Vous êtes sur la page 1sur 216

Graduate Texts in Mathematics 41

Editorial Board
J.H. Ewing F.W. Gehring P.R. Halmos
Graduate Texts in Mathematics

1 TAKEUTIIZARING. Introduction to Axiomatic Set Theory. 2nd ed.


2 OXTOBY. Measure and Category. 2nd ed.
3 SCHAEFFER. Topological Vector Spaces.
4 HILTON/STAMMBACH. A Course in Homological Algebra.
5 MAC LANE. Categories for the Working Mathematician.
6 HUGHES/PIPER. Projective Planes.
7 SERRE. A Course in Arithmetic.
8 TAKEUTIlZARING. Axiomatic Set Theory.
9 HUMPHREYS. Introduction to Lie Algebras and Representation Theory.
10 COHEN. A Course in Simple Homotopy Theory.
11 CONWAY. Functions of One Complex Variable. 2nd ed.
12 BEALS. Advanced Mathematical Analysis.
13 ANDERSON/FuLLER. Rings and Categories of Modules.
14 GOLUBITSKy/GUILLEMIN. Stable Mappings and Their Singularities.
15 BERBERIAN. Lectures in Functional Analysis and Operator Theory.
16 WINTER. The Structure of Fields.
17 ROSENBLATT. Random Processes. 2nd ed.
18 HALMOS. Measure Theory.
19 HALMOS. A Hilbert Space Problem Book. 2nd ed., revised.
20 HUSEMOLLER. Fibre Bundles. 2nd ed.
21 HUMPHREYS. Linear Algebraic Groups.
22 BARNES/MACK. An Algebraic Introduction to Mathematical Logic.
23 GREUB. Linear Algebra. 4th ed.
24 HOLMES. Geometric Functional Analysis and its Applications.
25 HEWITT/STROMBERG. Real and Abstract Analysis.
26 MANES. Algebraic Theories.
27 KELLEY. General Topology.
28 ZARISKIISAMUEL. Commutative Algebra. Vol. I.
29 ZARISKIISAMUEL. Commutative Algebra. Vol. II.
30 JACOBSON. Lectures in Abstract Algebra I: Basic Concepts.
31 JACOBSON. Lectures in Abstract Algebra II: Linear Algebra.
32 JACOBSON. Lectures in Abstract Algebra III: Theory of Fields and Galois Theory.
33 HIRSCH. Differential Topology.
34 SPITZER. Principles of Random Walk. 2nd ed.
35 WERMER. Banach Algebras and Several Complex Variables. 2nd ed.
36 KELLEy/NAMIOKA et al. Linear Topological Spaces.
37 MONK. Mathematical Logic.
38 GRAUERT/FRITZSCHE. Several Complex Variables.
39 ARVESON. An Invitation to C*-Algebras.
40 KEMENy/SNELL/KNAPP. Denumerable Markov Chains. 2nd ed.
41 ApoSTOL. Modular Functions and Dirichlet Series in Number Theory. 2nd ed.
42 SERRE. Linear Representations of Finite Groups.
43 GILLMAN/JERISON. Rings of Continuous Functions.
44 KENDIG. Elementary Algebraic Geometry.
45 LOEVE. Probability Theory I. 4th ed.
46 LOEVE. Probability Theory II. 4th ed.
47 MOISE. Geometric Topology in Dimensions 2 and 3.

continued after Indet


Tom M. Apostol

Modular Functions
and Dirichlet Series
in Number Theory
Second Edition

With 25 Illustrations

Springer-Verlag
New York Berlin Heidelberg
London Paris Tokyo Hong Kong
Tom M. Apostol
Department of Mathematics
California Institute of Technology
Pasadena, CA 91125
USA

Editorial Board
J.H. Ewing F.W. Gehring P.R. Halmos
Department of Department of Department of
Mathematics Mathematics Mathematics
Indiana University University of Michigan Santa Clara University
Bloomington, IN 47405 Ann Arbor, MI 48109 Santa Clara, CA 95053
USA USA USA

AMS Subject Classifications


IOA20, IOA45, 10045, IOH05, 10HIO, IOJ20, 30AI6

Library of Congress Cataloging-in-Publication Data


Apostol, Tom M.
Modular functions and Dirichlet series in number theory/Tom M. Apostol.-2nd ed.
p. cm.-(Graduate texts in mathematics; 41)
Includes bibliographical references.
ISBN 0-387-97127-0 (alk. paper)
I. Number theory. 2. Functions, Elliptic. 3. Functions, Modular. 4. Series,
Dirichlet. I. Title. II. Series.
QA241.A62 1990
512'.7-dc20 89-21760

Printed on acid-free paper.

© 1976, 1990 Springer-Verlag New York, Inc.


All rights reserved. This work may not be translated or copied in whole or in part without the
written permission of the publisher (Springer-Verlag New York, Inc., 17,:, Fifth Avenue, New
York, NY 10010, USA), except for brief excerpts in connection with reviews or scholarly
analysis. Use in connection with any form of information and retrieval, electronic adaptation,
computer software, or by sirrlilar or dissimilar methodology now known or hereafter developed
is forbidden.
The use of general descriptive names, trade names, trademarks. etc., in this publication, even if
the former are not especially identified, is not to be taken as a sign that such names, as
understood by the Trade Marks and Merchandise Marks Act, may accordingly be used freely
by anyone.

Typeset by Asco Trade Typesetting Ltd., Hong Kong.


Printed and bound by Edwards Brothers, Inc., Ann Arbor, Michigan.
Printed in the United States of America.

9 8 7 6 5 432 I

ISBN 0-387-97 J27-0 Springer-Verlag New York Berlin Heidelberg


ISBN 3-540-97127-0 Springer-Verlag Berlin Heidelberg New York
Preface

This is the second volume of a 2-volume textbook* which evolved from a


course (Mathematics 160) offered at the California Institute of Technology
during the last 25 years.
The second volume presupposes a background in number theory com-
parable to that provided in the first volume, together with a knowledge of
the basic concepts of complex analysis.
Most of the present volume is devoted to elliptic functions and modular
functions with some of their number-theoretic applications. Among the
major topics treated are Rademacher's convergent series for the partition
function, Lehner's congruences for the Fourier coefficients of the modular
functionj(r), and Heeke's theory of entire forms with multiplicative Fourier
coefficients. The last chapter gives an account of Bohr's theory of equivalence
of general Dirichlet series.
Both volumes of this work emphasize classical aspects of a subject which
in recent years has undergone a great deal of modern development. It is
hoped that these volumes will help the nonspecialist become acquainted
with an important and fascinating part of mathematics and, at the same
time, will provide some of the background that belongs to the repertory of
every specialist in the field.
This volume, like the first, is dedicated to the students who have taken
this course and have gone on to make notable contributions to number
theory and other parts of mathematics.

T.M.A.
January, 1976
* The first volume is in the Springer-Verlag series Undergraduate Texts in Mathematics under
the title Introduction to Analytic Number Theory.

v
Preface to the Second Edition

The major change is an alternate treatment of the transformation formula for


the Dedekind eta function, which appears in a five-page supplement to Chap-
ter 3, inserted at the end of the book (just before the Bibliography). Other-
wise, the second edition is almost identical to the first. Misprints have been
repaired, there are minor changes in the Exercises, and the Bibliography has
been updated.

T.M. A.
July, 1989
Contents

Chapter 1
Elliptic functions
1.1 Introduction 1
1.2 Doubly periodic functions 1
1.3 Fundamental pairs of periods 2
1.4 Elliptic functions 4
1.5 Construction of elliptic functions 6
1.6 The Weierstrass g;J function 9
1.7 The Laurent expansion of g;J near the origin 11
1.8 Differential equation satisfied by g;J 11
1.9 The Eisenstein series and the invariants 9z and 93 12
1.10 The numbers e b ez, e3 13
1.11 The discriminant .1 14
1.12 Klein's modular function J(T) 15
1.13 Invariance of J under unimodular transformations 16
1.14 The Fourier expansions of 9z(T) and 93(T) 18
1.15 The Fourier expansions of .1(T) and J(1:) 20
Exercises for Chapter 1 23

Chapter 2
The Modular group and modular functions
2.1 Mobius transformations 26
2.2 The modular group r 28
2.3 Fundamental regions 30
2.4 Modular functions 34
vii
2.5 Special values of J 39
2.6 Modular functions as rational functions of J 40
2.7 Mapping properties of J 40
2.8 Application to the inversion problem for Eisenstein series 42
2.9 Application to Picard's theorem 43
Exercises for Chapter 2 44

Chapter 3
The Dedekind eta function
3.1 Introduction 47
3.2 Siegel's proof of Theorem 3.1 48
3.3 Infinite product representation for .1( r) 50
3.4 The general functional equation for YJ(r) 51
3.5 Iseki's transformation formula 53
3.6 Deduction of Dedekind's functional equation from Iseki's
formula 58
3.7 Properties of Dedekind sums 61
~

3.8 The reciprocity law for Dedekind sums 62


3.9 Congruence properties of Dedekind sums 64
3.10 The Eisenstein series G 2 (r) 69
Exercises for Chapter 3 70

Chapter 4
Congruences for the coefficients of the modular function j
4.1 Introduction 74
4.2 The subgroup r o(q) 75
4.3 Fundamental region of r o(p) 76
4.4 Functions automorphic under the subgroup r o(P) 78
4.5 Construction of functions belonging to r o(P) 80
4.6 The behavior of fp under the generators of r 83
4.7 The function <p(r) = .1(qr)j.1(r) 84
4.8 The univalent function <I>(r) 86
4.9 Invariance of <I>(r) under transformations of r o(q) 87
4.10 The function j p expressed as a polynomial in <I> 88
Exercises for Chapter 4 91

Chapter 5
Rademacher's series for the partition function
5.1 Introduction 94
5.2 The plan of the proof 95
5.3 Dedekind's functional equation expressed in terms of F 96
5.4 Farey fractions 97

viii
5.5 Ford circles 99
5.6 Rademacher's path of integration 102
5.7 Rademacher's convergent series for pen) 104
Exercises for Chapter 5 110

Chapter 6
Modular forms with multiplicative coefficients
6.1 Introduction 113
6.2 Modular forms of weight k 114
6.3 The weight formula for zeros of an entire modular form 115
6.4 Representation of entire forms in terms of G4 and G 6 117
6.5 The linear space M k and the subspace Mk,o 118
6.6 Classification of entire forms in terms of their zeros 119
6.7 The Hecke operators Tn 120
6.8 Transformations of order n 122
6.9 Behavior of Tnfunder the modular group 125
6.1 0 Multiplicative property of Hecke operators 126
6.11 Eigenfunctions of Hecke operators 129
6.12 Properties of simultaneous eigenforms 130
6.13 Examples of normalized simultaneous eigenforms 131
6.14 Remarks on existence of simultaneous ~igenforms in M 2k, 0 133
6.15 Estimates for the Fourier coefficients of entire forms 134
6.16 Modular forms and Dirichlet series 136
Exercises for Chapter 6 138

Chapter 7
Kronecker's theorem with applications
7.1 Approximating real numbers by rational numbers 142
7.2 Dirichlet's approximation theor~m 143
7.3 Liouville's approximation theorem 146
7.4 Kronecker's approximation theorem: the one-dimensional
case 148
7.5 Extension of Kronecker's theorem to simultaneous
approximation 149
7.6 Applications to the Riemann zeta function 155
7.7 Applications to periodic functions 157
Exercisesfor Chapter 7 159

Chapter 8
General Dirichlet series and Bohr's equivalence theorem
8.1 Introduction 161
8.2 The half-plane of convergence of general Dirichlet series 161
8.3 Bases for the sequence of exponents of a Dirichlet series 166

ix
8.4 Bohr matrices 167
8.5 The Bohr function associated with a Dirichlet series 168
8.6 The set of values taken by a Dirichlet series f(s) on a line
U = Uo 170
8.7 Equivalence of general Dirichlet series 173
8.8 Equivalence of ordinary Dirichlet series 174
8.9 Equality of the sets Uf(uo) and Ug(u o) for equivalent
Dirichlet series 176
8.10 The set of values taken by a Dirichlet series in a neighborhood
of the line U = Uo 176
8.11 Bohr's equivalence theorem 178
8.12 Proof of Theorem 8.15 179
8.13 Examples of equivalent Dirichlet series. Applications of Bohr's
theorem to L-series 184
8.14 Applications of Bohr's theorem to the Riemann zeta function 184
Exercisesfor Chapter 8 187

Supplement to Chapter 3 190

Bibliography 196

Index of special symbols 199

Index 201

x
Elliptic functions

1.1 Introduction
Additive number theory is concerned with expressing an integer n as a sum
of integers from some given set S. For example, S might consist of primes,
squares, cubes, or other special numbers. We ask whether or not a given
number can be expressed as a sum of elements of S and, if so, in how many
ways this can be done.
Letf(n) denote the number of ways n can be written as a sum of elements
of S. We ask for various properties of f(n), such as its asymptotic behavior
for large n. In a later chapter we will determine the asymptotic value of the
partition function p(n) which counts the number of ways n can be written as a
sum of positive integers ~ n.
The partition function p(n) and other functions of additive number theory
are intimately related to a class of functions in complex analysis called
elliptic modular functions. They playa role in additive number theory analo-
gous to that played by Dirichlet series in multiplicative number theory. The
first three chapters of this volume provide an introduction to the theory of
elliptic modular functions. Applications to the partition function are given
in Chapter 5.
We begin with a study of doubly periodic functions.

1.2 Doubly periodic functions


A function j' of a complex variable is called periodic with period w if
f(z + w) = f(z)
whenever z and z + ware in the domain off If w is a period, so is nw for
every integer n. If W1 and W2 are periods, so is mW 1 + nW2 for every choice of
integers m and n.
1: Elliptic functions

Definition. A function f is called doubly periodic if it has two periods W1


and W2 whose ratio w 21w 1 is not real.

We require that the ratio be nonreal to avoid degenerate cases. For


example, if W 1 and W 2 are periods whose ratio is real and rational it is easy
to show that each of W 1 and W 2 is an integer multiple of the same period. In
fact, if w21w1 = alb, where a and b are relatively prime integers, then there
exist integers m and n such that mb + na = 1. Let w = mW 1 + nw 2 . Then
w is a period and we have

so W1 = bw and W2 = aw. Thus both W1 and W2 are integer multiples of W.


If the ratio w 21w 1 is real and irrational it can be shown thatfhas arbitrarily
small periods (see Theorem 7.12). A function with arbitrarily small periods
is constant on every open connected set on which it is analytic. In fact, at
each point of analyticity offwe have

f'(z) = lim f(z + zn) - f(z),


z"~o Zn

where {zn} is any sequence of nonzero complex numbers tending to O. If f


has arbitrarily small periods we can choose {zn} to be a sequence of periods
tending to O. Then j'(z + zn) = j'(z) and hence j"(z) = O. In other words,
f'(z) = 0 at each point of analyticity off, hence fmust be constant on every
open connected set in whichfis analytic.

1.3 Fundamental pairs of periods


Definition. Let JO have periods w 1 , W2 whose ratio w 21w 1 is not real. The
pair (w 1 , w 2) is called afundamental pair if every period of f is of the form
mW 1 + nw 2 , where m and n are integers.

Every fundamental pair of periods W1' W 2 determines a network of


parallelograms which form a tiling of the plane. These are called period
parallelograms. An example is shown in Figure l.la. The vertices are the
periods w = mW 1 + nw 2 . It is customary to consider two intersecting edges
and their point of intersection as the only boundary points belonging to the
period parallelogram, as shown in Figure l.lb.

Notation. If W1 and W2 are two complex numbers whose ratio is not real
we denote by 0(w 1 , W2), or simply by 0, the set of all linear combinations
mW 1 + nw 2 , where m and n are arbitrary integers. This is called the lattice
generated by W1 and W2'

2
1.3: Fundamental pairs of periods

(a) (b)

Figure 1.1

Theorem 1.1. I}' (WI' W 2 ) is a .fundamental pair oj' periods, then the triangle
with vertices 0, WI' W2 contains no further periods in its interior or on its
boundary. Conversely, any pair of periods with this property isfundamental.
PROOF. Consider the parallelogram with vertices 0, W b WI + W 2 , and W2'
shown in Figure 1.2a. The points inside or on the boundary of this parallel-
ogram have the form

Z = rJ..W 1 + f3W2'
where °:: ; °:: ;
rJ.. :::; 1 and 13 :::; 1. Among these points the only periods are 0,
and WI + W2' so the triangle with vertices 0, W b W 2 contains no
WI' W 2 ,
periods other than the vertices.

o o
(a) (b)

Figure 1.2

3
1: Elliptic functions

Conversely, suppose the triangle 0, w 1 , W2 contains no periods other


than the vertices, and let W be any period. We are to show that W = mW 1 +
nW 2 for some integers m and n. Since W2/W 1 is nonreal the numbers W1 and
W 2 are linearly independent over the real numbers, hence

where t 1 and t 2 are real. Now let [t] denote the greatest integer stand
write

Then
/
W - [t 1 ]W 1 - [t 2]W 2 = rl w l + r2w2.
If one of rl or r2 is nonzero, then rl W1 + r2 W2 will be a period lying inside
the parallelogram with vertices 0, Wb W2' W1 + W2 . But if a period w lies
inside this parallelogram then either w or W 1 + W 2 - w will lie inside the
triangle 0, W1 , W2 or on the diagonal joining W1 and W2, contradicting the
hypothesis. (See Figure 1.2b.) Therefore rl = r2 = 0 and the proof is
complete. D

Definition. Two pairs of complex numbers (w 1 , ( 2) and (w 1 ', w 2'), each with
nonreal ratio, are called equivalent if they generate the same lattice of
periods; that is, ifQ(w b (2) = Q(w 1 ', w 2 ').

The next theorem, whose proof is left as an exercise for the reader,
describes a fundamental relation between equivalent pairs of periods.

Theorem 1.2. Two pairs (w b ( 2) and (w 1 ', w 2') are equivalent if, and only if,
there is a 2 x 2 matrix (: ~) with integer entries and determinant
ad - be = ± 1, such that

or, in other words,


W2' = aW 2 + bw 1 ,
w 1 ' = eW 2 + dw 1 •

1.4 Elliptic functions


Definition. A function f is called elliptic if it has the following two properties:
(a) f is doubly periodic.
(b) f is meromorphic (its only singularities in the finite plane are poles).

4
1.4: Elliptic functions

Constant functions are trivial examples of elliptic functions. Later we


shall give examples of nonconstant elliptic functions, but first we derive some
fundamental properties common to all elliptic functions.

Theorem 1.3. A nonconstant ellipticfunction has afundamental pair of periods.


PROOF. Iff is elliptic the set of points where f is analytic is an open connected
set. Also, f has two periods with nonreal ratio. Among all the nonzero
periods of j' there is at least one whose distance from the origin is minimal
(otherwisefwould have arbitrarily small nonzero periods and hence would
be constant). Let W be one of the nonzero periods nearest the origin. Among
all the periods with modulus IW I choose the one with smallest nonnegative
argument and call it W l . (Again, such a period must exist otherwise there
would be arbitrarily small nonzero periods.) If there are other periods
with modulus Iwll besides W l and -W l , choose the one with smallest
argument greater than that of W l and call this W2. If not, find the next
larger circle containing periods =1= nW l and choose that one of smallest
nonnegative argument. Such a period exists since f has two noncollinear
periods. Calling this one W2 we have, by construction, no periods in the
triangle 0, W l , W 2 other than the vertices, hence the pair (0)1' W 2 ) is funda-
mental. [J

If f and g are elliptic functions with periods Wi and W 2 then their sum,
difference, product and quotient are also elliptic with the same periods. So,
too, is the derivative f'.
Because of periodicity, it suffices to study the behavior of an elliptic
function in any period parallelogram.

Theorem 1.4. If an elliptic function f has no poles in some period parallelogram,


then f is constant.
PROOF. Iffhas no poles in a period parallelogram, thenfis continuous and
hence bounded on the closure of the parallelogram. By periodicity, f is
bounded in the whole plane. Hence, by Liouville's theorem,fis constant. [J

Theorem 1.5. If an elliptic function f has no zeros in some period parallelogram,


then f is constant.
PROOF. Apply Theorem 1.4 to the reciprocaI1/! [J

Note. Sometimes it is inconvenient to have zeros or poles on the bound-


ary of a period parallelogram. Since a meromorphic function has only a
finite number of zeros or poles in any bounded portion of the plane, a period
parallelogram can always be translated to a congruent parallelogram with
no zeros or poles on its boundary. Such a translated parallelogram, with no
zeros or poles on its boundary, will be called a cell. Its vertices need not be
periods.

5
1: Elliptic functions

Theorem 1.6. The contour integral of an elliptic function taken along the
boundary of any cell is zero.
PROOF. The integrals along parallel edges cancel because of periodicity. D

Theorem 1.7. The sum of the residues of an ellipticfunction at its poles in any
period parallelogram is zero.
PROOF. Apply Cauchy's residue theorem to a cell and use Theorem 1.6. D
\
Note. Theorem 1.7 shows that an elliptic function which is not constant
has at least two simple poles or at least one double pole in each period
parallelogram.

Theorem 1.8. The number ofzeros ofan elliptic function in any period parallel-
ogram is equal to the number of poles, each counted with multiplicity.
PROOF. The integral

_1 f
f'(z) dz
2ni c j'(z) ,

taken around the boundary C of a cell, counts the difference between the
number of zeros and the number of poles insid'e the cell. But f' / f is elliptic
with the same periods as/, and Theorem 1.6 tells us that this integral is zero.
D

Note. The number of zeros (or poles) of an elliptic function in any period
parallelogram is called the order of the function. Every nonconstant elliptic
function has order ~ 2.

1.5 Construction of elliptic functions


We turn now to the problem of constructing a nonconstant elliptic function.
We prescribe the periods and try to find the simplest elliptic function having
these periods. Since the order of such a function is at least 2 we need a
second order pole or two simple poles in each period parallelogram. The
two possibilities lead to two theories of elliptic functions, one developed by
Weierstrass, the other by Jacobi. We shall follow Weierstrass, whose point
of departure is the construction of an elliptic function with a pole of order
2 at z = 0 and hence at every period. Near each period OJ the principal part
of the Laurent expansion must have the form

A B
---2+--·
(z - OJ) Z - OJ

6
1.5: Construction of elliptic functions

For simplicity we take A = 1, B = O. Since we want such an expansion near


each period w it is natural to consider a sum of terms of this type,

summed over all the periods w = mW l + nW2. For fixed z =1= w this is a
double series, summed over m and n. The next two lemmas deal with con-
vergence properties of double series of this type. In these lemmas we denote
by Q the set of all linear combinations mw 1 + nW2' where m and n are
arbitrary integers.

Lemma 1. If rJ. is real the infinite series

converges absolutely if, and only if, rJ. > 2.


PROOF. Refer to Figure 1.3 and let rand R denote, respectively, the minimum
and maximum distances from 0 to the parallelogram shown. If w is any of
the 8 nonzero periods shown in this diagram we have

r ~ Iwl ~ R (for 8 periods w).

Figure 1.3

In the next concentric layer of periods surrounding these 8 we have 2 ·8 = 16


new periods satisfying the inequalities

2r:::; Iwl :::; 2R (for 16 new periods w).

In the next layer we have 3 . 8 = 24 new periods satisfying

3r :::; Iwl :::; 3R (for 24 new periods w),

7
1: Elliptic functions

and so on. Therefore, we have the inequalities

~" ~ I~ I" ~ ~ for the first 8 periods w,

(2~t ~ I~I" ~ (2~)" for the next 16 periods w,


and so on. Thus the sum S(n) = L /OJI-(X, taken over the 8(1 + 2 + ... + n)
nonzero periods nearest the origin, satisfies the inequalities
8 2·8 n·8 8 2·8 n·8
R" + (2Rt + ... + (nR)" ~ S(n) ~ ? + (2r)" + ... + (nr)'"
or
8 n 1 8 n 1
R
(X L
k=l
k(X-l ~ S(n) ~ ex
r
L
k=l
k(X-l·

This shows that the partial sums S(n) are bounded above by 8'(~ - l)jr(X if
~ > 2. But any partial sum lies between two such partial sums, so all of the
partial sums of the series L /OJ ,- (X are bounded above and hence the series
converges if ~ > 2. The lower bound for S(n) also shows that the series
diverges if ~ ~ 2. D

Lemma 2. If ~ > 2 and R > 0 the series

L
Iwl > R
1"
(z - OJ)
converges absolutely and uniformly in the disk Iz I ~ R.
PROOF. We will show that there is a constant M (depending on R and ~)
such that, if ~ ~ '1, we have
1 M
(1) ---<-
Iz - OJI(X - IOJI(X

for all OJ with /OJ / > R and all z with Iz I ~ R. Then we invoke Lemma 1 to
prove Lemma 2. Inequality (1) is equivalent to

(2) Iz: WI" ~ ~.


To exhibit M we consider all OJ in n with IOJ I > R. Choose one whose
modulus is minimal, say IOJ I = R + d, where d > O. Then if Iz I s Rand
IOJ I ~ R + d we have

Iz: Wi 1 ~I ~ = 1- 1 -I~I ~ 1 - R ~ d'


8
1.6: The Weierstrass tJ function

and hence

where

M= ( l -
R
+d
)-a..
R
This proves (2) and also the lemma. D
As mentioned earlier, we could try to construct the simplest elliptic
function by using a series of the form
1
L
WEn
(Z-W)2'

This has the appropriate principal part near each period. However, the
series does not converge absolutely so we use, instead, a series with the
exponent 2 replaced by 3. This will give us an elliptic function of order 3.
Theorem 1.9. Let f be defined by the series
1
f(z) = L(
WEn z - OJ
)3'

Thenfis an ellipticfunction with periods OJ 1 , OJ 2 and with a pole oforder 3 at


each period OJ in Q.
PROOF. By Lemma 2 the series obtained by summing over IOJ I > R converges
uniformly in the disk Iz I s R. Therefore it represents an analytic function
in this disk. The remaining terms, which are finite in number, are also
analytic in this disk except for a 3rd order pole at each period OJ in the disk.
This proves thatfis meromorphic with a pole of order 3 at each OJ in Q.
Next we show thatfhas periods OJ 1 and OJ 2 • For this we take advantage
of the absolute convergence of the series. We have
1
f(z + wd = L ( + OJ
WEn z 1 - OJ
)3'

But OJ - OJ 1 runs through all periods in Q with OJ, so the series for f(z + OJ 1 )
is merely a rearrangement of the series for f(z). By absolute convergence we
have f(z + OJ 1 ) = f(z). Similarly, f(z + OJ 2 ) = f(z) so f is doubly periodic.
This completes the proof. 0

1.6 The Weierstrass 80 function


Now we use the function of Theorem 1.9 to construct an elliptic function
or order 2. We simply integrate the series for f(z) term by term. This gives us
a principal part - (z - OJ) - 2/2 near each period, so we multiply by - 2 to

9
1: Elliptic functions

get the principal part (z - w) - 2. There is also a constant of integration to


reckon with. It is convenient to integrate from the origin, so we remove the
term z - 3 corresponding to w = 0, then integrate, and add the term z - 2.
This leads us to the function
1 fZ -2
2
z
+ 0
L(
£1)*0 t - w
)3 dt.

Integrating term by term we arrive at the following function, called the


Weierstrass g;J function.

Definition. The Weierstrass g;J function is defined by the series

.f.J(z)
1 L
= 2z + w*o {II}
( Z - W
)2 - 2
W
.

Theorem 1.10. The function g;J so defined has periods W t and w 2 • It is analytic
except for a double pole at each period w in Q. Moreover g;J(z) is an even
function of z.
PROOF. Each term in the series has modulus

1 lIIW2-(Z-W)21IZ(2W-Z)1
1(z - W)2 - w2 2 = 2w (z - W)2 = w (z - W)2 .

Now consider any compact disk Iz I ::; R. There are only a finite number of
periods W in this disk. If we exclude the terms of the series containing these
periods we have, by inequality (1) obtained in the proof of Lemma 2,

where M is a constant depending only on R. This gives us the estimate

z(2w - z) I MR(2Iwl + R) MR(2 + R/lwl) 3MR


< < <--
I 2(z - W)2 -
w Iwl 4 - Iwl 3 - Iwl 3
since R < Iw I for w outside the disk Iz I::; R. This shows that the truncated
series converges absolutely and uniformly in the disk Iz I ::; R and hence
is analytic in this disk. The remaining terms give a second-order pole at
each w inside this disk. Therefore g;J(z) is meromorphic with a pole of order 2
at each period.
Next we prove that g;J is an even function. We note that

Since - w runs through all nonzero periods with w this shows that g;J( - z) =
g;J(z), so g;J is even.

10
1.8: Differential equation satisfied by f.J

Finally we establish periodicity. The derivative of ~ is given by


, 1
f.J (z) = - 2 L(
wen Z - W
)3·

We have already shown that this function has periods WI and ~2. Thus
~'(z + w) = ~'(z) for each period w. Therefore the function ~(z + w) - ~(z)
is constant. But when z = - w/2 this constant is ~(w/2) - ~(- w/2) = 0
since ~ is even. Hence ~(z + w) = ~(z) for each w, so ~ has the required
periods. 0

1.7 The Laurent expansion of 8'J near the


origin
Theorem 1.11. Let r = min {Iwl:w =I- O}. Thenfor 0 < Izi < r we have
1 00

(3) ~(z) = 2:
Z
+ L1 (2n + I)G2n+2z2n,
n=
where

(4) for n ~ 3.

PROOF. If 0 < Izi < r then Iz/wl < 1 and we have


1
--.,.(----.)-2
21
1
=
z
1 (
2
W
1+ L (n + 1)
00

n=1
(z )n)
-
W
,
W --
w
hence

Summing over all w we find (by absolute convergence)


1 00 1 1 00

~(z) = 2: + L (n + 1) L n+2 zn = 2: + L (n + I)G n + n


2z ,
z n=1 w*O W Z n=1
where Gn is given by (4). Since ~(z) is an even function the coefficients G 2n + 1
must vanish and we obtain (3). 0

1.8 Differential equation satisfied by 8'J


Theorem 1.12. The function ~ satisfies the nonlinear differential equation
[~'(Z)]2 = 4~3(Z) - 60G 4 SO(z) - 140G 6 .

PROOF. We obtain this by forming a linear combination of powers of ~ and


~'which eliminates the pole at z = O. This gives an elliptic function which has

11
1: Elliptic functions

no poles and must therefore be constant. Near z = 0 we have


2
8O'(z) = - 3 + 6G 4 z + 20G 6 z3 + ...,
z
an elliptic function of order 3. Its square has order 6 since
2 4 24G
[80'(z)] = -Z6 - -Z2-4 - 80G 6 + .. "
where + ... indicates a power series in z which vanishes at z = O. Now
4f.J3(Z) = ~ + 36G4 + 60G 6 + ...
Z6 Z2

hence
2 3 60G 4
[80'(z)] - 480 (z) = - -2- - 140G 6 + ...
z
so
[80'(z)]2 - 480 3(z) + 60G 4 80(z) = -140G 6 + ....
Since the left member has no pole at z = 0 it has no poles anywhere in a
period parallelogram so it must be constant. Therefore this constant must
be -140G 6 and this proves the theorem. D

1.9 The Eisenstein series and the invariants


92 and 93
Definition. If n ~ 3 the series

is called the Eisenstein series of order n. The invariants 92 and 93 are the
numbers defined by the relations

The differential equation for 80 now takes the form


[80'(z)]2 = 480 3(z) - g280(z) - g3'
Since only g2 and g3 enter in the differential equation they should determine
80 completely. This is actually so because all the coefficients (2n + 1)G 2n + 2
in the Laurent expansion of 8O(z) can be expressed in terms of g2 and g3.

Theorem 1.13. Each Eisenstein series Gn is expressible as a polynomial in g2


and g 3 with positive rational coefficients. In fact, if b(n) = (2n + l)G 2n + 2
we have the recursion relations
b(2) = g3/28,

12
and
n-2
(2n + 3)(n - 2)b(n) = 3 L b(k)b(n - 1 - k) for n ~ 3,
k=1

or equivalently,
m-2
(2m + l)(m - 3)(2m - 1)G 2m = 3 L (2r - 1)(2m - 2r - 1)G2rG2m-2r
r= 2

for m ~ 4.

PROOF. Differentiation of the differential equation for ~ gives another


differential equation of second order satisfied by ~,

(5) ~"(z) = 6~2(Z) - tg2.


Now we write ~(z) = z- 2 + L~= 1 b(n)z2n and equate like powers of z
in (5) to obtain the required recursion relations. D

Definition. We denote by eb e2' e3 the values of ~ at the half-periods,

The next theorem shows that these numbers are the roots of the cubic
polynomial4~3 - g2 ~ - g3·

Theorem 1.14. We have


4~3(Z) - g2 ~(z) - g3 = 4(~(z) - el)(~(z) - e2)(~(z) - e 3).

Moreover, the roots el' e2' e3 are distinct, hence g2 3 - 27g 32 #- o.


PROOF. Since ~ is even, the derivative ~' is odd. But it is easy to show that
the half-periods of an odd elliptic function are either zeros or poles. In fact,
by periodicity we have ~'( -tw) = ~'(w - tw) = ~'(tw), and since ~'is odd
we also have ~'( -tw) = - ~'(tw). Hence ~'(tw) = 0 if ~'(tw) is finite.
Since ~'(z) has no poles at tw 1, tw 2, t(w 1 + (2), these points must be
zeros of ~'. But ~' is of order 3, so these must be simple zeros of ~'. Thus
~' can have no further zeros in the period-parallelogram with vertices
0, WI' W2' WI + w 2 . The differential equation shows that each of these points
is also a zero of the cubic, so we have the factorization indicated.
Next we show that the numbers el' e2' e3 are distinct. The elliptic function
~(z) - el vanishes at z = tw 1. This is a double zero since ~'(tWl) = o.
Similarly, ~(z) - e2 has a double zero at tW2. If el were equal to e2, the
elliptic function ~(z) - el would have a double zero at tW I and also a double

13
1: Elliptic functions

zero at tOJ 2 , so its order would be ~ 4. But its order is 2, so e 1 =I- e2. Similarly,
e 1 =I- e 3 and e 2 =I- e 3.
If a polynomial has distinct roots, its discriminant does not vanish. (See
Exercise 1.7.) The discriminant of the cubic polynomial
4x 3 - g2 X - g3
is g2 3 - 27g 32. When x = ~(z) the roots of this polynomial are distinct so
the number g2 3 - 27g 32 =I- O. This completes the proof. 0

1.11 The discriminant ~

The number ~ = g 23 - 27 g 32 is called the discriminant. We regard the


invariants g2 and g3 and the discriminant ~ as functions of the periods WI
and W2 and we write

The Eisenstein series show that g2 and g3 are homogeneous functions of


degrees -4 and -6, respectively. That is, we have

g2(AW b A(2) = A-4 g2 (W b ( 2) and g3(AW 1, A( 2) = A-6 g3 (W b ( 2)


for any A =I- O. Hence ~ is homogeneous of degree -12,

~(AWb A(2) = A-12~(W1' ( 2).


Taking A = 1/w 1 and writing T = W2/W1 we obtain
g2(1, T) = W1 4 g2(W 1, ( 2), g3(1, T) = W 16 g3 (W 1, ( 2),
12
~(1, T) = W l ~(W1' ( 2).

Therefore a change of scale converts g2' g3 and ~ into functions of one


complex variable T. We shall label W1 and W2 in such a way that their ratio
T = W2/W1 has positive imaginary part and study these functions in the upper
half-plane Im(T) > O. We denote the upper half-plane Im(T) > 0 by H.
If T E H we write g2(T), g3(T) and ~(T) for g2(1, T) g3(1, T) and ~(1, T),
respectively. Thus, we have
+00 1
92(r) = 60
m, n
L= - (m + nT)4'
(m, n) * (0,0)ex;

+00 1
93(r) = 140 m.n~-oo (m + nr)6
(m, n) * (0,0)
and
~(T) = g2 3(T) - 27g 32(T).
Theorem 1.14 shows that ~(T) =I- 0 for all T in H.

14
1.12: Klein's modular function J(r)

1.12 Klein's modular function J(7:)


Klein's function is a combination of g2 'and g3 defined in such a way that,
as a function of the periods WI and W2' it is homogeneous of degree O.

Definition. If w 2 1w 1 is not real we define


g2 3(W 1 , (2)
J(w 1 , w 2 ) = ~( ).
W1 ,W2
Since g 23 and i\ are homogeneous of the same degree we have J(AW 1, A(2)
= J(Wb (2). In particular, if T E H we have
J(l, T) = J(Wl' ( 2).
Thus J(w 1 , (2) is a function of the ratio T alone. We write J(T) for J(l, T).

Theorem 1.15. The functions g2(T), g3(T), i\(T), and J(T) are analytic in H.

PROOF. Since i\(T) =I- 0 in H it suffices to prove that g2 and g3 are analytic
in H. Both {l" and g3 are given by double series of the form
+00 1
m.n~-oo
(m,n)*(O,O)
(m + mY
·'ith ex > 2. Let T = X + iy, where y > O. We shall prove that if ex > 2 this
series converges absolutely for any fixed T in H and uniformly in every strip
S of the form
S = {x + iy:lxl ~ A,y ~ iJ > O}.
(See Figure 1.4.) To do this we prove that there is a constant M > 0, depending
only on A and on iJ, such that
1 M
(6) - - - - < - - -a
1m + nTla - 1m + nil
for all T in S and all (m, n) =I- (0,0). Then we invoke Lemma 1.
To prove (6) it suffices to prove that
1m + nTI 2 > Kim + nil 2
for some K > 0 which depends only on A and iJ, or that
(7) (m + nx)2 + (ny)2 > K(m 2 + n2).
If n = 0 this inequality holds with any K such that 0 < K < 1. If n =I- 0
let q = min. Proving (7) is equivalent to showing that
(q + X)2 + y2
(8) 1 2 > K
+q
15
1: Elliptic functions

-----+----+----+------~ x
-A A

Figure 1.4

for some K > O. We will prove that (8) holds for all q, with
b2
K=-----
1 + (A + b)2
if Ix I s A and y ~ b. (This proof was suggested by Christopher Henley.)
If Iq I s A + b inequality (8) holds trivially since (q + X)2 ~ 0 and
y2 ~ b2. Iflql > A + b then Ix/ql < Ixl/(A + b) s A/(A + b) < 1 so

1+ ~I ~
1
q
1 -I~Iq > 1_ _ A
A+b
= _b
A+b
hence
qb
Iq + xl ~ A +b
and
(q + X)2 + y2 b2 q2
(9) 1 + q2 > (A + 15)2 1 + q2·
Now q2/(1 + q2) is an increasing function of q2 so
q2 (A + b)2
-->-----
1 + q2 - 1 + (A + b)2
when q2 > (A + b)2. Using this in (9) we obtain (8) with the specified K. D

1.13 In\rariance of J under unimodular


transformations
If WI' W2 are given periods with nonreal ratio, introduce new periods
WI" w 2 ' by the relations

16
1.13: Invariance of J under unimodular transformations

where a, b, c, d are integers such that ad - bc = 1. Then the pair (WI" W2')
is equivalent to (WI' W 2 ); that is, it generates the same set of periods Q.
Therefore g2(W I', w 2') = g2(W I , W2) and g3(W I', W2') = g3(Wb W2) since g2
and g3 depend only on the set of periods Q. Consequently, ~(WI" w 2 ') =
~(WI' W2) and J(WI', W2') = J(w l , w 2)·
The ratio of the new periods is
aW2 + bW I aT +b
CW2 + dW I CT + d'
where T = W2/WI. An easy calculation shows that
, (aT + b) ad - be Im(T)
Im(T) = 1m C't" +d = ICT + dl 2 Im(T) = leT + dl 2 '

Hence T' E H if and only if T E H. The equation


,
T =---
aT +b
CT d +
is called a unimodular transformation if a, b, c, d are integers with ad - bc = 1.
The set of all unimodular transformations forms a group (under composition)
called the modular group. This group will be discussed further in the next
chapter. The foregoing remarks show that the function J(T) is invariant
under the transformations of the modular group. That is, we have:

Theorem 1.16. If T E H and a, b, c, d are integers with ad - bc = 1, then


(aT + b)/(CT + d) E Hand

(10) aT
J ( CT
+
+d =
b) J(T).

Note. A particular unimodular transformation is T' = T + 1, hence (10)


shows that J( T + 1) = J( T). In other words, J( T) is a periodic function of T
with period 1. The next theorem shows that J(T) has a Fourier expansion.

Theorem 1.17. If T E H, J(T) can be represented by an absolutely convergent


FDurier series
00

(11) J(T) = L a(n)e21tint.


n= - 00

PROOF. Introduce the change of variable

Then the upper half-plane H maps into the punctured unit disk
D = {x:O < Ixl < I}.
17
1: Elliptic functions

(See Figure 1.5.) Each T in H maps onto a unique point x in D, but each
x in D is the image of infinitely many points in H. If T and T' map onto x
then e 21tit = e21tit' so T and T' differ by an integer.

D
H

Figure 1.5

If xED, let
f(x) = J(T)
where T is any of the points in H which map onto x. Since J is periodic with
period 1, J has the same value at all these points so .f(x) is well-defined.
Now fis analytic in D because

, d d dT , /dX J'(T)
f (x) = -d J(T) = -d J(T) -d = J (T) -d = 2 . 21tit'
X TXT nle

so f'(x) exists at each point in D. Since f is analytic in D it has a Laurent


expansion about 0,
00

f(x) = L a(n)x n ,
n= - 00

absolutely convergent for each x in D. Replacing x by e 21tit we see that J(T)


has the absolutely convergent Fourier expansion in (11). D

Later we will show that a_ n = 0 for n ~ 2, that a-I = 12- 3 , and that
the Fourier expansion of 12 3 J(T) has integer coefficients. To do this we first
determine the Fourier expansions of g2(T), g3(T) and ~(T).

1.14 The Fourier expansions of g2(r)


andg 3 (r)
Each Eisenstein series L(m,
n) #: (0,0) (m + nT) - k is a periodic function of T of

period 1. In particular, g2(T) and g3(T) are periodic with period 1. In this
section we determine their Fourier coefficients explicitly.
We recall that
1 1
g2(T) = 60 L
(m,n)#:(O,O) (m + nT)
4' g3(T) = 140 L
(m,n)#:(O, 0) (m + nT)

18
1.14: The Fourier expansions of g2(r) and g3(r)

These are double series in m and n. First we obtain Fourier expansions for
the simpler series
+00 1 +00 1
L
m= - 00 (m + nT)
4 and L
m= - 00 (m + nT)
6'

Lemma 3. If T E Hand n > 0 we have the Fourier expansions


1
- -8n L r3e21tirnt
4
L
+ 00 00

m= - 00 (m + nT)4 - 3 r= 1
and
1 8n 6
L L r5e21tirnt.
+ 00 00
6 = - -
m= - 00 (m + nT) 15 r= 1

PROOF. Start with the partial fraction decomposition of the cotangent:

L (-
n cot nT = -1 + +00 1- - -1).
T m=-oo T+m m
m*O
Let x = e21tit. If T E H then Ix I < 1 and we find
21tit
nT = nei . + 1 = ni -
cos -
n cot nT = n - x +-1 = - ni(-
x- + -1
-)
sin nT e 21tlt - 1 x-I 1- x 1- x

= - ni( I x + r=OI x
r=1
r r
) = - ni(l + 2 Ix
r=1
r
).

In other words, if T E H we have

-1 + L +00 (-
1- - -1) = - ni( 1 + 2 L e21tirt . 00 )

T m=-oo T+m m r=1


m*O
Differentiating repeatedly we find
00 1 00

(12) L = - (2ni)2 L re21tirt


m= - 00 (T + m)2 r =1
m*O
+00 1 00

-3! L 4 = -(2ni)4 L r3e21tirt


m=_oo(T+m) r=1

and

Replacing T by nT we obtain Lemma 3. o


19
1: Elliptic functions

Theorem 1.18. If T E H we have the Fourier expansions

g2(T) = -4n4 { 1 + 240 L (J3(k)e21tikt


00 }

3 k=1
and

g3('r) = 82~6 {1 - 504k~/T5(k)e2"ik}


where (J(J.(k) = Ldlk d(J..
PROOF. We write
+00 1
g2(r) = 60 m.n~-oo (m + nr)4
(m, n) =#= (0,0)

1 (1 1)}
=60 { L
+ 00

m= - 00
4
m
+ L L
00

n= 1 m= -
+ 00

00 (m + nT) 4+ (m - nT)4
m=#=O(n=O)

00 +00 1 }
= 60{ 2((4) + 2n~1 m=~oo (m + nr)4
2n 4 16n4
L1 L1 r
00 00 }
= 60 { - + - 3
x
nr
90 3 n= r=

where x = e 21tit . In the last double sum we collect together those terms for
which nr is constant and we obtain the expansion for g2(T). The formula
for g3(T) is similarly proved. 0

1.15 The Fourier expansions of ~(r) and J( r)

Theorem 1.19. 1fT E H we have the Fourier expansion


00

~(T) = (2n)12 L T(n)e21tint


n=1
where the coefficients T(n) are integers, with T(1) = 1 and T(2) = - 24.

Note. The arithmetical function T(n) is called Ramanujan's tau function.


Some of its arithmetical properties are described in Chapter 4.
PROOF. Let
00 00

A = L (J3(n)x n, B= L (J5(n)x n.
n=1 n=1
Then
64n 12
~(r) = g/(r) - 27g/(r) = ~ {(1 + 240A)3 - (1 - 504B)2}.

20
1.15: The Fourier expansions of L\(r) and J(r)

Now A and B have integer coefficients, and


(1 + 240A)3 - (1 - 504B)2 = 1 + 720A + 3(240)2 A 2 + (240)3 A 3 - 1
+ 1008B - (504)2 B 2
= 12 2 (5A + 7B)
+ 12 3(100A 2 - 147B 2 + 8000A 3 ).
But
00

5A + 7B = L {5a3(n) + 7a s(n)}x n
n=l
and

so
5d 3 + td s == 0 (mod 12).
3
Hence 12 is a factor of each coefficient in the power series expansion of
(1+ 240A)3 - (1 - 504B)2 so
12
~(r) = 64n
27
3
n=1
{12 f
r(n)e 21tinr } = (2n)12
n=1
r(n)e 21tinr f
where the r(n) are integers. The coefficient of x is 12 2 (5 + 7), so r(l) = 1.
Similarly, we find r(2) = - 24. D

Theorem 1.20. Ifr E H we have the Fourier expansion


00

12 3J(r) = e- 21tir + 744 + L c(n)e21tinr,


n= 1
where the c(n) are integers.

PROOF. We agree to write I for any power series in x with integer coefficients.
Then if x = e 21ti t: we have
g2
3
(r) = ~~n12(1 + 240x + 1)3 = ~~n12(1 + 720x + I),
~(r) = ~~n12{123x(1 - 24x + I)}
and hence
g2
3
(r) 1 + 720x + I 1
J(r) = ~(r) = 123x(1 _ 24x + 1) = 123x (1 + 720x + 1)(1 + 24x + 1)
so
1 00

12 3J(r) = - + 744 + L c(n)x n,


x n= 1
where the c(n) are integers. D
21
1: Elliptic functions

Note. The coefficients c(n) have been calculated for n ~ 100. Berwick
calculated the first 7 in 1916, Zuckerman the first 24 in 1939, and Van
Wijngaarden the first 100 in 1953. The first few are repeated here.
c(O) = 744
c(l) = 196, 884
c(2) = 21, 49~, 760
c(3) = 864, 299, 970
c(4) = 20, 245, 856, 256
c(5) = 333, 202, 640, 600
c(6) = 4, 252, 023, 300, 096
c(7) = 44, 656, 994, 071, 935
c(8) = 401, 490, 886, 656, 000
The integers c(n) have a number of interesting arithmetical properties. In
1942 D. H. Lehmer [20] proved that
(n + l)c(n) == 0 (mod 24) for all n ~ 1.
In 1949 Joseph Lehner [23] discovered divisibility properties of a different
kind. For example, he proved that
c(5n) == 0 (mod 25),
c(7n) == 0 (mod 7),
c(lln) == 0 (mod 11).
He also discovered congruences for higher powers of 5, 7, 11 and, in a later
paper [24] found similar results for the primes 2 and 3. In Chapter 4 we will
describe how some of Lehner's congruences are obtained.
An asymptotic formula for c(n) was discovered by Petersson [31] in 1932.
It states that
e47t.Jl1
c(n) '" M as n ~ 00.
v 2 n3 /4
This formula was rediscovered independently by Rademacher [37] in
1938.
The coefficients r(n) in the Fourier expansion of ~(r) have also been
extensively tabulated by D. H. Lehmer [19] and others. The first ten entries
in Lehmer's table are repeated here:
r(l) = 1 r(6) = -6048
r(2) = -24 r(7) = - 16744
r(3) = 252 r(8) = 84480
r(4) = -1472 r(9) = -113643
r(5) = 4830 r(10) = -115920.
Lehmer has conjectured that r(n) i= 0 for all n and has verified this for all
n < 214928639999 by studying various congruences satisfied by r(n). For
papers on r(n) see Section F35 of [27].

22
Exercises for Chapter 1

Exercises for Chapter 1


1. Given two pairs of complex numbers (Wi' W z ) and (Wi" W z') with nonreal ratios
WZ/w l and WZ'/w l '. Prove that they generate the same set of periods if, and only if,

there is a 2 x 2 matrix (: :) with integer entries and determinant ± 1 such that

2. Let S(O) denote the sum of the zeros of an elliptic function f in a period parallelo-
gram, and let S( (0) denote the sum of the poles in the same parallelogram. Prove
that S(O) - S( (0) is a period of I [Hint: Integrate ~f"(z)/f(z).]

3. (a) Prove that p(u) = p(v) if, and only if, U - v or u + v is a period of p.
(b) Let ai' ... , an and b 1, ... , bm be complex numbers such that none of the numbers
p(ai) - p(b j) is zero. Let

f(z) = }), [p(z) - p(ak)] III [p(z) - p(b,)].

Prove that f is an even elliptic function with zeros at ai' ... , an and poles at
b l , · · · , bm •

4. Prove that every even elliptic function f is a rational function of p, where the
periods of p are a subset of the periods of I

5. Prove that every elliptic function f can be expressed in the form

where R 1 and R z are rational functions and p has the same set of periods as f

6. Let f and 9 be two elliptic functions with the same set of periods. Prove that there
exists a polynomial P(x, y), not identically zero, such that
P[f(z), g(z)] = C
where C is a constant (depending on f and g but not on z).

7. The discriminant of the polynomial f(x) = 4(x - xd(x - xz)(x - X3) is the
product 16{(xz - Xd(X3 - XZ)(X3 - xd}z. Prove that the discriminant of f(x) =
4x 3 - ax - b is a 3 - 27 bZ •

8. The differential equation for p shows that p'(z) =0 if z = w l /2, wz/2 or


(Wi + w z)/2. Show that

p"(~,) = 2(el - e2)(el - e3)

and obtain corresponding formulas for p"(w z/2) and p"((w l + w z )/2).

23
1: Elliptic functions

9. According to Exercise 4, the function gu(2z) is a rational function of gu(z). Prove


that, in fact,

(2z) = {~l(z) + ~92}2 + 293 ~(z) = _ 2~(z) + !(~~(Z»)2.


~ 4~3(Z) - 92~(Z) - 93 4 ~ (z)

10. Let 01 1 and W2 be complex numbers with nonreal ratio. Letf(z) be an entire function
and assume there are constants a and b such that
f(z + wd = af(z), f(z + w2) = bf(z),
for all z. Prove that f(z) = AeBz , where A and B are constants.
11. If k ~ 2 and r E H prove that the Eisenstein series
G 2k(r) = L (m + nr)-2k
(m, n) * (0,0)
has the Fourier expansion

12. Refer to Exercise 11. If r E H prove that

G 2k( - l/r) = r 2k G 2k(r)


and deduce that
G 2k(i/2) = (- 4lG 2k(2i) for all k ~ 2,
G 2k(i) = 0 if k is odd,
G2k(e21ti/3) = 0 if k =1= 0 (mod 3).

13. Ramanujan's tau function r(n) is defined by the Fourier expansion


00

L\( r) = (2n)12 L r(n)e 21tint,


n=1

derived in Theorem 1.19. Prove that

where fog denotes the Cauchy product of two sequences,


n
(f g)(n) =
0 L f(k)g(n - k),
k=O

and O"a(n) = Ldln da for n ~ 1, with 0"3(0) = 2lo, 0"5(0) = - 5~4'


[Hint: Theorem 1.18.]
14. A series of the form L~= 1 f(n)x n/( 1 - x n ) is called a Lambert series. Assuming
absolute convergence, prove that
00 xn 00

L f(n) --n = L F(n)x n,


n=l I-x n=1

where
F(n) = L f(d).
dIn

24
Exercises for Chapter 1

Apply this result to obtain the following formulas, valid for Ix I < 1.
00 j1(n)xn 00 <p(n)x n x
(a) L
n=l
-1- n
- X
= X. (b) L
n=l
-_1n = - (
_1
X
)2'
X

00 na.xn 00 00 A(n)xn 00

(c) L --n = L O"a.(n)x n. (d) L --n = L x n2 .


n= 1 1 - x n= 1 n= 1 1- x n= 1

(e) Use the result in (c) to express g2(r) and g3(r) in terms of Lambert series in
x = e 21tir •
Note. In (a), j1(n) is the Mobius function; in (b), <p(n) is Euler's totient; and in (d),
A(n) is Liouville's function.

15. Let
00 n5xn
G(x) = "~1 1 - x"'

and let
n5x n
F(x) = L 1 + Xn
n=l
(n odd)

(a) Prove that F(x) = G(x) - 34G(x 2 ) + 64G(x 4 ).


(b) Prove that
31
504

(c) Use Theorem 12.17 in [4] to prove the more general result
::G n4k + 1 24k + I - 1
2:
n = 1
1 + e = 8k + 4 B 4k + 2'
n7r

(nodd)

25
The modular group
and modular functions

2.1 Mobius transformations


In the foregoing chapter we encountered unimodular transformations
, ar +b
r=--
cr +d
where a, b, c, d are integers with ad - bc = 1. This chapter studies such
transformations in greater detail and also studies functions which, like
J(r), are invariant under unimodular transformations. We begin with some
remarks concerning the more general transformations

(1) f(z) = az + b
cz +d
where a, b, c, d are arbitrary complex numbers.
Equation (1) definesj(z) for all z in the extended complex number system
C* = C U {oo} except for z = -dlc and z = 00. We extend the definition
ofjto all of C* by defining
a
and j(oo)=-,
c
with the usual convention that zlO = 00 if z i= O.
First we note that

(2) f(w) - f(z) = (ad - bc)(w - z)


(cw + d)(cz + d)'
which shows thatjis constant if ad - bc = O. To avoid this degenerate case
we assume that ad - bc i= O. The resulting rational function is called a

26
2.1: Mobius transformations

Mobius transformation. It is analytic everywhere on C* except for a simple


pole at z = -die.
Equation (2) shows that every Mobius transformation is one-to-one on
C*. Solving (1) for z in terms off(z) we find
df(z) - b
z=-----
-ef(z) + a'
so f maps C* onto C*. This also shows that the inverse function f -1 is a
Mobius transformation.
Dividing by w - z in (2) and letting w ~ z we obtain
, ad - be
f(z) = (cz + df'

°
hence f'(z) =I- at each point of analyticity. Therefore f is conformal every-
where except possibly at the pole z = -die.
Mobius transformations map circles onto circles (with straight lines
being considered as special cases of circles). To prove this we consider the
equation
(3) Azz + Bz + Bz + C = 0,
where A and C are real. The points on any circle satisfy such an equation
with A =I- 0, and the points on any line satisfy such an equation with A = 0.
Replacing z in (3) by (aw + b)/(ew + d) we find that w satisfies an equation
of the same type,
A'ww + B'w + B'w + C' = °
where A' and C' are also real. Hence every Mobius transformation maps a
circle or straight line onto a circle or straight line.
A Mobius transformation remains unchanged if we multiply all the
coefficients a, b, e, d by the same nonzero constant. Therefore there is no loss
in generality in assuming that ad - be = 1.
For each Mobius transformation (1) with ad - be = 1 we associate the
2 x 2 matrix

Then det A = ad - be = 1. If A and B are the matrices associated with


Mobius transformations f and g, respectively, then it is easy to verify
that the matrix prod uct AB is associated with the composition fog, where

(fog)(z) = f(g(z)). The identity matrix I = (~ ~) is associated with the


identity transformation

1z
f(z) = z = 02
+
+
°
l'

27
2: The modular group and modular functions

and the matrix inverse

A-I =(
-c
d -b)a
is associated with the inverse off,
dz - b
f-l(Z) = ---
-cz + a
Thus we see that the set of all Mobius transformations with ad - bc = 1
forms a group under composition. This chapter is concerned with an impor-
tant subgroup in which the coefficients a, b, c, d are integers.

2.2 The modular group r


The set of all Mobius transformations of the form

T
,
=---
aT +b
CT + d'
where a, b, c, d are integers with ad - bc = 1, is called the modular group and
is denoted by r. The group can be represented by 2 x 2 integer matrices

A = G~) with det A = 1,

provided we identify each matrix with its negative, since A and - A represent
the same transformation. Ordinarily we will make no distinction between
the matrix and the transformation. If A = (: ~) we write
aT +b
AT = CT + d·
The first theorem shows that r is generated by two transformations,

TT = T + 1 and ST =

Theorem 2.1. The modular group r is generated by the two matrices

T = (~ ~) and s= (~ -1)o·
That is, every A in r can be expressed in theform
A = T"lST"2S ... ST"k
where the ni are integers. This representation is not unique.

28
2.2: The modular group r

PROOF. Consider first a particular example, say

A = C; 2~)
We will express A as a product of powers of Sand T. Since S2 = I, only the
first power of S will occur.
Consider the matrix product

AT" = C~ 2DG ~) = C; 1;:: 2~)


Note that the first column remains unchanged. By a suitable choice of n
we can make 111n + 25/ < 11. For example, taking n = - 2 we find
lIn + 25 = 3 and

AT- 2 = (4 1).
11 3

Thus by multiplying A by a suitable power of T we get a matrix ( : ; ) with


Id I < /C /. Next, multiply by S on the right:

AT- S
2
= C; ~)(~ -~) = G -~;)
This interchanges the two columns and changes the sign of the second column.
Again, multiplication by a suitable power of T gives us a matrix with
Id I < /C /. In this case we can use either T 4 or T 3 . Choosing T 4 we find
1
AT- 2 ST 4 = (3 -4)(1 4) (1 0)
-11 0 1 = 3 1·
Multiplication by S gives

2 4
AT- ST S = (~ =~)
Now we multiply by T 3 to get

AT-2ST4ST3=(~ =~)(~ ~)=(~ -~)=S.


Solving for A we find
A = ST- 3 ST- 4 ST 2 .
At each stage there may be more than one power of T that makes Id I < Ie/
so the process is not unique.
To prove the theorem in general it suffices to consider those matrices

A = (: ;) in r with c ~ O. We use induction on c.


29
2: The modular group and modular functions

If e = 0 then ad = 1 so a = d = ± 1 and

A = (± ~ ±~) = (~ ±~) = T±b.


Thus, A is a power of T.
If e = 1 then ad - b = 1 so b = ad - 1 and

A = (~ ad d- 1) = (~ ~)(~ - ~)(~ ~) = raST d


Now assume the theorem has been proved for all matrices A with lower
left-hand element < e for some e ~ 1. Since ad - be = 1 we have (e, d) = 1.
Dividing d by e we get
d = eq + r, where 0 < r < e.
Then

and

AT-qS = (: -a q
; b)(~ -~) = (-a q
; b =:)
By the induction hypothesis, the last matrix is a product of powers of S
and T, so A is too. This completes the proof. D

2.3 Fundamental regions


Let G denote any subgroup of the modular group r. Two points rand r'
in the upper half-plane H are said to be equivalent under G if r' = Ar for
some A in G. This is an equivalence relation since G is a group.
This equivalence relation divides the upper half-plane H into a disjoint
collection of equivalence classes called orbits. The orbit Gr is the set of all
complex numbers of the form A r where A E G.
We select one point from each orbit; the union of all these points is
called a fundamental set of G. To deal with sets having nice topological
properties we modify the concept slightly and define a fundamental region
of G as follows.

Definition. Let G be a subgroup of the modular group r. An open subset


R G of H is called a fundamental region of G if it has the following two
properties:
(a) No two distinct points of R G are equivalent under G.
(b) If r E H there is a point r' in the closure of R G such that r' is equivalent
to r under G.

30
2.3: Fundamental regions

For example, the next theorem will show that a fundamental region R r
of the full modular group r consists of all r in H satisfying the inequalities
Irl> 1, Ir+il<l.
This region is the shaded portion of Figure 2.1.

u = -1-
r = u + iv, v > 0

/' It I = 1

-------+----r----r---~--_+_---- U
-1 o 1
"2

Figure 2.1 Fundamental region of the modular group

The proof will use the following lemma concerning fundamental pairs of
periods.

Lemma 1. Given WI" w 2' with W2'/W l ' not real, let
n= {mw l ' + nW2' : m, n integers}.
Then there exists afundamental pair (WI' (2) equivalent to (WI" W2') such
that

and such that

PROOF. We arrange the elements of n in a sequence according to increasing


distances from the origin, say
n = {O, Wb W2' ... }
where
O</Wl/~/W21~'" and argw n < argwn+l if /wn/=/Wn+l/'
31
2: The modular group and modular functions

Let WI = WI and let W 2 be the first member of this sequence that is not a
multiple of WI' Then the triangle with vertices 0, W b W 2 contains no element
of Q except the vertices, so (w b ( 2 ) is a fundamental pair which spans the
set Q. Therefore there exist integers a, b, c, d with ad - bc = ± 1 such that

(::) = (: ~)(::)
If ad - bc = -1 we can replace c by -c, d by -d, and WI by -WI and the
same equation holds, except now ad - bc = 1. Because of the way we have
chosen W b W2 we have
and
since WI ± W 2 are periods in Q occurring later than W 2 in the sequence. D
Theorem 2.2. If r' E H, there exists a complex number r in H equivalent to r'
under r such that
Irl ~ 1, Ir+ll~lrl and Ir-ll~lrl.

PROOF. Let WI' = 1, w 2 ' = r' and apply Lemma 1 to the set of periods
Q = {m + nr' : m, n integers}. Then there exists a fundamental pair W b W 2

with Iw2 1 ~ Iwtl, IW t ± w2 1 ~ Iw2 1. Let T = w 2 /w t . Then T = (: ~},


with ad - bc = 1 and
Irl ~ 1, Ir ± 11 ~ Irl. o
Note. Those r in H satisfying Ir ± 11 ~ Ir I are also those satisfying
Ir+rl~l.

Theorem 2.3. The open set


R r = {r E H : Ir I > 1, Ir + r I < I}
is a fundamental region for r. Moreover, if A E r and if Ar = r for some
r in R r , then A = 1. In other words, only the identity element has fixed
points in R r .

PROOF. Theorem 2.2 shows that if r' E H there is a point r in the closure of
R r equivalent to r' under r. To prove that no two distinct points of R r are

equivalent under r, let T' = AT where A = (: ~} We show first that


Im(r') < Im(r) if r E R r and c =1= O. We have

, Im(r)
Im(T) = leT + d1 2 •

32
2.3: Fundamental regions

IfrER r and c =1= 0 we have


Icr + dl =
2
(cr+ d)(cr + d) = c 2rr + cd(r + i) + d2 > c 2 - Icdl + d2 •
If d = 0 we find Icr + d 12 > c 2 2 1. If d =1= 0 we have
c 2 - Icdl + d2 = (Icl - Idl)2 + Icdl 2 Icdl 2 1
so again Icr + d 2 > 1. Therefore c =1= 0 implies cr + d 2 > 1 and hence
1 I 1

Im(r') < Im(r). In other words, every element A of r with c =1= 0 decreases
the ordinate of each point r in R r .
Now suppose both rand r' are equivalent interior points of R r . Then
, ar + b dr' - b
r =--- and r= .
cr + d -cr' + a
If c =1= 0 we have both Im(r') < Im(r) and Im(r) < Im(r'). Therefore c = 0
so ad = 1, a = d = ± 1, and

A=(: ~)=(±~ ±~)=T±b.


But then b = 0 since both rand r' are in R r so r = r'. This proves that no
two distinct points of R r are equivalent under r.
Finally, if Ar = r for some r in R r , the same argument shows that c = 0,
a = d = ± 1, so A = I. This proves that only the identity element has fixed
points in R r . 0

Figure 2.2 shows the fundamental region R r and some of its images under
transformations of the modular group. Each element of r maps circles into
circles (where, as usual, straight lines are considered as special cases of
circles). Since the boundary curves of R r are circles orthogonal to the real

-2 -1 1
-'! o 1
"2 2

Figure 2.2 Images of the fundamental region R r under elements of r

33
2: The modular group and modular functions

axis, the same is true of every image f(R r ) under the elements.f of r. The
set of all images f(R r ), where fEr, is a collection of nonoverlapping open
regions which, together with their boundary points, cover all of H.

2.4 Modular functions

Definition. A functionfis said to be modular if it satisfies the following three


conditions:
(a) f is meromorphic in the upper half-plane H.
(b) f(Ar) = f(r) for every A in the modular group r.
(c) The Fourier expansion off has the form
00

f(r) = L a(n)e21tint.
n= -m

Property (a) states thatfis analytic in H except possibly for poles. Property
(b) states that f is invariant under all transformations of r. Property (c) is
a condition on the behavior offat the point r = ioo. If x = e 21tit the Fourier
series in (c) is a Laurent expansion in powers of x. The behavior offat ioo is
described by the nature of this Laurent expansion near O. If m > 0 and
a( - m) =1= 0 we say that f has a pole of order m at ioo. If m ~ 0 we say f is
analytic at ioo. Condition (c) states that f has at worst a pole of order m at
ioo.
The function J is a modular function. It is analytic in H with a first order
pole at ioo. Later we show that every modular function can be expressed as
a rational function of J. The proof of this depends on the following property
of modular functions.

Theorem 2.4. Iff is modular and not identically zero, then in the closure of the
fundamental region R r , the number of zeros off is equal to the number of
poles.

Note. This theorem is valid only with suitable conventions at the boundary
points of R r . First of all, we consider the boundary of R r as the union of
four edges intersecting at four vertices p, i, P + 1, and ioo, where p = e 21ti / 3
(see Figure 2.3). The edges occur in equivalent pairs (1), (4) and (2), (3).
Iffhas a zero or pole at a point on an edge, then it also has a zero or pole
at the equivalent point on the equivalent edge. Only the point on the leftmost
edge (1) or (2) is to be counted as belonging to the closure of R r.
The order of the zero or pole at the vertex p is to be divided by 3; the order
at i is to be divided by 2; the order at ioo is the order of the zero or pole at
x = 0, measured in the variable x = e 21tit •

34
2.4: Modular functions

ooi
--~--
./
......

/
/ " ",

(1) (4)

(2) (3)

p p + 1

Figure 2.3

PROOF. Assume first that j has no zeros or poles on the finite part of the
boundary of R r . Cut R r by a horizontal line, Im(r) = M, where M > 0 is
taken so large that all the zeros or poles of j are inside the truncated region
which we call R. [Ifjhad an infinite number of poles in R r they would have
an accumulation point at ioo, contradicting condition (c). Similarly, since j
is not identically zero, j cannot have an infinite number of zeros in R r .]
Let oR denote the boundary of the truncated region R. (See Figure 2.4.)
Let Nand P denote the number of zeros and poles ofjinside R. Then

N - P = -1 i
-f'(r) dr = - 1
2ni oR f(r) 2ni
{f (1)
+ f (2)
+ f (3)
+ f
(4)
+ f}
(5)

where the path is split into five parts as indicated in Figure 2.5. The integrals
along (1) and (4) cancel because of periodicity. They also cancel along (2)
and (3) because (2) gets mapped onto (3) with a reversal of direction under

-} + iM , . . - - - - - - - - - - - - , ! + iM

R
t

p p + 1

Figure 2.4

35
2: The modular group and modular functions

(5)

(2) (3)

~ ~

Figure 2.5

the mapping u = S(r) = -l/r, or r = S-l U = S(u). The integrand remains


unchanged because f[S(u)] = f(u) implies f'[S(u)]S'(u) = f'(u) so

f'(r) dr = f'[S(u)] S'(u) du = f'(u) duo


f(r) f[S(u)] f(u)
Thus we are left with

N - P = _1
2ni
f
(5)
f'(r) dr.
f(r)
We transform this integral to the x-plane, x = e21tit • As r varies on the
horizontal segment r = u + iM, -t ~ u ~ t, we have

so x varies once around a circle K of radius e- 21tM about x = 0 in the negative


direction. The points above this segment are mapped inside K, so f has no
zeros or poles inside K, except possibly at x = O. The Fourier expansion
gives us

a-m + ... = F(x),


f (r) = --;n
x
say, with

= F'(x) ~:'
f'(r) dr = F'(x) dx.
f'(r) f(r) F(x)
Hence

N - P =~
2nl
f
(5)
f'(r) dr
f(r)
= - ~,[ F'(x) dx = -(N F
2nl JK F(x)
- PF) = PF - NF,

where N F and PF are the number of zeros and poles of F inside K.

36
2.4: Modular functions

PF
If there is a pole of order m at x =
- N F = m, and
° then m > 0, N F = 0, P F = m so

N = P + m.
°
Therefore jtakes on the value in R r as often as it takes the value 00.
If there is a zero of order n at x = 0, then m = - n so P F = 0, N F = n,
hence
N +n= P.
°
Again,jtakes the value in R r as often as it takes the value 00. This proves
the theorem ifJhas no zeros or poles on the finite part of the boundary of R r .
Ifjhas a zero or a pole on an edge but not at a vertex, we introduce detours
in the path of integration so as to include the zero or pole in the interior of R,
as indicated in Figure 2.6. The integrals along equivalent edges cancel as
before. Only one member of each pair of new zeros or poles lies inside the new
region and the proof goes through as before, since by our convention only
one of the equivalent points (zero or pole) is considered as belonging to the
closure of R r .

Figure 2.6

If j has a zero or pole at a vertex p or i we further modify the path of


integration with new detours as indicated in Figure 2.7. Arguing as above we
find

N - P = -1
2ni
{(II)
+
Ct
+ 1+
C3 C2
/
J,-1 2+iM}j"(7:)
1/2 + iM
-
j'(7:)
d7:

= _1
2ni
{(I + 1)+ 1
Ct C3 C2
}J'(7:) d7: + m
j(7:) ,

where x- m is the lowest power of x occurring in the Laurent expansion near


x = 0, x = e21tit .

37
2: The modular group and modular functions

-1 + iM , . . - - - - - - - - - - - - , t + iM

Figure 2.7
'.
p+ 1

Near the vertex p we write


f(r) = (r - p)kg(r), where g(p) # o.
The exponent k is positive iffhas a zero at p, and negative iffhas a pole at p.
On the path C 1 we write r - p = re i8 where r is fixed and rx ~ ~ nl2 e
where rx depends on r. Then
f'(r) = _k_ + g'(r)
f(r) r - p g(r)
and

-
1 f f'(r)
- - dr = -1 fa (k-. + g'(p + re
iO
. ))re iO'1 de
2ni Ct f(r) 2ni 1t/2 re+ re )
'O
g(p 'O

r fa g'(p + re ) i6 ~
i8
_ -krx' , _ _
- -- + - '0 e dO, where ex - 2 ex.
2n 2n 1t/2 g(p + ref )

As r -+ 0, the last term tends to 0 since the integrand is bounded. Also,


rx' -+ n/3 as r -+ 0 so
.
hm-
r~O
1
2ni
f Ct
f'(r) k
--dr = - - .
f(r) 6
Similarly,

lim _1 f f'(r) dr = - ~
r~O 2ni C3 f(r) 6

38
2.5: Special values of J

so
1
lim-
r-+O 2ni
(J. + J.
Ct C3
)f'(r) dr = - ~
f( r) 3·
Similarly, near the vertex i we write
f(r) = (r - i)'h(r), where h(i) # 0
and we find, in the same way,

lim _1
r-'O 2rci
J. C2
f'(r) dr = -
f( r) 2
i.
Therefore we get the formula
k I
N - P=m -"3 - 2·
Iffhas a pole at x = 0, and zeros at p and i, then m, k and I are positive and
we have
k I
N +"3 + 2 = P + m.
The left member counts the number of zeros offin the closure of R r (with the
conventions agreed on at the vertices) and the right member counts the
number of poles. Iff has a zero of order n at x = 0 then m = - n and the
equation becomes
k I
N + n + 3 + 2 = P.
Similarly, if f has a pole at p or at i the corresponding term k/3 or 1/2 is
negative and gets counted along with P. This completes the proof. D

Theorem 2.5. If f is modular and not constant, then for every complex c the
functionf - c has the same number of zeros as poles in the closure of R r .
In other words,ftakes on every value equally often in the closure of R r .
PROOF. Apply the previous theorem to f - c. D

Theorem 2.6. Iff is modular and bounded in H then f is constant.


PROOF. Sincefis bounded it omits a value sofis constant. D

2.5 Special values of J


Theorem 2.7. The function J takes every value exactly once in the closure of
R r . In particular, at the vertices we have
J(p) = 0, J(i) = 1, J(ioo) = 00.

There is afirst order pole at ioo, a triple zero at p, and J(r) - 1 has a double
zero at r = i.

39
2: The modular group and modular functions

PROOF. First we verify that 92(P) = 0 and 93(i) = o. Since p 3 = 1 and


p2 + P + 1 = 0 we have

~
60 92
(p) _ "
- tn (m + np)4 tn, ,(mp1+ np)4 -_ ~p4 tn" (mp 1+ n)4
3 2

1 1 1 1 1
= PJ,:n (n - m - mpt = PM~N (N + Mp)4 = 60p g2(P),

so g2(P) = O. A similar argument shows that g3(i) = O. Therefore


3 3
J( ) = 92 (P) = 0 and J(.) = 92 (i) = 1
p ~(p) 1 3
92 (i) .

The multiplicities are a consequence of Theorem 2.4. D

2.6 Modular functions as rational functions


of J
Theorem 2.8. Every rational function of J is a modular function. Conversely,
every modular function can be expressed as a rational function of J.
PROOF. The first part is clear. To prove the second, suppose f has zeros at
and poles at PI' ... , Pn with the usual conventions about multi-
Z l' ... , Zn
plicities. Let

g(7:) = Ii J(7:) - J(Zk)


k= 1 J (r) - J (p k)
where a factor 1 is inserted whenever Zk or Pk is 00. Then 9 has the same zeros
and poles asfin the closure of R r , each with proper multiplicity. Therefore
f /9 has no zeros or poles and must be constant, so f is a rational function
ofJ. D

2.7 Mapping properties of J


Theorem 2.7 shows that J takes every value exactly once in the closure
of the fundamental region R r . Figure 2.8 illustrates how R r is mapped by
J onto the complex plane.
The left half of R r (the shaded portion of Figure 2.8a) is mapped onto the
upper half-plane (shaded in Figure 2.8b) with the vertical part of the boundary
mapping onto the real interval ( - 00, 0]. The circular part of the boundary
maps onto the interval [0, 1], and the portion of the imaginary axis v > 1,
u = 0 maps onto the interval (1, + (0). Points in R r symmetric about the
imaginary axis map onto conjugate points in J(R r ). The mapping is con-
formal except at the vertices r = i and r = p where angles are doubled
and tripled, respectively.

40
2.7: Mapping properties of J

J
..
"-
/
P
/
I
"\
\
I
\
I
\
u
-1 0 1
0= J(p) 1 = J(i)

(a) (b)

Figure 2.8

These mapping properties can be demonstrated as follows. On the


imaginary axis in R r we have r = iv hence x = e 21tit = e- 21tv > 0, so the
Fourier series

shows that. J(iv) is real. Since J(i) = 1 and J(iv) ~ + 00 as v ~ + 00 the


portion of the imaginary axis 1 ~ v < + 00 gets mapped onto the real axis
1 ~ J(r) < + 00.
On the left boundary of R r we have r = -1 + iv, hence x = e 21tit =
e-21tve-1ti = _e-
21tv
< O. For large v (small x) we have J( -1 + iv) < 0 so
J maps the line u = -1 onto the negative real axis. Since J(p) = 0 and
J( (0) = 00, the left boundary of R r is mapped onto the line - 00 < J(r) ~ O.
As the boundary of R r is traversed counterclockwise the points inside R r
lie on the left, hence the image points lie above the real axis in the image
plane.
Finally, we show that J takes conjugate values at points symmetric about
the imaginary axis, that is,

J(r) = J(-i).
To see this, write r = u + iv. Then
x = e 21tit = e 21ti (u + iv) = e - 21tv e 21tiu

and

Thus rand - i correspond to conjugate points x and X, but the Fourier


series for J has real coefficients so J(r) and J( - i) are complex conjugates.

41
2: The modular group and modular functions

In particular, on the circular arc Ti = 1 we have - i = - liT, hence


J( - i) = J( -liT) = J(T) so J is real on this arc.

2.8 Application to the inversion problem for


Eisenstein series
In the Weierstrass theory of elliptic functions the periods Wt, W2 determine
the invariants g2 and g3 according to the equations

(4)

A fundamental problem is to decide whether or not the invariants g2 and g3


can take arbitrary prescribed values, subject only to the necessary condition
g2 3 - 27 g 3 2 =1= O. This is called the inversion problem for Eisenstein series
since it amounts to solving the equations in (4) for W t and W2 in terms of g2
and g3. The next theorem shows that the problem has a solution.

Theorem 2.9. Given two complex numbers a2 and a3 such that a2 3 - 27 a3 2 =1= O.
Then there exist complex numbers W t and W2 whose ratio is not real such
that
and
PROOF.We consider three cases: (1) a2 = 0; (2) a3 = 0; (3) a2a3 =1= O.
Case 1. If a2 = 0 then a3 =1= 0 since a2 3 - 27a32 =1= O. Let W t be any
complex number such that

and let W2 = PWt, where p = e 21ti /3. We know that g3(1, p) =1= 0 because
g2(1, p) = 0 and ~(1, p) = g2 3 - 27g 32 =1= O. Then
1
g2(W t , w 2) = g2(W b wtp) = - 4 g2(1, p) = 0 = a2
Wt

and

Case 2. If a3 = 0 then a2 =1= 0 and we take Wt to satisfy

42
2.9: Application to Picard's theorem

and let W2 = iw t . Then


1
g2(W b w 2) = g2(Wb iw t ) = -4 g2(1, i) = a2
Wt

and
1
g3(W b w 2) = g3(Wb iw t ) = -6 g3(1, i) = 0 = a3'
Wt

Case 3. Assume a2 =1= 0 and a3 =1= O. Choose a complex r with 1m r > 0


such that
3
a2
J(r) = 3 27 2'
a2 - a3
Note that J(r) =1= 0 since a2 =1= 0 and that
J(r) - 1 27a32
(5)
J(r) ~.

For this r choose Wt to satisfy

and let W2 = rw t . Then


g2(W t , w 2) _ W t -4g2 (1, r) _ 2 g2(1, r) _ a2
-g-3(-W-,-w--) - W -6
g3 (1, r) - Wt -g3-(1-,-r) - a3'
t 2 t

so

(6)

But we also have


J(r) - 1 27g 32(W t , w 2) 27(a3/a2)2 g2 2(W b w 2)
J(r) 3
g2 (W b W2) g2 3(Wb W2)

Comparing this with (5) we find that g2(W t , W2) = a2 and hence by (6) we
also have g3(Wb W2) = a3' This completes the proof. D

2.9 Application to Picard's theorem


The modular function J can be used to give a short proof of a famous theorem
of Picard in complex analysis.

Theorem 2.10. Every nonconstant entire function attains every complex value
with at most one exception.

Note. An example is the exponential function f(z) = eZ which omits


only the value O.

43
2: The modular group and modular functions

PROOF. We assume f is an entire function which omits two values, say a


and b, a =1= b, and show thatfis constant. Let

_ f(z) - a
g(Z ) - b .
-a
Then g is entire and omits the values 0 and 1.
The upper half-plane H is covered by the images of the closure of the
fundamental region R r under transformations of r. Since J maps the closure
of R r onto the complex plane, J maps the half-plane H onto an infinite-
sheeted Riemann surface with branch points over the points 0, 1 and 00
(the images of the vertices p, i and 00, respectively). The inverse function J- 1
maps the Riemann surface back onto the closure of the fundamental region
R r . Since J'(r) =1= 0 if r =1= p or r =1= i and since J'(p) == J'(i) == 0, each single-
valued branch of J - 1 is locally analytic everywhere except at 0 == J(p),
1 = J(i), and 00 = J( (0). For each single-valued branch of J- 1 the composite
function
h(z) = J- 1 [g(z)]
is a single-valued function element which is locally analytic at each finite
z since g(z) is never 0 or 1. Therefore h is arbitrarily continuable in the entire
finite z-plane. By the monodromy theorem, the continuation of h exists as a
single-valued function analytic in the entire finite z-plane. Thus h is an entire
function and so too is
qJ(z) = eih(z).

But 1m h(z) > 0 since h(z) E H so


IqJ(z) I == e-Imh(z) < 1.
Therefore qJ is a bounded entire function which, by Liouville's theorem,
must be constant. But this implies h is constant and hence g is constant since
g(z) = J[h(z)]. Thereforefis constant sincef(z) == a + (b - a)g(z). D

Exercises for Chapter 2


In these exercises, r denotes the modular group, Sand T denote its gen-
erators, S(r) = -l/r, T(r) = r + 1, and I denotes the identity element.
1. Find all elements A of r which (a) commute with S; (b) commute with ST.

2. Find the smallest integer n > 0 such that (ST)" = I.

3. Determine the point r in the fundamental region R r which is equivalent to


(a) (8 + 60/(3 + 20; (b) (10i + 11)/(6i + 12).

4. Determine all elements A of r which leave i fixed.

5. Determine all elements A of r which leave p = e21ti / 3 fixed.

44
Exercises for Chapter 2

QUADRATIC FORMS AND THE MODULAR GROUP

The following exercises relate quadratic forms and the modular group r. We
consider quadratic forms Q(x, y) = ax 2 + bxy + cy2 in x and y with real
coefficients a, b, c. The number d = 4ac - b2 is called the discriminant of
Q(x, y).
6. If x and yare subjected to a unimodular transformation, say

(1) x= lXX' + Py', y = yx' + 8y', where (; ~) E r,


prove that Q(x, y) gets transformed to a quadratic form Ql(X', y') having the same
discriminant. Two forms Q(x, y) and Ql(X', y') so related are called equivalent. This
equivalence relation separates all forms into equivalence classes. The forms in a given
class have the same discriminant, and they represent the same integers. That is, if
Q(x, y) = n for some pair of integers x and y, then Ql(X', y') = n for the pair of
integers x', y' given by (1).
In Exercises 7 thru 10 we consider forms ax 2 + bxy + cy2 with d > 0,
a > 0, and c > O. The associated quadratic polynomial
f(z) = az 2 + bz + c
has two complex roots. The root T with positive imaginary part is called the
representative of the quadratic form Q(x, y) = ax 2 + bxy + cy2.
7. (a) If d is fixed, prove that there is a one-to-one correspondence between the set
of forms with discriminant d and the set of complex numbers r with Im(r) > o.
(b) Prove that two quadratic forms with discriminant d are equivalent if and only if
their representatives are equivalent under r.

Note. A reduced form is one whose representative TERr. Thus, two


reduced forms are equivalent if and only if they are identical. Also, each
class of equivalent forms contains exactly one reduced form.
8. Prove that a form Q(x, y) = ax 2 + bxy + cy2 is reduced if, and only if, either
-a < b s a < cor 0 S b s a = c.
9. Assume now that the form Q(x, y) = ax 2 + bxy + cy2 has integer coefficients
a, b, c. Prove that for a given d there are only a finite number of equivalence classes
with discriminant d. This number is called the class number and is denoted by h(d).
Hint: Show that 0 < a S Jdi3 for each reduced form.
10. Determine all reduced forms with integer coefficients a, b, c and the class number
h(d) for each d in the interval 1 s d s 20.

CONGRUENCE SUBGROUPS

The modular group r has many subgroups of special interest in number


theory. The following exercises deal with a class of subgroups called con-
gruence subgroups. Let

and

45
2: The modular group and modular functions

be two unimodular matrices. (In this discussion we do not identify a matrix


with its negative.) If n is a positive integer write
A == B (mod n) whenever a == rJ., b == {3, c == y and d == {) (mod n).
This defines an equivalence relation with the property that
and
implies
AlB l == A 2 B 2 (mod n) and Al -
l
== A 2 -1 (mod n).
Hence
A == B (mod n) if, and only if, AB- l == I (mod n),
where I is the identity matrix. We denote by r(n) the set of all matrices in r
congruent modulo n to the identity. This is called the congruence subgroup
of level n (stufe n, in German).
Prove each of the following statements:
11. r(n) is a subgroup of r. Moreover, if BE r(n) then A -1 BA E r(n) for every A in r.
That is, r<n) is a normal subgroup of r.
12. The quotient group r /r<n) is finite. That is, there exist a finite number of elements of
r, say A 1, ... , A k , such that every B in r is representable in the form
B = AiB<n) where 1 sis k and B<n) E r<n).
The smallest such k is called the index of r<n) in r.
13. The index of r<n) in r is the number of equivalence classes of matrices modulo n.

The following exercises determine an explicit formula for the index.


14. Given integers a, b, c, d with ad - bc == 1 (mod n), there exist integers Ct, [3, y, b
such that Ct == a, [3 == b, y = c, b == d (mod n) with ab - [3y = 1.
15. If (m, n) = 1 and A E r there exists A in r such that
A == A (mod n) and A == I (modm).

16. Letf(n) denote the number of equivalence classes of matrices modulo n. Thenfis a
multiplicative function.
17. If a, b, n are integers with n ~ 1 and (a, b, n) = 1 the congruence
ax - by == 1 (mod n)
has exactly n solutions, distinct mod n.(A solution is an ordered pair (x, y) of integers.)
18. For each prime p the number of solutions, distinct mod pr, of all possible congruences
of the form
ax - by == 1 (mod pr), where (a, b, p) = 1,
is equal to f(pr).
19. If p is prime the number of pairs of integers (a, b), incongruent mod pr, which satisfy
the condition (a, b, p) = 1 is p2r- 2(p2 - 1).
20. f(n) = n 3 Ldln f.l(d)/d 2 , where f.l is the Mobius function.

46
The Dedekind eta function

3.1 Introduction
In many applications of elliptic modular functions to number theory the
eta function plays a central role. It was introduced by Dedekind in 1877
and is defined in the half-plane H = {r: Im(r) > O} by the equation
00

(1) 1J(r) = e1tit/12 Il (1 - e21tint).


n=1
The infinite product has the form Il
(1 - x n) where x = e 21tit . If r E H then
Ix I < 1 so the product converges absolutely and is nonzero. Moreover,
since the convergence is uniform on compact subsets of H, 1J(r) is analytic
onH.
The eta function is closely related to the discriminant L\(r) introduced
in Chapter 1. Later in this chapter we snow that
L\(r) = (2n)121J24(r).
This result and other properties of 1J(r) follow from transformation formulas
which describe the behavior of 1J(r) under elements of the modular group r.
For the generator Tr = r + 1 we have
00

(2) 1J(r + 1) = e1ti(t+ 1)/12 Il (1 - e 21tin (t+ 1») = e1ti /12 1J(r).
n=1
Consequently, for any integer b we have

(3) 1J(r + b) = e1tib/121J(r).


Equation (2) also shows that 1J24(r) is periodic with period 1.

47
3: The Dedekind eta function

For the other generator ST = -l/T we have the following theorem.


Theorem 3.1. If TE H we have

(4) '1( ~ 1) = (- h:)1/2'1(-r).

Note. We choose that branch of the square root function Z1/2 which is
positive when z > o.

This chapter gives two different proofs of (4). The first is a short proof of
C. L. Siegel [48] based on residue calculus, and the second derives (4) as a
special case of a more general functional equation which relates

r,(~)
+d CT
to r,(T) when

(: ~) E rand c > O.

(See Theorem 3.4.) A third proof, based on interchange of summation in a


conditionally convergent iterated series, is outlined in the exercises.

3.2 Siegel's proof of Theorem 3.1


First we prove (4) for T = iy, where y > 0, and then extend the result to all
Tin H by analytic continuation. If T = iy the transformation formula becomes
r,(i/y) = yl/2r,(iy), and this is equivalent to
log r,(i/y) - log r,(iy) = ! log y.
Now
log '1(iy) = -
ny
12 + log JI
00

(1 - e-
2nny
)

ny ny 00 00 e- 21tmny
L
00
- - +
12 n=1
log(l - e- 21tny ) = ---
12 LL
n= 1 m= 1 m
ny 00 1 e- 21tmy ny 00 1 1
- 12 - m~l;;; 1 - e- 2nmy -12 + m~l;;; 1 - e2nmy '
Therefore we are to prove that

1 1 lIn ( 1) 1
m~ 1 ;;; 1 - e2nmy m~ 1 ;;; 1 - e2nm/ y -
00 00

(5) - 12 Y - Y = - 2 log y.
This will be proved with the help of residue calculus.
For fixed y > 0 and n = 1,2, ... , let
1 nNz
F n(z) = - -8z cot niNz cot -y- ,

48
3.2: Siegel's proof of Theorem 3.1

-y o y

-i

Figure 3.1

where N = n + !. Let C be the parallelogram joining the vertices y, i, - y, - i


in that order. (See Figure 3.1.) Inside C, F n has simple poles at z = ik/N and
at z = ky/N for k = ± 1, ±2, ... , ±n. There is also a triple pole at z = 0
with residue i(y - Y- 1 )/24. The residue at z = ik/N. is
1 nik
8nk cot y.
Since this is an even function of k we have
n
n 1 nik
L Res Fn(z) = 2 L 8 k cot - .
k= -n z = ik/N k= 1 n Y
k,*O

But

cot
.0 _ cos iO _ . e- 8 + e8
1 - • .0 - 1 -8
sIn 1 e - e
8 - i e 28
e
28

_
1 1(1 - 1 _2)
+ 1= i e 28 •

Using this with () = nk/y we get


n 1 n 1 1 n 1 1
L
k= -n
Res F (z) = -
z=ik/N n 4ni
L -k - -2ni L ------,,---
k= 1 k 1- e k= 1
21tk Y
/ •
k,*O

Similarly
n in 1 i n1 1
L
k= -n
Res F (z) = - L - - - L --~
z=ky/N n 4n k= 1 k 2n k= lk 1 - e 21tkY •
k,*O

Hence 2ni times the sum of all the residues of Fn(z) inside C is an expression
whose limit as n ~ 00 is equal to the left member of(5). Therefore, to complete
the proof we need only show that

lim
n-+ 00
f
C
Fn(z) dz = -! log y.
49
3: The Dedekind eta function

On the edges of C (except at the vertices) the function zFn(z) has, as


n the limit -k on the edges connecting y, i and - y, - i, and the limit -t
-+ 00,
on the other two edges. Moreover, Fn(z) is uniformly bounded on C for all n
(because N = n + ! and y > 0). Hence by Arzehi's bounded convergence
theorem (Theorem 9.12 in [3]) we have f

lim
n-+oo
fC
Fn(z) dz = f lim zFn(z) dz
C n-+oo Z

=~{-fY.+Ji- ~-Y +f-i}dZ


8 -I y Ji _y Z

= ~ {- fY + Ji} dz
4 -I Y Z

= 4:1 {-(log y + 2 + ni) (ni2 - log y)} = - 21 log y.

This completes the proof. D

3.3 Infinite product representation for ~(L)

In this section we express the discriminant L\(T) in terms of 1J(T) and thereby
obtain a product representation of L\(T). The result makes use of the following
property of L\(T).

Theorem 3.2. If (: ~) E r then

d(:: : ~) = (CT + d) 12 d(T).


In particular,

L\(T + 1) = L\(T) and

PROOF. Since L\(w l , w 2) is homogeneous of degree -12 we have


L\(w b w 2) = WI - 12 L\(1, T) = WI - 12 L\(T),

where T = W2/Wl. Also,


L\(W b w 2) = L\(w l ', w 2')
if (Wb w 2 ) and (WI" w 2 ') are equivalent pairs of periods. Taking WI = 1,
W2 = T, WI' =
CT +
d, w 2 ' = aT + b, we find

L\(T) = L\(w b w 2) = L\(CT + d, aT + b) = (CT b)


aT +
+ d)-12L\ ( 1, - -. D
CT + d

50
3.4: The general functional equation for l1(r)

Theorem 3.3. If r E H and x = e 21tit we have

(6) L\(r) = (2n)121]24(r) = (2n)12 x n(1 -


00

n=1
X n)24.

Consequently,

L r(n)xn = x n(1
00 00

(7) - X
n)24 whenever Ix I < 1
n= 1 n= 1

where r(n) is Ramanujan's taufunction.

PROOF. Let f(r) = L\(r)/1] 24(r). Then f(r + 1) = f(r) and f( -l/r) = f(r),
so f is invariant under every transformation in r. Also, f is analytic and non-
zero in H because L\ is analytic and nonzero and 1] never vanishes in H.
Next we examine the behavior of f at ioo. We have

1]24(r) = e 21tit n(1 -


00

n=1
e21tint)24 = x n(1 -
00

n=1
X n)24 = x(l + I(x)),

where I(x) ,denotes a power series in x with integer coeffil ients. Thus, 1]24(r)
has a first order zero at x = 0. By Theorem 1.19 we also have the Fourier
expansion
00

(8) L\(r) = (2n)12 L r(n)x n = (2n)12 x (1 + I(x)).


n=1
Thus, near ioo the function f has the Fourier expansion

(9) f( ) = L\(r) = (2n)12 x (1 + I(x)) = (2 )12(1 I())


r 1]24(r) x(l + I(x)) n + x,
so f is analytic and nonzero at ioo. Therefore f is a modular function which
never takes the value 0, so f must be constant. Moreover, (9) shows that
this constant is (2n)12, hence L\(r) = (2n)121]24(r). This proves (6), and (7)
follows from (8). 0

3.4 The general functional equation for 1J(r)


Extracting 24th roots in the relation

L\(ar
cr
+ =
+d
b) (cr + d)12L\(r)
and using (6) we find that

11(:: : ~) = e(cr + d)1/2 11(r),


where e24 = 1. For many applications of 1](r) we require more explicit
information concerning e. This is provided in the next theorem.

51
3: The Dedekind eta function

Theorem 3.4 (Dedekind's functional equation). If (: :) E r, c > 0, and


,. E H, we have

(10) '1(:: : ~) = s(a, b, c, d){ - i(cr + dW/ 2'1(r)


where
s(a, b, c, d) = exp{ n{a l;cd + s( -d, C))}
and

(11) s(h, k) = L -kr (hr


k-l
r= 1
-k - -21) .
-k - [hrJ
Note. The sum s(h, k) in (11) is called a Dedekind sum. Some of its properties
are discussed later in this chapter.

We will prove Theorem 3.4 through a sequence of lemmas. First we note


that Dedekind's formula is a consequence of the following equation, obtained
by taking logarithms of both members of (10),

(12) log ( aT
cr'1 +d
+ b) = log + d + s( - d, c) ) +1 log{ -
'1( r) + ni(a----u;;- i(cr + d)}.
From the definition of l1(T) as a product we have
niT niT
L log(1 - e 21tlnt ) = - -
00 • 00
(13) log l1(T) =- + L A( -inT),
12 n=l 12 n=l

where A(X) is defined for Re(x) > 0 by the equation


00 e-21tmx
(14) A(X) = -log(1 - e- 21tx ) = L --.
m=l m
Equations (12) and (13) give us

Lemma 1. Equation (12) is equivalent to the relation

(15) 00
LA(-inT) = LA00 ( aT +
-in -- b) + -ni ( T ---
aT +b)
n =1 n =1 cT + d 12 CT +d
a + d + s( -d, c) ) +! log{ -
+ ni( ----u;;- i(cr + d)}.

We shall prove (15) as a consequence of a more general transformation


formula obtained by Sh6 Iseki [17] in 1957. For this purpose it is convenient
to restate (15) in an equivalent form which merely involves some changes
in notation.

52
3.5: Iseki's transformation formula

Lemma 2. Let z be any complex number with Re(z) > 0, and let h, k and H be
any integers satisfying (h, k) = 1, k > 0, hH == -1 (mod k). Then Equation
(15) is equivalent to the formula

(16) f A{~k
n= 1
(z - ih)} = f A{~k (~z -
n= 1
iH)}

n (z -
+ 21 log z - 12k 1) +
~ nis(h, k).

PROOF. Given (: ~) in r, with c > 0, and given r with Im(r) > 0, choose
z, h, k, and H as follows:
k = c, h = -d, H = a, z = -i(cr + d).
Then Re(z) > 0, and the condition ad - bc = 1 implies - hH - bk = 1, so
(h, k) = 1 and hH == -1 (mod k). Now b = -(hH + l)jk and iz = cr + d,
so
iz - d iz +h
r=--=--
c k
and hence

iz +h hH + 1 iz ( i)
ar +b= H -k- - k =k H + -; .
Therefore, since cr + d = iz, we have
ar
cr
+b= ~
+d k
(H + ~). z
Consequently

ar
r - cr
+ db = k1 (h
+ - H) + ki ( z 1) a +d
- ~ = - -c- + ki ( z 1)
- ~

so

;~ (r -::: ~) = -n{a 1;C ) - l~k (z - D·


d

Substituting these expressions in (15) we obtain (16). In the same way we


find that (16) implies (15). D

3.5 Iseki's transformation formula


Theorem 3.5 (Iseki's formula). If Re(z) > ° ° and ~ (1 ~ 1, ° ~ f3 ~ 1, let
00

(17) 1\((1, {3, z) = L {A((r + (1)z - i{3) + A((r + 1 - (1)z + i{3)}.


r=O

53
3: The Dedekind eta function

Then if either 0 :::; a :::; 1 and 0 < f3 < 1, or 0 < a < 1 and 0 :::; f3 :::; 1,
we have

Note. The sum on the right of (18), which contains Bernoulli polynomials
Bn(x), is equal to

PROOF. First we assume that 0 < a < 1 and 0 < f3 < 1. We begin with the
first sum appearing in (17) and use (14) to write
00 00 00 e21timp
(19) L ,1((r + a)z - if3) = L L - - e- 21tm (r+a)z.
r=O r=Om=l m

Now we use Mellin's integral for e- x which states that

1 fC+ ooi
(20) e- x = -2. r(s)x- S ds,
nl c- ooi

where c > 0 and Re(x) > O. This is a special case of Mellin's inversion formula
which states that, under certain regularity conditions, we have

1 fC + ooi

f
oo
q>(s) = XS-1tjJ(x) dx if, and only if, tjJ(x) = -2. q>(s)x- Sds.
o nl c- ooi

In this case we take q>(s) to be the gamma function integral,

ns) = LX' x S
- le- X dx

and invert this to obtain (20). (Mellin's inversion formula can be deduced
from the Fourier integral theorem, a proof of which is given in [3]. See also
[49], p. 7.) Applying (20) with x = 2nm(r + a)z and c = 3/2 to the last
exponential in (19) and writing f(c) for f~= ~~ we obtain

00

L ,1((r + a)z - if3) =


00 00

L L - - -2'
e21timP 1 f r(s){2nm(r + a)z} -s ds
r=O

= -.
1 f
r=O m=l m
r(s)
--S
nl

L
00
(3/2)

1
S
00 e21timp
L ----r+s ds
2nl (3/2) (2nz) r=O (r + a) m=l m

= -2.
1 f r(s)
-(2)S '(s, a)F(f3, 1 + s) ds.
nl (3/2) nz

54
3.5: Iseki's transformation formula

Here ((s, a) is the Hurwitz zeta function and F(x, s) is the periodic zeta function
defined, respectively, by the series
oc 1 00 e21timx

((s, tX) = L (r + a )" and F(x, s) = L -S-


r=O m=l m

where Re(s) > 1, 0 < a S 1, and x is real. In the same way we find

Loc A((r + 1 -
r= 0
a)z
1
+ i{3) = -2.
1tl
f(3/2)
r(s)
-(2)S ((s, 1 - a)F(1 - {3, 1 + s) ds,
1rZ

so (1 7) becomes

(21)
1
A(a, {3, z) = -2.
1tl
f
(3/2)
z - s<I>(a, {3, s) ds,

where
r(s)
(22) <1>(tX, 13, s) = (211:)' {((s, tX)F(f3, 1 + s) + ((s, 1 - tX)F(1 - 13, 1 + s)}.
Now we shift the line of integration from c = ~ to c = -1. Actually, we
apply Cauchy's theorem to the rectangular contour shown in Figure 3.2,

! + iT ~
! + iT

~
0

---..
1
!- iT t - iT

Figure 3.2

and then let T ---. 00. In Exercise 8 we show that the integrals along the
horizontal segments tend to 0 as T ---. 00, so we get

I3/2) - I-3/2) + R
where R is the sum of the residues at the poles of the integrand inside the
rectangle. This gives us the formula
1
A(a, {3, z) = -2.
1rl
f(- 3/2)
z -s<I>(a, {3, s) ds + R.

55
3: The Dedekind eta function

In this integral we make the change of variable u = - s to get it back in


the form of an integral along the 1 line. This gives us

(23) 1\(0:, {3, z) =


1
-2.
n1
f
(3/2)
zU<1>(o:, {3, - u) du + R.

Now the function <1> satisfies the functional equation


(24) <1>(0:, {3, - s) = <1>( 1 - {3, 0:, s).
This is a consequence of Hurwitz's formula for ((s, 0:) and a proof is outlined
in Exercise 7. Using (24) in (23) we find that
(25) 1\(0:, {3, z) = 1\(1 - {3, 0:, Z- 1) + R.
To complete the proof of Iseki's formula we need to compute the residue
sumR.
Equation (22) shows that <1>(0:, {3, s) has a first order pole at each of the
points s = 1, °
and -1. Denoting the corresponding residues by R(1),
R(O) and R( - 1) we find
r( 1) 1 oc (2Trinp - 2Trin p )
R(1) = -2- {F({3, 2) + F(1 - {3,2)} = -2 L ~ + _e- 2 -
nz nz n= 1 n n
1 oc e2Trinp 1 - (2n0 2 n
= -2
nz n
L - 2 - = -2
= - oc n
2'
nz.
B 2 ({3) = - B 2 ({3),
z
n*O

where we have used Theorem 12.19 of [4J to express the Fourier series as a
Bernoulli polynomial.
To calculate R(O) we recall that ((0,0:) = 1- 0:. Hence ((0, 1 - 0:) = 0: - 1
so
oc e2Trinp _ e - 2Trinp
R(O) = ((0, o:)F({3, 1) + ((0, 1 - o:)F(l - {3, 1) = (1 - 0:) L ----
n= 1 n
00 e2Trinp oc e2Trinp
= (1 - 0:) L -- = -B1(0:) L - - = 2niB 1 (0:)B 1 ({3),
n=-oo n n=-oc n
n*O n*O

where again we have used Theorem 12.19 of [4]. To calculate R( -1) we


write

R( -1) = Res z-s<1>(o:, {3, s) = lim (s + 1)z-s<1>(0:, {3, s)


s=-l s~-l

= lim( - s + 1)zs<1>( 0:, {3, - s).


s~l

Using the functional equation (24) we find

R(-1) = lim(1 - s)zs<1>(1 - {3, 0:, s) = -Res zS<1>(1 - {3, 0:, s).
s= 1

56
3.5: Iseki's transformation formula

Note that this is the same as R(I) == Res s = 1 z-S<l>(o:, {3, s), except that z is
replaced by - z - 1 , 0: by 1 - {3, and {3 by 0:. Hence we have
R( -1) == - nzB 2 (0:).
Thus

R = R( -1) + R(O) + R(l) = -nznto G)(iZ)-nB1-n(IX)Bi{3).

This proves Iseki's formula under the restriction a < 0: < 1, a < {3 < 1.
Finally, we use a limiting argument to show it is valid if a ~ 0: ~ 1 and
a < {3 < 1, or if a ~ {3 ~ 1 and a < 0: < 1. For example, consider the series
00 ex.; oc e2nimf3
L A((r + ex)z - i{3) == L L - - e- 2nm(r+a)z
r=O r=Om=l m
ex 2nimf3
L e-2nmrz
00
== I _e__ e-2nmaz
m=l m r=O

say, where
1 e- 2nmaz
.fa(m) == - 1 - 2nmz·
m -e

As m ~ 00, .fa(m) ~ a uniformly in 0: if a s 0: S 1. Therefore the series

00

L e2nimf3fa(m)
m=l

converges uniformly in 0: if a :::; 0: :::; 1, provided a< {3 < 1, so we can pass


a
to the limit 0: ~ + term by term. This gives us
00 ex.;

lim L A((r + o:)z - i{3) == L A(rz - i{3).


a-O+ r=O r=O

Therefore, if a < {3 < 1 we can let 0: ~ a+ in the functional equation. The


other limiting cases follow from the invariance of the formula under the
following replacements:

0: ~ 1- 0:, {3~1-{3

1
0: ~ {3, {3 ~ 1- 0:, z~-
Z

1
0: ~ 1 - {3, {3 ~ 0:, z ~-. D
z

57
3: The Dedekind eta function

3.6 Deduction of Dedekind's functional


equation from Iseki's formula
Now we use Iseki's formula to prove Equation (16) of Lemma 2. This, in turn,
will prove Dedekind's functional equation for 11(r).
Equation (16) involves integers hand k with k > O. First we treat the case
k = 1 for which Equation (16) becomes

(26) I A{n(z -
n= 1
ih)} = I A{n(~ - iH)} + ~2 log z - ~12 (z - ~).
n= 1 Z z
Since A(X) is periodic with period i this can be written as

(27) 00
- + -1log z -
L A(nz) = L A(n) 00 1)
-re ( z - - .
n=1 n=1 Z 2 12 z
We can deduce this from Iseki's formula (18) by taking f3 = 0 and letting
a ---+ 0 +. Before we let a ---+ 0 + we separate the term r = 0 in the first term
of the series on the left of (18) and in the second term of the series on the right
of (18). The difference of these two terms is A(az) - A(ia). Each of these tends
to 00 as a ---+ 0 + but their difference tends to a finite limit. We compute this
limit as follows:
. 1 - e-2n~
A(CXZ) - A(iCX) = log(l - e- Z"IlX) - log(l - e-z"",Z) = log 1 _ e-z"",z'

By L'Hopital's rule,

· 1 - e - 2ni(X 1· 2 rel
· ·l
11m _ 2n(Xz = 1m - =-
(X-+O + 1 - e (X-+O 2rez z

so

lim (A(CXZ) - A(icx)) = log ~ = ni - log Z.


(X-+O+ Z 2
Now when a ---+ 0+ the remaining terms in each series in (18) double up
and we obtain, in the limit,

rei
(28) - - log z + 2 L A(rz) = 2 L A(r)
00 rez re rei
- - - + - + -.
00

2 r= 1 r =1 Z 6 6z 2

This reduces to (27) and proves (16) in the case k = 1.


Next we treat the case k > 1. We choose rational values for a and f3 in
Iseki's formula (18) as follows. Take

Jl
a = k' where 1 ~ Jl ~ k - 1

58
3.6: Deduction of Dedekind's functional equation from Iseki's formula

and write

hJ1 = qk + v, where 1 ~ v ~ k - 1.

Now let

Note that v == hJ1 (mod k) so - Hv == - HhJ1 == J1 (mod k), and therefore


-Hv/k == J1/k (mod 1). Hence lJ. = J1/k == -Hv/k (mod 1) and f3 = v/k ==
hJ1/k (mod 1). Substituting in Iseki's formula (18) and dividing by 2 we get

Rewrite this as follows:

= ~ 00 { I(rk + V)G - iH)) (rk + k- V)G - iH))}


2 Jo A\ k +A k

Now sum both sides on J1 for J1 = 1, 2, ... , k - 1 and note that

{rk + J1:r = 0,1,2, ... ; J1 = 1,2, ... ,k - I} = {n:n =1= °


(modk)}

and similarly for the set of all numbers rk + k - J1. Also, since v == hJ1 (mod k),
as J1 runs through the numbers 1, 2, ... , k - 1 then v runs through the same

59
3: The Dedekind eta function

set of values in some other order. Hence we get

LCX) (n
A - (z - zh) = .) LCX) (n (1 .)) + -n(1- - z) L
A -: - - zH k- 1 J12
2
n= 1 k n= 1 k z 2 z p.= 1 k
n ~ 0 (mod k) n 'i 0 (mod k)

- -n - - z
2 Z
(1 ) L -k + -n (1- - )
k -1 J1

p.=1 12 z
z (k - 1)

+ ni L1- (V- - -1) - - L1V


k- J1
- +- ni k- ni
(k - 1)
p.= 1 k k 2 2 p.= 1 k 4

X
(
(k - 1) (2k - 1)
- 3(k - 1) + (k - 1)
)
+ ni L - (V- - -1)
k - 1 J1

k p.= 1 k k 2

I
n= 1
A(~k (~z - iH)) + ~
12
(z - ~)(1
z
~)
k
+ l/f !!. (~ - ~).
k k 2
- p.= 1
n$.O (mod k)

But V was defined by the equation hJ1 = qk + V, so we have


hJ1
k = q + k'
V
q=[~} ~= hJ1 _ [h J1 ]
k k k·
Therefore

L -J1 (V- - -
k-1
p.= 1 k k
1) =
2
L -J1(h- J1 - [h- J1 ] - -1) = s(h, k).
h-1
p.= 1 k k k 2
Therefore we have proved that

(29)
n~_CX)l
n 1- 0 (mod k)
A(-kn (z'- ih)) = n= 1
f:
n 1= 0 (mod k)
A(~k (~z - iH))

Add this to Equation (27) which corresponds to the case k = 1:

LCX) A(mz) = LCX) A(m)


- - - n( 1) + -log
z - - 1 z.
m= 1 m= 1 Z 12 z 2
This accounts for the missing terms in (29) with n == 0 (mod k), if we write
n = mk. When (27) is combined with (29) we get

f A(~k (z - ih)) = f A(~k (~z - iH)) - ~


n=1 12k
(z - ~)z + ~2 log z +
n= 1
nis(h, k).
60
3.7: Properties of Dedekind sums

This proves (16) which, in tum, completes the proof of Dedekind's functional
equation for Tl(T). For alternate proofs see p. 190 and [18], [35], and [45]. D

3.7 Properties of Dedekind sums


The Dedekind sums s(h, k) which occur in the functional equation for YJ(r)
have applications to many parts of mathematics. Some of these are described
in an excellent monograph on Dedekind sums by Rademacher and Grosswald
[38]. We conclude this chapter with some arithmetical properties of the sums
s(h, k) which will be needed later in this book. In particular, Theorem 3.11
plays a central role in the study of the invariance of modular functions under
transformations of certain subgroups of r, a topic discussed in the next
chapter.
Note. Throughout this section we assume that k is a positive integer and
that (h, k) = 1.
Dedekind sums are defined by the equation

(30) s(h, k) = kt1!.-k (hrk - [hrJ


r= 1 k
_~).
2
First we express these sums in terms of the function ((x)) defined by

((x) )-_{x0 - [x] -! ·ifI xfisIS·annot·Integer.


X
an integer,

This is a periodic function of x with period 1, and (( - x)) = - ((x)). Actually,


((x)) is the same as the Bernoulli periodic function B1 (x) discussed in [4],
Chapter 12. Since ((x)) is periodic and odd we find that

L ((~))
r mod k k
- 0

and, more generally,

L ((hr)) - 0 for (h, k) = 1.


r mod k k
Since

the Dedekind sums can now be represented as follows:

(31) s(h,k) = rm~k ((i))((h:)).


This representation is often more convenient than (30) because we can exploit
the periodicity of ((x)).

61
3: The Dedekind eta function

Theorem 3.6
(a) If h' == ± h (mod k), then s(h', k) = ± s(h, k), lvith the same sign as
in the congruence. Similarly, lve have:
(b) If hh == ± 1 (mod k) then s(h, k) = ± s(h, k).
(c) ~f h 2 + 1 == 0 (mod k), then s(h, k) = O.

PROOF. Parts (a) and (b) follow at once from (31). To prove (c) we note that
h2 + 1 == 0 (mod k) implies h == -h (mod k), where h is the reciprocal of
h mod k, so from (a) and (b) we get s(h, k) = -s(h, k) = O. D

For small values of h the sum s(h, k) can be easily evaluated from its
definition. For example, when h = 1 we find

s(l, k) =
k- 1
L - -- -
r(r 1) = 21 L r k- 1 2
- -
1L r
k- 1

r = 1 k k 2 k r = 1 2k r= 1

(k - 1)(2k - 1) k - 1 (k - 1)(k - 2)
---
6k 4 12k

Similarly, the reader can verify that

2 k) = (k - 1)(k - 5) if k is odd.
s( , 24k

In general there is no simple formula for evaluating s(h, k) in closed form.


However, the sums satisfy a remarkable reciprocity law which can be used
as an aid in calculating s(h, k).

3.8 The reciprocity law for Dedekind sums


Theorem 3.7 (Reciprocity law for Dedekind sums). If h > 0, k > 0 and
(h, k) = 1 we have

12hks(h, k) + 12khs(k, h) = h 2 + k 2 - 3hk + 1.

PROOF. Dedekind first deduced the reciprocity law from the functional
equation for log '1(7:). We give an arithmetic proof of Rademacher and
Whiteman [39], in which the sum L~= 1 ((hr/k))2 is evaluated in two ways.
First we have

(32) L ((hr))2
k
r= 1
-k = L r mod k
-k = r mod
((hr))2 L k ((r))2
-k = k-l L (r---
r= 1
1)2
k 2 .

62
3.8: The reciprocity law for Dedekind sums

We can also write

I
r= 1
((hr))2 =
k
kil (hrk _ [hrJ
r= 1 k
_ ~)2
2
1
= ki (h2~2 + [hrJ2 + ~ _ hr + [hrJ _ 2hr [hrJ)
r= 1 k k 4 k k k k

~)
k
= 2h
r= 1
f !:.- (hr _ [hrJ _
k k k 2
2

+L [hrJ([hrJ
-k -k +1 ) -2Lr
k-1

r= 1
h
k 4
1
+-2:1. k-1 2

r= 1
k-1

r= 1

Comparing this with (32) and using (30) we obtain

2 k
k- 1 [hrJ ([hrJ ) h + 1k- 1 2 1 - 1
(33) 2hs(h,k) + ~ -k -k + 1 = - k2 - ~ r - -k ~ r.
r-1 r-1 r-1

In the sum on the left we collect those terms for which [hr /k] has a fixed value.
Since 0 < r < k we have 0 < hr/k < h and we can write

(34) [~J= v - 1, where v = 1, 2, ... , h.

For a given v let N(v) denote the number of values of r for which [hr/k] =
v - 1. Equation (34) holds if, and only if

hr k(v - 1) kv
v-1<-<v or h < r < h'
k '

equality being excluded since (h, k) = 1 and 0 < r < k. Therefore, if


1 ::; v ::; h - 1, Equation (34) holds when r ranges from [k(v - 1)/h] + 1
to [kv/h], and hence

N(v) = [k; J- [k(V ~ 1) J if 1 ::; v ::; h - 1.

But when v = h the quotient kv/h = k and since r = k is excluded we have

N(h) =k_ 1_ [k(h; 1)]


63
3: The Dedekind eta function

Hence

(35) :t: [h{] ([h{] + 1) = vt (v - l)vN(v)

vt (v - l)v([k:] - [k(V ~ 1)]) - h(h - 1)


= V~l
h- 1 [k ] {(v -
: l)v - v(v + I)}
+ kh(h - 1) - h(h - 1)

[k ]
L v ~ + h(h - 1)(k - 1).
= -2 h-1
v= 1 h
Now we also have

2hs(k, h) = 2 L v(kV
h- 1
- -
[kVJ
-
1)
h- 1 [kV]
- - = -2 L v - + -2k h-L1 v2 - LV
h- 1

\'=1 h h 2 v=1 h h v=1 v=1

so (35) becomes

[h ] ([h ] ) 2k
r~l = 2hs(k,h) - h V~lV2 + V~lV + h(h - l)(k - 1).
k- 1 h- 1 h- 1
{ { + 1

We use this in (33) and multiply by 6k to obtain the reciprocity law. D

3.9 Congruence properties of Dedekind sums


Theorem 3.8. The number 6ks(h, k) is an integer. Moreover, if 0 = (3, k) we have
(a) 12hks(k, h) == 0 (mod Ok)
and
(b) 12hks(h, k) == h 2 + 1 (mod Ok).

PROOF. From (30) we find

6h k- 1 k- 1 [hrJ k- 1
(36) 6ks(h, k) = k r~lr2 - 6 r~/ k - 3 r~lr.
Since 6 L~: t r 2 = k(k - 1) (2k - 1) each term on the right of (36) is an
integer. Moreover, (36) shows that

6ks(h, k) == h(k - 1)(2k - 1) (mod 3)

so we have

(37) 12ks(h, k) == 2h(k - 1)(2k - 1) == h(k - 1)(k + 1) (mod 3).

64
3.9: Congruence properties of Dedekind sums

If 31 k then 3%h and (37) implies


12ks(h, k) == - h =1= 0 (mod 3).
If 3,rk then 31 (k - 1)(k + 1) and (37) implies
(38) 12ks(h, k) == 0 (mod 3).
In other words, 12ks(h, k) == 0 (mod 3) if, and only if, 3%k. Hence, inter-
changing hand k, we have
12hs(k, h) == 0 (mod 3) if, and only if, 3%h.
If 0 = 3 this implies (a) since (h, k) = 1. If 0 = 1, (a) holds trivially. Part (a),
together with the reciprocity law, gives (b) since k 2 - 3hk == 0 (mod Ok).
o
Note. Theorems 3.8(b) and 3.6(c) show that
s(h, k) = 0 if, and only if, h 2 + 1 == 0 (mod k).

Theorem 3.9. The Dedekind sums satisfy the congruence

(39) 12ks(h, k) == (k - 1)(k + 2) - 4h(k - 1) +4 I [ -k


2hrJ (mod 8).
r<k/2

If k is odd this becomes

(40) 12ks(h, k) == k - 1 + 4 I [2~rJ (mod 8).


r<k/2

PROOF. From (36) we obtain

12ks(h, k) = 2h(k - 1)(2k - 1) -


k-l [h
12r~/ :
J - 3k(k - 1)

= -2h(k - 1) + 4hk(k - 1) -
k-l [h
12r~/ :
J
+ k(k - 1) - 4k(k - 1).

Now we reduce the right member modulo 8. Since 4k(k - 1) == 0 (mod 8)


this gives us

12ks(h, k) == -2h(k - 1) -
k-l [h
4r~lr :
J+ k(k - 1) (mod 8)

I [hrJ
== (k - 1)(k - 2h) - 4 k-l - (mod 8)
r= 1 k
r odd

== (k - 1)(k - 2h) - 4kL:


- 1 [hrJ
k +4 L: [2hrJ
k (mod 8).
r= 1 r<k/2

65
3: The Dedekind eta function

The next to last term is equal to

-4 L [hrJ
k-1
- = 4L k-1 ((hr))
- - 4 L -hr + 2 L 1
k-1 k-1

r=1 k r=1 k r=1 k r=1

= 0 - 2h(k - 1) + 2(k - 1) = (k - 1)(2 - 2h).


Since
(k - 1)(k - 2h) + (k - 1)(2 - 2h) = (k - 1)(k + 2) - 4h(k - 1)
this proves (39).
When k is odd we have 4h(k - 1) == 0 (mod 8) and
(k - 1)(k + 2) = k 2
+k - 2 == k' - 1 (mod 8)
since k 2 == 1 (mod 8). Hence (39) implies (40). D

Theorem 3.10. If k = 2).k 1 where A > 0 and k 1 is odd, then for odd h z 1
we have
(41) 12hks(h, k) == h 2 + k 2 + 1 + 5k - 4k L [2~VJ (mod 2).+ 3).
v<h/2

PROOF. Since h is odd we can apply (40) to obtain, after multiplication by k,

12hks(k,h) == k(h - 1) + 4k L [ h2kV] (mod 2,\+3).


v <h/2

By the reciprocity law we have


12hks(h, k) = h 2 + k2 - 3hk + 1 - 12hks(k, h)

== h 2 + k 2 - 3hk + 1 - k(h - 1) - 4k L [2~V] (mod 2).+3)


v<h/2

== h 2 + k 2 + 1 + k - 4hk - 4k L [2kVJ (mod 2).+3).


v <h/2 h
Since h is odd we have 4k(h + 1) == 0 (mod 2'\ + 3) hence k - 4hk == 5k
(mod 2)' + 3) and we obtain (41). D
Finally, we obtain a property of Dedekind sums which plays a central
role in the study of the invariance of modular functions under transforma-
tions of certain subgroups of the modular group. This will be needed in
Chapter 4.
Theorem 3.11. Let q = 3, 5, 7 or 13 and let r = 24/(q - 1). Given integers
a, b, c, d with ad - bc = 1 such that c = c 1 q, where C 1 > 0, let

<5 = { s(a, c) a+
- ~ d} - {s(a, Cl) - + .
a12cl d}
Then rb is an even integer.

66
3.9: Congruence properties of Dedekind sums

PROOF. Taking k = c in Theorem 3.8(b) we find

12ac{S(a, c) - al;cd} == a 2 + 1 - a(a + d) == -bc (mod ec),

where 0 = (3, c). The same theorem with k = Cl = c/q gives, after multi-
plication by q,

12ac{S(a, Ct) - al~td} == qa 2


+ q - qa(a + d) == -qbc (mod etc),

where 0l = (3, c l ). Note that OliO so both congruences hold modulo 0lC.
Subtracting the congruences and multiplying by r we find
12acrb == r(q - 1)bc (mod OlC).
But r(q - 1)bc = 24bc == 0 (mod 0lC) so this gives
12acrb == 0 (mod OlC).
Now (a, c) = 1 since ad - bc = 1. Also, 12cb is an integer so we can cancel
a in the last congruence to get
(42) 12crb == 0 (mod OlC).
Next we show that we also have
(43) 12crb == 0 (mod 3c).
Assume first that q > 3. In this case 0 = (3, qc l) = (3, Cl) = 01 so (42)
becomes
12crb == 0 (mod Oc).
If 0 = 3 this gives (43). But if 0 = 1 then 3~c so 3~Cl and (38) implies
12cs(a, c) == 0 (mod 3) and 12cs(a, Cl) == 0 (mod 3). Hence
12crb == r(q - 1)(a + d) = 24(a + d) == 0 (mod 3),
which, together with (42), implies (43).
Now assume that q = 3 so r = 12. Then 0 = 3 and 0l is 1 or 3. IfO l = 3
we get (43) by the same argument used above, so it remains to treat the case
0 1 = 1. In this case 3~Cl so (38) implies 12c l s(a, c l ) == 0 (mod 3), hence
12cs(a, Cl) == 0 (mod 9).

Also,
12cb = 12cs(a, c) - (a + d) - 12cs(a, Cl) + 3(a + d)
== 12cs(a, c) + 2(a + d) (mod 9),
so
(44) 12racb = 12racs(a, c) + 2r(a 2 + ad) (mod 9).

67
3: The Dedekind eta function

But Theorem 3.8(b) gives us 12aes(a, e) == a 2 + 1 (mod 9) since 31 e. Hence


(44) becomes
12raeb == r(a 2 + 1) + 2ra 2 + 2rad (mod 9)
== 3ra 2 + r + 2r(1 + be) == 3r + 2rbe == 0 (mod 9)
since r = 12 and 9112e. This shows that
12raeb == 0 (mod 9).
Now 3~a since (a, e) = 1 so we can cancel a to obtain 12reb == 0 (mod 9)
which, with (42), implies (43).
Our next goal is to show that we also have
(45) 12erb == 0 (mod 24e)
since this implies rb is even and proves the theorem. To prove (45) we treat
separately the cases e odd and e even.
Case 1: e odd. Apply (40) with k = e to obtain

a + d} == e - 1
12e { s(a, e) - ~ + 4T(a, e) - (a + d) (mod 8)

where we have written

T(a, e) = L [2aVJ.
v<cj2 e
We only need the fact that T(a, e) is an integer. Applying (40) again with
k = e 1 = e/q and multiplying by q we have

12C{S(a, c l ) - al;C~} == C - q + 4qT(a, cd - q(a + d) (mod 8).

Subtracting the last two congruences and multiplying by r we find


12erb == r(q - 1) + r(q - 1)(a + d) == 0 (mod 8)
since r(q - 1) = 24 and 4r == 0 (mod 8). Combining this with (43) we obtain
(45) and the theorem is proved for odd e.
Case 2: e even. Write e = 2 Ay with y odd. Now a is odd since (a, e) = 1 so
if a ~ 1 we can apply Theorem 3.10 with k = e and h = a to obtain

2
12aC{S(a, c) - a l;Cd} == a + c2 + 1

+ 5e - 4eT(e, a) - a(a + d) (mod 2 A+ 3)


== e 2 + 5e - be - 4eT(e, a) (mod 2 A + 3 )
since ad - be = 1. Similarly,

12aC{S(a, c l ) - al~ld} = CC l + 5c - qbc - 4cT(c l , a) (mod 2.<+3).

68
3.LO: The Eisenstein series G2 (r)

Subtract, multiply by r and use the congruence 4cr == 0 (mod 2). + 3) to obtain
12carc5 == rcc1(q - 1) + r(q - l)bc == 0 (mod 2 A + 3 ).
Since a is odd we can cancel a to obtain
(46) 12crc5 == 0 (mod 2 A+ 3).
Now (43) states that 12crc5 == 0 (mod 3 ·2 Ay) which, together with (46)
implies (45) and proves the theorem for a ~ 1.
To prove it for a < 0, write c5 = c5(a) to indicate the dependence on a.
If a' = a + tc, where t is an integer, an easy calculation shows that
c5(a') - c5(a) = t(q - 1)/12 since s(a, c) = s(a', c) and s(a', c 1 ) = s(a, c 1 ). There-
fore rc5(a') - rc5(a) = 2t, an even integer. Choosing t so that a' ~ 1 we know
rc5(a') is even by the above argument, so rc5(a) is also even. This completes
the proof. D

3.10 The Eisenstein series G 2 ( r)


If k is an integer, k ~ 2, and if r E H the Eisenstein series
1
(47) G 2k (r) = L 2k
(m,n)*(O,O) (m + nr)
converges absolutely and has the Fourier expansion
2(2ni)2k 00 .
(48) G 2k(r) = 2((2k) + (2k _ I)! n~1(T2k-l(n)e2"'nt

where, as usual, (Jex(n) = Ldln dex . The cases k = 2 and k = 3 were worked out
in detail in Chapter 1, and the same argument proves (48) for any k ~ 2. If
k = 1 the series in (47) no longer converges absolutely. However, the series
in (48) does converge absolutely and can be used to define the function G 2 (r).

Definition. If r E H we define
00

(49) G 2 (r) = 2'(2) + 2(2ni)2 L (J(n)e21rinr.


n=l

If x = e 21rir the series on the right of (49) is an absolutely convergent


power series for Ix I < 1 so G 2 (r) is analytic in H. This definition also shows
that G 2 (r + 1) = G 2 (r).
Exercises 1 through 5 describe the behavior of G 2 under the other generator
of the modular group. They show that

(50) G2( ~ 1) = 2
r G 2 (r) - 2nir,

a relation which leads to another proof of the functional equation YJ( - l/r) =
(- ir)1/2YJ(r).

69
3: The Dedekind eta function

Exercises for Chapter 3


1. If r E H prove that
x x 1
(51) G 2 (r) = 2((2) + L L
n= - x m=-x (m + nr) 2'
n*O

Hint: Start with Equation (12) of Chapter 1, replace r by nr, where n > 0, and sum
over all n 2 1.

2. Use the series in (51) to show that

1) 1
(-=-r
oc x
(52) r- G 2
2
= 2((2) + L L 2'
m=-cx n=-x (m + nr)
n*O

the iterated series in (52) being the same as that in (51) except with the order of sum-
mation reversed. Therefore, proving (50) is equivalent to showing that

(53) L:
m= - x
L:
n= - x (m
1
+ nr)2
= L:
n= - x m= - x
I (m
1
+ nr)2
_ 2ni
r
n*O n*O

3. (a) In the gamma function integral r(z) = .r~ e-tt Z - 1 dt make the change of variable
t = ~1I. where ~ > O. to obtain the formula

(54) Il(-zf'(z) = L" e-·uu z


-
I
du,

and extend it by analytic continuation to complex ~ with Re(~) > O.


(b) Take z = 2 and ~ = - 2ni(m + nr) in (54) and sum over all 11 2 1 to obtain the
relation

L
OC 1
2 = - 8n
2 IX cos(2nmu)gr(U) du,
n=-x (nr + m) 0
n*O

where
x
giu) = LI L e2rrinru if u > 0
n=l

and

4. (a) Use Exercise 3 to deduce that

where
X

f(t) = L gr(t + k).


k=O

70
Exercises for Chapter 3

(b) The series on the right of (55) is a Fourier series which converges to the value
±{f(O +) + f(1- )}. Show that
-1 ex
. + Igr(k)
f(O+) = - 2
nlr k= 1
and that
ex ex
f(1- ) = I gr(k) = I a(n)e21tinr,
k= 1 n= 1

and then use (55) to obtain (50).

5. (a) Use the product defining 1](r) to show that

d
-4ni - log 1](r) = G 2(r).
dr
(b) Show that (50) implies

-d log 1] -
dr
(-1)
r
=
d 1](r)
-log
dr
Id + - -log( -ir).
2 dr
Integration of this equation gives 1]( -l/r) = C( - ir)1/21](r) for some constant C.
Taking r = i we find C = 1.
6. Derive the reciprocity law for the Dedekind sums s(h, k) from the transformation
formula for log 1](r) as given in Equation (12).

Exercises 7 and 8 describe properties of the function


r(s)
<I>((J(, 13, s) = (2n)' {((s, (J()F(f3, 1 + s) + ((s, 1 - (J()F(l - 13,1 + s)}
which occurs in the proof of Iseki's formula (Theorem 3.5). The properties
follow from Hurwitz's formula (Theorem 12.6 of [4]) which states that
r(s) . /2
+ e 1tlS
. /2
W - s, a) = (2n)' {e~ltIS F(a, s) F( -a, s)}.

7. (a) If 0 < a < 1 and Re (s) > 1, prove that Hurwitz's formula implies

- s) 1t1'( 1 )/2
F(a s)
,
= r(1
(2n)1 -
{eS
-s ((1 - - ,
s a) + e1tI.(S-l)/ 2s
/(1 - s 1 - a)}

(b) Use (a) to show that <J>(a, Ii, s) can be expressed in terms of Hurwitz zeta functions
by the formula

<I>(a, f3,~ = e 1tis/2{((s, a)(( -s, 1 - {3) + ((s, 1 - a)(( -s, fJ))
r(s)r( -s)
+ e- 1tis/2{(( -s, I - f3K(s, 1 - a) + (( -5, {3)((s, a)}
and deduce that <I>((X, {3, s) = <1>( 1 - {3, (X, - s).
8. This exercise gives an estimate for the modulus of the function z - s<I>((X, {3, s) which
occurs in the integral representation of A((X, fi, s) in the proof of Iseki's formula
(Theorem 3.5).

71
3: The Dedekind eta function

(a) Show that the formula of Exercise 7(b) implies


-nz- s .
z-s<I>(a, [3, s) = - .- {e- 1tls/2[((s, a)(( -s, [3) + ((s, 1 - a)(( -S, 1 - [3)J
s SIn ns
+ e1tis/2[((s, a)(( - s, 1 - [3) + ((s, 1 - a)(( - s, [3)J}.
(b) For fixed z with Iarg zI < n12, choose J > 0 so that Iarg zI ~ nl2 - J, and show
that if s = (J + it where (J ~ -~ we have
Iz-si = O(e 1tl (1t/2-b»),

where the constant implied by the O-symbol depends on z.


(c) If s = (J + it where (J ~ -~ and It I ~ 1, show that
1t1tl
1 (e- )
Is sin nsl = 0 -I-t-I '
and that

Ie1tiS/2 1 = O(e1tltl/2), Ie- 1tis /2 1 = O(e1tltl/2).

(d) If (J ~ -~ and It I ~ 1 obtain the estimate I((s, a) I = O( ItiC) for some c > 0
(see [4J, Theorem 12.23) and use (b) to deduce that
Iz-s<I>(a, [3, s)1 = O(/tI2C-le-/tlt)).
This shows that the integral of z-s<I>(~, [3, s) along the horizontal segments of the
rectangle in Figure 3.2 tends to 0 as T ~ 00.

PROPERTIES OF DEDEKIND SUMS

9. If k ~ 1 the equation

s(h, k) = rm~ k ((i))((h:))


is meaningful even if h is not relatively prime to k and is sometimes taken as the
definition of Dedekind sums. Using this as the definition of s(h, k) prove that
s(qh, qk) = s(h, k) if q > O.
10. If p is prime prove that
p- 1

(p + l)s(h, k) = s(ph, k) + L s(h + mk, pk).


m=O

11. For integers r, h, k with k ~ 1 prove that we have the finite Fourier expansion
hr)) I 2nhrv nv
(( -k = - -
k- 1
L .
SIn--cot-
2k \'= 1 k k

and derive the following expression for Dedekind sums:

1 k- 1 nhr nr
s(h, k) = - L cot - cot-.
4k r= 1 k k
12. This exercise relates Dedekind sums with the sequence {lI(Il)] of Fibonacci numbers
1, 1,2,3,5,8, ... , in which u(l) = u(2) = 1 and u(n + 1) = u(n) + u(n - 1).
(a) If h = u(2n) and k = u(2n + 1) prove that s(h, k) = o.
(b) If h = u(2n - 1) and k = u(2n) prove that 12hks(h, k) = h 2 + k 2 - 3hk + 1.

72
Exercises for Chapter 3

FORMULAS FOR EVALUATING DEDEKIND SUMS

The following exercises give a number of formulas for evaluating Dedekind


sums in closed form in special cases. Assume throughout that (h, k) = 1,
k ~ 1, h ~ 1.
13. If k == r (mod h) prove that the reciprocity law implies
12hks(h, k) = k2 - {12s(r, h) + 3}hk + h2 + 1.

Use the result of Exercise 13 to deduce the following formulas:


14. If k == 1 (mod h) then 12hks(h, k) = (k - l)(k - h2 - 1).

15. If k == 2 (mod h) then 12hks(h, k) = (k - 2)(k - !(h + 1)). 2

16. If k == -1 (mod h) then 12hks(h, k) = k 2 + (h 2 - 6h + 2)k + h2 + 1.

17. If k == r (mod h) and if h == t (mod r) where r ~ 1 and t = ± 1, then


h2 - t(r - l)(r - 2)h + r 2 + 1
12hks(h,k) = k2 - k + h2 + 1.
r
This formula includes those of Exercises 14 and 15 as special cases.
18. Show that the formula of Exercise 17 determines s(h, k) completely when r = 3
and when r = 4.
19. If k == 5 (mod h) and if h == t (mod 5), where t = ± 1 or ± 2, then
2
h k) k2 h + 4t(t - 2)(t + 2)h + 26 2
12hk s(, = - 5 k + h + 1.

20. Assume 0 < h < k and let ro, r 1, ... , r n+ 1 denote the sequence of remainders in the
Euclidean algorithm for calculating the gcd (h, k), so that

Prove that

1 n~1 {
2
"+ 1 r/ + rj_1 + I} (-1)" + 1
s(h,k)=- ~ (-I)J - .
12 j =1 rjrj-1 8
This also expresses s(h, k) as a finite sum, but with fewer terms than the sum in the
original definition.

73
Congruences for the coefficients
of the modular function j

4.1 Introduction
The function j(r) = 12 3 J(r) has a Fourier expansion of the form
1 00 •

j(r) = - + L c(n)x n , (x = e 2m !)
x n=O
where the coefficients c(n) are integers. At the end of Chapter 1 we mentioned
a number of congruences involving these integers. This chapter shows how
some of these congruences are obtained. Specifically we will prove that
c(2n) == 0 (mod 2 11 ),
c(3n) == 0 (mod 3 5 ),
c(5n) == 0 (mod 52),
c(7n) == 0 (mod 7).
The method used to obtain these congruences can be illustrated for the
modulus 52. We consider the function
00

f5(r) = L c(5n)x n

n=1

obtained by extracting every fifth coefficient in the Fourier expansion of j.


Then we show that there is an identity of the form
(1)
where the ai are integers and <I>(r) has a power series expansion in x = e 21ti !
with integer coefficients. By equating coefficients in (1) we see that each
coefficient of fs(r) is divisible by 25.
Success in this method depends on showing that such identities exist.
How are they obtained?

74
4.2: The subgroup r o(q)

Theorem 2.8 tells us that every modular function f is a rational function


of j. Sometimes this rational function is a polynomial in j with integer co-
efficients, giving us an identity of the form
f(r) = a1j(r) + a2j2(r) + ... + akjk(r).
However, the function fs(r) is not invariant under all transformations of
the modular group r and cannot be so expressed in terms of j(r). But we
shall find that fs(r) is invariant under the transformations of a certain
subgroup of r, and the general theory enables us to express fs(r) as a poly-
nomial in another basic function <l>(r) which plays the same role as j(r)
relative to this subgroup. This representation leads to an identity such as
(1) and hence to the desired congruence property.

The subgroup in question is the set of all unimodular matrices (: :)


with c == 0 (mod 5). More generally we shall consider those matrices in r
with c == 0 (mod q), where q is a prime or a power of a prime.

4.2 The subgroup r o(q)


Definition. If q is any positive integer we define r o(q) to be the set of all

matrices (: ~) in r with c == 0 (mod q).

It is easy to verify that r o(q) is a subgroup of r. The next theorem gives a


way of representing each element of r in terms of elements of r o(p) when
p is prime. In the language of group theory it shows that r o(p) is of finite
index in r.

Theorem 4.1. Let Sr = -l/r and Tr = r + 1 be the generators of the full


modular group r, and let p be any prime. Then for every V in r, V ~ r o(P),
there exists an element P in r o(P) and an integer k, 0 ~ k < p, such that
V = PST k •

PROOF. Given V = (~ ~) where C =1= 0 (mod p). We wish to find

p = G:). with c == 0 (mod p),

and an integer k, 0 ~ k < p, such that

75
4: Congruences for the coefficients of the modular function j

All matrices here are nonsingular so we can solve for ( : ~) to get

(: ~)=(~ ~)(~ -~rl =(~ ~)(_~ ~)=(~~=~ ~}


Choos~ k to be that solution of the congruence
kC == D (mod p) with 0 ~ k < p.
This is possible since C =1= 0 (mod p). Now take
c = kC - D, a = kA - B, b = A, d = C.
Then c == 0 (mod p) so PEro(p). This completes the proof. o
4.3 Fundamental region of r o(p)
As usual we write Sr = -l/r and Tr = r + 1, and let R r denote the funda-
mental region of r.

Theorem 4.2. For any prime p the set


p-l
Rr u U STk(R r )
k=O

is a fundamental region of the subgroup r o(p).


This theorem is illustrated for p = 3 in Figure 4.1.

PROOF. Let R denote the set


p-l
R = Rr u U STk(R r ).
k=O

We will prove

(i) if r E H, there is a V in r o(p) such that Vr belongs to the closure of R, and


(ii) no two distinct points of R are equivalent under r o(p).

To prove (i), choose r in H, choose r 1 in the closure of R r and choose


A in r such that Ar = r 1 . Then by Theorem 4.1 we can write

A-I = PW

where PEro(p) and W = I or W = ST k for some k, 0 ~ k ~ p - 1. Then


P = A- I W- 1 and p- I = WA. Let V = P- 1 • Then VEro(p) and

Vr = WAr = Wr i .

Since W = I or W = STk'l this proves (i).

76
4.3: Fundamental region of r o(p)

-1 o 1
2"

Figure 4.1 Fundamental region for r 0(3)

Next we prove (ii). Suppose!1 E R, ! 2 E R and v! 1 = ! 2 for some V in


r o(P). We will prove that! 1 = ! 2' There are three cases to consider:

(a) !1 ER r '!2 ERr· In this case!1 =!2 since VEr.


(b) ! 1 ERr, ! 2 E STk(R r ).
(c) ! 1 E STkl(R r ), ! 2 E ST k2(R r ).

In case (b), !2 = ST k !3 where !3 ERr. The equation

implies

V = ST k = 1 (0 -1)
k .

This contradicts the fact that V E r o(p).


Finally, consider case (c). In this case
and
where !1' and !2' are in R r . Since V!1 = !2 we have VST kl!I' = ST k2!2' so
VSTkl = ST k2,

77
4: Congruences for the coefficients of the modular function i

Since VE ro(p) this requires k 2 == k i (mod pl. But both k l , k 2 are in the
interval [0, p - IJ, so k 2 = k i . Therefore
V = SToS = S2 = I

and r I = r 2. This completes the proof. D


We mention (without proof) the following theorem of Rademacher [34]
concerning the generators of r o(p). (This theorem is not needed in the
later work.)

Theorem 4.3. For any prime p> 3 the subgroup ro(p) has 2[p112J +3
generators and they may be selected from the following elements:

where Tr = r + 1, Sr = -l/r, and


k -k' ( k'
v" = ST ST S = -(kk' + 1)

where kk' == -1 (mod pl. The subgroup r o(2) has generators T and VI~
the subgroup r 0(3) has generators T and V2 .

Here is a short table of generators:

p 2 3 5 7 11 13 17 19

Generators: T T T T T T T T
VI V2 V2 V3 V4 V4 V4 Vs
V3 Vs V6 Vs V7 V8
V8 Vg VI2
VIO VI3 VI3

4.4 Functions automorphic urlder the


subgroup r o(p)
We recall that a modular function f is one which has the following three
properties:

(a) f is meromorphic in the upper half-plane H.


(b) f(Ar) = f(r) for every transformation A in the modular group r.
(c) The Fourier expansion of f has the form
00

f(r) = L ane21tint.
n= -m

78
4.4: Functions automorphic under the subgroup r o(p)

If property (b) is replaced by


(b') f(Vr) = f(r) for every transformation V in r o(P),
thenfis said to be automorphic under the subgroup ro(p). We also say that
f belongs to r o(P)·
The next theorem shows that the only bounded functions belonging to
r o(p) are constants.
Theorem 4.4. If f is automorphic under r o(p) and bounded in H, then f is
constant.
According to Theorem 4.1, for every V in r there exists an element
PROOF.
P in r o(P) and an integer k, °
~ k ~ p, such that

V = PA k ,
where A k = ST k if k < p, and A p = I. For each k = 0, 1, ... , p, let
r k = {PA k : PEro(p)}.
Each set r k is called a right coset of r o(P). Choose an element ~ from the
coset r k and define a function fk on H by the equation
fk(r) = f(~ r).

Note that fp(r) = f(Pr) = f(r) since P E ro(P) and f is automorphic under
ro(P). The function value A(r) does not depend on which element ~ was
chosen from the coset r k because
A(r) = f(~r) = f(PAkr) = f(Akr)
and the element A k is the same for all members of the coset r k •
How does fk behave under the transformations of the full modular group?
If V E r then
fk(Vr) = f(~ Vr).
Now ~ V
such that
E r so there is an element Q in r o(P) and an integer m, ° ~ m ~ p,

Therefore we have

Moreover, as k runs through the integers 0, 1, 2, ... , p so does m. In other


words, there is a permutation (J of {O, 1,2, ... , p} such that
fk(Vr) = fa(k)(r) for each k = 0,1, ... ,p.
Now choose a fixed w in H and let

n {fk(r) -
p
cp(r) = f(w)}.
k=O

79
4: Congruences for the coefficients of the modular function j

Then if V E r we have
p p
cp(Vr) = n {fk(Vr) -
k=O
f(w)} = n {f(j(k)(r) -
k=O
f(w)} = cp(r),

so cp is automorphic under the full group r. Now cp is bounded in H (since


each fk is). Therefore~ cp omits some value hence, by Theorem 2.5, cp is
constant, so cp(r) = cp(w) for all r. But cp(w) = 0 because

n {fk(W) -
p
cp(w) = f(w)}
k=O
and the factor with k = p vanishes since fp = f Therefore cp(r) = 0 for all r.
Now take r = i. Then
p

o= n {A(i) -
k=O
f(w)}

hence some factor is o. In other words, f(w) = A(i) for some k. But w was
arbitrary so f can take only the values fo(i), ... ,fp(i). This implies that f is
constant. 0

4.5 Construction of functions belonging


to ro(p)
This section shows how to construct functions automorphic under the
subgroup r o(p) from given functions automorphic under r.

Theorem 4.5. If f is automorphic under r and if p is prime, let

fir) =! Pff(r + A).


p ),=0 P
Thenfp is automorphic under r o(p). Moreover, iff has the Fourier expansion
00

f(r) = L a(n)e21tinr
n= -m
then fp has the Fourier expansion
00

fp(r) = L a(np)e21tinr.
n= - [m/p]
PROOF. First we prove the statement concerning Fourier expansions. We have
1 p- 1 00

fp(r) = - L L a(n)e 21tin (r+).)/P


p ),=0 n=-m

1 p- 1
L L e21tin )./p.
00

=- a(n)e21tinr/p
p n= -m ),=0

80
4.5: Construction of functions belonging to r o(p)

But

L e21tin ;"/p = {O
p-I if p,(n
;"=0 P if pin
so
00 00

fp('r) = L a(n)e21tinr/p = L a(np)e21tinr.


n= -m n= -[m/p]
pin

This shows that fp has the proper behavior at the point r = ioo. Also,
fp is clearly meromorphic in H because it is a linear combination of functions
meromorphic in H.
Next we must show that
fp(Vr) = fp(r) whenever V E r o(p).
For this we use a lemma.
Lemma 1. If V E ro(p) and if
there exists an integer j1, 0
° ~
~ j1 ~
A ~ p - 1, let T;.,r = (r + A)jp. Then
p - 1 and a transformation ~ in
r O(p2) such that

Moreover, as A runs through a complete residue system modulo p, so does j1.

First we use the lemma to complete the proof of Theorem 4.5, then we
return to the proof of the lemma.
If V E r o(p) we have
P 1 P 1
1 - (Vr+A)
fp(Vr) = - L f = -1 L-
f(T;., Vr).
P ;"=0 P, P ;"=0

Now we use the lemma to write the last sum as


1 p- I 1 p- I
- L f (~ ~ r) = - L f (~ r) = fp( r).
PJL=o PJL=o
This proves that fp is invariant under all transformations in r o(p), so fp is
automorphic under r o(p). D

PROOF OF LEMMA 1. Let V = (: :). where c == 0 (mod p), and let A. be


given, 0 ~ A ~ p - 1. We are to find an integer j1, 0 ~ j1 ~ p - 1 and a
transformation ~ = (~ ~) such that ~ E r O(P2) and
T;.,V = WJLTJL.

81
4: Congruences for the coefficients of the modular function j

Since TA. = G~) we must satisfy the matrix equation

or

a + AC b + Ad) = (A AJl + BP)


( pc pd C CJl + Dp
with C == 0 (mod p2). Equating entries we must satisfy the relations
A = a + AC
(2) { C = pc

(3) {AJl + Bp = b + Ad
CJl + Dp = pd
with
C == 0 (mod p2) and AD - BC = 1.
Now (2) determines A and C. Since pic, we have C == 0 (mod p2). Substi-
tuting these values in (3) we must satisfy

(4) {(a + AC)Jl + Bp = b + Ad


CpJl + Dp = pd.
Choose Jl to be that solution of the congruence
Jla == b + Ad (mod p)
which lies in the interval 0 ~ Jl ~ P - 1. This is possible because ad - bc = 1
and pic imply p,ta. Note that distinct values of A mod p give rise to distinct
values of Jl mod p. Then, since pic we have
Jla + JlAc == b + Ad (mod p)
or
(a + AC)Jl == b + Ad (mod p).
Therefore there is an integer B such that
(a + AC)Jl + Bp = b + Ad.
Therefore the first relation in (4) is satisfied. The second relation requires
D = d - CJl. Thus, we have found integers Jl, A, B, C, D such that

Clearly AD - BC = 1 since all matrices in this equation have determinant


1 or p. This completes the proof of the lemma. D

82
4.6: The behavior of.f~ under the generators of r

4.6 The behavior of f p under the generators


ofr
Let Tr = r + 1 and Sr = -l/r be the generators of r. Since T E r o(p) we
have fp(Tr) = fp(r). The next theorem gives a companion result for fp(Sr).

Theorem 4.6. If f is automorphic under r and if p is prime, then


~ (- ~)
p r = f, (r) + ~p f(pr) - ~p f(~).
p P

To prove this we need another lemma.

Lemma 2. Let T). r = (r + A)/p. Thenfor each A in the interval 1 S ASP - 1


there exists an integer J.1 in the same interval and a transformation V in
r o(p) such that
T).S = VTl1 •
Moreover, as A runs through the numbers 1,2, ... ,p - 1, so does J.1.

PROOF OF LEMMA 2. We wish to find (: ~) in ro(p) such that

G;)(~ -~) = G~)(~ ~)


or

G-~) = (: :::~:)-
Take a = A, c = p and let J.1 be that solution of the congruence
,1,J.1 == -1 (mod p)
in the interval 1 S J.1 S P - 1. This solution is unique and J.1 runs through a
reduced residue system mod p with ,1,. Choose b to be that integer such that
aJ.1 + bp = - 1, and take d = - J.1. Then CJ.1 + dp = 0 and the proof is
complete. D
PROOF OF THEOREM 4.6. We have

Pfp( - ~) = Pi! f(sr + A) = f(sr) + Pf f(T;.Sr)


r ),=0 p P).= 1

= f(- ~) + Pi! f(VTI1 r) = f(rp) + Pi! f(TIL r) - f(~)


rp 11=1 11=0 P

= f(rp) + pfp(r) - fG)- 0

83
4: Congruences for the coefficients of the modular function.i

4.7 The function <p(r) = ~(qr)/~(r)

The number of poles of an automorphic function in the closure of its fun-


damental region is called its valence. A function is called univalent on a
subgroup G if it is automorphic under G and has valence 1. Such a function
plays the same role in G that J plays in the full group r.
It can be shown (using Riemann surfaces) that univalent functions exist
on G if and only if the genus of the fundamental region R G is zero. [This is
the topological genus of the surface obtained by identifying congruent edges
of R G • For example, the genus of R r is zero because R r is topologically
equivalent to a sphere when its congruent edges are identified.]
Our next goal is to construct a univalent function on the subgroup
r o(p) whenever the genus of r o(p) is zero. This will be done with the aid of
the discriminant ~ = g 2 3 - 27 g 3 2 •
We recall that ~(r) is periodic with period 1 and has the Fourier expansion
(Theorem 1.19)
00

~(r) = (2n)12 L r(n)e21tint


n=1

where the r(n) are integers with r(1) = 1 and r(2) = -24. However, ~(r) is
not invariant under all transformations of r. In fact we have

:
~G: ~) = (cr + d)12~(r) if (: ~) E r.
In particular,

~(r + 1) = ~(r) and ~( ~ 1) = r12~(r).


Even though ~(r) is not invariant under r it can be used to construct functions
automorphic under the subgroup r o(q) for each integer q.

Theorem 4.7. For afixed integer q, let

q>(r) = ~(qr) ifr E H.


~(r)

Then qJ is automorphic under ro(q). Moreover, the Fourier expansion ofqJ


has theform

where the bn are integers and x = e 21tit•

84
4.7: The function <p(r) = ~(qr)/~(r)

PROOF. First we obtain the Fourier expansion. We have

~(r) = (2n)12 n~1 r(n)x n = (2n)12x { 1 + n~1 r(n + 1)xn}.


where x = e 21tit • Hence

so

q>(r) = ~(qr) = x q- 1 1+ L:'~ 1r(n + 1)xnq = x q- 1(1 + f b x n)


~( r) 1 + L~= 1 r( n + 1)x n n= 1 n

where the bn are integers.


Now qJ is clearly meromorphic in H, and we will prove next that qJ is
invariant under r o(q).
If V = (: ~) E ro(q) then c = c q for some integer c
1 1• Hence

~( V r) = (cr + d) 12 ~( r) = (c 1 qr + d) 1 2 ~( r).
On the other hand,
aT + b a(qr) + bq
qVr = q - - d = ( ) d = W(qr),
cr + C 1 qr +
where
w= (aC1
q
b ).
d
But WE r because det W = ad - bc 1q = ad - bc = 1. Hence
~(qVr) = ~(W(qr)) = (c 1(qr) + d)12~(qr),
so

(Vr) = ~(qVr) = (c1qr + d)12~(qr) = (r).


qJ ~( Vr) (c 1 qr + d) 1 2 ~( r ) qJ

This completes the proof. D

Now qJ has a zero of order q - 1 at 00 and no further zeros in H. Next we


show that qJ does not vanish at the vertex r = 0 of the fundamental region
of r o(q). In fact, we show that qJ(r) -+ 00 as r -+ O.

Theorem 4.8. Ifr E H we have

qJ (-1) = 1
~ q12 qJ(r)·

Hence qJ(r) -+ 00 as r -+ O.

85
4: Congruences for the coefficients of the modular function j

PROOF. Since ~(-l/r) = r12~(r) we have

~( -:r) = (qr)12~(qr)
so

4.8 The univalent function <1>(r)


The function cp has a zero of order q - 1 at 00 and no further zeros so its
valence is q - 1. We seek a univalent function automorphic under r o(q)
and this suggests that we consider cpa, where r:J. = 1/(q - 1). The Fourier
expansion of cpa need not have integer coefficients, since

q>~(r) = x( 1 + n~l bnxnr


On the other hand we have the product representation

~(r) = (2n)12 x n00

n=1
(1 - Xn )24

so

where the coefficients dq(n) are integers. Therefore if a = 1/(q - 1) we have


00
n)24a
(5) q>~(r) = x ( 1 + n~l din)x

and the Fourier series for cpa(r) will certainly have integer coefficients if
24a is an integer, that is, if q - 1 divides 24. This occurs when q = 2, 3, 4, 5,
7, 9, 13, and 25.

Definition. If q - 1 divides 24 let a = 1/(q - 1) and r = 24a. We define the


function <D by the relations

a (~(qr))a (l1(qr))r
<I>(r) = q> (r) = ~(r) = 1](r) ·

86
4.9: Invariance of <1>( r) under transformations of r o(q)

The function <I> so defined is analytic and nonzero in H. The Fourier series
for <I> in (5) shows that <I> has a first order zero at 00 and that
1 1
<I>(r) = ~ + I(x),

where I(x) is a power series in x with integer coefficients.


Since qJ is automorphic under r o(q) we have qJ(Vr) = qJ(r) for every
element V of r o(q). Hence, extracting roots of order q - 1, we have
<1>( V r) = E<I>( r )
q
where E - 1 = 1. The next theorem shows that, in fact, E = 1 whenever
24/(q - 1) is an even integer and q is prime. This occurs when q = 2, 3, 5, 7,
and 13. For these values of q the function <I> is automorphic under r o(q).

4.9 Invariance of <l>(r) under transformations


of r o(q)
The properties of Dedekind sums proved in the foregoing chapter lead to a
simple proof of the invariance of the univalent function <I>(r).

Theorem 4.9. Let q = 2,3,5,7, or 13, and let r = 24/(q - 1). Then thefunction

(6) <1>( r) = (11( qr ))r


11(r)

is automorphic under the subgroup r o(q).


PROOF.Ifq = 2wehaver = 24 and <I>(r) = ~(qr)/~(r).Inthiscasethetheorem
was already proved in Theorem 4.7. Therefore we shall assume that q ~ 3.
Let V = (: ~) be any element of r o(q). Then ad - be = 1 and
c == 0 (mod q). We can suppose that c ~ O. If c = 0 then V is a power of
the translation Tr = r + 1, and since 11(r + 1) = e1ti/ 12 11(r) we find

<I>(r + 1) = (11(qr + q))r = e"ir(Q- 1l/12<1>(r) = <l>(r).


11(r + 1)
Therefore we can assume that c > 0 and that c = c1q, where C1 > O.
Dedekind's functional equation for 11(r) gives us

(7)
where

(8)

87
4: Congruences for the coefficients of the modular function.i

We also have

a(qr) + bq )
I](qVr) = I] ( c1(qr) + d = I](V1qr)

where

~I = (a CI
q
b )

Since VI E r we have

which, together with (7), gives us

<D(Vr) = (:~~;)'<D(r).
But (8) shows that (r.(V1 )/r.(V)t = e- rrirb~ where

(j,{a+d
= ~ + s(-d,c)} - {a+d
12c + s(-d,c 1 ) } .
1

Since ad - bc = 1 we have ad == 1 (mod c) and ad == 1 (mod c l ) so


s( -d, c) = -s(a, c) and s( -d, c l ) = -s(a, CI), and Theorem 3.11 shows that
rb is an even integer. Therefore e~1tirb = 1 and <I>(Vr) = <I>(r). D

4.10 The function j p expressed as a


polynomial in <l>

If p is prime and if f is automorphic under r, we have shown that the


function

fp(r) = ~ Pf f(r + A)
p ),=0 P
is automorphic under r o(p), and its Fourier coefficients consist of every pth
coefficient of f To obtain divisibility properties of the coefficients of jp(r)
we shall express j p as a polynomial in the function <1>.
In deriving the differential equation for the Weierstrass g;J function we
formed a linear combination of g;J, g;J2 and g;J3 which gave a principal part
near z = 0 equal to that of [g;J'(z)] 2. The procedure here is analogous.
Both functions jp and <I> have a pole at the vertex r = 0 of the fundamental
region of ro(p). We form a linear combination of powers of <I> to obtain a
principal part equal to that of jp.

88
4.10: The function j p expressed as a polynomial in <1>

To obtain the order of the pole of jp(r) at r = 0 we use Theorem 4.6


which gives us the relation

j p( - r ~) = j p(r) + ~p j(pr) - ~p j(~)


p

valid for prime p. Replacing r by pr in this formula we obtain

Theorem 4.10. If p is prime and r E H then

j (-
p ~)
pr = j p(pr) + ~p j(p2r) - ~p j(r).

Hence if x = e 21tit we have the Fourier expansion

pj (-
p
~)
pr
= x- p2 - X-I + I(x),
where I(x) is a power series in x with integer coefficients.
PROOF. We have
j(r) = X-I + c(O) + c(1)x + c(2)x 2 + ... ,
jp(r) = c(O) + c(p)x + c(2p)x 2 + ... ,
pjp(pr) = pc(O) + pc(p)x P + pc(2p)x 2p + ... ,
and

so

pjp( - ;r) = pjp(pr) + j(p2 r ) - j(r)

= x- p2 - X-I + I(x). o
Now we can express j p as a polynomial in <1>.

Theorem 4.11. Assume p = 2,3,5,7 or 13, and let


l1(Pr ))r 24
<I>(r) = ( '1(r) , where r = p _ 1.

Then there exist integers aI' ... , a p2 such that

PROOF. By Theorem 4.10 we have

pjp( - ;r) = x-
p2
- x- 1
+ I(x),

89
4: Congruences for the coefficients of the modular function j

and, since 12a = r/2, Theorem 4.8 gives us

pr/2cI>(_ ~) = _1_ = X-I + I(x).


pr cI>( r)
Let ljJ(r) = pr/2<D( -l/(pr)). Then the difference

pj (_
p
~)
pr
- {1jJ(r)}p
2

has a pole of order :s p2 - 1 at x = 0, and the Laurent expansion near x =0


has integer coefficients. Hence there is an integer b 1 such that

has a pole of order :s p2 - 2 at x = 0, and the Laurent expansion near x =0


has integer coefficients. In p2 steps we arrive at a function

f ( - -1 ) = Pl. ( - -1 ) - {1jJ(r)}P 2 -b 1 {IjJ(r)}P 2-1 - ... - b 2_ ljJ(r)


pr P pr P 1

which is analytic at x = 0 and has a power series expansion with integer


coefficients. Moreover, all the numbers b 1, ... , b p 2_1 are integers. Replacing
r by -1/(pr) we obtain
f(r) = pjp(r) - {pr/2cI>(r)}P 2 - b 1{pr/2cI>(r)}p2- 1 - ... - b p2_ 1{pr/2cI>(r)}.
Now f(r) is automorphic under r o(p) and analytic at each point r in H.
The function f is also analytic at the vertex r = 0 (by construction). Therefore
f is bounded in H so f is constant. But this constant is pc(O) since cI>( r)
vanishes at 00. Thus we find
pjp(r) = {pr/2cI>(r)}p2 + b 1{pr/2cI>(r)}p2- 1 + ... + b p2_1 {pr/2cI>(r)} + pc(O)
so j p( r) is expressible as indicated in (9). D
Theorem 4.12. The coefficients in the Fourier expansion of j(r) satisfy the
following congruences:
c(2n) == 0 (mod 2 11 )
c(3n) == 0 (mod 35)
c(5n) == 0 (mod 52)
c(7n) == 0 (mod 7).

PROOF. The previous theorem shows that for p = 2, 3, 5, 7 and 13 we have


c(pn) == 0 (mod p(r/2)-I),
where r = 24/(p - 1). Therefore we simply compute (r/2) - 1 to obtain the
stated congruences. Note that (r/2) - 1 = 0 when p = 13 so we get a trivial
congruence in this case. D

90
Exercises for Chapter 4

Note. By repeated application of the foregoing ideas Lehner [24] derived


the following more general congruences, valid for lJ. ~ 1:
c(2 Cl n) == 0 (mod 2 3Cl +8)
c(3 Cl n) == 0 (mod 32Cl + 3 )
C(5Cl n ) == 0 (mod 5Cl+I)

c(7 Cl n) == 0 (mod 7Cl ).

Since it is known that c(13) is not divisible by 13, congruences of the above
type cannot exist for 13. In 1958 Morris Newman [30] found congruences
of a different kind for 13. He showed that

c(13np) + c(13n)c(13p) + p-1c( p


13n) == 0 (mod 13),

where p-Ip == 1 (mod 13) and c(x) = 0 if x is not an integer. The congruences
of Lehner and Newman were generalized by Atkin and O'Brien [5J in 1967.

Exercises for Chapter 4


1. This exercise relates the Dedekind function 1J(r) to the Jacobi theta function 9(r)
defined on H by the eq uation
ex:: ex'

t9(r) = 1 + 2 L e1tin2t = L e1tin2t.


n=l

The definition shows that 9 is analytic in H and periodic with period 2.


Jacobi's triple product identity (Theorem 14.6 in [4J) states that

n (1 -
ex:: ex::
x 2n )(1 + X 2n - 1
z2)(1 + x 2n - 1z- 2) = L X
rn2 2rn
Z
n=l

if z =I 0 and Ix I < 1.
(a) Show that x and z can be chosen to give the product representation

:+(r) = n (1 -
00

e21tint)(1 + e(2n-1)1tit)2.
n=l

This implies that t9(r) is never zero in H.


(b) If r E H prove that

( q2(' ; 1)
ti(r) = .
l1(r + 1)
(c) Prove that :)( -l/r) = (- ir)1/2[)(r).
Hint: If Sr = -l/r, find elements A and B of r such that

Sr ; 1 = A(r ; 1) and Sr + 1= B(r + 1).

91
4: Congruences for the coefficients of the modular function j

2. Let G denote the subgroup of r generated by the transformations Sand T 2 , where


Sr = -l/r and Tr = r + 1.

(a) If (; ~) E G prove that a == d (mod 2) and b == c (mod 2).

(b) If V E G prove that there exist elements A and B of r such that

Vr 2+ 1= A( 1) T ; and Vr +1= B(r + 1).

(c) If (; ~) E G and c > 0 prove that


ar
9( cr
+
+db) = e(a, b, c, d){ -i(cr + d)}1/29(r),

where Ie(a, b, c, d) I = 1. Express e(a, b, c, d) in terms of Dedekind sums.

Exercises 3 through 8 outline a proof (due to Mordell [28]) of the multiplica-


tivity of Ramanujan's function r(n). We recall that

L r(n)e21tint = (21t)-12~(r) = e 21tit n (1


00 00

- e21timt)24.
n= 1 m= 1

3. Let p be a prime and let k be an integer, 1 ~ k ~ p - 1. Show that there exists an


integer h such that

and that h runs through a reduced residue system mod p with k.

4. If p is a prime, define

Fir) = p11~(pr) + -1 p-1


L~ (r-+-k) .
P k=O P
Prove that:

(-1)
(b) F p -r- = r 12 Fp(r).

Note: Exercise 3 will be helpful for part (b).


5. Prove that Fir) = r(p)~(r), where r(p) is Ramanujan's function.

6. Use Exercises 4 and 5 to deduce the formulas


(a) r(pn+1) = r(p)r(pn) - p11 r(pn-1) for n 2 1.

(b) r(pa n ) = r(p)r(pa-1 n ) - p11 r(pa-2 n ) for rt. 2 2 and (n, p) = 1.

7. If rt. is an integer, rt. 2 0, and if (n, p) = 1, let


g(rt.) = r(pa n ) - r(pa)r(n).

Show that g(rt. + 1) is a linear combination of g(rt.) and g(rt. - 1) for rt. 2 2 and deduce
°
that g(rt.) = for all rt..

92
Exercises for Chapter 4

8. Prove that

r(m)r(n) = L dllr(~~).
dl(rn,n)

In particular, when (m, n) = 1 this implies r(m)r(n) = r(mn).


21tit
9. If r E H and x = e prove that

{S04.t r
CT s(n)x' = {j(r) - 12 3
} '~1 r(n)x',
where (J 5(0) = - 1/504. Equate coefficients of x n to obtain the identity
n n-l
(504)2 L (Js(k)(Js(n - k) = r(n + 1) - 984r(n) + L c(k)r(n - k).
k=O k=1

10. Use Exercise 9 together with Exercise 10 of Chapter 6 to prove that


65520 n-l
- - {(Jll(n)- r(n)} = r(n+ 1)+ 24r(l1)+ L c(k)r(n- k).
691 k=1

This formula, due to Lehmer [20], can be used to determine the coefficients c(n)
recursively in terms of r(n). Since the right member is an integer, the formula also
implies Ramanujan's remarkable congruence

r(n) == (J 11 (n) (mod 691).

93
Rademacher's series for
the partition function

5.1 Introduction
The unrestricted partition function p(n) counts the number of ways a positive
integer n can be expressed as a sum of positive integers ~ n. The number of
summands is unrestricted, repetition is allowed, and the order of the sum-
mands is not taken into account.
The partition function is generated by Euler's infinite product

1
(1) F(x) = n
00

m= 1
--m
1- x
=
00

L p(n)x n,
n=O

where p(O) = 1. Both the product and series converge absolutely and repre-
sent the analytic function F in the unit disk Ix I < 1. A proof of (1) and other
elementary properties of p(n) can be found in Chapter 14 of [4J. This chapter
is concerned with the behavior of p(n) for large n.
The partition function p(n) satisfies the asymptotic relation

eK~
p(n) '" r; as n ---+ 00,
4n y 3

where K = n(2/3)1/2. This was first discovered by Hardy and Ramanujan [13J
in 1918 and, independently, by J. V. Uspensky [52] in 1920. Hardy and
Ramanujan proved more. They obtained a remarkable asymptotic formula
of the form
~
(2) p(n) = L Pk(n) + O(n- 1 / 4 ),
k<a...jn

94
5.2: The plan of the proof

where LI. is a constant and P 1 (n) is the dominant term, asymptotic to


eK y'rJI(4nJ"3). The terms P 2(n), P 3(n), ... are of similar type but with smaller
constants in place of K in the exponential. Since p(n) is an integer the finite
sum on the right of (2) gives p(n) exactly when n is large enough to insure
that the error term is less than 1/2. This is a rare example of a formula which
is both asymptotic and exact. As is often the case with asymptotic formulas
of this type, the infinite sum

(3)

diverges for each n. The divergence of (3) was shown by D. H. Lehmer [21]
in 1937.
Hans Rademacher, while preparing lecture notes in 1937 on the work of
Hardy and Ramanujan, made a small change in the analysis which resulted
in slightly different terms Rk(n) in place of the Pk(n) in (2). This had a profound
effect on the final result since, instead of (2), Rademacher obtained a con-
vergent series,
00

(4) p(n) = L Rk(n).


k= 1

The exact form of the Rademacher terms Rk(n) is described below in Theorem
5.10. Rademacher [35] also showed that the remainder after N terms is
O(n - 1/4) when N is of order .)ii, in agreement with (2).
This chapter is devoted to a proof of Rademacher's exact formula for
p(n). The proof is of special interest because it represents one of the crowning
achievements of the so-called "circle method" of Hardy, Ramanujan and
Littlewood which has been highly successful in many asymptotic problems
of additive number theory. The proof also displays a marvelous application
of Dedekind's modular function 1J(r).

5.2 The plan of the proof


This section gives a rough sketch of the proof. The starting point is Euler's
formula (1) which implies

= ~
k
F(x)
n+ 1 ~
p(k)x
n+ 1
·fO
1 <
IX I < 1,
X k=O X

for each n ~ O. The last series is the Laurent expansion of F(x)/x n + 1 in the
punctured disk 0 < Ix I < 1. This function has a pole at x = 0 with residue
p(n) so by Cauchy's residue theorem we have

p(n)
1
= -2.
1rl
J F(x)
n+i dx,
eX

95
5: Rademacher's series for the partition function

where C is any positively oriented simple closed contour which lies inside
the unit circle and encloses the origin. The basic idea of the circle method is
to choose a contour C which lies near the singularities of the function F(x).
The factors in the product defining F(x) vanish whenever x = 1, x 2 = 1,
3
x = 1, etc., so each root of unity is a singularity of F(x). The circle method
chooses a circular contour C of radius nearly 1 and divides C into arcs
C htk lying near the roots of unity e21rih/k, where °::;
h < k, (h, k) = 1, and
k = 1,2, ... , N. The integral along C can be written as a finite sum of integrals
along these arcs,
N k-l J
J C k~l (h,k)=
h~O 1
Ch.k·

On each arc Ch,k the function F(x) in the integrand is replaced by an elemen-
tary function t/Jh,k(X) which has essentially the same behavior as F near the
singularity e21rih/k. This elementary function t/Jh,k arises naturally from the
functional equation satisfied by the Dedekind eta function l1(r). The functions
F and 11 are related by the equation

F(e 21rit ) = e1rit/12/11(r),

and the functional equation for 11 gives a formula which describes the behavior
of F near each singularity e21rih/k. The replacement of F by t/Jh,k introduces
an error which needs to be estimated. The integrals of the t/Jh,k along Ch,k are
then evaluated, and their sum over h produces the term Rk(n) in Rademacher's
series.
In 1943 Rademacher [38] modified the circle method by replacing the
circular contour C by another contour in the r-plane, where x = e 21rit . This
new path of integration simplifies the estimates that need to be made and
clarifies the manner in which the singularities contribute to the final formula.
The next section expresses Dedekind's functional equation in terms of F.
Sections 5.5 and 5.6 describe the path of integration used by Rademacher,
and Section 5.7 carries out the plan outlined above.

5.3 Dedekind's functional equation expressed


in terms of F
Theorem 5.1. Let F(t) = l/n~= 1 (1 - t rn ) and let

2nih 2nz) , = exp (2niH 2n) ,


(5) x = exp( -k- - J:2 ' x -k- - ~

where Re(z) > 0, k > 0, (h, k) = 1, and hH == - 1 (mod k). Then


/

(6) F(x) = e1ris(h t k) -Z)1/2 exp ( - n - -nz-2) F(x')


(k 12z 12k .

96
5.4: Farey fractions

Note. If Iz I is small, the point x in (5) lies near the root of unity e21rih/k,
whereas x' lies near the origin. Hence F(x') is nearly F(O) = 1, and Equation
(6) gives the behavior of F near the singularity e21rih/k. Aside from a constant
factor, for small Iz I, F behaves like

Zl/2 expC;z)-

PROOF. If (: ~) E r with c > 0, the functional equation for 1Jet) implies

(7) 1J: ) = 1J(~') {-i(CT + d)p/2 exp{n{at;c


d+ S(-d,C»)}.
T

where T' = (aT + b)/(CT + d). Since F(e 21rit ) = e1rit/12/1J(T), (7) implies

(8) . = F(e 21rlt)


F(e 21rlt) ., exp (ni(T12
- T')) {- i(CT + d)} 1/2

xexp{n{a t~cd + S(-d,C»)}


Now choose
hH + 1 iz +h
a = H, C = k, d = -h, b = and T=-k-·
k

Then

, iz- 1 +H
T =----
k

and (8) becomes

2nih
F exp -
(
2nz))
( k- - - k
- - - -2n))z 1/2
= F ( exp(2niH
k kz

n nz }
x exp{ 12kz - 12k + nis(h, k) .

When z is replaced by z/k this gives (6). o

5.4 Farey fractions


Our next task is to describe the path of integration used by Rademacher.
The path is related to a set of reduced fractions in the unit interval called'
Farey fractions. This section describes these fractions and some of their
properties.

97
5: Rademacher's series for the partition function

Definition. The set of Farey fractions of order n, denoted by F n , is the set of


reduced fractions in the closed interval [0, IJ with denominators ~n,
listed in increasing order of magnitude.

EXAMPLES

F1: ¥, t
F2 : ¥, t, t
F 3: ¥, t, t, i, t
F 4: ¥, t, t, t, i, i, t
F 5: ¥,!, t, t, ~, t, ~, i, i, ~, t
F 6: ¥, i, !, !, t, ~, t, ~, i, i, ~, i, t
F 7: ¥,~, i, !, t, ~, t, i, ~, !, 4, ~, i, ~, i, ~, i, ~, t

These examples illustrate some general properties of Farey fractions.


F or example, F n c F n+ l' so we get F n+ 1 by inserting new fractions in F n.
If (alb) < (cld) are consecutive in F n and separated in F n+ b then the fraction
(a + c)/(b + d) does the separating, and no new ones are inserted between
alb and cld. This new fraction is called the mediant of alb and cld.

Theorem 5.2. If(alb) < (cld), their mediant (a + c)/(b + d) lies between them.
PROOF

a +c a bc - ad c a +c bc - ad
and
b + d - b= b(b + d) > 0 d- b +d= d(b + d) > O. 0

t
The above examples show that and ~ are consecutive fractions in F n
for n = 5, 6, and 7. This illustrates the following general property.

Theorem 5.3. Given 0 ~ alb < cld ~ 1. If bc - ad = 1 then alb and cld are
consecutive terms in F nfor the following values of n:
max(b, d) ~ n ~ b +d- 1.

PROOF. The condition be - ad = 1 implies that alb and eld are in lowest
terms. If max(b, d) s n then b ~ nand d ~ n so alb and cld are certainly in
F n. Now we prove they are consecutive if n ~ b + d - 1. If they are not
consecutive there is another fraction hlk between them, alb < hlk < cld.
But now we can show that k ~ b + d because we have the identity
(9) k = (bc - ad)k = b(ck - dh) + d(bh - ak).
But the inequalities alb < hlk < cld show that ck - dh ~ 1 and bh - ak ~ 1
so k ~ b + d. Thus, any fraction hlk that lies between alb and cld has
denominator k ~ b + d. Therefore, if n ~ b + d - 1, then alb and cld must
be consecutive in F n • This completes the proof. D
98
5.5 : Ford circles

Equation (9) also yields the following theorem.

Theorem 5.4. Given 0 ~ alb < cld ~ 1 with be - ad = 1, let hlk be the
mediant of alb and cld. Then alb < hlk < cld, and these fractions satisfy
the unimodular relations
bh - ak = 1, ck - dh = 1.

PROOF. Since hlk lies between alb and cld we have bh - ak ~ 1 and
ck - dh ~ 1. Equation (9) shows that k = b + d if, and only if, bh - ak =
ck - dh = 1. D

The foregoing theorems tell us how to construct F n + 1 from F n •

Theorem 5.5. The set F n + 1 includes F n • Eachfraction in F n + 1 which is not in


F n is the mediant of a pair of consecutive fractions in F n • Moreover, if
alb < cld are consecutive in any F n' then they satisfy the unimodular
relation bc - ad = 1.
PROOF. We use induction on n. When n = 1 the fractions Oil and III are
consecutive and satisfy the unimodular relation. We pass from F 1 to F 2 by
inserting the mediant 1/2. Now suppose alb and cld are consecutive in F n
and satisfy the unimodular relation bc - ad = 1. By Theorem 5.3, they will
be consecutive in F m for all m satisfying
max(b, d) ~ m ~ b +d- 1.
Form their mediant hlk, where h = a + c, k = b + d. By Theorem 5.4
we have bh - ak = 1 and ck - dh = 1 so hand k are relatively prime.
The fractions alb and cld are consecutive in F m for all m satisfying
max(b, d) ~ m ~ b + d - 1, but are not consecutive in F k since k = b + d
and hlk lies in F k between alb and cld. But the two new pairs alb < hlk and
hlk < cld are now consecutive in F k because k = max(b, k) and k = max(d, k).
The new consecutive pairs still satisfy the unimodular relations bh - ak = 1
and ck - dh = 1. This shows that in passing from F n to F n + 1 every new
fraction inserted must be the mediant of a consecutive pair in F n' and the new
consecutive pairs satisfy the unimodular relations. Therefore F n+ 1 has these
properties if F n does. 0

5.5 Ford circles


Definition. Given a rational number hlk with (h, k) = 1. The Ford circle
belonging to this fraction is denoted by C(h,' k) and is that circle in the
complex plane with radius 1/(2k 2 ) and center at the point (hlk) + il(2k 2 )
(see Figure 5.1).
Ford circles are named after L. R. Ford [9J who first studied their
properties in 1938.

99
5: Rademacher's series for the partition function

. 1
radIus = 2k 2

h
k

Figure 5.1 The Ford circle C(h, k)

Theorem 5.6. Two Ford circles C(a, b) and C(c, d) are either tangent to each
other or thev do not intersect. They are tangent if, and only if, bc - ad =
± 1. In particular, Ford circles of consecutive Farey fractions are tangent
to each other.
PROOF. The square of the distance D between centers is (see Figure 5.2)

a c
b d
Figure 5.2

whereas the square of the sum of their radii is

(r
1
+ R)2 = ( 2b 2 + 2d 2
1 )2 .
The difference D 2 - (r + R)2 is equal to

D
2
- (r + R)2 = (
ad - bc)2
bd
(1
+ 2b 2 - 2d 2
1)2 (12b
- 2
1)2
+ 2d 2
(ad - bC)2 - 1
-- b2 d 2 >0
-'

Moreover, equality holds if, and only if (ad - bC)2 = 1. D


100
5.5: Ford circles

Theorem 5.7. Let h 1 /k 1 < h/k < h 2 /k 2 be three consecutive Farey fractions.
The points of tangency of C(h, k) with C(h 1 , k 1) and C(h 2 , k 2 ) are the points
h k1 i
(Xl(h, k) = k- k(k 2 + k 12) + k 2 + k12
and
h k2 i
(X2(h,k) = k + k(k 2 + k 22) + k 2 + k/"
Moreover, the point ofcontact a 1 (h, k) lies on the semicircle whose diameter
is the interval [h1/k b h/k].
PROOF. We refer to Figure 5.3. Write a 1 for a 1 (h, k). The figure shows that

1
2k 2
1 b
a
2k 1 2 ,,
,,
---"'-::.----- ---- --- - --
" "- ,,
,,
,,
,,
"- ,
"
h
k
Figure 5.3

To determine a and b we refer to the similar right triangles and we get


1
a 2ki k12
so
~ - 1 1 - k + k12 '
2

k k1 2k 2 + 2k 1 2
Similarly, we find
1 1
b 2k2 - ~ _ k12 - k2 1 k 12 - k2
so = 2k2 k2 + k 1 2 "
1 1 1 - k 12 + k2 ' b
2k 2 2k 2 + 2k 1 2
These give the required formula for a b and by analogy we get the correspond-
ing formula for a 2 •

101
5: Rademacher's series for the partition function

To obtain the last statement, it suffices to show that the angle (J in Figure
5.3 is n12. For this it suffices to show that the imaginary part of rJ. I (h, k) is
the geometric mean of a and a', where
, h hi 1
and a =- - - - a=- - a.
a = k(k 2 + k I 2) k kl kk 1
(See Figure 5.4.) Now

, kl ( 1 kl )
aa = k(k2 + k12) kk l - k(k 2 + k 12)
2
kl ( k ) 1
= k (k + k I 2) k I (k + k I 2) = (k 2 + k I 2)2 '
2 2 2

and this completes the proof. o

a' a

Figure 5.4

5.6 Rademacher's path of integration


For each integer N we construct a path P(N) joining the points i and i + 1
as follows. Consider the Ford circles for the Farey series F v. If h1/k 1 < h/k
< h2 1k 2 are consecutive in F N , the points of tangency of C(h t , k 1 ), C(h, k),
and C(h 2 , k 2 ) divide C(h, k) into two arcs, an upper arc and a lower arc.
P(N) is the union of the upper arcs so obtained. For the fractions 0/1 and 1/1
we use only the part of the upper arcs lying above the unit interval [0, 1].

EXAMPLE. Figure 5.5 shows the path P(3).


Because of Theorem 5.7, the path P(N) always lies above the row of semi-
circles connecting adjacent Farey fractions in F N.
The path P(N) is the contour used by Rademacher as a path of integration.
It is convenient at this point to discuss the effect of a certain change of variable
on each circle C(h, k).

Theorem 5.8. The transformation

102
5.6: Rademacher's path of integration

i + 1

o "3
1 1
"2
2
"3

Figure 5.5 The Rademacher path P(3)

maps the Ford circle C(h, k) in the t-plane onto a circle K in the z-plane of
radius t about the point z = ! as center (see Figure 5.6). The points of
contact ltl(h, k) and lt 2 (h, k) of Theorem 5.7 are mapped onto the points
k2 kk
zl(h,k) = k 2 + k/ + i k 2 + lk/
and

The upper arc joining lt 1(h, k) with lt 2 (h, k) maps onto that arc of K which
does not touch the imaginary z-axis.

\
\
o 1
!

z-plane
Figure 5.6

103
5: Rademacher's series for the partition function

PROOF. The translation T - (h/k) moves C(h, k) to the left a distance h/k,
and thereby places its center at i/(2k 2). Multiplication by - ik 2 expands the
radius to 1/2 and rotates the circle through n/2 radians in the negative
direction. The expressions for z 1(h, k) and z2(h, k) follow at once. 0
Now we obtain estimates for the moduli of z 1 and z 2.
Theorem 5.9. For the points z 1 and z 2 of Theorem 5.8 we have
k k
(10) IZ1(h, k)1 = , IZ 2(h, k)1 = .
~k2 + k12 ~k2 + k 22
Moreover, ifz is on the chord joining ZI and Z2 we have
fik
(11) Izl<~,

if h 1 /k 1 < h/k < h 2/k 2 are consecutive in F N. The length of this chord does
not exceed 2fik/N.
PROOF. For IZ112 we have
k 4 + k 2k 2
2
k2
Iztl = (k 2 + kJ)2 k 2 + k 1 2"
There is a similar formula for Iz21 2. This proves (10). To prove (11) we note
that if z is on the chord, then Iz I ::::; max( Iz1I, Iz21 ), so it suffices to prove that
fik fik
(12) IZ ll < "N and IZ 21 < "N.
For this purpose we use the inequality relating the arithmetic mean and the
root mean square:
2
k +_
_ k 1 < (k + k 1 2)1/2
2 - 2 .
This gives us
(k 2 + k 2)1/2 > k + k 1 > N + 1 > !i
1 - fi - fi fi'
so (10) and (12) imply (11). The length of the chord is ::::; Iz11 + Iz21. D

5.7 Rademacher's convergent series for pen)


Theorem 5.10. If n ~ 1 the partition function p(n) is represented by the
convergent series

1 (X) d
p(n) =
ny
M L Ak(n)jk -dn
2 k= 1

104
5.7: Rademacher's convergent series for p(n)

where

A k(n) ~ eftis(h,k)-2ftinh/k.
= LJ
O~h<k
(h,k)= 1

PROOF. We have

(13) p(n)
1
= -2.
'Ttl
I
eX
F(x)
n+f dx where F(x) = n (1 -
00

m=l
X
m
)-l =
00

L p(n)x
n=O
n
;

C is any positively oriented closed curve surrounding x = 0 and lying inside


the unit circle. The change of variable

maps the unit disk Ix I ~ 1 onto an infinite vertical strip of width 1 in the
r-plane, as shown in Figure 5.7. As x traverses counterclockwise a circle of

x-plane
o
t-plane

Figure 5.7

radius e- 2ft with center at x = 0, the point r varies from i to i + 1 along a


horizontal segment. We replace this segment by the Rademacher path P(N)
composed of the upper arcs of the Ford circles formed for the Farey series
F N. Then (13) becomes

p(n) = r
J
i
i+ 1
F(e 2nh )e - 2ninr dr = r
Jp(N)
F(e 2nit)e - 2nint dr.

In this discussion the integer n is kept fixed and the integer N will later be
allowed to approach infinity. We can also write

fP(N) ktl O"~<k L(h'k)


(h,k)= 1
= t,.;, l(h'kl
where y(h, k) denotes the upper arc of the circle C(h, k), and Lh,k is an
abbreviation for the double sum over hand k.

105
5: Rademacher's series for the partition function

Now we make the change of variable

so that
h iz
t = k + k2 •
Theorem 5.8 shows that this maps C(h, k) onto a circle K of radius! about
z = ! as center. The arc y(h, k) maps onto an arc joining the points zl(h, k)
and z2(h, k) in Figure 5.6. We now have

p(n) = L J Z2
(h. k)F(ex p(21rih _ 2~Z)) ~ e- 2ninh/k e2nnz/k2
h,k zt{h,k) k k k
dz

Z
= L ik-2e-2ninh/k J 2(h.k) e2nnZ/k2F(exp(21rih _ 2~Z)) dz.
h,k zt{h,k) k k
Now we use the transformation formula for F (Theorem 5.1) which states
that
Z)1/2 ( 1r 1rZ )
F(x) = w(h, k) ( k exp 12z - 12k 2 F(x'),
where
21rih 21rZ)
x = exp ( -k- - k2 ' x, = exp (21riH
-k- - 21r) ,
-;-
and
w(h, k) = enis(h,k), (h, k) = 1.
hH == -1 (mod k),
Denote the elementary factor ZI/2 exp[,,-/(12z)- 1rz/{12k 2)J by \fJk{Z) and
split the integral into two parts by writing
F(x') = 1+ {F(x') - I}.
We then obtain
p(n) = L ik- 5 / 2W(h, k)e-2ninh/k(11(h, k) + 12(h,k))
h, k

where
Z 2(h'k)
1 1(h, k) =
Jzdh,k) qJk(z)e2nnz/k2 dz

and
Z

I 2(h, k) =
Jzt{h,2(h,k)k) qJ k(Z){F(exp (21riH
-k- - -21r)) - 1} e2nnz/k2 dz.
Z

We show next that 1 2 is small for large N. The path of integration in the
z-plane can be moved so that we integrate along the chord joining z 1 (h, k)
and z2(h, k). (See Figure 5.8.) We have already estimated the length of this

106
5.7: Rademacher's convergent series for p(n)

Figure 5.8

chord; it does not exceed 2J2k/N. On the chord itself we have Izi
~ max{lzll, IZ21} < J2k/N. Note also that the mapping w = l/z maps
the disk bounded by K onto the half-plane Re(w) ~ 1. Inside and on the
circle K we have 0 < Re(z) ~ 1 and Re(l/z) ~ 1, while on K itself we have
Re(l/z) = 1.
Now we estimate the integrand on the chord. We have

= Iz1
1
/
2
exp { 1~ ReG) - 1~2 Re(Z)}
x e2nltRe<zl/k21 m~1 p(m)e2ltiHm/ke- 2ltm/z I

~ Iz1 1/ 2 exp{~ Re(~)}e2nlt/k2 I: p(m)e-2ltmRe(1/z)


12 z m=l
00
< Izll/2e2n1t L p(m)e- 21t (m- O/24»Re O /z)

m=l
00

~ Izll/2e2n1t L p(m)e- 21t(m- O/24»


m=l
00
= Izll/2e2n1t L p(m)e- 21t ,(24m-1)/24
m=l
oc
< Izll/2e2n1t L p(24m - 1)e- 21t (24m-1)/24
m=l
oc
= Izll/2e2n1t L p(24m - 1)y24m-l (where y = e- 21t /24 )
m=l

107
5: Rademacher's series for the partition function

where
00

c = e 2nn I p(24m - 1)y24m- 1.


m=l
The number c does not depend on z or on N. (It depends on n, but n is fixed
in this discussion.) Since z is on the chord we have Iz I < ~ k/N so the
integrand is bounded by c2 1/4(k/N)1/2. The length of the path is less than
2~k/N, so altogether we find
11 2(h,k)1 < Ck 3 / 2N- 3 / 2

for some constant C, and therefore

I
Ih, k
ik-S/2W(h,k)e-21tinh/kI2(h,k)1 < f
k= 1 0
I
h< k
~
Ck- 1 N- 3 / 2
(h,k)= 1
N
~ CN- 3 / 2 1 I = CN- 1/2.
k=l
This means we can write
N
(14) p(n) = I I ik-S/2w(h,k)e-2ninh/kl1(h,k) + O(N- 1/2).
k=l O~h<k
(h,k)= 1
Next we deal with I 1(h, k). This is an integral joining zl(h, k) and z2(h, k)
along an arc of the circle K in Figure 5.8. We introduce the entire circle K
as path of integration and show that the error made is also O(N- 1 / 2 ). We have

1 1 (h,k) =
1 f
K(-)
-
0
Zdh' k)
-
JO
z2(h,k)
=
1 K(-)
-J 1 -J 2 ,

where K( - ) denotes that the integration is in the negative direction along K.


To estimate IJ 1 I we note that the length of the arc joining 0 to z 1(h, k) is less
than

Since Re(1/z) = 1 and 0 < Re(z) ~ 1 on K the integrand has absolute value

I\P (z)e2nnz/k21 = e2nnRe(z)/k2IzI1/2 exp{~ Re(~) _ _ 1r_2 Re(Z)}


k 12 z 12k
e2nn21/4k1/2en/12
< - -N-1 /-
- 2
--

so that

108
5.7: Rademacher's convergent series for p(n)

where C 1 is a constant. A similar estimate holds for IJ 21 and, as before,


this leads to an error term O(N- 1 / 2 ) in the formula for p(n). Hence (14)
becomes

p(n) = f
k =1 0 h < k
L
~
ik - 5/2 W (h, k)e - 2ninh/k
J
rK( - )
'1\(z)e2nnz/k2 dz + O(N- 1 / 2).
(h,k)= 1

Now we let N -+ 00 to obtain

where
Ak(n) = L e1tis(h, k) - 21tinhlk.
O~h<k
(h,k)= 1

The integral can be evaluated in terms of Bessel functions. The change of


variable
1 1
W=- , dz = - 2dw,
z W
gives us

_
p(n) -
1i k~/k(n)k
00 - 512 51 + ooi - 512
1- ooi W exp
{2
nw n( 1) I}
12 + k 2 n - 24 ~ dw.
Now put t = nw/12 and the formula becomes
n )3 /2 1 fC+ooi 5 2 {n 2 ( 1 ) I}
2n(12 k~lAk(n)k-5/2 2ni / exp t + 6k2 n - 24 t dt
00
p(n) = c-ooi C

where c = n/12. Now on page 181 of Watson's treatise on Bessel functions


[53] we find the formula
1 )V fC+ ooi
Iv(z) = ( 2:z . t-V-let+(Z2/4t) dt (if c > 0, Re(v) > 0),
2nl c- ooi

where I v(z) = i- vJ v(iz). Taking

and v = 3/2 we get


3/2 1 )-3 /4
~(1 ))
(12n)3 2k~lAk(n)k-5/2 n- (n
1
00
( n - 24
p(n) = (2n) 6-3/4k-3/2 13 / 2 kV"3\.n- 24.
1 )-3 /4
(
(2n) n - 24
(24)3/4
00

k~l Ak(n)k
-1 (n V"3~(
13 / 2 k
1 ))
\.n - 24 ·
109
5: Rademacher's series for the partition function

But Bessel functions of half odd order can be red uced to elementary functions.
In this case we have

13 / 2 (Z ) -_ ftz ~
d (sinh z) .
n z z

Introducing this in the previous formula we finally get Rademacher's


formula,

1 00 d
p(n) = -- L A (n)k
k
1 2
/ - D
nJ2k=l dn

Exercises for Chapter 5


1. Two reducedfractionsa/b andc/daresaid to besimilarlyorderedif(c - a)(d - b) ~ O.
Let al/b t < a2/b2 < ... denote the Farey fractions in F n •
(a) Prove that any two neighbors ai/bi and ai+ l/bi+ 1 are similarly ordered.
(b) Prove also that any two second neighbors ai/bi and ai+ 2/bi+ 2 are similarly ordered.
Note: Erdos [8J has shown that there is an absolute constant c > 0 such that the
kth neighbors a/bi and ai+k/bi+k in Fn are similarly ordered if 11 > ck.
2. If a, b, c, d are positive integers such that alb < c/d and if Aand J.1 are positive integers,
prove that the fraction

(}=Aa+J.1c
Ab + J.1d
lies between alb and c/d, and that (c - d(})/((}b - a) = ).. /J.1. When A = J.1, () is the
mediant of alb and c/d.
3. If bc - ad = 1 and n > max(b, d), prove that the terms of the Farey sequence F n
between alb and c/d are the fractions of the form (Aa + J.1c)/(Ab + J.1d) for which )..
and J.1 are positive relatively prime integers with Ab + J.1d ~ n. Geometrically, each
pair (A, J.1) is a lattice point (with coprime coordinates) in the triangle determined by
the coordinate axes and the line bx + dy = n. Neville [29] has shown that the
number of such lattice points is

3 n2
2 -b + O(n log n).
n d
This shows that for a given n, the number of Farey fractions between alb and c/d is
asymptotically proportional to l/(bd), the length of the interval [alb, c/d].

Exercises 4 through 8 relate Farey fractions to lattice points in the plane.


In these exercises, n ~ 1 and T" denotes the set of lattice points (x, y) in the
triangular region defined by the inequalities
1 ::::; x ::::; n, 1 ::::; y ::::; n, n + 1 ::::; x +y ::::; 2n.

110
Exercises for Chapter 5

Also, 1;,' denotes the set of lattice points (x, y) in 1;, with relatively prime
coordinates.
4. Prove that alb and c/d are consecutive fractions in the Farey sequence F n if, and only
if, the lattice point (b, d) E T~.

5. Prove that L(b.d)E r;, 1/(bd) = 1. Hint: Theorem 5.5.


6. Assign a weight f(x, y) to each lattice point (x, y) and let Sn be the sum of all the
weights in Tn,
Sn = L f(x, y).
(x, y) E Tn

(a) By comparing the regions T,. and T,. _ 1 for r ~ 2 show that
r-1 r-l

Sr - Sr-1 = f(r, r) + L {f(k, r) + f(r, k)} - L f(k, r - k),


k=1 k=l

and deduce that


n n r- 1 n r-l

Sn = L f(r, r) + L L {f(k, r) + f(r, k)} - L L f(k, r - k).


r=1 r=2 k=1 r=2k=1

Note: If f(x, y) = 0 whenever (x, y) > 1 this reduces to a formula of J. Lehner


and M. Newman [25],
r-1
(15) L f(x, y) = f(l, 1) + L L {f(k, r) + f(r, k) - f(k, r - k)}.
(x,Y)ET~ r=2 k=1
(k,r)= 1

This relates a sum involving Farey fractions to one which does not.

7. Let

1
S =
n (b, d)
L T~ bd(b + d)
E

(a) Use Exercise 5 to show that 1/(2n - 1) S Sn S 1/(n + 1).


(b) Choose f(x,y) = 1/(xy(x + y)) in (15) and show that
3 n r 1
Sn = - -
2
2
r
L= L=
1 k 1
2
r (r + k)
.

(k,r)= 1

When n ~ CJJ this gives a formula of Gupta [12],

00 r 1 3
L L
r= 1 k= 1
2
r (r + k)
=-.
4
(k,r)= 1

8. Exercise 7(a) shows that Sn ~ 0 as n ~ Cf:.;. This exercise outlines a proof of the
asymptotic formula

(16)
_ 12 log 2
Sn - 2 + 0
(log
2
n)
n n n
obtained by Lehner and Newman in [25].

111
5: Rademacher's series for the partition function

Let

Ar = L
r

k=l
2
r (r
1
+ k)
±I
k= 1 dl(r,k) r (r
/(d)
+ k)
,
(k,r)= 1
so that

r>n

(a) Show that

Ar
=" ~ _d_J.l(r_/d_)
L. L. 3
dl r h = 1 r (h + d)

and deduce that

Ar = log 2 -qJ(r)
3 + 0
r
(1 3 L IJ.l(d) I
r dlr
).
(b) Show that L~=l Ldlr IJ.l(d)/ = O(n log n) and deduce that

1 (lOg
L 3LIJ.l(d)/ = 0 - 2
n) .
r dlr
r> n n

(c) Use the formula Lr ~ n qJ(r) = 3n 2 /n 2 + O(n log n) (proved in [4J, Theorem 3.7) to
deduce that

(d) Use (a), (b), and (c) to deduce (16).

112
Modular forms with
multiplicative coefficients

6.1 Introduction
The material in this chapter is motivated by properties shared by the discrimi-
nant L\(r) and the Eisenstein series
1
G 2k(r) = L
(m,n):¢:(O, 0) (m + nr)
2k'

where k is an integer, k ~ 2. All these functions satisfy the relation

(1) f(:: : ~) = (c~ + d)' f(~),


where r is an integer and (: ~) is any element of the modular group r.
The function ~ satisfies (1) with r = 12, and G 2k satisfies (1) with r = 2k.
Functions satisfying (1) together with some extra conditions concerning
analyticity are called modular forms. (A precise definition is given in the
next section.)
Modular forms are periodic with period 1 and have Fourier expansions.
For example, we have the Fourier expansion,
00

~(r) = (2n)12 L r(n)e21tint,


n=l

where r(n) is Ramanujan's function, and


2(2ni)2k 00 2•
G2k(~) = 2(2k) + (2k _ 1)! n~/'2k-l (n)e ,,,nt,

where (J a(n) is the sum of the cxth powers of the divisors of n.

113
6: Modular forms with multiplicative coefficients

Both r(n) and O"(X(n) are multiplicative arithmetical functions; that is, we
have
(2) r(m)r(n) = r(mn) and
They also satisfy the more general multiplicative relations

(3) r(m)r(n) = L dllr(:~)


dl(m,n)

and

(4) O"(X(m)O"(X(n) = L d(XO"(X(md~)


dl(m,n)

for all positive integers m and n. These reduce to (2) when (m, n) = 1.
The striking resemblance between (3) and (4) suggests the problem of
determining all modular forms whose Fourier coefficients satisfy a multi-
plicative property encompassing (3) and (4). The problem was solved by
Hecke [16J in 1937 and his solution is discussed in this chapter.

6.2 Modular forms of weight k


In this discussion k denotes an integer (positive, negative, or zero), H denotes
the upper half-plane, H = {r: Im(r) > O}, and r denotes the modular group.

Definition. A function f is said to be an entire modular form of weight k if it


satisfies the following conditions:

(a) f is analytic in the upper half-plane H.

(b) f(:: : :) = (cr + d)kf(r) whenever (: :) E r.


(c) The Fourier expansion of f has the form
00

f(r) = L c(n)e21tint.
n=O

Note. The Fourier expansion of a function of period 1 is its Laurent


expansion near the origin x = 0, where x = e 21tit . Condition (c) states that
the Laurent expansion of an entire modular form contains no negative
powers of x. In other words, an entire modular form is analytic everywhere
in H and at ioo.

The constant term c(O) is called the value of f at ioo, denoted by f(ioo).
If c(O) = 0 the function f is called a cusp form (" Spitzenform" in German),
and the smallest r such that c(r) i= 0 is called the order of the zero of f at i 00.
It should be noted that the discriminant ~ is a cusp form of weight 12 with
a first order zero at i 00. Also, no Eisenstein series G2k vanishes at i 00.

114
6.3: The weight formula for zeros of an entire modular form

Warning. Some authors refer to the weight k as the "dimension - k"


or the "degree -k." Others write 2k where we have written k.

In more general treatments a modular form is allowed to have poles in


H or at i 00. This is why forms satisfying our conditions are called entire
forms. The modular function J is an example of a nonentire modular form of
weight 0 since it has a pole at ioo. Also, to encompass the Dedekind eta func-
tion there are extensions of the theory in which k is not restricted to integer
values but may be any real number, and a factor e(a, b, c, d) of absolute
value 1 is allowed in the functional equation (b). This chapter treats only
entire forms of integer weight with multiplier e = 1.
The zero function is a modular form of weight k for every k. A nonzero
constant function is a modular form of weight k only if k = O. An entire
modular form of weight 0 is a modular function (as defined in Chapter 2) and
since it is analytic everywhere in H, including the point ioo, it must be constant.
Our first goal is to prove that nonconstant entire modular forms exist
only if k is even and ~ 4. Moreover, they can all be expressed in terms of the
Eisenstein series G4 and G6. The proof is based on a formula relating the
weight k with the number of zeros of f in the closure of the fundamental
region of the modular group.

6.3 The weight formula for zeros of an entire


modular form
We recall that the fundamental region R r has vertices at the points p, i,
P + 1 and ioo. If f has a zero of order r at a point p we write r = N(p).

Theorem 6.1. Let f be an entire modularform ofweight k which is not identically


zero, and assume f has N zeros in the closure of the fundamental region
R r , omitting the vertices. Then we have the formula

(5) k = 12N + 6N(i) + 4N(p) + 12N(ioo).

PROOF. The method of proof is similar to that of Theorem 2.4 where we


proved that a modular function has the same number of zeros as poles in
the closure of R r . Since f has no poles we can write

N = _1 r f'(7:) d7:.
2ni JaR f(7:)
The integral is taken along the boundary of a region R formed by truncating
the fundamental region by a horizontal line y = M with sufficiently large M.
The path 8R is along the edges of R with circular detours made around the
vertices i, p and p + 1 and other zeros which might occur on the edges. By

115
6: Modular forms with multiplicative coefficients

calculating the limiting value of the integral as M -+ 00 and the circular


detours shrink to their centers we find, as in the proof of Theorem 2.4,
k 1 1
(6) N = 12 -"2 N(i) - 3 N(p) - N(ioo).

The only essential difference between this result and the corresponding
formula obtained in the proof of Theorem 2.4 is the appearance of the term
k112. This comes from the weight factor (cr + d)k in the functional equation
f(A(r)) = (cr + d)kf(r),
where A(r) = (ar + b)/(cr + d). Differentiation of this equation gives us
f'(A(r))A'(r) = (cr + d)kf'(r) ,+ kc(cr + d)k-l fer)
from which we find
f'(A(r))A'(r) = f'(r) + ~.
f(A(r)) fer) cr + d
Consequently, for any path y not passing through a zero we have

_1
2ni
f
A(y)
f'(u) du = _1
feu) 2ni
f
y
f'(r) dr
f(r)
+ _1
2ni
f
y
~ dr.
cr +d
Therefore the integrals along the arcs (2) and (3) in Figure 2.5 do not cancel
as they did in the proof of Theorem 2.4 unless k = O. Instead, they make a
contribution whose limiting value is equal to

-k
2ni
Ii ~dr = 2ni
p
-k -k (ni 2ni)
(log i-log p) = 2ni 2 - 3
k
= 12"

The rest of the proof is like that of Theorem 2.4 and we obtain (6), which
implies (5). D

From the weight formula (5) we obtain the following theorem.

Theorem 6.2

(a) The only entire modular forms of weight k = 0 are the constant
functions.
(b) Ifk is odd, ifk < 0, or ifk = 2, the only entire modular form of weight k
is the zero function.
(c) Every nonconstant entire modular.form has weight k ~ 4, where k is even.
(d) The only entire cusp form of weight k < 12 is the zero function.

PROOF. Part (a) was proved earlier. To prove (b), (c) and (d) we simply refer
to the weight formula in (5). Since each integer N, N(i), N(p) and N(ioo) is
nonnegative, k must be nonnegative and even, with k ~ 4 if k =1= O. Also,
if k < 12 then N(ioo) = 0 so f is not a cusp form unless f = O. D
116
6.4: Representation of entire forms in terms of G4 and G6

6.4 Representation of entire forms in terms


ofG 4 and G 6
In Chapter 1 it was shown that every Eisenstein series Gk with k > 2 is a
polynomial in G4 and G6 . This section shows that the same is true of every
entire modular form. Since the discriminant ~ is a polynomial in G4 and
G6 ,
L\ = g2 3 - 27g 3 2 = (60G 4)3 - 27(140G 6)2,
it suffices to show that all entire forms of weight k can be expressed in terms
of Eisenstein series and powers of ~. The proof repeatedly uses the fact that
the product fg of two entire forms f and g of weights WI and W2' respectively,
is another entire form of weight WI + W2' and the quotient f /g is an entire
form of weight WI - W2 if g has no zeros in H or at ioo.
Notation. We denote by M k the set of all entire modular forms of weight k.
Theorem 6.3. Let f be an entire modular form of even weight k ~ 0 and define
Go(r) = 1 for all r. Then f can be expressed in one and only one way as a
sum of the type
[k/12]
(7) f = L arGk-12r~r,
r=O
k-12r*2
where the ar are complex numbers. The cusp forms of even weight k are
those sums with ao = O.
PROOF. If k < 12 there is at most one term in the sum and the theorem can
be verified directly. If f has weight k < 12 the weight formula (5) implies
N = N(ioo) = 0 so the only possible zeros of f are at the vertices p and i.
For example, if k = 4 we have N(p) = 1 and N(i) = O. Since G4 has this
propertY,f/G 4 is an entire modular form of weight 0 and therefore is constant,
so f = aOG4 . Similarly, we find f = aOG k if k = 6,8 or 10. The theorem
also holds trivially for k = 0 (since f is constant) and for k = 2 (since the
sum is empty). Therefore we need only consider even k ~ 12.
We use induction on k together with the simple observation that every
cusp form in M k can be written as a product L\h, where hEMk-12'
Assume the theorem has been proved for all entire forms of even weight
<k. The form G k has weight k and does not vanish at ioo. Hence if
c = f(ioo)/Gk(ioo) the entire form f - cG k is a cusp form in M k so
f - cG k = L\h, where hEMk- 12' Applying the induction hypothesis to h
we have
[(k-12)/12] [k/12]
h= L brGk-12-12r~r = L br_lGk_12r~r-1.
r=O r= 1
k- 12 - 12r* 2 k-12r*2
117
6: Modular forms with multiplicative coefficients

Therefore f = cG k + ~h is a sum of the type shown in (7). This proves, by


induction, that every entire form of even weight k has at least one representa-
tion of the type in (7). To show there is at most one such representation we
need only verify that the products Gk-12r~r are linearly independent. This

as an exercise for the reader.


°
follows easily from the fact that ~(ioo) = but G 2r(ioo) =1= 0. Details are left
D

Since both ~ and G 2r can be expressed as polynomials in G4 and G 6 ,


Theorem 6.3 also shows that f is a polynomial in G4 and G6. The exact
form of this polynomial is described in the next theorem.

Theorem 6.4. Every entire modular form f of weight k is a polynomial in G4


and G6 of the type

(8) f = L ca,bG4aG6b
a,b
where the ca,bare complex numbers and the sum is extended over all integers
°
a ~ 0, b ~ such that 4a + 6b = k.
PROOF. °
If k is odd, k < or k = 2 the sum is empty and f is 0. If k = 0, f is
constant and the sum consists of only one term, co, 0. If k = 4, 6, 8 or 10
then each of the respective quotients f1G 4 , f1G 6 , flG 42 and fl(G 4 G6) is an
°
entire form of weight and hence is constant. This proves (8) for k < 12 or
k odd. To prove the result for even k ~ 12 we use induction on k.
Assume the theorem has been proved for all entire forms of weight < k.
Since k is even, k = 4m or k = 4m + 2 = 4(m - 1) + 6 for some integer
m ~ 3. In either case there are nonnegative integers rand s such that
k = 4r + 6s. The form g = G4rG 6s has weight k and does not vanish at ioo.
Hence if c = f(ioo)lg(ioo) the entire form f - cg is a cusp form in M k so
f - cg = ~h where hEM k- 12 . By the induction hypothesis, h can be
°
expressed as a sum as in (8), taken over all a ~ 0, b ~ such that 4a + 6b =
k - 12. Multiplication by ~ gives a sum of the same type with 4a + 6b = k.
Hence f = cg + ~h is also a sum of the required type and this proves the
theorem. D

6.5 The linear space M k and the


subspace Mk,o
The results of the foregoing section can be described in another way. Let
M k denote the set of all entire forms of weight k. Then M k is a linear space
over the complex field (since M k is closed under addition and under multi-
plication by complex scalars). Theorem 6.3 shows that M k is finite-dimen-
sional with a finite basis given by the set of products Gk-12r~r occurring
in the sum (7). There are [k/12] + 1 terms in this sum if k =1= 2 (mod 12),

118
6.6: Classification of entire forms in terms of their zeros

and one less term if k == 2 (mod 12). Therefore the dimension of the space
M k is given by the formulas

[ :2J if k == 2 (mod 12),


(9) dim M k =
[ lk
2
J+ 1 if k "# 2 (mod 12).

Another basis for M k is the set of products G4 aG 6 b where a 2 0, b 2 and


4a + 6b = k (see Exercise 6.12).
°
The set of all cusp forms in M k is a linear subspace of M k which we denote
by M k, o. The representation in Theorem 6.3 shows that
(10) dim Mk,o = dim M k - 1
since the cusp forms are those sums in (7) with ao = 0.
We also note that if k 2 12, f E M k , 0 if and only if f = I1h, where
hEMk- 12· Therefore the linear transformation T : M k- 12 --+ M k, 0 defined by
T(h) = ~h

establishes an isomorphism between Mk,o and M k- 12 . Consequently, if


k 2 12 we have
(11 )
The two formulas (11) and (10) imply
dim M k = 1 + dim M k - 12

if k ~ 12. This equation, together with the fact that dim M k = 1,0, 1, 1, 1, 1
when k = 0, 2, 4, 6, 8, 10, gives another proof of (9).
EXAMPLES. Formula (9) shows that
dim M k = 1 if k = 4, 6, 8, 10, and 14.
Corresponding basis elements are G4 , G6 , G4 2, G4 G6 , and G4 2G6 .
Formulas (11) and (9) together show that
dim Mk,o = 1 if k = 12, 16,18,20,22, and 26.
Corresponding basis elements are 11, I1G 4 , I1G 6 , I1G 4 2, I1G 4 G6 , and I1G 4 2G 6 •

6.6 Classification of entire forms in terms of


their zeros
The next theorem gives another way of expressing all entire forms in terms
of G 4 , G 6 , 11 and Klein's modular invariant J.

Theorem 6.5. Let f be an entire form of weight k and let z 1, ... , Z N denote the
N zeros o.f f in the closure o.f R r (omitting the vertices) with zeros of order

119
6: Modular forms with multiplicative coefficients

N(p), N(i) and N(ioo) at the vertices. Then there is a constant c such that

(12) 1(1:) = cG4(1:)N(P)G6(r)N(i)~(r)N(iOO)~(r)N


k=1
n {J(r) -
N
J(Zk)}'

PROOF. The product


N
g(r) = n {J(r) -
k=1
J(Zk)}

is a modular function with its only zeros in the closure of R r at Z1, . . . , ZN


and with a pole of order N at ioo. Since ~ has a first-order zero at ioo, the
product ~Ng is an entire modular form of weight 12N which, in the closure
of R r , vanishes only at ZI' ... , ZN' Therefore the product
h= G4N(p)G6N(i)~N(ioo)~Ng

has exactly the same zeros asfin the closure of R r . Also, h is an entire modular
form having the same weight as f since
k = 4N(p) + 6N(i) + 12N(ioo) + 12N.

Therefore f /h is an entire form of weight 0 so f /h is constant. This proves (12).


D

6.7 The Heeke operators Tn


Heeke determined all entire forms with multiplicative coefficients by intro-
ducing a sequence of linear operators T", n = 1, 2, ... , which map the linear
space M k onto itself. Heeke's operators are defined as follows.

Definition. For a fixed integer k and any n = 1, 2, ... , the operator T" is
defined on M k by the equation

(13)

In the special case when n is prime, say n = p, the sum on d contains


only two terms and the definition reduces to the formula

(14) (Tpf)('r) = pk-I f(pr) +~ Pilf(r + b).


P b=O P
The sum on b is the operator encountered in Chapter 4. It maps functions
automorphic under r onto functions automorphic under the congruence
subgroup r o(p).
We will show that T" maps each f in M k onto another function in M k •
First we describe the action of T" on the Fourier expansion of f

120
6.7: The Heeke operators T,.

Theorem 6.6. Iff E M k and has the Fourier expansion


00

f(1:) = L c(m)e21timt,
m=O
then 1;,f has the Fourier expansion
00

(15) (1;, f) (1:) = L Yn(m)e21timt,


m=O
where

(16) Yn(m) = ~
i..J dk-l c (mn)
J2 .
dl(n,m)

PROOF. From the definition in (13) we find


d-l 00

(Tnf)(1:) = nk - 1 L d- L L c(m)e21tim(nt+bd)/d2
k

din b=O m=O

= L L ~ )k-l c(m)e21timnt/d2 -1 L e21timb/d.


00 ( d-l

m=O din d d b=O


The sum on b is a geometric sum which is equal to d if dim, and is 0 otherwise.
Hence
(1;,f)(1:) = L L
00 (
~ )k- c(m)e21timnt/d2.
1

m= 0 din, dim d
Writing m = qd we have

(T,J)(r) =
q=O
f L (~)k- 1 c(qd)e21<iQnt/d,
din d
In the sum on d we can replace d by nld to obtain

(T"f)(r) = f
q=O
d din
L dk-lc(qn)e21tiqdt.
If x = e 21tit the last sum contains powers of the form x qd . We collect those
terms for which qd is constant, say qd = m. Then q = mid and dim so

(T" f)(r) = L L dk-l c(mn)


00
(j2 x m,
m=O dln,dlm

which implies (16). o


Our next task is to prove that 1;, maps M k into itself. For this purpose
we note that the definition of 1;, f can be written in a slightly different form.
We write n = ad and let
a1: + b
A1:=-d-·

121
6: Modular forms with multiplicative coefficients

Then (13) takes the form


1
(17) (T"f)(r) = nk - I
L d- k f(Ar) = - L ak f(Ar).
a ~ I , ad = n n a ~ I , ad = n
O~b<d O~b<d

The matrix (~ ~) which represents A has determinant ad = n. To deter-


mine the behavior of T"f under transformations of the modular group r
we need some properties of transformations with determinant n. These are
described in the next section.

6.8 Transformations of order n


Let n be a fixed positive integer. A transformation of the form
ar +b
Ar = er + d'
where a, b, e, d are integers with ad - be = n, is called a transformation of
order n. It can be represented by the 2 x 2 matrix

A = G~)
where, as usual, we identify each matrix with its negative.
We denote by r(n) the set of all transformations of order n. The modular
group r is r(l).
Two transformations Al and A z in r(n) are called equivalent, and we
write A I A z , if there is a transformation V in r such that
I'"V

Al = VAz·
The relation is obviously reflexive, symmetric, and transitive, and hence
I'"V

is an equivalence relation. Consequently, the set r(n) can be partitioned


into equivalence classes such that two elements of r(n) are in the same class
if, and only if, they are equivalent. The next theorem describes a set of
representatives.

Theorem6.7. In every equivalence class of r(n) there is a representative of


triangular form

a b)
(o where d > O.
d'

PROOF. Let A = (: ~) be an arbitrary element of r(n). If c = 0 there is


nothing more to prove. If e # 0 we reduce the fraction - ale to lowest terms.
That is, we choose integers rand s such that sir = -ale and (r, s) = 1.

122
6.8: Transformations of order n

Next we choose two integers p and q such that ps - qr = 1 and let

v= (~ ~}
Then V E rand

VA = (p q) (a b) (pa + qc pb + qd).
r s c d = ra + sc rb + sd
Since ra + sc = 0 and det(VA) = det V det A = n we see that VA E r(n) so
VA ~ A. Hence V A or its negative is the required representative. D

Theorem 6.8. A complete system qf nonequivalent elements in r(n) is given by


the set qf transformations ql' triangular form

(18)

where d runs through the positive divisors of n and, for each fixed d,
a = njd, and b runs through a complete residue system modulo d.
PROOF. Theorem 6.7 shows that every element in r(n) is equivalent to one
of the transformations in (18). Therefore we need only show that two such
transformations, say

At -
_(a0 1 b1 )
d and
l

are equivalent if, and only if,


(19) and
If(19) holds then b z = b 1 + qd 1 for some integer q and we can take

V = (~ ~}
Then V Al = A z so Al ~ A z .
Conversely, if A 1 ~ A z there is an element

in r such that A z = VA 1. Therefore

(20) (a z b z ) = (p q)(a 1 b 1) = (pal pb 1 + qd 1).


o dz r sOd1 ra 1 rb1 + sd 1
Equating entries we find ra 1 = 0 so r = 0 since a 1 # 0 because a1d1 = n ~ 1.
Now ps - qr = 1 so ps = 1 hence both p and s are 1 or both are -1. We can
assume p = s = 1 (otherwise replace V by - V). Equating the remaining
entries in (20) we find a z = aI' d z = d 1, b 2 = b 1 + qd 1, so bz == b 1 (mod d 1).
This completes the proof. D
123
6: Modular forms with multiplicative coefficients

Note. The sum in (17) defining 1;, f can now be written in the form
1
(21) (1;,f)(7:) = - L ak f(A7:),
n A
where A runs through a complete set of nonequivalent elements in r(n) of
the form described in Theorem 6.8. The coefficient a k is the kth power of
the entry in the first row and first column of A.

Theorem 6.9. If Al E r(n) and VI E r, then there exist matrices A 2 in r(n)


and V2 in r such that
(22)
Moreover, if

Ai = (~ ~:) and

for i = 1, 2, then we have


(23) al(Y2A27: + £5 2 ) = a2(YI7: + £5 1 ).
PROOF. Since det(A I VI) = det Al det VI = n, the matrix Al VI is in r(n) so,
by Theorem 6.7, there exists A 2 in r(n) and V2 in r such that (22) holds. To
verify (23) we first note that A 1 VI has the form .

and that

Equating entries in the second row we find

d 1 d 2 1'1 d2
Y2 = - - = -Yl
n al
and

124
6.9: Behavior of ~f under the modular group

since at d t = n. Hence
atY2 = d 2Yt and
and we obtain
at(Y2 A 2 r + £5 2 ) = at Y2 A 2 r + at£5 2

which proves (23). D

6.9 Behavior of Tn/under the modular group

Theorem 6.10. If f E M k and V = (~ ~) E r then


(24) (1;,f)(Vr) = (yr + £5)k(1;,f)(r).
PROOF. We use the representation in (21) to write
1
(T,. f)(r) =-
n
L at
Al
k f(A t r)
.

where At = ( at0 dht) and At runs through a complete set ofnoneqUlvalent


.
t
elements in r(n). Replacing r by Vr we find
1
(25) (T,.f)(Vr) = - L atkf(A t Vr).
nAt

By Theorems 6.7 and 6.9, there exist matrices

A2 = (~ ~:) in r(n) and V2 = G: ~:) in r


such that
and
Therefore
atkf(A t Vr) = at kf(V2A 2r) = atk(Y2A2 r + £5 2)kf(A 2 r)
= a2 k(yr + £5)k f(A 2 r)
since f E M k • Now as At runs through a complete set of nonequivalent
elements of r(n) so does A 2 • Hence (25) becomes
1
(T,.f)(Vr) = - (yr + £5)k L a2 k f(A z r) = (yr + £5)k(T,.f)(r). D
n A2

125
6: Modular forms with multiplicative coefficients

The next theorem shows that each Hecke operator 1;. maps M k into M k
and also maps Mk,o into Mk,o.

Theorem 6.11. If f E M k then 1;.f E M k. Moreover, if f is a cusp form then


1;. f is also a cusp form.
PROOF. If f E M k the definition of 1;. shows that T"f is analytic everywhere
in H. Theorem 6.6 shows that 1;.f has a Fourier expansion of the required
form and that 1;.f is analytic at ioo. And Theorem 6.10 shows that 1;.f has
the proper behavior under transformations of r. Finally, if f is a cusp form,
the Fourier expansion in Theorem 6.6 shows that T" f is also a cusp form. D

6.10 Multiplieative property of Heeke


operators
This section shows that any two Hecke operators 1;. and Tm defined on M k
commute with each other. This follows from a multiplicative property of the
composition Tm 1;.. First we treat the case in which m and n are relatively
prime.

Theorem 6.12. If (m, n) = 1 we have the composition property


(26)
PROOF. If f E M k we have
1
(1;.f)(r) = - L a k f(Ar),
n a~ 1,ad=n
O~b<d

where A = (~ :). Applying Tm to each member we have

1 1
{Tm(1;.(f))}(r)=- L rt k - L akf(BAr),
m IX ~ 1 , IXlJ = m n a ~ 1, ad = n
O~P<lJ O~b<d

where B = (~ ~} This can be written as


1
(27) {(Tm T,,)(f)} (r) =- L L (rta)k f(Cr),
mn IX ~ 1 , IXlJ = m a ~ 1, ad = n
O~P<lJ O~b<d

where

C = BA = (~ ~)(~ ~) = (~a ~b ~Pd}


As d and b run through the positive divisors of nand m, respectively, the
product db runs through the positive divisors of mn since (m, n) = 1. The

126
6.10: Multiplicative property of Hecke operators

linear combination rxb + f3d runs through a complete residue system


mod db as band f3 run through complete residue systems mod d and b,
respectively. Therefore the matrix C runs through a complete set of non-
equivalent elements of r(mn) and we see that (27) implies (26). D

The next theorem extends the composition property in (26) to arbitrary


m and n. For convenience in notation we write T(n) in place of 1;,.

Theorem 6.13. Any two Hecke operators T(n) and T(m) defined on M k commute
with each other. Moreover, we have the composition formula
(28) T(m)T(n) = L dk- 1 T(mnjd 2 ).
dl(m,n)

PROOF. Commutativity follows from (28) since the right member is symmetric
in m and n. If (m, n) = 1 formula (28) reduces to (26). Therefore, to prove
(28) it suffices to treat the case when m and n are powers of the same prime p.
First we consider the case m = p and n = pr, where r ~ 1. In this case we
are to prove that
(29)
We use the representation in (17) and note that the divisors of pr have the
form pt where 0 ~ t ~ r. Hence we have

(30) {T(pr)f}(r) = p-r L p(r-t)kf(pr-trt+ ht).


O:::;t::;r P
O:::;bt<pt
By (14) we have

{T(p)g}(r)=pk-l g(pr)+p-l p-l +


Lg - - ,
(r b)
b=O P
so when we apply T(P) to each member of (30) we find

{T(p)T(pr)f} (r) = pk-l-r L p(r-t)k f(pr+ I-t~ + Pht)


O:::;t:::;r P
O:::;bt<pt

+ P- 1 - r " (r - t)k p-l


" f (r-t
p r + bt + bPt) .
i...J P i...J t+l
O:::;t::;r b=O P
O:::;bt<pt
In the second sum the linear combination bt + bpt runs through a complete
residue system mod pt+ 1. Since r - t = (r + 1) - (t + 1) the second sum,
together with the term t = 0 from the first sum, is equal to {T(pr+ 1 )f} (r). In
the remaining terms we cancel a factor p in the argument of f, then transfer
the factor pk to each summand to obtain

{T(p)T(pr)f} (r) = {T(pr+ 1 )f} (r) + p-l-r L p(r+ I-t)k f(pr-t~_~ ht).
1 :::;t:::;r P
O:::;bt<pt

127
6: Modular forms with multiplicative coefficients

Dividing each bt by pt- 1 we can write


bt = qtpt-1 + rn
where 0 ~ rt < pt-1 and qt runs through a complete residue system mod p.
Since f is periodic with period 1 we have
pr-t r + bt) = (pr-t r + rt)
f( P
t- 1 f P
t- 1 '

so as qt runs through a complete residue system mod p each term is repeated


p times. Replacing the index t by t - 1 we see that the last sum is pk - 1 times
the sum defining {T(pr - 1) f} (r). This proves (29).
Now we consider general powers of the same prime, say m = pS and
n = pro Without loss of generality we can assume that r ~ S. We will use in-
duction on r to prove that

(31)

for all r and all s ~ r. When r = 1, (31) follows for all s ~ 1 from (29).
Therefore we assume that (31) holds for r and all smaller powers and all
s ~ r, and prove it also holds for r + 1 and all s ~ r + 1.
By (29) we have
T(P) T(p') T(pS) = T(pr + 1) T(pS) + pk - 1T(pr - 1) T(pS),
and by the induction hypothesis we have
r
T(p)T(pr)T(ps) = L pt(k-l)T(p)T(pr+s-2t).
t=O
Equating the two expressions, solving for T(pr+ 1 )T(pS) and using (29) in the
sum on t we find
r ,
T(pr+ 1 )T(pS) = L pt(k-1)T(pr+s+ 1- 2t) + L p(t+ 1)(k-1)T(p,+s-1-2t)
t=O t=O

By the induction hypothesis the last term cancels the second sum over t
except for the term with t = r. Therefore
r
T(pr+1)T(pS) = Lpt(k-1)T(pr+s+1-2t) + p(r+1)(k-1)T(ps-1-r)
t=O
r+ 1
-= L pt(k-1)T(pr+1+s-2t).
t=O
This proves (31) by induction for all r and all s ~ r, and also completes the
proof of (28). D

128
6.11: Eigenfunctions of Hecke operators

6.11 Eigenfunctions of Hecke operators


In Theorem 6.6 we proved that if f E M k and has the Fourier expansion
00

(32) f(1:) = L c(m)x m ,


m=O

where x = e 21tit , then 1;,f has the Fourier expansion


00

(33) (1;,f)(r) = L yn(m)x m ,


m=O

where

(34) Yn(m) = ~
i..J dk-l C(mn)
d2 •
dl(n,m)

When m = 0 we have (n, 0) = n so the constant terms of f and 1;,f are


related by the equation
(35) Yn(O) = L dk-1C(0) = (Tk-l(n)C(O)
din

for all n 2 1. Similarly, when m = 1 we find


(36) Yn(l) = c(n)
for all n 2 1.
The sum on the right of (34) resembles that which occurs in the multi-
plicative property of Ramanujan's function 1:(n) and the divisor functions
(Ja(n). These examples suggest we seek those formsffor which the transformed
function 1;, f has Fourier coefficients
(37) Yn(m) = c(n)c(m)
since this would imply the multiplicative property

c(n)c(m) = L dk - lC(;~).
dl(n,m)

The relation (37) is equivalent to the identity


T,. f = c(n) (
for all n 2 1. A nonzero function f satisfying a relation of the form
(38) 1;,f = A(n)f
for some complex scalar A(n) is called an eigenfunction (or eigenform) of the
operator Tn' and the scalar A(n) is called an eigenvalue of 1;,. If f is an
eigenform so is cf for every c # o.

129
6: Modular forms with multiplicative coefficients

EXAMPLES. If a linear operator T maps a I-dimensional function space V


into itself, then every nonzero function in Vis an eigenfunction of T. Formula
(9) shows that
dim M k = 1 if k = 4, 6, 8, 10 and 14,
so each Hecke operator T" has eigenforms in M k for each of these values of k.
For example, the respective Eisenstein series G4 , G6 , Gs , G IO and G I4 are
eigenforms for each T".
Similarly, formula (11) implies that
dim Mk,o = 1 if k = 12,16,18,20,22 and 26,
so each T" has eigenforms in M k, 0 for each of these values of k. The respective
cusp forms t1, t1G 4 , t1G 6 , t1G s , t1G IO and t1G 14 are eigenforms for each T".

If f is an eigenform for every Hecke operator T", n ~ 1, then f is called a


simultaneous eigenform. All the examples just mentioned are simultaneous
eigenforms.

6.12 Properties of simultaneous eigenforms


Theorem 6.14. Assume k is even, k ~ 4. If the space M k contains a simultaneous
eigenform f with Fourier expansion (32), then c(l) # O.

PROOF. The coefficient of x in the Fourier expansion of T"f is Yn(l) = c(n).


Since f is a simultaneous eigenform this coefficient is also equal to A(n)c(l), so
c(n) = A(n)c(l)
for all n ~ 1. If c(l) = 0 then c(n) = 0 for all n ~ 1 and f(1:) = c(O). But then
c(O) = 0 since k ~ 4, hence f = 0, contradicting the definition of eigenform.
This proves that c(l) # O. D

An eigenform with c(l) = 1 is said to normalized. If M k contains a simul-


taneous eigenform then it also contains a normalized eigenform since we
can always make c(l) = 1 by multiplying f by a suitable nonzero constant.
It is easy to characterize all cusp forms which are simultaneous eigenforms.
Since the zero function is the only cusp form of weight < 12 we need consider
only k ~ 12.

Theorem 6.15. Assume f E M k, 0 where k is even, k ~ 12. Then f is a simul-


taneous normalized eigenform if, and only if, the coe.fficients in the Fourier
expansion (32) sati~fy the multiplicative property

(39) c(m)c(n) = L dk - I c(;~)


dl<n,m)

for all m ~ 1, n ~ 1, in which case the coefficient c(n) is an eigenvalue ofT".

130
6.13: Examples of normalized simultaneous eigenforms

PROOF. The equation T" f == A(n).f is equivalent to the relation


(40) Yn(m) == A(n)c(m)
obtained by equating coefficients of x m in the corresponding Fourier expan-
sions. Since f is a cusp form so is T" f hence (40) is to hold for all m ~ 1
and n ~ 1. Now Yn(1) == c(n) so (40) implies A(n) == c(n) if c(1) == 1, and
hence Yn(m) == c(n)c(m). On the other hand, Equation (34) shows that (40)
is equivalent to (39) if c( 1) == 1. Therefore f is a normalized simultaneous
eigenform if, and only if, (39) holds for all m ~ 1, n ~ 1. D

6.13 Examples of normalized simultaneous


eigenforms
The discriminant L\ is a cusp form with Fourier expansion
00

L\(r) == (2n)12 L r(m)x m


m=1

where r(1) == 1. Therefore (2n)- 12 L\(r) is a normalized eigenform for each


T" with corresponding eigenvalue r(n). This also proves that Ramanujan's
function r(n) satisfies the multiplicative property in (3).
The next theorem shows that the only simultaneous eigenforms in M 2k
which are not cusp forms are constant multiples of the Eisenstein series G2k .

Theorem 6.16. Assume that.1' E M 2k, where k 2 2, and that f is not a cusp
form. Then f is a normalized simultaneous eigeYl:form (f, and only if,
(2k - I)!
(41) f(r) = 2(2ni)2k G2k(r).

PROOF. In the Fourier expansion (32) we have c(O) # 0 since f is not a cusp
form. The relation
(42) T" f == A(n)f
is equivalent to the relation
(43) Yn(m) == A(n)c(m)
obtained by equating coefficients of x m in the corresponding Fourier expan-
sions. When m == 0 this becomes

Yn(O) == A(n)c(O).
On the other hand, (35) implies Yn(O) = 0"2k-1(n)c(0) since f EM 2k. But
c(O) # 0, so Equation (42) holds if, and only if,
A(n) = 0"2k-1(n).

131
6: Modular forms with multiplicative coefficients

Using this in (43) we find that


Yn(m) = (J 2k - 1 (n)c(m).
When m = 1 this relation, together with (36), gives us
c(n) = (J 2k - 1 (n)c(I).
Therefore,! is a normalized simultaneous eigenform in M 2k if, and only if,
c(n) = (J 2k - 1 (n)
for all n ~ 1. Since the Eisenstein series G 2k has the Fourier expansion
2(2ni)2k 00
G2k(r) = 2((2k) + (2k _ I)! m~l 0"2k_l(m)x ,
m

the function in (41) is normalized and its Fourier expansion is given by


(2k - I)! 00
(44) f(r) = (2 yk ((2k)
nl
+ L 0"2k- l(m)x m.
m= 1
D

Note. Since
k+ 1 (2n)2k
((2k) = (-1) 2(2k)! B2k

where B k is the kth Bernoulli number defined by

x ~ Bk k
~1 = i..J -k'x,
e k=O •

the constant term in (44) is equal to -B 2k /(4k). (See [4J, Theorem 12.17.)
We can also write

Since the eigenvalue 2(n) in (42) is (J 2k _ 1 (n), Theorem 6.16 shows that the
divisor functions (Ja(n) satisfy the multiplicative property in Equation (4)
when (X = 2k - 1. Actually, they satisfy (4) for all real or complex (x, but
O'a(n) is the nth coefficient of an entire form only when (X is an odd integer ~ 3.

EXAMPLES. The problem of determining all entire noncusp forms whose


coefficients satisfy the multiplicative property (39) has been completely
settled by Theorem 6.16. For the cusp forms the problem has been reduced
by Theorem 6.15 to that of determining simultaneous normalized eigenforms
of even weight 2k ~ 12. We have already noted that the function (2n) -12 ~(r)
is the only simultaneous normalized eigenform of weight 2k = 12. Also
there is exactly one simultaneous normalized eigenform for each of the
weights
2k = 16, 18, 20, 22, and 26

132
6.14: Remarks on existence of simultaneous eigenforms in M 2k,O

since dim M 2k, 0 = 1 for these weights. The corresponding normalized


eigenforms are given by

(2n)- 1 2 L\( ). G 2 k - 12 (r ) ,00, ( ) n{ 1 2(2k - 12) " 00


( ) m}
r 2((2k-12) n~1rnx - B 2k - 12 m~1(J2k-13mx .

We define r(O) = 0 and (J2k-1(0) = - B 2k /(4k). Then the coefficients c(n) of


these eigenforms are given by the Cauchy product
4k - 24 n
c(n) = - B L r(m)0"2k-13(n - m).
2k-12 m=O
They satisfy the multiplicative property

c(m)c(n) = L d2k-1C(~:)
dl(m,n)

for all m ~ 1, n ~ 1.

6.14 Remarks on existence of simultaneous


eigenforms in M 2k,O
Let K = dim M 2k, 0 where 2k ~ 12. Then we' have

2k
[ 12
J- 1 if 2k == 2 (mod 12)
K=
2k
[ 12
J if 2k ¢ 2 (mod 12).

Let e(k) denote the number of linearly independent simultaneous eigenforms


in M 2k ,O. Clearly, e(k) ~ K. We have shown that e(k) = 1 when K = 1.
Hecke showed that e(k) = 2 when K = 2, and later Petersson [32] showed
that e(k) = K in all cases. He did this by introducing an inner product
(f, g) in M 2k, 0 defined by the double integral

(J, g) = ff j(r)g(r)v 2k - 2 du dv
Rr

extended over the fundamental region R r in the r = u + iv plane. Relative


to the Petersson inner product the Hecke operators are Hermitian, that is,
they satisfy the relation
(T" f, g) = (f, T"g)
for any two cusp forms in M 2k, o. Therefore, by a well known theorem of
linear algebra (see [2], Theorem 5.4) for each T" there exist K eigenforms
which form an orthonormal basis for M 2k, o. These need not be simultaneous
eigenforms for all the T,.. However, since the T,. commute with each other,
another theorem of linear algebra (see [10], Ch. IX, Sec. 15) shows that

133
6: Modular forms with multiplicative coefficients

M 2k, 0 has an orthonormal basis consisting of K simultaneous eigenforms.


Each of these can be multiplied by a constant factor to get a new basis of
simultaneous normalized eigenforms. (The new basis will be orthogonal
but need not be orthonormal.) Since the 1;. are Hermitian, the corresponding
eigenvalues are real. Details of the proofs of these statements can be found
in references [32], [26], or [11].

6.15 Estimates for the Fourier coefficients of


entire forms
Assume f is an entire form with Fourier expansion
00

(45) f(r) = L c(n)x n


,
n=O

where x = e 21tit . Write r = u + iv so that x = e-21tve21tiu. For fixed v > 0,


as u varies from 0 to 1 the point x traces out a circle C(v) of radius e - 21tv
with center at x = O. By Cauchy's residue theorem we have

(46) 1
c(n) = -2.
nl
I
C(v)
f(r) dx =
---,,-:tT
x
II 0
f(u + lV)X,
·-n
duo

We shall use this integral representation to estimate the order of magnitude


of Ic(n) I. First we consider cusp forms of weight 2k.

Theorem 6.17. Iff E M 2k ,O we have


c(n) = O(n k ).

PROOF. The series in (45) converges absolutely if Ix I < 1. Since c(O) = 0 we


can remove a factor x and write

II(r)1 = IXI!J1c(n)xn-11 ~ IxIJllc(n)lIxln-l.


If r is in R r , the fundamental region of r, then r = u + iv with v ~ fi/2
> 1/2, so Ixl = e- 21tv < e- 1t . Hence
If(r)1 ~ A Ixl = Ae- 21tv
where
00

A = L Ic(n)le-(n-l)1t.
n=1

This implies
(47)
Now define
g( r) = -t Ir - r I = v
134
6.15: Estimates for the Fourier coefficients of entire forms

ifrEH. Then
g(Ar) = ler + dl- 2g(r)

if A = (: ~)E r, so gk(Ar) = Icr + dl- 2kl(r). Therefore the product

qJ( r) = I f (r ) Igk(r) = I f (r) Ivk


is invariant under the transformations of r. Moreover, qJ is continuous
in R r , and (47) shows that qJ(r) ---+ 0 as v ---+ + 00. Therefore qJ is bounded
in R r and, since qJ is invariant under r, qJ is also bounded in H, say
IqJ(r) I ::; M

for all r in H. Therefore


If(r)1 ::; Mv- k
for all r in H. Using this in (46) we find

Ic(n) I :0:;; f If(u + iv)x- n Idu :0:;; Mv-klxl- n = Mv-ke-21Cnv.

This holds for all v > O. When v = lin it gives us


Ic(n)1 s Mn ke- 21t = O(n k).; o
Theorem 6.18. If f E M 2k and f is not a cusp form, then
(48) c(n) = O(n 2k - 1).

PROOF. If f = G 2k each coefficient c(n) is of the form CllT 2k -l(n) where Cl is


independent of n. Hence

Now

lT2k-l(n) = dLln(~)2k-l = n 2k - 1 L:-d2-~---1 :0:;; n


2k - 1 f -d2-~---1 = O(n
2k
-
1
),
din d= 1

so (48) holds if f = G2k ·


For a general noncusp form in M 2k, let 2 = f(ioo )/G 2k (ioo). Thenf - 2G 2k
is a cusp form so
f = 2G 2k +g
where gEM 2k, o. The Fourier coefficients of f are the sum of those of 2G 2k
and g so they have order of magnitude
o
135
6: Modular forms with multiplicative coefficients

Note. For cusp forms, better estimates for the order of magnitude of the
c(n) have been obtained by Kloosterman, Salie, Davenport, Rankin, and
Selberg (see [46]). It has been shown that
c(n) = O(n k -(1/4)+C)

for every 8 > 0, and it has been conjectured that the exponent can be further
improved to k - ! + 8. For the discriminant ~, Ramanujan conjectured the
sharper estimate
Ir(p) I :::; 2p ll/2

for primes p. This was recently proved by P. Deligne [7].

6.16 Modular forms and Dirichlet series


Hecke found a remarkable connection between each modular form with
Fourier series
00

(49) f(r) = c(O) + L c(n)e21tint


n= 1

and the Dirichlet series

(50) qJ(S) = f
n= 1
C(7)
n
formed with the same coefficients (except for c(O)). If f EM 2k then c(n) =
O(n k) iffis a cusp form, and c(n) = O(n 2k - 1 ) iffis not a cusp form. Therefore,
the Dirichlet series in (50) converges absolutely for (J = Re(s) > k + 1 if f is
a cusp form, and for (J > 2k if f is not a cusp form.

Theorem 6.19. If the coefficients c(n) sati~fy the multiplicative property

(51) c(m)c(n) = L d2k-lC(~~)


dl<m,n)

the Dirichlet series will have an Euler product representation of the form

1
(52) qJ(s) = n 1 -cpp
p
( ) -5 + P2k-l p -25'

absolutely convergent with the Dirichlet series.

PROOF. Since the coefficients are multiplicative we have (see [4], Theorem
11.7)

(53)

136
6.16: Modular forms and Dirichlet series

whenever the Dirichlet series converges absolutely. Now (51) implies


c(p)c(pn) = c(pn+ 1) + p2k-1 c(pn-1)
for each prime p. Using this it is easy to verify the power series identity

(1- c(P)x + p2k-l X2)(1 +J/(pn)x n) = 1

for all Ix I < 1. Taking x = p - s, we find that (53) reduces to (52). 0

EXAMPLE. For the Ramanujan function we have the Euler product represen-
tation
00 r(n) 1
Jl -;:;s = I] 1 - r(p)p-S + pll-2s

for (J > 7 since r(n) = O(n 6 ).

Hecke also deduced the following analytic properties of cp(s).

Theorem 6.20. Let <p(s) be the function defined for (J > k by the Dirichlet
series (50) associated with a modular form f(r) in M k having the Fourier
series (49), where k is an even integer ~ 4. Then <p(s) can be continued
analytically beyond the line (J = k with the following properties:
(a) If c(O) = 0, <p(s) is an entire function of s.
(b) If c(O) i= 0, <p(s) is analytic for all s except for a simple pole at s = k '
with residue
( -1)k/2 c(0)(2n)k
r(k)

(c) The ,function <p satisfies the functional equation

PROOF. From the integral representation for r(s) we have

if (J > O. Therefore if (J > k we can multiply both members by c(n) and sum
on n to obtain

(2n)-sr(s)q>(s) = fO {f(iy) - c(O)}yS-l dy.

137
6: Modular forms with multiplicative coefficients

Since f is a modular form in M k we have f(ijy) = (iy)k f(iy) so

(2n)-sr(s)qJ(s) = {>J{f(iY) - c(O)}yS-l dy + f{(iy)-kfG) - c(O)}yS-l dy

= foo {f(iy) - c(O)}yS-l dy + i- k foo f(iw)W k- s - 1 dw _ c(O)


l I S

= lOO {f(iy) - c(O)}yS-l dy

+ (-It/2 lOO {f(iw) - c(O)}W k- S- 1 dw

+ (-1)k/2 C(0) fOO Wk- s - 1 dw _ c(O)


1 S

= fOO {f(iy) - c(O)}(yS + (-1)k/2l- S) dy


1 y
1 (_I)k/2)
- c(O) ( ~+ ~.

Although this last relation was proved under the assumption that (J > k, the
right member is meaningful for all complex s. This gives the analytic continua-
tion of qJ(s) beyond the line (J = k and also verifies (a) and (b). Moreover,
replacing s by k - s leaves the right member unchanged except for a factor
(_I)k/2 so we also obtain (c). D

Hecke also proved a converse to Theorem 6.20 to the effect that every
Dirichlet series qJ which satisfies a functional equation of the type in (c),
together with some analytic and growth conditions, necessarily arises from
a modular form in M k • For details, see [15].

Exercises for Chapter 6


Exercises 1 through 6 deal with arithmetical functions f satisfying a relation
of the form

(54) f(m)f(n) = L rx(d)f(~~)


dl(m,n)

for all positive integers m and n, where rx is a given completely multiplicative


function (that is, rx(l) = 1 and rx(mn) = rx(m)rx(n) for all m and n). An arith-
metical function satisfying (54) will be called rx-multiplicative. We write
f = 0 if f(n) = 0 for all n.
1. Assume f is ~-multiplicative and f =1= o. Prove that f( 1) = 1. Also prove that cf is
a-multiplicative if, and only if, c = 0 or c = 1.

138
Exercises for Chapter 6

2. If f and g are ex-multiplicative, prove that f +g is ex-multiplicative if, and only if,
f = 0 or g = O.
3. Let II' ... , fk be k distinct nonzero ex-multiplicative functions. If a linear combination
k

f = I Cih
i=1
is also ex-multiplicative, prove that:
(a) The functions II' ... , Ik are linearly independent.
(b) Either all the Ci are 0 or else exactly one of the Ci is 1 and the others are O. Hence
either f = 0 or f = h for some i. In other words, linear combinations of ex-
multiplicative functions are never ex-multiplicative except for trivial cases.

4. If I is ex-multiplicative, prove that

a(n)f(m) = I Jl(d)f(mnd)f(!!-).
din d
5. If f is multiplicative, prove that f is ex-multiplicative if, and only if,

(55)
for all primes p and all integers k ~ 1.
6. The recursion relation (55) shows that f(pn) is a polynomial in f(p), say
f(pn) = Qn(f(P)).
The sequence {Qn(x)} is determined by the relations
Z
Ql(X) = x, Q2(X) = X - ex(p), Qr+ l(X) = xQr(x) - ex(p)Qr-1(X) for r 2: 2.
Show that
Qi2ex(p)I/Z X) = a(p)"/z Un(x),
where U n(x) is the Chebyshev polynomial of the second kind, defined by the relations
U 1(x) = 2x, Uz(x) = 4x z - 1, U r+1(x) = 2xU r(x) - U r- 1(x) for r 2: 1.

7. Let £Zk(r) = tG zk (r)/((2k). If x = eZ1tit verify that the Fourier expansion of £Zk(r)
has the following form for k = 2, 3, 4, 5, 6, and 7:
00

£4(r) = 1 + 240 I0"3(n)x n,


n=l
00

£6(r) = 1 - 504 IO"s(n)x n,


n= 1
00

£8(r) = 1 + 480 I0"7(n)x n,


n= 1
00

£10(r) = 1 - 264 I0"9(n)x n,


n=1

00

£14(r) = 1 - 24 IO"I3(n)x n.
n=1

139
6: Modular forms with multiplicative coefficients

Derive each of the identities in Exercises 8, 9, and 10 by equating coefficients


in appropriate identities involving modular forms.
n-1
8. 0"7(n) = 0"3(n) + 120 I 0"3(m)0"3(n - m).
m=l
n-1
9. 110"9(n) = 210"5(n) - 100"3(n) + 5040 I 0"3(m)0"5(n - m).
m=1

65 691 691 n-1


10. r(n) = -0"11(n) + -0"5(n)
756 756
- -
3 m= 1
I 0"5(m)0"5(n - m).

Show that this identity implies Ramanujan's congruence

r(n) == 0"11(11) (mod 691).


11. Prove that the products Gk-12r~r which occur in Theorem 6.3 are linearly inde-
pendent.

12. Prove that the products G4 G6 b are linearly independent, where a and b are non-
Q

negative integers such that 4a + 6b = k.

13. Show that the Dirichlet series associated with the normalized modular form

(2k - 1)! Y(2k ~ ( ) 21timt


f (r) = . 2k S ) + ~ 0" 2k - 1 me
(2nl) m= 1

is q>(s) = ((s)((s + 1 - 2k).

14. A quadratic polynomial 1 - Ax + Bx 2 with real coefficients A and B can be factored


as follows:
1 - Ax + Bx 2 = (1 - r 1 x)(1 - r2x).
Prove that '1
= 'Y. + if1 and '2
= }' - ip, where rJ., {3, I are real and {3()' - rJ.) = O.
Hence, if {3 # 0 the numbers, 1 and r2 are complex conjugates.

Note. For the quadratic polynomial occurring in the proof of Theorem


6.19 we have

where

and

Petersson conjectured that '1 and '2 are always complex conjugates. This
implies

and

When c(n) = r(n) this is the Ramanujan conjecture. The Petersson conjecture
was proved recently by Deligne [7].

140
Exercises for Chapter 6

15. This exercise outlines Riemann's derivation of the functional equation

(56) S
n- /
2rG}(s) = n(S-I)/ 2rC~ s)W - s)

from the functional equation (see Exercise 4.1)

(57) .9( ~ 1) = (_ ir)I/2.9(r)


satisfied by Jacobi's theta function
00

:J(r) = 1 + 2 L errin2t.
n=l
(a) If (J > 1 prove that

(S)2 2
n -S/2 r -n -s = foo e -rrn x x s/2- 1 dx
0

and use this to derive the representation

n- S / 2 rG)c(s) = faro r/J(x)x s2 1


/ - dx,

where 21/J(x) = 9(x)-1.


(b) Use (a) and (57) to obtain the representation

n-S/2r(:)((s) = __1_ + joo(XS/2-1 + X(1-s)/2-1)I/J(X) dx


2 s(s - 1) 1

for (J > 1.
(c) Show that the equation in (b) gives the analytic continuation of ((s) beyond the
line (J = 1 and that it also implies the functional equation (56).

141
Kronecker's theorem
with applications

7.1 Approximating real numbers by rational


numbers
Every irrational number fJ can be approximated to any desired degree of
accuracy by rational numbers. In fact, if we truncate the decimal expansion
of fJ after n decimal places we obtain a rational number which differs from
fJ by less than 10- n • However, the truncated decimals might have very large
denominators. For example, if
fJ = n - 3 = 0.141592653 ...
the first five decimal approximations are 0.1, 0.14, 0.141, 0.1415, 0.14159.
Written in the form alb, where a and b are relatively prime integers, these
rational approximations become
1 7 141 283 14159
10' 50' 1000' 2000' 100,000·
On the other hand, the fraction 1/7 = 0.142857 ... differs from fJ by less than
2/1000 and is nearly as good as 141/1000 for approximating fJ, yet its denomi-
nator 7 is very small compared to 1000.
This example suggests the following type of question: Given a real
number fJ, is there a rational number h/k which is a good approximation to
fJ but whose denominator k is not too large?
This is, of course, a vague question because the terms" good approxima-
tion" and "not too large" are vague. Before we make the question more
precise we formulate it in a slightly different way. If fJ - h/k is small, then
(kfJ - h)/k is small. For this to be small without k being large the numerator
kfJ - h should be small. Therefore, we can ask the following question:

142
7.2: Dirichlet's approximation theorem

Given a real number fJ and given e > 0, are there integers hand k such that
IkfJ - hi < e?
The following theorem of Dirichlet answers this question in the affirma-
tive.

7.2 Dirichlet's approximation theorem


Theorem 7.1. Given any real fJ and any positive integer N, there exist integers
hand k with 0 < k ~ N such that
1
(1) IkfJ - hi <-

PROOF. Let {x} = x - [x] denote the fractional part of x. Consider the
N + 1 real numbers
0, {fJ}, {2fJ}, ... , {NfJ}.
All these numbers lie in the half open unit interval 0 ~ {mfJ} < 1. Now
divide the unit interval into N equal half-open subintervals of length liN.
Then some subinterval must contain at least two of these fractional parts,
say {afJ} and {bfJ}, where 0 ~ a < b ~ N. Hence we can write
1
(2) I{bfJ} - {afJ} I < N·

But
{bfJ} - {afJ} = bfJ - [bfJ] - afJ + [afJ] = (b - a)fJ - ([bfJ] - [afJ]).
Therefore if we let
k=b-a and h = [bfJ] - [afJ]
inequality (2) becomes
1
IkfJ - hi < N' with 0 < k ~ N.

This proves the theorem. o


Note. Given e > 0 we can choose N > lie and (1) implies IkfJ - hi < E.

The next theorem shows that we can choose hand k to be relatively


prime.

Theorem 7.2. Given any real fJ and any positive integer N, there exist relatively
prime integers hand k with 0 < k ~ N such that
1
IkfJ - hi < N·

143
7: Kronecker's theorem with applications

PROOF. By Theorem 7.1 there is a pair h', k' with 0 < k' ~ N satisfying

h'l
IfJ - k
(3) 1
< Nk'·

Let d = (h', k'). If d = 1 there is nothing to prove. If d > 1 write h' = hd,
k' = kd, where (h, k) = 1 and k < k' ~ N. Then 11k' < 11k and (3) becomes

Ie - ~I < ~k' < ~k'


from which we find IkfJ - hi < liN. 0

Now we restate the result in a slightly weaker form which does not involve
the integer N.

Theorem 7.3. For every real fJ there exist integers hand k with k > 0 and
(h, k) = 1 such that

PROOF. In Theorem 7.2 we have 1/(Nk) ~ 1/k 2 because k ~ N. D

Theorem 7.4. If fJ is real, let S( fJ) denote the set of all ordered pairs of integers
(h, k) with k > 0 and (h, k) = 1 such that

Ie - ~I < :2'
Then S( fJ) has the following properties:
(a) S(fJ) is nonempty.
(b) If fJ is irrational, S(fJ) is an infinite set.
(c) When S(fJ) is infinite it contains pairs (h, k) with k arbitrarily large.
(d) If fJ is rational, S(fJ) is a finite set.

PROOF. Part (a) is merely a restatement of Theorem 7.3. To prove (b), assume
() is irrational and assume also that S(fJ) is finite. We shall obtain a contra-
diction. Let

.
a = mIn
(h, k) E 5(6)
IfJ - hi
- .
k
Since fJ is irrational, a is positive. Choose any integer N > 11a, for example,
N = 1 + [l/a]. Then liN < a. Applying Theorem 7.2 with this N we obtain
a pair of integers hand k with (h, k) = 1 and 0 < k ~ N such that

IfJ - ~Ik <_1


kN·

144
7.2: Dirichlet's approximation theorem

Now l/(kN) ~ 1/k 2 so the pair (h, k) E S(fJ). But we also have
1 1
-<-<~ so
kN- N '
contradicting the definition of~. This shows that S( fJ) cannot be finite if fJ is
irrational.
To prove (c) assume that all pairs (h, k) in S(fJ) have k ~ M for some M.
We will show that this leads to a contradiction by showing that the number
of choices for h is also bounded. If (h, k) E S(fJ) we have
1
IkfJ - hi < k~ 1,

so
Ihi = Ih - kfJ + kfJ I ~ Ih - kfJ I + IkfJ I < 1 + IkfJ I ~ 1 + M IfJ I·
Therefore the number of choices for h is bounded, contradicting the fact that
S( fJ) is infinite.
To prove (d), assume fJ is rational, say fJ = alb, where (a, b) = 1 and b > o.
Then the pair (a, b) E S(fJ) because fJ - alb = O. Now we assume that S(fJ)
is an infinite set and obtain a contradiction. If S(fJ) is infinite then by part (c)
there is a pair (h, k) in S(fJ) with k > b. For this pair we have

o < I~b -~I ~


k < k2 '
from which we find 0 < Iak - bh I < b/k < 1. This is a contradiction because
ak - bh is an integer. 0

Theorem 7.4 shows that a real number fJ is irrational if, and only if,
there are infinitely many rational numbers h/k with (h, k) = 1 and k > 0
such that

Ie - ~I < :2'
This inequality can be improved. It is easy to show that the numerator 1
can be replaced by ! (see Exercise 7.4). Hurwitz replaced by a smaller t
constant. He proved that fJ is irrational if, and only if, there exist infinitely
many rational numbers h/k with (h, k) = 1 and k > 0 such that

Ie - ~I < flk 2'

Moreover, the result is false if 1/.)5 is replaced by any smaller constant. (See
Exercise 7.5.) We shall not prove Hurwitz's theorem. Instead, we prove a
theorem of Liouville which shows that the denominator k 2 cannot be re-
placed by k 3 or any higher power.

145
7: Kronecker's theorem with applications

7.3 Liouville's approximation theorem


Theorem 7.5. Let fJ be a real algebraic number ofdegree n ~ 2. Then there is a
positive constant C(fJ), depending only on fJ, such that for all integers hand
k with k > 0 we have

(4)
IfJ - ~I > k
C(fJ)
kn '

PROOF. Since fJ is algebraic of degree n, fJ is a zero of some polynomial f(x)


of degree n with integer coefficients, say
n

f(x) = L arxr,
r=O

where f(x) is irreducible over the rational field. Since f(x) is irreducible it
has no rational roots so f(hlk) =1= 0 for every rational hlk.
Now we use the mean value theorem of differential calculus to write

(5) f(~) = f(~) - f(O) = f'(~)(~ - 0)'


where ~ lies between fJ and hlk. We will deduce (4) from (5) by getting an upper
bound for If'(~)1 and a lower bound for If(hlk)l. We have

f (k
h) = r
n (h)r
~oQ r k =
N
k"
where N is a nonzero integer. Therefore

(6)

which is the required lower bound. To get an upper bound for If'(~) I we let

If d > 1 then (4) holds with C(fJ) = 1, so we can assume that d < 1. (We
cannot have d = 1 since fJ is irrational.) Since ~ lies between fJ and hlk and
d < 1 we have I~ - fJ I < 1 so
I~I = IfJ + ~ - fJl ~ IfJl + I~ - fJl < IfJl + 1.
Hence
If'(~) I ~ A(fJ) < 1 + A(fJ),
where A(fJ) denote the maximum value of I f'(x) I in the interval Ix I ~ IfJ I + 1.
Using this upper bound for If'(~) I in (5) together with the lower bound in (6)
we obtain (4) with C(fJ) = 1/(1 + A(fJ)). D
146
7.3: Liouville's approximation theorem

A real number which is not algebraic is called transcendental. A simple


counting argument shows that transcendental numbers exist. In fact, the
set of all real algebraic numbers is countable, but the set of all real numbers
is uncountable, so the transcendental numbers not only exist but they form
an uncountable set.
It is usually difficult to show that some particular number such as e or n
is transcendental. Liouville's theorem can be used to show that irrational
numbers that are sufficiently well approximated by rationals are necessarily
transcendental. Such numbers are called Liouville numbers and are defined
as follows.
Definition. A real number () is called a Liouville number if for every integer
r ~ 1 there exist integers hr and kr with kr > 0 such that

(7) hrl < -


0< ()-- 1
I kr k/'

Theorem 7.6. E very Liouville number is transcendental.


PROOF. If a Liouville number () were algebraic of degree n it would satisfy
both inequality (7) and

for every r ~ 1, where C(()) is the constant in Theorem 7.5. Therefore


1
or 0 < C(()) < k r-n'
r

The last inequality gives a contradiction if r is sufficiently large. 0

EXAMPLE. The number


00

() = L
m=1
10m!

is a Liouville number and hence is transcendental. In fact, for each r ~ 1 we


can take k r = lor! and
r 1
hr = kr L 10m !'
m=1

Then we have
hr 00 1 1 00 1
o < () - k = L 10m!
r m=r+l
~ 10(r+l)! m=O
L 10m
10/9 1 10/9 1
=---=---<-
10(r+ I)! k/ lor! k/
so (7) is satisfied.

147
7: Kronecker's theorem with applications

Note. The same argument shows that L~=l am l0- m ! is transcendental if


am = 0 or 1 and am = 1 for infinitely many m.

We turn now to a generalization of Dirichlet's theorem due to Kronecker.

7.4 Kronecker's approximation theorem:


the one-dimensional case
Dirichlet's theorem tells us that for any real fJ and every e > 0 there exist
integers x and y, not both 0, such that
IfJx + yl < e.
In other words, the linear form fJx + y can be made arbitrarily close to 0 by a
suitable choice of integers x and y. If fJ is rational this is trivial because we
can make fJx + y = 0, so the result is significant only if fJ is irrational.
Kronecker proved a much stronger result. He showed that if fJ is irrational
the linear form fJx + y can be made arbitrarily close to any prescribed real
number r:J.. We prove this result first for rx in the unit interval. As in the proof
of Dirichlet's theorem we make use of the fractional parts {nfJ} = nfJ - [nfJ].

Theorem 7.7. IffJ is a given irrational number the sequence of numbers {nfJ}
°
is dense in the unit interval. That is, given any r:J., ~ rx ~ 1, and given any
e > 0, there exists a positive integer k such that
r {kfJ} - rxl < e.
Hence, ifh = [kfJ] we have IkfJ - h - rxl < e.
Note. This shows that the linear form fJx + y can be made arbitrarily
close to r:J. by a suitable choice of integers x and y.
PROOF. First we note that {nfJ} =1= {mfJ} if m =1= n because fJ is irrational.
Also, there is no loss of generality if we assume 0 < fJ < 1 since nfJ =
n[fJ] + n{fJ} and {nfJ} = {n{fJ}}.
Let e > 0 be given and choose any r:J., 0 ~ rx ~ 1. By Dirichlet's approxima-
tion theorem there exist integers hand k such that IkfJ - hi < e. Now
either kfJ > h or kfJ < h. Suppose that kfJ > h, so that 0 < {kfJ} < e. (The
argument is similar if kfJ < h.) Now consider the following subsequence of
the given sequence {nfJ}:
{kfJ}, {2kfJ}, {3kfJ}, ....
We will show that the early terms of this sequence are increasing. We have
kfJ = [kfJ] + {kfJ}, so mkfJ = m[kfJ] + m{kfJ}.
Hence

{mkO} = m{kO} if, and only if, {kO} < ~.


m
148
7.5: Extension of Kronecker's theorem to simultaneous approximation

Now choose the largest integer N which satisfies {kO} < liN. Then we have
1 1
N + 1 < {kO} < N'
Therefore {mkO} = m{kO} for m = 1, 2, , N, so the N numbers
{kO}, {2kO}, , {NkO}
form an increasing equally-spaced chain running from left to right in the
interval (0, 1). The last member of this chain (by the definition of N) satisfies
the inequality
N
- - 1 < {NkO} < 1,
N+
or
1
1- N +1< {NkO} < 1.

Thus {NkO} differs from 1 by less than 1/(N + 1) < {kO} < G. Therefore the
first N members of the subsequence {nkO} subdivide the unit interval into
subintervals of length < e. Since rx lies in one of these subintervals, the
theorem is proved. 0

The next theorem removes the restriction 0 :5; rx ~ 1.

Theorem 7.8. Given an)' real rx, any irrational 8.. and an)' f. > 0.. there exist
integers hand k with k > 0 such that
I kO - h - rxl < e.

PROOF. Write rx = [rx] + {rx}. By Theorem 7.7 there exists k > 0 such that
I{kO} - {rx} I < e. Hence
IkO - [kO] - (rx - [rxJ) 1< e
or
IkO - ([kO] - [rxJ) - rxl < e.
Now take h = [kO] - [rx] to complete the proof. o
7.5 Extension of Kronecker's theorem to
simultaneous approximation
We turn now to a problem of simultaneous approximation. Given n irrational
numbers 0b O2 , ••• , On' and n real numbers rxb rx2' ... , rx n, and given e > 0,
we seek integers hI' h 2 , ••• , hn and k such that
IkO i - hi - rxd < e for i = 1,2, ... , n.
149
7: Kronecker's theorem with applications

It turns out that this problem cannot always be solved as stated. For example,
suppose we start with two irrational numbers, say e i and 2e b and two real
numbers ~I and ~2, and suppose there exist integers h b h 2 and k such that
Ike 1 - hI - ~1 I < 8

and
12ke i - h2 - ~21 < 8.

Multiply the first inequality by 2 and subtract from the second to obtain
12h I - h2 + 2~1 - ~21 < 38.
Since 8, ~I and ~2 are arbitrary and h b h2 are integers, this inequality cannot
in general be satisfied. The difficulty with this example is that e i and 2e I are
linearly dependent and we were able to eliminate ei from the two inequalities.
Kronecker showed that the problem of simultaneous approximation can
always be solved if e 1, . . . , en are linearly independent over the integers;
that is, if
n

L Ciei = 0
i= 1

with integer multipliers C 1, ... , Cn implies C1 = ... = cn = O. This restriction


is compensated for, in part, by removing the restriction that the ei be
irrational. First we prove what appears to be a less general result.

Theorem 7.9 (First form of Kronecker's theorem). If ~I' ••• , ~n are arbitrary
real numbers, if e b . . . , en are linearly independent real numbers, and if
8 > 0 is arbitrary, then there exists a real number t and integers h b . . . , hn
such that
Ite i - hi - ~d < 8 for i = 1, 2, ... ,n.

Note. The theorem exhibits a real number t, whereas we asked for an


integer k. Later we show that it is possible to replace t by an integer k, but
in most applications of the theorem the real t suffices.

The proof of Theorem 7.9 makes use of three lemmas.

Lemma 1. Let {An} be a sequence of distinct real numbers. For each real t
and arbitrary complex numbers Co, ... , eN define
N
f(t) = L creitAr.
r=O

Then for each k we have

Ck = lim -1 JT f(t)e -ltAk


. dt.
T-+oo T 0

150
7.5: Extension of Kronecker's theorem to simultaneous approximation

PROOF. The definition of f(t) gives us


N
f(t)e - it).k = L crei().r- ).k)t.
r=O

Hence

from which we find

Now let T ~ 00 to obtain the lemma. D


Lemma 2. If t is real, let
n

(8) F(t) = 1 + L e21ti(tlJr-cxr),


r= 1

where CX l' . . . , CX n and 01 , ... , On are arbitrary real numbers. Let


L= sup IF(t)l.
-oo<t<+oo
Then the following two statements are equivalent:
(a) For every G > 0 there exists a real t and integers h 1 , ••• , hn such that
ItOr - CX r - hrI < G for r = 1, 2, ... , n.
(b) L = n + 1.
PROOF. The idea of the proof is fairly simple. Each term of the sum in (8)
has absolute value 1 so IF(t) I :::; n + 1. If (a) holds then each number tOr - CX r
is nearly an integer hence each exponential in (8) is nearly 1 so IF(t) I is nearly
n + 1. Conversely, if (b) holds then IF(t)1 is nearly n + 1 for some thence
every term in (8) must be nearly 1 since no term has absolute value greater
than 1. Therefore each number tOr - CX r is nearly an integer so (a) holds.
Now we transform this idea into a rigorous proof.
First we show that (a) implies (b). If (a) holds take e = 1/(2nk), where
k 2 1, and let t k be the corresponding value of t given by (a). Then
2n(t k Or - cx r) differs from an integer multiple of 2n by less than 11k so
1
cos 2n(t k Or - CX r) 2 cos k.
Hence
n 1
IF(t k ) I 2 1 + L cos 2n(t k Or - CX r) 2 1 + n cos-
r= 1 k

151
7: Kronecker's theorem with applications

and therefore L :2:: IF(t k ) I :2:: 1 + n cos(l/k). Letting k -+ 00 we find L :2:: n + 1.


Since L ::::; n + 1 this proves (b).
Now we assume (a) is false and show that (b) is also false. If(a) is false there
exists an G > 0 such that for all integers hI' ... , hn and all real t there is a k,
1 ::::; k ::::; n, such that
G
(9) ItO k - rt k - hkl :2:: 2n·

(We can also assume that G ::::; n/4 because if (a) is false for G it is also false for
every smaller G.) Let X r = tOr - rt r - hr. Then (9) implies 12nxkl :2:: Gso the
point 1 + e21tixk lies on the circle of radius 1 about 1 but outside the shaded
sector shown in Figure 7.1.

Figure 7.1

Now 11 + e if I < 2 so 11+ e if I = 2 - b for some b > O. Hence


11 + e21tixk I ::::; 11 + e if I = 2 - b,
so

IF(t) I = 11 + rtl e21tiXr IS 11 + e21tixk I + rt1I


r¢k
e21tixr I

::::; (2 - b) + (n - 1) = n + 1 - b.
Since this is true for all t we must have L ::::; n + 1- b < n + 1, contra-
dicting (b). D

Lemma 3. Let g = g(Xl' ... , x n ) be the polynomial in n variables given by

and write
(10) g p = 1 + "1..J arl, ... ,r x 1 rl
l1
..• X n r l1 ,

152
7.5: Extension of Kronecker's theorem to sin\ultaneous approximation

where p is a positive integer. Then the coefficients art ..... r" are positive
integers such that
(11) 1 + L art ..... r,1 =
+ n)P, (1
and the number of terms in (10) is at most (p + l)n.
PROOF. Since 1 + L art ..... r" = gP(I, 1, ... , 1) = (1 + n)P this proves (11).
Let 1 + N be the number of terms in (10). We shall prove that
(12) 1 + N ~ (p + It
by induction on n. For n = 1 we have

(1 + xl)P = 1 + (~)xl + (~)x/ + ... +x/


and the sum on the right has exactly p + 1 terms. Thus (12) holds for n = 1.
Ifn> 1 we have
gp = {(I + Xl + + x n- l ) + xn}P
= (1 + Xl + + xn-I)P + (i)(l + ... + xn_I)P-I xn + ... + x/,

so if there are at most (p + 1t - 1 terms in each group on the right there will
be at most (p + l)n terms altogether. This proves (12) by induction. D
PROOF OF KRONECKER'S THEOREM. Choosing F(t) as in Lemma 2 we have
n
F(t) = 1 + L e21ti(tOr-cxr).
r =: 1

By Lemma 2, to prove Kronecker's theorem it suffices to prove that


L= sup IF(t)l=n+l.
-oo<t<+oo

The pth power of F( t) is a sum of the type discussed in Lemma 1,


N
(13) f(t) = FP(t) = 1 + L creit).r,
r= 1

with Ao = 1 and Ar replaced by 2n(rl0l + ... + rnO n) if r ~ 1. The numbers


Ar are distinct because the Oi are linearly independent over the integers. The
coefficients C r in (13) are the integers art ..... r" of Lemma 3 multiplied by a
factor of absolute value 1. Hence (11) implies
N
(14) 1+ L Icrl = 1 + La rt ..... r" = (1 + n)P.
r= 1

By Lemma 1 we have

(15) C
r
= lim -1 f.T FP(t)e-lt).r
. dt.
T~oo T 0

153
7: Kronecker's theorem with applications

Now IF(t)1 ~ L so IFP(t) I ~ LP for all t, hence

I ~ L T

FP(t)e-it.l. r dtl :::;; ~ L T

U dt = U.

Hence (15) implies lerl ~ LP for each r, and (14) gives us


(1 + n)P ~ (N + I)LP ~ (p + l)nLP
by Lemma 3. Therefore

from which we find

IOg(n 1 1
) :::;; ~ log(p + 1).
Now let p ---+ 00. The last inequality becomes log[(n + 1)/L] ~ 0, so L 2
n + 1. But L ~ n + 1 hence L = n + 1, and this proves Kronecker's
theorem. []

The next version of Kronecker's theorem replaces the real number t by an


integer k.

Theorem 7.10 (Second form of Kronecker's theorem). If Ct 1 , ••• , Ct n are


arbitrary real numbers, if01, . .. ,On' 1 are linearly independent real numbers,
and if £ > 0 is given, then there exists an integer k and integers m 1 , •. , mn
sueh that

PROOF. We apply the first form of Kronecker's theorem to the system


Ct 1 , ••• , Ct n, 0 and {Ol}' {02}, ... , {On}' 1, with £/2 instead of £, where £ < 1.
Then there exists a real t and integers h 1, . . . , hn + 1 such that
£
It{OJ-hi-Ctil<2 fori=I,2, ... ,n

and

(16)

The last inequality shows that t is nearly equal to the integer hn + 1. Take
k = hn + 1 • Then (16) implies
Ik{OJ - hi - Ctd = It{OJ - hi - Ct i + (k - t){OJI
~ It{ OJ - hi - Ctd + Ik - t I < £.

154
7.6: Applications to the Riemann zeta function

Hence, writing {Oi} = 0i - [Oa, we obtain


Ik(Oi - [Oa) - hi - ad < G

or, what is the same thing,


IkO i - (hi + k[Oa) - (Xii < G.

Putting mi = hi + k[Oa we obtain the theorem. o


7.6 Applications to the Riemann zeta
function
With the help of Kronecker's theorem we can determine the least upper
bound and greatest lower bound of I'((J + it) I on any fixed line (J = constant,
(J > 1.

Definition. For fixed (J, we define


m((J) = infl '((J + it) I and M((J) = sup I'((J + it) I,
t t

where the infimum and supremum are taken over all real t.

Theorem 7.11. For eachfixed (J > 1 we have


'(2(J)
and m(O") = (0")'

PROOF. For (J > 1 we have I'((J + it)l ~ '((J) so M((J) = '((J), the supremum
being attained on the real axis. To obtain the result for m((J) we estimate the
reciprocal 11/ ((s) I. For (J > 1 we have

(17) 1_1
(s)
1- nil -
- p p
-si -< n (1 + P -a) --
p
'((J)
(20")'

Hence I'(s) I ~ '(2(J)/'((J) so m((J) 2 '(2(J)/'((J).


Now we wish to prove the reverse inequality m((J) ~ '(2(J)/'((J). The idea
is to show that the inequality
11 - p-si ~ 1 + p-a
used in (17) is very nearly an equality for certain values of t. Now
1 - p-s = 1 - p-a-it = 1 - p-ae-itlo gp = 1 + p-aei(-tlo gp -1t),

so we need to show that - t log p - 'It is nearly an even multiple of 2n for


certain values of t. For this we invoke Kronecker's theorem. Of course,
there are infinitely many terms in the Euler product for 1/'(s) and we cannot
expect to make - t log p - n nearly an even multiple of 2n for all primes p.
But we will be able to do this for enough primes to obtain the desired
inequality.

155
7: Kronecker's theorem with applications

Choose any c, 0 < c < nf2, and choose any integer n ~ 1. We apply
Kronecker's theorem to the numbers
-1
Ok = 2;" log Pk' k = 1,2, ... , n,

where Pl' ... , Pn are the first n primes. The Oi are linearly independent
because
n

L ai log Pi = 0 implies log(p 1 at ... Pn an) =0


i= 1
so Plat . . . Pn an = 1 hence each ai = O. We also take lJ. l = lJ.2 = ... = lJ. n = 1.
Then by Theorem 7.9 there is a real t and integers h l , . . . , hn such that
ItO k - lJ.k - hk I < cf(2n), which means
(18) I-t log Pk - n - 2nh k l < c.
For this t we have
+ Pk -O'e i(-tlo gpk -1t)
1 - Pk -s = 1 - Pk -O'e-itlogpk = 1
= 1 + Pk-O'cos(-tlogpk - n) + ipk-O'sin(-tlogpk - n),
so
11 - Pk - S I ~ 1 + Pk - 0' cos( - t log Pk - n).
But (18) implies
cosl-t log Pk - nl = cosl-t log Pk - n - 2nh k l > cos c,
since the cosine function decreases in the interval [0, nf2]. Hence
11 - Pk - S I > 1 + Pk - 0' cos c.
Now consider any partial product of the Euler product for If'(s). For
a given c and n there exists a real t (depending on c and on n) such that

(19) Ill1 (1 - Pk -')1 = kD1 11 - Pk -'I > kD1 (1 + Pk -a cos e).


Now

and hence, by the Cauchy condition for convergent products, there is an


no such that n ~ no implies

fI
Ik=n+l
11 - Pk -'I - 11 < c

or

n 11 -
00

1- c< Pk -si < 1 + c.


k=n+l
156
7.7: Applications to periodic functions

Using (19) with n ~ no we have


1 n o o n
Ws)l = }]1,1 - Pk-
s
'JI
1 1 - Pk-si > (1 - e)
I
a
(1 + Pk- cos e), }]I
This holds for n ~ no and a certain t depending on n and on c. Hence
1 l i n
- (
m (J
) = , f 1((
In t
. )1 = sup 1((
(J + It t
. )1 ~ (l - e)
(J + It k =1
n
(1 + Pk -a cos e).

Letting n -+ 00 we find
1
-(-) ~ (1 - c)
m (J
00

k= 1
n(1 + Pk -(1 cos c).

We will show in a moment that the last product converges uniformly for
o:: ;
c :::;; n/2. Therefore we can let c -+ 0 and pass to the limit term by term
to obtain
1 00 '((J)
-(1

m(O") ~ }]I
(1 + Pk ) = ((20")'

This gives the desired inequality m((J) :::;; '(2(J)/'((J).


To prove the uniform convergence of the product, we use the fact that
a product n (1 + fn(z)) converges uniformly on a set if, and only if, the
series Lfn(z) converges uniformly on this set. Therefore we consider the
L
series Pk cos c. But this is dominated by Pk
-(1 L n
-(1 :::;; L
= '((J) so the
-(1

convergence is uniform in the interval 0 :::;; c :::;; n/2, and the proof is complete.
D
7.7 Applications to periodic functions
We say that n complex numbers Wl' W 2 , ••• , W n are linearly independent
over the integers if no linear combination
alw l + a2 W2 + ... + an(Qn
with integers coefficients is 0 except when a l = a2 = ... = an = O. Other-
wise the numbers Wl' .. , W n are called linearly dependent over the integers.
Elliptic functions are meromorphic functions with two linearly indepen-
dent periods. In this section we use Kronecker's theorem to show that there
are no meromorphic functions with three linearly independent periods
except for constant functions.
Theorem 7.12. Let W l and W2 be periods off such that the ratio W2/Wl is
real and irrational. Then f has arbitrarily small nonzero periods. That is,
given c > 0 there is a period W such that 0 < IWI < c.
PROOF. We apply Dirichlet's approximation theorem. Let () = W 2 /W l '
Since () is irrational, given any c > 0 there exist integers hand k with k > 0
such that
c
1kO - hi < 1WI I'

157
7: Kronecker's theorem with applications

Multiplying by I w 1 1 we find
IkW2 - hw 1 I < c.
But w = kW 2 - hW 1 is a period of f with Iwl < c. Also, w =I- 0 since w21w 1
is irrational. 0

Theorem 7.13. Iff has three periods W1 , W2, W3 which are linearly independent
over the integers, then f has arbitrarily small nonzero periods.
PROOF. Suppose first that w21w 1 is real. If w 21w 1 is rational then W1 and
W2 are linearly dependent over the integers, hence W b W2 , W3 are also depen-
dent, contradicting the hypothesis. If W2 IW I is irrational, thenfhas arbitrarily
small nonzero periods by Theorem 7.12.
Now suppose w21w 1 is not real. Geometrically, this means that W1 and
W2 are not collinear with the origin. Hence W3 can be expressed as a linear
combination of W1 and W2 with real coefficients, say
W 3 = ~W1 + pw 2 , where ~ and p are real.
Now we consider three cases:
(a) Both ~ and p rational.
(b) One of ~, p rational, the other irrational.
(c) Both ~ and p irrational.
Case (a) implies w 1 , W2 , W3 are dependent over the integers, contradicting
the hypothesis.
For case (b), assume ~ is rational, say ~ = alb, and {3 is irrational. Then
we have

so

This gives us two periods bW 3 - aWl and bW2 with irrational ratio, hence f
has arbitrarily small periods. The same argument works, of course, if p is
rational and ~ is irrational.
Now consider case (c), both ~ and p irrational. Here we consider two
subcases.
(c l ) Assume ~ and p are linearly dependent over the integers. Then
there exist integers a and b, not both zero, such that a~ + bP = O. By sym-
metry, we can assume that b =I- O. Then p = -a~/b and
a
W3 = ~WI - b~wz, so bW 3 = ~(bWI - awz)·

Again we have two periods bW3 and bW I - aW2 with irrational ratio, so f
has arbitrarily small nonzero periods.
(cz) Assume ~ and p are linearly independent over the integers. Then by
Kronecker's theorem, given any c > 0 there exist integers h b h2 and k

158
Exercises for Chapter 7

such that
G
Ikf3 - h2 < 1 1+ IW l I + IW 2 "
Multiply these inequalities by Iwll, Iw 2 1, respectively, to get
l
IklXW 1 - hI WI I < 1 + IWGlw
l
I +' IW 2 I'
Since kW 3 = krJ.w l + kf3W2 we find, by the triangle inequality,
G(lwll + Iw 2 1)
Ikw3-hlWl-h2W21<1
+ IWt I + IW2 ,<G.
Thus kW3 - hlw l - h2w2 is a nonzero period with modulus <G. D

Note. In Chapter 1 we showed that a function with arbitrarily small


nonzero periods is constant on every open connected subset in which it is
analytic. Therefore, by Theorem 7.13, the only meromorphic functions
with three independent periods are constant functions.

Further applications of Kronecker's theorem are given in the next chapter.

Exercises for Chapter 7


1. Prove the following extension of Dirichlet's approximation theorem.
Given n real numbers 8 1 , ... , 8n and given an integer N ~ 1, there exist integers
h l , ... , h n and k, with 1 ~ k ~ N n , such that

1
Ik8 i - hi I < N for i = 1, 2, ... , n.

2. (a) Given n real numbers 8 1 " .• ,8n , prove that there exist integers h l , . . . , hn and
k > 0 such that

I
8i - ih'l < k 1l + l/n for i = 1,2, ... , n.

(b) If at least one of the 8i is irrational, prove that there is an infinite set of n-tuples
(hl/k, ... , hn/k) satisfying the inequalities in (a).

3. This exercise gives another extension of Dirichlet's approximation theorem. Given


m linear forms,

i = 1,2, ... , m,

in n + m variables Xl' ... , X n, Yl, ... ,Ym, prove that for each integer N > 1 there
exists integers Xl' ... , X n, Yl" .. , Ym such that

1
IL i I < N for i = 1, 2, ... , m

159
7: Kronecker's theorem with applications

and 0 < max{ lXII, ... , Ixn !} ~ N m / n• Hint: Let M j = aj1Xl + ... + ajnXn and
examine the points ({MIL ... , {M m }) in the unit cube in m-space, where {M j } =
M j - [M j ].

4. Let {} be irrational, 0 < () < 1. Then () lies between two consecutive Farey fractions,
say

~ < () <~.
b d
2
(a) Prove that either 0 - a/b < 1/(2b ) or c/d - () < 1/(2d 2 ).
(b) Deduce that there exist infinitely many fractions h/k with (h, k) = 1 and k > 0
such that

.
1
()_~1<_1
k 2k 2

5. Let a = (1 + )5)/2. This exercise shows that the inequality

(20)

has only a finite number of solutions in integers hand k with k > 0 if 0 < c < 1/)5.
(a) Let f3 = C/.. - )5 so that a and f3 are roots of the equation x 2 - x - I = O.
Show that for any integers hand k with k > 0 we have

and deduce that

(b) If (20) has infinitely many solutions h/k with k > 0, say h l/k l ' h2/k - , ... , show that
kn -+ X'; as n -+ x and use part (a) to prove that c :2: 1/)5.

6. In Lemma 2, define
L = lim sup IF(t) I instead of L = sup IF(t) I.
t~ + oc -oo<t<oo

Prove that the equation L = n + 1 is equivalent to the following statement: For


every G > 0 and every T > 0 there exists a real t > T and integers hi' ... , h n such that
It()i - hi - ad < G for every i = 1, 2, ... , n.
7. Prove that the multiplier t in the first form of Kronecker's theorem can be taken
positive and arbitrarily large. That is, under the hypotheses of Theorem 7.9, if T > 0
is given there exists a real t > T satisfying the n inequalities It()i - hi - ad < G.
Show also that the integer multiplier k in the second form of Kronecker's theorem
can be taken positive and arbitrarily large.

160
General Dirichlet series and
Bohr's equivalence theorem

8.1 Introduction
This chapter treats a class of series, called general Dirichlet series, which
includes both power series and ordinary Dirichlet series as special cases.
Most of the chapter is devoted to a method developed by Harald Bohr [6]
in 1919 for studying the set of values taken by Dirichlet series in a half-plane.
Bohr introduced an equivalence relation among Dirichlet series and showed
that equivalent Dirichlet series take the same set of values in certain half-
planes. The theory uses Kronecker's approximation theorem discussed in
the previous chapter. At the end of the chapter applications are given to the
Riemann zeta function and to Dirichlet L-functions.

8.2 The half-plane of convergence of general


Dirichlet series
Definition. Let {A(n)} be a strictly increasing sequence of real numbers such
that A(n) ~ + 00 as n ~ 00. A series of the form
00

L a(n)e-SA(n)
n=l

is called a general Dirichlet series. The numbers A(n) are called the
exponents of the series, and the numbers a(n) are called its coefficients.

As usual, we write s = (J + it where (J and t are real.

Note. When A(n) = log n then e-SA(n) = n- S and we obtain the ordinary
Dirichlet series L a(n)n- s • When A(n) = n the series becomes a power series
in x, where x = e- S •

161
8: General Dirichlet series and Bohr's equivalence theorem

A general Dirichlet series is analogous to the Laplace transform of a


function, Sg:: f(t)e- st dt. As a matter of fact, both Dirichlet series and Laplace
transforms are special cases of the Laplace-Stieltjes transform, S~ e- st drx(t).
When rx(t) has a continuous derivative rx'(t) = f(t) this gives the Laplace
transform of f When rx is a step function with jump a(n) at the point A(n)
the integral becomes the general Dirichlet series L a(n)e-SA(n). Much of
what we do here can be extended to Laplace-Stieltjes transforms, but we
shall not deal with these generalizations.
As is the case with ordinary Dirichlet series, each general Dirichlet
series has associated with it an abscissa (Jc of convergence and an abscissa
(Ja of absolute convergence. We could argue as in Chapter 11 of [4] to

prove the existence of (Jc and (Ja. Instead we give a different method of proof
which also expresses (Jc and (Ja in terms of the exponents A(n) and the
coefficients a(n).

Theorem 8.1. Assume that the series L a(n)e-SA(n) converges for some s with
positive real part, say for s = So with (Jo > O. Let
-I. logIL~=la(k)1
L - 1m sup '( ) .
n-oo An

Then L ::; (J 0 • Moreover, the series converges in the half-plane (J > L, and
the convergence is uniform on every compact subset of the half-plane
(J > L.

PROOF. First we prove that L ::; (J o. Let A(n) denote the partial sums of the
coefficients,
n
A(n) = L a(k).
k=l
Note that A(n) > 0 for all sufficiently large n. If we prove that for every
e > 0 we have
(1) logIA(n)1 < ((Jo + e)A(n)
for all sufficiently large n, then it follows that

logIA(n)1
A(n) < (Jo +e
for these n, so L ::; (Jo + e, hence L ::; (Jo. Now relation (1) is equivalent to the
inequality
(2)

To prove (2) we introduce the partial sums


n
S(n) = L a(k)e-SOA(k).
k=l

162
8.2: The half-plane of convergence of general Dirichlet series

The S(n) are bounded since the series Lk=


1 a(k)e-SOA(k) converges. Suppose

that IS(n) I < M for all n. To express A(n) in terms of the S(n) we use partial
summation:
n n
A(n) = L a(k) = L a(k)e-SOA(k)eSOA(k)
k=1 k=1
n
= L {S(k) - S(k - 1)}eSoA (k),
k=1
provided 8(0) = O. Thus
n n-l
A(n) = L S(k)eSOA(k) - L S(k)eSOA(k+ 1)
k=1 k=1
n-l
= L S(k) {esoA(k) - esoA(k+ 1)} + S(n)eSOA(n).
k=1

Hence
n-l
IA(n)1 < M L leSOA(k) - eSOA (k+1)1 + Me(]OA(n).
k=1

But
1
n-l jA(k+1) Sou n-l jA(k+1)
L IesoA(k) - esoA(k+ 1) I = L
n- 1
So e du
I
~ ISo I L e(]OU du
k= 1 k= 1 A(k) k= 1 A(k)

= ISol jA(nlel1OUdU = ~(eI10A(nl _ eI1OA(1l) < ~eI10A(nl.


A(I) 0'0 0'0

Thus

Now A(n) ~ 00 as n ~ 00 so

eeA(nl > M(l + 1;:1)


if n is sufficiently large. Hence for these n we have IA(n) I < e(UO+E:)A(n), which
proves (2) and hence (1). This proves that L ~ 0'0.
Now we prove that the series converges for all S with (J > L. Consider
any section of the series L
a(n)e-SA(n), say L~=a. We shall use the Cauchy
convergence criterion to show that this section can be made small when
a and b are sufficiently large. We estimate the size of such a section by using

163
8: General Dirichlet series and Bohr's equivalence theorem

partial summation to compare it to the partial sums A(n) = L~= 1 a(k). We


have
b b
L a(n)e-SA(n) = L {A(n) - A(n - l)}e- SA (n)
n=a n=a
b
= L A(n) {e-SA(n) - e-SA(n+ I)} + A(b)e-SA(b+ 1)
n=a

- A(a - l)e- SA (a).

This relation holds for any choice of s, a and b. Now suppose s is any complex
number with (J > L. Let e = !«(J - L). Then e > 0 and (J = L + 2e. By the
definition of L, for this e there is an integer N(e) such that for all n ~ N(e)
we have
logl A(n) I
A(n) < L + e.

We can also assume that A(n) > 0 for n ~ N(e). Hence


IA(n) I < e(L+f)A(n) for all n ~ N(e).
If we choose b ~ a > N(e) we get the estimate

The last two terms are e- fA (b+1) + e-fA(a) since L +e- (J = -c. Now we
estimate the sum by writing

le-SA(n) - e-SA(n+ 1)1 = -s


A(n+ 1)
e- su du
I ~ lsi fA(n+ 1)
e- au du
f A(n) A(n)
I

so

Lb e(L+f)A(n) Ie-SA(n) - e-SA(n+ 1)


b
I ~ Is I L e(L+E)A(n)
fA(n+ 1)
e- au du
n=a n=a A(n)

b fA(n+ 1) b fA(n+ 1)
~ lsi L e-aUe(L+f)U du = lsi L e-
fU
du
n=a A(n) n=a A(n)

A(b+ 1) II
= lsi e-fudu = ~(e-fA(a) - e- fA (b+l»).
f A(a) e
Thus we have

In=a
t a(n)e-SA(n) I ::s; ~ (e-·A(a) -
e
e-'A(b+ 0) + e-'A(b+ 1) + e-·A(a).

164
8.2: The half-plane of convergence of general Dirichlet series

Each term on the right tends to 0 as a -+ 00, so the Cauchy criterion shows
that the series converges for all s with (J > L. This completes the proof.
Note also that this proves uniform convergence on any compact subset of
the half-plane (J > L. D

Theorem 8.2. Assume the series La(n)e-SA(n) convergesfor some s with (J >0
but diverges for all s with (J < o. Then the number
_ 1· log IL~= 1 a(k) I
L - 1m sup
n-oo
'( )
A n

is the abscissa of convergence of the series. In other words, the series


converges for all s with (J > L and diverges for all s with (J < L.
PROOF. We know from Theorem 8.1 that the series converges for all s with
(J > L and that L cannot be negative. Let S be the set of all (J > 0 such that

the series converges for some s with real part (J. The set S is nonempty and
bounded below. Let (Jc be the greatest lower bound of S. Then (Jc > 0, Each
(J in S satisfies L :::; (J hence L :::; (Jc. If we had (Jc > L there would be a (J in

the interval L < (J < (Jc. For this (J we would also have convergence for all
s with real part (J (by Theorem 8.1) contradicting the definition of (J c. Hence
(J c = L. But the definition of (Jc shows that the series diverges for all s with

o :::; (J < L. By hypothesis it also diverges for all s with (J < O. Hence it
diverges for all s with (J < L. This completes the proof. D

As a corollary we have:

Theorem 8.3. Assume the series L a(n)e-SA(n) converges absolutely for some s
with (J > 0 but diverges for all s with (J < o. Then the number

_ 1· log L~= 1 Ia(k) I


(J a - 1m SUp '( )
n-oo An

is the abscissa of absolute convergence of the series.


PROOF. Let A be the abscissa of convergence of the series L la(n)le-SA(n).
Then, by Theorem 8.2,

_ 1· log L~= 1 Ia(k) I


A - 1m sup '( ) .
n-oo A n

We wish to prove that L la(n)le-aA(n) converges if (J > A and diverges if


(J < A. Clearly if (J > A then the point s = (J is within the half-plane of

convergence of L Ia(n) Ie-SA(n) so L Ia(n) Ie-GA(n) converges.


Now suppose L la(n)le-GA(n) converges for some (J < A. Then the series
L Ia(n) Ie-SA(n) converges absolutely for each s with real part (J so, in particular
it converges for all these s, contradicting the fact that A is the abscissa of
convergence of L Ia(n) Ie-SA(n) . D

165
8: General Dirichlet series and Bohr's equivalence theorem

8.3 Bases for the sequence of exponents of a


Dirichlet series
The rest of this chapter is devoted to a detailed study of Harald Bohr's
theory with applications to the Riemann zeta-function and Dirichlet's
L-series. The first notion we need is that of a basis for the sequence of
exponents of a Dirichlet series.

Definition. Let A = {)w(n)} be an infinite sequence of distinct real numbers. By


a basis of the set A we shall mean a finite or countably infinite sequence
B = {f3(n)} of real numbers satisfying the following three conditions:
(a) The sequence B is linearly independent over the rationals. That is,
for all m ~ 1, if

with rational multipliers rk' then each rk = O.


(b) Each A(n) is expressible as a finite linear combination of terms of B,
say
q(n)

A(n) = L rn,kf3(k)
k=1

where the r n, k are rational and the number of summands q(n) depends
on n. (By condition (a), if A(n) i= 0 this representation is unique.)
(c) Each f3(n) is expressible as a finite linear combination of terms of A,
say
m(n)

f3(n) = L tn,kA(k)
k=1

where the tn,k are rational and m(n) depends on n.

EXAMPLE 1. Let A be the set of all rational numbers. Then B = {I} is a basis.
EXAMPLE 2. Let A = {log n}. Then B = {log Pn} is a basis, where Pn is the
nth prime. It is easy to verify properties (a), (b) and (c). For independence
we note that
q

L rk log Pk = 0
k=l
implies rl = ... = rq = O.

To express each A(n) in terms of the basis elements we factor n and compute
log n as a linear combination of the logarithms of its prime factors. Property
(c) is trivially satisfied since B is a subsequence of A.

166
8.4: Bohr matrices

Theorem 8.4. Every sequence A has a subsequence which is a basis for A.


PROOF. Construct a basis as follows. For the first basis element take ,1(nl),
the first nonzero A (either ,1(1) or ,1(2)), and call this f3(I). Now delete the
remaining elements of A that are rational multiples of f3(I). If this exhausts
all of A take B = {f3( I)}. If not, let ,1(nz) denote the first remaining A, take
f3(2) = ,1(nz), and strike out the remaining elements of A which are rational
linear combinations of f3(I) and f3(2). Continue in this fashion to obtain a
sequence B = (f3(I), f3(2), ...) = (,1(nl)' ,1(nz), ...). It is easy to verify that B
is a basis for A. Property (a) holds by construction, since each f3 was chosen
to be independent of the earlier elements. To verify (b) we note that every A is
either an element of B or a rational linear combination of a finite number of
elements of B. Finally, (c) holds trivially since B is a subsequence of A. D

Note. Every sequence A has infinitely many bases.

8.4 Bohr matrices


It is convenient to express these concepts in matrix notation. We display the
sequences A and B as column matrices, using an infinite column matrix for A
and a finite or infinite column matrix for B, according as B is a finite or
infinite sequence.
We also consider finite or infinite square matrices R = (rij) with rational
entries. If R is infinite we require that all but a finite number of entries in each
row be zero. Such rational square matrices will be called Bohr matrices.
We define matrix addition and multiplication of two infinite Bohr
matrices as for finite matrices. Note that a sum or product of two Bohr
matrices is another Bohr matrix. Also, the product RB of a Bohr matrix R
with an infinite column matrix B is another infinite column matrix r.
Moreover, we have the associative property (R 1 R z )B = R1(RzB) if R 1
and R z are Bohr matrices and B is an infinite column matrix.
In matrix notation, the definition of basis takes the following form. B is
called a basis for A if it satisfies the following three conditions:
(a) If RB = 0 for some Bohr matrix R, then R = o.
(b) There exists a Bohr matrix R such that A = RB.
(c) There exists a Bohr matrix T such that B = T A.
The relation between two bases Band r of the same sequence A can be
expressed as follows:

Theorem 8.5. If A has two bases Band r, then there exists a Bohr matrix
A such that r = AB.
PROOF. There exist Bohr matrices Rand T such that r = TA and A = RB.
Hence r = T(RB) = (TR)B = AB where A = TR. D
167
8: General Dirichlet series and Bohr's equivalence theorem

Theorem 8.6. Let Band r be two bases for A, and write r = AB, A = RBB,
A = Rrr, where A, R B, R r ar~ Bohr matrices. Then R B = RrA.

Note. If we write AlB for R B, Air for R r and riB for A, this last equation
states that

PROOF. We have A = RBB and A = Rrr = RrAB. Hence RBB = RrAB,


so (R B - RrA)B = o. Since R B - RrA is a Bohr matrix and B is a basis, we
must have R B - RrA = O. D

8.5 The Bohr function associated with a


Dirichlet series
To every Dirichlet series f(s) = L~= 1 a(n)e-SA(n) we associate a function
F(Zb Z2' ...) of countably many complex variables Zl' Z2, ... as follows.
Let Z denote the column matrix with entries Z b Z 2, .... Let B = {p(n)}
be a basis for the sequence A = {A(n)} of exponents, and write A = RB,
where R is a Bohr matrix.

Definition. The Bohr function F(Z) = F(z b Z 2' ...) associated with f(s),
relative to the basis B, is the series
00

F(Z) = L a(n)e-(RZ)n,
n=l

where (RZ)n denotes the nth entry of the column matrix RZ.

In other words, if
q(n)

A(n) = L rn,kp(k)
k=l

then
00

F(Zl' Z2' ...) = L a(n)e-(r n ,l::l +ooo+rn,q(n)=q(n».


n=l

Note that the formal substitution Zm = spm gives Z = sB, RZ = sRB =


sA, so (RZ)n = sA(n) and hence
00

F(sB) = L a(n)e-SA(n) = f(s).


n=l

In other words, the Dirichlet series f(s) arises from F(Z) by a special choice
of the variables Z b Z 2, .... Therefore, if the Dirichlet series f(s) converges
for s = (J + it the associated Bohr series F(Z) also converges when Z = sB.

168
8.5: The Bohr function associated with a Dirichlet series

Moreover, if the Dirichlet series f(s) converges absolutely for s = a + it


then the Bohr series F(Z) converges absolutely for any choice of Z b Z2' ...
with Re Zn = af3(n) for all n. To see this we note that if Re Zn = af3(n) then
Re Z = aB so
00 00 00

L la(n)e-(RZ)nl = L la(n)le-a(RB)n = L la(n)le-a).(n).


n=l n=l n=l
To emphasize the dependence of the Bohr function on the basis B we
sometimes write A = RBB and
00

F B(Z) = L a(n)e-(RBZ)n.
n=l
Bohr functions F Band F r corresponding to different bases are related by
the following theorem.

Theorem 8.7. Let Band r be two bases for A and write r = AB for some
Bohr matrix A. Then

PROOF. By Theorem 8.6 we have


A = RBB = Rrr, where R B = RrA.
Hence
00 00

FB(Z) = L a(n)e-(RBZ)J1 = L a(n)exp{ -(RrAZ)n} = Fr(AZ). D


n=l n=l

Definition. Assume the Dirichlet series f(s) = L~= 1 a(n)e-s).(n) converges


absolutely for some s = a + it. We define Uf(a; B) to be the set of values
taken on by the associated Bohr function, relative to the basis B, when
Re Z = aB. Thus,

Uf(a; B) = {F(Z): Re Z = aB}.

The next theorem shows that this set is independent of the basis B.

Theorem 8.8. rr Band r are two bases for A then Uf( a; B) = Uf(a; r).
PROOF. Choose any value F B(Z) in Uf(a; B), so that Re Z = aB. By Theorem
8.7 we have F B(Z) = F r(AZ), where r = AB. But

ReAZ = ARe Z = AaB = aAB = ar


so F B(Z) E Uf(a; r). This proves Uf(a; B) ~ Uf(a; r), and a similar argument
givesUf(a;r)~ Uf(a;B). D
169
8: General Dirichlet series and Bohr's equivalence theorem

Note. Since U /(0'; B) is independent of the basis B we designate the set


U /(0'; B) simply by U /(0').

8.6 The set of values taken by a Dirichlet


series f(s) on a line (J = (J 0
This section relates the set U/( 0'0) with the set of values taken by the Dirichlet
series f(s) on the line a = 0'0.

Definition. If the Dirichlet series f(s) = L:= 1 a(n)e - sA(n) converges absolutely
for a = 0'0 we let
V/(ao) = {f(ao + it): -00 < t < +oo}
denote the set of values taken by f(s) on the line a = 0'0.

Since f(s) can be obtained from its Bohr function F(Z) by putting Z = aB,
it follows that V/(ao) ~ U /(0'0). Now we prove an inclusion relation in the
other direction. r

Theorem 8.9. Assume 0'0 > a a' where a a is the abscissa ofabsolute convergence
of a Dirichlet series f(s). Then the closure of V/(ao) contains U /(0'0). That
is, we have

V/(a o) ~ U /(0'0) ~ V/(a o), and hence U /(0'0) = V/(a o).


PROOF. The closure V/(a o) is the set of adherent points of V/(ao). We are
to prove that every point u in U /(0'0) is an adherent point of V/(a o). In
other words, given u in U /(0'0) and given e > 0 we will prove that there exists
a v in V/(a o) such that lu - vi < e. Since v = f(ao + it) for some t, we are
to prove that there exists a real t such that
If(ao + it) - ul < e.
Since u E U /(0'0) we have u = F(Zl' Z2' ...) where Zn = a o f3(n) + iYn. Hence
Z = 0'0 B + i 1': RZ = 0'0 RB + iR Y = 0'0 A + iR Y,
so

say. Therefore
00

u= L a(n)e-aoA(n)e-iJln.
n=l

On the other hand, we have


00

f(ao + it) = L a(n)e-aoA(n)e-itA(n),


n=l

170
8.6: The set of values taken by a Dirichlet series I(s) on a line (J = (Jo

hence
00

f((1o + it) - u = I a(n)e-UQA,(n)(e-itA,(n) - e- iJln ).


n=1

The idea of the proof from here on is as follows: First we split the sum into
two parts, I~ 1 + = I:=N + l' We choose N so the second part N + 1 is I:=
small, say its absolute value is < !8. This is possible by absolute convergence.
Then we show that the first part can be made small by choosing t properly.
The idea is to choose t to make every exponential e - itA,(n) very close to e - iJln
simultaneously for every n = 1,2., ... ., N. Then each factor e-iti.(n) - e- iJln
will be small, and since there are only N terms., the whole sum will be small.
Now we discuss the details. For the given 8, choose N so that

f a(n)e-ao;'(n)(e-it;'(n) - e- illn ) I < ~.


In=N+1 2
Then we have

This holds for any choice of t. We wish to choose t to make the first sum
<!8. Since IeitA,(n) I = 1 we can rewrite the sum in question as follows:
IntI a(n)e-ao;'(n)(e- it;'(n) - e- i ll n ) =I IntI e- it;'(n)a(n)e-ao;'(n)(l - ei(t;'/Jl-Iln») I
N
~ I I I
a(n) e-UQA,(n) lei(tA,(n)-Jln) - 11.
n=1

Let M = 1 + I~=1 /a(n)/e-UQA,(n). For the given 8 there is a <5 > 0 such
that

(3) /e ix _ 11 < _8_ iflxl < <5.


2M

Suppose we could choose a real t and integers k 1 , ••• , k N such that


(4) tA(n) - J-ln = 2nk n + Xn

where IX n I < <5 for n = 1, 2, ... , N. Then for this t we would have

By (3), this would give us


lei(tA,(n)-Jln) _ 11 < _8_
2M'

171
8: General Dirichlet series and Bohr's equivalence theorem

and hence

L la(n)le-UoA(n)lel(tA(n)-Jln) - 11 < - 8 L la(n)le-UoA(n) <-.


N . N 8

n= I 2M n= I 2
Thus, the proof will be complete if we can find t and integers k l , ... , k N to
satisfy (4). If the A(n) were linearly independent over the integers we could
apply Kronecker's theorem to A(I), ... A(N) and obtain (4). However the
A(n) are not necessarily independent so instead we apply Kronecker's
theorem to the following system:
f) b f) 2 , . . . , f) Q ' a b a 2, . . . , aQ'
where
f) = {3(n) Yn
n 2nD' an = 2nD·
The {3(n) are the elements of the basis B used to define F(Z), and the Yn are
the imaginary parts of the numbers Zn which determine u. The integers Q and
D are determined as follows. We express A in terms of B by writing
A(n) = rn, 1{3(I) + + rn,q(n){3(q(n)).
Then Q is the largest of the integers q(I), , q(N), and D is the least common
multiple of the denominators of the rational numbers ri, j that arise from the
A(n) appearing in the sum. There are at most q(l) + ... + q(N) such numbers
ri, j. The numbers f)n are linearly independent over the integers because B
is a basis.
By Kronecker's theorem a real t and integers hI, ... , hQ exist such that
b
Itf)k - ak - hkI < 2nDA '

where
N q(n)
A = L L Irn,jl.
n=1 j=l

For this t we have 12nDtf)k - 2nDa k - 2nDh k < bfA, or l

b
It{3(k) - Yk - 2nDh k l < A·
Therefore t{3(k) - Yk = 2nDh k + bk, where Ibkl < b/A. Now we can write
q(n) q(n)

tA(n) - J-ln = t L rn, j{3U) - Lr ll , jYj


j= I j= I
q(n) q(n)
= I rn,it {3(j) - y) = L rn,i2nDhj + b)
j= I j= I
q(n) q(n)
= 2n L hjDrn,j + L bjrn,j
j= I j= 1

172
8.7: Equivalence of general Dirichlet series

where k n is an integer and Ixnl < (bjA) LJ~)1 Irn,jl < b. But this means we
have found a real t and integers kJ, ... , k N to satisfy (4), so the proof is
complete. D

8.7 Equivalence of general Dirichlet series


Consider two general Dirichlet series with the same sequence of exponents
A, say
00 00

L a(n)e - s).(n) and L b(n)e-s).(n).


n=1 n=1
Let B = {f3(n)} be a basis for A and write A = RB, where R is a Bohr matrix.

Definition. We say the two series are equivalent, relative to the basis B, and
we write
W 00
L a(n)e-s).(n) "" L b(n)e-s).(n)
n=1 n=1
if there exists a finite or infinite sequence of real numbers Y = {Yn} such
that
b(n) = a(n)e ixn
where X = {x n } = R Y.

In other words, if we write


q(n)

).(n) = L r n ,kf3(k),
k=1
equivalence means that for some sequence {Yn} we have
q(n) )
b(n) = a(n) exp ( ik'f/n,kYk '

Theorem 8.10. Two equivalent Dirichlet series have the same abscissa of
absolute convergence. Moreover, the relation "" just defned is independent
of the basis B.
PROOF. Equivalence implies Ib(n) I = la(n)1 so the series have the same
abscissa of absolute convergence.
Now let Band r be two bases for A, and assume that two series are equiv-
alent with respect to B. We will show that they are also equivalent with
respect to r.
Write A = RBB. Then there is a sequence Y = {Yn} such that b(n) =
a(n)e ixn , where X = {x n } = R B Y. Now write A = Rrr. If we show that for
some sequence V = {v n } we have X = R r V then the two series will be
equivalent relative to r. The sequence
V = AY

173
8: General Dirichlet series and Bohr's equivalence theorem

has this property, where A is the Bohr matrix such that r = AB. In fact,
we have R r V = RrA Y = R B Y = X, since RrA = R B . This completes the
proof. D

Theorem 8.11. The relation ~ defined in the foregoing definition is an equiv-


alence relation. That is, it is reflective, symmetric, and transitive.
PROOF. Every series is equivalent to itself since we may take each Yn = O.
The corresponding X n will then be zero.
If b(n) = a(n)e ixn then a(n) = b(n)e- ixn . Since X = R B Y we have -X =
R B ( - Y) so the relation is symmetric.
To prove transitivity we may use the same basis throughout and assume
that b(n) = a(n)e ixn , where X = R B Y for some 1': and that a(n) = c(n)e iun ,
where U = R B V for some V Then b(n) = c(n)ei(Xn+u n) where
X + U = RB Y + RB V = R B( Y + V).
This completes the proof. D

8.8 Equivalence of ordinary Dirichlet series


Theorem 8.12. Two ordinary Dirichlet series

f a(~) and f b(~)n


n= 1 n n= 1

are equivalent if, and only ~f, there exists a completely multiplicative function
f such that
(a) b(n) = a(n)f(n) for all n ~ 1, and
°
(b) I f(P)1 = 1 whenever a(n) i= and p is a prime divisor ofn.
PROOF. For ordinary Dirichlet series the sequence of exponents A = {A(n)}
is {log n} and for a basis we may use the sequence B = {log Pn}, where Pn
denotes the nth prime. In fact, if we use the prime-power decomposition

(5) n= n
k=l
00

pkan,k

where each exponent an, k ~ 0, we he ~Te


00

log n = I an,k log Pk,


k=l

so the integer powers may be used as entries in the Bohr matrix R B for which
A = RBB. In the sum and product only a finite number of the an,k are
nonzero.
We note that, because of the fundamental theorem of arithmetic, the
numbers an, k defined by (5) have the property
(6)

174
8.8: Equivalence of ordinary Dirichlet series

Now let A(s) = L a(n)n- S , B(s) = I b(n)n- S • Suppose that A(s) '" B(s).
Then there exists a real sequence {Yk} such that

(7) b(n) = a(n)eXP{iJlan,kYk}

where the integers an, k are determined by equation (5). Define a function f
by the equation

f(n) = eXP{iJl an ,kYk}

Property (6) implies that f(mn) = f(m)f(n) for all m and n, so f is completely
multiplicative. Equation (7) states that b(n) = a(n)f(n), and the definition of
f shows that I f(n) I = 1 for all n, so conditions (a) and (b) of the theorem are
satisfied.
Now we prove the converse. Assume there exists a completely multi-
plicative function f satisfying conditions (a) and (b). We must show that
there is a real sequence {Yk} satisfying (7) for all n. First we consider those n
for which a(n) = O. Property (a) implies b(n) = 0, so equation (7) holds for
such n since both sides are zero no matter how we choose the real numbers
Yk. We shall now construct the sequence {Yk} so that equation (7) also holds
for those n for which a(n) =1= O.
Assume, then, that n is such that a(n) =1= O. We use the prime-power
decomposition (5) and the completely multiplicative property of f to write

n g(n, k),
00

(8) f(n) =
k=l

where

g(n, k) = {f(pk)an'k if Pk In,


1 otherwIse.
Condition (b) implies that I f(Pk) I = 1 for each prime divisor Pk ofn. Therefore
for such primes we may write

where Yk = argf(Pk)· The real numbers Yk have been defined for those k such
that the prime Pk divides some n with a(n) =1= O. For the remaining k (if any)
we define Yk = o. Thus, Yk is well-defined for every integer k ~ 1 and we
have
g(n, k) = exp(ia n , kYk)
for every k ~ 1. Equation (8) now becomes

f(n) = expL~l an,kYk}


175
8: General Dirichlet series and Bohr's equivalence theorem

This, together with property (a), shows that (7) holds for those n for which
a(n) i= O. Thus, (7) holds for all n so A(s) '" B(s). This completes the proof of
the theorem. D

8.9 Equality of the sets U/((Jo) and Ug((Jo) for


equivalent Dirichlet series
Theorem8.I3. Let .f(s) and g(s) be equivalent general Dirichlet series, each of
which converges absolutely for (J = (Jo. Then
Uf((Jo) = Ug((Jo).

PROOF. Let B = {f3(n)} be a basis for the sequence A of exponents. If


f(s) = L a(n)e-SA(n) and g(s) = L b(n)e-SA(n) then there is a real sequence
{Yk} such that

q(n) }
b(n) = a(n) exp{ -ik~/n.kYk .

The Bohr series of f and g are given by


00 {q(n)}
F(Zt,Z2"")=n~ta(n)exp -k~/n.kZk

and
ex; {q(n)}
G(Zt,Z2"")=n~tb(n)exp -k~/n.kZk'

Expressing the b(n) in terms of the a(n) we find

G(Zt, Z2"") = n~1 a(n) exp { - k~>n.k(Zk + iYk)} = F(zl + iYI' Z2 + iY2'" .).

Since the real part of Zn + iYn is the real part of Zn, both series take the same
set of values on the lines X n = (Jo f3(n). Hence Uf((JO) = Ug((Jo), as asserted.
D

8.10 The set of values taken by a Dirichlet


series in a neighborhood of the
line (J = (Jo
Definition. Let f(s) be a general Dirichlet series which converges absolutely
for (J > (J a. Given lJ > 0 and (J 0 such that (J 0 - lJ > (J a' we define the
set Wf ( (J 0; lJ) as follows:
Wf((Jo;lJ) = {f(s):(Jo -lJ < (J < (Jo + lJ, -00 < t < +oo}.

176
8.10: The set of values taken by a Dirichlet series in a neighborhood of the line (J = (J 0

That is, Wf(O"o; £5) is the set of values taken by f(s) in the strip
0"0 - £5 < 0" < 0"0 + £5.
Also, if 0"0 > 0"a we define
Wf(O"o) = n
o <c5<Q'o-Q'a
Wf(O"o; b).

Thus, Wf(O" 0) is the intersection of the sets of values taken by f(s) in all
such strips.

It is clear that Vf(O" 0) S; Wf(O" 0) since every value taken by f(s) on the line
0" =
0"0 is also taken in every strip containing this line. Of course, it may
happen that Vf(O"o) = Wf(O"o) or that Vf(O"o) =1= Wf(O"o).
In general, we have:

Theorem 8.14. Vf(O"o) S; Wf(O"o) S; Vf(O"o), hence Vf(O"o) = Wf(O"o).


PROOF. We remark that this proof is entirely function-theoretic and has
nothing to do with the concept of a basis.
We are to prove that every point in Wf (0"0) is in the closure of l/f( 0"0).
We will show that if WE Wf(O"o) then W is an adherent point of Vf(O"o). In
fact, we will prove that
W = lim f(O"o + it n)
n-oo

for some real sequence {t n }.


Since W E Wf ( 0"0) this means that W E Wf ( 0"0; £5) for all £5 > 0 such that
£5 < 0"0 - O"a. In particular, WE Wf(O"o; lin) for all n ~ no for some no.
This means that for n ~ no we have W = f(sn) where Sn = O"n + itn and
0"0 - (lin) < O"n < 0"0 + (lin). Using the numbers t n so determined, consider
the difference
~v - f(O"o + itn) = f(O"n + it n) - f(O"o + it n)
where n ~ no. We shall express this difference in terms of the derivative
f'(s). Now just as in the case of ordinary Dirichlet series, the function f(s)
defined by
ex:,
f(s) = L a(n)e-SA(n)
n=l

is analytic within its half-plane of absolute convergence. In fact in the proof


of Theorem 8.1 we showed that the series converges uniformly on every
compact subset of the half-plane 0" > O"c. Therefore the sum is analytic in
the half-plane 0" > O"c. Moreover, we can calculate the derivative f'(s) by
term-by-term differentiation, so
00

f'(s) = - L a(n)A(n)e-SA(n).
n=l

177
8: General Dirichlet series and Bohr's equivalence theorem

Hence if (J ~ (J 0 then s is the half-plane of absolute convergence and we get


00 00

If'(s)1 s L la(n)IIA(n)le-UA(n) = L la(n)le-uO'A(n)IA(n)le-(tr-uO')A(n)


n=1 n=1
where (Ja < (Jo' < (Jo. Now IA(n)le-(tr-tro')A(n) ~ 0 as n ~ 00 so, in particular,
this factor is less than 1 for large enough n. Hence
00

If'(s)1 s L la(n)le-tro'A(n). K
n=1
for some K, which shows that I f'(s) I is uniformly bounded in the region
(J ~ (Jo'. Let (Jo' = (Jo - 11no and let M be an upper bound for I f'(s) I in the
region (J ~ (Jo'. Then

Iw - f((1o + itn) I = I f((1n + itn) - f((1o + itn) I = If:" 1'((1 + itn) d(1!
M
S MI(Jn - (Jol s-
n

if n ~ no. Hence lim n -+ oo f((Jo + it n) = w, so w is an adherent point of


Vf( (J 0). This completes the proof. D

8.11 Bohr's equivalence theorem


We have just shown that Wf((Jo) ~ Vf((Jo). The next theorem shows that this
inclusion is actually equality.

Theorem 8.15. We have

The proof of Theorem 8.15 is lengthy and appears in Section 8.12. In


this section we show how Theorem 8.15 leads to Bohr's equivalence theorem.

Theorem 8.16 (Bohr's equivalence theorem). Let f and g be equivalent


Dirichlet series with abscissa of absolute convergence (Ja. Then in any
open half plane (J > C 1 ~ (Ja the functions f(s) and g(s) take the same set
ofvalues.
PROOF. Let S f((J 1) be the set of values taken by f(s) in the half-plane (J > (J 1.
Then
S f( (J 1) = U Vf ((J0)·
UO>U1
Now we prove that
S f((J 1) = U Wr((Jo)·
UO>U1
178
8.12: Proof of Theorem 8.15

First of all, we have Sf(0"1) ~ Ut1 0> t1 1 Wf(O"o) because Vf(O"o) ~ Wf(O"o). To
get inclusion in the other direction, assume WE Ut1 0> t1 1 Wf(O"o). Then
w E Wf ( 0"0) for some 0"0 > 0" 1. Hence W E Wf ( 0"0; b) for all <5 satisfying
o < b < 0"0 - O"a. In other words, f(s) takes the value w in every strip
0"0 - b < 0" < 0"0 + b if 0 < b < 0"0 - O"a. In particular, when b = 0"0 - O"b
we have 0"0 - <5 = 0"1 so f(s) = w for some s with 0" > 0"1. Hence WE Sf(0"1).
This proves that Ut10> t1 1 Wf(O"o) ~ Sf(0"1), so the two sets are equal. Therefore,
we also have
Sg(O" 1) = U ~(O"0)·
t10>t11

To prove Bohr's theorem it suffices to prove that


Wf(O"o) = ~(O"o)

whenever f and g are equivalent. But f '" g implies


Uf(O" 0) = U g(O" 0).

Hence Uf(O"o) = Ug(O"o). But, in view of Theorem 8.9, this means

Vf(O"o) = Vg(O"o)·

But Theorem 8.15 states that Vf(O"o) = Wf(O"o) and ~(O"o) = ~(O"o) so Bohr's
equivalence theorem is a consequence of Theorem 8.15. D

8.12 Proof of Theorem 8.15


To complete the proof of Bohr's equivalence theorem we need to prove
Theorem 8.15, which means we must establish the inclusion relation.
(9)
The proof of (9) makes use of two important theorems of analysis which we
state as lemmas:

Lemma 1 (Helly selection principle). Let {em, n} be a double sequence of real


numbers which is bounded, say
lem,nl < A for all m, n.
Then there exists a subsequence of integers nl < n z < ... with nr ~ 00
as r ~ 00, and a sequence {en} of real numbers such that for every
m = 1,2, ... , we have

lim em, n r = em.


r-+ 00

Note. The important point is that one subsequence {n k } works for every m.
To show the true import of the Lemma, let us see what we can deduce in a
trivial fashion. Display the double sequence as an infinite matrix. Consider

179
8: General Dirichlet series and Bohr's equivalence theorem

the first row: {0l,n}:==1. This is a bounded infinite sequence so it has an


accumulation point, say 01. Hence there is a subsequence {n r} such that
limr-+(X) 01,nr = 01. Similarly, for the second row there is an accumulation
point O2 and a subsequence nr' such that lim n -+ oo 02,nr' = 02, and so on. The
subsequence {n/} needed for O2 may be quite different from that needed for
01. Helly's principle says that one subsequence works simultaneously for all
rows.

PROOF OF LEMMA °
1. Let 1 be an accumulation point of the first row and
suppose the subsequence {n r (1)} has the property that

lim 0l,nr(l) = 01.


r-+ 00
In the second row, consider only those entries 02. n r (1). This is a bounded
°
sequence which has a convergent subsequence with limit 2 , say. Thus,
lim 02,n r(2) = O2
r-+ 00
where {n r(2)} is a subsequence of {n r(1)}. Repeat the process indefinitely. At
the mth step we have a subsequence {nr(m)} which is a subsequence of all
earlier subsequences and a number Om such that
lim Om, nr(m) = Om.
r-+ 00
Now define a sequence {n r } by the diagonal process:

That is, n1 is the first integer used in the first row, n2 the second integer used
in the second row, etc. Look at the mth row and consider the sequence
{Om,n r }. We assert that

r-+ 00
Since nr = nr(r), after the mth term in this row we have r > m so every integer
nr(r) occurs in the subsequence n/ m), so from this point on {n r} is a sub-
sequence of {nr(m)} hence 0m,n r ~ Om' as asserted. D

Lemma 2 (Rouche's theorem). Given two functions f(z) and g(z) analytic
inside and on a closed circular contour C. Assume
Ig(z)1 < If(z)1 on C.
Then f(z) and f(z) + g(z) have the same number of zeros inside C.
PROOF OF LEMMA 2. Let m=inf{lf(z)I-lg(z)l:zEC}. Then m>O
because C is compact and the difference If(z) I - Ig(z) I is a continuous
function on C. Hence for all real t in the interval 0 :::; t :::; 1 we have
If(z) + tg(z) I ~ If(z)1 - Itg(z) 1~ If(z)1 - Ig(z)1 ~ m > O.

180
8.12: Proof of Theorem 8.15

If 0 ~ t ~ 1 define a number cp(t) by the equation

cp(t) = _1 [f'(z) + tg'(z) dz.


2ni Jc f(z) + tg(z)
This number cp(t) is an integer, the number of zeros minus the number of
poles of the function f(z) + tg(z) inside C. But there are no poles, so cp(t) is
the number of zeros of f(z) + tg(z) inside C. But cp(t) is a continuous function
of t on [0, 1]. Since it is an integer, it is constant: cp(O) = cp(I). But cp(O) is the
number of zeros of f(z), and cp(l) is the number of zeros of f(z) + g(z). This
proves Rouche's theorem. 0

PROOF OF RELATION (9). Vf(O"o) £ Wf(O"o). Assume v E Vf(O"o). Then either


v E Vf(O"o) or v is an accumulation point of Vf(O"o). If v E Vf(O"o) then v E Wf(O"o)
since Vf(O"o) £ Wf(O"o). Hence we can assume that v is an accumulation point
of Vf(O"o), and v $ Vf(O"o). This means there is a sequence {t n } of real numbers
such that
v = lim f(O"o
n-+ 00
+ itn)'
We wish to prove that v E Wf ( 0" 0)' This means we must show that v E Wf ( 0" 0; b)
for every b satisfying 0 < b < 0"0 - O"a' In other words, if 0 < b < 0"0 - O"a
we must find an s = 0" + it in the strip
0"0 - b< 0" < 0"0 +b
such that f(s) = v. Therefore we are to exhibit an s in this strip such that
f(s) = lim f(O"o + itn)'
n-+ 00

Let us examine the numbers f(O"o + it m) for the given sequence {t n}. We
have
00

f(O"o + it m) = L a(n)e-tTOA(n). e-itmA(n).


n= 1

The products tmA(n) form a double sequence. There exists a double sequence
of real numbers 8n ,m such that
8n,m = tmA(n) + 2nk m,n' with 0 ~ 8n,m < 2n,
where km,n is an integer. If we replace tmA(n) by 8n,m in the series we don't
alter the terms, hence
00

f(O"o + it m) = L a(n)e-tToA(n)e-iOn,m.
n=l

By Lemma 1, there is a subsequence of integers {n r } and a sequence of real


numbers {8 m } such that
(10) lim 8m,nr
r-+ 00
= Om'

181
8: General Dirichlet series and Bohr's equivalence theorem

Use this sequence {Om} to form a new Dirichlet series


00

g(s) = L b(n)e-s).(n)
n=1
where

This has the same abscissa of absolute convergence as f(s). Now consider
the following sequence of functions:

f,.(s) = f(s + it n)
where {n r } is the subsequence for which (10) holds. We assert that

(a) f,.(s) -+ g(s) uniformly in the strip 0'0 - b < a < 0'0 + b, hence, in
particular, in the circular disk Is - 0'0 I < b.
(b) g(O'o) = v.
(c) There is a d, 0 < d < b, and an R such that fR(S) - v and g(s) - v have
the same number of zeros in the open disk Is - 0'01 < d.

If we prove (b) and (c) then fR(S) - v has at least one zero in the disk because
g(O'o) = v. But fR(S) = f(s + it nR ) and s + it nR is in the strip if s is in the disk,
so this proves the theorem. Now we prove (a), (b) and (c).

Proof of (a). We have

If,.(s) - g(s) I = IJla(n)e-SA(nl(e-i8n,n, - e- i8n )1

00

s L la(n)le-tr).(n)le-iOn,nr - e-iOnl
n=1
N
S L la(n)le-(tro-b»).(n)le-iOn,nr - e-iOnl
n=1
00

+2 L la(n)le-(tro-b»).(n).
n=N+ 1
Now if 8 > 0 is given there is a number N = N(8) such that
00 8
2 L la(n)le-(tro-b»).(n) < -,
n=N+ 1 2
because the series L:= 1 Ia(n) Ie-(tro-b»).(n) converges. In the finite sum L~= 1
we use the inequality

182
8.12: Proof of Theorem 8.15

to write

le-iOn,n r - e-iOnl <


- 10n, n - 0nI.
r

But if M(b) = 1 + L~=l la(n)le-(tTO-b)A(n), there is an integer ro = ro(8)


such that for every n = 1, 2, ... , N we have
8
/0n, nr - 0nI < - -
2M(b) if r ~ ro.

Therefore, if r ~ ro we have
8
I/"(s) - g(s) I ~ -- L la(n)/e-(tTO-b)A(n) + -8
N 8
< -
8
+- = 8.
2M( b) n= 1 2 2 2
Since ro depends only on 8 and on b this shows that /,.(s) ~ g(s) uniformly
in the strip 0'0 - b < a < 0'0 + b as r ~ 00. This proves (a).
Proof of (b). We use (a) to write
g(ao) = lim};.(a o) = limf(ao + itnJ = v.
r-oo r-oo

Proof of (c). Assume first that g is not constant. Since g(a0) = v there is a
positive d < b such that g(s) =I v on the circle
Is - 0'01 = d}.
C = {s:
Let M be the minimum value of Ig(s) - v I on C. Then M > O. Now choose R
so large that I fR(s) - g(s) I < M on C. This is possible by uniform convergence
of the sequence {fR(S)}, since the circle C lies within the strip 1(1 - 0'0 I < b.
Then on C we have
/fR(s) - g(s) I < M ~ Ig(s) - vi.
If G(s) = fR(s) - g(s) and F(s) = g(s) - v we have / G(s) / < /F(s) / on C with
F(s), G(s) analytic inside C. Therefore, by Rouche's theorem the functions
F(s) + G(s) and F(s) have the same number of zeros inside C. But F(s) + G(s)
= fR(s) - v, so fR(S) - v has the same number of zeros inside C as g(s) - v.
Now g(ao) = v so g(s) - v has at least one zero inside C. Hence fR(s) - v
has at least one zero inside C. As noted earlier, this completes the proof if g
is not constant.
To complete the proof we must consider the possibility that g(s) is constant
in the half-plane of absolute convergence. Then g'(s) = 0 for all s in this half-
plane, which means
00

g'(s) = - L A(n)b(n)e-SA(n) = O.
n= 1

But as in the case of ordinary Dirichlet series, if a general Dirichlet series


has the value 0 for a sequence of values of s with real parts tending to + 00
then all the coefficients must be zero. (See [4], Theorem 11.3.) Hence

183
8: General Dirichlet series and Bohr's equivalence theorem

A(n)b(n) = 0 for all n 2 1. Therefore b(n) = 0 with at most one exception,


say b(n 1 ), in which case A(n 1 ) = O. Therefore, since a(n) = b(n)e iOn , we must
have a(n) = 0 with at most one exception, say a(nl), and then A(nl) = O.
Hence the series for f (s) consists of only one term, f (s) = a(n l)e - sA(nd = a(n 1),
so f(s) itself is constant. But in this case the theorem holds trivially. D

8.13 Examples of equivalent Dirichlet series.


Applications of Bohr's theorem to
L-series
Theorem 8.17. Let k 2 1 be a given integer, and let Xbe any Dirichlet character
modulo k. Let L:=1 a(n)n - S be any Dirichlet series whose coefficients
have the following property:
a(n) -:1= 0 implies (n, k) = 1.
Then
i a(7) ~ i a(n)~(n).
n= 1 n n= 1 n
PROOF. Since these are ordinary Dirichlet series we may use "Theorem 8.12
to establish the equivalence. In this case we take f(n) = x(n). Then f is
completely multiplicative and condition (a) is satisfied. Now we show that
condition (b) is satisfied. We need to show that I f(p) I = 1 if a(n) -:1= 0 and
pIn. But a(n) -:1= 0 implies (n, k) = 1. Since pin we must have (p, k) = 1 so
If (P) I = IX(p) I = 1 since X is a character. Therefore the two series are
equivalent. D

Theorem 8.18. For a given modulus k, let ·Xl' ... , Xq>(k) denote the Dirichlet
characters modulo k. Then in any half-plane of the form (J > (J 1 2 1 the
set of values taken by the Dirichlet L-series L(s, X;) is independent of i.
PROOF. Applying the previous theorem with a(n) = Xl(n) we have
f Xl~n) ~ i Xl(n);(n)
n=l n n=l n
for every character X modulo k. Here we use the fact that Xl (n) -:1= 0 implies
(n, k) = 1. Thus each L-series L(s, X) is equivalent to the particular L-series
L(s, Xl). Therefore, by Bohr's theorem, L(s, X) takes the same set of values as
L(s, X1) in any open half-plane within the half-plane of absolute convergence.
D
8.14 Applications of Bohr's theorem to the
Riemann zeta function
OUf applications to the Riemann zeta function require the following identity
involving Liouville's function A(n) which is defined by the relations
A(I) = 1, A(PI al ... Pr ar ) = ( _1)a l + ... +a r •
184
8.14: Applications of Bohr's theorem to the Riemann zeta function

The function )..(n) is completely multiplicative and we have (see [4J, p. 231)

I
n=l
),(n)
nS
= ((2s)
((s)
if (J > 1
.

Theorem 8.19. Let A(n) denote Liouville's function and let

C(x) = L A(n).
n:5x n
Then if (J > 1 we have

((2s) = Joc C(x) dx. S


(s - 1)((s) 1 X

PROOF. By Abel's identity (Theorem 4.2 in [4J) we have

L 2-
A(n) = C(x)
S
+ s JX C(t) dt.
n:5 x n nS X 1
S
t +1

Keep (J > 0 aRd let x ~ 00. Then

C(x)
XS
= o(~ L
x(1 n:Sx
!) = O(IOg x) =
n x(1
0(1) as x~ 00,

so we find
~ A(n) _ x C(t)
~
n= 1 n
S+ 1 - S f1
S+I dt,
t
for (J > O.

Replacing s by s - 1 we get

I
n=l
A(7) = (s - 1)
n I t
foo C~) dt for (J > 1.

Since the series on the left has sum ((2s)/((s) the proof is complete. D

Now we prove a remarkable theorem discovered by P. Tunin [50] in


1948 which gives a surprising connection between the Riemann hypothesis
and the partial sums of the Riemann zeta function in the half-plane (J > 1.

Theorem 8.20. Let


n 1
(n(s) = L -ks '
k=l

If there exists an no such that (n(s) -:1= 0 for all n ~ no and all (J > 1, then
((s) -:1= 0 for (J > -!.
PROOF. First we note that the two Dirichlet series L~ = 1 k - S and L~ = 1 A(k)k - S
are equivalent because ).. is completely multiplicative and has absolute

185
8: General Dirichlet series and Bohr's equivalence theorem

value 1. Therefore~ by Bohr's theorem, (n(s) -=1= 0 for (J > 1 implies that
L~= 1 A(k)k -s -=1= 0 for (J > 1. But for s real we have

lim i
s- + 00 k = 1
,1(~) =
k
,1(1) = 1.

Hence for all real s > 1 we must have L~ = 1 A(k)k - s > O. Letting s ~ 1 +
we find

~ A(k) > 0
i...J if n ~ no·
k=1 k -
In other words, the function

(11 ) C(x) = L ,1(n)


n:S;x n

is nonnegative for x ~ no. Now we use the identity of Theorem 8.19,

valid for (J > 1. Note that the denominator (s - 1)'(s) is nonzero on the
real axis s > t, and '(2s) is finite for real s > t. Therefore, by the integral
analog of Landau's theorem (see Theorem 11.13 in [4]) the function on the
left is analytic everywhere in the half-plane (J > t. This implies that '(s) -=1= 0
for (J > t, and the proof is complete. D

Turan's theorem assumes that the sum C(x) in (11) is nonnegative for all
x ~ no. In 1958, Haselgrove [14] proved, by an ingenious use of machine
computation, that C(x) is negative for infinitely many values of x. Therefore,
Theorem 8.20 cannot be used to prove the Riemann hypothesis. Subse-
quently, Tunin [51] sharpened his theorem by replacing the hypothesis
C(x) ~ 0 by a weaker inequality that cannot be disproved by machine
computation.

Theorem 8.21 (Turan). Let C(x) = Ln:s;x A(n)jn. rr there exist constants
r:t. > 0, C > 0 and no such that

loga x
(12) C(x) > - c - -
fi
for all x ~ no, then the Riemann hypothesis is true.
PROOF. If G > 0 is given there exists an nl ~ no such that c loga x ~ xl: for
all x ~ nl so (12) implies
C(x) > - xl:- 1/2 •

186
Exercises for Chapter 8

Let A(x) = C(x) + xf,- 1/2, where G is fixed and 0 < G < t. Then A(x) > 0
for all x 2 n 1 • Also, for a > 1 we have

OO A(x)
- -x d -
- foo -C(x)
-x d + foo x f,-s-I/2 dX
f 1 X
S
1 x
S
1

'(2s) +s 1
1
= f( S,)
(s - 1)'(s) - 2 - G

say. Arguing as in the proof of Theorem 8.20, we find that the function f(s)
t
is analytic on the real line s > + G. By Landau's theorem it follows that
t
f(s) is analytic in the half-plane a > + G. This implies that '(s) =1= 0 for
a > t + G, hence '(s) =1= 0 for a > t since G can be arbitrarily small. D

Note. Since each function 'n(s) is a Dirichlet series which does not vanish
identically there exists a half-plane (J > 1 + an in which 'n(s) never vanishes.
(See [4], Theorem 11.4.) The exact value of an is not yet known. In his 1948
paper [50] Turan proved that, for all sufficiently large n, 'n(s) =1= 0 in the
half-plane a > 1 + 2(1og log n)jlog n, hence an ~ 2(log log n)jlog n for large
n. In the other direction, H. L. Montgomery has shown that there exists a
constant c > 0 such that for all sufficiently large n, 'n(s) has a zero in the half-
plane a > 1 + c(log log n)jlog n, hence an 2 c(1og log n)jlog n for large n.
The number 1 + an is also equal to the abscissa of convergence of the
Dirichlet series for the reciprocal 1j'n(s). If a > 1 + an we can write

1 _ f JIn(k)
'n(s) - k= 1 ~'
where /In(k) is the Dirichlet inverse of the function un(k) given by
1 if k ~ n,
un(k) = {0 ifk > n.

The usual Mobius function /l(k) is the limiting case of /In(k) as n -+ 00.

Exercises for Chapter 8


1. IfL a(n)e-SA(n) has abscissa of convergence (Jc < 0, prove that
. loglLk=n a(k) I
(J c = lIm sup A .
n--+oo (n)

2. Let (J c and (J a denote the abscissae of convergence and absolute convergence of a


Dirichlet series. Prove that
. log n
o ~ (Ja - (Jc ~ lIm sup A(n)' n--+ 00

This gives 0 ~ (Ja - (Jc ~ 1 for ordinary Dirichlet series.

187
8: General Dirichlet series and Bohr's equivalence theorem

3. If log njA(n) --+ 0 as n --+ 00 prove that


. 10gla(n)1
aa = a c = hm sup .
n-+ ex: A(n)

What does this imply about the radius of convergence of a power series?

4. Let {J-(n)} be a sequence of complex numbers. Let A denote the set of all points
s = a + it for which the series I a(n)e-SA(n) converges absolutely. Prove that A
is convex.

Exercises 5~ 6~ and 7 refer to the seriesf'(s) == L:::= 1 a{n)e-s).(n) with exponents


and coefficients given as follows

n 1 2 3 4 5

A(n) -1 - log 2 -1 -log2 -1 + log 2 0

1 1 1 1
a(n) 8 2" 4 -8
1
2"

n 6 7 8 9 10

A(n) 1 - log 2 log 2 1 log 3 1 + log 2


1 1 1 3 1
a(n) 8 -4 2" -4 -8

Also, a(n + 10) == -1 2- n


and A(n + 10) == (n + 1) log 3 for n ~ 1.
5. Prove that a a = - (log 2)jlog 3.
6. Show that the Bohr function corresponding to the basis B = (1, log 2, log 3) is
1 - 2e- Z3
F(Zb Z2' Z3) = cos(izd - !i sin(iz 2)(1 + cos(izd) + 2
-e
-Z3 '

if X3 > -log 2, Z1' 22 arbitrary.


7. Determine the set Uf(O). Hint: The points -1, 1 + i, 1 - i are significant.
8. Assume the Dirichlet series f(s) = I~= 1 a(n)e-s).(n) converges absolutely for a > aa'
If a > a a prove that

lim -
1 IT,
e (rT+lt)f(a + it) dt =
A {a(n) if A = A(n)
T-+ + y 2T - T 0 if A =1= A(I), A(2), ....
9. Assume the series f(s) = L~= 1 a(n)e-SA(n) converges absolutely for (J > aa > O. Let
v(n) = eA(n).

(a) Prove that the series g(s) = L~= 1 a(n)e-Sv(n) converges absolutely if a > O.
(b) If a > a a prove that

ns)f(s) = fC g(t)t S - 1
dt.

188
Exercises for Chapter 8

This extends the classic formula for the Riemann zeta function,

f(s)((s) = foco ~
e - 1
dt.

Hint: First show that f(s)e-SA(n) = So e-tv(n)t s - 1 dt.

189
Supplement to Chapter 3

Alternate proof of Dedekind' s functional equation


This supplement gives an alternate proof of Dedekind' s functional equation
as stated in Theorem 3.4:

Theorem. If A = (; ~) E r and c > 0, thenfor every T in H we have

(1)
aT
TJ ( CT
+ db) ==
+ {
e(A) - i(CT + d)}112TJ(T) ,
where

(2) e(A) = eXP{1Ti(a 1;Cd - s(d, C»)}


and s(d, c) is a Dedekind sum.

The alternate proof was suggested by Basil Gordon and is based on the fact
that the modular group r
has the two generators TT == T + 1 and ST =
- lIT. In Theorem 2.1 we showed that every A in r
can be expressed in
the form

where the n, are integers. But T = ST-1ST-1S, so every element of also r


has the form STtnlS · · · STtnk for some choice of integers m l , • • • , mk. The
idea of the proof is to show that if the functional equation (1) holds for a

particular transformation A = (~ ~) in r with c > 0 and with e(A) as

specified in (2), then it also holds for the products ATtn and AS for every

190
Supplement to Chapter 3

integer m. (See Lemma 3 below.) In Theorem 3.1 we proved that it holds


for S. Therefore, because every element of r has the form STmtS · · · STmk,
it follows that the functional equation (1) holds for every A with e > o.
The argument is divided into three lemmas that show that the general
functional equation is a consequence of the special case in Theorem 3. 1
together with three basic properties of Dedekind sums derived in Sections
3.7 and 3.8. The first two lemmas relate e(A) with e(AT m) and e(AS), where
T and S are the generators of the modular group.

Lemma 1. If A = (~ ~) E rand e > 0, then for every integer m we


have

PROOF. We have ATtn == b)(1O lm) == (ac am


(ac d em ++ db), so
s(AT
m
) = exp{ 7Ti(a+ ~;:. + d - seem + d, c»)}.
But s(em + d, c) == s(d, e) by Theorem 3.5(a), and hence, we obtain
Lemma 1. 0

Lemma 2. If A (~ ~) E rand e > 0, then we have

e -7Ti/4 e (A) if d > 0,


e(AS) == { e7T1 / 4 e(A) if d < O.
PROOF. We have

AS
== (ac db)(O1 -1)0 == (bd -a)
-c·

If d > 0, we represent the transformation AS by the matrix

AS = (~ -a),
-e
but if d < 0, then - d > 0 and we use the representation AS
-b
( -d
a).
e
For d > 0, we have

(3) seAS) = exp{ 7Ti(b1; / - s( - e, d))}


{7Ti(b1; / + s(e, d))}
= eX P

191
Supplement to Chapter 3

because s( - c, d) = - s(c, d) by Theorem 3.5(a). The reciprocity law for


Dedekind sums implies

c d 1 1
s(c, d) + s(d, c) = 12d + 12c - 4 + 12cd·

We replace the numerator in the last fraction by ad - bc and rearrange


terms to obtain

b-c a+d 1
Ud + s(c, d) = ~ - s(d, c) - 4·
Using this in (3), we find e(AS) = e - 7Ti/4 e(A).

-b
If d < 0, we use the representation AS = ( _ d ;) to obtain

(4)

In this case, - d > 0 and we use the reciprocity law in the form

c d I a d - bc
s(c, - d) + s( - d , c) = _ 12d - 12c - 4 - 12cd·

Rearrange terms and use s( - d, c) = - s(d, c) to obtain

-b+c a+d 1
-12d - s(c, -d) = ~ - s(d, c) + 4·
Using this in (4), we find that e(AS) e 7Ti/ 4 e(A). This completes the proof
of Lemma 2. o
Lemma 3. If Dedekind's functional equation
(5)

is satisfied for some A = (~ ~) in r with c > 0 and seA) given by (2),

then it is also satisfied for AT In and for AS. That is, (5) implies

(6)
and

(7)
whereas

(8)

192
Supplement to Chapter 3

PROOF. Replace T by TinT in (5) to obtain


'Y](ATmT) = e(A){ - i(cTmT + d)}1I2'Y](T mT)
= e(A){ - i(CT + mc + d)}1/2 e1Tim/12'Y](T).

Using Lemma 2 we obtain (6).


Now ~eplace T by ST in (5) to get
'Y](AST) = e(A){ - i(CST + d)}1/2'Y](ST).
Using Theorem 3.1 we can write this as
(9) 'Y](AST) = e(A){ - i(CST + d)}1/2 { - iT}I/2'Y](T).
If d > 0, we write
C dT - C
CST + d = -- + d = - _ .
T T '

hence,
- i(dT - c) 1Til2
- i(CST + d) = .
-IT
e- ,

and therefore, {- i(CST + d)}112 {- iT} 1/2 = e- 1T1 /4 { - i(dT - C)}1I2. Using this
in (9) together with Lemma 3, we obtain (7).
If d < 0, we write
_~ + d __ - dT +C
CST +d= T -T

so that in this case we have


- i( - dT + c) 1Til2
- i(CST + d) = _ iT e

and therefore, {- i(CST + d)}112 {- iT}1I2 = e1Ti/4 { - i( - dT + c)} 112 . Using


this in (9) together with Lemma 3, we obtain (8). D

Remark on the root of unity B(A)


Dedekind's functional equation (1), with an unspecified 24th root of unity
e(A), follows immediately by extracting 24th roots in t~e functional equation
for d(T). Much of the effort in this theory is directed at showing that the
root of unity e(A) has the form given in (2). It is of interest to note that a
simple argument due to Dedekind gives the following theorem:

Theorem. If (1) holds whenever A = (: ~) E rand c #- 0, then

d
B(A) = eXP{7Ti(a};C - f(d. C))}
for some rational number f(d, c) depending only on d and c.

193
Supplement to Chapter 3

PROOF. Let
aT +b , a'T + b'
AT=--- and AT= cT+d
CT + d

be two transformations in r having the same denominator CT + d. Then


ad-bc=1 and a'd - b'c = 1,
so both pairs a, b and a', b' are solutions of the linear Diophantine equation
xd - yc = 1.
Consequently, there is an integer n such that
a' = a + nc, b' = b + nd.

Hence,
(a + nC)T + (b + nd) aT + b A
A'T = CT+
d = ---d
CT+
+ n = T + n.
Therefore, we have
'11(A'T) = '11(AT + n) = e1Tin/12'11(AT) = e1Tin/12e(A){ - i(CT + d)}1/2'11(T),
because of (1). On the other hand, (1) also gives us
'11(A'T) = e(A'){ - i(CT + d)}1I2'11(T).
Comparing the two expressions for '11(A'T), we find e(A') e1Tin/12e(A). But
n = (a' - a)/c, so

tri(a' - a))
e(A') = exp ( 12c e(A),

or

( .')
tria, (. )
exp - 12c e(A) = exp - 12c e(A).
Tria

This shows that the product ex p (- ~~;)e(A) depends only on cand d. There-

fore, the same is true for the product

tri(a + d))
exp( 12c e(A).

This complex number has absolute value 1 and can be written as

tri(a + d))
exp( 12c e(A) = exp( -trif(d, c))

194
Supplement to Chapter 3

for some real number f(d, c) depending only on c and d. Hence,

Because 8 24 = 1, it follows that 12cf(d, c) is an integer, so f(d, c) is a


rational number. D

195
Bibliography

1. Apostol, Tom M. Sets of values taken by Dirichlet's L-series. Proc. Sympos. Pure
Math., Vol. VIII, 133-137. Amer. Math. Soc., Providence, R.I., 1965. MR 31
# 1229.
2. Apostol, Tom M. Calculus, Vol. II, 2nd Edition. John Wiley and Sons, Inc. New
York, 1969.
3. Apostol, Tom M. Mathematical Analysis, 2nd EditioQ. Addison-Wesley Publishing
Co., Reading, Mass., 1974.
4. Apostol, Tom M. Introduction to Analytic Number Theory. Undergraduate Texts
in Mathematics. Springer-Verlag, New York, 1976.
5. Atkin, A. O. L. and O'Brien, J. N. Some properties of p(n) and c(n) modulo powers
of 13. Trans. Amer. Math. Soc. 126 (1967),442-459. MR 35 # 5390.
6. Bohr, Harald. Zur Theorie der allgemeinen Dirichletschen Reihen. Math. Ann.
79 (1919), 136-156.
7. Deligne, P. La conjecture de Weil. I. Inst. haut. Etud sci., Publ. math. 43 (1973),
273-307 (1974). Z. 287, 14001.
8. Erdos, P. A note on Farey series. Quart. J. Math., Oxford Ser. 14 (1943), 82-85.
MR 5, 236b.
9. Ford, Lester R. Fractions. Amer. Math. Monthly 45 (1938), 586-601.
10. Gantmacher, F. R. The Theory of Matrices, Vol. 1. Chelsea Publ. Co., New York,
1959.
11. Gunning, R. C. Lectures on Modular Forms. Annals of Mathematics Studies, No.
48. Princeton Univ. Press, Princeton, New Jersey, 1962. MR 24 # A2664.
12. Gupta, Hansraj. An identity. Res. Bull. Panjab Univ. (N.S.) 15 (1964), 347-349
(1965). MR 32 # 4070.
13. Hardy, G. H. and Ramanujan, S. Asymptotic formulae in combinatory analysis.
Proc. London Math. Soc. (2) 17 (1918),75-115.
14. Haselgrove, C. B. A disproof of a conjecture of P6lya. Mathematika 5 (1958),
141-145. MR 21 #3391.
15. Heeke, E. Uber die Bestimmung Dirichletscher Reihen durch ihre Funktional-
gleichung. Math. Ann. 112 (1936), 664-699.

196
Bibliography

16. Hecke, E. Ober Modulfunktionen und die Dirichlet Reihen mit Eulerscher Produkt-
entwicklung. I. Math. Afll1.114 (1937),1-28; II. 316-351.
17. Iseki, Shoo The transformation formula for the Dedekind modular function and
related functional equations. Duke Math. J. 24 (1957),653-662. MR 19, 943a.
18. Knopp, Marvin I. Modular Functions in Analytic Number Theory. Markham
Mathematics Series, Markham Publishing Co., Chicago, 1970. MR 42 #198.
19. Lehmer, D. H. Ramanujan's function r(n). Duke Math. J. 10 (1943), 483-492.
MR 5, 35b.
20. Lehmer, D. H. Properties of the coefficients of the modular invariant J(r). Amer.
J. Math. 64 (1942), 488-502. MR 3, 272c.
21. Lehmer, D. H. On the Hardy-Ramanujan series for the partition function. J.
London Math. Soc. 12 (1937), 171-176.
22. Lehmer, D. H. On the remainders and convergence of the series for the partition
function. Trans. Amer. Math. Soc. 46 (1939),362-373. MR 1, 69c.
23. Lehner, Joseph. Divisibility properties of the Fourier coefficients of the modular
invariantj(r). Amer. J. Math. 71 (1949),136-148. MR 10, 357a.
24. Lehner, Joseph. Further congruence properties of the Fourier coefficients of the
modular invariarttj(r). Amer. J. Math. 71 (1949), 373-386. MR 10, 357b.
25. Lehner, Joseph, and Newman, Morris. Sums involving Farey fractions. Acta
Arith. 15 (1968/69),181-187. MR 39 # 134.
26. Lehner, Joseph. Lectures on Modular Forms. National Bureau of Standards,
Applied Mathematics Series, 61, Superintendent of Documents, U.S. Government
Printing Office, Washington, D.C., 1969. MR 41 # 8666.
27. LeVeque, William Judson. Reviews in Number Theory, 6 volumes. American Math.
Soc., Providence, Rhode Island, 1974.
28. Mordell, Louis J. On Mr. Ramanujan's empirical expansions of modular functions.
Proc. Cambridge Phil. Soc. /9(1917), 117-124.
29. Neville, Eric H. The structure of Farey series. Proc. London Math. Soc. 51 (1949),
132-144. MR 10, 681f.
30. Newman, Morris. Congruences for the coefficients of modular forms and for the
coefficients ofj(r). Proc. Amer. Math. Soc. 9 (1958),609-612. MR 20 # 5184.
31. Petersson, Hans. Uber die Entwicklungskoeffizienten der automorphen formen.
Acta Math. 58 (1932),169-215.
32. Petersson, Hans. Ober eine Metrisierung der ganzen Modulformen. Jher. Deutsche
Math. 49 (1939),49-75.
33. Petersson, H. Konstruktion der samtlichen L6sungen einer Riemannscher
Funktionalgleichung durch Dirichletreihen mit Eulersche Produktenwicklung. I.
Math. Ann. 116 (1939), 401-412. Z. 21, p. 22; II. 117 (1939),39-64. Z. 22, 129.
34. Rademacher, Hans. Ober die Erzeugenden von Kongruenzuntergruppen der
Modulgruppe. Abh. Math. Seminar Hamburg, 7 (1929), 134-148.
35. Rademacher, Hans. Zur Theorie der Modulfunktionen. J. Reine Angew. Math.
167 (1932), 312-336.
36. Rademacher, Hans. On the partition function pen). Proc. London Math. Soc. (2)
43 (1937), 241-254.
37. Rademacher, Hans. The Fourier coefficients of the modular invariant j( r). Amer.
J. Math. 60 (1938), 501-512.
38. Rademacher, Hans. On the expansion of the partition function in a series. Ann. of
Math. (2) 44 (1943),416-422. MR 5, 35a.
197
Bibliography

39. Rademacher, Hans. Topics in Analytic Number Theory. Die Grundlehren der
mathematischen Wissenschaften, Bd. 169, Springer-Verlag, New York-Heidel-
berg-Berlin, 1973. Z. 253.10002.
40. Rademacher, Hans and Grosswald, E. Dedekind Sums. Carus Mathematical
Monograph, 16. Mathematical Association of America, 1972. Z. 251. 10020.
41. Rademacher, Hans and Whiteman, Albert Leon. Theorems on Dedekind sums.
Amer. J. Math. 63 (1941),377-407. MR 2, 249f.
42. Rankin, Robert A. Modular Forms and Functions. Cambridge University Press,
Cambridge, Mass., 1977. MR 58 #16518.
43. Riemann, Bernhard. Gessamelte Mathematische Werke. B. G. Teubner, Leipzig,
1892. ErHiuterungen zu den Fragmenten XXVIII. Von R. Dedekind, pp. 466-
478.
44. Schoeneberg, Bruno. Elliptic Modular Functions. Die Grundlehren der mathema-
tischen Wissenschaften in Einzeldarstellungen, Bd. 203, Springer-Verlag, New
York-Heidelberg-Berlin, 1974. MR 54 #236.
45. Sczech, R. Ein einfacher Beweis der Transformationsforme I fur log T1(Z). Math.
Ann. 237 (1978), 161-166. MR 58 #21948.
46. Selberg, Atle. On the estimation of coefficients of modular forms. Proc. Sympos.
Pure Math., Vol. VIII, pp. 1-15. Amer. Math. Soc., Providence, R.I., 1965. MR 32
#93.
47. Serre, Jean-Pierre. A Course in Arithmetic. Graduate Texts in Mathematics, 7.
Springer-Verlag, New York-Heidelberg-Berlin, 1973.
48. Siegel, Carl Ludwig. A simple proofofl1( -l/r) == l1{r)JTTi. Mathematika 1 (1954),
4. MR 16, 16b.
49. Titchmarsh, E. C./ntroduction to the Theory ofFourier Integrals. Oxford, Clarendon
Press, 1937.
50. Tunin, Paul. On some approximative Dirichlet polynomials in the theory of the
zeta-function of Riemann. Danske Vid. Selsk. Mat.-Fys. Medd. 24 (1948), no. 17,
36 pp. MR 10, 286b.
51. Tunin, Paul. Nachtrag zu meiner Abhandlung •. On some approximative Dirichlet
polynomials in the theory of the zeta-function of Riemann." Acta Math. A cad.
Sci. Hungar. 10 (1959), 277-298. MR 22 # 6774.
52. Uspensky, J. V. Asymptotic formulae for numerical functions which occur in the
theory of partitions [Russian]. Bull. Acad. Sci. URSS (6) 14 (1920),199-218.
53. Watson, G. N. A Treatise on the Theory ofBessel Functions, 2nd Edition. Cambridge
University Press, Cambridge, 1962.

198
Index of special symbols

n(W t , W2) lattice generated by Wt and W2, 2


p(z) Weierstrass p-function, 10
Gn Eisenstein series of order n, n ~ 3, 12
G2 Eisenstein series of order 2, 69
g2,g3 invariants, 12
eb e 2,e3 values of f.J at the half-periods, 13
~(Wb w 2), ~(r) discriminant g~ - 27 g~ , 14
H upper half-plane Im('r) > 0, 14
J(r) Klein's modular function g~/~, 15
r(n) Ramanujan tau function, 20
Gcz(n) sum of the ath powers of divisors of n, 20
r modular group, 28
S, T generators of r, 28
RG fundamental region of sub-group G of r, 30
R fundamental region of r, 31
l1(r) Dedekind eta function, 47
s(h, k) Dedekind sum, 52
A(X) -log(1 - e- 21tx ), 52
1\((1, {3, z) Iseki's function, 53
'(s, a) Hurwitz zeta function, 55
F(x, s) periodic zeta function, 55
j(r) 12 3J(r), 74
ro(q) congruence subgroup of r, 75

lP-l (, + A)
fir) -If-, 80
P l=O P
('~(q')r(q-l)
<1>( r) 86
~(r) ,
9(r) Jacobi theta function, 91
p(n) partition function, 94

199
Index of special symbols

F(x) generating function for p(n), 94


Fn set of F arey fractions of order n, 98
Mk linear space of entire forms of weight k, 117
Mk,o subspace of cusp forms of weight k, 119
~ Hecke operator, 120
r(n) set of transformations of order n, 122
K dim M 2k,O, 133
E 2k (r) normalized Eisenstein series, 139
F(Z) Bohr function associated with Dirichlet series, 168
Vf(O"O) set of values taken by Dirichlet series f(s) on line 0" = 0"0, 170
'n(S) partial sums L k- s , 185
kSn

200
Index

A Bohr, equivalence theorem, 178


function of a Dirichlet series, 168
Abscissa, of absolute convergence, 165
matrix, 167
of convergence, 165
Additive number theorv, 1
Apostol, Tom M., 196
Approximation theorem, of Dirichlet, 143
c
of Kronecker, 148, 150, 154 Circle method, 96
of Liouville, 146 Class number of quadratic form, 45
Asymptotic formula for p(n), 94, 104 Congruence properties, of coefficients of
Atkin, A. O. L., 91,196 j(r), 22, 90
Automorphic function, 79 of Dedekind sums, 64
Congruence subgroup, 75
Cusp form, 114

B
Basis for sequence of exponents, 166 D
Bernoulli numbers, 132 Davenport, Harold, 136
Bernoulli polynomials, 54 Dedekind, Richard, 47
Berwick, W. E. H., 22 Dedekind function y/(r), 47
Bessel functions, 109 Dedekind sums, 52, 61
Bohr, Harald, 161, 196 Deligne, Pierre, 136, 140, 196

201
Index

Differential equation for p(z), 11


Dirichlet, Peter Gustav Lejeune, 143
Dirichlet's approximation theorem, 143 G
Dirichlet L-function, 184 92,93,12
Dirichlet series, 161 General Dirichlet series, 161
Discriminant L\(r), 14 Generators, of modular group r, 28
Divisor functions a:x(n), 20 of congruence subgroup r o(p), 78
Doubly periodic functions, 2 Grosswald, Emil, 61, 198
Gupta, Hansraj, Ill, 196

E
e t , e2 , e3 , 13 H
Eigenvalues of Hecke operators, 129 Half-plane H, 14
Eisenstein series Gn' 12 Half-plane, of absolute convergence, 165
recursion formula for, 13 of convergence, 165
Elliptic functions, 4 Hardy, Godfrey Harold, 94, 196
Entire modular forms, 114 Hardy-Ramanujan formula for p(n), 94
Equivalence, of general Dirichlet series, Haselgrove, C. B., 186, 196
173 Hecke, Erich, 114, 120, 133, 196, 197
of ordinary Dirichlet series, 174 Hecke operators Tn' 120
of pairs of periods, 4 HeIly, Eduard, 179
of points in the upper half-plane H, 30 Helly selection principle, 179
of quadratic forms, 45 Hurwitz, Adolf, 55, 145
Estimates for coefficients of modular Hurwitz approxiniation theorem, 145
forms, 134 Hurwitz zeta function, 55, 71
Euler, Leonhard, 94
Euler products of Dirichlet series, 136
I
Exponents of a general Dirichlet
series, 161 Invariants 92, 93' 12
Inversion problem for Eisenstein
series, 42
F Iseki, Sho, 52, 197
Farey fractions, 98 Iseki's transformation formula, 53
Ford, L. R., 99, 196
Ford circles, 99
Fourier coefficients of j(r), 21,74 J
divisibility properties of, 22, 74, 91 j(r), J(r), 74,15
Functional equation, for r,( r), 48, 52 Fourier coefficients of, 21
for 9(r), 91 Jacobi, Carl Gustav Jacob, 6, 91, 141
for '(s), 140 Jacobi theta function, 91, 141.
for A(ex, {3, z), 54 Jacobi triple product identity, 91
for <I>(ex, {3, s), 56, 71
Fundamental pairs of periods, 2
Fundamental region, of modular group K
r,31 Klein, Felix, IS
of subgroup r o(P), 76 Klein modular invariant J(r), 15

202
Index

Kloosterman, H. D., 136


Knopp, Marvin I., 197
Kronecker, Leopold, 148 N
Kronecker approximation theorem, 148, Neville, Eric Harold, 110, 197
150,154 Newman, Morris, 91, Ill, 197
Normalized eigenform, 130

L
o
Lambert, Johann Heinrich, 24
O'Brien, J. N., 91, 196
Lambert series, 24
Order of an elliptic function, 6
Landau, Edmund, 186
Lehmer, Derrick Henry, 22, 93, 95, 197
Lehmer conjecture, 22
Lehner, Joseph, 22, 91, Ill,} 97
p
LeVeque, William Judson, 197 p-function of Weierstrass, 10
Linear space M k of entire forms, 118 Partition function p(n), 1, 94
Linear subspace Mk,o of cusp forms, 119 Period, 1
Liouville, Joseph, 5, 146, 184 Period parallelogram, 2
Liouville approximation theorem, 146 Periodic zeta function, 55
Liouville function ).. (n), 25, 184 Petersson, Hans, 22, 133, 140, 197
Liouville numbers, 147 Petersson inner product, 133
Littlewood, John Edensor, 95 Petersson-Ramanujan conjecture, 140
Picard, Charles Emile, 43
Picard's theorem, 43
Product representation for L\(r), 51
M
Mapping properties of J(r), 40
Mediant,98 Q
Mellin, Robert Hjalmar, 54 Quadratic forms, 45
Mellin inversion formula, 54
Mobius, Augustus Ferdinand, 24, 27, 187
Mobius function, 24, 187 R
Mobius transformation, 27 Rademacher, Hans, 22, 62, 95, 102,
Modular forms, 114 104, 197
and Dirichlet series, 136 Rademacher path of integration, 102
Modular function, 34 Rademacher series for p(n), 104
Modular group r, 28 Ramanujan, Srinivasa, 20, 92, 94,
subgroups of, 46, 75 136,191
Montgomery, H. L., 187 Ramanujan conjecture, 136
Mordell, Louis Joel, 92, 197 Ramanujan tau function, 20, 22, 92,
Multiplicative property, of coefficients of 113, 131, 198
entire forms, 130 Rankin, Robert A., 136, 198
of Hecke operators, 126, 127 Reciprocity law for Dedekind sums, 62
of Ramanujan tau function, 93, 114 Representative of quadratic form, 45

203
Index

Riemann, Georg Friedrich Bernhard,


140, 155, 185, 198
Riemann zeta function, 20, 140, 155,
w
]85,189 Watson, G. N., 109, 198
Rouche, Eugene, 180 Weierstrass, Karl, 6
Rouche's theorem, 180 Weierstrass p-function, 10
Weight of a modular form, 114
Weight formula for zeros of an entire
s form, 115
Whiteman, Albert Leon, 62, 198
Salie, Hans, 136
Schoeneberg, Bruno, 198
Sczech, R., 61, 198 z
Selberg, Atle, 136, 198
Zeros, of an elliptic function, 5
Serre, Jean-Pierre, 198
Zeta function, Hurwitz, 55
Siegel, Carl Ludwig, 48, 198
periodic, 55
Simultaneous eigenforms, 130
Riemann, 140, 155, 185,189
Spitzenform, 114
Zuckerman, Herbert S., 22
Subgroups of the modular groups, 46, 75

T
Tau function, 20, 22, 92, 113, 131
Theta function, 91,141
Transcendental numbers, 147
Transformation of order n, 122
Transformation formula, of Dedekind,
48, 52
of Iseki, 54
Tunln, Paul, 185, 198
Tunin's theorem, 185, 186

u
Univalent modular function, 84
Uspensky, J. V., 94, 198

v
Valence of a modular function, 84
Van Wijngaarden, A., 22
Values, of J( r), 39
of Dirichlet series, 170
Vertices of fundamental region, 34

204
Graduate Texts in Mathematics
com"",(',/ from P(/~(' "

48 SACHS/WU. General Relativity for Mathematicians.


49 GRUENBERG/WEIR. Linear Geometry. 2nd ed.
50 EDWARDS. Fermat's Last Theorem.
51 KLINGENBERG. A Course in Differential Geometry.
52 HARTSHORNE. Algebraic Geometry.
53 MANIN. A Course in Mathematical Logic.
54 GRAVER/WATKINS. Combinatorics with Emphasis on the Theory of Graphs.
55 BROWN/PEARCY. Introduction to Operator Theory I: Elements of Functional Analysis.
56 MASSEY. Algebraic Topology: An Introduction.
57 CROWEUJFox. Introduction to Knot Theory.
58 KOBLITZ. p-adic Numbers, p-adic Analysis, and Zeta-Functions. 2nd ed.
59 LANG. Cyclotomic Fields.
60 ARNOLD. Mathematical Methods in Classical Mechanics. 2nd ed.
61 WHITEHEAD. Elements of Homotopy Theory.
62 KARGAPOLOV/MERZIJAKOV. Fundamentals of the Theory of Groups.
63 BOLLABAs. Graph Theory.
64 EDWARDS. Fourier Series. Vol. I. 2nd ed.
65 WELLS. Differential Analysis on Complex Manifolds. 2nd ed.
66 WATERHOUSE. Introduction to Affine Group Schemes.
67 SERRE. Local Fields.
68 WEIDMANN. Linear Operators in Hilbert Spaces.
69 LANG. Cyclotomic Fields II.
70 MASSEY. Singular Homology Theory.
71 FARKAS/KRA. Riemann Surfaces.
72 STILLWELL. Classical Topology and Combinatorial Group Theory.
73 HUNGERFORD. Algebra.
74 DAVENPORT. Multiplicative Number Theory. 2nd ed.
75 HOCHSCHILD. Basic Theory of Algebraic Groups and Lie Algebras.
76 lITAKA. Algebraic Geometry.
77 HECKE. Lectures on the Theory of Algebraic Numbers.
78 BURRIS/SANKAPPANAVAR. A Course in Universal Algebra.
79 WALTERS. An Introduction to Ergodic Theory.
80 ROBINSON. A Course in the Theory of Groups.
81 FORSTER. Lectures on Riemann Surfaces.
82 BOTT/Tu. Differential Forms in Algebraic Topology.
83 WASHINGTON. Introduction to Cyclotomic Fields.
84 IRELAND/RoSEN. A Classical Introduction to Modem Number Theory.
85 EDWARDS. Fourier Series: Vol. II. 2nd ed.
86 VAN LINT. Introduction to Coding Theory.
87 BROWN. Cohomology of Groups.
88 PIERCE. Associative Algebras.
89 LANG. Introduction to Algrebraic and Abelian Functions. 2nd ed.
90 BR0NDSTED. An Introduction to Convex Polytopes.
91 BEARDON. On the Geometry of Discrete Groups.
92 DIESTEL. Sequences and Series in Banach Spaces.
93 DUBROVIN/FoMENKO/NoVIKOV. Modem Geometry-Methods and Applications Vol. 1.
94 WARNER. Foundations of Differentiable Manifolds and Lie Groups.
95 SHIRYAYEV. Probability, Statistics, and Random Processes.
96 CONWAY. A Course in Functional Analysis.
97 KOBLITZ. Introduction to Elliptic Curves and Modular Forms.
98 BROCKER/TOM DIECK. Representations of Compact Lie Groups.
99 GRovE/BENSON. Finite Reflection Groups. 2nd ed.
100 BERG/CHRISTENSEN/RESSEL. Harmonic Anaylsis on Semigroups: Theory of Positive Definite
and Related Functions.
101 EDWARDS. Galois Theory.
102 VARADARAJAN. Lie Groups, Lie Algebras and Their Representations.
103 LANG. Complex Analysis. 2nd ed.
104 DUBROVIN/FoMENKO/NoVIKOV. Modem Geometry-Methods and Applications Vol. II.
105 LANG. SLiR).
106 SILVERMAN. The Arithmetic of Elliptic Curves.
107 OLVER. Applications of Lie Groups to Differential Equations.
108 RANGE. Holomorphic Functions and Integral Representations in Several Complex Variables.
109 LEHTO. Univalent Functions and Teichmiiller Spaces.
110 LANG. Algebraic Number Theory.
III HUSEMOLLER. Elliptic Curves.
112 LANG. Elliptic Functions.
113 KARATZAS/SHREVE. Brownian Motion and Stochastic Calculus.
114 KOBLITZ. A Course in Number Theory and Cryptography.
115 BERGERIGosTIAUX. Differential Geometry: Manifolds, Curves, and Surfaces.
116 KELLEy/SRINIVASAN. Measure and Integral, Volume 1.
117 SERRE. Algebraic Groups and Class Fields.
118 PEDERSEN. Analysis Now.
119 ROTlAAN. An Introduction to Algebraic Topology.
120 ZIEMER. Weakly Differentiable Functions: Sobolev Spaces and Functions of Bounded
Variation.

Vous aimerez peut-être aussi