• Aucun résultat trouvé

arXiv:1707.04804v1 [math.NT] 16 Jul 2017

N/A
N/A
Protected

Academic year: 2022

Partager "arXiv:1707.04804v1 [math.NT] 16 Jul 2017"

Copied!
38
0
0

Texte intégral

(1)

CRITICAL POINTS OF THE CLASSICAL EISENSTEIN SERIES OF WEIGHT TWO

ZHIJIE CHEN AND CHANG-SHOU LIN

ABSTRACT. In this paper, we completely determine the critical points of the normalized Eisenstein seriesE2(τ)of weight 2. AlthoughE2(τ) is not a modular form, our result shows thatE2(τ)has at most one critical point in every fundamental domain ofΓ0(2). We also give a criteria for a fundamental domain containing a critical point ofE2(τ). Furthermore, under the M ¨obius transformation ofΓ0(2)action, all critical points can be mapped into the basic fundamental domainF0 and their images are contained densely on three smooth curves. A geometric interpretation of these smooth curves is also given. It turns out that these smooth curves coincide with the degeneracy curves of trivial critical points of a multiple Green function related to flat tori.

CONTENTS

1. Introduction 1

2. Zeros of pre-modular forms 6

3. Existence and uniqueness ofτ(C) 16

4. Critical points ofη1(τ)or equivalentlyE2(τ) 24 5. Geometric interpretation and smoothness of the curves 34

Appendix A. Application 35

References 37

1. INTRODUCTION

The Jacobi theta functions, the Eisenstein series and the Weierstrass func- tions arise in numerous theories and applications of both mathematics and physics. Since their discovery in the early 19th century, the mathematical foundation of elliptic functions was subsequently developed. It turns out that, besides their applications in science, these special functions in the el- liptic function theory are rather deep objects by themselves.

The main goal of this paper is to completely locate all the critical points of the classical function η1(τ) or equivalently the normalized Eisenstein seriesE2(τ)of weight 2. Throughout the paper, we use the notationsR+ = (0,+), ω1 = 1, ω2 = τ, ω3 = 1+τ and Λτ = Z+Zτ, where τH= {τ|Imτ > 0}. Let℘(z) =℘(z|τ)be the Weierstrass℘-function with

1

arXiv:1707.04804v1 [math.NT] 16 Jul 2017

(2)

periodsΛτ, defined by

℘(z|τ):= 1

z2 +

ωΛτ\{0}

1

(zω)2ω12

. Letζ(z) =ζ(z|τ):=−Rz

℘(ξ|τ)be the Weierstrass zeta function, which is odd and has two quasi-periodsηk(τ):=2ζ(ω2k|τ),k=1, 2:

(1.1) η1(τ) =ζ(z+1|τ)−ζ(z|τ), η2(τ) =ζ(z+τ|τ)−ζ(z|τ). The well-known Legendre relation givesη2(τ) =τη1(τ)−2πi. In the liter- ature,η1(τ)is known as the Weierstrass eta function (cf. [1]), which is just a multiple of the normalized Eisenstein seriesE2(τ)of weight 2:

3

π2η1(τ) =E2(τ):= 3 π2

m=

n=

1 (+n)2 (1.2)

=1−24

n=1

bne2nπiτ, bn=

1d|n

d.

Conventionally, ∑ means to sum over (n,m) ∈ Z2\ {(0, 0)}. Besides, η1(τ)is also connected with Dedekind eta function

η(τ):=eπiτ12

n=1

(1−e2nπiτ)

through the following logarithmic differential formula (cf. [1, p.696]):

1 η(τ)η

(τ) = i

η1(τ).

Unlike the other Eisenstein series of weight 2kwithk2,E2(τ)is not a modular form. Its transformation under the action ofSL(2,Z)satisfies (1.3) E2

+b +d

= (+d)2E2(τ)− 6icπ (+d),

a b c d

SL(2,Z). Thus it is surprising that its critical points possess the following property.

Theorem 1.1. Let F be a fundamental domain ofΓ0(2). Then E2(τ)has at most one critical point in F.

HereΓ0(2)is the congruence subgroup ofSL(2,Z)defined by Γ0(2):= a b

c d

SL(2,Z)

c0 mod 2

.

Recently, there are some works studying the zeros ofE2(τ); see [9, 16] and references therein. As far as we know, there seems no results concerning the critical points ofE2(τ)in the literature.

(3)

In view of Theorem 1.1, a natural question is: What are those fundamental domains containing critical points? To answer this question, we introduce our

”basic” fundamental domainF0ofΓ0(2):

F0 :={τH|06 Reτ61 and|z12|> 1

2}. Note that givenγ=

a b c d

Γ0(2)/{±I2}(i.e. considerγandγto be the same),

γ(F0):=nγ·τ:= ++bdτF0

o= (−γ)(F0)

is another fundamental domain of Γ0(2). Moreover, γ(F0) = F0+m for somemZif and only ifc=0.

Theorem 1.2. Let F = γ(F0) be a fundamental domain of Γ0(2) with γ = a b

c d

Γ0(2)/{±I2}. Then F contains a critical point of E2(τ)if and only if c6=0.

By Theorem 1.2, we can transform every critical point of E2(τ) via the M ¨obius transformation ofΓ0(2)action to locate it in F0. Denotethe collec- tion of such corresponding points in F0byC, which consists of infinitely many points. A fundamental question is:What is the geometry of the setC?

Surprisingly, it turns out thatCwill locate onthree smooth curvesτ(C)in F0, which are parameterized byCR\ {0, 1}via the following identity

(1.4) C=τ2πi

η1(τpg2(τ)/12, τF0.

Hereg2(τ) =60G4(τ)is the well-known invariant coming from

(z|τ)2 =4℘(z|τ)3g2(τ)℘(z|τ)−g3(τ),

and G4(τ) is the Eisenstein series of weight 4. We will prove in Section 2 that for each CR\ {0, 1}, there isa unique τ(C) ∈ F0 such that (1.4) holds.1 Consequently, the parametrization (1.4) will give three smooth curves

C0:={τ(C)|C∈(0, 1)},

C:={τ(C)|C∈(−, 0)}, C+:= {τ(C)|C∈(1,+)}.

The relation between (1.4) andη1(τ)comes from the classical formula (see e.g. [1], or from Ramanujan’s formula: E2(τ) = πi6(E22E4))

(1.5) η1(τ) = i η1(τ)2121 g2(τ).

1Note that the RHS of (1.4) is actually a multi-valued function, please see Theorem 3.1 for the precise definition of this uniqueτ(C).

(4)

Theorem 1.3. Letτ(C)be defined by (1.4) for CR\{0, 1}. Then (1.6) C=

τ(cd)

a b c d

Γ0(2)/{±I2}with c6=0

⊂ C∪ C0∪ C+. Furthermore, the closure ofCin F0is precisely the union of the three smooth curves:

(1.7) C ∩F0 =C∪ C0∪ C+.

Remark1.4. In fact, we will proveτ(C) ∈ F˚0, where ˚F0 = F0\∂F0 denotes the set of interior points of F0. Given γ =

a b c d

Γ0(2)/{±I2} with c6=0, we will prove in Theorem 4.1 that the unique critical point ofE2(τ)in γ(F0)is precisely (

d c )+b (d

c )+dγ(F˚0). Givenγj =

aj bj

cj dj

Γ0(2)/{±I2} with cj 6= 0 such that γ1 6= ±γ2, we haveγ1(F˚0)∩γ2(F˚0) = (note that γ1(∂F0)∩γ2(∂F0)6=may happen) and so

a1τ(cd1

1 ) +b1

c1τ(cd1

1 ) +d1

6= a2τ(d2c

2 ) +b2

c2τ(cd2

2 ) +d2

.

Therefore, the map fromCto the set of critical points ofE2(τ)is one-to-one.

The above results completely locate all the critical points of the Eisenstein seriesE2(τ)or equivalentlyη1(τ). To the best of our knowledge, such fun- damental results have not appeared in the literature and are new. We be- lieve that they will have important applications. For example, we consider τ = 12+ibwithb > 0. Thenη1(τ) ∈ R. In order to study the behavior of the Green function on rhombus tori, Wang and the second author [12] con- sidered the monotone property ofη1(τ)and their numerical computation [12, Figure 2] suggests thatη1should increase from 0 to someb0and then decrease afterb0, but they can not prove this assertion in [12] because (1.2) implies

3

π2η1(12+ib) =1−24

n=1

(−1)nbne2nπb, bn =

1d|n

d>0,

from which it seems difficult to obtain the monotone property shown in [12, Figure 2]. Now this assertion is confirmed by the following corollary.

Corollary 1.5. There exists b0 ∈ (245,213)such that η1(12 +ib)is strictly in- creasing for b∈(0,b0)and strictly decreasing for b∈(b0,+).

One of our motivations of studying critical points ofη1(τ)comes from the Green function on flat tori. LetEτ := C/(Z+Zτ)be a flat torus and G(z) =G(z;τ)be the Green function on the torusEτ:

G(z;τ) =δ01

|Eτ| onEτ,

Z

Eτ

G(z;τ) =0,

(5)

whereδ0is the Dirac measure at 0 and|Eτ|is the area of the torus Eτ. See [12] for a detailed study ofG(z;τ). In [2, 13, 14], Chai, Wang and the sec- ond author introduced a multiple Green function Gn, nN. Geometri- cally, any critical point ofGnis closely related to bubbling phenomenon of nonlinear partial differential equations with exponential nonlinearities in two dimension; see [2, 14] for typical examples. Thus, understanding the critical points ofGnis important for applications.

For the casen=2, the multiple Green functionG2is defined by (1.8) G2(z1,z2;τ):=G(z1z2;τ)−2G(z1;τ)−2G(z2;τ), where 06= z1 6=z26=0. A critical point(a1,a2)ofG2satisfies

2∇G(a1;τ) =∇G(a1a2;τ), 2∇G(a2;τ) =∇G(a2a1;τ). Clearly if (a1,a2)is a critical point then so does (a2,a1), and we consider such two critical points to bethe same one. A critical point(a1,a2)is called a trivial critical pointif

{a1,a2}= {−a1,a2} in Eτ.

Recallω1 = 1,ω2 = τandω3 = 1+τ. It is known [14] that G2 has only five trivial critical points {(12ωi,12ωj)|i 6= j} and {(q±,−q±)|℘(q±|τ) =

±pg2(τ)/12}, and the Hassian at(q±,−q±)is given by detD2G2(q±,−q±;τ)

=3|g2(τ)|

4Imτ|℘(q±|τ) +η1(τ)|2Im

τ2πi

η1(τpg2(τ)/12

. (1.9)

From here and (1.4), we will prove in Section 5 that the three curves co- incide with degeneracy curves ofG2. The Hassian at(12ωi,12ωj)is related to the critical points of the classical function ek(τ) := ℘(ω2k|τ), {i,j,k} = {1, 2, 3}. We will study the critical points ofek(τ)in another paper.

Our proof of the existence and uniqueness ofτ(C)relies on apre-modular form Zr,s(2)(τ) of weight 3 introduced in [13]. See also [7]. For each pair (r,s)∈R2\12Z2,Z(r,s2)(τ)is defined by

Z(r,s2)(τ):=Zr,s(τ)33℘(r+|τ)Zr,s(τ)−℘(r+|τ), whereZr,s(τ)is introduced by Hecke [10]:

Zr,s(τ):=ζ(r+|τ)−1(τ)−2(τ). (1.10)

Indeed, if(r,s) ∈Q2,Zr,s(τ)is the well-known Eisenstein series of weight 1 with characteristic(r,s); see [8, p.139]. It is not difficult to see thatZr,s(τ) is a modular form of weight 1 with respect toΓ(N)if(r,s)is a N-torsion point, soZ(r,s2)(τ)is a modular form of weight 3. See Section 2. The impor- tance ofZr,s(2)(τ)lies on the fact that at any zeroτ0of Z(r,s2)(·), the pair(r,s) contains all the monodromy data of the classical Lam´e equation

(1.11) y′′(z) = [n(n+1)℘(z|τ0) +B]y(z), n=2

(6)

for someBC; see [13, Theorem 4.3]. Therefore, it is important to study the zero ofZr,s(2)(·), which has not been settled yet. In this paper, we study the zero structure of Z(r,s2)(·). Define four open triangles (see Figure 1 in Section 2):

(1.12)

0 :={(r,s)|0<r,s< 12, r+s > 12},

1 :={(r,s)| 12 <r <1, 0<s< 12, r+s>1},

2 :={(r,s)| 12 <r <1, 0<s< 12, r+s<1},

3 :={(r,s)|r>0, s>0, r+s< 12}.

Theorem 1.6. Let(r,s)∈ [0, 1]×[0,12]\12Z2. Then Zr,s(2)(τ) =0has a solution τ in F0 if and only if (r,s) ∈ △1∪ △2∪ △3. Furthermore, for any (r,s) ∈

1∪ △2∪ △3, the zeroτF0is unqiue and satisfiesτF˚0.

Remark thatZr,s(2)(τ) ≡ if (r,s) = (0, 0). To prove Theorems 1.2-1.3, we will ”blow up”Zr,s(2)(τ)by considering lims01

sZ(2Cs,s) (τ), CR, and the existence and uniqueness of τ(C)will follow from that of the zero of Z(2Cs,s) (τ)ass0.

The rest of this paper is organized as follows. Theorem 1.6 will be proved in Section 2. In Section 3, we apply Theorem 1.6 to prove the existence and uniqueness of τ(C). See Theorem 3.1. In Section 4, we give the detailed proofs of our main results Theorems 1.1-1.3 and Corollary 1.5. Some precise characterizations of the three curves (see Theorem 4.2) will also be given.

In Section 5, we introduce the relation between the three curves and the degeneracy curve ofG2and prove the smoothness of the curves. Finally in Appendix A, we give another application of Theorem 1.6.

2. ZEROS OF PRE-MODULAR FORMS

This section is devoted to the proof of Theorem 1.6. First we recall the modularity of g2(τ)and℘(z|τ). Given any

a b c d

SL(2,Z), it is well known that

(2.1) g2(++bd) = (+d)4g2(τ),

z

+d

++bd

= (+d)2℘(z|τ). From here we can obtain

ζ

z +d

++db

= (+d)ζ(z|τ), and so

(2.2) η2(++bd) η1(++bd)

!

= (+d) a b

c d

η2(τ) η1(τ)

.

In the rest of this paper, we will freely use the formulas (2.1)-(2.2).

(7)

As in [7, 13], for any(r,s)∈R2, we define pre-modular forms Zr,s(τ):=ζ(r+|τ)−1(τ)−2(τ)

=ζ(r+|τ)−(r+)η1(τ) +2πis, (2.3)

(2.4) Z(r,s2)(τ):=Zr,s(τ)33℘(r+|τ)Zr,s(τ)−℘(r+|τ).

Since ζ(z|τ) has simple poles at the lattice points Λτ, both Zr,s(τ) and Zr,s(2)(τ)≡provided(r,s)≡0 modZ2. If(r,s)∈ 12Z2\Z2, where

1

2Z2 :={(m2,n2)|m,nZ},

then (1.1) and the oddness ofζ(z|τ)implyZr,s(τ) ≡0 and soZ(r,s2)(τ)≡ 0, where we used℘(ω2k) = 0. Therefore, we only consider(r,s) ∈ R2\12Z2. Then bothZr,s(τ)and Zr,s(2)(τ)are holomorphic in H, and it is easy to see that the following properties hold:

(i) Zr,s(τ) =±Zm±r,n±s(τ)and henceZr,s(2)(τ) =±Zm(2±)r,n±s(τ)for any (m,n)∈Z2.

(ii) Zr,s(τ) = (+d)Zr,s(τ)and henceZr(2,s)(τ) = (+d)3Zr,s(2)(τ) for any γ =

a b c d

SL(2,Z), where τ = γ·τ := ++db and (s,r) = (s,rγ1.

In particular, when (r,s) ∈ QN is a N-torsion point for some NN3, where

(2.5) QN :=n kN1,kN2

gcd(k1,k2,N) =1, 0≤k1,k2N1o, andγΓ(N) := {γSL(2,Z)|γI2modN}, then(r,s) ≡ (r,s)mod Z2. In other words, if(r,s)∈ QN, then

Zr,s

+b +d

= (+d)Zr,s(τ), Zr,s(2)

+b +d

= (+d)3Zr,s(2)(τ)

hold for anyγ = a b

c d

Γ(N), namelyZr,s(τ)andZr,s(2)(τ)aremodular formsof weight 1 and 3, respectively, with respect to the principal congru- ence subgroup Γ(N). Due to this reason, Zr,s(τ) and Zr,s(2)(τ) are called pre-modular formsin this paper as in [13].

We are interested in the zero structures ofZr,s(τ)andZr,s(2)(τ)for(r,s)∈ R2\12Z2. By property (ii), we can restrictτin the fundamental domain F0

ofΓ0(2):

F0 :={τH|06 Reτ61 and|z12|> 1

2},

and by (i), we only need to consider(r,s) ∈ [0, 1]×[0,12]\12Z2. Recall the four open triangles defined in (1.12) (see Figure 1). Clearly[0, 1]×[0,12] =

3k=0k. The following result was proved in [6].

(8)

0 1

3 2

0 0.5 1

0.5

r s

FIGURE1. The four open trianlges△k.

Theorem A.[6]Let(r,s)∈ [0, 1]×[0,12]\12Z2. Then Zr,s(τ) =0has a solution τ in F0 if and only if(r,s) ∈ △0. Furthermore, for any (r,s) ∈ △0, the zero τF0is unqiue and satisfiesτF˚0.

In this paper, we will prove an analogous result forZr,s(2)(τ).

Theorem 2.1(=Theorem 1.6). Let(r,s)∈[0, 1]×[0,12]\12Z2. Then Zr,s(2)(τ) = 0has a solutionτin F0 if and only if(r,s) ∈ △1∪ △2∪ △3. Furthermore, for any(r,s)∈ △1∪ △2∪ △3, the zeroτF0is unqiue and satisfiesτF˚0.

UnlikeZr,s(τ), Theorem 2.1 shows an interesting phenomena forZr,s(2)(τ). For example,Zr,s(2)(τ)has zeros inF0for(r,s)∈ △1∪ △2, but it has no zeros inF0for(r,s)∈12.

The rest of this section is to prove Theorem 2.1. The reason why we choose the fundamental domain F0 will be clear from the proof, particu- larly Lemma 2.3. The basic strategy is similar to that of proving Theorem A in [6]. However, the argument is more involved and new techniques are needed. For example, for the same assertion of the pre-modular forms hav- ing no zero inF0for(r,s)∈ ∪3k=0k, it is a trivial consequence of the same assertion for(r,s)∈ △1∪ △2∪ △3in Theorem A; but obviously, this is not the case in Theorem 2.1.

First we need the following important results of PDE aspect.

Theorem B.[13, 4]

(1) [13]The mean field equation

(2.6) ∆u+eu=16πδ0 on Eτ :=C/(Z+Zτ)

has solutions if and only if there exists(r,s)∈ R2\12Z2such thatτis a zero of Zr,s(2)(·).

(2) [4]Ifτ∈ {eπi/3} ∪iR+, then equation (2.6) has no solutions.

(9)

Remark2.2. In [2, 13], Chai, Wang and the second author studied the fol- lowing singular Liouville equation

(2.7) ∆u+eu =8nπδ0 onEτ,

where nN. The solvability of (2.7) depends essentially on the moduli τof the flat torusEτ and is intricate from the PDE point of view. To settle this challenging problem, they studied it from the viewpoint of algebraic geometry. They developed a theory to connect this PDE problem with the Lam´e equation (1.11) and pre-modular forms. In particular, Wang and the second author [13] proved the existence of a pre-modular form Z(r,sn)(·) of weight n(n2+1) such that (2.7) onEτ has solutions if and only ifZ(r,sn)(τ) = 0 for some(r,s)∈R2\12Z2. Theorem B-(1) is a special case of this statement forn=2. Theorem B-(2) is a purely PDE result. We will see that Theorem B plays a crucial role in the proof of Lemma 2.3 and hence Theorem 2.1. This is the only place related to the PDE result.

Lemma 2.3. Let (r,s) ∈ [0, 1]×[0,12]\12Z2. Then Z(r,s2)(τ) 6= 0 for any τ ∈ {eπi/3} ∪(∂F0H).

Proof. It does not seem that this assertion could be obtained directly from the expression (2.4) ofZr,s(2)(τ). Indeed, this lemma is a consequence of the result ofnonlinear PDEs(i.e. Theorem B).

Givenτ ∈ {eπi/3} ∪(∂F0H). If τ ∈ {eπi/3} ∪iR+, then Theorem B implies Zr,s(2)(τ) 6= 0 for any (r,s) ∈ R2\12Z2. If τiR++1, then by applyingγ=

1 −1

0 1

in property (ii), we have thatτ1iR+and Zr,s(2)(τ) =Zr(+2)s,s(τ1)6=0 for any(r,s)∈R2\12Z2. If|τ12|= 12, then again by applyingγ=

1 0

1 1

in property (ii) we see that 1ττiR+and

(1−τ)3Zr,s(2)(τ) =Zr,r(2)+s(1ττ)6=0 for any(r,s)∈ R2\12Z2.

This completes the proof.

RecallingQN in (2.5), we define

(2.8) MN(τ):=

(r,s)QN

Z(r,s2)(τ).

By properties (i)-(ii), it is easy to see that MN(τ) is a modular form with respect toSL(2,Z)of weight 3|QN|(i.e. for anyγSL(2,Z), when (r,s) runs over all elements ofQN, then so does(r,s)after moduloZ2), where

|QN|=#QN. To apply the theory of modular forms, we recall the following classical formula. See e.g. [8, 15] for the proof.

(10)

Theorem C. Let f(τ) be a nonzero modular form with respect to SL(2,Z) of weight k. Then

(2.9)

τH\{i,ρ}

ντ(f) +ν(f) + 1

2νi(f) +1

3νρ(f) = k 12,

whereρ := eπi/3τ(f)denotes the zero order of f atτand the summation over τis performed modulo SL(2,Z)equivalence.

We note for each(r,s)∈ QN, there exists a unique(r, ˜˜ s)∈ QN such that (r, ˜˜ s) ≡ (−r,s) mod Z2. Then property (i) of Zr,s(2)(τ) gives Zr,˜(˜2s)(τ) =

Zr,s(2)(τ), which implies that

(2.10) ντ(MN)∈ 2N∪ {0} for any τH.

To apply Theorem C, we need the asymptotics ofZ(r,s2)(τ)as Imτ→+. Lemma 2.4. Let(r,s) ∈ [0, 1)×[0, 1)\12Z2 and q = e2πiτ withτF0. Then the asymptotics of Z(r,s2)(τ)at the three cuspsτ=0, 1,∞are as follows:

(a) AsF0τ,

Zr,s(2)(τ) =4π3is(1−s)(2s−1) +o(1) if s0,1212, 1, Z(r,s2)(τ) =−48π3sin(2πr)q+O(q2) if s=0,

Zr,s(2)(τ) =−12π3sin(2πr)q1/2+O(q) if s=1/2.

(b) AsF0τ0,

τlim0Zr,s(2)(τ) = if r0,12

12, 1 . (c) AsF0τ1,

τlim1Zr,s(2)(τ) = if (r+s)∈ 0,1212, 11,32.

Proof. By using theq-expansions of℘(z|τ)andZr,s(τ)(see (3.12)-(3.13) in Section 3), the asymptotics of Zr,s(2)(τ)as τ can be easily calculated.

Because the calculation is straightforward and is already done in [7, 13], we omit the details for (a) here.

The asymptotics ofZr,s(2)(τ)at the cusp 0 can be obtained by using prop- erty (ii) and the assertion (a). Lettingγ=

1 −1

1 0

leads to Z(r2+)s,r(ττ1) =τ3Z(r,s2)(τ).

(11)

WhenτF0andτ0, we have ττ1F0 and ττ1. Then applying (a) we obtain that asF0τ0,

Zr,s(2)(τ) = −1

τ3Z(2()r+s),r(ττ1)

= −1 τ3

3ir(1−r)(2r−1) +o(1) if r ∈(0,12)∪(12, 1). (2.11)

This proves (b).

Similarly, when τF0 and τ1, we have ττ1F0 and ττ10.

Applying property (ii) and (2.11) we obtain that asF0τ1, Zr,s(2)(τ) = 1

τ3Zr(+2)s,r(ττ1)

= −1 (τ1)3

3i(r+s)(1−rs)(2r+2s−1) +o(1) forr+s ∈ (0,12)∪(12, 1). The remaining caser+s ∈ (1,32)follows from Zr,s(2)(τ) =Zr(2)1,s(τ). This proves (c).

Lemma 2.4-(a) implies

(2.12) the vanishing order ofZr,s(2)(τ)at∞is

0 if s 6=0, 1/2 1 if s =0

1

2 if s=1/2 .

Recall△k,k=0, 1, 2, 3, defined in (1.12).

Lemma 2.5. Fix k ∈ {0, 1, 2, 3}. Then the number of zeros of Z(r,s2)(τ)in F0is a constant for(r,s)∈ △k.

Proof. Since(r,s)∈ △k, we haver,s,r+s 6∈ {0,12, 1,32}, so Lemma 2.4 (a)- (c) imply that

Zr,s(2)(τ)6→0 as F0τ, 0, 1

respectively. Together with Lemma 2.3 that Z(r,s2)(τ) 6= 0 on ∂F0H, it is easy to apply the argument principle to conclude that the number of zeros ofZ(r,s2)(τ)inF0is a constant for(r,s)∈ △k. Lemma 2.6. Let(r,s)∈Q3. Then Zr,s(2)(τ)6=0for anyτH.

Proof. Note that 3|Q3|=24. Sinces 6= 12and(13, 0),(23, 0)∈Q3, we see from Lemma 2.4-(a) (or (2.12)) thatM3(τ)∼ q2asF0τ, i.e. ν(M3) = 2.

Therefore, we deduce from (2.9) thatM3(τ)has no zeros inH.

(12)

-2.0 -1.5 -1 -0.5 0 0.5 1 1.5 2 0.5

1 1

.5

FIGURE2. F0= Fγ1(F)∪γ2(F). LetFbe a fundamental domain ofSL(2,Z)defined by2

(2.13) F :={τH|0Reτ1,|τ| ≥1,|τ1|>1} ∪ {eπi/3}. Define γ(F) := {γ·τ|τF} for any γSL(2,Z), then γ(F) is also a fundamental domain ofSL(2,Z). Let

(2.14) γ1:=

0 1

1 1

, γ2:=

1 −1

1 0

, then it is easy to prove that

(2.15) F0 =Fγ1(F)∪γ2(F).

See Figure 2, which is copied from [6]. Now we are in the position to prove Theorem 2.1.

Proof of Theorem 2.1. Recall Lemma 2.3 that Zr,s(2)(τ) 6= 0 for τ ∈ {ρ} ∪ (∂F0H). We divide the proof of Theorem 2.1 into several steps.

Step 1.We claim thatZr,s(2)(τ)has no zeros inF0for(r,s)∈ △0. Lemma 2.6 says thatZ(12)

3,13(τ)has no zeros in F0. Since(13,13) ∈ △0, our claim follows directly from Lemma 2.5.

Step 2. We claim thatZr,s(2)(τ)has a unique zero in F0 for(r,s) ∈ △1

2∪ △3. Since

(2.16) (56,13)∈ △1, (23,16)∈ △2, (16,16)∈ △3,

by Lemma 2.5 we only need to prove the claim for(r,s)∈ {(16,16),(23,16), (56,13)}. Note that M6(τ) is a modular form of weight 3|Q6| = 72. For

2Of course, the standard definition ofFshould beF :={τ H|0Reτ <1,|τ| ≥ 1,|τ1|>1} ∪ {eπi/3}, i.e. Reτ=1 is not needed. In this paper, to guarantee the validity of (2.15), it is more convenient for us to use the definition (2.13), which does not effect our following argument because Lemma 2.3 says that Zr,s(2)(τ) 6= 0 ifτ ∈ {τ H|Reτ = 1} ∪γ1({τH|Reτ=1})γ2({τH|Reτ=1})∂F0H.

Références

Documents relatifs

This leads us to observe that if one can resolve any singularities of the 1-dimensional foliation F into weakly reduced ones, and generalize the previous Brunella Theorem to

We investigate the discrete spectrum behaviour for the 2d Pauli operator with nonconstant magnetic field, perturbed by a sign-indefinite self-adjoint electric potential which

A proof of Mordell's conjecture [10] that a curve of genus ^2 has only a finite number of rational points would of course supersede the Siegel-Mahler theorem for such curves, but

Denote by H^ ^ „ the open subscheme of the Hilbert Scheme parametrizing the smooth irreducible curves of degree d and genus g in P&#34;.. The purpose of this paper is to prove that H^

Wüstholz: A New Approach to Baker’s Theorem on Linear Forms in Logarithms III, (New advances in transcendence theory, 1986 symposium, Durham), Cambridge University

It is plausible, however, that one might bound the number of algebraic families occurring in Corollary 4.4, and therefore get an effective bound on the number

These varieties are the loci of moduli of special curves distinguished by the presence of a Weier- strass point of a particular type.. These subloci of Mg (cf. 3) can

In a previous paper ([2]) we have determined the singularities of the generic polar curves of a sufficiently general irreducible algebroid curve 7 with prescribed