• Aucun résultat trouvé

Morita equivalence of smooth noncommutative tori

N/A
N/A
Protected

Academic year: 2022

Partager "Morita equivalence of smooth noncommutative tori"

Copied!
27
0
0

Texte intégral

(1)

c 2007 by Institut Mittag-Leffler. All rights reserved

Morita equivalence of smooth noncommutative tori

by

George A. Elliott

University of Toronto Toronto, Ontario, Canada

Hanfeng Li

State University of New York at Buffalo Buffalo, NY, U.S.A.

1. Introduction

Letn2 and denote by Tn the linear space of n×nreal skew-symmetric matrices. For eachθ∈Tn the correspondingn-dimensional noncommutative torus Aθ is defined as the universalC-algebra generated by unitariesU1, ..., Un satisfying the relation

UkUj=e(θkj)UjUk,

where e(t)=e2πit. Noncommutative tori are one of the canonical examples in noncom- mutative differential geometry [34], [10].

One may also consider the smooth versionAθ of a noncommutative torus, which is the algebra of formal series

cj1,...,jnU1j1... Unjn,

where the coefficient function Zn(j1, ..., jn)!cj1,...,jn belongs to the Schwartz space S(Zn). This is the space of smooth elements ofAθ for the canonical action ofTnonAθ. A notion of Morita equivalence ofC-algebras (as an analogue of Morita equivalence of unital rings [1, Chapter 6]) was introduced by Rieffel in [31] and [33]. This is now often called Rieffel–Morita equivalence. It is known that two unital C-algebras are Morita equivalent as unitalC-algebras if and only if they are Rieffel–Morita equivalent [2, Theorem 1.8]. Rieffel–Morita equivalentC-algebras share a lot in common such as equivalent categories of Hilbert C-modules, isomorphic K-groups, etc., and hence are usually thought of as having the same geometry.

In [36] Schwarz introduced the notion of complete Morita equivalence of smooth noncommutative tori, which includes Rieffel–Morita equivalence of the corresponding C-algebras, but is stronger, and has important application in M(atrix) theory, [36], [23].

This research was supported by a grant from the Natural Sciences and Engineering Research Council of Canada, held by the first named author.

(2)

A natural question is to classify noncommutative tori and their smooth counterparts up to the various notions of Morita equivalence. Such results have important application to physics [11], [36]. For n=2 this was done by Rieffel [32]. (For the earlier problem of isomorphism, see below.) In this case it does not matter what kind of Morita equivalence we are referring to: there is a (densely defined) action of the group GL(2,Z) on T2, and two matrices in T2 give rise to Morita equivalent noncommutative tori or smooth noncommutative tori if and only if they are in the same orbit of this action, and also if and only if the ordered K0-groups of the algebras are isomorphic. The higher-dimensional case is much more complicated and there are examples showing that the smooth counterparts of two Rieffel–Morita equivalent noncommutative tori may fail to be completely Morita equivalent [35].

In [35] Rieffel and Schwarz found a (densely defined) action of the group SO(n, n|Z) onTn generalizing the above GL(2,Z) action. Recall that O(n, n|R) denotes the group of linear transformations of the vector spaceR2npreserving the quadratic formx1xn+1+ x2xn+2+...+xnx2n, and that SO(n, n|Z) refers to the subgroup of O(n, n|R) consisting of matrices with integer entries and determinant 1.

Following [35], let us write the elements of O(n, n|R) in 2×2 block form:

g=

A B C D

.

ThenA,B,CandD are arbitraryn×nmatrices satisfying

AtC+CtA= 0, BtD+DtB= 0 and AtD+CtB=I. (1) The action of SO(n, n|Z) is then defined as

= (Aθ+B)(Cθ+D)1, (2)

wheneverCθ+Dis invertible. There is a dense subset ofTn on which the action of every g∈SO(n, n|Z) is defined [35, p. 291].

After the work of Rieffel, Schwarz and the second named author in [35], [36] and [24]

(see also [39]), it is now known that two matrices inTngive completely Morita equivalent smooth noncommutative tori (in the sense of [36]) if and only if they are in the same orbit of the SO(n, n|Z) action.

Phillips has been able to show that two simple noncommutative tori are Rieffel–

Morita equivalent if and only if their ordered K0-groups are isomorphic [29, Theo- rem 3.11]. Using the result in [24] and Phillips’s result, recently we have completed

(3)

the classification of noncommutative tori up to Rieffel–Morita equivalence [17]. In gen- eral, two noncommutative tori are Rieffel–Morita equivalent if and only if they have isomorphic ordered K0-groups and centers.

It remains to classify smooth noncommutative tori up to Morita equivalence as unital algebras. We shall consider a natural subset Tn⊆Tn which can be described both algebraically in terms of the properties of the algebra Aθ (see Notation 4.1 and Corollary 4.10) and number theoretically (see Proposition 4.11), and has the property that the complement Tn\Tn has Lebesgue measure zero (Proposition 4.3). The main result of this paper is the Morita equivalence classification of the algebras arising from the subsetTn. (In particular, we have solved the problem of classification up to Morita equivalence in the generic case.)

Theorem 1.1. (1) The set Tn is closed under Morita equivalence of the associated smooth noncommutative tori; in other words, if the algebras Aθ and Aθ are Morita equivalent and θ∈Tn, then θ∈Tn.

(2) Two matrices in Tn give rise to Morita equivalent smooth noncommutative tori if and only if they are in the same orbit of the SO(n, n|Z)action.

Denote by Tn the subset of Tn consisting of the θ’s such that Aθ is simple. The weaker form of Theorem 1.1 (1) withTn replaced byTn∩Tnis a consequence of [28] (see the discussion after the proof of Proposition 4.11).

Consider the subset T3 of T3 consisting of the θ’s such that the seven numbers consisting of 1, θ12, θ13, θ23, together with all products of any two of these four, are linearly independent over the rational numbers. In [35] Rieffel and Schwarz showed that for anyθ∈T3, the matricesθ and−θare not in the same orbit of the SO(n, n|Z) action, although, by the work of Qing Lin and the first named author on the structure of 3- dimensional simple noncommutative tori culminating in [25], the C-algebras Aθ and Aθ are isomorphic. It is easy to see that the complementT3\T3has Lebesgue measure zero. Our Theorem 1.1 (together with Proposition 4.3) shows that the complement T3\(T3∩T3) also has Lebesgue measure zero, and for any θ∈T3∩T3, the algebras Aθ andAθ are not Morita equivalent.

A related question is the classification of noncommutative tori and their smooth counterparts up to isomorphism. The case n=2 was done by Pimsner, Rieffel and Voiculescu [30], [32], and the simpleC-algebra case forn>2 was also done by Phillips [29, Theorem 3.12] (see [29,§0] for more of the history). There have been various results for the smooth algebra case with n>2, [15], [13], [7]. In particular, Cuntz, Goodman, Jorgensen, and the first named author showed that two matrices inTn∩Tn give isomor- phic smooth noncommutative tori if and only if the associated skew-symmetric bichar-

(4)

acters ofZn are isomorphic [13]. This result is essentially a special case of Theorem 1.1, and the proof of it obtained in this way (see Remark 5.8) is new.

Schwarz proved the “only if” part of Theorem 1.1 (2) in the context of complete Morita equivalence [36,§5]. His proof is based on the Chern character [8], [16], which is essentially a topological algebra invariant. In order to show that his argument still works in our situation, we have to show that a purely algebraic Morita equivalence between smooth noncommutative tori is automatically “topological” in a suitable sense. For this purpose and also for the proof of Theorem 1.1 (1), in§2 and§3 we show that any algebraic isomorphism between two “smooth algebras” (see Remark 2.5 below) is continuous and any derivation of a “smooth algebra” is continuous. These are the noncommutative ana- logues of the following well-known facts in classical differential geometry: any algebraic isomorphism between the smooth function algebras of two smooth manifolds corresponds to a diffeomorphism between the manifolds, and any derivation of the smooth function al- gebra corresponds to a (complexified) smooth vector field on the manifold. We introduce the setTn and prove Theorem 1.1 (1) in§4. Theorem 1.1 (2) will be proved in§5.

Throughout this paperAwill be aC-algebra, andAwill be a dense sub--algebra of A closed under the holomorphic functional calculus (after the adjunction of a unit) and equipped with a Fr´echet space topology stronger than theC-algebra norm topology.

Unless otherwise specified, the topology considered on A will always be this Fr´echet topology.

We thank the referees and Ryszard Nest for very helpful comments. This work was carried out while Hanfeng Li was at the University of Toronto.

2. Continuity of algebra isomorphisms

In this section we shall prove Proposition 2.2 and Theorem 2.3, which indicate that the topology of Anecessarily behaves well with respect to the algebra structure.

Lemma2.1. Let ϕ:A!Abe an R-linear map. If Aϕ!Ais continuous,then so is Aϕ!A.

Proof. We shall use the closed graph theorem [3, Corollary 48.6] to prove the conti- nuity ofAϕ!A. SinceAϕ!Ais continuous, the graph ofϕis closed with respect to the norm topology in the second coordinate. It is then also closed with respect to the Fr´echet topology in the second coordinate. It follows that Aϕ!A is continuous.

Proposition 2.2. In A the ∗-operation is continuous, and the multiplication is jointly continuous. In other words, A is a topological ∗-algebra.

(5)

Proof. By Lemma 2.1, the-operation is continuous and the multiplication is sepa- rately continuous inA. Proposition 2.2 follows because of the fact that if the multipli- cation of an algebra equipped with a Fr´echet topology is separately continuous then it is jointly continuous [40, Proposition VII.1].

The following result was proved by Gardner in the caseA=A[18, Proposition 4.1], and was given as Lemma 4 of [13] in the case of smooth noncommutative tori.

Theorem2.3. Every algebra isomorphism ϕ:A1 !A2 is an isomorphism of topo- logical algebras.

Proof. Our proof is a modification of that for [18, Proposition 4.1]. It suffices to show thatϕ1 is continuous. Fora∈A1 letr(a) denote the spectral radius ofa, inA1 and also inA1 [37, Lemma 1.2].

First, for anya∈A1 we have

a2=aa=r(aa) =r(ϕ(a)ϕ(a))ϕ(a)ϕ(a)ϕ(a)·ϕ(a).

Next, we use the closed graph theorem (cf. above) to show that ϕ ϕ1 is con- tinuous onA2 . Let{am}m∈N⊆A1 be such thatϕ(am)!0 andϕ(am)!ϕ(b) for some b∈A1 . By the preceding inequality we have

am2ϕ(am)·ϕ(am)!0·ϕ(b)= 0,

am−b2ϕ(am)−ϕ(b)·ϕ(am)−ϕ(b)!0·ϕ(b)= 0.

Thereforeb=0. This shows that the graph ofϕ ϕ1 is closed.

Finally, let us use the closed graph theorem to show that ϕ1 is continuous. Let {am}m∈N⊆A1 be such that ϕ(am)!0 and am!b for some b∈A1 . By continuity of ϕ ϕ1we haveϕ(am)!0. By the inequality derived in the second paragraph it follows thatam2!0. Thereforeb=0. This shows that the graph ofϕ1 is closed.

Question 2.4. Is every algebra isomorphismϕ:A1 !A2 continuous with respect to theC-algebra norm topology?

Remark 2.5. Let us say that a Fr´echet topology on an algebraAis asmooth topology if there is aC-algebraA and a continuous injective homomorphismϕ:A!Asuch that ϕ(A) is a dense sub--algebra of A closed under the holomorphic functional calculus.

Theorem 2.3 can be restated as thatevery algebra admits at most one smooth topology.

Example 2.6. Let p∈Mn(A) be a projection. Then pMn(A)p is a dense sub-

-algebra of pMn(A)p, and the relative topology on pMn(A)p is a Fr´echet topology stronger than the C-algebra norm topology. By [37, Corollary 2.3] the subalgebra Mn(A)⊆Mn(A) is closed under the holomorphic functional calculus. It follows that the subalgebrapMn(A)p⊆pMn(A)pis closed under the holomorphic functional calculus.

(6)

3. Continuity of derivations

Throughout this section we shall assume further thatAis closed under the smooth func- tional calculus. By this we mean that, after the adjunction of a unit, for anya∈(A)sa andf∈C(R) we havef(a)∈A. Our goal in this section is to prove Theorem 3.4.

Example 3.1. Let B be a C-algebra, and let Δ be a set of closed, densely defined

-derivations ofB. Fork=1,2, ...,, define

Bk:={b∈B: ifj < k+1 andδ1, ..., δjΔ thenbbelongs to the domain ofδ1... δj}. It is routine to check that Bk is a sub--algebra of B and has the Fr´echet topology determined by the seminorms

b−!δ1... δjb, j < k+1 and δ1, ..., δjΔ.

By [6, Lemma 3.2], ifBis unital, thenBkis closed under the smooth functional calculus.

In particular,Aθ is closed under the smooth functional calculus.

Example 3.2. LetGbe a discrete group equipped with a length functionl, that is, a nonnegative real-valued function on G such that l(1G)=0, l(g1)=l(g), and l(gh) l(g)+l(h) for all g and hin G. Consider the self-adjoint and closed unbounded linear operator Dl:l2(G)!l2(G) defined by (Dlξ)(g)=l(g)ξ(g) forg∈G. Then δl(a)=i[Dl, a]

defines a closed, unbounded-derivation fromB(l2(G)) into itself. By Example 3.1 the intersection of the domains of theδlk’s for allk∈Nis a Fr´echet sub--algebra ofB(l2(G)) closed under the smooth functional calculus. Denote by Sl(G) the intersection of this algebra with the reduced group algebra Cr(G). Then Sl(G) is a dense Fr´echet sub-- algebra of Cr(G) closed under the smooth functional calculus and containingCG. This construction is due to Connes and Moscovici [12, p. 384]. Recall that G is said to be rapidly decaying if there exists a length functionl onGsuch that the intersection of the domains of theDkl’s for allk∈N, which we shall denote byHl(G), is contained inCr(G) [12], [21]. In such a case, it is a result of Ji [20, Theorem 1.3] that Hl(G) coincides with Sl(G), and the Fr´echet topology is also induced by the seminorms Dkl(·)l2 for 0k<∞. Consequently, Hl(G) is closed under the smooth functional calculus if it is contained inCr(G).

Question3.3. IsMm(A) closed under the smooth functional calculus for allm∈N?

The following theorem was proved by Sakai in the case A=A and by Bratteli, Elliott and Jorgensen in the caseA=Aθ [6, Corollary 5.3.C2]) (cf. also [27, Theorem 1], [6, Theorem 3.1] and the proof of [13, Lemma 4]). (In fact we shall only use the result in the case A=Aθ —in contrast to Theorem 2.3 which is needed in a more general setting.)

(7)

Theorem 3.4. Every derivation δ:A!A is continuous.

Theorem 3.4 is useful for determining the derivations of various smooth (twisted) group algebras. As an example, let us determine the derivations of the smooth group al- gebra of the 3-dimensional discrete Heisenberg groupH3. This group is the multiplicative

group ⎧

⎪⎨

⎪⎩

⎜⎝

1 a c 0 1 b 0 0 1

⎟⎠:a, b, c∈Z

⎫⎪

⎪⎭.

It is also the universal group generated by two elementsU andV such that W=V U V1U1

is central. It is amenable [14, p. 200]. Note that it is finitely generated. The Fr´echet space Hl(H3) does not depend on the choice of generators if we use the word length function corresponding to finitely many generators. So we shall denote it by H(H3).

Using the word length function associated with the generatorsU andV, we have H(H3) =

p,q,r∈Z

ap,q,rUpVqWr

, (3)

where{ap,q,r}is in the Schwartz space S(Z3), and the Fr´echet topology onH(H3) is just the canonical Fr´echet topology onS(Z3). Note thatH3has polynomial growth, and thus H(H3) is contained in C(H3)=Cr(H3) [21, Theorem 3.1.7]. By Example 3.2 and Theorem 3.4 we know that every derivation ofH(H3) into itself is continuous, and hence is determined by the restriction onCH3. Define derivationsU andV onCH3by

U(U)=U, U(V)= 0, U(W)= 0,

V(U)= 0, V(V)=V, V(W)= 0,

and extend them continuously to H(H3) using (3). We shall denote these extensions also byU andV. It is a result of Hadfield that every derivationδ:CH3!H(H3) can be written uniquely asδ=zUU+zVV+ ˜δfor somezU andzV in the center ofH(H3) and some ˜δ as the restriction of some inner derivation ofH(H3) [19, Theorem 6.4]. It is also known that the center of H(H3) is just the smooth algebra generated by W, i.e.,

r∈ZarWr:{ar}r∈Z∈S(Z)

[19, Lemma 6.2]. Thus we get the following result.

Corollary 3.5. Every derivation δ:H(H3)!H(H3) can be written uniquely as δ=zUU+zVV+ ˜δ for some zU and zV in the center of H(H3) and some inner derivation δ˜of H(H3).

(8)

In view of Lemma 2.1, to prove Theorem 3.4 it suffices to prove the following lemma.

Lemma 3.6. Every derivation δ:A!A is continuous.

The proof of Lemma 3.6 is similar to that of [27, Theorem 1] and [6, Theorem 3.1].

For the convenience of the reader, we repeat the main arguments in our present setting below (in which the seminorms may not be submultiplicative).

Lemma 3.7. Let I be a closed (two-sided) ideal of A such that A/I is infinite- dimensional. Then there exists b∈(A)sa such that the image of b in A/I has infinite spectrum.

Proof. Without loss of generality, we may assume thatA is unital. Consider the quotient map ϕ:A!A/I. Note that A:=ϕ(A) is infinite-dimensional. Assume that every element inAsa=ϕ((A)sa) has finite spectrum (inA/I). We assert that there exist nonzero projectionsP1, ..., Pm, ...inAsasuch thatPjPk=0 for allj=k. Assume that we have constructed P1, ..., Pj for some j0 with the additional property that QjAQj is infinite-dimensional, where Qj=1j

s=1Ps. Choose an element b in (QjAQj)sa\RQj. Since b has finite spectrum (in A/I) we can find a nonzero projection P∈QjAQj with P=Qj. Since QjAQj is infinite-dimensional, it is easy to see that either PAP or (Qj−P)A(Qj−P) has to be infinite-dimensional. (Recall that A/I is a C-algebra.) We may now choose Pj+1 to be one ofP andQj−P in such a way thatQj+1AQj+1 is infinite-dimensional, whereQj+1=1j+1

s=1Ps. This finishes the induction step.

Since A is a Fr´echet space we can find a complete translation-invariant metric d on A giving the topology of A [3, Corollary 13.5]. Choose bm(A)sa such that ϕ(bm)=Pm. ChooseλmR\{0}such thatd(λmbm,0)2m. Note that

d m

n=k+1

λnbn,0

2k for all k < m.

So the series

n=1λnbn converges to some b in A. Since the convergence holds in particular inA, it follows thatϕ(b)=

n=1λnPn, the convergence being with respect to the norm topology on A/I. In particular, we have that λm!0. Since also λm=0, we see that ϕ(b) has infinite spectrum (in A/I). This contradicts the assumption to the contrary, which is therefore false. In other words, Asa=ϕ((A)sa) contains an element with infinite spectrum (inA/I).

Lemma 3.8. Let δ:A!Abe a derivation. Set

I={b∈A:a∈A!δ(ab)∈Ais continuous}

and denote by I the closure of I in A. Then I is a closed (two-sided)ideal of A,and A/I is finite-dimensional.

(9)

Proof. Clearly I is an ideal of A. So I is an ideal of A. Assume that A/I is infinite-dimensional. By Lemma 3.7 we can find a self-adjoint elementbofAsuch that the image ofbinA/Ihas infinite spectrum (inA/I). Choosing suitablefm∈C(R) and settingbm=fm(b) we obtain{bm}m∈N⊆Asuch thatb2j∈/I for allj∈Nandbjbk=0 for allj=k. We may assume thatbj1 andδ(bj)1 for allj∈N.

Letdbe as in the proof of Lemma 3.7. Sinceb2m∈I/ , there exists some εm>0 such that for any ε>0 we can find some a∈A with d(a,0)<ε and δ(ab2m)εm. The multiplication inAis continuous by Proposition 2.2. Thus there exists someε>0 such that d((m/εm)bbm,0)2m for any b∈A with d(b,0)<ε. Take an a as above for this ε and set am=(m/εm)a. Then d(ambm,0)2m and δ(amb2m)m. Note that dm

n=k+1anbn,0

2k for all k<m. So the series

n=1anbn converges in A, say toa. Then (with a second use of Proposition 2.2)

δ(a)+aδ(a)bm+aδ(bm)δ(abm)=δ(amb2m)m,

which is a contradiction. The assumption that A/I is infinite-dimensional is therefore not tenable. We must conclude thatA/I is finite-dimensional.

4. The generic set

In this section we shall define the setTn and prove the part (1) of Theorem 1.1.

Denote by Der(Aθ ) the linear space of derivationsδ:Aθ !Aθ . SetRn=L. We shall think ofZnas the standard lattice inL(so that Hom(G,R) in [6] is just ourL), and shall regardθas an element of2

L. Let us writeL⊗RC=LC. Recall thate(t)=e2πit. One may also describe theC-algebraAθ as the universalC-algebra generated by uni- taries{Ux}x∈Zn satisfying the relations

UxUy=σθ(x, y)Ux+y, (4)

whereσθ(x, y)=e((x·θy)/2). In this description the smooth algebraAθ becomes S(Zn, σθ),

the Schwartz space S(Zn) equipped with the multiplication induced by (4). There is a canonical action of the Lie algebra LC as derivations of Aθ , which is induced by the canonical action ofTn onAθ and is given explicitly by

δX(Ux) = 2πiX, xUx forX∈LC andx∈Zn.

(10)

Notation 4.1. Lete1, ..., en be a basis ofZn. Denote byTn the subset ofTn consist- ing of those θ’s such that everyδ∈Der(Aθ ) can be written asn

j=1ajδej+ ˜δ for some a1, ..., an in the center ofAθ and some inner derivation ˜δ.

Remark 4.2. For any θ∈Tn andδ∈Der(Aθ ) there is at most one way of writingδ asn

j=1ajδej+ ˜δfor somea1, ..., an in the center ofAθ and some inner derivation ˜δ(see Proposition 4.7).

One may identifyTn withRn(n1)/2in a natural way. We may therefore talk about Lebesgue measure onTn.

Proposition 4.3. The Lebesgue measure of Tn\Tn is zero.

Letθ:ZnZn!Tdenote the bicharacter ofZn corresponding toθ, i.e., θ(x∧y) =e(x·θy).

Recall thatAθ is simple ifθ is nondegenerate in the sense that if θ(g∧h)=1 for some g∈Zn and all h∈Zn, then g=0 [38, Theorem 3.7]. (In fact, the converse is also true, though we don’t need this fact here.) If θ is nondegenerate, and for every 0=gZn the functionh!|θ(g∧h)−1|1forθ(g∧h)=1 grows at most polynomially, thenθ∈Tn

[6, p. 185]. So Proposition 4.3 follows from the following lemma.

Lemma 4.4. Denote by Tn the set of θ∈Tn such that θ is nondegenerate and for every 0=gZn the function h!|θ(g∧h)−1|1 for θ(g∧h)=1 grows at most polyno- mially. Then Tn\Tn has Lebesgue measure zero.

Proof. Ifθ−θ∈Mn(Z), thenθ=θ. Hence Tn\Tn=

ηMn(Z)∩Tn

(η+Tn\Tn),

whereTnconsists ofθ=(θjk)∈Tnwith 0θjk<1 for all 1j <kn. Denote the Lebesgue measure onTnbyμ. It suffices to show thatμ(Tn\Tn)=0. If there is some polynomialf in 12n(n−1) variables such that

1<

1j<kn

θjkmjk−t f(m)

for allt∈Zand 0=m=(mjk)1j<knZn(n1)/2, then bothθis nondegenerate and the required growth condition is satisfied—in other words,θ∈Tn. Set

F(m) =

1j<kn

(m2jk+1)2,

(11)

and consider the set Zs,m,t=

θ∈Tn: 1

1j<kn

θjkmjk−t sF(m)

for everys∈N, 0=mZn(n1)/2 andt∈Z. Set

m,t

Zs,m,t=Ws.

Then everyθ∈Tn\

s∈NWs satisfies the above condition. Therefore, it suffices to show that

s∈NWs has measure zero. Set

mm,t12=0

Zs,m,t=Ws.

Thenμ(Ws)12n(n−1)μ(Ws) because of the symmetry between theθjk’s for 1j <kn.

Integrating the characteristic function ofZs,m,toverθ12first and then over the otherθjk’s for 1j <kn, (j, k)=(1,2), we get that

μ(Zs,m,t)2s1F(m)1|m121| form12=0 and|t||m|:=

1j<kn|mjk|, while Zs,m,t=∅ for|t|>|m|. It follows that

μ(Ws)2s1

m12m=0

F(m)1|m121| |m|2s1

v∈Z

1 v2+1

n(n1)/2

!0, as s!∞.

Consequently,μ

s∈NWs

=0.

We shall now give two other characterizations ofTn, one in Corollary 4.10, in terms of the properties of the algebra, and one in Proposition 4.11, in terms of the number- theoretical properties ofθ. We need the following well-known fact.

Lemma 4.5. An element a=

h∈ZnahUh is in the center of Aθ if and only if the support of the coefficients ah is contained in the subgroup

H={h∈Zn:θ(g∧h) = 1for all g∈Zn}.

In particular,the center of Aθ is Cif and only if θ is nondegenerate in the sense that H={0}.

(12)

Proof. The elementais in the center ofAθ exactly ifUga=aUg for allg∈Zn. Note thatUga(Ug)1=

h∈Znahθ(g∧h)Uh. Consequently,Uga=aUg for allg∈Zn if and only ifah=0 for allh∈Zn\H.

Recall that the topology considered on Aθ will always be the Fr´echet topology, unless otherwise specified.

Definition 4.6. Let us say that a derivation δ∈Der(Aθ ) is approximately inner if there is a sequence{am}m∈N⊆Aθ such that [am, a]!δ(a) (in the Fr´echet topology) for every a∈Aθ . Denote by ADer(Aθ ) the linear space of approximately inner derivations ofAθ .

By [6, Corollary 5.3.D2], everyδ∈Der(Aθ ) can be written uniquely asn

j=1ajδej+ ˜δ for somea1, ..., anin the center ofAθ and some ˜δ∈Der(Aθ ) such that there is a sequence {bm}m∈N⊆Aθ with[bm, a]−δ(a)˜ !0 for everya∈Aθ . In fact, we can require [bm, a]! δ(a) in the Fr´˜ echet topology.

Proposition4.7. Every δ∈Der(Aθ )can be written uniquely as n

j=1ajδej+ ˜δfor some a1, ..., an in the center of Aθ and some δ˜ADer(Aθ ). The bicharacter θ is nondegenerate if and only if every δ∈Der(Aθ ) can be written uniquely as δX+ ˜δ for some X∈LC and some δ˜ADer(Aθ ).

Denote byAFθ the linear span of{Ux}x∈Zn. This is a dense sub--algebra ofAθ. In the proof of [6, Corollary 5.3.D2], which itself is based on the proof of [6, Theorem 2.1], one sees easily that in the present case actually the sequence{bm}m∈Ncan be chosen in such a way that [bm, a]δ(a) in the Fr´echet topology for everya∈AFθ. Thus Proposition 4.7 follows from the following lemma and Lemma 4.5.

Lemma 4.8. Let δ, δ1, δ2, ...∈Der(Aθ )be such that

δm(a)!δ(a) for every a∈AFθ ⊆Aθ . Then δm(a)!δ(a)for every a∈Aθ .

Proof. The proof is similar to that of [6, Corollary 3.3.4]. Letj=(j1, ..., jk) with 1j1, ..., jknandk0. We say thatw=(w1, ..., ws) is a subtuple ofjifskand there is a strictly increasing mapf:{1, ..., s}!{1, ..., k} such thatwm=jf(m). Set

aj= supδws... δw1(a),

where the supremum runs over all subtuples wofj, fora∈Aθ . It suffices to show that δ(a)−δm(a)j!0 for every a∈Aθ andj.

(13)

For eachg=(q1, ..., qn)Zn set|g|=s

s=1|qs|. Set

M= supm(Uts)j:t= 1, ..., n, s=±1 and m= 1,2, ...}.

Using the derivation property ofδmwe haveδm(U1q1... Unqn)j(2π)kM|g|k+1for every g=(q1, ..., qn)Zn. For any finite subsetZ⊆Zn anda=

g∈ZnagUg∈Aθ set aZ=

gZ

agUg∈AFθ.

By Theorem 3.4, every derivation onAθ is continuous. Thusδ(a−aZ)j!0, asZ goes toZn. Also

δm(a−aZ)j

g∈Zn\Z

(2π)kM|g|k+1|ag|.

Soδm(a−aZ)j!0 uniformly, asZ goes toZn. By assumption δ(aZ)−δm(aZ)j!0, asm!∞. Thusδ(a)−δm(a)j!0, asm!∞.

Corollary 4.9. A derivation δ∈Der(Aθ ) is approximately inner if and only if there is a sequence {bm}m∈N⊆Aθ with [bm, a]−δ(a)!0 for every a∈Aθ .

Combining Lemma 4.5 and Proposition 4.7 we get the following result.

Corollary4.10. The set Tn consists of those θ’s such that every δ∈ADer(Aθ )is inner.

SetF(g)=max1jn|θ(g∧ej)1|forg∈Zn. ThenF vanishes exactly on the sub- groupH in Lemma 4.5. Denote byF1 the function onZn taking on the value F(g)1 atg /∈H and the value 0 atg∈H.

Proposition4.11. The set Tnconsists of those θ’s which are such that the function F1 grows at most polynomially.

Proof. In view of Corollary 4.10 it suffices to show that everyδ∈ADer(Aθ ) is inner if and only if the functionF1 grows at most polynomially.

By Proposition 4.7, Theorem 3.4, [6, Corollary 5.3.E2] and the proof of [6, Theo- rem 5.1] (see also the first paragraph of the proof of [6, Theorem 2.1]), the derivations δ∈ADer(Aθ ) are in bijective correspondence with those C-valued functions Q on Zn which are such thatQvanishes on H and the functionch:g!Q(g)(θ(g∧h)−1) on Zn is in the Schwartz space S(Zn) for every h∈Zn. Actually δ(Uh)=

gch(g)UhUg. Fur- thermore,δis inner if and only ifQ∈S(Zn). In this case,δ(·)=

gQ(g)Ug . Clearly, ch∈S(Zn) for every h∈Zn if and only if cej∈S(Zn) for every 1jn, and if and only if the functiong!Q(g)F(g) is in S(Zn). In other words, the derivationsδ∈ADer(Aθ )

(14)

are in bijective correspondence with thoseQ:Zn!C which are such thatQvanishes on H and the functiong!Q(g)F(g) is inS(Zn). Therefore, everyδ∈ADer(Aθ ) is inner if and only if the pointwise multiplication byF1sendsS(Zn) into itself. Using the closed graph theorem [3, Corollary 48.6] it is easy to see that if the pointwise multiplication by F1 sends S(Zn) into itself, then this map is continuous and hence F1 grows at most polynomially. Conversely, if F1 grows at most polynomially, then obviously the pointwise multiplication by F1 sendsS(Zn) into itself. Therefore every δ∈ADer(Aθ ) is inner if and only ifF1grows at most polynomially.

Denote byTn the subset ofTn consisting of theθ’s such that θ is nondegenerate.

Let us indicate how to deduce the weaker form of the part (1) of Theorem 1.1, withTn

replaced byTn∩Tn, from [28] using the topological or algebraic Hochschild cohomology of Aθ . Recall that if two unital algebras are Morita equivalent, then their algebraic Hochschild cohomologies are isomorphic [26]. By Theorem 2.3 and Example 2.6, if two unital smooth algebras are Morita equivalent, then their topological Hochschild coho- mologies are also isomorphic. Nest, in [28, Theorem 4.1], calculated the topological Hochschild cohomology Htop(Aθ ,(Aθ )top) of (the Fr´echet algebra)Aθ with coefficients in the topological dual (Aθ )top (the 2-dimensional case was calculated earlier by Connes in [9]). From [28, Theorem 4.1] it is easy to see that the combined condition that θ be nondegenerate and that θ satisfy the condition in Proposition 4.11 is equivalent to the condition that Htop(Aθ ,(Aθ )top) be finite-dimensional in every degree. Using the simple projective resolution ofAθ as anAθ -bimodule in [28,§3], one also finds that this happens if and only if the algebraic Hochschild cohomology Halg(Aθ ,(Aθ )alg) is finite- dimensional in every degree. Thus the above weak form of the part (1) of Theorem 1.1 follows from considering either the topological or algebraic Hochschild cohomology ofAθ . In the 2-dimensional case, when θ is nondegenerate, the condition in Proposi- tion 4.11 was called aDiophantine condition by Connes [9, p. 349].

In [5] Boca introduced a certain subset ofTn, the complement of which has Lebesgue measure zero and which is also described number theoretically. His set is contained inTn. We do not know whether his set is the same asTn∩Tnor not.

To prove Theorem 1.1 (1), we start with some general facts about the comparison of derivation spaces for Morita equivalent algebras. Let A be a unital algebra. LetE be a finitely generated projective right A-module and set End(EA)=B. If we take an isomorphism of rightA-modulesE!p(kA) for some projectionp∈Mk(A), where kAis the direct sum ofkcopies ofAas rightA-modules with vectors written as columns, then we have an induced isomorphismB!pMk(A)p.

Let δ∈Der(A). Recall [8] that a connection for (EA, δ) is a linear map :E!E

Références

Documents relatifs

Abstract In this short note we give a polynomial-time quantum reduction from the vectorization problem (DLP) to the parallelization problem (CDHP) for efficiently computable

The unitary maps ϕ 12 will be very useful in the study of the diagonal action of U(2) on triples of Lagrangian subspaces of C 2 (see section 3).. 2.4

In Section 1, following a general construction of the symplectic groupoids of reduced Poisson manifolds by means of symplectic reduction for symplectic groupoids, we obtain

is an isomorphism. The crucial point is to prove that its inverse.. is well defined.. The previous corollary allows us to embed the Brauer group into a bigger group

tion here is the condition of stability under pullbacks.. We write III meaning the geometric morphism given by I*-| ill .) Such will then be the objects of the

P A are precisely the finitely generated projective A-right modules. Every cartesian closed, bicomplete category V is an example. for the results of part

We just note that there are at most 4081 such classes, this number being the sum of automorphism orbits of cocycles running over the finite groups of order 32, and at least

First we construct an infinite sequence of rational curves on a generic quintic V such that each curve has negative normal bundle.. Then we deform to quintics with nodes in such a