Dokumen - Pub - Problems in The Theory of Modular Forms 978 981 10 2651 5 9811026513 978 93 80250 72 4
Dokumen - Pub - Problems in The Theory of Modular Forms 978 981 10 2651 5 9811026513 978 93 80250 72 4
Dokumen - Pub - Problems in The Theory of Modular Forms 978 981 10 2651 5 9811026513 978 93 80250 72 4
M. Ram Murty
Michael Dewar
Hester Graves
Problems in
the Theory of
Modular
Forms
HBA Lecture Notes in Mathematics
Series Editor
Sanoli Gun, Institute of Mathematical Sciences, Chennai, Tamil Nadu, India
Editorial Board
R. Balasubramanian, Institute of Mathematical Sciences, Chennai
Abhay G. Bhatt, Indian Statistical Institute, New Delhi
Yuri F. Bilu, Université Bordeaux I, France
Partha Sarathi Chakraborty, Institute of Mathematical Sciences, Chennai
Carlo Gasbarri, University of Strasbourg, Germany
Anirban Mukhopadhyay, Institute of Mathematical Sciences, Chennai
V. Kumar Murty, University of Toronto, Toronto
D.S. Nagaraj, Institute of Mathematical Sciences, Chennai
Olivier Ramaré, Centre National de la Recherche Scientifique, France
Purusottam Rath, Chennai Mathematical Institute, Chennai
Parameswaran Sankaran, Institute of Mathematical Sciences, Chennai
Kannan Soundararajan, Stanford University, Stanford
V.S. Sunder, Institute of Mathematical Sciences, Chennai
About the Series
The IMSc Lecture Notes in Mathematics series is a subseries of the HBA Lecture
Notes in Mathematics series. This subseries publishes high-quality lecture notes
of the Institute of Mathematical Sciences, Chennai, India. Undergraduate and
graduate students of mathematics, research scholars, and teachers would find this
book series useful. The volumes are carefully written as teaching aids and highlight
characteristic features of the theory. The books in this series are co-published with
Hindustan Book Agency, New Delhi, India.
Hester Graves
123
M. Ram Murty Hester Graves
Department of Mathematics and Statistics University of Michigan
Queen’s University Ann Arbor, MI, USA
Kingston, ON, Canada
Michael Dewar
Queen’s University
Kingston, ON, Canada
This work is a co-publication with Hindustan Book Agency, New Delhi, licensed for sale in
all countries in electronic form only. Sold and distributed in print across the world by
Hindustan Book Agency, P-19 Green Park Extension, New Delhi 110016, India. ISBN:
978-93-80250-72-4 © Hindustan Book Agency 2016.
© Springer Science+Business Media Singapore 2016 and Hindustan Book Agency 2016
This work is subject to copyright. All rights in this online edition are reserved by the Publishers, whether
the whole or part of the material is concerned, specifically the rights of reuse of illustrations, recitation,
broadcasting, and transmission or information storage and retrieval, electronic adaptation, computer
software, or by similar or dissimilar methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this
publication does not imply, even in the absence of a specific statement, that such names are exempt from
the relevant protective laws and regulations and therefore free for general use.
The publishers, the authors and the editors are safe to assume that the advice and information in this
book are believed to be true and accurate at the date of publication. Neither the publishers nor the
authors or the editors give a warranty, express or implied, with respect to the material contained herein or
for any errors or omissions that may have been made.
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xi
Preface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xiii
Part I Problems
1 Jacobi’s q-series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.1 The q-exponential function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2 Jacobi’s Triple Product Identity . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.3 Jacobi’s two-square theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.4 Jacobi’s four square theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.5 Supplementary problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
vii
viii Contents
Part II Solutions
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 289
Acknowledgements
xi
Preface
The earliest murmurs of a theory of modular forms can be traced back to the
work of Jacobi in 1829 when he wrote his famous treatise Fundamenta Nova
Theoriae Functionum Ellipticarum dealing with q-series and elliptic functions.
In some parenthetic sense, this was further developed by Riemann, Hur-
witz, Dedekind, Eisenstein, and Kronecker. However, it is in the work of
Ramanujan, in his celebrated paper [29] of 1916 in which he introduced the
τ -function, where we find the seeds of a comprehensive theory. There, Ra-
manujan studied the infinite product (in the variable q) given by
∞
Y 24
q (1 − q n ) (0.1)
n=1
xiii
xiv Preface
xvii
Part I
Problems
Chapter 1
Jacobi’s q-series
© Springer Science+Business Media Singapore 2016 and Hindustan Book Agency 2016 3
M.R. Murty et al., Problems in the Theory of Modular Forms, HBA Lecture Notes
in Mathematics, DOI 10.1007/978-981-10-2651-5_1
4 1 Jacobi’s q-series
so that one views q n −1 (or more precisely (q n −1)/(q −1)) as the q-analog of
the natural number n. Once this is understood, many of the functions in the
q-world become meaningful and exhibit remarkable structural properties.
1
(1 + x)E q1 (x) = .
Eq (x)
since for negative n, the product inside the sum is zero. Again by Exercise
1.1.3, replacing q by q 2 and replacing x by −q 2n+2 , we can write the product
inside the summation as
∞ 2
X (−1)m q m +m+2nm
.
m=0
(1 − q 2 ) · · · (1 − q 2m )
2 2m
q (m+n) xn+m .
m=0
(1 − q ) · · · (1 − q ) n=−∞
q
by Exercise 1.1.4 (where we have set q 2 for q and x for x). Putting everything
together gives the desired result. t
u
Exercise 1.2.2 (Euler’s pentagonal number theorem). Show that if |q| < 1
then
∞ ∞
Y X k(3k−1)
(1 − q n ) = (−1)k q 2 .
n=1 k=−∞
Proof. From the triple product identity, with x replaced by −x, we see that
∞ ∞
Y X 2
(1 − q 2n )(1 − xq 2n−1 )(1 − x−1 q 2n−1 ) = (−x)k q k .
n=1 k=−∞
The left hand side has a factor (1−xq) coming from n = 1, and consequently
vanishes when x = 1q . The same is true of the right hand side. Writing
∞
Y ∞
Y
(1 − xq 2n−1 ) = (1 − xq) (1 − xq 2n+1 ),
n=1 n=1
we obtain
∞ ∞
Y 1 X 2
(1 − q 2n )(1 − xq 2n+1 )(1 − x−1 q 2n−1 ) = (−x)k q k .
n=1
1 − xq
k=−∞
1
Putting x = q in the left hand side gives
∞
Y
(1 − q 2n )3 .
n=1
1
For the right hand side, we use l’Hopital’s rule to take the limit as x → q to
get
∞ ∞
1 X 2 X 2
− (−1)k kq 1−k+k = − (−1)k kq k −k .
q
k=−∞ k=−∞
We observe that the function f (k) = k 2 − k has the property that f (k) =
f (−(k − 1)). Thus, pairing up k and −(k − 1) in the sum, we get that it is
1.3 Jacobi’s two-square theorem 7
∞ ∞
X 2 X 2
−k −k
− {(−1)k k + (−1)k−1 (1 − k)}q k =− (−1)k (2k − 1)q k .
k=1 k=1
1
Changing q by q 2 gives the desired result. t
u
Exercise 1.2.4. Prove that
∞ ∞
X 2 X k(k+1)
(4n + 1)q 2n +n
= (−1)k (2k + 1)q 2 .
n=−∞ k=0
In this section, our goal is to obtain formulas for the number of ways a nat-
ural number m can be written as a sum of two squares. We recognize this as
the computation of the m-th coefficient in the power series expansion
∞
!2
n2
X
q .
n=−∞
We prove:
Theorem 1.3.1.
∞
!2 ∞
q 4n+1 q 4n+3
n2
X X
q =1+4 − .
n=−∞ n=0
1 − q 4n+1 1 − q 4n+3
∞
!2 ∞
q 4n+1 q 4n+3
n2
X X
q =1+4 − ,
n=−∞ n=0
1 − q 4n+1 1 − q 4n+3
Prove that x4 (n) is completely multiplicative. That is, show that x4 (mn) =
x4 (m)x4 (n).
Using results obtained for r2 (n), we will derive an explicit formula for r4 (n),
the number of ways of writing n as a sum of four squares. This formula is
due to Jacobi who derived it using the theory of elliptic functions. Here, we
will follow a method due to Ramanujan that is completely elementary and
based on the following exercise.
10 1 Jacobi’s q-series
qr
= ur (1 + ur ).
(1 − q r )2
Prove that
∞
X ∞
X
um (1 + um ) = nun .
m=1 n=1
qn
where un = 1−q n . Then L2 = T1 + T2 .
Corollary 1.4.5.
( ∞
)2 ∞
q 4n+1 q 4n+3 nq n
X X
1+4 − =1+8 .
n=0
1 − q 4n+1 1 − q 4n+3 n=1
1 − qn
n6≡0 (mod 4)
π
Proof. Put θ = 2 in Theorem 1.4.4. Then
1.4 Jacobi’s four square theorem 11
∞
1 X qn π
L= + sin n .
4 n=1 1 − q n 2
∞ ∞
!4 ∞
X
n
X
n2
X nq n
r4 (n)q = q =1+8 .
n=0 n=−∞ n=1
1 − qn
n6≡0 (mod 4)
where ∞
X
1 θ
S1 = cot un sin nθ
2 2 n=1
and
∞ X
X ∞
S2 = um un sin mθ sin nθ.
m=1 n=1
By Exercise 1.4.1,
1 θ 1 1
cot sin nθ = + cos θ + cos 2θ + · · · + cos(n − 1)θ + cos nθ
2 2 2 2
and
2 (sin mθ) (sin nθ) = cos(m − n)θ − cos(m + n)θ,
so that
∞
X 1 1
S1 = un + cos θ + cos 2θ + · · · + cos(n − 1)θ + cos nθ
n=1
2 2
and
∞ ∞
1 XX
S2 = um un {cos(m − n)θ − cos(m + n)θ} .
2 m=1 n=1
In other words,
2 ∞
2 1 θ X
L = cot + c0 + ck cos kθ
4 2
k=1
1.4 Jacobi’s four square theorem 13
uj uk−j = uk (1 + uj + uk−j )
and
uk+j + uj uk+j = uk (uj − uk+j )
so that
1 ∞ k−1
X 1X
c k = uk + (uj − uk+j ) − (1 + uj + uk−j ) .
2 2 j=1
j=1
The methods utilised above to determine formulas for r2 (n) and r4 (n)
can be extended for other values r2k (n) for certain values of k. We refer the
14 1 Jacobi’s q-series
reader to consult [11] for more examples. In [11], the student will also find a
discussion for finding formulas for rk (n) when k is odd and this is intimately
connected with the theory of modular forms of half-integral weight which
is beyond the scope of this book.
Show that τ (n) is odd if and only if n = (2m + 1)2 for some m.
Exercise 1.5.2. Let rk (n) be the number of ways of writing n as a sum of k squares.
Show that
Xn
rk (n) = ri (a)rk−i (n − a)
a=0
Deduce that
∞
X x
Lq (x) = n−x
.
n=1
q
Exercise 1.5.4. Show that if |x| < |q| and |q| > 1 then
xEq0 (−x)
Lq (x) = ,
Eq (−x)
Exercise 2.1.1. Let R be a commutative ring with identity. Show that the set
ab a, b, c, d ∈ R
SL2 (R) = :
cd ad − bc = 1
The (full) modular group SL2 (Z) plays a pivotal role in the theory of mod-
ular forms. One also considers PSL2 (Z) = SL2 (Z) /{±I}. The relationship
of SL2 (Z) to SL2 (R) is similar to the relationship of Z to R in the sense that
Z is a discrete subgroup of R and SL2 (Z) is a discrete subgroup of SL2 (R).
We will show below that SL2 (Z) is generated by the elements
0 −1 11
S= and T = .
1 0 01
© Springer Science+Business Media Singapore 2016 and Hindustan Book Agency 2016 15
M.R. Murty et al., Problems in the Theory of Modular Forms, HBA Lecture Notes
in Mathematics, DOI 10.1007/978-981-10-2651-5_2
16 2 The Modular Group
ab
Now let g = be any element of SL2 (Z). If c = 0, then ad = 1
cd
implies that a = d = ±1. In this case
0
1b 0
g=± = ±T b .
0 1
0 0
where b0 = ±b. Since S 2 = −I, either g = T b or S 2 T b . If c 6= 0, we proceed
as follows. Without loss of generality we may suppose |a| ≥ |c|, for oth-
erwise we can apply S to arrange this. By the Division Algorithm, we can
write a = cq + r with 0 ≤ r < |c|. Then T −q g has upper left entry r = a − qc
which is smaller than |c|. Applying S switches the rows (with a sign change)
and so we can iterate the process if r 6= 0. After a finite number of steps, we
are reduced to the case c = 0 and we are done. t
u
Exercise 2.1.3. Show that S has order 4, ST has order 6, and T has infinite order.
Exercise 2.1.4. Show that SL2 (Z) is generated by two elements of finite order,
namely S and ST of order 4 and 6 respectively.
φ : SL2 (Z) → C×
has image contained in the finite subgroup of C× consisting of 12th roots of unity.
Exercise 2.1.7. Suppose that (c, d, N ) = 1. Show that there are elements c0 =
c + tN and d0 = d + sN for some integers s, t such that (c0 , d0 ) = 1.
is a surjective homomorphism.
Exercise 2.2.2. Prove that Γ(N ) is a normal subgroup of SL2 (Z) of finite index.
Exercise 2.2.3. For any commutative ring R with identity, let GL2 (R) be the set
of 2 × 2 invertible matrices with entries in R. Show that GL2 (R) is a group.
Exercise 2.2.4. Let p be a prime. Show that the order of GL2 (Z/pZ) is given by
(p2 − 1)(p2 − p).
Exercise 2.2.5. Let p be a prime. Show that SL2 (Z/pZ) is a normal subgroup of
GL2 (Z/pZ) of index p − 1. Deduce that SL2 (Z/pZ) has order p(p2 − 1).
Exercise 2.2.6. Show that SL2 (Z/p2 Z) has size p4 (p2 − 1).
Exercise 2.2.7. Apply an induction argument to show that
SL2 (Z/pn Z)
Deduce that
Y 1
[SL2 (Z) : Γ(N )] = N 3 1− .
p2
p|N
is again surjective and the index of ker φN in SLn (Z) can be given explicitly.
Again a subgroup Γ ⊂ SLn (Z) is called a congruence subgroup if ker φN ⊂ Γ
18 2 The Modular Group
both of order N .
Exercise 2.2.11. In SL2 (Z), show that the group Γ generated by S and T 2 is a
congruence subgroup of index three.
It is easy to see that this is a group. Clearly Γ(N ) ⊂ Γ0 (N ) and so these are
congruence subgroups.
is a surjective homomorphism.
Thus we have
2.4 Groups acting on topological spaces 19
Let G be a group and X a topological space. We say that there is a left action
of G on X if, for each element g ∈ G, we have a continuous map x 7→ gx ∈ X
for x ∈ X satisfying the conditions:
(a) 1x = x ∀x ∈ X
(b) (gh)x = g(hx) ∀g, h ∈ G, x ∈ X.
From these axioms, we see that x 7→ g −1 x is the inverse map of x 7→ gx.
Thus each element g ∈ G gives rise to a topological automorphism of X.
It may be clearer to define an action as a pairing G × X → X, where we
use the notation g · x to denote the image of (g, x) under this mapping. We
suppress the ‘dot’ for aesthetic reasons.
For each element x ∈ X, we define the stabilizer subgroup Gx as
Gx := {g ∈ G : gx = x}.
Ggx = gGx g −1 .
Gx := {gx : g ∈ G} .
Exercise 2.4.2. If G acts transitively on X, show that all stabilizers are conjugate.
x 7→ π(x) = Gx.
In other words, we associate to each x its G-orbit. We put the strongest topol-
ogy on G \ X such that π is continuous. More precisely, a subset U of G \ X
is open if and only if π −1 (U ) is open in X. The topological space G \ X with
this topology is called the quotient space of X by G. If U is any open set of X,
we have
[
π −1 (π(U )) = gU
g∈G
is again a union of open sets. Thus π(U ) is open so that π is an open contin-
uous map of X onto G \ X.
Now suppose that G has the topological structure of a Hausdorff space
and the map G × G → G given by (g, h) 7→ gh−1 is a continuous map, we
say G is a topological group.
We say that a topological group G acts on a topological space X if, in
addition to conditions (a) and (b) defining group actions, we impose that the
map (g, x) 7→ gx is a continuous map with respect to the product topology
on G × X.
We will not prove but simply state the following important theorem. (The
reader can find a proof in Chapter 1 of [37].)
∀h ∈ H, ∀k ∈ K, hX = X = Xk.
`
Now as before, X is a disjoint union of its H orbits: X = i Hxi . Moreover,
K acts on the set of orbits via
Exercise 2.5.5. If SL2 (Z) has a non-congruence subgroup, show that it has in-
finitely many. (Hint: consider Γ ∩ Γ (p) for a non-congruence subgroup Γ
and rational primes p.)
Exercise 2.5.6. For each g ∈ SL2 (Z) and a subgroup Γ of finite index in SL2 (Z),
we define ng (Γ ) to be the least positive integer such that
11
T ng (Γ ) ∈ gΓ g −1 , T = as usual.
01
We say Γ is of class N if the least common multiple of the set of numbers {ng (Γ ) :
g ∈ SL2 (Z)} equals N . Show that Γ (N ) is of class N . Also show that the normal
closure of T N , namely
is of class N . (This exercise extends the notion of congruence level defined for
congruence subgroups to all subgroups of finite index, and is due to Fricke
and Wohlfahrt.)
Exercise 2.5.7. (a) Using the Chinese remainder theorem (or otherwise), show that
for given x, y, the following congruence
2.5 Supplementary problems 23
x(αy + β) ≡ γ mod m
Exercise 2.5.9. Let Γ be a subgroup in SL2 (Z) of class N . Show that Γ is a con-
gruence subgroup if and only if it is a congruence subgroup of level N . Deduce that
if Γ is a subgroup of SL2 (Z) of class N and the index [SL2 (Z) : Γ ] does not divide
[SL2 (Z) : Γ (N )], then Γ is not a congruence subgroup.
Exercise 2.5.10. Let F be a free group on n generators, say x1 , ..., xn . For a word
w ∈ F , define `i (w) to be the sum of all the exponents of xi occuring in w. For a
subset I of {1, ..., n}, and a positive integer `, define
Exercise 2.5.11. Recall that Γ (2) is generated by −I, T 2 , U 2 (see Exercise 2.2.10)
and that the subgroup generated by x1 = T 2 and x2 = U 2 is free of index 2 in
Γ (2).
(a) With notation as in the previous exercise, compute the class of the group
Γ` ({1}) and Γ` ({1, 2}).
(b) Show that Γ` ({1, 2}) is not a congruence subgroup if ` is not a power of 2.
Chapter 3
The Upper Half-Plane
The transformations
az + b
z 7→
cz + d
are called fractional linear transformations or Möbius transformations. Note that
the scalar matrices act trivially on H.
Exercise 3.1.2. Show that the group PSL2 (R) = SL2 (R)/ ± I acts faithfully on
H. That is, gz = z for all z ∈ H implies g = 1 ∈ PSL2 (R).
Exercise 3.1.3. Show that the action of PSL2 (R) on H is transitive, i.e. there is
only one orbit and for all z, w ∈ H, there exists some g ∈ PSL2 (R) such that
gz = w.
Exercise 3.1.4. Show that the map
z−i
z 7→
z+i
© Springer Science+Business Media Singapore 2016 and Hindustan Book Agency 2016 25
M.R. Murty et al., Problems in the Theory of Modular Forms, HBA Lecture Notes
in Mathematics, DOI 10.1007/978-981-10-2651-5_3
26 3 The Upper Half-Plane
φ(z) − a
ψ(z) =
b
maps i to i. But now, let h : H → D be given by
z−i 1 −i
h(z) = = z.
z+i 1 i
and so
ρ
ρ2 i
−1 − 12 0 1 1
2
3
1 ≥ |cz + d|2 = (x + d)2 + y 2 > (x + d)2 +
4
so that |x+d| < 1/2. Since |x| < 1/2, this means d = 0. But now 1 ≥ |cz+d| =
|z| which implies that z is not an interior point of F . t
u
Exercise 3.2.4. Suppose that z, w ∈ F are distinct points which are SL2 (Z)-
equivalent. Show that Im(z) = Im(w) and Re(z) = − Re(w).
Exercise 3.2.5. If F is the standard fundamental domain for SL2 (Z) and γ ∈
SL2 (Z), show that γF is again a fundamental domain for SL2 (Z).
One can prove that any discrete subgroup Γ of SL2 (R) has a fundamental
domain (see Theorem 1.6.2 of [22]). The next theorem allows us to construct
fundamental domains in certain cases.
Theorem 3.2.6. Let F be a fundamental domain for SL2 (Z) and let Γ be a congru-
ence subgroup
Sr with −I ∈ Γ. If g1 , . . . , gr are coset representatives of Γ such that
D = i=1 gi−1 F has a connected interior, then D is a fundamental domain for Γ.
Proof. It is clear that the boundary of each gi−1 F consists of a finite number
of smooth curves so that condition (c) in the definition of a fundamental
domain is satisfied. We need to verify (a) and (b). Now any z ∈ H is SL2 (Z)-
equivalent to some point w ∈ F . Thus w = τ z with τ ∈ SL2 (Z). We can
write τ = gi γ for some i and γ ∈ Γ. Hence γz ∈ gi−1 F so that γz ∈ D and (a)
is satisfied.
Now suppose two interior points of D are Γ-equivalent. In other words,
there is some γ ∈ Γ, and some interior point z ∈ D such that γz ∈ D is also
an interior point. The action of γ is a homeomorphism and so there exists an
open set U ⊂ D containing z such that γz ∈ γU ⊂ D. By the construction of
D, there exist indices 1 ≤ i, j ≤ r such that gi z, gj γz ∈ F . Let V = gi U ∩ F ◦ ,
30 3 The Upper Half-Plane
where F ◦ denotes the interior of F . The set V is open and non-empty. Take
any x ∈ V. Then
γgi−1 x ∈ γgi−1 V ⊆ γU ⊂ D.
Let k be such that γgi−1 x ∈ gk−1 F . Now x and gk γgi−1 x are SL2 (Z)-
equivalent interior points of F . Hence they are equal. The map w 7→ gk γgi−1 w
is an automorphism of H. Thus, there is an open set W ⊂ F ◦ containing x
such that gk γgi−1 W ⊂ F ◦ . But then every element of W is SL2 (Z)-equivalent
to some element of F ◦ . This can only happen if gk γgi−1 acts trivially on W.
As any two holomorphic functions agreeing on an open set must agree ev-
erywhere, we deduce that gk γgi−1 = ±I by Exercise 3.1.2. Now
gi Γ = gk γ{±I}Γ = gk Γ.
Exercise 3.2.7. Show that I, S, T −1 S are coset representatives for Γ0 (2) in SL2 (Z).
Deduce that a fundamental domain for Γ0 (2) is given by F ∪ SF ∪ ST F , where F
is the standard fundamental domain for SL2 (Z).
H? = H ∪ Q ∪ {i∞}.
That is, H? is obtained by adjoining all the rational numbers and i∞, called
the cusps. We topologize H? in the following way. For z ∈ H, we take the
usual fundamental system of neighborhoods. For a cusp s 6= i∞, we take all
sets of the form
{s} ∪ C ◦ ,
where C ◦ is the interior of a circle in H tangent to the real axis at s. If s = i∞,
we take all sets of the form
π : H? −→ Γ\H? ,
along with ( 10 01 ), give a complete set of right coset representatives for Γ0 (p) in
SL2 (Z).
Exercise 3.3.6. Let p be prime. Show that the set of matrices
k 1
, 0 ≤ k ≤ p − 1,
−1 0
|z − w|2
−1
d(z, w) = cosh 1+ ,
2 Im(z) Im(w)
Exercise 3.4.1. Prove that d(z, w) is SL2 (R)-invariant. That is, show that d(gz, gw) =
d(z, w) for all g ∈ SL2 (R).
One can also define an SL2 (R)-invariant measure on H which lifts to Γ\H?
for any subgroup Γ of SL2 (R). This is given by
dxdy
dµ := .
y2
dudv = (detJ)dxdy,
A tedious calculation shows that det J = (1/y)2 , from which the result fol-
lows. t
u
Exercise 3.4.5. Let F be the standard fundamental domain for SL2 (Z). Show that
Z Z
dxdy π
2
= .
F y 3
Exercise 3.5.1. Show that a set of coset representatives for Γ∞ in SL2 (Z) is given
by
??
: (c, d) = 1, c, d ∈ Z ,
cd
where the asterisks indicate that we can take any a, b such that ad − bc = 1.
Exercise 3.5.3. Let G = SL2 (R) act on H. Show that K, the stabilizer of i, consists
of matrices
cos θ sin θ
for θ ∈ R.
− sin θ cos θ
Then show that the upper half-plane is homeomorphic to G/K.
Chapter 4
Modular Forms of Level One
Uc = {z ∈ H : Im(z) > c}
of i∞ is mapped via z 7→ e2πiz into the punctured open disc centered at zero, of
radius e−2πc .
The function q(z) = e2πiz from Exercise 4.1.1 can be extended (continu-
ously) by sending i∞ to 0. Hence, the fundamental neighborhoods of i∞ are
mapped to the fundamental neigborhoods of the origin.
Since T = ( 10 11 ) ∈ SL2 (Z), condition (4.1) implies that f (z + 1) = f (z) for
all z ∈ H. Hence, there is a well-defined induced map, which we also call f ,
from the unit disc to C:
q 7→ f (z),
where z ∈ H is any point with q = e2πiz . If f (z) : H → C is holomorphic,
then f (q) will be holomorphic on the punctured unit disc and hence will
have a Laurent expansion centered at q = 0,
∞
X
f (q) = an q n ,
n=−∞
© Springer Science+Business Media Singapore 2016 and Hindustan Book Agency 2016 35
M.R. Murty et al., Problems in the Theory of Modular Forms, HBA Lecture Notes
in Mathematics, DOI 10.1007/978-981-10-2651-5_4
36 4 Modular Forms of Level One
∞
X
f (z) = an e2πizn
n=−∞
the Fourier series at i∞. We define the order of f at i∞ to be the index of the
first non-zero coefficient in the q-series:
vi∞ (f ) := inf {n ∈ Z : an 6= 0} .
f |γ = f, ∀γ ∈ SL2 (Z).
Exercise 4.1.3. Show that f |(γ1 γ2 ) = (f |γ1 )|γ2 for all γ1 , γ2 ∈ GL+
2 (R).
Let Mk (SL2 (Z)) denote the set of all modular forms of weight k on
SL2 (Z). Let Sk (SL2 (Z)) ⊆ Mk (SL2 (Z)) denote the set of cusp forms of
weight k on SL2 (Z). The use of the letter S for the space of cusp forms comes
from the German word “Spitzenform” which means cusp form.
(a) Both Mk1 (SL2 (Z)) and Sk1 (SL2 (Z)) are C-vector spaces.
(b) If f ∈ Mk1 (SL2 (Z)) and g ∈ Mk2 (SL2 (Z)), then f g ∈ Mk1 +k2 (SL2 (Z)).
(c) If f ∈ Mk1 (SL2 (Z)) and g ∈ Sk2 (SL2 (Z)), then f g ∈ Sk1 +k2 (SL2 (Z)).
Later, we will see that Mk (SL2 (Z)) and Sk (SL2 (Z)) are finite-dimensional
vector spaces. Consequently, methods of linear algebra can be applied to
study them and to construct “nice” bases for Mk (SL2 (Z)) and Sk (SL2 (Z)).
The previous exercise also allows us to speak about a graded C-algebras of
holomorphic modular forms and holomorphic cusp forms.
Exercise 4.1.6. Show that Mk (SL2 (Z)) = {0} for k odd.
which is positive definite. That is, A > 0, and AC − B 2 > 0. We may write
Q(m, n) in matrix form:
AB m
Q(m, n) = (m, n) .
BC n
A B ) is a real symmetric matrix, we can diagonalize it.
Since the matrix ( B C
Thus, we can find an orthogonal matrix P such that
AB λ1 0
P Pt = ,
BC 0 λ2
and if we put
u m
=P ,
v n
then we can write Q(m, n) = λ1 u2 + λ2 v 2 . Since Q is positive definite, the
eigenvalues λ1 , λ2 are real and positive. We see that
Q(m, n) ≥ λ(u2 + v 2 ),
where
λ = min(λ1 , λ2 ) > 0.
Moreover, P is an orthogonal transformation and thus preserves length.
Consequently, u2 + v 2 = m2 + n2 . This proves that
Q(m, n) ≥ λ(m2 + n2 )
38 4 Modular Forms of Level One
for some λ > 0. If z is a complex number, then |mz + n|2 is a positive definite
quadratic form in m, n.
These elementary observations from linear algebra allow us to construct
an important family of modular forms. For k > 2, define
X0 1
Gk (z) = , z ∈ H,
m,n
(mz + n)k
where the summation is over all (m, n) ∈ Z2 , with (m, n) 6= (0, 0). The dash
on the summation indicates this latter condition.
Exercise 4.2.1. Show that the series for Gk (z) converges absolutely for all z ∈ H
and for k > 2.
Gk (i∞) = 2ζ(k),
where ζ(s) denotes the Riemann zeta function. In addition, Gk (z) is a mod-
ular form of weight k since
Gk (z + 1) = Gk (z),
1
Gk − = z k Gk (z),
z
and as (m, n) ranges over elements of Z2 \ {(0, 0)}, so does (m, m + n). Sim-
ilarly,
1 X0 1
Gk − = zk k
= z k Gk (z)
z m,n
(nz − m)
since (n, −m) ranges over elements of Z2 \ {(0, 0)} as (m, n) ranges over
Z2 \{(0, 0)}. By Exercise 4.1.4, we see that Gk (z) is a modular form of weight
k for SL2 (Z).
4.2 Eisenstein series for SL2 (Z) 39
To study the behavior of Gk (z) near i∞, we now derive its q-expansion.
Recall
∞
z2
Y
sin πz = πz 1− 2 .
n=1
n
We can expand the inner sum using the Lipschitz formula (Theorem 4.2.2):
∞ ∞
(−2πi)k X X k−1 mn
Gk (z) = 2ζ(k) + 2 n q .
(k − 1)! m=1 n=1
X Bk xk ∞
x
x
= ,
e −1 k!
k=0
so that B0 = 1, B1 = − 12 , B2 = 16 , B3 = 0, B4 = − 30
1
, . . . . Show that
(2πi)2k B2k
2ζ(2k) = − , k ≥ 1.
(2k)!
Exercise 4.2.5. Prove that the Bernoulli numbers satisfy the recursion formula:
m−1
X
m+1
(m + 1)Bm = − Bk .
k
k=0
Prove that
∞
text X xn
t
= Bn (x) .
e − 1 n=0 n!
Using Exercise 4.2.4, we obtain:
Corollary 4.2.7. For every even k ≥ 4,
∞
2k X
Ek (z) = 1 − σk−1 (n)q n
Bk n=1
is a modular form of weight k for the full modular group SL2 (Z) satisfying
Ek (i∞) = 1.
Proof. Since Gk (z) = 2ζ(k)Ek (z), we can use Exercise 4.2.4 and Theorem
4.2.3 and the result is immediate. t
u
Here are some examples:
∞
X ∞
X
E4 (z) = 1 + 240 σ3 (n)q n , E10 (z) = 1 − 264 σ9 (n)q n ,
n=1 n=1
∞ ∞
X 65520 X
E6 (z) = 1 − 504 σ5 (n)q n , E12 (z) = 1 + σ11 (n)q n ,
n=1
691 n=1
X∞ ∞
X
E8 (z) = 1 + 480 σ7 (n)q n , E14 (z) = 1 − 24 σ13 (n)q n .
n=1 n=1
4.3 The valence formula 41
Exercise 4.2.8. Prove that Mk (SL2 (Z)) = CEk ⊕ Sk (SL2 (Z)) for all k ≥ 4.
Deduce that dim Mk (SL2 (Z)) = 1 + dim Sk (SL2 (Z)).
It is easily verified that
E43 − E62
∆ :=
1728
is a non-trivial cusp form of weight 12 for the full modular group. The
Fourier coefficients of ∆ define the celebrated Ramanujan τ -function:
∞
X
∆(z) = τ (n)q n = q − 24q 2 + 252q 3 + · · · .
n=1
shows that vz0 (f ) = vgz0 (f ) for all g ∈ SL2 (Z). In other words, vz0 (f ) de-
pends only on the orbit of z0 under SL2 (Z) so that we need only study vz (f )
for z in a fundamental domain of SL2 (Z). Recall that we have defined vi∞ (f )
to be the order of the zero at q = 0 in the q-expansion
∞
X
f (z) = an q n .
n=0
in the compact region {z ∈ F : Im(z) ≤ c} and this number is finite (see, for
example, Theorem 10.18 on page 208 of [33]).
Theorem 4.3.1 (The valence formula). Let f be a modular form, not identically
zero, of weight k for the full modular group SL2 (Z). Then
1 1 X0 k
vi∞ (f ) + vi (f ) + vρ2 (f ) + vz (f ) = ,
2 3 2
12
z6=i,ρ
z∈F
1
where the primed summation excludes points with Re z = 2 and points with both
√ πi
|z| = 1 and Re z > 0. Here i = −1 and ρ = e 3 .
We use primed summation in the valence formula to account for the fact
that half of the boundary of F is SL2 (Z)-equivalent to the other half.
Proof. We first give the proof in the case that f has no zeros on the bound-
ary of the standard fundamental domain F in H. Then, we indicate how to
modify the proof so as to treat the case when some zeros lie on the boundary.
iT
ρ
ρ2 i
−1 − 12 0 1 1
2
The contour has several segments. Since f (z + 1) = f (z), the two vertical
segments cancel each other. The horizontal segment maps (under z 7→ q =
e2πiz ) into the circle of radius e−2πT about q = 0 oriented clockwise and thus
4.3 The valence formula 43
contributes −vi∞ (f ) to the integral. The map z 7→ −1/z maps the arc from
ρ2 to i to the arc from ρ to i. Since
1
f − = z k f (z),
z
Hence
X k
vz (f ) = −vi∞ (f ) + .
12
z∈F
This gives the result in the case that f has no zeros on the boundary of F . We
now indicate how to modify this proof to allow for zeros on the boundary.
We modify C to create C 0 as follows. If there is a zero z1 with Re z1 = −1/2
and z1 6= ρ2 then we deform the contour as indicated in Figure 4.2. The arc
around T z1 is the transform by T of the arc around z1 . If there is a zero z2
with |z2 | = 1 and − 12 < Re z2 < 0, then we deform the contour as indicated
in Figure 4.3. The arc around Sz2 is the image under S of the arc around z2 .
We make as many such deformations as are necessary. Finally, we deform
the contour near the points ρ2 , i, and ρ so that it moves along circular arcs
of small enough radius > 0. See Figure 4.4. The argument principle now
gives
f 0 (z)
Z
1 X0
dz = vz (f ).
2πi C 0 f (z) 2
z6=i,ρ
z∈F
The modified vertical contours still cancel each other since their orientations
are opposite. Moreover, the arcs along |z| = 1 including the deformations
around points z2 and Sz2 , but excluding the arcs around ρ2 , i, and ρ, will
contribute k/12 once we take the limit → 0. To determine the contributions
44 4 Modular Forms of Level One
z1 T z1
−1 − 12 0 1 1
2
i Sz2
z2
ρ2 ρ
−1 − 12 0 1 1
2
We consider i first. If the arc were the full circle (oriented clockwise)
around i, the contribution would be −vi (f ). Since it is only a half-circle,
the arc is easily parametrized with the angle ranging from −π to 0 so that
the final contribution is − 12 vi (f ). The same is true of the arcs centered at ρ
4.3 The valence formula 45
i
ρ2 ρ
−1 − 12 0 1 1
2
and ρ2 . The arc around ρ is parametrized with the angle ranging from π/2 to
2π/3 (see Figure 4.5). The integral around the arc centered at ρ contributes
−vρ (f )/6. The same is true of the contribution from the arc centered at ρ2 .
But ρ and ρ2 are SL2 (Z)-equivalent so the final contribution is −vρ (f )/3. t
u
π
3
π
3
−1 0 1
Exercise 4.3.8. Show that Mk (SL2 (Z)) = CEk for k = 4, 6, 8, and 10.
(iii) Multiplication by ∆ defines an isomorphism of Mk−12 (SL2 (Z)) onto Sk (SL2 (Z))
for all k ∈ Z.
Proof. We have already noted that (i) and (ii) are immediate by the va-
lence formula (see Exercises 4.3.2, 4.3.3, 4.3.4, and 4.3.8). Clearly, if g ∈
Mk−12 (SL2 (Z)), then g∆ ∈ Sk (SL2 (Z)). If f ∈ Sk (SL2 (Z)), then g = f /∆
is a modular form of weight k − 12 since ∆ does not vanish on the upper
half-plane, implying g is analytic. This proves (iii). t
u
4.4 The dimension formula 47
Proof. The formula is true for k ≤ 10 by (i) and (ii) of Theorem 4.4.1. By
Exercise 4.2.8 and by part (iii) of Theorem 4.4.1, for k ≥ 12 we have
Corollary 4.4.4. The modular forms E4α E6β with 4α + 6β = k, and α, β integers
≥ 0, form a basis for Mk (SL2 (Z)).
Proof. We first show that any modular form f of weight k for SL2 (Z) can be
written as a linear combination of such elements. We proceed by induction
on k. This is clear for k ≤ 10 by Exercises 4.3.8-4.3.10. For general k, choose
positive integers α, β such that 4α + 6β = k. This is clearly possible since k is
even and we can find positive integers α, β such that 2α+3β = k/2. (Indeed,
if k/2 is even, we write 2α = k/2. If k/2 is odd, we write 2(α − 1) + 3 = k/2.)
Then g = E4α E6β has weight k and is non-zero at i∞. So there exists λ ∈ C
such that f −λg vanishes at i∞. Thus we can write f = λg +∆h for some h ∈
Mk−12 (SL2 (Z)) by (iii) of Theorem 4.4.1. Applying the induction hypothesis
to h shows that the monomials E4α E6β span Mk (SL2 (Z)). To show that these
elements are linearly independent, we suppose that
X
cαβ E4α E6β = 0.
4α+6β=k
Since E4 (ρ) = 0 and E6 (ρ) 6= 0, we see that cαβ with α = 0 must vanish.
The same argument shows that cαβ with β = 0 also vanishes, using the fact
E6 (i) = 0 and E4 (i) 6= 0. Thus, we have α ≥ 1, β ≥ 1 in the summation.
Hence X
cαβ E4α−1 E6β−1 E4 E6 = 0.
4(α−1)+6(β−1)=k−10
E43 − E62
Exercise 4.4.5. Prove that ∆ = has integer Fourier coefficients.
1728
∆ = q − 24q 2 + . . . ,
1
j(z) = + 744 + 196884q + 21493760q 2 + . . . (4.5)
q
If we write c(n) for the n-th Fourier coefficient of the j-function, then Peters-
son [28] showed in 1932 that
√
e4π n
c(n) ∼ √ ,
2n3/4
as n → ∞. Thus the integer coefficients c(n) grow at an exponential rate.
These coefficients are related to the representation theory of a sporadic sim-
ple group known as the Monster group that played a decisive role in the
classification of all finite simple groups. The j-function has many more
remarkable properties that we cannot adumbrate here. It is ubiquitous in
mathematics and one is compelled to study its beauty and character irre-
spective of any subdiscipline. We will content ourselves here with the fol-
lowing.
Theorem 4.5.1. Let F denote the fundamental domain from Theorem 3.2.2 and let
F 0 denote the set obtained by removing points with Re z = 12 and points with both
|z| = 1 and Re z > 0. The j-function z 7→ j(z) defines a bijection from F 0 onto C.
b c
a+ + =1
2 3
with a, b, c, integers ≥ 0 arises from
The reader may easily check that the valence formula also holds for mod-
ular functions, the proof being essentially identical. This permits an alter-
native proof of (i) ⇒ (iii) that treats zeros and poles on the same foot-
ing. Let f be a modular function, not identically zero. Assume first that
vi (f ) = vρ (f ) = 0. Let z1 , z2 , . . . , zr and w1 , w2 , . . . , ws denote, respectively,
the zeros and poles in H of f (z), listed according to multiplicity. Set
Qr
(j(z) − j(zi ))
g(z) = Qsi=1 .
i=1 (j(z) − j(wi ))
Exercise 4.6.1. Show that the Bernoulli numbers Bk vanish for odd k ≥ 3.
Exercise 4.6.2. Show that for k even and positive, (−1)k/2 Bk < 0.
Exercise 4.6.3. Prove that
Bk+1 (n + 1) − Bk+1
1k + 2k + · · · + nk = ,
k+1
for every natural number k.
Exercise 4.6.4. Prove that ζ(k) tends to 1 as k tends to infinity and use this to
show that √
|B2k | ∼ 4π e(k/πe)2k+1/2
as k tends to infinity.
Exercise 4.6.5. Show that if the Fourier coefficients of f ∈ Mk (SL2 (Z)) are inte-
gers, then f can be written as a polynomial in E4 , E6 , and ∆ with integer coeffi-
cients.
P∞
Exercise 4.6.6. Suppose f = n=0 an q n ∈ Mk (SL2 (Z)). Show that for any x ∈
R we have limy→∞ f (x+iy) = a0 . Show that for any x ∈ Q, if f is not a cusp form,
then limy→0 f (x + iy) = ∞, and if f is a cusp form, then limy→0 f (x + iy) = 0.
Deduce that the real line is a natural boundary for any modular form.
Exercise 4.6.7. Prove that E42 E6 = E14 .
4.6 Supplementary problems 51
Exercise 4.6.9. Let f and g be modular forms of weight k and ` respectively, for the
full modular group. Show that
kf g 0 − `f 0 g
dj
= −2πiE14 /∆.
dz
Exercise 4.6.11. Let rk = dim Mk (SL2 (Z)) and ` = k − 12rk + 12. For any
modular form f ∈ Mk (SL2 (Z)), show that
f E`−1 ∆1−rk
Exercise 4.6.13. Let f be a modular form of weight k for the full modular group
whose Fourier expansion at i∞ is given by
∞
X
f (z) = an e2πinz , an ∈ C.
n=0
Recall that by Corollary 4.4.4, there are unique cab ’s such that
X
f= cab E4a E6b , cab ∈ C.
4a+6b=k
Then, show that cab belong to F , whence all the Fourier coefficients belong to F .
Exercise 4.6.15. Suppose that f is a modular form of weight k whose Fourier ex-
pansion at i∞ is given by
∞
X
f (z) = an e2πinz , an ∈ Q.
n=0
16 ∗
r24 (n) = σ (n) + e24 (n),
691 11
where
128
e24 (n) = (−1)n−1 259τ (n) − 512τ (n/2) ,
691
with the last term appearing only if n is even. We will not prove this formula
in this book but refer the reader to Chapters 9 and 10 of [13] and [11] for
further details.
© Springer Science+Business Media Singapore 2016 and Hindustan Book Agency 2016 53
M.R. Murty et al., Problems in the Theory of Modular Forms, HBA Lecture Notes
in Mathematics, DOI 10.1007/978-981-10-2651-5_5
54 5 The Ramanujan τ -function
Our first goal is to prove that the left hand side of (5.1) is a cusp form of
weight 12 for the full modular group for q = e2πiz and z ∈ H. We follow a
method due to Hurwitz and study the Eisenstein series E2 (defined below)
to realize this goal.
The double series
∞ ∞
X X 0 1
G2 (z) := , z∈H (5.2)
m=−∞ n=−∞
(mz + n)2
1 1
am,n (z) = −
mz + n − 1 mz + n
for (m, n) 6= (0, 0), (0, 1). Let us observe that
1 −1
− am,n (z) = ,
(mz + n)2 (mz + n)2 (mz + n − 1)
where now the first double sum in the last equality is absolutely convergent
and the inner sum in the second double sum telescopes to zero. Therefore
∞
X X 1
G2 (z) = 2ζ(2) + − am,n (z) .
(mz + n)2
m6=0 n=−∞
so that
∞
z2
1 X X
G2 − = 2ζ(2) +
z n=−∞
(nz − m)2
m6=0
XX z2
= 2ζ(2)(1 + z 2 ) +
(nz − m)2
m6=0 n6=0
∞
X X z2
= 2ζ(2)z 2 + .
m=−∞ n6=0
(nz − m)2
∞ πn z
1 X π(n − 1) z
π cot − − π cot +
z z n−1 z n
n6=−∞
n6=0,1
X 1 1
X 1 1
+ − + − .
mz − 1 mz mz mz + 1
m6=0 m6=0
X 1 1
1 X
1 1
− =− +
mz − 1 mz + 1 z (1/z) + m (1/z) − m
m6=0 m6=0
2 π
=− π cot −z
z z
−2π π
= cot + 2.
z z
We now examine
∞ πn z
1 X π(n − 1) z
π cot − − π cot +
z z n−1 z n
n6=−∞
n6=0,1
N
1 X π(n − 1) πn z z
= lim π cot − π cot + − .
z N →∞ z z n n−1
n6=−N
n6=0,1
Keeping in mind that cotangent is an odd function, we easily see that this
telescopes to
1 π π(N + 1) πN z z
lim 2π cot − π cot − π cot + + − 2z
z N →∞ z z z N N +1
2π π 2π πN
= cot −2− lim cot .
z z z N →∞ z
e2πi(N/z) + 1
πN 1
lim cot = i lim = i + 2i lim 2πi(N/z) = i.
N →∞ z N →∞ e2πi(N/z) − 1 N →∞ e −1
This proves:
Theorem 5.1.1. For all z ∈ H,
1
G2 − = z 2 G2 (z) − 2πiz.
z
Exercise 5.1.2. Show that E2 (z) has the Fourier expansion at i∞ given by
∞
X
E2 (z) = 1 − 24 σ1 (n)q n ,
n=1
E43 − E62
Theorem 5.1.4 (Jacobi). The cusp form ∆ = ∈ S12 (SL2 (Z)) is Ra-
1728
manujan’s infinite product
∞
Y
∆=q (1 − q n )24 .
n=1
Noting that
∞
qn X
= q mn ,
1 − qn m=1
F 0 (z)
= 2πiE2 (z).
F (z)
F 0 (−1/z) F 0 (z)
2 6z
= 2πi z E2 (z) + = z2 + 12z.
F (−1/z) πi F (z)
F 0 (−1/z) 1 F 0 (z) 12
· 2 = + ,
F (−1/z) z F (z) z
we immediately see that F (−1/z) and z 12 F (z) have the same logarithmic
derivative. Thus F (−1/z) = λz 12 F (z) for some constant λ. Putting z = i,
we have λ = 1, as F (z) 6= 0 for z ∈ H. This completes the proof. t
u
Show that r
1 z
η − = η(z),
z i
where the square root is the branch with non-negative real part.
Exercise 5.1.8. The Serre derivative of a modular form f ∈ Mk (SL2 (Z)) is de-
fined as
1 0 k
θk f := f − E2 f.
2πi 12
Prove that θk : Mk (SL2 (Z)) → Mk+2 (SL2 (Z)).
d
For the differential operator D = q dq , we have
1
DP = (P 2 − Q),
12
1
DQ = (P Q − R),
3
1
DR = (P R − Q2 ).
2
Remark. The reader will note that if we replace q by e2πiz , then P , Q, and
R become E2 , E4 , and E6 respectively. The differential equations stated in
the theorem were first proved by Ramanujan in his 1916 paper using the
theory of elliptic functions. The differential equations have now become
central in the theory of modular forms and have even found application
in Nesterenko’s work in transcendental number theory. Our proof here will
involve only the basic property of the Serre derivative stated in Exercise
5.1.8.
Proof. Noting that if q = e2πiz , then
d 1 d
D=q = ,
dq 2πi dz
coefficients that
1 2 1
DP − P = − Q. t
u
12 12
Theorem 5.2.1. The number of orbits for the action of SL2 (Z) on Xm is σ1 (m),
and a set of orbit representatives can be taken as
ab
: ad = m, 0 ≤ b ≤ d − 1 .
0d
That is,
a a ab
Xm = SL2 (Z) .
0d
ad=m b (mod d)
d>0
Proof. First we show that every matrix in Xm appears in some orbit as given
in the theorem. Indeed, let
αβ
∈ Xm .
γ δ
such that
αβ rs ab ra rb + sd
= = .
γ δ tu 0d ta tb + ud
This suggests we take a = (α, γ), and r = α/a, t = γ/a. As (r, t) = 1, we can
find integers s, u such that
ru − ts = 1.
Then −1
ab rs αβ u −s αβ
= =
0d tu γ δ −t r γ δ
5.2 Hecke operators of level one 61
so that
b = βu − δs
d = −βt + δr.
then we get
a 0 b0
ar br + sd
= .
at bt + ud 0 d0
Thus, t = 0 and ru = det g = 1 so that r = ±1. As a, a0 > 0, we must have
r = 1 so that a = a0 . But then d = d0 since ad = a0 d0 = m. Finally, b + sd = b0
leads to s = 0 since 0 ≤ b ≤ d − 1 and 0 ≤ b0 ≤ d − 1. This completes the
proof. t
u
Proof. First we need to show that Tm (f )|γ = Tm (f ) for all γ ∈ SL2 (Z).
Fix γ ∈ SL2 (Z) and an orbit representative a0 db . Since SL2 (Z) Xm =
Xm SL2 (Z) (simply by taking determinants), we see that
0 0
ab 0 a b
γ=γ
0d 0 d0
a0 b 0
a 0 b0
ab 0
f γ =f γ =f .
0d 0 d0 0 d0
0 0
Moreover, the new representatives a0 db 0 obtained this way are again dis-
tinct as we range over ad = m, 0 ≤ b ≤ d − 1. Indeed, suppose also that
0 0
a 1 b1 a b
γ = γ 00
0 d1 0 d0
After cancelling the γ from both sides, we conclude (as in the proof of The-
orem 5.2.1) that this implies that γ 00 (γ 0 )−1 is the identity and that a0 db =
a 1 b1
0 d1 .
The fact that Tm (f ) is holomorphic at i∞ when f ∈ Mk (SL2 (Z)) and that
Tm (f ) vanishes at i∞ when f ∈ Sk (SL2 (Z)) will follow from the explicit
q-expansion which we develop in Theorem 5.2.4 below. t
u
Exercise 5.2.3. Show that
d−1
(
X
2πinb/d d if d | n
e =
b=0
0 otherwise
Proof. Writing
∞
X
f (z) = λ(n)e2πinz ,
n=0
and recalling the definition (4.2) of the slash operator for matrices in GL+
2 (R),
we see that
d−1 ∞
1 X m k X X
Tm (f )(z) = λ(n)e2πin(az+b)/d .
m d n=0
ad=m b=0
d>0
5.2 Hecke operators of level one 63
Interchanging the two inner sums and using Exercise 5.2.3, we obtain
X m k−1 X∞
Tm (f )(z) = λ(n)e2πinaz/d .
d n=0
ad=m
d>0 d|n
as claimed. t
u
Exercise 5.2.5. Prove that Tm (∆) = τ (m)∆. Deduce that τ (mn) = τ (m)τ (n)
for (m, n) = 1.
Thus, the theory of Hecke operators resolves the first two of Ramanujan’s
conjectures.
We will see below that L∆ (s) converges absolutely for Re(s) > 7 and ex-
tends to an entire function satisfying the functional equation
where Z ∞
Γ(s) = e−t ts−1 dt
0
is Euler’s Γ-function.
In his lectures on the work of Ramanujan, Hardy [13, p. 185] writes “The
functional equation (5.5) must have been familiar to Ramanujan, but I can-
not find an explicit statement of it either in his collected papers or in the
notebooks.” Apparently, equation (5.5) was proved by Wilton in 1929, but it
was Erich Hecke who recognized (5.5) and (5.4) as a special case of a general
theory of modular forms, Dirichlet series, and Euler products. This theory
we now call Hecke theory. In this section, we discuss some aspects of this
theory pertaining to level one.
5.3 Modular forms and Dirichlet series 65
|τ (p)| ≤ 2p11/2
is equivalent to the assertion that the roots of the quadratic polynomial 1 − τ (p)x +
p11 x2 are non-real.
Exercise 5.3.3. Let f ∈ Mk (SL2 (Z)). Show that y k/2 |f (z)| is invariant under
SL2 (Z), where y = Im(z).
k/2
at i∞, then af (n) = O n .
Proof. By the previous exercise, the continuous function φ(z) = y k/2 |f (z)| is
invariant under the modular group. Hence we consider z ∈ F , the standard
fundamental domain. As Im z → ∞, q → 0 and so since f is a cusp form,
Now fix y and vary x between 0 and 1. Then q = e2πi(x+iy) traverses counter-
clockwise the circle Cy of radius e−2πy , centered at zero. By Cauchy’s residue
theorem,
Z Z 1
1
af (n) = f (z)q −n−1 dq = f (x + iy)q −n dx.
2πi Cy 0
Proof. This is immediate from Hecke’s estimate for af (n) given by Theo-
rem 5.3.4. t
u
Theorem 5.3.7. Let f ∈ Sk (SL2 (Z)). The function Lf (s) extends to an entire
function and satisfies the functional equation
Proof. The estimate (5.6) shows that the integral (for fixed y0 > 0)
Z ∞
f (iy)y s−1 dy
y0
We change variables (y 7→ 1/y) in the first integral and the integral becomes
Z ∞
i
f y −s−1 dy.
1 y
so that
5.4 Ramanujan congruences 67
Z ∞ Z ∞
i −s−1 k
f y dy = i f (iy)y k−s−1 dy.
1 y 1
Thus Z ∞
−s
dy
(2π) Γ(s)Lf (s) = f (iy) y s + ik y k−s ,
1 y
from which the functional equation is immediate. t
u
There are many interesting arithmetical properties of the τ -function, too nu-
merous to expound here. We will content ourselves with a sampling of such
results and invite the reader to consult the article [45] by H.P.F. Swinnerton-
Dyer. Here is an example of an elementary congruence for the τ -function.
The function p(n) is the partition function and counts the number of ways
of writing n = λ1 + λ2 + · · · + λr with λ1 ≥ λ2 ≥ · · · ≥ λr ≥ 0, λi ∈ Z.
The partition function appears in many branches of mathematics ranging
from number theory, combinatorics, group theory, representation theory, to
mathematical physics. Ramanujan discovered that
68 5 The Ramanujan τ -function
p(5n + 4) ≡ 0 (mod 5)
p(7n + 5) ≡ 0 (mod 7)
p(11n + 6) ≡ 0 (mod 11),
and these were proved by him using the theory of elliptic functions. These
are often called Ramanujan congruences. We show below how the theory of
modular forms leads to “quick” proofs of these congruences. We follow the
treatment of Berndt [4].
Therefore,
∞ ∞
!
X Y
q p(n)q n
≡ DP (1 − q 25n )−1 (mod 5).
n=0 n=1
762048
Exercise 5.5.1. Prove that E62 = E12 − 691 ∆.
65520
P∞
Exercise 5.5.2. Show that E12 = 1 + 691 n=1 σ11 (n)q n .
Deduce that
n−1
X 65
σ11 (n) + O n6 .
σ5 (j)σ5 (n − j) =
j=1
(252)(691)
Exercise 5.5.7. If k1 > k2 ≥ 4 are even, prove that for any > 0,
n−1
X
σk1 −1 (j)σk2 −1 (n − j) = c(k1 , k2 )σk1 +k2 −1 (n) + O nk1 −1+ ,
j=1
Show that
αk+1 − β k+1
fk = .
α−β
If α = β, the right hand side is viewed as αk + αk−1 β + · · · + αβ k−1 + β k =
(k + 1)αk .
holds for all natural numbers m and n. Show that f is g-multiplicative if and only
if f is multiplicative and for all primes p and all natural numbers k ≥ 1, we have
Exercise 5.5.10. Let gk (n) = nk . Show that the Ramanujan τ -function is g11 -
multiplicative.
Show that X
G(m, n) = µ(d)F (m/d, n/d)
d|m,n
Exercise 5.5.14. Let A be the n×n matrix whose (i, j) entry is τ (ij). Show that the
determinant of A is given by n!11 µ(1)µ(2) · · · µ(n). In particular, the determinant
is zero if n ≥ 4.
Exercise 5.5.15. Show that τ (n) is odd if and only if n is an odd square.
5.5 Supplementary problems 71
Fk := E12r−k+2 ∆−r .
Exercise 5.5.17. Let j denote the j-function. Show that for any integer m ≥ 0,
dj
jm
dz
has a q-expansion without a constant term.
Exercise 5.5.18. Let f be a modular form of weight k for the full modular group
with q-expansion
f (z) = a0 + a1 q + a2 q 2 + · · ·
With Ckj defined as in Exercise 5.5.16, show that
Exercise 5.5.19. Suppose that k ≡ 2 (mod 4). With notation as in the previous
exercise, show that Ck0 6= 0.
Exercise 5.5.20. Suppose that k ≡ 0 (mod 4). With notation as in the previous
exercise, show that Ck0 6= 0.
Chapter 6
Modular Forms of Higher Level
H? = H ∪ Q ∪ {i∞}.
The reader may wish to stop and review Section 4.1 before proceeding.
The punctured fundamental neighborhood
Uc = {z ∈ H : Im(z) > c}
q 1/h 7→ f (z)
© Springer Science+Business Media Singapore 2016 and Hindustan Book Agency 2016 73
M.R. Murty et al., Problems in the Theory of Modular Forms, HBA Lecture Notes
in Mathematics, DOI 10.1007/978-981-10-2651-5_6
74 6 Modular Forms of Higher Level
∞
X
f (q 1/h ) = an q n/h .
n=−∞
We call
∞
X
f (z) = an e2πizn/h
n=−∞
where z ∈ H is any point such that q 1/h = e2πiz/h . Now (f |γ) will have a
Laurent expansion
X∞
(f |γ)(q 1/h ) = bn q n/h
n=−∞
1/h
centered at q = 0. We say that
∞
X
(f |γ)(z) = bn e2πizn/h (6.2)
n=−∞
1 h0
∈ γ −1 Γγ
0 1
or
−1 h0
∈ γ −1 Γγ
0 −1
holds and write the expansion of f |γ as
∞
X 0
(f |γ)(z) = cn e2πizn/h . (6.3)
n=−∞
Now we define the order of f at γ(i∞) to be the index of the least non-zero
coefficient in (6.3):
1
vγ(i∞) (f ) := inf n ∈ Z : cn 6= 0 .
2
Exercise 6.2.1. Let k ≥ 3 and N a natural number. Let g ∈ (Z/N Z)2 . Define the
Eisenstein series
76 6 Modular Forms of Higher Level
X 0 1
Gk,g (z) = , z∈H
(mz + n)k
(m,n)≡g (mod N )
where the summation is over all integers m, n satisfying the congruence condition
(m, n) ≡ g (mod N ) and the dash on the summation means we exclude (m, n) =
(0, 0) in the case g = (0, 0). For any γ ∈ SL2 (Z), show that Gk,g |γ = Gk,gγ .
Conclude that Gk,g ∈ Mk (Γ(N )).
Exercise 6.2.2. Show that Gk,g = (−1)k Gk,−g . Deduce that if k is odd and 2g ≡
(0, 0) (mod N ), then Gk,g = 0.
11 10
6.2.5. Show that Γ0 (4) is generated by T = ( 0 1 ), by ( 4 1 ), and by
Exercise
−1 0
0 −1 .
Exercise 2.3.3 shows that [SL2 (Z) : Γ0 (4)] = 6 and the reader may easily
verify that
10 10 0 −1 0 −1 0 −1 0 −1
, , , , , (6.4)
01 21 1 1 1 2 1 3 1 4
Exercise 6.2.6. Show that there are exactly three inequivalent cusps for the group
Γ0 (4). Find the widths of each cusp.
As in the previous section, let g = (a1 , a2 ) ∈ (Z/N Z)2 and consider Gk,g (z).
We derive the q-expansion of Gk,g (z). Let us write
6.3 Fourier expansions of higher level Eisenstein series 77
∞
X
Gk,g (z) = bk,g (0) + bk,g (n)e2πinz/N . (6.5)
n=1
Theorem 6.3.1. The Fourier coefficients bk,g (n) in (6.5) are given as follows:
(
0 if a1 6≡ 0 (mod N ),
bk,g (0) = P −k
m≡a2 (mod N ) m if a1 ≡ 0 (mod N ),
and
(−2πi)k X
bk,g (n) = dk−1 (sgn d)e2πia2 d/N
N k (k − 1)!
d|n
n/d≡a1 (mod N )
where the summation is over all divisors of n (positive and negative) and
(
+1 if d > 0,
sgn d =
−1 if d < 0.
Proof. The computation of bk,g (0) was done at the end of the first paragraph
of the solution to Exercise 6.2.1. We separate this term out and write
k
X X 1
Gk,g (z) = bk,g (0) + .
mz + n
m≡a1 (mod N ) n≡a2 (mod N )
m6=0
X X mz + a2 −k
−k −k
(mz + N j + a2 ) =N +j .
N
j∈Z j∈Z
As in the level one case, we use Lipschitz’s formula (Theorem 4.2.2) which
states that for k ≥ 2 and z ∈ H, we have
∞
X 1 (−2πi)k X k−1 2πinz
= n e .
(z + n)k (k − 1)! n=1
n∈Z
Since (mz + a2 )/N ∈ H if and only if m > 0, we first consider the cases with
m > 0 for which we have
−k ∞
X mz + a2 (−2πi)k X k−1 2πida2 /N 2πidmz/N
+j = d e e .
N (k − 1)!
j∈Z d=1
X m0 z − a2 −k ∞
(−2πi)k X k−1 −2πida2 /N 2πidm0 z/N
(−1)k −j = (−1)k d e e .
N (k − 1)!
j∈Z d=1
(−2πi)k X
bk,g (n) = dk−1 (sgn d) e2πia2 d/N
N k (k − 1)!
d|n
n/d≡a1 (mod N )
allowing for both positive and negative integers d that divide n in the sum-
mation. t
u
(−2πi)k X
bk,g (n) = dk−1 (sgn d).
N k (k − 1)!
d|n
n/d≡a1 (mod N )
Exercise 6.3.3. If g = (0, a2 ) ∈ (Z/N Z)2 , show that bk,g (n) = 0 if N - n, and
where ζ = e2πi/N . Deduce that Gk,g (z), with g = (0, a2 ), has a q-expansion of the
form
X∞
an q n , q = e2πiz .
n=0
Exercise 6.3.4. Prove that if g = (0, a2 ) ∈ (Z/N Z)2 , then Gk,g ∈ Mk (Γ1 (N ))
for k ≥ 3.
It is possible to derive the analogs of the valence and dimension formulas for
any congruence subgroup Γ of SL2 (Z). The general method used to derive
these formulas is via the Riemann-Roch theorem, which is usually a topic in
Riemann surface theory and algebraic geometry. Thus, we content ourselves
by merely stating the results and referring the interested reader to [9] for
further details. Here are the analogues of Theorem 4.3.1 and Corollary 4.4.2,
for an arbitrary congruence subgroup Γ of SL2 (Z).
We first need to define elliptic points. A point z ∈ H is called an elliptic
point for Γ if {±I} Γz is strictly larger than {±I}, where Γz = {γ ∈ Γ : γz =
z} is the stabilizer subgroup of z. In other words, z is an elliptic point for Γ
if and only if {±I}Γz 6= {±I}. We define the order of an elliptic point to be
|{±I}Γz /{±I}|.
Exercise 6.4.1. Show that the only elliptic points for SL2 (Z) lying in the standard
fundamental domain F are i, ρ = eπi/3 , and ρ2 .
Exercise 6.4.2. Find the stabilizer subgroups SL2 (Z)i and SL2 (Z)ρ in SL2 (Z) and
show that SL2 (Z)i / {±I} and SL2 (Z)ρ / {±I} have orders two and three, respec-
tively.
Exercise 6.4.3. If Γ has finite index in SL2 (Z), show that there are only finitely
many Γ-inequivalent elliptic points for Γ.
Exercise 6.4.5. Verify that Theorem 6.4.4 reduces to Theorem 4.3.1 in the case Γ =
SL2 (Z).
Theorem 6.4.6. (The dimension formula) Let Γ be a congruence subgroup of SL2 (Z)
and k an even non-negative integer. Then
and
80 6 Modular Forms of Higher Level
dim Mk (Γ) − ∞ if k ≥ 4
dim Sk (Γ) = g if k = 2 ,
0 if k = 0
then f = 0.
Proof (Sketch). If {±I} Γ = SL2 (Z), the result is clear from the valence for-
mula (Theorem 4.3.1). Let M = [SL2 (Z) : {±I} Γ] and 1 = γ1 , . . . , γM be a
complete set of coset representatives for {±I} Γ in SL2 (Z). Then, the func-
tion
M
Y
F (z) = (f |γj )(z)
j=1
is clearly an element of MkM (SL2 (Z)) since for γ ∈ SL2 (Z), the products γj γ,
1 ≤ j ≤ M , are again a complete set of coset representatives so that F |γ = F
for all γ ∈ SL2 (Z).
Consider one of the Γ-inequivalent
n cusps rj . Ifothis cusp has width h, then
(1) (h)
there are h coset representatives γrj , . . . , γrj ⊆ {γ1 , . . . , γM } such that
(s)
rj = γrj (i∞).Moreover, each of the corresponding Fourier expansions will
have the same order:
(f |γr(s)
j
)(z) = bs q vrj (f )/h + · · · .
By the level one case, we deduce that F = 0 and hence some f |γj = 0. But
then f = (f |γj )|γj−1 = 0|γj−1 = 0. t
u
6.5 The Jacobi four-square theorem revisited 81
∞
X ∞
X
Exercise 6.4.8. Let f = an q n and g = bn q n ∈ Mk (Γ) for some con-
n=0 n=0
gruence subgroup Γ of SL2 (Z). Suppose that an = bn for all n ≤ k[SL2 (Z) :
{±}Γ]/12. Show that f = g.
Exercise 6.4.9. For any congruence subgroup Γ of SL2 (Z), prove that
k
dim Mk (Γ) ≤ [SL2 (Z) : {±}Γ] + 1.
12
Exercise 6.4.10. Prove that
Exercise 6.4.12. Show that dim M2 (Γ0 (4)) = 2 and that {E2,2 , E2,4 } is a basis
for M2 (Γ0 (4)).
Exercise 6.5.1. Let η(z) denote the Dedekind η-function. Prove that
∞
X 2 η 5 (2z)
θ(z) := qn = ,
n=−∞
η 2 (z)η 2 (4z)
where q = e2πiz .
Then
82 6 Modular Forms of Higher Level
√
z
θ = 4z + 1 θ(z),
4z + 1
where the branch of the square root is the principal branch (whose image is contained
in the right half-plane).
Proof. We use Exercises 6.5.1 and 5.1.5. From the former we have
η 5 (2z)
θ(z) =
η 2 (z)η 2 (4z)
Finally,
−1 ! s
z 1 4z + 1 1
η =η − − −4 = i η − −4
4z + 1 z z z
s
−πi 4z + 1 1
=e 3 i η −
z z
s r
−πi 4z + 1 z
=e 3 i η(z).
z i
√
z
θ = 4z + 1 θ(z). t
u
4z + 1
Theorem 6.5.3 (Jacobi’s four square theorem). We have θ4 ∈ M2 (Γ0 (4)) and
X
r4 (n) = 8 d.
d|n
4-d
z
Proof. By Theorem 6.5.2, θ4 4z+1 = (4z + 1)2 θ4 (z). Since θ4 (z + 1) = θ4 (z)
and since Γ0 (4) is generated by ( 10 11 ) and ( 14 01 ) and −I by Exercise 6.2.5,
we deduce θ4 ∈ M2 (Γ0 (4)). By Exercise 6.4.12, E2,2 and E2,4 comprise a
basis for M2 (Γ0 (4)). Thus θ4 = aE2,2 + bE2,4 . A quick calculation comparing
coefficients of q-expansions of both sides of this equation leads to
1
θ4 = − E2,4 ,
3
which leads to the formula
r4 (n) = 8 σ(n) − 4σ(n/4) ,
Exercise 6.6.3. Let k ≥ 6 be even. Show that Sk (Γ0 (4)) = η 12 (2z)Mk−6 (Γ0 (4)).
k
Exercise 6.6.4. For even k ≥ 0, prove that dim Mk (Γ0 (4)) = 2 + 1.
Exercise 6.6.5. Let r12 (n)P be the number of ways of writing n as a sum of 12
∞
squares and let η 12 (2z) = n=1 an q n . Show that
n
r12 (n) = 8σ5 (n) − 512σ5 + 16an .
4
Exercise 6.6.6. Let f (z) ∈ Mk (Γ1 (N )) and suppose r ∈ Z is positive. Show that
f (rz) ∈ Mk (Γ1 (rN )). Moreover, if f (z) is a cusp form, then so is f (rz).
Chapter 7
The Petersson Inner Product
Exercise 7.1.1. Let f , g ∈ Mk (Γ) and let F be any fundamental domain for the
congruence subgroup Γ. Show that the integral
ZZ
dxdy
y k f (z)g(z) 2 < ∞
F y
© Springer Science+Business Media Singapore 2016 and Hindustan Book Agency 2016 85
M.R. Murty et al., Problems in the Theory of Modular Forms, HBA Lecture Notes
in Mathematics, DOI 10.1007/978-981-10-2651-5_7
86 7 The Petersson Inner Product
for Γ1 . Now
ZZ
1 dxdy
y k f (z)g(z) 2
SL2 (Z) : Γ1 F1 y
1 X ZZ dxdy
= y k f (z)g(z) 2
SL2 (Z) : Γ2 · Γ2 : Γ1 j −1
αj (F2 ) y
Z Z
1 X dxdy
= (Im αj−1 z)k f (αj−1 z)g(αj−1 z) 2 .
SL2 (Z) : Γ2 · Γ2 : Γ1 j F2 y
There are Γ2 : Γ1 summands and, since f, g ∈ Sk (Γ2 ), we find that they
are all equal to ZZ
dxdy
y k f (z)g(z) 2 ,
F2 y
so that the original integral is
ZZ
1 dxdy
y k f (z)g(z) ,
y2
SL2 (Z) : Γ2 F2
as desired. t
u
Exercise 7.1.4. Suppose α ∈ GL+ has integer entries. Let D = det α. If
2 (Q)
Γ(N ) ⊆ Γ for some N , show that Γ(N D) ⊆ α−1 Γα and Γ(N D) ⊆ αΓα−1 .
Exercise 7.1.5. Let α ∈ GL+
2 (Q) have integer entries. Show that there exists γ ∈
SL2 (Z) such that
ab
γ −1 α = ,
0d
where a and d are positive integers and b ∈ Z.
Exercise 7.1.6. Let f ∈ Mk (Γ) with Γ a congruence subgroup of SL2 (Z). Let α ∈
GL+2 (Q). Show that f |α is holomorphic at every cusp s ∈ Q ∪ {i∞}.
Exercise 7.1.7. Let f ∈ Sk (Γ) with Γ a congruence subgroup of SL2 (Z). Let α ∈
GL+2 (Q). Show that f |α vanishes at every cusp s ∈ Q ∪ {i∞}.
Theorem 7.1.10. Let f, g ∈ Sk (Γ) with Γ a congruence subgroup of SL2 (Z). Let
α ∈ GL+
2 (Q). Then
(f, g) = (f |α, g|α),
where the latter inner product is with respect to Γ0 = α−1 Γα ∩ Γ.
Proof. By Theorem 7.1.8, f |α, g|α ∈ Sk (Γ0 ) where Γ0 = α−1 Γα ∩ Γ. Let Γ00 =
Γ ∩ αΓα−1 . Then α−1 Γ00 α = Γ0 , and so if F 00 is a fundamental domain for Γ00 ,
then α−1 F 00 is a fundamental domain for Γ0 , by Exercise 7.1.9. Thus,
ZZ
1 dxdy
(f |α, g|α) = 0
(f |α)(z)(g|α)(z)y k 2
SL2 (Z) : Γ −1
α F 00 y
ZZ
1 dxdy
= 0
f (z)g(z)y k 2
SL2 (Z) : Γ F 00 y
= (f, g)
Exercise 7.1.11. Let α ∈ GL+ 2 (Q) have integer entries and determinant D. Let
α0 = Dα−1 . Show that for f, g ∈ Sk (Γ), with Γ a congruence subgroup of SL2 (Z),
we have
(f |α, g) = (f, g|α0 ).
Let Tn denote the nth Hecke operator for SL2 (Z), as developed in Section 5.2.
We prove that Tn Tm = Tm Tn so that the Tn ’s form a commuting family of
linear transformations on Mk (SL2 (Z)) and Sk (SL2 (Z)). We will use this fact
along with the multiplicative properties of the Hecke operators in the next
section to prove that the Tn ’s are Hermitian with respect to the Pettersson
inner product.
88 7 The Petersson Inner Product
∞
X
By Theorem 5.2.4, we have Tn (f ) = b(r)q r , where
r=0
X nr
b(r) = dk−1 λ .
d2
d|(n,r)
∞
X
Writing Tm Tn (f ) = c(r)q r , we have
r=0
X mr X X mnr
c(r) = ek−1 b = ek−1 dk−1 λ .
e2 d2 e2
e|(m,r) e|(m,r) d|( n, mr )
e2
As (m, n) = 1, the inner sum over divisors d of (n, mr/e2 ) is really a sum
over divisors d of (n, r/e) as (n, m/e) = 1. Therefore,
X X mnr
c(r) = ek−1 dk−1 λ 2 2
d e
e|(m,r) d|(n, re )
X X mnr
= (ed)k−1 λ 2 2 .
d e
e|(m,r) d|(n,r)
As d runs over divisors of (n, r) and e runs over divisors of (m, r), de runs
over divisors of (mn, r). Thus
X mnr
c(r) = tk−1 λ
t2
t|(mn,r)
Proof. Since the Hecke operators are defined in terms of the slash operator,
it is notationally convenient to introduce
j
−1 s−j
s pX
X p b
Tf
p s = ,
0 pj
j=0 b=0
7.2 Commutativity of the Hecke operators 89
where we view the right hand side as a “formal sum” of matrices. Then
k
ps(1− 2 ) Tps (f ) = f |Tf
ps .
Thus,
p−1
p0 X 1c
T
fp = + .
01 0p
c=0
Observe that for all a, b, d ∈ Z and all f ∈ Mk (SL2 (Z)), by the definition
of the slash operator (4.2) we have
ab a k/2 a b
f = f z+ , (7.1)
0d d d d
a0
f = f (z) . (7.2)
0a
Since f ∈ Mk (SL2 (Z)) is periodic, the right hand side of (7.1) depends only
on b (mod d), rather than b itself. In particular, we actually view Tfps as an
equivalence class of formal sums of matrices in which matrices are equiva-
lent if their corresponding slash operators are equal.
Although notationally Tps is a left action, Tf
ps acts from the right. Hence
we consider
j j
s pX −1 p−1 X
s pX −1
ps−j b
X s−j
p0 X 1c p b
T
fp Tf
ps = + . (7.3)
01 0 pj 0p 0 pj
j=0 b=0 c=0 j=0 b=0
Now
ps−j b ps+1−j bp
p0
=
01 0 pj 0 pj
and
ps−j b ps−j b + pj c
1c
=
0p 0 pj 0 pj+1
so that the first sum of (7.3) can be rewritten as
j
−1 s+1−j
s pX
ps+1 0
X p bp
+ , (7.4)
0 1 0 pj
j=1 b=0
In this last sum, the two inner summations can be combined to give
90 7 The Petersson Inner Product
pj+1
X−1 s−j b0
p
.
0 pj+1
b0 =0
T
fp Tf
ps = Tps+1 + pTps−1 .
] ]
s( k −1)
Since Tf
ps = Tps /p 2 , we immediately deduce the theorem. t
u
Tp r Tp s = Tp s Tp r .
It may be easier to prove Exercise 7.2.4 before Exercise 7.2.3. In fact, after
Exercise 7.2.4, Exercise 7.2.3 becomes a trivial statement about commuting
polynomials.
Exercise 7.2.4. Let p be prime. Show that there is a polynomial Ps (x) ∈ Z[x] such
that Tps = Ps (Tp ).
Exercise 7.2.5. Prove that for all m, n ≥ 1, the Hecke operators commute:
Tn T m = Tm T n .
Proof. By Theorem 7.2.1 and Exercise 7.2.4 we see that the algebra of Hecke
operators are generated by the Tp ’s, with p prime. Thus, to prove the theo-
rem, it suffices to show (Tp (f ), g) = (f, Tp (g)) for all primes p. By the defini-
tion of Tp ,
7.3 Hecke operators as Hermitian operators 91
p−1
( )
k
2 −1
p0 X 1b
Tp (f ) = p f + f
01 0p
b=0
and so by linearity
p−1
( )
k
2 −1
p0 X 1b
(Tp (f ), g) = p f ,g + f ,g
01 0p
b=0
and
1b p −b
f ,g = f, g .
0p 0 1
1 ≤ b ≤ p − 1, we observe that α := 01 −1
For each b ∈ SL2 (Z) and β :=
0 1
−1 −b ∈ SL2 (Z). Now by Theorem 7.1.10, we have
p −b p −b
f, g = f |β, g β
0 1 0 1
Exercise 7.3.4. Let f , g ∈ Sk (SL2 (Z)) be normalized eigenfunctions for all the
Hecke operators. Show that f = g or (f, g) = 0.
j j
!
X X
0 6= (fj+1 , fj+1 ) = αi fi , fj+1 = αi (fi , fj+1 ) = 0.
i=1 i=1
V1 = {f ∈ V : S1 f = λ1 f }
V = Cf1 ⊕ (Cf1 )⊥ ,
where
(Cf1 )⊥ = {g ∈ V : (f1 , g) = 0}.
7.5 Supplementary problems 93
Clearly, the Sj ’s preserve this space since for g ∈ (Cf1 )⊥ , we have (f1 , Sj g) =
(Sj f1 , g) = 0 since f1 is an eigenfunction for the Sj ’s. By the induction hy-
pothesis on the dimension of V , we see that (Cf1 )⊥ has an orthogonal basis
f2 , . . . , fr . This completes the proof. t
u
Theorem 7.4.2. The space Sk (SL2 (Z)) has an orthogonal basis of Hecke eigen-
forms.
Proof. We apply Theorem 7.4.1 to the commutative ring of Hecke operators
acting on the finite dimensional vector space Sk (SL2 (Z)) to deduce the re-
sult. t
u
Exercise 7.4.3. Show that Ek is an eigenfunction for all the Hecke operators acting
on Mk (SL2 (Z)).
Exercise 7.4.4. Show that (Ek , f ) = 0 for any f ∈ Sk (SL2 (Z)).
Exercise 7.4.5. Show that ∆k is not a Hecke eigenform for any k ≥ 2.
Exercise 7.5.1. Let 12 | k and d = dim Sk (SL2 (Z)). For each 1 ≤ j ≤ d, define
∞
2(d−j)
X
j
fj := ∆ E6 = a(j) n
n q .
n=1
(j) (j)
Verify that an = 0 for n < j and aj = 1. Conclude that the fj form a basis for
Sk (SL2 (Z)).
Exercise 7.5.2. Let k ≥ 4 be even and d = dim Sk (SL2 (Z)). Choose non-negative
integers a, b such that 12 6= 4a + 6b ≤ 14 and 4a + 6b ≡ k (mod 12). For each
1 ≤ j ≤ d, define
∞
2(d−j)+b a
X
fj := ∆j E6 E4 = a(j) n
n q .
n=1
(j) (j)
Verify that an = 0 for n < j and aj= 1. Conclude that the fj form a basis for
Sk (SL2 (Z)). (This basis is called the Miller basis in the literature.)
Exercise 7.5.3. Let k, d, a, and b be as in Exercise 7.5.2. Verify that there exists a
basis g0 , g1 , . . . , gd for Mk (SL2 (Z)) such that
∞
X
gi = q i + b(i)
n q
n
n=d+1
(i)
with bn ∈ Z for all n and all i. Conclude that if
94 7 The Petersson Inner Product
∞
X
f= cn q n ∈ Mk (SL2 (Z))
n=0
Exercise 7.5.4. Let d = dim Sk (SL2 (Z)) and fix a positive integer m. Suppose
∞
X
f= cn q n ∈ Mk (SL2 (Z))
n=0
Exercise 7.5.5. Let f be a Hecke eigenform for the full modular group. Show that
the Fourier coefficients of f are algebraic integers.
Chapter 8
Hecke Operators of Higher Level
with suitable γj in Γ2 .
Exercise 8.1.1. Let p be prime and Γ = SL2 (Z). Let Xp be the set of matrices with
integer entries of determinant p. Show that
p0
Xp = Γ Γ.
01
Xn = Xn (N, S × , S + )
ab × +
:= : a, b, c, d ∈ Z, ad − bc = n, N |c, a ∈ S , b ∈ S .
cd
© Springer Science+Business Media Singapore 2016 and Hindustan Book Agency 2016 95
M.R. Murty et al., Problems in the Theory of Modular Forms, HBA Lecture Notes
in Mathematics, DOI 10.1007/978-981-10-2651-5_8
96 8 Hecke Operators of Higher Level
Example 8.1.3. The standard Hecke congruence subgroups can all be writ-
ten in this form:
Γ1 (N ) = X1 (N, 1, Z)
Γ(N ) = X1 (N, 1, N Z)
Γ0 (N ) = X1 (N, (Z/N Z)× , Z).
Xm Xn ⊆ Xmn .
We define the nth Hecke operator on Mk (X1 ) as the map which sends f ∈
Mk (X1 ) to the finite sum
k
X
Tn (f ) := n 2 −1 f |αi . (8.2)
i
Proof. Suppose f ∈ Mk (Γ) and γ ∈ Γ. Then for any αi , since f |γαi = f |αi ,
we see that the map does not depend on the choice of orbit representatives
and is thus well-defined. Moreover,
8.1 Hecke operators for congruence subgroups 97
!
k
X
Tn (f ) γ = n 2 −1 f |αi γ
i
k
X
= n 2 −1 f |αi γ
i
= Tn (f ),
where the final step follows from (8.1) which shows that γ simply permutes
the orbits. By Exercise 7.1.6, each f |αi is holomorphic at every cusp Q ∪
{i∞} and hence the finite sum Tn (f ) is also holomorphic at every cusp. We
conclude that Tn (f ) ∈ Mk (Γ). If in addition f ∈ Sk (Γ), then by Exercise 7.1.7
we conclude that Tn (f ) ∈ Sk (Γ). The linearity of the map follows from the
linearity of the slash operator. t
u
Theorem 8.1.7. For each a ∈ (Z/N Z)× , fix σa ∈ SL2 (Z) such that
−1
a 0
σa ≡ (mod N ),
0 a
where the first disjoint union is taken over all positive integers a dividing n that are
coprime to N .
Proof. The terms on the right hand side are clearly contained in Xn . We first
show the union is a disjoint union. Suppose we had
a 1 b1 a 2 b2
γ 1 σa1 = γ 2 σa2 , γ1 , γ2 ∈ Γ1 (N ).
0 d1 0 d2
Now
αβ ??
A =
γ δ 0?
and has determinant n. Replacing A by ± 10 1j A if necessary, we may sup-
pose that
αβ ab a > 0,
A = ,
γ δ 0d 0 ≤ b ≤ d − 1.
Since αγ βδ ∈ Xn , considered modulo N this equality gives
1? ab
A ≡ (mod N )
0n 0d
A−1 = Bσa .
8.2 Nebentypus
Exercise 8.2.2. Show that the set of Dirichlet characters modulo N forms a group
of order φ(N ) under multiplication, where φ denotes the Euler function.
1 X
fχ = χ(d)f |γd ,
φ(N )
d∈(Z/N Z)×
1 X 1 X
fχ |γ = χ(d)f |γd γ = χ(d)f |γdδ
φ(N ) φ(N )
d∈(Z/N Z)× d∈(Z/N Z)×
1 X
= χ(dδ)χ(δ)f |γdδ .
φ(N )
d∈(Z/N Z)×
Now as d runs through coprime residue classes modulo N , so does dδ. Thus,
we can change variables in the sum and deduce that
fχ |γ = χ(δ)fχ .
and the inner sum is zero by Exercise 8.2.3 unless d = 1, in which case it is 1
so that X
f= fχ .
χ
To show the sum is a direct sum, we consider g ∈ Mk (Γ0 (N ), ψ). For any
character χ, we compute
(
1 X g if χ = ψ
gχ = ψχ(d) g =
φ(N ) 0 otherwise,
d∈(Z/N Z)×
then
X
g = gψ = gχ = 0. t
u
χ6=ψ
ψ
d−1
k
2 −1
X X ab
Tn (f ) = n χ(a) f .
0d
ad=n b=0
a>0
for a ≥ 1.
As before, we deduce that the Tn ’s form a commutative algebra of oper-
ators. We now state (without proof) the action of these Hecke operators on
the Petersson inner product.
Theorem 8.2.11. Let f ∈ Mk (Γ1 (N )), and g ∈ Sk (Γ1 (N )). For each a ∈
(Z/N Z)× , fix σa ∈ SL2 (Z) such that
−1
a 0
σa ≡ (mod N ),
0 a
102 8 Hecke Operators of Higher Level
Proof. This is clear from the theorem. Indeed, since cn is a root of unity, we
have (cn Tn (f ), g) = cn (Tn (f ), g) = cn χ(n)(f, Tn (g)) = cn cn 2 (f, Tn (g)) =
cn (f, Tn (g)) = (f, cn Tn (g)). t
u
Exercise 8.2.13. Show that Sk (Γ0 (N ), χ) has a basis of eigenforms for all the Tn
with (n, N ) = 1.
Exercise 8.2.13 only gives us a basis of eigenforms for all the Tn ’s under the
restriction that (n, N ) = 1. One would like a basis for all Tp ’s, including p|N .
Since all the Tp ’s commute, we know that the Tp ’s preserve each eigenspace.
If each eigenspace were one-dimensional, then we would deduce that the
basis is also a basis of eigenforms for Tp ’s with p|N . However, this need not
be the case. The problem arises from eigenforms arising from lower levels,
as the following examples show.
Solution. We have
r0 k
X X r0 ab
Tn f = n 2 −1 χ(a) f .
01 01 0d
ad=n b (mod d)
a>0
Now
r0 ab a br r0
=
01 0d 0 d 01
so that the sum becomes
k
2 −1
X X a br r0
n χ(a) f .
0 d 01
ad=n b (mod d)
a>0
8.4 Supplementary problems 103
Exercise 8.3.2. Show that ∆(z) and ∆(2z) are both eigenforms for the Tn ’s (n odd)
acting on S12 (Γ0 (2)), with the same eigenvalues.
Exercise 8.3.3. Show that ∆(z) and ∆(2z) are linearly independent in S12 (Γ0 (2)).
In 1970, Atkin and Lehner realized how to resolve this difficulty by fo-
cusing on forms “genuinely” of level N . Indeed, let d1 d2 = N . Suppose f ∈
Mk (Γ1 (d1 )). Then f ∈ Mk (Γ1 (N )). If g(z) := f (d2 z), then g ∈ Mk (Γ1 (N ))
by Exercise 6.6.6. The subspace of Sk (Γ1 (N )) spanned by forms obtained in
these two ways from f ∈ Sk (Γ1 (d)), as we range over proper divisors d|N
is called the space of oldforms, denoted Skold (Γ1 (N )). One can show that the
Tn ’s preserve this space. The orthogonal complement of this space with re-
spect to the Petersson inner product is called the space of newforms, denoted
Sknew (Γ1 (N )). It turns out that this space has the multiplicity one property:
the intersection of the eigenspaces for all the Tn ’s with (n, N ) = 1 is one-
dimensional and so is also an eigenspace for the Tp ’s with p|N . This fact al-
lows us to construct a complete theory of L-series and functional equations
analogous to the level one case.
0 −1 −1
Exercise 8.4.1. Let wN = N 0 . Prove that wN Γ0 (N )wN ⊂ Γ0 (N ).
Exercise 8.4.2. Show that if f ∈ Mk (Γ0 (N )), then f |wN ∈ Mk (Γ0 (N )).
Exercise 8.4.3. Show that if f ∈ Sk (Γ0 (N )), then f |wN ∈ Sk (Γ0 (N )).
Exercise 8.4.4. Show that if f ∈ Mk (Γ0 (N )), then (f |wN ) |wN = f . Thus wN is,
in fact, an involution.
Chapter 9
Dirichlet Series and Modular Forms
The general Dirichlet series is a discrete analog of the more familiar Laplace
transform of a function in the theory of complex variables. It is of the form
∞
X
an e−λn s (9.1)
n=1
with x = e−s . The reader is probably familiar with the theory of power
series from a first course in calculus. Power series have a wide variety of
applications, most notably to combinatorial problems.
Ordinary Dirichlet series (which we will now consider) arise with λn =
log n. Then (9.1) becomes
∞
X an
(9.2)
n=1
ns
and such series are useful in analytic number theory, especially in the con-
text of multiplicative number theory. Indeed, series of the form (9.2) were in-
troduced in 1839 by P.G.L. Dirichlet to prove his celebrated theorem on the
infinitude of primes in a given arithmetic progression. However, Dirichlet
studied such series only for a real variable s. It was Riemann who initiated
in 1860 an ardent study of such series for complex variables of s. In his fa-
mous memoir, he focused on the special case of (9.2) with an = 1 for all n
© Springer Science+Business Media Singapore 2016 and Hindustan Book Agency 2016 105
M.R. Murty et al., Problems in the Theory of Modular Forms, HBA Lecture Notes
in Mathematics, DOI 10.1007/978-981-10-2651-5_9
106 9 Dirichlet Series and Modular Forms
and introduced (what we now call) the Riemann zeta function ζ(s), which
for Re(s) > 1 is given by the Dirichlet series
∞
X 1
s
. (9.3)
n=1
n
Exercise 9.1.1. Show that the series (9.3) converges absolutely for Re(s) > 1.
Exercise 9.1.2. Given a sequence of complex numbers an , we suppose that for any
> 0, there is a constant C() such that |an | ≤ C()n for all natural numbers n.
Show that
∞
X an
n=1
ns
for Re(s) > 1. It is this connection that relates the distribution of prime
numbers with the zeros of the Riemann zeta function.
Exercise 9.1.4. Let f (n) be a multiplicative function. That is, f (mn) = f (m)f (n)
for (m, n) = 1. Assume that for any > 0, there is a constant C() such that
|f (n)| ≤ C()n . Show that for Re(s) > 1, we have
∞
f (p) f (p2 )
X f (n) Y
= 1 + + + · · · ,
n=1
ns p
ps p2s
Proposition 9.1.5 (Abel’s Lemma). Let m ∈ Z and suppose an and bn are se-
quences of complex numbers. Let
X
A(r) = an .
m≤n≤r
9.1 General Dirichlet series 107
Then
X X
an bn = A(r)br − A(m − 1)bm − A(n)(bn+1 − bn ).
m≤n≤r m≤n<r
Proof. We have
X X
a n bn = A(n) − A(n − 1) bn
m≤n≤r m≤n≤r
X X
= A(n)bn − A(n)bn+1 ,
m≤n≤r m−1≤n≤r−1
after a change of the index of summation. The two sums can be combined to
give X
A(r)br − A(m − 1)bm − A(n) (bn+1 − bn ) . t
u
m≤n<r
Exercise 9.1.6. Suppose that {an }∞ n=1 is a sequence of complex numbersP and f (u)
is a continuously differentiable function on the interval [1, x]. Let A(t) = n≤t an .
Prove that Z x
X
an f (n) = A(x)f (x) − A(t)f 0 (t)dt.
n≤x 1
√
Exercise 9.1.7. Let α > β > 0. Write z = x + iy, with x, y ∈ R, i = −1, and
x > 0. Show that
|z| −αx
e−αz − e−βz ≤ − e−βx .
e
x
Here is the main result regarding the region of convergence for a general
Dirichlet series.
converges for s = s0 , then it converges uniformly in any region of the form Re(s −
s0 ) ≥ 0, Arg(s − s0 ) ≤ α with α < π/2.
Proof. Without loss of any generality, we may assume s0 = 0 and that λn > 0
(since by our monotonicity condition for the λn ’s, only a finite number can
be negative and these terms converge for all values of s). We now apply
Abel’s lemma (Proposition 9.1.5). Since
∞
X
an
n=1
108 9 Dirichlet Series and Modular Forms
converges, we have that for any > 0, there is an N () such that for
X
A(r) = an ,
n≤r
we have
|A(r1 ) − A(r2 )| < for r1 , r2 > N ().
Applying Abel’s lemma with bn = e−λn s , we get
X X
an e−λn s = A(r)e−λr s −A(m−1)e−λm s − A(n) e−λn+1 s − e−λn s .
m≤n≤r m≤n<r
e−λm σ − e−λr σ
so that
X
−λn s |s|
an e ≤ 2 1 +
σ
m≤n≤r
|s|
≤ M,
σ
we have for m, r ≥ N () that
X
an e−λn s ≤ 2(M + 1).
m≤n≤r
In other words, in such a region, the series converges uniformly. But the re-
gion defined by the condition |s|/σ ≤ M is precisely the region cos arg(s) ≥
1/M . Since M can be arbitrarily large, this means arg(s) ≤ α with α < π/2.
t
u
Exercise 9.1.9. Suppose that
9.1 General Dirichlet series 109
X
A(x) = an = O(xδ )
n≤x
The following exercise gives the analytic continuation of ζ(s) to the region
Re(s) > 0.
Exercise 9.1.10. Show that
∞
{t}
Z
s
ζ(s) = −s dt,
s−1 1 ts+1
where {t} = t − btc and the integral converges absolutely for Re(s) > 0. Deduce
that lims→1+ (s − 1)ζ(s) = 1.
P∞ −λn s
Given a general Dirichlet series f (s) = n=1 an e , Theorem 9.1.8
shows that if it converges at s = s0 , then it converges for any s with
Re(s) > Re(s0 ), and the function thus defined is holomorphic in this region.
Thus, the set of values of s for which the Dirichlet series f (s) converges
contains a maximal open half-plane, called the half-plane of convergence. It is
convenient to view the empty set ∅ and the entire complex plane C as open
half-planes. If the half-plane of convergence is given by Re(s) > σ0 , we call
σ0 the abscissa of convergence. The half-plane of convergence of the Dirichlet
series
X∞
|an |e−λn s
n=1
σ 0 ≤ σ1 .
converges for any s with Re(s) > σ0 , we can apply Theorem 9.1.8 to deduce
that if f (s) converges at s0 , then f (s) → f (s0 ) when s → s0 in the region
Re(s) ≥ Re(s0 ) and arg(s − s0 ) ≤ α with α < π2 . This follows from the
uniform convergence and the fact that e−λn s tends to e−λn s0 as s → s0 .
Now
∞
X
f (k) (1) = (−1)k an λkn e−λn ,
n=1
converges. Thus, our Dirichlet series converges for s = − and hence for
Re(s) > −, by our earlier result. Hence the abscissa of convergence cannot
be zero.
Exercise 9.1.19. Let t0 ∈ R. With a = it0 , b = −it0 in the previous exercise, show
that
∞
X (σit0 (n))2 ζ(s)2 ζ(s − it0 )ζ(s + it0 )
= .
n=1
ns ζ(2s)
for some complex number w and k ∈ R. Then Lf (s) and Lg (s) extend analytically
to the entire complex plane apart from possible simple poles at s = 0 and k and
satisfy the functional equation
The right hand side is analytic for all s ∈ C apart from a simple pole at s = 0
with residue −a0 and s = k with residue −wb0 .
Changing f to g in the above identity shows that, upon using g(1/t) =
(1/w)tk f (t),
Z ∞ Z ∞
b0 (1/w)a0 dt dt
π −s Γ(s)Lg (s) = − − + (g(t)−b0 )ts + (f (t)−a0 )tk−s ,
s k−s 1 t 1 t
ˆ
In other words, F̂ (x) = F (−x).
Exercise 9.2.4. Show that Z ∞
2
e−πx dx = 1.
−∞
2 2
Exercise 9.2.5. For F (x) = e−πx , show that F̂ (u) = e−πu . In other words, F is
its own Fourier transform.
Exercise 9.2.6. Let F ∈ S and define G(t) = F (At + B) for real constants A, B,
with A 6= 0. Show that AĜ(t) = e2πiBt/A F̂ (t/A).
Theorem 9.2.7 (Poisson summation formula). Let F ∈ S. Then
X X
F (n) = F̂ (n).
n∈Z n∈Z
Proof. Let X
G(v) = F (n + v).
n∈Z
Since F̂ ∈ S, X
|F̂ (m)| < ∞
m∈Z
so X
G(v) = F̂ (m)e2πimv
m∈Z
by virtue of Exercise 9.2.10. We can then apply Theorem 9.2.1 with k = 1/2
to deduce that ζ(2s) extends analytically to the entire complex plane apart
from a possible simple pole at s = 0 and s = 1/2 and satisfies the functional
equation
1 1
π −s Γ(s)ζ(2s) = π −( 2 −s) Γ − s ζ(1 − 2s).
2
Changing s to s/2, we deduce the following.
Theorem 9.2.11 (Riemann, 1860). The function
Z ∞
dt
(ϑ(it) − 1) ts = 2π −s Γ(s)ζ(2s)
0 t
defined originally for Re(s) > 1/2, extends analytically to the entire complex plane
via the formula
Z ∞
−s 1 1 1 dt
2π Γ(s)ζ(2s) = − − 1 + (ϑ(it) − 1) (t2 + t 2 −s ) .
2 −s
s 1 t
and the right hand side is invariant under the map s → 1 − s. Moreover, the
expression is analytic for all s ∈ C except at s = 0 and s = 1 where it has a simple
pole.
Exercise 9.2.12. Show that (s − 1)ζ(s) extends to an entire function.
Exercise 9.2.13. Show that ζ(s) has simple zeros at s = −2n, for each positive
integer n.
9.2 The Poisson summation formula 117
Exercise 9.2.16. Let f ∈ S be an even function. That is, f (x) = f (−x). For any
x > 0, show that
∞ n ∞
X X 1 ˆ
f =x fˆ(nx) + xf (0) − f (0) .
n=1
x n=1
2
R∞
Exercise 9.2.17. Let f ∈ S. Set ζ(s, f ) = 0
f (x)xs dx
x . Show that the integral
converges for all s with Re(s) > 1.
ζ ? (s, f ) = ζ ? (1 − s, fˆ),
It is clear from the previous discussion that the essential tool in deriv-
ing the analytic continuation and functional equation for the Riemann zeta
function is the transformation law of Exercise 9.2.10. The function
X 2
Θ(z) = e2πin z
n∈Z
c
Exercise 9.2.19. Show that the symbol d is completely multiplicative in each of
the variables.
and c √
j(γ, z) := ε−1
d cz + d.
d
Then
Θ(γz) = j(γ, z)Θ(z) ∀γ ∈ Γ0 (4).
We will not prove this fact here, but only present it to indicate to the student
that our study has a natural entry into the theory of modular forms of half-
integral weight.
However, Riemann’s method of deriving the analytic continuation and
functional equation for ζ(s) extends to similar results for the Dirichlet L-
functions L(s, χ) defined as follows. Let χ : (Z/N Z)? → C be a Dirichlet
character. Then
∞
X χ(n)
L(s, χ) = .
n=1
ns
Exercise 9.2.20. Let d|N and let ψ be a character modulo d. Prove that
9.2 The Poisson summation formula 119
Y ψ(p)
L(s, χ0 ψ) = L(s, ψ) 1− .
ps
p|N
The previous exercise shows that to derive the analytic continuation and
functional equation for L(s, χ), we need only focus our attention on the
primitive characters χ (mod N ). The following exercises reduce the prob-
lem to a simple application of the Poisson summation formula. To this end,
we introduce for every character χ (mod N ), the Gauss sum τ (χ), defined as
N
X m
τ (χ) = χ(m)e , where e(t) = e2πit .
m=1
N
That is, the formula in the previous exercise holds without the condition (n, N ) = 1.
Proof. By virtue of Exercise 9.2.21, we need to show the identity also holds
for (n, N ) > 1. In this case, the left hand side is zero and so it suffices to
show the right hand size is zero. Write n/N = n1 /N1 with (n, N1 ) = 1. Thus
N1 |N and N1 < N . If N1 = 1, then n is a multiple of N so that e(mn/N ) = 1
and the right hand side is zero since
N
X
χ(m) = 0
m=1
and it suffices to prove that the inner sum is zero. Let us write
X
S(b) = χ(aN1 + b).
0≤a<N2
= S(b).
and as a ranges over 0 ≤ a < N2 , we see that ca ranges over the same set
(mod N2 ). Thus,
χ(c)S(b) = S(bc).
Since c ≡ 1 (mod N1 ), we write c = qN1 + 1 and so S(bc) = S(bqN1 + b) =
S(b) as S(b) is periodic with period N1 . As χ is a primitive character modulo
N , it is not periodic to any modulus N1 with N1 a proper factor of N . Thus,
there are integers c1 , c2 such that (c1 , N ) = (c2 , N ) = 1, c1 ≡ c2 (mod N1 )
and χ(c1 ) 6= χ(c2 ). From the preceding, upon setting c ≡ c1 c−1 2 (mod N ), we
have that c ≡ 1 (mod N1 ) so that χ(c)S(b) = S(b). But χ(c) = χ(c1 )χ(c2 ) 6= 1
and we conclude S(b) = 0 as desired. t
u
The previous
√ exercise shows that if χ is a primitive character modulo N
then |τ (χ)| = N .
Now suppose χ is an even character (that is, χ(−1) = 1). We have
Z ∞
dx
Γ(s) = e−x xs
0 x
∞
!
s Z ∞
−s/2
X 2 dx
π N s/2
Γ L(s, χ) = χ(n)e−n πx/N
xs/2
2 0 n=1
x
for Re(s) > 1. Since χ(−1) = 1 and χ(0) = 0, we can rewrite the sum in the
integrand as
∞
X 2
θ(x, χ) = χ(n)e−n πx/N .
n=−∞
By Theorem 9.2.22,
N
X mn
χ(n)τ (χ) = χ(m)e
m=1
N
so that
N ∞
X X 2
τ (χ)θ(x, χ) = χ(m) e−n πx/N +2πimn/N
.
m=1 n=−∞
so that
∞
N
1/2 X !
N X
−(nN +m)2 π/N x
τ (χ)θ(x, χ) = χ(m) e
x m=1 n=−∞
1/2 X
N 2
= χ(r)e−πr /N x
x
r∈Z
1/2
N 1
= θ ,χ . t
u
x x
Both integrals converge for all s ∈ C and thus we obtain analytic continuation of
L(s, χ) to the entire complex plane. Moreover,
−s/2 s/2
s
−(1−s)/2 (1−s)/2 1−s
π N Γ L(s, χ) = wχ π N Γ L(1 − s, χ)
2 2
Proof. Since
s Z ∞
−s/2 s/2 dx
2π N Γ L(s, χ) = θ(x, χ)xs/2
2 x
Z0 ∞ Z ∞
dx 1 dx
= θ(x, χ)xs/2 + θ , χ x−s/2
1 x 1 x x
Therefore
∞ ∞
N 1/2
s Z Z
dx 1−s dx
2π −s/2 N s/2 Γ L(s, χ) = θ(x, χ)xs/2 + θ(x, χ)x 2 .
2 1 x 1 τ (χ) x
t
u
Show that
Z ∞
s+1 s+1 s+1 s+1 dx
2π −( 2 )Γ N 2 L(s, χ) = ϑ1 (x, χ)x 2 .
2 0 x
9.3 L-functions attached to modular forms 123
Then Lf (s) and Lg (s) extend to entire functions and satisfy the functional equation
Λf (s) = ik Λg (k − s),
where
√ !s
N
Λf (s) = Γ(s)Lf (s)
2π
√ !s
N
Λg (s) = Γ(s)Lg (s).
2π
Proof. We have Z ∞
(2π)−s Γ(s)Lg (s) = g(it)ts−1 dt.
0
The assumption that f and g are cusp forms is not essential in the previ-
ous theorem, as shown in the following exercise.
0 −1
Exercise 9.3.2. Let f ∈ Mk (Γ0 (N ), χ) and g = f |w with w = . Sup-
N 0
posing that
9.3 L-functions attached to modular forms 125
∞
X
f (z) = an e2πinz
n=0
X∞
g(z) = bn e2πinz
n=0
a0 b0 i k
Λf (s) + +
s k−s
and
b0 a0 ik
Λg (s) + +
s k−s
are entire and satisfy the functional equation
Λf (s) = ik Λg (k − s).
Exercise 9.3.3. With f , g as in the previous exercise, show that Lf (0) = −a0 , Lg (0) =
−b0 ,
k
2πi
lim (s − k)Lf (s) = b0 √ /Γ(k)
s→k N
and k
2πi
lim (s − k)Lg (s) = a0 √ /Γ(k).
s→k N
Theorem 9.3.1 and Exercise 9.3.2 have a converse. The reader will note
that the integral expressions for
a0 b0 i k
Λf (s) + +
s k−s
and
b0 a0 ik
Λg (s) + +
s k−s
are bounded in any vertical strip of the form σ1 ≤ Re(s) ≤ σ2 .
126 9 Dirichlet Series and Modular Forms
A simple integration by parts shows that Γ(s + 1) = sΓ(s) and this allows
us to extend Γ(s) analytically to the entire complex plane with only simple
poles at s = 0, −1, −2, . . .
Stirling’s formula for the asymptotic behavior of Γ(s) as |s| → ∞ implies
that if σ is fixed, and |t| → ∞, then
√ π
|Γ(σ + it)| ∼ 2πe− 2 |t| |t|σ−1/2
(see, for example, Exercise 6.3.15 of [24]). This asymptotic formula allows
us to deduce that if Λf (s) and Λg (s) are bounded in every vertical strip,
then interval σ1 ≤ Re(s) ≤ σ2 , the series Lf (s) and Lg (s) are
for any fixed
1
O |t| 2 −σ eπ|t|/2 as |t| → ∞. A second result from complex analysis, namely
the Phragmén-Lindelöf theorem, allows us to sharpen this estimate. We state
this without proof but refer the reader to Theorem 8.1.13 of [24].
Theorem 9.3.4 (Phragmén-Lindelöf). Suppose that f (s) is analytic in the ver-
tical strip a ≤ Re(s) ≤ b, and that for some α ≥ 1,
α
|f (s)| = O e|t|
as |t| → ∞. If f (s) is bounded on the two vertical lines Re(s) = a and Re(s) = b,
then f (s) is bounded in the entire vertical strip.
We apply this theorem in our context via the following exercise.
Exercise 9.3.5. Suppose that f (s) is analytic in the vertical strip defined by a ≤
Re(s) ≤ b and that for some α ≥ 1,
α
f (s) = O e|t|
as |t| → ∞. If for some A> 0, f (s) = O |t|A on Re(s) = a and Re(s) = b, then
show that f (s) = O |t|A for all s in the vertical strip a ≤ Re(s) ≤ b.
In our context, both Lf (s) and Lg (s) are bounded in some fixed half-
plane. Moreover s(s − k)Lf (s) and s(s − k)Lg (s) are entire. By the functional
equation relating Λf (s) to Λg (k − s) and Stirling’s formula, we see that for
some A > 0,
Exercise 9.3.7. With f as in Exercise 9.3.2, show that for σ > 0 we have
Z σ+i∞
iy 1
f √ − a0 = Λf (s)y −s ds.
N 2πi σ−i∞
Λf (s) = ik Λg (k − s),
we have
Z −U +i∞
iy 1
f √ = ik Λg (k − s)y −s ds − ik b0 y −k .
N 2πi −U −i∞
which now holds for all z ∈ H by the principle of analytic continuation. Thus
f |wN = g. This proves the converse of Theorem 9.3.1 and Exercise 9.3.2.
Our discussion can be summarized by the following theorem, due essen-
tially to E. Hecke.
Theorem 9.3.8 (Hecke). Suppose f and g are given by the Fourier series
∞
X
f (z) = an e2πinz
n=0
X∞
g(z) = bn e2πinz
n=0
P∞ an
with an , bn bounded by O(nα ) for some α ≥ 0. Put Lf (s) = n=1 ns , Lg (s) =
P ∞ bn
n=1 ns and
√ !s
N
Λf (s) = Γ(s)Lf (s)
2π
√ !s
N
Λg (s) = Γ(s)Lg (s)
2π
and √ −k 1
f |w = Nz f −
Nz
where k is a fixed positive integer. Then the following are equivalent:
(a) g = f |w
(b) Λf (s) and Λg (s) extend to the entire complex plane and both
a0 b0 i k
Λf (s) + +
s k−s
and
b0 a0 i−k
Λg (s) + +
s k−s
are entire and bounded in every vertical strip and satisfy the functional equation
Λf (s) = ik Λg (k − s).
9.4 Twists of L-series attached to modular forms 129
where ad − bc = 1.
We use this to prove:
Theorem 9.4.2. Let f ∈ Mk (Γ0 (q), χ) where χ is a Dirichlet character of conduc-
tor q ? with q ? |q. Let ψ be a primitive Dirichlet character modulo r. If
∞
X
f (z) = an e2πinz ,
n=0
By Exercise 9.4.1,
Thus,
1 d2 u/r
X
τ (ψ)fψ | γ = ψ(u)χ(α)f
0 1
u (mod r)
Exercise 9.4.3. Let (q, r) = 1 and (u, r) = 1. Let d and w be integers such that
dr − quw = 1. Verify the matrix identity with N = qr2 ,
1 u/r 0 −1 0 −1 r −w 1 w/r
=r .
0 1 N 0 q 0 −qu d 0 1
fψ |wN = w(ψ)gψ
9.4 Twists of L-series attached to modular forms 131
where
w(ψ) = χ(r)ψ(q)τ (ψ)2 /r.
Given f ∈ Sk (Γ0 (q), χ) and ψ a primitive Dirichlet character modulo r,
we associate the “twisted” L-series:
∞
X an ψ(n)
Lf (s, ψ) =
n=1
ns
Λf (s, ψ) = ik w(ψ)Λg (k − s, ψ)
where
w(ψ) = χ(r)ψ(q)τ (ψ)2 /r
and g = f |wq , with wq = 0q −1
0 .
we have g = f N0 −1
0 . In the
special case N = 1, the functional equation
implies that g(z) = f 01 −1 0 . In particular if f = g, we see that the func-
0 −1
tional equation implies f 1 0 = f . Since f |( 10 11 ) = f, we deduce that
f |γ = f for all γ in the group generated by ( 10 11 ) and 01 −10 which is the full
modular group SL2 (Z) by Theorem 2.1.2. If N > 1, one functional equation
is not sufficient. However, in a celebrated paper A. Weil [46] showed the fol-
lowing holds, which we state without proof. An elegant and simple proof of
this theorem in the English language has been found by Razar [32] and we
invite the student to study it.
Theorem 9.4.6 (Weil). Let k be a positive even integer and χ a character modulo
q. Suppose f and g are given by the Fourier series
∞
X
f (z) = an e2πinz
n=0
X∞
g(z) = bn e2πinz ,
n=0
Suppose that Λf (s) and Λg (s) have meromorphic continuations for all s ∈ C and
a0 b0 i k
Λf (s) + + ,
s k−s
b0 a 0 ik
Λg (s) + +
s k−s
are entire and bounded in vertical strips and satisfy Λf (s) = ik Λg (k − s). Further,
let R be a set of prime numbers coprime with q which meets every coprime residue
class modulo c for any c ≥ 1. For each r ∈ R and every primitive character ψ
modulo r, suppose that Λf (s, ψ) and Λg (s, ψ) are entire and bounded in vertical
strips and satisfy the functional equation given as in Theorem 9.4.5 with w(ψ)
as specified there. Then f ∈ Mk (Γ0 (q), χ), g ∈ Mk (Γ0 (q), χ) and g = f |wq .
Moreover, if Lf (s) and Lg (s) converge absolutely on some line Re(s) = σ with
0 < σ < k, then f and g are cusp forms.
9.5 Supplementary problems 133
Exercise 9.5.2. For Re(c) > 0, let F (x) = e−c|x| . Show that
2c
F̂ (u) = .
c2 + 4π 2 u2
Exercise 9.5.3. Show that for Re(c) > 0,
∞
ec + 1 X 2c
= .
ec − 1 n=−∞ c2 + 4π 2 n2
Exercise 9.5.4. From the previous exercise, deduce Euler’s formula that
∞
X 1 π2
2
= .
n=1
n 6
Let ω1 and ω2 be two non-zero complex numbers which are linearly inde-
pendent over R. They generate a lattice L which is the additive group con-
sisting of elements of the form
mω1 + nω2 , m, n ∈ Z.
f (z + ω) = f (z) ∀z ∈ C, ∀ω ∈ L.
Sometimes, such functions are also called doubly periodic in the literature.
We can identify the space of such functions with meromorphic functions
on the complex torus C/L. The values of any such function are completely
determined by its values on the fundamental parallelogram which we take to
be the (closed) region in the complex plane enclosed by the parallelogram P
with vertices 0, ω1 , ω2 , and ω1 + ω2 :
P = {rω1 + sω2 : 0 ≤ r, s ≤ 1} .
It will be convenient to let P denote the interior of P along with the two
sides incident to the origin, including the origin itself:
© Springer Science+Business Media Singapore 2016 and Hindustan Book Agency 2016 135
M.R. Murty et al., Problems in the Theory of Modular Forms, HBA Lecture Notes
in Mathematics, DOI 10.1007/978-981-10-2651-5_10
136 10 Special Topics
Exercise 10.1.2. If two elliptic functions (for a fixed lattice L) have the same poles
with the same principal parts, show that they differ by a constant.
Exercise 10.1.3. If f is an elliptic function with no poles on the boundary of the
fundamental parallelogram P , show that the sum of the residues of f in P is zero.
Exercise 10.1.4. Let f be an elliptic function with respect to a lattice L. Show that
there is an α ∈ C such that f is holomorphic on the boundary of the translate α+P ,
where P is the fundamental parallelogram of L.
Exercise 10.1.5. Show that there is no non-constant elliptic function having ex-
actly one simple pole in the fundamental parallelogram P .
Exercise 10.1.6. Let f be an elliptic function. Show that the number of zeros of f
in the fundamental parallelogram P is equal to the number of poles of f (counted
with multiplicity).
The order of an elliptic function f is defined as the number of poles of f
(counted with multiplicity) in the fundamental parallelogram. By the previ-
ous exercise, the order of f is also equal to the number of zeros of f (counted
with multiplicity) in P .
Exercise 10.1.7. If f is an elliptic function of order m, then for any c ∈ C, show
that the number of solutions of f (z) = c with z ∈ P is exactly m (counted with
multiplicities).
The simplest non-constant elliptic function with respect to the lattice L is
the celebrated Weierstrass ℘ function defined by
1 X 1 1
℘(z) := 2 + − 2 .
z (z − ω)2 ω
06=ω∈L
P
One might naively hope for a simpler elliptic function like ω∈L 1/(z − ω),
but this does not converge. Moreover, Exercise 10.1.5 implies that 2 is the
smallest possible order for a non-constant elliptic function. We now address
the question of convergence of the series defining ℘(z).
Exercise 10.1.8. For any s > 2, show that the series
X 1
ωs
ω∈L
ω6=0
is absolutely convergent.
Exercise 10.1.9. Show that the sum
X 1 1
− 2
(z − ω)2 ω
06=ω∈L
By the solution to Exercise 10.1.9, the infinite sum defining ℘(z) converges
uniformly on compact subsets of C\L. Thus the sum defines a meromorphic
function on C. Now X 1
℘0 (z) = −2 3
ω∈L
(z − ω)
and it is clear that ℘0 (z) is a doubly periodic function, with periods ω1 and
ω2 . Hence for any ω ∈ L,
℘0 (z + ω) = ℘0 (z) ∀z ∈ C \ L.
Hence
℘ (z + ω) − ℘(z)
is a constant function of z. But for z = −ω/2, this term is ℘(ω/2) − ℘(−ω/2)
and since ℘(z) is an even function (an easy consequence of Exercise 10.1.9),
this term is zero. Hence we have proved:
Theorem 10.1.10. The Weierstrass ℘-function is an elliptic function of the lattice
L. Moreover, it is an even function of z. The derivative ℘0 (z) is an odd elliptic
function. Moreover, ℘(z) is holomorphic for all z ∈ P with z 6= 0. Its only poles are
at elements of L and these are all double order poles.
We now show that the field of elliptic functions on L is generated by ℘(z)
and ℘0 (z).
Theorem 10.1.11. Any even elliptic function on L is a rational function of ℘(z).
Any elliptic function f on L can be written as
f (z) = g ℘(z) + ℘0 (z)h ℘(z)
f (z) + f (−z)
fe (z) = ,
2
which is even, and
f (z) − f (−z)
fo (z) = ,
2
which is odd. Now
0 fo (z)
fo (z) = ℘ (z)
℘0 (z)
and f0 /℘0 is an even elliptic function.
Now suppose f is even. Then ordz0 (f ) = ord−z0 (f ) for any z0 ∈ C. We
now show that if 2z0 ∈ L, then ordz0 (f ) is even. To see this we differentiate
the equation f (z) = f (−z) to get
138 10 Special Topics
i
f (i) (z) = (−1) f (i) (−z),
so that if 2z0 ∈ L, then f (i) (z0 ) = f (i) (z0 − 2z0 ) = f (i) (−z0 ) and this implies
f (i) (z0 ) = 0 for all odd i. In other words, if 2z0 ∈ L, then ordz0 (f ) is even.
Now let Pe = P/ ± 1 be “half” of the fundamental parallelogram. Con-
sider the function Y nv
g(z) = ℘(z) − ℘(v)
v∈Pe
v6=0
where (
ordv (f ) if 2v ∈
6 L
nv =
ordv (f )/2 if 2v ∈ L.
For 2v 6∈ L, the term ℘(z) − ℘(v) has zeros at the distinct points z = ±v, and
a pole of order 2 at z = 0. Hence the zeros are simple and so
nv
ordv ℘(z) − ℘(v) = ordv (f ).
If v ∈ 2L, then ℘(z) − ℘(v) has a double zero at the point z = v = −v. In this
way, we see that f (z) and g(z) have exactly the same zeros and poles (with
the same orders) except possibly at z = 0. But by Exercise 10.1.6, they have
the same order at z = 0 also. Therefore f (z)/g(z) is a holomorphic elliptic
function which, by Exercise 10.1.1, is a constant. t
u
Prove that
∞
X
℘(z) = z −2 + (k + 1)Gk+2 z k
k=2
k even
Exercise 10.1.13. With notation as in the previous exercise, show that for z ∈ C\L,
where
g2 = 60G4 , g3 = 140G6
Curves of the form (10.1) are called elliptic curves. The cubic polynomial on
the right hand side has discriminant given by
∆ = g23 − 27g32
Exercise 10.1.14. Let L be a lattice and g2 , g3 defined as above. Show that all of the
complex solutions of the equation
y 2 = 4x3 − g2 x − g3
are given by ℘(z), ℘0 (z) where ℘(z) is the Weierstrass ℘-function attached to L
and z ranges over all complex numbers in C \ L.
The Weierstrass ℘-function satisfies a remarkable addition formula which
we record in the following theorem.
Theorem 10.1.15. Suppose z, w, z + w ∈ C \ L. If z 6= w, we have
2
℘0 (z) − ℘0 (w)
1
℘ (z + w) = −℘(z) − ℘(w) + .
4 ℘(z) − ℘(w)
When z = w, we have
2
℘00 (z)
1
℘ (2z) = −2℘(z) + .
4 ℘0 (z)
Proof. Let (x1 , y1 ) = (℘(z), ℘0 (z)) and (x2 , y2 ) = (℘(w), ℘0 (w)) be the corre-
sponding points on the elliptic curve y 2 = 4x3 − g2 x − g3 . Let y = ax + b be
the line through these two points. Then
℘0 (z) = a℘(z) + b
℘0 (w) = a℘(w) + b.
℘0 (u) − a℘(u) − b
has poles in the fundamental parallelogram P only at u = 0 and this pole has
order 3. By Exercise 10.1.6, it has three zeroes in P (counting multiplicities).
Working modulo L, we already have two of these zeros at u = z and u = w.
Let u = t be the third zero. Then
℘0 (t) = a℘(t) + b.
has roots x = ℘(z), ℘(w), and ℘(t). Since the sum of the roots of the cubic is
a2 /4, we get
a2
℘(z) + ℘(w) + ℘(t) = .
4
Since a is the slope of the line joining (x1 , y1 ) and (x2 , y2 ), we get
2
℘0 (z) − ℘0 (w)
1
℘(z) + ℘(w) + ℘(t) = .
4 ℘(z) − ℘(w)
Exercise 10.1.17. Suppose that ω1 , ω2 are linearly independent over R and that L
is the lattice spanned by ω1 , ω2 . Let ℘(z) be the associated ℘-function. Show that
℘ ω21 , ℘ ω22 , and ℘ ω1 +ω
2
2
are all distinct.
Proof. This statement is equivalent to showing that the roots of the associ-
ated cubic 4x3 − g2 (z)x − g3 (z) are all distinct. But this follows from Exer-
cise 10.1.17. t
u
It is clear that ∆0 (z) is a modular form of weight 12 for the full modular
group. It is related to the ∆(z) introduced in Chapter 4. Indeed, it is easily
verified that
12
∆0 (z) = (2π) ∆(z)
using our explicit knowledge of ζ(2k) given in Exercise 4.2.4. As before, we
let
E 3 (z)
j(z) = 4
∆(z)
and it is immediately checked that in our notation above,
g2 (z)3
j(z) = 1728 (10.2)
g2 (z)3 − 27g3 (z)2
However, now that we have related the j-function to elliptic curves, we will
elaborate on this relationship.
Exercise 10.2.2. If ω1 , ω2 are linearly independent over R and ac db ∈ SL2 (Z),
show that the lattice generated by aω1 + bω2 and cω1 +dω2 is the same as the lattice
generated by ω1 and ω2 .
0 0
L and ω1 , ω2 also generate the same
Exercise 10.2.3. If ω1 , ω2 generate the lattice
lattice L, show that there is a matrix ac db ∈ SL2 (Z) such that ω10 = aω1 + bω2
and ω20 = cω1 + dω2 .
y 2 = 4x3 − g2 x − g3 ,
we define
1728g23
j(E) = .
g23 − 27g32
One can show that j(E) is isomorphism invariant. In fact E1 ∼
= E2 over C if
and only if j(E1 ) = j(E2 ).
In Theorem 4.5.1 we proved that the map z 7→ j(z) is a bijection from
the standard fundamental domain F onto the complex numbers C. This fact
allows us to prove the following uniformization theorem.
Theorem 10.2.5. Given any two A, B ∈ C with A3 − 27B 2 6= 0, there exists a
lattice L such that its associated Weierstrass function satisfies
Proof. We consider two cases. The first case is when A = 0. The second case
is when A 6= 0.
In the first case (A, B) = (0, B) with B 6= 0. We choose z0 ∈ H so that
g2 (z0 ) = 0. This can be done since g2 (z) is, up to a scalar multiple, E4 (z)
and by Exercise 4.3.5 z = ρ works. Since ∆0 (z0 ) 6= 0, we have g3 (z0 ) 6= 0.
Choosing λ so that λ−6 g3 (z0 ) = B, the lattice L = [λ, λz0 ] has the required
property, by Exercise 10.2.4.
Now consider the case when A 6= 0. Let a = B 2 /A3 . Since A3 − 27B 2 6= 0,
we have a 6= 1/27. By Theorem 4.5.1, we can choose z0 ∈ H so that j(z0 ) =
1728/(1 − 27a). If B = 0, then we choose λ so that λ−4 g2 (z0 ) = A and we see
that L = [λ, λz0 ] works. If B 6= 0, we choose λ so that
Ag3 (z0 )
λ2 =
Bg2 (z0 )
and consider the lattice L = [λ, λz0 ]. From (10.2) we deduce that a =
g3 (z0 )2 /g2 (z0 )3 . Thus for the lattice L, the associated values of g2 and g3
are A and B respectively. Indeed,
g2 (z0 ) B 2 g2 (z0 )3
g2 = 4
= 2 =A
λ A g3 (z0 )2
and
g3 (z0 ) B 3 g2 (z0 )3
g3 = = =B
λ6 A3 g3 (z0 )2
as desired. t
u
The previous theorem shows that all elliptic curves over C have associ-
ated with them a lattice L and a Weierstrass ℘-function such that (℘(z), ℘0 (z))
10.3 Theta series and lattices 143
E43 (z)
j(z) = .
∆(z)
(·, ·) : V × V → R
has measure 1.
A lattice L of V is a subset of V which can be written as
L = ω1 Z + · · · + ωn Z
Recall that the dual vector space V ? of V is the vector space of linear function-
als on V . Since V is an inner product space these functionals are of the form
(·, v0 ) for some v0 ∈ V (Riesz representation theorem). We define the dual
lattice L? of L as the collection of elements v ? ∈ V ? such that v ? (ω) ∈ Z for
all ω ∈ L.
L? = ω1? Z + · · · + ωn? Z
we see that
10.3 Theta series and lattices 145
n X
X n
(ω, ω) = (ωi , ωj )xi xj
i=1 j=1
This proves:
Proof. From the proof immediately above, we easily see that the series repre-
senting θL (z) converges uniformly and absolutely on every compact subset
of H (since for any compact set in H, v is bounded from below.) Thus θL (z)
is holomorphic there. t
u
The matrix A = (ωi , ωj ) is called the Gram matrix associated to the
lattice L.
√
Exercise 10.3.3. Prove that µ(V /L) = det A.
Exercise 10.3.4. With A as above, show that the matrix A−1 is the Gram matrix of
the dual lattice L? .
To derive the modular transformation law for θL (z), we need the follow-
ing version of Poisson’s summation formula for lattices.
Let V be as above a vector space over R of dimension n. We view V as
being isomorphic to Rn with the usual Euclidean topology. A bounded func-
tion f : V → R is called rapidly decreasing if it can be partially differentiated
arbitrarily often and each of these partial derivatives are bounded.
The Fourier transform of f is a map fˆ : V ? → R defined by
Z
fˆ(x) = f (u)e−2πi(u,x) dµ(x)
V
146 10 Special Topics
where (u, x) denotes x(u). Once can show that fˆ is again a rapidly decreas-
ing function on V ? .
Theorem 10.3.5. For f rapidly decreasing,
X 1 X
f (ω) = fˆ(ω ? ).
µ(V /L) ? ?
ω∈L ω ∈L
for all z ∈ H. (Here, the square root is chosen so that it is positive for any positive
real number.)
Proof. By complex analysis, it suffices to prove the identity for purely imag-
inary z. In other words, it suffices to show that
X ? ? X
e−π(ω ,ω )/y = y n/2 µ (V /L) e−πy(ω,ω) .
ω ? ∈L? ω∈L
and a simple change of variables in the summation now gives the desired
result. t
u
A lattice L is said to be self-dual if L? = L. It is said to be even if ||ω||2 ∈ 2Z
for all ω ∈ L.
10.4 Special values of zeta and L-functions 147
Exercise 10.3.8. If L is self-dual show that the Gram matrix of L and L? have
determinant 1.
Exercise 10.3.9. Show that the theta function of a self-dual, even lattice is a modu-
lar form of weight n/2 for the full modular group SL2 (Z).
dk k!
− k
(π cot πz) = ψk (z) + (−1)k+1 ψk (−z) + (−1)k k+1 .
dz z
Exercise 10.4.3. Show that
dk
− (π cot πz) = ψk (z) + (−1)k+1 ψk (1 − z).
dz k
Let N be a natural number and let χ : (Z/N Z)∗ → C be a Dirichlet char-
acter and define the L-series L(s, χ) by
∞
X χ(n)
L(s, χ) := .
n=1
ns
N
X
(k − 1)!L(k, χ) = (−N )−k χ(a)ψk−1 (a/N ).
a=1
Exercise 10.4.5. If χ is an even character (that is, χ(−1) = 1) and k is even, show
that
N
X dk
−2(k − 1)!L(k, χ) = (−N )−k χ(a) k (π cot πz) .
a=1
dz z=a/N
Exercise 10.4.6. If χ is an odd character (that is, χ(−1) = −1) and k is odd, show
that
N
X dk
−2(k − 1)!L(k, χ) = (−N )−k χ(a) k (π cot πz) .
a=1
dz z=a/N
The gist of the preceding exercises is that the value of L(k, χ) when k
and χ have the same parity, is a non-zero algebraic multiple of π k . This fact
was first proved by Hecke [14] and he noted that this implies an interesting
result for real quadratic fields. Namely, if F is a real quadratic field, then
ζF (2m) is an algebraic multiple of π 4m . This motivated him to ask if such
a result holds generally for any totally real field F . That is, if F is totally
real of degree r over the rationals, then is it true that ζF (2m) is an algebraic
multiple of π 2rm ? This was anwered in the affirmative by Siegel and Klingen
and is now known as the Siegel-Klingen theorem.
The proof of the Siegel-Klingen theorem makes use of the theory of classi-
cal modular forms and Hilbert modular forms. Since a detailed explanation
of the proof is beyond the scope of this book, we content ourselves with
a brief outline which should enable the interested student to make a more
detailed study.
We summarise the contents of the appendix in [39]. For the relevant back-
ground in algebraic number theory, we refer the student to [25]. For a de-
tailed study of Hilbert modular forms, we suggest the book by Garrett [10].
Let F be a totally real number field of degree r and discriminant D. Let
x 7→ x(i) be an indexing of real embeddings of F and let OF be the ring of
integers of OF . The group SL2 (OF ) is called the Hilbert modular group and
it acts on Hr via the map:
Let k be a natural number (which is even if K has a unit of norm −1). For an
ideal a, we define the analogue of the Eisenstein series as
X
Fk (a, z) = N (a)k N (λz + µ)−k ,
(λ,µ)6=(0,0)
where the sum runs over a complete system of pairs of numbers in the ideal
a different from (0, 0) and not differing from one another by a factor which
is a unit (that is, not associated). One can show that the series converges
absolutely for k > 2 and for k = 2, one can apply a limit process (Hecke’s
trick). The function Fk (a, z) is an example of a Hilbert modular form in the
sense that
As OF is a lattice and its dual is the inverse different d−1 of OF (see [25]), one
has a Fourier expansion of Fk (a, z) of the form
r
(2πi)k
X
ζ(a, k) + D1/2−k σk−1 (a, ν)e2πiT r(zν) ,
(k − 1)!
ν∈d−1 ,ν0
X
σk−1 (a, ν) = sign(N (α)k )N ((α)ad)k−1 .
d−1 |(α)a|ν
The summation is over principal ideals (µ), (α) under the conditions given.
If we set all the variables z1 , ..., zr to z, then Fk (a, z) becomes a classical
modular form of weight rk. By Exercises 5.5.18, 5.5.19, and 5.5.20, we deduce
that ζ(a, k) is a an algebraic multiple of π kr . Since the Dedekind zeta function
ζF (k) can be written as a rational linear combination of the values ζ(a, k), the
Siegel-Klingen theorem follows from this.
Exercise 10.5.1. Let f be an elliptic function with period lattice L. Show that
X
ordz (f )z ∈ L
z∈C/L
150 10 Special Topics
Exercise 10.5.4. Show that for k ≥ 2, each G2k can be expressed as a polynomial
in G4 and G6 with positive rational coefficients.
Show that
∞
Y
ϑ(z) = (1 − e2πinz )(1 + e(2n−1)πiz )2 .
n=1
Exercise 10.5.6. With η being the Dedekind η-function and ϑ as in the previous
exercise, show that
ϑ(z) = η 2 ((z + 1)/2)/η(z + 1).
Exercise 10.5.7. Show that if j is the modular invariant, then j 0 (z) satisfies
az + b ab
j0 = (cz + d)2 j 0 (z) ∀ ∈ SL2 (Z)
cz + d cd
but that j 0 (z) is not a modular form of weight 2 for the full modular group.
Part II
Solutions
Chapter 1
Jacobi’s q-series
Iterating this relation and using that Eq ( qxn ) → 1 as n → ∞ gives the result.
t
u
© Springer Science+Business Media Singapore 2016 and Hindustan Book Agency 2016 153
M.R. Murty et al., Problems in the Theory of Modular Forms, HBA Lecture Notes
in Mathematics, DOI 10.1007/978-981-10-2651-5_11
154 1 Jacobi’s q-series
1
In Exercise 1.1.2, we replace q by q to obtain
∞
Y
E q1 (x) = (1 + q n x).
n=1
n(n+1)
on noting that q 1+2+···+n = q 2 . Thus
∞ ∞ n(n+1)
Y
n
X q 2 xn (1 + x)
(1 + q x) =
n=0 n=0
(1 − q) · · · (1 − q n )
∞
X xn n(n+1) n(n−1)
n
=1+ q 2 + q 2 (1 − q )
n=1
(1 − q) · · · (1 − q n )
∞ n
X q ( 2 ) xn
=1+ ,
n=1
(1 − q) · · · (1 − q n )
as required. t
u
1.1.4. Prove that if |q| > 1 then
1
(1 + x)E q1 (x) = .
Eq (x)
Therefore
1 x
E (x) =
1 E1 .
q (1 + x) q q
1.2 Jacobi’s Triple Product Identity 155
Iterating this gives the first result since as n → ∞, we have x/q n → 0, and
hence E1/q ( qxn ) → 1. In particular
∞ −1
Y x
E q1 (x) = 1+ n . (1.1)
n=0
q
(1 + x) E q1 (x) = Eq (x)−1 .
1.2.2 (Euler’s pentagonal number theorem). Show that if |q| < 1 then
∞ ∞
Y X k(3k−1)
n
(1 − q ) = (−1)k q 2 .
n=1 k=−∞
3 1
we replace q by q 2 and x by −q − 2 to get
∞ ∞
Y X (3k−1)k
(1 − q 3n )(1 − q 3n−2 )(1 − q 3n−1 ) = (−1)k q 2 .
n=1 k=−∞
Q∞
But the left hand side is simply n=1 (1 − q n ). t
u
We compare coefficients with the right hand side. When k = 0, the sum-
2
mand is 1. When k = 2n, the summand is (4n + 1)q 2n +n . When k = 2n − 1,
2
the summand is −(4n − 1)q 2n −n . t
u
Prove that x4 (n) is completely multiplicative. That is, show that x4 (mn) =
x4 (m)x4 (n).
By Corollary 1.3.2, we must show that the sum is equal to d1 (n) − d3 (n),
where di (n) is the number of divisors of n congruent to i (mod 4). By Exer-
cise 1.3.3, the sum is precisely equal to the desired quantity. t
u
Part (b) follows easily from two applications of the usual formula
and as
θ θ θ
cos θ = cos2 − sin2 = 2 cos2 − 1,
2 2 2
we see that
1 θ 1 + cos θ
cot sin θ = .
2 2 2
Now suppose the result has been proved for n. Observe that
1 θ 1 1
cot sin nθ + cos nθ + cos(n + 1)θ
2 2 2 2
cos θ2 sin nθ + cos nθ sin θ2 1
= θ
+ cos(n + 1)θ
2 sin 2 2
1
sin n + 2 θ 1
= + cos(n + 1)θ.
2 sin θ2 2
qr
= ur (1 + ur ).
(1 − q r )2
Prove that
∞
X ∞
X
um (1 + um ) = nun .
m=1 n=1
Since
∞
1 X
= qn ,
1 − q n=0
Hence,
∞ ∞ X ∞ ∞ ∞ ∞
X qm X
mn
X X
mn
X nq n
m 2
= nq = n q = ,
m=1
(1 − q ) m=1 n=1 n=1 m=1 n=1
1 − qn
as required. t
u
as required. t
u
Show that τ (n) is odd if and only if n = (2m + 1)2 for some m.
We use the fact that for any prime p and for 1 ≤ j ≤ pα − 1, the binomial
coefficient α
p
≡ 0 (mod p).
j
Thus (1 − q n )8 ≡ 1 − q 8n (mod 2) and hence
∞
Y ∞
Y
q (1 − q n )24 ≡ q (1 − q 8n )3 (mod 2).
n=1 n=1
1.5.2. Let rk (n) be the number of ways of writing n as a sum of k squares. Show
that
n
X
rk (n) = ri (a)rk−i (n − a)
a=0
∞
!k ∞
!i ∞
!k−i
X 2 X 2 X 2
qn = qn qn
k=−∞ k=−∞ k=−∞
Deduce that
160 1 Jacobi’s q-series
∞
X x
Lq (x) = n−x
.
n=1
q
We have n
X ∞ xn − x
x q
Lq (x) − Lq = n−1
q n=1
q
∞ n
X x
=
n=1
q
x
= .
q−x
Since |q| > 1, as n → ∞ we have x/q n → 0 and thus Lq (x/q n ) → 0. There-
fore, iteration leads to
x
x q x
Lq (x) = + + Lq
q − x q − xq q2
= ···
∞
X x
= n−x
. t
u
n=1
q
1.5.4. Show that if |x| < |q| and |q| > 1 then
xEq0 (−x)
Lq (x) = ,
Eq (−x)
we take the logarithmic derivative and use Exercise 1.5.3 to deduce the de-
sired result. t
u
Chapter 2
The Modular Group
2.1.1. Let R be a commutative ring with identity. Show that the set
ab a, b, c, d ∈ R
SL2 (R) = :
cd ad − bc = 1
© Springer Science+Business Media Singapore 2016 and Hindustan Book Agency 2016 161
M.R. Murty et al., Problems in the Theory of Modular Forms, HBA Lecture Notes
in Mathematics, DOI 10.1007/978-981-10-2651-5_12
162 2 The Modular Group
φ : SL2 (Z) → C×
has image contained in the finite subgroup of C× consisiting of 12th roots of unity.
By Exercise 2.1.4, SL2 (Z) is generated by S of order four and ST of order
six. The function φ must send these elements to a fourth root of unity and
a sixth root of unity, respectively.
√
Thus the image is contained in the group
1+ −3
generated by i and 2 , which is the group of 12th roots of unity. t
u
2.1.7. Suppose that (c, d, N ) = 1. Show that there are elements c0 = c + tN and
d0 = d + sN for some integers s, t such that (c0 , d0 ) = 1.
d + mN ≡ 1 (mod c1 ). (2.1)
is a surjective homomorphism.
2.2 Subgroups of the modular group 163
We must lift this to an element of SL2 (Z). By Exercise 2.1.7, since (c, d, N ) =
1, we can find c0 , d0 such that
c0 ≡ c (mod N )
d0 ≡ d (mod N )
We know that ad0 −bc0 = 1+jN for some j so the problem reduces to solving
jN + kd0 N − lc0 N = 0,
or equivalently,
j = lc0 − kd0 .
Since (c0 , d0 ) = 1, the extended Euclidean Algorithm produces k, l ∈ Z satis-
fying this requirement. t
u
2.2.2. Prove that Γ(N ) is a normal subgroup of SL2 (Z) of finite index.
and consequently normal. The image of this map is a finite set and the result
follows from the first isomorphism theorem. t
u
2.2.3. For any commutative ring R with identity, let GL2 (R) be the set of 2 × 2
invertible matrices with entries in R. Show that GL2 (R) is a group.
2.2.4. Let p be a prime. Show that the order of GL2 (Z/pZ) is given by (p2 − 1)(p2 − p).
column. Thus, there are p2 − p choices. Hence, the group GL2 (Z/pZ) has
(p2 − 1)(p2 − p) elements. t
u
2.2.5. Let p be a prime. Show that SL2 (Z/pZ) is a normal subgroup of GL2 (Z/pZ)
of index p − 1. Deduce that SL2 (Z/pZ) has order p(p2 − 1).
The map
so the image has size (p − 1). Thus SL2 (Z/pZ) has index p − 1 in GL2 (Z/pZ).
The cardinality is now deduced from Exercise 2.2.4. t
u
can be seen to be surjective by first lifting a matrix in SL2 (Z/pZ) to SL2 (Z)
by Exercise 2.2.1 and then reducing modulo p2 . The kernel of the map in
(2.2) consists of matrices of the form
1 + t1 p t 2 p
t3 p 1 + t4 p
2.2.7. Apply an induction argument to show that SL2 (Z/pn Z) has size p3n−2 (p2 − 1).
which is seen to be surjective by first lifting a matrix in SL2 (Z/pn Z) to SL2 (Z)
by Exercise 2.2.1 and then reducing modulo pn+1 . The kernel consists of
matrices
1 + t 1 pn t 2 p n
t 3 pn 1 + t 4 pn
2.2 Subgroups of the modular group 165
Deduce that
Y 1
3
[SL2 (Z) : Γ(N )] = N 1− 2 .
p
p|N
Suppose N = pn1 1 · · · pnt t where the pi are unique primes. By the Chinese
Remainder Theorem,
tion to Exercise 2.2.2, since Γ(N ) is the kernel of the natural map SL2 (Z) →
SL2 (Z/N Z), we deduce that [SL2 (Z) : Γ(N )] = | SL2 (Z/N Z)|. t
u
both of order N .
Clearly −I, T 2 , and U 2 are in Γ(2) so that the group generated by these
three elements is contained in Γ(2). We adapt the proof of Theorem 2.1.2 and
use the following modified division algorithm: given a and b, with b 6= 0,
there are q and r such that a = bq + r with |r| ≤ |b|/2. Observe that r can be
negative. Now choose any
ab
g= ∈ Γ(2).
cd
1m
If c = 0, then as before, g = ± for some m ∈ Z. Since g ∈ Γ(2), m is
0 1
even and hence of the form m = 2k. We can thus write g = ±T 2k ∈ h−I, T 2 i.
Now suppose c 6= 0. If |a| > |c|, with c 6= 0, we write a = (2c)q + r with
|r| ≤ |c| and note that
1 −2q ab r b − 2qd
T −2q g = = .
0 1 cd c d
As c is even, we must have r and c of opposite parity and so |r| < |c|. If
|a| < |c|, then as a is odd, it is not zero and we can write c = (2a)q + r with
|r| ≤ |a|. Thus
1 0 ab a b
U −2q g = = .
−2q 1 cd r d − 2qb
As before, we must have |r| < |a|, as a and r are of opposite parity. Using
T 2 and U 2 in this way, we can rearrange the lower left entry to be zero and
thus we are done. t
u
2.2.11. In SL2 (Z), show that the group Γ generated by S and T 2 is a congruence
subgroup of index three.
and use the correspondence theorem and work in SL2 (Z/2Z). The image of
Γ/Γ(2) in SL2 (Z/2Z) is the group generated by S (mod 2), which has order
2. Thus [Γ : Γ(2)] = 2. As Γ(2) has index 6 in SL2 (Z), it follows that Γ has
index 3 in SL2 (Z). t
u
2.3 The Hecke subgroups 167
is a surjective homomorphism.
The fact that the map is a homomorphism is easily checked. The surjec-
tivity follows from the fact (N, d) = 1 implies there are integers a, b such that
ad − bN = 1. Consequently, the matrix
a b
∈ Γ0 (N )
N d
Γ1 (N )/Γ(N ) ∼
= Z/N Z
so that from
we deduce
[SL2 (Z) : Γ(N )]
[SL2 (Z) : Γ1 (N )] = .
N
The result now follows from Exercise 2.2.8. t
u
168 2 The Modular Group
1
Q
2.3.4. Show that Γ0 (N ) has index N p|N 1+ p in SL2 (Z).
Γ0 (N )/Γ1 (N ) ∼
= (Z/N Z)×
and so
[SL2 (Z) : Γ1 (N )]
[SL2 (Z) : Γ0 (N )] = ,
φ(N )
where φ denotes Euler’s function. Since by Exercise 2.3.3 the numerator is
Y 1
N2 1− 2
p
p|N
and
Y 1
φ(N ) = N 1− ,
p
p|N
Ggx = gGx g −1 .
(a) We have
But double cosets HxK and HyK are either disjoint or equal. To see this,
suppose they share a common element h1 xk1 = h2 yk2 . Then x = h−1 −1
1 h2 yk2 k1 .
For any h3 xk3 ∈ HxK, we have
h3 xk3 = h3 h−1 −1
1 h2 y k2 k1 k3 ∈ HyK
and
m
a
H= H 1 hi .
j=1
170 2 The Modular Group
a m
aa
X= Hyi = H 1 hi y i
i∈I i∈I j=1
2.5.1. If Γ is a congruence subgroup of level N and γ ∈ SL2 (Z), show that γΓγ −1
is again a congruence subgroup of level N .
We know that
and hence [Γ(M )Γ(N ) : Γ(N )] = [Γ(M ) : Γ(M ) ∩ Γ(N )]. Thus, by the Tower
Rule for the indices of groups, we deduce that
2.5 Supplementary problems 171
[Γ(1) : Γ(N )]
[Γ(1) : Γ(M )Γ(N )] =
[Γ(M )Γ(N ) : Γ(N )]
[Γ(1) : Γ(N )]
=
[Γ(M ) : Γ(M ) ∩ Γ(N )]
[Γ(1) : Γ(N )][Γ(1) : Γ(M )]
= .
[Γ(1) : Γ(M ) ∩ Γ(N )]
as desired. t
u
2.5.4. Define
10
Γ1 (N ) = g ∈ SL2 (Z) : g ≡ (mod N ) .
?1
This is clear. t
u
2.5.5. If SL2 (Z) has a non-congruence subgroup, show that it has infinitely many.
(Hint: consider Γ ∩ Γ (p) for a non-congruence subgroup Γ and rational
primes p.)
2.5.6. For each g ∈ SL2 (Z) and a subgroup Γ of finite index in SL2 (Z), we define
ng (Γ ) to be the least positive integer such that
ng (Γ ) −1 11
T ∈ gΓ g , T = as usual.
01
We say Γ is of class N if the least common multiple of the set of numbers {ng (Γ ) :
g ∈ SL2 (Z)} equals N . Show that Γ (N ) is of class N . Also show that the normal
closure of T N , namely
172 2 The Modular Group
is of class N . (This exercise extends the notion of congruence level defined for
congruence subgroups to all subgroups of finite index, and is due to Fricke
and Wohlfahrt.)
Clearly, nI (Γ (N )) = N and as Γ (N ) is a normal subgroup of SL2 (Z), we
see that Γ (N ) is of class N . Moreover, as D(N ) ⊆ Γ (N ), and T N ∈ D(N ), it
follows that D(N ) is of class N . t
u
2.5.7.
(a) Using the Chinese remainder theorem (or otherwise), show that for given x, y,
the following congruence
x(αy + β) ≡ γ mod m
b1 = b + nx(by + d) ≡ 0 mod mN
c2 = c1 + dN z ≡ 0 mod mN
2.5.8. With notation as in Exercise 2.5.6, show that D(N )Γ (M N ) = Γ (N ) for all
positive integers M, N .
g 0 = L−d+1 T −1 L1−a T d
2.5.11. Recall that Γ (2) is generated by −I, T 2 , U 2 (see Exercise 2.2.10) and that
the subgroup F generated by x1 = T 2 and x2 = U 2 is free of index 2 in Γ (2).
(a) With notation as in the previous exercise, compute the class of the group
Γ` ({1}) and Γ` ({1, 2}).
(b) Show that Γ` ({1, 2}) is not a congruence subgroup if ` is not a power of 2.
(a) Let H denote any of the two groups Γ` ({1}) or Γ` ({1, 2}). Then T k ∈ H
if and only if T k ≡ I mod 2 and `1 (T k ) ≡ 0 mod `. This happens if and
only if k ≡ 0 mod 2 and k ≡ 0 mod `. That is, if and only if 2`|k if ` is
odd and or k|` if ` is even. Thus, H is of class 2` if ` is odd and ` if ` is
even.
(b) We have shown the following facts:
[SL2 (Z) : Γ` ({1, 2})] = [SL2 (Z) : Γ (2)][Γ (2) : F ][F : Γ` ({1, 2})] = 12`2 .
Since Γp ({1, 2}) ⊇ Γ` ({1, 2}) for any prime p dividing `, it follows that
Γp ({1, 2}) is a congruence subgroup if Γ` ({1, 2}) is a congruence sub-
group. Also, [SL2 (Z) : Γ (2p)] = 6p(p2 − 1) if p > 2 and 48 if p = 2. Since
` is not a power of 2, let p be an odd prime that divides `. If Γ` ({1, 2}) is
a congruence subgroup, then so is Γp ({1, 2}) by our remark. However,
since Γp ({1, 2}) is of level 2p, it follows that if Γp ({1, 2}) is a congruence
subgroup, then Γp ({1, 2}) ⊇ Γ (2p) so that 12p2 |6p(p2 − 1), a contradic-
tion. Thus, Γ` ({1, 2}) is not a congruence subgroup.
. t
u
Chapter 3
The Upper Half-Plane
Let g = ac db ∈ GL+
2 (R). Then it is easily checked that
In addition, the identity I acts trivially (as well as −I). The other axiom is
easily verified directly. t
u
3.1.2. Show that the group PSL2 (R) = SL2 (R)/ ± I acts faithfully on H. That is,
gz = z for all z ∈ H implies g = 1 ∈ PSL2 (R).
Let g = ac db and suppose that for all z ∈ H, we have
az + b
= z.
cz + d
Then cz 2 + (d − a)z − b = 0 for all z ∈ H which implies c = b = 0 and d = a
since a quadratic polynomial has at most two roots. Since the determinant
a2 = 1, we have a = ±1 so that g = ±I. t
u
3.1.3. Show that the action of PSL2 (R) on H is transitive, i.e. there is only one orbit
and for all z, w ∈ H, there exists some g ∈ PSL2 (R) such that gz = w.
It suffices to show that for any z ∈ H, there is a g ∈ PSL2 (R) such that
gz = i. Then given w ∈ H, there is also an element g1 ∈ PSL2 (R) such that
g1 (w) = i. Thus gz = g1 w and (g1−1 g)z = w.
© Springer Science+Business Media Singapore 2016 and Hindustan Book Agency 2016 175
M.R. Murty et al., Problems in the Theory of Modular Forms, HBA Lecture Notes
in Mathematics, DOI 10.1007/978-981-10-2651-5_13
176 3 The Upper Half-Plane
1 − |w|2
i(1 + w)(1 − w)
Im(z) = Im = > 0. t
u
|1 − w|2 |1 − w|2
3.1.6. Show that SL2 (Z) is a discrete subgroup of SL2 (R) ⊂ R4 , where R4 is en-
dowed with the Euclidean topology.
3.2.4. Suppose that z, w ∈ F are distinct points which are SL2 (Z)-equivalent.
Show that Im(z) = Im(w) and Re(z) = − Re(w).
Im(z)
Im(w) = = Im(z)
|cz + d|2
and
a ∓1 1
w= z = ±a − .
±1 0 z
As Sz = −1/z maps the arc of the unit circle with π/3 ≤ arg z ≤ π/2 to the
arc with π/2 ≤ arg z ≤ 2π/3, we see that either w = −1/z or that {z, w} =
{ρ, ρ2 }. Hence Re(w) = − Re(z), as required. t
u
3.2.5. If F is the standard fundamental domain for SL2 (Z) and γ ∈ SL2 (Z), show
that γF is again a fundamental domain for SL2 (Z).
3.2.7. Show that I, S, T −1 S are coset representatives for Γ0 (2) in SL2 (Z). Deduce
that a fundamental domain for Γ0 (2) is given by F ∪ SF ∪ ST F , where F is the
standard fundamental domain for SL2 (Z).
The group Γ0 (2) has index three in SL2 (Z) and it is easily verified that
I, S, T −1 S are coset representatives for Γ0 (2). To see that F ∪SF ∪ST F has a
connected interior, we need only sketch the region explicitly as in Figure 3.1
on page 178. t
u
3.3.2. If Γ is of finite index in SL2 (Z), show that the number of Γ-inequivalent
cusps is ≤ [SL2 (Z) : Γ].
178 3 The Upper Half-Plane
i
ρ
ρ2
SF
ST F
−1 − 12 0 1 1
2
Let s ∈ Q ∪ {i∞} be a cusp. By Exercise 3.3.1, there exists τ ∈ SL2 (Z) such
that s = τ (i∞). If g1 , . . . , gr are right coset representatives of Γ in SL2 (Z),
then s = τ (i∞) = γgj (i∞) for some γ ∈ Γ. Hence the set of Γ-inequivalent
cusps is contained in the set
along with ( 10 01 ), give a complete set of right coset representatives for Γ0 (p) in
SL2 (Z).
then −1
k0 1
0 −1 0 −1 0 −1
= ∈ Γ0 (p)
1 k 1 k0 1 k −1 0
which means k ≡ k0 (mod p).
Now given ac db ∈ SL2 (Z), if ac db 6∈ Γ0 (p) then we want to show that
there exists some k such that
ab 0 −1
∈ Γ0 (p) .
cd 1 k
so we need only find k so that ck ≡ d (mod p). This is always possible when
p - c. t
u
3.3.7. Let p be prime. Show that a complete set of inequivalent cusps for Γ0 (p) is
given by {0, i∞}.
180 3 The Upper Half-Plane
such that
ab
(i∞) = 0,
cd
6 0 and i∞ if c = 0. This means that our
then the left hand side is a/c if c =
matrix is of the form
0b
,
cd
which has determinant −bc and is consequently divisible by p, a contradic-
tion. t
u
3.4.1. Prove that d(z, w) is SL2 (R)-invariant. That is, show that d(gz, gw) =
d(z, w) for all g ∈ SL2 (R).
It suffices to show that
|z − w|2
Im(z) Im(w)
is SL2 (R)-invariant. Indeed, let g = ac db ∈ SL2 (R). Then by the solution to
Exercise 3.1.1,
Im(z) Im(w)
Im(gz) Im(gw) =
|cz + d|2 |cw + d|2
and
az + b aw + b z−w
− = ,
cz + d cw + d (cz + d)(cw + d)
since ad − bc = 1. The result is now immediate. t
u
3.4.2. Show that for x ≥ 1,
p
cosh−1 x = log(x + x2 − 1).
We have
(y − 1)2
−1
d(i, iy) = cosh 1+
2y
−1 1 1
= cosh y+
2 y
s 2 !
1 1 y 1 1
= log y+ + + 2−
2 y 4 4y 2
s
2
1 1 y 1
= log y+ + − .
2 y 2 2y
Thus
1 1 1 1
d(i, iy) = log y+ + y− .
2 y 2 y
If y > y1 , this is log y. If y < y1 , this is − log y. In either case, it can be written
as | log y|. t
u
3.4.5. Let F be the standard fundamental domain for SL2 (Z). Show that
Z Z
dxdy π
2
= .
F y 3
We compute easily Z Z
dxdy
= 1.
F :y≥1 y2
The integral
1
Z Z Z Z 1
dxdy 2 dy
=2 √ dx
F :y≤1 y2 0 1−x2 y2
Z 1
2 1
=2 −1 + √ dx
0 1 − x2
1
= −1 + 2 sin−1
2
π
= −1 + ,
3
182 3 The Upper Half-Plane
3.5.1. Show that a set of coset representatives for Γ∞ in SL2 (Z) is given by
??
: (c, d) = 1, c, d ∈ Z ,
cd
where the asterisks indicate that we can take any a, b such that ad − bc = 1.
The coset
ab a + nc b + nd
Γ∞ = :n∈Z .
cd c d
Thus, for fixed c, d with (c, d) = 1, the coset parametrizes all integer solu-
tions of ad − bc = 1, by the Euclidean algorithm. t
u
Since the measure dxdy/y 2 is SL2 (Z)-invariant, each integral is equal to π/3,
the area of F , by Exercise 3.4.5. t
u
3.5.3. Let G = SL2 (R) act on H. Show that K, the stabilizer of i, consists of matri-
ces
cos θ sin θ
for θ ∈ R.
− sin θ cos θ
Then show that the upper half-plane is homeomorphic to G/K.
If the matrix ac db ∈ SL2 (R) fixes i, then we must have
ai + b = i(ci + d) = −c + di.
3.5 Supplementary problems 183
Uc = {z ∈ H : Im(z) > c}
of i∞ is mapped via z 7→ e2πiz into the punctured open disc centered at zero, of
radius e−2πc .
If z = x + iy, then e2πiz = e−2πy e2πix . For fixed y and x ∈ R, the map
z 7→ e2πiz takes the line Im(z) = y in H to the circle centered at zero of
radius e−2πy . t
u
so that
j(γ1 γ2 , z) = (c1 a2 + d1 c2 )z + (c1 b2 + d1 d2 ).
Now j(γ2 , z) = c2 z + d2 and
a 2 z + b2
j(γ1 , γ2 z) = c1 + d1
c 2 z + d2
(c1 a2 + d1 c2 )z + (c1 b2 + d1 d2 )
=
c 2 z + d2
from which the result follows immediately. t
u
© Springer Science+Business Media Singapore 2016 and Hindustan Book Agency 2016 185
M.R. Murty et al., Problems in the Theory of Modular Forms, HBA Lecture Notes
in Mathematics, DOI 10.1007/978-981-10-2651-5_14
186 4 Modular Forms of Level One
Since (f |γ1 )(z) = (det γ1 )k/2 j(γ1 , z)−k f (γ1 z), we have
(f |γ1 ) γ2 (z) = (det γ2 )k/2 j(γ2 , z)−k (f |γ1 )(γ2 z)
= (det γ2 )k/2 j(γ2 , z)−k (det γ1 )k/2 j(γ1 , γ2 z)−k f (γ1 (γ2 z))
= f | (γ1 γ2 ) ,
We conclude that αf + βg ∈ Mk1 (SL2 (Z)). It is now clear that Mk1 (SL2 (Z))
and Sk1 (SL2 (Z)) are vector spaces.
Now for f ∈ Mk1 (SL2 (Z)) and g ∈ Mk2 (SL2 (Z)), if γ = ac db ∈ SL2 (Z),
then
(f g)(γz) = f (γz) · g(γz)
= (cz + d)k1 f (z) · (cz + d)k2 g(z)
= (cz + d)k1 +k2 (f g)(z).
Moreover, vi∞ (f g) = vi∞ (f ) + vi∞ (g). This proves (b) and (c).
4.2 Eisenstein series for SL2 (Z) 187
4.2.1. Show that the series for Gk (z) converges absolutely for all z ∈ H and for
k > 2.
We note that for z = x + iy, x, y ∈ R, we have
so that B0 = 1, B1 = − 12 , B2 = 16 , B3 = 0, B4 = − 30
1
, . . . . Show that
(2πi)2k B2k
2ζ(2k) = − , k ≥ 1.
(2k)!
2πiz
πz cot πz = πiz + .
e2πiz − 1
But the right hand side is
∞
X (2πiz)k
πiz + Bk . (4.1)
k!
k=0
upon writing
∞
z2 (z/n)2 X z 2k
= = .
n2 − z 2 1 − (z/n)2 n
k=1
Thus
∞
X
πz cot πz = 1 − 2 ζ(2k)z 2k .
k=1
4.2.5. Prove that the Bernoulli numbers satisfy the recursion formula:
m−1
X
m+1
(m + 1)Bm = − Bk .
k
k=0
Prove that
4.3 The valence formula 189
∞
text X xn
t
= Bn (x) .
e − 1 n=0 n!
4.2.8. Prove that Mk (SL2 (Z)) = CEk ⊕ Sk (SL2 (Z)) for all k ≥ 4. Deduce that
dim Mk (SL2 (Z)) = 1 + dim Sk (SL2 (Z)).
Suppose f ∈ Mk (SL2 (Z)). Let λ = f (i∞) denote the constant term in the
Fourier expansion. Clearly
f = λEk + f − λEk
where f −λEk ∈ Sk (SL2 (Z)) since it vanishes at the cusp i∞. The conclusion
follows. t
u
If k = 0, the left hand side of the valence formula is zero. However, for
any f (z) ∈ M0 (SL2 (Z)), the form f (z) − f (i) ∈ M0 (SL2 (Z)) clearly has a
zero at z = i and hence f (z) must be constant. t
u
The right hand side of the valence formula is 1/6. This is impossible un-
less f ≡ 0. t
u
If the weight of a modular form f is ≤ 10, then the right hand side of the
valence formula is ≤ 5/6 < 1. If there were a non-trivial cusp form of weight
≤ 10, there would be a contribution of at least 1 from the zero at i∞. t
u
E43 − E62
4.4.5. Prove that ∆ = has integer Fourier coefficients.
1728
We can write
E4 = 1 + 240A
E6 = 1 − 504B
where A, B are q-series with integer coefficients. Now
4.5 The j-function 191
Since 1728 = 123 , 240 = 12 × 20, and 504 = 12 × 42, we see that
and X
5σ3 (n) + 7σ5 (n) = (5d3 + 7d5 ).
d|n
By Exercises 4.3.5 and 4.3.6, we know that E4 (ρ) = 0 and E6 (i) = 0, and
that these are simple zeros. By the valence formula, there are no other zeros
of E4 and E6 in the fundamental domain F . Now
E4 (ρ)3
j(ρ) = =0
∆(ρ)
and
1728E4 (i)3
j(i) = = 1728.
E4 (i)3 − E6 (i)2
Moreover, it is clear that vρ (j) = 3vρ (E4 ) − vρ (∆) = 3, and since
4.5.3. Show that any modular function f has a finite number of zeros and poles in
F.
The set {z ∈ H : Im(z) > c} being a fundamental neighborhood of i∞,
and as the zeros and poles of any meromorphic function are isolated, we can
choose c sufficiently large such that i∞ is the only zero or pole of f in this
neighborhood. Now as the region
F ∩ {z ∈ H : Im(z) ≤ c},
xex x x(ex − 1) + x x
f (−x) = x
− 1 − = x
− 1 − = f (x)
e −1 2 e −1 2
is even. Hence Bk = 0 for odd k ≥ 3.
Alternatively, we may conclude that the Bk vanish for odd k ≥ 3 by com-
paring coefficients on both sides of equation (4.2) in the solution to Exer-
cise 4.2.4. t
u
4.6.2. Show that for k even and positive, (−1)k/2 Bk < 0.
This follows from (4.2) in the solution to Exercise 4.2.4 since ζ(2k) > 0. t
u
4.6.3. Prove that
Bk+1 (n + 1) − Bk+1
1k + 2 k + · · · + n k = ,
k+1
for every natural number k.
4.6 Supplementary problems 193
and the desired result now follows upon applying Exercise 4.2.6. t
u
4.6.4. Prove that ζ(k) tends to 1 as k tends to infinity and use this to show that
√
|B2k | ∼ 4π e(k/πe)2k+1/2
as k tends to infinity.
By the integral test, it is easily seen that ζ(k) = 1 + O(1/2k−1 ) so that ζ(k)
tends to 1 as k tends to infinity. The asymptotic for the Bernoulli numbers
follows upon using this together with Stirling’s formula:
√
(k)! ∼ (k/e)k 2πk.
t
u
4.6.5. Show that if the Fourier coefficients of f ∈ Mk (SL2 (Z)) are integers, then f
can be written as a polynomial in E4 , E6 , and ∆ with integer coefficients.
Write z = x + iy ∈ H. Then
194 4 Modular Forms of Level One
∞
X
lim |f (x + iy) − a0 | = lim an e2πi(x+iy)n
y→∞ y→∞
n=1
∞
X
≤ lim |an | e−2πyn
y→∞
n=1
= 0.
a b
If x ∈ Q, then there exists γ = c d ∈ SL2 (Z) such that γx = i∞. In
particular, cx + d = 0 and so
ax + b + aiy ax + b + aiy a ax + b i
γ(x + iy) = = = − .
cx + d + ciy ciy c c y
Since γ(x + iy) ∈ H, we must have (ax + b)/c < 0. Thus we set t :=
exp (2π(ax + b)/c) < 1. Writing m = vi∞ (f ) now gives
∞ ∞
f (γ(x + iy)) X an 2πi( ac − ax+b
cy i)(n−m) =
X an (n−m)/y
− 1 ≤ e t
am e2πiγ(x+iy)m n=m+1
a m n=m+1
a m
By modularity,
−k
f (z) = (cz + d) f (γz)
and hence
−k
lim f (x + iy) = lim (cx + iyc + d) f (γ(x + iy))
y→0 y→0
−k
= lim (iyc) f (γ(x + iy))
y→0
am e2πiγ(x+iy)m
= lim .
y→0 (iyc)k
am tm/y
lim f (x + iy) = k
lim k = 0
y→0 |c| y→0 y
The dimension formula shows that M14 (SL2 (Z)) is one-dimensional and
so is spanned by E14 . Since E42 E6 also has weight 14, we have E42 E6 = cE14
for some constant c. A quick verification using Bernoulli numbers and
Corollary 4.2.5 gives the result. t
u
4.6.8. Let rk = dim Mk (SL2 (Z)). Put ` = k − 12rk + 12 and E0 = 1. Show that
E` E14−` = E14
dj
= −2πiE14 /∆.
dz
We have
dj/dz = 3E42 E40 ∆−1 − ∆−2 E43 ∆0
so that
E43 − E62
d
2
∆ dj/dz = 3E42 E40 ∆ − E43
dz 1728
which equals
E42 E6
(2E4 E60 − 3E6 E40 ) .
1728
By the previous exercise, the expression in the brackets is a modular form of
weight 12. A quick verification shows that it is a cusp form and so must be a
suitable multiple of ∆. Comparing coefficients in the q-expansion immedi-
ately gives us the result. t
u
196 4 Modular Forms of Level One
4.6.11. Let rk = dim Mk (SL2 (Z)) and ` = k − 12rk + 12. For any modular form
f ∈ Mk (SL2 (Z)), show that
E4a E6b
such that 4a + 6b = k. From the previous exercise, E`−1 = E14−` /E14 , so that
f E`−1 is a linear combination of terms like
(a, b) ∈ {(0, 1), (1, 0), (1, 1), (2, 0), (2, 1)}.
f E`−1 ∆1−rk
4.6.13. Let f be a modular form of weight k for the full modular group whose Fourier
expansion at i∞ is given by
∞
X
f (z) = an e2πinz , an ∈ C.
n=0
Recall that by Corollary 4.4.4, there are unique cab ’s such that
X
f= cab E4a E6b , cab ∈ C.
4a+6b=k
Then, show that cab belong to F , whence all the Fourier coefficients belong to F .
Let F be the smallest field over Q containing the first dk Fourier coeffi-
cients of f , where dk is the dimension of Mk (SL2 (Z)) and k is the weight of
f . Then, by the previous exercise, all the Fourier coefficients are in F , which
is a finite extension of Q since it was obtained by adjoining finitely many
algebraic numbers. t
u
Using Corollary 4.4.4 and the earlier exercises, there are unique algebraic
numbers cab such that X
f= cab E4a E6b .
4a+6b=k
We first write
∞ ∞
1 X X 1
E2 (z) = 1 + .
ζ(z) m=1 n=−∞ (mz + n)2
Using the Lipschitz formula (Theorem 4.2.2) with k = 2 for the inner infinite
series, we get
∞ ∞
(−2πi)2 X X md
E2 (z) = 1 + dq .
ζ(2) m=1
d=1
© Springer Science+Business Media Singapore 2016 and Hindustan Book Agency 2016 199
M.R. Murty et al., Problems in the Theory of Modular Forms, HBA Lecture Notes
in Mathematics, DOI 10.1007/978-981-10-2651-5_15
200 5 The Ramanujan τ -function
Show that r
1 z
η − = η(z),
z i
where the square root is the branch with non-negative real part.
The product defining η(z) clearly converges to a non-zero value for any
z ∈ H and thus defines a holomorphic function on H. Taking the logarithmic
derivative, we find
∞
!
η 0 (z) πi X nq n
= 1 − 24 ,
η(z) 12 n=1
1 − qn
where q = e2πiz . As in the proof of Theorem 5.1.4, the right hand side is
easily seen to be πiE2 (z)/12. Therefore
η 0 (−1/z)
πi 1 πi 2 6z
= E2 − = z E2 (z) + .
η(−1/z) 12 z 12 πi
In other words
η 0 (−1/z) 1 πiE2 (z) 1
· = + .
η(−1/z) z 2 12 2z
so that
η 0 (−1/z) 1 η 0 (z) 1
· 2 = + .
η(−1/z) z η(z) 2z
√
But the right hand side
√ is clearly the logarithmic derivative of zη(z). There-
fore, η(−1/z) = λ zη(z) for some constant√λ. Putting z = i shows that
√
λ = 1/ i. Finally, since z ∈ H, it is clear that √zi = zi .
p
t
u
5.1.6. Show that
∆0 (z)
= 2πiE2 (z).
∆(z)
Since
∞
Y
∆(z) = e2πiz (1 − e2πinz )24 ,
n=1
Now
d az + b
= (cz + d)−2
dz cz + d
so that
az + b
f0 = (cz + d)k+2 f 0 (z) + k(cz + d)k+1 cf (z).
cz + d
We apply this to ∆ ∈ M12 (SL2 (Z)) to get
az + b
∆0 = (cz + d)14 ∆0 (z) + 12c(cz + d)13 ∆(z).
cz + d
Dividing through by
az + b
∆ = (cz + d)12 ∆(z),
cz + d
1 0 k
θk f := f − E2 f.
2πi 12
Prove that θk : Mk (SL2 (Z)) → Mk+2 (SL2 (Z)).
and
0 az + b
f = (cz + d)k+2 f 0 (z) + kc(cz + d)k+1 f (z).
cz + d
Putting these together, we find
az + b 1 0 az + b k az + b az + b
(θk f ) = f − E2 f
cz + d 2πi cz + d 12 cz + d cz + d
1 k
= (cz + d)k+2 f 0 (z) − E2 (z)f (z)
2πi 12
= (cz + d)k+2 (θk f )(z)
as required.
It remains to check that θk f is holomorphic at the cusp i∞. In a neighbor-
hood of i∞,
X∞
f (z) = an e2πizn
n=0
and so
∞
1 0 X
f (z) = an ne2πizn .
2πi n=0
By Exercise 5.1.2, E2 (z) also has a Fourier expansion at i∞, and putting the
expansions together easily shows that θk f is holomorphic at i∞. t
u
5.2 Hecke operators of level one 203
5.2.5. Prove that Tm (∆) = τ (m)∆. Deduce that τ (mn) = τ (m)τ (n) for (m, n) =
1.
By Theorem 5.2.2, Tm (∆) is again a cusp form of weight 12. Since S12 (SL2 (Z))
is one-dimensional, and spanned by ∆, we must have Tm (∆) = λm ∆ for
some constant λm . On the other hand, by Theorem 5.2.4, the n-th coefficient
in the q-expansion of Tm (∆) is
X mn
d11 τ = τ (mn)
d2
d|m,n
whenever (m, n) = 1. But Tm (∆) = λm ∆ and the right hand side of the
equation shows that the n-th coefficient is λm τ (n). Thus τ (mn) = λm τ (n).
Putting n = 1 shows τ (m) = λm from which the result follows. t
u
Again, comparing the n-th coefficient of both sides of the equation Tm (∆) =
τ (m)∆, we get mn
X
τ (m)τ (n) = d11 τ ,
d2
d|m,n
5.2.7. Let p be prime and a ≥ 1. Show that τ (pa+1 ) = τ (p)τ (pa ) − p11 τ (pa−1 ).
as required. t
u
204 5 The Ramanujan τ -function
where k = 16, 18, 20, 22, or 26, respectively. If (m, n) = 1, the sum becomes
af (mn). Hence af (mn) = λm af (n), whenever (m, n) = 1. Choosing n = 1,
we deduce af (m) = λm . The result now follows. t
u
5.2.9. Let p be a prime number. Show that
∞
X 1
τ (pn )xn = .
n=0
1 − τ (p)x + p11 x2
For n ≥ 2, the coefficient of xn in the product on the left hand side is τ (pn ) −
τ (p)τ (pn−1 ) + p11 τ (pn−2 ) which is zero by Exercise 5.2.7. For n = 1, the
coefficient of x is easily checked to be zero also, and as the coefficient of x0
is one, we are done. t
u
5.2.10. Prove that for n > 1,
n−1
X
(1 − n)τ (n) = 24 σ1 (j)τ (n − j).
j=1
By Exercise 5.1.6,
1 0
∆ (z) = ∆(z)E2 (z)
2πi
In other words,
∞ ∞ ∞
! !
X X X
nτ (n)q n = τ (n)q n 1 − 24 σ1 (n)q n .
n=1 n=1 n=1
5.3 Modular forms and Dirichlet series 205
This follows from Exercise 1.5.1 where it was shown that τ (n) is odd if
and only if n is an odd square. t
u
|τ (p)| ≤ 2p11/2
is equivalent to the assertion that the roots of the quadratic polynomial 1 − τ (p)x +
p11 x2 are non-real.
The discriminant of the quadratic is τ (p)2 − 4p11 and this is < 0 if and
only if |τ (p)| < 2p11/2 . We have strict inequality here since τ (p) is an integer
and p11/2 is irrational. t
u
5.3.2. Assuming Ramanujan’s conjecture that |τ (p)| ≤ 2p11/2 , prove that |τ (n)| ≤
n11/2 d(n) for every natural number n. Here d(n) = σ0 (n) is the number of divisors
of n.
Now
Y
|τ (n)| = τ (pα ) ≤ n11/2 d(n). t
u
pα ||n
5.3.3. Let f ∈ Mk (SL2 (Z)). Show that y k/2 |f (z)| is invariant under SL2 (Z),
where y = Im(z).
We have
az + b
f = (cz + d)k f (z)
cz + d
and
az + b Im(z)
Im =
cz + d |cz + d|2
from which the result follows. t
u
P∞ a (n)
and that af (0) 6= 0. Show that Lf (s) = n=1 fns has an analytic continuation
to the entire complex plane, except for a simple pole at s = 0 and s = k and satisfies
the functional equation
Put a0 = af (0). Let us note that f (iy) − a0 has exponential decay and we
can consider
Z ∞
(f (iy) − a0 )y s−1 dy = (2π)−s Γ(s)Lf (s)
0
as in the proof of Theorem 5.3.7. We split the integral into two parts:
Z 1 Z ∞
(f (iy) − a0 )y s−1 dy + (f (iy) − a0 )y s−1 dy.
0 1
5.4 Ramanujan congruences 207
In the first integral, we change variables y 7→ 1/y. Using the modular rela-
tion, we get
Z ∞
{ik y k (f (iy) − a0 ) + ik a0 y k − a0 }y −s−1 dy.
1
from which the analytic continuation and functional equation are now im-
mediate. t
u
The second congruence is obvious from the q-expansion of R. For the first,
we note that Q2 = E8 and
208 5 The Ramanujan τ -function
∞
X n7 q n
E8 = 1 + 480 .
n=1
1 − qn
Q 2 = E8 ≡ P (mod 7). t
u
Now
dP 2 dQ
(Q3 − R2 )2 ≡ 6q − 3q (mod 7)
dq dq
dP 2 dQ
≡ −q − 3q (mod 7),
dq dq
as required. t
u
By Theorem 5.1.4,
5.5 Supplementary problems 209
2 ∞ ∞
Q3 − R 2 (1 − q n )49
Y Y
= q2 (1 − q n )48 = q 2 .
1728 n=1 n=1
1 − qn
Hence we obtain
∞ ∞
!
X Y
3 2 2 2 n
(Q − R ) ≡ q p(n)q (1 − q 49n ) (mod 7),
n=0 n=1
Noting that for any q-series J with integer coefficients, the nth coefficient
of q dJ
dq is always divisible by n, we need only compare coefficients of q
7n
of
both sides of the congruence (5.1) to deduce the result. t
u
762048
5.5.1. Prove that E62 = E12 − 691 ∆.
Since E62 has weight 12, we can write it as αE12 +β∆ because M12 (SL2 (Z))
is spanned by E12 and ∆. Computing the coefficients of q 0 and q 1 leads to
the result. t
u
∞
5.5.2. Show that E12 = 1 + 65520 n
P
691 n=1 σ11 (n)q .
Deduce that
n−1
X 65
σ11 (n) + O n6 .
σ5 (j)σ5 (n − j) =
j=1
(252)(691)
Show that
αk+1 − β k+1
fk = .
α−β
If α = β, the right hand side is viewed as αk + αk−1 β + · · · + αβ k−1 + β k =
(k + 1)αk .
The recurrence for fk implies that we have the equality of power series
∞
X −1
fk xk = 1 − f x − gx2 .
k=0
Expanding the right hand side as a power series gives the desired result. t
u
holds for all natural numbers m and n. Show that f is g-multiplicative if and only
if f is multiplicative and for all primes p and all natural numbers k ≥ 1, we have
αpk+1 − βpk+1
f pk = . (5.2)
αp − βp
min(j,k)
X
f (pj )f (pk ) = g(p)i f (pj+k−2i ).
i=0
since g(p) = αp βp . The right hand side of the above equation is easily
summed as
j j
αpk+1 X j−i i βpk+1 X i j−i
α β − α β
αp − βp i=0 p p αp − βp i=0 p p
5.5.10. Let gk (n) = nk . Show that the Ramanujan τ -function is g11 -multiplicative.
ps(α+1) − 1
σs (pα ) = 1 + ps + · · · + psα = .
ps − 1
Show that X
G(m, n) = µ(d)F (m/d, n/d)
d|m,n
Since
τ (m) τ (n) X
2 d11
= τ (mn/d ) ,
m11/2 n11/2 (mn)11/2
d|m,n
we can apply the previous exercise to the functions F (m, n) = τ (m)τ (n)/(mn)11/2
and G(m, n) = τ (mn)/(mn)11/2 to deduce the result. t
u
5.5.14. Let A be the n × n matrix whose (i, j) entry is τ (ij). Show that the determi-
nant of A is given by n!11 µ(1)µ(2) · · · µ(n). In particular, the determinant is zero
if n ≥ 4.
We have
∞
X ∞
Y
τ (n)q n = q (1 − q n )24
n=1 n=1
t
u
Fk := E12r−k+2 ∆−r .
Since
∞
Y
∆−1 = q −1 (1 + q n + q 2n + · · · )24 ,
n=1
the Fourier coefficients in the q-expansion of any integer power of ∆ are in-
tegers. Writing k = 12t + δ with 0 ≤ δ ≤ 10, we have by the dimension
formula that r = t + 1 or t according as δ 6≡ 2(mod 12) or δ ≡ 2(mod 12) re-
spectively. Thus, 12r − k + 2 = 14 − δ or 2 − δ according as δ 6≡ 2(mod 12) or
δ ≡ 2(mod 12). The Eisenstein series E12r−k+2 has accordingly integer coef-
ficients. Thus, the product E12r−k+2 ∆−r also has integer coefficients. More-
over, it is easily checked that the leading coefficient is 1. t
u
5.5.17. Let j denote the j-function. Show that for any integer m ≥ 0,
dj
jm
dz
has a q-expansion without a constant term.
and
dj 1 dj m+1
jm = ,
dz m + 1 dz
the right hand side has no constant term from which the result is immediate.
t
u
5.5.18. Let f be a modular form of weight k for the full modular group with q-
expansion
5.5 Supplementary problems 215
f (z) = a0 + a1 q + a2 q 2 + · · ·
With Ckj defined as in Exercise 5.5.16, show that
Fk = E12r−k+2 ∆−r
1 dj
− W = f E14−` ∆−r = f Fk
2πi dz
(using Exercise 5.5.16). Comparing the constant term of both sides gives us
the desired result. t
u
5.5.19. Suppose that k ≡ 2 (mod 4). With notation as in the previous exercise, show
that Ck0 6= 0.
Ck0 is the constant term of Fk = E12r−k+2 ∆−r with r = dim Mk (SL2 (Z)).
Since k ≡ 2(mod 4), we have k ≡ 2t(mod 12) with t = 1, 3 or 5. Then 12r =
k − 2, k + 6, k + 2 respectively, so that 12r − k + 2 = 0, 8, 4 respectively. Thus,
E12r−k+2 = E0 , E42 , E4
respectively. As all the Fourier coefficients of E4 are positive and the same is
true for ∆−r , we deduce that all Ck0 , Ck1 , ..., Ckr are positive and so Ck0 > 0.
t
u
5.5.20. Suppose that k ≡ 0 (mod 4). With notation as in the previous exercise, show
that Ck0 6= 0.
216 5 The Ramanujan τ -function
−1 dj
Fk = E12r−k+2 ∆−r = −E12r−k+2 ∆1−r E14 ,
2πidz
by Exercise 4.6.10. Again, by Exercise 4.6.8, E` E14−` = E14 with ` = k −
12r + 12 so that
−1 dj dj
Fk = −E14−` ∆1−r E14 = −E`−1 ∆1−r .
2πidz 2πidz
As ` = 4t, we can rewrite this as
dj
Fk = −E4−t ∆1−r .
2πidz
Since E43 /∆ = j, it is not difficult to see that
3r + t − 3 3−t d∆−r
Vk := E .
(3 − t)r 4 2πidz
Since k > 2, we have 3r + t − 3 > 0 and the series E43−t begins with 1. More-
over, it has all strictly positive coefficients. Furthermore, the coefficients of
∆−r are all positive and when we differentiate and divide by 2πi, the coef-
ficients are all strictly negative, while the constant term is zero. Hence the
constant term of Vk is negative. That is, Ck0 < 0 and hence Ck0 6= 0. t
u
Chapter 6
Modular Forms of Higher Level
for some b ∈ Z. Suppose that the expansions at γ(i∞) and τ (i∞) are
X
f |τ = bn q n/N
X
f |γ = cn q n/N .
f |τ = f |δτ
1b
= f |γ| ±
01
X
k
= (±1) cn e2πi(z+b)n/N
X
= (±1)k cn e2πibn/N q n/N .
In particular, for all n we have bn = ±cn e2πibn/N . Thus, bn and cn are either
both zero or both non-zero. Hence the order of f is well-defined. t
u
© Springer Science+Business Media Singapore 2016 and Hindustan Book Agency 2016 217
M.R. Murty et al., Problems in the Theory of Modular Forms, HBA Lecture Notes
in Mathematics, DOI 10.1007/978-981-10-2651-5_16
218 6 Modular Forms of Higher Level
6.2.1. Let k ≥ 3 and N a natural number. Let g ∈ (Z/N Z)2 . Define the Eisenstein
series X 0 1
Gk,g (z) = , z∈H
(mz + n)k
(m,n)≡g (mod N )
where the summation is over all integers m, n satisfying the congruence condition
(m, n) ≡ g (mod N ) and the dash on the summation means we exclude (m, n) =
(0, 0) in the case g = (0, 0). For any γ ∈ SL2 (Z), show that Gk,g |γ = Gk,gγ .
Conclude that Gk,g ∈ Mk (Γ(N )).
The summands of Gk,g (z) are a proper subset of the summands of Gk (z).
We have already seen (Exercise 4.2.1) that series of this form converge ab-
solutely and uniformly on compact subsets of H for k > 2. Thus, Gk,g (z)
is holomorphic on the upper half-plane. Moreover, if g = (a1 , a2 ) with
a1 6≡ 0 (mod N ), then limIm z→∞ Gk,g (z) = 0. On the other hand, if a1 ≡ 0
(mod N ), then X
lim Gk,g (z) = n−k .
Im z→∞
n≡a2 (mod N )
since
ab
(m n) = ma + nc mb + nd
cd
and if we write gγ = (b1 , b2 ) the map (m, n) 7→ (m, n)γ sets up a one-to-one
correspondence between pairs of integers (m, n) ≡ g (mod N ) and pairs
(m0 , n0 ) ≡ gγ (mod N ). This proves Gk,g |γ = Gk,gγ for all γ ∈ SL2 (Z). Since
Gk,g (z) is holomorphic at i∞, it follows that it is holomorphic at all cusps. Fi-
nally, if γ ≡ I (mod N ), then gγ ≡ g (mod N ) so that Gk,g (z) ∈ Mk (Γ(N )).
t
u
6.2.2. Show that Gk,g = (−1)k Gk,−g . Deduce that if k is odd and 2g ≡ (0, 0)
(mod N ), then Gk,g = 0.
Putting γ = −I in Exercise 6.2.1, we see that Gk,g (z) = (−1)k Gk,−g (z).
To say 2g ≡ (0, 0) (mod N ) means g ≡ −g (mod N ) so that if k is odd,
Gk,g (z) ≡ 0. t
u
6.2.3. Let Γ be a congruence subgroup of SL2 (Z) of level N . If γ1 , . . . , γr are distinct
right coset representatives of Γ(N ) in Γ, show that
6.2 Eisenstein series for Γ(N ) 219
r
X
f := Gk,gγj ∈ Mk (Γ).
j=1
6.2.6. Show that there are exactly three inequivalent cusps for the group Γ0 (4). Find
the widths of each cusp.
three cusps: i∞, 1/2, and 0. To determine the width of γ(i∞), we look for
h ∈ Z such that
1h
γ(±I) γ −1 ∈ Γ0 (4).
01
Since −I ∈ Γ0 (4), we consider γ ( 10 h1 ) γ −1 for each of the coset representa-
tives γ. We first consider γ = ( 12 01 ). Now
1h −1 10 1h 1 0 1 − 2h h
γ γ = = .
01 21 01 −2 1 −4h 1 + 2h
It is clear that this matrix is in Γ0 (4) for all h ∈ Z. Hence, the cusp ( 12 01 ) i∞ =
1
2 has width 1.
Now consider γ = 01 −1
m for m = 1, 2, 3, or 4. We have
1h 0 −1 1h m 1 10
γ γ −1 = = ,
01 1 m 01 −1 0 h1
(−2πi)k X
bk,g (n) = dk−1 (sgn d).
N k (k − 1)!
d|n
n/d≡a1 (mod N )
where ζ = e2πi/N . Deduce that Gk,g (z), with g = (0, a2 ), has a q-expansion of the
form
X∞
an q n , q = e2πiz .
n=0
If N | n, we see that the summation formula for bk,g (n) in Theorem 6.3.1
shows that if we write n = N n0 , the summation over divisors of N n0 can be
re-written as a sum over divisors of n0 which reduces to the stated formula.
Finally, since the Fourier expansion is only supported on multiples of N , we
deduce that Gk,g has a traditional q-expansion, as given. t
u
6.3.4. Prove that if g = (0, a2 ) ∈ (Z/N Z)2 , then Gk,g ∈ Mk (Γ1 (N )) for k ≥ 3.
By Exercise 6.2.1, we have
Gk,g |γ = Gk,gγ ∀γ ∈ Γ1 (N )
and we need only observe that (0, a2 )γ ≡ (0, a2 ) (mod N ), to deduce the
desired result. t
u
6.3.5. Let
E2,N (z) := E2 (z) − N E2 (N z).
Show that E2,N ∈ M2 (Γ0 (N )).
It is clear that E2,N (z) is a holomorphic function on H. Now let γ =
a b ∈ Γ0 (N ). Then
c d
az + b az + b aN z + bN
E2,N = E2 − N E2 .
cz + d cz + d cz + d
Also,
a(N z) + bN 6c(cz + d)
E2 c = (cz + d)2 E2 (N z) +
N (N z) + d N πi
so that
az + b
E2,N = (cz + d)2 (E2 (z) − N E2 (N z))
cz + d
= (cz + d)2 E2,N (z).
such that
N 0 ab rs x y
= .
0 1 cd tu 0w
In particular, some computations as in the proof of Theorem 5.2.1 easily
show that we must have
x = gcd(c, N )
c = tx
N a = ux
xw = N
d = yt + uw.
Now
ab rs x y
N E2 (N z)| = (cz + d)−2 N E2 z
cd tu 0w
rσ + s
= (cz + d)−2 N E2
tσ + u
where
x y x y
σ := z= z+ .
0w w w
Observe that
w(tσ + u) = xtz + (yt + uw) = cz + d.
Moreover, by Exercise 5.1.7,
rσ + s 6t(tσ + u)
E2 = (tσ + u)2 E2 (σ) + .
tσ + u πi
Thus,
6.3 Fourier expansions of higher level Eisenstein series 223
ab 6t(tσ + u)
N E2 (N z)| = (cz + d)−2 N (tσ + u)2 E2 (σ) +
cd πi
2
−2 (cz + d) 6t(cz + d)
= (cz + d) N E2 (σ) +
w2 wπi
N 6c
= 2 E2 (σ) + .
w πi(cz + d)
Therefore,
ab ab ab
E2,N (z)| = E2 (z)| − N E2 (N z)|
cd cd cd
x
= E2 (z) − E2 (σ).
w
Now
x ∞
y y
σ1 (n)e2πin( w z+ w )
X x
E2 (σ) = E2 z+ = 1 − 24
w w n=1
∞
X
= 1 − 24 σ1 (n)e2πiny/w q nx/w .
n=1
In particular,
we see that E2,N (z)| ac db is holomorphic at i∞. Also, observe
that if ac db ∈ Γ0 (N ), then
N |c and sox x = N and w = 1. In particular, this
shows that E2,N (z)| ac db = E2 (z) − w E2 (z + y) = E2,N (z) is modular. t
u
Let σ := ( x0 wy ) z = w
x
z + wy . Now w(tσ + u) = xtz + yt + uw = cz + d as in
the solution to Exercise 6.3.5. Thus
ab ab ab
fN (z)| = f (z)| − N f (N z)|
cd cd cd
rs
= f (z) − N (cz + d)−k f σ
tu
= f (z) − N (cz + d)−k (tσ + u)k f (σ)
k
−k cz + d
= f (z) − N (cz + d) f (σ)
w
N x y
= f (z) − k f z+ .
w w w
P∞
If we assume that f (z) = n=0 an q n is the Fourier expansion of f at i∞,
then
∞
N x y N X
f z + = an e2πiny/w q xn/w
wk w w wk n=0
is clearly holomorphic at i∞ and hence so is fN (z)| ac db . t
u
6.4.1. Show that the only elliptic points for SL2 (Z) lying in the standard funda-
mental domain F are i, ρ = eπi/3 , and ρ2 .
Suppose that z ∈ F and ±I 6= ac db ∈ SL2 (Z) are such that
az + b
z= . (6.1)
cz + d
We must have c 6= 0 since otherwise ac db = ± ( 10 1b ) is a translation which
will only fix z when b = 0. Now by (6.1) above, we have
cz 2 + (d − a)z − b = 0
Now z ∈ H and so (a+d)2 −4 < 0 which can only happen when a+d =√0, ±1.
If a + d = 0 then a = −d and z = ac ± ci ∈ F . Thus |c|1
= Im z ≥ 23 and
hence |c| = 1. Now − 12 ≤ Re z ≤ 12 implies that a = 0. Hence z = i.
√ √ √
If a + d = ±1, then Im z = 2|c|3 . But z ∈ F implies that 2|c|3 ≥ 23 which
√
3
forces |c| = 1 and Im z = 2 . Thus z = ρ or ρ2 . t
u
6.4.2. Find the stabilizer subgroups SL2 (Z)i and SL2 (Z)ρ in SL2 (Z) and show that
SL2 (Z)i / {±I} and SL2 (Z)ρ / {±I} have orders two and three, respectively.
Suppose ±I 6= ac db ∈ SL2 (Z)i . From the solution to Exercise 6.4.1, we
know that 0 = a = −d and |c| = 1. Since ad − bc = 1, wededuce that b = −c
and hence that ac db = ± 01 −1 0 . Therefore SL2 (Z)i = ±I, ± 1 0
0 −1
and
so | SL2 (Z)i / ± I| = 2.
Now suppose that ±I 6= ac db ∈ SL2 (Z)ρ . From the solution to Exer-
cise 6.4.1, we know that |a + d| = 1 and |c| = 1. We also know that
√ p
1 3 a − d ± (a + d)2 − 4
− + i= .
2 2 2c
Hence Re ρ = −1/2 = (a − d)/2c. Thus c = d − a. It is easy to determine the
four possible combinations of a, c and d which satisfy these requirements.
In each case,
b can be determined from ad − bc = 1. We may conclude that
SL2 (Z)ρ = ±I, ± 01 −1 1 1
1 , ± −1 0 . In particular, | SL2 (Z)ρ / ± I| = 3. t
u
6.4.3. If Γ has finite index in SL2 (Z), show that there are only finitely many Γ-
inequivalent elliptic points for Γ.
6.4.5. Verify that Theorem 6.4.4 reduces to Theorem 4.3.1 in the case Γ = SL2 (Z).
By Exercises 6.4.1 and 6.4.2, the only SL2 (Z)-inequivalent elliptic points
are i and ρ and the appropriate stabilizers have orders 2 and 3 respectively.
This tallies with the left hand side of the formula in Theorem 4.3.1. Since
2 = 1, 3 = 1, g = 0, and ∞ = 1, this gives us k/12, which agrees with
Theorem 4.3.1. t
u
226 6 Modular Forms of Higher Level
∞
X ∞
X
6.4.8. Let f = an q n and g = bn q n ∈ Mk (Γ) for some congruence sub-
n=0 n=0
group Γ of SL2 (Z). Suppose that an = bn for all n ≤ k[SL2 (Z) : {±}Γ]/12. Show
that f = g.
k
dim Mk (Γ) ≤ [SL2 (Z) : {±}Γ] + 1.
12
Consider the linear map Mk (Γ) → Cr+1 , where r = k[SL2 (Z) : {±}Γ]/12,
X∞
given by f = an q n 7→ (a0 , a1 , . . . , ar ). Sturm’s bound (Theorem 6.4.7)
n=0
says that this map is injective. Hence dim Mk (Γ) ≤ r + 1. t
u
1
dim M4 (Γ0 (4)) ≤ [SL2 (Z) : Γ0 (4)] + 1 = 3.
3
Now E4 , E4,2 , E4,4 ∈ M4 (Γ0 (4)) and a quick check of their q-expansions at
i∞ shows that
E4 = 1 + 240q + 2160q 2 + · · ·
E4,2 = 1 − 16q + 112q 2 + · · ·
E4,4 = 1 − 16q 2 + 112q 4 + · · ·
and these are easily verified to be linearly independent over C. Thus, the
forms are a basis for M4 (Γ0 (4)). t
u
By Exercise 6.4.9,
dim M2 (Γ0 (2)) ≤ 3/2,
so that dim M2 (Γ0 (2)) ≤ 1. Since E2,2 ∈ M2 (Γ0 (2)) by Exercise 6.3.5, and as
E2,2 6= 0, we are done. t
u
6.4.12. Show that dim M2 (Γ0 (4)) = 2 and that {E2,2 , E2,4 } is a basis for M2 (Γ0 (4)).
6.6 Supplementary problems 227
By Exercise 6.4.9,
dim M2 (Γ0 (4)) ≤ 2.
Now E2,2 and E2,4 ∈ M2 (Γ0 (4)) by Exercise 6.3.5. It is easy to check that
E2,2 = −1 − 24q + · · ·
E2,4 = −3 − 24q + · · ·
where q = e2πiz .
so that η 12 (2z) transforms with weight 6 using Exercise 6.2.5. Now Γ0 (4) has
three inequivalent cusps, 0, 1/2 and i∞. Clearly
∞
Y 12
η 12 (2z) = q 1 − q 2n = q + ···
n=1
In particular, η 12 (2z) has a zero of order 1 at the cusp 1/2. We conclude that
η 12 (2z) ∈ S6 (Γ0 (4)). t
u
F := −E2,2
E2,4 − 3E2,2
G :=
48
so that F = 1 + 24q + · · · and G = q + · · · ∈ M2 (Γ0 (4)) are linearly indepen-
dent. Now
k
F 2 −i Gi = q i + · · ·
for i = 0, 1, . . . , k2 , are clearly linearly independent.
6.6.5. Let r12
P(n) be the number of ways of writing n as a sum of 12 squares and let
∞
η 12 (2z) = n=1 an q n . Show that
n
r12 (n) = 8σ5 (n) − 512σ5 + 16an .
4
By Exercise 6.6.4 applied to k = 6, we have dim M6 (Γ0 (4)) = 4. Moreover,
comprise a basis. Now θ12 (z) ∈ M6 (Γ0 (4)) and a quick comparison of the
q-expansions immediately leads to the result. t
u
6.6.6. Let f (z) ∈ Mk (Γ1 (N )) and suppose r ∈ Z is positive. Show that f (rz) ∈
Mk (Γ1 (rN )). Moreover, if f (z) is a cusp form, then so is f (rz).
By Equation (4.2) on page 36, f (rz) = r−k/2 f (z)|( 0r 10 ). Now the cusp
conditions (holomorphy/vanishing) follow from Exercises 7.1.6 and 7.1.7. It
remains to show that r 0
a b
f (z)|( 0 1 ) is invariant under the modular transforma-
tion. For any c d ∈ Γ1 (rN ), we have
! !
r0 ab ra rb a rb r0
f (z) = f (z) = f (z) .
01 cd c d c/r d 01
230 6 Modular Forms of Higher Level
a rb
Since rN |c, we see that c/r d ∈ Γ1 (N ). Hence
!
a rb r0 r0
f (z) = f (z) .
c/r d 01 01
7.1.1. Let f , g ∈ Mk (Γ) and let F be any fundamental domain for the congruence
subgroup Γ. Show that the integral
ZZ
dxdy
y k f (z)g(z) 2 < ∞
F y
The integral is
X ZZ dxdy
y k f (z)g(z) ,
j
y2
α−1
j (D)
Since one of the f, g ∈ Sk (Γ), we see that the integral is O y k e−2πy and
© Springer Science+Business Media Singapore 2016 and Hindustan Book Agency 2016 231
M.R. Murty et al., Problems in the Theory of Modular Forms, HBA Lecture Notes
in Mathematics, DOI 10.1007/978-981-10-2651-5_17
232 7 The Petersson Inner Product
Therefore, the Petersson inner product defines a Hermitian inner product on Sk (Γ).
7.1.4. Suppose α ∈ GL+2 (Q) has integer entries. Let D = det α. If Γ(N ) ⊆ Γ for
some N , show that Γ(N D) ⊆ α−1 Γα and Γ(N D) ⊆ αΓα−1 .
Note that α is not necessarily in SL2 (Z) and so Exercise 2.2.2 is not appli-
cable. The inverse α−1 does not necessarily have integer entries.
Let γ ∈ Γ(N D). Write γ = I + N Dβ for some 2 × 2 matrix β with inte-
ger entries. We want to show γ ∈ α−1 Γα, which is equivalent to showing
αγα−1 ∈ Γ. Now
and Dα−1 has integer entries. Therefore αγα−1 ∈ Γ(N ). A similar argument
shows α−1 γα ∈ Γ(N ). t
u
Let
r t
α=
su
and suppose that δ = (r, s). Put r1 = r/δ and s1 = s/δ. Since (r1 , s1 ) = 1, we
can find integers x, y such that r1 x + s1 y = 1. Now,
x y r t δ?
=
−s1 r1 su 0?
has the desired form since the determinants of both matrices on the left hand
side of the equation are positive. t
u
7.1.6. Let f ∈ Mk (Γ) with Γ a congruence subgroup of SL2 (Z). Let α ∈ GL+
2 (Q).
Show that f |α is holomorphic at every cusp s ∈ Q ∪ {i∞}.
A typical cusp s ∈ Q∪{i∞} is of the form s = β(i∞) for some β ∈ SL2 (Z).
We must show that for every β ∈ SL2 (Z), the function (f |α)|β = f |(αβ) is
holomorphic at i∞.
From (4.2) on page 36, we see that αβ can be multiplied by a positive
scalar without affecting f |αβ. Hence we may suppose that αβ has integer
entries. By Exercise 7.1.5, we can find γ ∈ SL2 (Z) such that
ab
αβ = γ ,
0d
7.1 The Hilbert space of cusp forms 233
is holomorphic. t
u
7.1.7. Let f ∈ Sk (Γ) with Γ a congruence subgroup of SL2 (Z). Let α ∈ GL+
2 (Q).
Show that f |α vanishes at every cusp s ∈ Q ∪ {i∞}.
This solution is identical to the previous solution, except that all Fourier
expansions now start at n = 1 instead of n = 0. t
u
From (4.2) on page 36, we see that g|α0 = g|α−1 . Thus, we need only
replace g with g|α−1 in Theorem 7.1.10 to deduce the result. t
u
234 7 The Petersson Inner Product
Tp r Tp s = Tp s Tp r .
Tpr Tpt+1 = Tpr Tpt Tp − pk−1 Tpt = Tpt Tp − pk−1 Tpt Tpr = Tpt+1 Tpr . t
u
7.2.4. Let p be prime. Show that there is a polynomial Ps (x) ∈ Z[x] such that
Tps = Ps (Tp ).
Tn Tm = Tm Tn .
When pi and qj are distinct primes, Tpsi i Tqtj = Tqtj Tpsi i by Theorem 7.2.1.
j j
When pi = qj , we have Tpsi i Tqtj = Tqtj Tpsi i by Exercise 7.2.3. Hence
j j
as desired. t
u
7.4 Basis of eigenforms 235
7.3.4. Let f , g ∈ Sk (SL2 (Z)) be normalized eigenfunctions for all the Hecke oper-
ators. Show that f = g or (f, g) = 0.
P∞ P∞
Write f (z) = n=1 an q n and g(z) = n=1 bn q n . Since f and g are nor-
malized, the Fourier coefficients are the eigenvalues of the Tn ’s. Thus,
7.4.3. Show that Ek is an eigenfunction for all the Hecke operators acting on
Mk (SL2 (Z)).
Notice that
b(0) = pk−1 a(0) + a(0) = pk−1 + 1 a(0),
and so, by comparing the constant terms, we see that the only possible eigen-
value is pk−1 + 1. Thus, we need to check Tp (f ) = (pk−1 + 1)f , but this is
directly verified by comparing the Fourier coefficients of both sides and us-
ing the multiplicativity of a(n) = σk−1 (n) as follows.
If (n, p) = 1, then by (7.1)
= pk−1 + 1 a(n),
as desired. t
u
7.4.4. Show that (Ek , f ) = 0 for any f ∈ Sk (SL2 (Z)).
It suffices to prove this assertion when f is a Hecke eigenform since
Sk (SL2 (Z)) has a basis of eigenforms by Theorem 7.4.2. Observe that Theo-
rem 7.3.1 actually holds when one of the forms is a cusp form and the other
is non-vanishing at i∞. Thus we have
(j) (j)
Verify that an = 0 for n < j and aj = 1. Conclude that the fj form a basis for
Sk (SL2 (Z)).
(j)
From the q-expansion ∆j = q j + · · · , it is evident that an = 0 for n < j
(j)
and aj = 1. From this, we immediately see that the fj ’s are linearly inde-
pendent over C. Moreover, the weight of fj is 12d which equals k by Exer-
cise 4.4.3. t
u
7.5 Supplementary problems 237
7.5.2. Let k ≥ 4 be even and d = dim Sk (SL2 (Z)). Choose non-negative integers
a, b such that 12 6= 4a + 6b ≤ 14 and 4a + 6b ≡ k (mod 12). For each 1 ≤ j ≤ d,
define
∞
2(d−j)+b a
X
fj := ∆j E6 E4 = a(j) n
n q .
n=1
(j) (j)
Verify that an = 0 for n < j and = 1. Conclude that the fj form a basis for
aj
Sk (SL2 (Z)). (This basis is called the Miller basis in the literature.)
7.5.3. Let k, d, a, and b be as in Exercise 7.5.2. Verify that there exists a basis g0 ,
g1 , . . . , gd for Mk (SL2 (Z)) such that
∞
X
gi = q i + b(i)
n q
n
n=d+1
(i)
with bn ∈ Z for all n and all i. Conclude that if
∞
X
f= cn q n ∈ Mk (SL2 (Z))
n=0
= q d−1 + O q d+1 .
P Now since we have a basis for Mk (SL2 (Z)), any modular form f =
cn q n ∈ Mk (SL2 (Z)) can be expressed as a linear combination of g0 , . . . ,
gd . In fact, the weights are exactly the initial coefficients c0 , . . . , cd . If these
are all integers, then clearly f is an integer linear combination of forms with
integral coefficients. t
u
7.5.4. Let d = dim Sk (SL2 (Z)) and fix a positive integer m. Suppose
∞
X
f= cn q n ∈ Mk (SL2 (Z))
n=0
f = c0 g0 + c1 g1 + · · · + cd gd
where the gi are the basis vectors from Exercise 7.5.3. Since m | ci for i = 0,
1, . . . , d, we have that
c cd
0
f =m g0 + · · · gd ∈ mZ[[q]]
m m
as desired. t
u
7.5.5. Let f be a Hecke eigenform for the full modular group. Show that the Fourier
coefficients of f are algebraic integers.
We show that the eigenvalues of the nth Hecke operator Tn are alge-
braic integers. Indeed, let f1 , . . . , fd be the Miller basis constructed in Exer-
cise 7.5.2. The fj ’s have Z-coefficients and Tn (fj ) can be written as a Z-linear
combination of the fj ’s. Thus, the matrix representing Tn with respect to the
Miller basis has Z-coefficients. Consequently, the characteristic polynomial
of Tn is monic with coefficients from Z, so the eigenvalues of Tn are alge-
braic integers. t
u
Chapter 8
Hecke Operators of Higher Level
8.1.1. Let p be prime and Γ = SL2 (Z). Let Xp be the set of matrices with integer
entries of determinant p. Show that
p0
Xp = Γ Γ.
01
and
10 0 −1 p0 0 1
= ,
0p 1 0 01 −1 0
so that
0 −1 p0 0 1 1b p0
σ = γ1 ∈Γ Γ. t
u
1 0 01 −1 0 01 01
© Springer Science+Business Media Singapore 2016 and Hindustan Book Agency 2016 239
M.R. Murty et al., Problems in the Theory of Modular Forms, HBA Lecture Notes
in Mathematics, DOI 10.1007/978-981-10-2651-5_18
240 8 Hecke Operators of Higher Level
Xm Xn ⊆ Xmn .
8.2 Nebentypus
8.2.1. Let χ be a Dirichlet character modulo N . For each γ = ac db ∈ Γ0 (N ),
define ψ(γ) = χ(d). Show that ψ(γ1 γ2 ) = ψ(γ1 )ψ(γ2 ).
ab ef ? ?
ψ =ψ = χ(cf + dh) = χ(dh) = χ(d)χ(h),
cd gh ? cf + dh
since χ is a homomorphism. t
u
8.2.2. Show that the set of Dirichlet characters modulo N forms a group under
multiplication of order φ(N ), where φ denotes the Euler function.
cyclic of orders 1 and 2 respectively, we see that in all cases there are φ(2α )
Dirichlet characters modulo 2α . tu
8.2 Nebentypus 241
p−1
k
2 −1
X 1b
Tp (f ) = p f .
0p
b=0
In the sum defining Tn , we see from Equation (8.3) on page 100 that this
is immediate. t
u
P∞ ∞
8.2.9. Let f = n=0 an q n ∈ Mk (Γ0 (N ), χ) and then write Tp (f ) = n=0 bn q n .
P
Show that
bn = apn + χ(p)pk−1 an/p ,
where we interpret an/p = 0 when p - n.
This is clear from Exercise 8.2.6. t
u
P∞
8.2.10. Let f = m=0 λ(m)q m ∈ Mk (Γ0 (N ), χ) be a normalized eigenfunction
for all the Tn ’s. Show that for p prime,
for a ≥ 1.
This again is immediate from Exercise 8.2.6 and the solution to Exer-
cise 8.2.7. We have Tn (f ) = λ(n)f . When n = pa , we compare coefficients of
q p of both sides. On the right hand side, we have λ(pa )λ(p). On the left hand
side, we have λ(pa+1 ) + χ(p)pk−1 λ(pa−1 ). t
u
8.2.13. Show that Sk (Γ0 (N ), χ) has a basis of eigenforms for all the Tn with
(n, N ) = 1.
This is immediate from Theorem 7.4.1, Corollary 8.2.12, and the fact that
the algebra of the Tn ’s is a commutative algebra. t
u
8.3.2. Show that ∆(z) and ∆(2z) are both eigenforms for the Tn ’s (n odd) acting
on S12 (Γ0 (2)), with the same eigenvalues.
This is clear from Example 8.3.1 applied to k = 12, χ = 1 and N = 2. t
u
8.3.3. Show that ∆(z) and ∆(2z) are linearly independent in S12 (Γ0 (2)).
We only need to look at the q-expansions:
∆(z) = q − 24q 2 + · · ·
∆(2z) = q 2 − 24q 4 + · · ·
0 −1 −1
8.4.1. Let wN = N 0 . Prove that wN Γ0 (N )wN ⊂ Γ0 (N ).
d −c/N
−1
Let ac db ∈ Γ0 (N ). Then wN ac db wN = −bN a
∈ Γ0 (N ). t
u
The solution is identical to that of Exercise 8.4.2 except that now we ap-
peal to Exercise 7.1.7 to deduce that f |wN vanishes at every cusp. t
u
8.4.4. Show that if f ∈ Mk (Γ0 (N )), then (f |wN ) |wN = f . Thus |wN is, in fact,
an involution.
= −N 0
2
We compute wN 0 −N . By Equation (4.2) on page 36, we see that
2
f |wN = f. t
u
Chapter 9
Dirichlet Series and Modular Forms
9.1.1. Show that the series (9.3) on page 106 converges absolutely for Re(s) > 1.
Writing s = σ + it, with σ > 1, and t ∈ R, we have
∞ ∞
X 1 X 1
s
≤ <∞
n=1
n n=1
nσ
Y −1 Y
1 1 1
1− s = 1 + s + 2s + · · · ,
p
p p
p p
© Springer Science+Business Media Singapore 2016 and Hindustan Book Agency 2016 245
M.R. Murty et al., Problems in the Theory of Modular Forms, HBA Lecture Notes
in Mathematics, DOI 10.1007/978-981-10-2651-5_19
246 9 Dirichlet Series and Modular Forms
where the product is over distinct prime numbers p. A typical term in the
Dirichlet series obtained by expanding the product on the right hand side is
1
s.
(pa1 1 · · · pakk )
αk
Since every natural number n can be written uniquely as n = pα 1 · · · pk
1
9.1.4. Let f (n) be a multiplicative function. That is, f (mn) = f (m)f (n) for
(m, n) = 1. Assume that for any > 0, there is a constant C() such that
|f (n)| ≤ C()n . Show that for Re(s) > 1, we have
∞
f (p) f (p2 )
X f (n) Y
= 1 + + + · · · ,
n=1
ns p
ps p2s
9.1.6. Suppose that {an }∞n=1 is a sequence of complex numbers and Pf (u) is a con-
tinuously differentiable function on the interval [1, x]. Let A(t) = n≤t an . Prove
that Z x
X
an f (n) = A(x)f (x) − A(t)f 0 (t)dt.
n≤x 1
Writing
Z n+1
f (n + 1) − f (n) = f 0 (t)dt,
n
we obtain
X X Z n+1
an f (n) = A(x)f (x) − A(n) f 0 (t)dt.
n≤x n≤x−1 n
9.1 General Dirichlet series 247
Since A(t) is a step function, we see that the latter sum can be written as
X Z n+1 Z x
A(t)f 0 (t)dt = A(t)f 0 (t)dt
n≤x−1 n 1
so that
Z β
e−αz − e−βz ≤ |z| e−tz dt
α
Z β
= |z| e−tx dt
α
|z| −αx
− e−βx .
= e t
u
x
9.1.9. Suppose that X
A(x) = an = O(xδ )
n≤x
A(x)
lim =0
x→∞ xs
since A(x) = O(xδ ). Thus
∞ Z ∞
X an A(t)dt
s
= s ,
n=1
n 1 ts+1
Since 0 ≤ {t} < 1, the integral converges absolutely for Re(s) > 0. Multiply-
ing by (s − 1) and taking limits as s → 1+ , we deduce
lim (s − 1)ζ(s) = 1. t
u
s→1+
which simplifies to the expression on the right hand side upon choosing a
common denominator. This example is due to S. Ramanujan. t
u
1 − a(p)b(p)c(p)d(p)T 2
(1 − a(p)c(p)T )(1 − a(p)d(p)T )(1 − b(p)c(p)T )(1 − b(p)d(p)T )
Thus Z ∞ Z ∞
m (n) dt
t F (t)dt < ∞.
1 1 t2
Similarly,
252 9 Dirichlet Series and Modular Forms
Z 1
tm F (n) (t)dt < ∞.
−∞
Finally
Z 1
tm F (n) (t)dt < ∞
−1
Thus ∞
dm n
Z
m
x F̂ (n)
(x) ≤ (2π) n−m
(t F (t)) e−2πitx dt .
−∞ dtm
The integral is a sum of terms of the form
Z ∞
tr F (s) (t)e−2πitx dt.
−∞
Let Z ∞
2
I= e−πx dx.
−∞
Then Z ∞ Z ∞
2
+y 2 )
I2 = e−π(x dxdy.
−∞ −∞
9.2 The Poisson summation formula 253
This is equivalent to
Z ∞
2
J(u) := e−π(x+iu) dx = 1.
−∞
9.2.6. Let F ∈ S and define G(t) = F (At + B) for real constants A, B, with
A 6= 0. Show that AĜ(t) = e2πiBt/A F̂ (t/A).
2
We apply the Poisson summation formula to the function F (t) = e−π(t+a) /x
.
By Exercise 9.2.8, the Fourier transform of F is
√ 2
xe2πita e−πxt .
is entire. Since sΓ(s) = Γ(s + 1), we see that sΓ(s/2) = 2Γ( 2s + 1) is analytic
at s = 0. Therefore ζ(s) is
analytic at s = 0. Since Theorem 9.2.11 gives an
expression for π −s/2 Γ 2s ζ(s) which is valid for all s and as Γ(s) is analytic
at s = 1, we see that ζ(s) has a simple pole at s = 1 which is also evident
from Exercise 9.1.10. t
u
9.2.13. Show that ζ(s) has simple zeros at s = −2n, for each positive integer n.
Since
Z ∞
−s/2 1 1 1 1−s dt
π Γ (s/2) ζ(s) = − + (ϑ(it) − 1)(ts/2 + t 2 ) ,
s−1 s 2 1 t
Since f ∈ S, xm f (n) (x) is bounded for all x ∈ R and each choice of inte-
gers m, n ≥ 0. In particular, x2Pf (x) is bounded from which we easily deduce
the absolute convergence of n∈Z f (nx). By Exercise 9.2.6 for the function
G(t) = f (xt), we have xĜ(t) = fˆ(t/x). Applying the Poisson summation
formula to G gives X X
G(n) = Ĝ(n).
n∈Z n∈Z
as desired. t
u
9.2.16. Let f ∈ S be an even function. That is, f (x) = f (−x). For any x > 0,
show that
∞ n ∞
X X 1 ˆ
f =x fˆ(nx) + xf (0) − f (0) .
n=1
x n=1
2
Changing x to 1/x and multiplying by x/2, we obtain the desired result after
rearranging terms. t
u
R∞
9.2.17. Let f ∈ S. Set ζ(s, f ) = 0 f (x)xs dx
x . Show that the integral converges
for all s with Re(s) > 1.
256 9 Dirichlet Series and Modular Forms
ζ ? (s, f ) = ζ ? (1 − s, fˆ),
so that
∞ ∞
! !
∞
fˆ(0)
Z X X dx 1 f (0)
ζ(s)ζ(s, f ) = s
f (nx)x + fˆ(nx)x1−s − + .
1 n=1 n=1
x 2 s 1−s
The integral converges absolutely for all s ∈ C and thus defines an entire
function. Moreover, the right hand side is invariant under the maps s →
1 − s, f → fˆ from which the result follows. t
u
c
9.2.19. Show that the symbol d is completely multiplicative in each of the vari-
ables.
This follows easily from the definition and the multiplicativity of the Leg-
endre symbol. t
u
9.2.20. Let d|N and let ψ be a character modulo d. Prove that
Y ψ(p)
L(s, χ0 ψ) = L(s, ψ) 1− s .
p
p|N
9.2 The Poisson summation formula 257
We have
Y −1
χ0 (p)ψ(p)
L(s, χ0 ψ) = 1−
p
ps
Y −1
ψ(p)
= 1− s
p
p-N
Y −1 Y
ψ(p) ψ(p)
= 1− s 1− s
p p
p-d p-d
p|N
Y ψ(p)
= L(s, ψ) 1−
ps
p|N
We have
N
X m
χ(n)τ (χ) = χ(m)χ(n)e .
m=1
N
Since (n, N ) = 1, let n0 be such that nn0 ≡ 1 (mod N ) so that setting h ≡ mn0
(mod N ), we get
N
X nh
χ(n)τ (χ) = χ(h)e .
N
h=1
By Theorem 9.2.22
N
X mn
χ(n)τ (χ) = χ(m)e
m=1
N
so that
N N
X X n(m1 − m2 )
|χ(n)|2 |τ (χ)|2 = χ(m1 )χ(m2 )e
m1 =1 m2 =1
N
258 9 Dirichlet Series and Modular Forms
By Exercise 9.2.9,
∞ ∞
X 2 X 2
e−n πy+2πinα
= y −1/2 e−π(n+α) /y
.
n=−∞ n=−∞
Show that
Z ∞
−( s+1
2 )
s+1 s+1 s+1 dx
2π Γ N 2 L(s, χ) = ϑ1 (x, χ)x 2 .
2 0 x
In the formula
s Z ∞
−s/2 s/2 −s 2
πx/N s/2 dx
π Γ N n = e−n x ,
2 0 x
we change s to s + 1 to get
Z ∞
− s+1 s+1 s+1 −s 2 s+1 dx
π 2 N 2 Γ n = ne−πn x/N x 2 .
2 0 x
Thus,
∞ N
!
X X mn 2
τ (χ)ϑ1 (x, χ) = n χ(m)e e−πn x/N
.
n=−∞ m=1
N
This is equal to
N ∞
X X 2
χ(m) ne−πn x/N +2πimn/N
.
m=1 n=−∞
so that
N
3/2 X ∞
N X m −πN (n+m/N )2 /x
τ (χ)ϑ1 (x, χ) = i χ(m) n+ e
x m=1 n=−∞
N
1/2 ∞
X
iN 2
= χ(n)ne−πn /N x
x3/2 n=−∞
1/2
iN 1
= ϑ1 ,χ .
x3/2 x
as required. t
u
By Exercise 9.2.27,
Z ∞
s+1
) N s+1 s+1 dx
2π −( 2 2 L(s, χ) = θ1 (x, χ)x 2 .
0 x
This is equal to
Z ∞ Z ∞
s+1 dx 1 s+1 dx
θ1 (x, χ)x 2 + θ1 , χ x− 2
1 x 1 x x
Both integrals are absolutely convergent for all s ∈ C. The functional equa-
tion is easily deduced by noting that changing s to 1 − s, changing χ to χ,
and multiplying by wχ , the expression remains invariant. t
u
a0 b0 i k
Λf (s) + +
s k−s
and
b0 a0 ik
Λg (s) + +
s k−s
are entire and satisfy the functional equation
Λf (s) = ik Λg (k − s).
As in the proof of Theorem 9.3.1 and Theorem 9.2.1, we have the easily
verified equality:
Z ∞
b0 i k
a0 it
Λf (s) + + = f √ − a0 ts−1 dt
s k−s 1 N
Z ∞
it
+ ik g √ − b0 tk−s−1 dt.
1 N
Both integrals converge for all s ∈ C since the an ’s and bn ’s have polynomial
growth. Similarly,
Z ∞
a0 i−k
b0 it
Λg (s) + + = g √ − b0 ts−1 dt
s k−s 1 N
Z ∞
−k it
+i f √ − a0 tk−s−1 dt.
1 N
262 9 Dirichlet Series and Modular Forms
has a simple pole at s = 0 with residue Lf (0). By Exercise 9.3.2, the expres-
sion
a0 b0 i k
Λf (s) + +
s k−s
is entire. Therefore Lf (0) + a0 = 0. Similarly Lg (0) = −b0 . From the same
expression, Λf (s) has a simple pole at s = k with residue b0 ik . Thus
√ k
from which we deduce lims→k (s − k)Lf (s) = b0 2πi/ N Γ(k)−1 , as re-
quired. The result for Lg (s) is similarly derived. t
u
9.3.5. Suppose that f (s) is analytic in the vertical strip defined by a ≤ Re(s) ≤ b
and that for some α ≥ 1, α
f (s) = O e|t|
as |t| → ∞. If for some A> 0, f (s) = O |t|A on Re(s) = a and Re(s) = b, then
show that f (s) = O |t|A for all s in the vertical strip a ≤ Re(s) ≤ b.
We apply the Phragmén-Lindelöf theorem to the function g(s) = f (s)/(s−
u)A where u > b. Then g is anlytic in the strip and bounded on Re(s) = a
and Re(s) = b. By Theorem 9.3.4, g(s) is bounded throughout the vertical
strip which completes the proof. t
u
9.3.6. Show that for any σ > 0,
Z σ+i∞
1
e−x = Γ(s)x−s ds.
2πi σ−i∞
9.4 Twists of L-series attached to modular forms 263
1
Z X (−1)j
Γ(s)x−s ds = xj .
2πi CT j!
0≤j<U
The integral is
Z σ+i∞
√ !s
1 N
Γ(s)Lf (s)y −s ds
2πi σ−i∞ 2π
∞ −s !
Z σ+i∞
X 1 2πny
= an Γ(s) √ ds
n=1
2πi σ−i∞ N
X∞ √
= an e−2πny/ N
n=1
iy
=f √ − a0 . t
u
N
where ad − bc = 1.
9.4.3. Let (q, r) = 1 and (u, r) = 1. Let d and w be integers such that dr−quw = 1.
Verify the matrix identity with N = qr2 ,
1 u/r 0 −1 0 −1 r −w 1 w/r
=r .
0 1 N 0 q 0 −qu d 0 1
−1
The left hand side is uN/r
N 0
. Using dr − quw = 1, we see that the right
hand side is
2
0 −1 r 0 0 −1 r 0 qru −1
r = =
q 0 −qu 1/r q 0 −qru 1 qr2 0
fψ |wN = w(ψ)gψ
where
w(ψ) = χ(r)ψ(q)τ (ψ)2 /r.
By a simple modification of Exercise 8.4.3, we see that g ∈ Sk (Γ0 (q), χ).
Since (q, r) = 1, we have for any (u, r) = 1, integers d and w such that
dr − quw = 1. By Exercise 9.4.3, with N = qr2 , we have the identity
1 u/r 0 −1 0 −1 r −w 1 w/r
=r
0 1 N 0 q 0 −qu d 0 1
so that
1 u/r 1 w/r 1 w/r
f wN = χ(d)g = χ(r)g .
0 1 0 1 0 1
where quw ≡ −1 (mod r). As u ranges over coprime residue classes modulo
r, so does w. Since quw ≡ −1 (mod r), we have ψ(quw) = ψ(−1) so that
9.5 Supplementary problems 265
= ψ(−q)χ(r)τ (ψ)gψ .
so that
rfψ |wN = ψ(q)χ(r)τ (ψ)2 gψ
which is the desired result. t
u
We have
X a
τ (χ) = χ(a)e −
N
a (mod N )
X b
= χ(−b)e
N
b (mod N )
= χ(−1)τ (χ). t
u
Since Z ∞
1
e−vx dx = ,
0 v
for Re(v) > 0, we get
1 1 2c
F̂ (u) = + = 2 ,
c + 2πi c − 2πiu c + 4π 2 u2
as required. t
u
The left hand side is essentially a geometric series and easily summed to be
∞
X 2e−c
1+2 e−cn = 1 +
n=1
1 − e−c
2 ec + 1
=1+ = c . t
u
ec −1 e −1
9.5.4. From the previous exercise, deduce Euler’s formula that
∞
X 1 π2
2
= .
n=1
n 6
By Exercise 9.5.3,
∞
2 1 1 X 1
1+ c − 2 =2 2 + 4π 2 n2
.
e −1 2c c n=1
c
The left hand side is easily seen to be a power series in c with rational coef-
ficients. Differentiating both sides k times and then setting c = 0 gives the
desired result. t
u
From the proof of Theorem 9.2.7, we can see that for any Schwartz func-
tion F , we have X X
F (n + v) = F̂ (n)e2πinv .
n∈Z n∈Z
since 0 ≤ v < 1. The two sums above are geometric sums and the result is
1
e−cv + ecv + e−cv
ec −1
from which the result is immediate. t
u
By Exercise 9.5.6,
1 X e2πinv 1
2 cosh cv
−cv
2
+ 2 2 2
= e + c .
c c + 4π n 2c e −1
n6=0
We need to compute
e−cv
1 cosh cv
lim − 2+ .
c→0 2c c e(ec − 1)
c2 v 2 c4 v 4
cosh cv = 1 + + + ···
2! 4!
and that
1 1 c 1
= 1− + c2 + · · ·
ec − 1 c 2 12
we see that the limit in question is
−cv
1 2 c2 v 2
e 1 1 c
lim − 2 + 2 1− + c + + ···
c→0 2c c c 2 12 2
c2 v 2 v2
1 1 1
= lim 1 − cv + + ··· − + +
c→0 2c 2 2c 12 2
1 1
= v2 − v + ,
2 6
as required. t
u
Chapter 10
Special Topics
10.1.2. If two elliptic functions (for a fixed lattice L) have the same poles with the
same principal parts, show that they differ by a constant.
The difference of the two elliptic functions is again elliptic and it is clearly
holomorphic. By the previous exercise, this is constant. t
u
10.1.4. Let f be an elliptic function with respect to a lattice L. Show that there is an
α ∈ C such that f is holomorphic on the boundary of the translate α + P , where P
is the fundamental parallelogram of L.
© Springer Science+Business Media Singapore 2016 and Hindustan Book Agency 2016 269
M.R. Murty et al., Problems in the Theory of Modular Forms, HBA Lecture Notes
in Mathematics, DOI 10.1007/978-981-10-2651-5_20
270 10 Special Topics
f 0 (z)
Z
1
dz = 0.
2πi ∂(α+P ) f (z)
is absolutely convergent.
The sum is
X 1 X 1
s = s .
(mω1 + nω2 ) ω2s ω1
mω +n
(m,n)6=(0,0) (m,n)6=(0,0) 2
ω1
+ n ≥ λ m2 + n2 .
m
ω2
Thus,
X 1 1 X 1
s ≤ s/2
ω1 λ (m2 + n2 )
(m,n)6=(0,0) m ω2 +n (m,n)6=(0,0)
1 1 R (2|w| + R) −3
2 − ≤ 2 = O |ω| ,
(z − ω) ω2 2
|ω | (|ω| − R)
Prove that
∞
X
℘(z) = z −2 + (k + 1)Gk+2 z k
k=2
k even
−1
obtained easily by differentiating the power series for (1 − x) . We insert
(10.1) into the series defining ℘(z). Reversing the order of summation (which
we can do by the absolute convergence of the series) yields
∞
X
℘(z) = z −2 + (k + 1)Gk+2 z k .
k=1
and so
Differentiating we obtain
and so
Since ℘0 (z) is an odd function, ℘0 (ω1 /2) = −℘0 (−ω1 /2). But ℘0 (z) is el-
liptic and so ℘0 (−ω1 /2) = ℘0 (−ω1 /2 + ω1 ) = ℘0 (ω1 /2). Thus ℘0 (ω1 /2) = 0.
Similarly, ℘0 (ω2/2) = 0and ℘0 ((ω1 + ω2 )/2) = 0. By Exercise 10.1.13, we de-
duce that ℘ ω21 , ℘ ω22 , and ℘ ω1 +ω 2
2
are the roots of the associated cubic.
t
u
10.1.17. Suppose that ω1 , ω2 are linearly independent over R and that L is the
latticespanned by ω1 , ω2 . Let ℘(z) be the associated ℘-function. Show that ℘ ω21 ,
Since (by the solution of the previous exercise) ℘0 (ω1 /2) = 0 , we see that
f1 (z) has a double order zero at z = ω1 /2. As ℘(z) has a double order pole at
z = 0 and as there are no other poles in the fundamental parallelogram, this
accounts for all of the zeros of f1 (z) by Exercise 10.1.6. If ℘(ω1 /2) = ℘(ω2 /2),
then z = ω2 /2 would be a zero of f1 (z) and hence ω1 ≡ ω2 (mod L), which
contradicts the linear independence of ω1 and ω2 over R. Thus, ℘(ω1 /2) and
℘(ω2 /2) are distinct.
A similar argument applies for the other 2-division points, namely ℘(ω2 /2)
and ℘((ω1 + ω2 )/2). t
u
ω1 = eω10 + f ω20
ω2 = gω10 + hω20 .
Thus 0
ω1 ef ω1 ef ab ω1
= =
ω2 gh ω20 gh cd ω2
and because ω1 , ω2 are linearly independent,
ef ab 10
=
gh cd 01
Similarly,
az + b
E60 (cz + d)−2 = 6(cz + d)5 cE6 (z) + (cz + d)6 E60 (z).
cz + d
Thus,
az + b az + b
6c(cz + d)7 E6 (z) + (cz + d)8 E60 (z)
f = 2E4
cz + d cz + d
az + b
4c(cz + d)5 E4 (z) + (cz + d)6 E40 (z)
− 3E6
cz + d
= (cz + d)12 f (z).
To see that it is a cusp form, we check its behaviour at i∞ and see that from
the q-expansion that it vanishes, since E40 (z) and E60 (z) both vanish there.
t
u
Hence
−1728∆2 j 0 = E42 E6 (2E4 E60 − 3E6 E40 ) .
By Exercise 10.2.7, E42 E6 = E14 . By Exercise 10.2.6, the term in the bracket is
a cusp form of weight 12. By considering the q-expansion, this cusp form is
easily seen to be 2πi(1728)∆ from which the result follows. t
u
Pn
Thus, f ? = j=1 f ? (ωj )ωj? and so L? ⊆ ω1? Z + · · · ωn? Z. t
u
√
10.3.3. Prove that µ(V /L) = det A.
t 2
so that if B = (bij ), the Gram matrix A is √ equal to B B. Thus (det B) =
det A and as det A > 0, we have det B = det A. At the same time, having
fixed our own basis e1 , . . . , en , the lattice spanned by ω1 , . . . , ωn is that lattice
spanned
nP by the columns of o B and the fundamental domain for this lattice
n
is j=1 λj ej : 0 ≤ λj ≤ 1 which has measure µ(V /L). Since this measure
is the volume of the parallelepiped spanned by the columns of B, we are
done. t
u
10.3.4. With A as above, show that the matrix A−1 is the Gram matrix of the dual
lattice L? .
Recall that 11
0 −1
the full modular group SL2 (Z) is generated by T = ( 0 1 ) and
S = 1 0 by Theorem 2.1.2. We first consider how θL (z) transforms under
T . Since L is even, we have
X X
θL (z + 1) = eπi(z+1)(ω,ω) = eπiz(ω,ω) = θL (z),
ω∈L ω∈L
10.3.8. If L is self-dual show that the Gram matrix of L and L? have determinant
1.
Since L = L? , µ(V /L) = µ(V /L? ) so that det A = det A−1 by Exer-
cise 10.3.4. Hence det A = ±1. Since A is positive definite, det A = 1. t
u
10.4 Special values of zeta and L-functions 279
10.3.9. Show that the theta function of a self-dual, even lattice is a modular form of
weight n/2 for the full modular group SL2 (Z).
From the solution to Exercise 10.3.7, we have that µ(V /L) = 1 and that
θL (z + 1) = θL (z). Now since 8 | n by Exercise 10.3.7, the result follows
immediately from Theorem 10.3.6,
1
θL − = z n/2 θL (z),
z
dk k!
− (π cot πz) = ψk (z) + (−1)k+1 ψk (−z) + (−1)k k+1
dz k z
From the partial fraction expansion of the cotangent function (see (4.3) in
Section 4.2), we have
∞
1 X 1 1
π cot πz = + + .
z n=1 z−n z+n
We may write
280 10 Special Topics
N
X X N
X ∞
X
L(k, χ) = χ(a) n−k = χ(a) (N j + a)−k
a=1 n≡a(mod N ) a=1 j=0
10.4.5. If χ is an even character (that is, χ(−1) = 1) and k is even, show that
N
X dk
−2(k − 1)!L(k, χ) = (−N )−k χ(a) (π cot πz) .
a=1
dz k z=a/N
which is
N
X
= (−N )−k χ(a)(ψk−1 (a/N ) + ψk−1 (1 − a/N ).
a=1
By exercise 10.4.3, the result follows. The last assertion is immediate upon
noting that
dk
(π cot πz)
dz k z=a/N
is an algebraic multiple of π k . t
u
10.4.6. If χ is an odd character (that is, χ(−1) = −1) and k is odd, show that
N
X dk
−2(k − 1)!L(k, χ) = (−N )−k χ(a) (π cot πz) .
a=1
dz k z=a/N
By exercise 10.4.3, the result follows. The last assertion is also immediate
upon noting (as before) that
dk
(π cot πz)
dz k z=a/N
is an algebraic multiple of π k . t
u
wf 0 (w)
Z
1
dw.
2πi ∂P f (w)
γ : [0, 1] → C
given by
γ(t) = g (1 − t)a + tb
282 10 Special Topics
10.5.2. With ℘ denoting the Weierstrass ℘-function as in Section 10.1, prove that
The result is immediate upon using Exercise 10.5.2 and comparing coef-
ficients of the Laurent expansions about z = 0 for the differential equation
satisfied by ℘00 (z). t
u
10.5.4. Show that for k ≥ 2, each G2k can be expressed as a polynomial in G4 and
G6 with positive rational coefficients.
Show that
∞
Y
ϑ(z) = (1 − e2πinz )(1 + e(2n−1)πiz )2 .
n=1
10.5.6. With η being the Dedekind η-function and ϑ as in the previous exercise,
show that
ϑ(z) = η 2 ((z + 1)/2)/η(z + 1).
but that j 0 (z) is not a modular form of weight 2 for the full modular group.
By Exercise 4.6.9 (or by direct verification), the first assertion is clear. j 0 (z)
is not a modular form since it is not holomorphic at infinity. t
u
A short guide for further reading
The readers will benefit much from reading Hardy’s lectures [13] and espe-
cially the chapter relating to Ramanujan’s work on the τ -function. Clearly,
this inspired Rankin’s doctoral thesis and his subsequent work on the topic.
A good introduction to the “q-world” discussed in Chapter 1 can be found
in [7].
The gentlest and most elegant introduction to the topic of modular forms
is the last section of Serre’s Course in Arithmetic [36]. There the student will
find the rudimentary notions needed to enter the topic in the case of the
full modular group. Another reasonably gentle introduction is [2]. Perhaps
the earliest books on the subject were Gunning [12] and Ogg [27]. Both of
these books are of reasonable size that the graduate student will not have
difficulty in studying the material in one semester. Zagier’s chapter in [5] is
also a short introduction to some of the modern themes.
For an amplified study from an analytic and classical perspective, Rankin’s
book [31] is an excellent source of material. Schoenberg’s book [35] also
gives a classical treatment that is accessible. A more amenable treatment
is Iwaniec [16] which assumes a bit more sophistication with analytic num-
ber theory. Lang’s monograph [20] blends some of the analytic and algebraic
themes and explores the theory of congruences from the standpoint of Ga-
lois representations. Miyake’s book [22] is also detailed and reasonably self-
contained for the advanced student. Sarnak [34] highlights some analytic
applications of the theory of modular forms but assumes the reader has a
modest understanding of the subject. Koblitz’s book [19] is also readable
and has the advantage of giving a quick introduction to the topic of half-
integral weight modular forms, a topic that was untouched in this book.
A geometric approach to the subject of modular forms is highlighted in
[9] as well as the online notes of Milne [21]. For the study of Hilbert mod-
ular forms, we recommend Garrett [10] as well as Bruinier’s chapter in [5].
For the topic of Siegel modular forms, a topic which we have not touched
in this book, we suggest the reader begin with van der Geer’s chapter [5].
Also, Siegel’s classic lectures notes from the Tata Institute [39] are an excel-
© Springer Science+Business Media Singapore 2016 and Hindustan Book Agency 2016 285
M.R. Murty et al., Problems in the Theory of Modular Forms, HBA Lecture Notes
in Mathematics, DOI 10.1007/978-981-10-2651-5
286 A short guide for further reading
References
1. G.E. Andrews, A simple proof of Jacobi’s triple product identity. Proc. Amer. Math.
Soc., 16 (1965), 333-334.
2. T. Apostol, Modular functions and Dirichlet series in number theory, Springer-
Verlag, 1976.
3. R. Askey, Ramanujan’s extensions of the Gamma and Beta functions. Amer. Math.
Monthly, 87 (1980), 346-359.
4. B. C. Berndt. Ramanujan’s congruences for the partition function modulo 5, 7, and
11. Int. J. Number Theory, 3 (2007), 349–354.
5. J. H. Bruinier, G. van der Geer, G. Harder, and D. Zagier, 1-2-3 of Modular Forms,
Springer, 2008.
6. J. H. Bruinier, K. Ono, and R. C. Rhoades. Differential operators for harmonic weak
Maass forms and the vanishing of Hecke eigenvalues. Math. Ann., 342(3):673–693,
2008.
7. Hei-Chi Chan, An invitation to q-series, World Scientific, 2011.
8. P. Deligne, La conjecture de Weil: I, Publications Mathématiques de l’IHES, 43
(1974), 273-307.
9. F. Diamond and J. Shurman. A first course in modular forms, Volume 228, Graduate
Texts in Mathematics, Springer-Verlag, New York, 2005.
10. P. Garrett, Holomorphic Hilbert modular forms, Wadsworth and Brooks/Cole
Mathematics Series, Chapman and Hall, CRC, 1989.
11. E. Grosswald, Representations of integers as sums of squares, Springer-Verlag,
New York, 1985.
12. R. Gunning, Lectures on modular forms, Princeton University Press, 1962.
13. G. H. Hardy. Ramanujan. Twelve lectures on subjects suggested by his life and
work, Cambridge University Press, Cambridge, England, 1940.
14. E. Hecke, Analytsche Arithmetik der Positiven Quadratischen Formen in Mathe-
matische Werke, pp. 789-918.
15. M. D. Hirschhorn, A simple proof of Jacobi’s two square theorem, Amer. Math.
Monthly, 92 (1985), 579-80.
16. H. Iwaniec, Topics in classical automorphic forms, Graduate Studies in Mathemat-
ics, Volume 17, American Mathematical Society,1997.
17. H. Iwaniec, Spectral methods in automorphic forms, Graduate Studies in Mathe-
matics, Volume 53, American Mathematical Society, 2002.
18. L. Kilford, Modular forms: a classical and computational introduction, Imperial
College Press, 2008.
19. N. Koblitz, Introduction to elliptic curves and modular forms, Springer, 1984.
20. S. Lang, Introduction to modular forms, Springer-Verlag, 1976.
21. J. Milne, Modular functions and modular forms, online course notes available at
http://www.jmilne.org/math/CourseNotes/mf.html.
22. T. Miyake, Modular forms, Springer-Verlag, Berlin, New York, 1989, 335 pp.
23. L.J. Mordell, On Mr. Ramanujan’s empirical expansions of modular functions, Pro-
ceedings of the Cambridge Philosophical Society, 19 (1917), 117-124.
References 287
24. M. R. Murty. Problems in analytic number theory, Volume 206, Graduate Texts in
Mathematics, Springer, New York, second edition, 2008. Readings in Mathematics.
25. M.R. Murty and J. Esmonde, Problems in algebraic number theory, Volume 190,
Graduate Texts in Mathematics, 2nd edition, Springer, 2005.
26. V. Kumar Murty, Ramanujan and Harish-Chandra, Math. Intelligencer, 15 (1993),
no. 2, 33-39.
27. A.P. Ogg, Modular forms and Dirichlet series, Benjamin, 1969.
28. H. Petersson. Über die Entwicklungskoeffizienten der automorphen Formen. Acta
Math., 58 (1932), 169–215.
29. S. Ramanujan, On certain arithmetical functions, Trans. Cambridge Philos. Soc., 22
(1916), 159-184.
30. R. A. Rankin, The modular group and its subgroups, The Ramanujan Institute,
Madras, 1969.
31. R.A. Rankin, Modular forms, Ellis Horwood, 1984.
32. M. Razar, Modular forms for G0 (N ) and Dirichlet series, Trans. Amer. Math. Soc.,
231 (1977), 489-495.
33. W. Rudin. Real and complex analysis, McGraw-Hill Book Co., New York, third
edition, 1987.
34. P. Sarnak, Some applications of modular forms, Cambridge University Press, 1990.
35. B. Schoenberg, Elliptic modular functions, Springer 1974.
36. J.-P. Serre, A course in arithmetic, Springer, 1970.
37. G. Shimura, Introduction to the Arithmetic Theory of Automorphic Functions,
Princeton University Press, 1971.
38. G. Shimura, Modular forms and beyond, Springer 2011.
39. C.L. Siegel, Advanced Analytic Number Theory, Tata Institute of Fundamental Re-
search, Bombay, 1980.
40. J. Silverman, The arithmetic of elliptic curves, Springer 1986.
41. J. Silverman, Advanced topics in the arithmetic of elliptic curves, Springer 1994.
42. M. Spivak, Calculus on manifolds, A modern approach to classical theorems of
advanced calculus, W. A. Benjamin, Inc., New York-Amsterdam, 1965.
43. W. Stein, Modular forms, a computational approach, Graduate Studies in Mathe-
matics, Volume 79, American Mathematical Society,2007.
44. B. Sury, The Congruence Subgroup Problem, Vol. 24, Texts and Readings in Math-
ematics, Hindustan Book Agency, India, 2003.
45. H. P. F. Swinnerton-Dyer. Congruence properties of τ (n). in Ramanujan revisited
(Urbana-Champaign, Ill., 1987), pages 289–311. Academic Press, Boston, MA, 1988.
46. A. Weil, Über die Bestimmung Dirichletscher Reihen durch Funkionalgleichungen,
Math. Annalen, 168 (1967), 149-156.
47. X. Wang and D. Pei, Modular forms with integral and half-integral weights,
Springer 2012.
Index
© Springer Science+Business Media Singapore 2016 and Hindustan Book Agency 2016 289
M.R. Murty et al., Problems in the Theory of Modular Forms, HBA Lecture Notes
in Mathematics, DOI 10.1007/978-981-10-2651-5
290 Index