• Aucun résultat trouvé

Comptes Rendus

N/A
N/A
Protected

Academic year: 2022

Partager "Comptes Rendus"

Copied!
18
0
0

Texte intégral

(1)

Comptes Rendus

Mathématique

YoungJu Choie, François Dumas, François Martin and Emmanuel Royer

Formal deformations of the algebra of Jacobi forms and Rankin–Cohen brackets

Volume 359, issue 4 (2021), p. 505-521

<https://doi.org/10.5802/crmath.193>

© Académie des sciences, Paris and the authors, 2021.

Some rights reserved.

This article is licensed under the

Creative Commons Attribution4.0International License.

http://creativecommons.org/licenses/by/4.0/

LesComptes Rendus. Mathématiquesont membres du Centre Mersenne pour l’édition scientifique ouverte

www.centre-mersenne.org

(2)

2021, 359, n4, p. 505-521

https://doi.org/10.5802/crmath.193

Algebra, Number theory /

Algèbre, Théorie des nombres

Formal deformations of the algebra of Jacobi forms and Rankin–Cohen brackets

Déformations formelles de l’algèbre des formes de Jacobi et crochets de Rankin–Cohen

YoungJu Choie a, François Dumasb, François Martinb and Emmanuel Royer b

aPohang University of Science and Technology, Department of Mathematics, Pohang, Korea

bUniversité Clermont Auvergne, CNRS, LMBP, F-63000 Clermont-Ferrand, France E-mails:yjc@postech.ac.kr (Y. Choie), Francois.Dumas@uca.fr (F. Dumas), Francois.Martin@uca.fr (F. Martin), Emmanuel.Royer@math.cnrs.fr (E. Royer)

Abstract.This work is devoted to the algebraic and arithmetic properties of Rankin–Cohen brackets allowing to define and study them in several natural situations of number theory. It focuses on the property of these brackets to be formal deformations of the algebras on which they are defined, with related questions on restriction-extension methods. The general algebraic results developed here are applied to the study of formal deformations of the algebra of weak Jacobi forms and their relation with the Rankin–Cohen brackets on modular and quasimodular forms.

Résumé.Ce travail est consacré aux propriétés algébriques et arithmétiques des crochets de Rankin–Cohen permettant de les définir et de les étudier dans plusieurs situations naturelles de la théorie des nombres. Il se concentre sur la propriété qu’ont ces crochets d’être des déformations formelles des algèbres sur lesquelles ils sont définis, avec des questions connexes sur les méthodes de restriction-extension. Les résultats algébriques généraux développés ici sont appliqués à l’étude des déformations formelles de l’algèbre des formes de Jacobi faibles et leur relation avec les crochets de Rankin–Cohen pour les formes modulaires et quasimodulaires.

2020 Mathematics Subject Classification.53D55, 17B63, 11F25, 11F11, 16W25.

Funding. The first author Y. Choie is partially supported by NRF 2018R1A4A 1023590 and NRF 2017R1A2B2001807 of the National Research Fund of Korea.

The second, third and fourth authors are partially funded by the project CAP 20–25 and thank the Pohang University of Science and Technology for providing very nice working conditions.

Manuscript received 4th December 2020, revised and accepted 19th February 2021.

(3)

Introduction

Appearing in the late 1950s in the study of modular forms, Rankin–Cohen brackets have under- gone considerable development in many related fields, giving rise to a very abundant literature in recent decades. The initial problem was to construct bi-differential operators in two variables in such a way that their evaluation at modular forms is still a modular form. The preservation of this property of SL(2,Z)-equivariance was the main objective of the generalizations proposed for other algebras of functions of arithmetic origin (such as quasimodular forms and Jacobi forms, see for example [3, 4, 13, 20]) with respect to the appropriate arithmetical parameters (weight, depth, index). It was also considered in various contexts related to Lie theory, representation the- ory or differential geometry, see for instance [7, 10, 11, 17]). But there is a second fundamental as- pect of the families of Rankin–Cohen brackets, namely the fact that they define formal associative deformations of the algebras considered and thus appear as alternative versions of the families of transvectants in the classical theory of invariants. This specific global property of Rankin–Cohen brackets is the focus of this paper. Number theory is the main framework of this work: both the motivations of the algebraization of the problem and the applications are related to modular, quasimodular and Jacobi forms.

The fact that Rankin–Cohen brackets define a formal deformation on modular forms is men- tioned as a final remark by Eholzer in Zagier’s article [23]. This fact encodes a large set of rela- tions between the arithmetic functions build from the Fourier coefficients of the modular forms.

Understanding this set of relations is indeed the very first motivation of the seminal work by Za- gier [23] (see also [18]). This property and the resulting links between Rankin–Cohen brackets and quantization procedures have given rise to many significant articles, among which we can cite Unterberger and Unterberger [21], Cohen, Manin and Zagier [5], Connes and Moscovici [6], Bieliavsky, Tang and Yao [1], with in Pevzner [19] an enlightening perspective on their different points of view. It is impossible to give here complete references on such a vast subject, but it is necessary for our study to mention that the article [6] gives as a corollary of more general results an explicit method to construct from any homogeneous derivationDon a graded algebraAfor- mal Rankin–Cohen brackets which give a deformation onA; this type of brackets correspond to the notion of standard RC algebra in [23]. This general process cannot be applied directly to the algebraM of modular forms because it is not stable by the complex derivative. That’s why Zagier introduced in [23] (see also [24]) a more subtle argument to define formal Rankin–Cohen brackets on the algebraM≤∞of quasimodular forms whose restriction toMgives precisely the classical Rankin–Cohen brackets. We give in this paper a formalisation of this extension-restriction argu- ment to the general framework of abstract differential algebras and use it to extend the classical Rankin–Cohen brackets onMinto a deformation of the algebra of weak Jacobi forms.

The text is organized in five sections. The first is devoted to recalling the basic notions and results on formal deformations and to formulate some of their corollaries in terms adapted to our study; we specify in particular in Propositions 3 and 5 the notion of formal Rankin–Cohen brackets associated with a homogeneous derivation of a graded or bigraded algebra.

In the second section, we give two possible strategies to construct significant formal defor- mations on a graded algebraA. The most direct one is to start with an homogeneous derivation DofAand to consider the associated formal Rankin–Cohen brackets. The second strategy is to embedAinto an algebraRand consider a suitable derivationD ofRwhich is not a derivation ofA, so that the formal Rankin–Cohen brackets onRassociated toDrestrict into a deformation ofA. This is the contents of Theorem 6. In the particular case whereA=M we apply the first strategy taking forDthe Serre derivative, and the second one withR=M≤∞andDthe complex derivative. We prove in Proposition 10 that the two formal deformations ofAobtained by these two approaches are not isomorphic. We develop these two strategies for weak Jacobi forms in the

(4)

rest of the paper.

The third section gathers useful notions on the weak Jacobi forms for the full modular group, according to [9]. The algebra of weak Jacobi formsJ is a polynomial algebra in four variables E4, E6, A and B overC, bigraded by the weight and the index, containing M =C[E4, E6] as a subalgebra.

We define in the fourth section a family of derivations ofJextending the Serre derivation on Mfrom which we deduce in Theorem 11 a family parameterized byC3of formal Rankin–Cohen brackets onJ. We classify in Theorem 14 these deformations ofJ up to modular isomorphism.

The last section deals with the natural problem of extending the classical Rankin–Cohen brackets onMto a deformation ofJ. To do this, we implement the extension-restriction method based on the Theorem 6. Here the considered extension is not a polynomial extension as in the case of the embedding ofM inM≤∞, but an extension by localization. More precisely we introduce the Laurent polynomial algebraK =C[E4, E6, A±1, B] which containsJ and also a copyQ=C[E4, E6, F2] ofM≤∞, where F2=B A1is a scalar multiple of the Weierstraß function.

We define in Theorem 15 a family parameterized byC2 of formal Rankin–Cohen brackets on K whose restriction toM are the classical Rankin–Cohen brackets onM, and determine in Theorem 18 the values of the parameters for which these brackets give deformations ofJ.

1. Algebraic results on formal deformations 1.1.

Formal deformations

In this section, we recall the basic properties of formal deformations and their isomorphisms.

Our main reference on this subject is [12, Chapter 13].

1.1.1. Definition and first properties

For any commutativeC-algebra A, let A[[ħ]] be the commutative algebra of formal power series in one variableħwith coefficients in A. Aformal deformationofAis a family (µj)j0of bilinear mapsµj:A×AAsuch thatµ0is the product ofAand such that the (non commutative) product onA[[ħ]] defined by extension of

∀(f,g)∈A2 f ?g=X

j≥0

µj(f,g)ħj (1) is associative. This associativity is reflected in

n≥0 ∀(f,g,h)∈A3

n

X

r=0

µnr(µr(f,g),h)=

n

X

r=0

µnr(f,µr(g,h)). (2) If (µj)j≥0is a formal deformation ofA, ifµ1is skew-symmetric andµ2is symmetric, thenµ1is a Poisson bracket onA.

1.1.2. Isomorphic formal deformations

Let (µj)j0and (µ0j)j0be two formal deformations ofA. They areisomorphicif there exists a C-linear bijective mapφ:AAsuch that

j≥0 ∀(f,g)∈A2 φ(µj(f,g))=µ0j(φ(f),φ(g)). (3) Assume thatµ1is skew-symmetric andµ2is symmetric. Formula (3) forj=0 andj=1 implies, in particular, thatφis an automorphism of the Poisson algebra (A,µ1). We denote by?and # the products onA[[ħ]] respectively associated to the formal deformations (µj)j0and (µ0j)j0. The C[[ħ]]-linear extensionφ:A[[ħ]]→A[[ħ]] satisfies

∀(f,g)∈A2 φ(f ?g)=φ(f) #φ(g). (4)

(5)

1.1.3. Example

It is well known that, if d and δare two C-derivations of Asatisfying δd =, then the sequence (Td,nδ)n0of formal transvectantsTd,nδ:A×AAdefined for anyf,gAby

Td,δn (f,g)=

n

X

r=0

(−1)r

r!(nr)!dn−rδr(f)drδn−r(g) (5) is a formal deformation ofA. The next paragraph is devoted to the more complicated situation where the two derivations don’t commute but generate the two dimensional non abelian Lie algebra.

1.2.

Connes–Moscovici deformations

We recall in the following proposition a particular case of a theorem due to Connes & Moscovici which provides a general method for constructing formal deformations.

Definition 1. Let A a commutativeC-algebra, andand D twoC-derivations of A satisfying

DD∆=D. (6)

The Connes–Moscovici deformation on A associated to(D,∆)is the sequence(CMD,n)n≥0of bilin- ear maps A×AA defined for any f,g∈A by

CMD,∆n (f,g)=

n

X

r=0

(−1)r

r!(n−r)!Dr(2∆+r)nr(f)Dnr(2∆+nr)r(g), (7) with convention1=IdAand for any function F:AA the Pochhammer notation:

F〈0〉=1 and F〈m〉=F(F+1)· · ·(F+m−1) for any m≥1. (8) Proposition 2. Let D andbe two derivations on A such thatDD∆=D. Then,(CMD,n)n≥0is a formal deformation of A.

Proof. See [6, eq. (1.5)], also [22] and [1].

The relationship between Connes–Moscovici deformations and transvectants has been exam- ined in different works, see for example [22, Section II.2.C], [1, Section 3], [16] and [15].

1.3.

Formal Rankin–Cohen brackets

Applying the above general result to graded situations gives rise to the following notion of formal Rankin–Cohen brackets.

Proposition 3. Let A=L

k∈NAk be a graded commutativeC-algebra, and D a derivation of A such that D(Ak)⊂Ak+2for any k≥0. Let us consider the sequence(FRCDn)n≥0of bilinear maps A×AA defined by bilinear extension of

FRCDn(f,g)=

n

X

r=0

(−1)r

Ãk+n−1 nr

`+n−1 r

!

Dr(f)Dn−r(g), (9) for any fAk,gA`. Then,

(1) (FRCDn)n0is a formal deformation of A.

(2) FRCDn(Ak,A`)⊂Ak+`+2nfor all n,k,`≥0.

(3) The associated Poisson bracket is defined by bilinear extension of FRCD1(f,g)=k f D(g)−`g D(f)for all fAk,gA`.

(6)

Proof. The linear map∆:AAdefined on each homogeneous component by∆(f)=12k f for any fAk is a derivation of Awhich satisfies ∆DD∆=D. We compute (2∆+r)〈n−r(f)=

(k+n−1)!

(k+r1)!f and (2∆+nr)〈r(g)=(`+(`+n−1)!nr1)!g. So

n≥0, CMD,n=FRCDn (10) and (1) follows from Proposition 2. Points (2) and (3) are obvious by construction.

Definition 4. The deformation(FRCDn)n0of A defined in Proposition 3 is called the sequence of formal Rankin–Cohen brackets on A associated to D.

The construction corresponding to Definition 4 appears in [23, Section 5]; the same article mentions (in a note added in proof ) a remark by Eholzer on the fact that it defines a formal deformation. We will need in Section 3 the following result, which is a parameterized version of the Proposition 3 for bigraded algebras.

Proposition 5. Let A=L

k,p∈ZAk,pbe a bigraded commutativeC-algebra, and D a derivation of A such that D(Ak,p)⊂Ak+2,pfor any k,p∈Z. For anyµ∈C, let us consider the sequence(FRCD,n µ)n≥0 of bilinear maps A×AA defined by bilinear extension of

FRCD,nµ(f,g)=

n

X

r=0

(−1)r

Ãk+µp+n−1 nr

`+µq+n−1 r

!

Dr(f)Dn−r(g), (11) for any fAk,p,gA`,q. Then,

(1) (FRCD,nµ)n0is a formal deformation of A.

(2) FRCD,µn (Ak,p,A`,q)⊂Ak+`+2n,p+qfor all n∈N, k,`,p,q∈Z. (3) The associated Poisson bracket is defined by bilinear extension of

FRCD,µ1 (f,g)=(k+µp)f D(g)−(`+µq)g D(f)for all fAk,p,gA`,q.

Proof. For anyµ∈C, we introduce the weighted Euler derivation ∆µ of A defined by linear extension of∆µ(f)=12(k+µp)f for all (k,p)∈Z2andfAk,p. It satisfies∆µDDµ=D. Next, we apply Proposition 2 and similar calculations to those in the proof of Proposition 3 prove that

CMD,n µ=FRCD,nµand give the results.

2. An extension-restriction method on formal deformations and Rankin–Cohen brackets on modular forms revisited

2.1.

Original problem

The classical Rankin–Cohen brackets were originally developed for modular forms and have since been used in a wide range of literature and applications. We refer for example to [24] as a primary reference. Recall that, iff andgare modular forms of respective weightskand`, the differential polynomial

RCn(f,g)=

n

X

r=0

(−1)r

Ãk+n−1 nr

`+n−1 r

!

Drz(f) Dn−rz (g) (12) is a modular form of weightk+`+2n. Here Dzis the usual normalized derivation2i1πzrelated to the complex variablez.

It is important to observe that, since the algebraM of modular forms is not stable by this derivation (the derivative of a modular form is not a modular form), the property of the sequence of classical Rankin–Cohen brackets defined by (12) to be a formal deformation ofM cannot be obtained by direct application of Proposition 3. Zagier has developed in [23] an argument to overcome this difficulty (see Paragraph 2.3 below). We extend it in the following theorem to the general formal framework of Connes–Moscovici deformations.

(7)

2.2.

A general argument about the restriction of some formal deformations

Our goal is to prove the following theorem.

Theorem 6. Consider a commutativeC-algebra R and a subalgebra A of R. Letandθbe two C-derivations of R such that∆θ−θ∆=θ. We assume that

(1) ∆(A)⊆A andθ(A)⊆A;

(2) there exists hA such as∆(h)=2h;

(3) there exists xR,xA such that∆(x)=x andθ(x)= −x2+h.

Then, the derivation D:=θ+2x∆of R satisfiesDD∆=D and the Connes–Moscovici deforma- tion(CMD,n)n≥0of R defines by restriction to A a formal deformation of A.

An obvious calculation shows that∆DD∆=D. We consider inR the Connes–Moscovici deformation (CMD,n)n≥0defined by (7). The proof of the theorem is based on the following two lemmas.

Lemma 7. The assumptions and notations are those of Theorem 6. We introduce for any integer n≥0the linear mapθ[n]:RR defined by

θ[n]=

n

X

`=0

(−1)n−`

Ãn

`

!

xn−`D`(2∆+`)〈n−`〉. (13) Then, for all n≥1, we have,

(1) θ[n+1]=θθ[n]+nhθ[n−1](2∆+n−1).

(2) θ[n](A)⊆A.

Proof. We directly check thatθ[1]=θandθ[2]=θ2+2h∆which shows (1) forn=1. We then proceed by induction. It follows fromθ(x)= −x2+hthat

θθ[n]=I+II+h·III with notations





 I=P

`(−1)n+1−`¡n

`

¢(n−`)xn−`+1D`(2∆+`)〈n−`〉, II=P

`(−1)n−`¡n

`

¢xn−`θD`(2∆+`)〈n−`〉, III=P

`(−1)n−`¡n

`

¢(n−`)xn−`−1D`(2∆+`)n−`〉. We have

III= −nX

`

(−1)n−1−`

Ãn−1

`

!

xn−1−`D`(2∆+`)〈n−1−`〉(2∆+n−1) so

θθ[n]+nhθ[n−1](2∆+n−1)=I+II.

We replaceθbyD−2x∆in II to obtain II=II1+II2with (II1= −P

`(−1)n−`¡ n

`−1

¢xn+1−`D`(2∆+`−1)〈n+1−`〉

II2= −2P

`(−1)n−`¡n

`

¢xn+1−`D`(2∆+`)〈n−`〉. We replace∆D`byD`∆+`D`to write II2=II21+II22with

(II22= −2P

`(−1)n−`¡n

`

¢`xn+1−`D`(2∆+`)〈n−`〉

II21= −2P

`(−1)n−`¡n

`

¢xn+1−`D`∆(2∆+`)〈n−`〉. By 2∆=2∆+(`−1)−(`−1) we finally find II21=II211+II212with

(II211= −P

`(−1)n−`¡n

`

¢xn+1−`D`(2∆+`−1)〈n+1−`〉

II212=P

`(−1)n−`¡n

`

¢(`−1)xn+1−`D`(2∆+`)n−`〉.

(8)

So we obtain

II1+II211=X

`

(−1)n+1−`

Ãn+1

`

!

xn+1−`D`(2∆+`−1)n+1−`〉

and

I+II22+II212=X

`

(−1)n+1−`

Ãn+1

`

!

(n+1−`)xn+1−`D`(2∆+`)n−`〉. Then, we compute

(2∆+`−1)n+1−`〉+(n+1−`)(2∆+`)n−`〉

=(2∆+`−1)(2∆+`)〈n−`〉+(n+1−`)(2∆+`)〈n−`〉

=(2∆+n)(2∆+`)〈n−`〉

=(2∆+`)〈n+1−`〉. We conclude

I+II=X

`

(−1)n+1−`

Ãn+1

`

!

xn+1−`D`(2∆+`)〈n+1−`〉,

which completes the proof of (1) of the lemma. As a consequence, we get (2).

Lemma 8. The assumptions and notations are those of Theorem 6. We introduce, for any integer n≥0, the bilinear mapΘn:R×RR defined by

Θn(f,g)=

n

X

r=0

(−1)r r!(n−r)!

¡θ[r](2∆+r)〈n−r〉¢ (f)¡

θ[n−r](2∆+nr)〈r¢

(g) (14)

for all(f,g)∈R2. Then, we have(CMD,n)n≥0=(Θn)n≥0.

Proof. We expressΘndepending onDusing (13), (14) and Lemma 7. We find Θn(f,g)=X

r,`,t

(−1)n+r−t−`

r!(n−r)!

Ãr

`

nr t

! xn−t−`

×D`(2∆+`)r−`〉(2∆+r)nr(f)Dt(2∆+t)nrt(2∆+nr)r(g).

Using (2∆+`)〈r−`〉(2∆+r)〈n−r=(2∆+`)〈n−`〉and (2∆+t)〈n−r−t〉(2∆+nr)〈r〉=(2∆+t)〈n−t〉, we obtain

Θn(f,g)=X

`,t

(−1)n−t

(n−t`)!`!t!D`(2∆+`)n−`〉(f)Dt(2∆+t)nt(g)xnt−`X

r

(−1)r−`

Ãnt` r`

! . The inner sum indexed byris equal to (1−1)n−t−`, so we obtain

Θn(f,g)=X

`

(−1)`

`!(n−`)!D`(2∆+`)〈n−`〉(f)Dn−`(2∆+n`)〈`〉(g)=CMD,n(f,g),

and the proof is complete.

Proof of Theorem 6. We know from Proposition 2 that (CMD,n )n≥0 is a formal deformation ofR but it is not clear thatCMD,n(f,g)∈ Afor f,gA. The sequence (Θn)n≥0 satisfies, by construction, thatΘn(f,g)∈Aforf,g∈Abut it is not clear from its definition that it is a formal deformation ofR. Thus, Theorem 6 follows from the equalityCMD,n =Θnproved in Lemma 8. Let us observe that the subalgebraAofRis stable by∆, is not necessarily stable byD, but is stable by

any bilinear applicationCMD,n .

Theorem 6 is formulated in terms of restriction from an algebraR to a subalgebraA. In the following number-theoretical applications, it will be applied in terms of extension fromAtoR either by polynomial extension (see Paragraph 2.3.2 for modular forms) or by localization (see Section 5 for weak Jacobi forms).

(9)

2.3.

Application to modular forms

A reference for details on this section is [13].

2.3.1. Basic facts and notations on modular and quasimodular forms

It is well known that theC-algebra of holomorphic modular formsM associated to the full modular group SL(2,Z) is the weight-graded polynomial algebra

M=C[E4, E6]= M

k∈2Z≥0 k6=2

Mk withMk= M

4i+6j=k

CEi4E6j (15)

where E4and E6are the Eisenstein series of respective weights 4 and 6. The Eisenstein series E2 is not a modular form but a quasimodular form (of weight 2 and depth 1) and the algebraM≤∞

of quasimodular forms can be described as the polynomial algebra M≤∞=M[E2]=C[E4, E6, E2]= M

k∈2Z≥0 k/2

M

s=0

Mk−2sEs2 (16) graded by the weightkand filtered by the depths(corresponding to the degree in E2).

We denote by Dzthe normalized complex derivative Dz=2i1πz. Ramanujan relations are Dz(E4)= −1

3(E6−E4E2), Dz(E6)= −1

2(E24−E6E2), Dz(E2)= − 1

12(E4−E22) (17) In particular the subalgebraMofM≤∞is not stable by the derivation DzofM≤∞. We introduce the Serre derivative, which is the derivation Se ofM≤∞defined by linear extension of

Se(f)=Dz(f)− k

12E2f for anyf ∈M≤∞of weightk. (18) We have

Se(E4)= −1

3E6, Se(E6)= −1

2E24, Se(E2)= − 1

12(E22+E4). (19) In particular M is stable by Se and the restriction of Se to M =C[E4, E6] is the derivative θ= −13E6E412E24E6.

2.3.2. Application of the extension-restriction method to modular forms

We apply here the general result of Theorem 6 to give another proof of the following well- known result.

Proposition 9. The sequence of classical Rankin–Cohen brackets(12)defines a formal deforma- tion of the algebraM.

Proof. We choose forAthe algebra of modular formsM=C[E4, E6], forθthe restriction toM of the Serre derivative Se and for∆the weight-derivative defined by∆(f)=k2f for anyf ∈Mk. We have∆θ−θ∆=θ. We introduce

h= − 1

144E4∈M and x= 1

12E2∈M≤∞.

We consider the polynomial extensionR=M[x], which is, by (16), the algebra of quasimodular formsM≤∞. We extend∆andθtoM≤∞by∆(x)=xandθ(x)= −x2+h. In other words

∆(E2)=E2 and θ(E2)= − 1 12E22− 1

12E4. Then, the extensions∆andθare the derivations ofM≤∞such that

∆(f)=k

2f for anyf ∈M≤∞of weightk, and θ=Se onM≤∞,

and they also satisfy∆θ−θ∆=θ. By (18) the derivationθ+2x∆ofM≤∞is equal to the derivative Dz. Applying Theorem 6, we deduce that the formal deformation (CMDz,)n≥0ofM≤∞defines

(10)

by restriction a formal deformation ofM. On the one handCMDnz,∆=FRCDnzonM≤∞by (10). On the other hand, the restriction ofFRCDnztoMis the classical Rankin–Cohen bracketRCndefined by (12). So we get a proof of the property that (RCn)n≥0is a formal deformation ofM. As observed at the end of the proof of Theorem 6, the algebraMis stable by∆, is not stable by Dz, but is stable by any Rankin–Cohen bracketsFRCDnz=CMDnz,∆.

2.3.3. Serre Rankin–Cohen brackets

Another strategy to overcome the fact thatM is not stable by the derivation Dzis to change the derivation and apply directly Proposition 3 to a weight 2 homogeneous derivation ofM, for instance the Serre derivation Se onM. So we define with (9) the Serre–Rankin–Cohen brackets (SRCn)n0=(FRCSen)n0onMby bilinear extension of

SRCn(f,g)=

n

X

r=0

(−1)r

Ãk+n−1 nr

`+n−1 r

!

Ser(f) Sen−r(g) (20) for (f,g)∈Mk×M`. This is by Proposition 3 a formal deformation ofMsatisfying

n≥0 SRCn(Mk,M`)⊂Mk+`+2n. (21) It follows from the definition thatSRC0=RC0andSRC1=RC1onM. The following proposition clarifies the relationship between the Serre–Rankin–Cohen brackets and the usual Rankin–Cohen brackets (12).

Proposition 10. The two formal deformations(RCn)n≥0and(SRCn)n≥0ofMare not isomorphic.

Proof. Ifϕis an isomorphism between the formal deformations (M, (SRC)n≥0) and (M, (RC)n≥0) then it is a Poisson automorphism of (M,RC1) sinceSRC1=RC1. Thenϕ=IdMby Proposition 7 of [8]. This is contradicted by

SRC2(f,g)=RC2(f,g)+ 1

288k`(k+`+2)f gE4 for (f,g)∈Mk×M`. (22) 2.3.4. Formal deformations on quasimodular forms

Using (19), the formal Rankin–Cohen brackets (SRCn)n≥0onMdefined by (20) extend canon- ically toM≤∞. Many other formal deformations ofM≤∞can be constructed using Proposition 2 and the systematic study of such analogues of Rankin–Cohen brackets onM≤∞is the subject of the article [8] under the additional arithmetical assumption of depth conservation. It is not sat- isfied by (SRCn)n≥0because for instanceSRC1(E2, E4)=13(E4E22−2 E6E2+E24) is of depth 2. How- ever, the deformations ofM≤∞appearing in Theorems B and D of [8] will play some role in the study of the deformations on the Jacobi forms in the following section.

3. The algebra of weak Jacobi forms 3.1.

The notion of weak Jacobi form

The aim of this part is to gather the notions we shall need on weak Jacobi forms. The main reference is [9]. LetH be the upper half plane,k an integer andma nonnegative integer. The Jacobi group SL(2,Z)J=SL(2,Z)nZ2acts on functionsf :H ×C→Cas follows: ifγa b

c d

¢∈ SL(2,Z) and (λ,µ)∈Z2, then

fkk,m

γ, (λ,µ)¢ )(τ,z)= (cτ+d)kexp

µ 2πim

µ

c(z+λτ+µ)2

+d +λ2τ+2λz+λµ

¶¶

f

µaτ+b

+d,z+λτ+µ +d

(11)

for all (τ,z)∈H×C.

A weak Jacobi form of weightkand indexmis a holomorphic functionΦ:H×C→Cinvariant by the actionkk,mof the Jacobi group and whose Fourier expansion has the shape

Φ(τ,z)=

+∞X

n=0

X

r∈Z r2≤4nm+m2

c(n,r)e2πi(nτ+r z). (23)

The vector spaceJk,mof such functions is finite dimensional. We identify functions onH×Cthat are not depending on the second variable with functions onH. On this subspace, the actionkk,0 of SL(2,Z)Jinduces the classical modular action of SL(2,Z) denoted by|k. The spaceJk,0=Mk

is the space of holomorphic modular forms of weightkon SL(2,Z) defined in 2.3.

The principal object of our study is the bigraded algebra of weak Jacobi forms J= M

k∈2Zm≥0

Jk,m. (24)

3.2.

Generators of the algebra of the weak Jacobi forms

The algebraJ is a polynomial algebra on two generators over the algebraM of modular forms.

We describe these two generators.

Form=0, we already mentioned in (15) that the Eisenstein series E4 and E6generate the algebra of modular forms:M =C[E4, E6]. Ifm6=0, the Eisenstein series Ek,mofJk,mof even weightk≥4 and indexmis

Ek,m(τ,z)=1 2

X

(c,d)∈Z2 (c,d)=1

X

λ∈Z

(cτ+d)−kexp µ

2iπm µ

λ2aτ+b

+d +2λzc z2 +d

¶¶

.

Let us define

Φ10,1=1441 (E6E4,1−E4E6,1)∈J10,1, Φ12,1=1441 (E24E4,1−E6E6,1)∈J12,1, and

∆= 1

1728(E34−E26)∈M12. We define inJthe elements

A=Φ10,1

∆ ∈J−2,1 and B=Φ12,1

∆ ∈J0,1. (25) By [9, Theorem 9.3], we have

J=M[A, B]=C[E4, E6, A, B].

Using the algorithm proven in [9, p. 39], we can compute the Fourier expansion ofΦ10,1andΦ12,1

and deduce the ones of A and B.

3.3.

A localization of the algebra of the weak Jacobi forms

We introduce the algebra

K =C[E4, E6, A±1, B]⊃J. (26) This is the localization ofJ with respect to the powers of A. The notions of weight and index naturally extend toK defining a bigraduation

K = M

k∈2Zm∈Z

Kk,m. (27)

(12)

We set

F2=B A−1, (28)

which satisfies

K =C[E4, E6, F2, A±1]=C[E4, E6, F2][A±1] and leads to introduce the subalgebra

Q=C[E4, E6, F2].

The elements ofQappear as the elements inK of index zero. The following table summarizes the weights and indices attached to these different generators.

E4 E6 A B F2

weight 4 6 -2 0 2

index 0 0 1 1 0

3.4.

Number-theoretic interpretation, relationship with quasimodular forms

The function F2has a number-theoretic meaning since F2= − 3

π2 (29)

whereis the Weierstraß function [9, Theorem 3.6] and henceQis the subalgebra generated by modular forms and the Weierstraß function

Q=M[]. (30) Another arithmetical point of view consists in seeingQ as a formal analogue to the algebra M≤∞=M[E2] of quasimodular forms. The algebra isomorphism involved is

ω:Q→M≤∞, P(E4, E6, F2)7→P(E4, E6, E2). (31) The degree related to F2of anyf ∈Qis the depth of the quasimodular formω(f) and we have

M⊂M≤∞'Q⊂K.

The isomorphismωand (30) emphasize that, from an algebraic point of view, the Weierstraß functionis similar to the Eisenstein series E2.

3.5.

Summary

We can summarize the relationships between the different function algebras under consideration by the following diagram:

J=C[E4, E6, A, B]  // K =C[E4, E6, F2, A±1]

M=C?[EOO 4, E6]  //

Q=C[E4, E6?OO, F2]'M≤∞

This is the framework for our study of formal deformations ofJ.

(13)

3.6.

Problem

Our goal is to construct families of Rankin–Cohen brackets on weak Jacobi forms,

(1) which are deformations of the algebraJ(this was not the case for some other construc- tion, appearing in the literature),

(2) which extend Rankin–Cohen brackets on modular forms,

(3) which are coherent with the weight and the index (that is preserve the index and increase the weight by two).

Two main methods can be used following the two points of view illustrated above in the case of modular forms.

(1) The first strategy is to start from a derivation ofJand to use the canonical construction of Proposition 5. This method gives rise in Section 4 to a family of deformations ofJ extending the Serre–Rankin–Cohen brackets onM (defined in Section 2.3.3).

(2) The second one is to apply the extension-restriction process of Theorem 6. So we start from a suitable derivationDof the extensionK ofJwhich doesn’t stabilizeJ. We thus construct in Section 5 a family of deformations stabilizingJ and extending the classical Rankin–Cohen brackets onM (defined in Section 2.1).

4. A first family of Rankin–Cohen deformations on weak Jacobi forms

In this section, we define and study a family of Rankin–Cohen deformations on J whose restriction toMis the sequence of Serre–Rankin–Cohen brackets.

4.1.

Construction of the deformations

We first define an extension of the Serre derivative to the weak Jacobi forms. For anya,b∈Cwe denote by Sea,bthe derivation ofJ=C[E4, E6, A, B] defined by

Sea,b(E4)= −1

3E6, Sea,b(E6)= −1

2E24, Sea,b(A)=aB, Sea,b(B)=bE4A . (32) This is by (19) the only way to extend the Serre derivation Se onM into a derivation ofJ preserving the index and increasing the weight by two.

We use the general process described in Section 1.3 to construct formal Rankin–Cohen brack- ets onJ.

Theorem 11. For all(a,b,c)∈C3, for any n≥0, let{·,·}[a,b,c]n be the bilinear map fromJ×J to Jdefined by bilinear extension of

{f,g}[a,b,c]n =

n

X

r=0

(−1)r

Ãk+c p+n−1 nr

`+c q+n−1 r

!

Sera,b(f) Sen−ra,b (g) (33) for all f ∈Jk,pand g∈J`,q. Then,

(1) The sequence³

{·,·}[a,b,c]n ´

n0is a formal deformation ofJ.

(2) We have{Jk,p,J`,q}[a,b,c]n ⊂Jk+`+2n,p+qfor all k,`∈2Z,p,q,n≥0.

(3) The subalgebra of modular formsM is stable by{·,·}[a,b,c]n and its restriction toM×Mis the Serre–Rankin–Cohen bracketSRCn.

Proof. For any (a,b)∈C2, the derivation Sea,b ofJ satisfies Sea,b(Jk,m)⊂Jk+2,m. Then, (1) and (2) follow from Proposition 5 since {·,·}[a,b,c]n =FRCSena,b,c. If f,g are modular forms of respective weightsk,`, we havep =q=0 in formula (33) which doesn’t depend onc in this case. Moreover Sea,b(f)=Se(f) and Sea,b(g)=Se(g) by (32) and (19). Hence {f,g}[a,b,c]n does not

depend on (a,b) and is by (20) equal toSRCn(f,g).

(14)

4.2.

Classification and separation results

The formal deformations ({·,·}[a,b,c]n )n≥0depend on three parameters. We can classify them, up to a suitable specialization of the notion of isomorphic deformations (see paragraph 1.1.2) with respect to the arithmetical context.

Definition 12. Two formal deformations³

{·,·}[a,b,c]n ´

n≥0and³

{·,·}[an0,b0,c0]´

n≥0ofJ are modular- isomorphic if there exists aC-linear bijective mapφ:J→J such that

(1) φ({f,g}[a,b,c]j )={φ(f),φ(g)}[aj 0,b0,c0]for all j≥0and f,g∈J.

(2) φpreserves the index and the weight of homogeneous weak Jacobi forms

In particular φ is a C-algebra automorphism of J and a Poisson isomorphism from (J, {·,·}[a,b,c]1 ) to (J, {·,·}[a10,b0,c0]).

Lemma 13. If two formal deformations ({·,·}[a,b,c]n )n≥0 and ({·,·}[an0,b0,c0])n≥0 are modular- isomorphic, then c=c0, and there existsξ∈Csuch that a0=ξa and b0=ξ−1b.

Proof. Let φ: J → J be as in Definition 12. It induces a Poisson isomorphism from

³J, {·,·}[a,b,c]1 )´ to³

J, {·,·}[a10,b0,c0]´

. We haveφ(M)⊂Mby (2) andRC1=SRC1={·,·}[a,b,c]1 onM. Then, the restriction ofφtoM is a Poisson modular automorphism of (M,RC1). By [8, Proposi- tion 7], this is the identity ofM.

Let f ∈Jk,p. Then,φ(f)∈Jk,p. The restriction of Sea,btoM is Se. The kernel of Se isC[∆] (see, for example, [8, Proposition 8]) andφ(∆)=∆. We deduce that

φ³

{f,∆}[a,b,c]1 ´

= −12∆φ¡

Sea,b(f)¢

and {φ(f),φ(∆)}[a10,b0,c0]= −12∆Sea0,b0¡ φ(f)¢ and hence

φ◦Sea,b=Sea0,b0◦φ. (34) It follows that, for allf ∈Jk,pandg∈J`,q, we have

{φ(f),φ(g)}[a10,b0,c0]=φ¡

(k+c0p)fSea,b(g)−(`+c0q)gSea,b(f)¢ and (1) leads to

(c0c)¡

p fSea,b(g)−q gSea,b(f)¢

=0.

We apply this equality tof =A E4andg=E6to obtainc0=c. Letλandµbe defined inCby φ(A)=λA andφ(B)=µB. Equation (34) applied to A and B gives=a0λandb0µ=. The

proof is complete settingξ=µ/λ.

Theorem 14. Let (a,b,c) ∈ C3. The formal deformation ({·,·}[a,b,c]n )n≥0 of J is modular- isomorphic to one of the following formal deformations,

(1) The formal deformation({·,·}[1,bn 0,c0])n≥0for some(b0,c0)∈C2. (2) The formal deformation({·,·}[0,1,cn 0])n≥0for some c0∈C. (3) The formal deformation({·,·}[0,0,cn 0])n0for some c0∈C.

These deformations are pairwise non modular-isomorphic for different values of the parameters.

Proof. For any (λ,µ)∈C2, denote byφλ,µtheC-algebra automorphism ofJ fixing E4and E6 and such thatφλ,µ(A)=λA andφλ,µ(B)=µB. By the same calculations as in proof of Lemma 13 we have, for any (a,b,a0,b0)∈C4,

(φλ,µ◦Sea,b=Sea0,b0◦φλ,µ) if and only if (aµ=a0λand=b0µ). (35) By (33), if one of the equivalent conditions of (35) is satisfied, then the formal deformations {·,·}[a,b,c]and {·,·}[a0,b0,c0]are isomorphic. Since it follows from (35) thatφa,1◦Sea,b=Se1,ab◦φa,1

for anya6=0, andφ1,b◦Se0,b=Se0,1◦φ1,bfor anyb6=0, the proof that ({·,·}[a,b,c]n )n≥0is modular- isomorphic to one of given formal deformations is complete. The separation of the different cases up to modular isomorphism follows from a direct application of Lemma 13.

Références

Documents relatifs

Thus each mathematical symbol (G, L,. .) has a single definite meaning inside each exercise, but might have different meanings from an exercise to another.. It is not required to

—The existence and construction of vector-valued modular forms (vvmf) for any arbitrary Fuchsian group G, for any representation ρ : G −→ GL d ( C ) of finite image can be

The main technique is a motivic variation of theorems of Mazur, Ogus, Illusie and Nygaard on the Katz conjecture (according to which the Newton polygon of Frobenius on

locally homogenous complex manifolds are the moduli spaces for that class of varieties... Returning to the situation in which 0393BD is a moduli space, in general 0393BD

Suppose R is a right noetherian, left P-injective, and left min-CS ring such that every nonzero complement left ideal is not small (or not singular).. Then R

Ce but est atteint selon la presente invention grace a un detecteur de pression caracterise par le fait que la jante de la roue est realisee en un materiau amagnetique, le capteur

9 As outlined in these brief summaries, the role of science in Professional Foul and Hapgood has much in common with Stoppard’s use of chaos theory in Arcadia, since catastrophe

Apart from discrete groups and duals of compact groups, the first test examples for a quantum Property RD should be the free quantum groups introduced by Wang [6], which are