Modular Functions and Dirichlet Series in Number Theory (Apostol) PDF
Modular Functions and Dirichlet Series in Number Theory (Apostol) PDF
EditOl·ial Board
s. Axler F.W. Gehring К.А. Ribet
Modular Functions
and Dirichlet Series
in Number Theory
Second Edition
With 25 Illustrations
Springer
Тот М. Aposto!
Department of Mathematics
Ca!ifornia Institute of Techno!ogy
Pasadena, СА 91125
USA
Еl!i/щiа! В{){т!
9 8 7 6 5 4 3
v
Preface to the Second Edition
T.M. A.
July, 1989
Contents
Chapter I
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 SJ function 9
1.7 The Laurent expansion of SJ near the origin 11
1.8 Differential equation satisfied by SJ 11
1.9 The Eisenstein series and the invariants g2 and g3 12
1.10 The numbers e 1 , e 2 , e 3 13
1.11 The discriminant ~ 14
1.12 Klein's modular function J(r) 15
1.13 Invariance of J under unimodular transformations 16
1.14 The Fourier expansions of g2(r) and gir) 18
1.15 The Fourier expansions of ~(r) and J(r) 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
VB
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 Ll(r) 50
3.4 The general functional equation for 11(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 !per) = Ll(qr)/Ll(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 II3
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 G6 II7
6.5 The linear space Mk 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
6.10 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 • O 133
6.15 Estimates for the Fourier coefficients of entire forms 134
6.16 Modular forms and Dirichlet series 136
Exercises for Chapt~r 6 138
Chapter 7
Kronecker's theorem with applications
7.1 Approximating real numbers by rational numbers 142
7.2 Dirichlet's approximation theorem 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
Exercises for 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 seriesf(s) on a line
a = ao 170
8.7 Equivalence of general Dirichlet series 173
8.8 Equivalence of ordinary Dirichlet series 174
8.9 Equality of the sets Uiao) and Uiao) for equivalent
Dirichlet series 176
8.10 The set of values taken by a Dirichlet series in a neighborhood
of the line a = ao 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
Exercises for Chapter 8 187
Bibliography 196
Index 201
x
1
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 S 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.
Notation. If WI and W2 are two complex numbers whose ratio is not real
we denote by il(WI' W2), or simply by il, the set of all linear combinations
mWI + nW2, where m and n are arbitrary integers. This is called the lattice
generated by WI and W2'
2
1.3: Fundamental pairs of periods
(a) (b)
Figure 1.1
o o
(a) (b)
Figure 1.2
3
I: Elliptic functions
where t 1 and t2 are real. Now let [t] denote the greatest integer:::;; t and
write
Then
W - [t 1]W 1 - [t 2]W 2 = r1w1 + r2w2'
If one ofr1 or r2 is nonzero, then r 1w 1 + r2w2 will be a period lying inside
the parallelogram with vertices 0, WI' W2, WI + W2 . But if a period w lies
inside this parallelogram then either w or WI + W2 - w will lie inside the
complete.
°
triangle 0, WI' W2 or on the diagonal joining WI and W2 , contradicting the
hypothesis. (See Figure 1.2b.) Therefore r1 = r2 = and the proof is
0
Definition. Two pairs of complex numbers (WI, ( 2) and (WI', wz'), each with
nonreal ratio, are called equivalent if they generate the same lattice of
periods; that is, if 0(W1' (2) = 0(w 1', W2').
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 (WI' (2) and (WI', wz') are equivalent if, and only if,
there is a 2 x 2 matrix (; : ) with integer entries and determinant
ad - bc = ± 1, such that
W/ = aW2 + bw l ,
WI' = CW2 + dWl'
4
1.4: Elliptic functions
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!,.
Because of periodicity, it suffices to study the behavior of an elliptic
function in any period parallelogram.
Theorem 1.6. The contour integral of an elliptic fimction taken along the
boundary of any cell is zero.
PROOF. The integrals along parallel edges cancel because of periodicity. 0
Theorem 1.7. The sum of the residues of an elliptic function at its poles in any
period parallelogram is zero.
PROOF. Apply Cauchy's residue theorem to a cell and use Theorem 1.6. 0
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 of zeros of an 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 /(z) ,
taken around the boundary C of a cell, counts the difference between the
number of zeros and the number of poles inside the cell. But f'1f is elliptic
with the same periods asf, and Theorem 1.6 tells us that this integral is zero.
o
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.
A B
.,-----:-;;-2
(z - w)
+- -.
z - OJ
6
1.5: Construction of elliptic functions
L~
W
well
",*0
Figure 1.3
This shows that the partial sums S(n) are bounded above by 8,(cx - l)/r" if
cx > 2. But any partial sum lies between two such partial sums, so all of the
L
partial sums of the series Iwl-" are bounded above and hence the series
converges if cx > 2. The lower bound for S(n) also shows that the series
diverges if cx ~ 2. D
L 1"
/w/>R (z - w)
(1)
for all w with Iwi> R and all z with IzI ~ R. Then we invoke Lemma 1 to
prove Lemma 2. Inequality (1) is equivalent to
(2) I-
z --
w
wi" > 1
-
-M'
To exhibit M we consider all w in n with Iwl > R. Choose one whose
modulus is minimal, say Iwl = R + d, where d > O. Then if Izl ~ Rand
Iw I ~ R + d we have
Iz : wi = 11 - ~I ~ 1 -I~I ~ 1 - R : d'
8
1.6: The Weierstrass f.J function
and hence
I ~I' > (1 __
W -
R ). =~M'
R+d
where
R )-.
M= ( 1 - - -
R +d .
This proves (2) and also the lemma. o
As mentioned earlier, we could try to construct the simplest elliptic
function by using a series of the form
1
L
well (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(
well Z - W
)3'
Thenfis an elliptic function with periods Wt, W2 and with a pole of order 3 at
each period W in n.
PROOF. By Lemma 2 the series obtained by summing over Iwi> R converges
uniformly in the disk Iz I :::; 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 w in the disk.
This proves thatfis meromorphic with a pole of order 3 at each w in n.
Next we show thatfhas periods Wt and W2' For this we take advantage
of the absolute convergence of the series. We have
But W - W t runs through all periods in n with w, so the series for f(z + wd
is merely a rearrangement of the series for f(z). By absolute convergence we
have f(z + Wt) = f(z). Similarly, f(z + W2) = f(z) so f is doubly periodic.
This completes the proof. 0
9
1: Elliptic functions
1
2"
z
+ i Z
0 ",*0
L (t --2w )3 dt.
Integrating term by term we arrive at the following function, called the
Weierstrass f.J function.
Theorem 1.10. The function f.J so defined has periods w! and W2 . It is analytic
except for a double pole at each period w in Q. Moreover f.J(z) is an even
function of z.
PROOF. Each term in the series has modulus
1 1 1 1w 2 - (z - W)21 1z(2w - z) 1
1(z - W)2 - w 2 = w 2(z _ W)2 = w 2(z _ W)2 .
Now consider any compact disk Iz I ~ R. There are only a finite numoer 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,
Since -w runs through all nonzero periods with w this shows that f.J( -z) =
f.J (z), so f.J is even.
10
1.8: Differential equation satisfied by f.J
p'(z) =- 2 L(
wen Z -
1 )3'
W
We have already shown that this function has periods WI and W2' Thus
p'(z + w) = p'(z) for each period w. Therefore the function p(z + w) - p(z)
is constant. But when z = -w/2 this constant is p(w/2) - p( -w/2) = 0
since p is even. Hence p(z + w) = p(z) for each w, so p has the required
periods. 0
(4) for n ~ 3.
hence
p(z) ----;;:;:z zn
z n=1 w*O W Z n=1
where Gn is given by (4). Since p(z) is an even function the coefficients G2n + 1
must vanish and we obtain (3). 0
11
1: Elliptic functions
is called the Eisenstein series of order n. The invariants g2 and g3 are the
numbers defined by the relations
or equivalently,
m-2
(2m + l)(m - 3)(2m - 1)G 2m = 3 2: (2r - 1)(2m - 2r - 1)G2rG2m-2r
r= 2
for m :::::: 4.
The next theorem shows that these numbers are the roots of the cubic
polynomial4 p 3 - g2 P - g3'
13
1: Elliptic functions
zero at 10)2' so its order would be ~ 4. But its order is 2, so el # e2' Similarly,
el # e3 and e2 # e3'
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 g/ - 27g/. When x = SO(z) the roots of this polynomial are distinct so
the number 9 2 3 - 279 3 2 # O. This completes the proof. D
+ co 1
g3(T) = 140 m,n~- ex; (m + m)6
(m,n)*(O,O)
and
~(T) = g/(T) - 27g/(r).
Theorem 1.14 shows that ~(r) # 0 for all T in H.
14
1.12: Klein's modular function J(T)
J( ) _ gz\w 1, w z )
W 1,WZ - A( ).
I..\W 1 ,W Z
Since gz 3 and ,1 are homogeneous of the same degree we have J(),w 1 , AWz)
= J(Wb w z). In particular, if r E H we have
15
1: Elliptic functions
--------+-----+-----~----~x
-A A
Figure 1.4
for some K > O. We will prove that (8) holds for all q, with
«5 2
K=---~
1 + (A + «5)2
if Ix I ~ A and y :2: «5. (This proof was suggested by Christopher Henley.)
If Iq I ~ A + «5 inequality (8) holds trivially since (q + X)2 :2: 0 and
y2 :2: «5 2.lflql > A + «5 then Ix/ql < Ixl/(A + «5) ~ A/(A + «5) < 1 so
11 + ~1:2:
q
1 -I~I >
q
1_ _ A
A+«5
=_«5
A+«5
hence
q«5
Iq + xl :2: A + «5
and
(9)
16
1.13: Invariance of J under unimodular transformations
where a, b, c, d are integers such that ad - be = 1. Then the pair (w 1 ', w z')
is equivalent to (W1' w z ); that is, it generates the same set of periods n.
Therefore gz(w 1', w z') = gz(w 1, wz) and g3(W 1', w z') = g3(W 1, w z ) since g2
and g3 depend only on the set of periods n. Consequently, L1(w 1', W2') =
L1(w 1, w 2) and J(w 1', W2') = J(W1' W2)'
The ratio of the new periods is
, w2' aW 2 + bW 1 ar +b
r
+ d'
=-=-~--
w 1' cW z + dW 1 cr
where r = WZ /w 1 . An easy calculation shows that
Im(r') = Im(ar + b) = ad - be Im(r) = Im(r) .
cr+d Icr+dl 2 Icr+dl z
Hence r' E H if and only if r E H. The equation
ar
r' = - - -
+b
cr +d
is called a unimodular transformation if a, b, c, d are integers with ad - be = 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(r) is invariant
under the transformations of the modular group. That is, we have:
(10) J( - b)
ar-+d = J(r).
cr +
Note. A particular unimodular transformation is r' = r + 1, hence (10)
shows that J(r + 1) = J(r). In other words, J(r) is a periodic function of r
with period 1. The next theorem shows that J(r) has a Fourier expansion.
L
00
Then the upper half-plane H maps into the punctured unit disk
D = {x: 0 < Ix I < I}.
17
1: Elliptic functions
(See Figure l.S.) Each t in B maps onto a unique point ~ in D, but each
x in D is the ima.f.pfinfinitely many points in B.·U t andr map onto x
then ellrit . . el*i't"sot.and r dift'er by an integer.
•
H
figure 1.5
IfxeD, let
f(x) == J(t)
where tis any of the points in B whieh map onto x. Since J is periodic with
period I, J has the same value at a11 these points so f(x) is well-defined.
Now f is anaJytic in D because
, d. d dt , IdX J'(t}
f (x) = d- J(t) = d- J(t) -d = J (t) -d = 21t1e. Zxit'
X txt
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=-oo(m + nr)4
and L
m= -00 (m + nr)
6·
L
n cot nr = - + +00 1
r "'=-00 r+m
(1
- - - -1).
m
",*0
1- +
r
L --
+00
"'=-00 r+m
(1 -- 1)
m
= -ni 1 (+ 2L
00 e21!irt) .
r=1
",*0
+ 00 1 00
-3' L - (2ni)4 L r3e21!irt
·"'=-00 (r + m)4 r= 1
and
+00 1 00
-5' L -(2ni)6 L rSe21tirt.
·m=-oo (r + m)6 r= 1
g2(') = 4n 4 {I
3
+ 240
k= 1
f (J3(k)e21tikr}
and
= 60{ I
m=-oo
~
m
f I (
+ n=lm=-oo (m+n,)
1 + (m-nr)
1 4 4)}
m*O(n=O)
00 +00 1 }
= 60 2((4) + 2n~' m=~oo +
{
(m nr)4
2n4 16n4
= 60 { - +- L L r3 xnr
00 00 }
90 3 n=1 r=1
where x = e 21tir • In the last double sum we collect together those terms for
which nr is constant and we obtain the expansion for g2(')' The formula
for g3(') is similarly proved. 0
L ,(n)e21tinr
00
L1(,) = (2n)12
n= 1
where the coefficients ,(n) are integers, with ,(1) = 1 and ,(2) = -24.
A = L (J3(n)xn, B = L (Js(n)xn.
n=1 n= 1
Then
20
1.15: The Fourier expansions of A(r) and J(r)
and
so
5d 3 + 7d s == 0 (mod 12).
Hence 12 3 is a factor of each coefficient in the power series expansion of
(1 + 240A)3 - (1 - 504B)2 so
12
i\(r) = 64n {123
27 n=1
f
r(n)e27rinr} = (2n)12
n=1
r(n)e 27rinr f
where the r(n) are integers. The coefficient of x is 122(5 + 7), so r(l) = 1.
Similarly, we find r(2) = - 24. 0
L c(n)e27rint,
00
PROOF. We agree to write I for any power series in x with integer coefficients.
Then if x = e 27rir we have
g/(r) = ~~nI2(1 + 240x + 1)3 = ~~nI2(1 + 720x + f),
i\(r) = ~~nI2{123x(1 - 24x + I)}
and hence
12 3J(r) = - + 744 +
x n=1
where the c(n) are integers. o
21
1: Elliptic functions
Note. The coefficients c(n) have been calculated for n ;S; 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 + 1)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(11n) == 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
e4 ",,"
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
.(4) = -1472 r(9) = - 113643
• (5) = 4830 .(10) = -115920.
Lehmer has conjectured that r(n) =1= 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 I
2. Let S(O) denote the sum of the zeros of an eJliptic function f in a period paraJlelo-
gram, and let S( 00) denote the sum of the poles in the same paraJlelogram. Prove
that S(O) - S(oo) is a period of f [Hint: Integrate z/,,(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 al,"" an and b l , ... , bm be complex numbers such that none of the numbers
pea;) - p(b) is zero. Let
Prove that f is an even eJliptic function with zeros at a I, ... , an and poles at
bl , .. ·, bm •
4. Prove that every even elliptic function f is a rational function of f.J, where the
periods of p are a subset of the periods of f
where R I and R 2 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 - XI)(X - X2)(X - X3) is the
product 16{(x2 - Xd(X3 - X2)(X3 - Xd}2. Prove that the discriminant of f(x) =
4X 3 - ax - b is a 3 - 27b 2 •
8. The differential equation for p shows that gJ'(z) = 0 if z = w l /2, w 2 /2 or
(WI + w 2 )/2. Show that
(m, 0) ~ (0,0)
and (J,(n) = Idlo d' for n 2': 1, with (J3(O) = 2io, (J5(O) = - 554'
[Hint: Theorem 1.18.J
14. A series of the form I:=
I f(ll)x nl(l - XO) is called a Lambert series. Assuming
absolute convergence, prove that
x"
I I
OC' 00
where
F(n) = I fed).
dlo
24
Exercises for Chapter 1
Apply this result to obtain the following formulas, valid for Ix I < 1.
oc 11(11 jxn cp(I1)X n X
L
oc
(aj --n=x. (b) n~1 1 _ xn (1 _ X)2'
n=11 - x
.l.(I1)xn
L a.(I1)x n. L - - n = L xn2.
00 00
(d)
n=1 n=1 1 - x n=1
(e) Use the result in (c) to express g2(T) and g3(T) in terms of Lambert series in
x = e 2 n:it.
Note. In (a), /1(11) is the Mobius function; in (b), cp(l1) is Euler's totient; and in (d),
.l.(I1) is Liouville's function.
15. Let
and let
F(xj = L 1 + xn
n::;;;1
(nodd)
(c) Use Theorem 12.17 in [4] to prove the more general result
~ 114k + 1 24k + 1 - 1
11
L
= I
1 + en .. = 8k + 4 B4k + 2 •
(n odd)
25
2
The modular group
and modular functions
(1) fez) = az +b
ez +d
where a, b, e, d are arbitrary complex numbers.
Equation (1) definesf(z) for all z in the extended complex number system
C* = C U {oo} except for z = -die and z = 00. We extend the definition
off to all of C* by defining
a
and f(oo) = -,
e
with the usual convention that zlO = 00 if z #- o.
First we note that
26
2.1: Mobius transformations
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
(f" g)(z) = f(g(z)). The identity matrix I = G~) is associated with the
identity transformation
1z +0
f(z) = z = Oz + l'
27
2: The modular group and modular functions
A- 1 = (
-c
d -b)a
is associated with the inverse ofJ,
dz - b
f-l(Z) =- --
-cz + a
Thus we see that the set of all Mobius transformations with ad - be = 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.
A = (: : ) 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
AT = aT + b.
cr +d
The first theorem shows that r is generated by two transformations,
1
Tr = T +1 and Sr = - - .
T
T = (~ !) and S = (~ -1)O·
That is, every A in r can be expressed in theform
A = Tn'STnzs ... STnk
28
2.2: The modular group r
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
-2
=
(4 1)
11 3'
AT
-2
S-
_ (4 1)(0 -1) _(1 -4)
11 3 1 0 - 3 -11 .
This interchanges the two columns and changes the sign ofthe second column.
Again, multiplication by a suitable power of T gives us a matrix with
Idl < lei. In this case we can use either T4 or T3. Choosing T4 we find
AT
-2
ST =
4 (1 -4)(1 1) G~).
3 -11 0 =
Multiplication by S gives
AT- 2 ST 4S = (~ -1)
-3 .
Now we multiply by T3 to get
If e = 0 then ad = 1 so a = d = ± 1 and
±b)1 = T±h .
Thus, A is a power of T
If e = 1 then ad - b = 1 so b = ad - 1 and
A = G 1) GnG -~)(~
ad; = ~) = TaST 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
-1)o = (-a q +b
r
-a).
-e
By the induction hypothesis, the last matrix is a product of powers of S
and T, so A is too. This completes the proof. 0
30
2.3: Fundamental regions
For example, the next theorem will show that a fundamental region Rr
of the full modular group r consists of all t in H satisfying the inequalities
It I > 1, It + fl < 1.
This region is the shaded portion of Figure 2.1.
t = U + iv, v> 0
----------~----.----+-----.----+----------u
-1 -t o
Figure 2.1 Fundamental region of the modular group
The proof will use the following lemma concerning fundamental pairs of
periods.
Let Wi = Wi and let W2 be the first member of this sequence that is not a
multiple of Wi' Then the triangle with vertices 0, Wi' W z contains no element
of n except the vertices, so (WI> w z ) is a fundamental pair which spans the
set n. Therefore there exist integers a, b, e, d with ad - be = ± 1 such that
(::) = G~)(::)
If ad - be = -1 we can replace c by -c, d by -d, and Wi by -WI and the
same equation holds, except now ad - be = 1. Because of the way we have
chosen WI> W z we have
and
since WI ± W 2 are periods in n occurring later than W2 in the sequence. 0
PROOF. Let WI' = 1, w 2 ' = r' and apply Lemma 1 to the set of periods
n = {m + nr': m, n integers}. Then there exists a fundamental pair WI' W z
with IW21 ~ Iwll, IWi ± wzi ~ IW21· Let r = WZ/w I· Then r = (: ~)r'
with ad - be = 1 and
Irl ~ 1, Ir ± 11 ~ Irl. o
N ate. Those r in H satisfying Ir ± 11 ~ Ir I are also those satisfying
Ir+ilsl.
PROOF. Theorem 2.2 shows that if r' E H there is a point r in the closure of
Rr equivalent to r' under r. To prove that no two distinct points of Rr are
equivalent under r, let r' = Ar where A = (: ~} We show first that
Im(r') < Im(r) if r E Rr and e ¥- O. We have
, Im(r)
Im(r) = Icr + dlz'
32
2.3: Fundamental regions
±1b) = T±b
.
But then b = 0 since both T and T' are in Rr so T = T'. This proves that no
two distinct points of Rr are equivalent under r.
Finally, if AT = T for some T 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 . D
Figure 2.2 shows the fundamental region Rr 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 Rr are circles orthogonal to the real
I T
33
2: The modular group and modular functions
axis, the same is true of every image f(Rr) under the elements f of r. The
set of all images f(R r ), where / E r, is a collection of nonoverlapping open
regions which, together with their boundary points, cover all of H.
L
00
Property (a) states that/is analytic in H except possibly for poles. Property
(b) states that / is invariant under all transformations of r. Property (c) is
a condition on the behavior offat the point r = ioo. If x = e2 1[it the Fourier
series in (c) is a Laurent expansion in powers of x. The behavior of/at ioo is
described by the nature of this Laurent expansion near O. If m > 0 and
a( -m) # 0 we say thatfhas a pole of order m at ioo. If m ~ 0 we say fis
analytic at ioo. Condition (c) states thatfhas at worst a pole of order mat
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 0/ the
fundamental region R r , the number 0/ 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 ail, we consider the boundary of Rr as the union of
four edges intersecting at four vertices p, i, P + 1, and ioo, where p = e2 1[i/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 i 00 is the order of the zero or pole at
x = 0, measured in the variable x = e 2nit •
34
2.4: Modular functions
-- .... --
ooi
".-
....
/ "" ",
(1) (4)
(2) (3)
p p +1
Figure 2.3
PROOF. Assume first that / has no zeros or poles on the finite part of the
boundary of Rr . Cut Rr by a horizontal line, Im(r) = M, where M > 0 is
taken so large that all the zeros or poles of/ are inside the truncated region
which we call R. [If/had an infinite number of poles in Rr they would have
an accumulation point at ioo, contradicting condition (c). Similarly, since/
is not identically zero, / 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 of/inside R. Then
N - P = _1
2ni
r f'(r) dr = _1
JOR /(r) 2ni
{f + f + f + f + f }
(1) (2) (3) (4) (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
-t +
-
iM . - - - - - - - - - - - - . , t + iM
R t
-
p p+i
Figure 2.4
35
2: The modular group and modular functions
(5)
j (I) (4) r
(2) (3)
~ ~
Figure 2.5
N - P = 2ni
1 f(5)
f'(r)
f(r) dr.
N - P = _1
2ni
f
(5)
f'(r) dr
f(r)
= - _1
2ni
i
JK
F'(x) dx = -(NF - PF) = PF - N F,
F(x)
where N F and PF are the number of zeros and poles of F inside K.
36
2.4: Modular functions
N = P + m.
°
Thereforeftakes on the value in Rr as often as it takes the value 00.
If there is a zero of order n at x = 0, then m = -n so PF = 0, NF = n,
hence
N +n= P.
°
Again,ftakes the value in Rr as often as it takes the value 00. This proves
the theorem iff has no zeros or poles on the finite part of the boundary of R r .
Iffhas 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 .
- t
Figure 2.6
N - P =1-
2ni
{(f + f) + f + f-
CI C3 C2
I 2 iM
/ +
1/2+iM
}f'(r}
-dr
f(r}
= -1
2ni
{(f + f) + f }f'(r}
CI
--
C3 C2 f(r}
dr + m,
-! + iM .------------, ! + iM
"
p + 1
Figure 2.7
_1_
2ni
f
CI
I'(r) dr = _1_
f(r) 2ni
fa
1[(2
(~+
+ re,9)
re,9
g'(p
g(p
+ re i9 ))re i9 i de
lim _1
2ni
r ... O
f CI
I'(r) dr
f(r)
= - ~6'
Similarly,
lim _1_
2ni
r ... O
f C3
I'(r) dr =
f(r)
k
6
38
2.5: Special values of J
so
lim _1
r~O 2ni
(1 + 1
c, C3
)f'(r) dr =
f(r)
k
lim _1 f f'(r) dr =
r~O 2ni C2 f(r) 2
Therefore we get the formula
k I
N - P = m - 3 - 2'
Ufhas 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 Rr (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 = then m = - n and the
equation becomes
°
I k
N +n +- + -
= P.
2 3
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. 0
Theorem 2.5. Iff is modular and not constant, then for every complex c the
function f - 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. o
Theorem 2.6. Iff is modular and bounded in H then f is constant.
PROOF. Since f is bounded it omits a value so f is constant. o
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
40
2.7: Mapping properties of J
v
Rr
J
..
-/
/
P ",\
I
I \
I \
\
u
-1 0 1
0= J(p) 1 = J(i)
(a) (b)
Figure 2.8
J(.) = J( -f).
and
Theorem 2.9. Given two complex numbers a2 and a3 such that a2 3 - 27a32 =1= O.
Then there exist complex numbers WI 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, be any
complex number such that
and let W2 = pw" where p = e211i/3. We know that 93(1, p) =1= 0 because
92(1, p)= 0 and L\(1, p) = 92 3 - 279/ =1= O. Then
1
92(WI, W2) = 92(W" wlP) = -492(1, p) = 0 = a2
WI
and
Wl4 = _92_(1_,_i)
42
2.9: Application to Picard's theorem
(6)
Comparing this with (5) we find that 92(W I, (2) = a2 and hence by (6) we
also have 93(W I, ( 2) = a3' This completes the proof. D
Theorem 2.10. Every nonconstant entire function attains every complex value
with at most one exception.
43
2: The modular group and modular functions
( ) _ /(z) - a
9z - b .
-a
Then 9 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 Rr under transformations of r. Since J maps the closure
of Rr 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'(T) "# 0 if T "# p or T "# i and since J'(p) = J'(i) = 0, each single-
valued branch of r 1 is locally analytic everywhere except at 0 = J(p),
1 = J(i), and 00 = J( (0). For each single-valued branch of r 1 the composite
function
h(z) = r 1 [g(z)]
44
Exercises for Chapter 2
CONGRUENCE SUBGROUPS
and
45
2: The modular group and modular functions
46
3
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 = {t: Im(t) > O} by the equation
TI (1
00
The infinite product has the form TI (1 - xn) where x = e 2nit . If t 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, l1{t) is analytic
onH.
The eta function is closely related to the discriminant Ll(t) introduced
in Chapter 1. Later in this chapter we snow that
Ll(t) = (211:)12 11 24(t).
This result and other properties of 1J(t) follow from transformation formulas
which describe the behavior of l1(t) under elements of the modular group r.
For the generator Tt = t + 1 we have
= eni(t+ 1)/12 TI (1
00
(3)
Equation (2) also shows that 11 24(t) is periodic with period 1.
47
3: The Dedekind eta function
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
l1(ar
cr
+
+d
b)
to l1(r) 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.
n=1
e- 2nny )
:n:y :n:y 00 en
L L
00
48
3.2: Siegel's proof of Theorem 3.1
-y y
-i
Figure 3.1
But
cot i(}
cos i(}
= -.-.-
sm I(} = i e
e- 8
8
+ e88 =
- e -
e28 + 1 1(
1 I - e28 .
i e28 - 1 = i -
2)
Using this with () = nk/y we get
n 1"1 I n l 1
L Res F (z) = - L - - - L ----"--,,..,-
k= -n z=ik/N n 4ni k= 1 k 2ni k= 1 k 1 - e21tk / y '
k,.O
Similarly
n in 1 i nl 1
L Res Fn(z) =-4 L -k--2 L-k-l-~2"~kY'
k = -" z =ky/N n k =1 n k =1 - e
k,.O
Hence 2ni times the sum of all the residues of Fiz) 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-+C()
f C
Fiz) dz =, -1 log y.
49
3: The Dedekind eta function
.f
hm
n--+oo C
Fn(z) dz = f. dz
hm zFn(z)-
C n-4co Z
1
=- {f
8
-
Y
-i
+ Ii -
Y
f- + f- i} -
i
Y
-y
dz
Z
= ~ {-
4
fY + Ii} dz
_I Y Z
50
3.4: The general functional equation for ~(r)
X"f4.
"~ 1
Consequently,
L r(n)x" = x n (1
00 00
PROOF. Let f(r) = L\(r)/I'/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 fat ioo. We have
I'/24(r) = e 21tit n (1 -
00
n~l
e 21ti"t)24 =x n (1 -
00
n~l
X")24 = x(l + I(x)),
where I(x) denotes a power series in x with integer coefficients. Thus, I'/24(r)
has a first order zero at x = O. By Theorem 1.19 we also have the Fourier
expansion
00
L\(ar
cr
+
+d
b) = (cr + d)12L\(r)
and using (6) we find that
I'/(ar
cr
+
+d
b) = e(cr + d)1/21'/(r),
(12) log 11(:: : !) = log I1(r) +n{a l;cd + s( -d, C)) +! log{ -i(cr + d)}.
From the definition of I1(r) as a product we have
nir nir
(13) log I1(r) = -
12
+ L log(1 -
00
n=l
•
e 2n •nt ) = - -
12
L A( -inr),
00
n=l
(15) LA(-mr)
00 •
= LA
00 (
-in +-
ar - b) + -ni ( r ---
ar + b)
n=l n=l cr+d 12 cr+d
a +d
+ ni( ~ + s( -d, c) ) +! log{ - i(cr + d)}.
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) I: A{~k
n= 1
(z - ih)} = I: A{~k (~z - iH)}
n= 1
n (z -
+ "21 log z - 12k ~1) + . 1rls(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 = e, h = -d, H = a, z = -i(cr + d).
Then Re(z) > 0, and the condition ad - be = 1 implies - hH - bk = 1, so
(h, k) = 1 and hH == -1 (mod k). Now b = -(hH + 1)/k and iz = er + d,
so
iz - d iz +h
r=--=--
e k
and hence
iz + h hH + 1 iz (
ar+ b = H -k- - k =I H+~. i)
Therefore, since er +d= iz, we have
ar +
er +d
b= ~k (H + ~).
z
Consequently
ar + b
r - cr + d =
1-
k (h H) + ki ( z - 1)
~ = - -e- +
a +d
ki ( Z - ~
1)
so
53
3: The Dedekind eta function
Note. The sum on the right of (18), which contains Bernoulli polynomials
Bix), is equal to
PROOF. First we assume that 0 < ex < 1 and 0 < {3 < 1. We begin with the
first sum appearing in (17) and use (14) to write
e27timfJ
L: A((r + ex)z - L: L - - e-27tm('+~)z.
ct:J ct:J ct:J
(19) i{3) =
,:0 ,:Om:J m
Now we use Mellin's integral for e- X which states that
(20) 1.
e- x = -2
m
IC
+
c-ct:Ji
ooi
r(s)x- S ds,
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
cp(s) = f
oo
xs-Jt/I(x) dx
.
if, and only if, t/I(x) = -2.
1 I C +ooi
cp(s)x- S ds.
o nl c- ct:Ji
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 + ex)z and c = 3/2 to the last
exponential in (19) and writing I(c) for I~~~: we obtain
L: A((r + ex)z -
00
i{3) = L: L -m- -2.m
00 00 e27timfJ 1 f. r(s){2nm(r + ex)z} -s ds
f.
,:0 ,:0 m: J (3/2)
1 r(s) 00 1 e 21timfJ
L (r + ex)S L -y:tS ds
00
= -2
m. (3/2)
-(2)S
nz ,:0 m:J m
= -2. f.
1 r(s)
-(2)S ((s, ex)F({3, 1 + s) ds.
m (3/2) nz
54
3.5: Iseki's transFormation Formula
Here ((s, 0:) is the Hurwitz zeta function and F(x, s) is the periodic zeta function
defined, respectively, by the series
ex 1 oc e27rimx
f. A((r + 1 -
r=O
o:)z + i{3) 1.
= -2
7rl
J
(312)
(2r(S))5 ((s, 1 - o:)F(1 - {3, 1
1[Z
+ 5) ds,
so (17) becomes
where
i + iT
~
-- i + iT
t
i-iT -- i-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
i312) - i - 312) +R
where R is the sum of the residues at the poles of the integrand inside the
rectangle. This gives us the formula
55
3: The Dedekind eta function
where we have used Theorem 12.19 of [4] to express the Fourier series as a
Bernoulli polynomial.
To calculate R(O) we recall that ((0, a) = 1 - a. Hence ((0, 1 - a) = a - 1
so
e 2nin f! _ e - 2ninfJ
L ----
CL
56
3.5: Iseki's transformation formula
Note that this is the same as R(l) = Res s = 1 z-S<l>(ex, {3, s), except that z is
replaced by - z - 1, ex by 1 - {3, and {3 by ex. Hence we have
R( -1) = - nzB2(ex).
Thus
L - - e- 2rrm.z L e-21tmrz
ex e2rrimp ex.
=
111=1 m r=O
(( e 21tim {J e - 2 rrmaz x
L
",=1
- - 1 _ e 2rrmz
m
L e 2rrimPJ.(m),
m=1
say, where
1 e - 2rrm.z
L
CJJ
e2rrimp J,(m)
m=1
ex ~ 1 - ex, {3~I-f3
ex ~ 1 - {3, z ~-.
z
o
57
3: The Dedekind eta function
(27) L A(nz) =
00
n= 1
LA
00
n= 1
- (n) + -21log z - -12n(z - -1)z .
Z
We can deduce this from Iseki's formula (18) by taking f3 = 0 and letting
0( -+ 0 +. Before we let 0( -+ 0 + we separate the term r = 0 in the first term
of the series on the left of(18) and in the second term ofthe series on the right
of (18). The difference of these two terms is A(o(Z) - A(iO(). Each of these tends
to 00 as 0( -+ 0+ but their difference tends to a finite limit. We compute this
limit as follows:
. 1 - e-2"i~
A(o(Z) - A(iO() = 10g(1 - e- 21t1 <%) - 10g(1 - e- 2,,<%Z) = log 1 _ e 2,,<%z'
By L'Hopital's rule,
. 1 - e- 2"i<% • 2ni i
hm 1
<%-0+ - e
2,,<%z = <%-0
hm -2 = -
nz Z
so
lim (A(O(Z) - A(iO(» = log ~ = ni - log z.
<%-0+ Z 2
Now when 0( -+ 0+ the remaining terms in each series in (18) double up
and we obtain, in the limit,
(28)
ni
- - log Z + 2 L
00
A(rz) = 2 A - - - L00 (r) nz + -n+-.
ni
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 0( and f3 in
Iseki's formula (18) as follows. Take
58
3.6: Deduction of Dedekind's functional equation from Iseki's formula
and write
Now let
Note that v == hf.1 (mod k) so - Hv == - HhJ1 == f.1 (mod k), and therefore
-Hv/k == 14k (mod 1). Hence ex = f.1/k == -Hv/k (mod 1) and f3 = v/k ==
hf.1/k (mod 1). Substituting in Iseki's formula (18) and dividing by 2 we get
Now sum both sides on f.1 for J1 = 1, 2, ... , k - 1 and note that
and similarly for the set of all numbers rk + k - f.1. Also, since v == hf.1 (mod k),
as f.1 runs through the numbers 1,2, ... , k - 1 then v runs through the same
59
3: The Dedekind eta function
n~1 A(~ (z - ih)) = n~1 A(k~ (-z1- iH)) + -rr2 (1-z - z) /1=L k-I
I
112
2"
k
n'tO (modk) ntO (modk)
- n(1
- - - Z L 11- + -rr(1- - )(k -
)k-1 Z 1)
2 Z /1= I k 12 Z
+ ni L - - - - - - L - + -
k-I Ii (v 1) ni k-I v rri
(k - 1)
/1=lk k 2 2/1=lk 4
<X)
L A( -n (1- - iH))
k Z
+ -rr ( Z
12
- -1) (
Z
1 - -1)
k
+ rri kL (v
- I -
P - - -1).
/1= I k k 2
n= I
n to (mod k)
h: = q + ~, q= [h: J I = h: - [~ J.
Therefore
(29)
m= I
f A(rnz) = m=f A(~) - ~12 (z - ~)z + ~2 log z.
I Z
This accounts for the missing terms in (29) with n == 0 (mod k), if we write
11 = rnk. When (27) is combined with (29) we get
This proves (16) which, in tum, completes the proof of Dedekind's functional
equation for 1](T). For alternate proofs see p. 190 and [18], [35], and [45]. 0
"" ((lir))
L.
r mod k k
_0 for (Ii, k) = 1.
Since
(31)
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 5(h', k) = ±5(h, k), with the same sign as
in the congruence. Similarly, we have:
(b) If hn =
± 1 (mod k) then s(n, k) = ± 5th, k).
(c) If h 2 + 1 = 0 (mod k), then 5th, k) = O.
PROOF. Parts (a) and (b) follow at once from (31). To prove (c) we note that
h2 + 1 =°
(mod k) implies h = -n
(mod k), where 11 is the reciprocal of
h mod k, so from (a) and (b) we get 5th, k) = -5th, k) = O. 0
For small values of h the sum 5th, k) can be easily evaluated from its
definition. For example, when h = 1 we find
s(l, k) =
r(r 1) 1 L r - -1 L r
L - -- -
k- 1
= 2:
k- 1 2 k- 1
r= 1 k k 2 k r= 1 2k r= 1
(2 k) = (k - l)(k - 5) if k is odd.
s , 24k
Theorem 3.7 (Reciprocity law for Dedekind sums). If h > 0, k > 0 and
(h, k) = 1 we have
PROOF. Dedekind first deduced the reciprocity law from the functional
equation for log I1(T). 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 - = L -
((hr))2 = L -
((r))2 = L (r---1)2
k- 1
r =1 k r mod k k r mod k k r=1 k 2 .
62
3.8: The reciprocity law for Dedekind sums
i
r= 1
((hl'))2 = kil (hI' _ [hl'J _~)2
k r= 1 k k 2
_
-
L (h2r2
k-l
-k 2+ [hrJ2
r= 1
1 hI'
-k +---+
4 k
[hrJ
-k -2hr
-k
[hrJ)
-k
L
= 2h k-l
r= 1
-kI' (hr-k - [hrJ
-k --21)
L [hrJ([hrJ
+ k-J - - + ) h L
1 - 2:
2
k-l
1'2 + -1 k-lL 1.
r= 1 k k k r= 1 4 r= 1
In the sum on the left we collect those terms for which [hr/k] has a fixed value.
I'
Since 0 < < k we have 0 < hr/k < h and we can write
For a given v let N(v) denote the number of values of I' for which [hr/k] =
v - 1. Equation (34) holds if, and only if
v-l<-<v
hI' or
k(v - 1) kv
k ' h <I'<h'
equality being excluded since (h, k) = 1 and 0 < I' < k. Therefore, if
1 S; v S; h - 1, Equation (34) holds when ranges from [k(v - 1)/h] + 1 I'
to [kv/h], and hence
But when v = h the quotient kv/h = k and since I' = k is excluded we have
N(h) = k _ 1_ [k(h; 1]
63
3: The Dedekind eta function
Hence
(35) L [hrJ([hrJ
k- I- - + 1) = Lh (v - l)vN(v)
r= I k k v= I
v~/v
h
- 1)v ([kVJ
h - [k(V h- I)J) - h(h - 1)
L
= h-I [kVJ
- {(v - 1)v - v(v + I)}
1'= I h
+ kh(h - 1) - h(h - 1)
= Lv~
-2 h-I [k J+ h(h - l)(k - 1).
v= I h
Now we also have
2hs(k, h) = 2 L v(kV
h-I
- -
[kVJ
- -
1)
- = - 2L v -
h-I [kVJ + -2k L v2 - h-I
h-I
LV
1'=1 h h 2 1'=1 h h,,=l v=1
so (35) becomes
k- I
II:
[h J([h J)
+ 1 = 2hs(k, h) -
:
2k
h V~I v2 + V~I V + h(h
h- 1 h- I
- l)(k - 1).
6h k- I k- I [hrJ k- I
(36) 6ks(h, k) = k r~/2 - 6 r~/ k - 3 r~/'
Since 6 I::
i r2 = k(k - 1)(2k - 1) each term on the right of (36) is an
integer. Moreover, (36) shows that
so we have
L [hrJ
== (k - l)(k - 2h) - 4 k-\ - (mod 8)
r= 1 k
r odd
65
3: The Dedekind eta function
-4 L [hr]
k-I
- = 4L k-I ((hr))
- - 4 L -hr + 2 L 1
k-I k-I
'= h2 + k 2 + 1 + k - 4hk - 4k L [ ~
2kV] (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). 0
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 = c1q, where C 1 > 0, let
66
3.9: Congruence properties of Dedekind sums
where 8 = (3, c). The same theorem with k = CI = c/q gives, after multi-
plication by q,
Also,
12c6 = 12cs(a, c) - (a + d) - 12cs(a, cd + 3(a + d)
== 12cs(a, c) + 2(a + d) (mod 9),
so
(44) 12rac6 = 12racs(a, c) + 2r(a 2 + ad) (mod 9).
67
3: The Dedekind eta function
a +
12e { s(a, e) - ~ d}
== e - 1 + 4T(a, c) - (a + d) (mod 8)
We only need the fact that T(a, e) is an integer. Applying (40) again with
k = el = c/q and multiplying by q we have
12ac{S(a, c) - a l;ed} == a 2 + c2 + 1
68
3.10: The Eisenstein series G2 (r)
Subtract. multiply by I' and use the congruence 4('1' == 0 (mod 2' + 3) to obtain
12earb == rcel(q - 1) + r(q - l)bc == 0 (mod 2-<+3).
Since a is odd we can cancel a to obtain
(46) 12erb == 0 (mod 2-<+ 3).
Now (43) states that 12erb == 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 b = b(a) to indicate the dependence on a.
If a' = a + te, where t is an integer, an easy calculation shows that
b(a') - b(a) = t(q - 1)/12 since s(a, c) = s(a', c) and s(a', el) = s(a, cd. There-
fore rb(a') - rb(a) = 2t, an even integer. Choosing t so that a' ~ 1 we know
rb(a') is even by the above argument, so rb(a) is also even. This completes
the proof. 0
where, as usual, a<x(n) = Ldln d<X. 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
+ 2(2ni)2 L a(n)e21tinr.
00
If x = e 21tir
the series on the right of (49) is an absolutely convergent
power series for Ixl < 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)
a relation which leads to another proof of the functional equation 1]( - l/r) =
(- ir)I/21](r).
69
3: The Dedekind eta function
H illt: Start with Equation (12) of Chapter 1, replace r by liT, where II > 0, and sum
over all n ~ 1.
(52) T- 2 G2 - (-1)T
= 2(2) + Loc
m=-xn=-oc(m+nr)
L
x 1 2'
the iterated series in (52) being the same as that in (5 1) except with the order of sum-
mation reversed, Therefore, proving (50) is equivalent to showing that
x oc 1 x x 1 2rri
(53) L L
m=_",'n=_",(m+nr)2
= L L
n=_cx:m=_oc(m+IH)2
--
T
I
n=_x(lH+m)2
1 = - 8rr 2 fox cos(2rrmll)9,(II) dtt,
n*O
where
xc
9,(11) = II Le 2 "in,u if II > 0
,,= 1
and
-\
9,(0) = lim 9,(11) = - , '
"-0+ 2mT
4. (a) Use Exercise 3 to deduce that
(55) L
Z
m=-, n=-x
x
L (nT
1
+ /11)2 = f r(t) cos(2rrmt) dr,
n*O
where
x
fir) = L g,(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
t{f(O+) + I(I- )}, Show that
_IX
1(0+) = ' + L9,(k)
-2
mT k=l
and that
d
-4rci -log I1(T) = G 2(T),
dT
-d log 11
ciT
(-I)
~
T
d
= -log
tiT
I1(T) Iti
+ - -Iog(
2 dT
-iT),
Integration of this equation gives 11( - liT) = C( - iT)l!2l1(r) for some constant C.
Taking T = i we find C = 1.
6. Derive the reciprocity law for the Dedekind sums s(l!, k) from the transformation
formula for log I1(T) as given in Equation (12),
7. (a) If 0 < a < I and Re (5) > I, prove that Hurwitz's formula implies
r(1 - 5) '1) 2
F(a, 5) = (2rc)1 s {err'l -s / W- S, a) + err,'( s- 1/2
) W - 5, I - a)l,
(b) Use (a) to show that <1>(1X, fi, 5) can be expressed in terms of Hurwitz zeta functions
by the formula
$(0:, fi, 5) , 2
-"-- = err ,,/ {(Is, 0:)(( -5, I - fJ) + ((5, I - 0:)(( -s,13))
ns)r( -s)
+ e - rris/2 {(( - s, I - fiKIs, I - 0:) + (( - s, fJK(s, 0:))
71
3: The Dedekind eta function
I (e- nlfl )
Is sin 1!51 = 0 -I-tl- ,
and that
-1
(d) If IJ :2: and It I :2: 1 obtain the estimate 1((5, a)1 = O(ltl') for some c > 0
(see [4], Theorem 12.23) and use (b) to deduce that
Iz-ScD(a, f3, 5)1 = O(ltI2e-Ie-lfl,').
This shows that the integral of z -ScD(~, f3, 5) along the horizontal segments of the
rectangle in Figure 3.2 tends to 0 as T --> Xo.
11. For integers r, h, k with k :2: I prove that we have the finite Fourier expansion
(( -hr)) = - -
I k- I
I .
SIn ~-
21!hrv 1!V
cot -
k 2k ,= I k k
I 1!hr 1!r
I
k-I
5(h, k) = - cot - cot-.
4k r= I k k
12. This exercise relates Dedekind sums with the sequence {u(Il)) of Fibonacci numbers
I, 1,2,3,5,8, ... , in which u(1) = u(2) = I and U(11 + 1) = U(I1) + !t(1I - I).
(a) If h = u(211) and k = u(211 + I) prove that 5(h, k) = O.
(b) If h = u(211 - I) and k = u(211) prove that 12hks(h, k) = h 2 + k 2 - 3hk + 1.
72
Exercises for Chapter 3
17. If k == I' (mod h) and if h == t (mod 1') where I' ;::: 1 and t = ± I, then
h2 - t(1' - 1)(1' - 2)h + 1'2 + 1
12hk5(h,k) = k 2 - k + h2 + 1.
I'
1'0 = k, 1'1 = h. I'j+ 1 == I'j_ 1 (mod 1'), 1:$ I'j+ 1 < I'j' I'n+ 1 = 1.
Prove that
s(lI,k)= ~
12j~1
nf {(-I)j+1 r/ +.r j _ 1
'jrj_1
2
+ I} _ (_I)" + I.
8
This also expresses 5(h, k) as a finite sum, but with fewer terms than the sum in the
original definition.
73
4
Congruences for the coefficients
of the modular function}
4.1 Introduction
The function j(,) = 12 3 J(,) has a Fourier expansion of the form
100.
j(,) = - + L c(n)xn, (x =e 21t1t )
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 211),
c(3n) == 0 (mod 35 ),
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
15(') = L c(5n)xn
n:l
74
4.2: The subgroup ro(q)
75
4: Congruences for the coefficients of the modular function j
~) to get
= 1 = (_ ~ ~) = (~~ =~ ~).
Choosp. k to be that solution of the congruence
kC == D (mod p) with 0:::; k < p.
This is possible since C¢:O (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 =- 1/r and Tr = r + 1, and let Rr denote the funda-
mental region of r.
We will prove
I T
-I -2
I
o
Figure 4.1 Fundamental region for r 0(3)
V = ST k = (~ -1)
k .
77
4: Congruences for the coefficients of the modular function i
Since VEro(p) this requires k2 == kl (mod pl. But both k l , k2 are in the
interval [0, p - 1], so k 2 = k I' Therefore
V = STOS = S2 = I
Theorem 4.3. For any prime p> 3 the subgroup ro(p) has 2[p/12] +3
generators and they may be selected from the following elements,'
V, - STkST-k'S _
k - -
(k' I)
-(kk' + I) -k'
where kk' == - 1 (mod p). The subgroup r 0(2) has generators T and VI;
the subgroup r 0(3) has generators T and V2 .
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 VB
VB V9 VI2
VIO VI3 VI3
L
00
f(r) = ane27tin,.
n= -m
78
4.4: Functions automorphic under the subgroup r o(p)
V = PA b
where Ak = ST k if k < p, and Ap = I. For each k = 0, 1, ... , p, let
r k = {PAk:PEro(p)}·
Each set r k is called a right coset of r o(P). Choose an element fk from the
coset r k and define a function fk on H by the equation
fk(T) = f(fk T).
Note that fp(T) = f(PT) = f(T) since PEr o(P) and f is automorphic under
ro(P). The function value h(T) does not depend on which element fk was
chosen from the coset r k because
fk(T) = f(fk T) = f(PA kT) = f(Ak T)
and the element Ak is the same for all members of the coset r k.
How doesfk behave under the transformations of the full modular group?
If V E r then
fk(VT) = f(fk VT).
Now fk V E
such that
r so there is an element Q in r o(P) and an integer m, °: ; m ::; p,
Therefore we have
79
4: Congruences for the coefficients of the modular function.i
Then if V E r we have
cp(Vr) = n {h(Vr) -
p
k=O
f(w)} = n {f,1(kir) -
p
k=O
f(w)} = cp(r),
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
n {h(O -
p
o= f(w)}
k=O
hence some factor is o. In other words, f(w) = h(O 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
= L a(n)e21tint
00
f(r)
n= -m
L
00
fp(r) = a(np)e21tint.
n= -[m/pl
1
L a(n)e 21t ;nt/ p L e 21t ;n)./p.
00 p-l
= -
p n=-m ),=0
80
4.5: Construction of functions belonging to r o(p)
But
if p,{'n
if pin
so
L L
00 00
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 offunctions
meromorphic in H.
Next we must show that
f/Vr) = fp(r) whenever V E r o(P).
F or this we use a lemma.
Lemma 1. If V E ro(P) and if 0 :$ .Ie :$ p - 1, let T).! = (r + .Ie)/p. Then
there exists an integer f1., 0 :$ f1. :$ P - 1 and a transformation in a;,
r O(P2) such that
Moreover, as .Ie runs through a complete residue system modulo p, so does f1..
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
PROOF OF LEMMA 1. Let V = (: ~), where c == 0 (mod p), and let .Ie be
or
{
A = a + AC
(2)
C = pc
(3) {All + Bp = b + Ad
CIl + Dp = pd
with
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)1l + Bp = b + Ad
CPIl + Dp = pd.
Choose Il to be that solution of the congruence
Ila == b + Ad (mod p)
which lies in the interval 0 S Il S P - 1. This is possible because ad - bc = 1
and pic imply p,( a. Note that distinct values of A mod p give rise to distinct
values of Il mod p. Then, since p Ie we have
Ila + IlAC == b + Ad (mod p)
or
(a + AC)1l == b + Ad (mod p).
Therefore there is an integer B such that
(a + AC)1l + Bp = b + Ad.
Therefore the first relation in (4) is satisfied. The second relation requires
D = d - ell. Thus, we have found integers II, A, B, C, D such that
f, (-
P
!)
r
= f, (r) + !p f(pr) - !P f(~).
P P
To prove this we need another lemma.
or
-1)o = (a
c
aJ1
CJ1
+ bP).
+ dp
Take a = A, C = P and let J1 be that solution of the congruence
AJ1 == -1 (mod p)
in the interval 1 ::; J1 ::; p - 1. This solution is unique and J1 runs through a
reduced residue system mod p with A. Choose b to be that integer such that
aJ1 + bp = -1, and take d = - J1. Then CJ1 + dp = 0 and the proof is
complete. 0
PROOF OF THEOREM 4.6. We have
( 1)
pfp - - = L f (Sr+A)
p-!
--' =
(sr) + L
f -
p-!
f(TlSr)
r l=O p P l= 1
83
4: Congruences for the coefficients of the modular function;
L ,(n)e 2 "int
00
~(,) = (2n)12
n= 1
where the T(n) are integers with T(1) = 1 and T(2) = - 24. However, ~(,) is
not invariant under all transformations of r. In fact we have
In particular,
so
<p(r) = ~(qr) =
~(r)
xq-I 1 + L:'=
1 r(n + l)x nq
1 +L:'=Ir(n+ l)x n
= X q- I(1 + f b xn)
n=I n
W=(a bq ).
CI d
But WE r because det W = ad - bClq = ad - bc = 1. Hence
~(qVr) = ~(W(qr)) = (cI(qr) + d)I2~(qr),
so
~(qVr) (clqr + d)I2~(qr)
<p(Vr) = ~(Vr) = (clqr + d)I2~(r) = <per).
<p ((ji'-1) 1
= qI2<p(r)'
Hence <per) -. 00 as r -. O.
85
4: Congruences for the coefficients of the modular function j
~( _ :r) = (qr)12~(qr)
so
~(r) = (2n)12x TI (1
00
- xn)24
n= 1
so
and the Fourier series for cp~(r) will certainly have integer coefficients if
24ex is an integer, that is, if q - 1 divides 24. This occurs when q = 2, 3, 4, 5,
7, 9, 13, and 25.
The function <1> so defined is analytic and nonzero in H. The Fourier series
for <1> in (5) shows that <1> has a first order zero at 00 and that
1 1
<1>(T) = ~ + lex),
where lex) is a power series in x with integer coefficients.
Since cp is automorphic under ro(q) we have cp(VT) = cp(T) for every
element V of r o(q). Hence, extracting roots of order q - 1, we have
<1>(VT) = c:<1>(T)
where c: q - 1 = 1. The next theorem shows that, in fact, c: = 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 <1> is automorphic under ro(q).
Theorem 4.9. Let q = 2,3,5,7,01' 13, and let I' = 24/(q - 1). Then thefimction
(8)
87
4: Congruences for the coefficients of the modular function.i
We also have
a(qr) + bq )
l1(qVr) = 11( cl(qr) + d = I1(Vl qr)
where
v
I
= (ac i
bdq) .
Since VI E r we have
. = {a~
(j
+ d + s(-d,c)} - {a12cI
+ d + s(-d,cd } .
fir) = ~
p
Pf f(r + A)
).=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 j p(r)
we shall express j p as a polynomial in the function <1>.
In deriving the differential equation for the Weierstrass f,J function we
formed a linear combination of f,J, f,J2 and f,J3 which gave a principal part
near z = 0 equal to that of [f,J'(z)Y The procedure here is analogous.
Both functions j p and <I> have a pole at the vertex r = 0 of the fundamental
region of r o(p). We form a linear combination of powers of <I> to obtain a
principal part equal to that of j P'
88
4.10: The function j p expressed as a polynomial in <ll
j (-
p
!)
r
= j (r)
P
+ !p j(pr) - !P j(::')
P
j (-
P
~)
pr
= j (pr)
P
+ !p j(p2r) - !P j(r).
Hence ifx = e 2nir we have the Fourier expansion
pj
p
(-~)
pr
= x- p2 _ x- 1 + I(x),
where I(x) is a power series in x with integer coefficients.
PROOF. We have
j(r) = x- 1 + c(O) + c(l)x + c(2)x 2 + ... ,
jp(r) = c(O) + c(p)x + C(2p)X2 + ... ,
pjp(pr) = pc(O) + pc(p)xP + pC(2p)X2p + ... ,
and
so
= x- p2 _ x- 1 + I(x). o
Now we can express jp as a polynomial in <1>.
89
4: Congruences for the coefficients of the modular function j
pj (-
P
~)
pT
- {1jJ(T)}p2
pj ( 1)
P
- -
pT
- {1jJ(T)}P 2 - b l {1jJ(T)}P , - I
f(-~)
pT
= pj (-~)
pTP
- {1jJ(T)}P' -bdljJ(TW'-1 - ... - b P'- I IjJ(T)
90
Exercises for Chapter 4
Since it is known that e(13) is not divisible by 13, congruences of the above
type cannot exist for 13. In 19S8 Morris Newman [30J found congruences
of a different kind for 13. He showed that
if z # 0 and Ix I < 1.
(a) Show that x and z can be chosen to give the product representation
oc
;1(r) = f1 (I - e 2 'in')(1 + e(2n-l)'i')2.
n=1
(
.1(T) =
~2(' : 1)
.
~(r + I)
(c) Prove that:j( -I/T) = (-ir)I/2:1(r).
Hint: If Sr = -I/T, find elements A and B of r such that
sr; 1 A(r : I)
= and Sr + 1 = B(r + I).
91
4: Congruences for the coefficients of the modular function j
2. Let G denote the subgroup of r generated by the transformations Sand T2, where
Sr = - l/r and Tr = r + 1.
I) ( cr +
ar b)
+ d = e(a, b, c, d){ - i(cr + d)} 1/21)( r),
_ e21timr)24.
n = 1 m=l
3. Let p be a prime and let k be an integer, 1 :0; k :0; P - 1. Show that there exists an
integer h such that
rl2d(r ; h) = d(krp~ 1)
and that h runs through a reduced residue system mod p with k.
4. If p is a prime, define
Fp(r) = p11d(pr) Ld
+ -1 p-I (r + k)
-- .
P k=O P
Prove that:
(-1)
(b) Fp -r- = r 12 Fir).
92
Exercises for Chapter 4
8. Prove that
r(m)r(n) = L dllr(m~).
dilm.n) d
In particular, when (m, n) = 1 this implies r(m)r(n) = r(mn).
9. If t E H and x = e 2 • iT prove that
{504JOO"s(n)xnf = {j(t)- 12 3 }J 1
r(I1)X n,
This formula, due to Lehmer [20], can be used to determine the coefficients c(1l)
recursively in terms of t(n). Since the right member is an integer. the formula also
implies Ramanujan's remarkable congruence
93
5
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
f1 L p(n)x",
00 00
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 [4]. This chapter
is concerned with the behavior of p(n) for large n.
The partition function p(n) satisfies the asymptotic relation
eKJR
p(n) '" Ii as n -+ 00,
4ny 3
where K = n(2/3)1/2. This was first discovered by Hardy and Ramanujan [13]
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
94
5.2: The plan of the proof
(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 ofthe Pk(n) in (2). This had a profound
effect on the final result since, instead of (2), Rademacher obtained a con"
vergent series,
00
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 .jn, 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 I1(r).
F(x) =
n+ 1
~
L...,
p(k)x k
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'
m
i F(x)
Ii'TI 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,
x 3 = 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
Ch,k lying near the roots of unity e21tih/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,
On each arc Ch,k the function F(x) in the integrand is replaced by an elemen-
tary function t/lh.k(X) which has essentially the same behavior as F near the
singularity e21tih/k. This elementary function t/lh,k arises naturally from the
functional equation satisfied by the Dedekind eta function I'/(r). The functions
F and 1'/ are related by the equation
96
5.4: Farey fractions
If
Note. Izl is small, the point x in (5) lies near the root of unity e27tihlk,
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 e27tihlk. Aside from a constant
factor, for small Iz I, F behaves like
Zl/2 ex p ( I;Z).
PROOF. If(: ~) E r with c > 0, the functional equation for '1(T) implies
(7) 1 _ 1.
'1(r) - '1(T') { -1(CT + d)} 1/2 exp {.(a~
1tl
+ d + s( -d, c))},
where r' = (ar + b)/(n + d). Since F(e 27tit ) = e7tit/121'1(r), (7) implies
a = H, c = k, d = - h, b =
hH + 1 and
iz + h
T=-k-·
k
Then
, iz- I +H
T=-~-
k
and (8) becomes
( (2ltih
F exp - - - -
k
2ltZ)) _- F(exp(2ltiH
k
- - - -2lt))z 1/2
k kz
1t
x exp{ 12kz - 12k
ltZ
+ nis(h, k) } .
When z is replaced by z/k this gives (6). o
97
5: Rademacher's series for the partition function
Definition. The set of Farey fractions of order n. denoted by Fn. is the set of
reduced fractions in the closed interval [0. 1] with denominators ::;n.
listed in increasing order of magnitude.
EXAMPLES
F1: ¥. t
F 2 : ¥. t. t
F 3: ¥. t. t. i. t
F4 : ¥.i.t.t.i.i.t
F 5: ¥. t. i. t. ~. t. i. j. t t t
F6: ¥.i.t.i.t.~.t.i.i.i.!.i.t
F7: ¥.+.i.!.i.~.t.~.i.t.4.i.i.~.i.!.i.~.t
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. D
Theorem 5.3. Given 0 ::; alb < cld ::; 1. If bc - ad = 1 then alb and cld are
consecutive terms in Fnfor the following values of n:
max(b. d) ::; n ::; b +d- 1.
PROOF. The condition bc - ad = 1 implies that alb and cld are in lowest
terms. If max(b. d) ::; n then b ::; nand d ::; n so alb and cld are certainly in
Fn. 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 Fn. This completes the proof. D
98
5.5: Ford circles
Theorem 5.4. Given 0 :s; alb < c/d :s; 1 with be - ad = 1, let h/k be the
mediant q{ alb and c/d. Then alb < h/k < c/d, and these fractions satisfy
the unimodular relations
bh - ak = 1, ck - dh = 1.
PROOF. Since h/k lies between alb and c/d 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
Theorem 5.5. The set Fn + 1 includes Fn' Each fraction in Fn + 1 which is not in
Fn is the mediant of a pair of consecutive fractions in Fn. Moreover, if
alb < c/d are consecutive in any Fn, then they satisfy the unimodular
relation bc - ad = 1.
PROOF. We use induction on n. When 11 = 1 the fractions 0/1 and 1/1 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 c/d are consecutive in Fn
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) :s; m :s; b + d - 1.
Form their mediant h/k, 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 c/d are consecutive in Fm for all m satisfying
max(b, d) :s; m :s; b + d - 1, but are not consecutive in Fk since k = b + d
and h/k lies in Fk between alb and c/d. But the two new pairs alb < h/k and
h/k < c/d are now consecutive in Fk 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 Fn to Fn + 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
. 1
radiUS = 2k2
h
k
Theorem 5.6. Two Ford circles qa, b) and qc, d) are either tangent to each
other or they 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)
D1 = ( ba - C)2
d +
(12b1- 2d1
1)1'
a c
b d
Figure 5.2
(r +
1
Rf = ( 2b1
1
+ 2d1 .
)2
The difference D2 - (r + R)l is equal to
Theorem 5.7. Let hdkl < h/k < h2/k2 be three consecutive Farey fractions.
The points of tangency of C(h, k) with C(h l , kd and C(h 2, k 2) are the points
h kl i
ttl(h, k) = k- k(k 2 + k 1 2) + k2 + k/
and
Moreover, the point of contact rtl(h, k) lies on the semicircle whose diameter
is the interval [hl/k l , h/k].
PROOF. We refer to Figure 5.3. Write rtl for rtl(h, k). The figure shows that
b
Q
2k/
,
---'--,-::,.----- ---- --- - --
"
;ii',
,
,,
,,
,
h
k
Figure 5.3
k/
so
Similarly, we find
b 20 - 2k;2 1 k/ - k 2
so b = 2k2 k 2 + k12·
1 1
2k2 2k2 + 2kl2
These give the required formula for rt l , and by analogy we get the correspond-
ing formula for rt 2 .
101
5: Rademacher's series for the partition function
To obtain the last statement, it suffices to show that the angle in Figure e
5.3 is n12. For this it suffices to show that the imaginary part of ril(h, k) is
the geometric mean of a and d, where
kl , h hi 1
a = k(k 2 + k 1 2 ) and a = k - k"; - a = kk 1 - a.
a
Figure 5.4
Z . 2( kh)
= -Ik r -
102
5.6: Rademacher's path of integration
i + I
o 1
:3
1
2
2
:3
maps the Ford circle C(h, k) in the T-plane onto a circle K in the z-plane of
radius! about the point z = ! as center (see Figure 5.6). The points of
contact rJ.1(h, k) and rJ.z{h, k) of Theorem 5.7 are mapped onto the points
k2 kk
zl(h,k) = k 2 + k 1 Z + i k Z + lk/
and
The upper arc joil1il1g rJ.1(h, k) with rJ.z(h, k) maps onto that arc of K which
does 110t touch the imaginary z-axis.
z-plane
Figure 5.6
103
5: Rademacher's series for the partition function
PROOF. The translation l' - (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 zl(h, k) and z2(h, k) follow at once. 0
Now we obtain estimates for the moduli of Z 1 and Z2 .
yfik
(11) Izi < t:/'
if hl/kl < h/k < h2/k2 are consecutive in F N' The length of this chord does
not exceed 2yfik/N.
PROOF. For IZ112 we have
e + k2k 2
IZ112 = (k 2 + kJf + k/' k2
There is a similar formula for IZ212. This proves (10). To prove (11) we note
that if z is on the chord, then IzI ~ max( Iz 1 I, Iz21), so it suffices to prove that
yfik yfik
(12) IZ11<N and IZ21<N'
For this purpose we use the inequality relating the arithmetic mean and the
root mean square:
k _
_ k1 + < (k2 + k 1 2)1/2
2 - 2 .
This gives us
(k 2 k 2)1/2 k + kiN + 1 N
+ 1 ~ yfi ~ yfi > yfi'
so (10) and (12) imply (11). The length of the chord is ~ IZI I + IZ21. 0
p(n)
1
= n yfi2 k=L1 Ak(n)jk -dn
00 d (Sinh{ ~ JfF1)f)
R 1
/1--
24
104
5.7: Rademacher's convergent series for p(n)
where
PROOF. We have
Xm)-l =
<Xl
L p{n)xn;
n=O
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
p(n) = r
J
i+ 1
F(e 21!i')e- 2 1!inr dr = r
Jp(N)
F{e 2"i')e- 2 "inr dr.
i
In this discussion the integer n is kept fixed and the integer N will later be
allowed to approach infinity. We can also write
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
h,k z,(h,k)
2nz))
-- - -
k k
2 2i
k
.
e-2nlnh/k e2nnz/k 2 dz
= L ik - 2 e - 2ninh/k f Z2
(h, k) e2nnz/k2 F(ex p(2nih _ 2n2z)) dz,
h,k zI(h,k) k k
Now we use the transformation formula for F (Theorem 5.1) which states
that
Z)I/2
F(x) = w(h, k) ( k
( n
exp 12z - 12e
nz )
F(x'),
where
2nih 2nz)
x = exp ( -k- - k2 ' x , = exp (2niH 2n),
-k- - ~
and
w(h, k) = enis(h, k), hH == - 1 (mod k), (h,k) = 1.
Denote the elementary factor Zl/2 exp[7l'/(l2z)- nz/(12k 2 )] by '1\(z) and
split the integral into two parts by writing
F(x') = 1 + {F(x') - I}.
We then obtain
p(n) = L ik-s/2w(h,k)e-Zninh/k(ll(h,k) + 12(h,k»
h, k
where
11(/1, k) = f Z2 (h,k)
zdh.k)
'1\(z )e2nnz/k2 dz
f
and
I 2(h, k) = (h,k) '1\(Z){F(ex p(2:iH _ 2n)) _ 1}e2nnz/k2 dz.
Z2
z,(h,k) Z
We show next that 12 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 dh, 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)
z tfh, k)
o\
Figure 5.8
chord; it does not exceed 2.j2k/N. On the chord itself we have Izl
::; max{IZ11, 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(1/z) ~ 1, while on K itself we have
Re{1/z) = 1.
Now we estimate the integrand on the chord. We have
L p(m)e- 21t(m-(1/24))
00
::; Izll/2e 2nrr
m=l
L p(m)e- 2rr(24m-l)/24
00
= Izll/2e 2nrr
m=1
L p(24m -
C£
107
5: Rademacher's series for the partition function
where
L p(24m -
00
c = e Zn1t 1)yZ4m-1.
m= 1
L L
N
(14) p(n) = ik-SIZw(h,k)e-Z1tinhlkl1(h,k) + O(N- 1/2 ).
k= 1 05h<k
(h,k)= 1
Next we deal with 1 1(h, k). This is an integral joining zl(h, k) and zz(h, k)
along an arc of the circle K in Figure 5.8. We introduce the entire circle K
i
as path of integration and show that the error made is also O(N- 1/ Z ). We have
11(h,k) = i i
K(-)
-
0
z llh, k)
-
fO
z2(h,k)
=
K(-)
-J 1 -J 2 ,
Since Re(1/z) = 1 and 0 < Re(z) :s; 1 on K the integrand has absolute value
so that
108
5.7: Rademacher's convergent series for p(n)
p(n) = I
k~ I Osh<k
I ik-5/2w(h,k)e-21tinhik r
JK(-)
\l'k(z)e2n1tz/k2 dz + O(N- 1/2).
(h, k) ~ I
p(n) = ik~IAk(n)k-5/2
00
JrK (_)ZI/2 exp {127
n + k2
2nz ( n - 1 )} dz,
24
where
Ak(n) = I e"is(h, k)- 21tinh/k.
Osh<k
(h, k)~ I
p(n) = 2nCn2Y/2 J/k(n)k- 5/2 2~i f_+:i it - 5/2 exp{t + 6nk22 (n - 2~)ndt
where e = n/12. Now on page 181 of Watson's treatise on Bessel functions
[53] we find the formula
} = {::2 (n - ;4)f/2
and v = 3/2 we get
)-3/4
1))
1
3 2(
n)3/2 n- / n - 24 (n ~(
k v"3 "n - 24.
-52
p(n) = (2n) ( 12
00
109
5: Rademacher's series for the partition function
But Bessel functions of half odd order can be reduced to elementary functions.
In this case we have
13 / 2 (z) _- ~z d (sinh
- -d ~~. z)
n Z Z
2. If a, b, c, d are positive integers such that alb < cld and if ;, and II are positive integers,
prove that the fraction
lies between alb and cld, and that (c - del/(eb - a) = ;./11. When A = fl, 0 is the
mediant of alb and cld.
3. If bc - ad = I and 11 > max(b, d), prove that the terms of the Farey sequence F"
between alb and cld are the fractions of the form (Aa + flC)/(Ab + luI) for which ).
and fl are positive relatively prime integers with ;,b + lid s 11. Geometrically. each
pair (A, fl) is a lattice point (with coprime coordinates) in the triangle determined by
the coordinate axes and the line bx + ely = 11. Neville [29] has shown that the
number of such lattice points is
3 112
2- + O(nlog/J).
TC bel
This shows that for a given n, the number of Farey fractions between alb and eld is
asymptotically proportional to l/(bd), the length of the interval [alb, ('1£1].
110
Exercises for Chapter 5
Also, T,,' denotes the set of lattice points (x, y) in T" with relatively prime
coordinates.
4. Prove that alb and cld are consecutive fractions in the Farey sequence F" if, and only
if, the lattice point (b, d) E T~.
(a) By comparing the regions T, and T,-l for I' ::::: 2 show that
r-l r-I
(15) L
(x. y) e T~
f(x, y) = f(1, I) +
r=2
L L k::;: 1
{.f(k,l') + f(l', k) - f(k, I' - k)}.
(k.r)= 1
This relates a sum involving Farey fractions to one which does not.
7. Let
1
S" =
(b, d)
L T~ bd(b + d) .
E
(a) Use Exercise 5 to show that 1/(2n - I) ::; S" ::; I/(n + I).
(b) Choose f(x,·),) = I/(xy(x + y)) in (15) and show that
3 "r 1
S" = - - 2L L 2 •
2 r= 1 k= 1 I' (I' + k)
(k,r)= 1
<X) r 1 3
r=
L1 L1 k=
1'2(1' + k) = 4'
(k.r)= 1
8. Exercise 7(a) shows that S" --> 0 as n --> x,. This exercise outlines a proof of the
asymptotic formula
(16) S" -
_ 12 log 2
2 +0
(lOg n)
7! n n2
obtained by Lehner and Newman in [25].
111
5: Rademacher's series for the partition function
Let
Ar = ~
k=l r2(r
1
+ k)
= I L Il(d)
k=l dllr.k)r 2(r + k)'
(k.r)= 1
so that
r>n
(a) Show that
Ar = L
d/r
f
h= 1
---.:--:-::-Il_(r/_d--::-)
r (h + d)
qJ(r)
Ar = log 2 - 3 + 0 ( 3"1 L 1J1(d) I) .
r r d/r
(b) Show that L~= 1 Ld/r IIl(d) I = O(n log n) and deduce that
1 (lOg
L 3" L IIl(d) I = 0 - 2
n) .
r> n r d/r n
(c) Use the formula Lr,; n q>(r) = 3n 2 /n 2 + O(nlog n)(proved in [4], Theorem 3.7)to
deduce that
112
6
Modular forms with
multiplicative coefficients
6.1 Introduction
The material in this chapter is motivated by properties shared by the discrimi-
nant ~(t) and the Eisenstein series
1
G 2k(t) = L
(m,II)"(O.O)
(m + nt )2k'
where k is an integer, k ~ 2. All these functions satisfy the relation
(1) at + b)
f ( ct + d = (Ct + d)'f(t),
L t(n)e2ltillt,
00
~(t) = (2n)12
n=1
where 11,,(n) is the sum of the ~th powers 'of the divisors of n.
113
6: Modular forms with multiplicative coefficients
Both r(n) and O"a(n) are multiplicative arithmetical functions; that is, we
have
(2) r(m)r(n) = r(mn) and O"a(m)O"a(n) = O"a(mn) whenever (m, n) = 1.
They also satisfy the more general multiplicative relations
and
(4)
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 [16] in 1937 and his solution is discussed in this chapter.
L c(n)e21tinr.
00
f(r) =
n~O
The constant term c(0) 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) =1= 0 is called the order of the zero of fat ioo.
It should be noted that the discriminant ~ is a cusp form of weight 12 with
a first order zero at ioo. Also, no Eisenstein series GZk vanishes at ioo.
114
6.3: The weight formula for zeros of an entire modular form
Theorem 6.1. Let f be an entire modular form of weight 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
N = _1
2n:i
f
oR
f'(r) dr.
f(r)
115
6: Modular forms with multiplicative coefficients
c.
f'(A(.»A'(.) 1'(.)
= --
~--,--,----,--,-- + -kc
-.
f(A(.» f(.) +d
Consequently, for any path y not passing through a zero we have
_1
2ni
f A(y)
f'(u) du = _1
f(u) 2ni
f
y
1'(.) d.
f(.)
+ _1
2ni
f ~d•.
y C't +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 ~d.
p = 2ni
-k -k (ni
(log i-log p) = 2ni 2" -"3
2ni)
=
k
12'
The rest of the proof is like that of Theorem 2.4 and we obtain (6), which
implies (5). 0
Theorem 6.2
(a) The only entire modular forms of weight k = 0 are the constant
functions.
= 2, the only entire modular form of weight k
(b) Ifk is odd, ifk < 0, or ifk
is the zero function.
(c) Every nonconstant entire modularform 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 =F O. Also,
if k < 12 then N(ioo) = 0 so f is not a cusp form unless f = O. 0
116
6.4: Representation of entire forms in terms of G4 and G6
117
6: Modular forms with multiplicative coefficients
°
need only verify that the products Gk - 12 ,!1' are linearly independent. This
follows easily from the fact that !1(ioo) = but G 2 ,(ioo) =f. 0. Details are left
as an exercise for the reader. 0
(8) f = L ca,bG4aG6b
a, b
°
where the Ca, b are 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 f/G 4 ,f/G 6 ,f/G42 and f1(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 = G/G 6s has weight k and does not vanish at ioo.
Hence if c = f(ioo)/g(ioo) the entire form f - cg is a cusp form in Mk so
f - cg =!1h 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 !1 gives a sum of the same type with 4a + 6b = k.
Hence f = cg + !1h is also a sum of the required type and this proves the
theorem. 0
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
Mk is given by the formulas
if k == 2 (mod 12),
(9)
if k 1= 2 (mod 12).
Another basis for Mk is the set of products G4 a G6 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 Mk - 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 = Ah, where
hEM k_ 12 . Therefore the linear transformation T : M k- 12 ...... M k, 0 defined by
T(h) = llh
establishes an isomorphism between M k, 0 and M k _ 12' Consequently, if
k 2:: 12 we have
(11) dim Mk,o = dim M k- 12 •
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 Mk = 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 , G/, G4 G6 , and G/G 6 ·
Formulas (11) and (9) together show that
dim Mk,o = 1 if k = 12,16,18,20,22, and 26.
Corresponding basis elements are ll, llG 4 , llG 6 , llG 4 2 , llG 4 G6 , and llG/G 6 .
Theorem 6.5. Let f be an entire form of weight k and let Zl' ... , ZN denote the
N zeros off in the closure of Rr (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
TI
N
(12) f(r) = cG4(r)N(p)G6(r)N(i)~(r)N(iOO)~(rt {J(r) - J(Zk)}'
k=l
PROOF. The product
N
g(r) = TI {J(r) - J(zd}
k=l
is a modular function with its only zeros in the closure of Rr at Zl' ... , 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 Zb ... , 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.
Definition. For a fixed integer k and any n = 1, 2, ... , the operator 1'" is
defined on Mk by the equation
(13)
120
6.7: The Heeke operators T"
where
(16)
(Tnf)(,) = nk - I
d-I
L d- L L C(m)e "im(nt+bd)/d
k
00
2 2
)k-I d-I
din b=O m=O
1
= L L
00 (
~ c(m)e 2 "imnt/d 2 - L e 2 "imb/d.
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.
()k-I
Hence
(T"f)(,) = L L
00
~ c(m)e 2 "imnt/d 2 •
m=O dln.dlm d
Writing m = qd we have
(T"f)(,) = f
Q=O din
L (~)k-I c(qd)e 2 "i nt/d.
d
Q
(Tnf)(,) = f Ld
q=Odln
k- 1 C(qn)e 2 "iQdt.
d
If x = e 2 "it 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
(Tn f)(,) = L L
00
121
6: Modular forms with multiplicative coefficients
(a
o
b)
d'
where d > o.
122
6.8: Transformations of order 11
v = (~ ~}
Then V Er and
(18)
where d runs through the positive divisors of n and, for each fixed d,
a = nld, and b runs through a complete residue system modulo d.
PROOF. Theorem 6.7 shows that every element in nn) is equivalent to one
of the transformations in (18). Therefore we need only show that two such
transformations, say
al
Al = ( 0 and
V = (~ i).
Then V A I = A 2 so A I '" A 2·
Conversely, if Al '" A2 there is an element
V = (~ ~)
in r such that A2 = VAI. Therefore
(20)
Note. The sum in (17) defining T"f can now be written in the form
1
(21) (T"f)(r) = -
n
L akf(Ar),
A
Ai = (~ ~:) and
_ 1
A2 -1 - - (d
n 0
z
d 1d 2 Y1 d2
Y2 = - - = - Y l
n a1
and
b2 = -d 1Y1 b 2 + d 1b 1a z = _ b2 Y1 + a2 b 1
n a1 a1
124
6.9: Behavior of T" f under the modular group
where A I = ( a0
l hi)
dI and A I runs through a complete set of noneqUlvalent .
elements in r(n). Replacing r by Vr we find
1
(25) (T"f)(Vr) = - L a/f(A I Vr).
n Al
By Theorems 6.7 and 6.9, there exist matrices
such that
and
Therefore
a/f(AIVr) = a/f(V2A 2r) = a/(Y2 A2r + b 2 tf(A 2 r)
= a/(yr + b)kf(A 2 r)
since f E M k • Now as Al runs through a complete set of nonequivalent
elements of r(n) so does A 2 . Hence (25) becomes
1
(T" f)( Vr) = - (yr + b)k L a/ f(A2 r) = (yr + b)k(T" f)(r). D
n A,
125
6: Modular forms with multiplicative coefficients
The next theorem shows that each Hecke operator 1'" maps Mk into Mk
and also maps Mk,o into Mk,o,
where
C = BA = (~ ~) ( ~ ~) = (~ ab ~ 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
Theorem 6.13. Any two Hecke operators T(n) and T(m) defined on M k commute
with each other. Moreover, we have the compositionjormu/a
(28) T(m)T(n) = L dk- 1 T(mn/d 2 ).
d!<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 pr where 0 s t S r. Hence we have
By (14) we have
+ p- 1- r L L ! (r-r
p(r - r)k p-l p r +r +b ~ + bp . r)
OsrSr b=O P
OSb,<p'
In the second sum the linear combination br + bpr runs through a complete
residue system mod pr+ 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
127
6: Modular forms with multiplicative coefficients
Equating the two expressions, solving for T(pr + 1 )T(pS) and using (29) in the
sum on t we find
r r
T(pr+l)T(ps) = Lpl(k-l)T(pr+S+l-U) + LP(I+l)(k-l)T(pr+s-I-2r)
1=0 1=0
_ pk-1T(pr-l)T(ps).
By the induction hypothesis the last term cancels the second sum over t
except for the term with t = r. Therefore
r
T(pr+l)T(pS) = Lpt(k-l)T(pr+s+1-21) + p(r+1)(k-l)T(ps-l-r)
1=0
r+ 1
= L p,(k-l)T(pr+l+S-U).
1=0
This proves (31) by induction for all r and all s ~ r, and also completes the
proof of (28). 0
128
6.11: Eigenfunctions of Hecke operators
where
c(n)c(m) = L dk-lC(;~).
dl(n,m)
129
6: Modular forms with multiplicative coefficients
for allm ;;::: I, n ;;::: I, ill which case the coeffiCient c(n) is an eigenvalue of T".
130
6.13: Examples of normalized simultaneous eigenforms
Theorem 6.16. Assume that I E M 2" where k ;:::; 2, alld that I is Ilot a cusp
form. Theil I is a normalized simultaneous eigen{orm if, alld only if,
. (2k - 1)!
(41 ) fir) = 2(2ni)2k G2k(r).
PROOF. In the Fourier expansion (32) we have c(O) =f. 0 since f is not a cusp
form. The relation
(42) T,J = A(n)f
On the other hand, (35) implies )In(O) = 0' 2k _ 1(1l)c(0) smce f EM 2k' But
dO) =f. 0, so Equation (42) holds if, and only if,
A(n) = 0'2k-l(Il).
131
6: Modular forms with multiplicative coefficients
the constant term in (44) is equal to - B2k/(4k). (See [4], Theorem 12.17.)
We can also write
Since the eigenvalue A,(n) in (42) is (J"2k-l(n), Theorem 6.16 shows that the
divisor functions (J"in) satisfy the multiplicative property in Equation (4)
when IX = 2k - 1. Actually, they satisfy (4) for all real or complex IX, but
(J"Cl(n) is the nth coefficient of an entire form only when IX is an odd integer ~ 3.
132
6.14: Remarks on existence of simultaneous eigenforms in M 2k. 0
(2 ) - 12 ~(T) . G2k -( )
12 T
00
'\' ()
n{ 1 2(2k - 12) '\'
00
( ) m}
1! 2((2k-12) n~ITnx - B 2k - 1Z mf~\O"Zk-13mx .
c(m)c(n) = L d2k-1C(~;)
dl(m.n)
for all m 2: 1, n 2: 1.
133
6: Modular forms with multiplicative coefficients
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) c(n) 1.
= -2
m
fC(v)
~~~
x
dx = flf(U + iv)x- n duo
0
= L Ic(n)le-(n-l)1t.
C1j
A
n=1
This implies
(47)
Now define
g( ,) = 11, - i I= v
134
6.15: Estimates for the Fourier coefficients of entire forms
iftEH. Then
g(At) = let + dl- 2g(t)
if A = (: ~) E r, so gk(At) = let + dl- 2kgk(t). Therefore the product
tp(t) = If(t)lgk(t) = If(t)lv k
is invariant under the transformations of r. Moreover, tp is continuous
in R r , and (47) shows that tp(t) -+ 0 as v -+ + 00. Therefore tp is bounded
in Rr and, since tp is invariant under r, tp is also bounded in H, say
Itp(t) I ~ M
for all t in H. Therefore
If(t)1 ~ Mv- k
for all t in H. Using this in (46) we find
Now
2k-l locI
_
0"2k-l(n) - L
din
(n)
d -_ n
2k - 1
L d 2k-l ~ n 2k - L d 2k -
din
1
d= 1
_
1 - O(n
2k - 1
),
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- O / 4 )+<)
for every 8 > 0, and it has been conjectured that the exponent can be further
improved to k - ! + 8. For the discriminant d, Ramanujan conjectured the
sharper estimate
Ir(p)1 :s;; 2pll/2
L c(n)e2rrint
00
(50) ( ) _ ~ c(n)
cps-L..,.-s
n= 1 n
formed with the same coefficients (except for c(O)). If f EM 2k then c(n) =
O(nk) if fis a cusp form, and c(n) = O(n 2k - 1) iff is not a cusp form. Therefore,
the Dirichlet series in (50) converges absolutely for (1 = Re(s) > k + 1 if f is
a cusp form, and for (1 > 2k if f is not a cusp form.
the Dirichlet series will have an Euler product representation of the form
(53)
136
6.16: Modular forms and Dirichlet series
for each prime p. Using this it is easy to verify the power series identity
EXAMPLE. For the Ramanujan function we have the Euler product represen-
tation
f r(n} = n 1 -
"=1 n p r(p)p
1
S + pll 2s
Theorem 6.20. Let q>(s) be the function defined for (j > k by the Dirichlet
series (50) associated with a modular form f(r) in Mk having the Fourier
series (49), where k is an even integer 2::: 4. Then q>(s) can be continued
analytically beyond the line (j = k with the following properties:
(a) If c(O) = 0, q>(s) is an entire function of s.
(b) If c(O) =F 0, q>(s) is analytic for all s except for a simple pole at s = k
with residue
( -1)k/2c(0)(271:Jk
r(k)
if (j > O. Therefore if (J > k we can multiply both members by c(n) and sum
on n to obtain
137
6: Modular forms with multiplicative coefficients
f:
Since f is a modular form in Mk we have f(ijy) = (iy)kf(iy) so
OO
+ (_1)k/2 fl {f(iw) - c(0)}W k - S- 1 dw
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 cp(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). 0
Heeke also proved a converse to Theorem 6.20 to the effect that every
Dirichlet series cp 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].
138
Exercises for Chapter 6
2. If I and 9 are IX-multiplicative, prove that I +9 is IX-multiplicative if, and only if,
! = 0 or 9 = O.
3. Let II' ... , Ik be k distinct nonzero IX-multiplicative functions. If a linear combination
i= 1
lX(n)!(I11) = I J1(d)!(l11nd)!(~).
dl" d
S. If I is multiplicative, prove that f is IX-multiplicative if, and only if,
(55)
Show that
Q,,(21X(p)I/1X) = lX(p),,/l U ,,(x),
where U ,,(x) is the Chebyshev polynomial of the second kind, defined by the relations
UI(x) = 2x, Ul(x) = 4Xl - I, U d1 (X) = 2xU,(x) - U,_I(X) for r;::O: 1.
7. Let E 2k(,) = -!G lk (,)/((2k). If x = e 2n ;, verify that the Fourier expansion of E lk (,)
has the following form for k = 2, 3, 4, 5, 6, and 7:
n=:1
Cf)
65520 cx·
Ed,) = 1 + - - IO"II(n)x",
691 ,,~I
oc
E I4 (,) = I - 24 IO"13(n)x".
,.,=1
139
6: Modular forms with multiplicative coefficients
n-I
65 691 691
I
n-I
10. r(n) = - O"l1(n) +- 0"5(11) - - 0"5(m)0"5(n - m).
756 756 3 m~1
12. Prove that the products G/G 6 b are linearly independent, where a and b are non-
negative integers such that 4a + 6b = k.
13. Show that the Dirichlet series associated with the normalized modular form
where
and
Peters son conjectured that /'1 and /'1 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
,9('r) = 1 + 2 Lenin".
"=1
(a) If (J > 1 prove that
n-S/2r(~)n-s = 1°Oe-nn'xxs/2-1 dx
n- 2rG),(s)
S/ = 1 00
t/I(x)x s/2 - I dx,
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
7
Kronecker's theorem
with applications
142
7.2: Dirichlet's approximation theorem
e
Given a real number and given 6 > 0, are there integers hand k such that
Ike - hi < 6?
The following theorem of Dirichlet answers this question in the affirma-
tive.
PROOF. Let {x} = x - [x] denote the fractional part of x. Consider the
N + 1 real numbers
0, {e}, {2e}, ... , {Ne}.
All these numbers lie in the half open unit interval 0 :::; {me} < 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 rae} and {be}, where 0 :::; a < b :::; N. Hence we can write
1
(2) I{be} - rae} I < N'
But
{be} - rae} = be - [be] - ae + Cae] = (b - a)() - ([be] - Cae]).
Therefore if we let
k=b-a and h = [be] - Cae]
inequality (2) becomes
1
Ike - hi < N' with 0 < k :::; N.
Theorem 7.2. Given any real e and any positive integer N, there exist relatively
prime integers hand k with 0 < k :::; N such that
1
Ike - 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
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
hili
If) - k <
Nk' < Nk'
Theorem 7.3. For every real f) there exist integers hand k with k > 0 and
(h, k) = 1 such that
PROOF. Part (a) is merely a restatement of Theorem 7.3. To prove (b), assume
f) is irrational and assume also that S(f)) is finite. We shall obtain a contra-
diction. Let
rJ. = mm . If) -
(h, k) E S(9)
hi
-.
k
Since f) is irrational, rJ. is positive. Choose any integer N > l/rJ., for example,
N = 1 + [1/rJ.]. Then liN < rJ.. 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
Now 1/(kN) ~ l/k2 so the pair (h, k) E S(8). But we also have
1 1
-<-<01 so
kN - N '
contradicting the definition of 01. This shows that S(8) cannot be finite if 8 is
irrational.
To prove (c) assume that all pairs (h, k) in S(8) 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(8) we have
1
1 k8 - hi < k~ 1,
so
1h 1= 1h - k8 + k8 1~ 1h - k81 + 1k8 1< 1 + 1k8 1~ 1 + MIG I·
Therefore the number of choices for h is bounded, contradicting the fact that
S(8) is infinite.
To prove (d), assume 8 is rational, say G = alb, where (a, b) = 1 and b > O.
Then the pair (a, b) E S(8) because 8 - alb = O. Now we assume that S(G)
is an infinite set and obtain a contradiction. If S(8) is infinite then by part (c)
there is a pair (h, k) in S(G) with k > b. For this pair we have
o < I~b ~I
k - ~
< k2 '
from which we find 0 < 1ak - bh 1 < b/k < 1. This is a contradiction because
ak - bh is an integer. 0
Theorem 7.4 shows that a real number 8 is irrational if, and only if,
there are infinitely many rational numbers h/k with (h, k) = 1 and k > 0
such that
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
(4) Ie - ~I
k
> C(e)
kn '
e
PROOF. Since is algebraic of degree n, e is a zero of some polynomial f(x)
of degree 11 with integer coefficients, say
where f(x) is irreducible over the rational field. Since f(x) is irreducible it
has no rational roots so f(h/k) #- 0 for every rational h/k.
Now we use the mean value theorem of differential calculus to write
e
where ~ lies between and h/k. We wilJ deduce (4) from (5) by getting an upper
bound for 1f'(~)1 and a lower bound for If(h/k)l. We have
(h) n
f k = r ~o ar k = kn
(h)r N
where N is a nonzero integer. Therefore
(6)
which is the required lower bound. To get an upper bound for If'(~) I we let
d= Ie - H
If d > 1 then (4) holds with C(e) = 1, so we can assume that d < 1. (We
e
cannot have d = 1 since is irrationa1.) Since ~ lies between and h/k and e
e
d < 1 we have I~ - I < 1 so
I~I = Ie + ~ - el s; lei + I~ - el < lei + 1.
Hence
1f'(~)1 s; A(e) < 1 + A(e),
where A(e) denote the maximum value of I f'(x) I in the interval Ix I S; Ie I + 1.
Using this upper bound for If'(~) I in (5) together with the lower bound in (6)
we obtain (4) with C(e) = 1/(1 + A(e)). D
146
7.3: Liouville's approximation theorem
(7)
Then we have
h '" 1 1 "'I
0< () - k r =
r
L
m==r+l
10m! ~ lO(r+l)! ~ 10m
m-O
10/9 1 10/9 1
- - - - - - - - <k:
- 10(r+ i)! - k: lor!
-
so (7) is satisfied.
147
7: Kronecker's theorem with applications
°
Note. The same argument shows that 2:~=1 amlO- m! is transcendental if
am = or 1 and am = 1 for infinitely many m.
In other words, the linear form (Jx + y can be made arbitrarily close to by a °
suitable choice of integers x and y. If (J is rational this is trivial because we
can make (Jx + y = 0, so the result is significant only if (J is irrational.
Kronecker proved a much stronger result. He showed that if (J is irrational
the linear form (Jx + y can be made arbitrarily close to any prescribed real
number 0:. We prove this result first for 0: in the unit interval. As in the proof
of Dirichlet's theorem we make use of the fractional parts {n(J} = n(J - [n(J].
Note. This shows that the linear form (Jx +y can be made arbitrarily
close to 0: by a suitable choice of integers x and y.
PROOF. First we note that {n(J} #- {m(J} if m #- n because (J is irrational.
Also, there is no loss of generality if we assume < (J < 1 since n(J =
n[(J] + n{(J} and {n(J} = {n{{}}}.
°
° °
Let I': > be given and choose any 0:, S 0: S 1. By Dirichlet's approxima-
tion theorem there exist integers hand k such that Ik(J - hi < 1':. Now
either k(J > h or k(J < h. Suppose that k(J > h, so that < {k(J} < 1':. (The °
argument is similar if k(J < h.) Now consider the following subsequence of
the given sequence {n(J}:
{k(J}, {2k(J}, {3k(J}, ....
We will show that the early terms of this sequence are increasing. We have
k(J = [k(J] + {k(J}, so mk(J = m[k(J] + m{k(J}.
Hence
Now choose the largest integer N which satisfies {kG} < liN. Then we have
1 1
- - < {kG} <-
N + 1 N·
Therefore {mkG} = m{kG} for m = 1,2, ... , N, so the N numbers
{kG}, {2kG}, ... , {NkG}
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 < {NkG} < 1,
N+
or
1
1- N + 1< {NkO} < 1.
Thus {NkO} differs from 1 by less than 1/(N + 1) < {kO} < c. Therefore the
first N members of the subsequence {nkO} subdivide the unit interval into
subintervals of length < B. Since a lies in one of these subintervals, the
theorem is proved. 0
Theorem 7.8. Givell any real a, allY irratiollal 8, alld all.\' > 0, there exist
°
f.
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 (Jl and 2(Jl, and two real
numbers 1X1 and 1X2' and suppose there exist integers hI, h2 and k such that
Ik(Jl - hI - 1X11 < f:
and
12k(J1 - h2 - 1X21 < f:.
Multiply the first inequality by 2 and subtract from the second to obtain
12hl - h2 + 21X1 - 1X21 < 3f:.
Since f:, 1X1 and 1X2 are arbitrary and hl, h2 are integers, this inequality cannot
in general be satisfied. The difficulty with this example is that (Jl and 2(Jl are
linearly dependent and we were able to eliminate (J 1 from the two inequalities.
Kronecker showed that the problem of simultaneous approximation can
always be solved if (Jl' ... , (In are linearly independent over the integers;
that is, if
n
L Ci(Ji = 0
i; 1
Theorem 7.9 (First form of Kronecker's theorem). If lXI' ••• , IXn are arbitrary
real numbers, if (Jl' ... , (In are linearly independent real numbers, and if
f: > 0 is arbitrary, then there exists a real number t and integers hI, ... , hn
such that
It(Ji - hi - IX;I < f: for i = 1,2, ... , n.
Lemma 1. Let {A·n} be a sequence of distinct real numbers. For each real t
and arbitrary complex numbers co, ... , CN define
N
f(t) = L creiO.•.
r;O
T~oo T °
150
7.5: Extension of Kronecker's theorem to simultaneous approximation
Hence
f Tf(t)e-i,J..k dt =
o
f
r=O
C r fTeiO.r-J..k)' dt
0
+ Ck T,
r*-k
where oc 1, ... , oc. and (}1,"" (). are arbitrary real numbers. Let
L= sup IF(t)l.
-oo<t<+oo
Then the following two statements are equivalent:
(a) For every e > 0 there exists a real t and integers h b ... , h. such that
It(}r - OCr - hrl < e forr = 1,2,oo.,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 t(}r - OCr
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) I 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 t(}r - OCr 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 tk be the corresponding value of t given by (a). Then
2n(t k (}r - ocr) differs from an integer multiple of 2n by less than 11k so
1
cos 2n(t k (}r - ocr) 2:: cos k'
Hence
151
7: Kronecker's theorem with applications
and therefore L 2:: IF(t k ) I 2:: I + n cos(l/k). Letting k --> CIJ we find L 2:: 11 + 1.
Since L .:::;; n + I this proves (b).
Now we assume (a) is false and show that (b) is also false. If (a) is false there
exists an e > 0 such that for all integers 11 b ... ,17 11 and all real t there is a k,
1 .:::;; k .:::;; n, such that
e
(9) Ite k - !Y.k - I1kl 2:: 2n'
(We can also assume that e .:::;; n/4 because if(a) is false for e it is also false for
every smaller e.) Let Xr = te r - !Y. r - hr. Then (9) implies 12nxk I 2:: e so the
point 1 + e2rrixk lies on the circle of radius 1 about 1 but outside the shaded
sector shown in Figure 7.1.
Figure 7.1
g = 1+ Xl + X z + ... + XII'
and write
(10) gP = 1 + "L arl, ... ,r l1
X 1 r, •.• X nrn ,
152
7.5: Extension of Kronecker's theorem to simultaneous approximation
where p is a positive integer. Then the coefficients art, .... ,,, are positive
integers such that
(11)
so if there are at most (p + l)n - 1 terms in each group on the right there will
be at most (p + l)n terms altogether. This proves (12) by induction. 0
PROOF OF KRONECKER'S THEOREM. Choosing F(t) as in Lemma 2 we have
n
F(t) = 1 + L e 21ti (rO,-a,l.
r= 1
By Lemma 1 we have
153
7: Kronecker's theorem with applications
Hence (15) implies lerl :s;; U for each r, and (14) gives us
(1 + n)p :s;; (N + I)U :s;; (p + l)nu
by Lemma 3. Therefore
°
arbitrary real numbers, iffJ 1 , ..• ,fJn , 1 are linearly independent real numbers,
and if e > is given, then there exists an integer k and integers m1, •. , mn
such that
IkfJ i - mi - ocd < e for i = 1,2, ... , 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{O;} - hi - (Xd = It{Oi} - hi - (Xi + (k - t){fJ i } I
:s;; It{Oi} - hi - (Xd + Ik - tl < e.
154
7.6: Applications to the Riemann zeta function
where the infimum and supremum are taken over all real t.
PROOF. For 0- > 1 we have 1((0- + it) I :s; ((0-) so M(o-) = ((0-), the supremum
being attained on the real axis. To obtain the result for m(o-) we estimate the
reciprocal 11/ ((s) I. For 0- > 1 we have
155
7: Kronecker's theorem with applications
Choose any e, 0 < e < n/2, and choose any integer 11 ~ 1. We apply
Kronecker's theorem to the numbers
-1
8k = 2nlogpk, k = 1,2, ... ,11,
where PI' ... , Pn are the first 11 primes. The 8i are linearly independent
because
n
L: ai log Pi =
i= I
° implies 10g(pI a, ... Pn a ") = 0
(19) I}]I (1
n
- Pk -S)
Inn
= )I
11 - Pk -si > }]I (1 + Pk -<1 cos e).
Now
1 00
fill
Ik=n+1 - Pk - S I - 11 < e
or
n 11 -
00
1- e< Pk - S I < 1 + e.
k=n+l
156
7.7: A pplications to periodic functions
-
1
m(l1)
2 (1 - e)
ro
k= I
n(1 + Pk -a cos e).
We will show in a moment that the last product converges uniformly for
o :$;
e :$; n/2. Therefore we can let e -> 0 and pass to the limit term by term
to obtain
_1_ > nro (1 + _a) _ ((11)
m(l1) - k= I Pk - ((2cr)"
157
7: Kronecker's theorem with applications
But W = kw z - hWI is a period of f with Iwi < B. Also, W =f. 0 since wz/w l
is irrational. 0
Theorem 7.13. Iff has three periods WI' Wz , W3 which are linearly independent
over the integers, then f has arbitrarily small nonzero periods.
PROOF. Suppose first that WZ/w l is real. If WZ/w l is rational then WI and
Wz are linearly dependent over the integers, hence WI' Wz , W3 are also depen-
dent, contradicting the hypothesis. If WZ/WI is irrational, thenfhas arbitrarily
small nonzero periods by Theorem 7.12.
Now suppose Wz /w l is not real. Geometrically, this means that WI and
Wz are not collinear with the origin. Hence W3 can be expressed as a linear
combination of WI and Wz with real coefficients, say
W3 = aWl + pw z , where a and p are real.
Now we consider three cases:
(a) Both a and p rational.
(b) One of a, p rational, the other irrational.
(c) Both a and Pirrational.
Case (a) implies WI' WZ , W3 are dependent over the integers, contradicting
the hypothesis.
For case (b), assume a is rational, say a = alb, and P is irrational. Then
we have
so
This gives us two periods bW3 - aWl and bw z with irrational ratio, hence f
has arbitrarily small periods. The same argument works, of course, if P is
rational and a is irrational.
Now consider case (c), both a and P irrational. Here we consider two
subcases.
(c l ) Assume a and P are linearly dependent over the integers. Then
there exist integers a and b, not both zero, such that aa + bP = O. By sym-
metry, we can assume that b =f. O. Then P = -aa/b and
so
Again we have two periods bW3 and bw! - aW 2 with irrational ratio, so f
has arbitrarily small nonzero periods.
(C2) Assume a and Pare linearly independent over the integers. Then by
Kronecker's theorem, given any B > 0 there exist integers hI> h2 and k
158
Exercises for Chapter 7
such that
2. (a) Given 11 real numbers B1 . . . . ,B", prove that there exist integers hl, ... ,h" and
k > 0 such that
(b) If at least one of the Bi is irrational, prove that there is an infinite set of n-tuples
(hdk, ... , h"/k) satisfying the inequalities in (a).
3. This exercise gives another extension of Dirichlet's approximation theorem. Given
m linear forms,
in n + m variables XI"'" X". Y1"'" Ym' prove that for each integer N > 1 there
exists integers X I" .. , X"' YI,' .. , Ym such that
1
ILi I < IV for i = 1, 2, ... , m
159
7: Kronecker's theorem with applications
2k2 .
l e-~I<-I
k
has only a finite number of solutions in integers hand k with k > 0 if 0 < c < 1/)5.
(a) Let fJ = ex - )5 so that ex and {3 are roots of the equation x 2 - x - I = O.
Show that for any integers hand k with k > 0 we have
(b) If (20) has infinitely many solutions hlk with k > 0, say h 11k 1, h 2lk· , ... ,show that
kn --> % as 11 --> % and use part (a) to prove that c ~ 11)5.
6. In Lemma 2, define
160
8
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.
L a(n)e-s).(n)
ro
n;1
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.
161
8: General Dirichlet series and Bohr's equivalence theorem
prove the existence of (J c and (J a' 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 -s).(n) converges for some s with
positive real part, say for s = So with (Jo > O. Let
L - I' 10gIL~=1 a(k)1
- 1m sup
n-+oo
1( )
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) 10gIA(n)1 < ((Jo + e)A(n)
for all sufficiently large n, then it follows that
log IA(n) I
A(n) < (Jo +e
for these n, so L ::; (J 0 + e, hence L ::; (J o. Now relation (l) is equivalent to the
inequality
(2)
162
8.2: The half-plane of convergence of general Dirichlet series
The Sen) are bounded since the series Lk'= 1 a(k)e-So)'(kl converges. Suppose
that ISen) I < M for all n. To express A(n) in terms of the Sen) we use partial
summation:
n n
A(n) = L a(k) = L a(k)e-So),(kleso).(kl
k=1 k=1
n
= L {S(k) - S(k - l)}eso)'(k),
k=1
provided S(O) = O. Thus
o 0-1
A(n) = L S(k)eso)'(kl - L S(k)eso),(k+ 1)
k= 1 k= 1
0-1
= L S(k){eso)'(k) - eso),(k+ I)} + S(n)eso).(nl.
k=1
Hence
n-l
IA(n)1 < M L leSo)'(k) - e So ).(k+l)1 + Metro).(n).
k=1
But
n-l
L Ieso)'(k) - eso).(k+ 1) I =
n- 1 1
L
So
f).(k+l) I
e SOU du :::;; ISo I L
n-l f).(k+l)
etrou du
k= 1 k= 1 )'(k) k= 1 )'(k)
Thus
Now A,(n) ~ 00 as n ~ 00 so
163
8: General Dirichlet series and Bohr's equivalence theorem
L A(n){e-SA(n) -
b
= e-SA(n+ I)} + A(b)e-SA(b+ 1)
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 forall n 2: N(e)
we have
10gIA(n)1
A(n) < L + e.
We can also assume that A(n) > 0 for n 2: N(e). Hence
IA(n)1 < e(L+.»).(n) for all n 2: N(e).
le-SA(n) - e- SA (n+l)1 = I -s
A(n+ 1)
e- SU du ::::;
I lsi fA(n+ 1)
e- UU du
f
A(n) A(n)
so
= lsi f )'(b+
).(a)
1)
e-' u du =
II
~ (e-fA(a)
e
_ e-d(b+ 1»).
Thus we have
t
In=aa(n)e-SA(n) I : : ; !.::!
e
(e-f).(a) - e-f)'(b+ 1») + e-f).(b+ I) + e-fA(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 0' > L. This completes the proof.
Note also that this proves uniform convergence on any compact subset of
the half-plane 0' > L. 0
Theorem 8.2. Assume the series La(n)e-SA(n) converges for some s with 0' >0
but diverges for all s with 0' < O. Then the number
As a corollary we have:
Theorem 8.3. Assume the series L a(n)e-SA(n) converges absolutely for some s
with 0' > 0 but diverges for all s with 0' < O. Then the number
_I' log Lk=l Ia(k) I
O'a - 1m sup '( )
n-oo ILn
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,
165
8: General Dirichlet series and Bohr's equivalence theorem
,l.(n) = L r n,k{3(k)
k=l
where the r n, k are rational and the number of summands q(n) depends
on n. (By condition (a), if ,l.(n) # 0 this representation is unique.)
(c) Each (3(n) is expressible as a finite linear combination of terms of A,
say
m(n)
(3(n) = L tn,k,l.(k)
k=l
EXAMPLE I. 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
1: r
k=l
k log Pk =0 implies so
To express each ,l.(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 8.5. 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 = T A and A = RB.
Hence r = T(RB) = (TR)B = AB where A = TR. 0
167
8: General Dirichlet series and Bohr's equivalence theorem
Theorem 8.6. Let Band f be two bases for A, and write f = AB, A = RBB,
A = Rrr, where A, R B, Rr arC? Bohr matrices. Then RB = RrA.
Note. If we write AlB for R B, Alf for Rr and fiB for A, this last equation
states that
A A f
B r B'
PROOF. We have A = RBB and A = Rrf = RrAB. Hence RBB = RrAB,
so (RB - RrA)B = O. Since RB - RrA is a Bohr matrix and B is a basis, we
must have RB - RrA = O. 0
Definition. The Bohr function F(Z) = F(Z1' Z2' ... ) associated with f(s),
relative to the basis B, is the series
L: a(n)e-(RZ)n,
00
F(Z) =
n= 1
where (RZ)n denotes the nth entry of the column matrix RZ.
In other words, if
gIn)
).(n) = L: r n,kf3(k)
k=1
then
oc
F(Z1,Z2"") = L: a(n)e-(rn.1=1+"·+rn,q(n)=q(n».
n=1
L: a(n)e-SA(n) = f(s).
00
F(sB) =
n=l
In other words, the Dirichlet series f(s) arises from F(Z) by a special choice
of the variables Z1' Z2, .... 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
L a(n)e-(RBZ)n .
00
F B(Z) =
• =1
Theorem 8.7. Let Band r be two bases for A and write r = AB for some
Bohr matrix A. Then
F B(Z) = F r(AZ).
FB(Z) = 0
n=1 n=1
Re AZ = A Re Z = AaB = aAB = ar
so F B(Z) E U f( a; r). This proves U f( a; B) £;; U f( a; r), and a similar argument
gives Uia; r) £;; Uia; B). 0
169
8: General Dirichlet series and Bohr's equivalence theorem
Definition. If the Dirichlet series f(s) = I:'; 1 a(n)e - sA(n) converges absolutely
for a = 0'0 we let
Vf(ao) = {f(a o + it): -00 < t < +oo}
denote the set of values taken by f(s) on the line a = 0'0'
Since f(5) can be obtained from its Bohr function F(Z) by putting Z = aB,
it follows that Vf(ao) £:; U f(a O)' Now we prove an inclusion relation in the
other direction. r
Theorem 8.9. Assume 0'0 > aa' where aa is the abscissa of absolute convergence
of a Dirichlet series f(s). Then the closure of Vf(ao) contains U f(aO)' That
is, we have
VAa o) £:; U f(a O) £:; Vf(a O)' and hence U f(a O) = VAa o).
PROOF. The closure VAa o) is the set of adherent points of Vf(a O)' We are
to prove that every point u in U f(a O) is an adherent point of VAa o). In
other words, given u in U Aa o) and given e > 0 we will prove that there exists
a v in Vf(a O) such that lu - vi < e. Since v = f(a o + it) for some t, we are
to prove that there exists a real t such that
If(a o + it) - ul < e.
Since u E Uf(a O) we have u = F(Zlo Z2,"') where Zn = a o{3(n) + iYn' Hence
Z = aoB + iY, RZ = aoRB + iRY = aoA + iRY,
so
(RZ)n = aoA(n) + i(RY)n = aoA(n) + illn,
say. Therefore
= I a(n)e-aoA(n)e-iltn.
co
U
n;l
L a(n)e-aOA(n)e-iIA(n),
co
f(a o + it) =
n;l
170
8.6: The set of values taken by a Dirichlet series f(s) on a line (J = (Jo
hence
L a(n)e-l1ol(nl(e-itl(nl -
00
The idea of the proof from here on is as follows: First we split the sum into
two parts, L~~ 1 + L~~ N + l' We choose N so the second part L~~ N + 1 is
small, say its absolute value is < !t:. 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 - itl(nl very close to e - ill"
simultaneously for every 11 = I, 2..... N. Then each factor e - itA(nl - e - ill"
will be small, and since there are only N terms. the whole sum will be small.
Now we discuss the details. For the given t:, choose N so that
I
f
n~N+l
a(n)e-I1OA(nl(e-itl(nl - e-ill")1 < ~.
2
Then we have
This holds for any choice of t. We wish to choose t to make the first sum
< !t:. Since leitA(n l I = 1 we can rewrite the sum in question as follows:
L la(n)le-l1oA(nllei(O.(nl-ll"l -
N
5 11.
n~ 1
Suppose we could choose a real t and integers kl' ... , kN such that
(4)
where Ixnl < b for n = 1,2, ... , N. Then for this t we would have
and hence
e
LN la(n)le- uo1(n)le (tl(n)-l'n) -
i 11 < - LN la(n)le-uo1(n) < -.e
n= I 2M n=1 2
Thus, the proof will be complete if we can find t and integers kl' ... , kN to
satisfy (4). If the A(n) were linearly independent over the integers we could
apply Kronecker's theorem to A(1), ... A(N) and obtain (4). However the
A(n) are not necessarily independent so instead we apply Kronecker's
theorem to the following system:
where
(J = p(n) Yn
n 2nD' CXn = 2nD'
The p(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, IP(1) + ... + rn,q(n)P(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(I) + ... + q(N) such numbers
ri, j ' The numbers (In 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
()
It(Jk - CXk - hkl < 2nDA'
where
N q(n)
A = L L Irn,jl.
n= j=I I
172
8.7: Equivalence of general Dirichlet series
where k., is an integer and Ix.1 < «(j/A) Lj<:\ Irn,jl < (j. But this means we
have found a real t and integers kl' ... , kN to satisfy (4), so the proof is
oo~~. 0
L a(n)e-s).(n)
cc
and
n=1 n=1
Let B = {f3(I1)} 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
L a(n)e-s).(n) '" L b(n)e-s).(n)
OC; 00
n=1 n=1
A(I1) = L rn,kf3(k),
k= 1
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)1 = 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 = {xn} = RB Y. Now write A = Rrr. If we show that for
some sequence V = {vn} we have X = Rr 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 Rr V = RrA Y = RB Y = X, since RrA = R B. This completes the
proof. 0
Theorem 8.11. The relation ~ defined in the foregoing definition is all 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 Xn will then be zero.
If b(n) = a(n)e iXn then a(n) = b(n)e- ixn . Since X = RB 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 = RB Y for some Y, and that a(n) = c(n)e illn ,
where U = RB V for some V. Then b(n) = c(n)ei(Xn+u n) where
X + U = RB Y + RB V = RB(Y + V).
This completes the proof. o
8.8 Equivalence of ordinary Dirichlet series
Theorem 8.12. Two ordinary Dirichlet series
(5)
so the integer powers may be used as entries in the Bohr matrix RB 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
where the integers a",k are determined by equation (5), Define a function f
by the equation
n g(n, k),
oc:
(8) fen) =
k=!
where
f(Pd = exp(iYd,
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",kYk)
for every k ~ 1. Equation (8) now becomes
f(n) = eXP{if,a",kYk}.
k= 1
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
Since the real part of Zn + iYn is the real part of Zn' both series take the same
set of values on the lines Xn = (Jo{J(n). Hence Uf((Jo) = Ug((Jo), as asserted.
o
176
8.10: The set of values taken by a Dirichlet series in a neighborhood of the line (J = (J0
That is, Wf(O'o; b) is the set of values taken by f(s) in the strip
0'0 - b< 0' < 0'0 + b.
Also, if 0'0 > O'a we define
WiO'o) = n
o <(j<O'o- O'a
WiO'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'o) ~ WiO'o) since every value taken by f(s) on the line
0' is also taken in every strip containing this line. Of course, it may
= 0'0
happen that ViO'o) = Wf(O'o) or that Vf(O'o) :t= Wf(O'o)'
In general, we have:
Theorem 8.14. Vf(O'o) !;;; Wf(O'o) !;;; 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(O'o) is in the closure of ViO'o).
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
L a(n)e-s).(n)
r>:;.
f(s) =
n=1
L a(n)A(n)e-s).(n).
00
/'(s) = -
n=1
177
8: General Dirichlet series and Bohr's equivalence theorem
L la(n)IIA,(n)le-a).(n) = L la(n)le-ao').(n)IA,(n)le-(a-ao')).(n)
00 00
II'(s) I ::::;
n=l n=l
where aa < a o' < ao. Now IA,(n) Ie-(a-ao')).(n) --+ 0 as n --+ 00 so, in particular,
this factor is less than 1 for large enough n. Hence
L
00
178
8.12: Proof of Theorem 8.15
First of all, we have SJ{O"I) s;;; U"o>", WJ{O"o) because VACTo) S;;; WACTo). To
get inclusion in the other direction, assume WE U"o>", WJ{CTo). Then
W E WJ { CT 0) for some CT 0 > CT l' Hence W E WJ { CT 0; 0) for all 0 satisfying
o < c5 < CT o - CTa • In other words, f{s) takes the value W in every strip
CT o - 0 < 0" < CT o + 0 irO < 0 < CTO - CT a • In particular, when 0 = 0"0 - CT 1 ,
we have CT o - 0 = CT 1 sof(s) = w for some s with CT > CT 1 • Hence WES J (CT 1 ).
This proves that U"o>", WJ(CTo) s;;; SJ(CT 1), so the two sets are equal. Therefore,
we also have
SiCTd = U Wg(CTo)·
0'0>0'1
Hence U ACTo) = U g(CT o)' But, in view of Theorem 8.9, this means
VACT o) = Yg(CTo).
But Theorem 8.15 states that VJ(CTo) = WJ(CTo) and Yg(CTo) = Wg(CT o) so Bohr's
equivalence theorem is a consequence of Theorem 8.15. D
Note. The important point is that one subsequence {nk} 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: {OI, n},~= I' This is a bounded infinite sequence so it has an
accumulation point, say ()I' Hence there is a subsequence {n r } such that
lim r _ oo ()I.nr = ()I' Similarly, for the second row there is an accumulation
point ()2 and a subsequence n,' such that lim n _ oo ()2,nr' = ()2' and so on. The
subsequence {n,'} needed for ()2 may be quite different from that needed for
()I' Helly's principle says that one subsequence works simultaneously for all
rows.
PROOF OF LEMMA 1. Let ()I be an accumulation point of the first row and
suppose the subsequence {n,!1)} has the property that
lim
r~<X)
()1,nr(l) = ()I'
In the second row, consider only those entries ()2. nr(I). This is a bounded
sequence which has a convergent subsequence with limit ()2, say. Thus,
lim () 2, n r (2) = () 2
r~oo
That is, n l 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
{()m,nJ. We assert that
Since nr = n,<r>, after the mth term in this row we have r > m so every integer
n,<r) occurs in the subsequence n,<m), so from this point on {n r} is a sub-
sequence of {nr(m)} hence ()m,nr ~ ()m' as asserted. 0
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 insidl! C.
PROOF OF LEMMA 2. Let m = inf{lf(z)1 - Ig(z)l:z E C}. Then m > 0
because C is compact and the difference 1f(z) 1 - Ig(z) 1 is a continuous
function on C. Hence for all real t in the interval 0 :s; t :s; 1 we have
If(z) + tg(z) 1 2 If(z) 1 - 1tg(z) 1 2 If(z)1 - Ig(z)1 2 m > O.
180
8.12: Proof of Theorem 8.15
We wish to prove that v E Wf(O'o). This means we must show that v E Wf(O'o; 6)
for every 6 satisfying 0 < 6 < 0'0 - O'a' In other words, if 0 < 6 < 0'0 - O'a
we must find an s = 0' + it in the strip
0'0 - 6< 0' < 0'0 +6
such that f(s) = v. Therefore we are to exhibit an s in this strip such that
f(s) = lim f(O' 0 + it.).
L a(n)e-CfoA(n). e-i1mA(n).
00
f(O'o + it m) =
n=1
The products tmA(n) form a double sequence. There exists a double sequence
of real numbers en, m such that
en,m = tmA(n) + 2nk m,n' with 0::::; e n. m < 2n,
where km," is an integer. If we replace tmA(n) by en,m in the series we don't
alter the terms, hence
+ it m) = L a(n)e - CfoA(n)e -
00
f(O' 0 i8 n ,,,,.
n= 1
g(s) = L b(n)e-s).(n)
n=l
where
This has the same abscissa of absolute convergence as f(s). Now consider
the following sequence of functions:
J..(s) = f(s + it n)
where {n r } is the subsequence for which (10) holds. We assert that
(a) J..(s) -+ g(s) uniformly in the strip (10 - 0 < (1 < (10 + 0, hence, in
particular, in the circular disk Is - 0"0 I < O.
(b) g«(1o) = v.
(c) There is a d, 0 < d < 0, 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 + itnR is in the strip if s is in the disk,
so this proves the theorem. Now we prove (a), (b) and (c).
0()
N
:S; L la(n)le-(<ro-d»).(n)le-iBn,nr - e-iBnl
n=l
L
0()
+2 Ia(n) Ie-(<ro-d»).(n).
n=N+ 1
Now if e > 0 is given there is a number N = N(e) such that
e
2 L la(n)le-(<ro-d»).(n) < -,
0()
n=N+1 2
because the series L:=l la(n)le-(<ro-d»).(n) converges. In the finite sum L~=l
we use the inequality
Ie - ib - e- ia I = I~ s:e- it dt I :S; Ib - aI
182
8.12: Proof of Theorem 8.15
to write
Ie-iBn ... , - e-iBnl -< 10n,n,. - 0n I.
Proof of (c). Assume first that 9 is not constant. Since g(O'o) = v there is a
positive d < b such that g(s) =I v on the circle
C = {s: Is - 0'0 I= d}.
Let M be the minimum value of Ig(s) - v Ion 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 10' - 0'01 < b.
Then on C we have
IfR(S) - g(s) I < M :s; Ig(s) - vi·
If G(s) = fR(S) - g(s) and F(s) = g(s) - v we have IG(s) I < IF(s) I 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(O'o) = 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 9
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
= - L A(n)b(n)e-Sl(n) = O.
00
g'(s)
n= 1
183
8: General Dirichlet series and Bohr's equivalence theorem
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)
~ ).(n) _ (2s) 'f
L. 5- r() 1(J>1.
n=l n .. s
C(x) = L A(n).
n$x n
Then if (J > 1 we have
(2s) = Joc C(x) dx
(s - 1)(s) 1 x5 •
"
L...
A(n) ~ _ C(x)
5 - 5 +s
JX C(t)1 dt. 5+
n$x n n x 1 t
Keep (J > 0 aRd let x --+ 00. Then
C(x) =
x5
o(~
x"
" !)
n'tx n
= O(IOg x) = 0(1) as x
x"
--+ 00,
so we find
.~
L... s:tT
n= 1
),(n)
n
=s I
X'
1
C(t)
--.-+r dt,
t
for (J > O.
Replacing s by s - 1 we get
I: A(7)
n= 1 nit
(s - 1) Joo C~) dt
= for (J > 1.
Since the series on the left has sum (2s)/(s) the proof is complete. D
If there exists an no such that (n(s) # 0 for all n ~ no and all (J > 1, then
(s) # 0 for (J > 1-
k - 5 and L~ = 1 A(k)k - 5
PROOF. First we note that the two Dirichlet series L~ = 1
are equivalent because A is completely multiplicative and has absolute
185
8: General Dirichlet series and Bohr's equivalence theorem
value 1. Therefore, by Bohr's theorem, 'l~) ::/; 0 for (J > 1 implies that
Lk=1 A(k)k- s ::/; 0 for (J > 1. But for s real we have
lim t
s- + 00 k= 1
A(~) = A(I) = 1.
k
Hence for all real s > 1 we must have Lk= 1 A(k)k- > O. S Letting s -+ 1+
we find
~ A(k) > 0
L.... if n ~ no·
k=1 k -
In other words, the function
valid for (J > 1. Note that the denominator (s - lK(s) is nonzero on the
real axis s > 1. 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
t.
left is analytic everywhere in the half-plane (J > This implies that '(s) ::/; 0
for (J > t, and the proof is complete. 0
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) = Lnsx A(n)/n. !f there exist constants
0( > 0, c > 0 and no such that
log« x
(12) C(x) > - c - -
Jx
for all x ~ no, then the Riemann hypothesis is true.
PROOF. If e > 0 is given there exists an n1 ~ no such that c log« x ::;; x' for
all x ~ n1 so (12) implies
C(x) > _X'- 1/2 .
186
Exercises for Chapter 8
Let A(x) = C(x) + X<-1/2, where E is fixed and 0 < E < 1. Then A(x) > 0
for all x ;:::: nl' Also, for (J > 1 we have
((2s) + ~ = f(s)
(s - 1)((s) s - "2 - E '
say. Arguing as in the proof of Theorem 8.20, we find that the function f(s)
is analytic on the real line S > 1 + E. By Landau's theorem it follows that
f(s) is analytic in the half-plane (J > 1 + E. This implies that ((s) # 0 for
(J > 1 + E, hence ((s) # 0 for (J > 1 since E 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 + (In in which (n(s) never vanishes.
(See [4], Theorem 11.4.) The exact value of (In is not yet known. In his 1948
paper [50] Tunin proved that, for all sufficiently large n, (n(s) # 0 in the
half-plane (J > 1 + 2(1og log n)/log n, hence (J n .::; 2(1og log n)/log 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 (J > 1 + c(log log n)/log n, hence (In ;;:: c(1og log n)/log n for large n.
The number 1 + (In is also equal to the abscissa of convergence of the
Dirichlet series for the reciprocall/(n(s). If (J > 1 + (In we can write
1 _ f
l1ik)
(n(s) - k~ 1 7 '
un(k) =
,
{I 0
~f k .::; n,
If k > n.
The usual Mobius function l1(k) is the limiting case of I1n(k) as n -> 00.
_ I'
(Je - 1m sup
log I D"=n a(k)1
1 '
n-", Il(n)
2. Let (Je and (Ja denote the abscissae of convergence and absolute convergence of a
Dirichlet series. Prove that
(Ja - °: ;. log n
(Je ::; hm sup ,l.(n)·
°: ;
n-a;
187
8: General Dirichlet series and Bohr's equivalence theorem
What does this imply about the radius of convergence of a power series?
4. Let Urn)} be a sequence of complex numbers. Let A denote the set of aU points
s = (J + it for which the series L a(n)e-sA(nl converges absolutely. Prove that A
is convex.
Exercises 5, 6, and 7 refer to the seriesfl\) = I:= I a(ll)e- SA (n) with exponents
and coefficients given as follows
n 1 2 3 4 5
a(n) 1
8
1
2" 4
1
-8
1 1
2"
n 6 7 8 9 10
a(n) 1
8 -4
1 1
2" - 43 -8
1
ifx3> -log2,zl,z2arbitrary.
7. Determine the set Vj(O). Hint: The points -1, 1 + i, 1 - i are significant.
8. Assume the Dirichlet series f(s) = L;'= I a(n)e-s!.(n) converges absolutely for (J > (Ja'
·
I1m -
1 fT eA(o+it}f( (J + It')d t = {a(n) if).=).(Il)
T~+-,;2T -T 0 ifX#).(1),).(2), ....
9. Assume the series f(s) = L::'=I a(ll)e- S A( n l converges absolutely for (J > (Ja > O. Let
v(n) = eArn).
(a) Prove that the series g(s) = L:= 1 a(Il)e-S\'ln) converges absolutely if (J > O.
(b) If (J > (Ja prove that
188
Exercises for Chapter 8
f'"
This extends the classic formula for the Riemann zeta function,
rs-I
r(s)((s) = -,- dr.
o e - 1
Hinr: First show that r(s)e-s).ln l = J~ e-'''ln1rs-1 cit.
189
Supplement to Chapter 3
where
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
A = T"IST"2S· •• ST''r,
where the n i are integers. But T = ST-IST-IS, so every element of r also
has the form ST'"IS ••• STInk 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
specified in (2), then it also holds for the products AT'" and AS for every
190
Supplement to Chapter 3
have
e-1Til4e(A) ifd>O,
e(AS) ={ e1Ti/4e(A) if d < O.
PROOF. We have
AS
= (ae b)(O
d 1
- 1) = (b
0 d
- 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
(=~ ~).
For d > 0, we have
b-e a+d 1
l2d + s(c, d) = ~ - s(d, c) - 4·
Using this in (3), we find e(AS) = e- 1Til4e(A).
In this case, -d'> 0 and we use the reciprocity law in the form
c d 1 ad - be
s(e, - d) + s( - d, e) = _ 12d - 12c - 4 - 12cd·
-b+e a+d 1
-12d - s(c, -d) = ~ - s(d, e) + 4·
Using this in (4), we find that e(AS) e1Ti/4e(A). This completes the proof
of Lemma 2. o
Lemma 3. If Dedekind's functional equation
(5)
then it is also satisfied for AT'" and for AS. That is, (5) implies
whereas
192
Supplement to Chapter 3
hence,
- i(dr - c)
r + d) --
- I.( CS . e -1TiI2 ,
-Ir
193
Supplement to Chapter 3
PROOF. Let
aT +b I a'T + b'
AT=--- and
CT + d AT= CT+d
Hence,
or
( ')
7TlQ
. e(A)I
exp - 12c = exp (. )
7TlQ
- 12c e(A).
This shows that the product exp( - ~~:)e(A) depends only on C and d. There-
fore, the same is true for the product
exp ( -
7Ti(a +
12c
d») e(A).
This complex number has absolute value 1 and can be written as
exp(
7Ti(a +
12c
d») e(A) = exp(-7Tif(d, c»
194
Supplement to Chapter 3
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 Edition. 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 Wei!. I. Inst. haut. Etud sci., Publ. math. 43 (1973),
273-307 (1974). Z. 287, 14001.
8. Erdos, P. A note on Farey series. Quart. 1. 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 oj Matrices, Vol. I. Chelsea Pub!. Co., New York,
1959.
II. 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. Hecke, 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. Ann. JJ4 (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 ,(n). Duke Math. J. 10 (1943), 483-492.
MR 5, 35b.
20. Lehmer, D. H. Properties of the coefficients of the modular invariant J(,). 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
invariant}(,). Amer. J. Math. 71 (1949), 136-148. MR 10, 357a.
24. Lehner, Joseph. Further congruence properties of the Fourier coefficients of the
modular invariantj(')' 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. 19 (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(')' Proc. Amer. Math. Soc. 9 (1958), 609-612. MR 20 #5184.
31. Petersson, Hans. Ober die Entwicklungskoeffizienten der automorphen formen.
Acta Math. 58 (1932),169-215.
32. Petersson, Hans. Ober eine Metrisierung der ganzen Modulformen. Jber. Deutsche
Math. 49 (1939), 49-75.
33. Petersson, H. Konstruktion der samtlichen Losungen 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 p(n). Proc. London Math. Soc. (2)
43 (1937), 241-254.
37. Rademacher, Hans. The Fourier coefficients of the modular invariantj(r). Amer.
J. Math. 60 (1938),501-512.
38. Rademacher, Hans. On the expansion of th'e 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. Erliiuterungen 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 Transformationsformel fUr log 1](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 proofoflJ( - liT) = IJ(T)JTii. Mathematika I (1954),
4. MR 16, 16b.
49. Titchmarsh, E. C.lntroduction to the Theory of Fourier Integrals. Oxford, Clarendon
Press, 1937.
50. Tunin, Paul. On some approximative Dirichlet polynomials in the theory of the
zeta-function of Riemann. DallSke Vid. Selsk. MaI.-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. JO (1959),277-298. MR 22 #6774.
52. Uspensky, J. V. Asymptotic formulae for numerical functions which occur in the
theory of partitions [Russian]. Bull. A cad. Sci. URSS (6) 14 (1920), 199-218.
53. Watson, G. N. A Treatise on the Theory of Bessel Functions, 2nd Edition. Cambridge
University Press, Cambridge, 1962.
198
Index of special symbols
!p(t)
Ip-I
- IJ- ,
('+A) 80
P P
l=O
( l1(qt)Y/(Q-I)
<1>( t) 86
l1(t) ,
9(t) Jacobi theta function, 91
p(n) partition function, 94
199
Index of special symbols
200
Index
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 '1(r), 47
Bessel functions, 109 Dedekind sums, 52, 61
Bohr, Harald, 161,196 Deligne, Pierre, 136, 140, 196
201
Index
E
e l ,e 2 ,e3 ,13 H
Eigenvalues of Hecke operators, 129 Half-plane H, 14
Eisenstein series G., 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 pen), 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 T., 120
of pairs of periods, 4 Helly, 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 approximation theorem, 145
forms, 134 Hurwitz zeta function, 55, 71
Euler, Leonhard, 94
Euler products of Dirichlet series, 136
Exponents of a general Dirichlet
I
series, 161 InvariantsB2,B3,12
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 ofj(-r), 21, 74 J
divisibility properties of, 22, 74, 91 j(-r), J(-r), 74, 15
Functional equation, for "I(-r), 48, 52 Fourier coefficients of, 21
for .9(-r), 91 Jacobi, Carl Gustav Jacob, 6, 91,141
for C(s), 140 Jacobi theta function, 91, 141.
for t\(0(, p, z), 54 Jacobi triple product identity, 91
for clI(lX, p, s), 56, 71
Fundamental pairs of periods, 2
Fundamental region, of modular group K
r. 31 Klein, Felix, 15
of subgroup r o(P), 76 Klein modular invariant J(-r), 15
202
Index
Kloostennan. 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
Lambert, Johann Heinrich, 24
o
Lambert series. 24 O'Brien, J. N., 91,196
Landau, Edmund. 186 Order of an elliptic function, 6
Lehmer. Derrick Henry. 22, 93, 95, 197
Lehmer conjecture, 22
Lehner, Joseph, 22, 91, Ill,) 97
p
LeVeque, William Judson, 197 f.J-function of Weierstrass, 10
Linear space Mk of entire forms, 118 Partition function pen), 1, 94
Linear subspace M k. 0 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 A(t), 51
M
Mapping properties of J(t), 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 pen), 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
s form, 115
Whiteman, Albert Leon, 62, 198
Sa lie, 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, I 13, 131
Theta function, 91,141
Transcendental numbers, 147
Transformation of order n, 122
Transformation formula, of Dedekind,
48,52
of Iseki, 54
Tunin, 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
(comillued from page ii)