Modular Forms Script
Modular Forms Script
Modular Forms
and Elliptic Curves
DRAFT, Release 1.50
I prepared these notes for the participants of the lectures. The lecture took place at
the mathematical department of LMU (Ludwig-Maximilians-Universität) at
Munich. The first time during the winter term 2017/18 and then during the winter
term 2020/21. Some parts of the notes rely on notes of an unpublished lecture
of O. Forster.
Compared to the oral lecture in class these written notes contain some additional
material.
I thank all participants - in particular W. Hensgen - for pointing out some errors
and for their proposals for improvement. Please report to
any further errors or typos, adding the version of the lecture notes.
Release notes:
• Release 1.50: Update acknowledgment.
• Release 1.49: Chapter 5. Remark 5.15 moved from Chapter 7. Chapter 7.
Revision.
• Release 1.48: Chapter 6. Revision, Lemma 6.21 added.
• Release 1.47: Chapter 5. Proposition 5.9 added.
• Release 1.46: Chapter 3 and Chapter 2. Revision.
• Release 1.45: Chapter 4. Revision.
• Release 1.44: Chapter 5. Revision.
• Release 1.43: Chapter 5. Theorem 5.25 proof corrected.
• Release 1.42: Chapter 5. Theorem 5.25 corrected. Chapter 6 minor revisions.
• Release 1.41: Chapter 7. Section 7.2 added. Chapter 5. Theorem 5.23 expanded.
Theorem 5.25 expanded and corollary integrated.
• Release 1.40: Chapter 7. Section 7.1 added.
• Release 1.39: Chapter 4. Proposition 4.25 expanded.
• Release 1.38: Chapter 6. Sections 6.1 and 6.2 added.
• Release 1.37: Chapter 5. Section 5.2, Lemma 5.24 replaced by new version,
subsequent Corollary added, Theorem 5.25 adapted, renumbering, Section 5.3
added.
• Release 1.36: Chapter 5. Section 5.2, Proposition 5.28 added, minor revision.
• Release 1.35: Chapter 5. Section 5.2 added.
• Release 1.34: Chapter 5. Section 5.1, minor revision.
• Release 1.33: Chapter 4. Example 4.39 expanded.
• Release 1.32: Chapter 5. Section 5.1 added.
• Release 1.31: Chapter 4. Section 4.2, Definition 4.19 expanded, Lemma 4.30
and Theorem 4.31 clarified, Remark 4.27 added, renumbering, minor revision.
Section 4.3 added.
3
• Release 1.30: Chapter 4. Section 4.2, Example 4.36 added, minor revision.
• Release 1.29: Chapter 3, Lemma 3.4, exponent of determinant changed.
Chapter 4, Example 4.4 added. Renumbering, Section 4.2 added.
• Release 1.28: Chapter 4. Proposition 4.8, proof corrected. Section 4.1, minor
revisions.
• Release 1.27: Chapter 2, Remark 2.30, formulas corrected by extended
Legendre symbol and Kronecker symbol. Chapter 4, Proposition 4.6 added,
minor revisions, renumbering. List of results updated.
• Release 1.26: Chapter 3, proof of Lemma 3.4: formula corrected. Chapter 4,
minor revisions
• Release 1.25: Chapter 4, Section 4.1 added.
• Release 1.24: Minor revisions.
• Release 1.23: Minor revisions.
• Release 1.22: Chapter 3 completed.
• Release 1.21: Section 3.2, some typos corrected.
• Release 1.20: Minor revisions.
• Release 1.19: Minor revisions.
• Release 1.18: Chapter 2, Remark 3.5 added.
• Release 1.17: Minor revisions.
• Release 1.16: Chapter 2, Remark 2.8, Example 2.11, Lemma 2.26 corrected.
Chapter 3, Section 3.1 reordered.
• Release 1.15: Chapter 3, Remark 3.7 added, Remark 3.7 expanded, minor
revisions, renumbering.
• Release 1.14: Minor revisions.
• Release 1.13: Chapter 3, Section 3.2 added. List of results added.
• Release 1.12: Chapter 1, Proposition 1.19 added, renumbering. Chapter 2,
Remark 2.17 added, renumbering. Chapter 3, Section 3.1 added.
• Release 1.11: Chapter 2 completed.
• Release 1.10: Chapter 2, Section 2.2 correction of some typos.
• Release 1.9: Chapter 1, revision of Corollary 1.9 and 1.10, Lemma 1.13,
Corollary 1.15 and Theorem 1.21. Further minor revisions.
• Release 1.8: Chapter 2 minor revision.
• Release 1.7: Chapter 2 minor revision, some additions.
• Release 1.6: Chapter 2 minor revision.
• Release 1.5: Chapter 2, Sections 2.1 and 2.2 added.
• Release 1.4: Chapter 1, minor revision.
• Release 1.3: Introduction added.
• Release 1.2: Chapter 1, minor revision.
• Release 1.1: Complete revision, starting with Chapter 1.
Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
PARI files . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1 Elliptic functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.1 The field of elliptic functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.2 The Weierstrass ℘-function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
1.3 Abel’s theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
v
vi Contents
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 277
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 281
Introduction
Modular Forms and elliptic curves are a classical domain from mathematics. At
least, since the proof of Fermat’s last conjecture the domain attracts widespread at-
tention. The domain of Modular Forms integrates the three mathematical disciplines
Complex Analysis, Algebraic Geometry, and Algebraic Number Theory.
• Arithmetic geometry: Elliptic curves can be defined over different fields, e.g.
over Q or F p and over the ring Z.
The algebraic point of view identifies complex elliptic curves with smooth cubic
hypersurfaces of P2 . Hence elliptic curves are the first ones in the series of cubic
hypersurfaces in complex projective space Pn , n ≥ 2. The higher dimensional
cubics are also challenging examples, see [29, Chap.V, §4], [31].
Γ ×H →
− H
The quotients
1
2 Introduction
Γ0 (N) \H
of the restricted action of the Hecke congruence subgroups
Γ0 (N) ⊂ Γ
Chapter 4 makes a new start from the viewpoint of algebraic geometry: Complex
tori embed as elliptic curves into the complex projective plane. They can be repre-
sented as non-singular cubic curves, defined as zero set of a Weierstrass polynomial.
The main tool to represent a complex torus as an elliptic curve is the Weierstrass ℘-
function and its differential equation. Conversely, each complex elliptic curve arises
as embedding of a complex torus.
Choosing different fields for the coefficents of the Weierstrass polynomials al-
lows to consider elliptic curves defined over C, Q, Z or even over the finite fields F p .
The focus of the chapter are some relations between complex analytic geometry, al-
gebraic geometry and arithmetic geometry.
In the second, more advanced part of the lecture notes Chapter 6 deals with
deeper applications of the theory of modular forms to algebraic number theory. The
chapter investigates the relation between tori with complex multiplication and imag-
inary quadratic number fields. The main role is played by the modular invariant j.
As an application the chapter proves a lower bound of the class number.
The final Chapter 7 gives an outlook to the modularity theorem for elliptic curves,
which plays the dominant role in Wiles’ proof of the Fermat conjecture. A second
section gives an outlook to the moonshine relation between the Fourier coefficients
of the modular j-invariant and the dimensions of the irreducible representations of
the monster group, which at day culminates in the work of Borcherds.
It is also the aim of these lecture notes to illustrate their results and the outlook
by a series of numerical calculations using a computer algebra system.
PARI files
Chapter 1:
• Weierstrass_p_function_08: Laurent expansion of the
Weierstrass ℘-function and its derivative, cf. Remark 1.14.
Chapter 2
• modular_curve_as_covering: Genus, degree and further parameters of
modular curve X0 (N), cf. Remark 2.30.
Chapter 3
• Congruence_subgroup_19: Dimension of Mk (Γ0 (pn )) via Riemann-Roch,
cf. Remark 3.27.
Chapter 4
• Torus_to_Weierstrass_equation_01: The plane cubic of an
embedded torus, cf. Example 4.4
3
4 PARI files
Chapter 5
• Hecke_matrix_02: Hecke operator T2 ∈ End(S28 (Γ )), cf. Example 5.14.
Chapter 6
• modular_polynomial_01: Modular polynomials of different levels, cf.
Example 6.9.
Chapter 7
• Elliptic_curve_taniyama_10: Illustration of Wiles’ theorem by some
numerical examples, cf. Example 7.7.
\r filename
For more information see the PARI manual from the PARI homepage.
Part I
General Theory
6
The chapter starts with complex analysis in the plane. Subsequently the results are
translated into the language of Riemann surface. Here they are combined with some
results from algebraic topology. A final section applies the Riemann-Roch theorem
to prove Abel’s theorem about divisors on complex tori.
The present section deals with complex analysis in the plane, i.e. we study holomor-
phic and meromorphic functions on domains in the complex plane C.
The trigonometric functions sin and cos are examples of periodic holomorphic
functions, while tan and cot are periodic meromorphic. All periods of these func-
tions are integer multiples of respectively 2π and π. Elliptic functions are meromor-
phic with period group isomorphic to Z ⊕ Z.
f : G\S → C
9
10 1 Elliptic functions
Only seemingly Definition 1.1 excludes the poles from the definition of period-
icity. A pole z0 of f is an isolated singularity of f . Therefore the function f expands
into a convergent Laurent series around z0
∞
f (z) = ∑ an · (z − z0 )n
n≥n0
with coefficients
1 Z
f (z)
an = dz for suitable ε > 0.
2πi (z − z0 )n+1
|z−z0 |=ε
If f has the period ω then the Laurent expansions of f around z0 and z0 + ω are the
same because the integrands attain the same value at z and z + ω. Hence periodicity
of f includes also the poles.
If a meromorphic function f has a period ω 6= 0 one can ask for the set of all
periods. Apparently elements of the form kω, k ∈ Z are also periods of f . How
many independent periods of f exist?
ii) If Γf were not discrete then a convergent sequence (ωn )n∈N of pairwise distinct
periods ωn ∈ Γf would exist.
1.1 The field of elliptic functions 11
Hence for an arbitrary but fixed z0 ∈ G, which is not a pole of f , and for all n ∈ N
f (z0 ) = f (z0 + ωn ).
Γ ⊂ (C, +)
3. Rank = 2:
Γf = Zω1 + Zω2
is also named the period lattice of f , see Figure 1.1. The subset
P = {λ1 ω1 + λ2 ω2 : 0 ≤ λ1 , λ2 < 1}
Which elliptic functions do exist? Proposition 1.7 shows that it is not interesting
to study the subclass of holomorphic elliptic functions. Therefore we will focus on
1.1 The field of elliptic functions 13
meromorphic functions with poles. See Proposition 1.8, and its corollaries about the
value attainment of elliptic functions.
f :C→C
is constant.
Proof. An entire function f is holomorphic and attends the maximum of its modulus
on its closed period parallelogram which is a compact set. Being a bounded entire
function, f is constant according to Liouville’s theorem, q.e.d.
∑ resζ ( f ) = 0.
ζ ∈P
Proof. The boundary ∂ P comprises only finitely many poles of f . Therefore we can
choose a number a ∈ C such the boundary of the translated period parallelogram
Pa := a + P
1 Z
f (z) dz = ∑ resζ ( f ) = ∑ resζ ( f ).
2πi ∂ Pa ζ ∈P ζ ∈P
a
The integrand on the left-hand side is doubly periodic. Therefore the integration
along opposite sides of the period parallelogram cancels and the whole integral van-
ishes, q.e.d.
2. There are no elliptic functions f with only one zero of order = 1 modulo the
period lattice Γf and no other zero mod Γf .
Corollary 1.10 (Counting poles and zeros of elliptic functions). Any non-constant
elliptic function f attains modulo each lattice
Λ ⊂ Γf
all values a ∈ C ∪ {∞} with the same multiplicity. In particular, f has mod Λ the
same number of poles and zeros taken with multiplicity.
Proof. The case a 6= ∞ reduces to the second claim by considering the function f − a.
It has the same poles as f . Therefore it suffices to show that f has the same number
of zeros and poles, taken with multiplicity. Moreover it suffices to prove the theorem
for Λ = Γf .
q.e.d.
Λ := Z ω1 + Z ω2 .
1
GΛ ,k := ∑ 0 ωk
ω∈Λ
Proof. We choose the exhaustion of Λ by the sequence (Λn )n∈N of disjoint sets of
indices of increasing modulus
i.e.
˙
[
Λ= Λ .
n∈N n
Then
card Λn = 8n.
A suitable constant c exists with |ω| ≥ c · n for all n ∈ N and for all ω ∈ Λn . There-
fore !
1 8·n 8 1
∑ |ω|k ≤ ck · nk ≤ ck · nk−1 .
ω∈Λn
For k > 2
∞ 1
∑ nk−1 < ∞.
n=1
We will study elliptic functions with period lattice Λ . According to Corollary 1.9
a candidate f for the most simple example would have only one pole mod Λ , the
pole having order = 2 and residue = 0. The Laurent expansion of f around z0 = 0
would start
1
f (z) = 2 + a1 · z + O(2).
z
Being doubly periodic, f would have poles exactly at the points ω ∈ Λ . The follow-
ing Theorem 1.12 shows that such a function actually exists.
1 1
∼ .
(z − ω)2 ω 2
16 1 Elliptic functions
Therefore the series with these summands alone does not converge. We will show
that the difference
1 1
2
− 2
(z − ω) ω
is proportional to
1
.
ω3
Therefore the series ℘(z) converges.
Note: TeX reserves the separate symbol “backslash wp” to denote the Weierstrass
function ℘.
holds
|ω|
− |z| ≥ 0.
2
The triangle inequality
|z|2 |z| 4 · R2 R R
≤ 4· + 8 · ≤ +8· = 10 ·
|ω|4 |ω|3 2 · R · |ω|3 |ω|3 |ω|3
independent from |z| < R.
1.2 The Weierstrass ℘-function 17
converges absolutely. For |z| ≤ R we decompose the series with respect to the sum-
mation over ω ∈ Λ
" !# " !#
1 1 1 1 1
+ − + −
z2 ω∈Λ 0∑
,|ω|<2R
(z − ω)2 ω 2 ∑
ω∈Λ ,|ω|≥2R
(z − ω)2 ω 2
The first summand is meromorphic with poles exactly at the points z ∈ Λ ∩ D2R (0).
The second summand converges absolutely and compactly on D2R (0) due to Lemma 1.11,
and defines a holomorphic function on D2R (0).
Because R can be choosen arbitrary, the series ℘ is compact convergent and de-
fines a meromorphic function on C with pole set Λ .
−2 −2 1
℘ 0 (z) = 3
+ ∑ 3
= −2 ∑ 3
z ω∈Λ 0
(z − ω) ω∈Λ (z − ω)
℘ 0 (z + ω0 ) = ℘ 0 (z)
℘(z + ω j ) = ℘(z) + c j
z := −ω j /2,
18 1 Elliptic functions
Hence
c j = 0, q.e.d.
Λ = Zω1 + Zω2
and denote by ( )
ω1 ω2 ω1 + ω2
Z(ω1 ,ω2 ) := , ,
2 2 2
the set of half-periods of its fundamental period parallelogram with respect to
(ω1 , ω2 ). The derivative ℘ 0 of the Weierstrass function ℘ of Λ has mod Λ
• a pole of order = 3 at 0 ∈ C and no other poles,
Proof. Consider u ∈ Z(ω1 ,ω2 ) . According to Theorem 1.12 the function ℘ has no
pole at u. The derivative ℘ 0 has the same pole set as ℘. Because ℘ is even, the
function ℘ 0 is an odd elliptic function
Therefore
℘ 0 (u) = 0,
i.e. each point of Z(ω1 ,ω2 ) is a zero of ℘ 0 within the fundamental period parallelo-
gram. Hence ℘ 0 has at least three zeros mod Λ .
Moreover ℘ 0 like ℘ has a single pole at 0 ∈ C mod Λ . The pole of ℘ 0 has
order = 3. Corollary 1.10 implies that ℘ 0 has exactly three zeros mod Λ . As a
consequence, the points of Z(ω1 ,ω2 ) represent exactly the zeros of ℘ 0 , and each zero
has order = 1, q.e.d.
Note that the determination of the two zeros mod Λ of℘is much more delicate [19].
Remark 1.14 (Weierstrass function ℘). Figure 1.2 shows the Laurent expansion
around 0 ∈ C of the Weierstrass function ℘ and its derivative ℘ 0 for the two
lattices Λ
1.2 The Weierstrass ℘-function 19
Theorem 1.12 shows that the ℘-function of the lattice Λ has all points from Λ as
periods. Corollary 1.15 shows the other direction: There are no additional periods,
i.e. Γ℘ = Λ .
Corollary 1.15 (Period lattice of ℘ and ℘ 0 ). The Weierstrass function ℘ of the
lattice Λ and its derivative ℘ 0 have the period lattice Λ , i.e.
Λ = Γ℘ = Γ℘ 0
Proof. We know
Λ ⊂ Γ℘ ⊂ Γ℘ 0
20 1 Elliptic functions
Here the first inclusion has been shown in Theorem 1.12, while the second inclusion
is obvious. We show
Γ℘ 0 ⊂ Λ :
Let ω ∈ Γ℘ 0 be a period of ℘ 0 . Then
Either ω/2 is not a pole of℘ 0 . Then ω/2 is a zero of℘ 0 and 1.13 implies ω/2 ∈ Λ /2
or ω ∈ Λ .
Or ω/2 is a pole of ℘ 0 . Then Theorem 1.12 implies ω/2 ∈ Λ , and in
particular ω ∈ Λ , q.e.d.
1 ∞
℘(z) = + ∑ a2k · z2k
z2 k=1
1 1
−
(z − ω)2 ω 2
−1 1 1/ω ∞
= = = (1/ω) · ∑ (1/ω)ν · zν
z − ω ω − z 1 − (z/ω) ν=0
shows
1 ∞ ∞ 1
ν ν−1
= (1/ω) · ∑ ν · (1/ω) · z = ∑ (ν + 1) · ν+2 · zν
(z − ω)2 ν=1 ν=0 ω
and
1 1 ∞ 1
2
− 2 = ∑ (ν + 1) ν+2 · zν .
(z − ω) ω ν=1 ω
1.2 The Weierstrass ℘-function 21
Because ℘ is even and the series converges absolutely, see Theorem 1.12, we obtain
after rearrangement and using that ℘ is even
! !
1 1 1 1 ∞ 1
℘(z) = 2 + ∑ − = 2 + ∑ (ν + 1) ∑ ν+2 zν =
z ω∈Λ 0 (z − ω)2 ω 2 z ν=1 ω∈Λ 0
ω
!
1 ∞ 1
= 2 + ∑ (2k + 1) ∑ z2k , q.e.d.
z k=1 ω∈Λ 0
ω 2k+2
g2 := gΛ ,2 := 60 · GΛ ,4 , g3 := gΛ ,3 := 140 · GΛ ,6
1 1
℘(z) = + ∑ (2k + 1) · G2k+2 · z2k = 2 + 3 · G4 · z2 + 5 · G6 · z4 + O(6)
z2 k≥1 z
2
℘ 0 (z) = − + 6 · G4 · z + 20 · G6 · z3 + O(5)
z3
4 24 · G4
℘ 0 (z)2 = − − 80 · G6 + O(2)
z6 z2
1
℘(z)2 = + 6 · G4 + 10 · G6 · z2 + O(4)
z4
1 6 · G4 3 · G4 1 9 · G4
℘(z)3 = 6
+ 2 + 10 · G6 + 2 + 5 · G6 + O(2) = 6 + 2 + 15 · G6 + O(2)
z z z z z
Therefore all summands of order < 1 of the sum
2
℘ 0 − 4 ·℘3 + g2 ·℘+ g3
cancel and
2
℘ 0 − 4 ·℘3 + g2 ·℘+ g3 = O(2).
The left-hand side is an elliptic function. The right-hand side shows that the function
is even holomorphic and vanishes at 0 ∈ C. According to Proposition 1.7 the left-
hand side is constant, i.e. zero, q.e.d.
22 1 Elliptic functions
g2 := 60 · GΛ ,4 , g3 := 140 · GΛ ,6 .
F(℘ 0 ) = 0 ∈ C(℘).
The unique pole of f mod Λ has order k ≥ 2 due to Corollary 1.9. We prove
f ∈ C(℘,℘ 0 )
For the induction step assume k > 2. On one hand, for even k = 2m a suitable
constant c ∈ C exists such that f − c ·℘ m has only poles of order < k. Therefore
f − c ·℘ m ∈ C(℘,℘ 0 )
f − c ·℘ m−1 ·℘ 0
1.2 The Weierstrass ℘-function 23
f − c ·℘ m−1℘ 0 ∈ C(℘,℘ 0 ).
In both cases
f ∈ C(℘,℘ 0 ).
• Eventually, we consider the case of an arbitrary pole set of f .
Then f has mod Λ only finitely many poles zν ∈ P\Λ , ν = 1, ..., n. The function ℘
is holomorphic in P \ Λ . Hence for suitable constants k1 , ..., kn ∈ N the function
n
f (z) · ∏ (℘(z) −℘(zν ))kν
ν=1
Therefore
f ∈ C(℘,℘ 0 ), q.e.d.
Λ = Zω1 + Zω2 ⊂ C
GΛ (T ) := 4T 3 − g2 · T − g3 ∈ C[T ], g2 := 60 · GΛ ,4 , g3 := 140 · GΛ ,6
has three pairwise distinct zeros: The values ℘(ω) ∈ C at the half-period points
of Λ
ω ∈ ZΛ := {ω1 /2, ω2 /2, (ω1 + ω2 )/2}
As a consequence, the discriminant
1
discr(GΛ ) = · (g32 − 27 · g23 ) ∈ C
24
of the polynomial GΛ (T ) is non-zero, in particular
g32 − 27 · g23 6= 0.
Proof. 1. Pairwise distinct zeros: The differential equation of℘ 0 from Theorem 1.17
implies
℘ 0 (ω)2 = 4 ·℘(ω)3 − g2 ·℘(ω) − g3 = GΛ (℘(ω))
24 1 Elliptic functions
℘(ω) 6= ℘(ω 0 ).
The zeros and poles of an elliptic functions cannot be prescribed in an arbitrary way:
Abel’s Theorem states a sufficient and necessary condition.
Proposition 1.20 (Zero and pole divisor). Consider an elliptic function and denote
by A = (ai )i=1,...,n its family of zeros in P and by B = (bi )i=1,...,n its family of poles
in P. Then n ≥ 2 and
∑ ai − ∑ bi ∈ Λ .
i=1,...,n i=1,...,n
Note that each point p of A and B appears as often as its multiplicity as zero or pole
of f indicates. By Corollary 1.9 holds n ≥ 2, and by Corollary 1.10 both families A
and B have the same cardinality.
1.3 Abel’s theorem 25
Pa := P + a
of the fundamental parallelogram such that f has neither zeros nor poles on the
boundary ∂ Pa . With a counter-clockwise orientation of ∂ Pa the residue theorem
gives !
1 Z f 0 (z) f 0 (z)
z· dz = ∑ res p z · .
2πi ∂ Pa f (z) p∈Pa f (z)
In a neighbourhood of p
f 0 (z) k
= + f1 (z)
f (z) z − p
with f1 holomorph. As a consequence
and !
f 0 (z) n n
∑ res p z· = ∑ ai − ∑ bi .
p∈Pa f (z) i=1 i=1
Because
26 1 Elliptic functions
f (A) = f (B)
an integer m2 ∈ Z exists with
1 Z f 0 (z)
· z· dz = m1 · ω1 + m2 · ω2 ∈ Λ , q.e.d.
2πi f (z)
∂ Pa
Theorem 1.21 (Abel’s theorem). Consider n ≥ 2 and two finite families of complex
points
A := (ai )i=1,...,n , B := (bi )i=1,...,n ⊂ C
with A and B pairwise disjoint mod Λ . Then are equivalent:
• There exists an elliptic function f ∈ M (Λ ) which has mod Λ the zero set A and
the pole set B
Proof. Proposition 1.20 proves that the zero set and the pole set of f satisfy the con-
dition stated above. In order to prove the opposite direction of the theorem assume
two sets A and B with the property stated above. Without restriction we may assume
n
∑ (a j − b j ) = 0
j=1
has
– a simple zero at a1 and at a2 if a1 6≡ a2 mod Λ , and a zero of order = 2 at
a1 ≡ a2 otherwise
• a1 ∈ Λ : Then b1 ∈
/ Λ . Set
1
f :=
℘−℘(b1 )
• b1 ∈ Λ : Then a1 ∈
/ Λ . Set
f := ℘−℘(a1 )
Hence f is elliptic with prescribed zeros and poles.
Then mod Λ
(a1 − b1 ) + ... + (an−1 − bn−1 ) + (an − a˜n ) ≡ 0
and
(ãn + an+1 ) − (bn + bn+1 ) ≡ 0.
Depending on the position of ãn relative Λ we distinguish the following cases:
28 1 Elliptic functions
• ãn 6≡ a j for all j = 1, ..., n: First we exclude the two exceptional cases ãn ≡ bn
and ãn ≡ bn+1 : They lead to respectively
f2 ∈ M (Λ )
• ãn ≡ an : Then
(a1 − b1 ) + ... + (an−1 − bn−1 ) ≡ 0
By induction assumption two elliptic functions exist
f1 ∈ M (Λ )
with zeros (a1 , ..., an−1 ) and poles (b1 , ..., bn−1 ) and
f2 ∈ M (Λ )
and
(a1 + an+1 ) − (bn + bn+1 ) ≡ 0.
Similarly to the previous case by induction assumption two elliptic functions
exist
f1 ∈ M (Λ )
with zero set (a2 , ..., an ) and pole set (b1 , ..., bn−1 ) and
f2 ∈ M (Λ )
f := f1 · f2 ∈ M (Λ )
has the prescribed zeros (a1 , ..., an+1 ) and poles (b1 , ..., bn+1 ), q.e.d.
1.3 Abel’s theorem 29
Corollary 1.22 (Elliptic functions with pole of order = 3). An elliptic function f ∈ M (Λ )
with a pole mod Λ at 0 ∈ Λ of order = 3 and no other poles mod Λ , has exactly
three zeros a1 , a2 , a3 mod Λ . They satisfy
a1 + a2 + a3 = 0 mod Λ .
H := {τ ∈ C : Im τ > 0}
the set of positively oriented bases of R2 . The set R of all lattices Λ ⊂ C is the set
of (left) equivalence classes of M
SL(2, Z)\M
31
32 2 The modular group Γ and its Hecke congruence subgroups Γ0 (N)
Proof. The invertible matrices γ ∈ M(2 × 2, Z) have determinant ±1. Because both
bases are positively oriented we have det γ = 1, q.e.d.
F : C/Λ1 → C/Λ2
between two tori with F(0) = 0. Then a uniquely determined number a ∈ C exists
such that the following diagram commutes
µa
C C
p1 p2
F
C/Λ1 C/Λ2
F̃ : C →
− C,
2.1 The moduli space of complex tori and group actions 33
F̃(z + ω) − F̃(z) ∈ Λ2 .
F̃(z + ω) − F̃(z)
F̃ 0 (z + ω) − F̃ 0 (z) = 0.
F : T1 → T2
µa · Λ1 ⊂ Λ2 .
Λτ = Z · 1 + Z · τ and Λτ 0 = Z · 1 + Z · τ 0
a τ +b
τ0 =
c τ +d
for a matrix
ab
∈ SL(2, Z).
cd
Λ = Zω1 + Zω2
2.1 The moduli space of complex tori and group actions 35
Λτ := Z · 1 + Z · τ
due to
1
· Λ = Λτ .
ω1
According to Corollary 2.4 the two tori T and T 0 := C/Λτ are biholomorphic
equivalent.
2. i) Assume
a τ +b ab
τ0 = γ · τ = with γ := ∈ SL(2, Z).
c τ +d cd
Set
ω2 τ
:= γ · , i.e. ω2 = aτ + b, ω1 = cτ + d.
ω1 1
The base change implies
Z · ω1 + Z · ω2 = Z · 1 + Z · τ = Λτ .
As a consequence
1
· Λτ = Λτ 0
ω1
because
ω2
τ0 = .
ω1
The lattices Λτ and Λτ 0 are similar. Hence the tori C/Λτ and C/Λτ 0 are biholo-
morphic equivalent according to Corollary 2.4.
ii) For the opposite direction assume that two tori with normalized lattices
exists with 0
τ ατ
=γ· ,
1 α
i. e.
a · ατ + b · α a · τ + b
τ0 = = , q.e.d.
c · ατ + d · α c · τ + d
Theorem 2.5 shows the consequences when two points of H are related by an
element of the group SL(2, Z). We formalize this relation by the concept of a group
acting on a topological space.
φ : G×X →
− X, (g, x) 7→ gx,
also written g(x) := gx, satisfying the following properties: For all g1 , g2 ∈ G and
for all x ∈ X
• ex = x
• g1 (g2 x) = (g1 g2 )x
Gx := {g ∈ G : gx = x} ⊂ G.
{gx ∈ X : g ∈ G} ⊂ X
Two points of X are equivalent with respect to the group action if they belong
to the same orbit. The set of equivalence classes is the orbit space G\X, the
set of orbits of the action. The canonical map to the orbit set is denoted
π :X →
− G\X, x 7→ [x].
2.1 The moduli space of complex tori and group actions 37
• The action is transitive if the whole set X is a single orbit, i.e. if for any two
points x1 , x2 ∈ X exists g ∈ G with x2 = gx1 .
Apparently all points of the same orbit have conjugated isotropy groups because
Ggx = g · Gx · g−1 .
Theorem 2.5 deals with the specific group action from Definition 2.7.
Definition 2.7 (Action of SL(2, Z) on H). The SL(2, Z)-left action on H is the map
aτ + b ab
Φ : SL(2, Z) × H → H, (γ, τ) 7→ γ(τ) := , γ := .
cτ + d cd
then
Im τ
Im γ(τ) = det(γ) ·
|cτ + d|2
As a consequence
τ ∈ H =⇒ γ(τ) ∈ H.
Hence the map from Definition 2.7 extends to an action
aτ + b
GL(2, R)+ × H →
− H, γ(τ) :=
cτ + d
which apparently restricts to an action of all subgroups of GL(2, R)+ , in particu-
lar to Γ ⊂ GL(2, R)+ .
38 2 The modular group Γ and its Hecke congruence subgroups Γ0 (N)
Due to Theorem 2.5 the biholomorphic equivalence classes of complex tori cor-
respond bijectively to the points of the orbit space SL(2, Z)\H. Hence the orbit
space is named the moduli space of 1-dimensional complex tori.
2. Action of Γ on H∗ : We have
Moreover, due to
az + b a + (b/z)
=
cz + d c + (d/z)
we can extend γ to the argument ∞ by setting
∞ if d 6= 0 and c = 0
γ(∞) :=
a
otherwise
c
Set
H∗ := H ∪
˙ Q∪
˙ {∞}
Then the extension
Φ : SL(2, Z) × H∗ →
− H∗ , (γ, z) 7→ γ(z),
is well-defined.
Note. The proof of Theorem 2.28 will provide H∗ with a specific topology which
induces on Q ⊂ H a topology different from the subspace topology of Q ⊂ C and
also different neighbourhoods of ∞ ∈ H∗ .
Γ := SL(2, Z)
ΦN : Γ0 (N) × H∗ → H∗ .
2.1 The moduli space of complex tori and group actions 39
3. Points τ ∈ H with non-trivial isotropy group, i.e. Γτ ) {±id}, are named elliptic
points of Γ0 (N). Half the order of the isotropy group
ord Γ0 (N)τ
hτ :=
2
is named the period of τ. An orbit is named an elliptic orbit if at least one point
of the orbit - and hence all points - are elliptic.
4. The orbits of the points τ ∈ Q ∪ {∞} are named the cusps of Γ0 (N). The set
of cusps is denote by cusp(Γ0 (N)). The width (Deutsch: “Breite”) of the cusp
passing through
τ ∈ Q ∪ {∞}
is the index of the isotropy groups
hτ := [Γτ : Γ0 (N)τ ]
−id ∈ Γ0 (N)τ
The elements
±id ∈ Γ0 (N)
are the only elements which act trivially on H. Therefore some authors apply the
name modular group to the quotient
We will not follow this convention. Because one can study also the action of
certain subgroups of SL(2, Z) which do not contain the element −id.
3. Corollary 2.18 will show that the isotropy groups of all elliptic points from Def-
inition 2.9 are finite groups. But the isotropy group Γ∞ ⊂ Γ is an infinite group,
more specific
11
Γ∞ =< ±T > with translation T := ∈Γ
01
40 2 The modular group Γ and its Hecke congruence subgroups Γ0 (N)
4. Theorem 2.28 will show the importance to consider the group action not only on
the open upper half-plane H but also on its “closure” H∗ .
Example 2.11 (Cusps of Γ , Γ0 (2), Γ0 (4) and Γ0 (11)). See [17, Sect. 3.8], and the
PARI commands mfcusps, mfnumcusps, mfcuspwidth.
1. For N ∈ N the number ε∞ (Γ0 (N)) of cusps is
∑ φ (gcd(d, N/d))
d|N
with the sum extending over all positive divisors d of N, and φ denoting the Euler
function defined as
Φ(n) := card (Z/nZ)∗
2. The group Γ has a single cusp
• Orbit of the point 0:
Q ∪ {∞}
with width = 1. All rationals p/q ∈ Q belong to the Γ -orbit of ∞, see
Remark 2.8.
with width = 2.
with width = 1.
with width = 4.
with width = 1.
with width = 1.
gcd(a, b, c) = 1.
gcd(a + r · b, c) = 1.
r := ∏ p
p|c,p-a
42 2 The modular group Γ and its Hecke congruence subgroups Γ0 (N)
q | gcd(a + r · b, c).
gcd(a, b, c, ) 6= 1,
a contradiction.
and
SL(2, Z/NZ) := {B ∈ M(2 × 2, Z/NZ) : det B = 1 ∈ (Z/NZ)∗ }
and the residue morphism
π :Γ →
− SL(2, Z/NZ), A 7→ A.
det A = 1 ∈ Z/NZ
implies
αδ − β γ ≡ 1 mod N,
2.1 The moduli space of complex tori and group actions 43
and therefore
gcd(γ, δ , N) = 1.
Lemma 2.12 provides an integer r ∈ Z satisfying
gcd(γ, d) = 1
for
d := δ + rN.
On one hand, from the determinant
αd − β γ = αδ − β γ + αrN = 1 + sN
gcd(γ, d) = 1
provides an equation
γy − dx = s
with suitable integers x, y ∈ Z.
In particular, all Hecke congruence subgroups have finite index in the modular
group.
Proof. See also [53, Chap. 1.6].
44 2 The modular group Γ and its Hecke congruence subgroups Γ0 (N)
id ψ
Apparently
[Γ0 (N) : Γ (N)] = card(im ψ) = φ (N) · N
with the Euler function φ . The exact sequence of the first line of the diagram implies
and the factorisation of the SL(2, −)-groups due to the Chinese remainder theorem
k
e
SL(2, Z/NZ) ' ∏ SL 2, Z/p j j Z .
j=1
One has
2.1 The moduli space of complex tori and group actions 45
!
e
3e 1
card SL 2, Z/p j j Z = p j j 1− 2 ,
pj
see [53, Chap. 1.6] and [17, Exerc. 1.2.3]. For a proof see also [28, Cor. 2.8] and use
for the reduction from GL to SL the exact sequence of the determinant
det
1→ − GL (2, Z/pe Z) −→ (Z/pe Z)∗ →
− SL (2, Z/pe Z) → − 1.
And also
[Γ : Γ (N)] = card SL(2, Z/NZ),
see part i). It follows
k
e
[Γ : Γ (N)] = ∏ card SL 2, Z/p j j Z =
j=1
! !
k 1 k 1
3e 2 ej
=∏ pj j
1− 2 = N · ∏ pj · 1− 2 =
j=1 pj j=1 pj
! ! !
k 1 1 k 1
e
= N2 · ∏ p j · 1 −
j
1+ = N 2 · φ (N) · ∏ 1 +
j=1 pj pj j=1 pj
Hence
[Γ : Γ (N)]
[Γ : Γ0 (N)] = =
[Γ0 (N) : Γ (N)]
!
2 k 1
N · φ (N) · ∏ j=1 1 + !
pj 1
= = N ·∏ 1+ , q.e.d.
N · φ (N) p|N
p
[Γ : Γ0 (2)] = 3
for
−1 1
U :=
−1 0
46 2 The modular group Γ and its Hecke congruence subgroups Γ0 (N)
[Γ : Γ0 (4)] = 6
The section continues the study of the moduli space of complex tori. We denote by
aτ + b ab
Φ : Γ × H → H, (γ, τ) 7→ γ(τ) := , γ :=
cτ + d cd
the continuous Γ -left action introduced in Remark 2.8. Our final aim is to provide
the orbit space
Y := Γ \H
with an analytic structure and to compactify the resulting Riemann surface to obtain
a compact Riemann surface X. The latter will be named the modular curve of the
action Φ. We proceed along the following steps:
• Constructing a Hausdorf topology on Y , see Theorem 2.21.
To provide the orbit space of a group action with a suitable topological or differ-
entiable structure is always a subtle task. In the present context the problem is inten-
sified by the fact that the modular group does not operate freely, see Theorem 2.16:
There are elliptic points. The difficulty consist in dealing with the elliptic points
with respect to the Hausdorff property and with respect to complex charts on the
quotient.
−τ
0 −1
S := ∈ Γ , S(τ) = −1/τ = 2 .
1 0 |τ|
2.2 Topology of the orbit space of the Γ -action 47
• Translation by one:
1 1
T := ∈ Γ , T (τ) = τ + 1.
0 1
Φ : Γ ×H →
− H, (γ, τ) 7→ γ(τ),
D := {τ ∈ H : |τ| > 1 and −1/2 ≤ Re τ < 1/2}∪{τ ∈ H : |τ| = 1 and −1/2 ≤ Re τ ≤ 0},
3. The action Φ has exactly two elliptic points τ0 ∈ D. Their isotropy groups are
cyclic:
• τ0 = i with
Γi =< S >⊂ Γ , ord Γi = 4,
• τ0 = ρ := e2πi/3 with
Fig. 2.1 Fundamental domain D of the modular group Γ shaded, from [29]
Figure 2.1 shows the fundamental domain D of the modular group. In order to
exclude pairs of congruent points within D one has to exclude part of the boundary
of D. This convention is followed by [29] but not by all references. Most references
name fundamental domain the closure D.
Fig. 2.2 Some translates of the fundamental domain of Γ , from [52, Chap. VII, Fig. 1]
2.2 Topology of the orbit space of the Γ -action 49
Figure 2.2 shows some translates of the fundamental domain of Γ under the
elements T, S ∈ Γ and their products.
Proof. The three parts of the subsequent proof do not correspond one to one to the
three parts of the theorem. Note that the subgroup
{±id} ⊂ Γ
τ 0 := (S ◦ γ0 )(τ).
In both cases τ 0 ∈ D.
ii) Isotropy groups and fundamental domain: Consider two points τ, τ 0 ∈ D with
ab
τ 0 = γ(τ), γ = ∈Γ.
cd
We show τ = τ 0 , and the latter equation implies γ ∈ Γτ .
Im τ
Im τ 0 = .
|c · τ + d|2
implies
|c · τ + d| ≤ 1.
√
Because Im τ ≥ 3/2 only the cases c = 0, ±1 are possible:
• Case 1, c = 0: Then d = ±1 and a = d. Without restriction a = d = 1, hence
τ 0 = τ + b, b ∈ Z.
a·τ −1
τ0 = = a − (1/τ), a ∈ Z.
τ
Due to τ ∈ D the equation
|c · τ + d| = |τ| ≤ 1
1/τ = τ and τ 0 + τ = a ∈ Z
which implies
Re τ 0 + Re τ = a ∈ Z and Im τ 0 = Im τ.
Therefore: Either a = 0 and τ = i which implies
τ 0 = −(1/τ) = i = τ
and
0 −1
γ =± = ±S ∈ Γi .
1 0
Or a = −1 and τ = ρ which implies
τ 0 = −1 − (1/τ) = ρ = τ
and
−1 −1
γ =± = ±(ST )2 ∈ Γρ .
1 0
Case 2b, d = ±1: If d = ±1 then
2.2 Topology of the orbit space of the Γ -action 51
|c · τ + d| = |τ ± 1| ≤ 1.
a·ρ +b
ad − bc = a − b = 1 and τ 0 = = a + ρ,
ρ +1
hence a = 0, τ 0 = ρ = τ and
0 −1
γ= = ST ∈ Γρ .
1 1
τ0 := 2 · i ∈ D.
Due to part i) an element γ0 ∈ G exists with
(γ0 ◦ γ)(τ0 ) ∈ D.
τ0 = (γ0 ◦ γ)(τ0 ) ∈ D,
i.e.
γ0 ◦ γ ∈ Γτ0 = {±id}
which implies γ ∈ G, q.e.d.
Remark 2.17 (Presentation of the modular group and fundamental domain of its
subgroups).
1. Consider the element
−1 1
U := −T S = ∈ SL(2, Z)
−1 0
G =< S, U > .
52 2 The modular group Γ and its Hecke congruence subgroups Γ0 (N)
Therefore Theorem 2.16 implies that the two elements S and U generate the
group SL(2, Z). One can show that all relations of the generators are generated
by
S4 = U 3 = S2U · (US2 )−1 = id,
i.e. that
SL(2, Z) =< S, U; S4 ,U 3 , S2U · (US2 )−1 >
is a presentation of SL(2, Z), see [36, Kap. II, §2 Bemerk.].
2. Denote by
PSL(2, Z) := SL(2, Z)/{±id}
the corresponding projective-linear group, which acts faithful on H, and by
π : SL(2, Z) →
− PSL(2, Z), A 7→ A,
the canonical quotient map. The projective-linear group has the two cyclic sub-
groups
H1 :=< S >' Z/2Z and H2 :=< U >' Z/3Z.
One can show
2 3
PSL(2, Z) = H1 ∗ H2 =< S,U; S ,U >
as the free product, see the short argument from [3] and also [52, Chap. VII, § 1, Rem.].
of the induced K-action, which derives from the standard fundamental domain D
of Γ . In the following we will always choose fundamental domains D(K) ⊂ H
which originate as translates of D. These fundamental domains are measurable
subsets of H.
Φ : Γ0 (N) × H →
− H
has finitely many elliptic orbits. Each elliptic point τ ∈ H has the finite isotropy
group
Γ0 (N)τ = Γτ ∩ Γ0 (N).
2.2 Topology of the orbit space of the Γ -action 53
Proof. Theorem 2.16 shows: The action of Γ has exactly two elliptic points i, ρ ∈ D,
and both have finite isotropy groups. The index formula from Proposition 2.14 im-
plies the finiteness of
[Γ : Γ0 (N)] =: k.
According to Remark 2.17 each finite coset decomposition
k
[
Γ= γ j · Γ0 (N), γ j ∈ Γ ,
j=1
e.g. see Figure 2.3. For each j = 1, ..., k and each point τ ∈ γ −1 j (D) the group mor-
phism
− Γγ j (τ) , γ 7→ γ j · γ · γ −1
Γ0 (N)τ → j ,
injects Γ0 (N)τ into Γγ j (τ) . Hence the elliptic points of Γ0 (N) contained in D(Γ0 (N))
are contained in the finite set
k k
γ −1 γ −1
[ [
j (i) ∪ j (ρ), q.e.d.
j=1 j=1
Figure 2.3 shows the fundamental domain of Γ0 (13). The fundamental domain
D(Γ0 (13)) splits into 14 disjoint translates of D according to the index formula from
Proposition 2.14
[Γ : Γ0 (13)] = 14
For further explanations of Figure 2.3 see [17, Fig. 3.1 and Exerc. 3.1.4].
54 2 The modular group Γ and its Hecke congruence subgroups Γ0 (N)
Fig. 2.3 Fundamental domain of Γ0 (13), adapted from [17, Fig. 3.1]
The following Lemma 2.19 about the local behaviour of the action Φ covers the
main step for proving the Hausdorff property of the quotient topoplogy with respect
to the quotient map
H→ − Γ \H.
γ(U) ∩U 6= 0/ =⇒ γ ∈ Γτ .
Fig. 2.4 Local behaviour of the action of the modular group, see Lemma 2.19
Figure 2.4 visualizes Lemma 2.19: If an element γ ∈ Γ maps a point of the dis-
tinguished neighbourhood U1 of τ1 to the distinguished neighbourhhood U2 of τ2 ,
then γ maps τ1 to τ2 . The add-on states: If γ ∈ Γ does not fix τ, then γ moves all
points from U to the complement of U.
γ(V1 ) ∩V2 6= 0.
/
The idea is to consider the extension of the group action of Γ to SL(2, R), see
Remark 2.8,
aτ + b ab
Φ : SL(2, R) × H → H, (γ, τ) 7→ γ(τ) := , γ := ,
cτ + d cd
One checks
γ(i) = i ⇐⇒ a = d and b = −c,
hence
ab
∈ SO(2, R).
cd
The action Φ of SL(2, R) is transitive because any point τ ∈ H belongs to the
orbit of i: If
τ = x + iy
then γτ (i) = τ for the matrix
x
√
y √
y
= √1 · y x ∈ SL(2, R).
γτ :=
y 01
1
0 √
y
γ
e1 e2
γe−1
1
γe2
i i
γ(V 1 ) ∩V 2 6= 0/
and
2.2 Topology of the orbit space of the Γ -action 57
Because γe and its inverse γe−1 depend continuously on e ∈ H, the latter set on
the right-hand side is the continuous image of the compact set
V 2 × SO(2, R) ×V 1
G∩Γ
2. Shrinking V1 and V2 : According to the first part only finitely many γ ∈ Γ exist
with
γ(V1 ) ∩V2 6= 0.
/
In particular, the set
is finite. Therefore, after finitely many steps we can shrink the neighbourhoods V1
and V2 to neighbourhoods U1 and U2 with
τ = τ1 = τ2 .
fixes at most one point z ∈ H: If z ∈ H and γ(z) = z then z satisfies the quadratic
equation
cz2 + (d − a)z − b = 0
If
c 6= 0 or d 6= a or b 6= 0
the equation has at most two solutions, counted with multiplicity. The same
quadratic equation is also satisfied by z. Because z 6= z, the equation has no fur-
ther solution. Hence there is no
z0 ∈ H, z0 6= z,
γ ∈ Γτ 0 , γ 6= ±id.
Because
τ 0 ∈ γ(U) ∩U
the previous part of the proof implies
γ ∈ Γτ
Lemma 2.20 (Openness of the canonical projection onto the orbit space). De-
note by
p:H→ − Y := Γ \H
the canonical projection onto the orbit space of the action
Φ : Γ ×H →
− H
and provide Y with the corresponding quotient topology. Then p is an open map.
p−1 (U) ⊂ H
is open.
For each arbitrary but fixed γ ∈ Γ its orbit map, the restriction
Φ(γ, −) : H →
− H, τ 7→ γ(τ),
Φ(γ −1 , −) : H →
− H, τ 7→ γ −1 (τ).
In order to prove the openness of p we consider an open set U ⊂ H and prove the
openness of p(U) ⊂ Y : The inverse image
p−1 (p(U)) =
[
γ(U)
γ∈Γ
p(U) ⊂ Y
Theorem 2.21 (Hausdorff topology of the orbit space). The quotient topology on
the orbit space
Y := Γ \H
is Hausdorff.
p(τ1 ) 6= p(τ2 )
Lemma 2.20 implies that the sets p(U j ) ⊂ Y, j = 1, 2 are open. Hence they are
disjoint open neighbourhoods of respectively
With Theorem 2.21 the present section has obtained its first goal. Its second goal
is to embed the result into the general context: Which properties ensure that a set
of equivalence classes becomes a Hausdorff space when provided with the quotient
topology with respect to the canonical projection? We introduce the general concept
of a proper group action in Definition 2.22 and prove the Hausdorff criterion from
Lemma 2.24.
G × X → X, (g, x) 7→ g(x),
Note. The image of the map σ from Definition 2.22 is the equivalence relation
R ⊂ X ×X
which the group action induces on X.
Corollary 2.23 (Proper action of the modular group). The action of the modular
group
Φ : Γ ×H → H
is a proper group action.
Proof. Consider a compact subset
K ⊂ H × H.
σ −1 (K) ⊂ Γ × H
(γν , τν )ν∈N
2.2 Topology of the orbit space of the Γ -action 61
is convergent. Set
τν ∈ U1 and γν (τν ) ∈ U2 ,
hence
γν (U1 ) ∩U2 6= 0.
/
The lemma implies
γν (x1 ) = x2
The set
{γ ∈ Γ : γ(x1 ) = x2 }
has the same cardinality as the isotropy group Γx1 . The latter is a finite group due to
Theorem 2.16. A subsequence (γν )ν∈N becomes stationary, hence convergent, q.e.d.
p:X →
− X/R
p(x) 6= p(y)
p(U) ∩ p(V ) = 0,
/ q.e.d.
With the new concepts and results the proof of the Hausdorff property from
Theorem 2.21 reads as follows:
Corollary 2.25 (Hausdorff topology of the orbit space). The quotient topology on
the orbit space
Y := Γ \H
is Hausdorff.
Proof. We apply Lemma 2.24: First, Lemma 2.20 states that p is an open map.
Secondly, we have
R = σ (Γ × H) ⊂ H × H
According to Corollary 2.23 the map
σ : Γ ×H → H×H
is proper, and therefore a closed map. Therefore Lemma 2.24 implies: The orbit
space Y , provided with the quotient topology, is a Hausdorff space, q.e.d.
p−1 (y) ⊂ H, y ∈ Y,
The elliptic points of the group action are those points which complicate the intro-
duction of a complex structure on the orbit space Y . To overcome this difficulty we
make a detailed study of the group action in the neighbourhood of an elliptic point.
Fortunately, here the action restricts to the action of the isotropy group and the lat-
ter turns out as a group of rotations. The statement of the following Lemma 2.26 is
trivial for non-elliptic points.
For elliptic points it relies on the Lemma of Schwarz. Lemma 2.26 states: Locally
the isotropy group of an elliptic point τ ∈ H with period hτ acts as the group of
rotations of a disk around the origin with angles a multiple of 2π/hτ .
Lemma 2.26 (Isotropy group acting as group of rotations). For any point τ ∈ H
exist
• a neighbourhood Uτ ⊂ H of τ such that for all γ ∈ Γ
γ(Uτ ) ∩Uτ 6= 0/ =⇒ γ ∈ Γτ ,
γ˜τ := λτ ◦ γ ◦ λτ−1 : ∆τ → ∆τ
γ
Uτ γ(Uτ )
λτ λτ
γ˜τ
∆τ ∆τ
τ ∈ {i, ρ}.
Denote by
64 2 The modular group Γ and its Hecke congruence subgroups Γ0 (N)
z−τ
λτ : P1 → P1 , λτ (z) := ,
z−τ
the fractional linear automorphism of P1 with λ (H) = ∆ and
λτ (τ) = 0, λτ (τ) = ∞.
γ
H H
λτ λτ
γ˜τ
∆ ∆
The conjugate
γ̃τ := λτ ◦ γ ◦ λτ−1 : ∆ → ∆
has the fixed point 0 ∈ ∆ . The non-trivial isotropy groups are
because in particular
|S 0 (i)| = |(ST )0 (ρ)| = 1.
The Lemma of Schwarz implies that
γ̃τ : ∆ →
− ∆
is a rotation
γ̃τ (z) = ek·(2πi/hτ ) · z, k ∈ {0, 1, ..., hτ − 1}
We choose a disk ∆τ ⊂ ∆ such that the neighbourhood of τ ∈ H
Uτ := λτ−1 (∆τ )
satisfies the properties from Lemma 2.19. Then restricting the above digram proves
the claim, q.e.d.
In order to obtain a complex structure on the orbit space, the Hausdorff space
Y = Γ \H,
2.3 Modular curves X(Γ0 (N)) as compact Riemann surfaces 65
(1/2) · ord Γτ = hτ =: k.
z 7→ zk
foliates the k different segments of the disk ∆τ , which are eqivalent under the Γτ -action,
to an open set Wτ ⊂ C, see Figure 2.5, and induces a chart
φτ
p(Uτ ) −→ Wτ .
Fig. 2.5 The three equivalent segments of the disk ∆τ of an elliptic point τ ∈ H with hτ = 3
Proposition 2.27 (Analytic structure of the orbit space). On the orbit space
Y = Γ \H
exists an analytic structure such that Y becomes a Riemann surface and the canon-
ical projection
p:H→Y
becomes a holomorphic map.
Considered from a general point of view the complex manifold Y is the coarse
moduli space of complex tori.
66 2 The modular group Γ and its Hecke congruence subgroups Γ0 (N)
powτ : ∆τ → C, z 7→ zhτ ,
and set
Wτ := powτ (∆τ )
Eventually, define a chart of Y around y = p(τ) ∈ Y
φτ : p(Uτ ) → Wτ
p|Uτ
Uτ p(Uτ )
λτ φτ
powτ
∆τ Wτ
The map φτ is
• well-defined and bijective according to Lemma 2.26 because for all γ ∈ Γ \ Γτ
γ(Uτ ) ∩Uτ = 0.
/
The different inverse images under p|Uτ of a given point from p(Uτ ) are identi-
fied by powτ ◦ λτ ,
ii) Explicit form of the transition functions: We show that the family
is a complex atlas of Y .
p(Uτ1 ) ∩ p(Uτ2 ) 6= 0.
/
2.3 Modular curves X(Γ0 (N)) as compact Riemann surfaces 67
Consider a point
x ∈ p(U1 ) ∩ p(U2 ),
see Figure 2.6.
68 2 The modular group Γ and its Hecke congruence subgroups Γ0 (N)
φ2,1 := φ2 ◦ φ1−1
We have to prove: Taking the h1 -root does not prevent the holomorphy of the map.
iii) Holomorphy of the transition functions: We prove the non-trivial case h1 > 1,
i.e. we assume that the orbit x ∈ X is elliptic.
• Case φ1 (x) = 0 ∈ W1 , i.e. τ̃1 = τ1 :
The point
τ1 = τ̃1
is elliptic. Therefore also τ̃2 = γ(τ˜1 ) ∈ U2 is elliptic. Because τ2 is the only
elliptic point in U2 we obtain τ˜2 = τ2 , and
τ2 = γ(τ1 ) and h := h2 = h1 .
→ P1 , z 7→ (z : 1),
C ,−
– while the vector (1 0)> ∈ C2 represents the point (1 : 0) ∈ P1 and the point
∞ ∈ C ∪ {∞}.
As a consequence
a0
A=
0d
and
h h
φ2,1 (q) = (λ2 ◦ γ ◦ λ1−1 )(q1/h ) = A(q1/h ) = ((a · q1/h )/d)h = (a/d)h · q.
We reduce this case to the other two cases. Choose τ3 ∈ D with p(τ3 ) = x and
consider the chart
φτ3 : Uτ3 → Wτ3
with the shorthand
U3 := Uτ3 , φ3 := φτ3 ,W3 := Wτ3 .
In a suitable neighbourhood W of x with
we have
φ1,2 = φ1,3 ◦ φ3,2 .
The right-hand side is holomorphic due to part i) and ii). Therefore φ1,2 is
holomorphic, q.e.d.
The important steps in the proof of Proposition 2.27 and its prerequisites:
The upper half-plane has a canonical boundary point, the point ∞. The extension of
the group action from Remark 2.8 shows that one has to consider also the rational
line Q, but not R \ Q, as part of the boundary.
When extending the Euclidean topology of H to a topology on H∗ , one has to
define the neighbourhoods in H∗ of the additional points from
Q ∪ {∞}.
∞ ∈ P1
from the Riemann sphere, alias projective space. Instead one takes the sets
i.e. the complements of sets with bounded imaginary part. Secondly, one translates
neighbourhoods of ∞ to neighbourhoods of a rational q ∈ Q by fractional-linear
transformations γ ∈ Γ with
γ(∞) = q.
The resulting neighbourhood basis of a point q ∈ Q are disks in H with q a point on
the Euclidean boundary of the disk, see Figure 2.7. The second step is determined
by the requirement, that the modular group acts as group of homeomorphisms
on H∗ . The new topology of H∗ induces on Q the discrete topology. Apparently the
latter is finer than the topology induced on Q by the Euclidean topology of R.
Fig. 2.7 Neighbourhoods of infinity and of some rational points from H∗ , from [17, Fig. 2.5]
Theorem 2.28 (The modular curve of the modular group). Consider the action
of the modular group Γ on the extended upper half-plane
72 2 The modular group Γ and its Hecke congruence subgroups Γ0 (N)
H∗ := H ∪ Q ∪ {∞}.
H H∗
p p∗
Y X
q := a/c ∈ Q, gcd(a, c) = 1,
the sets
a∗
γ(UR ), R > 0, γ = ∈Γ,
c∗
because
γ(∞) = q
due to Remark 2.8. These neighbourhood bases together with the open sets of H
generate a topology on H∗ with H ⊂ H∗ an open subspace.
Note. Because fractional linear transformations from Γ take lines to circles or lines
the neighbourhoods of γ(UN ) are circles tangent to the real axis at the distinguished
point a/c ∈ Q. Hence neighbourhoods of ∞ and of rational points are completely
different from their counterpart with respect to the Euclidan topology, see
Figure 2.7.
X := Γ \H∗
p∗ : H∗ → X.
Then p∗ is an open map: The proof is the same as the corresponding proof for
Lemma 2.20. The proof uses only the fact that the quotient arises from a group
action.
the formula
Im τ Im τ
Im γ(τ) = 2
= ≤
|c · τ + d| (c · Re τ + d)2 + c2 · (Im τ)2
1
c=0
2 c=0 Im τ
≤ Im τ · d ≤ ≤ max{Im τ, 1/Im τ}
1 1 c 6= 0
c 6= 0
2
c · (Im τ)2 Im τ
shows
Im γ(τ) ≤ max{Im τ, 1/(Im τ)},
independently from γ. Therefore a relatively compact neighbourhood
U1 ⊂⊂ H
of τ1 and a neighbourhood
U2 ⊂ H∗
of τ2 = ∞ exist such that for all γ ∈ Γ
γ(U1 ) ∩U2 = 0.
/
As a consequence
p∗ (U1 ) ∩ p∗ (U2 ) = 0.
/
• τ1 , τ2 ∈ H: This case has been considered already by Corollary 2.25.
H
iii) Compactness of X: Denote by D ⊂ H the closure of D with respect to H.
Then
74 2 The modular group Γ and its Hecke congruence subgroups Γ0 (N)
D := D ∪ {∞} ⊂ H∗
H
D ⊂ H∗ .
One has
Y = X \ {p∗ (∞)}.
Therefore Y ⊂ X is an open subset.
U∞ := U2 = {τ ∈ H : Im τ > 2} ∪ {∞}.
We define (
e2πi·τ 6 ∞
if τ =
pow∞ : U∞ → ∆ , τ →
7
0 if τ = ∞
and set
W∞ := pow∞ (U∞ ) ⊂ ∆ .
For two points τi ∈ U∞ , i = 1, 2,
p∗ |U∞
U∞ p∗ (U∞ )
pow∞ φ∞
W∞
One shows analogously to the proof of Proposition 2.27 that the map φ∞ is a home-
omorphism.
Eventually we have to show that the charts of Y and the chart φ∞ are compatible, It
suffices to prove the compatibility of the chart
φ∞ : p∗ (U∞ ) → W∞
φτ : p(Uτ ) = p∗ (Uτ ) → Wτ
around an arbitrary point p(τ), τ ∈ H. We may assume that Uτ does not contain an
elliptic point, and therefore that
φτ ◦ (p∗ |Uτ )
is injective and biholomorphic. Hence the transition function of the two charts is
holomorphic.
X ' P1
a formal proof will be given in Corollary 3.16. This second proof uses the modular
invariant j from the theory of modular forms.
76 2 The modular group Γ and its Hecke congruence subgroups Γ0 (N)
Definition 2.29 (Modular curve of Γ ). The compact Riemann surface from Theo-
rem 2.28
X(1) := X(Γ ) := Γ \H∗
is named the modular curve of Γ or simply the modular curve.
Remark 2.30 (Modular curves of the congruence subgroups). Generalizing the con-
struction of the modular curve
X(1) = Γ \H∗
from Theorem 2.28 one can introduce for each Hecke congruence subgroup
Γ0 (N) ⊂ Γ , N ∈ N,
see [17, Prop. 2.4.2ff.] The resulting compact Riemann surface X0 (N) is named the
modular curve of Γ0 (N).
→ Γ (1)
Γ0 (N) ,−
f : X0 (N) →
− X(1)
b( f )
g(X0 (N)) = 1 + − deg f
2
with
b( f ) = ∑ b( f ; x)
x∈X0 (N)
A detailed investigation of the elliptic points and of the cusps of the action
of Γ0 (N) shows the genus formula
deg f ε2 ε3 ε∞
g(X0 (N)) = 1 + − − −
12 4 3 2
with ε2 , ε3 and ε∞ denoting the number of elliptic points of respectively period 2
and 3 and the number of cusps of the action of Γ0 (N), see [17, Theor. 3.1.1].
ε∞ = ∑ Φ(gcd(d, N/d)),
d|N
see Example 2.11, and ε2 and ε3 are the numbers of solutions in Z/NZ of the
equations respectively
x2 + 1 = 0 and x2 − x + 1 = 0.
Using the quadratic residue symbol for n ∈ {−1, −3} and extended to all primes p
0 if p = 2
−1
:= 1 if p ≡ 1 mod 4
p
−1 if p ≡ 3 mod 4
and
0
if p = 3
−3
:= 1 if p ≡ 1 mod 3
p
−1 if p ≡ 2 mod 3
one obtains, see [17, Cor. 3.7.2, Exerc. 3.7.6] and [53, Theor. 1.43]:
−1
∏ p|N 1 + p if 4 - N
ε2 =
0 otherwise
and
78 2 The modular group Γ and its Hecke congruence subgroups Γ0 (N)
−3
∏ p|N 1 + p if 9 - N
ε3 =
0 otherwise
3. Estimate of the genus g(X0 (N)): As a corollary of the genus formula the constants
from part 2) satisfy the estimate
ε2 2ε3
2 · g(X0 (N)) − 2 + + + ε∞ > 0.
2 3
4. Modular curves of prime level: For a prime p ∈ N and k := p + 1 the modular
curve X0 (p) has the genus
$ %
k
− 1 if k ≡ 2 mod 12
12
g(X0 (p)) =
$ %
k
12 otherwise
In particular,
5. Numerical example: See the PARI-file “modular curve as covering”. For the
computation of the extended quadratic residue symbol PARI provides the com-
mand
kronecker(a,p).
The Kronecker symbol is defined for arbitrary integer a ∈ Z and arbitrary primes
including p = 2 by extending the Legendre symbol:
• Odd prime p:
a
kronecker(a, p) := for odd p
p
• Even prime p = 2:
(
0 if a even
kronecker(a, 2) := 2
(−1)(a −1)/8 otherwise
• Even prime p = 2:
−1 −3
kronecker(−4, 2) = 0 = and kronecker(−3, 2) = −1 =
2 2
Hence
∏ p|N (1 + kronecker(−4, p))
if 4 - N
ε2 =
0 otherwise
and
∏ p|N (1 + kronecker(−3, p))
if 9 - N
ε3 =
0 otherwise
X0 (p) →
− X(1)
2. On the other hand, they expand into a Fourier series around the point ∞ with
rational Fourier coefficients after normalization.
The interplay of these two properties is coined “The magic of modular forms” [65].
Therefrom arises a series of remarkable identities between the Fourier coefficients
of modular forms. In many cases these identities imply interesting statements about
arithmetic functions. We will see examples in Section ?? and Chapter 6.
81
82 3 The algebra of modular forms
We start from the general context of representation theory of the group Γ , i.e. we
introduce a Γ -action on certain vector spaces of holomorphic differential forms.
W 1 := H 0 (H, ΩH1 )
holds
dτ
Φ 1 (ω, γ)(τ) := (γ ∗ ω)(τ) = ( f ◦ γ)(τ)
(cτ + d)2
2. For even k ∈ N set
⊗k/2
Ω k/2 := ΩH and W k/2 := H 0 (H, Ω k/2 )
and
Φ k/2 : W k/2 × Γ → W k/2 , (ω, γ) 7→ Φ k/2 (ω, γ) := γ ∗ ω.
For
ab
γ= ∈ Γ , ω = f (dτ)k/2 ∈ W k/2 and τ ∈ H
cd
holds
(dτ)k/2
Φ k/2 (ω, γ)(τ) := ((γ ∗ )k/2 ω)(τ) = ( f ◦ γ)(τ)
(cτ + d)k
3. For even k ∈ N,
ab
γ= ∈ Γ and ω = f (dτ)k/2 ∈ W k/2
cd
3.1 Modular forms and cusp forms 83
1
γ ∗ ω = ω ⇐⇒ f (τ) = ( f ◦ γ)(τ) for all τ ∈ H.
(cτ + d)k
Note. In the context of modular forms Ω k/2 denotes the tensor product of Ω 1 not
the exterior product.
Proof. 1. We have
Moreover
!
aτ + b a(cτ + d) − (aτ + b)c 1
(dγ)(τ) = d = 2
dτ = dτ
cτ + d (cτ + d) (cτ + d)2
Therefore
dτ
(Φ 1 ( f dτ, γ))(τ) := (γ ∗ ( f dτ))(τ) = f (γ(τ)) (dγ)(τ) = ( f ◦ γ)(τ)
(cτ + d)2
2. For even k > 2 the proof is analogous by applying the derivation separately for
each tensor factor dτ, q.e.d.
Proposition 3.1 shows: For even k ∈ N the invariant differential forms from the
fixed space of the Γ -representation
f dτ k/2 ∈ (W k/2 )Γ
are characterized by holomorphic coefficient functions f ∈ O(H) satisfying
1
f (τ) = ( f ◦ γ)(τ) ·
(cτ + d)k
for all
ab
γ= ∈ Γ and τ ∈ H.
cd
If these coefficent functions extend holomorphically to ∞ ∈ H∗ they will be named
modular functions of weight = k. The factors
1
(cτ + d)k
from their transformation law will be isolated and termed factors of automorphy,
see Definition 3.3.
84 3 The algebra of modular forms
1
= 1.
cτ + d
Hence
and the latter equality means that the function f has the period = 1. Hence f expands
into a convergent Fourier series
∞
f (τ) = fˆ(q) = ∑ an · qn , q = e2πi·τ ,
n=−∞
1 2πi·z
+ e−2πi·z .
cos(2πz) = e
2
The holomorphic map
H → ∆ ∗ , τ 7→ e2πiτ ,
surjects to the punctured unit disk
f
H C
fˆ
∆∗
Set
∞
fˆ(q) := f (τ) = ∑ an · qn , q = e2πi·τ .
n=−∞
3.1 Modular forms and cusp forms 85
f (∞) := fˆ(0).
which extends the group action of the modular group Γ to GL(2, R)+ . The map
V := { f : H →
− C}
V × GL(2, R)+ →
− V, ( f , γ) 7→ f [γ]k ,
with
f [γ]k (τ) := f (γ(τ)) · h(γ, τ)−k · (det γ)k−1 , τ ∈ H,
is a right group action of GL(2, R)+ on the vector space V .
In the definition of f [γ]k from Lemma 3.4 the exponent of det γ is not uni-
form in the literature. We use the exponent (k − 1) from [17, Exerc. 1.2.11],
while [35, Chap. III, §3] uses the exponent k/2. The choice of the exponent (k − 1)
will later simplify Definition 5.10.
Note that in Lemma 3.4 the group GL(2, R)+ is written on the right-hand side of
the vector space V .
Left-hand side:
τ A(τ) B(A(τ))
B· A· = B· · h(A, τ) = · h(B, A(τ)) · h(A, τ).
1 1 1
Equating with the right-hand side gives for the lower row
Remark 3.5 (Meromorphy and holomorphy at the cusps). The Fourier expansion
from Definition 3.2 is well-defined for holomorphic functions f : H →
− C wich are
invariant with respect to the translation T . The modular group Γ has only a sin-
gle cusp, the orbit comprising ∞ and all rationals q ∈ Q. But Hecke congruence
subgroups Γ0 (N), N ≥ 2, have several cusps, each comprising a different subset of
rationals q ∈ N.
a
Fix a Hecke congruence subgroup Γ0 (N), a weight k ∈ Z, and a rational q = ∈ Q.
c
How to extend a holomorphic function
f :H→
− C
with
f [γ]k = f
for all γ ∈ Γ0 (N) to the point q ∈ Q in a meromorphic or holomorphic way?
− PSL(2, Z) := Γ /{±1}, α 7→ α.
Γ→
then
Γ0 (N)q ' α −1 · Γ0 (N)q · α ⊂ Γ∞
or
−1
Γ0 (N)q ' α · Γ0 (N)q · α ⊂ Γ ∞ ' Z
For
ab N −1 ∗ ∗
α= define γ := α · T · α = ∈ Γ0 (N)q
cd −Nc2 ∗
Then γ ∈ Γ0 (N)q . Because < T N > ⊂ Γ∞ is a subgroup of finite index, the same
holds for
< γ > ⊂ Γq and for < γ > ⊂ Γ q
We obtain
1 ≤ hq := [Γq : Γ0 (N)q ] = [Γ q : Γ0 (N)q ] < ∞
The fundamental object of the whole theory is the modular form. We intro-
duce modular forms of the modular group Γ respectively its Hecke congruence
subgroups Γ0 (N) as the coefficient functions of the invariant differential forms from
Proposition 3.1 which extend to ∞ ∈ H∗ . It turns out that modular forms are mero-
morphic differential forms on the modular curve X0 (N), see Theorem 3.17.
f (∞) = 0
• On denotes by
Ak (Γ ) ⊃ Mk (Γ ) ⊃ Sk (Γ )
the vector spaces of automorphic forms with respect to Γ of weight k, the
subspace of modular forms, and the subspace of cusp forms.
f [γ]k = f ,
We denote by
Ak (Γ0 (N)) ⊃ Mk (Γ0 (N)) ⊃ Sk (Γ0 (N))
the complex vector space of respectively automorphic forms, modular forms, and
cusp forms of weight k with respect to Γ0 (N).
Definition 3.6, part 2) and 3) generalize part 1): For γ ∈ Γ holds det γ = 1, hence
1
f [γ]k (τ) = ( f ◦ γ)(τ) ·
(cτ + d)k
Note that Definition 3.6 uses the weight convention from [17, Chap. 1.1]. Some
sources use a different weight convention. In the literature also the notation to dis-
criminate between modular functions and modular forms is not uniform.
Remark 3.7 (Modular forms of odd weight). A modular form of a congruence sub-
group Γ0 (N), N ∈ N, of odd weight k ∈ Z satisfies for all τ ∈ H and for the particular
element
−1 0
γ= ∈ Γ0 (N)
0 −1
the equation
We will now construct for the modular group Γ = SL(2, Z) explicit modular
forms Gk of even weight k ≥ 4 and a distinguished holomorphic function G2 , which
behaves in a different way under the Γ -action. To achieve the construction we
need the convergence of certain infinite series. Lemma 3.8 provides the necessary
prerequisites.
3.2 Eisenstein series and the algebra of modular forms of Γ 91
satisfies the prerequisites of the rearrangement theorem, i.e. there exists a con-
stant K such that for all finite subsets A, B ⊂ N∗ holds
∑ d · |q|md ≤ K
m∈A,d∈B
is absolutely convergent.
• Fourier series
cos(πτ) 1+q ∞
π · cot(πτ) = π = (−iπ) · = −πi − 2πi ∑ qm .
sin(πτ) 1−q m=1
1 1
cos πτ = (eiπτ + e−iπτ ) and sin πτ = (eiπτ − e−iπτ )
2 2i
which imply
1+q ∞ ∞
= (1 + q) ∑ qm = 1 + 2 ∑ qm
1−q m=0 m=1
1 (−2πi)k ∞ k−1 m
∑ k
= ·∑m q .
m∈Z (τ + m) (k − 1)! m=1
Hence
1 1 1
∑ (τ + n)k = (−1)k · ∑ (−τ − n)k = (−1)k · ∑ (−τ + n)k
n∈Z n∈Z n∈Z
1 (−2πi)k
(−1)k · ∑ (−τ + n)k = · (−1)k · ∑ rk−1 · q−r
n∈Z (k − 1)! r=1
lim |1 − qm |2 = 1.
m→∞
Set
∞ qm
K := ∑
m=1 (1 − qm )2
As a consequence
∑ d · qmd ≤ K.
m∈A,d∈B
Hence w.l.o.g. we may assume q real with 0 ≤ q < 1, which proves the claim.
The latter series is convergent: For k ≥ 4 any rearrangement of the original double
series is permitted. For k = 2 the result of part 2) allows to apply the rearrange-
ment theorem to the double series
∞
∑ r · qrm , |q| < 1.
m,r=1
Gk (∞) = 2ζ (k).
Here
∞
ζ (s) := ∑ (1/ns ), Re s > 1,
n=1
(2πi)k ∞
Gk (τ) = 2ζ (k) + 2
(k − 1)! ∑ σk−1 (n) · qn , q = e2πiτ .
n=1
Gk (τ) := GΛτ ,k
The series GΛ ,k is absolute convergent due to Lemma 1.11, hence allows arbitrary
rearrangement. With respect to the translation
T : H → H, τ 7→ T (τ) := τ + 1,
we obtain
Gk (τ) = Gk (τ + 1).
With respect to the reflection
−1
S : H → H, τ 7→ S(τ) :=
τ
we have
1 1 0 1
Gk (S(τ)) · h(S, τ)−k = · Gk (−(1/τ)) = k · ∑ = Gk (τ).
τk τ (c,d)∈Z2 (−c(1/τ) + d)k
Because S and T generate the modular group, Lemma 3.4 implies for all γ ∈ Γ and
for all τ ∈ H
Gk (τ) = Gk (γ(τ)) · h(γ, τ)−k , i.e. Gk = Gk [γ]k
0 1 1 1
Gk (τ) = ∑ = ∑ k+ ∑ =
(c,d)∈Z2
(cτ + d)k d∈Z∗ d c∈Z∗ ,d∈Z (cτ + d)k
2 (2πi)k ∞
= 2ζ (k) + · ∑ σk−1 (n) qn , q = e2πi·τ .
(k − 1)! n=1
The Fourier expansion, a convergent power series, proves the holomorphy of Gk in
H and at the point ∞, q.e.d.
In order to get rid of the transcendent values in the Fourier expansion of Gk one
divides by the value of the Riemann ζ -function.
Definition 3.10 (Normalized Eisenstein series). For even k ≥ 4 one defines the
normalized Eisenstein series
Gk
Ek := ∈ Mk (Γ )
2ζ (k)
Corollary 3.11 (Normalized Eisenstein series). For each k ≥ 4 all Fourier coeffi-
cients of the normalized Eisenstein series are rational, more precisely
2k ∞
Ek (τ) = 1 − · ∑ σk−1 (n) · qn ∈ Mk (Γ ), q = e2πi·τ
Bk n=1
Proof. 1. Euler relation of Bernoulli numbers: See [20, Kap. VII, Satz 4.1]. We
relate the cotangent function to the Bernoulli numbers (Bn )n∈N . The Bernoulli
numbers are the generators of the power series
z ∞ Bn n
=: ∑ ·z .
ez − 1 n=0 n!
We compute
imply !µ !
2
∞1 ∞ z2
πz · cot(πz) = 1 + 2z · ∑ − 2 · ∑ =
ν=1 ν µ=0 ν 2
!
∞ ∞ 1
= 1 − 2 ∑ ∑ 2µ z2µ
µ=1 ν=1 ν
πz · cot(πz)
Gk (τ) 2k ∞
Ek (τ) := = 1 − · ∑ σk−1 (n) · qn ∈ Mk (Γ ).
2ζ (k) Bk n=1
All Bernoulli numbers are rational due to the recursion formula for N ∈ N∗
N
N +1
∑ n · Bn = 0 with B0 = 1.
n=0
one has
(2, π 2 /6, 1/6, −24)
(4, π 4 /(2 · 32 · 5), −1/30, 240)
(6, π 6 /(33 · 5 · 7), 1/42, −504)
(8, π 8 /(2 · 33 · 52 · 7), −1/30, 480)
(10, π 10 /(35 · 5 · 7 · 11), 5/66, −264)
(12, 691 · π 12 /(36 · 53 · 72 · 11 · 13), −691/2730, 65520/691)
For k ∈ {4, 6, 8} the numerical value of Bk implies that all Fourier coefficients
of Ek are not only rational but are integers.
2. Weight k ≥ 4: For even k > 2 the Eisenstein series Gk relate to the lattice
constants GΛ ,k of an arbitrary lattice with positively oriented basis
Λ = Zω1 + Zω2
1
GΛ ,k = Gk (τ).
ω1k
0 1
∑ (mτ + n)2
(m,n)∈Z2
3.2 Eisenstein series and the algebra of modular forms of Γ 99
is not absolutely convergent. Anyhow, in both cases some specific orders of sum-
mation provide certain finite values. In the 2-dimensional case we define the
Eisenstein series
!
∞ 1 1
G2 (τ) := 2 · ∑ 2 + ∑ ∑ 2
, τ ∈ H.
n=1 n m∈Z∗ n∈Z (mτ + n)
Lemma 3.8, part 3) implies that the sum on the right-hand side is convergent with
Fourier series
∞
G2 (τ) = 2 · ζ (2) − 8π 2 · ∑ σ1 (n) · qn , q = e2πi·τ .
n=1
The equality
ζ (2) = π 2 /6
implies for the normalized Eisenstein series of weight k = 2
G2 (τ) ∞
E2 (τ) := = 1 − 24 · ∑ σ1 (n) · qn , q = e2πi·τ .
2ζ (2) n=1
Proposition 1.19 dealt with the discriminant of the cubic polynomial from the dif-
ferential equation of the Weierstrass function ℘ of a lattice. Definition 3.13 shows:
These discriminants are the values of a modular form.
Definition 3.13 (Discriminant form). The discriminant form is the modular form
of Γ of weight k = 12 defined as
∆ := g32 − 27 · g23 : H → C
100 3 The algebra of modular forms
g2 := 60 · G4 , g3 := 140 · G6 .
In Definition 3.13 the letters g2 and g3 designate modular functions, not the con-
stants from Theorem 1.17, which denote the value at the period τ of a given normal-
ized lattice
Λ = Z · 1 + Z · τ.
Proposition 3.14 (Fourier coefficients of the dicriminant form). All Fourier co-
efficients of the normalized discriminant form
∆
∈ M12 (Γ )
(2π)12
are integers:
∆ (q) ∞
12
= ∑ τn · qn ∈ Z{q}, q = e2πi·τ ,
(2π) n=0
with τ0 = 0 and τ1 = 1.
In particular ∆ ∈ S12 (Γ ) is a cusp form of weight 12, and the zero at ∞ ∈ H∗ has
order = 1.
G4 ∞
E4 = 32 · 5 · 4
= 1 + 240 · ∑ σ3 (n) · qn
π n=1
33 · 5 · 7 G6 ∞
E6 = · 6 = 1 − 504 · ∑ σ5 (n) · qn .
2 π n=1
We have
22 4 23
g2 = 60 · G4 = π · E4 and g3 = 140 · G6 = 3 π 6 · E6 .
3 3
Hence
26 3 3
12 E4 − E6
2 3 2
12 E4 − E6
∆ = g32 − 33 g23 = π 12 · (E − E6
2
) = (2π) · = (2π) ·
33 4 26 · 33 1728
3.2 Eisenstein series and the algebra of modular forms of Γ 101
and eventually
∆ E43 − E62 E43 − E62
= = .
(2π)12 26 · 33 1728
We set as shorthand
∞
Xk (q) := ∑ σk (n) · qn
n=1
and obtain
∆ (1 + 240 · X3 )3 − (1 − 504 · X5 )2
=
(2π)12 26 · 33
When denoting by
Q = Q(q)
the numerator of the latter fraction, then
we get
Q(q) = 3 · 240 · q + 1008 · q + O(2) = 1728 · q + O(2)
which proves τ0 = 0 and τ1 = 1.
5 · X3 + 7 · X5
5 · d3 + 7 · d5 ≡ 0 mod 12
According to the Chinese remainder theorem the claim reduces to the two congru-
ences
5 · d 3 + 7 · d 5 ≡ 0 mod 3 and 5 · d 3 + 7 · d 5 ≡ 0 mod 4
102 3 The algebra of modular forms
or
d 5 ≡ d 3 mod 3, mod 4.
The latter two congruences reduce to
• If d 3 6≡ 0 mod 3 then d 2 ≡ 1 mod 3,
Definition 3.15 (Modular invariant j). The modular invariant j is the automor-
phic function, defined as the quotient of two modular forms of weight = 12,
g32 g3
j := 1728 · = 1728 · 3 2 2 .
∆ g2 − 27g3
Corollary 3.16 (Modular curve and modular invariant j). The modular invari-
ant is a biholomorphic map
'
→ P1
j : X(1) −
which maps the cusp of X(1) to ∞ ∈ P1 .
∆
(2π)12
G4 (∞) = 2ζ (4) 6= 0
from Theorem 3.9 show that j has a pole of order = 1 at the point ∞. Because j is
even an automorphic function, it is a meromorphic function on the modular curve,
i.e. a holomorphic map
'
→ P1 .
j : X(1) −
Because all fibres of j have the same multiplicity = 1, the map j is biholomorphic,
q.e.d.
3.2 Eisenstein series and the algebra of modular forms of Γ 103
Continuing with the investigation started in Proposition 3.1 we now set out to
show: Modular forms of weight k are those meromorphic differential forms on the
modular curves X0 (N) which are multiples of a certain divisor which depends on
the level N and the weight k. Hence modular forms are the holomorphic sections of
certain line bundles on the modular curves. Applying the Riemann-Roch theorem
we draw several conclusions. In detail:
deg OD := deg D.
deg K = 2g(X) − 2.
dim H 0 (X, OD ) = 0.
• We derive a formula for the dimension of the vector spaces Mk (Γ ). Section 3.3
will generalize the computation to Mk (Γ0 (N)).
• Chapter 4 will show: Any complex elliptic curve arises from a complex torus.
More precisely, the elliptic curve is parametrized by the two meromorphic func-
tions ℘Λ and ℘Λ0 defined on a suitable torus C/Λ , see Theorem 4.28.
p : H∗ → X := Γ \H∗ ,
skipping the ∗-superscript used in the notation from Theorem 2.28. We introduce
the class notation
[τ] = p(τ) ∈ X.
In addition we use the floor function: For x ∈ R
Theorem 3.17 (Modular forms are meromorphic differential forms on the modular
curve). Consider an even weight k ≥ 0.
1. Differential forms: The map of vector spaces is injective
α : Mk (Γ ) → H 0 X, M ⊗O Ω ⊗k/2 , f 7→ ω f ,
The attachment
U 7→ ΩΓ ,k (U), U ⊂ X open,
with
n o
ΩΓ ,k (U) := ω ∈ H 0 (U, M ⊗O Ω ⊗k/2 ) : ω 6= 0 and div ω ≥ −Dk |U ∪ {0}
ΩΓ ,k ⊂ M ⊗ O Ω ⊗k/2
LΓ ,k := O(k/2)·K+Dk ∈ Pic(X)
3. Meromorphic differential forms on the modular curve: The image of the map α
from part 1) is
α(Mk (Γ )) = H 0 (X, ΩΓ ,k ).
Proof. Cf. [50, §4]. The numbers of the different parts of the proof do not corre-
spond one-to-one to the sections of the theorem.
3.2 Eisenstein series and the algebra of modular forms of Γ 105
f (τ) (dτ)k/2 := p∗ ω f :
One considers first the form ω f outside the three points which are the orbits of
the two elliptic points and the cusp. Secondly one extends ω f holomorphically into
these exceptional points by using Riemann’s extension theorem for bounded holo-
morphic functions.
q:U →
− V ⊂C
and represent
ω|U = f · (dq)k/2
Then
ordx ω := ordx f ,
independently from the choice of the chart. To prove the order relation for t ∈ {i, ρ, ∞}
we use the local form of
p : H∗ → X
and the charts around t from Theorem 2.28 and Proposition 2.27. We denote by τ
points from Ut ⊂ H and by q points from Wt ⊂ C:
• For t = ∞ ∈ H∗ we recall the chart pow∞ around t and the chart φ∞ around the
cusp [∞] from the diagram
106 3 The algebra of modular forms
p|U∞
U∞ p(U∞ )
pow∞ φ∞
W∞
For τ ∈ U∞ (
e2πiτ τ=6 ∞
q(τ) := pow∞ (τ) :=
0 τ =∞
We denote by τ the points on U∞ ⊂ H and by q the points on W∞ ⊂ ∆ ⊂ C. Then
which implies
1 (dq)k/2
(dτ)k/2 = ·
(2πi)k/2 qk/2
The order relation relates
to
ord[∞] ω f = ord((φ∞−1 )∗ ω f ; q = 0) = ord(h; q = 0)
with
h(q) (dq)k/2 := (φ∞−1 )∗ ω f
The commutative diagram
On one hand
p∗ ω f = f (τ) (dτ)k/2
On the other hand,
pow∗∞ (φ∞−1 )∗ ω f = pow∗∞ h(q) (dq)k/2 = h(q(τ)) · (2πi)k/2 · q(τ)k/2 (dτ)k/2
Hence
f (τ) = h(q(τ)) · (2πi)k/2 · q(k(τ))k/2
and
fˆ(q) = h(q) · (2πi)k/2 · qk/2
which implies
ord( fˆ; q = 0) = ord(h; q = 0) + k/2
3.2 Eisenstein series and the algebra of modular forms of Γ 107
or
ord∞ f = ord[∞] ω f + k/2
p|Ut
Ut p(Ut )
λt φt
powt
∆t Wt
The map
'
λt : Ut −
→ ∆t
is biholomorphic. For the order calculation we assume w.l.o.g.
λt = id and Ut = ∆t
e := ht
For τ ∈ Ut
q(τ) := powt (τ) := τ e
Therefore
If
ω f (q) = g(q) (dq)k/2
then
As a consequence
f (τ) = g(q(τ)) · τ (k/2)·(e−1) · ek/2
By definition
ord[t] ω f = ord0 g
Hence
Summing up
f (τ) (dτ)k/2 = p∗ ω
iii) Line bundle: It is well known that the sheaves of multiples of a divisor are line
bundles. The sheaf of multiples of the sum of two divisors is the tensor product of
the sheaves of multiples of each summand, see [63], q.e.d.
The formulas in the proof of Theorem 3.17, part 2) show: While a modular
form f on H∗ is holomorphic when considered as a function, the representing dif-
ferential form ω f on the modular curve X may have poles. Actually, it is the divisor
of ω f which encodes the information about f .
In the language of line bundles, modular forms of weight = k are the sections of
the line bundle LΓ ,k on the modular curve:
Mk (Γ ) ' H 0 (X, LΓ ,k )
3.2 Eisenstein series and the algebra of modular forms of Γ 109
S2 (Γ ) = 0.
ω f := α( f ) ∈ H 0 (X(1), ΩΓ ,k )
ord[τ] ω f = ord[τ] f ≥ 0,
in particular for k = 2:
The calculation from Example 3.18 carries over to arbitrary even weights k ∈ Z.
110 3 The algebra of modular forms
For the dimension of the vector space Sk (Γ ) of cusp forms see Lemma 3.23.
Proof. We have
Mk (Γ ) ' H 0 (X, LΓ ,k ).
For k = 0 we have on the compact Riemann surface X ' P1
• Residue β = 0:
1 − 6α + 4α + 3α = 1 + α = 1 + bk/12c
• Residue β = 2:
= 1−6α −β /2+4α +bβ /3c+3α +bβ /4c = 1+α +(−β /2+bβ /3c+bβ /4c) =
= 1 + α = 1 + bk/12c
q.e.d.
In particular
dim M0 = 1, dim M2 = 0, dim M4 = dim M6 = dim M8 = dim M10 = 1, dim M12 = 2.
The non-zero vector spaces in the last line are spanned by the Eisenstein series G4 , G6
and their products. Only M12 contains a non-zero cusp form, namely the discrimi-
nant modular form ∆ .
Apparently the numerical value dim H 0 (X, LΓ ,k ) can be obtained without applying
the theorem of Riemann-Roch and Serre duality in the proof of Corollary 3.19: On
the modular curve X ' P1 each line bundle L is determined by its degree m, and
for m ≥ 0
dimH 0 (P1 , L ) = m + 1.
We gave the proof in the above form, because Theorem 3.26 will generalize the
proof to the modular curves X0 (N) of the Hecke congruence subgroups.
Corollary 3.20 (Weight formula for modular forms). Consider an even integer k ≥ 4
and a modular form f ∈ Mk (Γ ). Then the sum of the orders of f satisfies the follow-
ing equation:
1 1 0 k
ord∞ f + ordi f + ordρ f + ∑ ordτ f =
2 3 τ 12
Here the symbol ∑0τ denotes the summation over all points τ ∈ D \ {i, ρ}.
Proof. We recall the relation between the order of f ∈ Mk (Γ ) at a point τ and the
order of the differential form ω f ∈ H 0 (X, M ⊗O Ω ⊗k/2 ) at the point [τ] ∈ X estab-
lished during the proof of Theorem 3.17:
The modular curve X has genus g = 0 according to Corollary 3.16. Therefore the
line bundle ΩX has degree = −2, and the tensor product ΩX ⊗k/2 has degree
= (−2) · k/2 = −k
which implies
deg div(ω f ) = −k.
As a consequence
1 1 0
ord∞ f + ordi f + ordρ f + ∑ ordτ f =
2 3 τ
! ! !
k k k 0
= ord[∞] ω f + + ord[i] ω f + + ord[ρ] ω f + + ∑ ord[τ] ω f =
2 4 3 τ
k k k k k k k
= deg div(ω f ) + + + = −k + + + = , q.e.d.
2 4 3 2 4 3 12
Corollary 3.21 (Specific values and Fourier coefficents of the modular invariant j).
g2 3
j = 1728 ·
g2 − 27g3 2
3
g32
j = 26 · 33 ·
∆
3.2 Eisenstein series and the algebra of modular forms of Γ 113
∆ (τ) ∞
12
= ∑ τn · qn ∈ Z{q}, q = e2πi·τ .
(2π) n=1
We obtain
!
g3 26 3 12 1
j = 26 · 33 · 2 = 26 · 33 · ·E ·π · =
∆ 33 4 ∆
E4 (q) = 1 + O(q)
from Corollary 3.11 and the fact, that E4 has integer Fourier coefficents, q.e.d.
f (τ) = ∑ cn · qn , c−k 6= 0,
n≥−k
• and all coefficients of P(X) are Z-linear combinations of the Fourier coeffi-
cients c−k , ..., c−1 , c0 .
2. The field of automorphic functions, which equals the field of meromorphic func-
tions on the modular curve X, is
A0 (Γ ) = C( j).
114 3 The algebra of modular forms
Thanks to Corollary 3.16 the second part of Corollary 3.22 is a disguised version of
the well-known fact that the field of meromorphic functions on P1 is C(z), the field
of rational functions in one complex variable.
Proof. 1. The proof goes by induction on k: If k = 0 than f is a modular func-
tion, hence equals the constant c0 . For the induction step k − 1 7→ k consider the
function
g := f − c−k · jk .
The function g has at most a pole of order = k − 1 at ∞. Therefore the induction
assumption applies to g.
the respective zero set and pole set of f , counted with multiplicity. Both sets have
the same cardinality because a non-constant holomorphic map between compact
Riemann surfaces attains each value with the same multiplicity, see [63]. Then
n j − j(xν )
g := ∏ ∈ A0 (Γ )
ν=1 j − j(yν )
has the same zero set and pole set as f . The quotient f /g has no poles and zeros
on H ∪ {∞}. According to part 1) a constant a ∈ C exists with
f
a=
g
or
f = a · g ∈ C( j), q.e.d.
ε : Mk (Γ ) → C, f 7→ f (∞),
µ∆ : Mk−12 (Γ ) → Mk (Γ ), f 7→ f · ∆ ,
the multiplication with the discriminant form. Then for even k ≥ 4: The following
sequence of vector space morphisms is exact
∆ µ ε
0 → Mk−12 (Γ ) −→ Mk (Γ ) →
− C → 0.
In particular:
3.2 Eisenstein series and the algebra of modular forms of Γ 115
• For even k ≥ 4:
Mk (Γ ) = Sk (Γ ) ⊕ C · Gk ,
and the multiplication by the discriminant modular form ∆ defines a vector space
isomorphism
'
µ∆ : Mk−12 (Γ ) −
→ Sk (Γ ).
• For k < 0:
dim Mk = 0
As a consequence, the formulas from Corollary 3.19 allow to compute the di-
mension of Sk (Γ ). A partial generalization of Lemma 3.23 will be proved in
Corollary 3.28.
Proof. i) Exactness on the right-hand side: For even k ≥ 4 the Eisenstein series Gk ∈ Mk (Γ )
satisfies
Gk (∞) = 2 · ζ (k) 6= 0.
Therefore ε is surjective.
ii) Exactness in the middle: If f ∈ Mk (Γ ) with ε( f ) = 0, i.e.
f ∈ Sk (Γ )
then the quotient
f
g :=
∆
is holomorphic on H, because ∆ has no zero on H due to Proposition 1.19. In
addition g is holomorphic at the point ∞ because ∆ has a zero of order 1 at ∞ ac-
cording to Proposition 3.14. Hence g ∈ Mk−12 (Γ ). In addition, the composition
ε ◦ µ∆ = 0
is obvious.
iii) Exactness on the left-hand side: Apparently the map µ∆ is injective.
iv) As a consequence
dim Mk (Γ ) = 0
of modular forms of the module group Γ is commutative and freely generated by the
two Eisenstein series
G4 ∈ M4 (Γ ) and G6 ∈ M6 (Γ ).
The subalgebra of cusp forms
M
S∗ (Γ ) := S2k (Γ ) ⊂ M∗ (Γ )
k≥0
k = 4·α +6·β
The claim follows by repeating the argument with g ∈ Mk−12 (Γ ), and using
M4 (Γ ) = C · G4 and M6 (Γ ) = C · G6
3.3 Generalization to Hecke congruence subgroups Γ0 (N) 117
ii) Freely generated: The generators G4 and G6 are even algebraically independent.
Assume the existence of a polynomial P ∈ C[T1 , T2 ] with
P(G4 , G6 ) = 0.
We may assume that all monomials of P(G4 , G6 ) have the same weight. The case
∗
P(G4 , G6 ) = a · Gm
4 + G6 · P̂(G4 , G6 ), a ∈ C ,
P(G4 , G6 ) = G4 G6 · P̂(G4 , G6 ) = 0.
deg f ε2 ε3 ε∞
g(X0 (N)) = 1 + − − −
12 4 3 2
Moreover, we recall the canonical projection from the extended half-plane
p : H∗ →
− X0 (N) := Γ0 (N)\H∗
The aid to represent modular forms of Mk (Γ0 (N)), which are defined as functions
on H∗ , as meromorphic differential forms on the modular curve X0 (N) are the divi-
sors
$ % $ %
k k k
Dk := · ∑ x+ · ∑ x + · ∑ x ∈ Div(X0 (N)), even k ≥ 0.
4 x∈X (N) 3 x∈X (N) 2 x∈X (N)
0 0 0
x elliptic x elliptic x cusp
hx =2 hx =3
Lemma 3.25 (Order relation). Consider a Hecke congruence subgroup Γ0 (N) and
an even weight k ≥ 0.
1. Definition of α: The map
− H 0 X0 (N), M ⊗ O Ω ⊗k/2 , g 7→ ωg ,
α : Mk (Γ0 (N)) →
is well-defined.
2. Order relation: The order of a modular form g and the order of the meromorphic
differential form
ωg = α(g)
relate as
ord[τ] ωg + k/2
[τ] cusp
ordτ g =
hτ · ord[τ] ωg + k/2 · (hτ − 1) τ ∈ H
g(τ) (dτ)k/2
Y0 (N) := p(H)
Moreover Γ0 (N) has only finitely many cusps, see Example 2.11. For
arbitrary q ∈ Q and
β ∈ Γ with q = β (∞)
each modular form g ∈ Mk (Γ0 (N)) has a well-defined order
A cumbersome calculation by using charts, see [17, Sect. 3.3.], shows that the order
relations from the proof of Theorem 3.17 carry over: These formula show that the
map α is well-defined. The map α is injective, and via pullback of differential forms
also surjective, q.e.d.
120 3 The algebra of modular forms
of degree $ % $ %
k k k
deg Dk = · ε2 + · ε3 + · ε∞ ≥ 0
4 3 2
satisfies:
• Meromorphic differential forms:
ΩN,k ⊂ M ⊗ O Ω ⊗k/2
defined by
p∗ ωg = g(τ) (dτ)k/2 .
Here p : H∗ →
− X0 (N) denotes the canonical projection. Then
3.3 Generalization to Hecke congruence subgroups Γ0 (N) 121
• Cohomology: For k ≥ 2
H 1 (X0 (N), LN,k ) = 0
• Dimension of vector space of sections: For k ≥ 2
and for k = 0
Proof. We set
g := g(X0 (N))
i) Estimating the degree of (k/2) · K + Dk : By definition
LN,k = O(k/2)·K+Dk .
and the genus formula for X0 (N) we obtain for the degree of the corresponding
divisor for k ≥ 2
! !
k k−2 k−2 k
deg ((k/2) · K + Dk ) ≥ · (2g − 2) + · ε2 + · ε3 + · ε∞ =
2 4 3 2
!
k−2 ε2 2
= (2g − 2) + · 2g − 2 + + · ε3 + ε∞ + ε∞ =
2 2 3
!
k − 2 deg f
= (2g − 2) + · + ε∞ > 2g − 2
2 6
ii) Vanishing of dim H 1 (X0 (N), LN,k ): We apply Serre’s duality theorem in the form
∨
H 1 (X0 (N), LN,k ) = H 0 (X0 (N), LN,k ⊗ O Ω )∨
Hence
∨
dim H 1 (X0 (N), LN,k ) = dim H 0 (X0 (N), LN,k ⊗OΩ) = 0
122 3 The algebra of modular forms
iii) Riemann-Roch theorem for LN,k : The Riemann-Roch theorem proves the di-
mension formula
Remark 3.27 (Modular forms of Hecke congruence subgroups). Figure 3.1 shows
different modular curves as coverings of P1 and the dimension of their vector spaces
of modular forms, see PARI file ”Congruence subgroup 19”. The last column shows
the numerical value
dim H 0 (X(Γ0 (N)), LN,k )
computed by applying the Riemann-Roch theorem to the line bundle LN,k .
Apparently
M0 (Γ0 (N)) ' C, N ∈ N,
which implies
3.3 Generalization to Hecke congruence subgroups Γ0 (N) 123
Γ0 (N), N ∈ N.
2. For weight k = 2 cusp forms are exactly the holomorphic differential forms on
the modular curve. Hence
D∞
k := ∑ (−1) · x + Dk ∈ Div(X0 (N))
x∈X0 (N)
x cusp
LN,k
∞
= O(k/2)·K+D∞k ∈ Pic(X0 (N)
with
deg LN,k
∞
= −ε∞ + deg LN,k .
The order relations imply for even k ≥ 2
H 0 X0 (N), LN,k
∞
=
= {h ∈ Mk (Γ0 (N)) : ordt h ≥ 1 for all t ∈ H∗ with [t] cusp} = Sk (Γ0 (N))
1. k ≥ 4: The formula from the proof of Theorem 3.26 shows for even k ≥ 4
124 3 The algebra of modular forms
k − 2 deg f
deg LN,k
∞
≥ 2g − 2 + · > 2g − 2
2 6
because deg f ≥ 1. Hence H 1 X0 (N), LN,k
∞ = 0 because
∞ ∨
deg (LN,k ) ⊗ O Ω 1 < −(2g − 2) + (2g − 2) = 0,
D∞
2 = 0 ∈ Div(X0 (N)),
which implies
LN,2
∞
= OK = Ω 1
and
S2 (X0 (N)) = H 0 (X0 (N), Ω ), q.e.d.
Chapter 4
Elliptic curves
The present chapter ties quite closely with Section 2.1 on tori. We now bridge com-
plex analysis and algebraic geometry.
Section 4.1 shows that each complex torus is an algebraic variety in the complex
projective space P2 . The variety is the zero set of a cubic polynomial.
Section 4.2 introduces elliptic curves as projective-algebraic varieties of genus = 1
defined over arbitrary subfields of C. We show that elliptic curves correspond bijec-
tively to the smooth zero set of cubic polynomials, the Weierstrass polynomials. The
passage to algebraic geometry allows to investigate the points of elliptic curves with
coordinates in distinguished subfields of C.
Pn := (Cn+1 \ {0})/ ∼
π : Cn+1 \ {0} →
− Pn , z 7→ [z].
125
126 4 Elliptic curves
U := (Ui )i=0,...,n
Here the “hat” indicates omission of the term. The transformation between two
charts φ j and φi with i < j is the map
ψi j := φi ◦ φ j−1 : φ j (Ui ∩U j ) →
− φi (Ui ∩U j ),
!
w0 wi−1 wi+1 w j−1 1 w j+1 wn
w0 , ..., w j−1 , 1̂, w j+1 , ..., wn 7→ , ..., , 1̂, , ..., , , , ...,
wi wi wi wi wi wi wi
Theorem 4.3 shows once more the importance of the Weierstrass function ℘ and of
its differential equation for ℘ 0 in the theory of complex tori.
Theorem 4.3 (The embedding theorem for tori). Consider a complex torus
T := C/Λ
with lattice
Λ = Zω1 + Zω2 .
Set
g2 := 60 · GΛ ,4 , g3 := 140 · GΛ ,6 ,
and denote by ℘ = ℘Λ the Weierstrass function of Λ . The map
4.1 Embedding tori as plane cubic hypersurfaces 127
(
2 (1 : ℘(p) : ℘ 0 (p)) if p 6= 0
φ : T → P , p 7→
(0 : 0 : 1) if p = 0
Proof. The subsequent proof of Theorem 4.3 relies heavily on the properties of ℘
and ℘ 0 from Section 1.2.
φ (T \ {0}) ⊂ Ea f f := {(x, y) ∈ C2 : y2 = 4 · x3 − g2 · x − g3 }
Ea f f ⊂ C2 ⊂ P2
is the projective algebraic variety E ⊂ P2 , see Figure 4.1, the zero set in P2 of
Fhom (X0 , X1 , X2 ),
lim φ (z) = (0 : 0 : 1) := O ∈ P2
z→0
− P1
℘: T →
℘0 : T →
− P1
z 6= w ∈ T \ {0}
with
(℘(z),℘ 0 (z)) = (℘(w),℘ 0 (w)) ∈ C2 .
Then
℘(z) = ℘(w) implies w = −z
due to Corollary 1.10 and because ℘ is an even function due to Theorem 1.12.
Moreover
z = [ω/2], ω ∈ Λ and w = −z ∈ T
are two distinct classes of half-period points of Λ . But ℘ attains at z the value ℘(z)
with multiplicity = 2, hence ℘ assumes the value ℘(z) = ℘(w) with
multiplicity ≥ 4, a contradiction.
E \ Ea f f = O = φ (0).
− P1 ,
℘: T →
℘(z) = x.
℘(−z) = x.
From
y2 = 4x3 − g2 · x − g3 = 4℘(z)3 − g2 ·℘(z) − g3 = ℘ 0 (z)2
follows:
• Either
y = ℘ 0 (z) and φ (z) = (1 : x : y).
• Or
y = −℘ 0 (z) = ℘ 0 (−z) and φ (−z) = (1 : x : y)
130 4 Elliptic curves
v) Smoothness of E:
• A point (x0 , y0 ) ∈ C2 is a singular point of Ea f f iff it satisfies the three equations
∂F ∂F
(x0 , y0 ) = 2y0 = 0 and (x0 , y0 ) = −12x02 + A = 0
∂y ∂x
Introducing the cubic poynomial
∆ f = A3 − 27B2 .
We have
φ2 (E ∩U2 ) = {(u, v) ∈ C2 : f (u, v) = 0}
with
f (u, v) := u − (4 · v3 − A · u2 · v − B · u3 ).
Then the partial derivatives are
∂f ∂f
(u, v) = 1 − (2 · A · u · v − 3 · B · u2 ) and (u, v) = −12 · v2 + A · u2
∂u ∂v
hence
∇ f (0, 0) = (1, 0) 6= 0.
The non-singular projective algebraic curve E = φ (T ) ⊂ P2 is connected, hence
a compact Riemann surface with the induced analytic structure as a
hypersurface of P2 .
φ :T →
− E
between Riemann surfaces is a bijective holomorphic map, hence it is
biholomorphic, q.e.d.
The embedding result of Theorem 4.3 is a particular case of the general question:
How to obtain holomorphic maps
− Pn
X→
Definition 4.5 (The twisted line bundle on Pn ). The twisted line bundle O(1)
on Pn is defined with respect to the standard atlas of Pn by the cocycle
g = (gi j ) ∈ Z 1 (U , O ∗ )
with
zj
gi j (z0 : ... : zn ) := , 0 ≤ i 6= j ≤ n.
zi
Here sk ∈ H 0 (Pn , O(1)) is defined with respect to the i-th standard chart by the
holomorphic functions
zk
sk,i : Ui →
− C, sk,i (z0 : ... : zn ) :=
zi
Proof. The map defined above on the standard charts provide a section
because
z j zk zk
gi j · sk, j = · = = sk,i , q.e.d.
zi z j zi
O(k) := O(1)⊗k , k ∈ N,
mx ⊂ Ox
Proof. We set
φ := φL .
i) Definition: We consider the definition of φ in a suitable open neighbourhood U of x.
On U we choose a trivialization to identify the line bundle L with the struc-
ture sheaf O. Hence sections from H 0 (U, L ) become holomorphic functions.
Because L is globally generated there is an index j ∈ {0, ..., n} with
σ j (x) 6= 0.
L |U ' O|U.
ii) Pullback of the twisted line bundle: For each i = 0, ..., n the sets
Xi := φ −1 (Ui ) = {x ∈ X : (σi )x ∈
/ mx Lx }
O(1)(Ui ) →
− φ∗ (L )(Ui ) = L (Xi ) induced by si |Ui 7→ σi |Xi
O(1) →
− φ∗ L
φ ∗ (O(1)) →
− L,
because the two functors φ ∗ and φ∗ are adjoint, see [29, Chap. II, § 5], i.e. there is a
canonical group isomorphim
x ∈ Xi ⊂ X and y := φ (x) ∈ Ui ⊂ Pn
Theorem 4.9 provides a geometric criterion for the map provided by a globally
generated line bundle L to be a closed embedding.
or vice versa.
4.1 Embedding tori as plane cubic hypersurfaces 135
dL ,x : {σ ∈ H 0 (X, L ) : σx ∈ mx Lx } →
− mx Lx /m2x Lx , σ 7→ [σx ],
is surjective.
Remark 4.10 (Separating points and tangent vectors). We explain the meaning of
the two statements from Theorem 4.9.
1. Separating points: The homogeneous coordinates of a given point in Pn do not
fix the value of the components. But it is determined whether a component is
zero or not zero. Therefore for each pair of points p 6= q ∈ X
φL (p) 6= φL (q)
if at least one component of the left-hand side is different from zero while the
same component on the right-hand side is zero.
dL ,x : {σ ∈ H 0 (X, L ) : σx ∈ mx Lx } →
− Tx1 ⊗Ox Lx
σ ∈ H 0 (X, L ) with σx ∈ mx Lx
σ = f · σ1
dL ,x (σ ) = dx f ⊗ σ1 ∈ Tx1 ⊗Ox Lx
136 4 Elliptic curves
dO(1),y
{s ∈ H 0 (Pn , O(1)) : s(y) = 0} TP1n ,y ⊗OPn ,y O(1)y
φ∗ φT∗1
dL ,x
{σ ∈ H 0 (X, L ) : σ (x) = 0} 1 ⊗
TX,x OX,x Lx
y := φ (x)
Definition 4.11 (Very ample line bundle). A globally generated line bundle L on
a compact Riemann surface X is very ample if the induced map
− Pn
φL : X →
is an embedding.
Notation 4.12. For a line bundle L on a Riemann surface X and a divisor D ∈ Div(X)
we denote by
LD := L ⊗O OD
the sheaf of meromorphic sections of L which are multiples of the divisor −D. The
sheaf LD is isomorphic to a line bundle.
H 0 (X, LD ), D ∈ Div(X),
• For any two not necessarily distinct points p, q ∈ X the corresponding point di-
visors
P := 1 · p, Q := 1 · q ∈ Div(X)
satisfy
dim H 0 (X, L−(P+Q) ) = dim H 0 (X, L ) − 2.
Proof. i) Assume the validity of the dimension formula: The formula implies for any
two point divisors P, Q ∈ Div(X)
H 0 (X, L ) →
− L p /m p L p ' C, σ 7→ [σ p ],
− Pn
φL : X →
is well-defined.
• The equation
H 0 (X, L−(P+Q) ) ( H 0 (X, L−P )
implies that for any pair of distinct points p, q ∈ X there exists a section
i.e. satisfying
σ (p) = 0 but σ (q) 6= 0.
Hence the sheaf L separates points.
• The dimension formula shows for any point p ∈ X with point divisor P ∈ Div(X)
H 0 (X, L−P ) →
− mX,p L p →
− mX,p L p /m2X,p L p
due to
dimC (mX,p /m2X,p ) = 1.
Therefore L separates tangent vectors.
ii) Assume L very ample: Theorem 4.9 implies that L separates points and
tangent vectors. Separating points implies for all point divisors P 6= Q ∈ Div(X)
L := OD ∈ Pic(X)
(1,℘,℘ 0 ),
which gives a second proof for the embedding from Theorem 4.3.
Genus
Structure sheaf OX
Line bundle
Very ampleness
Divisor
Cohomology
Riemann-Roch Theorem
Serre Duality
The table relates concepts from complex analytix geometry on the left-hand side
with concepts from algebraic geometry on the right-hand side. But not any compact
manifold, and not even any compact Kaehler manifold has a counter part in Alge-
braic Geometry. Hence not any instance of a concept on the left-hand side of the
table induces a correponding instance on the right-hand side. On both sides there
are objects which cannot be studied by methods from the other side.
4.2 Elliptic curves over subfields of C 141
Remark 4.16 recalls varieties as the basic objects of classical algebraic geometry.
Varieties are irreducible by definition, hence they are defined by prime ideals. But
different from manifolds varieties may have singularities. Hence varieties are the
analogue of reduced and irreducible complex spaces. We distinguish between their
field k of definition and the algebraic closure k where the points of the variety have
their coordinates. As a topological space varieties will be equipped with the Zariski
topology.
Remark 4.16 (Algebraic variety with structure sheaf). Consider a perfect field k and
denote by k the algebraic closure of k.
1. Variety: An affine algebraic variety X is the zero set
n
X = V (a) ⊂ An (k) = k
The ring
R := k[X1 , ..., Xn ]/a
is the affine coordinate ring of X.
X = V (a) ⊂ Pn (k)
of a prime ideal
a ⊂ k[X0 , ..., Xn ],
142 4 Elliptic curves
3. Zariski topology: Closed sets of An (k) are the affine algebraic subvarieties and
their finite unions. The Zariski topology on an affine algebraic variety X ⊂ An (k)
is the subspace topology induced from the Zariski topology of An (k). A basis of
the Zariski topology of X are the sets
Dh := {x ∈ X : h(x) 6= 0}, h ∈ R.
Closed sets of Pn (k) are the projective algebraic subvarieties. The Zariski topol-
ogy on a projective algebraic variety X ⊂ Pn (k) is the subspace topology induced
from the Zariski topology of Pn (k).
Non-empty open subsets of an irreducible space are irreducible itself, and are
also dense.
4. Structure sheaf and sheaf of rational functions: Consider an affine algebraic va-
riety X and an open set U ⊂ X. A function
f :U →k
g
f= with g, h ∈ R, h locally without zeros in X.
h
Note that the quotient representation of f is only required to hold locally. The
sheaf OX of regular functions on an affine algebraic variety X is the contravariant
functor
U 7→ OX (U) := { f : U → k| f regular }, U ⊂ X open,
with the restriction of functions. The sheaf OX is named the structure sheaf of X.
Sections over a basis open set Dh ⊂ X form the ring
O(Dh ) = Rh := {g/hn : g ∈ R, n ∈ N}
OX (U), U ⊂ X open,
with Q(OX (U)) the quotient field of the integral domain OX (U). Sections of MX (U)
are locally quotients
g/h, g, h ∈ R, h 6= 0.
For a non-empty open set U ⊂ X holds
independent from U. The field Q(R) is named the function field k(X) of X. The
sheaf
M∗ ⊂ M
is the multiplicative subsheaf of non-zero rational functions.
The definitions carry over to projective algebraic varieties, beaucse the latter
have an atlas of affine algebraic varieties.
C ⊂ P2 (k)
7. Affine schemes: From a modern viewpoint the primary property of a variety is not
the point set
X ⊂ An (k),
the primary object is the ideal
a ⊂ k[X1 , ..., Xn ]
144 4 Elliptic curves
(X := Spec R, OX )
Here Spec R is a topological space with points the prime ideals of R. The structure
sheaf OX has as stalks the localizations
Rp , p ⊂ R prime ideal.
• Secondly, the algebraically closed field k, used to visualize K/k as a point set X
from a geometrical view point. Also the condition on smoothness refers to k.
But the set X ⊂ An (k) respectively X ⊂ Pn (k) is not the only point set attached
to K/k. Definition 4.17 introduces point sets X(K) attached to all intermediate
fields k ⊂ K ⊂ k.
Definition 4.17 (K-valued point pf X/k). Consider a projective algebraic variety X/k
with
X ⊂ Pn (k).
Then for any field
k⊂K
the set of K-valued points of X is defined as
X(K) := X ∩ Pn (K) =
4.2 Elliptic curves over subfields of C 145
φ : X an →
− X,
see [46]. As a consequence: For a given complex analytic space Y each morphism
f :Y →
− X
f an : Y →
− X an ,
f an f
φ
X an X
The analytification φ : X an →
− X from the analytification defines for each x ∈ X an
a pull-back
146 4 Elliptic curves
φx : OX,φ (x) →
− OX an ,x
which allows to compare the local ring of regular functions on X and the local
ring of holomorphic functions on X an . As a consequence one obtains compari-
son theorems between the algebraic properties of the local rings of the structure
sheaves of X and X an , see [46, Chap. 2].
φX
X an X
f an f
φY
Y an Y
Again, one obtains comparison theorems between the algebraic properties of the
morphisms
f : X → Y and f an : X an →
− Y an ,
see [46, Chap. 3].
− X an
φ :X →
from the analytification of X defines the functor analytical inverse image as the
pull-back of module sheaves over the structure sheaf:
φ ∗ : Shea f O →
− Shea f O , F 7→ φ ∗ F
X X an
and
− G , 7→ φ ∗ f : φ ∗ F →
f :F → − φ ∗G .
The stalk at a point x ∈ X an of the inverse image sheaf is easy to calculate
f :X →
−Y
4.2 Elliptic curves over subfields of C 147
D= ∑ nx · x
x∈C
with integers nx ∈ Z satisfying nx = 0 for all but finitely many indices. For a
point p ∈ C the corresponding point divisor is denoted
P := 1 · p
deg D := ∑ nx ∈ Z
x∈C
The additive group of divisors on C/k is denoted Div(C), its subgroup Div0 (C) ⊂ Div(C)
contains the classes of divisors of degree = 0.
D = div f
Cl(C) := Div(C)/Prin(C)
3. A divisor
D= ∑ nx · x ∈ Div(C)
x∈C
is defined over k if it is invariant under the action of the Galois group Gk/k , i.e. if
for all σ ∈ Gk/k
D = Dσ := ∑ nx · xσ ∈ Div(C)
x∈C
dim H 0 (C, O) = 1
hold literally in the same form as on compact Riemann surfaces. Also the canonical
divisor K ∈ Div(C) has degree
deg K = 2g(C) − 2.
Hence we will apply these results also in the context of algebraic geoemtery without
further mentioning.
Definition 4.20 (Elliptic curve over k). An elliptic curve defined over k is a
pair (E, O) with
• a projective algebraic curve E/k of genus g = 1
Note. In Definition 4.20 “O” denotes the upper-case letter O, not the digit 0. We
denote by
O∞ := 1 · O ∈ Div(E)
4.2 Elliptic curves over subfields of C 149
If the base field k and the k-valued point O are not relevant for the context, we will
often denote an elliptic curve just by E.
Note that there exist smooth projective algebraic curves C/Q of genus g = 1
without rational points, i.e. with C(Q) = 0.
/ An example is the Selmer
curve C ∈ P2 (C) defined by the homogeneous cubic equation
3X 3 + 4Y 3 + 5Z 3 = 0.
The fact g = 1 for an elliptic curve E allows a simple application of the Riemann-
Roch theorem to compute the vector spaces of multiples of a divisor on E.
Lemma 4.21 (Divisors on an elliptic curve). Consider an elliptic curve E and a
divisor D ∈ Div(E). Then
0 if deg D < 0
0 if deg D = 0, D not principal
dim H 0 (E, OD ) =
1 if deg D = 0, D principal
if deg D > 0
deg D
Proof. The Riemann-Roch theorem together with Serre duality imply for any
divisor D ∈ Div(E)
= 1 − g + deg D = deg D
Because an elliptic curve E has genus g = 1, its canonical divisor has degree
deg K = 2g − 2 = 0.
As a consequence
D principal ⇐⇒ OD ' O
150 4 Elliptic curves
On the other hand, for non-principal D the line bundle OD is not trivial. Hence each
regular section of s ∈ H 0 (E, OD ) has a zero. If s 6= 0, then deg D = 0 implies that s
has also at least a pole, contradicting the regularity of s. Therefore s = 0, q.e.d.
The following Lemma 4.22 will be used in the proof of Theorem 4.23. It indi-
cates why the existence of a k-valued point is requested for an elliptic curve defined
over k.
Lemma 4.22 (Divisors defined over a subfield). Consider a projective algebraic
curve C/k and a divisor
D = ∑ nx · x ∈ Div(C)
x∈X
Elliptic curves are by definition subvarieties of a projective space Pn (k). Theorem 4.23
proves that they are even plane curves, i.e. that they embed as hypersurfaces
into P2 (k).
Theorem 4.23 (Elliptic curves are plane cubics). For an elliptic curve (E/k, O)
exist two rational functions x, y ∈ k(E) with poles only at O such that the map
(
2 (1 : x(p) : y(p)) p 6= O
φ : E → P (k), p 7→
(0 : 0 : 1) p=O
D := 3 · O∞ ∈ Div(E).
and
dim H 0 (E, OD−(P+Q) ) = dim H 0 (E, OD ) − 2.
For n ≥ 0
∗
H 0 (E, On·O∞ ) ) = { f ∈ k (E) : f has at most a pole at O of order = n} ∪ {0}
exists with
• y ∈ H 0 (E, O3·O∞ ) ⊂ k(E) has a pole of order = 3 at O and no other pole,
(1, x, y, x2 , xy, x3 , y2 )
of seven rational functions from the 6-dimensional vector space H 0 (E, O6·O∞ ) is
linearly dependent. Therefore the functions satisfy a linear relation with coefficients
from k. Among them the two functions x3 and y2 are the only ones having a pole
at O ∈ E of order = 6. Hence their coefficients in the linear relation are not zero.
After division we may assume that their coefficients are = ±1 and
y2 + a1 xy + a3 y = x3 + a2 x2 + a4 x + a6
152 4 Elliptic curves
in the single point (0 : 0 : 1) with multiplicity = 3. Its affine part is the affine variety
of the polynomial F(X,Y ) ∈ k[X,Y ]
F(X,Y ) = (Y + b1 )(Y + b2 )
with polynomials b1 , b2 ∈ k[X]. Therefore Caff and also C are irreducible. Hence the
non-constant regular map
φ :E →C
is surjective, cf. [29, Chap. II, Prop. 6.8], an analogue of the theorem about non-
constant proper holomorphic maps between Riemann surfaces. As a consequence
In the following we will identify an abstract elliptic curve (E/k, O) with the plane
cubic from Theorem 4.23. We say that (E/k, O), more precisely its affine part, is
defined over k by the Weierstrass equation F(X,Y ) = 0.
4.2 Elliptic curves over subfields of C 153
with coefficients
a1 , a2 , a3 , a4 , a6 ∈ k.
1. The equation
F(X,Y ) = 0,
i.e.
Y 2 + a1 XY + a3Y = X 3 + a2 X 2 + a4 X + a6
is named the Weierstrass equation over k defined by F.
2. The discriminant of F is
with
b2 := a21 + 4a2 , b4 := 2a4 + a1 a3 , b6 := a23 + 4a6
and
b8 := a21 a6 + 4a2 a6 − a1 a3 a4 + a2 a23 − a24 .
3. The j-invariant of F is
c34
jF := , c4 := b22 − 24bb .
∆F
Starting from an elliptic curve over a field k the resulting Weierstrass polynomial
is not unique. There exist coordinate transformations defined over k which transform
a Weierstrass polynomial over k to a different form, in particular to a more simple
form.
Proposition 4.25 (Weierstrass equation in short form, discriminant and j-invariant).
Consider a Weierstrass polynomial
with constants
r, s,t ∈ k, u ∈ k∗
154 4 Elliptic curves
with discriminant
∆F 0 = −24 · (4A3 + 27B2 ) ∈ k
and the same j-invariant
jF 0 = jF
2. Transforming Weierstrass polynomials in short form: Each linear coordinate
transformation over k
0 2 0
X X u 0 X
7→ := · .
Y0 Y 0 u3 Y0
F(X,Y ) = Y 2 − (X 3 + A · X + B) ∈ k[X,Y ].
A = u4 · A0 , B = u6 · B0 , ∆F = u12 · ∆F 0 .
(4A)3 (4A0 )3
jF := 1728 · = 1728 · =: jF 0
∆F ∆F 0
3. The j-invariant: Two Weierstrass polynomials in short form have the same j-invariant
if and only if they are related by an isomorphism as in part 3) defined over k -
not necessarily over k.
4. For each algebraic number j0 ∈ k exists an elliptic curve E/k( j0 ) with j-invariant
equal to j0 .
For the proof of Proposition 4.25 see [56]: For part 1) cf. Chap. III, §1; for part 2)
cf. Chap. III, Rem. 1.3; for part 3) cf. Chap. III, Prop. 1.4(b); for part 4) cf.
Chap. III, Prop. 1.4(c).
C = V (Fhom ) ⊂ P2 (k)
F(X,Y ) = Y 2 − (X 3 + AX + B) ∈ k[X,Y ]
Remark 4.27 (Cubic projective-algebraic curves and elliptic curves). The adjunc-
tion formula for non-singular hypersurfaces implies: Any projective algebraic cu-
bic curve C ⊂ P2 has genus = 1. Hence C/k is an elliptic curve if C has at least
one k-valued point, see [29, Chap. V, Example. 1.5.1].
Theorem 4.28 is the converse of Theorem 4.3. it shows that the complex points of
an elliptic curve, defined over a subfield of C, can be parametrized by the
Weierstrass function ℘ and its derivative ℘ 0 of a suitable torus. The proof relies on
the result of Theorem 4.23 that elliptic curves are plane cubics which satisfy a
Weierstrass equation.
A very coarse analogue of the uniformization is the parametrization of the points of
the unit circle by the values of the exponential function e2πi·t on the parameter
interval [0, 1].
Theorem 4.28 (Uniformization of elliptic curves E/C by tori). For any elliptic
curve E defined over C exists a torus T = C/Λ and a biholomorphic map
'
Φ :T −
→ E(C).
Here E(C) ⊂ Pn (C) is provided with the induced Euclidean topology and analytic
structure from the complex manifold Pn (C) = Pn .
156 4 Elliptic curves
Proof. Due to Theorem 4.23 we may assume that E is defined as a plane curve,
defined by the homogenization of a Weierstrass polynomial in short form
2
F(X,Y 0 ) := Y 0 − (X 3 + a4 · X + a6 ) ∈ C[X,Y 0 ].
The substitution
Y 0 = (1/2) ·Y
shows that E can also be defined by the Weierstrass polynomial
F(X,Y ) = Y 2 − (4 · X 3 − A · X − B)
∆F = A3 − 27 · B3
Λ = Zω1 + Zω2 ⊂ C
A = 60 · GΛ , 4 and B = 140 · GΛ , 6 .
A3
j(τ) = 1728 · .
∆F
If g2 (τ) = 0 then τ = ρ according to Corollary 3.20. We choose
Λ := Λρ .
Λ := Λi .
g32 (τ) A3
j(τ) = 1728 · = 1728 ·
g32 (τ) − 27g23 (τ) ∆F
and
∆F = A3 − 27B2
4.2 Elliptic curves over subfields of C 157
imply
g32 (τ) A3
= .
g32 (τ) − 27g23 (τ) A3 − 27 · B2
Clearing the denominators by crosswise multiplication implies
!3 !2
A B
=
g2 (τ) g3 (τ)
We choose ω1 ∈ C with
!4 !6
A 1 B 1
= and = .
g2 (τ) ω1 g3 (τ) ω1
ω2 := τ · ω1 , i.e. τ = ω2 /ω1 .
Λ := Zω1 + Zω2 .
and similarly
GΛ ,6 = (1/140) · B,
i.e.
A = gΛ ,2 and B = gΛ ,3 .
The map
− P2
φ : C/Λ →
from Theorem 4.3 maps the torus C/Λ biholomorphically to the cubic hypersurface
with Weierstrass polynomial
Example 4.29 (Elliptic curves and cubics with a singularity). The following Figures 4.5 - 4.8
show some elliptic curves.
While Figure 4.9 and Figure 4.10 show the singular cubics node and the Neil
parabola. They have a single singularity, namely at the origin respectively a node
and a cusp. The corresponding table from Figure 4.4 shows the invariants. The role
of the conductor will be explained in Definition 4.40 later.
158 4 Elliptic curves
For the computation and the plots see the PARI files Elliptic_curve_02
and Elliptic_curve_plot_06.
Elliptic curve
4.7 Y 2 = X3 − X 1728 26 25
4.8 Y 2 = X3 + 3 0 −24 · 35 24 · 35
Singular cubic
4.9 Y 2 = X3 + X2 − 0 −
4.10 Y 2 = X3 − 0 −
Fig. 4.5 Y 2 = X 3 − 3X + 3
Fig. 4.6 Y 2 = X 3 + X
Fig. 4.7 Y 2 = X 3 − X
160 4 Elliptic curves
Fig. 4.8 Y 2 = X 3 + 3
Fig. 4.9 Y 2 = X 3 + X 2
Fig. 4.10 Y 2 = X 3
4.2 Elliptic curves over subfields of C 161
Lemma 4.30 prepares the proof of Theorem 4.31 and the Definition 4.32 of the
group structure of an elliptic curve (E/k, O). The lemma shows that for a pair (p, q)
of distinct points from E the divisor
D := P − Q ∈ Div(E)
cannot be a principal divisor. Note that the lemma gives also a second proof of
Corollary 1.9.
Lemma 4.30 (No single pole of order one). Consider an elliptic curve (E/k, O)
defined over a field k. There is no ratiomal function f ∈ k(E)∗ with
div f = P − Q
div f = 0 ∈ Div(E)
Lemma 4.30 shows: The obstructions for a divisor D ∈ Div0 (E) to be a principal
divisor are divisors of type P − Q or even of type P − O∞ . A divisor D ∈ Div0 (E) is
the divisor of a rational function up to a divisor of type
P − O∞ .
Theorem 4.31 (The class group of degree zero divisors on an elliptic curve).
Consider an elliptic curve (E/k, O).
162 4 Elliptic curves
D ∼ P − O∞
σ : Div0 (E) → E, D 7→ p.
2. The map σ is surjective. It induces a bijective map from the group of divisor
classes of degree = 0
'
σ : Cl0 (E) −
→E
with inverse
'
E−
→ Cl0 (E), p 7→ [P − O∞ ].
div f ≥ −(D + O∞ ).
Because
deg (div f ) = 0 but deg(−(D + O∞ )) = −1,
there exists a point p ∈ E with
div f = −(D + O∞ ) + P
Then
0 ∼ −(D + O∞ ) + P i.e. D ∼ P − O∞ .
If for a different point
q ∈ E, q 6= p,
also
D ∼ Q − O∞
then
0 = D − D ∼ P − Q.
∗
Hence for a suitable rational function g ∈ k (E)
P − Q = div g,
σ (D) := p ∈ E
is well-defined.
4.2 Elliptic curves over subfields of C 163
D := P − O∞ ∈ Div0 (E).
iii) Injectivity of σ :
Definition 4.32 (Group structure of an elliptic curve). Consider an elliptic curve (E/k, O)
and denote by (Cl0 (E), +) the group of divisor classes of degree zero. The bijective
map from Theorem 4.31
σ : Cl0 (E) →
− E, [P − O∞ ] 7→ p,
carries over the group structure from (Cl0 (E), +) to E: The map
⊕ : E ×E →
− E, p ⊕ q := σ ([P − O∞ ] + [Q − O∞ ]),
defines an Abelian additive group (E, ⊕) with neutral element the k-valued point O ∈ E(k).
Corollary 4.33 (Abel’s Theorem for elliptic curves). Consider an elliptic curve (E/k, O).
A divisor
D = ∑ n p · P ∈ Div0 (E)
p∈E
Here the last equation refers to the group structure ⊕ on E. The product n p ∗ p
means:
• Taking n p -times the sum of a point p ∈ E if n p ≥ 0, and
Proof. Because
0 = deg D = ∑ np
p∈E
we have
D = D− ∑ nP · O∞ = ∑ n p · (P − O∞ ) ∈ Div0 (E)
p∈E p∈E
Proposition 4.34 (Group structure of tori and elliptic curves). For a complex
torus T = C/Λ the biholomorphic map from Theorem 4.3
φ :T →
− E(C)
ρ : (T, ⊕) →
− (Div0 (T ), +), p 7→ 1 · p − 1 · 0T ,
D1 + D2 − D = 1 · p1 + 1 · p2 − 1 · (p1 ⊕ p2 ) − 1 · 0T ∈ Div0 (T )
p1 ⊕ p2 (p1 ⊕ p2 ) 0T = 0T ∈ Λ
ρ(p) = 1 · p − 1 · 0T ∈ Div0 (T )
ρ :T →
− Cl0 (T )
4.2 Elliptic curves over subfields of C 165
ρ :T →
− Cl0 (T )
is injective.
defines a point ! !
M M
p := na · a nb · b ∈T
a∈A b∈B
Then
ρ(p) = 1 · p − 1 · 0T ∈ Div0 (T )
Due to Abel’s Theorem the difference
! !! !
M M
ρ(p)−D = 1 na · a nb · b −1·0T − ∑ na · a − ∑ nb · b ∈ Div0 (T )
a∈A b∈B a∈A b∈B
2. Due to part 1) and Definition 4.32 the group structures respectively on domain
and codomain of
φ :T →− E(C)
are determined by the divisor class groups. Because φ is an analytical isomor-
phism it is a group isomorphism, q.e.d.
See Figure 4.11: Two add two points p, q ∈ E \ {O} consider the line L(p, q)
passing through p and q. The line intersects E in a third point r ∈ E. The second
line L(r, O) intersects E in a third point, which is p ⊕ q. For more details see [63].
166 4 Elliptic curves
2. Mordells theorem: The group structure on E(Q) provides also the set E(K)
of K-valued points with a group structure for any intermediate field
Q ⊂ K ⊂ Q.
E(Q) = E(Q)tors × Zr .
Example 4.36 (Group structure of an elliptic curve). Consider an elliptic curve E/Q.
The PARI file Elliptic_curve_torsion_group investigates the torsion
group of E(Q) for the elliptic curve E/Q with Weierstrass polynomial
y2 = x3 − 43x + 166
see Figure 4.12. It shows that the torsion group E(Q)tors has a factor isomorphic
to Z/7Z. For more information, in particular for the equality
E(Q)tors = Z/7Z
see [54, Chap. VII, Theor. 7.5]. It is conjectured that the rank r of the free part of
the group E(Q) equals the analytic rank of E/Q.
168 4 Elliptic curves
The present section starts the investigation of elliptic curves defined over Q. Each
elliptiv curve E/Q can also be defined by a distinguished Weierstrass polynomial
with integer coefficients, called a global minimal polynomial, see Definition 4.37
and Remark 4.38.
Entering into the field of arithmetic geometry one may reduce the integer equa-
tion modulo any prime p ∈ Z, each time obtaining an elliptic curve defined over a
finite field F p . Considered from a general point of view: An elliptic curve defined
over Q is a family of not necessarily smooth curves defined over all prime fields of
prime characteristic. Or even more general, it is a map to Spec Z, and all but but
finitely many fibres are elliptic curves over prime fields. Reduction modulo p is a
powerful tool to investigate the arithmetic information encoded in an elliptic curve
defined over Q.
Before investigating the family of elliptic curves one has to eliminate redundant
prime powers of the discriminant which otherwise would distort the result. The dis-
criminants of the different Weierstrass polynomials of an elliptic curve differ by
a non-zero rational. After multiplying with a principal nominator they differ by a
non-zero integer. Hence the reduction of the discriminants modulo a prime p ∈ Z
may differ. To generalize the discriminant criterion from Proposition 4.26 one has
to find a unique minimal discriminant. The corresponding Weierstrass polynomial
is named global minimal.
Definition 4.37 (Global minimal Weierstrass polynomial). For an elliptic curve E/Q
a Weierstrass polynomial F ∈ Z[X,Y ] with integer coefficients is global minimal if
its discriminant
∆F = ± ∏ pv(p) , v(p) ∈ N∗ ,
p prime
is minimal for each given prime factor p of ∆F in the following sense: The elliptic
curve E/Q cannot be defined by a Weierstrass polynomial Fp (X,Y ) ∈ Z[X,Y ] with
discriminant ∆Fp satisfying
ord p (∆Fp ) < v(p).
Each elliptic curve over k = Q has a global minimal Weierstrass polynomial. The
result generalizes to elliptic curves over arbitrary number fields with class num-
ber = 1, see [56, Chap. VIII, Cor. 8.3] and more specific [33, Theor. 10.3]. A
global minimal Weierstrass polynomial is not necessarily in short form.
3. PARI command: The Pari command ellminimalmodel provides the global mini-
mal model for elliptic curves over number fields k with class number = 1.
F(X,Y ) = Y 2 − (X 3 + 15.625)
Its discrimant
∆F = −24 · 33 · 512
is not minimal because
512 |∆F
The global minimal Weierstrass polynomial is
Fmin (X,Y ) = Y 2 − (X 3 + 1)
with discriminant
∆Fmin = −24 · 33 .
4.3 Elliptic curves over finite fields 171
with discriminant
∆F = 212 · 73
is global minimal, despite of the exponent 12 of the particular prime factor p = 2.
3. Consider the elliptic curve E/Q with Weierstrass polynomial in short form
1
F(X,Y ) = Y 2 − (X 3 + · X + 1) ∈ Q[X,Y ]
5
which is not global minimal because it has a non-integer cefficient. To obtain a
global minimal Weierstrass polynomial, we make the transformation
−2 0
X u ·X
= −3 0 with u := 5
Y u ·Y
2 3 1 2 3
Y 0 = X 0 + · u−2 · u6 · X 0 + u6 i.e. Y 0 = X 0 + 53 · X 0 + 56
u
Hence E/Q is also defined by the Weierstrass polynomial
E = (E/Q, O)
Fhom,p ∈ F p [X0 , X1 , X2 ]
with coefficients from the finite field F p . The Weierstrass polynomial defines the
projective variety
cond(E) = ∏ pe p
p
with the same prime factors p as the discriminant ∆F . The exponents of the prime
factors satisfy
e p := 1 ⇐⇒ E is semistable at p
and for p > 3
e p := 2 ⇐⇒ E is unstable at p,
For p = 2, 3 see also [54, App. C, §16].
4.3 Elliptic curves over finite fields 173
with discriminant
∆F = −24 · 33 · 512 ∈ Z.
Note 15.625 = 56 .
ii) Therefore we have to apply first the coordinate transformation with the matrix
2
1/u 0
, u := 5,
0 1/u3
with discriminant
∆F 0 = −24 · 33 ∈ Z.
Reducing the coefficients of F 0 mod p = 5 creates the Weierstrass polynomial
2 3
F50 (X 0 ,Y 0 ) = Y 0 − (X 0 + 1) ∈ F5 [X 0 ,Y 0 ]
with discriminant
∆F 0 = 3 ∈ F5 .
5
E p (F pn ), n ∈ N∗ ,
of orders of the groups of F pn -valued points of the elliptic curve E p . These groups
collect those points from E p ⊂ P2 (F p ) with all coordinates in the finite fields F pn .
We recall the following result of Hasse for the case n = 1:
√
1 + p − card E p (F p ) ≤ 2 · p.
For a proof see [33, Theor. 10.5].
1+ p
a p (E) := 1 + p − card E p (F p )
between the average number and the actual number of F p -valued points of E.
3. The sequence
(a p (E)) p prime
card E p (F pn )
for n ≥ 2. The latter will serve as the generators of the ζ -function of E, see
Definition 4.44.
176 4 Elliptic curves
1
ζ (s) = ∏
p 1 − p−s
3. For each N ∈ N expand the finite product by using the geometric series and the
unique prime factor decomposition of integers
!
1 1 1 1
∏ 1 − p−s = ∏ 1 + ps + p2s + ... = ∑ 0 ns
p<N p<N
with the sum on the right-hand side taken over all positive integers with all prime
factors less than N. The limit N → ∞ proves the claim, q.e.d.
From the viewpoint of the ring Z the Riemann ζ -function is a global object.
Because it considers all primes p ∈ Z and collects as a product the local information
encoded in the geometric series of p
1 ∞ n
= ∑ 1/p−s
1 − p−s n=0
We now switch from the primes p of Z to the elliptic curves E p /F p , which are the
fibres of a given elliptic curve E/Q over its stable primes p. First we study the
local information, i.e. the information encoded in E p /F p . The curve E p contributes
as local information at the prime p the sequence of cardinalities of points with
values in the finite fields of characteristic p
card E p (F pn ), n ∈ N∗ .
Remark 4.45 (Properties of the ζ -function of an elliptic curve over F p ). The ζ -function
of an elliptic curve E p /F p
Z(E p , T ) ∈ Q[[T ]]
has the following properties:
1. Rationality: The ζ -function is a rational function over the ring Z, more precisely:
L(E p , T )
Z(E p , T ) = ∈ Z(T )
(1 − T ) · (1 − p · T )
The proof of these properties is due to Weil, [56, Chap. V, Theor. 2.2].
Weil also stated and proved a generalization for projective algebraic curves
curves C p /F p of arbitrary genus. If the curve C p is the reduction mod p of a curve
C/Q
then Weil’s theorem shows the neat relation between the curve C p and the complex
curve X(C) ⊂ Pn : The dimension of the polynomial in the numerator of
the ζ -function of C p is the first Betti number of the compact Riemann surface X/C.
The L-series L(E) of an elliptic curve E/Q collects for each prime p ∈ Z the
local information about the reduction E p . Hence the L-series is a global object.
Definition 4.46 (L-series of an elliptic curve over Q). The L-series of an elliptic
curve E/Q is the function
1
L(E, −) : {s ∈ C : Re s > 3/2} → C, L(E, s) := ∏
p Z(E p , p−s )
rank E(Q) = r
with r equal to the analytic rank defined as the order of the zero of the L-series
at s = 1
r := ord(L(E); 1).
The PARI script Elliptic_curve_04_2 computes the analytic rank for
some elliptic curves E/Q of higher analytic rank, see Figure 4.15.
1 1
−s
= ζ (s) and = ζ (s − 1)
1− p 1 − p1−s
They do not provide new information about the elliptic curve.
3. If one compares the L-series L(E, s) of an elliptic curve E/Q with the Riemann ζ -function ζ (s)
on the level of the two product representations
1 1
L(E, s) = ∏ −s
and ζ (s) = ∏ −s
p Z(E p , p ) p 1− p
1
Z(E p , p−s ) corresponds to
1 − p−s
The geometric series on the right hand side depends only on the prime p, while
the corresponding ζ -function on the left-hand side depends on p and on the re-
duction of the given elliptic curve E/Q.
180 4 Elliptic curves
f : N∗ →
− N∗
f (n · m) = f (n) · f (m)
and the formula for the finite geometric series imply the product formula
k·e1 k·eN
σk (n) = pk1 + pk·2
1 + ... + p1 · ... · pkN + pk·2
N + ... + pN =
181
182 5 Introduction to Hecke theory and applications
ej
! (e j +1)k
N ν N pj −1
=∏ ∑ pkj =∏
j=1 ν=1 j=1 pkj − 1
And the product formula implies for coprime n, m ∈ N∗
Corollary 5.3 (Eisenstein series are multiplicative). For each even k ≥ 4 the
Fourier coefficents of the Eisenstein series
− Bk ∞
· Ek (τ) = ∑ cν · qν , q = e2πi·τ ,
2k ν=0
Proof. We reduce the statement of the corollary to Proposition 5.2: Corollary 3.11
shows
2k ∞
Ek (τ) = 1 − · ∑ σk−1 (ν) · qν , q = e2πi·τ
Bk ν=1
and
− Bk − Bk ∞
· Ek (τ) = + ∑ σk−1 (ν) · qν ,
2k 2k ν=1
∆
(2π)12
Hecke operators are linear maps on modular forms, more precisely: For each
even weight k ∈ N the Hecke algebra
Hk (Γ ) ⊂ End(Mk (Γ ))
Besides the two sequences of levels N ∈ N∗ from Definition 2.9 and weights k ∈ Z
from Definition 3.6 we now introduce a third sequence m ∈ N∗ . It is the sequence
of the sets of integral matrices with determinant = m, see Definition 5.4. These ma-
trices generalize the modular group Γ with respect to the determinant. They serve
to define the Hecke operators which average the values of modular forms of a given
weight and level. In addition, these integral matrices will serve in Section 6.2 to
average as class invariants distinguished values of the j-invariant.
Φm : Γ × Γm → Γm , (A, M) 7→ A · M
its restriction. The respective orbits sets are denoted as the left quotients
Remark 5.5 (Restriction of the Γ -action to the primitive matrices). The subset
Γm,prim ⊂ Γm
I := (αa + β c, αb + β d, γa + δ c, γb + δ d) ⊂ Z
(1) = (a, b, c, d) ⊂ I
which implies
(1) = I
Hence the coefficients of the product matrix A · M are coprime, i.e.
Lemma 5.6 shows that each orbit of the group action can be represented by a
triangular matrix which facilitates many calculations.
Lemma 5.6 (Orbit sets of the left action). Consider a positive integer m ∈ N∗ .
1. A bijective map exists
' ab
Γ \Γm −
→ V (m) := ∈ Γm : 0 ≤ b < d
0d
to a subset of upper triangular matrices. The orbit set of Φm has cardinality σ1 (m).
ψ(m) := m · ∏ (1 + (1/p))
p|m
p prime
α · δ − β · γ = 1.
Then 0 0
a b
A1 · M =
0 d0
and
det (A1 · M) = det M = a0 · d 0 = m.
In particular d 0 6= 0 and possibly after replacing A1 by −A1 even a0 , d 0 > 0.
A2 (k) := T k · A1 ∈ Γ .
Then 0 0 0 0
a b + k · d0
1k a b
A2 (k) · M = · = .
01 0 d0 0 d0
We choose k0 ∈ Z such that
0 ≤ b0 + k0 · d 0 < d 0
and set
A := A2 (k0 ).
ii) The construction from part i) determines uniquely the element
A · M ∈ V (m) :
186 5 Introduction to Hecke theory and applications
Ai · M =: Mi ∈ V (m), i = 1, 2.
Then
M = A−1 −1 −1
1 · M1 = A2 · M2 or A21 · M1 = M2 , A21 := A2 · A1 .
Set
αβ ai bi
A21 =: and Mi := , i = 1, 2.
γ δ 0 di
From
αβ a b a b
· 1 1 = 2 2
γ δ 0 d1 0 d2
follows:
• γ · a1 = 0, hence γ = 0.
• As a consequence a1 = a2 , d1 = d2 , and b1 + β · d1 = b2 .
p : Γm → V (m), M 7→ A · M,
v) The equality
card V (m) = σm (1)
follows at once from the explicit representation of the elements of V (m): Each
coefficient d ≥ 0 satisfies d|m and allows d different coefficients b.
5.1 Hecke operators of the modular group and their eigenforms 187
2. The representation of the orbit set Γ \Γm,prim follows from part 1) by restriction.
For the cardinality of V (m, prim) see [38, Chap. 5, §1].
2. The set
V (m, prim) · γ ⊂ Γm,prim
is a complete set of representatives of the orbit set Γ \Γm,prim .
Proof. 1. Consider a given γ ∈ Γ . We show: For arbitrary but fixed M ∈ V (m) exist
an element
M1 ∈ V (m)
and a unique matrix A ∈ Γ such that
A · M = M1 · γ.
V (m) →
− V (m), [M] 7→ [M · γ],
2. Analogously, q.e.d.
Φle f t : Γ × Γp →
− Γp , (γ, M) 7→ γ · M
188 5 Introduction to Hecke theory and applications
1. Besides the set V (p) from Lemma 5.6 the set of symmetric matrices
p0 1 b
Vsym := ∪ : 0≤b< p
01 b p + b2
V ] := {M ] : M ∈ V }, M ] := (det M) · M −1 ,
2. i) Let
Φright : Γp × Γ →
− Γp , (M, γ) 7→ M · γ,
be the corresponding right action of Γ . For each complete system V of represen-
tatives of Φle f t the set of transposed elements
V > := {v> : v ∈ V }
v := (det w) · w−1 ∈ Γp .
Then
v = w] and det v = (det w)2 · det w−1 = det w and w = (det v) · v−1
A1 , A2 ∈ Γm,prim , m ∈ N∗ ,
A2 = γ1 · A1 · γ2 .
Proposition 5.9 is a special case of the Smith normal form, a result which holds for
matrices over principal ideal domains.
Definition 5.10 (Hecke operators of the modular group). For any positive integer m ≥ 1
the m-th Hecke operator is the C-linear endomorphism of the algebra of modular
forms
190 5 Introduction to Hecke theory and applications
Tm : M∗ (Γ ) → M∗ (Γ )
which is defined on Mk (Γ ), k ∈ N, as
Mk (Γ ) → Mk (Γ ), f 7→ Tm f := ∑ f [M]k .
Γ \Γm
Γ × Γm →
− Γm .
The value f [M]k does not depend on the choice of a representative from the
orbit Γ · M because due to Lemma 3.4: For all α ∈ Γ
f [α · M]k = ( f [α]k )[M]k = f [M]k
E.g., the summation in Definition 5.10 of the Hecke operator Tm may extend over
the elements of V (m) or over the elements of V (m) · γ, γ ∈ Γ . Note T1 = id.
f ∈ Mk (Γ ) =⇒ Tm f ∈ Mk (Γ ).
bn = ∑ rk−1 · amn/r2 , n ∈ N,
r>0
r∈(m,n)
in particular
b0 = a0 · σk−1 (m) and b1 = am
3. The ideal of cusp forms is stable under all Hecke operators, i.e. for each m ∈ N∗
Tm (S∗ (Γ )) ⊂ S∗ (Γ ).
5.1 Hecke operators of the modular group and their eigenforms 191
Due to Theorem 5.11, part 1) we will use the same notation Tm for a Hecke operator
and its restriction to a single modular space of fixed weight.
Proof. 1. To prove that Tm f is a modular form of the same weight as f we have to
verify the properties from Definition 3.6:
• Appararently Tm f is holomorphic on H.
= ∑ f [N]k = ∑ f [M]k = Tm f .
N∈V (m)·γ M∈V (m)
ii) (
2πib·(n/d) d if d|n
∑ e =
0≤b<d 0 otherwise
The first case follows from
e2πib·(n/d) = 1
for each of the d summands with 0 ≤ b < d. The second case follows from
the formula of the finite geometric series with argument e2πi·(n/d) :
d−1 e2πi·d·(n/d) − 1
∑ e2πib·(n/d) = e2πi·(n/d) − 1
=0
b=0
192 5 Introduction to Hecke theory and applications
iii) We have
! !
aτ + b ∞ aτ + b
f (M(τ)) = f = ∑ an · exp 2πin · =
d n=0 d
! !
∞ aτ b
= ∑ an · exp 2πin · · exp 2πin ·
n=0 d d
Hence
d−1 ∞
∑ f (M(τ)) = ∑ d · and · qn(m/d) , q = exp(2πinτ),
b=0 n=0
because
!
d−1 d−1 ∞
2πin(aτ)/d 2πin(b/d)
∑ f (M(τ)) = ∑ ∑ an · e ·e =
b=0 b=0 n=0
!
∞ d−1 ∞
2πin(aτ)/d 2πin(b/d)
= ∑ an · e · ∑e = ∑ and · e2πinaτ · d =
n=0 b=0 n=0
∞ ∞
= d · ∑ and · qna = d · ∑ and · qn(m/d)
n=0 n=0
Due to part ii) the third last sum retains only summands with index = nd.
We now use the representation of the orbit set Γ \Γm from Lemma 5.6 and
insert the Fourier series of f . We obtain due to the formula from part iii)
(Tm f )(τ) = ∑ f [M]k (τ) = mk−1 · ∑ h(M, τ)−k · f (M(τ)) =
M∈V (m) M∈V (m)
!k−1 !
k−1 1 ∞ m ∞
=m ∑ d k · ∑ and · qn(m/d) · d = ∑ d ∑ and · q n(m/d)
=
d>0 n=0 d>0 n=0
d|m d|m
!k−1 !
m ∞
= ∑ d ∑ an · q(n/d)·(m/d)
d>0 n=0
d∈(m,n)
The substitution
provides
!2
d2 · t d tm
n= = tm = .
m m r2
5.1 Hecke operators of the modular group and their eigenforms 193
Hence
∞
(Tm f )(τ) = rk−1 · a(tm)/r2 q t
∑ ∑
t=0 r>0
r|(m,t)
with the inner sum being finite. Apparently, the Fourier series has no coeffi-
cients with negative index. Hence Tm f is holomorphic at ∞ ∈ H∗ .
2. For the proof see the Fourier series from the previous part.
The following corollary specializes the general formula from Theorem 5.11. It
gives the formula for the Fourier coefficients of the Hecke transform for prime in-
dices.
Corollary 5.12 (Fourier coefficients with prime index of Hecke transforms).
Consider a modular form f ∈ Mk (Γ ) with Fourier coefficients (an )n∈N . Then the
Fourier coefficients (bn )n∈N of its Hecke transform Tm f satisfy for each prime
index p
(
amp if p - m
bp =
amp + pk−1 · am/p if p|m
Proof. The claim follows from the third formula for the general case in Theorem 5.11:
• If p - m then r|(p, m) implies r = 1.
∆ ∞
12
= ∑ τm · qm
(2π) m=1
Tm (∆ ) = τm · ∆ .
Proof. According to Corollary 3.19 and Lemma 3.23 the vector space S12 (Γ ) of
cusp forms of weight k = 12 is 1-dimensional. It has the discriminant modular
form ∆ from Definition 3.13 as a basis. Hence ∆ is an eigenvector of each Hecke
operator Tm , m ≥ 1. To calculate the corresponding eigenvalue we recall the Fourier
expansion of the normalized discriminant modular form from Proposition 3.14:
∆ (τ) ∞
= ∑ τn · qn , τ1 = 1.
(2π)12 n=1
Hence by comparing the linear term of the Fourier series on both sides
Tm (∆ ) = τm · ∆ , q.e.d.
Example 5.14 (Hecke operators). Consider the vector space of cusp forms S28 (Γ ).
Due to Theorem 3.19 and Lemma 3.23
dim S28 (Γ ) = 2.
f1 := ∆ · E44 and f2 := ∆ 2 · E4
T2 : S28 (Γ ) →
− S28 (Γ ).
The endomorphism has two distinct eigenvalues, hence it can be diagonalized. Fig-
ure 5.1 shows
• the Fourier series of f1 and f2
5.1 Hecke operators of the modular group and their eigenforms 195
• the eigenvalues of T2
The result from Figure 5.1 has been computed by the PARI script
Hecke_matrix_02.
We give an analogue to the Hecke operators from the theory of quadrics. The
preface of the textbook [17] chooses the analogue as an introduction to the theory
of modular forms. Corollary 5.13 represents the discriminant modular form ∆ as si-
multaneous eigenform of all Hecke operators (Tm )m≥1 acting on the complex vector
space
196 5 Introduction to Hecke theory and applications
V := S12 (Γ ).
The corresponding eigenvalues are the Fourier coefficients of ∆ . Remark 5.15 trans-
lates the law of quadratic reciprocity from algebraic number theory into a similar
shape.
The law of quadratic reciprocity relates the Legendre symbol of two odd primes
p 6= q according to the formula
q
p
p ≡ 1 mod 4 or q ≡ 1 mod 4
p
=
q
− q
p ≡ 3 mod 4 and q ≡ 3 mod 4
p
F(X) := X 2 − d ∈ Z[X]
{x ∈ C : F(x) = 0}.
Fp (X) := X 2 − d p ∈ F p [X]
The number 1 in the definition is the average number of solutions mod p. The
values a p (Q) extend from prime indices to an (Q), n ∈ N, and satisfy the product
rule
anm (Q) = an (Q) · am (Q), n, m ∈ N∗ .
For odd prime p the number a p (Q) equals the Legendre symbol
5.1 Hecke operators of the modular group and their eigenforms 197
d
a p (Q) = .
p
Due to the Law of Quadratic Reciprocity and its two supplements the Legendre
symbol depends only on the residue class p mod 4d.
Set
N = NQ := 4|d|,
and consider the multiplicative group
G := (Z/NZ)∗ ,
V := { f : G → C}.
fQ : G → C, f (n) := an (Q).
Tp ( fQ ) = a p (Q) · fQ .
is commutative. And Theorem 5.26 will show that the Hecke operators are selfad-
joint with respect to a certain Hermitian scalar product on the cusp forms. Hence all
Hecke operators on cusp forms can be simultaneously diagonalized.
Definition 5.16 (Hecke algebra of weight k). The Hecke algebra of even weight k ∈ N
Heckek (Γ ) ⊂ End(Mk (Γ ))
is the subalgebra spanned by the family (Tm |Mk (Γ ))m∈N of Hecke operators.
To prepare the proof of Theorem 5.18 about the Hecke algebra we consider for two
matrices
aj bj
Mj = ∈ V (n j ), j = 1, 2
0 dj
the components of the product
a1 a2 a1 b2 + b1 d2
M1 · M2 = ∈ Γn ·n
1 2
0 d1 d2
Lemma 5.17 (Restsystem for V (n1 · n2 )). For j = 1, 2 consider two coprime num-
bers
n j ∈ N∗ , j = 1, 2.
1. For given positive numbers a j , d j with n j = a j · d j the map of restsystems
is bijective.
V (n1 ) ×V (n2 ) →
− Γn1 ·n2 , (M1 , M2 ) 7→ M1 · M2 ,
Proof. 1. It suffices to show that φ is injective because its domain and codomain
have the same finite cardinality. Therefore assume
i.e.
a1 · b2 + b1 · d2 ≡ a1 · b̃2 + b̃1 · d2 mod d1 d2
Then
a1 (b2 − b̃2 ) ≡ (b̃1 − b1 )d2 mod d1 d2
which implies
a1 (b2 − b̃2 ) ≡ 0 mod d2
Because
gcd(n1 , n2 ) = 1 =⇒ gcd(a1 , d2 ) = 1
therefore
b1 − b̃1 ≡ 0 mod d1 , i.e. b1 ≡ 1̃2 mod d1
which implies
b2 = b̃2 .
Analogously b1 = b̃1 .
2. In addition to part 1) one uses the bijectivity of the maps of divisor sets
Theorem 5.18 (The Hecke algebras are commutative). For each weight k ∈ N
the Hecke algebra Heckek (Γ ) is commutative. It has the following properties:
1. For coprime n, m ∈ N∗ Hecke operators are multiplicative:
Tm ◦ Tn = Tm·n
In particular
Tpr ∈ Z[Tp ]
3. The algebra Heckek (Γ ) is generated by the family Tp , p prime.
2. The proof is by induction. First one replaces Lemma 5.17 by two analogous re-
sults about restsystems. Secondly, one evaluates the formula for the Fourier coef-
ficients of Hecke transforms from Theorem 5.11 and Corollary 5.12, cf. [36, Kap. IV, §2.2].
3. Due to part 1) the family (Tm )m∈N of all Hecke operators generates the same
algebra as the family of Hecke operators with all prime power indices
Tpr , p prime , r ∈ N.
Due to part 2) for each fixed prime p the family Tpr , r ∈ N, generates the same
algebra as the single Hecke operator Tp .
The commutativity of the Hecke algebras follows from the parts 1) and 3) , q.e.d.
Next we introduce a Hermitian scalar product on the vector space of cusp forms of
a given congruence subgroup Γ0 (N).
Lemma 5.20 (Invariance of the hyperbolic volume form). The hyperbolic vol-
ume form is GL(2, R)+ -invariant, i.e. for each γ ∈ GL(2, R)+ holds
γ ∗ dµ = dµ
(γ ∗ dµ)(τ) = dµ(τ) :
i
dx ∧ dy = dτ ∧ dτ
2
5.2 The Petersson scalar product 201
which implies
i dτ ∧ dτ
dµ(τ) = ·
2 (Im τ)2
and
i dγ(τ) ∧ dγ(τ)
(γ ∗ dµ)(τ) = ·
2 (Im γ(τ))2
The proof of Proposition 3.1 and the formulas from Remark 2.8
Im τ dτ
Im γ(τ) = det(γ) · and dγ(τ) = det(γ) ·
|cτ + d|2 (cτ + d)2
imply
(γ ∗ dµ)(τ) = dµ(τ), q.e.d.
independent from the chosen left coset decomposition of Γ with respect to Γ0 (N).
Ω :H→
− C
Z 1/2
dx 1/2
= √ = arcsin x = π/6 − (−π/6) = π/3
−1/2 1 − x2 −1/2
follows from
• the index formula Proposition 2.14,
If one splits H as the disjoint union of the sets A(F2 ), A ∈ K, and observes
that ±1 act as identity on H, which introduces the factor = 1/2, then
Z Z Z
Ω dµ = (1/2) · ∑ Ω dµ = (1/2) · ∑ Ω dµ =
F1 A∈K A(F2 )∩F1 A∈K F2 ∩A (F1 )
−1
Z Z
= (1/2) · ∑ Ω dµ = Ω dµ, q.e.d.
A∈K A(F1 )∩F2 F2
Definition 5.22 (Petersson scalar product). For each level N ∈ N and each even
weight k ∈ N, k ≥ 0, the Petersson scalar product on the cusp space Sk (Γ0 (N)) is
the C-sesquilinear map
1 Z
< f , g >Γ0 (N) := · f (τ) · g(τ) · (Im τ)k dµ(τ)
vol D(Γ0 (N)) D(Γ0 (N))
Note
vol D(Γ0 (N)) < ∞ because [Γ : Γ0 (N)] < ∞.
Definition 5.22, see [17, Chap. 5, Def. 5.4.1], implies: If
Γ1 := Γ0 (N1 ) ⊂ Γ2 := Γ0 (N2 ) and f , g ∈ Sk (Γ2 )
then
< f , g >Γ1 =< f , g >Γ2
While [36, Kap. IV, 3.2], which considers only the modular group Γ , employs in
Definition 5.22 instead before the integral the factor = 1.
Theorem 5.23 (Petersson scalar product). The map from Definition 5.22
φ :H→
− C
To prove that the Petersson scalar product of two modular forms f , g ∈ Sk (Γ0 (N)) is
well-defined, consider the continuous function
i) To prove the boundedness of φ it suffices to prove for any γ ∈ Γ that the restriction
(φ ◦ γ)|D
for a suitable h ∈ N, show that both cusp forms are of type O(qh ) in the limit Im τ → ∞.
Then
|qh | = |e2πiτ/h | = e−2π·(Im τ)/h =⇒ O(q2h ) · (Im τ)k = o(qh )
ii) The function φ is Γ -invariant because for each γ ∈ Γ and τ ∈ H according to
Definition 3.6
φ (γ(τ)) = f (γ(τ)) · g(γ(τ)) · (Im γ(τ))k =
k
= f [γ]k (τ) · h(γ, τ)k · g[γ]k (τ) · h(γ, τ) · (Im τ)k · |h(γ, τ)|−2k =
= f [γ]k (τ) · g[γ]k (τ) · (Im τ)k = f (τ) · g(τ) · (Im τ)k = φ (τ)
Hence part i) and Proposition 5.21, part 3) imply that the Petersson scalar product
is independent from the choice of a fundamental domain, which finishes the proof,
q.e.d.
Note. The proof of Theorem 5.23 shows that the Hermitian scalar
product < f , g >Γ0 (N) is also defined for two modular forms f , g ∈ Mk (Γ0 (N)) if at
least one of them is a cusp form.
Recall from Remark 2.8 the operation of GL(2, Q)+ and its subgroups on H
1 Z Z
· Ω dµ = Ω dµ
[Γ (Ω ) : K] D(K) D(Γ (Ω ))
Γ̃ (p) := {L ∈ Γ : L ≡ ±1 mod p}
has finite index. In addition, define for each pair of continuous functions
h1 , h2 : H →
− C
5.2 The Petersson scalar product 205
the function
Proof. Cf. [36, Kap. IV, §3.3]. Because k is fixed during the whole computation we
simplify the notation [−] := [−]k .
1. Consider a fundamental domain F of Γ (Ω ). Then each decomposition of Γ (Ω )
into the finite number [Γ (Ω ) : K] of left classes with respect to K
˙
Mν−1 · K
[
Γ (Ω ) =
ν
hence Z Z Z
Ω dµ = Ω dµ = ∑ Ω dµ =
D(K) G ν Mν (F )
Z Z
=∑ Ω dµ = [Γ (φ ) : K] · Ω dµ
ν F D(Γ (Ω ))
L = ±1 + p · A, A ∈ M(2 × 2, Z).
Hence
L0 M = ML and L0 ∈ Γ
which implies
( f [M])[L] = ( f [L0 ])[M] = f [M].
Because Γ̃ (p) ⊂ Γ also
g[L] = g.
The two L-invariant elements
f [M] and g
satisfy for τ ∈ H the transformation formula
We compute
= f [M](τ) · h(L, τ)k · g[L](τ) · h(L, τ)k = f [M](τ) · g(τ) · |h(L, τ)|2k
Multiplying by (Im L(τ))k both sides of the transformation formula for
f [M] and g
and using the transformation formula for the imaginary part of the argument
Im L(τ) = (Im τ) · h(L, τ)−2
shows
( f [M] · g)(L(τ)) · (Im L(τ))k = ( f [M] · g)(τ) · (Im τ)k
or
Ω ( f [M], g) ◦ L = Ω ( f [M], g)
Because L ∈ Γ̃ (p) is arbitrary we have shown
The finiteness of the index of Γ̃ (p) in the fixgroups follow from the finiteness
[Γ : Γ̃ (p)] < ∞
5.2 The Petersson scalar product 207
If L ∈ Γ̃ (kp) then
M · L · M ] ≡ ± M · 1 · M ] mod kp = ± p · 1 mod kp
1
M·L· · M ] ≡ ± 1 mod k i.e. M · L · M −1 ∈ Γ̃ (k).
det M
Now setting k = 1 shows
MΓ̃ (p)M −1 ⊂ Γ̃ ,
while setting k = p shows
As a consequence
Hence
[Γ : Γ̃ (p)] = [Γ : K] = [Γ : M −1Γ̃ (p)M],
in particular independent from M ∈ Γp , q.e.d.
Theorem 5.25 (Selfadjointness of the Hecke operators). For each even weight k ≥ 0
the Hecke operators
Tm ∈ End(Sk (Γ )), m ∈ N,
are self-adjoint with respect to the Petersson scalar product, i.e. for all f , g ∈ Sk (Γ )
Proof. Due to Theorem 5.18 it suffices to prove the claim for m = p prime. Due to
Lemma 5.8 there exists a complete set V (p) of representatives of the Γ -action
Φle f t : Γ × Γp →
− Γp
such that also V (p)] , the set of adjoints of the matrices from V (p), is a complete
set of representatives.
by proving for γ ∈ Γ Z
ΩM,le f t dµ exists :
γ(D)
5.2 The Petersson scalar product 209
Similarly to the proof of Theorem 5.23, part ii) one checks on D an analogue of the
transformation formula. First,
Set
γle f t := A−1 ∈ Γ ,
then
M · γ = γle f t · L.
Secondly, using that f and g are modular forms with respect to Γ
= Ω ( f [L]k , g) = ΩL,le f t
As a consequence
Z Z Z
ΩM,le f t dµ = ΩM,le f t ◦ γ dµ = ΩL,le f t dµ
γ(D) D D
Because L is an upper triangular matrix the Fourier series of the cusp form f can
easily be evaluated at L(τ), τ ∈ H, while the automophy factor h(L, −) is a
constant and det L = p. An analogous estimate as in the proof of Theorem 5.23,
part ii) shows the boundedness of the integrand on D. Hence
Z Z
ΩM,le f t dµ = ΩL,le f t dµ < ∞
γ(D) D
because
Γ̃ (p) ⊂ Γ (ΩM,le f t )
has finite index according to Lemma 5.24, part 2). Concerning the indices
Lemma 5.24, part 3) implies
hence also
Γ (ΩM,le f t ) : Γ̃ (p) = Γ (ΩM,le f t ) : M −1Γ̃ (p)M
210 5 Introduction to Hecke theory and applications
Z
= Γ (ΩM,le f t ) : M −1Γ̃ (p)M ·
ΩM,le f t dµ =
D(Γ (ΩM,le f t ))
Z Z Z
ΩM,le f t dµ = ΩM0 ,right ◦ M dµ = ΩM0 ,right dµ
M −1 F M −1 F F
iv) Left-right switching of the Hecke operator:
Z Z
< Tp f , g >= (Tp f · g)(τ) · (Im τ)k dµ = Ω (Tp f , g) dµ
D D
For any pair of cusp forms Tp f and g the integrand is invariant under the action
of Γ as checked during the proof of Theorem 5.23, part i). Hence continuing
Z
1 Z
Ω (Tp f , g) dµ = · Ω (Tp f , g)dµ =
D [Γ : Γ̃ (p)] F
1 Z
1 Z
= · ∑ ΩM,le f t dµ = · ∑ ΩM,le f t dµ
[Γ : Γ̃ (p)] F M∈V (p) [Γ : Γ̃ (p)] M∈V (p) F
Using the result from part iii) and the property of V (p)] the calculation continues
1 Z
1 Z
· ∑ ΩM,le f t dµ = · ∑ ΩM0 ,right dµ =
[Γ : Γ̃ (p)] M∈V (p) F [Γ : Γ̃ (p)] M∈V (p) F
1 Z
1 Z
· ∑ ΩM0 ,right dµ = · Ω ( f , Tp g) dµ =
[Γ : Γ̃ (p)] F M∈V (p) [Γ : Γ̃ (p)] F
Z Z
= Ω ( f , Tp g) dµ = ( f · Tp g)(τ) · (Im τ)k dµ =< f , Tp g >, q.e.d.
D D
Theorem 5.26 (Eigenforms of the Hecke algebra of the modular group). For
each even weight k ≥ 0 the cusp space Sk (Γ ) has a basis of eigenforms of the Hecke
algebra Heckek (Γ ).
Proof. First, Theorem 5.18 implies that Heckek (Γ ) is commutative. Secondly, The-
orem 5.25 implies that all Hecke operators Tm , m ∈ N, are selfadjoint endomor-
phisms of the unitary vector space (Sk (Γ ), < −, − >). Hence the claim follows due
to a result from linear algebra about simultaneous diagonalizability of a family of
pairwise commuting, self-adjoint endomorphisms, q.e.d.
5.2 The Petersson scalar product 211
Remark 5.27 (Hecke theory of congruence subgroups). For each level N ∈ N and
even weight k ≥ 0 one can define Hecke operators on modular forms of Mk (Γ0 (N)).
The corresponding Hecke algebras are commutative, and each vector space of cusp
forms of fixed weight has an eigenbasis with respect to the family
Tm , m ∈ N and gcd(N, m) = 1.
For a proof see [34, Chap. 4, Theor. 4.22], [17], [33, Chap. IX, §6].
One should not consider a Hecke congruence subgroup Γ0 (N) as an isolated object
of a fixed level N. Instead one should focus on the whole family of Hecke
congruence subgroups
Γ0 (N), N ∈ N
For each weight k ∈ N and positive integer m ∈ N the modular spaces of different
levels
Mk (Γ0 (M)) and Mk (Γ0 (m · M))
are related.
Proposition 5.28 (Changing levels (N/m) 7→ N). Consider a weight k ∈ N, a
level M ∈ N and a positive integer m ∈ N. Set
m0
µm := ∈ Γm
0 1
The map
j : Mk (Γ0 (M)) →
− Mk (Γ0 (m · M)), f 7→ f [µm ]k ,
is a well-defined injection and restricts to an inclusion of cusp forms.
Under the assumptions of Proposition 5.28 one has apparently also the injection
→ Mk (Γ0 (m · N)).
Mk (Γ0 (N)) ,−
then
1 1 0 m0 a b/m
µm−1 · γ · µm = · ·γ · = ∈ GL(2, Q)+
m 0m 0 1 cm d
212 5 Introduction to Hecke theory and applications
Claim: For N := m · M
Γ 0 ∩ Γ0 (M) = Γ0 (N).
The inclusion
Γ 0 ∩ Γ0 (M) ⊂ Γ0 (N)
is obvious because
M|c =⇒ m · M|c · m, i.e. N|c · m
To prove the opposite inclusion
Γ0 (N) ⊂ Γ 0 ∩ Γ0 (M) :
Each matrix
a b a b
β= = ∈ Γ0 (N)
cN d cmM d
has the form
a mb
β= µm−1 · γ · µm with γ := .
cM d
Here
det γ = det β = 1, hence γ ∈ Γ0 (M).
As a consequence
A · M ∈ Γm
is triangular. Set
αle f t := A−1 ∈ Γ
3. Weak modularity: Assume f ∈ Mk (Γ0 (M)) and set
g := f [µm ]k .
β = µm−1 · γ · µm
5.2 The Petersson scalar product 213
4. Holomorphy at the cusps: According to the definition one has to show: For each
αle f t ∈ Γ
such that
µm · α = αle f t · θ
with a triangular matrix
ab
θ= ∈ M(2 × 2, Z) ∩ GL(2, Q)+
0d
Hence
g[α]k = f [µm · α]k = f [αle f t · θ ] = ( f [αle f t ]k )[θ ]k
From the Fourier expansion
∞
f [αle f t ]k = ∑ an · q̂n , q̂ = e2πi·τ/h
n=0
Hence g[α]k is holomorphic at ∞. And the last Fourier series shows that j maps
cusp forms to cusp forms.
If we fix the level N then each prime factor p|N defines a level changing
p0
(N/p) 7→ N, using µ p = ∈ Γp .
01
2. The orthogonal complement of Skold (Γ0 (N) with respect to the Petersson scalar
product is named the subspace
of newforms.
Apparently, by the same mappings one can also define the oldspace of modular
forms as a subspace fo Mk (Γ0 (N)). But unfortunately the Petersson scalar product
does not extend from the cusp forms to all modular forms. Hence there is no
orthogonal complement of subspace of old forms.
According to Definition 5.29 one has to consider for the splitting the single prime
factor p = 2 and the cusp forms Sk (Γ0 (1)), i.e. the cusp forms of the full modular
group Γ = Γ0 (1).
∆2 : H →
− C, τ 7→ ∆ (2 · τ)
all cusp forms in S12 (Γ0 (2)) are old forms induced by S12 (Γ ).
• Congruence subgroup Γ0 (4): Similarly Figure 5.3 considers the modular group Γ0 (4)
of level N = 4. For the splitting one has to consider Sk (Γ0 (2)) and the level
change 2 7→ 4. As expected the oldforms of Γ0 (4) contain all cusp forms induced
from the cusp forms of Γ0 (2).
• Congruence subgroup Γ0 (11): The final Figure 5.4 considers the modular group Γ0 (11)
of level N = 11. For the splitting one has to consider the level change 1 7→ 11.
The old forms of Γ0 (11) derive from the modular forms ∆ and j(∆ ).
The present section draws some conclusion from the existence of simultaneous
eigenforms of the Hecke operators. First we prove a conjecture of Ramanujan con-
cerning a certain numerical function. Secondly, we derive the four squares theorem.
Definition 5.31 (Ramanujan τ-function). The Fourier coefficients (τn )n∈N of the
normalized discriminant form
∆ (τ) ∞
= ∑ τn · qn , q = e2πi·τ ,
(2π)12 n=1
τ : N → Z, n 7→ τ(n) := τn .
∆ (τ)
= q − 24q2 + 252q3 − 1.472q4 + 4.830q5 − 6.048q6 + O(7)
(2π)12
Note. There is a typo in equation (104) referring to the factor 2. It should read
|τ(p)| ≤ 2 · p11/2
218 5 Introduction to Hecke theory and applications
The conjecture was proved inter alia one year later by Mordell [42]. Figure 5.6
shows Mordell’s introduction from his paper.
5.3 Numerology: Lagrange, Jacobi, Ramanujan, Mordell 219
Proof. 1. Consider an arbitrary, but fixed m ∈ N. Corollary 5.13 from Hecke theory
derives the eigenvalue equation
!
∆ ∆
Tm 12
= τ(m) ·
(2π) (2π)12
When equating for each fixed n ∈ N the terms of order n on both sides of the
equation the coefficent formula from Theorem 5.11 implies:
For coprime integers m, n we have (m, n) = 1 and the summation reduces to the
single term for r = 1:
τ(m) · τ(n) = τ(m · n)
2. We set m = pk and n = p in the product formula from part 1). Due to
(pk , p) = p
Remark 5.33 (Growth estimation of the Ramanujan τ-function). The final conjec-
ture of Ramanujan about the τ-function is the growth estimation
see Figure 5.5 equation (104) (after correction). This conjecture lies much deeper
than Ramanujan’s other conjectures. Deligne proved the growth condition of the τ-function
as a consequence of his proof of the Weil conjectures in 1974.
5.3 Numerology: Lagrange, Jacobi, Ramanujan, Mordell 221
The next application uses information encoded by modular forms from M2 (Γ0 (4)).
The “Four squares theorem” answers the question: How many possibilities exist to
represent a positive integer as the sum of four integer squares?
Fermat recalls in a letter to a friend from 1659 those results from arithmetic
which he considers his most important contribution to this field. One of these results
is the claim that any natural number can be written as the sum of four squares. The
first proof of this theorem has been given by Lagrange in 1770. The more refined
version of Theorem 5.38 is due to Jacobi in 1834. For the history of these issues
see [47].
Note that r(n, k) counts k-tuples with positive and negative components as dif-
ferent, and distinguishes also k-tuples with the same components, but in different
order.
We will see that the representation of integers as sum of squares is related to the
congruence subgroup Γ0 (4), the relevant space of modular forms is M2 (Γ0 (4)).
Proposition 5.36 (Generating function as modular form). The generating function θ (−, 4)
is a modular form of weight 2 of the congruence subgroup Γ0 (4), i.e.
Proof. Scetch, see cf. [17, Chap.1, §2; Chap. 4, §9], [36].
Then follows
! ! v ! !
u
τ −1 u 1 1
ϑ =ϑ = t2i +1 ·ϑ − −1 =
4τ + 1 4(−1/(4τ) − 1) 4τ 4τ
v ! ! v !
√
u u
u 1 1 u 1
= t2i +1 ·ϑ − = t2i + 1 · (−2iτ) · ϑ (τ) = 4τ + 1 · ϑ (τ),
4τ 4τ 4τ
and exponention
ϑ (−, 4) = ϑ 4
shows the transformation formula
!
τ
ϑ , 4 = (4τ + 1)2 · ϑ (τ, 4) i.e. ϑ (−, 4)[γ]2 = ϑ (−, 4)
4τ + 1
shows
ϑ (τ, 4) = (1 + 2 · q)4 + O(2) = 1 + 8 · q + O(2), q.e.d.
Theorem 5.37 (The modular form G2,N ∈ M2 (Γ0 (N))). Denote by G2 the Eisen-
stein series from Remark 3.12. For each level N ≥ 1 set
N0
µN := ∈ M(2 × 2, Z) ∩ GL(2, Q)+
0 1
The function
G2,N := G2 − G2 [µN ]2 : H →
− C
is a modular function
G2,N ∈ M2 (Γ0 (N))
Proof. By definition
i) Holomorphy: Due to Remark 3.12 the functions G2 and hence also G2,N are holo-
morphic in H ∪ {∞}.
Here β ∈ Γ because
c ≡ 0 mod N =⇒ c/N ∈ Z
Then
224 5 Introduction to Hecke theory and applications
Then
((G2 [µN ]2 )[α]2 )(τ) = (G2 [µN · α]2 )(τ) = (G2 [β · µN ]2 )(τ) =
and
h(α, τ) = cτ + d
which implies
h(β , µN (τ)) = h(α, τ)
Hence
((G2 [µN ]2 )[α]2 )(τ) = (G2 [µN ]2 )(τ) − 2πi · c · h(α, τ)−2
r(n, 4) = 8 · ∑ d
d|n, 4-d
i) Dimension: dim M2 (Γ0 (4)) = 2 according to Theorem 3.26, see also Figure 5.3.
ii) Basis: According to Theorem 5.37 for each N ≥ 2 the Eisenstein series
The Fourier expansion of G2 from Remark 3.12 provides for the two Eisenstein
series G2,2 and G2,4 the Fourier expansions
π2 ∞ π 2
G2,2 (τ) = − 1 + 24 · ∑ d qn = − (1 + 24q + O(2))
3 ∑ 3
n=1 d|n
d odd
and
! !
∞
2
G2,4 (τ) = −π 1+8· ∑ ∑ d q n
= −π 2 (1 + 8q + O(2)).
n=1 d|n, 4-n
Comparing terms of order 0 and 1 in the Fourier expansions shows that both
Eisenstein series are linearly independent. Hence the family (G2,2 , G2,4 ) is a basis
of M2 (Γ0 (4)). In particular, it generates θ (−, 4).
r(n, 4) = 8 · ∑ d, q.e.d.
d|n, 4-d
Example 5.39 (Modular forms via ϑ -functions). Figure 5.7 shows for some pairs (k, e)
the numbers
r(n, k, e).
Here r(n, k, e) ∈ N is the number of different representations of n as a sum of k
summands, each an e-power. These numbers are the Fourier coefficients of suitable
modular forms. The case
r(n, 4, 2) =: r(n, 4)
has been investigated in Theorem 5.38. The figure shows the output of the PARI-
script Modular_forms_theta_01.
226 5 Introduction to Hecke theory and applications
The modular j-invariant parametrizes the set of compex tori by the values of the
affine part of the modular curve X(Γ ) ' P. All complex analytic properties of a
given torus with normalized period lattice (1, τ) are encoded in the value j(τ) ∈ C
of the modular j-invariant. But j(τ) is too coarse to discriminate between the arith-
metic properties of the members of a class of biholomorphic equivalent tori.
Section 6.2 will investigate for tori with complex multiplication: Which arith-
metic information can be obtained when considering besides j(τ) also the values
j(α(τ)), α ∈ Γm,prim , m ∈ N∗ ?
References for the present chapter are [66, Chap. 6.1] and [50].
As a first arithmetic property of complex tori and elliptic curves we observe: Com-
plex tori with normalized lattice generated by (1, τ) relate for certain numbers τ ∈ H
to an imaginary quadratic field. How to characterize tori with those arithmetic prop-
erties? The answer is given by Theorem 6.6.
229
230 6 Application to imaginary quadratic fields
2. The Galois group Gal(K/Q) of a number field K is the group of all field auto-
morphism of K which pointwise fix the elements of Q.
n = [K : Q]
1. Denote by
(α j ) j=1,...,n
a Q-basis of K. For each α ∈ K the multiplication
µα : K →
− K, x 7→ α · x,
µα (α j ) = ∑ α jk · αk , j = 1, ..., n.
k=1,...,n
∆K := det(trK (α j · αk )) ∈ Q
The discriminant depends on the choice of the basis: It transforms with the square
of the determinant of the base change matrix.
OK := {x ∈ K : x integral over Z}
and " √ #
∆K + ∆K
OK = Z
2
7. Each number field K has a primitive element, i.e.
K = Q(α)
Note. Due to historical reasons the symbol O has two different meanings. In the
context of complex analysis O also denotes the complex structure sheaf.
(T, 0) = C/Λ
is a holomorphic map
f : (T, 0) →
− (T, 0)
with f (0) = 0.
Recall from Proposition 2.3 that each endomorphism f of a torus fits into a
commutative diagram
µα
C C
p p
f
C/Λ C/Λ
T = C/Λ
constitutes with respect to addition and composition a ring End(T ), the endomor-
phism ring of T . The map
End(T ) →
− {α ∈ C : α · Λ ⊂ Λ }, f 7→ α f ,
is an isomorphism of rings.
If not stated otherwise by the above ring isomorphism we identify the endomor-
phism ring End(T ) with a subring of C. Because each torus T is an Abelian group
its endomorphism ring End(T ) comprises at least the group Z:
Z ⊂ End(T ) ⊂ C.
End(T ) = Z
is the general case, justifying a specific name for tori with additional endomor-
phisms.
Definition 6.5 (Complex multiplication (CM)).
1. A complex number τ ∈ H is a CM-point and its class [τ] ∈ X(Γ ) in the modular
curve is a CM-class, if the corresponding torus
T = C/Λ , Λ = Z ⊕ Z · τ,
satisfies
Z ( End(T )
2. The values j(τ) ⊂ C at CM-points τ ∈ H are named singular moduli or singular
values of the modular invariant j.
Theorem 6.6 states a remarkable relation between tori and number fields. It re-
veals the arithmetic structure encoded in a complex torus with complex multiplica-
tion.
Theorem 6.6 (Complex multiplication and imaginary quadratic fields). Con-
sider a point τ ∈ H.
1. Then: τ is a CM-point ⇐⇒ Q(τ) is an imaginary quadratic number field.
T := C/Λτ , Λτ := Z + Z · τ,
End(T ) ⊂ C
an order of K.
Note that any order of an imaginary quadratic number field is a lattice.
Proof. 1. i) Consider a CM-point τ ∈ H. By definition exists α ∈ C \ Z with
α · Λτ ⊂ Λτ .
We obtain a matrix
ab
γ= ∈ M(2 × 2, Z)
cd
with
1 1
α· =γ· .
τ τ
234 6 Application to imaginary quadratic fields
We obtain
α = a + b · τ, α · τ = c + d · τ,
which implies b 6= 0, because α ∈
/ Z, and
b · τ 2 + (a − d) · τ − c = 0.
Because τ ∈ H is not a real number, the number field Q(τ) is imaginary quadratic.
ii) Assume that τ ∈ H belongs to an imaginary quadratic number field. Then for
suitable integers a, b, c ∈ Z with a 6= 0 because τ ∈ H
a · τ 2 + b · τ + c = 0.
From
1 0 a 1
(aτ) · = ·
τ −c −b τ
follows aτ ∈ End(T ), ατ ∈
/ Z, and T has complex multiplication,
with
1 1
α· =γ·
τ τ
Therefore α is an eigenvalue of γ and satisfies the eigenvalue equation, which is
an integral equation:
α 2 − (d + a) · α + ab − cd = 0.
Z ( End(T ) ⊂ OK
Because OK is free of rank = 2 the same holds true for End(T ), and the
subring End(T ) is a sublattice of OK , i.e. an order, q.e.d.
We consider in detail the two elliptic points of the modular group Γ , see
Theorem 2.16:
Example 6.7 (CM-points).
1. Elliptic point i: The point i ∈ H determines the normalized lattice of Gaussian
integers
6.1 Imaginary quadratic fields and tori with complex multiplication 235
Λi = Z + Z · i
The torus T = C/Λi has the endomorphism ring
End(T ) = Λi
j(i) = 1728 = 26 · 33 ∈ Z.
Λρ = Z + Z · ρ
End(T ) = Λρ
End(T ) = Λρ
j(ρ) = 0 ∈ Z.
T = C/Λ2ρ
Λ2ρ = Z + Z · 2ρ
α ∈ Λ2ρ .
ρ 2 + ρ + 1 = 0.
236 6 Application to imaginary quadratic fields
j(2ρ) = 54000 ∈ Z.
The present section takes up the study of imaginary quadratic fields K from the view-
point of modular forms. We relate complex multiplication to the value attainment of
the modular j-function: Theorem 6.14 proves that the value j(τ) ∈ C is an algebraic
integer for any CM-class [τ] ∈ X(Γ ). The main tool to derive the necessary proper-
ties of the modular j-invariant are the modular polynomials from Definition 6.8.
The modular polynomials are a family of polynomials (Fm (X))m∈N∗ which extract
the properties of the j-invariant j ∈ M (H) from its transformation behaviour with
respect to the operation of the modular group Γ
Γ × Γm,prim →
− Γm,prim
There is the following analogy to the family of Hecke operators (Tm )m∈N∗ : The
latter extract the Fourier coefficients of a given modular f ∈ Mk (Γ ) from the
transformation behaviour of f with respect to the operation of the modular group Γ
Γ × Γm →
− Γm
on the set of matrices Γm . The Hecke operators are the family of linear maps with
Tm ( f ) := ∑ f [M]k
M∈V (m)
Fm (X) = ∏ (X − j ◦ M) ∈ M (H)[X]
M∈V (m,prim)
6.2 Modular polynomials 237
defined as the product over a complete set V (m, prim) of representatives. Note the
similarity between the definition f [M]k for modular forms from Definition 3.4 and
the expression j ◦ M for the automorphic function j.
ψ(m) := m · ∏ (1 + (1/p))
p|m
p prime
Definition 6.8 (Modular polynomials). Consider m ∈ N∗ and the complete set V (m, prim)
of representatives of Γ \Γm,prim .
1. Each matrix M ∈ V (m, prim) induces the automorphism of the upper half-plane
M : H → H, τ 7→ Mµ (τ).
Fm (X) := ∏ (X − jM ) ∈ M (H)[X].
M∈V (m,prim)
Note. The modular invariant is constant along the orbits of the action of the
modular group Γ . Therefore the value of jM in Definition 6.8 does not depend
on representing a coset from Γ \Γm,prim by the matrix M ∈ V (m, prim) or by the
matrix γ · M, γ ∈ Γ .
1. m = 2:
ψ(2) = 3
and
F2 (X, j) = X 3 +
+(− j2 + 1488 ∗ j − 162000) ∗ X 2 +
+(1488 ∗ j2 + 40773375 ∗ j + 8748000000) ∗ X+
+( j3 − 162000 ∗ j2 + 8748000000 ∗ j − 157464000000000)
2. m = 3:
ψ(3) = 4
and
F3 (X, j) = X 4 +
+(− j3 + 2232 ∗ j2 − 1069956 ∗ j + 36864000) ∗ X 3 +
+(2232∗ j3 +2587918086∗ j2 +8900222976000∗ j +452984832000000)∗X 2 +
+(−1069956∗ j3 +8900222976000∗ j2 −770845966336000000∗ j +1855425871872000000000)∗X+
+( j4 +36864000∗ j3 +452984832000000∗ j2 +1855425871872000000000∗ j)
The roots of the modular polynomial Fm (X) ∈ M (H)[X] are exactly the class
invariants jM ∈ V (m, prim) of the operation
Γ × Γm,prim →
− Γm,prim
We denote by sµ , j = 1, ..., ψ(m), the elementary symmetric functions in ψ(m)
variables. Then up to sign the functions
is the coefficient of the term X ψ(m)−1 . Theorem 6.10 now takes a closer look on
these coefficients and proves: They are polynomials in the modular j-invariant with
integer coeffcients, i.e. the coefficients belong to the subring
Z[ j] ⊂ M (H).
Proof. For arbitrary but fixed index µ = 1, ..., ψ(m) we consider the coefficient
Fm (X) ∈ M (H)[X].
We prove step by step the following properties of the coefficient c ∈ M (H) until
arriving at c ∈ Z:
i) Weakly modular: According to Corollary 5.7 right multiplication with an
element γ ∈ Γ permutes the orbits of the Γ -left operation Φm
{[M1 ], ..., [Mψ(m) ]} = {[M1 · γ], ..., [Mψ(m) · γ]}.
Hence right multiplication does not change the value of the elementary symmetric
functions: For all j = 1, ..., ψ(m)
c(τ) = ∑ aν · qν , q = e2πiτ .
ν∈Z
We have to show that the singularity q = 0 is a pole of c. For the proof we estimate
the growth of c: According to Corollary 3.21 the modular invariant has the Fourier
expansion
1
j(τ) = + P(q)
q
with P ∈ Z{q} a convergent power series with integer coefficients. For an upper
triangular matrix
ab
α= ∈ V (m, prim)
0d
the Fourier series of the class invariant j ◦ α derives from the Fourier series of j by
substituting the variable q = e2πiτ by
The equality
ad = m
implies
2
e2πiα(τ) = ζmab · (q1/m )a with ζm := e2πi/m
a primitive m-th root of unity. In particular
1
|( j ◦ α)(τ)| ≤ + O(1)
|q|a/d
1
|c(τ)| ≤ M ·
|q|k
iv) Integrality c ∈ OK [ j]: Due to the part already proved, Corollary 3.22 implies
that c is a polynomial with complex coefficients in the modular invariant, i.e.
c ∈ C[ j].
We show that even
c ∈ OK [ j]
with
K := Q(ζm )
the m-th cyclotomic number field. For the proof we recall from part iii) the substi-
tution of q by
6.2 Modular polynomials 241
2
ζm ab · (q1/m )a .
As a consequence the class invariant j ◦ α has a Fourier expansion with respect
to q1/m , namely
1
1/m a2 ab
j(α(τ)) = 2 ab
+ P (q ) · ζm ,
(q1/m )a · ζm
which shows that the Fourier coefficents of c belong to K. According to Corollary 3.22
the coefficients of the polynomial c ∈ C[ j] are Z-linear combinations of the Fourier
coefficients of c. Hence the coefficients are Z-linear combinations of the m-th roots
of unity. Therefore the coefficients of c ∈ C[ j] are algebraic integers from OK ,
i.e. c ∈ OK [ j].
with
0 ≤ bχ < d and bχ ≡ e · b mod d.
Because the function c is fixed under the operation of Gal(K/Q), its Fourier co-
efficients belong to the fixed field Q. Due to part iv) the Fourier coefficients even
belong to
Q ∩ OK = Z, q.e.d.
Fm (X) ∈ Z[ j][X]
on the diagonal {X = j}, i.e. we consider the polynomials of the single variable j
242 6 Application to imaginary quadratic fields
Φm ( j) := Fm ( j, j) = ∏ ( j − jM ) ∈ Z[ j].
M∈V (m,prim)
Example 6.11 (The polynomials Φm ( j) ∈ Z[ j]). We take up Example 6.9 with Figure 6.2,
which shows some polynomials Φm . The PARI script Modular_polynomial_02
computes their factorization:
1. m = 2:
2. m = 3:
−769939996672000000 ∗ j2 + 3710851743744000000000 ∗ j
splits into the factors
3. The example Φ5 ( j) shows: In general Φm ( j), m prime, does not split into linear
factors over Z.
ck ck−1
Φm ( j)(τ) = + + higher terms
qk qk−1
Because j has a pole at τ = ∞ of order = 1, the coefficient ck of the Fourier series is
also the leading coefficient of Φm as polynomial in j. The coefficient is the product
of the highest negative coefficients of the factors
j − jµ , µ = 1, ..., ψ(m).
1
j(τ) = + P(q)
q
implies !
aτ + b e−2πi(b/d)
a/d 2πi(b/d)
jµ (τ) = j = + P q · e
d qa/d
with jµ corresponding to the matrix
ab
∈ V (m, prim).
0d
Note that the case a = d is excluded because m = a · d is squarefree. In any case the
highest negative coefficent of
j − jµ
is a root of unity, and the same holds for the coefficient ck of the product
244 6 Application to imaginary quadratic fields
Φ= ∏ ( j − jM ).
M∈V (m,prim)
Proposition 6.13 will show that Fm ( j, X) is symmetric in both variables, i.e. the
coefficients satisfy
crs = csr .
As a consequence, the highest exponent of j in Fm ( j, X) is ψ(m).
Fm ( j, X) = Fm (X, j).
K := Q(R) = Q( j).
Fm (X) = ∏ (X − jM ) ∈ R[X]
M∈V (m,prim)
They are pairwise distinct which can be shown by comparing their Fourier coeffi-
cients. To prove that Fm (X) is an irreducible polynomial we show that the Galois
group
Gal(L/K)
6.2 Modular polynomials 245
Elements of the Galois group can be obtained as follows: Any matrix γ ∈ Γ defines
the automorphism
L = K( j1 , ..., jψ(m) ) →
− L, f 7→ f ◦ γ.
Consider two class invariants
jµ , jν , 1 ≤ µ, ν ≤ ψ(m)
Mµ , Mν ∈ V (m, prim)
jν = j µ ◦ γ
γ 0 · Mν = Mµ · γ,
which is achieved by Proposition 5.9, part 2). Therefore Gal(L/K) operates transi-
tively on the roots of Fm (X), and the latter polynomial is irreducible over K.
ii) Common root of Fm ( j, X) and Fm (X, j): We consider the two polynomials
jM := j ◦ M
with
m0
β := ∈ Vm,prim
0 1
Also f ( jβ ) = 0 because β ∈ Vm,prim . Hence both polynomials f (X), g(X) ∈ R[X]
have the common root jβ .
iii) Symmetry: Due to Lemma 6.12 and part i) the polynomial f (X) ∈ K[X] is
irreducible with leading coefficient ±1. Hence f (X) is the minimal polynomial of
the field extension K ⊂ K[ jβ ], and g(X) is a multiple of f (X), i.e.
Fm ( j, X) = H(X, j) · H( j, X) · Fm ( j, X),
H(X, j) · H( j, X) = 1,
and
H(X, j) = H( j, X) = ±1.
In case H(X, j) = −1 the previous equation
Fm ( j, X) = H(X, j) · Fm (X, j)
would imply
Then
Fm ( j, j) = 0
which contradicts the fact that Φm ( j) = Fm ( j, j) has leading coefficient ±1 accord-
ing to Lemma 6.12. Therefore
H(X, j) = 1
and
Fm ( j, X) = Fm (X, j), q.e.d.
We now derive from the theory of modular polynomials that all singular values
j(τ) are algebraic integers.
Theorem 6.14 (Singular values of the modular j-invariant). For each CM-point τ ∈ H
the j-value j(τ) ∈ C is an algebraic integer. More precisely: There exists m ∈ N∗
such that j(τ) is a root of the polynomial Φm (X) ∈ Z[X].
OK ⊂ K
OK = Z ⊕ Z · z
with a Z-basis (z, 1), z ∈ H. For admissible values of z see Remark 6.2.
N(ξ ) = ξ ξ ∈ Z
OK →
− OK , x 7→ ξ · x.
Denote by
ab
α= ∈ M(2 × 2, Z) ∩ GL(2, R)+
cd
its matrix with respect to the basis (z, 1) of OK . Then
z z
ξ· =α·
1 1
or in components
ξ ·z az + b
=
ξ ·1 cz + d
• The first form shows: The matrix
α ∈ M(2 × 2, Z)
m := det α = ξ ξ = N(ξ ) ∈ Z
a·w+b
H→
− H, w 7→ α(w) :=
c·w+d
the matrix matrix α ∈ GL(2, R)+ satisfies
a·z+b ξ ·z
α(z) = = =z
c·z+d ξ ·1
ii) j(z) is an algebraic integer: Lemma 5.6 implies the existence of a matrix
Φm,prim : Γ × Γm,prim →
− Γm,prim , (A, M) 7→ A · M,
w.l.o.g. µ = 1. Then
j ◦ α = j ◦ α1
because the value of j is constant along each orbit of the Γ -operation. Due to part i)
Φm ( j) = ∏ ( j − jM ) ∈ Z[ j]
M∈V (m,prim)
vanishes at z ∈ H:
Φm ( j)(z) = Φm ( j(z)) = 0.
The polynomial
Φm ( j) ∈ Z[ j]
has integer coefficients, and leading coefficient ±1 due to Lemma 6.12. Hence
0 = Φm ( j(z))
Z[ j(z)] ⊂ C
τ ∈ K = Q+Q·z
has a representation
az + b
τ = β (z) :=
d
with a primitive matrix
ab
β= ∈ Γm,prim , m := a · d.
0d
Alike to the argument from part ii) the Γ -orbit of β passes through a matrix αµ ∈ V (m, prim)
which proves
j(τ) = j(β (z)) = j(αµ (z)) = jµ (z).
Theorem 6.10 implies that jµ is integral over the ring Z[ j]. Hence
250 6 Application to imaginary quadratic fields
j(τ) = jµ (z)
iv) j(τ) is an algebraic integer: When combining part iii) and part ii) the
transitivity of integral dependence proves the integrality of j(τ) over Z, i.e. τ is an
algebraic integer, q.e.d.
Example 6.15 (Singular values of the modular j-function). Continuing Example 6.11
we consider a CM-class [τ] ∈ X(Γ ) and choose the level
m = 2.
Φ2 ( j) ∈ Z[ j]
Which property is expressed by the three factors, why are the values j ∈ {8000, 1728, −3375}
related?
α j ∈ Γ2,prim , j = 1, 2, 3,
•
1 1
α2 = has fixed point z2 = i
−1 1
which defines the lattice
6.3 Fractional ideals and the class number formula 251
√
Λ2 :=< 1, i >= OK2 ⊂ K2 := Q( −1)
•
√
0 −1
α3 = has fixed point z3 = i 2
2 0
which defines the lattice
√ √
Λ3 :=< 1, i 2 >= OK3 ⊂ K3 := Q( −2)
One knows that all three fields have class number hK = 1, which will confirm
Corollary 6.23.
A further application of the modular j-invariant to number theory is the class num-
ber formula for imaginary quadratic fields.
We first recall the divisor sequence of a Riemann surface, e.g., cf. [63].
Remark 6.16 (Class group of a compact Riemann surface). The divisor sequence
on a compact Riemann surface X
− O∗ →
1→ − M∗ →
− D→
− 0
because
252 6 Application to imaginary quadratic fields
Div(X)
H 0 (X, O ∗ ) = C∗ , H 0 (X, D) = Div(X), Cl(X) := and H 1 (X, M ∗ ) = 0
im[M ∗ (X) → Div(X)]
between the group of divisor classes and the Picard group of isomorphism classes
of holomorphic line bundles on X.
For number fields an analogue of divisors are fractional ideals, see [4, Chap. 9].
d · a ⊂ OK .
a = (α) := OK · α
a·b
a−1 := {x ∈ K : x · a ⊂ OK }
Remark 6.18 (Fractional ideals). Consider a number field K with number ring OK .
6.3 Fractional ideals and the class number formula 253
a 6= {0} ⊂ K
the inverse a−1 is also a fractional ideal, see [4, Theor. 9.8]. It satisfies
a · a−1 = (1)
3. With respect to multiplication the set of fractional ideals is an Abelian group J(K).
Its quotient by the subgroup of principal fractional ideals is named the ideal class
group or simply class group of K
Cl(K) := J(K)/im[K ∗ →
− J(K)]
a2 = a · a1 for a suitable a ∈ K ∗ .
− OK∗ →
1→ − K∗ →
− J(K) →
− Cl(K) →
− 1
hK = 1
A theorem from algebraic number theory states the finiteness of the class number
hK := ord Cl(K)
see [40, Chap. 5, Cor. 2], [32, Chap. 12, §2]. The result is an analogue to the fact
that the Neron-Severi group
5. Comparing the exact sequence of a number field with the exact divisor sequence
of a Riemann surface from Remark 6.16 one easily observes the analogy between
corresponding objects:
between the ideal class group Cl(K) and the set of classes of biholomorphically
equivalent complex tori with ring of endomorphisms OK . The bijection is induced
by the attachement
6.3 Fractional ideals and the class number formula 255
Λ ⊂K⊂C
we may assume
Λ = Z+Z·ω
a normalized lattice because the ideal class is determined up to an element of K.
Due to Proposition 6.4 the torus
T := C/Λ
End(T ) = {α ∈ C : α · Λ ⊂ Λ }
OK ⊂ End(T ).
α ·1 = α ∈ Λ.
OK = End(T ) = {α ∈ C : α · Λτ ⊂ Λτ }
provides a matrix
ab
∈ M(2 × 2, Z)
cd
with
1 ab 1
ω· = · .
τ cd τ
Hence
ω = a + b · τ and ω · τ = c + d · τ
256 6 Application to imaginary quadratic fields
1, ω, τ, τ · ω ∈ Λτ
b · Λτ ⊂ Z + Z · b · τ ⊂ Z + Z · ω = OK
iii) The two constructions from part i) and ii) are inverse to each other, q.e.d.
f :Λ →
− R
R = f (Z + Z · τ) = Z + Z · f (τ) = Z + Z · τ, q.e.d.
OK = Z + Z · τ
The complex number j(τ) ∈ C is an algebraic integer due to Theorem 6.14.
Hence Q( j(τ)) is a number field. We show the estimate
[Q( j(τ)) : Q] ≤ hK .
[Q( j(τ)) : Q] = hK
See also [55, Chap. II, Theor. 4.3]. The book uses the notation
OK = Z + Z · τ, τ ∈ H,
its number ring. Then the class number hK satisfies the estimate
[Q( j(τ)) : Q] ≤ hK .
T := C/Λ
ii) The number field Q( j(τ)): Theorem 6.6 implies that the complex number τ ∈ H
from part i) is a CM-point, and Theorem 6.14 concludes that j(τ) ∈ C is an
algebraic integer. Define the number field
L := Q( j(τ))
and denote by
m := [L : Q]
258 6 Application to imaginary quadratic fields
φi : L →
− C, i = 1, ..., m,
with pairwise distinct values φi ( j(τ)). Each extends to an element of the Galois
group Gal(C/Q). That’s made possible by extending elements from the Galois
group over Q along algebraic field extensions and along a pure transcendental field
extension: A given element
φ ∈ {φi : i = 1, ..., m}
φ 1 ∈ Gal(Lsp /Q)
with Lsp ⊂ C the splitting field of L over Q. The element φ 1 extends to an element
φ 2 ∈ Gal(Q/Q)
φ 3 ∈ Gal(K/Q).
σ ∈ Gal(C/Q)
σi ∈ Gal(C/Q), i = 1, ..., m,
iii) Functorial action of the Galois group: For the elliptic curve
E 7→ E σ
6.3 Fractional ideals and the class number formula 259
− E 0 defines a
from part ii) extends to a functor, i.e. each morphism Φ : E →
canonical morphism
− E0
σ
Φσ : Eσ →
with
id σ = id and (Φ ◦Ψ )σ = Φ σ ◦Ψ σ
For the proof represent a given holomorphic map
− E 0 , Φ(O) = O,
Φ :E →
F := µa : C →
− C, z 7→ a · z,
via the corresponding uniformization of the elliptic curves by complex tori C/Γ
and C/Γ 0 from Theorem 4.28.
F
C C z 6= 0 a · z, z 6= 0
Φ
E E0 (1 : ℘Γ (z) : ℘Γ0 (z)) (1 : ℘Γ 0 (a · z) : ℘Γ0 0 (a · z))
− E0
Φ :E →
because
a·Γ ⊂ Γ 0
and the functions ℘Γ 0 and its derivative have the period Γ 0 . The morphism Φ is
locally represented by a pair of rational maps
!
p1 p2
,
q1 q2
− End(E σ ), Φ 7→ Φ σ ,
End(E) →
End(E σ ) = End(E).
F(X,Y ) = Y 2 − (4 · X 3 − g2 · X − g3 )
σ (g2 )3
j(E σ ) = 1728 · = σ ( j(τ))
σ (g2 )3 − 27 · σ (g3 )2
( j(E σi ) = σi ( j(τ))i=1,...,m
(E σi )i=1,...,m
with these j-invariants are pairwise not biholomorphic equivalent. For i = 1, ..., m
holds
End(E σi ) = OK
due to part iii). Proposition 6.20 implies that the corresponding ideal classes
in Cl(K) are pairwise distinct. Their number m is bounded by the class number
hK = card Cl(K).
Hence
m = [L : Q] ≤ hK , q.e.d.
OK = Z + Z · τ
Then j(τ) ∈ Z.
Proof. The proof follows from Theorem 6.14 and Theorem 6.22, q.e.d.
Chapter 7
Outlook: Modular elliptic curves, monstrous
moonshine
The relation between elliptic curves E/Q and modular forms is given by Wiles’
theorem, the proof of the Shimura-Taniyama-Weil conjecture for semistable elliptic
curves. Wiles’s modularity theorem has been subsequently extended to all elliptic
curves E/Q by Breuil, Conrad, Diamond and Taylor. The modularity theorem re-
lates for each elliptic curve E/Q the numbers
A first version of the relation between rational elliptic curves and the theory of
modular forms is Conjecture 7.4.
X0 (N) →
− E(C)
For the proof of the modularity theorem, Theorem 7.2, by Wiles in the stable and
semistable case see the symposion [15]. Concerning the remaining cases subse-
quently proved by several authors see the announcement [16] and the proof in [10].
261
262 7 Outlook: Modular elliptic curves, monstrous moonshine
When comparing rational elliptic curves E/Q and cusp forms f ∈ Sk (Γ0 (N)) one
compares the L-series of both mathematical objects. For the elliptic curve E/Q
recall the L-series L(E, −) arising from the numbers (a p (E)) p prime , see
Definitions 4.44 and 4.46, and also Remark 4.45.
1 1
s
· Γ (s) · L( f , s) = (−1)k/2 · · Γ (k − s) · L( f , k − s)
(2π) (2π)k−s
The textbook [17] does not prove Theorem 7.2. But it gives an excellent
introduction to the problem and it highlights different views onto the theorem. In
joint work with his coworkers, one of the authors of [17] proved in [10] the final
step of the modularity theorem.
1728 · g32
jE = ∈Q
g32 − 27 · g23
7.1 Modular elliptic curves 263
X0 (N)Q →
− E/Q,
3. Jacobi torus: For each elliptic curve E/C with rational modular invariant
1728 · g32
jE = ∈Q
g32 − 27 · g23
Jac(X0 (N)) →
− E,
from the Jacobi torus of a modular curve X0 (N), N ∈ N, which is a group homo-
morphism, see [17, Theor. 6.1.3].
4. Cusp form, version with coefficients: For each elliptic curve E/Q with conductor NE ∈ N
exists a new form
∞
f (q) = ∑ an ( f ) · qn ∈ S2 (Γ0 (NE ))
n=1
a p ( f ) = a p (E),
Recall from Definition 4.40 for an elliptic curve E/Q with global minimal Weier-
strass polynomial F ∈ Z[X,Y ] the conductor
cond E = ∏ pe p ∈ Z,
p
5. Cusp form, version with L-series: For each elliptic curve E/Q with conductor NE ∈ N
exists a new form
∞
f (q) = ∑ an ( f ) · qn ∈ S2 (Γ0 (NE ))
n=1
264 7 Outlook: Modular elliptic curves, monstrous moonshine
with
L( f , −) = L(E, −),
see [17, Theor. 8.8.3]
X0 (N) →
− E(C)
X0 (N) →
− E(C)
exists, then the map is induced by a new form f ∈ S2new (Γ0 (N)).
3. If the new form f ∈ S2new (Γ0 (N)) from the previous step has integer-valued
Fourier coefficients then
L( f , −) = L(E, −)
and N is the conductor of E.
X n +Y n = Z n , n ∈ N, n ≥ 3,
For the particular exponents n = 3, 4 the Fermat conjecture was proved by Euler.
In addition it was known: The general case follows from the case of prime expo-
nents n = l ≥ 5. Before 1995 the conjecture was open for general primes l ≥ 5.
Y 2 = X · (X − al ) · (X − cl ),
today named the Frey-Hellegouarch curve. The relation between non-trivial so-
lutions of the Fermat equation and points on elliptic curves had already been
considered by Hellegouarch, see [30]. The curve EFH has discriminant
∆ = 24 · (abc)2l
c34
j= with c4 = 24 · (a2l − (ac)l + c2l )
∆
The global minimal Weierstrass polynomial of EFH has conductor
N= ∏ p
p|abc
Serre and Ribet proved that the existence of EFH would contradict the validity of
the Shimura-Taniyama-Weil conjecture. Hence Wiles’ modularity theorem, The-
orem 7.2, implies that the Frey-Hellegouarche curve EFH does not exist. Hence
the Fermat conjecture is true.
A reference for all issues of this remark is [33, Chap. XII, § 1 and 4].
Example 7.7 (Modularity theorem). Figure 7.1 shows different elliptic curves E/Q
defined by a minimal Weierstrass polynomial, their conductor N, and a sample of
coefficients of their L-series and of the Fourier coefficients of the new-forms in
see Corollary 3.28. The numerical result from the PARI script Elliptic_curve_taniyama_10
confirms the Modularity Theorem that a rational elliptic curves arises from a mod-
ular form.
266 7 Outlook: Modular elliptic curves, monstrous moonshine
Fig. 7.1 Relation between rational elliptic curves and new forms
For an introduction see [8] and [59], on a more advanced level see [24] and [25].
H 6= {e} and H 6= G
If a finite group G has a proper normal subgroup H ( G then one will try to in-
vestigate G by studying the smaller groups H and G/H. Proposition 7.9 generalizes
this construction. It formalizes in which sense simple groups are the building blocks
of all finite groups.
7.2 Monstrous moonshine 267
Proposition 7.9 (Jordan-Hölder theorem). Each finite group G has a finite, strictly
increasing sequence of subgroups, a composition series,
G0 = {e} ⊂ G1 ⊂ ... ⊂ Gn = G
Qi := Gi /Gi−1
All composition series of G have the same lenght. Each two composition series of G
have - up to permutation - the same simple quotients counted with multiplicity.
For a proof see [38, Chap. IV, §4].
Example 7.10 (Composition series). The cyclic groups C12 has the following dif-
ferent composition series
{0} ⊂ C3 ⊂ C6 ⊂ C12
{0} ⊂ C2 ⊂ C6 ⊂ C12
{0} ⊂ C2 ⊂ C4 ⊂ C12
Each composition series of C12 has the same family of simple quotients (C2 ,C2 ,C3 )
up to permutation.
ρ :G→
− GL(V )
G ×V →
− V, (g, v) 7→ g · v := ρ(g)(v).
trace ρ : G →
− C, χρ (g) := trace ρ(g),
The monster has 194 conjugacy classes and therefore exactly 194 irreducible
representations.
• and on the left-hand side the Fourier coefficients c(n) of the q-expansion of the
modular j-invariant.
What later was named “monstrous moonshine” started with the observations from
Figure 7.2
1 = 1 and 196.884 = 1 + 196.883
by McKay in 1978. The formula states
270 7 Outlook: Modular elliptic curves, monstrous moonshine
relating two coefficients of the q-expansion of the modular j-invariant to the dimen-
sions of the two lowest irreducible M-modules. As Borcherds remarks “moonshine
is not a poetic term referring to light from the moon. It means foolish or crazy idea
(Quatsch in German).”
Γ0 (N) ⊂ G
φ : G × H∗ →
− H∗
genus X(G) = 0
H := G ∩ Γ
[Γ : H] ≤ [Γ : Γ0 (N)] < ∞.
7.2 Monstrous moonshine 271
For a moonshine group G of genus X(G) = 0 the modular curve X(G) is P1 , and JG
is a distinguished generator of the field of meromorphic functions
M (X(G)) = C(JG ).
1
JΓ (q) = j(q) − 744 = + 1960 884 · q + 210 4930 760 · q2 + 8640 2990 970 · q3 + O(4)
q
1
JΓ0 (2) = + 276 · q − 20 048 · q2 + 110 202 · q3 + O(4)
q
Γ0 (2)+ ⊂ SL(2, R)
1
JΓ0 (2)+ = + 40 372 · q + 960 256 · q2 + O(3)
q
V\ = Vn\
M
n∈N
satisfying: The graded components Vn\ split as the direct sum of irreducible M-modules
with a certain multiplicity. The M-module V \ is characterized by the family of
series, later named McKay-Thompson series
see [60].
(V0 , ρ1 ), (V1 = {0}), (V2 , ρ1 ⊕ ρ1960 883 ), (V3 , ρ1 ⊕ ρ1960 883 ⊕ ρ210 2960 876 )
The series generated by the dimensions (dim Vn )n∈N has the Fourier expansion
∞
∑ (dim Vn+1 ) · qn =
n=−1
1
= J(τ) = + 1960 884 · q + 210 4930 760 · q2 + 210 4930 760 · q3 + O(4), q = e2πi·τ .
q
The open problem was to find explicitly a representation with the properties from
the Conway-Norton conjecture, see Remark 7.17. Borcherds proved in 1992
Theorem 7.18, see [6].
Elliptic functions attain each value with equal multiplicity (Cor. 1.10)
273
274 List of results and some outlooks
Hausdorff topologyy of the orbit space of the Γ -action (Theor. 2.21, Cor. 2.25)
Modular forms as meromorphic diff. forms on the modular curve (Theor. 3.17)
The algebra of modular forms and the ideal of cusp forms (Theor. 3.24)
The group of classes of degree zero divisors of an elliptic curve (Theor. 4.31)
For CM-points τ ∈ H the singular values j(τ) ∈ C are algebraic integers (Theor.
6.14)
The ideal class group and tori with complex multiplication (Prop. 6.20)
The class number formula and the modular j-invariant (Theor. 6.22)
277
278 References
45. Ramanujan, Srinivasa: On certain arithmetical functions. Trans. Cambridge Philosophical So-
ciety. 22(9) (1916), p. 159-184
46. Raynaud, Michèle: Géometrie Algébrique et Géometrie Analytique. Séminaire de Géometrie
Algébrique du Bois Marie 1960-61, Exposé XII. https://arxiv.org/pdf/math/
0206203.pdf Call June 2020
47. Scharlau, Winfried; Opolka Hans: Von Fermat bis Minkowski. Eine Vorlesung über Zahlen-
theorie. Springer, Berlin (1980)
48. Schultz, Dan: Notes on Modular Forms.
https://faculty.math.illinois.edu/\%7eschult25/ModFormNotes.
pdf Call 27.10.2020
49. Serre, Jean-Pierre: Géométrie Algébrique et Géométrie Analytique. Annales de l’Institut
Fourier. IV (1955-56), p. 1-42
50. Serre, Jean-Pierre: II Modular Forms. (1957/58). In: [9].
51. Serre, Jean-Pierre: Représentations linéaires des groupes finis. Herman, Paris (1967)
52. Serre, Jean-Pierre: A Course in Arithmetic. Springer, New York (1973)
53. Shimura, Goto: Introduction to the Arithmetic Theory of Automorphic Functions. Iwanami
Shoten and Princeton University Press, o.O. (1971)
54. Silverman, Joseph: The Arithmetic of Elliptic Curves. Springer, New York (1986)
55. Silverman, Joseph: Advanced Topics in the Arithmetic of Elliptic Curves. Springer, New York
(1994)
56. Silverman, Joseph; Tate, John: Rational points on Elliptic Curves. Springer, New York (1992)
57. stackexchange. https://math.stackexchange.com/
58. Szpiro, Lucien: Séminaire sur les pinceaux arithmétiques: La conjecture de Mordell.
Astérisque 127. Société Mathématiques de France, Paris (1985)
59. Tatitscheff, Valdo: A short introduction to Monstrous Moonshine. Preprint (2019)
https://www.researchgate.net/publication/331008545_A_short_
introduction_to_Monstrous_Moonshine
60. Thompson, J. G.: Some Numerology between the Fischer-Griess monster and the Elliptic
Modular Function. Bull. Lond. Math. Soc. 11 (1979), 352-353
61. van der Geer, Gerard: Siegel Modular Forms. arXiv:math/0605346 [math.AG]
62. Wehler, Joachim: Complex Analysis. Lecture Notes (2019) http://www.mathematik.
uni-muenchen.de/\%7ewehler/20181218_Funktionentheorie_Script.
pdf
63. Wehler, Joachim: Riemann Surfaces. Lecture Notes (2020) http://www.mathematik.
uni-muenchen.de/\%7ewehler/20190530_RiemannSurfacesScript.pdf
64. Zagier, Don: Zetafunktionen und quadratische Körper. Eine Einführung in die höhere Zahlen-
theorie. Springer, Berlin (1981)
65. Zagier, Don: https://www.youtube.com/watch?v=zKt5L0ggZ3o (2015)
66. Zagier, Don: Elliptic Modular Forms and Their Applications.
https://people.mpim-bonn.mpg.de/zagier/files/doi/10.1007/
978-3-540-74119-0_1/fulltext.pdf Call 16.7.2020
Index
281
282 Index