Sunteți pe pagina 1din 133

Hossein Movasati

A Differential Introduction to
Modular Forms and Elliptic
Curves
The text is under construction and it is very messy. DO NOT PRINT IT

February 15, 2018

Publisher
Contents

1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.1 Fibonacci sequence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2 Fermat’s last theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.3 Aritmetic modularity theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.4 Beyond elliptic curves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.5 Prereuisites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

2 Modular forms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.1 Elliptic functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.2 The modular group and its action . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.3 Slash operator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.4 Weierstrass ℘-function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.5 Differential equation of ℘ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.6 Eisenstein series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.7 Fourier expansion of Eisenstein series . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.8 The Eisenstein series E2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.9 The algebra of moduler forms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.10 Ramanujan relations between eisenstein series . . . . . . . . . . . . . . . . . . 24
2.11 The product formula for discriminant . . . . . . . . . . . . . . . . . . . . . . . . . . 25
2.12 The j function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
2.13 Poincaré metric . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.14 The numbers e1 , e2 , e3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.15 Growth of coefficients . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
2.16 Dedekind eta function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.17 Modular forms as k-fold differential . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
2.18 Petersson Scalar product . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

3 Elliptic curves and integrals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33


3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
3.2 Elliptic integrals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
3.3 Elliptic curves in Weierstrass format . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

v
vi Contents

3.4 Picard-Lefschetz theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37


3.5 Weierstrass uniformization theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
3.6 Sketch of the proof of Theorem (2.9.1) . . . . . . . . . . . . . . . . . . . . . . . . . 39
3.7 Some identities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
3.8 Schwarz function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
3.9 CM elliptic curves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
3.10 Fourier expansions and elliptic integrals . . . . . . . . . . . . . . . . . . . . . . . . 42

4 Rudiments of Algebraic Geometry of curves . . . . . . . . . . . . . . . . . . . . . . 45


4.1 Curves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
4.2 Charts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
4.3 Schemes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
4.4 Singular and smooth curves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
4.5 Resultant and discriminant . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
4.6 Discriminant . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
4.7 The main property of the discriminant . . . . . . . . . . . . . . . . . . . . . . . . . 51
4.8 Descriminant of projective curves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
4.9 Curves of genus zero . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
4.10 Curves of genus bigger than one . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
4.11 Elliptic curves in Weierstrass form . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
4.12 Elliptic curves in Weierstrass form . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
4.13 Real geometry of elliptic curves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
4.14 Complex geometry of elliptic curves . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
4.15 The group law in elliptic curves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
4.16 Divisors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
4.17 Riemann-Roch theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
4.18 Weierstrass form revised . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
4.19 Moduli of elliptic curves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
4.20 The addition formula for ℘ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
4.21 Why Schemes? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

5 Mordell-Weil Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
5.1 Mordell-Weil theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
5.2 Weak Mordell-Weil theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69

6 Torsions and isogeny . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71


6.1 Torsion points . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
6.2 Isogeny . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
6.3 Isogeny II . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74

7 Hecke operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
7.1 Hecke operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
7.2 Hecke and cusp forms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
7.3 Proof of Theorem 6.3.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
Contents vii

8 Riemann zeta function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83


8.1 Riemann zeta function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
8.2 The big Oh notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
8.3 Gamma function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
8.4 Mellin transform . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
8.5 Analytic extension . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
8.6 Functional equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
8.7 Second proof for functional equation . . . . . . . . . . . . . . . . . . . . . . . . . . 88
8.8 Zeta and primes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
8.9 Other zeta functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
8.10 Dedekind Zeta function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
8.11 L-function of cusp forms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
8.12 Hecke’s L-functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
8.13 L-function of CY-modular forms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94

9 Congruence groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
9.1 Congruence groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
9.2 Weil pairing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
9.3 Moduli spaces of elliptic curves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
9.4 Modular forms for congruence groups . . . . . . . . . . . . . . . . . . . . . . . . . 99
9.5 q-expansion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
9.6 Transcendental degree of modular forms . . . . . . . . . . . . . . . . . . . . . . . 102

10 Elliptic curves as Diophantine equations . . . . . . . . . . . . . . . . . . . . . . . . . . 105


10.1 Finite fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
10.2 Zeta functions of elliptic curves over finite fields . . . . . . . . . . . . . . . . 105
10.3 Nagell-Lutz Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
10.4 Mazur theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
10.5 One dimensional algebraic groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
10.6 Reduction of elliptic curves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
10.7 Zeta functions of curves over Q . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
10.8 Hasse-Weil conjecture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
10.9 Birch Swinnerton-Dyer conjecture . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
10.10Congruent numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
10.11p-adic numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115

11 Theta series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117


11.1 Two-squares theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
11.2 Poisson summation formula . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121

12 Online supplemental items . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123


12.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
12.2 How to start? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
12.3 Ramanujan differential equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
viii Contents

Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
Preface

There are so many books on moduler forms and elliptic curves that it might seem
useless to add another one. None of these books approach modular forms from the
point of view of differential equations and this differentiate the present book from
others. This has resulted in a tremendous generalization of modular forms whose
origin partially comes from many q-expansions computations in theoretical physics
and in particular string theory.

Hossein Movasati
January 2019
Rio de Janeiro, RJ, Brazil

1
Chapter 1
Introduction

Our main interest on modular forms is the fact that they are generating functions
for many unexpected counting in mathematics. Why generating functions are useful
might be explained with the simple example of Fibonacci numbers.

1.1 Fibonacci sequence

The Fibonacci sequence is defined in the following way

Fn+2 = Fn+1 + Fn , n > 0, F0 = 1, F1 = 1 (1.1)

Few elements of this sequence are

1, 1, 2, 3, 5, 8, 13, 21, . . .

Once you have a sequence of natural numbers in mathematics, it is recommanded


to put it in a generating function:

F (q) : = q + q2 + 2q3 + 3q4 + . . . + Fnqn + · · ·

At the beginning this is just a formal power series, however, soon it will become
clear that its a convergence and radius of convergence carries many information of
the sequence Fn itself. For now, let us do the following manipulation:
∞ ∞
F (q) = ∑ Fn qn = q + ∑ (Fn−1 + Fn−2 ) qn
n=0 n=2
2
= q + q · F (q) + q F (q)

which implies that


q
F (q) = (1.2)
1 − q − q2

3
4 1 Introduction

Therefore, F (q) converges to a rational function. In order to find the radius of con-
vergence of a rational function, we have to find the roots of its denominator:

q q (α − β )−1 (α − β )−1
F (q) = = = −
1 − q − q2 (1 − α · q) (1 − β · q) 1−α ·q 1−β ·q

αn − β n
∞  
= ∑ qn
n=0 α −β

1
 √   √ 
where α = 2 1 + 5 , β = 12 1 + 5 . We conclude that
 √ n  √ n
1+ 5
2 − 1−2 5
Fn = √
5
Which at first glance looks strange because we have found a formula for the in-
tegernFn in terms
o of square root of 5. Since the radius of convergence of F (q) is
1
min |α| , |β1 | = max {|α| , |β |} = |α|, we conclude that

Fn 1 1 √ 
lim = lim Fnn = 1+ 5
n→∞ Fn−1 n→∞ 2
This number is called the golden rati or the golden number.
Exercise 1 Show that  n  
11 Fn+1 Fn
= .
10 Fn Fn−1

1.2 Fermat’s last theorem

Modular forms and elliptic curves are firmly rooted in the fertil grounds of number
theory. As a proof of the mentioned fact and as an introduction to the present text we
mention the followings: For p prime, the Fermat last theorem ask for a non-trivial
integer solution for the Diophantine equation

ap + bp + cp = 0

For a hypothetical solution (A, B,C) = (a p , b p , c p ) of the Fermat equation with


abc 6= 0, Gerhart Frey considered the elliptic curve

EA,B,C : y2 = x(x − A)(x + B)

From this one construct a modular form fA,B,C and a Galois representation with
certain properties and then one proves that such objects does not exist. During this
passage one encounters the Modularity conjecture which claims that every elliptic
1.3 Aritmetic modularity theorem 5

curve over Q is modular. Roughly speaking this means that every elliptic curve over
Q appears in the Jacobian of of a modular curve of level N. Another formulation of
modularity property is by using L functions which generalizes the famous Riemann
zeta function

1
ζ (s) := ∑ s
n=1 n

Riemann hypothesis claims that all the non-trivial zeros of ζ lies on ℜ(s) = 12 and it
has strong consequences on the growth of prime number. For the L functions associ-
ated to elliptic curves one has the Birch-Swinnerton Dyer conjecture which predicts
the rank of an elliptic curve to be the order of vanishing of the corresponding L-
function at s = 1.

1.3 Aritmetic modularity theorem

Modular forms as generating functions have many fascinating and mysterious ap-
plications. Aritmetic modularity theorem is one of these. In many books and articles
we find the expression
“Let E be an elliptic curve over Z” .
This has an intrinsic definition, that for now, we don’t want to get into its details.
We content ourselves with the example.

E : y2 + y = x3 − x2

which the reader might consider it as a Diophantine equation, that is, we are inter-
ested to find x and y in the ring of integers, the field of rational numbers, finite fields,
etc. Let p be a prime number (don’t take the Grothendieck’s prime) i We count the
number of solutions N p of E modulo the prime p.

P Solutions Np
2 (0,0),(0,1), (1,0), (1,1) 4
3 (0,0),(0,2), (1,0), (1,2) 4
5 (0,0),(0,4), (1,0), (1,4) 4
7 (0,0),(0,6), (1,0), (1,6), ( , ),. . . 9
11 (0,0),(0,10), (1,0), (1,10),( , ),. . . 10

In total we have to substitute p2 pairs (x, y) , x, y = 0, 1, . . . , p − 1 inside E and verify


whether modulo prime p it is zero or not. The first four solutions in the above table
have to do with the fact that over integers E has already four solutions.

(0, 0) , (0, −1) , (1, 0) , (1, −1)

iA. Grothendieck (1928-2014) is one of the founders of modern Arithmetic Algebraic Geometry.
Once he was asked to give an example of a prime number and he answered: 57.
6 1 Introduction

A priori, if we have computed N2 , N3 , N5 , . . . , N11 , this doesn’t give any clue how to
find the number N13 . We have to check 132 cases again. In modern language, we say
that, we are counting the number of F p -rational points of E and we write

N p : = #E (F p )

Here F p := {0, 1, 2, . . . , p − 1} is the finite field with p elements.


Exercise 2 Find N p for all p 6 20.
The theory of modular forms, and in particular arithmetic modularity theorem, says
that there is a closed formula for the generating function of N p ’s. This is as follows.
Let
1

η (q) = q 24 ∏ (1 − qn ) (1.3)
n−1

be the Dedekind eta function. We consider it as a formal product. Let


2
F (q) = η (q)2 η q11
2 2·11
∞ ∞ 2
= q 24 + 24
∏ (1 − qn )2 ∏ 1 − q11n
n−1 n−1
= q − 2q − q + 2q + q + 2q6 − 2q7 − 2q9 − 2q10 + q11 − 2q12 + 4q13 + . . .
2 3 4 5

= ∑ fn qn
n=1

The arithmetic modularity theorem tells us that

Np = p − f p (1.4)

and f is a modular form. More precisely,


“ f is a weight 2 new form for Γ0 (11)”
One of the aims of the present text is to understand this statement. This phenomena
is a part of a general theorem:
Theorem 1.3.1 (Arithmetic modularity theorem) (A. Wiles, R.Taylor, C. Breuil,
B. Conrad, F. Diamond) For any elliptic curve E over Q, there is a modular form
f = ∑∞ n
n=1 f n q such that (1.4) holds for all except a finite number of primes.

A precise statement, together with other equivalent versions will be presented in this
text.
Exercise 3 Show that the radius of convergence of the Dedekind η function is 1.
1.5 Prereuisites 7

1.4 Beyond elliptic curves

There is a tremendous amount of effort to generalize the arithmetic modularity the-


orem beyond elliptic curves. Here we give an example taken from [Schuett2013].
Let us consider the Fermat quartic surface

X = X24 ⊆ P3 : x04 + x14 + x24 + x34 = 0

We count the number of solutions of this Diophantine equation over the field F p , p 6=
2
#X(F p ) : = # [x0 : x1 : x2 : x3 ]|x04 + x14 + x24 + x34 = 0 .


Here, [x0 : x1 : x2 : x3 ] is the equivalence class

(x0 : x1 : x2 : x3 ) ∼ (y0 , y1 , y2 , y3 ) ⇔ ∃a ∈ F p − {0}

such that xi = ayi , ∀i = 0, 1, 2, 3. It turns out that

#X(F p ) = 1 + b p + h · p + p2

where
∞ ∞
η(4τ)6 = q ∏ (1 − q4n )6 = ∑ bn qn
n=1 n=1

h = 5 + 3χ−1 (p) + 6 · (χ2 (p) + χ−2 (p))


and χa (p) : = ( ap ) is the Legendre symbol.
Exercise 4 Verify the above affirmation for p = 3, 5, . . . what goes wrong for p = 2?

1.5 Prereuisites

It is assumed that the reader has a basic knowledge in algebraic geometry of curves
and complex analysis in one variable.
Chapter 2
Modular forms

2.1 Elliptic functions

Definition 2.1.1 A lattice Λ in C is a discrete subgroup of (C, +) which generate it


as an R−vector space
It follows easily that

Λ = Zω1 + Zω2 = {nω1 + mω2 |n, m ∈ Z}

where ω1 , ω2 ∈ C, ω1 , ω2 6= 0, Im(ω2 /ω1 ) 6= 0. By changing the order of ω1 , ω2 , if


necessary, we can assume that
ω1
Im(τ) > 0, τ := (2.1)
ω2
A lattice in general is equipped with a bilinear map Λ × Λ → Z. In our case it is
skew-symmetric, that is,

< a, b >= − < b, a > ∀a, b ∈ Λ

and so < ω1 , ω1 >= 0, < ω2 , ω2 >= 0. Therefore,

< ω2 , ω1 >:= 1 (2.2)

determines< ·, · > uniquely. The choice of ω1 , ω2 with (2.1) and hence with (2.2) is
also called an orientation of Λ . If we choose another basis ω10 , ω20 with (2.2) then
 0  
ω1 ω1
= A , A ∈ SL(2, Z)
ω20 ω2

where   
ab
ad − bc = 1, a, b, c, d, ∈ Z
SL(2, Z) := (2.3)
cd

9
10 2 Modular forms

Fig. 2.1 Lattice

Let P be the space of lattice in C. The group Gm = C∗ = (C − {0}, ·) acts on P from


the right
P × C∗ → P
(Λ , λ ) 7−→ λ · Λ := Zλ ω1 + Zλ ω2
For a lattice Λ the associate complex tori is simply

E := C/Λ

this means that two points z1 , z2 ∈ C are equivalent if z1 − z2 ∈ Λ .


The set C/Λ is an example of a Riemann surface or complex manifold. It is
called real torus of dimension two or complex torus of dimension one. It has the
structure of an abelian group which inherits from (C, +). Still we do not call it an
elliptic curve as this name is reserved for a similar object in algebraic geometry.
In mathematics when we have a space, then we start to study the set of its func-
tions. In our case, we are interested on meromorphic functions on C/Λ . This is
Translated into functions

f :C→C
f (z + ω) = f (z) ∀z ∈ C, ω ∈ Λ

Since Λ is generated by ω1 , ω2 , (2.4) is equivalent to

f (z + ω1 ) = f (z)
∀z ∈ C
f (z + ω2 ) = f (z)
2.1 Elliptic functions 11

Fig. 2.2 Torus

that is f is double periodic. We may also view f as a function of Λ ∈ P. In this


case we write f (z) = f (z,Λ ). Since P is equipped with a C∗ -action, it is natural to
characterize functions f with

f (λ z, λΛ ) = λ −a f (z,Λ ), ∀λ ∈ C∗ (2.4)

for some fixed a ∈ Z.


Definition 2.1.2 A meromorphic function f with the property (2.4) is called an
elliptic function (of weight a).
Let us consider an elliptic function f and write its Laurent series at z = 0,
+∞
f (z,Λ ) = ∑ fn (Λ )Z n
n=−∞

The coefficients fn (Λ ) are functions of the lattice Λ and it is easy to see that (2.4)
is equivalent to the following functional equations for fn (Λ )’s

fn (λΛ ) = λ −a−n fn (Λ ) ∀λ ∈ C∗

This is as follows
12 2 Modular forms

+∞
f (λ z, λΛ ) = ∑ fn (λΛ )(λ z)n
n=−∞
!
+∞
= λ −a ∑ fn (Λ )zn
n=−∞

Definition 2.1.3 A meromorphic function f on the space P of lattices is called a


meromorphic modular form of weight n ∈ Z if

f (λΛ ) = λ −n f (Λ ) ∀λ ∈ C∗ , Λ ∈ P

Therefore, from a meromorphic elliptic function of weight a we get meromorphic


modular form fn of weight n + a. If we evaluate a meromorphic modular form f of
weight n on lattices Λ = Zτ + Z with τ in the upper half plane H and regard them as
a function in τ, we get a meromorphic function f in H with the following functional
equation.
aτ + b
(cτ + d)−n f ( ) = f (τ), ∀τ ∈ H. (2.5)
cτ + d

2.2 The modular group and its action

The upper half plane is defined as

H := {τ ∈ C| Im(τ) > 0}.

The following group acts on H by Möbius transformation


  
a b
SL(2, R) := ad − bc = 1, a, b, c, d, ∈ R
cd
SL(2, R) × H −→ H
(A, τ) 7−→ Aτ := aτ+b
cτ+d
 
ab
where A = ∈ SL(2, R). This follows from
cd

Im(τ) det(A)
Im(Aτ) = (2.6)
|cτ + d|2

for any matrix A with real entries. An element A ∈ SL(2, R) acts as identity on H if
it is ±I, where  
10
I=
01
is the indentity matrix. Therefore, it is usefull to define
2.4 Weierstrass ℘-function 13

PSL(2, R) = SL(2, R)/ ± I

The protagonist of the present text is the group SL(2, Z) defined in (2.3). Let τ ∈ H
and assume that it has non-trivial stablizer under the action of SL(2, Z), that is, there
is A ∈ SL(2, Z), A 6= ±I such
√ that Aτ = τ. Since det A = +1 and Im(τ) > 0 we get
a−d+i (a+d)2 −4
|a + d| < 2 and τ = c .
Exercise 5 Show that the group SL(2, Z) is generated by
   
11 0 −1
T := , S := . (2.7)
01 1 0

2.3 Slash operator

For A ∈ GL(2, R) and a modular form f of weight k we define the slash operator
 
∗∗
f |k A := (det A)k−1 (cτ + d)−k f (Aτ), A = .
cd

In some books, the powe k − 1 of det Ais different. For instance, in Chapter 7 the
slash operator is different. For A ∈ SL(2, R) this will not make difference.
Proposition 2.3.1 If B ∈ GL(2, R)and f is a meromoephic modular form of weight
k for some group Γ ∈ GL(2, R) then f |k B is a modular form of the same weight for
B−1Γ B.

Proof. This follows from

( f |k B)|k B−1 AB = ( f |k A)|k B = f |k B

t
u

2.4 Weierstrass ℘-function

When a discrete group Γ acts on a space M and we want to construct functions on


the quotient space

Γ \M := M/ ∼ χ ∼ y ⇔ χ = Ay for som A ∈ Γ

the first recepie is to start with a function f˜ on M and define the formal sum

f (τ) = ∑ f˜(Aτ) (2.8)


A∈Γ
14 2 Modular forms

If we don’t care about the convergence of f , the we can check that it is invariant
under the action of Γ . For any B ∈ Γ we have

f (Bτ) = ∑ f˜(ABτ)
A∈Γ
= ∑ f˜(Aτ) = f (τ)
A∈Γ

Note that we have used the fact that the multiplication by B from the right induces a
bijection Γ → Γ . We get the function

fˇ : Γ \M → C, fˇ([τ]) = f (τ)

which we denot it again by f = fˇ If Γ is finite then (2.8) is a finite sum an so f is


well-defineed, however, in general such a sum might not be convergent. In Our case,
the lattice Λ as an additive group acts on C

Λ × C → C, (λ , z) 7−→ z + λ

Theorem 2.4.1 Show that the number of zeros of a non-constant elliptic function
counted in C/Λ is equal to the number of poles, counted with multiplicity, and it is
bigger than or equal to 2.

Proof. See [Apostol 1989], page 5-6.

For our purpose we start with f˜(z) = z−a , a ∈ Z and define

fa (z) = fa (z, ω) : = ∑ (z + ω)−a


ω∈Λ
= ∑ (z + nω1 + mω2 )−a
(n,m)∈Z2

Proposition 2.4.1 The infinite series fa (z) converges absolutely for a ∈ N with a >
3

Proof. (See lemma 2 page 8 of [Apostol]). We can assume that the sum is taken for
all ω ∈ Λ , |ω| > R and |z| < R. There is a constant M depending on R and a such
that
1 M ∀ω ∈ Λ , |ω| > R
|z−ω|a 6 |ω|a ∀z ∈ C, |z| < R
1
Therefore, it is enough to prove that the sum ∑ω∈Λ |ω|a is convergent. Let r and R
be the minimum and maximum distanced of 0 from the parallelogram formed by
±ω1 + ±ω2 . Then the parallelogram Pn formed by four vertices n(±ω1 ± ω2 ) has
the minimum and maximum distances nr and nR, respectively, from 0. Moreover, it
2.4 Weierstrass ℘-function 15

has 8n points of the lattice. Therefore, the sum Sn corresponding to all parallelograls
P1 , P2 , . . . , Pn satisfies
8 8
(1 + 2a−1 + · · · + na−1 ) ≤ Sn ≤ a (1 + 2a−1 + · · · + na−1 )
Ra r
t
u

Exercise 6 Show that f2 does not converge in any z ∈ C, z ∈



Note that
∂ fa
= −a · fa+1
∂z

Definition 2.4.1 The Weierstrass ℘ function (read P) is


1 1 1
℘(z,Λ ) = ℘(z) := 2
+∑ 2
− 2
z ω6=0 (z − ω) ω

Proposition 2.4.2 ℘ is convergent

Proof. (See [Apostol] Theorem 1.10). We have



1 z(2ω−z) M·R·(2|ω|+R)
(z−ω)2 − ω12 = (z−ω)2 ω 2 6 |ω|2 |ω|2

M·R·(2+R/|ω|) 3MR
6 |ω|3
6 |ω|3

where we used the notation in the proof Proposition (2.4.2) t


u

It is easy to see that


℘(−z) = ℘(z)
that is ℘ is an even function.
Exercise 7 Show that there is no function f (ω), ω ∈ Λ such that

1
∑ + f (ω)
ω∈Λ z−ω

is convergent
The function ℘ has poles at the points of Λ . We write the Laurent expansion of ℘
at z = 0
Theorem 2.4.2 We have

1
℘(z) = + ∑ (2n + 1)G2n+2 · zn
z2 n=1
16 2 Modular forms

where
1
G2n+2 = ∑ 2n+2
· (2.9)
ω6=0 ω
and
0 < |z| < r := min{|ω| | ω 6= 0}· (2.10)
Proof. For z in (2.10) we have ωz < 1 and

!
∞  z n
1 1 1
2
= 2 z 2 = 1 + ∑ (n + 1)
(z − ω) ω (1 − ω ) ω2 n=1 ω

and so

1 1 n+1
2
− 2 = ∑ n+2 n
(z − ω) ω n=1 ω ·z
Summing over ω ∈ Λ , ω 6= 0,we get the result. t
u
The notation
g2 = 60G4 , g3 = 140G6 (2.11)
is frequently used.

2.5 Differential equation of ℘

Theorem 2.5.1 The function ℘ satisfies the differential equation

℘0 (z)2 = 4℘(z)3 − 60G4 x − 140G6 (2.12)

Proof. Let f (z) be the difference of both sides of (2.12). This is clearly an elliptic
function with possible poles at z ∈ Λ . We show that f is holomorphic at z = 0 and
so f = 0. We have

℘0 (z) = −2
z3
+ 6G3 · z + 20G6 · z3 + . . .
℘0 (z)2 = z46 − 24G
z2
4
− 80G6 + . . .
3 4 36G 4
4℘(z) = z6 + z2 + 60G6 + . . .

hence
℘0 (z)2 = 4℘(z)3 = − 60G
z2
4
− 140G6 + . . .
℘0 (z)2 3
= 4℘(z) = +60G4℘(z) = −140G6 + . . . t
u

Definition 2.5.1 An elliptic function f is a meromorphic function in C such that


It is double periodic, that is, there is two Z linearly independent complex numbers
ω1 , ω2 ∈ Z such that

f (z + ω1 ) = f (z), f (z + ω2 ) = f (z)
2.6 Eisenstein series 17

If f is a non-constant elliptic function, we can easily see that Im(ω1 , /ω2 ) 6= 0.


Exercise 8 ([Apostol] page 23, Exercise 5) Prove that every elliptic function f can
be written as
R1 [℘(z)] +℘0 (z)R2 [℘(z)]
where R1 , R2 are rational functions and ℘ has the same set of periods as f .
Exercise 9 ([Apostol] page 24, Exercise 9)

(℘(z)2 + 14 g2 )2 + 2g3 ·℘(z)


℘(2z) =
4℘3 (2) − g2℘(z) − g3
1 ℘00 (z) 2
 
= −2℘(z) +
4 ℘0 (z)

Exercise 10 Show that


1
℘00 (z) = 6℘(z) − g2
2
Can you compute ℘00 (z) in terms of ℘(z),℘0 (z)

2.6 Eisenstein series

The series

1
Gn := ∑ n
, n even, n > 4
ω∈Λ , ω6=0 ω
are called Eisenstein series. From the convergence of ℘ we can derive the conver-
gence of Gn ’s. They satisfy

Gn (λΛ ) = λ −n G(Λ ), ∀λ ∈ C∗ (2.13)

we usually define
1
Gn (τ) = Gn (Zτ + Z) = ∑ , τ ∈H
(a,b)∈Z2 , (a,b)6=(0,0)
(a + bτ)n

Exercise 11 Show that G2 doesn’t converge at any point τ ∈ H. Show also that
Gn ≡ 0, for n an odd number.
From the functional equation (2.13) we deduce the following: For all A ∈ SL(2, Z)
18 2 Modular forms
 
aτ + b
Gn (Aτ) = G
cτ + d
 
aτ + b
=G Z+Z
cτ + d
= (cτ + d)+n G((aτ + b)Z + (cτ + d)Z)
= (cτ + d)n G(τZ + Z)
= (cτ + d)n G(τ)

Later we will see that


lim Gn (τ) exist
Im(τ)→∞

This motivate us to define (holomorphic) modular forms.


Definition 2.6.1 Let k ∈ Z be an integer and f be a meromorphic function on the
upper half plane. Then f is called a meromorphic modular form for the subgroup
Γ ⊆ SL(2, R) if
   
−k aτ + b ab
(cτ + d) f = f (τ) ∀ ∈Γ (2.14)
cτ + d cd
 
11
Let us assume that ∈ Γ . Then
01

f (τ + 1) = f (τ)

Therefore, f defines a meromorphic function g in the punctured disc:

D∗ = z ∈ C| |z| < 1\{0}

which is defined by
f (τ) = f˜(q), where q := e2πiτ
The map H → D∗ is depicted in Figure (*)
We write the laurent expansion of f˜ at q = 0.
n=+∞
f˜(q) = ∑ fn · qn , fn ∈ C (2.15)
n=−∞

This is also called the Fourier expansion of f˜.


Definition 2.6.2 A meromorphic modular form for Γ = SL(2, Z) apart from the
property (2.14) is meromorphic at i∞, that is, in (2.15) we have some M ∈ Z such
that fn = 0 for all n 6 M. It is called weakly modular form if it is holomorphic in
H and possibly meromorphic at i∞. It is called (holomorphic) modular form if it is
holomorphic in H ∪ {i∞}.
From now we use the letter f for f˜ too, and write the q−expansion of a modular
form as
2.7 Fourier expansion of Eisenstein series 19

Fig. 2.3 q map

∞ ∞
f= ∑ fn · qn = ∑ fn · e2πiτ
n=0 n=0

It will be clear from the text whether we consider f as a function of τ or q.

2.7 Fourier expansion of Eisenstein series

We know that the Eisenstein series are weakly holomorphic moduler forms. In this
section we show that they are holomorphic at i∞ and so they are moduler forms. The
computation in this section can be found in [koblitz] page 110, [Apostol] page 18,
Serre page 91.
Recall the Bernouli numbers
x ∞
xk
ex − 1
= ∑ Bk · k!
k=0

For instance, B0 = 1, B1 = −1 1 −1
2 , B2 = 6 , B4 = 30 , B6 =
1
42
It is easy to see that for any odd k > 3 we have Bx = 0
Theorem 2.7.1 The Eisenstein series Gk has the following q-expansion
!
2k ∞ n
Gk = 2ζ (k) 1 − ∑ σk−1 (n)q
Bk n=1

1
where ζ (k) := ∑∞
n=1 nk is the Riemann’s zeta function and

σa (n) := ∑ da·
d|n
20 2 Modular forms

We follow [Koblitz] page 110. Let us first state the main ingredient of the proof
of Theorem 2.7.1
Proposition 2.7.1 We have

(2πi) k Bk
ζ (k) = − k even > 2 (2.16)
2 k!
1 (−2πi)k ∞
∑ (a + n)k = (k − 1)! ∑ nk−1 e2πina , k ∈ N, k ≥ 2, a ∈ C − Z (2.17)
n∈Z n=0

Proof. We have the following product formula for sine function


∞ 
a2

sin(πa) = πa ∏ 1 − 2 , a ∈ C (2.18)
n=1 n

Take the logarithmic derivative if (2.18) and we have


∞  
1 1 1
π · cot(πa) = + ∑ + , (2.19)
a n=1 a + n a − n

The left hand side of this is


eπia + e−πia 2πi
π · cot(πa) = πi −πia
= πi + 2πia
e −e
πia e −1
!

= πi − 2πi ∑ e2πina (2.20)
n=0

In (2.19) multiply both side with a and set x := 2πia.


x x x x
+ x = 1+ ∑ +
2 e −1 n=1 (x + 2πin) (x − 2πin)
!
−x k  x k
∞ ∞  
x
= 1+ ∑ ∑ −
n=1 2πin k=0 2πin 2πin
∞ ∞
xk+1
= 1−2 ∑ ∑
n=1 k=1
(2πin)k+1
k odd

xk+1
= 1−2 ∑
k=1
(2πi)k+1 ζ (k + 1)
k odd

We get the well-know formula (2.16). We differentiate (2.19) and (2.20) with respect
to a, k − 1 times, and we get
2.7 Fourier expansion of Eisenstein series 21

1
(−1)(−2) . . . (−k + 1) ∑
(a + n)k
= −(2πi)k
∑ nk−1 e2πina
n∈Z n=0

which is (2.17).

Proof (of Theorem 2.7.1). Take a = mτ, m ∈ Z, m 6= 0, and we have

1 (−2πi)k ∞ k−1 nm
∑ (mτ + n)k = ∑n q
(k − 1)! n=1
n∈Z

The result follows immediately:


∞ +∞
1
Gk = 2ζ (k) + 2 ∑ ∑ (2.21)
m=1 n=−∞ (mτ + n)k
!
(−2πi)k ∞
= 2ζ (k) 1 + ∑ nk−1 qnm t
u (2.22)
ζ (k)(k − 1)! m,n=1

We know that the Eisenstein series Gk for k ≥ and odd number is identically zero
and hence the equality (2.21) is valid for k ≥ 4 an even number. However, note that
the equality (2.22) is valid also for k ≥ 3 an odd number. In this case the number
πk
ζ (k)
is conjecturally a transcendetal number. For this reason we may take (2.21) as
the definition of Gk which works also for odd k ≥ 3.
Exercise 12 ∗ Describe the functional equation of Gk (τ), k ≥ 3 under the action of
SL(2, Z). See the Master thesis of H. Bachmann.
Exercise 13 Prove the product formula for sine in (2.18). Hint: both side have the
same zero set.
We will use the following new notation

2k ∞
Ek = Gk /2ζ (k) = 1 − ∑ σk−1 (n)qn
Bk n=1
(2.23)
!

E4 = 1 + 240 ∑ σ3 (n)qn (2.24)
n=1
!

E6 = 1 − 504 ∑ σ5 (n)qn (2.25)
n=1
!

n
E8 = 1 + 480 ∑ σ7 (n)q (2.26)
n=1
22 2 Modular forms

2.8 The Eisenstein series E2

We follow [Koblitz]p. 112. We define


∞ ∞ 0
1
G2 (τ) = ∑ ∑
m=−∞ n=−∞ (mτ + n)2

Where 0 means that if m = 0 then n 6== 0. The arguments in §2.7 shows that the
inner sums converge for any m and τ ∈ H and then the other sum converges. Here,
the order of summation is important.
Exercise 14 Show that the double sum
1
∑ (mτ + n)2
(n,m)∈Z2 \(0,0)

doesn’t converge.
In a similar way as in §2.7 we get

G2 (τ) = 2ζ (2)E2 (τ), E2 (τ) = 1 − 24 ∑ σ1 (n)qn
n=1

Proposition 2.8.1 We have


 
−2 aτ + b 12 c
(cτ + d) E2 − E2 (τ) = (2.27)
cτ + d 2πi cτ + d
 
ab
for all ∈ SL(2, Z)
cd

Proof. First we note that if we define

f k2 A := (cτ + d)−2 f (Aτ) − c(cτ + d)−1

for a holomorphic function f on H then

( f k2 A) k2 B = f k2 AB

Since PSL(2, Z) is generated by T and S and the (2.27) is trivial for T , it is


enough to verify (2.27) for S, that is
 
−2 −1 12 1
τ E2 = E2 (τ) +
τ 2πi τ
t
u
2.9 The algebra of moduler forms 23

2.9 The algebra of moduler forms

Theorem 2.9.1 The Eisenstein series E4 and E6 are algebraically independent over
C, that is, there is no polynomial P(X,Y ) with coefficients in C such that P(E4 , E6 ) =
0. Moreover any holomorphic moduler form f of weight k can be written uniquely
as
f = P(E4 , E6 )
where P is a homogeneous polynomial of degree k in the ring

C[X,Y ], weight (X) = 4, weight (Y ) = 6 (2.28)

There is a clasical proof of Theorem (2.9.1) which can be found in almost all books
on modular forms. In §3.6 we will give a new proof which is inspired by the author’s
study of the generalized period domain in [Movasati2008]. The proof is based on
the study of elliptic integrals and Gauss-Manin connection.
Note that a homogeneous polynomial P of degree k in the ring (2.28) is of the
form
P(X,Y ) = ∑ ai, j X iY j ai, j ∈ C, i, j ∈ N0
4i+6 j=k

and so P(X,Y ) is clearly a moduler form of weight k for SL(2, Z). Let

M = M(SL(2, Z)) := ⊕k∈Z Mk

be the algebra of moduler forms. By Theorem (2.9.1), for k ∈ Z, k 6 2 or k odd we


have Mk = 0 and
M = C[E4 , E6 ], Mk = C[E4 , E6 ]k
weight (E4 ) = 4, weight(E6 ) = 6

Exercise 15 Show that

E42 = E8 , E4 E6 = E10 , E6 · E8 = E14

and derive the corresponding equalities for 6k (n). For instance


n−1
67 (n) = 63 (n) + 120 ∑ 63 (n − m)
m=1

The dimension of the space of modulex forms Mk is listed below:

k 0 2 4 6 8 10 12 14 16 18 k k + 12
dimMk 1 0 1 1 1 1 2 1 2 2 d d +1
Note that
dimMk = ]{(x, y) ∈ N2 /4x + 6y = k}
It is easy to see that
24 2 Modular forms

1
∑ dimMk qk = (1 − q4 )(1 − q6 ) . (2.29)
k=0

2.10 Ramanujan relations between eisenstein series

The derivation of a modular form with respect to τ is no more a modular form.


Instead, we have
Proposition 2.10.1 We have the following map

∂f k
Mk → Mk+2 , f 7→ − 2πi E2 · f . (2.30)
∂τ 12
which is called the Serre derivative of f
Proof. Let g be the Serre derivative of f . We have to show that g ∈ Mk+2 . Only the
functional equation of g is non-trivial:

k
g(Aτ) = f 0 (Aτ) − (2πi) f (Aτ)E2 (Aτ)
12
= ···
= (cτ + d)k+2 · f (τ)
 
ab
for A = ∈ SL(2, Z).
cd
We have
∂ ∂
:= 2πiq .
∂τ ∂q
and sometime it is useful to divide the Serre derivative over 2πi and redefine it
f 7→ q ∂∂ qf − 12
k
E2 f .

Proposition 2.10.2 we have the following equalities between Eisenstein series and
their derivatives  ∂E 1 2
 q ∂ q = 12 (E2 − E4 )
2

q ∂∂Eq4 = 13 (E2 E4 − E6 ) , (2.31)
q ∂∂Eq6
 1 2
= 2 (E2 E6 − E4 )

Proof. The proof of the second and third equality follows from the Serre derivative.
The proof of the first equality follows in a similar way. We need to prove that f (τ) :=
12 ∂ E2 2
− 2πi ∂ τ + E2 is a modular form of weight 4 and its constant term is 1. Therefore,
by Theorem 2.9.1 it must be E4 . t u
2.11 The product formula for discriminant 25

2.11 The product formula for discriminant

Recall that
π4 π6
ζ (4) = , ζ (6) = ,
90 945
We define
(2π)12 3
∆ := t23 − 27t32 = (2ζ (4)60E4 )3 − 27(2ζ (6)140)2 E62 = (E − E62 )
1728 4
and we have

1
(2π)−12 ∆ = (E43 −E62 ) = ( ∑ τ(n)qn = q−24q2 +252q3 −1472q4 +4830q5 +· · · =
1728 n=1

τ(n) is called the Ramanujan function of n.


Proposition 2.11.1 We have

1
(E43 − E62 ) = q ∏ (1 − qn )24 (2.32)
1728 n=1

Proof. This follows from Ramanujan relations between Eisenstein series. The log-
arithmic derivative of both sides in (2.32) is E2 . t
u

Exercise 16 ([Koblitz] III, §2, 4)

1. Show that
(2πi)−12 · 26 · 35 · 72
E12 − E62 = ·∆
691
2. From this derive an expression for τ(n) in terms of σ11 and σ5
3. Show that

τ(n) ≡ σ11 (n)( mod 691)


The following was conjectured by Ramanujan and proved by Deligne in [Deligne]
as a by-product of his proof of the Weil conjectures.
Theorem 2.11.1 We have 11
|τ(n)| < n 12 σ · (n),
where σ0 (n) is the number of divisors of n.
Another interesting function is
∞ ∞
1728 · q 1
F(q) := = ∏ = ∑ Pn qn
E43 − E62 n=1 (1 − qn ) n=0
26 2 Modular forms

It can be easily checked that Pn is the unrestricted partition function, that is, Pn is the
number of ways a positive integer n can be expressed as a sum of positive integers:

n = a1 + a2 . . . + ak , ak ∈ N.
There is no restriction on k, order of ai ’s, and repetion of ai ’s is allowed. For more
information see Apostol’s book [Apostol] Chapter 5.

2.12 The j function

The following

t23 ∞
j(τ) = 1728 3 2
= ∑cn qn
t2 − 27t3 n=−1
E3 1
= 1728 3 4 2 = + 744 + 196884q + 21493760q2 + 864299970q3 + · · ·
E4 − E6 q

is called the j-function, or Klein’s modular function. It is holomorphic in H and has


a pole of order one at i∞. From the functional equation of Eisenstein series it follows
that j is invariant under the action of SL(2, Z):
 
aτ + b ab
j( ) = j(τ), ∀ ∈ SL(2, Z). (2.33)
cτ + d cd

Theorem 2.12.1 The map j : SL(2, Z)\H → C is one to one and surjective.
There is a beautiful history behind the j-function. According to [Apostol], Berwick
in 1916 calculated the first seven coefficients of j, Zuckerman the first 24 in 1939,
and Van Wijngaarden the first 100 in 1953. The only reason for computing such
numbers, seems to be only the joy of playing with them and their mysteriousness.
In [Apostol] we find also some divisibility properties of cn ’s due to D.H. Lehmer in
1942 and J. Lehner in 1949. An asymptototic formula due to Petersson in 1932 and
Rademacher in 1932 is also reported in [Apostol].
In 1978 MacKay noticed that 196884 = 196883 + 1 and 196883 is the number
of dimensions in which the Monster group can be most simply represented. Based
on this observation J.H. Conway and S.P. Norton in 1979 formulated the Monstrous
moonshine conjecture which relates all the coefficients in the j-function to the rep-
resentation dimensions of the Monster group. In 1992 R. Borcherds solved this con-
jecture and got fields medal, see [Gan06] for more information on this conjecture.
There proof does not give any clue why elliptic curves must have something to do
with the Monster group, and so the mystery involved around it still exists. For in-
stance, in a private conversation J.H. Conway expressed the fact that the proof for
him is not satisfactory.
2.14 The numbers e1 , e2 , e3 27

2.13 Poincaré metric

In the upper half plane H one usually use the Poincaré metric:

dx ⊗ dx + dy ⊗ dy dz ⊗ d z̄
ds = = ℑ(ω), ω :=
y ℑ(z)

ω is called the Hermitian form.


Exercise 17 The Poincaré metric is invariant under the action of SL(2, R). Hint:
Show that
d(Az) ⊗ d(Az) dz ⊗ d z̄
=
ℑ(Az) ℑ(z)
The volume form of the Poincaré metric is given by the imaginary part of ω:

−dx ⊗ dy + dy ⊗ dx dx ∧ dy 1 dz ∧ d z̄
dV = ℑ(ω) = = =
y y −2i ℑ(z)

We use the Poincaré metric to prove the convergency of Eisenstein series.

2.14 The numbers e1 , e2 , e3

The following functions are defined in [Apostol]


ω1 
e1 := ℘ Zω1 + Zω2 , 2

ω2 
e2 := ℘ Zω1 + Zω2 , 2

ω1 +ω2 
e3 := ℘ Zω1 + Zω2 , 2

Proposition 2.14.1 The numbers e1 , e2 , e3 are distinct and we have

4℘3 (z) − g2 ·℘(z) − g3 = 4 ℘(z) − e1 ℘(z) − e2 ℘(z) − e3


  
(2.34)

Proof. Since ℘(z) is even, ℘0 (z) is odd. Therefore,


       
1 1 1 1
−℘0 ω = ℘0 − ω = ℘0 ω − ω = ℘0 ω ∀ω ∈ Λ
2 2 2 2

This implies that ω21 , ω22 , ω1 +ω


2
2
are roots of ℘0 (z). The function ℘0 (z) has a pole of
order 3 at z = 0 ∈ E and so the mentioned three points are simple roots of ℘0 (z). The
differential equation of ℘(z), implies that e1 , e2 , e3 are roots of the left hard side of
2.34 Now e1 , e2 , e3 are distinct. We showthat e1 6= e2 . The elliptic function℘(z)−ei
has a double root at ω21 , because ℘0 12 ωi = 0, all these for i = 1, 2. If e1 = e2 then
this function must have pole order > 4 at z = 0, which is a contradiction. t u
28 2 Modular forms

Now let us regard ei ’s as functions in τ ∈ H and redefine


   
 τ 1 τ +1
e1 = e1 Zτ + Z, , e2 = e2 Zτ + Z, , e3 = e3 Zτ + Z,
2 2 2

and consider them as holomorphic functions in τ. [PUT HERE t1 ,t2 ,t3 of Darboux-
Halphen, We have ti = ei + g1 i = 1, 2, 3, See [Quasi-Moduler forms, I]].

2.15 Growth of coefficients

We follow [Serre].
Theorem 2.15.1 (Hecke) If f is a holomorphic cusp form for a group SL(2, Z) of
weight k then
fn = O(nk )

where f = ∑ fn qn is the q-expansion of f .
n=1

Proof. Cauchy residue formula implies that

1 dq
Z
fn = f (q)q−n
2πi δ q

Where δ is a small circle turning around q = 0 ∈ C anticlockwise. This is


Z Z 1
−2πnτi
fn = f (τ)e dτ = e 2πyn
f (x + iy)e−2πinx dx (2.35)
0
 k

The function f (τ)(Imτ) 2 is invariant under the action of SL(2, Z), and so, it
gives us a function in SL(2, Z)\H. Since

| f (τ)| ∼ O(q) ∼ O(e−2πy )

This function sis bounded when y → ∞, and so, there exists a constant M such that
k
| f (τ)| 6 My− 2 ∀ ↑∈ H

Putting this in 2.35 we have


k
| fn | 6 e2πyn · M · y− 2
1
Here, y can be any positive number, we put y = n and get the desired result. 

As a corollary we get
2.16 Dedekind eta function 29

Proposition 2.15.1 Let f be a holomorphic modular form of weight k for SL(2, Z).
Then
fn = O(nk−1 )
Proof. Let σk be the Einstein series of weight k, and λ ∈ C, be constant such that
λ σk + f is a cusp form. The proposition follow from theorem 2.15.1 and the asymp-
totic behaviour of Forier coefficients of σk :

σk−1 (n) = ∑ d k−1 = O(nk−1 )


d|n

Because
 k−1
σk−1 (n) d
=∑
nk−1 d|n
n
 k−1 ∞
1 1
=∑ 6 ∑ k−1 = ξ (k − 1) < ∞
d|n
d d=1 d


Since n 2k
compared to nk−1
is negligible, we get the result. Deligne in [Deligne
1973] has shown that for a cusp form f , we have
 k 1 
fn = O n 2 − 2 σ0 (n) (2.36)

Where σ0 (n) is the number of positive divisors of n. This implies that


 k 1 
fn = O n 2 − 2 +ε , ∀ε > 0

According to [Milne], Deligne in 1969 proved that 2.36 follows from the Weil con-
jectures for varieties over finite fields and he proved the Weil conjectures in 1973.

2.16 Dedekind eta function

According to [Apostol] the Dedekind eta function

2πiτ

1 − e2πinτ

η(τ) := e 12

n=1

was introduced by Dedekind in 1877. We know that

∆ (τ) = 2πi)12 η 24

and the functional equation of ∆ with respect to the action of SL(2, Z). Taking 12th
root of this we get
30 2 Modular forms
 
aτ + b 1
η = ε(cτ + d) 2 η(τ)
cτ + d
for som ε which is a 12th root of unity and depends only on A and τ. In fact
Theorem 2.16.1 (Dedekind functional equation) We have

aτ + b
  1
2
η = ε(A) − i(cτ + d) η(τ)
cτ + d
 
ab
For all A = ∈ SL(2, Z) with c > 0, where
cd
  
a+d
ε(A) : = exp πi + S(−d, c)
2c
k−1   
γ h 1
S(d, c) : = ∑ γ· −
γ=1 k r 2
a : = a − [a]
 
0 1
For a proof see [Apostol] Theorem 3.4. For A = we get
−1 0
 
−1 1
η = (−iτ) 2 η(τ)
τ

2.17 Modular forms as k-fold differential

Let f be a meromorphic function in H. Show that the k-fold differential

f (τ) dτ ⊗ dτ ⊗ · · · ⊗ dτ
| {z }
k−times

is invariant under SL(2, Z), ( and hence it induces a k-fold differential in SL(2, Z)\H)
if and only if f is a meromorphic modular form of weight 2k for SL(2, Z)

2.18 Petersson Scalar product

we follow [Lang] page 37. Our main theorem in this section is that the space of
cusps of weight k for SL(2, Z) has a basis of eigenforms. For this we have to put a
Hermitian form on Sk . This will be the Petersson scalar product. Let Γ ⊆ SL(2, Z)
be a sugroup of finite index and let f , g be two holomorphic functions on H. We
define
2.18 Petersson Scalar product 31

1 dxΛ dy
Z
h f , gi := f (τ) g(τ) yk
[SL(2, Z) : Γ ] F y2
dxΛ dy i dτΛ dτ
where τ = x + iy. Note that y2
= 2 Im(τ)

dτΛ dτ
1. The Im(τ)2
is invariant GL+ (2, R) action. This follows from

d(Aτ) = (cτ + d)−2 det(A) · dτ


(2.37)
Im(Aτ) = Im(τ) · det(A)|cτ + d|2

Let us define
k
f |k A = f (Aτ)(cτ + d)−k (det A) 2
i dτΛ dτ
Ω ( f , g) : = f (τ) g(τ) Im(τ)k ·
2 Im(τ)2

We have 
Ω f |k A, g|k A = Ω ( f , g)(Aτ) (2.38)
Which follows from 2.37.
2. We will take f , g two modular form of weight k for a subgroup Γ ⊆ SL(2, Z) of
definite index. In this way for A ∈ Γ , Ω ( f , g) is invariant under the action of Γ
and so it gives a differential form on Γ \H. The integration is over this space/a
fundamental domain of Γ
3. The factor [SL(2, Z) : Γ ] is inserted so that h f , gi becomes independent of the
group Γ , that is if f , g are modular forms of weight k for Γ1 ⊆ Γ2 then h f , giΓ1 =
h f , giΓ2
4. we have to check that the integral is convergent. For this we have to assume that
f , g are cusp forms.
The fundamental domain for Γ is a union of finite number of translates of the
classical fundamental domain Iˇ for SL(2, Z)

F= UAF̌
A∈Γ \SL(2,Z)

It is enough to show that th integral converges for Im(τ) → +∞. For a cusp form
f we have
| f (τ)| << e−cIm(τ)
for sufficiently large Im(τ), and so, if one for g is a cusp form, then the integral
converges
For A ∈ GL+ (2, R) Let us define

A0 := A−1 · det(A)

Theorem 2.18.1 Let f , g be cusp forms of weight k. Let also A ∈ GL+ · (2, Q). We
have
32 2 Modular forms

1. h f |k A, g|k Ai = h f , gi
2. h f |k A, gi = h f , g|k Ai
3. The scalar product in 2.38 depends only on double cosets Γ AΓ

Proof.
1
Z 
h f |k A, g|k Ai = Ω f |k A, g|k A
[SL(2, Z)Γ ] F
Z Z
= 00 Ω ( f , g)(Aτ) = 00 Ω ( f , g)
F AF

Here, F is a fundamental domain for Γ , and so AF is a fundamental domain for


AΓ A−1 . We take smaller Γ such that

Γ , AΓ A−1 ⊆ SL(2, Z)

This gives us
[SL(2, Z) : Γ ] = [SL(2, Z) : AΓ A−1 ]
and the first part follows.
The second part follows from the first part and

g|k A0 = g|k A−1




Note that det(A0 ) = det(A) and here we use the new definition of the slash operator.

Theorem 2.18.2 The Hecke operators acting on Sk SL(2, Z) are Hermitian with
respect to the Petersson scalar product that is

hTn f , gi = h f , Tn gi

Proof. For A ∈ Mat n (2, Z) we have

h f |k A, gi = h f , g|k A0 i

SL(2, Z)Mat n (2, Z) → SL(2, Z)Mat n (2, Z)


The map
A 7→ A0
is a bijection


Proposition 2.18.1 The eigenvalues of Tn0 s are totally real algebraic numbers 2.
The space of cusp form has a basis of eigenforms.
Chapter 3
Elliptic curves and integrals

Although most of the seminars I couldn’t understand, after 10 times I started to get
something and that something could be very useful for my development in mathe-
matics or even to physics eventually, (S.-T. Yau in Kavli IPMU News No. 33 March
2016).

3.1 Introduction

In this chapter we study elliptic curves over complex numbers and the corresponding
elliptic integrals.

3.2 Elliptic integrals

We start with an elliptic integral of the form


Z b
dx
p , (3.1)
a p(x)

where p(x) is a polynomial of degree 3 and with three distinct real roots, and a, b
are two consecutive elements among the roots of p and ±∞. For instance, the poly-
nomial p(x) := 4x3 − t2 x − t3 , t2 ,t3 ∈ C has three distinct roots if and only if
∆ := 27t23 − t32 6= 0. If p(x) has repeated roots one can compute it easily.
Exercise 18 Compute the indefinite integral
Z
dx
p , (3.2)
p(x)

33
34 3 Elliptic curves and integrals

Fig. 3.1 The elliptic curve y2 = p(x) in the four dimensional space C2 .

where p is a polynomial of degree 1 and 2. Compute it also when p is of degree 3 but


it has double roots. These integrals are computable because y2 = P(x) is a rational
curve! Let p be a polynomial of degree 3 and with three real roots t1 < t2 < t3 . Show
that two of the four integrals
Z t1 Z t2 Z t3 Z +∞
dx dx dx dx
p , p , p , p ,
−∞ p(x) t1 p(x) t2 p(x) t3 p(x)

can be computed in terms of the other two.


In many calculus books we find tables of integrals and there we never find a
formula for elliptic integrals. Already in the 19th century, it was known that if we
choose p randomly (in other words for generic p) such integrals cannot be calculated
in terms of until then well-known functions. For particular examples of p we have
some formulas calculating elliptic integrals in terms of the values of the Gamma
function on rational numbers.
Exercise 19 For particular examples of polynomials p of degree 3, there are some
formulas for elliptic integrals 3.2 in terms of the values of the Gamma function on
rational numbers. For instance, verify the equality
+∞
Γ ( 17 )Γ ( 27 )Γ ( 74 )
Z
dx
√ = √ . (3.3)
7 x3 − 35x − 98 2πi −7
In [Wal06] page 439 we find also the formulas
3.2 Elliptic integrals 35
1
Γ ( 1 )2
Z
dx
√ = 4 31 ,
0 1 − x3 23 32 π
Z 1
dx Γ ( 14 )2
√ = 3 1 .
0 x − x3 22 π 2
These formulas can be also derived using the software Mathematica. The Chowla-
Selberg theorem, see for instance Gross’s articles [Gro78, Gro79], describes this
phenomenon in a complete way. The right hand side of (3.3) can be written in terms
of the Beta function which is more natural when one deals with the periods of alge-
braic differential forms.
The fact that we only need two of the integrals in (3.1) in order to calculate the
others, can be easily seen by considering the integration in the complex domain
x ∈ C, in which we may discard the assumption that p has only real roots. The
integration is done over a path γ in the x ∈ C domain which connects two points
in the set of roots of p and ∞, and avoids other roots except at its start and end
points. An amazing fact that we learn in a complex analysis course is that if the path
γ moves smoothly, without violating the properties as before, then the value of the
integral does not change. This is certainly the origin of homotopy theory, or at least
one of them. The next step in the study of elliptic integrals is the invention of the y
variable which is basically the square root of p(x):

E := (x, y) ∈ C2 | y2 = p(x) .

(3.4)

This is called an elliptic curve in Weierstrass form.

Theorem 3.2.1 The set E as a toplogical space is a compact torus mines one point,
see Figure 3.1.

Proof.

We add another point O to E and will call it the point at infinity. We write Ē =
E ∪ {O} and sometimes by abuse of notation use the same letter E for Ē. If we write
the equation of E in homogeneous coordinates [x : y : z] ∈ P2 then

O = [0 : 1 : 0].

We define H1 (E, Z) as the abelization of the fundamental group of E, that is, the
quotient of the fundamental group of E by its subgroup generated by commutators:

H1 (E, Z) := π1 (E, b)/[π1 (E, b), π1 (E, b)], (3.5)

where for a group G, [G, G] is the subgroup of G generated by the commutators


aba−1 b−1 , a, b ∈ G. It turns out that the integrals
Z
ω, δ ∈ H1 (E, Z).
δ
36 3 Elliptic curves and integrals

are well-defined. The following Proposition can be proved without the help of The-
orem 3.2.1.
Proposition 3.2.1 The abelian group H1 (E, Z) is free of rank 2, and hence, it is
isomorphic to (Z2 , +).

Proof (Sketch). We prove that the non-abelian group π1 (E, b) is free and it is gen-
erated by two elements. Let π : E → C be the projection into x-coordinate and a be
the x-coordinate of b. The map π is 2 to 1, except at (t1 , 0), (t2 , 0), (t3 , 0), where ti ’s
are the root of p(x). An element δ of the homotopy group π1 (E, b) can be identified
with δ := π(γ) p∈ π1 (C\{t1 ,t2 ,t3 }, a) which has this property that the multivalued
function y := p(x) along γ is one valued. The closed paths γ1 and γ2 in Figure 3.1
are in the image of π, let us say π(δi ) = γi , i = 1, 2, and δ1 , δ2 generate π1 (E, b)
freely. t u

Another important ingredient of H1 (E, Z) is the skew symmetric intersection bilin-


ear map
H1 (E, Z) × H1 (E, Z) → Z
The cycles δ1 and δ2 can be choosen in such a way that hδ1 , δ2 i = −1.
Proposition 3.2.2 Let δ1 and δ2 be generators of H1 (E, Z) with hδ1 , δ2 i = −1. Then
the quotient R dx
δ1 y
τ := R dx
δ2 y

has positive imaginary part.

Proof (sketch). First we note that ω := dx


y restricted to E is holomorphic even at the
infinity point O. The statement follows from
√ √ Z
Z Z Z Z 
−1 ω ω− ω ω = −1 ω ∧ ω̄ > 0,
δ1 δ2 δ2 δ1 E

where ω = dx y . The equality follows from Stokes theorem for the complement of δ1

and δ2 in E. In local holomorphic coordinate system z = x1 + −1x2 in E we have
ω = f (z)dz, f (z) holomorphic, and

ω ∧ ω̄ = | f (z)|2 dz ∧ d z̄ = − −1| f (z)|2 dx1 ∧ dx2 .

t
u

Definition 3.2.1 The lattice of elliptic integrals is


Z
dx dx
Z Z
=Z +Z ω2 ,
H1 (E,Z) y δ1 y δ

where δ1 , δ2 is a basis of H1 (E, Z) with hδ1 , δ2 i = −1.


3.4 Picard-Lefschetz theory 37

3.3 Elliptic curves in Weierstrass format

We consider again the familly of elliptic curves

Et = Et2 ,t3 = (x, y) ∈ C2 y2 = 4x3 − t2 x − t3 , t ∈ T



(3.6)
T := C2 − {(t2 ,t3 ) ∈ C2 | ∆ = 0}, ∆ := t23 − 27t32 . (3.7)

The curve Et is called an elliptic curve in the Weierstrass format.

3.4 Picard-Lefschetz theory

By Ehresmann’s theorem the fibration Et ,t ∈ T is a C∞ bundle over T , i.e. it is lo-


cally trivial. This is the basic stone for the Picard-Lefschetz theory (see for instance
[Mov04] and the references there). It gives us the following linear action:

π1 (T, b) × H1 (Eb , Z) → H1 (Eb , Z)

where b ∈ T is a fixed point In order to calculate it we proceed as follows: First


we choose two cycles δ1 , δ2 ∈ H1 (Eb , Z). For the fixed parameter t2 6= 0, define the
function f in the following way:

f : C2 → C, (x, y) 7→ −y2 + 4x3 − t2 x.


q
t3
The function f has two critical values given by t˜3 , tˇ3 = ± 272 . In a regular fiber Et =
f −1 (t3 ) of f one can take two cycles δ1 and δ2 such that hδ1 , δ2 i = 1 and δ1 (resp.
δ2 ) vanishes along a straight line connecting t3 to t˜3 (resp. tˇ3 ). The corresponding
anti-clockwise monodromy around the critical value t˜3 (resp tˇ3 ) can be computed
using the Picard-Lefschetz formula:

δ1 7→ δ1 , δ2 7→ δ2 + δ1 ( resp. δ1 7→ δ1 − δ2 , δ2 7→ δ2 ).

It is not hard to see that the canonical map π1 (C\{t˜3 , tˇ3 }, b) → π1 (T,t) induced by
inclusion is an isomorphism of groups and so the image of the monodromy group
written in the basis δ1 and δ2 is:
   
10 1 −1
hA1 , A2 i = SL(2, Z), where A1 := , A2 := .
11 0 1
   
0 1 1 1
Note that g1 := A−1 2 A −1 −1
1 A 2 = , g2 := A−1 −1
1 A2 = and SL(2, Z) =
−1 0 −1 0
hg1 , g2 | g21 = g32 = −Ii, where I is the identity 2 × 2 matrix.

Exercise 20 Discuss the Picard-Lefschetz theory as above for the Legendre family
of elliptic curves:
38 3 Elliptic curves and integrals

y2 = x(x − 1)(x − λ )

3.5 Weierstrass uniformization theorem

Let
t := (g2 , g3 ) = (60G4 (Λ ), 140G6 (Λ )),
where G4 and G6 are complex numbers defined in (2.9). From Proposition 2.14.1 it
follows that g32 − 27g23 6= 0 and so t ∈ T, where T defined in (3.7) (do not use the
product formula for the discriminant in (2.32)). Let Et be the corresponding elliptic
curve in (3.6). From the differential equation of the Weierstrass ℘ function, see
Theorem 2.5.1, it follows that we have a map

f : C/Λ −→ Et
(3.8)
z 7−→ (℘(Λ , z),℘0 (Λ , z)), f (0) = O.

Theorem 3.5.1 The map (3.8) is an isomorphism of sets.


Actually, f is an isomorphism of Riemann surfaces. Moreover, we will see that Et
has a structure of a group and it is also a morphism of groups.
Proof. The heart of the proof is Theorem 2.4.1. Let (x, y) ∈ Et . The elliptic function
℘(z) − x has a pole of order two at z = 0. Therefore, it must have two zeros z1 , z2 in
C\Λ . By the differential equation of ℘, we know that {℘0 (z1 ),℘0 (z2 )} = {y, −y}.
If y 6= 0 then there is exactly one of zi ’s, let us say z1 , such that ℘0 (z1 ) = y. If y = 0
then ℘ has a zero of multiplicity 2 at z1 and hence z1 = z2 in C/Λ . This argument
proves that f is one to one and surjective. t u
Let δ1 and δ2 be closed paths in C/Λ which are the images of the vectors ω1 , ω2 ∈ C
under the canonical map C → C/Λ . We also use the same notation for their images
in Et under the map (3.8). We have

dx
Z
= ωi , i = 1, 2. (3.9)
δi y

In particular, the lattice of elliptic integrals for Et as above is Λ . The inverse of f in


(3.8) can be given explicitly as follows:

f −1 : Et2 ,t3 → C/Λ (3.10)


Z P
dx
f −1 (P) :=
O y

The integration can be interpreted in the following way. We take a path in the x-
plane which connects the infity to the x-coordinate of P. We also choose a branch of
dx √dx . In geometric terms, this is to say, we connect P to the point at infinity
y = P(x)
3.6 Sketch of the proof of Theorem (2.9.1) 39

O of E and we integrate dx
y over this path. The map (3.10) is well-defined. We claim
−1
that f is the inverse of f .
Z f (z̃)
dx
f −1 ◦ f (z̃) =
O y
Z z̃ Z z̃
d℘(z)
= = dz = z̃
O ℘0 (z) O

Let P be the space of lattices.


Theorem 3.5.2 The map given by

p : C2 \{t23 − 27t32 = 0} → P

dx
Z
(t2 ,t3 ) 7→
H1 (Et ,Z) y

is well-defined and it is a biholomorphism which satisfies

p(t2 λ −4 ,t3 λ −6 ) = λ p(t2 ,t3 ), λ ∈ C, λ 6= 0.

Its inverse p−1 is given by the Eisenstein series:

Λ → (g2 (Λ ), g3 (Λ )) = (60G4 , 140G6 )

Proof. The fact that p is well-defined follows from Theorem 3.2.2. From Theorem
3.8 we can easily derive p ◦ p−1 = Id which implies that p−1 is injective. The equal-
ity p−1 ◦ p = Id is equivalent to the fact that p is injective. This in turn is equivalent
to the injectivity of j : SL(2, Z)\H → C.[Proof of this fact!!!!]
Exercise 21 Ex. 6.2, 6.4,6.6,6.7,6.14 of [Sil92].

3.6 Sketch of the proof of Theorem (2.9.1)

We regard modular forms as functions on the space of lattices P. Under the bijec-
tion p any moduler form of weight k becomes a holomorphic function f in T with
the property
f (t2 λ 4 ,t3 λ 6 ) = λ k f (t2 ,t3 ) (3.11)
We use the growth condition of f and prove that f is a polynomial of degree k in
the weighted ring

C[t2 ,t3 ], weight (t2 ) = 4, weight (t3 ) = 6.

Since the lattices associated to τ and τ +1 are the same, we get a map D−{0} → P,
where D is the disc of radius 1 and center 0 in C. We compose it with p−1 and get
the map
40 3 Elliptic curves and integrals

Fig. 3.2 Discriminant

i : D → C2 , q 7→ (g2 (q), g3 (q))


Note that this map is even defined at 0 ∈ D and the image of 0 is in the discrim-
inant locus {∆ = 0}. The growth condition of f implies that f |Im(i) extends as a
holomorphic function at i(0). Now, the C∗ -action in C2 implies that f extends to
a holomorphic function in C2 \{(0, 0)}. By Hartogs theorem we conclude that f is
holomorphic in C2 . We write the Taylor series of f at (0, 0) and (3.11) implies the
des

3.7 Some identities

Let Et , t = (t2 ,t3 ) be an elliptic curve in Weierstrass format and let P = (x(P), y(P))
be a point of Et . In Theorem 3.5.1 in order to prove f ◦ f −1 = Id, we need to prove
that:
!−2  !−2 
ZcP Z P −2 Z P ZcP
dx dx dx dx
x(P) = +  ∑ − − 
O y O y O y O y
Z P
−3
dx
y(P) = (−2) ∑ O y

where the sum is taken over all, except one, non-homotopic paths in E wich connect
O to P and OP means integration over this path. In the above formulas the integration
R
3.8 Schwarz function 41

over the exceptional path is denoted by OP . It is easy to see that this doesn’t depend
Rc

on the choice of exceptional path.


The Eisenstein series can be written in the following way. Let δ0 ∈ H1 (E, Z) be
a primitive element, that is, it is not divisable by ab integer. We have
Z −k
ζ (k) · ∑ ω = Gk (E, ω), k ≥ 4, k even (3.12)
δ
δ a monodromy of δ0

where the sum runs in all monodromies δ ∈ H1 (E, ω) of δ0 . Since the monodromy
group of the Weierstrass family is SL(2, Z), we can also take the sum over all prim-
itive elements of H1 (E, Z). The sums
Z −k
∑ ω , k≥3 (3.13)
δ
δ a monodromy of δ0 , hδ ,δ0 i>0

are related to the discussion after Theorem 2.7.1. These functions seem to give an
embedding of the universal cover of C2 \{27t23 − t32 = 0} inside some affine space.
The following sum might be also interesting:
Z −k
∑ ω . (3.14)
δ
δ a clockwise monodromy of δ0

3.8 Schwarz function

Let us take the Legendre family of elliptic curves and consider the Schwarz function
R dx
δ1 y
λ 7→ R dx
∈H
δ2 y

It is multivalued beacuse of the choice of δ1 , δ2 . In order to get a one valued function


we restrict λ to H and choose cycles δi , i = 1, 2 such that the projection of δ1 (resp.
δ2 ) to the x-palne is a cycle around 0 (resp. 1 )and λ . In this way the Shwarz function
is a biholomorphisim. Its analytic continuations around λ = 0, 1 corresponds to the
Picard-Lefschetz transformation of δ1 and δ2 .
Exercise 22 Show that the image of the Schwarz function is the region depicted in
3.3i . The analytic continuations of the Schwarz map gives us the trianglization of
H.

i Reproduced from Wikipedia


42 3 Elliptic curves and integrals

Fig. 3.3 Fundamental domain

3.9 CM elliptic curves

Recall that P is a complex manifold whose points are lattices Λ = Zω1 + Zω2 . In
fact one can reinterpret it as follows:
Exercise 23 P is the moduli of triples (E, ω, p), where E is a Riemann surface of
genus one, p ∈ E, and ω is a holomorphic differential form on E. The canonical
action of a ∈ C∗ on P corresponds to the multiplication of ω with a−1 .
Definition 3.9.1 The endomorphisim group of an elliptic curve E = EΛ is the set of
all holomorphic maps EΛ → EΛ which are in addition homomorphisims of groups.
It is in one to one correspondance with

End(E) = {α ∈ C | αΛ ⊂ Λ }

We have Z ⊂ End(E) and we say that E is CM if the inclusion is strict.


Later, we will encounter two spcial CM elliptic curves as follows:
Exercise 24 Classify all elliptic curves E with α ∈ End(E) which is not multipli-
cation by ±1 and is an isomorphism. More precisely show that we have only two
such elliptic curve √
−1 + i 3
E = Ehz,1i , z = i, .
2
Exercise 25 Ex. 8,9,10,12 of Koblitz.

3.10 Fourier expansions and elliptic integrals

Let Ez , z ∈ P0 be a family of elliptic curves with a singular fiber at z = 0. Further,


assume that the monodromy

H1 (Ez , Z) → H1 (Ez , Z)

For z near 0 and in a basis δ1 , δ2 ∈ H1 (E, Z), hδ1 , δ2 i = −1 is given by


3.10 Fourier expansions and elliptic integrals 43
    
δ1 11 δ1

δ2 00 δ2

Moreover, for a holomorphic differential form


Z
I2 : = ω = holomorphic at z = 0 and I2 (0) 6= 0
δ2

ln z + Ie1
I1 =
2πi
I1 = holomorphic at z = 0 and Ie1 (0) 6= 0
e

For an explicit example see [Movasati 2012]. The mirror map is defined in the fol-
lowing way !
I1 1 Ie1
τ= = ln z +
I2 2πi Ie2
For z near to O and taking the branch of ln z with 0 < Im(ln z) < 2π, τ is near i∞ and
it is in the band
{τ ∈ C|0 6 Re(τ) 6 1, Im(τ) > 0}
For a modular form f we use 2.36 and we get
   
I1 (z) −2πin II1 (z) I1
Z
fn = f e 2 d (z)
γ I2 (z) I2
−n
m
˜
  ∞ 
I1 Z I1 I2 θ I1 − I1 · θ I2 dz
Z
= f −∑ −n
γ I2 m=0 m! I2 I22 Z

(−n)m  
= ∑ Residue gn, m(z), z = 0 , (3.15)
m=0 m!
 m
I1 I2 · θ I1 − Ii θ I2 I˜1
 
gn, m(z) := f Z−n
I2 I22 I2
 
In many concrete examples, f II12 turns out to be a rational function in I2 , θ I2 , Z.
More over, for large m, gn, m(z) turns out to to be holomorphic at z and so the sum
in 3.15 is a actually finite.
Chapter 4
Rudiments of Algebraic Geometry of curves

Throughout the present text we work with a field k of arbitrary characteristic and
not necessarily algebraically closed. By k̄ we mean the algebraic closure of k. The
main examples that we have in mind are

Z
k = Q, R, C, F p := , k̃(t)
pZ
and a number field. A number field k is a field that contains Q and has finite
dimension, when considered as a vector space over Q. A function field k̃(t) =
1 ,t2 ,··· ,ts )
k̃(t1 ,t2 , . . . ,ts ) over a field k̃ is the field of rational functions a(t
b(t ,t ,...,ts ) , where a and
1 2
b are polynomials in indeterminates t1 ,t2 , . . . ,ts and with coefficients in k̃. Later, we
will also use the field of p-adic numbers.

4.1 Curves

Let k be a field and k[x, y] be the space of polynomial in two variables x, y and with
coefficients in k. The n dimensional affine space over k is by definition

An (k) = k × k × · · · × k, n times

and the projective n dimensional space is

Pn (k) := An+1 (k) − {(0, 0, · · · , 0)}/ ∼

a ∼ b if and only if ∃λ ∈ k, a = λ b
We will consider the following inclusion

An (k) → Pn (k), (x1 , x2 , · · · , xn ) 7→ [x1 ; x2 ; · · · ; xn ; 1]

45
46 4 Rudiments of Algebraic Geometry of curves

and call Pn the compactification of An . The projective space at infinity is defined to


be
Pn−1 n n
∞ (k) = P (k) − A (k) = {[x1 ; x2 ; · · · ; xn ; xn+1 ] | xn+1 = 0}.

For simplicity, in the case n = 1, 2 and 3 we use x, (x, y) and (x, y, z) instead of
x1 , x2 , . . ..
Any polynomial f ∈ k[x, y] defines an affine curve

C(k) := {(x, y) ∈ k2 | f (x, y) = 0}.

The most famous Diophantine curve is give by f = xn + yn − 1. We denote it by Fn .


Remark 4.1.1 The set C(k) may be empty, for instance take k = Q, f = x2 +y2 +1.
This means that the identification of a curve with its points in some field is not a
good treatment of curves. One of the starting points of the theory of schemes is this
simple observation.
For f ∈ k[x, y] we define the homogenization of f
x y
F(x, y, z) = zd f ( , ), d := deg( f ).
z z

F defines a projective plane curve in P2 (k):

C̄(k) := {[x; y; z] ∈ P2 (k) | F(x, y, z) = 0}.

Note that
∀c ∈ k, (x, y, z) ∈ k3 , F(cx, cy, cz) = cd F(x, y, z).
One has the injection
C(k) → C̄(k), (x, y) 7→ [x; y; 1]
and for this reason one sometimes says that C̄(k) is the compactification of C(k).
Let g be the las homogeneous piece of the polynomial f . By definition it is a homo-
geneous polynomial of degree d. The points in

C̄(k) −C(k) = {[x; y] ∈ P1∞ | g(x, y) = 0}

are called the points at infinity of C(k). The set of points at infinity of F̄n is empty if
n is even and it is {[1; −1]} if n is odd.

4.2 Charts

From now on we use the notation C to denote the curve C̄ in the previous section. We
simply say that the curve C in an affine chart is given by f (x, y) = 0. The projective
space P2 is covered by three canonical charts:

αi : A2 (k) ,→ P2 (k)
4.3 Schemes 47

α1 (x, y) = [x; y; 1], α2 (x, z) = [x; 1; z], α3 (y, z) = [1; y; z].


and the curve in each chart is repectively given by

f1 (x, y) := F(x, y, 1) = 0, f2 := F(x, 1, z) = 0, and f3 := F(1, y, z) = 0.

We are also going to use the notion of an arbitrary curve over k from algebraic
geometry of schemes. Roughly speaking, a curve C over k means C over k̄ and the
ingredient polynomials of C are defined over k. The reader who is not familiar with
those general objects may follow the text for affine and projective curves as above.
The set C(k) is now the set of k-rational points of C.

4.3 Schemes

We defined P2 (k) and C(k) without defining P2 and C. In this section we fill this
gap and we explain the rough idea behind the definition of the schemes P2 and C.
By the affine scheme A2 we simply think of the ring k[x, y]. Open subsets of A2
are given by the localization of k[x, y]. We will need two open subsets of A2 given
respectively by
1 1
k[x, y, ] and k[x, y, ]
y x
By the projective scheme P2 we mean three copies of A2 , namely

k[x, y], k[x, z], k[y, z]

together with the isomorphism of affine subsets:


1 1 x 1
k[x, y, ] ∼
= k[x, z, ], x 7→ , y 7→ (4.1)
y z z z
1 1 1 y
k[x, y, ] ∼
= k[y, z, ], x 7→ , y 7→
x z z z
1 1 1 z
k[x, z, ] ∼
= k[y, z, ], x →7 , z→7
x y y y
The best way to see these isomorphisims is, for instance: we look at an element of
k[x, y] as a function on the first chart A2 (k) and for (a, b) in this chart we use the
identities
a 1 b 1
[a; b; 1] = [ ; 1; ] = [1; ; ].
b b a a
We think of the the scheme C in the same way as P2 , but replacing k[x, y] with
k[x, y]/h f1 i and so on. Here h f1 i is the ideal k[x, y] generated by a single element
f1 . We can also think of C in the same way as P2 but with the following additional
relations between variables:
48 4 Rudiments of Algebraic Geometry of curves

f1 (x, y) = 0 in k[x, y]

f2 (x, z) = 0 in k[x, z]
and
f3 (y, z) = 0 in k[y, z].

Remark 4.3.1 The above discssion does not use the fact that k is a field. In fact, we
can use an arbitrary ring R instead of k. In this way, we say that we have an scheme
C over the ring R.
The function field of the projective space P2 is defined to be

k(P2 ) := k(x, y) ∼
= k(x, z) ∼
= k(y, z).

where the isomorphisms are given by (4.1). The field of rational function on the
curve C is the field of fractions of the ring k[x, y]/h f1 i. Using the isomorphism (4.1),
this definition does not depend on the chart with (x, y) coordinates. We can also think
of k(C) as k(x, y) but with the relation f1 (x, y) = 0 between the variables x, y. Any
f ∈ k(C) induces a map
C(k) → k
that we denote it by the same letter f .

4.4 Singular and smooth curves

Definition 4.4.1 We say that an affine curve given by f (x, y) = 0 is singular if there
is a point (a, b) ∈ k̄2 such that

f (a, b) = fx (a, b) = fy (a, b) = 0.

where fx is the derivation of f with respect to x and so on. The point (a, b) is called
a singularity of the affine curve. A projective curve is singular if one of its affine
charts is singular. For a curve C, affine or projective, we denote by Sing(C) ⊂ C(k̄)
the set of singular points of C.

4.5 Resultant and discriminant

Let f , g ∈ Z[x]. In the order to compute the resultant of f , g we proceed as follows.


We take one of f and g, let us say f , and define

V := Q[x]/ f (x)·Q[x]
4.6 Discriminant 49

This is a vector space of dimension deg( f ) and it has the basis

1, x, x2 , . . . , xdeg f −1 (4.2)

Now consider the linear map

A: V →V
A P(x) := P(x) · g(x)

We write this map in the basis 4.2 and compute its characteristic polynomial

F(s) := det(A − sI)

It might be better from computational point of view to find the minimal polynomial
F(s) of A, that is, a polynomial of minimal degree and with coefficients in Q such
that
F(A) = 0
It follows that  
F g(x) = P(x) · f (x)

for some polynomial P(x) ∈ Q[x]. We have

P(x) · f (x) + Q(x) g(x) = F(0) (4.3)


 
F(0) − F g(x)
where Q(x) = ∈ Q[x].
g(x)
After multiplying 4.3 with a natural number we get an equality

P(x) · f (x) + Q(x) g(x) = ∆

where ∆ ∈ Z, P, Q ∈ Z[x]. We call ∆ the resultant of f , g. The resultant of f (x), f 0 (x)


is called the discriminant of f .

4.6 Discriminant

Let R be a ring and k be the field of fractions of k.


Definition 4.6.1 Let us be given a polynomial f ∈ R[x, y, . . .], where (x, y, . . .) is a
multi variable. The discriminant ideal of f contains all element ∆ ∈ R such that

∆ = f a1 + fx a2 + fy a3 + · · · , for some a1 , a2 , . . . ∈ R[x, y, . . .].

When the discriminant ideal is principal we denote by ∆ its generator and we call
it the discriminant of the polynomial f . It is defined up to units of the ring R, and
50 4 Rudiments of Algebraic Geometry of curves

hence in the case R = Z it is defined up to sign. In this case we fix this ambiguity by
assuming that ∆ is positive.
Exercise 26 For
f = y2 − x3 − t4 x − t6 , t4 ,t6 ∈ R; (4.4)
show that the discriminant ideal is generated by

∆ = 2(4t43 + 27t62 ),

The corresponding a1 , a2 and a3 in this case are given by

a1 = 2(27x3 − 27y2 + (27t4 )x + (−27t6 )),

a2 = 2(−9x4 + (−15t4 )x2 + (−4t42 )), a3 = −54x3 y + 27y3 + (−54t4 )xy.


In this case, ∆ is the resultant of the polynomials P(x) and Px (x), where P = x3 +
t4 x + t6 .
In all the case below, the discriminant ideal is principal and we have calculated its
generator.
f = y2 − x3 − t4 x − t6 − t2 x2 + t1 xy + t3 y. (4.5)
∆ = (t16 t6 − t15 t3 t4 + t14 t2 t32 + 12t14 t2 t6 − t14 t42 − 8t13 t2 t3 t4 − t13 t33 − 36t13 t3 t6 + 8t12 t22 t32 + 48t12 t22 t6 − 8t12 t2 t42

+30t12 t32 t4 − 72t12 t4 t6 − 16t1 t22 t3 t4 − 36t1 t2 t33 − 144t1 t2 t3 t6 + 96t1 t3 t42 + 16t23 t32 + 64t23 t6 − 16t22 t42

−72t2 t32 t4 − 288t2 t4 t6 + 27t34 + 216t32 t6 + 64t43 + 432t62 ).

a1 = 432x3 − 432y2 + (−432t1 )xy + (432t2 )x2 + (−432t3 )y + (432t4 )x + (−t16 − 12t14 t2 + 36t13 t3 − 48t12 t22

+72t12 t4 + 144t1 t2 t3 − 64t23 + 288t2 t4 − 216t32 − 432t6 )

a2 = −144x4 + (−48t12 − 192t2 )x3 + (−t14 − 8t12 t2 − 120t1 t3 − 16t22 − 240t4 )x2 + (−t15 )y + (6t14 t2 − 20t13 t3 +

24t12 t22 − 40t12 t4 − 80t1 t2 t3 + 32t23 − 160t2 t4 )x + (t14 t4 + 4t13 t2 t3 + 8t12 t2 t4 − 16t12 t32

+8t1 t22 t3 − 64t1 t3 t4 + 16t22 t4 − 64t42 ).

a3 = −432x3 y + 216y3 + (−144t1 )x4 + (324t1 )xy2 + (54t12 − 432t2 )x2 y + (−3t13 − 120t1 t2 − 216t3 )x3 +

(324t3 )y2 + (108t1 t3 − 432t4 )xy + (−t15 + 4t13 t2 − 21t12 t3 + 8t1 t22 − 96t1 t4 − 216t2 t3 )x2 + (t16 + 6t14 t2 − 18t13 t3 + 24t12 t22

−36t12 t4 − 72t1 t2 t3 + 32t23 − 144t2 t4 + 162t32 )y + (−t15 t2 + t14 t3 + 2t13 t4 + 4t12 t2 t3 + 8t1 t2 t4 − 27t1 t32 − 216t3 t4 )x

+(−t15 t4 + t14 t2 t3 − 4t13 t2 t4 − t13 t32 + 8t12 t22 t3 + 14t12 t3 t4 − 8t1 t22 t4 − 36t1 t2 t32 + 32t1 t42 + 16t23 t3 − 72t2 t3 t4 + 27t33 );

This modulo 2 is:

∆ = t14t2t32 + t15t3t4 + t16t6 + t13t33 + t14t42 + t34

a1 = t16 , a2 = t14 x2 + t15 y + t14t4


a3 = t15 x2 + t13 x3 + t16 y + t15t2 x + t14t3 x + t12t3 x2 + t14t2t3 + t15t4 + t13t32 + t1t32 x + t33
For the case
f = y2 − x3 − t4 x − t6 − t2 x2 . (4.6)
we have
∆ = 2(4t23t6 − t22t42 − 18t2t4t6 + 4t43 + 27t62 );
4.8 Descriminant of projective curves 51

a1 = 2(27x3 − 27y2 + (27t2 )x2 + (27t4 )x + (−4t23 + 18t2t4 − 27t6 ));


a2 = 2(−9x4 + (−12t2 )x3 + (−t22 − 15t4 )x2 + (2t23 − 10t2t4 )x + (t22t4 − 4t42 ));
a3 = −54x3 y + 27y3 + (−54t2 )x2 y + (−54t4 )xy + (4t23 − 18t2t4 )y;
Modulo 3 this is:
∆ = t22t42 − t23t6 − t43
a1 = t23 , a2 = t22 x2 + t23 x + t2t4 x − t22t4 + t42 , a3 = t23 y.

4.7 The main property of the discriminant

The main property of the singular ideal is


Proposition 4.7.1 Let I be any maximal ideal of R and so R/I is a field. The affine
variety f = 0 is singular over R/I if and only if the discriminant ideal is a subset of
I.
Let us describe our main examples for the above theorem.
1. R = k and so I = {0}.
2. R = Z. This is a principal ideal domain and so the discriminat ideal is generated
by some ∆ ∈ N and I is generated by some prime p ∈ N. In this case, f = 0 is
singular over F p if and only if p | ∆ .
3. R = k[t] and for a ∈ ks , I is the ideal of R generate by ti − ai , i = 1, 2, . . . , s.
Let also assume that the discriminant ideal is generated by ∆ . In this case, the
curve f = 0 with the evaluation of the parameters t = a is singular if and only if
∆ (a) = 0.

Proof. We can assume that R = k is a field.

4.8 Descriminant of projective curves

Let C ⊂ P2 be a curve over the ring R. We define its discriminant ideal to be the
ideal generated by all discriminant ideals of C in some affine coordinates.
Exercise 27 Let us take the curve y2 z − x3 − 17z3 = 0 over Z. Its discriminant in
the affine coordinates z = 1, respectively y = 1, is 2 · 3 · 172 , respectively 24 . Its
discriminant is 24 · 3 · 72 (see the tex file of this text for the corresponding Singular
code).
52 4 Rudiments of Algebraic Geometry of curves

4.9 Curves of genus zero

Let f ∈ k[x, y] and let C be the curve induced by f = 0 in P2 . Let us assume that f
is of degree 2 and it is irreducible over k̄. Further, assume that C(k) has at least one
point P. Note that if we look C in an affine chart then this point can be a point at
infinity. The following procedure finds all the points of C(k). We fix a line L in P2 ,
for instance take y = 0. For any point X ∈ L(k), we connect X to P by a line L0 and
find the second intersction f (X) of L0 with L. Since P ∈ C(k), we have f (X) ∈ C(k)
and we get a bijection
f : L(k) → C(k)

Exercise 28 Use the above geometric argument and find all k-rational points of the
Diophantine equation
tx2 + sy2 = t + s.
for some s,t ∈ k.
The argument discussed in the previous paragraph works if f ∈ k[x, y] is irreducible
over k̄ and has degree 3 and the induced curve C in P2 is singular. In this case, we can
prove that the singularity P of C is defined over k, that is P ∈ C(k), and it is unique.
This point serves us as the point with the same name in the previous paragraph.
Exercise 29 Find all the k-rationa points of the Diophantine equation

y2 − x3 − t4 x − t6 = 0

for some t4 ,t6 ∈ k with 4t43 + 27t62 = 0.

4.10 Curves of genus bigger than one

Let C/Q be a smooth projective curve of degree d in P2 , i.e. its defining polynomial
is of degree d. Its genus is by definition

(d − 1)(d − 2)
g(C) :=
2
The main objective of the Diophantine theory is to describe the set C(Q) for the
curves defined over Q. The most famous example is the Fermat curve given by the
polynomial f = xn + yn − 1. The machinery of algebraic geometry is very useful
to distinguish between various types of Diophantine equations. For instance, one
can describe the rational points of genus zero curves. Genus one curves are called
elliptic curves and the study of their rational points is the objective of the present
text. For higher genus we have a conjecture of Mordell around 1922 which is proved
by Faltings in 1982:
4.12 Elliptic curves in Weierstrass form 53

A non-singular projective curve of genus> 1 and defined over Q has only finitely
many Q-rational points.
In fact, the above theorem is true even for number fields. For instance the above
theorem says that the Fermat curve Fn has a finite number of Q-rational points.
However, it does not say something about the nature of Fn (Q). Mordell’s conjecture
for function fields was proved by Y. Manin in 1963, see [Man63] .

4.11 Elliptic curves in Weierstrass form

In this chapter, we define what is an elliptic curve over a field k, we describe the
Weierstrass format of an elliptic curve and we define the group structure of an ellip-
tic curve. If the reader is not familiar with the notion of an a curve over a field, he
can use the curves in p2 which we worked out in the previous section.
We are ready to give the definition of an elliptic curve:
Definition 4.11.1 An elliptic curve over k is a pair (E, O), where E is a genus one
complete smooth curve and O is a k-rational point of E, that is, O ∈ C(k).
Therefore, by definition an elliptic curve over k has at least one k-rational point. A
smooth projective curve of degree 3 is therefore an elliptic curve if it has a k-rational
point. For instance, the Fermat curve

F3 : x3 + y3 = z3

is an elliptic curve over Q. It has Q-rational points [0, 1, 1] and [1, 0, 1]. However

E : 3x3 + 4y3 + 5z3 = 0

has not Q-rational points and so it is not an elliptic curve defined over Q. It is an
interesting fact to mention that E(Q p ) for all prime p and E(R) are not empty. This
example is due to Selmer (see [Cas66, Sel51]).

4.12 Elliptic curves in Weierstrass form

An elliptic curve in the Weierstrass form E is the affine curve given by the polyno-
mial
Et2 ,t3 : y2 − x3 − t2 x − t3 , t2 ,t3 ∈ k,
∆ := 2(4t23 + 27t32 ) 6= 0,
and in particular k is not of charachteristic 2. In homogeneous coordinates it is
written in the form
zy2 − x3 − t2 xz2 − t3 z3 = 0.
54 4 Rudiments of Algebraic Geometry of curves

It has only one point at infinity, namely [0; 1; 0], which is considered as the marked
point in the definition of an elliptic curve. It is in fact a smooth point of Ē which is
tangent to the projective line at infinity of order 3 and [0; 1; 0] is the only intersection
point of the line at infinity with Ē. If char(k) = 2 then the curve Et2 ,t3 is always
singular. We have already seen in Proposition 4.7.1 that ∆ = 0 if and only if the
corresponding curve is singular.

4.13 Real geometry of elliptic curves

For a projective smooth curve C defined over R the set C(R) has many connected
components, all of them topologically isomorphic to a circle. We call each of them
an oval.
For an elliptic curve E defined over R we want to analyze the topology of E(R).
For simplicity (in fact because of Proposition 4.18.1 which will be presented later)
we assume that E = Et2 ,t3 is in the Weierstrass form. For (t2 ,t3 ) ∈ R2 let ∆ = 2(4t23 +
27t32 ) be the discriminant of the elliptic curve E. We have:
1. If ∆ < 0 then E(R) has two connected components, one is a closed path in R2 ,
which we call it an affine oval, and the other a closed path in P2 (R). We call it a
projective oval.
2. If ∆ > 0 then E(R) has only one component which is a projective oval.
3. If ∆ = 0 and t3 < 0 then E(R) is an α-shaped path in R2 (∞-shaped path in
P2 (R)). In this case, we say that E has a real nodal singularity.
4. If ∆ = 0 and t3 > 0 then E(R) is a union of a point and a projective oval. In this
case, we say that E has a complex nodal singularity.
5. If t2 = t3 = 0 then E(R) look likes a broken line in R2 . In this case, we say that
E has a cuspidal singularity.
Note that E(R) intersects the line at infinity only at [0; 1; 0]. To see/prove all the
topological statements above, it is enough to take an example in each class and
draw the corresponding E(R). Note that in the (t2 ,t3 )-space each set defined by the
above items is connected and the topology of E(R) does not change in each item
(see Figure 4.1 and 4.2, the correspondence between the values of t2 ,t3 and Et2 ,t3 (R)
are done by colours).

4.14 Complex geometry of elliptic curves

Let C be a smooth projective curve of genus g over a subfield k of C. It can be


shown that C(C) is a compact (Riemann) surface with g(C) wholes (a sphere with
g handles).
4.15 The group law in elliptic curves 55

1.6

0.8

-2.4 -1.6 -0.8 0 0.8 1.6 2.4

-0.8

-1.6

Fig. 4.1 Elliptic curves: y2 − x3 − t2 x − t3 = 0

2.5

-5 -2.5 0 2.5 5

-2.5

Fig. 4.2 The discriminant curve 4t23 + t32 = 0

Exercise 30 Give a proof of the above statement using the followings. 2. Any com-
pact Riemann surface is diffeomorphic to a sphere with some handles. 2. Riemann-
Hurwitz formula.
In genus one case, therefore, the set C(C) is torus.
Exercise 31 For a smooth elliptic curve E over R and in the Weierstrass form de-
scribe the real curves E(R) inside the torus (Hint: Use the Riemann-Hurwitz for-
mula).

4.15 The group law in elliptic curves

Let C be a smooth cubic curve in P2 . Let also P, Q ∈ C(k) and L be the line in P2
connecting two points P and Q. If P = Q then L is the tangent line to C at P. The
line L is defined over k and it is easy to verify that the third intersection R := PQ of
C(k̄) with L(k̄) is also in C(k). Fix a point O ∈ C(k) and call it the zero element of
C(k). Define
P + Q = O(PQ)
56 4 Rudiments of Algebraic Geometry of curves

For instance, for an elliptic curve in the weierstrass form take O = [0; 1; 0] the point
at infinity. By definition O + O = O.
Theorem 4.15.1 The above construction turns C(k) into a commutative group.
Proof. The only non-trivial piece of the proof is the associativity property of +:

(P + Q) + R = P + (Q + R)

The proof constitute of three pieces:


1. Let Pi = [xi ; yi ; zi ] be 8 points in P2 (k̄) such that the vectors (xi3 , · · · , z3i ) ∈ k̄10 of
monomials of degree 3 in xi , yi , zi are linearly independent. A cubic polynomial F
passing through all Pi ’s corresponds to a vector a ∈ k̄10 such that Pi · a = 0 and so the
space of such cubic polynomials is two dimensional. This means that there is two
cubic polynomial F and G such that any other cubic polynomial passing through
Pi ’s is of the form λ F + µG and so it crosses a ninth point too.
2. We apply the first part to the eight points O, P, Q, R, PQ, QR, P + Q, Q + R
and conclude that (P + Q)R = P(Q + R). Note that from these 8 points it crosses
there cubic polynomials: C, the product of lines through (0, PQ, P + Q), (R, Q, QR),
(P(Q + R), P, Q + R) and the product of the lines (0, QR, Q + R), (PQ, Q, P), (P +
Q, R, (P + Q)R):  
P+Q PQ O
 R Q QR 
(P + Q)R, P(Q + R) P Q + R
(each column or row corresponds to aline).
3. The morphisms C ×C ×C → C, (P, Q, R) 7→ (P +Q)+R, P +(Q +R) coincides
in a Zariski open subset and so they are equal.
Exercise 32 ([Sil92] p. 60) On the elliptic curve

E : y2 = x3 + 17

over Q. We have points

P1 = (−2, 3), P2 = (−1, 4), P3 = (2, 5), P4 = (4, 9), P5 = (8, 23)P6 = (43, 282)

P7 = (52, 375), P8 = (5234, 378661)


verify:
P5 = 2P1 , P4 = P1 − P3 , 3P1 − P3 = P7 ,
Prove that E(Q) is freely generated by P1 and P3 and there are only 16 integral
points ±Pi , i = 1, 2, . . . , 8 (see [Nag35]).
Exercise 33 [Kob93], p. 35, Problem 4b: For the elliptic curve En : y2 = x3 − n2 x
find an explicit formula for the x coordinates of inflection points.
Exercise 34 [Kob93], p. 36, Problem 7: How many elements of En (R) or of order
2, 3 and 4? Describe geometrically where these points are located.
4.16 Divisors 57

Exercise 35 [Kob93], p. 36, Problem 9: For an elliptic curve over R prove that
E(R) (as a group) is isomorphic to R/Z or R/Z × Z/2Z.
Exercise 36 [Kob93], p. 36, Problem 11:

4.16 Divisors

Let E/k be an elliptic curve over a field k. A divisor in E and defined over k is a
formal finite sum D := ∑i ni pi , ni ∈ Z, pi ∈ E(k̄) such that it is invariant under the
Galois group Gal(k̄/k), that is,

σ (D) = D, ∀σ ∈ Gal(k̄/k)

where
σ (∑ ni pi ) := ∑ ni σ (pi )
i i

The set of divisors over k, let us denote it by Div(E/k) form an abelian group in a
natural way. For any rational function f ∈ k(E) we define

div( f ) := ∑ ni pi
i

where f is of order ni at pi . It is a divisor defined over k. The set of such divisors


form an abelian group which we denote it by Div(k(E)). The Picard-Group of E is
defined to be
Pic(E) := Div(E)/Div(k(E)).
The Chern class map is defined in the following way

c : Pic(E) → Z, c(∑ ni pi ) := ∑ ni
i i

we define
Pic0 (E) := ker(Pic(E) → Z)
We have a canonial map

E(k) → Pic0 (E), P 7→ P − O (4.7)

Proposition 4.16.1 The map 4.7 is an isomorphism of groups.

Proof. First of all we notice that it is a group morphism. Just for this proof we denote
by ⊕ the addition structure in E(k). Let L1 , respectively L2 , be the equation of the
line in P2 passing through P, Q, PQ, respectively O, PQ, P ⊕ Q. We have LL12 ∈ k(E)
with the divisor
58 4 Rudiments of Algebraic Geometry of curves

P + Q + PQ − O − PQ − P ⊕ Q
and so in Pic0 (E) we have P − O + Q − O = P ⊕ Q − O.

4.17 Riemann-Roch theorem

Definition 4.17.1 We say that a divisor D = ∑i ni pi is positive and write D ≥ 0 if


all coefficients ni are non negative integers. In a similar way we define D ≤ 0.
For a divisor D on a curve C/k define the linear system

P(D) = { f ∈ k(C), f 6= 0 | div( f ) + D ≥ 0} ∪ {0}

and
l(D) = dimk (P(D)).

Theorem 4.17.1 (Riemann-Roch theorem) Let C be a smooth curve over k.

l(D) − l(K − D) = deg(D) − g + 1,

where K is the canonical divisor and g is the genus of C.


We only need to know that the canonical divisor satisfies:

deg(K) = 2g − 2

and so for deg(D) > 2g − 2, equivalently deg(K − D) < 0, we have

l(D) = deg(D) − g + 1. (4.8)

4.18 Weierstrass form revised

In this section we prove that any elliptic curve can be realized as a certain curve in
P2 . The following proposition is proved in [Sil92], III, Proposition 3.1.
Proposition 4.18.1 Let E be an elliptic curve over a field k. There exist functions
x, y ∈ k(E) such that the map

E → P2 , a 7→ [x(a); y(a); 1]

give an isomorphism of E/k onto a curve given by

y2 + a1 xy + a3 y = x3 + a2 x2 + a4 x + a6 , a1 , · · · , . . . , a6 ∈ k
4.19 Moduli of elliptic curves 59

sending O to [0; 1; 0]. If further char(k) 6= 2, 3 we can assume that the image curve
is given by
y2 = 4x3 − t2 x − t3 , t2 ,t3 ∈ k,t23 − 27t32 6= 0.
We call x and y the weierstrass coordinates of of E.
Proof. Using Riemann-Roch theorem and in particular (4.8) with g = 1 and D = nO
we get l(D) = n. For n = 2 we can choose x, y ∈ k(E) such that 1, x form a basis
of P(2O) and 1, x, y form a basis of P(3O). The function x (resp. y) has a pole
of order 2 (resp. 3) at O. Now P(6O) has dimension 6 and 1, x, y, x2 , xy, y2 , x3 ∈
P(6O). It follows that there is a relation

ay2 + a1 xy + a3 y = bx3 + a2 x2 + a4 x + a6 , a1 , · · · , . . . , a6 , a, b ∈ k.

Note that ab 6= 0, otherwise every term would have a different pole order at O and so
all the coefficients would vanish. Multiplying x, y with some constants and dividing
the whole equation with another constant, we get the desired equation. The map
induced by x and y is the desired map (check the details).
If char(k) 6= 2, 3 we make the change of variables x0 = x, y0 = y − a21 x and we
eliminate xy term. A change of variables x0 = x − a32 , y0 = y − a23 will eliminate x2
and y terms.
Exercise 37 Write the following elliptic curves in the Weierstrass form:

y2 = x4 − 1, O = [0; 1; 0]

x3 + y3 = 1, O = [0; 1; 1]

4.19 Moduli of elliptic curves

Now we can state what is the moduli of elliptic curves.


Proposition 4.19.1 Assume that char(k) 6= 2, 3. Two elliptic curves Et2 ,t3 and Et 0 ,t 0
2 3
are isomorphic if and only if there exists λ ∈ k, λ 6= 0 such that

t20 = λ 4t2 , t30 = λ 6t3

The isomorphism is given by

(x, y) 7→ (λ 2 x, λ 3 y).

Proof. Let (x, y) and (x0 , y0 ) be two sets of Weierstrass coordinate functions on an
elliptic curve Et2 ,t3 . It follows that {1, x} and {1, x0 } are both bases of P(2O), and
similarly {1, x, y} and {1, x0 , y0 } are both bases for P(3O). Writing x0 , y0 in terms
of x, y and substituting in the equation of Et 0 ,t 0 we get the first affirmation of the
2 3
proposition. The second affirmation is easy to check.
60 4 Rudiments of Algebraic Geometry of curves

Combining Proposition 4.18.1 and Proposition 4.19.1 we conclude that the moduli
space of elliptic curves over a field of characteristic 6= 2, 3 is

M1 (k) := A2 (k) − {(t2 ,t3 ) | 4t23 + 27t32 = 0} / ∼




where

(t2 ,t3 ) ∼ (t20 ,t30 ) if and only if ∃λ ∈ k, λ 6= 0, (t20 ,t30 ) = (λ 4t2 , λ 6t3 ).

If k is algebraically closed then this is the set of k-rational points of the weighted
projective space P2,3 (k) mines a point induced in P2,3 by ∆ = 0. In this case the
j-invariant of elliptic curves

1728 · 4t23
j : M1 (k) → A(k), j[t2 ;t3 ] =
4t23 + 27t32

is an isomorphism and so the moduli of elliptic curves over k is A1 (k). However,


note that if k is not algebraically closed then j has non-trivial fibers. For instance,
all the elliptic curves
y 2 = x 3 − t3 , t3 ∈ Q
are isomorphic over Q̄ but not over Q.
If j0 6= 0, 1728 consider the elliptic curve:
36 1
E j0 : y2 + xy = x3 − x− .
j0 − 1728 j0 − 1728

It satisfies j(E) = j0 .

4.20 The addition formula for ℘

For a fixed z0 ∈ C, z0 6= 0, let us consider ℘(z + z0 ) which is double periodic in z,


and hence, it is a rational function in ℘ and ℘0

Proposition 4.20.1 We have

1 ℘0 (z) −℘0 (y) 2


 
℘(z + y) = −℘(z) −℘(y) (4.9)
4 ℘(z) −℘(y)

Proof. Let f (z) be the difference between the left and the right hand sides of 4.9.
Its only possible poles are in
z = 0, ±y
We examine the Lauerent expansion of f (z) at the point z = 0 and see that it is
holomorphic at z = 0 and there it vanishs. In a similar way, it has no poles at z = y
4.20 The addition formula for ℘ 61

and so, at worst it has a simple pole at z = −y. Since f is double periodic we get the
result.
We let y goes to z and we get

1 ℘00 (z) 2
 
℘(2z) = − 2℘(z).
4 ℘0 (z)

If we make derivation of 4.9 with respect to z we get a similar formula for ℘0 (x + y).
A better organization of this formula is

℘0 (z) |

℘(z),
℘0 (y) | = 0

℘(y),
−℘(x + y) ℘0 (z + y), |

Note that we can state the same formulas in the algebraic context, that is, for an
elliptic curve E over the field k with the Weierstrass coordinates x and y we have
 2
1 y(P) − y(Q)
x(P + Q) = − x(P) − y(Q) (4.10)
4 x(P) − x(Q)

x(P),
y(P) |
x(Q),
y(Q) | = 0 (4.11)
−x(P + Q) y(P + Q), |
!2
1 6x(P) − 12 g2
x(2P) = − 2x(P) (4.12)
4 y(P)

Even though, we have proved 4.10, 4.11, 4.12 in the complex context, they
are valid for an elliptic curve over an arbitrary field. The argument for instance for
4.11 is as follows: We consider g2 , g3 , x(P), y(P), x(Q), y(Q), x(P + Q), y(P + Q) as
variables. The equation 4.11 is in fact a collection of equations:
 2 2
 y(P) 2 = 4x(P) 2− g2 x(P) − g3

 y(Q) = 4x(Q) − g2 · x(Q) − g3


The equation of line x passing through P, Q
The equation of P + Q ∈ x




y(P + Q)2 = 4x(P + Q)3 − g2 · x(P + Q) − g3

 
1
This is a variety defined over Z
6
62 4 Rudiments of Algebraic Geometry of curves

4.21 Why Schemes?

One of the fundamental observation in Grothendieck’s revolution of Algebraic Ge-


ometry, replacing varieties with schemes, is that

Spec(Z) := { prime ideals of Z} ' {2, 3, 5, . . . , p, . . .}

is like a parameter, for instance, the parameter λ in the legendry family of elliptic
curves:
Eλ : y2 = 4χ(2 − 1)(χ − λ ), λ ∈ C
We consider Eλ as a curve over the ring

C[λ ] := polynomials in λ with C − coe f f icients.

The prime ideals of C[λ ] are



Spec(C[λ ]) := {(λ − λ0 )C[λ ] λ0 ∈ C} ' C

For a prime ideal P ⊆ C[λ ], the residue field is naturally isomorphic to C

C[λ ]/(λ −λ0 )C[λ ] −


g→C

p(λ ) 7−→ P(λ0 )


The process of substituting λ with λ0 , can be interpreted as considering Eλ over
the residue field. In a similar way for an elliptic curve over Z, for instance

E : y2 = x3 + 1

The process of working modulo a prime number p is the same as considering E over
the residue field
F p = Z/ pZ
Chapter 5
Mordell-Weil Theorem

We have seen that for an elliptic curve over Q the set E(Q) is an abelian group and
so
E(Q)/E(Q)tors
where
E(Q)tors := {x ∈ E(Q) | nx = 0, for some n ∈ N},
is a freely generated Z-module.
Theorem 5.0.1 For an elliptic curve E over a number field k the group E(k) of
k-rational points is finitely generated abelian group.
The above theorem for k = Q was proved by Mordell. Its generalization for an
arbitrary number field was proved by André Weil and it is known as the Mordell-
Weil theorem. For the proof see [Lan78a, Hus04].
Let us take k = Q. The above theorem implies that the set of torsion points
E(Q)tors of E(Q) is finite and there is a number r ∈ N such that

E(Q) ∼
= Zr ⊕ E(Q)tors .

The non-negative integer r is called the rank of E(Q).

5.1 Mordell-Weil theorem

The free part of E(k), even for k = Q, is mysterious. We do not know whether
there exist an elliptic curve of arbitrary rank or not. Until 1977, only, elliptic curves
of rank 6 9 was known. This is somehow trivial. Then, Mestre rank 6 15, Nago
1992-1994, rank=17, 20, 21, Fermigier, 1992-1997 rank 19, 20, Martin-McMiller,
1998-2000, rank=23, 24, and finally in 2006, N.Elkies found an elliptic curve of
rank 28. This is the largest rank known until now. There are results saying that with
probability 1, an elliptic curve E/Q has rank 0 or 1.

63
64 5 Mordell-Weil Theorem

Proof of Mordell’s theorem consist of two steps



E(Q)/2E(Q) is finite (Weak Mordell)
Existence of a hight function

Theorem 5.1.1 (descent theorem). Let Γ be a commutative group. Suppose that


there is a function
h : Γ → [0, ∞)
such that

1. For any real number M, the set P ∈ Γ |h(P) 6 M is finite.


2. For every P0 ∈ Γ , there is a constant k0 such that

h(P + P0 ) 6 2h(P) + k0 ∀P ∈ Γ

3. There is a constant k such that

h(2P) 6 4h(P) − k ∀P ∈ Γ

Then Γ is finitely generated.

Proof. Let Γ /2Γ = Q1 , Q2 , . . . , Qn . It means that for every P ∈ Γ ∃ Qi1 , depending


on P such that
P − Qi1 = 2P2 ∈ 2Γ
We repeat this for P2 . 

 P − Qi1 = 2P1
 P1 − Qi − 2P2

2
..


 .
Pm−1 − Qim = 2Pm

This implies that

P = Qi1 + 2Qi2 + 22 Qi2 + . . . + 2m−1 Qim + 2m Pm

Now we use the height function . We have

h(P − Qi ) 6 2h(P) + k0 ∀P ∈ Γ i = 1, . . . , n

Where k0 is the maximum of ki ’s attached to each Qi in property 2 of h. We use


property 3 of h and we have
5.1 Mordell-Weil theorem 65

4h(Pj ) 6 h(2Pj ) + k = h Pj−1 − Q j−1 + k
6 2h(Pj−1 ) + k0 + k ⇒
1  k + k0
h(Pj ) 6 h Pj−1 +
2 4
3 1 
= h(Pj−1 ) − h(Pj−1 ) − (k + k0 )
4 4
Therefore, if h(Pj−1 ) > k + k0 then
3
h(Pj ) 6h(Pj−1 )
4
We do this process until for some m h(Pm ) 6 k + k0 and so by property 1 of h the
set of such Pm is finite. We conclude that Γ is generated by

Q1 , Q2 , . . . , Qn , {P ∈ Γ h(P) 6 k + k0 }


Now, let us construct the heught function for E(Q). Let P = (x, y) ∈ E(Q). Write

x := mn , (m, n) = 1
H(P) := H(x) = max{|m|, |n|}
h(P) := log H(P)
 
m n
Let P = (x, y) = ,
e2 e3
, m, n, e ∈ Z. We have

|n|2 6 |m|3 + |t2 | e4 |m| + |t3 | e6


6 k2 H(P)3 ⇒
3
|n| 6 kH(P) 2
Where k is a constant number which only depends on E. Let P = (x, y), P0 = (x0 , y0 ).
We would like to estimate h(P + P0 ).

 2
y − y0
h(P + P0 ) = − x − x0
x − x0
(y − y0 )2 − (x − x0 )2 (x + x0 )
=
(x − x0 )2
Ay + Bx2 +Cx + D
=
Ex2 + Fx + G

Here, we have used y2 = x3− t2 x −


 t3 and A, B, . . . , G are constants depending only
m n
on E and P0 . Now set P = e2 , e3 , (m, e2 ) = 1 and (n, e3 ) = 1
66 5 Mordell-Weil Theorem

H(P + P0 ) 6 max Ane + Bm2 +Cme2 + De4 , Em2 + Fme2 + Ge4




But, we know that


1 3
|e| 6 H(P) 2 , |n| 6 k · H(P) 2 , |m| 6 H(P)

Which implies that


H(P + P0 ) 6 constant. H(P)2
Where the constant term depends only on E and P0 . Taking the logarithm of this, we
have the property 2 of h.
Let us prove property 3 of h. We have

f 0 (x)
x(2P) = − 2x
4 f (x)

Where y2 = x3 − t2 x − t3 = f (x) is the elliptic curve E We assume that 2P 6= 0

P(x)
x(2P) = , deg P(x) = 4, deg Q(x) = 3
Q(x)

Where the coefficients of P, Q only depends on the elliptic curve.


Let 4(4 t23 + 27 t32 ) This is the resultant of P(x)andQ(x). This means that

f1 (x)P(x) + f2 (x)Q(x) = ∆

f1 , f2 ∈ Z[x], deg f1 , deg f2 6 3

We need also
g1 (x)P(x) + g2 (x)Q(x) = ∆ · x7

g1 , g2 ∈ Z[x], deg g1 , deg g2 6 3


This can be considered as the resultant of x4 P 1x , x4 Q 1x
 

Let x = ab , (a, b) = 1 and

F(a, b)  
x(2P) = , δ := F(a, b), G(a, b)
G(a, b)

Therefore,

f1 (a, b) F(a, b) + f2 (a, b) G(a, b) = 4∆ b7


(5.1)
g1 (a, b) F(a, b) + g2 (a, b) G(a, b) = 4∆ a7

This gives
δ |4∆ and |δ | 6 |4∆ |
and so
5.1 Mordell-Weil theorem 67

H(2P) > max { |F(a, b)|, |G(a, b)| } / |4∆ |


On the other hands

|4∆ b7 | 6 2 max{ | f1 |, | f2 | } max { |F|, |G| }


(5.2)
|4∆ a7 | 6 2 max{ |g1 |, |g2 | } max { | f |, |G| }

Where f1 , f2 . . .are evaluated at (a, b).


Now f1 , f2 , g1 , g2 are polynomials of degree 6 3.

max { | f1 |, | f2 |, |g1 |, |g2 | } 6 C max { |a|3 , |b|3 } (5.3)


Where C is a constant which only depends on E. Combining 5.2 and 5.3 we get

|4∆ | max { |a7 |, |b7 | } 6 2C · max{ |a3 |, |b3 | } max {F(a, b), G(a, b)}

and so
max{|F(a,b)|,|G(a,b)|} max{|F(a,b)|,|G(a,b)|}
H(2P) = δ > 4∆

> (2C)−1 max{|a|, |b|} = (2C)−1 H(P)


For more on descent procedure see [Silvermann I] page 199 chapter VIII.
Proposition 5.1.1 Let E be an elliptic curve over Q. We have

h(P1 + P2 ) + h(P1 − P2 ) 6 2h(P1 ) + 2h(P2 ) + k (5.4)

Where k only depends on E.


For a proof see [Silvermann I] page 216
Proposition 5.1.2 (Tate). Let E/Q be an elliptic curve

1
ĥ(P) := lim 4−N h(2N P)
2 N→∞
exists.
Proof. Taken from [Silvermann I] page 288. We show that the sequence in lim is
Cauchy. In 5.4 we pu P = P1 = P2

h(2P) − 4h(P) 6 k

Where k is a constant depending only on k. For N > M > 0 integers, we have


N−1
−N N −M M −n−1 n+1 −n n

4 h(2 P) − 4 h(2 P) = ∑ 4 h(2 P) − 4 h(2 P)
n=M
N−1 N−1
6 ∑ 4−n−1 h(2n+1 P) − 4h(2n P) 6 ∑ 4−n−1 C

n=M n=M

C
6 4M+1
68 5 Mordell-Weil Theorem

Definition 5.1.1 ĥ is called the canonical or Neron-Tate height of E.


Theorem 5.1.2 (Neron-Tate) The canonical height ĥ satisfies
1. For all P, Q ∈ E(Q)

ĥ(P + Q) + ĥ(P − Q) = 2h̆(P) + 2ĥ(Q)

2. For all P ∈ E(Q) and m ∈ Z

ĥ(M, P) = m2 ĥ(P)

3. ĥ is a quadratic form in E(Q), that is, ĥ is even and the paring

h· , ·i : E(Q) × E(Q) → R
hP, Qi = ĥ(P, Q) − ĥ(P) − ĥ(Q)

is bilinear.
4. For P ∈ E(Q) we have ĥ(P) > 0 and

ĥ(P) = 0 ⇔ P is a torsion point.

5. 2ĥ − h is bounded on E(Q).

Proof. Taken from [Silvermann I] page 229


5. We have
−N
C
4 h(2N P) − 4−M h(2M P) 6

4M+1
Take M = 0 and let N → ∞. We get
C
|2ĥ(P) − h(P)| 6
4
1. We have
h(P + Q) + h(P − Q) = 2h(P) + 2h(Q) + O(1)
replace P, Q by 2N P, 2N Q and multiply it with 12 4−N and then N → ∞.
2. We know that h(mP) − m2 h(P) is bounded. Again replace P by 2−N P, multiply
with 4−N and N → ∞
The rest is left to the reader.
Let P1 , P2 , . . . Pγ be a basis of the free part of E(Q) The regulator of E/Q is
defined
RE/Q := det hPi , Pj i
5.2 Weak Mordell-Weil theorem 69

5.2 Weak Mordell-Weil theorem

Theorem 5.2.1 Let E be an elliptic curve over a number field k. Then E(k)/mE(k)
is finite for all m ∈ N.
We give a proof of weak Mordell-Weil theorem for k = Q, the family

Ea, b : y2 = x3 + ax2 + bx

and for E(Q)/2E(Q). We follow Husemller, chapter 6 3, in which he use an explicit


2-isogeny in chapter 4 5.
Chapter 6
Torsions and isogeny

6.1 Torsion points

Using Weierstrass theorem we have seen a correspondence

The space of lattices −


g→ the space of pairs (E, ω) (6.1)

Where E is an elliptic curve over C and ω is a regular differential 1-form. Since we


have not defined these objects intrinsicly, then E must be taken in the Weierstrass
format an ω = dx y . In the right hand side of 6.1 we can talk about pairs defined over
an arbitrary field.
Definition 6.1.1 A pair E, ω is called an enhanced elliptic curve. There will be
other enhancements, and in order to reduce confusion, we say that E is enhanced
with a regular differential 1-form.

Definition 6.1.2 For an elliptic curve E 0 over a field of characteristic zero k, the set
of m-torsion in
E[m] = {P ∈ E(k) | mP = O}
This is a subgroup of E(k).
Proposition 6.1.1 We have an embedding of groups

E[m] ,→ Z/mZ × Z/mZ

and for k an algebraically closed field, this is an isomorphism.

Proof. We first prove the second part for k = C. We can assume that E = C/Λ in
this case
1
E[m] ' Λ /Λ ' Z/mZ × Z/mZ
m
Now, we prove the second part. Since k is a field of characteristic zero and E uses a
finite numbers of elements of E we can assume that there is an embedding of fields

71
72 6 Torsions and isogeny

σ : k ,→ C

Let Eσ be the elliptic curve over C obtained from E and regarding its coefficients as
complex numbers. We have an embedding of groups

E[m] ,→ Eσ [m] ' Z/mZ × Z/mZ

For k an algebraically closed field, the algebraic equation mP = O has m2 distinct


solution over C, and it is defined over k. This means that all these solutions are
defined over k and the result follows.


In particular, for an elliptic curve E over Q we have

E[m] ' Z/mZ × Z/mZ

Let
E[2] = {(t1 , 0), (t2 , 0), (t3 , 0), O}
For families of elliptic curves with a 3-torsion point see [Husemller] Chapter 4, 2.

6.2 Isogeny

Let Λ̌ ⊆ Λ ⊆ C be two lattices and let

N := #Λ /Λ̌

This gives a map of tori


f : Ê → E (6.2)
Which is induced by the identity map C → C. Here, E = C/Λ and Ě = C/Λ̌ . This
is actually a holomorphic map between two Riemann surfaces. It is as a morphism
of groups. We have
f ∗ ω = ω̌,
where ω and ω̌ are the differential form dz induced in Ě, respectively. Furthermore

Λ
f −1 [z] = [z] +

Λ̂
which means that f is a N to 1 map with no ramification points.
Definition 6.2.1 f as in 6.2 is called an isogeny of degree N.
Definition 6.2.2 Let N := #Λ /Λ̌ . we have

NΛ ⊆ Λ̌ ⊆ Λ
6.2 Isogeny 73

and this gives us the maps


g f
E −→ Ĕ −→ E
[z] 7−→ [Nz]
[z] −→ [z]

In the level of differential forms

Nω ←[ ω̌ ←[ ω

g is called the dual isogeny of f . Note that both F0 g, g0 f are multiplication by N


map.

Exercise 38 Let G be a finite abelian group generated by at most two elements.


There are unique d1 , d2 ∈ N such that

G ' Z/d1 Z ⊕ Z/d1 d2 Z

Conclude that any isogeny E1 7→ E2 can be uniquely written as

α β
E1 −→ E1 −→ E2
[z1 ] → [d1 z1 ]

Where α is a multiplication by d1 and ker(β ) is cyclic of order d2


Let a ∈ C∗ and Λ ⊆ C be a lattice. Let also

Ě = C/aΛ , E = C/Λ

We have the map


fa : E → Ě [z] 7→ [az]
which is a bijection and its inverse is given by fa−1 . Moreover

fa∗ ω = aω̌

Under the mentioned isomorphism, the lattices aΛ and Λ corresponds to (E, aω), (E, ω)
Exercise 39 Show that the number of sublattices Λ̌ ⊆ Λ of a fixed lattice Λ with #Λ /Λ̌ =
n is
σ (n) = ∑ d
d|n

Hinit: Take a basis of Λ 0 and Λ and show that

⊆ Λ ,#Λ
Λ̆  −→SL(2,
 /Λ̆= ng   Z)\Mat n (2, Z)
a b ω 1 ab
Λ 0 generated by ←−[
cd ω2 cd

This quotient has representatives


74 6 Torsions and isogeny
 
ab
0 6 b 6 d − 1, a · d = n
cd

Note that for n = P prime the set of such lattices is

Λ̌ = ZPω1 + Zω2 , Z(ω1 + ω2 ) + ZPω2


0 6 b 6 P−1

6.3 Isogeny II

Let E be an elliptic curve in the Weierstrass format and defined over Q, that is, t2 ,t3 ∈
Q. We have
G := Λ /Λ̌ ⊆ E[n]
G ⊆ E(Q)
We way think of Ĕ as the quotient E/G . The elliptic curve Ĕ is also defined over
(Q) and the reason is as follows.
The pull-back of ℘(z,Λ ), ℘(z,Λ ) by the map of isogeny is an elliptic function
with respect to the lattice Λ̆ . Therefore, we have
 
℘(z,Λ ) = P ℘(z, Λ̌ ),℘0 (z, Λ̌ )
 
℘0 (z,Λ ) = Q ℘(z, Λ̌ ),℘0 (z, Λ̆ )

where P and Q are rational functions in two variables and with coefficients in C
Theorem 6.3.1 If E is defined over Q then the isogeny
 
Ě → E (x, y) → P(x, y), Q(x, y)

is defined over Q, that is Ĕ and P, Q are defined over Q.


This will be proved in §7.3 when we introduce Hecke operator.
Chapter 7
Hecke operators

In this chapter we introduce one of the fundamental features of modular forms which
is responsible for many arithmetic properties. This is namely the Heck operators
acting on the space of modular forms. There are many text books covering this topic
perfectly, see for instance Apostol’s book [Apo90] Chapter 6. We will adopt a more
geometric approach suitable for the same topic in the context of algebraic geometry
of elliptic curves. The first application of Hecke operators is the following.
Theorem 7.0.2 The numbers τ(n) are multiplicative, that is for all n, m ∈ N with
(n, m) = 1 we have
τ(n · m) = τ(n) · τ(m)

7.1 Hecke operators

So far, we have interpreted modular forms as functions in theree spaces:


1. The Poicaré upper half plan H,
2. the space P of lattices,
3. and the affine space A2C with the coordinate system (t2 ,t3 ).
In this section for each natural number n we want to define the Hecke operator

Tn : Mk → Mk

which is a linear map. It is given by one of the following equivalent definitions:


1. For f : H → C a modular form of weight k we have
s
Tn ( f ) = ∑ f |k Ai ,
i=1

where {[A1 ], [A2 ], · · · [As ]} = SL(2, Z)/Mat n (2, Z).

75
76 7 Hecke operators

2. For f : P → C a modular form of weight k we have

Tn ( f )(Λ ) = nk−1 ∑ f (Λ 0 ),
Λ0

where Λ 0 runs through all sublattices Λ 0 ⊂ Λ of index n. This means that


#(Λ /Λ 0 ) = n.
3. For f a homogeneous polynomial of degree k in C[tt ,t3 ], deg(t2 ) = 4, deg(t3 ) = 6
we have
Tn ( f )(t2 ,t3 ) = nk−1 ∑ f (t 0 ),
t0

where t 0 = (t20 ,t30 ) runs through all parameters for which there is an isogeny α :
Et 0 → Et such that α ∗ ( dx dx
y ) = y and deg(α) = n.

Exercise 40 Prove the equivalence of the above definitions.


Exercise 41 Prove that each equivalence class in SL(2, Z)/Mat n (2, Z) is repre-
sented exactly by one of the matrices
 
d b n
, d|n, 0 ≤ b < .
0 dn d

Using the above exercise we know that the action of the Hecke operator Tn on a
modular form of weight k defined on the upper half plane is given by

aτ + b
Tn ( f )(τ) = nk−1 ∑ d −k f ( ) (7.1)
a·d=n, 0≤b≤d−1 d

Proposition 7.1.1 For two natural numbers n.m and Hecke operators Tn , Tm ∈
Mk → Mk prove that
Tn ◦ Tm = ∑ d k−1 T nm2 . (7.2)
d
d|(n,m)

In particular, for n and m coprime we have

Tn ◦ Tm = Tnm

and for p a prime number

Tp ◦ Tpe = Tpe+1 + pk−1 Tpk−1 . (7.3)

Proof. We follow [Silvermann] page 69.


The idea of the proof is more and less in [Silvermann I] page 62. Let n, m ∈
N and d|(n, m) be fixed. We prove that for pairs of isogenies.

α β
E1 −−→ E2 −−→ E, deg α = n, deg β = m
℘ nm
There is a unique isogeny E1 −−→ E, deg℘ = d2
such that
7.1 Hecke operators 77

℘◦ [d] = β ◦ α

and for h fixed we have d pairs of such isogenies α, β


This decomposition is inspired by the identity
 nm 
σ (n) · σ (m) = ∑ d · σ
d|(n,m)
d2

If this is the case, then


 
Tn ◦ Tm f (E, ω) = (n, m)k−1 ∑ f E1 , (β ◦ α)∗ ω
E1 α E2 β E

− → −
 
= ∑ (nm)k−1 d ∑ f E1 , (℘◦ [d])∗ ω
d|(n,m) E1 ℘ E


−k
= ∑ (nm)k−1 d f (E1 , ℘∗ ω)
d|(n,m)
nm
= ∑ d k−1 T
d|(n,m)
d2

we have used [d]∗ ω = dω and f (E, d∗) = d −k f (E, ∗)


in the order to prove the affirmation on isogenies we prove the corresponding
affirmation on lattices. Let Λ ⊆ C be a fixed lattices and
Λ2 Λ
Λ1 ⊆ Λ2 ⊆ Λ # = n, # =m
Λ1 Λ2

For (n, m) = 1, Λ2 is uniquely characterized by

Λ2 = {x ∈ Λ | nx ∈ Λ1 }

and the affirmation is trivial. We fin a subgroup

Gd := Z/dZ × Z/dZ ⊆ Λ /Λ1

and define
Λ3 = pull-back of Gd by Λ → Λ /Λ1
We have
dΛ3 ⊆ Λ1 ⊆ Λ3
and the index in both inclusions dΛ3 ⊆ Λ3 and Λ1 ⊆ Λ3 is d 2 . Therefore Λ3 = Λ1 .
We have to show that there are d such Gd and hence Λ3 .
t
u

The formula (??) is summarized in the following formal equality:


78 7 Hecke operators

∑ Tn n−s = ∏(1 − Tp p−s + pk−1−2s )−1
n=1 p

Let f be a modular form with the Fourier expansion:



f (z) = ∑ an qn , q = e2iπz .
n=0


For m ∈ N, we have Tm f (z) = ∑ bn qn , where
n=0

bn = ∑ d k−1 amn/d 2 .
d| gcd(m,n)

Exercise 42 Can you show by algebraic geometric methods that for fixed t ∈
C2 \{∆ = 0} the set of parameters t 0 with

dx dx
α : Et 0 → Et , α ∗ = , deg(α) = n
y y
is finite.
 
λ 0
Let A be an element in the group generated by GL+ (2, R) and { | λ ∈ C∗ } ∼
=
  0 λ
ω1
C∗ . Let also ω = ∈ P. Then
ω2

Aω ∈ P.

Using the map p in Weierstrass uniformization theorem, we can translate the above
process to the (t2 ,t3 )-space. Namely, for each t ∈ C2 − {∆ = 0} and a basis of
the homology δ1 , δ2 ∈ H1 (Et , Z) with hδ1 , δ2 i = 1 and A as above we have a local
holomorphic map t 7→ α(t). If we choose another basis of H1 (Et , Z) obtained by the
previous one by an element B ∈ SL(2, Z), then we have a new period matrix ABω.
This is equal to Aω in P if and only if

ABA−1 ∈ SL(2, Z).

We conclude that if we choose representatives for the quotient

(A · SL(2, Z) · A−1 ∩ SL(2, Z))\(A · SL(2, Z) · A−1 ) = {[Ai ] | i = 1, 2, . . . , s}

then to each Ai we have associated a local map αi (t).


Exercise 43 Let Ai = ASi A−1 , Si ∈ SL(2, Z). Show that

SL(2, Z)/Mat n (2, Z) = {[ASi ] | i = 1, 2, . . . , s}

and vice-versa.
7.2 Hecke and cusp forms 79

Exercise 44 Let n ∈ N. Are there polynomials pn,i (t2 ,t3 ), i = 2, 3 such that pn :=
(pn,2 , pn,3 ) leaves ∆ = 0 invariant and

Tn ( f )(t) = f (p−1
n (t)), f ∈ Mk ,

where f (X) = ∑x∈X f (x).

7.2 Hecke and cusp forms

Theorem 7.2.1 Let f be a cusp form of weight k and suppose that f is an eigen
form for all Hecke operators and fn = 1. Then

Tn f = fn · f

Proof. We have Tn f = λn · f and so


λn fm = (Tn f )m = ∑ d k−1 fdnm2
d|(n,m)
We put m = 1 and get λn = fn
1. f is a cusp form with f1 = 1
Definition 7.2.1 A normalized eigen function is a modular form f with
2. Tn f = fn · f
Exercise 45 (1.25 Silvermann II page 92) 1. Ek is an eigenform for all Hecke
operators. 2. if f is an eigen form for all Hecke operators and f is not a cusp
form then f is a multiple of Ek
Theorem 7.2.2 Let f (τ) = ∑∞ n
n=1 f n q be a normalized eigen function of weigh k.
Then
fn m = fn fm (n, m) = 1
f pn · f p = f pn+1 + pk−1 f pn−1 p prime
Proof. Tn f = fn · f and the theorem follows from [taken from Silvermann page 79]
Our main example is

∆ = qπ(1 − qn )24 = ∑ τ(n)qn


n=1

We have
τ(n)τ(m) = τ(nm) (n, m) = 1
(7.4)
τ(P)τ(P ) = τ(P ) + P11 τ(Pn−1 )
n n+1

 
This follows from the fact that S12 SL(2, Z) is generated by ∆
The identities were conjectured by Ramanujan and proved by Mordell.
 
Exercise 46 Can you find a basis of Sn SL(2, Z) , n = 14, 16, 18 wich are nor-
malized eigen forms
80 7 Hecke operators

7.3 Proof of Theorem 6.3.1

First we prove that Ě is defined over Q.

In order to
 see the analogy between
 holomorphic and algegraic context we define
0
(x, y) = ℘(z,Λ ), ℘ (z,Λ )
 
ˇ 0 (z, Λ̌ )
(x̌, y̌) = ℘(z, Λ̌ ), ℘
f −1 (O) and f −1 (P) both contains N points and let
 
x̌ f −1 (O) = {∞, a2 , a3 , ..., as } ⊆ Q
  (7.5)
x̌ f −1 (P1 ) = {b1 , b2 , ..., br } ⊆ Q

Since Ě is defined over Q its torsion points are also defined over Q and the inclusions
Q above follows. There is no repetition among ai 0 s (resp.bi 0 s). Note that the map

Ě → Ĕ, X → −X
leaves both sets f −1 (O) and f −1 (P1 ) invariant. For f −1 (P1 ) this follows from P1 =
−P1 . Moreover
x̌(x) = x̌(−x), X ∈ Ě
It follows that s = r. Analysing the set of poles and zeros of (x − e1 ) of we get
r
∏ (x̌ − bi )
i=1
(x − e1 )o f = a r
∏ (x̌ − ai )
i=2

For some a ∈ C. Since f sends torsions of Ě to E evaluating (x − e1 )o f at any tor-


sion points of Ě different from 7.5 we get a ∈ Q.

In a similar way, we have

yo f ∏(x̌ − ci )
=
y̌ ∏(x̌ − di )

Where ci , di 0 s are algebraic numbers obtained by evaluating x̌ at f −1 E[2] .




Let f ∈ Mk be modular form defined over Q.
 
P(x) = ∏ x − f (Λ̌ )
Λ̌ ⊆ Λ , #Λ /Λ̌ =n
7.3 Proof of Theorem 6.3.1 81

Fig. 7.1 Isogeny

N
The coefficients of P(x) are symmetric polynomials in N = σ (n) quantities ∑ xi , ∑ xi ji ∑ xi x j xk , ...
i=1 i< j i< j<k
We can write these as polynomials with coefficients in Q of the quantities
N
∑ xim , m = 1, 2, 3, ...
i=1

These are Tn f m . it is enough to prove that Hecke operators send modular forms
defined over Q to modular forms defined over Q. This follows from the computation
of q-expansion of (Tn f )0 s.
We know that f Ě(Q)tors to E(Q)tors , therefore for an elliptic function P on E
with poles and zeros on torsion points of E, the pull-back of P by f has poles and
zeros in torsion points of Ĕ.   
2
We write ℘0 (z,Λ )2 = 4 ℘(z,Λ ) − e1 ℘(z,Λ ) − e2 ℘(z,Λ ) − e3
℘(z,Λ ) − e1 has a pole of order two at O and a zero of multiplicity 2 at a 2-torsion
poin. The pull-back of this function gives us the desired result.
Chapter 8
Riemann zeta function

In this chapter we introduce the Riemann zeta function. We will follow mainly the
Riemann’s original article ([Rie]) and the book [Edw01] which explain an historical
account on Riemann’s paper. Euler considered the zeta function

1
ζ (s) := ∑ ns
n=1

for real s and Riemann introduced ζ (s) for complex s and its extension as a mero-
morphic function to the whole s-plane. In particular, it was known before Riemann
that
π2 π4 π6 π8
ζ (2) = , ζ (4) = , ζ (6) = , ζ (8) =
6 90 945 9450

8.1 Riemann zeta function

In this section we are going to study the first of all Zeta function, namely Riemann
zeta function:

1
ζ (s) := ∑ s
n=1 n

We mainly use the Riemman’s original article [Rie].


Proposition 8.1.1 The series ζ (s) converges for all s ∈ C with ℜ(s) > 1 and

1
ζ (s) = ∏ .
p (1 − p−s )

where p runs over all primes.

Proof. We have |n−s | = nℜs and so it is enough to prove the proposition for s ∈
R, s > 1. We have

83
84 8 Riemann zeta function

1 x−s+1 +∞ 1
Z ∞
∑ ns ≤ x−s = | = if s > 1
n=2 1 −s + 1 1 s−1

Again we assume that s is a real number bigger than 1. We have p−s < 1 and so

(1 − p−s ) = ∑ p−ms .
m=0

By unique factorization theorem

∏ (1 − p−s )−1 = ∑ n−s + RN (s).


p≤N n≤N

Clearly

RN (s) ≤ ∑ n−s .
n=N+1

Since ζ (s) converges we have RN (s) → 0 as N → ∞ and the result follows.

8.2 The big Oh notation

In this section for a ∈ R ∪ {±∞} we define the interval Ia to be a small one sided
neighborhood of a. If a ∈ R this means that Ia = (a, a + ε) or = (a − ε, a) for some
small ε and for a = +∞ this means Ia = (b, +∞) for a big positive number b and for
a = −∞ this means Ia = (−∞, −b) for a big positive number b.
Definition 8.2.1 Let f , g be two complex valued function in Ia we write

f = O(g) or f ∼x→a g
f (x)
to say that g(x) is bounded near a, that is there exists a constant M such that

| f (x) |≤ M | g(x) |, ∀x ∈ Ia .

For three complex valued functions f , g and h in Ia we write f (x) = h(x) + O(g(x))
if f (x) − h(x) = O(g(x)).
We mainly use the following convergence criterion: Let f be a complex valued
continuous function in Ia , a ∈ R and

f ∼x→a (x − a)s , s ∈ R.

For a = ±∞ we assume f ∼x→a xs . We can extend this assumption to s = ±∞. For


instance for a ∈ R the expression f ∼x→a (x − a)+∞ means

∀s ∈ R+ , f ∼x→a (x − a)s
8.3 Gamma function 85

For a, s ∈ R the integral Z


f (x)dx
Ia
R
converges if s > −1. If a = ±∞, s ∈ R then the integral Ia f (x)dx converges if
s < −1.

8.3 Gamma function

The gamma function is defined by


Z ∞
Γ (s) = xs−1 e−x dx, ℜ(s) > 0
0

It converges because near 0

xs−1 e−x ∼x→0 xℜ(s)−1

and so for ℜ(s) > 0 the integral near zero converges. Near infinity it is alway con-
vergent because
xs−1 e−x ∼x→+∞ x−n−1 , ∀n ∈ N.
We have
Γ (s) = (s − 1)Γ (s − 1) (8.1)
because Z ∞
Γ (s) = − xs−1 de−x = · · · = (s − 1)Γ (s − 1)
0
Since Γ (1) = 1 this implies that

Γ (n) = (n − 1)!, n ∈ N

and so the Γ -function is the interpolation of the factoriel function.


The equalities

Γ (s + 1) Γ (s + n + 1)
Γ (s) = = ··· = , n∈N
s s(s + 1) · · · (s + n)

enables us to continue Γ analytically onto all of the complex s-palne with poles of
simple order at s = 0, −1, −2, . . .. We have also Euler’s reflection formula
π
Γ (1 − s)Γ (s) = (8.2)
sin (πs)

which shows that Γ has not zeros. We have also the equality
86 8 Riemann zeta function


 
1
Γ (s) Γ s+ = 21−2s π Γ (2s).
2

Remark 8.3.1 Gauss introduced the notation Π (s) = Γ (s + 1) which is used in


Riemann’s original article. The notation Γ is due to Legendre, see [Edw01] p. 8.

8.4 Mellin transform

The content of this section is taken from Wikipedia. The Mellin transform of a
function f defined on R+ is
dx
Z ∞
xs f (x) .
0 x
If f (x) is locally integrable along the positive real line, and

f (x)x→0+ = O(xu ) and f (x)x→+∞ = O(xv )

then its Mellin transform converges in the fundamental strip [−u, −v].
By definition the Γ function is the Mellin transform of e−x .

8.5 Analytic extension

From the definition of Γ -function it follows:


Γ (s)
Z ∞
= xs−1 e−nx dx, ℜ(s) > 1 (8.3)
ns 0

Taking sum for n = 1, 2, . . . we obtain


Z ∞ s−1
x dx
Γ (s)ζ (s) = , ℜ(s) > 1.
0 ex − 1

One can see that near +∞ we have ex1−1 = O(x−∞ ) and near 0 we have ex1−1 =
O(x−1 ). Therefore, the convergence strip for the above integral is ℜ(s) ∈ (1, +∞)
(see the section on Mellin transform).
Now, we consider the integral

(−x)s dx
Z +∞
I(s) :=
+∞ ex − 1 x

Here, we have taken the branch of (−x)s = es ln(−x) in C\R+ such that ln(−x) for
negative x is a real number. The path of integration begins at +∞, moves to the left
up to the positive axis, circles the origin once in the counterclockwise direction, and
returns down to the positive real axis to +∞. Now the above integral is convergent
8.6 Functional equation 87

for all s and it gives an entire function in s. A simple calculation show that it is equal
to
xs dx
Z ∞
I(s) = (eπis − e−πis ) x
, ℜ(s) > 1
0 e −1 x
In particular, this shows that I(s) vanishes in s = 2, 3, . . .. Therefore, we get

(−x)s dx
Z +∞
2i sin(πs)Γ (s)ζ (s) =
+∞ ex − 1 x
and by (8.2)
(−x)s dx
Z +∞
Γ (1 − s)
ζ (s) =
2πi +∞ ex − 1 x
This equality shows that
1. ζ (s) extends to a meromorphic function on the s palne
2. it has a unique pole in s = 1. The point s = 1 is a simple pole of ζ .
3. It vanishs in s = −2, −4, −6, . . ..
The third item follows from the following: The Bernoulli numbers Bk are given by
−1 1 −1 1 1 5
x ∞
xk 2 6 2 30 4 42 6 32 8 66 10
= ∑ Bk = 1 + x + x + x + x + x + x +···
ex − 1 k=1 k! 1! 2! 4! 6! 8! 10!

For s ∈ Z, s ≤ 0 we have

Γ (1 − s) ∞
Bm xm+s−2 B1−s
Z
ζ (s) = − ( ∑ (−1)s−1 )dx = −Γ (1 − s)(−1)s−1
2πi |x|=ε m=0 m! (1 − s)!

8.6 Functional equation

By Cauchy theorem for s with ℜ(s) < 0 we get



2 sin(πs)Γ (s)ζ (s) = (2π)s ((−i)s−1 + is−1 ) ∑ ns−1
n=1

In other words the function


s s
Γ ( )π − 2 ζ (s)
2
remains invariant under s 7→ 1 − s. By analytic continuation this holds in the whole
s-plane.
Exercise 47 Find the values of ζ for even positive integers:

(2π)2n (−1)n+1 B2n


ζ (2n) =
2 · (2n)!
88 8 Riemann zeta function

8.7 Second proof for functional equation

s
In the equality (8.3) we make the change of variables n → n2 π and s → 2 and we
obtain:
Γ ( 2s ) − 1
Z ∞
s 2
s
π 2 = x 2 −1 e−n πx dx, ℜ(s) > 1
n 0
Define

2
ψ(x) := ∑ e−n πx
n=1

and so
s
Z ∞
1 s
ζ (s)Γ ( )π − 2 = x 2 −1 ψ(x)dx, ℜ(s) > 1 (8.4)
2 0

Exercise 48 Prove that ψ converges and


1 1
1 + 2ψ(x) = x− 2 (1 + 2ψ( ). (8.5)
x
Moreover, the above integral converges for ℜ(s) > 1. We have the theta series
∞ 1 2
θ3 (z) := ∑ q 2 n , q = e2πiz , z ∈ H
n=−∞

which is related to ψ by
θ3 (z) = 1 + 2ψ(−iz).
Using (8.5) one can see that

s dx 1
Z ∞
1 s 1−s
ζ (s)Γ ( )π − 2 = (x 2 + x 2 )ψ(x) − , s∈C (8.6)
2 1 x s(1 − s)

for which the right hand side is convergent for all s ∈ C. We multiply the above
equality by s(s−1)
2 and define
s s
ξ (s) := Γ ( + 1)(s − 1)π − 2 ζ (s)
2
which is an entire function and satisfies ξ (s) = ξ (1 − s).
Exercise 49 From 8.4 we have
Z i∞
s s 1 s
i 2 ζ (s)Γ ( )π − 2 = z 2 −1 ψ(−iz)dz, ℜ(s) > 1
2 0

In the above formula, use the Schwarz map of the family of elliptic curves y2 − x3 +
3x − t = 0 and write ζ as an integral in the t-domain. For this you have to calculate
8.8 Zeta and primes 89
R dx
δ y
ψ( R 1 dx ) as a polynomial in elliptic integrals. In particulat, one may prove in this
δ2 y
way that ζ (s) for s integer is a period.

8.8 Zeta and primes

In this section we sketch the relation between primes and the zeros of ζ .
Theorem 8.8.1 The sum
s s
∑(ln(1 − ρ ) + ln(1 − 1 − ρ ))
ρ

where ρ runs over all roots of ξ with ℑ(ρ) > 0, converges absolutely. In particular,
s
ξ (s) = ξ (0) ∏(1 − ).
ρ ρ

where the infinite product is taken in an order which pairs each root ρ with 1 − ρ.
We have
1
Z ∞
ln ζ (s) = ∑ ∑ p−ns = s J(x)x−s−1 dx, ℜ(s) > 1 (8.7)
p n n 0

where
1 1 1
J(x) := ( ∑ + ∑ ).
2 pn <x n pn ≤x n
The order of summation in 8.7 is unimportant because it is absolutely convergent.
Now we use the Fourier inversion
Z a−i∞
1 ds
J(x) = ln ζ (s)xs , a>1 (8.8)
2πi a+i∞ s
we have also
s s s
ln(ζ (s)) = ln(ξ (0)) + ∑ ln(1 − ) − lnΓ ( + 1) + ln π − ln(s − 1). (8.9)
ρ ρ 2 2

The direct substitution of 8.9 in (8.8) leads to divergent integrals. For this reason we
write Z a−i∞
1 d ln ζ (s) s
J(x) = ( )x ds, a > 1 (8.10)
2πi ln(x) a+i∞ ds s
Now the substitution of (8.9) gives us convergent integrals and a formula for J(x):

dt
Z ∞
J(x) = Li(x) − ∑ (Li(xρ ) + ln(x1−ρ )) + + ln(ξ (0)), x > 1,
ℑ(ρ)>0 x t(t 2 − 1) ln(t)
90 8 Riemann zeta function

where Z x
dt
Li(x) := .
0 ln(t)
Now we derive a formula for
π(x) = ∑1
p<x

using

µ(n) 1
π(x) = ∑ J(x n ),
n=1 n
where µ(n) is zero if n is divisable by a prime square, 1 if n is a product of an even
number of distinct primes, and −1 otherwise.

8.9 Other zeta functions

There are many generalizations of the Riemann Zeta function. One of them is al-
ready used in §10.2. Bellow we explain how the zeta functions of a curve over a
finite field is a generalization of the Riemann zeta function.
Consider a plane affine curve C : f (x, y) = 0, f ∈ Fp (x, y) defined over the field
F p . In analogy with the Riemann zeta function we define

1
ζ (C, s) = ∏ . (8.11)
p (1 − (Np)−s )

where p runs over all non-zero prime ideals of F p [C] := F p [x, y]/h f (x, y)i. Here Np
is the order of the quotient F p [C]/p. Since such a quotient is a finite integral domain
it is a field and hence it has pn elements. We define deg(p) := n. This allows us to
redefine the zeta function as follows:
1
Z(C, T ) = ∏ .
p (1 − T deg(p) )

with ζ (C, s) = Z(C, p−s ).


Proposition 8.9.1 Let C be a curve ove the finite field F p . We have

#C(F pr ) r
Z(C, T ) = exp( ∑ T )
r=1 r

For a proof see [Mil96] p. 90.


Exercise 50 Calculate the zeta functions of A1 and P1 .
Exercise 51 Calculate the zeta function of degenerated elliptic curves:

y2 = x3 + ax + b, 4a3 + 27b2 = 0.
8.11 L-function of cusp forms 91

8.10 Dedekind Zeta function

In this section we give a summary of Dedekind Zeta functions. Let k be anumber


field and Ok be its ring of integers.
Theorem 8.10.1 The integer ring of a number field is a Dedekind domain, i.e every
ideal a ⊂ Ok in a unique way can be written

a = pα1 1 pα2 2 · · · pαs s ,

where α1 , . . . , αs ∈ N0 and p1 , . . . , ps are prime ideals.


A character χ on Ok is a map from the set of non-zero ideals of Ok to C such that it
is multiplicative:
χ(a1 a2 ) = χ(a1 )χ(a2 ).
We mainly use the Character χ ≡ 1. Formally, the Dedekind Zeta function is defined
in the following way:

χ(a) 1
ζk (s) = ∑ s
=∏ −s
, N(a) = #(Ok /a),
a N(a) p 1 − χ(p)N(p)

where the sum is running in non-zero ideals of Ok and the product is running in the
prime ideals of Ok .
Exercise 52 Discuss the convergence of the Dedekind Zeta function (put χ ≡ 1).
Exercise 53 Discuss the fact that the integer ring of a number field is not necessarily
a unique factorization domain/principal
√ ideal domain. Give examples of irreducible
but not prime elements. Hint Q( −5)
√ √
2.3 = (1 + −5)(1 − −5)
√ √ √ √
h6i = h2, 1 + −5ih2, 1 − −5ih3, 1 + −5ih3, 1 − −5i
Exercise 54 The ring of Gaussian integers is Z[i]. Prove that the prime ideals of
Z[i] are of two types:

p = hpi, ifp ≡ 3(4), = ha + ibi, if a2 = b2 = p ≡ 1(4).

In the second case we say that p splits in Z[i]. Show that the only units of Z[i] are
±1, ±i. Show also that the only ideal which ramifies is p = h1 + ii, i.e. p2 = hpi.

8.11 L-function of cusp forms

We follows [Silvermann II] page 80-84


We first define L-functions for the full modular group .
92 8 Riemann zeta function

Let f = ∑ fn qn f1 = 1, be a normalized eigenfunction of weight k. Then
n=1

fmn = fm fn , (n, m) = 1
f pe · f p = f pe+1 + pk−1 f pe−1 e>1

Proposition 8.11.1 We have



fn 1
L( f , s) := ∑ = (8.12)
n=1 ns P ∏
prime 1 − f p p−s + pk−1−2s

for Re(s) > 2k + 1

Proof. The convergence follows from fn ∼ nk and the same convergence statement
for Riemanns zeta function

L( f , s) = ∑ fn · n−s = ∏ ∑ f pe · p−es
P prime e>◦

we have
 
(1 − f p p−s + pk−1−2s ) ∑ f pe · p−es = ∑ f pe p−es − ∑ f p · f pe p−s−es + ∑ f pe pk−1−2s−es
e>◦ e>◦ e>◦ e>◦
| {z } | {z } | {z }
A B C
= A +C − ∑ > 1( f peH + pk−1 f pe−1 p−s−es − f p · p−s )
e
= A +C − (A − 1 − f p p−s ) −C − f p−s = 1


The product in 8.12 is also called the Euler product of L-function.
Theorem 8.11.1 (Hecke): Let f be a cusp form of weight k for SL(2, Z) Then
1. L( f , s) has an analytic extension to C
k
2. R( f , s) := (2π)−sΓ (s)( f , s). Then R( f , k − s) = (−1) 2 R( f , s)
k
Note that R is symmetric with respect to Re(s) = 2

We prove this theorem in a general context introduced by Hecke, see §8.12.

8.12 Hecke’s L-functions

The following is due to [Hecke1936]. Let us consider a series of the form


−1 −1 −1
f = f 0 + f 1 qλ + f2 · q2·λ + . . . + fn qnλ +...
8.12 Hecke’s L-functions 93

Where fi ∈ C and λ ∈ R+ . we assume that f is convergent in the unit disk and hence
if we set q = e2πiτ then it defines a holomorphic function

f :H→C

We further assume that f satisfies


 −1 
f = γ · (−iτ)k f (τ)
τ
for some γ = ±1 and k ∈ Q
The L-function attached to f is

fn
L( f , s) = ∑
n=1 ns

which converges in the region Re(s) > a if we assume that fn ∼ na . We have

Γ (s)
Z ∞
= xs−1 e−nx dx
ns 0

which implies
Z ∞  ∞ 
L( f , s)Γ (s) = xs−1 ∑ fn · e−nx dx
0 n=1

we make the change of variables x → 2πx · λ −1 and get

L( f , s) · Γ (s)
Z ∞
R( f , s) := = xs−1 f˜(τ)dx , τ := ix
(2πλ −1 )s 0

Where f = f0 + f˜. Note that the functional equation of f˜ is


 
˜f −1 = f0 · (γxk − 1) + γxk f˜(τ)
τ

We have
Z 1  
˜f −1 x−2 dx + xs−1 f˜(τ)dx
Z ∞ Z ∞
1−s
R( f , s) = + = x
0 1 1 τ
Z ∞ Z ∞
= γx−1−s+k f˜(τ)dx + xs−1 f˜(τ)dx + f0 x−1−s (γxk − 1)dx
1 1
 
1
Z ∞
−1−s+k

s−1 ˜ γ
= γx +x f (τ)dx + f0 · −
1 −s −s + k

We have used Re(s) > k, Re(s) > 0 in order to compute the last integral. the first
integral is a holomorphic entire function in s ∈ C and so we conclude that R( f , s)
extends to a meromorphic function in s ∈ C with possible poles at s = 0 and s = k.
Moreover, it satisfies
94 8 Riemann zeta function

R( f , k − s) = γ R( f , s)
Here is the place, where we use γ = ±1
Riemann has used the theta series
+∞ 1 2
∞ 1 2
θ3 = ∑ q2n = 1+2 ∑ q2n
n=−∞ n=1

In this case  
−1 1
θ3 = (−iτ) 2 θ3 (τ)
τ
and so k = 12 , γ = +1, λ = 2. In this case

1
L(θ3 , s) = 2 ∑ = 2ξ (2s)
(n2 )s

and so we get the functional equation of ξ (s)

Note that we could also use Einstein series and define.


∞ ∞
σk−1 (n) 1
L(Ek , s) = ∑ = ∑ ∑
n ns d=1 m=1 d s−k ms
= ξ (s − k) ξ (s)

and so we will get again the functional equation of ξ (s)


Exercise 55 Use the functional equation of E2
 
−1 12
E2 = τ 2 E2 (τ) + τ
τ 2πi

and describe L(E2 , s).

8.13 L-function of CY-modular forms

The theory of Calabi-Yaw forms developed in [Mov16, ?] still lives its infancy.
The lack of applications is one the big obstacles for developing the theory further.
Looking at the history of (elliptic) modular forms, the author feel himself like E.
Hecke who is one of the main resposables for the arithmetic of modular forms,
however , he didnt see the most amazing arithmetic applications of modular forms
in his life time. In the present text, we would like to develope of L-functions attached
to CY-modular forms.
Recall the classical L-function attached to a cusp form f .
8.13 L-function of CY-modular forms 95


Z ∞ Z
L( f , s) = f (τ)τ s · =
o τ ℘

We regard ℘ as path ℘1 −℘2 , where ℘1 connects i∞ to i and ℘2 is


 
0 1
℘2 = S ·℘1 , S =
−1 0
Which connects i to O. insome sense, L is attached to S and its fixed point i.
Now, consider A = −1 1 1 , with A3 = −I, AP = P.
0
Under the iteration of A we have
i∞ → −1 → 0 → i∞
−1
τ → τ+1
−τ → τ+1 → τ

Furthermore, we have the functional equations


   
τ +1 k −1
f = (−τ) f (τ) f = (τ + 1)k f (τ)
−τ τ +1
Let δ1 be the path Re(P) which connects i∞ to P, δ2 = A δ1 , δ3 = Aδ2 , see Figure
Let also

℘1 = δ1 − δ2 , ℘2 = δ2 − δ3 , ℘3 = δ3 − δ1
(8.13)
Z
Li := f (τ)τ S−1 , i = 1, 2, 3
℘i

We have
L1 + L2 + L3 = 0
The integrals in 8.13 are convergent at i∞, −1, 0 respectively.
For τ = it we have
| f (it)| 6 M · e−2πt
and
lim e−t · t m = 0 ∀m > 1
t→+∞

Therefore
Z Z ∞ Z +∞
S−1 S−1
t Re(s)−1+1

f (τ)τ dτ 6
P f (it)t dt 6 M
P

1
96 8 Riemann zeta function

Fig. 8.1 L-function


Chapter 9
Congruence groups

In this chapter we work with modular forms for a congruence subgroup of SL(2, Z).
One of the most well-known applications of such modular forms is the so-called
arithmetic modularity theorem.

9.1 Congruence groups

We have seen that SL(2, Z) appears as the monodromy group of the Weierstrass
familly of elliptic curves. If we take other families of elliptic curves and compute
the corresponding mondromy group then we will get subgroups of SL(2, Z) of finite
index. Congruence groups are the most well-known subgroups of SL(2, Z). Let N
be a positive integer number. Define
   
10
Γ (N) := A ∈ SL(2, Z)|A ≡ ( mod N)
01

It is the kernel of the canonical homomorphism of groups SL(2, Z) → SL(2, Z/NZ).

Definition 9.1.1 A subgroup Γ ⊂ SL(2, Z) is called a congruence subgroup of level


N if
(N) ⊂Γ
1. Γ 
1N
2. A A−1 ∈ Γ ∀A ∈ SL(2, Z)
0 1
Our main examples are
   
∗∗
Γ0 (N) := A ∈ SL(2, Z) A ≡
( mod N)
0∗
   
1∗
Γ1 (N) := A ∈ SL(2, Z) A ≡
( mod N)
01

97
98 9 Congruence groups

For a description of a fundamental domain for the action of Γ0 (p), p a prime number,
see Apostol’s book [Apo90] Theorem 4.2.
Definition 9.1.2 A holomorphic function f : H → C is called a moduler form of
weight k for Γ if
1. f |k A = f ∀A ∈ Γ

lim f |k A = exists and < ∞


2. For all A ∈ SL(2, Z)
Im(τ) → +∞

9.2 Weil pairing

Let Let Λ = Zω1 + Zω2 ⊂ C be a lattice with Im( ω 1


ω2 ) > 0 and let E = C/Λ be the
corresponding elliptic curve. For a natural number N ∈ N an N-torsion [z] ∈ E is an
element with Nz = 0 in E, or equivalently Nz ∈ Λ . The subgroup of E of N-torsions
is
1
Λ
E[N] : = {z ∈ E|nz = 0} ' N
 Λ 
aω1 + bω2 Z Z Z
' a, b ∈ ' ×
N NZ NZ NZ

The following definition of Weil pairing is taken from [Silverman II], page 89.
Definition 9.2.1 Let E be an elliptic curve. The Weil pairing is

eN E[N] × E[N] → µN
 
aω1 + bω2 cω1 + dω2 ad−bc
eN , = e2πi N
N N

Here 2πik
µN := {e N | k ∈ Z}.

Exercise 56 Prove that the above definition is well-defined.


Theorem 9.2.1 Let

Y0 (N) := Γ0 (N)\H, Y1 (N) := Γ1 (N)\H, Y (N) := Γ (N)\H.

1. The set Y0 (N) is the moduli space of pairs (E,C), where E is a complex elliptic
curve and C is a cyclic subgroup of E of order N.
2. The set Y1 (N) is the moduli space of pairs (E, p), where E is a complex elliptic
curve and p is a point of E of order N.
9.4 Modular forms for congruence groups 99

3. The set Y (N) is the moduli space of pairs (E, (p, q)), where E is a complex elliptic
curve and (p, q) is a pair of points of E that generates the N-torsion subgroup of
2πi
E with Weil pairing eN (p, q) = e N .

Proof. We only prove item 2. The others are left to the reader.

Exercise 57 Prove items 1 and 3 above.

9.3 Moduli spaces of elliptic curves

We consider the following moduli spaces, which we call them period domains. The
difference between these moduli spaces and those in theorem (9.2.1) is the presence
of the differential form ω together with E. Recall that by integration the pair (E, ω)
is identified with a lattice Λ ⊂ C.

P := moduli of elliptic curves (E, ω)


P1 (N) := moduli of (E, ω, z), z ∈ E[N]
P(N) := moduli of (E, ω, {z1 , z2 }), z1 , z2 ∈ E[N], eN (z1 , z2 ) = ζN
P0 (N) := moduli of (E, ω,C), C ⊆ E[N] cyclic group of order N.

Let f be an elliptic function of weigh k with poles at 0. For instance, we use f =


℘,℘0 which are of weight 2 and 3, respectively. We know the following functions
on period domains

f1,N : P1 (N) → C, f1,N (E, z) = f (Λ , z)


fNi : P(N) → C, fN (E, z1 , z2 ) = f (Λ , zi ), i = 1, 2
f0,N : P0 (N) → C, f0,N (E,C) = f (Λ , z) = ∑ f (Λ , z)
z∈C

Any function g as above has the following functional equation:

g(E, aω, ∗) = a−k g(E, ω, ∗), ∀a ∈ C∗ . (9.1)

where k is the wight of the elliptic function f .

9.4 Modular forms for congruence groups

We consider the following maps


100 9 Congruence groups

1
i : H → P1 (N) τ 7→ (E, ω, )
 N 
τ 1
i : H → P(N) τ 7→ E, ω, ,
N N 
 1 2 N −1 
i : H → P0 (N) τ 7→ E, ω, , ,...,
N N N
here E = C/Zτ + Z and ω = dz
Proposition 9.4.1 Let f be an elliptic function of weigh k with poles at 0. Then
the pull-back of f1,N , fN , f0,N by i is a holomorphic modular form of weight k for
Γ1 (N),Γ (N), and Γ0 (N) respectively.

Proof. We only prove the P1 (N) case


   
aτ + b aτ + b 1
f1,N = f Z + Z,
cτ + d cτ + d N
cτ + d
= (cτ + d)+k f (Z(aτ + b) + Z(cτ + d),
N
1
= (cτ + d)+k f (Zτ + Z, )
N
 
ab
For all ∈ Γ1 (N). Now we have to show the growth condition. For this it is
cd
enough to assume that f = ℘ or ℘0 . In these cases the affirmation follows from
1 1
lim ℘(Zτ + Z, z) = ∑ 2
− 2
ℑ(τ)→∞ m∈Z (z − m) m
1
lim ℘0 (Zτ + Z, z) = −2 ∑
ℑ(τ)→∞ m∈Z (z − m)3

t
u

Exercise 58 Show that


 1
#[SL(2, Z) : Γ1 (N)] = deg(P1 (N) → P) = n2 Π 1 − 2
p
 1
#[SL(2, Z) : Γ (N)] = deg(P(N) → P) = n Π 1 − 2
3
p
 1
#[SL(2, Z) : Γ0 (N)] = deg(P0 (N) → P) = n Π 1 −
p

9.5 q-expansion

Let f be a modular form for a congruence group Γ of level N. It follows that for all
A ∈ SL(2, Z), f |k A has a q-expansion. Let
9.5 q-expansion 101
2πiτ
qN := e N

We have    
11 N 1N
T= ,T = ∈ Γ (N)
00 0 1
and so
( f |k A)T N = f |k (AT N A−1 )A = f |k A
This implies that

f |k A = ∑ an qnN , an ∈ C.
n=0

Proposition 9.5.1 If f is a modular form of weight k for SL(2, Z) then g(τ) :=


f (Nτ) is a modular form of weight k for Γ0 (N).
 
k−1 N0
Proof. First note that N · g = f |k and so g(τ) is a priori a modular form
0 1
for  −1  
N0 N0
SL(2, Z) .
0 1 0 1
 
ab
For A = ∈ SL(2, Z) we have
cd
 −1    
a N −1 b
  
N 0 ab N0
=
0 1 cd 0 1 N ·c d

which means that g is modular for Γ0 (N).

Exercise 59 Compute the q-expansions of ℘1,N , ℘N , ℘0,N .


We can construct modular functions for congruence groups by division for in-
stance:
∆ (Nτ)
ϕ(τ) :=
∆ (τ)
is a modular function for the group Γ0 (N). This follows from the functional equation
of ∆ (τ) is follows that ϕ is holomorphic in H It is also holomorphic at i∞:

qN ∏ (1 − qnN )24 ∞
!
n−1 N−1 n
ϕ(τ) = ∞ =q 1 + ∑ bn q
q ∏ (1 − q) n=1
n−1

and actually it has a zero of order N −1 at q = 0. There are no non-constant holomor-


phic functions on a compact Riemann surface, and so, ϕ as a function on Γ0 (N)\H
is necessarily meromorphic. Since in Γ0 (N)\H ∪ {i∞} it has only one zero which is
102 9 Congruence groups

of order N − 1 doing q-expansions in other cusps we must have poles. For this we
use
∑ ord∗ ( f ) = 0
∗∈Γ0 (N)\Q

Note that the Fourier coefficients bn of ϕ are integers. For more information on ϕ
see Apostol’s book [Apostol], section 4.7

9.6 Transcendental degree of modular forms

Let Γ ⊆ SL(2, Z) be a subgroup of finite index a. Let also f ∈ Mk (Γ ). We define


a
∑ ga−i · X i = ∏ (X − f |k A) , g0 := 1 (9.2)
i=0 A∈Γ \SL(2,Z)

Proposition 9.6.1 We have gi ∈ Mk·i (SL(2, Z)).

Proof. Let P(X) be the right hand side of 9.2. Then for B ∈ SL(2, Z) we have


P(X)|k B = ∏ X − ( f |k A)|k B = P(X)
A∈Γ \SL(2,Z)

Therefore, the coefficients of P(X) has the correct functional equation. The finite
growth of gi ’s follow from the finite growth of f |k A’s for all A ∈ Γ \SL(2, Z). t
u

Conversely, let us be given gi ∈ Mk·i SL(2, Z) , i = 1, 2, ..a, and define
a
P(X) = ∑ ga−i · X i
i=0

The resultant of P(X) is a homogeneous polynomial of degree 2 · k · a in

Q[g1 , g2 , . . . , ga ], weight(gi ) = ki

This is a weight 2ka modular form for SL(2, Z). Assume that this resultant has no
zeros. Since H is simply connected, we can find holomorphic functions f1 , f2 , .., fa :
H → C such that
P(X) = (X − f1 )(X − f2 ) · · · (X − fa )
We have the representation

χ : SL(2, Z) → GL(a, Z)

whose image is isomorphic to the permutation group in a elements and such that
9.6 Transcendental degree of modular forms 103
 
f1 |k A  
f1
 f2 |k A   f2 
 ..  = χ(A)  , ∀A ∈ SL(2, Z). (9.3)
 
 .   
fa |A f a

Here χ(A) is just a permutation matrix in 1, 2, .., a. Let



Γi := {A ∈ SL(2, Z) χ(A)ei = ei }

1 , 0, . . . , 0]tr .
where ei = [0, 0, . . . , |{z}
i−th place

Proposition 9.6.2 We have fi ∈ Mk (Γi ).

Proof. This is a direct consequence of 9.3 and the definition of Γi . t u



Exercise 60 Show that P(X) is irreducible over Mk SL(2, Z) [X] if and only if an
orbit of χ in {1, 2, . . . , a} is the whole set. It follows that if P(X) is irreducible over
Mk SL(2, Z) [X]. Then

{ f1 , f2 , . . . , fa } = { fi |k A, A ∈ Γ \SL(2, Z)}
Γi := A−1 · Γ1 A A ∈ Γ \SL(2, Z)

The following question is natural to ask: under which conditions on gi ’s, Γ1 is a


congruence group?We have proved:
Theorem 9.6.1 The ring of modular forms for congruence groups is of transcen-
dental degree 2. More, precisely any modular form for a congruence group is in the
algebraic closure of C(E4 , E6 ).
Chapter 10
Elliptic curves as Diophantine equations

10.1 Finite fields

A finite field, as its name indicates, is a field with finite cardinality. By definition of
a field and finiteness property, the characteristic of a finite field is a prime number
p > 1. Finite fields are completely classified as follows:
1. The order of a finite field of characteristic p is pn for some n ∈ N.
2. There is a unique (up to isomorphism of fields) finite field with pn , n ∈ N ele-
ments.
3. For a prime number the finite field with cardinality p is simply the quotient

Z
F p := .
pZ
4. For q = pn , n ∈ N the finite field with cardinality pn is denoted by Fq . It is the
spliting field of the polynomial xq − x over F p .
5. Every finite integral domain is a field and in particular
6. Let f (T ) be a monic irreducible polynomial of degree n in F p [T ]. Then the quo-
tient Fq [T ]/h f i is a finite field with pn elements.
7. Let f (x, y) ∈ F p [x, y] be a polynomial and I be a non zero prime ideal of R :=
F p [x, y]/h f i. Then the quotient R/I is a finite field.
For more on finite fields the reader is referred to [Jac85].

10.2 Zeta functions of elliptic curves over finite fields

Let V be an affine or projective variety defined over Fq . The zeta function of V is


defined to be the formal power series in T :

105
106 10 Elliptic curves as Diophantine equations

#V (Fqr ) r
Z(V, T ) = exp( ∑ T )
r=1 r

Theorem 10.2.1 Let E be an elliptic curve defined over F p . Then

1 + 2aE T + pT 2
Z(E, T ) = . (10.1)
(1 − T )(1 − pT )

where aE is an integer depending only on E. Moreover, the Riemann hypothesis


holds for E, i.e. the only zeros of

ζ (C, s) := Z(E, q−s )

are in the line ℜ(s) = 12 .


Let
1 − 2aE T + qT 2 = (1 − αT )(1 − β T )
and so
α + β = 2aE , αβ = q (10.2)
Note that α and β are algebraic integers:
q
α, β = aE ± a2E − q.

We take the logarithmic derivative of both sides of (10.1) and one easily finds the
equalities
#E(F pr ) = pr + 1 − α r − β r , r = 1, 2, 3, . . .
For r = 1 we obtain
#E(F p ) = p + 1 − 2aE
We conclude that for elliptic curves over a finite field Fq the number of Fq -rational
points determine the number of Fqr -rational points.
Concerning the Riemann hypothesis, we note that it is equivalent to the inequal-
ity:

|#E(F p ) − p − 1| < 2 p. (10.3)
1
The Riemann hypothesis holds if and only if |α| = |β | = p 2 . If these equalities
happen then

|#E(F p ) − p − 1| = |2aE | = |α + β | < 2 p.
(the equality cannot occur because p is prime). Conversely, if (10.3) happens then
a2E − p < 0 and so the roots of the polynomial 1 − 2aE T + qT 2 are complex conju-
1
gate, β = ᾱ and so |α| = |β | = p 2 .
The general reuslt as in (10.1) was conjectured by André Weil [Wei49] and was
proved by P. Deligne (see for instance [Kat76] for an exposition of Deligne results).
10.4 Mazur theorem 107

10.3 Nagell-Lutz Theorem

In this section we state Nagell-Lutz theorem which gives a finite set of of possibili-
ties for a torsion point of an elliptic curve.
Theorem 10.3.1 (Nagell-Lutz Theorem) Let E be an elliptic curve with the Weier-
strass equation:

y2 = x3 + t2 x + t3 , t2 ,t3 ∈ Z, ∆ := 4t23 + 27t32 6= 0.

Then for all non-zero torsion points P = (a, b) ∈ E(Q) we have:


1. The coordinates of P are in Z, i.e. a, b ∈ Z.
2. If P is of order greater than 2, then b2 divides ∆ .
3. If P is of order 2 then b = 0 and a3 + t2 a + t3 = 0.
A proof can be found in [Sil92], p. 221 or in [ST92] p.56.
Exercise 61 [Mil96], Exercise 8.11. For four of the following elliptic curves com-
pute the torsion subgroup.
y2 = x3 + 2, · · ·
See the reference above for the list of elliptic curves.

10.4 Mazur theorem

Theorem 10.4.1 (Mazur, [Maz77, Maz78]) Let E be an elliptic curve over Q. Then
the torsion subgroup E(Q)tors is one of the following fifteen groups:

Z/NZ, 1 ≤ N ≤ 10, or N = 12

Z/2Z × Z/2NZ, 1 ≤ N ≤ 4
Note that the above theorem implies that for an elliptic curve over Q we have always:

#(E(Q)tors ) ≤ 16.

It is natural to conjecture that: If E is an elliptic curve over a number field k, the


order of the torsion subgroup of E(k) is bounded by a constant which depends only
on the degree of k over Q. This is known uniform boundedness conjecture (UBC).
It is proved by S. Kamienny in [Kam92] for all quadratic fields and by L. Merel in
[Mer96] for all number fields.
For the proof of all the statements above one needs the notion of modular curves
X0 (N) and modular forms which will be introduced in the forthcoming chapters.
108 10 Elliptic curves as Diophantine equations

10.5 One dimensional algebraic groups

We follow [Mil96] p. 23. When elliptic curves degenerate we find the following
algebraic groups:
1. The additive group Ga := (A1 (k), +).
2. The multiplicative group Gm := (A(k), ·).
3. Twisted multiplicative group Gm [a].
Exercise 62
Gm [a] ∼
= Gm [ac2 ], a, c ∈ k − 0,
Exercise 63
Gm [a](Fq ) = q + 1.
By Bezout theorem a cubic curve E in P2 has a unique singular point (if there are
two singularities then the line connecting that points meets the curve in 4 points
counted with multiplicities). The singular point is defined over k because it is fixed
under the action of the Galois group Gal(k̄/k). Let S be the singular point of of E
and
E ns (k) := E(k)\{S}.
The same definition of group law for elliptic curves applies for E ns and it turns out
that E ns as a group and:
Exercise 64 If the elliptic curve E is given by the Weierstrass form

y2 = x3 + t4 x + t6 , t2 ,t3 ∈ k, ∆ = 2(4t43 + 27t62 ) = 0.

then E ns is isomorphic to the three one dimensional group described above:

E ns (k) ∼
= Gm (k) or Gm [c](k), or Ga (k)

mentioned in the lectures. Does we need char(k) 6= 2, 3?


Exercise 65 For char(k) = 3 (resp. char(k) = 2) we have to consider the case (4.6)
(resp. (4.4)). Discuss the reduction modulo 2 and 3 in such cases.

10.6 Reduction of elliptic curves

We take an elliptic curve in the Weierstrass form

y2 = x3 + t4 x + t6 , t2 ,t3 ∈ Q, ∆ := 2(4t43 + 27t62 ) 6= 0.

and by change of coordinates (x, y) 7→ (c2 x, c3 y), c ∈ Q we assume that |∆ | is mini-


mal. For p prime different from 2 and 3 we have the curve E/F p and the reduction
map
10.7 Zeta functions of curves over Q 109

E(Q) → E(F p ).
1. Good reduction. If p does not divide ∆ then E/F p is an elliptic curve.
2. Cuspidal reduction/additive reduction. The reduced curve E/F p has a cusp as a
singularity and so its non-singular part is an additive group. If char(k) 6= 2, 3 this
case happens if and only if p | ∆ , and p | 2t4t6 .
3. Nodal reduction/split multiplicative. The reduced curve E ns /F p is a multiplica-
tive group.
4. Nodal reduction/nonsplit multiplicative. The reduced curve E ns /F p is a twisted
multiplicative group.
Exercise 66 Reduction moulo 3 of the above elliptic curve in Weierstrass form is
singular if and only if t4 = 0. In the singular case it is always a cusp. In reduction
modulo 2 the elliptic curve E/F2 is always singular and its singular point is S =
(t4 ,t6 ). Find the four groups E ns (F2 ) corresponding to the four choice of (t4 ,t6 ).
Exercise 67 Let E/Q : y2 + y = x3 − x2 + 2x − 2. Show that 1. the primes of bad
reduction for E are p = 5 and 7. 2. The reduction at p = 5 is additive, while the
reduction at p = 7 is multiplicative. 3. NE/Q = 175.

10.7 Zeta functions of curves over Q

We follow [Mil96] p. 102. The non-complete zeta function of a smooth curve E :


f (x, y) = 0, f ∈ Z[x, y] is defined to be

ζS (E, s) = ∏ ζ (E/F p , s).


p6∈S

where S is a finite number of prime numbers such that E/F p is singular.


Exercise 68 Can you justify the definition of the zeta function of a variety over Q
by interpreting it as a Euler product, the one similar to (8.11).
In the case of elliptic curves it is natural to define
1
LS (E, s) := ∏ s 1−2s
p6∈S 1 + (#(E(F p )) − p − 1)p + p

and so we have
ζS (s)ζS (s − 1)
ζS (E, s) =
LS (E, s)

Proposition 10.7.1 The product ζS (E, s) and hence LS (E, s) converges for ℜ(s) > 23

Proof. It is direct consequence of the Riemann hypothesis for elliptic curves over
finite fields and the convergence of the Riemann zeta function(see [Mil96] p.102).
110 10 Elliptic curves as Diophantine equations

We we want to define the complete L function by adding bad prime numbers p ∈ S.


We define
 2
 1 + (#(E(F p )) − p − 1)T + pT p good

1−T modulo p we have split multiplicative reduction

L p (T ) =

 1 + T modulo p we have non-split multiplicative reduction
1 modulo p we have additive reduction

We have defined this in such a way that

#E ns (F p )
L p (p−1 ) =
p
Now we define the L-function of an elliptic curve E over Q:
1
L(E, s) = ∏ .
p L p (p−s )

10.8 Hasse-Weil conjecture

The conductor of an elliptic curve over Q is defined to be

NE/Q = ∏ p fp
p bad

where f p = 1 if E has multiplicative reduction at p, f p = 2 if p 6 |2, 3 and E has


additive reduction at p. For the case in which we have additive reduction modulo
p = 2, 3 we have f p ≥ 2, f p ∈ N and f p depends on wild ramification in the action
of the inertia group at of Gal(Q̄/Q) on the Tate module of E.
Exercise 69 Discuss the case p = 2, 3 in the above definition. [Mil96] is also talking
about a formula of Ogg f p = ord p ∆ + 1 − m p using Néron models. Can you obtain
some information on this.
Define s
2
Λ (E, s) := NE/Q (2π)−sΓ (s)L(E, s)

Theorem 10.8.1 (Hasse-Weil conjecture for elliptic curves) The function Λ (E, s)
can be analytically continued to a meromorphic function on the whole C and it
satisfies the functional equation

Λ (E, s) = ±Λ (E, 2 − s).

This theorem was first proved for CM elliptic curves by Deuring 1951/1952. It is
proved in its generality by the works of Eichler and Shimura, Wiles, Taylor, Dia-
mond and others.
10.10 Congruent numbers 111

10.9 Birch Swinnerton-Dyer conjecture

For the functional equation of L the value s = 1 is in the middle, i.e. it is the fixed
point of s 7→ 2 − s.
Conjecture 10.9.1 (BSD conjecture) For an elliptic curve E over Q, the function
L(E, s) is holomorphic at s = 1 and its order of vanishing at s = 1 is the rank of the
elliptic curve E.
A weak form of this conjecture is not also proved:
Conjecture 10.9.2 (weak BSD conjecture) L(E, 1) = 0 if and only if E has in-
finitely many rational points.
For papers on BSD conjecture see [CW77, BSD63, BSD65, Tun83, Lan78b]
[Ser89], [Mor69].

10.10 Congruent numbers

A natural number n is said to be congruent if it is the area of a right triangle whose


sides have rational length. In other words we are looking for the Diophantine equa-
tion:
1
Cn : x2 + y2 = z2 , n = xy
2
in Q, where x, y and z are the sides of the triangle. Consider the affine curve Cn /Q
in A3 defined by the above equations. It intersects the projective space at infinity in
4 points:
[x; y; z; w] = [0; ±1; 1; 0], [±1; 0; 1; 0].
Let
C : y2 = x4 − n2 , En : y2 = x3 − n2 x.
We have morphisms
z x2 − y2
Cn → C, (x, y, z) 7→ ( , )
2 4
and
C → En , (x, y) 7→ (x2 , xy)
defined over Q.
Proposition 10.10.1 A necessary and sufficient condition for the point (x, y) ∈
En (Q) be in the image of Cn (Q) → En (Q) is that
1. x to be a square and that
2. its denominator be divisible by two
3. and its numerator has no common factor with n.
The proof is simple and is left to the reader (see [Kob93]).
112 10 Elliptic curves as Diophantine equations

Exercise 70 Let C̄n be the projectivization of Cn in P3 . Is C̄n smooth? If yes deter-


mine its genus.
Exercise 71 Ex. 1,2,3,4 of Koblitz, page 5.

We want to analyze the torsion points of

En : y2 = x3 − n2 x

By definition of the group structure of En we know that

O, (0, 0), (0, ±n)

are 2-torsions of En . Following the lines of [Kob93] p. 44 Proposition 4, we want to


prove:
Proposition 10.10.2 We have

En (Q)tors = {O, (0, 0), (0, ±n)}

and so #En (Q)tors = 4.

Proof. Let us first give the strategy of the proof. Let E/Q be an a elliptic curve in
the Weierstrass form and let p > 2 be a prime number which does not divide the
discriminant of E. By a linear change of variable (x, y) 7→ (a2 x, a3 y) we can assume
that the ingredient coefficients of E are in Z. Let Ē/F p be the elliptic curve obtained
from E by considering the coefficients of E modulo p. The main ingredient of the
proof is the reduction map
E(Q) → Ē(F p ),
which is a group homomorphism. Note that by our assumption on p, Ē/F p is not
singular. This is an injection of E(Q)tors inside E(F p ) for all but finitely many p and
so for such primes m := #E(Q)tors divides #E(F p ). In fact, we have not yet proved
that E(Q)tors is finite (a corollary of Mordell-Weil theorem). Therefore, we take a
finite subgroup G of #E(Q)tors and prove that the reduction map restricted to G is
an injection and so m := #G divides #E(F p ). From another side, we prove that for
E = En :
#En (F p ) = p + 1, ∀p prime p ≡ −1 mod 4 (10.4)
Therefore, for all but finitely many primes p ≡ −1 mod 4 we have p ≡ −1 mod m.
This implies that m = 4. Therefore, every finite subgroup of E(Q)tors is of order 4.
Since all the elements of E(Q)tors are torsion, we conclude that #En (Q)tors = 4.
Now let us prove that the reduction map induces an injection in a finite subgroup
G of E(Q)tors . Two points P = [x; y; z], Q = [x0 ; y0 ; z0 ] ∈ E(Q) are the same after
reduction if and only if
xy0 − x0 y, xz0 − x0 z, yz0 − y0 z (10.5)
are zero modulo p. For all pairs P, Q in G, the number of numbers (10.5) is finite and
so there are finitely many primes dividing at least one of them. For all other primes
10.10 Congruent numbers 113

p, we have the injection of G in E(F p ) by the reduction map. The proof of (10.4) is
done in the next proposition.
Proposition 10.10.3 Let q = p f , p 6 |2n. Suppose that q ≡ −1 mod 4. Then there
are q + 1 Fq points on the elliptic curve En : y2 = x3 − n2 x.
Proof. Consider the map

f : Fq → Fq , f (x) = x3 − n2 x

f is an odd function, i.e. f (−x) = − f (x), and −1 is not in its image (this follows
from the hypothesis on p). It follows that the index of the multiplicative group F2q −
{0} in Fq − {0} is two and so for all x ∈ Fq − {0} exactly one of x or −x is square
and so for all x ∈ Fq − {0, n, −n} exactly one of f (x) or f (−x) is square. Each such
a pair (x, y), y = f (x) gives us two points (x, y), (x, −y) ∈ En (Fq ) and so in total we
have 3 + 2 q−12 points in En (Fq ).

Proposition 10.10.4 The natural number n is congruent if and only if En (Q) has
non-zero rank.
Proof. If n is a congruent number then by Proposition 10.10.1, En has Q-rational
point with x-coordinate in (Q+ )2 . The x coordinates of 2-torsion points in the affine
chart x, y are 0, ±n. The fact that n is square free and Proposition 10.10.2 implies
that such a rational point is of infinite order.
Conversely, suppose that P is a rational point of infinite order in En . Then by
Exercise 72, the One can
Exercise 72 ([Kob93], p. 35, Ex. 2c) If P is a point not of order 2 in En (Q), then
the x-coordinate of 2P is a square of rational number having an even denominator.
By Proposition 10.10.1, 2P comes from a point in Cn (Q) and hence n is a congruent
number.
Exercise 73 ([Kob93], p. 49-50) Ex. 4,5,6, 7,9.
Let us now state the main result in §10.2 for the elliptic curve En related to the
congruent numbers. The Legendre symbol is defined for integers a and positive odd
primes p by 
   0 if p divides a
a
= 1 for some x ∈ Z, a ≡ x2 mod p
p
−1 otherwise

Proposition 10.10.5 In the zeta function of En : y2 = x3 − n2 x defined over F p , p a


prime p 6 |2n, we have:
( √
i p  if p ≡ 3( mod 4) in this case aEn = 0
α=
 
2k + p + 2ki if p ≡ 1( mod 4) in this case aEn = 2k + np
n

In the second case k is determined by the fact that α ᾱ = p


114 10 Elliptic curves as Diophantine equations

Exercise 74 ([Mil96], Ex. 19.12) Let E be the elliptic curve

E : y2 = x3 − 4x2 + 16

Consider also the formal power series given by



F(q) = q ∏ (1 − qn )2 (1 − q11n )2 = q − 2q2 − q3 + 2q4 + · · ·
n=1

1. Compute N p = #E(F p ) for all primes 3 ≤ p ≤ 13.


2. Calculate the coefficient of Mn of qn in F(q) for n ≤ 13.
3. Compute the sum M p + N p for p prime p ≤ 13.
4. Formulate a conjecture on the sum M p + n p . Can you prove it?.
The bad prime numbers for the elliptic curve En : y2 = x3 − nx are those which
divide 2n. For p | 2n, p 6= 2 or p = 2, 2|n we have an additive reduction. For p = 2
and p 6 |n we have apparently a multiplicative reduction: y2 = x3 + x. The singular
point in this case is S = (1, 0) and E ns (F2 ) = {O, (0, 0)} which is isomorphic to
(A(F2 ), +) and so it is additive.
The conductor of En is:
 4 2
2 n if n is even
NEn /Q =
25 n2 if n is odd

In Theorem 10.8.1 the root number ± is determined in the following way:



+1 if n ≡ 1, 2, 3 (8)
−1 if n ≡ 5, 6, 7 (8)

Reformulating Proposition 10.10.5 and using Exercise 54 we have:

(1 − T )(1 − pT )Z(En /F p , T ) = ∏ (1 − (αp T )deg(p) )


p|hpi

where
 √

 i p if p = hpi
 a + ib if p splits, where a + ib is the unique generator of p
αp =

 which is congruent to ( np ) mod 2 + 2i.
0 p | 2n

The L function of En is

L(En , s) = ∏ (1 − (αp )deg(p) (Np)−s )−1


p⊂Z[i] prime

Now Z[i] is a Dedekind domain and so we can define a unique map χ from the ideals
deg(p)
of Z[i] to C such that χn (p) = αp . Therefore
10.11 p-adic numbers 115

L(En , s) = ∏ (1 − χ(p)(Np)−s )−1 = ∑ χn (a)(Na)−s


p⊂Z[i] prime a⊂Z[i]

where the sum is taken over all non-zero ideals.

10.11 p-adic numbers

By definition a p-adic integer is an element in the inverse limit of

· · · → Z/p3 Z → Z/p2 Z → Z/pZ.

One can show that a p-adic integer is identified with a formal series

a1 p + a2 p2 + a3 p3 + · · · , ai ∈ {0, 1, 2, . . . , p − 1}.

The set of p-adic integers is denoted by Z p :

Z p := lim Z/Zpn .
∞←n

Z p is a ring without zero divisor, i.e. if ab = 0, a, b ∈ Z p then either a = 0 or b = 0.


The field Q p of p-adic numbers is the quotient field of Z p . The ring Z of integers is
a subring of Z p in a natural way and so Q p is a field extension of Q, Q ⊂ Q p .
Exercise 75 Show that the Diphantine equation x3 + y3 − 3 = 0 has not a solution
in Q3 and hence it has not a solution in Q.

There is another way to define p-adic numbers. Any non-zero rational number a can
be expressed in the form a = pr mn with m, n ∈ Z and not divisable by p. We define

1
ord p (a) := r, |a| p := , |0| p := 0
pr
We have
1.
|a| p = 0 if and only if a = 0
2.
|ab| p = |a| p |b| p , a, b ∈ Q.
3.
|a + b| p ≤ max{|a| p , |b| p } and so ≤ |a| p + |b| p
Therefore,
d p (a, b) := |a − b| p
is a metric on Q. The field Q p of p-adic numbers is the completion of Q with
respect to d p . We have a cononical matric, call it again d p , on Q p which extends the
116 10 Elliptic curves as Diophantine equations

previous one on Q ⊂ Q p (this inclusion is given by sending a ∈ Q to the constant


Cauchy sequence a, a, . . .). The same construction with the usual norm of Q, i.e.
d(a, b) = |a − b| yields to the field of real numbers R.
Exercise 76 Prove that the two definitions of Q p presented above are equivalent.
Prove also
Z p = {a ∈ Q p | |a| p ≤ 1} = the closure of Z in Q p
Chapter 11
Theta series

Jacobi’s theta function is the followingg infinite sum


+∞

θ (z, τ) = ∑ e2πinz+πin
n=−∞

Proposition 11.0.1 The Jacobi’s theta function satisfies:


1. It is a holomorphic function in C × H
2. θ (z + 1, τ) − θ (z, τ)
3. θ (z + τ, τ) − θ (z, τ) e−πiτ e−2πiz
4. θ (z, τ) − 0 for z = 12 + τ2 + n + mτ, n, m ∈ Z

Proof. 1. For |Z| < M and Im(τ) > to we have


∞ ∞
2πinZ+πin2 τ 2
∑ e 6 C ∑ e2πnM e−πn t0
n=−∞ n=0

for some C ∈ R+ . This shows that θ is converges is C × H.


2. This is immediate
3.
+∞
2 +2n)τ
θ (z + τ, τ) = ∑ e2πinz eπi(n
n=−∞
+∞

= ∑ e2πi(n+1)z eπi(n+1) e−πiτ e−2πiz
n=−∞
= θ (z, τ) · e−πiτ e−2πiz

4.   ∞
1 τ 2 +n)τ
θ + ,τ = ∑ (−1)n eπi(n
2 2 n=−∞

For n > 0 the terms corresponding to n and −n − 1 cancel each other. 

117
118 11 Theta series

We will frequently use the followings:


∞ 1 2
θ3 (τ) = θ (0, τ) ∑ q2n
n=−∞
∞ 1 2
1
∑ (−1)n q 2 n

θ4 (τ) = θ 2,τ
n=−∞
1 +∞ 1 1 2
= q− 8 ∑ q 2 (n+ 2 )
τ

θ2 (τ) = θ 2,τ
n=−∞

Theorem 11.0.1 We have


∞  1
 1

θ (z, τ) = ∏ (1 − qn ) 1 + qn− 2 e2πiz 1 + qn− 2 e−2πiz (11.1)
n=1

Proof. Let π(z, τ) be the right hand side of 11.1. We prove that π(z, τ) is a holo-
morphic function in C × H and satisfies the same properties as of θ (z, τ) in
1. For the convergence we use the criterion for convergence of infinite products.
 1
 1
  
(1 − qn ) 1 − qn− 2 e2πiz 1 − qn− 2 e2πiz = 1 + O |9|n−1 e2π|z|


and ∑ |9|n converges.
n=1
2. This is immediate
3.
∞  1
 3

π(z, τ) = ∏ (1 − qn ) 1 − qn+ 2 e2πiz 1 − qn− 2 e−2πiz
n=1
 1

1 − q− 2 e−2πiz
=  1
 π(z, τ)
1 − q 2 e2πiz

1+χ
We have 1+χ −1 = χ, χ 6= −1 and 3 follows.

4. The product vanishes at a point (z, τ) if


 
1 1
±Z + n − τ ∈ Z+
2 2

which gives us the result.


(z,τ)
Now, let us prove the equality 11.1. Let F(z, τ) = θπ(z,τ) . This as a function in
z is double periodic and has no poles. Therefore, it is constant as a function in z,
Therefore C(τ) = θ (z, τ)/π(z, τ).
We put z = 12
11 Theta series 119
∞ 1 2
∑ (−1)n q 2 n
n=−∞
C(τ) = ∞    (11.2)
1 1
∏ 1 − q 2 n 1 − qn− 2
n=1
1
We put z = 4

+∞ 2
∑ (−1)n q2n
n=−∞
C(τ) = ∞ (11.3)
∏ (1 − q2n ) (1 − q4n−2 )
n=1
k
11.2 and 11.3 imply C(4τ) = C(τ) for all τ ∈ H. Since q4 → 0 when k → ∞ we
can conclude that C(τ) = 1 

We have
η(τ)5
θ3 (τ) = 2 (11.4)
η 21 τ η(2τ)2
2
η 12 τ
θ4 (τ) = (11.5)
η(τ)
2η(2τ)2
θ2 (τ) = (11.6)
η(τ)

Taken from Wikipedia and [Oliver]. Note that θ2 θ3 θ4 = 2η(τ)3 .


Theorem 11.0.2 For τ ∈ H and z ∈ C we have
  r
−1 τ πiτz2
θ z, = e θ (zτ, τ)
τ i
q
Here we have chosen a branch of τi , τ ∈ H such that for imaginary τ, it is positive.

Proof. It is enough to prove the formula z = α ∈ R and τ = it,t ∈ R+ . This is exactly


the Poisson summation formula.


we get the following functional equations for θ2 , θ3 , θ4
  r
−1 τ
θ3 = θ3 (τ) (11.7)
τ i
  r
−1 τ
θ4 = ξ8 · θ2 (τ)
τ i
  r
−1 τ
η = · η(τ)
τ i
120 11 Theta series

Let f (τ) = θ3 (8τ)8

l f (τ + 1) = f (τ)
   
−1 τ 4
f = f (τ)
4τ 2

Which says that f is a modular form for the group


   −1 
11 0
, 1 2 ⊆ SL(2, Q)
01 2 0

11.1 Two-squares theorem

For k ∈ N a = (a1 , a2 , . . . ak ) ∈ Zk

γk , a(n) = #{(x1 , x2 , . . . , xk ) ∈ Zk a1 x12 + a2 x22 + · · · + ak xk2 = n}


Then

θ (2a1 τ) θ (2a2 τ) · · · θ (2ak τ) = ∑ γk , a(n)qn . (11.8)
n=0

Therefore, in order to find formulas for γk , a(n) we have to study the analytic
function in the left hand side of 11.8.
Let d1 (n) denotes the number of divisors of n of the form 4k + 1, and d3 (n) the
number of divisors of n of the form 4k + 3.
Theorem 11.1.1 For n > 1
 
γ2 (n) = 4 d1 (n) − d3 (n) (11.9)

The generating function of the right hand side of 11.9 is


+∞
1
C(τ) : = 2 ∑ n + q−n
n=−∞ q

qn
= 1+4 ∑
n=1 1 + q2n

qn q3n
= 1+4 ∑ 4n

n=1 1−q 1 − q4n

Therefore, in order to prove 11.1.1, we have to prove that

θ32 (2τ) = C(τ)


11.2 Poisson summation formula 121

For the rest of the proof see [Stein] page 299.

11.2 Poisson summation formula

Let ϕ : R → R be any continuous function which decreases rapidly, let us say

ϕ(x) ∼ |x|−c c > 0

as x → +∞. Then the Fourier transform of ϕ is


Z
ϕ̌(y) := ϕ(x) e−2πixy dx
R

The Poisson summation formula says that

∑ ϕ(x + n) = ∑ ϕ̌(x + γ) (11.10)


n∈Z γ∈Z

Proof. (Taken from Zagier) The growth condition on ϕ(x) ensures that the left hand
side of 11.10 converges to a continuous function φ . This function satisfies

Φ(x + 1) = Φ(x)

and so Φ has Fourier expansion


Z 1
Φ(x) = ∑ cγ · e2πiγx cγ = Φ(x) e−2πiγx dx
γ∈Z 0

substituting cγ in Φ(x) we get


Z 1 ∞ ∞ Z n+1
cγ = ∑ ϕ(x + n) e−2πiγ(x+n) dx = ∑ ϕ(x) e−2πiγx dx
0 n=−∞ n=−∞ n
Z ∞
= ϕ(x) e−2πiγx dx
−∞

This gives us
Z 
∑ ϕ(x + n) = ∑ ϕ(t) e−2πiγt dt e2πiγx
n∈Z γ∈Z R

which is the desired statement. 


Chapter 12
Online supplemental items

12.1 Introduction

In this chapter we explain many procedures of the library foliation.lib of


Singular, [GPS01], a computer programming language for polynomial computa-
tions. We also explain few other softwares which are useful when one deal with
computational aspects of modular forms and elliptic curves.

12.2 How to start?

One has to run Singular in the same directory, where foliation.lib lies. Then
in Singular’s command line one has to type:
LIB "foliation.lib";
In order to get an example and help of a command, for instance PeriodMatrix,
one has to type respectively:
example PeriodMatrix;
help PeriodMatrix;

In this chapter I will only sketch few procedures related to the topic of the present
text. For more information the reader might consult the help and example of each
procedure.

12.3 Ramanujan differential equation

123
References

Apo90. Tom M. Apostol. Modular functions and Dirichlet series in number theory, volume 41 of
Graduate Texts in Mathematics. Springer-Verlag, New York, second edition, 1990.
BSD63. B. J. Birch and H. P. F. Swinnerton-Dyer. Notes on elliptic curves. I. J. Reine Angew.
Math., 212:7–25, 1963.
BSD65. B. J. Birch and H. P. F. Swinnerton-Dyer. Notes on elliptic curves. II. J. Reine Angew.
Math., 218:79–108, 1965.
Cas66. J. W. S. Cassels. Diophantine equations with special reference to elliptic curves. J. London
Math. Soc., 41:193–291, 1966.
CW77. J. Coates and A. Wiles. On the conjecture of Birch and Swinnerton-Dyer. Invent. Math.,
39(3):223–251, 1977.
Edw01. H. M. Edwards. Riemann’s zeta function. Dover Publications Inc., Mineola, NY, 2001.
Reprint of the 1974 original [Academic Press, New York; MR0466039 (57 #5922)].
Gan06. Terry Gannon. Moonshine beyond the Monster. Cambridge Monographs on Mathematical
Physics. Cambridge University Press, Cambridge, 2006. The bridge connecting algebra,
modular forms and physics.
GPS01. G.-M. Greuel, G. Pfister, and H. Schönemann. S INGULAR 2.0. A Computer Algebra
System for Polynomial Computations, Centre for Computer Algebra, University of Kaiser-
slautern, 2001. http://www.singular.uni-kl.de.
Gro78. Benedict H. Gross. On the periods of abelian integrals and a formula of Chowla and
Selberg. Invent. Math., 45(2):193–211, 1978. With an appendix by David E. Rohrlich.
Gro79. Benedict H. Gross. On an identity of Chowla and Selberg. J. Number Theory, 11(3 S.
Chowla Anniversary Issue):344–348, 1979.
Hus04. Dale Husemöller. Elliptic curves, volume 111 of Graduate Texts in Mathematics.
Springer-Verlag, New York, second edition, 2004. With appendices by Otto Forster, Ruth
Lawrence and Stefan Theisen.
Jac85. Nathan Jacobson. Basic algebra. I. W. H. Freeman and Company, New York, second
edition, 1985.
Kam92. S. Kamienny. Torsion points on elliptic curves and q-coefficients of modular forms.
Invent. Math., 109(2):221–229, 1992.
Kat76. Nicholas M. Katz. An overview of Deligne’s proof of the Riemann hypothesis for varieties
over finite fields. In Mathematical developments arising from Hilbert problems (Proc.
Sympos. Pure Math., Vol. XXVIII, Northern Illinois Univ., De Kalb, Ill., 1974), pages 275–
305. Amer. Math. Soc., Providence, R.I., 1976.
Kob93. Neal Koblitz. Introduction to elliptic curves and modular forms, volume 97 of Graduate
Texts in Mathematics. Springer-Verlag, New York, second edition, 1993.
Lan78a. Serge Lang. Elliptic curves: Diophantine analysis, volume 231 of Grundlehren der
Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences].
Springer-Verlag, Berlin, 1978.
Lan78b. Serge Lang. Sur la conjecture de Birch-Swinnerton-Dyer (d’après J. Coates et A. Wiles).
In Séminaire Bourbaki, 29e année (1976/77), volume 677 of Lecture Notes in Math., pages
Exp. 503, pp. 189–200. Springer, Berlin, 1978.
Man63. Ju. I. Manin. Rational points on algebraic curves over function fields. Izv. Akad. Nauk
SSSR Ser. Mat., 27:1395–1440, 1963.
Maz77. B. Mazur. Modular curves and the Eisenstein ideal. Inst. Hautes Études Sci. Publ. Math.,
(47):33–186 (1978), 1977.
Maz78. B. Mazur. Rational isogenies of prime degree (with an appendix by D. Goldfeld). Invent.
Math., 44(2):129–162, 1978.
Mer96. Loı̈c Merel. Bornes pour la torsion des courbes elliptiques sur les corps de nombres.
Invent. Math., 124(1-3):437–449, 1996.
Mil96. J. S. Milne. Elliptic curves. Lecture notes, www.jmilne.org, 1996.
Mor69. L. J. Mordell. Diophantine equations. Pure and Applied Mathematics, Vol. 30. Academic
Press, London, 1969.

124
Mov04. Hossein Movasati. Abelian integrals in holomorphic foliations. Rev. Mat. Iberoameri-
cana, 20(1):183–204, 2004.
Mov16. Hossein Movasati. Gauss-Manin connection in disguise: Calabi-Yau modular forms.
To appear in Surveys of Modern Mathematics, IP, Boston. Available online at
http://w3.impa.br/∼hossein/myarticles/GMCD-MQCY3.pdf, 170 pages, 2016.
Nag35. Trygve Nagell. Solution de quelques problèmes dans la théorie arithmétique des cubiques
planes du premier genre. Technical report, Skr. Norske Vid.-Akad., Oslo 1935, No.1, 1-25
, 1935.
Rie. B. Riemann. Ueber die anzahl der primzahlen unter einer gegebenen grösse. Monat. der
Königl. Preuss. akad. der Wissen. zu Berlin, 1859, 671-680, also, Gesmmelte math. Werke,
1892, 145-155.
Sel51. Ernst S. Selmer. The Diophantine equation ax3 + by3 + cz3 = 0. Acta Math., 85:203–362
(1 plate), 1951.
Ser89. Jean-Pierre Serre. Lectures on the Mordell-Weil theorem. Aspects of Mathematics, E15.
Friedr. Vieweg & Sohn, Braunschweig, 1989. Translated from the French and edited by
Martin Brown from notes by Michel Waldschmidt.
Sil92. Joseph H. Silverman. The arithmetic of elliptic curves, volume 106 of Graduate Texts in
Mathematics. Springer-Verlag, New York, 1992. Corrected reprint of the 1986 original.
ST92. Joseph H. Silverman and John Tate. Rational points on elliptic curves. Undergraduate
Texts in Mathematics. Springer-Verlag, New York, 1992.
Tun83. J. B. Tunnell. A classical Diophantine problem and modular forms of weight 3/2. Invent.
Math., 72(2):323–334, 1983.
Wal06. Michel Waldschmidt. Transcendence of periods: the state of the art. Pure Appl. Math. Q.,
2(2):435–463, 2006.
Wei49. André Weil. Numbers of solutions of equations in finite fields. Bull. Amer. Math. Soc.,
55:497–508, 1949.

125
Index

Dedekind eta function, 6 Modular form, 12

Point at infinity, 35
Elliptic curve, 35
Elliptic function, 11
Serre derivative, 24

Lattice of elliptic integrals, 36 Weierstrass form, 35

127

S-ar putea să vă placă și