vector-valued automorphic forms and vector bundles › ... › 3 ›...

116
Vector-valued automorphic forms and vector bundles by Hicham Saber Thesis submitted to the Faculty of Graduate and Postdoctoral Studies In partial fulfillment of the requirements For the Ph.D. degree in Mathematics Department of Mathematics and Statistics Faculty of Science University of Ottawa c Hicham Saber, Ottawa, Canada, 2015

Upload: others

Post on 25-Jun-2020

5 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Vector-valued automorphic forms andvector bundles

by

Hicham Saber

Thesis submitted to the

Faculty of Graduate and Postdoctoral Studies

In partial fulfillment of the requirements

For the Ph.D. degree in

Mathematics

Department of Mathematics and Statistics

Faculty of Science

University of Ottawa

c© Hicham Saber, Ottawa, Canada, 2015

Page 2: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Abstract

In this thesis we prove the existence of vector-valued automorphic forms for

an arbitrary Fuchsian group and an arbitrary finite dimensional complex

representation of this group. For small enough values of the weight as well as

for large enough values, we provide explicit formulas for the spaces of these

vector-valued automorphic forms (holomorphic and cuspidal).

To achieve these results, we realize vector-valued automorphic forms as

global sections of a certain family of holomorphic vector bundles on a certain

Riemann surface associated to the Fuchsian group. The dimension formulas

are then provided by the Riemann-Roch theorem.

In the cases of 1 and 2-dimensional representations, we give some ap-

plications to the theories of generalized automorphic forms and equivariant

functions.

ii

Page 3: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Acknowledgements

I would like to thank everyone who has helped me in any way, particularly my

supervisor Dr. Abdellah Sebbar. I would also like to thank the Department

of Mathematics and Statistics for its financial support.

iii

Page 4: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Contents

Introduction 1

1 Preliminaries 8

1.1 Vector Bundles . . . . . . . . . . . . . . . . . . . . . . . . . . 8

1.1.1 Basic properties . . . . . . . . . . . . . . . . . . . . . . 9

1.1.2 Divisors . . . . . . . . . . . . . . . . . . . . . . . . . . 12

1.1.3 The Riemann-Roch theorem. . . . . . . . . . . . . . . . 17

1.2 Fuchsian groups and their Riemann surfaces. . . . . . . . . . . 19

1.2.1 Classification of Mobius transformations . . . . . . . . 20

1.2.2 Fuchsian groups . . . . . . . . . . . . . . . . . . . . . . 22

1.2.3 The Riemann surface of a Fuchsian group. . . . . . . . 25

1.2.4 Congruence subgroups of PSL(2,Z) . . . . . . . . . . . 27

1.3 Automorphic forms. . . . . . . . . . . . . . . . . . . . . . . . . 29

1.3.1 Definition of automorphic forms. . . . . . . . . . . . . 29

1.3.2 Examples of modular forms . . . . . . . . . . . . . . . 32

2 Vector-valued automorphic forms and vector bundles 34

2.1 Notations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

2.2 The family EΓ,R,k of vector bundles . . . . . . . . . . . . . . . 37

iv

Page 5: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

2.2.1 The unbranched case. . . . . . . . . . . . . . . . . . . 37

2.2.2 The elliptic case. . . . . . . . . . . . . . . . . . . . . . 40

2.2.3 The cuspidal case. . . . . . . . . . . . . . . . . . . . . 43

2.2.4 Construction of the 1−cocycle. . . . . . . . . . . . . . 46

2.3 Behavior at the cusps . . . . . . . . . . . . . . . . . . . . . . . 50

2.4 The divisor of a vector-valued automorphic form . . . . . . . 51

2.5 Mk(Γ, R) and Sk(Γ, R) as global sections of vector bundles . . 57

3 The dimension Formula 69

3.1 The degree of EΓ,R,k . . . . . . . . . . . . . . . . . . . . . . . 69

3.2 The holomorphic degree of a vector bundle. . . . . . . . . . . 75

3.3 The dimension of Mk(Γ, R) and Sk(Γ, R) . . . . . . . . . . . 81

3.4 Finite image representations . . . . . . . . . . . . . . . . . . . 87

4 Applications 90

4.1 Generalized automorphic forms . . . . . . . . . . . . . . . . 90

4.2 ρ−equivariant functions . . . . . . . . . . . . . . . . . . . . . 95

4.2.1 Differential equations . . . . . . . . . . . . . . . . . . . 96

4.2.2 The correspondence . . . . . . . . . . . . . . . . . . . . 99

4.2.3 Examples . . . . . . . . . . . . . . . . . . . . . . . . . 103

v

Page 6: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Introduction

This thesis deals with the topic of vector-valued automorphic forms. These

are considered as a natural extension of classical automorphic forms and are

defined as follows. Let SL(2,R) be the group of real 2 × 2 matrices with

determinant one, and PSL(2,R) = SL(2,R)/±I. The group SL(2,R) acts

on the Poincare half-plane

H = z ∈ C : Im z > 0

by Mobius transformations, i.e.,

γz =az + b

cz + d, z ∈ H , γ =

(a b

c d

)∈ SL(2,R).

If Γ is a discrete subgroup of SL(2,R), and R is an n-dimensional complex

representation of Γ, that is a homomorphism

R : Γ −→ GL(n,C) ,

then an unrestricted vector-valued automorphic form for Γ of multiplier R

and integral weight k is a meromorphic map F : H −→ Cn satisfying

F (γ · z) = Jkγ (z)R(γ) F (z) , z ∈ H , γ ∈ Γ ,

where Jγ(z) = cz+d if γ =

(∗ ∗c d

). When F is meromorphic (resp. holomor-

phic) at the cusps of Γ, then we say that F is a vector-valued automorphic

form (resp. holomorphic vector-valued automorphic form) for Γ of multiplier

1

Page 7: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

R and weight k. This definition can be extended to real weights k using the

notion of multiplier system, see [23]. In this thesis we content ourselves to

the case of integral weights.

The theory of vector-valued automorphic forms has been around for a long

time, first, as a generalization of the classical theory of scalar automorphic

forms, then as natural objects appearing in mathematics and physics. For

instance, Selberg suggested vector-valued forms as a tool to study modular

forms for finite index subgroups of the modular group [49] and they also

appear as Jacobi forms in the work of M. Eichler and D. Zagier [10], and in

[57]. In the last decade, there has been a growing interest in the study of

vector-valued forms, and several important results have been obtained by M.

Knopp, G. Mason and other authors [4, 17, 30, 31, 35, 36, 37].

In order to prove the existence of vector-valued automorphic forms or to

compute the dimensions of their vector spaces, people used a wide range of

tools coming from geometry, spectral theory, differential equations and clas-

sical complex analysis. For instance, J. Fischer [14], D. A. Hejhal [22, 23],

and W. Roelcke [47, 48] used the Selberg trace formula to compute the di-

mensions of the spaces of holomorphic vector-valued automorphic forms, but

they imposed the tight condition of the unitarity of the representation R, or

more generally the multiplier system. This condition was also imposed by

R. Borcherds [6, 7] for the metaplectic group of a general Fuchsian group of

the first kind. In the special case of the modular group SL(2,Z), G. Mason

[37] employed linear differential equations to construct vector-valued auto-

morphic forms of negative weight. In the paper [17], T. Gannon used the

Riemann-Hilbert problem to prove the existence of vector-valued automor-

phic forms for SL(2,Z), with a possible generalization to the case of Γ being a

Fuchsian group of genus zero as announced there. Classical methods such as

Poincare series have been used by M. Knopp, and G. Mason [30] to solve the

existence problem, but always restricted to the special case of Γ = SL(2,Z).

A different approach based on the Riemann-Roch theorem has been used by

C. Meneses [39], but only when the weight is two and the representation is

2

Page 8: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

unitary. E. Freitag [15] also used the Riemann-Roch theorem in the case

when Γ has a finite index subgroup Γ0 such that the image of the elements

of Γ0 by the multiplier system are simultaneously diagonalizable. In other

words, up to a change of basis, the components of a vector-valued automor-

phic form for Γ are simply classical automorphic forms for Γ0. A sketch of a

proof has been given in the case of unitary multiplier system.

These different restrictions on the representation R or the group Γ have

the disadvantage of excluding many important cases which are highly relevant

to the theory of automorphic forms. For example, weight two vector-valued

automorphic forms for the symmetric power representations [34], which are

not unitary, play an essential role in the theory of periods of modular forms,

see Shimura’s work in [56]. Also, in [42] it is shown that the theory of quasi-

modular forms [38] can be covered by vector-valued modular forms with

respect to the symmetric tensor representations. In the case of 2-dimensional

representations, one misses the theory of equivariant functions developed

by A. Sebbar and others, see [11, 12, 51, 54]. Even in the 1-dimensional

case, these restrictions are strict enough to exclude the emerging theory of

generalized modular forms, see [28, 25, 26, 29, 33, 44, 45].

What we propose in this thesis is to prove the existence of vector-valued

automorphic forms in full generality, i.e., without any restriction on the group

Γ or the representation. Also, to show that given a discrete subgroup Γ of

SL(2,R), an n-dimensional complex representation R of Γ, and an integer

k ∈ Z, the dimensions dΓ,R,k, sΓ,R,k of the spaces Mk(Γ, R), Sk(Γ, R) of

holomorphic, cusp vector-valued automorphic forms for Γ of multiplier R

and weight k are infinite when Γ is Fuchsian group of the second kind, and

they are finite when Γ is Fuchsian group of the First kind. In this last case,

we provide two rational numbers k+Γ,R, and k−Γ,R depending on Γ and R such

that:

(A) For k < k−Γ , we have dΓ,R,k = 0, and sΓ,R,k = 0.

(B) For k > k+Γ , we have explicit formulas for dΓ,R,k and sΓ,R,k expressed in

3

Page 9: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

terms of k, n, and some invariants of Γ and R.

Our method consists of associating to the triplet (Γ, R, k) a holomorphic

vector bundle EΓ,R,k on a certain, well known, Riemann surface XΓ. Then

we prove that the global sections of EΓ,R,k, when lifted to H, correspond to

vector-valued automorphic forms for Γ of multiplier R and weight k. In case

XΓ is noncompact, we show that dΓ,R,k and sΓ,R,k are both infinite. When XΓ

is compact, then we use the Riemann-Roch theorem to give explicit formulas

for dΓ,R,k and sΓ,R,k as explained in (A) and (B). The existence of vector-

valued automorphic forms will be proved by showing that dΓ,R,k and sΓ,R,k,

when they are finite, are asymptotically equivalent to k, that is

dΓ,R,k

k−→ nmΓ

2, as k →∞,

andsΓ,R,k

k−→ nmΓ

2, as k →∞,

where mΓ is some positive constant depending on Γ, see Remark 3.1.5.

It should be pointed that in a joint work with A. Sebbar [50], we proved

the existence of vector-valued automorphic forms in the case of subgroups

of PSL(2,R) rather than subgroups of SL(2,R). This was achieved as fol-

lows: we used the solution to the Riemann-Hilbert problem for non-compact

Riemann surfaces [13] to attach to each pair (Γ, R) of a discrete subgroup Γ

of PSL(2,R) and an n-dimensional complex representation R of Γ, a holo-

morphic vector bundle EΓ,R on XΓ in such a way that the global sections

of EΓ,R correspond to vector-valued automorphic forms for Γ of multiplier R

and weight 0. Then by the Kodaira vanishing theorem for compact Riemann

surfaces [19], in case XΓ is compact, or by the fact that XΓ is a Stein variety

[21] in case it is not compact, we proved the existence of n linearly indepen-

dent vector-valued automorphic forms for Γ of multiplier R and a certain

integral weight k.

The 2-dimensional case is of special interest since it has a connection

to ρ-equivariant functions. More precisely, given a discrete subgroup Γ of

4

Page 10: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

SL(2,R) and a representation ρ : Γ −→ GL(2,C), a meromorphic function h

on H is called a ρ−equivariant function with respect to Γ if

h(γz) = ρ(γ)h(z) for all γ ∈ Γ , z ∈ H , (0.0.1)

where the action on both sides is by linear fractional transformations. When

ρ is the defining representation of Γ, that is ρ(γ) = γ for all γ in Γ, h is

simply an equivariant function, see (4.2.1).

ρ-equivariant functions and 2−dimensional vector-valued automorphic

forms are connected in the following way: if F = (f1, f2)t is an unrestricted

vector-valued automorphic form for Γ of multiplier ρ and weight k, then

one can check that if f2 is nonzero, then the function hF = f1/f2 is a ρ-

equivariant function for Γ. The main result of our joint work [52] with A.

Sebbar is that the converse is also true: for any ρ−equivariant function there

exists an unrestricted vector-valued automorphic form F = (f1, f2)t for Γ of

multiplier R and some weight k ∈ Z such that:

h = f1/f2.

In this thesis we enhance this result by proving that for any 2-dimensional

representation ρ of Γ, there exists a ρ−equivariant function of the form

hF = f1/f2,

where F = (f1, f2)t is a holomorphic vector-valued automorphic form for Γ

of multiplier R and a certain nonnegative weight k.

The 1-dimensional case corresponds to the theory of generalized auto-

morphic forms. Due to their similarity with classical automorphic forms and

their richness, the generalized automorphic forms are receiving an increasing

interest, specially after the work of M. Knopp and G. Mason in [28], see

[25, 26, 29, 33, 44, 45]. A meromorphic function f on H is called a general-

ized automorphic form of weight k for a discrete subgroup Γ of SL(2,Z) if

for all γ ∈ Γ and z ∈ H, we have

f(γ · z) = µ(γ) Jkγ (z) f(z)

5

Page 11: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

where µ : Γ −→ C is a character. In addition, f should be meromorphic

at the cusps. The fundamental difference between generalized automorphic

forms and classical automorphic forms is that the character µ need not be

unitary. The existence theorem of vector-valued automorphic forms in di-

mension 1 guarantees the existence of generalized automorphic forms for an

arbitrary character of an arbitrary discrete group Γ. Our main contribution

to this theory will be the computation of the dimensions of the spaces of

holomorphic and cusp generalized automorphic forms in the sense of (A) and

(B).

This thesis is organized as follows: in chapter one we shall review some

classical results from the theories of holomorphic vector bundles on Riemann

surfaces, Fuchsian groups, and automorphic forms.

In chapter two we realize vector-valued automorphic forms as global sec-

tions of holomorphic vector bundles. More precisely, given a Fuchsian sub-

group Γ of SL(2,R), a finite-dimensional representation R of Γ, and an in-

teger k such that the pair (R, k) is simple, see Definition 2.1.2, we construct

a holomorphic vector bundle EΓ,R,k over the Riemann surface XΓ = Γ\H∗

associated to Γ. The spaces Mk(Γ, R), Sk(Γ, R) of holomorphic, cusp vector-

valued automorphic forms for Γ of multiplier R and weight k will be respec-

tively isomorphic to H0(XΓ,O(EΓ,R,−k)) and H0(XΓ,O(−DS,R,−k+EΓ,R,−k)),

where DS,R,−k is a certain holomorphic line bundle depending on the cusp-

idal points of XΓ. As a result, we deduce that the dimensions dΓ,R,k, sΓ,R,k

of Mk(Γ, R), Sk(Γ, R) are finite if and only if Γ is Fuchsian group of the first

kind, see Theorem 2.5.9.

In chapter three we introduce the notion of the holomorphic degree of a

holomorphic vector bundle, which consists of associating to each holomorphic

vector bundle E , defined on a compact Riemann surface, an integer d(E) in

a such a way that:

d(E) < 0 =⇒ H0(X,O(E)) is trivial.

We will give some basic properties of the holomorphic degree and some useful

6

Page 12: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

bounds of d(E). Applying the Riemann-Roch theorem to the holomorphic

vector bundles EΓ,R,−k and −DS,R,−k+EΓ,R,−k, we prove the results mentioned

in (A) and (B).

The last chapter provides applications of our results in the special cases

of 1 and 2-dimensional representations. In particular, we compute the di-

mensions of the spaces of holomorphic and cusp generalized automorphic

forms. Also, we show the existence of ρ−equivariant functions, and prove

their parametrization by 2-dimensional unrestricted vector-valued automor-

phic forms of multiplier ρ. This parametrization relies on the existence of

global solutions to a certain second degree differential equation, as well as on

the fact that the Schwarz derivative of a ρ−equivariant function is a weight 4

unrestricted automorphic form for Γ. We end the chapter by constructing ex-

amples of ρ−equivariant functions when ρ is the monodromy representation

of second degree ordinary differential equations.

7

Page 13: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Chapter 1

Preliminaries

The aim of this chapter is to give a review of some basic facts from the theories

of holomorphic vector bundles on Riemann surfaces, Fuchsian groups, and

automorphic forms.

1.1 Vector Bundles

The content of this section is entirely based on the references:

[2]: Chapter 1: §1.5, §1.8.

[13]: Chapter 1: §29, §30, and Chapter 2: §16.

[19] Chapter 0: §0.5, Chapter 1: §1.1, and Chapter 1: §2.1 .

[20] Chapter C.

[?] Chapter 6: §6.5.

X will be a Riemann surface, that is, a one-dimensional complex manifold,

and n a positive integer.

8

Page 14: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

1.1.1 Basic properties

Definition 1.1.1. Let X be a Riemann surface, E be a complex manifold,

and p : E −→ X be a holomorphic map. Suppose that each fiber Ex :=

p−1(x) has the structure of an n−dimensional vector space over C. Then

p : E −→ X, or simply E , is called a holomorphic vector bundle of rank n on

X if every point x ∈ X has an open neighborhood U such that :

1. There exists a biholomorphic map

φU : EU := p−1(U) −→ U × Cn

taking Ex to x × Cn for every x ∈ U .

2. For every x ∈ U , the map φU is a linear isomorphism from Ex to

x × Cn ∼= Cn.

The map φU : EU −→ U ×Cn is called a trivialization of E over U . A vector

bundle of rank 1 is called a line bundle.

Definition 1.1.2. Let E and F be two holomorphic vector bundles on X

of rank n and m respectively. A holomorphic map Ψ : E −→ F is called a

homomorphism (resp. isomorphism) if for all x ∈ X

1. Ψ(Ex) ⊆ Fx

2. Ψx := Ψ|Ex : Ex −→ Fx is a linear homomorphism (resp. isomorphism).

E and F are called isomorphic if there exists an isomorphism between them.

In particular, E is called a trivial holomorphic vector bundles on X if it is

isomorphic to the trivial holomorphic vector bundel X × Cn.

Note that for any pair of trivializations φU and φV we have a holomorphic

map

gUV : U ∩ V −→ GL(n,C),

9

Page 15: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

such that

φUV := φU φ−1V : (U ∩ V )× Cn −→ (U ∩ V )× Cn

is given by

φUV (x, t) = (x, gUV (x)t) for every x ∈ (U ∩ V )× Cn.

The map gUV is called a transition function for E relative to the trivializa-

tions φU and φV . Also, one has the relation

gUV gVW = gUW on U ∩ V ∩W. (1.1.1)

Definition 1.1.3. Let U be an open subset of X. Let O(U) (resp. M(U))

denotes the group of holomorphic (resp. meromorphic) functions on U , and

GL(n,O(U)) (resp. GL(n,M(U)) ) the group of all n × n invertible ma-

trices with coefficients in O(U) (resp. M(U) ). Together with the natural

restrictions when V ⊆ U , they define sheaves of groups O, M, GL(n,O),

and GL(n,M) on X ( GL(n,O) and GL(n,M) are not abelian for n ≥ 2).

If U = (Ui)I is an open cover of X, we define Z1(U ,GL(n,O)) to be

the set of all 1−cocycles with values in GL(n,O) with respect to U , i.e., all

families (gUV )U with

gUV ∈ GL(n,O(U ∩ V ))

and verifying (1.1.1).

Remark 1.1.1. Note that for n ≥ 2 the set Z1(U ,GL(n,O)) is not a group

with respect to component-wise multiplication.

As seen above, any holomorphic vector bundle gives rise to one a 1−cocycle

with respect to an open cover U of X by means of the transition functions.

Conversely, we have:

Theorem 1.1.2. Let U = (Ui)I be an open cover of X, and (gUV )U be an

element of Z1(U ,GL(n,O)). Then up to isomorphism, there exits a unique

holomorphic vector bundle E on X of rank n having transition functions

(gUV )U .

10

Page 16: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Also, we have

Theorem 1.1.3. Let U = (Ui)I be an open cover of X, and (gUV )U , (hUV )U

be elements of Z1(U ,GL(n,O)). Suppose that there exist elements

fU ∈ GL(n,O(U)), U ∈ U , such that

hUV = fU gUV f−1V on U ∩ V ,

then (gUV )U and (hUV )U represent the same holomorphic vector bundle.

In principale, all operations on vector spaces induce operations on vector

bundles. For example, if E and F are holomorphic vector bundles over X of

rank n and m having transition functions (gUV )U and (hUV )U one can define:

1. The dual bundle E∗, given by the transition functions

g∗UV (x) = (g−1UV )T (x).

2. The direct sum E ⊕ F , given by the transition functions

gUV (x)⊕ hUV (x) ∈ GL(Cn ⊕ Cm).

3. The tensor product E ⊗ F , given by the transition functions

gUV (x)⊗ hUV (x) ∈ GL(Cn ⊗ Cm).

4. The exterior product∧k E , given by the transition functions

k∧gUV (x) ∈ GL(∧kCn).

In particular,∧n E is a line bundle given by the transition functions

det(gUV )(x) ∈ GL(1,C) = C∗,

also called the determinant bundle of E .

11

Page 17: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

We now come to the notion of a section over a holomorphic vector bundle.

Definition 1.1.4. Let p : E −→ X be a holomorphic vector bundle of rank

n, and U be an open subset of X. A meromorphic section σ of E over U is a

meromorphic map σ : U −→ E such that p σ = idU , i.e., σ(x) ∈ Ex for all

x ∈ U\Pσ, where Pσ is the set of poles of σ in U . If U = X, then σ is called

a global meromorphic section of E .

If σ has no poles, it is called a holomorphic section.

If φU is a trivialization of E over U , then we can associate to each holo-

morphic section σ of E over U a unique meromorphic map fU : U −→ Cn

such that

φU(σ(x)) = (x, fU(x)), for all x ∈ U\Pσ.

If φV is a trivialization of E over V , then we have

fU = gUV fV on (U ∩ V )\Pσ,

where gUV is the transition function of E relative to φU and φV . Thus we

have

Proposition 1.1.4. In terms of trivializations EU : U −→ U × Cn, a mero-

morphic section σ of E over⋃U∈U U , corresponds to a collection (fU)U of

meromorphic maps fU : U −→ Cn such that

fU = gUV fV on (U ∩ V )\Pσ,

where the gUV , U, V ∈ U , are transition functions of E relative to (φU)U .

1.1.2 Divisors

In this subsection, we define the notion of a divisor of a section of a holo-

morphic vector bundle and give its basic properties.

12

Page 18: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Let U be an open subset of C, and F : U −→ Cn be a meromorphic map.

Then for any w in H, there exists an open disc Dw centered at w, and a

unique integer nW such that

F (z) = (z − w)nwGw(z), z ∈ Dw,

where Gw is a holomorphic map on Dw satisfying Gw(w) 6= 0. The integer

nw is called the order of F at w and is denoted by ordw(F ).

Remark 1.1.5. If Ψ: Dw −→ GL(n,C) is a holomorphic map, Then

ordw(F ) = ordw(ΨF ).

Suppose that U is an open cover of X. Let E be a holomorphic vector

bundle over X having transition functions (gUV )U , σ be a meromorphic sec-

tion of E over X, and (fU)U be its corresponding collection of meromorphic

maps. Let P ∈ U and take an open neighborhood UP of P in U such that

under a coordinate chart ϕ for X, we have ϕ(UP ) = D and ϕ(P ) = 0, where

D is the unit disc. The map FU,P = fU ϕ−1 is then meromorphic on D, and

we define the order νP (σ) of σ at P by

νP (σ) = ord0(FU,P ). (1.1.2)

According to [58], the integer νP (σ) is invariant under coordinate changes.

Hence, it is enough to show that νP (σ) is independent of U . Indeed, if P lies

in U ∩ V , then we have

fU = gUV fV on (U ∩ V )\Pσ.

Since gUV ∈ GL(n,O(U ∩ V )), by the above remark, we have

ordP (fU) = ordP (fV ).

The divisor of σ is defined by the formal sum

div(σ) =∑P∈X

νP (σ)P. (1.1.3)

13

Page 19: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Similarly, the divisor of a meromorphic function f on X is given by

div(f) =∑P∈X

νP (f)P, (1.1.4)

where νP (f) is the order of f at P . More generally

Definition 1.1.5. A divisor D on X is a locally finite formal sum

D =∑P∈X

dP P, dP ∈ Z,

where locally finite means that for any point P in X, there exists a neigh-

borhood U of P such that dP 6= 0 for only finitely many P in U . If U is an

open subset of X, we set

D|U =∑P∈U

dP P.

Div(X) will be the abelian group of all divisors on X. For D,D′ ∈Div(X),

set D ≥ D′ if dP ≥ d′P for all P ∈ X. A divisor D is called effective if

D ≥ 0.

The support Supp(D) of D will be the closure in X of the set

P ∈ X | dP 6= 0.

Remark 1.1.6. When X is compact, a locally finite sum is simply a finite

sum. Hence Supp(D) is a finite set.

Now, we focus on the relationship between divisors and line bundles. Let

D be a divisor on X. By the local finiteness assumption on D, one can find

an open cover U of X and functions fU ∈M(U), U ∈ U , such that

div(fU) = D on U.

Since fU and fV have the same divisor D|U∩V on U ∩ V , we see that

gUV := fU/fV ∈ O∗(U ∩ V ) = GL(1,O(U ∩ V )).

14

Page 20: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

It is clear that (gUV )U is an element of Z1(U ,GL(n,O)). Hence it defines a

line bundleD onX, and by construction, σD = (fU)U is a global meromorphic

section of D such that

div(σD) = D.

If the set of line bundles constructed in the above way is denoted by

AD, then it can be shown that any two line bundles in AD are naturally

isomorphic. We use this fact to identify all elements of AD to one line bundle

that will be called the associated line bundle to the divisor D, and will be

denoted by |D|.

Suppose that L is a holomorphic line bundle given by transition functions

(gUV )U . If L has a nonzero global meromorphic section σ = (fU)U , then by

Proposition 1.1.4 we have

gUV = fU/fV ,

that is L = |div(σ)|. From this we deduce that a line bundle is associated

to a divisor (resp. an effective divisor) if and only if it has a nonzero global

meromorphic (resp. holomorphic) section.

Theorem 1.1.7. Any holomorphic line bundle L on X has a nontrivial

global meromorphic section. In other words it is of the form L = |D| for

some divisor D on X.

Recall that when X is compact, the support of any divisor on X is a finite

set.

Definition 1.1.6. The degree of a divisor D =∑

P∈X dP P, on a compact

Riemann surface X is defined by

deg(D) =∑P∈X

dP .

We have the following

Proposition 1.1.8. Suppose that X is compact. If f is a meromorphic

function on X, then

deg(div(f)) = 0.

15

Page 21: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

The degree of a holomorphic line bundle over a compact Riemann surface

X is defined as follows. Let U be an open cover of X, L a holomorphic line

bundle over X having transition functions (gUV )U , σ a meromorphic section

of E over X, and (fU)U its corresponding collection of meromorphic maps.

Such σ always exists according to Theorem 1.1.7. If σ1 = (hU)U is an other

meromorphic section of L, then from the relations

fU = gUV fV , hU = gUV hV on U ∩ V

we see that

hU/fU = hV /fV , on U ∩ V.

Hence f := σ1/σ is a global meromorphic function on X, and we have

div(σ1) = div(σ) + div(f).

Therefore

deg(div(σ1)) = deg(div(σ)) + deg(div(f)) = deg(div(σ)),

since deg(div(f)) = 0 according to Theorem 1.1.8. Thus, the divisors of

global meromorphic sections have the same degree.

Definition 1.1.7. Let X be a compact Riemann surface, and let E be a

holomorphic vector bundle over X, and L be a holomorphic line bundle over

X.

1. The degree of L is defined by

deg(L) = deg(div(σ)),

where σ is any global meromorphic section of L.

2. The degree of E is defined by

deg(E) = deg(det(E)),

where det(E) is the determinant bundle of E .

16

Page 22: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

The degree has the following useful property.

Proposition 1.1.9. Let X be a compact Riemann surface, and let E and Fbe a holomorphic vector bundles over X, and L be a holomorphic line bundle

over X. We have

deg(E ⊕ F) = deg(E) + deg(F),

and

deg(L ⊗ E) = n deg(L) + deg(E),

where n is the rank of E.

1.1.3 The Riemann-Roch theorem.

The goal of this subsection is to state the Riemann-Roch theorem for holo-

morphic vector bundles over compact Riemann surfaces.

Let E be a holomorphic vector bundle over X of rank n, and D a divisor

on X, σ a meromorphic section of E over X, and (fU)U its corresponding

collection of meromorphic maps. For any open subset U of X set O(E)(U)

(resp. M(E)(U) ) to be the O(U)−module of holomorphic (resp. mero-

morphic) sections of E over U . Together with the natural restrictions when

V ⊆ U , they define sheaves O(E) and M(E) of O−modules on X. The

vector space H0(X,O(E)) (resp. H0(X,M(E))) of global sections of O(E))

(resp. M(E))) is exactly the vector space of global holomorphic (resp. mero-

morphic) sections on X.

Remark 1.1.10. The dimension of the C−vector space H0(X,O(E)) will be

denoted by h0(E) when it is finite.

Example 1.1.11. If (U, zU), U ∈ U , is a covering of X by coordinate neigh-

borhoods, then on U ∩ V the function

gUV = dzU/dzV

17

Page 23: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

lies in GL(1,O(U ∩ V )) = O∗(U ∩ V ), and it is clear that

(gUV )U ∈ Z1(U ,GL(1,O)).

The attached holomorphic line bundle is called the canonical line bundle of

X, and will be denoted by KX .

The sheaf O(KX) is isomorphic to the sheaf ΩX of holomorphic 1−forms

on X, the latter is defined by taking Ω(U) to be the O(U)−module of holo-

morphic 1−forms on U together with the natural restrictions. Thus the

sections of KX over U can be identified with holomorphic 1−forms on U . We

will adopt this identification for the rest of this thesis.

Similarly, if for an open subset U of X we set

OD(E)(U) = σ ∈M(E)(U) | div(σ) ≥ −D|U,

then OD(E)(U) is an O(U)−module, and together with the natural restric-

tions, we get a sheaf OD(E) of O−modules on X. By definition, we have

H0(X,OD(E)) = σ ∈ H0(X,M(E)) | div(σ) ≥ −D.

Moreover, if |D| is the associated line bundle to the divisor D, and σ is a

section over |D| such that div(σ) = D, then the multiplication by σ gives an

isomorphism of sheaves

OD(E) −→ O(|D| ⊗ E) : f 7→ σf.

This induces the isomorphism

H0(X,OD(E)) −→ H0(X,O(|D| ⊗ E)) : f 7→ σf, (1.1.5)

which we use to identify H0(X,OD(E)) with H0(X,O(|D| ⊗ E)). We have

the following fundamental theorem:

Theorem 1.1.12. Suppose that X is compact, and let E be a holomorphic

vector bundle over X of rank n. Then H0(X,O(E)) is a finite dimensional

C−vector space.

18

Page 24: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Definition 1.1.8. The genus g of a compact Riemann surface X is defined

to be the dimension of the space of holomorphic 1−forms on X, that is

g = h0(KX).

Finally, the Riemann-Roch theorem for holomorphic vector bundles on a

compact Riemann surfaces reads as follows:

Theorem 1.1.13. Suppose that X is compact and having genus g. If E is a

holomorphic vector bundle over X of rank n, then

h0(E)− h0(KX ⊗ E∗) = deg(E)− n(g − 1).

Remark 1.1.14. To ease the notations, in the sequel, the tensor product

L ⊗ E of a vector bundle E and a line bundle L over X will be denoted by

L + E .

1.2 Fuchsian groups and their Riemann sur-

faces.

All the content of this section is taken from the references:

[5]: Chapter 4.

[24]: Chapter 2.

[46] Chapter 1: §1.4 .

[55] Chapter 1.

We give the classification of Mobius transformations, and some standard

results on Fuchsian groups and their Riemann surfaces.

19

Page 25: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

1.2.1 Classification of Mobius transformations

Let C = C ∪ ∞ and a, b, c, d be complex numbers with ad − bc 6= 0, then

the map g : C→ C defined by

g(z) =az + b

cz + d

is called a Mobius transformation. The set M of all Mobius transformations

equipped with the composition of maps is a group. Moreover, the map given

by A 7→ gA, where

gA(z) =az + b

cz + d, A =

(a b

c d

),

is a group homomorphism φ : GL(2,C)→M, and

kerφ =

(a 0

0 a

), a 6= 0

.

In general, we shall be more concerned with the restriction of φ to SL(2,C).

The kernel of this restriction is

kerφ ∩ SL(2,C) = −I, I.

Hence M is isomorphic to PSL(2,C) = SL(2,C)/−I, I.

Let A =

(a b

c d

)∈ GL(2,C) and tr(A) = a+ d, then the function

tr2(A)

det(A)

is invariant under the transformation A → λA, λ 6= 0, and so it induces a

function on M, namely

Tr(g) =tr2(A)

det(A),

where A is any matrix which projects on g. Notice that Tr(g) is invariant

under conjugation.

20

Page 26: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

We now proceed to classify Mobius transformations. First, we introduce

some normalized Mobius transformations. For each k 6= 0 in C, we define

mk by

mk(z) = kz if k 6= 1,

and

m1(z) = z + 1.

Notice that for all k 6= 0, we have

Tr(mk) = k +1

k+ 2.

If g 6= I is any Mobius transformation, we denote by α ∈ C its fixed point

if it has a unique one, and by α and β in C, α 6= β, if it has two. Now let h

be any Mobius transformation such that

h(α) =∞, h(β) = 0, h(g(β)) = 1 if g(β) 6= β,

then

hgh−1(∞) =∞, hgh−1(0) =

0 if g(β) = β

1 if g(β) 6= β .

If g fixes α and β, then hgh−1 fixes 0 and ∞ and thus hgh−1 = mk for some

k 6= 1. If g fixes α only, then hgh−1 fixes ∞ only and hgh−1(0) = 1 and thus

hgh−1 = m1. Therefore, any nonidentity Mobius transformation is conjugate

to one of the standard form mk.

Definition 1.2.1. Let g 6= I be any Mobius transformation. We say that

1. g is parabolic if and only if g is conjugate to m1(equivalently g has a

unique fixed point in C);

2. g is loxodromic if and only if g is conjugate to mk for some k satisfying

|k| 6= 1 (g has exactly two fixed points in C);

3. g is elliptic if and only if g is conjugate to mk for some k satisfying

|k| = 1 (g has exactly two fixed points in C).

21

Page 27: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

It is convenient to subdivide the loxodromic class by reference to invariant

discs rather than fixed points.

Definition 1.2.2. Let g be a loxodromic transformation. We say that g is

hyperbolic if g(D) = D for some open disc or half-plane D in C. Otherwise

g is said to be strictly loxodromic.

Now, using the fact that Tr(g) is invariant under conjugation, we can give

a complete classification of all Mobius transformations. Indeed, we have the

following results.

Theorem 1.2.1. Let f and g be two Mobius transformations, neither of

which is the identity. Then Tr(f) = Tr(g) if and only if f and g are conju-

gate.

Theorem 1.2.2. Let g 6= I be any Mobius transformation. Then

1. g is parabolic if and only if Tr(g) = 4;

2. g is elliptic if and only if Tr(g) ∈ [0, 4);

3. g is hyperbolic if and only if Tr(g) ∈ (4,+∞);

4. g is strictly loxodromic if and only if Tr(g) 6∈ [0,+∞).

1.2.2 Fuchsian groups

Let H = z ∈ C, Im(z) > 0 be the Poincare upper half-plane, SL(2,R) be

the group of real 2 × 2 matrices with determinant one, R := R ∪ ∞, and

H := H ∪ R. If A =

(a b

c d

)∈ SL(2,R) and gA(z) =

az + b

cz + dis the associated

Mobius transformation, then gA maps H into H. Indeed, for z in H

Im(gA(z)) =Im(z)

|cz + d|2. (1.2.1)

Also, it is clear that gA maps R into R.

22

Page 28: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Let MR denote the group of real Mobius transformations. As we have

seen, the map φ : SL(2,R) → MR given by A −→ gA is a surjective group

homomorphism, and kerφ = ±1 (1 here stands for I2, this will be used for

the rest of the thesis). Thus MR can be identified with

PSL(2,R) = SL(2,R)/±1.

Recall that an automorphisms of H is a biholomorphic bijection from Hto H.

Theorem 1.2.3. The group of automorphisms of H is MR = PSL(2,R).

The group PSL(2,C) inherits the topology of SL(2,C), which is a topo-

logical space with respect to the standard norm of C4. A subgroup Γ of

PSL(2,C) is called discrete if the subspace topology on Γ is the discrete

topology. Thus Γ is discrete if and only if: For any sequence Tn in Γ,

Tn → T ∈ GL(2,C) implies Tn = T for all sufficiently large n. Giving

PSL(2,R) ⊆ PSL(2,C) the subspace topology of PSL(2,C), we have

Definition 1.2.3. A discrete subgroup of PSL(2,R) is called a Fuchsian

group.

Example 1.2.4. The modular group PSL(2,Z) = SL(2,Z)/±1 and its

subgroups are Fuchsian groups.

For γ in PSL(2,R), let Fix(γ) be the set of fixed points of γ in H. By

Theorem 1.2.2, γ cannot be strictly loxodromic, and one can show that

1. If γ is elliptic, then Fix(γ) = zγ, for some zγ in H.

2. If γ is parabolic, then Fix(γ) = sγ, for some sγ in R.

3. If γ is hyperbolic, then Fix(γ) = aγ, bγ, for some aγ, bγ in R.

Definition 1.2.4. Let Γ be a Fuchsian subgroup of PSL(2,R), and z be a

point of H. Then:

23

Page 29: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

1. The point z is called cuspidal or a cusp (resp. elliptic, hyperbolic) if it

is the fixed point of a parabolic (resp. elliptic, hyperbolic) element of

Γ.

2. The set of all elliptic (resp. cuspidal) points of Γ will be denoted by

EΓ (resp. CΓ).

3. The stabilizer Γz of z in Γ is the subgroup defined by

Γz = γ ∈ Γ | γ.z = z.

We have

Proposition 1.2.5. Let Γ be a Fuchsian subgroup of PSL(2,R), and z be a

point of H. Then

1. For any γ ∈ PSL(2,R), we have

(γΓγ−1)γ.z = γ(Γz)γ−1.

In particular, if γ ∈ Γ, then Γγ.z = γ(Γz)γ−1.

2. Γz is cyclic.

3. If z ∈ H, then Γz is finite and Γz = ±1 if z is not elliptic.

Definition 1.2.5. Let z be a point of H. The cardinality of Γz will be called

the order of z, and will be denoted by nz. In particular, nz = 1 if z is not

elliptic.

Remark 1.2.6. Since the conjugate of an elliptic (resp. parabolic) element

of PSL(2,C) is also elliptic (resp. parabolic), we see that Γ acts on EΓ and

CΓ.

24

Page 30: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

1.2.3 The Riemann surface of a Fuchsian group.

In this subsection, Γ will be a Fuchsian subgroup of PSL(2,R), and

H∗ := H ∪ CΓ.

Since Γ acts on CΓ, it then acts on H∗, and so we can form the quotient

X = XΓ := Γ\H∗. Our aim here is to give X a structure of a Riemann

surface.

First we define a topology on H∗ as follows. H is given its usual topology.

For a cusp s 6= ∞, we take as a fundamental system of neighborhoods all

sets of the form:

s ∪ the interior of a circle in H tangent to the real axis at s

If∞ is cusp, then a fundamental system of neighborhoods of∞ are the sets:

∞ ∪ z ∈ H | =(z) > c,

for all positive numbers c. Endowed with this topology, H∗ becomes a Haus-

dorff space, and Γ acts on H∗ by homeomorphisms.

We give X a structure of a Riemann surface in the following way. Let φ

denotes the natural projection map of H∗ onto X = Γ\H∗, p be an arbitrary

point of H∗, and P = φ(p). Then

Lemma 1.2.7. There exists an open neighborhood U of p in H∗ such that

Γp = γ ∈ Γ | (γ.U) ∩ U 6= ∅.

Hence we have a natural injection of Γp\U −→ X, and Γp\U is an open

neighborhood of P in X. If P is neither an elliptic point nor a cusp, then ΓP

contains only the identity, so that the map U −→ Γp\U is a homeomorphism.

We take the pair (Γp\U, φ−1) as a member of the local charts defining the

complex structure on X.

25

Page 31: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

In case p is elliptic, setting

α(z) =z − pz − p

, z ∈ H,

then α(H) is the unit disc D, α(p) = 0, and αΓpα−1 is the group

〈σ′p : w 7−→ ζpw〉,

where ζp is a primitive np−th root of unity, np is the order of p. Then we

define a map ψ : Γp\U −→ C by

ψ(φ(z)) = α(z)np .

We see that ψ is a homeomorphism onto an open subset of C. Thus we

include (Γp\U, ψ) in our local charts.

Finally, if p is a cusp of Γ, and α ∈ SL(2,R) is such that α · s =∞, then

we have

αΓpα−1 = 〈th : z 7−→ z + h〉,

h being a positive number. We define a homeomorphism ψ of Γp\U onto an

open subset of C by

ψ(φ(z)) = exp(2πiα(z)/h),

and include (Γp\U, ψ) in the local charts.

The atlas formed by these local charts gives X the structure of a Riemann

surface.

Definition 1.2.6. Keeping the above notations, we have :

1. A point P in X will be called elliptic, if it corresponds to an elliptic

point of Γ. The function t = wnp will be called the local parameter at

P (or at p ).

2. A point P in X will be called a cusp, if it corresponds to a cusp of Γ.

The function qh := exp(2πiz/h) will be called the local parameter at

P (or at p ).

26

Page 32: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

An important class of Fuchsian groups consists of Fuchsian groups of the

first kind; they are defined as follows:

Definition 1.2.7. Γ is called a Fuchsian group of the first kind, if XΓ = Γ\H∗

is compact. A subgroup Γ1 of SL(2,R) is a Fuchsian group of the first kind,

if its associated subgroup Γ1 := ±1\Γ1.±1 of PSL(2,R) is of the first

kind.

As we have seen above, Γ acts on the sets EΓ (resp. CΓ) of elliptic (resp.

cuspidal) points of Γ. Set

EΓ := Γ\EΓ, SΓ := Γ\CΓ. (1.2.2)

Proposition 1.2.8. If Γ is of the first kind, then EΓ and SΓ are both finite

sets. Their cardinality will be respectively denoted by eΓ and cΓ.

Also, we have following useful result:

Proposition 1.2.9. Suppose that Γ is of the first kind, then any finite index

subgroup of Γ is also a Fuchsian group of the first kind.

Suppose that p ∈ H is an elliptic fixed point of Γ. By Proposition 1.2.5,

the orders of γ.p and p are equal for all γ ∈ Γ. Therefore, the order of an

elliptic point can be lifted to X.

Definition 1.2.8. If P ∈ X is not a cuspidal point, then we define the order

of P to be the order of any point in H corresponding to P .

1.2.4 Congruence subgroups of PSL(2,Z)

The modular group PSL(2,Z) and its finite index subgroups form a very rich

category of Fuchsian groups of the first kind. Here we list some well known

examples.

27

Page 33: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Let

T =

(1 1

0 1

), S =

(0 −1

1 0

), P = ST =

(0 −1

1 1

).

Then S is elliptic of order 2 with fixed point i, and P is elliptic of order 3 and

fixes ρ = e2πi/3. Moreover PSL(2,Z) is generated by S and T , or equivalently,

by S and P .

An important class of subgroups of PSL(2,Z) are the so-called congruence

subgroups. A subgroup Γ of PSL(2,Z) is called a congruence subgroup if it

contains some principal congruence subgroup Γ(n) for some positive integer

n, where Γ(n) is defined by

Γ(n) = γ ∈ SL(2,Z), γ ≡ ±1 mod n/±1.

The smallest such n is called the level of Γ.

One easily shows that Γ(n) is a normal subgroup of PSL(2,Z) of finite

index and that the conjugates in PSL(2,Z) of a congruence subgroup are also

congruence subgroups.

Example 1.2.10. Important examples of congruence subgroups of PSL(2,Z)

are:

Γ1(n) = γ ∈ SL(2,Z), γ ≡ ±

(1 ∗0 1

)mod n/±1,

Γ0(n) =

(a b

c d

)∈ SL(2,Z), c ≡ 0 mod n

/±1,

Γ0(n) =

(a b

c d

)∈ SL(2,Z), b ≡ 0 mod n

/±1.

It is clear that Γ(n) ⊆ Γ1(n) ⊆ Γ0(n) and that Γ0(n), Γ1(n), Γ0(n) are of

level n. Note that Γ0(n) is conjugate to Γ0(n) by(n 00 1

)

28

Page 34: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

1.3 Automorphic forms.

In this section we give the definition of automorphic forms and provide some

well known examples.

1.3.1 Definition of automorphic forms.

In this subsection we follow The second chapter of [55].

Let k ∈ Z, γ ∈ GL+(2,R), Γ be a Fuchsian, and f be a meromorphic

function on H, we set

f |kγ = (det γ)k2 J−kγ f γ , (1.3.1)

where Jγ(z) = cz + d if γ =

(∗ ∗c d

). We have for all α, γ ∈ Γ,

Jαγ(z) = Jα(γz)Jγ(z), z ∈ H. (1.3.2)

Notice that for k odd, Jk−γ = −Jkγ , hence f |k(−γ) = −f |kγ. If k is even, then

f |k(−γ) = f |kγ.

Definition 1.3.1. Let Γ be a Fuchsian subgroup of SL(2,R), and k be an in-

teger. A meromorphic function f on H is called an unrestricted automorphic

form of weight k for Γ, if it satisfies

(i) f |kγ = f for all γ ∈ Γ.

If in addition

(ii) f is meromorphic at every cusp of Γ,

then f is called an automorphic form of weight k for Γ. When k = 0, f is

called an automorphic function.

29

Page 35: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

By a cusp of Γ we mean a cusp of its associated subgroup Γ := ±1\Γ.±1of PSL(2,R). The last condition in the definition means the following. If Γ

has no cusp, then this condition is empty. In case Γ has a cusp s, then the

stabilizer of s in Γ is equal to Γs for some subgroup Γs ⊆ Γ. Moreover, if

α =

(a b

c d

)∈ SL(2,R) is such that α · s =∞, then we have

αΓsα−1±1 =

⟨±(

1 h

0 1

)⟩,

with h being a positive real number referred to as the cusp width at s. From

(i) we deduce that fs := f |k(α−1) is invariant under the group αΓα−1.

Case I: k is even. Here fs is invariant under the translation

th : z 7−→ z + h,

hence fs = Gs(qh) for some meromorphic function Gs in the punctured disc

: 0 < |qh| < r, where r is a positive number, and qh := exp(2πiz/h). We say

that f is meromorphic at the cusp s if Gs is meromorphic at qh = 0, this

means that f , when it is nonzero, has a Laurent expansion∑n≥ns

anqnh , 0 < |qh| < r,

where ns is the order ords(Gs) of Gs at 0. We define the order of f at s by

ords(f) = ord0(Gs).

Case II: k is odd. If Γ contains −1, then condition (i) implies that

f = −f . Hence there is no nontrivial automorphic form of weight k for Γ.

Therefore we may assume that −1 6∈ Γ. Then αΓsα−1 is generated by

(1 h0 1

),

or by −(

1 h0 1

). We say that the cusp s is regular or irregular, accordingly. If s

is regular, then the meromorphy at s is treated in the same way as in Case

I. If s is irregular, then fs(z + h) = −fs. Letting

gs(z) = exp(−πiz/h)fs, z ∈ H,

then gs is invariant under the translation th : z 7−→ z + h, hence gs = Gs(qh)

for some meromorphic function Gs in the punctured disc : 0 < |qh| < r,

30

Page 36: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

where r is a positive number, and qh := exp(2πiz/h). We say that f is

meromorphic at the cusp s if Gs is meromorphic at qh = 0. We define the

order of f at s by ords(f) = ord0(Gs) + 1/2.

f is said to be holomorphic (resp. cuspidal) at s if ords(f) ≥ 0 (resp.

ords(f) > 0). An unrestricted automorphic form which is meromorphic at

any point of H∗, is called a meromorphic automorphic form. An automor-

phic form which is holomorphic at any point of H∗, is called a holomorphic

automorphic form. A holomorphic automorphic form which is cuspidal at

any cusp is called a cusp automorphic form, or simply a cusp form. The vec-

tor space of automorphic forms of weight k for Γ will be denoted by Ak(Γ).

The vector spaces of unrestricted, holomorphic, cusp automorphic forms of

weight k for Γ will be respectively denoted by Gk(Γ), Mk(Γ), and Sk(Γ).

If Γ is Fuchsian group of the first kind, then Mk(Γ) and Sk(Γ) have finite

dimensions. Indeed, we have

Theorem 1.3.1. Let Γ be a Fuchsian group of the first kind, and k be an

even integer. Let g be the genus of XΓ = Γ\H∗, e and c be the number of

elliptic and cuspidal points of XΓ respectively. If n1, . . . , ne are the orders of

the elliptic points, then

dimMk(Γ) =

(k − 1)(g − 1) + (k/2)c+∑e

i=1[k(ni − 1/2ni)] (k > 2)

g + c− 1 (k = 2, c > 2)

g (k = 2, c = 0)

1 (k = 0)

0 (k < 0),

31

Page 37: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

and

dimSk(Γ) =

(k − 1)(g − 1) + (k/2− 1)c+∑e

i=1[k(ni − 1/2ni)] (k > 2)

g (k = 2)

1 (k = 0, m = 0)

1 (k = 0, m > 0)

0 (k < 0),

where [.] denotes the integral part of a real number.

1.3.2 Examples of modular forms

The main reference for this subsection is the first chapter of [38]. When

working with the modular group PSL(2,Z) and its subgroups, the word au-

tomorphic is replaced by modular.

We start our examples by the well known Eisenstein series. They are

defined for every even integer k ≥ 2 and z ∈ H by

Gk(z) =∑n

∑m

′ 1

(nz +m)k,

where the symbol∑ ′ means that the summation is over the pairs (m,n)

different from (0, 0). Notice that this series is not absolutely convergent for

k = 2. If we normalize the Eisenstein series by letting

Ek = 2ζ(k)Gk,

where ζ is the Riemann Zeta function, then we get

Ek(z) = 1− 2k

Bk

∞∑n=1

σk−1(n)qn, q = e2πiz.

Here Bk is the k-th Bernoulli number and σk−1(n) =∑

d|n dk−1. The most

familiar Eisenstein series are:

E2(z) = 1− 24∞∑n=1

σ1(n)qn,

32

Page 38: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

E4(z) = 1 + 240∞∑n=1

σ3(n)qn,

E6(z) = 1− 504∞∑n=1

σ5(n)qn.

For k ≥ 4, the series Ek are holomorphic modular forms of weight k. The

Eisenstein series E2 is holomorphic on H and at the cusps, but it is not a

modular form as it does not satisfy the modularity condition. The Eisenstein

series E2 is an example of a quasimodular form and plays an important role

in the theory of vector valued modular forms, see [37]. Moreover, E2 satisfies

E2(z) =1

2πi

∆′(z)

∆(z), (1.3.3)

where ∆ is the weight 12 cusp form for PSL(2,Z) given by

∆(z) = q∏n≥1

(1− qn)24.

The Eisenstein series satisfy the Ramanujan relations

6

πiE ′2 = E2

2 − E4 , (1.3.4)

3

2πiE ′4 = E4E2 − E6 , (1.3.5)

1

πiE ′6 = E6E2 − E2

4 . (1.3.6)

The Dedekind j-function given by

j(z) =E3

4 − E26

is a modular function and it generates the function field of modular functions

for PSL(2,Z).

33

Page 39: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Chapter 2

Vector-valued automorphic

forms and vector bundles

The aim of this chapter is to realize vector-valued automorphic forms as

global sections of holomorphic vector bundles. More precisely, given a Fuch-

sian subgroup Γ of SL(2,R), a finite-dimensional representation R of Γ,

and an integer k such that the pair (R, k) is simple, see Definition 2.1.2,

we construct a holomorphic vector bundle EΓ,R,k over the Riemann surface

XΓ = Γ\H∗ associated to Γ. The global meromorphic sections of EΓ,R,k will

correspond to vector-valued automorphic forms of weight k and multiplier

R. More interestingly, the spaces Mk(Γ, R), Sk(Γ, R) of holomorphic, cusp

vector-valued automorphic forms of weight k and multiplier R, will be respec-

tively isomorphic to H0(XΓ,O(EΓ,R,−k)) and H0(XΓ,O(−DS,R,−k+EΓ,R,−k)),

where DS,R,−k is a certain holomorphic line bundle depending on the cuspidal

points of XΓ. As an immediate consequence, we deduce that Mk(Γ, R) and

Sk(Γ, R) are finite-dimensional C−vector spaces if and only if Γ is Fuchsian

group of the first kind, see Theorem 2.5.9.

34

Page 40: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

2.1 Notations

For the remainder of this thesis, n is a positive integer, Γ is a Fuchsian

subgroup of SL(2,R), R is a representation of Γ in GL(n,C). We will denote

the matrix R(γ), γ ∈ Γ, by Rγ. R∗ will be the adjoint representation to R,

that is

R∗γ = (Rγ−1 )T ,

where (.)T is the transpose operator. Γ will be the image of Γ in PSL(2,R),

that is Γ = ±1\Γ.±1. E will be the set of elliptic fixed points of Γ and

C the set of its cusps in R ∪ ∞. We define

H∗ := H ∪ C , S := Γ\C, E := Γ\E,

X = X(Γ) := Γ\H∗ , X ′ = X ′(Γ) := Γ\(H− E) ,

Y ′ = Y ′(Γ) := H− E.

We have an unbranched covering map

π : Y ′ −→ X ′,

and since Γ acts properly and discontinuously on Y ′, then the group of cov-

ering transformations is

Deck (Y ′/X ′) = Γ.

Moreover, this is a Galois covering in the sense that for all y1 and y2 in Y ′

with π(y1) = π(y2), there exists σ ∈ Γ such σ(y1) = y2.

Also, when Γ does not contain −1, we can identify Γ with Γ via the

canonical map, and so Γ will be viewed as a group of matrices as well as a

group of Mobıus transformations.

If k is an integer, then we define the slash operator |k on meromorphic

maps F : H −→ Cm, m being an integer, by

F |kγ = J−kγ F γ , γ ∈ GL+(2,R) , (2.1.1)

where Jγ(z) = cz + d if γ =

(∗ ∗c d

).

35

Page 41: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Definition 2.1.1. Let k be an integer. An unrestricted vector-valued au-

tomorphic form for Γ of multiplier R and weight k is a meromorphic map

F : H −→ Cn satisfying

F |kγ = Rγ F.

The C-vector space of these maps will be denoted by Gk(Γ, R).

If R = R1 ⊕ R2 is the direct sum of two representations R1 and R2,

then it is clear that Gk(Γ, R) = Gk(Γ, R1)⊕Gk(Γ, R1), and when Γ contains

−1, then after [3], the representation R will be called even (resp. odd) if

R(−1) = In (resp. R(−1) = −In). Since −1 commutes with Γ, then modulo

a conjugation in GL(n,C), any representation R may be decomposed into a

direct sum R = R+ ⊕ R− of even and odd representations. It follows from

Definition 2.1.1 that for an even (resp. odd) representation R there are no

nontrivial unrestricted vector-valued automorphic forms of odd (resp. even)

weight. Therefore

Gk(Γ, R) =

Gk(Γ, R+) if k is even,

Gk(Γ, R−) if k is odd.

This shows that it suffices to treat the case when both R and k are even, or

when they are both odd.

To avoid trivial cases, we adopt the following definition:

Definition 2.1.2. The pair (R, k), k being an integer, is called simple if

1. Γ does not contain −1, or

2. Γ contains −1, R and k have the same parity.

Remark 2.1.1. In the sequel, we only consider the integers k such that

(R, k) is a simple pair. In particular, R is either even or odd when −1 ∈ Γ.

36

Page 42: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Definition 2.1.3. The integer ε = ε(R) defined by

ε =

0 if R is even, or −1 6∈ Γ

1 if R is odd

is called the parity index of R.

Notice that ε is the smallest nonnegative integer k such that (R, k) is a

simple pair.

Remark 2.1.2. 1. In this work, almost all the constructed objects will

depend on at least three parameters, namely : Γ, R, and the weight k.

When the context is clear, we will only keep the relevant indices.

2. If the context is clear, an object defined on Y ′ which is invariant under

the covering transformations, will be considered as an object defined

on X ′.

2.2 The family EΓ,R,k of vector bundles

The goal of this section is to associate to each triplet (Γ, R, k), with (R, k)

being a simple pair, a holomorphic vector bundle EΓ,R,k on X. This will

be achieved by constructing a 1-cocycle in Z1(U ,GL(n,OX)), where U is

an open cover of X. Then EΓ,R,k will be its associated holomorphic vector

bundle.

2.2.1 The unbranched case.

Let (R, k) be a simple pair, we construct a matrix-valued automorphic form

of weight k and multiplier R on the unbranched covering Y ′ of X ′. More

precisely, we have

37

Page 43: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Theorem 2.2.1. There exists a holomorphic map

Ψk = ΨR,k : Y ′ −→ GL(n,C)

such that for all σ ∈ Γ, we have

Ψk|k σ = Ψk Rσ−1 .

Proof. Since π : Y ′ −→ X ′ is an unbranched Galois covering, there exists an

open covering U = (Ui)i∈I of X ′ and homeomorphisms

φi = (π, ηi) : π−1(Ui) −→ Ui × Γ

which are compatible with the action of Γ in the sense that φi(y) = (x, σ)

implies φ(τy) = (x, τσ). In other words, the mapping ηi : π−1(Ui) −→ Γ

satisfies ηi(τy) = τηi(y) for all y ∈ π−1(Ui) and τ ∈ Γ.

On Yi = π−1(Ui) we define Ψk,i : Yi −→ GL(n,C) by

Ψk,i(y) = J−kηi(y)−1(y)Rηi(y)−1 , y ∈ Yi,

which is well defined since the expression

J−kσ Rσ

is well defined on Γ for a simple pair (R, k). Moreover, Ψk,i is holomorphic

since it is locally constant, and so Ψk,i ∈ GL(n,O(Yi)). Now, if y ∈ Yi and

σ ∈ Γ, then

Ψk,i(σy) = J−kηi(σy)−1(σy)Rηi(σy)−1

= J−kηi(y)−1σ−1(σy)Rηi(y)−1σ−1

= J−kσ−1(σy)J−kηi(y)−1(y)Rηi(y)−1Rσ−1

= Jkσ (y)Ψk,i(y)Rσ−1 .

38

Page 44: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Hence, for σ ∈ Γ,

Ψk,i|k σ = Ψk,iRσ−1 on Yi.

Therefore, Ψk,i has the required property on Yi. If we set

Fk,ij = FR,k,ij = Ψk,iΨ−1k,j ∈ GL(n,O(Yi ∩ Yj)) ,

then for all y ∈ Yi ∩ Yj we have Fk,ijσ = Fk,ij, that is, the Fk,ij is invariant

under covering transformations and hence may be considered as an element

of GL(n,O(Ui ∩ Uj)). Thus, we have a cocycle

(Fk,ij) ∈ Z1(U ,GL(n,O)) .

As X ′ is a noncompact Riemann surface, there exist elements

Fk,i = FR,k,i ∈ GL(n,O(Ui))

such that

Fk,ij = Fk,iF−1k,j on Ui ∩ Uj , (2.2.1)

see [13]. We now look at the Fk,i as elements of GL(n,O(Yi)) that are

invariant under covering transformations and set

Ψi = F−1k,i Ψk,i ∈ GL(n,O(Yi)) .

Then, for every σ ∈ Γ, we have

Ψk,i|k σ = (Fk,iσ)−1Ψk,i|kσ = F−1i Ψk,iRσ−1 = ΨiRσ−1 .

Moreover, on Yi ∩ Yj we have

(Ψk,i)−1Ψk,j = Ψ−1

k,iFk,iF−1k,j Ψk,j = Ψ−1

k,iFk,ijΨk,j = Ψ−1k,iΨk,iΨ

−1k,jΨk,j = 1.

Thus, the Ψk,i’s define a global function Ψk ∈ GL(n,O(Y ′)) with

Ψk|k σ = ΨkRσ−1 , σ ∈ Γ .

39

Page 45: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

We see that the ΨR,k constructed above depends on the choice of the

elements FR,k,i ∈ GL(n,O(Ui)), i ∈ I, let us call it Ψk,F . Suppose that

GR,k,i ∈ GL(n,O(Ui)), i ∈ I, is another choice, then by (2.2.1), we have

Fk,iF−1k,j = Fk,ij = Gk,iG

−1k,j on Ui ∩ Uj .

Hence, the element AF,G ∈ GL(n,O(X ′)) given by

AF,G|Ui = G−1k,iFk,i on Ui

is well-defined and we

Ψk,G = AF,GΨk,F . (2.2.2)

Now, we take an arbitrary choice of the elements FR,k,i ∈ GL(n,O(Ui)),

i ∈ I, and we fix it for the rest of the thesis. Its corresponding map will be

our ΨR, k.

Remark 2.2.2. From the proof of Theorem 2.2.1, we see that

F ∗R,k,i(F∗R,k,j)

−1 = FR∗,−k,i F−1R∗,−k,j on Ui ∩ Uj ,

and hence, as in the above discussion, we have

(ΨR,k)∗ = ∆R,k ΨR∗,−k,

for some ∆R,k ∈ GL(n,O(X ′)).

2.2.2 The elliptic case.

Now, let e be an elliptic fixed point of Γ, then the stabilizer of e in Γ is equal

to Γe for some subgroup Γe ⊆ Γ. Moreover, Γe = 〈σe〉 with σe = γe, for some

γe in Γe. If e denotes the complex conjugate of e, then we have a character

µe on Γe defined by

µe(γ) = Jγ(e), γ ∈ Γe, (2.2.3)

40

Page 46: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

such that µe(−1) = −1, when −1 ∈ Γ. Also, the holomorphic function on Hgiven by

fe(z) =1

z − e, z ∈ H,

verifies

fe|1γ = µe(γ)fe, γ ∈ Γe.

Lemma 2.2.3. There exists a meromorphic map Ψk,e = ΨR,k,e in H, having

a pole only at e, such that for all γ ∈ Γe, we have

Ψk,e|kγ = Ψk,eRγ−1 .

Proof. If

α(z) =z − ez − e

, z ∈ H,

then α(H) is the unit disc D, α(e) = 0, and αΓeα−1 is the group

〈σ′e : w 7−→ ζew〉

where ζe is a primitive ne−th root of unity, ne being the order of e.

Let R = Rk,e := µ−ke R, then R is a well defined representation of Γe. Since

Rneγe = 1, then Rγe is diagonalizable and for some Ak,e = AR,k,e ∈ GL(n,C)

we have

Ak,eRσeA−1k,e = diag (λ1, . . . , λn), λnej = 1 .

Write λj = ζmk,je , 0 ≤ mk,j ≤ ne − 1, and for w ∈ D∗ set

Φk,e(w) = ΦR,k,e(w) := diag (w−mk,1 , . . . , w−mk,n) .

We have

Φk,e(σ′ew) = Φk,e(ζew)

= diag (ζ−mk,1e w−mk,1 , . . . , ζ

−mk,ne w−mk,n)

= diag (λ−11 w−mk,1 , . . . , λ−1

n w−mk,1)

= Φk,e(w)diag (λ1, . . . , λn)

= Φk,e(w)Ak,eRσ−1eA−1k,e .

41

Page 47: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Therefore,

Φk,e(σ′e)Ak,e = Φk,eAk,eRσ−1 ,

If we set

Φk,e = (Φk,eα)Ak,e ,

then for all σ ∈ Γe,

Φk,e(σ) = Φk,eRσ−1 .

Now, let

Ψk,e = fke Φk,e,

then for all γ ∈ Γe,

Ψk,e|kγ = µke(γ)fke Φk,eR(γ)−1 = µke(γ)Ψk,eR(γ)−1 .

But for γ ∈ Γe we have

R(γ)−1 = µ−ke (γ−1)Rγ−1 .

Hence for all γ ∈ Γe, we have

Ψk,e|kγ = Ψk,eRγ−1

as desired.

Notice that the construction of ΨR,k,e depends on the choice of the AR,k,e.

We fix a choice for the matrix AR,k,e once for all, and its corresponding map

will be our ΨR,k,e .

Remark 2.2.4. The relationship between (ΨR,k,e) and ΨR∗,−k,e is given by

the following: Since

AR,k,e(µ−ke (γe)Rγe

)A−1R,k,e = diag (ζ

mR,k,1e , . . . , ζ

mR,k,ne ),

where (mR,k,j)j is a sequence of integers in [0, ne), then

A∗R,k,e(µke(γe)R

∗γe

)(A∗R,k,e)

−1 = diag (ζne−mR,k,1e , . . . , ζ

ne−mR,k,ne ).

42

Page 48: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

But

diag (ζne−mR,k,1e , . . . , ζ

ne−mR,k,ne ) = AR∗,−k,e

(µke(γe)R

∗γe

)A−1R∗,−k,e.

This implies that the matrix

BR,k,e = A∗R,k,eA−1R∗,−k,e

commutes with

diag (ζne−mR,k,1e , . . . , ζ

ne−mR,k,ne ),

hence with ΦR∗,−k,e, and we have

(ΨR,k,e)∗ = (α(z))ne BR,k,e ΨR∗,−k,e.

Definition 2.2.1. With the above notations, we define the k−order of R at

e to be

νk,e(R) = −(1/ne)n∑j=1

mR,k,j,

where the integers 0 ≤ mR,k,j ≤ ne − 1 are such that

AR,k,e(µ−ke (γe)Rγe )A−1

R,k,e = diag (ζmR,k,1e , . . . , ζ

mR,k,ne ).

Since it only depends on the class of e modulo Γ, we define the k−order of

R at an elliptic point P ∈ X to be

νk,P (R) = νk,e(R),

with e ∈ H is an elliptic point corresponding to P .

2.2.3 The cuspidal case.

Similarly, If s is a cusp of Γ, then the stabilizer of s in Γ is equal to Γs

for some subgroup Γs ⊆ Γ. Also, Γs = 〈γs〉 with γs in Γs, and we have a

character µs on Γs defined by

µs(γ) = Jγ(s), γ ∈ Γs, (2.2.4)

43

Page 49: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

such that µs(−1) = −1, when −1 ∈ Γ. Notice that for γ =

(aγ bγcγ dγ

)∈ Γs,

a simple computation shows that

s = (aγ − dγ)/2cγ.

This implies that

µs(γ) = Trace(γ)/2, γ ∈ Γs, (2.2.5)

which is also valid for s =∞. Since the trace of γ is ±2, we see that

µs(γ) = ±1 γ ∈ Γs.

Moreover, the holomorphic function on H defined by

fs(z) =

1z−s , z ∈ H, if s 6=∞

1 if s =∞ ,

verifies

fs|1γ = µs(γ)fs, γ ∈ Γs.

If α =

(a b

c d

)∈ SL(2,R) is such that α · s = ∞, then, using the notations

of [55], we have

αΓsα−1 · ±1 = 〈th : z 7−→ z + h〉,

with h being a positive real number referred to as the cusp width of s. Also,

from s = α−1 · ∞ = −d/c, we have fs = cα /Jα for some nonzero constant

cα ∈ C.

Lemma 2.2.5. [50] There exists a holomorphic map Ψk,s = ΨR,k,s in H,

such that for all γ ∈ Γs, we have

Ψk,s|kγ = Ψk,sRγ−1 .

44

Page 50: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Proof. Let R = Rk,s := µ−ks R, then R is a well-defined representation of Γs.

Take Bk,s = BR,k,s ∈M(n,C) such that

Rγs = exp(2πiBk,s).

If we set

Φk,s(z) = exp(−2πi

z

hBk,s

), z ∈ H ,

we have

Φk,s(thz) = exp

(−2πi

z + h

hBk,s

)= exp

(−2πi

z

hBk,s − 2πiBk,s

)= R(γs)−1Φk,s(z)

= Φk,s(z)R(γs)−1 .

Hence if

Ψk,s = fks Φk,sα = ckα Φk,s|kα,

then Ψk,s has the required property.

Remark 2.2.6. Note that Ψk,s depends on the choice of Bk,s. In the sequel

we take Bk,s such that

0 ≤ <(λ) < 1

for every eigenvalue λ of Bk,s, which is always possible, see [18]. With this

choice, one can check that [18, 16]

BR∗,−k,s = In − (BR,k,s)t.

Therefore

(ΦR,k,s)∗ = qh(α)ΦR∗,−k,s,

where qh(α(z)) = exp (2πiα(z)/h), z ∈ C.

45

Page 51: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Definition 2.2.2. Let Bk,s = BR,k,s ∈M(n,C) such that

µ−ks (γs)Rγs = exp(2πiBk,s),

and for all eigenvalues λ of Bk,s

0 ≤ <(λ) < 1 .

Then the number

νk,R(s) = −Trace(Bk,s)

is called the k−order of R at s. Since it only depends on the class of s modulo

Γ, we define the k−order of R at a cuspidal point P ∈ X to be

νk,P (R) = νk,s(R),

where s ∈ H∗ is a cusp corresponding to P .

Definition 2.2.3. If for a point P ∈ X \ (E ∪ S ) we set νk,P (R) = 0, then

the k−divisor of R is defined by

DR,k =∑P∈X

νk,P (R)P.

2.2.4 Construction of the 1−cocycle.

We now come to the main construction in this section. Write

E ∪S = aii≥1

for the discrete closed set of classes of cusps and elliptic fixed points in X.

In particular, they correspond to inequivalent points ai in H∗. For each ai

we choose a neighborhood in X in the following way:

If ai is a cusp, let Ui be a neighborhood of ai in X given by

Ui = (Γai\Dai) ∪ ai,

46

Page 52: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

where Dai is a horocycle in H tangent at ai [55] (if ai =∞, this horocycle is

a half-plane). Here Γai is infinite cyclic.

If ai is an elliptic fixed point, then Ui is given by

Ui = Γai\Dai ,

where Dai is an open disc in H centered at ai. Here Γai is a finite cyclic

group.

Further, these neighborhoods can be taken such that for ai 6= aj, we have

Ui ∩ Uj = ∅. Also, for each ai choose a chart z for X such that z(Ui) = Dand z(ai) = 0 where D is the unit disc.

Set U0 = X ′, V0 = H − E = π−1(X ′) = Y ′, and Vi = π−1(Ui − ai),i ≥ 1. Then U = (Ui)i≥0 is an open covering of X, and we have

Vi =⊔

γ∈Γ/Γai

Zγ,

where γ is a representative of γ in Γ/Γai , and Zγ is a connected component

of Vi. If Ψk,0 = ΨR,k is the map constructed in Theorem 2.2.1, then we have

the following.

Proposition 2.2.7. [50] If ai is a cusp, then there exists a holomorphic map

Ψk,i : Vi −→ GL(n,C)

such that Ψk,i Ψ−1k,0 is invariant under Deck (Zγ/Ui − ai) for all connected

components Zγ of Vi.

Proof. We have

Vi = ΓDai =⊔

γ∈Γ/Γai

γDai =⊔

γ∈Γ/Γai

Dγai .

Here we have a disjoint union of connected components Zγ = γDai of Vi and

γ is a chosen representative of γ in Γ/Γai . Then

Zγ −→ Ui − ai

47

Page 53: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

is a universal covering with the covering transformations given by the fun-

damental group

π1(Ui − ai) = Γγai .

By Theorem 2.2.1 and Lemma 2.2.5, the maps Ψk,γai and Ψk,0 on Zγ have

the same automorphic behaviour with respect to Γγai , and so Ψk,γai Ψ−1k,0 is

invariant under Γγai , hence under Γγai . The map Ψk,i defined on Vi by

Ψk,i|Zγ = Ψk,γai

is then a holomorphic map on Vi such that Ψk,i Ψ−1k,0 is invariant under

Deck (Zγ/Ui − ai) for all connected components Zγ of Vi.

Proposition 2.2.8. [50] If ai is an elliptic fixed point, then there exists a

holomorphic map

Ψk,i : Vi −→ GL(n,C)

such that Ψk,i Ψ−1k,0 is invariant under Deck (Zγ/Ui − ai) for all connected

components Zγ of Vi. Moreover, Ψk,i is meromorphic at each point of π−1ai.

Proof. If ai is the class of an elliptic fixed point ai in H, then

Ui − ai = Γai\D∗ai ,

where D∗ai = Dai − ai is the punctured disc. Here, the stabilizer Γai of

cyclic of finite order ki. Furthermore, we have

Vi =⊔

γ∈Γ/Γai

γD∗ai =⊔

γ∈Γ/Γai

Zγ ,

where the Zγ = γD∗ai are the connected components of Vi. In the meantime,

the group of covering transformations of the unbranched covering Zγ −→ D∗aiis given by the fundamental group

π1(Ui − ai) = Γγai .

48

Page 54: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Indeed, it is isomorphic to the covering

D∗ −→ D∗

z 7−→ zki

for which the covering transformations are given by the group 〈σi : z 7−→ ζiz〉where ζi is a primitive ki−th root of unity, ki is the order of ai.

By Theorem 2.2.1 and Lemma 2.2.3, the maps Ψk,γai and Ψk,0 on Zγ

have the same automorphic behavior with respect to Γγai , and so Ψk,γai Ψ−1k,0

is invariant under Γγai , hence under Γγai . The map Ψk,i defined on Vi by

Ψk,i|Zγ = Ψk,γai

is then a holomorphic map on Zγ = γD∗ai , and meromorphic at γai ∈ γDai .

Hence, Ψk,i is holomorphic in Vi and meromorphic at each point of π−1ai.Also, Ψk,i Ψ

−1k,0 is invariant under Deck (Zγ/Ui − ai) for all connected com-

ponents of Zγ of Vi.

Now define the 1−cocycle

Fk,ij = Ψk,iΨ−1k,j ∈ GL(n,O(Vi ∩ Vj)) if i 6= j and Fk,ii = id . (2.2.6)

Theorem 2.2.9. [50] The 1−cocycle (Fk,ij) defines a 1−cocycle (Fk,ij) in

Z1(U ,GL(n,O)) such that Fk,ij = π∗Fk,ij.

Proof. We need to prove that the Fk,ij’s defined in (2.2.6) can be considered

as elements of GL(n,O(Ui ∩ Uj)). By construction, for i 6= 0 and j 6= 0 we

have

Ui ∩ Uj = Vi ∩ Vj = ∅ , U0 ∩ Ui = Ui − ai , V0 ∩ Vi = Vi .

Thus we only need to prove that Ψk,0Ψ−1k,i ∈ GL(n,O(Vi)) defines an el-

ement of GL(n,O(Ui − ai)), that is to show that Ψk,0Ψ−1k,i is invariant

under the action of Deck (Zγ/Ui − ai) which is a direct consequence of

49

Page 55: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Proposition 2.2.7 and Proposition 2.2.8. Therefore, there exists an element

(Fk,ij) ∈ Z1(U ,GL(n,O)) such that Fk,ij = π∗Fk,ij.

According to Theorem 1.1.2, a 1−cocycle in Z1(U ,GL(n,O)) gives rise

to a holomorphic vector bundle on X.

Definition 2.2.4. We define EΓ,R,k to be the holomorphic vector bundle

p : EΓ,R,k −→ X of rank n whose transition functions are the (Fk,ij). If the

context is clear EΓ,R,k will be denoted by Ek.

If R is the trivial character χ0 sending all elements of Γ to 1, then EΓ,χ0,k

is a line bundle which will be denoted by Lχ0,k = Lk. Since χ0 is an even

representation, then Lk is defined only for even k’s when −1 ∈ Γ, and for

any k when −1 6∈ Γ.

2.3 Behavior at the cusps

In this section we shall make explicit the behavior of unrestricted vector-

valued automorphic forms at cusps.

Let s be a cusp of Γ. As we have seen in §2.2.3, the stabilizer of s in

Γ is equal to Γs for some subgroup Γs ⊆ Γ. Also, Γs = 〈γs〉 with γs in Γs.

Moreover, if α ∈ SL(2,R) is such that α · s =∞, then we have

αΓsα−1 = 〈th : z 7−→ z + h〉,

h being the cusp width of s.

Let F be an unrestricted vector-valued automorphic form for Γ of mul-

tiplier R and weight k, Fs := Ψ−k,sF , and Fs := Fsα−1. By Lemma 2.2.5,

Fs is invariant under Γs, hence under Γs. Therefore Fs is invariant under

αΓsα−1, and so

Fsth = Fs .

50

Page 56: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Hence Fs(z) = Gs(qh) for some meromorphic map

Gs : qh; 0 < |qh| < r −→ Cn,

where r is a positive number, and qh := exp(2πiz/h). We say that F is

meromorphic at the cusp s if Gs is meromorphic at qh = 0, this means that

Gs, when it is nonzero, has a Laurent expansion∑n≥ns

anqnh , 0 < |qh| < r,

where ns is the order ords(Gs) of Gs at 0. We define the order of F at s by

ords(F ) := ord0(Gs).

F is said to be holomorphic at s if ords(F ) ≥ 0. An unrestricted vector-

valued automorphic form which is meromorphic (resp. holomorphic) at every

point of H∗, is called a vector-valued automorphic form (resp. holomorphic

vector-valued automorphic form). The vector space of vector-valued automor-

phic forms for Γ of multiplier R and weight k will be denoted by Ak(Γ, R),

and its subspace of holomorphic vector-valued automorphic forms will be

denoted by Mk(Γ, R).

2.4 The divisor of a vector-valued automor-

phic form

The goal of this section is to associate to each vector-valued automorphic

form F a divisor div(F ) defined on the Riemann surface X.

Suppose that F is a vector-valued automorphic form for Γ of multiplier

R and weight k. Then by Remark 1.1.5, we have ordγw(F ) = ordw(F ) for

any w ∈ H∗, γ ∈ Γ. This means that the order of F can be lifted to X.

If P ∈ X corresponds to a point w ∈ H, then set νP (F ) = ordw(F )/nw,

where nw is the order of the point w, see Definition 1.2.5. If P ∈ X corre-

sponds to a cusp s, then set νP (F ) = ords(F ). We define the divisor of F on

51

Page 57: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

X to be the formal sum

div(F ) =∑P∈X

νP (F )P. (2.4.1)

From this we see that F is holomorphic in H∗ exactly when div(F ) ≥ 0.

Proposition 2.4.1. we have

Mk(Γ, R) =F ∈ Ak(Γ, R) | div(F ) ≥ 0

.

As usual, one can define a cusp vector-valued automorphic form of mul-

tiplier R and weight k, to be an element F of Mk(Γ, R) vanishing at all the

cusps of Γ. More precisely, suppose that s is cusp, then using notations of

Lemma 2.2.5, and Definition 2.2.2, we have

µks(γs)Rγs = exp(2πiBR,−k,s),

such that the eigenvalues λ−k,s,j of BR,k,s satisfy the condition

0 ≤ <(λ−k,s,j) < 1, 0 ≤ j ≤ n.

In terms of the local parameter qh at s, we have

F |kα−1 = ckα exp

(2πi

z

hBR,−k,s

)Gs(qh). (2.4.2)

By the Jordan decomposition theorem, we can write BR,−k,s in the form

BR,−k,s = DR,−k,s +NR,−k,s,

where DR,−k,s is diagonalizable matrix, NR,−k,s is nilpotent matrix, and

DR,−k,sNR,−k,s = NR,−k,sDR,−k,s.

Therefore

exp(

2πiz

hBR,−k,s

)= exp

(2πi

z

hNR,−k,s

)exp

(2πi

z

hDR,−k,s

).

52

Page 58: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Let nR be the smallest integer such that NnRR,−k,s = 0. From the definition of

the exponential of a matrix, we see that the entries of the matrix

exp(

2πiz

hNR,−k,s

)are polynomials in z of degree less than nR.

Since DR,−k,s is diagonalizable matrix, we have

DR,−k,s = AR,−k,s diag (λ−k,s,1, . . . , λ−k,s,nR)A−1R,−k,e,

for some A = AR,k,e ∈ GL(n,C). By (2.4.2), we have

A−1F |kα−1 =

ckα

(A−1 exp

(2πi

z

hNR,−k,s

)A)

exp(

2πiz

hdiag(λ−k,s,1, . . . , λ−k,s,nR)

)A−1Gs(qh).

Set

Fs = A−1 F |kα−1, Hs = A−1Gs(qh), Q(z) = ckαA

−1 exp(2πiz

hNR,−k,s)A.

Then it is clear that ord0(Hs) = ord0(G), and that the entries of Q(z) are

polynomials in z. Also, we have

exp(

2πiz

hdiag(λ−k,s,1, . . . , λ−k,s,nR)

)=

diag(

exp(2πiλ−k,s,1z

h), . . . , exp(2πiλ−k,s,nR

z

h))

:=

diag(qλ−k,s,1h , . . . , q

λ−k,s,nRh

)With these notations, we get

Fs(z) = Q(z) diag(qλ−k,s,1h , . . . , q

λ−k,s,nRh

)Hs(qh).

We say that F vanishes at s if the right hand side of this equation vanishes

when z → i∞ (or qh → 0).

53

Page 59: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

We want to describe this condition in terms of the order of F at s, i.e., in

terms of ords(F ) = ord0(Gs) = ord0(Hs). Let

δR,−k,s = min<(λ−k,s,j), 0 ≤ j ≤ n.

Suppose that δR,−k,s > 0. Since the entries of Q(z) are polynomials in z, it

suffices to ask that Hs is holomorphic in qh (ord0(Hs) ≥ 0) in order to get

the vanishing of F at s. If δR,−k,s = 0, we need that ords(F ) = ord0(Hs) ≥ 1.

This leads to the following definition.

Definition 2.4.1. Let BR,k,s = Bk,s ∈M(n,C) such that

µ−ks (γs)Rγs = exp(2πiBk,s),

and

0 ≤ <(λk,s) < 1 ,

for all eigenvalues λk,s of Bk,s. Let

δR,k,s = min <(λk,s), λk,s is an eigenvalue of Bk,s .

Then

1. The number

ρk,s(R) =

0 if δR,k,s > 0

1 if δR,k,s = 0

is called the k−cuspidal order of R at s. Since it only depends on the

class of s modulo Γ, we define the k−cuspidal order of R at a cuspidal

point P ∈ X to be

ρk,P (R) = ρk,s(R),

where s ∈ H∗ is a cusp corresponding to P .

2. The k−cuspidal Divisor of R is defined by

DS,R,k =∑P∈S

ρk,P (R)P.

The associated line bundle |DS,R,k| to the divisor DS,R,k will be denoted

by DS,R,k.

54

Page 60: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

We see that an element F in Ak(Γ, R) vanishes at a cuspidal point s of Γ

if and only if νP (F ) ≥ ρ−k,s(R), in this case we say that F is cuspidal at s.

An element F in Mk(Γ, R) is called a cusp vector-valued automorphic form of

multiplier R and weight k, if it is cuspidal at all cusps of Γ. The vector space

of cusp vector-valued automorphic forms for Γ of multiplier R and weight

k will be denoted by Sk(Γ, R). We summarize the above discussion in the

following result

Proposition 2.4.2. In terms of divisors, Sk(Γ, R) can be written as

Sk(Γ, R) = F ∈Mk(Γ, R) | div(F ) ≥ DS,R,−k .

Now, suppose that χ is a character on Γ and that f ∈ Al(Γ, χ). If

F ∈ Ak(Γ, R), then f F ∈ A(k+l)(Γ, χR). In the following, we shall express

the divisor of fF in terms of div(f) and div(F ). Set R = χR. If P ∈ X

corresponds to a point of H, then it is clear that

νP (fF ) = νP (f) + νP (F ).

If P ∈ X corresponds to a cusp s, the above formula is no longer valid.

Indeed, using notations of Lemma 2.2.5, and Definition 2.2.2, we have

µks(γs)Rγs = exp(2πiBR,−k,s),

with eigenvalues λk,s of BR,k,s verifying 0 ≤ <(λk,s) < 1. Similarly

µls(γs)χγs = exp(2πibχ,−l,s),

with 0 ≤ <(bχ,l,s) < 1. This implies that

µ(k+l)s (γs)Rγs = exp(2πi(bχ,−l,sIn +BR,−k,s).

Since the eigenvalues of bχ,l,sIn +BR,k,s are the eigenvalues of BR,k,s plus the

scalar bχ,l,s, we see that

BR,−(k+l),s = (bχ,−l,sIn +BR,−k,s) + diag (−m1, . . . ,−mn),

55

Page 61: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

where mj ∈ 0, 1, 1 ≤ j ≤ n.

Using notations of Lemma 2.2.5, then in terms of the local parameter qh

at s, we have

c−kα exp(−2πiz

hBR,−k,s)F |kα−1 = Gs(qh),

c−lα exp(−2πiz

hbχ,−l,s) f |l(α−1) = gs(qh),

c−(k+l)α exp(−2πi

z

hBR,−(k+l),s) (fF )|

k+l(α−1) = Gs(qh),

and by definition, we have

νP (F ) = ord0(Gs), νP (f) = ord0(gs), νP (fF ) = ord0(Gs).

From this we see that

c−(k+l)α exp

(−2πi

z

h(bχ,−l,sIn +BR,−k,s)

)(fF )|

k+l(α−1) = gs(qh)Gs(qh),

and so

exp(−2πi

z

h(bχ,−l,sIn +BR,−k,s −BR,−(k+l),s)

)Gs(qh) = gs(qh)Gs(qh).

Hence

diag (qm1h , . . . , qmnh )Gs(qh) = gs(qh)Gs(qh).

Let gs,j (resp. gs,j), 1 ≤ j ≤ n, be the components of Gs (resp. Gs). Then

for some 1 ≤ j0, j1 ≤ n, we have

mj0 + ord0(Gs) = ord0(gs) + ord0(gs,j0),

mj1 + ord0(gs,j1) = ord0(gs) + ord0(Gs).

From the definition of the order, we deduce that

ord0(gs) + ord0(Gs) ≤ mj0 + ord0(Gs)

mj1 + ord0(Gs) ≤ ord0(gs) + ord0(Gs),

56

Page 62: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

which yield

mj1+ ≤ ( ord0(gs) + ord0(Gs)− ord0(Gs) ) ≤ mj0 ,

Hence

0 ≤ νP (f) + νP (F )− νP (fF ) ≤ 1.

Thus, we have proved

Theorem 2.4.3. Let χ be a character, and f be an element of Al(Γ, χ). If

F ∈ Ak(Γ, R), then f F ∈ A(k+l)(Γ, χR), and we have

div(fF ) = div(f) + div(F )−Df,F ,

where

Df,F =∑Q∈S

aQ,f,F Q

is a divisor having support in S with coefficients aQ,f,F ∈ 0, 1.

2.5 Mk(Γ, R) and Sk(Γ, R) as global sections of

vector bundles

We now come to the main construction of this chapter. We establish the

correspondence between vector-valued automorphic forms of weight k and

multiplier k for Γ, and global sections of the the holomorphic vector bundles

Ek = EΓ,R,k on the Riemann surface X.

In this section we keep the notations of §2.2.4 used to construct the vector

bundle Ek. Recall that the transition functions of Ek are the elements of the

1−cocycle (Fk,ij) ∈ Z1(U ,GL(n,O)), and that Fk,ij = π∗Fk,ij, with

Fk,ij = Ψk,iΨ−1k,j ∈ GL(n,O(Vi ∩ Vj)) if i 6= j and Fk,ii = id ,

where Ψk,i is the map constructed in Proposition 2.2.7 and Proposition 2.2.8.

Now, suppose that G is a collection (Gi)Vi of meromorphic functions Gi on

57

Page 63: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Vi satisfying

Gi = Fk,ijGj on Vi ∩ Vj. (2.5.1)

If G is invariant under Γ, that is

Giγ = Gi, for all γ ∈ Γ, (2.5.2)

then there exists a collection G = (Gi)Ui of meromorphic functions Gi on Ui

such that

Gi = π∗Gi on Vi.

Note that, by construction, Vi is Γ-invariant. Since Fk,ij = π∗Fk,ij, then, by

(2.5.1), we have

Gi = Fk,ijGj on Ui ∩ Uj.

Hence by Proposition 1.1.4, we deduce that G = (Gi)Ui is a meromorphic

section of Ek.

Conversely, if G = (Gi)Ui is a meromorphic section of Ek, then

G = π∗G = (π∗Gi)Vi

clearly satisfies (2.5.1) and (2.5.2). Thus we have proved the following

Proposition 2.5.1. There is a correspondence between the global meromor-

phic sections of Ek, and the collections (Gi)Vi of meromorphic functions Gi

on Vi satisfying

Gi = Fk,ijGj on Vi ∩ Vj,

and

Giγ = Gi, for all γ ∈ Γ.

Theorem 2.5.2. We have a linear isomorphism between A−k(Γ, R) and

H0(X,M(Ek)) given by

F 7→ ΘF = (Ψk,iF )Vi ,

58

Page 64: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

with inverse

Θ 7→ FΘ = (Ψ−1k,iπ

∗Θi)Vi ,

where Θ = (Θi)Ui .

Proof. By Proposition 2.5.1, the only nontrivial part is to prove that FΘ is

globally defined on H. For this, it suffices to show that

Ψ−1k,iπ

∗Θi = Ψ−1k,jπ

∗Θj on Vi ∩ Vj.

Since π∗Fk,ij = Ψk,iΨ−1k,j on Vi ∩ Vj, we have

Ψ−1k,iπ

∗Θi = Ψ−1k,iπ

∗(Fk,ijΘj)

= Ψ−1k,iπ

∗(Fk,ij)π∗(Θj)

= Ψ−1k,iFk,ijπ

∗Θj

= Ψ−1k,iΨk,iΨ

−1k,jπ

∗Θj

= Ψ−1k,jπ

∗Θj .

The next step is to find the relationship between the divisor of an element

F of A−k(Γ, R) and its corresponding element ΘF of H0(X,M(Ek)), see

(1.1.3) and (2.4.1). First, suppose that P ∈ X ′ corresponds to a point

z0 ∈ V0, then in a local chart ΘF is just Ψ0F . Since Ψ0 lies in GL(n,O(V0)),

then by Remark 1.1.5, we have νp(ΘF ) = νp(F ).

In case P corresponds to an elliptic point e of order ne, then using the no-

tations in the proof of Lemma 2.2.3 and letting α(z) = w, the local parameter

at P is t = wne . Letting ΘF = G(t), we have in terms of w

G(wne) = Ψe(α−1w)Ak,e F (α−1w) =

fke (α−1w) diag (w−m1 , . . . , w−mn)Ak,e F (α−1w),

which implies that

ne νP (ΘF ) = ord0(Fe), (2.5.3)

59

Page 65: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

where Fe is defined near w = 0 by

Fe(w) = diag (w−m1 , . . . , w−mn)Ak,eF (α−1w). (2.5.4)

If Fe = (Ak,eF )α−1, then it is clear that ord0(Fe) = orde(F ). Moreover,

if ni denotes the order of the ith component of Fe, 1 ≤ i ≤ n, then we have

ord0(Fe) = ni0 −mi0 ,

for some index i0, 1 ≤ i0 ≤ n . Since orde(F ) = ord0(Fe) = nj for some j,

then we have

ord0(Fe) ≤ nj −mj = orde(F )−mj.

In other words,

mj ≤ orde(F )− ord0(Fe).

Let me = maxmi, 1 ≤ i ≤ n,

F ∗(w) := diag (wme−m1 , . . . , wme−mn)(w−ord0(Fe)Fe(w)),

and

Fe(w) := w−ord0(Fe)Fe(w).

Then F ∗ is holomorphic in w, and by (2.5.4), we have

Fe(w) = w(ord0(Fe)−ord0(Fe)−me) F ∗(w).

If ord0(Fe) − ord0(Fe) − me > 0, then Fe(0) = 0 which contradicts the

definition of the order of Fe at 0, and so ord0(Fe)− ord0(Fe) ≤ me. Hence

mj ≤ orde(F )− ord0(Fe) ≤ me. (2.5.5)

We are led to the following definition.

Definition 2.5.1. The elliptic error τe(F ) of F at the elliptic point e ∈ H is

defined by

τe(F ) =orde(F )− ord0(Fe)

ne.

60

Page 66: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Notice that by (2.5.5), we have

0 ≤ τe(F ) ≤ 1− 1/ne.

Since τe(F ) depends only on the class of e modulo Γ, then:

1. If P ∈ X corresponds to an elliptic point e ∈ H, then we define

τP (F ) = τe(F ).

2. If Θ ∈ H0(X,M(Ek)), then we define the elliptic error of Θ at an

elliptic point P of X by

τ(Θ) = τ(FΘ),

where FΘ is the element of A−k(Γ, R) constructed in Theorem 2.5.2.

3. We set

τ(F ) = τ(ΘF ) =∑P∈E

τP (F )P =∑P∈E

τP (ΘF )P.

Using the fact that ne νP (F ) = orde(F ) and (2.5.3), we have

νP (ΘF ) = νP (F )− τP (F ). (2.5.6)

Finally, suppose that P corresponds to a cusp s. Since F has weight

−k, then from §2.3, we see that the behavior of F at s is by definition the

behavior of ΘF at P . Therefore

νP (ΘF ) = νP (F ). (2.5.7)

Combining (2.5.6) and (2.5.7), we get the following.

Theorem 2.5.3. Let F be an element of A−k(Γ, R), and ΘF be its corre-

sponding element in H0(X,M(Ek)). We have

div(ΘF ) = div(F )− τ(F ).

61

Page 67: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

At this point we need to give an explicit formula for the line bundle

Lk = EΓ,χ0,k for even integers k, where χ0 is the trivial character sending all

elements of Γ to 1, see Definition 2.2.4.

It is well-known that to each automorphic form f of even weight k is

associated a (k/2)−form η = f(dz)k/2 on X, see [55].

Proposition 2.5.4. [55] Let f be a nonzero automorphic form for Γ of even

weight k, and η = f(dz)k/2 be the associated (k/2)−form on X. Then

div(f) = div(η) + (k/2)

(∑P∈E

(1− 1/ne)P +∑Q∈S

Q

).

Definition 2.5.2. As in [55], we define the integral part of a rational divisor

D =∑P∈X

aP P , ap ∈ Q,

by

[D] =∑P∈X

[aP ]P,

where [x] is the integral part of x ∈ R.

We have the following useful lemma.

Lemma 2.5.5. [55] Let A =∑

P∈X aP P , ap ∈ Q, be a rational divisor on

X. Then for an integral divisor D on X, i.e., having coefficients in Z, we

have

D ≥ −A ⇐⇒ D ≥ −[A].

Remark 2.5.6. Notice that for Θ in H0(X,M(Ek)) and FΘ the correspond-

ing element of A−k(Γ, R), see Theorem 2.5.3, we have

div(Θ) = [div(FΘ).]

Recall that the line bundle associated to a divisor D on X is denoted by

|D|, see §1.1.2.

62

Page 68: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Proposition 2.5.7. Let KX be the canonical bundle of X,

DS =∑P∈S

P, D′E =∑P∈E

(1− 1/nP )P,

and DS = |DS|. Then for k even, we have

L−k = (k/2)KX + (k/2)DS + | [(k/2)D′E] |.

In particular

L−2 = KX +DS.

Proof. Let Θ be an element of H0(X,M(L−k)), and FΘ be the corresponding

element in Ak(Γ, R). By Proposition 2.5.7 and Remark 2.5.6, we have

div(Θ) = [div(FΘ)] = [ div(ηFΘ) + (k/2)

(∑P∈E

(1− 1/ne)P +∑Q∈S

Q

)],

that is

div(Θ) = div(ηFΘ) + (k/2)DS + [(k/2)D′E].

Hence

|div(Θ)| = |div(ηFΘ)| + (k/2)DS + | [(k/2)D′E] |.

Since L−k = |div(Θ)| and (k/2)KX = |div(ηFΘ)|, we have

L−k = (k/2)KX + (k/2)DS + | [(k/2)D′E] |,

as desired. When k = 2, the divisor (k/2)[D′E] = [D′E] is equal to the zero

divisor, and hence L−2 = KX +DS.

We now come to the main theorem of this chapter. Recall that the

(−k)−cuspidal divisor of R is given by DS,R,−k =∑

P∈S ρ−k,P (R)P , see

Definition 2.4.1.

63

Page 69: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Theorem 2.5.8. If DS,R,−k = |DS,R,−k| is the line bundle associated to

DS,R,−k, then we have

Mk(Γ, R) ∼= H0(X,O(E−k)),

and

Sk(Γ, R) ∼= H0(X,O(−DS,R,−k + E−k)) ∼=Θ ∈ H0(X,O(E−k)) | div(Θ) ≥ DS,R,−k

.

Proof. Let Θ be an element of H0(X,M(E−k)), and FΘ be its corresponding

element in Ak(Γ, R). By Theorem 2.5.3, we have

div(Θ) = div(FΘ)− τ(Θ).

According to Definition 2.5.1, the coefficients of τ(Θ) are in [0, 1), hence

[τ(Θ)] = 0. By Lemma 2.5.5, we have

div(FΘ) ≥ 0 ⇐⇒ div(Θ) ≥ −τ(Θ) ⇐⇒ div(Θ) ≥ −[τ(Θ)] = 0.

Similarly, since DS,R,−k is an integral divisor, we have

div(FΘ) ≥ DS,R,−k ⇐⇒ div(Θ)−DS,R,−k ≥ −τ(Θ) ⇐⇒

div(Θ)−DS,R,−k ≥ −[τ(Θ)] = 0 ⇐⇒ div(Θ) ≥ DS,r,−k .

Now, the result is a straightforward consequence of (1.1.5), Theorem 2.5.2,

Proposition 2.4.1, and Proposition 2.4.2.

As a first consequence we have

Theorem 2.5.9. Let Γ be a Fuchsian group. Then for any simple pair (R, k),

k ∈ Z, the dimensions of Mk(Γ, R) and Sk(Γ, R) are finite if and only if Γ is

of the first kind.

64

Page 70: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Proof. If Γ is of the first kind, then X is a compact Riemann surface.

From Theorem 1.1.12 and Theorem 2.5.8 we deduce that the dimensions

of Mk(Γ, R) and Sk(Γ, R) are finite.

If Γ is not of the first kind, then X is a noncompact Riemann surface.

This implies that any holomorphic vector bundle V on a X is trivial, see

[13]. Therefore H0(X,O(V)) is isomorphic to O(V)n, where n is the rank

of V . But the dimension of O(V) is infinite since any effective divisor is

the divisor of a holomorphic function on X, see [13]. Hence H0(X,O(V) is

an infinite-dimensional vector space. By the Theorem 2.5.8, Mk(Γ, R) and

Sk(Γ, R) can be realized as global holomorphic sections of some holomorphic

vector bundles on X, hence they have infinite dimensions.

From now on Γ will be a Fuchsian group of the first kind, and so X will

be a compact Riemann surface. We set

1. gX for the genus of X.

2. cX = |S| for the number of cuspidal points of X, which is finite since

Γ is of the first kind.

3. eX = |E| for the number of elliptic points of X, which is finite since Γ

is of the first kind.

4. dΓ,R,k = dk and sΓ,R,k = sk for the dimension of Mk(Γ, R) and Sk(Γ, R)

respectively.

Recall that the parity index of R is denoted by ε = ε(R), see Defini-

tion 2.1.3.

Theorem 2.5.10. Let DE =∑

P∈E P , and DE = |DE| be its associated

holomorphic line bundle. Then for any line bundle L on X, we have

h0(−DS + L+ L−(k−ε) + E−ε) ≤ h0(L+ E−k) ≤ h0(DE + L+ L−(k−ε) + E−ε).

65

Page 71: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Proof. Let L be a divisor on X such that L = |L|. From the definition of the

parity index, we see that the space Ak−ε(Γ) of automorphic forms of weight

k − ε for Γ is nontrivial. We fix a choice of a nonzero element f ∈ Ak−ε(Γ),

and we let α be its corresponding element of H0(X,M(L−(k−ε))). Then by

(1.1.5), H0(X,O(L+ E−k) is isomorphic toΘ ∈ H0(X,M(E−k)) | div(Θ) ≥ −L

,

which is, by Theorem 2.5.3 and Theorem 2.5.2, isomorphic to

F ∈ Ak(Γ, R) | div(F )− τ(F ) ≥ −L.

Since the multiplication by f gives an isomorphism

Aε(Γ, R) −→ Ak(Γ, R) : G 7→ f G,

we have

F ∈ Ak(Γ, R) | div(F )− τ(F ) ≥ −L ∼=

G ∈ Aε(Γ, R) | div(fG)− τ(fG) ≥ −L.

On the other hand, using Theorem 2.4.3, we have

div(fG) = div(f) + div(G)−Df,G,

where

Df,G =∑P∈S

aP,f,G P

is a divisor having support in S with coefficients aP,f,G ∈ 0, 1. HenceG ∈ Aε(Γ, R) | div(fG)− τ(fG) ≥ −L

=

G ∈ Aε(Γ, R) | div(f) + div(G)− τ(fG)−Df,G ≥ −L,

which is, by Theorem 2.5.3 and Theorem 2.5.2, isomorphic toΘ ∈ H0(X,M(E−ε)) | div(α)+τ(f)+div(Θ)+τ(GΘ)−τ(fGΘ)−Df,GΘ

≥ −L

66

Page 72: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

which is in turn equal toΘ ∈ H0(X,M(E−ε)) | div(α)+div(Θ)+L−Df,GΘ

≥ −(τ(f)+τ(Θ)−τ(fGΘ)).

By Lemma 2.5.5, the latter space is equal toΘ ∈ H0(X,M(E−ε)) | div(α)+div(Θ)+L−Df,GΘ

≥ −[τ(f)+τ(GΘ)−τ(fGΘ)]

which is equal toΘ ∈ H0(X,M(E−ε)) | div(α)+div(Θ)+L ≥ Df,GΘ

−[τ(f)+τ(GΘ)−τ(fGΘ)]

=: Vk,L .

According to Definition 2.5.1, the divisors τ(f), τ(GΘ), and τ(fGΘ) have

coefficients in [0, 1). Hence, the divisor τ(f) + τ(GΘ) − τ(fGΘ) has co-

efficients in (−1, 2 ). If Θ1 is the element of H0(X,M(E−k)) correspond-

ing to the element fGΘ of Ak(R,Γ), then from the relation div(fGΘ) =

div(f) + div(GΘ)−Df,GΘ, and Theorem 2.5.3 applied to GΘ, we have

div(Θ1) + τ(fGΘ) = div(α) + τ(f) + div(Θ) + τ(GΘ)−Df,GΘ,

which is equivalent to

div(Θ1)− div(α)− div(Θ) +Df,GΘ= τ(f) + τ(GΘ)− τ(fGΘ).

Therefore the coefficients of τ(f) + τ(GΘ)− τ(fGΘ) are all integers, and so

lie in 0, 1. Hence

0 ≤ [τ(α) + τ(Θ)− τ(fGΘ)] ≤∑P∈E

P = DE.

By Theorem 2.4.3, the divisor Df,GΘhas support in S with coefficients

in 0, 1, and so Df,GΘ≤ DS. Therefore, we have

Θ ∈ H0(X,M(E−ε)) | div(α) + div(Θ) + L ≥ DS

⊆ Vk,L,

67

Page 73: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

and

Vk,L ⊆

Θ ∈ H0(X,M(E−ε)) | div(α) + div(Θ) + L ≥ −DE

.

But (1.1.5) implies thatΘ ∈ H0(X,M(E−ε)) | div(α) + div(Θ) + L ≥ DS

∼=H0(X,O(−DS + L+ L−(k−ε) + E−ε)),

and Θ ∈ H0(X,M(E−ε)) | div(α) + div(Θ) + L ≥ −DE

∼=H0(X,O(DE + L+ L−(k−ε) + E−ε)).

By Theorem 2.5.8, we conclude that

h0(−DS + L+ L−(k−ε) + E−ε) ≤ h0(L+ E−k) ≤ h0(DE + L+ L−(k−ε) + E−ε).

Taking L = −DS,R,−k or the trivial line bundle, then by Theorem 2.5.8,

we get the following.

Corollary 2.5.11. We have

h0(−DS + L−(k−ε) + E−ε) ≤ dk ≤ h0(DE + L−(k−ε) + E−ε).

h0(−DS−DS,R,−k+L−(k−ε) +E−ε) ≤ sk ≤ h0(−DS,R,−k+DE+L−(k−ε) +E−ε).

The above two results will be of fundamental use in the next chapter

when computing the dimensions of Mk(Γ, R) and Sk(Γ, R).

68

Page 74: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Chapter 3

The dimension Formula

3.1 The degree of EΓ,R,k

Our main tool to compute the dimensions of Mk(Γ, R) and Sk(Γ, R) is the

Riemann-Roch theorem applied to the holomorphic vector bundles E−k =

EΓ,R,−k and −DS,R,−k + E−k, where DS,R,−k is the holomorphic line bundle

associated to the (−k)−cuspidal divisor of R, see Definition 2.4.1. Therefore

we need to compute the degree of E−k.

Let Θ1, . . . ,Θn be global meromorphic sections of the holomorphic vector

bundle E−k such that

Θ = Θ1 ∧ . . . ∧Θn (3.1.1)

is a nonzero section of the holomorphic line bundle∧n E−k (the determinant

bundle of E−k ). These always exist according to [50]. By Theorem 2.5.2,

each Θi corresponds to an element Fi of Ak(Γ, R). Hence F = F1 ∧ . . . ∧ Fnis an element of Ank(Γ, det(R)).

Recall that for an integer l, the space of automorphic forms of weight l

and trivial character for Γ is denoted by Al(Γ). Set χ = det(R), and let g

be a fixed element of A2nk(Γ). Then F1 = F 2/g lies in A0(Γ, χ2), and so

f = F ′1/F1 is a weight 2 automorphic form for Γ. Write η = fdz for its

69

Page 75: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

corresponding 1−form on X. To find the degree of E−k, we follow the same

strategy used in [43]. First, we express the residues of η in terms of the orders

of Θ and g, then use the fact that the sum of residues of η is 0.

Lemma 3.1.1. Let f be a weight 2 automorphic form, and η be its associated

1−form on X. Suppose that P ∈ X corresponds to x ∈ H∗, then

1. If x ∈ H, then ResP (η) = (1/nP )Resx(f), where nP is the order of P .

2. If x is a cusp s, and ∑n≥ns

an(f) qnh

is the expansion of f at s, qh being the local parameter at s, and h the

cusp width at s, then we have

(2πi/h)ResP (η) = a0(f).

Proof. Suppose that x ∈ H. If

α(z) =z − xz − x

, z ∈ H,

then α(H) is the unit disc D, α(e) = 0. With α(z) = w, the local parameter

at P is t = wnx , where nx = nP is the order of P . Set β := α−1. By

definition, Resx(f) is the residue of the 1−form f(z)dz on H. Writing

β∗(f(z)dz) = h(w)dw

and using the fact that the residue of a differential form is invariant under

changes of local parameter, we deduce that

Resx(f) = Res0(h). (3.1.2)

if η = h1(t)dt, then according to [55], we have

h(w) = nxwnx−1 h1(wnx). (3.1.3)

70

Page 76: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Suppose that the t−expansion of h1 near t = 0 is given by

h1(t) =∑n≥n0

cn tn,

where n0 = ord0(h1), then

h1(wnx) =∑n≥n0

cnwnnx .

Using (3.1.3), we find

Res0(h) = nx Res0(h1) = nx ResP (η),

and from (3.1.2), we conclude that

ResP (η) = (1/nP )Resx(f).

As for 2., if x is a cusp s ∈ H∗ with cusp width h, then if we take

α =

(a b

c d

)∈ SL(2,R) such that α · s = ∞, the local parameter at P is

given by qh = exp(2πiz/h). Suppose that η = F (qh)dqh, then according to

[55], we have

(2π/h)F (qh) = q−1h

∑n≥ns

an(f) qnh . (3.1.4)

Thus

(2π/h) Res0(F ) = a0(f).

Since Res0(F ) = ResP (η), we conclude that

(2πi/h)ResP (η) = a0(f).

For non-cuspidal points of X, we have the following

Lemma 3.1.2. If P ∈ X ′, then

ResP (η) =

2 νP (Θ) − 2 ν−k,P (R) − νP (g) if P is an elliptic point

2 νP (Θ) − νP (g) otherwise.

71

Page 77: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Proof. Let x ∈ H be above P . Then

ResP (η) = Resx(f) = νx(F1) = 2νx(F )− νx(g) = 2νP (F )− νP (g),

and so it suffices to express νP (F ) in terms of νP (Θ). Suppose that x is an

elliptic fixed point e of Γ. Using the notations in the proof of Lemma 2.2.3

and letting α(z) = w, the local parameter at P is t = wne , where ne = nP is

the order of P . If Θi = Gi(t), then by the correspondence in Theorem 2.5.2,

we have in terms of w

Gi(wne) = Ψ−k,e(α

−1w)A−k,e Fi(α−1w)

= f−ke (α−1w) diag (w−m−k,1 , . . . , w−m−k,n)A−k,e Fi(α−1w).

If Θ = G(t), then G = G1 ∧ . . . ∧Gn and

G(wne) = f−nke (α−1w)w−(∑nj=1m−k,j) det(A−k,e)F (α−1w).

Hence

νP (Θ) = νP (F )− (1/ne)n∑j=1

m−k,j = νP (F ) + ν−k,P (R).

In case x is not an elliptic point, then in a local chart Θi is given by

Ψ−k,0Fi, and so Θ = det(Ψ−k,0)F . Since Ψ−k,0 ∈ GL(n,O(Y ′)), we have

νp(Θ) = νp(F ).

As for the cusps, we have:

Lemma 3.1.3. Suppose that P ∈ X corresponds to a cusp s ∈ H∗, then

ResP (η) = 2 νP (Θ) − 2 ν−k,P (R) − νP (g).

Proof. Let α =

(a b

c d

)∈ SL(2,R) such that α · s = ∞, then the local

parameter at P is given by qh = exp(2πiz/h), h being the cusp width at

72

Page 78: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

s. Suppose that Θi = Gi(qh) and Θ = G(t), then using notations from the

proof of Lemma 2.2.5, we have

Gi(qh) = f−ks (α−1z) exp(−2πi

z

hB−k,s

)Fi(α

−1z).

This, together with the fact that for a square matrix A,

det( exp(A) ) = exp( trace(A) ),

gives

G(qh) = f−nks (α−1z) exp(−2πi

z

htrace(B−k,s)

)F (α−1z) = .

G(qh) = f−nks (α−1z) qν−k,P (R)

h F (α−1z),

As −trace(B−k,s) is ν−k,P (R).

Since s = α−1 · ∞ = −d/c, we have fs(α−1z) = cα Jα−1(z) for some

nonzero constant cα ∈ C. Hence

G(qh) = cα qν−k,P (R)

h F |nk

(α−1).

Applying the logarithmic derivative on both sides we get

2πi qhh

(∂qhG)(qh)

G(qh)=

2πi ν−k,P (R)

h+

(F |nk

(α−1) )′

F |nk

(α−1),

and since f = 2F ′/F − g′/g, we get

f |2(α−1) = 2(F |

nk(α−1) )

F |nk

(α−1)− ( g|

2nk(α−1) )

g|2nk

(α−1).

Therefore

f |2(α−1) =4πi qhh

(∂qhG)(qh)

G(qh)− 4πi ν−k,P (R)

h− ( g|

nk(α−1) )

g|nk

(α−1).

Comparing the constant terms of the qh−expansions and using Lemma 3.1.1,

we have

(2πi/h)ResP (η) = (4πi/h)νP (Θ) − (4πi/h)ν−k,P (R) − (2πi/h)νP (g)

and hence the formula.

73

Page 79: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Combining the preceding two lemmas with the fact that∑p∈X

ResP (η) = 0,

as X is compact, we get

div(Θ) = DR,−k + (1/2) div(g),

where DR,−k is the (−k)−divisor of R, see Definition 2.2.3. If ηg is the

(nk)−form associated to g, then by Proposition 2.5.4 we have

div(Θ) = DR,−k + (1/2) div(ηg) + (nk/2)(D′E +DS).

Hence

deg(E−k) = deg( div(Θ) ) =

deg(DR,−k) + nk(gX − 1) + (nk/2)( deg(D′E) + deg(DS) ),

where gX is the genus of the compact Riemann surface X. We summarize

the above in the following:

Theorem 3.1.4. Let cX = |S| be the number of cuspidal points of X, gX be

the genus of X. Then the degree of E−k is given by

deg(E−k) =nk

2

(2gX − 2 + cX +

∑P∈E

(1− 1/nP )

)+ deg(DR,−k).

Remark 3.1.5. The number mX defined by

mX = 2gX − 2 + cX +∑P∈E

(1− 1/nP )

is positive, see [55], and 2πmX is called the hyperbolic area of X. Also, it

can be shown that

mX ≥ 1/42,

see [55].

74

Page 80: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

According to Definition 2.2.3 the k−divisor of R is given by

DR,k =∑P∈X

νk,P (R)P.

Set

DR,k =∑P∈X

<(νk,P (R))P. (3.1.5)

The above theorem implies that the degree of DR,k is a rational number.

Hence

deg(DR,k) = deg(DR,k). (3.1.6)

But from Definition 2.2.1, and Definition 2.2.2, we see that

−n < <(νk,P (R)) ≤ 0,

for all P ∈ E ∪S, and since the support of DR,k is by definition in E ∪S,

we conclude that

−n (cX + eX) < deg(DR,k) ≤ 0.

Now, using (3.1.6), we find

Proposition 3.1.6. Let DR,k be the k−divisor of R, we have

−n (cX + eX) < deg(DR,k) ≤ 0.

3.2 The holomorphic degree of a vector bun-

dle.

Recall that the dimensions of Mk(Γ, R) and Sk(Γ, R) are respectively denoted

by

75

Page 81: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

dk = dΓ,R,k and sk = sΓ,R,k. Applying the Riemann-Roch theorem to the

holomorphic vector bundles E−k and −DS,R,−k + E−k yields

h0(E−k)− h0(KX + E∗−k) = deg(E−k)− n(gX − 1),

h0(−DS,R,−k + E−k)− h0(KX +DS,R,−k + E∗−k) =

deg(−DS,R,−k + E−k)− n(gX − 1),

where E∗−k is the dual holomorphic vector bundle of E−k. Combining this

with Theorem 2.5.8 gives:

dk = deg(E−k)− n(gX − 1) + h0(KX + E∗−k),

sk = deg(−DS,R,−k + E−k)− n(g − 1) + h0(KX +DS,R,−k + E∗−k).

Therefore we need to compute h0(KX + E∗−k) and h0(KX +DS,R,−k + E∗−k).

The key point in computing dk and sk in the case of automorphic forms

(i.e., R is the trivial character χ0) relies on the vanishing of these two di-

mensions, which is in fact equivalent to the non-existence of holomorphic au-

tomorphic forms of negative weight. In higher dimensions, this is no longer

valid since for some representations vector-valued automorphic forms of neg-

ative weight do exist, see (4.2.7). To handle this problem we shall introduce

the notion of the holomorphic degree of a vector bundle, which associates to

each holomorphic vector bundle E , defined over a compact Riemann surface,

an integer d(E) in a such way that:

d(E) < 0 =⇒ H0(X,O(E)) is trivial.

We will give some basic properties of the holomorphic degree, and some useful

bounds of d(E) that will serve in the next section.

For this section X will be an arbitrary compact Riemann surface of genus

g.

76

Page 82: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Proposition 3.2.1. Let E be a vector bundle over a compact Riemann sur-

face X. There exists a constant CE ∈ Z such that

deg(div(Θ)) ≤ CE for all Θ ∈ H0(X,M(E)).

Proof. If no such constant exists, then we will have a sequence

(Θn)n∈N ∈ H0(X,M(E))

such that deg(div(Θn)) > n. Let Dn = div(Θn), and Dn = |Dn| be its

associated holomorphic line bundle. By the Riemann-Roch theorem, we have

h0(Dn) = deg(Dn) + h0(KX −Dn)− (g − 1).

If we take n > deg(KX) = 2(g − 1), then h0(KX −Dn) = 0, and so

h0(Dn) = deg(Dn)− (g − 1).

Hence

h0(Dn) > n− (g − 1). (3.2.1)

Let L(Dn) = f ∈ M(X) | div(f) + Dn ≥ 0. Then the injective homomor-

phism

L(Dn) −→ H0(X,M(E)) : f 7→ fΘn

has its image in H0(X,O(E)). Indeed, if f ∈ L(Dn), then

div(fΘn) = div(f) + div(Θn) = div(f) +Dn ≥ 0.

Hence L(Dn) imbeds in H0(X,O(E)). Now, using the standard fact that

L(Dn) is isomorphic to H0(X,O(Dn)), we deduce that H0(X,O(Dn)) imbeds

in H0(X,O(E)). Therefore h0(E) ≥ h0(Dn). If n > 2(g − 1), then according

to (3.2.1), we have

h0(E) > n− (g − 1),

Thus h0(E) =∞, which contradicts Theorem 1.1.12.

As a consequence, maxdeg(div(Θ)); Θ ∈ H0(X,M(E)) is finite.

77

Page 83: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Definition 3.2.1. Let E be a holomorphic vector bundle over a compact

Riemann surface X, then the holomorphic degree d(E) of E is defined by

d(E) = maxdeg(div(Θ)) : Θ ∈ H0(X,M(E)).

Here are some basic properties of the holomorphic degree:

Proposition 3.2.2. Let X be a compact Riemann surface, E a holomorphic

vector bundle over X and L a holomorphic line bundle over X. Then we

have

1. d(L) = deg(L).

2. d(L+ E) = d(L) + d(E).

3. If d(E) < 0, then h0(E) = 0.

Proof. By definition of d(L), there exits an element α ∈ H0(X,M(L)) such

that

d(L) = deg(div(α)) = deg(L)

as L is a line bundle, which proves 1.

To prove 2., take α ∈ H0(X,M(L)) and Θ ∈ H0(X,M(E)) such that

deg(div(α)) = d(L), deg(div(Θ)) = d(E),

then αΘ ∈ H0(X,M(L+ E)) and we have

deg(div(αΘ)) = deg(div(α)) + deg(div(Θ)) = d(L) + d(E).

Let Θ1 ∈ H0(X,M(L+ E)). Then Θ1/α ∈ H0(X,M(E)) and we have

deg(div(Θ1)) = deg(div(α)) + deg(div(Θ1/α)) ≤

deg(div(α)) + deg(div(Θ)) = d(L) + d(E).

78

Page 84: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Therefore

d(L+ E) ≤ d(L) + d(E) = deg(div(αΘ)) ≤ d(L+ E).

Hence d(L+ E) = d(L) + d(E).

As for 3., suppose that d(E) < 0. If there exists a nonzero element

Θ ∈ H0(X,O(E)), then we will have

0 ≤ deg(div(Θ) ≤ d(E),

which is impossible. Hence h0(E) = 0.

We have the following upper bound of d(E):

Theorem 3.2.3. Let E be a holomorphic vector bundle over X of holomor-

phic degree d(E). We have

d(E) ≤ h0(E) + g − 1.

Proof. Take an element Θ of H0(X,M(E)) such that deg(div(Θ)) = d(E),

and set D = div(Θ) and D = |D|. As in the proof of Proposition 3.2.1, we

have an embedding of H0(X,O(D)) in H0(X,O(E)). Therefore

h0(D) ≤ h0(E).

By the Riemann-Roch theorem, we have

h0(D) = deg(D) + h0(KX −D)− (g − 1),

which implies that

deg(D)− (g − 1) ≤ h0(D).

Since deg(D) = deg(D) = d(E), we have

d(E)− (g − 1) ≤ h0(D) ≤ h0(E),

hence

d(E) ≤ h0(E) + g − 1.

79

Page 85: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Now we will give a lower bound for d(E). Applying the Riemann-Roch

theorem to the holomorphic vector bundle E , we get

h0(E) = h0(KX + E∗) + deg(E)− n(g − 1),

where n is the rank of E . Thus

deg(E)− n(g − 1) ≤ h0(E). (3.2.2)

Let P be any point of X, and L = |P | be the line bundle associated to the

divisor P . If E1 is defined by

E1 = −(d(E) + 1)L+ E ,

then we have

d(E1) = −(d(E) + 1)d(L) + d(E) = −(d(E) + 1) + d(E) = −1.

Hence, by Proposition 3.2.2, we have h0(E1) = 0. By (3.2.2) applied to E1,

we have

deg(E1)− n(g − 1) ≤ h0(E1) = 0.

Using Proposition 1.1.9, we get

−n(d(E) + 1) deg(L) + deg(E)− n(g − 1) ≤ 0,

and since deg(L) = 1, we conclude that

(1/n) deg(E)− g ≤ d(E).

Thus we have proved:

Theorem 3.2.4. Let E be a rank n holomorphic vector bundle over X of

holomorphic degree d(E). We have

deg(E)

n− g ≤ d(E).

80

Page 86: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

3.3 The dimension of Mk(Γ, R) and Sk(Γ, R)

We now come to one of the main results of this thesis. The goal is to give

formulas for the dimensions of Mk(Γ, R) and Sk(Γ, R). We will show that

there are some constants k+R,X and k−R,X such that:

1. For k < k−R,X , Mk(Γ, R) will be trivial.

2. For k > k+R,X , we will have an explicit formula for the dimensions of

Mk(Γ, R) and Sk(Γ, R) expressed in terms of some invariants of R and

X.

Recall that

DE =∑P∈E

P, DS =∑P∈S

P,

DR,−k is the (−k)−divisor of R, and that DS,R,−k is the (−k)−cuspidal

divisor of R, see Definition 2.2.3 and Definition 2.4.1. Their associated line

bundles are respectively denoted by DE, DS, DR,−k, and DS,R,−k. Also, recall

that

mX = 2gX − 2 + cX +∑P∈E

(1− 1/nP ),

and that the parity index of R is denoted by ε = ε(R) according to Defini-

tion 2.1.3.

Applying the Riemann-Roch theorem to the holomorphic vector bundles

E−k and −DS,R,−k + E−k, we get:

h0(E−k)− h0(KX + E∗−k) = deg(E−k)− n(gX − 1),

h0(−DS,R,−k+E−k)−h0(KX+DS,R,−k+E∗−k) = deg(−DS,R,−k+E−k)−n(gX−1).

Combining this with Theorem 2.5.8 gives:

dk = deg(E−k)− n(gX − 1) + h0(KX + E∗−k),

sk = deg(−DS,R,−k + E−k)− n(gX − 1) + h0(KX +DS,R,−k + E∗−k),

81

Page 87: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

where E∗−k is the dual holomorphic vector bundle of E−k. We have the fol-

lowing formula for E∗−k.

Lemma 3.3.1. If R∗ is the adjoint representation of R, then

E∗R,−k = DE +DS + ER∗,k .

Proof. This is a direct consequence of the Remarks (2.2.2, 2.2.4, 2.2.6), and

Theorem 1.1.3. The only nontrivial part is the contribution of the line bun-

dles DE = |DE| and DS = |DS|. Using notations of Remark 2.2.4 and

Remark 2.2.6, this can be justified by the fact that the local parameter at an

elliptic (resp. cuspidal) point P ∈ X corresponding to e ∈ H (resp. s ∈ H∗ )

is t = wne (resp. exp(2πiz/h) ), and that the section defining DE (resp. DS)

is given near P by the function t (resp. qh ), see [19].

As a consequence we have:

Proposition 3.3.2. We have

dk = deg(E−k)− n(gX − 1) + h0(KX +DE +DS + ER∗,k),

and

sk = deg(−DS,R,−k +E−k)−n(gX−1)+h0(KX +DE +DS +DS,R,−k +ER∗,k).

We will show that for k large enough, the numbers h0(KX +DE +DS +

ER∗,k) and h0(KX +DE +DS +DS,R,−k + ER∗,k) vanish, thus providing exact

formulas for the dimensions dk and sk. We only treat the first case, the

second one follows in a similar way. Using Theorem 2.5.10, we have

h0(KX +DE +DS + ER∗,k) ≤ h0(KX + 2DE +DS + L(k+ε) + ER∗,−ε).

Hence, by Proposition 3.2.2, it suffices to show that

d(KX + 2DE +DS + L(k+ε) + ER∗,−ε) < 0,

82

Page 88: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

which is equivalent to

d(KX) + 2 d(DE) + d(DS) + d(L(k+ε)) + d(ER∗,−ε) < 0,

that is

(2gX − 2 + 2eX + cX) + d(L(k+ε)) + d(ER∗,−ε) < 0. (3.3.1)

By Theorem 3.1.4, the degree of L(k+ε) is given by

d(L(k+ε)) = − (k + ε)mX

2+ deg(Dχ0,k+ε),

where χ0 is the trivial character sending all elements of Γ to 1, see Defini-

tion 2.2.4. By definition, the support of the divisor

Dχ0,k+ε =∑P∈X

νk+ε,P (χ0)P

lies in E ∪S, with coefficients in (−1, 0]. Therefore

deg(Dχ0,k+ε) ≤ 0.

Hence, to get (3.3.1), it suffices to have

d(ER∗,−ε) + (2gX − 2 + cX + 2 eX) − (k + ε)mX

2< 0 ,

or equivalently,

2

mX

(d(ER∗,−ε) + 2gX − 2 + cX + 2 eX)− ε < k.

Now, using Theorem 3.1.4 and Proposition 3.3.2, we get

Theorem 3.3.3. Let Γ be a Fuchsian group of the first kind, and X be the

associated compact Riemann surface. If

k+X,R =

2

mX

( d(ER∗,−ε) + 2gX − 2 + cX + 2 eX )− ε,

then

83

Page 89: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

1. For k > k+R,X , the dimension of Mk(Γ, R) is given by

dk =n kmX

2− n (gX − 1) + deg(DR,−k).

2. For k > k+X,R + (2/mX) deg(DS,R,−k) , the dimension of Sk(Γ, R) is

given by

sk = dk − n deg(DS,R,−k).

If we relax the lower bound of k, we get the following:

Corollary 3.3.4. For

k ≥ 84(h0(ER∗,−ε) + gX + cX + 2 eX

),

we have

1. The dimension of Mk(Γ, R) is given by

dk =n kmX

2− n (gX − 1) + deg(DR,−k).

2. The dimension of Sk(Γ, R) is given by

sk = dk − n deg(DS,R,−k).

Proof. By Remark 3.1.5, we have

( 2/mX ) ≤ (2) 42 = 84.

Also, by Theorem 3.2.3, we have

d(ER∗,−ε) ≤ h0(ER∗,−ε) + (gX − 1).

By definition, DS,R,−k ≤ DS, and so deg(DS,R,−k) ≤ deg(DS) = cX . Com-

bining these inequalities with the fact that (2gX − 2 + cX)/mX ≤ 1, and

ε ∈ 0, 1, we have

k+X,R ≤ k+

X,R + (2/mX) deg(DS,R,−k) ≤ k+X,R + (2/mX)cX =

84

Page 90: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

(2/mX) (d(ER∗,−ε) + 2gX − 2 + 2cX + 2 eX)− ε ≤

(2/mX)(h0(ER∗,−ε) + (gX − 1) + cX + 2 eX

)+ 2/(nmX)(2gX−2 + cX) ≤

84(h0(ER∗,−ε) + (gX − 1) + cX + 2 eX

)+ (2/n) ≤

84(h0(ER∗,−ε) + gX + cX + 2 eX

).

We conclude using Theorem 3.3.3.

Now, recall from Proposition 3.1.6 that

−n (cX + eX) < deg(DR,k) ≤ 0,

so that | deg(DR,k)| < n (cX + eX). Since 0 ≤ deg(DS,R,−k) ≤ cX , then by

Corollary 4.1.3, we have

Corollary 3.3.5. Let Γ be a Fuchsian group of the first kind. The dimensions

dk of Mk(Γ, R), and sk of Sk(Γ, R) are asymptotically equivalent to k. More

precisely, we havedkk−→ nmX

2, as k →∞,

andskk−→ nmX

2, as k →∞.

Combining this result with Theorem 2.5.9, we have

Theorem 3.3.6. Holomorphic, cusp vector-valued automorphic forms exist

for any Fuchsian group Γ and any n−dimensional representation R of Γ.

Using the upper bounds of dk and sk from Corollary 2.5.11, we have

Theorem 3.3.7. Let Γ be a Fuchsian group of the first kind, X be the asso-

ciated compact Riemann surface. If

k−X,R =−2

mX

( d(ER,−ε) + eX )− ε,

then

85

Page 91: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

1. For k < k−R,X , the dimension dk of Mk(Γ, R) is zero.

2. For k < k−X,R + (2/mX) deg(DS,R,−k) , the dimension sk of Sk(Γ, R)

is zero.

Proof. By Corollary 2.5.11, we have

dk ≤ h0(DE + L−(k−ε) + E−ε),

sk ≤ h0(−DS,R,−k +DE + L−(k−ε) + E−ε).

Hence it suffices to show that

d(DE + L−(k−ε) + E−ε) < 0, (3.3.2)

d(−DS,R,−k +DE + L−(k−ε) + E−ε) < 0. (3.3.3)

We only treat the first case. As in the proof of the Theorem 3.3.3, we

have

d(DE + L−(k−ε) + ER,−ε) =

d(DE) + d(L−(k−ε)) + d(ER,−ε) ≤

eX +(k + ε)mX

2+ d(E−ε).

To get (3.3.2), it is enough to have

eX +(k + ε)mX

2+ d(E−ε) < 0,

that is

k <−2

mX

( d(ER,−ε) + eX )− ε = k−X,R .

As in Corollary 4.1.3, we have the following simpler upper bound for k:

Corollary 3.3.8. If

k < − 84(h0(ER,−ε) + (gX − 1) + eX

)− 1,

then Mk(Γ, R) and Sk(Γ, R) are trivial.

86

Page 92: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Remark 3.3.9. Notice that for values of k in the interval [k−X,R , k+X,R] we

have:

dk ≥ deg(E−k)− n(gX − 1),

and

sk ≥ deg(−DS,R,−k + E−k)− n(gX − 1).

This a direct consequence of Proposition 3.3.2.

3.4 Finite image representations

Due to its richness and simplicity, the case of finite image representations

has been treated by many authors, see [3, 15]. Our aim in this section is to

provide bounds for the weight k simpler than k+X,R and k−X,R . For example, we

will see that k−X,R can be replaced by 0. For the sake of clarity and to avoid

technical details, we restrict ourselves to the case of even representations,

that is ε = ε(R) = 0.

Let

ΓR := ker(R), ΓR := ±1\ΓR.±1, XR := ΓR\H∗.

Since the image of R is finite, ΓR has finite index in Γ. Therefore, ΓR is a

Fuchsian group of the first kind, and XR is a compact Riemann surface, see

Proposition 1.2.9. Suppose that F ∈ Ak(Γ, R), by definition we have

F |kγ = Rγ F, γ ∈ Γ ,

and so for γ ∈ ΓR we get

F |kγ = F.

Hence all components of F are weight k automorphic forms for ΓR, that

is F ∈ (Ak(ΓR) )n. As ΓR has finite index in Γ, Γ and ΓR have the same

cusps, and the cuspidal behavior of F is the same with respect to Γ and ΓR.

Therefore we have the following two embedding

Mk(Γ, R) −→ (Mk(ΓR) )n : F −→ F, (3.4.1)

87

Page 93: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Sk(Γ, R) −→ (Sk(ΓR) )n : F −→ F. (3.4.2)

Since Mk(ΓR) is trivial for k < 0, we deduce that Mk(Γ, R) and Sk(Γ, R)

are both trivial for k < 0.

The next step is to simplify the bound

k+X,R =

2

mX

(d(ER∗,−ε) + 2gX − 2 + cX + 2 eX)− ε

given by Theorem 3.3.3 (here ε = 0). Suppose that d(ER∗,−ε) − (gX−1) > 0,

and take Θ1 ∈ H0(X,M(ER∗,0)) such that deg(div(Θ1)) = d(ER∗,0). Set

D = div(Θ1) and D = |D| its associated holomorphic line bundle. By the

Riemann-Roch theorem, we have

h0(D) = deg(D) + h0(KX −D)− (gX − 1) ≥ d(ER∗,−ε) − (gX − 1) > 0.

Hence H0(X,O(D)) is nontrivial. Let δ be an element of H0(X,M(D))

such that div(δ) = D, and take σ ∈ H0(X,O(D)). Then f = σ/δ is a

meromorphic function on X satisfying

div(f) + D = div(σ) ≥ 0.

Therefore,

div(f Θ1) = div(f) + D ≥ 0,

and hence Θ = f Θ1 ∈ H0(X,O(ER∗,0)). Let F be the element of M0(Γ, R∗)

corresponding to Θ. Since the elements of M0(ΓR) correspond to holomorphic

functions on the compact Riemann surface XR, we have M0(ΓR∗) ⊆ Cn.

By the embedding of M0(Γ, R∗) in M0(Γ∗R), see (3.4.1), we deduce that Θ

is a nonzero vector in Cn. Therefore, div(F ) is the trivial divisor. Since

[div(F )] = div(Θ) and deg( div(f) ) = 0, we have

0 = deg( div(Θ) ) = deg( div(Θ1) ) = deg(D) = d(ER∗,0).

We conclude that if d(ER∗,−ε) − (gX − 1) > 0, then d(ER∗,0) = 0, and so

gX = 0. Hence, in all cases

d(ER∗,0) ≤ gX .

88

Page 94: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Using the fact that mX ≥ 2gX − 2 + cX , we see that

k+X,R ≤ 2 +

2

mX

(gX + 2 eX) .

Now, by Theorem 3.3.3 we conclude that

Theorem 3.4.1. Let Γ be a Fuchsian group of the first kind, X be the associ-

ated compact Riemann surface, and R be an even finite image representation.

If

fX,R = 2 +2

mX

(gX + 2 eX) ,

then

1. For k > fR,X , the dimension of Mk(Γ, R) is given by

dk =n kmX

2− n (gX − 1) + deg(DR,−k).

2. For k > fX,R + (2/mX) deg(DS,R,−k) , the dimension of Sk(Γ, R) is

given by

sk = dk − n deg(DS,R,−k).

3. For k < 0, Mk(Γ, R) and Sk(Γ, R) are both trivial.

As we have seen in the discussion preceding Theorem 3.4.1, M0(Γ, R) is a

subset of Cn. Combining this with the definition of vector valued automor-

phic forms of weight 0 and multiplier R, and the fact that the elements of

S0(Γ, R) ⊆M0(Γ, R) ⊆ Cn vanish at the cusps, we have

Corollary 3.4.2. Let Γ be a Fuchsian group of the first kind, X be the associ-

ated compact Riemann surface, and R be an even finite image representation.

Then S0(Γ, R) is trivial, and we have

M0(Γ, R) = v ∈ Cn |Rγ(v) = v for all γ ∈ Γ.

89

Page 95: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Chapter 4

Applications

4.1 Generalized automorphic forms

Since the work of M. Knopp and G. Mason in [28], the theory of gen-

eralized automorphic forms has been receiving an increasing interest, see

[25, 26, 29, 32, 33, 44, 45]. We propose here to compute the dimensions of

the vector spaces of generalized automorphic forms. In this section R will be

a 1−dimensional representation (a character) and will be denoted by χ. The

holomorphic vector bundle EΓ,χ,k is then a line bundle, and will be denoted

by IΓ,χ,k = Ik.

The generalized automorphic forms look like classical modular forms with

a multiplier system except for the fact that the character χ need not be

unitary. They can be defined as 1−dimensional vector valued automorphic

forms. More precisely,

Definition 4.1.1. Let (χ, k), k ∈ Z, be a simple pair. Then

1. An element of Ak(Γ, χ) is called a generalized automorphic form for Γ

of multiplier χ and weight k.

2. An element of Gk(Γ, χ) (resp. Mk(Γ, χ), Sk(Γ, χ) ) is called an unre-

90

Page 96: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

stricted (resp. holomorphic, cusp) generalized automorphic form for Γ

of multiplier χ and weight k.

Recall that the dimensions of Mk(Γ, χ) and Sk(Γ, χ) are denoted by

dΓ,χ,k = dk and sΓ,χ,k = sk respectively. Also, according to Definition 2.4.1,

DS,χ,−k denotes the holomorphic line bundle associated to the k−cuspidal

divisor of χ

DS,χ,k =∑P∈S

ρk,P (χ)P.

To compute dk and sk, we follow the same steps as in §3.3. Indeed, applying

the Riemann-Roch theorem to I−k and −DS,χ,−k + I−k yields

h0(I−k)− h0(KX + I∗−k) = deg(I−k)− (gX − 1),

h0(−DS,χ,−k+I−k)−h0(KX+DS,χ,−k+I∗−k) = deg(−DS,χ,−k+I−k)−(gX−1),

where I∗−k is the dual holomorphic line bundle of E−k. Combining this with

Theorem 2.5.8, and the fact that I∗−k = −I−k gives:

dk = deg(I−k)− (g − 1) + h0(KX − I−k),

sk = deg(−DS,R,−k + I−k)− (g − 1) + h0(KX +DS,R,−k − I−k).

We will show that for k large enough, the numbers h0(KX − I−k) and

h0(KX +DS,χ,−k−I−k) vanish. Thus we find an exact formula for the dimen-

sions dk and sk. As usual, we only treat the first case. By Proposition 3.2.2,

h0(KX − I−k) is zero if

d(KX − I−k) = d(KX)− d(I−k) < 0,

that is

2(gX − 1)− d(I−k) < 0.

By Theorem 3.1.4, the degree of I−k is given by

d(I−k) = (k/2)mX + deg(Dχ,−k).

91

Page 97: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Hence, d(KX − I−k) < 0 if and only if

2(gX − 1) − (k/2)mX − deg(Dχ,−k) < 0.

By definition, the support of the divisor

Dχ,−k =∑P∈X

ν−k,P (χ)P

lies in E ∪S, with coefficients in (−1, 0], and so

− deg(Dχ,−k) ≤ cX + eX .

Therefore, to get d(KX − I−k) < 0 if suffices to have

2(gX − 1) − (k/2)mX + cX + eX < 0,

in other words,

(2/mX) ( 2(gX − 1) + cX + eX) < k.

But

(2/mX) ( 2(gX − 1) + cX + eX) = mX +∑P∈E

1

nP.

Thus we conclude that h0(KX − I−k) = 0 if

2 +2

mX

∑P∈E

1

nP< k.

Using Theorem 3.1.4 and Proposition 3.3.2, we have

Theorem 4.1.1. Let Γ be a Fuchsian group of the first kind, X be its compact

Riemann surface, and let

mX = 2gX − 2 + cX +∑P∈E

(1− 1/nP ),

l+X,χ = 2 +2

mX

∑P∈E

1

nP.

Then:

92

Page 98: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

1. For k > l+X,χ , the dimension of Mk(Γ, χ) is given by

dk = (k/2)mX − (gX − 1) + deg(Dχ,−k).

2. For k > l+X,χ + (2/mX) deg(DS,χ,−k) , the dimension of Sk(Γ, χ) is

sk = dk − deg(DS,χ,−k).

We have the following simplification for the lower bound of k.

Corollary 4.1.2. Suppose that gX ≥ 1. Then

1. For k > 4, the dimension of Mk(Γ, χ) is

dk = (k/2)mX − (gX − 1) + deg(Dχ,−k).

2. For k > 6, the dimension of Sk(Γ, χ) is

sk = dk − deg(DS,χ,−k).

Proof. First we show that mX is greater than cX and∑

P∈E(1/nP ). Indeed,

if gX ≥ 1, then 2(gX − 1) ≥ 0, and hence all the terms in the expression of

mX are nonnegative. This implies that mX ≥ cX . If Γ has an elliptic point

P , then by definition nP ≥ 2. Therefore∑P∈E

1

nP≤∑P∈E

(1− 1/nP ) ≤ mX .

If Γ is torsion-free, this inequality is obviously satisfied, since∑

P∈E(1/nP )

is interpreted as 0. Now, we have

l+χ,X = 2 + (2/mX)∑P∈E

(1/nP ) ≤ 2 + 2 = 4.

Also, since DS,R,−k ≤ DS, we have deg(DS,R,−k) ≤ deg(DS) = cX , and so

l+χ,X + (2/mX) deg(DS,χ,−k) ≤ 4 + (2/mX) cX ≤ 4 + 2 = 6.

The formulas follow from Theorem 4.1.1.

93

Page 99: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

When gX is zero, we have the following.

Corollary 4.1.3. Suppose that gX = 0. If cX ≥ 2, then

1. For k > 4, the dimension of Mk(Γ, χ) is

dk = (k/2)mX + deg(Dχ,−k) + 1.

2. For k > 4 + max6, 4/mX, the dimension of Sk(Γ, χ) is

sk = dk − deg(DS,χ,−k).

Moreover, we have max6, 4/mX ≤ 168.

Proof. Suppose that gX = 0. If cX ≥ 2, then all the terms in the expression

of mX are nonnegative. As in the preceding proof, we conclude that l+χ,X ≤ 4.

Since deg(DS,R,−k) ≤ cX , we have

l+χ,X + (2/mX) deg(DS,χ,−k) ≤ 4 + (2/mX) cX .

Suppose that cX > 2. Since mX ≥ (cX − 2), we have

(2/mX) cX ≤ 2cX/(cX − 2) = 2 + 4/(cX − 2) ≤ 6.

Hence

l+χ,X + (2/mX) deg(DS,χ,−k) ≤ 4 + max6, 4/mX.

The result follows from Theorem 4.1.1.

By Remark 3.1.5, we have

( 4/mX ) ≤ (4) 42 = 168.

Hence max6, 4/mX ≤ 168.

Remark 4.1.4. In the remaining case, that is gX = 0 and cX < 2, one can

show that Γ is generated by elliptic elements, and so the image of χ is fi-

nite. Therefore, our generalized automorphic forms are classical automorphic

forms.

94

Page 100: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Rewriting Theorem 3.3.7 in the 1-dimensional case gives the following:??

Theorem 4.1.5. Let Γ be a Fuchsian group of the first kind, X be its compact

Riemann surface, and

mX = 2gX − 2 + cX +∑P∈E

(1− 1/nP ).

Then:

1. For k < 0, the dimension dk of Mk(Γ, χ) is zero.

2. For k < (2/mX) deg(DS,χ,−k) , the dimension sk of Sk(Γ, χ) is zero.

Corollary 4.1.6. If k < 0, then Mk(Γ, χ) and Sk(Γ, χ) are trivial.

4.2 ρ−equivariant functions

Throughout this section, Γ will be a Fuchsian subgroup of SL(2,R), and

ρ : Γ −→ GL(2,C) will be a 2-dimensional complex representation of Γ.

Definition 4.2.1. A meromorphic function h on H is called a ρ−equivariant

function with respect to Γ if

h(γ · z) = ρ(γ) · h(z) for all z ∈ H , γ ∈ Γ ,

where the action on both sides is by Mobius transformations. The set of

ρ−equivariant functions for Γ will be denoted by Eρ(Γ).

In the case ρ is the defining representation of Γ, that is ρ(γ) = γ for all

γ ∈ Γ, then elements of Eρ(Γ) are simply called equivariant functions. These

were studied extensively in [11, 12, 51, 54] and have various connections to

modular forms, quasi-modular forms, and elliptic functions. In particular,

one shows that the set of equivariant functions for a Fuchsian group Γ without

the trivial one h0(z) = z has a vector space structure isomorphic to the space

95

Page 101: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

of weight 2 unrestricted automorphic forms for Γ. Nontrivial examples are

constructed from automorphic forms. Indeed, if f is a weight k automorphic

form for Γ, then

hf (z) = z + kf(z)

f ′(z)

is equivariant for Γ. These are referred to as the rational equivariant functions

[12].

The first main result is that ρ−equivariant functions are parameterized by

2-dimensional unrestricted vector-valued automorphic form of multiplier ρ.

More precisely, if F = (f1, f2)t is an unrestricted vector-valued automorphic

form of multiplier ρ and an arbitrary weight such that f2 is nonzero, then

hF (z) = f1(z)/f2(z) is a ρ-equivariant function. We will show that, in fact,

every ρ−equivariant function arises in this way. To achieve this parametriza-

tion, we use the fact that the Schwarz derivative of a ρ−equivariant function

is a weight 4 unrestricted automorphic form for Γ, as well as the existence of

global solutions to a certain second degree differential equation. The second

main result of this section is that the ρ−equivariant functions always exist.

This will follow from Theorem 2.5.9 and Theorem 3.3.5.

We end this section by constructing examples of ρ−equivariant functions,

specially when ρ is the monodromy representation of second degree ordinary

differential equations.

4.2.1 Differential equations

Let D be a domain in C and let f be a meromorphic function on D. Its

Schwarz derivative, S(f), is defined by

S(f) =

(f ′′

f ′

)′− 1

2

(f ′′

f ′

)2

.

This is an important tool in projective geometry and differential equations.

The main properties that will be useful to us are summarized as follows ( see

[40] for more details):

96

Page 102: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Proposition 4.2.1. We have

1. If y1 and y2 are two linearly independent solutions to a differential

equation y′′ + Qy = 0 where Q is a meromorphic function on D, then

S(y1/y2) = 2Q.

2. If f and g are two meromorphic functions on D, then S(f) = S(g) if

and only if f =ag + b

cg + dfor some

(a b

c d

)∈ GL(2,C).

3. S(f γ)(z) = (cz + d)4S(f) provided γ · z ∈ D, where γ =

(∗ ∗c d

).

In particular, we have

Proposition 4.2.2. If f is a ρ-equivariant for Γ, then S(f) is an unrestricted

automorphic form of weight 4 for Γ.

Now, consider the second order ordinary differential equation (ODE)

x′′ + Px′ +Qx = 0,

where P and Q are holomorphic functions on D. This ODE has two linearly

independent holomorphic solutions on D if D is simply connected. For a

fixed z0 ∈ D, set

y(z) = x(z) exp

(∫ z

z0

1

2P (w)dw

).

The above ODE reduces to an ODE in normal form

y′′ + gy = 0 , (4.2.1)

with

g = Q− 1

2P ′ − 1

4P 2.

When the domain D is not simply connected, we may not expect to find

global solutions to (4.2.1) on D. However, under some conditions on g, global

solutions do exist as it is illustrated in the following theorem which will be

crucial for the rest of this section.

97

Page 103: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Theorem 4.2.3. Let D be a domain in C. Suppose h is a nonconstant

meromorphic function on D such that S(h) is holomorphic in D, and let

g = 12S(h). Then the differential equation

y′′ + gy = 0

has two linearly independent holomorphic solutions in D.

Proof. Let Ui, i ∈ I be a covering of D by open discs with dimV (Ui) = 2

for all i ∈ I where V (Ui) denotes the space of holomorphic solutions to

y′′ + gy = 0 on Ui. Choose Li and Ki to form a basis for V (Ui). Using

property (1) of Proposition 4.2.1, we have S(Ki/Li) = 2g = S(h) on Ui.

Now, using property (2) of Proposition 4.2.1, we have Ki/Li = αi · h for

αi ∈ GL(2,C). On the other hand, on each connected component W of

Ui ∩ Uj, we have

(Ki, Li)t = αW (Kj, Lj)

t , αW ∈ GL(2,C) ,

since each of (Ki, Li) and (Kj, Lj) is a basis of V (W ). Hence, on W we have

Ki

Li= αW ·

Kj

Lj,

and therefore

αih = αWαjh .

It follows that

αiα−1j = αW

as h is meromorphic and nonconstant and thus it takes more than three

distinct values on the domain D. Therefore, αW does not depend on W .

Moreover, on Ui ∩ Uj we have

α−1i (Ki, Li)

t = α−1j (Kj, Lj)

t . (4.2.2)

If we define f1 and f2 on Ui by

(f1, f2)t = α−1i (Ki, Li)

t ,

98

Page 104: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

then using (4.2.2), we see that f1 and f2 are well defined all over D and they

are two linearly independent solutions to y′′+ gy = 0 on all of D as they are

linearly independent over Ui.

4.2.2 The correspondence

The aim of this subsection is to prove the existence of ρ−equivariant functions

as well as their parametrization by unrestricted vector-valued automorphic

forms of multiplier ρ.

We start with the following proposition.

Proposition 4.2.4. Let F (z) = (f1(z), f2(z))t be a nonzero unrestricted

vector-valued automorphic form for ρ of a certain weight. If f2 is nonzero,

then hF := f1/f2 is a ρ−equivariant function for Γ.

Proof. Suppose that F is of weight k, k ∈ Z, and that ρ(γ) =

(aγ bγcγ dγ

),

γ ∈ Γ. Since F |kγ = ρ(γ)F , we have

f1(γ) = Jkγ (aγ f1 + bγ f2) , (4.2.3)

and

f2(γ) = Jkγ (cγ f1 + dγ f2) . (4.2.4)

This implies thatf1(γ)

f2(γ)=

aγ f1 + bγ f2

cγ f1 + dγ f2

,

that is

hF (γ) =aγ hF + bγcγ hF + dγ

= ρ(γ) · hF .

The slash operator has the following useful property known as Bol’s iden-

tity.

99

Page 105: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Proposition 4.2.5. [27] Let r be a nonnegative integer, F (z) a complex

function and γ ∈ SL(2,C), then

(F |−rγ)(r+1)(z) = F (r+1)|r+2γ(z) .

As a consequence, we have

Corollary 4.2.6. Let r be a nonnegative integer, g an unrestricted automor-

phic form of weight 2(r+ 1) for Γ, and D a domain in H that is stable under

the action of Γ. Denote by Vr(D) the solution space on D to the differential

equation

f (r+1) + gf = 0 .

Suppose that f ∈ Vr(D). Then for all γ ∈ Γ, we have f |−rγ ∈ Vr(D).

Proof. Suppose that f ∈ Vr(D). Using Bol’s identity and the fact that

g|2(r+1)

= g, we have

(f |−rγ)(r+1) + g (f |−rγ) = f (r+1)|(r+2)

γ + g (f |−rγ) =

J−(r+2)γ

(f (r+1)(γ) + (g|

2(r+1)γ) f(γ)

)=(f (r+1) + g f

)|(r+2)

γ = (0) |(r+2)

γ = 0.

Corollary 4.2.7. The operator |−r provides a representation ρr of Γ in

GL(Vr(D)). Moreover, if f1, f2,. . . ,fr+1 form a basis of Vr (if the basis

exists), then

F = (f1, f2, . . . , fr+1)t

is an unrestricted vector valued automorphic form of multiplier ρr and weight

−r for Γ.

Recall from Definition 2.1.1 that the space of unrestricted vector-valued

automorphic forms for Γ of multiplier ρ and weight k ∈ Z, is denoted by

Gk(Γ, ρ). Our first main result of this section is the following.

100

Page 106: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Theorem 4.2.8. The map

G−1(Γ, ρ) −→ Eρ(Γ)

F 7→ hF

is surjective.

Proof. Suppose that h is a ρ−equivariant function for Γ. According Proposi-

tion 4.2.2, its Schwarz derivative S(h) is an unrestricted automorphic form of

weight 4 for Γ. Let g = 12S(h) and D the complement in H of the set of poles

of g. Then D is a domain that is stable under Γ since g is an unrestricted

automorphic form for Γ.

Using the same notation as in the previous section, we have, for r = 1,

S(f1/f2) = S(h) where f1, f2 are two linearly independent solutions in

V (D) provided by Theorem 4.2.3. Hence, by Proposition 4.2.1

f1

f2

= α · h , α ∈ GL(2,C) .

Also, using Corollary 4.2.7 with r = 1, we deduce that F1 = (f1, f2)t is a

vector-valued unrestricted automorphic form for Γ of multiplier ρ1 and weight

−1. Therefore,

α−1ρ1α = ρ .

Hence F = α−1F1 lies in G−1(Γ, ρ), and hF = h on D. Since g has only

double poles, then by looking at the form of the solutions near a singular

point, and using the fact that f1 and f2 are holomorphic and thus single-

valued, we see that f1 and f2 can be extended to meromorphic functions on

H.

Remark 4.2.9. If there exists an unrestricted automorphic form f of weight

k + 1 for Γ, then (f1, f2) −→ (ff1, ff2) yields an isomorphism between

G−1(Γ, ρ) and Gk(Γ, ρ), and therefore, the above surjection in the theorem

extends to Gk(Γ, ρ) whenever it is nontrivial.

101

Page 107: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

We now prove the existence of ρ−equivariant functions. Recall that a

representation is called decomposable if it is the direct sum of two represen-

tations. If ρ is decomposable, then ρ = χ1 ⊕ χ2 for some two characters χ1

and χ2 on Γ. If k is a large enough integer, then by Theorem 2.5.9 and Theo-

rem 3.3.5, Mk(Γ, χ2) is nontrivial. Let f2 be a nonzero element of Mk(Γ, ρ2),

then for any f1 ∈ Mk(Γ, ρ1), we have F (z) = (f1(z), f2(z))t ∈ Mk(Γ, ρ).

Hence, by Proposition 4.2.4, hF = f1/f2 is a ρ−equivariant function for Γ.

Suppose that ρ is indecomposable. As usual, modulo a conjugation in

GL(2,C), the representation ρ is either even, odd, or a direct sum ρ = ρ+⊕ρ−of an even and an odd representations. Since ρ is indecomposable, the latter

case is excluded. Hence, by Theorem 2.5.9 and Theorem 3.3.5, there exists

k0 ∈ N such that for all k ≥ k0, Mk(Γ, ρ) is nontrivial.

We now show that for a certain integer k ≥ k0, there exists an element

F (z) = (f1(z), f2(z))t ∈ Mk(Γ, ρ) such that f2 is nonzero. Suppose the

converse is true, then for k ≥ k0, any nontrivial element F ∈ Mk(Γ, ρ)

has the form F = (f1, 0)t for some nonzero holomorphic function f1 on H.

According to (4.2.3), we have for all γ ∈ Γ

f1|kγ = aγ f1, (4.2.5)

where ρ(γ) =

(aγ bγcγ dγ

). Using the fact that |k is a linear operator, we deduce

that the map

α : Γ −→ C∗ γ 7→ aγ

is character on Γ. Thus f1 ∈ Gk(Γ, α). Also, from (4.2.4), we see that cγ = 0

for all γ ∈ Γ, and so F and f1 have the same cuspidal behavior with respect

to Γ. Since F ∈ Mk(Γ, ρ), we conclude that f1 ∈ Mk(Γ, α), and hence we

have an embedding of Mk(Γ, ρ) into Mk(Γ, α) given by F = (f1, 0)t −→ f1.

Consequently, for all k ≥ k0, we have

dΓ,ρ,k ≤ dΓ,α,k, (4.2.6)

where dΓ,ρ,k and dΓ,α,k respectively denote the dimensions of Mk(Γ, χ) and

102

Page 108: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Mk(Γ, α). But from Theorem 3.3.5, we know that

dΓ,ρ,k

k−→ mX , as k →∞,

anddΓ,α,k

k−→ mX

2, as k →∞.

where by see Remark 3.1.5

mX = 2gX − 2 + cX +∑P∈E

(1− 1/nP ) ≥ 1/42

Hence for k large enough, we have

dΓ,ρ,k

k>

dΓ,α,k

k,

that is

dΓ,ρ,k > dΓ,α,k,

which contradicts (4.2.6). Therefore, for a certain integer k ≥ k0, there exists

an element F (z) = (f1(z), f2(z))t ∈ Mk(Γ, ρ) such that f2 is nonzero. Then

by Proposition 4.2.4, hF = f1/f2 is a ρ−equivariant function for Γ.

Thus we have proved the following existence theorem for ρ−equivariant

functions.

Theorem 4.2.10. There exists a ρ−equivariant function hF = f1/f2, where

F (z) = (f1(z), f2(z))t ∈Mk(Γ, ρ) for a certain nonnegative integer k.

4.2.3 Examples

In this subsection, we shall illustrate our parametrization by constructing

examples of unrestricted vector valued automorphic forms, and their corre-

sponding ρ−equivariant functions.

Recall that the Γ−equivariant functions are simply the ρ0−equivariant

functions for Γ, with ρ0(γ) = γ for all γ ∈ Γ. One can easily see that

F0(z) =

(z

1

), z ∈ H

103

Page 109: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

is an unrestricted vector-valued automorphic form for Γ of multiplier ρ0 and

weight −1, and

hF0(z) = z, z ∈ H.

As we mentioned before, if we take a weight k unrestricted automorphic

form f for Γ, then

hf (z) = z + kf(z)

f ′(z), ∈ H

is an equivariant function for Γ. This corresponds to the (k− 1) unrestricted

vector-valued automorphic form for Γ of multiplier ρ0 given by

F (z) =

(k f + z f ′

f ′

), z ∈ H. (4.2.7)

Our next example is the monodromy representation of a differential equa-

tion. Indeed, let U be a domain in C such that C \ U contains at least two

points. The universal covering of U is then H as it cannot be P1(C) because

U is noncompact and it cannot be C because of Picard’s theorem.

Let π : H −→ U be the covering map. We consider the differential

equation on U

y′′ + Py′ +Qy = 0, (4.2.8)

where P and Q are two holomorphic functions on U . This differential equa-

tion has a lift to Hy′′ + π∗Py′ + π∗Qy = 0 . (4.2.9)

Let V be the solution space to (4.2.9) which is a 2-dimensional vector space

since H is simply connected. Let γ be a covering transformation in Deck(H/U)

which is isomorphic to the fundamental group π1(U) and let f ∈ V . Then

γ∗f = f γ−1 is also a solution in V . This defines the monodromy represen-

tation of π1(U):

ρ : π1(U) −→ GL(V ) .

If f1 and f2 are two linearly independent solutions in V , we set F = (f1, f2)t.

Then we have

F γ = ρ(γ)F,

104

Page 110: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

that is F ∈ G0 (π1(U), ρ). Therefore, the quotient f1/f2 is a ρ−equivariant

function on H for the group π1(U), which is a torsion-free Fuchsian group.

105

Page 111: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

Bibliography

[1] M.F. Atiyah, Vector bundles on an elliptic curve, Proc. London Math.

Soc. 7 (1957) 414–452.

[2] W. Barth, C. Peters: A. Van de Ven, Compact Complex Surfaces,

Ergebnisse der Mathematik und ihrer Grenzgebiet 3. Folge. Band 4 A

Series of Modem Surveys in Mathematics, Springer-Verlag, 1984.

[3] P. Bantay, The dimension of spaces of vector-valued modular forms

of integer weight, Lett. Math. Phys (2013) 103:1243–1260.

[4] P. Bantay and T. Gannon. Vector-valued modular functions for the

modular group and the hypergeometric equation, Commun. Number

Th. Phys. 1 (2008) 637–666.

[5] A. F. Beardon, The geometry of discrete groups, Springer-Verlag,

New-York, 1983.

[6] R. Borcherds, The Gross-Kohnen-Zagier theorem in higher dimen-

sions, Duke Math. J. 97, no. 2, 219–233 (1999), Correction: Duke

Math. J. 105, no. 1, 183–184 (2000).

[7] R. Borcherds, Reflection groups of Lorentzian lattices, Duke Math. J.

104, no. 2, 319–366 (2000).

106

Page 112: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

[8] C. Dong, H. Li, G. Mason, Modular-invariance of Trace Functions

in Orbifold Theory and Generalized Moonshine, Comm. Math. Phys.

214 (2000), 1–56.

[9] W. Eholzer, N.-P. Skoruppa, Modular invariance and uniqueness of

conformal characters, Comm. Math. Phys. 174 No. 1 (1995), 117–136.

[10] M. Eichler and D. Zagier, The Theory of Jacobi Forms, Prog. Math.

55 (Birkhauser, Boston, 1985).

[11] A. Elbasraoui, A. Sebbar, Equivariant forms: Structure and geometry.

Canad. Math. Bull. Vol. 56 (3), (2013) 520–533.

[12] A. Elbasraoui, A. Sebbar, Rational equivariant forms. Int J. Number

Th. 08 No. 4(2012), 963–981.

[13] O. Forster, Lectures on Riemann surfaces, Graduate Texts in Mathe-

matics, 81. Springer-Verlag, New York, 1991.

[14] J. Fischer, An approach to the Selberg trace formula via the Selberg

zeta function, Lecture Notes in Mathematics, 1253, Springer Verlag,

Berlin, Heidelberg, New York (1987)

[15] E. Freitag, Riemann surfaces: Sheaf Theory, Riemann Surfaces, Au-

tomorphic Forms. CreateSpace Independent Publishing Platform, 1

edition (2014)

[16] J. H. Gallier, Logarithms and Square Roots of Real Matrices (2008).

Technical Reports (CIS). Paper 876.

[17] T. Gannon, The theory of vector-valued modular forms for the mod-

ular group, arXiv:1310.4458

[18] F.R. Gantmacher, The Theory of Matrices, Vol. I. AMS Chelsea, First

edition, 1977.

107

Page 113: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

[19] P. Griffiths, J. Harris, Principles of algebraic geometry. Wiley Classics

Library. John Wiley & Sons, Inc., New York, 1994.

[20] R. Gunning, Introduction to holomorphic functions of several vari-

ables (volume I) Function Theory. Wadsworth & Brooks/Cole, 1990.

[21] R. Gunning, Introduction to holomorphic functions of several vari-

ables (volume III) Homological Theory. Wadsworth & Brooks/Cole,

1990.

[22] D. A. Hejhal, The Selberg trace formula for PSL(2,R), Vol. 1. Lecture

Notes in Mathematics 548, Springer-Verlag, Berlin-New York, 1976.

[23] D. A. Hejhal, The Selberg trace formula for PSL(2,R), Vol. 2. Lecture

Notes in Mathematics 1001, Springer-Verlag, Berlin-New York, 1983.

[24] S. Katok, Fuchsian groups. The University of Chicago Prees, Chicago

and London, 1992.

[25] W. Kohnen, G. Mason, On generalized modular forms and their ap-

plications. Nagoya Math. J. 192 (2008), 119–136.

[26] W. Kohnen, Y. Martin, On product expansions of generalized modular

forms. Ramanujan J (2008) 15: 103–107.

[27] M. Knopp, Polynomial automorphic forms and nondiscontinuous

groups, Trans. Amer. Math. Soc. 123 (1966), 506-520. MR 34:341.

[28] M. Knopp, G. Mason, Generalized modular forms. J. Number Theory

99 (2003), no. 1, 1–28.

[29] M. Knopp, G. Mason, Parabolic generalized modular forms and their

characters. Int. J. Number Theory 5 (2009), no. 5, 845–857.

[30] M. Knopp, G. Mason, Vector-valued modular forms and Poincare se-

ries, Illinois J. of Math. 48 (2004), 1345–1366.

108

Page 114: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

[31] M. Knopp, G. Mason, Logarithmic vector-valued modular forms, Acta

Arith. 147 (2011) 261–282.

[32] M. Knopp, J. Lehner and W. Raji, Eichler cohomology for generalized

modular forms, Int. J. Number Theory 5 (2009) 1049–10.

[33] M. Knopp, W. Raji, Eichler cohomology for generalized modular

forms II. Int. J. Number Theory Vol. 6, no. 5 (2010) 1083–1090

[34] M. Kuga, G. Shimura, On vector differential forms attached to auto-

morphic forms, J. Math. Soc. Japan 12 (1960), 258–270.

[35] C. Marks, Irreducible vector-valued modular forms of dimension less

than six, Illinois J. Math. 55 (2011) 1267–1297.

[36] C. Marks, G. Mason, Structure of the module of vector-valued mod-

ular forms, J. London Math. Soc. 82 (2010)32–48.

[37] G. Mason, Vector-valued modular forms and linear differential oper-

ators, Int J. Number Th. 3 (2007) 377–390.

[38] F. Martin, R. Emmanuel, Formes modulaires et periodes, Seminaires

Et Congres, collection SMF, num 12.

[39] C. Meneses, On vector-valued Poincare series of weight 2.

arXiv:1408.6494v2 [math.CV].

[40] J. McKay; A. Sebbar Fuchsian groups, Automorphic functions and

Schwarzians. Math. Ann. 318 (2), (2000) 255–275.

[41] M. Miyamoto, Modular invariance of vertex operator algebras satis-

fying C2-cofiniteness, Duke Math. J. 122 No. 1 (2004), 51–91.

[42] Y. Choie, M. H. Lee, Symmetric tensor representations, quasimodular

forms, and weak Jacobi forms. arXiv:1007.4590v1 [math.NT].

109

Page 115: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

[43] V. B. Mehta, C. S. Seshadri, Moduli of vector bundles on curves with

parabolic structures, Math. Ann. 248 (1980), 205–239.

[44] W. Raji, Fourier coefficients of generalized modular forms of negative

weight. Int. J. Number Theory 5 (2009), no. 1, 153–160.

[45] W. Raji, Construction of generalized modular integrals, Funct. Ap-

prox. Comment. Math. 41(2) (2009) 105–112.

[46] R. Rankin, Modular Forms and Functions, Cambridge Univ. Press,

Cambridge, 1977.

[47] W. Roelcke, Das Eigenwertproblem der automorphen Formen in der

hyperbolischen Ebene I, Math. Ann. 167(1966) pp.292-337.

[48] W. Roelcke, Das Eigenwertproblem der automorphen Formen in der

hyperbolischen Ebene II, Math. Ann. 168(1967) pp.261-324.

[49] A. Selberg, On the estimate of Fourier coefficients of modular forms,

Proc. Sympos. Pure Math VIII (Amer. Math. Soc. Providence, 1965)

pp. 1–15.

[50] H. Saber, A. Sebbar, Vector-valued automorphic forms and vector

bundles. arXiv:1312.2992v3 [math.NT].

[51] H. Saber, Equivariant functions for the Mobius subgroups and appli-

cations. arXiv:1412.8100 [math.NT].

[52] H. Saber, A. Sebbar, Equivariant functions and vector-valued modular

forms. Intl J. Number Th. 10, 949 (2014)

[53] H. Saber, A. Sebbar, On the critical points of modular forms. J. Num-

ber Theory 132 (2012) 1780–1787.

[54] A. Sebbar, A. Sebbar, Equivariant functions and integrals of elliptic

functions. Geom. Dedicata 160 (1), (2012) 37–414.

110

Page 116: Vector-valued automorphic forms and vector bundles › ... › 3 › Saber_Hicham_2015_thesis.pdf · 2017-01-31 · Vector-valued automorphic forms and vector bundles by Hicham Saber

[55] G. Shimura, Introduction to the Arithmetic Theory of Automorphic

Functions (Princeton University Press, 1971).

[56] G. Shimura, Sur les integrales attaches aux formes automorphes, J.

Math. Soc. Japan 11 (1959), 291–311.

[57] N.-P. Skoruppa, Uber den Zusammenhang zwischen Jacobiformen und

Modulformen halbganzen Gewichts, Ph.D. Thesis, Universitat Bonn

(1984).

[58] G. Springer, Introduction to Riemann surfaces, Addison-Wesley,

Reading, 1957.

111