Elliptic Curves, Modular Forms, and Their L-Functions
Elliptic Curves, Modular Forms, and Their L-Functions
Elliptic Curves,
Modular Forms,
and Their
L-functions
Álvaro Lozano-Robledo
Elliptic Curves,
Modular Forms,
and Their L-functions
Álvaro Lozano-Robledo
c 2011 by the American Mathematical Society. All rights reserved.
The American Mathematical Society retains all rights
except those granted to the United States Government.
Printed in the United States of America.
∞ The paper used in this book is acid-free and falls within the guidelines
established to ensure permanence and durability.
Visit the AMS home page at http://www.ams.org/
10 9 8 7 6 5 4 3 2 1 16 15 14 13 12 11
A mis padres, que me enseñaron todo lo importante,
a mi abuela, por su sonrisa que no aparece en fotografías,
a Marisa, por lograr que siempre me supere,
y a “Sally”, que nacerá pronto.
Contents
Preface xi
Chapter 1. Introduction 1
§1.1. Elliptic curves 1
§1.2. Modular forms 7
§1.3. L-functions 11
§1.4. Exercises 15
vii
viii Contents
§2.12. Exercises 69
Bibliography 189
Index 193
Preface
This book grew out of the lecture notes for a course on “Elliptic
Curves, Modular Forms and L-functions” that the author taught at
an undergraduate summer school as part of the 2009 Park City Mathe-
matics Institute. These notes are an introductory survey of the theory
of elliptic curves, modular forms and their L-functions, with an em-
phasis on examples rather than proofs. The main goal is to provide
the reader with a big picture of the surprising connections among
these three types of mathematical objects, which are seemingly so
distinct. In that vein, one of the themes of the book is to explain
the statement of the modularity theorem (Theorem 5.4.6), previously
known as the Taniyama-Shimura-Weil conjecture (Conjecture 5.4.5).
In order to underscore the importance of the modularity theorem, we
also discuss in some detail one of its most renowned consequences:
Fermat’s last theorem (Example 1.1.5 and Section 5.5).
It would be impossible to give the proofs of the main theorems
on elliptic curves and modular forms in one single course, and the
proofs would be outside the scope of the undergraduate curriculum.
However, the definitions, the statements of the main theorems and
their corollaries can be easily understood by students with some stan-
dard undergraduate background (calculus, linear algebra, elementary
number theory and a first course in abstract algebra). Proofs that are
accessible to a student are left to the reader and proposed as exercises
xi
xii Preface
(1) There are not that many books on these subjects at the
undergraduate level. However, Silverman and Tate’s book
[SiT92] is an excellent introduction to elliptic curves for
undergraduates. Washington’s book [Was08] is also acces-
sible for undergraduates and emphasizes the cryptography
applications of elliptic curves. Stein’s book [Ste08] also has
an interesting chapter on elliptic curves.
Álvaro Lozano-Robledo
Chapter 1
Introduction
Notation:
N = {1, 2, 3, . . .} is the set of natural numbers.
Z = {. . . , −3, −2, −1, 0, 1, 2, 3, . . .} is the ring of integers.
n : m, n ∈ Z, n = 0} is the field of rational numbers.
Q = {m
R is the field of real numbers.
C = {a + bi : a, b ∈ R, i2 = −1} is the field of complex numbers.
In this chapter, we introduce elliptic curves, modular forms and L-
functions through examples that motivate the definitions.
1
2 1. Introduction
The reader can provide a proof (see Exercise 1.4.3). For example,
the curve E : y 2 = x3 − 25x has a point (−4, 6) that corresponds to
the triangle ( 23 , 20 41 1681 62279
3 , 6 ). But E has other points, such as ( 144 , 1728 )
that corresponds to the triangle
1519 4920 3344161
, ,
492 1519 747348
which also has area equal to 5. See Figure 2.
Today, there are partial results toward the solution of the congru-
ent number problem, and strong results that rely heavily on famous
(and widely accepted) conjectures, but we do not have a full answer
yet. For instance, in 1975 (see [Ste75]), Stephens showed that the
Birch and Swinnerton-Dyer conjecture (which we will discuss in Sec-
tion 5.2) implies that any positive integer n ≡ 5, 6 or 7 mod 8 is a
1.1. Elliptic curves 5
and, if n is even,
n
#{(x, y, z) ∈ Z3 : = 4x2 + y 2 + 32z 2 }
2
1 n
= #{(x, y, z) ∈ Z3 : = 4x2 + y 2 + 8z 2 } .
2 2
Moreover, if the Birch and Swinnerton-Dyer conjecture is true, then,
conversely, these equalities imply that n is a congruent number.
S4 (n) = 8 d,
d|n, 4d
−1 −1
S6 (n) = 22ν+4 − 4 d2 ,
g d
d|g
d3 if n is odd,
S8 (n) = 16 · d|n
d|n d −2
3
d|g
3
d if n is even.
−1 −1
S6 (4) = 2·2+4
2 −4 · 1 = (28 − 4) · 1 = 252.
2
1 1
10 1. Introduction
The formulas for Sk (n) given above are derived using the theory of
modular forms, as follows. We define a formal power series Θ(q) by
∞
2
Θ(q) = qj
j=−∞
and, for k ≥ 2, consider the power series expansion of the kth power
of Θ:
⎛ ⎞k
∞
= ⎝ qj ⎠
2
(Θ(q))k
j=−∞
∞
∞
a21 a2k
= q ··· q = cn q n .
a1 =−∞ ak =−∞ n≥0
1.3. L-functions
An L-function is a function L(s), usually given as an infinite series of
the form
∞ ∞
an a2 a3
L(s) = an n−s = s
= a1 + s + s + · · ·
n=1 n=1
n 2 3
where g(s) is a function with g(1) finite, and φ is the Euler φ-function.
Therefore, there cannot be a finite number of primes of the form
p ≡ a mod N .
by
∞
χn (a)
L(s, χn ) = .
a=1
as
We are ready to write down the formula for S3 (n), due to Gauss,
Dirichlet and Shimura (there are also formulas for S5 (n), due to Eisen-
stein, Smith, Minkowski and Shimura, and a formula for S7 (n), also
due to Shimura). For simplicity, let us assume that n is odd and
square free (for the utmost generality, please check [Shi02]):
0 if n ≡ 7 mod 8,
S3 (n) = √
24 n
π L(1, χn ) otherwise.
The reader is encouraged to investigate this problem further by at-
tempting Exercises 1.4.6 and 1.4.7.
1.4. Exercises
Exercise 1.4.1. Use the divisibility properties of integers to show
that the only solutions to y 2 = x(x + 1)(x + 2) with x, y ∈ Z are
(0, 0), (−1, 0) and (−2, 0). (Hint: If a and b are relatively prime and
ab is a square, then a is a square and b is a square.)
Exercise 1.4.2. Find all the Pythagorean triples (a, b, c), i.e., a, b, c ∈
Z and a2 + b2 = c2 , such that b2 + c2 = d2 for some d ∈ Z. In other
words, find all the integers a, b, c, d such that (a, b, c) and (b, c, d)
are both Pythagorean triples. (Hint: You may assume that y 2 =
x(x + 1)(x + 2) has no rational points other than (0, 0), (−1, 0) and
(−2, 0).)
Exercise 1.4.3. Prove Proposition 1.1.3; i.e., show that f ((a, b, c)) is
a point in En , that g((x, y)) is a triangle in Cn and that f (g((x, y))) =
(x, y) and g(f ((a, b, c))) = (a, b, c).
has an Euler product; i.e., prove the following formal equality of series
∞
1 1
= .
n=1
n s
p prime
1 − p−s
Elliptic curves
(2.1) f (x1 , x2 , . . . , xr ) = 0
1The contents of this chapter are largely based on the article [Loz05], in Spanish.
17
18 2. Elliptic curves
2.2. Definition
Definition 2.2.1. An elliptic curve over Q is a smooth cubic projec-
tive curve E defined over Q with at least one rational point O ∈ E(Q)
that we call the origin.
(2.3) aX 3 + bX 2 Y + cXY 2 + dY 3
+eX 2 + f XY + gY 2 + hX + jY + k = 0
2.2. Definition 21
but with the understanding that in this new model we may have left
out some points of E at infinity (i.e., those points [X, Y, 0] satisfying
Eq. 2.2).
In general, one can find a change of coordinates that simplifies
Eq. 2.3 enormously:
From now on, we will often work with an elliptic curve in this form.
Notice that a curve E given by a Weierstrass equation y 2 = x3 +Ax+
B is non-singular if and only if 4A3 + 27B 2 = 0, and it has a unique
point at infinity, namely [0, 1, 0], which we shall call the origin O or
the point at infinity of E.
Sometimes we shall use a more general Weierstrass equation
y 2 + a1 xy + a3 y = x3 + a2 x2 + a4 x + a6
with ai ∈ Q (we will explain the funky choice of notation for the
coefficients later), but most of the time we will work with equations
of the form y 2 = x3 + Ax + B. It is easy to come up with a change
of variables from one form to the other (see Exercise 2.12.3).
22 2. Elliptic curves
X 3 + Y 3 = dZ 3
with O = [1, −1, 0]. The reader should verify that E is a smooth
curve. We wish to find a Weierstrass equation for E and, indeed, one
given by
can find a change of variables ψ : E → E
P+Q
y = − 23 (x − 5)
y 2 = x3 − 25x.
Plugging the first equation into the second one, we obtain an equation
4 185 100
x3 − x2 − x− = 0,
9 9 9
which factors as (x − 5)(x + 4)(9x + 5) = 0. The first two factors
are expected, since we already knew that P = (5, 0) and Q = (−4, 6)
are in L ∩ E. The third point of intersection must have x = − 59 ,
y = − 23 (x − 5) = 100 5 100
27 and, indeed, R = (− 9 , 27 ) is a point in
26 2. Elliptic curves
right triangle
1519 4920 3344161
(a, b, c) = , , .
492 1519 747348
Q = a1 P 1 + a2 P 2 + · · · + an P n
for some ai ∈ Z.
Example 2.4.4. Consider the elliptic curve E/Q given by the Weier-
strass equation
y 2 + y = x3 − 7x + 6.
The set of rational points E(Q) for this elliptic curve is infinite. For
instance, the following points are on the curve:
(1, 0), (2, 0), (0, −3), (−3, −1), (8, −22), (−2, −4), (3, −4),
(3, 3), (−1, −4), (1, −1), (0, 2), (2, −1), (−2, 3), (−1, 3),
1 13 25 91 26 28 7 17
, , ,− , − , , , ,....
4 8 9 27 9 27 9 27
At a first glance, it may seem very difficult to describe all the points
on E(Q), including those listed above, in a succinct manner. However,
the Mordell-Weil theorem tells us that there must be a finite set of
points that generate the whole group. Indeed, it can be proved that
2.4. The group structure on E(Q) 29
For a proof, see [Sil86], Ch. VIII, Corollary 7.2, or [Mil06], Ch.
II, Theorem 5.1.
Example 2.5.6. Let E/Q : y 2 = x3 − 2, so that A = 0 and B = −2.
The polynomial x3 − 2 does not have any rational roots, so E(Q)
does not contain any points of order 2. Also, 4A3 + 27B 2 = 27 · 4.
Thus, if (x(P ), y(P )) are the coordinates of a torsion point in E(Q),
then y(P ) is an integer and y(P )2 divides 27 · 4. This implies that
y(P ) = ±1, ±2, ±3, or ±6. In turn, this implies that x(P )3 = 3, 6,
11 or 38, respectively. However, x(P ) is an integer, and none of 3, 6,
11 or 38 are a perfect cube. Thus, E(Q)torsion is trivial (i.e., the only
torsion point is O).
Example 2.5.7. Let p ≥ 2 be a prime number and let us define a
curve Ep : y 2 = x3 + p2 . Since x3 + p2 = 0 does not have any rational
roots, Ep (Q) does not contain points of order 2. Let P be a torsion
point on Ep (Q). The list of all squares dividing 4A3 + 27B 2 = 27p4
is short, and by the Nagell-Lutz theorem the possible values for y(P )
are:
y = ±1, ±p, ±p2 , ±3p, ±3p2 , and ± 3.
Clearly, (0, ±p) ∈ Ep (Q) and one can show that those two points and
O are the only torsion points; see Exercise 2.12.8. Thus, the torsion
subgroup of Ep (Q) is isomorphic to Z/3Z for any prime p ≥ 2.
2.6. Elliptic curves over finite fields 35
ΔE = −16(4A3 + 27B 2 ).
f (x, y) = y 2 + a1 xy + a3 y − x3 − a2 x2 − a4 x − a6 ,
f (x, y) − f (x0 , y0 )
= λ1 (x − x0 )2 + λ2 (x − x0 )(y − y0 ) + λ3 (y − y0 )2 − (x − x0 )3
= ((y − y0 ) − α(x − x0 )) · ((y − y0 ) − β(x − x0 )) − (x − x0 )3
y − y0 = α(x − x0 ), y − y0 = β(x − x0 ).
of points (how many?). Thus, the number Nq := |E(F q )|, i.e., the
number of points on E over Fq , is finite. The following theorem
provides a bound for Nq . This result was conjectured by Emil Artin
(in his thesis) and was proved by Helmut Hasse in the 1930’s:
be
Theorem 2.6.11 (Hasse; [Sil86], Ch. V, Theorem 1.1). Let E
an elliptic curve defined over Fq . Then
√ √
q + 1 − 2 q < Nq < q + 1 + 2 q,
q )|.
where Nq = |E(F
Remark 2.6.13. Suppose that E/Q is an elliptic curve that has bad
reduction at a prime p. How many points does the singular curve
E have over Fp ? The reader can find the answer to this question in
Exercise 5.7.1.
does not divide 13, it follows that E(Q)[5] must be trivial. Similarly,
one can show that E(Q)[7] is trivial, and we conclude that E(Q)torsion
is trivial.
However, notice that P = (1, 2) ∈ E(Q) is a point on the curve.
Since we just proved that E does not have any points of finite order,
it follows that P must be a point of infinite order, and, hence, we have
shown that E has infinitely many rational points: ±P, ±2P, ±3P, . . ..
In fact, E(Q) ∼= Z and (1, 2) is a generator of its Mordell-Weil group.
1 H(2N · P )
h(P ) = lim .
2 N →∞ 4N
Note: here 2N ·P means multiplication in the curve, using the addition
law defined in Section 2.4, i.e., 2 · P = P + P , 22 · P = 2P + 2P , etc.
Example 2.7.2. Let E : y 2 = x3 + 877x, and let P be a generator
of E(Q). Here are some values of 12 · H(24N ·P ) :
N
1
· H(P ) = 47.8645312628 . . .
2
1 H(2 · P )
· = 47.7958126219 . . .
2 4
1 H(22 · P )
· = 47.9720107996 . . .
2 42
1 H(23 · P )
· = 47.9636902383 . . .
2 43
1 H(24 · P )
· = 47.9901607777 . . .
2 44
1 H(25 · P )
· = 47.9901600133 . . .
2 45
1 H(26 · P )
· = 47.9901569227 . . .
2 46
1 H(27 · P )
· = 47.9901419861 . . .
2 47
1 H(28 · P )
· = 47.9901807594 . . . .
2 48
The limit is in fact equal to h(P ) = 47.9901859939..., well below the
value |ΔE |1/2 = 207, 773.12....
For the proofs of these properties, see [Sil86], Ch. VIII, Thm.
9.3, or [Mil06], Ch. IV, Prop. 4.5 and Thm. 4.7.
As we mentioned at the beginning of this section, we can calculate
upper bounds on the rank of a given elliptic curve (see [Sil86], p. 235,
exercises 8.1, 8.2). Here is an example:
Theorem 2.7.4 ([Loz08], Prop. 1.1). Let E/Q be an elliptic curve
given by a Weierstrass equation of the form
E : y 2 = x3 + Ax2 + Bx, with A, B ∈ Z.
Let RE be the rank of E(Q). For an integer N ≥ 1, let ν(N ) be the
number of distinct positive prime divisors of N . Then
RE ≤ ν(A2 − 4B) + ν(B) − 1.
More generally, let E/Q be any elliptic curve with a non-trivial point
of 2-torsion and let a (resp. m) be the number of primes of additive
(resp. multiplicative) bad reduction of E/Q. Then
RE ≤ m + 2a − 1.
Example 2.7.5. Pierre de Fermat proved that n = 1 is not a con-
gruent number (see Example 1.1.2) by showing that x4 + y 4 = z 2 has
no rational solutions. As an application of the previous theorem, let
us show that the curve
E1 : y 2 = x3 − x = x(x − 1)(x + 1)
only has the trivial solutions (0, 0), (±1, 0), which are torsion points
of order 2. Indeed, the minimal discriminant of E1 is ΔE1 = 64.
Therefore p = 2 is the unique prime of bad reduction. Moreover, the
reader can check that the reduction at p = 2 is multiplicative. Now
thanks to Theorem 2.7.4 we conclude that RE1 = 0 and E1 only has
torsion points. Finally, using Proposition 2.6.15 or Theorem 2.5.5,
we can show that the only torsion points are the three trivial points
named above.
46 2. Elliptic curves
part of E(Q), then the determinant of H({Pi }ri=1 ) is called the elliptic
regulator of E/Q.
(2) If det(H) = 0, then the points {Pi }ri=1 are linearly indepen-
dent and the rank of E(Q) is ≥ r.
The previous example points to the fact that there may be a ho-
momorphism between points on E(Q) and triples (a, b, c) of rational
numbers modulo squares, or square-free parts of rational numbers;
formally, we are talking about Q× /(Q× )2 × Q× /(Q× )2 × Q× /(Q× )2 .
Here, the group Q× /(Q× )2 is the multiplicative group of non-zero
rational numbers, with the extra relation that two non-zero rational
numbers are equivalent if their square-free parts are equal (or, equiv-
alently, if their quotient is a perfect square). For instance, 3 and 12
2 2 25
represent the same element of Q× /(Q× )2 because 12 25 = 3 · 5 . The
following theorem constructs such a homomorphism. Here we have
adapted the proof that appears in [Was08], Theorem 8.14.
52 2. Elliptic curves
2ab 2Ab 1
(2.10) 2
−B− = − ,
c c c2
2ab b3 Ab
(2.11) + 3− = 0.
c2 c c
e3 − e1 = δ1 X 2 − δ1 δ2 Z 2 ,
C(δ1 , δ2 ) :
e3 − e2 = δ2 Y 2 − δ1 δ2 Z 2 ,
2.10. Homogeneous spaces 61
e1 − e2 = δ2 Y 2 − δ1 X 2 ,
C(δ1 , δ2 ) :
e3 − e2 = δ2 Y 2 − δ1 δ2 Z 2 .
The space C(δ1 , δ2 ) is the intersection of two conics, and it may have
rational points or not. If (δ1 , δ2 , δ3 ) is in the image of δ, however,
then the space C(δ1 , δ2 ) must have a rational point; i.e., there are
X, Y, Z ∈ Q that satisfy the equations of C(δ1 , δ2 ). Moreover, if
X0 , Y0 , Z0 ∈ Q are the coordinates of a point in C(δ1 , δ2 ), then
P = (e1 + δ1 X02 , δ1 δ2 X0 Y0 Z0 ).
δ2 Y 2 − δ1 X 2 = 34,
C(δ1 , δ2 ) :
δ2 Y 2 − δ1 δ2 Z 2 = 68
(1) ((δ1 , δ2 , δ3 ) = (1, 1, 1)). The point at infinity (i.e., the origin)
is sent to (1, 1, 1) via δ, i.e., δ(O) = (1, 1, 1).
(2) (δ1 < 0 and δ2 < 0). The equation δ2 Y 2 − δ1 δ2 Z 2 = 68
cannot have solutions (in Q or R) because the left-hand side
is always negative for any X, Z ∈ Q.
(3) (δ1 > 0 and δ2 < 0). The equation δ2 Y 2 − δ1 X 2 = 34
cannot have solutions (in Q or R), because the left-hand
side is always negative.
(4) (δ1 = −1, δ2 = 34). The space C(−1, 34) has a rational
point (X, Y, Z) = (0, 1, 1), which maps to T1 = (0, 0) on
E(Q) via Eq. (2.12).
(5) (δ1 = −34, δ2 = 2). The space C(−34, 2) has the rational
point (X, Y, Z) = (1, 0, 1), which maps to T2 = (−34, 0) on
E(Q) via Eq. (2.12).
(6) (δ1 = 34, δ2 = 17). If δ(T1 ) = δ((0, 0)) equals (−1, 34, −34),
and δ(T2 ) = (−34, 2, −17), then
δ(T1 + T2 ) = δ(T1 ) · δ(T2 ) = (−1, 34, −34) · (−34, 2, −17) = (34, 17, 2).
Thus, the space C(34, 17) must have a point that maps
back to T1 + T2 = (34, 0). Indeed, C(34, 17) has a point
(X, Y, Z) = (1, 2, 0) that maps to (34, 0) via Eq. (2.12).
2.10. Homogeneous spaces 63
(7) (δ1 = −1, δ2 = 2). The space C(−1, 2) has a rational point
(X, Y, Z) = (4, 3, 5), which maps to P = (−16, −120) on
E(Q) via Eq. (2.12). P is a point of infinite order.
(8) ((δ1 , δ2 ) = (1, 17), (34, 1), or (−34, 34)). These are the pairs
that correspond to (−1, 2) · γ, with γ = (−1, 34), (−34, 2)
or (34, 17). Therefore, the corresponding spaces C(δ1 , δ2 )
must have rational points that map to P + T1 , P + T2 and
P + T1 + T2 , respectively.
(9) (δ1 = −2, δ2 = 2). The space C(−2, 2) has a rational point
(X, Y, Z) = (1, 4, 3), which maps to Q = (−2, −48) on E(Q)
via Eq. (2.12). Q is a point of infinite order.
(10) ((δ1 , δ2 ) = (2, 17), (17, 1), or (−17, 34)). These are the pairs
that correspond to (−2, 2) · γ, with γ = (−1, 34), (−34, 2)
or (34, 17). Therefore, the corresponding spaces C(δ1 , δ2 )
must have rational points that map to Q + T1 , Q + T2 and
Q + T1 + T2 , respectively.
(11) ((δ1 , δ2 ) = (2, 1), and (−2, 34), (−17, 2), or (17, 17)). Since
(−1, 2) and (−2, 2) correspond to P and Q, respectively,
then (−1, 2) · (−2, 2) = (2, 1) corresponds to P + Q. The
other pairs correspond to (−2, 2) · γ, with γ = (−1, 34),
(−34, 2) or (34, 17). Therefore, the corresponding spaces
C(δ1 , δ2 ) must have rational points that map to P + Q + T1 ,
P + Q + T2 and P + Q + T1 + T2 , respectively.
(12) (δ1 = 1, δ2 = 2). The space C(1, 2) does not have ratio-
nal points (see Exercise 2.12.21). In fact, it does not have
solutions in Q2 , the field of 2-adic numbers.
(13) ((δ1 , δ2 ) = (2, 2), (17, 2), (34, 2), (−1, 1), (−2, 1), (−17, 1),
(−34, 1), (−1, 17), (−2, 17), (−17, 17), (−34, 17), (1, 34),
(2, 34), (17, 34), (34, 34)). The corresponding spaces C(δ1 , δ2 )
do not have rational points. For instance, suppose C(2, 2)
had a point. Then (2, 2, 1) would be in the image of δ.
Since (2, 1, 2) is in the image of δ (we already saw above
that C(2, 1) has a point), then (2, 1, 2) · (2, 2, 1) = (1, 2, 2)
would also be in the image of δ, but we just saw (in the pre-
vious item) that (1, 2, 2) is not in the image of δ. Therefore,
64 2. Elliptic curves
times (−1, 82, −82), (−82, 2, −41), (82, 41, 2), which do have
points by a previous item in this list.
How about all the other possible pairs (δ1 , δ2 )? Consider (−1, 2, −2)
and its homogeneous space:
2Y 2 + X 2 = 82,
C(−1, 2) :
2Y 2 + 2Z 2 = 164.
Let us show that there are solutions to C(−1, 2) over R, Q2 and Q41 :
√ √
• (Over R). The point (0, 41, 41) is a point on C(−1, 2)
defined over R.
• (Over Q41 ). Let Y0 = 1 and put f (X) = X 2 − 80, g(Z) =
Z 2 − 81. By Hensel’s Lemma (see Appendix D.1 and Corol-
lary D.1.2), it suffices to show that there are α0 , β0 ∈ F41
such that
f (α0 ) = g(β0 ) ≡ 0 mod 41 and f (α0 ), g (β0 ) ≡ 0 mod 41.
The reader can check that the congruences α0 ≡ 11 mod 41
and β0 ≡ 9 mod 41 work. Thus, there are α, β ∈ Q41 such
that f (α) = 0 = g(β). Hence, (X0 , Y0 , Z0 ) = (α, 1, β) is a
point on C(−1, 2) defined over Q41 , as desired.
• (Over Q2 ). Let X0 = 0 and put f (Y ) = Y 2 −41. Let α0 = 1.
Then f (α0 ) = −40, f (α0 ) = 82 and
3 = ν2 (−40) > ν2 (822 ) = ν2 (22 · 412 ) = 2.
Thus, by Hensel’s Lemma (Theorem D.1.1; see also Ex-
ample D.1.4), there is α ∈ Q2 such that f (α) = 0, or
α2 = 41. Hence, the point (X0 , Y0 , Z0 ) = (0, α, α) is a point
on C(−1, 2) defined over Q2 , as desired.
One can also show that, in fact, C(−1, 2) has a point over Qp for
all p ≥ 2. Therefore, we cannot deduce any contradictions working
locally about whether C(−1, 2) has a point over Q. A computer search
does not yield any Q-points on C(−1, 2). Therefore, our method
breaks at this point, and we cannot determine whether there is a
point on E(Q) that comes from C(−1, 2).
It turns out that C(−1, 2) does not have rational points (but this
is difficult to show). This type of space, a space that has solutions
66 2. Elliptic curves
everywhere locally (Qp , R) but not globally (Q) is the main obstacle
for the descent method to fully work.
Notice that, indeed, the elements of Sel2 (E/Q) listed above form
a subgroup of Γ × Γ ⊂ (Q× /(Q× )2 )2 . Since all the elements of
Sel2 (E/Q) have rational points, we conclude that Sel2 (E/Q) equals
E(Q)/2E(Q) and
X2 (E/Q) = Sel2 (E/Q)/(E(Q)/2E(Q)) = {C(1, 1)},
i.e., X2 is the trivial subgroup in this case.
2.12. Exercises
Exercise 2.12.1. Let f (x) = a0 xn + a1 xn−1 + . . . + an , with ai ∈ Z.
Prove that if x = pq ∈ Q, with gcd(p, q) = 1, is a solution of f (x) = 0,
then an is divisible by p and a0 is divisible by q.
Exercise 2.12.2. Let C be the conic defined by x2 − 2y 2 = 1.
(1) Find all the rational points on C. (Hint: the point O = (1, 0)
belongs to C. Let L(t) be the line that goes through O and
has slope t. Since C is a quadratic and L(t) ∩ C contains
at least one rational point, there must be a second point of
intersection Q. Find the coordinates of Q in terms of t.)
√ √ √
(2) Let α = 1 + 2. Calculate α2 = a + b 2 and α4 = c + d 2
and verify that (a, b) and (c, d) are integral points
√ on C :
x2 − 2y 2 = 1. (Note: in fact, if α2n = e + f 2, then
(e, f ) ∈ C and the coefficients of α2n+1 are a solution of
x2 − 2y 2 = −1.)
(3) (This problem is only for those who already know√about
continued fractions.) Find the continued fraction of 2 and
find the first 6 convergents. Use the convergents to find
three distinct (positive) integral solutions of x2 − 2y 2 = 1,
other than (1, 0). (Note: the reader should remind herself
or himself how to find the continued fraction and conver-
gents by hand, then check his or her answer using Sage; see
Appendix A.4.)
Exercise 2.12.3. Let C/Q be an affine curve.
(1) Suppose that C/Q is given by an equation of the form
(2.13) C : xy 2 + ax2 + bxy + cy 2 + dx + ey + f = 0.
Find an invertible change of variables that takes the equa-
tion of C onto one of the form xy 2 +gx2 +hxy+jx+ky+l = 0.
(Hint: consider a change of variables X = x + λ, Y = y).
(2) Suppose that C /Q is given by an equation of the form
(2.14) C : xy 2 + ax2 + bxy + cx + dy + e = 0.
Find an invertible change of variables that takes the equa-
tion of C onto one of the form y 2 + αxy + βy = x3 + γx2 +
70 2. Elliptic curves
(2) Let y 2 = x(x + 5)(x + 10) and P = (−9, 6). Find nP for
n = 1, . . . , 12. Compare the x-coordinates of nP with the
list given at the end of Example 1.1.1, and write down the
next three numbers that belong in the list.
Exercise 2.12.12. Prove parts (1) and (3) of Theorem 2.8.5. (Hint:
use Definition 2.8.4 and Proposition 2.7.3.)
Modular curves
77
78 3. Modular curves
−1 1 2 3
−1
is a lattice, and every lattice is given in this way. The lattice gener-
ated by w1 , w2 ∈ C is denoted by w1 , w2 . We will insist on a positive
orientation of our basis; i.e., we require w1 /w2 to have positive imag-
inary part. In other words, w1 /w2 belongs to the upper half complex
plane H, where
H = {a + bi ∈ C : b > 0}.
4i
3i
2i
−1 1 2 3
−i
of the square F is identified with the opposite side modulo the lattice
L. Thus, C/Z[i] is indeed a torus when considered as a surface.
Proposition 3.1.6. Let L = w1 , w2 and L = w1 , w2 be lattices
with oriented bases (i.e., w1 /w2 and w1 /w2 ∈ H).
(1) L = L if and only if there is a matrix M ∈ SL(2, Z) such
1
that w1
w
=M w w2 .
2
The proofs of (most of) the proposition and corollary are left as
exercises (Exercises 3.7.2 and 3.7.3). We will, however, need to rely
on the following fact from complex analysis without giving a proof:
if a map ψ : C/L → C/L is an analytic isomorphism, then there
is α ∈ C× such that L = αL and ψ(z mod L) = αz mod L . See
[Sil86], Ch. VI, Theorem 4.1 for more details.
Remark 3.1.8. Let L = w1 , w2 and L = w1 , w2 be two arbitrary
lattices. Then, the map ψ : C/L → C/L , given by
ψ(λw1 + μw2 mod L) = λw1 + μw2 mod L
for any 0 ≤ λ, μ < 1, is a bijection of sets (indeed, ψ is a bijection
between the fundamental domains of C/L and C/L ). In fact, ψ is
also an isomorphism of abelian groups. However, in general, this map
is not analytic.
Example 3.1.9. Let L = Z[i] = i, 1 with w1 = i and w2 = 1. Let
3 5
M= ∈ SL(2, Z).
1 2
3.1. Elliptic curves over C 81
1 i 3i+5
Put w 1
w2
=M w w2 = M 1 = i+2 . Then Z[i] = i, 1 = 5 + 3i, 2 +
i. Indeed, it is clear that 5 + 3i, 2 + i ∈ i, 1. Moreover,
therefore i, 1 ⊆ 5 + 3i, 2 + i, and so they are equal lattices. Now,
define L = 5i + 13
5 , 1 = 2+i 5 + 3i, 2 + i = 2+i Z[i]. By Proposition
1 1
Here, we will not worry too much about convergence, but the
worried reader may be relieved to know that G2k (L) is absolutely
convergent for k > 1 and ℘(z, L) converges uniformly on every com-
pact subset of C − L (the worried reader can find a proof of the
3.2. Functions on lattices and elliptic functions 83
See Exercise 3.7.4 for a proof of the first part of the theorem (part
(2) is shown in [Sil86], Theorem VI.3.5). Theorem 3.2.4 shows that
there is a map:
℘ (z, L)
(3.1) Φ : C/L → EL (C), z mod L → ℘(z, L), .
2
It turns out that the map Φ has all the “nice” properties that one
would hope for: it is a complex analytic isomorphism of abelian
groups. Moreover, if E/C : y 2 = x3 + Ax + B is an elliptic curve,
then there is a lattice L ⊂ C such that Φ : C/L ∼
= E(C). This result
is usually called the uniformization theorem:
Y (1) = H/Γ(1)
{z = a + bi ∈ C : b > 0}
= .
{z ∼ z if and only if z = M z for some M ∈ SL(2, Z)}
−aτ − b aτ + b
(−M )τ = = = M τ.
−cτ − d cτ + d
topology of Y (1) and X(1) (for instance, see [Sil94], Ch. 1, §2 or see
[DS05], Sections 2.1 and 2.2 for a more generalized approach). The
formal construction of this point at infinity is the following.
We define an extended upper half plane by H∗ := H ∪ P1 (Q), i.e.,
the union of H and a copy of a projective line over Q. The points
in the projective line of the form {[s, 1] : s ∈ Q} are simply the
rational points of the real axis in the complex plane (so [s, 1] stands
for s + 0 · i ∈ Q ⊂ R). The remaining point of P1 (Q) is the point at
infinity ∞ := [1, 0]. If the reader needs to review the basics about the
projective line, see Appendix C.1.
We also extend the action of Γ(1) to all of H∗ as follows. Let
M = (a, b; c, d) ∈ SL(2, Z) and let [s, t] ∈ P1 (Q). Then we define
X(1) = H∗ /Γ(1)
{z = a + bi ∈ C : b > 0} ∪ {s ∈ Q} ∪ {∞}
= .
{z ∼ z if and only if z = M z for some M ∈ SL(2, Z)}
Example 3.6.1. Let p ≥ 2 and let X0 (p) = H∗ /Γ0 (p). Then X0 (p)
has exactly two cusps. The points 0 = [0, 1] and ∞ = [1, 0] are
inequivalent in X0 (p) and are representatives of the two non-trivial
cusps. See Exercise 3.7.10.
Remark 3.6.4. One aspect of modular curves that is not at all ob-
vious is the fact that modular curves have algebraic models; i.e., if
Γ is a congruence subgroup, then H∗ /Γ is a compact Riemann sur-
face and it has a model as a projective algebraic curve over C, given
by polynomial equations. The modular curves for Γ0 (N ) have the
surprising property that they have a canonical model defined over
Q. The reason is that the modular j-invariant function j(z) (see Ex-
ample 4.1.10) and the function j(N z) satisfy an algebraic equation
FN (j(z), j(N z)) = 0, with FN (x, y) ∈ Q[x, y], which gives an alge-
braic model for X0 (N ). However, this is typically a highly singular
model, which can be transformed into a non-singular model for the
modular curve. For instance,
(1) The curve X0 (11) = H∗ /Γ0 (11) has a model y 2 + y = x3 −
x2 − 10x − 20 (notice that it is an elliptic curve!).
(2) The curve X1 (11) = H∗ /Γ1 (11) has a model y 2 +y = x3 −x2 .
(3) The curve X0 (14) has a model y 2 + xy + y = x3 + 4x − 6.
(4) The curve X1 (14) has a model y 2 + xy + y = x3 − x.
(5) The curve X1 (13) has a model y 2 + (x3 − x2 − 1)y = x2 − x.
This is not an elliptic curve (it has genus 2, not 1). The
examples above (1)-(4) are nice but the model of a modular
curve will be often much more complicated than a cubic.
12 otherwise,
(p − 1)(p − 11)
genus(X1 (p)) = 1 + , and
24
(p2 − 1)(p − 6)
genus(X(p)) = 1 + ,
24
where [x] is the greatest integer ≤ x.
The genus formulas for X(p), X0 (p) and X1 (p) are consequences
of the Hurwitz and Riemann-Hurwitz genus formulas (see Exercises
3.1.4, 3.1.5 and 3.1.6 of [DS05] or see Chapter 1 of [Shi73] for proofs).
The list of all modular curves X0 (N ) with genus 0, 1 or 2 can be found
in [Maz72]. For more general genus formulas see [DS05], Section 3.1.
3.7. Exercises
Exercise 3.7.1. Let a, b, c, d ∈ R, τ ∈ C and τ ∈
/ R. Show that:
(1) The imaginary part of τ = aτ +b
cτ +d is Im(τ ) =
(ad−bc) Im(τ )
|cτ +d|2 .
a b
(2) If M = ∈ SL(2, Z) and τ ∈ H, then M τ ∈ H.
c d
Exercise 3.7.2. In this exercise we study the relationships between
different bases of a lattice.
(1) Let L = i, 1 be the lattice of Gaussian integers Z[i]. Let
a, b, c, d be integers such that ad − bc = 1. Show that the
lattice L generated by w1 = ai + b and w2 = ci + d is also
Z[i].
(2) More generally, let L be a lattice generated by w1 and w2 ∈
C with w1 /w2 ∈ H. Let
a b
M=
c d
be a matrix in SL(2, Z), i.e., a, b, c, d ∈ Z and ad − bc = 1.
1
Let w 1
w2
= M w w2 , where the operation here is the usual
matrix multiplication of vectors, i.e., w1 = aw1 + bw2 and
3.7. Exercises 95
w2 = cw1 + dw2 . Show that w1 /w2 ∈ H and the lattice
generated by w1 and w2 is also L. (Hint: do Exercise 3.7.1.
Also, notice that M is an invertible matrix.)
(3) Conversely, suppose that L = w1 , w2 = w1 , w2 , for some
wi , wi ∈ C, such that w1 /w2 , w1 /w2 ∈ H. Show that there
1
is a matrix M ∈ SL(2, Z) such that w 1
w
=M ww2 .
2
Exercise 3.7.6. The goal of this problem is to show that the relation
that appears in Definition 3.3.1 is indeed an equivalence relation. Let
M = (a, b; c, d) ∈ SL(2, Z), τ, τ ∈ H and define M τ = aτ +b
cτ +d . We say
that τ ∼ τ if there is a matrix M ∈ SL(2, Z) such that τ = M τ .
Show that:
(1) (Reflexive) τ ∼ τ for all τ ∈ H;
(2) (Symmetric) if τ ∼ τ , then τ ∼ τ for all τ, τ ∈ H;
(3) (Transitive) if τ ∼ τ and τ ∼ τ , then τ ∼ τ for all
τ, τ, τ ∈ H.
Exercise 3.7.10. Let p be a prime and let X0 (p) = H∗ /Γ0 (p). Show
that X0 (p) has exactly two cusps. In particular, show that if [s, t] ∈
3.7. Exercises 97
Modular forms
Remark 4.1.2. Let Γ = SL(2, Z) and let X(1) = H∗ / SL(2, Z). Then
a modular function for SL(2, Z) is a function f : H∗ → C ∪ {∞} that
99
100 4. Modular forms
Remark 4.1.4. Exercise 4.5.5 shows that condition (2) in the defi-
nition of weakly modular function (resp. modular form below); i.e.,
Definition 4.1.3 above (resp. 4.2.1 below), is equivalent to saying
that f (z)(dz)k is a differential k-form, invariant under the action of
SL(2, Z) (resp. congruence subgroup Γ).
Then G2k (z) is a modular form of weight 2k for SL(2, Z), and the
value of G2k at the cusp ∞ of X(1) is equal to 2ζ(2k), where ζ(s) is
the Riemann zeta function.
For a proof of this fact, see [Kob93], Ch. III, Prop. 6. The
values of ζ(2k) can be computed in terms of Bernoulli numbers:
∞
1 (2πi)2k B2k
ζ(2k) = = − for all k ≥ 1.
n=1
n2k 2(2k)!
Next, we list a number of properties satisfied by modular forms.
Proposition 4.1.8. Let k, k ≥ 2 be integers.
(1) Suppose that f and g are modular forms of weight k for
SL(2, Z). Then, for all λ, μ ∈ C, the function λf (z) + μg(z)
is also a modular form of weight k for SL(2, Z). Therefore,
the set of all modular forms of weight k for SL(2, Z) is a
vector space over C.
4.1. Modular forms for the modular group 103
Example 4.1.10. Let g2 (z) = −15G4 (z), g3 (z) = −35G6 (z), and
define
Δ(z) = −16 4(g2 (z))3 + 27(g3 (z))2 .
Then Δ is a modular form of weight 12 for SL(2, Z). The modular
form Δ is usually called the modular discriminant. Δ(z) has a simple
zero at ∞ and no other zeros. The function
(4g2 (z))3
j(z) = 1728
Δ(z)
is a modular function of weight 0 (as in Definition 4.1.1) but it is not
a modular form. j(z) is analytic on H but it is not analytic at ∞
because Δ(z) has a zero at ∞ but g2 (z) does not vanish at ∞, so j(z)
has a pole. The function j(z) is called the modular j-invariant.
Example 4.1.11. The modular forms f1 (z) = E10 (z) and f2 (z) =
E4 (z) · E6 (z) are both in the space M10 (SL(2, Z)). A priori, they
are distinct modular forms. However, if we knew the dimension of
M10 (SL(2, Z)) and the dimension was 1, then there should be a linear
relationship between both f1 and f2 . Thus, we need to know the
dimensions of spaces of modular forms!
For a proof, see [DS05], Theorem 3.5.2; [Ser77], Ch. VII, The-
orem 4; or [Kob93], Ch. III, Proposition 9.
Example 4.1.13. Let f1 (z) = E10 (z) and f2 (z) = E4 (z) · E6 (z),
which are both in the space M10 (SL(2, Z)). By Theorem 4.1.12, the
dimension of M10 as a C-vector space is 1. Therefore, there exists
λ ∈ C such that f1 (z) = λf2 (z). However, both f1 (z) and f2 (Z)
are normalized modular forms, meaning that their first non-zero co-
efficient of their q-expansions equals 1. Hence, comparing their q-
expansions, we conclude that λ = 1 and E10 (z) = E4 (z)E6 (z).
The equality we just deduced, together with the q-expansion of
the Eisenstein series (Prop. 4.1.7), can be rephrased as:
20 8 12
1− σ9 (n)q n = (1 − σ3 (k)q k )(1 − σ5 (h)q h ).
B10 B4 B6
n≥1 k≥1 h≥1
Using the fact that S12 (SL(2, Z)) is 1-dimensional (by Theorem 4.1.12),
one can show the following surprising congruence:
τ (n) ≡ σ11 (n) ≡ d11 mod 691 for all n ≥ 1.
0<d|n
∞
fM (z) = cn (qN )n .
n=−m
Notice also that X(Γ) only has a finite number of cusps, say
s1 , s2 , . . . , sn , so one only needs to check the condition in part (a) of
Defn. 4.2.1 for a finite number of matrices M1 , M2 , . . . , Mn such that
Mi sends si to ∞.
See Exercise 4.5.8 for a proof of the invariance under Γ(N ) and
Γ1 (N ).
Remark 4.2.9. The Eisenstein series are very useful because most
of the spaces we are discussing in this book have a basis formed by
Eisenstein series, and we can calculate their q-expansions. For pre-
cise statements see [Miy06], Chapter 7, or [Ste07], Chapter 5 (in
particular, see Section 5.3).
f (q) = q − 2q 2 − q 3 + 2q 4 + q 5 + 2q 6 − 2q 7 − 2q 9 − 2q 10 + O(q 11 )
12 36 48 84 72 144 6
g(q) = 1 + q + q 2 + q 3 + q 4 + q 5 + q + O(q 7 ),
5 5 5 5 5 5
where q = e2πiz . Thus, we deduce that S2 (Γ0 (11)) is 1-dimensional,
generated by f (q).
f (q) = q + q 3 − 2q 4 − q 7 − 2q 9 + 3q 11 − 2q 12 − 4q 13 + O(q 16 )
g(q) = q 2 + 2q 3 − 2q 4 + q 5 − 3q 6 − 4q 9 − 2q 10 + 4q 11 + O(q 12 )
2 8 14 16
h(q) = 1 + q + 2q 2 + q 3 + q 4 + 4q 5 + 8q 6 + q 7 + O(q 8 ),
3 3 3 3
where, once again, q = e2πiz . Thus, we deduce that S2 (Γ0 (37)) is
2-dimensional, generated by f (q) and g(q).
110 4. Modular forms
wN : Sk (Γ0 (N )) → Sk (Γ0 (N ))
defined by
√
−k −1
(wN (f ))(z) = i ·k
Nz ·f .
Nz
Therefore, wN 2
= Id and the eigenvalues of wN are +1 or −1. Thus,
Sk (Γ0 (N )) can be expressed as the direct sum of the eigenspace cor-
responding to +1 plus the eigenspace corresponding to −1; i.e., if we
define spaces
Vm (f ) = an q mn and Um (f ) = an q n/m .
n≥0 n≡0 mod m
It is a crucial fact in the theory that one can find a basis for
Sknew (Γ1 (N ))
such that every element of the basis is an eigenform.
Recall that we defined spaces of new forms and old forms (Definitions
4.2.6 and 4.3.2) such that
Sk (Γ(N )) = Sknew (Γ(N )) ⊕ Skold (Γ(N )) = Skold (Γ(N ))⊥ ⊕ Skold (Γ(N ))
where new and old forms are orthogonal with respect to the Petersson
inner product. We define Sknew (Γ1 (N )) = Sk (Γ1 (N )) ∩ Sknew (Γ(N )),
and define similarly the old space for Γ1 (N ).
1 if i = j;
fi , fj = wN (fi ) = ±fi and Tn fi = λn,i fi ,
0 if i = j
where ·, · is the Petersson inner product for all n ≥ 1 and all 1 ≤
i ≤ d, for some eigenvalues λn,i ∈ C.
4.5. Exercises
Exercise 4.5.1. The goal of this problem is to show that the modu-
larity condition (2) in Definition 4.1.3 works “as one would hope” un-
der matrix multiplication. Let M = (a, b; c, d) and M = (a , b ; c , d )
be matrices in a congruence subgroup Γ ≤ SL(2, Z), and let M M =
(a , b ; c , d ) be their product. Let k ≥ 1 and let f (z) be a function.
Show that:
a z+b a (M z)+b
(1) (M M )z = M (M z), i.e., c z+d = c (M z)+d , where M z =
az+b
cz+d ;
(2)
a (M z) + b
(cz + d)−k (c (M z) + d )−k f
c (M z) + d
a z + b
= (c z + d )−k f .
c z + d
Exercise 4.5.2. Let f (z) be a weakly modular function of weight k
for SL(2, Z) with k odd. Show that f (z) = 0 for all z ∈ H. (Hint:
show that f (z) = −f (z) for all z ∈ H.)
where ζ(s) = ∞ 1
n=1 ns is the Riemann zeta function. (Hint:
you may assume that the convergence is uniform [can you
prove this?], and so the limit can be brought inside the sum-
mation.)
Exercise 4.5.7. Let Δ(z) be the modular discriminant form, and let
E2k (z) be the normalized Eisenstein series of weight 2k for SL(2, Z).
(1) Show that M = M12 (SL(2, Z)) is a 2-dimensional vector
space, and {Δ(z)} is a basis of the cusp forms S12 (SL(2, Z)).
(2) Show that E12 and E62 belong to M , and
(2π)−12 · 26 · 35 · 72
E12 − E62 = λΔ, where λ = .
691
(3) Use the q-expansions of E6 , E12 and Δ to write an expression
for τ (n) in terms of σ5 (n) and σ11 (n).
(4) Show that τ (n) ≡ σ11 (n) ≡ 0<d|n d11 mod 691 for all n ≥
1.
Sk (Γ0 (N )).
(3) Show that wN (wN (f )) = f for all f ∈ Sk (Γ0 (N )). Hence
wN (wN ) = Id, the eigenvalues of wN are all ±1, and Sk (Γ0 (N ))
factors into the direct sum of eigenspaces
Sk (Γ0 (N )) = Sk+ (Γ0 (N )) ⊕ Sk− (Γ0 (N )).
Exercise 4.5.12. Let N ≥ 1, δ ∈ Z and let M = (a, b; c, d) ∈ Γ0 (N )
with δ ≡ d mod N .
(1) Show that, for any f ∈ Mk (Γ1 (N )), the modular form δf
is also modular of weight k under the action of Γ1 (N ) (where
δ is as in Definition 4.4.3).
(2) Let M = (a , b ; c , d ) ∈ Γ0 (N ) be another matrix with
δ ≡ d ≡ d mod N , and let f ∈ Mk (Γ1 (N )). Show that
(cz + d)−k f (M z) = (c z + d )−k f (M z)
4.5. Exercises 121
Show that
(1) If (n, m) = 1, then σn σm = σnm .
(2) If p is a prime with gcd(N, p) = 1, then
σp σpr = σpr+1 + χ(p)pk−1 σpr−1 .
Chapter 5
L-functions
y 2 + a1 xy + a3 y = x3 + a2 x2 + a4 x + a6
123
124 5. L-functions
where δp is a technical invariant (see [Sil94], Ch. IV, §10; the invari-
ant δp describes whether there is wild ramification in the action of
the inertia group at p of Gal(Q/Q) on the Tate module Tp (E)).
where the product is over all primes and the exponents fp are defined
as above.
0.8
0.6
0.4
0.2
1 2 3
E0 : y 2 + y = x3 − x2 − 10x − 20, E 1 : y 2 + y = x3 − x
E2 : y 2 + y = x3 + x2 − 2x, E3 : y 2 + y = x3 − 7x + 6.
0.6
0.4
0.2
1 2 3
−0.2
0.6
0.4
0.2
1 2 3
E/Q : y 2 = x3 − 1156x
RE 2, P = (−16, 120), Q = (−2, 48)
|X| 1
ΩE 0.8993583214 . . .
Reg(E/Q) det H({P, Q}) = 7.0996751824 . . .
E(Q)torsion Z/2Z × Z/2Z ∼
= (0, 0), (34, 0)
p≥2 cp c2 · c17 = 4 · 4
We can also calculate the value L(E, 1) and the values of the deriva-
tives L (E, 1) and L (E, 1); i.e., we can approximate numerically
these values. For instance, one can use Sage (see Appendix A.3).
For a technical description of the algorithms involved, see [Dok04].
Once we have calculated these values, we can write the first few terms
132 5. L-functions
E/Q : y 2 = x3 − 6724x
RE 0
|X| 4
ΩE 0.5791156343 . . .
Reg(E/Q) det H({}) = 1
E(Q)torsion Z/2Z × Z/2Z ∼
= (0, 0), (81, 0)
p≥2 cp c2 · c41 = 4 · 4
Thus, the first part of the BSD conjecture is the statement that
the analytic rank equals the (algebraic) free rank of the Mordell-Weil
group E(Q).
Example 5.2.7. Let E/Q : y 2 = x3 −1572 x. Recall that Proposition
1.1.3 says that the rational points on E/Q with y = 0 give right
triangles of area 157, so if we find a single non-trivial point on E we
prove that n = 157 is a congruent number (as defined in Example
1.1.2).
Comparing values of Λ(s) and Λ(2 − s), we calculate the root
number w = w(E/Q) = −1. Thus, the parity conjecture suggests
that E(Q) has odd rank, therefore ≥ 1, and so E(Q) must be infinite.
However, a computer search only yields the trivial 2-torsion points
(0, 0), (157, 0) and (−157, 0). We can calculate values of L(E, s) and
its derivatives at s = 1 and write down an approximate Taylor ex-
pansion:
L(E, s) ≈ (11.4259445007) · (s − 1) − (49.9773214816) · (s − 1)2 + · · · .
134 5. L-functions
167661624456834335404812111469782006
y(P ) =
150201095200135518108761470235125
and the canonical height of P is precisely 27.3004446469 . . ., as pre-
dicted by the Birch and Swinnerton-Dyer conjecture.
where q = e 2πiz
for some coefficients an ∈ C.
Definition 5.3.1. The L-function attached to a cusp form f (z) =
n≥1 an q ∈ Sk (Γ0 (N )) is defined by
n
an a2 a3
L(f, s) = an n−s = = a1 + s + s + · · · .
ns 2 3
n≥1 n≥1
10x − 20. Are they truly the same L-series? Further calculations
show that all terms agree as we increase the precision. We will see
that the Taniyama-Shimura-Weil conjecture 5.4.5, i.e., the modularity
theorem, implies that L(E, s) = L(f, s). Notice that the conductor
of E/Q is precisely N = 11, as we saw in Example 5.1.8.
Example 5.3.3. Let N = 37 and k = 1. In Example 4.2.12 we
described the space S2 (Γ0 (37)) with basis elements {f, g} given by
the q-expansions
f (q) = q + q 3 − 2q 4 − q 7 − 2q 9 + 3q 11 − 2q 12 − 4q 13 + O(q 16 ),
g(q) = q 2 + 2q 3 − 2q 4 + q 5 − 3q 6 − 4q 9 − 2q 10 + 4q 11 + O(q 12 ).
The L-functions attached to f and g are
1 2 1 2 3 2 4
L(f, s) = 1 + s − s − s − s + s − s − s + . . . ,
3 4 7 9 11 12 13
1 2 2 1 3 4 2 4
L(g, s) = + s − s + s − s − s − s + s + ....
2s 3 4 5 6 9 10 11
Now, let EA and EB be the elliptic curves of conductor 37 described
in Example 5.1.8. Then
1 2 1 2 3 2 4
L(EB , s) = 1 + s − s − s − s + s − s − s + . . .
3 4 7 9 11 12 13
and, indeed, we shall see that L(f, s) = L(EB , s). How about EA ?
2 3 2 2 6 1 6 4 5
L(EA , s) = 1 − s
− s + s − s + s − s + s + s − s +...
2 3 4 5 6 7 9 10 11
so L(EA , s) = L(g, s) or L(f, s). Is there some form F (z) ∈ Sk (Γ0 (37))
such that L(EA , s) = L(F, s)? If so, F (q) must be a linear combina-
tion λ · f (q) + μ · g(q) for some λ, μ ∈ C. After a quick look at the
first few coefficients of the q-expansions of f and g, and those of the
series L(EA , s), one can check that, if some F works, then it must be
F (q) = f (q) − 2g(q), and indeed
(f −2g)(q) = 1−2q 2 −3q 3 +2q 4 −2q 5 +6q 6 −q 7 +6q 9 +4q 10 −5q 11 +O(q 12 )
and so
2 3 2 2 6 1 6 4 5
L(f − 2g, s) = 1 − − + − + − + + − +...
2s 3s 4s 5s 6s 7s 9s 10s 11s
Once again, we shall see that the Taniyama-Shimura-Weil conjecture
implies the equality L(f − 2g, s) = L(EA , s).
5.4. The Taniyama-Shimura-Weil conjecture 137
E : y 2 = x(x − ap )(x + bp )
would be semistable with conductor NE = |abc (see Exercise
5.7.5) and would satisfy some other technical properties. Moreover,
Frey claimed that such a curve E/Q could not be modular; i.e.,
there is no weight 2 normalized newform f ∈ S2 (Γ0 (NE )) such that
142 5. L-functions
L(f, s) = L(E, s). The problem with the modularity of E was made
precise by Serre, and Ribet [Rib90] proved in 1986 that, indeed, E
cannot be modular.
Finally, in 1995, Wiles [Wil95] and Taylor and Wiles [TW95]
proved the Taniyama-Shimura-Weil conjecture 5.4.5 for all semistable
elliptic curves E/Q. Therefore, E : y 2 = x(x − ap )(x + bp ) would have
to be modular if it existed. Hence, neither E nor the aforementioned
solution (a, b, c) to xp + y p = z p can exist, and Fermat’s last theorem
holds.
5.7. Exercises
Exercise 5.7.1. Let E/Q be an elliptic curve and let p ≥ 2 be a
prime. Define E ns (Fp ) to be the set of all non-singular points on
E(Fp ), and write Npns = |E ns (Fp )|. For instance, if p is a prime of
good reduction, then E ns (Fp ) = E(Fp ) and Npns = Np = p + 1 − ap .
144 5. L-functions
y 2 + a1 xy + a3 y = x3 + a2 x2 + a4 x + a6
147
148 A. PARI/GP and Sage
Q = [−1, 0]. Let us find P + Q and 2P (the answers are [2, −3] and
[0, −1] respectively). The commands are:
E = EllipticCurve([0,0,0,0,1]);
Ep = plot(E, -1,2.5,thickness=2);
p1=(2,3); p2=(0,1); p3=(-1,0); p4=(0,-1); p5=(2,-3);
L1=line([p1,p3],rgbcolor=(1,0,0));
L2=line([p5,p3],rgbcolor=(1,0,0));
L3=line([p4,p3],rgbcolor=(1,0,0));
L4=line([p2,p5],rgbcolor=(1,0,0));
L5=line([p4,p1],rgbcolor=(1,0,0));
T1=text(’P’,[2,3.5]); T2=text(’2P’,[0.15,1.5]);
T3=text(’3P’,[-1,.5]); T4=text(’4P’,[0.15,-1.5]);
T5=text(’5P’,[2,-3.5]);
P=point([p1,p2,p3,p4,p5],pointsize=30,
rgbcolor=(0,0,0));
PLOT=Ep+T1+T2+T3+T4+T5+L1+L2+L3+L4+L5+P; show(PLOT)
150 A. PARI/GP and Sage
Q = E(2,3);
Qplot = plot(Q, pointsize=30)+plot(2*Q, pointsize=30);
show(Qplot)
E=EllipticCurve([0,5,0,0,35]);
prime_divisors(E.discriminant())
factor(E.discriminant()).
Then one can use the command kodaira_type() to find out the
precise type of reduction: I0 is good reduction; Ij, where j > 0 is
some positive number, means bad multiplicative reduction; II, III,
IV or Ij*, for j ≥ 0, or II*,III*,IV* mean additive reduction. For
an explanation of the terminology of Kodaira symbols, see [Sil86],
Appendix C, §15. For our example E : y 2 = x3 + 5x2 + 35, we obtain
E=EllipticCurve([0,0,0,0,15625]);
prime_divisors(E.discriminant()) returns [2,3,5] but
E.kodaira_type(5) returns I0 (i.e., good).
A.1.6. The free part and the rank. It also follows from the
Mordell-Weil theorem that the free part (here free is the opposite
of torsion) of the group of points E(K) on an elliptic curve (again
over a number field K) is generated by a finite number of points P1 ,
P2 , ..., PR of infinite order. The number R of generators (of infinite
order) is called the rank of E(K). There is no known algorithm that
will always terminate and provide the rank and a set of generators.
However, the so-called “descent algorithm” will terminate in certain
cases (the descent procedure is an algorithm if X is finite, and we
conjecture that X is always finite). The following commands com-
pute lower and upper bounds for the rank and, in some cases, if they
coincide, provide the rank of the curve. There are also commands to
calculate generators; however, in many situations, the resulting points
will only generate a group of finite index in E(K) (the software will
warn you when this may be the case). Some of the algorithms take
an optional argument of a bound B.
In Sage, the command E.selmer_rank_bound() gives an upper
bound of the rank, and E.rank(), E.gens() try to find, respectively,
the rank and generators modulo torsion... but the computer may not
succeed! When these commands are called, Sage is using an algorithm
of Cremona in the background (see [Cre97]).
A.1. Elliptic curves 153
In GP use ellheight(E,P);
In Sage simply use P.height(), where P is a point on E.
In GP use S = [P1,P2,P3];
H=ellheightmatrix(E,S); matdet(H);
In Sage use E.height_pairing_matrix([P1,P2,P3]),
L=E.period_lattice()
154 A. PARI/GP and Sage
and a basis for the period lattice is found simply using L.basis().
Using PARI/GP, one can start from a lattice and obtain the associ-
ated elliptic curve, as follows:
L=[1,I];
elleisnum(L,4) returns G4 (L),
which equals 2268.8726415...,
elleisnum(L,6) returns G6 (L),
which equals -3.97...E-33, i.e., 0,
thus, L corresponds to an elliptic curve
y 2 = x3 − (34033.089 . . .)x.
The elliptic curve y 2 = x3 − (34033.089 . . .)x is isomorphic to E/Q :
y 2 = x3 − x over C. Thus, C/1, i ∼
= E(C).
You can call the generators by using the suffix [0], [1], etc. Here
are some examples:
A = SL2Z([1,1,0,1]);
G = SL2Z.gens() returns two matrices
0 −1 1 1
G[0] = , G[1] = .
1 0 0 1
A.3. L-functions
Let E/Q be an elliptic curve, and let L(E, s) be the Hasse-Weil L-
function associated to E, as in Definition 5.1.1. This L-function is
defined in Sage using the command
L=E.lseries()
or one can use L=E.lseries().dokchitser() to use Dokchitser’s
algorithms to calculate values ([Dok04]). Once we have defined L =
L(E, s), we can evaluate L. For example:
E=EllipticCurve([1,2,3,4,5]);
L=E.lseries();
L(1) which returns 0,
L(1+I) = -0.485502124065793 + 0.627256178203893*I.
The value L(E, 1) = 0 is predicted in this case by the Birch and
Swinnerton-Dyer conjecture (Conjecture 5.2.1), since the rank of E is
> 0 (in fact, the rank is 1). One can also plot L(E, x) when x takes
A.3. L-functions 157
real values (because L(E, x) is real valued for x ∈ R). For instance,
the graph in Figure 2 was created with the following lines of code:
E0=EllipticCurve([0,-1,1,-10,-20]);
L0=E0.lseries().dokchitser();
P0=plot(lambda x: L0(x).real(),0, 3);
show(P0,xmin=-0.5, ymin=0, dpi=150).
If you want to create a PDF file with your graph, you can use
E=EllipticCurve([1,2,3,4,5]);
M=E.period_lattice();
Then M.omega returns ΩE = 2.78074001376673 . . ..
Complex analysis
C = {a + bi : a, b ∈ R, i2 = −1}.
159
160 B. Complex analysis
The set of all complex numbers together with the operations of addi-
tion and multiplication form a field (see Exercise B.7.1).
There are two other operations on complex numbers that occur
often: complex conjugation and calculating the modulus, or absolute
value. The complex conjugate of α = a+bi is α = a−bi. The modulus
or absolute value of α is
√ & &
|α| = α · α = (a + bi)(a − bi) = a2 + b2 .
Notice that, for any α, β ∈ C, we have
α + β = α + β, α · β = α · β, |αβ| = |α||β|, and |α + β| ≤ |α| + |β|.
∂f ∂f ∂u ∂v ∂v ∂u
f (z) = = (−i) = + i= − i.
∂x ∂y ∂x ∂x ∂y ∂y
The last equality implies that the real and imaginary parts of every
analytic function must satisfy the following differential equations:
∂u ∂v ∂u ∂v
(B.2) = and =− .
∂x ∂y ∂y ∂x
These are called the Cauchy-Riemann differential equations.
infinity that we use in real analysis (“very, very far along the positive
x-axis”). In fact, in R, the limit limx→0 1/x is undefined (as the value
may be ±∞ depending on how we approach 0), but in C, the limit
limz→0 1/z = ∞ simply means that if z is close to 0, then 1/z is far
from 0 (in some direction, not necessarily along the x-axis).
Suppose that f (z) is some complex-valued function that is not
defined at α but is analytic in a neighborhood of α. How can f fail to
be analytic at α? The function f (z) may have a removable singularity
(e.g., sin(z)/z), an essential singularity (e.g., sin(1/z)) or a pole (e.g.,
1/z). Here we will only discuss poles in some detail (for a complete
discussion, see [Ahl79], Ch. 4, §3).
Example B.3.4. Let p(z) and q(z) be polynomials in C[z] such that
p and q have no common factors. Then the rational function p(z)/q(z)
is a meromorphic function with isolated poles at the zeros of q(z).
and
(3) has an essential singularity at α if there are infinitely many
n < 0 such that cn = 0.
If z = x + yi with x, y ∈ R:
ns = elog(n)s = elog(n)x+log(n)yi
= elog(n)x (cos(log(n)y) + sin(log(n)y)i)
= nx (cos(log(n)y) + sin(log(n)y)i).
B.7. Exercises
Exercise B.7.1. The goal of this exercise is to prove that C is a field.
(1) Show that any non-zero complex number α = a + bi has a
multiplicative inverse which is also a complex number α−1 =
c + di with c, d ∈ R.
(2) Convince yourself that C is a field; i.e., justify why C satisfies
each of the field axioms.
Exercise B.7.2. Let α be a complex number. Show that α ∈ R if
and only if α = α.
Exercise B.7.3. Show that f (z) = z is analytic on C. Also, show
that gn (z) = z n is analytic on C for every n ≥ 1 and that the deriva-
tive is gn (z) = nz n−1 .
Exercise B.7.4. Let f (z) be a complex-valued function that is ana-
lytic in a region U ⊆ C.
(1) Show that f is also continuous at every point of U (i.e.,
limh→a f (h) = f (a) for every a ∈ U ).
(2) Let f (z) = u(z)+v(z)i, where u(z) and v(z) are real-valued.
Show that u and v are continuous on U .
170 B. Complex analysis
Projective space
171
172 C. Projective space
√ √
For instance, ( 2, 2) ∼ (1, 1). We denote by [x, y] the set of all
vectors (a, b) such that (x, y) ∼ (a, b):
[x, y] = {(a, b) : a, b ∈ R such that (a, b) = (0, 0) and (x, y) ∼ (a, b)}.
Finally, we define the real projective line by
P1 (R) = {[x, y] : x, y ∈ R with (x, y) = (0, 0)}.
If you think about it, P1 (R) is the set of all lines through the origin
(each class [x, y] consists of all points — except the origin — on the
line that goes through (x, y) and (0, 0)). The important thing to
notice is that if [x, y] ∈ P1 (R) and y = 0, then (x, y) ∼ ( xy , 1), so the
class of [x, y] contains a unique representative of the form (a, 1) for
some a = xy ∈ R. This allows the following decomposition of P1 (R):
P1 (R) = {[x, 1] : x ∈ R} ∪ {[1, 0]}.
The set of points {[x, 1]} are in bijection with R and, therefore, form
a real line. The point [1, 0], which is the only point in P1 (R) that
does not belong to the real line {[x, 1]}, is called the point at infinity
(see Figure 1).
|2, 3| |1, 1|
3
2 |3, 2|
{|x, 1|}
1
|1, 0|
−3 −2 −1 1 2 3
−1
−2
−3
Notice that, as before, the class [x, y, z] contains all the points
in the line in R3 that goes through (x, y, z) and (0, 0, 0) except the
origin. We define the projective plane to be the collection of all such
lines:
P2 (R) = {[x, y, z] : x, y, z ∈ R such that (x, y, z) = (0, 0, 0)}.
If z = 0, then (x, y, z) ∼ ( xz , yz , 1). Thus,
P2 (R) = {[x, y, 1] : x, y ∈ R} ∪ {[a, b, 0] : a, b ∈ R}.
The points of the set {[x, y, 1] : x, y ∈ R} are in 1-to-1 correspondence
with the real plane R2 , and the points in {[a, b, 0] : a, b ∈ R} are called
the points at infinity and form a P1 (R), a projective line.
One interesting consequence of the definitions is that any two
parallel lines in the real plane {[x, y, 1]} intersect at a point at infinity
[a, b, 0]. Indeed, let L : y = mx + b and L : y = mx + b be distinct
parallel lines in the real plane. If points in the real plane {[x, y, 1]}
correspond to lines in R3 , then lines in the real plane correspond to
planes in R3 :
L = {[x, y, z] : mx − y + bz = 0}, L = {[x, y, z] : mx − y + b z = 0}.
What is L ∩ L ? The intersection points are those [x, y, z] such that
mx − y + bz = mx − y + b z = 0, which implies that (b − b )z = 0.
Since L = L , we have b = b and, therefore, we must have z = 0.
Hence
L ∩ L = {[x, mx, 0] : x ∈ R} = {[1, m, 0]},
and so the intersection consists of a single point at infinity: [1, m, 0].
C : f (x, y) = 0
C : F (x, y, z) = 0,
: F (x, y, z) = 0 is
where F (x, y, z) = z d · f xz , yz . Conversely, if C
a curve in the projective plane, then C : F (x, y, 1) = 0 is a curve in
the affine plane. In this case, C is the projection of C onto the chart
z = 1; we may also look at other charts, e.g., x = 1, which would
yield a curve C : F (1, y, z) = 0.
176 C. Projective space
10
7.5
2.5
−0
−2.5
−5
−7.5
−10
−1 0 1 2 3 4 5 6
10
7.5
2.5
−0
−2.5
−5
−7.5
−10
−5 −4 −3 −2 −1 −0 1 2 3 4 5
1.5
0.5
−0
−0.5
−1
−1.5
−2
−1 −0.75 −0.5 −0.25 −0 0.25 0.5 0.75 1
are the first few terms of a 3-adic integer; notice that all the coordi-
nates are coherent with the previous terms under congruences modulo
powers of 3. The vector (2, 2, 2, 2, . . .) is another element of Z3 (which
we will denote simply by 2).
179
180 D. The p-adic numbers
and
(an )∞ ∞ n ∞
n=1 · (bn )n=1 = (an · bn mod p )n=1 .
The reader should check that the sum and product of two coherent
vectors is also coherent under congruences and, therefore, a new ele-
ment of Zp . These operations make Zp a commutative ring with iden-
tity element 1 = (1, 1, 1, 1, . . .) and zero element 0 = (0, 0, 0, 0, . . .).
For any prime p ≥ 2, the p-adic integers contain a copy of Z,
where the integer m is represented by the element
m = (m mod p, m mod p2 , m mod p3 , . . .).
For example, the number 200 in Z3 is given by
200 = (2, 2, 11, 38, 200, 200, 200, 200, 200, 200, . . .).
Thus, we may write Z ⊆ Zp (see Exercise D.2.1). However, there are
elements in Zp that are not in Z, so Z Zp . Here is an example for
p = 7: we are going to show that Z7 , unlike Z,√contains an element
whose square is 2 (which we will denote by “ 2”). Indeed, 2 is a
quadratic residue in Z/7Z, and 2 has two square roots, namely 3 and
4 modulo 7. A standard theorem of number theory shows that, hence,
2 is in fact a quadratic residue modulo 7n for all n ≥ 1. Thus, there
exist integers an such that a2n ≡ 2 mod pn for all n ≥ 1. Moreover, it
can also be shown that, if an is chosen, then there is an+1 mod pn+1
with a2n+1 ≡ 2 mod pn+1 and an+1 ≡ an mod pn (we say that an can
be lifted to Z/pn+1 Z; see Exercise D.2.2). Indeed, here are the first
few coordinates of an element α of Z7 such that α2 = (2, 2, 2, . . .):
α = (3, 10, 108, 2166, 4567, . . .).
√
Thus, α should be regarded as “ 2 ” inside Z7 , and −α is another
square root of 2.
The usual integers, Z, are not a field because not every element
has a multiplicative inverse (only ±1 have inverses!). Similarly, the
p-adic integers Zp do not form a field either; e.g., p = (p, p, p, . . .)
is not invertible in Zp , but many elements of Zp are invertible. For
D.1. Hensel’s lemma 181
and the general case of Hensel’s lemma applies. Hence, there exists a
2-adic solution to x2 + 7 = 0.
D.2. Exercises
Exercise D.2.1. Show that if q and t are distinct integers (in Z),
then their representatives in Zp for any prime p ≥ 2, given by q =
(q mod pn )∞ n ∞
n=1 and t = (t mod p )n=1 , are also distinct in Zp .
Parametrization of
torsion structures
The point (0, 0) is a torsion point of the maximal order in the group.
The discriminant Δa,b of Ea,b is always assumed to be non-zero.
185
186 E. Parametrization of torsion structures
Ea,b /Q Δa,b = 0 G
y 2 = x3 + ax2 + bx a2 b2 − 4b3 = 0 Z/2Z
y 2 + axy + by = x3 a3 b3 − 27b4 = 0 Z/3Z
y 2 = x(x + a)(x + b) 0 = a = b = 0 Z/2Z ⊕ Z/2Z
t(t−1)(2t−1) t3 (t−1)(2t−1)
a= t2 −3t+1 b= (t2 −3t+1)2 Z/10Z
t(1−2t)(3t2 −3t+1) 2t2 −2t+1
a= (t−1)3 b = −a · t−1 Z/12Z
a=0 b = t2 − 1
16 Z/2Z ⊕ Z/4Z
−2(t−1) (t−5)
2
a= 10−2t
t2 −9 b= (t2 −9)2 Z/2Z ⊕ Z/6Z
(2t+1)(8t2 +4t+1) (2t+1)(8t2 +4t+1)
a= 2(4t+1)(8t2 −1)t b= (8t2 −1)2 Z/2Z ⊕ Z/8Z
Notice, however, that this does not imply that the torsion sub-
group of Et,t (Q) is identical to Z/5Z. For instance, let t = 12. The
torsion subgroup of the elliptic curve
E12,12 : y 2 − 11xy − 12y = x3 − 12x2
is isomorphic to Z/10Z. The point (0, 0) is a point of order 5, but the
point (−6, −18) has exact order 10.
Example E.1.2. According to Figure 2, each curve in the family
y 2 + (1 + t − t2 )xy + (t2 − t3 )y = x3 + (t2 − t3 )x2
has a torsion point of exact order 7, namely P = (0, 0), as long as
the discriminant Δ = t7 (t − 1)7 (t3 − 8t2 + 5t + 1) is non-zero, which
can only happen for the rational values t = 0 and t = 1. By Mazur’s
Theorem 2.5.2, the only possible torsion subgroup for an elliptic curve
over Q that contains Z/7Z as a subgroup is Z/7Z itself. Thus, the
torsion subgroup of each elliptic curve in this family is exactly Z/7Z.
Similarly, if Ea,b is an elliptic curve in one of the families in Figure
2 that correspond to G in the list
Z/7Z, Z/9Z, Z/10Z, Z/12Z, Z/2Z ⊕ Z/6Z, or Z/2Z ⊕ Z/8Z,
then the torsion subgroup of Ea,b (Q) must be exactly G.
Bibliography
189
190 Bibliography
193
194 Index