kfkF α,c
p
.
To show the reverse inequality, we choose a nonnegative infinitely differentiable function χ on R
such that χ(s) = 0 if s < 1
. For i = 1, . . . , d, we define χi on Rd as
2√d
follows:
and χ(s) = 1 if s ≥ 1√d
χi(ξ) =
1
χ(ξ1)ξ1β + . . . + χ(ξd)ξdβ
χ(ξi)ξiβ
(2πiξi)β
,
whenever the first denominator is positive, say, when ξ ≥ 1. Then for any j ≥ 1, χiϕj is a
well-defined infinitely differentiable function on Rd\{ξ : ξi = 0} and
ϕ(j) =
dXi=1
χiDi,βϕ(j).
Inhomogeneous Triebel-Lizorkin spaces
21
Then by Theorem 2.1, we have
kfkF α,c
p ≤ kϕ0 ∗ fkp +
. kϕ0 ∗ fkp +
22jα χi ∗ ϕj ∗ Dβ
2(cid:13)(cid:13)p
i f2(cid:1) 1
2(cid:13)(cid:13)(cid:0)Xj≥1
2jβkχi(2j+k·)ϕkHσ
dXi=1(cid:13)(cid:13)(cid:0)Xj≥1
dXi=1
2jβkχi(2j+k·)ϕkHσ
−2≤k≤2
sup
j≥1
2 (Rd) = kφi(2k·)ϕkHσ
2 (Rd),
However,
where
22j(α−β)ϕj ∗ Dβ
2(cid:13)(cid:13)p
i f2(cid:1) 1
.
φi(ξ) =
χ(2jξ1)ξ1β + . . . + χ(2jξd)ξdβ
1
χ(2jξi)ξiβ
(2πiξi)β
.
Since all partial derivatives of φiϕ(2k·), of order less than a fixed integer, are bounded uniformly in
2 (Rd) are bounded from above by a constant independent
j, k and i, and the norm of φiϕ(2k·) in H σ
of j, k and i. Then we deduce
kfkF α,c
p . kϕ0 ∗ fkp +
≤ kϕ0 ∗ fkp +
The assertion is proved.
dXi=1(cid:13)(cid:13)(cid:0)Xj≥1
dXi=1
kDβ
22j(α−β)ϕj ∗ Dβ
2(cid:13)(cid:13)p
i f2(cid:1) 1
i fkF α−β,c
p
.
(cid:3)
Definition 3.5. For α ∈ R, we define F α,c
∞ (Rd,M) as the space of all f ∈ S′(Rd;M) such that
kϕ0 ∗ fkN + sup
Q<1(cid:13)(cid:13)(cid:13)
We endow the space F α,c
∞ (Rd,M) with the norm:
= kϕ0 ∗ fkN + sup
kfkF α,c
∞
1
QZQ Xj≥− log2(l(Q))
Q<1(cid:13)(cid:13)(cid:13)
1
QZQ Xj≥− log2(l(Q))
22jαϕj ∗ f (s)2ds(cid:13)(cid:13)(cid:13)
1
2
M
< ∞.
22jαϕj ∗ f (s)2ds(cid:13)(cid:13)(cid:13)
1
2
.
M
p
Proposition 3.6. Let 1 ≤ p < ∞, α ∈ R and q be the conjugate index of p. Then the dual space
of F α,c
(Rd,M) coincides isomorphically with F −α,c
(Rd,M).
Proof. First, we show that J α is an isomorphism between F α,c
end, we use the discrete Carleson characterization of bmoc(Rd,M) in [36, Corollary 5.13]:
(3.3)
∞ (Rd,M) and bmoc(Rd,M). To this
1
1
2
q
kfkbmoc ≈ kφ ∗ fkN + sup
where Φ ∈ S(Rd) and φ ∈ H σ
(3.3) to J αf :
2 (Rd) satisfying (1.7). By taking φ = ϕ0 and Φ = J−αϕ, we apply
,
M
QZQ Xj≥− log2(l(Q))
Q<1(cid:13)(cid:13)(cid:13)
QZQ Xj≥− log2(l(Q))
Q<1(cid:13)(cid:13)(cid:13)
QZQ Xj≥− log2(l(Q))
Q<1(cid:13)(cid:13)(cid:13)
Φj ∗ f (s)2ds(cid:13)(cid:13)(cid:13)
(J−αϕ)j ∗ (J αf )(s)2ds(cid:13)(cid:13)(cid:13)
22jαϕj ∗ f (s)2ds(cid:13)(cid:13)(cid:13)
M
1
1
1
2
1
2
M
kJ αfkbmoc ≈ kϕ0 ∗ fkN + sup
= kϕ0 ∗ fkN + sup
= kfkF α,c
∞
.
Since J α is also an isomorphism between F α,c
1-bmoc duality and the hc
hc
with equivalent norms.
p-hc
p
(Rd,M) and hc
p(Rd,M) for any 1 < p < ∞, by the
(Rd,M)
(cid:3)
(Rd,M)∗ = F −α,c
q
q duality in Theorem 1.1, we see that F α,c
p
22
R. Xia and X. Xiong
3.2. Interpolation. Now we indicate a complex interpolation result of Triebel-Lizorkin spaces.
It is deduced from the interpolation of local Hardy and bmo spaces in Theorem 1.2, and the
boundedness of complex order Bessel potentials on them.
Proposition 3.7. Let α0, α1 ∈ R and 1 < p < ∞. Then
(cid:0)F α0,c
∞ (Rd,M), F α1,c
1
(Rd,M)(cid:1) 1
p
= F α,c
p
(Rd,M), α = (1 −
1
p
)α0 +
α1
p
.
Proof. Let f ∈ F α,c
p(Rd,M). Therefore, according
to Theorem 1.2 (1), there exists a continuous function on the strip {z ∈ C : 0 ≤ Rez ≤ 1}, analytic
in the interior, such that J αf = F ( 1
(Rd,M). By Proposition 3.4, we have J αf ∈ hc
p
p ) ∈ hc
p(Rd,M) and
sup
sup
t∈R kF (it)kbmoc < ∞ and
t∈R kF (1 + it)khc
1
< ∞.
We consider Bessel potentials of complex order. For z ∈ C, define Jz(ξ) = (1 + ξ2) z
be the associated Fourier multiplier. We set
2 , and J z to
For any t ∈ R,
and
keF (it)kF
keF (1 + it)kF
t∈R keF (it)kF
This will imply that f = eF ( 1
F α,c
sup
p
J−(1−z)α0−zα1 F (z).
p )2
eF (z) = e(z− 1
∞ ≈ e−t2+ 1
1 ≈ e−t2+(1− 1
α0 ,c
α1,c
p2 kJ it(α0−α1)F (it)kbmoc
p )2
kJ it(α0−α1)F (1 + it)khc
1
.
α1,c
1
sup
α0,c
∞
< ∞ and
∞ (Rd,M), F α1,c
t∈R keF (1 + it)kF
(Rd,M)(cid:1) 1
∞ (Rd,M), F α1,c
p ) ∈(cid:0)F α0,c
(Rd,M) ⊂(cid:0)F α0,c
1
1
p
(Rd,M)(cid:1) 1
p
. Hence,
.
< ∞.
We claim that J it is a bounded Fourier multiplier on hc
bmoc(Rd,M) too. Therefore, we will have
1(Rd,M), so by duality, it is bounded on
By duality, we will get the reverse inclusion for the Calder´on's second interpolation (·,·)
the inclusion (·,·) 1
we will obtain the desired assertion.
1
p . Then by
1
p between two kinds of complex interpolations (see [2, Theorem 4.3.1]),
p ⊂ (·,·)
Now, we prove the claim. First, we easily check that Jit is d-times differentiable on Rd \ {0},
and for any m ∈ Nd
0 and m1 ≤ d, we have
Next, we check that (with Jit(2kξ) = (1 + 2kξ2)
max
−2≤k≤2kJit(2k·)ϕkHd
2
sup(cid:8)ξm1DmJit(ξ) : ξ 6= 0(cid:9) . (1 + t)d.
and kJit(ϕ(0) + ϕ(1))kHd
. (1 + t)d
it
2 ),
2
. (1 + t)d.
By (3) in Proposition 3.3, if we take (ϕj )j≥0 to be the Littlewood-Paley decomposition on Rd
satisfying (0.2) and (0.3), we have
and
kJ itfkhc
1 ≈(cid:13)(cid:13)(Xj≥0
1 ≈(cid:13)(cid:13)(Xj≥0
−2≤k≤2 kJit(2k·)ϕkHd
kfkhc
2
1
2(cid:13)(cid:13)1
Jit ∗ ϕj ∗ f2)
ϕj ∗ f2)
.
1
2(cid:13)(cid:13)1
Then, we apply Theorem 2.7 with ρj = ϕj, φj (2j·) = Jit, and α = 0, σ = d,
2(cid:9)kfkhc
. max(cid:8) max
,kJit(ϕ(0) + ϕ(1))kHd
kJ itfkhc
1
1
The claim is proved.
. (1 + t)dkfkhc
1
.
(cid:3)
Inhomogeneous Triebel-Lizorkin spaces
23
Remark 3.8. The real interpolation of the couple(cid:0)F α,c
of(cid:0)F α1,c
that of Hardy spaces (see Theorem 1.2) and Proposition 3.4. But if α1 6= α2, the real interpolation
in this paper, and refer the reader to [40] for similar results on homogeneous Triebel-Lizorkin (and
Besov) spaces.
(Rd,M)(cid:1) follows easily from
(Rd,M)(cid:1) will give Besov type spaces. We will not consider this problem
∞ (Rd,M), F α2,c
∞ (Rd,M), F α,c
1
1
1
3.3. Triebel-Lizorkin spaces with α > 0. The following result shows that when α > 0, the
F α,c
(Rd,M)-norm can be rewritten as the sum of two homogeneous norms. Recall that for a fixed
Schwartz function ϕ in (0.1), the functions ϕj's determined by bϕj(ξ) = ϕ(2−jξ), j ∈ Z give a
homogeneous Littlewood-Paley decomposition on Rd satisfying (0.4).
Proposition 3.9. Let α > 0. If 1 ≤ p < ∞, then
22jα ϕj ∗ f2)
,
∀ f ∈ F α,c
p
(Rd,M).
If 1 ≤ p ≤ 2,
kfkF α,c
+∞Xj=−∞
p ≈ kϕ0 ∗ fkp +(cid:13)(cid:13)(
+∞Xj=−∞
p ≈ kfkp +(cid:13)(cid:13)(
1
2(cid:13)(cid:13)p
2(cid:13)(cid:13)p
,
1
∀ f ∈ F α,c
Proof. Firstly, we prove the first equivalence. By the definition of the F α,c
22jα ϕj ∗ f2)
kfkF α,c
p
(Rd,M).
-norm, it is obvious that
To prove the reverse inequality, it suffices to show:
22jα ϕj ∗ f2)
kfkF α,c
p . kϕ0 ∗ fkp +(cid:13)(cid:13)(
+∞Xj=−∞
+∞Xj=1
2(cid:13)(cid:13)p . kϕ0 ∗ fkp +(cid:13)(cid:13)(
1
22jα ϕj ∗ f2)
22jα ϕj ∗ f2)
.
1
2(cid:13)(cid:13)p
0Xj=−∞
(cid:13)(cid:13)(
p
.
1
2(cid:13)(cid:13)p
By the support assumption of ϕ, we have ϕ(0) = 1 for any ξ ≤ 1. Thus, when j < 0,
Then
(3.4)
ϕ(2j·) = ϕ(2j·)ϕ(0).
ϕj ∗ f = ϕj ∗ ϕ0 ∗ f.
By the triangle inequality, (3.4) and [39, Lemma 1.7], we obtain
0Xj=−∞
(cid:13)(cid:13)(
2jα ϕj ∗ f2)
1
2(cid:13)(cid:13)p .
2jαk ϕj ∗ ϕ0 ∗ fkp + k ϕ0 ∗ fkp
.
−1Xj=−∞
−1Xj=−∞
0Xj=−∞
+∞Xj=1
. kϕ0 ∗ fkp +(cid:13)(cid:13)(
.
2jαk ϕjk1kϕ0 ∗ fkp +(cid:13)(cid:13)ϕ(ϕ0 + ϕ1 + ϕ2) ∗ f(cid:13)(cid:13)p
+∞Xj=1
2jαkϕ0 ∗ fkp +(cid:13)(cid:13)(
22jαϕj ∗ f2)
2(cid:13)(cid:13)p
1
22jαϕj ∗ f2)
1
.
2(cid:13)(cid:13)p
2(cid:13)(cid:13)p gives rise to an equivalent
1
Therefore, we have proved that kϕ0 ∗ fkp +(cid:13)(cid:13)(P+∞j=1 22jαϕj ∗ f2)
(Rd,M) when α > 0.
norm on F α,c
p
Now let us deal with the second equivalence. For any 1 ≤ p ≤ 2 and α > 0, we have
(Rd,M) ⊂ hc
. Combined with the equiva-
p(Rd,M) ⊂ Lp(N ). Therefore kfkp . kfkF α,c
F α,c
lence obtained above, we see that
p
p
kfkp +(cid:13)(cid:13)(
+∞Xj=−∞
22jα ϕj ∗ f2)
1
2(cid:13)(cid:13)p . kfkF α,c
p
.
The reverse inequality can be easily deduced by the fact that kϕ0 ∗ fkp ≤ kϕ0k1kfkp.
(cid:3)
24
R. Xia and X. Xiong
We also have a continuous counterpart of Proposition 3.9. For any ε ≥ 0, we define ϕε =
F−1(ϕ(ε·)).
Corollary 3.10. Let 1 ≤ p ≤ 2 and α > 0. Then, for any f ∈ F α,c
ε−2α ϕε ∗ f2 dε
)
kfkF α,c
ε
p
p ≈ kfkp +(cid:13)(cid:13)(cid:13)(Z ∞
0
4. Characterizations
(Rd,M),
.
1
2(cid:13)(cid:13)(cid:13)p
In this section we give two kinds of characterizations of the Triebel-Lizorkin spaces defined
previously: one is done by directly replacing the function ϕ in Definition 3.1 by more general
convolution kernels; the other is described by Lusin square functions. Since the local Hardy spaces
are included in the family of inhomogeneous Triebel-Lizorkin spaces, these two characterizations
can be seen as extensions as well as improvements of those in [36] for local Hardy spaces, listed in
Theorems 1.4 and 1.5. The multiplier theorems in section 2 will play a crucial role in this section.
4.1. General characterizations. We have seen in section 3.1 that the definition of Triebel-
Lizorkin spaces is independent of the choice of ϕ satisfying (0.1). In this section, we will show
that this kernel is not even necessarily a Schwartz function.
Let σ > d
2 and Φ(0), Φ be two complex-valued infinitely differentiable functions defined respec-
tively on Rd and Rd\{0}, which satisfy
Φ(0)(ξ) > 0 if ξ ≤ 2,
2−kα0kΦ(0)(2k·)ϕkHσ
sup
k∈N0
2
< ∞,
1
2 ≤ ξ ≤ 2,
2−kα0kΦ(2k·)ϕkHσ
(1 + s2)σF−1(Φϕ(0)I−α1 )(s)ds < ∞.
< ∞,
2
Φ(ξ) > 0 if
sup
k∈N0
ZRd
(4.1)
and
(4.2)
write
Then
(4.4)
Let Φ(j) = Φ(2−j·) for j ≥ 1, and Φj be the function whose Fourier transform is equal to Φ(j)
Recall that here I−α1 (ξ) is the symbol of the Riesz potential I−α1 = (−(2π)−2∆) −α1
for any j ∈ N0.
Theorem 4.1. Let 1 ≤ p < ∞ and α ∈ R. Assume that α0 < α < α1, α1 ≥ 0 and Φ(0), Φ satisfy
conditions (4.1) and (4.2) respectively. Then for any f ∈ S′(Rd; L1(M) + M), we have
(4.3)
.
,
1
2
kfkF α,c
p ≈(cid:13)(cid:13)(Xj≥0
22jαΦj ∗ f2)
2(cid:13)(cid:13)p
where the relevant constants are independent of f .
Proof. We follow the pattern of the proof of [40, Theorem 4.17]. Denote the norm on the right
hand side of (4.3) by kfkF α,c
Step 1. Let ϕk = 0 (and so is ϕ(k)) if k < 0. Given a positive integer K, for any j ∈ N0, we
p,Φ
.
Φ(j) =
K−1Xk=−∞
Φj ∗ f = Xk≤K−1
Φ(j)ϕ(j+k) +
∞Xk=K
Φj ∗ ϕj+k ∗ f + Xk≥K
Φ(j)ϕ(j+k).
Φj ∗ ϕj+k ∗ f.
Temporarily we take for granted that the second series is convergent not only in S′(Rd; L1(M)+M)
but also in F α,c
(Rd,M), which is to be settled up in the last step. Then we obtain
p
kfkF α,c
p,Φ ≤ I + II + III,
Inhomogeneous Triebel-Lizorkin spaces
25
where
The term II is easy to deal with. By (0.5) and (4.1), we obtain
K−1Xk=0
kΦ0 ∗ ϕk ∗ fkp =
kΦ0 ∗ (ϕk−1 + ϕk + ϕk+1) ∗ ϕk ∗ fkp
22jαΦj ∗ ϕj+k ∗ f2)
kΦ0 ∗ ϕk ∗ fkp,
22jαΦj ∗ ϕj+k ∗ f2)
,
1
2(cid:13)(cid:13)p
2(cid:13)(cid:13)p
.
1
I = Xk≤K−1(cid:13)(cid:13)(Xj≥1
II = Xk≤K−1
III = Xk≥K(cid:13)(cid:13)(Xj≥0
K−1Xk=0
K−1Xk=0
.
kϕk ∗ fkpkΦ0 ∗ (ϕk−1 + ϕk + ϕk+1)k1
. sup
k∈N0
2−kα0kΦ(0)(2k·)ϕkHσ
2
. CKkfkF α,c
p
.
K−1Xk=0
2k(α0−α)k2kαϕk ∗ fkp
Let us treat the terms I and III separately. By the support assumption of ϕ(k) and the property
that it is equal to 1 when ξ ≤ 1, for k ≤ K − 1, we have
Φ(ξ)ϕ(0)(2−Kξ)
Φ(ξ)ϕ(k)(ξ) =
(4.5)
ξα1
= 2kα1 η(ξ)ρ(k)(ξ),
ξα1 ϕ(k)(ξ)
where η, ρ are defined by
η(ξ) =
Let η(j) = η(2−j·), j ∈ Z. For j ≥ 1, define ηj = F−1(η(j)). Then for any j ≥ 1, we have
Φ(ξ)ϕ(0)(2−Kξ)
ξα1
and ρ(ξ) = ξα1 ϕ(ξ).
Φj ∗ ϕj+k ∗ f = 2kα1 ηj ∗ ρj+k ∗ f.
Now we are ready to estimate I. Applying Theorems 2.2 and 2.7 twice, we get
(4.6)
1
1
1
, max
22(j+k)αηj ∗ ρj+k ∗ f2)
I = Xk≤K−1
2k(α1−α)(cid:13)(cid:13)(Xj≥1
2(cid:13)(cid:13)p
= Xk≤K−1
2k(α1−α)(cid:13)(cid:13)( Xj≥k+1
2(cid:13)(cid:13)p
22jαηj−k ∗ ρj ∗ f2)
. Xk≤K−1
2(cid:9)
2k(α1−α) max(cid:8)kη(−k)(ϕ(0) + ϕ(1))kHσ
−2≤ℓ≤2kη(−k−ℓ)ϕkHσ
·(cid:13)(cid:13)(Xj≥0
2(cid:13)(cid:13)p
22jαρj ∗ f2)
. Xk≤K−1
2k(α1−α) max(cid:8)kη(−k)(ϕ(0) + ϕ(1))kHσ
2(cid:9)(cid:13)(cid:13)(Xj≥0
· max(cid:8)kIα1 (ϕ(0) + ϕ(1))kHσ
= Xk≤K−1
2k(α1−α) max(cid:8)kη(−k)(ϕ(0) + ϕ(1))kHσ
2(cid:9)kfkF α,c
· max(cid:8)kIα1 (ϕ(0) + ϕ(1))kHσ
2(cid:9)
−2≤ℓ≤2kη(−k−ℓ)ϕkHσ
2(cid:13)(cid:13)p
22jαϕj ∗ f2)
2(cid:9)
−2≤ℓ≤2kη(−k−ℓ)ϕkHσ
,kIα1 ϕkHσ
,kIα1 ϕkHσ
, max
, max
.
p
2
2
2
2
2
1
Let us deal with all the factors in the last term of the above inequality. Firstly, when α1 = 0, it
2 (Rd). Secondly, we treat the case α1 > 0.
< ∞. Next, we deal with the term Iα1 (ϕ(0) + ϕ(1)), which can
is obvious that Iα1 ϕ ∈ H σ
First, it is easy to see that kIα1 ϕkHσ
2 (Rd) and Iα1 (ϕ(0) + ϕ(1)) ∈ H σ
2
26
R. Xia and X. Xiong
be reduced to Iα1 ϕ(0) by dilation. Let N be a positive integer such that α1 > 1
N . If the dimension
d is odd, we consider the function F (z) = e(z− N +2
N )zϕ(0), which is continuous
on the strip {z ∈ C : 0 ≤ Re(z) ≤ 1}, and analytic in the interior. A direct computation shows
that supt∈R kF (it)kH
2 , we
have
< ∞ and supt∈R kF (1 + it)kH
< ∞. Then for θ =
ξα1− 1
N + 1
2
1
N +1
N +(1+ 1
2N +2 )2
2− 1
> 1
2 − 1
+ 1
2
d
2
2
d
2
1
2
F (θ) = Iα1 ϕ(0) ∈ H σ
(Rd), H
2 . On the other hand, if d is even, set F (z) = e(z− 1
< ∞, and that supt∈R kF (1 + it)kH
2 (Rd) =(cid:0)H
2 − 1
2
d
2
2
d
2 + 1
2
2N )2
d
2
2
+1
d
2
2
for some σ > d
check that supt∈R kF (it)kH
have
F (θ) = Iα1 ϕ(0) ∈ H
d
2 + 1
2
2N
(Rd) =(cid:0)H
d
2
2 (Rd), H
d
2 +1
2
Thus, for any α1 > 0, we can always choose a positive σ > d
2 (Rd). Finally,
we have to estimate kη(−k)ϕkHσ
2 uniformly in k, which will yield the convergence
of the last sum in (4.6) by dilation again. To this end, note that by (4.2), η is integrable on Rd,
then we use the Cauchy-Schwarz inequality in the following way:
2 such that Iα1 ϕ(0) ∈ H σ
2 and kη(−k)ϕ(0)kHσ
2 ϕ(0). We can also
2N , we
,
(Rd)(cid:1)θ
ξN α1z+ α1
< ∞. Then for θ = 1
(Rd)(cid:1)θ
.
F−1(η(−k)ϕ)(s)2 = ZRd
η(t)F−1(ϕ)(s − 2kt)dt2
≤ kηk1ZRd η(t) · F−1(ϕ)(s − 2kt)2dt.
(1 + s − 2kt2)σF−1(ϕ)(s − 2kt)2dsdt
Hσ
2
For k ≤ K − 1, we have
kη(−k)ϕk2
≤ kηk1ZRd
. kηk1ZRd
≤ 2Kσkηk1ZRd
≤ Cϕ0,σ,K(cid:0)ZRd
(4.7)
The other term kη(−k)ϕ(0)kHσ
(4.8)
(1 + s2)σF−1(η(−k)ϕ)(s)2ds
=ZRd
(1 + s2)σZRd η(t) · F−1(ϕ)(s − 2kt)2dtds
(1 + 2kt2)ση(t)ZRd
(1 + t2)ση(t)dtZRd
(1 + t2)ση(t)dt(cid:1)2
I . CΦ,ϕ(0),α1,α,KZRd
η = I−α1 Φ[ϕ(0)(2−K·) − ϕ(0)] + I−α1 Φϕ(0).
2 is dealt with in the same way.
(1 + t2)ση(t)dtkfkF α,c
.
p
.
(1 + s2)σF−1(ϕ)(s)2ds
Going back to the estimate of I, by the previous inequalities, we obtain
In order to return from η back to ϕ0, we write
Since I−α1 Φ(ϕ(0)(2−K·) − ϕ(0)) is an infinitely differentiable function with compact support, we
have
(1 + t2)σF−1(I−α1 Φ(ϕ(0)(2−K·) − ϕ(0)))(t)dt = C′Φ,ϕ(0) ,α1,α,K < ∞.
Then (4.2) implies that
ZRd
(1 + t2)ση(t)dt . C′Φ,ϕ(0),α1,α,K +ZRd
ZRd
Therefore,
(1 + s2)σF−1(I−α1 Φϕ(0))(s)ds < ∞.
Step 2. Now it remains to estimate the third term III. Let H be a Schwartz function such that
I . kfkF α,c
p
.
(4.9)
supp H ⊂ {ξ ∈ Rd :
1
4 ≤ ξ ≤ 4}
and H(ξ) = 1 if
1
2 ≤ ξ ≤ 2.
Inhomogeneous Triebel-Lizorkin spaces
27
Let H (k) = H(2−k·). For k ≥ K, we have
Φ(ξ)ϕ(k)(ξ) =
(4.10)
and
(4.11)
Φ(0)(ξ)ϕ(k)(ξ) =
Φ(ξ)
ξα0
H (k)(ξ)ϕ(k)(ξ)ξα0 ,
Φ(0)(ξ)
ξα0
H (k)(ξ)ϕ(k)(ξ)ξα0 .
For any j ∈ N0, we keep using the notation Φj = F−1(Φ(j)) and Hj = F−1(H (j)). Thus, we have
Φj ∗ ϕj+k ∗ f = 2kα0 (I−α0 Φ)j ∗ Hj+k ∗ (Iα0 ϕ)j+k ∗ f.
Since both H and ϕ vanish near the origin, by Theorems 2.2 and 2.7, we obtain
Therefore,
III = Xk≥K
= Xk≥K
Xk≥K
2k(α0−α)(cid:13)(cid:13)(cid:0)Xj≥k
max(cid:8)2−kα0 max
· Xk≥K
2k(α0−α)kfkF α,c
. sup
k∈N0
.
p
2(cid:13)(cid:13)p
22(j+k)α(I−α0 Φ)j ∗ Hj+k ∗ (Iα0 ϕ)j+k ∗ f2(cid:1) 1
2(cid:13)(cid:13)p
22jα(I−α0 Φ)j−k ∗ Hj ∗ (Iα0 ϕ)j ∗ f2(cid:1) 1
2k(α0−α)(cid:13)(cid:13)(cid:0)Xj≥0
2k(α0−α)(cid:13)(cid:13)(cid:0)Xj≥k
2(cid:13)(cid:13)p
22jα(I−α0 Φ)j−k ∗ Hj ∗ (Iα0 ϕ)j ∗ f2(cid:1) 1
−2≤ℓ≤2kI−α0 Φ(2k+ℓ·)H(2ℓ·)ϕkHσ
.
2
2(cid:9)
, 2−kα0kI−α0 Φ(0)(2k·)H(ϕ(0) + ϕ(1))kHσ
Then by (2.6), (4.1) and (4.2), we have, for any −2 ≤ ℓ ≤ 2,
2−kα0kI−α0 Φ(2k+ℓ·)H(2ℓ·)ϕkHσ
2 ZRd
≤ 2−kα0kΦ(2k+ℓ·)ϕkHσ
. 2−kα0kΦ(2k+ℓ·)ϕkHσ
2 ≤ sup
k∈N0
2
(1 + t2)σF−1(I−α0 H(2ℓ·))(t)dt
2−kα0kΦ(2k·)ϕkHσ
2
< ∞,
(4.12)
and
(4.13)
2−kα0kI−α0 Φ(0)(2k·)H(ϕ(0) + ϕ(1))kHσ
= 2−kα0kI−α0Φ(0)(2k·)H
ϕ(2−ℓ′·)kHσ
2
2
1Xℓ′=−2
. 2−kα0
≤ 2−kα0
kI−α0Φ(0)(2k+ℓ′·)H(2ℓ′·)ϕkHσ
2
1Xℓ′=−2
1Xℓ′=−2
2−kα0kΦ(0)(2k·)ϕkHσ
2 ZRd
kΦ(0)(2k+ℓ′·)ϕkHσ
< ∞.
2
. sup
k∈N0
Then we get
(1 + t2)σF−1(I−α0 H(2ℓ·))(t)dt
III ≤ CΦ,α0,α,KkfkF α,c
Combining this estimate with those of I and II, we finally get
.
p
kfkF α,c
p,Φ
. kfkF α,c
p
.
Step 3. We turn to the reverse inequality. Note that ϕ(0)(ξ) = 1 when ξ ≤ 1, then by (4.1)
and (4.2), for any j ∈ N0, we write
(4.14)
ϕ(j)(ξ) = ϕ(j)(ξ) ϕ(0)(2−j−M ξ) =
ϕ(j)(ξ)
Φ(j)(ξ)
ϕ(0)(2−j−M ξ)Φ(j)(ξ),
28
R. Xia and X. Xiong
where M is a positive integer to be chosen later. By Theorems 2.2 and 2.7,
kfkF α,c
p =(cid:13)(cid:13)(cid:0)Xj≥0
. max(cid:8) max
·(cid:13)(cid:13)(cid:0)Xj≥0
.(cid:13)(cid:13)(cid:0)Xj≥0
2(cid:13)(cid:13)p
22jαϕj ∗ f2(cid:1) 1
−2≤ℓ≤2kΦ−1(2ℓ·)ϕ(2ℓ·)ϕkHσ
2(cid:13)(cid:13)p
22jα(ϕ0)j+M ∗ Φj ∗ f2(cid:1) 1
2(cid:13)(cid:13)p
22jα(ϕ0)j+M ∗ Φj ∗ f2(cid:1) 1
,
2
2(cid:9)
,k(Φ(0))−1ϕ(0)(ϕ(0) + ϕ(1))kHσ
where (ϕ0)j+M is the Fourier inverse transform of ϕ(0)(2−j−M·). Let h = 1 − ϕ(0). Write
ϕ(0)(2−j−M ξ)Φ(j)(ξ) = Φ(j)(ξ) − h(j+M)(ξ)Φ(j)(ξ). Then, we have
kfkF α,c
p . kfkF α,c
p,Φ
+(cid:13)(cid:13)(cid:0)Xj≥0
2(cid:13)(cid:13)p
22jαhj+M ∗ Φj ∗ f2(cid:1) 1
,
where the relevant constant depends only on p, σ, d and ϕ(0). Applying the arguments in the
estimate of III, (4.10) with h(M)Φ in place of Φ and (4.11) with h(M)Φ(0) in place of Φ(0), we
deduce
≤ C1 sup
k≥M
(cid:13)(cid:13)(cid:0)Xj≥0
2(cid:13)(cid:13)p
22jαhj+M ∗ Φj ∗ f2(cid:1) 1
2−kα0 max(cid:8) max
−2≤ℓ≤2kh(2k−M+ℓ·)Φ(2k+ℓ·)ϕkHσ
· Xk≥M
2k(α0−α)kfkF α,c
2−kα0 max(cid:8) max
2M(α0−α)
1 − 2α0−αkfkF α,c
−2≤ℓ≤2kh(2k−M+ℓ·)Φ(2k+ℓ·)ϕkHσ
= sup
k≥M
· C1
,
p
p
2
2
2(cid:9)
,kh(2k−M·)Φ(0)(2k·)ϕkHσ
2(cid:9)
,kh(2k−M·)Φ(0)(2k·)ϕkHσ
where C1 is a constant which depends only on p, σ, d, H and α0. Now we replace h in the above
Sobolev norm by 1 − ϕ(0):
kh(2k−M+ℓ·)Φ(2k+ℓ·)ϕkHσ
2 ≤ kΦ(2k+ℓ·)ϕkHσ
2 + kϕ(0)(2k−M+ℓ·)Φ(2k+ℓ·)ϕkHσ
2
.
The support assumptions of ϕ(0) and ϕ imply that when k ≥ M , ϕ(0)(2k−M+ℓ·)ϕ 6= 0 if and only
if k + ℓ = M or k + ℓ = M + 1. Then by (2.6), we have
kϕ(0)(2k−M+ℓ·)Φ(2k+ℓ·)ϕkHσ
2 ≤ C2kΦ(2k+ℓ·)ϕkHσ
2
,
where C2 depends only on ϕ(0), σ and d. Thus,
kh(2k−M+ℓ·)Φ(2k+ℓ·)ϕkHσ
2 ≤ (1 + C2)kΦ(2k+ℓ·)ϕkHσ
2
.
Similarly, we have
Putting all the estimates that we have obtained so far together, we get
kh(2k−M·)Φ(0)(2k·)ϕkHσ
2 ≤ (1 + C2)kΦ(0)(2k·)ϕkHσ
2
.
kfkF α,c
p ≤ C3(cid:16)C1(1 + C2)
p,Φ(cid:17),
+ kfkF α,c
2M(α0−α)
1 − 2α0−α sup
k≥M
2−kα0 max{kΦ(2k·)ϕkHσ
2
,kΦ(0)(2k·)ϕkHσ
2 }kfkF α,c
p
where the three constants C1, C2, C3 are independent of M , so we could take M large enough to
make sure the multiple of kfkF α,c
2 , so that we have
above is less than 1
p
kfkF α,c
p . kfkF α,c
p,Φ
.
Inhomogeneous Triebel-Lizorkin spaces
29
Step 4. We now settle the convergence issue of the second series in (4.4). For every j ≥ 0,
Φj ∗ ϕj+k ∗ f is an L1(M) + M-valued tempered distribution on Rd. We now show that the series
converges in S′(Rd; L1(M) + M). By (4.12) and (4.13), for any L > K, we have
kΦj ∗ ϕj+k ∗ fkp
2jα
LXk=K
. kIα0 ϕkHσ
. kfkF α,c
.
p
2 Xk≥K
2k(α0−α) sup
k∈N0
max(cid:8)2−kα0kΦ(2k·)ϕkHσ
2
, 2−kα0kΦ(0)(2k·)ϕkHσ
2(cid:9)kfkF α,c
p
Therefore, for any j ≥ 0, Pk≥K+1 Φj ∗ ϕj+k ∗ f converges in Lp(N ), so in S′(Rd; L1(M) + M)
too. In the same way, we can show that the series also converges in F α,c
the proof.
(Rd,M), which completes
(cid:3)
p
The following is the continuous analogue of Theorem 4.1. We use similar notation for continuous
parameters: given ε > 0, Φε denotes the function whose Fourier transform is Φ(ε) = Φ(ε·).
Theorem 4.2. Keep the assumption of the previous theorem. Then for f ∈ S′(Rd; L1(M) + M),
we have
(4.15)
kfkF α,c
p ≈ kΦ0 ∗ fkp +(cid:13)(cid:13)(cid:13)(cid:0)Z 1
0
2(cid:13)(cid:13)(cid:13)p
ε (cid:1) 1
ε−2αΦε ∗ f2 dε
.
Proof. This proof is very similar to the previous one. We keep the notation there and only point
out the necessary modifications. First, we need to discretize the integral on the right hand side of
(4.15). There exist two constants C1, C2 such that
C1
∞Xj=0
22jαZ 2−j
2−j−1 Φε ∗ f2 dε
ε ≤Z 1
0
ε−2αΦε ∗ f2 dε
ε ≤ C2
∞Xj=0
22jαZ 2−j
2−j−1 Φε ∗ f2 dε
ε
.
By approximation, we can assume that f is good enough so that each integral over the interval
(2−j−1, 2−j) can be approximated uniformly by discrete sums. Instead of Φ(j)(ξ) = Φ(2−jξ), we
have now Φ(ε)(ξ) = Φ(εξ) with 2−j−1 < ε ≤ 2−j. We transfer the split (4.5) into:
Φ(ε)(ξ)ϕj+k(ξ) =
Φ(2−j · 2jεξ)ϕ(0)(2−Kξ)
2−jξα1
2−jξα1 ϕj+k(ξ).
Thus,
with
Φε ∗ ϕj+k ∗ f = 2kα1 ηj ∗ ρj+k ∗ f
η(ξ) =
Φ(2jεξ)ϕ(0)(2−K ξ)
ξα1
and ρ(ξ) = ξα1 ϕ(ξ).
We proceed as in step 1 of the previous theorem, where we transfer (4.7) to the present setting:
kη(−k)ϕkHσ
2
(1 + t2)ση(t)dt
. Cϕ(0) ,σ,kZRd
= Cϕ(0) ,σ,kZRd
(1 + t2)σ(cid:12)(cid:12)F−1(cid:0)I−α1 Φ(δj·)ϕ(0)(2−K·)(cid:1)(t)(cid:12)(cid:12)dt
j ZRd
j 2−K·)(cid:1)(t)(cid:12)(cid:12)dt,
(1 + t2)σ(cid:12)(cid:12)F−1(cid:0)I−α1 Φϕ(0)(δ−1
≤ Cϕ(0) ,σ,kδα1
30
R. Xia and X. Xiong
where δj = 2jε and 1
2 < δj ≤ 1. The last integral is estimated as follows:
ZRd
j 2−K·)(cid:1)(t)(cid:12)(cid:12)dt
(1 + t2)σ(cid:12)(cid:12)F−1(cid:0)I−α1 Φϕ(0)(δ−1
≤ZRd
(1 + t2)σF−1(I−α1 Φϕ(0))(t)dt
+ZRd
(1 + t2)σ(cid:12)(cid:12)F−1(cid:0)I−α1 Φ[ϕ(0) − ϕ(0)(δ−1
≤ZRd
(1 + t2)σF−1(I−α1 Φϕ(0))(t)dt
2 <δ≤1ZRd
+ sup
1
j 2−K·)](cid:1)(t)(cid:12)(cid:12)dt
(1 + t2)σ(cid:12)(cid:12)F−1(cid:0)I−α1 Φ[ϕ(0) − ϕ(0)(δ−12−K·)](cid:1)(t)(cid:12)(cid:12)dt.
0
. Xk≤K−1
Xk≤K−1(cid:13)(cid:13)(cid:13)(cid:0)Z 1
Note that the above supremum is finite since I−α1 Φ[ϕ(0)−ϕ(0)(δ−12−K·)] is a compactly supported
and infinitely differentiable function whose inverse Fourier transform depends continuously on δ.
Then it follows that for 2−j−1 ≤ ε ≤ 2−j,
2(cid:13)(cid:13)(cid:13)p
ε (cid:1) 1
ε−2αΦε ∗ f2 dε
kΦ0 ∗ fkp +(cid:13)(cid:13)(cid:13)(cid:0)Z 1
We make similar modifications in step 2 of the previous theorem and then establish the third part.
Moreover, by the previous theorem, kΦ0 ∗ fkp . kfkF α,c
2(cid:13)(cid:13)(cid:13)p
ε (cid:1) 1
ε−2αΦε ∗ f2 dε
there exists 2 < a ≤ 2√2 such that Φ(ξ) > 0 on {ξ : a−1 ≤ ξ ≤ a}. Then for j ≥ 1, Rj = {ε :
a−12−j+1 < ε ≤ a2−j−1} are disjoint sub intervals on (0, 1], and ϕ(j)
Φ(ε) is well-defined for any ε ∈ Rj.
We slightly modify (4.14) as follows: for any ε ∈ Rj , we have
For the reverse inequality, we follow the argument in step 3 in the previous proof. By (4.2),
2k(α1−α)kfkF α,c
p . kfkF α,c
. Thus, we have proved
. kfkF α,c
p
0
.
.
p
p
ϕ(j)(ξ) = ϕ(j)ϕ(0)(2−j−K ξ) =
ϕ(j)(ξ)
Φ(ε)(ξ)
ϕ(0)(2−j−K ξ)Φ(ε)(ξ),
j ∈ N0.
Since for any −2 ≤ ℓ ≤ 2,
kΦ−1(2−jε−12ℓ·)ϕ(2ℓ·)ϕkHσ
2 ≤ sup
2a−1≤δ≤ a
2
and
k(Φ(0))−1(2−jε−1·)ϕ(0)(ϕ(0) + ϕ(1))kHσ
2 ≤ sup
2a−1≤δ≤ a
2
kΦ−1(δ2ℓ·)ϕ(2ℓ·)ϕkHσ
2
< ∞
k(Φ(0))−1(δ·)ϕ(0)(ϕ(0) + ϕ(1))kHσ
2
< ∞,
we follow the argument in step 3 in the previous theorem to get
kfkF α,c
p .(cid:13)(cid:13)(cid:0)Xj≥0
. kΦ0 ∗ fkp +(cid:13)(cid:13)(cid:13)(cid:0)Z 1
22jαZRj (ϕ0)j+k ∗ Φε ∗ f2(cid:1) 1
2(cid:13)(cid:13)p
2(cid:13)(cid:13)(cid:13)p
ε (cid:1) 1
ε−2αΦε ∗ f2 dε
0
+(cid:13)(cid:13)(cid:16)Xj≥0
22jαZRj hj+k ∗ Φε ∗ f2 dε
2(cid:13)(cid:13)(cid:13)p
ε (cid:1) 1
.
(cid:3)
The remaining of the proof follows step 3 with necessary modifications.
We now concretize the general characterization in the previous theorem to the case of Poisson
kernel. Recall that P denotes the Poisson kernel of Rd and
Pε(f )(s) =ZRd
Pε(s − t)f (t)dt,
(s, ε) ∈ Rd+1
+ .
The following theorem improves [35, Section 2.6.4] even in the classical case: [35, Section 2.6.4]
requires k > d + max{α, 0} for the Poisson characterization while we only need k > max{α, 0}.
The proof of this theorem is similar to but easier than that of [40, Theorem 4.20], since we assume
k > 0 here; we omit the details. The key ingredient is the improvement of the characterization of
Hardy spaces in terms of Poisson kernel given in [38, Theorem 1.5]
Inhomogeneous Triebel-Lizorkin spaces
31
Theorem 4.3. Let 1 ≤ p < ∞, α ∈ R, and k ∈ N such that k > max{α, 0}. Assume that Φ(0)
satisfies (4.1). Then for f ∈ S′(Rd; L1(M) + M), we have
kfkF α,c
p ≈ kΦ0 ∗ fkp +(cid:13)(cid:13)(cid:13)(cid:0)Z 1
0
ε2(k−α)(cid:12)(cid:12) ∂k
2(cid:13)(cid:13)(cid:13)p
∂εk Pε(f )(cid:12)(cid:12)2 dε
ε (cid:1) 1
.
4.2. Characterizations via Lusin functions. We are going to give some characterizations for
Triebel-Lizorkin spaces via Lusin square functions. As what we did in the previous part of this
section, we still use Fourier multiplier theorems as our main tool. But now we have to rely on the
Hilbertian (instead of ℓ2) versions of the Fourier multiplier theorems.
The following characterization, via Lusin square functions associated to ϕ given by the condition
(0.1), is a special case of the characterization in Theorem 1.5. We keep using the notation ϕj being
the function whose Fourier transform is equal to ϕ(2−j·) for j ∈ N, and ϕ0 being the function
whose Fourier transform is equal to 1 −Pj≥1 ϕ(2−j·).
Proposition 4.4. For 1 ≤ p < ∞ and f ∈ F α,c
(4.16)
p
kfkF α,c
p ≈ kϕ0 ∗ fkp +(cid:13)(cid:13)(cid:13)(Xj≥1
(Rd,M), we have
2j(2α+d)ZB(0,2−j ) ϕj ∗ f (· + t)2dt)
.
1
2(cid:13)(cid:13)(cid:13)p
(Rd,M), by the lifting property in Proposition 3.4, we have J αf ∈
Proof. For any f ∈ F α,c
p(Rd,M). Then, we apply the discrete characterization in Theorem 1.5 with φ = J−αϕ0 and
hc
Φ = I−αϕ to J αf ,
p
Following the argument in the proof of (2.9), we can prove
kfkF α,c
p ≈ kJ αfkhc
p ≈ kϕ0 ∗ fkp + ksc,D
I−αϕ(J αf )kp.
Moreover, we can easily check that
.
I−αϕ(I αf )(cid:13)(cid:13)p
(cid:13)(cid:13)sc,D
I−αϕ(J αf )(cid:13)(cid:13)p ≈(cid:13)(cid:13)sc,D
I−αϕ(I αf )(cid:13)(cid:13)p =(cid:13)(cid:13)(cid:13)(Xj≥1
(cid:13)(cid:13)sc,D
p ≈ kϕ0 ∗ fkp +(cid:13)(cid:13)(cid:13)(Xj≥1
2j(2α+d)ZB(0,2−j ) ϕj ∗ f (· + t)2dt)
2(cid:13)(cid:13)(cid:13)p
2j(2α+d)ZB(0,2−j ) ϕj ∗ f (· + t)2dt)
2(cid:13)(cid:13)(cid:13)p
.
1
1
kfkF α,c
Therefore, we conclude
The assertion is proved.
.
(cid:3)
From the above Lusin square function by ϕ, we can deduce Lusin type characterizations with
general convolution kernels by the aide of Theorems 2.5 and 2.9.
Theorem 4.5. Let 1 ≤ p < ∞ and α ∈ R. Assume that α0 < α < α1, α1 ≥ 0 and Φ(0), Φ satisfy
conditions (4.1), (4.2). Then for any f ∈ S′(Rd; L1(M) + M), we have
kfkF α,c
p
(Rd,M) ≈ kΦ0 ∗ fkp +(cid:13)(cid:13)(cid:13)(Xj≥1
where the equivalent constant is independent of f .
2j(2α+d)ZB(0,2−j ) Φj ∗ f (· + t)2dt)
,
1
2(cid:13)(cid:13)(cid:13)p
Proof. This proof is very similar to that of Theorem 4.1. The main target is to replace the standard
test functions ϕ0 and ϕ in Proposition 4.4 with Φ0 and Φ satisfying (4.1) and (4.2). This time we
need to use the Lusin type multiplier theorem i.e. Theorem 2.5, instead of Theorem 2.2. For the
special case p = 1, we apply Theorem 2.9 instead of Theorem 2.7.
(cid:3)
Using a similar argument as in Theorem 4.2, we also have the following continuous analogue of
the above theorem. This is the general characterization of Triebel-Lizorkin spaces by Lusin square
functions. Recall thateΓ = {(t, ε) ∈ Rd+1
+ : t < ε < 1}.
32
R. Xia and X. Xiong
Theorem 4.6. Keep the assumption in the previous theorem. Then for any L1(M) + M-valued
tempered distribution f on Rd, we have
kfkF α,c
p
(Rd,M) ≈ kΦ0 ∗ fkp +(cid:13)(cid:13)(cid:13)(cid:0)ZeΓ
2(cid:13)(cid:13)(cid:13)p
εd+1(cid:1) 1
ε−2αΦε ∗ f (· + t)2 dtdε
.
5. Smooth atomic decomposition
1
This section is devoted to the study of atomic decomposition of F α,c
(Rd,M). We aim to
decompose F α,c
(Rd,M) into atoms which have good enough size, smooth and moment conditions.
To proceed in an orderly way step by step, we begin with the special case α = 0, i.e., the space
hc
1(Rd,M) below does not lead to the one for general
1(Rd,M). Even though the result for hc
F α,c
(Rd,M) are
already contained in those for hc
1(Rd,M). The main results in this section will be very useful in
our forthcoming paper [37] on mapping properties of pseudo-differential operators.
(Rd,M) directly, the main ingredients to obtain smooth decomposition for F α,c
1
1
1
5.1. Smooth atomic decomposition of hc
1(Rd,M). In the classical theory, the smooth atoms
have been widely studied and have played a crucial role when studying the mapping properties of
pseudo-differential operators acting on local Hardy spaces, or more generally, on Triebel-Lizorkin
spaces. Details can be found in [3], [4] and [35].
In this subsection, we will show that in our
operator-valued case, the atoms in Theorem 1.6 can also be refined to be infinitely differentiable.
As in the classical case, the theory of tent spaces will be of great service in our proof of smooth
atomic decomposition theorem. Tent spaces in the operator-valued setting have been introduced
in [18] and [19] first; see also [38] for further complement. For our use, we study the local version
of tent spaces defined in [36]. For any function defined on the strip S = Rd × (0, 1) with values in
L1(M) + M, whenever it exists, we define
First, we introduce a lemma concerning the atomic decomposition of the tent space T c
1 (Rd,M).
, s ∈ Rd.
2
For 1 ≤ p < ∞, we define
equipped with the norm kfkT c
Ac(f )(s) =(cid:16)ZeΓ f (t + s, ε)2 dtdε
εd+1(cid:17) 1
p (Rd,M) = {f : Ac(f ) ∈ Lp(N )}
T c
p (Rd,M) = kAc(f )kp.
A function a ∈ L1(cid:0)M; L2(S, dsdε
ε )(cid:1) is called a T c
• τ(cid:16)RT (Q) a(s, ε)2 dsdε
ε (cid:17) 1
≤ Q− 1
2 .
• supp a ⊂ T (Q) for some cube Q in Rd with Q ≤ 1;
1 -atom if
Let T c
2
(5.1)
f =
λj aj,
1,at(Rd,M) be the space of all f : S → L1(M) admitting a representation of the form
∞Xj=1
∞Xj=1
where the aj's are T c
the following norm
1 -atoms and λj ∈ C such that P∞j=1 λj < ∞. We equip T c
1,at(Rd,M) with
kfkT c
1,at = inf{
λj : f =
λjaj; aj's are T c
1 -atoms, λj ∈ C}.
∞Xj=1
Lemma 5.1. We have T c
Proof. In order to prove T c
satisfies kakT c
1
1,at(Rd,M) = T c
1,at(Rd,M) ⊂ T c
. 1. By the support assumption of a, we have
1 (Rd,M) with equivalent norms.
1 (Rd,M), it is enough to show that any T c
1 -atom a
kakT c
1 =(cid:13)(cid:13)Ac(a)(cid:13)(cid:13)1 = τZRd(cid:16)Z 1
0 ZB(t,ε) a(s, ε)2 dsdε
εd+1(cid:17) 1
2 τ(cid:16)ZRdZ 1
0 ZB(t,ε) a(s, ε)2 dsdε
2 τ(cid:16)ZT (Q) a(t, ε)2 dtdε
ε (cid:17) 1
. Q
d Q
εd+1
= c
1
2
2
1
1
2
dt
. 1.
2
dt(cid:17) 1
Inhomogeneous Triebel-Lizorkin spaces
33
Then by the duality T c
1 (Rd,M)∗ = T c
∞(Rd,M) (see [36]), we have T c
∞(Rd,M) ⊂ T c
1,at(Rd,M)∗.
Now let Q be a cube in Rd with Q ≤ 1. If f ∈ L1(cid:0)M; Lc
2(T (Q), dsdε
ε )(cid:1), then
is a T c
2(T (Q), dsdε
2kfk−1
a = Q− 1
1 -atom supported in T (Q). Hence,
ε )(cid:1)f
ε )(cid:1).
1,at induces a continuous functional on L1(cid:0)M; Lc
L1(cid:0)M;Lc
2kfkL1(cid:0)M;Lc
1,at ≤ Q
1,at(Rd,M) for every cube Q. Therefore, every continuous func-
Thus, L1(cid:0)M; Lc
tional ℓ on T c
than or equal to Q 1
1,at)∗ . Let Q0 be the cube centered at the origin with side length 1 and
Qm = Q0 + m for each m ∈ Zd. Then Rd = ∪m∈Zd Qm. Consequently, we can choose a sequence
of functions (gm)m∈Zd such that
ε )(cid:1) with norm smaller
ε )(cid:1) ⊂ T c
2(T (Q), dsdε
2(T (Q), dsdε
2kℓk(T c
kfkT c
2(T (Q), dsdε
1
ℓ(a) = τZT (Qm)
a(s, ε)g∗m(s, ε)
dsdε
ε
,
∀ T c
1 -atom a with supp a ⊂ T (Qm),
and
1,at)∗ .
2(T (Qm), dsdε
kgmkL∞(cid:0)M;Lc
ℓ(a) = τZS
Let g(s, ε) = gm(s, ε) for (s, ε) ∈ T (Qm). Then, we have
dsdε
ε )(cid:1) ≤ kℓk(T c
∀ T c
It follows that, for any cube Q with Q ≤ 1, gT (Q) ∈ L∞(cid:0)M; Lc
ε )(cid:1) ≤ Q
2(T (Qm), dsdε
1,at(Rd,M)∗ ⊂ T c
which implies g ∈ T c
1,at(Rd,M)∗ with equivalent norms. Finally, by the density of T c
T c
get the desired equivalence.
kgT (Q)kL∞(cid:0)M;Lc
∞(Rd,M). Hence, T c
a(s, ε)g∗(s, ε)
ε
,
1
1 -atom a.
2(T (Q), dsdε
ε )(cid:1) and
1,at)∗ ,
2kℓk(T c
∞(Rd,M). Therefore, T c
1,at(Rd,M) in T c
∞(Rd,M) =
1 (Rd,M), we
(cid:3)
The following Lemma shows the connection between T c
modelled on the classical argument of [3, Theorem 6].
p (Rd,M) and hc
p(Rd,M). The proof is
Lemma 5.2. Fix a Schwartz function Φ on Rd satisfying:
Φ is supported in the cube with side length 1 and centered at the origin;
(5.2)
RRd Φ(s)ds = 0;
Let πΦ be the map given by
Φ is nondegenerate in the sense of (1.2).
Then πΦ is bounded from T c
πΦ(f )(s) =Z 1
0 ZRd
p (Rd,M) to hc
dtdε
,
Φε(s − t)f (t, ε)
p(Rd,M) for any 1 ≤ p < ∞.
ε
s ∈ Rd.
Proof. For any 1 < p < ∞, let q be its conjugate index. By Theorem 1.1, it suffices to estimate
τR πΦ(f )(s)g∗(s)ds, for any g ∈ hc
q(Rd,M). Note that
τZRd
πΦ(f )(s)g∗(s)ds = τZRdZ 1
= τZRdZ 1
0
0
Φε(s − t)f (t, ε)
dtdε
ε
g∗(s)ds
f (t, ε)(eΦε ∗ g)∗(t)
dεdt
ε
,
34
R. Xia and X. Xiong
where eΦ(s) = Φ(−s). Then by the Holder inequality,
cd(cid:12)(cid:12)(cid:12)τZRdZ 1
cd(cid:12)(cid:12)(cid:12)τZRdZeΓ
πΦ(f )(s)g∗(s)ds(cid:12)(cid:12)(cid:12) =
(cid:12)(cid:12)(cid:12)τZRd
1
1
=
. kAc(f )kpksc
pkgkhc
. kfkT c
q
eΦ
.
0 ZB(s,ε)
f (t, ε)(eΦε ∗ g)∗(t)
f (s + t, ε)(eΦε ∗ g)∗(s + t)
(g)kq
dεdt
εd+1
dεdt
εd+1
ds(cid:12)(cid:12)(cid:12)
ds(cid:12)(cid:12)(cid:12)
Now we deal with the case p = 1. The argument below is based on the atomic decompositions
1(Rd,M) and T c
of hc
1 -atom to
a bounded multiple of an hc
1 based on some cube Q with Q ≤ 1.
Since Φ is supported in the unit cube, it follows from the definition of πΦ that πΦ(a) is supported
1 (Rd,M). By Lemma 5.1, it is enough to show that πΦ maps a T c
in 2Q. Moreover, it satisfies the moment cancellation thatR πΦ(a)(s)ds = 0 since bΦ(0) = 0. So it
remains to check that πΦ(a) satisfies the size estimate. To this end, we use the Cauchy-Schwarz
inequality and the Plancherel formula (0.7),
1-atom. Let a be an atom in T c
2
dε
kπΦ(a)kL1(M;Lc
2(Rd)) = τ(cid:0)ZRd \πΦ(a)(ξ)2dξ(cid:1) 1
= τ(cid:16)ZRd Z 1
ε 2dξ(cid:17) 1
0 bΦ(εξ)ba(ξ, ε)
≤ τ(cid:16)ZRdZ 1
ε Z 1
dξ(cid:17) 1
0 bΦ(εξ)2 dε
0 ba(ξ, ε)2 dε
≤ τ(cid:16)ZT (Q) a(s, ε)2 dsdε
ε (cid:17) 1
≤ Q− 1
2 .
ε
2
2
2
(5.3)
(5.4)
where
(5.5)
(5.6)
(5.7)
Therefore we obtain the boundedness of πΦ from T c
1,at(Rd,M).
Now we are able to refine the smoothness of the atoms given in Theorem 1.6.
1,at(Rd,M) to hc
(cid:3)
Theorem 5.3. For any f ∈ L1(M; Rc
can be represented as
d) + L∞(M; Rc
d), f belongs to hc
1(Rd,M) if and only if it
• the bj's are infinitely differentiable atoms supported in 2Q0,j with Q0,j = 1. For any
0, there exists a constant Cγ which depends on γ satisfying
multiple index γ ∈ Nd
• the gj's are infinitely differentiable atoms supported in 2Qj with Qj < 1, and such that
f =
∞Xj=1
(µjbj + λjgj),
τ(cid:0)Z2Q0,j Dγ bj(s)2ds(cid:1) 1
2 ≤ Cγ;
2 . Qj− 1
2
and
gj(s)ds = 0;
Z2Qj
• the coefficients µj and λj are complex numbers such that
(µj + λj) < ∞.
τ(cid:0)Z2Qj gj(s)2ds(cid:1) 1
∞Xj=1
1(Rd,M).
Moreover, the infimum of (5.7) with respect to all admissible representations gives rise to an
equivalent norm on hc
Proof. Since the bj's and gj's are atoms in hc
can be represented as in (5.4) and
1(Rd,M), it suffices to show that any f ∈ hc
1(Rd,M)
∞Xj=1
(µj + λj) . kfkhc
1
.
Inhomogeneous Triebel-Lizorkin spaces
35
To begin with, we construct a smooth resolution of the unit on Rd. Let κ be a radial, real and
infinitely differentiable function on Rd which is supported in the unit cube centered at the origin.
Moreover, we assume thatbκ(0) > 0. We take bΦ = · 2bκ, which can be normalized as:
= 1,
ξ ∈ Rd\{0}.
And we define
(5.8)
0
ε
Z ∞
bΦ(εξ)2 dε
bφ(ξ) = 1 −Z 1
0 bΦ(εξ)2 dε
ε
,
ξ ∈ Rd.
By the Paley-Wiener theorem, bΦ can be extended to an analytic function bΦ(z) of d complex
variables z = (z1, . . . , zd), and for any λ > 0, there exists a constant Cλ such that
√d
holds for any z = ξ1 + iξ2. Therefore,
Z 1
0 bΦ(εz)2 dε
4 +
eε( λ
2 )ξ2(ξ12 + ξ22)
bΦ(z) ≤ Cλe( λ
λZ 1
2 +√d)ξ2ε3dε · (ξ12 + ξ22)2
ε ≤ C2
λZ 1
2 +√d)ξ2(ξ12 + ξ22)2
ε3dε · e( λ
≤ C2
2 +√d)ξ2(1 + ξ12)2(1 + ξ22)2
λe( λ
≤ C2
λe(λ+2√d)ξ2(1 + ξ1)4.
≤ C2
0
0
Now applying the Paley-Wiener-Schwartz theorem to distributions, we obtain that φ is a distribu-
tion with support in {s ∈ Rd : s ≤ 2√d}. On the other hand, if we define its value at the origin
as 0, the function R 1
ε is an infinitely differentiable function on Rd, which ensures that φ
is a Schwartz function. Thus, supp φ ⊂ {s ∈ Rd : s ≤ 2√d}. By (5.8), we arrive at the following
0 bΦ(ε·)2 dε
decomposition of f :
(5.9)
f = φ ∗ f +Z 1
0
Φε ∗ Φε ∗ f
dε
ε
.
We first deal with φ ∗ f . By Theorem 1.6, we obtain an atomic decomposition of f :
(5.10)
where the aj's are hc
1-atoms andPj eµj . kfkhc
f =Xj eµjaj,
φ ∗ f =Xj eµj φ ∗ aj.
1. Thus,
2Q0 (with Q0 the unit cube centered at the origin), andPk∈Zd X0(s − k) = 1 for every s ∈ Rd.
We now show that every φ∗ aj can be decomposed into smooth atoms supported in cubes with side
length two. Let X0 be a nonnegative infinitely differentiable function on Rd such that suppX0 ⊂
See [32, Section VII.2.4] for the existence of such X0. Set Xk = X0(· − k). Then Xk is supported
in the cube 2Qk = k + 2Q0, and all Xk's form a smooth resolution of the unit:
(5.11)
1 = Xk∈Zd Xk(s),
∀ s ∈ Rd.
Take a to be one of the atoms in (5.10) supported in Q. Since φ has compact support, i.e. there
exists a constant C such that supp φ ⊂ CQ0, then φ ∗ a is supported in (C + 1)Q0. Thus, we get
the decomposition
φ ∗ a =
NXk=1
bk with bk = Xk · (φ ∗ a),
36
R. Xia and X. Xiong
where N is a positive integer depending only on the dimension d and C. For any β, γ ∈ Nd
β ≤ γ if βj ≤ γj for every 1 ≤ j ≤ d. Then, by the Cauchy-Schwarz inequality, for any k,
0, denote
2
2
τ(cid:0)ZRd Dγ bk(s)2ds(cid:1) 1
2 .Xβ≤γ
.Xβ≤γ
≤Xβ≤γ(cid:0)ZQZ2Qk Dβφ(s − t)2dsdt(cid:1) 1
τ(cid:0)Z2Qk Dβφ ∗ a(s) · Dγ−βXk(s)2ds(cid:1) 1
τ(cid:0)Z2Qk ZRd
Dβφ(s − t)a(t)dt2ds(cid:1) 1
2 τ(cid:0)ZQ a(t)2dt(cid:1) 1
φ ∗ f =Xj
with bj as desired. Furthermore,Pj µj . kfkhc
2 ≤ 1,
. Q
µjbj,
1.
1
definition of the tent space and Theorem 1.4 that Φε ∗ f ∈ T c
. kfkhc
kφ ∗ fk1 + kΦε ∗ fkT c
1
1
.
1 (Rd,M) and
2 · τ(cid:0)ZQ a(t)2dt(cid:1) 1
2
where the relevant constants depend only on γ, φ and X0. Thus, we have proved that φ ∗ f can be
decomposed as follows:
Now it remains to deal with the second term on the right hand side of (5.9). It follows from the
1 -atoms based on cubes with side length less than or equal to 1. For each
By Lemma 5.1, we decompose Φε ∗ f as follows:
(5.12)
Φε ∗ f (s) =
∞Xj=1
where the eaj's are T c
eaj(s, ε) based on Qj in (5.12), we set
gj(s) =Z 1
(5.13)
0
λjeaj(s, ε) with
dε
ε
∞Xj=1
λj . kΦε ∗ fkT c
1
,
We observe from the proof of Lemma 5.2 that gj is a bounded multiple of an hc
1-atom supported
in 2Qj with vanishing mean. Moreover, gj is infinitely differentiable. Thus, gj satisfies (5.6) with
relevant constant depending only on Φ. Combining (5.12) and (5.13), we obtain the decomposition
Φε ∗eaj(s, ε)
= πΦeaj(s),
∀s ∈ Rd.
Z 1
0
Φε ∗ Φε ∗ f
dε
ε
=
λj gj,
∞Xj=1
1. The proof is complete.
withP∞j=1 λj . kfkhc
5.2. Atomic decomposition for F α,c
(Rd,M).
For every l = (l1,··· , ld) ∈ Zd, µ ∈ N0, we define Qµ,l in Rd to be the cubes centered at 2−µl, and
with side length 2−µ. For instance, Q0,0 = [− 1
2 )d is the unit cube centered at the origin. Let Dd
be the collection of all the cubes Qµ,l defined above. We write (µ, l) ≤ (µ′, l′) if
(Rd,M). Now we turn to the general space F α,c
2 , 1
(cid:3)
1
1
µ ≥ µ′
and Qµ,l ⊂ 2Qµ′,l′ .
For a ∈ R, let a+ = max{a, 0} and [a] the largest integer less than or equal to a. Recall that
0 and J α is the Bessel potential
γ1 = γ1 + ··· + γd for γ ∈ Nd
of order α.
Definition 5.4. Let α ∈ R, and let K and L be two integers such that
K ≥ ([α] + 1)+ and L ≥ max{[−α],−1}.
for s ∈ Rd, β ∈ Nd
1 ··· sβd
0, sβ = sβ1
d
• supp b ⊂ 2Q0,k;
(1) A function b ∈ L1(cid:0)M; Lc
• τ(cid:0)RRd Dγ b(s)2ds(cid:1) 1
(2) Let Q = Qµ,l ∈ Dd, a function a ∈ L1(cid:0)M; Lc
2 ≤ 1, ∀γ ∈ Nd
2(Rd)(cid:1) is called an (α, 1)-atom if
• supp a ⊂ 2Q;
0 , γ1 ≤ K.
2(Rd)(cid:1) is called an (α, Q)-subatom if
Inhomogeneous Triebel-Lizorkin spaces
37
• τ(cid:0)RRd Dγ a(s)2ds(cid:1) 1
2 ≤ Q α
• RRd sβa(s)ds = 0, ∀β ∈ Nd
2(Rd)(cid:1) is called an (α, Qk,m)-atom if
(3) A function g ∈ L1(cid:0)M; Lc
τ(cid:0)ZRd J αg(s)2ds(cid:1) 1
d − γ1
0 , β1 ≤ L.
2 . Qk,m− 1
d , ∀γ ∈ Nd
0 , γ1 ≤ K;
and g = X(µ,l)≤(k,m)
(5.14)
2
dµ,laµ,l,
for some k ∈ N0 and m ∈ Zd, where the aµ,l's are (α, Qµ,l)-subatoms and the dµ,l's are complex
numbers such that
(cid:0) X(µ,l)≤(k,m)
dµ,l2(cid:1) 1
2 ≤ Qk,m− 1
2 .
Remark 5.5. If L < 0, the third assumption of an (α, Q)-subatom means that no moment
cancellation is required.
In the second assumption of an (α, 1)-atom b and that of an (α, Q)-
subatom a, it is tacitly assumed that b and a have derivatives up to order K. For such a, we can
define a norm by
kak∗ = sup
γ1≤K kDγakL1(cid:0)M;Lc
2(Rd)(cid:1).
Then the convergence in (5.14) is understood in this norm, and we will see that the atom g in
(5.14) belongs to F α,c
1
(Rd,M).
Remark 5.6. In the classical case, the first size estimate in (5.14) is not necessary.
words, if g =P(µ,l)≤(k,m) dµ,laµ,l with the subatoms aµ,l's and the complex numbers dµ,l's such
that(cid:0)P(µ,l)≤(k,m) dµ,l2(cid:1) 1
2 , then g satisfies that estimate in (5.14) automatically. We
refer the readers to [35] for more details. Unfortunately, in the current setting, we are not able to
prove this estimate, so we just add it in (5.14) for safety.
2 ≤ Qk,m− 1
In other
The following is our main result on the atomic decomposition of F α,c
(Rd,M). The idea comes
from [35, Theorem 3.2.3], but many techniques used are different from those of [35, Theorem 3.2.3]
due to noncommutativity.
Theorem 5.7. Let α ∈ R and K, L be two integers fixed as in Definition 5.4. Then any f ∈
F α,c
1
1
(Rd,M) can be represented as
(5.15)
f =
(5.16)
where the bj's are (α, 1)-atoms, the gj's are (α, Q)-atoms, and µj, λj are complex numbers with
∞Xj=1(cid:0)µjbj + λj gj(cid:1),
∞Xj=1
(µj + λj) < ∞.
∞Xj=1
(µj + λj) . kfkF α,c
1
.
Moreover, the infimum of (5.16) with respect to all admissible representations is an equivalent
norm in F α,c
Proof. Step 1. First, we show that any f ∈ F α,c
(Rd,M) admits the representation (5.15) and
(Rd,M).
1
1
The proof of this part is similar to the proof of Theorem 5.3. Let κ be the Schwartz function
that N ≥ max{L, α}, then Φ can be normalized as follows:
defined in the proof of Theorem 5.3. We take bΦ = · Nbκ, where N is a positive even integer such
Z ∞
bΦ(εξ)2 dε
(J−αbΦ)(2−jξ)2 andP∞j=−∞
Since −α+N ≥ 0, bothP∞j=−∞
(J−αbΦ)(2−jξ)2 are rapidly decreasing
∞Xj=−∞
(J−αbΦ)(2−jξ)2 < ∞
and infinitely differentiable functions on Rd \ {0}. So we have
(5.17)
∀ ξ ∈ Rd \ {0}.
= 1,
ε
0
38
and
(5.18)
with
Applying the Paley-Wiener-Schwartz theorem, we get a compactly supported function Φ0 ∈ S such
that
Denote by Φε the Fourier inverse transform of Φ(ε·). For any f ∈ F α,c
(5.19)
1
.
(Rd,M), we have
Let us deal with the two terms on the right hand side of (5.19) separately.
The term Φ0∗f is easy to treat. If α ≥ 0, Proposition 3.3 ensures that F α,c
Then we can repeat the first part of the proof of Theorem 5.3: for any f ∈ F α,c
admits the decomposition
1
(Rd,M) ⊂ hc
1(Rd,M).
(Rd,M), Φ0 ∗ f
1
R. Xia and X. Xiong
(I−αbΦ)(2−jξ)2 < ∞.
0 bΦ(εξ)2 dε
∞Xj=−∞
bΦ0(ξ) = 1 −Z 1
f = Φ0 ∗ f +Z 1
Φε ∗ Φε ∗ f
ε
0
.
dε
ε
Φ0 ∗ f =Xj
µjbj,
Xj
µj . kfkhc
1
. kfkF α,c
1
,
µjbj,
. Then
J [α]Φ0 ∗ f =Xj
Φ0 ∗ f =Xj
µjJ−[α]bj.
where the bj's, together with their derivatives Dγ bj's, satisfy (5.5) with some constants Cγ de-
pending on γ. When K is fixed, we can normalize the bj's by maxγ1≤K Cγ, then the new bj's
are (α, 1)-atoms. If α < 0, by Propositions 3.3 and 3.4, we have J [α]f ∈ F α−[α],c
1. Then
J [α]Φ0 ∗ f admits the decomposition
⊂ hc
1
withPj µj . kJ [α]fkhc
1
. kfkF α,c
1
If −[α] is even, it is obvious that supp J−[α]bj ⊂ supp bj. Moreover, for any γ ∈ Nd
γ1 ≤ K, we have
0 such that
τ (ZRd Dγ J−[α]bj(s)2ds)
1
2 . Xγ′1≤K−2[α]
τ (ZRd Dγ′bj(s)2ds)
1
2 ≤ CK .
We normalize J−[α]bj by this constant CK depending on K, then we can make it an (α, 1)-atom.
When −[α] is odd, it suffices to replace [α] in the above argument by [α] − 1, and then we get the
desired decomposition.
It follows from
Step 2. Now we turn to the second term on the right hand side of (5.19).
1 (Rd,M) and
Theorem 4.6 and the definition of the tent space that ε−αΦε ∗ f ∈ T c
By Lemma 5.1, we have
(5.20)
1
kε−αΦε ∗ fkT c
∞Xj=1
ε−αΦε ∗ f (s) =
. kfkF α,c
1
.
λj bj(s, ε),
Φε ∗ f (s) =
∞Xj=1
λjaj(s, ε)
where the bj's are T c
εαbj(s, ε), we obtain
1 -atoms based on the cubes Qj's with Qj ≤ 1. Then, if we set aj(s, ε) =
and
(5.21)
In particular,
(5.22)
For every aj, we set
(5.23)
supp aj ⊂ T (Qj)
and τ(cid:16)ZT (Qj )
gj(s) = πΦ(aj)(s) =Z 1
0
ε (cid:17) 1
ε−2αaj(s, ε)2 dsdε
2
≤ Qj− 1
2 .
Φε ∗ aj(s, ε)
dε
ε
.
Then supp gj ⊂ 2Qj. We arrive at the decomposition
=
Φε ∗ Φε ∗ f
dε
ε
Z 1
0
λj gj.
∞Xj=1
Now we show that every gj is an (α, Qkj ,mj )-atom. Firstly, for any Qj, there exist kj ∈ N0 and
s ∈ Rd such that
2−kj−1 ≤ l(Qj) ≤ 2−kj
and cQj = l(Qj)s.
Take mj = [s] ∈ Zd, where [s] = ([s1],··· , [sd]). Then, we easily check that
(5.24)
Qj ⊂ 2Qkj ,mj , Qkj ,mj ∈ Dd.
Next, by the argument similar to that in (5.3) and by (5.22), we have
Inhomogeneous Triebel-Lizorkin spaces
39
∞Xj=1
λj . kε−αΦε ∗ fkT c
1
. kfkF α,c
1
.
If α ≤ 0, it is clear that
τ(cid:0)ZRd I απΦ(aj)(s)2ds(cid:1) 1
τ(cid:0)ZRd J απΦ(aj)(s)2ds(cid:1) 1
τ(cid:0)ZRd J απΦ(aj)(s)2ds(cid:1) 1
If α > 0, we have
2 .
2
≤ Qj− 1
2 . Qkj ,mj− 1
2 . Qkj ,mj− 1
2 . τ(cid:16)ZT (Qj )
ε (cid:17) 1
ε−2αaj(t, ε)2 dtdε
2 ≤ τ(cid:0)ZRd I απΦ(aj)(s)2ds(cid:1) 1
2 . Qj− 1
2 . τ(cid:0)ZRd πΦ(aj)(s)2ds(cid:1) 1
2 + τ(cid:0)ZRd I απΦ(aj)(s)2ds(cid:1) 1
. τ(cid:16)ZT (Qj ) aj(t, ε)2 dtdε
ε (cid:17) 1
+ Qj− 1
. τ(cid:16)ZT (Qj )
ε (cid:17) 1
ε−2αaj(t, ε)2 dtdε
≤ 2Qj− 1
2 . Qkj ,mj− 1
+ Qj− 1
2 .
2 .
2
2
2
2
2
Then we get, for any α ∈ R,
(5.25)
τ(cid:0)ZRd J αgj(s)2ds(cid:1) 1
2 = τ(cid:0)ZRd J απΦ(aj )(s)2ds(cid:1) 1
2 . Qkj ,mj− 1
2 .
Finally, we decompose the slice T (Qj) ∩ {2−µ−1 ≤ ε ≤ 2−µ} into (d + 1)-dimensional dyadic cubes
whose projections on Rd belong to Dd, and with side length 2−µ, µ ∈ N0. Let bQ be one of those
dyadic cubes with side length 2−µ and Q be its projection on Rd. Let
a(s) =Z bQ
Φε(s − t)aj(t, ε)
dtdε
.
ε
By the support assumption of Φ, it follows that
Then
supp a ⊂ 2Q,
supp a ⊂ 2Qj ⊂ 4Qkj ,mj .
ba(ξ) =Z 2−µ+1
2−µ
bΦ(εξ)F(cid:0)aj(·, ε)1Q(cid:1)(ξ)
dε
ε
.
40
R. Xia and X. Xiong
τ(cid:0)Z a(s)2ds(cid:1) 1
Since DβbΦ(0) = 0 for any β1 ≤ N , we obtain
Furthermore, by the Cauchy-Schwarz inequality, we have
∀β1 ≤ L.
ZRd
(−2πis)βa(s)ds = Dβba(0) = 0,
2 = τ(cid:16)Z5Q(cid:12)(cid:12)Z 2−µ+1
ZQ
ε (cid:12)(cid:12)(cid:12)
· τ(cid:16)Z 2−µ+1
2(cid:16)Z 2−µ+1
ZQ
ε (cid:17) 1
. τ(cid:16)Z 2−µ+1
ZQ aj(s, ε)2 dsdε
ε (cid:17) 1
Φε(s − t)aj(t, ε)
ε−2d dtdε
. Q
dtdε
2−µ
2−µ
2−µ
2−µ
1
2
2
2
2
ds(cid:17) 1
ZQ aj(t, ε)2 dtdε
ε (cid:17) 1
2
α
d τ(cid:16)Z 2−µ+1
2−µ
. Q
Similarly, we have
γ
τ(cid:0)Z D
a(s)2ds(cid:1) 1
2 ≤ C′γQ
The above discussion gives
ε (cid:17) 1
ε−2αaj(s, ε)2 dsdε
2
.
ZQ
d τ(cid:16)Z 2−µ+1
2−µ
α
d − γ1
ZQ
ε (cid:17) 1
ε−2αaj(s, ε)2 dsdε
2
.
(5.26)
where each aj
µ,l,
dj
µ,laj
gj = X(µ,l)≤(kj ,mj )
γ1≤K{C′γ}τ(cid:16)Z 2−µ+1
ZQµ,l
2 ≤ Cτ(cid:16)ZT (Qj )
µ,l2(cid:1) 1
dj
µ,l is an (α, Qµ,l)-subatom. The normalizing factor is given by
ε (cid:17) 1
ε−2αaj(s, ε)2 dsdε
ε (cid:17) 1
ε−2αaj(s, ε)2 dsdε
(cid:0) X(µ,l)≤(kj ,mj)
dj
µ,l = max
2), we get
2−µ
2
2
By the elementary fact that ℓ2(L1(M)) ⊃ L1(M; ℓc
(5.27)
.
≤ CQkj ,mj− 1
2 ,
where C is independent of f . We may assume C = 1, otherwise, we can put C in (5.20) in the
numbers λj , which does not change (5.21). In summary, (5.24), (5.25), (5.26) and (5.27) ensure
that gj is an (α, Qkj ,mj )-atom.
Step 3. Now we show the reverse assertion: if f is given by (5.15), then f ∈ F α,c
1
(Rd,M) and
kfkF α,c
1
.
(µj + λj).
∞Xj=1
To this end, we have to show that every (α, 1)-atom b and every (α, Q)-atom g belong to F α,c
and
1
(Rd,M)
Let b be an (α, 1)-atom in F α,c
α ≤ 0, by Proposition 3.3, hc
3.4, we have
1 ⊂ F α,c
1
1
1
. 1
kbkF α,c
(Rd,M). We observe that b is also an atom in hc
and kgkF α,c
. 1.
1
. Then, we have kbkF α,c
1
. kbkhc
1
1(Rd,M). For
. 1. If α > 0, by Proposition
Note that for any 1 ≤ i ≤ d, DK
we have
1(Rd,M). Since α − K < 0, by Proposition 3.3,
1 ≈ kϕ0 ∗ bk1 +
kbkF α,c
i b is an atom in hc
dXi=1
kDK
i bkF α−K,c
1
.
kbkF α,c
1
. kϕ0 ∗ bk1 +
dXi=1
kDK
i bkhc
1
. 1.
Inhomogeneous Triebel-Lizorkin spaces
41
On the other hand, let g be an (α, Qk,m)-atom in the sense of Definition 5.4. We may use the
discrete general characterization of F α,c
1
(Rd,M) given in Theorem 4.1, i.e.
1
kgkF α,c
22jαΦj ∗ g2)
∞Xj=0
1 ≈(cid:13)(cid:13)(
j=0 andP∞j=k. When j ≥ k, by the support assumption of Φ, we
2(cid:13)(cid:13)1
.
have supp Φj ∗ g ⊂ 5Qk,m. If α ≥ 0, by (5.18), (5.14) and the Plancherel formula (0.7), we obtain
We splitP∞j=0 into two partsPk−1
τ(cid:0)Z5Qk,m
∞Xj=k
22jαΦj ∗ g(s)2ds(cid:1) 1
If α < 0, by (5.17), (5.14) and the Plancherel formula (0.7) again, we have
τ(cid:0)Z5Qk,m
∞Xj=k
22jαΦj ∗ g(s)2ds(cid:1) 1
2
2
2
2 .
2 ≤ Qm,k− 1
2 = τ(cid:0)Z5Qk,m
∞Xj=k
(I−αΦ)j ∗ I αg(s)2ds(cid:1) 1
≤ τ(cid:0)ZRd
∞Xj=k
(I−αbΦ)(2−jξ)2Iαbg(ξ)2dξ(cid:1) 1
. τ(cid:0)ZRd Iαbg(ξ)2dξ(cid:1) 1
2 = τ(cid:0)ZRd I αg(s)2ds(cid:1) 1
≤ τ(cid:0)ZRd J αg(s)2ds(cid:1) 1
2 ≤ τ(cid:0)Z5Qk,m
∞Xj=k
22jαJ−αΦj ∗ J αg(s)2ds(cid:1) 1
≤ τ(cid:0)ZRd
∞Xj=k
(J−αbΦ)(2−jξ)2Jαbg(ξ)2dξ(cid:1) 1
. τ(cid:0)ZRd Jαbg(ξ)2dξ(cid:1) 1
2 = τ(cid:0)ZRd J αg(s)2ds(cid:1) 1
2(cid:13)(cid:13)1 . 1.
eg = 2k(α−d)g(2−k·).
≤ Qm,k− 1
2 .
2
2
2
j=0 , we begin with a technical modification of g. Let
It follows that
1
22jαΦj ∗ g2)
∞Xj=k
(cid:13)(cid:13)(
In order to estimate the sumPk−1
Then it is easy to see thateg is an (α, Q0,m)-atom. Moreover, we have
which implies that
Φj ∗ g = 2k(d−α)Φj−k ∗eg(2k·),
2(cid:13)(cid:13)1 ≤(cid:13)(cid:13)(
22jαΦj ∗eg2)
−1Xj=−∞
1
(5.28)
k−1Xj=0
(cid:13)(cid:13)(
22jαΦj ∗ g2)
where (Φ0)−k denotes the inverse Fourier transform of the function Φ(0)(2k·). In other words, we
can assume, by translation, that the atom g is based on a cube Q with side length 1 and centered
2(cid:13)(cid:13)1 + 2−kαk(Φ0)−k ∗egk1,
at the origin. Then, let us estimate the right hand side of (5.28) with g instead ofeg.
By the triangle inequality, we have
1
−1Xj=−∞
(cid:13)(cid:13)(
22jαΦj ∗ g2)
2jατZRd Φj ∗ g(s)ds
1
2(cid:13)(cid:13)1 ≤
≤
−1Xj=−∞
−1Xj=−∞ X(µ,l)≤(0,0)
dµ,l 2jατZRd Φj ∗ aµ,l(s)ds.
42
R. Xia and X. Xiong
M + α > 0 and L ≥ M − 1. By the moment cancellation of aµ,l, we have
Φj ∗ aµ,l(s)
Now we estimate 2jατRRd Φj ∗ aµ,l(s)ds for every (µ, l) ≤ (0, 0). Let M = [−α] + 1. Then
= 2jdZ2Qµ,l(cid:2)Φ(2js − 2jt) − Φ(2js − 2j2−µl)(cid:3)aµ,l(t)dt
= 2j(d+M) Xβ1=M
Φj ∗ aµ,l(s)2 . Xβ1=M
(1 − θ)M DβΦ(cid:0)2js − 2j(θt + (1 − θ)2−µl)(cid:1)aµ,l(t)dθ dt.
(1 − θ)2MDβΦ(cid:0)2js − 2j(θt + (1 − θ)2−µl)(cid:1)2dθdt
Z2Qµ,l
(2−µl − t)βZ 1
22j(d+M)Z2Qµ,lZ 1
It follows that
M + 1
β!
0
0
·Z2Qµ,l t − 2−µl2Maµ,l(t)2dt.
2
.
If Φj ∗ aµ,l(s) 6= 0, then we have 2js − 2jt ≤ 1 for some t ∈ 2Qµ,l. Hence, Φj ∗ aµ,l(s) = 0 if
s − 2−µl > 3 · 2−j−1√d. Consequently,
2jατZRd Φj ∗ aµ,l(s)ds
−1Xj=−∞
2j(d+M+α)τ(cid:0)Z2Qµ,l t − 2−µl2Maµ,l(t)2dt(cid:1) 1
−1Xj=−∞
· Xβ1=MZs−2−µl≤3·2−j−1√d(cid:16)Z2Qµ,lZ 1
−1Xj=−∞
−1Xj=−∞
= 2−µ(α+M) −1Xj=−∞
(1 − θ)2MDβΦ(2js − 2j(θt + (1 − θ)2−µl))2dθdt(cid:17) 1
2 τ(cid:0)Z2Qµ,l aµ,l(t)2dt(cid:1) 1
2Zs−2−µl≤3·2−j−1√d
2j(d+M+α) · 2−jd · 2−µ(α+M)Qµ,l
2j(d+M+α) · 2−µMQµ,l
2 . 2−µ(α+M)Qµ,l
2j(M+α)Qµ,l
1
2 .
ds
ds
.
.
1
2
0
1
1
2
Similarly, we also have
and
Therefore, kgkF α,c
1
. 1. The proof is complete.
Thus, by the Cauchy-Schwarz inequality, we get
2 ≤ 2−µ(α+M)Qµ,l
1
1
22jαΦj ∗ g2)
2(cid:13)(cid:13)1 ≤
−1Xj=−∞
(cid:13)(cid:13)(
2−kατZRd (Φ0)−k ∗ aµ,l(s)ds . 2−k(M+α)2−µ(α+M)Qµ,l
22jατZRd Φj ∗ g(s)ds
−1Xj=−∞
∞Xµ=0
2−µ(α+M)(cid:0)Xl
2(cid:0)Xl
dµ,l2(cid:1) 1
∞Xµ=0
2−kαk(Φ0)−k ∗ gk1 .
2−µ(α+M) < ∞.
2−µ(α+M) < ∞,
.
.
∞Xµ=0
2
Qµ,l(cid:1) 1
1
2 .
(cid:3)
Inhomogeneous Triebel-Lizorkin spaces
43
We close this section by a very useful result of pointwise multipliers, which can be deduced from
the above atomic decomposition. Let k ∈ N and Lk(Rd,M) be the collection of all M-valued
functions on Rd such that Dγ h ∈ L∞(N ) for all γ with 0 ≤ γ1 ≤ k.
Corollary 5.8. Let α ∈ R and let k ∈ N be sufficiently large and h ∈ Lk(Rd,M). Then the map
f 7→ hf is bounded on F α,c
Proof. First, consider the case α > 0. We apply the atomic decomposition in Theorem 5.7 with
K = k and L = −1.
In this case, no moment cancellation of subatoms is required. We can
easily check that, multiplying every (sub)atom in Definition 5.4 by h, we get another (sub)atom.
Moreover,
(Rd,M)
1
(5.29)
khfkF α,c
1 ≤ Xγ≤k
sup
s∈Rd kDγh(s)kM · kfkF α,c
1
.
The case α ≤ 0 can be deduced by induction. Assume that (5.29) is true for α > N ∈ Z. Let
and
can be represented as f = J 2g = (1 − (2π)−2∆)g with g ∈ F α+2,c
α > N − 1. Any f ∈ F α,c
kfkF α,c
. Since
hf = (1 − (2π)−2∆)(hg) + ((2π)−2∆h)g + (2π)−2∇h · ∇g,
1 ≈ kgkF α+2,c
1
1
1
we deduce
khfkF α,c
1
. k(1 − (2π)−2∆)(hg)kF α,c
1
+ k(∆h)gkF α,c
1
+
(5.30)
. kgkF α+2,c
1
+ k(∆h)gkF α+2,c
1
+
dXi=1
k∂ih · ∂igkF α+1,c
1
.
dXi=1
k∂ih · ∂igkF α,c
1
If k ∈ N is sufficiently large, we have
. kgkF α+2,c
Continuing the estimate in (5.30), we obtain
k(∆h)gkF α+2,c
1
1
khfkF α,c
1
. kgkF α+2,c
1
+Xi
which completes the induction procedure.
,
k∂ih · ∂igkF α+1,c
1
. k∂igkF α+1,c
1
.
k∂igkF α+1,c
1
. kgkF α+2,c
1
. kfkF α,c
1
,
(cid:3)
Acknowledgements. The authors are greatly indebted to Professor Quanhua Xu for having
suggested to them the subject of this paper, for many helpful discussions and very careful reading
of this paper. The authors are partially supported by the the National Natural Science Foundation
of China (grant no. 11301401).
References
[1] T. Bekjan, Z. Chen, M. Perrin and Z. Yin. Atomic decomposition and interpolation for Hardy spaces of non-
commutative martingales. J. Funct. Anal. 258 (2010), no. 7, 2483-2505.
[2] J. Bergh and J. Lofstrom. Interpolation Spaces: An Introduction. Springer, Berlin, 1976.
[3] R. Coifman, Y. Meyer and E. M. Stein. Some new function spaces and their applications to Harmonic analysis.
J. Funct. Anal. 62 (1985), no. 2, 304-335.
[4] M. Frazier and B. Jawerth. The ϕ-transform and applications to distribution spaces. In Function spaces and
applications, M. Cwikel et al. eds., Springer Lect. Notes in Math. 1302 (1988), 223-246.
[5] M. Frazier and B. Jawerth. A discrete transform and decompositions of distribution spaces. J. Funct. Anal. 93
(1990), no. 1, 34-170.
[6] M. Frazier, R. Torres and G. Weiss. The boundedness of Caldern-Zygmund operators on the spaces F α,q
p
. Rev.
Mat. Iberoam. Vol. 4 (1988), 41-72.
[7] J. Garc´ıa-Cuerva and J.L. Rubio de Francia. Weighted norm inequalities and related topics. North-Holland
Mathematics Studies, 116. Notas de Matemtica, 104. North-Holland Publishing Co., Amsterdam, 1985. x+604
pp.
[8] D. Goldberg. A local version of real Hardy spaces. Duke Math. J. 46 (1979), no. 1, 27-42.
[9] L. Grafakos. Classical Fourier analysis. Second Edition. Springer-Verlag, New York, 2008.
[10] G. Hong, L. D. L´opez-Snchez, J. M. Martell, and J. Parcet. Caldern-Zygmund operator associated to matrix-
valued kernels. Int. Math. Res. Not. 2014, no. 5, 1221-1252.
[11] G. Hong and T. Mei. John-Nirenberg inequality and atomic decomposition for noncommutative martingales. J.
Funct. Anal. 263 (2012), no. 4, 1064-1097.
44
R. Xia and X. Xiong
[12] Y. Jiao, F. Sukochev, D. Zanin and D. Zhou. Johnson-Schechtman inequalities for noncommutative martingales.
J. Funct. Anal. 272 (2017), no. 3, 976-1016.
[13] M. Junge. Doob's inequality for non-commutative martingales. J. Reine Angew. Math. 549 (2002), 149-190.
[14] M. Junge, C. Le Merdy and Q. Xu. H∞-functional calculus and square functions on noncommutative Lp-spaces.
Ast´erisque No. 305 (2006), vi+138 pp.
[15] M. Junge and Q. Xu. Non-commutative Burkholder/Rosenthal Inequalities, I and II. Ann. Prob. 31 (2003), no.
2, 948-995 and Israel J. Math. 167 (2008), 227-282.
[16] M. Junge and Q. Xu. Noncommutative maximal ergodic theorems. J. Amer. Math. Soc. 20 (2007), 385-439.
[17] M. Junge and T. Mei. Noncommutative Riesz transforms - a probabilistic approach. Amer. J. Math. 132 (2010),
611-681.
[18] T. Mei. Operator-valued Hardy Spaces. Mem. Amer. Math. Soc. 188 (2007), no. 881, vi+64 pp.
[19] T. Mei. Tent spaces associated with semigroups of operators. J. Funct. Anal. 255 (2008), no. 12, 3356-3406.
[20] M. Musat. Interpolation between non-commutative BMO and non-commutative Lp-spaces. J. Funct. Anal. 202
(2003), no. 1, 195-225.
[21] Yu. V. Netrusov. Embedding theorems for Lizorkin-Triebel spaces. (Russian) Zap. Nauchn. Sem. Leningrad.
Otdel. Mat. Inst. Steklov. (LOMI) 159 (1987), 103-112. translation in J. Soviet Math. 47 (1987), no. 6, 2896-2903.
[22] J. Parcet. Pseudo-localization of singular integrals and noncommutative Calder´on-Zygmund theory. J. Funct.
Anal. 256 (2009), no. 2, 509-593.
[23] J. Parcet and N. Randrianantoanina. Gundy's decomposition for non-commutative martingales and applications.
Proc. London Math. Soc. 93 (2006), 227-252.
[24] G. Pisier. Noncommutative vector valued Lp spaces and completely p-summing maps. Ast´erisque. 247 (1998),
vi+131 pp.
[25] G. Pisier and Q. Xu. Non-commutative Martingale Inequalities. Comm. Math. Phys. 189 (1997), 667-698.
[26] G. Pisier and Q. Xu. Noncommutative Lp-spaces. Handbook of the geometry of Banach spaces, Vol. 2, 1459-
1517, North-Holland, Amsterdam, 2003.
[27] N. Randrianantoanina. Noncommutative martingale transforms. J. Funct. Anal. 194 (2002), 181-212.
[28] N. Randrianantoanina. Conditional square functions for noncommutative martingales. Ann. Prob. 35 (2007),
1039-1070.
[29] N. Randrianantoanina and L. Wu. Noncommutative Burkholder/Rosenthal inequalities associated with convex
functions. Ann. Inst. Henri Poincar´e Probab. Stat. 53 (2017), no. 4, 1575-1605.
[30] N. Randrianantoanina, L. Wu and Q. Xu. Noncommutative Davis type decompositions and applications. arXiv:
1712.01374.
[31] E. M. Stein. Singular integrals and differentiability properties of functions. Princeton Mathematical Series, No.
30 Princeton University Press, Princeton, N.J. 1970.
[32] E. M. Stein. Harmonic analysis: real-variable methods, orthogonality, and oscillatory integrals. Princeton Math-
ematical Series, 43. Monographs in Harmonic Analysis, III. Princeton University Press, Princeton, NJ, 1993.
[33] E. M. Stein and G. Weiss. Introduction to Fourier analysis on Euclidean spaces. Princeton Mathematical Series,
No. 32. Princeton University Press, Princeton, N.J., 1971.
[34] H. Triebel. Theory of function spaces. Modern Birkhuser Classics. Birkhuser/Springer Basel AG, Basel, 2010.
[35] H. Triebel. Theory of function spaces. II. Monographs in Mathematics, 84. Birkhuser Verlag, Basel, 1992.
[36] R. Xia and X. Xiong. Operator-valued local Hardy spaces Arxiv: 1803.10321.
[37] R. Xia and X. Xiong. Pseudo-differential operators with operator-valued kernel. Preprint.
[38] R. Xia, X. Xiong and Q. Xu. Characterizations of operator-valued Hardy spaces and applications to harmonic
analysis on quantum tori. Adv. Math. 291 (2016), 183-227.
[39] X. Xiong, Q. Xu and Z. Yin. Function spaces on quantum tori. C. R. Math. Acad. Sci. Paris 353 (2015), no.
8, 729-734.
[40] X. Xiong, Q. Xu and Z. Yin. Sobolev, Besov and Triebel-Lizorkin spaces on quantum tori. Mem. Amer. Math.
Soc. 252 (2018), no. 1203, vi+118 pp.
[41] Q. Xu. Noncommutative Lp-spaces and martingale inequalities. Book manuscript, 2007.
Laboratoire de Math´ematiques, Universit´e de Franche-Comt´e, 25030 Besanc¸on Cedex, France, and
Instituto de Ciencias Matem´aticas, 28049 Madrid, Spain
E-mail address: [email protected]
Department of Mathematics and Statistics, University of Saskatchewan, Saskatoon, Saskatchewan,
S7N 5E6, Canada
E-mail address: [email protected]
|
1807.08371 | 1 | 1807 | 2018-07-22T21:31:07 | Column extreme multipliers of the Free Hardy space | [
"math.OA",
"math.FA"
] | The full Fock space over $\mathbb C ^d$ can be identified with the free Hardy space, $H^2 (\mathbb B ^d _\mathbb N)$ - the unique non-commutative reproducing kernel Hilbert space corresponding to a non-commutative Szeg\"{o} kernel on the non-commutative, multi-variable open unit ball $\mathbb B ^d _\mathbb N := \bigsqcup _{n=1} ^\infty \left( \mathbb C^{n\times n} \otimes \mathbb C ^d \right) _1$.
Elements of this space are free or non-commutative functions on $\mathbb B ^d _\mathbb N$. Under this identification, the full Fock space is the canonical non-commutative and several-variable analogue of the classical Hardy space of the disk, and many classical function theory results have faithful extensions to this setting. In particular to each contractive (free) multiplier $B$ of the free Hardy space, we associate a Hilbert space $\mathcal H(B)$ analogous to the deBranges-Rovnyak spaces in the unit disk, and consider the ways in which various properties of the free function $B$ are reflected in the Hilbert space $\mathcal H(B)$ and the operators which act on it. In the classical setting, the $\mathcal H(b)$ spaces of analytic functions on the disk display strikingly different behavior depending on whether or not the function $b$ is an extreme point in the unit ball of $H^\infty(\mathbb D)$. We show that such a dichotomy persists in the free case, where the split depends on whtether or not $B$ is what we call {\it column extreme}. | math.OA | math |
COLUMN EXTREME MULTIPLIERS OF THE FREE HARDY
SPACE
MICHAEL T. JURY AND ROBERT T.W. MARTIN
Abstract. The full Fock space over Cd can be identified with the free Hardy
space, H 2(Bd
N) - the unique non-commutative reproducing kernel Hilbert space
corresponding to a non-commutative Szego kernel on the non-commutative,
multi-variable open unit ball Bd
n=1 (cid:0)Cn×n ⊗ Cd(cid:1)1.
N := F∞
Elements of this space are free or non-commutative functions on Bd
N. Under
this identification, the full Fock space is the canonical non-commutative and
several-variable analogue of the classical Hardy space of the disk, and many
classical function theory results have faithful extensions to this setting.
In
particular to each contractive (free) multiplier B of the free Hardy space, we
associate a Hilbert space H(B) analogous to the deBranges-Rovnyak spaces
in the unit disk, and consider the ways in which various properties of the free
function B are reflected in the Hilbert space H(B) and the operators which
act on it.
In the classical setting, the H(b) spaces of analytic functions on
the disk display strikingly different behavior depending on whether or not the
function b is an extreme point in the unit ball of H∞(D). We show that such
a dichotomy persists in the free case, where the split depends on whtether or
not B is what we call column extreme.
1. Introduction
The classical Hardy space, H 2(D), can be defined as the Hilbert space of all analytic
functions on D whose Taylor series at 0 have square summable coefficients (and with
inner product equal to the ℓ2 inner product of these Taylor coefficients). Equivalently,
H 2(D) = H(k), is the unique reproducing kernel Hilbert space (RKHS) of functions on D
corresponding to the positive sesqui-analytic kernel function k : D × D → C:
k(z, w) :=
1
1 − zw∗ ;
z, w ∈ D,
the Szego kernel. The operator of multiplication by z on H 2(D) is called the shift, and
it is easily seen to be isomorphic to the unilateral shift on ℓ2(N0), where N0 denotes
the non-negative integers. Proofs of many deep results in classical Hardy space theory
ultimately appeal to the fact that S is the universal cyclic pure isometry (recall the Wold
decomposition says that any isometry is unitarily equivalent to a direct sum of shifts and
a unitary operator).
From the viewpoint of reproducing kernel theory and operator theory, the canonical
(commutative) multi-variable analogue of the Hardy space is then the Drury-Arveson
space, H 2
d := H(k), where now k : Bd × Bd → C is:
1 − zw∗ ;
the multi-variable Szego kernel, and zw∗ := z1w∗
d = (w, z)Cd . (Here, Bd := (Cd)1,
the multi-variable open unit ball.) The appropriate analogue of the shift in this setting
1 + ...zdw∗
k(z, w) :=
z, w ∈ Bd,
1
Date: July 24, 2018.
1
2
MICHAEL T. JURY AND ROBERT T.W. MARTIN
is the Arveson d−shift, S = (S1, ..., Sd) : H 2
d , (Sjh)(z) := zj h(z); z =
(z1, ..., zd) ∈ Bd. This is a (row) partial isometry (from d copies of H 2
d into itself), but
no longer an isometry, and this defect is the source of several differences between the
single and several-variable theories. Faithful analogues of classical Hardy space results
typically seem to exist, but often new (and often more complicated) proof techniques and
approaches are required [23].
d ⊗ Cd → H 2
An alternative approach to extending Hardy space theory from one to several variables
would be to seek analogues of Hardy space results for a several-variable shift. Namely, the
natural multi-variable analogue of ℓ2(N0) is ℓ2(Fd), where Fd is the free monoid (unital
semi-group) of all words in the d letters {1, ..., d}, and with unit equal to the empty word,
∅, containing no letters. This monoid can be identified with a simple directed tree starting
at single node and with d branches at each node (clearly F1 ≃ N0). There is a natural
d−tuple of shifts, L := (L1, ..., Ld) on ℓ2(Fd) which are defined by
Lkeα := ekα,
where {eα}α∈Fd is the canonical orthonormal basis of ℓ2(Fd). It is easy to see that each
Lk is a pure isometry and the Lk have pairwise orthogonal ranges L∗
In
particular, the row L = (L1, ..., Ld) : ℓ2(Fd) ⊗ Cd → ℓ2(Fd) is an isometry from d copies
of ℓ2(Fd) into itself which we call the left free shift. The Popescu-Wold decomposition
for row isometries shows that L has the same universal property as the shift S: any row
isometry (an isometry from d copies of a Hilbert space into itself) is isomorphic to the
direct sum of several copies of L and a row unitary (an onto row isometry).
kLj = δk,jI.
The left free shifts Lk are of course non-commuting, and it would appear that one loses
the analytic function theory interpretation of the shift as acting as multiplication by the
independent variable on a space of analytic functions. Surprisingly, this is not the case: the
fields of non-commutative function theory [13, 1, 18, 20, 19], and the recently developed
theory of non-commutative reproducing kernel Hilbert spaces (NC-RKHS) [3] have shown
that ℓ2(Fd) is canonically isomorphic to the free Hardy space, H 2(Bd
N) of non-commutative
or free holomorphic functions on a certain non-commutative multi-variable open unit ball,
Bd
N (we will introduce these objects and this theory in an upcoming subsection).
The Drury-Arveson space H 2
d can be identified with a subspace of H 2(Bd
N) which is
co-invariant and cyclic for both the left and right free shifts:
H 2
d ≃ _z∈Bd
Kz ⊆ H 2(Bd
N),
the span of all the kernel vectors at level one. This subspace is the orthogonal complement
of the range of both a right inner and a left inner free multiplier. For example, if d = 2,
H 2
d ≃ Ran(cid:18) 1
√2
(L1L2 − L2L1)(cid:19)⊥
,
and this shows that the theory of H 2
d should be closer in analogy to that of the theory of
model subspaces of H 2(D). In particular, commutative Drury-Arveson space analogues of
all of the results of this paper (and those of [8]) can be easily obtained by compression.
In recent work, we have extended Hardy space results including the concept of Aleksandrov-
Clark measure, the theory of Clark's unitary perturbations, and equivalent characteriza-
tions of extreme points from one to several variables.
In particular, the reference [8]
extends the theory of Clark measures and Clark peturbations to the non-commutative
setting of the full Fock space over Cd (which can be identified with ℓ2(Fd)) using the
theory of free formal reproducing kernel Hilbert spaces [4].
The goal of this paper is to develop non-commuative analogues of our recent results
on extreme points of the closed unit ball of the multiplier algebra of Drury-Arveson space
[11, 12]. We will also extend and re-cast the main results of [8] in the modern language
COLUMN EXTREME MULTIPLIERS OF THE FREE HARDY SPACE
3
of NC-RKHS. In particular we give a number of equivalent characterizations of so-called
column extreme multipliers of the free Hardy space.
2. Preliminaries
All Hilbert space inner products will be conjugate linear in their first argument. If X
is a Banach space, (X)1 and [X]1 denote the open and closed unit balls of X, respectively.
2.1. The full Fock space. Recall that the full Fock space over Cd, F 2
of all tensor powers of Cd:
d , is the direct sum
F 2
d
:= C ⊕(cid:16)Cd ⊗ Cd(cid:17) ⊕(cid:16)Cd ⊗ Cd ⊗ Cd(cid:17) ⊕ ···
=
∞Mk=0(cid:16)Cd(cid:17)k·⊗
.
Fix an orthonormal basis {e1, ..., ed} of Cd. The left creation operators L1, ..., Ld are the
operators which act as tensoring on the left by these basis vectors:
and similarly the right creation operators Rk; 1 ≤ k ≤ d are defined by tensoring on the
right
Lkf := ek ⊗ f ;
f ∈ F 2
d ,
Rkf := f ⊗ ek.
d ⊗ Cd into F 2
The left and right free shifts are the row operators L := (L1, ..., Ld) and R := (R1, ..., Rd)
d . Both L, R are in fact row isometries: L∗L = IF 2 ⊗ Id = R∗R.
which map F 2
It follows that the component shifts are also isometries with pairwise orthogonal ranges.
The orthogonal complement of the range of L or R is the vacuum vector 1 which spans the
the subspace C =: (Cd)0·⊗ ⊂ F 2
d is then {eα}α∈Fd
where eα = Lα1 = Rα1 and Fd is the free unital semigroup or monoid on d letters. Here,
if α = i1 ·· · in ∈ Fd, we use the standard notation Lα = Li1 Li2 ··· Lin .
Recall here that the free monoid, Fd, on d ∈ N letters, is the multiplicative semigroup
of all finite products or words in the d letters {1, ..., d}. That is, given words α := i1...in,
β := j1...jm, ik, jl ∈ {1, ..., d}; 1 ≤ k ≤ n, 1 ≤ l ≤ m, their product αβ is defined by
concatenation:
d . A canonical orthonormal basis for F 2
αβ = i1...inj1...jm,
and the unit is the empty word, ∅, containing no letters. Given α = i1 ··· in, we use the
standard notation α = n for the length of the word α. The transpose map † : Fd → Fd,
defined by
i1 ··· id = α 7→ α† := id ··· i1,
is an involution.
Define L∞
d
:= Alg(I, L)−W OT , R∞
d
right) free
analytic Toeplitz algebra (W OT denotes weak operator topology). The transpose unitary,
U† : F 2
d , and it is easy to verify
that
d , defined by eα 7→ eα† is a unitary involution of F 2
:= Alg(I, R)−W OT , the left (resp.
d → F 2
U†LkU ∗
† = Rk,
so that adjunction by U† implements a unitary isomorphism between L∞
d and R∞
d .
2.2. The free Hardy space. It will be convenient to view F 2
d as a non-commutative
reproducing kernel Hilbert space (NC-RKHS) [3] of freely non-commutative (holomorphic)
functions on the non-commutative open unit ball [13]:
Bd
N :=
Bd
n;
∞Gn=1
Bd
n :=(cid:16)Cn×n ⊗ Cd(cid:17)1
.
4
MICHAEL T. JURY AND ROBERT T.W. MARTIN
Elements of Bd
vector space V ,
n are viewed as strict row contractions on Cn. Recall that for any complex
The NC unit ball Bd
under direct sums, and one writes:
N is an example of a NC set: A set Ω ⊆ Vnc is an NC set if it is closed
Vn := V ⊗ Cn×n =: V n×n.
Vnc :=G Vn;
Ω =:G Ωn;
f : Ωn → Cn×n;
A function f : Ω → Cnc is called a NC or free function if:
Ωn := Ω ∩ Vn.
f respects the grading,
and if X ∈ Cn×m, Z ∈ Xn, W ∈ Xm obey ZX = XW , then,
f respects intertwinings.
As shown in [3], F 2
variable NC unit ball Bd
to the NC-Szego kernel: K : Bd
f (Z)X = Xf (W );
d = H 2(Bd
N, i.e. H 2(Bd
N × Bd
Z αP (W ∗)α†
K(Z, W )[P ] := Xα∈Fd
N) can be viewed as the free Hardy space of the multi-
N) = Hnc(K) is the unique NC-RKHS corresponding
N → L(Cnc) defined by:
;
Z ∈ Bd
n, W ∈ Bd
m, P ∈ Cn×m.
See [3] for the full definition and theory of NC kernels.
respects the grading and intertwinings in both arguments [3, Section 2.3].
In particular, any NC kernel
One can show that elements of H 2(Bd
automatically) holomorphic free functions on Bd
is Fr´echet and Gateaux differentiable at any point Z ∈ Bd
series expansion (Taylor-Taylor series) about any point.
N) := Hnc(K) are locally bounded (and hence
N [13, Chapter 7]. That is, any f ∈ H 2(Bd
N)
N and f has a convergent power
(Generally any) Hnc(K) is formally defined as the Hilbert space completion of the
linear span:
_Z∈Bd
K{W, x, u}(Z)y := K(Z, W )[yu∗]x;
K{Z, y, v},
n, y,v∈Cn
Completion is with respect to the inner product:
where the K{W, x, u} are the free functions on Bd
N, K{W, x, u} : Bd
n → Cn×n, defined by:
W ∈ Bd
m, Z ∈ Bd
n; u, x ∈ Cm, y ∈ Cn.
hK{Z, y, v}, K{W, x, u}i := (y, K(Z, W )[vu∗]x)Cn ;
Z ∈ Bd
n, v, y ∈ Cn; W ∈ Bd
m, u, x ∈ Cm.
These point evaluation vectors have a familiar reproducing property: K(Z, y, v) is the
unique vector in H 2(Bd
N) such that for any f ∈ H 2(Bd
N),
(2.1)
hK{Z, y, v}, fi = (y, f (Z)v)Cn .
For any Z ∈ Bd
N)) as
follows: Any A ∈ Cn×n can be written as a linear combination of the rank one outer
products
n one can also define a natural kernel map KZ ∈ L(Cn×n, H 2(Bd
yv∗ =
y1
...
yn
(cid:2)v1,
···
, vn(cid:3) ;
y ∈ Cn, v∗ ∈ (Cn)∗.
Then we define KZ on rank one matrices yv∗ by the formula
KZ(yv∗) := K{Z, y, v} ∈ H 2(Bd
N).
(2.2)
Let us check that KZ is well defined: the vectors y and v determining a rank one matrix yv∗
are unique up to the scaling y → λy, v → λ
v where λ is any nonzero complex number.
From the reproducing formula (2.1), it is evident that the vector K{Z, y, v} is invariant
−1
COLUMN EXTREME MULTIPLIERS OF THE FREE HARDY SPACE
5
under such a scaling, and so the formula (2.2) is unambiguous. If we view Cn×n as a Hilbert
space equipped with the normalized trace inner product, then KZ : Cn×n → H 2(Bd
N)
extends to a bounded linear map, and its Hilbert space adjoint is the point evaluation
map at Z:
The free Hardy space and the full Fock space are canonically isomorphic: Define U :
F 2
d → H 2(Bd
N) by:
K∗
Z F = F (Z) ∈ Cn×n.
x := Xα∈Fd
fx(Z) := Xα∈Fd
xαLα1
Z αxα;
U
7→ fx ∈ H 2(Bd
N),
Z ∈ Bd
N.
The inverse, U −1, acts on kernel vectors as:
(2.3)
K{Z, y, v}
U −1
7→ x[Z, y, v] := Xα∈Fd hZ αv, yi Lα1 ∈ F 2
d .
2.3. Left and Right free multipliers. As in the classical setting, given a NC-RKHS
Hnc(K) on an NC set Ω (e.g. Bd
N), it is natural to consider the left and right multiplier
algebras
MultL(Hnc(K)), MultR(Hnc(K))
N) := MultL(H 2(Bd
of NC functions on Ω which left or (resp.) right multiply Hnc(K) into itself. Namely,
N is said to be a left free multiplier if, for any f ∈ H 2(Bd
a free function F on Bd
N),
N). Similarly, G is called a right free multiplier if f G ∈ H 2(Bd
F f ∈ H 2(Bd
N) for all
f ∈ H 2(Bd
N). As in the classical setting, the left and right free multiplier algebras,
H ∞
L (Bd
N) are weak operator toplogy (WOT)-closed unital
operator algebras. Moreover, adjunction by the canonical unitary U defines a unitary
∗−isomorphism of the left and right free analytic Toeplitz algebras L∞
d onto these
N). As in the classical setting of H ∞(D), the
left and right free multiplier algebras of H 2(Bd
multiplier norm of any F ∈ H ∞
N) can be computed as the supremum norm on the NC
unit ball:
N)), H ∞
d , R∞
R (Bd
L (Bd
kFk := sup
Z∈Bd
N
kF (Z)k.
The left and right Schur classes, Ld, Rd are then defined as the closed unit balls of these
left and right multiplier algebras (equivalently as the closed unit balls of L∞
d , R∞
d ).
Observe that if F is a left free multiplier then,
hK{Z, y, v}, F fi = hy, F (Z)f (Z)vi
= hK{Z, F (Z)∗y, v}, fi ,
so that
(2.4)
(2.5)
and similarly, if G is a right free multiplier,
(M L
F )∗K{Z, y, v} = K{Z, F (Z)∗y, v},
(M R
G )∗K{Z, y, v} = K{Z, y, G(Z)v}.
Alternatively, using the kernel maps KZ , we can write:
(M L
F )∗KZ(yv) = KW (F (Z)∗yv∗),
and
(M R
G )∗KZ = KZ(yv∗G(Z)∗).
One can check that if, e.g., right multiplication by G(Z) is a right free multiplier then
((M R
G )∗KZ )∗((M R
G )∗KW ) = K(Z, W )[G(Z) · G(W )∗].
In particular, free holomorphic F (Z), G(Z) belong to the left or right Schur classes if
and only if
K F (Z, W )[·] := K(Z, W ) − F (Z)K(Z, W )[·]F (W )∗
6
or
MICHAEL T. JURY AND ROBERT T.W. MARTIN
K G(Z, W )[·] := K(Z, W ) − K(Z, W )[G(Z)[·]G(W )∗]
are CPNC kernels, respectively. These NC kernels are called the left or right free deBranges-
Rovnyak kernels of F, G (resp.) and in this case the corresponding NC-RKHS Hnc(K F ) =:
H L(F ), Hnc(K G) =: H R(G) are the left and right free deBranges-Rovnyak spaces of
F, G.
2.4. Left vs. Right. Any element F ∈ L∞
series:
FαLα;
F ∼ F (L) := Xα∈Fd
Fα := hLα1, F 1i .
d can be identified with the left free Fourier
That is, F is identified with its symbol :
f := F 1 = Xα∈Fd
FαLα1 ∈ F 2
d ,
and we say that F (L) = M L
f acts as left multiplication by f = F 1. In general the free
Fourier series does not converge in SOT or WOT, but the Ces`aro sums converge in the
strong operator toplogy (SOT) to F [6].
Similarly, in the operator valued setting, any F ∈ L∞
f , where the symbol, f ∈ F 2
d ⊗ L(H, J ) is defined by
M L
d ⊗ L(H,J ) is written F = F (L) =
f := F (1 ⊗ IH) =Xα
Lα1 ⊗ Fα;
Fα ∈ L(H,J ).
In this case the operator-valued free holomorphic function F (Z) takes values in (C)nc ⊗
L(H,J ).
We can also identify any G ∈ R∞
d with its symbol:
g := G1 = Xα∈Fd
GαLα1,
then we can view G as right multiplication by g(Z),
Alternatively, we can write
G = M R
g(Z).
g = Xα∈Fd
Gα† Rα1,
so that
G = M R
g(Z) = g†(R), where
g†(Z) := Xα∈Fd
Gα† Z α.
That is, if G ∈ R∞
G(Z), then M R
sums converge SOT to G).
d acts as right multiplication by the free NC holomorphic function
G(Z) is identified with the right free Fourier series G†(R) (whose Ces`aro
Remark 2.5. In the right operator-valued setting, suppose that G(R) := g(R) ⊗ X ∈
R∞
d ⊗ L(H, J ) and F := f (R) ⊗ Y ∈ R∞
d , X ∈ L(H,J ), and
Y ∈ L(J ,K). If H = F G, then observe that
d ⊗ L(J ,K) with f, g ∈ R∞
H †(Z) = g†(Z)f †(Z) ⊗ Y X.
This extends to a 'right product' for arbitrary operator-valued free holomorphic functions
on Bd
N, H(Z) = F (Z)•R G(Z). In the scalar-valued setting this simply reduces to F (Z)•R
G(Z) = G(Z)F (Z).
COLUMN EXTREME MULTIPLIERS OF THE FREE HARDY SPACE
7
2.6. Operator-valued free multipliers. It will also be convenient to consider operator-
valued (left and right) free multipliers between vector-valued free Hardy spaces. Namely, if
H is an auxiliary Hilbert space, one can consider the NC-RKHS H 2(Bd
N )⊗H of H-valued
NC functions on Bd
N. This NC-RKHS has the operator-valued CPNC kernel:
K(Z, W ) ⊗ IH,
and is spanned by the elements
with inner product defined by
K{Z, y, v}h := K{Z, y, v} ⊗ h,
h ∈ H,
n, W ∈ Bd
hK{Z, y, v}h, K{W, x, u}gi := (y, K(Z, W )[vu∗]x)Cn · hh, giH ,
m, v, y ∈ Cn and u, x ∈ Cm. We will write H ∞
L (Bd
for h, g ∈ H, Z ∈ Bd
N)⊗ L(H, J )
in place of MultL(H 2(Bd
N)⊗J ), the W OT−closed left multiplier space between
these vector-valued free Hardy spaces. (That is, we write H ∞
N) ⊗ L(H, J ) in place of
the weak operator topology closure of this algebraic tensor product). The operator-valued
Schur classes, Ld(H,J ), Rd(H,J ) are then the closed unit balls of these operator-valued
left and right (resp.) multiplier spaces.
N)⊗H, H 2(Bd
L (Bd
If A ∈ Ld(H,J ), consider the operator-valued CPNC kernel:
K A(Z, W )[·] = K(Z, W )[·] ⊗ IJ − A(Z)(K(Z, W )[·] ⊗ IH)A(W )∗.
Here, K is the free Szego kernel. This is the left free deBranges-Rovnyak CPNC ker-
nel of A, and the corresponding NC-RKHS, Hnc(K A) =: H L(A) is called the left free
deBranges-Rovnyak space of A. Namely, H L(A) is the closed linear span of vectors of
the form K A{Z, y, v}g whose inner product is defined by:
DK A{Z, y, v}g, K A{W, x, u}fEH L(A)
:= (y ⊗ g, (K(Z, W )[vu∗] ⊗ IJ − A(Z)(K(Z, W )[vu∗] ⊗ IH)A(W )∗) x ⊗ f )Cn⊗J .
Z : Cn×n ⊗ J → H L(A) for the
On the other hand, if B ∈ Rd(H, J ), then the right free deBranges-Ronvyak space
In this vector-valued setting, for Z ∈ Bd
Z (yv∗ ⊗ h) = K A{Z, y, v}h.
kernel map K A
n we write K A
H R(B) is spanned by the vectors K B{Z, v, y}g with inner product:
DK B{Z, y, v}g, K B{W, x, u}fEH R(B)
:= (cid:16)y ⊗ g, K B(Z, W )[vu∗ ⊗ IH]x ⊗ f(cid:17)Cn⊗J
K B(Z, W ) = K(Z, W ) ⊗ IJ − (K(Z, W ) ⊗ IJ )[B†(Z)(· ⊗ IH)B†(W )∗].
It is not difficult to see that free operator-valued holomorphic functions A, B on Bd
N belong
to the left or right free Schur classes if and only if the above NC deBranges-Rovnyak kernels
are (completely) positive.
Remark 2.7. (Right Product) If F ∈ MultR(K1, K2)⊗ L(H, J ) is a right operator-valued
multiplier between vector-valued NC-RKHS on (say) the open unit NC ball Bd
N, then one
can easily verify that for any g ∈ J ,
(M R
F )∗K2 {Z, y, v} g = K1{Z, y, F (Z) •R v}g,
where for any f ∈ Hnc(K1) we define:
hK1{Z, y, F (Z) •R v}g, fi = hy ⊗ h, F (Z) •R f (Z)vi .
Also, in the above, given any element f of a H−valued NC-RKHS Hnc(K), note that
f (Z) ∈ Cn×n ⊗ H, so that f (Z)v is to be interpreted as an element of Cn ⊗ H.
8
MICHAEL T. JURY AND ROBERT T.W. MARTIN
2.8. Coefficient evaluation and free formal RKHS. Let K be an operator-valued
CPNC kernel on Bd
1 = Bd converges absolutely
on Bd
N, and uniformly on compacta:
N whose Taylor-Taylor series about 0 ∈ Bd
K(Z, W )[P ] = Xα,β∈Fd
Kα,βZ α[P ](W ∗)β†
;
Kα,β ∈ L(H).
Any right or left (operator-valued) deBranges-Rovnyak kernel has this property, for ex-
ample. The coefficient kernel function K(·,·) : Fd × Fd → L(H) is then an operator-valued
free formal kernel in the sense of [4, 3] (see also [8] which develops free Aleksandrov-Clark
theory using the free formal RKHS setup).
If F (Z) :=Pα FαZ α ∈ Hnc(K), then for any α ∈ Fd, the linear H−valued map defined
α)∗ : H → Hnc(K) will be called the
is bounded. The Hilbert space adjoint Kα := (K∗
coefficient kernel map, and one always has
α(F ) := Fα ∈ H,
by coefficient evaluation:
K∗
Kα,β = K∗
αKβ ∈ L(H),
and
Kβ(Z) = Xα∈Fd
Kα,βZ α.
Observe that Kα,β is a positive kernel function in the classical sense on the discrete set
Fd.
2.9. The free Herglotz-Schur classes.
Definition 2.10. The left free Herglotz-Schur class, L +
valued free NC functions on Bd
N.
d , is the set of all accretive matrix-
If H ∈ L +
follows that
d , then H(Z) is an accretive matrix (positive semi-definite real part), and it
is a contractive free function on Bd
N, i.e., B ∈ Ld. Conversely, given such a B,
BH (Z) := (H(Z) − I)(H(Z) + I)−1,
HB(Z) := (I + B(Z))(I − B(Z))−1,
has non-negative real part. That is, this fractional linear transformation, the Cayley
Transform, defines a bijection between Ld and L +
d . The right free Herglotz-Schur class,
R +
d under the transpose map, †, and we similarly define the operator-
valued free Herglotz-Schur classes L +
d (H, H) as the image of Ld(H) under
Cayley transform.
d is the image of L +
d (H) = L +
Remark 2.11. By [10, Lemma 3.2, Lemma 3.3], if B ∈ Rd or Ld, then 1 − B is outer
(and necessarily invertible on Bd
N). Moreover, the free Herglotz-Schur classes are contained
in the free Smirnov classes, which can be identified with closed, densely-defined (generally
unbounded) right and left multipliers of H 2(Bd
N) [10]. In particular, by [10, Corollary 3.13,
Corollary 3.15], if B ∈ Ld or A ∈ Rd, then the free polynomials belong to the domains of
and are cores for both HB(L)∗ and HA(R)∗.
3. Free Aleksandrov-Clark measures
Let Ad :=(cid:0)Wα∈Fd Lα(cid:1)−k·k, the free disk algebra. In the case where d = 1, L = S (the
shift), and we recover the classical disk algebra A1 = A(D) of bounded analytic functions
on D with continous extensions to the unit circle T = ∂D.
COLUMN EXTREME MULTIPLIERS OF THE FREE HARDY SPACE
9
Recall the Herglotz representation formula for Herglotz functions (analytic functions
with non-negative real part) on the disk: If H is a Herglotz function, then there is a unique
finite positive Borel measure µ on T so that:
H(z) = iIm (H(0)) +ZT
1 + zζ
1 − zζ
µ(dζ).
As discussed above, any such H ∈ S + := L +
1 is the Cayley transform of some contractive
analytic Schur class function b ∈ S := [H ∞(D)]1 = L1, H = Hb = 1+b
1−b , and the measure
µ =: µb is called the Aleksandrov-Clark measure of b. This defines bijections (modulo
imaginary constants) between S , S +, and the set of all finite positive Borel measures on
T.
Any finite positive Borel measure µ on T can be identified with a positive linear func-
tional on the disk algebra operator system: A1 + A∗
1:
µ(Sn) :=ZT
ζ nµ(dζ).
The theory of closed, densely-defined operators affiliated to the shift [24, 22], implies that
if b ∈ S , then multiplication by Hb ∈ S + is a closed, densely-defined (and accretive)
operator on H 2(D), and that W Sn1 is a core for M ∗
It is then easy to verify the
following formula for µb:
Hb .
(3.1)
µb(Sn) =
1
2(cid:0)hHb(S)∗1, Sn1iH 2 + h1, Hb(S)∗Sn1iH 2(cid:1) ,
where m denotes normalized Lebesgue measure on T. Equivalently, using that any kernel
vector kz, for z ∈ D, is necessarily an eigenvector for Hb(S)∗ with eigenvalue Hb(z)∗, one
can check that
(3.2)
Hb(z) = iIm (Hb(0)) + µb(cid:0)(I + zS∗)(I − zS∗)−1(cid:1) .
1
1
=
d
µB(Lα)
:=
This Clark functional formula (3.1) extends verbatim to the non-commutative several-
variable setting:
Definition 3.1. Let (A, B) ∈ Rd × Ld be a transpose-conjugate (A = B†) free Schur
class pair. The Clark functional of (A, B) is the self-adjoint linear functional: µA = µB :
W Lα1 +W(Lα)∗ → C defined by:
In the above, recall that the free monomials always belong to the domain of HA(R)∗,
d in place of its norm closure:
2(cid:16)hHA(R)∗1, Lα1iF 2
2(cid:16)HB(0)δα,∅ + h1, HA(R)∗Lα1iF 2
d(cid:17) .
see Remark 2.11. To simplify notation, we will write Ad +A∗
Ad + A∗
Lemma 3.2. µB extends by continuity to a positive, bounded linear functional on the
norm-closed operator system Ad + A∗
d.
Proof. That Lα1 belongs to the domain of HA(R)∗ (and that, in fact, free polynomials
are a core for HA(R)∗) follows from [10, Corollary 3.9, Corollary 3.10, Remark 3.12]. It
d(cid:17)
+ h1, HA(R)∗Lα1iF 2
is easy to check that µB is positive on Wα Lα +Wα(Lα)∗. Since this is a positive linear
functional, its norm is given by
d = (Ad + A∗
d)−k·k.
kµBk = µB(I) = Re ((HB)∅) = Re (HB(0)) < ∞.
Hence, µB extends by continuity to a bounded positive linear functional on the free disk
operator system.
(cid:3)
10
MICHAEL T. JURY AND ROBERT T.W. MARTIN
Theorem 3.3. The map B 7→ µB is a bijection from Ld onto the set of positive linear
functionals on the free disk system, and one has the free Herglotz formula: For any Z ∈ Bd
n,
HB(Z) = iIm (HB(0n)) + (idn ⊗ µB)(cid:0)(In×F 2 + ZL∗)(In×F 2 − ZL∗)−1)(cid:1) .
In the above,
and
Also note that:
ZL∗ := (Z ⊗ IF 2)(In ⊗ L)∗ = Z1 ⊗ L∗
1 + ... + Zd ⊗ L∗
d,
In×F 2 := In ⊗ IF 2
d
.
(I − ZL∗)−1 =
∞Xk=0
(ZL∗)k = Xα∈Fd
Z α ⊗ (L∗)α.
Proof. Let HB(Z) = Pα HαZ α be the Taylor-Taylor series of HB(Z) about Z = 0 ∈
Bd = Bd
1. Consider (idn ⊗ µB)(cid:0)(I − ZL∗)−1(cid:1):
(idn ⊗ µB)(cid:0)(I − ZL∗)−1(cid:1) = Xα
= Xα
Z α 1
Z αµB((L∗)α)
1, 1E + δα†,∅H ∗
∅(cid:17)
hLγ , 1i Hβ†
1
2(cid:16)DHA(R)∗Lα†
2Xα
Z α Xγβ=α†
2Xα
Z αHα
1
InH ∗
∅ +
HB(0n)∗ +
HB(0n)∗ +
HB(Z).
1
2
=
=
=
1
2
1
2
1
2
In the above, note that HA(R) = U†HB(L)U ∗
the fact that if
† , so that HA(R) = M R
. We also used
H
†
B (Z)
is any free holomorphic function so that M R
Lα1 ∈ Dom((M R
F (Z))∗), and
F (Z) is densely-defined, then the monomials
F (Z) =Xα
Z αFα,
(M R
F (Z))∗Lα1 = Xγβ=α
Lγ 1F ∗
β ,
see for example [10, Corollary 3.13] and [8, Lemma 2.3]. Using that (I+ZL∗)(I−ZL∗)−1 =
2(I − ZL∗)−1 − I, we obtain:
(idn ⊗ µB)(cid:0)(I + ZL∗)(I − ZL∗)−1(cid:1) = HB(0n)∗ + HB(Z) − Re (HB(0n)) ,
and the formula follows.
Conversely, starting with a positive linear functional on the free disk system, this
Herglotz formula defines a free Herglotz function, and by Cayley transform we obtain a
free Schur function whose Clark functional is the original functional. This shows B 7→ µB
is surjective.
(cid:3)
Remark 3.4. Replacing F 2
d ⊗ H where H is a separable or finite-dimensional
Hilbert space, the above results are easily extended to the operator-valued setting of
B ∈ Ld(H). In this operator-valued setting we define the Clark map µB : Ad +A∗
d → L(H)
by the formula:
d with F 2
µB(Lα) :=
1
2
H∅δα,∅IH +
1
2
(h1, ·1i ⊗ idH) (HA(R)∗(Lα ⊗ IH)) .
COLUMN EXTREME MULTIPLIERS OF THE FREE HARDY SPACE
11
Again this defines a bijection between the operator-valued free Schur classes, the operator-
valued free Herglotz-Schur classes, and L(H)−valued completely positive maps on the free
disk system.
Remark 3.5. Theorem 3.3 was first obtained by Popescu in [19, Section 5, Theorem 5.3]
(with a different, but equivalent formula for the Clark functional). Given B ∈ Ld, the
Clark functional µB can also be defined in terms of the Fourier series coefficients of the
free Herglotz function HB, as in [8, Section 4].
4. Free Cauchy transforms
In his seminal paper on unitary perturbations of the shift (see [21] for the fully general,
non-inner case), D.N. Clark showed that there is a canonical isometry, the weighted Cauchy
Transform, Fb, from H 2(µb), the closure of the analytic polynomials in L2(µb) (the Hilbert
space of functions on T which are square-integrable with respect to µb), onto H (b), the
deBranges-Rovnyak space of b ∈ S = L1 [5]: For any polynomial, p ∈ H 2(µb),
(Fbp) (z) := (I − b(z))ZT
1
1 − zζ ∗
p(ζ)µb(dζ).
One can also define an unweighted Cauchy Transform, Cb, from H 2(µb) onto H+(Hb) :=
H(K Hb ), the Herglotz space of b, the unique RKHS corresponding to the positive sesqui-
analytic Herglotz kernel:
K Hb (z, w) :=
1
2
Hb(z) + Hb(w)∗
1 − zw∗
With a bit of algebra, one can verify that
;
z, w ∈ D.
K Hb (z, w) = (I − b(z))−1kb(z, w)(I − b(w)∗)−1,
where kb is the deBranges-Rovnyak kernel of b. The theory of RKHS then implies that
the multiplier
Ub := M(I−b) : H+(Hb) → H (b),
1
is an isometry of the Herglotz space onto the deBranges-Rovnyak space of b. The Cauchy
Transform Cb : H 2(µb) → H+(Hb) is the linear map defined by:
1 − zζ ∗ p(ζ)µb(dζ),
(Cb(p)) (z) :=ZT
and this extends to an isometry of H 2(µb) onto the Herglotz space of b so that Fb = UbCb.
In [8], we extended the notions of Cauchy Transform and weighted Cauchy Transform
to the (operator-valued and) free setting using the theory of free formal RKHS. Here we
describe Cauchy transforms in the setting of NC-RKHS: Assume that B ∈ Ld(H) or that
A ∈ Rd(H) are in the left or right operator-valued free Schur classes and that A = B†
so that µA = µB. The free left Herglotz space H L
+ (HB) = Hnc(K L) is the NC-RKHS
corresponding to the free left Herglotz kernel:
K L(Z, W )[P ] :=
1
2
(HB(Z)(K(Z, W )[P ] ⊗ IH) + (K(Z, W )[P ] ⊗ IH)HB(W )∗) ,
where K is the free Szego kernel. As in the classical theory, it is straightforward to verify
that
is an onto isometric left free multiplier. If A ∈ Rd(H), then the right free Herglotz space,
H R
+ (HA)
onto H R(A).
+ (HA) is defined similarly, and UA = M R
+ (HB) → H L(B),
(I−A†(Z)) is an isometric multiplier of H R
UB := M L
(I−B) : H L
We can expand this kernel in a formal power series (actually a convergent Taylor-Taylor
series about 0):
K L(Z, W )[P ] :=Xα,β
K L
α,β Z αP (W ∗)β†
,
12
MICHAEL T. JURY AND ROBERT T.W. MARTIN
where
K L
α,β := µB((Lα†
)∗Lβ†
),
is the free left formal Herglotz kernel defined in [8, Proposition 4.5]. In the right case, if
A ∈ Rd(H) one simply defines
K R
α,β := µA((Lα)∗Lβ).
As described in [8], given a transpose-conjugate pair (A, B) ∈ Rd(H) × Ld(H), the
appropriate generalization of the (analytic part of the) Clark measure space H 2(µb) is the
Stinespring-Gelfand-Naimark-Segal (S-GNS) space or free Hardy space of µB : Ad +A∗
d →
L(H), F 2(µB). This is defined as the Hilbert space completion of Ad ⊗H (modulo vectors
of zero length) with respect to the pre-inner product:
DLα ⊗ h, Lβ ⊗ gEµB
:=Dh, µB(cid:16)(Lα)∗Lβ(cid:17) gEH
.
The semi-Dirichlet property: (Ad)∗Ad = Ad + A∗
defined inner product, and the left regular representation: Lα 7→ πB(L)α where
d (norm closure) ensures this is a well-
πB(Lα)p(L) ⊗ h := Lαp(L) ⊗ h,
is completely isometric, unital, and extends to a ∗-representation of the Cuntz-Toeplitz
C ∗−algebra. We will set ΠB
k := πB(Lk), so that ΠB = πB(L) is the S-GNS row isometry
on F 2(µB). This also provides a S-GNS formula for µB:
where [I⊗]B : H → F 2(µB) is the bounded embedding:
µB(Lα) = [I⊗]∗
BπB(L)α[I⊗]B,
[I⊗]Bh := I ⊗ h.
The left and right Cauchy transforms, CL : F 2(µB) → H L
+ (HA) are then defined by
H R
+ (HB) and CR : F 2(µA) →
(4.1)
CL(Lα ⊗ h) := K L
α† h,
and,
CR(Lα ⊗ h) := K R
α h.
Observe that if HB(Z) =Pα HαZ α, then HA(Z) =Pα Hα† Z α. One can then calcu-
1
late that:
In the above, we write β ≥ α if β = αγ, and β > α if β ≥ α and β 6= α. Similarly,
K L
α,β =
α,β =
K R
1
(β†\α†)†
2 H ∗
β† > α†
2 H(α†\β†)† α† > β†
Re (H∅)
α = β
else.
0
1
(β\α)†
2 H ∗
β > α
1
2 H(α\β)† α > β
α = β
Re (H∅)
else.
0
These formulas follow easily from the Clark map formula. For example if β > α (and
H = C) then
K R
α,β = µA(Lβ\α)
=
= 0 +
1
2(cid:16)DHA(R)∗1, Lβ\α1E +D1, HA(R)∗Lβ\α1E(cid:17)
2 Xγλ=β\α
h1, Lγ 1i H ∗
λ†
1
=
(H)∗
(β\α)†.
1
2
The above formulas allow one to alternatively define the Clark map of B in terms of the
Fourier series coefficients of the Herglotz functions HB, as was done in [8, Section 4].
(cid:3)
.
COLUMN EXTREME MULTIPLIERS OF THE FREE HARDY SPACE
13
Lemma 4.1. The left free Cauchy transform acts as:
(CLp(L)1)(Z) = (idn ⊗ µB)(cid:0)(In×F 2 − ZL∗)−1(In ⊗ p(L))(cid:1) .
Proof. It suffices to check on monomials, so take p(L) = Lβ†
1. Then, the above becomes:
Z αµB((Lα†
)∗Lβ†
) = K L
α (Z),
Xα
as claimed.
Lemma 4.2. Given any Z ∈ Bd
, (I − Z∗L)−1(y ⊗ 1)(cid:17)Cn
n and v, y ∈ Cn,
(Z αv, y)Cn Lα†
1
K L{Z, y, v} = CLXα
d
DK L{Z, y, v}, FE = (y, F (Z)v)Cn
= CL(cid:16)v ⊗ IF 2
Proof. For any F ∈ H L,+(HB), we have that
= Xα
= Xα
= *Xα
(y, Z αv)Cn Fα
(y, Z αv)CnDK L
(Z αv, y)Cn K L
The above proves that K L {Z, y, v} = Pα (Z αv, y)Cn K L
CLLα†
1. Hence we have that:
α , FE
α , F+ .
K L{Z, y, v} = CLX (Z αv, y)Cn Lα†
1
, (Z∗)α†
= CLX(cid:16)v ⊗ IF 2
= CL(cid:16)v ⊗ IF 2
d
d
⊗ Lα†
(y ⊗ 1)(cid:17)Cn
, (I − Z∗L)−1(y ⊗ 1)(cid:17)Cn
.
α , and by definition, K L
α =
(cid:3)
)(cid:17)
Remark 4.3. We also have the formula:
K L {W, x, u} (Z) = Xα
(W αu, x)Cn (idn ⊗ µB)(cid:16)(I − ZL∗)−1(In ⊗ Lα†
⊗ Lα†!
= ((·u, x)Cn ⊗ µB) (I − ZL∗)−1Xα
= ((·u, x)Cn ⊗ µB)(cid:0)(I − ZL∗)−1(I − LW ∗)−1(cid:1) .
µb((I − zL∗)−1(I − Lw∗)−1) = K b(z, w),
(W ∗)α†
The above is the free version of the commutative Cauchy transform formula,
from [11, Proposition 2.6, Subsection 2.8]. Here K b(z, w) is the positive Herglotz kernel
for b in the Schur class of contractive Drury-Arveson space multipliers.
If A = B† so that µA = µB and F 2(µA) = F 2(µB), then the weighted free Cauchy
transforms F L, F R, are defined as:
F L = M L
(I−B)CL,
and, F R = M R
(I−A†)CR,
and these are isometries of F 2(µB) onto H L(B) and H R(A), respectively.
14
MICHAEL T. JURY AND ROBERT T.W. MARTIN
4.4. Cauchy Transform of the Stinespring-GNS representation. As in the com-
mutative setting, if B ∈ Ld(H) we define
V B := CLπµB (L)(CL)∗,
a row isometry on the left Herglotz space H L
Proposition 4.5. The range R of the row isometry V B is:
+ (HB).
and for any Z ∈ Bd
(V B
R :=_(cid:16)K HB{Z, y, v} − K HB{0n, y, v}(cid:17) = _α6=∅
n, v, y ∈ Cn, and j = 1, . . . , d,
j )∗(cid:16)K HB{Z, y, v} − K HB{0n, y, v}(cid:17) = K HB{Z, y, Zj v}
K HB
α ,
(so that the span of all such vectors is dense in H L
The image of Ran(cid:0)V B(cid:1) under (CL)∗ is F 2
non-constant free monomials in F 2(µB). If F ∈ H L
then there is a f ∈ H so that for any Z ∈ Bd
n,
+ (HB) ⊗ Cd).
0 (µB) = Wα6=∅ Lα ⊗ H, the closure of the
+ (HB) is orthogonal to Ran(cid:0)V B(cid:1),
F (Z) = In ⊗ f,
i.e. F ≡ f is constant-valued.
Proof. By the proof of [10, Lemma 3.14], for any α ∈ Fd, one can find jointly nilpotent
Z ∈ Bd
n and v, y ∈ Cn with n = α + 1 so that
(cid:16)Z β v, y(cid:17)Cn
= δα,β .
It then follows from Lemma 4.2 and the definition of left free Cauchy transform that
K L{Z, v, y} = K L
α . This shows that the two formulas for R above are the same. By
definition, CL(Lα ⊗ h) = K L
α† h, and it follows that the image of R under inverse Cauchy
transform is F 2
0 (µB) = Ran (πB(L)). Since V B and ΠB are unitarily equivalent under
Cauchy transform, it follows that R = Ran(cid:0)V B(cid:1).
+ (HB) is orthogonal to
Ran(cid:0)V B(cid:1), set f := (K HB
If F ∈ H L
0 = DK HB {Z, y, v}, FE −DK HB{0n, y, v}, FE
)∗F ∈ H. Then, for any Z ∈ Bd
n and v∗, y ∈ Cn,
0
= (y ⊗ IH, F (Z)v) − (In ⊗ f ) (y, v)Cn ,
and it follows that F (Z) = In ⊗ f .
The second claim is a straightforward calculation: for each j = 1, . . . , d,
(V B
j )∗(cid:16)K L{Z, y, v} − K L{0n, y, v}(cid:17) = CLπ(Lj)∗Xα6=∅
(Z αv, y)Cn Lα†
1
= CLXβ (cid:16)Z βjv, y(cid:17)Cn
= K L{Z, y, Zj v}.
Lβ†
1
Since V B is an isometry, the above shows that the closed span of ⊕d
all of H L
j=1K L{Z, y, Zj v} is
(cid:3)
+ (HB) ⊗ Cd.
5. Gleason solutions
5.1. The free setting. Fix A ∈ Rd(H, J ). Exactly as in the commutative setting, we
define:
COLUMN EXTREME MULTIPLIERS OF THE FREE HARDY SPACE
15
Definition 5.2. A linear map X : H R(A) → H R(A) ⊗ Cd is called a Gleason solution
for H R(A) if:
(5.1)
Such an X is contractive if
Z(Xf )(Z) = f (Z) − f (0n)
∀ Z ∈ Bd
n.
(5.2)
and extremal if equality holds.
X ∗X ≤ I − K A
0 (K A
0 )∗,
Similarly, a linear map A : H → H R(A) ⊗ Cd is called a Gleason solution for A if
ZA(Z) = A†(Z) − A(0)In
∀ Z ∈ Bd
N.
and extremal if equality holds.
A∗A ≤ IH − A(0)∗A(0),
(5.3)
A is contractive if
(5.4)
Define:
(5.5)
X := L∗ ⊗ IJH R(A),
and
A := (L∗ ⊗ IJ )A.
The right free deBranges-Rovnyak space, H R(A), is left shift co-invariant, L∗
j A ∈ H R(A),
and X, A obey the contractivity conditions of Gleason solutions for H R(A), A, respec-
tively [2, Proposition 4.2]. It is also easy to check that X, A are Gleason solutions. For
example, given f =Pα fαLα1, it is clear that
( Xj f )(Z) =Xα
fαLα\j1,
where we set Lα\j = Lβ if β = jα, and = 0 else. It follows that
fαZ α = f (Z) − f (0n).
Z( Xf )(Z) =Xα6=∅
Also note that the defining formula (5.1) for a Gleason solution for H R(A) is equivalent
to:
which can be re-arranged to:
(L∗K A
Z )∗(Xf ) = (K A
Z )∗f − (K A
0n )∗f,
(5.6)
Z = K A
0n .
Given A ∈ Rd(H,J ), the support of A is defined to be
(I − X ∗L∗)K A
n; v∗,y∈Cn(cid:16)y ⊗ IH, A†(Z)∗v∗ ⊗ J(cid:17) ⊆ H.
_Z∈Bd
(5.7)
supp(A) :=
Proposition 5.3. Suppose that A ∈ Rd(H,J ). A linear map X : H R(A) → H R(A)⊗Cd
is a contractive Gleason solution for H R(A) if and only if there is a contractive Gleason
solution A : H → H R(A) ⊗ Cd for A so that,
XK A{W, x, u} = K A{W, W ∗x, u} − A(cid:16)u, A†(W )∗x(cid:17)Cm ∈ L(J , H R(A) ⊗ Cd).
X is extremal if A is extremal. Conversely A is extremal if X is extremal and supp(A) =
H.
This is a free analogue of [11, Theorem 4.4]. Since the proof is (formally) analogous,
we prove only the sufficiency.
Remark 5.4. The expression (cid:0)u, A†(W )∗x(cid:1)Cm is to be interpreted as taking values in
m, A†(W )∗ ∈ Cm×m ⊗ L(J ,H). Namely, for any g ∈ J , the
L(J ,H) since, for W ∈ Bd
above formula can be written:
XK A{W, x, u}g = K A{W, W ∗x, u}g − A(cid:16)u ⊗ IH, A†(W )∗x ⊗ g(cid:17) .
16
MICHAEL T. JURY AND ROBERT T.W. MARTIN
Proof. Let A be a contractive Gleason solution for A. We wish to show that the formula
in the proposition statement defines a contractive Gleason solution for H R(A). To prove
this, it is sufficient to check that Formula (5.1) holds on kernel vectors. Namely, it suffices
to show that
(5.8)
DK A{Z, Z∗y, v}, XK{W, x, u}E =(cid:16)y, Z(XK A{W, x, u})(Z)v(cid:17)Cn
In the above we have used the compact notation:
= (cid:16)y,(cid:16)K A(Z, W )[vu∗] − K A(0n, W )[vu∗](cid:17) x(cid:17)Cn ∈ L(J ).
DK A{Z, Z∗y, v}, XK{W, x, u}E :=
dXj=1DK A{Z, Z∗
and we will continue to use this throughout. Calculate:
j y, v}, Xj K{W, x, u}E ,
DK A{Z, Z∗y, v}, XK {W, x, u}E =DK A{Z, Z∗y, v}, K A{W, W ∗x, u}E
−DK A{Z, Z∗y, v}, AE(cid:16)u, A†(W )∗x(cid:17)Cm
= (cid:16)y, ZK A(Z, W )[v∗u]W ∗x(cid:17)Cn − (y, ZA(Z)v∗)Cn(cid:16)u∗, A†(W )∗x(cid:17)Cm
= (cid:16)y,(cid:16)K A(Z, W )[vu∗] − vu∗ + A†(Z)vu∗A†(W )∗(cid:17) x(cid:17)Cn
= (cid:16)y,(cid:16)K A(Z, W )[vu∗] − vu∗ + A†(Z)vu∗A†(W )∗(cid:17) x(cid:17)Cn
+(cid:16)y, A(0n)vu∗A†(W )∗x(cid:17)Cn
= (cid:16)y,(cid:16)K A(Z, W )[vu∗] − K A(0n, W )[vu∗](cid:17) x(cid:17)Cn ∈ L(J ).
1 = Bd, since A† is
In the above, note that A†(0n) = A(0n) = A(0)In = A∅In where 0 ∈ Bd
a free function. This proves that X is a Gleason solution. To see that X is contractive,
again calculate on kernel vectors:
kXK A{Z, y, v}k2 = DK A{Z, Z∗y, v}, K A{Z, Z∗y, v}E
−(cid:16)y, (A†(Z) − A†(0n))v(cid:17)Cn(cid:16)u, A†(W )∗x(cid:17)Cm
−(cid:16)y, A†(Z)vu∗A†(W )∗x(cid:17)Cn
2
−DK A{Z, Z∗y, v}, AE(cid:16)A†(Z)v, y(cid:17)Cn
−(cid:16)y, A†(Z)v(cid:17)CnDA, K A{Z, Z∗y, v}E + kAk2(cid:12)(cid:12)(cid:12)(cid:16)A†(Z)v, y(cid:17)Cn(cid:12)(cid:12)(cid:12)
≤ (cid:16)y, ZK A(Z, Z)[vv∗]Z∗y(cid:17)Cn −(cid:16)y, (A†(Z) − A(0n))v(cid:17)Cn(cid:16)A†(Z)v, y(cid:17)Cn
−(cid:16)y, A†(Z)v(cid:17)Cn(cid:16)(A†(Z) − A(0n))v, y(cid:17)Cn
+(I − A(0)∗A(0))(cid:16)y, A(Z)vv∗A†(Z)∗y(cid:17)Cn
−2(cid:16)y, A†(Z)vv∗A†(Z)∗y(cid:17)Cn
+(I − A(0)∗A(0))(cid:16)y, A†(Z)vv∗A†(Z)∗y(cid:17)Cn
= (cid:16)y, K A(Z, Z)[vv∗]y(cid:17)Cn − (y, vv∗y)Cn +(cid:16)y, A(0n)vv∗A†(Z)∗y(cid:17)Cn
+(cid:16)A(0n)vv∗A†(Z)∗y, y(cid:17)Cn − A(0)∗A(0)(cid:16)y, A†(Z)vv∗A†(Z)∗y(cid:17)Cn
= (cid:16)y, K A(Z, Z)[vv∗]y(cid:17)Cn −(cid:16)y, (vv∗ − A†(Z)vv∗A†(Z)∗)y(cid:17)Cn
+(cid:16)y, A(0n)vv∗A†(Z)∗y(cid:17)Cn
(5.9)
.
+(cid:16)A(0n)vv∗A†(Z)∗y, y(cid:17)Cn
COLUMN EXTREME MULTIPLIERS OF THE FREE HARDY SPACE
17
Observe that equality holds in the above if A is extremal. Compare this to:
0 (K A
0 )∗)K A{Z, y, v}E
DK A{Z, y, v}, (I − K A
=(cid:16)y, K A(Z, Z)[vv∗]y(cid:17)Cn − kK A{Z, y, v}(0)k2
= kK A{Z, y, v}k2 −(cid:12)(cid:12)(cid:12)(cid:16)y, K A(0n, Z)v(cid:17)Cn(cid:12)(cid:12)(cid:12)
= kK A{Z, y, v}k2 −(cid:12)(cid:12)(cid:12)(cid:16)y,(cid:16)In − A(0n)A†(Z)∗(cid:17) v(cid:17)Cn(cid:12)(cid:12)(cid:12)
= kK A{Z, y, v}k2 − (y, vv∗y)Cn + 2Re(cid:16)(y, v)Cn(cid:16)y, A(0)A†(Z)∗∗(cid:17)Cn(cid:17)
−A(0)∗A(0)(cid:16)y, A†(Z)∗v(cid:17)Cn(cid:16)A†(Z)∗v, y(cid:17)Cn
2
2
,
which is the same (up to elementary manipulations) as Equation (5.9) above. This proves
that X ∗X ≤ I − K A
0 )∗ so that X is a contractive Gleason solution (which will be
extremal if A is).
(cid:3)
0 (K A
Theorem 5.5. Suppose that A ∈ Rd(H). Then A := (L∗ ⊗ IH)A and X := (L∗ ⊗
IH)H R(A) are the unique contractive Gleason solutions for A and H R(A), respectively.
The proof uses similar arguments to those of [11, Section 4].
Lemma 5.6. ([8, Proposition 6.2]) The contractive Gleason solution A = (L∗ ⊗ IH)A is
given by the formula
A = F RΠ∗
A[I⊗]A(I − A(0))
= UA(V A)∗K HA
0
(I − A(0)),
where ΠA = πA(L) is the row isometry obtained from the S-GNS representation of the free
Clark measure µA.
Recall that UA := M R
V A is the row isometry on H L
(I−A†(Z)) is the unitary multiplier of H R
+ (HA) onto H R(A), and
+ (HA) defined in Subsection 4.4.
+ (HA) → H R
+ (HA) ⊗ Cd by:
Proof. (of Theorem 5.5) Let A be any contractive Gleason solution for A. Define a linear
map D∗ : H R
(5.10) D∗K HA {Z, y, v} := K HA{Z, Z∗y, v} + M R
Recall U ∗
construction,
(I−A†(Z))−1 , the unitary right multiplier of H R(A) onto H R
(I−A†(Z))−1 A(I − A(0))−1 (v, y)Cn .
+ (HA). By
A := M R
D∗ (K {Z, y, v} − K{0n, y, v}) = K HA{Z, Z∗y, v}
= (V A)∗ (K {Z, y, v} − K{0n, y, v}) .
We claim that D∗ is a contraction:
kD∗K HA {Z, y, v}k2 = kK HA{Z, Z∗y, v}k2
AAE (I − A(0))−1 (v∗, y)Cn + c.c.
+DK HA{Z, Z∗y, v},U ∗
+ (y, v)Cn (I − A(0)∗)−1 hA, Ai (I − A(0))−1 (v, y)Cn
≤ kK HA{Z, Z∗y, v}k2 +DK HA{Z, Z∗y, (I − A†(Z))−1v}, AE (I − A(0))−1 (v, y)Cn
+c.c. + K HA (0, 0) (y, vv∗y)Cn ,
(5.11)
18
MICHAEL T. JURY AND ROBERT T.W. MARTIN
where c.c. denotes the complex conjugate of the previous term. The cross-term becomes:
1
=
DK HA{Z, Z∗y, (I − A†(Z))−1v}, AE (I − A(0))−1 (v, y)Cn
= (cid:16)y, (I − A(0n))−1(A†(Z) − A(0n))(I − A†(Z))−1v(cid:17)Cn
2(cid:16)y, (H †
kK HA{Z, Z∗y, v}k2 +
A(0n))(vv∗)y(cid:17)Cn
2(cid:16)y,(cid:16)H †
A(Z)vv∗ + vv∗H †
A(Z) − H †
1
.
A(Z)∗(cid:17) y(cid:17)Cn
(v, y)Cn
It follows that Equation (5.11 ) becomes:
= kK HA{Z, y, v}k2.
This proves that D∗ is a contractive extension of (V A)∗ so that D is a row contractive
extension of V A (by, for example, [11, Lemma 2.3]). However V A is a row isometry and
has no non-trivial extensions. Hence, D = V A, and Equation (5.10) and Lemma 5.6 then
imply that A = (L∗ ⊗ IH)A = A. It follows also that X is unique, by Proposition 5.3, so
that X = (L∗ ⊗ IH)H R(A).
(cid:3)
If B ∈ Ld(H), the formula is similar:
(5.12)
B = M L
(I−B(Z))(V B)∗K HB
0
(I − B(0)).
6. Column Extreme
Recall that B ∈ Ld(H) or A ∈ Rd(H) is said to be quasi-extreme if F 2
0 (µB) = F 2(µB),
i.e. if and only if
I ⊗ H ⊆ F 2
0 (µB),
see [8] and [11, Definition 3.19]. This concept of quasi-extreme was first introduced for
contractive scalar multipliers of the Drury-Arveson space in [9], extended to operator-
valued multipliers b ∈ Sd(H) in [11], and to the 'rectangular setting' of arbitrary b ∈
Sd(H,J ) in [15]. (Here Sd(H,J ) denotes the Schur class of contractive operator-valued
multipliers between vector-valued Drury-Arveson spaces.) The main result of [12] shows
that a more descriptive name for this property could be column extreme (CE), and we
will use this new terminology for the remainder of this paper.
Definition 6.1. A Schur class B ∈ Ld(H,J ) is column extreme (CE) if there is no
non-zero A ∈ Ld(H, J ) so that the column:
A(cid:19) ∈ Ld(H,J ⊗ C2),
(cid:18)B
is also Schur class. Column extreme for the right Schur class is defined analogously.
Remark 6.2. Observe that the definition of column extremity can be recast as follows:
B is column extreme if and only if the only multiplier A satisfying the inequality
(6.1)
M L∗
A M L
A ≤ I − M L∗
B M L
B
is A = 0. The existence of such A for given B was considered by Popescu [17], who showed
that a nonzero A exists if and only if e(I − M L∗
B ) > −∞, where e(·) is the so-called
entropy of a multi-analytic Toepliz operator as defined in [17]. However, it seems to be
difficult to compute the entropy for arbitrary B (or even to decide if it is finite or not).
Regarding the equivalences in Theorem 6.4 below, it is not hard to see from the definition
of the entropy invariant, that e(I − M L∗
B ) = −∞ is equivalent to our condition (5), so
that the equivalence of (1) and (5) is essentially contained in [17, Corollary 1.2].
B M L
B M L
COLUMN EXTREME MULTIPLIERS OF THE FREE HARDY SPACE
19
In this general 'rectangular' setting, it will often be convenient to consider the square
completion [B], of B: The above column-extreme property is clearly invariant under
conjugation by isometries; a given B ∈ Ld(H,J ) is CE if and only if B′ = V BW ∗ is CE,
where W : H → H′, and V : J → J ′ are fixed onto isometries. It follows that we can
assume, without loss of generality, that H ⊆ J or J ( H, and complete B to a 'square'
[B] ∈ Ld(J ) or in Ld(H), respectively by adding columns or rows of zeros:
[B] :=
(cid:2)B 0J ⊖H(cid:3) H ( J
0H⊖J(cid:21)
(cid:20) B
J ( H.
Remark 6.3. Observe that if H ( J then H L(B) = H L([B]) so that the unique
contractive Gleason solution for B is given by [B]H, where [B] is the unique contractive
Gleason solution for [B]. It is clear that B is extremal if and only if [B] is extremal in
this case.
In the second case where J ( H we have that H L([B]) = H L(B)L(cid:0)F 2
and the unique contractive Gleason solution for [B] is given by:
d ⊗ (H ⊖ J )(cid:1),
[B] = (L∗ ⊗ IH)(cid:18) B
0H⊖J(cid:19) =(cid:18) B
0H⊖J(cid:19) ,
where 0H⊖J : H ⊖ J → H L(B) ⊗ Cd maps everything to the zero element. In this case
it is clear that
and it follows as before that B is extremal if and only if [B] is extremal.
Theorem 6.4. Given B ∈ Ld(H, J ), the following are equivalent:
B =(cid:0)I,
0(cid:1) [B],
(1) B is column extreme.
(2) The unique contractive Gleason solution B = R∗B : H → H L(B) ⊗ Cd, for B is
(3) The unique contractive Gleason solution X = R∗H R(B) for H R(B) is extremal,
extremal.
and H = supp(B).
(I − B(0))H ⊆ Ran(cid:16)V [B](cid:17).
H[B]
(4) K
0
(5) B has the Szego extremal property: I ⊗ (I − B(0))H ⊆ F 2
0 (µ[B]).
(6) There is no non-zero H-valued constant function H ≡ h ∈ H L
(7) There is no non-zero h ∈ H so that Bh ∈ H L(B).
+ (H[B]).
If B = [B] is square, then the above are equivalent to:
(8) πµB (L) (equivalently V B) is a Cuntz row isometry.
In the above, recall that F 2
defined in Equation (5.7).
0 (µB) =Wα6=∅ Lα ⊗ IH ⊆ F 2(µB), and the support of B was
Remark 6.5. In the classical (single-variable, scalar-valued) setting, the equivalent state-
ments in the above theorem recover several characterizations of extreme points of the Schur
class, S , of contractive analytic functions on D:
Theorem. Given b ∈ S , the following are equivalent:
(0) b is an extreme point.
(1) b is column extreme.
(2) S∗b is extremal, i.e. kS∗bkH (b) = 1 − b(0)2.
(3) X := S∗H (b) is extremal, i.e., X ∗X = I − kb
(5) H 2(µb) = H 2
(7) b does not belong to H (b).
(8) All Clark perturbations of S∗H (b) are unitary.
0 (µb).
0(kb
0)∗.
20
MICHAEL T. JURY AND ROBERT T.W. MARTIN
Conditions (1) and (2) are characterizations of extreme points of S which follow from
results of Sarason, [21, Chapter III, Chapter IV]. The spaces F 2(µB), and F 2
0 (µB) are
multi-variable non-commutative analogues of the spaces H 2(µb), and H 2
0 (µb), the clo-
sure of the analytic polynomials and closed span of the non-constant analytic monomials
0 (µb) = H 2(µb)
(resp.) in L2(µb), for b ∈ S . In the classical setting the condition that H 2
is equivalent to b being an extreme point of the Schur class. This follows from the
Szego-Kolmogoroff-Kreın distance formula for the distance from the constant function
1 to H 2
0 (µb), the formula for the Radon-Nikodym derivative of µb with respect to normal-
ized Lebesgue measure, and other classical facts, see [7, Chapter 4, Chapter 9], and the
discussion in [8, Section 1]. Item (7) is again a result of Sarason [21, Chapter IV], and
the final item is equivalent to the well-known fact that b is extreme if and only if all of
the Clark perturbations of the restricted backward shift S∗H (b) are unitary, see e.g. [14].
Corollary 6.8 will prove that any column-extreme B ∈ Ld or Rd is necessarily an extreme
point, whether the converse holds is an open problem.
We will need the following free or NC analogue of a result from vector-valued RKHS
theory, [16, Theorem 10.17] (see the proof of [12, Proposition 5.1]):
Lemma 6.6. Let Hnc(K) be a vector-valued NC-RKHS on a NC set Ω. A (vector-valued)
free NC function f on Ω belongs to Hnc(K) if and only if λ2K(Z, W )− f (Z)(·)f (W )∗ ≥ 0
is a (operator-valued) CPNC Kernel on Ω for some λ2 > 0. The norm of f is the infimum
of all such λ.
Lemma 6.7. If B ∈ Ld(H,J ) and C := [B] ∈ Ld(K), then Bh ∈ H L(B) for h ∈ H if
and only if F ≡ (I − C(0)∗)−1h ∈ H is a constant function in H L
Proof. If Bh ∈ H L(B) set f := (I − C(0)∗)−1h ∈ H. Then,
+ (HC).
(K B
0 f )(Z) = In ⊗ f − B(Z)(In ⊗ B(0)∗(I − C(0)∗)−1h)
= In ⊗ f − B(Z)C(0)∗(I − C(0)∗)−1h
= In ⊗ f − B(Z)(I − C(0)∗)−1h + B(Z)h
= (In − B(Z))f + B(Z)h.
Since B(Z)h = B(Z)(In⊗h) ∈ H L(B), we conclude that F ≡ f ∈ H belongs to H L
(it is the image of (I − B)f under the canonical unitary multiplier).
In ⊗ h, then M L
(I−B(Z))H = F ∈ H L(C), and also
+ (HC) is such that H ≡ h ∈ H, i.e. H(Z) =
This argument is reversible:
if H ∈ H L
+ (HC)
so that
0 (Z)h = In ⊗ h − B(Z)(In ⊗ B(0)∗h),
K B
(K B
0 − F )(Z) = B(Z)(I − B(0)∗),
and we conclude that B(I − C(0)∗)h = B(I − B(0)∗)h ∈ H L(C).
If J ⊇ H so that
[B] ∈ Ld(J ), then H L(C) = H L(B). Otherwise if H ) J then H L(C) = H L(B) ⊕
d ⊗(H⊖J ). Since C(Z)h = B(Z)h for any h ∈ H, and B(Z) ∈ Cn×n⊗ L(H,J ), Lemma
F 2
6.6 implies that there is a λ2 > 0 so that
(cid:18)B(Z)g(B(W )g)∗
0
0
0H⊖J(cid:19) = C(Z)g(C(W )g)∗
≤ λ2K C (Z, W ) = λ2(cid:18)K B(Z, W )
where g := (I − B(0)∗)h = (I − C(0)∗)h ∈ H. Comparing top left entries, Lemma 6.6
again implies that B(Z)(I − B(0)∗)h = Bg ∈ H L(B).
K(Z, W ) ⊗ (H ⊖ J )(cid:19) ,
0
0
(cid:3)
The proof of equivalence of the first two items is the most involved, so we will first
establish the equivalence of the remaining items.
COLUMN EXTREME MULTIPLIERS OF THE FREE HARDY SPACE
21
Proof. (of equivalence of items (2) − (10) in Theorem 6.4) To simplify notation, we will
write C = [B] ∈ Ld(K) (with K = H or J ).
(2) ⇔ (3). This was proven as part of Proposition 5.3 which relates Gleason solutions for
H L(B) and Gleason solutions for B. Also note that if supp(B) ( H, then B will have a
matrix representation of the form:
(cid:0)B′,
so that for any Schur A′ ∈ Ld,
A(cid:19) ∈ Ld(H ⊕ C,J ⊕ C),
0 A′(cid:19) =:(cid:18)B
(cid:18)B′
0(cid:1) ,
0
is Schur and B is not CE in this case.
(2) ⇔ (4). There are two cases to consider. If C = [B] ∈ Ld(H) then H L(B) ⊕ F 2
(H ⊖ J ) = H L(C) and B is extremal if and only if C is. If C is extremal then
d ⊗
I − C(0)∗C(0) = C∗C
0
= (I − C(0)∗)(K HC
)∗V C (V C )∗K HC
≤ (I − C(0)∗)K HC (0, 0)(I − C(0))
= I − C(0)∗C(0),
0
(I − C(0))
and this happens if and only if V C is a co-isometry. This establishes the equivalence in
this case. Alternatively, if H ⊆ J then B = CH so that B will be extremal if and only if
I − B(0)∗B(0) = B∗B
= PHC∗CPH
= PH(I − C(0)∗)(K HC
≤ PH(I − C(0)∗C(0))PH = I − B(0)∗B(0).
)∗V C(V C )∗K HC
0
0
(I − C(0))PH
0
0
(I − C(0))H = K HC
0 (µC ) onto Ran(cid:0)V C(cid:1), and which maps I ⊗ g ∈ F 2(µC ) to K HC
(I − B(0))H ⊆ Ran(cid:0)V C(cid:1), and
The above holds if and only if K HC
this proves the equivalence in the second case.
(4) ⇔ (5). This follows immediately from the fact that the Cauchy transform CL :
F 2(µC ) → H L
+ (HC) is an onto isometry which intertwines ΠC = πC (L) and V C , which
takes F 2
g, see Proposition
4.5.
(4) ⇒ (6). Assume that (6) does not hold so that there is a constant H−valued function
+ (HC). Set f := (I − B(0)∗)h ∈ H so that h = (I − C(0)∗)−1f . If (4)
H ≡ h ∈ H in H L
(I − C(0))f ∈ Ran(cid:0)V C(cid:1) ⊥ H so that
holds then K HC
0 =DK HC
(I − B(0))f , HE =(cid:10)(I − C(0))f , (I − C(0)∗)−1f(cid:11) = kfk2.
We conclude that f = 0 so that h = 0 (since B(0)∗ is a pure contraction). This shows
that (4) cannot hold.
(6) ⇔ (7). This equivalence is an immediate consequence of Lemma 6.7.
(7) ⇒ (4). Our proof will be a bit circuitous: First consider the following condition (4)′:
If F ∈ H L
0
0
0
+ (HC) is constant valued, F ≡ f ∈ K, then
PH(I − B(0)∗)f = 0.
We claim that (4)′ ⇒ (4). Condition (4)′ implies that
PH(I − B(0)∗)(K HC
and taking the adjoint of this expression gives:
0
)∗(I − V V ∗) = 0,
(I − V V ∗)K HC
0
(I − B(0))PH = 0,
which is condition (4). It remains to show that (7) ⇒ (4)′, and this will be accomplished
If (4)′ does not hold then there is a f ∈ K so
by demonstrating the contrapositive.
+ (HC), and PH(I − B(0)∗)f 6= 0. By Lemma 6.7,
that F (Z) = In ⊗ f belongs to H L
22
MICHAEL T. JURY AND ROBERT T.W. MARTIN
(I − C(0)∗)f = g is such that Cg ∈ H L(C). There are two cases:
If H ⊆ J then
In this case Cg = BPHg ∈ H L(B), and by assumption, PHg =
H L(C) = H L(B).
PH(I − B(0)∗)f 6= 0. We conclude that (7) does not hold in this case. In the second case
J ⊆ H, and g = (I− B(0)∗)f ∈ H is such that Cg = Bg ∈ H L(C) = H L(B)⊕ F 2
d ⊗(H⊖
J ). Since B(Z) ∈ Cn×n ⊗ L(H,J ), one can apply Lemma 6.6 (as in the proof of Lemma
6.7) to show that Bh ∈ H L(B), and again (7) does not hold. Hence (7) ⇒ (4)′ ⇒ (4).
Assuming now that B = [B] = C ∈ Ld(H) is square, item (6) is equivalent to the
statement that V C is a Cuntz (onto) row isometry, and since V C is unitarily equivalent
to ΠC = πC (L) via Cauchy transform, it follows that (6) ⇔ (8).
(cid:3)
The proof of (1) ⇔ (2) is the free and operator-valued extension of the main result of
[12], and the argument is formally analogous.
Proof. ((1) ⇒ (2) of Theorem 6.4) Suppose B ∈ Ld(H,J ) is not column-extreme so that
A(cid:19) is Schur.
there is a non-zero A ∈ Ld(H, J ) so that the two-component column C :=(cid:18)B
Without loss of generality, we can assume that 0 6= A∅ = A(0) ∈ L(H,J ). The argument
is as in [12, Lemma 5.2]: If A∅ = 0 choose α ∈ Fd of minimal length so that Aα 6= 0, and
set: eA := (Lα)∗ ⊗ IJ A. Then
(cid:18)B
eA(cid:19) =(cid:18)I
C =(cid:0)R∗ ⊗ IJ , R∗ ⊗ IJ(cid:1)(cid:18)B
is also Schur and satisfies eA∅ = Aα 6= 0. The unique contractive Gleason solution, C for
A(cid:19) ,
(Lα)∗ ⊗ IJ(cid:19)(cid:18)B
A(cid:19) ,
A(cid:19) =(cid:18)B
where B, A are the unique contractive Gleason solutions for B, A. Observe that
C∗C ≤ I − C(0)∗C(0) = I − B(0)∗B(0) − A(0)∗A(0) < I − B(0)∗B(0),
C is:
0
0
since we can assume A(0) 6= 0.
H L(C) for any h ∈ H, where 1 ≤ j ≤ d, there is a tj > 0 so that as CPNC kernels,
Now we apply the argument of [12, Proposition 5.1]: By Lemma 6.6, since each Cj h ∈
∗
∗(cid:19)
j(cid:18)K B(Z, W )
∗
∗
∗(cid:19) ,
(Cj(Z)h)(Cj(W )h)∗ = (cid:18)(Bj(Z)h)(Bj (W )h)∗
∗
≤ t2
j K C (Z, W ) = t2
and one can take tj := kCj hk. It follows that for any h ∈ H,
kBhk2 = hh, B∗Bhi
t2
j
dXj=1
≤
= hh, C∗Chi
≤ hh, (I − C(0)∗C(0))hi
= hh, (I − B(0)∗B(0) − A(0)∗A(0))hi .
This proves that
B∗B < I − B(0)∗B(0),
so that B is not extremal.
(cid:3)
COLUMN EXTREME MULTIPLIERS OF THE FREE HARDY SPACE
23
The proof of (2) ⇒ (1) will employ the colligation and transfer-function theory of [2].
We briefly recall the pertinent facts: A colligation is any contractive linear map:
U :=
A1 B1
...
...
Ad Bd
C D
=:(cid:18)A B
C D(cid:19) :(cid:18)K
H(cid:19) →(cid:18)K ⊗ Cd
J (cid:19) .
The transfer-function of the contractive colligation U is the function BU defined on the
free unit ball by:
BU (Z) := D + C(I − ZA)−1ZB ∈ Cn×n ⊗ L(H, J );
Z ∈ Bd
n,
where ZA := Z1A1 + ...ZdAd. The theory of [2] shows that a free function B on Bd
belongs to the left free Schur class if and only if B = BU is the transfer function of some
contractive colligation U (see [2, Theorem 3.1]). Moreover, any B ∈ Ld(H,J ) always has
the (left) canonical deBranges-Rovnyak colligation
N
CdBR DdBR(cid:19)
UdBR :=(cid:18)AdBR BdBR
AdBR := R∗H L(B), BdBR := R∗B, CdBR := (K B
constructed by choosing K := H L(B) and
0 )∗, and DdBR := B(0),
(so that B is recovered as the transfer function of this colligation) see [2, Theorem 4.3].
Similarly, if A ∈ Rd(H,J ) then A(R) = M R
A†(Z) is such that the free holomorphic func-
tion A†(Z) can be recovered as the transfer function of the (right) canonical deBranges-
Rovnyak colligation given by choosing K := H R(A), and
AdBR := L∗H R(A), BdBR := L∗A, CdBR := (K A
0 )∗, and DdBR := A(0).
Proof. (of (2) ⇒ (1)) We give the proof for right free multipliers A ∈ Rd(H, J ). Assuming
A is not extremal, we choose 0 ≤ a∅ ∈ L(H) satisfying:
∅ = I − A(0)∗A(0) − A∗A.
a2
As in the proof of Proposition 5.3 one can calculate that,
(6.2)
= (cid:16)y, K A(Z, W )[vu∗]x(cid:17)Cn − (y, ZA(Z)v)Cn(cid:16)u, A†(W )∗x(cid:17)Cm
DX ∗XK A{Z, y, v}, K A {W, x, u}E
−(cid:16)A†(Z)∗y, v(cid:17)Cn
(W A(W )u, x)Cm +(cid:16)y, A†(Z) (vu∗ ⊗ A∗A) A†(W )∗x(cid:17)Cn
.
∅, the above becomes:
In the above K A {Z, y, v} is a bounded linear map from J into H R(A), so that the
above inner product is L(J )−valued (we have omitted vectors to simplify the notation).
Applying the definition of a2
= DK A {Z, y, v}, K A {W, x, u}E −(cid:16)y, (vu∗ − A†(Z)vu∗A†(W )∗)x(cid:17)Cn
−(cid:16)y, (A†(Z) − A†(0n))vu∗A†(W )∗x(cid:17)Cn −(cid:16)y, A†(Z)vu∗(A†(W )∗ − A†(0m)∗)x(cid:17)Cn
+(cid:16)y, A†(Z)(cid:0)vu∗ ⊗ (I − A(0)∗A(0) − a2
∅)(cid:1) A†(W )∗x(cid:17)Cn
= DK A {Z, y, v}, K A {W, x, u}E −(cid:16)y, A†(Z)(vu∗ ⊗ a2
− (y, vu∗x)Cn +(cid:16)y, A(0n)vu∗A†(W )∗x(cid:17)Cn
+(cid:16)y, A†(Z)vu∗A(0m)∗x(cid:17)Cn −(cid:16)y, A†(Z)(vu∗ ⊗ A(0)∗A(0))A†(W )∗x(cid:17)Cn
∅)A†(W )∗x(cid:17)Cn
(6.4)
(6.3)
.
24
MICHAEL T. JURY AND ROBERT T.W. MARTIN
On the other hand, one can calculate that (up to a change of sign) line (6.3) + line (6.4)
in the above are equal to:
0 (K A
DK A
0 )∗K A {Z, y, v}, K A {W, x, u}E ,
and it follows that
(6.5)
DX ∗XK A {Z, y, v}, K A {W, x, u}E =D(I − K A
−(cid:16)y, A†(Z)(vu∗ ⊗ a2
∅)A†(W )∗x(cid:17)Cn
GA(Z, W )[P ] := A†(Z)[·] ⊗ a2
If we define the L(J )−valued CPNC kernel:
.
∅A†(W )∗,
0 (K A
0 )∗)K A {Z, y, v}, K A {W, x, u}E
then Equation (6.5) implies that GA ≤ K A as CPNC kernels so that, by Lemma 6.6,
(where (Aa∅)(Z) = A†(Z)a∅). Moreover, Equation (6.5) further implies that
Aa∅ : H → H R(A),
(6.6)
I − X ∗X = K A
0 (K A
0 )∗ + Aa∅(Aa∅)∗.
This is the appropriate analogue of the formula from [12, Proposition 3.2]. To complete
the proof, we apply the transfer function theory of [2]. We define:
U :=
X
0 )∗
(K A
−(Aa∅)∗
H (cid:19) →(cid:18)H R(A) ⊗ Cd
J ⊗ C2 (cid:19) .
A
a∅
A(0)
:(cid:18)H R(A)
a(cid:19) ∈ Rd(H,J ⊗ C2),
(cid:18)A
The top 2 × 2 block of U is the canonical deBranges-Rovnyak colligation with transfer
function equal to A. It follows that if the above U is contractive, then its transfer function
will have the form:
for some non-zero a ∈ Rd(H, J ). We will prove that, in fact, U is an isometry:
(6.7)
0 )∗ + Aa∅(Aa∅)∗ X ∗A + K A
0 (K A
∗
∅(cid:19) .
0 A(0) − Aa2
∅
A∗A + A(0)∗A(0) + a2
U ∗U =(cid:18)X ∗X + K A
By previous formulas the diagonal entries are equal to IH R(A) and IH, respectively, and it
remains to show that the top right (and hence also the bottom left) component vanishes.
This can be verified as follows:
DK A {Z, y, v} f , X ∗AhE =DK A{Z, Z∗y, v}f − A(cid:16)v ⊗ IH, A†(Z)∗y ⊗ f(cid:17), AhE
= (cid:16)y ⊗ f, (A†(Z) − A(0n))v ⊗ h(cid:17) −(cid:16)(v ⊗ IH, A†(Z)∗y ⊗ f ), (I − A(0)∗A(0) − a2
∅)h(cid:17)
= (cid:16)y ⊗ f,(cid:16)A†(Z) − A(0n) − A†(Z) + A†(Z) ⊗ (A(0)∗A(0) + a2
∅)(cid:17) v ⊗ h(cid:17) .
On the other hand,
(6.8)
DK A {Z, y, v} f , K A
∅hE
0 A(0)h − Aa2
= (y ⊗ f, v ⊗ A(0)h) −(cid:16)y ⊗ f, A†(Z)v ⊗ A(0)∗A(0)h(cid:17) −(cid:16)y ⊗ f, A†(Z)v ⊗ a2
∅h(cid:17) .
Adding these expressions together gives 0, which proves the off-diagonal component van-
ishes.
(cid:3)
Corollary 6.8. If B ∈ Ld(H, J ) is column extreme, then it is an extreme point.
COLUMN EXTREME MULTIPLIERS OF THE FREE HARDY SPACE
25
Proof. This is the same contrapositive proof as in [12, Corollary 1.2]: If B is not extreme
then there is a non-zero A ∈ Ld(H,J ) so that both B ± A are Schur class, which implies:
(M L
B−A)∗M L
B−A ≤ I,
and (M L
B+A)∗M L
B+A ≤ I.
Averaging these inequalities gives:
(M L
A )∗M L
A + (M L
B )∗M L
B ≤ I, i.e., (cid:18)B
A(cid:19) ∈ Ld(H,J ⊗ C2),
so that B is not column-extreme.
(cid:3)
The next two corollaries were established in the proof of (2) ⇒ (1) of Theorem 6.4
above:
Corollary 6.9. Given A ∈ Rd(H, J ), define a∅ ∈ L(H)+ by the formula:
∅ := I − A(0)∗A(0) − A∗A ≥ 0.
a2
Then for any h ∈ H, Aa∅h ∈ H R(A), so that Aa∅ : H → H R(A).
Given A ∈ Rd(H, J ), define DA ⊆ H as the linear space of all h ∈ H such that Ah ∈
H R(A), and let A : DA → H R(A) be the linear transformation Ah := Ah ∈ H R(A).
Here, we write A = A(R) so that (Ah)(Z) = A†(Z)h. The previous Corollary 6.9 shows
that Ran (a∅) ⊆ Dom( A).
Corollary 6.10. Given A ∈ Rd(H,J ), we have the identity:
I − X ∗X = K A
0 (K A
0 )∗ + ( Aa∅)( Aa∅)∗.
Claim 6.11. The linear transformation A is closed.
Proof. Suppose (hn) ⊂ DA, hn → h ∈ H, and Ahn → F ∈ H R(A). It is easy to see that
for any Z ∈ Bd
n,
F (Z) = lim
n
A†(Z)hn = A†(Z)h,
so that F = Ah, proving that h ∈ DA and Ah = F .
(cid:3)
Since A is a closed linear transformation, it follows by general facts that A∗ A is densely-
defined in the Hilbert space DA and positive semi-definite on a domain Dom( A∗ A) ⊆ DA
(which is a dense in DA).
Proposition 6.12. Given any h ∈ Dom( A∗ A),
∅(I + A∗ A)h = h,
a2
and Dom( A) = DA = Ran (a∅). Viewing A as a closed linear transformation from DA ⊆
DA → H,
∅ = (IDA + A∗ A)−1.
a2
This is a free and operator-valued analogue of [12, Lemma 3.3].
Proof. Given h ∈ Dom( A∗ A) ⊆ DA, we calculate (I − X ∗X)Ah ∈ H R(A) in two different
ways: First, since Ah ∈ H R(A), XAh = Ah, and
DK A {Z, y, v} g, (I − X ∗X)AhE =(cid:16)y ⊗ g, A†(Z)v ⊗ h(cid:17)
−DK A{Z, Z∗y, v}g, AhE +DA(cid:16)v ⊗ IH, A†(Z)∗y ⊗ g(cid:17), AhE
= (y ⊗ g, A(0n)v ⊗ h) +(cid:16)y ⊗ g, A†(Z)v ⊗ A∗Ah(cid:17) .
((I − X ∗X)Ah) (Z) = A(0)(In ⊗ h) + A†(Z)(In ⊗ A∗Ah).
Equivalently, for any Z ∈ Bd
n,
26
MICHAEL T. JURY AND ROBERT T.W. MARTIN
Secondly, apply the identity (6.6) to obtain:
(cid:16)K A
0 (K A
0 )∗Ah + Aa∅(Aa∅)∗Ah(cid:17) (Z)
Equating these two expressions yields:
∅
A∗ Ah)
= K A(Z, 0)A(0)h + A†(Z)(In ⊗ a2
= A(0n)(I ⊗ h)A†(Z)(cid:16)In ⊗ a2
A†(Z)(cid:16)In ⊗ a2
A∗ Ah − In ⊗ A(0)∗A(0)h(cid:17) .
A∗ Ah − In ⊗ A(0)∗A(0)h(cid:17) = A†(Z)(In ⊗ A∗Ah).
∅
∅
Using that A∗A = I − A(0)∗A(0) − a2
∅ yields:
a2
∅
A∗ Ah − A(0)∗A(0)h = (I − A(0)∗A(0) − a2
∅)h,
and solving for h gives:
∅(I + A∗ A)h = h.
a2
On the other hand, Corollary 6.9 shows that Ran (a∅) ⊆ Dom( A) = DA. For any g ∈
Dom( A∗ A) and any h ∈ H,
This can be re-arranged as:
= (cid:10)g, a2
= (cid:10)g, a2
∅hE =(cid:10)g, (I − a2
hg, hi = D(I + A∗ A)g, a2
∅hE
∅h(cid:11) +D A∗ Ag, a2
∅hE
∅h(cid:11) +D Ag, Aa∅hE .
∅)h(cid:11) ,
∅h ∈ Dom( A∗), and that
∅)h,
D Ag, Aa2
for any g ∈ Dom( A∗ A) and h ∈ H. Since A is a closed linear transformation, Dom( A∗ A)
is a core for A, and the above then implies that Aa2
∅h = (I − a2
A∗ Aa2
which is equivalent to
(I + A∗ A)a2
∅h = h,
and we conclude that a2
∅ = (I + A∗ A)−1. This proves that Ran(cid:0)a2
∅(cid:1) = Dom( A∗ A),
and by the polar decomposition for closed operators, Ran (a∅) = Dom(pI + A∗ A) =
Dom(p A∗ A) = Dom( A).
Corollary 6.13. Given A ∈ Rd(H,J ), and F ∈ H R(A), we have that Lj F ∈ H R(A) if
j F ∈ Dom( A).
and only if A∗
(cid:3)
Proof. The proof is formally identical to that of the commutative analogue of this result
in [9, Corollary 4.5]. Consider, for g ∈ J ,
DK A {Z, y, v} g, X ∗
j FE = DK A{Z, Z∗
j y, v}g, FE −DAj(cid:16)v ⊗ IH, A†(Z)∗y ⊗ g(cid:17), FE
Now A∗
j F ∈ H R(A)
so that also Zj F (Z) ∈ H R(A). Conversely if Zj F (Z) ∈ H R(A), then the above formula
shows that AA∗
= (y ⊗ g, Zj F (Z)v) −(cid:16)y ⊗ g, A†(Z)v ⊗ A∗
j F(cid:17) .
j F ∈ Ran (a∅), then AA∗
j F ∈ H, so by the previous theorem, if A∗
(cid:3)
j F ∈ H R(A).
Corollary 6.14. Given A ∈ Rd(H, J ), if Ah ∈ H R(A) for every h ∈ H, then H R(A)
is L-invariant.
COLUMN EXTREME MULTIPLIERS OF THE FREE HARDY SPACE
27
Corollary 6.15. Given any A ∈ Rd(H, J ), we have that H R(A) is L−invariant if and
only if Ran (A∗) ⊆ Ran (a∅) = Dom( A). This happens if and only if A is densely-defined
and there is a 0 < r < 1 such that
or, equivalently, there is a 0 < ρ < 1 so that
r(I − A(0)∗A(0)) ≤ (I + A∗ A)−1,
A∗A ≤ ρ(I − A(0)∗A(0)).
Proof. By Corollary 6.13, H R(A) is L−invariant if and only if Ran (A∗) ⊆ Ran (a∅). By
the Douglas Factorization Lemma, this happens if and only if there is a λ2 > 0 so that
Re-arranging gives
A∗A∗ ≤ λ2a2
∅ = λ2(I − A∗A − A(0)∗A(0)).
A∗A ≤
λ2
1 + λ2 (I − A(0)∗A(0)).
If Dom( A) = Ran (a∅) is not dense, then since Ran (A∗) = Ran (A∗A), we have that there
is a non-zero h ∈ H so that a∅h = 0 (recall that a∅ ≥ 0 so that Ran (a∅)⊥ = Ker (a∅)). It
then follows that
and one cannot have A∗A ≤ λ2a2
dense range. In this case,
∅ in this case. Hence A∗A ≤ λ2a2
∅ implies Dom( A) has
A∗Ah = (I − A(0)∗A(0) − a2
∅)h 6= 0,
I − a2
∅ − A(0)∗A(0) ≤ λ2a2
∅
⇒ I − (I + A∗ A)−1 − A(0)∗A(0) ≤ λ2(I + A∗ A)−1
⇒
1 + λ2 (I − A(0)∗A(0)) ≤ (I + A∗ A)−1.
1
(cid:3)
6.16. Clark Intertwining. Fix A ∈ Rd(H). As in [8, Theorem 6.3], one can verify that
the weighted Cauchy transform F R : F 2(µA) → H R(A) intertwines the adjoint of the
Stinespring-GNS row isometry ΠA = πA(L) with a perturbation of the restricted backward
left free shift, L∗ ⊗ IHH R(A) = X (and this is a rank-one perturbation in the case where
+ (HA) onto H R(A)
H = C). Equivalently, the unitary multiplier UA := M R
(I−A†(Z)) of H R
intertwines the adjoint of the isometry V A with a perturbation of X:
0 )∗ = F RΠ∗
L∗ ⊗ IHH R(A) + A(I − A(0))−1(K A
A(F R)∗
= UA(V A)∗U ∗
A.
Any U ∈ L(H) yields a different free Clark functional µAU ∗ . Since H R(AU ∗) = H R(A),
it follows that every U ∈ L(H) gives a different perturbation of the restricted backward
left free shift. In particular, if A is column extreme, each of these perturbations will be a
Cuntz unitary (an onto row isometry). In the classical (d = 1, H = C) case, one recovers
Clark's perturbations of the backward shift.
References
[1] J. Agler and J.E. McCarthy. Global holomorphic functions in several non-commuting vari-
ables. arXiv:1305.1636, 2013.
[2] Joseph A. Ball, Vladimir Bolotnikov, and Quanlei Fang. Schur-class multipliers on the Fock
space: de Branges-Rovnyak reproducing kernel spaces and transfer-function realizations. In
Operator theory, structured matrices, and dilations, volume 7 of Theta Ser. Adv. Math.,
pages 85 -- 114. Theta, Bucharest, 2007.
[3] Joseph A. Ball, Gregory Marx, and Victor Vinnikov. Noncommutative reproducing kernel
Hilbert spaces. J. Funct. Anal., 271(7):1844 -- 1920, 2016.
28
MICHAEL T. JURY AND ROBERT T.W. MARTIN
[4] Joseph A. Ball and Victor Vinnikov. Formal reproducing kernel Hilbert spaces: the commu-
tative and noncommutative settings. In Reproducing kernel spaces and applications, volume
143 of Oper. Theory Adv. Appl., pages 77 -- 134. Birkhauser, Basel, 2003.
[5] Douglas N. Clark. One dimensional perturbations of restricted shifts. J. Analyse Math.,
25:169 -- 191, 1972.
[6] Kenneth R. Davidson, Jiankui Li, and David R. Pitts. Absolutely continuous representations
and a Kaplansky density theorem for free semigroup algebras. J. Funct. Anal., 224(1):160 --
191, 2005.
[7] Kenneth Hoffman. Banach spaces of analytic functions. Dover Publications, Inc., New York,
1988. Reprint of the 1962 original.
[8] M. T. Jury and R. T. W. Martin. Non-commutative Clark measures for the free and abelian
Toeplitz algebras. J. Math. Anal. Appl., 456(2):1062 -- 1100, 2017.
[9] Michael T. Jury. Clark theory in the Drury-Arveson space. J. Funct. Anal., 266(6):3855 -- 3893,
2014.
[10] M.T. Jury and R.T.W. Martin. The free Smirnov classes. 2018. In Preparation.
[11] M.T. Jury and R.T.W. Martin. Aleksandrov-Clark theory for Drury-Arveson space. Integral
Equations and Operator Theory, to appear. arXiv:1608.04325.
[12] M.T. Jury and R.T.W. Martin. Extremal multipliers of the Drury-Arveson space. Proceedings
of the American Mathematical Society, to appear. arXiv:1608.04327.
[13] Dmitry S. Kaliuzhnyi-Verbovetskyi and Victor Vinnikov. Foundations of free noncommuta-
tive function theory, volume 199 of Mathematical Surveys and Monographs. American Math-
ematical Society, Providence, RI, 2014.
[14] R. T. W. Martin. Representation of simple symmetric operators with deficiency indices (1, 1)
in de Branges space. Complex Anal. Oper. Theory, 5(2):545 -- 577, 2011.
[15] R.T.W. Martin and A. Ramanantoanina. A Gleason solution model for row contractions.
arXiv:1612.07972, 2016.
[16] Vern I. Paulsen and Mrinal Raghupathi. An introduction to the theory of reproducing ker-
nel Hilbert spaces, volume 152 of Cambridge Studies in Advanced Mathematics. Cambridge
University Press, Cambridge, 2016.
[17] Gelu Popescu. Entropy and multivariable interpolation. Mem. Amer. Math. Soc.,
184(868):vi+83, 2006.
[18] Gelu Popescu. Free holomorphic functions on the unit ball of B(H )n. J. Funct. Anal.,
241(1):268 -- 333, 2006.
[19] Gelu Popescu. Noncommutative transforms and free pluriharmonic functions. Adv. Math.,
220(3):831 -- 893, 2009.
[20] Gelu Popescu. Free holomorphic functions on the unit ball of B(H )n. II. J. Funct. Anal.,
258(5):1513 -- 1578, 2010.
[21] Donald Sarason. Sub-Hardy Hilbert spaces in the unit disk, volume 10 of University of
Arkansas Lecture Notes in the Mathematical Sciences. John Wiley & Sons, Inc., New York,
1994. A Wiley-Interscience Publication.
[22] Donald Sarason. Unbounded operators commuting with restricted backward shifts. Oper.
Matrices, 2(4):583 -- 601, 2008.
[23] O. Shalit. Operator theory and function theory in Drury -- Arveson space and its quotients. In
Handbook of Operator Theory, pages 1125 -- 1180. Springer, 2015.
[24] Daniel Su´arez. Closed commutants of the backward shift operator. Pacific J. Math.,
179(2):371 -- 396, 1997.
University of Florida
E-mail address: [email protected]
University of Cape Town
E-mail address: [email protected]
|
1312.5195 | 2 | 1312 | 2015-03-30T02:36:20 | Strong pure infiniteness of crossed products | [
"math.OA",
"math.DS"
] | Consider an exact action of discrete group $G$ on a separable $C^*$-algebra $A$. It is shown that the reduced crossed product $A\rtimes_{\sigma, \lambda} G$ is strongly purely infinite - provided that the action of $G$ on any quotient $A/I$ by a $G$-invariant closed ideal $I\neq A$ is element-wise properly outer and that the action of $G$ on $A$ is $G$-separating (cf. Definition 4.1). This is the first non-trivial sufficient criterion for strong pure infiniteness of reduced crossed products of $C^*$-algebras $A$ that are not $G$-simple. In the case $A=\mathrm{C}_0(X)$ the notion of a $G$-separating action corresponds to the property that two compact sets $C_1$ and $C_2$, that are contained in open subsets $C_j\subseteq U_j \subseteq X$, can be mapped by elements of $g_j\in G$ onto disjoint sets $\sigma_{g_j}(C_j)\subseteq U_j$, but we do not require that $\sigma_{g_j}(U_j)\subseteq U_j$. A generalization of strong boundary actions on compact spaces to non-unital and non-commutative $C^*$-algebras $A$ (cf. Definition 6.1) is also introduced. It is stronger than the notion of $G$-separating actions by Proposition 6.6, because $G$-separation does not imply $G$-simplicity and there are examples of $G$-separating actions with reduced crossed products that are stably projection-less and non-simple. | math.OA | math | STRONG PURE INFINITENESS OF CROSSED PRODUCTS
E. KIRCHBERG AND A. SIERAKOWSKI
Abstract. Consider an exact action of discrete group G on a separable C *-algebra
A. It is shown that the reduced crossed product A ⋊σ,λ G is strongly purely infinite
-- provided that the action of G on any quotient A/I by a G-invariant closed ideal
I 6= A is element-wise properly outer and that the action of G on A is G-separating
(cf. Definition 4.1). This is the first non-trivial sufficient criterion for strong pure
infiniteness of reduced crossed products of C *-algebras A that are not G-simple. In
the case A = C0(X) the notion of a G-separating action corresponds to the property
that two compact sets C1 and C2, that are contained in open subsets Cj ⊆ Uj ⊆ X,
can be mapped by elements of gj ∈ G onto disjoint sets σgj (Cj ) ⊆ Uj, but we do not
require that σgj (Uj) ⊆ Uj. A generalization of strong boundary actions [18] on com-
pact spaces to non-unital and non-commutative C *-algebras A (cf. Definition 6.1) is
also introduced. It is stronger than the notion of G-separating actions by Proposi-
tion 6.6, because G-separation does not imply G-simplicity and there are examples
of G-separating actions with reduced crossed products that are stably projection-less
and non-simple.
5
1
0
2
r
a
M
0
3
]
.
A
O
h
t
a
m
[
2
v
5
9
1
5
.
2
1
3
1
:
v
i
X
r
a
Contents
1.
Introduction
2. Preliminaries
3. Proper outerness and ideal structure
4. Strongly purely infinite crossed products
5. The case of commutative C *-algebras
6. Strong boundary actions versus G-separating actions
Acknowledgments
Appendix A.
Date: October 11, 2018.
2010 Mathematics Subject Classification. Primary: 46L35; Secondary: 19K99, 46L80, 46L55.
1
2
4
5
11
14
17
27
27
References
29
1. Introduction
In this paper we pursue the study of C *-dynamical systems with applications in
classification via equivariant KK-theory. It was shown by the first named author that
for any two separable nuclear strongly purely infinite C *-algebras, both with primitive
ideal space isomorphic to a T0-space X, the algebras are isomorphic if and only if
they are KKX -equivalent. It is however far from understood when C *-algebra crossed
products A ⋊σ,λ G associated to C *-dynamical systems are strongly purely infinite in
terms of properties of the action σ, in particular in the non-simple case. Our main
focus of this work is such characterisation for crossed products that are either simple
or more generally contain ideals coming from arbitrary G-invariant ideals of the algebra
A on which the group G acts.
We begin (in Section 2) by introducing crossed products and by proving the notation
used throughout the paper.
In Section 3 we look at results related to the ideal structure of crossed products. It
was shown in [23] that residually properly outer (Definition 4.1) and exact (Definition
3.5) actions σ : G → Aut(A) on a separable C *-algebra A have the property that the
lattice of (closed) ideals of the reduced crossed product A⋊σ,λ G is naturally isomorphic
to the lattice of G-invariant ideals of A (by the map I 7→ A ∩ I). We refine this result
by showing that for any exact and residually properly outer action σ of a discrete
group G on a separable or commutative C *-algebra A the set A+ is a filling family
(Definition A.3) for A ⋊σ,λ G (which implies that I 7→ A ∩ I is injective, see Remark
A.5 for details).
In Section 4 we introduce the notion of G-separating actions (Definition 4.1). We
show in Theorem 4.3 that for any exact and residually properly outer action σ of a
discrete group G on a separable or commutative C *-algebra A and for any filling family
F ⊆ A+, the crossed product A ⋊σ,λ G is strongly purely infinite if and only if F has
the diagonalization property (Definition A.9) in A ⋊σ,λ G. Applying the work [15] we
obtain (in Proposition 4.5) an equivalent characterisation of G-separating actions, from
which we can deduce that A+ has the diagonalization property whenever the action on
A is G-separating. By evoking [15] once again we prove our main result:
2
Theorem 1.1. Suppose that (A, G, σ) is a C*-dynamical system, where G is discrete
and A is separable or commutative.
If the action σ of G on A is exact (Def. 3.5), residually properly outer (Def. 3.1)
and G-separating (Def. 4.1), then A ⋊σ,λ G is strongly purely infinite.
In Section 5 we look at actions on commutative C *-algebras. Here we characterise
the notion of G-separating action purely in terms of the underlying geometry. More
specifically we consider actions α of a discrete group G on a locally compact Hausdorff
space X, and denote by σ the induced action on A := C0(X). We show (in Lemma
5.1) that the action of G on A is G-separating if and only if the following holds: For
every open U1, U2 ⊆ X and compact K1, K2 ⊆ X with K1 ⊆ U1, K2 ⊆ U2, there exist
g1, g2 ∈ G such that αg1(K1) ⊆ U1, αg2(K2) ⊆ U2, αg1(K1) ∩ αg2(K2) = ∅. This result
is what motives the choice of our terminology "G-separating". As a consequence of
this characterisation we obtain in Corollary 5.2 a characterisation of when a crossed
product C0(X) ⋊σ,λ G is a strongly purely infinite C *-algebra in terms of condition on
α.
In the final Section 6 we consider actions that produce simple and strongly purely infi-
nite crossed products. We introduce (Definition 6.2 and 6.1) the notion of n-majorizing
(n ≥ 1) and n-covering actions (n ≥ 2), the later for actions on unital C *-algebras.
These two notions aim to refine results on simple purely infinite crossed products in
[18, 11] where the notion of strong boundary actions (Definition A.1) and n-fillings ac-
tions (Definition A.2) was introduced. We prove in Remark 6.7 our notions are weaker:
Any n-filling action on a unital C *-algebra A is n-covering, and for any action α on
a compact spaces X with more than two points (on which strong boundary actions
are defined) the action α is a strong boundary action if and only if its adjoint action
σ on C(X) is 1-majorizing. Both our notions are G-simple. Therefore we call the
1-majorizing actions on not-necessarily unital or commutative C*-algebras also strong
boundary actions. Despite our weaker assumptions we are able to prove:
Theorem 1.2. Suppose that the C*-dynamical system (A, G, σ) with discrete G is n-
majorizing (Def. 6.1) for some n ≥ 1 or n-covering (Def. 6.2) for some n ≥ 2, the
latter if A is unital. If the action σ is element-wise properly outer (Def. 3.1), and A
is separable or commutative, then A ⋊σ,λ G is simple strongly purely infinite.
In Section 6 we also look at how the different properties relate to each other. In
Lemma 6.3 we show that each n-covering action (for n ≥ 2) on a unital C *-algebra
3
A action is n-majorizings, and the latter properly (for n ≥ 1) implies that the action
is (n + 1)-covering. In Proposition 6.6 we prove that that any 1-majorizing action on
a non-unital C *-algebra A is automatically G-separating.
In Remark 6.7 we prove
that any action on a unital commutative C *-algebra A is n-filling if and only if it is n-
covering, and for n = 2 this is again equivalent to a strong boundary (i.e., 1-majorizing)
action.
We end with a number of remarks, including a proof of the fact that our notions
of G-separating, n-majorizing and n-covering actions can be expressed in terms of
projections when A has real rank zero (see Remark 6.9).
We hope that the study of crossed products -- even those for actions of amenable
discrete groups on locally compact Polish spaces -- can help to detect possible differences
between strong and weak pure infiniteness. This paper is a very first step in this
direction, and gives a sufficient criterium by conditions on the action that implies
strong pure infiniteness of reduced crossed products.
2. Preliminaries
We let A+ denote the set of positive elements in a C *-algebra A. We denote the
positive and the negative part of a selfadjoint element a ∈ A by a+ := (a + a)/2 ∈ A+
and a− := (a − a)/2) ∈ A+ , where a := (a∗a)1/2. If a ∈ A+, then (a − ε)+, the
positive part of a − ε1 in M(A), is again in A+. Here M(A) is the multiplier algebra
of A. This notation will be used also for functions f : R → R, then e.g. (f − ε)+(ξ) =
max(f (ξ) − ε , 0) . A subset F ⊆ A+ is invariant under ε-cut-downs if for each a ∈ F
and ε ∈ (0,kak) we have (a − ε)+ ∈ F . The minimal unitalisation of A is denoted A.
Restriction of a map f to X is denoted fX. We let Cc(0,∞]+ denote the set of all
non-negative continuous functions ϕ on [0,∞) with ϕ[0, η] = 0 for some η ∈ (0,∞),
such that limt→∞ ϕ(t) exists.
Remarks 2.1. (i) Suppose that a, b ∈ A+ and ε > 0 satisfy k a − bk < ε . Then the
positive part (b−ε)+ ∈ A of (b−ε·1) ∈ M(A) can be decomposed into d∗ad = (b−ε)+
with some contraction d ∈ A ([14, lem. 2.2]).
(ii) Let τ ∈ [0,∞) and 0 ≤ b ≤ a + τ · 1 (in M(A)), then for every ε > τ there is a
contraction f ∈ A such that (b − ε)+ = f ∗a+f . (See [14, lem. 2.2] and [3, sec. 2.7].)
We abbreviate C *-dynamical systems by (A, G, σ) with discrete groups G. We de-
note by e the unit of G, and consider only closed and two-sided ideals of A. The reduced
4
(resp. the full) crossed product associated to (A, G, σ) is denoted by A ⋊σ,λ G (resp.
A ⋊σ G). The norm on A ⋊σ,λ G will be sometimes written as k · kλ if it is necessary to
distinguish it from other norms. Let I(A) denote the lattice of ideals in a C *-algebra
A. The map η : A → A ⋊σ G means the natural embedding into the full crossed prod-
uct. Let πλ : A ⋊σ G → A ⋊σ,λ G be the natural epimorphism. We will sometimes
suppress the canonical inclusion maps η : A → A ⋊σ G and πλ ◦ η : A ⊆ A ⋊σ,λ G.
Let U denote the canonical unitary representation U : G → M(A ⋊σ G). Notice here
that the linear span of η(A)U(G) is is a dense *-subalgebra of A ⋊σ G. We denote by
Uλ : G → M(A ⋊σ,λ G) the regular representation for some more precise explanations.
The same happens with ηλ := πλ ◦ η.
The set Cc(G, A) consists of the maps f : G → A with finite support F := G\f −1(0).
There is a natural linear embedding of Cc(G, A) into A ⋊σ G by canonical identification
of f : G → A (of finite support) with an element of A⋊σG : Let F ⊆ G be a finite subset,
with f (g) = 0 for g 6∈ F . Then f will be identified with the element Pg∈F η(ag)U(g)
of A ⋊σ G , where ag := f (g). In this way Cc(G, A) becomes a *-subalgebra of A ⋊σ G
that contains A. The natural C *-morphism πλ : A ⋊σ G → A ⋊σ,λ G is faithful on
Cc(G, A), and we do not distinguish between
πλ(X
η(ag)U(g)) = X
g∈F
g∈F
ηλ(ag)Uλ(g)
and Pg∈F η(ag)U(g). In particular, η(a) ∈ A⋊σ G and ηλ(a) ∈ A⋊σ,λ G will be denoted
simply by a ∈ A, and Uλ(g) might be denoted U(g).
We now recall the conditional expectation E : A ⋊σ G → η(A) ∼= A : The map
Ealg : Cc(G, A) → A , Pg∈F agU(g) 7→ ae , extends by continuity to a faithful con-
ditional expectation Eλ : A ⋊σ,λ G → A.
In particular Eλ is a completely positive
contraction, Eλ(A ⋊σ,λ G) = A, and Eλ(b) = 0 imply b = 0 for b ∈ (A ⋊σ,λ G)+.
Since A is also contained in its full crossed product A ⋊σ G, we can use the natural
epimorphism A ⋊σ G → A ⋊σ,λ G to define E by E := Eλ ◦ πλ as a (not necessarily
faithful) conditional expectation from A ⋊σ G onto its C *-subalgebra A.
3. Proper outerness and ideal structure
In this section we look at conditions on a C *-dynamical system (A, G, σ) ensuring
that the set A+ is a filling family for A ⋊σ,λ G in the sense of Definition A.3. This in
particular implies that the is a on-to-one correspondence between ideals of A⋊σ,λ G and
G-invariant ideals of A, but (as wee shall see) also applies to the verification of when
5
a crossed product is strongly purely infinite. Proper outerness of the automorphisms
σt of A defining the action σ turns out also to be a crucial ingredient. We recall the
definition below.
Definition 3.1. Suppose that (A, G, σ) is a C *-dynamical system and that G is
discrete. The action σ will be called element-wise properly outer if, for each g ∈
G\{e}, the automorphism σg of A is properly outer in the sense of [6, def. 2.1], i.e.,
k σgI − Ad(U)k = 2 for any σg-invariant non-zero ideal I of A and any unitary U in
the multiplier algebra M(I) of I. See also [19, thm. 6.6(ii)].
We call here an action σ residually properly outer if, for every G-invariant ideal
J 6= A of A, the induced action [σ]J of G on A/J is element-wise properly outer.
Remarks 3.2. (i) Notice that element-wise proper outerness passes to subgroups, i.e.,
for each subgroup H ⊆ G the system (A, H, σH) is element-wise properly outer on
A if (A, G, σ) is element-wise properly outer. But residual proper outerness does not
necessarily pass to subgroups. The system (A, H, σH) is not necessarily residually
properly outer, if (A, G, σ) is residual proper outer, because possibly there could be
more H-invariant ideals than G-invariant ideals of A.
(ii) If A is non-commutative, then topological freeness of (A, G, σ) in sense of [1, def. 1]
is -- at least formally -- stronger than the assumption of element-wise proper outerness
of (A, G, σ) in Definition 3.1 (cf. [1, prop. 1]). We do not know examples where they
actually differ. Thus, for non-commutative A, "essential freeness" of the corresponding
action of G on bA in the sense of [23, def. 1.17] (inspired by [22, def. 4.8]) is -- formally
-- stronger than our residual proper outerness of (A, G, σ).
(iii) If G is countable and acting on C0(X), one can show -- using the Baire property of
X -- that elementwise proper outerness is the same as the requirement (for the action
α of G on X with σg(f ) := f ◦ αg−1) that points with trivial fix-point subgroup (trivial
isotropy) are dense in X, i.e., Definition [23, def. 1.17] holds. We can reformulate this
as: stability subgroups of non-empty open subsets are trivial.
Remark 3.3. We recall [19, lem. 7.1.] (cf. also [17, lem. 3.2]):
If α1, α2, ..., αn are properly outer automorphisms of a separable C*-algebra A, there
is, for each a0, a1, a2, ..., an ∈ eA, with 0 6= a0 ≥ 0, and each ε > 0, an element x ∈ A+
with kxk = 1 such that
kxa0xk > ka0k − ε ,
kxaiαj(x)k < ε ,
6
1 ≤ i, j ≤ n .
If A is commutative, i.e., A ∼= C0(X) for X = bA ⊆ A∗, then it is not necessary
to suppose that A is separable in the quoted lemma of D. Olesen and G. Pedersen
(Compare also [7]): An automorphism σ ∈ Aut(A) is properly outer, if and only if,
for every open subset ∅ 6= U ⊆ X there exists y ∈ U with bσ(y) 6= y. Thus, for every
finite set S ⊆ Aut(A) of properly outer automorphisms, every non-empty open subset
U ⊆ X contains a non-empty open subset V ⊆ U with bσ(V ) ∩ V = ∅ for all σ ∈ S. If
one takes U := a−1
0 (ka0 k−ε,∞) and non-empty V ⊆ U as above, then each x ∈ C0(X)
with kxk = 1 and support in V satisfies kxa0xk > ka0k − ε and xσ(x) = 0 for σ ∈ S.
The following Lemma 3.4 is a suitable modification of proofs of [19, lem.7.1,thm.7.2].
It has been proved in [1] under the stronger assumption that the action σ is topologically
free, and part (iii) has been shown in [12, thm. 4.1] even to be equivalent to the
topological freeness of the action if A is commutative and unital and G is amenable.
Compare also Remark 5.3 for a "residual" version.
Lemma 3.4. Suppose that A is separable or commutative, and that the action of G on
A is element-wise properly outer.
(i) For every b ∈ (A ⋊σ G)+ with E(b) 6= 0 and ε > 0 there exist x ∈ A+ satisfying
that
kxk = 1 , kxbx − xE(b)xk < ε , kxE(b)xk > kE(b)k − ε .
This holds also for b ∈ (A ⋊σ,λ G)+ and Eλ in place of E.
(ii) If h : A ⋊σ G → L(H) is a *-representation such that hA is faithful, then
kh(b)k ≥ kE(b)k for all b ∈ (A ⋊σ G)+.
In particular, the kernel of h is contained in the kernel Iλ of the natural
epimorphism πλ : A ⋊σ G → A ⋊σ,λ G.
(iii) Every non-zero ideal of A ⋊σ,λ G, has non-zero intersection with A.
Proof. (i): Let b ∈ (A ⋊σ G)+ with E(b) 6= 0, and ε > 0.
Let a0 := E(b). Since Cc(G, A) is dense in A ⋊σ G, there exists a′ = c0 +
Pm
j=1 ajU(gj) ∈ Cc(G, A) with g−1
6= e for i 6= j ∈ {1, . . . , m},
i gj
and ka′ − bk < ε/6. Since E is a contraction, it follows that kb − ak < ε/3 and
E(a) = a0 = E(b) for g0 := e and a := a0 + (a′ − c0) = Pm
j=0 ajU(gj). By [19, lem. 7.1]
and Remark 3.3 there exists x ∈ A+ with kxk = 1, kxE(a)xk > kE(a)k − ε/3m, and
kxajσgj (x)k < ε/3m for gj 6= e, j = 1, . . . , m . In particular,
6= e and gj
kxE(b)xk = kxE(a)xk > kE(a)k − ε = kE(b)k − ε .
7
Since kxajU(gj)xk = kxajσgj (x)k we get in A ⋊σ G that
kxajσgj (x)k ≤ ε/3 .
kx(a − E(a))xk ≤ X
gj6=e
Thus, in the full crossed product A ⋊σ G we have
kxbx − xE(b)xk ≤ kx(b − a)xk + kx(a − E(a))xk + kx(E(a) − E(b))xk < ε .
The same arguments work for b ∈ (A ⋊σ,λ G)+ and Eλ in place of E.
(ii): The restriction of h to A ⊆ A ⋊σ G is faithful, hence kh(a)k = kak for all
a ∈ A. Let b ∈ (A ⋊σ G)+ be given. If E(b) = 0 then kh(b)k ≥ kE(b)k. If E(b) 6= 0,
then select x ∈ A+ as in (i). It follows that kh(xE(b)x)k = kxE(b)xk ≥ kE(b)k − ε .
On the other hand, kh(b)k ≥ kh(x)h(b)h(x)k = kh(xbx)k and ε > kxbx − xE(b)xk ≥
kh(xbx) − h(xE(b)x)k. Thus kh(b)k + ε ≥ kh(xE(b)x)k, and kh(b)k + 2ε ≥ kE(b)k for
all ε > 0.
Since E = Eλ ◦ πλ, we have b ∈ (A ⋊σ G)+ and E(b) = 0 implies that b is contained
in the kernel of πλ : A ⋊σ G → A ⋊σ,λ G . In particular, if h : A ⋊σ G → L(H) is any
*-representation with kh(b)k ≥ kE(b)k for all b ∈ (A ⋊σ G)+, then the kernel h−1(0) of
h is contained in the kernel of the natural epimorphism πλ : A ⋊σ G → A ⋊σ,λ G.
(iii): Let I a closed ideal of A ⋊σ,λ G with I∩A = {0}, consider (A ⋊σ,λ G)/I as a C *-
subalgebra of some L(H), and let h : A ⋊σ G → L(H) the corresponding representation
with kernel h−1(0) = J := π−1
λ (0). Then h is faithful on A and, therefore,
satisfies π−1
(cid:3)
λ (0) ⊇ h−1(0). It follows I = πλ(h−1(0)) = {0}.
λ (I) ⊇ π−1
Definition 3.5 ([23, def. 1.2]). Suppose that (A, G, σ) is a C *-dynamical system with
locally compact G. The action σ of G on A is exact, if, for every G-invariant ideal J
in A, the sequence 0 → J ⋊σJ,λ G → A ⋊σ,λ G → A/J ⋊[σ]J ,λ G → 0 is short-exact.
Remarks 3.6. (i) Recall that a locally compact group G is exact, if and only if, every
action σ : G → Aut(A) is exact. If G is discrete, then this is equivalent to the exactness
of the C *-algebra C ∗
λ(G), cf. [16]. This applies to all amenable groups G, e.g. G = Z.
Under Definition 3.5 each minimal (= G-simple) action is exact. In particular, non-
exact discrete groups can have exact (and faithful) actions.
(ii) Let F denote the (small) Thompson group and ρ : F → Homeo(R) the minimal
action of F (or only of its commutator subgroup F ′) on the real line R as described
by Haagerup and Picioroaga in [10, rem. 2.5.]. One can use ρ to construct a F -
separating, non-minimal and exact action α of F (or F ′) on the disjoint union of two
8
lines X := R ∪ (i + R) ⊆ C if one considers the restriction α(g) := β(g)X to X of the
action g ∈ F → β(g) on C given by β(g)(s + it) := ρ(g)(s) + it ( 1 ). It is at present
unknown whether the Thompson group F is exact or not, cf. [2, 8, 9].
(iii) It is not known if Gromov's examples of non-exact groups can have non-exact
actions on commutative C *-algebras. It is likely that it has to do with still missing
non-trivial geometric conditions for G-actions on locally compact spaces X that are
equivalent to the exactness of the adjoint action σ : G → Aut(C0(X)) given by σg(f ) :=
f ◦ αg−1. Therefore we use the trivial and non-geometric definition and define α to be
exact if its adjoint action σ on C0(X) is exact.
Remark 3.7. Combination of Lemma 3.4(iii) and of the exactness of an action σ : G →
Aut(A) on a separable or commutative C *-algebra A shows that the lattice of (closed)
ideals of the reduced crossed product A ⋊σ,λ G is naturally isomorphic to the lattice
of G-invariant ideals of A (by the map J 7→ A ∩ J), if σ is exact and residual properly
outer. (See [23, Remark 2.23] for details.)
Theorem 3.8. Let (A, G, σ) a C*-dynamical system, with discrete G and separable or
commutative A. If the action σ of G on A is exact and residually properly outer, then
the elements of A+ build a filling family for A ⋊σ,λ G in the sense of Definition A.3.
Proof. We show that for every hereditary C *-subalgebra D of A ⋊σ,λ G and every
(closed) ideal I of A ⋊σ,λ G with D 6⊆ I there exist f ∈ A+ and z ∈ A ⋊σ,λ G such that
z∗z ∈ D and zz∗ = f 6∈ I.
Suppose that D is a hereditary C * -- subalgebra of A ⋊σ,λ G and that I is an ideal
of A ⋊σ,λ G with D 6⊆ I. Let J := I ∩ A, then J is a G-invariant ideal of A with
J ⋊σJ,λ G ⊆ I and g ∈ G 7→ [σg]J is an element-wise properly outer action on A/J by
our assumptions on σ. We denote this action by α, i.e., αg(a + J) := σg(a) + J.
By Remark 3.7, the exactness and residual proper outerness of σ : G → Aut(A) allow
a natural identification
(A ⋊σ,λ G)/I = (A/J) ⋊α,λ G .
Since D 6⊆ I implies D+ 6⊆ I, there exists d ∈ D+, d /∈ I. The epimorphism
πI : A ⋊σ,λ G → (A ⋊σ,λ G)/I is equal to the quotient map πJ from A ⋊σ,λ G onto
(A/J)⋊α,λ G ∼= (A⋊σ,λ G)/I (under natural identifications). We denote the conditional
1 This action is not topologically free.
9
expectation Eλ : (A/J) ⋊α,λ G → A/J (temporarily) by E and define
b := πI(d) ,
and
ε :=
1
4 kE(b)k > 0 .
Lemma 3.4(i) gives an element x ∈ (A/J)+ such that
kxk = 1, kxbx − xE(b)xk < ε,
kxE(b)xk > kE(b)k − ε =
By Remark 2.1(i), there is a contraction y ∈ (A/J) ⋊α,λ G such that
3
4 kE(b)k .
y∗xbxy = (xE(b)x − ε)+ ∈ (A/J)+ .
Note that y∗xbxy 6= 0 , because
k(xE(b)x − ε)+k = kxE(b)xk −
1
4 kE(b)k >
1
2 kE(b)k = 2ε .
Since πIA = πJA and (xE(b)x − ε)+ ∈ (A/J)+, there is exists c ∈ A+ such that
πJ(c) = (xE(b)x − ε)+. Since πJ (= πI ) is surjective, there exists a contraction
w ∈ A ⋊σ,λ G with πJ (w) = xy. We obtain that
c = w∗dw + v
for some v ∈ I. The set Cc(G, J) is dense in I, because I = J ⋊σ,λ G and G is discrete.
This allows us to see, that JIJ is dense in I. It follows that {e ∈ J+ ; kek < 1} is an
approximate unit of I. In particular, there exists e ∈ J+ with kv − evk < ε.
Let 1 denote the unit of eA ⋊σ,λ G, then A ⋊σ,λ G is an ideal of eA ⋊σ,λ G. With
g := (1 − e) ∈ eA+, kgk ≤ 1 we get
kgw∗dwg − gcgk = kgvgk ≤ kv − evk < ε =
1
4 kE(b)k .
Since gzg = z + eze − (ze + ez) and πI (e) = πJ (e) = 0, we have πI(gzg) = πI(z) for
all z ∈ A ⋊σ,λ G.
By Remark 2.1(i), there exists a contraction h ∈ A ⋊σ,λ G such that
h∗(gw∗dwg)h = (gcg − ε)+ ∈ A+ .
With z := (d1/2wgh)∗ we have that z∗z ∈ D and zz∗ = (gcg − ε)+ =: f ∈ A+ . Finally,
we see from πI (gcg) = πI(c) that
kπI(f )k = kπI((gcg) − ε)+)k = (kπI(gcg)k − ε)+ = (kπI(c)k − ε)+
= (k(xE(b)x − ε)+k − ε)+ = kxE(b)xk −
Hence, f 6∈ I.
10
1
2kE(b)k >
1
4kE(b)k > 0 .
(cid:3)
4. Strongly purely infinite crossed products
In this section we prove out main result Theorem 1.1. We start with the definition
of an G-separating action.
Definition 4.1. Suppose that (A, G, σ) is a C *-dynamical system with discrete group
G. The action of G on A is G-separating if for every a, b ∈ A+ , c ∈ A, ε > 0, there
exist elements s, t ∈ A and g, h ∈ G such that
k s∗a s − σg(a)k < ε , k t∗b t − σh(b)k < ε and k s∗c tk < ε .
(1)
Remarks 4.2. (i) Notice that Definition A.6 and Remark A.7 immediately implies
that every action σ : G → Aut(A) is G-separating if A itself is strongly purely infinite:
Take h = g = e ∈ G. If the contractions s, t ∈ A satisfy the defining inequalities (7) of
strongly p.i. algebras A then they also satisfy the inequalities (1).
(ii) G-separating actions on a locally compact space X are not necessarily minmal. One
can show that above mentioned example of an exact and non-minimal action of the
(small) Thompson group F on two parallel lines R∪ (i + R) ⊆ C is also F -separating.
(iii) The existence of a G-separating action on A imposes requirement on A itself,
e.g. the cases a = b = c = p and a = b = c = 1 with ε = 1/3 in inequalities (1)
show that A can not contain minimal non-zero projections p ∈ A and that 1A must be
properly infinite in A if A is unital. Therefore, C *-algebras, that are commutative and
unital, can not have any G-separating actions.
(iv) Further variations of the concepts that we introduce here are possible, e.g. one
could start with conditions that are weaker than conditions for G-separating actions.
Also one could require the existence of n ∈ N such that for a, b ∈ A+ and ε > 0 there is
a solution d1, . . . , dn ∈ A and g1, . . . , gn ∈ G of the inequality (3) in Definition 6.1 of n-
majorizing actions whenever b is in the smallest G-invariant closed ideal that contains
a. Or one could attempt to replace the filling family F := A+ by smaller filling families
F ⊆ A+ and require more elaborate local matrix diagonalization formulas involving
also G-translates, cf. Definition A.8.
Combing Theorem 3.8 with Theorem A.13 we obtain the following result
Theorem 4.3. Let G be a discrete group acting by σ : G → Aut(A) on a separable
or commutative C*-algebra A. Suppose that the action is residually properly outer
11
(cf. Def. 3.1) and exact (cf. Def. 3.5). Let F ⊆ A+ be a filling family for A. Then the
following are equivalent:
(i) The crossed product A ⋊σ,λ G is strongly purely infinite.
(ii) The family F has the diagonalization property in A ⋊σ,λ G.
Proof. Let B := A ⋊σ,λ G. The assumptions ensure that A+ is a filling family for B
by Theorem 3.8. Since F is filling for A, F is also filling for B by Lemma A.4.
(i)⇒(ii): If B is s.p.i., then B+ has the diagonalization property (see Definition
A.9) in B, cf. [14, lem. 5.7]. This implies that our family F ⊆ A+ ⊆ B+ has the
diagonalization property in B.
(ii)⇒(i): Since our family F ⊆ A+ is filling for B, and since F has the diagonalization
property in B, we get that B is s.p.i. by Theorem A.13.
(cid:3)
Remark 4.4. Let (A, G, σ) a C *-dynamical system.
(i) For each a1, a2 ∈ A+, x, d1, d2 ∈ A, g0, g1, g2 ∈ G and s1 := d1U(g1), s2 :=
0 g2g2), c := xU(g0), b1 := a1, and b2 := σg0(a2) the following
(d2)U(g−1
σg−1
equalities hold:
0
ks∗
j ajsj − ajk = kd∗
j bjdj − σgj (bj)k
and
ks∗
1cs2k = kd∗
1xd2k .
(ii) With g0 = e in (i) the equalities reduce to:
ks∗
j ajsj − ajk = kd∗
j ajdj − σgj (aj)k
and
ks∗
1cs2k = kd∗
1cd2k .
Proposition 4.5. Suppose that (A, G, σ) is a C*-dynamical system with discrete G.
The following properties (i) -- (ii) are equivalent:
(i) The action of G on A is G-separating in the sense of Definition 4.1.
(ii) For every a1, a2 ∈ A+, c ∈ A ⋊σ,λ G and ε > 0 there exist d1, d2 ∈ A and
g1, g2 ∈ G such that the elements sj = djU(gj) of Cc(G, A) satisfy, for j = 1, 2,
(2)
k s∗
jajsj − aj k < ε , and k s∗
1cs2 k < ε .
Proof. (ii)⇒(i): If we take c ∈ A, a1 := a and a2 := b for a, b ∈ A+ and ε > 0, then
(ii) implies, using Remark 4.4, that there exist d1, d2 ∈ A and g1, g2 ∈ G such that
kd∗
1cd2k < ε, so the inequalities (1) of Definition 4.1 are
satisfied with d1, d2, g1, g2 in place s, t, g, h.
j ajdj − σgj (aj)k < ε and kd∗
12
(i)⇒(ii): Define C := {dU(g) ; d ∈ A , g ∈ G} and S := C . Select any ε0 > 0.
Clearly, the closed linear span of C is equal to A ⋊σ,λ G. If we can show that F := A+ ,
C and S satisfy the assumptions (i)-(iii) of Lemma A.11 -- with A ⋊σ,λ G in place of A
-- , then it follows from Lemma A.11 that for every a1, a2 ∈ A+, c ∈ A ⋊σ,λ G and ε > 0
there exist d1, d2 ∈ A and g1, g2 ∈ G such that sj = djU(gj) ∈ S fulfil (2), which in
turn gives (ii).
It is evident that our C and S satisfy properties (ii) and (iii) in Lemma A.11. Since
A+ is closed under ε-cut-downs, property (i) becomes automatic if each pair (a1, a2),
with a1, a2 ∈ A+, has the matrix diagonalization property of Definition A.8 with respect
to S and C:
If a1, a2 ∈ A+, c = xU(g0) ∈ C with x ∈ A, g0 ∈ G, and ε > 0 are given, then we
define b1 := a1, b2 := σg0(a2). (If we instead of ε are given positive ε1, ε2 and τ , set
ε := min(ε1, ε2, τ ).) Since the action σ is G-separating, we can find d1, d2 ∈ A and
g1, g2 ∈ G with kd∗
1xd2k all strictly below
ε. Remark 4.4 provides elements sj ∈ C satisfying (2). Thus (a1, a2) has the matrix
diagonalization property with respect to S and C.
2b2d2 − σg2(b2)k and kd∗
1b1d1 − σg1(b1)k, kd∗
(cid:3)
Theorem 4.6. Let (A, G, σ) a C*-dynamical system, with discrete G. Suppose that
A+ is a filling family for A ⋊σ,λ G and that the action of G on A is G-separating. Then
A ⋊σ,λ G is strongly purely infinite.
Proof. By Theorem A.13 it remains to show that A+ has the diagonalization property
in A ⋊σ,λ G. Since A+ is closed under ε-cut-downs Lemma A.10 applies, and therefore
it is enough to show that each pair (a1, a2) with a1, a2 ∈ A+, has the matrix diagonal-
ization property in A ⋊σ,λ G. But this follows from the G-separation property of the
action σ by Proposition 4.5.
(cid:3).
Proof of Theorem 1.1: By Theorems 3.8 and 4.6 the assumptions imply that
A ⋊σ,λ G is strongly purely infinite.
(cid:3)
Remark 4.7. Suppose that (A, G, σ) is a C *-dynamical system and that G is discrete.
Then a family F ⊆ A+ ⊆ A ⋊σ,λ G which is invariant under ε-cut-downs has the
diagonalization properly in A ⋊σ,λ G , if and only if, for every a1, a2 ∈ F , c ∈ Cc(G, A)
and > 0 , there exist s1, s2 ∈ Cc(G, A) such that, for j = 1, 2,
k s∗
jajsj − aj k < ε
and
13
k s∗
1cs2 k < ε .
This follows from Lemma A.10, Lemma A.11 and the fact that Cc(G, A) is dense in
A ⋊σ,λ G .
Remark 4.8. Notice that for an exact locally compact group G the reduced group C *-
algebra C ∗
λ(G) is an exact C *-algebra (cf. [16, p. 171]). By Theorem A.12, the minimal
C *-tensor product A ⊗min B of a s.p.i. C *-algebra A with an exact C *-algebra B is
again s.p.i. Hence, if G is an exact locally compact group, σ(g) := idA is the trivial
action on a s.p.i. C *-algebra A then A ⋊σ,λ G ∼= A ⊗min C ∗
λ(G) is s.p.i.
This shows that there is room for refinements of our sufficient conditions on the
actions that imply strong pure infiniteness of the reduced crossed products: Here the
action σ is even not element-wise properly outer, but satisfies the G-separation property
and is an exact action by C *-exactness of C ∗
λ(G).
5. The case of commutative C *-algebras
The case of G-actions on commutative C *-algebras allows some topological interpre-
tation. The next lemma has inspired our choice of the name G-separating in Definition
4.1. Notice that we do not require αgj (Uj) ⊆ Uj in its part (ii).
Lemma 5.1. Suppose that (A, G, σ) is a C*-dynamical system, that A ∼= C0(X) is
commutative, and that the action σ of G on C0(X) is induced by the action α of G on
for f ∈ A, g ∈ G). Then the following properties are
X ∼= bA (i.e., σg(f ) := f ◦ α−1
g
equivalent:
(i) The action of G on A is G-separating, i.e., for every a, b ∈ A+ , c ∈ A, ε > 0,
there exist elements d1, d2 ∈ A and g1, g2 ∈ G such that
k d∗
1ad1 − σg1(a)k < ε , k d∗
2bd2 − σg2(b)k < ε and k d∗
1cd2 k < ε .
(ii) For every open U1, U2 ⊆ X and compact K1, K2 ⊆ X with K1 ⊆ U1, K2 ⊆ U2,
there exist g1, g2 ∈ G such that
αg1(K1) ⊆ U1, αg2(K2) ⊆ U2, αg1(K1) ∩ αg2(K2) = ∅ .
Proof. (ii)⇒(i): Let a, b ∈ A+, c ∈ A and ε > 0. We use assumption (ii) on
U1 := a−1(ε/4,∞) = {x ∈ X ; a(x) > ε/4} , U2 := {x ∈ X ; b(x) > ε/4} ,
K1 := {x ∈ X ; a(x) ≥ ε/2} , K2 := {x ∈ X ; b(x) ≥ ε/2} ,
14
and find g1, g2 ∈ G such that
αg1(K1) ⊆ U1 , αg2(K2) ⊆ U2 , αg1(K1) ∩ αg2(K2) = ∅ .
Since a, b ∈ C0(X)+, we have that U1 ⊆ a−1[ε/4,∞) and U2 ⊆ b−1[ε/4,∞) are compact
subsets of X.
Since the compact sets αg1(K1) and αg2(K2) are disjoint, applications of Tietze
extension theorem gives elements e1, e2 ∈ A+ with kejk ≤ 2/√ε and a contraction
f = f ∗ ∈ A such that
e1U1 = a−1/2U1 , e2U2 = b−1/2U2 , fαg1(K1) = −1 , fαg2(K2) = 1 .
Let f+, f− ∈ A+ be the canonical decomposition f = f+ − f− with f+f− = 0. We
define
d1 := e1(σg1(a) − ε/2)1/2
1cd2 = 0 because f+f− = 0.
Then d∗
+ f− and d2 := e2(σg2(b) − ε/2)1/2
+ f+ .
Since (σg1(a) − ε/2)+(x) 6= 0 implies a(αg−1
(x) ∈ K1, and
x ∈ αg1(K1) ⊆ U1 ⊆ U1. It implies that f−(x) = 1 and e1(x) = a−1/2(x). We obtain
that
(x)) > ε/2, we get αg−1
1
1
1ad1 = e2
d∗
1a(σg1(a) − ε/2)+(f−)2 = (σg1(a) − ε/2)+ .
In a similar way we see that d∗
2bd2 = (σg2(b) − ε/2)+.
(i)⇒(ii): Let U1, U2 ⊆ X be open and K1, K2 ⊆ X compact subsets such that
K1 ⊆ U1 and K2 ⊆ U2. We can assume that the intersection K1 ∩ K2 is non-empty.
There exists an open set W with a compact closure W such that
K1 ∪ K2 ⊆ W ⊆ W ⊆ U1 ∪ U2 .
By the Tietze extension theorem, there are contractions a, b, c ∈ A+ such that
aK1 = 1 , bK2 = 1 , cW = 1 ,
supp(a) ⊆ U1 ∩ W,
supp(b) ⊆ U2 ∩ W,
supp(c) ⊆ U1 ∪ U2 .
We apply assumption (i) to a, b, c and ε := 1/4, and get elements d1, d2 ∈ A and
g1, g2 ∈ G such that
kd∗
1ad1 − σg1(a)k < 1/4 , kd∗
2bd2 − σg2(b)k < 1/4 , kd∗
1cd2k < 1/4 .
If x ∈ W , then c(x) = 1 and d1(x)d2(x) ≤ kd∗
1cd2k < 1/4. Thus,
V1 := {x ∈ W ;
d1(x) ≥ 1/2} , V2 := {x ∈ W ; d2(x) ≥ 1/2}
15
1
1
(x) ∈ K1 and σg1(a)(x) = a(αg−1
are disjoint sets. If x ∈ αg1(K1), then αg−1
(x)) = 1. It
follows d1(x)2a(x) ≥ 3/4 and d1(x) > 1/2 (the latter because 1 ≥ a(x) > 0). Thus,
x ∈ U1∩ W and x ∈ V1. It shows αg1(K1) ⊆ U1∩ V1. In a similar way we get αg2(K2) ⊆
U2 ∩ V2. It implies αg1(K1) ⊆ U1 , αg2(K2) ⊆ U2 and that αg1(K1) ∩ αg2(K2) = ∅ . (cid:3)
The following condition (i) in Corollary 5.2 is satisfied if the action α has the residual
version of the topological freeness in sense of [1, def. 1], see e.g. the essential freeness
defined in [23, def. 1.17] (inspired by [22, def. 4.8]) when G is countable.
Corollary 5.2. Let G be a discrete group, α : G → Homeo(X) an action of G on a
locally compact Hausdorff space X. Suppose that
(i) for every closed G-invariant subset Y of X and every e 6= g ∈ G the set
{y ∈ Y : αg(y) = y} has empty interior,
(ii) the action σ : G → Aut(C0(X)), given by σg(f ) := f ◦ (αg)−1, is exact on the
C*-algebra C0(X), and
(iii) the action α is G-separating, i.e., by Lemma 5.1, for every U1, U2 ⊆ X open
and K1, K2 ⊆ X compact such that K1 ⊆ U1, K2 ⊆ U2, there exist g1, g2 ∈ G
such that
αg1(K1) ⊆ U1, αg2(K2) ⊆ U2, αg1(K1) ∩ αg2(K2) = ∅ .
Then C0(X) ⋊σ,λ G is a strongly purely infinite C*-algebra.
Proof. Let A := C0(X). It is easy to see that property (i) implies that the action on
any quotient A/I by a G-invariant closed ideal I 6= A is element-wise properly outer.
Now Theorem 1.1 applies to A ⋊σ,λ G.
(cid:3)
The following remark shows that in case of commutative A and discrete amenable
G several of the previously considered properties are equivalent.
Remark 5.3. If A is commutative and G is a discrete amenable group that acts on A
by σ, then the following properties are equivalent:
(i) A separates the ideals of A ⋊σ G, i.e., I 7→ A ∩ I is an injective map from
I(A ⋊σ G) into I(A) (see [23, def. 1.9]).
(ii) The action σ : G → Aut(A) is residually properly outer (Definition 3.1).
(iii) The family F := A+ is filling for A ⋊σ G (Definition A.3).
Proof. (i)⇒(ii): By [12, thm. 4.1] (in case of unital A, and [1, thm. 2] for the general
-- non-unital -- case) the separation property implies that the adjoint action of G on the
16
Gelfand spectrum of A is topologically free, which is equivalent to element-wise proper
outerness by [1, prop. 1].
This applies also to the quotients (A/J) ⋊[σ]J ,λ G, because the property (i) passes to
quotients by amenability of G. See also [23, thm. 1.13].
(ii)⇒(iii): Since amenable G are exact, the residual proper outerness of the action
implies that F := A+ is filling for A ⋊σ,λ G by Theorem 3.8.
(iii)⇒(i): By Remark A.5, the subalgebra A separates the ideals of B := A ⋊σ,λ G if
(cid:3)
F := A+ is filling for B.
Asking G to be amenable can be weakened to exactness of σ and A ⋊σ,λ G ∼= A ⋊σ G.
One might also expect nuclearity of A ⋊σ,λ G would suffice in place of amenability of
G (this is know at least in the unital case).
6. Strong boundary actions versus G-separating actions
In this section we prove our Theorem 1.2. We state with the definition of n-
majorizing and n-covering actions.
Definition 6.1. Let n ∈ N and A a non-zero C *-algebra, that is not isomorphic to
a subalgebra of Mn+1(C) if A is unital. An action σ : G → Aut(A) will be called an
n-majorizing action of G on A, if, for every non-zero a ∈ A+, every non-invertible
b ∈ A+ and every ε > 0, there exist d1, . . . , dn ∈ A and g1, . . . , gn ∈ G such that
nX
j=1
k
d∗
j σgj (a) dj − bk < ε .
(3)
Definition 6.2. Let n ∈ N, n ≥ 2 ( 2 ). Suppose that A is a unital C *-algebra, that is
not isomorphic to a *-subalgebra of Mn(C). An action σ of G on A is an n-covering
action if, for every non-zero a ∈ A+, and every ε > 0, there exist d1, . . . , dn ∈ A and
g1, g2, . . . , gn ∈ G and such that
nX
(4)
d∗
j σgj (a) dj − 1 k < ε .
k
j=1
The following lemma denies the existence of non-zero "socles" in C *-algebras A that
admit n-majorizing or n-covering actions considered in Definitions 6.1 and 6.2.
2 If n = 1 then property (4) holds if and only if A is a unital simple purely infinite C *-algebra.
17
Lemma 6.3. Let (A, G, σ) C*-dynamical system.
Consider the following properties (α) or (β) of (A, G, σ) depending on n ∈ N:
(α) There is n ≥ 1 such that, for each non-zero a ∈ A+, non-invertible b ∈ A+ and
ε > 0, there exist d1, . . . , dn ∈ A and g1, . . . , gn ∈ G that satisfy the inequality
(3) in Definition 6.1.
(β) A is unital, and there is n ≥ 2 such that, for each non-zero a ∈ A+ and ε > 0,
there exist d1, . . . , dn ∈ A and g1, . . . , gn ∈ G that satisfy the inequality (4) in
Definition 6.2.
If A is unital and (A, G, σ) satisfies (α) then it satisfies (β) with n replaced by n+1, and
if (A, G, σ) satisfies (β) then it satisfies (α) -- with same n ∈ N. If (A, G, σ) satisfies
(α) or (β), then the algebra A is G-simple, i.e., A and {0} are the only G-invariant
closed ideals of A.
If A contains a projection p 6= 0 with pAp = C· p, then A is a C*-subalgebra of Mn+1
(respectively of Mn) if (A, G, σ) has property (α) (respectively has property (β)).
The shift action σ of the cyclic group Zn+1 on A := C(Zn+1) satisfies (α) for n ∈ N
and is element-wise properly outer.
Proof. If A is unital and σ has property (α), then let b := 1 − (kak−1a)3, and find
d1, . . . , dn and g1, . . . , gn that satisfy the inequality (3). If we let gn+1 := e and dn+1 :=
kak−3/2a, then a and g1, . . . , gn, gn+1 satisfy (4).
It shows that actions on unital A
that satisfy property (α) are also actions that satisfy (β) with n + 1 in place of n. If
(A, G, σ) satisfies (β) and non-zero elements a, b ∈ A+ are given with kbk = 1, then
d1b1/2, . . . , dnb1/2 and g1, . . . , gn is a solution of the inequality (3) in Definition 6.1 if
d1, . . . , dn and g1, . . . , gn satisfy the inequality (4) of Definition 6.2.
The properties (α) and (β) imply that {0} and A are the only G-invariant closed
ideals of A : If A is non-unital in case (α), then the approximation, as expressed by the
inequalities (3), shows that for each non-zero a ∈ A+ the smallest closed G-invariant
ideal of A containing a contains each b ∈ A+.
If A is unital and the actions has
property (β) then each G-invariant closed ideal of A contains 1. If A is unital and the
C*-dynamical system (A, G, σ) satisfy property (α) then it satisfies property (β) with
n replaced by n + 1. Thus, again, A and {0} are the only closed G-invariant ideals.
18
From now on, we suppose that there exists a projection 0 6= p ∈ A+ with pAp = Cp.
We call those projections "minimal", even if minimal non-zero projections of a C *-
algebra A do not have the property pAp = Cp in general, e.g. the unit of the Jiang-Su
algebra Z is a minimal projection. We show that this assumption of the existence of
such p ∈ A, together with the assumption that σ satisfies (α), implies that A is unital.
Thus A satisfies (β) with n + 1 in place of n . Then we derive that property (β) and
the existence of such p ∈ A imply that A is a C *-subalgebra of Mn.
It is obvious that the ideal socle(A) generated by those "minimal" projections is
(universally) invariant under all automorphisms of A, i.e., α(socle(A)) = socle(A) for
all α ∈ Aut(A). This happens also for the closure J of socle(A). Thus, J must be
G-invariant. It follows that J = A using socle(A) 6= ∅.
It is not difficult to see, that J is isomorphic to the c0-direct sum of a family of
algebras K(Hτ ) of compact operators on suitable Hilbert spaces Hτ , and that p is a
rank-one projection on some Hτ0. Let H denote the Hilbert space sum of the Hilbert
spaces Hτ . Then A becomes isomorphic to a non-degenerate C *-subalgebra of the alge-
bra of compact operators K(H) on H, in a way that each minimal non-zero projection
p ∈ A becomes a rank one projection on H. This happens also for all σg(p). Recall
that every projection in A ⊆ K(H) has finite rank in L(H). Since A is a C *-subalgebra
of K(H), A is in particular an AF-algebra, and -- therefore -- contains an approximate
unit (qτ ) consisting of an upward directed net of projections of finite rank in H.
We show that A must be unital if (A, G, σ) satisfies (α) in addition: Suppose that A
is not unital, then none of the projections (qτ ) are invertible in A. Therefore, we can
take b := qτ , a := p and ε = 1/2 in (3). It follows that each qτ has linear rank ≤ n.
This implies that the approximate unit (qτ ) must be constant qτ = e for all τ ≥ τ1
with suitable τ1. Then e ∈ A is necessarily the unit element of A, in contradiction to
our assumption that A is not unital.
It follows that A must be unital, and -- as above observed -- the action σ satisfies
property (β) with n replaced by n + 1.
If A is unital and (A, G, σ) satisfies property (β), then we take a := p and ε := 1/2
in inequalities (4). It shows that the rank of 1A in its representation is ≤ n. Thus A is
a C *-subalgebra of Mn in case (β).
The crossed product C(Zn+1) ⋊ Zn+1 is naturally isomorphic to Mn+1. Hence, by
If a ∈
Remark 5.3, the action of Zn+1 on C(Zn+1) is element-wise properly outer.
19
A+ := C(Zn+1)+ is non-zero and b ∈ A+ is not invertible, then there are non-zero
minimal projections p, q ∈ A+ and δ > 0 such that a ≥ δp and b ≤ kbk · (1 − q). Select
g1, . . . , gn ∈ Zn+1 with Pn
j=1 σgj (a) ≥ 1 − q.
Thus, there exists T ∈ (1 − q)A+(1 − q) with T (Pn
j=1 σgj (a))T = 1 − q. Then a, b,
j=1 σgj (p) = 1 − q. It implies that δ−1Pn
dj := T b1/2, j = 1, . . . , n and g1, . . . , gn satisfy the inequality (3) for each ε > 0.
(cid:3)
Remark 6.4. Let B be a non-zero simple C *-algebra. In preperation for the proof of
Theorem 1.2 we display here a number of properties equivalent to strong pure infinite-
ness of B.
(i) B is strongly purely infinite.
(ii) Each non-zero element of B+ is properly infinite in sense of [13].
(iii, n) There exists n ∈ N such that, for each non-zero elements a, b ∈ B+ and ε > 0,
there exists n elements d1, . . . , dn ∈ B with
kd∗
1ad1 + · · · + d∗
nadn − bk < ε ,
(5)
and B is not isomorphic to Mk for any k ≤ n.
(iv) B is locally purely infinite in sense of [3, def. 1.3], i.e., each non-zero hereditary
C *-subalgebra of B contains a non-zero stable C *-subalgebra.
(v) B is purely infinite in the sense of J. Cuntz [5, p. 186], i.e., each non-zero
hereditary C *-subalgebra contains an infinite projection.
Proof. Property (ii) implies (i) by [3, thm. 5.8] and (i) implies (ii) by [14, prop. 5.4].
Property (iii,n = 1) is equivalent to (ii) by [13, thm. 4.16]. The properties (iii,n = 1),
(iv) and (v) are equivalent by [3, prop. 3.1].
(iii)⇒(ii): Suppose that B is elementary, i.e., that B ∼= K(H) for some Hilbert space
H. By (iii), applied on some rank-one projection a := p ∈ B+, b ∈ B+, and ε = 1/2, it
follows that each element of b ∈ B+ has rank ≤ n. Thus, H has dimension k ≤ n, and
B ∼= Mk. But the latter case was excluded in (iii,n). Therefore B is non-elementary
and hence has the global Glimm halving property in sense of [3, def. 2.6]. This is easy
to see for non-elementary simple B (or use Glimm halving [20, lem. 6.7.1]). Since B
is simple property (iii, n) ensures that B satisfies property (i) of [3, def. 1.2] of pi(n).
Therefore, [3, prop. 4.14] says that B is pi(1). Since B is simple and B 6= C, there are
no non-zero characters on B. In particular pi(1) ensures that B is purely infinite in
the sense of [13, def. 4.1]. By [13, thm. 4.1] we obtain property (ii).
(cid:3)
20
Proof of Theorem 1.2: By Lemma 6.3, A is G-simple. Thus the action σ is
automatically exact by Definition 3.5. Since σ is element-wise properly outer by as-
sumption, it is now also residually properly outer, and Theorem 3.8 applies. It says
that F := A+ is a filling family in A ⋊σ,λ G. In particular, A separates the ideals of
the reduced crossed product. It shows that B := A ⋊σ,λ G is simple.
If b ∈ B+ with kbk = 1, then there exists non-zero z ∈ B such that z∗z ≤ b
and zz∗ ∈ A, because A+ is filling for B: One way to see this is to use property (i) of
Definition A.3 on a′ := (3b−2)+, b′ := (3b−1)+−a′ and c′ := 3b−a′−b′ to get elements
zj, d ∈ B. This imply that z∗
i zi = c′ec′ ≤ kc′kkekc′ ≤ 3kc′kkek b, where
e 6= 0 because d∗ed = a′ 6= 0.
j zj ≤ e = Pi z∗
Take δ ∈ (0,kzk2/2). Then 0 6= z(z∗z − δ)+z∗ = ϕ(zz∗) ∈ A+ for some suitable
ϕ ∈ Cc(0,∞]+.
We consider three cases: (i) the action is n-majorizing and A is non-unital, (ii) the
action is n-majorizing and A is unital and (iii) the action is n-covering and A is unital.
Since G is discrete, A is a non-degenerate C *-subalgebra of B. In particular, A+
contains an approximate unit (eν) of positive contractions in A+ for B, which we will
use for case (i). In case (ii)-(iii) let eν := 1, where 1 is the unit of A. Define m := n + 1
for case (ii) and m := n for case (i) and (iii). By each of the Definitions 6.1 and
j (z(z∗z − δ)+z∗)fjk < ε for fj := U(g−1
j )dj in B.
6.2 (and using Lemma 6.3 to get property (β) in case (ii) with n replaced by m),
for each ε > 0 and eν ∈ A+ there are d1, . . . , dm ∈ A and g1, . . . , gm ∈ G such that
keν − Pj f ∗
By Remark 2.1(ii), there exists a contraction d0 ∈ B with d∗
the elements yj := d0z∗fj ∈ B satisfy keν−Pm
to a ∈ B+, we get that B has the following property:
For any two non-zero elements b, a ∈ B+ and ε > 0 there exists m elements
d1, . . . , dm ∈ B such that k a − Pm
j bdj k < ε. A simple C *-algebra B with
this property is strongly purely infinite by Remark 6.4 if B is not isomorphic to Mk for
some k ≤ m. The case that B is isomorphic to a C *-subalgebra of Mm has been ex-
cluded by the definitions: For case (i) we know A is non-unital, hence B is non-unital.
For case (ii)-(iii) we know A is not isomorphic to a C *-subalgebra of Mm, hence B is
not isomorphic to a C *-subalgebra of Mm.
(cid:3)
0bd0 = (z∗z − δ)+. Then
j byjk < ε. Since a1/2eνa1/2 converges
j=1 y∗
j=1 d∗
Lemma 6.5. The following are equivalent for C*-algebras B.
(i) B does not contain a non-zero projection p ∈ B+ with pBp = C · p.
21
(ii) For every non-zero hereditary C*-subalgebra D of B, each maximal commutative
C*-subalgebra C of D has perfect Gelfand spectrum bC. i.e., bC does not contain
an isolated point.
(iii) For every a1, a2 ∈ B+\{0} and c ∈ B there are b1, b2 ∈ B+\{0} with b1cb2 = 0,
and bj ≤ aj (j = 1, 2).
Proof. (iii)⇒(i): Let p∗ = p = p2 ∈ B\{0}, put ak := c := p for k = 1, 2. Then there
are non-zero b1, b2 ∈ (pBp)+ with b1b2 = 0. Thus pBp 6= C · p.
(i)⇒(ii): Let D 6= {0} a hereditary C *-subalgebra of B, and C a maximal commuta-
tive C *-subalgebra of D. Suppose that bC is not perfect. Then bC contains an isolated
point χ. The point χ corresponds to a projection 0 6= p ∈ C+ ⊆ D+ with pBp = C · p,
see [21, lem. 7.14].
(ii)⇒(iii): It is easy to see that a commutative C *-algebra C with a perfect spectrum
bC contains non-zero contractions e1, e2 ∈ C+ with e1e2 = 0, because the locally compact
Hausdorff space bC must in particular contain two different points (6= ∞).
Given c ∈ B and non-zero a1, a2 ∈ B+, let dj := (aj − kajk/2)+ and x := d1/2
1 cd1/2
.
Notice that 0 6= dj ∈ B+, and that 0 6= d1/2
j ≤ aj for all non-zero contractions
0 ≤ yj ∈ djBdj.
If x 6= 0, consider the hereditary
C *-subalgebra D := x∗Bx = x∗xBx∗x that is generated by x∗x, and is contained in
2Bd2. Let C be a maximal commutative C *-subalgebra of D with x∗x ∈ C. Then C
d∗
contains non-zero contractions e1, e2 ∈ D+ with e1e2 = 0 = e1(x∗x)1/2e2.
If x = 0 then take bj
j yjd1/2
:= dj.
2
It is well-known (and easy to see) that the polar decomposition x = v(x∗x)1/2 in B∗∗
1Bd1. Thus f := ve1v∗ ∈ xBx∗ and has the
2 e2d1/2
has the property that vDv∗ = xBx∗ ⊆ d∗
property f xe2 = ve1(x∗x)1/2e2 = 0. It follows that b1 := d1/2
satisfy b1cb2 = d1/2
and b2 := d1/2
1 f d1/2
1 f xe2d1/2
2 = 0 and 0 6= bj ≤ aj.
1
2
(cid:3)
Proposition 6.6. Let (A, G, σ) a C*-dynamic system with non-zero non-unital A. If
the action σ of G on A is an 1-majorizing action in the sense of Definition 6.1, then
A is G-simple and σ is G-separating for A.
Proof. The algebra A is G-simple and A does not contain a projection p 6= 0 with
pAp = C · p by Lemma 6.3. To show that σ is G-separating let a1, a2 ∈ A+\{0}, c ∈ A
and ε > 0. By Lemma 6.5, there exist b1, b2 ∈ A+\{0} with b1cb2 = 0 and bj ≤ aj
(j = 1, 2). Using (twice) that the action is 1-majorizing we find ej ∈ A, hj ∈ G for
22
j = 1, 2 such that k e∗
we get kd∗
j ajdj − σgj (aj)k < ε and d∗
j σhj(cid:0)bjajbj(cid:1) ej − aj k < ε . With gj := h−1
j
1cd2 = 0, i.e., σ is G-separating.
and dj := bjσgj (ej)
(cid:3)
Remark 6.7. Suppose that (A, G, σ) is a C *-dynamical system, A is unital and com-
mutative, and G is discrete. Then the following properties (i) -- (iv) of the action σ are
equivalent:
(i) The action is 2-covering in sense of Definition 6.2.
(ii) The corresponding (adjoint) action bσ, on bA is a strong boundary action in the
sense of Definition A.1.
(iii) The action is 1-majorizing in sense of Definition 6.1.
(iv) The action is 2-filling in sense of Definition A.2, and A is not isomorphic to a
subalgebra of M2(C).
We do not know if, also for non-commutative and unital A, every 2-covering action is
a 1-majorizing action, or a 2-filling action.
Proof. We show more general implications, except for (i)⇒(ii). In particular we show
that an action σ on a unital abelian C *-algebra A is n-filling if and only if it is n-
covering provided that the (linear) dimension of A is greater than n.
g
(i)⇒(iv) (for A unital, commutative, any n):
Suppose that A ∼= C(X), and take any n ≥ 2. Let α denote the action of G on X
inducing σ, i.e., σg(f ) = f ◦ α−1
for f ∈ A and g ∈ G. Since the action α is minimal by
Lemma 6.3, Remark [11, rem. 0.4] shows it suffices to prove that for each non-empty
subset U of X there exist g1, . . . , gn ∈ G, such that αg1(U)∪ αg2(U)∪· · ·∪ αgn(U) = X.
Let such U be given. Select non-zero a ∈ A+ with support contained in U. By (i) there
exist g1, . . . , gn ∈ G and d1, . . . , dn ∈ A such that Pn
2 . In particular
for each x ∈ X, σgj (a)(x) is non-zero for some j, so x ∈ αgi(U).
j σgj (a)dj ≥ 1
j=1 d∗
(iv)⇒(i) (for A unital, commutative/non-commutative, any n):
Suppose that A is unital, and take any n ≥ 2. Let 0 6= a ∈ A+. Using (iv) there are
g1, . . . , gn ∈ G and δ > 0 such that D := Pn
j=1 σgj (a) ≥ δ1 , and A is not isomorphic
to a C *-subalgebra of Mn. Thus, D is invertible in A and Pn
j σgj (a)dj = 1 for
dj := D−1/2 in A.
j=1 d∗
(iii)⇒(i) (for A unital, commutative/non-commutative, any n):
Each n-majorizing actions on unital A is an (n + 1)-covering action by Lemma 6.3.
23
(i)⇒(ii) (for A unital, commutative, one n):
Suppose that A ∼= C(X) and let α denote the action of G on X inducing σ. The
equivalence of (i) and (iv), shows that for given 0 6= a ∈ A+, there exists g ∈ G and
δ > 0 such that a + σg(a) ≥ δ1.
Let U ⊆ X open and non-empty. There is 0 6= a ∈ C(X)+ with support a−1(0,∞) ⊆
U. Choose h ∈ G and δ > 0 with a + σh(a) ≥ δ1. It implies that σh(a)(x) > 0 for all
x ∈ X\U. Thus, αg(x) ∈ U for all x ∈ X\U and g := h−1, i.e., there exists g ∈ G with
αg(X\U) ⊆ U.
Given non-empty open subsets U and V of X. We let W := U ∩ V if U ∩ V 6= ∅.
Then g ∈ G with αg(X\W ) ⊆ W satisfies αg(X\U) ⊆ V . If U ∩ V = ∅ then we find
g, h ∈ G with αg(X\U) ⊆ U ⊆ X\V and αh(X\V ) ⊆ V . Then αhg(X\U) ⊆ V .
The space X contains more than two points because A is not isomorphic to a C *-
subalgebra of M2(C). Thus, (X, G, α) satisfies the conditions of Definition A.1 of a
strong boundary action.
(ii)⇒(iii) (for A unital/non-unital, commutative, one n):
We show (iii) using possibly less than (ii): Let X be a locally compact space that is
not necessarily compact and contains more than 2 points. Let α an action of G on
X with the property that, for every compact subset K ⊆ X with K 6= X and each
non-empty open subset U ⊆ X, there exists g ∈ G with αg(K) ⊆ U. Then the adjoint
action σ of α on A := C0(X) is an 1-majorizing action of G on A.
Indeed: Let 0 6= a ∈ A+, b ∈ A+ non-invertible, and ε > 0. Put δ := ε/3. Then,
considered as functions on X, they have the property that U := a−1(kak/2,∞) is non-
empty and open and K := b−1[δ,∞) is compact. Find h ∈ G with αh(K) ⊆ U. Then
x ∈ K ⇔ b(x) ≥ δ implies that for g := h−1 we get σg(a)(x) = a(αh(x)) > kak/2 . It
follows kd∗σg(a)d − bk < ε with d ∈ A+ given by d(x) := σg(a)(x)−1/2(b(x) − 2δ)1/2
+ for
x ∈ K and d(x) := 0 for x ∈ X\K.
(cid:3)
Remarks 6.8. (i) Let α be an action of a discrete group G on a locally compact
Hausdorff space X with more than two points, and σ the induced action on A := C0(X).
By Remark 6.7 the following properties are equivalent for X compact:
(1) The action α is a strong boundary action (Definition A.1) in the sense of [18]:
For each pair of non-empty open subsets U and V of X there exists g ∈ G with
U ∩ αg(V ) = X.
24
(2) For any compact set K 6= X and any open set U 6= ∅ there exist t ∈ G such
that αt(K) ⊆ U.
(3) For every non-zero a ∈ A+, every non-invertible b ∈ A+ and every ε > 0, there
exist d ∈ A and g ∈ G such that kd∗σg(a)d − bk < ε.
Clearly, this can not be the case if X is locally compact but is not compact. In general
(i.e., when X is compact or non-compact) we know (1)⇒(2)⇒(3). Properties (2)-(3)
are both candidates for a generalised notion of a strong boundary action, however only
(3) applies when A is non-commutative.
(ii) The notion of a strong boundary action (Definition A.1) is defined on compact
Hausdorff spaces with more than two points. In view of Remark 6.8(i) and Remark 6.7,
we call the 1-majorizing actions on not-necessarily unital or commutative C *-algebras
also strong boundary actions.
(iii) Suppose that a discrete group G acts by a topologically free action α on a compact
Hausdorff space X, and that X contains more than two points. It was shown in [18,
thm. 5] that the crossed product C(X) ⋊σ,λ G is purely infinite provided that the
action -- in addition -- is a strong boundary action. Since topological freeness implies σ
is element-wise properly outer (by [1, prop. 1]) we conclude that, with the terminology
of Remark 6.8(ii), [18, thm. 5] is a special case of Theorem 1.2.
(cid:3)
(iv) Let α be an action on a non-compact locally compact Hausdorff space X with more
than two points and σ the induced action. It was shown in Proposition 6.6 that σ is
G-separating if σ is a strong boundary (i.e., 1-majorizing) action. A simpler argument
applies if we assume that for any compact set K 6= X and any non-empty open set
U ⊆ X there exist g ∈ G such that αg(K) ⊆ U:
Proof. Since any finite subset M of X is compact, it can be moved by suitable αg into
any non-empty open subset U of X. In particular X is perfect and each non-empty
open set V ⊆ X contains at least two non-empty open disjoint subsets V1 and V2. Let
K1 ⊆ U1 and K2 ⊆ U2 with Kj compact (hence Kj 6= X) and Uj open. If U1 and U2
are disjoint, then we can take g1 = g2 = e in Lemma 5.1(ii). If V := U1 ∩ U2 6= ∅,
then consider the above disjoint non-empty open subsets Vj ⊆ V . By assumption,
there exist g1, g2 ∈ G with αgj (Kj) ⊆ Vj ⊆ Uj. Thus, the adjoint action σ of α is
G-separating.
(cid:3)
(v) Suppose that a discrete group G acts by a topologically free action α on a non-
compact locally compact Hausdorff space X, and that X contains more than two points.
25
Then the crossed product C0(X) ⋊σ,λ G is purely infinite provided that the following
property holds: for any compact set K 6= X and any non-empty open set U ⊆ X there
exist t ∈ G such that αt(K) ⊆ U. This follows as a corollary of Theorem 1.1, also of
Theorem 1.2 or of Corollary 5.2.
Proof. We must verify the following properties according to each of the listed results:
(1.1) The action σ is exact, residually properly outer and G-separating.
(1.2) The action σ is 1-majorizing, and element-wise properly outer.
(5.2) The action σ is exact, G-separating and fulfills (*): For every closed G-invariant
subset Y of X and every g 6= e the set {y ∈ Y : αg(y) = y} has empty interior.
By Remark 6.8(i) we know the action σ is a strong boundary (i.e., 1-majorizing) action.
Hence A is G-simple, cf. Lemma 6.3. In particular it follows that the action α on X
is minimal. This reduces property (*) to the definition of topological freeness, cf. [1,
p.120]. The minimality of the action α implies that the corresponding adjoint action
σ : G → Aut(C0(X)) is exact, and that it becomes residually properly outer if it is
element-wise properly outer. But σ is element-wise properly outer if and only if α
is a topological free action (see [1, p.120] or [23, cor 2.22]). It remains to show σ is
G-separating, but this is already contained in Remark 6.8(iv).
(cid:3)
(vii) It is an important point that a strong boundary action is often G-separating and
(in fact always) minimal, but the notion of a G-separating action is not typically related
to minimality. Consequently, working with G-separating actions allows us to consider
ideal-related classification of non-simple strongly purely infinite crossed products.
Remark 6.9. If A has real rank zero, then one can restrict the conditions in Definitions
4.1, 6.1 and 6.2 to projections p, q ∈ A in place of the elements a, b ∈ A+.
Proof. Case of Definition 4.1: Let a1, a2 ∈ A+, c ∈ A, ε > 0 and define
δ := ε/(1 + ka1k + ka2k) .
By [4], Dj
:= ajAaj contains an approximate unit consisting of non-zero projec-
tions. Thus, there are projections pj ∈ Dj such that kaj − a1/2
j k < δ. Use [13,
prop. 2.7(i)] and the comment following [13, prop. 2.6] to select zj ∈ Dj satisfying
j aj zj = pj. Let c′ := z∗
z∗
Suppose that there exists ej ∈ A, gj ∈ G such that
and ke∗
j pjej − σgj (pj)k < δ
1c′ e2k < δ .
pj a1/2
j
1cz2.
ke∗
26
j
) and dj
j
:= zjejσgj (a1/2
j
) . They satisfy d∗
j ajdj = v∗
j e∗
j pja1/2
1c′e2v2, we get kd∗
1e∗
j pjejvj,
j ajdj − σgj (aj)k < (1 + kajk)δ ≤ ε . Since
). Thus, kd∗
1cd2k < δ(ka1k · ka2k)1/2 ≤ δ(ka1k + ka2k) ≤ ε .
:= σgj (a1/2
Define vj
j σgj (pj)vj = σgj (a1/2
v∗
1cd2 = v∗
d∗
Case of Definitions 6.1 and 6.2: Let a1, a2 ∈ A+ and ε > 0, with a1 6= 0 and a2 not
invertible in A (respectively a2 = 1). We can assume ε ≤ 1. Define δ := ε/(1 + ka2k) .
Choose pj, zj ∈ Dj := ajAaj as above with ka1/2
j ajzj = pj. Then
p1 6= 0 and p2 is not invertible in A (i.e., p2 6= 1) if a2 is not invertible, otherwise p2 = 1:
If p2 is invertible then 1 ∈ a2Aa2, so a2 is invertible. Conversely, if a2 is invertible then
kp2 − 1k < ε/2, so p2 is invertible.
If there are e1, . . . , en ∈ A and g1, . . . , gn ∈ G with k p2 − Pj e∗
dj := σgj (z1)eja1/2
j σgj (p1)ej k < δ , then
j −ajk < δ and z∗
j pja1/2
(cid:3)
satisfies k a2 − Pj d∗
2
j σgj (a1)dj k < (1 + ka2k)δ ≤ ε .
Acknowledgments
Parts of this work were conducted while the second named author was at the Fields
Institute from 2009 to 2012. It is with great pleasure we forward our thanks to the
Fields Institute and in particular Professor George Elliott for all the support. This
research was also supported by the Australian Research Council.
Appendix A.
In this appendix we have included a few recent definitions and results that are
frequently cited throughout this paper. The results quoted from [15] are available as
preprint.
Definition A.1 ([18]). Let α be an action of a discrete group G on a compact spaces
X with at least three points. The action α is as strong boundary action if for every pair
U and V of non-empty open subsets of X there exists t ∈ G such that αt(X \ U) ⊆ V .
Definition A.2 ([11]). An action σ of a discrete group G on a unital C *-algebra A
is n-filling (n ≥ 2) if, for all b1, . . . , bn ∈ A+, with kbjk = 1 for each j, and all ε > 0,
there exist g1, . . . , gn ∈ G such that Pn
Definition A.3 ([15]). Let F be a subset of A+. The set F is a filling family for A,
if F satisfies the following equivalent conditions (i) and (ii).
j=1 σgj (bj) ≥ 1 − ε.
27
(i) For every a, b, c ∈ A with 0 ≤ a ≤ b ≤ c ≤ 1, with ab = a 6= 0 and bc = b,
there exists z1, z2, . . . , zn ∈ A and d ∈ A with zj(zj)∗ ∈ F , such that ec = e and
d∗ed = a for e := z∗
1z1 + . . . + z∗
nzn.
(ii) For every hereditary C *-subalgebra D of A and every primitive ideal I of A
with D 6⊆ I there exist f ∈ F and z ∈ A with z∗z ∈ D and zz∗ = f 6∈ I.
Lemma A.4 ([15]). Suppose that A ⊆ B is a C*-subalgebra of B and F ⊆ A+ is a
subset of A+. If F is filling for A, and A+ is filling for B, then F is a filling family
for B.
Remark A.5 ([15]). Let A ⊆ B be C *-algebras and F ⊆ A+. If F := A+ ⊆ B is
filling for B, then the map I ∈ I(B) 7→ I ∩ A ∈ I(A) is injective, i.e., A separates the
closed ideals of B.
Definition A.6 ([15]). A C *-algebra A is strongly purely infinite (for short: s.p.i. )
if, for every a, b ∈ A+ and ε > 0, there exist elements s, t ∈ A such that
k s∗a2s − a2 k < ε , k t∗b2t − b2 k < ε and k s∗abtk < ε .
(6)
Remark A.7 ([15]). A C *-algebra A is strongly purely infinite if and only if for every
a, b ∈ A+, c ∈ A and ε > 0, there exist contractions s, t ∈ A such that
k s∗as − ak < ε , k t∗bt − bk < ε and k s∗ctk < ε .
(7)
Definition A.8 ([15]). Let S ⊆ A be a multiplicative sub-semigroup of a C *-algebra
A and C ⊆ A a subset of A. An n-tuple (a1, . . . , an) of positive elements in A has the
matrix diagonalization property with respect to S and C, if for every [aij] ∈ Mn(A)+
with ajj = aj and aij ∈ C (for i 6= j) and εj > 0, τ > 0 there are elements s1 , . . . , sn ∈ S
with
ks∗
j ajj sj − ajjk < εj ,
(8)
If S = C = A then this is the matrix diagonalization property of (a1, . . . , an) as defined
in [14, def. 5.5], and we say that (a1, . . . , an) has matrix diagonalization (in A).
i aijsjk < τ for i 6= j .
and ks∗
Definition A.9 ([15]). Let F be a subset of A+. The family F has the (matrix)
diagonalization property (in A) if each finite sequence a1, . . . , an ∈ F has the matrix
diagonalization property (in A) of Definition A.8.
Lemma A.10 ([15]). Suppose that F ⊆ A+ is invariant under ε-cut-downs, i.e., that
for each a ∈ F and ε ∈ (0,kak) we have (a − ε)+ ∈ F .
Then the family F has the matrix diagonalization property, if and only if, each pair
of elements in F has the matrix diagonalization property of Definition A.8.
28
Lemma A.11 ([15]). Let ε0 > 0 and non-empty subsets F ⊆ A+, C ⊆ A be given,
and let S ⊆ A be a (multiplicative) sub-semigroup of A that satisfies s∗
2Cs1 ⊆ C for all
s1, s2 ∈ S. Suppose that the following properties hold:
(i) For every ε0 > δ > 0, the pair ((a1 − δ)+, (a2 − δ)+) the matrix diagonalization
property with respect to S and C of Definition A.8.
(ii) ϕ(a1)cϕ(a2) ∈ C for each c ∈ C and ϕ ∈ Cc(0,∞]+.
(iii) ϕ(a1)s, ϕ(a2)s ∈ S for each s ∈ S and ϕ ∈ Cc(0,∞]+.
Then, for every c ∈ span(C), a1, a2 ∈ F , ε0/2 ≥ ε > 0, and τ > 0, there exists
s1, s2 ∈ S that fulfil ksjk2 ≤ 2kajk/ε and
ks∗
1a1s1 − a1k < ε ,
ks∗
2a2s2 − a2k < ε
and ks∗
1cs2k < τ .
(9)
Theorem A.12 ([15]). The minimal tenor product of a strongly purely infinite and an
exact C*-algebra is strongly purely infinite.
Theorem A.13 ([15]). Suppose that A+ contains a filling family F that has the diag-
onalization property (in A). Then A is strongly purely infinite.
References
[1] R.J. Archbold and J.S. Spielberg, Topologically free actions and ideals in discrete C*-dynamical
systems, Proc. Edinburgh Math. Soc. (2) 37 (1994), 119 -- 124. 6, 7, 16, 17, 25, 26
[2] G. Arzhantseva, V. Guba, and M. Sapir, Metrics on diagram groups and uniform embeddings in
Hilbert space, Comment. Math. Helv. 81 (2006), 911 -- 929. 9
[3] E. Blanchard and E. Kirchberg, Non-simple purely infinite C*-algebras: the Hausdorff case, J.
Funct. Anal. 207 (2004), 461 -- 513. 4, 20
[4] L.G. Brown and G.K. Pedersen, C*-algebras of real rank zero, J. Funct. Anal. 99 (1991), 131 -- 149.
26
[5] J. Cuntz, K-theory for certain C*-algebras, Ann. of Math. 113 (1981), 181 -- 197. 20
[6] G.A. Elliott, Some simple C*-algebras constructed as crossed products with discrete outer auto-
morphism groups, Publ. Res. Inst. Math. Sci. 16 (1980), 299 -- 311. 6
[7] R. Exel, M. Laca, and J. Quigg, Partial dynamical systems and C*-algebras generated by partial
isometries, J. Operator Theory 47 (2002), 169 -- 186. 7
[8] D.S. Farley, Proper isometric actions of Thompson's groups on Hilbert space, Int. Math. Res.
Notes 45 (2003), 2409 -- 2414. 9
[9] E. Guentner and J. Kaminker, Exactness and uniform embeddability of discrete Groups, J. London
Math. Soc.(2) 70 (2004), 703 -- 718. 9
[10] U. Haagerup and G. Picioroaga, New presentations of Thompson's groups and applications, J.
Operator Theory 66 (2011), 217 -- 232. 8
29
[11] P. Jolissaint and G. Robertson, Simple purely infinite C*-algebras and n-filling actions, J. Funct.
Anal. 175 (2000), 197 -- 213. 3, 23, 27
[12] S. Kawamura and J. Tomiyama, Properties of topological dynamical systems and corresponding
C*-algebras, Tokyo J. Math. 13 (1990), 251 -- 257. 7, 16
[13] E. Kirchberg and M. Rørdam, Non-simple purely infinite C*-algebras, Amer. J. Math. 122 (2000),
637 -- 666. 20, 26
[14]
, Infinite non-simple C*-algebras: absorbing the Cuntz algebras O∞, Adv. Math. 167
(2002), no. 2, 195 -- 264. 4, 12, 20, 28
[15] E. Kirchberg and A. Sierakowski, Filling families and strong pure infiniteness, preprint 2014. 2,
27, 28, 29
[16] E. Kirchberg and S. Wassermann, Exact groups and continuous bundles of C*-algebras,
Math. Ann. 315 (1999), 169 -- 203. 8, 14
[17] A. Kishimoto, Outer automorphisms and reduced crossed products of simple C*-algebras,
Comm. Math. Phys. 81 (1981), 429 -- 435. 6
[18] M. Laca and J. Spielberg, Purely infinite C*-algebras from boundary actions of discrete groups,
J. Reine Angew. Math. 480 (1996), 125 -- 139. 1, 3, 24, 25, 27
[19] D. Olesen and G.K. Pedersen, Applications of the Connes spectrum to C*-dynamical systems. III,
J. Funct. Anal. 45 (1982), 357 -- 390. 6, 7
[20] G.K. Pedersen, C*-algebras and their automorphism groups, LMS Monographs, vol. 14, Academic
Press Inc., London, 1979. 20
[21] M. Rørdam, F. Larsen, and N. Laustsen, An introduction to K-theory for C*-algebras, London
Mathematical Society Student Texts 49, Cambridge University Press, Cambridge, 2000. 22
[22] J. Renault, The ideal structure of groupoid crossed product C*-algebras, With an appendix by
Georges Skandalis. J. Operator Theory 25 (1991), 3 -- 36. 6, 16
[23] A. Sierakowski, The ideal structure of reduced crossed products, Munster J. Math. 3 (2010), 223 --
248. 2, 6, 8, 9, 16, 17, 26
Institut fur Mathematik, Humboldt Universitat zu Berlin, Unter den Linden 6, D --
10099 Berlin, Germany
[email protected]
School of Mathematics & Applied Statistics, University of Wollongong, Faculty of
Engineering & Information Sciences, 2522 Wollongong, Australia
[email protected]
30
|
1012.4435 | 2 | 1012 | 2011-04-13T16:18:04 | Integrable representations of involutive algebras and Ore localization | [
"math.OA",
"math.RA",
"math.RT"
] | Let $\mathcal A$ be a unital algebra equipped with an involution $(\cdot)^\dagger$, and suppose that the multiplicative set $\mathcal S\subseteq \mathcal A$ generated by the elements of the form $1 + a^\dagger a$ satisfies the Ore condition. We prove that: (i) Cyclic representations of $\mathcal A$ admit an integrable extension (acting on a possibly larger Hilbert space), and (ii) Integrable representations of $\mathcal A$ are in bijection with representations of the Ore localization $\mathcal A\mathcal S^{-1}$ (which we prove to be an involutive algebra). This second result is a limited converse to a theorem by Inoue asserting that representations of symmetric involutive algebras are integrable. | math.OA | math |
Integrable representations of involutive
algebras and Ore localization
Rodrigo Vargas Le-Bert∗
Instituto de Matem´atica y F´ısica, Universidad de Talca
Casilla 747, Talca, Chile
April 5, 2011
Abstract
Let A be a unital algebra equipped with an involution (·)†, and suppose
that the multiplicative set S ⊆ A generated by the elements of the form
1 + a†a contains only regular elements and satisfies the Ore condition. We
prove that:
• Ultracyclic representations of A admit an integrable extension (acting
on a possibly larger Hilbert space).
• Integrable representations of A are in bijection with representations
of the Ore localization AS−1 (which we prove to be an involutive
algebra).
This second result can be understood as a restricted converse to a theorem
by Inoue asserting that representations of symmetric involutive algebras
are integrable.
2010 MSC: 16S (primary); 46L (secondary).
1 Introduction
Unbounded operator algebras appear in several important domains, such as
quantum field theory, representations of Lie algebras and quantum groups.
Consequently, there has long been an interest in developing their theory, which
despite that has grown rich in technical details and relatively poor in applica-
tions (see [14, 6, 1, 7] for complete expositions, [2] for a physical applications
survey, and [12] for applications in other fields). Among the causes for this fact
∗[email protected]. Supported by Fondecyt Postdoctoral Grant No3110045.
1
lie the inherent difficulties in the representation theory of general involutive
algebras, which we summarize as follows.
Let A be a unital algebra equipped with an involution (·)† and H a Hilbert
space.
1. Speaking of a representation presupposes that sum, product and involu-
tion are well-defined; however, unbounded operators are not defined all
over H and cannot be blindly added or composed. Hence, one usually
postulates the existence of an invariant domain V ⊆ H for the operators
which will represent A, thus allowing for a pointwise definition of their
sum and product. Letting
L†(V) = { a : V → V V ⊆ D (a∗) and a∗V ⊆ V } ,
one obtains an algebra with involution a† = a∗V, and then can define a
representation to be a morphism π : A → L†(V).
2. Operators in L†(V) are closable, for their adjoint is densely defined.
Therefore, the natural algebraic operations with them are the so-called
strong sum and strong product, see Definition 2.3. The problem arises from
this, together with the fact that each a ∈ L†(V) may admit more than
one closed extension. As a consequence, pointwise operations become
ambiguous in a sense. They are too weak a version of the strong ones.
For a representation π, this means that
¯π(a + b) ⊆ ¯π(a) + ¯π(b),
¯π(ab) ⊆ ¯π(a) ¯π(b),
¯π(a†) ⊆ ¯π(a)∗,
and equality does not hold in general.
Simply put, L†(V) is not an object that truly captures the algebraic structure
of closed operators, thus allowing, even in the simplest cases, for the appearance
of ill-behaved representations. The following two examples are archetypical of
good and bad behaviour.
Example 1.1. Let G be a Lie group with Lie algebra g. Representations of
the universal enveloping algebra A = U(g) that arise by differentiation from
unitary representations of G are called integrable (a complete exposition is found
in [14]). They are characterized by the fact that they respect the involution:
¯π(a†) = ¯π(a)∗, ∀a ∈ A,
a property which is taken as the definition of integrability in the general case.
Integrability is also very important when A is the observable algebra of a
quantum system.
Example 1.2. Let M be any properly infinite von Neumann algebra acting on a
separable Hilbert space. Then, there exists a representation π of A = C[x, y], the
algebra of polynomials in two commuting hermitian variables, such that π(x)
2
and π(y) are essentially self-adjoint and their spectral projections generate M.
Such representations are not integrable. For more details, see [13, Section 5]
and [14, Section 9.4].
The second example shows that conditions must be imposed on represen-
tations, in order to exclude pathological behaviour and make them useful in
practice. Now, during the last decade, two general approaches to the well-
behaved representations have been developed. In what follows we provide a
brief description of both (a complete discussion is found in [1]). After that, we
will be in position to comment on one important limitation that they share, and
on how it is overcome in an approach that we propose in this paper.
The first one, of an algebraic flavor, is due to Schm udgen [15]. It generalizes
the notion of integrable representations of universal enveloping algebras U(g),
integrable representations of U(g) determine unitary
in the following way:
representations of the corresponding simply connected Lie group G. Those,
in turn, are equivalent to representations of the Banach algebra L1(G). Now, it
turns out that an integrable representation of U(g) can be recovered from the
corresponding representation of C∞
0 (G) ⊆ L1(G) by making use of the natural
action of U(g) on C∞
0 (G). The generalization to arbitrary involutive algebras
goes as follows: say that the involutive algebra A and the normed involutive
algebra A0 are compatible if there is an action of A on A0 such that
(a · x)†y = x†(a† · y), ∀a ∈ A, ∀x, y ∈ A0.
Then, non-degenerate, continuous representations π0 of A0 determine what we
define to be the well-behaved representations π of A by
π(a)π0(x) = π0(a · x),
a ∈ A, x ∈ A0.
The second approach, of an analytic flavor, was proposed by Bhatt, Inoue
and Ogi [3].
It is based on unbounded C*-seminorms. Given an involutive
algebra A, an unbounded C*-seminorm is a C*-seminorm p : A0 → R, where
A0 is a subalgebra of A. The kernel I of such a seminorm is a bilateral ideal
of A0, and the completion with respect to p of A0/I is a C*-algebra which we
denote by A. Now, let
N = { x ∈ A0 ax ∈ A0, ∀a ∈ A } ,
which is a left ideal of A. Each representation π0 : A → B(H ) induces a
representation π : A → L†(V), where
V = Span(cid:8) π0(x + I)ξ (cid:12)(cid:12)(cid:12)
x ∈ N, ξ ∈ H (cid:9) ,
by the simple formula π(a)π0(x + I)ξ = π0(ax + I)ξ. Observe, however, that
V might not be dense in H and, when this is the case, we say that the repre-
sentation is well-behaved. One sees that well-behaved representations satisfy
k ¯π(a)k = p(a), for all a ∈ A0 (in general, k ¯π(a)k ≤ p(a)).
3
This second approach is more general than the first one, and can be fur-
ther generalized to the case of partial involutive algebras [17]. The precise
relationship between the two approaches has been worked out in [11]. The
limitation referred to above is more apparent in the first one: obtaining well-
behaved representations of A = U(g) requires knowledge of representations
of A0 = C∞
0 (G). If one is to obtain representations of G from representations of
U(g), this is clearly going in the wrong direction. The same happens with the
second approach, as is best seen with an example: take A = C[x], the commu-
tative free algebra on one hermitian generator, and consider its well-behaved
representation by multiplication operators on L2(R). There is no subalgebra of
A on which the corresponding norm is finite. Thus, obtaining such a simple
representation by the second approach requires enlarging C[x] to contain at
least one function of exponential decay -- a procedure which is, again, taking as
given something that, in several practical cases, should come as a result.
Our approach is precisely based on enlarging the algebra A to contain
inverses to the elements of the form 1 + a†a, thus obtaining a so-called symmetric
involutive algebra. It is known that representations of the latter are integrable [9].
Our main result is Theorem 3.15, which says that when the multiplicative set
S generated by { 1 + a†a a ∈ A } satisfies the Ore condition (see Section 3.2),
integrable representations of A are in bijection with representations of the Ore
localization AS−1.
The paper is organized as follows: the second section is a short reminder
of necessary background, and our results are proved in the third section. For
convenience, an elementary appendix on closable operators is included.
2 Short reminder on representation theory
This section contains no new results. It is included only for the convenience of
non-expert readers, and it will serve to introduce our notations, too.
2.1 Ultracyclic representations and GNS construction
Let V be a complex vector space and a ∈ L(V). We denote their corresponding
algebraic duals by
V† = (cid:8) f : V → C (cid:12)(cid:12)(cid:12)
f is complex linear (cid:9) ,
a† : f ∈ V† 7→ f a ∈ V†.
Now, let A be a unital involutive algebra with involution (·)†, to be represented
by operators in L(V). Allowing for a correspondence between involution and
algebraic duality requires a choice of antilinear inclusion V ֒→ V†, which we
assume given by an embedding V ֒→ H as a dense subspace in a Hilbert space.
4
Definition 2.1. Given a dense subspace V of a Hilbert space H , denote by
L†(V) = { a : V → V V ⊆ D (a∗) , a∗V ⊆ V } ,
which is an involutive algebra with involution a† = a∗V. A representation of the
involutive algebra A is a morphism π : A → L†(V) of involutive algebras. We
say that π is ultracyclic if it admits an ultracyclic vector, that is, an Ω ∈ V such
that V = π(A)Ω.
Definition 2.2. Let A be an involutive algebra and Ah the real vector subspace
of its hermitian elements. We say that x ∈ Ah is positive if it belongs to the cone
Π(A) =
n
X
i=1
λia†
i ai (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
λi > 0, ai ∈ A
⊆ Ah.
The cone of positive elements allows one to define an order relation on A.
This, in turn, gives an order relation on its algebraic dual: we say that f ∈ A†
is positive if
f (x) ≥ 0, ∀x ∈ Π(A).
Recall that positivity implies:
1.
f (a†b) = f (b†a).
2. The Cauchy-Schwartz inequality f (a†b)2 ≤ f (a†a) f (b†b).
We say that f ∈ A† is a state if it is positive and f (1) = 1. We denote the set of
states of A by Σ(A).
Given a representation π : A → L†(V), one obtains a state f ∈ A† from any
vector Ω ∈ V with kΩk = 1 by the formula
f (a) = hΩ, π(a)Ωi.
The GNS construction allows a recovery of both ππ(A)Ω (modulo unitary con-
jugation) and Ω (modulo a phase factor) from f , thus establishing a correspon-
dence between ultracyclic representations and states. We proceed to describe
it; for a complete treatment, see [14].
Let f ∈ Σ(A), and consider the set I = { a ∈ A f (a†a) = 0 } . Using the
Cauchy-Schwartz inequality it is easily seen that I is a left ideal. The quotient
V = A/I is densely embedded in its Hilbert space completion H with respect
to the scalar product
h[a], [b]i = f (a†b),
a, b ∈ A,
where [·] denotes the equivalence class in V of its argument. The GNS repre-
sentation π is simply given by the canonical left A-module structure of V. It
admits the ultracyclic vector Ω = [1] ∈ V.
5
2.2 Integrability
As mentioned in the introduction, the algebraic operations of L†(V) are not
satisfactory from an analytic point of view. Indeed, every a ∈ L†(V) is closable
because its adjoint a∗ is densely defined. Now, while closed operators cannot
always be added or composed, the natural operations when they can are the
following.
Definition 2.3. Let A, B ∈ C(H ), the set of closed, densely defined operators on
a Hilbert space H . If D = D (A) ∩ D (B) is dense and AD + BD is closable, we
define their strong sum
A + B = AD + BD ∈ C(H ).
Analogously, if D = B−1D (A) is dense and ABD is closable, we define their
strong product
AB = ABD ∈ C(H ).
Remark 2.4. When operating with closed operators we will always mean strong
operations. This should cause no confusions, because we will always write
elements of C(H ) in uppercase, and elements of L†(V) in lowercase.
Proposition 2.5. Let a, b ∈ L†(V). If A and B are their respective closures, then A + B
and AB exist.
Proof. Let D = D (A) ∩ D (B). One has that
h(A + B)ξ, ηi = hξ, (a† + b†)ηi, ∀ξ ∈ D, ∀η ∈ V,
whence a† + b† ⊆ (AD + BD)∗ and A + B exists. Analogously, if D = B−1D (A),
hABξ, ηi = hξ, b†a†ηi, ∀ξ ∈ D, ∀η ∈ V,
whence b†a† ⊆ (ABD)∗ and AB exists.
(cid:3)
Remark 2.6. The analytic procedure of closure breaks down the algebraic struc-
ture of L†(V). Indeed,
a + b ⊆ ¯a + ¯b,
ab ⊆ ¯a¯b,
a† ⊆ a∗, ∀a, b ∈ LH (V),
and equalities do not hold in general.
Thus, given a representation π : A → L†(V), one should not expect that ¯π
be a representation too, where
¯π : a ∈ A 7→ π(a) ∈ C(H ).
Integrable representations are, almost by definition, those for which this is
actually the case.
6
Definition 2.7. We say that a representation π : A → L†(V) is integrable if
¯π(a†) = ¯π(a)∗, for all a ∈ A.
Proposition 2.8. Let π : A → L†(V) be a representation. One has that π is
integrable if, and only if, ¯π(A) ⊆ C(H ) is an involutive algebra and ¯π : A → ¯π(A)
is a morphism.
Proof. Sufficiency is obvious. For necessity, let a, b ∈ A and A = ¯π(a), B = ¯π(b).
From the proof of Proposition 2.5,
A + B = (AD + BD)∗∗ ⊆ π(a† + b†)∗ = ¯π(a + b),
where D = D (A) ∩ D (B) and the last equality follows from integrability. Anal-
ogously,
AB = (ABD)∗∗ ⊆ π(b†a†)∗ = ¯π(ab),
where D = B−1D (A).
(cid:3)
2.3 Symmetric involutive algebras
We have seen that representations which are not integrable do not really deserve
their name from an analytic point of view, and that the problem originates in
the fact that L†(V) is not a good replacement for B(H ) in generalizing the
theory of C*-algebras to unbounded operator algebras. Now, the need for
such a generalization has long been recognized and, over time, the notion of
partial involutive algebras [1] has emerged as the safest candidate (because of its
generality). In that approach, one actually gives up the structure of algebra,
fully acknowledging the fact that strong sum and strong product are not defined
all over C(H ) × C(H ). Here, we will limit ourselves to work with subsets of
C(H ) which are involutive algebras with respect to strong sum, strong product
and operator involution. Two known examples are:
the set of measurable
operators affiliated with a von Neumann algebra admitting a normal, semifinite
trace [16]; and the so-called symmetric involutive algebras [9], on which we base
our approach to the well-behaved representations of involutive algebras.
Definition 2.9. A unital, involutive algebra is said to be symmetric if 1 + a†a is
invertible, for all a ∈ A.
There is a generalization to non-unital algebras, see [7], for instance. Repre-
sentations π : A → L†(V) of symmetric involutive algebras enjoy several good
properties, among which we mention:
• They are integrable [10]. We will revisit this in the course of this paper.
• They are direct sums of cyclic representations. We remark that this is not
always the case if A is any involutive algebra, see [14, Corollary 11.6.8].
7
3 Integrable representations and Ore localization
3.1 Integrability and symmetry
All over this subsection, π : A → L†(V) will be a representation and
n
Y
i=1
S =
(cid:16)1 + a†
i ai(cid:17) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
ai ∈ A
.
If π is integrable, it follows immediately from Propositions 2.8 and 3.21 that ¯π(s)
is invertible, for all s ∈ S. Here we show that these two properties are actually
equivalent, and then give an alternative proof of Inoue's result asserting that
representations of symmetric involutive algebras are integrable.
Lemma 3.1. Let s = Qn
i=1(1 + a†
i ai) ∈ S and
S =
n
Y
i=1
(1 + A∗
i Ai), Ai = ¯π(ai).
If ¯π(s) is surjective, then ¯π(s) = S. In particular, ¯π(s) is invertible and has a bounded
inverse.
Proof. Observe that
S = (1 + A∗
1A1)(1 + A∗
2A2)D2 · · · (1 + A∗
i+1Ai+1)−1Di and D1 = D(cid:16)1 + A∗
nAn)Dn
1A1(cid:17). It follows that S is
where Di+1 = (1 + A∗
closed (for it is invertible with bounded inverse). Since π(s) ⊆ S, we conclude
that ¯π(s) ⊆ S, too. But, by hypothesis, ¯π(s) is surjective, whence it does not
admit any injective, strict extension and must be equal to S.
(cid:3)
Corollary 3.2. If ¯π(s) is surjective for all s ∈ S, then
¯π(st) = ¯π(s) ¯π(t), ∀s, t ∈ S.
Proof. It is a straightforward consequence of Lemma 3.1.
(cid:3)
Lemma 3.3. Let a ∈ A and s ∈ S. If ¯π(t) is surjective for all t ∈ S, then ¯π(s)V is a
core for both ¯π(a) and ¯π(a)∗.
Proof. Let t = 1+a†a ∈ S. From Lemma 3.1 and Proposition 3.21 in the appendix,
we see that
¯π(t)−1 : (H , k k) → (cid:16)D ( ¯π(a)) , k k ¯π(a)(cid:17)
8
is continuous and has dense range. Therefore, ¯π(t)−1D will be a core for ¯π(a),
for any dense D ⊆ H . Now, consider D = ¯π(ts)V. It is a dense subspace, for
¯π(ts) is surjective and V is a core for it. Using Corollary 3.2, we conclude that
¯π(t)−1 ¯π(ts)V = ¯π(s)V
is a core for ¯π(a). The fact that it is also a core for ¯π(a)∗ follows from the same
argument, applied this time to t = 1 + aa†. Indeed, now
¯π(t)−1 : (H , k k) → (cid:16)D ( ¯π(a)∗) , k k ¯π(a)∗(cid:17)
is continuous and has dense range.
(cid:3)
Corollary 3.4. π is integrable if, and only if, ¯π(s) is surjective, for all s ∈ S.
Proof. Indeed, given a ∈ A, ¯π(a†) ⊆ ¯π(a)∗. But, being V a core for both of them,
they must actually coincide.
(cid:3)
We finish this subsection with the following result. As mentioned before,
our main result is a partial converse.
Proposition 3.5. Suppose that A ⊆ B, where B is an involutive algebra such that the
elements of S are invertible in it. If π admits an extension π : B → L†(V), then it is
integrable.
Proof. Given s = Qn
the proof of Lemma 3.1,
i=1(1 + a†
i ai), we have to prove that ¯π(s) is surjective. As in
¯π(s) ⊆ S =
n
Y
i=1
(1 + A∗
i Ai), Ai = ¯π(ai).
Now, let ξ ∈ H and consider a sequence { vn } ⊆ V converging to ξ. Since S−1 is
bounded, un = S−1vn converges to, say, η ∈ H . But, by hypothesis, s is invertible
in B, and
un = S−1vn = S−1π(s) π(s−1)vn = π(s−1)vn,
so that un → η and π(s)un = vn → ξ. By closedness, ¯π(s)η = ξ, as needed.
(cid:3)
Remark 3.6. As a consequence, we recover the following result due to Inoue [9]:
if A is symmetric, then its representations are integrable.
9
3.2 Ore localization for involutive algebras
Let A be a unital ring and S ⊆ A a multiplicative subset (that is, such that
SS = S. Note that, in particular, 1 ∈ S). Proposition 3.5 leads to study the
problem of adjoining inverses to the elements of S. It is not hard to see that
there exists a ring AS, called universal localization of A at S, which is a universal
solution to this. Now, when localizing, unexpected things can happen (such as
ending up with a trivial ring) and it is good to have conditions controlling AS
and, above all, enabling one to make calculations in it. From our point of view,
having a manageable localization is important to investigate two problems:
1. The extension of states from A to AS. By GNS construction, this amounts
to obtaining integrable extensions of cyclic representations.
2. The possibility that integrable representations actually extend to repre-
sentations of AS, providing a converse to Proposition 3.5.
In this section we revisit the Ore construction, which deals with the particular
case in which AS is the non-commutative analogue of a ring of fractions. A
complete introduction can be found in [8]. The novelty here is that we show that
the Ore construction carries over smoothly to the case of involutive algebras.
Definition 3.7. Let A be a ring. We say that a subset S ⊆ A is a right (resp. left)
Ore set if it is multiplicative and
∀(a, s) ∈ A × S, ∃(b, t) ∈ A × S, at = sb (resp. ta = bs),
and we say that it is an Ore set if it is both a left and a right Ore set.
Remark 3.8. If s and t are invertible, the equation at = sb can be rewritten as
s−1a = bt−1. In intuitive terms, the Ore property is establishing the possibility of
writing "non-commutative fractions" indifferently from the right or from the
left.
We say that a ring morphism f : A → B is S-inverting if f (s) is invertible
in B, for all s ∈ S. Then, consider the following universal problem for the
S-inverting morphism ℓ:
A
ℓ
f
AS
f
B,
where f is any S-invertible morphism. If S ⊆ A is an Ore set, the solution can
be constructed as follows. On the set A × S define the equivalence relation
(a, s) ∼ (b, t) ⇔ ∃u, v ∈ A, (au, su) = (bv, tv) ∈ A × S,
10
and denote by [a, s] the equivalence class of (a, s). Linear combination and
multiplication of elements [a1, s1], [a2, s2] ∈ (A × S)/∼ are defined as follows:
• Choose (b, t) ∈ A × S such that s1t = s2b. Then,
λ[a1, s1] + [a2, s2] = [λa1t, s1t] + [a2b, s2b] = [λa1t + a2b, s1t].
• Choose (b, t) ∈ A × S such that a2t = s1b. Then,
[a1, s1] · [a2, s2] = [a1b, s2t]
(cid:16) = [a1b, s1b][a2t, s2t]. (cid:17)
The last equality is enclosed in parentheses because it does not necessarily
make sense; we include it because it does motivate the definition.
It can be checked that this definitions equip AS−1 = (A × S)/∼ with a unital
ring structure, whose unit is [1, 1]. The morphism ℓ : A → AS−1 is defined
by a 7→ [a, 1]. This turns out to give an inclusion if S contains uniquely regular
elements -- that is, elements s such that
0 ∈ {as, sa} ⇒ a = 0, ∀a ∈ A.
An analog construction to that of AS−1 can be done if S ⊆ A is a left Ore
set, and in the "bilateral" case those two localizations coincide, modulo an
isomorphism whose restriction to (the image of) A is trivial.
Remark 3.9. Before going further, we note a simple property of the multipli-
cation. Given (a, s) ∈ A × S, let (b, t) ∈ A × S be such that at = sb. Then, by
definition, [1, s][a, 1] = [b, t]. Now, given any u ∈ A such that us ∈ S, one still
has that uat = usb, and therefore
[1, us][ua, 1] = [b, t] = [1, s][a, 1].
We are ready to treat the case of a unital involutive algebra A, which is
not dealt with in ring theory and constitutes our humble contribution to the
subject. Suppose that S ⊆ A is an Ore subset such that S† = S (observe that, in
this case, the left and right Ore conditions are equivalent). On AS−1 define the
following operation:
[a, s]† = [1, s†][a†, 1].
By Remark 3.9, this depends uniquely on the equivalence class of (a, s) because,
given u ∈ A such that su ∈ S,
[au, su]† = [1, u†s†][u†a†, 1] = [1, s†][a†, 1] = [a, s]†.
Proposition 3.10. (·)† : AS−1 → AS−1 is an involution, and ℓ : A → AS−1 is a
morphism of unital involutive algebras.
11
Proof. In fact, given (a1, s1), (a2, s2) ∈ A × S:
• Let (b, t) ∈ A × S be such that s1t = s2b. We have that
(cid:16)λ[a1, s1] + [a2, s2](cid:17)†
= [λa1t + a2b, s1t]† = [1, t†s†
1][ ¯λt†a†
1
+ b†a†
2, 1].
On the other hand, using Remark 3.9,
¯λ[a1, s1]† + [a2, s2]† = [1, s†
2][b†a†
2, 1]
1, 1] + [1, s†
1][ ¯λa†
1][ ¯λt†a†
1][ ¯λt†a†
2][a†
2, 1]
1, 1] + [1, b†s†
+ b†a†
2, 1].
1
= [1, t†s†
= [1, t†s†
Hence, (·)† is antilinear.
• Let (b, t) ∈ A × S be such that a2t = s1b. We have that
(cid:16)[a1, s1][a2, s2](cid:17)†
On the other hand,
= [a1b, s2t]† = [1, t†s†
2][b†a†
1, 1].
[a2, s2]†[a1, s1]† = [1, s†
= [1, s†
2, 1][1, s†
1][a†
2][a†
2][1, t†][b†, 1][a†
1, 1] = [1, s†
1, 1] = [1, t†s†
2, s†
2][a†
2][b†a†
1][a†
1, 1].
1, 1]
• Finally,
[a, s]†† = (cid:16)[1, s†][a†, 1](cid:17)†
= [a†, 1]†[1, s†]† = [a, 1][1, s] = [a, s].
(cid:3)
From now on, we will write as−1 instead of [a, s] for the elements of AS−1.
3.3 Integrable representations and representations of AS−1
i=1(1 + a†
that S = n Qn
All over this subsection A will be a unital involutive algebra and we will assume
ai ∈ A o contains only regular elements and satisfies
the Ore condition. We start by proving that every ultracyclic representation
of A admits an integrable extension, which is a simple consequence of the
following fact.
i ai) (cid:12)(cid:12)(cid:12)
Lemma 3.11. Π(A) is cofinal in Π(AS−1).
Proof. Given (a, s) ∈ A × S, we will prove that (s−1a)†s−1a ≤ a†a (recall that all the
elements of AS−1 can be written in the form s−1a). In order to do so, note the
following order property of involutive algebras, which is a direct consequence
of the definition: if x ≤ y, then a†xa ≤ a†ya. Therefore, it suffices to show that
(s−1)†s−1 ≤ 1, for all s ∈ S -- which we do next.
12
Let b ∈ A. We have that 1 ≤ 1 + 2b†b + (b†b)2 = (1 + b†b)2, from which
(cid:16)(1 + b†b)−1(cid:17)†
(1 + b†b)−1 = (1 + b†b)−2 ≤ 1.
Now, suppose that
s = (1 + b†
1b1)(1 + b†
2b2) · · · (1 + b†
nbn) ∈ S,
bi ∈ A.
Inductively, we see that
(s−1)†s−1 = (1 + b†
1b1)−1 · · · (1 + b†
nbn)−1(1 + b†
nbn)−1 · · · (1 + b†
1b1)−1 ≤ 1.
(cid:3)
Corollary 3.12. Every ultracyclic representation π : A → L†(V) admits an inte-
grable extension π : A → L†( V).
Proof. Let f ∈ Σ(A) be the state producing, by GNS construction, the represen-
tation π. Since Ah is cofinal in the real ordered vector space (AS−1)h, the restric-
tion fh = f Ah admits a positive extension fh to all of (AS−1)h (see, for instance,
[5, Theorem 9.8]) which, in turn, uniquely determines a state f ∈ Σ(AS−1) by
f (x + iy) = fh(x) + i fh(y),
x, y ∈ Ah.
By GNS construction, we get an ultracyclic representation π : AS−1 → L†( V),
where
V = (AS−1)/ { a ∈ AS−1 f (a†a) = 0 } .
The inclusion A ⊆ AS−1 induces an inclusion V ⊆ V, which is well defined
because
{ a ∈ A f (a†a) = 0 } ⊆ { a ∈ AS−1 f (a†a) = 0 } .
The restriction of π to A gives the desired extension, which is integrable thanks
to Proposition 3.5.
(cid:3)
Remark 3.13. Results of this kind are not unknown in some important cases. For
instance, in [4] Borchers and Yngvason work out the case of Borchers algebras,
appearing naturally in quantum field theory.
Next we prove our main result. We start with the following crucial lemma.
Lemma 3.14. Suppose that π is integrable, and let (a, s), (b, t) ∈ A × S be such that
ta = bs. Then,
¯π(a) ¯π(s)−1 ⊇ ¯π(t)−1 ¯π(b).
In particular, ¯π(s)−1V ⊆ D ( ¯π(a)) , for all (a, s) ∈ A × S.
Proof. Indeed, starting from π(t)π(a) = π(b)π(s) we find, pre-multiplying by
¯π(t)−1 and post-multiplying by ¯π(s)−1π(s)V, that
¯π(a) ¯π(s)−1π(s)V = ¯π(t)−1 ¯π(b)π(s)V.
13
Now, given ξ ∈ D ( ¯π(b)), by Lemma 3.3 there exists a sequence [scn] → ξ such
that [bscn] → ¯π(b)ξ. Since ¯π(t)−1 is bounded, we see that
¯π(a)[cn] = ¯π(a) ¯π(s)−1[scn] = ¯π(t)−1[bscn]
converges. Therefore, ¯π(s)−1ξ ∈ D ( ¯π(a)) and ¯π(a) ¯π(s)−1ξ = ¯π(t)−1 ¯π(b)ξ, as
needed.
(cid:3)
Theorem 3.15. There exists a bijective correspondence between integrable representa-
tions of A and representations of AS−1.
Proof. By Proposition 3.5, representations of AS−1 induce, by restriction, inte-
grable representations of A. In the other direction, let π : A → L†(V) be an
integrable representation. We extend it to AS−1 as follows. Define
V = span (cid:18)[
s∈S
¯π(s)−1V(cid:19) ⊇ V.
Note that, by Lemma 3.14,
π(a) := ¯π(a) V belongs to L†( V), which means verifying that:
V ⊆ D ( ¯π(a)), for all a ∈ A. We will show that
1. ¯π(a) V ⊆ V. Indeed, given s ∈ S, let (b, t) ∈ A × S be such that ta = bs.
One has that ¯π(a) ¯π(s)−1V = ¯π(t)−1 ¯π(b)V ⊆ ¯π(t)−1V ⊆ V.
V ⊆ D(cid:16)( ¯π(a) V)∗(cid:17) = D ( ¯π(a)∗). By integrability, ¯π(a)∗ = ¯π(a†) and we
already know that V ⊆ D(cid:16) ¯π(a†)(cid:17).
2.
3. ( ¯π(a) V)∗ V ⊆ V. Again, follows by integrability and the already proved
fact that ¯π(a†) V ⊆ V.
By the universal property of the Ore localization, the morphism π will factor
through AS−1 if we prove, given s ∈ S, that π(s) is invertible in L†( V). This
means verifying that ¯π(s)−1 V ∈ L†( V). But, by integrability,
V is invariant
under ¯π(s)−1:
¯π(s)−1 V = [
t∈S
¯π(s)−1 ¯π(t)−1V = [
t∈S
¯π(ts)−1V ⊆ V.
The same calculation shows that V is invariant under ( ¯π(s)−1)∗ = ¯π(s†)−1, and
the conclusion follows.
(cid:3)
Remark 3.16. An important question arises: do integrable representations al-
ways admit an extension to the universal localization AS? Our proof makes
essential use of the Ore condition, which conceivable might not always hold
for subsets of C(H ) which are algebras.
14
Appendix: short reminder on unbounded operators
Let H be a Hilbert space and a : D (a) ⊆ H → H a linear operator, not
necessarily bounded. We say that a is closed if its graph
G(a) = { ξ ⊕ aξ ξ ∈ H } ⊆ H ⊕ H
is closed, and that a is closable if G(a) is a graph, namely if
0 ⊕ η ∈ G(a) ⇒ η = 0, ∀η ∈ H .
In this case, there exists a unique operator ¯a such that G(a) = G( ¯a), which is
called the closure of a.
Remark 3.17. If a closed operator was defined all over H , by the closed graph
theorem it would be bounded. While this is not always the case, it is expected
that closed operators inherit some of the good behaviour of bounded operators.
In order to define the adjoint a∗ of the unbounded operator a, we consider
the linear function
fξ = hξ, a(·)i : D (a) → C.
If fξ had an extension ¯fξ ∈ H ∗, then a∗ξ should be its inverse image via the
antilinear isomorphism ι : H → H ∗ induced by the inner product. Define,
whence,
D (a∗) = (cid:8) ξ ∈ H (cid:12)(cid:12)(cid:12)
∃C(ξ) > 0, ∀η ∈ D (a) , hξ, aηi ≤ C(ξ)kηk (cid:9) .
By Riesz's theorem, if ξ ∈ D (a∗), then a∗ξ = ι−1( ¯fξ) exists.
Remark 3.18. D (a∗) has no reason to be dense in H .
A geometrical interpretation for the adjoint operator can be obtained through
the isometry
J : ξ ⊕ η ∈ H ⊕ H 7→ (−η) ⊕ ξ ∈ H ⊕ H .
Indeed, it is easily shown that G(a∗) = (cid:16)J G(a)(cid:17)⊥
Another consequence is the following proposition.
. In particular, a∗ is closed.
Proposition 3.19. D (a∗) is dense in H if, and only if, a is closable. Moreover, in this
case a∗∗ = ¯a.
An operator with dense domain a : D (a) ⊆ H → H is hermitian if
hξ, aηi = haξ, ηi, ∀ξ, η ∈ D (a) .
This means that a∗ is an extension of a, which we denote by a ⊆ a∗. In particular,
D (a) ⊆ D (a∗) and therefore a is closable. Since the closure of a is a minimal
closed extension, we have that
a ⊆ ¯a ⊆ a∗.
15
If ¯a = a∗, we say that a is essentially self-adjoint and that ¯a is self-adjoint -- but, of
course, this is not always the case.
Now we consider closed operators.
In order to distinguish them from
arbitrary unbounded operators, we write them in capital letters.
Definition 3.20. A subspace V ⊆ H is said to be a core for the closed operator
A : D (A) ⊆ H → H if V ⊆ D (A) and AV = A.
By definition, a core V for A is such that { ξ ⊕ Aξ ξ ∈ V } is dense in G(A).
Now, the isomorphism of vector spaces
ξ ∈ D (A) 7→ ξ ⊕ Aξ ∈ G(A)
enables one to pass the inner product of H ⊕ H to D (A): we obtain
hξ, ηiA = hξ, ηi + hAξ, Aηi.
In terms of this space, V is a core for A if, and only if, it is dense in (D (A) , k kA).
We conclude the appendix with the statement of an important result -- see [5],
for example, for a proof.
Proposition 3.21. Let A : D (A) ⊆ H → H be a closed operator. One has that:
1. 1 + A∗A : A−1D (A∗) → H is a bijection, and its inverse
(1 + A∗A)−1 : H → A−1D (A∗) ⊆ D (A)
is bounded as an operator (H , k k) → (D (A) , k kA).
2. D (A∗A) = A−1D (A∗) is a core for A.
References
[1] J. P. Antoine, A. Inoue, and C. Trapani, Partial ∗-algebras and their operator
realizations, Kluwer Academic Publishers, 2002.
[2] F. Bagarello, Algebras of unbounded operators and physical applications: a
survey, Rev. Math. Phys. 19 (2007), 231 -- 271.
[3] S. J. Bhatt, A. Inoue, and H. Ogi, Unbounded C*-seminorms and unbounded
C*-spectral algebras, J. Operat. Theor. 45 (2001), no. 1, 53 -- 80.
[4] H. J. Borchers and J. Yngvason, Partially conmmutative moment problems,
Math. Nachr. 145 (1990), 111 -- 117.
[5] J. B. Conway, A course in functional analysis, Springer -- Verlag, 1994.
16
[6] D. A. Dubin and M. A. Hennings, Quantum mechanics, algebras and distribu-
tions, Pitman research notes in mathematics, vol. 238, Longman scientific
and technical, 1990.
[7] M. Fragoulopoulou, Topological algebras with involution, North-Holland
mathematical studies, vol. 200, Elsevier, 2005.
[8] K. R. Goodearl and R. B. Warfield, An introduction to noncommutative noethe-
rian rings, second ed., Cambridge university press, 2004.
[9] A. Inoue, On a class of unbounded operator algebras, Pacific J. Math. 65 (1976),
no. 1, 77 -- 95.
[10]
[11]
, Unbounded representations of symmetric ∗-algebras, J. Math. Soc.
Japan 29 (1977), no. 2, 219 -- 232.
, Well-behaved ∗-representations of ∗-algebras, Acta Univ. Oulu. Ser. A
Sci. Rerum Natur. 408 (2004), 107 -- 117.
[12] A. Mallios and M. Haralampidou, Topological algebras and applications,
American mathematical society, 2005.
[13] R. T. Powers, Self-adjoint algebras of unbounded operators, Commun. Math.
Phys. 21 (1971), 85 -- 124.
[14] K. Schm udgen, Unbounded operator algebras and representation theory,
Birkhauser, 1990.
[15]
, On well-behaved unbounded representations of ∗-algebras, J. Operat.
Theor. 48 (2002), no. 3, 487 -- 502.
[16] I. Segal, A non-commutative extension of abstract integration, Ann. Math. 57
(1953), no. 3, 401 -- 457.
[17] C. Trapani, Unbounded C*-seminorms, biweights, and ∗-representations of ∗-
algebras: a review, Int. J. Math. Math. Sci. 2006 (2006), 1 -- 34.
17
|
1106.5523 | 2 | 1106 | 2011-07-15T14:18:58 | Divisibility properties for C*-algebras | [
"math.OA"
] | We consider three notions of divisibility in the Cuntz semigroup of a C*-algebra, and show how they reflect properties of the C*-algebra. We develop methods to construct (simple and non-simple) C*-algebras with specific divisibility behaviour. As a byproduct of our investigations, we show that there exists a sequence $(A_n)$ of simple unital infinite dimensional C*-algebras such that the product $\prod_{n=1}^\infty A_n$ has a character. | math.OA | math |
Divisibility properties for C ∗-algebras
Leonel Robert∗ and Mikael Rørdam∗
Abstract
We consider three notions of divisibility in the Cuntz semigroup of a C ∗-algebra,
and show how they reflect properties of the C ∗-algebra. We develop methods to
construct (simple and non-simple) C ∗-algebras with specific divisibility behaviour.
As a byproduct of our investigations, we show that there exists a sequence (An) of
An has a
simple unital infinite dimensional C ∗-algebras such that the product Q∞
character.
n=1
1
Introduction
A unital embedding of a matrix algebra Mm(C) into a unital C ∗-algebra A can exist only
if the equation mx = [1A] has a solution x ∈ K0(A). Thus, only C ∗-algebras in which the
class of the unit in K0 is m-divisible admit a unital embedding of Mm(C). Whereas all von
Neumann algebras (with no central summand of type In for n finite) have this divisibility
property for all m, the same is not true for C ∗-algebras, even for the simple ones. C ∗-
algebras can fail to have non-trivial projections. Even if they have many projections, as
in the real rank zero case, one cannot expect to solve the equation mx = [1A] exactly in
K0(A). This paper is concerned with different weaker notions of divisibility, phrased in
terms of the Cuntz semigroup of the C ∗-algebra, and with how they relate to embeddability
properties of the C ∗-algebra. Instead of solving the equation mx = [1A] for x ∈ K0(A),
one should look for less restrictive notions of divisibility. One can try, for example, to solve
the inequalities mx ≤ h1Ai ≤ nx in the Cuntz semigroup of A for fixed positive integers
m and n (typically with m < n). We say that A is (m, n)-divisible if one can solve this
inequality. This is one of three divisibility properties we shall consider in this paper. We
show that there is a full ∗-homomorphism from CMm(C), the cone over Mm(C), into A if
and only if A is (m, n)-divisible for some n.
Let us mention three embedding problems that served as motivation for this paper.
Let A be a unital C ∗-algebra with no non-zero finite dimensional representations. Can one
always find an embedding of some unital simple infinite dimensional C ∗-algebra into A?
Can one always find an embedding of CM2(C) into A whose image is full in A? Can one
∗This research was supported by the Danish National Research Foundation (DNRF) through the Centre
for Symmetry and Deformation
1
always find two positive mutually orthogonal full elements in A? An affirmative answer
to the former problem will imply an affirmative anwer to the second problem, which is
known as the "Global Glimm Halving problem". An affirmative answer to the Global
Glimm Halving problem will imply an affirmative answer to the last mentioned problem.
We suspect that all three problems may have negative answers.
The second and the third problem led us to consider two new notions of divisibility
properties. In more detail, we say that A is weakly (m, n)-divisible if there are elements
x1, . . . , xn in Cu(A) such that mxj ≤ h1Ai ≤ x1 + · · · + xn. Weak divisibility measures
the rank of A in the sense that A is weakly (m, n)-divisible for some n if and only if A
has no non-zero representations of dimension < m. In particular, A has no non-zero finite
dimensional representations if and only if for every m there is n such that A is weakly
(m, n)-divisible. We say that A is (m, n)-decomposable if there are elements y1, . . . , ym in
Cu(A) such that y1 + · · · + ym ≤ h1Ai ≤ nyj. For a given m, A is (m, n)-decomposable for
some n if and only if A contains m pairwise orthogonal, pairwise equivalent full positive
elements.
It was shown in [DHTW09] that there exists a simple unital infinite dimensional C ∗-al-
gebra which does not admit a unital embedding of the Jiang-Su algebra Z. This answered
in the negative a question posed by the second named author. It is implicit in [DHTW09]
that this simple C ∗-algebra has bad divisibility properties, cf. Remark 3.14. This leads us
to a useful observation, which loosely can be formulated as follows: if A and B are unital
C ∗-algebras, and if there is a unital ∗-homomorphism from A to B, then the divisibility
properties of B are no worse than those of A. In other words, if A has better divisibility
properties than B, then you can not unitally embed A into B.
Comparability in the Cuntz semigroup is concerned with the extent to which one can
conclude that x ≤ y if the "size" of x (e.g., measured in terms of states) is (much) smaller
than the "size" of y. Comparability and divisibility are probably the two most fundamental
properties of the Cuntz semigroup. Good comparability and divisibility properties are
necessary and sufficient conditions in Winter's theorem, [Win], to conclude that a simple,
separable, unital C ∗-algebra with locally finite nuclear dimension tensorially absorbs the
Jiang-Su algebra. Also, good comparability and divisibility properties are both necessary
and sufficient conditions to ensure that the Cuntz semigroup of a simple, separable, unital,
exact C ∗-algebra A is (naturally) isomorphic to Aff(T (A)) ⊔ V (A), cf. [PT07], [BPT08],
and [ERS].
The existence of simple C ∗-algebras with bad comparability properties was discovered
by Villadsen, [Vil98], in the mid 1990's. This discovery was the first indication that the
Elliott conjecture could be false (in general), and it was also the first example of a simple
C ∗-algebra exhibiting "infinite dimensional" behaviour. Villadsen's example in [Vil98] has
been generalized extensively by several authors (including Villadsen himself) to exhibit
simple C ∗-algebras with various kinds of unexpected behaviour, including many ways of
failing to have good comparability properties. However,
little work has been done to
construct simple C ∗-algebras with bad divisibility behaviour, and the literature does not
contain systematic ways of producing such examples. In this paper we show that there is
a duality between comparability and divisibility (Lemma 6.1), and we use this duality to
2
construct examples of simple and non-simple C ∗-algebras with bad divisibility behaviour.
We use Lemma 6.1 to obtain a result that concerns the structure of C ∗-algebras that
arise as the tensor product of a sequence of unital (simple non-elementary) C ∗-algebras.
Each such C ∗-algebra will of course have non-trivial central sequences. Dadarlat and Toms
n=1 A of a fixed unital C ∗-algebra
A contains a unital copy of an AHS-algebra without characters, then it automatically
absorbs the Jiang-Su algebra. It is not known if this condition always is satisfied, even
n=1 A has the Corona
Factorization Property for every unital A without characters (and in particular for every
unital simple C ∗-algebra A 6= C). In the other direction we give, in Section 7, an example of
n=1 An,
proved in [DT09] that if the infinite tensor power N∞
when A is simple and non-elementary. We show in Section 6 that N∞
a sequence of simple unital infinite dimensional C ∗-algebras whose tensor power, N∞
does not absorb (or admit an embedding of) the Jiang-Su algebra.
sional C ∗-algebras such that Q∞
Non-divisibility of a C ∗-algebra can be interpreted as a degree of inhomogeneity (or
"lumpiness") of the C ∗-algebra. Simple C ∗-algebras are sometimes thought of as being
very homogeneous, as for example in [KOS03]. From this point of view it may at first be
surprising that a simple infinite dimensional C ∗-algebra can fail to have good divisibility
properties. We show that there exists a sequence (An) of simple, unital, infinite dimen-
n=1 An (and also the associated ultrapowers of (An)) has a
character. None of the C ∗-algebras An can have a character (being simple and not equal
to C), however we can show that they posses "almost characters" as defined in Section 8.
In Section 9 we consider what one might call "super-divisibility", which leads to a
(new) notion of infiniteness of positive elements (and which implies that a multiple of the
given element is properly infinite). We use this to reformulate the Corona Factorization
Property of semigroups considered in [OPR]. We study variations of examples, originally
due to Dixmier and Douady, and answer in this way two questions from [KR00] in the
negative: The sum of two properly infinite positive elements need not be properly infinite,
and the multiplier algebra of a C ∗-algebra which has a properly infinite strictly positive
element need not be properly infinite.
2 Preliminaries
Let A be a C ∗-algebra and let Cu(A) denote the Cuntz semigroup of A, i.e., the set of
Cuntz equivalence classes of positive elements in A ⊗ K endowed with suitable order and
addition structures. Given a positive element a in A ⊗ K, we shall denote by hai the
Cuntz class of a.
In [CEI08], Coward, Elliott and Ivanescu give an alternative picture
of the Cuntz semigroup where Cu(A) consists of suitable equivalence classes of countably
generated Hilbert C ∗-modules over A. When using the Hilbert modules picture of Cu(A),
we shall denote the equivalence class of a countably generated Hilbert module H by [H].
We present here some well-known definitions and facts about the Cuntz semigroup.
First of all, we shall frequently use the axioms of the category Cu, of which Cu(A) is always
an object (see [CEI08]). An ordered abelian semigroup S is an object in the category Cu
if
3
(A1) every increasing sequence in S has a supremum,
(A2) for every u ∈ S there exists a sequence (ui) in S such that ui ≪ ui+1 and supi ui = u,
(A3) if u′ ≪ u and v′ ≪ v, then u′ + v′ ≪ u + v,
(A4) if (ui) and (vi) are increasing sequences then supi ui + supi vi = supi(ui + vi).
Recall that u ≪ v in S if whenever v = supi vi for some increasing sequence (vi) in S, then
u ≤ vi for some i. An element u ∈ S is called compact if u ≪ u.
We also note the following two additional properties of the Cuntz semigroup of a C ∗-
algebrawhich are note listed among the axioms of Cu. The first of them asserts that the
Cuntz semigroup of a C ∗-algebra almost has the Riesz Decomposition Property, and the
second states that its order relation is almost the algebraic order.
(P1) if u′ ≪ u ≤ v + w, then there exist v′ and w′, with v′ ≤ u, v and w′ ≤ u, w, and such
that u′ ≪ v′ + w′.
(P2) if u′ ≪ u ≤ v, then there exists w such that u′ + w ≤ v ≤ u + w.
For the proofs of these facts, see [Rob11, Proposition 5.1.1] for the first and [RW10, Lemma
7.1 (i)] for the second.
We will also make use of the sequential continuity with respect to inductive limits of
the functor Cu(·) proved in [CEI08] (see also the proof of [ERS, Theorem 4.8]). It can be
stated as follows:
Proposition 2.1 ([CEI08]). Let A = lim−→(Ai, ϕi,j) be a sequential inductive limit of C ∗-al-
gebras.
(i) For every u ∈ Cu(A) there exists an increasing sequence (ui)∞
i=1 with supremum u
and such that each ui belongs to Sj Im(Cu(ϕj,∞)).
(ii) If u, v ∈ Cu(Ai) are such that Cu(ϕi,∞)(u) ≤ Cu(ϕi,∞)(v), then for every u′ ≪ u
there exists j such that Cu(ϕi,j)(u′) ≤ Cu(ϕi,j)(v).
Remark 2.2 (The cone over a matrix algebra). Let m be a positive integer, and let
CMm(C) denote the cone over Mm(C), i.e., the C ∗-algebra of all continuous functions
f : [0, 1] → Mm(C) that vanish at 0.
For each i, j = 1, 2, . . . , m, let eij denote the (i, j)th matrix unit in Mm(C), and denote
by eij ⊗ ι the function t 7→ teij in CMm(C). Then (eii ⊗ ι)m
i=1 are positive contractions in
CMm(C) which are pairwise equivalent and orthogonal. (We say here that two positive
elements a and b are equivalent, denoted a ∼ b, if a = xx∗ and b = x∗x for some element
x in the ambient C ∗-algebra.)
We recall the following well-known universal property of CMm(C) (see for example
[RW10, Propositions 2.3 and 2.4]): Let A be any C ∗-algebra and let a1, a2, . . . , am be pos-
itive contractions in A. Then there exists a ∗-homomorphism ϕ : CMm(C) → A satisfying
ϕ(ejj ⊗ ι) = aj if and only if a1, a2, . . . , am are pairwise orthogonal and pairwise equivalent
in A.
4
The following lemma is well-known:
Lemma 2.3. Let A be a C ∗-algebra and let a, b1, b2, . . . , bn be positive elements in A. Then:
i=1hbii if and only if for each ε > 0 there exist d1, d2, . . . , dn ∈ A such that
i=1hbii ≤ hai if and only if for each ε > 0 there exist mutually orthogonal positive
elements a1, a2, . . . , an in aAa such that ai ∼ (bi − ε)+ for all i.
i=1 dibid∗
i .
(a − ε)+ = Pn
(i) hai ≤ Pn
(ii) Pn
Proof. (i). If hai ≤ Pn
i=1hbii, then a - b1 ⊕ b2 ⊕ · · · ⊕ bn, whence
(a − ε)+ = d∗(b1 ⊕ b2 ⊕ · · · ⊕ bn)d =
n
Xi=1
d∗
i bidi
for some d = (d1, d2, . . . , dn)t ∈ Mn,1(A). The converse statement is trivial.
(ii). Suppose that Pn
for some d = (d1, d2, . . . , dn) in M1,n(A). Thus d∗
all i. Put ai = a1/2did∗
The converse statement is trivial.
i=1hbii ≤ hai. Then (b1 − ε)+ ⊕ (b2 − ε)+ ⊕ · · · ⊕ (bn − ε)+ = d∗ad
i adi = (bi − ε)+ for
i a1/2. It is now straightforward to verify that the ai's are as desired.
j adi = 0 if j 6= i and d∗
Here is another lemma that we will use frequently:
Lemma 2.4. Let a and b be positive elements in A ⊗ K with kak ≤ 1, and let m ∈ N. The
following are equivalent:
(i) mhai ≤ hbi,
(ii) for each ε > 0 there exist mutually orthogonal positive elements b1, b2, . . . , bm in
b(A ⊗ K)b such that hbii = h(a − ε)+i for all i.
(iii) for each ε > 0 there exists a ∗-homomorphism ϕ : CMm(C) → b(A ⊗ K)b such that
hϕ(e11 ⊗ ι)i = h(a − ε)+i.
Proof. The implications (iii) ⇒ (ii) ⇒ (i) are clear, cf. Remark 2.2 and Lemma 2.3. Let
us show that (i) implies (iii). Let ε > 0 be given. By Lemma 2.3 (ii) there are mutually
orthogonal positive elements b1, b2, . . . , bm in b(A ⊗ K)b such that each bj is equivalent to
(a − ε)+. By the universal property of the cone CMm(C), see Remark 2.2, there is a ∗-ho-
momorphism ϕ : CMm(C) → b(A ⊗ K)b satisfying ϕ(ejj ⊗ ι) = bj. Hence (iii) holds.
3 Three divisibility properties
Definitions and basic properties
Definition 3.1. Let A be a C ∗-algebra and fix u ∈ Cu(A). Let m, n ≥ 1 be integers.
Then:
5
(i) u is (m, n)-divisible if for every u′ ∈ Cu(A) with u′ ≪ u there exists x ∈ Cu(A) such
that mx ≤ u and u′ ≤ nx.
The least n such that u is (m, n)-divisible is denoted by Divm(u, A), with Divm(u, A) =
∞ if no such n exists.
(ii) u is (m, n)-decomposable if for every u′ ∈ Cu(A) with u′ ≪ u there exist elements
x1, x2, . . . , xm ∈ Cu(A) such that x1 + x2 + · · · + xm ≤ u and u′ ≤ nxj for all
j = 1, 2, . . . , m.
The least n such that u is (m, n)-decomposable is denoted by ∂ivm(u, A), with
∂ivm(u, A) = ∞ if no such n exists.
(iii) u is weakly (m, n)-divisible if for every u′ ∈ Cu(A) with u′ ≪ u there exist elements
x1, x2, . . . , xn ∈ Cu(A) such that mxj ≤ u for all j = 1, 2, . . . , m and u′ ≤ x1 + x2 +
· · · + xn.
The least n such that u is weakly (m, n)-divisible is denoted by divm(u, A), with
divm(u, A) = ∞ if no such n exists.
Remark 3.2. In the case that u in Definition 3.1 is compact (e.g., when A is unital and
u = h1Ai), the conditions above read a little easier:
(i) u is (m, n)-divisible if there exists x ∈ Cu(A) such that mx ≤ u ≤ nx.
(ii) u is (m, n)-decomposable if there exist elements x1, x2, . . . , xm ∈ Cu(A) such that
x1 + x2 + · · · + xm ≤ u ≤ nxj for all j = 1, 2, . . . , m.
(iii) u is weakly (m, n)-divisible if there exist elements x1, x2, . . . , xn ∈ Cu(A) such that
mxj ≤ u ≤ x1 + x2 + · · · + xn.
The three divisibility properties above are related as follows:
Proposition 3.3. Let m, n ∈ N and u ∈ Cu(A). Then
divm(u, A) ≤ Divm(u, A),
∂ivm(u, A) ≤ Divm(u, A),
divm(u, A) ≤ ∂ivm(u, A)m.
Proof. The two first inequalities are clear (take xi = x in both cases).
To prove the last inequality, suppose that u is (m, n)-decomposable. We show that u
is weakly (m, nm)-divisible. Let u′ ≪ u and find u′′ such that u′ ≪ u′′ ≪ u. There exist
i=1 xi ≤ u and u′′ ≤ nxi for all i. We proceed to
elements x1, . . . , xm in Cu(A) such that Pm
find elements
y(i1, . . . , ik), y(i1, . . . , ik) ∈ Cu(A),
k = 1, . . . , m, ij = 1, . . . , n,
satisfying
(a) y(i1, . . . , ik) ≪ y(i1, . . . , ik),
6
(b) y(i1, . . . , ik−1) ≪ Pn
(c) y(i1, . . . , ik−1) ≤ Pn
i=1 y(i1, . . . , ik−1, i) if k ≥ 2, and u′ ≪ Pn
i=1 y(i1, . . . , ik−1, i) if k ≥ 2, and u′ ≤ Pn
(d) y(i1, . . . , ik) ≤ xk,
i=1 y(i),
i=1 y(i),
(e) y(i1, . . . , ik−1, ik) ≤ y(i1, . . . , ik−1) if k ≥ 2, and y(i) ≤ u′′.
The elements above are constructed inductively after k using the following fact:
(∗) if x′ ≪ x ≤ nz in Cu(A), then there exist y1, . . . , yn ∈ Cu(A) such that x′ ≪ Pn
yi ≤ x, and yi ≤ z,
i=1 yi,
which follows from Property (P1) of the Cuntz semigroup stated in the previous section.
Take first k = 1. The existence of y(i), with i = 1, . . . , n, satisfying (b), (d) and (e)
follows from (∗) applied to u′ ≪ u′′ ≤ nx1. The existence of y(i) ≪ y(i) satisfying (a)
and (c) then follows from Axiom (A2) of the Cuntz semigroup from the previous section.
Assume that 2 ≤ k ≤ m and that y(i1, i2, . . . , ik−1) and y(i1, i2, . . . , ik−1) have been found.
The existence of y(i1, . . . , ik−1, i), with i = 1, . . . , n, satisfying (b), (d) and (e) follows from
(∗) applied to
y(i1, . . . , ik−1) ≪ y(i1, . . . , ik−1) ≤ nxk.
(To see that the latter inequality holds, note that y(i1, . . . , ik−1) ≪ u′′, which follows by
repeated use of (e).) The existence of y(i1, i2, . . . , ik) satisfying (a) and (c) follows from
Axiom (A2).
We claim that the nm elements (cid:0)y(i1, . . . , im)(cid:1) witness the weak (m, nm)-divisibility of
u. Indeed, it follows from (d) and (e) that y(i1, . . . , im) ≤ xj for all j = 1, . . . , m, whence
m·y(i1, . . . , im) ≤ x1 + x2 + · · · + xm ≤ u.
It follows from (b) and (c) that the sum of the elements y(i1, . . . , im) is greater than or
equal to u′.
If any of the divisibility numbers Divm(u, A), ∂ivm(u, A), and divm(A) is less than m, then
u (or a multiple of u) must be properly infinite, as shown below. We shall pursue this and
related questions in more detail in Section 9.
Proposition 3.4. Let A be a C ∗-algebra and let u ∈ Cu(A).
(i) If u is properly infinite, then Divm(u, A) = 1 for all integers m ≥ 1.
(ii) If 1 ≤ n < m are integers and if u is either (m, n)-divisible, (m, n)-decomposable or
weakly (m, n)-divisible, then nu is properly infinite, i.e., nu = 2nu.
(iii) If 1 ≤ n < m are integers and if u is compact and (m, n)-divisible, then u is properly
infinite.
7
Proof. (i). If u is properly infinite, then mu ≤ u for all m, whence Divm(u, A) = 1 .
(ii). Assume that u is weakly (m, n)-divisible and take u′ ≪ u. Then there exist
x1, . . . , xn such that mxi ≤ u for all i, and u′ ≤ Pn
n
i=1 xi. Thus,
mu′ ≤
Xi=1
mxi ≤ nu.
As this holds for all u′ ≪ u, we get ((m − n) + n)u = mu ≤ nu. This entails that
(k(m − n) + n)u ≤ nu for all positive integers k, whence ℓu ≤ nu for all ℓ ≥ n.
In
particular, 2nu ≤ nu, which implies that nu is properly infinite.
Next, suppose that u is (m, n)-decomposable and let u′ ≪ u. Then there exist
x1, . . . , xm such that Pm
i=1 xi ≤ u and u′ ≤ nxi for all i. Thus,
mu′ ≤ n
m
Xi=1
xi ≤ nu.
Arguing as before, we conclude that nu is properly infinite.
Finally note that if u is (m, n)-divisible, then it is both (m, n)-decomposable and weakly
(m, n)-divisible, whence nu is properly infinite.
(iii). Since Divm(u, A) = n < m and u ≪ u, there exists x such that mx ≤ u ≤ nx.
Arguing as above this implies that nx is properly infinite.
Remark 3.5. By functoriality, each ∗-homomorphism ϕ : A → B between C ∗-algebras A
and B induces a morphism Cu(ϕ) : Cu(A) → Cu(B) which preserves order, addition, and
the relation of compact containment. Thus, for each u ∈ Cu(A), and with v = Cu(ϕ)(u),
we have:
Divm(v, B) ≤ Divm(u, A),
∂ivm(v, B) ≤ ∂ivm(u, A),
divm(v, B) ≤ divm(u, A).
In particular, if A and B are unital C ∗-algebras, and if Divm(h1Bi, B) > Divm(h1Ai, A)
for some m (or if the corresponding inequality holds for one of the other two divisibility
numbers), then one can not find a unital embedding of A into B. Divisibility numbers thus
serve as an obstruction for embedding a unital C ∗-algebra with nice divisibility properties
into a unital C ∗-algebra with less nice divisibility properties.
The three divisibility properties behave well with respect to inductive limits thanks to the
sequential continuity of the functor Cu(·):
Proposition 3.6. Let A = lim−→(Ai, ϕi,j) be a sequential inductive limit of C ∗-algebra. Let
u ∈ Cu(A1) and, for each i, denote by ui ∈ Cu(Ai) and u∞ ∈ Cu(A) the images of u in
Cu(Ai) and Cu(A), respectively. Then:
Divm(u∞, A) ≤ inf
i
Divm(ui, Ai),
∂ivm(u∞, A) ≤ inf
i
∂ivm(ui, Ai),
If u is compact (i.e., if u ≪ u), then the above inequalities are equalities.
divm(u∞, A) ≤ inf
i
divm(ui, Ai).
8
Proof. We will only prove the statements above in the former case; the proofs for the two
other cases are similar.
The inequalities Divm(u∞, A) ≤ Divm(ui, Ai), with i = 1, 2, . . . , follow from Re-
mark 3.5. Suppose now that u is compact. Set Divm(u∞, A) = n. Then there exists
x ∈ Cu(A) such that mx ≤ u∞ ≤ nx. By Proposition 2.1 (i) and compactness of u∞,
it follows that x is the image of some y ∈ Cu(Ai) for some i. By Axiom (A2) of the
Cuntz semigroup and by compactness of u∞ there exists y′ ≪ y in Cu(Ai) such that
u∞ ≤ n Cu(ϕi,∞)(y′). Since the ui's are compact, Proposition 2.1 (ii) implies that there
exists j > i such that
m Cu(ϕi,j)(y′) ≤ uj ≤ n Cu(ϕi,j)(y′).
Thus uj is (m, n)-divisible in Cu(Aj).
Definition 3.7. Let A be a σ-unital C ∗-algebra. Then A contains a strictly positive
element. This element represents a class in Cu(A), which is independent of the choice of
the strictly positive element, and which we shall denote by hAi. If A is unital, then hAi =
h1Ai. We shall write divm(A), ∂ivm(A), and Divm(A) for divm(hAi, A), ∂ivm(hAi, A), and
Divm(hAi, A), respectively.
If A and B are unital C ∗-algebras such that there exist unital ∗-homomorphisms A → B
and B → A, then, by Remark 3.5, we must have
Divm(A) = Divm(B),
∂ivm(A) = ∂ivm(B),
divm(A) = divm(B)
for all m. This applies in particular to the situation where A is any unital C ∗-algebra and
B = A ⊗ D for some unital C ∗-algebra D which has a character. In general, if D is any
unital C ∗-algebra, possibly without characters, the divisibility numbers associated with
A ⊗ D are smaller than or equal to those of A.
Examples and remarks
Let us first examine the divisibility numbers for matrix algebras:
Example 3.8. Let m ≥ 2 and k ≥ 2 be integers. Using that
an elementary algebraic argument yields that
(cid:0)Cu(Mk(C)), h1i(cid:1) ∼= (cid:0){0, 1, 2, 3, . . . , ∞}, k(cid:1),
divm(Mk(C)) = ∂ivm(Mk(C)) = Divm(Mk(C)),
and
Here ⌈·⌉ and ⌊·⌋ are the "ceiling" and "floor" functions. In particular, Divm(Mk(C)) = m
if and only if m k, and Divm(Mk(C)) = m + 1 if m∤ k and m(m − 1) ≤ k.
m ⌋(cid:25) ,
if m ≤ k,
if m > k.
(3.1)
(cid:24) k
Divm(Mk(C)) =
⌊ k
∞,
9
Definition 3.9 (The rank of a C ∗-algebra). Let A be a C ∗-algebra. Let rank(A) denote
the smallest positive integer n for which A has an irreducible representation on a Hilbert
space of dimension n, and set rank(A) = ∞ if A has no finite dimensional (irreducible)
representation.
Note that rank(A) = 1 if and only if A has a character. We remind the reader about the
following classical result due to Glimm:
Proposition 3.10 (Glimm). Let A be a (not necessarily unital) C ∗-algebra. Then there is
a non-zero ∗-homomorphism CMn(C) → A if and only if A admits at least one irreducible
representation on a Hilbert of dimension ≥ n.
It follows from Remark 3.5 that divm(A), ∂ivm(A), and Divm(A) are greater than or equal
to Divm(Mn(C)) if rank(A) = n. In particular, these three quantities are infinite when
m > rank(A).
Example 3.11 (Simple C ∗-algebras). If A is a simple, unital, infinite dimensional C ∗-al-
gebra, then Divm(A), ∂ivm(A), and divm(A) are finite for all positive integers m. Indeed,
by the assumption that A is infinite dimensional, it follows that there is a non-zero x ∈
Cu(A) such that mx ≤ h1Ai. As every simple unital C ∗-algebra is algebraically simple,
it follows that h1Ai ≤ nx for some positive integer n, i.e., h1Ai is (m, n)-divisible. Hence
Divm(A) ≤ n, which entails that also ∂ivm(A) ≤ n and divm(A) ≤ n.
Example 3.12. The dimension drop C ∗-algebra Zp,q, associated with the positive integers
p and q, is defined to be
Zp,q = {f ∈ C([0, 1], Mp ⊗ Mq f (0) ∈ Mp ⊗ C1q, f (1) ∈ C1p ⊗ Mq}.
Note that rank(Zp,q) = min{p, q}. It was shown in [Rør04, Lemma 4.2] (and its proof)
that Divm(Zm,m+1) = m + 1. By Remark 3.5, it follows that if Zm,m+1 maps unitally into
A, then Divm(A) ≤ m + 1. Moreover, as shown in [RW10, Proposition 5.1], if A is a unital
C ∗-algebra of stable rank one, then Divm(A) ≤ m + 1 if and only if Zm,m+1 maps unitally
into A.
Remark 3.13 (Almost divisibility). The property "almost divisibility" of a C ∗-algebra is
expressed by saying that Divm(A) ≤ m + 1 for all integers m ≥ 1. If every dimension drop
algebra Zm,m+1 maps unitally into A, or if the Jiang-Su algebra maps unitally into A, then
A is almost divisible.
Remark 3.14 (Non-embeddability of the Jiang-Su algebra). It was shown in [DHTW09]
that there is a simple unital infinite dimensional nuclear C ∗-algebra A such that the dimen-
sion drop C ∗-algebra Z3,4, and hence the Jiang-Su algebra Z, do not embed unitally into A.
The divisibility properties of A were not explicitly mentioned in [DHTW09], but it is easily
seen (using Lemma 6.1, that is paraphrased from [Rør04, Lemma 4.3]) that Div3(A) > 4.
We shall in Section 7 give further examples of simple unital infinite dimensional C ∗-algebras
where the divisibility numbers attain non-trivial values.
10
Remark 3.15 (Real rank zero C ∗-algebras). It was shown in [PR04, Proposition 5.7] that
if A is a unital C ∗-algebra of real rank zero then A admits a unital embedding of a finite
dimensional C ∗-algebra of rank at least n if and only if rank(A) ≥ n. Combining this with
Remark 3.8 we see that Divm(A) ≤ m + 1 whenever A is a unital C ∗-algebra of real rank
zero and with rank(A) ≥ m(m − 1). In particular, every unital C ∗-algebra A of real rank
zero and with rank(A) = ∞ is almost divisible.
Kirchberg considered in [Kir06] a covering number of a unital C ∗-algebra B. Let us recall
the definition:
Definition 3.16 (Kirchberg). Let m ∈ N. The covering number of a unital C ∗-alge-
bra B, denoted by cov(B, m), is the least positive integer n such that there exist finite
dimensional C ∗-algebras F1, F2, . . . , Fn with rank(Fi) ≥ m, ∗-homomorphisms ϕi : CFi →
B, and d1, d2, . . . , dn ∈ B such that 1B = Pn
i=1 d∗
i ϕi(1Fi ⊗ ι)di.
Kirchberg's covering number cov(B, m) relates to our divm(B) as follows.
Proposition 3.17. Let B be a unital C ∗-algebra and let m be a positive integer.
(i) cov(B, m) is the least n for which there exist x1, x2, . . . , xn ∈ Cu(B) such that
xi ≤ h1Bi ≤ x1 + x2 + · · · + xn,
xi =
ki
Xj=i
mijyij
(3.2)
for some integers mij ≥ m, some positive integers ki, and some yij ∈ Cu(A).
(ii) cov(B, m) ≤ divm(B) ≤ (2m − 1) cov(B, m).
Proof. (i). Assume that n ≥ cov(B, m) and let Fi, ϕi : CFi → B, and di ∈ B be as in
j=1 Mmij (C) with mij ≥ m. Let e(ij) be a one-dimensional
projection in Mmij (C). It then follows from Lemma 2.3 that the elements
Definition 3.16. Write Fi = Lki
xi = hϕi(1Fi ⊗ ι)i,
yij = hϕi(e(ij) ⊗ ι)i
satisfy the relations in (3.2).
Cu(B) satisfying (3.2). Put Fi = Lki
Suppose, conversely, that n ≥ 1 is chosen such that there are elements xi and yij in
j=i mijyij ≤
h1Bi it follows from Lemma 2.3 (ii) that there are mutually orthogonal positive elements
aijr in B, where 1 ≤ i ≤ n, 1 ≤ j ≤ ki, 1 ≤ r ≤ mij, such that haijri = yij. We can
further assume that the r positive elements aij1, . . . , aijr are pairwise equivalent. It then
follows from the universal property of the cone over a matrix algebra (see Remark 2.2)
that there are ∗-homomorphisms ϕi : CFi → B such that hϕi(e(ij) ⊗ ι)i = yij, where e(ij)
is a one-dimensional projection in the summand Mmij (C) of Fi. The existence of di ∈ B
j=1 Mmij (C). By the assumption that Pki
with 1B = Pn
i=1 d∗
i ϕi(1Fi ⊗ ι)di follows from Lemma 2.3 (i). Thus cov(B, m) ≤ n.
(ii). To prove the first inequality, assume that divm(B) = n < ∞ and take y1, . . . , yn
such that myj ≤ h1Bi ≤ y1 + · · · + yn. Then (3.2) holds with ki = 1 and xi = myi.
11
Assume next that cov(B, m) = n < ∞, and find elements xi and yij satisfying the
relations in (3.2). Upon replacing yij with an integral multiple of yij we can assume that
m ≤ mij < 2m for all i and j. Let zik, 1 ≤ k ≤ 2m − 1, be the sum of a suitable subset of
j=i mijyij = xi. The (2m − 1)n elements (zik) will then
k=1 zik = Pki
the yij's such that P2m−1
witness that divm(B) ≤ (2m − 1)n.
4 The asymptotic divisibility numbers
One can collect the sequence of divisibility numbers (cid:0)Divm(A)(cid:1)∞
A into a single divisibility number as follows:
m=2 of a unital C ∗-algebra
Div∗(A) = lim inf
m→∞
Divm(A)
m
.
In a similar way one can define ∂iv∗(A) and div∗(A). Propositions 4.1 and 4.2 below hold
verbatim for those quantities as well. However, to keep the exposition bounded, we only
treat the case of "Div".
It follows from Proposition 3.4 that Div∗(A) = 0 if and only if A is properly infinite
and that Div∗(A) ≥ 1 if A is not properly infinite.
Proposition 4.1. Let A be a unital C ∗-algebra.
(i) Divm(A) ≤ m Div∗(A) + 1 for all integers m ≥ 2.
(ii) Div∗(A) = limm→∞ Divm(A)/m (the limit always exists, but is possibly equal to ∞).
(iii) If A is not properly infinite, then Div∗(A) = 1 if and only if Divm(A) ≤ m + 1 for
all integers m ≥ 2.
It follows from Proposition 3.4 and from (iii) above, that A is almost divisible if and only
if Div∗(A) ≤ 1 (i.e., if and only if Div∗(A) = 0 or Div∗(A) = 1).
Proof. (i). If Div∗(A) = ∞ there is nothing to prove. Assume that 1 ≤ Div∗(A) < ∞. Let
m ≥ 2 be given. Let L be the smallest integer strictly greater than m Div∗(A). We show
that Divm(A) ≤ L. Choose α > 1 and a positive integer r0 such that
α
r0 + 1
r0
m Div∗(A) ≤ L.
By the definition of Div∗(A) there is k ≥ r0m such that ℓ := Divk(A) ≤ αk Div∗(A). Take
x ∈ Cu(A) such that kx ≤ h1Ai ≤ ℓx. Write k = rm+ d, with 0 ≤ d < m and r ≥ r0. Also,
write ℓ = tr − d′, with 0 ≤ d′ < r and t ≥ 1. Put y = rx ∈ Cu(A). Then my ≤ h1Ai ≤ ty.
12
With ⌈ · ⌉ denoting the ceiling function, we have
Divm(A) ≤ t = ⌈
ℓ
r
⌉
k − d
k
= l ℓ
≤ lα
≤ lα
≤ lα
mm
m Div∗(A)m
m Div∗(A)m
m Div∗(A)m ≤ L.
k − d
r + 1
r0 + 1
r
r0
(ii). It follows from (i) that
lim supm→∞ Divm(A)
m
≤ Div∗(A) =
lim inf m→∞ Divm(A)
m
,
and so the claims follows.
(iii). The "if" part is trivial, and the "only if" part follows from (i).
We proceed to discuss how Div∗( · ) behaves under forming matrix algebras:
Proposition 4.2. Let A be a unital C ∗-algebra.
(i) Div∗(Mn(A)) ≤ Div∗(A) for all integers n ≥ 2.
(ii) If Cu(A) is almost unperforated, then Div∗(Mn(A)) = Div∗(A) for all integers n ≥ 2.
Proof. (i) follows from Remark 3.5 (as A embeds unitally into Mn(A)).
(ii). Assume that Cu(A) is almost unperforated. We show first that
Div∗(A) ≤
n + 1
n − 1
Div∗(Mn(A))
(4.1)
for all n ≥ 2. To see this take any integer m ≥ 2, and use Proposition 4.1 (i) to see that
ℓ := Divm(Mn(A)) ≤ m Div∗(Mn(A)) + 1. Write m = r(n + 1) + d and ℓ = t(n − 1) − d′,
where r and t are positive integers, 0 ≤ d < n + 1, and 0 ≤ d′ < n − 1.
Identify Cu(Mn(A)) with Cu(A) in the canonical way, where h1Mn(A)i ∈ Cu(Mn(A)) is
identified with nh1Ai. Under this identification we can find x ∈ Cu(A) such that mx ≤
nh1Ai ≤ ℓx. In particular,
(n + 1)rx ≤ nh1Ai ≤ (n − 1)tx,
which by the assumption that Cu(A) is almost unperforated implies that rx ≤ h1Ai ≤ tx.
This shows that
Divr(A)
r
≤
t
r
ℓ
n − 1
n − 1m ≤ r−1(cid:0)
= r−1l ℓ
≤ r−1(cid:16)m Div∗(Mn(A)) + 1
+ 1(cid:1)
+ 1(cid:17)
Div∗(Mn(A)) + r−1 n
n − 1
n + 1
n − 1
n − 1
≤
13
Div∗(Mn(A)) + r−1 n
n − 1
.
Now, r → ∞ as m → ∞, and so (4.1) follows by letting m tend to infinity.
To complete the proof of (ii), take n ≥ 2. By (i) and (4.1) we have:
Div∗(A) ≤
kn + 1
kn − 1
Div∗(Mnk(A)) ≤
kn + 1
kn − 1
Div∗(Mn(A))
for all k ≥ 1, which shows that Div∗(A) ≤ Div∗(Mn(A)).
We have previously remarked that Divm(A) = ∞ whenever m > rank(A).
It follows
that Div∗(A) = ∞ whenever rank(A) < ∞, i.e., whenever A admits a non-zero finite
dimensional representation.
Remark 4.3. It can happen that Div∗(Mn(A)) < Div∗(A). Take for example A such
that Mn(A) is properly infinite, but A itself is not properly infinite, cf. [Rør03]. Then
Div∗(Mn(A)) = 0 and Div∗(A) ≥ 1.
It is an important open problem if Div∗(A) ≤ 1 (i.e., if A is almost divisible) for every
(simple) unital infinite dimensional C ∗-algebra A for which Cu(A) is almost unperforated.
5 Finite-, infinite-, and ω-divisibility
The property that any of the divisibility numbers Divm(A), ∂ivm(A), and divm(A) is finite,
when A is a unital C ∗-algebra, has interpretations in terms of structural properties of the
C ∗-algebra A. We have already noted that the divisibility numbers always are finite when
A is a simple C ∗-algebra, and the corresponding structural properties of the C ∗-algebra are,
as we shall see, trivially satisfied for simple C ∗-algebras. The correct definition of "finite
divisibility" in the non-unital case is what we call (m, ω)-divisibility as defined below.
Definition 5.1. Let A be a C ∗-algebra, let u ∈ Cu(A), and let m be a positive integer.
Then:
(i) u is (m, ω)-divisible if for all u′ ∈ Cu(A) with u′ ≪ u there exists x ∈ Cu(A) such
that mx ≤ u and u′ ≤ nx for some positive integer n.
(ii) u is (m, ω)-decomposable if for all u′ ∈ Cu(A) with u′ ≪ u there exist elements
x1, x2, . . . , xm ∈ Cu(A) such that x1 + x2 + · · · + xm ≤ u and u′ ≤ nxj for some
positive integer n and for all j.
(iii) u is weakly (m, ω)-divisible if for all u′ ∈ Cu(A) with u′ ≪ u there exist elements
x1, x2, . . . , xn in Cu(A) such that mxj ≤ u for all j and u′ ≤ x1 + x2 + · · · + xn.
Remark 5.2. If u in Definition 5.1 is compact, then u is (m, ω)-divisible, (m, ω)-decompo-
sable, respectively, weakly (m, ω)-divisible if and only if Divm(u, A) < ∞, ∂ivm(u, A) < ∞,
respectively, divm(u, A) < ∞, cf. Remark 3.2.
14
In the next result we express (m, ω)-divisibility in terms of structural properties of the
C ∗-algebra. Part (iii) is almost contained in [Kir06] (see [Kir06, Definition 3.1] and [Kir06,
Remark 3.3 (7)] and compare with Definition 3.16 and Proposition 3.17). Recall the defi-
nition of the rank of a C ∗-algebra from Definition 3.9.
Theorem 5.3. Let A be a σ-unital C ∗-algebra and let e be a strictly positive element of
A. (If A is unital, we can take e to be the unit of A.) Put u = hei = hAi.
(i) u is (m, ω)-divisible if and only if for every ε > 0 there exists a ∗-homomorphism
ϕ : CMm(C) → A such that (e − ε)+ belongs to the closed two-sided ideal generated
by the image of ϕ.
(ii) u is (m, ω)-decomposable if and only if for every ε > 0 there exist mutually orthogonal
positive elements b1, b2, . . . , bm in A such that (e − ε)+ belongs to the closed two-sided
ideal generated by bi for each i.
(iii) The following are equivalent:
(a) u is weakly (m, ω)-divisible,
(b) rank(A) ≥ m,
(c) there exist ∗-homomorphisms ϕi : CMm(C) → A, i = 1, 2, . . . , n, for some n,
such that (e − ε)+ belongs to the closed two-sided ideal generated by the union
of the images of the ϕi's.
Proof. (i). Let us assume that u is (m, ω)-divisible. Let ε > 0. Find x ∈ Cu(A) and a
positive integer n such that mx ≤ u and h(e − ε/2)+i ≤ nx. Choose a positive element
a in A ⊗ K such that x = hai, and choose η > 0 such that h(e − ε)+i ≤ nh(a − η)+i.
By Lemma 2.4 there exists ϕ : CMm(C) → A such that hϕ(e11 ⊗ ι)i = h(a − η)+i. Then
h(e − ε)+i ≤ nhϕ(e11 ⊗ ι)i which implies that (e − ε)+ belongs to the closed two-sided ideal
generated by the image of ϕ.
Suppose conversely that for every ε > 0 there exists ϕ : CMm → A such that (e − ε)+
is in the closed two-sided ideal generated by ϕ(e11 ⊗ ι). Set hϕ(e11 ⊗ ι)i = x. Then mx ≤ u
by Lemma 2.4, while h(e − 2ε)+i ≤ nx for some positive integer n. This shows that u is
(m, ω)-divisible.
(ii). "Only if". Let ε > 0 and suppose that b1, b2, . . . , bm in A exist with the stipulated
properties. Set hbji = xj ∈ Cu(A). Then
x1 + x2 + · · · + xm = hb1 + b2 + · · · + bmi ≤ u.
Since (e − ε)+ belongs to the closed two-sided ideal generated by bj, h(e − 2ε)+i ≤ nxj for
some integer n ≥ 1. It follows that u is (m, ω)-decomposable.
"If". If u = hei is (m, ω)-decomposable and if ε > 0, then there are positive elements
a1, a2, . . . , am in A ⊗ K such that ha1i + ha2i + · · · + hami ≤ u and h(e − ε/2)+i ≤ nhaji
for some positive integer n. Choose η > 0 such that h(e − ε)+i ≤ nh(aj − η)+i for all j.
By Lemma 2.3 (ii) there are pairwise orthogonal positive elements b1, b2, . . . , bm in A such
15
that bj ∼ (aj − η)+. Then the closed two-sided ideal generated by bj contains (e − ε)+ for
each j.
(iii). (a) ⇒ (b). Assume that u is weakly (m, ω)-divisible. Suppose that A has an
irreducible representation π : A → B(Ck) = Mk(C) of finite positive dimension k. Then
π is necessarily surjective. Since (m, ω)-divisibility is preserved by ∗-homomorphisms (cf.
Remark 3.5), we conclude that Mk(C) is weakly (m, ω)-divisible. But then k ≥ m, cf.
Example 3.8. Hence (b) holds.
(b) ⇒ (c). Assume that (b) holds. Let (ϕi)i∈I be the family of all non-zero ∗-ho-
momorphisms ϕi : CMm(C) → A and let I be the closed two-sided ideal in A generated
by the images of all ϕi's. Thus each ϕi maps CMm(C) into I. We claim that I = A.
Assume, to reach a contradiction, that I 6= A. By the assumption that rank(A) ≥ m,
all irreducible representations of A/I have dimension at least m. It follows from Glimm's
lemma (Proposition 3.10) that there is a non-zero ∗-homomorphism CMm(C) → A/I,
which by projectivity lifts to a ∗-homomorphism ϕ : CMm(C) → A. But the image of ϕ is
not contained in I, which is a contradiction.
For each finite subset F of I consider the closed two-sided ideal IF of A generated by
Si∈F ϕi(CMm(C)). Then A is the closure of the union of the upwards directed family of
ideals (IF ). Hence, for each ε > 0, there is a finite subset F of I such that (e − ε)+ belongs
to IF . Thus (c) holds.
(c) ⇒ (a). Assume that (c) holds. Set zi = hϕi(e11 ⊗ ι)i for i = 1, 2, . . . , n. Then
mzi ≤ u for all i. Moreover, (a − ε)+ belongs to the algebraic ideal generated by the n
elements ϕi(e11 ⊗ ι), whence h(a − ε)+i ≤ Pn
j=1 njzj for suitable positive integers nj. Put
N = P nj and let x1, x2, . . . , xN be a listing of the elements z1, . . . , zn, with zj repeated
nj times. Then mxj ≤ u and (a − ε)+ ≤ x1 + x2 + · · · + xN . This shows that u is weakly
(m, ω)-divisible.
The theorem above can be simplified in the case where u is compact, and in particular in
the case where A is unital:
Corollary 5.4. Let A be a unital C ∗-algebra, and let m be a positive integer. Then:
(i) Divm(A) < ∞ if and only if there exists a ∗-homomorphism ϕ : CMm(C) → A whose
image is full in A.
(ii) ∂ivm(A) < ∞ if and only if there exist full, pairwise orthogonal positive elements
b1, b2, . . . , bm in A.
(iii) The following are equivalent:
(a) divm(A) < ∞,
(b) rank(A) ≥ m,
(c) there exist ∗-homomorphisms ϕi : CMm(C) → A, i = 1, 2, . . . , n for some n,
such that the union of their images is full in A.
Propositions 3.10 (Glimm), Proposition 3.3, and Corollary 5.4 (i) immediately imply:
16
Corollary 5.5. If A is a unital, infinite dimensional, simple C ∗-algebra, then the three
divisibility numbers Divm(A), ∂ivm(A), and divm(A) are finite for every integer m ≥ 1.
Let us also note what it means to have infinite divm( · ) numbers:
Corollary 5.6. Let A be a unital C ∗-algebra.
(i) A admits a character if and only if div2(A) = ∞.
(ii) A admits a finite dimensional representations if and only if divm(A) = ∞ for some
integer m ≥ 2.
Remark 5.7 (The Global Glimm Halving Problem). Glimm's lemma (Proposition 3.10)
says that there exists a non-zero ∗-homomorphism from CMn(C) into a C ∗-algebra A if and
only if A admits an irreducible representation of dimension at least n. It is not known how
"large" one can make the image of such a ∗-homomorphism. In particular, it is not known
for which C ∗-algebras A one can find a ∗-homomorphism CMn(C) → A whose image is full
in A (i.e., the image is not contained in any proper closed two-sided ideal in A). For n = 2
this problem is known as the "Global Glimm Halving Problem" (see [BK04a], [BK04b] and
[KR02]). A unital C ∗-algebra A is said to have the Global Glimm Halving Property if there
is a ∗-homomorphism CM2(C) → A with full image.
More specifically, one can ask if any (unital) C ∗-algebra, which admits no finite di-
mensional representation, satisfies the Global Glimm Halving Property. In view of Corol-
lary 5.4, this problem for unital C ∗-algebras A may be restated as follows: Does Div2(A) =
∞ imply that divm(A) = ∞ for some positive integer m? For a non-unital C ∗-algebra A,
the one can restate the problem in the following way: Does Div2(A) = ∞ imply that hAi
fails to be (m, ω)-divisible for some positive integer m.
It is shown in [KR02] that if A is a weakly purely infinite C ∗-algebra, then A is purely
infinite if and only if all hereditary sub-C ∗-algebras of A have the Global Glimm Halving
(It is easy to see that the rank of any weakly purely infinite C ∗-algebra is
Property.
infinite.) It is an open problem if all weakly purely infinite C ∗-algebras are purely infinite.
Remark 5.8. Let A be a unital C ∗-algebra. It follows from Proposition 3.3 (and also from
Corollary 5.4) that
Divm(A) < ∞ =⇒ ∂ivm(A) < ∞ =⇒ divm(A) < ∞
for all positive integers m. None of the two reverse implications hold in general.
For each integer m ≥ 2 consider the Bott projection p in C(S2m) ⊗ K. This projection
has (complex) dimension m and it has no non-trivial sub-projections. The unital C ∗-al-
gebra A = p(C(S2m) ⊗ K)p is a homogeneous C ∗-algebra of rank m. Hence divm(A) <
∞. Suppose that ∂ivm(A) < ∞. Then, by Corollary 5.4 (ii), there would exist full,
pairwise orthogonal, positive elements b1, . . . , bm in A. This would entail that each bj is
one-dimensional in each fiber of A, and hence that fj.bj is a one-dimensional projection
for some fj ∈ C(S2m). But this contradicts the fact that p has no proper subprojections.
17
To see that ∂ivm(A) < ∞ does not imply Divm(A) < ∞, consider the C ∗-algebra
B = C(S2)⊗K, and take the (one-dimensional) Bott projection p ∈ B and a trivial (m−1)-
dimensional projection q ∈ B such that p and q are orthogonal. Put A = (p + q)B(p + q).
It follows from a K-theoretical argument that p + q cannot be written as the sum of m
pairwise orthogonal and equivalent projections (because [p + q] is not divisible by m in
K0(A)). The unit of A can be written as the sum of m (necessarily full) projections, so
∂ivm(A) < ∞. Assume that Divm(A) < ∞. Then, by Corollary 5.4 (i), there is a ∗-ho-
momorphism ϕ : CMm(C) → A whose image is full in A. As explained in Remark 2.2,
this entails that there exist full, pairwise orthogonal, pairwise equivalent, positive elements
b1, b2, . . . , bm in A. Arguing as in the paragraph above, we can assume that each bj is in
fact a projection. But that contradicts the fact that 1A = p+q is not the sum of m pairwise
equivalent projections.
6 Divisibility and comparability
Let A and B be C ∗-algebras. Then there is a natural bi-additive map
Cu(A) × Cu(B) → Cu(A ⊗ B),
(x, y) 7→ x ⊗ y,
defined as follows: If x = hai and y = hbi with a a positive element in A ⊗ K and b a
positive element in B ⊗ K, then x ⊗ y = ha ⊗ bi, where we identify (A ⊗ K) ⊗ (B ⊗ K) with
A ⊗ B ⊗ K. Note that x1 ⊗ y1 ≤ x2 ⊗ y2 if x1 ≤ x2 and y1 ≤ y2.
Part (i) of the following result was (implicitly) proved in [Rør04, Lemma 4.3], and was
used to prove that Cu(A ⊗ Z) is almost unperforated for all unital C ∗-algebras A.
Lemma 6.1. Let A and B be unital C ∗-algebras and let 1 ≤ m < n be integers.
(i) Let x, y ∈ Cu(A) be such that nx ≤ my. If B is (m, n)-divisible, then x ⊗ h1Bi ≤
y ⊗ h1Bi.
(ii) Let x1, x2, . . . , xm, y ∈ Cu(A) be such that nxj ≤ y for all j. If B is (m, n)-decompo-
sable, then
(x1 + x2 + · · · + xm) ⊗ h1Bi ≤ y ⊗ h1Bi.
(iii) Let x, y1, y2, . . . , yn ∈ Cu(A) be such that x ≤ myj for all j. If B is weakly (m, n)-
divisible, then
x ⊗ h1Bi ≤ (y1 + y2 + · · · + yn) ⊗ h1Bi.
Proof. (i). Take z ∈ Cu(B) such that mz ≤ h1Bi ≤ nz. Then
x ⊗ h1Bi ≤ x ⊗ nz = nx ⊗ z ≤ my ⊗ z = y ⊗ mz ≤ y ⊗ h1Bi.
18
(ii). Take z1, z2, . . . , zm ∈ Cu(B) such that z1 + z2 + · · · + zm ≤ h1Bi ≤ nzj. Then
(x1 + x2 + · · · + xm) ⊗ h1Bi ≤ x1 ⊗ nz1 + x2 ⊗ nz2 + · · · + xm ⊗ nzm
= nx1 ⊗ z1 + nx2 ⊗ z2 + · · · + nxm ⊗ zm
≤ y ⊗ (z1 + z2 + · · · + zm)
≤ y ⊗ h1Bi.
(iii). Take z1, z2, . . . , zn ∈ Cu(B) such that mzj ≤ h1Bi ≤ z1 + z2 + · · · + zn. Then
(y1 + y2 + · · · + yn) ⊗ h1Bi ≥ y1 ⊗ mz1 + y2 ⊗ mz2 + · · · + yn ⊗ mzn
= my1 ⊗ z1 + my2 ⊗ z2 + · · · + myn ⊗ zn
≥ x ⊗ (z1 + z2 + · · · + zn)
≥ x ⊗ h1Bi.
The lemma above can loosely be paraphrased as follows: Good divisibility properties of B
ensure good comparability properties of A ⊗ B, and bad comparability properties of A ⊗ B
entail bad divisibility properties of B.
We proceed to show that infinite tensor products of (suitable) unital C ∗-algebras cannot
have very bad comparability properties.
Lemma 6.2. Let (cid:0)Ak(cid:1)∞
∞. Then
k=1 be a sequence of unital C ∗-algebras such that N := supk div2(Ak) <
divm(
∞
Ok=1
Ak) ≤ N n,
for all integers m ≥ 2, where n is any integer satisfying 2n ≥ m.
Proof. Take n such that 2n ≥ m. For each k, find z(k)
2z(k)
i=1 z(k)
i
. Given a multi-index (i1, i2, . . . , in) ∈ {1, . . . , N}n, put
i ∈ Cu(Ak), i = 1, 2, . . . , N, such that
i ≤ h1Aki ≤ PN
zi1,i2,...,in = z(1)
i1 ⊗ z(2)
i2 ⊗ · · · ⊗ z(n)
in ∈ Cu(cid:16)
n
Ok=1
Ak(cid:17).
Then
m · zi1,i2,...,in ≤ 2nzi1,i2,...,in = (2z(1)
ii ) ⊗ (2z(2)
i2 ) ⊗ · · · ⊗ (2z(n)
in )
≤ h1A1 ⊗ 1A2 ⊗ · · · ⊗ 1Ani
i (cid:17) ⊗(cid:16)
z(1)
N
Xi=1
i (cid:17) ⊗ · · · ⊗(cid:16)
z(2)
N
Xi=1
i (cid:17)
z(n)
zi1,i2,...,in.
N
≤ (cid:16)
Xi=1
Xi1,i2,...,in=1
=
N
19
Recall that a C ∗-algebra A has the Corona Factorization Property if and only if all full
projections in M(A ⊗ K) are properly infinite. It was shown in [OPR] that if A is a sepa-
rable C ∗-algebra, then A and all its closed two-sided ideals have the Corona Factorization
Property if and only if for every integer m ≥ 2, and for all x′, x, y1, y2, . . . in Cu(A) such
that x′ ≪ x and x ≤ myj for all j, one has x′ ≤ y1 + y2 + · · · + yN for some integer N ≥ 1.
Proposition 6.3. Let (cid:0)Ak(cid:1)∞
k=1 be a sequence of unital C ∗-algebras such that
sup
div2(Ak) < ∞.
k
It follows that the C ∗-algebra N∞
N∞
Proof. Put B = N∞
k=1 Ak and all its closed two-sided ideals have the Corona
Factorization Property. In particular, if A is a unital C ∗-algebra without characters then
k=1 A and all its closed two-sided ideals have the Corona Factorization Property.
We shall view (Bn)∞
is dense in B, and we shall identify B with Bn ⊗ Dn for all n.
k=1 Ak and Dn = N∞
k=1 Ak, and for each n ≥ 1 put Bn = Nn
n=1 as an increasing sequence of sub-C ∗-algebras of B such that S∞
k=n+1 Ak.
n=1 Bn
Let m ≥ 1 be an integer and let x′, x, y1, y2, y3, . . . in Cu(B) be such that x′ ≪ x and
x ≤ myj for all j. By Lemma 6.2 there is a positive integer N such that divm(Dn) ≤ N
for all n. We show that x′ ≤ y1 + y2 + · · · + yN . This will prove that B has the Corona
Factorization Property.
Repeated use of Proposition 2.1 (i) and (ii) shows that there exists a positive integer
n, and elements x′′, y′
1, y′
x′ ≤ x′′ ⊗ h1Dni ≤ x,
2, . . . , y′
N in Cu(Bn) such that
y′
j ⊗ h1Dni ≪ yj
in Cu(B);
x′′ ≪ my′
j
in Cu(Bn),
where x 7→ x ⊗ h1Dni denotes the canonical embedding Cu(Bn) → Cu(B). We can now
apply Lemma 6.1 (iii) to deduce that
x′ ≤ x′′ ⊗ h1Dni ≤ (y′
1 + y′
2 + · · · + y′
N ) ⊗ h1Dni ≤ y1 + y2 + · · · + yN
as desired.
7 Obstructions to Divisibility
A trivial obstruction to (weak) divisibility of a C ∗-algebra is its rank: divm(A) < ∞ if
and only if m ≤ rank(A) (by Corollary 5.4 (iii)). In this section we shall discuss ways of
obtaining homogeneous C ∗-algebras with large rank and large weak divisibility constant.
We use these techniques to construct unital simple C ∗-algebras with large weak divisibility
constants.
We remark first that Lemma 6.1 provides non-trivial obstructions to divisibility in B.
Indeed, it follows by that lemma that if there exists a unital C ∗-algebra A and x, y ∈ Cu(A)
such that nx ≤ my but x ⊗ h1Bi (cid:2) y ⊗ h1Bi, then Divm(B) > n. Similarly, if there exist
x1, . . . , xm, y in Cu(A) such that nxj ≤ y for all j while
(x1 + x2 + · · · + xm) ⊗ h1Bi (cid:2) y ⊗ h1Bi,
20
then ∂ivm(B) > n. Finally, if there exist x, y1, . . . , yn ∈ Cu(A) such that x ≤ myj for all j
while
x ⊗ h1Bi (cid:2) (y1 + y2 + · · · + yn) ⊗ h1Bi,
then divm(B) > n.
We introduce below another way to obtain bad divisibility behavior:
Lemma 7.1. Let u, v ∈ Cu(A) be compact elements. If Div2(u + v, A) ≤ N then there
exist x1, x2, . . . , xN in Cu(A) such that 2xi ≤ v for all i and
2v ≤ v + (2N + 1)u +
N
Xi=1
2xi.
Proof. By assumption there exists x such that 2x ≤ u + v ≤ Nx. By compactness of u + v
we can find x′′ ≪ x′ ≪ x such that u + v ≤ Nx′′. Since x′ ≪ x ≤ u + v, it follows from
Property (P1) of the Cuntz semigroup (see Section 2) (leaving u unchanged) that there
exists v1 such that
x′ ≤ u + v1,
1 ≪ v1 such that x′′ ≤ u + v′
v1 ≤ x,
v1 ≤ v.
1. Apply (P2) to the relation
As x′′ ≪ u + v1 there is v′
v′
1 ≪ v1 ≤ v to obtain v2 satisfying
v′
1 + v2 ≤ v ≤ v1 + v2.
By compactness of v we can find v′
2 ≪ v2 such that v ≤ v1 + v′
2. Now,
2 ≪ v2 ≤ u + v ≤ Nx′′ ≤ Nu + Nv′
v′
1,
and so we can use (P1) (leaving Nu unchanged) to find x1, . . . , xN such that
v′
2 ≤ Nu +
N
Xj=1
xj,
xj ≤ v′
1,
xj ≤ v2.
It follows that 2xj ≤ v′
1 + v2 ≤ v and that
2v ≤ 2v1 + 2v′
2 ≤ 2x + 2Nu + 2
N
Xj=1
xj ≤ v + (2N + 1)u + 2
N
Xj=1
xj,
as desired.
The corollary below illustrates how the preceding lemma can be used to find elements with
bad divisibility properties:
Corollary 7.2. Let X be a compact Hausdorff space and suppose that p ∈ C(X) ⊗ K is
a projection such that [1] (cid:2) (2N + 1)[p] in K0(C(X)), where 1 denotes the unit of C(X).
Then Div2(h1i + hpi, C(X)) > N.
21
Proof. Consider the compact elements u = hpi and v = h1i of Cu(C(X)). Note that
2x ≤ h1i implies x = 0 and that 2h1i (cid:2) h1i + (2N + 1)hpi. The desired conclusion now
follows from Lemma 7.1.
The relation [1] (cid:2) (2N + 1)[p] in K0(C(X)) is satisfied whenever the (2N + 1)-fold direct
sum of p with itself is a projection with non-trivial Euler class (as explained in more detail
below). It is known that for each integer d ≥ 1 and for each positive integer N there exist
X and p ∈ C(X) ⊗ K such that (2N + 1)[p] has non-trivial Euler class and p has rank d.
The unital C ∗-algebra A = (p ⊕ 1)(cid:0)C(X) ⊗ K(cid:1)(p ⊕ 1) with this choice of X and p will then
satisfy rank(A) = d and Div2(A) > N.
We will now give a different method for constructing elements with large divisibility
constants and large rank, where we get upper bounds and where we also can give sharper
lower bounds for the weak divisibility constant. Let S2 denote the 2-dimensional sphere.
Let p denote the "Bott-projection" in C(S2) ⊗ M2 ⊆ C(S2) ⊗ K, i.e., the projection
associated to the Hopf line bundle over S2. For each 1 ≤ j ≤ N, let pj ∈ C((S2)N ) ⊗ K be
given by
pj(x1, x2, . . . , xN ) = p(xj),
(x1, . . . , xN ) ∈ (S2)N .
Since h1i ≤ 2hpi in Cu(C(S2)), we have
Nh1i ≤ 2D
N
Mi=1
piE.
As another obstruction to weak divisibility, we shall use the following corollary of Lemma 6.1
(iii), cf. the remarks at the beginning of this section, applied to the relations h1i ≤ 2hpii,
i = 1, 2 . . . , N.
Corollary 7.3. Let X be a locally compact Hausdorff space, and let q ∈ C0(X) ⊗ K be a
projection. Let (pi)N
i=1 be the projections in C((S2)N )⊗K defined in the preceding paragraph.
Suppose that
q ⊗ 1 - q ⊗
N
Mi=1
pi.
Then div2(hqi, C0(X)) > N.
Let us now give examples of projections to which the corollary above can be applied.
We will make use of characteristic classes of vector bundles. Recall that projections in
C(X) ⊗ K, with X compact and Hausdorff, give rise to vector bundles over X:
if p is
a projection, then ηp = (Ep, X, π), with Ep = {(x, v) ∈ X × l2(N) p(x)v = v} is the
vector bundle associated to p. Up to Murray-von Neumann equivalence of projections
and isomorphism of vector bundles, this correspondence is a bijection. We denote by
e(ηp) ∈ H ∗(X), or simply e(p), the Euler class of ηp. For the cartesian product of spheres
(S2)N we have (e.g., by the Kunneth formula) that
H ∗((S2)N ) ∼= C[z1, z2, . . . , zN ]/(z2
N ).
1, z2
2, . . . , z2
22
With this identification, the Euler classes of the projections pi ∈ C((S2)N ) ⊗ K defined
earlier can be shown to be e(pi) = zi.
Proposition 7.4. Let X be a compact Hausdorff space and let q ∈ C(X)⊗K be a projection
such that e(q)N 6= 0. Then
div2(h1 ⊕ qi, C(X)) > N.
Proof. By Corollary 7.3 it suffices to show that
(1X ⊕ q) ⊗ 1(S2)N - (1X ⊕ q) ⊗
N
Mi=1
pi
(7.1)
in Cu(C(X) ⊗ C((S2)N )), where 1X denotes the unit in C(X). (In the formulation of the
proposition above, we denoted 1X simply by 1.) Observe that the trivial rank 1 projection
is a subprojection of the projection on the left-hand side of (7.1). Thus, it suffices to show
that the right side of (7.1) has non-zero Euler class.
Set rank(q) = k. For each positive integer i, let ci(q) ∈ H 2i(X) denote the ith char-
acteristic class of q (so that ck(q) = e(q)). By the Kunneth Theorem ([MS74, Theorem
A.6]), we can identify H ∗(X × (S2)N ) with H ∗(X) ⊗ H ∗((S2)N ). Then
e(cid:0)(1X ⊕ q) ⊗
N
Mi=1
pi(cid:1) = e(cid:0)1X ⊗
N
N
Mi=1
N
N
Mi=1
pi(cid:1)
e(q ⊗ pi)
pi(cid:1) e(cid:0)q ⊗
Yi=1
Xj=0
ck−j(q)e(pi)j
e(pi) e(q)N 6= 0.
e(1X ⊗ pi)
e(pi)
k
N
Yi=1
=
=
=
N
Yi=1
Yi=1
Yi=1
N
In the above computation we have used that e(q⊗p) = Pk
j=0 ck−j(q)e(p)j, for q a projection
of rank k and p a projection of rank 1. To obtain the last equality we have used that
e(pi)2 = 0 for all i.
Let us now give examples of families of projections to which the above proposition can be
applied. We shall here and in the following, whenever p is a projection (in a C ∗-algebra)
and n is a positive integer, let n·p denote the n-fold direct sum, p ⊕ p ⊕ · · · ⊕ p, (in a matrix
algebra over the given C ∗-algebra) of the projection p.
Example 7.5. Let N be a positive integer, and let CPN denote the 2N-dimensional
complex projective space. Let η denote the tautological line bundle over CPN and pη the
rank 1 projection associated to it. It is known that e(pη) = z2 ∈ C[z2]/(z2N ), where we
23
have identified H ∗(CPN ) with C[z2]/(z2N ). Let d, d′ be positive integers such that dd′ < N.
Then e(d·pη)d′ = z2dd′
6= 0. It follows that
(cid:22)N − 1
d (cid:23) < div2(h1 ⊕ d·pηi, C(CPN )) ≤ Div2(h1 ⊕ d·pηi, C(CPN )) ≤ (cid:24)N + d + 1/2
⌊d/2⌋
(cid:25) .
Indeed, the first inequality follows from Proposition 7.4 and the calculations made above.
The second inequality follows from Proposition 3.3. The last inequality can be proved as
follows: Put x = ⌊d/2 ⌋hpηi. Then 2x ≤ h1 ⊕ d · pηi. By a classical result about vector
bundles (see [Hus94, Chapter 9, Proposition 1.1]) we have that 1 - k ·pη if 2N ≤ 2k − 1.
It follows that
or, equivalently, that h1 ⊕ d·pηi ≤ nx, if n ⌊d/2 ⌋ ≥ N + d + 1/2.
1 ⊕ d·pη - n ⌊d/2 ⌋ pη,
Simple C ∗-algebras with bad divisibility
In this and the following two subsections we give examples of unital simple C ∗-algebras
with bad divisibility behaviour. They use the Euler class obstruction described in the
following example.
Example 7.6. Let d be a positive integer. Following the notation in [Rør03], for each set
I = {i1, i2, . . . , ik} ⊆ {1, 2, . . . , d}, let pI be the one-dimensional projection in C((S2)d)⊗K
given by
pI(x) = pi1(x) ⊗ pi2(x) ⊗ · · · ⊗ pik (x),
x ∈ (S2)d,
where pi is as defined above Corollary 7.3. It is shown in [Rør03, Proposition 4.5] that if
I1, I2 . . . , Ir are subsets of {1, 2, . . . , d} that admit a matching (i.e., Si∈F Ii ≥ F for all
subset F of {1, 2, . . . , r}) then the Euler class of pI1 ⊕ pI2 ⊕ · · · ⊕ pIr is non-zero.
The examples constructed in this and the following two subsectons are built on the same
template described in the following lemma (which is a variation of one of Villadsen's con-
structions). We retain the terminology from the example above throughout the rest of this
section.
j=1 be a sequence of pairwise disjoint subsets of N. Choose dn large
enough so that all Jj, j = 1, 2, . . . , n, are contained in the set {1, 2, . . . , dn}. (This, of
j=1 Jj.) Consider the projection qn of rank
Lemma 7.7. Let (cid:0)Jj(cid:1)∞
course, can be accomplished by taking dn = Pn
2n in C((S2)dn) ⊗ K given by
qn = 1 ⊕ pJ1 ⊕ 2·pJ2 ⊕ · · · ⊕ 2n−1·pJn.
It follows that there is a simple unital AH-algebra A which is the inductive limit of the
sequence
q1(cid:0)C((S2)d1) ⊗ K(cid:1)q1
where the connecting mappings ϕn are unital.
ϕ1
/ q2(cid:0)C((S2)d2) ⊗ K(cid:1)q2
ϕ2
/ · · ·
/ A,
24
/
/
/
Proof. Set Xn = (S2)dn and An = qn(cid:0)C(Xn) ⊗ K(cid:1)qn. Write
Xn+1 = Xn × (S2)dn+1−dn,
let πn : Xn+1 → Xn be the projection mapping, and let πm,n : Xn → Xm denote the
composition map πm ◦ πm+1 ◦ · · · ◦ πn−1. Choose xn ∈ Xn for each n such that the set
{πm,n(xn) n ≥ m} is dense in Xm for all m ≥ 1.
Define a ∗-homomorphism ϕ0
n : C(Xn, K) → C(Xn+1, K) by
ϕ0
n(f )(x) = f (πn(x)) ⊕ (cid:0)f (xn) ⊗ pJn+1(x)(cid:1),
f ∈ C(Xn, K),
x ∈ Xn+1,
where we in a suitable way have identified K ⊕ (K ⊗ K) with a subalgebra of K. We also
identify C(X) ⊗ K with C(X, K). We make another identification: if
J ⊆ {1, 2, . . . , dn} ⊆ {1, 2, . . . , dn+1},
then the projection pJ is defined both in C(Xn) ⊗ K and in C(Xn+1) ⊗ K and pJn = pJn ◦ πn
(where the former occurrence of pJn is viewed as an element in the former algebra, and the
latter in the latter). We shall use the same notation for the two projections. Taking these
identification a step further, we have qn = qn ◦ πn and that qn+1 = qn ⊕ 2n ·pJn+1. (These
identification hold, strictly speaking, only up to conjugation with an inner automorphism
on K.) In this notation we get
ϕ0
n(qn)(x) = qn(πn(x)) ⊕ (cid:0)qn(xn) ⊗ pJn+1(x)(cid:1)
n(qn) = qn+1 (possibly after composing ϕ0
= qn(πn(x)) ⊕ rank(qn)·pJn+1(x) = qn+1(x),
for all x ∈ Xn+1, i.e., ϕ0
phism on K). This shows that ϕ0
∗-homomorphism that arises in this way, i.e., ϕn is the restriction of ϕ0
co-restriction to An+1), and let A be the inductive limit of the sequence
n with an inner automor-
n maps An unitally into An+1. Let ϕn denote the unital
n to An (and the
A1
ϕ1
/ A2
ϕ2
/ A3
ϕ3
/ · · ·
/ A.
Let ϕm,n : Am → An denote the composition map ϕn−1◦ϕn−2◦· · ·◦ϕm when m ≤ n. One can
check that ϕm,n(f )(x) is non-zero for all x ∈ Xn if f is a function in Am = qmC(Xm, K)qm
which is non-zero on at least one point in the set {πm,k(xk) m ≤ k ≤ n}. By the choice
of the points xn, it follows that for each m and for each non-zero f in Am there is n ≥ m
such that ϕm,n(f ) is full in An (i.e., that ϕm,n(f )(x) 6= 0 for all x ∈ Xn). This entails that
A is simple.
Lemma 7.8. Let N be a positive integer. In the notation of Lemma 7.7 choose the sequence
(Jj)∞
j=1 such that Jj = N ·2n−1. It then follows that
div2(cid:0)hqni, C((S2)dn)(cid:1) > N,
Div2(cid:0)hq2i, C((S2)d2(cid:1) ≤ 3N + 4.
for all n.
25
/
/
/
/
Proof. We use Proposition 7.4 to prove the first claim. It suffices to show that the Euler
class of the projection N·pJ1 ⊕ 2N·pJ2 ⊕ · · · ⊕ 2n−1N·pJn is non-zero. But this follows from
[Rør03, Proposition 4.5], cf. Example 7.6 above, and from the choice of the sets Jn.
To prove the second claim, put x = hpJ2i and note that 2x ≤ hq2i. It follows from
[Dup76, Proposition 1] that q2 - M ·pJ2 if M − 4 ≥ (2d2 − 1)/2 = 3N − 1/2. This shows
that hq2i ≤ (3N + 4)x.
Theorem 7.9. For each positive integer N there exists a simple unital infinite dimensional
AH-algebra A such that N < div2(A) ≤ Div2(A) ≤ 3N + 4.
Proof. Let A be the simple C ∗-algebra constructed in Lemma 7.7 based on the choice of
j=1 made in Lemma 7.8. Then A is the inductive limit of the sequence of C ∗-algebras
(Jj)∞
An = qn(cid:0)C((S2)dn) ⊗ K(cid:1)qn with unital connecting mappings. It follows from Lemma 7.8
that div2(An) > N for all n, and that Div2(A2) ≤ 3N + 4.
By Proposition 3.6 and Remark 3.5,
div2(A) = inf
n∈N
div2(An) > N,
and Div2(A) ≤ Div2(A2) ≤ 3N + 4.
Remark 7.10 (Initial objects). Suppose that C is a class of unital C ∗-algebra. An element
A in C is an inital object in C if there exists a unital ∗-homomorphism A → B for every B
in C.
It is well-known that the Cuntz algebra O∞ is an initial object in the class of unital
properly infinite C ∗-algebras. In fact, a unital C ∗-algebra is properly infinite if and only if
it contains O∞ as a unital sub-C ∗-algebra. Every properly infinite unital sub-C ∗-algebra of
O∞ is then also an initial object in the class of unital properly infinite C ∗-algebras. Hence
the Cuntz-Toeplitz algebras, Tn, n ≥ 2, are initial objects and so are all unital Kirchberg
algebras A for which the assignment [1A] 7→ 1 extends to a homomorphism K0(A) → Z.
It was shown in [ER06] that also the class of unital C ∗-algebras of real rank zero and
of infinite rank has initial objects. One can even find initial objects to this class which
are simple AF-algebras (necessarily with infinite dimensional trace simplex). It follows in
particular that the class of unital simple infinite dimensional C ∗-algebras of real rank zero
has initial objects.
Clearly, C is an initial object in the category of all unital C ∗-algebras, and so is any
unital C ∗-algebra that admits a character. (Note that we do not require the unital ∗-ho-
momorphism A → B to be injective.)
The corollary below shows that initial objects do not exist in the general non-real rank
zero case.
Corollary 7.11. The class of unital simple infinite dimensional C ∗-algebras and the class
of unital C ∗-algebras of infinite rank do not have initial objects. In fact, there is no unital
C ∗-algebra without characters that maps unitally into every unital simple infinite dimen-
sional C ∗-algebra.
26
Proof. If A is a unital C ∗-algebra that maps unitally into every unital simple infinite
dimensional C ∗-algebra, then div2(A) ≥ div2(B) for all unital simple infinite dimensional
C ∗-algebras B, cf. Remark 3.5, whence div2(A) = ∞ by Theorem 7.9. On the other hand,
if A has no character, then div2(A) < ∞ by Corollary 5.6.
The asymptotic divisibility numbers
We can give a lower and an upper bound on the asymptotic divisibility constant (discussed
in Section 4) for the C ∗-algebra considered above:
Corollary 7.12. Let N be a positive integer, and let A be the simple AH-algebra con-
structed in Theorem 7.9 associated with N. It follows that
(N − 1)/2 < Div∗(A) ≤ 2N + 2.
Proof. By Proposition 4.1 we get that Div∗(A) ≥ (Div2(A) − 1)/2 > (N − 1)/2. To
prove the reserve inequality, take any positive integer n and put m = 2n−1. We show that
Divm(An) ≤ (2N + 2)m, where An is as in the proof of Theorem 7.9. In the notation of
Lemma 7.8, let x = hpJni, put u = hqni, and recall that qn is the unit of the C ∗-algebra
An. By the definition of qn (in Lemma 7.8) it follows that mx ≤ u. As
dim((2 + 2N)m·pJn) − dim(qn) = (2N + 2)m − 2n = 2nN ≥
dim(Xn) − 1
2
,
if follows from [Dup76, Proposition 1] that (2 + 2N)m·pJn - qn, whence u ≤ (2N + 2)mx.
This proves that Divm(An) ≤ (2N + 2)m.
It follows from Remark 3.5 that Divm(A) ≤ (2N + 2)m whenever m is a power of 2,
and this entails that
Div∗(A) = lim inf
m→∞
Divm(A)/m ≤ lim inf
n→∞
Div2n−1(A)/2n−1 ≤ 2N + 2,
as desired.
We can use Lemma 7.7 to construct a unital, simple AH-algebra A such that Div∗(A) = ∞.
The proof requires the following sharpening of Corollary 7.3 that may have independent
interest.
Corollary 7.13. Let X be a locally compact Hausdorff space, and let q ∈ C0(X) ⊗ K
be a projection. Let m and N be positive integers, let I1, I2, . . . , IN be pairwise disjoint
subsets of N with Ii = m − 1 for all i, and let (pIi)N
i=1 be the associated projections in
C((S2)(m−1)N ) ⊗ K defined in Example 7.6. Suppose that
Then divm(hqi, C0(X)) > N.
q ⊗ 1 - q ⊗
N
Mi=1
pIi.
27
Proof. Apply Lemma 6.1 (iii) to x = h1i and yi = hpIii, and note that x ≤ myi, cf.
Example 7.6.
Lemma 7.14. Let (Jj)∞
Then, in the notation of Lemma 7.7, we have
j=1 be a sequence of pairwise disjoint subsets of N with Jj = 22j−1 j.
2k k < div2k(cid:0)hqni, C((S2)dn)(cid:1) < ∞
if n ≥ k, and that div2k(cid:0)hqni, C((S2)dn)(cid:1) = ∞ if n < k.
it suffices to show that qn ⊗LN
Proof. We use Corollary 7.13 with N = 2k k and m = 2k to prove the first claim. As 1 - qn
i=1 pIi has non-trivial Euler class, when I1, . . . , IN are as in
Corollary 7.13. Write
qn ⊗
N
Mi=1
pIi =
n
Mj=0
N
Mi=1
2max{0,j−1}·pJj ⊗ pIi =
n
Mj=0
N
Mi=1
2max{0,j−1}·pJj∪Ii.
qn ⊗ LN
As explained in Example 7.6, to prove non-triviality of the Euler class of the projection
i=1 pIi one needs to verify the combinatorial fact that the family of sets (Jj ∪ Ii),
j = 0, . . . , n, i = 1, . . . , N, and with the set Jj ∪ Ii repeated 2max{0,j−1} times, satisfies the
Marriage Lemma condition.
By first exhausting the elements in the sets Ii, and using that Pk−1
j=0 2max{0,j−1} = 2k−1 <
Ii, it suffices to show that the family of sets (Jj), j = k, . . . , n, with each set repeated
2j−1N = 2j+k−1k times, satisfies the Marriage Lemma condition. However, this holds
because Jj = 22j−1j ≥ 2j+k−1k when j ≥ k.
The second claim follows from the fact that the dimension of the projection qn is 2n
and that divm(hqni, C((S2)dn)) = ∞ whenever m > dim(qn).
Theorem 7.15. There is a simple unital infinite dimensional AH-algebra A which satisfies
Div∗(A) = ∞.
Proof. Let A be the simple AH-algebra contructed in Lemma 7.7 with respect to the choice
of (Jj) from Lemma 7.14. Recall that A is an inductive limit of a sequence of unital C ∗-
algebras A1 → A2 → · · · , where
An = qn(cid:0)C((S2)dn) ⊗ K(cid:1)qn.
It therefore follows from Lemma 7.14 that div2k (An) > 2k k, when n ≥ k, and div2k(An) =
∞ when n < k. This entails that Div2k(A) > 2k k by Proposition 3.6. Finally, by Propo-
sition 4.1 (ii),
Div∗(A) = lim sup
k→∞
Div2k (A)/2k = ∞.
As remarked in Section 4, if A is any unital C ∗-algebra, then Div∗(A) = 0 if and only if A
is properly infinite, and Div∗(A) ∈ [1, ∞] otherwise. Moreover, Div∗(A) = 1 if and only if
A is almost divisible (and not properly infinite). In other words, the range of the invariant
28
Div∗( · ) is contained in the set {0} ∪ [1, ∞], and Div∗(A) ≤ 1 if and only if A is almost
divisible.
We can easily produce examples of simple, unital, infinite dimensional C ∗-algebras A
such that Div∗(A) = 0 (eg., A could be a Cuntz algebra), or such that Div∗(A) = 1 (eg., A
is any simple, unital, infinite-dimensional C ∗-algebra of real rank zero, cf. Example 3.15).
The theorem above provides an example of a simple, unital, infinite dimensional C ∗-algebra
A where Div∗(A) = ∞.
It follows from Corollary 7.12 that Div∗( · ) attains infinitely many values in the interval
(1, ∞), when restricted to the class of unital, simple, infinite dimensional C ∗-algebras, and
that the possible values of Div∗( · ) in this interval is upwards unbounded. We do not know
if all values in the interval (1, ∞) are thus attained. For that matter we cannot exhibit
any number in the interval (1, ∞) which for sure is the value of Div∗(A) for some simple
unital infinite dimensional C ∗-algebra A.
Divisibility of infinite tensor products
C ∗-algebra (or a unital C ∗-algebra without characters), cf. [DT09].
We end this section by giving yet another class of examples of simple unital C ∗-algebras
j=1 Aj,
where the Aj's are unital simple infinite dimensional C ∗-algebras. In particular, such C ∗-
algebras need not absorb the Jiang-Su algebra tensorially. It remains an open problem if
j=1 A absorbs the Jiang-Su algebra whenever A is a simple unital infinite dimensional
with bad divisibility properties. The ones we construct below are of the form N∞
N∞
neous C ∗-algebras of rank two such that N∞
tally into N∞
Jiang-Su algebra does not embed into N∞
It was shown in [HRW07, Example 4.8] that there exists a sequence (An) of homoge-
n=1 An does not absorb the Jiang-Su algebra
tensorially. (It is an easy conseqence of this that the Jiang-Su algebra cannot embed uni-
n=k An for some k.) Of course, one can regroup the tensor factors An to get
a new sequence (Bn) of unital C ∗-algebras each of which has infinite rank and where the
n=1 Bn. It is not known if every unital C ∗-algebra
of infinite rank admits an embedding of a unital simple infinite dimensional C ∗-algebra. If
it were true, then Theorem 7.17 below would follow from [HRW07, Example 4.8] .
We introduce some notation to keep track of the combinatorics. Define a total order
on the set N × N0 by
(k, j) ≤ (ℓ, i) ⇐⇒ k + j < ℓ + i
or
(k + j = ℓ + i and k < ℓ).
For each (k, j) ∈ N × N0 and for each integer m ≥ k let S(m; k, j) denote the set of all
m-tuples (i1, i2, . . . , im) ∈ Nm
0 such that ik = j and (ℓ, iℓ) < (k, j) for all ℓ 6= k.
Lemma 7.16. Let N ≥ 1 be an integer. For each integer k ≥ 1, let (J (k)
of subsets of N such that J (k)
i = ∅ whenever (k, j) 6= (ℓ, i), and such that
j ∩ J (ℓ)
)∞
j=1 be a sequence
j
J (k)
j
= max
m≥k X(i1,...,im)∈S(m;k,j)
N
m
Yt=1
2max{it−1,0}.
29
(The quantity on the right-hand side is finite because S(m; k, j) is finite for all (k, j) and
all m ≥ k, and S(m; k, j) = ∅ when m > k + j.) Let d(k)
n ) ⊗ K
be as defined in Lemma 7.7 associated with the sequence (J (k)
n ∈ C((S2)d(k)
n ∈ N and q(k)
)∞
j=1. It then follows that
j
div2(cid:16)(cid:10)q(1)
n ⊗ q(2)
n ⊗ · · · ⊗ q(m)
n (cid:11), C((S2)d(1)
n ) ⊗ C((S2)d(2)
n ) ⊗ · · · ⊗ C((S2)d(m)
n )(cid:17) > N,
for all positive integers n and m.
Proof. Let T (n, m) be the set of all non-zero m-tuples (i1, i2, . . . , im) ∈ Nm
such that
0
ik ≤ n for all k = 1, 2, . . . , m. Adopt the convention J (k)
0 = ∅ for all k and let p∅ denote
the trivial (= constant) one-dimensional projection. We can then express the projection
q(1)
n ⊗ q(2)
n ⊗ · · · ⊗ q(m)
as follows:
n
1 ⊕ X(i1,...,im)∈T (n,m)(cid:0)
m
Yt=1
2max{it−1,0}(cid:1) · pJ (1)
i1
⊗ pJ (2)
i2
⊗ · · · ⊗ pJ (m)
im
.
By Proposition 7.4 it suffices to show that the Euler class of the projection
X(i1,...,im)∈T (n,m)(cid:0)N
m
Yt=1
2max{it−1,0}(cid:1) · pJ (1)
i1
⊗ pJ (2)
i2
⊗ · · · ⊗ pJ (m)
im
is non-zero, or, equivalently, that the Euler class of the projection
X(i1,...,im)∈T (n,m)(cid:0)N
m
Yt=1
2max{it−1,0}(cid:1) · pJ (1)
i1
∪J (2)
i2
∪···∪J (m)
im
i2 ∪ · · · ∪ J (m)
i2 ∪ · · · ∪ J (m)
i1 ∪ J (2)
im is repeated N ·Qm
i2 ∪ · · · ∪ J (m)
i1 ∪ J (2)
is non-zero. By [Rør03, Proposition 4.5], cf. Example 7.6, it suffices to show that the
family of sets J (1)
im , where (i1, . . . , im) ∈ T (n, m) and where the set
i1 ∪ J (2)
J (1)
Construct a matching by selecting the matching elements for the N · Qm
t=1 2max{it−1,0}
copies of the set J (1)
, where (k, j) is the largest
of the elements (1, i1), (2, i2), . . . , (m, im). To check that this work, i.e., to see that J (k)
is
large enough, let T (n, m; k, j) be the set of those m-tuples (i1, i2, . . . , im) in T (n, m) for
which
t=1 2max{it−1,0} times, admits a matching.
im inside the subset J (k)
j
j
(k, j) = max{(1, i1), (2, i2), . . . , (m, im)}.
Then T (n, m; k, j) ⊆ S(m; k, j), and so it follows by the assumption on J (k)
j
that
J (k)
j
≥
X(i1,...,im)∈T (n,m;k,j)
N
m
Yt=1
2max{it−1,0}.
The suggested matching is therefore possible.
30
C ∗-algebra A = N∞
The theorem below shows that an infinite tensor product of simple unital infinite di-
mensional C ∗-algebras does not necessarily have good divisibility properties. Any such
n=1 An will have (many) non-trivial central sequences, i.e., the central
sequence algebra Aω ∩ A′ with respect to an ultrafilter ω on N is non-trivial. For example,
CMm(C) embeds into Aω ∩ A′ for all m, albeit, not necessarily with full image. However,
in the example contructed below, one cannot embed the Jiang-Su algebra into Aω ∩ A′.
Theorem 7.17. For each integer N > 2, there exist a sequence (An) of unital simple
infinite dimensional C ∗-algebras (in fact, AH-algebras) such that
div2(cid:16)
∞
On=1
An(cid:17) > N.
In particular, N∞
a unital embedding of Z.
n=1 An does not absorb the Jiang-Su algebra Z, and it does not even admit
j (cid:1)∞
the sequence (cid:0)J (k)
Proof. Let Ak be the simple unital AH-algebra constructed in Lemma 7.7 associated with
j=1 from Lemma 7.16. Then Ak is an inductive limit of a sequence,
Ak(1) → Ak(2) → · · · , of unital homogenous C ∗-algebras with unital connecting maps,
where
Ak(n) = q(k)
n (cid:16)C((S2)d(k)
n ) ⊗ K(cid:17)q(k)
n .
n=1 An is the inductive limit of the sequence
A1 → A1 ⊗ A2 → A1 ⊗ A2 ⊗ A3 → · · ·
The infinite tensor product N∞
with unital connecting maps. It therefore suffices to show that div2(cid:0)Nm
every m, cf. Proposition 3.6. Now, Nm
Ok=1
Ak(3) → · · · ,
Ak(1) →
Ak(2) →
Ok=1
k=1 Ak is the inductive limit of the sequence
m
Ok=1
m
m
k=1 Ak(cid:1) > N for
with unital connecting mappings, and so, again by Proposition 3.6, it suffices to show that
div2(cid:16)
m
Ok=1
Ak(n)(cid:17) > N
for every m and n. The latter is precisely the content of Lemma 7.16.
8 Ultrapowers
In this section we show that our divisibility properties behave well with respect to taking
direct products and ultrapowers of sequences of unital C ∗-algebras. This has the surprising
31
consequence that such products and ultrapowers may admit characters even if all the C ∗-
algebras in the ingoing sequence are unital, simple and infinite dimensional.
We define the notion of "almost characters" and show that the existence of such is re-
lated to the invariant div2( · ). It follows in particular that simple unital infinite dimensional
C ∗-algebras can have almost characters.
First we need some technical lemmas:
Lemma 8.1. Let A be a unital C ∗-algebra and let (Iλ) be an upward directed family of
ideals of A. Set S Iλ = I. It follows that
Divm(A/I) = inf
λ
Divm(A/Iλ),
∂ivm(A/I) = inf
λ
∂ivm(A/Iλ)
for all positive integers m.
divm(A/I) = inf
λ
divm(A/Iλ)
Proof. The inequality "≤" in all three cases follows from Remark 3.5 since we have a unital
∗-homomorphism A/Iλ → A/I for each λ. We prove the reverse inequality "≥" only in the
case of ∂ivm( · ); the proofs of the other two instances are similar.
Set ∂ivm(A/I) = n, and let us show that ∂ivm(A/Iλ) ≤ n for some λ. Find x1 . . . , xm
in Cu(A/I) be such that
x1 + x2 + · · · + xm ≤ h1i ≤ nxj
for all j. Find positive contractions a1, . . . , am in A ⊗ K such that xj = hbji, where
bj ∈ A/I ⊗ K is the image of aj under the quotient mapping A → A/I. Find ε > 0 such
that h1i ≤ nh(bj −ε)+i for all j. It follows from [KR00, Lemma 4.12] that there are positive
elements c, c′
m in I ⊗ K such that
1, . . . , c′
h(a1 − ε/2)+i + · · · + h(am − ε/2)+i ≤ h1i + hci,
h1i ≤ nh(aj − ε)+i + hc′
ji
for all j. There is δ > 0 such that
h(a1 − ε)+i + · · · + h(am − ε)+i ≤ h1i + h(c − δ)+i,
h1i ≤ nh(aj − ε)+i + h(c′
j − δ)+i.
Since S Iλ is dense in I, it follows that (c−δ)+ and (c′
j −δ)+ all belong to Iλ⊗K for some
λ. Let zj ∈ Cu(A/Iλ) be the Cuntz class of the image of the element (aj − ε)+ under the
quotient mapping A → A/Iλ. Then z1 +· · ·+zm ≤ h1i ≤ nzj, whence ∂ivm(A/Iλ) ≤ n.
For each ε > 0, let hε : R+ → [0, 1] be a continuous functions such that hε(0) = 0 and
hε(t) = 1 when t ≥ ε.
Lemma 8.2. Let A be a unital C ∗-algebra. Let b1, b2, . . . , bn be positive elements in A such
j=1hbji ≥ h1Ai. Then, for some ε > 0, there are contractions yj in A such that
that Pn
n
Xj=1
y∗
j hε(bj)yj = 1A.
32
Proof. By assumption, and by compactness of h1Ai, there are elements vj ∈ A such that
j (bj − ε)+vj is invertible for some ε > 0, and so there are
+ wj and notice
j (bj − ε)+wj = 1A. Put yj = (bj − ε)1/2
j bjvj = 1A. Thus Pn
Pn
j=1 v∗
j=1 v∗
elements wj ∈ A such that Pn
that hε(bj)(bj − ε)+ = (bj − ε)+ for all j. Thus
j=1 w∗
n
Xj=1
y∗
j hε(bj)yj =
n
Xj=1
y∗
j yj = 1A,
which shows that the yj's are contractions with the desired properties.
Lemma 8.3. Let A be a unital C ∗-algebra, and let m, n be positive integers.
(i) A is weakly (m, n)-divisible if and only if there exist positive contractions aij and
contractions yj in A, j = 1, 2, . . . , n and i = 1, 2, . . . , m, such that a1j, a2j, . . . , amj
are pairwise equivalent and orthogonal for all j, and such that 1A = Pn
(ii) A is (m, n)-decomposable if and only if there exist pairwise orthogonal positive con-
tractions ai and contractions yij in A, j = 1, 2, . . . , n and i = 1, 2, . . . , m, such that
j=1 y∗
j a1jyj.
Pn
j=1 y∗
ijaiyij = 1A for all i.
(iii) A is (m, n)-divisible if and only if there exist pairwise equivalent and pairwise or-
thogonal positive contractions ai and contractions yj in A, j = 1, 2, . . . , n and i =
1, 2, . . . , m, such that Pn
j=1 y∗
j a1yj = 1A.
Proof. We view A as being a sub-C ∗-algebra of A ⊗ K which is identified with the corner
1A(A ⊗ K)1A.
(i). "If". Put xj = ha1ji = haiji ∈ Cu(A). Then
mxj = h
m
Xi=1
n
n
y∗
Xj=1
aiji ≤ h1Ai = (cid:10)
j a1jyj(cid:11) ≤
j ≪ xj such that h1Ai ≤ x′
Xj=1
1+x′
hy∗
j a1jyji ≤
n
Xj=1
ha1ji =
n
Xj=1
xj.
"Only if". Choose x′
n. Use Lemma 2.3 (ii) to find
positive contractions bij in A, j = 1, 2, . . . , n and i = 1, 2, . . . , m, such that b1j, b2j, . . . , bmj
are pairwise orthogonal and equivalent for all j and such that x′
j ≤ hbiji ≤ xj. It then
follows that 1A - b11 ⊕b12 ⊕· · ·⊕b1n, and so it follows from Lemma 8.2 that there are ε > 0
j hε(b1j)yj = 1A. The contractions yj together
2+· · ·+x′
and contractions yj in A such that Pn
with the positive contractions aij = hε(bij) are then as desired.
j=1 y∗
(ii). "If". Put xi = haii ∈ Cu(A). Then
x1 + · · · + xm = (cid:10)a1 + · · · + am(cid:11) ≤ h1Ai = (cid:10)
n
Xj=1
y∗
ijaiyij(cid:11) ≤
n
Xj=1
hy∗
ijaiyiji ≤ nxi
for all i.
"Only if". Choose x′
i ≪ xi such that nx′
orthogonal and equivalent positive contractions b1, b2, . . . , bm in A such that x′
i ≥ h1Ai. Use Lemma 2.3 (ii) to find pairwise
i ≤ hbii ≤ xi.
33
Proceed as in the proof of "only if" in part (i) to obtain elements ai = hε(bi) (for a suitable
ε > 0) and yij with the desired properties.
(iii). "If". Put x = ha1i = haii ∈ Cu(A). Then
mx = (cid:10)a1 + a2 + · · · + ami ≤ h1Ai ≤ (cid:10)
n
Xj=1
y∗
j a1yj(cid:11) ≤
n
Xj=1
hy∗
j a1yji ≤ nx.
"Only if". Choose x′ ≪ x such that h1Ai ≤ nx′. By Lemma 2.3 (ii) there are pairwise
orthogonal and pairwise equivalent positive contractions b1, b2, . . . , bm in A such that x′ ≤
hbji ≤ x. We can now follow the proof of "only if" in part (i) to obtain elements aj = hε(bj)
(for a suitable ε > 0) and yj with the desired properties.
bounded sequences (ak), with ak ∈ Ak.
If (Ak) is a sequence of C ∗-algebras, then we denote by Q∞
cω({Ak}) the closed two-sided ideal in Q∞
limω kakk = 0. Finally, denote the quotient Q∞
k=1 Ak the C ∗-algebra of all
If ω is a (free) filter on N, then denote by
k=1 Ak consisting of those sequences (ak) for which
k=1 Ak/cω({Ak}) by Qω Ak.
Proposition 8.4. Let (An) be a sequence of unital C ∗-algebras. Then, for all integers
m ≥ 2 and for any free filer ω on N we have:
Divm(Ak).
∂ivm(Ak).
divm(Ak).
ω
k=1 Ak(cid:17) = supk Divm(Ak), Divm(cid:16)Qω Ak(cid:17) = lim sup
k=1 Ak(cid:17) = supk ∂ivm(Ak),
∂ivm(cid:16)Qω Ak(cid:17) = lim sup
k=1 Ak(cid:17) = supk divm(Ak),
divm(cid:16)Qω Ak(cid:17) = lim sup
(i) Divm(cid:16)Q∞
(ii) ∂ivm(cid:16)Q∞
(iii) divm(cid:16)Q∞
We have unital ∗-homomorphisms Q∞
We show next that Divm(cid:16)Q∞
Proof. We only prove (i). The proofs of (ii) and (iii) are very similar.
ω
ω
k=1 Ak → An for all n. Therefore the inequality
k=1 Ak(cid:17) ≤ supk Divm(Ak). Let n be a positive integer such
that Divm(Ak) ≤ n for all k. Then, by Lemma 8.3 (i), for each k we can find positive
contractions a(k)
in Ak, for j = 1, 2, . . . , n and i = 1, 2, . . . , m, such
that a(k)
mj are pairwise orthogonal and equivalent for all j, and such that
ij and contractions y(k)
2j , . . . , a(k)
1j , a(k)
j
"≥" holds in the first identity in (i) (and also in (ii) and (iii)), cf. Remark 3.5.
1Ak =
n
Xj=1
(y(k)
j )∗a(k)
1j y(k)
j
.
Put
aij = (a(k)
ij ) ∈
∞
Yk=1
Ak,
yj = (y(k)
j ) ∈
∞
Yk=1
Ak.
34
Then a1j, a2j, . . . , amj are pairwise orthogonal and equivalent, and Pn
to the unit of Q∞
whence Divm(cid:16)Q∞
∗-homomorphism Qk∈I Ak → Qω Ak for each I ∈ ω. We can therefore use Remark 3.5 and
k=1 Ak. By Lemma 8.3 (i), this shows that Q∞
k=1 Ak(cid:17) ≤ n.
To prove the second part of (i) note first that we have a natural unital (surjective)
j=1 y∗
j a1jyj is equal
k=1 Ak is (m, n)-divisible,
the first identity in (i) to conclude that
Divm(cid:0)Yω
Ak(cid:1) ≤ Divm(cid:0)Yk∈I
Ak(cid:1) = sup
k∈I
Divm(Ak),
which shows that Divm(cid:0)Qω Ak(cid:1) ≤ lim supω Divm(Ak).
We proceed to prove the reverse inequality. For each I ∈ ω consider the ideal J(I) in
Q∞
k=1 Ak consisting of those sequences (ak) for which ak = 0 for all k ∈ I. Then
cω({Ak}) = [I∈ω
J(I),
(where ω is ordered by reverse inclusion). We can now use Lemma 8.1 and the first identity
in (i) to conclude that
Divm(cid:0)Yω
Ak(cid:1) = inf
I∈ω
Divm(cid:16)(cid:0)
∞
Yk=1
Ak(cid:1)/J(I)(cid:17) = inf
I∈ω
Divm(cid:0)Yk∈I
Ak(cid:1)
= inf
I∈ω
sup
k∈I
Divm(Ak) = lim sup
ω
Divm(Ak).
If we combine the proposition above with Corollary 5.6 (i) we obtain:
Corollary 8.5. Let (Ak) be a sequence of unital C ∗-algebras such that limk→∞ div2(Ak) =
∞. Then Q∞
k=1 Ak has a character, and so does Qω Ak for each free filter ω on N.
If we combine the corollary above with Theorem 7.9, then we obtain the following surprising
fact:
Corollary 8.6. There is a sequence (Ak) of unital simple infinite dimensional C ∗-algebras
such that Q∞
k=1 Ak and Qω Ak have characters for each free filter ω on N.
Clearly, none of the C ∗-algebras Ak in the corollary above can have a character. However,
they have "almost characters" in the sense defined below. This is one way of understanding
how the product C ∗-algebra can have a character when none of the individual C ∗-algebras
has one.
Definition 8.7. Let N ≥ 2 be an integer and let ε > 0. A unital C ∗-algebra A is said to
have (N, ε)-characters if for every N-tuple u1, u2, . . . , uN of unitaries in A there exists a
state ρ on A such that ρ(uj) ≥ 1 − ε for j = 1, 2, . . . , N.
35
A state ρ on a unital C ∗-algebra is a character if and only if ρ(u) = 1 for all unitary
elements u ∈ A. Most simple C ∗-algebras that we know of do not have (2, ε)-characters
for small ε > 0. For example, if A is a C ∗-algebra which contains unitaries u, v such that
kuvu∗v∗ − λ1Ak < η for some λ ∈ T and for some η < 1 − λ, then A does not admit any
(2, ε)-character for some small enough ε > 0. Indeed, if ρ is a state on A such that ρ(u)
and ρ(v) are close to 1, then ρ(uvu∗v∗) is close to 1.
Proposition 8.8. A unital C ∗-algebra has a character if and only if it has (N, ε)-characters
for all pairs (N, ε), where N ≥ 2 is an integer and ε > 0.
Proof. The "only if" part is trivial. Assume that A is a unital C ∗-algebra that has (N, ε)-
characters for all pairs (N, ε). For each finite subset F of the unitary group of A and for
each ε > 0, let S(F, ε) denote the set of states ρ on A such that ρ(u) ≥ 1 − ε for all u ∈ F .
It follows that T(F,ε) S(F, ε) is non-empty,
Then, by assumption, S(F, ε) is non-empty.
and any state in this intersection is a character.
Proposition 8.9. Let (Ak) be a sequence of unital C ∗-algebras, and let ω be a free ultrafilter
ε > 0 there exists I ∈ ω such that Ak has (N, ε)-characters for each k ∈ I.
on N. Then Qω Ak has a character if and only if for each integer N ≥ 2 and for each
Proof. We prove first the "if" part. By Proposition 8.8 it suffices to show that Qω Ak has
(N, ε)-characters for all (N, ε). Fix (N, ε) and find I ∈ ω such that Ak has (N, ε)-characters
for each k ∈ I. Let u1, . . . , uN be unitaries in Qω Ak, and let (u(k)
j ) ∈ Q∞
k=1 Ak be a lift
of uj. Then for each k ∈ I there is a state ρk on Ak such that ρk(u(k)
j ) ≥ 1 − ε for
j = 1, 2, . . . , N. Choose arbitrary states ρk on Ak for k /∈ I and define a state ρ on Qω Ak
by ρ(x) = limω ρk(xk), where (xk) ∈ Q∞
k=1 Ak is a lift of x. (A priori, ρ defines a state on
Q∞
k=1 Ak, and one checks that it vanishes on the ideal cω({Ak}).) Then
ρk(u(k)
ρk(u(k)
j ) ≥ 1 − ε,
ρ(uj) = lim
k→ω
j ) ≥ inf
k∈I
for j = 1, 2, . . . , N, which shows that Qω Ak has (N, ε)-characters.
Suppose next that Qω Ak has a character ρ. Fix (N, ε), and let J be the set of those
k ∈ N for which Ak does not have (N, ε)-characters. For each k ∈ J choose unitaries u(k)
j
in Ak, j = 1, 2, . . . , N, such that there is no state ρ′ on Ak for which ρ′(u(k)
j ) ≥ 1 − ε for all
j = 1, 2, . . . , N. Choose arbitrary unitaries u(k)
j ∈ Ak for k /∈ J, and let uj be the unitary
element (u(k)
j ) in Q∞
by the unitaries πω(uj), where πω is the quotient mapping Q∞
for all x = (xk) ∈ Q∞
k=1 Ak. Let B be the (separable) sub-C ∗-algebra of Qω Ak generated
k=1 Ak → Qω Ak. By [Kir06,
Lemma 2.5] there is a sequence ρk of pure states on Ak such that ρ(πω(x)) = limω ρk(xk)
k=1 Ak with πω(x) ∈ B. Now,
1 = ρ(πω(uj)) = lim
ω
ρk(u(k)
j ) = lim inf
k→ω
ρk(u(k)
j ) = sup
I∈ω
ρk(u(k)
j ).
inf
k∈I
It follows that there exists I ∈ ω such that ρk(u(k)
j ) ≥ 1 − ε for all k ∈ I and for all
j = 1, 2, . . . , N. This entails that I ∩ J = ∅. Hence Ak has (N, ε)-characters for all
k ∈ I.
36
We can relate the existence of (N, ε)-characters on a C ∗-algebra A to the divisibility quan-
tity div2(A).
Theorem 8.10. For each pair (N, ε), where N ≥ 2 is an integer and ε > 0, there exists an
integer n ≥ 2 such that every unitalC ∗-algebra A which satisfies div2(A) ≥ n has (N, ε)-
characters. Conversely, for every integer n ≥ 2 there exists a pair (N, ε), where N ≥ 2 is
an integer and where ε > 0, such that every unital C ∗-algebra A which has (N, ε)-characters
satisfies div2(A) ≥ n.
Proof. Suppose that the first claim were false. Then there would exist a pair (N, ε) and
a sequence (An) of unital C ∗-algebras such that div2(An) ≥ n and none of the An's have
(N, ε)-characters. However, if ω is any free ultrafilter N, then Qω An has a character
by Corollary 8.5, whence An has (N, ε)-characters for each n in some subset I ∈ ω, a
contradiction.
Suppose next that the second statement were false. Then there would exist an integer
n ≥ 2 and a sequence (Ak) of unital C ∗-algebras such that Ak has (k, 1/k)-characters
but div2(Ak) < n. Let ω be a free ultrafilter on N. It then follows from Proposition 8.9
that Qω Ak has a character. Hence div2(Qω Ak) = ∞, whence limω div2(Ak) = ∞ by
Proposition 8.4, a contradiction.
Corollary 8.11. For each pair (N, ε), where N ≥ 2 is an integer and ε > 0, there exists
a unital simple infinite dimensional C ∗-algebra which has (N, ε)-characters.
We end this section by giving several equivalent formulation of some well-known open
problems for C ∗-algebras. Recall that a C ∗-algebra A has the Global Glimm Halving
property if there is a ∗-homomorphism CM2(C) → A whose image is full in A.
Proposition 8.12. The following statements are equivalent:
(i) Every unital C ∗-algebra that has no finite dimensional representation has the Global
Glimm Halving property.
(ii) For all unital C ∗-algebras A, if divm(A) < ∞ for all m ≥ 2, then Div2(A) < ∞.
(iii) For every sequence (Ak) of unital C ∗-algebras, if supk divm(Ak) < ∞ for all m, then
supk Div2(Ak) < ∞.
Proof. (i) ⇔ (ii). A has no finite dimensional representations if and only if rank(A) ≥ m
for all m, which by Corollary 5.4 (iii) is equivalent to divm(A) < ∞ for all m.
It was
shown in Corollary 5.4 (i) that the Global Glimm Halving property holds for A if and only
if Div2(A) < ∞.
(ii) ⇒ (iii). Given a sequence (Ak) of unital C ∗-algebras such that supk divm(Ak) < ∞
k=1 Ak. Then divm(A) = supk divm(Ak) < ∞ by
Proposition 8.4. Thus Div2(A) < ∞, which implies that supk Div2(Ak) = Div2(A) < ∞,
again by Proposition 8.4.
for all m. Consider the C ∗-algebra A = Q∞
(iii) ⇒ (ii) is trivial: Take Ak = A for all k.
37
Proposition 8.13. The following statements are equivalent:
(i) All unital C ∗-algebras A that have no finite dimensional representation contain two
positive full elements that are orthogonal to each other.
(ii) For all unital C ∗-algebras A, if divm(A) < ∞ for all m ≥ 2, then ∂iv2(A) < ∞.
Proof. (i) ⇔ (ii). As in the proof of Proposition 8.12, A has no finite dimensional rep-
resentations if and only if divm(A) < ∞ for all m.
It was shown in Corollary 5.4 (ii)
that A contains two positive full elements that are orthogonal to each other if and only if
∂iv2(A) < ∞.
(ii) ⇒ (iii) is similar to the proof of (ii) ⇒ (iii) in Proposition 8.12. (iii) ⇒ (ii) is
trivial.
9
Infinite elements
Following [KR00], a Cuntz class u in the Cuntz semigroup of a C ∗-algebra A is said to
be properly infinite if it satisfies u = 2u (whence u = ∞ · u). Similarly, a countably
generated Hilbert module over A is properly infinite if its Cuntz class is properly infinite.
We saw in Proposition 3.4 how infiniteness of a Cuntz class can arise from a certain type
of divisibility property. In this section we shall investigate this and related phenomena
further with emphasis on the following property:
Definition 9.1. Let A be a C ∗-algebra, let n ≥ 1 be an integer, and let u be an element in
i=1 xi ≤ u
Cu(A). We say that u is (ω, n)-decomposable if there exist x1, x2, . . . such that P∞
and u ≤ nxi for all i.
If u is (ω, n)-decomposable, then u is (m, n)-decomposable for all m.
Proposition 3.4 (ii), it follows that nu is properly infinite.
In particular, by
The condition in the definition above can be reformulated in several different ways:
Lemma 9.2. Let A be a C ∗-algebra, let n ≥ 1 be an integer, and let u be an element of
Cu(A). Then the following conditions are equivalent:
(i) u is (ω, n)-decomposable,
(ii) there exist x1, x2, . . . such that P∞
(iii) there exist x1, x2, . . . and y1, y2, . . . such that P∞
(iv) there exist x1, x2, . . . such that P∞
and ∞·u ≤ ∞·sup yi,
i=1 xi ≤ u and nP∞
j=k xj = ∞·u for all k.
i=1 xi ≤ u and nxi = ∞·u for all i,
i=1 xi ≤ u, yi−1 ≤ yi ≤ nxi for all i,
38
Proof. (i) ⇒ (ii). Suppose that x1, x2, . . . satisfy P∞
partition of the natural numbers into infinite sets. Then the elements x′
that condition (ii) holds.
i=1 xi ≤ u ≤ nxj. Let {Ii}∞
i = Pj∈Ii
i=1 be a
xj witness
To get (ii) ⇒ (iii), set yi = ∞·u for all i and choose (xi)∞
(iii)⇒ (iv). Let (xi)∞
i=1 and (yi)∞
i=1 be as in (iii). Then P∞
the left side of this inequality decreases as k increases. Thus, P∞
k′ ≥ k. Taking the supremum over all k′ ≥ k we get P∞
i=k xi ≥ ∞·sup yi ≥ ∞·u.
(iv) ⇒ (i). Suppose that x1, x2, . . . satisfy the condition in (iv). Let (ui)∞
i=1 be such
that ui ≪ ui+1 for all i and supi ui = u. Then there exists a sequence 1 = k0 < k1 < · · ·
such that the elements x′
i=1 be a partition
of the natural numbers into infinite sets. Then the elements x′′
x′
j satisfy the
condition in Definition 9.1.
i ≥ ui for all i. Let {Ii}∞
i = Pj∈Ii
i = Pki−1
xj satisfy x′
j=ki−1
i=1 that satisfies (ii).
i=k xi ≥ ∞·yk. Observe that
i=k xi ≥ ∞ · yk′ for all
It was shown in [OPR] that the Corona Factorization Property for a C ∗-algebra is equivalent
to a condition for its Cuntz semigroup, that we here shall refer to as (CFP4S). A complete
ordered abelian semigroup is said to have (CFP4S) if whenever (xi)∞
i=1 is a full sequence,
x′ ≪ x1, and (yi)∞
i=1 is such that myi ≥ xi for all i and some m, then there exists n such
i=1 yi ≥ x′. Recall that a full sequence is one that is increasing and such that sup xi
is a full element, i.e., ∞ · sup xi is the largest element of the semigroup (which we shall
denote by ∞).
that Pn
In Section 6 we discussed a related notion, called the strong Corona Factorization
Property, and its analog for the Cuntz semigroup.
The proposition below relates the (CFP4S) with the notion of (ω, m)-divisibility. In
fact, it is a consequence of this proposition that a semigroup in the category Cu has a full
elements which is (ω, m)-divisible and not properly infinite if and only if the semigroup
does not satisfy (CFP4S).
Proposition 9.3. The following four conditions are equivalent for any object S in the
category Cu.
(i) S has property (CFP4S).
(ii) For every sequence (yi)∞
i=1 in S, if there is a full sequence (xi)∞
i=1 in S such that
myi ≥ xi for some m and for all i, then P∞
i=1 yi = ∞.
(iii) For every sequence (yi)∞
∞.
P∞
(iv) For every sequence (yi)∞
i=1 yi = ∞.
i=1 in S, if myi = ∞ for some m and for all i, then P∞
i=1 in S, if P∞
i=n myi = ∞ for some m and for all n, then
i=1 yi =
(v) For every full element y in S, if y is (ω, m)-decomposable for some m, then y is
properly infinite (whence y = ∞).
39
Proof. (i) ⇒ (ii). Apply the (CFP4S) to the tail sequences (xi)∞
i=n and (yi)∞
i=n. Then we
i=n yi for all x′ ≪ xn, whence ∞ = supn xn ≤ P∞
(ii) ⇒ (i). If (xi) is a full sequence and if (yi) is another sequence such that xi ≤ myi,
i=1 yi for some n by the
i=1 yi = ∞ by (ii). In particular, if x′ ≪ x1, then x′ ≤ Pn
i=1 yi.
(ii) ⇒ (iii). Suppose that (ii) holds and that (yi) is a sequence in S such that myi = ∞
i=n yi ≤ P∞
definition of compact containment.
get x′ ≤ PN
then P∞
for all i. Then ∞ = P∞
i=1 yi ≤ y and myi = ∞ · y = ∞ for all i. But then P∞
P∞
i=n myi = ∞ for some m and for all n. Put y = P∞
P∞
(v) ⇒ (iv). Suppose that (v) holds and let (yi)∞
i=1 yi by (ii) with xi = ∞ for all i.
whence y = ∞.
(iii) ⇒ (v). Suppose that (iii) holds, and that y ∈ S is full and (ω, m)-decomposable.
Then y satisfies condition (ii) of Lemma 9.2, so there exists a sequence (yi) such that
i=1 yi = ∞ because (iii) holds,
i=1 be a sequence in S such that
i=1 yi. We must show that y = ∞.
We know that my = ∞, so y is full.
It is easy to see that y satisfies condition (iv) of
Lemma 9.2, so y is (ω, m)-decomposable. Hence y = ∞ by the assumption that (v) holds.
i=1 be a full sequence in S, let m ≥ 1 be a positive integer, and let
(iv) ⇒ (ii). Let (xi)∞
(yi) be such that myi ≥ xi for all i. Then
∞
Xi=n
myi ≥
∞
Xi=n
xi ≥
∞
Xi=k
xi ≥ ∞·xk
for all k ≥ n. As ∞ = supk ∞·xk, we conclude that P∞
i=1 yi = ∞, and in particular that x1 ≤ P∞
entails that P∞
i=1 yi.
i=n myi = ∞ for all n. By (iv) this
In the following example we describe a Cuntz semigroup with an element u that is (ω, 2)-
decomposable but not properly infinite. In particular, 2u is properly infinite while u is
not. This example is well known in other contexts, and it was discussed in the paragraph
preceding Corollary 7.3.
Example 9.4. Let X = (S2)∞ be a countable cartesian product of 2-dimensional spheres,
and let pi ∈ C(X, K) be the one-dimensional projection arising as the pull back of a non-
trivial rank one projection p in C(S2) ⊗ K along the ith coordinate projection X → S2,
cf. the comments above Corollary 7.3. Let e be a trivial one-dimensional projection. Then
i=1 pj for all N, because the Euler class of the projection on the right-hand side is
Put xi = hpii, v = hei, and put u = P∞
i=1 xi. Then v (cid:2) u, v ≤ 2xi, and u + u = ∞·v =
j=i xi = ∞·v = ∞·u, and so u is (ω, 2)-decomposable; but u is not properly
non-trivial.
e - LN
∞·u. Hence 2P∞
infinite.
We now look more closely at the properties of (ω, n)-decomposable elements.
Proposition 9.5. Let (ai)∞
a C ∗-algebra A such that P∞
that Pj≥i nhaii = ∞ for all i. Then A ⊗ B is stable for every σ-unital C ∗-algebra B with
i=0 be a sequence of mutually orthogonal positive elements in
i=0 ai converges to a strictly positive element in A. Assume
rank(B) ≥ n.
40
Proof. Set P∞
i=0 ai = a ∈ A and let b be a strictly positive element in B. Notice that a ⊗ b
is a strictly positive element of A ⊗ B. In order to prove stability of A ⊗ B we will use
the stability criterion obtained in [HR98]: A ⊗ B is stable if for every ε > 0 there exists a
positive element c in A which is orthogonal to (a ⊗ b − ε)+ and satisfies h(a ⊗ b − ε)+i ≤ hci.
Arguing as in the proof of (iv) ⇒ (i) in Lemma 9.2 we may assume that nhaii = ∞ for
all i. By Theorem 5.3 (iii), rank(B) ≥ n is equivalent to weak (n, ω)-divisibility for hbi.
Thus there exist a sequence (xi)∞
i=1 xi = ∞.
We can form a new sequence (x′
i=1 in which each xi appears repeated infinitely often.
i≥j xi = ∞ for all j. Find
i=1 in Cu(B) such that nxi ≤ hbi for all i and P∞
In this way we may assume without loss of generality that P∞
positive elements bi in B such that xi = hbii and kbik ≤ 2−i. Then
i)∞
(a ⊗ b − ε)+ =
∞
Xi=1
(ai ⊗ b − ε)+ =
N
(ai ⊗ b − ε)+,
Xi=1
for some integer N ≥ 1. Set c = Pi>N ai ⊗ b. Then c is orthogonal to (a ⊗ b − ε)+. Also,
hai ⊗ bi = haii ⊗ hbi ≥ n haii ⊗ hbii = ∞·hai ⊗ hbii
for each i. Hence
hci = Xi>N
hai ⊗ bi = Xi>N
∞·hai ⊗ hbii = ∞.
Thus, h(a ⊗ b − ε)+i ≤ ∞ = hci. This shows that A ⊗ B is stable.
The proposition above can be applied to the C ∗-algebra A = P (C(X) ⊗ K)P arising from
i=1 pi ∈ M(C(X) ⊗ K). The C ∗-algebra A is not stable (because
e /∈ A while e ∈ M2(A)), but A ⊗ B is stable for every C ∗-algebra B that does not have a
character by Proposition 9.5 and Example 9.4.
Example 9.4 with P = L∞
The example obtained in [Rør97] of a simple C ∗-algebra A of stable rank 1 such that
Mn(A) is stable, while Mn−1(A) is not stable, likewise satisfies the hypotheses of Proposi-
tion 9.5. In fact, to the authors knowledge, every example of a C ∗-algebra that tensored
with Mn(C) becomes stable also has the stronger property of becoming stable after being
tensored with any C ∗-algebra that has no representations of dimension less than n. This
raises the following question:
Question 9.6. Is there a C ∗-algebra A such that M2(A) is stable but A ⊗ B is not stable
for some C ∗-algebra B without characters?
Proposition 9.7. The following statements are equivalent for every C ∗-algebra A with a
strictly positive element a.
(i) hai is (ω, n)-decomposable.
(ii) A contains a full hereditary subalgebra B such that B ⊗C is stable for any C ∗-algebra
C such that rank(C) ≥ n.
41
(iii) A contains a full hereditary subalgebra B such that Mn(B) is stable.
Proof. (i) ⇒ (ii). Let (xi)∞
i=1 xi ≤ hai and nxi = ∞. Let b ∈ A ⊗ K be
strictly positive and let bi ∈ A ⊗ K be mutually orthogonal elements such that hbii = xi.
We can find mutually orthogonal positive elements ai in A such that haii ≤ hbii, nhaii ≥
i=1 ai is convergent. It then follows from Proposition 9.5 that
i=1 be such that P∞
h(b − 1/i)+i, and such that P∞
(ii) holds when B is the hereditary sub-C ∗-algebra generated by P∞
(ii) ⇒ (iii) is clear.
(iii) ⇒ (i). Since A is σ-unital, and B is stably isomorphic to A, B is σ-unital too. Let
i=1 ai.
b be a strictly positive element in B.
Use [OPR, Lemma 5.3] to find a sequence (bk) of pairwise orthogonal positive elements
in B such that h(b − 1/k)+i ≤ nh(b − 1/k)+i ≤ nhbki for all k. Then condition (iii) of
Lemma 9.2 is satisfied with u = hbi, xk = hbki, and yk = h(b − 1/k)+i, whence hbi is
(ω, n)-decomposable.
Finally, by the fact that hbi ≤ hai ≤ ∞·hbi, it follows by the equivalence of (i) and (ii)
in Lemma 9.2 that hai is (ω, n)-decomposable.
Definition 9.8. Let n ∈ N and u ∈ Cu(A). We call u weakly (ω, n)-divisible if for every
u′ ≪ u there exist xi ∈ Cu(A), i = 1, 2, . . . , n, such that ∞·xi ≤ u and u′ ≤ Pn
Observe that if u is weakly (ω, n)-divisible, then u is weakly (m, n)-divisible for all m ∈ N,
whence nu is properly infinite by Proposition 3.4.
i=1 xi.
We will next give an example of a Cuntz semigroup element that is weakly (ω, 2)-
divisible but not properly infinite. This example needs some preparatory results. Let us
first recall an example given by Dixmier and Doaudy in [DD63].
Example 9.9 (Dixmier -- Douady, [DD63, §17]). Let B∞ denote the closed unit ball of
l2(N) endowed with the weak topology. Let l2(B∞) denote the C(B∞)-Hilbert module
of continuous maps from B∞ to l2(N). We will construct a countably generated C(B∞)-
Hilbert module D such that l2(B∞) ֒→ D ֒→ l2(B∞) but D ≇ l2(B∞).
Let x : B∞ → l2(N) ⊕ Ce be given by
x(z) = z +p1 − kzk2 · e, for z ∈ B∞.
Consider the C(B∞)-module D0 of functions from B∞ to l2(N) ⊕ Ce that have the form
y + xλ, with y ∈ l2(B∞) and λ ∈ C(B∞). The module D0 is a pre-Hilbert C*-module over
C(B∞) when endowed with the pointwise inner product. Indeed, if y1 + xλ1 and y2 + xλ2
are vectors in D0 then
hy1 + xλ1, y2 + xλ2i = hy1, y2i + hy1, ziλ2 + λ1hz, yi + λ1λ2 ∈ C(B∞).
Let D denote the completion of D0 with respect to the norm induced by its C(B∞)-valued
inner product. Observe that l2(B∞) ֒→ D0 ⊆ D. Since D is countably generated, we also
have that D ֒→ l2(B∞) by Kasparov's stabilization theorem. Let us see that D ≇ l2(B∞).
Consider E ⊆ D, the orthogonal complement of {x}. Then E = E0, where
E0 = {y + xλ ∈ D hy(z), zi + λ(z) = 0 for all z ∈ B∞}.
(9.1)
42
It was implicitly shown by Dixmier and Douady, and explicitly pointed out by Blanchard
and Kirchberg ([BK04a, Proposition 3.6]), that for any v ∈ E there exists z ∈ B∞ such
that hv, vi(z) = 0. That is, every section of E vanishes at some point (we will reprove this
fact in Proposition 9.10 below). Notice that
D = E + x · C(B∞) ∼= E ⊕ C(B∞).
It can be deduced from this that D ≇ l2(B∞) (see [DD63, Proposition 19]).
Let B3 denote the unit ball in R3. Let f ∈ M2(B3)+ be defined as
f (x, y, z) =
1
2 (cid:18) 1 + z x − iy
1 − z (cid:19) .
x + iy
(The function f is a homeomorphism from B3 to the set of positive elements of M2(C) with
trace 1. On the boundary 2-sphere of B3 it agrees with the tautological rank 1 projection.)
Consider the C(B3)-module associated to f :
F := f (cid:18)C(B3)
C(B3)(cid:19).
(9.2)
Proposition 9.10. Let B∞ and B3 be as before. Let X = Qi∈I Xi, where each Xi is either
B∞ or B3 and the index set I is non-empty. For each i, let Hi be the pull-back along the
projection map πi : X → Xi of either the module E defined in Example 9.9 or the module
F defined in (9.2). Finally, let H be the C(X)-module defined by H = Li∈I Hi. Then
C(X) does not embed in H as a C(X)-module (i.e., for every v ∈ H there exists z ∈ X
such that hv, vi(z) = 0).
Notice that if every Xi agrees with B3, the above proposition can be proven using standard
methods in algebraic topology (e.g., characteristic classes). Indeed, it suffices to restrict
to the boundary 2-sphere of each Xi and use that on that set F is the tautological rank 1
projective module. It is the inclusion of the spaces B∞ in the definition of X that forces
us to use a different route in the proof.
Proof. Let v ∈ H, and write v = Pi∈I vi, with vi ∈ Hi. In order to show hv, vi(z) = 0
for some z ∈ X, it suffices to prove this for v belonging to a dense submodule of H. For
suppose that (v(n)) is a sequence in H such that v(n) → v and hv(n), v(n)i(zn) = 0 for
some zn ∈ X. Then by the compactness of X there exists a subsequence (znk) such that
znk → z ∈ X, and so hv, vi(z) = 0. Thus, we may assume that the index set I is finite.
Furthermore, for the indices i such that Hi = π∗
i, where
H ′
i ⊆ Hi is the pull back along πi of the dense submodule E0 defined in (9.1).
i (E), we may assume that vi ∈ H ′
In the sequel, we assume that I = {1, 2, . . . , n}, Xi = B∞ for i = 1, 2, . . . , n1, and
Xi = B3 for i = n1 + 1, . . . , n, where n1 ≤ n.
We will argue by contradiction that hv, vi(z) = 0 for some z ∈ X. Suppose that hv, vi
is invertible, and assume without loss of generality that hv, vi = 1. Observe that, for each
43
i ≤ n1, vi is a function from X into the unit ball of l2(N) ⊕ C, while for n1 < i ≤ n
the entry vi is a function from X into the unit ball of C ⊕ C (let us denote it by B4).
Let h0 : l2(N) ⊕ C → l2(N) denote the projection onto the first direct summand and let
h1 : B4 → B3 denote the Hopf fibration (extended to the unit ball):
h1(z0, z1) := (2z0z1, z02 − z12).
Let λ : [0, 1] → [0, 1] be such that λ(0) = 0, λ(t) = 1 for t ∈ [ 1
Define h0, h1 : B4 → B3 by
n, 1], and λ is linear in [0, 1
n ].
h0(w) = h0(cid:16) λ(w)
w
w(cid:17),
h1(w) = −h1(cid:16)λ(w)
w
w(cid:17).
Consider the continuous map Φ : X → X given by the vector of functions
Φ := (h0 ◦ v1, h0 ◦ v2, . . . , h0 ◦ vn1, h1 ◦ vn1+1, . . . , h1 ◦ vn).
Since X is a compact convex subset of the vector space (l2(N))n1 × (R3)n−n1, the map Φ
has a fixed point by the Schauder fixed point theorem. Let z := (zi)n
i=1 ∈ X be a fixed
point of Φ. Since kv(z)k = 1, we must have kvi(z)k ≥ 1
n for at least one index i. Notice
that both h0 and h1 map all vectors of norm at least 1/n into the unit sphere of either B∞
or B3. It follows that the fixed point z satisfies kzik = 1 and zj = 0 for all j 6= i.
There are two cases to consider: i ≤ n1 and i > n1. Suppose that i ≤ n1. The general
form of vi ∈ H ′
kzik = 1, we have
i is f + (zi +p1 − kzik2e)α, for some f : X → l2(N) and α ∈ C(X). Since
vi(z) = f (z) + α(z)zi = zi.
But hf (z), zii + α(z) = 0. This contradicts that kzik = 1.
Suppose that i > n1. Since zi
7→ vi(· · · , zi, · · · ), with zi ∈ S2, is a section of the
tautological bundle on S2, we have h1 ◦ vi(z) = zi whenever zi ∈ S2.
It follows that
h1 ◦ vi(z) = −zi. But h1 ◦ vi(z) = zi, by the fixed point property of z. This again
contradicts that kzik = 1.
We are now prepared to give examples of weakly (ω, 2)-divisible elements which are not
properly infinite.
Example 9.11. Let X = B∞ × B3 and consider the Hilbert module H = π∗
2(F ),
described in the statement of the previous proposition. We have shown that [C(X)] (cid:2) [H].
In particular, [H] is not properly infinite (since it is full). Let us show that [H] is weakly
(ω, 2)-divisible. Consider the open sets U := B∞ × B+
3 and
B−
3 are (open) upper and lower hemispheres of B3 that together cover B3. We claim that
l2(U) ֒→ HC0(U) and l2(V ) ֒→ HC0(V ). Indeed,
3 and V := B∞ × B−
3 , where B+
1(E) ⊕ π∗
HC0(U) = π∗
= π∗
1(E)C0(U) ⊕ π∗
1(E)C0(U) ⊕ π∗
2(F )C0(U)
2(F C0(B+
3 )).
44
But F C0(B+
3 ) ∼= C0(B+
3 ) ⊕ C0(B+
3 \S2). Therefore,
HC0(U) = π∗
= π∗
1(E)C0(U) ⊕ C0(U) ⊕ F ′
1(E ⊕ C(B∞))C0(U) ⊕ F ′.
But l2(B∞) ֒→ D = E ⊕ C(B∞). Thus, l2(U) ֒→ HC0(U). Symmetrically, we have
that l2(V ) ֒→ HC0(V ). It follows that [HC0(U)] and [HC0(V )] are properly infinite, and
[H] ≤ [HC0(U)] + [HC0(V )]. Thus, [H] is weakly (ω, 2)-divisible.
Remark 9.12. The previous example answers a question posed in [KR00, Question 3.10]:
If a and b are properly infinite positive elements, is a + b properly infinite? In the language
of Hilbert modules, this question asks whether H is properly infinite if H = H1 + H2, and
H1, H2 ⊆ H are properly infinite submodules of H. We obtain a counterexample taking H
as in the previous example, H1 = HC0(U) and H2 = HC0(V ).
Example 9.13. In this example we answer (in the negative) the following question, posed
in [KR00, Question 3.4]: if [H] is properly infinite, is the unit of B(H) a properly infinite
projection? Let X = B∞ × (B3)∞ and consider the Hilbert C(X)-module
H = C(X) ⊕ π∗
1(E) ⊕
∞
Mi=2
π∗
i (F ).
The module C(X) ⊕ π∗
Since l2(C(B∞)) embeds in D, l2(C(X)) embeds in C(X) ⊕ π∗
1(E) is the pull back along π1 of the Dixmier-Douady module D.
1(E). Thus, [H] is properly
infinite. Also, the direct sum of the module L∞
i (F ) with itself gives l2(C(X)) (because
F ⊕ F contains C(B3) as a direct summand). Therefore, H ⊕ H ∼= l2(C(X)). However, H
is not isomorphic to l2(C(X)), because every section of π∗
i (F ) vanishes, and
so adding the trivial rank 1 module to it cannot yield the trivial Hilbert module l2(C(X))
(see the proof of D ≇ l2(C(B∞)) in [DD63, Proposition 19]). It follows that H ⊕ H is not
a direct summand of H, i.e., the unit of B(H) is not properly infinite.
1(E) ⊕L∞
i=2 π∗
i=2 π∗
References
[BK04a]
E. Blanchard and E. Kirchberg, Global Glimm halving for C ∗-bundles, J. Operator
Theory 52 (2004), no. 2, 385 -- 420.
[BK04b]
, Non-simple purely infinite C ∗-algebras: the Hausdorff case, J. Funct. Anal.
207 (2004), no. 2, 461 -- 513.
[BPT08]
N. Brown, F. Perera, and A. S. Toms, The Cuntz semigroup, the Elliott conjecture,
and dimension functions on C ∗-algebras, J. Reine Angew. Math. 621 (2008), 191 --
211.
[CEI08]
K. T. Coward, G. A. Elliott, and C. Ivanescu, The Cuntz semigroup as an invariant
for C ∗-algebras, J. Reine Angew. Math. 623 (2008), 161 -- 193.
45
[DD63]
J. Dixmier and A. Douady, Champs continus d'espaces hilbertiens et de C ∗-alg`ebres,
Bull. Soc. Math. France 91 (1963), 227 -- 284.
[DHTW09] M. Dadarlat, I. Hirshberg, A. Toms, and W. Winter, The Jiang-Su Algebra does not
always embed, Math. Res. Lett. 16 (2009), no. 1, 23 -- 26.
[DT09]
M. Dadarlat and A. Toms, Z-stability and infinite tensor powers of C ∗-algebras, Adv.
Math. 220 (2009), no. 2, 341 -- 366.
[Dup76]
M. J. Dupr´e, Classifying Hilbert bundles. II, J. Funct. Anal. 22 (1976), no. 3, 295 -- 322.
[ER06]
[ERS]
[HR98]
G. A. Elliott and M. Rørdam, Perturbation of Hausdorff moment sequences, and an
application to the theory of C ∗-algebras of real rank zero, Operator Algebras: The
Abel Symposium 2004, Abel Symp., vol. 1, Springer, Berlin, 2006, pp. 97 -- 115.
G. A. Elliott, L. Robert, and L. Santiago, The cone of lower semicontinuous traces
on a C*-algebra, American J. Math., to appear.
J. Hjelmborg and M. Rørdam, On stability of C ∗-algebras, J. Funct. Anal. 155 (1998),
no. 1, 153 -- 170.
[HRW07]
I. Hirshberg, M. Rørdam, and W. Winter, C0(X)-algebras, stability and strongly
self-absorbing C ∗-algebras, Math. Ann. 339 (2007), no. 3, 695 -- 732.
[Hus94]
[Kir06]
D. Husemoller, Fibre Bundles, 3rd. ed., Graduate Texts in Mathematics, no. 20,
Springer Verlag, New York, 1966, 1994.
E. Kirchberg, Central sequences in C ∗-algebras and strongly purely infinite C ∗-
algebras, Operator Algebras (Berlin) (S. Neshveyev C. Skau O. Bratteli, ed.), Abel
Symp., vol. 1, Springer, 2006, pp. 175 -- 232.
[KOS03]
A. Kishimoto, N. Ozawa, and S. Sakai, Homogeneity of the pure state space of a
separable C ∗-algebra, Canad. Math. Bull. 46 (2003), no. 3, 365 -- 372.
[KR00]
E. Kirchberg and M. Rørdam, Non-simple purely infinite C ∗-algebras, American J.
Math. 122 (2000), 637 -- 666.
[KR02]
, Infinite non-simple C ∗-algebras: absorbing the Cuntz algebra O ∞, Advances
in Math. 167 (2002), no. 2, 195 -- 264.
[MS74]
[OPR]
[PR04]
[PT07]
J.W. Milnor and J.D. Stasheff, Characteristic classes, no. 76, Princeton Univ Pr,
1974.
E. Ortega, F. Perera, and M. Rørdam, The Corona Factorization property, Stability,
and the Cuntz semigroup of a C*-algebra., Int. Math. Res. Not. IMRN. To appear.
F. Perera and M. Rørdam, AF-embeddings into C ∗-algebras of real rank zero, J.
Funct. Anal. 217 (2004), no. 1, 142 -- 170.
F. Perera and A. S. Toms, Recasting the Elliott conjecture, Math. Ann. 338 (2007),
no. 3, 669 -- 702.
46
[Rob11]
L. Robert, The cone of functionals on the Cuntz semigroup, 2011.
[Rør97]
M. Rørdam, Stability of C ∗-algebras is not a stable property, Documenta Math. 2
(1997), 375 -- 386.
[Rør03]
, A simple C ∗-algebra with a finite and an infinite projection, Acta Math. 191
(2003), 109 -- 142.
[Rør04]
, The stable and the real rank of Z-absorbing C ∗-algebras, Internatinal J.
Math. 15 (2004), no. 10, 1065 -- 1084.
[RW10]
M. Rørdam and W. Winter, The Jiang-Su algebra revisited, J. Reine Angew. Math
642 (2010), 129 -- 155.
[Vil98]
[Win]
J. Villadsen, Simple C ∗-algebras with perforation, J. Funct. Anal. 154 (1998), no. 1,
110 -- 116.
W. Winter, Nuclear dimension and Z-stability of pure C ∗-algebras, Invent. Math.,
to appear.
Department of Mathematical Sciences, University of Copenhagen, Univer-
sitetsparken 5, DK-2100 Copenhagen Ø, Denmark
E-mail address: [email protected]
Department of Mathematical Sciences, University of Copenhagen, Univer-
sitetsparken 5, DK-2100 Copenhagen Ø, Denmark
E-mail address: [email protected]
47
|
1102.4494 | 1 | 1102 | 2011-02-22T13:13:03 | Noncommutative maximal ergodic inequality for non-tracial L1-spaces | [
"math.OA"
] | We extend the noncommutative L1-maximal ergodic inequality for semifinite von Neumann algebras established by Yeadon in 1977 to the framework of noncommutative L1-spaces associated with sigma-finite von Neumann algebras. Since the semifnite case of this result is one of the two essential parts in the proof of noncommutative maximal ergodic inequality for tracial Lp-spaces (1<p<infinity) by Junge-Xu in 2007, we hope our result will be helpful to establish a complete noncommutative maximal ergodic inequality for non-tracial Lp-spaces in the future. | math.OA | math |
Noncommutative maximal ergodic inequality for
non-tracial L1-spaces
Qin Zhang ∗
Graduate School of Mathematical Sciences, The University of Tokyo
Abstract
We extend the noncommutative L1-maximal ergodic inequality for semifinite von
Neumann algebras established by Yeadon in 1977 to the framework of noncommutative
L1-spaces associated with σ-finite von Neumann algebras. Since the semifinite case of
this result is one of the two essential parts in the proof of noncommutative maximal
ergodic inequality for tracial Lp-spaces (1 < p < ∞) by Junge-Xu in 2007, we hope
our result will be helpful to establish a complete noncommutative maximal ergodic
inequality for non-tracial Lp-spaces in the future.
1
Introduction
Theory of von Neumann algebras is regarded as the noncommutative measure and inte-
gration theory (Chapter IX, [28]), so it is natural to consider extensions of classical ergodic
theorems for spaces of measurable functions to the framework of noncommutative spaces
associated with von Neumann algebras. Such an extension topic appeals to many math-
ematicians and they had interesting results even from 1970's (for example, the pioneering
works of [17] and [30]). Since then many classical mean ergodic theorems and ergodic theo-
rems of other types were successfully transformed to the noncommutative context which is a
semifinite von Neumann algebra or a non-commutative Lp-space associated with a semifinite
von Neumann algebra, and some authors even considered a general von Neumann algebra or
a non-tracial Lp-space associated with it. Among these noncommutative ergodic theorems,
however, the problem of finding a noncommutative analogue of the famous Dunford-Schwartz
maximal ergodic inequality ([7]) was left open until the appearance of Junge-Xu's prominent
∗Address: 3-8-1 Komaba, Meguro, Tokyo, 153-8914, Japan. The author is supported by the
Japanese Government Scholarship (No.052111). Email Address: [email protected]
1
paper in 2007. The main obstacle in this problem is that it is difficult to define the supremum
of a sequence of operators even in the finite-dimensional Hilbert space cases, although it is
straightforward to take the supremum of a sequence of measurable functions. For this rea-
son, many powerful techniques in classical ergodic theory involving maximal functions seem
no longer available in the research of noncommutative ergodic results. This difficulty was
overcome in Junge-Xu's work ([15]) by using the noncommutative vector valued Lp-space
theory developed by Pisier and Junge (see [21] for the case of hyperfinite von Neumann alge-
bras and [14] for that of general ones). Junge-Xu established noncommutative version of the
Dunford-Schwartz maximal ergodic inequality first for non-commutaive Lp-spaces associated
with a semifinite von Neumann algebra, or in other words, tracial Lp-spaces (Theorems 4.1,
[15]). In order to state their result we need some notations as follows.
Let M be a semifinite von Neumann algebra equipped with a normal semifinite faithful
trace τ , and let Lp(M) be the associated noncommutative Lp-space (see, for example, [19]
for a detailed definition). The symbol T : M → M denotes a linear map which satisfies the
following conditions.
(J1) T is a contraction on M: T x ≤ x for all x ∈ M, where · means the ∞-norm,
or in other words, the usual operator norm.
(J2) T is positive: T x ≥ 0 if x ≥ 0.
(J3) τ ◦ T ≤ τ : τ (T (x)) ≤ τ (x) for all x ∈ L1(M) ∩ M+.
Theorem 1.1. (Theorem 4.1 in [15]). Let T be a linear map on a semifinite von Neumann
algebra M with (J1) -- (J3), then T extends naturally to a contraction on Lp(M) for 1 ≤ p <
∞. Put
Sn ≡ Sn(T ) =
1
n + 1
then for every p, 1 < p < ∞, we have
T k,
nXk=0
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)sup
n
+Sn(x)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)p
≤ cp xp , for all x ∈ Lp(M),
for some positive constant cp depending only on p.
Note that sup+
n Sn(x)p is the notation for {Sn(x)}n∈NLp(M ; l∞) used in [15], where
Lp(M;
l∞) is the Pisier-Junge's noncommutative vector valued Lp-space defined by the
space of all sequences x = {xn}n∈N in Lp(M) (1 ≤ p ≤ ∞) such that each sequence admits a
2
factorization of the following form: there are a, b ∈ L2p(M) and y = {yn}n∈N ⊂ L∞(M) = M
such that xn = aynb for all n ∈ N and the norm is defined by
xLp(M ; l∞) ≡ inf{a2p sup
n∈N
yn b2p},
where the infimum runs over all such factorizations as above.
For details of such spaces, we refer the reader to [21] and [14].
But considering only semifinite von Neumann algebras is not enough sometimes. It was
stated in [28] that most of factors arising from physics are of type III, which are of course
not semifinite. Another fact is that it was shown by Pisier ([22]) that OH cannot completely
embed in a semifinite L1(M). More and more recent works concern the type III case or
need results on noncommutative Lp-spaces associated with a not necessarily semifinite von
Neumann algebra.
For a general von Neumann algebra, there are several equivalent constructions of non-
commutative Lp-spaces associated with it, and the important ones include [2, 6, 9, 11, 13,
16, 29] and they are all based on the Tomita-Takesaki theory. Since a general von Neumann
algebra does not necessarily admit a normal semifinite faithful trace, any of these (equivalent)
constructions are called non-tracial Lp-spaces.
The first non-tracial Lp-spaces are Haagerup's ones, and just for Haagerup's Lp-spaces,
Junge-Xu established the non-commutative Dunford-Schwartz maximal ergodic inequality as
a non-tracial extension of Theorem 1.1 above (see Theorem 7.4, [15]). Their method is to use
an early result of Haagerup named reduction method to approximate Haagerup's Lp-spaces
by simifinite ones. We state this theorem here.
Theorem 1.2. (Theorem 7.4 in [15]). Suppose M is a (σ-finite) von Neumann algebra with
a normal faithful state ϕ, let T be a linear map on M satisfying the following properties
(H1) -- (H4).
(H1) T is a contraction on M: T x ≤ x for all x ∈ M.
(H2) T is completely positive.
(H3) ϕ(T (x)) ≤ ϕ(x) for all x ∈ M+.
(H4) T ◦ σϕ
t = σϕ
t ◦ T for all t ∈ R.
3
Then T extends naturally to a contraction on Haagerup's Lp(M) for 1 ≤ p < ∞. Put
Sn ≡ Sn(T ) =
1
n + 1
nXk=0
T k.
Then for every p, 1 < p < ∞, we have
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)sup
n
for some positive constant cp depending only on p.
≤ cp xp , for any x ∈ Lp(M),
+Sn(x)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)p
This theorem was also established in [10] (Theorem 7.9).
We may compare the statement of this theorem with that of Theorem 1.1. It is clear
that the assumption (J2) is reinforced to the assumption (H2), and (H4) is new, the others
are natural extensions. Since the modular automorphism group στ
t
trivial, we do not think the condition (H4) is unnatural any more when we consider the
induced by a trace τ is
non-tracial cases. However, we feel that the "complete positivity" assumption of (H2) seems
a little bit stronger. Although it is remarked in [15], page 425, that the complete positivity
is unlikely to be really necessary, they did not give a proof to release this assumption, and
they still used complete positivity assumption in [10].
We would like to remark that Haagerup's reduction method used in [10] is really pow-
erful to extend the noncommutative martingale inequalities from the tracial cases to the
non-tracial ones without any change in the statement form for almost all those results. This
is because a conditional expectation in the framework of von Neumann algebras is certainly
completely positive ([1] and [27]).
In noncommutative ergodic theorems, "complete pos-
itivity" seems a little restrictive, moreover, many important positive but not completely
positive state-preserving transforms describing open quantum evolution are now considered
by mathematical physicists (see, for example, [24, 25, 12, 5, 26]).
So we do not think Theorem 1.2 is a complete answer for the problem of non-tracial
extension of Theorem 1.1. Recalling the proof of Theorem 1.1 in [15], we know that this proof
consists of two essential parts, one is the noncommutative maximal ergodic inequality for
semifinite L1(M) established by Yeadon, and the other one is the noncommutative semifinite
Marcinkiewicz interpolation theorem in [15]. A natural and possible way of thought for the
problem of a complete non-tracial extension of Theorem 1.1 is to get non-tracial extensions
of these two parts. The main result of our paper is just to provide the non-tracial extension
of maximal ergodic inequality for the non-tracial L1(M). We have been unable to get a
noncommutative Marcinkiewicz interpolation theorem for the non-tracial Lp-spaces so far
4
and we will explain the reason in the next section. We hope the result in this paper will
be helpful to establish a complete noncommutative maximal ergodic inequality for the non-
tracial Lp-spaces in the future.
It is well-known that all the von Neumann algebras encountered in quantum statistical
mechanics and quantum field theory are σ-finite ([4], p.84), probably for this reason, only
the noncommutative Lp-spaces associated with a σ-finite von Neumann algebra are consid-
ered in Theorem 1.2 instead of general non-tracial situation and we think this restriction to
the σ-finite case is of full meaning. Our framework is the Hilsum spatial Lp-spaces associ-
ated with a σ-finite von Neumann algebra which are certainly isometrically isomorphic to
Haagerup's ones associated with a σ-finite von Neumann algebra. On the detailed construc-
tion of the Hilsum spatial Lp-spaces and the close relation between those spaces and complex
interpolation theory, we will give a short description on them in the beginning of the next
section.
Let M be a σ-finite von Neumann algebra represented on a Hilbert space H, hence we
may assume M admits a normal faithful state ϕ, and ψ is a normal faithful state on the
commutant M ′ of M, let Lp(M; ψ) denote the Hilsum spatial Lp-spaces with respect to ψ.
Our main result is the following theorem.
Theorem 1.3. (Theorem 2.1 in the next section). Assume M is a σ-finite von Neumann
algebra which admits a normal faithful state ϕ and T is a linear map on M satisfying the
following conditions.
(1) T is a contraction of M, i.e., T x ≤ x for all x ∈ M.
(2) T is positive, i.e., T y ≥ 0 if y ∈ M+.
(3) ϕ(T (y)) ≤ ϕ(y), for all y ∈ M+,
Then T extends to a positive linear contraction on L1(M; ψ). For any a ∈ L1
+(M; ψ) and
any λ > 0, and any n ∈ N, there exists a projection en ∈ M such that
enSr(a)en ≤ λenden for all r ∈ {0, 1, ..., n},
and
where Sr(a) = 1
Furthermore, for any a ∈ L1
r+1Pr
ϕ(1 − en) ≤
2
λZ adψ,
k=0 T k(a), d = dϕ
dψ is the spatial derivative and 1 is the identity of M.
+(M; ψ), there exists a projection e ∈ M such that for any r ∈ N,
Z eSr(a)edψ ≤ 4λ, and ϕ(1 − e) ≤
2
λZ adψ.
5
We recall that in the semifinite case, the L1-norm of the tracial L1-space L1(M) is
defined by ·1 = τ ( · ), and in the spatial L1-space L1(M; ψ), the L1-norm is defined by
·1 = (R ·dψ), and we know that when M is semifinite, L1(M; ψ) is equivalent to the tracial
L1(M), so this is the reason we useR ·dψ in the statement of our theorem. In the semifinite
case, Yeadon's result ([30]) provided a bound on eSr(a)e, and this is because in semifinite
case, the elements in Lp-spaces are unbounded (also including bounded) operators affiliated
with the von Neumann algebra, hence we may cut such unbounded operators by using some
projection in the von Neumann algebra. However, for the non-tracial case, elements in the
Hilsum (or other equivalent) Lp-spaces are never affiliated with the von Neumann algebra
M, so eSr(a)e cannot be majorized by any bounded operator, and the result in our theorem
is the best one we can hope in the non-tracial case.
Such a result also has its own value. In the commutative case we know that the following
Chebyshev type inequality is quite important in measure theory and probability theory,
µ({x : sup
n∈N
sn(f )(x) > λ}) ≤
c
λ
f 1 ,
and our result is a non-commutative extension of this inequality for σ-finite von Neumann
algebras.
2 Main Part
Now we introduce the definition of the Hilsum spatial Lp-spaces. We assume M is a
general von Neumann algebra (because the Hilsum Lp-spaces are defined for general von
Neumann algebras) which admits a normal semifinite faithful weight ϕ and furthermore, M
is represented on a Hilbert space H and we have a normal semifinite faithful weight ψ on
the commutant M ′ of M.
A vector ξ ∈ H is said to be ψ-bounded if there exists a positive constant c such that
yξ2 ≤ cψ(y ∗y) for any y ∈ nψ, where nψ = {y ∈ M ′ ψ(y ∗y) < ∞}. We let D(H, ψ) be
the subspace of H consisting of all ψ-bounded vectors. Then for any ξ ∈ D(H, ψ), Rψ(ξ)
is the unique bounded operator from Hψ (the GNS representation Hilbert space induced
by ψ) to H such that Rψ(ξ)ηψ(y) = yξ, where ηψ is the canonical injection of nψ into Hψ,
and θψ(ξ, ξ) = Rψ(ξ)Rψ(ξ)∗ ∈ M. The map ξ → φ(θψ(ξ, ξ)) defines a lower semicontinuous
positive form on D(H, ψ), where φ is any normal semifinite weight on M. Then the positive
self-adjoint operator associated with this form is called the spatial derivative dφ
dψ defined by
Connes ([6]).
6
The Hilsum spatial Lp-space Lp(M; ψ) is defined as (1 ≤ p < ∞)
Lp(M; ψ) =
=
R apdψ < ∞
The Lp-norm is defined by ·p = (R · pdψ)
operator norm.
a is a closed densely defined operator on H with
polar decomposition a = ua such that u ∈ M
and ap = dφ
dψ for some φ ∈ M +
∗
a is a closed densely defined operator on H and
(− 1
p )-homogeneous with respect to ψ such that
1
p . If p = ∞, L∞(M; ψ) = M with the usual
For the definition of homogeneity with respect to ψ and the detailed properties of dφ
dψ and
Lp(M; ψ), we refer the reader to [6, 11, 29].
Concerning the Hilsum spatial Lp-spaces, Terp's paper ([29]) deeply revealed the close
relation between the complex interpolation theory on (M, M∗) and Lp(M; ψ).
The subspace L of M consists of x ∈ M for which there exists a ϕx ∈ M∗ such that
for any y, z ∈ nϕ : (z∗y, ϕx) = hJπϕ(x)∗Jηϕ(y), ηϕ(z)i,
where J is the associated modular conjugation in Hϕ, (·, ·) denotes the duality between M
and M∗, and h , i is the scalar inner product on Hϕ .
For x ∈ L, we put xL = max{x , ϕx}, where the norm for ϕx means the functional
norm.
The normed space (L, ·L) is a Banach space and it can be embedded naturally into M
and M∗ by x 7→ x : L → M and x 7→ ϕx : L → M∗.
By transposition of the above two embeddings we have the injections M → L∗ and
M∗ → L∗ given by
for all y ∈ M and
(x, y)(L,L∗) = (y, ϕx)(M,M∗), x ∈ L
(x, φ)(L,L∗) = (x, φ)(M,M∗), x ∈ L
for all φ ∈ M∗, where L∗ means the dual of (L, ·L).
The following diagram commutes,
DDDDDDDD
={{{{{{{{
L∗
M
>
AAAAAAAA
L
M∗
7
!
!
>
=
and L = M ∩ M∗ when M and M∗ are considered as subspaces of L∗ (see Section 1, [29]),
and L is σ-weakly dense in M, the embedding of L in M∗ is weakly and norm dense in M∗
(Corollary 5, [29]).
Hence (M, M∗) is turned into a compatible pair of Banach spaces in the complex inter-
polation sense (Section 2.3, [3]), and Terp proved that for 1 < p < ∞, the Hilsum spatial
Lp-spaces Lp(M; ψ) are just the complex interpolation spaces of M∗ and M. Accurately
speaking, Lp(M; ψ) = C 1
(M, M∗) (Theorem 36 in [29]).
p
Let T be a linear map on M satisfying the following conditions.
(1) T is a contraction of M, i.e., T x ≤ x for all x ∈ M.
(2) T is positive, i.e., T y ≥ 0 if y ∈ M+.
(3) ϕ(T (y)) ≤ ϕ(y) for all y ∈ L+.
From now on, we will concentrate on the σ-finite cases.
Let M be a σ-finite von Neumann algebra acting standardly on the Hilbert space H,
M admits a normal faithful state ϕ, ψ is a normal faithful state on M ′, and T is a linear
transform satisfying the above conditions (1) − (3).
We note that for any x ∈ M, we have x ∈ mϕ = span{x ∈ M+ ϕ(x) < ∞} as ϕ is a
state, and we know from Note (2), p.329 in [29] that mϕ ⊆ L, combined with the fact that
L is defined to be a linear subspace of M, we have mϕ = M = L in the σ-finite cases.
That is to say, condition (3) above can be replaced by the following one when M is
σ-finite.
(3) ϕ(T (y)) ≤ ϕ(y) for all y ∈ M+.
Moreover, the embedding of M = L into L1(M; ψ) is given by the map M → d
2 ,
where d = dϕ
dψ is the spatial derivative of ϕ with respect to ψ (see Section 2.3 and Theorem 27
in [29]). Such an embedding is equivalent to the embedding M → M η ⊆ M∗ with η = 1
2 (the
symmetric case) in Definition 7.1 of [16] and Lp(M; ψ) is equivalent to Kosaki's Lp-spaces
2 (Definition 7.2, [16]). Moreover, since the Hilbert space is standard,
C 1
the Hilsum space Lp(M; ψ) is now as the same as the Araki-Masuda Lp-space in [2].
(M η, M∗) with η = 1
p
1
1
2 M d
Then for a general p, 1 ≤ p < ∞, we may define the following map Tp as
for any x ∈ M.
Tp : d
1
2p M d
1
2p → d
1
2p M d
1
2p
1
d
2p xd
1
2p 7→ d
1
2p T (x)d
1
2p ,
8
We claim that the map T1 defined above extends naturally to a positive contraction of
L1(M; ψ) → L1(M; ψ), and it will still be denoted by T1. To show this claim, we need the
Lemma 5.2 in [10]. Although this lemma is stated and proved for Haagerup's L1-spaces, it is
still valid in the framework of the Hilsum spatial L1-spaces through isometric isomorphism.
We state it here for L1(M; ψ). Let x ∈ M and x is self-adjoint, then
= inf{ϕ(a) + ϕ(b) x = a − b, a, b ∈ M+}.
1
2 xd
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)d
1
2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)1
The proof of this result for L1(M; ψ) is essentially the same as that of Lemma 5.2 in [10]
when replacing D there by the spatial derivative d. Then we will use this result and follow
the method of Lemma 5.3 in [10] to show our claim. Let y ∈ M+, then d
T1(d
2 ≥ 0, hence T1 is also positive. By condition (3),
2 ≥ 0, so
2 T (y)d
2 ) = d
2 yd
2 yd
1
1
1
1
1
1
Now assume x ∈ M and x is self-adjoint, for any ε > 0, there exist a, b ∈ M+ such that
x = a − b and
It follows that
1
1
1
1
2 ad
2 T (y)d
2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)1
=Z d
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)d
=ϕ(T (y)) ≤ ϕ(y) =(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)d
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)d
+(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)d
2 )(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)1
≤ (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)d
2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)1
≤(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)d
2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)1
2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)1
2 T (a)d
2 xd
2 bd
1
1
1
1
1
1
2 T (y)d
1
1
1
.
2 yd
2 dψ =Z T (y)ddψ
2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)1
≤(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)d
2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)1
+(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)d
2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)1
≤(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)d
2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)1
2 T (b)d
+ ε.
2 xd
2 xd
1
1
1
1
1
1
2 xd
1
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)T1(d
2 )(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)1
1
+ ε.
1
2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)1
1
1
1
1
1
2 xd
2 xd
. And since d
for any x ∈ M is self-adjoint since ε is arbitrary. Fi-
That is (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)T1(d
nally, decomposing any x ∈ M into its real and imaginary parts, we get (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)T1(d
2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)1
2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)d
2 is norm dense in L1(M; ψ), T1 extends to a bounded pos-
itive map on L1(M; ψ) with T1 ≤ 2. Thus it remains to reinforce the norm bound 2 to
1 which is a linear map on M = L1(M; ψ)∗.
1. To this end, we consider the adjoint map T ∗
1 attains its norm at the identity 1 of M, i.e.,
Since T1 is positive, T ∗
T1 = T ∗
1 (1) ≤ 1. Indeed, by condition
(3),
1 (1). Hence we are reduced to showing T ∗
1 is also positive and T ∗
2 )(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)1
1 = T ∗
2 M d
1
2 xd
≤
1
1
1 (1)d
1
2 yd
1
2 dψ =Z T1(d
1
2 yd
1
2 )dψ = ϕ(T (y)) ≤ ϕ(y) =(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)d
2 in L1
2 M+d
1
1
1
2 yd
1
2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)1
for any y ∈ M+. We get T ∗
1 (1) ≤ 1 by the density of d
+(M; ψ), and hence
Z T ∗
2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)1
≤(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)d
1
1
2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)1
1
2 T (x)d
1
2 xd
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)d
for any x ∈ M and our claim follows.
9
Combined with condition (1) and the abstract Riesz-Thorin Theorem, we obtain that
1
2p T (x)d
1
2p xd
,
for 1 < p < ∞, and x ∈ M.
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)d
1
2p(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)p
≤(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)d
1
2p(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)p
1
1
p
2p M d
(M∗, M) (Theorem 36 in [29], or Kosaki's
Moreover, since for 1 < p < ∞, Lp(M; ψ) = C 1
noncommutative interpolation theorem in [16]), in accordance with Theorem 4.2.2(a) in [3],
2p is ·p-norm dense in Lp(M; ψ). Therefore, the map Tp defined above
we have that d
also extends naturally to a positive contraction of Lp(M; ψ) → Lp(M; ψ) for each p and we
still denote it by Tp (1 < p < ∞). From the idea in [29], M = L∞(M; ψ), M∗ = L1(M; ψ) and
Lp(M; ψ), 1 < p < ∞, can be regarded as injective subspaces in M + M∗ (factually it is just
M∗ for the σ-finite cases, see Definition 7.2 in [16]). In this situation it is easily seen that the
maps T, Tp (1 < p < ∞) and T1 defined above coincide on L. For this reason, we may have a
linear map on M + M∗, and the restriction of this map on M, Lp(M; ψ) (1 < p < ∞) and M∗
will be T , Tp (1 < p < ∞) and T1, respectively when considering M, Lp(M; ψ) (1 < p < ∞)
and M∗ in M + M∗. Since this map on M + M∗ is deduced by T on M, it is viewed as the
extension of T on M + M∗, and we still denote it by T .
Then we arrive at the stage to show our main result. From above we know that the
restriction of T on L1(M; ψ) satisfies the following conditions.
• T (a)1 ≤ a1, for all a ∈ L1
• T is positive, i.e., T (a) ≥ 0 if a ∈ L1
+(M; ψ).
+(M; ψ).
Theorem 2.1. We assume T is a linear transform on a σ-finite von Neumann algebra M
satisfying the conditions (1) − (3) above, and T extends to a linear positive contraction on
L1(M; ψ). Then for any a ∈ L1
+(M; ψ) and any λ > 0, and any n ∈ N, there exists a
projection en ∈ M such that
enSr(a)en ≤ λenden for all r ∈ {0, 1, ..., n},
and
ϕ(1 − en) ≤
2
λZ adψ,
where Sr(a) = 1
Furthermore, for any a ∈ L1
k=0 T k(a), d = dϕ
dψ is the spatial derivative and 1 is the identity of M.
+(M; ψ), there exists a projection e ∈ M such that for any r ∈ N,
r+1Pr
Z eSr(a)edψ ≤ 4λ, and ϕ(1 − e) ≤
2
λZ adψ.
10
compute this value explicitly as follows.
Proof. For any x ∈ M+, R T (a)xdψ is a positive linear functional on L1(M; ψ) since T
R T (a)xdψ takes positive values for any a ∈ L1
Indeed, the property M = (L1(M; ψ))∗ implies that each x ∈ M+ gives a
is positive.
positive linear functional on L1(M; ψ), and T (a) is still positive, hence the functional action
+(M; ψ). Or in other words, we may also
2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) T (a)1,
2 ∈ M+, in accordance with Proposition 8(4) in [11],(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)x
2 T (a) ∈ L1(M; ψ). The Proposition in page 159 of [11] yields that R T (a)xdψ =
2 dψ. For any family {ξα} in D(H, ψ) such that Pα θψ(ξα, ξα) = 1, hence from
R x
the definition ofR ·dψ in page 163 of [6], we have
2 T (a)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)1
≤(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)x
Since x
2 T (a)x
i.e., x
1
1
1
1
1
1
1
2 T (a)x
1
hx
2 T (a)x
1
2 ξα, ξαi
1
2 dψ =Xα
Z x
=Xα
1
1
hT (a)
2 x
2 ξα, T (a)
1
2 x
1
2 x
≥ 0.
1
2 ξαi =Xα (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)T (a)
2
1
2 ξα(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
+(M; ψ) and x ∈ M+.
That is to say,R T (a)xdψ ≥ 0 for any a ∈ L1
p + 1
If the reader is more familiar with the properties of Haagerup's Lp-spaces, we may recall
the Proposition 1.20 in [9], i.e., let p, q ∈ [1, ∞], 1
q = 1 and let a ∈ Lq(M), then
a ≥ 0 if and only if tr(ab) ≥ 0 for any b ∈ Lp(M)+, and Hilsum's spatial Lp-spaces are
get the positivity of the functional at the beginning of this proof.
isometrically isomorphic to Haagerup's Lp-spaces, by replacing tr(·) byR ·dψ, we may also
Therefore there exists some ex ∈ M+ such that R T (a)xdψ = R aexdψ again by the fact
M = (L1(M; ψ))∗. If we denoteex by eT (x) for each x ∈ M+, then
and eT extends linearly to be a linear transform on M such that eT (x) ≥ 0 if x ≥ 0.
Z T (a)xdψ =Z aeT (x)dψ,
Also we note that for any x ∈ M+,
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)eT (x)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) = sup(cid:26)Z aeT (x)dψ a1 ≤ 1, a ∈ L1
= sup(cid:26)Z T (a)xdψ a1 ≤ 1, a ∈ L1
+(M; ψ)(cid:27)
+(M; ψ)(cid:27)
≤ x ,
where the last inequality is because T (a)1 ≤ a1 for any a ∈ L1
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)eT (x)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ x can also be obtained from the fact that T is a contraction on L1(M; ψ) and eT
is just the adjoint map of T .
+(M; ψ). This result
11
Let n ≥ 1 be fixed, we put
K = {(x0, x1, . . . , xn) xr ∈ M+ for 0 ≤ r ≤ n and
xr ≤ 1},
nXr=0
then K is σ-weakly compact in M × M × · · · × M of (n + 1)-copies.
1
1
1
1
2 xd
2 = (x
is given by µ1(x) = d
We know from Section 2.3 in [29] that the embedding map µ1 of L+ into L1(M; ψ)+
dψ is the spatial
dψ dψ = ϕx(1) < ∞ since ϕx is
a positive normal linear functional on M. And for any x ∈ M+, as M = L in σ-finite case,
2 ) for x ∈ L+, where d = dϕ
derivative. Hence we may consider the valueR d
R d
2 dψ < ∞ for any x ∈ M+ is well-defined.
Therefore for any a ∈ L+
1 (M; ψ), we may define a function g on K by
2 =R dϕx
2 )∗(x
2 xd
2 xd
2 d
2 d
1
1
1
1
1
1
g((x0, x1, ..., xn)) =
(r + 1)Z Sr(a)xrdψ − λ
nXr=0
1
1
2 dψ.
2 xrd
(r + 1)Z d
nXr=0
+(M; ψ), so R Sr(b)xrdψ is well-
Note that Sr(b)xr ∈ L1(M; ψ) as xr ∈ M and Sr(b) ∈ L1
defined and takes finite values for each 0 ≤ r ≤ n. We recall that K is σ-weakly compact
and g is σ-weakly continuous on K, hence the finite maximum value of g is attained for some
choice of (x0, x1, . . . , xn) in K. Then we explain the reason that g is σ-weakly continuous.
dψ xrdψ = φr(xr) for
2 dψ =
dψ xrdψ = ϕ(xr), for xr ∈ M, is σ-weakly continuous on M for ϕ is a normal
For each r ∈ N∪{0}, there exists a φr ∈ M∗ such thatR Sr(a)xrdψ =R dφr
xr ∈ M, and φr is normal, i.e., φr is σ-weakly continuous on M. Also we haveR d
R dxrdψ =R dϕ
We let 1−Pn
on n). For any x ∈ M+ with x ≤ zn, we havePn
g((x0, x1, . . . , xn)) ≥ g((xr + δ(r, r0)x)r=0,1,...,n), where δ(r, r0) =(1 r = r0
r=0 xr = zn (we choose this notation because the positive operator zn depends
r=0 xr + x ∈ M+ and
r=0 xr + x ≤ 1 andPn
state on M. Thus g is σ-weakly continuous on K.
hence for any fixed r0 ∈ {0, 1, . . . , n},
0 r 6= r0.
2 xrd
1
1
As a result, we have
nXr=0
nXr=0
≥
2 dψ
1
1
2 xrd
nXr=0
(r + 1)Z d
(r + 1)Z Sr(a)xrdψ − λ
(r + 1)Z Sr(a)xrdψ + (r0 + 1)Z Sr0(a)xdψ−
(r + 1)Z d
nXr=0
2 dψ − λ(r0 + 1)Z d
2 xrd
2 xd
1
1
1
1
2 dψ.
− λ
12
Then we get
(r0 + 1)Z Sr0(a)xdψ ≤ λ(r0 + 1)Z d
1
2 xd
1
2 dψ for any r0 ∈ {0, 1, . . . , n}.
From the above inequality, for any x ∈ M+, x ≤ zn, any r in {0, 1, . . . , n}, we have
(2.1)
(2.2)
Take y = (yr)r=0,1,...,n with
and we havePn−1
and for any a ∈ L1
Therefore eT (1) ≤ 1 and we getPn−1
As a result, we obtain that g(x0, x1, . . . , xn) ≥ g(y). That is to say,
0
1
1
2 xd
2 dψ.
r = n,
n−1Xr=0
+(M; ψ), we have
Z Sr(a)xdψ ≤ λZ d
yn =(eT (xr+1) 0 ≤ r ≤ n − 1,
r=0 eT (xr+1) ≤ 1. Indeed, because eT is linear,
n−1Xr=0 eT (xr+1) = eT (
xr+1) ≤ eT (1),
Z aeT (1)dψ =Z T (a)dψ = T (a)1 ≤ a1 .
r=0 eT (xr+1) ≤ 1, so y is in K.
(r + 1)Z d
(r + 1)Z Sr(a)xrdψ − λ
nXr=0
nXr=0
(r + 1)Z T (Sr(a))xr+1dψ − λ
n−1Xr=0
n−1Xr=0
2eT (xr+1)d
(r + 1)Z T (Sr(a))xr+1dψ
(r + 1)Z Sr(a)xrdψ −
n−1Xr=0
nXr=0
(r + 1)Z d
(r + 1)Z d
n−1Xr=0
nXr=0
2eT (xr+1)d
(r + 1)Z Sr(a)xrdψ −
(r + 1)Z T (Sr(a))xr+1dψ
(r + 1)Z d
2 xrd
2 dψ
2 xrd
2 dψ − λ
≥λ
1
1
≥
1
1
1
n−1Xr=0
1
2 dψ.
1
1
2 dψ.
Hence from the above inequality, it follows
We compute the left hand side of (2.2), and it equals
nXr=0
=Z S0(a)x0dψ +
=Z S0(a)x0dψ +
nXr=1
n−1Xr=0
(r + 1)Z Sr(a)xrdψ −
n−1Xr=0
(r + 2)Z Sr+1(a)xr+1dψ −
(r + 1)Z T (Sr(a))xr+1dψ
(r + 1)Z T (Sr(a))xr+1dψ.
n−1Xr=0
13
Since (r + 2)Sr+1(a) = a + T (a) + · · · + T r(a) + T r+1(a) and (r + 1)T Sr(a) = T (a) +
T 2(a) + · · · + T r(a) + T r+1(a), the left hand side of (2.2) equals
Therefore, inequality (2.2) becomes
nXr=0Z axrdψ ≥ λ
So we obtain the following inequality,
≥λ
from the following inequalities,
1
1
1
1
2eT (xr+1)d
1
1
1
1
2 xrd
2 dψ
2 dψ ≥ 0,
2 xrd
2 dψ − λ
2 xrd
2 dψ − λ
rZ d
1
2eT (xr)d
1
nXr=0Z d
nXr=1
n−1Xr=0Z axr+1dψ
nXr=1Z axrdψ =
nXr=0Z axrdψ.
(r + 1)Z d
n−1Xr=0
Z S0(a)x0dψ +
=Z ax0dψ +
(r + 1)Z d
nXr=0
nXr=0Z axrdψ − λ
rZ d
nXr=0
Z d
2eT (xr)d
= sup{Z d
2eT (xr)d
= sup{Z xrT (d
= sup{Z d
≤ sup{Z d
=Z d
nXr=0Z axrdψ ≥ λ
0 sdps be the spectral decomposition of 1 −Pn
nXr=0Z d
2 T (x)dψ x ∈ M+, x ≤ 1}
2 xrd
2 xdψ x ∈ M+, x ≤ 1}
2 xdψ x ∈ M+, x ≤ 1}
2 )dψ x ∈ M+, x ≤ 1}
2 xrd
2 dψ.
2 xrd
2 dψ,
1
1
2 xrd
1
1
2 xd
1
2 dψ
1
1
1
1
1
1
1
1
0 ≤ ym ≤ 1 − pm−1 ≤ m(1 −
xr).
nXr=0
14
1
2 dψ.
(2.3)
r=0 xr, suppose
Hence we have
for any a ∈ L1
+(M; ψ).
Letting 1 −Pn
r=0 xr = R 1
y ∈ M+, y ≤ 1. Writing ym = (1 − pm−1)y(1 − pm−1) for each m ∈ N, we have
Hence by (2.1),
for 0 ≤ r ≤ n.
Z Sr(a)ymdψ ≤ λZ d
1
2 ymd
1
2 dψ
hence enSr(a)en ≤ λenden for 0 ≤ r ≤ n because M = (L1(M; ψ))∗.
r=0 xr ∈ M+, we have
Taking the limit as m → ∞, and putting en = 1 − p0, we get
1
1
2 dψ
2 enyend
Z enSr(a)enydψ =Z Sr(a)enyendψ
≤ λZ d
= λZ endenydψ,
Therefore for r = 0, we get enaen ≤ λenden, and sincePn
xr)dψ ≤ λZ den(
nXr=0
xr)dψ ≥ λZ d(1 − en)(
xr)dψ,
nXr=0
nXr=0
0 sdps = 0, we obtain that 1 − en = (1 − en)(Pn
nXr=0
xr)dψ.
Z aen(
Z a(1 − en)(
r=0 xr = p0R 1
Z d
1
λZ a(1 − en)dψ ≤
1
λZ adψ.
1
2 (1 − en)d
1
2 dψ ≤
Since p01 − p0Pn
we get
which together with (2.3) gives
r=0 xr). Hence
From the above procedure, we get en for each n ∈ N such that ϕ(1 − en) ≤ 1
enSr(a)en ≤ λenden for any r in {0, 1, . . . , n}. Choose a subnet enk which converges weakly
to some h ∈ M with 0 ≤ h ≤ 1.
λR adψ and
We assume that Sr(a) ∈ L1
∗ for each r ∈ N. Since
M acts standardly on H and each φr is normal and hence a vector state, for any fixed r ∈ N,
there exists a vector ξr in H such that φr = ωξr,ξr . As a result,
+(M; ψ) corresponds to some φr ∈ M +
Z hSr(a)hdψ = φr(h2) = hh2ξr, ξri = hξr2 .
The subnet enk converges to h in the weak operator topology, and for each r ∈ N, we have
enkξr → hξr in the weak topology of H when k → ∞. Since the Hilbert space norm · is
lower semicontinuous relative to this topology, we have
hξr2 ≤ lim inf
k→∞
enkξr2 .
(2.4)
15
Hence, we have
Combined with
Z hSr(a)hdψ ≤ lim inf
k→∞
enk ξr2 .
dϕ
dψ
for k ∈ N such that nk is larger than r, we conclude that
enkξr2 = φr(enk) =Z enk Sr(a)enkdψ
≤λZ enk denkdψ = λZ enk
Z hSr(a)hdψ ≤ λ lim
Taking the spectral decomposition for h =R 1
M, let g =R 1
s−1des. Then for each r ∈ N,
k→∞
1
2
dψ = λϕ(enk),
ϕ(enk) = λϕ(h),
since ϕ is a normal state on M, i.e., ϕ is weakly continuous on the unit ball of M.
0 sdes, and let e = 1 − e 1
2
∈ M for e 1
2
is in
Z eSr(a)edψ =Z ghSr(a)hgdψ ≤ 4Z hSr(a)hdψ,
thanks to e = gh and g ≤ 2.
Therefore we obtain that
Z eSr(a)edψ ≤ 4Z hSr(a)hdψ ≤ 4λϕ(h) ≤ 4λϕ(1) = 4λ.
Moreover, for the reason 1 − e = e 1
2
≤ 2(1 − h), we get
ϕ(1 − e) ≤ 2ϕ(1 − h) = 2 lim
k→∞
ϕ(1 − enk ) = 2 lim
k→∞Z d
1
2 (1 − enk)d
1
2 dψ ≤
2
λZ adψ,
also because ϕ is a normal state on M.
We should point out that we learned much from Professor Kosaki and Professor Xu in
the proof of this theorem.
We turn to the case that M is semifinite, L1(M; ψ) is equivalent to the tracial L1(M).
Moreover, since the construction of the Hilsum Lp-spaces is independent on the choice of
the normal semifinite faithful weight ψ in the isometrically isomorphic sense, hence we may
choose a special case that ψ(·) = ϕ(J · J), then the spatial derivative dϕ
dψ becomes △ϕ, i.e.,
the modular operator associated with ϕ and in the semifinite case, ϕ can be assumed to be a
16
normal faithful semifinite trace. Considering the modular automorphism group induced by
a trace is trivial, we get △ϕ = 1 in this case, then the result in the above theorem is for any
a ∈ L1
+(M; ψ) and any λ > 0, and any n ∈ N, there exists a projection en ∈ M such that
enSr(a)en ≤ λen for all r ∈ {0, 1, ..., n} and ϕ(1 − en) ≤ 2
in [30], Sr(a)
hSr(a)
λR adψ, then by Yeadon's method
2 h weakly as k → ∞. In fact, for ξ1 ∈ H, ξ2 ∈ D(Sr(a)
2 enk ξ1, ξ2i = henkξ1, Mr(a)
2 ξ2i → hhξ, Sr(a)
2 enk → Sr(a)
2 ), we have
2 ξ2i, and
1
1
1
1
1
1
1
(cid:12)(cid:12)(cid:12)hhξ1, Sr(a)
2 enk(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) = enkSr(a)enk
hSr(a)
1
1
Since(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)Sr(a)
Hence it follows
2 ξ2i(cid:12)(cid:12)(cid:12) = lim
k→∞(cid:12)(cid:12)(cid:12)hSr(a)
2 ≤ λ
1
1
1
2 enkξ1, ξ2i(cid:12)(cid:12)(cid:12) ≤ λ
2 if nk ≥ r, so that hξ1 ∈ D(Sr(a)
1
2 ) and
1
2 ξ1 ξ2 .
2 hξ1, ξ2i = hhξ1, Sr(a)
1
2 ξ2i = lim
k→∞
hSr(a)
1
2 enkξ1, ξ2i.
2
1
2 hξ1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
2
≤ lim
k→∞(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)Sr(a)
1
2 enk ξ1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
λhenkξ1, ξ1i = λhhξ1, ξ1i,
henkSr(a)enkξ1, ξ1i ≤ lim
k→∞
= lim
k→∞
hhSr(a)hξ1, ξ1i =(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)Sr(a)
, g =R 1
2
i.e., hSr(a)h ≤ λh for each r ∈ N. Taking the spectral decomposition for h =R 1
0 sdes, and
s−1des. Then we have eSr(a)e ≤ λghg ≤ 2λe, this is just the result of
let e = 1 − e 1
1
2
Yeadon's Theorem 1 in [30].
Then we introduce the conceptions of "type" and "weak type" for the action of T on
M + M∗. Such conceptions "type" and "weak type" appeared in the classical real analysis
first and have been widely used in classical ergodic theorems. They are modified by Junge-
Xu (pp 396 -- 397, [15]), for the framework of noncommutative Lp-spaces associated with a
semifinite von Neumann algebra. Here we rewrite them for the transform T we constructed
above in the framework of Lp(M; ψ)'s (1 ≤ p ≤ ∞).
For each n ∈ N, Sn is a linear map on M + M∗ satisfying the conditions (1) − (3).
Thus S = (Sn)n∈N is a map which sends a positive element in Lp(M; ψ) for some fixed p,
1 ≤ p ≤ ∞ to a sequence of positive elements in Lp(M; ψ). Here we identify L1(M; ψ) with
M∗ and L∞(M; ψ) with M.
We say that S is of type (p, p), (1 ≤ p ≤ ∞) if there is a positive constant c such that
+(M; ψ) satisfying ap ≤ c xp and Sn(x) ≤ a, for any
+(M; ψ), there is a ∈ Lp
for any x ∈ Lp
n ∈ N.
17
Then we say that S is of weak type (p, p), (1 ≤ p < ∞), if there is a positive constant c
such that for any x ∈ Lp
+(M; ψ), and any λ > 0 there is a projection e ∈ M such that
λ (cid:19)p
ϕ(1 − e) ≤(cid:18) c xp
and eSn(x)e ≤ λ1, for any n ∈ N,
where 1 is the identity of the von Neumann algebra M.
Yeadon's Theorem shows that S = (Sr)r∈N is of weak type (1, 1) when M is semifinite.
Indeed, by condition (1), we get that for each r ∈ N, if x ∈ M+ = L∞
It is obvious that S = (Sr)r∈N is of type (∞, ∞) for an arbitrary von Neumann algebra
M.
+ (M; ψ), we
have T r(x) ≥ 0 for n ∈ N and T r(x) ≤ x and thus Sr(x) ≤ x. Hence if we put
a = x 1 ∈ L∞
+ (M; ψ), we have a = x and Sr(x) ≤ a, for all r ∈ N.
As we have mentioned in the previous section, such a weak type conception is no longer
appropriate for the non-tracial cases. We give a pre-version of pre-weak type here, we hope
it is of some meaning.
We say that S is of pre-weak type (p, p), (1 ≤ p < ∞) if there is a positive constant c
such that for any x ∈ Lp
+(M; ψ), and any λ > 0 there is a projection e ∈ M such that
ϕ(1 − e) ≤(cid:18) c xp
λ (cid:19)p
and eSn(x)ep ≤ λ, for any n ∈ N,
where 1 is the identity of the von Neumann algebra M. Theorem 2.1 shows that S = (Sr)r∈N
is of pre-weak type (1, 1) for a σ-finite von Neumann algebra.
If we would like to obtain a satisfactory non-tracial extension of Theorem 1.1, one possible
method is to consider real interpolation theory (possible non-tracial real interpolation theory
for pre-weak type, though we are not very sure about the existence of such theory and it
is still in exploring) of von Neumann algebras, because the real interpolation theory always
provides us the type for midpoints from the weak type assumption of endpoints. But as
is pointed out by Junge-Xu (pp 385 -- 386, [15]), in contrast with the classical theory, the
noncommutative nature of weak type (1, 1) inequalities seems a priori unsuitable for classical
interpolation arguments. More accurately speaking, Pisier-Xu gave a counterexample saying
that the complex interpolation space Lp(M) may not coincide with the real interpolation
space Lp,p(M) for non-tracial von Neumann algebras if we establish non-commutative real
interpolation theory verbatim from classical one (p.1472, Example 3.3, [23]). We had tried
several ways to modify the definition of real interpolation construction in order to suit well
to the von Neumann algebra theory and the complex interpolation of Terp ([29]) at the same
time. Unfortunately, we have not obtained any valid method for this matter so far. Let us
18
point out that a key obstacle in this work is the absense of generalized singular numbers
for the non-tracial cases. The generalized singular number function ([8]) is a powerful tool
when dealing with the τ -measurable operators associated with a semifinite von Neumann
algebra, and the noncommutative tracial Lp-spaces are just consisting of such operators.
In the non-tracial cases, whether there exists such a counterpart theory which contains the
generalized singular number theory by [8] has not been sufficiently understood now. We do
not know whether there is any other method available and we will continue to explore this
problem in the future.
Finally, we give three examples as applications of Theorem 2.1.
Example 2.2. Let (Ω, F, µ) be a finite measure space and N be a σ-finite von Neumann
algebra equipped with a normal faithful state ϕ1. We consider the von Neumann algebra
tensor product (M, ϕ2) = (L∞(Ω), µ)⊗(N, ϕ1), where ϕ2 is a normal faithful state since µ is
finite and ϕ1 is a normal faithful state. For 1 ≤ p < ∞, the corresponding noncommutative
Lp(M; ψ2) is just Lp(Ω, Lp(N; ψ1)), the usual Lp-space of strongly measurable p-integrable
functions on Ω with values in Lp(N; ψ1), where ψ1 (resp. ψ2) is a normal faithful state
on the commutant of N (resp. M), and we may choose ψ1 (resp. ψ2) to be associated
with ϕ1 (resp. ϕ2) by the Tomita-Takesaki theory. Now let S be a linear map on L∞(Ω)
satisfying conditions (1) − (3) (with M = L∞(Ω) there), then T = I ⊗ S is a linear map on
M verifying the same conditions (with M = L∞(Ω)⊗N there). From Theorem 2.1, for any
a ∈ L1(Ω, L1
+(N; ψ1)) and any λ > 0, and any n ∈ N, there exists a projection en ∈ M such
that
and
enSr(a)en ≤ λenden for all r ∈ {0, 1, ..., n},
ϕ2(1 − en) ≤
2
λZ adψ2,
where Sr(a) = 1
M. Furthermore, for any a ∈ L1(Ω, L1
for any r ∈ N,
k=0 I ⊗ Sk(a), d = dϕ2
r+1Pr
dψ2
is the spatial derivative and 1 is the identity of
+(N; ψ1)), there exists a projection e ∈ M such that
Z eSr(a)edψ2 ≤ 4λ, and ϕ2(1 − e) ≤
2
λZ adψ2.
Example 2.3. Let M be a von Neumann algebra with a normal faithful state ϕ, and let N
be any von Neumann subalgebra of M. The generalized conditional expectation ε : M → N
relative to ϕ defined by Accardi-Cecchini is given as ε(x) = JN PN Jπϕ(x)JPN JN for any
19
x ∈ M (see [1]).
(1) − (3). Therefore, Theorem 2.1 implies that for any a ∈ L1
any n ∈ N, there exists a projection en ∈ M such that
If we regard ε to be a linear map from M to M, ε satisfies conditions
+(M; ψ) and any λ > 0, and
enSr(a)en ≤ λenden for all r ∈ {0, 1, ..., n},
and
ϕ(1 − en) ≤
2
λZ adψ,
where Sr(a) = 1
Furthermore, for any a ∈ L1
k=0 T k(a), d = dϕ
dψ is the spatial derivative and 1 is the identity of M.
+(M; ψ), there exists a projection e ∈ M such that for any r ∈ N,
r+1Pr
Z eSr(a)edψ ≤ 4λ, and ϕ(1 − e) ≤
2
λZ adψ.
If N is globally invariant under the modular automorphism group σϕ
t , then ε will be the
conditional expectation in the sense of [27], i.e., a projection of norm one. As a projection
is idempotent, we have enε(a)en ≤ λenden andR eε(a)edψ ≤ 4λ in this case.
Example 2.4. Let {(Mi, ϕi)}i∈I be a family of von Neumann algebras and each ϕi is a normal
faithful state. Let (M, ϕ) = ∗i∈I(Mi, ϕi) be the von Neumann algebra reduced free product
(see [20]), and hence ϕ is a normal faithful state on M. Put M ◦
i = {x ∈ Miϕi(x) = 0}, then
Mi = C1Mi ⊕ M ◦
i , and let Ti : Mi → Mi be defined by
TiC1Mi
= IC1Mi
and TiM ◦
i = exp(−1)IM ◦
i
,
and {Ti}i∈I defines a positive linear map T on M by free product, and T is uniquely deter-
mined by its action on monomials:
T (x1x2 · · · xn) = Ti1(x1)Ti2(x2) · · · Tin(xn) = exp(−n)x1x2 · · · xn,
for any x1, x2, · · ·, xn with xk ∈ M ◦
ik and i1 6= i2 6= · · · 6= in. The map T is called the free
product of the family {Ti}i∈I. Then T satisfies conditions (1) − (3) with respect to M since
each Ti satisfies conditions (1) − (3) with respect to Mi. Hence T extends to a positive linear
map on L1(M; ψ) and by Theorem 2.1, for any a ∈ L1
+(M; ψ) and any λ > 0, and any n ∈ N,
there exists a projection en ∈ M such that
enSr(a)en ≤ λenden for all r ∈ {0, 1, ..., n},
and
ϕ(1 − en) ≤
2
λZ adψ,
20
where Sr(a) = 1
Furthermore, for any a ∈ L1
k=0 T k(a), d = dϕ
r+1Pr
Z eSr(a)edψ ≤ 4λ, for any r ∈ N, and ϕ(1 − e) ≤
dψ is the spatial derivative and 1 is the identity of M.
+(M; ψ), there exists a projection e ∈ M such that
2
λZ adψ.
21
Acknowledgements
The author would like to take this opportunity to express his highest respect and appreci-
ation to his supervisor Professor Y.Kawahigashi, who led him into the realm of von Neumann
algebra theory and paid great care and patience for him in the past five years. And the au-
thor is also very grateful to Professor H.Kosaki in Kyushu University and Professor Q.Xu in
Universit´e de Franche-Comt´e from whom he learned many techniques on noncommutative
Lp-spaces through e-mail correspondence, and they kindly pointed out mistakes in the previ-
ous version of this paper. And he also thanks for Professor N.Ozawa and Professor Y.Ogata
and other ones in their group, since he benefited greatly from seminars and discussions in the
group. Finally, his thanks go to Mr. Zhao in his research room, for much help from Mr. Zhao.
References
[1] L. Accardi, C. Cecchini, Conditional expectations in von Neumann algebras and a the-
orem of Takesaki, J. Funct. Anal. 45 (1982) 245-273. MR0647075(84j:46088)
[2] H. Araki, T. Masuda, Positive cones and Lp-spaces for von Neumann algebras, Publ.
Res. Inst. Math. Sci. 18 (1982) 759-831. MR0677270(84h:46082)
[3] J. Bergh,
J. Lofstrom,
Interpolation Spaces, Springer-Verlag, Berlin,
1976.
MR0482275(58:2349)
[4] O. Bratteli, D. Robinson, Operator Algebras and Quantum Statistical Mechanics. I, C ∗-
and W ∗-algebras, symmetry groups, decomposition of states, in: Texts and Monographs
in Physics, Springer-Verlag, New York, 1987. MR0887100(88d:46105)
[5] H. A. Carteret, D. R. Terno, K.
Zyczkowski, Dynamics beyond completely positive
maps: some properties and applications, Phys. Rev. A 77 (2008) 042113.1-042113.8.
MR2491057(2010d:82072)
[6] A. Connes, On the spatial theory of von Neumann algebras, J. Funct. Anal. 35(1980)
153-164. MR0561983(81g:46083)
[7] N. Dunford, J. T. Schwartz, Linear Operators, I, General Theory, in: Applied Mathe-
matics, Vol. 7, Interscience Publishers, Inc., New York, 1958. MR0117523(22:8302)
22
[8] T. Fack, H. Kosaki, Generalized s-numbers of τ -measurable operators, Pacific J. Math.
123(1986) 269-300. MR0840845(87h:46122)
[9] U. Haagerup. Lp-spaces associated with an arbitrary von Neumann algebra,
In:
Alg`ebres d'op´erateurs et leurs applications en physique math´ematique, Proc. Col-
loq., Marseille, 1977, Volume 274 of Colloq. Internat. CNRS, 175-184. Paris, 1979.
MR0560633(81e:46050)
[10] U. Haagerup, M. Junge, Q. Xu, A reduction method for noncommutative Lp-spaces and
applications, Trans. Amer. Math. Soc. 362 (2010) 2125-2165. MR2574890
[11] M. Hilsum, Les espaces Lp d'une alg`ebre de von Neumann d´efinies par la deriv´ee spatiale,
J. Funct. Anal. 40 (1981) 151-169. MR0609439(83c:46053)
[12] P. Horodecki, R. Augusiak, M. Demianowicz, General construction of noiseless networks
detecting entanglement with the help of linear maps, Phys. Rev. A 74 (2006) 052323.1-
052323.6. MR2288089(2007k:81023)
[13] H.
Izumi, Constructions of non-commutative Lp-spaces with a complex pa-
Internat. J. Math. 8 (1997) 1029-1066.
rameter arising from modular actions,
MR1484866(99a:46114)
[14] M. Junge, Doob's inequality for non-commutative martingales, J. Reine Angew. Math.
549 (2002) 149-190. MR1916654(2003k:46097)
[15] M. Junge, Q. Xu, Noncommutative maximal ergodic theorems, J. Amer. Math. Soc. 20
(2007) 385-439. MR22767755(2007k:46109)
[16] H. Kosaki, Applications of the complex interpolation method to a von Neumann algebra:
noncommutative Lp-spaces, J. Funct. Anal. 56 (1984) 29-78. MR0735704(86a:46085)
[17] E. C. Lance, Ergodic theorems for convex sets and operator algebras, Invent. Math. 37
(1976) 201-214. MR0428060(55:1089)
[18] T. Mei, Operator-valued Hardy spaces, Mem. Amer. Math. Soc. 188, no.881, American
Mathematical Society, 2007. MR2327840(2010d:46085)
[19] E. Nelson, Notes on non-commutative integration, J. Funct. Anal. 15 (1974) 103-116.
MR0355628(50#8102)
23
[20] A. Nica, R. Speicher, Lectures on the combinatorics of free probability, in: London
Mathematical Society Lecture Notes Series, 335, Cambridge University Press, Cam-
bridge, 2006. MR2266879(2008k:46198)
[21] G. Pisier, Non-commutative vector valued Lp-spaces and completely p-summing maps,
Ast´erisque 247 (1998). MR1648908(2000a:46108)
[22] G. Pisier, The operator Hilbert space OH and type III von Neumann algebras, Bull.
London. Math. Soc. 36 (2004) 455-459. MR2069007(2005c:46082)
[23] G. Pisier, Q. Xu, Non-commutative Lp-spaces. In: Handbook of the Geometry of Banach
Spaces, Vol. 2, 1459-1517, North-Holland, Amsterdam, 2003. MR1999201(2004i:46095)
[24] A. Royer, Families of positivity preserving but not completely positive superoperators,
Phys. Lett. A 336 (2005) 295-310. MR2119628(2005j:81022)
[25] A. Shaji, E. C. G. Sudarshan, Who's afraid of not completely positive maps? Phys.
Lett. A 341 (2005) 48-54. MR2144129(2006g:82029)
[26] S. J. Szarek, E. Werner, K. Zyczkowski, Geometry of sets of quantum maps: a generic
positive map acting on a high-dimensional system is not completely positive, J. Math.
Phys. 49 (2008) 032113.1-032113.21. MR2406781(2009b:81025)
[27] M. Takesaki, Conditional expectations in von Neumann algebras, J. Funct. Anal. 9
(1972) 306-321. MR03033078(46#2445)
[28] M. Takesaki, Theory of Operator Algebras, II, In: Encyclopaedia of Mathematical
Sciences, 125, Operator Algebras and Non-commutative Geometry, 6, Springer-Verlag,
Berlin, 2003. MR1943006(2004g:46079)
[29] M. Terp, Interpolation spaces between a von Neumann algebra and its predual, J.
Operator. Theory 8 (1982) 327-360. MR0677418(85b:46075)
[30] F. J. Yeadon, Ergodic theorems for semifinite von Neumann algebras I, J. London.
Math. Soc. 16 (1977) 326-332. MR0487482(58:7111)
24
|
1310.7880 | 1 | 1310 | 2013-10-29T17:06:59 | Radial multipliers on arbitrary amalgamated free products of finite von Neumann algebras | [
"math.OA",
"math.GR"
] | Let $(M_i)_{i}$ be a (finite or infinite) family of finite von Neumann algebras with a common subalgebra $P$. When $\varphi:\IN\rightarrow\IC$ is a function, we define the radial multiplier $M_\varphi$ on the amalgamated free product $M=M_1\free_P M_2\free_P\ldots$ setting $M_{\varphi}(x)=\varphi(n)x$ for every reduced expression $x$ of length $n$. In this paper we give a sufficient condition on $\varphi$ to ensure that the corresponding radial multiplier $M_\varphi$ is a completely bounded map, and moreover we give an upper bound on its completely bounded norm. Our condition on $\varphi$ does not depend on the choice of von Neumann algebras $(M_i)_i$ and $P$. This result extends earlier results by Haagerup and M\"oller, who proved the same statement for free products without amalgamation, and M\"oller showed that the same statement holds when $P$ has finite index in each of the $M_i$. | math.OA | math |
RADIAL MULTIPLIERS ON ARIBITRARY AMALGAMATED FREE
PRODUCTS OF FINITE VON NEUMANN ALGEBRAS
STEVEN DEPREZ
Abstract. Let (Mi)i be a (finite or infinite) family of finite von Neumann algebras with a
common subalgebra P . When ϕ : N → C is a function, we define the radial multiplier Mϕ
on the amalgamated free product M = M1 ∗P M2 ∗P . . . setting Mϕ(x) = ϕ(n)x for every
reduced expression x of length n. In this paper we give a sufficient condition on ϕ to ensure
that the corresponding radial multiplier Mϕ is a completely bounded map, and moreover
we give an upper bound on its completely bounded norm. Our condition on ϕ does not
depend on the choice of von Neumann algebras (Mi)i and P . This result extends earlier
results by Haagerup and Moller, who proved the same statement for free products without
amalgamation, and Moller showed that the same statement holds when P has finite index in
each of the Mi.
Introduction
Let Γ be a countable group. A function f : Γ → C is called a Herz-Schur multiplier if the
map mf : C Γ → C Γ that is defined by mf (ug) = f (g)ug extends to a σ-weakly continuous
completely bounded map mf : L Γ → L Γ. The Herz-Schur norm kf kHS of f is the completely-
bounded norm of mf .
The free groups Fn (2 ≤ n ≤ ∞) come equiped with a natural word-length function ·,
associated to the canonical generators. Given a map ϕ : N → C, we can look at the associated
radial function fϕ : Fn → C, that is given by fϕ(g) = ϕ(g). Haagerup and Szwarc [HSS]
showed that fϕ is a Herz-Schur multiplier if and only if the Hankel matrix that is given by
Hϕ = (ϕ(n + m) − ϕ(n + m + 2))n,m is trace class. In that case, there exist constants c± ∈ C
such that ϕ(n) = ϕ0(n) + c+ + (−1)nc− where ϕ0(n) converges to 0. The Herz-Schur norm
of fϕ is bounded by
In fact, this inequality becomes an equality if n = ∞, and Haagerup and Szwarc give an exact
expression for the Herz-Schur norm, also when n is finite.
kfϕkHS ≤ kHϕk1 + c+ + c− .
We say that a function ϕ : N → C is in class C if the Hankel matrix Hϕ is trace class, and we
will write kϕkC = kHϕk + c+ + c−.
corresponding author, University of Copenhagen, [email protected].
Supported by ERC Advanced Grant no. OAFPG 247321.
Supported by the Danish National Research Foundation through the Centre for Symmetry and Deformation.
(DNRF92).
Department of mathematics, Copenhagen university, Universitetsparken 5, 2500, Copenhagen O.
1
2
STEVEN DEPREZ
Wysoczanski [W] studied radial multipliers on free products Γ = ∗i Γi of groups. A free
product comes equiped with a natural word-length function g 7→ g, so it makes sense to talk
about radial multipliers on free products. Wysoczanski showed that a radial function fϕ is a
Herz-Schur multiplier if the Hankel matrix Kϕ = (ϕ(n + m) − ϕ(n + m + 1))n,m is trace class.
If this is the case, we say that ϕ is in class C′. If ϕ ∈ C′, then the limit c = limn→∞ ϕ(n)
exists and the Hankel matrix eKϕ = (ϕ(n + m + 1) − ϕ(n + m + 2))n,m is also trace class.
Wysoczanski showed that the Herz-Schur norm of fϕ is bounded by
kfϕkHS ≤ kϕkC ′ = kKϕk1 +(cid:13)(cid:13)(cid:13)eKϕ(cid:13)(cid:13)(cid:13)1
+ c .
When all the groups Γi have the same (possibly infinite) cardinality, he can explicitly compute
the Herz-Schur norm of fϕ.
Wysoczanski's result has been extended to an operator algebraic framework by Haagerup
and Moller [HM]. Namely, when M = ∗i Mi is a free product of von Neumann algebras, then
M is σ-weakly densely spanned by the reduced words, i.e. words of the form x1 . . . xn where
xk ∈ Mik \ C 1 and i1 6= i2 6= . . . 6= in. Given a function ϕ : N → C, we consider the radial
multiplier mϕ that is given by mϕ(x1 . . . xn) = ϕ(n)x1 . . . xn for every reduced word x1 . . . xn.
Haagerup and Moller showed that mϕ is a completely bounded map if the Hankel matrix Kϕ
defined above is trace class. They also showed that in that case kmϕkcb ≤ kϕkC ′. Recently,
Moller extended this result to include amalgamated free products where the amalgam has
finite index in each of the factors Mi [M].
The present paper extends this result to arbitrary amalgamated free products of finite von
Neumann algebras. We give an overview of the techniques used. It is instructive to first look
at the ideas behind Wysoczanski's result.
The Haagerup-Szwarc result can be restated as a result about radial Schur multipliers on
trees, as we explain now. In this form, the result was published by Haagerup, Steenstrup and
Szwarc [HSS], but is based on earlier unpublished notes by Haagerup and Szwarc. Let X be
a countable set, and consider a map k : X × X → C. The Schur product k ⋆ T of k with a
bounded operator T ∈ B(ℓ2(X)) is given by
hδx, (k ⋆ T )δyi = k(x, y) hδx, T δyi .
We say that k is a Schur multiplier is mk : T 7→ k ⋆ T is a well-defined completely bounded
map from B(ℓ2(X)) to itself, and its Schur norm is kkkS = kmkkcb. Suppose now that
d : X × X → R+ is a metric on X. We say that a Schur multiplier k : X × X → C is a radial
Schur multiplier if it is of the form k(x, y) = ϕ(d(x, y)) for some function ϕ : R+ → C.
The concepts of Schur multipliers and Herz-Schur multipliers are closely related: a function
f : Γ → C on a group Γ is a Herz-Schur multiplier if and only if the function defined by
k(g, h) = f (g−1h) is a Schur multiplier, and moreover kf kHS = kkkS.
Let T be a regular tree, which we identify with its vertex set. We consider T as a metric space
with the usual metric d : T × T → N, given by the length of the shortest path between two
vertices. Haagerup, Steenstrup and Szwarc showed that a radial function kϕ : T × T → C,
RADIAL MULTIPLIERS ON VON NEUMANN ALGEBRAS
3
defined by kϕ(x, y) = ϕ(d(x, y)) is a radial Schur multiplier if and only if ϕ ∈ C, and moreover
its Schur norm is bounded by kkϕkS ≤ kϕkC.
Using this restatement, the proof of Wysoczanski's result goes as follows. Observe that
Γ = ∗i Γi is the fundamental group of the following graph of groups:
{e}
Γ2
Γ1
Γn
Consider the action of Γ on the Bass-Serre tree T associated to this graph of groups. Observe
that the vertices of the tree T are indexed by Γ ⊔ Γ/Γ1 ⊔ . . . ⊔ Γ/Γn. Consider the vertex
v0 ∈ T that corresponds to the identity in Γ. For every group element g ∈ Γ, we see that
the distance between v0 and g · v0 is exactly twice the word-length of g. Let ϕ : N → C be a
function. Consider the function ψ : N → C that is defined by ψ(n) = ϕ(n/2) if n is even, and
ψ(n) = 0 if n is odd. The function ψ is in class C if and only if ϕ is in class C′, and moreover
kψkC = kϕkC ′. Consider the radial functions k = kψ : T × T → C and f = fϕ : Γ → C as
before. We define an isometry u : ℓ2(Γ) → ℓ2(T ) ⊗ ℓ2(Γ) by the formula u(δg) = δg−1v0 ⊗ δ−1
g .
Let L(Γ) act on ℓ2(Γ) by left multiplication, i.e. ugδh = δgh. We observe that the multipliers
mf and mk are related by mf (x) = u∗(mk ⊗ id)(uxu∗)u for all x ∈ L(Γ). In particular, we
see that
kmf kcb ≤ kmkkcb ≤ kψkC = kϕkC ′ .
At no point in the argument above did we use the fact that Γ was a free product as opposed
to an amalgamated free product. We prove our result about amalgamated free products of
von Neumann algebras using a similar strategy. For this, we have to find a good replacement
for the "fundamental group of a graph of groups". We think that the right notion is that of
the "relative Gaussian construction", which we describe in short below, and in more detail in
section 2. They come naturally with a notion of a "reduced word", so it makes sense to talk
about radial multipliers. We show that, for an important subclass of the "relative Gaussian
constructions", a radial multiplier mϕ is completely bounded, whenever ϕ is in class C, and
kmϕkcb ≤ kϕkC. This is only true for a subclass of the "relative Gaussian constructions", as
it fails for the classical Gaussian construction.
The relative Gaussian construction is a strong generalization of Voiculescu's free Gaussian
construction [V]. Many generalizations of the free Gaussian construction have been intro-
duced before. For example, there are the q-Gaussian constructions [BS1]. The construction
that is closest to our "relative Gaussian construction" is given by Shlyakhtenko's "A-valued
semicircular systems" [S]. One can describe the "relative Gaussian construction" as being a
tracial A-valued semicircular system, but deformed in a way similar to the q-Gaussian con-
structions. As with the free Gaussian construction and its generalizations, we first define a
4
STEVEN DEPREZ
kind of Fock space and creation and annihilation operators on them. Then we define the "rel-
ative Gaussian construction" as the algebra generated by certain creation and annihilation
operators.
For the construction of the relative Fock space, we need three pieces of data:
• A von Neumann algebra M with a faithful normal state τ .
• A bimodule H over M .
• A self-adjoint M -M bimodular contraction F : H ⊗M H → H ⊗M H that satisfies the
braid relation
(F ⊗ 1H )(1H ⊗ F )(F ⊗ 1H ) = (1H ⊗ F )(F ⊗ 1H)(1H ⊗ F ),
as operators on H ⊗M H ⊗M H.
Consider the n-fold Connes tensor product bimodule H (n) of H with itself. By convention,
we set H (0) = M . We define a number of M -M bimodular operators on H (n). Let σ ∈ Sn
be a permutation on n elements. It is well-known that σ can be decomposed as σ = t1 . . . tn
where each ti = (ki, ki + 1) is a transposition of two consecutive elements. Moreover, the
shortest such decomposition is unique, up to applying the braid relation
(k, k + 1)(k, k − 1)(k, k + 1) = (k, k − 1)(k, k + 1)(k, k − 1).
We assume that σ = t1 . . . tn is such a shortest decomposition. Put
F(k,k+1) = −1H (k−1) ⊗ F ⊗ 1H (n−k−1) ,
and observe that Fσ = Ft1 . . . Ftn is well-defined. Define an operator Dn by
Dn = Xσ∈Sn
Fσ.
Bozejko and Speicher showed in [BS2] that such an operator Dn is always a positive operator.
We define a new inner product h·, ·iF = h·, Dn·i on H (n). Denote by H (n)
the M -M -bimodule
that we obtain from H (n) by completion and separation with respect to this new inner product
h·, ·iF . The relative Fock space is now
F
F M (H, F ) =Mn≥0
H (n)
F
In F M (H, F ), the vector 1 ∈ L2(M ) ⊂ F M (H, F ) plays a special role. It will be called the
vacuum vector and is denoted by Ω to avoid confusion. For every ξ ∈ H, we define a left
creation operator L(ξ) on F M (H, F ), by the formula
L(ξ)η = ξ ⊗ η ∈ H (n+1) ⊂ H (n+1)
F
for all η ∈ H (n).
This creation operator is in general a closable unbounded operator. We still denote its closure
by L(ξ). The adjoint L(ξ)∗ is called an annihilation operator. When F = 0, then we get
Dn = 1 for all n ∈ N, and the creation operators satisfy the relation
L(ξ)∗L(η) = hξ, ηiM .
In this case, we obtain the Fock space from Shlyakhtenko's A-valued semicircular elements.
If M = C and F (ξ ⊗ η) = qη ⊗ ξ for some −1 < q < 1, we obtain the Fock space of the
RADIAL MULTIPLIERS ON VON NEUMANN ALGEBRAS
5
(−q)-Gaussian construction, and the creation and annihilation operators satisfy the relation
L(ξ)∗L(η) = hξ, ηi + qL(ξ)L(η)∗.
In general, a similar relation holds, but it is slightly more complicated to write down. See
section 2, proposition 2.4. In any case, the algebra TM (H, F ) = BM (F M (H, F )) of bounded
right-M -linear operators on F M (H, F ) is the w∗-closed linear span of the left action of M
and of (the spectral projections of the self-adjoint part of) operators of the form
L(ξ1) . . . L(ξn)L(ηm)∗ . . . L(η1)∗ where ξ1, . . . , ξn, η1, . . . , ηm ∈ H.
For the relative Gaussian construction, we need one more piece of data: an anti-unitary
J : H → H that satisfies J(xξy) = y∗J(ξ)x∗, and that is compatible with F is some way.
See section 2 for more details. Then we define Γ′′
M (H, F, J) to be the von Neumann algebra
on F M (H, F ) that is generated by the left action of M and all the operators of the form
L(ξ) + L(Jξ) with ξ ∈ H.
Obviously the free Gaussian construction can be obtained from our construction by taking
M = C, F = 0 and where J is given by complex conjugation on ℓ2(S) for a countable set S.
More generally, the q-Gaussian construction can be obtained by setting F (ξ ⊗ η) = −qη ⊗ ξ.
The tracial cases of Shlyakhtenko's A-valued semicircular systems can be obtained by setting
F = 0 in our general construction.
For us, a more important class of examples is given by amalgamated free products and HNN
extensions. If M1 ∗P M2 is an amalgamated free product, then we set M = M1 ⊕ M2, and
H = L2(M1) ⊗P L2(M2) ⊕ L2(M2) ⊗P L2(M1)
The anti-unitary J interchanges the two components of the direct sum above. The operator
F is the projection onto the closed subspace
L2(M1) ⊗P L2(M1) ⊕ L2(M2) ⊗P L2(M2)
= L2(M1) ⊗P L2(P ) ⊗P L2(M1) ⊕ L2(M2) ⊗P L2(P ) ⊗P L2(M2)
⊂ L2(M1) ⊗P L2(M2) ⊗P L2(M1) ⊕ L2(M2) ⊗P L2(M1) ⊗P L2(M2)
= H (2).
Then we obtain that
Γ′′
M (H, F, J) = M2(C) ⊗ (M1 ∗P M2).
More generally, suppose we are given a graph Γ of von Neumann algebras. A graph consists of
a set V of vertices, and a set E of edges, together with source and target maps s, t : E → V ,
and with an involution · : E → E that reverses each edge, so s(e) = t(e), t(e) = s(e) and
e = e. A graph of von Neumann algebras is a graph (V, E) together with a family of von
Neumann algebras (Mv)v∈V and a family of von Neumann subalgebras Pe ⊂ Ms(e) (e ∈ E)
and ∗-isomorphisms αe : Pe → Pe such that αe = α−1
e . Moreover, we assume that there are
normal conditional expectations Ee : Ms(e) → Pe.
6
STEVEN DEPREZ
Then we set M =Lv∈V Mv and we define H to be the completion of
L2(Ms(e)) ⊗Pe L2(Mt(e)),
Me∈E
where the inclusion of Pe into Mt(e) is given by the isomorphism αe : Pe → Pe ⊂ Mt(e). The
Jordan subspace is spanned (as a real vector space) by elements of the form
x ⊗ y + y∗ ⊗ x∗ ∈ Ms(e) ⊗Pe Mt(e) + Mt(e) ⊗Pe Ms(e).
The operator F is the projection onto the closed subspace of H (2) that is spanned by the
spaces of the form
Mt(e) ⊗Pe Pe ⊗Pe Mt(e).
We also call Γ′′
M (H, F, J) the fundamental von Neumann algebra of the given graph of von
Neumann algebras. With this terminology, we do not get that the group von Neumann
algebra of a fundamental group of a gaph of groups is exactly the fundamental von Neumann
algebra of the corresponding graph of von Neumann algebras. Instead the fundamental von
Neumann algebra is the amplification of the group von Neumann algebra of the fundamental
group. In fact, the fundamental von Neumann algebra is the groupoid von Neumann algebra
of the fundamental groupoid of the graph of groups.
Observe that in both of the previous examples, the contraction F is in fact a projection and
moreover 1 ⊗ F commutes with F ⊗ 1 on H (3). We can not extend Haagerup-Moller's result
to arbitrary relative Gaussian constructions, but we can extend it to all the cases where the
contraction F is a projection such that 1 ⊗ F commutes with F ⊗ 1:
Theorem 0.1 (see theorem 3.1). Let (M, τ ) be a tracial von Neumann algebra and let H
be a Hilbert M -M bimodule. Assume that F is a projection onto an M -M subbimodule
of H ⊗M H such that F ⊗ 1 commutes with 1 ⊗ F on H ⊗M H ⊗M H. Let ϕ : N → C
be a function in class C defined above. Then there is a unique completely bounded map
Φϕ : BM (F M (H, F )) → BM (F M (H, F )) that satisfies
Φϕ(L(ξ1) . . . L(ξn)L(ηm)∗ . . . L(η1)) = ϕ(n + m)L(ξ1) . . . L(ξn)L(ηm)∗ . . . L(η1)
Φϕ(x) = ϕ(0)x
kΦϕkcb ≤ kϕkC .
for all x ∈ M acting on the left
The result about amalgamated free products follows by restricting this completely bounded
map to a corner of a subspace.
1. Preliminaries and Notation
In this paper, the set of natural numbers N includes 0. When we want to refer to the natural
numbers excluding 0, we write N1.
RADIAL MULTIPLIERS ON VON NEUMANN ALGEBRAS
7
1.1. Permutation groups. We consider the permutation group Sn to be the group of per-
mutations of the set {1, . . . , n}. This permutation group Sn (1 ≤ n < ∞) is generated by the
transpositions of consecutive elements t1 = (1, 2), . . . , tn−1 = (n − 1, n). A complete set of
relations for these generators is given by
(1)
(2)
(3)
t2
i = e
titj = tjti
titi+1ti = ti+1titi+1
for all 1 ≤ i ≤ n − 1
whenever j − i ≥ 2
for all 1 ≤ i ≤ n − 2
The permutation group Sn is a Coxeter group with Coxeter system (Sn, {ti}), and most of
the results mentioned below are true in the more general context of Coxeter groups.
We write σ for the length of an element σ ∈ Sn with respect to the generating set {t1, . . . , tn−2}.
Observe that every element σ ∈ Sn is represented by a word ti1 . . . tis of minimal length, and
that this word is unique up to relations (2) and (3). The length σ can also be computed as
the number of inversions of σ, i.e. the number of pairs 1 ≤ i < j ≤ n such that σ(i) > σ(j).
Let π be a partition of {1, ldots, n} into consecutive sets, i.e. π is of the form
(4)
π =({1, . . . , k1}, {k1 + 1, . . . , k1 + k2}, . . . ,(1 +
ki, . . . ,
s−1Xi=1
ki))
sXi=1
for some k1, . . . , ks ≥ 1 withPi ki = n. Then we can consider the subgroup Sπ of permuta-
tions that preserve the partition π, i.e.
Sπ = {σ ∈ Sn σ(X) = X ∀X ∈ π} ∼= Sk1 ⊕ . . . ⊕ Sks .
Such a subgroup is called a parabolic subgroup of Sn. Every left coset C of Sπ contains a
unique element σC of minimal length. So every element σ ∈ C can be written uniquely as a
product σ = σCσ0 where σ0 ∈ Sπ. This decomposition satisfies σ = σC + σ0. The set of
all element σC is denoted by Vπ. This set Vπ can also be described directly in terms of the
partition π:
Vπ = {σ ∈ Sn σX is increasing for every X ∈ π} .
When π is given explicitly in the form (4), we also write Vk1,...,ks = Vπ. It is convenient to
allow ki = 0 for some i. In that case, the corresponding set in π is empty.
For σ1 ∈ Sn and σ2 ∈ Sm, we write σ1 × σ2 for the permutation in Sn+m that is given by
(σ1 × σ2)(i) =(σ1(i)
if i ≤ n
σ2(i − n) + n if i > n
The identity element in Sn is denoted by idn, or just id when n is clear from the context. We
write σk,l for the permutation in Sk,l that satisfies
(5)
σk,l(i) =(i + l
i − k
if i ≤ k
if i > k
Lemma 1.1. With this notation, we get the following relation between sets of the form
Vk1,...,ks:
8
STEVEN DEPREZ
• For every n, m ∈ N1 and for all k ≥ −n, we get
Vn+k,m =
min(n,m)Gl=max(−k,0)
(Vn−l,l × Vk+l,m−l)(idn−l ×σk+l,l × idm−l).
Proof. The first two assertions follow immediately from the fact that every element of σ ∈
Vn,m,k is the unique element of shortest length in the coset σ(Sn × Sm × Sk).
The last assertion is most easily proven with the characterisation of Vn,m in terms of the order
on {1, . . . , n + m}. The argument is best explained using the following picture:
idn−l ×σk+l,l × idm−l
σ1 × σ2
m
n+k
l
k+l
n
m+k
The elements σ ∈ Vn+k,m are precisely the permutations that are increasing on each of
{1, . . . , n + k} and {n + k + 1, . . . , n + k + m} separately. Set l to be the number of elements of
{n + k + 1, . . . , n + k + m} that are mapped into {1, . . . , n}. These elements from necessarily
an initial segment {n + k + 1, . . . , n + k + l}, because σ is increasing on {n + k + 1, . . . , n +
k + m}. Because σ is a permutation, it maps the elements {n − l + 1, . . . , n + k} into
the set {n + 1, . . . , m + n + k}.
It is now clear that σ can be uniquely decomposed as
σ = (idn−l ×σk+l,l×idm−l)(σ1 ×σ2) where σ1 ∈ Vn−l,l and σ2 ∈ Vk+l,m−l. Conversely, it is easy
to see that every such element is increasing on {1, . . . , n+k} and {n+k+1, . . . , n+k+m}. (cid:3)
1.2. Bimodules. Let M be a von Neumann algebra. A left M -module is a Hilbert space H
with a normal representation λ : M → B(H). Given a normal tracial state τ on M , we can
perform the GNS construction: on M we consider the inner product given by hx, yi = τ (x∗y).
The completion of M with respect to this inner product is the GNS construction of M and is
written as L2(M, τ ). We also write kxk2
2 = τ (x∗x). The Hilbert space L2(M, τ ) comes with
a natural map M ∋ x 7→ x ∈ L2(M, τ ), and this map has dense range. There is a natural
which trace we use, we just write L2(M ). Left M -modules over von Neumann algebras are
representation of M on L2(M, τ ) that is given by λ(x)y =cxy. If it is clear from the context
not very interesting: every left M -module H is isomorphic toLi L2(M, τ )pi where the pi are
projections in M and τ is any faithful normal tracial state on M .
RADIAL MULTIPLIERS ON VON NEUMANN ALGEBRAS
9
An M -M bimodule is a Hilbert space H with two normal representations λ : M → B(H)
and ρ : M op → B(H) such that λ(x) commutes with ρ(y) for all x ∈ M and y ∈ M op.
We will write xξy = λ(x)ρ(yop)ξ for all x, y ∈ M . The GNS construction L2(M, τ ) is an
M -M -bimodule where ρ is given by ρ(xop)y =cyx.
Let M be a von Neumann algebra and fix a trace τ on M . Let H be an M -M -bimodule and
ξ ∈ H a vector. Then we can define an unbounded operator l(ξ) : L2(M ) → H by the formula
l(ξ)x = ξx. We say that a vector ξ is left-bounded if l(ξ) is a bounded operator. When ξ, η ∈
H are left-bounded vectors, then we can define an operator l(ξ)∗l(η) : L2(M, τ ) → L2(M, τ ).
This operator commutes with the right representation of M on L2(M, τ ), so it is of the form
λ(x) = l(ξ)∗l(η) for some x ∈ M . We denote this x ∈ M by hξ, ηiM . The map h·, ·iM is called
the (right) M -valued inner product on H. It is easy to see that this inner product satisfies
hxξy, ηziM = y∗ hξ, x∗ηiM z for all x, y, z ∈ M , and hξ, ηi∗
M = hη, ξiM . Moreover, the norm of
ξ is given by kξk2 = τ (hξ, ξiM ).
We also consider the right multiplication operator r(ξ) that is given by r(ξ)x = xξ. A vector
ξ is said to be right-bounded if r(ξ) is a bounded operator. For right-bounded operators
ξ, η ∈ H, we can also consider the operator r(ξ)∗r(η) on L2(M ). This operator now commutes
with the left representation of M , so it is of the form ρ(xop)∗ for some x ∈ M . This element
x is denoted by Mhξ, ηi. The map Mh·, ·i is called the left M -valued inner product on H. It
satisfies the relation Mhxξz, yηi = x Mhξ, ηz∗i y∗.
We say that a vector is bi-bounded if it is both left and right-bounded. We denote the space
of all bi-bounded vectors in H by
◦
H. The space of all bi-bounded vectors is dense in H.
Let H, K be two M -M bimodules. The Connes tensor product H ⊗M K is the M -M bimodule
◦
H ⊗alg
◦
K, we define an inner product by
defined as follows. On the algebraic tensor product
the formula
hξ ⊗ η, ζ ⊗ θi = hη, hξ, ζiM θi = hξ Mhη, θi , ζi = τ ( Mhη, θi∗ hξ, ζi).
◦
The Connes tensor product H ⊗M K is the result of separation and completion of
K
with respect to this inner product. Observe that when either ξ is left-bounded or η is right-
bounded, the elementary tensor ξ ⊗ η ∈ H ⊗M K is well-defined. These elementary tensors
satisfy the relation ξx ⊗ η = ξ ⊗ xη for all x ∈ M .
◦
H ⊗alg
Lemma 1.2. Let M be a finite von Neumann algebra and let H, K be M -M bimodules
• The set of left-bounded vectors in H with the property that hξ, ξiM is a projection in
M , densely spans H.
• The set of elementary tensors ξ ⊗ η where hξ, ξiM and Mhη, ηi are the same projection
in M , densely spans H ⊗M K.
Proof. To prove the first point, we will show that every left-bounded vector ξ ∈ H is the sum of
two vectors ξ1, ξ2 with the property that hξi, ξiiM is a projection for i = 1, 2. Set x =phξ, ξiM
and denote ξn = ξ( n
nx+1 increases to p = χ(0,∞)(x).
In particular, it follows that xn converges to p in k·k2 and hence is a k·k2-Cauchy sequence.
But kξn − ξmk = kxn − xmk2 for all n, m ∈ N. So also the sequence (ξn)n is a Cauchy
nx+1 ). Observe that xn :=phξn, ξniM = nx
10
STEVEN DEPREZ
sequence, and hence has a limit η. This limit satisfies hη, ηiM = p and moreover, ξ = ηx.
It suffices now to observe that x is the linear combination x = 1
2 kxk (u+ + u−) of the two
unitaries u± = x
kxk ± iq1 − x2
kxk2 .
For the second point, we will show that every elementary tensor product ξ ⊗ η where ξ is
left-bounded and η is right-bounded, can be written as a linear combination of two elementary
tensors ξi ⊗ ηi (i = 1, 2) with the property that hξi, ξiiM and Mhηi, ηii are the same projection
(i = 1, 2). As in the proof of the first point, we find vectors ζ ∈ H and θ ∈ K and x ∈ M such
that ξ ⊗ η = ζ ⊗ xθ and such that p = hζ, ζiM and q = Mhθ, θi are projections. We can assume
that p is the smallest projection in M that satisfies px = x and similarly that q is minimal with
the property that xq = x. Write the polar decomposition of x by x = v x. Then we know
that vv∗ = p, v∗v = q and x ∈ M is a positive element. We write x = 1
2 kxk (u+ + u−)
where u± = x
kxk2 as before. Observe that u± commutes with x and hence
with its support q. In particular, we get that v± = vu± satisfies v±v∗
±v± = q.
Now we see that the vectors ξ1 = ξ2 = ζ, η1 = v+θ and η2 = v−θ satisfy our condition:
ξ ⊗ η = ξ1 ⊗ η1 + ξ2 ⊗ η2.
(cid:3)
kxk ± ir1 − x2
± = p and v∗
2. The relative Gaussian Construction
In this section we define the relative Gaussian construction, giving formal definitions for all
the concepts. For a more accessible account of the construction, we refer to the introduction.
For the rest of this section, fix the following data.
• a finite von Neumann algebra M with a trace τ
• an M -M bimodule H
• a self-adjoint M -M bimodular contraction F : H ⊗M H → H ⊗M H that satisfies the
braid relation
(F ⊗ 1H )(1H ⊗ F )(F ⊗ 1H ) = (1H ⊗ F )(F ⊗ 1H)(1H ⊗ F ).
We denote the n-fold Connes tensor product by H (n) = H ⊗M H . . . ⊗M H. By convention
we write H (0) = L2(M ) and H (1) = H. We will write the identity operator of H (n) by 1n.
On each H (n), we define M -M bimodular operators(cid:16)F (n)
σ (cid:17)σ∈Sn
in the following way.
We freely use the notations introduced in subsection 1.1. For a transposition t = (i, i + 1) of
consecutive numbers, we set
If σ = ti1 . . . tik is a decomposition of minimal length, then the M -M bimodular operator
F (n)
t = −1i−1 ⊗ F ⊗ 1n−i−1.
σ = F (n)
F (n)
ti1
. . . F (n)
tik
does not depend on the choice of the (minimal) decomposition. When n, m ∈ N, we also write
Fn,m = Fσn,m where σn,m is as defined by (5) in subsection 1.1.
RADIAL MULTIPLIERS ON VON NEUMANN ALGEBRAS
11
We set
D(n) = Xσ∈Sn
F (n)
σ .
By [BS2, theorem 1.1], this operator D(n) is positive definite. By convention, we set D(0) = 1
and D(1) = 1H .
We define a new inner product on H (n) by the formula hξ, ηiF = (cid:10)ξ, D(n)η(cid:11). Apply sep-
aration/completion to this new inner product and denote the resulting M -M bimodule by
H (n)
and observe that this is a bounded M -M
bimodular operator that has dense range. We denote the identity operator on H (n)
F by idn.
F . Consider the natural map I (n) : H (n) → H (n)
F
Lemma 2.1. For every n, m ∈ N, we get the following results
• The identity operator from on H (n+m) extends uniquely to a bounded operator
In,m : H (n)
F ⊗M H (m)
F → H (n+m)
F
.
• The operator Fn,m on H (n+m) extends uniquely to a contraction
Fn,m : H (n)
F ⊗M H (m)
F → H (m)
F ⊗M H (n).
Proof. To prove the first point, observe that D(n+m) = En,m(cid:0)D(n) ⊗ D(m)(cid:1). In particular,
n (cid:19)2(cid:16)D(n) ⊗ D(m)(cid:17)2
(cid:16)D(n+m)(cid:17)2
n,mEn,m(cid:16)D(n) ⊗ D(m)(cid:17) ≤(cid:18)n + m
=(cid:16)D(n) ⊗ D(m)(cid:17) E∗
.
Since the square root is an operator-monotone function, we also get that
and hence that
D(n+m) ≤(cid:18)n + m
≤(cid:18)n + m
n (cid:19)(D(n) ⊗ D(m))
n (cid:19) kξk
(n)
F ⊗M H
(n+m)
F
H
H
kξk
.
(m)
F
for all ξ ∈ H (n+m). So indeed, the identity operator on H (n+m) extends uniquely to a bounded
operator
In,m : H (n) ⊗M H (m) → H (n+m).
The norm of this operator is less than(cid:0)n+m
n (cid:1).
As to the last point, remark that
σn,m(σ1 × σ2) = (σ2 × σ1)σn,m
for all permutations σ1 ∈ Sn and σ2 ∈ Sm. Moreover, counting the number of inversions, it is
easy to see that
σn,m(σ1 × σ2) = σn,m + σ1 + σ2 = (σ2 × σ1)σn,m .
12
STEVEN DEPREZ
It follows that Fn,m(cid:0)D(n) ⊗ D(m)(cid:1) =(cid:0)D(m) ⊗ D(n)(cid:1) Fn,m. In particular, we see that F ∗
commutes with D(n) ⊗ D(m) and hence with the square root(cid:0)D(n) ⊗ D(m)(cid:1) 1
2 . So we see that
n,mFn,m
kFn,mξk
H
(m)
F ⊗M H
(n)
F
n,m(cid:16)D(m) ⊗ D(n)(cid:17) Fn,mξE
n,mFn,m(cid:16)D(n) ⊗ D(m)(cid:17) ξE
=Dξ, F ∗
=Dξ, F ∗
=(cid:28)ξ,(cid:16)D(m) ⊗ D(n)(cid:17) 1
≤Dξ,(cid:16)D(n) ⊗ D(m)(cid:17) ξE
2 F ∗
= kξk
.
H
(n)
F ⊗M H
(m)
F
2 ξ(cid:29)
n,mFn,m(cid:16)D(m) ⊗ D(n)(cid:17) 1
for all ξ ∈ H (n+m)
F
. So Fn,m extends uniquely to a contraction
Fn,m : H (n)
F ⊗M H (m)
F → H (m)
F ⊗M H (n)
F .
Definition 2.2. The relative Fock space of M , H and F is defined to be
(cid:3)
We denote the algebraic direct sum by
F M (H, F ) =Mi
algMi
M (H, F ) =
F 0
H (n)
F .
H (n)
F .
Now we can define creation and annihilation operators. Let T : H (n)
right M -modular operator. Then we define a right-M -modular operator L0(T ) : F 0
F 0
M (H, F ) by the relation that
F → H (m)
F
be a bounded
M (H, F ) →
L0(T )η = Im,k(T ⊗ idk)I ∗
n,kη
L0(T )η = 0
whenever η ∈ H (n+k)
F
whenever η ∈ H (k)
F with k < n.
Each of the operators Im,k is bounded, but as k → ∞, their norm may tend to ∞. This is the
case for the classical Gaussian construction. In such a case, the operators L0(T ) need not be
bounded. But the operators L0(T ) are closable because they have a densely defined adjoint:
L0(T ∗) ⊂ L0(T )∗.
Definition 2.3. The creation operator on F M (H, F ) associated to T is the closure L(T ) of
the densely defined operator L0(T ).
For every right-bounded vector in ξ ∈ H, we get that the formula l(ξ)x = ξx defines a bounded
right-M -modular map l(ξ) : L2(M ) → H. The operator L(l(ξ)) is called the creation operator
of ξ. We also denote this operator by L(ξ). The operator L(l(ξ)∗) = L(ξ)∗ is the corresponding
annihilation operator.
RADIAL MULTIPLIERS ON VON NEUMANN ALGEBRAS
13
Proposition 2.4. The creation operators satisfy the following composition rule: for bounded
right-M -modular operators S : H (n1)
and T : H (n2)
, we get that
F → H (m2)
F
F → H (m1)
F
(6)
L(S)L(T ) =
min(n1,m2)Xk=0
In particular, the space
L
Im1,m2−k(S ⊗ idm2−k)(Ik,n1−k ⊗ idm2−k)
(idk ⊗Fm2−k,n1−k)
(I ∗
k,m2−k ⊗ idn1−k)(T ⊗ idn1−k)In2,n1−k
.
TM (H, F ) = span{L(T ) T : H (n)
F → H (m)
F
for some n, m ∈ N}
is a ∗-algebra whose commutant is the right action of M on F M (H, F ).
Proof. It is clear from the definition of In,m that, for all n, m, l ∈ N,
In+m,l(In,m ⊗ idl) = In,m+l(idn ⊗Im,l),
since both are the unique extension of the identity operator to a bounded linear map from
H (n)
F to H (n+m+l)
F ⊗M H (m)
F ⊗M H (l)
F
.
Translating lemma 1.1 in terms of the operators In,m and Fn,m yields the following relation
I ∗
n1,l−n1Im2,l−m2 =
min n1,m2Xk=max(0,n1+m2−l)
(Ik,n1−k ⊗ Im2−k,l+k−n1−m2)
(idk ⊗Fm2−k,n1−k ⊗ idl+k−n1−m2 )
n1−k,l+k−n1−m2)
k,m2−k ⊗ I ∗
(I ∗
It is clear that both the left hand side and the right hand side of (6) give 0 when they are
evaluated in an η ∈ H (l)
F with l < n2 or l < n2 + n1 − m2. Let l ≥ max(n1, m2) and take a
vector η ∈ H (l−m2+n2). Then we compute that
L(S)L(T )η = Im1,l−n1(S ⊗ idl−n1)I ∗
n1,l−n1Im2,l−m2(T ⊗ idl−m2)I ∗
n2,l−m2η
=
min n1,m2Xk=max(0,n1+m2−l)
Im1,l−n1(S ⊗ idl−n1)(Ik,n1−k ⊗ Im2−k,l+k−n1−m2)
(idk ⊗Fm2−k,n1−k ⊗ idl+k−n1−m2)
(I ∗
k,m2−k ⊗ I ∗
n1−k,l+k−n1−m2)(T ⊗ idl−m2)I ∗
n2,l−m2η
=
=
as claimed.
min n1,m2Xk=max(0,n1+m2−l)
L
min(n1,m2)Xk=0
Im1+m2−k,l+k−n1−m2(Im1,m2−k ⊗ idl+k−n1−m2)
(S ⊗ idl−n1)(Ik,n1−k ⊗ idl−n1)
(idk ⊗Fm2−k,n1−k ⊗ idl+k−n1−m2)
(I ∗
k,m2−k ⊗ idl−m2)(T ⊗ idl−m2)
(I ∗
n2,n1−k ⊗ idl+k−n1−m2)I ∗
n1+n2−k,l+k−m2−n1η
Im1,m2−k(S ⊗ idm2−k)(Ikn1 − k ⊗ idm2−k)
(idk ⊗Fm2−k,n1−k)
(I ∗
k,m2−k ⊗ idn1−k)(T ⊗ idn1−k)In2,n1−k
η.
14
STEVEN DEPREZ
It remains to show that the commutant of TM (H, F ) is just the right action of M .
It is
clear that right multiplication by M commutes with every operator in TM (H, F ). Suppose
that T ∈ B(F M (H, F )) satisfies T L(S) ⊂ L(S)T for all bounded right-M -modular operators
S : H (n)
F be a left-bounded vector. Then we see that
F → H (m)
F . Let ξ ∈ H (k)
T ξ = T L(l(ξ))Ω = L(l(ξ))T Ω,
so T is completely determined by its value T Ω. On the other hand,
hξ, T Ωi = hΩ, L(l(ξ)∗)T Ωi = hΩ, T L(l(ξ)∗)Ωi = 0
for every ξ ∈ H (k)F with k > 0. Hence, T Ω ∈ L2(M ). Since T is a bounded operator, it is
clear that T Ω = x for some x ∈ M , so T is given by right multiplication by x.
(cid:3)
We write T ′′
M (H, F ) for the bicommutant of TM (H, F ), i.e. all the bounded linear operators
that commute with the right action of M . As a von Neumann algebra, this is not a very
M (H, F ) ∼= B(ℓ2(N)) ⊗ M , but later on we will consider more
interesting object, since T ′′
interesting subalgebras.
Consider the orthogonal projection P : F M (H, F ) → L2(M ), and observe that P T P ∗ ∈
B(L2(M )) commutes with the right action of M , for all T ∈ T ′′
M (H, F ). So we can define a
map E : T ′′
M (H, F ) → M by the formula E(T ) = P T P ∗.
Lemma 2.5. The map E : T ′′
faithful.
M (H, F ) → M is a normal conditional expectation, but not
Proof. It is clear that E is a normal unital completely positive map. The map E is also M -M
bimodular because P intertwines the left actions of M . To show that E is not faithful, take
any nonzero bounded right-M linear map T : H (n)
F → L2(M ) with n 6= 0. Then we see that
L(T ∗)P L(T ) ∈ T ′′
M (H, F ) is a non-zero positive bounded operator with
E(L(T ∗)P L(T )) = 0.
(cid:3)
The relative Gaussian construction will be a special kind of subalgebra of T ′′
M (H, F ) on which
E is faithful. In order to define the relative Gaussian construction, we need one more piece
of data, namely an anti-unitary operator J : H → H that satisfies the following relations
• J is an involution, i.e. J 2 = id1.
• J intertwines the left and right representations of M on H, i.e.
J(xξy) = y∗J(ξ)x∗ for all x, y ∈ M and ξ ∈ H.
• J is compatible with F in the sense that
(7)
(l(ζ)∗ ⊗ id1)F (Jη ⊗ ξ) = J(l(ξ)∗) ⊗ 1)F (η ⊗ ζ),
for all bi-bounded vectors ξ, ζ ∈ H and all vectors η ∈ H.
RADIAL MULTIPLIERS ON VON NEUMANN ALGEBRAS
15
Then we want to define the algebraic relative Gaussian construction ΓM (H, F, J) to be the
subalgebra of TM (H, F ) that is generated by the left action of M and by elements of the form
W (ξ) = L(ξ)+L(Jξ)∗ where ξ is a bi-bounded vector in H. The relative Gaussian construction
Γ′′
M (H, F, J) will then be the von Neumann algebra generated by ΓM (H, F, J). Observe that
Jξ is a left-bounded vector if and only if ξ is right-bounded. More precisely, l(Jξ)x = Jr(ξ)x∗.
In particular, we get that hJξ, JηiM = Mhξ, ηi. We consider the elements of M to be reduced
words of length 0 and elements of the form W (ξ) are interpreted as reduced words of length 1.
These reduced words satisfy xΩ = x ∈ L2(M ) ⊂ F M (H, F ) and W (ξ)Ω = ξ. We will define
longer "reduced words" W (ξ) for vectors ξ ∈ H (n)
F by a generalization of the Wick formula.
The operator W (ξ) will be the unique element in ΓM (H, F, J) that satisfies W (ξ)Ω = ξ.
We extend J to an anti-unitary operator eJ on F M (H, F ). First, define an anti-unitary
operator J (n) on H (n) by the relation that J (n)(ξ1 ⊗ . . . ⊗ ξn) = Jξn ⊗ . . . ⊗ Jξ1. Observe that
this does indeed define an anti-unitary because by induction we get
DJ (n)(ξ1 ⊗ . . . ⊗ ξn), J (n)(η1 ⊗ . . . ⊗ ηn)E
=DJ (n−1)(ξ1 ⊗ . . . ⊗ ξn−1), hJξn, JηniM J (n−1)(η1 ⊗ . . . ⊗ ηn−1)E
=DJ (n−1)(ξ1 ⊗ . . . ⊗ ξn−1), J (n−1)(η1 ⊗ . . . ⊗ ηn−1 Mhηn, ξni)E
= h(η1 ⊗ . . . ⊗ ηn−1) Mhηn, ξni , ξ1 ⊗ . . . ⊗ ξn−1i
= hη1 ⊗ . . . ⊗ ηn, ξ1 ⊗ . . . ⊗ ξni
By convention, we set J (1) = J and J (0) x =cx∗. The extended operator
is now the direct sum of the operators J (n).
eJ : F M (H, F ) → F M (H, F )
Lemma 2.6. We get the following relation between J (n) and F (n)
σ
for σ ∈ Sn:
(8)
J (n)F (n)
σ J (n) = Fγσγ−1 ,
where γ ∈ Sn is defined by γ(i) = n − i + 1.
Proof. Once we prove that J (2) commutes with F , it is clear that J (n)(idk−1 ⊗F idn−k−1)J (n) =
idn−k−1 ⊗F ⊗ idk−1 for all 1 ≤ k ≤ n − 1. So (8) holds for the generators of Sn, and hence
for all of Sn. The fact that J (2) commutes with F follows from our compatibility relation
between F and J: observe that (8) is equivalent to the fact that, for all bi-bounded vectors
ξ1, ξ2, η1, η2, we have that
hξ1 ⊗ ξ2, F (Jη1 ⊗ η2)i = hF (η1 ⊗ ξ1), η2 ⊗ Jξ2i .
16
STEVEN DEPREZ
Now we see that
Dξ1 ⊗ ξ2, J (2)F (η1 ⊗ η2)E = hF (η1 ⊗ η2), Jξ2 ⊗ Jξ1i
=Dξ1 ⊗ ξ2, F J (2)(η1 ⊗ η2)E
= hη2 ⊗ ξ1, F (Jη1 ⊗ Jξ2)i
= hF (η2 ⊗ ξ1), Jη1 ⊗ Jξ2i
(cid:3)
Definition 2.7. Let n, m ∈ N and let ξ ∈ H (n+m)
F → H (m)
Sn,m(ξ) : H (n)
bounded vectors η ∈ H (n).
F
F
by setting Sn,m(ξ)η = J (m)(l(η)∗ ⊗ idm)I ∗
. We define an unbounded operator
n,mJ (n+m)ξ for all left-
We say that a vector ξ ∈ H (n)
F
k = 0, . . . , n. The space of all bounded vectors in H (n)
F
is denoted by
◦
H (n)
F .
is bounded if Sk,n−k(ξ) is a bounded operator for all integers
When ξ ∈ H = H (1)
F , then we see that S0,1(ξ) = l(ξ) while S1,0(ξ) = l(Jξ)∗. In particular, a
F is bounded if and only if it is bi-bounded, and a vector ξ ∈ L2(M ) = H (0)
vector ξ ∈ H = H (1)
is bounded if and only if ξ = x for some x ∈ M . Observe that Sn,m(xξy + η) = xSn,m(ξ)y +
Sn,m(η). We can also describe Sn,m(ξ) by the relation that
F
hη, Sn,m(ξ)ζi =DIm,n(cid:16)η ⊗ J (n)ζ(cid:17) , ξE .
In the special case where F = 0, we see that
L(Sn,m(ξ1 ⊗ . . . ⊗ ξn ⊗ η1 ⊗ . . . ⊗ ηm)) = L(ξ1) . . . L(ξn)L(Jη1)∗ . . . L(Jηm)∗,
for all sets of bi-bounded vectors ξ1, . . . , ξn, η1, . . . , ηm ∈ H.
Definition 2.8. Let ξ ∈ H (n)
F
be a bounded vector. Then we define the word of ξ to be
(9)
W (ξ) =
nXk=0
L(Sk,n−k(ξ))
Observe that W (ξ)Ω = ξ. For a bi-bounded vector ξ ∈ H, the formula above reads W (ξ) =
L(ξ) + L(Jξ)∗. When F = 0, the formula above reduces to the usual Wick formula
W (ξ1 ⊗ . . . ⊗ ξn) =
nXk=0
L(ξ1) . . . L(ξk)L(Jξk+1)∗ . . . L(Jξn)∗.
For this reason, we still call (9) the Wick formula for the relative Gaussian construction.
We define bilinear maps ⊠k :
F
×
◦
H (k+m)
◦
→ H (n+m)
H (n+k)
ξ ⊠k η = In,m(Sk,n(ξ) ⊗ idm)I ∗
F
F
k,mη,
by the formula
RADIAL MULTIPLIERS ON VON NEUMANN ALGEBRAS
17
or equivalently,
for all ζ ∈ H (n)
F
hζ ⊗ θ, ξ ⊠k ηi =DI ∗
and θ ∈ H (m)
n,kJ (n+k)ξ ⊗ θ, J (n)θ ⊗ I ∗
k,mηE
bounded vectors ξ ∈
F . Observe that J (n+m)(ξ ⊠ η) = J (n+k)η ⊠ J (m+k)ξ for all
, η ∈
◦
H (m+k)
.
F
◦
H (n+k)
F
In the case where F = 0, these bilinear maps can also be described by
(ξ1 ⊗ ξ2) ⊠k (η1 ⊗ η2) = ξ1 ⊗DJ (k)ξ2, η1EM
η2,
◦
H (k) and η ∈
◦
H (m).
◦
H (n), η1, ξ2 ∈
for all ξ1 ∈
Lemma 2.9. Let ξ ∈ H (n)
F
be a bounded vector. Then J (n)ξ is still a bounded vector and
W (ξ)∗ = W (Sξ).
Let ξ ∈ H (n)
F
vectors, for all k = 0, . . . , min(n, m). Moreover, we get the following product formula
be bounded vectors. Then the vectors ξ ⊠k η are still bounded
and η ∈ H (m)
F
W (ξ)W (η) =
min(n,m)Xk=0
W (ξ ⊠k η).
In particular, the space
ΓM (H, F, J) = span{W (ξ) ξ ∈ H (n)
F
is bounded, n ∈ N}
is a ∗-algebra. We call this ∗-algebra the algebraic relative Gaussian construction.
Proof. In order to prove this theorem, it will be convenient to use the following variants of
Sn,m and ⊠k. Let K, L be M -M bimodules. For a vector ξ ∈ K ⊗M L, we define an operator
TL,K(ξ) : L → K by the formula
hη, TL,K (ξ)ζi =(cid:10)η ⊗ ζ, ξ(cid:11) ,
for all η ∈ K and for all left-bounded vectors ζ ∈ L. This is not necessarily a bounded
operator, but it is a closable, densely define unbounded operator. Indeed, TL,K is closable
because its adjoint is given by
TL,K(ξ)∗ ⊃ TK,L(ξ),
where we identified K ⊗M L with L ⊗M K. We denote the closure of TL,K still by TL,K. We
say that ξ is bounded for the tensor product decomposition K ⊗M L if this operator TL,K(ξ)
is bounded.
Let K, L1, L2 be M -M bimodules. For vectors ξ ∈ L1 ⊗M K and η ∈ K ⊗M L2 that are
bounded in these tensor product decompositions, we write ξ ⊡K η = (TK,L1 ⊗ idL2)η, or
equivalently, ξ ⊡K η is the unique vector in L1 ⊗M L2 that satisfies
hζ1 ⊗ ζ2, ξ ⊡K ηi =(cid:10)ξ ⊗ ζ2, ζ1 ⊗ η(cid:11) .
18
STEVEN DEPREZ
Observe that the anti-unitary operator J : H → H can also be interpreted as an M -M
bimodule isomorphism J : H → H. This is also true for the operators J (n) on H (n)
F . The
original Sn,m and ⊠k can easily be expressed in terms of TL,K and ⊡K.
Sn,m(ξ) = T
H
(n)
F ,H
(m)
F
((idm ⊗J (n))I ∗
m,nξ)
for ξ ∈ H (n+m)
F
ξ ⊠k η = In,m(((id ⊗J (k))I ∗
n,kξ) ⊡ (I ∗
k,mη))
for ξ ∈ H (n+k)
F
and η ∈ H (k+m)
F
.
Moreover, a vector ξ ∈ H (n)
F
the tensor product decomposition H (n)
is bounded if and only if (idk ⊗J (n−k))E∗
, for all k = 0, . . . , n.
F ⊗M H (n−k)
F
n,n−kξ is bounded for
It is now clear that, whenever ξ ∈ H (n)ξ is a bounded vector, then J (n)ξ is still a bounded
vector, and Sk,n−k(ξ)∗ = Sn−k,k(J (n)ξ). It follows that W (J (n)ξ) = W (ξ)∗.
We prove the product formula in three steps.
step 1: For all n, m and for all bi-bounded vectors ξ, θ ∈ H (m) and all vectors η ∈ H (n), we
get
(10)
J (n)(l(ξ)∗ ⊗ idn)Fn,m(η ⊗ ζ) = (l(ζ)∗ ⊗ idn)Fn,m(J (n)η ⊗ ξ).
Remark that (10) can be rewritten to the relation that
DFn,m(η ⊗ ζ), ξ ⊗ J (n)θE =Dζ ⊗ θ, Fn,m(J (n)η ⊗ ξ)E ,
for all bi-bounded vectors ξ, ζ ∈ H (m)
F . Observe that, for fixed
bi-bounded vectors ξ, ζ, the left and right hand sides of the equation are continuous in η, θ,
so we only have to check this for dense sets of η, θ ∈ H (n)
F .
and all vectors η, θ ∈ H (n)
F
First we observe that the relation (10) is symmetric in n, m: if Fn,m satisfies (10), then we
see that, for all bi-bounded vectors ξ, ζ ∈ H (n)
F
and η, θ ∈ H (m)
F
DFm,n(η ⊗ ζ), ξ ⊗ J (m)θE =Dη ⊗ ζ, Fn,m(ξ ⊗ J (m)θ)E
=DFn,m(J (n)ξ ⊗ η), J (m)θ ⊗ J (n)ζE
=Dζ ⊗ θ, Fm,n(J (m)η ⊗ ξ)E ,
where we used the fact that F ∗
n,m = Fm,n = J (n+m)Fn,mJ (n+m). So also Fm,n satisfies (10).
We will prove (10) by induction on m and n. So it suffices to show that Fn+k,m satisfies (10),
whenever Fn,m and Fk,m satisfy (10). Note that
Fn+k,m = (idm ⊗In,k)(Fn,m ⊗ idk)(idn ⊗Fk,m)(I ∗
n,k ⊗ idm).
Let ξ, ζ ∈ H (m)
be bi-bounded vectors and let η1, θ2 ∈ H (n), η2, θ1 ∈ H (k) be vectors such that
hη1, η1iM = p = Mhη2, η2i and such that hθ1, θ1iM = q = Mhθ2, θ2i for projections p, q ∈ M .
F
RADIAL MULTIPLIERS ON VON NEUMANN ALGEBRAS
19
Let (χi)i be an orthonormal basis for pH (m)
F q. Then we compute that
D(Fn,m ⊗ idk)(idn ⊗Fk,m)(η1 ⊗ η2 ⊗ ζ), ξ ⊗ J (n)θ2 ⊗ J (k)θ1E
D(Fn,m ⊗ idk)(η1 ⊗ χi ⊗ J (k)θ1), ξ ⊗ J (n)θ2 ⊗ J (k)θ1E
=Xi
D(idn ⊗Fk,m)(η1 ⊗ η2 ⊗ ζ), η1 ⊗ χi ⊗ J (k)θ1E
=Xi DFn,m(η1 ⊗ χi), ξ ⊗ J (n)θ2EDFk,m(η2 ⊗ ζ), χi ⊗ J (k)θ1E
=Xi Dζ ⊗ θ1, Fk,m(J (k)η2 ⊗ χi)EDχi ⊗ θ2, Fn,m(J (n)η1 ⊗ ξ)E
=Xi
DJ (k)η2 ⊗ χi ⊗ θ2, (idk ⊗Fn,m)(J (k)η2 ⊗ J (n)η1 ⊗ ξ)E
=Dζ ⊗ θ1 ⊗ θ2, (Fk,m ⊗ idn)(idk ⊗Fn,m)(J (k)η2 ⊗ J (n)η1 ⊗ ξ)E
Dζ ⊗ θ1 ⊗ θ2, (Fk,m ⊗ idn)(J (k)η2 ⊗ χi ⊗ θ2)E
This implies that Fn+k,m satisfies (10).
step 2: Let n, m ∈ N and let L1, L2, K be M -M bimodules. Then for all vectors ξ ∈
L1 ⊗M H (n)
F ⊗M L2 that are bounded in all the tensor product
decompositions above, we get that
F ⊗M K and η ∈ K ⊗M H (m)
(11) T
L2⊗M H
(n)
F ,L1⊗M H
(m)
F
((idL1 ⊗ id
H
(m)
F
⊗J (n) ⊗ idL2
)(idL1 ⊗Fn,m ⊗ idL2
)(ξ ⊡K η))
=(cid:16)T
K⊗M H
(n)
F ,L1(cid:16)(idL1 ⊗J (n) ⊗ idK )ξ(cid:17) ⊗ idm(cid:17) (idK ⊗Fm,n)(cid:16)T
L2,K⊗M H
(η) ⊗ idn(cid:17) .
(m)
F
The left-hand side in (11) is a closed operator while the right-hand side is bounded, so we
only have to check (11) on a dense subset of the domain of the left-hand side. Let ζ1 ∈ H (n)
F ,
ζ2 ∈ L2, θ1 ∈ L1 and θ2 ∈ H (m) be bi-bounded vectors. Then we have to show that
Dθ1 ⊗ θ2 ⊗ J (n)ζ2 ⊗ ζ1, (idL1 ⊗Fn,m ⊗ idL2)(ξ ⊡K η)E
=DT
F (cid:16)(idL1 ⊗J (n) ⊗ idK)ξ(cid:17) θ1 ⊗ θ2, (idK ⊗Fm,n)(cid:16)T
L1,K⊗M H
(n)
L2,K⊗M H
(η)ζ1 ⊗ ζ2(cid:17)E .
(m)
F
For fixed bi-bounded vectors ζ1, ζ2, θ1, θ2, both the left hand side and the right hand side
of the above expression is continuous in ξ, provided η is bounded for both tensor product
decompositions of K ⊗M H (m)
F ⊗M L2. The same is true about continuity in the variable ξ.
Hence we can assume that ξ = ξ1 ⊗ ξ2 ⊗ ξ2 and η = η1 ⊗ η2 ⊗ η3 where all of ξ1, ξ2, ξ3, η1, η2, η3
are bi-bounded vectors.
20
STEVEN DEPREZ
Now the left hand side reduces to
Dθ1 ⊗ θ2 ⊗ J (n)ζ2 ⊗ ζ1, (idL1 ⊗Fn,m ⊗ idL2)(ξ ⊡K η)E
=Dθ1 ⊗ θ2 ⊗ J (n)ζ2 ⊗ ζ1, (idL1 ⊗Fn,m ⊗ idL2)(ξ1 ⊗ ξ2 ⊗(cid:10)ξ3, η1(cid:11)M η2 ⊗ η3)E
=Dhξ1, θ1iM θ2 ⊗ J (n)(ζ2) M(cid:10)ζ1, η3(cid:11) , Fn,m(ξ2 ⊗(cid:10)ξ3, η1(cid:11)M η2)E
=Dhξ1, θ1iM θ2 ⊗ J (n)(hη3, ζ1iM ζ2), Fn,m(ξ2 ⊗(cid:10)ξ3, η1(cid:11)M η2)E
The right hand side reduces to
(n)
F
L1,K⊗M H
((idL1 ⊗J (n) ⊗ idK )ξ)θ1 ⊗ θ2, (idK ⊗Fm,n)(T
DT
=Dξ3 ⊗ J (n)ξ2 ⊗ hξ1, θ1iM θ2, (idK ⊗Fm,n)(η1 ⊗ η2 ⊗ hη3, ζ1iM ζ2)E
=DJ (n)ξ2 ⊗ hξ1, θ1iM θ2, Fm,n((cid:10)ξ3, η1(cid:11)M η2 ⊗ hη3, ζ1iM ζ2)E
=DFn,m(J (n)ξ2 ⊗ hξ1, θ1iM θ2),(cid:10)ξ3, η1(cid:11)M η2 ⊗ hη3, ζ1iM ζ2E
L2,K⊗M H
(η)ζ1 ⊗ ζ2)E
(m)
F
These expressions are equal by step 1.
step 3: An elementary but rather tedious computation using step 2 and proposition 2.4
shows that ξ ⊠k η is indeed a bounded vector for all k, whenever ξ ∈ H (n)
are
bounded vectors. Moreover we get that
F , η ∈ H (m)
F
W (ξ)W (η) =
min(n,m)Xk=0
W (ξ ⊠k η).
Definition 2.10. The relative Gaussian construction Γ′′
bra generated by the spectral projections of the operators in ΓM (H, F, J).
M (H, F, J) is the von Neumann alge-
(cid:3)
It follows from the result above that
for Γ′′
M (H, F, J). We define a state ϕ on Γ′′
◦
H (n)
F is dense in H (n)
F . In particular, Ω is a cyclic vector
M (H, F, J) by the relation that ϕ(x) = hΩ, xΩi.
Theorem 2.11. The state ϕ defined above is a faithful normal trace on Γ′′
satisfies ϕ = τ ◦ E.
faithful.
In particular, the conditional expectation E : Γ′′
M (H, F, J) that
M (H, F, J) → M is
Proof. We know already that Ω is a cyclic vector for Γ′′
M (H, F, J). Moreover, we see that
eJW (ξ)eJΩ = eJξ = W (eJξ)Ω. Hence it suffices to show that W (ξ) commutes with eJW (η)eJ .
RADIAL MULTIPLIERS ON VON NEUMANN ALGEBRAS
21
For each n ∈ N, we know that the space
W (ξ) are bounded when restricted to H (n)
is dense in H (n)
◦
H (n)
F
F . So it suffices to show that
F , and moreover, the operators
for all bounded vectors ξ ∈ H (n)
F . We compute that
F
and ζ ∈ H (k)
F , η ∈ H (m)
W (ξ)eJW (η)eJ W (ζ)Ω = eJ W (η)eJW (ξ)W (ζ)Ω,
eJW (η)eJ W (ζ)Ω = eJ
W (η ⊠i J (k)ζ)Ω
=
W (ζ ⊠i J (m)η)Ω
min(m,k)Xi=0
min(m,k)Xi=1
= W (ζ)W (J (m)η)Ω
So we find that
W (ξ)eJW (η)eJ W (ζ)Ω = W (ξ)W (ζ)W (J (m)η)Ω
=
=
W (ξ ⊠i ζ)W (η)Ω
min(n,k)Xi=0
min(n,k)Xi=0
= eJW (η)eJ W (ξ)W (ζ)Ω.
eJ W (η)eJW (ξ ⊠i ζ)Ω
It follows that W (ξ) commutes with eJ W (η)eJ, and hence that ϕ is a faithful normal trace. (cid:3)
3. Proof of the Main Result
This section is devoted to the proof of the following theorem. The proof resembles very closely
the proof of the main result of [HM], but our proof is a little less technical because the radial
structure of relative Gaussian constructions is easier to handle.
Theorem 3.1 (see theorem 0.1). Let M be a von Neumann algebra and let H be a Hilbert
M -M bimodule. Assume that F is a projection onto an M -M subbimodule of H ⊗M H such
that F ⊗ 1 commutes with 1 ⊗ F on H ⊗M ⊗M H. Let ψ : N → C be a function in class
C defined above. Then there is a unique ultraweakly continuous, completely bounded map
Φψ : T ′′
M (H, F ) that satisfies
M (H, F ) → BM T ′′
Φψ(L(T )) = ψ(n + m)L(T )
for all bounded right-M -linear operators T : H (n)
norm is less than kΦψkcb ≤ kψkC .
F → H (m)
F . Moreover, the completely bounded
22
STEVEN DEPREZ
We prove the theorem by a series of lemmas. For this section, fix a tracial von Neumann
algebra (M, τ ), an M -M bimodule H and an M -M bimodular projection F : H ⊗H → H ⊗H
that satisfies the relation
(F ⊗ id)(id ⊗F ) = (id ⊗F )(F ⊗ id).
The first lemma provides a way to estimate the completely bounded norm of an operator of
the form Φ(T ) =Pi uiT vi. It was proven by Christensen and Sinclair in [CS].
the formula Φ(T ) = Pi uiT v∗
Lemma 3.2 (see [CS, Corollary 6.2]). Let H, K be Hilbert spaces and (ui), (vi) sequences
i vi are bounded operators, then
of bounded operators from H to K.
i defines a completely bounded map Φ : B(H) → B(K). The
completely bounded norm of Φ is bounded by kΦk2
If Pi uiu∗
i and Pi v∗
cb ≤ kPi uiu∗
i k kPi viv∗
i k.
Whenever x ∈ ℓ∞(N), we define a radial multiplication operator Mx on F M (H, F ) by the
relation that Mxη = x(n)η whenever η ∈ H (n)
F . On ℓ∞(N), we consider the one-directional
shift S, which is defined by S(x)n = xn−1 when n > 0 and S(x)0 = 0. We denote the shift in
the other direction by S∗.
Remark 3.3. It is clear that the radial multiplication operators are M -M bimodular and they
satisfy the following relations with the creation operators.
MxL(T )My = Mx((S ∗)nSmy)L(T ) = L(T )My(Sn(S ∗)mx)
for all bounded right-M modular operators T : H (n)
F → H (m)
F .
We consider sequences z, rk ∈ ℓ∞(N) that are defined by zn = (−1)n and rk(n) = 1 whenever
n ≥ k and rk(n) = 0 otherwise. Observe that u = Mz is a unitary and that the qk = Mrk are
projections.
As a left M -module, we can write H in the form M H ∼=Li M L2(M, ϕ)pi. We write ξi for
the vector corresponding to 1pi in the i-th component of the direct sum above. Observe that
Mhξi, ξji = δi,jpi.
Lemma 3.4. Define a completely positive map ρ : B(F M,ϕ(H, F )) → B(F M,ϕ(H, F )) by the
formula ρ(T ) =Pi R(ξi)T R(ξi)∗. This operator satisfies
ρl(L(T )) = L(Im,l(T ⊗ idl)I ∗
n,l) = L(T )qn+l,
for all bounded right-M modular operators T : H (n)
subunital completely positive map, because 1 = L(id0) and q1 = L(id1) is a projection.
F → H (m)
and l ≥ 0. In particular, ρ is a
F
F → H (m)
Proof. We first prove lemma 3.4 with l = 1. Let T : H (n)
F
modular operator. It is clear that ρ(L(T ))η = 0 = L(Im,1(T ⊗ id)I ∗
be a bounded right-M -
n,1)η whenever η ∈ H k for
RADIAL MULTIPLIERS ON VON NEUMANN ALGEBRAS
23
some k < n + 1. Let η ∈ H n+k+1 for some k ≥ 0. We compute that
Im+k,1(idm+k ⊗r(ξi))Im,k(T ⊗ id1)I ∗
n,k(idn+k ⊗r(ξi)∗)I ∗
n+k,2η
ρ(L(T ))η =Xi
Xi
Im,k,1(T ⊗ idk ⊗r(ξi)r(ξi)∗)I ∗
n,k,1η
= Im,k,1(T ⊗ idk ⊗ id1)I ∗
k,1))I ∗
= Im,k+1(T ⊗ (Ik,1I ∗
and L(T ⊗ id)η = Im,1,k(T ⊗ id1 ⊗ idk)I ∗
1,k))I ∗
m,k+1.
and L(T )qn+1η = Im,k+1(T ⊗ idk+1)I ∗
= Im,k+1(T ⊗ (I1,kI ∗
n,1,k
n,k,1
n,k+1
n,k+1
When F is a projection such that F ⊗ id1 commutes with id1 ⊗F , then we get that D(n) :
H (n) → H (n) is simply the projection onto the orthogonal complement of the closed linear
span of the subspaces of the form H (i−1) ⊗M F (H (2)) ⊗M H (n−i−1) for i = 1, . . . , n − 1.
So H (n)
is a closed subspace of H (n). The operator In,m is simply the projection from
H (n)
n,m = idn+m, so
it follows that L(T ⊗ id)η = L(T )qn+1η. This is true for all η ∈ H k, for all k ∈ N, so
ρ(L(T )) = L(T ⊗ id) = L(T )qn+1.
onto its closed subspace H (n+m)
. Hence we see that In,mI ∗
F ⊗M H (m)
F
F
F
Now let l > 1. By induction we get that
ρl(L(T )) = L(Im,1,...,1(T ⊗ id1 ⊗ . . . ⊗ id1)In,1,...,1)
= L(Im,l(T ⊗ I1,...,1I ∗
1,...,1)I ∗
n,l.
By the same argument as above, we see that 1,...,1I ∗
1,...,1 = idl, so indeed
ρl(L(T )) = L(Im,l(T ⊗ idl)I ∗
n,l) = L(T )qn+l.
(cid:3)
Lemma 3.5. Let x, y ∈ ℓ2(N), and define an M -M -bimodular map Φx,y : T ′′
T ′′
M (H, F ) by the formula
M (H, F ) →
Φx,y(T ) =Xn≥0
(S ∗)ny +Xn≥1
M(S ∗)nxT M ∗
MSnxρn(T )M ∗
Sny.
Then this operator is completely bounded with completely bounded norm kΦx,ykcb ≤ kxk2 kyk2
that satisfies
(12)
Φx,y(L(T )) = h(S∗)mx, (S∗)nyi L(T ).
lemma 3.2, we have that Pi uiu∗
Proof. Observe that Φx,y is given in the form of lemma 3.2, and that in the notation of
i = Φy,y(1). So, once we show
(12), it follows from lemma 3.2 that Φx,y is completely bounded and its norm is bounded by
kΦx,ykcb ≤ kxk2 kyk2.
i = Φx,x(1) and Pi viv∗
24
STEVEN DEPREZ
Let T : H (n)
F → H (m)
F
be a bounded right-M -linear operator. Then we see that
Φx,y(L(T )) =Xk≥0
(S ∗)k y +Xk≥1
M(S ∗)k xL(T )M ∗
MSkxρk(L(T ))M ∗
Sk y
= L(T )Xk≥0
= L(T )Mf ,
MSn(S ∗)k+mxM ∗
(S ∗)k y +Xk≥1
MSn+k(S ∗)m+kSkxM ∗
Skyqn+k
where the function f is given by
f (l) = 0
x(l + m + k)y(l + n + k) +
if l < n
x(l − k + m)y(l + n − k)
if l ≥ 0
lXk=1
f (n + l) =Xk≥0
=Xk≥l
x(k + m)y(k + n) +
x(k + m)y(k + n)
l−1Xk=0
= h(S∗)ny, (S∗)mxi .
Since L(T ) = L(T )qn, we see that indeed
Φx,y(L(T )) = L(T ) h(S∗)ny, (S∗)mxi .
(cid:3)
proof of theorem 3.1. It is well-known that every trace class operator T ∈ B(ℓ2 N) can be
written as a sum of rank one operators T =Pn xn ⊗ yn with xn, yn ∈ ℓ2(N) and where the
sum converges in trace norm. Moreover, the trace norm of T isPn kxnk2 kynk2. In particular,
for the Hankel matrix Hψ, we find sequences of vectors xn, yn ∈ ℓ2(N) such that
ψ(k + l) − ψ(k + l + 2) = Hψ(k, l) =Xn
xn(k)yn(l) for all k, l ∈ N,
and such that kHψk1 =Pn kxnk2 kynk2. Moreover, we see that
(ψ(k + 2m) − ψ(k + 2m + 2)).
So we also see that
ψ(k) = c+ + (−1)kc− +Xm
ψ(k + l) = c+ + (−1)k+lc− +Xm,n
xn(k + m)yn(l + m)
= c+ + (−1)k+lc− +Xn D(S∗)lyn, (S∗)kxE
RADIAL MULTIPLIERS ON VON NEUMANN ALGEBRAS
25
Now, we set Φψ = c+ + c− Adu +Pk Φxk,yk , where u = Mz with z(k) = (−1)k as before. Let
be a bounded right-M -linear operator. Then it follows from lemma 3.5 that
F
T : H (n)
F → H (m)
Φψ(L(T )) = c+ + c−L(T )MSm(S ∗)nzMz +Xk
= c+ + c−(−1)n+m +Xn
= ψ(n + m)L(T )
Φxk,yk (L(T ))
h(S∗)my, (S∗)nxi! L(T )
So Φψ is indeed the map we were searching for. Moreover Φψ is completely bounded and its
completely bounded norm is bounded by
kΦψkcb ≤ c+ + c− +Xk
kxkk2 kykk2 = kψkC .
(cid:3)
4. Amalgamated free product von Neumann algebras
In this section we deduce the following theorem from theorem 3.1.
Theorem 4.1. Let (Mi)i be a (finite or countably infinite) family of tracial von Neumann
algebras with a common subalgebra P . Denote by M the amalgamated free product over
P of the von Neumann algebras Mi. Let ψ : N → C be a function in class C′ defined in
the introduction. Then there is a unique ultraweakly continuous completely bounded map
Ψψ : M → M that satisfies
Ψψ(x1 . . . xn) = ψ(n)x1 . . . xn for every reduced word x1 . . . xn ∈ M.
Moreover, the completely bounded norm of Ψψ is bounded above by kΨψkcb ≤ kψkC ′.
Proof. Consider the graph of von Neumann algebras that is depicted below
M (1)
M (2)
P
P
P
P
M (i)
26
STEVEN DEPREZ
As in the introduction, we can consider the fundamental von Neumann algebra fM of this
graph. In other words, we set fM = Γ′′
(H, F, J) where
eP
Mi
eP = P ⊕Mi
H =Mi
J(x) =cx∗ ∈ P L2(Mi)Mi
Mi
L2(Mi)P ⊕ P L2(Mi)Mi
L =Mi
⊂Mi
P P P ⊕Mi
P ⊕Mi,j
P M (i)
= H (2).
for all x ∈ Mi
L2(Mi)P and vice-versa.
and where F is the projection onto the eP -eP Hilbert subbimodule
M (j)M (i) ⊗P M (i)
M (i)
M (i)M (i) ⊗P M (j)
M (j)
Denote by p the central projection in eP ⊂ fM that corresponds to the term P in the direct
sum eP = P ⊕Li M (i). Then it follows that
(13) p F eP (H, F )p ∼= P L2(P )P ⊕Mi
P L2(Mi ⊖P )P ⊕Mi6=j
P L2(Mi ⊖P )⊗P L2(Mj ⊖P )P ⊕. . . ,
a priori just as P -P bimodules. Observe that each term L2(Mi) that appears in the direct
sum above is a direct summand of H (2)
F . The P -P bimodule on the right
is precisely L2(M ). We denote by U : p F eP ,ϕ(H, F )p → L2(M ) the unitary that implements
the identification in (13). It is now easy to see that
F rather than of H (1)
for every y ∈ Mi ⊖ P .
upW (y)pu∗ = y ∈ M,
In particular, we get that α = Adu∗ : M → pfM p is an isomorphism that maps a reduced word
and ψ(2n + 1) = 0 for all n ∈ N. Then we know that ψ ∈ C and its norm is(cid:13)(cid:13)(cid:13) ψ(cid:13)(cid:13)(cid:13)C
x1 . . . xn of length n to the reduced word pW (x1⊗. . .⊗xn)p that has length 2n. Let ψ : N → C
be a function in class C′. We consider the function ψ : N → C that is defined by ψ(2n) = ψ(n)
= kψkC ′.
M (H, F )) that
By theorem 3.1, we find an ultraweakly continuous map Φ ψ : T ′′
satisfies
M (H, F ) → T ′′
Φ ψ(L(T )) = ψ(n + m)L(T )
for all bounded right-M modular operators T : H (n)
F → H (m)
F . Moreover, the completely
bounded norm of Φ ψ is bounded by(cid:13)(cid:13)(cid:13)Φ ψ(cid:13)(cid:13)(cid:13)cb
≤ kψkC.
RADIAL MULTIPLIERS ON VON NEUMANN ALGEBRAS
27
In particular, we get that Φ ψ(W (ξ)) = ψ(n)W (ξ) for all bounded vectors ξ ∈ H (n)
F . Moreover,
Φ ψ is M -M -bimodular, and in particular, Φ ψ(pfM p) ⊂ pfM p. We define Φψ : M → M by the
formula Φψ(x) = α−1(Φ ψ(α(x))). We conclude that
Φψ(x1 . . . xn) = ψ(2n)x1 . . . xn = ψ(n)x1 . . . xn
for all reduced words x1 . . . xn in the amalgamated free product decomposition M = M1 ∗P M2 ∗P . . ..
Moreover, the completely bounded norm of Φψ is bounded by
kΦψkcb ≤(cid:13)(cid:13)(cid:13)Φ ψ(cid:13)(cid:13)(cid:13)cb
≤(cid:13)(cid:13)(cid:13) ψ(cid:13)(cid:13)(cid:13)C
= kψkC ′ .
(cid:3)
So Φψ satisfies the conditions of theorem 3.1
References
[BS1] M. Bozejko and R. Speicher. An example of a generalized Brownian motion. Comm. Math. Phys.,
137(3):519 -- 531, 1991.
[BS2] M. Bozejko and R. Speicher. Completely positive maps on Coxeter groups, deformed commutation
relations, and operator spaces. Math. Ann., 300(1):97 -- 120, 1994.
[CS] E. Christensen and A. M. Sinclair. A survey of completely bounded operators. Bull. London Math. Soc.,
21(5):417 -- 448, 1989.
[HM] U. Haagerup and S. Moller. Radial multipliers on reduced free products of operator algebras. J. Funct.
Anal., 263(8):2507 -- 2528, 2012.
[HSS] U. Haagerup, T. Steenstrup, and R. Szwarc. Schur multipliers and spherical functions on homogeneous
[M]
[S]
[V]
[W]
trees. Internat. J. Math., 21(10):1337 -- 1382, 2010.
S. Moller. Radial multipliers on amalgamated free products of II1 factors. preprint. ArXiV:1306.5540.
D. Shlyakhtenko. A-valued semicircular systems. J. Funct. Anal., 166(1):1 -- 47, 1999.
D. Voiculescu. Symmetries of some reduced free product C ∗-algebras. In Operator algebras and their
connections with topology and ergodic theory (Bu¸steni, 1983), volume 1132 of Lecture Notes in Math.,
pages 556 -- 588. Springer, Berlin, 1985.
J. Wysocza´nski. A characterization of radial Herz-Schur multipliers on free products of discrete groups.
J. Funct. Anal., 129(2):268 -- 292, 1995.
|
1306.3411 | 3 | 1306 | 2015-03-26T13:01:19 | Bilinear Ideals in Operator Spaces | [
"math.OA",
"math.FA"
] | We introduce a concept of bilinear ideal of jointly completely bounded mappings between operator spaces. In particular, we study the bilinear ideals $\mathcal{N}$ of completely nuclear, $\mathcal{I }$ of completely integral, $\mathcal{E}$ of completely extendible bilinear mappings, $\mathcal{MB}$ multiplicatively bounded and its symmetrization $\mathcal{SMB}$. We prove some basic properties of them, one of which is the fact that $\mathcal{I}$ is naturally identified with the ideal of (linear) completely integral mappings on the injective operator space tensor product. | math.OA | math |
BILINEAR IDEALS IN OPERATOR SPACES
VERÓNICA DIMANT AND MAITE FERNÁNDEZ-UNZUETA
Abstract. We introduce a concept of bilinear ideal of jointly completely bounded mappings
between operator spaces. In particular, we study the bilinear ideals N of completely nuclear,
I of completely integral, E of completely extendible bilinear mappings, MB multiplicatively
bounded and its symmetrization SMB. We prove some basic properties of them, one of which
is the fact that I is naturally identified with the ideal of (linear) completely integral mappings
on the injective operator space tensor product.
1. Introduction and Preliminaries
Let V, W and X be operator spaces. If we consider the underlying vector space structure, the
relations
(1)
Bil(V × W, X)
ν
≃ L(V ⊗ W, X)
ρ
≃ L(V, L(W, X))
hold through the two natural linear isomorphisms ν, ρ. In order for ν and ρ to induce natural
morphisms in the operator space category, it is necessary to have appropriately defined an oper-
ator space tensor norm on V ⊗ W and specific classes of linear and bilinear mappings. This is the
case, for instance, of the so called projective operator space tensor norm k · k∧, the completely
bounded maps and the jointly completely bounded bilinear mappings, where ν and ρ induce the
following completely bounded isometric isomorphisms:
J CB(V × W, X) ≃ CB(Vb⊗W, X) ≃ CB(V, CB(W, X)).
There are many possible ways to provide V ⊗ W with an operator space tensor norm and, of
course, to define classes of mappings. Several authors, inspired by the success that the study of
the relations between tensor products and mappings has had in the Banach space setting, have
systematically study some analogous relations for operator spaces. This is the case, for instance,
of the completely nuclear and completely integral linear mappings (see [7, Section III]).
In this paper we follow this approach as well, but with the attention focused on the relations
involving ν, the isomorphism in (1) which concerns bilinear mappings. In Section 2 we introduce
the notion of an ideal of completely bounded bilinear mappings and study its general properties.
In Section 3 we define the ideals of completely nuclear and completely integral bilinear mappings.
The main result proved here is that the ideal of completely integral bilinear mappings is naturally
identified with the ideal of completely integral linear mappings on the injective operator space
tensor product, that is I(V × W, X) ∼= LI(V
∨
⊗ W, X) (see Theorem 3.8). This implies that,
2010 Mathematics Subject Classification. 47L25,47L22, 46M05.
Key words and phrases. Operator spaces, Bilinear mappings, Bilinear ideals.
The first author was partially supported by CONICET PIP 0624 and PAI-UDESA 2011. The second author
was partially supported by CONACYT 182296.
1
2
VERÓNICA DIMANT AND MAITE FERNÁNDEZ-UNZUETA
contrary to the result for Banach spaces, the relation I(V × W ) ∼= LI(V, W ∗) does not always
hold. Indeed, it holds if and only if W is locally reflexive.
η
(cid:18)V
⊗ W(cid:19)∗
The ideal E of bilinear completely extendible mappings is introduced in Section 4. We prove in
Proposition 4.4 that E gives rise, through duality, to an operator space tensor product η such that
∼= E(V ×W ). In Section 5 we consider the ideal SMB of symmetrized multiplicatively
bounded mappings, which is the symmetrization of the ideal MB of multiplicatively bounded
mappings. The following theorem summarizes the inclusion relations among all these bilinear
ideals:
Theorem 1.1. Let V, W and X be operator spaces. Then, we have the following complete
contractive inclusions:
(a) N (V × W, X) ⊂ I(V × W, X) ⊂ MB(V × W, X) ⊂ SMB(V × W, X) ⊂ J CB(V × W, X).
(b) I(V × W, X) ⊂ E(V × W, X) ⊂ J CB(V × W, X).
(c) MB(V × W, L(H)) ⊂ SMB(V × W, L(H)) ⊂ E(V × W, L(H))) ⊂ J CB(V × W, L(H)).
In Section 6 we prove the inclusions and provide examples to distinguish the ideals.
We now recall some basic concepts about operator spaces, mainly with respect to bilinear
operators and tensor products. For a more complete presentation of these topics, see [2, 7, 14].
All vector spaces considered are over the complex numbers. For a linear space V , we let Mn×m(V )
denote the set of all the n × m matrices of elements in V . In the case n = m, the notation is
simplified to set Mn×n(V ) = Mn(V ).
If V is the scalar field we just write Mn×m and Mn,
respectively. For α ∈ Mn×m, its norm kαk will be considered as an operator from ℓm
2 to ℓn
2 .
Given v = (vi,j) ∈ Mn(V ) and w = (wk,l) ∈ Mm(V ), v ⊕ w ∈ Mn+m(V ) stands for the matrix
v ⊕ w = (vi,j)
0
0
(wk,l) ! .
A matrix norm k · k on a linear space V is an assignment of a norm k · kn on Mn(V ), for each
n ∈ N. A linear space V is an operator space if it is endowed with a matrix norm satisfying:
M1 kv ⊕ wkn+m = max{kvkn, kwkm}, for all v ∈ Mn(V ) and w ∈ Mm(V ).
M2 kαvβkm ≤ kαk · kvkn · kβk, for all v ∈ Mn(V ), α ∈ Mm×n and β ∈ Mn×m.
We usually omit the subindex n in the matrix norms and simply denote k · k instead of k · kn. The
inclusion Mn×m(V ) ֒→ Mmax{n,m}(V ) naturally endows the rectangular matrices with a norm.
Throughout the article, V , W , X, Y , Z, U1, U2 will denote operator spaces where the underlying
normed space is complete (i.e. it is a Banach space).
Every linear mapping ϕ : V → W induces, for each n ∈ N, a linear mapping ϕn : Mn(V ) →
Mn(W ) given by
ϕn(v) = (ϕ(vi,j)) , for all v = (vi,j) ∈ Mn(V ).
It holds that kϕk = kϕ1k ≤ kϕ2k ≤ kϕ3k ≤ .... The completely bounded norm of ϕ is
defined by
kϕkcb = sup
n∈N
kϕnk.
BILINEAR IDEALS IN OPERATOR SPACES
3
We say that ϕ is completely bounded if kϕkcb is finite, that ϕ is completely contractive
if kϕkcb ≤ 1 and that ϕ is a complete isometry if each ϕn : Mn(V ) → Mn(W ) is an isometry.
It is easy to see that k·kcb defines a norm on the space CB(V, W ) of all completely bounded linear
mappings from V to W . The natural identification Mn (CB(V, W )) ∼= CB (V, Mn(W )) provides
CB(V, W ) with the structure of an operator space. Also, since V ∗ = CB(V, C), the dual of an
operator space is again an operator space.
In contrast to the linear case, a bilinear mapping φ : V × W → X naturally induces not one,
but two different bilinear mappings in the matrix levels. Some authors (see, for instance [7, 18])
use the name "complete boundedness" for the first notion and "multiplicative boundedness" or
"matrix complete boundedness" for the second one, while others [2, 3, 20] use the name "jointly
complete boundedness" for the first concept and "complete boundedness" for the second one. In
order to avoid confusion, we will not use the name "complete boundedness" for bilinear mappings.
So, given a bilinear mapping φ : V × W → X, consider the associated bilinear mapping
φn : Mn(V ) × Mn(W ) → Mn2(X) defined, for each n ∈ N, as follows:
φn(v, w) = (φ(vi,j, wk,l)) , for all v = (vi,j) ∈ Mn(V ), w = (wk,l) ∈ Mn(W ).
When their norms are uniformly bounded, that is, when
kφkjcb ≡ sup
n∈N
kφnk < ∞,
we say that φ is jointly completely bounded. It is plain to see that k · kjcb is a norm on the
space J CB(V × W, X) of all jointly completely bounded bilinear mappings from V × W to X.
As in the linear setting, the identification
Mn (J CB(V × W, X)) ∼= J CB (V × W, Mn(X)) .
provides J CB(V × W, X) with an operator space structure.
The second way to naturally associate φ with a bilinear mapping φ(n) : Mn(V ) × Mn(W ) →
Mn(X), for each n ∈ N, involves the matrix product and it is given by
φ(n)(v, w) = nXk=1
φ(vi,k, wk,l)! , for all v = (vi,j) ∈ Mn(V ), w = (wk,l) ∈ Mn(W ).
We say that φ is multiplicatively bounded if
kφkmb = sup
n∈N
kφ(n)k < ∞.
Again, it is easily seen that k·kmb is a norm on the space MB(V ×W, X) of all multiplicatively
bounded bilinear mappings from V × W to X. The identification
Mn (MB(V × W, X)) ∼= MB (V × W, Mn(X))
endows MB(V × W, X) with matrix norms that give the structure of an operator space.
We finish this section recalling three basic examples from the theory of tensor products of
operator spaces (the general notion is in Definition 2.3): the operator space projective tensor
norm, the operator space injective tensor norm and the operator space Haagerup tensor norm.
4
VERÓNICA DIMANT AND MAITE FERNÁNDEZ-UNZUETA
Consider two operator spaces V and W . The definition of the first norm uses the fact that
each element u ∈ Mn(V ⊗ W ) can be written as:
(2)
u = α(v ⊗ w)β
with v ∈ Mp(V ), w ∈ Mq(W ), α ∈ Mn×p·q, β ∈ Mp·q×n, for certain p, q ∈ N, where v ⊗ w is the
p · q × p · q-matrix given by
(3)
v ⊗ w =
v1,1 ⊗ w1,1
...
v1,1 ⊗ wq,1
· · ·
· · ·
vp,1 ⊗ w1,1
...
vp,1 ⊗ wq,1
· · ·
...
· · ·
· · ·
· · ·
· · ·
...
· · ·
v1,1 ⊗ w1,q
...
v1,1 ⊗ wq,q
· · ·
· · ·
vp,1 ⊗ w1,q
...
vp,1 ⊗ wq,q
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
v1,p ⊗ w1,1
...
v1,p ⊗ wq,1
· · ·
· · ·
vp,p ⊗ w1,1
...
vp,p ⊗ wq,1
· · ·
...
· · ·
· · ·
· · ·
· · ·
...
· · ·
v1,p ⊗ w1,q
...
v1,p ⊗ wq,q
· · ·
· · ·
vp,p ⊗ w1,q
...
vp,p ⊗ wq,q
The Haagerup tensor norm is defined as:
kukh = inf {kvk · kwk : u = v ⊙ w, v ∈ Mn×r(V ), w ∈ Mr×n(W ), r ∈ N} ,
while the Haagerup tensor product V
h
⊗ W is the completion of (V ⊗ W, k · kh).
For any operator spaces V and W , k · k∨ and k · k∧ are, respectively, the smallest and the
largest operator space cross norms on V ⊗ W . In particular, for each u ∈ Mn(V ⊗ W ) it holds
that
kuk∨ ≤ kukh ≤ kuk∧.
The operator space projective tensor norm of u ∈ Mn(V ⊗ W ) is defined as
kuk∧ = inf{kαk · kvk · kwk · kβk : all representations of u as in (2)}.
The operator space injective tensor norm of u ∈ Mn(V ⊗ W ) is defined as
kuk∨ = sup {k(f ⊗ g)n(u)k : f ∈ Mp(V ∗), g ∈ Mq(W ∗), kf k ≤ 1, kgk ≤ 1} .
The operator space projective tensor product Vb⊗W and the operator space in-
∨
⊗ W are the completion of (V ⊗ W, k · k∧) and the completion
jective tensor product V
of (V ⊗ W, k · k∨), respectively.
There is a natural completely isometric identification:
So, in particular:
J CB(V × W, X) ∼= CB(Vb⊗W, X) ∼= CB(V, CB(W, X)).
J CB(V × W ) ∼= (Vb⊗W )∗ ∼= CB(V, W ∗).
∨
⊗ W )∗ with a subset of bilinear mappings is done later, in Proposition
Every u ∈ Mn(V ⊗ W ) can be written as u = v ⊙ w, for certain matrices v ∈ Mn×r(V ) and
The identification of (V
3.11.
w ∈ Mr×n(W ), where
v ⊙ w = rXk=1
vi,k ⊗ wk,j! .
BILINEAR IDEALS IN OPERATOR SPACES
5
The Haagerup tensor product is naturally associated with multiplicatively bounded bilinear
operators through the following identifcations:
MB(V × W, X) ∼= CB(V
h
⊗ W, X)
and MB(V × W ) ∼= (V
h
⊗ W )∗.
Remark 1.2. We will use repeatedly along the text the following extension property for completely
bounded linear mappings (see [7, Theorem 4.1.5]): if V is a subspace of an operator space W and
H is a Hilbert space, then every completely bounded linear map ϕ : V → L(H) has a completely
bounded extension ϕ : W → L(H) with kϕkcb = kϕkcb.
Equivalently, this can be stated as in [14, Theorem 1.6]: if V , W are operator spaces, H, K
are Hilbert spaces such that V is a subspace of L(H) and W is a subspace of L(K), then every
completely bounded linear map ϕ : V → W has a completely bounded extension ϕ : L(H) →
L(K) with kϕkcb = kϕkcb.
2. Bilinear ideals
The linear structure and the closedness by compositions are the basic properties required of a
subset of maps, in order to have a suitable relation between mappings spaces and tensor products.
These will be, precisely, the defining properties of a bilinear ideal (see Definition 2.2). To deal
with compositions, we need first to prove the following estimate:
Lemma 2.1. Let φ ∈ Mn (J CB(V × W, X)), r1 ∈ CB(U1, V ), r2 ∈ CB(U2, W ), s ∈ CB(X, Y ).
Then sn ◦ φ ◦ (r1, r2) is jointly completely bounded and
ksn ◦ φ ◦ (r1, r2)kjcb ≤ kskcb · kφkjcb · kr1kcb · kr2kcb.
Proof. Let ψ = sn ◦ φ ◦ (r1, r2). It is easy to see that
ψm = sn·m2 ◦ φm ◦ ((r1)m, (r2)m).
Thus, for every m,
kψmk ≤ ksn·m2k · kφmk · k(r1)mk · k(r2)mk ≤ kskcb · kφkjcb · kr1kcb · kr2kcb,
and the conclusion follows.
(cid:3)
In accordance with the definition of an operator space ideal of linear mappings (see [6] and
[7]), we introduce:
Definition 2.2. An operator space bilinear ideal A is an assignment, to each group of three
operator spaces V , W and X, of a linear subspace A(V × W, X) of J CB(V × W, X) containing
all finite type continuous bilinear maps, together with an operator space matrix norm k · kA such
that:
(a) For all φ ∈ Mn(A(V × W, X)), kφkjcb ≤ kφkA.
(b) For all φ ∈ Mn(A(V × W, X)), r1 ∈ CB(U1, V ), r2 ∈ CB(U2, V ), s ∈ CB(X, Y ), the
matrix sn ◦ φ ◦ (r1, r2) belongs to Mn(A(U1 × U2, Y )) and
ksn ◦ φ ◦ (r1, r2)kA ≤ kskcb · kφkA · kr1kcb · kr2kcb.
We now introduce the notion of tensor norm for operator spaces.
6
VERÓNICA DIMANT AND MAITE FERNÁNDEZ-UNZUETA
Definition 2.3. We say that α is an operator space tensor norm if α is an operator space
matrix norm on each tensor product of operator spaces V ⊗ W that satisfies the following two
conditions:
(a) α is a cross matrix norm, that is, α(v ⊗ w) = kvk · kwk, for all v ∈ Mp(V ), w ∈ Mq(W ),
p, q ∈ N.
(b) α fullfils the "completely metric mapping property":
for every r1 ∈ CB(U1, V ), r2 ∈
CB(U2, W ), the operator r1 ⊗ r2 : (U1 ⊗ U2, α) → (V ⊗ W, α) is completely bounded and
kr1 ⊗ r2kcb ≤ kr1kcb · kr2kcb.
α
⊗ W the completion of (V ⊗ W, α).
We denote by V
This notion is, in principle, less restrictive than the one introduced in [3, Definition 5.9], which
the authors called "uniform operator space tensor norm". Whenever the linear isomorphism
determined by (3) (the so called algebraic shuffle isomorphism) Mp(V ) ⊗ Mq(W ) → Mpq(V ⊗ W )
extends to a complete contraction Mp(V )⊗α Mq(W ) → Mpq(V ⊗α W ), both notions coincide [20].
That is the case of the three tensor norms defined above (projective, injective and Haagerup).
The proof that these main examples satisfy the definition, as well as the fact that the projective
tensor norm k · k∧ is the largest operator space tensor norm, can be found in [7].
Every operator space tensor norm determines, through ν in (1), an operator space bilinear
ideal according to the following identification: Given V , W , X operator spaces, let
Aα(V × W, X) ∼= CB(V
α
⊗ W, X).
Proposition 2.4. Let α be an operator space tensor norm. Then Aα is an operator space bilinear
ideal.
α
CB(V
Proof. From the relation CB(V
subspace of J CB(V × W, X). Also, it is clear that all finite type continuous bilinear mappings
belong to Aα(V × W, X).
⊗ W, X) ⊂ CB(Vb⊗W, X), it follows that Aα(V × W, X) is a
⊗ W, X)(cid:17) ∼=
α
α
α
⊗ W ) → Mm(Mn(X)) has norm
On the other hand, φ also belongs to Mn(J CB(V × W, X)) and it has an associated matrix of
keφmk = supneφm(u) : u ∈ Mm(V ⊗ W ), α(u) ≤ 1o .
(a) Let φ ∈ Mn(Aα(V × W, X)) then its linear associated eφ belongs to Mn(cid:16)CB(V
⊗ W, Mn(X)). This says that kφkAα = keφkcb = supm keφmk.
The mapping eφm : Mm(V
linear mappings φ ∈ Mn(cid:0)CB(Vb⊗W, X)(cid:1) ∼= CB(Vb⊗W, Mn(X)). This implies that
and the mapping φm : Mm(Vb⊗W ) → Mm(Mn(X)) has norm
For each u ∈ Mm(V ⊗ W ), eφm(u) = φm(u) and α(u) ≤ kuk∧. Then, for every m, kφmk ≤
keφmk, and thus kφkjcb ≤ kφkAα.
kφmk = sup(cid:8)φm(u) : u ∈ Mm(V ⊗ W ), kuk∧ ≤ 1(cid:9) .
kφkjcb = kφkcb = sup
m
kφmk,
BILINEAR IDEALS IN OPERATOR SPACES
7
(b) For φ ∈ Mn(Aα(V ×W, X)), leteφ ∈ Mn(cid:16)CB(V
mappings. For any r1 ∈ CB(U1, V ), r2 ∈ CB(U2, W ) and s ∈ CB(X, Y ), the following equality
holds.
α
⊗ W, X)(cid:17) be its associated matrix of linear
A direct computation gives the required inequality.
ksn ◦ φ ◦ (r1, r2)kAα = ksn ◦eφ ◦ (r1 ⊗ r2)kcb
(cid:3)
Example 2.5. Since MB(V × W, X) ∼= CB(V
is an operator space bilinear ideal.
h
⊗ W, X), from Proposition 2.4 we obtain that MB
With similar arguments to those used to prove Proposition 2.4, we obtain:
Proposition 2.6. Let α be an operator space tensor norm and B be an operator space ideal of
linear mappings. Given the operator spaces V , W and X, let AB
α (V × W, X) be the operator
space determined by the identification
(4)
α (V × W, X) ∼= B(V
AB
α
⊗ W, X).
Then, AB
α is an operator space bilinear ideal.
3. Completely nuclear and completely integral bilinear mappings
In [7, Sections 12.2 and 12.3] the definitions of completely nuclear and completely integral
linear mappings are presented. We now introduce and study the analogous bilinear concepts.
We will see that they define operator space bilinear ideals. Theorem 3.8 provides a concrete
identification of the integral bilinear ideal as in (4). On the contrary, from Proposition 3.12, it
will follow that the nuclear bilinear ideal can not be described in such a way.
In order to properly define the notion of nuclearity in the context of bilinear mappings on
operator spaces, we need to state first some natural mappings. Let
Θ : (V ∗ ∨
⊗ W ∗)
∨
⊗ X ֒→ J CB(V × W, X)
∨
be the natural complete isometry obtained as a composition of the natural complete isometries
V ∗
and (V ∗
∨
⊗ X (see [7, Proposition 8.1.2 and Proposition 8.1.5]). Let
⊗ X ֒→ CB(Vb⊗W, X) ∼= J CB(V × W, X)
⊗ W ∗ ֒→ (Vb⊗W )∗,
∨
⊗ W ∗)
∨
∨
(Vb⊗W )∗
⊗ X ֒→ (Vb⊗W )∗
Φ : (V ∗b⊗W ∗)b⊗X → (V ∗ ∨
⊗ W ∗)
∨
⊗ X
be the canonical complete contraction and let
With such a Ψ:
Ψ = Θ ◦ Φ : (V ∗b⊗W ∗)b⊗X → J CB(V × W, X).
Definition 3.1. A bilinear mapping φ ∈ J CB(V × W, X) is completely nuclear if it belongs to
the image of Ψ. The operator space structure in the set of completely nuclear bilinear mappings
N (V × W, X), is given by the identification of the image of Ψ with the quotient of its domain by
its kernel. That is,
N (V × W, X) ∼= (V ∗b⊗W ∗)b⊗X/ ker Ψ.
8
VERÓNICA DIMANT AND MAITE FERNÁNDEZ-UNZUETA
Proposition 3.2. N is an operator space bilinear ideal.
Proof. By definition N (V × W, X) is a linear subspace of J CB(V × W, X) and the contention of
finite type elements is plain. The injective mapping N (V × W, X) → J CB(V × W, X) induced
on the quotient by the complete contraction Ψ, has norm less or equal than Ψ, and so, it is again
a complete contraction. Hence, kφkjcb ≤ kφkN and (a) is proved.
(b) Let Ψ denote the quotient map induced by Ψ. Given φ ∈ Mn (N (V × W, X)), r1 ∈
CB(U1, V ), r2 ∈ CB(U2, W ) and s ∈ CB(X, Y ), consider the following diagram:
((r∗
1 ⊗r∗
Mn(cid:0)(V ∗b⊗W ∗)b⊗X(cid:1)
2 )b⊗Y(cid:1)
Mn(cid:0)(U ∗
1b⊗U ∗
2 )⊗s)n
Ψn
Ψn
Mn (N (V × W, X))
/ Mn (N (U1 × U2, Y )) ,
where the right vertical arrow is the mapping φ 7→ sn ◦ φ ◦ (r1, r2). It is immediate to check that
the mappings are well defined and that the diagram commutes. In particular, sn ◦ φ ◦ (r1, r2)
belongs to Mn (N (U1 × U2, Y )). If u ∈ Mn(cid:0)(V ∗b⊗W ∗)b⊗X(cid:1) is such that Ψn(u) = φ it holds
sn ◦ φ ◦ (r1, r2) = sn ◦ Ψn(u) ◦ (r1, r2) = Ψn (((r∗
2) ⊗ s)n (u)) .
1 ⊗ r∗
The estimate we are looking for follows from the fact that the inequality
ksn ◦ φ ◦ (r1, r2)kN ≤ k ((r∗
1 ⊗ r∗
2) ⊗ s)n (u)kMn((U ∗
1 b⊗U ∗
2 )b⊗Y )
(cid:3)
holds for every u such that Ψn(u) = φ.
Definition 3.3. We say that a bilinear mapping φ ∈ J CB(V × W, X) is completely integral if
kφkI = sup {kφF1×F2kN : F1 ⊂ V, F2 ⊂ W of finite dimension} < ∞.
Let I(V × W, X) be the space of all completely integral bilinear mappings from V × W to X.
We consider in I(V × W, X) the matrix norm given by
kφkI = sup {kφF1×F2kN : F1 ⊂ V, F2 ⊂ W of finite dimension} ,
for every φ ∈ Mn (I(V × W, X)). It is easy to see that this norm endowed I(V × W, X) with
the structure of an operator space.
Proposition 3.4. Let V, W , X be operator spaces and let φ ∈ Mn (N (V × W, X)). Then
The first inequality also holds for φ ∈ Mn (I(V × W, X)).
kφkjcb ≤ kφkI ≤ kφkN .
Proof. For φ ∈ Mn (I(V × W, X)), consider finite dimensional spaces F1 ⊂ V and F2 ⊂ W .
Since kφF1×F2kjcb ≤ kφF1×F2kN and
kφkjcb = sup {kφF1×F2kjcb : F1 ⊂ V, F2 ⊂ W of finite dimension}
we obtain that
kφkjcb ≤ kφkI .
/
/
/
BILINEAR IDEALS IN OPERATOR SPACES
9
Now, if φ ∈ Mn (N (V × W, X)) and we denote by j1 : F1 ֒→ V and j2 : F2 ֒→ W the canonical
(completely contractive) embeddings, it is clear that
kφF1×F2kN = kφ ◦ (j1, j2)kN ≤ kφkN · kj1kcb · kj2kcb = kφkN .
(cid:3)
Proposition 3.5. I is an operator space bilinear ideal.
Proof. By definition I(V × W, X) is a linear subspace of J CB(V × W, X). Finite type continuous
bilinear maps are obviously contained in I(V × W, X). Condition (a) was already proved above.
(b) Let φ ∈ Mn (I(V × W, X)), r1 ∈ CB(U1, V ), r2 ∈ CB(U2, V ) and s ∈ CB(X, Y ). For finite
dimensional spaces F1 ⊂ U1 and F2 ⊂ U2 let j1 : F1 ֒→ U1 and j2 : F2 ֒→ U2 be the canonical
(completely contractive) embeddings. We have
ksn ◦ φ ◦ (r1, r2)F1×F2kN = ksn ◦ φ ◦ (r1j1, r2j2)kN ≤ kskcb · kφkI · kr1kcb · kr2kcb.
(cid:3)
A pointwise limit of completely nuclear bilinear contractions is not necessarily completely
nuclear, but it is always integral. This result is in the following two lemmas and will be used
several times. The statements given here are simpler than their linear analogues given in [7,
Lemma 12.2.7 and Lemma 12.3.1].
Lemma 3.6. Let (φλ) and φ in Mn (N (F1 × F2, Mm)), where F1 and F2 are finite dimensional
operator spaces. Suppose that there exists a constant C such that kφλkMn(N (F1×F2,Mm)) ≤ C for
all λ and that φλ(x, y) → φ(x, y) for every (x, y) ∈ F1 × F2. Then, kφkMn(N (F1×F2,Mm)) ≤ C.
Proof. Take {x1, . . . , xk} and {y1, . . . , yl} vector bases of F1 and F2, respectively, and denote by
{x∗
l } the corresponding dual bases. Since
1, . . . , x∗
k} and {y∗
1, . . . , y∗
kφλ − φkMn(N (F1×F2,Mm)) ≤ Xi,j
≤ Xi,j
Hence, the result follows.
kφλ(xi, yj) − φ(xi, yj)kMn·m · kx∗
i ⊗ y∗
j kN (F1×F2)
kφλ(xi, yj) − φ(xi, yj)kMn·m · kx∗
i k · ky∗
j k → 0.
(cid:3)
Lemma 3.7. Suppose that φ ∈ Mn (J CB(V × W, Mm)) and that there exists a net (φλ) ⊂
Mn (N (V × W, Mm)) with
kφλkMn(N (V ×W,Mm)) ≤ C, for all λ
and
φλ(v, w) → φ(v, w), for all v ∈ V, w ∈ W.
Then, φ belongs to Mn (I(V × W, Mm)) and kφkMn(I(V ×W,Mm)) ≤ C.
Proof. For a given pair of finite dimensional subspaces F1 ⊂ V and F2 ⊂ W , the net (φλF1×F2)
and the map φF1×F2 satisfy the hypothesis of the previous lemma. Thus, kφF1×F2kMn(N (V ×W,Mm)) ≤
C. This implies that φ is completely integral and kφkMn(I(V ×W,Mm)) ≤ C.
(cid:3)
φλ =Xi,j
we have
φλ(xi, yj) x∗
i ⊗ y∗
j
and
φ(xi, yj) x∗
i ⊗ y∗
j
φ =Xi,j
10
VERÓNICA DIMANT AND MAITE FERNÁNDEZ-UNZUETA
For the classes of completely nuclear and completely integral mappings, it is necessary to recall
the linear definitions in order to make precise the relationship between bilinear mappings on
operator spaces and linear mappings on operator space tensor products. A complete exposition
of this topic is provided in [7, Chapter 12]. A linear mapping ϕ : V → W is said to be
completely nuclear, ϕ ∈ LN (V, W ), if it belongs to the image of the canonical completely
contractive mapping
The operator space structure of LN (V, W ) is given by the identification
⊗ W ֒→ CB(V, W ).
LΨ : V ∗b⊗W → V ∗ ∨
LN (V, W ) ∼= V ∗b⊗W/ ker LΨ.
A linear mapping ϕ : V → W is said to be completely integral, ϕ ∈ LI(V, W ), if the
completely nuclear norms of all its restrictions to finite dimensional subspaces of V are bounded.
The operator space matrix norm on LI(V, W ) is given by
kϕkMn(LI (V,W )) = sup {kϕF kLN : F ⊂ V of finite dimension} ,
for each ϕ ∈ Mn (LI(V, W )).
So, the relation we were seeking states the following:
Theorem 3.8. For every three operator spaces V, W and X, there is a complete isometry
I(V × W, X) ∼= LI(V
∨
⊗ W, X).
An analogous relation in the Banach space setting holds, and it is crucial in the study of
the bilinear integral mappings (see [19]). The proof for operator spaces is, however, quite more
involved.
We prove first the particular case of Theorem 3.8 when X is the finite dimensional operator
space of n × n-matrices Mn. The operator space dual/pre-dual of Mn is the space Tn of n × n-
matrices where the norm is given by
kαkTn = trace(α).
Remark 3.9. A version of "Goldstine's theorem" holds in operator spaces: If u ∈ Mn(V ∗∗) with
kuk ≤ 1, then there exists a net (uλ) ∈ Mn(V ) such that kuλk ≤ 1, for all λ and ϕn(uλ) → u(ϕ),
for all ϕ ∈ V ∗ (see [7, Proposition 4.2.5]).
Proposition 3.10. There is a complete isometry I(V × W, Mn) ∼= LI(V
∨
⊗ W, Mn).
Proof. Since Mn = T ∗
that there is a completely isometric identity
n is a finite-dimensional operator space, from [7, Corollary 12.3.4] we get
Thus, the result will be proved once we see that there is a complete isometry
LI(V
∨
⊗ W, Mn) ∼=(cid:18)(V
I(V × W, Mn) ∼=(cid:18)(V
∨
⊗ W )
∨
⊗ W )
.
∨
⊗ Tn(cid:19)∗
⊗ Tn(cid:19)∗
∨
.
To that end, consider the following applications:
BILINEAR IDEALS IN OPERATOR SPACES
11
, which is the canonical completely isometric
isomorphism given by the identification
• S : J CB(V × W, Mn) → (cid:0)(Vb⊗W )b⊗Tn(cid:1)∗
• bΨ : (V ∗b⊗W ∗)b⊗Mn → N (V × W, Mn), the quotient map.
• Ω : (V ∗b⊗W ∗)b⊗Mn →(cid:18)(V
J CB(V × W, Mn) ∼= CB(Vb⊗W, Mn) ∼=(cid:0)(Vb⊗W )b⊗Tn(cid:1)∗
⊗ Tn(cid:19)∗
∨
⊗ W )
∨
∨
⊗ W
∨
⊗ Tn)∗
n → (V
.
7→ (v ⊗ w ⊗ φ 7→ v∗(v)w∗(w)φ∗(φ)),
V ∗ × W ∗ × T ∗
(v∗, w∗, φ∗)
, the linearization of the trilinear mapping
which is completely contractive.
∨
⊗ Tn(cid:19)∗
∨
⊗ W )
• Φ∗ : (cid:18)(V
(Vb⊗W )b⊗Tn → (V
֒→ (cid:0)(Vb⊗W )b⊗Tn(cid:1)∗
range, Φ∗ results an injective complete contraction.
, which is the transpose mapping of Φ :
∨
⊗ W )
∨
⊗ Tn. Since Φ is a complete contraction and it has dense
Replicating the argument of the linear case we use the previous mappings to construct a com-
mutative diagram:
yS
N (V × W, Mn) ⊆
I(V × W, Mn)
⊆ J CB(V × W, Mn)
bΨx
(V ∗b⊗W ∗)b⊗Mn
∨
∨
⊗ W )
Ω−→ (cid:18)(V
⊗ Tn(cid:19)∗
Snuc : N (V × W, Mn) →(cid:18)(V
Φ∗
−→ (cid:0)(Vb⊗W )b⊗Tn(cid:1)∗
⊗ Tn(cid:19)∗
The injectivity of both SN and Φ∗ yields that ker(Ω) = ker(bΨ). This allows us to define:
in such a way that Snuc◦bΨ = Ω and Φ∗◦Snuc = SN . The mapping Snuc is a complete contraction.
Let us suppose now that φ ∈ I(V × W, Mn) with kφkI(V ×W,Mn) ≤ 1. We want to see
that S(φ) is continuous with respect to the injective tensor norm of (V ⊗ W ) ⊗ Tn. Given
u ∈ (V ⊗ W ) ⊗ Tn with kuk∨ ≤ 1, there exist finite-dimensional spaces Vu ⊂ V and Wu ⊂ W
such that u ∈ (Vu ⊗ Wu) ⊗ Tn. Let us call jVu : Vu ֒→ V and jWu : Wu ֒→ W the canonical
inclusions, then
∨
⊗ W )
∨
hS(φ), ui = hSnuc(φ ◦ (jVu , jWu )), ui.
Therefore,
hS(φ), ui ≤ kSnuc(φ ◦ (jVu , jWu))k(cid:18)(Vu
∨
⊗Wu)
∨
⊗Tn(cid:19)∗ · kuk
∨
⊗Wu)
∨
⊗Tn
(Vu
≤ kφ ◦ (jVu, jWu )kN (Vu×Wu,Mn) · kuk∨
≤ kφkI(V ×W,Mn) ≤ 1.
Thus, S determines a contractive mapping
Sint : I(V × W, Mn) →(cid:18)(V
∨
⊗ W )
∨
⊗ Tn(cid:19)∗
.
Through a similar argument it can be seen that Sint is also a complete contraction.
12
VERÓNICA DIMANT AND MAITE FERNÁNDEZ-UNZUETA
Let us show now that Sint is a complete isometry. For that, get φ ∈ Mm (I(V × W, Mn)) such
that k(Sint)m(φ)k
Mm(cid:18)(cid:18)(V
∨
⊗W )
∨
⊗Tn(cid:19)∗(cid:19) ≤ 1. We have to prove that kφkMm(I(V ×W,Mn)) ≤ 1.
Since (Sint)m(φ) ∈ Mm(cid:18)(cid:18)(V
∨
⊗ Tn ֒→
J CB(V ∗ × W ∗, Tn) is a complete isometry, by Remark 1.2, (Sint)m(φ) extends to ^(Sint)m(φ) ∈
CB(J CB(V ∗ × W ∗, Tn), Mm) preserving the norm. Now, we have completely isometric identifi-
cations
⊗ Tn(cid:19)∗(cid:19) ∼= CB((V
∨
⊗ Tn, Mm) and (V
∨
⊗ W )
∨
⊗ W )
∨
⊗ W )
∨
, Mm(cid:17)
CB(J CB(V ∗ × W ∗, Tn), Mm) ∼= CB(CB(V ∗b⊗W ∗, Tn), Mm) ∼= CB(cid:16)(cid:0)(V ∗b⊗W ∗)b⊗Mn(cid:1)∗
and we thus know that k ^(Sint)m(φ)kMm(((V ∗b⊗W ∗)b⊗Mn)∗∗) ≤ 1. Hence, by Remark 3.9, there exists
a net (uλ) in Mm(cid:0)(V ∗b⊗W ∗)b⊗Mn(cid:1) with kuλk ≤ 1 such that, for all ϕ ∈(cid:0)(V ∗b⊗W ∗)b⊗Mn(cid:1)∗
∼= Mm(cid:16)(cid:0)(V ∗b⊗W ∗)b⊗Mn(cid:1)∗∗(cid:17) ,
ϕm(uλ) → ^(Sint)m(φ)(ϕ).
,
In particular, for any v ∈ V , w ∈ W and α ∈ Tn,
((v ⊗ w) ⊗ α)m(uλ) → ^(Sint)m(φ)((v ⊗ w) ⊗ α) = (Sint)m(φ)((v ⊗ w) ⊗ α).
Looking into the coordinates of this matrix limit, with the notation uλ = (uk,l
λ )k,l and φ =
(φk,l)k,l, we obtain
λ )(v, w), αi = ((v ⊗ w) ⊗ α)(uk,l
λ ) → Sint(φk,l)((v ⊗ w) ⊗ α) = hφk,l(v, w), αi,
for every (v, w) ∈ V × W , α ∈ Tn and k, l ∈ {1, . . . , m}. Thus, for each pair (v, w) ∈ V × W , the
converges weakly to φk,l(v, w). Being Mn a finite dimensional space, this
convergence turns out to be strong and now we can also forget the coordinates and look at the
hbΨ(uk,l
net(cid:16)bΨ(uk,l
λ )(v, w)(cid:17)k,l
whole picture again. So we have bΨm(uλ)(v, w) → φ(v, w), for all (v, w) ∈ V × W .
Since bΨ is a complete contraction, we know kbΨm(uλ)kMm(N (V ×W,Mn)) ≤ 1 and with an ap-
It only remains to prove that Sint is surjective. Let f ∈(cid:18)(V
⊗ Tn(cid:19)∗
. The surjectivity
of S tells us that there exists φ ∈ J CB(V × W, Mn) such that Φ∗(f ) = S(φ). Moreover, for finite
dimensional spaces F1 ∈ V and F2 ∈ W with canonical inclusions j1 : F1 ֒→ V and j2 : F2 ֒→ W
it holds
pealing to Lemma 3.7 we derive that kφkMm(I(V ×W,Mn)) ≤ 1.
∨
⊗ W )
∨
Φ∗(f ◦ (j1, j2)) = S(φ ◦ (j1, j2)).
Since φ ◦ (j1, j2) belongs to N (F1 × F2, Mn) ∼= I(F1 × F2, Mn) it is clear that Sint(φ ◦ (j1, j2)) =
f ◦ (j1, j2).
Hence,
kφ ◦ (j1, j2)kN (F1×F2,Mn) = kφ ◦ (j1, j2)kI(F1×F2,Mn) = kSint(φ ◦ (j1, j2))k
= kf ◦ (j1, j2)k ≤ kf k.
Thus, φ ∈ I(V × W, Mn) with kφkI(V ×W,Mn) ≤ kf k.
(cid:3)
BILINEAR IDEALS IN OPERATOR SPACES
13
Now we can prove the general result I(V × W, X) ∼= LI (V
∨
⊗ W, X) :
Proof of Theorem 3.8. Let φ ∈ I(V × W, X) and consider the associated linear application
Lφ : V ⊗ W → X.
We begin by proving that Lφ is completely bounded from (V ⊗ W, ∨) to X. This will allows us to
extend Lφ to V
∨
⊗ W . For that, we need to find a common bound for the norms of the mappings
(Lφ)n : Mn(V ⊗ W, ∨) → Mn(X).
Let u ∈ Mn(V ⊗ W ). By [7, Lemma 2.3.4], there exists ξ ∈ CB(X, Mn) with kξkcb ≤ 1 satisfying
k(Lφ)n(u)kMn(X) = kξn ((Lφ)n(u)) kMn(Mn) = k(ξ ◦ Lφ)n(u)kMn(Mn) = k(Lξ◦φ)n(u)kMn(Mn).
Since ξ ◦ φ : V × W → Mn is completely integral, we know from Proposition 3.10 that Lξ◦φ
belongs to LI(V
∨
⊗ W, Mn). Thus, Lξ◦φ ∈ CB(V
∨
⊗ W, Mn) and therefore,
k(Lξ◦φ)n(u)kMn(Mn) ≤ kLξ◦φkcb · kuk
Mn(V
∨
⊗W )
≤ kξkcb · kφkI(V ×W,X) · kuk
Mn(V
.
∨
⊗W )
This yields that Lφ ∈ CB(V
Lφ belongs to Mn(cid:18)LI(V
∨
∨
⊗W, X). Let us prove now that, indeed, given φ ∈ Mn (I(V × W, X)),
⊗ W, X)(cid:19). To that end we need to compute the nuclear norms of its
∨
F1
∨
⊗ F2)∗ ∼= F ∗
∨
⊗ F2, X). Thus,
restrictions to finite dimensional spaces. Let F ⊂ V
exist finite dimensional subspaces F1 ∈ V and F2 ∈ W such that F ⊂ F1
isometry (F1
LN (F1
∨
⊗ W be a finite dimensional subspace. There
∨
⊗ F2. The complete
2 (see, for instance, [7, (15.4.1)]) yields that N (F1 × F2, X) ∼=
1b⊗F ∗
kLφF kMn(LN (F,X)) ≤(cid:13)(cid:13)(cid:13)Lφ
⊗F2(cid:13)(cid:13)(cid:13)Mn(cid:18)LN (F1
⊗F2,X)(cid:19) = kφF1×F2kMn(N (F1×F2,X)) ≤ kφkMn(I(V ×W,X))
Hence, it follows that Lφ ∈ Mn(cid:18)LI(V
⊗ W, X)(cid:19) and kLφk
To prove the opposite contention, consider L ∈ Mn(cid:18)LI (V
⊗F2(cid:13)(cid:13)(cid:13)Mn(cid:18)LN (F1
kφF1×F2kMn(N (F1×F2,X)) =(cid:13)(cid:13)(cid:13)Lφ
L is Lφ, for some φ ∈ Mn (J CB(V × W, X)) . The same argument as above shows that for any
finite dimensional subspaces F1 ∈ V and F2 ∈ W ,
Mn(cid:18)LI (V
⊗W,X)(cid:19) ≤ kφkMn(I(V ×W,X)).
⊗ W, X)(cid:19). It is plain to see that
Consequently, φ ∈ Mn (I(V × W, X)) and kφkMn(I(V ×W,X)) ≤ kLφk
⊗F2,X)(cid:19) ≤ kLφk
Mn(cid:18)LI (V
∨
⊗W,X)(cid:19).
∨
F1
(cid:3)
∨
∨
∨
∨
∨
Mn(cid:18)LI (V
∨
⊗W,X)(cid:19).
The scalar valued case. Let V and W be operator spaces and let ν be the linear isomorphism
in (1). As a corollary of Theorem 3.8 we have that ν induces the following complete isometry:
Proposition 3.11. I(V × W ) ∼= (V
∨
⊗ W )∗.
In contrast, in the case of the nuclear bilinear ideal we have:
Proposition 3.12. The following are equivalent:
14
VERÓNICA DIMANT AND MAITE FERNÁNDEZ-UNZUETA
(i) There exists an operator space tensor norm α such that N (V × W ) ∼= (V
(ii) N (V × W ) = I(V × W ).
α
⊗ W )∗.
In this case, α coincides with the injective operator space tensor norm.
Proof. (i) follows from (ii) by Proposition 3.11. To prove the other implication, recall that
k · k∨ ≤ k · kα for any operator space tensor norm α. Thus, if (i) holds for some α, then
I(V × W ) ∼= (V
α
⊗ W )∗ ∼= N (V × W ).
∨
⊗ W )∗ ⊂ (V
(cid:3)
It is worth noticing that there are examples of completely integral scalar valued bilinear
mappings which are not completely nuclear (see Example 6.1). Thus, the completely nuclear
bilinear ideal is not of the type described in Proposition 2.6.
Something more can be said about a tensorial representation of N (V × W ). First, recall the
following definition
Definition 3.13. An operator space V is said to have the operator space approximation
∨
⊗ V and for every ε > 0 there exists a finite rank
property (OAP) if for every u ∈ K(H)
mapping T on V such that ku − (I ⊗ T )(u)k < ε.
By [7, Theorem 11.2.5], V has OAP if and only if the canonical inclusion Vb⊗W ֒→ V
∨
⊗ W is
one-to-one, for every operator space W (or just for V ∗). Recall that the standard translation of
this result to the Banach space setting was also valid. As a direct consequence we can state the
following:
Proposition 3.14. If V ∗ or W ∗ has OAP then there is a complete isometry:
N (V × W ) ∼= V ∗b⊗W ∗.
As an example we can consider a reflexive operator space V such that its dual V ∗, looked
as a Banach space has the (Banach) approximation property but as an operator space V ∗ has
not OAP (see [1, 11] for examples of such spaces). In this case, the space of (Banach) nuclear
bilinear forms on V × V ∗ has a canonical representation as a projective tensor product while the
space of completely nuclear bilinear forms has not:
N B(V × V ∗) ∼= V ∗ ⊗π V ∗∗
and
N (V × V ∗) 6∼= V ∗b⊗V ∗∗.
Remark 3.15. The argument in Proposition 3.14 can be easily extended to the vector valued
case. Hence, we have
whether two of the three spaces V ∗, W ∗ and X have OAP.
N (V × W, X) ∼= (V ∗b⊗W ∗)b⊗X,
Looking at the equivalence J CB(V ×W ) ≃ CB(V, W ∗) and taking into account the situation in
the Banach space setting, we question about the existence of an operator space identification for
completely nuclear bilinear/linear mappings and for completely integral bilinear/linear mappings.
For the nuclear case, a careful look to the definitions of the spaces of completely nuclear
bilinear and linear mappings, easily gives the following.
Proposition 3.16. N (V × W ) ∼= LN (V, W ∗).
BILINEAR IDEALS IN OPERATOR SPACES
15
The situation for completely integral mappings is quite different: since LI(V, W ∗) is not
∨
⊗ W )∗ [7, Section 12.3] then neither the spaces I(V × W ) and
always completely isometric to (V
LI(V, W ∗) are always completely isometric. In the Banach space setting, the space of integral
bilinear forms from two Banach spaces is isometrically isomorphic to the space of integral linear
mappings from one of the spaces to the dual of the other (see, for instance, [17, Proposition
3.22]). The hidden reason behind this different behavior is the Principle of Local Reflexivity,
which is valid for every Banach space while its operator space version does not always hold (see
[7, Section 14.3] or [14, Definition 18.1] for a precise definition).
Indeed, [7, Theorem 14.3.1]
along with Proposition 3.11 give us the statement below.
Proposition 3.17. Let W be an operator space.Then the following are equivalent:
(i) W is locally reflexive.
(ii) For every operator space V , there is a complete isometry I(V × W ) ∼= LI(V, W ∗).
4. Completely extendible bilinear mappings
Within the scope of Banach spaces, the non-validity of a Hahn-Banach theorem for multilinear
mappings and homogeneous polynomials motivates the study of the 'extendible' elements (those
that can be extended to any superspace). We propose and study here a version of this concept for
bilinear mappings between operator spaces. Our approach was strongly inspired by the results
and arguments of [4] (see also [9]).
Definition 4.1. A mapping φ ∈ J CB(V × W, Z) is completely extendible if for any operator
spaces X and Y such that V ⊂ X, W ⊂ Y there exists a jointly completely bounded extension
φ : X × Y → Z of φ.
By the Representation Theorem for operator spaces (see, for instance [7, Theorem 2.3.5]), any
operator space can be seen, through a complete isometry, as a subspace of certain L(H).
Given V and W , let us denote the complete isometries that realize these spaces by
ΩV : V → L(HV )
and
ΩW : W → L(HW ).
Following the idea of [4, Theorem 3.2], we obtain:
Proposition 4.2. A jointly completely bounded mapping φ : V × W → Z is extendible if and
only if it can be extended to L(HV ) × L(HW ). In this case, if φ0 is such an extension, then for
every X ⊃ V and Y ⊃ W there exists an extension φ : X × Y → Z with kφkjcb ≤ kφ0kjcb.
Proof. Let φ0 : L(HV ) × L(HW ) → Z be an extension of φ. By Remark 1.2, ΩV and ΩW have
complete contractive extensions ΩV : X → L(HV ) and ΩW : Y → L(HW ). Then, φ : X ×Y → Z
given by
φ(x, y) = φ0(ΩV (x), ΩW (y)),
for all x ∈ X, y ∈ Y,
extends φ and
kφkjcb ≤ kφ0kjcb · kΩV kcb · kΩW kcb = kφ0kjcb.
(cid:3)
16
VERÓNICA DIMANT AND MAITE FERNÁNDEZ-UNZUETA
Let
E(V × W, Z) = {φ ∈ J CB(V × W, Z) : φ is extendible} .
It is clear that E(V × W, Z) is a subspace of J CB(V × W, Z). Moreover, it is an operator space
if we consider the following norm: for each φ ∈ Mn (E(V × W, Z)), let kφkE be the infimum of
the numbers C > 0 such that for all X ⊃ V and Y ⊃ W there exists φ ∈ Mn (J CB(X × Y, Z))
which extends φ, kφkjcb ≤ C. The previous proposition tells us that we can define equivalently
kφkE = inf {kφ0kjcb : φ0 extension of φ to Mn (J CB(L(HV ) × L(HW ), Z))} .
Proposition 4.3. E is an operator space bilinear ideal.
Proof. Since continuous functionals are completely extendible, it is clear that all finite type
continuous bilinear mappings belong to this subspace.
(a) For any φ ∈ Mn (E(V × W, Z)) we know that kφkjcb ≤ kφ0kjcb for every extension φ0 ∈
Mn (J CB(L(HV ) × L(HW ), Z)). Thus, kφkjcb ≤ kφkE .
(b) Consider φ ∈ Mn (E(V × W, Z)), r1 ∈ CB(U1, V ), r2 ∈ CB(U2, W ) and s ∈ CB(Z, Y ).
Since φ is a matrix of completely extendible maps, given ε > 0, there exists an extension
φ0 ∈ Mn (J CB(L(HV ) × L(HW ), Z)) such that kφ0kjcb ≤ kφkE + ε.
According to Remark 1.2, let R1 : L(HU1) → L(HV ) and R2 : L(HU2) → L(HW ) be completely
bounded extensions of r1 and r2, respectively, with kr1kcb = kR1kcb and kr2kcb = kR2kcb. Then,
sn ◦ φ0 ◦ (R1, R2) is an extension of sn ◦ φ ◦ (r1, r2) to L(HU1) × L(HU2) and
ksn ◦ φ0 ◦ (R1, R2)kjcb ≤ kskcb · kφ0kjcb · kR1kcb · kR2kcb ≤ kskcb · (kφkE + ε) · kr1kcb · kr2kcb.
Therefore, sn ◦ φ ◦ (r1, r2) ∈ Mn (E(U1 × U2, Z)) and ksn ◦ φ ◦ (r1, r2)kE ≤ kskcb · kφkE · kr1kcb ·
kr2kcb.
(cid:3)
Motivated by what is done in the Banach space setting (see [4, Corollary 3.9] or [9, Proposition
3]), we now define an operator space tensor norm η such that for any V, W , the dual operator space
η
⊗ W )∗ coincides with the scalar-valued completely extendible bilinear mappings E(V × W ).
(V
To that end, consider the tensor product of the canonical operator space inclusions where the
range is endowed with the operator space projective tensor norm:
Let η be the operator space tensor norm in V ⊗ W induced by this application. Thus, for any
u ∈ Mn (V ⊗ W ),
ΩV ⊗ ΩW : V ⊗ W → L(HV )b⊗L(HW ).
η(u) = k(ΩV ⊗ ΩW )n(u)k∧.
It is plain to see that η is an operator space matrix norm that does not depend on the
representations of ΩV and ΩW but just on the operator space structure of V and W . Also, since
ΩV and ΩW are complete isometries it easily follows that η is a cross matrix norm. Moreover, it
can be proved evidently that η is an operator space tensor norm according to Definition 2.3.
Let V
η
⊗ W denote the completion of (V ⊗ W, η).
Proposition 4.4. There is a complete isometry
∼= E(V × W ).
(cid:18)V
η
⊗ W(cid:19)∗
BILINEAR IDEALS IN OPERATOR SPACES
17
η
Since V
η
⊗ W(cid:19)∗
and denote by φ the associated bilinear form, φ : V × W → C.
Proof. Let ϕ ∈ (cid:18)V
of L(HV )b⊗L(HW ). By Remark 1.2, ϕ can be extended to ϕ0 : L(HV )b⊗L(HW ) → C with
⊗ W ֒→ L(HV )b⊗L(HW ) is a complete isometry, we can see V
kϕ0kcb = kϕkcb. It is easy to see that the bilinear map φ0 : L(HV ) × L(HW ) → C associated to
ϕ0 is an extension of φ. Also,
η
⊗ W as a subspace
Then, φ is completely extendible and kφkE ≤ kϕk.
kφ0kjcb = kϕ0kcb = kϕkcb.
Reciprocally,
let φ ∈ E(V × W ) and denote its linear associated by ϕ : V ⊗ W → C.
Let φ0 : L(HV ) × L(HW ) → C be an extension of φ and consider its linear associated ϕ0 ∈
. Thus, for each u ∈ V ⊗ W ,
ϕ(u) = ϕ0(ΩV ⊗ ΩW )(u) ≤ kϕ0kcb · k(ΩV ⊗ ΩW )(u)k∧ = kϕ0kcb · kukη.
This implies that ϕ is η-continuous and so it can be extended continuously to V
η
⊗ W . Hence,
(cid:0)L(HV )b⊗L(HW )(cid:1)∗
ϕ ∈(cid:18)V
⊗ W(cid:19)∗
The isometry between(cid:18)V
η
that the isometry is complete.
with kϕk ≤ kφkE .
η
⊗ W(cid:19)∗
and E(V × W ) is now proved and a similar argument shows
(cid:3)
5. The symmetrized multiplicatively bounded bilinear ideal
Given a bilinear mapping φ : V × W → Z, its transposed φt : W × V → Z is defined by the
relation φt(w, v) = φ(v, w). We will say that an operator space bilinear ideal A is symmetric
when satisfies that if φ ∈ A(V × W, Z) then φt ∈ A(W × V, Z) with kφkA = kφtkA.
The bilinear ideals J CB, N , I and E are clearly symmetric, while MB is not (see Example
6.2).
Definition 5.1. A bounded bilinear mapping φ : V ×W → Z is symmetrized multiplicatively
bounded, φ ∈ SMB(V × W, Z) if it can be decomposed as φ = φ1 + φ2 with φ1 ∈ MB(V × W, Z)
and φt
2 ∈ MB(W × V, Z).
The space SMB(V ×W, Z) is equiped with an operator space structure through the identifica-
tion with the sum MB(V ×W, Z)+ tMB(W ×V, Z) in the sense of operator spaces interpolation
theory (see [12, Chapter 2]). In this way, the norm of a matrix φ ∈ Mn(SMB(V × W ; Z)) is
given by
kφksmb = inf(cid:8)k(φ1, φ2)kMn(MB(V ×W,Z)⊕1
tMB(W ×V,Z)) : φ = φ1 + φ2(cid:9) .
Proposition 5.2. SMB is a symmetric operator space bilinear ideal.
Proof. By means of [12, Proposition 2.1] it is easy to see that whenever A1 and A2 are operator
space bilinear ideals then the same holds for A1 + A2. Hence, this is valid for SMB = MB +
tMB.
(cid:3)
18
W
VERÓNICA DIMANT AND MAITE FERNÁNDEZ-UNZUETA
We denote by (V
h
⊗ V . Appealing again to interpolation theory, we can see (V
h
⊗ V ) the set of elements u in V
h
⊗ W ) ∩ (W
h
⊗ W ) ∩ (W
h
⊗ W such that ut belongs to
h
⊗ V ) as an operator
space with the structure inherited by the canonical inclusion in (V
h
⊗ W ) ⊕∞ (W
h
⊗ V ).
The completely isometric identity (X ∩ Y )∗ ∼= X∗ + Y ∗ [12, page 23] applied to our case says:
SMB(V × W ) ∼=(cid:16)(V
h
⊗ W ) ∩ (W
h
⊗ V )(cid:17)∗
,
completely isometrically.
In the vector-valued case, there is also some interplay between the
space of symmetrized multiplicatively bounded bilinear mappings and the intersection of both
Haagerup tensor products:
Proposition 5.3. Let V , W and Z be operator spaces. Then:
h
⊗ V ), Z) is a complete contraction.
h
⊗ W ) ∩ (W
(a) The inclusion SMB(V × W, Z) ֒→ CB((V
(b) If Z = L(H) there is a complete isomorphism
SMB(V × W, L(H)) ∼= CB((V
h
⊗ W ) ∩ (W
h
⊗ V ), L(H)).
Proof. (a) Composing the restriction with the usual identification we naturally have the following
complete contractions:
MB(V ×W, Z) ֒→ CB((V
h
⊗W )∩(W
h
⊗V ), Z)
and
tMB(W ×V, Z) ֒→ CB((V
h
⊗W )∩(W
h
⊗V ), Z).
Thus, the classical interpolation property (see [12, Proposition 2.1]) gives that the mapping
SMB(V × W, Z) = MB(V × W, Z) + tMB(W × V, Z) ֒→ CB((V
h
⊗ W ) ∩ (W
h
⊗ V ), Z)
is also a complete contraction.
h
⊗ W ) ∩ (W
h
h
⊗ W ) ∩ (W
h
(b) In the case Z = L(H), let us see that the injective mapping of (a) is actually a surjective
to prove that the bilinear associate φ belongs to Mn (SMB(V × W, L(H))) with kφk ≤ 2kLφk.
⊗ V ), L(H))(cid:17). We have
complete isomorphism. For that, consider Lφ ∈ Mn(cid:16)CB((V
Since Lφ ∈ Mn(cid:16)CB((V
⊗ V ), L(H n)(cid:17) and
an extension Leφ ∈ CB(cid:16)(V
⊗ V ), L(H n)(cid:17) with the same completely bounded
norm. Then, we should have that the bilinear associated to Leφ is written as φ1 + φ2 with
kφ1kMB(V ×W,L(H n)) ≤ kLeφk and kφt
2kMB(W ×V,L(H n)) ≤ kLeφk. Hence,
⊗ V ), L(H))(cid:17) ∼= CB(cid:16)(V
h
⊗ V ) is completely isometrically contained in (V
h
⊗ W ) ⊕∞ (W
h
⊗ W ) ⊕∞ (W
h
⊗ V ) there is
h
⊗ W ) ∩ (W
h
h
⊗ W ) ∩ (W
(V
h
kφ1kMB(V ×W,L(H n)) + kφt
2kMB(W ×V,L(H n)) ≤ 2kLφk.
Now, the usual identification MB(V × W, L(H n)) ∼= Mn (MB(V × W, L(H))) yields:
kφkMn(SMB(V ×W,L(H))) ≤ kφ1kMn(MB(V ×W,L(H))) + kφt
2kMn(MB(W ×V,L(H)))
≤ 2kLφk
Mn(cid:16)CB((V
h
⊗W )∩(W
h
⊗V ),L(H))(cid:17).
(cid:3)
BILINEAR IDEALS IN OPERATOR SPACES
19
The case of scalar valued mappings is of special interest and was extensively studied in the
literature in relation with the so called Non-commutative Grothendieck's Theorem. In the next
section there is a briefly exposition of this.
We thank the referee for suggesting us to study the symmetrized multiplicatively bounded
mappings and for his/her very valuable comments.
6. Proof of Theorem 1.1 and Examples
Now we study the relationships between the bilinear ideals: we prove the inclusion relations
that always hold, and provide examples that distinguish them when they are different.
Proof of Theorem 1.1(a). It is clear, by definition, that every completely nuclear bilinear map-
ping is completely integral. Also, the fact that k · k∨ is smaller than k · kh implies that
LI (V
∨
⊗ W, X) ⊂ CB(V
∨
⊗ W, X) ⊂ CB(V
h
⊗ W, X).
h
Moreover, since I(V × W, X) ∼= LI (V
⊗ W, X), we
obtain that I(V × W, X) ⊂ MB(V × W, X). From the very definition of SMB, the relation
MB(V × W, X) ⊂ SMB(V × W, X) always holds.
∨
⊗ W, X) and MB(V × W, X) ∼= CB(V
All these inclusions are strict as we can see in the following examples.
Recall that in the Banach space setting, a classical example of an integral non-nuclear bilinear
mapping is φ : ℓ1 × ℓ1 → C given by φ(x, y) =Pn xnyn. For operator spaces, a similar example
works.
Example 6.1. A completely integral bilinear form which is not completely nuclear.
Let us consider the operator space τ (ℓ2) of trace class operators from ℓ2 to ℓ2. Naturally, each
element x ∈ τ (ℓ2) is identified with an infinite matrix (xs,t).
We define a bilinear map φ : τ (ℓ2) × τ (ℓ2) → C by
φ(x, y) =Xs
xs,s · ys,s
The bilinear map φ is jointly completely bounded but not completely nuclear. Indeed, by Propo-
sition 3.16, if φ is completely nuclear so is Lφ : τ (ℓ2) → L(ℓ2) given by
Lφ(x) =
.
x1,1
0
0
0
...
0
x2,2
0
· · ·
· · ·
0
x3,3
0
· · ·
· · ·
0
x4,4
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
Lφ could not be completely nuclear because it is not compact [7, Proposition 12.2.1].
Now we want to see that φ is completely integral. Invoking Lemma 3.7, we want to estimate
the completely nuclear norms of the mappings φm : τ (ℓ2) × τ (ℓ2) → C given by
φm(x, y) =
xs,s · ys,s.
mXs=1
20
VERÓNICA DIMANT AND MAITE FERNÁNDEZ-UNZUETA
For each s ∈ N, let us denote by εss the element in L(ℓ2) represented by the matrix with a
number 1 in position (s, s) and numbers 0 in all the other places. Recall that N (τ (ℓ2) × τ (ℓ2)) ∼=
defined in Section 3. Since
L(ℓ2)b⊗L(ℓ2)/ ker Ψ, where Ψ : L(ℓ2)b⊗L(ℓ2) → J CB(τ (ℓ2) × τ (ℓ2)) is the canonical mapping
we have kφmkN ≤ kPm
mXs=1
φm = Ψ mXs=1
2m Xδ∈{−1,1}m mXs=1
s=1 εss ⊗ εssk∧. In order to compute this norm, consider the following
δsεss! ⊗ mXs=1
εss ⊗ εss! ,
δsεss! .
usual way of expressing it:
εss ⊗ εss =
(5)
1
It is easy to prove that for vectors v1, . . . , vp in any operator space V we have the following
representation:
vj ⊗ vj = α · ((v1 ⊕ · · · ⊕ vp) ⊗ (v1 ⊕ · · · ⊕ vp)) · β,
pXj=1
εss ⊗ εss(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)∧
where α ∈ M1×p2, β ∈ Mp2×1 and both α and β have '1' in p of the places and '0' in the others.
Applying this representation to the expression (5), we obtain
≤
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
mXs=1
others. Since kαk = kβk = 2m/2 and kPm
where α ∈ M1×22m , β ∈ M22m×1 and both α and β have '1' in 2m of the places and '0' in the
s=1 εss ⊗
εssk∧ ≤ 1. Hence, kφmkN ≤ 1 (in fact, it is equal to 1) and by Lemma 3.7, φ is completely
integral with kφkI = 1.
1
2m kαk · max
· kβk,
δsεss(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
δ∈{−1,1}m(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
mXs=1
s=1 δsεsskL(ℓ2) = maxs δs = 1, we derive kPm
L(ℓ2)
2
Example 6.2. A multiplicatively bounded bilinear mapping which is not completely integral / A
symmetrized multiplicatively bounded bilinear mapping which is not multiplicatively bounded.
Let H be a Hilbert space and denote by Hc the column space associated to H. An exam-
ple of non commutativity of Haagerup tensor product is given through the canonical complete
isometries (see, for instance [7, Propositions 9.3.1, 9.3.2 and 9.3.4]):
h
⊗ (Hc)∗ ∼= Hc
∨
⊗ (Hc)∗ ∼= K(H)
Hc
and
(Hc)∗ h
A close look to these mappings allows us to state that the application
⊗ Hc ∼= (Hc)∗b⊗Hc ∼= τ (H).
Hc ⊗ (Hc)∗ → (Hc)∗ h
v ⊗ w 7→ w ⊗ v
⊗ Hc
could not be extended as a completely bounded mapping defined on Hc
h
⊗ (Hc)∗. Consider
φ : (Hc)∗ × Hc → (Hc)∗
h
⊗ Hc
and
φt : Hc × (Hc)∗ → (Hc)∗
h
⊗ Hc
(w, v)
7→
w ⊗ v
(v, w)
7→ w ⊗ v.
It turns out that φ is multiplicatively bounded while φt is not. Hence, φ could not be completely
integral (because the ideal of completely integral bilinear mappings is symmetric). Therefore, φ
BILINEAR IDEALS IN OPERATOR SPACES
21
is multiplicatively bounded but not completely integral and φt is symmetrized multiplicatively
bounded but not multiplicatively bounded.
We also see in [8, Example 3.6], or in Example 6.5 below, that the bilinear ideals SMB and
J CB do not coincide.
Proof of Theorem 1.1(b). The ideal of completely extendible bilinear mappings cannot be placed
as a link in the chain of inclusions in Theorem 1.1 (a):
It contains the ideal of completely
integral bilinear operators (see arguments below), but it has not a relation with the ideal of
multiplicatively bounded bilinear mappings holding for every operator space. Examples 6.4 and
6.6 prove this. We will see, though, that in the particularly relevant cases when the range is C
or L(H) there are relations between them.
In the Banach space setting, Grothendieck-integral bilinear mappings are always extendible [5,
Proposition 7]. Let us see that an analogous contention holds in the operator space framework.
Pisier (personal communication) made us realize that completely integral linear mappings being
completely 2-summing are hence completely extendible [13, Proposition 6.1]. This linear result
allows us to derive the bilinear one.
Indeed, from Theorem 3.8, we know I(V × W, X) ∼= LI(V
∨
⊗ W, X). Now, the previous linear
∨
⊗ W, X) ⊂ LE (V
inclusion gives us LI (V
k · k∨ is smaller than η, and LE is an ideal, we have LE (V
η
conclusion follows once we see that given any ϕ ∈ LE (V
⊗ W, X), its associated bilinear mapping
φ : V × W → X belongs to E(V × W, X).
η
⊗ W, X). Now, the
∨
⊗ W, X). Also, since the operator space tensor norm
∨
⊗ W, X) ⊂ LE (V
The extendibility of ϕ along with the inclusion V
⊗ W ֒→ L(HV )b⊗L(HW ) produce that, for
any ε > 0 there exists a completely bounded linear mapping ϕ0 : L(HV )b⊗L(HW ) → X that
extends φ with
kϕkLE ≤ kϕ0kcb ≤ kϕkLE + ε.
η
It is clear now that the bilinear map associated to ϕ0, φ0 : L(HV ) × L(HW ) → X, is an extension
of φ that satisfies
Hence, φ is completely extendible with kφkE ≤ kϕkLE .
kφkE ≤ kφ0kjcb = kϕ0kcb ≤ kϕkLE + ε.
Therefore, (b) in Theorem 1.1 is proved: I(V × W, X) ⊂ E(V × W, X) ⊂ J CB(V × W, X).
Examples 6.4 and 6.6 below, will show that both inclusions could be strict.
It is known [20, page 45] that multiplicatively bounded bilinear mappings with range L(H) are
completely extendible. This can also be seen as a consequence of Arvenson-Wittstock extension
theorem for completely bounded mappings (Remark 1.2) along with the fact that the Haagerup
tensor norm preserves complete isometries. Moreover, the inclusion MB(V × W, L(H)) ⊂ E(V ×
W, L(H)) is a complete contraction. Since E is a symmetric ideal, appealing once more to [12,
Proposition 2.1] we derive the complete contractive inclusion
SMB(V × W, L(H)) ⊂ E(V × W, L(H)),
which proves (c) in Theorem 1.1.
22
VERÓNICA DIMANT AND MAITE FERNÁNDEZ-UNZUETA
We do not know whether this last inclusion is strict. Actually, for scalar-valued bilinear map-
pings we do know that the equality isomorphically holds. This is a consequence of Grothendieck's
Theorem for C∗-algebras. In [15] one may find a broad exposition on the topic. For the moment
let us recall just some relevant results in a terminology according to our presentation. Pisier and
Shlyakhtenko [16] obtain the result for exact operator spaces (and also for C ∗-algebras satisfying
some conditions). In [16, Theorem 0.4] they prove:
Theorem (Pisier-Shlyakhtenko). If V and W are exact operator spaces, then the following iso-
morphism holds:
SMB(V × W ) = J CB(V × W ).
Haagerup and Musat [8] prove the theorem for general C ∗-algebras. Combining [8, Theorem
1.1] with [8, Lemma 3.1] (which relies on Pisier and Shlyakhtenko's result) produces:
Theorem (Haagerup-Musat). If A and B are C ∗-algebras, then the following isomorphism holds:
SMB(A × B) = J CB(A × B).
As a consequence, for any operator spaces V and W the following (Banach space) isomorphism
holds:
SMB(V × W ) = E(V × W ).
Indeed, let φ ∈ E(V × W ). For V → L(HV ) and W → L(HW ) complete isometries and
ε > 0, let ψ : L(HV ) × L(HW ) → C be a jointly completely bounded extension of φ with
kψkjcb ≤ kφkE + ε. By Haagerup-Musat's Theorem (for A = L(HV ) and B = L(HW )), ψ can
be decomposed as ψ = ψ1 + ψ2, with ψ1 ∈ MB(L(HV ) × L(HW )), ψt
2 ∈ MB(L(HW ) × L(HV ))
and kψ1kmb + kψt
2kmb ≤ Kkψkjcb. Restricting the domains of ψ1 and ψ2 to V × W , we complete
the proof.
A predual version of the last expression reads as
η
⊗ W = (V
V
h
⊗ W ) ∩ (W
h
⊗ V )
isomorphically.
It is worth noticing that Oikhberg and Pisier in [10] proved that the sum of these Haagerup
h
⊗ W ) + (W
h
⊗ V ) is completely isometric to the "maximal" tensor product
tensor products (V
V
µ
⊗ W which was introduced and studied in that article.
Let us now show that the other two inclusions of Theorem 1.1 (c) are strict. We have already
distinguished the space of multiplicatively bounded bilinear forms from its symmetrized relative.
These spaces may be different even when the range is L(H). To construct an example, first we
need an easy observation:
Remark 6.3. Let φ : V × W → X be a jointly completely bounded bilinear mapping and
j : X → Y be a complete isometry. Then, φ is multiplicatively bounded if and only if j ◦ φ is
multiplicatively bounded.
Indeed, for any v ∈ Mn(V ) and w ∈ Mn(W ), since (j ◦ φ)(n)(v, w) = jn(cid:0)φ(n)(v, w)(cid:1) we have
(cid:13)(cid:13)(j ◦ φ)(n)(v, w)(cid:13)(cid:13) =(cid:13)(cid:13)jn(cid:0)φ(n)(v, w)(cid:1)(cid:13)(cid:13) =(cid:13)(cid:13)φ(n)(v, w)(cid:13)(cid:13) .
Thus, kj ◦ φkmb = kφkmb.
BILINEAR IDEALS IN OPERATOR SPACES
23
Example 6.4. A symmetrized multiplicatively bounded bilinear mapping with range L(H), which
is not multiplicatively bounded / A completely extendible bilinear mapping which is not completely
integral.
We recover the mappings φ and φt of Example 6.2. Denoting by V = (Hc)∗
h
⊗ Hc, we consider
ΩV : V → L(HV ) the usual completely isometric inclusion. Now, let ψ = ΩV ◦ φ : (Hc)∗ × Hc →
L(HV ). The previous remark and the fact that φt is not multiplicatively bounded, imply that
ψt = ΩV ◦ φt : Hc × (Hc)∗ → L(HV ) neither is multiplicatively bounded.
On the other hand, φ ∈ MB((Hc)∗ × Hc, (Hc)∗
h
⊗ Hc) and so ψ ∈ MB((Hc)∗ × Hc, L(HV )).
Hence, ψt ∈ SMB((Hc)∗ × Hc, L(HV )).
Example 6.5. A jointly completely bounded bilinear mapping (with range C) which is not ex-
tendible (and hence not symmetrized multiplicatively bounded).
Consider a non-complemented copy of ℓ2 in L(H), and let V be the operator space determined
by ℓ2 with the matrix structure inherited from L(H). Let
φ : V × V ∗ →
((ai)i, (bi)i)
7→ P∞
C
i=1 aibi.
φ is jointly completely bounded but there is not a jointly completely bounded extension of φ
defined on L(H) × V ∗, since this extension would give rise to a bounded projection on L(H)
onto that copy of ℓ2.
Now we prove that the inclusion of the space of multiplicatively bounded bilinear mappings
(and hence symmetrized multiplicatively bounded) into the space of completely extendible bilin-
ear mappings is not longer true when the range space is an arbitrary operator space.
For that, it is convenient to introduce the concept of completely extendible linear mapping.
We say that a mapping ϕ ∈ CB(V, Z) is completely extendible if for any operator space X
such that V ⊂ X, there exists a completely bounded extension ϕ : X → Z of ϕ. The set of
completely extendible linear mappings from V to Z is denoted by LE (V, Z).
Following the same steps as in the proofs of Proposition 4.2 it is obtained that ϕ ∈ CB(V, Z)
is completely extendible if and only if it can be extended to L(HV ) and that LE (V, Z) is an
operator space with the norm given by
kϕkLE = inf{kϕ0kcb : ϕ0 extension of ϕ to Mn (CB(L(HV ), Z))},
for every ϕ ∈ Mn (LE (V, Z)).
As in Proposition 4.3 it is also obtained that LE is a (linear) mapping ideal.
Example 6.6. A multiplicatively bounded bilinear mapping which is not extendible.
Let V be the operator space of Example 6.5. The canonical mapping V
h
⊗ C → V is a complete
isometry. Hence, its associated bilinear map φ : V ×C → V is multiplicatively bounded. However,
since id : V → V is not extendible, φ neither is so.
Acknowledgements. The first author wishes to thank the Centro de Investigación en Matemáti-
cas (Guanajuato) for its kind hospitality during the months of January and February 2012, when
this work was initiated.
24
VERÓNICA DIMANT AND MAITE FERNÁNDEZ-UNZUETA
References
[1] Arias, Alvaro. Operator Hilbert spaces without the operator approximation property. Proc. Amer. Math.
Soc. 130 (2002), no. 9, 2669 -- 2677.
[2] Blecher, David P.; Le Merdy, Christian. Operator algebras and their modules: an operator space approach.
Oxford University Press, USA, 2005.
[3] Blecher, David P.; Paulsen, Vern I. Tensor products of operator spaces. J. Funct. Anal. 99 (1991), no. 2,
262 -- 292.
[4] Carando, Daniel. Extendible polynomials on Banach spaces. J. Math. Anal. Appl. 233 (1999), no. 1, 359 -- 372.
[5] Carando, Daniel; Lassalle, Silvia. Extension of vector-valued integral polynomials. J. Math. Anal. Appl. 307
(2005), no. 1, 77 -- 85.
[6] Effros, Edward G.; Junge, Marius; Ruan, Zhong-Jin. Integral mappings and the principle of local reflexivity
for noncommutative L1-spaces Ann. of Math. 151 (2000), 59-92.
[7] Effros, Edward G.; Ruan, Zhong-Jin. Operator spaces. London Mathematical Society Monographs. New
Series, 23. The Clarendon Press, Oxford University Press, New York, 2000.
[8] Haagerup, Uffe; Musat, Magdalena. The Effros-Ruan conjecture for bilinear forms on C∗-algebras. Invent.
Math. 174 (2008), no. 1, 139-163.
[9] Kirwan, Pádraig; Ryan, Raymond. Extendiblity of homogeneous polynomials on Banach spaces. Proc. Amer.
Math. Soc. 126 (1998), 1023 -- 1029.
[10] Oikhberg, Timur; Pisier, Gilles. The "maximal" tensor product of operator spaces. Proc. Edinburgh Math.
Soc. (2) 42 (1999), no. 2, 267 -- 284.
[11] Oikhberg, Timur; Ricard, Éric. Operator spaces with few completely bounded maps. Math. Ann. 328 (2004),
no. 1-2, 229 -- 259.
[12] Pisier, Gilles. The operator Hilbert space OH, complex interpolation and tensor norms. Mem. Amer. Math.
Soc. 122 (1996), no. 585.
[13] Pisier, Gilles. Non-commutative vector valued Lp-spaces and completely p-summing maps. Astérisque 247
(1998).
[14] Pisier, Gilles. Introduction to operator space theory. London Mathematical Society Lecture Note Series, 294.
Cambridge University Press, Cambridge, 2003.
[15] Pisier, Gilles. Grothendieck's theorem, past and present. Bull. Amer. Math. Soc. (N.S.) 49 (2012), no. 2,
237-323.
[16] Pisier, Gilles; Shlyakhtenko, Dimitri. Grothendieck's theorem for operator spaces. Invent. Math. 150 (2002),
no. 1, 185-217.
[17] Ryan, Raymond. Introduction to tensor products of Banach spaces. Springer Monographs in Mathematics.
Springer-Verlag London, Ltd., London, 2002.
[18] Schreiber, Bertram M. Operator spaces: Basic theory and applications. Advanced Courses of Mathematical
Analysis III, World Scientific Publishing, 2008.
[19] Villanueva, Ignacio. Integral mappings between Banach spaces. J. Math. Anal. Appl. 279 (2003), no. 1,
56 -- 70.
[20] Wittstock, Gerd, et al. "What are operator spaces." An online dictionary. URL: http://www. math. uni-sb.
de/ ag-wittstock/projekt2001. html (2001).
Departamento de Matemática, Universidad de San Andrés, Vito Dumas 284, (B1644BID) Vic-
toria, Buenos Aires, Argentina and CONICET.
E-mail address: [email protected]
Centro de Investigación en Matemáticas (Cimat), A.P. 402 Guanajuato, Gto., México
E-mail address: [email protected]
|
1304.6337 | 1 | 1304 | 2013-04-23T16:19:40 | A Resolvent Approach to the Real Quantum Plane | [
"math.OA"
] | Let $q\neq \pm 1$ be a complex number of modulus one. This paper deals with the operator relation $AB=qBA$ for self-adjoint operators $A$ and $B$ on a Hilbert space. Two classes of well-behaved representations of this relation are studied in detail and characterized by resolvent equations. | math.OA | math |
A Resolvent Approach to the Real Quantum Plane
VASYL OSTROVSKYI, KONRAD SCHM UDGEN
Abstract. Let q 6= ±1 be a complex number of modulus one. This paper
deals with the operator relation AB = qBA for self-adjoint operators A and B
on a Hilbert space. Two classes of well-behaved representations of this relation
are studied in detail and characterized by resolvent equations.
AMS Subject Classification (2000). 47D40, 81R50, 47B25
Key words: real quantum plane, q-commutation relations
1. Introduction
The algebraic relation ab = qba is a basic ingredient of the theory of quantum
groups. Let us assume for a moment that this relation holds for a complex number
q and some elements a and b of a unital ∗-algebra with involution x → x+. There
are three important cases in which this relation is invariant under the involution.
The first one is when a is unitary (that is, a+a = aa+ = 1) and b is hermitian (that
is, b+ = b), while in the second case we have a = b+. In both cases q is real. From
an operator-theoretic point of view these two cases are closely related (for instance,
by taking the polar decomposition of a in the second case). In the third case a and
b are hermitian and q is of modulus one. All three cases occur in the definitions
of real forms of quantum groups and quantum algebras, see e.g.
[7, Subsections
6.1.7, 9.2.4, 9.2.5]. The present paper deals with operator representations of the
relation ab = qba in this third case. The corresponding ∗-algebra generated by a
and b is the coordinate algebra of the real quantum plane [13] and of the quantum
ax + b-group [14].
The general operator relation ab = qba has been studied in many papers such as
[9], [2], [11], [12], [10], [15], [6], [3].
Throughout this paper q is a fixed complex number of modulus one such that
q2 6= 1 and A and B are self-adjoint operators on a Hilbert space H. We write
(1)
q = e−iθ0, where 0 < θ0 < π.
Our aim is to study the operator relation
(2)
AB = qBA.
It turns out that this simple operator relation leads to unexpected technical dif-
ficulties and interesting operator-theoretic phenomena. If A and B are bounded
and AB = 0, then (2) is obviously satisfied. Let us call representations of (2) with
AB = 0 trivial. Since q2 6= 1, these are the only representations of (2) given by
bounded operators (see [2] or [9]). Operator representations of algebraic relations
have been extensively studied in [9], but the methods developed therein lead only
to trivial representations of (2). Further, as noted in [11, p. 1031], in contrast to
Lie algebra relations the method of analytic vectors fails for the relation (2).
1
2
VASYL OSTROVSKYI, KONRAD SCHM UDGEN
Representations of (2) by unbounded self-adjoint operators A and B have been
investigated in [11] and [12]. Some classes of well-behaved representations of (2)
have been introduced and classified in [12]. The present paper is devoted to an
approach to the operator relation (2) that is based on the resolvents of the self-
adjoint operators A and B. For two classes C0 and C1 (see Definition 2) of well-
behaved representations this approach is developed in detail.
This paper is organized as follows. In Section 2 we give a number of reformu-
lations of the operator relation (2) in terms of the resolvent Rλ(A) and B, the
resolvent Rµ(B) and A, and the resolvents Rλ(A) and Rµ(B), and we study the
largest linear subspace Dq(A, B) on which relation (2) holds. In Section 4 the two
classes C0 and C1 of well-behaved representations of relation (2) are defined and
investigated in detail. All irreducible pairs of these classes are built of self-adjoint
operators eαQ and eβP on the Hilbert space L2(R), where Q = x, P = i d
dx
and α, β ∈ R. We prove that the weak resolvent forms qRλq(A)B ⊆ BRλ(A)
and Rµ(B)A ⊆ qARµq(B) of relation (2) hold for all pairs {A, B} of these classes
and for all complex numbers λ resp. µ outside certain critical sectors. Section 5
contains the main results of this paper. These are various theorems which char-
acterize (under additional technical assumptions) well-behaved representations, es-
pecially pairs {A, B} of the classes C0 and C1, by weak resolvent relations such as
qRλq(A)B ⊆ BRλ(A).
Let A := eαQ, B := eβP , and q := e−iαβ, where α, β ∈ R. As shown in Section
4, the resolvent relations (12) and (13) are satisfied on L2(R) if αβ < π and λ, µ
are not in the critical sector S(q)+. From Propositions 1 and 2 it follows that
for arbitary numbers α, β, λ, µ the relations (12) and (13) holds for vectors of the
closures of subspaces (B − µI)(A− λI)D0 and (A− λqI)(B − µqI)D0, respectively,
where D0 = Lin {e−εx2+γx; ε > 0, γ ∈ C}. In Section 6 the orthogonal complements
of these two subspaces are explicitely described and the resolvent actions on these
complements are computed.
Some technical preliminaries are contained in Section 3. Amongs these are prop-
erties of the operator eβP and a formula for fractional powers of sectorial operators.
Let us collect some basic notations on operators. Let T be a densely defined
closed operator on a Hilbert space. We denote its domain by D(T ), its resolvent
set by ρ(T ) and its resolvent (T − λI)−1 by Rλ(T ). Let UT be the phase operator
occuring in the polar decomposition T = UTT of the operator T . The symbol
L2(R) stands for the L2-space with respect to the Lebesgue measure on R.
2. General considerations on the relation (2)
The following two propositions contain some simple reformulations of equation
(2) in terms of the resolvents of the self-adjoint operators A and B.
Proposition 1. Suppose that λ, λq ∈ ρ(A) and µ, µq ∈ ρ(B).
then
(i) If D is a linear subspace of D(AB) ∩ D(BA) and (2) holds for all f ∈ D,
(3)
BRλ(A)g = qRλq(A)Bg
for all g ∈ E := (A − λI)D and E is a linear subspace of D(B).
satisfied for all g ∈ E, then (2) holds for all f ∈ D := Rλ(A)E.
(ii) If E is a linear subspace of D(B) such that Rλ(A)g ∈ D(B) and (3) is
A Resolvent Approach to the Real Quantum Plane
3
(iii) If E is a linear subspace of D(B) and (3) holds for all g ∈ E, then
(4)
Rλ(A)Rµ(B)h = qRµq(B)Rλq(A)h + µλq(q − 1)Rµq(B)Rλq(A)Rλ(A)Rµ(B)h
(iv) If F is a linear subspace of H such that (4) holds for all h ∈ F , then (3) is
for all h ∈ F := (B − µI)E.
fulfilled for all g ∈ E := Rµ(B)F .
Proof. (i): Clearly, (2) implies that
(A − λqI)Bg = qB(A − λI)g
for f ∈ D. Hence, for all vectors of the form g = (A − λI)f , where f ∈ D, we have
Rλ(A)g ∈ D(B) and
Rλq(A)(A − λqI)BRλ(A)g = qRλq(A)B(A − λI)Rλ(A)g,
so that
BRλ(A)g = qRλq(A)Bg.
(iii): Let g ∈ E. From equation (3) we obtain
(B − µqI)Rλ(A)g = (BRλ(A) − µqRλ(A))g = (qRλq(A)B − µqRλ(A))g
= (qRλq(A)(B − µI) + µqRλq(A) − µqRλ(A))g
= (qRλq(A)(B − µI) + µq(λq − λ)Rλq(A)Rλ(A))g.
(5)
Setting h = (B − µI)g, we have g = Rµ(B)h. Inserting this into (5) and applying
Rµq(B) to both sides yields (4) for h ∈ (B − µI)E.
(ii) and (iv) follow by reversing the preceding arguments of proofs of (i) and (iii),
respectively.
(cid:3)
Using the equalities Rλ(qA) = qRλq(A) and Rµ(qB) = qRµq(B) one can rewrite
(3) in the form
and (4) as
BRλ(A)g = Rλ(qA)Bg
qRλ(A)Rµ(B)h = Rµ(qB)Rλ(qA)h + µRµ(qB)(qRλ(qA) − Rλ(A))Rµ(B)h.
In a similar manner the following proposition is derived.
Proposition 2. Suppose that λ, λq ∈ ρ(A) and µ, µq ∈ ρ(B).
then
(i) If equation (2) is satisfied for all f of a linear subspace D ⊆ D(AB)∩D(BA),
Rµ(B)Ag = qARµq(B)g
(6)
for all g ∈ E := (B − µqI)D and E is a subspace of D(A).
for all g ∈ E, then (2) is true for all f ∈ D := Rµq(B)E.
(ii) If E is a linear subspace of D(A) such that Rµq(B)g ∈ D(A) and (6) holds
(iii) If E is a linear subspace of D(A) and (6) is satisfied for all g ∈ E, then
(7)
Rλ(A)Rµ(B)h = qRµq(B)Rλq(A)h + µλq(q − 1)Rλ(A)Rµ(B)Rµq(B)Rλq(A)h
for all h ∈ F := (A − λqI)E.
satisfied for all g ∈ E := Rλq(A)F .
(iv) If equation (7) holds for all h of a linear subspace F ⊆ H, then (6) is
4
VASYL OSTROVSKYI, KONRAD SCHM UDGEN
Comparing Propositions 1 and 2, especially formulas (4) and (7), we obtain
Corollary 3. Let λ, λq ∈ ρ(A) and µ, µq ∈ ρ(B). If equation (2) holds on a linear
subspace D ⊆ D(AB)∩D(BA), then the operators Rλ(A)Rµ(B) and Rµq(B)Rλq(A)
commute on the linear space (A − λqI)(B − qµI)D ∩ (B − µI)(A − λI)D.
Without further assumptions the linear subspace (A − λI)D of D(B) is neither
a core for B nor the subspace (B − µI)E is dense in H. Note that (2) for all f ∈ D
implies that (4) holds for all vectors h ∈ (B − µI)(A − λI)D and (7) is valid for
h ∈ (A − λqI)(B − µqI)D.
Definition 1. Dq(A, B) := {f ∈ D(BA) ∩ D(AB) : ABf = qBAf}.
Obviously, Dq(A, B) is the largest linear subspace of H on which relation (2)
holds. Of course, for arbitary self-adjoint operators A and B it may happen
Dq(A, B) = {0}. From Proposition 1 we immediately obtain the following de-
scriptions of the space Dq(A, B):
Dq(A, B) = Rλ(A){g ∈ D(B) : Rλ(A)g ∈ D(B) and BRλ(A)g = qRλq(A)Bg}
= Rλ(A)Rµ(B){h ∈ H : Rλ(A)Rµ(B)h = qRµq(B)Rλq(A)h
+ µλq(q − 1)Rµq(B)Rλq(A)Rλ(A)Rµ(B)h}.
Similarly, Proposition 2 leads to the following descriptions of Dq(A, B):
Dq(A, B) = Rµq(B){g ∈ D(A) : Rµq(B)g ∈ D(A) and Rµ(B)Ag = qARµq(B)g}
= Rλ(A)Rµ(B)(cid:8)h ∈ H : Rλ(A)Rµ(B)h = qRµq(B)Rλq(A)h
+ µλq(q − 1)(A)Rλ(A)Rµ(B)Rµq(B)Rλqh(cid:9).
In particular, we have
Dq(A, B) ⊂ Rλ(A)D(B) ∩ Rµq(B)D(A).
The operator relation (2) is obviously equivalent to the the relation
(8)
Hence Dq(A, B) = Dq(B, A).
BAf = qABf.
If equation (3) holds for all vectors g of the whole domain D(B), that is, if
(9)
qRλq(A)B ⊆ BRλ(A),
we shall say that relation (9) is the weak A-resolvent form of equation (2) for
λ, λq ∈ ρ(A). If equation (6) holds for all vectors g of the domain D(B), that is, if
(10)
Rµ(B)A ⊆ qARµq(B),
we say that relation (10) is the weak B-resolvent form of equation (2) for µ,
µq ∈ ρ(B). Setting ν = µq relation (10) can be rewritten as
(11)
qRνq(B)A ⊆ ARν (B).
The form (11) of the weak B-resolvent relation of (2) corresponds to the weak
A-resolvent form of equation (8) which is obtained by interchanging A and B and
replacing q by q .
Further, if equation (4) is satisfied for all h ∈ H, that is, if
(12)
Rλ(A)Rµ(B) = qRµq(B)Rλq(A) + µλq(q − 1)Rµq(B)Rλq(A)Rλ(A)Rµ(B),
A Resolvent Approach to the Real Quantum Plane
5
The resolvent relations (12) and (13) can be rewritten as
then equation (4) is called the (A, B)-resolvent form of equation (2) for λ, λq ∈
ρ(A) and µ, µq ∈ ρ(B). Likewise, if equation (7) holds for all h ∈ H, that is, if
(13)
Rλ(A)Rµ(B) = qRµq(B)Rλq(A) + µλq(q − 1)Rλ(A)Rµ(B)Rµq(B)Rλq(A),
then equation (7) is called the (B, A)-resolvent form of equation (2) for λ, λq ∈
ρ(A) and µ, µq ∈ ρ(B).
(cid:16)Rµq(B)Rλq(A) −
λ2µ2q(q − 1)2 I,
(cid:16)Rλ(A)Rµ(B) +
λ2µ2q(q − 1)2 I,
respectively. They hold for all vectors from the subspaces (B−µI)(A−λI)Dq (A, B)
and (A− λqI)(B − µqI)Dq(A, B), respectively. In Section 6 we derive for a class of
representations of (2) the form of resolvent relations on the complements of these
subspaces.
I(cid:17)(cid:16)Rλ(A)Rµ(B) +
I(cid:17)(cid:16)Rµq(B)Rλq(A) −
I(cid:17) = −
I(cid:17) = −
µλq(q − 1)
µλq(q − 1)
1
1
1
µλ(q − 1)
1
µλ(q − 1)
1
1
Proposition 4. The weak A-resolvent form (9) is equivalent to the (A, B)-resolvent
form (12) of equation (2). The weak B-resolvent form (10) and the (B, A)-resolvent
form (13) of (2) are equivalent.
Proof. First suppose that (9) holds. This means that (3) is satisfied for all vectors
g ∈ D(B). Therefore, by Proposition 1(ii), equation (3) holds for h ∈ (B−µI)D(B).
Since µ ∈ ρ(B), (B − µI)D(B) is equal to H which yields (12).
Conversely, assume that (12) is fulfilled. Let g ∈ D(B). We set h = (B − µqI)g
in (4). Since the ranges of resolvents of B are contained in the domain of B, the
vector in (4) is in D(B), so we can apply the operator B − µI to both sides of (4).
Then we obtain (3) which proves (9).
The equivalence of (13) and (10) follows by a similar reasoning.
(cid:3)
The next proposition collects a number of basic facts concerning the weak resol-
vent identities.
Proposition 5. Suppose that λ, λq ∈ ρ(A) and µ, µq ∈ ρ(B).
(i) qRλq(A)B ⊆ BRλ(A) if and only if Dq(A, B) = Rλ(A)D(B).
(ii) Rµ(B)A ⊆ qARµq(B) if and only if Dq(A, B) = Rµq(B)D(A).
(iii) qRλq(A)B ⊆ BRλ(A) if and only if qRλ(A)B ⊆ BRλq(A).
(iv) Rµ(B)A ⊆ qARµq(B) if and only if Rµq(B)A ⊆ qARµ(B).
(v) If qRλq(A)B ⊆ BRλ(A), then Dq(A, B) is a core for A.
(vi) If Rµ(B)A ⊆ qARµq(B) then Dq(A, B) is a core for B.
Proof. We carry out the proofs of (i), (iii), and (v). The proofs of (ii), (iv), and
(vi) follows by a similar reasoning.
(i): Throughout this proof let us set DB := Rλ(A)D(B).
First suppose that qRλ(A)B ⊆ BRλ(A). Obviously, DB ⊂ D(A). The inclusion
DB ⊂ D(B) follows from qRλ(A)Bf = BRλ(A)f , f ∈ D(B). Further, we have
BDB ⊂ D(A) since qRλ(A)Bf = BRλ(A)f and Rλ(A)Bf ⊆ D(A), f ∈ D(B).
Also, ADB ⊂ D(B), since A− λI maps DB onto D(B). Therefore, DB ⊂ Dq(A, B).
Since g = (A − λI)f ∈ D(B) for any f ∈ Dq(A, B), we see that f = Rλ(A)g, so
that Dq(A, B) ⊆ DB. Thus, DB = Dq(A, B).
Conversely, assume that DB = Dq(A, B). Then (A − λI)Dq(A, B) = D(B) and
by Proposition 1(i), we have qRλ(A)Bf = BRλ(A)f for all f ∈ D(B).
6
VASYL OSTROVSKYI, KONRAD SCHM UDGEN
(iii): Suppose that qRλq(A)B ⊆ BRλ(A). Since Rλq(A) is bounded, we have
(Rλq(A)B)∗ = B∗(Rλq(A))∗ = BRλq(A) and hence
q BRλq(A) = (qRλq(A)B)∗ ⊇ (BRλ(A))∗ ⊇ Rλ(A)B,
so that qRλ(A)B ⊆ BRλq(A).
The converse direction follows by applying the same implication once again.
(v): Since Dq(A, B) = Rλ(A)D(B) by (i), (A − λI)Dq(A, B) = D(B) is dense in
(cid:3)
H. Hence Dq(A, B) is a core for A.
An immediate consequence of Proposition 5 is the following corollary.
Corollary 6. Let λ, λq ∈ ρ(A) and µ, µq ∈ ρ(B). Assume that qRλ(A)B ⊆
BRλ(A) and Rµ(B)A ⊆ qARµq(B). Then
(14)
Dq(A, B) = Rλ(A)D(B) = Rµq(B)D(A)
and Dq(A, B) is a core for A and B.
Corollary 7. Suppose that µ, µq, µq2 ∈ ρ(B).
Rµq(B)A ⊆ qARµq2 (B), then Dq2 (A2, B) is a core for B.
Proof. From the assumptions we derive Rµ(B)A2 ⊆ qARµq(B)A ⊆ q2A2Rµq2 (B),
that is, the weak B-resolvent form for the relation A2B = q2BA2 is satisfied.
Therefore, Dq2 (A2, B) is a core for B by Proposition 5(vi).
(cid:3)
If Rµ(B)A ⊆ qARµq(B) and
The next proposition shows how the resolvent relations (12) and (13) follow from
the essential self-adjointness of a certain symmetric operator.
Let us fix a, b ∈ R and choose the branch of the square root such that
(15)
q1/2 = q1/2 .
We define an operator T with domain D(T ) := Dq(A, B) by
(16)
T f = ¯q1/2(A − aq1/2)(B − bq1/2)f +
Lemma 8. The operator T is symmetric.
¯q1/2 − q1/2
2
abf, f ∈ D(T ).
Proof. Clearly, T f = (cid:0)¯q1/2AB − bA − aB + q1/2+¯q1/2
derive
2
ab(cid:1)f . Using this formula we
ab(cid:1)f, gi
ab(cid:1)gi
ab(cid:1)gi = hT f, gi
q1/2 + ¯q1/2
2
q1/2 + ¯q1/2
q1/2 + ¯q1/2
2
2
hT f, gi = h(cid:0)¯q1/2AB − bA − aB +
= hf,(cid:0)q1/2BA − bA − aB +
= hf,(cid:0)¯q1/2AB − bA − aB +
for f, g ∈ D(T ), that is, T is symmetric.
Proposition 9. Assume that ab 6= 0 and q2 6= 1. If the operator T is essentially
self-adjoint, then both resolvent relations (12) and (13) hold on H for λ = a¯q1/2,
µ = b¯q1/2 and the operator Rbq1/2 (B)Raq1/2 (A) is normal.
(cid:3)
A Resolvent Approach to the Real Quantum Plane
7
Proof. Setting
τ =
¯q1/2 − q1/2
2
ab,
the operator T can be rewritten as
T f = ¯q1/2(A − aq1/2)(B − bq1/2)f + τ f = q1/2(B − b¯q1/2)(A − a¯q1/2)f − τ f
for f ∈ D(T ). Therefore, since T is essentially self-adjoint and τ is purely imagi-
nary and nonzero (by the assumptions ab 6= 0 and q2 6= 1), the set
F0 := (T − τ I)D(T ) = (A − aq1/2)(B − bq1/2)D(T )
is dense in H. By Proposition 2,(i) and (iii), equation (7) is satisfied for λ = a¯q1/2,
µ = b¯q1/2 and all vectors h ∈ F0. Since F0 is dense and all resolvent operators are
bounded, equation (7) holds for all h ∈ H. That is, we have
Ra¯q1/2 (A)Rb¯q1/2 (B) = qRbq1/2 (B)Raq1/2 (A)
(17)
+ ab(q − 1)Ra¯q1/2 (A)Rb¯q1/2 (B)Rbq1/2 (B)Raq1/2 (A).
Thus, the (B, A)-resolvent relation (13) is satisfied.
Similarly, we conclude that
F1 := (T + τ I)D(T ) = (B − b¯q1/2)(A − a¯q1/2)D(T )
is dense in H and equation (4) holds for λ = a¯q1/2, µ = b¯q1/2 and h ∈ F1 by
Proposition 1,(i) and (iii), and hence for all vectors h ∈ H. That is, the (A, B)-
resolvent relation (12) is valid and we have
Ra¯q1/2 (A)Rb¯q1/2 (B) = qRbq1/2 (B)Raq1/2 (A)
(18)
+ ab(q − 1)Rbq1/2 (B)Raq1/2 (A)Ra¯q1/2 (A)Rb¯q1/2 (B).
Comparing (17) and (18) we conclude that
Ra¯q1/2 (A)Rb¯q1/2 (B)Rbq1/2 (B)Raq1/2 (A) = Rbq1/2 (B)Raq1/2 (A)Ra¯q1/2 (A)Rb¯q1/2 (B)
which means that the operator Rbq1/2 (B)Raq1/2 (A) is normal.
(cid:3)
3. Operator-theoretic preliminaries
We denote by P = i d
dx the momentum operator and by Q = x the position
operator acting on the Hilbert space L2(R) with respect to the Lebesgue measure
on R. Fix β > 0.
Lemma 10. (i) Suppose that f (z) is a holomorphic function on the strip Iβ :=
{z ∈ C : 0 < Imz < β} such that
(19)
−∞ f (x + iy)2 dx < ∞.
0<y<β Z +∞
sup
Set fy(x) := f (x + iy). Then the limits f0 := limy↓0 fy(x) and fβ := limy↑β fy(x)
exist in L2(R) and we have f0 ∈ D(eβP ) and eβP f0 = fβ.
such that f0 := limy↓0 fy(x) in L2(R) and eβP f0 = fβ.
(ii) For each function f0 ∈ D(eβP ) there exists a unique function f as in (i)
Proof. [12, Lemma 1.1].
(cid:3)
8
VASYL OSTROVSKYI, KONRAD SCHM UDGEN
If f is a function as in Lemma 10(i), we write simply f (x) for f0(x) and f (x+ iβ)
for fβ(x). Then the operator eβP acts by
(20)
(eβP )(x) = f (x + iβ), f ∈ D(eβP ).
For a nonzero complex number q we denote by S(q)+ the closed sector in the
plane with opening angle less than π between the positive x-axis and the half-line
through the origin and q and set S(q) := S(q)+ ∪ (−S(q)+).
q
¯q S(q)+
S(q)
q
¯q
Figure 1. The sectors S(q)+ and S(q)
We fix two reals α, β such that β > 0 and 0 < αβ < π. Put q := e−iαβ. Now
we define positive selfadjoint operators A and B on the Hilbert space L2(R) by
A := eαQ and B := eβP .
(21)
Corollary 11. If f ∈ D(BA) ∩ D(B), then f ∈ D(AB) and ABf = qBAf .
Proof. Using the description of the domain D(B) = D(eβP ) in Lemma 10 and
formula (20) we derive
(qBAf )(x) = qB(eαxf (x)) = e−αiβeα(x+iβ)f (x + iβ) = eαxf (x + iβ) = (ABf )(x).
(cid:3)
Clearly, the linear space
D0 = Lin {e−εx2+γx; ε > 0, γ ∈ C}
is contained in D(A) ∩ D(B) and it is invariant under A and B and also under the
Fourier transform and its inverse. By Corollary 11, we have D0 ⊆ Dq(A, B). As
noted in [12], D0 is a core for both selfadjoint operators A and B.
Proposition 12. Suppose that λ ∈ C\S(q)+. Then λ and λq are in ρ(A) and
qRλq(A)B ⊆ BRλ(A).
Proof. If z runs through the strip {z : 0 ≤ Im z ≤ β}, then the number eαz fills
the sector S(q)+. Hence the infimum of the function eαz − λ on the strip Iβ is
equal to the distance of λ from S(q)+. In particular, this infimum is positive, since
λ /∈ S(q)+.
Let f ∈ D(B) and let f (z) be the corresponding holomorphic function from
Lemma 10. Since eαz − λ has a positive infimum on the strip Iβ, the function
g(z) = (eαz − λ)−1f (z) is holomorphic on Iβ and it satisfies condition (19) as well,
because f does. Therefore, from Lemma 10 we conclude that g ∈ D(B) and
(BRλ(A)f )(x) = (Bg)(x) = g(x + iβ) = (eα(x+iβ) − λ)−1f (x + iβ)
= q(eαx − λq)−1f (x + iβ) = (qRλq(A)Bf )(x).
(cid:3)
A Resolvent Approach to the Real Quantum Plane
9
Another technical ingredient used below is Balakrishnan's theory of fractional
powers of nonnegative operators on Banach spaces [1], see e.g. [5].
Suppose that T is a closed linear operator on a Banach space such that
(−∞, 0) ⊆ ρ(T ) and sup {λ(T + λI)−1 : λ > 0} < ∞.
(22)
Then, for any γ ∈ C, 0 < Re γ < 1, the Balakrishnan operator J γ (see [5], p. 57)
is defined by
(23)
J γf =
sin(ε + it)π
π
Z ∞
0
λγ−1(T + λI)−1T f dλ,
f ∈ D(J γ) := D(T ).
Here the integral is meant as an improper Riemann integral of a continuous function
on (0, +∞) with values in the underlying Banach space. The operator J γ (or its
closure) is considered as a power of the operator T with exponent γ.
For our investigations the following special case is sufficient.
Proposition 13. Suppose that A is a positive self-adjoint operator on a Hilbert
space H such that ker A = {0} and let ϑ ∈ R, ϑ < π. Let T denote the normal
operator eiϑA in H. Then, for any 0 < ε < 1, t ∈ R and f ∈ D(T ) = D(A) we
have
(24)
T ε+itf = eiϑεe−ϑtAε+itf =
sin(ε + it)π
π
Z ∞
0
λε+it−1(T + λI)−1T f dλ,
where the operators T ε+it and Aε+it are defined by the spectral functional calculus.
Proof. Using that ϑ < π and A ≥ 0 it is easily verified that the operator T satisfies
the conditions stated in (22). Hence formula (23) for the Balakrishnan operator
J ε+it holds. For the normal operator T the closure of the operator J ε+it is just the
power T ε+it defined by the functional calculus (see Example 3.3.2 in [5]), where the
principal branch of the complex power has to be taken. Further, since ϑ < π, we
have T ε+it = eiϑεe−ϑtAε+it. Hence formula (24) follows from (23).
(cid:3)
Lemma 14. If A is a positive self-adjoint operator with trivial kernel, then
lim
ε→+0
Aεf = f
for
f ∈ D(A).
Proof. By the spectral calculus we have
kAεf − fk2 = Z ∞
0
λε − 12 dhE(λ)f, fi.
Passing to the limit ε → +0 and using Lebesgue's dominated convergence theorem
(by the assumption f ∈ D(A)) we obtain the assertion.
(cid:3)
4. Two Classes of Well-behaved Representations of Relation (2)
In this section we describe some well-behaved representations of relation (2). For
this we also restate some results from [12].
Recall that q = e−iθ0 and 0 < θ0 < π by (1). Set θ1 := θ0 − π if θ0 > 0,
θ1 := θ0 + π if θ0 < 0. Then we also have
−q = e−iθ1
and 0 < θ1 < π.
If A = 0 or if B = 0, then Dq(A, B) = D(B) resp. Dq(A, B) = D(A) and it is
obvious that the pair {A, B} satisfies the relation (2) and the resolvent relations
(3) and (4). We call pairs of the form {0, B} and {A, 0} trivial representations of
relation (2).
10
VASYL OSTROVSKYI, KONRAD SCHM UDGEN
Interesting representations of relation (2) are the classes C0 and C1 defined as
follows.
Definition 2. Suppose that ker A = ker B = {0}. We say that the pair {A, B} is
a representation of the class C0 if
(25)
and that the pair {A, B} is a representation of the class C1 if
AitB ⊆ eθ0tBAit , t ∈ R, and UAB ⊆ BUA.
AitB ⊆ eθ1tBAit , t ∈ R, and UAUB = −UBUA.
(26)
Definition 3. The trivial pairs {A2, 0}, {0, B2} and pairs {A0, B0} and {A1, B1}
of the classes C0 and C1, respectively, and orthogonal direct sum of such pairs are
called well-behaved representations of relation (2).
Remarks. 1. Note that the class C0 defined above is precisely the class C0 in
[12], while the class C1 according to Definition 2 corresponds to C1 if θ0 < 0 and to
C−1 if θ0 > 0 in [12].
2. Suppose that {A, B} is a well-behaved representation of relation (2). If A ≥ 0
and ker A = ker B = {0}, then UA = I and {A, B} is a pair of the class C0. Further,
if A ≥ 0, then the well-behaved representation {A, B} cannot have an orthogonal
summand of the class C1.
3. As it is usual for relations having unbounded operator representations there
are many "bad" unbounded representations of relation (2).
In [11] pairs of self-
adjoint operators A and B have been constructed for which Dq(A, B) is a core for
A and B, but the pair {A, B} is not a well-behaved representation of relation (2)
and it is not in one of classes Cn, n ∈ Z, defined in [12].
Let us describe all pairs of the classes C0 and C1 up to unitary equivalence. We
fix real numbers α, α1, β, β1, where β > 0, β1 > 0, such that
αβ = θ0
α1β1 = θ1, where
q = e−iθ0 and −q = e−iθ1.
Let u, v be two commuting self-adjoint unitaries on K. We define self-adjont
(27)
and
Let K be a Hilbert space.
operators A and B on the Hilbert space H = K ⊗ L2(R) a by
(28)
and self-adjoint operators A1 and B1 on the Hilbert space H1 = (K ⊕ K) ⊗ L2(R)
by the operator matrices
A0 = u ⊗ eαQ, B0 = v ⊗ eβP
(29)
A1 = (cid:18)eα1Q
0
−eα1Q(cid:19) , B1 = (cid:18) 0
eβ1P
eβ1P
0 (cid:19) .
Proposition 15. The pairs {A0, B0} and {A1, B1} belong to the classes C0 and
C1, respectively. Each pair of the class C0 resp. C1 is unitarily equivalent to a pair
{A0, B0} resp. {A1, B1} of the form (28) resp. (29).
Corollary 16. Up to unitary equivalence there are precisely five nontrivial ir-
reducible well-behaved representations of relation (2). These are the fours pairs
{A = ε1eαQ, B = ε2eβP} on L2(R), where ε1, ε2 ∈ {+1,−1}, and the pair {A, B}
on C2 ⊗ L2(R) given by (29) with K = C.
Any well-behaved representation {A, B} of relation (2) satisfying ker A = ker B =
{0} is a direct orthogonal sum of these representations.
A Resolvent Approach to the Real Quantum Plane
11
Corollary 17. Let {A, B} be a well-behaved representation of relation (2) for which
ker A = ker B = {0}. Then there is a linear subspace D ⊆ D(A) ∩ D(B) such that
(i) AD = D, BD = D, and AitD = D, BitD = D for t ∈ R,
(ii) D is a core for A and B,
(iii) ABf = BAf for f ∈ D.
Corollary 18. A pair {A, B} is a well-behaved representation (resp. of the class
C0 or C1) of relation (2) if and only if {B, A} is a well-behaved representation (resp.
of the class C0 or C1) of relation (8).
Proposition 15 and Corollaries 16 -- 18 are contained in [12, Section 2].
The next proposition is essentially used in the proofs of various theorems in
Section 5.
Proposition 19. Let k = 0, 1. Suppose that ker A = ker B = {0} and Dq(A, B)
is a core for B. If
for
t ∈ R,
AitB ⊆ eθktBAit
(30)
then {A, B} is a pair of the class Ck.
Proof. Putting qk := (−1)kq we have qk = eiθk. By (30), Proposition 2.3 in [12]
applies to the pair {A, B} and the relation AB = qkBA. Hence there exists
a linear subspace D of Dqk (A, B) such that D = AD is a core for B. Then
ABg = qkBAg for g ∈ D by the definition of Dqk (A, B). Since A is self-adjoint
and ker A = {0}, UA is self-adjoint unitary and A = AUA.
Let f ∈ Dq(A, B) and g ∈ D. Using the preceding facts we derive
hUAf, BAgi = hUAf, qkABgi = hf, qkUAABgi = hf, qkABgi = hqkAf, Bgi
= h(−1)kqBAf, gi = h(−1)kABf, gi = h(−1)kUABf,Agi.
Since D = AD is a core for B, from the preceding equality we conclude that
UAf ∈ D(B) and BUAf = (−1)kUABf for f ∈ Dq(A, B). By assumption Dq(A, B)
is a core for B, so the latter implies that UAB ⊆ (−1)kBUA.
Since UA is a self-adjoint unitary, we get UABUA ⊆ (−1)kB, that is, the self-
adjoint operator (−1)kB is an extension of the self-adjoint operator UABUA on
H. This is only possible if UABUA = (−1)kB. From the latter it follows that
UABUA = B and hence BUA = UAUABUA = BUA. Therefore,
UAUBB = UAB ⊆ (−1)kBUA = (−1)kUBBUA = (−1)kUBUAB.
Since ker B = {0}, the range of B is dense in H, so we get UAUB = (−1)kUAUB.
Thus, {A, B} is in Ck.
(cid:3)
(31)
Now let us return to the weak resolvent equations
qRλq(A)B ⊆ BRλ(A),
qRµq(B)A ⊆ ARµ(B).
(32)
For the trivial representations {0, B} resp. {A, 0} they are obviously fulfilled for
all λ 6= 0 resp. µ 6= 0. The classes C0 and C1 are treated in the next theorem.
Theorem 20. (i) If {A, B} is a pair of the class C0 for the relation (2), then (31)
and (32) are satisfied for all λ ∈ C\S(q) and µ ∈ C\S(q). If in addition A ≥ 0
resp. B ≥ 0, then (31) resp. and (32) holds for λ ∈ C\S(q)+ resp. µ ∈ C\S(q)+.
12
VASYL OSTROVSKYI, KONRAD SCHM UDGEN
(ii) If {A, B} is a pair of the class C1 for the relation (2), then (31) and (32) are
fulfilled for λ ∈ C\S(−q) and µ ∈ C\S(−q).
Proof. By Corollary 18 it suffices to prove all assertions for the first relation (31).
Clearly, (31) is preserved under orthogonal direct sums. Therefore, by Corollary
16, it is sufficient to prove (31) for the corresponding irreducible representations
listed in Corollary 16.
(i): Let {A0 = eαQ, B0 = eβP} and suppose that λ /∈ S(q)+. Then, by Propo-
sition 12, relation (31) is valid. Then, obviously, (31) holds also for the pair
{A0,−B0}. Since Rz(−A0) = −R−z(A0), it follows that (31) is satisfied for the
pairs {−A0,±B0} provided that λ /∈ −S(q)+.
(ii): We have to show that (31) holds for the pair {A1, B1} given by (29). Put
A := eα1Q, B := eα1P and q := −q. Then, since
(33)
Rz(A) = (cid:18)Rz(A)
0
0
Rz(−A)(cid:19) , B = (cid:18) 0 B
B 0(cid:19) ,
the weak resolvent relation qRλq(A)B ⊆ BRλ(A) reduces to the relations
qRλq(A)B ⊆ BRλ(−A),
qRλq(−A)B ⊆ BRλ(A).
Since Rz(−A) = −R−z(A) and q := −q, the latter equalities are equivalent to
qR(−λ)q(A)B ⊆ BR(−λ)(A),
qRλq(A)B ⊆ BRλ(A).
But these relations follow from Proposition 12, now applied to q := −q.
(cid:3)
5. Characterizations of Classes C0 and C1 by Weak Resolvent
Identities
Recall that A and B always denote self-adjoint operators acting on a Hilbert
space H and that the linear subspace Dq(A, B) was defined in Definition 1.
Let {A, B} be a pair of class C0. If λ > 0, then −λ is not in the sector S(q)+,
so the weak resolvent identity qR−λq(A)B ⊆ BR−λ(A) holds by Theorem 20(i).
Similarly, if µ ∈ Ri, µ 6= 0, and θ < π
2 , then µ /∈ S(q)+ and hence qRµq(A)B ⊆
BRµ(A). The following two theorems state some converses of these assertions.
Theorem 21. Let A is a positive operator such that ker A = {0}. Suppose that
0 < θ0 < π and the domain Dq(A, B) is a core for B. Assume that
qR−λq(A)B ⊆ BR−λ(A) for λ > 0.
Then the pair {A, B} is an orthogonal direct sum of a trivial representation {A2, 0}
and a pair {A0, B0} of the class C0.
Proof. Let f ∈ Dq(A, B). Cleary, the positive self-adjoint operator A and the nor-
mal operator qA (because of q = e−iθ0 with θ0 < π) satisfy the assumptions of
Proposition 13. By the definition of Dq(A, B), the vectors f and Bf are in D(A),
so formula (24) applies to the operator T = A and the vector Bf and also to the
operator T = qA and the vector f . The assumptions qR−λq(A)B ⊆ BR−λ(A) and
ABf = qBAf imply that
(qA + λI)−1(qA)Bf = B(A + λI)−1Af.
Next we apply Proposition 13 to the operator T = qA. Since q = eiθ0 with θ0 < π,
the assumptions of Proposition 13 are fulfilled. Interchanging the closed operator B
A Resolvent Approach to the Real Quantum Plane
13
and the integral in formula (24) (by considering the integral as a limit of H-valued
Riemann sums) we therefore obtain
(34)
(qA)ε+itBf = BAε+itf for f ∈ Dq(A, B), t ∈ R.
By the first equality in (24) and the relation Aε+itf = AεAitf we have
(qA)ε+itBf = (eiθ0A)ε+itBf = eiεθ0e−θ0t AitAεBf,
so by (34) we obtain
eiεθ0e−θ0t AitAεBf = BAεAitf.
(35)
Recall that f ∈ D(A) and AitBf ∈ D(A) by the assumption f ∈ Dq(A, B). Passing
to the limit ε → +0 in (35) by using Lemma 14 and the fact that the operator B
is closed it follows that AitBf = eθ0tB Aitf for all f ∈ Dq(A, B). Since Dq(A, B)
is a core for B by assumption, we conclude that AitB ⊆ eθtBAit for all t ∈ R. The
latter implies that Ait leaves the closed subspace H2 := ker B invariant. Hence H2
is reducing for A and B, so we have B = 0⊕ B0 and A = A2 ⊕ A0 on H = H2 ⊕H⊥
such that ker B0 = {0} and (A0)itB0 ⊆ eθ0tB0(A0)it for t ∈ R. Since A ≥ 0 and
ker A = {0}, the pair {A0, B0} belongs to C0.
Theorem 22. Suppose that 0 < θ0 < π
Dq2 (A2, B) is a core for B and
2 and ker A = {0}. Assume that
2
(cid:3)
qRµiq(A)B ⊆ BRµi(A) for µ ∈ R, µ 6= 0.
Then {A, B} is an orthogonal sum of a trivial representation {A2, 0} and a pair
{A0, B0} such that {A0, B0} belongs to the class C0. If in addition Dq(A, B) is a
core for B, then {A0, B0} is a pair of the class C0.
Proof. From the relation qRµiq(A)B ⊆ BRµi(A) it follows that each resolvent
Rµi(A) and its adjoint R−µi(A) leaves the closed linear subspace H2 := ker B
invariant. This implies that H2 is a reducing subspace for Rµi(A) and therefore for
the operator A. Obviously, H2 reduces B. Hence the pair {A, B} is an orthogonal
sum of a trivial representation {A2, 0} and a pair {A0, B0} such that ker B0 = {0}.
For notational simplicity let us assume already that ker B = {0}. Our aim is to
prove that {A, B} is in C0.
First we recall a simple operator-theoretic fact: If T is a closed operator such
that ν,−ν ∈ ρ(T ), then ν2 ∈ ρ(T 2) and
(36)
Rν2 (T 2) = Rν(T )R−ν(T ) =
1
2ν
(Rν (T ) − R−ν(T )).
Suppose now that λ > 0. Putting µ = √λ, we have (iµ)2 = −λ. Let f ∈ D(B).
Using the identity (36) twice, for ν = µiq and for ν = µi, and the assumptions
qR±µiq(A)B ⊆ BR±µi(A), we obtain
q2R−λq2 (A2)Bf =
(37)
=
q2
2µiq
1
2µi
(Rµiq(A) − R−µiq(A))Bf
B(Rµi(A) − R−µi(A))f = BR−λ(A2)f.
Thus, since q2 = e−2iθ0, 2θ0 < π and Dq2 (A2, B) is a core for B by assumption,
the pair {A2, B} satisfies all assumptions of Theorem 21 for the relation A2B =
q2BA2. Therefore, by this theorem we have (A2)itB ⊆ e2θ0tB(A2)it for t ∈ R, so
that AisB ⊆ eθ0sBAis for s ∈ R. Hence the pair {A, B} belongs to C0 (see e.g.
14
VASYL OSTROVSKYI, KONRAD SCHM UDGEN
Remark 2 in Section 4). Further, if in addition Dq(A, B) is a core for B, it follows
from Proposition 19 that {A, B} is in the class C0.
(cid:3)
The next theorem contains a characterization of the class C1. Recall that for the
class C1 the weak resolvent relation (31) holds for λ ∈ C\S(−q) by Theorem 20(ii).
Theorem 23. Suppose that 0 < θ0 < π and ker A = ker B = {0}. Assume that
both domains Dq2 (A2, B) and Dq(A, B) are cores for the operator B and that there
exists a number p ∈ C\S(−q) such that
qRµpq(A)B ⊆ BRµp(A)
for all µ ∈ R, µ 6= 0.
Then the pair {A, B} belongs to the class C1.
Proof. Without loss of generality we can choose the number p ∈ C\S(−q) of modu-
lus one and contained in the open sector with angle less than π between the positive
x-axis and the half-line through the origin and q. We modify some arguments that
have been used already in the proofs of Theorems 21 and 22.
−¯q
−q
S(−q)
q
p
¯q
S(−q)
q
¯q
S(−q)
−¯q
S(−q)
p
−q
Figure 2. The S(−q) sector and admissible ranges for p as
0 < arg q < π
2 < arg q < π
2 and as π
Using the identity (36) the assumption qRµpq(A)B ⊆ BRµp(A) implies that
q2Rλp2q2 (A2)B ⊆ BRλp2 (A2), λ > 0.
(38)
Let f ∈ Dq2 (A2, B). From (38) and the relation A2Bf = q2BA2f we derive
(39)
(−p2q2A2 + λI)−1(−p2q2A2)Bf = B(−p2A2 + λI)−1(−p2A2)f, λ > 0.
Our aim is to apply Lemma 13, especially formula (24) therein, to the operators
−p2q2A2 and −p2A2. To fullfill the assumptions of Lemma 13 it is crucial to write
the number −p2q2 and −p2 (of modulus one) in the form eiϑ with ϑ ∈ (−π, π).
Let us write p as p = eiψ with 0 < ψ < π and let s(ψ) denote the sign of ψ. Note
that q = e−iθ and θ < π. Also we recall that by definition we have θ1 = θ − π if
θ > 0 and θ1 = θ + π if θ < 0. We shall prove that
(40)
(41)
−p2 = ei(−2ψ+s(ψ)π)
−p2q2 = ei(2θ1−2ψ+s(ψ)π)
and − 2ψ + s(ψ)π ∈ (−π, π),
and 2θ1 − 2ψ + s(ψ)π ∈ (−π, π).
First suppose that ψ > 0. Then π > θ > ψ > 0 by the choice of the number p.
Hence −2ψ + π ∈ (−π, π) and −p2 = −(e−iψ)2 = ei(−2ψ+π). Further,
2θ1 − 2ψ + π = 2(θ − π) − 2ψ + π = 2(θ − ψ) − π ∈ (−π, π)
and −p2q2 = −(e−iψ)2(eiθ)2 = ei(2θ−2ψ−π) = ei(2θ1−2ψ+π).
A Resolvent Approach to the Real Quantum Plane
15
Next we treat the case ψ < 0. Then 0 > ψ > θ > −π by the definition of p.
Therefore, −2ψ − π ∈ (−π, π) and −p2 = −(e−iψ)2 = ei(−2ψ−π). Moreover,
2θ1 − 2ψ − π = 2(θ + π) − 2ψ − π = 2(θ − ψ) + π ∈ (−π, π)
and −p2q2 = −(e−iψ)2(eiθ)2 = ei(2θ−2ψ+π) = ei(2θ1−2ψ−π). This proves (40) and
(41) in both cases.
By (40) and (41) it follows from the first equality of formula (24) that
(42)
(−p2A2)ε+itg = ei(−2ψ+s(ψ)π)ε e−(−2ψ+s(ψ)π)t (A2)ε+itg,
(−p2q2A2)ε+itg = ei(2θ1−2ψ+s(ψ)π)ε e−(2θ1−2ψ+s(ψ)π)t (A2)ε+itg
(43)
for any g ∈ D(A2). Since f, Bf ∈ D(A2), the second equality of formula (24) yields
(44)
(−p2q2A2)ε+itBf = B(−p2A2)ε+itf.
e−2θ1t(A2)itBf = B(A2)itf, t ∈ R,
Inserting (42) with g = f and (43) with g = Bf into (44) and passing to the limit
ε → +0 we obtain
and hence AisBf = eθ1sBAisf , s ∈ R, for all f ∈ Dq2 (A2, B). Since Dq2 (A2, B)
is a core for B by assumption, the latter implies that AisB ⊆ eθ1sBAis for s ∈ R.
Therefore, since we also assumed that Dq(A, B) is a core for B, the assumptions
of Proposition 19 are fulfilled with k = 1, so the pair {A, B} belongs to the class
C1.
(cid:3)
Related characterizations of the class C1 can be also given by requiring the weak
resolvent relation for A rather than A. The following theorem is a sample of such
a result.
Theorem 24. Suppose that 0 < θ0 < π and ker A = ker B = {0}. Suppose that
the domains Dq(A, B) and D−q(A, B) are cores for B and
−qRλq(A)B ⊆ BR−λ(A)
for λ > 0.
Then the pair {A, B} is in C1.
Proof. The assumptions of Theorem 24 imply that the pair {A, B} satisfies the
assumptions of Theorem 21 for the relation (2) with q = e−iθ0 replaced by −q =
e−iθ1. Note that 0 < θ1 < π, since we assumed that 0 < θ0 < π. Therefore, by
Theorem 21, we have AitB ⊆ eθ1tBAit for all t ∈ R. Since Dq(A, B) is a core for
B, it follows from Proposition 19 that the pair {A, B} is in C1.
(cid:3)
The crucial assumption in the preceding Theorems 21, 22, and 23 is that the weak
A-resolvent identity (9) holds on some line that intersects the critical sector only at
the origin. In addition there have been technical assumptions such as ker A = {0}
and the requirement that Dq(A, B) resp. Dq2 (A2, B) is a core for the operator B.
These technical assumptions can be avoided if we assume in addition that the weak
B-resolvent identity (10) holds for some points of the resolvent set ρ(B).
Theorem 25. Let A is a positive operator. Suppose that 0 < θ0 < π and there
exist a number ν ∈ ρ(B) such that νq, ν q ∈ ρ(B),
(45)
q Rνq(B)A ⊆ ARν(B)
and q Rν q(B)A ⊆ ARν (B).
Assume that
(46)
qR−λq(A)B ⊆ BR−λ(A)
for all λ > 0.
16
VASYL OSTROVSKYI, KONRAD SCHM UDGEN
Then the pair {A, B} is an orthogonal direct sum of trivial representations {A2, 0}
and {0, B2} and a pair {A0, B0} of the class C0.
Proof. From (45) it follows that the operator Rν (B) and its adjoint Rν (B) leave
the closed subspace G2 := ker A invariant. Therefore, G2 is reducing for Rν (B)
and hence for B. Since G2 ≡ ker A is obviously reducing for A, the pair {A, B}
decomposes as an orthogonal sum of pairs {0, B2} and { A, B} such that ker A =
{0}. Further, by Proposition 5(vi), (45) implies that Dq(A, B) is core for B. Hence
Dq( A, B) is core for B. Clearly, (46) leads to the relation R−λq( A) B ⊆ BR−λ( A)
for λ > 0. Thus, the pair { A, B} satisfies the assumptions of Theorem 21 which
gives the assertion.
Theorem 26. Suppose that 0 < θ0 < π
that νq, ν q, ν q2 ∈ ρ(B) and
(47)
2 and there exists a number ν ∈ ρ(B) such
for λ = νq , ν, ν.
(cid:3)
Assume that
q Rλq(B)A ⊆ ARλ(B)
qRµiq(A)B ⊆ BRµi(A) for all µ ∈ R, µ 6= 0.
Then {A, B} is an orthogonal sum of trivial representations and a pair {A0, B0}
belonging to the class C0.
Proof. Arguing in a similar manner as in the proof of Theorem 25 the assertion
is reduced to Theorem 22. We sketch only the necessary modifications: From the
relations (47), applied for λ = ν, ν, it follows that ker A is reducing for the pair
{A, B}. By Proposition 5(vi) and Corollary 7, the relations (47), applied with
λ = ν, νq, imply that the domains Dq(A, B) and Dq2 (A2, B) are cores for B. (cid:3)
Theorem 27. Suppose that π
2 < θ0 < π and there exist numbers ν ∈ ρ(B) and
p ∈ C\S(−q) such that νq, ν q, ν q2 ∈ ρ(B),
(48)
for λ = νq , ν, ν,
for all µ ∈ R, µ 6= 0.
(50)
Then {A, B} is an orthogonal sum of trivial representations and a pair {A1, B1} of
the class C1.
Proof. As in the proof of Theorem 26 it follows from (48) that ker A is reducing
for the pair {A, B} and that Dq(A, B) and Dq2(A2, B) are cores for B. Likewise
the relations qRp q(A)B ⊆ BRp(A) and qRpq(A)B ⊆ BRp(A) (by (49) and (50))
imply that ker B is reducing for the pair {A, B}. Using these facts the assertion
is derived from Theorem 23.
(cid:3)
Remarks. By Theorem 20, the weak B-resolvent relation qRµq(B)A ⊆ ARµ(B)
is satisfied for a pair {A, B} of the class C0 if µ ∈ C\S(q) and for a pair {A, B}
in C1 if µ ∈ C\S(−q). From this result it follows that for the corresponding pairs
in Theorems 25 -- 27 the assumptions (45), (47), and (48) can be fulfilled. That is,
if {A, B} is a pair of the class C0 we can choose ν ∈ C\S(q) such that ν ∈ C\S(q)
(then (45) holds) and if in addition 0 < θ0 < π
2 there exists ν ∈ C\S(q) such that
νq, ν ∈ C\S(q) (which implies (47)). If the pair {A, B} is in C1 and π
2 < θ0 < π
we can find ν ∈ C\S(−q) such that νq , ν ∈ C\S(−q) (these conditions imply (48))
(49)
q Rλq(B)A ⊆ ARλ(B)
qRp q(A)B ⊆ BRp(A),
qRµpq(A)B ⊆ BRµp(A)
A Resolvent Approach to the Real Quantum Plane
17
and there exists p ∈ C\S(−q) such that p ∈ C\S(−q) (then (46) and (50) are
satisfied by Theorem 20(ii)).
6. Deficiency subspaces and their dimensions
Let A and B be positive self-adjoint operators with trivial kernels acting on a
Hilbert space H. Then, by Theorems 20 and 21 and by Proposition 5, the pair
{A, B} belongs to the class C0 of representations of equation (2) if and only if the
resolvent relations (12) and (13) are satisfied for all λ, µ ∈ C such that λ, µ /∈ S(q)+.
Moreover, up to unitary equivalence, the only irreducible such pair is {eαQ, eβP},
where αβ < π and q = e−iαβ; see Section 3. In this section we study the resolvent
relations for the pair {eαQ, eβP} in the general case, that is, we do not assume that
αβ < π or λ, µ /∈ S(q)+.
First let us fix some assumptions and notations that will be kept throughout this
section. Suppose that α, β ∈ R and q := e−iαβ 6= ±1. We consider the pair
{A := eαQ, B := eβP}
of self-adjoint operators on the Hilbert space H := L2(R). Recall that A and B act
by
(51)
(Af )(x) = eαxf (x)
and (Bf )(x) = f (x + iβ)
for all functions f of the dense domain
D0 = Lin {e−εx2+γx; ε > 0, γ ∈ C}.
Since (2) is satisfied for f ∈ D0 (by (51)), we know from Section 2 that (4) holds
for all h ∈ (B− µI)(A− λI)D0 and that (7) holds for all h ∈ (A− λqI)(B− µqI)D0.
Therefore, in order to find the resolvent equations in the present case it suffices to
describe the form of resolvent equations on the orthogonal complements of spaces
(B − µI)(A− λI)D0 and (A− λqI)(B − µqI)D0, respectively. This will be achieved
by the formulas at the end of this section.
Definition 4.
HA(λ, µ) = {ψ ∈ H : ψ ⊥ (B − µI)(A − λI)D0},
HB(λ, µ) = {η ∈ H : η ⊥ (A − λqI)(B − µqI)D0}.
Assume that λ, µ ∈ C r (R+ ∪ ¯qR+), where R+ = [0, +∞), and write
λ = er+is, µ = eu+iv,
r > 0, u > 0, s < π, v < π.
Further, we let
so that q = e−iθ0, and set
αβ = θ0 + 2πm, m ∈ Z, θ0 ∈ (−π, π),
ε1 = sign(v), ε2 = sign(v − θ0), ε3 = sign(s), ε4 = sign(s − θ0).
Theorem 28. (i) The vector space HA(λ, µ) is spanned by the functions
ψj(x) =
¯µ−ix/βeπ(1+ε1)x/β
e2πx/β − ¯λ2π/αβ e−4iπ2j/αβ
,
where j ∈ Z and 0 < (s − 2πj)/αβ < 1, and its dimension is
ε3 − ε4
(cid:17).
dimHA(λ, µ) = sign(αβ)(cid:16)m +
2
18
VASYL OSTROVSKYI, KONRAD SCHM UDGEN
(ii) The space HB(λ, µ) is the linear span of functions
ηk(x) =
(¯µ¯q)−ix/βe2πkx/β
.
eαx − ¯λ¯q
where k ∈ Z and 0 < (θ0 + 2πk − v)/αβ < 1, and it has the dimension
dimHB(λ, µ) = sign(αβ)(cid:16)m +
ε1 − ε2
2
(cid:17).
Proof. (i): First we study the space HB(λ, µ). Let η ∈ HB(λ, µ). Applying (51) to
φ(x) = e−εx2+itx ∈ D0 we compute
((A − λq)(B − µq)φ)(x) = (eαx − λq)(e−ε(x+iβ)2+it(x+iβ) − µqe−εx2+itx)
= (eαx − λq)(e−2iβεx+εβ2−tβ − µq)e−εx2+itx.
Therefore, since η ⊥ (A − λq)(B − µq)φ(x), we obtain
η(x) (eαx − λq)(e−2iβεx+εβ2−tβ − µq)e−εx2+itx dx
0 = ZR
= eεβ2−tβ ZR
− µqZR
ei(t−2βε)x η(x) (eαx − λq) e−εx2
eitx η(x) (eαx − λq) e−εx2
dx
dx
which implies that
(52)
eεβ2−tβgε(t − 2βε) = µq gε(t),
gε(t) = ZR
eitx η(x) (eαx − λq) e−εx2
dx.
Since µ = eu+iv, v < π, it is easily seen that the function
Gε(t) = e− t2
4ε − u+i(v−θ0 )
2εβ
t
satisfies (52). Therefore, any solution of (52) has the form Gε(t)Hε(t), where Hε is
a periodic function on R with period −2εβ, that is, Hε(t − 2εβ) = Hε(t) for t ∈ R.
The crucial step of this proof is contained in the following lemma.
Lemma 29. Hε is a trigonometric polynomial, that is, there are an integer l ≥ 0
and numbers ck ∈ C, k = −l, . . . , l such that Hε(t) = Pl
k=−l dke
iπkt
εβ .
Proof. We have
(53) Hε(t) = (Gε(t))−1gε(t) = e
t2
4ε + u+i(v−θ0 )
2εβ
tZR
eitx η(x) (eαx − λq) e−εx2
dx.
Because the above arguments are valid for complex numbers t as well, Hε becomes
a periodic function on a whole complex plane C.
is also in L2(R) for any
τ > 0. Therefore, the integral in (53) is an entire function on the complex plane C.
Since η ∈ L2(R), the function eτ xη(x) (eαx − λq) e−εx2
We show that Hε is of exponential type. For any s ∈ R we have
Hε(t + is) = e
4ε + u+i(v−θ0)
(t+is)2
2εβ
t2
−s2
4ε + ut−(v−θ0 )s
2εβ
t2
−s2
4ε + ut−(v−θ0 )s
2εβ
= e
≤ e
ZR
(t+is)(cid:12)(cid:12)(cid:12)
ei(t+is)x η(x) (eαx − λq) e−εx2
ZR
(cid:12)(cid:12)(cid:12)
eitx η(x) (eαx − λq) e−εx2−sx dx(cid:12)(cid:12)(cid:12)
kηk(cid:16)ZR eαx − λq2 e−2εx2−2sx dx(cid:17)1/2
.
dx(cid:12)(cid:12)(cid:12)
A Resolvent Approach to the Real Quantum Plane
19
Since
ZR eαx − λq2 e−2εx2−2sx dx = r π
2ε
e
for 0 ≤ t ≤ 2εβ we have
s2
2ε(cid:0)λ2 + e
α(α−2s)
2ε − (λq + ¯λ¯q)e
α(α−4s)
8ε
(cid:1),
Hε(t + is) ≤ M e− (v−θ0 )s
2εβ kηk(cid:16)r π
2ε(cid:0)λ2 + e
α(α−2s)
2ε − (λq + ¯λ¯q)e
α(α−4s)
8ε
(cid:1)(cid:17)1/2
,
t2
4ε + ut
2εβ on the interval [0, 2εβ]. Since Hε(t + is)
where M is the supremum of e
is periodic in t, the latter estimate holds for all t ∈ R. Hence Hε(t + is) has
exponential growth. Thus, Hε is an entire periodic function of exponential type.
By a result from complex analysis (see, e.g., [8, p. 334]), such a function Hε has to
be a trigonometric polynomial
Hε(t) =
l
Xk=−l
cke
iπkt
βε .
(cid:3)
By Lemma 29 the equality gε(t) = Gε(t)Hε(t) takes the form
ZR
eitx η(x) (eαx − λq) e−εx2
dx = e− t2
4ε − u+i(v−θ0 )
2εβ
l
t
Xk=−l
cke
iπkt
βε .
Applying the Fourier transform, we obtain
η(x) (eαx − λq) e−εx2
e−itxgε(t) dt
1
=
2π ZR
= e−εx2
eixu/βe−(v−θ0)x/β
l
Xk=−l
dke2πkx/β,
where
Obviously, the factor e−εx2
dk =
cke(u+i(v−θ0−2kπ))2/(4εβ2).
√ε
√π
cancels, so we get
eixu/βe−(v−θ0)x/β
(54)
η(x) =
eαx − λq
Let us introduce the functions
l
Xk=−l
dke
2πkx
β =
(µq)ix/β
eαx − λq
l
Xk=−l
dke2πkx/β.
(¯µ¯q)−ix/βe2πkx/β
e−iux/βe(θ0−v+2πk)x/β
ηk(x) :=
=
eαx − ¯λ¯q
.
eαx − ¯λ¯q
¯dkηk(x).
k=−l
Then we have η(x) = Pl
Next we want to decide which functions ηk belong to L2(R). Let us begin with
the case where α > 0. Then the function ηk(x) behaves like e(θ0−v+2πk)x/β as
x → −∞ and like e(θ0−v+2πk−αβ)x/β as x → +∞, so that ηk ∈ L2(R) if and only
0 < (θ0 − v + 2πk)/β < α. In the case where α < 0 a similar reasoning shows that
ηk ∈ L2(R) if and only α < (θ0 − v + 2πk)/β < 0. Thus, in both cases we have
ηk ∈ L2(R) if and only if if and only if 0 < (θ0 − v + 2πk)/αβ < 1, that is,
v − θ0 < 2πk < v + 2πm for αβ > 0,
for αβ < 0,
v + 2πm < 2πk < v − θ0
20
VASYL OSTROVSKYI, KONRAD SCHM UDGEN
or equivalently,
1 + ε2
2 ≤ k ≤ m −
1 − ε1
2
for αβ > 0, m +
1 + ε1
2 ≤ k ≤
ε2 − 1
2
for αβ < 0.
The functions ηk are obviously linearly independent. From the asymptotic be-
¯dkηk(x) is in L2(R) if and only each ηk
haviour of ηk it follows that η(x) = Pl
with nonvanishing coefficient ¯dk is in L2(R). Therefore, we obtain
k=−l
dimHB = sign(αβ)(cid:16)m +
ε1 − ε2
2
(cid:17).
(ii): Now we turn to the space HA(λ, µ). Let ψ ∈ L2(R). Recall that ψ ∈
HA(λ, µ) if and only if
ψ ⊥ (B − µI)(A − λI)D0 = (eβP − µI)(eαQ − λI)D0
For the Fourier transform F we have eαQ = F e−αPF −1 and eβP = F eβQF −1.
Hence the latter is equivalent to
ψ ⊥ F (eβQ − µI)(e−αP − λI)F ∗D0.
Since D0 is invariant under the Fourier transform, ψ ∈ HA(λ, µ) if and only if the
inverse Fourier transform F ∗ψ of ψ satisfies
F ∗ψ ⊥ (eβQ − µI)(e−αP − λI)D0.
From the proof of (i) we already know that ψ ∈ HA(λ, µ) if and only if F ∗ψ is a
linear combination of functions
where k ∈ Z and 0 < (s − 2πk)/αβ < 1. The condition on k can be rewritten as
φk(x) =
(¯λ)ix/αe−2πkx/α
,
eβx − ¯µ
1 + ε4
2 − m ≤ k ≤
2 ≤ k ≤
ε4 − 1
1 + ε3
ε3 − 1
2
2 − m for αβ < 0.
for αβ > 0,
Therefore, we have
dimHA = sign(αβ)(cid:16)m +
ε3 − ε4
2
(cid:17).
To calculate the Fourier transform of φk, we shall apply the following formula
(see, e.g., [4, 3.311.9])
(55)
ZR
e−δx
e−x + γ
dx =
πγδ−1
sin πδ
,
arg γ < π, 0 < Re δ < 1.
After some computations using (55) we obtain F φk(x) = Ckψk(x), where
i√2πei(u−iv−iπ(1−ε1))(r−is+i2πk)/αβ
Ck =
ψk(x) =
¯µβ
¯µ−it/βeπ(1+ε1)t/β
(e2πt/β − ¯λ2π/αβe4iπ2k/αβ )
.
,
(cid:3)
The following corollary restates the result on the dimensions of Theorem 28 in
an important special case.
A Resolvent Approach to the Real Quantum Plane
21
Corollary 30. Suppose that {A, B} is an irreducible nontrivial representation of
the class C0 for relation (2). Then
dimHA(λ, µ) = 0
dimHB(λ, µ) = 0
for λ /∈ S(q)+,
for µ /∈ S(q)+,
dimHA(λ, µ) = 1
dim HB(λ, µ) = 1
for λ ∈ S(q)+,
for µ ∈ S(q)+.
Proof. By Corollary 16, {A, B} is unitarily equivalent to a pair {δ1eαP , δ2eβQ},
where δ1, δ2 ∈ {1,−1} and αβ = θ0, θ0 < π. Hence the assertion follows at once
from Theorem 28.
(cid:3)
We now continue the considerations towards the modified resolvent relations.
Proposition 31. (i) R¯µ(B)R¯λ(A) −
(ii) R¯λ¯q(A)R¯µ ¯q(B) +
1
¯q¯λ¯µ(1−¯q) I maps HA(λ, µ) into HB(λ, µ).
1
¯λ¯µ(1−¯q) I maps HB(λ, µ) into HA(λ, µ).
Proof. (i): Indeed, for φ ∈ D0 we have
(B − µI)(A − λI)φ = ¯q(A − λqI)(B − µqI)φ + λµ(1 − q)φ.
Therefore, for η ∈ HB(λ, µ) we derive
1
(cid:10)(cid:0)R¯µ(B)R¯λ(A) −
¯λ¯µ(1 − ¯q)
= hR¯µ(B)R¯λ(A)η, (B − µI)(A − λI)φi −
I(cid:1)η, (B − µI)(A − λI)φ(cid:11)
¯λ¯µ(1 − ¯q) hη, (B − µI)(A − λI)φi
1
q
= −
¯λ¯µ(1 − ¯q) hη, (A − λqI)(B − µqI)φi = 0,
¯λ¯µ(1−¯q) I(cid:1)η ∈ HA(λ, µ).
1
that is, (cid:0)R¯µ(B)R¯λ(A) −
(ii) is proved in a similar manner.
From Proposition 31 it follows that the operator
(56)
(cid:16)R¯µ(B)R¯λ(A) −
1
¯λ¯µ(1 − ¯q)
I(cid:17)(cid:16)R¯λ¯q(A)R¯µ¯q(B) +
1
¯q¯λ¯µ(1 − ¯q)
maps the subspace HA(λ, µ) into itself and that the operator
I(cid:17)(cid:16)R¯µ(B)R¯λ(A) −
(cid:16)R¯λ¯q(A)R¯µ¯q(B) +
(57)
1
¯q¯λ¯µ(1 − ¯q)
1
¯λ¯µ(1 − ¯q)
(cid:3)
I(cid:17)
I(cid:17)
maps HB(λ, µ) into itself. In the special case when HA(λ, µ) = HB(λ, µ) = {0}
we know from Section 2 that the corresponding operators in (56) and (57) are both
equal to −q¯λ−2 ¯µ−2(1 − ¯q)−2 I on the whole Hilbert space.
22
VASYL OSTROVSKYI, KONRAD SCHM UDGEN
In the general case some lengthy but straightforward computations using (55)
and the relations eαQ = F e−αPF −1, eβP = F eβQF −1, lead to the formulas
(cid:16)R¯λ¯q(A)R¯µ¯q(B) +
I(cid:17)ψj(x)
1
¯λ¯µ¯q(1 − ¯q)
¯λ2π(m−(1−ε1)/2)/αβe−4iπ2(m−(1−ε1)/2)j/αβ
=
m−(1−ε1)/2
Xl=(1+ε2)/2
¯λ−2πl/αβ e4iπ2lj/αβ ηl(x),
¯λ¯q(1 − ¯q)¯µ
1
(cid:16)R¯µ(B)R¯λ(A) −
= −2π¯λ−2π(m−k+(ε1−1)/2)/αβ
¯λ¯µ(1 − ¯q)
I(cid:17)ηk(x)
αβ¯λ¯µ(1 − ¯q)
Therefore, we finally obtain
m−(ε4+1)/2
Xj=(1−ε3)/2
e4iπ2j(m−k+(ε1−1)/2)/αβψj(x).
1
1
=
=
(cid:16)R¯µ(B)R¯λ(A) −
1
¯λ¯µ(1 − ¯q)
−2π
αβ¯λ2 ¯µ2 ¯q(1 − ¯q)2
I(cid:17)(cid:16)R¯λ¯q(A)R¯µ¯q(B) +
¯λ¯µ¯q(1 − ¯q)
I(cid:17)ψj(x)
m−1−(ε2−ε1)/2
m−(ε4+1)/2
Xk=0
Xl=(1−ε3)/2
e−4iπ2kj/αβ e4iπ2lk/αβ ψl(x),
(cid:16)R¯λ¯q(A)R¯µ¯q(B) +
1
¯λ¯µ¯q(1 − ¯q)
−2π
m−(ε4+1)/2
I(cid:17)(cid:16)R¯µ(B)R¯λ(A) −
Xl=(1+ε2)/2
Xj=(1−ε3)/2
m−(1−ε1)/2
αβ¯λ2 ¯µ2 ¯q(1 − ¯q)2
I(cid:17)ηk
¯λ¯µ(1 − ¯q)
e−4iπ2j(k−l)/αβ ¯λ2π(k−l)/αβ ηl(x).
The preceding two equations are the versions of the resolvent equations for basis
elements of the subspaces HA(λ, µ) and HB(λ, µ), respectively.
Acknowledgments
The results of this paper were obtained during research visits of the first author
at Leipzig University supported by the DFG grant SCHM 1009/5-1. Excellent
working conditions and warm hospitality are acknowledged. The authors express
their gratitude to Prof. Lyudmila Turowska and Prof. Andrew Bakan for helpful
discussions of topics related to this research.
References
[1] Balakrisnan, A.V., Fractional powers of closed operators and the semigroup generated by
them, Pacific J. Math. 10(1960),419 -- 437.
[2] Brooke, J. A., Busch, P. and Pearson, D. B., Commutativity up to a factor of bounded
operators in complex Hilbert space, R. Soc. Lond. Proc. Ser. A Math. Phys. Eng. Sci. 458
(2002), 109 -- 118.
[3] Cho, M., Harte, R. and Ota, Sch., Commutativity to within scalars on Banach space, Funct.
Anal. Approx. 3 (2011), 69 -- 77.
[4] Gradshteyn, I.S. and Ryzhik, I.M. Table of Integrals, Series and Products. 7-th ed. Acad.
Press, 2007.
[5] Martinez Carracedo, C. and Sanz Alix, M., The Theory of Fractional Powers of Operators,
North-Holland, Amsterdam, 2001.
[6] Mortad, M.H., Commutativity up to a factor: more results and the unbounded case. Z. Anal.
Anwend. 29(2010), 303-307.
[7] Klimyk, A. and Schmudgen, K., Quantum Qroups and Their Representations, Springer, 1997.
A Resolvent Approach to the Real Quantum Plane
23
[8] A.V.Markushevich, The theory of analytic functions. Vol. II. (Russian)
[9] Ostrovskyi, V. and Samoilenko, Yu., Introduction to the Theory of Representations of Finitely
Presented ∗-Algebras, Gordon and Breach, London, 1999.
[10] Ota, S. and Szafraniec, F.H., Notes on q-normal operators, Studia Math. 165(2004), 295 -- 301.
[11] Schmudgen, K.: Operator representations of R2
q, Publ. RIMS Kyoto Univ. 29(1992), 1030 --
1061.
[12] Schmudgen, K.: Integrable representations of R2
q, Xq,γ , and SLq(2, R), Commun. Math.
Phys. 159(1994), 27 -- 237.
[13] Schmudgen, K.: The quantum quarter plane and the real quantum plane, Inter. J. Math.
13(2002), 279 -- 321.
[14] Woronowicz, S. L. and Zakrzewski, S., Quantum 'ax + b' group. Rev. Math. Phys. 14 (2002),
797 -- 828.
[15] Yang, J. and Du, H.-P., A note on commutativity up to a factor of bounded operators, Proc.
Amer. Math. soc. 132 (2004), 1713 -- 1720.
Institute of Mathematics, Ukrainian Academy of Sciences, Tereshchenkivska 3, Kyiv
01601, Ukraine
E-mail address: [email protected]
Mathematisches Institut, Universitat Leipzig, Augustusplatz 9/10, 04109 Leipzig,
Germany
E-mail address: [email protected]
|
1811.06923 | 1 | 1811 | 2018-11-16T17:09:12 | Constructing KMS states from infinite-dimensional spectral triples | [
"math.OA",
"math.DS",
"math.KT",
"math.SP"
] | We construct KMS-states from $\mathrm{Li}_1$-summable semifinite spectral triples and show that in several important examples the construction coincides with well-known direct constructions of KMS-states for naturally defined flows. Under further summability assumptions the constructed KMS-state can be computed in terms of Dixmier traces. For closed manifolds, we recover the ordinary Lebesgue integral. For Cuntz-Pimsner algebras with their gauge flow, the construction produces KMS-states from traces on the coefficient algebra and recovers the Laca-Neshveyev correspondence. For a discrete group acting on its Stone-\v{C}ech boundary, we recover the Patterson-Sullivan measures on the Stone-\v{C}ech boundary for a flow defined from the Radon-Nikodym cocycle. | math.OA | math |
Constructing KMS states from infinite-dimensional spectral
triples
Magnus Goffeng∗, Adam Rennie‡, Alexandr Usachev∗
∗ Department of Mathematical Sciences,
Chalmers University of Technology and University of Gothenburg,
Gothenburg, Sweden
‡School of Mathematics and Applied Statistics,
University of Wollongong, Northfields Ave
Wollongong, Australia
November 19, 2018
Abstract
We construct KMS-states from Li1-summable semifinite spectral triples and show that
in several important examples the construction coincides with well-known direct construc-
tions of KMS-states for naturally defined flows. Under further summability assumptions
the constructed KMS-state can be computed in terms of Dixmier traces. For closed ma-
nifolds, we recover the ordinary Lebesgue integral. For Cuntz-Pimsner algebras with their
gauge flow, the construction produces KMS-states from traces on the coefficient algebra
and recovers the Laca-Neshveyev correspondence. For a discrete group acting on its Stone-
Cech boundary, we recover the Patterson-Sullivan measures on the Stone- Cech boundary
for a flow defined from the Radon-Nikodym cocycle.
Contents
1 Introduction
1.1 Main results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.2 Connection to some earlier work . . . . . . . . . . . . . . . . . . . . . . . . . .
1.3 Structure of the paper . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.4 Notations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.5 Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2 Preliminaries
2.1.1
2.2 Examples
2.1 Semifinite spectral triples and summability . . . . . . . . . . . . . . . . . . . .
Semifinite spectral triples from unbounded Kasparov modules . . . . . .
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.1 Dirac operators on closed manifolds
. . . . . . . . . . . . . . . . . . . .
2.2.2 Graph C ∗-algebras . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.3 Cuntz-Pimsner algebras . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.4 Group C ∗-algebras . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2
3
5
7
7
8
8
8
12
13
14
17
19
24
∗email: [email protected], [email protected], [email protected]
1
3 KMS states constructed from Li1-summable spectral triples
3.1 The positive part of the spectrum and heat traces
. . . . . . . . . . . . . . . .
3.2 To-plitz or not To-plitz . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.3 Modular spectral triples and modular index theory . . . . . . . . . . . . . . . .
4 The KMS-state φω and Dixmier traces
5 The KMS-state φω in examples
5.1 Dirac operators on closed manifolds
. . . . . . . . . . . . . . . . . . . . . . . .
5.2 Graph C ∗-algebras . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.3 Group C ∗-algebras . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6 KMS-states on Cuntz-Pimsner algebras with their gauge action
6.1 KMS-states on Cuntz-Pimsner algebras from traces on the coefficient algebra .
6.2 The Toeplitz construction vs the Laca-Neshveyev correspondence . . . . . . . .
6.3 Obstructions to bi-Hilbertian bimodule structures . . . . . . . . . . . . . . . . .
26
27
31
37
38
43
43
47
48
50
50
53
57
1
Introduction
The construction of the JLO cocycle [26, 36, 37] from θ-summable spectral triples [14] has
from the start been closely linked with the idea of KMS states. A θ-summable spectral triple
(A, H, D) on a C ∗-algebra A gives rise to a state φ(a) := Tr(ae−D2
) on A and under suitable
conditions this is a KMS-state on the saturation of A by the R-action defined from the wave
operators eitD2
. By [37] the JLO-cocycle can be defined starting from this KMS-state. On the
other hand, [14] shows that a finitely summable spectral triple (A, H, D) on a C ∗-algebra A
defines a tracial state on A. Similar constructions were studied in [63].
The idea since then has been to understand the measure theory associated to θ-summable
spectral triples in terms of 'twisted traces', and more specifically KMS states. Indeed this idea
was present early in the development, [37]. Two viewpoints make it interesting to study states
associated with spectral triples having specified summability degrees: the associated states
obstructs summability degrees, and the states provide a notion of measure theory.
In this paper we present a construction of KMS states from Li1-summable spectral triples. By
definition, a spectral triple (A, H, D) is Li1-summable if and only if e−tD is trace class for t
large enough -- a slight strengthening of being θ-summable.
It is an important observation that large classes of examples of θ-summable spectral triples are
also Li1-summable.
For the spectral triple defined from a Dirac operator on a closed manifold, our construction
recovers the Lebesgue integral. For Cuntz-Pimsner algebras we also relate our construction to
previous work of Laca and Neshveyev [45], and the authors [31, 56]. We also examine spectral
triples arising from certain Hilbert space valued cocycles on discrete groups.
In the examples we consider, the KMS-states are associated to flows that are well-suited to the
geometries. This is usually not the case for the KMS-state φ(a) = Tr(ae−D2
) associated with
a θ-summable spectral triple. It is our hope that our construction provides a more natural
approach to the KMS-states appearing in the JLO-cocycle and that in the future it will have
a bearing on the index theory of Li1-summable spectral triples.
2
1.1 Main results
We now state our main results. All our results make sense for general semifinite spectral triples,
and so we fix a semifinite trace T for this discussion.
First, we state the main technical construction of KMS-states from Li1-summable spectral
triples. After that, we state the implications of this construction to more specific examples.
We use the notation PD for the non-negative spectral projection of D, i.e. PD := χ[0,∞)(D). If
for some βD ≥ 0, T(PDe−tD) is finite for t > βD and diverges as t ց βD, we say that D
has positive T-essential spectrum. We define the C ∗-algebra AD as the saturation of A under
the action of the wave group eitD, that is
AD := C ∗ (∪t∈Rσt(A)) , where σt(a) := eitDae−itD.
At this stage, we formulate our results in terms of AD. In Subsection 3.2 we refine the con-
struction to a smaller C ∗-algebra. In examples, the construction often applies to A directly.
Recall from [4, Definition 5.3.1] that a state φ on an R-C ∗-algebra σ : R y A is said to be
KMS at inverse temperature β if φ(ab) = φ(σ−iβ(b)a) for a, b from an R-invariant norm dense
∗-subalgebra of A. If φ is a state on an R-von Neumann algebra σ : R y A we say that it is
KMS if the same condition holds on an R-invariant σ-weakly dense ∗-subalgebra of A.
The following theorem is the main result of the paper.
Theorem 1. Let (A, H, D, N, T) be a unital Li1-summable semifinite spectral triple such that
D has positive T-essential spectrum (see Definition 3.1 on page 27) and is β-analytic (see
Definition 3.18 on page 33). Define
βD := inf{t > 0 : T(PDe−tD) < ∞}.
For any extended limit ω ∈ L∞(βD,∞)∗ as t → βD (see Definition 3.14 on page 32), we define
the state φω on AD as
φω(a) := lim
t→ω
T(PDae−tD)
T(PDe−tD)
.
Then φω is a KMS-state at inverse temperature βD for the R-action defined from σt.
particular, if βD = 0 then φω is a tracial state on A.
If βD = 0, and there is a decreasing function ψ : [0,∞) → (0,∞) with regular variation
of index −1, satisfying the conditions (4.2) and (4.3), and for some d > 0 we have that
µT(t, PDD) ∼ ψ(t)−1/d as t → ∞, then for any exponentiation invariant extended limit ω as
t → ∞,
In
φω(a) = Tω,ψ(PDa(1 + D2)−d/2),
where ω is an extended limit as t → 0 defined in Theorem 4.9 (see page 41), and Tω,ψ is the
Dixmier trace defined from T and ω on the weak ideal Lψ(N) := {T ∈ KN : µT(t, T ) = O(ψ(t))}.
The first part of this result can be found as Corollary 3.21 (see page 34) in the body of the
text and the second part as Corollary 4.10 (see page 42).
Remark 1.1. If βD = 0, any unital Li1-summable semifinite spectral triple with T(PD) = ∞
has positive T-essential spectrum and is β-analytic. Therefore, Theorem 1 shows that any
unital Li1-summable semifinite spectral triple (A, H, D, N, T) with T(PDe−tD) < ∞ for t > 0
and T(PD) = ∞ gives rise to a tracial state on A. This extends a result of Voiculescu [63,
Proposition 4.6]. For details on this case, see Theorem 3.22 (see page 34).
3
The following three results compute the KMS-state in specific examples.
Theorem 2. Let M be a closed Riemannian manifold, A := C ∞(M ), D be a Dirac operator
on a Clifford bundle S → M and H := L2(M, S). Then the KMS-state φω constructed in
Theorem 1 is independent of ω and is a tracial state on C(M ) that takes the form
φω(a) =ZM− a dV,
where dV denotes the volume measure defined from the Riemannian metric on M and Z− the
normalized integral.
This result appears as Theorem 5.1 (see page 43) in the body of the text.
Theorem 3. Let A be a unital C ∗-algebra, E be a strictly W-regular fgp bi-Hilbertian bimodule
(see Definitions 2.21 and 2.25 on pages 19 and 21, respectively) and (OE, ΞA, D) the associated
unbounded (OE, A)-cycle as in [31]. If τ is a positive trace on A, then the semifinite spectral
triple (OE, ΞA ⊗A L2(A, τ ), D ⊗ 1A, (End∗
Moreover, if τ is critical for E (see Definition 6.2 on page 51), the assumptions in Theorem
1 are satisfied and the state φω is KMS for the gauge action on OE. If τ satisfies the Laca-
Neshveyev condition for α ≥ 0 (see Definition 6.7 on page 53), then φω is independent of ω and
takes the form φω = φLN,τ where φLN,τ is the KMS-state defined from τ via the Laca-Neshveyev
correspondence.
A(ΞA) ⊗ 1)′′, Trτ ) is Li1-summable.
This result is found in Section 6 (starting on page 50). We also discuss extensions of these
results to more general A−A-correspondences in Subsection 6.3 (starting on page 57) dispensing
the assumption of strict W-regularity.
Theorem 4. Let Γ be a discrete group and c : Γ → H0 a Hilbert space valued proper 1-
cocycle defining a length function of at most exponential growth. The semifinite spectral triple
(A, H, D, N, T) constructed from c in Subsection 2.2.4 is an Li1-summable semifinite spectral
triple on Cb(Γ) ⋊ Γ. Moreover, if c is critical (see Definition 3.10 on page 30) the assumptions
of Theorem 1 are satisfied and the associated KMS-state φω on C(∂SCΓ) ⋊ Γ is given by
where µω is a quasi-invariant Patterson-Sullivan measure on the Stone-Cech boundary ∂SCΓ.
The state φω extends to a KMS-state on the von Neumann algebra L∞(∂SC Γ, µω)⋊Γ where it
is KMS with inverse temperature 1 for the R-action defined from the Radon-Nikodym cocycle
φωXg∈Γ
agλg =Z∂Γ
ae dµω,
σtXg∈Γ
agλg :=Xg∈Γ(cid:18) dg∗µ
dµ (cid:19)it
agλg.
This result appears as Theorem 5.10 (see page 49) below. Our method extends to proper
quasi-cocycles, and as such would allow for the construction of KMS-states from semifinite
spectral triples with possible K-homological content on a-TT-menable groups.
Remark 1.2. We will prove that the spectral triple of a length function (which is K-homologically
trivial) gives rise to the same KMS state as that appearing in Theorem 4.
4
1.2 Connection to some earlier work
Here we show how our approach relates to some results obtained by Connes in [15, Section
IV.8.α, Theorem 4]. Connes proves that θ-summable Fredholm modules can be lifted to θ-
summable spectral triples. We show that Connes' result can be extended to Lis-summability
for 0 < s ≤ 1, and discuss obstructions to summability properties of K-homology classes. For
terminology and notations concerning summability and operator ideals, the reader is referred
forward to Subsection 2.1.
Recall [17, 6] that a semifinite Fredholm module is a collection (A, H, F, N, T) where A acts
on the Hilbert space H by operators from N and F ∈ N is an operator with a(F − F ∗), a(F 2 −
1), [F, a] ∈ KT for all a ∈ A. We say that (A, H, F, N, T) is unital if A acts unitally. A
unital semifinite Fredholm is said to be Lis-summable if [F, a] ∈ Lis(T) for all a ∈ A and
F 2 − 1, F − F ∗ ∈ Li2s(T). If the same conditions holds with Lis(T) replaced by Lp(T), and
Li2s(T) by Lp/2(T), we say that (A, H, F, N, T) is p-summable. If (A, H, F, N, T) is a semifinite
Fredholm module we say that a semifinite spectral triple (A, H, D, N, T) is a lift if F −sign(D) ∈
KT.
Theorem 5. Let s ∈ (0, 1] and (A, H, F, N, T) be a unital semifinite Lis-summable Fredholm
module with F 2 = 1 and F = F ∗. Assume that A is countably generated. Then there is a
self-adjoint operator D affiliated with N making (A, H, D, N, T) into a unital semifinite Lis-
summable spectral triple with
Moreover, (A, H, D, N, T) satisfies that a Dom(D1/s) ⊆ Dom(D1/s) and [D1/s, a] has a
bounded extension for all a ∈ A.
F = FD := DD−1.
This theorem is found in [15, Section IV.8.α, Theorem 4] in the special case s = 1/2 and
N = B(H). We will not give the full details of the proof in the general case, but merely
indicate how Connes' proof extends. The starting point of Connes' proof is a reduction to the
case that A contains F and is generated by a countable group of unitaries Γ generated by a
countable set of unitaries (uµ)µ∈N. This argument extends to a general von Neumann algebra
N. Connes introduces the operator
G :=Xµ∈N
[F, uµ]∗[F, uµ]
2µk[F, uµ]∗[F, uµ]kLi1
.
Since [F, uµ] ∈ Li1/2 for all µ, the series converges in Li1. The proof proceeds by using an
average procedure Θ over the group Γ applied to G and Connes proves that D := F Θ(G)−1/2
fulfils the statement of the theorem. For general s ∈ (0, 1], the proof goes mutatis mutandis
using the operator
Gs :=Xµ∈N
([F, uµ]∗[F, uµ])2s
2µk([F, uµ]∗[F, uµ])2skLi1 ∈ Li1(T),
and setting D := F Θ(G)−s.
In the special case s = 1, we obtain that (A, H, F, N, T) lifts to a unital semifinite Lis-summable
for all a ∈ A the commutator
spectral triple (A, H, D, N, T) which is Lipschitz regular, i.e.
[D, a] is bounded.
5
The lifting theorem for Lis-summable spectral triples (Theorem 5) stands in sharp contrast to
the finitely summable setup, or even the Li(0),s-summable setup. The two upcoming theorems
show that a statement as in Theorem 5 could not extend to the ideal Li(0),1.
Theorem 6. Let A be a unital C ∗-algebra with no tracial states and (A, H, D, N, T) be a unital
semifinite spectral triple on A defining a non-trivial class in KK1(A, KT). Then PD(i± D)−1 /∈
Li(0),1(T).
Proof. Consider a unital C ∗-algebra A and an Li(0),1-summable unital semifinite spectral triple
(A, H, D, N, T) on A defining a non-trivial class in KK1(A, KT). In particular, T(PD) = ∞;
otherwise PD ∈ KT which contradicts the non-triviality of the KK1(A, KT)-class defined
by (A, H, D, N, T). By Remark 1.1 (see page 3) all assumptions of Theorem 1 reduces to
T(PD) = ∞ in the Li(0),1-summable case. Therefore, the existence of Li(0),1-summable unital
semifinite spectral triples on A being non-trivial in KK implies that A admits a tracial state.
This argument shows that if A admits no tracial states, it admits no Li(0),1-summable unital
semifinite spectral triple. In fact, a careful inspection of the results used show that as soon
as there is a unital semifinite spectral triple on A with PD(i ± D)−1 ∈ Li(0),1(T), there is an
associated tracial state on A. The theorem follows.
There are several C ∗-algebras carrying no traces, for instance any purely infinite C ∗-algebra.
Using Theorems 1 and 6, we will give an example of a finitely summable Fredholm module that
can not lift to an Li(0),s-summable spectral triple. In particular, lifting of finite summability
and Li(0),s-summability fails in general.
Theorem 7. There is a C ∗-algebra A with K 1(A) 6= 0, such that for a dense ∗-subalgebra
A ⊆ A we can represent any x ∈ K 1(A) by a Fredholm module (A, Hx, Fx) with F 2
x = 1,
F ∗
x = Fx and for any a ∈ A, [Fx, a] is of finite rank. Moreover, any lift (A, Hx, Dx) of
(A, Hx, Fx) will satisfy that (1 + D2
x)−1/2 /∈ Li(0),1(H).
Proof. Consider the Cuntz algebra A = ON and A the ∗-algebra generated by isometries
S1, S2, . . . , SN ∈ ON with orthogonal ranges. It is a well known fact that K 1(ON ) ∼= Z/(N −
1)Z 6= 0. By [29], we can represent the generator of K 1(ON ) ∼= Z/(N − 1)Z by the Fredholm
module (A, L2(ON , φ), 2P − 1) where φ is the KMS-state on ON and P is the orthogonal
projection onto the closed linear span of Sµ, where µ ranges over all finite words on the
alpabet {1, . . . , N}. By the results of [29, Section 2.2], [2P − 1, a] = 2[P, a] is finite rank for all
a ∈ A. The first statement of the theorem follows.
There are no tracial states on ON since
1
N − 1
NXj=1
S∗
j Sj −
NXj=1
SjS∗
j =
1
N − 1
[S∗
j , Sj].
NXj=1
1ON =
1
N − 1
(N − 1)1ON =
We can now deduce the second statement of the theorem from Theorem 6.
Remark 1.3. It is not of importance that K 1(ON ) is torsion for the argument in Theorem 7
to work. In [29], non-torsion examples satisfying the conclusions of Theorem 7 can be found.
The reader should also note that the proof of Theorem 7 obstructs all lifts (A, Hx, Dx) of
(A, Hx, Fx) with PDx (1 + D2
x)−1/2 ∈ Li(0),1(H).
6
In the nonunital case, the techniques of [5] will likely be required. The substantial technical
considerations in the nonunital case goes beyond this paper, and is left to future work.
1.3 Structure of the paper
Section 2 recalls the basics of (unital) semifinite spectral triples and their summability. We
also recall our main examples from the literature in this section for later use.
Section 3 presents our construction of KMS states from Li1-summable spectral triples. We
close the section by discussing connections to modular spectral triples. We consider the case
βD = 0 in Section 4 and compute the tracial states constructed in Section 3 by means of
Dixmier traces. In Section 5 we apply the techniques of Section 3 to the examples.
The final Section 6 examines the construction of KMS states for Cuntz-Pimsner algebras. In
this case we apply our ideas to derive obstructions to the existence of fgp bi-Hilbertian bimodule
structures compatible with the underlying correspondence of the Cuntz-Pimsner algebra.
1.4 Notations
A(X)
N
T
KN
End∗
KA(X)
A
A′
A
AD
PD := χ[0,∞)(D)
FD := 2PD − 1
µT(·, T )
nT(·, T )
Lis(T), Li(0),s(T)
Lψ(T), L(0),ψ(T)
Tω,ψ(T )
T(A)
L∞(a,∞), a ≥ 0
C0(a,∞), a ≥ 0
limt→ω f (t)
ℓ∞(N)
limk→ω xk
Lg
f ∼ g
semifinite von Neumann algebra
positive, faithful, normal, semifinite trace on N
ideal of T-compact operators
C ∗-algebra of adjointable endomorphisms of an A-Hilbert C ∗-module X
C ∗-algebra of compact endomorphisms of an A-Hilbert C ∗-module X
∗-algebra
the commutant of an algebra A
C ∗-closure of an algebra A
saturation of A under the action of the wave group eitD
non-negative spectral projection of an operator D
singular values function of an operator T affiliated with N
distribution function of an operator T affiliated with N
ideals of compact operators in Definition 2.6 on page 9
ideals of compact operators in Definition 2.6 on page 9
Dixmier trace on Lψ(T)
set of positive traces on a unital C ∗-algebra A
space of essentially bounded functions on (a,∞) equipped with
the essential supremum norm
subspace of L∞(a,∞) of all continuous functions vanishing at infinity
value of an extended limit ω on a function f
space of bounded sequences equipped with the supremum norm
value of an extended limit ω on a sequence x
transfer operator defined by formula (2.13) on page 23
for two functions or sequences f and g if f = g + o(f ) and g = f + o(g)
7
1.5 Acknowledgements
A. R. thanks the Gothenburg Centre for Advanced Studies in Science and Technology for
funding and the University of Gothenburg and Chalmers University of Technology for their
hospitality in 2017 when this work was begun. M. G. and A. U. were supported by the Swedish
Research Council Grant 2015-00137 and Marie Sklodowska Curie Actions, Cofund, Project
INCA 600398. The authors acknowledge the support of the Erwin Schrodinger Institute where
part of this work was conducted.
The authors are grateful to Alan Carey, Heath Emerson and Bram Mesland for inspiring
discussions. We also thank Branimir ´Ca´ci´c for sharing his construction of semifinite spectral
triples from proper group cocycles, and Edward McDonald for references on previous work in
that direction.
2 Preliminaries
Before entering into the body of the paper, we recall some basic definitions that we will require
and provide some examples that motivated this work. These examples will be studied further
in the later sections of the paper. The results will be formulated for semifinite spectral triples.
We do however remark that there are several examples of 'vanilla' spectral triples that will be
used throughout the paper.
2.1 Semifinite spectral triples and summability
To set the stage for the paper, we summarize the basic definitions and properties of semifinite
spectral triples. The reader familiar with semifinite spectral triples and symmetrically normed
operator ideals can skip this subsection.
We let N denote a semifinite von Neumann algebra and we fix a positive, faithful, normal,
semifinite trace T on N. The T-compact operators are denoted by KN. The C ∗-algebra KN
can be defined as the norm closed ideal generated by the projections E ∈ N with T(E) < ∞.
Equivalently, one can define KN := {T ∈ N : µT(t, T ) = o(1) as t → ∞} where the singular
value function µT(t, T ) is defined as
µT(t, T ) := inf(cid:8)kT (1 − E)kN : where E ∈ N is a projection with T(E) ≤ t(cid:9).
Definition 2.1. A semifinite spectral triple (A, H, D, N, T) consists of
(2.1)
• A ∗-algebra A represented on a Hilbert space H as operators in N ⊆ B(H), that is, we
have a specified ∗-homomorphism π : A → N.
• A densely defined self-adjoint operator D : dom D ⊂ H → H which is affiliated with N
such that for all a ∈ A we have a · domD ⊂ domD and
1. [D, π(a)] := Dπ(a) − π(a)D initially defined on dom(D) is bounded in operator
norm.
2. π(a)(1 + D2)−1/2 ∈ KN.
8
Remark 2.2. Sometimes we write a spectral triple as a collection of three objects (A, H, D).
In this case, it is implicitly assumed that N = B(L2(M, S)) and T is the standard trace.
Remark 2.3. If in addition to the data (A, H, D, N, T) we have specified an operator γ ∈ B(H)
with γ = γ∗, γ2 = 1, Dγ + γD = 0 on Dom(D), and for all a ∈ A we have γπ(a) = π(a)γ, we
call the semifinite spectral triple even, or sometimes graded. If γ has not been specified, we
say that the semifinite spectral triple is odd, or ungraded. This distinction plays an important
role in the topological properties of the spectral triple, but since this paper deals with measure
theory it will not play a role in this paper.
Remark 2.4. We will nearly always dispense with the representation π, treating A as a subal-
gebra of N ⊆ B(H).
Remark 2.5. In the sequel we assume that the algebra A is unital and that 1 ∈ A acts as
In particular, the operator (1 + D2)−1/2 is a T-compact
the identity of the Hilbert space.
operator. To emphasize this assumption, we refer to the data (A, H, D, N, T) as a unital
semifinite spectral triple.
Examples of semifinite spectral triples often satisfy a finer summability structure, i.e. a re-
finement of the condition (1 + D2)−1/2 ∈ KN. We formulate such conditions in terms of
symmetrically quasi-normed operator ideals. We will use the Schatten ideals, the Li-ideals and
more generally weak ideals.
Definition 2.6. Let N denote a semifinite von Neumann algebra and T a positive, faithful,
normal, semifinite trace on N. For parameters p, d ∈ [1,∞) and s > 0 we define the following
operator ideals.
• Lp(T) := {T ∈ KN : µT(·, T ) ∈ Lp(0,∞)}.
• L(d,∞)(T) := {T ∈ KN : µT(·, T ) = O(t−1/d) as t → ∞}.
• Lis(T) := {T ∈ KN : µT(t, T ) = O((log(t))−s) as t → ∞}.
• Li(0),s(T) := {T ∈ KN : µT(t, T ) = o((log(t))−s) as t → ∞}.
• If ψ : [0,∞) → (0,∞) is a decreasing function satisfying that supt>0
the associated weak ideal
and its separable subspace
Lψ(T) := {T ∈ KN : µT(t, T ) = O(ψ(t))},
L(0),ψ(T) := {T ∈ KN : µT(t, T ) = o(ψ(t))},
ψ(t)
ψ(2t) < ∞, we define
The condition supt>0
quasi-Banach space in the quasi-norm
ψ(2t) < ∞ guarantees that Lψ(T) is a vector space, and in fact even a
ψ(t)
kTkLψ := sup
t>0
µT(t, T )
ψ(t)
.
Note that Lis(T) = Lψ(T) and Li(0),s(T) = L(0),ψ(T) for ψ(t) := (log(2 + t))−s. It is immediate
from the definition that Lp(T) ⊆ Li(0),s(T) for any p and s. More generally, if ψ1, ψ2 : [0,∞) →
ψj (2t) < ∞, then Lψ1(T) ⊆ Lψ2(T)
(0,∞) are two decreasing functions satisfying that supt>0
as soon as ψ1 = O(ψ2).
ψj (t)
9
Remark 2.7. Our definition of symmetrically normed operator ideals in the semifinite setting
differs slightly from the standard definition unless N is atomic. In the usual definition, the sym-
metrically normed operator ideals are defined from operators affiliated with N that potentially
are unbounded. Since we only use bounded operators from these ideals, we have incorporated
this fact in our definition.
Definition 2.8. Let (A, H, D, N, T) be a unital semifinite spectral triple.
• (A, H, D, N, T) is said to be p-summable if (1 + D2)−1/2 ∈ Lp(T).
• (A, H, D, N, T) is said to be (d,∞)-summable if (1 + D2)−1/2 ∈ L(d,∞)(T).
• (A, H, D, N, T) is said to be Lis-summable if (1 + D2)−1/2 ∈ Lis(T).
• (A, H, D, N, T) is said to be Li(0),s-summable if (1 + D2)−1/2 ∈ Li(0),s(T).
• (A, H, D, N, T) is said to be ψ-summable if (1 + D2)−1/2 ∈ Lψ(T).
The standard terminology in the literature for the special case s = 1/2 is to refer to Li1/2-
summability as weak θ-summability and to Li(0),1/2-summability as θ-summability. Since
L(d,∞)(T) ⊆ Lp(T) for all p > d, (d,∞)-summability refines p-summability.
The notion of (d,∞)-summability is a noncommutative generalization of being d-dimensional
as the spectral triple defined from a Dirac operator on a closed d-dimensional manifold (as
in Subsection 2.2.1) is (d,∞)-summable. We shall see an abundance of Li1-summable, truly
noncommutative, examples where p-summability and (d,∞)-summability fails for all p and d.
The notion of ψ-summability generalizes both (d,∞)-summability and Li1-summability, and
appears naturally in examples of (semi-) group actions on manifolds (see Subsection 5.1 and
[19, 32]). We will make use of this notion in Section 4 where certain conditions on ψ allows
one to compute the tracial state defined from a ψ-summable unital semifinite spectral triples
in terms of Dixmier traces on Lψ(T).
Remark 2.9. It is readily verified that Lis-summability is equivalent to
T(e−tD1/s
) < ∞,
for t > t0 for some critical value t0,
and that Li(0),s-summability is equivalent to
T(e−tD1/s
) < ∞,
for t > 0.
In particular, (A, H, D, N, T) is θ-summable if and only if
T(e−tD2
) < ∞,
for all t > 0.
Historically, θ-summability has been studied more in depth than Li1-summability. This can
in part be explained from the two facts that the JLO-cocycle only requires θ-summability and
classically, the heat operator e−tD2
is geometrically more interesting than e−tD to study on a
manifold. The two operators e−tD2
and e−tD can be compared by explicit integral formulas, see
[28, Chapter 4]. We will exploit the observation that large classes of examples of θ-summable
spectral triples are also Li1-summable.
10
In the bulk of the paper, we are interested in computing asymptotics of heat traces of the
form T(Be−tD1/s
) for B ∈ N as t approaches a critical value. When (A, H, D, N, T) is Li(0),s-
summable, the critical value of t is 0, and in several classical examples (e.g. on closed manifolds)
the heat trace T(Be−tD1/s
) admits an asymptotic expansion. The following result is useful for
relating heat trace asymptotics to zeta function asymptotics in the case of the nice behaviour
appearing when t0 = 0.
Theorem 2.10. Let s ∈ (0, 1]. Let (A, H, D, N, T) be an Li(0),s-summable semifinite spectral
triple and B ∈ N. The following are equivalent.
1. There are constants pheat > 0, ǫ > 0 and cheat
T(Be−tD1/s
) = cheat
B t−spheat
B ∈ C such that
+ O(t−spheat+ǫ),
as t → 0.
2. The ζ-function ζ(z; B,D1/s) := T(BD−z/s) is well-defined for large Re(z) and there
B ∈ C and a function f = f (z) holomorphic in the region
are constants pζ > 0, ǫ′ > 0, cζ
Re(z) > spζ − ǫ′ such that
ζ(z; B,D1/s) =
1
Γ(spζ)
cζ
B
z − spζ + f (z).
In this case, pheat = pζ and cheat
it holds for all s ∈ (0, 1].
If B = 1 and either of the conditions above hold, then (A, H, D, N, T) is pζ-summable.
B. Moreover, if the conditions above hold for one s ∈ (0, 1],
B = cζ
The proof of Theorem 2.10 follows by noting that Γ(z)ζ(z; B,D1/s) is the Mellin transform
of T(Be−tD1/s
) and using [33, Proposition 5.1].
Recall that we use the notation PD := χ[0,∞)(D) for the non-negative spectral projection of D.
Let us state a fundamental lemma on the commutators of A with the function of D defined by
FD := 2PD − 1.
Note that FD differs from the phase DD−1 by the T-finite kernel projection of D.
Lemma 2.11. Let (A, H, D, N, T) be a semifinite spectral triple. Then for any a ∈ A
[FD, a] ∈ KN.
Moreover if (A, H, D, N, T) is unital, then if (A, H, D, N, T) is p-summable, then [FD, a] ∈
Lp(T) for all a ∈ A, and if (A, H, D, N, T) is Lis-summable, then [FD, a] ∈ Lis(T) for all
a ∈ A. More generally, if (A, H, D, N, T) is ψ-summable then [FD, a] ∈ Lψ(T) for all a ∈ A.
Proof. The proof of the operator inequality −k[D, a]kD−1 ≤ [FD, a] ≤ k[D, a]kD−1 for
invertible D and a = −a∗ is found in the proof of [59, Proposition 1]. The assertion follows
from the definition of p-, Lis- and ψ-summability, resp.
In the non-invertible case we replace (A, H, D, N, T) by
(cid:18)(cid:18)A 0
0(cid:19) , H ⊕ H, Dµ =(cid:18)D µ
µ −D(cid:19) , M2(N), T ⊗ TrM2(cid:19) ,
0
(2.2)
11
for µ ∈ [0, 1]. When µ > 0 we are back in the invertible case. In [6, Proposition 2.25] it is
shown that for any a ∈ A we have the norm limit
[FDµ, a]
µ→0
(cid:18)[FD, a] 0
0(cid:19) = lim
0
and so [FD, a] is compact. Indeed, the proof of [6, Proposition 2.25] shows that if (1 + D2)−1/2
is in a symmetrically quasi-normed ideal J of T-compact operators then [FD, a] ∈ J for all
a ∈ A. Again the assertion follows from the definitions of summability.
2.1.1 Semifinite spectral triples from unbounded Kasparov modules
For several kinds of C ∗-algebras one can capture the noncommutative geometry through an
unbounded Kasparov module. This is a bivariant generalization of spectral triples. Localizing
an unbounded Kasparov module in a positive trace gives rise to a semifinite spectral triple as
in Theorem 2.12 below. Several of the examples in this paper arises in this way. We briefly
recall this construction, which has been informally used for some years.
Let A and B be unital C ∗-algebras. A unital unbounded (B, A)-Kasparov module is a collection
(B, X, D) where
• B ⊆ B is a dense ∗-subalgebra,
• X is an A-Hilbert C ∗-module carrying a left action of B as adjointable operators,
• D is an A-linear, densely defined, self-adjoint, regular operator on X with A-compact
resolvent (i ± D)−1 ∈ KA(X) and
• for a ∈ B the operator [D, a] is defined on dom(D) and is bounded in the norm on X.
If τ is a positive trace on A, we write L2(X, τ ) := X ⊗A L2(A, τ ) where L2(A, τ ) is the
GNS-representation associated with τ .
For ξ, η ∈ X, we write Θξ,η for the rank one operator Θξ,η(ν) = ξ(ην)A. The von Neumann
algebra Nτ (X) := (End∗
A(X)⊗1A)′′ ⊆ B(L2(X, τ )) coincides with the weak closure of the set of
operator spanned by {Θξ,η ⊗ 1A : ξ, η ∈ X} and carries a positive, normal, semifinite, faithful
trace Trτ characterized by Trτ (Θξ,η ⊗ 1A) := τ ((ηξ)A), see [45, Section 3]. The following
theorem also appears in [49].
Theorem 2.12. Let (B, XA, D) be a unital unbounded (B, A)-Kasparov module and τ : A → C
a faithful norm densely defined norm lower semicontinuous tracial weight. Then the data
(B, L2(X, τ ), D⊗ 1, Nτ (X), Trτ ) defines a semifinite spectral triple. The von Neumann algebra
is Nτ (X) = (End∗
A(X)⊗ 1)′′ and Trτ : N → C is the (positive faithful semifinite normal) trace
dual to the normal extension of τ to A′′ ⊆ B(L2(A, τ )).
Proof. The operator D ⊗ 1 is self-adjoint by [46, Proposition 9.10]. The commutant of Nτ (X)
in X ⊗A L2(A, τ ) is the algebra A′′ (acting by right multiplication). Every unitary in A′′ thus
preserves the domain of D⊗1 and so D⊗1 is affiliated to Nτ (X). Plainly commutators of D⊗1
with B remain bounded. Since we start with an unbounded Kasparov module, the operator
12
(1 + D2)−1/2 ∈ KA(X). So we can approximate (1 + D2)−1/2 in norm by finite rank operators
Pj Θxj,yj and we can take the xj, yj ∈ X. Hence (1 + (D ⊗ 1)2)−1/2 = (1 + D2)−1/2 ⊗ 1 is in
the norm closure of the finite trace operators in Nτ (X), and so T-compact.
A conceptual viewpoint is that (B, L2(X, τ ), D ⊗ 1A, (End∗
A(X) ⊗ 1A)′′, Trτ ) is a semi-finite
refinement of the unbounded Kasparov product of (B, X, D) with the Morita morphism A →
KA(X) and the ∗-homomorphism KA(X) → K
A(X)⊗1A)′′. There is a close relationship
between the semifinite index and the Kasparov product, described in [6, 39].
(End∗
We remark at this stage that there is to date no general theory of symmetrically quasi-normed
operator ideals in Hilbert C ∗-modules, and, as such, no satisfactory way of describing summa-
bility. In concrete applications, it is possible to circumvent this problem by choosing a frame
on X, implicitly using the machinery of [52]. In the examples of most relevance to this paper
the spectrum of D is discrete, and we can use the following proposition to study summability.
Proposition 2.13. Let (B, X, D) be a unital unbounded (B, A)-Kasparov module, where D
has discrete spectrum σ(D) ⊆ R, and τ be a positive trace on A. Set Pλ := χ{λ}(D) ∈ KA(X)
for λ ∈ σ(D). Then it holds that
1. (B, L2(X, τ ), D ⊗ 1A, Nτ (X), Trτ ) is Lis-summable if and only if
Xλ∈σ(D)
e−tλ1/s
Trτ (Pλ) < ∞,
for t large enough.
2. (B, L2(X, τ ), D ⊗ 1A, Nτ (X), Trτ ) is p-summable if and only if
Xλ∈σ(D)
(1 + λ2)−p/2 Trτ (Pλ) < ∞.
The proof follows from the following formula:
Trτ (f (D)) = Xλ∈σ(D)
f (λ) Trτ (Pλ),
which holds for every positive Borel function f .
2.2 Examples
To give some further context before entering into the main construction of this paper, let us
recall some well known examples that we will further explore later on in the paper. The focus
in our presentation is on Li1-summability and heat traces. We remark that the constructions
in this subsection are rather lengthy, and the reader familiar with the literature can at a first
read restrict themself to glancing through this subsection.
13
2.2.1 Dirac operators on closed manifolds
The prototypical example of a spectral triple arises from Dirac operators on a closed Rieman-
if S → M
nian manifold M . We can work with a rather general type of Dirac operators:
is a Clifford module on M and /D is a first order elliptic operator acting on C ∞(M, S) being
symmetric in the L2-inner product and /D2 is a Laplacian type operator1, we say that /D is a
Dirac operator. In this case, the closure of /D in its graph norm defines a self-adjoint oper-
ator on L2(M, S) that we by an abuse of notation also denote by /D. It is well-known that
(C ∞(M ), L2(M, S), /D) is a spectral triple on C ∞(M ). We summarize the main properties of
its heat traces in the following proposition.
Proposition 2.14. Let M be an n-dimensional Riemannian closed manifold, /D a Dirac op-
erator on M , and (C ∞(M ), L2(M, S), /D) the associated spectral triple. This spectral triple is
(n,∞)-summable and for any classical zero-th order pseudo-differential operator A on S with
principal symbol a ∈ C ∞(S∗M, End(S)) and every s ∈ (0, 1] we have
TrL2(M,S)(Ae−t /D1/s
) = Γ(sn + 1)cs
nt−snZS∗M
TrS(a) dV + O(t−sn+ǫ),
as t → 0,
for some dimensional constant cn > 0 and ǫ > 0. Here TrS(a) ∈ C ∞(S∗M ) denotes the
fibrewise trace of a.
The dimensional constant cn is determined by the Weyl law for /D describing the ordered
sequence (λk( /D))k∈N of eigenvalues as
λk( /D) =
1
cnvol(M )1/nrank(S)1/n k1/n + O(k1/n−ǫ0),
for some ǫ0 > 0. The Weyl law is proven in many places, for instance [27]. The general heat
trace asymptotics follows from Theorem 2.10 and [27]. We return to this example below in
Example 3.5 (see page 28) and Subsection 5.1 (see page 43).
Later on in the paper, we will make use of a modification of the spectral triple coming from
a Dirac operator that is also compatible with non-isometric semigroup actions. Similar con-
structions were previously considered in [19, 32].
Definition 2.15. Let ψ : [0,∞) → (0,∞) be a positive measurable function.
• We say that ψ is regularly varying of index ρ if for all λ > 0
lim
t→∞
ψ(λt)
ψ(t)
= λρ.
(2.3)
• We say that ψ is smoothly regularly varying of index ρ if ψ ∈ C ∞ and for any k ∈ N,
tkψ(k)(t)
ψ(t)
lim
t→∞
= ρ(ρ − 1)··· (ρ − k + 1).
(2.4)
1I.e. the symbol of /D2
coincides with the Riemannian metric as a function on T ∗M .
14
ψ(2t) < ∞ and there is an associated weak ideal Lψ
A regularly varying function satisfies supt>0
as in Definition 2.6. By [32, Lemma 7.1], any smoothly regularly varying function ψ satisfies
t ψ(t) = O(ψ(t)(1 + t2)−k/2) so if ψ additionally is bounded, ψ belongs to the Hormander
that ∂k
class S0.
ψ(t)
Smooth regular variation is a strengthening of having regular variation -- a condition used
below in Section 4 in the context of defining and computing Dixmier traces. See more also
in [32]. By [3, Theorem 1.8.2], any regularly varying function asymptotically behaves like a
smoothly regularly varying function. Smooth regular variation allows for defining associated
classes of pseudo-differential operators and computing Dixmier traces of geometric operators
by means of a Connes trace theorem, see [32, Section 7 and 9].
For a decreasing smoothly varying function ψ : [0,∞) → (0,∞) with limt→0 ψ(t) = 0, we define
the self-adjoint operator
/Dψ := F /Dψ( /Dn)−1.
It follows from [32, Proposition 10.1] that /Dψ ∈ L0
ψ−1(M, S) (see [32] for the meaning of this
symbol) and its ψ-principal symbol is cS(ξ)ξ−1ψ(ξ)−1, where cS : T ∗M → End(S) denotes
Clifford multiplication. The next result follows from [32, Proposition 10.3].
Proposition 2.16. Let M be an n-dimensional Riemannian closed manifold, /D a Dirac op-
erator on M , and ψ as above with
ψ(t)−1 = O(t1/n),
as t → ∞.
Then (C ∞(M ), L2(M, S), /Dψ) is a ψ-summable spectral triple whose associated K-homology
class coincides with that of (C ∞(M ), L2(M, S), /D).
If ψ(t)−1 = o(t1/n), then for any a ∈
C ∞(M ), the operator [ /Dψ, a] is compact with
µ(t, [ /Dψ, a]) = O(t−1/nψ(t)−1),
as t → ∞.
We return to the problem of computing heat traces involving /Dψ below in the Section 4 (see
page 38) and Subsection 5.1 (see page 43).
The spectral triple (C ∞(M ), L2(M, S), /Dψ) does in some cases extend to a crossed product by
a semigroup action. For simplicity, we consider an action of N by a local diffeomorphism g :
M → M . Following [32], we say that g acts conformally if there is a function cg ∈ C ∞(M, R>0)
such that g∗gM = cggM where gM denotes the Riemannian metric on M . We say that g lifts
to the Clifford bundle S → M if there is a unitary Clifford linear morphism ug : g∗S → S. For
simplicity, we assume M to be connected and define N := #g−1({x}) for some x ∈ M . Since
g is a local diffeomorphism and M is connected, N is independent of the choice of x.
If g : M → M is a surjective local diffeomorphism, acting conformally and lifting to S, we can
define the isometry
Vg : L2(M, S) → L2(M, S),
Vgξ := cn/4
g N −1/2ug(ξ ◦ g).
(2.5)
The isometry Vg satisfies the following for a ∈ C(M ):
VgaV ∗
g = (a ◦ g)VgV ∗
g ,
and V ∗
g aVg = Lg(a),
15
where Lg(a)(x) := Pg(y)=x a(y). See more in [19, Proposition 8.3]. Define A as the ∗-
algebra generated by C ∞(M ) and Vg. By [19, Proposition 8.6], the C ∗-closure of A is
the image in a representation on L2(M, S) of the Cuntz-Pimsner algebra OEg defined from
Eg := C(M )Vg (see more in Example 2.30 (see page 23) and Example 6.6 (see page 52)).
The space Eg = C(M )Vg ⊆ B(L2(M, S)) is considered as a bimodule over C(M ) and is an
fgp bi-Hilbertian C(M )-bimodule because Eg can be identified with C(M ) with the bimodule
structure (af b)(x) = a(x)f (x)b(g(x)) for a, b, f ∈ C(M ), for details, see Example 2.30 below
on page 23 or [19, Section 8].
Definition 2.17. Let ψ : [0,∞) → (0,∞) be a decreasing function and g : M → M a local
diffeomorphism of a Riemannian manifold.
• We say that ψ and g are compatible if there is a constant C ≥ 0 such that for all x ∈ M
and ξ ∈ T ∗
x M , the differential Dg satisfies
(cid:12)(cid:12)ψ((Dg)T
x ξ) − ψ(ξ)(cid:12)(cid:12) ≤ Cψ(ξ)2.
• We say that g acts isometrically if Dg acts isometrically on each fibre.
The following results poses restrictions on a function ψ compatible with a local diffeomorphism
which acts non-isometrically.
Proposition 2.18. Let ψ : [0,∞) → (0,∞) be a decreasing function with limt→∞ ψ(t) = 0
and g : M → M a compatible local diffeomorphism of a Riemannian manifold. If ψ has regular
variation, then either g acts isometrically or ψ has regular variation of index 0.
Proof. Assume that g acts non-isometrically, we shall prove that in this case ψ has regular
ψ(rt)
ψ(t) = 1 for some
variation of index 0. Since ψ is decreasing, it suffices to show that limt→∞
r 6= 1 by [32, Proposition 2.15]. If g acts non-isometrically, there is a point x ∈ M and a unit
vector ξ0 ∈ T ∗
x M such that (Dg)T
x ξ0g(x) 6= 1. Set
r := (Dg)T
x ξ0g(x) 6= 1.
Since ψ is compatible with g, there is a constant C ≥ 0 such that
and by setting ξ = tξ0, we arrive at the inequality
(cid:12)(cid:12)ψ((Dg)T
x ξ) − ψ(ξ)(cid:12)(cid:12) ≤ Cψ(ξ)2
ψ(rt) − ψ(t) ≤ Cψ(t)2.
After dividing by ψ(t), taking the limit t → ∞ and using that limt→∞ ψ(t) = 0 we arrive at
the desired equality limt→∞
ψ(rt)
ψ(t) = 1.
A prototypical example of a function with regular variation of index 0 is ψ(t) := log(1 + t).
This function is compatible with any conformal local diffeomorphism and has been considered
in the context of constructing spectral triples in [19, 29, 30, 32, 50].
16
Proposition 2.19. Let M be an n-dimensional Riemannian closed manifold, /D a Dirac op-
erator on S → M , g : M → M a surjective local diffeomorphism acting conformally and lifting
to S and ψ : [0,∞) → (0,∞) a decreasing function compatible with g, having smooth regular
variation and satisfying limt→∞ ψ(t) = 0 and ψ(t)−1 = O(t1/n) as t → ∞. Let A denote the
∗-algebra generated by C ∞(M ) and the isometry Vg from Equation (2.5) (see page 15). Then
(A, L2(M, S), /Dψ) is a unital ψ-summable spectral triple.
This result follows in the same manner as in [19, Section 8] and [32, Theorem 10.6]. Clearly, if
ψ(t)−1 = O(log(t)) as t → ∞, then (A, L2(M, S), /Dψ) is Li1-summable. The reader should note
that the K-homology class [(A, L2(M, S), /Dψ)] ∈ K ∗(OEg ), where OEg is the Cuntz-Pimsner
algebra of the module Eg = C(M )Vg, discussed later. The class [(A, L2(M, S), /Dψ)] restricts to
[(C ∞(M ), L2(M, S), /D)] ∈ K ∗(C(M )). Thus the class [(C ∞(M ), L2(M, S), /D)] ∈ K ∗(C(M ))
obstructs [(A, L2(M, S), /Dψ)] ∈ K ∗(OEg ) being a Kasparov product with the Cuntz-Pimsner
boundary extension in KK 1(OEg , C(M )) -- we will return to study this boundary extension
and its associated semifinite spectral triples below in Subsection 2.2.3.
2.2.2 Graph C ∗-algebras
A class of examples carrying interesting noncommutative geometries with a well-studied set of
KMS-states is that of graph C ∗-algebras. With a finite directed graph G = (V, E), with edge
set E and vertex set V , one associates a C ∗-algebra C ∗(G) [54]. For simplicity we suppose
that we have no sources nor sinks. The C ∗-algebra C ∗(G) is generated by partial isometries
(Se)e∈E and projections (pv)v∈V satisfying the relations
S∗
e Se = pr(e),
and pv = Xs(e)=v
SeS∗
e ,
where r(e) denotes the range of the vertex e and s(e) its source. The C ∗-algebra C ∗(G)
can be described as a Cuntz-Pimsner algebra in several ways, a class of C ∗-algebras carrying
noncommutative geometries that we will study in the next subsection. In this subsection, we
focus on a construction of noncommutative geometries along an orbit in the infinite path space
of G -- a construction based in the model of C ∗(G) as a groupoid C ∗-algebra. The associated
noncommutative geometries come from [29]. As explained in [31] the groupoid model can be
seen as a Cuntz-Pimsner model of C ∗(G) using the one-sided infinite path space
ΩG := {x = e1e2 ··· ∈ E N : s(ej) = r(ej+1) ∀j}.
The path space ΩG is a compact Hausdorff space in the subspace topology ΩG ⊆ E N. It carries
a shift mapping σG : ΩG → ΩG, σG(e1e2e3 ··· ) := e2e3 ··· . The shift mapping is a surjective
local homeomorphism. We can define an ´etale groupoid GG over ΩG by
GG :=(cid:8)(x, n, y) ∈ ΩG × Z × ΩG : ∃k ≥ max(0,−n) such that σn+k
G (x) = σk
The range mapping is defined by r(x, n, y) = x, the source mapping as s(x, n, y) := y and the
product by
G(y)(cid:9).
(x, n, y)(y, m, z) := (x, n + m, z).
17
The topology of GG is uniquely determined by declaring the groupoid to be ´etale and the
mappings (x, n, y) 7→ n and
κG(x, n, y) := min{k ≥ max(0,−n) : σn+k
G (x) = σk
G(y)},
to be continuous. There is an isomorphism πG : C ∗(G) → C ∗(GG) determined by defining
πG(Se) to be the characteristic function of the set {(x, 1, σG(x)) : x ∈ Ce} where Ce :=
{e1e2 ··· ∈ ΩG : e1 = e}. See [18].
Define the function ψ0 : {(n, k) ∈ Z × N : n + k ≥ 0} → Z by
k = 0
k > 0.
ψ0(n, k) :=(n,
−n − k,
For a point y ∈ ΩG, we define the discrete set
Vy := d−1({y}) = {(x, n) : (x, n, y) ∈ GG}.
The set Vy is the union of all forward orbits of all backward orbits of y under σG where we keep
track of the lag n. Since Vy is a fibre of the domain mapping, point evaluation in y induces a
representation πy : C ∗(G) → B(ℓ2(Vy)). If G is primitive C ∗(G) is simple and πy is faithful.
At the level of the generators,
πy(Se)δ(x,n) =(δ(ex,n+1),
0,
r(x) = s(e),
r(x) 6= s(e).
Proposition 2.20. Define the operator Dy densely on ℓ2(Vy) as the self-adjoint operator with
Dyδ(x,n) := ψ0(n, κ(x, n, y))δ(x,n).
The triple (Cc(GG, ℓ2(Vy), Dy) is an Li1-summable spectral triple. Moreover, under the isomor-
phism K 1(C ∗(G)) ∼= ZE/(1− Aedge)ZE, of odd K-homology with the cokernel of the edge adja-
cency matrix, the class [(Cc(GG, ℓ2(Vy), Dy)] is mapped to the element δe mod (1 − Aedge)ZE
where δe ∈ ZE denotes the basis element corresponding to e ∈ E.
Proof. It is proven in [29, Theorem 5.2.3] that (Cc(GG, ℓ2(Vy), Dy) is a spectral triple whose
K-homology class corresponds to the element δe mod (1 − Aedge)ZE under K 1(C ∗(G)) ∼=
ZE/(1− Aedge)ZE. It remains to prove that (Cc(GG, ℓ2(Vy), Dy) is Li1-summable. We compute
that
Tr(e−tDy) = X(x,n)∈Vy
However, if (x, n) is such that κG(x, n, y) = k the path x is determined by y except for its first
n + k steps so #{(x, n) : κG(x, n, y) = k} ≤ En+k ≤ elog(E)(n+k). We conclude that
e−t(n+κG(x,n,y)) =Xn∈Z
#{(x, n) : κG(x, n, y) = k}e−t(n+k).
∞Xk=max(0,−n)
Tr(e−tDy) < ∞ if
t > log(E).
We return to this example below in Example 3.6 (see page 28) and Subsection 5.2 (see page
47).
18
2.2.3 Cuntz-Pimsner algebras
In this subsection we consider the construction of semi-finite spectral triples on Cuntz-Pimsner
algebras -- a broad class of examples which include both Cuntz-Krieger algebras and crossed
products by Z. Quite general techniques for constructing spectral triples for these algebras
were developed in a series of papers (in rough chronological order) [29, 56, 31, 30, 57]. We
consider the set up of [56] and [31], which provide a means of lifting data from the (unital)
coefficient algebra of a bi-Hilbertian bimodule to its Cuntz-Pimsner algebra.
We start with a unital, separable C ∗-algebra A, and a finitely generated projective (fgp) bi-
Hilbertian bimodule E over A, i.e. a module fulfilling the conditions of the following definition.
Definition 2.21. An fgp bi-Hilbertian bimodule E over A is an A-bimodule equipped with
the following structures:
• E has both left and right A-valued inner products which induce equivalent norms on E.
• The left and right actions are both injective and adjointable.
• E is finitely generated and projective as both a left and right module.
To separate the left and the right structures, we write EA when we want to emphasize the
right module structure and (··)A for the right inner product. Similarly, AE denotes the left
module defined from E and A(··) the left inner product.
The algebraic Fock space Falg
space FE is defined as the right A-Hilbert C ∗-module completion of Falg
algebra TE ⊆ End∗
for µ ∈ Falg
E is the algebraic direct sum of the A-modules E⊗Ak. The Fock
E . The Cuntz-Toeplitz
A(FE) is the C ∗-algebra generated by the creation operators Tµξ := µ ⊗ ξ
E . The Cuntz-Pimsner algebra OE is defined from the short exact sequence
0 → KA(FE) → TE → OE → 0.
We call this short exact sequence the defining extension of OE.
A set (ej)N
e = Pj ej(eje)A, and similarly for a left frame (fk)N
j=1 ⊂ E of vectors is a frame for the right module EA if for all e ∈ E we have
k=1. Since E is finitely generated and
projective, there exists left and right frames and we can for simplicity assume that they have
the same cardinality. For e and f in the right Hilbert module EA, we denote the associated
rank-one operator by Θe,f := e(f·). Then the frame condition can be expressed as
Θej,ej = IdE
NXj=1
and similarly for fσ. The frame (ej)N
a multi-index and eρ = eρ1 ⊗ ··· ⊗ eρk .
We define the right Watatani index of E⊗k as the element of A given by
j=1 induces a frame for E⊗k
A , namely (eρ)ρ=k where ρ is
eβk = Xρ=k
A(eρeρ) = Xρ′=k−1
A(eρ′ eβeρ′).
(2.6)
19
The right Watatani index is positive, central and since the left action is injective, also invertible.
Therefore βk is a well defined self-adjoint central element in A. The key assumptions we make
concern the asymptotic behaviour of the right Watatani indices. In [56, Section 3.2] we define
an A-bilinear functional Φ∞ : OE → A. This functional gives us an A-valued inner product on
OE. The construction of Φ∞ begins by defining
Φk : End∗
A(E⊗k) → A,
A(T eρeρ).
(2.7)
Φk(T ) = Xρ=k
Here we use the notation End∗
A(E⊗k) for the C ∗-algebra of A-linear adjointable operators on
E⊗k. It follows from [41, Lemma 2.16] that Φk does not depend on the choice of frame. We
note that eβk = Φk(IdE⊗k). Since Φk is independent of the choice of frame, so is eβk . We
extend the functional Φk to a mapping End∗
A(FE) → A by compressing along the orthogonally
complemented submodule E⊗k ⊆ FE. To obtain a good "limiting functional" Φ∞(T ) :=
limk→∞ Φk(T )e−βk on the Cuntz-Toeplitz algebra, we impose the following condition on the
Watatani indices.
Definition 2.22. Let E be an fgp bi-Hilbertian bimodule over the unital C ∗-algebra A. We
say that E is W-regular if for every k ∈ N and ν ∈ E⊗k there exists a ν ∈ E⊗k satisfying
ke−βnνeβn−k − νkE⊗k → 0 as n → ∞.
In [56] the reader can find several examples of Cuntz-Pimsner algebras for which a stronger
version of W-regularity as defined in Definition 2.22 holds. There are no known examples of
modules that are not W-regular. When E is W-regular, [56, Proposition 3.5] guarantees that
limk→∞ Φk(T )e−βk is well defined for T from the ∗-algebra generated by the set of creation
operators {Tν : ν ∈ Falg
E } and is continuous in the C ∗-norm. In Section 6.3 we shall see that
there are ways around W-regularity, and even the existence of an A-valued left inner product,
when constructing semifinite spectral triples giving rise to KMS-states.
We thus obtain a unital positive A-bilinear functional Φ∞ : TE → A. The functional Φ∞
annihilates the compact endomorphisms, and descends to a well-defined functional on the
Cuntz-Pimsner algebra OE. By an abuse of notation, we also denote this functional by Φ∞ :
OE → A. Since Φk and eβk do not depend on the choice of frame, neither does Φ∞. We define
the inner product
S1, S2 ∈ OE.
When computing these inner products, the following fact is useful.
(S1S2)A := Φ∞(S∗
1 S2),
Lemma 2.23. Let E be a W-regular fgp bi-Hilbertian bimodule. For homogeneous elements
µ, ν ∈ Falg
E we have
Φ∞(SµS∗
ν ) = lim
k→∞
A(µe−βk νeβk−ν) = A(µν).
(2.8)
In particular, if S ∈ OE is homogeneous of degree n 6= 0, then Φ∞(S) = 0.
Completing OE modulo the vectors of zero length (with respect to Φ∞) yields a right A-
Hilbert C ∗-module that we denote by ΞA. The module ΞA carries a left action of OE given by
extending the multiplication action of OE on itself. By considering the linear span of the image
of the generators Sν, ν ∈ Falg
E , inside the module ΞA, we obtain an isometrically embedded
and complemented copy of the Fock space. We let Q be the projection on this copy of the
Fock space.
20
Theorem 2.24 (Proposition 3.14 of [56]). Let E be a W-regular fgp bi-Hilbertian bimodule
over a unital C ∗-algebra A. The tuple (OE, ΞA, 2Q−1) is an odd Kasparov module representing
the class of the defining extension
0 → KA(FE) → TE → OE → 0.
(2.9)
To construct an unbounded representative of (OE, ΞA, 2Q − 1), we will add an additional
assumption regarding the fine structure of the operation ν 7→ ν in the definition of W-regularity
(see Definition 2.22). Assuming W-regularity, we can define the operator qk : E⊗k → E⊗k by
qkν := ν = lim
n→∞
e−βnνeβn−k .
Definition 2.25. Let E be an fpg bi-Hilbertian bimodule over the unital C ∗-algebra A. We
say that E is strictly W-regular if it is W-regular and for any k, we can write qk = ckPk =
Pkck where Pk ∈ End∗
A(E⊗k) is a (necessarily A-bilinear) projection and ck is given by left-
multiplication by an element in A.
Remark 2.26. As with W-regularity, the reader can in [56] find several examples of Cuntz-
Pimsner algebras for which strict W-regularity holds. There are no known examples of modules
that are not strictly W-regular.
Remark 2.27. If there is a decomposition qk = ckPk as in the definition of strict W-regularity,
[31, Lemma 3.8] shows that it is unique and of a very specific form. Indeed, each ck is central,
invertible and ck = Φk(Pk)−1. Strict W-regularity is readily verified in practice using [31,
Lemma 3.8]. For instance, if β1 is central for the module action on E, ck = e−βk = e−kβ1 is
central for the module action on E and Pk = 1E⊗k .
When E is strictly W-regular, an unbounded self-adjoint regular operator Dψ on ΞA is con-
structed in [31] making (OE, ΞA, Dψ) into an unbounded Kasparov module representing the
KK-class of the defining extension (2.9). The operator Dψ is of the form
Dψ =Xn∈Z Xr≥max{0,n}
ψ(n, r)Pn,r
where ψ : Z × N → [0,∞) is a function with certain Lipschitz properties (see [31, Remark
3.20]), and the Pn,r are projections on finitely generated projective subspaces, [31]. While the
particular choice of function ψ does not matter much, we will take the function
ψ(n, r) =(cid:26) n
n = r
.
−(2r − n) otherwise
r=max(0,n)Pn,r is the projection onto the A-linear span of {SµS∗
The projections Pn,r form a sequence of mutually orthogonal projections satisfying that the
direct sum ⊕r0
ν : µ − ν =
In particular, Pn,n is the projection onto E⊗n for n ∈ N. More
n, max(0, n) ≤ r ≤ r0}.
precisely, the projections are defined by
Pn,r :=(Qn,r − Qn,r−1,
(2.10)
Qn,r,
r > max{0, n}
r = max{0, n}
21
where the projections Qn,r are defined in terms of the right frame (ej)N
(fj)N
j=1 as
j=1 and the left frame
Qn,r := Xρ−σ=n, ρ=r
ΘW
eρ,c
−1/2
σ
PF fσ
,W
eρ,c
−1/2
σ
PF fσ
(2.11)
where PF = ⊕Pk is the projection on the Fock module coming from Definition 2.25. Here we
have written Wξ,η ∈ ΞA for the element defined from SξS∗
η ∈ OE where ξ ∈ E⊗r and η ∈ E⊗k.
For details, see [31, Lemma 3.10 and Proposition 3.11].
To obtain a semifinite spectral triple, we localize (OE, ΞA, Dψ) in a positive trace on A. Fol-
lowing Proposition 2.12, we consider the semifinite spectral triple
(OE , L2(ΞA, τ ), Dψ, Nτ (ΞA), Trτ ).
Here L2(ΞA, τ ) := ΞA ⊗A L2(A, τ ) and Trτ is the dual trace on Nτ (ΞA) := (End∗
which satisfies Trτ (Θe,f ) = τ ((fe)A), [45, 62].
Lemma 2.28. Assume that the fgp bi-Hilbertian bimodule E is strictly W-regular and that τ
is a positive trace on A. Then the semifinite spectral triple
A(ΞA) ⊗ 1)′′
is Li1-summable.
(OE, L2(ΞA, τ ), Dψ, Nτ (ΞA), Trτ )
Remark 2.29. The assumptions on the existence of limiting behaviour for the Watatani indices
are really just for convenience here. These assumptions relate to existence and behaviour of
norm limits, but we have passed to the 'measurable setting' and so really only need weak limits.
We will explore this point of view in Subsection 6.3.
Proof. We need to prove that the following expression is finite for t large enough:
Trτ (e−tDψ ) =Xn∈Z Xr≥max{0,n}
e−tψ(n,r) Trτ (Pn,r).
By definition (see (2.10)), Pn,r = Qn,r − Qn,r−1 when r > max(0, n) and Pn,r = Qn,r when
r = max(0, n) and Qn,r is defined as in (2.11). Using the computations of [31, Lemma 2.8], we
see that
eρ,c−1/2
σ PF fσW
eρ,c−1/2
σ PF fσ
(2.12)
τ(cid:18)(W
τ ◦ Φ∞(cid:18)S
Trτ (Qn,r) = Xρ−σ=n, ρ=r
= Xρ−σ=n, ρ=r
= Xρ−σ=n, ρ=r
c−1/2
σ PF fσ
S∗
eρSeρS∗
τ (A(PF fσPF fσ(eρeρ)A)) .
)A(cid:19)
σ PF fσ(cid:19)
c−1/2
Using the fact that the elements of the frame have norm bounded by 1, we see that
Trτ (Qn,r) ≤ Xρ−σ=n, ρ=r
≤ Xρ−σ=n, ρ=r
τ (A(PF fσPF fσ(eρeρ)A))
k(A(PF fσPF fσ(eρeρ)A)kA ≤ N 2r−n ≤ elog(N )ψ(n,r),
22
where N is the number of elements in the left frame and the right frame. We can now estimate
(cid:12)(cid:12)(cid:12)Trτ (e−tDψ )(cid:12)(cid:12)(cid:12) =(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xn∈Z Xr>max{0,n}
+(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xn∈Z Xr=max{0,n}
≤2Xn∈Z Xr≥max{0,n}
≤2Xn∈Z Xr≥max{0,n}
e−(t−log(N ))ψ(n,r) < ∞,
e−tψ(n,r) Trτ (Qn,r − Qn,r−1)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
e−tψ(n,r) Trτ (Qn,r)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
e−tψ(n,r) Trτ (Qn,r)
if t > log(N ).
We return to Cuntz-Pimsner algebras below in Example 3.7 (see page 29) and Section 6 (see
page 50). Let us discuss a special case of Cuntz-Pimsner algebras arising on a commutative
coefficient algebra.
Example 2.30. Let Y be a compact Hausdorff space and g : Y → Y a surjective local
homeomorphism. We consider the module Eg = C(Y ) with the bimodule action
(af b)(x) = a(x)f (x)b(g(x)),
a, b ∈ C(Y ), f ∈ Eg.
This is an fgp bi-Hilbertian bimodule in the inner products
C(Y )(f1, f2) = f1f2
and (f1f2)C(Y ) := Lg(f1f2),
where Lg : C(Y ) → C(Y ) is the transfer operator
Lg(f )(x) := Xy∈g−1(x)
f (y).
(2.13)
The Cuntz-Pimsner algebra OEg can be realized as a groupoid C ∗-algebra as in [18] (see also
[19, Theorem 3.2]) over the solenoid
X = {x = y1y2 ··· ∈ Y N : g(yk+1) = yk ∀k}.
The case that X equipped with the shift mapping is a Smale space was studied in [19].
j=1 where ej = √χj for a partition of unity (χj)N
The module Eg has a right frame (ej)N
subordinate to an open covering (Uj)N
j=1 of Y such that gUj is injective for all j. Using this
partition of unity, one sees that βk = 0 for all k. It follows that Eg is a strictly W-regular
module.
The case that g : M → M was a surjective local diffeomorphism acting conformally was con-
sidered in Subsection 2.2.1 (see page 14). However, the spectral triple considered Proposition
2.19 on OEg differs greatly from the semifinite spectral triples considered in Lemma 2.28 -- the
latter are in the image of the boundary mapping in KK1(OEg , C(M )) defined from Equation
(2.9) while the former is not if [ /D] 6= 0 ∈ K ∗(C(M )).
j=1
23
Another class of examples already considered arises from a finite graph G as in Subsection 2.2.2
(see page 17) where the shift mapping σG : ΩG → ΩG is a surjective local homeomorphism
and C ∗(G) ∼= OEσG
. The spectral triples in Proposition 2.20 (see page 18) arises from the
construction of Lemma 2.28 by taking the trace τ : C(ΩG) → C to be defined from point
evaluation in y.
2.2.4 Group C ∗-algebras
We now turn our attention to examples coming from the reduced group C ∗-algebra of a discrete
group. A well known construction associates a spectral triple with a length function on the
group, we consider this example and a semifinite modification thereof which is possible for
a-T-menable groups, i.e. groups with the Haagerup property. The methods extend to a-TT-
menable groups, a class of groups containing all hyperbolic groups, see more in [51, Chapter
7.2].
Let Γ denote a countable discrete group. Recall that a length function ℓ : Γ → R≥0 is a
function satisfying ℓ(e) = 0 for e ∈ Γ the identity, and ℓ(γγ′) ≤ ℓ(γ) + ℓ(γ′) for all group
elements γ, γ′ ∈ Γ. We say that ℓ is a proper length function if ℓ is a proper function, i.e. the
set {γ ∈ Γ : ℓ(γ) ≤ R} is finite for any R ≥ 0. If there exists a constant β ≥ 0 such that
#{γ ∈ Γ : ℓ(γ) ≤ R} = O(eβR) as R → ∞ we say that (Γ, ℓ) has at most exponential growth.
Define the operator Dℓ densely on ℓ2(Γ) as the self-adjoint operator with
Dℓδγ := ℓ(γ)δγ .
The space of compactly supported functions cc(Γ) is a core for Dℓ. We define the ∗-algebra
cb(Γ) ⋊alg Γ as the ∗-algebra generated by multiplication operators (by bounded functions on
Γ) cb(Γ) ⊆ B(ℓ2(Γ)) and all left translation operators.
Proposition 2.31. Let ℓ be a proper length function on Γ. The triple (cb(Γ) ⋊alg Γ, ℓ2(Γ), Dℓ)
is a spectral triple defining the trivial class in the K-homology of the C ∗-algebra cb(Γ) ⋊r Γ.
Moreover, if (Γ, ℓ) has at most exponential growth the spectral triple (cb(Γ) ⋊alg Γ, ℓ2(Γ), Dℓ) is
Li1-summable.
Proof. Since ℓ is proper, it is clear that Dℓ has compact resolvent and if (Γ, ℓ) has at most
exponential growth, then there is C > 0 such that
Tr(e−tDℓ) =Xγ∈Γ
e−tℓ(γ) =
∞Xn=0
#{γ ∈ Γ : ℓ(γ) = n}e−tn ≤ C
∞Xn=0
e−(t−β)n =
C
1 − eβ−t < ∞.
To show that (cb(Γ) ⋊alg Γ, ℓ2(Γ), Dℓ) is a spectral triple, it remains to show that cb(Γ) ⋊alg Γ
preserves the domain of Dℓ and has bounded commutators with Dℓ. Domain preservation is
clear. For an element aλγ ∈ cb(Γ) ⋊alg Γ and a function f ∈ cc(Γ) we compute that
[Dℓ, aλγ]f (g) = a(g)(ℓ(g) − ℓ(γ−1g))f (γ−1g).
It follows that
k[Dℓ, aλγ]kB(ℓ2(Γ) ≤ kakcb(Γ) sup
g∈Γ ℓ(g) − ℓ(γ−1g) ≤ kakcb(Γ)ℓ(γ).
24
The K-homology class of (cb(Γ) ⋊alg Γ, ℓ2(Γ), Dℓ) is trivial. We shall now consider a topolog-
ically more interesting semifinite spectral triple that can be constructed on groups with the
Haagerup property. We are grateful to Branimir ´Ca´ci´c for sharing this construction with us.
Similar ideas appeared in [38, Appendix B].
Let Γ be a discrete group with the Haagerup property. Then there is a proper isometric action
of Γ on a real Hilbert space HΓ. By the Mazur-Ulam theorem there exists an orthogonal
representation
on the Hilbert space HΓ and a proper cocycle cΓ for πΓ, meaning that cΓ : Γ → H is a proper
function satisfying the cocycle identity
πΓ : Γ → O(HΓ)
The cocycle identity allows us to define a length function on Γ by
cΓ(γ1γ2) = cΓ(γ1) − πΓ(γ1)cΓ(γ2).
(2.14)
ℓ(γ) := kcΓ(γ)kHΓ.
Since cΓ is proper, so is ℓ.
Remark 2.32. The existence of a proper isometric action of a group Γ on a Hilbert space is
equivalent to Γ having the Haagerup property, also known as a-T-menability. Our construction
extends to the case when there exists an orthogonal representation πΓ : Γ → O(HΓ) and a
proper quasi-cocycle cΓ : Γ → HΓ. That is, when
Q(cΓ) := sup
γ1,γ2 kcΓ(γ1γ2) − cΓ(γ1) + πΓ(γ1)cΓ(γ2)kHΓ < ∞.
In this case, ℓ(γ) := kcΓ(γ)kHΓ could fail to be a length function but still satisfies ℓ(γγ′) ≤
ℓ(γ) + ℓ(γ′) + Q(cΓ) which suffices for our purposes. The existence of a proper quasi-cocycle on
a Hilbert space is equivalent to Γ being a-TT-menable. Hyperbolic groups are a-TT-menable.
For notational simplicity, we restrict our attention to cocycles.
Definition 2.33. Let Cℓ(HΓ) denote the the complex Clifford algebra of HΓ and assume
that cS : Cℓ(H) → B(SH) is a representation of Cℓ(H). We say that a unitary representation
πS : Γ → U (SH) is a lift of πΓ to SH if for all v ∈ H and g ∈ Γ we have
πS(g)cS (v)πS (g−1) = cS(πΓ(g)v).
(2.15)
When H is finite dimensional this is just the well-known Clifford algebra, but when H is infinite
dimensional we refer to [8, 65] for a description of this algebra.
For a representation SH of Cℓ(H) we consider the new Hilbert space ℓ2(Γ, SH). Assuming that
πS lifts πΓ to SH the Hilbert space ℓ2(Γ, SH) carries a representation of Γ defined by
π : Γ → U (ℓ2(Γ, SH)),
(cid:0)π(g)f(cid:1)(γ) = πS(g)f (g−1γ).
On the Hilbert space ℓ2(Γ, SH) define a self-adjoint operator Dc by declaring
(Dcf )(γ) := cS(cΓγ))f (γ),
f ∈ cc(Γ, SH).
Since cS(v)2 = kvk2
H for all v ∈ H, the domain for Dc can be deduced from
(D2
c f )(γ) = ℓ(γ)2f (γ).
25
(2.16)
(2.17)
The compatibility requirement Equation (2.15) and cocycle property Equation (2.14) imply
that for g, γ ∈ Γ we have
πS(g)cS (cΓ(γ)) = cS (π(g)cΓ(γ))πS (g) = cS(cΓ(g))πS (g) + cS(cΓ(gγ))πS (g).
Then the commutator of Dc and a group element is
([Dc, π(g)]f )(γ) = cS(cΓ(γ))πS(g)f (g−1γ) − πS(g)cS (cΓ(g−1γ))f (g−1γ)
= cS(cΓ(g))πS (g)f (g−1γ) = (cS(cΓ(g))π(g))(f )(γ).
Hence the commutators between Dc and group elements are bounded. It is moreover clear that
these commutators lie in N0 ⋊ Γ where N0 = cS(Cℓ(H))′′. We define N as the von Neumann
algebraic tensor product B(ℓ2(Γ)) ¯⊗N0.
Finally,
(1 + D2
c )−1 ∈ K(ℓ2(Γ)) ⊗ 1 ⊂ K(ℓ2(Γ)) ⊗ Kτ
where Kτ is the compacts in N0 for a choice of normalized positive trace τ . Let Trτ be the
c )−1 ∈ KTrτ . Finally, if ℓ
trace on N defined from the trace τ on N0. We conclude that (1 + D2
has at most exponential growth then Trτ (e−tDc) =Pγ∈Γ e−tℓ(γ) < ∞ for t large enough. As
such, (i ± Dc)−1 ∈ Li1 if ℓ has at most exponential growth. We conclude the following result.
Proposition 2.34. Assume that cS : Cℓ(H) → B(SH) is a representation of Cℓ(H) and that
the unitary representation πS : Γ → U (SH) lifts πΓ to SH. Let cb(Γ) be the (continuous)
bounded functions on Γ, and define a representation of cb(Γ) ⋊ Γ on ℓ2(Γ, SH) by
πS(aλg)f (γ) = a(γ)[π(g)f ](γ),
where π is as in (2.16). Then the triple (cb(Γ) ⋊alg Γ, ℓ2(Γ, SH), Dc, N, Trτ ) is a semifinite
spectral triple which is Li1-summable if ℓ has at most exponential growth.
We return to the example of this subsubsection in Example 3.9 (see page 30) and Subsection
5.3 (see page 48).
Example 2.35. The following example of a proper group cocycle shows the construction's
geometric advantage compared to only using a length function. Consider the trivial action of
the discrete group Γ = Zn on HΓ = Rn. The inclusion Zn ֒→ Rn is additive and proper, and
therefore a proper group cocycle for the trivial action. The semifinite spectral triple associated
with a finite dimensional Clifford representation cS : Rn → EndC(S) can when restricted to
C ∗(Zn) ∼= C(Tn) be identified with the semifinite spectral triple (C ∞(Tn), L2(Tn, S), /DTn ⊗
1S, B(L2(Tn)) ⊗ Cℓn, Trτ ) using Fourier theory on the dual torus Tn = cZn. We observe that
to extract K-homological content from this construction we need to specify a grading if n is
even. In particular, it is unclear how to interpret the construction above in K-homology when
HΓ is infinite-dimensional.
3 KMS states constructed from Li1-summable spectral triples
This section contains the fundamental technical construction of the paper. Starting from a
semifinite Li1-summable spectral triple, we use the associated algebra of Toeplitz operators to
construct an action from the operator D and a KMS-state from the operator D.
26
3.1 The positive part of the spectrum and heat traces
Throughout this section we suppose that (A, H, D, N, T) is a unital semifinite spectral triple.
The spectral triple can be semifinite, in which case we let T denote the given positive faithful
normal semifinite trace. In general T is not unique, and coincides with a non-zero multiple of
the operator trace in the "usual" non-semifinite case N = B(H). We write KN for the compacts
for T. Again, KN coincides with the usual compacts in the case of the type I factor N = B(H).
We write N+ for the von Neumann algebra PDNPD. By an abuse of notation, we write T also
for the induced faithful normal semifinite trace on N+. The T-compacts on N+ will be denoted
by K+
N.
Definition 3.1. The operator D is said to have positive T-essential spectrum if for some
β ∈ [0,∞), we have T(PDe−tD) < ∞ for t > β and T(PDe−tD) ր ∞ as t ց β.
Proposition 3.2. Let D be a self-adjoint densely defined operator affiliated with N satisfying
that (1 + D2)−1/2 ∈ Li1(T). We define the number
βD := inf{t > 0 : T(PDe−tD) < ∞}.
(3.1)
Then βD ∈ [0,∞) and D has positive T-essential spectrum if and only if
lim
tցβD
T(PDe−tD) = ∞.
In particular, if βD = 0 then D has positive T-essential spectrum if and only if T(PD) = ∞.
Proof. By definition, if (1 + D2)−1/2 ∈ Li1 then T(e−tD) < ∞ for t large enough. Therefore,
T(PDe−tD) = T(PDe−tD) < ∞ for t large enough and βD := inf{t > 0 : T(PDe−tD) < ∞} will
be a number in [0,∞). By definition, T(PDe−tD) < ∞ for t > βD and if limtցβD T(PDe−tD) =
∞ then D has positive T-essential spectrum with β = βD. Conversely, if D has positive T-
essential spectrum there is a β ∈ [0,∞) with T(PDe−tD) < ∞ for t > β and T(PDe−tD) ր ∞
as t ց β, and in this case it is clear that β = βD.
Remark 3.3. We will often impose the assumption of positive T-essential spectrum. If for some
β ∈ [0,∞), T((1 − PD)etD) < ∞ for t > β and T((1 − PD)etD) ր ∞ as t ց β, we can equally
well use −D in our construction.
Proposition 3.4. Let D be a self-adjoint densely defined operator affiliated with N satisfying
that (1 + D2)−1/2 ∈ Li1(T). Then D has positive T-essential spectrum if and only if both of the
following conditions fail:
1. PD has finite T-trace.
2. There exists a p > 0 such that PDe−D ∈ Lp(T) \ ∩q>pLq(T).
The reader should note that conditions 1. and 2. are mutually exclusive.
Proof. If D has positive T-essential spectrum, then clearly 1. fails. Also 2. fails if D has positive
T-essential spectrum because condition 2. is equivalent to βD = p and limtցp T(PDe−tD) being
finite.
27
fails, then either βD = 0 and limtց0 T(PDe−tD) must be
Conversely, if Condition 1. and 2.
infinite not to violate T(PD) being infinite or βD > 0 and the set {p > 0 : PDe−D ∈ Lp(T)} is
open (due to condition 2. failing) showing that T(PDe−tD) ր ∞ as t ց βD.
Example 3.5. Dirac operators on closed manifolds, as considered in Subsection 2.2.1 (see
page 14), have positive essential spectrum.
In this case, we can compute βD = 0 and the
leading term in the heat trace asymptotics using Proposition 2.14. If (C ∞(M ), L2(M, S), /D)
is the spectral triple associated with a Dirac operator, Proposition 2.14 implies that
TrL2(M,S)(PDe−t /D) = n!cnt−nZS∗M
TrS(p /D)dV + O(t−n+ǫ),
as t → 0,
2 (cS(ξ) + 1) for x ∈ M and ξ ∈ S∗
where p /D is the principal symbol of the zeroth order pseudo-differential operator P /D. A
direct computation shows that p /D(x, ξ) = 1
xM . Here
cS : T ∗M → End(S) denotes Clifford multiplication. More generally, Proposition 2.14 allows
us to conclude that for a ∈ C ∞(M ),
TrL2(M,S)(PDae−t /D) = n!cnt−nZS∗M
= n!cnt−nZMZS∗
= cnt−nZM
TrS(p /D)a dV + O(t−n+ǫ) =
xM (ξ)dV (x) + O(t−n+ǫ) =
a(x)TrS(p /D(x, ξ)) dVS∗
a dV + O(t−n+ǫ),
as t → 0,
xM
for a new constant cn > 0, depending only on the dimension of M and the rank of S.
In
the last equality we used that p /D − 1/2 is an antisymmetric function under the involution
(x, ξ) 7→ (x,−ξ) of S∗M and therefore
ZS∗
xM
TrS(p /D(x, ξ)) dVS∗
TrS(p /D(x,−ξ)) dVS∗
xM =
xM =ZS∗
xM
=
rank(S)
2
ZS∗
xM
dVS∗
xM =
rank(S)πdim(M )/2
Γ(dim(M )/2)
.
Example 3.6. The spectral triples for graph C ∗-algebras from Proposition 2.20 (see page 18)
also have positive essential spectrum. We compute βD and the heat trace asymptotics assuming
that G is primitive. In this example, PDy is the projection onto the subspace ℓ2(Vy ∩ κ−1
G (0)).
We use the notation
y := Vy ∩ κ−1
V+
G (0) = {(x, n) ∈ Vy : n ≥ 0, σn
G(x) = y}.
(3.2)
. Note that
G(x) = y then x is uniquely determined by y and a finite path σ = σ1σ2 ··· σn with x = σy.
G ({y}).
The space PDy ℓ2(Vy) is therefore spanned by the orthonormal basis (δ(x,n))(x,n)∈V+
if σn
Note that paths of the form x = σy, with s(σn) = r(y), exhaust all possible x ∈ σ−n
Using that PDy Dyδ(x,n) = nδ(x,n) for (x, n) ∈ V+
y , we compute that
y
Tr(PDy e−tDy) =
∞Xn=0 Xx∈σ−n
G ({y})
e−tn =
∞Xn=0
#{σ ∈ En : s(σn) = r(y)}e−tn.
28
Let A denote the edge adjacency matrix of G and rσ(A) its spectral radius. If G is primitive,
(i.e. all entries of Ak are positive for some integer k > 0) we let w ∈ CE its ℓ2-normalized
Perron-Frobenius vector. It follows from [56, Lemma 3.7] that there is an α0 ∈ [0, 1) such that
#{σ ∈ En : s(σn) = r(y)} = kwkℓ1 wr(y)rσ(A)n + O((α0rσ(A))n),
as n → ∞. We can conclude that there is a function f holomorphic in Re(t) > log rσ(A)+log α0
such that
Tr(PDy e−tDy) = kwkℓ1 wr(y)
1 − rσ(A)e−t + f (t).
Therefore, Tr(PDy e−tDy) −
log α0 whenever G is primitive.
kwkℓ1 wr(y)
t−log(rσ(A)) has a holomorphic extension to Re(t) > log rσ(A) +
More generally, if G is primitive, the method above shows that for two finite paths µ and ν we
can compute that
Tr(PDy SµS∗
ν e−tDy) =δµ,ν
kS∗
G ({y})
ℓ2e−tn
µδ(x,n)k2
∞Xn=0 Xx∈σ−n
∞Xn=µ
#(cid:8)σ ∈ En−µ : r(σ1) = s(µ), s(σn−µ) = r(y)(cid:9)e−tn
µ−1Xn=0 Xx∈σ−n
kS∗
µδ(x,n)k2
ℓ2e−tn
G ({y})
=
+ δµ,ν
=δµ,νws(µ)wr(y)
e−tµ
1 − rσ(A)e−t + δµ,ν fµ,y(t),
for a function fµ,y holomorphic in Re(t) > log rσ(A) + log α0. We conclude that
Tr(PDy SµS∗
ν e−tDy) − δµ,νwd(µ)wr(y)
rσ(A)−µ
t − log(rσ(A))
,
has a holomorphic extension to Re(t) > log rσ(A) + log α0 whenever G is primitive. As such,
βDy = log(rσ(A)) and Dy has positive essential spectrum.
Example 3.7. Let OE be a Cuntz-Pimsner algebra defined from a strictly W-regular (recall
Definition 2.25) finitely generated and projective bi-Hilbertian bimodule EA and a positive
trace τ on the coefficient algebra A. The semifinite spectral triple considered in Lemma 2.28
(see page 22) also has positive Trτ -essential spectrum assuming a criticality condition on τ
that we formulate below (see Definition 6.2 on page 51). The heat trace asymptotics are
slightly more involved, and we compute these explicitly in Subsection 6.2 under a condition
on τ previously studied by Laca-Neshveyev [45] in the context of KMS-states. However, for a
general τ we can proceed as in the proof of Lemma 2.28 to deduce the following.
Proposition 3.8. For any strictly W -regular fgp bi-Hilbertian bimodule E over the unital
C ∗-algebra and a positive trace τ on A,
Trτ (PDψ e−tDψ ) =
e−tnτ∗(E⊗An),
∞Xn=0
29
where τ∗ : K0(A) → R denotes the map induced by τ on K-theory. In particular, Trτ (PDψ e−tDψ )
does not depend on the choice of inner products on E but only on τ and the bimodule structure
on E.
Proof. We compute that
Trτ (PDe−tDψ ) =
∞Xn=0
e−tψ(n,n) Trτ (Pn,n) =
e−tn Trτ (Qn,n) =
∞Xn=0
∞Xn=0 Xρ=n
e−tnτ ((eρeρ)A).
On the other hand, τ∗(E⊗An) = (τ ⊗ TrMN(n))(pE⊗An) where pE⊗An ∈ MN (n)(A) is a projection
representing E⊗An. Using the choice of frame (ej)N
j=1, we can take N (n) := N n and pE⊗An :=
((eµeν )A)µ=ν=n. In this choice of representing projection,
τ∗(E⊗An) = (τ ⊗ TrMN(n))(pE⊗An) = TrMN(n)((τ ((eµeν )A)µ=ν=n)) = Xρ=n
τ ((eρeρ)A).
This computation shows that it is in general difficult to compute βD. In this case βD depends
on the asymptotic properties of the sequence (τ∗(E⊗An))n∈N as n → ∞.
For a simple tensor σ ∈ E⊗m write σ = σσ, where the initial segment σ will be of a length
understood from context (σ = µ in the next computation). With this notation, we can
ν ∈ OE, where µ ∈ E⊗k, ν ∈ E⊗l are simple tensors.
compute our functional on a typical SµS∗
We find
Trτ (PDSµS∗
ν e−tDψ) = δµ,ν
= δµ,ν
= δµ,ν
∞Xn=0
∞Xn=µ Xσ=n
∞Xn=µ Xσ=n
e−tn Trτ (SµS∗
ν Qn,n)
(3.3)
e−tnτ ((S∗
µeσS∗
ν eσ)E⊗(n−µ))
e−tnτ ( ( (µeσ)Eµeσ (νeσ)Eµeσ)E⊗(n−µ)).
Example 3.9. Consider a length function ℓ on a countable group Γ as in Subsection 2.2.4.
Definition 3.10. We define the critical value of (Γ, ℓ) as
β(Γ, ℓ) := inf{t ≥ 0 :Xγ∈Γ
e−tℓ(γ) < ∞}.
IfPγ∈Γ e−tℓ(γ) ր ∞ as t ց β(Γ, ℓ), we say that ℓ is critical.
It follows directly from Definition 3.1 that the operator Dℓ appearing in Proposition 2.31 (see
page 24) has positive essential spectrum as long as ℓ is critical. The heat trace of an element
aλg ∈ cb(Γ) ⋊alg Γ is given by
Tr(PDℓ aλge−tDℓ) = Tr(aλge−tDℓ) = δe,gXγ∈Γ
a(γ)e−tℓ(γ).
30
Similar computations can be carried out for the semifinite spectral triple constructed in Propo-
sition 2.34 using a Hilbert space valued cocycle cΓ (see page 26). Note that
Therefore, the heat trace of an element aλg ∈ cb(Γ) ⋊alg Γ is given by
+ 1(cid:19) .
PDc f (g) =
1
kcΓ(g)kHΓ
2(cid:18) cS(cΓ(g))
2Xγ∈Γ(cid:28)δγ, τ(cid:18) cS(cΓ(g))
δe,gXγ∈Γ
kcΓ(g)kHΓ
1
a(γ)e−tℓ(γ) =
2
Trτ (PDc aλge−tDc) =
=
1
1
2
+ 1(cid:19) a(g−1γ)δgγ(cid:29) e−tℓ(γ) =
Tr(PDℓ aλge−tDℓ).
Here we use that τ (cS(v)) = 0 for any v ∈ HΓ which holds due to the fact that we can pick a
w ∈ HΓ orthogonal to v and compute that
τ (cS (v)) = τ (cS (w)cS (v)cS (w)) = −τ (cS (w)2cS(v)) = −τ (cS(v)).
If the length function ℓ(γ) := kcΓ(γ)kHΓ associated with cΓ is critical, we say that cΓ is
critical. We conclude that the semifinite spectral triple from Proposition 2.34 has positive
essential spectrum if cΓ is critical.
3.2 To-plitz or not To-plitz
We proceed under the same working conditions as in the previous section to construct states
from spectral triples. Recall that N+ = PDNPD and that K+
N = PDKNPD.
Definition 3.11. Let (A, H, D, N, T) be a semifinite spectral triple. We define the Toeplitz
algebra of (A, H, D, N, T) as
The saturated Toeplitz algebra of (A, H, D, N, T) is defined by
TA := PDAPD + K+
N ⊆ N+.
TA,D := C ∗ [s∈R
eisDTAe−isD! = C ∗ [s∈R
eisDPDAPDe−isD + K+
N! ⊆ N+.
Proposition 3.12. The Toeplitz algebra TA of a unital semifinite spectral triple (A, H, D, N, T)
is a C ∗-algebra.
Proof. It follows from Lemma 2.11 that [PD, a] ∈ KN for all a ∈ A. Therefore, the mapping
βD : A → N+/K+
N,
is a ∗-homomorphism and βD(A) ⊆ N+/K+
the preimage of βD(A) under the quotient mapping N+ → N+/K+
algebra.
a 7→ PDaPD mod K+
N,
N is a closed C ∗-subalgebra. By definition, TA is
N and is therefore a C ∗-
The reader should note that Proposition 3.12 also holds in the non-unital setting because it
only relies on Lemma 2.11 which holds non-unitally.
31
Proposition 3.13. The saturated Toeplitz algebra of (A, H, D, N, T) carries an R-action σ+ :
R → Aut(TA,D) defined by
σ+
s (T ) := PDeisDT e−isDPD = PDeisDT e−isDPD,
s ∈ R, T ∈ TA,D.
This proposition is a consequence of that TA,D is constructed as the saturation of TA under
the action σ+ extended to N+. We note that if we identify TA,D with a subalgebra of N, we
can also write σ+
s (T ) := eisDT e−isD = eisDT e−isD.
Define a one-parameter family of states (φt,0)t>β on TA,D by
φt,0(T ) :=
T(PDT e−tD)
T(PDe−tD)
.
We shall compose the family (φt,0)t>β with an "extended limit" as t → β:
Definition 3.14. An extended limit as t → β is a state ω ∈ L∞(β,∞)∗ such that ω(f ) = 0
whenever limt→β f (t) = 0. For an extended limit ω and f ∈ L∞(β,∞), we write
Let ω be any extended limit and define
lim
t→ω
f := ω(f ).
φω,0 : TA,D → C,
φω,0(T ) := ω ◦ φt,0(T ) = lim
t→ω
φt,0(T ).
Lemma 3.15. Let (A, H, D, N, T) be a semifinite Li1-summable spectral triple with positive
T-essential spectrum (see Definition 3.1 on page 27). For any extended limit ω, the functional
φω,0 is a state on TA,D. Moreover φω,0(T ) = 0 for all T ∈ K+
N.
Proof. It is immediate that φω,0 is a state. For the statement that φω,0(T ) = 0 for all T ∈ K+
N,
we observe that since φω,0 is a state, it is also norm-continuous. It therefore suffices to prove
that φω,0(T ) = 0 for all projections T ∈ N+ with T(T ) < ∞. For such T , we can estimate that
φt,0(T ) =
T(PDT e−tD)
T(PDe−tD) ≤
T(T )
T(PDe−tD)
.
Since, T(PDe−tD) ր ∞ as t ց β, it follows that limt→β φt,0(T ) = 0. We conclude that for all
projections T ∈ N+, with T(T ) < ∞, and any extended limit ω as t → β, ω ◦ φt,0(T ) = 0.
Due to Lemma 3.15, we can make the following definition.
Definition 3.16. Define the C ∗-algebra AD := TA,D/K+
N and the state φω on AD as
φω(T mod K+
N) := φω,0(T ).
The state φω,0 also restricts to a state on TA, and Lemma 3.15 implies that φω,0TA descends
to a state on A via the ∗-epimorphism βD : A → TA/K+
N (see the proof of Proposition 3.12
on page 31). To analyse the situation of the state on AD versus that on A, we consider the
following ideal
I := {a ∈ A : PDaPD ∈ K+
N}
32
so that PDIPD = PDAPD ∩ K+
N. Since TA ⊆ TA,D, we obtain a commuting diagram
0
0
/ K+
N
=
/ K+
N
TA
A/I
/ 0
/ TA,D
/ AD
/ 0,
with exact rows. The mapping A/I → AD is indeed injective by the four lemma. We identify
A/I with a subalgebra of AD. The induced mapping γ : A → AD is compatible with the states
φω and φω,0 in the sense that φω,0(PDaPD) = φω(γ(a)) for a ∈ A.
The R-action σ+
sition follows from the construction of TA,D as the saturation of TA under σ+.
s (T ) := eisDT e−isD on TA,D induces an R-action on A/I. The following propo-
Proposition 3.17. Let β ∈ R. The algebra AD carries an R-action σ : R → Aut(AD) defined
by declaring the quotient mapping TA,D → AD to be equivariant. The C ∗-algebra AD is the
saturation of A/I under the action σ, i.e. AD is generated in N+/K+
N by ∪s∈Rσs(A/I).
The aim of our construction is to obtain a KMS state on A, or, failing that, on A/I. As a
first step we introduce conditions ensuring that we at least get a KMS state on TA,D, and so
on AD. To this end we make the following assumption
Definition 3.18. Let (A, H, D, N, T) be a semifinite spectral triple. We say that a subset
S ⊂ A is analytically generating at β if the set PDSPD + K+ generates the Toeplitz algebra
TA as a C ∗-algebra and satisfies that eβDPDSPDe−βD ⊂ N+.
We say that the semifinite spectral triple (A, H, D, N, T) is β-analytic if it admits an analyti-
cally generating set at β.
We note that this condition is empty if β = 0. The condition of being β-analytic is just
requiring that we have enough analytic elements in TA,D to verify the KMS condition. Indeed
we have the following.
Proposition 3.19. Let (A, H, D, N, T) be a semifinite spectral triple and β ∈ R. Consider the
following statements:
i) The semifinite spectral triple (A, H, D, N, T) is β-analytic.
ii) There is a dense σ+-invariant subspace T 0
A,D ⊆ TA,D of elements satisfying that for any
t (T ) has a bounded holomorphic extension
A,D the function fT : R → N, fT (t) := σ+
T ∈ T 0
to the strip {z ∈ C : Im(z) ∈ (−β, 0)}.
iii) There is a dense σ-invariant subspace A0
D ⊆ AD of elements satisfying that for any
a ∈ AD the function fa : R → N, fa(t) := σt(a) has a bounded holomorphic extension to
the strip {z ∈ C : Im(z) ∈ (−β, 0)}.
It holds that i) implies ii) which implies iii). If I := {a ∈ A : PDaPD ∈ K+
i).
N} = 0, iii) implies
33
/
/
/
/
/
_
_
/
/
/
/
/
Proof. It is clear that i) implies ii) since for an analytically generating set S at β we can take
A,D to be the σ+-invariant ∗-algebra generated by PDSPD ∪ FD, where FD ⊆ K+
T 0
N is the dense
two sided ideal in N generated by the spectral projections of D over compact intervals in R.
The implication ii)⇒iii) is seen from taking A0
If I = 0, the set S = A0
a bounded operator in N. Thus eβDPDSPDe−βD ⊂ N+ and PDSPD + K+
TA. We conclude that S is analytically generating at β and that iii) implies i) if I = 0.
D ∩ A is dense in A, and for every s ∈ S we find that eβDse−βD is
N plainly generates
A,D ∩ K+
N).
D := T 0
A,D/(T 0
Proposition 3.20. Let (A, H, D, N, T) be a βD-analytic Li1-summable spectral triple with pos-
itive T-essential spectrum. For any extended limit ω as t → βD, the state φω,0 is a βD-KMS
state on TA,D for the one-parameter group σ+.
Proof. It follows from Proposition 3.19 that the dense subalgebra T 0
A,D ⊆ TA,D consists of
βD-analytic elements of TA,D. The twisted trace property relative to the one-parameter group
σ+ holds on T 0
A,D by direct computation: for all T1, T2 ∈ T 0
A,D
φω(T1T2) = lim
t→ω
T(T1T2e−tD)
T(PDe−tD)
= lim
t→ω
T(etDT2e−tDT1e−tD)
T(PDe−tD)
= φω(σ−iβ(T2)T1).
By definition, φω,0 is a βD-KMS-state for σ+.
Corollary 3.21. Let (A, H, D, N, T) be a βD-analytic Li1-summable spectral triple with positive
T-essential spectrum. For any extended limit ω as t → βD, the state φω is a βD-KMS state on
AD for the one-parameter group σ, defined in Proposition 3.17.
Proof. This follows from Proposition 3.19 and Proposition 3.20 because φω,0 vanishes on the
compacts and the fact that φω is induced from φω,0.
In practice, for a βD-analytic Li1-summable spectral triple with positive T-essential spectrum,
we will want to check that in fact TA,D = TA.
In this case AD = A/I and φω induces a
KMS-state on A/I. In the often occuring special case PDAPD ∩ K+
N = 0, we obtain a KMS
state on the algebra A. In practice, these things are all checkable and we will do so in several
examples in the subsequent sections.
The special case β = 0 has been addressed by Voiculescu, [63, Proposition 4.6] under the
assumption that D is positive. Exponential β-compatibility is superfluous when βD = 0. By
Proposition 3.2, when βD = 0, positive T-essential spectrum is equivalent to T(PD) = ∞.
Theorem 3.22. Let (A, H, D, N, T) be a unital Li1-summable semifinite spectral triple with
βD = 0 and T(PD) = ∞. Then A has a tracial state. Indeed, for any extended limit ω as
t → 0,
is a tracial state.
φω(a) := lim
t→ω
T(PDae−tD)
T(PDe−tD)
Theorem 3.22 applies to unital Li(0),1-summable semifinite spectral triples with T(PD) = ∞
since Li(0),1-summability implies βD = 0.
34
Remark 3.23. If the heat trace has an asymptotic expansion as in Theorem 2.10, then Theorem
3.22 can be further simplified. Assume that there is a p > 0 and that for any a ∈ A, there is a
φ0(a) ∈ C such that φ0(1) 6= 0 and
T(PDae−tD) = φ0(a)t−p + O(t−p+ǫ),
as t → 0,
for some ǫ > 0 (which can depend on a). Since φ0(1) 6= 0, it follows that φ0 is continuous in
the C ∗-norm on A and φω(a) = φ0(a)
φ0(1) for all a ∈ A. In fact, φ0(a) = Γ(p)Resz=pζ(z; PDa,D).
The construction of the state φω in Corollary 3.21 involves the operator PD. We shall now
provide a result allowing us to remove PD from the definition of φω in the presence of certain
symmetries on the spectral triple. The result provides a checkable set of conditions to compute
φω by means of asymptotics of e−tD.
Lemma 3.24. Let (A, H, D, N, T) be a unital Li1-summable semifinite spectral triple and β ≥ 0
a number such that T(e−tD) < ∞ for t > β and T(e−tD) ր ∞ as t ց β. Assume that there
exists self-adjoint operators γ1, . . . , γN ∈ N ∩ A′ such that
1. PN
2. γj Dom(D) ⊆ Dom(D) and [D, γj]+ := Dγj + γj D has a bounded extension to H;
3. For j = 1, . . . , N and some ǫ > 0, the function t 7→ e−tDγjetD extends to a norm
j is invertible;
j=1 γ2
continuous function from the interval [β, β + ǫ) to N with
e−tDγjetD = γj.
lim
t→β
Then βD = β and D has positive T-essential spectrum. Moreover, for any extended limit ω as
t → βD,
φω(a) = lim
t→ω
T(ae−tD)
T(e−tD)
.
Example 3.25. Before proceeding with the proof of the lemma, we give some examples of how
the operators γ1, . . . , γN ∈ N ∩ A′ can arise. The most trivial instance is when (A, H, D, N, T)
is even, in which case the grading γ will satisfy the conditions of Lemma 3.24.
Here is a more geometric example. Let (C ∞(M ), L2(M, S), /D) denote the spectral triple de-
fined from a Dirac operator on a closed Riemannian manifold M as in Proposition 2.14 (see
page 14). If we take a collection X1, . . . , XN ∈ C ∞(M, T M ) of vector fields spanning the tan-
gent bundle T M in all points, the collection of Clifford multiplication operators γj := cS(Xj),
j = 1, . . . , N is readily verified to satisfy the conditions of Lemma 3.24. This construction ex-
tends to semi-finite spectral triples defined from the fibrewise Dirac operator of a Riemannian
spinc submersion π : M → B (see more in [40]) and a measure on B by taking X1, . . . , XN to
be vertical vector fields spanning the vertical tangent bundle ker dπ in all points of M .
Proof. Define C := k(PN
j )−1k and recall that FD := DD−1 modulo a finite trace projec-
j=1 γ2
35
tion. For any self-adjoint a ∈ A, we estimate
T(FDae−tD) =
C
2
γj(FDa + aFD)e−tDγj =
γ2
1
2
+
−
+
=
C
2
C
2
C
2
C
2
C
2
= −
T
T(cid:16)(FDa + aFD)e−tD(cid:17) ≤
NXj=1
γj(FDa + aFD)γje−tD +
T
NXj=1
γj(FDa + aFD)(e−tDγjetD − γj)e−tD =
T
NXj=1
j (FDa + aFD)e−tD +
T
NXj=1
γj([FD, γj]+a + a[FD, γj]+)γje−tD
T
NXj=1
γj(FDa + aFD)(e−tDγjetD − γj)e−tD ≤
T
NXj=1
T
γj(FDa + aFD)(e−tDγjetD − γj)e−tD
NXj=1
γj([FD, γj]+a + a[FD, γj]+)γje−tD = o(T(e−tD)),
τ
NXj=1
γj(FDa + aFD)(e−tDγjetD − γj)e−tD = o(T(e−tD)),
τ
NXj=1
≤ − T(cid:16)FDae−tD(cid:17) −
T
NXj=1
+
C
2
C
2
γj([FD, γj]+a + a[FD, γj]+)γje−tD
Since [D, γj ]+ is bounded, [FD, γj]+ is compact and an approximation argument by finite T-
rank operators shows that
as t ց β. By norm continuity of t 7→ e−tDγjetD we can also deduce that
as t ց β. In conclusion, for a self-adjoint a,
T(FDae−tD) = −T(FDae−tD) + o(T(e−tD)).
We can conclude that τ (FDae−tD) = o(T(e−tD)) as t ց β. Since 2PD = FD + 1, we have for
any a ∈ A that
T(ae−tD) = 2T(PDae−tD) + o(T(e−tD)),
as t ց β.
36
In particular β = βD and T(PDe−tD) ր ∞ as t ց β. We compute that
T(ae−tD)
T(e−tD)
=
T(PDae−tD)
T(PDe−tD)
+ o(1),
as t ց β.
In particular, for any extended limit ω as t → β,
T(ae−tD)
T(e−tD)
lim
t→ω
= lim
t→ω
T(PDae−tD)
T(PDe−tD)
.
This concludes the proof of the lemma.
3.3 Modular spectral triples and modular index theory
Modular spectral triples and their (equivariant) index theory were considered in [10, 12, 7],
with the definition laid out most clearly in [58, Definition 2.1]. These were defined in order
to study the (equivariant) index theory of KMS weights associated to periodic flows, so one
might wonder how modular spectral triples fit into our scheme.
Given a KMS state ψ : B → C with inverse temperature β for a one-parameter group σ :
R → Aut(B) on a unital C ∗-algebra and a faithful expectation onto the fixed point algebra
Φ : B → Bσ, we can emulate the constructions that inspired the definition of modular spectral
triples.
First we construct the right C ∗-module X over Bσ by completing B in the norm coming from
the inner product
(xy)Bσ = Φ(x∗y).
Then we can use [45] to construct Trψ : KBσ (X) → C the trace dual to ψBσ . The action σ
induces a one parameter unitary group on L2(X, ψ), and we let D be the generator of this one
parameter group. By [7], when the action σ is periodic, the data
(B, L2(X, ψ), D, KBσ (X)′′, φD),
where φD(a) := Trψ(e−βD/2ae−βD/2), gives us a modular spectral triple. The operator D is
affiliated to the semifinite algebra (K(X)σφD )′′
Proposition 3.26. Let (A, H, D, N, T) be a semifinite spectral triple such that that for all
t > β
T(PDe−tD) < ∞
and limtցβ T(PDe−tD) = ∞. Assume that PDAPD ∩ KN = {0} and the spectral triple is β-
analytic. Then we obtain the KMS state φω : AD → C, and provided that D has discrete
spectrum we obtain an expectation Φ : AD → Aσ
Provided that the group σ is periodic we then obtain a finitely summable modular spectral triple
D onto the fixed point subalgebra.
(AD, L2(AD, φω), D, (KAσ
D (AD))′′, φD),
where the operator D generates the one-parameter group induced by σ on L2(AD, φω) and
φD := Trφω (e−βD/2 · e−βD/2). The spectral dimension is 1.
37
Proof. The existence of the KMS state φω comes from Corollary 3.21.
In general the action σ is real, but assuming that the operator D has discrete spectrum, the
action will factor through a (compact) torus. To see this, one takes a rational basis of the
eigenvalues (possibly an infinite basis), and takes a product over the circles corresponding to
these individual actions.
Consequently, by averaging over this torus, there is an expectation Φ : N → Nσ onto the fixed
point algebra for t 7→ (T 7→ eitDT e−itD). Of course D is affiliated to the fixed-point algebra.
Finally if the action σ is periodic then [7] proves that we have a modular spectral triple, and
that φD((1 + D2)−s/2) < ∞ for s > 1.
For circle actions there is a local index formula in twisted cyclic theory, but for real actions
factoring through a torus there is not.
One serious issue that comes up in this more general setting is the compactness of the resolvent
of D, and determining summability. We leave this issue to another place.
4 The KMS-state φω and Dixmier traces
Dixmier traces. For a decreasing function ψ : [0,∞) → (0,∞) we denote Ψ(t) :=R t
In the present section we discuss a relation between the trace φω from Theorem 3.22 and
0 ψ(s)ds.
Let Lψ(T) be the principal ideal defined as in Definition 2.6 (see page 9). Let ET be the
spectral projection of an operator T affiliated with N and let nT(s, T ) := T(ET (s,∞)) be its
distribution function.
The following result extends [48, Lemma 12.6.3].
Lemma 4.1. Let ψ be a regularly varying function of index −1. Let T ∈ Lψ(T) be strictly
positive and µT(t, T ) ∼ ψ(t), t → ∞. For every q > 0 we have
T(e−T −q/t) ∼ Γ(1 +
1
q
)ψ−1(t− 1
q ), t → ∞.
Proof. The assumption µT(t, T ) ∼ ψ(t) implies µT(t, T q) ∼ [ψ(t)]q, t → ∞. Since the distribu-
1
tion function is an inverse of the singular values function, it follows that nT(s, T q) ∼ ψ−1(s
q ),
s → 0+. Next we have T(ET −q (t)) ∼ nT(1/t, T q), t → ∞. Thus,
q ), t → ∞.
T(ET −q (t)) ∼ ψ−1(t− 1
Since ψ is varying regularly with index −1, [3, Theorem 1.5.12] implies that ψ−1 varies regularly
with index −1, too. Thus, T(ET −q ) varies regularly with index 1
q . Writing the heat trace as a
Laplace transform
T(e−T −q/t) =Z ∞
0
e−z/t dT(ET −q (z))
and using the Karamata theorem [44, Chapter IV, Theorem 8.1] we obtain
T(e−T −q/t) ∼ Γ(1 +
1
q
)ψ−1(t− 1
q ), t → ∞.
38
Let Pa : L∞(0,∞) → L∞(0,∞) be the exponentiation operator defined by (Paf )(t) :=
f (ta), t > 0.
Definition 4.2. [11] An extended limit ω as t → ∞ on L∞(0,∞) is said to be exponentiation
invariant if
lim
t→ω
(Paf )(t) = lim
t→ω
f (t)
for every f ∈ L∞(0,∞) and every a > 0.
Definition 4.3. Let ψ : [0,∞) → (0,∞) be a regularly varying function of index −1. For any
extended limit ω as t → ∞ on L∞(0,∞) a linear extension of the weight
Tω,ψ(T ) := lim
t→ω
µT(s, T ) ds,
0 ≤ T ∈ Lψ(T)
0
1
Ψ(t)Z t
is said to be a Dixmier trace on Lψ(T).
Remark 4.4. Usually Dixmier traces are defined on Lorentz ideals corresponding to the function
Ψ (which are strictly larger than Lψ(T)) by exactly the same formula as in Defninition 4.3 (see
e.g. [20, 15, 48]). Then, Dixmier traces on Lψ(T) are restrictions of those on Lorentz ideal to
Lψ(T). Since we do not deal with Lorentz ideals here, it is convenient to define Dixmier traces
directly on Lψ(T).
Also, it should be pointed out that on Lorentz ideals to define Dixmier traces one needs an
additional assumption on ω: either dilation invariance [20, 48] or exponentiation invariance
[25, 61]. As it was shown in [60, Theorem 17] these requirements are redundant on Lψ(T).
The proof of the following theorem is the same as that of [48, Theorem 8.5.1] and thus omitted.
Note however, that in [48] the result was proved for Lorentz ideals and required dilation
invariance of the extended limit ω. For the case of Lψ(T) one can refer to [60, Lemma 15] to
remove this assumption.
Theorem 4.5. Let f ∈ C 2[0,∞) be a bounded function such that f (0) = f ′(0) = 0. Let
T ∈ Lψ(T) be positive and let B ∈ N. For every extended limit ω as t → ∞ on L∞(0,∞) we
have
t→ω(cid:18) 1
Ψ(t)Z t
lim
1
T(f (sT )B)
ds
s2(cid:19) =Z ∞
0
f (s)
ds
s2 · lim
t→ω(cid:18) 1
Ψ(t)Z t
1
T(e−(sT B)−1
)
ds
s2(cid:19) .
Below we will need the relation between generalised heat kernels and Dixmier traces on Lψ(T),
which was proved in [25] under the additional assumption that
AΨ(α) := lim
t→∞
Ψ(tα)
Ψ(t)
exists for every α > 0.
(4.1)
Recall the notation Ψ(t) =R t
0 ψ(s)ds. The corresponding (natural) assumption on ψ is that
α · lim
t→∞
ψ(tα)tα−1
ψ(t)
exists for every α > 0.
(4.2)
The number appearing in Equation (4.2) equals AΨ(α). Note that condition (4.2) implies the
condition (4.1) and regular variation of ψ with index −1 [25, Proposition 1.7].
39
Let H : L∞(0,∞) → L∞(0,∞) be the Cesaro mean defined as follows:
(Hf )(t) :=
f (s) ds, t > 0.
1
tZ t
0
Let MΨ : L∞(0,∞) → L∞(0,∞) be the Cesaro mean twisted by Ψ, that is
(MΨf )(t) : = [(H(f ◦ Ψ−1)) ◦ Ψ](t) =
Lemma 4.6. If ψ satisfies condition (4.2), then
1
Ψ(t)Z t
0
f (s)ψ(s) ds, t > 0.
MΨ ◦ Pa − Pa ◦ MΨ : L∞(0,∞) → C0(0,∞)
for every a > 0.
Proof. For f ∈ L∞(0,∞) we have
Ψ(t)Z t
(MΨ ◦ Paf )(t) =
1
0
f (sa)ψ(s) ds =
1
Ψ(t)Z ta
0
f (s)ψ(s1/a)
s1/a−1
a
ds.
Using conditions (4.1) and (4.2) we obtain
(MΨ ◦ Paf )(t) ∈ AΨ(a)AΨ(
1
a
)
1
Ψ(ta)Z ta
0
f (s)ψ(s) ds + C0.
Direct calculations show that AΨ(a)AΨ( 1
a ) = 1. Thus,
(MΨ ◦ Pa − Pa ◦ MΨ)f ∈ C0.
Lemma 4.7. If ψ : [0,∞) → (0,∞) is a decreasing function with regular variation of index
−1. Then for any d > 0, as t → ∞,
ψ(t) = o((log(t))−d).
Proof. Since ψ is decreasing and limt→∞
Cǫ > 0 such that for t ∈ (2k, 2k+1],
ψ(2t)
ψ(t) = 2−1, we can for any ǫ ∈ (0, 1) find a constant
ψ(t) ≤ Cǫ(2 − ǫ)−k.
For details, see [32, Proposition 2.21]. Therefore, ψ(t) = o(t−1/2) and the lemma follows.
For the next result we need to assume the (stronger) condition, that the function ψ satisfies
t2ψ(t)
ψ−1(1/t)
lim
t→∞
= c,
(4.3)
for some constant c > 0.
Remark 4.8. For every k ∈ Z, the functions ψ(t) = logk t
tion (4.3). The functions ψ = Ψ′ with Ψ(t) = elogβ t do not satisfy (4.3) for any β > 0.
and ψ(t) = logk(log t)
t·log t
t
satisfy condi-
40
In the following result we use a singular values function of an unbounded operator affiliated
with N. For such operators the formula (2.1) cannot be used. The singular values function of
an operator T affiliated with N is defined [24] as follows:
µT(t, T ) := inf{s ≥ 0 : nT(s, T ) ≤ t}.
Theorem 4.9. Let d > 0 and ψ : [0,∞) → (0,∞) be a decreasing function satisfying condi-
tions (4.2) and (4.3). Assume that (A, H, D, N, T) is a unital semifinite spectral triple such
that
1. T is infinite;
2. D is positive;
3. µT(t, D) ∼ ψ(t)−1/d.
Then (A, H, D, N, T) is Li(0),1-summable with positive T-essential spectrum. Moreover, for
every a ∈ AD and every exponentiation invariant extended limit ω as t → ∞ we have
φω(a) = Tω,ψ(a(1 + D2)−d/2),
where ω := ω◦(JMΨJ) and J : L∞(0,∞) → L∞(0,∞) is defined as the pullback along t 7→ t−1.
Proof. Condition (4.2) on ψ implies that ψ has regular variation of index −1 so ψ(t) =
o((log(t))−d) by Lemma 4.7. Therefore, µT(t, D) = o(log(t)) and the semifinite spectral triple
(A, H, D, N, T) is Li(0),1-summable. Since T is infinite, T(e−tD) ր ∞ as t ց 0. The operator
D therefore has positive T-essential spectrum. Thus,
φω(a) = lim
t→ω
T(ae−tD)
T(e−tD)
is a trace on A by Theorem 3.22. The assumption on D implies that µT(t, (1+ D2)−d/2) ∼ ψ(t),
t → ∞. Applying Lemma 4.1 with T = (1 + D2)−d/2 and q = 1/d yields
Using the properties of extended limits, the definitions of Pa and J and Lemma 4.6, we obtain
Γ(1 + d)
lim
t→ω
T(e−D/t) ∼ Γ(1 + d)ψ−1(t−d), t → ∞.
(J ◦ MΨ) T(ae−D/t)
ψ−1(t−d)!
(J ◦ MΨ ◦ Pd) T(ae−(tD−d)−1/d
(J ◦ Pd ◦ MΨ) T(ae−(tD−d)−1/d
ψ−1(1/t)
ψ−1(1/t)
lim
t→ω
lim
t→ω
Γ(1 + d)
1
1
1
Γ(1 + d)
φω(a) =
=
=
)
)
!
! .
(4.4)
Using the definition of MΨ and assumption (4.3) we obtain
MΨ T(ae−(tD−d)−1/d
ψ−1(1/t)
)
! =
T(ae−(sD−d)−1/d
)
ψ−1(1/s)
ψ(s) ds
T(ae−(sD−d)−1/d
)
ds
s2 + C0(0,∞).
(4.5)
0
1
Ψ(t)Z t
Ψ(t)Z t
1
0
41
∈
Since ω is exponentiation invariant extended limit, combining (4.4) and (4.5) we obtain
φω(a) =
1
Γ(1 + d) ·
1/t→ω(cid:18) 1
Ψ(t)Z t
lim
1
T(ae−(sD−d)−1/d
)
ds
s2(cid:19) .
Now we apply Lemma 4.5 twice with T = D−d, f (x) = e−x−1/d
Since
and then with f (x) = e−x−1
.
we obtain
φω(a) = lim
1/t→ω
0
Z ∞
J(cid:18) 1
Ψ(t)Z t
1
e−x−q dx
x
1
q
= Γ(1 +
),
T(ae−(sD−d)−1
)
ds
s2(cid:19) = Tω,ψ(cid:16)a(1 + D2)−d/2(cid:17) ,
where the last equality follows from [25, Theorem 4.7].
We can now provide sufficient conditions on a general Li(0),1-summable unital semifinite spec-
tral triples ensuring a relation between the trace φω of Theorem 3.22 and Dixmier traces.
Corollary 4.10. Let (A, H, D, N, T) be a unital semifinite spectral triple with T(PD) = ∞.
Assume that there is a number d > 0 and a decreasing function ψ : [0,∞) → (0,∞) with
regular variation of index −1 satisfying conditions (4.2) and (4.3) and
µ(t, PDD) ∼ ψ(t)−1/d.
Then, βD = 0 and for any exponentiation invariant extended limit ω as t → ∞ and a ∈ AD,
φω(a) = Tω,ψ(PDa(1 + D2)−d/2),
where ω is as in Theorem 4.9.
The corollary follows immediately from Theorem 4.9 applied to the unital semifinite spectral
triple (T, PDH, PDD, N+, T) where T is the ∗-algebra generated by PDAPD.
Example 4.11. Let us revisit the spectral triple constructed in Proposition 2.16 (see page
15). If ψ : [0,∞) → (0,∞) is a smoothly varying function with limt→0 ψ(t) = 0 and ψ(t)−1 =
O(t1/n) as t → ∞, and /D a Dirac operator on a Riemannian closed n-dimensional manifold
M , a ψ-summable spectral triple (C ∞(M ), L2(M, S), /Dψ) was constructed in Proposition 2.16,
where /Dψ := F /Dψ( /Dn)−1. The Weyl law for /D and Theorem 2.10 applied to B = P /D shows
that the order of the spectral asymptotics of /D coincides with the order of the spectral
asymptotics of P /D /D so µ(t, P /D /Dψ) ∼ cψ(t)−1 for some constant c > 0.
If ψ has regular variation of index −d for a d > 0, Corollary 4.10 shows that the tracial state
on C(M ) defined from the spectral triple (C ∞(M ), L2(M, S), /Dψ) takes the form
φω(a) = c′ Trω,ψ(P /Daψ( /Dn)1/d),
for some proportionality constant c′. Applying Connes' trace theorem as in [32, Theorem 9.1],
it follows that
φω(a) =ZM− adV,
42
where dV denotes the Riemannian volume measure on M andZM− the normalized integral.
The computation above requires ψ to have strictly negative index of regular variation. We note
that by Proposition 2.18 (see page 16), the computation above can only extend to the spectral
triple from Proposition 2.19 (see page 17) on a crossed product by a local diffeomorphism if
the local diffeomorphism acts isometrically.
5 The KMS-state φω in examples
We are now in a state where we can compute the KMS-states associated with the spectral triples
considered in Subsection 2.2 (see page 13). The computations for the KMS-states associated
with the unbounded Kasparov cycle on Cuntz-Pimsner algebras considered in Subsection 2.2.3
(see page 19) are more involved and dedicated a separate section, Section 6 (see page 50).
5.1 Dirac operators on closed manifolds
For a closed manifold M with a Dirac operator /D acting on a Clifford bundle S → M ,
we consider the spectral triple (C ∞(M ), L2(M, S), /D) as in Proposition 2.14 (see page 14).
The following theorem can be deduced immediately from Example 3.5 (see page 28) or from
Corollary 4.10 and Connes' trace theorem for pseudo-differential operators.
Theorem 5.1. Let (C ∞(M ), L2(M, S), /D) be the spectral triple associated with a Dirac op-
erator on a closed manifold, ω an extended limit as t → 0 and φω the associated tracial state
from Theorem 3.22 (see page 34). The trace φω is the normalized volume integral on M , i.e.
for a ∈ C(M ),
φω(a) =ZM− a dV.
For completeness, let us describe the Toeplitz algebras TC(M ), TC(M ), /D and the flow σ on
C(M ) /D in this example. We remark that since φω is a trace in this case, the flow induced
from our construction is irrelevant for the study of φω but it could nevertheless serve to clarify
the constructions of Subsection 3.2. The relevant algebras are all contained in the C ∗-algebra
C∗(M, S) defined as the C ∗-closure of the ∗-algebra Ψ0
Ψ0
cl(M, S) of zeroth order classical pseudo-
differential operators acting on L2(M, S). It is well known that Ψ0
C∗(M, S) fits into a short
exact sequence
0 → K(L2(M, S)) → Ψ0
C∗(M, S)
symb
−−−→ C(S∗M, End(S)) → 0,
where symb denotes the continuous extension of the principal symbol mapping Ψ0
C ∞(S∗M, End(S)). Since P /D is a projection in Ψ0
Ψ0
cl(M, S) →
cl(M, S), we can consider the C ∗-algebra
C∗,+(M, S) := P /DΨ0
C∗(M, S)P /D and we obtain a short exact sequence
0 → K(P /DL2(M, S)) → Ψ0
C∗,+(M, S)
symb
−−−→ C(S∗M, End(p /DS)) → 0,
where p /D ∈ C ∞(S∗M, End(S)) denotes the principal symbol of P /D. The algebras TC(M ) and
43
TC(M ), /D are characterized by the following commuting diagram with exact rows
0
0
0
/ K(P /DL2(M, S))
TC(M )
=
/ K(P /DL2(M, S))
TC(M ), /D
=
symb
symb
C(M )
C(M ) /D
/ 0
/ 0
/ K(P /DL2(M, S))
/ Ψ0
C∗,+(M, S)
symb /
/ C(S∗M, End(p /DS))
/ 0,
The composition of the mappings in the right most column coincides with the pull back ho-
momorphism C(M ) → C(S∗M ) composed with the inclusion C(S∗M ) ⊆ C(S∗M, End(p /DS)).
To describe the flow σ on C(M ) /D, we describe it on C ∞(S∗M, End(p /DS)). Surjectivity
of the principal symbol mapping implies that any a ∈ C ∞(S∗M, End(p /DS)) is the symbol
of an operator A ∈ P /DΨ0
cl(M, S)P /D. By Egorov's theorem [22] (see also [35, Section IV]
and [21]), eis /DAe−is /D is again an element of P /DΨ0
cl(M, S)P /D and the expression σs(a) :=
symb(eis /DAe−is /D) gives a well defined flow on C ∞(S∗M, End(p /DS)). Again by Egorov's
theorem, using that σs(a) = symb(eis /DAe−is /D) for A ∈ P /DΨ0
cl(M, S)P /D, we have that
σs(a) = g∗
s (a) where gs : S∗M → S∗M is the Hamiltonian flow associated with the symbol
ξ of /D. On the cosphere bundle, this Hamiltonian flow coincides with the geodesic flow.
We conclude that the flow σ is induced by geodesic flow and C(M ) /D ⊆ C(S∗M ) is a closed
subalgebra invariant under geodesic flow. This discussion should be compared with that in
[16].
As in Subsection 2.2.1, we consider a local diffeomorphism g : M → M . Assuming that g acts
conformally and lifts to S → M , it is readily verified that g is compatible with the decreasing
. We use the notation /Dlog := /Dψ for this particular choice of ψ.
function ψ(t) :=
Note that
log(1+t2/n)
1
/Dlog = F /D log(1 + /D2) and e−t /Dlog = (1 + /D2)−t.
By Proposition 2.19, we arrive at a spectral triple (A, L2(M, S), /Dlog) where A is the ∗-algebra
generated by C ∞(M ) and an isometry Vg. Let us compute KMS-state constructed from
(A, L2(M, S), /Dlog). Before diverting into this computation, we recall that the C ∗-closure
of A coincides with the image of a representation of OEg . As such, we can write elements of
A as linear span of elements of the form SµS∗
ν where µ = a1Vg ··· akVg and ν = b1Vg ··· blVg,
where a1, . . . , ak, b1, . . . , bl ∈ C ∞(M ).
Proposition 5.2. Set S := C ∞(M ) ∪ C ∞(M )Vg ⊆ A. For any β ∈ R, the set S is an
analytically generating set at β for (A, L2(M, S), /Dlog).
Proof. For notational convenience, write D = /Dlog. The set S generates A, so PDSPD + K
generates TA. For any β ∈ R, and a ∈ S
e−βDPDaPDeβD = (1 + /D2)−βP /DaP /D(1 + /D2)β.
The proposition follows from that Dom((1 + /D2)β) = H 2β(M, S) as Banach spaces and any
a ∈ S extends to a continuous operator on the Sobolev spaces H 2β(M, S) for all β.
44
/
/
/
/
/
_
_
/
/
/
/
/
/
_
_
/
/
/
/
Theorem 5.3. Let M be a connected n-dimensional Riemannian manifold, /D a Dirac operator
on S → M , and g : M → M a local diffeomorphism acting conformally and lifting to S. Then
the spectral triple (A, L2(M, S), /Dlog) is Li1-summable, has positive essential spectrum with
βD = n/2 and is n/2-analytic.
Assume for all m ∈ N+, that the set of fixed points
{x ∈ M : gm(x) = x},
has measure zero with respect to the volume measure. Then the state φω on A constructed from
Corollary 3.21 (see page 34) is independent of ω and takes the form
φω(SµS∗
ν ) = δµ,νZM− Lg(cn/2
g
kLg(cn/2
b∗
g
k−1Lg(··· Lg(cn/2
b∗
g
b∗
1a1)a2)··· ak−1)ak) dV,
(5.1)
For µ = a1Vg ··· akVg and ν = b1Vg ··· blVg. Here dV denotes the Riemannian volume form.
The state φω viewed as a state on OEg via its representation on L2(M, S) is KMS for the
action defined by γt(aVg) := citn/2
aVg with inverse temperature 1.
g
We note that the state Theorem 5.3 is a KMS-state on a Cuntz-Pimsner algebra, but not for
its gauge action. If cg(x) < 1 for some x ∈ M , the generator of γ is not positive on Eg and the
Laca-Neshveyev correspondence [45] does not apply in the context of OEg with the action γ.
In the case k = l = 0, the formula (5.1) should be interpreted as φω(a) = ZM− a dV for
a ∈ C ∞(M ). This special case follows from Connes' trace theorem.
Proof. For t > n/2, standard techniques of pseudo-differential operators show that the integral
kernel Kt of the trace class operator P /D(1 + /D2)−t belongs to C(M × M, END(S))∩ C ∞(M ×
M \ ∆M , END(S)). Here ∆M ⊆ M × M denotes the diagonal and END(S) denotes the big
endomorphism bundle defined by END(S)(x,y) := Hom(Sx, Sy) for (x, y) ∈ M × M . By [19,
Proposition 8.3], V ∗
g takes the form
where LS,g is the L2-extension of the operator
V ∗
g = N 1/2LS,gc−n/4
g
,
LS,g : C(M, S) → C(M, S), LS,gξ(x) := Xg(y)=x
ug(y)−1ξ(y).
Take a1, . . . , ak, b1, . . . , bl ∈ C ∞(M ) and write µ = a1Vg ··· akVg and ν = b1Vg ··· blVg. We
introduce the notation aj := ajcn/4
. We can compute for t > n/2 that
and bj = bjc−n/4
g
g
TrL2(M,S)(P /DSµS∗
ν e−t /Dlog) = TrL2(M,S)(a1Vg ··· akVgV ∗
g b∗
l V ∗
g b∗
g ··· V ∗
1P /D(1 + /D
2
)−t)
= N −(k−l)/2 TrL2(M,S)(a1cn/4
g ugg∗LS,gc−n/4
g
b∗
l ··· c−n/4
g
LS,gb∗
1P /D(1 + /D
2
)−t)
aj(xj )
k+lYj=k+1
bj−k(xj )∗ Kt(xk+1, x1)dV (x),
= N −(k−l)/2ZM
g ugg∗ ··· akcn/4
kYj=1
X
(x1,...,xk+l)∈M(x,k,l)
45
where M (x, k, l) ⊆ M k+l is defined as the k + l-tuples (x1, . . . , xk+l) such that for j = 1, . . . , k,
xj = gj−1(x), xk = g(xk+l) and for j = k + 1, . . . k + l − 1, g(xj ) = xj+1. Note that M (x, k, l)
is finite for all x, because g is a local homeomorphism, and that x1 = x for (x1, . . . , xk+l) ∈
M (x, k, l).
Define M 0(x, k, l) ⊆ M (x, k, l) as the k + l-tuples (x1, . . . , xk+l) where xk+1 = x1. The set
M 0(x, k, l) can be characterized as the k + l-tuples (x1, . . . , xk+l) with x = x1 = xk+1 and
xk = gl(xk+l) such that for j = 1, . . . , k, xj = gj−1(x), and for j = k + 1, . . . k + l − 1,
xj+1 = g(xj). In particular, if M 0(x, k, l) is non-empty then gk(x) = gl(x). In other words,
(x1, . . . , xk+l) ∈ M 0(x, k, l) if and only if gk(x) = gl(x), and xj = gj−1(x) for j = 1, . . . , k
and xk+j = gj−1(x) for j = 1, . . . , l. Therefore, M 0(x, k, l) contains at most one element.
In particular, since M 0(x, k, l) is non-empty then gk(x) = gl(x) and the set of fixed points
{x ∈ M : gm(x) = x} has measure zero for all m ∈ N+ by assumptions, then
M 0(x, k, l) = ∅
if k 6= l a.e. in x.
As t approaches n/2, the integral kernel Kt localizes (up to lower order term) to the diagonal
and the leading order terms come from the sum over M 0(x, k, l). The Weyl law for /D2 and an
explicit pseudo-differential computation of the principal symbol of P /D(1 + /D2)−t implies that
there is a constant c and an ǫ > 0 only depending on /D such that
TrL2(M,S)(P /DSµS∗
ν e−t /Dlog) =
=cδk,l(t − n/2)−1ZM
X
(x1,...,xk+l)∈M 0(x,k,l)
kYj=1
aj(xj )
k+lYj=k+1
bj−k(xj)∗ dV (x) + fµ,ν(t),
where fµ,ν is holomorphic on a neighborhood of the intervall [n/2 − ǫ, n/2].
Recall the notation aj := ajcn/4
we write
and bj := bjc−n/4
. For k = l and (x1, . . . , xk+l) ∈ M 0(x, k, k)
g
g
aj(gj−1(x))b∗
j (gj−1(x)) =
2kYj=k+1
aj(xj)(cid:17)
(cid:16) kYj=1
kYj=1
=
kYj=1
bj−k(xj)∗ =
j (gj−1(x)) =(cid:16)[a1b∗
aj(gj−1(x))b∗
1][g∗(a2b∗
2)][(g2)∗(a2b∗
2)]··· [(gk)∗(akb∗
k)](cid:17) (x).
, we can partially integrate RM ag∗(b)dV =
By the same argument that V ∗
g = N −1/2LS,gc−n/4
g
RM Lg(cn/2
g a)bdV for a, b ∈ C(M ). By partially integrating k − 1 times we deduce that for
some function fµ,ν holomorphic on a neighborhood of the intervall [n/2 − ǫ, n/2].
TrL2(M,S)(P /DSµS∗
ν e−t /Dlog) =
=cδk,l(t − n/2)−1ZM
= cδk,l(t − n/2)−1ZM
[a1b∗
1][g∗(a2b∗
2)][(g2)∗(a2b∗
Lg(cn/2
g
b∗
kLg(cn/2
g
k)]dV + fµ,ν(t) =
2)]··· [(gk)∗(akb∗
b∗
1a1)a2)··· ak−1)ak)dV + fµ,ν(t)
b∗
k−1Lg(··· Lg(cn/2
g
We conclude that formula (5.1) holds.
It remains to show that φω defines a KMS-state on OEg for the action defined by γt(aVg) :=
citn/2
aVg. Let τ denote the tracial state on C(M ) defined from integrating against the volume
g
46
form and L ∈ End∗
products shows that for µ = a1Vg ⊗ ··· ⊗ akVg, ν = b1Vg ⊗ ··· ⊗ bkVg ∈ E
computation
C(M )(Eg) the generator of γt, i.e. L = n
2 log(cg). Some yoga with inner
, we have the
⊗C(M )k
g
φω(SµS∗
ν ) =ZM
Lg(cn/2
g
kLg(cn/2
b∗
g
k−1Lg(··· Lg(cn/2
b∗
g
b∗
1a1)a2)··· ak−1)ak)dV =
= τ(cid:0)b1Vg ⊗ ··· ⊗ bkVge−L(a1Vg) ⊗ ··· ⊗ e−L(akVg)(cid:1)E
= φω(S∗
ν γi(Sµ)).
⊗C(M )k
g
We conclude that φω(ab) = φω(bγi(a)) for a, b ∈ A and φω is a KMS-state on OEg in the action
γ.
5.2 Graph C ∗-algebras
For a finite graph G we consider the spectral triple on C ∗(G) constructed in Proposition 2.20
from the choice of a point y ∈ ΩG in the infinite path space. We will assume that G is primitive,
in which case C ∗(G) is simple. For an element (x, n) ∈ Vy and a finite path µ, we compute
that
eisDy PDy SµPDy e−isDy δ(x,n) = eisµPDy SµPDy δ(x,n).
ν ) = eis(µ−ν)SµS∗
It follows that σt(SµS∗
ν and σ coincides with the gauge action on the graph
C ∗-algebra C ∗(G). We can conclude that C ∗(G) = C ∗(G)Dy is closed under the flow σ. The
reader can recall from Example 3.6 (see page 28) that PDy ℓ2(Vy) = ℓ2(V+
y is defined
y )) is the C ∗-algebra generated by the creation
in Equation (3.2). Moreover, TC∗(G) ⊆ B(ℓ2(V+
operators
Teδ(x,n) :=(δ(ex,n),
y ) where V+
0,
r(e) = s(x),
r(e) 6= s(x).
Proposition 5.4. Let β ∈ R. The set S = {Se : e ∈ E} ⊆ Cc(GG) is an analytically generating
set at β for (Cc(GG), ℓ2(Vy), Dy).
Proof. Since PDy SePDy = Te, it is clear that S satisfies that PDy SPDy + K(ℓ2(V+
y ) generates
the Toeplitz algebra TC∗(G) as a C ∗-algebra. Moreover, eβDy PDSePDe−βDy = eβPDSePD is
bounded and the proposition follows.
We conclude the following theorem from Example 3.6.
Theorem 5.5. Let G be a finite primitive graph with edge adjacency matrix A and y ∈ ΩG.
For any extended limit ω as t → log rσ(A), the KMS-state φω on C ∗(G) associated with the
spectral triple (Cc(GG), ℓ2(Vy), Dy) (see Proposition 2.20 on page 18) as in Corollary 3.21 (see
page 34) is given by
φω(SµS∗
ν ) = δµ,ν
rσ(A)−µ,
ws(µ)
kwkℓ1
where ν and µ are finite paths and w ∈ CE is the ℓ2-normalized Perron-Frobenius vector. The
state φω is KMS for the gauge action and its inverse temperature is log(rσ(A)).
The KMS-state φω on C ∗(G) in Theorem 5.5 is the unique KMS-state by [23]. Numerous
authors present constructions of this state and for more general graphs, eg [1, 2, 13, 42].
47
5.3 Group C ∗-algebras
For the reduced group C ∗-algebra of a countable discrete group we considered two types of
(semifinite) spectral triples in Subsection 2.2.4 (see page 24). We now compute the associated
KMS-states.
We fix a length function ℓ on the discrete countable group Γ. For technical simplicity, we
assume that Γ is an exact group ensuring that Γ acts amenably on its Stone-Cech boundary
∂SCΓ (see [53]). We assume that (Γ, ℓ) is of at most exponential growth and that ℓ is critical (see
Definition 3.10 on page 30). Let (cb(Γ) ⋊alg Γ, ℓ2(Γ), Dℓ) denote the associated Li1-summable
spectral triple as in Proposition 2.31. Since Dℓ ≥ 0, we have that
Tcb(Γ)⋊rΓ = cb(Γ) ⋊r Γ + K(ℓ2(Γ)) = cb(Γ) ⋊r Γ.
The last equality follows from that c0(Γ) ⋊r Γ = K(ℓ2(Γ)). We conclude that we have a short
exact sequence
0 → K(ℓ2(Γ)) → Tcb(Γ)⋊r Γ → C(∂SCΓ) ⋊r Γ → 0.
The flow σ+ on Tcb(Γ)⋊rΓ = cb(Γ) ⋊r Γ is given on an element aλg ∈ cb(Γ) ⋊ Γ by
s (aλg) = eis(ℓ(·)−ℓ(g−1·)aλg,
σ+
(5.2)
and we conclude that Tcb(Γ)⋊rΓ is invariant under σ+. Therefore Tcb(Γ)⋊rΓ = Tcb(Γ)⋊rΓ,Dℓ.
Proposition 5.6. Let β ∈ R. The ∗-algebra cb(Γ) ⋊alg Γ is an analytically generating set at β
for (cb(Γ) ⋊alg Γ, ℓ2(Γ), Dℓ).
Proof. Note that PDℓ = 1 because Dℓ is positive. For aλg ∈ cb(Γ) ⋊alg Γ, we compute that
eβDℓ aλge−βDℓ = e−β(ℓ(·)−ℓ(g−1·))aλg.
Since ke−β(ℓ(·)−ℓ(g−1·)akcb(Γ) ≤ eβℓ(g−1)kakcb(Γ), it holds that eβDℓ aλge−βDℓ ∈ cb(Γ) ⋊alg Γ and
the proposition follows.
Proposition 5.7. Let β ∈ R. The ∗-algebra cb(Γ) ⋊alg Γ is an analytically generating set at β
for (cb(Γ) ⋊alg Γ, ℓ2(Γ, SH), Dc, N, Trτ ).
Proof. For aλg ∈ cb(Γ) ⋊alg Γ, we compute for f ∈ ℓ2(Γ, SH) that
eβDc πS(aλg)e−βDcf (γ) = e−β(ℓ(γ)−ℓ(g−1γ))a(γ)[π(g)f ](γ).
The estimate ke−β(ℓ(·)−ℓ(g−1·)akcb(Γ) ≤ eβℓ(g−1)kakcb(Γ) shows that eβDc πS(aλg)e−βDc ∈
π(cid:0)cb(Γ) ⋊alg Γ(cid:1). Therefore,
eβDc PDc(cb(Γ) ⋊alg Γ)PDc e−βDc
= PDc eβDc(cb(Γ) ⋊alg Γ)e−βDcPDc ⊆ PDc(cb(Γ) ⋊alg Γ)PDc ⊆ N+,
and the proposition follows.
48
Definition 5.8. If ω is an extended limit as t → β(Γ, ℓ), and ℓ is critical, we define the
Patterson-Sullivan measure µω on the Stone- Cech boundary ∂SCΓ as
Z∂SC Γ
a dµω := lim
t→ωPγ∈Γ a(γ)e−tℓ(γ)
Pγ∈Γ e−tℓ(γ)
,
for a function a ∈ C(∂SCΓ) and where a ∈ cb(Γ) is any function with a = a mod c0(Γ).
Remark 5.9. It is possible to define the Patterson-Sullivan measure µω as an extended weak*-
limit of the family of probability measures on Γ
µt = Pγ∈Γ δγe−tℓ(γ)
Pγ∈Γ e−tℓ(γ)
.
In the literature, Patterson-Sullivan measures are usually defined as weak* accumulation points
of (µt)t>β(Γ,ℓ) but we allow for a slightly more general construction with extended limits. A
priori, µω is a probability measure on the Stone- Cech compactification of Γ. Since the support
of µω is contained in the closed subspace ∂SCΓ we consider µω as a measure on ∂SCΓ.
Theorem 5.10. Let Γ be a discrete group and φω the KMS-state on C(∂SCΓ) ⋊r Γ constructed
as in Corollary 3.21 (see page 34) using an extended limit ω as t → β(Γ, ℓ) and either of the
following two semifinite spectral triples:
• The spectral triple
(cb(Γ) ⋊alg Γ, ℓ2(Γ), Dℓ)
associated with a critical length function of at most exponential growth as in Proposition
2.31 (see page 24).
• The semifinite spectral triple (cb(Γ) ⋊alg Γ, ℓ2(Γ, SH), Dc, N, Trτ ) associated with a critical
proper Hilbert space valued cocycle of at most exponential growth as in Proposition 2.34
(see page 26).
Then φω is given in terms of the Patterson-Sullivan measure µω by
φω(aλg) = δe,gZ∂SC Γ
a dµω.
The state φω is KMS at inverse temperature β(Γ, ℓ) for the flow on C(∂SCΓ) ⋊r Γ induced by
the action σ+ on cb(Γ) ⋊r Γ given in Equation (5.2). Moreover, φω extends to a KMS-state at
inverse temperature 1 on the von Neumann algebra L∞(∂SCΓ, µω)⋊Γ with its Radon-Nikodym
flow
σRN
s
(aλg) =(cid:18) dg∗µω
dµω (cid:19)is
aλg.
Proof. By the computations of Example 3.9 (see page 30), the spectral triple associated with
a length function as in Proposition 2.31 have the same heat traces as the semifinite spectral
triple associated with a proper Hilbert space valued cocycle as in Proposition 2.34. In both
cases, Example 3.9 shows that for aλg ∈ cb(Γ) ⋊alg Γ we have
φt,0(aλg) = δe,gPγ∈Γ a(γ)e−tℓ(γ)
Pγ∈Γ e−tℓ(γ)
.
49
It follows that φω(aλg) = δe,gR∂SC Γ a dµω in both cases.
To relate φω to the Radon-Nikodym flow, we first show that µω is strictly positive, i.e. that
µω(U ) > 0 for any open set U ⊂ ∂SCΓ. For any open set U ⊂ ∂SCΓ, the translates (γU )γ∈Γ
cover ∂SC Γ. If µω(U ) = 0, then by quasi-invariance µω(γU ) = 0 which contradicts µω being a
probability measure.
The fact that µω is strictly positive ensures that the Radon-Nikodym derivatives dg∗µω
dµω
defined and strictly positive. The mapping g 7→ dg∗µω
is a cocycle, i.e. for h, g ∈ Γ,
are well
dµω
d(gh)∗µω
dµω
=
dg∗µω
dµω
dµω (cid:21) .
(g−1)∗(cid:20) dh∗µω
The proof is completed by computing that for aλg, bλh ∈ C(∂SCΓ) ⋊alg Γ, we have the identity
φω(aλgbλh) = δh,g−1φω(ah∗(b)) = δh,g−1Z∂SC Γ
ah∗(b)dµω =
g∗(a)bd(h∗µω) =
dµω =
bg∗(a)
dh∗µω
dµω
h∗ (g∗(a)b) dµω = δh,g−1Z∂SC Γ
= δh,g−1Z∂SC Γ
= δh,g−1Z∂SC Γ
b(h−1)∗ a(cid:18) dg∗µω
dµω (cid:19)−1! dµω =
= δh,g−1Z∂SC Γ
aλg! = φω(cid:0)bλhσRN
= φω bλh(cid:18) dg∗µω
dµω (cid:19)−1
s=i (aλg)(cid:1)
(h−1)∗(cid:18) dg∗µω
dµω (cid:19) dh∗µω
= 1.
dµω
In the third last identity we used the cocycle identity implying that if hg = e, then
Remark 5.11. The reader should note that φωC∗
r (Γ) coincides with the ℓ2-trace.
6 KMS-states on Cuntz-Pimsner algebras with their gauge action
In this section we consider the constructions of Corollary 3.21 in a broad class of examples which
include both Cuntz-Krieger algebras and crossed products by Z. Here we use the techniques
from Section 3 in conjunction with those from Subsection 2.2.3 to analyse the KMS states on
Cuntz-Pimsner algebras, and compare them to the Laca-Neshveyev correspondence establishing
a bijection between KMS-states on Cuntz-Pimsner algebras and tracial states on its coefficient
algebra.
6.1 KMS-states on Cuntz-Pimsner algebras from traces on the coefficient algebra
Firstly, we shall show that a critical positive trace on A (see Definition 6.2 below) gives rise to
a KMS-state on the Cuntz-Pimsner algebra OE assuming that E is strictly W-regular. Recall
Lemma 2.28 (see page 22) giving the construction of the Li1-summable semifinite spectral triple
(OE, L2(ΞA, τ ), Dψ, Nτ (ΞA), Trτ ), where Nτ (ΞA) := (End∗
A(ΞA) ⊗A 1)′′.
50
Lemma 6.1. Let E be a strictly W-regular fgp bi-Hilbertian bimodule over A, β ∈ R and τ a
positive trace on A. The set S = {Se : e ∈ E} ⊂ OE is an analytically generating set at β for
(OE, L2(ΞA, τ ), Dψ, Nτ (ΞA), Trτ ).
Proof. Since TOE,D is precisely the Toeplitz algebra TE, it is immediate that the set of operators
{P SeP : e ∈ E} generates TOE ,D. The analyticity condition on the Fock space is likewise
obvious from the computation
eβDψ PDψ SePDψ e−βDψ = eβPDψ SePDψ .
Definition 6.2. Let E be an A-bimodule which is fgp from the right and τ a positive trace
on A. We define the critical value of (E, τ ) as
β(E, τ ) := inf{t ≥ 0 :
∞Xn=0
τ∗(E⊗An)e−tn < ∞}.
We say that τ is critical for E if
lim
tցβ(E,τ )
∞Xn=0
τ∗(E⊗An)e−tn = ∞.
Remark 6.3. Note that the critical value of a trace and the notion of it being critical only
depends on [E] ∈ KK0(A, A), in fact only on the sequence (ch0(E⊗An))n∈N ⊆ HC0(A) in cyclic
homology. Just as in the proof of Lemma 2.28 we obtain the estimate 0 ≤ β(E, τ ) ≤ log(N ),
where N is the number of elements in the left frame and the right frame.
It follows from Definition 3.1 and Proposition 3.8 that the Li1-summable semifinite spectral
triple (OE , L2(ΞA, τ ), Dψ, Nτ (ΞA), Trτ ) of Lemma 2.28 has positive Trτ -essential spectrum if
and only if τ is critical.
By construction, the projection PDψ is the projection onto the Fock module FE and therefore
Nτ (ΞA)+ = (End∗
A(FE)⊗A1)′′ is a subalgebra of B(L2(FE, τ )). We let N denote the number op-
erator on L2(FE, τ ) -- the self-adjoint operator defined from NE⊗n⊗AL2(A,τ ) = nIdE⊗n⊗AL2(A,τ ).
The next proposition follows from the definition of the Toeplitz algebra of a semifinite spectral
triple.
Proposition 6.4. The Toeplitz algebra TOE of the semifinite spectral triple
(OE, L2(ΞA, τ ), Dψ, Nτ (ΞA), Trτ )
is given by
TOE = TE ⊗A 1A + K(End∗
A(FE )⊗A1)′′ ,
where TE ⊆ End∗
A(FE) is the Cuntz-Toeplitz algebra of E. Moreover, the action σ+ preserves
TOE and is generated by the number operator in the sense that for µ ∈ Ek, ν ∈ E⊗l and
K ∈ K
A(FE )⊗A1)′′, we have
(End∗
σ+
s (TµT ∗
ν + K) = eis(µ−ν)TµT ∗
ν + eisN Ke−isN .
An immediate consequence of Proposition 6.4 is that OE = TOE /K
action σ on OE coincides with the gauge action σs(SµS∗
theorem is readily deduced from the computations of Example 3.7 (see page 29).
ν ) = eis(µ−ν)SµS∗
A(FE ))′′ and that the
ν . The following
(End∗
51
Theorem 6.5. Let E be a strictly W-regular fgp bi-Hilbertian bimodule over A. For any
positive trace τ on A which is critical for E and any extended limit ω as t → β(E, τ ), the KMS-
state φω on OE associated with the semifinite spectral triple (OE , L2(ΞA, τ ), Dψ, Nτ (ΞA), Trτ )
as in Theorem 3.21 (see page 34) is given by
φτ,ω(SµS∗
ν ) = δµ,νe−β(E,τ )µ lim
n=0 TrE⊗n
τ
t→ωP∞
P∞
((νµ)Eµ) e−tn
,
n=0 TrE⊗n
τ
(1)e−tn
where ν ∈ E⊗k and µ ∈ E⊗l. The state φτ,ω is KMS for the gauge action on OE and its
inverse temperature is β(E, τ ).
Proof. By Lemma 6.1 the semifinite spectral triple (OE, L2(ΞA, τ ), Dψ, Nτ (ΞA), Trτ ) is β-
analytic for any β ∈ R. By Definition 3.1 and Proposition 3.8 it has positive Trτ -essential
spectrum. Thus the state φτ,ω as in Theorem 3.21 is KMS for the gauge action on OE with
inverse temperature β(E, τ ). Using the computation in Equation (3.3), we see that
φτ,ω(SµS∗
ν ) = δµ,ν lim
t→ωP∞
n=µPσ=n e−tnτ(cid:0)( (µeσ)Eµ eσ (νeσ)Eµ eσ)E⊗(n−µ)(cid:1)
P∞
n=0Pρ=n τ ((eρeρ)A)e−tn
,
for ν ∈ E⊗k and µ ∈ E⊗l. However, this expression can be vastly simplified using that φτ,ω is
KMS. Using
φτ,ω(SµS∗
ν ) = δµ,νe−β(E,τ )µφτ,ω(S∗
ν Sµ) = δµ,νe−β(E,τ )µφτ,ω((νµ)E⊗µ ),
we obtain
φτ,ω(SµS∗
= δµ,νe−β(E,τ )µ lim
ν ) = δµ,νe−β(E,τ )µφω((νµ)E⊗µ)
n=0Pσ=n e−tnτ(cid:0)( eσ (νµ)Eµ eσ)E⊗n(cid:1)
t→ωP∞
P∞
n=0Pρ=n τ ((eρeρ)A)e−tn
t→ωP∞
((νµ)Eµ) e−tn
P∞
= δµ,νe−β(E,τ )µ lim
n=0 TrE⊗n
n=0 TrE⊗n
(1)e−tn
.
τ
τ
The reader should note that in the formula computing φτ,ω, it is only the right inner product
on E that appears.
Example 6.6. Let us consider the construction from Theorem 6.5 in a specific example. As
in Example 2.30 (see page 23), we consider a compact Hausdorff space Y , a surjective local
homeomorphism g : Y → Y and the associated bimodule Eg. Let us compute φτ,ω starting
from a positive trace τ on C(Y ), i.e. a positive measure λ on Y . By the argument of Theorem
6.5, the KMS-condition on φτ,ω guarantees that it suffices to describe φτ,ω(a) for a ∈ C(Y ).
By the Riesz representation theorem,
for a probability measure λω. We compute that for a ∈ C(Y ),
TrE⊗n
τ
(a) = Xσ=n
Ln
g (a) dλ =ZY
a d[(Ln
g )∗λ].
a dλω,
φτ,ω(a) =ZY
τ ((eσaeσ)C(Y )) =ZY
52
λω = lim
t→ω P∞
P∞
n=0 e−tn(Ln
g )∗λ
.
n=0 e−tn[(Ln
g )∗λ](Y )
From Theorem 6.5, we conclude that λω is given by an extended weak* limit of measures
The KMS-condition on φτ,ω translates into (Lg)∗λω = eβ(E,τ )λω which is readily verified for
the measure λω.
We remark that the construction above is reminiscent of the method in [64] to construct
equilibrium measures. In the case that g is mixing, i.e. for all open subsets U, V ⊆ Y there is
an N ≥ 0 such that gn(U ) ∩ V 6= ∅ for all n ≥ N , then there exists a unique KMS-state on
OEg (see [19, Theorem 6.1]). In particular, for mixing g, the KMS-state φτ,ω on OEg does not
depend on the choice of trace τ .
6.2 The Toeplitz construction vs the Laca-Neshveyev correspondence
In the previous subsection we saw that there is a mapping from the set of positive critical
traces on a C ∗-algebra A to the set of KMS-states on the Cuntz-Pimsner algebra OE when E
is strictly W-regular. As Example 6.6 shows, this mapping is not injective in general, but it
is surjective in some cases (e.g. when g is mixing). We now compare our construction to the
bijective correspondence between a certain set of tracial states on A with KMS-states on the
Cuntz-Pimsner algebra OE first discovered by Laca-Neshveyev [45].
Definition 6.7. The positive trace τ : A → C satisfies the Laca-Neshveyev condition for α ≥ 0
if
where La denotes the left action of a on E.
TrE
τ (La) = eατ (a),
For notational simplifity we often write TrE
τ (La). Given a positive trace
τ : A → C satisfying the Laca-Neshveyev condition, it was proven by Laca-Neshveyev [45,
Theorem 2.1 and 2.5] that the expression
τ (a) instead of TrE
φLN,τ (SµS∗
ν ) := δµ,νe−αµτ ((νµ)A),
defines an α-KMS state on OE. Moreover, Laca-Neshveyev proved that the construction τ 7→
φLN,τ is a bijection between tracial states on A satisfying the Laca-Neshveyev condition for
α ≥ 0 and α-KMS states on OE.
The work of Laca-Neshveyev [45] gives more context to the construction in Theorem 6.5 (see
page 52). For a unital C ∗-algebra A, we let T(A) denote the set of positive traces on A. If E is
an A − A-correspondence which is finitely generated and projective as a right module, we can
also define CTE,α(A) as the set of positive critical traces τ with β(E, τ ) = α. We also define
LNE,α(A) as the set of positive traces satisfying the Laca-Neshveyev condition.
Following [45, Discussion proceeding Definition 2.3], we define
FE,α : T(A) → T(A), FE,ατ (a) = TrE
τ (a)e−α.
(6.1)
Proposition 6.8. For any positive trace τ on a unital C ∗-algebra and an A-A-correspondence
E which is fgp from the right, it holds that
E,ατ (a) = e−αn TrE⊗n
F n
τ
(a), n ∈ N+.
53
computation shows that
Proof. By definition, FE,ατ (a) = e−αPN
E,ατ (a) = e−αn Xσ=n
F n
τ (eσaeσ)A = e−αn TrE⊗n
τ
(a).
j=1 τ (ejaej )A for a right frame (ej)N
j=1 of E. A direct
Proposition 6.9. Let E be an A-A-correspondence which is fgp from the right. Then τ ∈ T(A)
is a fixed point of FE,α if and only if τ satisfies the Laca-Neshveyev condition from Definition
6.7 (see page 53).
Proposition 6.9 is a direct consequence of Definition 6.7 and the formula (6.1).
Proposition 6.10. Let E be an A-A-correspondence which is fgp from the right. If τ ∈ T(A)
is a tracial state satisfying the Laca-Neshveyev condition for α ≥ 0, then
τ∗(E⊗An) = eαnF n
E,ατ (1) = eαn.
Proof. The proposition follows from the computation
τ∗(E⊗An) = TrE⊗n
τ
(1) = eαnF n
E,α(τ )(1),
and Proposition 6.9.
Proposition 6.11. Let E be an A-A-correspondence which is fgp from the right and α ≥ 0.
Then the following holds:
1. A positive trace τ ∈ T(A) is critical for α if and only if the positive trace
St
E,ατ :=
∞Xn=0
e−tnF n
E,ατ,
is finite for t > 0 and satisfies St
inclusion of sets
E,ατ (1) → ∞ as t → 0.
In particular, we have an
2. For any extended limit ω as t → 0, the mapping
LNE,α(A) ⊆ CTE,α(A).
Sω
E,α : CTE,α(A) → LNE,α(A), Sω
E,ατ := lim
t→ω
E,ατ
St
St
E,ατ (1)
,
is well defined. Moreover, (Sω
states in LNE,α(A).
E,α)2τ =
Sω
E,ατ
E,ατ (1) and Sω
Sω
E,α surjects onto the set of tracial
Proof. Statement 1 is an immediate consequence of Proposition 6.10. The first part of state-
ment 2 follows from the computation
St
E,ατ (1) − FE,αSt
E,ατ (1) = τ (1) = o(St
E,ατ (1)),
as t → 0 for τ ∈ CTE,α(A).
The second part of statement 2 follows from the fact that Proposition 6.9 implies that for
τ ∈ LNE,α(A),
St
E,ατ := (1 − e−t)−1τ.
54
Remark 6.12. For Example 6.6, the mapping FE,α takes the form FE,ατ = e−αL∗
gτ . In par-
ticular, the computations of Example 6.6 is a special case of the constructions in Proposition
6.11. We shall see that this holds in general below in Proposition 6.14.
Using our previous results, Proposition 3.8 (see page 29) and Proposition 6.10 (see page 54),
we can deduce a computation of heat traces.
Proposition 6.13. Let E be a strictly W-regular fpg bi-Hilbertian bimodule over a unital C ∗-
algebra A. If τ is a tracial state on A satisfying the Laca-Neshveyev condition for α ≥ 0,
then
Trτ (PDe−tDψ) =
1
1 − eα−t .
In particular, for any tracial state τ satisfying the Laca-Neshveyev condition for α ≥ 0 the
semifinite spectral triple (OE, L2(ΞA, τ ), Dψ, Nτ (ΞA), Trτ ) has positive Trτ -essential spectrum
with βD = β(E, τ ) = α.
Proof. We compute that
Trτ (PDe−tDψ ) =
τ∗(E⊗An)e−tn =
∞Xn=0
∞Xn=0
e−(t−α)n =
1
1 − eα−t .
In the first step we used Proposition 3.8 and in the second step we used Proposition 6.10.
We can now reformulate Theorem 6.5 in terms of the map FE,α and the constructions of
Proposition 6.11.
Proposition 6.14. Assume that E is a strictly W-regular fgp bi-Hilbertian bimodule over A.
Let α ≥ 0, ω be an extended limit as t → α and τ ∈ CTE,α(A) a critical trace. The KMS-state
φτ,ω defined from Theorem 6.5 takes the following form:
φτ,ω(SµS∗
ν ) = δµ,νe−αµSωα
E,ατ ((νµ)A), µ, ν ∈ Falg
E ,
where ωα is the extended limit at 0 obtained from translating ω by α.
Proof. The KMS-condition on OE reduces the proof to showing that φτ,ω(a) = Sωα
E,ατ (a) for
a ∈ A, just as in the proof of Theorem 6.5. Using Proposition 6.8 we can compute for a ∈ A
that
φτ,ω(a) = lim
n=0 TrE⊗n
n=0 TrE⊗n
τ
τ
(a)e−tn
(1)e−tn
t→ω0P∞
P∞
= lim
t→ωP∞
P∞
n=0 F n
n=0 F n
E,ατ (a)e−(t−α)n
E,ατ (1)e−(t−α)n
= Sωα
E,ατ (a).
Let KM Sα(OE ) denote the set of α-KMS states on OE for the gauge action and LNSE,α(A)
for the set of tracial states satisfying the Laca-Neshveyev condition.
Theorem 6.15. Let α ≥ 0 and let ω be an extended limit as t → α. Assume that E is a
strictly W-regular fgp bi-Hilbertian bimodule over A. The mapping
LNSE,α(A) → KM Sα(OE),
τ 7→ φτ,ω,
defined from Theorem 6.5, and revisited in Proposition 6.14, is a well defined bijection of sets.
More precisely, φτ,ω : OE → C is a KM Sα state for the gauge action, which is independent of
ω and coincides with φLN,τ .
55
Proof. By Theorem 6.5 and Proposition 6.13, the mapping τ 7→ φτ,ω is a well defined mapping
from the set of positive traces on A satisfying the Laca-Neshveyev condition for α and α-KMS-
states on OE for the gauge action. By the Laca-Neshveyev correspondence, the KMS-state
φτ,ω is uniquely determined by the trace φτ,ωA. We can therefore deduce that φτ,ω = φLN,τ
and the Theorem upon proving the identity φτ,ωA = τ . This statement follows immediately
from Proposition 6.14 and the second part of Proposition 6.11.
There are some quasi-invariance assumptions on traces that allows us to compare the KMS-
states constructed in Theorem 6.5 and the Laca-Neshveyev correspondence to a simpler con-
struction involving Φ∞. While the construction of Φ∞ depends on the left inner product on E,
the quasi-invariance condition we impose also depends on the left inner product. The following
quasi-invariance condition is a refinement of the notion of E-invariant functionals from [56].
Definition 6.16. Let α ≥ 0 and E a finitely generated projective bi-Hilbertian bimodule. We
say that a positive trace τ on A is α-quasi-invariant with respect to E and the extended limit
ω0 ∈ ℓ∞(N)∗ if for all n ∈ N and µ, ν ∈ E⊗n we have
e−αµτ ((νµ)A) = lim
k→ω0
τ (Φk(TµT ∗
ν )e−βk ) = lim
k→ω0
τ (A(µνeβk−ν)e−βk ).
(6.2)
If τ is α-quasi-invariant with respect to E and some extended limit, we simply say that τ is
α-quasi-invariant with respect to E.
Note, that Φk are defined on page 20 (formula (2.7) and the paragraph after).
Observe that if E is W-regular then the limit in the definition of quasi-invariance exists, and
so is independent of the extended limit ω0.
Remark 6.17. Just like in [56, Lemma 4.2], if E is full as a right module, then any positive
functional τ : A → C which is quasi-invariant in the sense of Definition 6.16 is a positive trace.
To see this, observe that for all µ, ν ∈ Falg
E and a ∈ A, the centrality of the Watatani indices
eβk ∈ A (see formula (2.6) for the definition) gives
e−αµτ ((νµ)Aa) = e−αµτ ((νµa)A) = lim
τ (A(µaνeβk−ν )e−βk )
τ (A(µνa∗eβk−ν)e−βk ) = e−αµτ ((νa∗µ)A)
k→ω0
= lim
k→ω0
= e−αµτ (a(νµ)A).
Example 6.18. We consider the module Eg over C(Y ) defined from a surjective local home-
omorphism g : Y → Y as in Example 6.6 (see page 52). In this case, βn = 0 for all n and
quasi-invariance of a positive trace τ given by a positive measure λ on Y is equivalent to the
condition
(Lg)∗λ = eαλ.
Another computation shows that this condition is equivalent to the Laca-Neshveyev condition.
In particular, for the module Eg, quasi-invariance is equivalent to satisfying the Laca-Neshveyev
condition.
Example 6.19. Let us consider the Cuntz algebra ON defined as the Cuntz-Pimsner algebra
of the C-bimodule CN . In this case, (νµ)A = A(µν) for all µ, ν ∈ (CN )⊗n and βn = n log(N ).
Therefore, quasi-invariance of a trace τ on C is equivalent to
eατ = N τ.
56
That is, any non-zero trace on C is log(N )-quasi invariant.
Our immediate aim is to connect the quasi-invariance of Definition 6.16 with the condition
imposed by Laca-Neshveyev, [45].
Lemma 6.20. Let E be an fgp bi-Hilbertian bimodule. Suppose that τ satisfies the α-quasi-
invariance condition of Definition 6.16 with respect to E. Then τ satisfies the Laca-Neshveyev
condition for α.
Proof. This is a computation using Definition 6.16 of quasi-invariance and a frame (ej) for EA.
So
τ ((ejaej )A) = eα lim
k→ω0
Taej T ∗
ej )e−βk )
TrE
τ (La) = Trτ(cid:18)Xj
aΘej,ej(cid:19) =Xj
τ (Xj
= eα lim
k→ω0
τ (Φk(Xj
aA(ejej eβk−1)e−βk ) = eα lim
k→ω0
τ (aeβk e−βk ) = eατ (a).
If τ : A → C satisfies the α-quasi-invariance condition of Definition 6.16, then we can rewrite
the Laca-Neshveyev KMS state φLN,τ : OE → C as
φLN,τ (TµT ∗
ν ) = δµ,νe−αµτ ((µν)A) = lim
k→ω0
τ (Φk(TµT ∗
ν )e−βk ).
This computation proves the following. Note that the next result does not require any W-
regularity from the module E.
Proposition 6.21. Let E be an fgp bi-Hilbertian bimodule and consider an α-quasi-invariant
positive trace τ : A → C with respect to E. The state on OE defined by
SµS∗
ν 7→ lim
k→ω0
τ (Φk(TµT ∗
ν )e−βk ),
(6.3)
is α-KMS for the gauge action on OE and coincides with φLN,τ .
ν ) := limk→∞ Φk(TµT ∗
If E is W-regular, the definition Φ∞(SµS∗
Equation (6.3) coincides with τ ◦ Φ∞. We conclude the following.
Corollary 6.22. Let τ : A → C be an α-quasi-invariant positive trace with respect to an fgp
bi-Hilbertian bimodule E. Assume that E is a W-regular. Then, the state τ ◦ Φ∞ on OE is
α-KMS for the gauge action on OE and coincides with φLN,τ .
ν )e−βk shows that the state in
By Theorem 6.15 and Corollary 6.22, we have that φω,τ = τ ◦ Φ∞ for any α-quasi-invariant
positive trace τ and extended limit ω at α assuming that E is strictly regular.
6.3 Obstructions to bi-Hilbertian bimodule structures
In the two previous subsections, we assumed our A-bimodule E to be an fgp bi-Hilbertian
bimodule, and imposed the additional assumption of strict W-regularity (see Definitions 2.22
and 2.25). The assumptions allowed us to construct a semifinite spectral triple from a Kasparov
module relying on both the left and the right inner product on E. Instead, we can just use [45]
57
to proceed from a KMS-state directly to a semifinite spectral triple whose associated KMS-
state as in Corollary 3.21 coincides with the original KMS-state.
In order to compare the
indirect approach for strictly W-regular modules to the direct approach from the KMS-state
we will need to extend our module to von Neumann algebra coefficients, and along the way
we derive obstructions to having the structure of a strictly W-regular bi-Hilbertian bimodule
structure on an A − A-correspondence.
We suppose that we have a finitely generated projective right A-module EA, with A unital,
and carrying a unital left action of A. Let τ : A → C be a faithful positive trace satisfying the
Laca-Neshveyev condition for α ≥ 0 (see Definition 6.7 on page 53), and define the associated
KMS-state on OE by
φLN,τ (SµS∗
ν ) := δµ,ντ ((νµ)A)e−αµ, µ, ν ∈ Falg
E .
By construction, φLN,τA = τ .
The Cuntz-Pimsner algebra OE acts on the GNS-space L2(OE, φLN,τ ) by left multiplication
and, since φLN,τ restricts to a trace on A, A acts by both left and right multiplication on
L2(OE, φLN,τ ). For notational simplicity, we identify τ with its normal extension to A′′. Note
that A′′ is independent of whether we take the bicommutant in L2(OE, φLN,τ ) or L2(A, τ ) and
by faithfulness of τ we can identify A with its image under the GNS-representation and obtain
an inclusion A ⊆ A′′.
Proposition 6.23. Let E be an fgp right A-Hilbert C ∗-module with a left unital action of
A, τ a faithful positive trace on A satisfying the Laca-Neshveyev condition and let P0 :
L2(OE, φLN,τ ) → L2(A, τ ) ⊂ L2(OE , φLN,τ ) denote the orthogonal projection. It then holds
that
A′′ = P0O′′
EP0.
Proof. It is clear that A′′ ⊆ P0O′′
write T = P0T0P0 where T0 is the WOT-limit of a net Tλ =Pj Sµλ,j S∗
EP0. To prove the converse inclusion, take T ∈ P0O′′
P0TλP0 = Xj:µλ,j=νλ,j=0
EP0 and
νλ,j ∈ OE. We have that
Sµλ,j S∗
νλ,j ,
so P0TλP0 ∈ A and T ∈ A′′.
We can define a conditional expectation
Φ∞ : O′′
E → A′′,
Φ∞(S) := P0SP0,
which is well defined by Proposition 6.23. Using the expectation Φ∞ we can define a right
module ΞA′′ by completing OE in the norm defined by the inner product
(S1S2)A′′ := Φ∞(S∗
1 S2), S1, S2 ∈ OE.
It is clear that L2(OE, φLN,τ ) = L2(ΞA′′, τ ). The construction of ΞA′′ does not require EA to
be biHilbertian, just an A-A-correspondence. The following result follows from the relations
defining the Cuntz-Pimnser algebra and the fact that Φ∞ is a conditional expectation.
58
Proposition 6.24. For µ ∈ E⊗k and ν ∈ E⊗l,
Φ∞(S∗
µSν) = δµ,ν(µν)A.
In particular, the map µ 7→ Sµ extends to an A′′-linear isometric embedding FE ⊗A A′′ → ΞA′′
of the Fock module.
We now turn to describing the WOT-closure of E inside O′′
A-sub-bimodule of OE via µ 7→ Sµ.
Lemma 6.25. Let A be a unital C ∗-algebra. Let EA be a finitely generated projective right
A-module with a unital left action, and suppose that τ : A → C satisfies the Laca-Neshveyev
condition on E for α ≥ 0. Then
E. We will identify E with an
WOT
is an A′′-bimodule. Moreover, the following hold:
E′′ := E
⊂ O′′
E
1. As right A′′-modules, E′′ ∼= E ⊗A A′′ and the isomorphism is an isomorphism of A′′-
Hilbert C ∗-modules when equipping E′′ with the right inner product
(µν)A′′ := Φ∞(S∗
µSν), µ, ν ∈ E′′.
2. The right A′′-Hilbert C ∗-module E′′ is finitely generated and projective.
3. If E is finitely generated and projective as a left A-module, then E′′ is finitely generated
and projective as a left A′′-module.
4. If the implication
holds, the expression
P0ee∗P0 = 0 ⇒ e = 0 ∀e ∈ E′′,
(6.4)
gives a left inner product on E′′ making it into a bi-Hilbertian bimodule. The right
Watatani index of E′′ is 1 and qk = Id(E′′)⊗k for all k (and so is invertible).
A(ef ) := P0SeS∗
f P0
5. If E′′ is a finitely generated projective module from the left and the implication (6.4)
holds, then E′′ is a strictly W -regular fgp bi-Hilbertian bimodule over A′′.
Proof. The initial statement of the lemma is clear since E is an A-sub-bimodule of OE, so its
WOT closure is a bimodule over A′′. To prove statement 1. we note that E⊗AA′′ ∼= EA′′ ⊆ O′′
E.
Moreover, using that E is finitely generated and projective, it follows that E′′ = EA′′ and
therefore E′′ ∼= E ⊗A A′′ follows. Statement 1. now follows from Proposition 6.24. Statement
2. follows from statement 1. because E is finitely generated and projective over A. Statement
3. is proven in a similar way as Statement 1., indeed if E is finitely generated and projective
as a left A-module then E′′ ∼= A′′ ⊗A E as a left A-module.
Statement 4. is less trivial. Assuming that the implication (6.4) holds, it is straight-forward to
verify that the left and right actions are compatible, i.e. that the A-action from the left/right is
adjointable for the right/left inner product. For E′′ to be a bi-Hilbertian bimodule it remains
59
to show that the norm arising from the left inner product is equivalent to the norm arising
from the right inner product. For any e ∈ E, we compute that
kA(ee)kA = kP0SeS∗
e P0kL2(A,τ ) = kS∗
e P0SekL2(E∗,φτ ) = kS∗
e SekL2(E∗,φτ ) =
(6.5)
are equivalent.
where the norm of S∗
e Se is attained by S∗
= k(ee)AkL2(E∗,φτ ) = k(ee)AkA
e P0Se on L2(E∗, φτ ) by the assumption that A(··) is
positive definite. Equation (6.5) shows that the two normspk(··)AkA andpkA(··)kA on E
To finalize the proof of statement 4., we compute the right Watatani index of E, which exists be-
cause E′′ is finitely generated projective by statement 2. We compute on L2(A, τ ) ⊂ L2(OE, φτ )
that
ha,Xj
P0π(1)(Θej,ej )P0ai = φτ (a∗P0π(1)(IdE)P0a) = φτ (a∗1Aa) = τ (a∗a).
A(ejej)ai = ha,Xj
By the faithfulness of τ and Cuntz-Pimsner covariance we can now deduce that the right
Watatani index is equal to 1A. This immediately implies that qk = Id(E′′)⊗k for all k.
Finally, statement 5. follows from that under the stated assumptions, E′′ is an fgp bi-Hilbertian
bimodule (using statements 1., 2. and 4.) and by statement 4. qk satisfies the condition in
Definition 2.25, so E′′ is strictly W-regular.
Remark 6.26. There are examples of A-A-correspondences E that are fgp from the right but
not fgp from the left such that E′′ is fgp from the left. These examples come from (certain)
self-similar dynamical systems, see [43].
Here is a simple example. Let A = C([0, 1]),
γ1 : [0, 1] → [0, 1] γ1(x) = x/2,
γ2 : [0, 1] → [0, 1] γ2(x) = 1/2 + x/2,
and E = C({(γ1(x), x) : x ∈ [0, 1]} ∪ {(γ2(x), x) : x ∈ [0, 1]}). The correspondence structure is
defined for a, b ∈ A and e ∈ E by
(a · e · b)(γj(x), x) = a(γj(x))e(γj (x), x)b(x)
e1(γj(x), x)e(γj (x), x).
and (e1e2)A(x) = Xj=1,2
The graphs of γ1 and γ2 in [0, 1]× [0, 1] are disjoint and their respective characteristic function
χ1 and χ2 are elements of E. One checks directly that {χ1, χ2} is a right frame for EA, and
since (χ1χ2) = 0, there is an isomorphism of right Hilbert C ∗-modules EA ∼= C[0, 1] ⊕ C[0, 1].
We conclude that EA is fgp from the right. Using the frame {χ1, χ2}, we can identify AE ∼=
C[0, 1/2]⊕C[1/2, 1] as a left C[0, 1]-module. As a left module, AE is therefore finitely generated
but clearly not projective as the rank of Ex := E/C0([0, 1]\{x})E is discontinuous at x = 1/2.
We shall now see that it is even impossible for a left inner product compatible with the right
inner product to exist. Let 0 ≤ φ ∈ A be 1 on [0, 1/2]. Then if we have a compatible left
A-valued inner product
A(χ1χ2)(x) = A(φ · χ1χ2)(x) = φ(x)A(χ1χ2)(x) = A(χ1χ2)(x)φ(x) = A(χ1φ · χ2)(x).
60
Taking the infimum over such φ we see that the support of A(χ1χ2) is contained in {1/2}.
Then for arbitrary a, b ∈ A
A(aχ1 + bχ2aχ1 + bχ2)
= aA(χ1χ1)a∗ + bA(χ2χ2)b∗ + a(1/2)b∗(1/2)A(χ1χ2) + b(1/2)a∗(1/2)A(χ2χ1).
From here one can show that any inner product taking values in the continuous functions takes
values in the functions vanishing at 1/2. Then one shows that the associated norm can not be
equivalent to the right inner product.
The situation is better for E′′. We consider the trace τ (a) :=R 1
0 a(x)dx on C[0, 1]. A short
computation shows that TrE
τ = 2τ so τ satisfies the Laca-Neshveyev condition for α = log(2)
and extends to a KMS-state on OE at inverse temperature log(2). It is readily verified that
C[0, 1]′′ = L∞[0, 1], E′′ ∼= L∞[0, 1] ⊕ L∞[0, 1] as a right module and E′′ = L∞[0, 1/2] ⊕
L∞[1/2, 1] as a left module. In particular, E′′ is an fgp bi-Hilbertian bimodule over L∞[0, 1].
The same discussion applies to any 'graph separated' iterated function system satisfying the
open set condition. See [43] for more details.
Theorem 6.27. Let EA be a right A-Hilbert C ∗-module with a unital left action and assume
that E is finitely generated both as a left and a right A-module and E′′ is finitely generated and
projective both as a left and a right A′′-module. Then EA has a left inner product such that
E is an fgp bi-Hilbertian bimodule, has finite right Watatani index and is W -regular with qk
invertible for all k if and only if Φ∞ is faithful and Φ∞(OE ) ⊆ A.
We remark that if E is W-regular and qk is invertible for all k, then E is strictly W-regular by
[31, Lemma 3.8].
Proof. First suppose that Φ∞ is faithful and Φ∞(OE) ⊆ A. Lemma 6.25 shows that E′′ has
a left inner product making E (not just E′′) bi-Hilbertian with right Watatani index 1A and
(invertible) qk = Id(E)⊗k . Since E, both as a left and a right module, is finitely generated and
admits a Hilbert C ∗-module structure it is also projective.
Conversely, if E is an fgp bi-Hilbertian bimodule with finite right Watatani index and is W-
regular with invertible qk, then [56] proves that the map Φ∞(SµS∗
ν ) = A(µqν(ν)) is a (faithful)
conditional expectation. That it agrees with Φ∞ is a computation.
Remark 6.28. The issue with qk being non-invertible is as follows. Since qk = q1⊗ q1⊗···⊗ q1,
the non-invertibility occurs with q1. Supposing OE to be strictly W-regular, we can define the
right module ΞA as the completion of OE for the norm coming from Φ∞. Then Φ∞(SeS∗
f ) =
A(eq1(f )), and so if q1 is not invertible, we do not get a left inner product in this way. See [56,
Example 3.10] for an example where we have strict W-regularity with q1 not being invertible.
Heuristically, one should view the passage to ΞA as erasing the information about the left inner
product on E, corresponding to the kernel of q1. On the other hand, if qk is invertible for all k
then we can replace our left inner products A(··)E⊗k
and obtain an equivalent
inner product structure with right Watatani index 1.
by A(·qk(·))E⊗k
We can now relate our constructions above back to KMS-states.
61
Proposition 6.29. Let A be a unital C ∗-algebra, E a finitely generated projective right A-
Hilbert C ∗-module with a unital adjointable left A-action, α ≥ 0 and τ a positive trace on A.
We assume that this data satisfies the following conditions:
• τ satisfies the Laca-Neshveyev condition.
• The A′′-bimodule E′′ is finitely generated and projective from the left.
• The implication (6.4) (see page 59) holds.
We define a semifinite spectral triple (OE, L2(OE, φLN,τ ), Dψ, N, T) using that L2(OE , φLN,τ ) =
L2(OE′′, φLN,τ ) = L2(ΞA′′, τ ) and pulling back the semi-finite spectral triple defined from the
fgp bi-Hilbertian A′′-bimodule E′′ as in Lemma 2.28 along the inclusion OE → OE′′. This
semi-finite spectral triple is Li1-summable, α-analytic, has positive essential T-spectrum with
βDψ = α and its associated KMS-state for the gauge action (as in Corollary 3.21 on page 34)
coincides with φLN,τ .
References
[1] A. An Huef, M. Laca, I. Raeburn, A. Sims, KMS states on the C ∗-algebras of finite
graphs, J. Math. Anal. and App., 405 (2), (2013), 388-399.
[2] A. an Huef, I. Raeburn, Equilibrium States on Graph Algebras, In: Carlsen T.M., Larsen
N.S., Neshveyev S., Skau C. (eds) Operator Algebras and Applications. Abel Symposia,
vol 12. Springer, Cham, (2016).
[3] N. H. Bingham, C. M. Goldie, J. L. Teugels, Regular variation, volume 27 of Ency-
clopedia of Mathematics and its Applications. Cambridge University Press, Cambridge,
1989.
[4] O. Bratteli, D. W. Robinson, Operator algebras and quantum statistical mechanics. 2.
Equilibrium states. Models in quantum statistical mechanics. Second edition. Texts and
Monographs in Physics. Springer-Verlag, Berlin, 1997.
[5] A. Carey, V. Gayral, A. Rennie, F. Sukochev, Integration on locally compact noncom-
mutative spaces, J. Funct. Anal., 263 (no. 2) (2012), 383 -- 414.
[6] A. Carey, V. Gayral, A. Rennie, F. Sukochev, Index theory for locally compact noncom-
mutative geometries, Mem. Amer. Math. Soc. 231 no. 1085, (2014), vi+130 pp.
[7] A. Carey, S. Neshveyev, R. Nest, A. Rennie, Twisted cyclic theory, equivariant KK-
theory and KMS states, J. reine angew. Math., 650 (2011), 161 -- 191.
[8] A. Carey, D. O'Brien, Automorphisms of the infinite dimensional Clifford algebra and
the Atiyah-Singer mod 2 index theorem, Topology, 22 no. 4 (1983), 437 -- 448.
[9] A. Carey, J. Phillips, Spectral flow in Fredholm modules, eta invariants and the JLO
cocycle, K-Theory 31 no. 2 (2004), 135 -- 194.
[10] A. Carey, J. Phillips, A. Rennie, Twisted cyclic theory and an index theory for the gauge
invariant KMS state on Cuntz algebras, J. K-Theory, 6 no. 2 (2010), 339 -- 380.
62
[11] A. L. Carey, J. Phillips, F. A. Sukochev, Spectral flow and Dixmier traces, Adv. Math.,
173 (2003), 68 -- 113.
[12] A. Carey, A. Rennie, K. Tong, Spectral flow invariants and twisted cyclic theory for the
Haar state on SUq(2), J. Geom. Phys., 59 (2009), 1431 -- 1452.
[13] J. Christensen, K. Thomsen, Finite digraphs and KMS states, J. Math. Anal. Appl. 433
(2016), no. 2, 1626 -- 1646.
[14] A. Connes, Compact metric spaces, Fredholm modules, and hyperfiniteness, Ergodic
Theory Dynam. Systems 9 (1989), no. 2, pp. 207 -- 220.
[15] A. Connes, Noncommutative geometry, Academic Press, Inc., San Diego, CA, 1994.
[16] A. Connes, Geometry from a spectral point of view, Lett. Math. Phys. 34 (1995), 203 --
238.
[17] A. Connes, J. Cuntz, Quasi homomorphismes, cohomologie cyclique et positivit´e, Com-
mun. Math. Phys. 114 (1988), 515 -- 526.
[18] V. Deaconu, Generalized solenoids and C ∗-algebras, Pacific J. Math. 190 (1999), 247 --
260.
[19] R. J. Deeley, M. Goffeng, B. Mesland, and M. F. Whittaker, Wieler solenoids, Cuntz-
Pimsner algebras and K-theory, Ergodic Theory Dynam. Systems 38 (2018), no. 8,
2942 -- 2988.
[20] J. Dixmier, Existence de traces non normales, C. R. Acad. Sci. Paris S´er. A-B 262
(1966), A1107 -- A1108.
[21] J. J. Duistermaat, I. M. Singer, Order-preserving isomorphisms between algebras of
pseudo-differential operators, Comm. Pure Appl. Math. 29 (1976), no. 1, 39 -- 47.
[22] J. V. Egorov, The canonical transformations of pseudodifferential operators, Uspehi
Mat. Nauk 24 1969 no. 5 (149) 235 -- 236.
[23] M. Enomoto, M. Fujii, Y. Watatani, KMS states for gauge action on OA, Math. Japon.,
29 (4). (1984), 607 -- 619.
[24] T. Fack, H. Kosaki, Generalised s-numbers of τ -measurable operators, Pacific J. Math.,
123 (1986), 269 -- 300.
[25] V. Gayral, F. Sukochev, Dixmier traces and extrapolation description of noncommuta-
tive Lorentz spaces. J. Funct. Anal., 266 (10):6256 -- 6317, 2014.
[26] E. Getzler, A. Szenes, On the Chern character of a theta-summable Fredholm module,
J. Funct. Anal. 84 (1989), no. 2, 343 -- 357.
[27] P. Gilkey, Invariance theory, the heat equation, and the Atiyah-Singer index theorem,
2nd edition, CRC Press, 1995.
[28] H. Gimperlein, G. Grubb, Heat kernel estimates for pseudodifferential operators, frac-
tional Laplacians and Dirichlet-to-Neumann operators, J. Evol. Equ. 14 (2014), no. 1,
49 -- 83.
63
[29] M. Goffeng, B. Mesland, Spectral triples and finite summability on Cuntz-Krieger alge-
bras, Doc. Math. 20 (2015), 89 -- 170.
[30] M. Goffeng, B. Mesland, Spectral triples on ON , Conference proceedings, MATRIX-
program "Refining C ∗-algebraic invariants for dynamics using KK-theory" in Creswick,
Australia, 2016.
[31] M. Goffeng, B. Mesland, A. Rennie, Shift-tail equivalence and an unbounded represen-
tative of the Cuntz-Pimsner extension, Ergodic Theory Dynam. Systems 38 (2018), no.
4, 1389 -- 1421.
[32] M. Goffeng, A. Usachev, Dixmier traces and residues on weak operator ideals.
arXiv:1710.08260
[33] G. Grubb, R. T. Seeley, Zeta and eta functions for Atiyah-Patodi-Singer operators, J.
Geom. Anal. 6 no. 1 (1996), 31 -- 77.
[34] N. Higson, J. Roe, Analytic K-Homology, Oxford Mathematical Monographs. Oxford
Science Publications. Oxford University Press, Oxford, 2000. xviii+405 pp.
[35] L. Hormander, Fourier integral operators. I. Acta Math. 127 (1971), no. 1-2, 79 -- 183.
[36] A. Jaffe, A. Lesniewski, and K. Osterwalder, Quantum K-theory. I. The Chern charac-
ter, Comm. Math. Phys. 118 no. 1 (1988), 1 -- 14.
[37] A. Jaffe, A. Lesniewski, and K. Osterwalder, On super-KMS functionals and entire
cyclic cohomology, K-Theory 2 no. 6 (1989), 675 -- 682.
[38] M. Junge, T. Mei, J. Parcet, Noncommutative Riesz transforms?dimension free bounds
and Fourier multipliers, J. Eur. Math. Soc. (JEMS) 20 (2018), no. 3, 529 -- 595.
[39] J. Kaad, R. Nest, A. Rennie, KK-theory and spectral flow in von Neumann algebras,
J. K-Theory, 10 (no. 2) (2012), 241 -- 277.
[40] J. Kaad, W. D. van Suijlekom, Riemannian submersions and factorization of Dirac
operators, arXiv:1610.02873.
[41] T. Kajiwara, C. Pinzari, Y. Watatani, Jones index theory for Hilbert C ∗-bimodules and
its equivalence with conjugation theory, J. Funct. Anal. 215 (2004), 1 -- 49.
[42] T. Kajiwara, Y. Watatani, KMS states on finite-graph C ∗-algebras, Kyushu J. Math.
67 (1) (2013), 83 -- 104.
[43] T. Kajiwara, Y. Watatani, C ∗-algebras associated with self-similar sets, J. Operator
Theory, 56 (2) (2006), 225 -- 247.
[44] J. Korevaar, Tauberian theory, volume 329 of Grundlehren der Mathematischen
Wissenschaften [Fundamental Principles of Mathematical Sciences]. Springer-Verlag,
Berlin, 2004. A century of developments.
[45] M. Laca, S. Neshveyev, KMS states of quasi-free dynamics on Pimsner algebras, J.
Funct. Anal. 211 (2004), 457 -- 482.
64
[46] E. C. Lance, Hilbert C ∗-modules. A toolkit for operator algebraists. London Mathemat-
ical Society Lecture Note Series, 210. Cambridge University Press, Cambridge, 1995.
x+130 pp.
[47] S. Lord, A. Rennie, J. V´arilly, Riemannian manifolds in noncommutative geometry, J.
Geom. Phys. 62 (2012), 1611 -- 1638.
[48] S. Lord, F. Sukochev, D. Zanin, Singular Traces: Theory and Applications, volume 46
of Studies in Mathematics. De Gruyter, 2012.
[49] L. MacDonald, A. Rennie, The Godbillon-Vey invariant in equivariant KK-theory,
arXiv:1811.04603.
[50] B. Mesland, M. H. Sengun, Hecke operators in KK-theory and the K-homology of
Bianchi groups, to appear in Journal of Noncommutative Geometry, arXiv:1610.06808.
[51] M. Mimura, Rigidity Theorems For Universal And Symplectic Universal Lattices,
PhD-thesis University of Tokyo, 2011, https://www.math.tohoku.ac.jp/∼mimura-
mas/mimura phdthesis.pdf
[52] V. Nistor, A bivariant Chern-Connes character, Ann. of Math. (2) 138 no. 3 (1993),
555 -- 590.
[53] N. Ozawa, Amenable actions and exactness for discrete groups, C. R. Acad. Sci. Paris
Sr. I Math. 330 (2000), no. 8, 691 -- 695.
[54] I. Raeburn, Graph Algebras, Published for the Conference Board of the Mathematical
Sciences, Washington, DC, 2005, vi+113.
[55] I. Raeburn, D. P. Williams, Morita Equivalence and Continuous-Trace C ∗-algebras,
Mathematical Surveys and Monographs 60, American Mathematical Society, Provi-
dence, RI, 1998.
[56] A. Rennie, D. Robertson, A. Sims, The extension class and KMS states for Cuntz --
Pimsner algebras of some bi-Hilbertian bimodules, Journal of Topology and Analysis,
09 no. 02, (2017) 297 -- 327.
[57] A. Rennie, D. Robertson, A. Sims, Poincar´e duality for Cuntz-Pimsner algebras of
bimodules, arXiv:1804.08114.
[58] A. Rennie, R. Senior, The resolvent cocycle in twisted cyclic cohomology and a local
index formula for the Podle´s sphere, J. Noncommut. Geom., 8 (no. 1) (2014), 1 -- 43.
[59] E. Schrohe, M. Walze, J.-M. Warzecha, Construction de triplets spectraux ´a partir de
modules de Fredholm, C. R. Acad. Sci. Paris S´er. I Math. 326 no. 10 (1998), 1195 -- 1199.
[60] A. Sedaev, F. Sukochev, Dixmier measurability in Marcinkiewicz spaces and applica-
tions, J. Funct. Anal. 265 no. 12 (2013), 3053 -- 3066.
[61] F. Sukochev, A. Usachev, D. Zanin, Dixmier traces generated by exponentiation invari-
ant generalised limits, J. Noncommut. Geom. 8 no. 2 (2014), 321 -- 336.
65
[62] M. Takesaki, Theory of operator algebras. I. Reprint of the first (1979) edition. En-
cyclopaedia of Mathematical Sciences, 124. Operator Algebras and Non-commutative
Geometry, 5. Springer-Verlag, Berlin, 2002. xx+415 pp.
[63] D. Voiculescu, On the existence of quasicentral approximate units relative to normed
ideals. Part I, J. Funct. Anal. 91 (1990), 1 -- 36.
[64] P. Walters, Convergence of the Ruelle operator for a function satisfying Bowen's condi-
tion, Trans. Amer. Math. Soc. 353 no. 1 (2001), 327 -- 347.
[65] G. P. Wene, The Clifford algebra of an infinite-dimensional space, J. Math. Phys. 30
(1989), 249 -- 251.
66
|
1604.08774 | 3 | 1604 | 2017-04-03T16:35:54 | Just-infinite C*-algebras | [
"math.OA",
"math.GR"
] | By analogy with the well-established notions of just-infinite groups and just-infinite (abstract) algebras, we initiate a systematic study of just-infinite C*-algebras, i.e., infinite dimensional C*-algebras for which all proper quotients are finite dimensional. We give a classification of such C*-algebras in terms of their primitive ideal space that leads to a trichotomy. We show that just-infinite, residually finite dimensional C*-algebras do exist by giving an explicit example of (the Bratteli diagram of) an AF-algebra with these properties.
Further, we discuss when C*-algebras and *-algebras associated with a discrete group are just-infinite. If $G$ is the Burnside-type group of intermediate growth discovered by the first named author, which is known to be just-infinite, then its group algebra $C[G]$ and its group C*-algebra $C^*(G)$ are not just-infinite. Furthermore, we show that the algebra $B = \pi(C[G])$ under the Koopman representation $\pi$ of $G$ associated with its canonical action on a binary rooted tree is just-infinite. It remains an open problem whether the residually finite dimensional C*-algebra $C^*_\pi(G)$ is just-infinite. | math.OA | math |
Just-infinite C -algebras
Rostislav Grigorchuk∗, Magdalena Musat: and Mikael Rørdam†
Dedicated to Efim Zelmanov on the occasion of his 60th birthday
Abstract
By analogy with the well-established notions of just-infinite groups and just-infinite
(abstract) algebras, we initiate a systematic study of just-infinite C -algebras, i.e., infi-
nite dimensional C -algebras for which all proper quotients are finite dimensional. We
give a classification of such C -algebras in terms of their primitive ideal space, that leads
to a trichotomy. We show that just-infinite, residually finite dimensional C -algebras
do exist by giving an explicit example of (the Bratteli diagram of) an AF-algebra with
these properties.
Further, we discuss when C -algebras and -algebras associated with a discrete group
are just-infinite. If G is the Burnside-type group of intermediate growth discovered by
the first-named author, which is known to be just-infinite, then its group algebra CrGs
and its group C -algebra C pGq are not just-infinite. Furthermore, we show that the
algebra B " πpCrGsq under the Koopman representation π of G associated with its
canonical action on a binary rooted tree is just-infinite. It remains an open problem
whether the residually finite dimensional C -algebra C
π pGq is just-infinite.
1
Introduction
A group is said to be just-infinite if it is infinite and all its proper quotients are finite.
Just-infinite groups arise, e.g., as branch groups (including the Burnside-type group of
intermediate growth discovered by the first named author, see [19]). A trichotomy describes
the possible classes of just-infinite groups, see [22, Theorem 3]. Each finitely generated
infinite group has a just-infinite quotient. Therefore, if we are interested in finitely generated
infinite groups satisfying a certain exotic property preserved by homomorphic images, if such
a group exists, then one is also to be found in the class of just-infinite groups.
The purpose of this paper is to investigate just-infinite dimensional C -algebras, defined
to be infinite dimensional C -algebras for which all proper quotients by closed two-sided
ideals are finite dimensional. (In the future, we shall omit "dimensional" and refer to these
C -algebras as just-infinite. The well-established notion of infiniteness of a unital C -al-
gebra, that is, its unit is Murray-von Neumann equivalent to a proper subprojection, is
unrelated to our notion of just-infiniteness.) Analogous to just-infiniteness in other cate-
gories, any infinite dimensional simple C -algebra is just-infinite for trivial reasons. It is
also easy to see that if a C -algebra A contains a simple essential closed two-sided ideal
I such that A{I is finite dimensional, then A is just-infinite. (A closed two-sided ideal in
a C -algebra is essential if it has non-zero intersection with every other non-zero closed
∗The first named author was supported by NSF grant DMS-1207699 and NSA grant H98230-15-1-0328
†The second and third named authors were supported by the Danish National Research Foundation
(DNRF) through the Centre for Symmetry and Deformation at University of Copenhagen, and The Danish
Council for Independent Research, Natural Sciences.
1
two-sided ideal.) Hence, e.g., any essential extension of the compact operators on an infinite
dimensional separable Hilbert space by a finite dimensional C -algebra is just-infinite.
We give in Theorem 3.10 a classification of just-infinite C -algebras into three types,
depending on their primitive ideal space. In more detail, if A is a separable just-infinite
C -algebra, then its primitive ideal space is homeomorphic to one of the T0-spaces Yn,
0 ď n ď 8, defined in Example 3.7. The case n " 0 corresponds to A being simple, while
the case 1 ď n ă 8 occurs when A is an essential extension of a simple C -algebra by
a finite dimensional C -algebra with n simple summands. If the primitive ideal space of
a separable just-infinite C -algebra A is infinite, then it is homeomorphic to the T0-space
Y8, and in this case A is residually finite dimensional (i.e., there is a separating family of
finite dimensional representations of A). The C -algebra A has an even stronger property,
described in Section 2, that we call strictly residually finite dimensional. We refer the reader
to the survey paper [6] for a more comprehensive treatment of residually finite groups and
residually finite dimensional group C -algebras.
To our knowledge, residually finite dimensional, for short RFD, just-infinite C -alge-
bras have not been previously considered in the literature. A priori it is not even clear
that they exist. This issue is settled in Section 4, where we construct a RFD just-infinite
unital AF-algebra, by giving an explicit description of its Bratteli diagram. Residually finite
dimensional C -algebras have been studied extensively, see for example [18], [15], [14], [6],
[13]. They are always quasidiagonal (see, e.g., [12] or [13, Chapter 7]). Interesting classes
of C -algebras, such as the full group C -algebras of the free groups and subhomogenous
C -algebras, are RFD. Among RFD C -algebras, the just-infinite ones are distinguished by
having the smallest possible ideal lattice.
In Section 5, we show that unital, separable, RFD just-infinite C -algebras need not
be AF-algebras, nor nuclear, or even exact. Using a construction of Dadarlat from [14], we
show that the just-infinite, residually finite dimensional AF-algebra constructed in Section 4
contains a RFD just-infinite, non-nuclear sub-C -algebra. Moreover, this AF-algebra is
contained in a non-exact C -algebra, which, likewise, is RFD and just-infinite.
Just-infiniteness for C -algebras is less prevalent than the corresponding property in the
category of groups. Not every infinite dimensional C -algebra has a just-infinite quotient,
since, for example, no abelian C -algebra is just-infinite; cf. Example 3.3. There seems to
be no natural condition ensuring that a C -algebra has a just-infinite quotient.
We discuss in Section 6 when a group C -algebra is just-infinite, depending on properties
of the group. We prove that the full group C -algebra C pGq of a discrete group G is just-
infinite if and only if its group algebra CrGs has a unique (faithful) C -norm and it is
-just-infinite, i.e., is just-infinite as a -algebra. The former property holds trivially for
any locally finite group. We do not know of any non-locally finite group for which CrGs
has unique C -norm. We show that there are locally finite just-infinite groups G, for which
the group C -algebra C pGq and the group algebra CrGs are just-infinite. If the reduced
group C -algebra C
λ pGq must be simple, or G must be
λpGq coincides with the full group C -algebra C pGq. It seems
amenable, in which case C
plausible that the group C -algebra associated with unitary representations other than the
universal or the left-regular one might be just-infinite for a larger class of groups.
λ pGq is just-infinite, then either C
If the group algebra CrGs of a group G is -just-infinite, then G must be just-infinite,
but the converse does not hold. Indeed, we show in Section 7 (Theorem 7.10) that CrGs
is not -just-infinite whenever G is a branch group, while there are many branch groups
which are just-infinite, e.g., the group G of intermediate growth mentioned above. We show
that the image B " πpCrGsq of CrGs under the Koopman representation π of G, associated
with the canonical action of G on a binary rooted tree, is just-infinite. We leave open the
2
question whether or not the C -completion C
π pGq of B is just-infinite. In the affirmative
case, this would provide an example of a RFD just-infinite C -algebra arising from a group.
2 Preliminaries
As we shall later describe just-infinite C -algebras in terms of their primitive ideal space,
and as the interesting cases of just-infinite C -algebras are those that are residually finite
dimensional, we review in this section the relevant background.
2.1 The primitive ideal space of a C -algebra
A C -algebra A is said to be primitive if it admits a faithful irreducible representation
on some Hilbert space. It is said to be prime if, whenever I and J are closed two-sided
ideals in A such that I X J " 0, then either I " 0, or J " 0.
It is easy to see that
every primitive C -algebra is prime, and it is a non-trivial result that the converse holds
for all separable C -algebras; cf. [31, Proposition 4.3.6]. However, there are (complicated)
examples of non-separable C -algebras that are prime, but not primitive, see [37].
A closed two-sided ideal I in a C -algebra A is said to be primitive if I ‰ A and I
is the kernel of an irreducible representation of A on some Hilbert space. The primitive
ideal space, PrimpAq, is the set of all primitive ideals in A. A closed two-sided ideal I
of A is primitive if and only if the quotient A{I is a primitive C -algebra. In particular,
0 P PrimpAq if and only if A is primitive. The primitive ideal space is a T0-space when
equipped with the hull-kernel topology, which is given as follows: the closure F of a subset
F Ď PrimpAq consists of all ideals I P PrimpAq which contain ŞJPF J. If A is primitive,
so that 0 P PrimpAq, then t0u " PrimpAq. In the commutative case, the primitive ideal
space is the usual spectrum: PrimpC0pXqq is homeomorphic to X, whenever X is a locally
compact Hausdorff space. The following fact will be used several times in the sequel:
Remark 2.1. Each finite dimensional C -algebra A is (isomorphic to) a direct sum of full
matrix algebras,
A – Mk1pCq ' Mk2pCq ' ' MknpCq,
for some positive integers n, k1, k2, . . . , kn. As each matrix algebra is simple, PrimpAq can
be naturally identified with the set t1, 2, . . . , nu, equipped with the discrete topology. The
primitive ideal space is Hausdorff in this case.
A closed subset F of a T0-space X is said to be prime if, whenever F 1 and F 2 are closed
subsets of X such that F Ď F 1 Y F 2, then F is contained in one of F 1 and F 2. The closure
of any singleton is clearly prime. A spectral space is a T0-space for which the converse holds:
each closed prime subset is the closure of a singleton. The results listed in the proposition
below can be found in [30, Sect. 5.4], or [31, Sect. 4.3]:
Proposition 2.2. Let A be a C -algebra.
(i) If A is unital, then PrimpAq is a compact1 T0-space.
(ii) Let I P PrimpAq. Then tIu is closed in PrimpAq if and only if I is a maximal proper
ideal in A, i.e., if and only if the quotient A{I is simple.
(iii) If A is separable, then PrimpAq is a second countable spectral space.
1A (possibly non-Hausdorff) topolotical space is said to be compact if it has the Heine-Borel property: each
open cover can be refined to a finite open cover. Sometimes this property is referred to as quasi-compactness.
3
By Remark 2.1, the only finite dimensional C -algebras which are primitive are those which
are isomorphic to full matrix algebras. Hence, the following holds:
Proposition 2.3. Let A be a separable C -algebra, and let I P PrimpAq be such that A{I
is finite dimensional. Then A{I – MkpCq, for some k P N, and tIu is closed in PrimpAq.
A T0-space X is said to be totally disconnected if there is a basis for its topology consisting of
compact-open sets. If the projections in a C -algebra A separate its ideals, then PrimpAq is
totally disconnected. In this situation, we have the following result, which will be discussed
in more detail in Section 4:
Theorem 2.4 (Bratteli–Elliott, [11]). Let X be a second countable, totally disconnected
spectral space. Then X is homeomorphic to PrimpAq, for some separable AF-algebra A. If
X is compact, then A can be taken to be unital.
Recall that an AF-algebra is a C -algebra which is the completion of an increasing union
of finite dimensional sub-C -algebras.
We end this section by recalling that there is a one-to-one correspondence between open
subsets U of PrimpAq and closed two-sided ideals IpUq of A, given by
IpUq " čJPPrimpAqzU
J,
with the convention that IpHq " 0 and IpPrimpAqq " A. Moreover,
PrimpA{IpUqq " PrimpAqzU ,
(2.1)
(2.2)
for each open subset U of PrimpAq. Consequently, each closed subset of PrimpAq is the
primitive ideal space of a quotient of A (see [31, Theorem 4.1.3]). Note furthermore that if
J P PrimpAq, then J " IpUq, where U is the complement of the closure of tJu.
2.2 Residually finite dimensional C -algebras
This section is devoted to discussing residually finite dimensional C -algebras and their
primitive ideal spaces. We also introduce the class of so-called strictly residually finite
dimensional C -algebras, and describe them in terms of their primitive ideal space.
A C -algebra A is said to be residually finite dimensional (RFD), if it admits a separating
family of finite dimensional representations. The finite dimensional representations can be
taken to be irreducible and pairwise (unitarily) inequivalent. (We note that two irreducible
finite dimensional representations are equivalent if and only if they are weakly equivalent,
i.e., they have the same kernel.)
Assume that tπiuiPI is a family of irreducible and pairwise inequivalent finite dimensional
representations of a C -algebra A. Let ki be the dimension of the representation πi, and
identify the image of πi with MkipCq. We then get a -homomorphism
ΨI " àiPI
πi : A Ñ źiPI
MkipCq.
Note that ΨI is injective if and only if ŞiPI Kerpπiq " t0u, which again happens if and only
if tKerpπiq : i P Iu is a dense subset of PrimpAq. Therefore, the following lemma holds; cf.
Proposition 2.3:
Lemma 2.5. A C -algebra A is RFD if and only if PrimpAq contains a dense subset P
such that A{I is a full matrix algebra, for each I P P.
4
Observe that tIu is closed in PrimpAq, for each I P P, by Proposition 2.2 (ii).
separable, then one can choose the set P in the lemma above to be countable.
If A is
Since the ideals Kerpπiq are maximal and pairwise distinct (by the assumed inequiva-
lence of the finite dimensional representations πi, which implies that they are also weakly
inequivalent), it follows from the Chinese Remainder Theorem that the map
ΨF " àiPF
πi : A Ñ źiPF
MkipCq
is surjective, for each finite subset F of I.
Definition 2.6. A unital C -algebra A is said to be strictly residually finite dimensional
(strictly RFD ) if there exists an infinite family tπi : A Ñ MkipCquiPI of irreducible, pairwise
inequivalent, finite dimensional representations of A such that the map
ΨT " àiPT
πi : A Ñ źiPT
MkipCq
(2.3)
is injective, for each infinite subset T of I.
The following characterizes strictly RFD C -algebras in terms of their primitive ideal space:
Proposition 2.7. A unital separable C -algebra A is strictly RFD if and only if there exists
an infinite subset P of PrimpAq such that each of its infinite subsets is dense in PrimpAq,
and such that A{I is finite dimensional, for each I P P.
Note that if such a subset P of PrimpAq exists, then each infinite subset of P has the same
properties, and hence one can take P to be countably infinite.
Proof. Suppose first that A is a strictly RFD unital separable C -algebra witnessed by an
infinite family tπi : A Ñ MkipCquiPI of irreducible, pairwise inequivalent, finite dimensional
representations. Set
P " tKerpπiq : i P Iu Ď PrimpAq.
If I " Kerpπiq P P, then A{I – πipAq is finite dimensional. Let T be an infinite subset of I,
then 0 " KerpΨT q " ŞiPT Kerpπiq. Therefore tKerpπiq : i P T u is dense in PrimpAq.
Suppose conversely that A is a unital separable C -algebra for which there exists an
infinite subset P " tIiuiPI of PrimpAq satisfying the hypotheses. For each i P I, find an
irreducible representation πi : A Ñ BpHiq with Kerpπiq " Ii. Then πipAq – A{Ii is finite
dimensional, so Hi is finite dimensional and πipAq " BpHiq – MkipCq, where ki " dimpHiq.
Let T be an infinite subset of I. The associated map ΨT then satisfies
KerpΨT q " čiPT
Kerpπiq " čiPT
Ii " 0,
by the assumption that tIiuiPT is dense in PrimpAq. This shows that A is strictly RFD.
3 Just-infinite C -algebras: A classification result
By analogy with the notion of just-infiniteness in the category of groups and of abstract
algebras, see [29], we define a C -algebra to be just-infinite as follows:
Definition 3.1. A C -algebra A is said to be just-infinite if it is infinite dimensional, and
for each non-zero closed two-sided ideal I in A, the quotient A{I is finite dimensional.
5
Lemma 3.2. Every just-infinite C -algebra is prime.
Proof. Let A be a just-infinite C -algebra, and let I and J be two non-zero closed two-sided
ideals in A. Consider the natural homomorphism π : A Ñ A{I ' A{J. By the assumption
that A is just-infinite, the image is finite dimensional. It follows that π cannot be injective,
so I X J " Kerpπq ‰ 0.
Example 3.3. The group Z is just-infinite, and it is the only abelian just-infinite group.
It is also known, see [22, Proposition 3(a)], that every finitely generated infinite group has
a just-infinite quotient.
The corresponding statements for C -algebras are false: No commutative C -algebra is
just-infinite, since no commuative C -algebra other than C is prime. This also shows that
no commutative C -algebra has a just-infinite quotient.
It is well-known that every unital C -algebra has a maximal proper closed two-sided
ideal, and hence a quotient which is simple. If, moreover, such a simple quotient is infinite
dimensional, then it is just-infinite. There seems to be no satisfactory description of unital
C -algebras having an infinite dimensional simple quotient.
Lemma 3.4. Each non-zero closed two-sided ideal in a just-infinite C -algebra is essential
and infinite dimensional.
Proof. It is easy to see that a C -algebra is prime if and only if each non-zero closed two-
sided ideal is essential, so the first statement of the lemma follows from Lemma 3.2.
If a closed two-sided ideal I in a C -algebra A has a unit e, then e is a central projection in
A and I " Ae. Thus Ae and Ap1 ´ eq are closed two-sided ideals in A with zero intersection.
So if I is essential, then Ap1 ´ eq " 0 and I " A. Now, as each finite dimensional C -algebra
has a unit, we see that no non-zero closed two-sided ideal in a just-infinite C -algebra can
be finite dimensional.
The class of just-infinite C -algebras does not have good permanence properties. In fact,
almost all natural operations on C -algebras (such as passing to sub-C -algebras, extensions,
passing to ideals and quotients, taking inductive limits, Morita equivalence, forming crossed
products by suitable groups) fail to be consistent with the class of just-infinite C -algebras.
However, the following permanence-type properties of just-infinite C -algebras do hold:
Proposition 3.5.
(i) If B is an infinite dimensional hereditary2 sub-C -algebra of a just-infinite C -algebra
In particular, each non-zero closed two-sided ideal in a
A, then B is just-infinite.
just-infinite C -algebra is again just-infinite.
(ii) Let 0 Ñ I Ñ A Ñ Q Ñ 0 be a short exact sequence of C -algebras, where I is an
essential ideal in A. Then A is just-infinite if and only if Q is finite dimensional and
I is just-infinite.
Proof. (i). Let J be a non-zero closed two-sided ideal in B, and let I be the (necessarily
non-zero) closed two-sided ideal in A generated by J. Then J " B X I, so B{J is isomorphic
to a (hereditary) sub-C -algebra of A{I. The latter is finite dimensional, so B{J must also
be finite dimensional.
2A sub-C -algebra B of a C -algebra A is hereditary if whenever b P B and a P A are such that 0 ď a ď b,
then a P B.
6
The second part of (i) follows from the fact that each closed two-sided ideal in a C -
algebra is a hereditary sub-C -algebra, together with Lemma 3.4, which ensures that each
non-zero closed two-sided ideal in A must be infinite dimensional.
(ii). Suppose that I is just-infinite and Q is finite dimensional. Let J be a non-zero
ideal in A. Then we have a short exact sequence 0 Ñ I{pI X Jq Ñ A{J Ñ Q{πpJq Ñ 0,
where π : A Ñ Q is the quotient mapping. Now, I X J is non-zero (since I is an essential
ideal), so I{pI X Jq is finite dimensional. This implies that A{J is finite dimensional, being
an extension of two finite dimensional C -algebras.
Conversely, if A is just-infinite, then Q, which is isomorphic to the quotient A{I, is
finite dimensional (because I is non-zero). The ideal I cannot be finite dimensional (since
otherwise A would be finite dimensional), so it follows from (i) that I is just-infinite.
The observation made above that the class of just-infinite C -algebras is not closed under
Morita equivalence, can be justified as follows. If A is a just-infinite C -algebra and if K
denotes the C -algebra of compact operators on a separable Hilbert space, then A b K is
just-infinite if and only if A is simple (since all proper non-zero quotients of Ab K are stable,
and therefore infinite dimensional).
Remark 3.6 (Hereditary just-infiniteness). If G is a residually finite group and all its normal
subgroups of finite index, including G itself, are just-infinite, then G is said to be hereditarily
just-infinite. Just-infinite branch groups are residually finite, but not hereditarily just-
infinite; cf. [22, Section 6], so a finite index normal subgroup of a just-infinite group need
not be just-infinite. We shall say more about hereditarily just-infinite groups and just-
infinite branch groups in Examples 6.12 and 6.13 and in Theorem 7.10.
It follows from Proposition 3.5 above that just-infinite C -algebras automatically have a
property analogous to being hereditarily just-infinite for groups: Any non-zero closed two-
sided ideal in a just-infinite C -algebra is itself just-infinite. Note also that the following
three conditions for a closed two-sided ideal I in a just-infinite C -algebra A are equiva-
lent (cf. Lemma 3.4 and the definition of being just-infinite): I is non-zero, I is infinite
dimensional, and I has finite co-dimension in A.
We proceed to describe the primitive ideal space of a just-infinite C -algebra. They turn
out to be homeomorphic to one of the T0-spaces in the following class:
Example 3.7. For each n P t0, 1, 2, . . . , 8u, consider the T0-space Yn defined to be the
disjoint union Yn " t0u Y Y 1
n has
countably infinitely many elements, if n " 8. Equip Yn with the topology for which the
closed subsets of Yn are precisely the following sets: H, Yn, and all finite subsets of Y 1
n.
n is a set with n elements, if n is finite, and Y 1
n, where Y 1
We shall usually take Y 1
n to be t1, 2, . . . , nu, if 1 ď n ă 8, and Y 1
8 to be N.
The spaces Yn have the following axiomatic properties:
Lemma 3.8. A (non-empty) second countable T0-space X is homeomorphic to Yn, for some
n P t0, 1, 2, . . . , 8u, if and only if it the following conditions hold, for some point x0 P X:
(A) tx0u is dense in X,
(B) txu is closed, for all x P Xztx0u,
(C) each infinite subset of X is dense in X.
Moreover, if X is any T0-space satisfying conditions (A), (B) and (C) above, then
7
(i) the closed subsets of X are the following sets: H, X, and all finite subsets of Xztx0u,
(ii) X is second countable if and only if X is countable,
(iii) each subset of X is compact (in particular, X is totally disconnected),
(iv) X is a spectral space.3
Proof. The spaces Yn satisfy conditions (A), (B), and (C) with x0 " 0. We show below that
(A), (B) and (C) imply (i)–(iv). Any second countable T0-space X satisfying (i) and (ii) is
homeomorphic to Yn, where n is the cardinality of X 1 " Xztx0u. Indeed, X is countable by
(ii), and any bijection f : X Ñ Yn, with f px0q " 0, is a homeomorphism by (i).
Let now X be a T0-space satisfying (A), (B) and (C). We show that (i), (ii), (iii) and
(iv) hold. Set X 1 " Xztx0u. It follows from (B) that each finite subset of X 1 is closed, and
so are H and X. Conversely, if F is a closed subset of X and if F ‰ X, then F must be a
finite subset of X 1 by (A) and (C). Hence (i) holds.
Suppose now that X is second countable and X ą 1. Let tUnu8
for some n, whence x P Fn. Thus X 1 is contained in the countable set Ť8
n"1 be a basis for the
topology on X consisting of non-empty open sets. For n ě 1, set Fn " XzUn, and observe
that Fn is finite (or empty) by (i). Let x P X 1. Then Xztxu is open by (i), so Un Ď Xztxu
n"1 Fn, so (ii) holds.
Let K be an arbitrary subset of X and let tWiuiPI be an open cover of K. Take any
i0 P I such that Wi0 is non-empty. Then the set F " XzWi0 is finite. Hence F X K is
covered by finitely many open sets from the collection tWiuiPI which, together with Wi0,
gives a finite open cover of K. This proves (iii).
Finally, let F ‰ H be a closed subset of X which is prime. If F " X, then F is the
closure of tx0u. If F ‰ X, then F is a finite subset of X 1, by (i). Write F " ŤxPF txu, and
note that each singleton txu, x P F , is closed. Hence F can have at most one element, so it
is in particular the closure of a singleton. This proves that X is a spectral space.
Lemma 3.9. Let A be a separable C -algebra. The following hold:
(i) PrimpAq is homeomorphic to Yn, for some n P t0, 1, 2, . . . , 8u, if and only if the
following three conditions hold:
(a) A is primitive,
(b) A{I is simple, for each non-zero primitive ideal I in A,
(c) if PrimpAq is infinite, then ŞIPP I " 0, for each infinite subset P of PrimpAq.
(ii) If PrimpAq is infinite and A satisfies (b) and (c), then it automatically satisfies (a).
If A{I is finite dimensional, for each non-zero I P PrimpAq, then condition (b) holds.
(iii) If A is just-infinite, then PrimpAq " Yn, for some n P t0, 1, 2, . . . , 8u.
Proof. (i). It follows from Proposition 2.2 (iii) that PrimpAq is second countable. It therefore
suffices to show that the conditions (a), (b) and (c) are equivalent to items (A), (B) and
(C) of Lemma 3.8 (with X " PrimpAq and x0 " 0). By definition, A is primitive if and
only if 0 P PrimpAq, so (a) is equivalent to (A). The equivalence of (b) and (B) follows from
Proposition 2.2 (ii), while the equivalence of (c) and (C) follows from the fact that a subset
P of PrimpAq is dense if and only if ŞIPP I " 0.
(ii). Suppose that PrimpAq is infinite and that (b) and (c) are satisfied. We assert that
(a) holds, as well. By [31, Proposition 4.3.6] it suffices to check that PrimpAq is prime, i.e.,
3See definition above Proposition 2.2
8
whenever PrimpAq " F Y G, where F and G are closed subsets of PrimpAq, then one of F
and G is equal to PrimpAq. However, one of F and G must be infinite, and hence dense in
PrimpAq by (c) (which is equivalent to (C)), and therefore one of F and G must be equal
to PrimpAq. The remaining assertion follows from Proposition 2.3.
(iii). Suppose that A is just-infinite. To see that PrimpAq – Yn, for some n, it suffices
by (i) and (ii) to check that (b) and (c) hold. Moreover, we conclude from (ii) that (b)
holds because A{I must be finite dimensional, for each non-zero ideal I. Suppose that P is
an infinite subset of PrimpAq, and set J " ŞIPP I. For each finite subset F of Pzt0u, let
JF " ŞIPF I Ě J. Then A{JF is isomorphic to ÀIPF A{I, which has dimension at least
F, so A{J also has dimension at least F. As this holds for all finite subsets F of P, we
conclude that A{J must be infinite dimensional. Hence J " 0, since A is just-infinite. This
proves that P is dense in PrimpAq, so (c) holds.
Just-infinite C -algebras are classified as follows (to be compared with [22, Theorem 3]):
Theorem 3.10. Let A be a separable C -algebra. Then A is just-infinite if and only if
PrimpAq is homeomorphic to Yn, for some n P t0, 1, 2, . . . , 8u, and each non-faithful irre-
ducible representation of A is finite dimensional. (If n " 0, we must also require that A is
infinite dimensional; this is automatic when n ě 1.) Moreover:
(α) PrimpAq " Y0 if and only if A is simple. Every infinite dimensional simple C -algebra
is just-infinite.
(β) PrimpAq " Yn, for some integer n ě 1, and A is just-infinite, if and only if A con-
tains a simple non-zero essential infinite dimensional ideal I0 such that A{I0 is finite
dimensional. In this case, n is equal to the number of simple summands of A{I0.
(γ) The following conditions are equivalent:
(i) A is just-infinite and PrimpAq " Y8,
(ii) A is just-infinite and RFD,
(iii) PrimpAq is an infinite set, all of its infinite subsets are dense, and A{I is finite
dimensional, for each non-zero I P PrimpAq,
(iv) PrimpAq is an infinite set, the direct sum representation ÀiPT πi is faithful for
each infinite family tπiuiPT of pairwise inequivalent irreducible representations of
A, and each non-faithful irreducible representation of A is finite dimensional.
We shall occasionally refer to a just-infinite C -algebra as being of type (α), (β) and (γ),
respectively, if it satisfies the corresponding condition in the theorem above. In view of the
theorem, we shall also, more frequently, refer to a just-infinite C -algebra of type (γ) as a
RFD just-infinite C -algebra.
Proof. If A is just-infinite and separable, then PrimpAq " Yn, for some n P t0, 1, 2 . . . , 8u,
by Lemma 3.9, and each non-faithful irreducible representation of A is finite dimensional
(by the definition of being just-infinite).
Suppose conversely that PrimpAq " Yn, for some n P t0, 1, 2, . . . , 8u, and that each non-
faithful irreducible representation of A is finite dimensional. We show that A then must be
just-infinite (if it is also infinite dimensional). This is clear if n " 0, since A is simple in
this case. This also shows that (α) holds.
9
Suppose that 1 ď n ď 8. Since Yn is non-Hausdorff, when n ą 0, and the primitive
ideal space of any finite dimensional C -algebra is Hausdorff; cf. Remark 2.1, A must be
infinite dimensional. Write
PrimpAq " t0u Y tIjun
j"1.
(3.1)
Any non-zero proper ideal J of A is the intersection of the primitive ideals in A that contain
it. By Lemma 3.9 (c), any intersection of infinitely many distinct primitive ideals of A is
zero. Hence J " ŞjPF Ij, for some, necessarily finite, subset F of N (or of t1, 2, . . . , nu, if
n ă 8). Now, A{J is isomorphic to ÀjPF A{Ij, and each quotient A{Ij is finite dimensional
by assumption, whence A{J is finite dimensional. This shows that A is just-infinite.
We proceed to verify the claims in (β) and (γ).
(β). The "if" part follows from Proposition 3.5 (ii). Moreover, PrimpAq consists of 0 (cf.
Lemma 3.2) and the kernels of the maps onto the n simple summands of A{I0, so PrimpAq
has cardinality n ` 1. Also, PrimpAq is homeomorphic to Yk, for some k, by Lemma 3.9
(iii), and by cardinality considerations, we conclude that k " n.
some n P N. Retain the notation set forth in (3.1), and let I0 " Şn
Let us prove the "only if" part. Suppose that A is just-infinite and PrimpAq " Yn, for
j"1 Ij. In the notation
from (2.1), we have I0 " Ipt0uq (observe that t0u is an open subset of PrimpAq, when
n ă 8). We deduce that I0 is non-zero and simple. Each non-zero ideal in a primitive C -
algebra is essential, so I0 is an essential ideal in A, by Lemma 3.2. Since A is just-infinite,
A{I0 is finite dimensional. Finally, by (2.2),
PrimpA{I0q " PrimpAqzt0u " tI1, I2, . . . , Inu,
and since A{I0 is finite dimensional, n is the number of direct summands of A{I0; cf.
Remark 2.1.
(γ). (i) ñ (iii). If A is just-infinite, then A{I is finite dimensional, for each non-zero ideal
I in A; and if PrimpAq " Y8, then each infinite subset of PrimpAq is dense (by Lemma 3.9
(i)(c)), and PrimpAq is an infinite set.
(iii) ñ (ii). The assumptions in (iii) imply that A is RFD; cf. Lemma 2.5. If π is a
non-faithful irreducible representation of A, then Kerpπq " I is a non-zero primitive ideal
in A, so πpAq – A{I is finite dimensional. To conclude that A is just-infinite we show
that PrimpAq is homeomorphic to Y8. For this it suffices to verify conditions (b) and
(c) of Lemma 3.9 (i). Item (b) holds because A{I is finite dimensional, for each non-zero
I P PrimpAq; cf. Lemma 3.9 (ii).
Item (c) is equivalent to condition (C) in Lemma 3.8,
which holds by assumption.
(ii) ñ (i). If A is RFD, then A cannot be just-infinite of type (α) or (β), so PrimpAq
must be homeomorphic to Y8.
(iii) ñ (iv). We already saw that (iii) implies that A is just-infinite, and hence that
each non-faithful irreducible representation is finite dimensional. Let tπiuiPT be an infinite
family of pairwise inequivalent irreducible representations of A. Since tKerpπiq : i P T u is an
infinite set, and hence by assumption a dense subset of PrimpAq, it follows that the kernel
of ÀiPT πi, which is equal to ŞiPT Kerpπiq, must be zero.
(iv) ñ (iii). Let P be an infinite subset of PrimpAq, and choose pairwise inequivalent
irreducible representations tπiuiPT of A such that P " tKerpπiq : i P T u. The assumptions
in (iv) now yield
0 " Ker´àiPT
πi¯ " čiPT
Kerpπiq " čIPP
I,
which implies that P is dense in PrimpAq.
10
If I is a non-zero primitive ideal in A, then I " Kerpπq, for some (non-faithful) irreducible
representation of A, so A{I – πpAq is finite dimensional.
The following result follows immediately from Proposition 2.7 and Theorem 3.10:
Corollary 3.11. Each separable RFD just-infinite C -algebra is strictly RFD.
We note that not all strictly RFD C -algebras are just-infinite; cf. Section 4.3 below.
Corollary 3.12. The primitive ideal space of a separable just-infinite C -algebra is count-
able. Moreover, any RFD just-infinite separable C -algebra has countably infinitely many
equivalence classes of finite dimensional irreducible representations.
Proof. The first claim follows from Lemma 3.9 (iii). The second claim follows from Theo-
rem 3.10 (γ), by the fact that there is a one-to-one correspondence between weak equivalence
classes of irreducible representations and the primitive ideal space of a separable C -algebra
(given by mapping an irreducible representation to its kernel), and by the fact, observed
earlier, that two finite dimensional irreducible representations are unitarily equivalent if
they are weakly equivalent.
Remark 3.13. It is shown in Theorem 3.10 that a separable C -algebra A is just-infinite
if and only if the following two conditions hold: PrimpAq " Yn, for some n P t0, 1, 2, . . . , 8u
and each non-faithful irreducible representation is finite dimensional. These two conditions
are independent, i.e., none of them alone implies that A is just-infinite, as shown below.
If X is a Hausdorff space and k is a positive integer, then all irreducible representations
of MkpCpXqq have dimension k, and PrimpMkpCpXqqq " X. If X is not a point, then X is
not homeomorphic to Yn, for any n, because Yn is non-Hausdorff for all n ą 0. Therefore
MkpCpXqq is not just-infinite.
For each n P t0, 1, 2, . . . , 8u, there is a unital AF-algebra whose primitive ideal space is
Yn, by Theorem 2.4 and Lemma 3.8. The AF-algebras obtained in this way may or may
not have the property that each non-faithful irreducible representation is finite dimensional.
Tensoring such an AF-algebra by a UHF-algebra, we obtain a unital separable C -algebra
whose primitive ideal space is Yn, and which has no finite dimensional irreducible represen-
tations. Therefore, it is not just-infinite.
We show in the next example and in Section 4 below that each space Yn can be realized
as the primitive ideal space of a just-infinite AF-algebra.
Example 3.14 (Existence of just-infinite C -algebras). Any simple infinite dimensional
C -algebra is just-infinite of type (α) (and there are many examples of such, both in the
unital and the non-unital case).
To exhibit examples of just-infinite C -algebras of type (β), let n P N, and let F be
a finite dimensional C -algebra with n simple summands, e.g., F " C ' ' C with n
summands. Let H be an infinite dimensional separable Hilbert space, and let π : BpHq Ñ
BpHq{K be the quotient mapping onto the Calkin algebra, where as before K denotes the
compact operators on H. Let τ : F Ñ BpHq{K be a unital injective -homomorphism. Set
A " π´1`τ pF q Ď BpHq.
(3.2)
Then K is a simple essential ideal in A and A{K is isomorphic to F . Hence A is just-infinite
of type (β), and PrimpAq " Yn; cf. Theorem 3.10 (β). Since A is an extension of two
AF-algebras, it is itself an AF-algebra.
11
Each just-infinite C -algebra A arising as in (3.2) above is of type I : for each irreducible
representation of A on a Hilbert space H, the image of A contains the compact operators on
H. Conversely, a separable C -algebra A of type I is just-infinite if and only if it is isomorphic
to the compact operators K on a separable Hilbert space, or it is of the form described in
(3.2) above for some finite dimensional C -algebra F . Indeed, if A is separable, just-infinite
and of type I, then A is prime by Lemma 3.2, hence primitive (because it is separable), so
it admits a faithful irreducible representation ρ on some (separable) Hilbert space. Being of
type I, ρpAq contains the compact operators K. If ρpAq ‰ K, then the quotient B :" ρpAq{K
is finite dimensional, because A is just-infinite, so A – ρpAq " π´1pBq is as in (3.2).
It requires more work to establish the existence of RFD just-infinite C -algebras, i.e., those
of type (γ). This will be done in Section 4.
Remark 3.15 (Characteristic sequences of just-infinite C -algebras). Let A be a unital
separable just-infinite C -algebra. If A is non-simple, then PrimpAq " Yn, for some n P
t1, 2, . . . , 8u. Let tIjun
j"1 be the non-zero primitive ideals of A. Then A{Ij – Mkj pCq,
for some kj P N; cf. Proposition 2.3. The resulting n-tuple, or sequence, tkj un
j"1 (as an
unordered set) is an invariant of A, which we shall call the characteristic sequence of A.
that such a sequence tπjun
irreducible representations of A. Equivalently, tπjun
For each j, choose an irreducible representation πj : A Ñ Mkj pCq with kernel Ij. We say
j"1 is an exhausting sequence of pairwise inequivalent non-faithful
j"1 is an exhausting sequence of pairwise
inequivalent non-faithful irreducible representations of A if PrimpAqzt0u " Kerpπjq : j "
1, 2, . . . , n(, and Kerpπjq ‰ Kerpπiq when i ‰ j.
If n P N and if I0 is the (unique) simple essential ideal in A, then (as in the proof of
Theorem 3.10) we have the following isomorphisms
A{I0 –
n
àj"1
A{Ij –
n
àj"1
Mkj pCq.
(3.3)
It follows from Example 3.14 (and Remark 2.1) that for all positive integers k1, k2, . . . , kn,
there exists a just-infinite C -algebra A, which is necessarily an AF-algebra, such that
(3.3) holds with I0 " K. This argument shows in particular that each finite characteristic
sequence tkjun
j"1, where n P N, is realized by a just-infinite AF-algebra (of type (β)).
We end this section by showing that the characteristic sequence tkju8
j"1 of a RFD just-
infinite C -algebra must tend to infinity. The proof of this fact involves results about
subhomogeneous C -algebras. Recall that a C -algebra is said to be subhomogeneous if it
is isomorphic to a sub-C -algebra of MkpCpXqq, for some compact Hausdorff space X, and
some k P N. The next proposition is well-known, but we include a brief proof for the sake
of completeness.
Proposition 3.16. For a C -algebra A, the following conditions are equivalent:
(i) A is subhomogeneous,
(ii) the bidual A of A is isomorphic to Àn
j"1 Mkj pCpΩjqq, for some positive integers
n, k1, k2, . . . , kn, and some (extremally disconnected) compact Hausdorff spaces Ω1,
Ω2, . . . , Ωn,
(iii) there exists a positive integer k such that each irreducible representation of A has
dimension at most k,
12
(iv) there exist a positive integer k and a separating family tπiuiPT of irreducible represen-
tations of A such that each πi has dimension at most k.
Proof. The implication (ii) ñ (i) holds because A is a sub-C -algebra of A. If (iii) holds,
then A, which is a von Neumann algebra, cannot have central summands of type In, for
n ą k, or of type II or III. Therefore (ii) holds. The implication (i) ñ (iv) follows easily
from the definition of subhomogeneity. Suppose now that (iv) holds, and that there exists
an irreducible representation of A of dimension strictly greater than k (possibly infinite
dimensional). By (a version of) Glimm's lemma, see, e.g., [33, Proposition 3.10], there is a
non-zero -homomorphism ρ : C0pp0, 1sq b Mk`1 Ñ A. However, there is no non-zero -ho-
momorphism C0pp0, 1sq b Mk`1 Ñ BpHq when dimpHq ď k, so it follows that πi ρ " 0, for
each i P T . As the family tπiuiPT is separating, we conclude that ρ " 0, a contradiction.
Lemma 3.17. No separable subhomogeneous C -algebra is just-infinite.
Proof. Let A be a separable just-infinite C -algebra. Then A is prime; cf. Lemma 3.2,
hence primitive, and so A admits a faithful irreducible representation. Such a representa-
tion cannot be finite dimensional, because A is infinite dimensional. Hence A cannot be
subhomogeneous; cf. Proposition 3.16.
Proposition 3.18. Let A be a separable RFD just-infinite C -algebra with characteristic
sequence tkju8
j"1. Then limjÑ8 kj " 8.
Proof. Let I1, I2, . . . be the non-zero primitive ideals of A, and for each j, let πj be an
irreducible representation of A whose kernel is Ij, such that the dimension of πj is kj. We
must show that for each k, Tk :" tj P N : kj ď ku is finite. Suppose that the set T " Tk is
infinite. Then the -homomorphism ΨT " ÀjPT πj is injective, which implies that tπjujPT
is a separating family of irreducible representations of A, each of which having dimension
less than or equal to k. Then Proposition 3.16 implies that A is subhomogeneous, but this
is impossible by Lemma 3.17.
4 Examples of RFD just-infinite AF-algebras
We construct an example of a RFD just-infinite AF-algebra. By Theorem 3.10, its primitive
ideal space must be Y8. The existence of a unital AF-algebra whose primitive ideal space
is homeomorphic to Y8 follows from Theorem 2.4 (Bratteli–Elliott). To conclude that
such an AF-algebra is just-infinite, we must also ensure that its non-faithful irreducible
representations are finite dimensional; cf. Theorem 3.10. This is accomplished by taking a
closer look at the construction by Bratteli and Elliott, done in Proposition 4.1 below.
4.1 Construction of a RFD just-infinite AF-algebra
Recall that a Bratteli diagram is a graph pV, Eq, where V " Ť8
n"1 En
(disjoint unions), all Vn and all En are finite sets, and where each edge e P En connects
a vertex v P Vn to a vertex in w P Vn`1. In this case, we write speq " v and rpeq " w,
thus giving rise to the source and the range maps s, r : E Ñ V . It was shown by Bratteli,
[10], that there is a bijective correspondence between Bratteli diagrams (modulo a natural
equivalence class of these) and AF-algebras (modulo Morita equivalence).
n"1 Vn and E " Ť8
An ideal in a Bratteli diagram pV, Eq is a subset U Ď V with the following properties:
• for all e in E, if speq belongs to U , then so does rpeq,
13
• for all v in V , if trpeq e P s´1pvqu is contained in U , then v belongs to U .
The ideal lattice of an AF-algebra associated with a given Bratteli diagram is isomorphic
to the ideal lattice of the Bratteli diagram, see [17] or [16]. The following proposition is
contained in [11]:
Proposition 4.1 (Bratteli–Elliott). Let X be a second countable, compact, totally discon-
nected T0-space. Let G1, G2, . . . be finite families of compact-open subsets of X such that:
(i) X " ŤGPGn
G, for each n ě 1,
(ii) for each n ě 1, Gn`1 is a refinement of Gn, i.e., each set in Gn`1 is contained in a set
in Gn, and each set in Gn is the union of sets from Gn`1,
8
(iii)
ďn"1
Gn is a basis for the topology on X.
Consider the Bratteli diagram for which the vertices at level n are the sets in Gn, and where
there is one edge from G P Gn to G1 P Gn`1 if G1 Ď G, and none otherwise. Then there is
a one-to-one correspondence between open subsets of X and ideals of the Bratteli diagram,
given as follows: the ideal in the Bratteli diagram associated with an open subset U of X
consists of all vertices G P Ť8
n"1 Gn for which G Ď U .
If, in addition, X is a spectral space, and if A is an AF-algebra associated with the
Bratteli diagram constructed above, then PrimpAq is homeomorphic to X.
In the following, we construct a sequence G1, G2, G3, . . . of finite families of compact-open
subsets of X satisfying the conditions of Proposition 4.1 in the case where X " Y8. Recall
that Y8 " t0u Y N, that the open subsets of Y8 are H, Y8, and all co-finite subsets of N,
and that every subset of Y8 is compact. For all n ě 1, set
Fn,k " t1, 2, . . . , nuztku,
Gn,k " Y8zFn,k,
for 1 ď k ď n, and let
Gn " tGn,1, Gn,2, . . . , Gn,nu.
Observe that each Gn,k is open and (automatically) compact. Moreover, the sets tGnu8
satisfy conditions (i), (ii) and (iii) in Proposition 4.1. Furthermore, for 1 ď k ď n,
n"1
Gn`1,k Ď Gn,k,
Gn`1,n`1 Ď Gn,k.
No other inclusions between sets in Gn`1 and sets in Gn hold. Therefore, the Bratteli diagram
associated with this sequence of compact-open subsets of Y8 as in Proposition 4.1 is:
‚
‚
‚
‚
‚
...
‚
❆❆❆❆❆❆❆
PPPPPPPPPPPPPP
❆❆❆❆❆❆❆
❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯
PPPPPPPPPPPPPP
❆❆❆❆❆❆❆
❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲
❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯
PPPPPPPPPPPPPP
❆❆❆❆❆❆❆
‚
‚
‚
‚
‚
...
‚
‚
...
‚
...
C
C ' C
C ' C ' M2pCq
C ' C ' M2pCq ' M4pCq
(4.1)
C ' C ' M2pCq ' M4pCq ' M8pCq
‚
...
14
The sequence of finite dimensional C -algebras on the right-hand side, equipped with unital
connecting mappings given by the Bratteli diagram, defines a unital AF-algebra A, associ-
ated with the Bratteli diagram. The one-to-one correspondence between (non-empty) open
subsets G Ď Y8 " t0u Y N and ideals U pGq of the Bratteli diagram above is given as follows:
U pGq " tGn,k Gn,k Ď Gu " tGn,k k P G, n ě max Y8zGu.
E.g., U pY8zt1, 3uq " tGn,k n ě 3, k ‰ 1, 3u and U pY8ztjuq " tGn,k n ě j, k ‰ ju, j ě 1.
The quotient of the AF-algebra A by the ideal in A corresponding to U pGq is given by
the Bratteli diagram that arises by removing U pGq from the original diagram. The two
pictures below show the ideal U pGq (in blue) and the Bratteli diagram of the quotient (in
red) in the cases where G " Y8zt2u, respectively, G " Y8zt1, 3u:
‚
‚
‚
‚
‚
...
‚
❆❆❆❆❆❆❆
PPPPPPPPPPPPPP
❆❆❆❆❆❆❆
❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯
PPPPPPPPPPPPPP
❆❆❆❆❆❆❆
❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲
❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯
PPPPPPPPPPPPPP
❆❆❆❆❆❆❆
‚
‚
‚
‚
‚
‚
...
‚
...
‚
...
‚
‚
‚
‚
‚
...
‚
...
‚
❆❆❆❆❆❆❆
PPPPPPPPPPPPPP
❆❆❆❆❆❆❆
❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯
PPPPPPPPPPPPPP
❆❆❆❆❆❆❆
❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲
❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯
PPPPPPPPPPPPPP
❆❆❆❆❆❆❆
‚
‚
‚
‚
‚
‚
...
‚
...
‚
...
‚
...
The quotient of A by the ideal in A corresponding to U pGq, with G " Y8zt2u, is the AF-
algebra associated to the red part of the Bratteli diagram, which is C. The quotient of A
in the case where G " Y8zt1, 3u is similarly seen to be C ' M2pCq.
By construction, and by Proposition 4.1, we have PrimpAq – Y8. In more detail, we
have PrimpAq " t0u Y tI1, I2, I3, . . . u, where Ij is the primitive ideal in A corresponding to
the ideal U pY8ztjuq of the Bratteli diagram. Arguing as in the two examples above, we see
that A{Ij – MkpjqpCq, where kp1q " kp2q " 1 and kpjq " 2j´1, for j ě 2. Hence A{I is finite
dimensional, for each non-zero primitive ideal I of A. It now follows from Theorem 3.10
that A is just-infinite and RFD, as desired. The characteristic sequence of A is precisely
the sequence tkpjqu8
j"1 defined above.
One can modify the Bratteli diagram in various ways to construct new RFD just-infinite
AF-algebras with other characteristic sequences. For example, one can delete the first
n ´ 1 rows and let row n correspond to an arbitrary finite dimensional C -algebra with
n summands. (The remaining finite dimensional C -algebras are then determined by the
one chosen and by the Bratteli diagram.) One is also allowed to change the multiplicity
of the edges connecting the vertex at position pn, kq, 1 ď k ď n, to the vertex at position
pn ` 1, n ` 1q. In these examples, the characteristic sequences all grow exponentially. By
Proposition 3.18, we know that they must tend to infinity. This leaves open the following:
Question 4.2. What are the possible characteristic sequences tkj u8
C -algebras?
j"1 of RFD just-infinite
15
4.2 The dimension group
We compute the dimension group pK0pAq, K0pAq`, r1sq of the just-infinite AF-algebra A
constructed above (associated with the Bratteli diagram (4.1)).
Recall that the dimension group, pH, H `, vq, associated with the Bratteli diagram (4.1)
is the inductive limit of the ordered abelian groups
αnpx1, x2, . . . , xnq " px1, x2, . . . , xn, x1 ` x2 ` ` xnq,
px1, . . . , xnq P Zn,
Z
α1
/ Z2 α2
/ Z3 α3
/ ,
where v P H ` is the image of 1 in the first copy of Z. It follows from standard theory of
AF-algebras that pK0pAq, K0pAq`, r1sq is isomorphic to pH, H `, vq. We proceed to identify
the latter more explicitly.
Let śjPN
Z denote the (uncountable) group of all sequences x " txju8
j"1 of integers,
equipped with the usual order: x ě 0 if and only if xj ě 0, for all j ě 1. Let G be the
countable subgroup of śjPN
j"1 for which the identity
xj`1 " x1 `x2 ` `xj holds eventually, and equip G with the order inherited from śjPN
Z.
Set u " p1, 1, 2, 4, 8, q. We show below that pH, H `, vq – pG, G`, uq. In conclusion,
Z consisting of those sequences txju8
pK0pAq, K0pAq`, r1sq – pH, H `, vq – pG, G`, uq.
For this, define ρn : Zn Ñ G by
ρnpx1, x2, . . . , xnq " px1, x2, . . . , xn, xn`1, xn`2, . . . q,
where xj`1 " x1 ` x2 ` ` xj, for all j ě n. Then ρn`1 αn " ρn, for all n, and each ρn
is positive. It follows that the ρn's extend to a positive group homomorphism ρ : H Ñ G.
Each ρn is injective, so ρ is injective.
To complete the proof that ρ is an order isomorphism, we show that ρpH `q " G`. Take
j"1 P G`, and let n ě 1 be such that xj`1 " x1 ` x2 ` ` xj, for all j ě n. Then
x " txju8
x " ρnpx1, x2, . . . , xnq
" x1ρnpepnq
" x1ρpf pnq
1
2 q ` ` xnρnpepnq
n q
q ` ` xnρpf pnq
n q,
1 q ` x2ρnpepnq
q ` x2ρpf pnq
2
1 , epnq
2 , . . . , epnq
where epnq
are the
corresponding images in H ` Ď H. This shows that x P ρpH `q. Finally, ρpvq " ρ1p1q " u,
as wanted.
n is the standard basis for pZnq` Ď Zn, and f pnq
, . . . , f pnq
, f pnq
n
1
2
Unital AF-algebras are completely classified by their ordered K0-group, together with the
position of the class of the unit. It is therefore an interesting question to classify, or charac-
terize, those dimension groups which are the K0-group of a RFD just-infinite AF-algebra.
In the light of the computation above, one may first wish to consider those dimension
Z. In addition, one should assume that G
Z, for each finite subset
Z. The dimension group
groups G which are (ordered) subgroups of ś8
is a subdirect product of ś8
F of N, where ϕF is the canonical projection of ś8
Z, in the sense that ϕF pGq " śjPF
Z onto śjPF
considered above has this property.
j"1
j"1
j"1
16
/
/
/
4.3 A strictly RFD C -algebra which is not just-infinite
It was shown in Corollary 3.11 that all RFD just-infinite C -algebras are strictly RFD.
We show here that the converse does not hold, by constructing an example of a unital
AF-algebra which is strictly RFD and not just-infinite.
Let us first describe the example at the level of its primitive ideal space. Let X be the
disjoint union of two copies of Y8, i.e., X " X1 > X2, where X1 " X2 " Y8. Equip X
with the following topology: A non-empty subset U of X is open if and only if U X X1 is
non-empty and open, and U X X2 is open. That this indeed defines a topology on X follows
from the fact that the intersection of any two non-empty open subsets of X1 is non-empty,
or, equivalently, that the set X1 is prime.
Observe that X2 is an infinite closed subset of X. Hence X2 is a non-dense infinite subset
of X. This shows that X cannot be the primitive ideal space of a just-infinite C -algebra;
cf. Lemma 3.9. The set X1, on the other hand, is an open and dense subset of X, and each
infinite subset of X1 is dense in X1, and therefore also dense in X.
The space X is the primitive ideal space of the unital AF-algebra B whose Bratteli
diagram is given as follows (ignoring at first the coloring of the vertices and edges):
‚
‚
‚
‚
‚
...
‚
❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜❜
❆❆❆❆❆❆❆
⑥⑥⑥⑥⑥⑥⑥
❝❝❝❝❝❝❝❝❝❝❝❝❝❝❝❝❝❝❝❝❝❝❝❝❝❝❝❝❝❝❝❝❝❝❝❝❝❝❝❝❝❝❝❝❝❝❝
PPPPPPPPPPPPPP
♥♥♥♥♥♥♥♥♥♥♥♥♥♥
❆❆❆❆❆❆❆
⑥⑥⑥⑥⑥⑥⑥
❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯
❡❡❡❡❡❡❡❡❡❡❡❡❡❡❡❡❡❡❡❡❡❡❡❡❡❡❡❡❡❡❡❡❡
PPPPPPPPPPPPPP
✐✐✐✐✐✐✐✐✐✐✐✐✐✐✐✐✐✐✐✐
♥♥♥♥♥♥♥♥♥♥♥♥♥♥
❆❆❆❆❆❆❆
⑥⑥⑥⑥⑥⑥⑥
❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲
❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯
❣❣❣❣❣❣❣❣❣❣❣❣❣❣❣❣❣❣❣❣❣❣❣❣❣❣❣
PPPPPPPPPPPPPP
✐✐✐✐✐✐✐✐✐✐✐✐✐✐✐✐✐✐✐✐
❥❥❥❥❥❥❥❥❥❥❥❥❥❥❥❥❥❥❥
♥♥♥♥♥♥♥♥♥♥♥♥♥♥
❆❆❆❆❆❆❆
⑥⑥⑥⑥⑥⑥⑥
‚
‚
‚
‚
‚
‚
‚
...
‚
‚
...
‚
‚
...
‚
‚
...
‚
‚
...
‚
‚
...
‚
...
‚
...
‚
‚
‚
‚
‚
...
The left-hand half of this Bratteli diagram is an essential ideal in the Bratteli diagram4
and therefore corresponds to an essential ideal I of the AF-algebra B. The right-hand half
is the Bratteli diagram of the quotient B{I. Hence B{I is equal to the RFD just-infinite
AF-algebra A described in Section 4.1, and I is Morita equivalent to A. Hence B cannot
be just-infinite. For each k ě 1, let Uk be largest ideal of the Bratteli diagram which does
not contain any vertex from the kth column of the left-hand half of the Bratteli diagram.
Furthermore, let Ik be the ideal of B corresponding to the ideal Uk.
To illustrate this definition, in the diagram above, the ideal U3 is marked in blue and
the Bratteli diagram of the quotient B{I3 is marked in red The quotient B{I3 is seen to be
isomorphic to M4pCq.
In general, for each k ě 1, we see that B{Ik is a full matrix algebra, and (hence) that
this, observe that Uk contains no vertices from the top k ´ 1 rows of the left-hand half of the
each Ik is a primitive ideal. Moreover, ŞkPT Ik " 0, for each infinite subset T of N. (To see
Bratteli diagram, or from the top k ´ 2 rows of the right-hand half. Hence ŞkPT Uk " H,
for each infinite subset T of N.)
This shows that B is a strictly RFD AF-algebra which is not just-infinite.
4An ideal U in a Bratteli diagram is said to be essential, if U X V ‰ H, for all non-empty ideals V .
17
5 Subalgebras and superalgebras
In this section, which is addressed to specialists in C -algebras, we investigate when subalge-
bras and superalgebras of just-infinite C -algebras are again just-infinite, and we also show
that not all RFD just-infinite C -algebras are nuclear, or even exact. The third named
author thanks Jose Carrion for his suggestion to use Theorem 5.3 below of Dadarlat to
conclude that there are non-nuclear, and even non-exact, RFD just-infinite C -algebras.
Recall that a C -algebra A has real rank zero if each self-adjoint element in A is the norm
limit of self-adjoint elements in A with finite spectra. A commutative C -algebra CpXq has
real rank zero if and only if X is totally disconnected (or, equivalently, dimpXq " 0). Real
rank zero is therefore viewed as a non-commutative analog of being zero-dimensional. A
C -algebra has real rank zero if it has "sufficiently many projections". Each closed two-
sided ideal of a C -algebra of real rank zero again has real rank zero and, as a consequence,
is generated by its projections.
We denote by IdealpAq the lattice of closed two-sided ideals in A. If B is a sub-C -alge-
bra of A, then there is a natural map Φ : IdealpAq Ñ IdealpBq, given by ΦpIq " I X B. The
map Φ is, in general, neither injective nor surjective, but it is both in the special situation
of the lemma below. We use the symbol p „A q to denote that p and q are Murray-von
Neumann equivalent projections, relatively to the C -algebra A.
Lemma 5.1. Let B Ď A be unital C -algebras of real rank zero, and suppose that there is a
-homomorphism κ : A Ñ B such that κppq „A p, for all projections p P A, and κpqq „B q,
for all projections q P B. Then the map Φ : IdealpAq Ñ IdealpBq is a lattice isomorphism.
Proof. We first show that Φ is injective. Let I ‰ I 1 P IdealpAq be given. Since A has
real rank zero, and ideals in A are generated by their projections, there exists a projection
p P I such that p R I 1 (or vice versa). Set q " κppq „ p. Then q P I X B " ΦpIq, but
q R I 1 X B " ΦpI 1q. Hence ΦpIq ‰ ΦpI 1q.
Let now J P IdealpBq be given, and let I " AJ A be the closed two-sided ideal in A
generated by J. Then, clearly, J Ď I X B " ΦpIq. To see that ΦpIq Ď J, it suffices to
show that each projection q in ΦpIq belongs to J. Being a projection in I, q belongs to the
j"1 aj xjbj for some aj, bj P A and
xj P J. The conditions on κ, together with the fact that B is a C -algebra of real rank zero,
imply that κ maps J into itself, so
algebraic two-sided ideal in A generated by J, so q " řn
q „B κpqq "
n
ÿj"1
κpajqκpxjqκpbj q P J.
This shows that q belongs to J, as desired.
Lemma 5.2. Let A be a unital separable RFD just-infinite C -algebra of real rank zero,
and let tπnu8
n"1 be an exhausting5 sequence of pairwise inequivalent non-faithful irreducible
representations of A.
(i) Suppose that B is a unital sub-C -algebra of A such that the map Φ : IdealpAq Ñ
IdealpBq is an isomorphism, and such that each projection in A is equivalent to a
n"1 is an
projection in B.
exhausting sequence of pairwise inequivalent non-faithful irreducible representations
of B, and that πnpBq " πnpAq, for all n.
In particular, A and B have the same
characteristic sequence.
It follows that B is just-infinite and RFD, that tπnBu8
5See Remark 3.15.
18
(ii) Suppose that C is a unital C -algebra of real rank zero which contains A and is asymp-
totically homotopy equivalent6 to A. Suppose also that Φ : IdealpCq Ñ IdealpAq is
an isomorphism.7 It follows that C is just-infinite and RFD with an exhausting se-
quence tνnu8
n"1 of pairwise inequivalent non-faithful irreducible representations for
which KerpνnAq " Kerpπnq and νnpCq – πnpAq, for all n. In particular, A and C
have the same characteristic sequence.
Proof. (i). The lattice isomorphism Φ : IdealpAq Ñ IdealpBq restricts to a homeomorphism
PrimpAq Ñ PrimpBq, and so PrimpBq is homeomorphic to PrimpAq, which again is homeo-
morphic to Y8. Moreover,
PrimpBqzt0u " tKerpπnq X B n P Nu " tKerpπnBq n P Nu.
Let I be a non-zero primitive ideal of B. Then I " KerpπnBq, for some n, and B{I is
isomorphic to πnpBq, which is a subalgebra of the finite dimensional C -algebra πnpAq, so
B{I is finite dimensional. It now follows from Theorem 3.10 that B is just-infinite.
Let us also show that πnpBq " πnpAq, for all n. Since πnpBq Ď πnpAq and both C -al-
gebras are full matrix algebras, it suffices to show that πnpBq contains a minimal projection
in πnpAq. Let e P πnpAq be such a projection and lift it to a projection p P A (which is
possible because A is assumed to have real rank zero). Find a projection q P B which is
equivalent to p. Then πnpqq is equivalent to e, which implies that πnpqq itself is a minimal
projection in πnpAq.
(ii). As in (i), the given lattice isomorphism Φ : IdealpCq Ñ IdealpAq restricts to a
homeomorphism PrimpCq Ñ PrimpAq, so PrimpCq is homeomorphic to Y8.
Given n ě 1, let Jn P PrimpCq be such that ΦpJnq " Kerpπnq. Since Φ is an isomorphism,
each non-zero primitive ideal in C is of this form. Identify πnpAq with MkpCq, for some
positive integer k. Find an irreducible representation νn : C Ñ BpHq on some Hilbert space
H, with Kerpνnq " Jn. Then Kerpπnq " ΦpJnq " Jn X A " KerpνnAq. Let ι : MkpCq Ñ
BpHq be the inclusion mapping making the following diagram commutative:
A
πn
C
νn
MkpCq
ι
/ BpHq
We show that dimpHq " k, which by Theorem 3.10, will imply that C is just-infinite. It
will also imply that ι is an isomorphism, and that πnpAq – νnpAq " νnpCq " BpHq.
It is clear that dimpHq ě k. Suppose that dimpHq ą k. Then we can find pairwise
orthogonal non-zero projections f1, f2, . . . , fk`1 in νnpCq. (Indeed, νnpCq acts irreducibly
on H, so if dimpHq is finite, then νnpCq " BpHq.
If dimpHq is infinite, then νnpCq is
infinite dimensional and of real rank zero.
In either case, one can find the desired pro-
jections.) Since C has real rank zero, we can lift the projections f1, f2, . . . , fk`1 to mutu-
ally orthogonal projections p1, p2, . . . , pk`1 in C. Applying the asymptotic homomorphism
C Ñ A to the projections p1, p2, . . . , pk`1, and using that Ck`1 is semiprojective (see [8]),
we obtain mutually orthogonal projections q1, q2, . . . , qk`1 in A. Since the asymptotic ho-
momorphism C Ñ A composed with the inclusion mapping A Ñ C is homotopic to the
6This means that there exists an asymptotic morphism C Ñ A, so that the asympotic morphism C Ñ C
(obtained by composing it with the inclusion mapping A Ñ C) is homotopic to the identity on C in the
category of asymptotic morphism. See also [14].
7In fact, the assumptions on A and C imply that Φ is an isomorphism. This can be shown along the same
lines as the proof of Lemma 5.1.
19
/
/
/
identity mapping on C, we further get that qj is equivalent to pj, for each j. In particular,
pι πnqpqjq " νnpqjq „ νnppjq " fj, for each j, so πnpqjq is non-zero. But MkpCq does not
contain k ` 1 mutually orthogonal non-zero projections. This proves that dimpHq " k.
We shall combine Lemma 5.2 with the following results due to Dadarlat:
Theorem 5.3 (Dadarlat, [14, Theorem 11 and Proposition 9]). Let A be a unital AF-algebra
not of type I. Then:
(i) A contains a unital non-nuclear sub-C -algebra B of real rank zero and stable rank
one, for which there exists a unital -monomorphism κ : A Ñ B such that ι κ is
homotopic to idA and κ ι is asymptotically homotopic to idB, where ι is the inclusion
mapping B Ñ A. Moreover, Φ : IdealpAq Ñ IdealpBq is an isomorphism of lattices.
(ii) A is contained in a unital separable non-exact C -algebra C of real rank zero and
stable rank one, which is asymptotically homotopy equivalent to A, and for which
Φ : IdealpCq Ñ IdealpAq is an isomorphism of lattices.
The statements (i) and (ii) that Φ is an isomorphism between the ideal lattices of A and B,
respectively, of C and A, are included in the quoted results of Dadarlat, and it also follows
from Lemma 5.1 in the situation considered in (i).
To apply Theorem 5.3, we need the following:
Lemma 5.4. A separable just-infinite C -algebra is of type I if and only if either it is
isomorphic to K, the compact operators on a separable infinite dimensional Hilbert space, or
it is an essential extension of K by a finite dimensional C -algebra. In the former case, A is
just-infinite of type (α), and in the latter case A is just-infinite of type (β); cf. Theorem 3.10.
In particular, no just-infinite C -algebra of type (γ), i.e., RFD, is of type I.
Proof. Let A be a separable just-infinite C -algebra of type I. By Lemma 3.2, A is prime, and
hence primitive, so we can find a faithful irreducible representation π of A on a separable,
necessarily infinite dimensional, Hilbert space H. Since A is a C -algebra of type I, the
algebra K of compact operators on H is contained in the image of π. Hence I " π´1pKq
is a non-zero closed two-sided ideal in A, which is isomorphic to K. As A is just-infinite,
either I " A, or A{I is finite dimensional.
Corollary 5.5. Let A be a unital separable RFD just-infinite AF-algebra, and let tπnu8
n"1
be an exhausting sequence of pairwise inequivalent non-faithful irreducible representations
of A. It follows that A contains a unital non-nuclear RFD just-infinite sub-C -algebra B
of real rank zero such that tπnBu8
n"1 is an exhausting sequence of pairwise inequivalent
non-faithful irreducible representations of B, and πnpBq " πnpAq, for all n.
Proof. By Lemma 5.4, we can now apply Theorem 5.3 (i) to find a sub-C -algebra B of A
with the properties listed therein. Each projection p P A is equivalent to a projection in B.
Indeed, set q " κppq P B. Then q " pι κqppq is homotopic (and hence equivalent) to p.
The desired conclusion now follows from Lemma 5.2 (i).
Corollary 5.6. Let A be a unital separable RFD just-infinite AF-algebra. Then A is con-
tained in a separable non-exact unital RFD just-infinite C -algebra C of real rank zero,
equipped with an exhausting sequence of pairwise inequivalent non-faithful irreducible repre-
sentations tνnu8
n"1, such that their restrictions to A form an exhausting sequence of pairwise
inequivalent non-faithful irreducible representations for A, and νnpAq " νnpCq, for all n.
20
Proof. This follows immediately from Lemma 5.4, Theorem 5.3 (ii) and Lemma 5.2 (ii).
The two above corollaries, in combination with the existence of a RFD just-infinite AF-
algebra (see Section 4), now yield the following:
Corollary 5.7. There exist non-nuclear exact RFD just-infinite C -algebras, and there also
exist non-exact RFD just-infinite C -algebras.
It is shown in [32, Theorem 4.3] that each unital C -algebra A of real rank zero contains a
unital AF-algebra B such that each projection in A is equivalent to a projection in B, and
such that Φ : IdealpAq Ñ IdealpBq is an isomorphism. Together with Lemma 5.2 (i), this
proves the following:
Proposition 5.8. Let A be a unital separable RFD just-infinite C -algebra of real rank zero,
and let tπnu8
n"1 be an exhausting sequence of pairwise inequivalent non-faithful irreducible
representations of A. It follows that A contains a unital RFD just-infinite AF-sub-C -alge-
bra B such that tπnBu8
n"1 is an exhausting sequence of pairwise inequivalent non-faithful
irreducible representations of B, and πnpBq " πnpAq, for all n.
By combining Corollary 5.5 with Proposition 5.8, one obtains the following fact: Suppose
that A is a unital separable RFD just-infinite C -algebra of real rank zero, and tπnu8
n"1 is an
exhausting sequence of pairwise inequivalent non-faithful irreducible representations. Then
there is a strictly decreasing sequence A Ą A1 Ą A2 Ą A3 Ą of unital sub-C -algebras
Ak of A such that each Ak is a RFD just-infinite C -algebra, and πnpAkq " πnpAq, for
all k and n. (In fact, every other C -algebra in the sequence tAku can be taken to be an
AF-algebra and the remaining ones to be non-nuclear.)
In particular, a unital separable RFD just-infinite C -algebra of real rank zero can never
be minimal in the sense that it contains no proper RFD just-infinite sub-C -algebras.
6 Just-infiniteness of group C -algebras
We discuss in this section when C -algebras associated with groups are just-infinite.
The group algebra CrGs of a group G is in a natural way a -algebra in such a way that
each group element g P G becomes a unitary in CrGs, and it can be completed to become
a C -algebra, usually in many ways. The universal C -algebra of G, denoted by C pGq, is
the completion of CrGs with respect to the maximal C -norm on CrGs. Each unitary rep-
resentation π of the group G on a Hilbert space gives rise to unital -representations (again
denoted by π) of the -algebras CrGs and C pGq on the same Hilbert space. Respectively,
each unital -representation π of C pGq restricts to a -representation of CrGs, and if this
restriction is faithful, then it creates a C -norm } }π on this algebra. Each C -norm on
CrGs arises in this way, where by a C -norm on CrGs we mean a (faithful) norm such that
the completion of CrGs with respect to this norm is a C -algebra.
Given a unitary representation π of G, we let C
π pGq denote the completion of πpCrGsq.
This is equal to the completion of CrGs with respect to the norm } }π, if π is faithful
on CrGs. The reduced group C -algebra of G, denoted by C
λpGq, arises in this way from
the left-regular representation λ of G on ℓ2pGq. It is well-known that the maximal and the
reduced C -norms on CrGs are equal, i.e., C pGq " C
λ pGq, if and only if G is amenable
(see [13, Theorem 2.6.8]). It is also well-known (see, e.g., [13, Exercise 6.3.3]) that if the
reduced group C -algebra C
λ pGq has a finite-dimensional representation, then G must be
amenable. Hence the following holds:
21
Proposition 6.1. Let G be a group and suppose that C
C
λ pGq is simple, or G is amenable.
λ pGq is just-infinite. Then either
Whereas CrGs always has one maximal C -norm, there may or may not be a minimal C -
λ pGq is a simple
norm on CrGs, depending on the group G.
C -algebra, then the norm } }λ on CrGs is minimal.
If G is C -simple, i.e., if C
Proposition 6.2. Let G be a group, and let π be a representation of G which gives a faithful
representation of CrGs. If C
π pGq is just-infinite, then } }π is a minimal C -norm on CrGs.
Proof. Any C -norm on CrGs which is smaller than } }π arises from a unitary representation
ν of G on a Hilbert space, which factors through C
π pGq. Since ν is injective on CrGs, the
image νpC
π pGqq cannot be finite dimensional, so ν is injective, and hence isometric, on
π pGq. (Recall that each injective -homomorphism between C -algebras automatically is
C
isometric.) The norm arising from ν is therefore equal to the norm arising from π.
If G is infinite and if C
then C
π pGq is simple, for some unitary representation π of the group G,
π pGq is just-infinite and } }π is a minimal norm on CrGs; cf. Proposition 6.2.
The group algebra CrGs is said to be -just-infinite if each -representation of CrGs
either is injective, or has finite dimensional image. Note that -just-infinite is a formally
weaker condition than "just-infinite", as CrGs can have non-self-adjoint two-sided ideals.
Proposition 6.3. Let G be an infinite group. Then C pGq is just-infinite if and only if
CrGs is -just-infinite and CrGs has a unique C -norm.
Proof. Suppose first that C pGq is just-infinite. Let π be a unital -representation of CrGs,
and extend it to a -representation of C pGq. Then π is either injective on C pGq, or
πpC pGqq is finite dimensional. If π is injective on C pGq, then it is also injective on CrGs,
while if πpC pGqq is finite dimensional, then so is πpCrGsq. Hence CrGs is -just-infinite.
Each C -norm on CrGs arises as } }π, for some -representation π of C pGq which is
faithful on CrGs. Thus πpC pGqq is infinite dimensional, so π must be injective on C pGq.
This entails that } }π is the maximal norm on CrGs, and thus the only C -norm on CrGs.
Suppose now that CrGs has a unique C -norm, and that CrGs is -just-infinite. Let
π be a non-faithful unital -representation of C pGq. If the restriction of π to CrGs were
faithful, then it would induce a C -norm on CrGs, which by uniqueness would be equal to
the maximal C -norm on CrGs. This contradicts that π is non-faithful on C pGq. Hence
π is not faithful on CrGs, whence πpCrGsq is finite dimensional. In this case, πpC pGqq is
equal to πpCrGsq. This proves that C pGq is just-infinite.
Corollary 6.4. Let G be a group for which C pGq is just-infinite. Then G is amenable,
and hence C pGq " C
λpGq is nuclear.
Proof. It follows from Proposition 6.3 that the reduced and the maximal norm on CrGs
coincide, so G is amenable.
Corollary 6.5. For each group G, if C pGq is just-infinite, then CrGs is -just-infinite,
which in turn implies that G is just-infinite.
Proof. The first implication follows from Proposition 6.3. To see that the second implication
holds, suppose that CrGs is -just-infinite, and let N be a non-trivial normal subgroup
of G. The quotient map G Ñ G{N lifts to a necessarily non-injective -homomorphism
CrGs Ñ CrG{N s. Hence CrG{N s must be finite dimensional, whence G{N is finite.
22
None of the reverse implications above hold; cf. Examples 6.6 and 7.3.
Example 6.6. The group algebra CrZs is -just-infinite, and the group Z is just-infinite;
but C pZq is not just-infinite, and CrGs has no minimal C -norm.
Proof. Each unital -representation π of CrZs on a Hilbert space H admits a natural fac-
torization CrZs Ñ CpKq Ñ BpHq, where K Ď T is the spectrum of the unitary operator
u " πp1q, and where CpKq Ñ BpHq is injective. It is easy to see that π is faithful on CrZs
if and only if K is an infinite set. If π is not faithful, then K is finite, which entails that
πpCrZsq is finite dimensional. This shows that CrZs is -just-infinite.
As there is no minimal closed infinite subset of T, there is no minimal C -norm on CrZs,
and we conclude from Proposition 6.3 that C pZq is not just-infinite. This conclusion also
follows from Example 3.3.
Proposition 6.7. If G is a locally finite group, then CrGs has a unique C -norm.
Proof. Each element x P CrGs is a linear combination of finitely many elements from G,
and each finitely generated subgroup of G is finite, by assumption. Hence there is a finite
subgroup H of G such that x P CrHs Ď CrGs. Now, CrHs is a (finite dimensional) C -al-
gebra, so it has a unique C -norm. Thus any two C -norms on CrGs must agree on x. As
x was arbitrarily chosen, we conclude that CrGs has a unique C -norm.
Question 6.8. Let G be a group and suppose that CrGs has a unique C -norm. Does it
follow that G is locally finite?
The augmentation ideal of the full group C -algebra C pGq of a group G is the kernel of the
trivial representation C pGq Ñ C. If G is infinite and if the augmentation ideal is simple,
or, more generally, just-infinite, then C pGq is just-infinite by Proposition 3.5 (ii), since the
augmentation ideal always is essential when G is infinite.
There are locally finite groups whose augmentation ideal is simple, such as Hall's univer-
sal groups, see [9] and [27]. It follows from Theorem 3.10 that C pGq is just-infinite of type
(β), for any such group G. It is easy to see that if an amenable group G has simple augmen-
tation ideal, then it must be simple; however, simple groups (even locally finite ones) need
not have simple augmentation ideal: the infinite alternating group A8 is a counterexample.
Lemma 6.9. Let G be a residually finite group for which C pGq is just-infinite. Then
C pGq is RFD (and hence of type (γ); cf. Theorem 3.10).
Proof. Let tNiuiPI be a decreasing net of finite index normal subgroups of G with ŞiPI Ni "
teu, and consider the -homomorphism
Φ : C pGq Ñ źiPI
C pG{Niq.
It suffices to show that Φ is injective; and by the assumption that C pGq is just-infinite, it
further suffices to show that the image of Φ is infinite dimensional. The latter follows from
the fact that G is infinite (as C pGq is just-infinite) and (hence) that supiPI G : Ni " 8.
Question 6.10. Does there exist an infinite, residually finite group G such that C pGq is
just-infinite?
If such a group G exists, then C pGq will be a RFD just-infinite C -algebra by Lemma 6.9.
If the answer to Question 6.8 is affirmative, then G must be locally finite. This leads to the
following:
23
Question 6.11. Does there exist an infinite, residually finite, locally finite (necessarily
just-infinite) group G such that CrGs is -just-infinite?
If such a group G exists, then C pGq will be a RFD just-infinite C -algebra by Lemma 6.9,
Proposition 6.7 and Proposition 6.3. After the first version of this paper was made public,
Question 6.11 has been answered in the affirmative in [7].
Just-infinite groups are divided into three disjoint subclasses (the trichotomy for just-
infinite groups), see [22, Section 6]: The non-residually finite ones (which contain a finite
index normal subgroup N which is the product of finitely many copies of a simple group),
branch groups (see more about those in Theorem 7.10 below), and the hereditarily just-
infinite groups, i.e., the residually finite groups for which all finite index normal subgroups
are just-infinite. It is shown in Theorem 7.10 below that if G is a just-infinite branch group,
then CrGs is not -just-infinite, whence C pGq is not just-infinite. Hence, if there exists
a residually finite group G for which C pGq is just-infinite (and hence also RFD), then G
must be hereditarily just-infinite.
Consider the following three (classes of) examples of hereditarily just-infinite groups:
the integers Z, the infinite dihedral group D8, and PSLnpZq, for n ě 3. As shown below,
if G is any of these groups, then C pGq is not just-infinite. Moreover, there is no unitary
representation π of G such that C
π pGq is RFD and just-infinite. If G " Z, then this claim
follows immediately from Example 3.3. In the two examples below we discuss the situation
for the two other (classes of) groups.
Example 6.12 (PSLnpZq, n ě 3). The groups PSLnpZq, n ě 3, are renowned for being
the first examples of infinite groups with Kazdan's property (T), as first shown by Kazdan.
For a different and nice proof by Shalom, see [34]. They are residually finite, as witnessed
by the finite quotient groups PSLnpZ{N Zq, N P N; and they are hereditarily just-infinite
by Margulis' normal subgroup theorem. Bekka–Cowling–de la Harpe proved in [4] that
PSLnpZq is C -simple, for all n ě 2.
In particular, PSLnpZq is an ICC group (all its
non-trivial conjugacy classes are infinite).
We conclude from these facts that the C -algebra C
λ pPSLnpZqq is just-infinite (being
simple) for all n ě 2, while the full group C -algebra C pPSLnpZqq is not just-infinite,
because PSLnpZq is non-amenable, for n ě 2.
Bekka proved in [3] that the set of extremal characters on PSLnpZq, for n ě 3, is
a countably infinite set consisting of the trivial character δe and a sequence tδku8
k"1 of
characters, each of which factors through a finite quotient, PSLnpZ{N Zq, of PSLnpZq for
a suitable integer N (depending on k). Recall that each (extremal) character on a group
corresponds to an extremal trace on its full group C -algebra. The trivial character δe
on PSLnpZq corresponds to the canonical trace τ0 on C pPSLnpZqq; while for k ě 1, the
character δk corresponds to a trace, denoted by τk, whose GNS-representation πτk is finite
dimensional. Bekka also shows that τk Ñ τ0 in the weak topology.
Furthermore, observe that C pPSLnpZqq has a just-infinite quotient, namely the simple
λ pPSLnpZqq. However, as shown below, there is no unitary representation π
C -algebra C
of PSLnpZq such that C
π pPSLnpZqq is RFD and just-infinite.
Indeed, assume that π is such a unitary representation of PSLnpZq. As in Remark 3.15,
let tπju8
j"1 be an exhausting sequence of pairwise inequivalent non-faithful irreducible rep-
resentations of C
π pPSLnpZqq. Then ρj " πj π, j ě 1, is a sequence of pairwise inequivalent
non-faithful irreducible representations of C pPSLnpZqq. Hence ρj is equivalent to πτkpjq, for
some kpjq ě 1, by the above mentioned result of Bekka. Suppose now that x P C pPSLnpZqq
belongs to the kernel of π. Then, for all j ě 1, πτkpjqpxq " 0, so τkpjqpxxq " 0. It follows
that τ0pxxq " limjÑ8 τkpjqpxxq " 0, so λpxq " 0. This shows that λ is weakly contained
24
in π. We conclude that the left-regular representation λ factors through π, so the simple
C -algebra C
π pPSLnpZqq. Each simple quotient of a RFD
just-infinite C -algebra is finite dimensional, so C
π pPSLnpZqq cannot both be RFD and
just-infinite.
λpPSLnpZqq is a quotient of C
Example 6.13 (The infinite dihedral group D8). The infinite dihedral group D8 is an
example of a hereditarily residually finite just-infinite group, see [22], and it is isomorphic
to the free product Z2 Z2, which is an amenable group (of linear growth). The group C -
algebra C pZ2 Z2q is known to be a sub-C -algebra of M2pCpr0, 1sqq (being the universal
unital C -algebra generated by two projections), and is hence subhomogeneous (cf. Proposi-
tion 3.16). Clearly, any quotient of a subhomogeneous C -algebra is again subhomogeneous,
so we conclude from Lemma 3.17 that C pZ2 Z2q is not just-infinite, and neither is any of
its quotients.
7 Algebras associated with groups of intermediate growth
In this section we present some results concerning algebras associated with the 3-generated
infinite torsion group constructed in [19], which we here will denote by G. This group is
a simple example of a group of Burnside-type, and it is investigated more deeply in [20]
and many other papers (see the surveys [23], [24], and the references therein). Among
its unusual properties, most notably G is of intermediate growth (between polynomial and
exponential), and, as a consequence, it is amenable, but not elementary amenable, thus
answering questions by Milnor and Day, respectively; cf. [20]. Furthermore, G is a just-
infinite group of branch type (and hence residually finite), and moreover, it is a self-similar
group (i.e., a group generated by states of a Mealy-type automaton).
There are indications that various C -algebras associated with G (including C pGq and
some of its quotients, discussed below) may be new types of C -algebras with properties
unseen yet in the theory of operator algebras. Our main conjecture in this direction is the
following:
Conjecture 7.1. The (self-similar RFD) C -algebra C
representation π of G is just-infinite.
π pGq generated by the Koopman
The Koopman representation π of G will be described below, along with the notion of self-
π pGq is a just-infinite C -algebra of
similarity.
type (γ) as described in Theorem 3.10; cf. Lemma 6.9, and it is the first example of such a
C -algebra associated with a group.
If the conjecture above is correct, then C
Recall that the group G is generated by four elements a, b, c, d satisfying the relations
1 " a2 " b2 " c2 " d2 " bcd " σk`padq4 " σk`padacacq4,
for k " 0, 1, 2, . . . , where the permutation σ is given by the substitutions:
(7.1)
a Ñ aca,
b Ñ d,
c Ñ b,
d Ñ c.
This presentation was found by I. Lysenok in [28], and it is a minimal presentation (in the
sense that no relator in (7.1) can be deleted without changing the group, see [21]). In fact,
G is generated by 3 elements, as d " bc. The set t1, b, c, du is the Klein group Z{2Z ' Z{2Z.
For our purposes it will be most important to know that G has a faithful self-similar
action by automorphisms on an infinite binary rooted tree T , as shown, in part, here:
25
‚
rrrrrrrrrrrr
✿✿✿✿✿✿✿
✑✑✑✑✑✑
✲✲✲✲✲✲
‚
‚
‚
v0 ‚
✴✴✴✴✴✴
‚
‚
✑✑✑✑✑✑
‚
▲▲▲▲▲▲▲▲▲▲▲▲
☎☎☎☎☎☎☎
✲✲✲✲✲✲
✑✑✑✑✑✑
‚
‚
‚
‚
v1
❃❃❃❃❃❃❃
✎✎✎✎✎✎
‚
‚
✲✲✲✲✲✲
‚
(7.2)
The generators a, b, c, d act on T as follows: The root of the tree (marked in red) is a common
fixed point. The generator a just permutes the two vertices v0 and v1 at the first level and
acts trivially inside the subtrees T0 and T1 with roots v0 and v1, respectively. The generators
b, c, d fix the vertices v0 and v1 (and hence leave the subtrees T0 and T1 invariant), and they
are defined recursively by:
bT0 " a,
bT1 " c,
cT0 " a,
cT1 " d,
dT0 " 1,
dT1 " b,
(7.3)
when identifying the subtrees T0 and T1 with T in the natural way, and where 1 stands for
the identity automorphism. For more details on this definition, and other definitions of G,
we refer to [19, 20, 23, 24]. The relations (7.3) imply that G is a self-similar group in the
sense that it has a natural embedding
ψ : G Ñ G ≀ pZ{2Zq – pG Gq ¸ pZ{2Zq,
(7.4)
where Z{2Z " te, εu acts on G G by permuting the two copies of G (e is the identity
element and ε is a transposition). The embedding ψ is given as follows:
ψpaq " p1, 1qε " ε, ψpbq " pa, cqe " pa, cq, ψpcq " pa, dqe " pa, dq, ψpdq " p1, bqe " p1, bq.
To further illlustrate this action of G on the tree T it is convenient to label the vertices of
the nth level, Vn, of T by the set t0, 1un and equip each Vn with the lexicographic ordering:
♠♠♠♠♠♠♠♠♠♠♠♠♠♠♠♠♠
❈❈❈❈❈❈❈❈
☛☛☛☛☛☛
✸✸✸✸✸✸
01
0
④④④④④④④④
✸✸✸✸✸✸
001
00
☛☛☛☛☛☛
000
H
◗◗◗◗◗◗◗◗◗◗◗◗◗◗◗◗◗
④④④④④④④④
✸✸✸✸✸✸
☛☛☛☛☛☛
10
(7.5)
1
❈❈❈❈❈❈❈❈
☛☛☛☛☛☛
110
11
✸✸✸✸✸✸
111
010
011
100
101
The action of the group G on T yields an action of G by homeomorphisms on the boundary
BT of T , which consists of geodesic rays joining the root H with infinity. The boundary BT
can in a natural way be identified with the Cantor set t0, 1uN of infinite binary sequences
equipped with the Tychonoff topology.
n"1 t 1
Let µ " Ś8
2 u be the uniform Bernoulli measure on BT . It is invariant with respect
to the action of the entire group AutpT q of automorphisms on T , and hence with respect
to the action of G on T . The topological dynamical system pG, BT q can be converted into
2 , 1
26
a metric dynamical system pG, BT, µq which is ergodic (while pG, BT q is minimal), because
the action of G on each level Vn is transitive, see [25, Proposition 6.5].
Let π be the (unitary) Koopman representation of G on the Hilbert space L2pBT, µq
given by `πpgqfpxq " f pg´1xq, where f P L2pBT, µq, g P G, and x P BT . We denote the
image of the group algebra CrGs under the representation π by B, and we let as usual C
denote the completion of CrGs with respect to the norm induced by π.
π pGq
The following theorem carries some evidence in support of Conjecture 7.1.
Theorem 7.2. Let G " xa, b, c, dy be the infinite torsion group of intermediate growth from
above, let π be the Koopman representation of G, and let B " πpCrGsq. Then:
(i) B is self-similar, infinite dimensional and RFD.
(ii) C
π pGq is self-similar, infinite dimensional, RFD, and it posseses a faithful trace.
(iii) The natural surjection π : CrGs Ñ B is not injective, whence CrGs is not -just-infinite.
(iv) B is just-infinite.
The notions of self-similarity of the algebras B and C
rem 7.2 above is proved at the end of this section.
π pGq will be explained below. Theo-
The type of just-infinite algebras (also called "thin algebras") considered above were
studied for the first time by Sidki in [35]. The group used by Sidki was the Gupta–Sidki
3-group H, and the algebra was defined over a field F3 in a rather involved way as a certain
inductive limit. If one considers the "Koopman" representation of H over the field F3, then
the image of the group algebra F3rHs will be isomorphic to Sidki's thin algebra.
The C -algebra generated by the Koopman representation of the group G (considered
in this section) was considered in [2], and so was the algebra B, even though it was not
explicitly defined there. Vieira, [36], used Sidki's approach to define a "thin algebra" of the
group G over the field F2, and proved that it is just-infinite.
Thin algebras under the name "Tree enveloping algebras" were considered by Bartholdi
in [1]. He defines algebras, similar to the algebra B in Theorem 7.2, however, over arbitrary
fields. He considers a vector space with a basis consisting of all points of the boundary of the
rooted tree, and then defines an algebra as the image of the group algebra in the algebra of
endomorphisms of this huge vector space. One can show that if the field is complex numbers
and the group is the group G, then Bartholdi's algebra is isomorphic to the algebra B we
are considering here.
In [1, Theorem 3.9], a sufficient condition is given for the tree enveloping algebra to be
just-infinite. This condition is satisfied in the case of the group G.
Example 7.3. As mentioned above, the group G is just-infinite. We can therefore deduce
from Theorem 7.2(iii) that just-infiniteness of a group G does not imply that its complex
group algebra CrGs is -just-infinite.
Self-similarity of graphs, Hilbert spaces, representations and algebras
Let X " tx1, x2, . . . , xdu be an alphabet on d ě 2 letters, let X " Ů8
n"0 X n be the set of
words over X, and let T " TX be the d-arnery rooted tree whose vertices are in bijection
with the elements of X (so that the nth level Vn of T corresponds to X n). The action of
an arbitrary group G on T by automorphisms induces an action G ñ X . This action is
said to be self-similar if for all g P G and all x P X, there are h P G and y P X such that
27
gpxwq " yhpwq, for all w P X . If this holds, then for every v P X , there exists a unique
h P G satisfying, for all w P X ,
gpvwq " gpvqhpwq.
(7.6)
The element h is called the section (or restriction) of g in v, and is denoted by h " gv. For
example, for the group G " xa, b, c, dy under examination, we have; cf. (7.3), that
av0 " av1 " 1,
bv0 " a,
bv1 " c,
cv0 " a,
cv1 " d,
dv0 " 1,
dv1 " b.
Let H be a separable infinite dimensional Hilbert space, and fix an integer d ě 2. A
unitary operator u : H Ñ H d " H ' H ' ' H is called a d-similarity of H. Each
d-similarity arises from d isometries s1, . . . , sd on H, satisfying řd
ÿj"1
upξ1, . . . , ξdq "
upξq " ps
1 ξ, . . . , s
d ξq,
d
sjξj,
j"1 sjs
j " 1, as follows:
for ξ, ξ1, . . . , ξd P H. Observe that s1, . . . , sd define a representation of the Cuntz algebra Od,
and that every representation of Od is obtained in this way. For each x " pxi1 , xi2 , . . . , xinq P
X consider the isometry on H given by Sx " si1 si2 . . . sin, and observe that SxSy " Sxy.
A unitary representation ρ of a group G on a Hilbert space H is said to be self-similar with
respect to the d-similarity ψ considered in (7.4) above, if
ρpgqSx " Syρphq,
(7.7)
for all g, h P G and all x, y P X satisfying gpxwq " yhpwq, for all w P X . In other words,
ρpgqSx " Sgpxqρpgxq, for all g P G and x P X .
The image Bρ " ρpCrGsq, where ρ is a self-similar representation, is called a self-similar
(abstract) algebra. The C -algebra C
ρ pGq associated with a self-similar representation ρ is
called a self-similar C -algebra. One of the features of the self-similar algebra Bρ (or of the
C -algebra C
ρ pGq) is the existence of the unital embedding
ψρ : Bρ Ñ MdpBρq,
b ÞÑ
s
1 bs1
...
s
dbs1
s
1bsd
...
s
dbsd
,
‹‚
(7.8)
It follows from (7.7) that s
for b P Bρ.
j Bρsi Ď Bρ, for all i, j. The embedding ψρ is
typically not surjective. Nonetheless, it has many interesting and non-trivial features, see,
for example, Lemma 7.6 below.
In the case of our main example G " xa, b, c, dy and of the Koopman representation π of
G on H " L2pBT, µq, we have an explicit self-similarity H Ñ H ' H arising from the two
isometries s0, s1 on H defined by
psif qpxq " f pixq,
(7.9)
for i " 1, 2, where f P L2pBT, µq and x P BT , and where ix P BT " t0, 1uN is the word
obtained by putting the letter i in front of the word x. The resulting embedding ψπ : B Ñ
M2pBq is given as follows on the generators:
ψπp¯aq " 0 1
1 0 , ψπp¯bq " ¯a 0
0 ¯c , ψπp¯cq " ¯a 0
¯d , ψπp ¯dq " 1 0
0 ¯b ,
0
(7.10)
(as can be deduced from (7.7) and (7.8)), where we have introduced the notation ¯g " πpgq,
for g P G. (The Koopman representation is faithful on G, so the map g ÞÑ ¯g is injective, but
the Koopman representation is not faithful on CrGs; cf. Theorem 7.2, so it is pertinent to
distinguish between g and πpgq.)
28
More on the Koopman representation
What we are going to present here is known in the more general situation of groups acting
on rooted trees, [2, 5, 23]. Consider the binary rooted tree T (as described in (7.2) and
(7.3)), and the Koopman representation π of the group G " xa, b, c, dy on L2pBT, µq.
For each n ě 1, let vn,1, vn,2, . . . , vn,2n be the order preserving enumeration of the set
i"1 En,i,
Vn " t0, 1un (equipped with the lexicographic ordering); cf. (7.5), and write BT " Ů2n
where En,i is the set of infinite words in BT " t0, 1uN that start with vn,i. Set
Hn " spantχEn,i i " 1, 2, . . . , 2nu Ď H " L2pBT, µq,
which is a subspace of dimension 2n. Since En,i " En`1,2i´1 Y En`1,2i, we see that Hn Ď
Hn`1. Moreover, as the cylinder sets En,i, n ě 1, 1 ď i ď 2n, form a basis for the topology
on BT , it follows that Ť8
n"1 Hn is dense in H.
The subspaces Hn are π-invariant. Let πn be the restriction of π to Hn, for n ě 1.
Observe that πn is unitarily equivalent to the representation of G on ℓ2pVnq arising from its
action on the nth level Vn of the tree T . More specifically, identify Hn with ℓ2pVnq via the
isomorphism that identifies χEn,i with δvn,i. Write Hn`1 " Hn ' H K
n denote the
restriction of π to H K
n has dimension 2n. It is shown in the appendix of [5]
that the representation πK
n of G is irreducible, for each n ě 1. Thus we have decompositions
n . Note that H K
n , and let πK
H " C '
8
àn"0
H K
n ,
π " 1 '
8
àn"0
πK
n ,
(7.11)
of the Hilbert space H and of the representation π into irreducible representations, where
we identify H0 with C, and π0 with the trivial representation 1.
The proof of Theorem 7.2
Proof of Theorem 7.2 (i): Recall from (7.9) that we have isometries s0, s1 on the Hilbert
space H " L2pBT, µq satisfying the Cuntz relation s0s
1 " 1. The range of the isometry
si is L2pBTi, µiq, where T0 and T1 are the subtrees of T with roots v0 and v1, respectively; cf.
(7.2), and where µ0 and µ1 are the normalized restrictions of µ to the subsets BT0 and BT1,
respectively, (making them probability measures). The Koopman representation π is self-
similar with respect to the 2-similarity of H given by the isometries s0, s1, so B " πpCrGsq
is self-similar.
0 `s1s
By (7.11) and irreducibility of the representations πK
n , we see that B is a subalgebra of
M :" C '
8
źn"0
M2npCq,
(7.12)
with the property that the projection of B onto each summand in (7.12) is surjective. Hence
B is infinite dimensional and RFD. This completes the proof of (i).
Proof of Theorem 7.2 (ii): It follows from (7.11) that the inclusion of B into M is isometric,
when B is equipped with the norm arising from the Koopman representation π. Thus C
π pGq,
which is the completion of B with respect to this norm, embeds into M . Hence C
π pGq is
RFD. Moreover, it is infinite dimensional because it contains the infinite dimensional algebra
B, and it is self-similar because the Koopman representation π is self-similar. Finally, M
has a faithful trace, for example the one given by
αjτjpxjq,
τ pxq "
8
ÿj"´1
29
where x " px1, x0, x1, . . . q P M , τn is the normalized trace on M2npCq, for each n ě 1 (and
τ´1 and τ0 are the normalized traces on C), and tαju8
j"´1 is any sequence of strictly positive
π pGq has a faithful trace, being a sub-C -algebra of M .
numbers summing up to 1. Hence C
Proof of Theorem 7.2 (iii): The first claim of (iii) is proved in the lemma below, and the
second claim follows from the first claim and the fact, proved in (i), that B is infinite
dimensional.
The result below can be found in [26]. We include its proof for completeness of the exposition.
Lemma 7.4. p1 ´ dqap1 ´ dq is a non-zero element in the kernel of π : CrGs Ñ B.
Proof. We observe first that z :" a ´ da ´ ad ` dad is non-zero in CrGs. Indeed, if z " 0,
then a ` dad " da ` ad, which can happen only if either a " da and dad " ad, or a " ad
and dad " da. Both are impossible, because d ‰ e. (It is also easy to see, for example using
the action of G of the tree T , that the four elements a, da, ad, dad are pairwise distinct.)
By (7.10) (and retaining the notation ¯g " πpgq, for g P G), we have
ψπpπpzqq " ψπ`p1 ´ ¯dq¯ap1 ´ ¯dq " 0
0
0 1 ´ ¯b " 0,
0
0 1 ´ ¯b0 1
1 00
where ψπ is the embedding of B into M2pBq arising from self-similarity. As ψπ is injective,
this implies that πpzq " 0.
The proof of part (iv) of Theorem 7.2 is somewhat lengthy and is divided into several
lemmas. The proof mimics the proof of the fact that G is a just-infinite group, as well as the
idea from the proof of [22, Theorem 4] showing that a proper quotient of an arbitrary branch
group is virtually abelian. In our situation, the following statement from [22, Proposition 8]
is useful:
Proposition 7.5. The normal subgroup K " xpabq2yG has finite index 16 in G, and it is of
self-replicating type, written K K ă K, i.e., K K Ď ψpKq, where ψ is given by (7.4).
Let ψ : G Ñ pG Gq ¸ Z{2Z be as defined in (7.4). For each m ě 1, the stabilizer subgroup
StGpmq of G, with respect to the action of G on the tree T , is the set of elements g P G
In particular, if g P StGp1q, then
that fix all vertices at level m, i.e., all vertices in Vm.
ψpgq P G G. The group K is a subgroup of StGp1q.
It is also shown in [19] that G itself is self-replicating (or recurrent), in the sense that
StGp1q ďS G G, where ďS is subdirect product. This means that the group homomorphisms
StGp1q
ψ
/ G G
πj
/ G,
where πj, j " 0, 1, are the coordinate homomorphisms, are surjective.
Let ∆ be the ideal in B generated by the set t¯k ´ 1 k P Ku Ď B, where K is as in
Proposition 7.5 above. Then B{∆ has dimension at most 16. To see this, let tt1, t2, . . . , t16u
be representatives of the cosets of K in G. For each g P G, there exist i in t1, 2, . . . , 16u and
k in K such that g " tik " ti ` tipk ´ 1q, so ¯g P ¯ti ` ∆. This shows that B{∆ is the linear
span of the elements ¯t1 ` ∆, ¯t2 ` ∆, . . . , ¯t16 ` ∆.
Let ψπ : B Ñ M2pBq be as defined in (7.8), and let ψn
π : B Ñ M2npBq " B b M2n pCq
denote the "nth iterate of ψπ", in the sense that
ψn
π " pψπ b idM2n´1 pCqq pψπ b idM2pCqq ψπ.
The homomorphisms ψn
π are not surjective, but the following holds:
30
/
/
Lemma 7.6. For each n ě 1, M2np∆q Ď ψn
π p∆q.
Proof. The lemma follows easily by induction on n, once the base step n " 1 has been
verified. So let us show that M2p∆q Ď ψπp∆q.
It follows from Proposition 7.5 that for each k P K we can find k1 P K such that
ψpk1q " pk, 1q. Hence
ψπp ¯k1q " ¯k 0
0 1 ,
ψπp ¯k1 ´ 1q " ¯k ´ 1 0
0 .
0
Let x, x1 P ∆ be such that
ψπpx1q " x 0
0 0 .
(7.13)
(7.14)
Since G is self-replicating; cf. the comments below Proposition 7.5, we can for each f P G
find g P StGp1q and h P G, such that ψpgq " pf, hq. Then
ψπp¯gx1q " ¯f x 0
0 ,
0
ψπpx1¯gq " x ¯f
0
0
0 .
Together with (7.13), this shows that
∆ 0
0 Ď ψπp∆q.
0
(7.15)
If x, x1 P ∆ are such that (7.14) holds, then
ψπpx1aq " 0 0
x 0 ,
ψπpax1q " 0 x
0 0 ,
ψπpax1aq " 0 0
0 x .
Together with (7.15), this completes the proof.
Lemma 7.7. dimpB{∆2q ď G : rK, Ks ă 8.
Proof. Let ∆1 be the ideal in B generated by the set t¯k ´ 1 k P rK, Ksu. Exactly as in the
argument above, showing that the dimension of B{∆ is at most G : K " 16, we see that
the dimension of B{∆1 is at most G : rK, Ks. Now, K is finitely generated, and so is the
quotient K{rK, Ks, which, moreover, is an abelian torsion group. Hence K{rK, Ks is finite,
so G : rK, Ks " G : KK : rK, Ks is finite. For all k1, k2 P K,
which shows that ∆1 Ď ∆2. This proves the lemma.
2 `pk1 ´ 1qpk2 ´ 1q ´ pk2 ´ 1qpk1 ´ 1q P ∆2,
rk1, k2s ´ 1 " k´1
1 k´1
One more property of G, that we are going to exploit, is the so-called contracting property,
already used in [19]. Let g denote the length of g P G with respect to the canonical
generating set ta, b, c, du. With ψ : G Ñ pG Gq ¸ Z{2Z as defined in (7.4), and g P G, we
have ψpgq " pg0, g1qη, where g1, g2 P G and η P te, ǫu. By [19], see also [23, Lemma 3.1],
gi ď
g ` 1
2
,
(7.16)
for i " 0, 1. In particular, gi ă g if g ě 2. The set of elements g P G for which g ď 1 is
equal to N " t1, a, b, c, du, which is called the nucleus of G.
We can repeat this process and obtain for each g P G and v P t0, 1un a section gv "
gv P G (defined underneath (7.6)), such that ψpgvq " pg0v, g1vqηv, where ηv P te, ǫu and
giv ď pgv ` 1q{2, for i " 0, 1. It follows that, for each g P G, there exists n ě 1 such that
gv P N , for all v in t0, 1un. By the construction of the self-similarity map ψπ : B Ñ M2pBq,
this leads to the following:
31
Lemma 7.8. For each x in B, there exists n ě 1 such that the 2n 2n matrix ψn
M2npBq has entries in the linear span of the element in the nucleus ¯N " t1, ¯a, ¯b, ¯c, ¯du.
π pxq P
Next we will prove:
Lemma 7.9. Let J be a non-zero ideal in B. There is m ě 1 so that M2m p∆2q Ď ψm
π pJq.
Proof. Let x be a non-zero element in J. Suppose that there exists m ě 1 such that one of
π pxq is a non-zero scalar λ. Denote by epmq
the 2m 2m entries, say the ps, tqth entry, of ψm
,
i, j " 1, 2, . . . , 2m, the standard matrix units of M2m pCq. Then, upon identifying M2m pBq
with B b M2m, we have
ij
pp b epmq
is qψm
π pxqpq b epmq
tj q " λpq b epmq
ij
,
(7.17)
for all p, q P B and all i, j " 1, 2, . . . , 2m. It follows from (7.17) and from Lemma 7.6 that
pq b epmq
π pJq, for all p, q P ∆. We conclude that z b epmq
π pJq,
for all z P ∆2 and all i, j " 1, 2, . . . , 2m, and hence that M2mp∆2q Ď ψm
ij belongs to ψm
belongs to ψm
ij
π pJq.
To complete the proof, we show below that one of the entries of ψm
π pxq is a non-zero
scalar, for some m ě 1.
Let n ě 1 be as in Lemma 7.8 (associated with our given x P B). Write ψn
π pxq "
s,t"1 with xs,t P B. By the choice of n, we deduce that xs,t belongs to the span of
π pxq is non-zero, so we can find s, t
pxs,tq2n
¯N " t1, ¯a, ¯b, ¯c, ¯du, for all s, t. Since ψn
such that xs,t is non-zero. Write
π is injective, ψn
xs,t " ρ 1 ` ξ¯a ` β¯b ` γ¯c ` δ ¯d,
for suitable ρ, ξ, β, γ, δ P C. Observe that, by (7.10),
ψπpxs,tq " pβ ` γq¯a ` δ ` ρ
ξ
ξ
β¯c ` γ ¯d ` δ¯b ` ρ .
(7.18)
The proof is now divided into three cases:
1). Assume that ξ ‰ 0. In this case both off diagonal entries of ψπpxs,tq are non-zero
pxq, at least one
scalars, and since ψπpxs,tq is a sub-matrix of the 2n`1 2n`1 matrix ψn`1
of the entries of ψn`1
pxq is a non-zero scalar.
π
π
2). Assume that either β ` γ ‰ 0, or δ ` ρ ‰ 0. Use (7.10) to compute the 2 2 matrix
ψπppβ ` γq¯a ` δ ` ρq " δ ` ρ β ` γ
δ ` ρ .
β ` γ
By assumption, one of the scalar entries in this matrix is non-zero. Further, it is a sub-
πpxs,tq and hence a sub-matrix of the 2n`2 2n`2 matrix
matrix of the 4 4 matrix ψ2
ψn`2
pxq. Thus at least one of the matrix entries of ψn`2
pxq is a non-zero scalar.
π
π
3). Assume that ξ " β ` γ " δ ` ρ " 0. Then
ψπpβ¯c ` γ ¯d ` δ¯b ` ρq " pβ ` γq¯a ` δ ` ρ
0
0
β¯c ` γ ¯d ` δ¯b ` ρ " 0
0 β¯c ´ β ¯d ` δ¯b ´ δ ,
0
and
ψπpβ¯c ´ β ¯d ` δ¯b ´ δq " pβ ` δq¯a ` β ` δ
0
0
β ¯d ´ β¯b ` δ¯c ´ δ .
32
π
If β ` δ ‰ 0, then, as in step 2), ψπppβ ` δq¯a ` β ` δq is a non-zero scalar 2 2 matrix,
which is a sub-matrix of the 16 16 matrix ψ4
πpxs,tq, whence at least one of the entries of
ψn`4
pxq is a non-zero scalar.
If β ` δ " 0, then β ‰ 0 (because xs,t ‰ 0), so β¯c ´ β ¯d ` δ¯b ´ δ " βp¯c ´ ¯d ´ ¯b ` 1q, and
ψπp¯c ´ ¯d ´ ¯b ` 1q " 0
¯b ´ ¯c ´ ¯d ` 1 .
¯d ´ ¯b ´ ¯c ` 1 , ψπp ¯d ´ ¯b ´ ¯c ` 1q " 2 ´ 2¯a
0
0
0
0
Arguing as in step 2), we see that ψπp2 ´ 2¯aq is a non-zero scalar 2 2 matrix, which is
a sub-matrix of the 32 32 matrix ψ5
pxq is a
non-zero scalar.
πpxs,tq, so at least one of the entries of ψn`5
π
We are now ready to complete the proof of Theorem 7.2.
Proof of Theorem 7.2 (iv): Let J be a non-zero ideal in B. Use Lemma 7.9 to find n ě 1
such that M2np∆2q Ď ψn
π is injective, it follows that
π pJq. Since ψn
dimpB{Jq " dimpψn
π pBq{ψn
π pJqq ď dimpM2n pB{∆2qq " 22n dimpB{∆2q ă 8,
by Lemma 7.7. This completes the proof.
We end our paper by showing that if G is a residually finite group for which CrGs is -just-
infinite, then G is hereditarily just-infinite (see also the discussion at the end of Section 6).
Indeed, if CrGs is -just-infinite, then G is just-infinite, by Corollary 6.5. By the trichotomy
for just-infinite groups, [22, Section 6], G must be either a branch group or hereditarily
just-infinite, and the theorem below rules out the former possibility.
We remind the reader about some facts concerning branch groups (see also [22]). Con-
sider a spherically homogeneous rooted tree T " T ¯m, where ¯m " tmnu8
n"0, is the branching
index of the tree (each mn ě 2 is an integer). For each vertex v in the kth level of the tree
T , let Tv be the sub-tree of T consisting of all vertices "below" v, so that Tv is a rooted tree
with root v and branching index tm1
nu8
n"0, where m1
n " mn`k.
Suppose that G is a group that acts on such a spherically homogeneous rooted tree T .
Then G fixes the root of the tree and hence leaves each level of the tree invariant. The rigid
stabilizer of a vertex v P T , denoted by ristGpvq, is the subgroup of G consisting of all g P G
which act trivially outside Tv (and fix v). The rigid stabilizer, ristGpnq, at level n P N is
the subgroup of G generated by the rigid stabilizers ristGpvq of all vertices v at level n. It
is easy to see that ristGpnq is, in fact, the direct product of the groups ristGpvq, where v is
a vertex at level n.
A group G is said to be a branch group if it admits a faithful action on such a spherically
homogeneous rooted tree T " T ¯m, such that the index G : ristGpnq is finite, for all n P N,
and such that T acts transitively on each level of the tree.
Theorem 7.10. If G is a branch group, then CrGs is not -just-infinite, whence C pGq is
not just-infinite.
Proof. Fix an action of G on a spherically homogeneous rooted tree T " T ¯m satisfying the
above mentioned conditions. Let π be the Koopman representation of G into the unitary
n"0 µn, and µn is the uniform prob-
ability measure on the set t1, 2, . . . , mnu. Denote also by π the associated -representation
CrGs Ñ BpHq.
group of the Hilbert space H " L2pBT, µq, where µ " Ś8
We show that π : CrGs Ñ BpHq is not injective, and that πpCrGsq is infinite dimensional.
This will imply that CrGs is not -just-infinite, and hence (by Corollary 6.5) that C pGq
33
is not just-infinite. Since G acts level transitively on T , we conclude that G is infinite and
that πpCrGsq is infinite dimensional.
is isomorphic to Śm
Let m " mp0q and let v1, v2, . . . , vm be the vertices at the first level of the tree T (below
the root of the tree). The condition that G : ristGp1q is finite implies that ristGp1q, which
j"1 ristGpvjq, is infinite. Moreover, by level transitivity of the action of
G on T , the rigid stabilizers ristGpvjq are pairwise conjugate, so they are, in particular,
non-trivial. We can therefore choose gj P ristGpvjq, for j " 1, 2, such that gj ‰ 1. Observe
that p1 ´ g1qp1 ´ g2q " 1 ´ g1 ´ g2 ` g1g2 is non-zero in CrGs, because g1 ‰ 1 and g2 ‰ 1.
For i " 1, 2, . . . , m, let Xi be the subset of BT consisting of words that start with vi, i.e.,
Xi " BTvi , so that BT is the disjoint union of the sets X1, X2, . . . Xm. Set Hi " L2pXi, µq.
i"1 Hi. Let Pi be the projection from H onto Hi. Since gj acts trivially on the
sub-trees Tvi, for i ‰ j, we conclude that Pi commutes with πpgjq for i " 1, 2, . . . , m and
j " 1, 2, and Piπpgjq " Pi, when j ‰ i. Hence πp1 ´ gjqPi " 0, for i ‰ j. It follows that
Then H " Àm
πpp1´g1qp1´g2qq " πpp1´g1qp1´g2qq
m
ÿi"1
Pi " πpp1´g1qp1´g2qqP2 " πp1´g1qP2πp1´g2q " 0,
so π : CrGs Ñ BpHq is not injective, as wanted.
The theorem above (and its proof) contains item (iii) of Theorem 7.2, since G is a branch
group. As in the conclusion of Theorem 7.2, it can happen, at least for some just-infinite
branch groups G (for example, when G " G), that πpCrGsq is just-infinite.
It may also
π pGq is a RFD just-infinite C -alge-
happen, for some just-infinite branch groups G, that C
bra, where π as above is the Koopman representation of G arising from its action on a tree.
We conjecture that C
π pGq is a RFD just-infinite C -algebra.
References
[1] L. Bartholdi, Branch rings, thinned rings, tree enveloping rings, Israel J. Math. 154 (2006),
93–139.
[2] L. Bartholdi and R. I. Grigorchuk, On the spectrum of Hecke type operators related to some
fractal groups, Proc. Steklov Inst. Math. 231 (2000), no. 4, 1–41, English translation.
[3] M. Bekka, Operator-algebraic superridigity for SLnpZq, n ě 3, Invent. Math. 169 (2007), no. 2,
401–425.
[4] M. Bekka, M. Cowling, and P. de la Harpe, Simplicity of the reduced C -algebra of PSLpn, Zq,
Internat. Math. Res. Notices (1994), no. 7, 285ff., approx. 7 pp. (electronic).
[5] M. B. Bekka and P. de la Harpe, Irreducibility of unitary group representations and reproducing
kernels Hilbert spaces, Expo. Math. 21 (2003), no. 2, 115–149, Appendix by the authors in
collaboration with Rostislav Grigorchuk.
[6] M. B. Bekka and N. Louvet, Some properties of C -algebras associated to discrete linear groups,
C -algebras (Munster, 1999), Springer, Berlin, 2000, pp. 1–22.
[7] V. Belyaev, R. Grigorchuk and P. Shumyatsky, On just-infiniteness of locally finite groups and
their C -algebras, Bull. Math. Sciences, to appear, arXiv:1606.07648.
[8] B. Blackadar, Shape theory for C -algebras, Math. Scand. 56 (1985), no. 2, 249–275.
[9] K. Bonvallet, B. Hartley, D. S. Passman, and M. K. Smith, Group rings with simple augmenta-
tion ideals, Proc. Amer. Math. Soc. 56 (1976), 79–82.
[10] O. Bratteli, Inductive limits of finite dimensional C -algebras, Trans. Amer. Math. Soc. 171
(1972), 195–234.
34
[11] O. Bratteli and G. A. Elliott, Structure spaces of approximately finite-dimensional C -algebras.
II, J. Funct. Anal. 30 (1978), no. 1, 74–82.
[12] N. P. Brown, On quasidiagonal C -algebras, Operator algebras and applications, Math. Soc.
Japan, Tokyo, Adv. Stud. Pure Math. 38 (2004), 19–64.
[13] N. P. Brown and N. Ozawa, C -algebras and finite-dimensional approximations, Graduate Stud-
ies in Mathematics, vol. 88, American Mathematical Society, Providence, RI, 2008.
[14] M. Dadarlat, Nonnuclear subalgebras of AF-algebras, American J. Math. 122 (2000), no. 3,
581–597.
[15] M. Dadarlat, Residually finite dimensional C -algebras and subquotients of the CAR algebra.,
Math. Res. Lett. 8 (2001), no. 4, 545–555.
[16] K. R. Davidson, C -algebras by Example, Fields Institute Monographs, Amer. Math. Soc.,
Providence, R.I., 1996.
[17] E. G. Effros, Dimensions and C -algebras, CBMS Regional Conference Series in Mathematics,
vol. 46, Amer. Math. Soc., Washington, D.C., 1981.
[18] R. Exel and T. Loring, Finite-dimensional representations of free product C -algebras, Internat.
J. Math. 3 (1992), no. 4, 469–476.
[19] R. I. Grigorchuk, On Burnside's problem on periodic groups, Funktsional. Anal. i Prilozhen. 14
(1980), no. 1, 53–54.
[20]
[21]
[22]
[23]
[24]
, Degrees of growth of finitely generated groups and the theory of invariant means, Izv.
Akad. Nauk SSSR Ser. Mat. 48 (1984), no. 5, 939–985.
, On the system of defining relations and the Schur multiplier of periodic groups generated
by finite automata, Groups St. Andrews 1997 in Bath, I, London Math. Soc. Lecture Note Ser.,
vol. 260, Cambridge Univ. Press, Cambridge, 1999, pp. 290–317.
, Just infinite branch groups, New horizons in pro-p groups, Progr. Math., vol. 184,
Birkhauser Boston, Boston, MA, 2000, pp. 121–179.
, Solved and unsolved problems around one group, Infinite groups: geometric, combina-
torial and dynamical aspects, Progr. Math., vol. 248, Birkhauser, Basel, 2005, pp. 117–218.
, Some problems of the dynamics of group actions on rooted trees, Tr. Mat. Inst. Steklova
273 (2011), no. Sovremennye Problemy Matematiki, 72–191.
[25] R. I. Grigorchuk, V. V. Nekrashevich, and V. I. Sushchanskiı, Automata, dynamical systems, and
groups, Tr. Mat. Inst. Steklova 231 (2000), no. Din. Sist., Avtom. i Beskon. Gruppy, 134–214.
[26] R. I. Grigorchuk and V. V. Nekrashevych, Self-similar groups, operator algebras and Schur
complement, J. Mod. Dyn. 1 (2007), no. 3, 323–370.
[27] F. Leinen and O. Puglisi, Cofined subgroups in periodic simple finitary linear groups, Israel J.
Math. 128 (2002), 285–324.
[28] I. G. Lysenok, A set of defining relations for the Grigorchuk group, Mat. Zametki 38 (1985),
no. 4, 503–516, 634.
[29] D. McCarthy, Infinite groups whose proper quotien groups are finite, I, Comm. Pure and Applied
Math. XXI (1968), 545–562.
[30] G. J. Murphy, C -algebras and operator theory, Academic Press, London, 1990.
[31] G. K. Pedersen, C -algebras and their automorphism groups, Academic Press, London, 1979.
[32] F. Perera and M. Rørdam, AF-embeddings into C -algebras of real rank zero, J. Funct. Anal.
217 (2004), no. 1, 142–170.
[33] L. Robert and M. Rørdam, Divisibility properties for C -algebras, Proc. Lond. Math. Soc. (3)
106 (2013), no. 6, 1330–1370.
35
[34] Y. Shalom, Bounded generation and Kazhdan's property (T), Inst. Hautes ´Etudes Sci. Publ.
Math. (1999), no. 90, 145–168 (2001).
[35] S. Sidki, A primitive ring associated to a Burnside 3-group, J. London Math. Soc. (2) 55 (1997),
no. 1, 55–64.
[36] A. C. Vieira, Modular algebras of Burnside p-groups, Mat. Contemp. 21 (2001), 287–304, 16th
School of Algebra, Part II (Portuguese) (Bras´ılia, 2000).
[37] N. Weaver, A prime C -algebra that is not primitive, J. Funct. Anal. 203 (2003), no. 2, 356–361.
Rostislav Grigorchuk
Mathematics Department
Texas A& M University
College Station, TX 77843-3368
USA
[email protected].
Magdalena Musat
Department of Mathematical Sciences
University of Copenhagen
Universitetsparken 5, DK-2100, Copenhagen Ø
Denmark
[email protected]
Mikael Rørdam
Department of Mathematical Sciences
University of Copenhagen
Universitetsparken 5, DK-2100, Copenhagen Ø
Denmark
[email protected]
36
|
1708.03471 | 2 | 1708 | 2017-08-29T13:01:33 | A bicategorical interpretation for relative Cuntz--Pimsner algebras | [
"math.OA"
] | We interpret the construction of relative Cuntz-Pimsner algebras of correspondences in terms of the correspondence bicategory, as a reflector into a certain sub-bicategory. This generalises a previous characterisation of absolute Cuntz-Pimsner algebras of proper correspondences as colimits in the correspondence bicategory. | math.OA | math |
A BICATEGORICAL INTERPRETATION FOR
RELATIVE CUNTZ -- PIMSNER ALGEBRAS
RALF MEYER AND CAMILA F. SEHNEM
Abstract. We interpret the construction of relative Cuntz -- Pimsner algebras
of correspondences in terms of the correspondence bicategory, as a reflector
into a certain sub-bicategory. This generalises a previous characterisation of
absolute Cuntz -- Pimsner algebras of proper correspondences as colimits in the
correspondence bicategory.
1. Introduction
Many important C∗-algebras may be described as (relative) Cuntz -- Pimsner
algebras, see [12, 16, 17]. These are defined by triples (A,E, J), where A is a
C∗-algebra, E is a C∗-correspondence from A to itself, that is, a Hilbert A-module
with a nondegenerate left action of A by adjointable operators, ϕ: A → B(E), and
J / A is an ideal that acts on E by compact operators, that is, ϕ(J) ⊆ K(E). The
Cuntz -- Pimsner covariance condition is only required on J.
We view the correspondence E as a generalised endomorphism of A. If E comes
from an automorphism α of A, then the relative Cuntz -- Pimsner algebra for J = A
is naturally isomorphic to the crossed product A (cid:111)α Z. So we may view Cuntz --
Pimsner algebras as analogues of crossed products for automorphisms. This is made
precise in [2] by viewing both crossed products and Cuntz -- Pimsner algebras as
colimits of diagrams in the bicategory of C∗-correspondences. The interpretation
of Cuntz -- Pimsner algebras in [2] is limited, however, to proper correspondences,
that is, ϕ(A) ⊆ K(E), and the "absolute" case J = A. This article is concerned
with another bicategorical interpretation of the Cuntz -- Pimsner algebra construction,
which needs no properness and extends to the relative case.
Our results use the equivalence between C∗-algebras with a T-action and Fell
bundles over Z, see [1]. The spectral decomposition of a T-action β on a C∗-algebra B
gives a Fell bundle (Bn)n∈N over the group Z whose section C∗-algebra C∗((Bn)n∈N)
is canonically isomorphic to B; namely,
Bn := {b ∈ B : βz(b) = zn · b}
for n ∈ Z with the multiplication, involution and norm from B. Conversely, the
section C∗-algebra of any Fell bundle over Z carries a canonical gauge action of T.
The Fell bundle underlying a Cuntz -- Pimsner algebra is semi-saturated, that is,
Bn · Bm = Bn+m if n, m ≥ 0 (or if n, m ≤ 0). Here and below, X · Y means the
closed linear span of {x · y : x ∈ X, y ∈ Y }. By the results of [1], a semi-saturated
Fell bundle is determined by its fibres B0 and B1: B0 is a C∗-algebra, B1 is a Hilbert
B0-bimodule, and the crossed product for the Hilbert B0-bimodule B1 is isomorphic
to the section C∗-algebra of the Fell bundle generated by B0 and B1.
Thus we split the construction of Cuntz -- Pimsner algebras with their canonical
T-action into two steps. The first builds the Hilbert bimodule O1
J,E, the
second takes the crossed product for this Hilbert bimodule. When we include the
J,E over O0
Key words and phrases. relative Cuntz -- Pimsner algebra; correspondence bicategory.
The second author was supported by CNPq (Brazil).
1
2
RALF MEYER AND CAMILA F. SEHNEM
J,E ,O1
gauge action, then the second step is reversible using the spectral decomposition.
This article interprets the first step in the construction as a reflector to a sub-
bicategory. A Hilbert bimodule is a C∗-correspondence with an additional left
inner product, which is unique if it exists. Thus Hilbert bimodules form a full
sub-bicategory in the correspondence bicategory. We describe a bicategory whose
objects are the triples (A,E, J) needed to define a relative Cuntz -- Pimsner algebra.
Those triples where E is a Hilbert bimodule and J is Katsura's ideal for E form
a full sub-bicategory. We show that the construction of (O0
J,E) is a reflector
onto this sub-bicategory. Roughly speaking, a reflector approximates a given object
by an object in the sub-bicategory in the optimal way. More precisely, it is a left
(bi)adjoint to the inclusion of the sub-bicategory.
We gradually work up to such bicategorical considerations. Section 2 deals
with known properties of relative Cuntz -- Pimsner algebras. We also discuss their
Fell bundle structure coming from the gauge action, and we show that the Cuntz --
Pimsner algebra OJ,E is the crossed product of its gauge-fixed point algebra O0
J,E
by the Hilbert O0
J,E. Section 2 culminates in a result about the
functoriality of relative Cuntz -- Pimsner algebras, which goes back to an idea of
Schweizer [18]. We correct his idea and extend it to the relative case by defining
proper covariant correspondences between triples (A,E, J) so that they induce
correspondences between the associated relative Cuntz -- Pimsner algebras.
J,E-bimodule O1
T. The objects of C
This construction is upgraded in Section 3 to a homomorphism of bicategories
N
pr to the T-equivariant correspondence
(or "functor") from a certain bicategory C
pr are the triples (A,E, J) needed to define a rela-
N
bicategory C
tive Cuntz -- Pimsner algebra, the arrows are the proper covariant correspondences
introduced in Section 2, and the 2-arrows are isomorphisms of covariant correspon-
dences. Whereas Schweizer reduces to ordinary categories by identifying isomorphic
correspondences, bicategories are crucial for our purposes, as in [2].
N
Then we define a sub-bicategory C
pr,∗ by restricting to Hilbert bimodules instead
of correspondences. We prove a crucial statement about covariant correspondences,
namely, that proper covariant correspondences (A,E, J) → (B,G, IG) are "equivalent"
J,E , IE) → (B,G, IG) for all (B,G, IG)
to proper covariant correspondences (O0
pr,∗, that is, for a Hilbert B-bimodule G and Katsura's ideal IG.
N
in C
Section 4 introduces the bicategorical language to understand this fact: it says
J,E ,O1
that a certain arrow
υ(A,E,J) : (A,E, J) → (O0
J,E ,O1
J,E , IE)
is a universal arrow from (A,E, J) to C
N
pr,∗. The existence of universal arrows implies
an adjunction (see [8]). So general bicategory theory upgrades the "equivalence"
pr,∗ ⊆ C
N
N
observed above to our main statement, namely, that the sub-bicategory C
pr
pr → C
N
N
is reflective and that the reflector homomorphism C
pr,∗ acts on objects
by mapping (A,E, J) to (O0
J,E). We describe this reflector in detail
pr,∗ → C
T
N
and show that its composite with the crossed product homomorphism C
T described in
is the relative Cuntz -- Pimsner algebra homomorphism C
Section 3. The definitions of bicategories, homomorphisms, transformations, and
modifications are recalled in the appendix, together with some examples related to
the correspondence bicategory.
pr,∗ ⊆ C
N
J,E ,O1
J,E , IO1
2. Preliminaries
In this section, we recall basic results on Cuntz -- Pimsner algebras, and their gauge
action and Fell bundle structure. We correct and generalise an idea by Schweizer on
the functoriality of Cuntz -- Pimsner algebras for covariant correspondences.
A BICATEGORICAL INTERPRETATION FOR CUNTZ -- PIMSNER ALGEBRAS
3
2.1. Correspondences. Let F1, F2 be Hilbert B-modules. Let B(F1,F2) be the
space of adjointable operators from F1 to F2. Let ξihη ∈ B(F1,F2) for ξ ∈ F2
and η ∈ F1 be the generalised rank-1 operator defined by ξihη(ζ) := ξhη ζiB. Let
K(F1,F2) be the closed linear span of ξihη for ξ ∈ F1 and η ∈ F2. Elements
of K(F1,F2) are called compact operators. We abbreviate B(F) := B(F,F) and
K(F) := K(F,F) if F = F1 = F2.
Lemma 2.1. Let E1 ⊆ F1 and E2 ⊆ F2 be Hilbert B-submodules. There is a unique
map K(E1,E2) → K(F1,F2) that maps ξihη ∈ K(E1,E2) to ξihη ∈ K(F1,F2) for
all ξ ∈ E2, η ∈ E1. This map is injective.
Definition 2.2. Let A and B be C∗-algebras. A correspondence from A to B is a
Hilbert B-module F with a nondegenerate left action of A through a ∗-homomorphism
ϕ: A → B(F). A correspondence is proper if ϕ(A) ⊆ K(F). It is faithful if ϕ is
injective. We write F : A (cid:32) B to say that F is a correspondence from A to B.
Definition 2.3. A Hilbert A, B-bimodule is a (right) Hilbert B-module F with
a left Hilbert A-module structure hh··iiA such that hhξ ηiiAζ = ξhη ζiB for all
ξ, η, ζ ∈ F.
If F is a Hilbert A, B-bimodule, then A acts by adjointable operators on F
and B acts by adjointable operators for the left Hilbert A-module structure, that
is, hhξb ηiiA = hhξ ηb∗iiA for all ξ, η ∈ F and all b ∈ B. In particular, E is an A, B-
bimodule. The next lemma characterises which correspondences may be enriched to
Hilbert bimodules:
Lemma 2.4 (see [7, Example 1.6]). A correspondence F : A (cid:32) B carries a Hilbert
A, B-bimodule structure if and only if there is an ideal I / A such that left action
on F restricts to a ∗-isomorphism I ∼= K(F). In this case, the ideal I and the left
inner product are unique, and I = hhF FiiA.
Definition 2.5. Let F1,F2 : A (cid:32) B be C∗-correspondences. A correspondence
isomorphism F1 ⇒ F2 is a unitary A, B-bimodule isomorphism from F1 to F2. We
write "⇒" because these isomorphisms are the 2-arrows in bicategories that we are
going to construct.
Let F be a Hilbert B-module and let ϕ: A → B(F) be a ∗-homomorphism. For
ξ ∈ E, we define an operator
Tξ : F → E ⊗ϕ F,
ξ (ζ ⊗ η) = ϕ(hξ ζi)η on elementary tensors, see [17]. Hence
It is adjointable with T ∗
ζ = ξihζ ⊗ 1,
TξT ∗
ζ Tξ = ϕ(hζ ξi),
T ∗
η 7→ ξ ⊗ η.
where ξihζ ⊗ 1 is the image of ξihζ under the canonical map B(E) → B(E ⊗ϕ F),
T 7→ T⊗1. Hence the operator Tξ for ξ ∈ E is compact if and only if ϕ(hξ ξi) = T ∗
ξ Tξ
is compact.
Lemma 2.6 ([17, Corollary 3.7]). Let J := ϕ−1(K(F)) / A and let T ∈ K(E). The
operator T ⊗ 1 on E ⊗A F is compact if and only if T ∈ K(E · J) (see Lemma 2.1
for the inclusion K(E · J) ⊆ K(E)).
In particular, if ϕ(A) ⊆ K(F), then T ⊗ 1 ∈ K(E ⊗ϕ F) for all T ∈ K(E).
over A. By convention, E⊗0 := A. Let E+ :=L∞
2.2. C∗-algebras of correspondences. Let E : A (cid:32) A be a correspondence over A.
Let ϕ: A → B(E) be the left action. Let E⊗n be the n-fold tensor product of E
n=0 E⊗n be the Fock space of E,
see [17]. Define
ξ : E⊗n → E⊗n+1,
tn
η 7→ ξ ⊗ η,
RALF MEYER AND CAMILA F. SEHNEM
4
for n ≥ 0 and ξ ∈ E; this is the operator Tξ above for F = E⊗n. The operators tn
combine to an operator tξ ∈ B(E+), that is, tξE⊗n = tn
ξ . Let ϕ∞ : A → B(E+) be
the obvious representation by block diagonal operators and let t∞ : E → B(E+) be
the linear map ξ 7→ tξ.
Definition 2.7. The Toeplitz C∗-algebra TE of E is the C∗-subalgebra of B(E+)
generated by ϕ∞(A) + t∞(E).
ξ
Let J be an ideal of A with ϕ(J) ⊆ K(E). Let P0 be the projection in B(E+) that
is the identity on A ⊆ E+ and that vanishes on E⊗n for n ≥ 1. Then J0 := ϕ∞(J)P0
is contained in TE. The ideal in TE generated by J0 is equal to K(E+J) ⊆ K(E+).
Definition 2.8 ([16, Definition 2.18]). The relative Cuntz -- Pimsner algebra OJ,E
of E with respect to J is TE /K(E+J).
IE := ϕ−1
E (K(E)) ∩ (ker ϕE)⊥,
The following three cases are particularly important. First, if J = {0}, then
OJ,E is the Toeplitz C∗-algebra TE. Secondly, if J = ϕ−1(K(E)) and ϕ is injective,
then OJ,E is the algebra OE defined by Pimsner [17]. Third, if J is Katsura's ideal
(2.9)
then OIE ,E is Katsura's Cuntz -- Pimsner algebra as defined in [12].
Proposition 2.10. Katsura's ideal IE in (2.9) is the largest ideal J in A with
ϕ(J) ⊆ K(E) for which the canonical map A → OJ,E is injective.
Proof. That πIE is injective is [12, Proposition 4.9]. The ideal IE is maximal with
this property because any ideal J / A with ϕ(J) ⊆ K(E) and J 6⊆ (ker ϕ)⊥ must
contain a ∈ J with ϕ(a) = 0. Then ϕ∞(a) ∈ ϕ∞(J) · P0 becomes 0 in OJ,E.
(cid:3)
Definition 2.11. Let E : A (cid:32) A be a correspondence and B a C∗-algebra. A
representation of E in B is a pair (π, t), where π : A → B is a ∗-homomorphism,
t: E → B is a linear map, and
(1) π(a)t(ξ) = t(ϕ(a)ξ) for all a ∈ A and ξ ∈ E;
(2) t(ξ)∗t(η) = ϕ(hξ ηiA) for all ξ, η ∈ E.
These conditions imply t(ξ)π(a) = t(ξa) for all ξ ∈ E and a ∈ A.
In particular, (ϕ∞, t∞) is a representation of E in the Toeplitz C∗-algebra TE.
This representation is universal in the following sense:
Proposition 2.12. Any representation (π, t) of E in a C∗-algebra B is of the
form (π ◦ ϕ∞, π ◦ t∞) for a unique ∗-homomorphism π : TE → B. Conversely,
(π ◦ ϕ∞, π ◦ t∞) is a representation of E for any ∗-homomorphism π : TE → B.
Lemma 2.13. For any representation (π, t) of E, there is a unique ∗-homomorphism
π1 : K(E) → B with π1(ξihη) = tξt∗
Proposition 2.14 ([16, Theorem 2.19]). The representation π of TE associated to
a representation (π, t) of E factors through the quotient OJ,E of TE if and only if
(2.15)
In this case, we call the representation covariant on J.
η for all ξ, η ∈ E.
π(a) = π1(ϕ(a))
for all a ∈ J.
Let (πJ , tJ) be the canonical representation of E in OJ,E. Proposition 2.14 says
that (πJ , tJ) is the universal representation of E that is covariant on J.
Proposition 2.16. A representation (π, t) in B is covariant on J if and only if
π(J) ⊆ t(E) · B.
A BICATEGORICAL INTERPRETATION FOR CUNTZ -- PIMSNER ALGEBRAS
5
Proof. Let a ∈ J. Then π1(ϕ(a)) is contained in the closed linear span of t(E)t(E)∗
and hence in t(E)·B. So π(a) ∈ t(E)·B is necessary for π(a) = π1(ϕ(a)). Conversely,
assume π(a) ∈ t(E) · B for all a ∈ J. We have π(a) · t(ξ) = t(ϕ(a)ξ) = π1(ϕ(a))t(ξ)
for all ξ ∈ E (see [12, Lemma 2.4]). Hence (π(a) − π1(ϕ(a))) · t(E) · B = 0. Since
π(a∗), π1(ϕ(a∗)) ∈ t(E) · B, we get (π(a) − π1(ϕ(a))) · (π(a) − π1(ϕ(a)))∗ = 0. This
(cid:3)
is equivalent to π(a) = π1(ϕ(a)).
2.3. Gauge action and Fell bundle structure. Let E : A (cid:32) A be a correspon-
dence and let J / A be an ideal with ϕ(J) ⊆ K(E). If (π, t) is a representation of E
that is covariant on J, then so is (π, z·t) for z ∈ T. This operation on representations
comes from an automorphism of the relative Cuntz -- Pimsner algebra OJ,E by its
universal property. These automorphisms define a continuous action γ of T on OJ,E,
called the gauge action. Let
On
J,E := {b ∈ OJ,E : γz(b) = znb for all z ∈ T}
and (On
J,E · Om
J,E ⊆ On+m
J,E
J,E)∗ = O−n
for n ∈ Z be the nth spectral subspace. These spectral subspaces form a Fell
bundle over Z, that is, On
J,E for all n, m ∈ Z.
In particular, for J = {0} we get a gauge action on TE and corresponding spectral
subspaces T nE ⊆ TE. Explicitly, the gauge action on TE comes from the obvious
N-grading on E+: if x ∈ TE, then x ∈ T nE if and only if x(E⊗k) ⊆ E⊗n+k for all
k ∈ N; this means xE⊗k = 0 if k + n < 0. And OnE,J is the image of T nE in OE,J.
Lemma 2.17. Let n ∈ Z. The subspace On
J,E in OJ,E is the closed linear span of
tJ(ξ1)tJ(ξ2)··· tJ(ξk) · t∗
J(η1) for ξi, ηj ∈ E, k − l = n. If n ∈ N,
then
J(ηl)··· t∗
J(η2)t∗
J,E ∼= E⊗n ⊗A O0
On
J,E
J,E. The Fell bundle (Ok
T
J,E · Ol
as a correspondence A (cid:32) O0
is, Ok
Proof. Let b ∈ On
of monomials tJ(ξ1)tJ(ξ2)··· tJ(ξk) · t∗
J,E = Ok+l
J,E if k, l ≥ 0.
J,E and let > 0. Then b is -close to a finite linear combination b
Z
J(η1) with k, l ∈ N. Define
x ∈ OJ,E .
J(ηl)··· t∗
z−nγz(x) dz,
pn(x) :=
J(η2)t∗
J,E)k∈Z is semi-saturated, that
This is a contractive projection from OJ,E onto On
we have kb − pn(b)k ≤ as well.
J(ηl)··· t∗
tJ(ξ1)tJ(ξ2)··· tJ(ξk)·t∗
Hence On
The monomials generating Ok+l
J,E is the closed linear span of such monomials with k − l = n.
J,E · Ol
J(η2)t∗
J,E for k, l ≥ 0 are obviously in Ok
J,E. Since pn(b) = b and kpnk ≤ 1,
Inspection shows that pn maps a monomial
J(η1) to itself if k−l = n and kills it otherwise.
the first statement immediately implies the last one. There is an isometric A,O0
bimodule map
J,E. Hence
J,E-
E⊗n ⊗A O0
J,E → On
J,E ,
ξ1 ⊗ ··· ⊗ ξn ⊗ y 7→ tJ(ξ1)··· tJ(ξn) · y.
The first statement implies that its image is dense, so it is unitary.
The Fell bundle (On
J,E)n∈Z need not be saturated, that is, On
J,E · O−n
J,E.
from O0
Theorem 2.18. The relative Cuntz -- Pimsner algebra is T-equivariantly isomorphic
J,E by the Hilbert O0
to the crossed product of O0
J,E and to the full or
reduced section C∗-algebra of the Fell bundle (On
Proof. The Fell bundle (On
of Abadie -- Eilers -- Exel [1] imply our claims.
J,E)n∈Z is semi-saturated by Lemma 2.17. Now the results
(cid:3)
J,E-bimodule O1
J,E)n∈Z.
(cid:3)
J,E may differ
6
RALF MEYER AND CAMILA F. SEHNEM
J,E-bimodule O1
Theorem 2.18 splits the construction of relative Cuntz -- Pimsner algebras into
two steps. The first builds the Hilbert O0
J,E, the second takes the
crossed product for this Hilbert bimodule. A Hilbert bimodule G on a C∗-algebra B
is the same as a Morita -- Rieffel equivalence between two ideals in B or, briefly, a
partial Morita -- Rieffel equivalence on B (this point of view is explained in [4]). The
crossed product O0
J,E generalises the partial crossed product for a partial
automorphism. Many results about crossed products for automorphisms extend to
Hilbert bimodule crossed products. In particular, the standard criteria for simplicity
and detection and separation of ideals are extended in [14].
Proposition 2.19. The following conditions are equivalent:
J,E (cid:111) O1
(1) the map πJ : A → O0
J,E is an isomorphism;
(2) the map ϕ: J → K(E) is an isomorphism;
(3) the correspondence E comes from a Hilbert bimodule and J = IE.
IG ,G as a Hilbert B-bimodule.
IG ,G (cid:111) O1
IG ,G ∼= G ⊗A O0
IG ,G, and the isomorphism O1
Proof. If J = IE is Katsura's ideal, then everything follows from [12, Proposi-
tion 5.18]. So it remains to observe that (1) and (2) fail if J 6= IE. Lemma 2.4 shows
that E comes from a Hilbert bimodule if and only if there is an ideal I in A so that
ϕI : I → K(E) is an isomorphism. In this case, I is the largest ideal on which ϕ
restricts to an injective map into K(E). So I = IE. Thus (2) ⇐⇒ (3).
If J 6⊆ IE, then A → OJ,E is not injective by Proposition 2.10. So (1) implies
J ⊆ IE. If J ⊆ IE and (1) holds, then the map A → OIE ,E is still surjective because
OIE ,E is a quotient of OJ,E, and it is also injective by Proposition 2.10. Hence
OIE ,E = OJ,E. This implies K(E+IE) = K(E+J) and hence IE = J because of the
direct summand A in E+.
(cid:3)
Proposition 2.20. Let G be a Hilbert B-bimodule and let IG be Katsura's ideal
for G. Then OIG ,G ∼= B (cid:111) G T-equivariantly.
Proof. Theorem 2.18 identifies OIG ,G ∼= O0
IG ,G. Proposition 2.19 gives
B ∼= O0
IG ,G from Lemma 2.17 implies
that G ∼= O1
(cid:3)
2.4. Functoriality of relative Cuntz -- Pimsner algebras. Schweizer [18] has
defined "covariant homomorphisms" and "covariant correspondences" between self-
correspondences and has asserted that they induce ∗-homomorphisms and corre-
spondences between the associated Toeplitz and absolute Cuntz -- Pimsner algebras.
For the proof of functoriality for covariant correspondences he refers to a preprint
that never got published. In fact, there are some technical pitfalls. We correct his
statement here, and also add a condition to treat relative Cuntz -- Pimsner algebras.
Throughout this subsection, let E : A (cid:32) A and G : B (cid:32) B be correspondences
and let JA ⊆ ϕ−1(K(E)) and JB ⊆ ϕ−1(K(G)) be ideals.
Definition 2.21. A covariant correspondence from (A,E, JA) to (B,G, JB) is a
pair (F, V ), where F is a correspondence A (cid:32) B with JA · F ⊆ F · JB and V is
a correspondence isomorphism E ⊗A F ⇒ F ⊗B G. A covariant correspondence is
proper if F is proper.
Proposition 2.22. A proper covariant correspondence (F, V ) from (A,E, JA) to
(B,G, JB) induces a proper T-equivariant correspondence OF ,V : OJA,E (cid:32) OJB ,G.
Schweizer [18] claims this also for non-proper correspondences, and he allows V
to be a non-adjointable isometry. In fact, a pair (F, V ) where V is only a non-
adjointable isometry induces a correspondence between the Toeplitz C∗-algebras. It
is unclear, however, when this correspondence descends to one between the absolute
or relative Cuntz -- Pimsner algebras. And we need F to be proper. Alternatively, we
may require E instead of F to be proper. This situation is treated in [2].
A BICATEGORICAL INTERPRETATION FOR CUNTZ -- PIMSNER ALGEBRAS
7
Proof. We use the canonical ∗-homomorphism πJB : B → OJB ,G to view OJB ,G
as a proper correspondence B (cid:32) OJB ,G. Thus FO := F ⊗B OJB ,G becomes
a proper correspondence A (cid:32) OJB ,G, that is, a Hilbert OJB ,G-module with a
representation π : A → K(FO). The T-action on OJB ,G induces a T-action on FO
because πJB(B) ⊆ O0
JB ,G. We are going to define a map t: E → K(FO) such
that (π, t) is a representation of (A,E) on FO that is covariant on JA. Then
Proposition 2.14 yields a representation π : OJA,E → K(FO). This is the desired
correspondence OJA,E (cid:32) OJB ,G.
There is an isometry µG : G ⊗B OJB ,G ⇒ OJB ,G, ζ ⊗ y 7→ t∞(ζ) · y, of correspon-
dences B (cid:32) OJB ,G. Usually, it is not unitary. We define an isometry
1⊗µG===⇒ F ⊗B OJB ,G = FO.
V ! : E ⊗A FO = E ⊗A F ⊗B OJB ,G V ⊗1===⇒ F ⊗B G ⊗B OJB ,G
It yields a map t from E to the space of bounded operators on FO by t(ξ)(η) := V !(ξ⊗
η). To show that t(ξ) is adjointable, we need that FO is a proper correspondence
A (cid:32) OJB ,G: then Tξ ∈ K(FO,E ⊗A FO), and composition with V ! maps this into
K(FO) by Lemma 2.1. So even t(ξ) ∈ K(FO) for all ξ ∈ E.
We claim that the pair (π, t) is a representation. We have π(a)t(ξ) = t(ϕ(a)ξ)
because V ! is a left A-module map. And t(ξ1)∗t(ξ2) = π(hξ1 ξ2i) holds because
ht(ξ1)η1 t(ξ2)η2i = hV !(ξ1 ⊗ η1) V !(ξ2 ⊗ η2)i
= hξ1 ⊗ η1 ξ2 ⊗ η2i = hη1 π(hξ1 ξ2i)η2i.
If JA = 0, then we are done at this point, and we have not yet used that V is
unitary. So the Toeplitz C∗-algebra of a correspondence remains functorial for
proper covariant correspondences where V is not unitary.
It remains to prove that π is covariant on JA. By Proposition 2.16, this is
equivalent to π(JA)(FO) ⊆ t(E)(FO). And JB · OJB ,G ⊆ tJB(G) · OJB ,G holds
because the canonical representation of (B,G) on OJB ,G is covariant on JB. Since
JA · F ⊆ F · JB by assumption,
JA · FO ⊆ F ⊗ JB · OJB ,G ⊆ F ⊗ tJB(G) · OJB ,G = (1 ⊗ µG)(F ⊗B G ⊗B OJB ,G).
Since V is unitary, we may rewrite this further as V !(E ⊗A F ⊗B OJB ,G) = t(E)·FO.
This finishes the proof that (π, t) is covariant on JA. The operators t(ξ) for ξ ∈ E
(cid:3)
are homogeneous of degree 1 for the T-action. Thus π is T-equivariant.
Example 2.23. Let A = B and J = JA = JB 6= {0} and let E ⊆ G be an A-invariant
Hilbert submodule. Then the identity correspondence F = A with the inclusion
map E ⊗A F ∼= E ,→ G ∼= G ⊗B F is a covariant correspondence in the notation of
Schweizer. There is indeed a canonical ∗-homomorphism TE → TG. But it need
not descend to the relative Cuntz -- Pimsner algebras because ϕG(a) ∈ K(G) for
a ∈ J need not be the extension of ϕE(a) ∈ K(E) given by Lemma 2.1. So the
Cuntz -- Pimsner covariance conditions for OE,J and OG,J may be incompatible. We
ask V to be unitary to avoid this problem.
Example 2.24. Turn O0
J,E with the
obvious left action of A. The proper correspondence O0
J,E with the
isomorphism from Lemma 2.17 is a proper covariant correspondence from E : A (cid:32) A
J,E : O0
with the ideal J to O1
J,E (cid:32) O0
J,E. It remains to
show that J ·O0
J,E ·IO1
J,E ⊆ O0
J,E = IO1
J,E is a Hilbert bimodule, Katsura's
ideal is equal to the range ideal of the left inner product, that is, the closed linear
span of xy∗ for all x, y ∈ O1
J,E. This contains K(E) for x, y ∈ E, which in turn
contains J by the Cuntz -- Pimsner covariance condition on J, see Proposition 2.14.
So J · O0
J,E) is
J,E, into a proper C∗-correspondence A (cid:32) O0
J,E : A (cid:32) O0
J,E with Katsura's ideal IO1
J,E. Since O1
J,E , IO1
J,E ⊆ IO1
J,E. The relative Cuntz -- Pimsner algebra of (O0
J,E ,O1
J,E by Proposition 2.19. The correspondence O0
J,E (cid:32) O0
J,E associated to the
J,E.
RALF MEYER AND CAMILA F. SEHNEM
8
again O0
covariant correspondence above is just the identity correspondence on O0
Remark 2.25. If JA = 0 or JB = ϕ−1(K(G)), then the condition JA · F ⊆ F · JB for
covariant transformations (A,E, JA) → (B,G, JB) always holds and so may be left
out. This is clear if JA = 0. Let JB = ϕ−1(K(G)). Since F is proper, JA acts on
E ⊗A F ∼= F ⊗B G by compact operators by Lemma 2.6. Again by Lemma 2.6, this
implies JA ⊆ K(F · JB). Thus JA · F ⊆ F · JB.
Example 2.26. Covariant correspondences are related to the T-pairs used by Kat-
sura [13] to describe the ideal structure of relative Cuntz -- Pimsner algebras. For this,
we specialise to covariant correspondences out of (A,E, J) where the underlying
correspondence comes from a quotient map A → A/I. That is, F = A/I : A (cid:32) A/I
for an ideal I / A. When is this part of a covariant correspondence from (A,E, J)
to (A/I,E0, J0) for some E0, J0?
There are natural isomorphisms E ⊗A F ∼= E/EI and F ⊗A/I E0 ∼= E0 as corre-
spondences A (cid:32) A/I. So the only possible choice for E0 is E0 := E/EI with a left
A/I-action which gives the canonical A-action when composed with the quotient
map A → A/I. Such a correspondence E/EI : A/I (cid:32) A/I exists if and only if E is
positively invariant, that is, IE ⊆ EI. Assume this to be the case.
An ideal J0 / A/I is equivalent to an ideal I0 / A that contains I. For a covariant
correspondence, we require JF ⊆ FJ0, which means that J ⊆ I0. And in order for
(A/I,E0, J0) to be an object of C
pr, the ideal J0 or, equivalently, I0, should act by
N
compact operators on E0 := E/EI.
It is unique up to an
automorphism of E/EI, that is, a unitary operator on E/EI that also commutes
with the left action of A or A/I, but this shall not concern us. So we get a
covariant correspondence in this case. This induces a correspondence from OJ,E
to OJ0,E0 by Proposition 2.22. Actually, our covariant correspondence is a covariant
homomorphism, and so the correspondence from Proposition 2.22 comes from a
T-equivariant ∗-homomorphism, which turns out to be surjective. So a pair of
ideals (I, I0) as above induces a T-equivariant quotient or, equivalently, a T-invariant
ideal in OJ,E.
If I0/I
contains elements that act by 0 on K(E/EI), then the map A/I → OJ0,E0 is not
injective by Proposition 2.10. Then we may enlarge I without changing the relative
Cuntz -- Pimsner algebra. When we add the condition that no non-zero element
of I0/I acts by a compact operator on E/E · I, then we get exactly the T-pairs
with J ⊆ I0 of [13]. The T-pairs (I, I0) with J ⊆ I0 correspond bijectively to
gauge-invariant ideals of OJ,E by [13, Proposition 11.9].
Sometimes different pairs (I, I0) produce the same quotient of OJ,E.
Then there is an isomorphism E ⊗A F ∼= F ⊗A E0.
3. Bicategories of correspondences and Hilbert bimodules
We are going to enrich the relative Cuntz -- Pimsner algebra construction to a
homomorphism (functor) from a suitable bicategory of covariant correspondences
to the T-equivariant correspondence bicategory. Most of the work is already done
in Proposition 2.22, which describes how this homomorphism acts on arrows. It
remains to define the appropriate bicategories and write down the remaining data
of a homomorphism.
The correspondence bicategory of C∗-algebras and related bicategories have
been discussed in [2,4 -- 6]. We recall basic bicategorical definitions in the appendix
for the convenience of the reader. Here we go through these notions much more
quickly. Let C be the correspondence bicategory. It has C∗-algebras as objects,
A BICATEGORICAL INTERPRETATION FOR CUNTZ -- PIMSNER ALGEBRAS
9
C∗-correspondences as arrows, and correspondence isomorphisms as 2-arrows. The
composition is the tensor product ⊗B of C∗-correspondences.
Given any bicategory D, there is a bicategory CD with homomorphisms D → C
as objects, transformations between these homomorphisms as arrows, and mod-
ifications between these transformations as 2-arrows (see the appendix for these
notions). There is also a continuous version of this for a locally compact, topological
bicategory D. In particular, we shall use the T-equivariant correspondence bicat-
T. Its objects are C∗-algebras with a continuous T-action. Its arrows are
egory C
T-equivariant C∗-correspondences, and 2-arrows are T-equivariant isomorphisms of
C∗-correspondences.
When D is the monoid (N, +), we may simplify the bicategory CD, see [2,
Section 5]. An object in it is equivalent to a C∗-algebra A with a self-correspondence
E : A (cid:32) A. An arrow is equivalent to a covariant correspondence (without the
condition JAF ⊆ FJB), and a 2-arrow is equivalent to an isomorphism between two
N where
N
covariant correspondences. The bicategory C
pr that we need is a variant of C
we add the ideal J and allow only proper covariant correspondences as arrows.
N
pr has the following data (see Definition A.1):
• Objects are triples (A,E, J), where A is a C∗-algebra, E : A (cid:32) A is a
C∗-correspondence, and J ⊆ ϕ−1(K(E)) is an ideal.
• Arrows (A,E, J) → (A1,E1, J1) are proper covariant correspondences (F, u)
from (A,E, J) to (A1,E1, J1), that is, F is a proper correspondence A (cid:32) A1
with JF ⊆ FJ1 and u is a correspondence isomorphism E ⊗AF ⇒ F ⊗A1 E1.
• 2-Arrows (F0, u0) ⇒ (F1, u1) are isomorphisms of covariant correspondences,
that is, correspondence isomorphisms w: F0 ⇒ F1 for which the following
diagram commutes:
Definition 3.1. The bicategory C
E ⊗A F0
1E⊗w
E ⊗A F1
u0 +
u1 +
F0 ⊗A1 E1
w⊗1E1
3 F1 ⊗A1 E1
• The vertical product of 2-arrows
w0 : (F0, u0) ⇒ (F1, u1),
w1 : (F1, u1) ⇒ (F2, u2)
pr((A,E, J), (A1,E1, J1)).
N
is the usual product w1 · w0 : F0 → F2. This is indeed a 2-arrow from
(F0, u0) to (F2, u2). And the vertical product is associative and unital.
Thus the arrows (A,E, J) → (A1,E1, J1) and the 2-arrows between them
form a category C
• Let (F, u): (A,E, J) → (A1,E1, J1) and (F1, u1): (A1,E1, J1) → (A2,E2, J2)
be arrows. Their product is (F1, u1)◦(F, u) := (F⊗A1F1, u•u1), where u•u1
is the composite correspondence isomorphism
E ⊗A F ⊗A1 F1
• The horizontal product for a diagram of arrows and 2-arrows
u⊗1F1−−−−→ F ⊗A1 E1 ⊗A1 F1
1F⊗u1
−−−−→ F ⊗A1 F1 ⊗A2 E2.
(A,E, J)
(F ,u)
w
(eF ,eu)
3(A1,E1, J1)
(F1,u1)
w1
(eF1,eu1)
3(A2,E2, J2)
3
+
+
3
+
+
3
• The unit arrow on the object (A,E, J) is the proper covariant correspondence
(A, ιE), where A is the identity correspondence, that is, A with the obvious
A-bimodule structure and the inner product hx yi := x∗y, and ιE is the
canonical isomorphism
E ⊗A A ∼= E ∼= A ⊗A E
10
RALF MEYER AND CAMILA F. SEHNEM
is the 2-arrow
(A,E, J)
(F⊗A1F1,u•u1)
w⊗w1
(eF⊗A1eF1,eu•eu1)
2 (A2,E2, J2).
This horizontal product and the product of arrows combine to composition
bifunctors
pr((A,E, J), (A1,E1, J1)) × C
N
C
pr((A1,E1, J1), (A2,E2, J2))
N
→ C
pr((A,E, J), (A2,E2, J2)).
N
built from the right and left actions of A on E.
• The associators and unitors are the same as in the correspondence bicategory.
Thus they inherit the coherence conditions needed for a bicategory.
pr → C
T that maps each object (A,E, J)
N
Theorem 3.2. There is a homomorphism C
to its relative Cuntz -- Pimsner algebra and is the construction of Proposition 2.22 on
arrows.
Proof. The construction in Proposition 2.22 is "natural" and thus functorial for
isomorphisms of covariant correspondences, and it maps the identity covariant
correspondence to the identity T-equivariant correspondence on the relative Cuntz --
Pimsner algebras. Let (F, u): (A,E, J) → (A1,E1, J1) and (F1, u1): (A1,E1, J1) →
(A2,E2, J2) be covariant correspondences and let OF ,u and OF1,u1 be the associated
correspondences of relative Cuntz -- Pimsner algebras. By definition, OF ,u ⊗OJ1 ,E1
OF1,u1 and OF⊗A1F1,u•u1 are equal to (F ⊗A1 OJ1,F1)⊗OJ1 ,F1 (F1 ⊗A2 OJ2,F2) and
(F ⊗A1 F1) ⊗A2 OJ2,F2 as T-equivariant correspondences A (cid:32) OJ2,F2. Associators
and unit transformations give a canonical T-equivariant isomorphism between
these correspondences. This isomorphism also intertwines the representations of E.
Hence it is an isomorphism of correspondences OJ,F (cid:32) OJ2,F2. These canonical
isomorphisms satisfy the coherence conditions for a homomorphism of bicategories
(cid:3)
in Definition A.5.
J,E (cid:111) O1
The relative Cuntz -- Pimsner algebra OJ,E is the crossed product O0
J,E by
Theorem 2.18. So OJ,E with the gauge T-action and the Hilbert O0
J,E-bimodule
O1
J,E contain the same amount of information. We now study the construction that
sends (A,E, J) to the Hilbert O0
J,E. The appropriate bicategory of
Hilbert bimodules is a sub-bicategory of C
pr,∗ ⊆ C
N
N
Definition 3.3. Let C
pr be the full sub-bicategory whose objects are triples
(B,G, IG), where G is a Hilbert B-bimodule and IG is Katsura's ideal for G, which
is also equal to the range ideal hhG Gii of the left inner product on G. The arrows
N
N
and 2-arrows among objects of C
pr, including the condition
pr,∗ are the same as in C
IEF ⊆ FIG for covariant correspondences.
J,E-bimodule O1
N
pr:
pr → C
N
T
When we restrict the relative Cuntz -- Pimsner algebra construction C
N
pr,∗, we get the (partial) crossed product construction for Hilbert bimodules by
to C
,
,
2
A BICATEGORICAL INTERPRETATION FOR CUNTZ -- PIMSNER ALGEBRAS
11
T.
J,E ,O1
Proposition 2.20. Thus Theorem 3.2 also completes the crossed product for Hilbert
pr,∗ → C
N
bimodules to a functor C
The map that sends (A,E, J) to (O0
pr → C
N
N
J,E) is part of a functor C
pr,∗
pr,∗ → C
T, gives the
N
which, when composed with the crossed product functor C
relative Cuntz -- Pimsner algebra functor of Theorem 3.2. We do not prove this now
because it follows from our main result. The key step is the following universal
property of (O0
Proposition 3.4. Let (A,E, J) and (B,G, IG) be objects of C
N
pr and C
tively. Let
N
pr,∗, respec-
J,E ,O1
J,E , IO1
J,E , IO1
J,E):
υ(A,E,J) : (A,E, J) → (O0
J,E ,O1
J,E , IO1
J,E)
be the covariant correspondence from Example 2.24. Composition with υ(A,E,J)
induces a groupoid equivalence
(cid:0)(A,E, J), (B,G, IG)(cid:1) ' C
pr,∗(cid:0)(O0
N
N
pr
C
J,E), (B,G, IG)(cid:1).
J,E ,O1
J,E , IO1
Recall that C
pr ((A,E, J), (A1,E1, J1)) for objects (A,E, J) and (A1,E1, J1) of C
N
N
pr
denotes the groupoid with arrows (A,E, J) → (A1,E1, J1) as objects and 2-arrows
among them as arrows.
Proof. We begin with an auxiliary construction. Proposition 2.20 identifies OIG ,G ∼=
IG ,G ∼= G, O−1
IG ,G ∼= G∗
B(cid:111)G as Z-graded C∗-algebras. In particular, O0
(B,G, IG). It induces a proper, T-equivariant correspondence OF ,V =L
as Hilbert B-bimodules. Let (F, u) be a proper covariant correspondence (A,E, J) →
n∈Z OnF ,V
from OJ,E to OIG ,G by Proposition 2.22. By construction, OnF ,V = F ⊗B On
IG ,G.
IG ,G ∼= F ⊗B G. The
Thus O0F ,V = F ⊗B O0
left action on OF ,V is a nondegenerate, T-equivariant ∗-homomorphism OJ,E →
K(OF ,V ). So O0
J,E acts on OF ,V by grading-preserving operators. Restricting to the
J,E → K(O0F ,V ) ∼= K(F).
degree-0 part, we get a nondegenerate ∗-homomorphism O0
Let F # be F viewed as a correspondence O0
IG ,G ∼= F ⊗B B ∼= F and O1F ,V = F ⊗B O1
IG ,G ∼= B and O1
J,E (cid:32) B in this way.
We now construct an isomorphism of correspondences
J,E F # ⇒ F # ⊗B G.
J,E ⊗O0
u# : O1
We need two descriptions of u#. The first shows that it is unitary, the second that
it intertwines the left actions of O0
J,E. The first formula for u# uses Lemma 2.17,
which gives unitary Hilbert B-module maps
J,E F # ∼= E ⊗A O0
J,E F # ∼= E ⊗A F.
O1
J,E ⊗O0
J,E ⊗O0
J,E ⊗O0
∼= F # ⊗B G.
J,E O0F ,V → O1F ,V
J,E F # = O1
J,E-linear because the isomorphism F # ⊗B On
Composing with u: E ⊗A F ⇒ F ⊗B G gives the desired unitary u#. The second
formula for u# restricts the left action of OJ,E on OF ,V to a multiplication map
O1
J,E ⊗O0
(3.5)
IG ,G ∼= OnF ,V is by
This is manifestly O0
right multiplication and so intertwines the left actions of O0
J,E. The map in (3.5)
maps tJ(ξ) ⊗ η 7→ u(ξ ⊗ η) for all ξ ∈ E, η ∈ F. This determines it by Lemma 2.17.
So both constructions give the same map u#.
J,E · F # ⊆ F# · IG holds, so that the pair (F #, u#) is a proper
J,E) to (B,G, IG). The ideal IO1
J,E is
J,E. Using the Fell bundle structure,
covariant correspondence from (O0
J,E , IO1
equal to the range of the left inner product on O1
we may rewrite this as O1
J,E · O−1
J,E · O−1
J,E. Thus
J,E · O0F ,V ⊆ O1
F ,V = E · O−1
F ,V .
We claim that IO1
F ,V = E · O0
J,E · O0F ,V = O1
J,E · O−1
J,E · O−1
J,E ,O1
IO1
12
The product E · O−1
Proposition 2.22. So E · O−1
RALF MEYER AND CAMILA F. SEHNEM
F ,V uses the representation of E on OF ,V built in the proof of
E ⊗A F ⊗B G∗ ∼= F ⊗B G ⊗B G∗ = F · IG.
F ,V is the image of the map
J,E , IO1
J,E ,O1
J,E ,O1
J,E) to (B,G, IG).
J,E·O0F ,V ⊆ F·IG as claimed. We have turned a proper covariant correspondence
So IO1
(F, u) from (A,E, J) to (B,G, IG) into a proper covariant correspondence (F #, u#)
from (O0
J,E ∼= O0
Conversely, take a proper covariant correspondence (F, u) from (O0
J,E)
to (B,G, IG). Composing it with υ(A,E,J) gives a proper covariant correspondence
from (A,E, J) to (B,G, IG). We now simplify this product of covariant correspon-
J,E in υ(A,E,J) is O0
dences. The underlying correspondence A → O0
J,E, and the
isomorphism E ⊗A O0
J,E = O1
J,E is the one from Lemma 2.17.
We identify the tensor product O0
J,E F with F by the canonical map. Thus
the product of (F, u) with υ(A,E,J) is canonically isomorphic to a covariant cor-
respondence (F [, u[) with underlying correspondence F [ = F : A (cid:32) B with the
left A-action through πJ : A → O0
J,E. The isomorphism u[ : E ⊗A F [ ⇒ F [ ⊗B G
J,E F ⇒ F ⊗B G with the
is the composite of the given isomorphism u: O1
isomorphism E ⊗A O0
J,E from Lemma 2.17.
Now let (F, u) be a proper covariant correspondence from (A,E, J) to (B,G, IG).
J,E O1
J,E ⊗O0
J,E ∼= O1
J,E ⊗Ou
J,E ⊗O0
J,E , IO1
J,E , IO1
J,E ,O1
(F #[, u#[) = (F, u).
J,E is the original action of A. The isomorphism E ⊗A O0
We claim that
(3.6)
By construction, the underlying Hilbert B-module of F #[ is F. We even have
F #[ = F as correspondences A (cid:32) B, that is, the left O0
J,E-action on F # composed
J,E ∼= O1
with πJ : A → O0
J,E
is used both to get u# from u and to get u#[ from u#. Unravelling this shows that
u#[ = u.
Now we claim that the map that sends a proper covariant correspondence
J,E) → (B,G, IG) to (F [, u[) is injective. This claim and (3.6)
(F, u): (O0
imply (F [#, u[#) = (F, u), that is, our two operations are inverse to each other. To
prove injectivity, we use Proposition 2.22 to build a correspondence OF ,u : OJ,E (cid:32)
OIG ,G from (F, u). This correspondence determines (F, u): we can get back F as
its degree-0 part because OIG ,G = B (cid:111) G, and because u and the left O0
J,E-module
structure on F are both contained in the left OJ,E-module structure on OF ,u.
An OJ,E-module structure on OIG ,G is already determined by a representation
of (A,E). Since On
IG ,G, this representation is determined by its
IG ,G ∼= B. And (F [, u[) determines the representation of (A,E) on B.
restriction to O0
Thus (F [, u[) determines (F, u).
The constructions of (F #, u#) and (F [, u[) are clearly natural for isomorphisms
(cid:0)(A,E, J), (B,G, IG)(cid:1) ∼= C
of covariant correspondences. So they form an isomorphism of categories
J,E), (B,G)(cid:1).
pr,∗(cid:0)(O0
IG ,G = O0
IG ,G · On
J,E ,O1
J,E , IO1
N
C
pr
N
One piece in this isomorphism is naturally equivalent to the functor that composes
(cid:3)
with υ(A,E,J). Hence this functor is an equivalence of categories, as asserted.
4. The reflector from correspondences to Hilbert bimodules
We now strengthen Proposition 3.4 using some general results on adjunctions of
homomorphisms between bicategories. We first recall the related and better known
results about ordinary categories and functors.
C(c, x) → B(b, R(x)),
f 7→ R(f) ◦ υ,
A BICATEGORICAL INTERPRETATION FOR CUNTZ -- PIMSNER ALGEBRAS
13
Let C and B be categories. Let R: C → B be a functor and b ∈ obB. An
object c ∈ obC with an arrow υ : b → R(c) is called a universal arrow from b to R
if, for each x ∈ obC and each f ∈ B(b, R(x)), there is a unique g ∈ C(c, x) with
R(g) ◦ υ = f. Equivalently, the maps
(4.1)
are bijective for all x ∈ obC. The functor R has a left adjoint L: B → C if and only
if such universal arrows exist for all x ∈ obC. The left adjoint functor L: B → C
is uniquely determined up to natural isomorphism. It maps b 7→ c on objects, and
the isomorphisms (4.1) become natural in both b and x when we replace c by L(b).
An adjunction between L and R may also be expressed through its unit and counit,
that is, natural transformations L◦ R ⇒ idC and idB ⇒ R◦ L such that the induced
transformations L ⇒ L◦R◦L ⇒ L and R ⇒ R◦L◦R ⇒ R are unit transformations.
A subcategory C ⊆ B is called reflective if the inclusion functor R: C → B has a
left adjoint L: B → C. The functor L is called reflector. The case we care about is
a bicategorical version of a full subcategory. If C ⊆ B is a full subcategory, then we
may choose L ◦ R to be the identity functor on C and the counit L ◦ R ⇒ idC to be
the unit natural transformation.
Fiore [8] carries the story of adjoint functors over to homomorphisms between
2-categories (which he calls "pseudo functors"), that is, bicategories where the
associators and unitors are identity 2-arrows. The bicategories we need are not
2-categories. But any bicategory is equivalent to a 2-category by MacLane's Coher-
ence Theorem. Hence Fiore's definitions and results apply in bicategories as well.
We shorten notation by speaking of "universal" arrows and "adjunctions" instead of
"biuniversal" arrows and "biadjunctions." A 2-category is also a category with some
extra structure. So leaving out the prefix "bi" may cause confusion in that setting.
But it will always be clear whether we mean the categorical or bicategorical notions.
Definition 4.2 ([8, Definition 9.4]). Let B and C be bicategories, R: C → B a
homomorphism, and b ∈ obB. Let c ∈ obC and let g : b → R(c) be an arrow in B.
The pair (c, g) is a universal arrow from b to R if, for every x ∈ obC, the following
functor is an equivalence of categories:
g∗ : C(c, x) → B(b, R(x)),
f 7→ R(f) · g, w 7→ R(w) • 1g.
Universal arrows are called left biliftings by Street [19].
We can now reformulate Proposition 3.4:
Proposition 4.3. Let (A,E, J) ∈ ob C
N
pr. The covariant correspondence υ(A,E,J)
J,E ,O1
J,E , IE) is a universal arrow from (A,E, J) to the inclusion
from (A,E, J) to (O0
pr,∗ → C
(cid:3)
N
N
pr.
homomorphism C
There are two alternative definitions of adjunctions, based on equivalences between
morphism categories or on units and counits. These are spelled out, respectively,
by Fiore in [8, Definition 9.8] and by Gurski in [11, Definition 2.1]. We shall use
Fiore's definition.
Definition 4.4 ([8, Definition 9.8]). Let B and C be bicategories. An adjunction
between them consists of
• two homomorphisms L: B → C, R: C → B;
• equivalences of categories
ϕb,c : C(L(b), c) ' B(b, R(c))
for all b ∈ obB, c ∈ obC;
14
RALF MEYER AND CAMILA F. SEHNEM
• natural equivalences of functors
C(L(b1), c1)
f∗
C(L(b2), c1)
g∗
/ C(L(b2), c2)
ϕb1,c1
B(b1, R(c1))
/ B(b2, R(c1))
/ B(b2, R(c2))
g∗
ϕb2,c2
f∗
for all arrows f : b2 → b1, g : c1 → c2 in B and C.
These are subject to a coherence condition. In brief, the functors ϕb,c and the
natural equivalences form a transformation between the homomorphisms
Bop × C ⇒ Cat,
(b, c) 7→ C(L(b), c), B(b, R(c)).
Here Cat is the bicategory of categories, see Example A.2.
Theorem 4.5 ([8, Theorem 9.17]). Let B and C be bicategories and let R: C → B
be a homomorphism. It is part of an adjunction if and only if there are universal
arrows from c to R for each object c ∈ obC.
More precisely, let cb ∈ obC and υb : b → R(cb) for b ∈ obC be universal arrows
from b to R. Then there is an adjoint homomorphism L: B → C that maps b 7→ cb
on objects. In particular, this assignment is part of a homomorphism of bicategories.
Theorem 4.6 ([8, Theorem 9.20]). Two left adjoints L, L0 : B ⇒ C of R: C → B are
equivalent, that is, there are transformations L ⇒ L0 and L0 ⇒ L that are inverse
to each other up to invertible modifications.
Using these general theorems, we may strengthen Proposition 3.4 (in the form of
pr,∗ ⊆ C
N
N
pr is reflective, that is, the inclusion
pr,∗ → C
pr → C
N
N
N
N
pr has a left adjoint (reflector) L: C
pr,∗. On
Proposition 4.3) to an adjunction theorem:
Corollary 4.7. The sub-bicategory C
homomorphism R: C
objects, this adjoint homomorphism maps
(A,E, J) 7→ (O0
J,E ,O1
J,E , IO1
J,E).
The homomorphism L is determined uniquely up to equivalence by Theorem 4.6.
So we have characterised the construction of relative Cuntz -- Pimsner algebras
pr,∗ ⊆ C
N
N
in bicategorical terms, as the reflector for the full sub-bicategory C
pr.
By Corollary 4.7, the relative Cuntz -- Pimsner algebra construction is part of a
homomorphism L: C
Corollary 4.8. The relative Cuntz -- Pimsner algebras OJ,E and OJ1,E1 are Morita
equivalent if there is a Morita equivalence F between E and E1 as in [15, Definition 2.1]
with J · F ⊆ F · J1.
pr → C
N
N
pr,∗. For instance, this implies the following:
The proof of Theorem 4.5 also describes the adjoint functor. We now describe
pr → C
N
N
the reflector L: C
pr,∗ explicitly, thereby explaining part of the proof of
Theorem 4.5. Much of the work in this proof is needed to check that various
diagrams of 2-arrows commute. We do not repeat these computations here.
The homomorphism L maps (A,E, J) 7→ (O0
J,E) on objects. Let
(A,E, J) and (A1,E1, J1) be objects of C
pr and let (F, u): (A,E, J) → (A1,E1, J1) be
N
proper covariant correspondences. We use the notation of the proof of Proposition 3.4
∼= O1
and write ¯ιE1 for the canonical isomorphism E1 ⊗A1 O0
O0
J1,E1
from Lemma 2.17, which is the covariance part of υ(A1,E1,J1). Let
J1,E1 ⊗O0
J,E ,O1
J,E , IO1
J1,E1
J1,E1
L(F, u): (O0
J,E ,O1
J,E , IO1
J,E) → (O0
J1,E1 ,O1
L(F, u) :=(cid:0)(F ⊗A1 O0
J1,E1, IO1
J1,E1
),
J1,E1)#, (u • ¯ιE1)#(cid:1) .
/
/
/
o
w
/
/
(F0 ⊗A1 O0
J1,E1) ⊗O0
given by the left action of O0
of Proposition 2.22.
(F1 ⊗A2 O0
J1,E1
J1,E1 on (F1 ⊗A2 O0
J2,E2) ∼= (F0 ⊗A1 F1) ⊗A2 O0
J2,E2
A BICATEGORICAL INTERPRETATION FOR CUNTZ -- PIMSNER ALGEBRAS
15
In other words, we first compose (F, u) with υ(A1,E1,J1) to get a covariant corres-
pondence (F ⊗A1 O0
) and then
apply the equivalence in Proposition 3.4.
J1,E1, u • ¯ιE1) from (A,E, J) to (O0
The construction on covariant correspondences above is clearly "natural", that
is, functorial for isomorphisms. Explicitly, L maps an isomorphism of covariant
correspondences w: (F, u) ⇒ (F0, u0) to
J1,E1,O1
J1,E1 , IO1
J1 ,E1
L(w) := (w ⊗ 1O0
J1 ,E1
)# : L(F, u) ⇒ L(F0, u0).
#
J,E ,O1
J,E ,O1
(A,E,J) : (O0
J,E , IE) → (O0
To make L a homomorphism, we also need compatibility data for units and
composition of arrows. The construction of L above maps the identity covariant
correspondence on (A,E, J) to υ
J,E , IE). This is
canonically isomorphic to the identity covariant correspondence on (O0
J,E , IE)
because the equivalence in Proposition 3.4 is by composition with υ(A,E,J). This is
the unit part in our homomorphism L.
Let (F, u): (A,E, J) → (A1,E1, J1) and (F1, u1): (A1,E1, J1) → (A2,E2, J2) be
proper covariant correspondences. Then the homomorphism L contains isomor-
phisms of covariant correspondences
(4.9)
which are natural for isomorphisms of covariant correspondences and satisfy some
coherence conditions when we compose three covariant correspondences or compose
with identity covariant correspondences. We take λ to be the isomorphism
λ(cid:0)(F, u), (F1, u1)(cid:1): L(F, u) ◦ L(F1, u1) ⇒ L(cid:0)(F, u) ◦ (F1, u1)(cid:1),
J,E ,O1
J2,E2) that is constructed in the proof
The proof of Theorem 4.5 builds λ using only the universality of the arrows υ(A,E,J).
By the equivalence of categories in Proposition 3.4, whiskering (horizontal composi-
tion) with υ(A,E,J) maps isomorphisms as in (4.9) bijectively to isomorphisms
(4.10)
L(F1, u1) ∼= (F, u) ◦ (F1, u1) ◦ υ(A2,E2,J2) and υ(A,E,J) ◦ L(cid:0)(F, u) ◦ (F1, u1)(cid:1) ∼=
The construction of L implies υ(A,E,J) ◦ L(F, u) ◦ L(F1, u1) ∼= (F, u) ◦ υ(A1,E1,J1) ◦
(cid:0)(F, u) ◦ (F1, u1)(cid:1) ◦ υ(A2,E2,J2), where we disregard associators. Hence there is
υ(A,E,J) ◦ L(F, u) ◦ L(F1, u1) ⇒ υ(A,E,J) ◦ L(cid:0)(F, u) ◦ (F1, u1)(cid:1).
a canonical isomorphism of covariant correspondences as in (4.10). This Ansatz
produces the same isomorphisms λ as above. We have now described the data of the
homomorphism L. Fiore's arguments in [8] show that it is indeed a homomorphism.
Proposition 4.11. The composite of L and the crossed product homomorphism
pr,∗ → C
N
T of Theorem 3.2.
C
Proof. Our homomorphisms agree on objects by Proposition 2.19. The proof of
Proposition 3.4 constructed the covariant correspondence (F #, u#) by taking the
degree-0 part in the correspondence constructed in the proof of Proposition 2.22.
Thus we may build a natural isomorphism between the functors in question out of
the nondegenerate left action of O0
(cid:3)
pr → C
N
T is naturally isomorphic to the homomorphism C
J1,E1 on OJ1,E1.
So the reflector L lifts the Cuntz -- Pimsner algebra homomorphism C
T to a
N
homomorphism with values in C
pr,∗. Such a lifting should exist because a Hilbert
bimodule and its crossed product with the T-action determine each other.
An adjunction also contains "natural" equivalences of categories ϕb,c : C(L(b), c) '
B(b, R(c)), where naturality is further data, see Definition 4.4. In the case at hand,
pr → C
N
16
RALF MEYER AND CAMILA F. SEHNEM
these equivalences are exactly the equivalences of categories
(cid:0)(A,E, J), (B,G, IG)(cid:1) ' C
pr,∗(cid:0)(O0
N
υ∗
(A,E,J) : C
N
pr
J,E), (B,G, IG)(cid:1).
J,E ,O1
J,E , IO1
in Proposition 3.4. Their naturality boils down to the canonical isomorphisms of
correspondences υ(A,E,J) ◦ L(F, u) ∼= (F, u) ◦ υ(A,E,J), which we have already used
above to describe the multiplicativity data λ in the homomorphism L.
Finally, we relate our adjunction to the colimit description of Cuntz -- Pimsner
algebras in [2]. Let C and D be categories. Let CD be the category of functors D → C,
which are also called diagrams of shape D in C. Identify C with the subcategory of
"constant" diagrams in CD. This subcategory is reflective if and only if all D-shaped
diagrams in C have a colimit, and the reflector maps a diagram to its colimit.
This remains true for the bicategorical colimits in [2]: by definition, the colimit
of a diagram is a universal arrow to a constant diagram. In our context, a constant
N
diagram in C
pr is an object of the form (B, B, B) that is, the Hilbert B-bimodule is
N
the identity bimodule and J = B as always for objects of C
pr,∗. Since the condition
J · F ⊆ F · B always holds, the ideal J plays no role, compare Remark 2.25.
A proper covariant correspondence (A,E, J) → (B, B, B) is equivalent to a
proper correspondence F : A (cid:32) B with an isomorphism E ⊗A F ⇒ F because
F ⊗B B ∼= F. As shown in [2], such a pair is equivalent to a representation (ϕ, t) of
the correspondence E on F that is nondegenerate in the sense that t(E) · F = F.
The properness of F means that ϕ(A) ⊆ K(F), which implies t(E) ⊆ K(F).
It is shown in [2] that all diagrams of proper correspondences of any shape have
a colimit. This is probably false for diagrams of non-proper correspondences, such
as the correspondence '2(N): C (cid:32) C that defines the Cuntz algebra O∞. The way
around this problem that we found here is to enlarge the sub-bicategory of constant
diagrams, allowing diagrams of Hilbert bimodules. In addition, we added an ideal J
to have enough data to build relative Cuntz -- Pimsner algebras.
Since the sub-bicategory C ⊆ C
N
pr of constant diagrams is contained in C
N
pr,∗, we
pr,∗ as follows. Let (A,E, J) be
N
may relate universal arrows to objects in C and C
J,E ,O1
pr. Then υ(A,E,J) : (A,E, J) → (O0
N
J,E) is a universal arrow
an object of C
J,E , IO1
N
to an object of C
pr,∗ by Proposition 4.3. The universality of υ(A,E,J) implies that a
universal arrow from (A,E, J) to a constant diagram factors through υ(A,E,J), and
J,E ,O1
that an arrow from (O0
J,E) to a constant diagram is universal if and only
if its composite with υ(A,E,J) is universal. In other words, the diagram (A,E, J) has
a colimit if and only if (O0
J,E) has one, and then the two colimits are
the same. We are dealing with the same colimits as in [2] because the ideal J in
(A,E, J) plays no role for arrows to constant diagrams.
Appendix A. Bicategories
J,E , IO1
J,E ,O1
J,E , IO1
We recall some basic definitions from bicategory theory, following [3,10]. We also
give a few examples with Sections 3 and 4 in mind.
Definition A.1. A bicategory B consists of the following data:
• a set of objects obB;
• a category B(x, y) for each pair of objects (x, y); objects of B(x, y) are
called arrows (or morphisms) from x to y, and arrows in B(x, y) are called
2-arrows (or 2-morphisms); the category structure on B(x, y) gives us a unit
2-arrow 1f on each arrow f : x → y, and a vertical composition of 2-arrows:
w0 : f0 ⇒ f1 and w1 : f1 ⇒ f2 compose to a 2-arrow w1 · w0 : f0 ⇒ f2;
• composition functors
◦: B(y, z) × B(x, y) → B(x, z)
A BICATEGORICAL INTERPRETATION FOR CUNTZ -- PIMSNER ALGEBRAS
17
for each triple of objects (x, y, z); this contains a horizontal composition of
2-arrows as displayed below:
f0
x
w0
; y
f1
g0
w1
g1
g0·f0
7→ x
; z
w1•w0
5 z.
g1·f1
• a unit arrow 1x ∈ B(x, x) for each x;
• natural invertible 2-arrows (unitors) rf : f · 1x ⇒ f and lf : 1y · f ⇒ f for
• natural isomorphisms
all f ∈ B(x, y);
B(x, y) × B(y, z) × B(z, w)
(1,◦)
B(x, y) × B(y, w)
(◦,1)
a
◦
B(x, z) × B(z, w)
◦
B(x, w);
that is, natural invertible 2-arrows, called associators,
a(f1, f2, f3): (f3 · f2) · f1 ⇒' f3 · (f2 · f1),
where f1 : x → y, f2 : y → z and f3 : z → w.
This data must make the following diagrams commute:
((f4 · f3) · f2) · f1
(f4 · f3) · (f2 · f1)
3 f4 · (f3 · (f2 · f1))
(f4 · (f3 · f2)) · f1
3 f4 · ((f3 · f2) · f1),
(f2 · 1y) · f1
f2 · (1y · f1)
f2 · f1,
where f1, f2, f3, and f4 are composable arrows, and the 2-arrows are associators
and unitors and horizontal products of them with unit 2-arrows.
We write "·" or nothing for vertical products and "•" for horizontal products.
Example A.2. Categories form a bicategory Cat with functors as arrows and
natural transformations as 2-arrows. Here the composition of morphisms is strictly
associative and unital, that is, Cat is even a 2-category.
Example A.3. A category C may be regarded as a bicategory in which the categories
C(x, y) have only identity arrows.
Example A.4. The correspondence bicategory C is defined in [6] as the bicategory with
C∗-algebras as objects, correspondences as arrows, and correspondence isomorphisms
as 2-arrows. The unit arrow 1A on a C∗-algebra A is A viewed as a Hilbert
A-bimodule in the canonical way. The A, B-bimodule structure on F provides the
unitors A ⊗A F ⇒ F and F ⊗B B ⇒ F for a correspondence F : A (cid:32) B. The
associators (E ⊗A F) ⊗B G ⇒ E ⊗A (F ⊗B G) are the obvious isomorphisms.
Definition A.5. Let B,C be bicategories. A homomorphism F : B → C consists of
• a map F : obB → obC between the object sets;
• functors Fx,y : B(x, y) → C(F 0(x), F 0(y)) for all x, y ∈ obB;
#
#
;
#
#
;
)
)
5
/
/
/
/
/
7
+
3
+
+
K
S
+
3
%
-
18
RALF MEYER AND CAMILA F. SEHNEM
• natural transformations
B(y, z) × B(x, y)
(Fy,z,Fx,y)
◦
ϕxyz
B(x, z)
Fx,z
C(F(y), F(w)) × C(F(x), F(y))
◦
C(F(x), F(z))
for all triples x, y, z of objects of B; explicitly, these are natural 2-arrows
ϕ(f1, f2): Fy,z(f2) · Fx,y(f1) ⇒ Fx,z(f2 · f1);
• 2-arrows ϕx : 1F (x) ⇒ Fx,x(1x) for all objects x of B.
This data must make the following diagrams commute:
(A.6)
(Fz,w(f3) · Fy,z(f2)) · Fx,y(f1)
ϕ(f2,f3)•1Fx,y(f1)
a0
Fz,w(f3) · (Fy,z(f2) · Fx,y(f1))
1Fz,w(f3)•ϕ(f1,f2)
Fy,w(f3 · f2) · Fx,y(f1)
ϕ(f1,f3·f2)
Fx,w((f3 · f2) · f1)
Fx,w(a)
Fz,w(f3) · Fx,z(f2 · f1)
ϕ(f2·f1,f3)
3 Fx,w(f3 · (f2 · f1));
Fx,y(f1) · Fx,x(1x) ϕ(1x,f1)
3 Fx,y(f1 · 1x)
(A.7)
1Fx,y(f1)•ϕx
Fx,y(f1) · 1F (x)
r0
Fx,y(f1)
Fy,y(1y) · Fx,y(f1) ϕ(f1,1y)
(A.8)
ϕy•1Fx,y(f1)
1F (y) · Fx,y(f1)
l0
Fx,y(f1)
Fx,y(rf1 )
Fx,y(f1);
3 Fx,y(1y · f1)
Fx,y(lf1 )
Fx,y(f1).
Example A.9. A semigroup P may be viewed a category with one object and P as its
set of arrows. It may be viewed as a bicategory as well as in Example A.3. A homo-
morphism from P to C is equivalent to an essential product system (A, (Ep)p∈P op, µ)
over P op as defined by Fowler [9]. The condition (A.6) says that the multiplication
'→ Eqp are associative. The conditions (A.7) and (A.8) mean
maps µp,q : Ep ⊗A Eq
that µ1,p(a ⊗ ξ) = ϕp(a)ξ and µp,1(ξ ⊗ a) = ξa for a ∈ A, ξ ∈ Ep.
A morphism f : x → y in a bicategory B induces functors
f∗ : B(c, x) → B(c, y),
f∗ : B(y, c) → B(x, c)
for c ∈ obB by composing arrows with f and composing 2-arrows horizontally
with 1f on one side (this is also called whiskering with f).
Definition A.10. Let F, G: B ⇒ C be homomorphisms. A transformation α: F ⇒
G consists of
• morphisms αx : F(x) → G(x) for all x ∈ obB;
• natural transformations
B(x, y)
Fx,y
/ C(F(x), F(y))
Gx,y
C(G(x), G(y))
αx,y
∗
αx
αy∗
/ C(F(x), G(y)),
/
/
/
/
/
7
+
3
+
+
+
3
K
S
+
+
3
K
S
/
p
x
/
A BICATEGORICAL INTERPRETATION FOR CUNTZ -- PIMSNER ALGEBRAS
that is, 2-arrows αx,y(f): αyFx,y(f) ⇒ Gx,y(f)αx for all x, y ∈ obB.
This data must make the following diagrams commute:
19
αz(Fy,z(g)Fx,y(f))
1•ϕF (f,g) +
3 αzFx,z(gf)
αx,z(gf)
3 Gx,z(gf)αx
(αzFy,z(g))Fx,y(f)
αy,z(g)•1
(Gy,z(g)αy)Fx,y(f) k
ϕG(f,g)•1
(Gy,z(g)Gx,y(f))αx
Gy,z(g)(αyFx,y(f))1•αx,y(f)+
3 Gy,z(g)(Gx,y(f)αx).
αxFx,x(1x)
1αx•ϕF
x
αx1F (x)
r
αx,x(1x)
Gx,x(1x)αx
x •1αx
ϕG
3 αx
l−1
1G(x)αx;
Example A.11. Let G be a group. A transformation between homomorphisms G → C
consists of a correspondence F : A (cid:32) B and isomorphisms αs : Es ⊗A F ' F ⊗B Gs
so that the following diagrams commute for all s, t ∈ G:
(Es ⊗A Et) ⊗A F
s,t⊗1
w1
3 Est ⊗A F
αst
3 F ⊗B Gst
1⊗w2
s,t
F ⊗B (Gs ⊗B Gt)
Es ⊗A (Et ⊗A F)
1⊗αt
Es ⊗A (F ⊗B Gt) k
(Es ⊗A F) ⊗B Gt
αs⊗1
3 (F ⊗B Gs) ⊗B Gt.
This is called a correspondence of Fell bundles (see [6, Proposition 3.23]).
Definition A.12. Let α, β : F ⇒ G be transformations between homomorphisms.
A modification ∆: α (cid:86) β is a family of 2-arrows ∆x : αx ⇒ βx such that for every
2-arrow w: f1 ⇒ f2 for arrows f1, f2 : x → y, the following diagram commutes:
αyFx,y(f1)
∆y•Fx,y(w)
3 βyFx,y(f2)
αx,y(f1)
βx,y(f2)
Gx,y(f1)αx
Gx,y(w)•∆x +
3 Gx,y(f2)βx
References
[1] Beatriz Abadie, Søren Eilers, and Ruy Exel, Morita equivalence for crossed products by Hilbert
C∗-bimodules, Trans. Amer. Math. Soc. 350 (1998), no. 8, 3043 -- 3054, doi: 10.1090/S0002-
9947-98-02133-3. MR 1467459
[2] Suliman Albandik and Ralf Meyer, Colimits in the correspondence bicategory, Münster J.
Math. 9 (2016), 51 -- 76, doi: 10.17879/45209432019. MR 3549542
[3] Jean Bénabou, Introduction to bicategories, Reports of the Midwest Category Seminar,
Springer, Berlin, 1967, pp. 1 -- 77, doi: 10.1007/BFb0074299. MR 0220789
Math. 47 (2017), no. 1, 53 -- 159, doi: 10.1216/RMJ-2017-47-1-53. MR 3619758
[4] Alcides Buss and Ralf Meyer, Inverse semigroup actions on groupoids, Rocky Mountain J.
[5] Alcides Buss, Ralf Meyer, and Chenchang Zhu, Non-Hausdorff symmetries of C∗-algebras,
, A higher category approach to twisted actions on C∗-algebras, Proc. Edinb. Math.
Math. Ann. 352 (2012), no. 1, 73 -- 97, doi: 10.1007/s00208-010-0630-3. MR 2885576
[6]
Soc. (2) 56 (2013), no. 2, 387 -- 426, doi: 10.1017/S0013091512000259. MR 3056650
K
S
+
K
S
s
+
3
K
S
k
s
+
k
s
K
S
+
+
K
S
s
+
3
+
K
S
+
20
RALF MEYER AND CAMILA F. SEHNEM
[7] Siegfried Echterhoff, Steven P. Kaliszewski, John Quigg, and Iain Raeburn, A categorical
approach to imprimitivity theorems for C∗-dynamical systems, Mem. Amer. Math. Soc. 180
(2006), no. 850, viii+169, doi: 10.1090/memo/0850. MR 2203930
[8] Thomas M. Fiore, Pseudo limits, biadjoints, and pseudo algebras: categorical founda-
tions of conformal field theory, Mem. Amer. Math. Soc. 182 (2006), no. 860, x+171, doi:
10.1090/memo/0860. arXiv: math/0408298. MR 2229946
[9] Neal J. Fowler, Discrete product systems of Hilbert bimodules, Pacific J. Math. 204 (2002),
no. 2, 335 -- 375, doi: 10.2140/pjm.2002.204.335. MR 1907896
[10] John W. Gray, Formal category theory: adjointness for 2-categories, Lecture Notes in Math-
ematics, vol. 391, Springer-Verlag, Berlin-New York, 1974. doi: 10.1007/BFb0061280 MR
0371990
available at http://www.tac.mta.ca/tac/volumes/26/14/26-14abs.html. MR 2972968
[11] Nick Gurski, Biequivalences in tricategories, Theory Appl. Categ. 26 (2012), No. 14, 349 -- 384,
[12] Takeshi Katsura, On C∗-algebras associated with C∗-correspondences, J. Funct. Anal. 217
, Ideal structure of C∗-algebras associated with C∗-correspondences, Pacific J. Math.
(2004), no. 2, 366 -- 401, doi: 10.1016/j.jfa.2004.03.010. MR 2102572
[13]
230 (2007), no. 1, 107 -- 145, doi: 10.2140/pjm.2007.230.107. MR 2413377
[16]
[14] Bartosz Kosma Kwaśniewski and Ralf Meyer, Aperiodicity, topological freeness and pure out-
erness: from group actions to Fell bundles, Studia Math. (2016), accepted. arXiv: 1611.06954.
[15] Paul S. Muhly and Baruch Solel, On the Morita equivalence of tensor algebras, Proc. London
, Tensor algebras over C∗-correspondences:
Math. Soc. (3) 81 (2000), no. 1, 113 -- 168, doi: 10.1112/S0024611500012405. MR 1757049
representations, dilations, and
C∗-envelopes, J. Funct. Anal. 158 (1998), no. 2, 389 -- 457, doi: 10.1006/jfan.1998.3294. MR
1648483
[17] Mihai V. Pimsner, A class of C∗-algebras generalizing both Cuntz -- Krieger algebras and
crossed products by Z, Free probability theory (Waterloo, ON, 1995), Fields Inst. Commun.,
vol. 12, Amer. Math. Soc., Providence, RI, 1997, pp. 189 -- 212. MR 1426840
[18] Jürgen Schweizer, Crossed products by C∗-correspondences and Cuntz -- Pimsner algebras,
C∗-Algebras (Münster, 1999), Springer, Berlin, 2000, pp. 203 -- 226, doi: 10.1007/978-3-642-
57288-3. MR 1798598
[19] Ross Street, Fibrations in bicategories, Cahiers Topologie Géom. Différentielle 21 (1980),
no. 2, 111 -- 160. MR 574662
E-mail address: [email protected]
E-mail address: [email protected]
Mathematisches Institut, Georg-August-Universität Göttingen, Bunsenstrasse 3 -- 5,
37073, Göttingen, Germany
|
1107.3965 | 1 | 1107 | 2011-07-20T13:08:24 | An Ergodic Dilation of Completely Positive Maps | [
"math.OA"
] | We shall prove the following Stinespring-type theorem: there exists a triple $(\pi,\mathcal{H},\mathbf{V})$ associated with an unital completely positive map $\Phi:\mathfrak{A}\rightarrow \mathfrak{A}$ on C* algebra $\mathfrak{A}$ with unit, where $\mathcal{H}$ is a Hilbert space, $\pi:\mathfrak{A\rightarrow B}(\mathcal{H})$ is a faithful representation and $\mathbf{V}$ is a linear isometry on $\mathcal{H}$ such that $\pi(\Phi(a)=\mathbf{V}^*\pi(a)\mathbf{V}$ for all $a$ belong to $\mathfrak{A}$. The Nagy dilation theorem, applied to isometry $\mathbf{V}$, allows to construct a dilation of ucp-map, $\Phi$, in the sense of Arveson, that satisfies ergodic properties of a $\Phi $-invariante state $\phi$ on $\mathfrak{A}$, if $\Phi$ admit a $\phi $-adjoint. | math.OA | math |
An Ergodic Dilation of Completely Positive Maps
Carlo Pandiscia
Abstract
We shall prove the following Stinespring-type theorem: there exists a triple (π, H, V) associated
with an unital completely positive map Φ : A → A on C*-algebra A with unit, where H is a Hilbert
space, π : A → B(H) is a faithful representation and V is a linear isometry on H such that π(Φ(a) =
π(a)V for all a belong to A. The Nagy dilation theorem, applied to isometry V, allows to construct
V∗
a dilation of ucp-map, Φ, in the sense of Arveson, that satisfies ergodic properties of a Φ-invariante
state ϕ on A, if Φ admit a ϕ-adjoint.
1 Introduction
A discrete quantum process is a pair (M, Φ) consisting of a von Neumann algebra M and a normal unital
completely positive map Φ on M. In this work we shall prove that any quantum process is possible dilate
to quantum process where the dynamic Φ is a *-endomorphism of a larger von Neumann algebra.
In dynamical systems, the process of dilation has taken different meanings. Here we adopt the following
definition (See Ref. Muhly-Solel [6]):
Suppose M acts on Hilbert space H, a dilation of a quantum process (M, Φ) is a quadruple (R, Θ, K, z)
where (R, Θ) is a quantum process with R acts on Hilbert space K and Θ is a homomorphism (i.e.
*-endomorphism on von Neumann algebra R) with z : H → K isometric embedding such that:
• zMz∗ ⊂ R and z∗Rz ⊂ M;
• Φn(a) = z∗Θn(zaz∗)z
for all a ∈ M and n ∈ N;
• z∗Θn(X)z = Φn(z∗Xz) for all X ∈ R and n ∈ N.
Many authors in the past have been applied to problems very similar to the one we described above.
We remember the work of Arveson [2] on the Eo-semigroups, of Baht-Parthasarathy on the dilations of
nonconservative dynamical semigroups [3] and finally, the most recent work of Mhulay-Solel [6].
We shall prove the existence of dilation using the Nagy theorem for linear contraction (See Fojas-Nagy
Ref.[7]) and of a particular covariat representation obtained through the Stinespring's theorem for com-
pletely positive maps (See Stinespring Ref.[10]).
We recall that a covariant representation of discrete quantum process (M, Φ) is a triple (π, H, V) where
π : M → B(H) is a normal faithful representation on the Hilbert space H and V is an isometry on H
such that for a ∈ M and a ∈ N,
π(Φn(a)) = Vn∗π(a)Vn.
Since the covariant representation is faithful and normal, we identify the von neuman algebra M with
π(M) and in sec. 3 we construct a dilation of the quantum process (π(M), Ψ) where Ψ is the following
completely positive map Ψ(π(x)) = π(Φ(x)) for all n ∈ M.
In fact, if the triple (bV,bH, z) is the minimal unitary dilation of isometry V, we can construct a von
Neumann algebrascM ⊂ B(bH) with following properties: bV∗cMbV ⊂cM and z∗cMz = M.
Of fundamental importance to quantum process theory, is the ϕ-adjointness properties. The dynamic Φ
admit a ϕ-adjoint (See Kummerer Ref.[4]) relative to the normal Φ-invariant state ϕ on M, if there is a
normal unital completely positive map Φ♮ : M → M such that for a, b ∈ M,
ϕ(Φ(a)b) = ϕ(aΦ♮(b)).
1
The relationship between reversible process, modular operator and ϕ-adjointness has been studied by
Accardi-Cecchini in [1] and Majewski in [5].
In sec. 4 we shall prove that our dilation satisfies ergodic properties of a Φ-invariante state ϕ on M if
the dynamic Φ admit a ϕ-adjoint.
More precisely, let (R, Θ) be our dilation of quantum process (M, Φ), we shall prove that if
lim
n→∞
1
n + 1
nXk=0
ϕ(aΦk(b)) − ϕ(a)ϕ(b) = 0,
for all a, b ∈ M, we have
lim
n→∞
1
n + 1
nXk=0
ϕ(z∗XΘk(Y )z) − ϕ(z∗Xz)ϕ(z∗Y z) = 0,
for all X, Y ∈ R.
For generality, we will work with concrete unital C*-algebras A and unital completely positive map Φ
(briefly ucp-map). The results obtained are easily extended to the quantum process (M, Φ).
Before introducing the proof about existence of dilation of discrete quantum process, it is necessary to
recall the fundamental Nagy dilation theorem, subject of the next section.
2 Nagy dilation theorem
If V is a linear isometry on Hilbert space H, there is a triple (bV,bH, Z) where bH is a Hilbert space,
Z : H→bH is a lineary isometry, while bV is an unitary operator on bH such that for n ∈ N,
(1)
with the following minimal properties:
bVnZ = ZVn,
bH = _k∈ZbVkZH.
bVn =(cid:12)(cid:12)(cid:12)(cid:12)
Vn C(n)
0 Wn (cid:12)(cid:12)(cid:12)(cid:12) ,
2
For our purposes it is useful to recall here the structure of the unitary minimal dilation of a contraction
(See Fojas-Nagy Ref.[7]).
Let K be a Hilbert space, by l2(K) we denote the Hilbert space {ξ : N → K : Pn≥0
We now get the orthogonal projection F = I − VV∗ and the following Hilbert space bH = H ⊕ l2(FH)
and define the following unitary operator on the Hilbert space bH:
0 W (cid:12)(cid:12)(cid:12)(cid:12) ,
where for each j ∈ N we have set with Πj : l2(FH) → H the canonical projections:
bV =(cid:12)(cid:12)(cid:12)(cid:12)
kξ(n)k2 < ∞}.
V FΠ0
while W : l2(FH) → l2(FH) is the linear operator
Πj(ξ0, ξ1...ξn...) = ξj ,
W(ξ0, ξ1...ξn...) = (ξ1, ξ2...),
for all (ξ0, ξ1...ξn...) ∈ l2(FH).
If Z : H → bH is the isometry defined by ZΨ = Ψ ⊕ 0 for all Ψ ∈ H, it's simple to prove that the
relationships 1 and 2 are given.
We observe that for each n ∈ N we have
(2)
(3)
where C(n) : l2(FH) → H are the following operators:
C(n) =
Furthermore, for each n, m ∈ N we obtain:
Vn−j FΠj−1,
n ≥ 1.
nPj=1
ΠnWm = Πn+m and ΠnWm∗
since
Wm∗(ξ0, ξ1...ξn..) = (0, 0....0,
while for each k and p natural number, we obtain:
0
ΠpC(k)∗ =(cid:26) FV(k−p−1)∗
}
FV(k−1)∗
k−time
z
Ψ......FV
C(k)∗Ψ = (
k > p
elsewhere
∗
Ψ, FΨ, 0, .0..).
{
since
for all Ψ ∈ H.
0
n < m
=(cid:26) Πn−m n ≥ m
z}{ξ0 , ξ1...),
m+1
,
(4)
(5)
(6)
3 Invariant algebra
Let be A ⊂ B(H) a C*-algebras with unit and V an isometry on Hilbert space H such that
V∗AV ⊂ A.
sition:
If (bV,bH, Z) denotes the minimal unitary dilation of the isometry V we shall prove the following propo-
Proposition 1. There exists a C*-algebra with unit bA ⊂ B(bH) such that:
1 - ZAZ∗ ⊂bA and Z∗bAZ ⊂ A,
2 - bV∗bAbV ⊂bA,
3 - Z∗bV∗XbVZ = V∗Z∗XZV,
4 - Z∗bV∗(ZAZ∗)bV = V∗AV,
first of all we want to consider some special operators on Hilbert space H.
for all X ∈bA,
for all A ∈ A.
3.1 The gamma operators associated to pair (A, V )
The sequences of elements of type α = (n1, n2....nr, A1, A2...Ar), with nj ∈ N and Aj ∈ A for all
j = 1, 2...r, are called strings of A of length r and weight
ni.
For each α string of A, we associate the following operators of B(H):
nPi=1
α) = A1Vn1 · · · ArV and
(α = Vn∗
r Ar · · · Vn∗
1 A1,
ni and l(α) = r, while n) denote the set operators α) with
·
α = n and usually
furthermore
.
α =
nPi=1
n)A =nα)A : A ∈ A and α-string of A with
·
α = no .
The symbols (n and A (n have the same obvious meaning of above.
3
Proposition 2. Let α and β are strings of A for each R ∈ A we have:
and with a simple calculation
Proof. For each m, n ∈ N and R ∈ A we have:
·
α ≥
·
α <
·
β
·
β
,
·
·
if
α −
·
β −
β(cid:12)(cid:12)(cid:12)(cid:12)
α(cid:19) A if
β(cid:19) .
(α R β) ∈
A(cid:18) ·
(cid:12)(cid:12)(cid:12)(cid:12)
α) R β) ∈(cid:12)(cid:12)(cid:12)(cid:12)
RVn∈(cid:26) V(m−n)∗
Vm∗
·
α +
·
A m ≥ n
AV(n−m) m < n
(7)
(8)
(9)
·
·
·
α <
Let α = (m1, m2....mr, A1, A2...Ar) and β = (n1, n2....ns, B1, B2...BS) strings of A, we obtain:
eα <
r Ar · · · Vm∗
(α R β) = Vm∗
·
β it follows that
while if
In fact if m1 ≥ n1 we obtain:
1 A1RB1Vn1 · · · BsVns = (eα I(cid:12)(cid:12)(cid:12)eβ(cid:17)
where eα and eβ are strings of A with l (eα) + l(cid:16)eβ(cid:17) = l (α) + l (β) − 1. Moreover if
R1B2Vn2 · · · BsVns = (eα I(cid:12)(cid:12)(cid:12)eβ(cid:17) ,
2 A2R1V(n1−m1)B2 · · · BsVns = (eα I(cid:12)(cid:12)(cid:12)eβ(cid:17) ,
1 A1RB1Vn1 , eα = (m1 − n1, m2....mr, R1, A2...Ar) and eβ = (n2....ns, B2...BS).
1 A1RB1Vm1 , eα = (m2....mr, A2...Ar) and eβ = (n1 − m1, n2....ns, R1, B2...BS).
where R1 = Vm∗
Then by induction on number ν = l(α) + l(β) we have the relationship 7.
where R1 = Vn∗
If m1 < n1 we can write:
r Ar · · · A2V(m1−n1)∗
(α R β) = Vm∗
(α R β) = Vm∗
r Ar · · · Vm∗
eβ.
·
α ≥
·
α ≥ 1, we define the linear operators:
For each α string of A with
·
β we have
·
eα ≥
·
eβ
Γ(α) = (αFΠ ·
α−1
,
that will be the gamma associated operators to the pair (A, V).
Proposition 3. For each α and β strings of A with
satisfy the following relationship:
·
α,
·
β ≥ 1, the gamma operators associated to (A, V)
Γ(α) · Γ(β)∗ ∈ A.
Proof. We obtain:
in fact
since we have (α V ∈(cid:16) ·
∗
,
0
Π
α−1
·
β−1
·
β
·
β
(α F β)
·
α =
·
α 6=
Γ(α) · Γ(β)∗ = (αFΠ ·
(α F β) = (α (I − VV∗) α) = (α I α) − (α VV∗ α) ∈ A,
F β) =
α − 1(cid:12)(cid:12)(cid:12) while V∗ α) ∈(cid:12)(cid:12)(cid:12)
α − 1(cid:17) and by relationship 7 follows that:
α − 1(cid:12)(cid:12)(cid:12) I(cid:12)(cid:12)(cid:12)
(cid:16) ·
α − 1(cid:17) ⊂ A.
·
·
4
We have an operator system Σ of B(l2(FH)) this is:
Σ =(cid:8)T ∈ B(l2(FH)) : Γ1TΓ∗
We observe that I ∈ Σ and Γ∗
2 ∈ A f or all gamma operators Γi associated to (A, V(cid:9) .
1AΓ2 ∈ Σ for all gamma operators Γi. Moreover Σ is a norm closed,
(10)
while it is a weakly closed if A is a W*-algebra.
3.2 The napla operators
For each α, β strings of A, A ∈ A and k ∈ N we define the napla operators of B(l2(FH)):
∆k(A, α, β) = Π
∗
·
α+k
Fα)A(βFΠ ·
β+k
.
For each h, k ≥ 0 we obtain the following results:
∆k(A, α, β)∗ = ∆k(A∗, β, α),
and
∆k(A, α, β)·∆h(B, γ, δ) =
In fact we have:
0
∆k (R, α, ϑ) k +
∆h (R, ϑ, δ)
k +
.
β = h +
.
β = h +
k +
.
γ, h − k ≥ 0, with
.
γ, k − h > 0, with
.
.
β 6= h +
γ,
.
ϑ =
.
ϑ =
.
δ + h − k and R ∈ A
.
δ + k − h and R ∈ A
(11)
∆k (A, α, β) · ∆h (B, γ, δ) = Π
∗
·
α+k
F α) A (β FΠ ·
β+k
Π
·
γ+h
∗
F γ) B (δ FΠ ·
δ+h
and if k +
.
β 6= h +
.
γ follows that Π ·
∗
Π
get h ≥ k, and we obtain
.
β =
.
γ + h − k ≥
= 0, while if k +
·
γ+h
.
γ . Moreover by relationship 7
.
β = h +
β+k
.
γ, without losing generality we can
β −
(β F γ) ∈ A(cid:16) .
A (β F γ) B (δ ∈ A(cid:18) ·
.
γ(cid:12)(cid:12)(cid:12)
.
β −
δ +
·
δ +
.
β −
.
γ and a R ∈ A such that:
.
γ(cid:12)(cid:12)(cid:12)(cid:12) ,
then
there exists ϑ string of A with
·
ϑ =
Since
·
ϑ =
·
δ + h − k we have:
∆k (A, α, β) · ∆h (B, γ, δ) = Π
∗
·
α+k
A (β F γ) B (δ = R (ϑ .
F α) R (ϑ FΠ ·
δ+h
∗
= Π
·
α+k
F α) R (ϑ FΠ ·
ϑ+k
= ∆k (R, α, ϑ) .
Proposition 4. The linear space Xo generated by napla operators, is a *-subalgebra of B(cid:0)l2 (FH)(cid:1)
included in the operator systems Σ defined in 10.
Proof. From relationship 11 the linear space Xo is a *-algebra. Moreover for each gamma operators Γ (α)
and Γ (β) we obtain:
Γ (α) ∆k (A, γ, δ) Γ (β)∗ = (α FΠ ·
α−1
since by the relationships 7 and 8 we have
∗
Π
·
γ+k
F γ) A (δ FΠ ·
δ+k
Π ·
β−1
F β) ∈ A,
(α FΠ ·
α−1
Π
∗
·
γ+k
F γ) A (δ FΠ ·
δ+k
Π ·
β−1
F β) ∈(cid:26) (k + 1 A k + 1)
0
.
α − 1 =
.
γ + k,
.
β − 1 =
.
δ + k
elsewhere
5
In fact if
.
α =
.
γ + k + 1 we can write:
(α FΠ ·
α−1
since
∗
Π
·
γ+k
F γ) = (α F γ) = (α I γ) − (α VV∗ γ) ∈ A (k + 1
while if
.
β =
.
δ + k + 1 we obtain
(α I γ) ∈ A (k + 1 and (α VV∗ γ) ∈ A (k + 1
(δ FΠ ·
δ+k
Π ·
β−1
F β) ∈ (k + 1 A.
Corollary 1. The *-algebra Xo and the operator systems Σ are W-invariant:
W∗XoW ⊂ Xo and W∗ΣW ⊂ Σ.
Proof. Let be T belong to Σ, for each gamma operators Γ (α) and Γ (β) we have:
Γ (α) (W∗TW) Γ (β)∗ = (α FΠ ·
α−1
W∗TWΠ ·
β−1
F β) =
= (α FΠ ·
α−2
TΠ ·
β−2
F β) ∈ AV∗Γ1 (αo) TΓ2 (βo) VA ⊂ V∗AV ⊂A.
where αo and βo are strings of A with
In fact let α = (m1, m2....mr, A1, A2...Ar) by definition of gamma operator, there is i ≤ r with mi ≥ 1
such that
.
αo =
.
α − 1 and
.
βo =
.
β − 1.
(α FΠ ·
α−2 =A1 · · · ·AiV∗(αo FΠ ·
α−2 =A1 · · · ·AiV∗Γ (αo) ,
where αo = (0, ..0, mi − 1, mi+1..mr, A1, A2...Ar) with
·
αo =
·
α − 1.
Let X be the closure in norm of the *-algebra Xo. Since Σ is a norm closed set, we have X ⊂ Σ while
if A is a von Neumann algebra of B (H) then Σ is weakly closed and X′′
o ⊂ Σ.
Proposition 5. The set
S =(cid:26)(cid:12)(cid:12)(cid:12)(cid:12)
A Γ1
Γ∗
2 T (cid:12)(cid:12)(cid:12)(cid:12) : A ∈ A, T ∈ X and Γi are gamma op.of (A, V)(cid:27) ,
(12)
is an operator system of B(cid:16)bH(cid:17) such that:
Furthermore
where C ∗ (S) is the C*-algebra generated by the set S.
bV∗SbV ⊂ S.
bV∗C ∗ (S)bV ⊂ C ∗ (S) ,
Proof. We obtain:
bV∗SbV =(cid:12)(cid:12)(cid:12)(cid:12)
V∗AV
C (1)∗ AV + W∗Γ∗
2V C (1)∗ AC (1) + W∗Γ∗
V∗AC (1) + V∗Γ1W
2C (1) + C (1)∗ Γ1W + W∗TW (cid:12)(cid:12)(cid:12)(cid:12) ,
where the operators V∗Γ (α) W and V∗AC (1) are gamma operators associated to pair (A, V), while
C (1)∗ AC (1), C (1)∗ Γ (α) W, and W∗T W are operators belonging to X.
In fact we have the following relationships:
V∗AC (1) = V∗AFΠ0 = Γ (ϑ) with ϑ = (1, A) .
6
while if α = (m1, m2..mr, A1, A2...Ar) we obtain:
V∗Γ (α) W = V∗ (α FΠ ·
α−1
W =Γ (ϑ) ,
with ϑ = (m1 + 1, m2..mr, A1, A2...Ar) since Π ·
Furthermore
W = Π ·
α
.
α−1
C (1)∗ AC (1) = Π
∗
0FAFΠ0 = ∆0 (A, α, β) with α = β = (0, I)
while
C (1)∗ Γ (α) W = Π
∗
0F (α FΠ ·
α−1
∗
W = Π
0F γ) (α FΠ ·
α+0
= ∆0 (I, γ, α) with γ = (0, I) .
We observe that the *-algebra A∗ (S) generated by the operator system S is given by
Now we can easily prove proposition 1.
A∗ (S) =(cid:12)(cid:12)(cid:12)(cid:12)
A
XΓ∗A
AΓX
X (cid:12)(cid:12)(cid:12)(cid:12) .
(13)
Proof. We get C ∗ (S), the C*-algebra generated by S defined in 12, by the definition ZAZ∗ ⊂ S then
Moreover for X ∈ C ∗(S) we have:
Z∗C ∗ (S) Z ⊂ A.
Z∗bV∗XbVZ = VZ∗XZV,
family F with inclusion is partially ordered set, then for Zorn lemma's exists a minimal element that we
since bVZ = ZV.
Let be F the family of C*-subalgebras bB with unit of C ∗(S) such that ZAZ∗ ⊂ bB and bV∗bBbV ⊂ bB. The
shall denote with bA.
4 Stinespring's theorem and dilations
We examine a concrete C*-algebra A of B(H) with unit and an ucp-map Φ : A → A. By the Stinespring
theorem for the ucp-map Φ, we can deduce a triple (VΦ, σΦ, LΦ) constituted by a Hilbert space LΦ, a
representation σΦ : A → B(LΦ) and a linear contraction VΦ : H → LΦ such thata for ∈ A,
Φ(a) = V∗
ΦσΦ(a)VΦ.
(14)
We recall that on the algebraic tensor A ⊗ H we can define a semi-inner product by
ha1 ⊗ Ψ1, a2 ⊗ Ψ2iΦ = hΨ1, Φ (a∗
1a2) Ψ2iH ,
for all a1, a2 ∈ A and Ψ1, Ψ2 ∈ H furthermore the Hilbert space LΦ is the completion of the quotient
space A⊗ΦH of A ⊗ H by the linear subspace
{X ∈ A ⊗ H : hX, XiΦ = 0}
with inner product induced by h·, ·iΦ. We shall denote the image at a ⊗ Ψ ∈ A ⊗ H in A⊗ΦH by a⊗ΦΨ,
so that we have
ha1⊗ΦΨ2, a2⊗ΦΨ2iLΦ = hΨ1, Φ (a∗
1a2) Ψ2iH ,
for all a1, a2 ∈ A and Ψ1, Ψ2 ∈ H.
Moreover σΦ (a) (x⊗ΦΨ) = ax ⊗Φ Ψ, for each x⊗ΦΨ ∈ LΦ and VΦΨ = 1⊗ΦΨ for each Ψ ∈ H.
Since Φ is unital map, the linear operator VΦ is an isometry with adjoint V∗
Φ defined by
V∗
Φa⊗ΦΨ = Φ(a)Ψ,
7
for all a ∈ A and Ψ ∈ H.
We recall that the multiplicative domain of the ucp-map Φ : A → A is the C*-subalgebra of A such
defined:
DΦ = {a ∈ A : Φ(a∗)Φ(a) = Φ(a∗a) and Φ(a)Φ(a∗) = Φ(aa∗)},
we have the following implications (See Paulsen Ref.[9]):
a ∈ DΦ if and only if Φ(a)Φ(x) = Φ(ax) and Φ(x)Φ(a) = Φ(xa) for all x ∈ A.
Proposition 6. The ucp-map Φ is a multiplicative if and only if VΦ is an unitary. Moreover if x ∈ D (Φ)
we have:
σΦ (x) VΦV∗
Φ = VΦV∗
ΦσΦ (x) .
Proof. For each Ψ ∈ H we obtain the following implications:
a⊗ΦΨ = 1⊗ΦΦ (a) Ψ ⇐⇒ Φ (a∗a) = Φ (a∗) Φ (a) ,
since
ka⊗ΦΨ − 1⊗ΦΦ (a) Ψk = hΨ, Φ (a∗a) Ψi − hΨ, Φ (a∗) Φ (a) Ψi .
Furthermore, for each a ∈ A and Ψ ∈ H we have VΦV∗
Φa⊗ΦΨ = 1⊗ΦΦ (a) Ψ.
Now we prove the following Stinespring-type theorem (See Zsido Ref.[11]):
Proposition 7. Let A be a concrete C*-subalgebra with unit of B (H) and Φ : A → A an ucp-map, then
there exists a faithful representation (π∞, H∞) of A and an isometry V∞ on Hilbert Space H∞ such that
for a ∈ A,
V∗
∞π∞ (a) V∞ = π∞ (Φ (a)) ,
(15)
where
and (Vn, σn+1, Hn+1) is the Stinespring dilation of Φn for every n ≥ 0,
σ0 = id,
Φn = σn ◦ Φ
H∞ =
∞Mj=0
and
Hj ,
Hj = A⊗Φj−1
Hj−1,
for j ≥ 1 and H0 = H;
(16)
V∞(Ψ0, Ψ1, Ψ2, ...) = (0, V0Ψ0, V1Ψ1, ...)
for each (Ψ0, Ψ1, Ψ2, ...) ∈ H∞.
Furthermore the map Φ is a homomorphism if and only if V∞V∗
′
∞ ∈ π∞ (A)
.
Proof. By the Stinespring theorem there is triple (V0, σ1, H1) such that for each a ∈ A we have Φ(a) =
V∗
0σ1(a)V0. The application a ∈ A → σ1(Φ(a)) ∈ B(H1) is a composition of cp-maps therefore it is
also a cp map. Set Φ1(a) = σ1(Φ(a)). By appling the Stinespring's theorem to Φ1, we have a new triple
(V1, σ2, H2) such that Φ1(a) = V∗
1σ2(a)V1. By induction for n ≥ 1 we define Φn(a) = σn(Φ(a)) and we
have a triple (Vn, σn+1, Hn+1) such that Vn : Hn → Hn+1 and Φn(a) = V∗
We get the Hilbert space H∞ defined in 16 and the injective representation of the C*-algebra A on H∞ :
nσn+1(a)Vn.
π∞(a) =Mn≥0
σn(a)
with σ0(a) = a, for each a ∈ A.
Let V∞ : H∞ → H∞ be the isometry defined by
V∞(Ψ0, Ψ1....Ψn...) = (0, V0Ψ0, V1Ψ1....VnΨn...),
(17)
(18)
8
for all Ψi ∈ Hi with i ∈ N.
The adjoint operator of V∞ is
V∗
∞(Ψ0, Ψ1, ....Ψn...) = (V∗
0Ψ1, V∗
1Ψ2....V∗
n−1Ψn...)
(19)
for all Ψi ∈ Hi with i ∈ N, therefore
V∗
∞π∞ (a) V∞Ln≥0
Ψn = Ln≥0
= Ln≥0
V∗
nσn+1 (a) VnΨn = Ln≥0
σn (Φ (a)) Ψn = π∞ (Φ (a))Ln≥0
Ψn.
Φn (a) Ψn =
We notice that En = VnV∗
n be the orthogonal projection of B (Hn−1), we have:
E (Ψ0, Ψ1...Ψn..) = (0, E0Ψ1, E1Ψ2, ...EnΨn+1...) .
Finally for the proof of the last statement we only need to note that x belong to multiplicative domains
D (Φ) if and only if we have:
π∞ (x) V∞V∗
∞ = V∞V∗
∞π∞ (x) .
.
Remark 1. Let (M, Φ) be a quantum process, the representation π∞(a) : M → B(H∞) defined in
proposition 7 is normal, since the Stinespring representation σΦ : A → B(LΦ) is a normal map. Then
(π∞, H∞, V∞) is a covariant representation of quantum process.
4.1 Dilations of ucp-Maps
If (H∞, π∞, V∞) is the Stinespring representation of proposition 7, we have that V∗
∞π∞ (A) V∞ ⊂
for all X ∈bA.
π∞ (A) and by proposition 1 there exists a C*-algebra with unit of B(cid:16)bH(cid:17) such that:
1 - Zπ∞ (A) Z∗ ⊂bA,
2 - Z∗bAZ = π∞ (A) ,
3 - Z∗bV∗XbVZ = Vπ∞ (Z∗XZ) V,
Furthermore, we have a homomorphism bΦ :bA →bA thus defined
bΦ(X) = bV∗XbV
for all X ∈bA, such that for A ∈ A, X ∈bA and n ∈ N we have:
Φn(A) = Z∗bΦn(ZAZ∗)Z
Z∗bΦn(X)Z = Φn(Z∗XZ).
and
Φ : A → A.
The quadruple (bΦ,bA, H, Z) with the above properties, is said to be a multiplicative dilation of ucp-map
Remark 2. It is clear that these results are easily extended to the von Neumann algebras M with Φ
normal ucp-map. In this way we obtain a dilation of discrete quantum process (M, Φ).
5 Ergodic properties
Let A be a concrete C*-algebra of B (H) with unit, Φ : A → A an ucp-map and ϕ a state on A such that
ϕ ◦ Φ = ϕ. We recall (See N.S.Z. Ref.[8]) that the state ϕ is a ergodic state, relative to the ucp-map Φ, if
lim
n→∞
1
n + 1
nXk=0
[ϕ(aΦk(b)) − ϕ(a)ϕ(b)] = 0,
9
for all a, b ∈ A, while is weakly mixing if
lim
n→∞
1
n + 1
nXk=0
ϕ(aΦk(b)) − ϕ(a)ϕ(b) = 0,
for all a, b ∈ A.
We observe that by the Stinepring-type theorem 7 we can assume, without losing generality, that A is a
concrete C*-algebra of B (H), and that there is a linear isometry V on H such that:
Φ (A) = V∗AV for all A ∈ A.
Then(cid:16)bV,bH, Z(cid:17) is the minimal unitary dilation of (V, H) and the C*-algebra bA defined in proposition
1 is included in B(bH).
We want to prove the following ergodic theorem, for dilation ucp-map (bΦ,bA, H, Z) previously defined:
Proposition 8. If the ucp-map Φ admits a ϕ-adjoint and ϕ is a ergodic state, we obtain:
lim
N→∞
1
N + 1
NXk=0
while if ϕ is weakly mixing:
[ϕ(Z∗XbVk∗
Y bVkZ) − ϕ(Z∗XZϕ(Z∗Y Z)] = 0,
lim
N→∞
1
N + 1
NXk=0
ϕ(Z∗XbVk∗
Y bVkZ) − ϕ(Z∗XZ)ϕ(Z∗Y Z) = 0,
obtain:
for all X, Y ∈bA.
If we write every element X of B(cid:16)bH(cid:17) in matrix form X =(cid:12)(cid:12)(cid:12)(cid:12)
ϕ(cid:16)Z∗XbVk∗
Y bVkZ(cid:17) = ϕ(cid:0)X1,1VkY1,1Vk(cid:1) + ϕ(cid:0)X1,2C (k)∗ Y1,1Vk(cid:1) + ϕ(cid:16)X1,2Wk∗
Y2,1Vk(cid:17)
Lemma 1. Let X ∈ A∗(S), the *-algebra generated by operator system S defined in 12 and Y ∈bA,
X1,1 X1,2
X2,1 X2,2 (cid:12)(cid:12)(cid:12)(cid:12) with bH = H ⊕ l2 (FH) we
and the proof of previous proposition is an easy consequence of the following lemma:
a] if ϕ is an ergodic state we have:
(20)
(21)
b] if ϕ is weakly mixing we have:
1
lim
N→∞
N + 1
lim
N→∞
1
N + 1
ϕ(cid:16)X1,2C (k)∗ Y1,1Vk + X1,2Wk∗
NPk=0
NPk=0(cid:12)(cid:12)(cid:12)ϕ(cid:16)X1,2C (k)∗ Y1,1Vk + X1,2Wk∗
=( A (γ F α) B (β FΠ ·
Y2,1Vk(cid:17) = 0,
Y2,1Vk(cid:17)(cid:12)(cid:12)(cid:12) = 0.
β+m
β+m
0
Now we observe taht there is a natural number ko such that for each k > ko we obtain:
X1,2Wk∗
Y2,1Vk = 0
10
Proof. Since X ∈ A∗ (S) we can assume that X1,2 = AΓ (γ) ∆m (B, α, β) with A, B ∈ A and γ string of
A. Then:
X1,2 = A (γ FΠ ·
γ−1
Π
∗
·
α+m
F α) B (β FΠ ·
.
γ − 1 =
.
α + m
elsewhere
(22)
In fact we have that
for all (ξ0, ξ1...ξn..) ∈ l2 (FH) then Πβ+mWk∗
It follows that:
k−time
Wk∗
(ξ0, ξ1...ξn...) =
ϕ(cid:16)X1,2C (k)∗ Y1,1Vk + X1,2Wk∗
0, ...0 , ξ0, ξ1... ,
z}{
Y2,1Vk(cid:17) = lim
= 0 for all k >
.
β + m.
1
1
N + 1
lim
N→∞
NPk=0
ϕ(cid:0)X1,2C (k)∗ Y1,1Vk(cid:1) ,
Then we compute only the term ϕ(cid:0)X1,2C (k)∗ Y1,1Vk(cid:1) and by relationship 22 we can write that:
NPk=0
N + 1
N→∞
X1,2C (k)∗ Y1,1Vk = A (γ F α) B (β FΠ ·
C (k)∗ Y1,1Vk
β+m
moreover by relationship 6 for k >
.
β + m we have:
C (k)∗ = FV(k−β−m−1)∗
,
Π ·
β+m
it follows that
X1,2C (k)∗ Y1,1Vk = A (γ F α) B (β FV(k−β−m−1)∗
Y1,1Vk = A (γ F α) B (β FΦ(k−β−1) (Y1,1) Vβ+m+1.
Since
·
γ =
·
α + m + 1, by relationship 7 we obtain:
it follows that there exists a ϑ string of A with
·
ϑ =
·
β + m + 1 and an operator R ∈ A, such that
A (γ F α) B (β ∈ A(cid:16) .
β + m + 1(cid:12)(cid:12)(cid:12) ,
A (γ F α) B (β = R (ϑ .
Then
X1,2C (k)∗ Y1,1Vk = R (ϑ FΦ(k−β−1) (Y1,1) Vβ+m+1.
If we set ϑ = (n1, n2, ...nr, A1,A2, ....Ar) . we have n1 + n2 + ... + nr =
.
β + m + 1 and
R (ϑ FΦ(cid:18)k−
·
β−1(cid:19) (Y1,1) V
·
β+m+1 = RVn∗
r ArVn∗
r−1 Ar−1 · · · A2Vn∗
1 A1FΦ(cid:18)k−
·
β−1(cid:19) (Y1,1) V
·
β+m+1 =
where
We have:
= RΦnr (ArΦnr−1 (Ar−1 · · · Φn2 (A2Rk))) ,
Rk = Φnr (Ar) Φ(k−β−1) (Y1,1) − Φnr−1(cid:16)Φ (Ar) Φ(k−β) (Y1,1)(cid:17) .
♮
♮
ϕ(cid:0)X1,2C (k)∗ Y1,1Vk(cid:1) = ϕ (RΦnr (ArΦnr−1 (Ar−1 · · · Φn2 (A2Rk)))) =
(R) ArΦnr−1 (Ar−1 (· · ·Φn2 (A2Rk) )(cid:17) =
= ϕ(cid:16)Φnr
(cid:16)Φnr
= ϕ(cid:16)Φnr−1
= ϕ(cid:16)Φn2
♮ (cid:16)Φn3
♮ (cid:16)Φn3
(cid:16)Φnr
(R) Ar(cid:17) · · · A3(cid:17) A2Rk =
(R) Ar(cid:17) Ar−1 (Ar−2 · · · A3Φn2 (A2Rk)(cid:17) =
(R) Ar(cid:17) · · · A3(cid:17) A2Rk(cid:17)
and replacing Rk, we obtain:
Φn2
♮
· · · Φnr−1
(cid:16)Φnr
· · · Φnr−1
♮
♮
♮
♮
♮
♮
11
♮
♮
♮
♮
·
β(cid:19) (Y1,1)!!(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Then:
1
♮
♮
♮
♮
♮
♮
♮
♮
♮
♮
♮
1
N + 1
N + 1
1
N + 1
♮
♮
=
−
= Φn2
−Φn2
♮
♮
♮
lim
N→∞
· · · Φnr−1
· · · Φnr−1
♮
♮
· · · Φnr−1
♮
♮
while
lim
N→∞
· · · Φnr−1
♮
N + 1
1
It follows that :
♮
· · · Φnr−1
· · · Φnr−1
♮
· · · Φnr−1
· · · Φnr−1
If the state ϕ is ergodic we have:
(cid:16)Φnr
♮
♮
· · · Φnr−1
(R) Ar(cid:17) · · · A3(cid:17) A2Φn1 (A1) Φ(k−β−1) (Y1,1) −
♮ (cid:16)Φn3
(cid:16)Φnr
(R) Ar(cid:17) · · · A3(cid:17) A2Φn1−1(cid:0)Φ (A1) Φ(k−β) (Y1,1)(cid:1) .
♮ (cid:16)Φn3
(cid:16)Φnr
ϕ(cid:0)X1,2C (k)∗ Y1,1Vk(cid:1) =
= ϕ(cid:16)Φn2
♮ (cid:16)Φn3
(cid:16)Φnr
(R) Ar(cid:17) · · · A3(cid:17) A2Φn1 (A1) Φ(k−β−1) (Y1,1)(cid:17) −
(R) Ar(cid:17) · · · A3(cid:17) A2Φn1−1(cid:0)Φ (A1) Φ(k−β) (Y1,1)(cid:1)(cid:17).
−ϕ(cid:16)Φn2
♮ (cid:16)Φn3
(cid:16)Φnr
ϕ(cid:0)X1,2C (k)∗ Y1,1Vk(cid:1) =
NPk=0
(R) Ar(cid:17) · · · A3(cid:17) A2Φn1 (A1) Φ(k−β−1) (Y1,1)(cid:17) −
ϕ(cid:16)Φn2
♮ (cid:16)Φn3
(cid:16)Φnr
NPk=0
(R) Ar(cid:17) · · · A3(cid:17) A2Φn1−1(cid:0)Φ (A1) Φ(k−β) (Y1,1)(cid:1)(cid:17).
ϕ(cid:16)Φn2
♮ (cid:16)Φn3
(cid:16)Φnr
NPk=0
(R) Ar(cid:17) · · · A3(cid:17) A2Φn1−1(cid:0)Φ (A1) Φ(k−β) (Y1,1)(cid:1)(cid:17) =
ϕ(cid:16)Φn2
♮ (cid:16)Φn3
(cid:16)Φnr
NPk=0
(R) Ar(cid:17) · · · A3(cid:17) A2Φn1 (A1)(cid:17) ϕ (Y1,1) =
= ϕ(cid:16)Φn2
♮ (cid:16)Φn3
(R) Ar(cid:17) · · · A3(cid:17) A2(cid:17) A1(cid:17) ϕ (Y1,1)
= ϕ(cid:16)Φn1
♮ (cid:16)Φn2
♮ (cid:16)Φn3
(cid:16)Φnr
(cid:16)Φnr
♮ (cid:16)Φn3
(cid:16)Φn2
ϕ(cid:16)Φn1−1
NPk=0
(R) Ar(cid:17) · · · A3(cid:17) A2(cid:17) Φ (A1)(cid:17) ϕ (Y1,1) =
(cid:16)Φnr
♮ (cid:16)Φn3
(cid:16)Φn2
= ϕ(cid:16)Φn1−1
(R) Ar(cid:17) · · · A3(cid:17) A2(cid:17)(cid:17) A1(cid:17) ϕ (Y1,1) ,
= ϕ(cid:16)Φ♮(cid:16)Φn1−1
(cid:16)Φn2
♮ (cid:16)Φn3
(cid:16)Φnr
ϕ(cid:0)X1,2C (k)∗ Y1,1Vk(cid:1) = 0.
NPk=0
kY1,1Vk(cid:1)(cid:12)(cid:12) =(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
β−1(cid:19) (Y1,1)! − ϕ BΦn1−1 Φ (A1) Φ(cid:18)k−
ϕ BΦn1 (A1) Φ(cid:18)k−
(R) Ar(cid:17) · · · A3(cid:17) A2.
♮ (cid:16)Φn3
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
β−1(cid:19) (Y1,1)! − ϕ BΦn1−1 Φ (A1) Φ(cid:18)k−
ϕ BΦn1 (A1) Φ(cid:18)k−
β−1(cid:19) (Y1,1)! − ϕ (BΦn1 (A1)) ϕ (Y1,1)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
≤(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
ϕ BΦn1 (A1) Φ(cid:18)k−
β(cid:19) (Y1,1)!! − ϕ (BΦn1 (A1)) ϕ (Y1,1)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
+(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
ϕ BΦn1−1 Φ (A1) Φ(cid:18)k−
(R) Ar(cid:17) · · · A3(cid:17) A2(cid:17) Φ (A1) Φ(k−β) (Y1,1)(cid:17) =
Adding and subtracting the element ϕ (BΦn1 (A1)) ϕ (Y1,1) we can write:
β(cid:19) (Y1,1)!!(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
In weakly mixing case, using the previous results, we obtain:
(cid:12)(cid:12)ϕ(cid:0)X1,2C∗
♮
♮
· · · Φnr−1
(cid:16)Φnr
♮
where B = Φn2
then we obtain
· · · Φnr−1
♮
· · · Φnr−1
♮
· · · Φnr−1
♮
♮
♮
lim
N→∞
N + 1
1
♮
♮
·
·
+
♮
♮
N + 1
♮
1
·
·
≤
.
·
12
Moreover
·
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
ϕ BΦn1−1 Φ (A1) Φ(cid:18)k−
=(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
ϕ Φn1−1
(B) Φ (A1) Φ(cid:18)k−
♮
·
1
lim
N→∞
N + 1
and
lim
N→∞
1
N + 1
and by the weakly mixing properties we obtain:
♮
,
=
β(cid:19) (Y1,1)!! − ϕ (BΦn1 (A1)) ϕ (Y1,1)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
β(cid:19) (Y1,1)! − ϕ(cid:16)Φn1−1
NXk=0(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
ϕ BΦn1 (A1) Φ(cid:18)k−
NXk=0(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
ϕ Φn1−1
(B) Φ (A1)(cid:17) ϕ (Y1,1)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
β−1(cid:19) (Y1,1)! − ϕ (BΦn1 (A1)) ϕ (Y1,1)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(B) Φ (A1)(cid:17) ϕ (Y1,1)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
β(cid:19) (Y1,1)! − ϕ(cid:16)Φn1−1
(B) Φ (A1) Φ(cid:18)k−
= 0,
·
·
♮
♮
= 0.
Finally, the proof of proposition 8 is a simple result of the previous lemma.
References
[1] L. Accardi and C. Cecchini: Conditional expectations in von Neumann algebras and a theorem of
Takesaki,J. Funct. An.45 (1982) 245-273.
[2] W. Arveson: Non commutative dynamics and Eo-semigroups. Monograph in mathematics. Springer-
Verlag (2003).
[3] B.V. Bath and K.R. Parthasarathy: Markov dilations of nonconservative dynamical semigroups and
quantum boundary theory. Annales de lI. H. P., section B, tome 31, No 4 (1995) 601-651
[4] B. Kummerer: Markov dilations on W*-algebras. - J. Funct. Anal. 63 (1985), 139-177 .
[5] W.A. Majewski: On the relationship between the reversibility of dynamics and balance conditions -
Annales de l'I. H. P. section A, tome 39, no.1 (1983), 45-54.
[6] P.S. Muhly and B. Solel: Quantum Markov Processes (correspondeces and dilations). - Int. J. Math
Vol.13, No. 8 (2002), 863-906.
[7] B.Sz. Nagy and C. Foia¸s: Harmonic analysis of operators on Hilbert space - Regional Conference
Series in Mathematics, n.19 (1971).
[8] c. Niculescu, A. Stroh and L.Zsid´o: Non commutative extensions of classical and multiple recurrence
theorems - J. Operator Theory 50 (2002), 3-52.
[9] V.I. Paulsen: Completely bounded maps and dilations - Pitman Research Notes in Mathematics 146,
Longman Scientific & Technical, 1986.
[10] F. Stinesring: Positive functions on C* algebras - Proc. Amer. Math. Soc. 6 (1955) 211-216.
[11] L. Zsido: Personal communication - 2008.
13
|
Subsets and Splits
No saved queries yet
Save your SQL queries to embed, download, and access them later. Queries will appear here once saved.