+ All Categories
Home > Documents > HRG Master's Project

HRG Master's Project

Date post: 09-Feb-2017
Category:
Upload: hugh-geller
View: 33 times
Download: 6 times
Share this document with a friend
77
Ramanujan type congruences for the Klingen-Eisenstein series A Project Presented to the Graduate School of Clemson University In Partial Fulfillment of the Requirements for the Degree Master’s of Science Mathematics by Hugh Roberts Geller December 2016 Accepted by: Dr. Jim Brown, Committee Chair Dr. Kevin James Dr. Hui Xue
Transcript
Page 1: HRG Master's Project

Ramanujan type congruences for the Klingen-Eisensteinseries

A Project

Presented to

the Graduate School of

Clemson University

In Partial Fulfillment

of the Requirements for the Degree

Master’s of Science

Mathematics

by

Hugh Roberts Geller

December 2016

Accepted by:

Dr. Jim Brown, Committee Chair

Dr. Kevin James

Dr. Hui Xue

Page 2: HRG Master's Project

Table of Contents

Title Page . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . i

1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.1 Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2

2 Siegel Modular Forms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42.1 Siegel Modular Group . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42.2 Modular Forms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72.3 Klingen Eisenstein Series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152.4 R-modules of modular forms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 252.5 The Hecke Algebra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33

3 Results of T. Kikuta and S. Takemori . . . . . . . . . . . . . . . . . . . . . . . . . 523.1 Definitions, Remarks, and Main Results . . . . . . . . . . . . . . . . . . . . . . . . . 523.2 Lemmas and Their Proofs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 533.3 Proof of the main theorem and its corollary . . . . . . . . . . . . . . . . . . . . . . . 603.4 Numerical Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 623.5 New Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67

4 Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72

Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74

ii

Page 3: HRG Master's Project

Chapter 1

Introduction

The following work is a Master’s project meant as introduction to the rich and complex

topic that is Siegel modular forms. In this first chapter we introduce the motivation of the paper

by T. Kikuta and S. Takemori cited as [9] as well as introducing some notation that will be used

consistently throughout this write-up.

In Chapter 2, we introduce the concept of a Siegel modular form. We introduce several

concepts and relate them to the degree one case, which is more commonly referred to as elliptic

modular forms. The reasons for doing this are two-fold. First, we wish to demonstrate how Siegel

modular forms act a generalization of the well-known elliptic modular forms. Second, we do this in

order to understand the work of [9].

In Chapter 3 we will directly address the findings of [9]. We fill in the details of their proofs

in order to fully understand their stated main theorem and its corollary. In short, the paper studies

congruences of the Fourier coefficients of Siegel modular forms. In particular, they define what it

means for a modular form to be a (mod pm) cusp form. We define this precisely in Definition 3.1.1

but for now it suffices to say that a degree n Siegel modular form is a (mod pm) cusp form if every

Fourier coefficient of rank 0 ≤ r ≤ n− 1 vanishes modulo pm.

The main theorem states that given a number field K for nearly every prime ideal p in

the ring of algebraic integers, (mod pm) cusp forms are congruent to true cusp forms of the same

weight. The precise theorem is as follows

Main Theorem ([9]) There exists a finite set Sn(K) of prime ideals in K depending on n such

1

Page 4: HRG Master's Project

that the following holds: For a prime ideal p of O not contained in Sn(K) and a (mod pm) cusp

form F ∈Mk(Γn)Opwith k > 2n, there exists G ∈ Sk(Γn)Op

such that F ≡ G (mod pm).

The corollary gives conditions under which we can take what is known as a Hecke eigen

cusp form and force a constant multiple of its attached Klingen-Eisenstein series to be congruent to

a true cusp form modulo pm. Precisely, it states

Main Corollary ([9]) Let k > 2n be even and f ∈ Sk(Γr)Kf, n > r, a Hecke eigenform. For the

Klingen-Eisenstein series [f ]n−r attached to f , we choose a prime ideal p in OKfwith p 6∈ Sn(Kf )

such that υ(n)p ([f ]n−r) = υp(Φ[f ]n−r) −m, m ∈ Z≥1. Then there exists F ∈ Sk(Γn)Op

such that

α[f ]n−r ≡ F (mod pm) for some 0 6= α ∈ pm.

Kikuta and Takemori assert that this is a generalization of Ramanujan’s congruence

σ11(n) ≡ τ(n) (mod 691).

They make this assertion since σ11(n) denotes the nth Fourier coefficient of the Eisenstein series

of weight 12, which we denote in Section 3.4 as ψ(1)12 , and τ(n) is the nth Fourier coefficient of

Ramanujan’s ∆ function [9]. More precisely, they are assert that they are generlizing the result that

691ψ(1)12 is a (mod 691) cusp form of degree 1 that is congruent modulo 691 to the Ramanujan’s ∆

function, the unique full level, degree 1, weight 12 cusp form.

1.1 Notation

We fix the following notation.

Let R be a commutative ring with identity.

Let Matn(R) be the set of n× n matrices with entries in R.

We denote the tranpose of a matrix γ ∈ Matn(R) by tγ.

For matrices A,B ∈ Matn(R) ⊂ Matn(C) we define A[B] = tBAB.

For V ∈ GLn(R) we define U(V ) =

V 0n

0ntV −1

.

For tS = S ∈ Matn(R) we define T (s) =

In S

0n In

.

2

Page 5: HRG Master's Project

Let K be a number field with ring of integers OK .

For a Hecke Eigenform f , let Kf be the number field generated by adjoining the Hecke

eigenvalues of f .

3

Page 6: HRG Master's Project

Chapter 2

Siegel Modular Forms

2.1 Siegel Modular Group

For n ∈ Z≥1 define J2n =

0n In

−In 0n

where 0n is the n×n zero matrix and In is the n×n

identity matrix. Note, we will write J for J2n when the dimension is clear from the context. Using

this matrix, we define the general symplectic group GSp2n(R) by

GSp2n(R) = {γ ∈ Mat2n(R) : tγJγ = µ(γ)J, µ(γ) ∈ R×}.

We call µ(γ) the similitude factor of γ. Therefore, writing γ =

A B

C D

, where A,B,C,D ∈

Matn(R), yields the following relations; tAC = tCA, tDB = tBD, tAD − tCB = µ(γ)In,

tDA− tBC = µ(γ)In, and γ−1 = µ(γ)−1

tD −tB

−tC tA

.

We will work with various subgroups of GSp2n(R). For R ⊆ R are the interested in the

subgroup matrices with positive similitude factor, denoted GSp+2n(R). We are particularly interested

in the subgroup of symplectic, 2n× 2n matrices with integral entries. We refer to this group as the

Siegel modular group (of degree n) and write

Γn = Sp2n(Z) = ker{µ : GSp2n(Z)→ R×}.

4

Page 7: HRG Master's Project

Note that µ(γ) = 1 if and only if tγJγ = J.

We will be working over the Siegel upper half-space of degree n and which is given by

Hn = {Z ∈ Matn(C) : tZ = Z, Z = X + iY, X, Y ∈ Matn(R), Y > 0}

where Y > 0 means Y is positive definite. For the degree 1 case, we see that this matches the

definition of the upper half plane. We define an action of GSp+2n(R) on Hn by

Z 7→ γ · Z = (AZ +B)(CZ +D)−1

where γ =

A B

C D

∈ GSp+2n(R). For this to be a well defined group action we must check the

following:

1. CZ +D is invertible;

2. γ · Z ∈ Hn;

3. I2n · Z = Z;

4. (γγ′) · Z = γ · (γ′ · Z) for all γ, γ′ ∈ GSp+2n(R) and all Z ∈ Hn.

To check the first condition we use the following identity

t(CZ +D)(AZ +B)− t

(AZ +B)(CZ +D) = (tDA− tBC)Z +t Z(tCB − tAD)

= µ(γ)InZ − µ(γ)InZ

= µ(γ)(Z − Z)

= 2iµ(γ)Y.

We now recall that a n × n matrix is invertible if and only if it gives a bijective map on an n-

dimensional vector space to itself. Thus, if CZ + D is not invertible then there exists a non-zero

5

Page 8: HRG Master's Project

vector v ∈ Cn such that (CZ +D)v = 0. This would then yield

tvY v =−i2µ(γ)−1t(v)

(t(CZ +D)(Az +B)− t

(AZ +B)(CZ +D))v

=−i2µ(γ)−1

(t((CZ +D)v)(AZ +B)v − t

(v)t(AZ +B)(CZ +D)v

)= 0,

but this contradicts the fact that Y is assumed to be positive definite. Therefore no such v ∈ Cn

can exist and we conclude CZ +D is invertible.

We now must show that γ · Z ∈ Hn. We start by checking for symmetry

t(CZ +D)(γ · Z − t

(γ · Z))(CZ +D) =t(CZ +D)(AZ +B)− t

(AZ +B)(CZ +D)

= Z − tZ

= 0n.

Since CZ +D is invertible this implies γ · Z is symmetric. Similarly, we have

t(CZ +D)Im(γ · Z)(CZ +D) =

−i2t(CZ +D)

(γ · Z − t

(γ · Z))

(CZ +D)

=−i2

(t(CZ +D)(AZ +B)− t

(AZ +B)(CZ +D))

= µ(γ)Y

> 0n.

The final inequality holds since Y is positive definite and γ ∈ GSp+2n(R) means µ(γ) > 0. We thus

conclude that γ · Z ∈ Hn.

The third condition is clear since

I2n · Z = (InZ + 0n)(0nZ + In)−1 = Z.

6

Page 9: HRG Master's Project

For the last condition let γ =

A B

C D

and γ′ =

A′ B′

C ′ D′

be in GSp+2n(R). We then get

γ · (γ′ · Z) = γ ·((A′Z +B′)(C ′Z +D′)−1

)=

(A(A′Z +B′)(C ′Z +D′)−1 +B

) (C(A′Z +B′)(C ′Z +D′)−1 +D

)−1

= (A(A′Z +B′) +B(C ′Z +D′)) (C(A′Z +B′) +D(C ′Z +D′))−1

= ((AA′ +BC ′)Z + (AB′ +BD′)) ((CA′ +DC ′)Z + (CB′ +DD′))−1

= (γγ′) · Z.

Thus, our definition for γ · Z defines a group action of GSp+2n(R) of Hn. Since Γn and

GSp+2n(R) are subgroups of GSp+

2n(R) (for R a subring of R), we get both subgroups give a group

action on Hn via the same action.

2.2 Modular Forms

Using the group action defined in Section 2.1, we define the weight k slash operator on

functions f : Hn → C by

(f |kγ)(Z) := µ(γ)nk−n(n+1)

2 j(γ, Z)−kf(γ · Z), where j(γ, Z) = det(CZ +D)

for γ =

A B

C D

∈ GSp+2n(R). We now give the definition of a Siegel modular form.

Definition 2.2.1 A classical Siegel modular form of weight k, degree n > 1, and level Γn is a

holomorphic function f : Hn → C such that for all γ ∈ Γn we have (f |kγ)(Z) = f(Z).

The collection of modular forms of weight k and level Γn is a C-vector space which we denote

Mk(Γn) (when the degree is clear we will simply write Mk). We note that by Theorem 2 of Chapter

2, Section 4 of [11] that this vector space is finite-dimensional.

For degree 1 classical Siegel modular forms we are simply working with elliptic modular

forms and require f to be holomorphic at the cusps, that is the orbit of i∞ under the group action.

Initially this was a requirement for any degree but Max Koecher proved in 1954 that for degree

n > 1 we automatically get the function is holomorphic at the cusps.

7

Page 10: HRG Master's Project

Furthermore, f has Fourier expansion

f(Z) =∑

0≤T∈Λn

a(T ; f)qT , qT = exp{2πitr(TZ)}, Z ∈ Hn,

where

Λn = {T = (tij) ∈ Matn(Q) : tT = T and tii, 2tij ∈ Z}.

That is, the sum runs over all semi-positive definite, half-integral matrices. This series is uniformly

convergent on compact subsets of Hn ([23], page 9).

It is important to note that the Fourier coefficients are given by

a(T ; f) =

∫X mod 1

f(Z)e−2πitr(TZ)dX

where dX =∏i≤jdXi,j . Also, X mod 1 means Xi,j mod 1 for i ≤ j and Xi,j is the (i, j)-th entry of

X. We note that these coefficients are independent of Y as demonstrated on pages 32 and 33 of

[25]. We will revisit the Fourier expansion once we have a definition for the subspace of cusp forms.

To motivate the idea of weight k, degree n, level Γn cusp forms we introduce the Siegel

operator

Φ : Mk(Γn) → Mk(Γn−1)

f 7→ limt→∞

f

Zit

, Z ∈ Hn−1, t ∈ R.

It is trivial to see this defines a linear map due to basic limit laws. Page 11 of [23] explains

how to identify Φ(f) in Mk(Γn−1).

It is crucial that we acknowledge that Φ depends on the n though the notation does not

reflect this. It should be clear from the context which degree we are working with. The notation

can be difficult to follow when we consider the iterative mapping given by

Mk(Γn)Φ→Mk(Γn−1)

Φ→ · · · Φ→Mk(Γr+1)Φ→Mk(Γr)

since each Φ above is differs from the rest. To take forms from Mk(Γn) to Mk(Γr) as in the sequence

above we use Φn−r.

8

Page 11: HRG Master's Project

Using this operator, we define Sk(Γn) := ker Φ to be the subspace of cusp forms. To see

why this makes sense we visit the case of elliptic modular forms, that is, take f ∈Mk(Γ1). We then

have

Φ(f) = limt→∞

f(it)

= limt→∞

(a0(f) +

∞∑i=1

ai(f) · e2πi(it)

)

= a0(f) +

∞∑i=1

limt→∞

(ai(f) · e−2πt

)= a0(f) +

∞∑i=1

ai(f)(

limt→∞

e−2πt)

= a0(f).

Note that since the Fourier expansion is uniformly convergent we can use the dominated convergence

theorem to exchange the limit and summation in line 3.

Thus, f ∈ ker Φ if and only if a0(f) = 0, which agrees with the elliptic definition of cusp

forms. Revisiting the formula for the Fourier coefficients allows us to determine when exactly

f ∈ ker Φ ⊂Mk(Γn).

Let f ∈Mk(Γn), T ′ ∈ Λn−1, Z ′ ∈ Hn−1, and X ′ = Re(Z ′). We have

a(T ′; Φ(f)) =

∫X′ mod 1

(Φ(f))(Z ′)e2πitr(T ′Z′)dX ′

=

∫X′ mod 1

limt→∞

f

Z ′it

e2πitr(T ′Z′)dX ′

= limt→∞

∫X mod 1

f

Z ′it

e

2πitr

T′

0

Z′

it

dX

= a

T ′

0

; f

.

9

Page 12: HRG Master's Project

To get the third equality use Theorem 4.3 of [23] to get

∣∣∣∣∣∣∣fZ ′

it

e2πitr(T ′Z′)

∣∣∣∣∣∣∣ =

∣∣∣∣∣∣∣fZ ′

it

∣∣∣∣∣∣∣∣∣∣e2πitr(T ′Z′)

∣∣∣ ≤Mt · 1 = Mt

where Mt = M is fixed for large enough t. Thus we can apply the dominated convergence theorem

in order to exchange the limit and integral. In the same line, we switch from X ′ to X where

X = Re

Z ′it

=

X ′0

.

Thus, f ∈ Sk(Γn) if and only if a

T ′

0

; f

= 0 for all T ′ ∈ Λn−1. We can strengthen

the result to f ∈ Sk(Γn) if and only if a(T ; f) = 0 for all 0 ≤ T ∈ Λn such that 0 6< T . To get this

we use the identity

a(tUTU ; f) = det(U)−ka(T ; f), for all 0 ≤ T ∈ Λn, and U ∈ GLn(Z).

It is important to note the following consequence of this relation.

Lemma 2.2.2 ([23], Corollary 4.2) A classical Siegel Modular form of weight k with kn ≡ 1(mod 2)

vanishes.

Proof Let f ∈Mk(Γn) with kn ≡ 1 mod 2 and let 0 ≤ T ∈ Λn. Observe thatt(−In)T (−In) = T .

Thus if kn ≡ 1(mod 2), then we have

a(T ; f) = a(t(−In)T (−In); f)

= det(−In)−ka(T ; f)

= (−1)−nka(T ; f)

= −a(T ; f).

Thus we have a(T ; f) = 0 for all T ∈ Λn and therefore f vanishes. �

Using this result, we see that every Siegel modular form of odd weight, even degree, and full level

must be a cusp form. To see this, suppose f ∈ Mk(Γn) with k odd. Applying the Siegel operator,

10

Page 13: HRG Master's Project

we see that Φ(f) ∈Mk(Γn−1) but

k(n− 1) ≡ 1 mod 2

and hence f ∈ ker Φ, proving it is a cusp form. Consequently the main theorem of [9] is trivial for

k odd. Moving forward, unless otherwise specified, we take k even.

Since we know how Φ interacts with the fourier coefficients of a form f ∈Mk(Γn), we now

consider the more general result for Φn−r. The closed form of the iterative application is given by

Φn−r : Mk(Γn) → Mk(Γr)

f 7→ limt→∞

f

Z ′itIn−r

, Z ′ ∈ Hr, t ∈ R

where In−r is the n − r × n − r identity matrix. To see the motivation behind this definition,

consider the Fourier coefficients, a(T ′; Φn−r(f)), of a form f ∈ Mk(Γn). The affect of Φn−r on a

form f ∈Mk(Γn) is shown by its affect on the Fourier coefficients

a(T ′; Φn−r(f)) = a

T ′

0n−r

; f

.

Using this and taking r ≥ 1, we have

a(T ′; ΦΦn−r(f)) = a

T ′

01

; Φn−r(f)

= a

T ′

01

0n−r

; f

= a

T ′

0n−r+1

; f

= a(T ′; Φn−r+1(f)).

Thus, for any f ∈Mk(Γn) and r ≥ 1 the Fourier coefficients of ΦΦn−r(f) and Φn−r+1(f) coincide.

Since the coefficients of a power series, and in particular those of a Fourier expansion, uniquely

11

Page 14: HRG Master's Project

determine the function, we must have Φ ◦ Φn−r = Φn−r+1. Furthermore, induction now grants us

Φr−j ◦ Φn−r = Φn−j for 0 ≤ j ≤ r ≤ n.

By Corollary 5.5 of [23], it is known that Φ is surjective for even k > 2n, so by the first

isomorphism theorem of vector spaces we have

Mk(Γn)/Sk(Γn) = Mk(Γn)/ ker Φ 'Mk(Γn−1).

Since Mk(Γn) is a C-vector space we can express it as the direct sum of Sk(Γn) and its compliment,

Sk(Γn)c. We wish to address the structure of Sk(Γn)c. To do so we introduce the Petersson inner

product.

Definition 2.2.3 Let f, g ∈Mk(Γn) with at least one a cusp form. We define the Petersson inner

product as

〈f, g〉 :=

∫Γn\Hn

f(Z)g(Z) det(Y )kdνn(Z)

where dνn(Z) = det(Y )−(n+1)∏i≤j dxijdyij, Z = X + iY , and xij , yij are the i, jth entry of X,Y ,

respectively.

Lemma 2.2.4 For n ∈ Z≥1 the measure dνn is invariant under the action of GSp+2n(R) on Hn.

That is, dνn(γZ) = dνn(Z) for all γ ∈ GSp+2n(R).

Proof Let g = GSp+2n(R) and define X∗ = Re(γZ) and Y ∗ = Im(γZ). The Jacobian for the change

of variables of Z 7→ gZ is given by

det

(∂(X∗, Y ∗)

∂(X,Y )

)where

∂(X∗, Y ∗)

∂(X,Y )=

∂X∗

∂X∂Y ∗

∂X

∂X∗

∂Y∂Y ∗

∂Y

.

Since the action of γ is holomorphic on Hn we have the Cauchy-Riemann equations

∂X∗

∂X=∂Y ∗

∂Yand

∂X∗

∂Y= −∂Y

∂X.

Using these relations we see that

12I − i

2I

− i2I

12I

∂X∗

∂X∂Y ∗

∂X

∂X∗

∂Y∂Y ∗

∂Y

I iI

iI I

=

∂X∗

∂X − i∂X∗

∂Y 0

0 ∂X∗

∂X + i∂X∗

∂Y

=

∂(gZ)∂Z 0

0 ∂(gZ)∂Z

.

12

Page 15: HRG Master's Project

Consequently, we have

det

(∂(X∗, Y ∗)

∂(X,Y )

)= det

(∂(γZ)

∂Z

)det

(∂(γZ)

∂Z

).

We now write dZ = (dzij) and d(gZ) for the matrices of the differentials of Z and γZ,

respectively. Let γ =

A B

C D

, then (γZ)(CZ +D) = (AZ +D). We have

d(γZ)(CZ +D) + (γZ)CdZ = AdZ.

Multiplying both sides on the left byt(CZ +D) produces

t(CZ +D)AdZ = d(γZ)[CZ +D] +

t(CZ +D)(γZ)CdZ

= d(γZ)[CZ +D] +t(CZ +D)

t((AZ +B)(CZ +D)−1)CdZ

= d(γZ)[CZ +D] +t(AZ +B)CdZ.

Rewriting, we have

d(γZ)[CZ +D] =t(CZ +D)AdZ − t

(AZ +B)CdZ

= Z(tCA− tAC)dZ + (tDA− tBC)dZ

= µ(γ)dZ since γ ∈ GSp+2n(R).

We then have

d(γZ) =t(CZ +D)−1µ(γ)dZ(CZ +D)−1 = dZ

[√µ(γ)(CZ +D)−1

].

From page 9 of [11], it is known that the determinant of the linear map Z 7→ tMZM = Z[M ] on

symmetric matrices, Z, is given by detMn+1. Since∂(γZ)

∂Zmeasures changes in γZ as Z varies and

we have d(γZ) = dZ[√

µ(γ)(CZ +D)−1], we obtain

det

(∂(γZ)

∂Z

)= det(

√µ(γ)(CZ +D)−1)n+1 = µ(γ)n(n+1)/2 det(CZ +D)−(n+1).

13

Page 16: HRG Master's Project

Similarly, we obtain

det

(∂(γZ)

∂Z

)= µ(γ)n(n+1)/2det(CZ +D)

−(n+1).

Combining all the results have

dνn(γZ) = det(Y ∗)−(n+1) det

(∂(X∗, Y ∗)

∂(X,Y )

)∏i≤j

dxijdyij

=

(µ(γ)n det(Y )

|det(CZ +D)|2

)−(n+1)

det

(∂(γZ)

∂Z

)det

(∂(γZ)

∂Z

)∏i≤j

dxijdyij

=|det(CZ +D)|2n+2

µ(γ)n(n+1) det(Y )n+1µ(γ)n(n+1)|det(CZ +D)|−2n−2

∏i≤j

dxijdyij

= det(Y )−(n+1)∏i≤j

dxijdyij

= dνn(Z).

Therefore the measure is invariant under the action of GSp+2n(R) on Hn.

Furthermore, the integral defined in the Petersson inner product converges when f or g is

a cusp form. Due to the orthogonal decomposition,

Mk(Γn) = Sk(Γn)⊕Sk(Γn)⊥,

we take Sk(Γn)⊥ as our choice for Sk(Γn)c. Since Φ is surjective for k > 2n, we obtain Φ(Sk(Γn)⊥

)=

Mk(Γn−1) = Sk(Γn−1)⊕Sk(Γn−1)⊥. We then define Mkn,n−1 = Sk(Γn)⊥ ∩Φ−1

(Sk(Γn−1)

). Iter-

ating this process we obtain the following subspaces due to [15] using the notation of [10]

Mkn,r =

Sk(Γn) r = n

{f ∈ Sk(Γn)⊥ : Φ(f) ∈Mkn−1,r} 0 ≤ r < n.

More precisely, since Mk(Γn) is finite-dimensional we are able to use induction to identify these

spaces as

Mkn,r = ker Φn+r+1/ ker Φn+r.

Our identification makes it clear that these spaces are pair-wise disjoint but we explicitly show this

for clarity. Consider r1 6= r1 and, without loss of generality, let 0 ≤ r1 < r2 < n. We then have the

14

Page 17: HRG Master's Project

following

Mkn,r1 ∩Mk

n,r2 = {f ∈ Sk(Γn)⊥ : Φ(f) ∈Mkn−1,r1 ∩Mk

n−1,r2}

= {f ∈ Sk(Γn)⊥ : Φn−r2(f) ∈Mkr2,r1 ∩Mk

r2,r2}

= {f ∈ Sk(Γn)⊥ : Φn−r2(f) ∈Mkr2,r1 ∩Sk(Γr2)}.

However, since r2 > r1 we have Mkr2,r1 = {g ∈ Sk(Γr2)⊥ : Φ(g) ∈ Mk

r2−1,r1}. By orthogonality

of Mkr2,r1 and Mk

r2,r2 = Sk(Γr2) for r1 < r2, we must then have Mkr2,r1 ∩Sk(Γr2) = {0}, showing

Φn−r2(f) = 0. However, by the definition of Mkn,r2 , anything that vanishes under Φn−r2 must be 0.

Therefore we conclude Mkn,r1 ∩Mk

n,r2 = {0} for r1 6= r2. Note that we did not need to check the

cases 0 ≤ r1 < r2 = n since the result would follow trivially from the definition of Mkn,r.

Since k > 2n we have Φn is surjective and maps Mk(Γn) to Mk(Γ0) = Mk0,0 = C we can

write

Mk(Γn) = Mkn,0 + Mk

n,1 + · · ·+ Mkn,n−1 + Mk

n,n.

This combined with the subspaces being pairwise disjoint we get

Mk(Γn) = Mkn,0 ⊕Mk

n,1 ⊕ · · · ⊕Mkn,n−1 ⊕Mk

n,n

= Mkn,0 ⊕Mk

n,1 ⊕ · · · ⊕Mkn,n−1 ⊕Sk(Γn).

Thus, given f ∈Mk(Γn) we have a unique fr ∈Mkn,r such that f =

∑nr=1 fr. For each fr

we have Φn−r(fr) = fgr ∈ Mkr,r = Sk(Γr), noting that Φn−n is just the identity operator yielding

fn = gn. Therefore, we have a (n+ 1)-tuple of cusp forms (g0, g1, . . . , gn) associated to f . The next

section will highlight the significance of this association.

2.3 Klingen Eisenstein Series

In the previous section we saw how to take a Siegel modular form of arbitrary degree and

obtain a tuple of cusp forms. This, however, supposes Siegel modular forms of degree greater than

1 exist. In this section we give a construction of degree n Siegel modular forms from degree r forms.

We will then revisit the association of a modular form with a tuple of cusp forms.

15

Page 18: HRG Master's Project

For 0 ≤ r < n, we define the Klingen parabolic subgroup

∆n,r :=

∗ ∗

0n−r,n+r ∗

∈ Γn

.

We construct the Klingen Eisenstein series associated to f ∈ Sk(Γr) by

Ekn,r(f)(Z) =∑

γ∈∆n,r\Γn

j(γ, Z)−kf((γ · Z)∗), Z ∈ Hn,

where (γ · Z)∗ is the upper left r × r matrix of γ · Z = (AZ +B)(CZ +D)−1.

As an example, take n = 1, r = 0 and f = 1 ∈Mk(Γ0) . We have

Ek1,0(1)(z) =∑

a b

c d

∈∆1,0\Γ1

(cz + d)−k.

We will show that this is precisely the level 1, weight k Eisenstein series for elliptic modular forms.

To see this we only need to show there is a bijection between cosets of ∆1,0 in Γ1 and relatively

prime pairs (c, d) ∈ Z2 up to sign; that is, (c, d) and (−c,−d) correspond to the same coset but

(c,−d) produces a different coset.

We start by giving a more explicit description of ∆1,0 ⊂ Sp2(Z) = SL2(Z):

∆1,0 =

∗ ∗

01−0,1+0 ∗

∈ Γ1

=

±1 α

0 ±1

∈ Γ1 : α ∈ Z

.

If γ =

a b

c d

∈ SL2(Z), then we have gcd(c, d) = 1 and

∆1,0γ =

±a+ αc ±b+ αd

±c ±d

∈ Γ1 : α ∈ Z

.

Since −I2 ∈ ∆1,0 we can map the pair (−c,−d) to (c, d). For simplicity, we focus on the pair (c, d).

16

Page 19: HRG Master's Project

Now suppose γ′ =

a′ b′

c d

∈ SL2(Z), then (a′, b′) is a solution to the Diophantine equation

xc− yd = 1.

If we consider (a, b) as our initial solution the equation, then it is an elementary result that any

solution is of the form

(a+ ct, b+ dt) where t ∈ Z,

and thus there exists some α ∈ Z such that a′ = a+ αc and b′ = b+ αd. Consequently we have

γ′ =

a+ αc b+ αd

c d

=

1 α

0 1

a b

c d

∈ ∆1,0γ.

Therefore, we have a bijection of sets given by

∆1,0\Γ1 ←→{

(c, d) ∈ Z2 : gcd(c, d) = 1}

∆1,0

a b

c d

←→ (c, d)

and hence

Ek1,0(1)(z) =∑

γ∈∆1,0\Γ1

(cz + d)−k =∑

gcd(c,d)=1

(cz + d)−k,

which is precisely the level 1, weight k Eisenstein series for elliptic modular forms.

Moreover, we have the following theorem,

Theorem 2.3.1 ([23], Theorem 5.4) Let n ≥ 1 and 0 ≤ r ≤ n and k > n + r + 1 be integers

with k even. Then for every cusp form f ∈ Sk(Γr) the series Ekn,r(f) converges to a classical Siegel

modular form of weight k in Mk(Γn) and satisfies Φn−r(Ekn,r(f)) = f .

Using [10] we first show the convergence of Ekn,r(f) and then address the iterated Siegel

operator. To do this we show the series converges absolutely-uniformly. For f ∈ Sk(Γr) we know by

proposition 4 (p. 45) of [11] that the quantity det(Y )k/2f(Z) is bounded for all X + iY = Z ∈ Hr.

17

Page 20: HRG Master's Project

Thus, up to some constant, we have the majorant Ekn,r(f)(Z) in Hn by

Gkn,r(Z) =∑

γ∈∆n,r\Γn

det(Im(γ · Z)∗)−k2 |j(γ, Z)|−k.

We study this series on vertical strips of the Siegel upper half-space given by

Bn(d) = {Z ∈ Hn : tr(X2) ≤ d−1, Y ≥ dIn}

for d > 0. On these strips we have the following lemma.

Lemma 2.3.2 ([10], p. 33) For n ≥ 1 and d > 0 there exists c > 0, dependent only on n and d,

so that

det(Im(γ · Z)∗)|j(γ, Z)|2 ≥ cdet(Im(γ · iIn)∗)|j(γ, iIn)|2

for all Z ∈ Bn(d), γ ∈ Sp2n(R), and 0 ≤ r ≤ n.

Note that r in the lemma is in reference to (Z)∗ meaning take the upper-left r × r matrix

of Z. Combining this lemma with the decomposition

η =

η∗ 0

0 η2

Ir η3

0 In−r

where γ · Z = ξ(Z) + iη(Z) we can obtain the uniform convergence of Gkn,r(Z) by studying

Gkn,r(iIn) =∑

γ∈∆n,r\Γn

det(Im(γ · iIn)∗)−k2 |j(γ, iIn)|−k

=∑

γ∈∆n,r\Γn

det(η∗(iIn))−k2 |j(γ, iIn)|−k

=∑

γ∈∆n,r\Γn

det(η2(iIn))k2 det(η(iIn))−

k2 |j(γ, iIn)|−k

=∑

γ∈∆n,r\Γn

det(η2(iIn))k2 det(In)−

k2

(|j(γ, iIn)|−2

)− k2 |j(γ, iIn)|−k

=∑

γ∈∆n,r\Γn

det(η2(iIn))k2 .

18

Page 21: HRG Master's Project

Note that the second to last equality arise from the identity

t(CZ +D)η(Z)(CZ +D) =

t(CZ +D)Im(γ · Z)(CZ +D) = Y

for γ ∈ Γn, noting that η changes if we change γ. In [10] we see why this is absolutely convergent

and consequently get that Ekn,r(f) converges to a form in Mk(Γn).

We now consider Φn−r(Ekn,r(f)) and note that the absolute convergence of the series

Ekn,r(f)(Z) =∑

γ∈∆n,r\Γn

j(γ, Z)−kf((γ · Z)∗), Z ∈ Hn,

allows us to apply Φn−r term by term in the summation. The first element in the sum we consider

is any representative

γ =

Ar,r Ar,n−r Br,r Br,n−r

An−r,r An−r,n−r Bn−r,r Bn−r,n−r

Cr,r Cr,n−r Dr,r Dr,n−r

0n−r,r 0n−r,n−r 0n−r,r Dn−r,n−r

∈ ∆n,r ⊂ Γn

and since this matrix is in Γn we have the following relations

Ar,r Br,r

Cr,r Dr,r

∈ Γr, An−r,n−rtDn−r,n−r = In−r,

and Ar,n−rtDn−r,n−r = 0r,n−r = Cr,n−r

tDn−r,n−r.

Since Dn−r,n−r is invertible we must have Ar,n−r = Cr,n−r = 0r,n−r. Thus a representative of ∆n,r

produces the following; Ar,rZ∗ +Br,r Br,n−r

An−r,rZ∗ +Bn−r,r itAn−r,n−r +Bn−r,n−r

Cr,rZ∗ +Dr,r Dr,n−r

0n−r,r Dn−r,n−r

−1

=

Ar,rZ∗ +Br,r Br,n−r

An−r,rZ∗ +Bn−r,r itAn−r,n−r +Bn−r,n−r

(Cr,rZ

∗ +Dr,r)−1 −(Cr,rZ

∗ +Dr,r)−1Dr,n−r(Dn−r,n−r)

−1

0n−r,r (Dn−r,n−r)−1

19

Page 22: HRG Master's Project

=

(Ar,rZ∗ +Br,r)(Cr,rZ

∗ +Dr,r)−1 ∗

∗ ∗

.

Next, we see that

det

Cr,rZ∗ +Dr,r Dr,n−r

0n−r,r Dn−r,n−r

= det(Cr,rZ∗ +Dr,r) det(Dn−r,n−r) = ±det(Cr,rZ

∗ +Dr,r)

sinceDn−r,n−r and its inverse are integral matrices. Using these results, writing g(Z) = j(γ, Z)−kf((γ·

Z)∗), and recalling that k is even, we have

Φn−r(g(Z)) = limt→∞

j

γ,Z∗

itIn−r

−k

f

γ ·

Z∗itIn−r

= limt→∞

det

C Z∗ 0r,n−r

0n−r,r itIn−r

+D

−k

f

(Ar,rZ

∗ +Br,r)(Cr,rZ∗ +Dr,r)

−1 ∗

∗ ∗

= (±det(Cr,rZ∗ +Dr,r))

−kf

(Ar,rZ

∗ +Br,r)(Cr,rZ∗ +Dr,r)

−1 ∗

∗ ∗

∗ , γ ∈ ∆n,r

= det(Cr,rZ∗ +Dr,r)

−kf((Ar,rZ∗ +Br,r)(Cr,rZ

∗ +Dr,r)−1) since k is even.

= f(Z∗) since

Ar,r Br,r

Cr,r Dr,r

∈ Γr.

We will now show that the iterated Siegel operator forces the other terms in Ekn,r(f) to 0.

To get this result, we return to Gkn,r; by showing the summands become zero away from ∆n,r we

show the same applies for Ekn,r(f). Writing Z∗ = X∗ + iY ∗, we have

Z =

Z∗ 0r,n−r

0n−r,r itIn−r

and Y =

Y ∗ 0r,n−r

0n−r,r tIn−r

.

Rewriting the identity

t(CZ +D)η(Z)(CZ +D) =

t(CZ +D)Im(γ · Z)(CZ +D) = Y

20

Page 23: HRG Master's Project

produces the expression

η∗−1(Z) 0

0 η−12 (Z)

= η(Z)−1

= (CZ +D)(Y −1)t(CZ +D)

= (CZ +D)

(Y ∗)−1 0r,n−r

0n−r,r −it−1In−r

t(CZ +D)

=

∗ ∗

∗ Y ∗−1[(Z∗tCn−r,r + tDn−r,r)] + t−1[(ittCn−r,n−r + tDn−r,n−r)]

.

Thus we have

η−12 (Z) = Y ∗−1[(Z∗tCn−r,r + tDn−r,r)] + t−1[(ittCn−r,n−r + tDn−r,n−r)]

allowing us to simplify the following expression

det(Im(γ · Z)∗)|j(γ, Z)|2 = det(η∗)|j(γ, Z)|2

=det(η(Z))

det(η2)|j(γ, Z)|2

= det(η−12 ) det(Y )

= det(η−12 ) det(Y ∗)tn−r

= det(tη−12 ) det(Y ∗)

= det(Y ∗)P (t),

where P (t) = det(tY ∗−1[Z∗tCn−r,r + tDn−r,r] + In−r[ittCn−r,n−r + tDn−r,n−r]). Thus using this

equality we have

Gkn,r(Z) =∑

γ∈∆n,r\Γn

(det(Y ∗)P (t))− k

2 .

Since the Siegel operator uses the limit as t → ∞ and we know the series is absolutely convergent,

we apply the limit term-wise. Moreover, since det(Y ∗) is independent of t we need only focus on

limt→∞

P (t)−k2 . Clearly, P (t) is a polynomial in t and thus vanishes for deg(P ) > 0. Thus we will

21

Page 24: HRG Master's Project

assume deg(P ) = 0 and show this only occurs for γ ∈ ∆n,r. To see this we note that for t ≥ 0, both

tY ∗−1[Z∗tCn−r,r + tDn−r,r] and In−r[ittCn−r,n−r + tDn−r,n−r]

are positive semi-definite since Y ∗−1 and In−r are positive definite. Furthermore their sum must be

positive definite since the left-hand side of

det(Im(γ · Z)∗)|j(γ, Z)|2 = det(Y ∗)P (t)

is non-zero for all Z ∈ Hn and γ ∈ ∆n,r. As a consequence we have

det(In−r[ittCn−r,n−r + tDn−r,n−r]) ≤ P (t)

for all t and where we have equality when 0r,r = Y ∗−1[Z∗tCn−r,r + tDn−r,r]. Since Y ∗−1 is positive

definite, this occurs when 0n−r,r = Z∗tCn−r,r+ tDn−r,r which holds if and only if 0n−r,r = Cn−r,r =

Dn−r,r. This then leaves

P (t) = det(In−r[ittCn−r,n−r + tDn−r,n−r])

= det(t2Cn−r,n−rtCn−r,n−r +Dn−r,n−r

tDn−r,n−r),

which still produces a polynomial in t unless 0 = Cn−r,n−r. Thus we have limt→∞

P (t)−k2 vanishes

unless γ ∈ ∆n,r\Γn has 0 = Cn−r,r = Dn−r,r and 0 = Cn−r,n−r. This is precisely when γ ∈ ∆n,r.

Therefore all terms in the series in Gkn,r(Z), and thus in Ekn,r(Z), with γ away from ∆n,r vanish for

Z =

Z∗ 0

0 itIn−r

as t→∞. Thus for Z∗ ∈ Hr, 0 ≤ r ≤ n, f ∈ Sk(Γn), and k > n+ r + 1 even

22

Page 25: HRG Master's Project

we have:

Φn−r(Ekn,r(f))(Z∗) = limt→∞

∑γ∈∆n,r\Γn

det

C Z∗ 0r,n−r

0n−r,r itIn−r

+D

−k

f

γ ·

Z∗ 0r,n−r

0n−r,r itIn−r

=∑

γ∈∆n,r\Γn

limt→∞

det

C Z∗ 0r,n−r

0n−r,r itIn−r

+D

−k

f

γ ·

Z∗ 0r,n−r

0n−r,r itIn−r

= (±det(Cr,rZ∗ +Dr,r))

−kf

(Ar,rZ

∗ +Br,r)(Cr,rZ∗ +Dr,r)

−1 ∗

∗ ∗

∗ , γ ∈ ∆n,r

= det(Cr,rZ∗ +Dr,r)

−kf((Ar,rZ∗ +Br,r)(Cr,rZ

∗ +Dr,r)−1) since k is even.

= f(Z∗) since

Ar,r Br,r

Cr,r Dr,r

∈ Γr.

Thus we see that Ekn,r( · ) is a right inverse to Φn−r when we restrict to degree r cusp forms. From

here we use the notation of [12] and write

[f ]nr := Ekn,r(f) for f ∈ Sk(Γr).

We do this in order to extend the definition of the operator [ · ]nr to all of Mk(Γr). For f = [fj ]rj

with fj ∈ Sk(Γj) we define [f ]nr = [fj ]nj . Using this notion as well as our desire for a linear operator,

we take f ∈ Mk(Γr) and get the r + 1-tuple of cusp forms, (f0, · · · , fr), mentioned at the end of

Section 2.2 and define

[f ]nr =r∑j=0

[fj ]nj =

r∑j=0

Ekn,j(fj).

Note that this agrees with our construction in the previous section since we have seen Φn−j([fj ]nj ) =

fj .

Using this relation we introduce the notation

En,j,k =[Sk(Γj)

]nj.

We then have the decomposition

[Mk(Γn)]nr =

r⊕j=0

En,j,k.

23

Page 26: HRG Master's Project

It is important to note that this decomposition is one into orthogonal subspaces with respect to the

Petersson inner product. However, to see this we use Maass’ extension of the inner product [15,

p. 96] that says for f =∑rj=0[fj ]

nj and g =

∑rj=0[gj ]

nj we write

〈f, g〉M :=

n∑j=0

⟨Φn−j([fj ]

nj ),Φn−j([gj ]

nj )⟩

=

n∑j=0

〈fj , gj〉

which is a finite sum of convergent inner products, since both fj , gj are cusp forms for all j.

To see that this agrees with the original definition we consider 〈f, g〉 and 〈f, g〉M where

at least one of f and g is a cusp form. Without loss of generality, let g ∈ Sk(Γn) and write

f =∑nj=0[fj ]

nj . Since f ∈ Mk(Γn) = Sk(Γn) ⊕ Sk(Γn)⊥ and thus can also write f = fc + fp

where fc ∈ Sk(Γn) and fp ∈ Sk(Γn)⊥. Recall that at the end of Section 2.2 that we showed the

(n+ 1)-tuple of cusp forms (f0, f1, . . . , fn) associated with f is unique, hence we must have fn = fc.

We see that

〈f, g〉M = 〈fn, g〉+

n−1∑j=0

〈Φn−j([fj ]nj ), 0〉 since f ∈ Sk(Γn)

= 〈fn, g〉+

n−1∑j=0

0

= 〈fn, g〉

= 〈fc, g〉

= 〈fc, g〉+ 〈fp, g〉 since fp ∈ Sk(Γn)⊥

= 〈fc + fp, g〉

= 〈f, g〉.

Now consider f ∈ En,`,k and g ∈ En,j,k where ` 6= j and, without loss of generality, 0 ≤ ` <

24

Page 27: HRG Master's Project

j ≤ n. Then both forms have expressions f = [f`]n` and g = [gj ]

nj , and thus we have

〈f, g〉M =⟨Φn−`([f`]

n` ),Φn−`([g`]

n` )⟩

+⟨Φn−j([fj ]

nj ),Φn−j([gj ]

nj )⟩

= 〈f`, g`〉+ 〈fj , gj〉

= 〈f`, 0〉+ 〈0, gj〉 by definition of f, g

= 0.

Thus we see that f ∈ En,`,k and g ∈ En,j,k are orthogonal for 0 ≤ ` < j ≤ n. For the specific case of

n = j we have

En,n,k = [Sk(Γn)]nn = Sk(Γn),

thus showing degree n cusp forms are orthogonal to all Klingen-Eisenstein series.

2.4 R-modules of modular forms

Recall that Mk(Γn) is a finite-dimensional C-vector space.

Given a subring R ⊂ C we wish to consider the R-module

Mk(Γn)R = {f ∈Mk(Γn) : a(T ; f) ∈ R for all T ∈ Λn}

and the corresponding cusp forms

Sk(Γn)R = Sk(Γn) ∩Mk(Γn)R.

It is natural to ask what information on Mk(Γn) can be obtained by studying Mk(Γn)R. In

[21], Theorem 3 implies

Mk(Γn) ∼= Mk(Γn)Q ⊗Q C

for k even and gives a singular condition for the result to hold when k is odd. We do not state the

condition as [20] shows it holds for all odd k ∈ Z and at no other point will we use the condition.

Note, this isomorphism is often exploited and sometimes the two sets are written as being equal

instead of isomorphic. We also note that this isomorphism demonstrates that Mk(Γn)Q is a finite

dimensional Q-vector space since tensoring with C is merely an extension of scalars. Therefore, if

25

Page 28: HRG Master's Project

{f1, . . . , fm} is Q-basis of Mk(Γn)Q then it also acts as a C-basis of Mk(Γn). Thus, for f ∈Mk(Γn)

we identify it in Mk(Γn)Q ⊗Q C by

f =

m∑j=1

αjfj ←→m∑j=1

(fj ⊗ αj).

Using this identification we prove the following lemma.

Lemma 2.4.1 If k ∈ Z>0 and K is a number field, then

Mk(Γn)Q ⊗Q K ∼= Mk(Γn)K .

Proof Let {f1, . . . , fm} be as defined above and consider the following map extended linearly

ψ : Mk(Γn)Q ⊗Q K −→ Mk(Γn)K

f ⊗ c 7−→ cf.

It is clear that this is a linear map of K-modules. To see this map is surjective, let f ∈Mk(Γn)K ⊂

Mk(Γn). There exists {αj}mj=1 ⊂ C such that f =∑mj=1 αjfj . Note that this is equivalent to

a(T ; f) =

m∑j=1

αja(T ; fj) for all T ∈ Λn.

Since the fj form a basis, it is possible to take a subset {Ti}mi=1 ⊂ Λn such that a(Ti; f) 6= 0 for

some i and det(A) 6= 0 where A = (a(Ti, fj)) ∈ Matm(Q). Supposing otherwise would imply linear

dependence amongst the fj , contradicting the definition of a basis. Setting α =t(α1 · · ·αm) and

a =t(a(T1; f) · · · a(Tm; f)) produces Aα = a. Let Aj denote the matrix produced by replacing the

jth column of A with a. By Cramer’s rule, for all j, we then have

αj =det(Aj)

det(A)∈ K since det(A) ∈ Q,det(Aj) ∈ K.

Consequently∑mj=1(fj ⊗ αj) ∈ Mk(Γn)Q ⊗Q K and is thus in the preimage of f , demonstrating

that ψ is surjective.

Now suppose∑ri=1(hi ⊗ αi) ∈ kerψ. Combining this with the fact that {f1, . . . , fm} is a

26

Page 29: HRG Master's Project

basis, we obtain

0 =

r∑i=1

αihi

=

r∑i=1

αi

m∑j=1

ci,jfj

where ci,j ∈ Q

=

m∑j=1

(r∑i=1

αici,j

)fj .

By definition of a basis we must then have∑ri=1 αici,j = 0 for all j. Using this produces

r∑i=1

(hi ⊗ αi) =

r∑i=1

m∑j=1

ci,jfj ⊗ αi

=

r∑i=1

m∑j=1

(ci,jfj ⊗ αi)

=

m∑j=1

r∑i=1

(fj ⊗ αici,j)

=

m∑j=1

(fj ⊗

r∑i=1

αici,j

)

=

m∑j=1

(fj ⊗ 0)

= 0.

Therefore kerψ = {0}, showing that the map is injective. Combining this with the surjectivity we

have that ψ is an isomorphism. �

Revisiting [21], we get the following lemma and introduce a corollary to highlight our in-

tended use.

Lemma 2.4.2 ([21], Theorem 1) Let Q denote an algebraic closure of Q, and O the ring of all

algebraic integers in Q. Then, for every f ∈ Mk(Γn)Q, there exists c ∈ Z>0 such that cf ∈

Mk(Γn)O.

Corollary 2.4.3 Let K be a number field and consider a subring K ⊇ R ⊇ OK . Then, for every

f ∈Mk(Γn)R, there exists c ∈ Z>0 such that cf ∈Mk(Γn)OK.

Proof Let f ∈ Mk(Γn)R ⊂ Mk(Γn)Q. By Lemma 2.4.2 there exists c ∈ Z>0 such that cf ∈

27

Page 30: HRG Master's Project

Mk(Γn)O. However, we also have that cf ∈Mk(Γn)R ⊂Mk(Γn)K , and thus

cf ∈Mk(Γn)O ∩Mk(Γn)K = Mk(Γn)O∩K = Mk(Γn)OK. �

We note that c is not unique; in fact, if we define the ideal

(f)K := {α ∈ OK : αf ∈Mk(Γn)OK}

of OK then it is easy to see that {0} ( cOK ⊆ (f)K for any c satisfying the result of Corollary

2.4.3. Since we know the ideals (f)K are strictly larger than the zero-ideal, it is worth asking if

there exist rings, R, satisfying OK ⊆ R ⊆ K such that R× ∩ (f)K 6= ∅ for all f ∈Mk(Γn)R. Taking

R = K is clearly true and later we will see that for a prime ideal p ⊆ OK , that the localization with

respect to p, Op, also satisfies this property. Knowing that the set of rings satisfying this property

is non-empty warrants the consideration of the following proposition.

Proposition 2.4.4 For all rings R satisfying OK ⊆ R ⊆ K such that R× ∩ (f)K 6= ∅ for all

f ∈Mk(Γn)R, we have

Mk(Γn)R ∼= Mk(Γn)OK⊗OK

R

as R-modules.

Proof We define the following map of R-modules

ϕR : Mk(Γn)R −→ Mk(Γn)OK⊗OK

R

f 7−→ uf ⊗ u−1,

where we take u ∈ R× ∩ (f)K , thus u−1 is defined. Note that the choice of u does not matter since

for any u, v ∈ R× ∩ (f)K we have u, v ∈ OK since (f)K ⊆ OK and hence

uf ⊗ u−1 = uf ⊗ vv−1u−1

= uvf ⊗ (uv)−1

= vf ⊗ uu−1v−1

= vf ⊗ v−1.

28

Page 31: HRG Master's Project

Thus we see the map is well-defined. Now observe that if we take f to be the 0 function and u = 1

then we get ϕ(0) = 0⊗ 1. For arbitrary f ∈Mk(Γn)R and αβ ∈ R ⊆ K with α, β ∈ OK and β ∈ R×

we have,

α

βϕR(f) =

α

β(uf ⊗ u−1)

= uf ⊗ α

βu−1

= uαf ⊗ (uβ)−1

= (uβ)

βf

)⊗ (uβ)−1

= ϕR

βf

).

We wish to extend ϕR linearly by defining

ϕR

∑j=1

fj

=∑j=1

ϕR(fj).

To see this is well defined let f1, f2, f3 ∈Mk(Γn)R such that f3 = f1 + f2. Let u1, u2, u3 be chosen

so that ui ∈ R× ∩ (fi)K . Note that the units of any ring is a multiplicatively closed set and thus

any product ue11 ue22 u

e33 ∈ R× for any ej ∈ Z. Thus we see that

ϕR(f1 + f2) = ϕR(f3)

= u3f3 ⊗ u−13

= u3f3 ⊗ (u1u2)(u1u2u3)−1

= (u1u2u3)(f1 + f2)⊗ (u1u2u3)−1

=(u2u3(u1f1)⊗ (u1u2u3)−1

)+(u1u3(u2f2)⊗ (u1u2u3)−1

)=

(u1f1 ⊗ u−1

1

)+(u2f2 ⊗ u−1

2

)= ϕR(f1) + ϕR(f2).

Since ϕR is a well-defined linear map, it is natural to ask about its kernel and image. Since R is a

subring of a field it is an integral domain, thus we have

0 = ϕR(f) = uf ⊗ u−1

29

Page 32: HRG Master's Project

if and only if one of the following three situations occurs:

1. u = 0,

2. u−1 = 0,

3. f = 0.

We note that this is an if and only if relation since for a pure tensor to be zero there must

be an element in the equivalence class such that one of the two elements in the tensor vanishes.

However, since R is an integral domain and f ∈Mk(Γn)R, we have uf vanishes if and only if u = 0

or f = 0. Since u ∈ R× we must have u 6= 0. By the same rational we also have u−1 6= 0. Thus we

must have f = 0, demonstrating kerϕR = {0}, that is, ϕR is injective.

Consequently, we know that Mk(Γn)R is isomorphic to the image of ϕR. Let

∑j=1

(fj ⊗

αjβj

)∈Mk(Γn)OK

⊗OKR

with αj , βj ∈ OK and βj ∈ R×. We obtain a preimage through the following process.

∑j=1

fj ⊗αjβj

=∑j=1

αjfj ⊗ β−1j

=∑j=1

βjαjβjfj ⊗ β−1

j

=∑j=1

ϕR

(αjβjfj

)

= ϕR

∑j=1

αjβjfj

.

Thus ϕR is an isomorphism and we have proven the proposition. �

Taking R = Q, we obtain Mk(Γn)Q ∼= Mk(Γn)Z ⊗Z Q. Combining this with Lemma 2.4.1

yields

Mk(Γn)K ∼= Mk(Γn)Q ⊗Q K

∼= Mk(Γn)Z ⊗Z Q⊗Q K

∼= Mk(Γn)Z ⊗Z K.

30

Page 33: HRG Master's Project

This is necessary result for Mizumoto’s proof of the following lemma from [17].

Lemma 2.4.5 ([17], Lemma A.4) Let {ω1, . . . , ω`} ⊂ OK be an integral basis of K; that is, for

any α ∈ K there exists {βi}`i=1 ⊂ Q such that α =∑i=1

βiωi. Then

Mk(Γn)OK=⊕i=1

Mk(Γn)Z · ωi

for n, k ∈ Z>0.

Proof The first step is to show that

Mk(Γn)OK=∑i=1

Mk(Γn)Z · ωi.

The inclusion ⊃ is obvious so we turn our attention to ⊂. Let f ∈ Mk(Γn)OK. Since we identify

Mk(Γn)Z ⊗Z K = Mk(Γn)K via

f =

m∑j=1

αjfj ←→m∑j=1

(fj ⊗ αj) with m ∈ Z>0,

there exist f1, · · · , fr ∈ Mk(Γn)Z and ai ∈ K such that f =∑ri=1 aifi. Furthermore, since

{ω1, . . . , ω`} ⊂ OK is an integral basis of K there exist cij ∈ Q such that

ai =∑j=1

cijωj .

Thus we have

f =

r∑i=1

aifi

=

r∑i=1

∑j=1

cijωj

fi

=∑j=1

(r∑i=1

cijfi

)ωj

=∑j=1

hjωj .

31

Page 34: HRG Master's Project

Since the hj ’s are Q-linear combinations of f1, · · · , fr ∈Mk(Γn)Z, we have that all hj ∈Mk(Γn)Q.

Now for all T ∈ Λn we have

a(T ; f) =∑j=1

a(T ;hj)ωj .

Since {ω1, . . . , ωg} is an integral basis of K and a(T ;ϕ) ∈ OK , we must have a(T ;hj) ∈ Z for all

j since the representation fo a(T ;ϕ) is unique with respect to the basis {ω1, . . . , ω`} and OK =

Z[ω1, . . . , ω`] as a free Z-module. Consequently, we have hj ∈ Mk(Γn)Z for all j; thus giving the

summation we set out to demonstrate.

The next step is to show that this summation is in fact a direct sum. Suppose f = 0,

then a(T ; f) = 0 for all T ∈ Λn. Since {ω1, . . . , ω`} is a basis, the only linear combination of these

elements that produces 0 is the trivial one, and thus a(T ;hj) = 0 for all T ∈ Λn and all j. Therefore

hj = 0 for all j, showing that the sum is actually a direct sum, yielding the desired result. �

We note this is equivalent to Mk(Γn)OK

∼= Mk(Γn)Z ⊗Z OK through the linear extension

of the map

f =

g∑j=1

hjωj 7→g∑j=1

(hj ⊗ ωj).

To understand the work of [9] requires an analogous result localizing both Z and OK at prime ideals.

To get this result we need to show that R = Op satisfies the requirements of Proposition 2.4.4. To

see this we use the following lemma.

Lemma 2.4.6 ([17], Lemma A.3) For any f ∈ Mk(Γn)Op with n, k ∈ Z>0, there exists a p-unit

u ∈ OK such that uf ∈Mk(Γn)OK.

In other words, for all f ∈ Mk(Γn)Op there exists u ∈ (OK − p) ⊂ O×p such that uf ∈

Mk(Γn)OK. Thus we have u ∈ O×p ∩ (f)K giving the following corollary to Proposition 2.4.4.

Corollary 2.4.7 Let K be a number field with ring of integers OK . Let Op be the localization of

OK at the prime ideal p ⊆ OK . Then

Mk(Γn)Op∼= Mk(Γn)OK

⊗OKOp.

Proof Set R = Op, the localiztion of OK at the prime ideal p. By Lemma 2.4.6 we have seen for

all f ∈Mk(Γn)Op that O×p ∩ (f)K 6= ∅. Thus we apply Proposition 2.4.4 to get the result.

32

Page 35: HRG Master's Project

We immediately have the following corollary.

Corollary 2.4.8 Let p be prime and Z(p) the integers localized at (p). Then

Mk(Γn)Z(p)

∼= Mk(Γn)Z ⊗Z Z(p).

Proof Set K = Q and apply Corollary 2.4.7. �

Corollary 2.4.9 Let Op be the localization of OK at the prime ideal p ⊆ OK . Then

Mk(Γn)Op∼= Mk(Γn)Z(p)

⊗Z(p)Op,

where (p) = Z ∩ p.

Proof Recall for rings R ⊂ S that S ∼= R ⊗R S. Since (p) ⊂ p and Z ⊂ OK , we have Z(p) ⊂ Op.

Using this observe that

Mk(Γn)Op∼= Mk(Γn)OK

⊗OKOp by Corollary 2.4.7

∼=(Mk(Γn)Z ⊗Z OK

)⊗OK

Op by Lemma 2.4.5

∼= Mk(Γn)Z ⊗Z Op

∼= Mk(Γn)Z ⊗Z(Z(p) ⊗Z(p)

Op)

∼= Mk(Γn)Z(p)⊗Z(p)

Op by Corollary 2.4.8. �

2.5 The Hecke Algebra

The natural question to ask now is whether or not the operators Φn−r and [·]nr preserve any

special properties of modular forms. To answer that question we need to introduce the following.

Definition 2.5.1 For N ∈ Z>0, the subgroup

Γn(N) :=

A B

C B

∈ Γn : A ≡ D ≡ In (mod N), B ≡ C ≡ 0n (mod N)

of Γn is called the principle subgroup of level N .

33

Page 36: HRG Master's Project

Observe that Γn(N) is the kernel of the surjective group homomorphism

Γn = Sp2n(Z) −→ Sp2n(Z/NZ)A B

C D

7−→

A B

C D

(mod N).

Thus we have Γn(N) / Γn as well as

[Γn : Γn(N)] = #Sp2n(Z/NZ) <∞.

Definition 2.5.2 A subgroup Γ ⊆ Γn such that Γn(N) ⊂ Γ ⊂ Γn is called a congruence subgroup of level N .

Technically, if N is a multiple of N then we have Γn(N) ⊂ Γn(N) and we have any congru-

ence subgroup of level N is also a congruence subgroup N . To remove any ambiguity, we will refer

to Γ as a congruence subgroup of level N if N is minimal with respect to Γn(N) ⊂ Γ ⊂ Γn. We

further note that Γn(1) = Γn.

Since the principle congruence subgroups have finite index in Γn, it is easy to see that any

congruence subgroup also has finite index. To see this observe that if Γn(N) ⊆ Γ ⊂ Γn then

[Γn : Γ] ≤ [Γn : Γ][Γ : Γn(N)] = [Γn : Γn(N)] <∞.

These definitions and observations are necessary to show that we can write

ΓngΓn =∐i

Γngi, g, gi ∈ GSp+2n(Q).

We start by demonstrating that ΓngΓn can be expressed as such a union. To do so, observe

that ΓngΓn = {γ1gγ2 : γ1, γ2 ∈ Γn} and consequently we have the orbit space Γn\ΓngΓn = {Γngγ :

γ ∈ Γn}. Now suppose that we have γ1, γ2 such that Γngγ1∩Γngγ2 6= ∅. There must exist h1, h2 ∈ Γn

such that h1gγ1 = h2gγ2. This is equivalent to γ1 = g−1h−11 h2gγ2 and thus

Γngγ1 = Γngg−1h−11 h2gγ2

= Γnh−11 h2gγ2

= Γngγ2.

34

Page 37: HRG Master's Project

Therefore we get the union is disjoint by taking distinct orbits Γngi = Γngγi. To see that the union

is finite we use the notation Γ = g−1Γng ∩ Γn and utilize analogues to the techniques used to get

Lemmas 5.1.1 and 5.1.2, respectively, of [5].

Lemma 2.5.3 Let g ∈ GSp+2n(Q), then Γ is a congruence subgroup.

Proof Since g is a rational matrix, there exists N ∈ Z>0 such that Ng, Ng−1 ∈ Mat2n(Z). Note,

by definition, we have Γn(N) ⊂ Γn. Set N = N3 and observe that

gΓn(N)g−1 ⊂ g (I2n +N ·Mat2n(Z)) g−1

= g(I2n + N3 ·Mat2n(Z)

)g−1

= I2n + (Ng)(N ·Mat2n(Z)

)(Ng−1)

⊂ I2n + N ·Mat2n(Z).

Furthermore, it is clear that every element of gΓn(N)g−1 has determinant equal to 1. Combining

results we have gΓn(N)g−1 ⊂ Γn(N) and, equivalently,

Γn(N) ⊂ g−1Γn(N)g ⊂ g−1Γng.

Since, by definition, Γn(N) ⊂ Γn we therefore have

Γn(N) ⊂ g−1Γng ∩ Γn = Γ

and conclude that Γ is a congruence subgroup. �

Lemma 2.5.4 The set {γi} is a set of coset represeentatives of Γ\Γn if and only if {gi} = {gγi} is

a set of orbit representatives for Γn\ΓngΓn.

Proof Consider the map of sets

Γn −→ Γn\ΓngΓn

γ 7−→ Γngγ.

35

Page 38: HRG Master's Project

This map is surjective since for any Γnα ∈ Γn\ΓngΓn there exist γ1, γ2 ∈ Γn such that

Γnα = Γnγ1gγ2 = Γngγ2

which is the image of γ2. Now suppose Γngγ1 = Γngγ2; then we must have γ1γ−12 ∈ g−1Γng. We

also have by definition that γ1γ−12 ∈ Γn. Consequently γ1γ

−12 ∈ g−1Γng ∩ Γn = Γ, that is, two

elements of Γn produce the same orbit in Γn\ΓngΓn if they differ by an element of Γ. Thus the full

preimage of Γngγ is the coset Γγ, demonstrating that there is a bijection between the coset space

and orbit space. Consequently {Γγi} is a complete set of cosets if and only if {Γngγi} = {Γngi} is

a complete set of orbits, therefore demonstrating the desired result. �

Since Γ is a congruence subgroup of Γn, as demonstrated by Lemma 2.5.3, we have the set

of coset representatives, {γi}, of Γ\Γn is finite. Therefore, by Lemma 2.5.4, we have {gi} = {gγi}

is a finite set of orbit respresentatives for Γn\ΓngΓn. We then conclude that ΓngΓn =∐i

Γngi is a

finite disjoint union.

Definition 2.5.5 For g ∈ GSp+2n(Q), we define the weight k, ΓngΓn operator by

f [ΓngΓn]k :=∑i

f |kgi

where the gi are defined by ΓngΓn =∐i

Γngi.

The above is an example of a Hecke operator. More generally, a Hecke operator is any

element of the Hecke algebra.

Definition 2.5.6 The Hecke algebra, Tn, is the set of finite formal sums of double cosets ΓngΓn,

g ∈ GSp+2n(Q) with coefficients in C. That is

Tn =

m∑j=1

cjΓngjΓ

n : cj ∈ C, gj ∈ GSp+2n(Q)

.

We will write T for Tn except for the cases when we are moving from the space of modular

forms with degree n to degree r; in those cases we will use the exponent notation to be clear as to

which space we are working with. Further note that the above definition gives T as a C-module.

36

Page 39: HRG Master's Project

Given two double cosets

ΓngΓn =

ν(g)∐i=1

Γngi and Γng′Γn =

ν(g′)∐j=1

Γng′j

where ν(g) = |Γn\ΓngΓn| = [Γn : Γ], we have

(ΓngΓn) · (Γng′Γn) =∑

ΓnhΓn⊂ΓngΓng′Γn

c(g, g′;h)ΓnhΓn

where c(g, g′;h) is the number of pairs gi, g′j such that gig

′j ∈ Γnh. In Lemma 3.1.5 of [1], Andri-

anov proves this formula comes from wanting multiplication of double cosets to correspond with

composition of Hecke operators, that is, the desire to have

(f [ΓngΓn]k)[Γng′Γn]k = f [(ΓngΓn) · (Γng′Γn)]k.

Since we have c(g, g′;h) ∈ Z for all g, g′, h it is easy to see that we can form an R-algebra for

Z ⊆ R ⊆ C, a subring, by defining

T(R) :=

m∑j=1

aj [ΓngjΓn]k : m ∈ Z≥0, aj ∈ R, gj ∈ GSp+

2n(Q)

.

From this definition, is easy to see that taking R = Q and considering the map

T −→ T(Q)⊗Q Cm∑j=1

aj [ΓngjΓn] 7−→

m∑j=1

([ΓngjΓn]⊗ aj)

yields an isomorphism as C-algebras. Moving forward, we will use T ∈ T to denote an arbitrary

Hecke operator. That is, for T ∈ T(R) we have

T =∑j=1

aj [ΓngjΓn]k

37

Page 40: HRG Master's Project

with aj ∈ R. For T acting on a modular form, f , we write T f where

T f := f∑j=1

aj [ΓngjΓn]k :=

∑j=1

ajf [ΓngjΓn]k .

Since T is a linear operator, it is natural to ask if it has any eigenvectors; that is, are there any

f ∈ Mk(Γn) such that T f = λ(T ; f)f for some λ(T ; f) ∈ C? The short answer is yes, and of

particular interest are eigen modular forms; modular forms satisfying T f = λ(T ; f)f for all T ∈ T.

Most of the results on Hecke operators will be shown on a single arbitrary coset with the

general result being obtained by extending it linearly. Since we take g ∈ GSp+2n(Q) there exists

a ∈ Z>0 such that ag ∈ GSp+2n(Q)∩Mat2n(Z), therefore if we are working in T(R) with R ⊃ Q then

we can write ΓngΓn = 1aΓn(ag)Γn. This is desirable due to the following proposition.

Proposition 2.5.7 ([1], Lemma 3.3.6) Let γ ∈ GSp+2n(Q) ∩ Mat2n(Z), then the double coset

ΓnγΓn has a unique representative of the form

g = diag(a1, . . . , an, d1, . . . , dn)

with aj , dj ∈ Z satisfying aj > 0, ajdj = µ(γ) for all j. Furthermore an|dn, aj |aj+1 for j =

1, . . . , n− 1.

Proof Andrianov proves the result by induction on the degree n. For the case n = 1 we have

Γ1 = SL2(Z) and GSp+2 (Q) = GL+

2 (Q). Let γ =

a b

c d

∈ GSp+2 (Q) ∩Mat2(Z). We then have

µ(γ) = ad− bc. Using lemma 3.2.2 of [1] we have

Γ2γΓ2 = SL2(Z)γSL2(Z) = SL2(Z)

e1 0

0 e2

SL2(Z)

with e|e2 and note that this is the just the matrix of elementary divisors of γ over Z. Setting a1 = e1

and d1 = e2 we get a1|d1 and a1d1 = ad− bc = µ(γ), thus showing the base case.

We now assume proposition 2.5.7 holds for all degrees n < m and consider the case n = m.

38

Page 41: HRG Master's Project

To utilize the induction hypothesis we will show that

ΓnγΓn = Γn

a1 01,n−1 0 01,n−1

0n−1,1 A4 0n−1,1 B4

0 01,n−1 d1 01,n−1

0n−1,1 C4 0n−1,1 D4

Γn

where

A4 B4

C4 D4

∈ GSp+2(n−1)(Q) ∩Mat2(n−1)(Z). To do this, we first assume that γ is primitive

since if it is not that we set a1 = gcd(γ), i.e., a1 is the greatest common divisor of the entries of γ

and instead would work with γ′ where γ = a1γ′.

Let δ = δ(γ) be the minimum of the greatest common divisors coming from the columns of

γ. Let i denote the column corresponding to δ and note that we can assume i ≤ n since we would

otherwise consider γJ2n. We can further assume that i = 1 since we can use a permutation matrix

V ∈ SLn(Z) to interchange the 1st and ith columns via γ

V 0

0 tV −1

. Now that we have the first

column of γ produces the values δ, we assume δ = 1. In this case, we use Lemma 1.3.9 of [1] to pick

a representative of Γnγ such that the first column is given by t

(1 0 . . . 0

)∈ R2n. Using the

block form, we may now assume that we have

γ

1 A2 B1 B2

0n−1,1 A4 B3 B4

0 C2 D1 D2

0n−1,1 C4 D3 D4

and observe that for W =

1 −A2

0n−1,1 In−1

we have

γ

W 0n

0ntW−1

=

a1 01,n−1 B1 B2

0n−1,1 A4 B3 B4

0 C3 D1 D2

0n−1,1 C4 D3 D4

.

39

Page 42: HRG Master's Project

It is important to note that the subblocks forming the blocks B,C,D are not necessarily the same

as the ones we start with. Now assuming that we chose γ to be of the form we just obtained, we

consider the product

γ

In S

0n In

where S =

−B1 −B2

−tB2 0n−1

.

This produces a matrix of the form

1 01,n−1 0 01,n−1

0n−1,1 A4 B3 B4

0 C2 D1 D2

0n−1,1 C4 D3 D4

.

Using the relation AtB = BtA produces B3 = 0n−1,1. Similarly, tAC = tCA produces C2 = 01,n−1.

Combing these facts with the formulas AtD − BtC = µ(γ)In and tAD − tCB = µ(γ)In produces

D1 = µ(γ), D2 = 01,n−1, D3 = 0n−1,1, and

A4 B4

C4 D4

∈ Γn−1. Thus, our induction hypothesis on

the degree n would complete the proof for δ = 1. We now consider the case δ > 1. In this case, we

can follow the same construction up to a representative of the form

δ A2 B1 B2

0n−1,1 A4 B3 B4

0 C2 D1 D2

0n−1,1 C4 D3 D4

.

We then use right multiplication by the matrices U

1 −A2

0n−1,1 In−1

, T

0 −B2

−tB2 0n−1

, and T

−B1 01,n−1

01,n−1 0n−1

where the entries of A2− δA2, B2− δB2, and B1− δB1 are all in the set {1, 2, . . . , δ}. It should also

be noted that the B1 reference is the B1 block of

γU

1 −A2

0n−1,1 In−1

T

0 −B2

−tB2 0n−1

,

i.e., it is necessary to reduce A2 and B2 before B1.

40

Page 43: HRG Master's Project

We denote this new reduced matrix by γ0 and observe that since the the first entry of every

column is less than or equal to δ that we must have δ(γ0) ≤ δ(γ) = δ. If δ(γ0) = δ(γ) = δ, then

we must have δ divides all the entries of γ; however, this is a contradiction since we assumed γ is

primitive and δ > 1. We must then take 1 ≤ δ(γ0) < δ and we repeat the process on γ0. We then

iterate the process until δ(γ0) = and we apply the earlier case of δ = 1; that is, we now can pick a

representative for γ satisfying

ΓnγΓn = Γn

1 01,n−1 0 01,n−1

0n−1,1 A4 0n−1,1 B4

0 01,n−1 µ(γ) 01,n−1

0n−1,1 C4 0n−1,1 D4

Γn,

with γ∗ =

A4 B4

C4 D4

∈ GSp+2(n−1)(Q)∩Mat2(n−1)(Z) with µ(γ∗) = µ(γ). Our induction hypothesis

then gives the unique representative diag(a2, . . . , an, d2, . . . , dm) of the double coset Γn−1γ∗Γn−1

satisfying ai|ai+1, an|dn, and aidi = µ(γ∗) = µ(γ). Moreover, there exist ξ =

ξ1 ξ2

ξ3 ξ4

and

ξ′ =

ξ′1 ξ′2

ξ′3 ξ′4

such that ξγ∗ξ′ = diag(a2, . . . , an, d2, . . . , dm). We then have

1 01,n−1 0 01,n−1

0n−1,1 ξ1 0n−1,1 ξ2

0 01,n−1 µ(γ) 01,n−1

0n−1,1 ξ3 0n−1,1 ξ4

1 01,n−1 0 01,n−1

0n−1,1 A4 0n−1,1 B4

0 01,n−1 µ(γ) 01,n−1

0n−1,1 C4 0n−1,1 D4

1 01,n−1 0 01,n−1

0n−1,1 ξ′1 0n−1,1 ξ′2

0 01,n−1 µ(γ) 01,n−1

0n−1,1 ξ′3 0n−1,1 ξ′4

=

1 01,n−1 0 01,n−1

0n−1,1 diag(a2, . . . , an) 0n−1,1 0n−1

0 01,n−1 µ(γ) 01,n−1

0n−1,1 0n−1 0n−1,1 diag(d2, . . . , dn)

.

Thus, by induction we have the result for primitive γ.

Recall that if γ is not primitive then we take a1 to be the greatest common divisor of the

41

Page 44: HRG Master's Project

entries of γ. We then have γ = a1γ′ where γ′ is primitive and thus Γnγ′Γn has a unique representative

given by diag(1, a′2, . . . , a′n, µ(γ′), d′2, . . . , d

′n). We observe that µ(γ) = µ(a1γ

′) = a21µ(γ′). Combing

this with ΓnγΓn = a1Γnγ′Γn, we have the unique representative

diag(a1, a1a′2, . . . , a1a

′n, a1µ(γ′), a1d

′2, . . . , a1d

′n) = diag(a1, . . . , an, d1, . . . , dn).

We will always choose g in ΓngΓn as the unique diagonal matrix described above. Doing

so shows that ΓngΓn = Γng∨Γn where g∨ = µ(g)g−1. This comes from combining the identity

tgJg = µ(g)J with the fact that we pick g to be diagonal and so

g = tg = µ(g)Jg−1J−1 = Jg∨J−1 ∈ Γng∨Γn.

More generally we have g∨ = J tgJ−1 and (gγ)∨ = γ∨g∨. Using these relations we obtain the

following.

∐Γngi = ΓngΓn

= Γng∨Γn

= (ΓngΓn)∨

= Jt(ΓngΓn)J−1

= JΓntgΓnJ−1

=∐

JΓngiΓ−1

=∐

(Γngi)∨

=∐

g∨i (Γn)∨

=∐

g∨i Γn.

Since∐

Γngi = ΓngΓn = Γng∨Γn =∐

g∨j Γn, we have

∑f |kgi =

∑f |kg∨j . (2.1)

42

Page 45: HRG Master's Project

Recall that the summation is finite. It is useful to see the action of the slash operator

associated with g∨i . The action is given by

(f |kg∨i )(Z) = j(g∨i , Z)−kµ(g∨i )nk−n(n+1)

2 f(g∨i Z)

= µ(gi)−nkj(g−1

i , Z)−kµ(gi)nk−n(n+1)

2 f(µ(gi)In · (g−1i Z))

= j(g−1i , Z)−kµ(gi)

−n(n+1)2 f(g−1

i Z). (2.2)

We could further analyze this function; however, the current formulation is sufficient for showing

that the Hecke operators are hermitian with respect to the Petersson inner product. In showing this,

we will make the change of variable ωi = giZ when necessary and will write Y ∗i = Im(giZ) = Im(ωi).

Note that

〈f1 [ΓngΓn] , f2〉 =⟨∑

f1|kgi, f2

⟩=

∑〈f1|kgi, f2〉

=∑ ∫

Γn\Hn

det(Y )k (f1|kgi) (Z)f2(Z)dνn(Z)

=∑ ∫

Γn\Hn

det(Y )kµ(gi)nk−n(n+1)

2 j(gi, Z)−kf1(giZ)f2(Z)dνn(Z)

=∑ ∫

Γn\Hn

j(gi, Z)k det(Y ∗i )kµ(gi)−n(n+1)

2 f1(giZ)f2(Z)dνn(Z)

=∑ ∫

Γn\giHn

j(gi, g−1i ωi)

k det(Y ∗i )kµ(gi)−n(n+1)

2 f1(ωi)f2(g−1i ωi)dνn(g−1

i ωi)

=∑ ∫

Γn\Hn

det(Y ∗i )kf1(ωi)µ(gi)−n(n+1)

2 j(g−1i , ωi)−kf2(g−1

i ωi)dνn(ωi)

=∑ ∫

Γn\Hn

det(Y ∗i )kf1(ωi)(f2|kg∨i ) (ωi)dνn(ωi) by (2.2)

=∑〈f1, f2|kg∨i 〉

=⟨f1,∑

f2|kg∨i⟩

= 〈f1, f2 [ΓngΓn]〉 due to (2.1).

43

Page 46: HRG Master's Project

To obtain the fifth equality we use the identity

t(CiZ +Di)Y

∗i (CiZ +Di) = µ(gi)Y.

Taking the derivative of both sides then produces

det(Y )k = µ(gi)−nk det(Y ∗)kj(gi, Z)kj(gi, Z)k.

To obtain the seventh equality we use the identity

j(γδ, Z) = j(γ, δZ)j(δ, Z)

which allows us to write the following.

j(gi, g−1i ωi)

k = j(gi, g−1i ωi)

k

=j(gig

−1i , ωi)

k

j(g−1i , ωi)k

=j(I2n, ωi)

k

j(g−1i , ωi)k

= j(g−1i , ωi)

−k.

Thus we see that any Hecke operator represented by a single coset is hermitian. The result extends

to all Hecke opertors due to linearity.

Since the Hecke operators are Hermitian with respect to the Petersson inner product on

cusp forms, a finite dimensional C-vector space, we know from linear algebra that Sk(Γn) then has

a basis of orthogonal eigenforms.

Another reason to choose g to be diagonal, is we can then write it as a finite product of

diagonal matrices,∏p gp, where each gp has diagonal entries restricted to powers of p. Combining

this with the following proposition allows us to write

ΓngΓn =∏p

ΓngpΓn.

44

Page 47: HRG Master's Project

Proposition 2.5.8 ([1], Proposition 3.3.9) Let M , M ′ ∈ Γn. Suppose that the quotients

gcd

(a1(M)

d1(M),a1(M ′)

d1(M ′)

)= 1,

then

(ΓnMΓn) (ΓnM ′Γn) = ΓnMM ′Γn.

To clarify, the notation a1(M) and d1(M) reference, respectively, the first and the (n+ 1)st

diagonal entries of the unique representative of the double coset ΓnMΓn from Proposition 2.5.7. It

is also important to note that Andrianov proves this for arbitrary level, Γn(N), and in that case the

double cosets are of the form Γn0 (N)MΓn0 (N) where

Γn0 (N) =

A B

C D

∈ Γn : C ≡ 0 (mod N)

⊇ Γn(N).

We note that for full level we have Γn0 (1) = Γn = Γn(1). We can apply this proposition to our case

since the entries of gp and gp′ are coprime for p 6= p′. We can use this relation to identify

ΓngΓn =∏p

ΓngpΓn ←→

⊗p prime

ΓngpΓn

where we gp is defined as the gp in the finite product for p|µ(g) and I2n for p - µ(g). Through this

mapping we obtain the decomposition

T(Q) =⊗p

Tp,

where Tp is the set of formal sums with coefficients in Q of double cosets ΓngpΓn satisfying µ(gp) = pn

for some n ∈ Z. By Theorem 3.3.23 of [1] we know that Tp is generated by elements of the form

45

Page 48: HRG Master's Project

T (p) = T (n)(p) = Γn

In 0

0 pIn

Γn,

Ti(p2) = T

(n)i = Γn

In−i 0 0 0

0 pIi 0 0

0 0 p2In−i 0

0 0 0 pIi

Γn

for i = 1, . . . , n, and T(n)n (p2)−1 = [ΓnpI2nΓn]

−1which we show is in T by showing it is in T(Q).

Working over Q we see that

ΓnI2nΓn = Γn(pI2n)(p−1I2n)Γn

= ΓnpI2nΓnΓnp−1I2nΓn

= Tn(p2)Γnp−1I2nΓn.

We are able to split the double coset by Proposition 2.5.8 since

gcd

(a1(pI2n)

d1(pI2n)

a1(p−1I2n)

d1(p−1I2n)

)= gcd

(p

p,p−1

p−1

)= gcd(1, 1) = 1.

This shows that over Q we have

Tnn (p2)−1 = Γnp−1I2nΓn = p−1ΓnI2nΓn ∈ T(Q).

We now consider Theorem 2 of [26] which tells us that there exists an epimorphism from

C[T (p), T(n)1 (p2), . . . , T

(n)n (p2)] → C[T (p), T

(n−1)1 (p2), . . . , T

(n−1)n−1 (p2)] given by T 7→ T ∗ where T is

Hecke operator satisfying ΦT = T ∗Φ. Taking [ΓngpΓn] ∈ Tp we then have

[ΓngpΓn] = p−` [ΓngpΓ

n]

where p` is the largest power of p occuring in the denominator of the entries of gp and gp = p`gp. If

46

Page 49: HRG Master's Project

gp = I2n then we have

Φ(f [ΓngpΓ

n]k)

= Φ(p−`f [ΓnI2nΓn]k

)= p−`Φ (f [ΓnI2nΓn]k) = p−` (Φ(f))

[Γn−1I2(n−1)Γ

n−1]k

since Φ is C-linear, Mk(Γn) is invariant under [ΓnI2nΓn]k, and Mk(Γn−1) is invariant under[Γn−1I2(n−1)Γ

n−1]k. If gp 6= I2n then it is of a form on which we can apply Theorem 3.3.23 of [1]

to get that it is generated by elements of the form T (p) and T(n)i (p). Thus the linearity of Φ gives

the result for single cosets as well as extends to formal sums of cosets, thereby giving the following

proposition.

Proposition 2.5.9 There exists a mapping T 7→ T ∗ of T(n)p to T(n−1)

p with the following properties.

• Φ(T f) = T ∗Φ(f) for all f ∈Mk(Γn) and T ∈ T(n)p ,

• the mapping is an epimorphism.

Since for a general diagonal matrix g as described in Proposition 2.5.7 we have the the

decomposition ΓngΓn =∏p ΓngpΓ

n, we also have the expression

f [ΓngΓn]k = f

[∏p

ΓngpΓn

]k

=((f [Γngp1Γn]k

)· · ·)

[ΓngpmΓn]k .

We only concern ourselves with the full set of distinct primes, p1, . . . , pm, occuring in the entries of

g since for p not in that list gp = I2n and acts trivially on the modular form. Using Proposition

2.5.9, let T ∗p be the element of T(n−1)p such that

Φ(f [ΓngpΓ

n]k)

= T ∗p (Φ(f)) .

Using this we have

Φ (f [ΓngΓn]k) = Φ(((

f [Γngp1Γn]k)· · ·)

[ΓngpmΓn]k)

= T ∗pm(Φ(((

f [Γngp1Γn]k)· · ·) [

Γngpm−1Γn]k

))...

= T ∗pm(· · ·(T ∗p1 (Φ(f))

))= (T ∗pm · · · T

∗p1) (Φ(f)) .

47

Page 50: HRG Master's Project

Since Φ is a linear operator this result extends linearly to all of T(n)(Q). Furthermore, since the ring

we are looking at the algebra over only determines the coefficients of the formal sums, the linearity

of Φ shows that the result holds for T(n)(R) for Q ⊆ R ⊆ C. That is to say, for all T ∈ T(n)(R)

there exists T ∗ ∈ T(n−1)(R) such that Φ(T f) = T ∗(Φ(f)). We can extend this result via induction

to obtain Φ`(T f) = T ∗(Φ`(f)) for ` ≥ 0 where T ∈ T(n)(R) and T ∗ ∈ T(`)(R). This is necessary to

prove the following theorem.

Theorem 2.5.10 ([12],Theorem 1) Let f ∈ Mk(Γr) be an eigenform for some fixed r ≥ 0 and

even k > n+ r + 1 with n ≥ r. Then [f ]nr ∈Mk(Γn) is also an eigenform.

Proof We first claim that En,j,k is stable under T(n). Observe that if g ∈ Ej,j,k, then f ∈ Sk(Γj).

Since Sk(Γj) has a basis of eigenforms. Expressing f as linear combination of the basis shows that

T f ∈ Sk(Γj) for all T ∈ T(r). We now suppose for induction that the result hold for f ∈ Er,j,k for

r < n < 12k and consider the case f ∈ En,j,k.

Consider T T(n). We then have T f = [gj ]nj + f0 for some gj ∈ Sk(Γj) and f0 orthogonal to

En,j,k. We then consider Φ(T f) = [gj ]n−1j + Φ(f0) and note that it is equal to

Φ(T f) = T ∗Φ(f) = [hj ]n−1j

since Φ(f) ∈ En−1,j,k which is stable under T(n−1) by our induction hypothesis. Combining the

equations yields 0 = [gj ]n−1j − [hj ]

n−1j + Φ(f0) = [gj − hj ]n−1

j + Φ(f0). Since f0 is orthogonal to

En,j,k, we have Φ(f0) is orthogonal to En−1,j,k. Thus we must have gj = hj and Φ(f0) = 0. Since we

now have f0 ∈ Sk(Γn), it suffices to also show that f0 ∈ Sk(Γn)⊥ to obtain En,j,k is stable under

T(n). To see this, consider any ψ ∈ Sk(Γn) and observe that 〈T f, ψ〉 = 〈f, T ψ〉 = 0 since Hecke

operators are hermitian, Sk(Γn) is stable under T(n), and En,j,k is orthogonal to Sk(Γn). We then

use this to obtain

0 = 〈T f, ψ〉 = 〈[gj ]nj , ψ〉+ 〈f0, ψ〉 = 〈f0, ψ〉

for all ψ ∈ Sk(Γn) since 〈[gj ]nj , ψ〉 = 0. Thus, f0 ∈ Sk(Γn)⊥ but since it is also a cusp form we must

have f0 = 0 and T f = [gj ]nj ∈ En,j,k. Thus by induction, we have En,j,k is stable under T(n) for all

j ≤ n < 12k.

We now consider f ∈ Mk(Γr) satisfying the assumptions of the theorem. We then have

f =∑rj=0[fj ]

rj and consequently [f ]nr =

∑rj=0[fj ]

nj . Since En,j,k is stable under T(n), for each j

48

Page 51: HRG Master's Project

there exists some gj satisfying T [fj ]nj = [gj ]

nj , yielding

T [f ]nr =

r∑j=0

[gj ]nj and Φn−r(T [f ]nr ) =

r∑j=0

[gj ]rj .

We also have

Φn−r(T [f ]nr ) = T ∗Φn−r([f ]nr ), T ∗ ∈ T(r)

= T ∗f

= λ(T ∗; f)f

= λ(T ∗; f)r∑j=0

[fj ]rj

=

r∑j=0

[λ(T ∗; f)fj ]rj .

Combining these expressions and using the linearity of [·]rj yields

0 =

r∑j=0

[gj − λ(T ∗; f)fj ]rj .

Applying Φr to both sides of the expression yields 0 = gr − λ(T ∗; f)fr, that is, gr =

λ(T ∗; f)fr and 0 =∑r−1j=0[gj − λ(T ∗; f)fj ]

rj . Applying Φr−1 then produces gr−1 = λ(T ∗; f)fr−1.

Continuing this process, we obtain gj = λ(T ∗; f)fj and consquently

T [f ]nr =

r∑j=0

[gj ]nj =

r∑j=0

[λ(T ∗; f)fj ]nj = λ(T ∗; f)

r∑j=0

[fj ]nj = λ(T ∗; f)[f ]nr .

Therefore [f ]nr is an eigenform with λ(T ; [f ]nr ) = λ(T ∗; f) for every T ∈ T(n). �

To make use of this result we need the following theorem.

Theorem 2.5.11 ([13], Theorem 1) Let f ∈ Sk(Γn) be an eigenform. The field Kf = Q({λ(T ; f) :

T f = λ(T ; f)f, T ∈ T(n)}) is a totally real number field.

Proof Let T(n) = HomC(T(n),C). For λ ∈ T(n) we define

Mk(Γn;λ) = {f ∈Mk(Γn) : T f = λ(T ; f)f}

49

Page 52: HRG Master's Project

and

m(λ) = dimC (Mk(Γn;λ)) ≥ 0.

Let Λ(T) = {λ ∈ T : m(λ) ≥ 1}. Then we have M = #Λ(T) ≤ dimC (Mk(Γn)). Write Q(λ) =

Q(λ(T(Q))). Now take σ ∈ Aut(C) and for every g =∑T≥0 a(T ; g)qT ∈ Mk(Γn) set σ(g) =∑

T≥0 σ(a(T ; g))qT . We know from Section 2.4 that Mk(Γn)Q ⊗Q C ∼= Mk(Γn), thus we identify

g ∈Mk(Γn) with∑i gi ⊗ ci ∈Mk(Γn)Q ⊗Q C. Since Q is fixed by every field automorphism of C

we have

σ(g) = σ(∑

gi ⊗ ci)

with ci ∈ C, gi ∈Mk(Γn)Q

=∑

σ(gi)⊗ σ(ci)

=∑

gi ⊗ σ(ci) ∈Mk(Γn)Q ⊗Q C ∼= Mk(Γn).

For λ ∈ T we also define σ(λ)(T ) = σ(λ(T )) for T ⊗ 1 ∈ T(Q)⊗Q C ∼= T. It is also known

from Equation 2.4 of [27] that Mk(Γn)Q is T(Q)-stable. Consequently, for g ∈Mk(Γn;λ) we have

T σ(g) = T σ(∑

gi ⊗ ci)

with ci ∈ C, gi ∈Mk(Γn)Q

= T(∑

gi ⊗ σ(ci))

=∑T gi ⊗ σ(ci)

= σ(∑

T gi ⊗ ci)

= σ(T∑

gi ⊗ ci)

= σ(T g)

= σ(λ(T ; g)g)

= σ(λ)(T )σ(g).

We see that if g ∈ Mk(Γn;λ) then σ(g) ∈ Mk(Γn;σ(λ)). Thus, for any σ ∈ Aut(C)

and λ ∈ Λ(T) then σ(λ) ∈ Λ(T) and σ(Q(λ)) = Q(σ(λ)). Since the automorphisms of C act as

permutations on the elements of Λ(T), which is finite, we see that Q(Λ(T))/Q is Galois, where

Q(Λ(T)) is the composite field. Furthermore, EndQ(Q(λ)) ⊆ Gal(Q(Λ(T))/Q) ⊂ Sym(m), yielding

#EndQ(Q(λ)) ≤ #Gal(Q(Λ(T))/Q) ≤M !

50

Page 53: HRG Master's Project

showing that Q(λ) is a finite extension.

To get Q(λ) ⊆ R we recall that all T ∈ T are hermitian with respect to the Petersson inner

produce and thus have only real eigenvalues.

Lastly, for f ∈ Sk(Γn) an eigenform we have λ(T ; f) ∈ Λ(T) and thus have Kf :=

Q({λ(T ; f)}) is a totally real number field. �

We conclude that by stating that it is known by Theorem A(n,r) (found in Appendix A) of

[18] that if f ∈ Sk(Γr) is an eigenform then [f ]nr ∈Mk(Γn)Kf.

51

Page 54: HRG Master's Project

Chapter 3

Results of T. Kikuta and S.

Takemori

3.1 Definitions, Remarks, and Main Results

Let K be an algebraic number field and O = OK its ring of integers. For a prime ideal

p ⊂ O, let Op be the localization of O at p.

Definition 3.1.1 Let F ∈Mk(Γn)Op . We say F is a mod pm cusp form if Φ(F ) ≡ 0 mod pm.

Main Theorem There exists a finite set Sn(K) of prime ideals in K depending on n such that

the following holds. For a prime ideal p of O not contained in Sn(K) and a mod pm cusp form

F ∈Mk(Γn)Opwith k > 2n, there exists G ∈ Sk(Γn)Op

such that F ≡ G mod pm.

Note: The set Sn(K) is defined at the end of Section 3.2. It is presented along with proofs

of some of its properties. We also note that the theorem is trivial for odd weight k because there

are no non-cusp forms of odd weight.

Letting νp be the normalized additive valuation with respect to p, i.e., νp(p) = 1. For

F ∈Mk(Γn)K we define the following;

νp(F ) := min{νp(a(T ;F )) : T ∈ Λn}

ν(n′)p (F ) := min{νp(a(T ;F )) : T ∈ Λn, rank(T ) = n′} for 0 ≤ n′ ≤ n.

52

Page 55: HRG Master's Project

Main Corollary Let k > 2n be even and f ∈ Sk(Γr)Kf, n > r, a Hecke eigenform. For the

Klingen-Eisenstein series [f ]nr attached to f , suppose there exists a prime ideal p in OKfwith p 6∈

Sn(Kf ) such that ν(n)p ([f ]nr ) = νp(Φ([f ]nr ))−m, for some m ∈ Z≥1. Then there exists F ∈ Sk(Γn)Op

such that α[f ]nr ≡ F mod pm for some 0 6= α ∈ pm.

3.2 Lemmas and Their Proofs

By Theorem 2.3 of [6, pp. 150] we know that⊕1≤k

Mk(Γn)Z is finitely generated; that is,

there exists a set of generators {ψ1, · · · , ψt} such that

⊕k≥1

Mk(Γn)Z ∼= Z[ψ1, · · · , ψt]

as Z-algebras. By Corollary 2.4.7, we have that Mk(Γn)Z(p)

∼= Mk(Γn)Z ⊗Z Z(p) for any prime p.

Combining these facts we get

⊕k≥1

Mk(Γn)Z(p)=

⊕k≥1

(Mk(Γn)Z ⊗Z Z(p)

)

=

⊕k≥1

Mk(Γn)Z

⊗Z Z(p)

= Z[ψ1, · · · , ψt]⊗Z Z(p)

= Z(p)[ψ1, · · · , ψt].

This allows us to introduce the following lemma.

Lemma 3.2.1 Let M be a natural number. For any prime p, the algebra⊕k>M

Mk(Γn)Z(p)is finitely

generated over Z(p).

Proof Let {ψ1, · · · , ψt} be the set of generators of the Z(p)-algebra⊕

kMk(Γn)Z(p), where ψi ∈

Mki(Γn)Z(p)

. For each i, let αi ∈ Z≥0 be the smallest integer such that the weight of ψαii , αiki, is

strictly greater than M . We prove that⊕k>M

Mk(Γn)Z(p)is generated over Z(p) by monomials of the

form

ψα11 , · · · , ψαt

t , (3.1)

ψi11 · · ·ψitt (i1k1 + · · ·+ itkt > M, 0 ≤ ij < 2αj). (3.2)

53

Page 56: HRG Master's Project

Note that any F ∈ Mk(Γn)Z(p)can be written as a linear combination of monomials

ψa11 · · ·ψatt with i1k1 + · · ·+ itkt = k. Thus it suffices to show that F = ψa11 · · ·ψ

att can be generated

by monomials of the form (3.1) and (3.2) for k > M .

For each ai we can write ai = qiαi + ri for qi, ri ∈ Z with 0 ≤ ri < αi. Thus we write

F = ψa11 · · ·ψatt = (ψr11 · · ·ψ

rtt )

t∏i=1

(ψαii )

qi .

If r1k1 + · · · rtkt > M then we are done, so we suppose the contrary. Then there must be some

qj ≥ 1, otherwise we would have

a1k1 + · · · atkt = r1k1 + · · · rtkt ≤M

contradicting F ∈⊕k>M

Mk(Γn)Z(p). Thus we rewrite our expression of F as

F = (ψr11 · · ·ψrj−1

j−1 ψrj+αj

j ψrj+1

j+1 · · ·ψrtt )(ψαj

j

)qj−1∏i 6=j

(ψαii )

qi .

It is clear that(ψαj

j

)qj−1∏i 6=j (ψαi

i )qi is generated by monomials of the form (3.1), so it remains to

show that ψr11 · · ·ψrj−1

j−1 ψrj+αj

j ψrj+1

j+1 · · ·ψrtt is of the form (3.2).

First, each ri 6= rj satisfies 0 ≤ ri < αi < 2αi. For rj we have 0 ≤ rj +αj < αj +αj = 2αj .

Thus each individual exponent satisfies its respective bounds. Now observe that the weight of the

monomial is given by

(rj + αj)kj +∑i 6=j

riki = αjkj +

t∑i=0

riki

≥ αjkj

> M

by the definition of αj . Thus ψr11 · · ·ψrj−1

j−1 ψrj+αj

j ψrj+1

j+1 · · ·ψrtt is of the form (3.2) and consequently

F is generated by a finite product of monomials of the form of (3.1) and (3.2). �

Lemma 3.2.2 Let M be a natural number. Assume that⊕k>M

Mk(Γn)Z(p)= Z(p)[f1, · · · , fs] with

fi ∈Mki(Γn)Z(p)

. Then we have⊕k>M

Mk(Γn)Op = Op[f1, · · · , fs] for any prime ideal p above p.

54

Page 57: HRG Master's Project

Proof Let Z(p)[x1, . . . , xs] be a polynomial ring and let I be the kernel of the Z(p)-algebra morphism

ϕ : Z(p)[x1, · · · , xs]→ Z(p)[f1, · · · , fs]

defined by ϕ(xi) = fi for 1 ≤ i ≤ s. By Corollary 2.4.8 we have Mk(Γn)Z(p)⊗Z(p)

Op ∼= Mk(Γn)Op ,

thus it follows that

⊕k>M

Mk(Γn)Op∼=

⊕k>M

(Mk(Γn)Z(p)

⊗Z(p)Op)

=

(⊕k>M

Mk(Γn)Z(p)

)⊗Z(p)

Op

∼= (Z(p)[x1, · · · , xs]/I)⊗Z(p)Op

∼= Op[x1, · · · , xs]/IOp

∼= Op[f1, · · · , fs].�

Lemma 3.2.3 For k > 2n, the restricted Siegel Φ-operator ΦK : Mk(Γn)K → Mk(Γn−1)K is

surjective.

Proof To get this result, we utilize the fact that C is faithfully flat over any number field K. By

Theorem 7.2 of [16] any ring S ⊃ K is said to be faithfully flat if given any K-module, M , we have

M⊗KS 6= 0. Since K is a number field, we have C ⊃ K. Furthermore, K is a field so any K-module,

M , is actually a K-vector space. Therefore M ⊗K C is an extension of scalars and thus a non-trivial

C vector space. Since C is faithfully flat over K, we know that any sequence of K-modules is exact

if and only if the corresponding sequence obtained by tensoring with C is exact. We now consider

the following diagram.

0 −→ Sk(Γn)id−→ Mk(Γn)

Φ−→ Mk(Γn−1) −→ 0

∼=↓ ∼=↓ ∼=↓

Sk(Γn)K ⊗K C id⊗1−→ Mk(Γn)K ⊗K C ΦK⊗1−→ Mk(Γn−1)K ⊗K C

We know the first sequence is short exact and wish to get the second one is as well by showing the

55

Page 58: HRG Master's Project

two sequences are isomorphic. This amounts to showing

Mk(Γn)Φ−→ Mk(Γn−1)

∼=↓ ∼=↓

Mk(Γn)K ⊗K C ΦK⊗1−→ Mk(Γn−1)K ⊗K C

commutes. To see this, we consider a basis {f1, · · · , fm} of Mk(Γn)K and with the standard identi-

fication obtain {f1⊗1, · · · , fm⊗1} is a basis of Mk(Γn)K⊗K C. Thus we see that {f1, · · · , fm} also

serves as a basis of Mk(Γn). Since linear maps of vector spaces are completely determined by how

they act on a basis we will show that diagram commutes on an arbitrary basis element fi. Going

down and across we obtain the following composition of mappings

fi 7−→ fi ⊗ 1 7−→ ΦK(fi)⊗ 1.

Going across and then down we obtain the composition given by

fi 7−→ Φ(fi) 7−→ Φ(fi)⊗ 1.

Since Φ essentially strips away the Fourier coefficients of fi with full rank, we have Φ(fi) ∈

Mk(Γn−1)K and more importantly Φ(fi) = ΦK(fi). Hence we have the diagram commutes on

a basis and hence every element. Moreover, we now have that the two sequences are isomorphic,

proving that the following sequence is exact

0 −→ Sk(Γn)K ⊗K C id⊗1−→ Mk(Γn)K ⊗K C ΦK⊗1−→ Mk(Γn−1)K ⊗K C −→ 0.

Since we know C is faithfully flat over K we then have

0 −→ Sk(Γn)Kid−→ Mk(Γn)K

ΦK−→ Mk(Γn−1)K −→ 0.

Therefore, we conclude that the restricted Siegel operator ΦK : Mk(Γn)K →Mk(Γn−1)K is surjec-

tive. �

In order to define Sn(K), we first make the two following oberservations. First, by Lem-

56

Page 59: HRG Master's Project

mas 3.2.1 and 3.2.2 we have⊕

2n<k

Mk(Γn−1)Op is a finitely generated Op-algebra. Restricting

to⊕

2n<k∈2ZMk(Γn−1)Op , we find that we still have a finitely generated Op-algebra though the

number of generators may increase, i.e., if f1 and f2 are both generators with odd weights k1

and k2 then they may need to be replaced with f21 , f1f2, and f2

2 having respective weights 2k1,

k1 + k2, and 2k2. The other observation is that given a set of generators {f1, . . . , fs} of the algebra⊕2n<k∈2ZMk(Γn−1)Op , Lemma 3.2.3 tells us that Φ−1

K (fi) 6= ∅ for 1 ≤ i ≤ s.

Definition 3.2.4 Let Sn(K) be the set of all primes ideals of p ⊂ OK such that given a set of

generators {f1, . . . , fs} of⊕

2n<k∈2ZMk(Γn−1)Op , there exists some i such that for all Fi ∈ Φ−1

K (fi)

we have νp(Fi) < 0.

Lemma 3.2.5 Whether or not p ∈ Sn(K) depends on n but not on the set of generators of⊕2n<k∈2Z

Mk(Γn−1)Op .

Proof Suppose for some p ⊂ O we have two sets of generators, {f1, . . . , fs} and {g1, . . . , gr}, of⊕2n<k∈2Z

Mk(Γn−1)Op . It suffices to show that if for 1 ≤ i ≤ s we take Fi ∈ Mk(Γn)Op with

Φ(Fi) = fi, then for every 1 ≤ j ≤ r we can get Gj ∈ Mk(Γn)Op such that Φ(Gj) = gj . To see

this observe that gj ∈⊕

2n<k∈2ZMk(Γn−1)Op = Op [f1, . . . , fs] means there exists a polynomial

Pj ∈ Op[x1, . . . , xs] such that

gj = Pj(f1, . . . , fs).

Taking Fi ∈Mk(Γn)Op with Φ(Fi) = fi, set

Gj = Pj(F1, . . . , Fs).

Recall the limit of a product of functions if the product of the limits, provided the limit of each

function exists. Keeping this in mind, we observe that

Φ(f · g) = limt→∞

fZ

it

g

Zit

=

limt→∞

f

Zit

limt→∞

g

Zit

= (Φ(f))(Φ(g)).

57

Page 60: HRG Master's Project

Since Pj(x1, . . . , xs) is a polynomial in s variables we have

Φ(Gj) = Pj(Φ(F1), . . . ,Φ(Fs)) = P (f1, . . . , fs) = gj .

We now show that if there exists some i such that νp(Gi) < 0 for any choice of Gi ∈ Φ−1K (gi) then

there must be some j such that νp(Fj) < 0 for all Fj ∈ Φ−1K (fj).

Since νp is an additive valuation and Pi(x1, . . . , xs) is a finite sum of monomials in s vari-

ables, one of the monomials, say αs∏`=1

xe`` with 0 6= α ∈ Op, must satisfy

νp (P (F1, . . . , Fs)) ≥ νp

s∏`=1

F e``

).

Consequently, we have

0 > νp(Gi)

≥ νp (P (F1, . . . , Fs))

= νp

s∏`=1

F e``

)

= νp +

s∑`=1

e`νp(F`)

≥s∑`=1

e`νp(F`) since α ∈ Op.

By definition of a polynomial, we must have e` ≥ 0 for all `. Hence there must exist some j such

that νp(Fj) < 0. Since Fj ∈ Φ−1K (fj), the only way to choose a different form Fj ∈ Φ−1

K (fj) is if we

have Fj = Fj + H for some H ∈ Sk(Γn), and hence Φ(Fj) = Φ(Fj) = fj . Using this relation we

observe that

Φ(Pi(F1, . . . , Fj−1, Fj , Fj+1, . . . , Fs)) = Pi(Φ(F1), . . . ,Φ(Fj−1),Φ(Fj),Φ(Fj+1), . . . ,Φ(Fs))

= Pi(f1, . . . , fj−1, fj , fj+1, . . . , fs)

= gi.

58

Page 61: HRG Master's Project

Thus Pi(F1, . . . , Fj−1, Fj , Fj+1, . . . , Fs) ∈ Φ−1K (gi) and thus we have

νp

(Pi(F1, . . . , Fj−1, Fj , Fj+1, . . . , Fs)

)< 0.

If we are able to choose Fj such that νp(Fj) ≥ 0 then we do so and there exists another form in

the set {F1, . . . , Fs} with negative valuation. In that case we would repeat this process. In a worst

case scenario, we would repeat this process no more than s− 1 times in order to isolate a form with

negative valuation which cannot be replaced with another that has non-negative valuation. Thus

without loss of generality we can assume that our first choice Fj cannot be replaced by a form with

non-negative valuation. That is, any choice of Fj ∈ Φ−1K (fj) satisfies νp(Fj) < 0. Thus we see that

p ∈ Sn(K) is independent of the set of generators. �

Lemma 3.2.6 For d = [K : Q] <∞ we have

{p : p ∩ Z = (p) ∈ Sn(Q), p - d} ⊆ Sn(K) ⊆ {p : p ∩ Z = (p) ∈ Sn(Q)}.

Proof We start by proving the second containment. Let p ∈ Sn(K) and {f1, . . . , fs} be a set

of generators of⊕

2n<k∈2ZMk(Γn−1)Z(p). By Lemma 3.2.2 we have

⊕2n<k∈2Z

Mk(Γn−1)Op =

Op[f1, . . . , fs]. By definition, since p ∈ Sn(K), there exists an i with 1 ≤ i ≤ s such that for

all Fi ∈ Φ−1K (fi) we have νp(Fi) < 0. In particular, for all Fi ∈ Φ−1

Q (fi) ⊆ Φ−1K (fi) we have

ν(p)(Fi) ≤ νp(Fi) < 0. Thus (p) ∈ Sn(Q) and p ∈ {p : p ∩ Z = (p) ∈ Sn(Q)}.

We now prove the first containment. Let {f1, . . . , fs} be as before and further suppose that

p ∩ Z = (p) ∈ Sn(Q) with p - d = [K : Q]. Now suppose for contradiction that for every 1 ≤ i ≤ s

there exists Fi ∈ Φ−1K (fi) such that νp(Fi) ≥ 0. Consider the forms given by

Gi =∑

σ∈Emb(K,C)

Fσi ∈Mk(Γn)Q.

Since νp(Fi) ≥ 0 we have ν(p)(Gi) ≥ 0 and thus Gi ∈Mk(Γn)Z(p). Furthermore, since any embed-

ding of K into C fixes Q and since we have fi ∈ Mk(Γn)Z(p)⊂ Mk(Γn)Q we obtain Φ(Fσi ) = fi.

Thus

Φ(Gi) = Φ

∑σ∈Emb(K,C)

Fσi

=∑

σ∈Emb(K,C)

Φ(Fσi ) =∑

σ∈Emb(K,C)

fi = dfi.

59

Page 62: HRG Master's Project

It follows that d−1Gi ∈ Φ−1

Q (fi). By our initial assumption, p - d so ν(p)(d−1Gi) ≥ 0, contradicting

that p ∩ Z = (p) ∈ Sn(Q). �

3.3 Proof of the main theorem and its corollary

We now have the necessary information to prove Theorem 3.1.2, the main theorem of [9].

We recall the statement is as follows.

Main Theorem There exists a finite set Sn(K) of prime ideals in K depending on n such that

the following holds. For a prime ideal p of O not contained in Sn(K) and a mod pm cusp form

F ∈Mk(Γn)Opwith k > 2n, there exists G ∈ Sk(Γn)Op

such that F ≡ G mod pm.

Proof Recall that the case of odd weight is trivial; thus we focus on even weight. Let {f1, . . . , fs}

be a set of generators of⊕

2n<k∈2ZMk(Γn−1)Op . Suppose that F ∈Mk(Γn)Op is a mod pm cusp

form, that is,

Φ(F ) ≡ 0 mod pm.

It follows by definition of νp that there exists some γ ∈ pm such that γ−1Φ(F ) ∈ Mk(Γn−1)Op .

Note that there exists some T0 ∈ Λn−1 such that a(T0; Φ(F )) = νp(Φ(F )), in which case we can

take γ = a(T0; Φ(F )). Since

γ−1Φ(F ) ∈Mk(Γn−1)Op ⊂⊕

2n<k∈2ZMk(Γn−1)Op

there exists a polynomial P ∈ Op[x1, . . . , xs] such that

γ−1Φ(F ) = P (f1, . . . , fs).

Since p ∈ Sn(K), for every 1 ≤ i ≤ s there exists some Fi ∈ Φ−1K (fi) such that νp(Fi) ≥ 0. Thus

F ∈ γP (F1, . . . , Fs) + ker ΦK

and specifically, there exists some G ∈ ker ΦK = Sk(Γn)K such that

F = γP (F1, . . . , Fs) +G.

60

Page 63: HRG Master's Project

In particular, since Fi ∈ Mk(Γn)Op for all i, we then have P (F1, . . . Fs) ∈ Mk(Γn)Op . Since

F ∈Mk(Γn)Op we then obtain G ∈ ker ΦK = Sk(Γn)Op ⊂ Sk(Γn)K . Therefore we have

F = γP (F1, . . . , Fs) +G

≡ 0 +G mod pm

≡ G mod pm. �

With the main theorem of [9] proved, we move onto their main corollary.

Main Corollary Let k > 2n be even and f ∈ Sk(Γr)Kf, n > r, a Hecke eigenform. For the

Klingen-Eisenstein series [f ]nr attached to f , suppose there exists a prime ideal p in OKfwith

p 6∈ Sn(Kf ) such that ν(n)p ([f ]nr ) = νp(Φ([f ]nr )) − m, for some m ∈ Z≥1. Then there exists F ∈

Sk(Γn)Of,p, where Of,p is the localization of OKf

at p, such that α[f ]nr ≡ F mod pm for some

0 6= α ∈ pm.

Proof Let f ∈ Sk(Γr)Kf, then by [18] we know [f ]nr ∈Mk(Γn)Kf

and [f ]n−1r ∈Mk(Γn−1)Kf

. We

further suppose that p /∈ Sn(Kf ) such that ν(n)p ([f ]nr ) = νp(Φ([f ]nr ))−m, for some constant m ∈ Z≥1.

By Corollary 2.4.3, there exists some c ∈ Z>0 such that c[f ]n−1r ∈ Mk(Γn−1)OKf

. It is acceptable

to weaken the choice of c so that we have c[f ]n−1r ∈ Mk(Γn−1)O

f,pwhere Of,p is the localization

of OKfat p. Either way, we obtain

νp(c[f ]n−1r ) = νp(cΦ([f ]nr )) ≥ 0.

Moreover, we have

ν(n)p (c[f ]nr ) = νp(cΦ([f ]nr ))−m ≥ −m.

We now take some 0 6= β ∈ pm and observe that

ν(n)p (βc[f ]nr ) = νp(β) + ν

(n)p (c[f ]nr ) = νp(β) + νp(cΦ([f ]nr ))−m ≥ m−m = 0.

We now set α = βc, noting that we have α ∈ pm, and observe that α[f ]nr ∈Mk(Γn)Of,p

. Further-

more, we have

νp(Φ(α[f ]nr )) = νp(βc[f ]n−1r ) = νp(β) + νp(c[f ]n−1

r ) ≥ m.

61

Page 64: HRG Master's Project

This tells us that α[f ]n−1r vanishes (mod pm), demonstrating that α[f ]nr ∈ Mk(Γn)O

f,pis a

(mod pm) cusp form. We now apply the Main Theorem to get F ∈ Sk(Γn)Of,p

such that

α[f ]nr ≡ F (mod pm).

3.4 Numerical Examples

We now consider the numerical examples of [9]. To start, we review their work giving a

superset of S3(Q). Let ψ(n)k be the normalized weight k Siegel-Eisenstein series; that is,

ψ(n)k (Z) = [1]n(Z) =

∑γ∈∆n,0\Γn

j(γ, Z)−k

with a(

0n;ψ(n)k

)= 1 where 0n is the n × n 0 matrix. It is known that ψ

(n)k ∈ Mk(Γn)Q and in

particular, by [7], that

χ10 = − 43867

212 · 35 · 52 · 7 · 53

(2)4 ψ

(2)6 − ψ(2)

10

)and

χ12 =131 · 593

213 · 37 · 53 · 72 · 337

(3272(ψ

(2)4 )3 + 2 · 53(ψ

(2)6 )2 − 691ψ

(2)12

)are normalized cusp forms in M10(Γ2)Q and M12(Γ2)Q, respectively. In [8] these forms are further

normalized to

X10 = −4χ10 and X12 = 12χ12

so that a

1 1/2

1/2 1

;Xk

= 1 for k = 10, 12. Furthermore, by Theorem 1 of [8] we have

Xk ∈ Sk(Γ2)Z for k = 10, 12.

Theorem 3.4.1 ([19], Theorem 4.3) Assume p ≥ 5. If F ∈ Mk(Γ2)Z(p), then there exists a

unique polynomial P ∈ Z(p)[x1, x2, x3, x4] such that

F = P(ψ

(2)4 , ψ

(2)6 , X10, X12

).

Note that [19] uses the notation χ10 and χ12 instead of our X10 and X12, but states earlier

62

Page 65: HRG Master's Project

in the paper that Xk is normalized at a

1 1/2

1/2 1

;Xk

= 1 for k = 10, 12. We note that this

is equivalent to ⊕k∈2Z

Mk(Γ2)Z(p)= Z(p)

(2)4 , ψ

(2)6 , X10, X12

].

By [24], we have ψ(3)k ∈ Φ−1

Q(ψ

(2)k

)for k > 4 since ψ

(3)k = [1](3) and 1 ∈ Sk(Γ0). We define

F10 =43867

210 · 35 · 52 · 7 · 53

(3)4 ψ

(3)6 − ψ(3)

10

)F12 =

131 · 593

211 · 36 · 53 · 72 · 337

(3272(ψ

(3)4 )3 + 2 · 53(ψ

(3)6 )2 − 691ψ

(3)12

)

and immediately see that ΦQ(F10) = X10 and ΦQ(F12) = X12. Thus it is clear that S3(Q) is

contained in the set of primes p satisfying ν(p)

(3)k

)< 0 for k = 4, 6 or ν(p)(Fk) < 0 for k = 10, 12.

For the primes satisfying ν(p)

(3)k

)< 0 for k = 4, 6 we refer to Bocherer [2]. For ν(p)(Fk) < 0 for

k = 10, 12 it is clear that the primes we wish to avoid are contained in {2, 3, 5, 7, 53, 337} since we

are not guaranteed that there is a choice in Φ−1

Q (Xk) where any of those primes no longer occur in

the denominater of a Fourier coefficient.

We now review some of the numerical examples of Corollary 3.1.3 as given by [9] for the

case of degree 2. For simplicty, we will drop the exponent notation for the Siegel-Eisenstein sereies;

that is, we will write ψk = ψ(2)k . Furthermore, we let ∆ ∈ S12(Γ1) be Ramanujan’s delta function.

To simplify notation, we adopt the standard notation (m, r, n) to represent

m r/2

r/2 n

∈ Λ2.

Combining Theorem 1.1 and Proposition 2.2 of [4] we get the following theorem from [9].

Theorem 3.4.2 ([9], Theorem 4.1) Let p ≥ 5 be a prime, k even, and F ∈Mk(Γ(2))Z(p). If

a((m, r, n);F ) ≡ 0 (mod p)

for every 0 ≤ m,n ≤ k10 and r with 4mn− r2 ≥ 0, then we have F ≡ 0 (mod p).

The authors of [9] then state the theorem is easily modified to obtain the following.

Theorem 3.4.3 Let K be a number field with ring of integers O. Let p be a prime ideal such that

63

Page 66: HRG Master's Project

p - 2 · 3. Suppose k is even and F ∈Mk(Γ2)Op with

a((m, r, n);F ) ≡ 0 (mod p)

for every 0 ≤ m,n ≤ k10 and r with 4mn− r2 ≥ 0, then we have F ≡ 0 (mod p).

From these two theorems it becomes apparent that it suffices to check the following Fourier

coefficents based off of the even weight k of a degree two Siegel modular form.

T = (1, 0, 1), (1, 1, 1) if 10 ≤ k ≤ 18,

T = (1, 0, 1), (1, 0, 2), (1, 1, 1), (1, 1, 2), (2, 0, 2), (2, 1, 2), (2, 2, 2) if 20 ≤ k ≤ 28.

Though the theorems above call for more Fourier coefficients to be checked, in the case of even

weight k, we have the coefficients for (m, r, n), (n, r,m), (m,−r, n), and (n,−r,m) are all the same.

Weight 12 Example

Consider Ramanujan’s ∆-function and set f12 = 7∆ ∈ S12(Γ1). We then have [f12] is a mod

7 cusp form. By the main corollary, there exists a cusp form F12 ∈ S12(Γ2) such that [f12] ≡ F12

mod 7. To confirm this numerically we consider F12 := X12 ∈ S12(Γ2) and observe the following

congruences.

T = (m, r, n) a(T ; [f12]) a(T ;F12) Modulo 7

(1,0,1) 1242 10 3

(1,1,1) 92 1 1

Using Theorem 3.4.2 we see that [f12]− F12 vanishes modulo 7 and thus

[f12] ≡ F12 mod 7.

Weight 16

Consider the quadratic x2 − x − 12837 and one of its roots a; set K = Q(a). By [14] that

dimS16(Γ1) = 1 and thus posses a unique newform C16 ∈ S16(Γ1)Z with a(1;C16) = 1. Setting

f16 = 72 · 11 · C16 we obtain a cusp form f16 ∈ S16(Γ1) satisfying a(1; f16) = 72 · 11. Since C16

64

Page 67: HRG Master's Project

has integer Fourier coefficients, we must than have 72 · 11 divides all the coefficients of f16, i.e., f16

vanishes (mod 72). We take p = (7, a+4) and immediately see that [f16] is a mod p2 cusp form since

Φ([f16]) = f16. Moreover, by [14], we know that the space of degree 1, weight 30, newforms is spanned

by two Hecke eigenforms with coefficients in Q(α) where α satisfies 0 = x2 − 8640x − 454569984.

We note that −192a + 4416 also satisfies this quadratic. From [14], we take the newform that has

−192a + 4416 as its eigenvalue for the T (2) Hecke operator, for elliptic modular forms. We denote

this eigenform as g30 ∈ S30(Γ1) and take F16 ∈ S16(Γ2) to be its Saito-Kurokawa lift when we

normalize it as below.

T = (m, r, n) a(T ; [f16]) a(T ;F16) Modulo p2

(1,0,1) 5394 80a+3600 4

(1,1,1) 124 8a+1248 26

We apply Theorem 3.4.3 repeatebly where our first application of yields [f16] − F16 ≡ 0 (mod p),

i.e., there exists some form g ∈ M16(Γ2) such that [f16] − F16 = αg for some α ∈ p\p2. We then

apply the theorem again, this time to α−1([f16] − F16). We note that this is well defined since

[f16]− F16 vanishes (mod p), and since the required coefficients are congruent (mod p2) we must

have α−1([f16]− F16) ≡ 0 (mod p). Thus applying the theorem repeatedly yields

[f16] ≡ F16 (mod p2).

Weight 20

It is known that dimS20(Γ1) = 1 and thus, via scaling, we have f20 ∈ S20(Γ1) such that

a(1; f20) = 11 · 712. This shows that [f20] is a mod 712 cusp form. From [14], we take the weight 20,

degree 2 cusp form listed as “interesting” and scale it by -38 to obtain F20 ∈ S20(Γ2) with explicit

formula

F20 = 38

(2)4 ψ

(2)6 X10 +

(2)4

)2

X12 − 1785600X210

).

We have the following table of Fourier coefficients.

65

Page 68: HRG Master's Project

T = (m, r, n) a(T ; [f20]) a(T ;F20) Modulo 712

(1,0,1) 10386 304 304

(1,0,2) 1925356716 198816 2217

(1,1,1) 76 76 76

(1,1,2) 162929376 4256 4257

(2,0,2) 1238800286736 -335343616 3868

(2,1,2) 385264596000 278989920 816

(2,2,2) 9084897120 -63912960 1879

Thus we see from Theorem 3.4.3 that [f20]− F20 vanishes (mod 712) and hence

[f20] ≡ F20 (mod 712).

Weight 22

We consider weight 22 since dimS22(Γ1) = 1 and thus can obtain a cusp form f22 satisfying

a(1; f22) = 7 · 13 · 17 · 61 · 103. Thus we have [f22] is a mod 61 cusp form. We set F22 ∈ S22(Γ2) to

be the cusp form from [14] scaled by 29 . Thus we have

F22 =2

9

(−61

(2)4

)3

X10 − 5(ψ

(2)4

)2

X10 + 30ψ(2)4 ψ

(2)6 X12 + 80870400X10X12

).

We have the following table of Fourier coefficients.

T = (m, r, n) a(T ; [f22]) a(T ;F22) Modulo 61

(1,0,1) -179610 96 35

(1,0,2) -133169475780 -1728 41

(1,1,1) -740 -8 53

(1,1,2) -8620265280 -10752 45

(2,0,2) 54428790246720 -313368576 14

(2,1,2) 15093047985984 142287360 41

(2,2,2) 223472730240 17725440 60

Thus, by Theorem 3.4.2, we see that [f22]−F22 vanishes (mod 61), allowing us to conclude

that

[f22] ≡ F22 (mod 61).

66

Page 69: HRG Master's Project

3.5 New Examples

In order to understand the nature of the congruences given by the main corollary of [9] we

provide some more examples. These examples are restricted to degree 2 forms, so we simplify our

notation by writing ψ4 = ψ(2)4 and ψ6 = ψ

(2)6 . Throughout this section it is important to note that

any form ψa4ψb6X

c10X

d12 with c > 0 or d > 0 is a cusp form since

Φ(ψa4ψb6X

c10X

d12) = Φ (ψ4)

aΦ (ψ6)

bΦ (X10)

cΦ (X12)

d.

It is important to recall that⊕k∈2Z

Mk(Γ2)Z(p)= Z(p)

(2)4 , ψ

(2)6 , X10, X12

].

Weight 12

We start by highlighting why Takemori and Kikuta chose f12 = 7∆. It is known by [14]

that

[∆]21 =1

1728ψ3

4 −1

1728ψ2

6 +288

7X12 =

1

26 · 33ψ3

4 −1

26 · 33ψ2

6 +25 · 32

7X12.

This shows that ν(2)7 ([∆]21) = −1 while ν7(Φ([∆]21)) = ν7(∆) = 0 since a(1; ∆) = 1 and ∆ ∈ S12(Γ1).

Hence, we take 7 ∈ (7) in order to get

7[∆]21 =7

1728ψ3

4 −7

1728ψ2

6 + 288X12 ≡ X12 (mod 7).

Weight 16

Let C16 ∈ S16(Γ1) be the unique normalized newform of weight 16 and degree 1. Note that

this corresponds to the choice, f16, in [9] through the equation f16 = 72 · 11C16. By [14] we have the

following Klingen-Eisenstein series scaled so that a((1, 0, 0); [C16]21) = 1.

[C16]21 =1

26 · 33ψ4

4 −1

26 · 33ψ4ψ

26 +

23280

539ψ6X10 −

1152

77ψ4X12

=1

26 · 33ψ4

4 −1

26 · 33ψ4ψ

26 +

24 · 3 · 5 · 97

72 · 11ψ6X10 −

27 · 32

7 · 11ψ4X12

It is immediate that we have

49[C16]21 ≡ 3ψ6X10 (mod 7);

67

Page 70: HRG Master's Project

however, to match the corollary we must consider 49[C16]21 (mod 49). In that case we have

49[C16]21 ≡ 45ψ6X10 − 28ψ4X12 (mod 49).

We need to show that the right-hand side does not vanish, otherwise this is a trivial example. Using

Takemori’s code from GitHub associate with his paper [22] we have a((1, 0, 1);ψ6X10) = −2 and

a((1, 0, 1);ψ4X12) = 10, demonstrating that

a((1, 0, 1); 45ψ6X10 − 28ψ4X12) ≡ 45a((1, 0, 1);ψ6X10)− 28a((1, 0, 1);ψ4X12) (mod 49)

≡ 22 (mod 49).

Thus 49[C16]21 ≡ 45ψ6X10 − 28ψ4X12 (mod 49) is a non-vanishing example.

We can perform the saem analysis using the prime 11. In that case we obtain

11[C16]21 ≡ 3ψ6X10 + 2ψ4X12.

As before, we wish to show that the right-hand side does not vanish. To do so, we consider the

following.

a((1, 0, 1); 3ψ6X10 + 2ψ4X12) ≡ 3a((1, 0, 1);ψ6X10) + 2a((1, 0, 1);ψ4X12) (mod 11)

≡ 3 (mod 11)

This demonstrates that any choice of prime p such that ν(2)p (f) = νp(Φ(f)) −m for some

m ∈ Z>0 can be used to construct an example.

Weight 18

Let C18 ∈ S18(Γ1) be the unique normalized newform of weight 18 and degree 1. We then

have

[C18]21 =1

1728ψ3

4ψ6 −1

1728ψ3

6 −11448

143ψ2

4X10 +108144

1001ψ6X12

=1

26 · 33ψ3

4ψ6 −1

26 · 33ψ3

6 −23 · 33 · 53

11 · 13ψ2

4X10 +24 · 32 · 571

7 · 11 · 13ψ6X12

68

Page 71: HRG Master's Project

We then have [C18]21 ∈ M18(Γ2)Q and f18 := 1729728[C18]21 ∈ M18(Γ2)Z. We can then identify

these forms respectively in M18(Γ2)K and M18(Γ2)OKwhere K = Q(α) ' Q[x]/(x2 − x− 589050).

Let F18 ∈ S18(Γ2) be the unique Maass form given by [14] with explciit formula

F18 = 1295ψ24X10 + (α− 575)ψ6X12 ∈ S18(Γ2)OK

.

Before checking for congruences we note that Q(α) = Q(√d) where d = 2356201 is the discriminant

of x2 − x − 589050. This equality allows us to utilize Theorem 8.3 of [3] to see that the ideals

(7), (11), (13) ⊂ OK all split. Using Corollary 8.4 of [3] we then have (7, α), (11, α+10), (13, α+4) ⊂

OK are prime ideals lying above (7), (11), and (13), respectively.

The first case we consider is congruences with respect to p1 = (7, α). Observe that we have

the following table of Fourier coefficients.

T = (m, r, n) a(T ; f18) a(T ;F18) Modulo p1

(1,0,1) -2643840 10α− 8340 -3

(1,1,1) -34560 α+ 720 -1

Using Theorem 3.4.3 we then have f18 − F18 ≡ 0 (mod p1) which is equivalent to

1729728[C18]21 ≡ F18 (mod p1).

The second case we consider is p2 = (11, α+10). For this case, we use 3f18 = 5189184[C18]21.

T = (m, r, n) a(T ; 3f18) a(T ;F18) Modulo p2

(1,0,1) -7931520 10α− 8340 -5

(1,1,1) -103680 α+ 720 -3

Using Theorem 3.4.3 we then have 3f18 − F18 ≡ 0 (mod p2) which is equivalent to

5189184[C18]21 ≡ F18 (mod p2).

The last case we consider for weight 18 is p3 = (13, α+ 4) and −11f18 = −19022058[C18]21.

T = (m, r, n) a(T ;−11f18) a(T ;F18) Modulo p3

(1,0,1) 29082240 10α− 8340 5

(1,1,1) 380160 α+ 720 1

69

Page 72: HRG Master's Project

By Theorem 3.4.3 we then have F18 + 11f18 ≡ 0 (mod p3) which is equivalent to

−19022058[C18]21 ≡ F18 (mod p3).

Weight 20

Let C20 ∈ S20(Γ1) be the unique normalized newform of weight 20 and degree 1. We then

have

[C20]21 =1

26 · 33ψ5

4 −1

26 · 33ψ2

4ψ26 +

25 · 3 · 5 · 23 · 421

11 · 712ψ4ψ6X10 −

25 · 32 · 977

712ψ2

4X12 −214 · 37 · 53 · 17

11 · 712X2

10.

We consider the case p = 11 we have

11[C20]21 ≡ 7psi4ψ6X10 + 3X210 (mod 11).

Using [22] we see that a((1, 0, 1);ψ4ψ6X10) = −2 and a((1, 0, 1);X210) = 0. We see that

a((1, 0, 1); 7ψ4ψ6X10 + 3X210) = −14 ≡ 8 (mod 11)

showing that the right-hand side does not vanish. We go even further by considering the field

Q(β) ∼= Q[x]/(x2− x− 15934380). Since x2− x− 15934380 ≡ x(x− 1) (mod 11) we know 11 splits.

We consider the prime ideal p = (11, β) and the form g20 := −95819328[C20]21. We set G20S20(Γ2)

to be the unique Maass form given by [14] with explicit formula

G20 = (10β + 40100)ψ4ψ6X10 + (−7β − 21560)ψ24X12 + 6749568000X2

10.

Using [14] we obtain the following table of Fourier coefficients.

70

Page 73: HRG Master's Project

T = (m, r, n) a(T ; 5g20) a(T ;G20) Modulo p

(1,0,1) -89735040 −90β − 295800 1

(1,0,2) -16635082026240 −27036β − 78025680 4

(1,1,1) -656640 3β + 18540 5

(1,1,2) -1407709808640 −5544β − 19679520 -3

(2,0,2) -10703234477399040 −10036800β + 11326214400 -5

(2,1,2) -3328686109440000 −3997080β − 40732178400 -1

(2,2,2) -78493511116800 287712β + 8190149760 1

By Theorem 3.4.3 we have 5g20 −G20 ≡ 0 (mod p) which is equivalent to

−479096640[C20]21 ≡ G20 (mod p).

71

Page 74: HRG Master's Project

Chapter 4

Future Work

One possible direction for future work is given in [9]. The authors pose the following

question.

Question: Given degree n, does there exist an explicit bound Cn such that

maxSn(Q) < Cn?

They pose this question since we know S2(Q) ⊂ {2, 3} and S3(Q) ⊂ {2, 3, 5, 7, 53, 337}. We the

significance of such a bound is seen in Lemma 3.2.6, where Sn(K) is related back to Sn(Q). Hence,

knowing a bound on on primes in Sn(Q) would provide more information about the primes occuring

in Sn(K) for arbitrary number field K.

Another direction would be to pursue the same result but for degree n, level N , weight k

modular forms given by Mk(Γn(N)). For the results used in [9] to hold we have two main sticking

points:

1. Does Theorem 2.3 of [6] hold for arbitrary level? That is; do we have

⊕k≥1

Mk(Γn(N))Z = Z[φ1, . . . , φm]

for some m ∈ Z≥1 and generators φ1, . . . , φm?

2. For which k do we have

Mk(Γn(N)) = Mk(Γn(N))Q ⊗Q C?

72

Page 75: HRG Master's Project

To answer the first question requires developing a background in algebraic geometry. We can tackle

the second question by combining Theorem 1 of [21] with the result from the same paper stating

Mk(Γn(N)) = Mk(Γn(N))Z ⊗Z C

for all even k as well as for any odd k satisfying

0 6= A(Γn(N),Q) =

{f

g| f ∈Mk+m(Γn(N))Q, g ∈Mm(Γn(N))Q,m ∈ Z≥0

}.

This should follow from [20] giving A(Γn,Q) 6= 0 due to theta functions.

A final direction worth considering would be the case of Hilbert modular forms as they are

another multivariables generalization of elliptic modular forms. Many of Shimura’s results on Siegel

modular forms hold for Hilbert modular forms as demonstrated by [20]. The approach may have to

differ as the main theorem of [9] depended on using the Siegel operator, in particular its surjectivity,

to transition between degree n− 1 and degree n Siegel modular forms.

73

Page 76: HRG Master's Project

Bibliography

[1] Anatolij N. Andrianov. Quadratic forms and Hecke operators, volume 286 of Grundlehren derMathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Springer-Verlag, Berlin, 1987.

[2] S. Bocherer. Uber die fourierkoeffizienten der siegelschen eisensteinreihen. Manuscr. Math.,45:273–288, 1985.

[3] K. Conrad. Factoring in quadratic fields. http://www.math.uconn.edu/ kcon-rad/blurbs/gradnumthy/quadraticgrad.pdf.

[4] T. Kikuta D. Choi, Y. Choie. Sturm type theorem for siegel modular forms of genus 2 modulop. Acta Arithmetica, 158(2):129–139, 2013.

[5] Fred Diamond and Jerry Shurman. A first course in modular forms, volume 228 of GraduateTexts in Mathematics. Springer-Verlag, New York, 2005.

[6] Gerd Faltings and Ching-Li Chai. Degeneration of abelian varieties, volume 22 of Ergebnisseder Mathematik und ihrer Grenzgebiete (3) [Results in Mathematics and Related Areas (3)].Springer-Verlag, Berlin, 1990. With an appendix by David Mumford.

[7] J. Igusa. On siegel modular forms of genus two. American Journal of Mathematics, 84:175–200,1962.

[8] J. Igusa. On siegel modular forms of genus two. American Journal of Mathematics, 101:149–183,1979.

[9] T. Kikuta and S. Takemori. Ramanujan type congruences for the Klingen-Eisenstein series.ArXiv e-prints, February 2014.

[10] H. Klingen. Zum darstellungssatz fur siegelsche modulformen. Math. Zeit., 102:30–43, 1967.

[11] Helmut Klingen. Introductory lectures on Siegel modular forms, volume 20 of Cambridge Studiesin Advanced Mathematics. Cambridge University Press, Cambridge, 1990.

[12] N. Kurokawa. On eisenstein series for siegel modular groups. Proc. Japan Acad., 57(A):51–55,1981.

[13] N. Kurokawa. On siegel eigenforms. Proc. Japan Acad., 57(A):47–50, 1981.

[14] The LMFDB Collaboration. The l-functions and modular forms database. http://www.lmfdb.org, 2016. [Online; accessed 24 October 2016].

[15] H. Maass. Die primzahlen in der theorie de siegelschen modukfunktionen. Math. Annalen,124:87–122, 1951.

74

Page 77: HRG Master's Project

[16] Hideyuki Matsumura. Commutative ring theory, volume 8 of Cambridge Studies in AdvancedMathematics. Cambridge University Press, Cambridge, 1986. Translated from the Japanese byM. Reid.

[17] S. Mizumoto. Congruences for fourier coefficients of lifted siegel modular forms i: Eisensteinlifts. Abh. Math. Sem. Univ. Hamburg, 75:97–120, 2005.

[18] Shin-ichiro Mizumoto. Poles and residues of standard L-functions attached to Siegel modularforms. Math. Ann., 289(4):589–612, 1991.

[19] S. Nagaoka. Note on mod p siegel modular forms. Mathematische Zeitschrift, 235:405–420,2000.

[20] G. Shimura. On some arithmetic properties of modular forms of one and several variables.Annals of Mathematics, 102(3):491–515, 1975.

[21] G. Shimura. On the fourier coefficients of modular forms of several variables. Nachr. Akad.Wiss Gottingen Math.-Phys., 17:261–268, 1975.

[22] S. Takemori. On the computation of siegel modular forms of degree 2 with sage. RIMSKokyuroku, 1907-18, 2014.

[23] Gerard van der Geer. Siegel modular forms and their applications. In The 1-2-3 of modularforms, Universitext, pages 181–245. Springer, Berlin, 2008.

[24] R. Weissauer. Stabile Modulformen und Eisenstein Reihen, Lect. Notes in Math., volume 1219.Springer, 1987.

[25] D. Zantout. On the Cuspidality of Maass-Gritsenko and Mixed Level Lifts. PhD thesis, ClemsonUniversity, 2013.

[26] N. A. Zharkovskaya. The siegel operator and hecke operators. Funct. Anal., 8:113–120, 1974.

[27] N. A. Zharkovskaya. On the connection of the eigenvalues of hecke operators and the fourier coef-ficients of eigenfunctions for siegel’s modular forms of genus n. Math. USSR Sbornik, 25(4):549–557, 1975.

75


Recommended