• Aucun résultat trouvé

Orientation of quantum Cayley trees and applications

N/A
N/A
Protected

Academic year: 2021

Partager "Orientation of quantum Cayley trees and applications"

Copied!
38
0
0

Texte intégral

(1)

angewandte Mathematik

( Walter de Gruyter Berlin New York 2005

Orientation of quantum Cayley trees and applications

By Roland Vergnioux at Mu¨nster

Abstract. We introduce the quantum Cayley graphs associated to quantum discrete groups and study them in the case of trees. We focus in particular on the notion of quan- tum ascending orientation and describe the associated space of edges at infinity, which is an outcome of the non-involutivity of the edge-reversing operator and vanishes in the classical case. We end with applications to Property AO and K -theory.

0. Introduction

The original motivation of this paper is Cuntz’ result on the K-amenability of free groups [7], and the geometric proof of this result given by the more general paper of Julg and Valette [8] on groups acting on trees with amenable stabilizers. Natural quantum ana- logues of the free groups are the free quantum groups defined by Wang and van Daele [16]

and studied by Banica [5]. Moreover, equivariant KK-theory can be generalized to the case of coactions of Hopf C -algebras [2], and the notion of K -amenability carries over to this quantum framework without di‰culty [17]. It is therefore natural to ask whether free quantum groups are K -amenable.

To apply the method of Julg and Valette in this framework, one needs a quantum geometric object to play the role of the tree acted upon by the quantum group under con- sideration. In the case of amalgamated free products of amenable discrete quantum groups, the construction of a quantum analogue of the Bass-Serre tree was achieved in [18] and could be used to prove the K -amenability of these amalgamated free products. In the case of the free quantum groups, the needed objects should be generalizations of the Cayley graphs of the free groups. The main goal of this paper is to define a notion of Cayley graph for discrete quantum groups, and to study its geometric properties in the case of the free quantum groups.

We will give in the last section applications of this study: a proof of the property of

Akemann and Ostrand and the construction of a KK-theoretic element g for free quantum

groups. To prove that these quantum groups are K -amenable, it remains to prove that

g ¼ 1. We refer the reader to the last section for more historical remarks and references

about Property AO and K -amenability.

(2)

The paper is organized as follows:

1. In the first section, we recall some notation and formulae concerning discrete quantum groups and classical graphs.

2. The second section is a technical one about fusion morphisms of free quantum groups, and its results are only used in the proofs of Sections 6 and 7. The reader will probably like to skip over this section at first.

3. In the third section, we give the definition of the Cayley graphs of discrete quan- tum groups and state some basic results about them.

4. We then restrict ourselves to the case of Cayley trees. We introduce and charac- terize this notion in the fourth section, where we also study the natural ascending orienta- tion of such a tree.

5. In the fifth section, we study more precisely the space of geometric edges of a quantum Cayley tree and we find that the projection of ascending edges onto geometric ones is not necessarily injective.

6. We show more precisely in the sixth section that the obstruction to this injectivity is the existence of a natural space of (geometric) edges at infinity, which vanishes in the classical case.

7. In the seventh section, we equip this space with a natural representation of the free quantum group under consideration, thus turning it into an interesting geometric object on its own.

8. Finally the last section deals with applications, as explained above.

Acknowledgments. Most of the results of Sections 2–5 and, in the special case of the orthogonal free quantum groups, of Sections 6, 7 and 8.2, were included in my PhD thesis [17] at the University Paris 7. I would like to thank my advisor, Prof. G. Skandalis, for having directed me to the beautiful paper [8] of Julg and Valette, and for his precious help in the redaction of my thesis.

The generalization and continuation of these results, as well as the redaction of the present article, were completed during a one-year stay at the University of Mu¨nster, where I could benefit from a postdoctoral position. It is a pleasure to thank Prof. J. Cuntz for his friendly hospitality and for the very stimulating atmosphere of his team.

1. Notation

The general framework of this paper will be the theory of compact quantum groups

due to Woronowicz [22]. In fact we will use it, from the dual point of view, as a theory of

discrete quantum groups. Let us fix the notation for the rest of the paper. The starting ob-

ject is a unital Hopf C -algebra ðS; dÞ such that dðS Þð1 n SÞ and dðSÞðS n 1Þ are dense in

S n S. Such a Hopf C -algebra will be called a Woronowicz C -algebra. One of the key

(3)

results of the theory is the existence of a unique Haar state h on ðS ; dÞ [20]. We will put d 2 ¼ ðid n dÞd ¼ ðd n idÞd and similarly d 3 ¼ ðid n id n dÞd 2 .

We denote by C the category of corepresentations of ðS; dÞ on finite-dimensional Hilbert spaces, by Irr C a set of representatives of irreducible corepresentations modulo equivalence, and by 1 C ¼ id C n 1 S the trivial corepresentation. We will denote by H a and v a A BðH a Þ n S the Hilbert space and the corepresentation associated to an object a A C.

The category C is equipped with direct sum, tensor product and conjugation operations: for the first and the second ones we refer to [20], and we give now some precisions about the third one which is slightly more involved.

Let ðe i Þ be an orthonormal basis of H a . The conjugate object a of a A C is charac- terized, up to isomorphism, by the existence of a conjugation map j a : H a ! H a , z 7! z which is an anti-isomorphism such that t a : 1 7! P

e i n e i and t a 0 : z n x 7! ðzjxÞ are resp.

elements of Morð1; a n aÞ and Morða n a; 1Þ. We put F a ¼ j a j a and we say that j a is nor- malized if Tr F a ¼ Tr F a 1 . This positive number, which is also equal to kt a ð1Þk 2 , does not depend on the normalized map j a . It is called the quantum dimension of a and is denoted by M a . When a is in Irr C, we can assume that a is in Irr C, and the possible conjugation maps j a only di¤er by a scalar. We have then a ¼ a, and if a 3 a one can choose nor- malized conjugation maps j a ; j a such that j a j a ¼ 1. If a ¼ a one has j a 2 ¼ G 1 for every normalized j a .

The coe‰cients of the corepresentations v a span a dense subspace S H S which turns out to be a Hopf *-algebra. We denote by m its multiplication, and by e : S ! C and k : S ! S its co-unit and its antipode. Notice that k is not involutive in general. In this regard, an important role is played by a family ð f z Þ z AC of multiplicative linear forms on S , which are also related to the non-triviality of the modular properties of h. We will need in this paper the following formulae in the Hopf *-algebra S:

E x A S ðid n eÞ dðxÞ ¼ ðe n idÞ dðxÞ ¼ x;

ð1Þ

E x A S m ðid n kÞ dðxÞ ¼ m ðk n idÞ dðxÞ ¼ eðxÞ1:

ð2Þ

Let L h : S ! H be the GNS construction of the Haar state h, denote by l : S ! BðHÞ the corresponding GNS representation and by S red its image. The Kac system of the compact quantum group ðS; dÞ is given by the following formulae, where

f ? x :¼ ðid n f ÞdðxÞ is the convolution product of f A S and x A S : V : ðL h n L h Þðx n yÞ 7! ðL h n L h Þ

dðxÞ1 n y ð3Þ ;

U : L h ðxÞ 7! L h

f 1 ? kðxÞ ð4Þ :

Let us recall the following notation and formulae from the general theory of multiplicative unitaries [3]. The unitary V A BðH n H Þ is multiplicative, meaning that V 12 V 13 V 23 ¼ V 23 V 12 , and for any o A BðHÞ one puts LðoÞ ¼ ðo n idÞðV Þ and rðoÞ ¼ ðid n oÞðV Þ. On the other hand, U is an involutive unitary on H such that V V ~ ¼ Sð1 n U ÞV ð1 n UÞS and V V ^ ¼ SðU n 1ÞV ðU n 1ÞS are again multi- plicative unitaries. Moreover the irreducibility property holds:

Sð1 n U ÞV 3

¼ 1 or,

equivalently, V V V ^ V V ~ ¼ ðU n 1ÞS. Here S denotes the flip operator, and we use the leg

numbering notation.

(4)

The reduced C -algebra S red coincides with the closure of L BðHÞ

in BðHÞ, and we similarly denote by S S ^ the closure of r

BðH Þ

. Both can be made Hopf C -algebras by the following formulae:

d red ðsÞ ¼ V ðs n 1ÞV ¼ V V ^ ð1 n sÞ V V ^ ; ð5Þ

ddð ^ ^ ssÞ ¼ V ð1 n ssÞV ^ ¼ V V ~ ð ss ^ n 1Þ V V ~ : ð6Þ

Notice that the reduction homomorphism l : S ! S red induces then an isomorphism between the dense Hopf *-algebras of both Woronowicz C -algebras. Besides, the unitary V lies in Mð S S ^ n S red Þ and we have the following commutation relations inside BðH Þ:

½S red ; US red U ¼ ½ S S; ^ U S SU ^ ¼ 0. There is also a full version of S [3] and we will say that S is a full Woronowicz C -algebra when it coincides with its full version.

Finally, the structure of the dual C -algebra S S ^ is very easy to describe: it is iso- morphic to the direct sum over a A Irr C of the matrix C -algebras BðH a Þ. We will denote by p a A BðH Þ the corresponding minimal central projections of S S, except the one associated ^ to the trivial corepresentation 1 C which will be denoted by p 0 .

Let us recall some facts about free quantum groups. The definition was given in [19], [16]: let n f 2 be an integer, and Q an invertible matrix in M n ðCÞ, the C -algebra A u ðQÞ is then the universal unital C -algebra generated by n 2 elements u i; j and the relations that make U ¼ ðu i; j Þ and QUQ 1 ¼ Qðu i; j ÞQ 1 A M n

A u ðQÞ

unitary. The C -algebra A o ðQÞ is defined similarly with the relations making U unitary and QU Q 1 equal to U . We will write S ¼ A o ðQÞ or A u ðQÞ when there is no need to distinguish the unitary and orthogonal versions. It is easy to see that S carries a unique Woronowicz C -algebra structure ðS; dÞ for which U is a corepresentation.

The corepresentation theory of A u ðQÞ was fully described in [5] in the following way.

The set of representatives Irr C can be identified with the free monoid on two generators u and u in such a way that the corepresentation associated to u is equivalent to U and the following recursive rules hold:

au n ua 0 ¼ auua 0 l a n a 0 ; au n ua 0 ¼ auua 0 l a n a 0 ; au n ua 0 ¼ auua 0 ; au n ua 0 ¼ auua 0 ; au ¼ ua; au ¼ ua:

The corepresentation theory of A o ðQÞ is even simpler. We assume in this case that QQ is a scalar matrix, otherwise the fundamental corepresentation U is not irreducible. The set Irr C can then be identified with N in such a way that the corepresentation associated to a 1

is equivalent to U and the fusion and conjugation rules read as in the representation theory of SUð2Þ:

a k n a l ¼ a jklj l a jkljþ2 l l a kþl2 l a kþl ; a k ¼ a k :

Let us finally fix some terminology concerning classical graphs. Following [14], a

graph g will be given by a set of vertices v, a set of edges e, an endpoints map e : e ! v v

(5)

and a reversing map y : e ! e which should be an involution such that e y ¼ s e. In this paper we denote by s the flip map for spaces and C -algebras. If e is injective, the graph g ¼ ðv; e; e; yÞ is isomorphic to the graph

v; eðeÞ; i can ; s

, which we will call the simplicial realization of g—although it only comes from a simplicial complex when it has no loops, i.e. when eðeÞ doesn’t meet the diagonal.

The set of geometric, or non-oriented, edges of g is the quotient e g of e by the relation a @ yðaÞ. An orientation of the graph is a subset e þ H e such that e is the disjoint union of e þ and yðe þ Þ. The quotient map evidently induces a bijection between any orientation and the set of geometric edges. When g is a tree endowed with an origin a 0 , we denote by j j the distance to a 0 and the ascending orientation of g is the set of edges a such that eðaÞ ¼ ða; bÞ with j bj > jaj.

Let D be a finite subset of a discrete group G such that 1 B D and D 1 ¼ D. The directional picture of the Cayley graph associated to ðG; DÞ is given by v ¼ G, e ¼ G D, eða; gÞ ¼ ða; agÞ and yða; gÞ ¼ ðag; g 1 Þ. Its simplicial realization will be called the simplicial picture of the Cayley graph.

2. Complements on fusion morphisms

In [4], [5] a full description of the involutive semi-ring structure of the core- presentation theory of A u ðQÞ and A o ðQÞ was given by means of the fusion and conjuga- tion rules on the set of irreducible objects up to equivalence. In this section we choose concrete representatives for the irreducible objects and compute explicitly isometric mor- phisms realizing the ‘‘basic’’ fusion rules. This section is a technical one and its results are only used in Sections 6 and 7: we advise the reader interested in quantum Cayley graphs to skip to the next section.

In the case of A u ðQÞ, with Q A GL n ðCÞ and n f 2, let us choose g ¼ u or u, and put g 2l ¼ g, g 2lþ1 ¼ g. We will mainly be interested in the corepresentations a k ¼ ggg g k (k terms) and a k;k

0

¼ g g k n g kþ1 g kþk

0

. As a matter of fact, the fusion rules of A u ðQÞ reduce to the relations a kþ1; k

0

þ1 ¼ a kþk

0

þ2 l a k;k

0

and trivial tensor products. In the or- thogonal case, we will also put g k ¼ g ¼ a 1 for every k A N and a k; k

0

¼ a k n a k

0

, to simplify the exposition.

Let us now choose concrete corepresentation spaces H k and H k for the classes a k ; a k . We first take H 0 ¼ C, equipped with the corepresentation 1 C n 1 S , and H g ¼ H g ¼ C n , equipped with the corepresentations U or U. For any k A N we denote by H g n k the tensor product corepresentation H g n H g n n H g

k

, and we define H k to be its unique sub-corepresentation equivalent to a k . We proceed in the same way inside H g nk

k

and

H g nk n H g n k

0

k

to get corepresentation spaces H k and H k; k

0

representing a k and a k; k

0

. We

will denote by t k ; t k and t d the morphisms associated to normalized conjugation maps of

a k ; a k and d A fg; gg respectively—we can and will assume in this section that j g j g ¼ G 1,

and we denote by H 1 the opposite sign. We put m k ¼ M a

k

¼ M a

k

and we call ðm k Þ k the

sequence of quantum dimensions of the quantum group. Let us gather simple facts about

them in the following lemma:

(6)

Lemma 2.1. (1) Denote by T l : H g n k ! H g n kþ2 the morphism id n l n t g

lþ1

n id n kl . We have then H kþ2 ¼ T k

l¼0

Ker T l H H g n kþ2 . (2) We have ðid n t

d Þðt d n idÞ ¼ G id H

d

and t d t d ¼ m 1 id C for d A fg; gg.

(3) For any k A N we have m k f dim H k , with equality i f f F k is the identity. More- over the equality m 1 ¼ 2 happens only in the three cases

A o

0 1

1 0

; A o

0 1 1 0

and A u

1 0 0 1

;

up to isomorphism.

(4) Put m 1 ¼ 0. The sequence of quantum dimensions satisfies the induction equations m 1 m k ¼ m kþ1 þ m k1 for k A N. Moreover m 0 ¼ 1 and m 1 is the geometric mean of Tr Q Q and TrðQ 1 .

Proof. The equality of Point (1) is true when k ¼ 2 because H g n H g is the ortho- gonal direct sum of t g ðCÞ and a subspace equivalent to a 2 , and the general result follows by induction because H kþ1 ¼ H 1;k X H k; 1 . The proof of Point (2) is an easy calculation. For Point (3), denote by a (resp. h) the arithmetic (resp. harmonic) mean of the eigenvalues of F k : the normalization condition of j k shows that a ¼ h 1 so that

m k ¼ a dim H k ¼ ffiffiffiffiffiffiffiffi p a=h

dim H k f dim H k :

For the equality case m 1 ¼ 2, see [5]. The induction equation of Point (4) relies on the fu- sion rule a k n a 1 ¼ a k1 l a kþ1 which implies that Sð j k n j 1 Þ and j k1 l j kþ1 are normal- ized conjugation maps for the same corepresentation [21]. Finally, the formula for m 1 holds because the matrix Q defines in the canonical base of C n a (non-normalized) conjugation map for H 1 , by definition of A u ðQÞ [5]. r

We want now to give the explicit expression of an isometric morphism from H p;p

0

to H pþ1; p

0

þ1 , for any p; p 0 A N. Note that there is an evident morphism T : H p; p

0

! H pþ1; p

0

þ1

given by the formula

T ¼ ðp pþ1 n p p

0

þ1 Þ ðid H

p

n t g

pþ1

n id H

p0

Þ;

where p k denotes the orthogonal projection of H g n k onto H k . However T is not isometric, and its definition does not allow to compute easily the image of a vector x A H p; p

0

. In Proposition 2.2 we give an explicit and simple expression of T , which allows us to compute its polar decomposition in Proposition 2.3. From this we finally deduce Lemmas 2.4 and 2.5 which will be used in the proofs of Lemmas 6.3 and 7.1 respectively.

We use more precisely the following ‘‘basic morphisms’’ from H g npþp

0

to H g npþp

0

þ2 , indexed by l A ½½0; p and l 0 A ½½0; p 0 :

T l; l

0

¼ ðid npþ1 n t g

pþ2

n id np

0

þ1 Þ ðid npl n t g

plþ1

n id nlþl

0

n t g

pþl0þ1

n id np

0

l

0

Þ:

If A ¼ ða l;l

0

Þ is a ð p þ 1Þ ðp 0 þ 1Þ matrix, we will write T A ¼ P

a l;l

0

T l;l

0

. Besides, we have

by Lemma 2.1 a simpler expression of T l; l

0

when l or l 0 equals zero:

(7)

T l; 0 ¼ G ðid n pl n t g

plþ1

n id n p

0

þl Þ;

T 0; l

0

¼ G ðid n pþl

0

n t g

pþl0þ1

n id n p

0

l

0

Þ:

Proposition 2.2. (1) There is at most one matrix A, up to a scalar factor, such that T A

restricts to a non-zero morphism from H p; p

0

to H pþ1;p

0

þ1 . If this is the case, one can assume that a 0; 0 ¼ 1 and one has then T A ¼ T .

(2) The following matrix A satisfies the conditions of Point (1):

a l;l

0

¼ ð H 1Þ lþl

0

m pl m p

0

l

0

m p m p

0

:

Proof. (1) It is not hard to check that the family ðT l; l

0

Þ is free, even when restricted to H p;p

0

. Hence it su‰ces to prove that an admissible T A is necessarily a multiple of T . First of all, Point (1) of Lemma 2.1 shows that we have

y j T l;l

0

ðxÞ

¼ 0 for any y A H pþ1; p

0

þ1

and ðl; l 0 Þ 3 ð0; 0Þ. Hence if T A ðxÞ A H pþ1;p

0

þ1 we obtain kT A ðxÞk 2 ¼ a 0;0

T A ðxÞ j T 0;0 ðxÞ

¼ a 0; 0

T A ðxÞ j TðxÞ ð7Þ :

In particular a 0;0 must be non-zero, and therefore we can assume that it equals 1. To con- clude we observe that the irreducible subspaces of H p;p

0

(resp. H pþ1; p

0

þ1 ) are pairwise in- equivalent, so that the morphisms T A and T must be proportional on each irreducible subspace of H p;p

0

, and (7) finally shows that the corresponding proportionality coe‰cients all equal 1.

(2) We will express the condition that T A ðxÞ should be in H pþ1;p

0

þ1 for any x A H p; p

0

using Point (1) of Lemma 2.1: for any k A ½½1; p and k 0 A ½½1; p 0 we should have T k;0 T A ðxÞ ¼ T 0; k

0

T A ðxÞ ¼ 0. We therefore compute, for l A ½½0; p and l 0 A ½½0; p 0 :

G T k; 0 T l; l

0

ðxÞ ¼ ðid n pk n t g

pkþ1

n id n k1 n t g

p

n id n p

0

þ1 Þ

ðid npl n t g

plþ1

n id nlþl

0

n t g

pþl0þ1

n id np

0

l

0

ÞðxÞ

¼

0 if k e l 2 or k f l þ 2;

G T 1; 0 T 0; l

0

ðxÞ if k ¼ l 1 or l þ 1;

m 1 T 1; 0 T 0; l

0

ðxÞ if k ¼ l ðsee Lemma 2:1Þ:

8 >

<

> :

As a result, we get the following su‰cient conditions on A:

E k A ½½1; p; l 0 A ½½0; p 0 m 1 a k;l

0

G ða k1;l

0

þ a kþ1; l

0

Þ ¼ 0;

if one agrees to put a pþ1;l

0

¼ 0. We recognize the induction equations satisfied by the sequence ðm pi Þ 0eiepþ1 , up to a sign change. Therefore these conditions mean that the columns of A should be proportional to

ð H 1Þ p m p ; . . . ; H m 1 ; 1

. Symmetrically the conditions T 0;k

0

T A ðxÞ ¼ 0 are equivalent to the lines of A being proportional to ð H 1Þ p

0

m p

0

; . . . ; H m 1 ; 1

. The matrix of the statement satisfies these conditions, hence the

associated morphism T A maps H p;p

0

to H pþ1; p

0

þ1 . r

(8)

Proposition 2.3. Let q A ½½0; minð p; p 0 Þ and denote by G H H p; p

0

the subspace equi- valent to a pþp

0

2q . One has then

kT jG k 2 ¼ m pþ1 m p

0

m pq m p

0

q1

m p m p

0

:

Proof. Let z A G be a unit vector. Because G is irreducible and T is a morphism, it is enough to compute the number N p; q p

0

¼ kTðzÞk 2 . Of course we will use the expression T ¼ T A of Proposition 2.2. We start from the formula (7) and notice that T l; l

0

ðzÞ is or- thogonal to T 0;0 ðzÞ A H p n H g

p

n H g

p

n H p

0

whenever l f 1 or l 0 f 1. Hence

kT A ðzÞk 2 ¼

T A ðzÞ j T 0;0 ðzÞ

¼ P

l;l

0

¼0; 1

a l; l

0

T l;l

0

ðzÞ j T 0;0 ðzÞ :

When l or l 0 equals zero, we can use the formulae for T l;l

0

T 0; 0 ðzÞ obtained in the proof of Proposition 2.2. The term l ¼ l 0 ¼ 1 will be a recursive one. Let us denote by T k;k 0

0

and T 0 ¼ ðp p n p p

0

Þ T 0; 0 0 the morphisms analogous to T k; k

0

and T for the inclusion H p1; p

0

1 ! H p;p

0

. We remark that

T 0;0 T 1; 1 ¼ G T 0; 0 ðid n p1 n t g

p

n t g

p

n id n p

0

1 ÞT 0; 0 0 ¼ T 0; 0 0 T 0; 0 0 ; so that

T 1; 1 ðzÞ j T 0; 0 ðzÞ

¼ kT 0; 0 0 ðzÞk 2 ¼ kT 0 ðzÞk 2 . Putting all together, we get the relation

N p;p q

0

¼ m 1 m p1

m p

m p

0

1

m p

0

þ m p1 m p

0

1

m p m p

0

N p1; q1 p

0

1

, m p

0

ðm p N p; q p

0

m pþ1 Þ ¼ m p

0

1 ðm p1 N p1;p q1

0

1 m p Þ:

Hence the left-hand side quantity is invariant under simultaneous shifts of the three indices p; p 0 and q. Note that the above relation is still valid when q ¼ 0 if one puts N k;k 1

0

¼ 0 for any k and k 0 : as a matter of fact, in this case z lies in H pþp

0

and in particular T 0 ðzÞ ¼ 0.

One can therefore shift q þ 1 times the indices and obtain the desired identity:

m p

0

ðm p N p;p q

0

m pþ1 Þ ¼ m p

0

q1 m pq : r Lemma 2.4. Let H k1; 1; k H H g nk1 n H g

k

n H g n k

k

be the tensor product of the respective subspaces equivalent to a k1 ; g k and a k , with k A N . Let t A MorðH k2 ; H k1; 1 Þ be an injection and denote by

— G 1 the subspace of ðt n idÞðH k2;k Þ H H k1; 1; k equivalent to a 2l ,

— G 2 the subspace of H k1; kþ1 H H k1; 1;k equivalent to a 2l . Then the norm of the orthogonal projection from G 1 onto G 2 equals

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 1 m l m l1

m k m k1

r

.

Proof. Let A ¼ ða l;l

0

Þ and T ¼ T A be the matrix and the morphism of Proposition

2.2 in the case ð p; p 0 Þ ¼ ðk 2; kÞ. We denote by A 0 ¼ ða l;l 0

0

Þ the matrix given by a l;0 0 ¼ a l; 0

(9)

and a l; 0 l

0

¼ 0 if l 0 f 1, and we remark that we have then T A

0

¼ ðT 0 n idÞ, where T 0 is the morphism of Proposition 2.2 for ð p; p 0 Þ ¼ ðk 2; 0Þ. Hence if x A H g n2k2 is a vector in the subspace of H k2;k equivalent to a 2l , we have T A

0

ðxÞ A G 1 and T A ðxÞ A G 2 . The orthogonal projection of G 1 onto G 2 being a morphism, it is a multiple of an isometry, so that its norm equals

T A

0

ðxÞ j T A ðxÞ

kT A

0

ðxÞk kT A ðxÞk ¼ kT A ðxÞk 2

kT A

0

ðxÞk kT A ðxÞk ¼ kTðxÞk kðT 0 n idÞðxÞk ;

because the terms T l; l

0

ðxÞ with l 0 f 1 are orthogonal to T A ðxÞ. We finally compute the value of the last quotient thanks to Proposition 2.3, with ðp; p 0 ; qÞ ¼ ðk 2; k; k 1 lÞ and ð p; p 0 ; qÞ ¼ ðk 2; 0; 0Þ:

kTðxÞk 2

kðT 0 n idÞðxÞk 2 ¼ m k1 m k m l1 m l

m k2 m k

m k2

m k1 ¼ 1 m l m l1

m k m k1 : r Lemma 2.5. Let H 1; k; k

0

H H g n H g n k n H g nk

0

k

be the tensor product of the respective subspaces equivalent to g; gg g k (k terms) and g k g kþ1 g kþk

0

1 (k 0 terms), with k; k 0 A N . Let t A MorðH k1 ; H 1 n H k Þ be an injection and denote by

— G 1 the subspace of ðt n idÞðH k1;k

0

Þ H H 1;k; k

0

equivalent to a kþk

0

1 ,

— G 2 the subspace of H 1; kþk

0

H H 1; k; k

0

equivalent to a kþk

0

1 . Then the norm of the orthogonal projection from G 1 to G 2 equals

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 1 m k

0

1

m kþk

0

1 m k

r

.

Proof. Like in the previous proof we will use the morphisms T : H 0; kþk

0

1 ! H 1; kþk

0

and T 0 : H 0;k1 ! H 1;k studied in Proposition 2.2. We notice that G 1 (resp. G 2 ) is the image of H kþk

0

1 by ðT 0 n idÞ (resp. T), so that the norm of the projection we are interested in is given by

TðxÞ j ðT 0 n idÞðxÞ

kTðxÞk kðT 0 n idÞðxÞk ¼ kTðxÞk kðT 0 n idÞðxÞk ;

for the same reason as above. Proposition 2.3 with ð p; p 0 ; qÞ ¼ ð0; k þ k 0 1; 0Þ and ð0; k 1; 0Þ gives then

kTðxÞk 2

kðT 0 n idÞðxÞk 2 ¼ ðm 1 m kþk

0

1 m kþk

0

2 Þm k1

m kþk

0

1 ðm 1 m k1 m k2 Þ ¼ m kþk

0

m k1

m kþk

0

1 m k

:

The result follows then from the identity m kþk

0

1 m k ¼ m kþk

0

m k1 þ m k

0

1 , which is easy to prove by induction, or by noticing that the irreducible subobjects of H kþk

0

1;k are the same as for H kþk

0

;k1 , up to the one equivalent to H k

0

1 . r

3. Quantum Cayley graphs

In this section we introduce the notion of Cayley graph for discrete quantum groups.

In fact the classical notion can be generalized into two di¤erent directions, coming from the

(10)

two di¤erent pictures introduced in Section 1. The quantum generalization of the simplicial picture is still a classical graph. On the contrary, the l 2 -spaces of the directional picture give rise in the quantum case to a quantum object, in the spirit of non-commutative geometry.

In the following definition, we use freely the notation of Section 1. In particular, S and S S ^ are the dual Hopf C -algebras of a compact quantum group—S being unital—, H is the GNS space of the Haar state of S, p a is the minimal central projection of S S ^ corre- sponding to an irreducible corepresentation a A Irr C and p 0 ¼ p 1

C

.

Definition 3.1. Let S be a Woronowicz C -algebra and p 1 a central projection of S S ^ such that Up 1 U ¼ p 1 and p 0 p 1 ¼ 0.

(1) The classical Cayley graph g associated with ðS ; p 1 Þ is given in simplicial form by v ¼ Irr C and e ¼ fða; a 0 Þ A v 2 j ddð ^ p a

0

Þð p a n p 1 Þ 3 0g.

(2) The hilbertian quantum Cayley graph associated with ðS; p 1 Þ is the 4-uplet ðH ; K; E; YÞ where K ¼ H n p 1 H, E ¼ V jK A BðK ; H n HÞ and Y ¼ V V ~ ð1 n UÞ jK A BðK Þ.

Let us introduce some more objects associated with this quantum graph. We denote by the linear form on p 1 H defined by

L h ðxÞ

¼ L h

eðxÞ .

(3) The source and target operators of the hilbertian quantum Cayley graph are E 1 ¼ ðid n Þ and E 2 ¼ E 1 Y A BðK ; HÞ.

(4) The quantum l 2 -space of geometric edges is K g ¼ KerðY þ idÞ.

Remarks 3.2. (1) The central projections p 1 that match the hypotheses of Definition 3.1 are sums of projections p a over finite subsets D H Irr C such that D ¼ D and 1 C B D.

The elements of e are then the ordered pairs of vertices ða; a 0 Þ for which there exist g A D such that a 0 H a n g. Note that this set of edges is symmetric, thanks to the equivalence a 0 H a n g , a H a 0 n g (Jacobi duality). In this paper, the classical Cayley graph will mainly be used as a tool for the study of the quantum one.

(2) The hilbertian quantum Cayley graph will be more useful for our purposes be- cause he naturally carries representations of the discrete quantum group under consider- ation: the C -algebra S acts on H via the GNS representation, and we let it act trivially on p 1 H. Moreover the operators Y, E 1 and E 2 commute to these representations, and in par- ticular K g is also endowed with a natural representation of S. The commutation properties to the action of S S ^ will be examined in Proposition 3.7.

(3) The identity V V V ^ V V ~ ¼ ðU n 1ÞS provides us with another expression for the re- versing operator: Y ¼ ðS V V VÞ ^ . Moreover the identity V V ~ ¼ ð J J ^ n J Þ V V ~ ð J J ^ n J Þ, where J; J J ^ are the modular conjugations of S and S S ^ [10], shows that ð J J ^ n J JÞYð ^ J J ^ n J J ^ Þ ¼ Y ¼ Y 1 . But the main fact about the reversing operator is its non-involutivity in the quantum case, see Proposition 3.4—in fact in this proposition it is enough to consider the restriction of V ~

V ð1 n UÞ to H n p 1 H , as soon as D generates C . r

Example 3.3 (classical case). Suppose S ¼ C ðGÞ for some discrete group G, with

the co-commutative coproduct given by dðgÞ ¼ g n g for all g A G H C ðGÞ. Then Irr C

(11)

identifies with G in such a way that v g F id C n g for every g A G, and the tensor product (resp. the conjugation) of corepresentations then coincides with the product (resp. the in- verse) of G. In particular, the inclusion a 0 H a n g reduces in this case to an equality a 0 ¼ ag, so that the classical graph of Definition 3.1 is nothing but the simplicial picture of the Cayley graph associated to ðG; DÞ, with D ¼ D .

Besides, one has H ¼ l 2 ðGÞ, S red ¼ C red ðGÞ, S S ^ ¼ c 0 ðGÞ and the projections p a corre- spond to the characteristic functions 1 a A S S ^ of the points of G. Moreover one has the fol- lowing expressions for the Kac system of ðS; dÞ: V ð1 a n 1 b Þ ¼ 1 a n 1 ab and U ð1 a Þ ¼ 1 a

1

. From this it is easy to see that the hilbertian quantum graph of Definition 3.1 is nothing but the l 2 -object associated to the directional picture ðv; e; e; yÞ of the Cayley graph of ðG; DÞ.

The only non-trivial check concerns the reversing operator: according to Proposition 3.4, one has S V V V ^ ¼ V SV ¼ E SE so that Y ¼ Y is the classical reversing operator. r

Proposition 3.4. Let ðH ; V ; U Þ be an irreducible Kac system [3]. Then the multi- plicative unitary V is co-commutative i f f V V ^ ¼ SV S i f f V V ~ ð1 n U Þ is involutive.

Proof. The direct implications are easy to check in the underlying locally compact groups. Conversely, assume that V V ^ ¼ SV S. Then for any x ¼ ðid n oÞðV Þ A S S ^ red , one also has x ¼ U ðid n oÞðV ÞU A U S S ^ red U H S S ^ red 0 , hence S S ^ red is commutative. Replacing V by V ~

V one gets the dual version of this result: if V V ~ ¼ SV S, then V is commutative. Now, V ~

V ð1 n UÞ is involutive i f f V V ~ ð1 n U Þ ¼ ð1 n U Þ V V ~ i f f Sð1 n U Þ V V ~ ð1 n U ÞS ¼ S V V ~ S, which implies by the previous ‘‘dual’’ statement that V V ~ is commutative, hence V is co- commutative. r

Let us give now alternative expressions for the reversing, source and target operators in terms of the Hopf *-algebra structure of S , and study the intertwining properties of these operators relatively to the representations of the dual Hopf C -algebra S S. ^

Lemma 3.5. Let x; y A S H S red , we have V ~

V ð1 n U Þ ðL h n L h Þðx n yÞ ¼ ðL h n L h Þ ðid n kÞ

ðx n 1ÞdðyÞ :

Proof. In this proof we will write Y in place of V V ~ ð1 n U Þ, although we do not necessarily restrict ourselves to K . We have

Y ðL h n L h Þðx n yÞ ¼ ðx n idÞ Y ðL h n L h Þð1 n yÞ;

so that it su‰ces to consider the case when x ¼ 1. Let us use the expressions (3) and (4) of U and V :

Y ðL h n L h Þð1 n yÞ ¼ Sð1 n U ÞV ð1 n U Þ ðL h n L h Þ

f 1 ? kðyÞ n 1

¼ Sð1 n U Þ ðL h n L h Þ d

f 1 ? kð yÞ : It is easy to check that dð f z ? aÞ ¼

id n ð f z ?Þ dðaÞ

, hence Y ðL h n L h Þð1 n yÞ ¼ Sð1 n UÞ ðL h n L h Þ

id n ð f 1 ?Þ d

kð yÞ

¼ Sð1 n UÞ ðL h n L h Þ

k n ð f 1 ?Þk

sdðyÞ:

(12)

One recognizes then ð1 n U Þ 2 ¼ 1 n 1:

Y ðL h n L h Þðx n yÞ ¼ S ðL h n L h Þ ðk n idÞsdðyÞ

¼ ðL h n L h Þ ðid n kÞdð yÞ: r

Proposition 3.6. Let be the linear form on L h ðSÞ defined by L h ¼ L h e. We have the following identities:

(1) E 1 ¼ ðid n Þ V and E 2 ¼ ð n idÞ V on K X ðL h n L h ÞðS n SÞ.

(2) E 2 ðL h n L h Þðx n yÞ ¼ L h ðxyÞ for x n y A S n S, and E 2 Y ¼ E 1 . Proof. The co-unit e being multiplicative, we have for all x and y in S

ðid n eÞ

dðxÞð1 n yÞ

¼ eðyÞðid n eÞdðxÞ ¼ eð yÞx ¼ ðid n eÞðx n yÞ;

hence E 1 ¼ ðid n Þ V . In the same way one can write, using the identity e k ¼ e and Equation (1):

ðid n eÞðid n kÞ

ðx n 1Þdð yÞ

¼ ðid n eÞ

ðx n 1Þdð yÞ

¼ xy ¼ ðe n idÞ

dðxÞð1 n yÞ

;

which yields E 2 ðL h n L h Þðx n yÞ ¼ L h ðxyÞ and E 2 ¼ ð n idÞ V , thanks to the defini- tion of E 2 and the expression of Y given by Lemma 3.5. Now, using these results and Equation (2), we can proceed to the last computation, where m : S n S ! S denotes the multiplication of S :

E 2 Y ðL h n L h Þðx n yÞ ¼ L h

x

mðid n kÞdð yÞ

¼ eðyÞL h ðxÞ ¼ E 1 ðL h n L h Þðx n yÞ: r Proposition 3.7. Let us define p p ^ 2 : ^ S S n2 ! LðHÞ by the formula p ^

p 2 ðx n x 0 Þ ¼ xðUx 0 UÞ. Similarly, let us denote by p p ^ 4 : ^ S S n 4 ! LðK Þ the homomorphism such that p p ^ 4 ðx n y n y 0 n x 0 Þ ¼ ðx n yÞðUx 0 U n Uy 0 U Þ, and let us put dd ^ 0 ¼ p p ^ 4 ð1 n 1 n ddÞ, ^ so that dd ^ 0 ðxÞ ¼ ðU n U ÞS ddðxÞSðU ^ n U Þ. One has then, for any x A S S: ^

(1) Y ðx n 1Þ ¼ ddðxÞ ^ Y, (2) Y ð1 n xÞ ¼ ð1 n UxU Þ Y, (3) Y dd ^ 0 ðxÞ ¼ ðUxU n 1Þ Y,

(4) E 2 ddðxÞ ¼ ^ x E 2 and E 2 dd ^ 0 ðxÞ ¼ UxU E 2 .

Hence Y intertwines the representations p p ^ 4 ðid n id n ddÞ ^ and p p ^ 4 ð dd ^ n id n idÞ of S ^

S n S S ^ n S S on K. In particular ^ Y commutes to p p ^ 4 dd ^ 3 . Similarly, E 2 intertwines the repre-

sentations p p ^ 2 and p p ^ 4 ð dd ^ n ddÞ ^ of S S ^ n S S ^ .

(13)

Proof. Point (1) results from the identity ddðxÞ ¼ ^ V V ~ ðx n 1Þ V V ~ . Writing E 2 ¼ ðid n Þ Y 1 , it implies the first relation of Point (4). For Point (3), one uses the formula Y ¼ ðS V V VÞ ^ and the fact that V commutes to U S SU ^ n 1:

ðUxU n 1ÞY ¼ ðUxU n 1ÞV V V ^ S ¼ V ðUxU n 1Þ V V ^ S

¼ V V V ^ ðU n UÞ V V ~ ðx n 1Þ V V ~ ðU n U ÞS

¼ V V V ^ ðU n UÞ ddðxÞðU ^ n UÞS ¼ V V V ^ S dd ^ 0 ðxÞ ¼ Y dd ^ 0 ðxÞ:

Composing on the left by ðid n Þ, one obtains the second relation of Point (4). For Point (2), simply notice that V V ~ is in MðUSU n S S ^ Þ, and hence commutes to 1 n U S SU ^ . r

4. Ascending orientation

In the case of a classical tree, the ascending orientation associated to a chosen origin defines a subspace K þ of the l 2 -space of edges K. The aim of this section is to introduce and study such a subspace in the case of quantum Cayley graphs. The next definition relies on the links between the quantum and classical Cayley graphs, the latter one being en- dowed with the origin 1 C .

Definition 4.1. Let S be a Woronowicz C -algebra and p 1 a central projection of S S ^ such that Up 1 U ¼ p 1 and p 0 p 1 ¼ 0. Assume that the classical Cayley graph associated with ðS; p 1 Þ is a tree, and denote by j j the distance to the origin 1 C in this tree.

(1) For any n A Nnf0; 1g we put p n ¼ P p a

jaj ¼ n A Zð S SÞ. ^ (2) We call p ?þ ¼ P

ð p n n p 1 Þ ddð ^ p nþ1 Þ and p þ? ¼ P

ð p n n p 1 Þ dd ^ 0 ð p nþ1 Þ the left and right ascending projections. Put p ? ¼ 1 p ?þ , p ? ¼ 1 p þ? .

(3) We call p þþ ¼ p þ? p the ascending projection of the quantum Cayley tree, and we denote by K þþ its image. We define similarly

p þ ¼ p þ? p ? ; p þ ¼ p ? p and p ¼ p ? p ? ; K þ ¼ p þ K ; K þ ¼ p þ K and K ¼ p K :

Remarks 4.2. (1) We have jaj ¼ 1 , a A D and jaj ¼ 0 , a ¼ 1 C . In particular the first point of Definition 4.1 is consistent with the notation p 0 and p 1 used in Definition 3.1.

(2) Take a A Irr C with jaj ¼ n þ 1. We have ddð ^ p a Þ ¼ P ddð ^ p a Þð p b n p b

0

Þ, where the sum goes over the ordered pairs ð b; b 0 Þ such that a H b n b 0 . Hence

ddð ^ p a Þð p n n p 1 Þ ¼ ddðp ^ a Þðp a

0

n p 1 Þ;

where a 0 is the vertex preceding a in the classical Cayley graph g. Hence we get the fol- lowing expression of p ?þ , in terms of the classical ascending orientation e þ of g:

p ¼ P

ða

0

; aÞ A e

þ

V ð p a

0

n p a ÞV ðid n p 1 Þ:

(14)

Recall that V plays the role of the endpoints operator, which implements in the co- commutative case the equivalence between the simplicial and directional pictures of the Cayley graph.

(3) Let J (resp. J J ^ ) be the modular conjugation on H induced by the involution of S (resp. S S ^ ). We know from [10] that U ¼ J JJ, ^ ½J; p n ¼ 0 and ðJ n JÞ ddð ^ p n ÞðJ n JÞ ¼ S ddð ^ p n ÞS.

From this we can deduce the following relation between p þ? and p ?þ : p þ? ¼ ð J J ^ n J JÞ ^ p ð J J ^ n J JÞ: ^

Hence p þ? and p come from the same projection of Mð S S ^ n S S ^ Þ acting respectively on the left and on the right of K. In particular they commute and are equal in the co-commutative case. r

In the next proposition we examine the links between the ascending projections and the reversing and target operators. The first point of the proposition shows that the re- versing operator switches the left and right versions of the quantum ascending projections:

this is the reason why it is necessary to use both p ?þ and p þ? in the general case.

Proposition 4.3. We use the hypothesis and notation of Definition 4.1.

(1) We have p ? ¼ Yp þ? Y and p ? ¼ Y p Y. More precisely:

Yp þ? ð p n n idÞ ¼ ð p nþ1 n idÞ p ? Y;

Yp ? ð p n n idÞ ¼ ð p n1 n idÞ p Y:

(2) We have E 2 p þ ¼ E 2 p þ ¼ 0 and

p n E 2 ¼ E 2 ð p n1 n idÞp þþ þ E 2 ðp nþ1 n idÞ p :

Proof. (1) We put u ¼ AdðU Þ. Using the formulae V V ~ ðp a n 1Þ V V ~ ¼ ddð ^ p a Þ and V ð1 n p a ÞV ¼ ddð ^ p a Þ for the dual coproduct, and the fact that p n commutes to U, we see that

V ~

V ð p n n 1Þ V V ~ ¼ ðU n 1ÞSV SðUp n U n 1ÞSV SðU n 1Þ

¼ ðu n idÞs

V ð1 n p n ÞV

¼ ðu n idÞs ddð ^ p n Þ : We use this expression in conjunction with the definition of Y:

Yp þ? ð p n n 1ÞY ¼ V V ~ ðp n n p 1 Þðu n idÞs ddð ^ p nþ1 Þ V V ~

¼ ð1 n p 1 Þ ddð ^ p n Þ V V ~ ðu n idÞs ddð ^ p nþ1 Þ V V ~ ¼ ddð ^ p n Þð p nþ1 n p 1 Þ:

To prove that the last expression is equal to p ? ðp nþ1 n p 1 Þ, it is enough to check that we

(15)

obtain p ? by summing it over n. But ð p n n p 1 Þ ddðp ^ n

0

Þ vanishes as soon as n 0 3 n G 1, so that

P ð p nþ1 n p 1 Þ ddð ^ p n Þ þ p ?þ ¼ P

ð p nþ1 n p 1 Þ ddðp ^ n Þ þ ðp n n p 1 Þ ddðp ^ nþ1 Þ

¼ P

ð p n n p 1 Þ P ddð ^ p n

0

Þ

¼ id K :

(2) The last point of Proposition 3.7 shows that p k E 2 ð p l n p 1 Þ equals E 2 ddð ^ p k Þðp l n p 1 Þ. In particular E 2 ð p n n p 1 Þ p ?þ ¼ p nþ1 E 2 p ?þ and similarly

E 2 ð p n n p 1 Þ p ? ¼ p n1 E 2 p ? : On the other hand one has, using Propositions 3.6 and 4.3:

E 2 ðp n n p 1 Þp ? ¼ E 1 Yðp n n p 1 Þp ? ¼ E 1 ð p n1 n p 1 Þp ?þ Y

¼ p n1 E 1 p ?þ Y ¼ p n1 E 2 p ? ; and similarly

E 2 ð p n n p 1 Þ p þ? ¼ p nþ1 E 2 p þ? :

As a result E 2 ðp n n p 1 Þp þ equals both p nþ1 E 2 p þ and p n1 E 2 p þ , so that it must vanish, and in the same way E 2 ð p n n p 1 Þ p þ ¼ 0. In particular p n E 2 ¼ p n E 2 ð p þþ þ p Þ, and the last statement of the proposition results then from the identities

p n E 2 p ¼ E 2 ðp n1 n p 1 Þ p and p n E 2 p ? ¼ E 2 ð p nþ1 n p 1 Þ p ? that we proved above. r

Until now we have used a very minimal notion of ‘‘tree’’ for our quantum Cayley graphs, namely the fact that the corresponding classical Cayley graph should be a classical tree. However this notion is too weak for our purposes, because it doesn’t take into account multiplicity issues that appear in the quantum case. More precisely, let us define the ‘‘full’’

classical Cayley graph G associated to ðS ; p 1 Þ in the following way:

v ¼ Irr C ; e ¼ fða; a 0 ; g; iÞ j g A D ; a 0 H a n g with multiplicity order ig;

eða; a 0 ; g; iÞ ¼ ða; a 0 Þ; yða; a 0 ; g; iÞ ¼ ða 0 ; a; g; iÞ:

Here D stands for the set of corepresentations associated with p 1 , like in Remark 3.2.1. The image of G by e is the classical Cayley graph g of Definition 3.1, but the map e needs not to be injective in general. The component g of an edge ða; a 0 ; g; iÞ is called the direction of the edge. In the rest of this paper, we will assume that the full classical Cayley graph G with origin 1 C is a ‘‘directional tree’’, meaning that it is a tree and that the ascending edges starting from a given vertex have pairwise di¤erent directions.

In Lemma 4.4 we state some basic results about classical Cayley graphs and give a

corepresentation-theoretic formulation of the extra assumptions introduced above. Propo-

sition 4.7 shows that our framework is the right one for the study of free quantum groups,

(16)

i.e. free products of orthogonal and unitary free quantum groups [16], [5]. Finally we prove that the quantum ascending orientation K þþ H K behaves nicely in this framework: the target operator induces a bijection between ascending edges and vertices orthogonal to the origin, exactly like in the classical case.

Lemma 4.4. Let S be a Woronowicz C -algebra and p 1 a central projection of S S such ^ that Up 1 U ¼ p 1 and p 0 p 1 ¼ 0. Assume that the classical Cayley graph g is a tree and denote by ða n gÞ þ (resp. ða n gÞ ) the sum of the subobjects of ða n gÞ which are further from (resp.

closer to) 1 C than a.

(1) For every a A Irr C one has jaj ¼ jaj in g.

(2) The full classical Cayley graph G is a directional tree i f f

— for all a A Irr C and g A D, ða n gÞ þ is irreducible or zero and

— for all a A Irr C and g 3 g 0 A D , ða n gÞ þ and ða n g 0 Þ þ are inequivalent or zero.

(3) We assume that G is a directional tree. For any ða; gÞ A Irr C D, one has ða n gÞ þ ¼ 0 i f f dim g ¼ 1 and a is the target of an ascending edge with direction g.

(4) If G is a directional tree and ða; bÞ is an ascending edge then dim b f dim a, with equality i f f the corresponding direction g A D has dimension 1.

Proof. (1) For this first point g does not need to be a tree. Because D ¼ D, it is enough to prove the following property: jaj e n i f f there exist elements g 1 ; . . . ; g n A D such that a H g 1 n n g n . We proceed by induction over n: for n ¼ 0 the property is satisfied because a H 1 C , a ¼ 1 C . Assume now that the property is satisfied for a given n f 0 and consider an a A Irr C such that jaj ¼ n þ 1. By definition of g there exist b A v and g A D such that j bj ¼ n and a H b n g, and the induction hypothesis for b gives the desired in- clusion a H g 1 n n g n n g. Assume conversely that a H g 1 n n g nþ1 and let ð b k Þ be a maximal orthogonal family of irreducible subobjects of g 1 n n g n . Because a is irre- ducible the inclusion a H L

ðb k n g nþ1 Þ implies that a H b k n g nþ1 for some k. By induc- tion hypothesis one has jb k j e n, hence jaj e n þ 1.

(2) Recall that the endpoints map e induces a morphism from G onto g, the latter one being a tree. Therefore G is a tree i f f e is injective, and it is enough to check it on the ascending orientation e þ H e: this leads to the condition that the subobjects ða n gÞ þ , for a given a, should have pairwise di¤erent subobjects without multiplicity. The tree G is then directional with respect to the origin 1 C i f f the corepresentations ða n gÞ þ have at most one subobject.

(3) and (4) We proceed again by induction on the distance to the origin: let ða; bÞ

be an ascending edge with direction g and assume that dim b f dim a, with equality

i f f dim g > 1. Take g 0 A D , the assumption on G shows that ð b n g 0 Þ ¼ ð b n g 0 Þ þ or

ðb n g 0 Þ ¼ a l ðb n g 0 Þ þ . In the first case, which can only happen when ðb n g 0 Þ þ 3 0, one

has clearly dimðb n g 0 Þ þ f dim b with equality i f f dim g 0 ¼ 1. On the other hand, we are in

(17)

the second case i f f g 0 ¼ g, because of the equivalence a H b n g 0 , b H a n g 0 . Moreover one has then ðb n g 0 Þ þ ¼ 0 i f f dim b dim g 0 ¼ dim a, which is equivalent to dim g 0 ¼ 1 by induction hypothesis. If on the contrary dim g 0 ¼ dim g > 1, the strict case of the induction hypothesis gives

dimðb n g 0 Þ þ ¼ dim b dim g 0 dim a f 2 dim b dim a > dim b: r

Proposition 4.5. Let S be a full Woronowicz C -algebra and p 1 a central projection of S ^

S such that Up 1 U ¼ p 1 and p 0 p 1 ¼ 0. If the full classical Cayley graph G is a directional tree, then

— S is a free product of a finite number of free Woronowicz C -algebras A o ðQ i Þ and A u ðQ j 0 Þ, with Q i Q i A C id and Q j 0 invertible,

— p 1 is the sum of the central supports of the respective fundamental corepresentations of these Woronowicz C -algebras.

Conversely the full classical Cayley graph G of any such pair ðS; p 1 Þ is a directional tree.

Proof. The classical graph g being a tree, the set Irr C of its vertices lies in one-to- one correspondence with the set of paths without half-turns starting from the origin 1 C . Because g is isomorphic to the full graph G, these paths are characterized by the finite se- quences of the directions they follow. Finally, Lemma 4.4 shows that the finite sequences ðg i Þ of elements of D that arise in such a way are exactly the ones that fulfill the condition g iþ1 3 g i or dim g iþ1 > 1 for each i.

For every pair fg; gg H D with g ¼ g (resp. g 3 g), the universal property of free quantum groups gives a Hopf homomorphism from some A o ðQÞ (resp. A u ðQÞ) onto S, where Q is a matrix such that QQ A C id (resp. is invertible). By universality of free products, one obtains then a surjective Hopf homomorphism F : F ! S, where F is some finite free product of free quantum groups. By definition, for each factor A o ðQÞ; A u ðQÞ H F the fundamental corepresentation U and its conjugate are mapped by id n F onto the corresponding pair fg; gg H D.

On the other hand, the starting remarks on the structure of g show that Irr C is the monoid generated by D and the relations fgg ¼ gg ¼ 1 j g A D ; dim g ¼ 1g. Hence F induces a bijection between Irr C and the set Irr F of irreducible corepresentations of F (up to equivalence)—see [19], [4], [5] for the description of Irr F and notice that A o ðQÞ and A u ðQÞ are respectively isomorphic to C ðZ=2ZÞ and C ðZÞ when dim Q ¼ 1. This proves, using [21] and the fact that we are dealing with full Woronowicz C -algebras, that F is injective.

The statement that G is a directional tree for any free product of free quantum groups follows easily from the above mentioned description of Irr F. r

Example 4.6 (free quantum groups). Let us picture the simplest cases of Proposition

4.5. When dim Q > 1 and QQ A C id, the classical Cayley graph of A o ðQÞ endowed with its

fundamental corepresentation is the half line with vertices at the integers. When dim Q > 1

and Q is invertible, the classical Cayley graph of A u ðQÞ is drawn in Figure 1. r

(18)

Proposition 4.7. Let S be a Woronowicz C -algebra and p 1 a central projection of S S ^ such that Up 1 U ¼ p 1 and p 0 p 1 ¼ 0. Assume that the classical Cayley graph G of ðS; p 1 Þ is a directional tree. Then the restriction of E 2 to K þþ is injective and its image is ð1 p 0 ÞH.

Proof. Let a A Irr C and g A D be such that jaj ¼ n and ða n gÞ þ 3 0. The subspace ð p a n p g ÞK is irreducible with respect to the representation p p ^ 4 of S S ^ n 4 , and equivalent to ða n gÞ n ða n gÞ. By definition, p ?þ ðp a n p g Þ ¼ ddð ^ p nþ1 Þð p a n p g Þ, so that p ?þ ð p a n p g ÞK is equivalent to ða n gÞ þ n ða n gÞ with respect to the representation p p ^ 4 ð dd ^ n id n idÞ of S ^

S n 3 . Similarly, and thanks to the first point of Lemma 4.4, the subspace p þ? ðp a n p g ÞK is equivalent to ða n gÞ n ða n gÞ þ with respect to the representation p p ^ 4 ðid n id n ddÞ. ^ Finally, p þþ ð p a n p g ÞK is equivalent to ða n gÞ þ n ða n gÞ þ for the representation p ^

p 4 ð dd ^ n ddÞ ^ of S S ^ n S S, and therefore irreducible by hypothesis. ^

Recall now from Proposition 3.7 that E 2 intertwines p p ^ 4 ð dd ^ n ddÞ ^ and p p ^ 2 . Hence the restriction of E 2 to ð p a n p g ÞK þþ is a multiple of an isometry, and one can compute the corresponding norm by considering the image of particular vectors, for instance characters.

One has by Proposition 3.6

E 2 p þþ ðw a n w g Þ ¼ E 2 p ?þ ðw a n w g Þ ¼ E 2 ddð ^ p nþ1 Þðw a n w g Þ

¼ p nþ1 E 2 ðw a n w g Þ ¼ p nþ1 ðw a n g Þ

¼ p nþ1 ðw ða n gÞ

þ w ða n gÞ

þ

Þ ¼ w ða n gÞ

þ

:

The norm in H of the character of an irreducible corepresentation equals 1 (cf. [20], th. 5.8), so that one eventually gets the following lower bound for the norm of E 2 p þþ ðp a n p g Þ:

kE 2 p þþ ð p a n p g Þk ¼ kw ðangÞ

þ

k

kp þþ ðw a n w g Þk f kw ðangÞ

þ

k

kw a n w g k ¼ 1:

ð8Þ

To conclude, let us remark that E 2 p þþ maps the respective orthogonal subspaces p þþ ðp a n p g ÞK onto the subspaces p ða n gÞ

þ

H , which are pairwise di¤erent by hypothesis, hence orthogonal, and whose sum equals ð1 p 0 ÞH. The operator E 2 is therefore injective and has dense image in ð1 p 0 ÞH, but this image is closed by (8). r

Remark 4.8. We will need in Section 8.1 to have a slightly more general and more precise result than (8). Let H be the algebraic direct sum of the subspaces p k H, and let E 2 be the operator defined on H n H in the same way as E 2 , that is, by Proposition 3.6,

u

1

C

u

Figure 1. Classical Cayley graph of the unitary free quantum group.

(19)

coming from the multiplication of S. Let us also extend p ?þ to H n H by putting P ?þ ¼ P ddð ^ p nþk Þð p n n p k Þ: we have then E 2 P ?þ ðid n p 1 Þ ¼ E 2 p ?þ ¼ E 2 p þþ . The first ar- guments of the preceding proof are still valid if one replaces g with any b A Irr C: as a matter of fact the hypothesis implies that a n b contains at most one subobject d with jdj ¼ jaj þ jbj. If such a d exists one gets the following generalization of (8):

kE 2 P ?þ ðp a n p b Þk ¼ k ddð ^ p d Þðw a n w b Þk 1 :

Let us remark that w a (resp. t a ð1Þ) generates the invariant line of p a H (resp. H a n H a ) with respect to the action of S S. Moreover one has ^ kw a k ¼ 1 and kt a ð1Þk ¼ ffiffiffiffiffiffiffi

M a

p , so that w a in fact corresponds to t a ð1Þ= ffiffiffiffiffiffiffi

M a

p in the isomorphism p a H F H a n H a , up to a phase factor. Consequently, in the isomorphism p a H n p b H F H a n H b n H b n H a the vector ddð ^ p d Þðw a n w b Þ corresponds to

ddð ^ p d Þ n id n id

t a n b ð1Þ

= ffiffiffiffiffiffiffiffiffiffiffiffiffiffi M a M b

p ¼ t d ð1Þ= ffiffiffiffiffiffiffiffiffiffiffiffiffiffi M a M b

p ;

if one isometrically identifies H d with the equivalent subspace of H a n H b . We therefore get the following exact formula, from which (8) can be recovered by noticing that M d e M a n M b :

kE 2 P ?þ ð p a n p b Þk ¼

ffiffiffiffiffiffiffiffiffiffiffiffiffiffi M a M b

M d

r

: r

5. Geometric edges

In this section we will study the Hilbert space K g ¼ KerðY þ idÞ when the classical Cayley graph G is a directional tree. We consider Proposition 4.7 as an evidence that K þþ

provides a good notion of ‘‘quantum ascending edges’’, and we would similarly like to know whether K g provides a good notion of ‘‘quantum geometric edges’’. By this we mean that there should be exactly one geometric edge for each ascending edge, which can be more rigorously expressed in the hilbertian framework by the fact that the restriction

p þþ : K g ! K þþ should be invertible.

Of course the study of K g ¼ KerðY þ idÞ is closely related to the problem of the non-involutivity of the reversing operator Y. The next proposition provides a ‘‘weak in- volutivity’’ property which we will use for the proof of Theorem 5.3, as well as a technical corollary obtained in Lemma 5.2. Notice that ðp þþ þ p ÞK behaves as a subspace of

‘‘quasi-classical’’ quantum edges in this regard.

Proposition 5.1. Let S be a Woronowicz C -algebra and p 1 a central projection of S S ^ such that Up 1 U ¼ p 1 and p 0 p 1 ¼ 0. Assume that the classical Cayley graph G of ðS; p 1 Þ is a directional tree. Then we have, for all n A N:

ðp þþ þ p ÞY n ð p þþ þ p Þ ¼ ðp þþ þ p ÞY n ðp þþ þ p Þ :

Proof. Inserting id ¼ p þþ þ p þ þ p þ þ p between the occurrences of Y G 1 in

Y Gn and developing, the statement of the theorem becomes an equality between two sums

(20)

of terms looking like p

00

;

00

Y G 1 p

1

;

10

Y G 1 Y G 1 p

n

;

n

. We will in fact prove that these terms are pairwise equal: for i ; 0 i A fþ; g with i A ½½0; n, one has

p

00

;

00

Yp

1

;

10

Y Yp

n1

;

n10

Yp

n

;

n

ð9Þ

¼ p

0

0

;

00

Y 1 p

0

1

;

1

Y 1 Y 1 p

0

n1

;

n1

Y 1 p

n

;

n

:

Let us proceed by induction over n A N, calling ‘‘rank 0’’ the trivial equality p

0

;

0

p ; ¼ p

0

;

0

p ; . Choose n f 1. As a first step, assume that there exists k A ½½1; n 1 such that k ¼ k 0 . Then the conclusion results straightforwardly from two applications of the induction hypothesis at ranks k and n k, with ð i ; 0 i Þ 0eiek and ð i ; 0 i Þ keien respectively.

We assume now that i ¼ 0 i for each i. If one side of (9) is non-zero, we necessarily have ð i ; i 0 Þ ¼ ð 0 0 ; n Þ for all indices i: as a matter of fact, Proposition 4.3 shows that the equalities iþ1 ¼ 0 i are required for the products in (9) not to vanish. In particular, we have then 0 0 ¼ n . This proves that the equalities from the first step are su‰cient to get the identities p Y n p ¼ p Y n p and p þþ Y n p þþ ¼ p þþ Y n p þþ . Moreover, taking the adjoint allows to switch from 0 0 ¼ 1 to 0 0 ¼ 1, so that it only remains to prove the equality

p þþ Yp þ Y Yp þ Yp ¼ p þþ Y 1 p þ Y 1 Y 1 p þ Y 1 p :

By adding terms from the first step we rather focus on the following equivalent equality:

p þþ Y n p þ p Y n p ¼ ? p þþ Y n p þ p Y n p :

Using the fact from Proposition 4.7 that the target operator E 2 is injective on K þþ , we can compose on the left by E 2 and use Proposition 4.3 to get another equivalent equality:

E 2 Y n p ¼ E 2 Y n p . But this is true since we have, from Proposition 3.6 and the defini- tion of E 2 : E 2 Y 2 ¼ E 1 Y ¼ E 2 , hence E 2 Y 2k ¼ E 2 Y 2k ¼ E 2 and E 2 Y ¼ E 2 Y 1 . r

Lemma 5.2. Let S be a Woronowicz C -algebra and p 1 a central projection of S S such ^ that Up 1 U ¼ p 1 and p 0 p 1 ¼ 0. Assume that the classical Cayley graph G of ðS; p 1 Þ is a di- rectional tree. Then there exists a unique unitary operator W : K þ ! K þ such that

E k A N Wð p þk p þþ ¼ ðp þ Y 1 Þ k p þþ : Moreover we have Wp þ Y ¼ p þ Y 1 W and p Y ¼ p Y 1 W on K þ .

Proof. Let X (resp. X 0 ) be the operator from K þþ n l 2 ðNÞ to K þ (resp. K þ ) de- fined by X ðx n e k Þ ¼ 2 k ð p þk x (resp. X 0 ðx n e k Þ ¼ 2 k ð p þ Y 1 Þ k x). Thanks to the coe‰cients 2 k , the operators X and X 0 are bounded, and it is easy to see that their ad- joints are resp. given by

X ¼ P

2 k T k p þþ ðY 1 p þ Þ k and X 0 ¼ P

2 k T k p þþ ðYp þ Þ k ;

where we put T k ðxÞ ¼ x n e k for any x A K þþ . Let z be an element of Ker X , for every k

(21)

and n we have p þþ ðY 1 p þ Þ k ð p n n idÞz ¼ 0. In particular ðY 1 p þ Þ n ð p n n p 1 Þz vanishes:

by Proposition 4.3 it is an element of ð p 0 n idÞK, which is contained in K þþ . By a finite descending induction on k A ½½0; n, we deduce that

ðY 1 p þ Þ k ðp n n p 1 Þz ¼ p þþ ðY 1 p þ Þ k ð p n n p 1 Þz þ YðY 1 p þ Þ kþ1 ð p n n p 1 Þz vanishes, and in particular ðp n n idÞz ¼ 0 for any n. Hence X is injective and X has dense image. In the same way, X 0 has dense image.

To prove the existence and the uniqueness of W, which is characterized by the iden- tity WX ¼ X 0 , it is therefore enough to show that kX hk ¼ kX 0 hk for any h A K þþ n l 2 ðNÞ, or as well, that X X ¼ X 0 X 0 . We will work on each subspace K þþ n e i separately, and we are thus led to prove for every k and l the equality

p þþ ðY 1 p þ Þ l ð p þk p þþ ¼ p þþ ðYp þ Þ l ð p þ Y 1 Þ k p þþ ; ð10Þ

which can also be written

p þþ Y 1 p þ p þ Y 1 ð1 p ÞYp þ p þ Yp þþ

¼ p þþ Yp þ p þ Yð1 p ÞY 1 p þ p þ Y 1 p þþ :

We proceed by induction on minðk; lÞ and distribute ð1 p Þ on both sides: the terms coming from p are equal thanks to Equation (9) of Proposition 5.1, and the terms com- ing from 1 are equal by induction hypothesis. When kl ¼ 0 but ðk; lÞ 3 ð0; 0Þ, both sides of (10) vanish, and when k ¼ l ¼ 0, (10) is trivial.

Because X has dense image, it su‰ces to check the equalities Wp þ Y ¼ p þ Y 1 W and p Y ¼ p Y 1 W on the image of ð p þ YÞ k p þþ , for every k. The first one follows immediately from the definition of W:

ðWp þ YÞð p þk p þþ ¼ Wð p þkþ1 p þþ ¼ ð p þ Y 1 Þ kþ1 p þþ and

ð p þ Y 1 WÞð p þk p þþ ¼ p þ Y 1 ðp þ Y 1 Þ k p þþ ¼ ðp þ Y 1 Þ kþ1 p þþ : For the second one, we use furthermore Equation (9) from Proposition 5.1:

ð p Y 1 WÞð p þk p þþ ¼ p Y 1 ðp þ Y 1 Þ k p þþ

¼ ð p YÞð p þk p þþ : r

Theorem 5.3. Let S be a Woronowicz C -algebra and p 1 a central projection of S S ^ such that Up 1 U ¼ p 1 and p 0 p 1 ¼ 0. Assume that the classical Cayley graph G of ðS; p 1 Þ is a directional tree. Then the orthogonal projection from K g to K þþ is injective and its image is given by

p þþ K g ¼ fz A K þþ j b h A K þ ðid þ p þ YÞðhÞ ¼ p þ Yzg:

Références

Documents relatifs

It is an open question as to whether families of sparse CSS codes exist with a minimum distance that grows at least as N α for α &gt; 1/2, even for Quantum codes with dimension 1..

The deterministic source-complete graphs are the graphs such that for any label word u and for any vertex s, there is a unique path from s labeled by u. As G is deterministic, [[ ]]

We will use either restrictions of Schubert classes from the ambient Grassmannian G(4, 7), or when convenient, the classes σ 8 , σ 7 and σ 0 6 of points, lines and β-planes in CG,

A kind of randomness can be viewed as a property of trajectories within classical dynamical systems, namely as unpre- dictability in finite time, over approximated physical measure

Of course over the real numbers, only the first type is relevant: the multiplicative norm on the real Cayley algebra being positive definite, it restricts on any four

We prove that the trace of a transient branching random walk on a planar hyperbolic Cayley graph has a.s.. continuum many ends and no

After recalling some notation about discrete quantum groups and classical Cayley graphs, we intro- duce in Section 2 the notion of path cocycle for discrete quantum groups and

2 Quantum Cayley trees The classical picture Quantum Cayley graphs Quantum path cocycles..