Você está na página 1de 205

An Introduction to the

Theory of L-functions
J

orn Steuding (W urzburg University)


A course given at Universidad Autonoma de Madrid, 2005/06
-1 1 2 3
-1.5
-1
-0.5
0.5
1
1.5
Contents
Preface iii
Chapter 1. The classical L-functions of Dirichlet, Riemann & co. 1
1.1. Motivation: prime numbers 1
1.2. Riemanns zeta-function 7
1.3. Dirichlet L-functions 14
1.4. The prime number theorem 24
1.5. Tauberian theorems a general approach 35
1.6. The explicit formula 49
Chapter 2. Zero-distribution of the Riemann zeta-function 66
2.1. The Riemann hypothesis 66
2.2. The approximate functional equation 72
2.3. Power moments 78
2.4. Hardys theorem: zeros on the critical line 84
2.5. Density theorems 87
2.6. Universality and self-similarity. 95
Chapter 3. Modular forms and Hecke theory 105
3.1. The functional equation for zeta and more 105
3.2. The zeta-function at the integers 117
3.3. Hamburgers theorem 121
3.4. Modular forms 124
3.5. Heckes converse theorem 129
3.6. Shimura-Taniyama-Wiles 134
Chapter 4. The Selberg class an axiomatic approach 143
4.1. Denition and rst observations 143
4.2. The structure of the Selberg class 146
4.3. The Riemannvon Mangoldt formula 149
4.4. Primitivity and Selbergs conjectures 160
4.5. Hecke L-functions 167
4.6. Artin L-functions and Artins conjecture 173
4.7. Langlands program 185
Bibliography 191
ii
Preface
This course provides an introduction to the theory of L-functions, a topic
which plays a central role in number theory since Dirichlets proof of the
prime number theorem in arithmetic progressions in 1837 and Riemanns
famous path-breaking paper in 1859. L-functions are generating functions
formed out of local data associated with either an arithmetic object or with an
automorphic form. These functions are special examples of so-called Dirichlet
series; all of them have in common that besides their series representation
they can also be described by an Euler product, i.e., a product taken over
prime numbers. The famous Riemann zeta-function
(s) =

n=1
1
n
s
=

p
_
1
1
p
s
_
1
may be regarded as the prototype. L-functions encode in their value-
distribution information about the underlying arithmetical or algebraic struc-
ture that is often not obtainable by elementary or algebraic methods, e.g. the
classical prime number theorem which states that the number (x) of primes
p x is asymptotically equal to the integral logarithm, resp.
lim
x
(x)
log x
x
= 1.
Another example is Dirichlets analytic class number formula which measures
the deviation from unique prime factorization in the ring of integers of qua-
dratic number elds. Two of the seven millennium problems are questions
about L-functions: the famous Riemann hypothesis (all non-real zeros of (s)
lie on the critical line Re s =
1
2
) and the conjecture of Birch & Swinnerton-
Dyer (the rank of the Mordell-Weil group of an elliptic curve is equal to the
order of the zero of the associated L-function L
E
(s) at s = 1).
We want to give an overview of the variety of L-functions, their importance
for number theory and allied elds, and recent progress toward old and new
problems. After introducing the classical examples, as (s) and Dirichlet L-
functions, and studying basic properties, we concentrate on three main lines
of investigation in detail. First, we give a rather detailed account on studies
of the zero-distribution of Riemanns zeta-function. We shall prove that (s)
has innitely many zeros on the critical line and further that there cannot be
too many zeros o the critical line. This supports the Riemann hypothesis.
It is believed that a proof of the Riemann hypothesis for the zeta-function
should easily carry over to other L-functions and, indeed, most of the tech-
niques in the second chapter can be generalized; however, these techniques
alone will probably not be sucient for a proof of the Riemann hypothesis.
iii
Second, there is Heckes theory which links modular forms and Dirichlet se-
ries with functional equation (Wiles et al. proof of the Shimura-Taniyama
conjecture, including Fermats last theorem, marks one of the highlights of
this approach); here we shall meet further examples of L-functions and learn
new techniques going beyond the theory of the nineteenth century (or those
designed to deal with the zeta-function). Finally, we study the axiomatic
approach initiated by Selberg with its far-reaching consequences on many
number theoretical problems as, for example, Artins conjecture on the holo-
morphy of Artin L-functions subject to the truth of Selbergs orthogonality
conjecture.
There is another quite remarkable line of investigation, namely the impact
of Random Matrix Theory, i.e., the recent idea to model L-functions by large
unitary random matrices; this approach is motivated by Montgomerys cele-
brated pair correlation conjecture and computations observing that the near-
est neighbour spacing for the nontrivial zeros of (s) seems to be amazingly
close (statistically the same?) to those for the eigenangles of the Gaussian
Unitary Ensemble. These observation have restored some hope to an old idea
of Hilbert and Polya that the Riemann hypothesis follows from the existence
of a self-adjoint Hermitian operator whose spectrum of eigenvalues corre-
sponds to the nontrivial zeros of the zeta-function. First it was our intention
to give a brief account of these ideas in the notes too; however, by lack of
time we did not include this approach here. We hope to add this approach
in a later version of these notes.
The course is aimed at doctoral students and non-experts which want to
learn the fundamentals of this subject. Of course, it is far beyond the scope
of this course to prove all relevant results, for instance, the rather technical
converse theorem of Weil (or the Shimura-Taniyama-Weil conjecture which
I hardly understand myself). However, we want to sketch the main ideas in
order to obtain a rst impression on the theory of L-functions, to learn its
big picture-questions and the modern approaches with which these objects
are studied. These notes contain more material than that presented in the
classroom (where we had two hours per week); furthermore, we have added
many exercises (the advanced marked with an asterisk) with the aim to give
the interested reader the possibility to get in touch with the basic objects
and to practise the presented techniques.
I am very grateful to Fernando Chamizo, Keith Conrad, Ernesto Girondo,
Fernando Holgado, Rasa Steuding, and Adrian Ubis for valuable comments
and corrections.
Jorn Steuding, Madrid, January 2006.
iv
CHAPTER 1
The classical L-functions of Dirichlet, Riemann & co.
The main theme in this introductory chapter are prime numbers. Questions
about primes had been a driving force for number theory ever since their
discovery by the ancient Greeks. Prime number distribution is intimately
linked with analytic objects, so-called L-functions. In this rst chapter we
will introduce some classical examples: the Riemann zeta-function, Dirichlet
L-functions, and Dedekind zeta-functions. The particular case of Riemanns
zeta-function, the prototype of an L-function, will be discussed in detail. We
shall learn rst fundamental properties, prove the celebrated prime number
theorem, and get to know the big open conjectures as, for example, the famous
Riemann hypothesis. For further reading we refer to Apostol [2], Iwaniec &
Kowalski [101], and Titchmarsh [200].
1.1. Motivation: prime numbers
A prime number is a positive integer n > 1 without proper divisors (in N).
The prime numbers are the multiplicative atoms of the integers: any positive
integer can be written as a unique product of powers of distinct primes (up
to the order of the factors). This fact is called the unique prime factorization
of the integers. Euclid (Prop. 20 in Elements 9; around 300 B.C.) proved
that there are innitely many prime numbers as follows: if 2, p
1
, . . . , p
n
are
prime numbers, then the number
Q := 2 p
1
. . . p
n
+ 1
has a prime divisor q dierent from 2, p
1
, . . . , p
n
(since otherwise q would
divide any linear combination of Q and q, in particular, +1).
An analytic version of the unique prime factorization is given by the identity

nN
1
n
s
=

p
_
1
1
p
s
_
1
, (1.1)
where the product is taken over all primes (a proof will be given later). Both,
the series and the product converge for s > 1 (also this will be proved below).
The identity between the series and the product was discovered by Euler [51]
in 1737. It gives a rst glance on the intimate connection between the prime
numbers and certain objects in analysis. A rst immediate consequence is
Eulers proof of the innitude of the primes. Assuming that there were only
1
2 Chapter 1 Classical L-functions
nitely many primes, the product in (1.1) is nite, and therefore convergent
throughout the whole complex plane, contradicting the fact that the series
reduces to the divergent harmonic series as s 1+. Hence, there exist
innitely many prime numbers. This argument might be slightly more com-
plicated than Euclids elementary proof but, as we shall see later, the analytic
access yields much deeper knowledge on the distribution of the prime num-
bers. In fact, the series in (1.1) denes the famous Riemann zeta-function
which encodes many arithmetic information in its value distribution.
In view of the innitude of the primes it is natural to ask how they are
distributed among the integers. It was the young Gauss who conjectured in
1791 (see Tagebuch, Werke, vol. 10.1) for the number (x) of primes p x
the asymptotic formula
(1.2) (x) Li (x),
where
(1.3) Li (x) :=
_
x
0
du
log u
:= lim
0+
__
1
0
+
_
x
1+
_
du
log u
is the logarithmic integral. This would imply that, in rst approximation,
the number of primes x is asymptotically
x
log x
, and so the primes form a
set of zero density in N. It is recorded that Gauss came to his conjectural
asymptotic formula by calculating the number of primes up to several mil-
lions. However, there is also a heuristic argument in favor for his conjecture
by exploiting identity (1.1). For this aim we cut the product and the series
at x (assuming that this still leads to an asymptotic identity as x ) and
let s = 1. This yields

nx
1
n

px
_
1
1
p
_
1
= exp
_

px
log
_
1
1
p
_
_
= exp
_

px
1
p
+ O(1)
_
.
By the well-known asymptotics for the truncated harmonic series,
(1.4)

nx
1
n
= log x + C + O
_
1
x
_
,
where C := lim
N

nN
1
n
log N = 0.577 . . . is the Euler-Mascheroni
constant, we get
(1.5)

px
1
p
log log x.
Section 1.1 Prime numbers 3
This formula is indeed true and was rst obtained by Euler [51] in the form
1
2
+
1
3
+
1
5
+ . . . = log log ;
however, his proof had some gaps and the rst waterproof argument is due to
Mertens [145]. Certainly, this asymptotic formula cannot be deduced from
Euclids proof. In particular, it shows that the sum over the reciprocals of the
prime numbers diverges, indicating that there are quite many primes (more
than squares since

n
1/n
2
< ). Using the Stieltjes integral (resp. partial
summation, a technique we meet later in detail), we also nd

px
1
p
=
_
x
2
1
u
d(u)
_
x
2
(u)
u
2
du.
Inserting Gauss conjectural asymptotics (1.2) shows that this is indeed of
the same size as predicted by Eulers formula (1.5). Clearly, this is not a
proof but it might suggest that (1.2) indicates the correct order for the prime
counting function (x).
Further evidence was found by Chebyshev [32, 33] around 1850 who proved
by elementary means that for suciently large x
0.921 . . . (x)
log x
x
1.055 . . . .
Moreover, he showed that if the limit
lim
x
(x)
log x
x
exists, the limit is equal to one, which supports relation (1.2). For a proof of
these results and also for more details on the history of the theory of prime
number distribution we refer to Narkiewicz [159].
There exist plenty of problems concerning prime numbers which are easy
to formulate but rather dicult to solve. Here is a short list of four famous
problems concerning the distribution of prime numbers.
Does there exist an exact formula for the number (x) of primes
p x? Is there an explicit formula for the nth prime number?
Given a positive integer B 2, are there innitely many pairs of
consecutive prime numbers having a dierence B? (For B = 2
this is the famous twin prime conjecture!)
Can any positive number be written as the sum of three primes?
Can any even integer greater than 2 be written as the sum of two
primes? (The second question is the open Goldbach conjecture!)
Is there always a prime number in between two squares of positive
integers? (Having a view on the rst primes we might expect a
positive answer.)
4 Chapter 1 Classical L-functions
We shall discuss the state of art of these problems later in these notes; we
may regard them as indicator what can be done and what cannot be done
with present day methods.
Another natural question is how the prime numbers are distributed in
residue classes (of course, this makes only sense for classes a mod m with
coprime a, m). One may try to mimic Euclids proof of the innitude of
primes and, indeed, one can show that there are innitely many prime num-
bers p 1 mod 4; however, one cannot succeed in proving the same for the
residue class 3 mod 5. M.R. Murty [153] gave a characterization of all prime
residue classes a mod m for which a Euclid-type proof exists; he showed that
a necessary and sucient condition is that a
2
1 mod m.
In 1837, Dirichlet proved that there are innitely many primes in any prime
residue class. His ingenious argument relies on a family of identities similar
to (1.1) and analytic properties of the appearing series, named Dirichlet L-
functions. His approach is regarded as the beginning of analytic number
theory and it also marks the beginning of the theory of L-functions; it is
legend that the capital L in the word L-function stands for one of his
initials (Peter Gustav Lejeune Dirichlet).
For short, the idea of analytic number theory can be described as follows:
given an arithmetic function,
f : N C , n f(n),
one hopes to get arithmetic information about f by studying the analytic
behaviour of the generating function
L
f
(s) :=

n=1
f(n)
n
s
;
in honour of Dirichlets contribution the generating series are called Dirichlet
series. It turns out that this is a rather fruitful concept. The set / of
arithmetic functions forms a commutative ring with respect to the standard
addition (f + g)(n) := f(n) + g(n) and the convolution (multiplication)
(f g)(n) :=

d|n
f(d)g(n/d),
where, as usual, we write d [ n if the integer d divides the integer n, and d n
otherwise. These operations correspond to the addition and the multiplica-
tion of the associated Dirichlet series:
L
f+g
(s) = L
f
(s) + L
g
(s) and L
fg
(s) = L
f
(s) L
g
(s).
The set T of associated Dirichlet series forms a ring isomorphic to /, and
convolution identities in arithmetic (which play a centrale role in elementary
Section 1.1 Prime numbers 5
number theory) correspond one-to-one to product identities of Dirichlet se-
ries. This leads via (formal) dierentiation to new identities for arithmetic
functions from old ones. Furthermore, in many cases one can exhibit number
theoretical information from identities for the associated Dirichlet series and
their analytic behaviour.
In number theory we are often concerned with multiplicative arithmetic
functions; their associated Dirichlet series can be written as an innite prod-
uct over the prime numbers and this is the essential property of an L-function.
In the next section we will present the prototype.
Exercise 1. Let x 3. Prove that there are more than (log 2)
1
log log x many
prime numbers p x.
Hint: use Euclids proof and induction.
Exercise 2. Prove that there are intervals of arbitrary length in (0, +) free of
prime numbers.
Exercise 3. Show that, for x > 1,
Li (x) =
_
x
1
_
1
1
u
_
du
log u
+ log log x + C
=
x
log x
N

k=0
k!
(log x)
k
+ O
_
x
(log x)
N+2
_
,
where C is a constant; in fact, one can show that C is the Euler-Mascheroni
constant, can you?.
Exercise 4. Prove

px
1
p
log log x + O(1);
this is half way to the asymptotic formula (1.5).
Hint: Start to show the inequalities

px
_
1
1
p
_

nx
1
n
>
_
x
1
du
u
.
Eulers -function (n) counts the number of prime residue classes mod n, i.e.,
(n) = 1 a n : gcd(a, n) = 1.
Exercise 5. i) Show that (n) = n 1 if and only if n is prime.
ii) Prove that
(n) = n

p|n
_
1
1
p
_
.
Hint: consider rst n = p
k
with p prime.
6 Chapter 1 Classical L-functions
Exercise 6. Let q > 1 and x = qk + r with 0 r < q be positive integers. Prove
that
(x) q + k(q) + r 2q +
_
x
q
_
(q)
and
limsup
x
(x)
x

(q)
q
.
Deduce that (x) = o(x). (This argument is due to Fousserau, cf. [159]).
The sieve of Eratosthenes is a very ecient algorithm to produce a list of all
prime numbers below a given magnitude:
Exercise 7. * Make a list of all positive integers 2 n x and mark all proper
multiples of prime numbers p

x. Then the number of unmarked numbers is
(x). Why?
ii) Prove
(1.6) (x) (y) x

py
_
1
1
p
_
+ O(2
y
)
for any 1 y

x (maybe with help of some literature, e.g., [159]).


iii) Use ii) to show that, for suciently large x,
(x)
x
log log x
.
Hint: recall Exercise 4.
Exercise 8. i) Prove that there are innitely many primes p 1 mod 4 and 3
mod 4 (one case is rather tricky and involves the theory of quadratic residues).
Hint: For one case one may use a fact from the theory of quadratic residues:
the congruence X
2
1 mod p with a prime p ,= 2 is solvable if and only if
p 1 mod 4.
ii) What can be done for the prime residue classes mod 6 and mod 10?
The Mobius -function is dened by setting (1) = 1, (n) = (1)

if n is the
product of distinct primes, and (n) = 0 otherwise, i.e., if n has a quadratic
prime divisor.
Exercise 9. i) Show that

d|n
(d) =
_
1 if n = 1,
0 otherwise,
ii) Prove the M obius inversion formula: for two arithmetic functions f and g, the
statement
f(n) =

d|n
g(d)
is equivalent to
g(n) =

d|n
(d)f(n/d).
Section 1.2 Riemanns zeta-function 7
Exercise 10. Prove all claims about the commutative ring / of arithmetic func-
tions (for the commutativity one needs M obius inversion formula). What is the
neutral element in this ring with respect to convolution? Prove the isomorphy be-
tween the ring of arithmetic functions / and the ring T of associated Dirichlet
series! Finally, give a characterization of the invertible elements in these rings!
Hint: for some help see [2].
1.2. Riemanns zeta-function
Let s = + it with , t R and i :=

1 be a complex variable (this


mixture of greek and latin letters have become tradition since their use in
Landaus papers). The Riemann zeta-function is given by
(s) =

n=1
1
n
s
. (1.7)
This series was studied ever since the fundamentals of calculus were laid.
One of the most famous question in the early 18th century was about the
value of (2) found by Euler in 1737. Euler considered only real s in his
studies but Riemann was the rst to investigate the Riemann zeta-function
as a function of a complex variable. In his only one but outstanding paper
[175] on number theory from 1859, Riemann outlined how Gauss conjecture
(1.2) could be proved by using the function (s). As a matter of fact, it is
the complex-analytic point of view that allows to get deeper knowledge about
the zeta-function (and which therefore was unattainable for Euler). However,
at Riemanns time the theory of functions was not developed so far, but the
open questions concerning the zeta-function pushed the research in this eld
quickly forward.
1.2.1. The half-plane of absolute convergence. It is easily seen (by
Riemanns integral test) that the series (1.7) dening zeta converges abso-
lutely for > 1. Since, for
0
> 1,

n=1
1
n
s

n=1
1
n

0
1 +

n=2
_
n
n1
du
u

0
= 1 +
_

1
u

0
du = 1 +
1

0
1
,
the series in question converges uniformly in any compact subset of the half-
plane of absolute convergence > 1. A well-known theorem of Weierstrass
states that the limit of a uniformly convergent sequence of analytic functions
is analytic (see Titchmarsh [199], 2.8). Hence, (s) is analytic for > 1.
This reasoning holds far more general for Dirichlet series: in general, Dirichlet
8 Chapter 1 Classical L-functions
series converge in half-planes (provided that they do converge) and dene
analytic functions in their half-plane of uniform convergence.
Recall from the introduction identity (1.1) linking the prime numbers and
the zeta-function. The product over the primes is called Euler product in
honour of its discoverer. Our next aim is to verify this fundamental Euler
product representation. Let > 1. In view of the unique prime factorization
of the integers and the geometric series expansion,

px
_
1
1
p
s
_
1
=

px
_
1 +
1
p
s
+
1
p
2s
+ . . .
_
=

n
p|npx
1
n
s
.
Since

n=1
1
n
s

n
p|npx
1
n
s

<

n>x
1
n


_

x
du
u

=
x
1
1
tends to zero as x , we obtain identity (1.1) by sending x . Sum-
ming up, we have just proved
Theorem 1.1. (s) is analytic for > 1 and satises in this half-plane the
identity
(1.8) (s) =

n=1
1
n
s
=

p
_
1
1
p
s
_
1
.
Later we shall see more identities between Dirichlet series and Euler products,
each of them will allow us to study a certain arithmetic object (encoded in
the Euler product) by means of analysis (via Dirichlet series).
1.2.2. Riemanns memoir - proven facts. Now we study Riemanns
famous memoir [175]. Actually, he proved only two statements. First of all,
Riemann showed that the function
(s)
1
s 1
is entire; thus, (s) has an analytic continuation throughout the whole com-
plex plane except for a simple pole at s = 1 with residue 1 (corresponding
to the divergent harmonic series). Secondly, Riemann proved the functional
equation for the zeta-function: for all s C,
(1.9)

s
2

_
s
2
_
(s) =

1s
2

_
1 s
2
_
(1 s).
This shows a point symmetry for the function dened by the left-hand side
with respect to the point s =
1
2
. In view of the Euler product (1.8) it is easily
seen that (s) has no zeros in the half-plane > 1. Using the functional
Section 1.2 Riemanns zeta-function 9
equation (1.9), it turns out that (s) vanishes in < 0 exactly at the so-
called trivial zeros
(2n) = 0 for n N,
all of them being simple. This follows from some basic properties of
the Gamma-function. By Gauss product representation for the Gamma-
function,
(1.10) (z) = lim
N
N!N
z
z(z + 1)(z + 2) . . . (z + N)
,
(z) has simple poles for z = 0, 1, 2, . . . and no zeros at all. In order to
compensate the poles of (
s
2
) in (1.9) for s = 2n, (s) has to vanish there.
The behaviour of (s) is quite well understood in all of the complex plane
but the so-called critical strip 0 1 (which justies to call this strip
critical).
-14 -12 -10 -8 -6 -4 -2
-0.1
-0.075
-0.05
-0.025
0.025
0.05
0.075
Figure 1. The graph of (s) for s [14.5, 0].
All other zeros of (s) are said to be nontrivial, and it comes out that they
are all non-real (and that there location is in fact a nontrivial task). We
denote the nontrivial zeros by = + i. Obviously, they have to lie in
the critical strip 0 1. The functional equation, in addition with the
identity
(s) = (s),
show some symmetries of (s). In particular, the nontrivial zeros of (s)
have to be distributed symmetrically with respect to the real axis and the so-
called critical line =
1
2
. It was Riemanns ingenious contribution to number
theory to point out how the distribution of these nontrivial zeros is linked to
the distribution of prime numbers.
1.2.3. Analytic continuation. To set the stage for the further discus-
sion of Riemanns memoir, we shall sketch a proof of his rst result concerning
the meromorphic continuation of (s). At s = 1 the series dening the zeta-
function reduces to the harmonic series. For an analytic continuation for (s)
we have to seperate this singularity. For this purpose we shall make use of
10 Chapter 1 Classical L-functions
Lemma 1.2. Let
1
<
2
< . . . be a divergent sequence of real numbers,
dene for
n
C the function A(u) :=

nu

n
, and let F : [
1
, ) C
be a continuous dierentiable function. Then

nx

n
F(
n
) = A(x)F(x)
_
x

1
A(u)F

(u) du.
This switch from a sum to an integral is called Abels partial summation. It
is an important technical tool in analytic number theory: often integrals are
easier to handle than sums. The reader who is familiar with the Riemann-
Stieltjes integral may skip the proof.
Proof. We have
A(x)F(x)

nx

n
F(
n
) =

nx

n
(F(x) F(
n
))
=

nx
_
x
n

n
F

(u) du.
Since
1

n
u x, interchanging integration and summation yields the
assertion.
Now we apply partial summation to nite pieces of the Dirichlet series
dening zeta. Let N < M be positive integers and > 1. Then, applying
Lemma 1.2 with F(u) = u
s
,
n
= 1 and
n
= n, yields

N<nM
1
n
s
= M
1s
N
1s
+ s
_
M
N
[u]
u
s+1
du
= M
1s
N
1s
+ s
_
M
N
[u] u
u
s+1
du + s
_
M
N
du
u
s
=
1
s 1
(N
1s
M
1s
) + s
_
M
N
[u] u
u
s+1
du;
here, as usual, we write [u] for the largest positive integer less than or equal
to u. Sending M , we obtain
(1.11) (s) =

nN
1
n
s
+
N
1s
s 1
+ s
_

N
[u] u
u
s+1
du.
Since 1 < [u] u 0, it follows that the integral exists for any s with > 0
(and any value for N). Thus we have proved
Theorem 1.3. For > 0,
(s) =
s
s 1
+ s
_

1
[u] u
u
s+1
du.
Section 1.2 Riemanns zeta-function 11
Hence, (s) has an analytic continuation to the half-plane > 0 except for a
simple pole at s = 1 with residue 1.
By the functional equation (1.9) we obtain a meromorphic continuation for
the zeta-function to the whole complex plane (however, we postpone the proof
of the functional equation to Chapter 2). Taking into account properties of
the Gamma-function it turns out that the only singularity of (s) is the simple
pole at s = 1. This proves Riemanns rst statement (subject to the validity
of (1.9)).
1.2.4. Riemanns memoir - the conjectures. More spectacular than
Riemanns proven results are his conjectures. First of all, for the number
N(T) of nontrivial zeros = + i with 0 < T (counted according
multiplicities) he conjectured the asymptotic formula
N(T)
T
2
log
T
2e
;
this was proved in 1895/1905 by von Mangoldt [141, 142] who found more
precisely
N(T) =
T
2
log
T
2e
+ O(log T). (1.12)
Hence there are innitely many nontrivial zeros and their frequency increases
with their imaginary parts. Riemanns second conjecture was about the hor-
izontal distribution of the nontrivial zeros. Riemann worked with (
1
2
+ it)
and wrote
...und es ist sehr wahrscheinlich, dass alle Wurzeln reell sind.
Hiervon ware allerdings ein strenger Beweis zu w unschen; ich
habe indess die Aufsuchung desselben nach einigen uchtigen
vergeblichen Versuchen vorlaug bei Seite gelassen...
which means that very likely all roots t are real, i.e., all nontrivial zeros lie on
the so-called critical line =
1
2
. This is the famous, yet unproved Riemann
hypothesis. It had been in Hilberts famous list of 23 problems for the 20th
century and it is now one of the seven millennium problems. It should be
noticed that Riemann also calculated the rst three zeros (i.e., with respect
to their imaginary parts in the upper half-plane, ordered by their size); the
rst one is =
1
2
+ i 14.134 . . ..
Further, Riemann conjectured that there exist some constants A, B such
that
(1.13)
1
2
s(s 1)

s
2

_
s
2
_
(s) = exp(A + Bs)

_
1
s

_
exp
_
s

_
.
12 Chapter 1 Classical L-functions
-1 1 2 3
-1.5
-1
-0.5
0.5
1
1.5
Figure 2. The values of (1/2 + it) for 0 t 50.
His nal conjecture relates the prime numbers with the zeros of the zeta-
function. The so-called explicit formula states that
(x) +

n=2
(x
1
n
)
n
= Li(x)

=+i
>0
_
Li(x

) + Li(x
1
)
_
(1.14)
+
_

x
du
u(u
2
1) log u
log 2
for any x 2 not being a prime power (otherwise a term
1
2k
has to be added
on the left hand-side, where x = p
k
); the appearing logarithmic integral has
to be dened carefully by analytic continuation from (1.3). This was proved
in 1895 by von Mangoldt [141] whereas the last but one conjecture was
proved by Hadamard [72]. The explicit formula follows from both product
representations of (s), the Euler product on one side and the Hadamard
product over the zeros on the other side.
Riemanns ideas led to the rst proof of Gauss conjecture (1.2), the cele-
brated prime number theorem, by Hadamard [73] and de La Vallee Poussin
[202] (independendly) in 1896. Later in this chapter we will prove the prime
number theorem and all of Riemanns conjectures (the Hadamard product
representation and the explicit formula in this chapter, the functional equa-
tion in the following chapter, and the Riemann-von Mangoldt formula in
Chapter 3) except his hypothesis. However, rst we travel back in time
and study Dirichlets approach to the problem of prime number distribution
in arithmetic progressions.
Exercise 11. Deduce from the prime number theorem in the form (x) x/ log x
that

px
log p
p
= log x + O(1) and

px
1
p
= log log x + O(1).
Hint: partial summation.
Section 1.2 Riemanns zeta-function 13
Exercise 12. The following evaluation of (2) by elementary means is due to
Calabi: verify
3
4

n=1
1
n
2
=

m=0
1
(2m + 1)
2
=

m=0
_
1
0
_
1
0
x
2m
y
2m
dxdy
=
_
1
0
_
1
0

m=0
(xy)
2m
dxdy =
_
1
0
_
1
0
dxdy
1 x
2
y
2
.
Use the transformation
x =
sin u
cos v
and y =
sinv
cos u
in order to compute the appearing double integral above and deduce
(2) =

n=1
1
n
2
=

2
6
.
Exercise 13. * i) This provides an alternative analytic continuation for the zeta-
function: Prove
(1.15) (s) =
1
1 2
1s

n=1
(1)
n1
n
s
and show that the alternating series on the right-hand side converges for > 0.
Thus, in view of the functional equation (1.9), this yields a meromorphic contin-
uation to the whole complex plane. Where are possible singularities? None in the
half-plane of convergence but a simple pole at s = 1 of residue 1 since 1 2
1s
vanishes for s = 1 and 1
1
2
+
1
3
. . . = log 2; however, the other zeros of 1 2
1s
do not lead to singularities why?.
Hint: consider the series
_
_
_

n=0 mod 3
2

n0 mod 3
_
_
_
1
n
s
.
ii) Use (1.15) to show that (s) < 0 for 0 s < 1.
Exercise 14. Show that [(s)[ 2[s[ for
1
2
.
Exercise 15. Show that the multiplicity of any nontrivial zero = + i is
bounded above by log [[.
Hint: use the Riemann-von Mangoldt formula (1.12).
Exercise 16. Find representations in terms of the zeta-function for
(1.16) L

(s) =

n=1
(n)
n
s
and L

(s) =

n=1
(n)
n
s
,
where is Eulers -function and (n) :=

d|n
1 is the divisor function.
14 Chapter 1 Classical L-functions
1.3. Dirichlet L-functions
A special role in number theory is played by multiplicative arithmetic func-
tions and their associated generating series. Multiplicative functions respect
the multiplicative structure of N: an arithmetic function f is called multi-
plicative if f(1) ,= 0 and
f(m n) = f(m) f(n)
for all coprime integers m, n; if the latter identity holds for all integers, f
is said to be completely multiplicative. The generating Dirichlet series as-
sociated with a completely multiplicative function has, at least in a formal
way, an Euler product representation similar to the one for the Riemann
zeta-function. However, in this section we shall specify to a concrete family
of completely multiplicative functions introduced by Dirichlet [47] in 1837
in order to prove that there are innitely many primes in any prime residue
class.
1.3.1. Characters. A character is a non-trivial group homomorphism
from a nite (for the sake of simplicity) abelian group G onto C

. By the
structure theorem for nite abelian groups any such group G is the direct
product of cyclic groups. Later we will be concerned with the the multi-
plicative group of the ring of residue classes mod q, i.e., the group of prime
residue classes modulo q,
(Z/qZ)

:= a mod q : gcd(a, q) = 1.
By the chinese remainder theorem,
(Z/qZ)

p|q
(Z/p
(q;p)
Z)

,
where (q; p) denotes the exponent of the prime p in the prime factorization of
the integer q. In this case the decomposition into a product of cyclic groups is
much easier to obtain. Gauss proved that the group of residue classes modulo
q is cyclic if and only if q = 2, 4, p

or 2p

, where p ,= 2; a generator of such


a cyclic group (Z/qZ)

is called a primitive root mod q. In the case q = 2

one has
(Z/2

Z)

= 1 5
(which leads to a cyclic group if = 1, 2, since then 1 5 mod 2
2
). In any
case, the group of prime residue classes mod q is a product of nitely many
cyclic groups.
For the rst we shall argue more generally. Assume that
G =
r

j=1
G
j
with G
j
= g
j
.
Section 1.3 Dirichlet L-functions 15
In particular, any g G has a unique representation of the form
g =
r

j=1
g
t
j
j
with 0 < t
j

j
,
where
j
= G
j
is the group order of G
j
. Since a character on G is a group
homomorphism, i.e.,
(a b) = (a) (b) for all a, b G,
it follows that
(g) =
r

j=1
(g
j
)
t
j
for g =
r

j=1
g
t
j
j
.
Therefore, a character is uniquely determined by its values on the generators.
By a theorem of Lagrange, the order of any element of a nite abelian group
is a divisor of the group order (in the particular case of the group of prime
residue classes this is an older theorem of Fermat and Euler). Hence,
1 = (1) = (g

j
j
) = (g
j
)

j
,
and thus (g
j
) is an
j
-th root of unity, i.e.,
(g
j
) = exp
_
2i
k
j

j
_
for some k
j
Z with 0 < k
j

j
.
Consequently, there are at most
1
. . .
r
many characters on G. On
the contrary, any choice of k
1
, . . . , k
r
with 0 < k
j

j
denes via (g
j
) =
exp(2i
k
j

j
) such a character. Hence, the number of characters on G is equal
to the group order G =
1
. . .
r
.
We may dene the product of two characters mod q by setting
( )(g) = (g) (g);
this gives the set of characters mod q the structure of a group, the character
group (resp. dual group) of G, for short

G. Its unit element, the principal
character, is the character constant 1 and is denoted by
0
. Since [(g)[ = 1,
the inverse of a character

G is given by
(g) = (g) = (g)
1
.
Given

k
(g
j
) =
_
exp
_
2i
1

j
_
if j = k,
1 otherwise,
the mapping g
j

j
is an isomorphism between G and its character group

G. We illustrate these observations with the example G = (Z/5Z)

:
16 Chapter 1 Classical L-functions

0

1

2

3
1 2
0
+1 +1 +1 +1
2 2
1
+1 -1 +i -i
4 2
2
+1 +1 -1 -1
3 2
3
+1 -1 -i +i
We nd 2

=
2
(of course, here we can also replace 2 by 3 or
2
by
3
).
1.3.2. The orthogonality relations. Next we shall prove the impor-
tant orthogonality relations for characters, the heart of Dirichlets method.
Lemma 1.4. For g G,
1
G

G
(g) =
_
1 if g = 1,
0 otherwise,
and, for

G,
1
G

gG
(g) =
_
1 if =
0
,
0 otherwise.
Proof. Given ,=
0
, there exists an element h G with (h) ,= 1. Since
with g also gh runs through G, we get

gG
(g) =

gG
(gh) = (h)

gG
(g).
Hence,

gG
(g) = 0. The case =
0
is trivial. The second formula
follows in a similar way or, alternatively, via the isomorphism G

=

G.
Using Lemma 1.4 with g
1
a in place of g resp. with instead of and
noting that (g
1
a) = (g)(a), we obtain
Lemma 1.5. For g, a G,
1
G

G
(g)(a) =
_
1 if g = a,
0 otherwise,
and, for ,

G,
1
G

gG
(g)(g) =
_
1 if = ,
0 otherwise.
Now we restrict to groups of prime residue classes (Z/qZ)

. Via the natural


embedding of (Z/qZ)

in Z we can dene characters mod q on the whole


of Z by setting
(n) =
_
(n + qZ) if gcd(n, q) = 1,
0 otherwise.
Section 1.3 Dirichlet L-functions 17
The new objects are called Dirichlet characters mod q. The function n
(n) is completely multiplicative; moreover, it is a q-periodic function on Z,
i.e., (n + q) = (n) for any n Z. Notice that (Z/qZ)

= (q). The
orthogonality relation for characters takes therefore the form: if a and q are
coprime, then
1
(q)

mod q
(a)(n) =
_
1 if n a mod q,
0 otherwise.
(1.17)
With this tool we can sieve prime residue classes from the set of positive
integers. In view of the divergence of the sum of the reciprocals of the primes
we shall investigate the formal identity

pa mod q
1
p
=
1
(q)

mod q
(a)

p
(p)
p
. (1.18)
If we can prove the divergence of the expression on the right-hand side, then
there are innitely many prime numbers p a mod q. Of course, this makes
only sense if we assume a and q to be coprime.
1.3.3. Dirichlets prime number theorem for arithmetic progres-
sions. For > 1, the Dirichlet L-function L(s, ) associated with a character
mod q is given by
L(s, ) =

n=1
(n)
n
s
=

p
_
1
(p)
p
s
_
1
;
the proof of the identity between the Dirichlet series and the Euler product
follows along the lines of Theorem 1.1. In the special case of the principal
character
0
mod q we obtain
(1.19) L(s,
0
) =

pq
_
1
1
p
s
_
1
= (s)

p|q
_
1
1
p
s
_
;
in particular, we may regard (s) as the Dirichlet L-function to the principal
character
0
mod 1., and also for larger moduli q the Dirichlet L-function to
principal characters have a similar analytic behaviour as the zeta-function.
Theorem 1.6. Let mod q be a character ,=
0
. Then, the series

n=1
(n)n
s
converges in > 0 and uniformly in any compact subset;
in particular, L(s, ) is analytic in > 0.
Notice that the series dening L(s, ) cannot converge absolutely in 1
(and hence the Euler product representation for L(s, ) is not valid inside
the critical strip).
18 Chapter 1 Classical L-functions
Proof. Clearly, A(x) :=

nx
(n) 1. Partial summation shows

N<nM
(n)
n
s
=
A(M)
M
s

A(N)
N
s
+ s
_
M
N
A(u)
u
s+1
du
_
1 +
[s[

_
N

.
This implies the convergence; the other assertions of the theorem follow as
in the case of the zeta-function.
In particular, L(s, ) is regular in s = 1 if and only if ,=
0
. In view
of the Euler product representation there are no zeros of L(s, ) in s >
1. Consequently, we can dene the logarithm of Dirichlet L-functions (by
choosing any of its branches). We nd, for > 1,
log L(s, ) =

k=1
(p)
k
kp
ks
=

p
(p)
p
s
+ O(1). (1.20)
In view of (1.18) we shall show that Dirichlet L-functions L(s, ) do not
vanish at s = 1.
Theorem 1.7. For any character , we have L(1, ) ,= 0.
This statement is the dicult part in Dirichlets argument [47]; however,
here we shall not give his original innovative but rather complicated proof for
which he developed the analytic class number formula, an identity relating
the value L(1, ) as a nite sum with certain non-zero invariants on classes
of quadratic forms (for details of this approach we refer to Narkiewicz [159]).
We shall follow an argument of Mertens from 1897.
Proof. We may assume that is not the principal character. Let s > 1. It
follows from (1.20) and the orthogonality relation for characters (1.17) that
1
(q)

mod q
(a) log L(s, ) =

k=1
p
k
a mod q
1
kp
ks
0.
In particular, for a = 1,
(1.21)

mod q
L(s, ) 1.
Since L(s,
0
) has a simple pole at s = 1 (inherited from (s) by (1.19)) and,
by Theorem 1.6, all other L(s, ) are regular, it follows from (1.21) that there
is at most one character for which L(1, ) = 0. Since
L(1, ) = L(1, )
such a character has to be real, i.e., = .
Section 1.3 Dirichlet L-functions 19
Now suppose is real. Then we dene f = 1, resp. f(n) =

d|n
(d)
(resp. L
f
(s) = (s)L(s, )). Obviously, f is multiplicative. We nd f(p
k
) = 1
if p divides q; otherwise, if p does not divide q, then
f(p
k
) =
_
_
_
k + 1 if (p) = +1,
1 if (p) = 1 and k 0 mod 2,
0 if (p) = 1 and k 1 mod 2.
It follows that f(n) 0 and f(m
2
) 1. Therefore,

nN
2
f(n)
n
1
2

mN
f(m
2
)
m

mN
1
m
,
which diverges, as N . On the contrary, partial summation implies

nN
2
f(n)
n
1
2
=

dN
(d)
d
1
2

b
N
2
d
1
b
1
2
+

bN
1
b
1
2

N<d
N
2
b
(d)
d
1
2
= 2NL(1, ) + O(1). (1.22)
Since the left-hand side diverges to +, this yields L(1, ) ,= 0. This proves
the theorem.
In order to prove the innitude of primes in prime residue classes a mod q,
we introduce in (1.18) a variable s > 1. By (1.20), we have

pa mod q
1
p
s
=
1
(q)

mod q
(a)

p
(p)
p
s
=
1
(q)
log L(s,
0
) +
1
(q)

=
0
(a) log L(s, ) + O(1).
Sending s 1+, the rst term on the right-hand side diverges by (1.19),
and the second term converges with regard to Theorem 1.7. Hence, the series
on the left-hand side is divergent. Thus we have proved Dirichlets prime
number theorem for arithmetic progressions:
Theorem 1.8. Any prime residue class contains innitely many prime num-
bers.
We resume: the divergence of the series over all reciprocals of primes p
a mod q with coprime a and q was shown by exploiting the pole of L(s,
0
)
at s = 1, so via (1.19) once more the pole of the zeta-function (as in Eulers
proof of the innitude of primes). As we shall see later on, much of the
machinery developed for the zeta-function in order to prove Gauss conjecture
(1.2), the celebrated prime number theorem, can (with slight modications)
also be applied to Dirichlet L-functions. This will lead us to the following
generalization of the prime number theorem: let (x; a mod q) denote the
20 Chapter 1 Classical L-functions
number of primes p x in the residue class a mod q; then, for a coprime
with q,
(x; a mod q)
1
(q)
(x). (1.23)
This shows that the primes are uniformly distributed in the prime residue
classes.
In 1853, Chebyshev claimed (in a letter to Fuss, cf. [67]) that there are,
in some sense, more primes in the residue class 3 mod 4 than in the class
1 mod 4, e.g., there are 4808 primes of the rst type and only 4783 of the
second type below 100 000 and this bias seems to hold if we count more and
more primes. However, this claim is not true: Littlewood [136] showed that
there are arbitrarily large values of x such that
(x; 1 mod 4) (x; 3 mod 4)
1
2
x
1
2
log x
log log x.
Nevertheless, assuming the generalized Riemann hypothesis (which will be
explained in the following paragraph), Rubinstein & Sarnak [176] proved
that Chebyshevs claim holds for more than 99.59% of the values of x. In
general it is expected that such a phenomenon can be observed for any pair
of prime residue classes a, b mod q with a being a quadratic residue and b not
and that in the prime number race the primes p b mod q dominate over
those in a mod q. For a nice survey on this theme see Granville & Martin
[67].
1.3.4. Analytic theory of Dirichlet L-functions. Let be a char-
acter mod q. It is possible that for values of n coprime with q the character
(n) may have a period less than q. If so, we say that is imprimitive, and
otherwise primitive. If q is prime, then every character mod q is primitive.
If

is a primitive character modq

and q a multiple of q

, then we can
construct via
(n) =
_

(n) if gcd(n, q) = 1,
0 if gcd(n, q) > 1,
a character mod q, and is induced by

. We illustrate this by the


following example:
n mod 10 1 2 3 4 5 6 7 8 9 10

(n) = +1 +i i 1 0 +1 +i i 1 0
(n) = +1 0 i 0 0 0 +i 0 1 0
Every imprimitive character is induced by a primitive one. Two characters
are non-equivalent if they are not induced by the same character. If

mod
Section 1.3 Dirichlet L-functions 21
q

is a primitive character which induces another character mod q, then


L(s, ) = L(s,

p|q
_
1

(p)
p
s
_
. (1.24)
Being twists of the Riemann zeta-function with multiplicative characters,
Dirichlet L-functions share many properties with the zeta-function. For in-
stance, there is an analytic continuation to the whole complex plane, only
with the dierence that L(s, ) is regular at s = 1 if and only if is non-
principal (see Theorem 1.7). Furthermore, L-functions to primitive charac-
ters satisfy a functional equation of the Riemann-type; namely,
(1.25)
_
q

_s+
2

_
s +
2
_
L(s, ) =
()
i

q
_
q

_1+s
2

_
1 + s
2
_
L(1 s, ),
where :=
1
2
(1 (1)) and
(1.26) () :=

a mod q
(a) exp
_
2ia
q
_
is the Gaussian sum attached to . One nds a setting for the zeros which is
quite similar to the one for zeta: the trivial zeros are those which correspond
to poles of the Gamma-factors in the functional equation; all other zeros
are said to be nontrivial and they lie in the critical strip. Also for Dirich-
let L-functions it is expected that the analogue of the Riemann hypothesis
holds; more precisely: all nontrivial zeros of a Dirichlet L-function L(s, )
to a primitive character are conjectured to lie on the critical line. The re-
striction to primitive characters is made to exclude the zeros of the factor

p|q
(1

(p)p
s
) in (1.24), which all lie on the line = 0.
Exercise 17. i) Let f be a multiplicative arithmetic function. Prove the formal
identity

n=1
f(n)
n
s
=

k=0
f(p
k
)
p
ks
.
Moreover, if f is completely multiplicative, then

n=1
f(n)
n
s
=

p
_
1
f(p)
p
s
_
1
.
ii) Assume that f(n) n
c
for some non-negative constant c. Show that F(s) :=

n=1
f(n)n
s
converges in some half-plane >
a
and denes there an analytic
function; nd an explicit value for the abscissa of convergence.
Exercise 18. Prove that (n), (n), and (n) are multiplicative functions. Are
they also completely multiplicative? Can you prove Euler product representations
22 Chapter 1 Classical L-functions
for the associated Dirichlet series, i.e., for the functions in (1.16) as well as for

n=1
(n)n
s
?
Exercise 19. For an odd prime p, the Legendre symbol modulo p is dened by
_
a
p
_
=
_
_
_
+1 if X
2
a mod p is solvable,
0 if p [ a,
1 otherwise.
Prove that the Legendre symbol is a character mod p.
Hint: the squares in (Z/pZ)

form a subgroup of index 2.


Exercise 20. Determine all characters modq for q = 10, 12, 16. Compute the
structure of the corresponding character groups.
The mean-value of arithmetic functions can often be computed by counting lat-
tice point subject to some side-conditions. One of the basic techniques is Dirichlets
hyperbola method.
Exercise 21. * i) For the divisor function (n) =

d|n
1 show that

nx
(n) = xlog x + x(2C 1) + O
_
x
1
2
_
,
where C is the Euler-Mascheroni constant.
Hint: note that the left hand side counts the number of integral lattice points under
a hyperbola and write for this

bdx
1 =

bdx
d

x
1 +

bdx
b

x
1

b,d

x
1.
ii) Verify all steps in identity (1.22).
Exercise 22. * Let a and q be coprime. Prove that

pa mod q
1
p
s

1
(q)
1
s 1
.
and

(q)

px
pa mod q
1
p
log log x

1.
Can you use the latter estimate to nd an upper bound for the least prime p
a mod q?
Exercise 23. * Let be the non-principal character modulo 4. Observe that the
factors in the Euler product

p=2
_
1
(p)
p
_
are greater than 1 for primes p 3 mod 4 and less than 1 for p 1 mod 4. What
is the value of this product? How can this value be used as support for Chebyshevs
claim on the existence of more primes p 3 mod 4 than p 1mod; 4?
Section 1.3 Dirichlet L-functions 23
As a matter of fact, Euler already had an analytic proof for the innitude of
primes in the prime residue classes mod 4 (see Weil [211]). His argument shall be
recovered in the following
Exercise 24. * Let denote the non-principal character mod 4. Prove that
2 =

p
p + (p)
p (p)
.
Deduce that
1
2
log 2 =

p
(1)
(p)
p
+
1
3

p
(1)
(p)
p
3
+ . . . .
Use Maple or Mathematica in order to nd that

p
(1)
(p)
p
= 0.33498 . . . ;
deduce that there are innitely many primes in any prime residue class mod 4.
Exercise 25. * i) Let be a character modulo q and denote by () the associated
Gauss sum. Show that, for n and q coprime,
(n)() =

a mod q
(a) exp
_
an
q
_
;
if is primitive, then this identity holds for all n.
ii) For a primitive character mod q, prove that [()[
2
= q.
Hint: use i) (or search for help in [2]).
The Polya-Vinogradov inequality states that characters cannot be constant on a
long sequence of consecutive integers:
Exercise 26. * Let be a non-principal character modulo q. Prove that

nN
(n)

2q
1
2
log q.
Hint: use the previous exercise to substitute the appearing character by trigonomet-
ric expressions.
A function has at most one Dirichlet series representation:
Exercise 27. * i) Assume that
A(s) =

n=1
a(n)
n
s
and B(s) =

n=1
b(n)
n
s
are two Dirichlet series converging in some half-plane >
a
. Prove that if there
is a region in this half-plane for which A(s) = B(s), then a(n) = b(n) for all n.
ii) Deduce from i) that any convergent Dirichlet series has a zero-free half-plane.
24 Chapter 1 Classical L-functions
1.4. The prime number theorem
It was Riemanns contribution which led to the proof of Gauss conjecture
(1.2), the prime number theorem. After substantial work by von Mangoldt
and others Hadamard [73] and de la Vallee-Poussin [202] gave the rst proof
(independently) in 1896. It is legend that everyone who nds a new proof
will become one hundred years old and, indeed, both Hadamard and de la
Vallee-Poussin lived almost a century. The aim of this section is to prove
Theorem 1.9. There exists a positive constant c such that, for x 2,
(x) = Li(x) + O
_
xexp
_
c(log x)
1
9
__
.
The integral logarithm can be approximated by x/ log x; however, this is a
less good approximation to (x) as the following table illustrates.
x (x) Li(x) error in % x/ log x error in %
10
3
168 178 5.95 145 14
10
6
78498 78628 0.1656 72382 7.8
10
9
50847534 50849235 0.003345 48254942 5.1
10
12
37607912018 37607950281 0.0001017 36191206825 3.8
Out of technical reasons we prefer to work with the logarithmic derivative
of (s) (instead of log (s) as Riemann did). Logarithmic dierentiation of
the Euler product (1.8) gives for > 1

(s) =

n=1
(n)
n
s
,
where
(n) :=
_
log p if n = p
k
,
0 otherwise,
is the von Mangoldt -function. Since (s) does not vanish in the half-plane
> 1, the logarithmic derivative is analytic for > 1. As we shall see below
all desired information on (x) is encoded in
(x) :=

nx
(n) =

px
log p + O
_
x
1
2
_
. (1.27)
The idea of proof is simple. Partial summation gives
(1.28)

(s) = s
_

1
(x)
x
s+1
dx.
If we could transform this into a formula in which (x) is isolated and given
in terms of a complex integral over the zeta-function, then we might hope to
nd an asymptotic formula for (x) by contour integration methods. Indeed,
such a transformation exists (Perrons formula); however, this alone is not
sucient. In order to prove Gauss conjecture we shall also need knowledge
Section 1.4 The prime number theorem 25
on the analytic behaviour of the zeta-function on and in neighbourhood of
the line = 1.
1.4.1. A zero-free region. First of all we shall establish a zero-free
region for (s) which covers the abscissa of absolute convergence = 1. In
this delicate problem we follow (with slight modications) the ideas of de La
Vallee-Poussin (see also Titchmarsh [200]).
In the sequel we shall only argue for s = + it from the upper half-plane;
with regard to (s) = (s) all estimates below can be reected with respect
to the real axis.
Lemma 1.10. For t 8, 1
1
2
(log t)
1
2,
(s) log t and

(s) (log t)
2
.
Proof. Let 1 (log t)
1
3. If n t, then
[n
s
[ = n

n
1(log t)
1
= exp
__
1
1
log t
_
log n
_
n.
Thus, (1.11) implies
(s)

nt
1
n
+ t
1
log t.
The estimate for

(s) follows immediately from Cauchys formula

(s) =
1
2i
_
(z)
(z s)
2
dz,
where the integration is taken over the circle [zs[ =
1
2
(log t)
1
; alternatively,
one can perform (carefully) dierentiation of (1.11).
In view of the Euler product (1.8) we have, for > 1,
[( + it)[ = exp(Re log (s)) = exp
_

p,k
cos(kt log p)
kp
k
_
.
Since
(1.29) 17 + 24 cos + 8 cos(2) = (3 + 4 cos )
2
0,
it follows that
()
17
[( + it)[
24
[( + 2it)[
8
1. (1.30)
This inequality is the main idea for our following observations. In view of
the simple pole of (s) at s = 1 we have for small > 1
()
1
1
.
26 Chapter 1 Classical L-functions
Assuming that (1 + it) has a zero for t = t
0
,= 0, it would follow that
[( + it
0
)[ 1,
leading to
lim
1+
()
17
[( + it
0
)[
24
= 0,
in contradiction to (1.30). Thus, the zeta-function has no zeros on the 1-line:
(1 + it) ,= 0 for t R.
Actually, this non-vanishing argument should be compared with Mertens
proof of L(1, ) ,= 0. It can be shown that the non-vanishing of (1 + it) is
equivalent to Gauss conjecture (1.2), i.e., the prime number theorem with-
out error term, and we shall prove this equivalence in the following section.
However, here we are interested in a prime number theorem with error term.
For this purpose we have to enter the critical strip.
A simple renement of the argument allows a lower estimate for the absolute
value of (1+it): for t 1 and 1 < < 2, we deduce from (1.30) and Lemma
1.10
1
[( + it)[
()
17
24
[( + 2it)[
1
3
( 1)

17
24
(log t)
1
3
.
Furthermore, with Lemma 1.10,
(1 + it) ( + it) =
_

1

(u + it) du [ 1[(log t)
2
. (1.31)
Hence
[(1 + it)[ [( + it)[ c
1
( 1)(log t)
2
c
2
( 1)
17
24
(log t)

1
3
c
1
( 1)(log t)
2
,
where c
1
, c
2
are certain positive constants. Choosing a constant B > 0 such
that A := c
2
B
17
24
c
1
B > 0 and putting = 1 + B(log t)
8
, we obtain
[(1 + it)[
A
(log t)
6
. (1.32)
This lower bound we shall use for an estimate to the left of the line = 1.
Lemma 1.11. We have
(s) ,= 0 for 1 min1, (log t)
8
;
more precisely, there exists a positive constant c
3
such that
(1.33) [( + it)[
c
3
(log t)
6
.
Section 1.4 The prime number theorem 27
Proof. In view of Lemma 1.10 estimate (1.31) holds for 1 (log t)
8

1. Using (1.32), it follows that
[( + it)[
A c
1

(log t)
6
,
where the right-hand side is positive for suciently small . This yields
Lemma 1.11.
The largest known zero-free region for the zeta-function was found by Vino-
gradov [204] and Korobov [121] (independently). Using Vinogradovs inge-
nious method for exponential sums, they proved
(1.34) (s) ,= 0 in 1
c
(log [t[)
1
3
(log log [t[)
2
3
for some positive constant c and suciently large [t[; for a proof see Ivic [98].
However, it is still unknown whether there exists any > 0 such that (s)
does not vanish for > 1 . No progress here for almost half a century!
1.4.2. Perrons formula. The next ingredient in the proof of the prime
number theorem is
Lemma 1.12. For positive real numbers c, y, T, dene
I(y, T) =
1
2i
_
c+iT
ciT
y
s
s
ds
and
(y) =
_
_
_
0 if 0 < y < 1,
1
2
if y = 1,
1 if y > 1.
Then
[I(y, T) (y)[ <
_
y
c
min1, (T[ log y[)
1
if y ,= 1,
c/T otherwise.
The expression (y) is a good approximation to the integral I(y, T) since
(1.35) I(y, ) = lim
T
I(y, T) =
1
2i
_
c+i
ci
y
s
s
ds = (y),
and the error term is rather small.
Proof. For y = 1 and s = c + it, we nd
I(1, T) =
1
2
_
T
T
dt
c + it
=
1

_
T
0
c
c
2
+ t
2
dt =
1
2

1

_

T/c
du
1 + u
2
,
where we have used the fact that
_
U
0
du
1 + u
2
= arctanU
28 Chapter 1 Classical L-functions
and arctanU tends to

2
as U . Now it is easy to deduce the desired
estimate for [I(1, T) (1)[.
Now assume that 0 < y < 1 and r > c. Since the integrand is analytic in
> 0, Cauchys theorem implies, for T > 0,
I(y, T) =
1
2i
__
riT
ciT
+
_
r+iT
riT
+
_
c+iT
r+iT
_
y
s
s
ds.
For = r we have
[
y
s
s
[
y
r
r

1
r
.
Hence, as r ,
I(y, T) =
_

1
2i
_
+iT
c+iT
+
1
2i
_
iT
ciT
_
y
s
s
ds,
resp.
[I(y, T)[
1
T
_

c
y

d
y
c
T[ log y[
.
This is the estimate for 0 < y < 1.
Finally, if y > 1, then we integrate over the rectangular contour with
corners c iT, r iT, analogously. In this case the pole of the integrand
at s = 0 with residue
Res
s=0
y
s
s
= lim
s0
y
s
s
s = 1
gives the value (y) = 1 for the integral in question; the error estimate follows
as in the previous case.
We apply this lemma to the logarithmic derivative of the zeta-function.
Lets assume that x , Z and c > 1. Then
_
c+i
ci

n=1
(n)
n
s
x
s
s
ds =

n=1
(n)
_
c+i
ci
_
x
n
_
s
ds
s
;
here interchanging integration and summation is allowed by the absolute
convergence of the series. In view of Lemma 1.12 with T (i.e., (1.35))
it follows that

nx
(n) =
1
2i
_
c+i
ci

n=1
(n)
n
s
x
s
s
ds,
resp.
(1.36) (x) =
1
2i
_
c+i
ci
_

(s)
_
x
s
s
ds.
This is Perrons formula and, of course, it holds in a far more general setting
for arbitrary Dirichlet series in the half-plane of absolute convergence. How-
ever, for applications it is often useful to work with integrals over compact
Section 1.4 The prime number theorem 29
line segments. Lemma 1.12 yields
(x) =
1
2i
_
c+iT
ciT

(s)
x
s
s
ds + error(x, T, c),
where
error(x, T, c)
x
c
T

n=1
(n)
n
c
[ log
x
n
[
.
We split the series on the right-hand side as follows
_
_
_

|nx|>
x
4
+

|nx|
x
4
_
_
_
(n)
n
c
[ log
x
n
[
.
Since [ log
x
n
[
1
is bounded by a constant in the rst sum and log x in the
second one, we get
(x) =
1
2i
_
c+iT
ciT

(s)
x
s
s
ds +
+O
_
x
c
T

(c)

+
x(log x)
2
T
+ log x
_
. (1.37)
1.4.3. Final steps of the proof. Now we are in the position to prove
Gauss conjecture (1.2), the celebrated prime number theorem. Here we shall
combine our observation from the previous two sections.
In order to nd an asymptotic formula for the integral in (1.37) we move
the path of integration to the left. By the theorem of residues we shall obtain
contributions from the poles of the integrand, i.e.,
the zeros of (s) inside the contour,
the pole of (s) at s = 1, and
the pole of
x
s
s
at s = 0 (if surrounded by the contour);
the latter quantity is independent of x and therefore a constant. For our
purpose it is sucient to include only the pole at s = 1; however, later, when
we are going to prove the explicit formula, we have to include all appearing
poles. In view of the zero-free region of Lemma 1.11 we put c = 1 + with
= (log T)
8
, where is given by Lemma 1.11, and integrate over the
boundary of the rectangle 1 given by the corners 1 iT. By this choice
(s) does not vanish in and on the boundary of 1. The calculus of residues
30 Chapter 1 Classical L-functions
implies
_
c+iT
ciT
_

(s)
_
x
s
s
ds
=
__
1iT
1+iT
+
_
1+iT
1iT
+
_
1+iT
1+iT
__

(s)
_
x
s
s
ds
+2iRes
s=1
_

(s)
_
x
s
s
.
For the logarithmic derivative of (s) we have

(s) =
d
ds
log (s) =
1
s 1
+ O(1)
as s 1. Thus, we obtain for the residue at s = 1
Res
s=1
_

(s)
_
x
s
s
= lim
s1
(s 1)
_
1
s 1
+ O(1)
_
x
s
s
= x;
this will turn out to be the main term. It remains to bound the integrals.
For the horizontal integrals we nd with regard to Lemma 1.11
_
1+iT
1iT
_

(s)
_
x
s
s
ds
x
1+
T
.
Further, for the vertical integral,
_
1+iT
1iT
_

(s)
_
x
s
s
ds x
1
(log T)
9
.
Collecting together, we deduce from (1.37)
(x) = x + O
_
x
1+
T
+ x
1
(log T)
9
+
x(log x)
2
T
+ log x
_
.
Choosing T = exp(
1
10
(log x)
1
9
), we arrive at
(x) = x + O
_
xexp(c(log x)
1
9
)
_
for some positive constant c. Now it easily follows from (1.27) that also
(1.38) (x) :=

px
log p = x + O
_
xexp(c(log x)
1
9
)
_
.
Applying partial summation, we nd
(x) =

px
log p
1
log p
=
(x)
log x

_
x
2
(u)
d
du
1
log u
du
=
x
log x

_
x
2
u
d
du
1
log u
du + O
_
xexp
_
c(log x)
1
9
__
.
Section 1.4 The prime number theorem 31
Now partial integration shows that the rst two terms on the right-hand side
are equal to the integral logarithm (up to a constant); this nishes the proof
of the prime number theorem 1.9.
Reviewing the proof we see that the simple pole of the zeta-function is not
only the key in Eulers proof of the innitude of primes but also gives the
main term of the asymptotic formula in the prime number theorem.
In view of the largest known zero-free region (1.34) one can obtain the
following stronger form of the prime number theorem:
(1.39) (x) = Li(x) + O
_
xexp
_
c
(log x)
3
5
(log log x)
1
5
__
.
1.4.4. A probabilistic model and its limits. The prime numbers,
which on rst sight seem to be randomly distributed among the positive
integers, satisfy a strong distribution law! The prime number theorem allows
the following probabilistic interpretation: the probability that a given positive
integer n is prime is (asymptotically) equal to
1
log n
. We may use this inter-
pretation in order to make some heuristics about prime numbers of a special
shape.
The Mersenne numbers are given by M
p
= 2
p
1, where p is prime; no-
tice that if the exponent p is not prime, one can easily factor 2
p
1. For
the Mersenne numbers there exist a very simple (and fast) primality test.
Consider the following iteration
s := 4, for i from 3 to p do s := s
2
2 mod (2
p
1).
The Lucas-Lehmer test states that M
p
is prime if and only if the iteration
yields the result s = 0 (the test is simple; however, its proof is rather involved;
see [81]). The sequence of iterated values of s (not reduced mod M
p
) starts
with
s = 4 14 = 2 7 194 37 634 = 2 31 607,
from which we can read the rst two Mersenne primes 7 and 31.
1
It is
unknown whether there are innitely many Mersenne primes; however, we
might be optimistic: using the probabilistic model, a number M
p
is prime
with probability
1
log M
p

1
p log 2
,
1
The currently largest known prime number is a Mersenne prime, naemly M
30 402 457
found by Cooper & Boone in December 2005 (see http://www.mersenne.org/prime.htm for
its 9 152 052 digits and the Great Internet Mersenne Prime Search, initiated by Woltman).
32 Chapter 1 Classical L-functions
and hence the expectation value for the number of Mersenne primes is
1
log 2

p
1
p
,
which is divergent.
In the 1920s, Hardy & Littlewood developed some heuristics for more ad-
vanced questions. We illustrate their reasoning with a famous open problem.
Two numbers p and p + 2 are said to be twin primes if both p and p + 2
are prime numbers. It is a long-standing conjecture that there are innitely
many twin primes. Hardy & Littlewood [80] gave a conjectural asymptotic
formula for the number of twin primes as follows. According to our proba-
bilistic model we observe: given that n is prime, if one is supposed that n+2
to be random, its chance of being prime would be
1
log(n + 2)

1
log n
too, and so the probability of primality of both n and n+2 would be (log n)
2
.
However, if n is prime, then n + 2 can fall into n 1 residue classes mod p
for any prime p ,= n, of which p 2 are non-zero. Thus the chance that p
does not divide n + 2 is (p 2)/(p 1) rather than (p 1)/p as it would be
if n + 2 were random. Hence, we have to expect a correction factor
(p 2)/(p 1)
(p 1)/p
= 1
1
(p 1)
2
for each odd prime p; clearly, the oddest prime p = 2 is not a twin. Since
half of the integers is odd but with n also n + 2 is odd, we further have to
multiply with a factor 2. Hence, it is natural to conjecture that the number
of twin primes n, n + 2 with n x is asymptotically equal to
2

p=2
_
1
1
(p 1)
2
_
x
(log x)
2
as x . Computations support this conjecture. By his extension of Er-
atostheness sieve method, Brun [27] showed that the number of twin primes
below x is bounded above by O(x/(log x)
2
) which implies the convergence of
the series over the reciprocals of twin primes:

px
p+2 prime
1
p
< ,
in contrast to the divergence of the sum of reciprocals of all primes. Bruns
result indicates that almost all primes are not twin primes. A more general
conjecture is the one of Bateman & Horn [13] on prime values of polynomials
which seems to be far out of reach with present day methods.
Section 1.4 The prime number theorem 33
In 1936, Cramer [42] introduced the following model for the distribution
of prime numbers:
Let U
1
, U
2
, U
3
, . . . be an innite series of urns containing black
and white balls, the chance of drawing a white ball from U
n
being
1
log n
for n > 2 while the composition of U
1
and U
2
may
be arbitrarily chosen. We now assume that one ball is drawn
from each urn, so that an innite series of alternately black
and white balls is obtained. If P
n
denotes the number of the
urn from which the nth white ball in the series was drawn, the
numbers P
1
, P
2
, . . . will form an increasing sequence of integers,
and we shall consider the class C of all possible sequences (P
n
).
Obviously the sequence S of prime numbers (p
n
) belongs to this
class. We shall denote by (x) the number of those P
n
which
are x, thus forming an analogy to the ordinary notation (x)
for the number of primes p
n
x. (. . .) As a matter of fact, it
may be shown that, with probability 1, the relation
limsup
x
[(x) Li (x)[

2x
_
log log x
log x
= 1
is satised. With respect to the corresponding dierence (x)
Li (x) in the prime number problem, it is known that, if the
Riemann hypothesis is assumed, the true maximum order of
this dierence lies between the functions

x
log x
and

xlog x. It
is interesting to nd that the order of the function occurring in
the denominator in the above equation falls inside this interval
of indetermination.
Cramer used this model in order to conjecture an asymptotic formula for
the largest gap between consecutive primes. Denote by p
n
the nth prime in
ascending order. Cramer was led to conjecture that
max
pnx
(p
n+1
p
n
) (log x)
2
.
This seems to be a good guess but only little is known in this direction.
Recently, a related problem was solved by Goldston, Pintz & Yildirim [65].
They showed that there exist
liminf
n
p
n+1
p
n
log p
n
= 0.
Their method depends on the level of distribution of primes in arithmetic
progressions. Assuming the Elliott-Halberstam conjecture (which is to com-
plicated to be given here), they also proved that there are innitely often
primes diering by 16 or less. This is a remarkable progress towards the twin
prime conjecture!
34 Chapter 1 Classical L-functions
Another question in prime number distribution theory asks for which func-
tions (x) does
(x + (x)) (x)
(x)
log x
hold as x ? Huxley [92] proved that one can choose (x) = x
7
12
+
;
under assumption of the Riemann hypothesis one can replace the exponent
by
1
2
+ . Assuming Riemanns hypothesis, Selberg [178] proved that the
asymptotic formula is true for almost all values of x provided (x)/(log x)
2
tends with x to innity; here the notion for almost all values of x means
that the set of exceptional x u has Lebesgue measure o(u) as u .
Cramers probabilistic model predicts that one can relax Selbergs condition
to (x) (log x)
2
. However, this was disproved by Maier [140] who showed
by his celebrated matrix method that
limsup
x
(x + (x)) (x)
(x)/ log x
> 1 and liminf
x
(x + (x)) (x)
(x)/ log x
< 1,
where (x) = (log x)

for > 1. Hence, the local distribution of primes does


not follow this simple probabilistic model. This observation started the search
for further violations of Cramers model as well as for suitable modications.
A nice survey on this topic is Granville [66].
Exercise 28. Prove (1.27) and (1.38). Further, ll the gaps left in the proof of
Lemma 1.12.
Exercise 29. Try other trigonometric identities like (1.29) in order to obtain a
better error term in the prime number theorem.
Exercise 30. Prove that if p
n
is the n-th prime number, then
p
n
nlog n.
Show that this also implies the prime number theorem without error term.
Exercise 31. Prove a Perron type formula for arbitrary Dirichlet series in their
half-plane of absolute convergence.
Exercise 32. * Prove the following prime number theorem for arithmetic pro-
gressions: Let a and q be xed positive and coprime integers. Then there exists a
positive constant c, depending only on a mod q, for which
(x, a mod q) =
1
(q)
Li (x) + O
_
xexp
_
c(log x)
1
9
__
.
Exercise 33. * Denote by d
k
(n) the number of representations of the positive
integer n as a product of k positive integers.
i) Show that, for > 1,
(s)
k
=

n=1
d
k
(n)
n
s
.
Section 1.5 Tauberian theorems 35
ii) Prove that

nx
d
k
(n) = xP
k
(log x) + error,
where P
k
(X) is a polynomial of degree k 1 in X, equal to Res
s=1
(s)
k x
s
s
and
the error term is reasonably small. In the case k = 2 compare with the result of
Exercise 21 i).
The numbers F
k
= 2
2
k
+ 1 with k = 0, 1, 2, . . . , are called Fermat numbers.
Gauss showed that the regular n-gon can be constructed only by use of ruler and
compass if and only if n is a power of 2 times a product of distinct prime Fermat
numbers.
Exercise 34. Compute the rst 10 Fermat numbers and test whether they are
prime. Using the heuristics of the previous section, state a conjecture on the num-
ber of prime Fermat numbers.
The famous open Goldbach conjecture claims that any even integer greater than
or equal to 4 can be written as a sum of two prime numbers.
Exercise 35. Use the Hardy & Littlewood heuristics
i) to nd an asymptotic formula for the number of representations of a large positive
integer as a sum of two primes, and
ii) to state a conjectural asymptotic formula for the number of primes of the form
p = n
2
+ 1 below x.
Recently, Green & Tao [69] proved a famous conjecture, namely, that the set of
prime numbers contains arbitrarily long arithmetic progressions (see also Greens
survey [68]).
Exercise 36. * Use the probabilistic model in the form
Prob.(n is prime : 1 n N)
to show that
Expect. (1 n, d N : n, n + d, . . . , n + (k 1)d are all prime)
N
2
(log N)
k
,
under the assumption that the events that n + jd and n + d for j ,= are all
independent. Can you make the asymptotics more precise by a reasoning a la
Hardy & Littlewood?
1.5. Tauberian theorems a general approach
In number theory we are often faced with the following problem: given a
sequence of complex numbers a(n), we want to know the behaviour of the
summatory function

nx
a(n). The prime number theorem may be a rst
example and thus one may try to study the generating Dirichlet series by use
of Perrons formula. This approach can be streamlined and the output can
36 Chapter 1 Classical L-functions
be found in so-called Tauberian theorems, developed by Hardy & Littlewood,
Ikehara, Wiener, and many others. A good overview on this rich theory gives
Korevaar [120].
1.5.1. The theorem of Wiener-Ikehara. Abel proved that

n=0
a(n) = 1 implies that

n=0
a(n)x
n
tends to 1 as x 1. In 1897,
Tauber [196] proved that the converse implication holds if na(n) = o(1).
After Tauber plenty of similar results were proven, many of them with direct
applications to number theory (created with number theoretical motivation
in mind). Following Bochner, resp. Chandrasekharan [29] here we shall
prove the Tauberian theorem of Ikehara [95] and Wiener [212]:
Theorem 1.13. Let A(x) be a non-negative, non-decreasing function of x
[0, ). Suppose that the integral
_

0
A(x) exp(sx) dx
converges to the function f(s) and that f(s) is analytic in the half-plane
1, except for a simple pole at s = 1 with residue 1. Then
lim
x
A(x) exp(x) = 1.
Proof. Dene B(x) = A(x) exp(x). First we shall prove that, for any
positive ,
(1.40) lim
y
_
y

B
_
y
v

_
_
sin v
v
_
2
dv = .
For > 1, we have
f(s) =
_

0
A(x) exp(sx) dx and
1
s 1
=
_

0
exp((1 s)x) dx.
Thus,
F(s) := f(s)
1
s 1
=
_

0
(B(x) 1) exp((1 s)x) dx.
By assumption F(s) is analytic for 1. Now dene F

(t) = F(1 + + it)


for > 0. For > 0, we obtain
_
2
2
F

(t)
_
1
[t[
2
_
exp(iyt) dt
=
_
2
2
_
1
[t[
2
_
exp(iyt) (1.41)

__

0
(B(x) 1) exp(( + it)x) dx
_
dt.
Section 1.5 Tauberian theorems 37
Next we want to interchange the order of integration on the right-hand side.
Since A(x) is non-negative and non-decreasing, for real s and x > 0,
f(s) A(x)
_

x
exp(su) du =
A(x) exp(sx)
s
,
resp. A(x) sf(s) exp(sx). Since f(s) is analytic for > 1, this implies
A(x) = O(exp(sx)) for any s > 1 and
B(x) exp(x) = A(x) exp((1 + )x) = o(1)
for every > 0. It follows that the integral
_

0
(B(x) 1) exp(( + it)x) dx
converges uniformly for 2 t 2. Thus, we can interchange the order
of integration in (1.41) and obtain
_

0
(B(x) 1) exp(x)
__
2
2
exp(i(y x)t)
_
1
[t[
2
_
dt
_
dx.
This leads with (1.41) to
_
2
2
F

(t)
_
1
[t[
2
_
exp(iyt) dt
= 2
_

0
(B(x) 1) exp(x)
(sin((y x)))
2
(y x)
2
dx. (1.42)
Since F(s) is analytic in 1, it follows that F

(t) tends to F(1 + it) as


0, uniformly for 2 t 2. Moreover,
lim
0
_

0
exp(x)
(sin((y x)))
2
(y x)
2
dx =
_

0
(sin((y x)))
2
(y x)
2
dx.
We deduce
lim
0
_

0
B(x) exp(x)
(sin((y x)))
2
(y x)
2
dx =
_

0
B(x)
(sin((y x)))
2
(y x)
2
dx.
By (1.42),
1
2
_
2
2
F(1 + it)
_
1
[t[
2
_
exp(iyt) dt (1.43)
=
_

0
(B(x) 1)
(sin((y x)))
2
(y x)
2
dx.
The Riemann-Lebesgue lemma states that
lim
y
_

f(x) exp(ixy) dx = 0
38 Chapter 1 Classical L-functions
for any absolutely integrable function f. Thus, letting y , the left-hand
side of (1.43) tends to zero while
(1.44) lim
y
_

0
(sin((y x)))
2
(y x)
2
dx = lim
y
_
y

_
sin v
v
_
2
dv = .
Hence,
lim
y
_

0
B(x)
(sin((y x)))
2
(y x)
2
dx = ;
this proves (1.40).
In order to prove the theorem we have to show
(1.45) 1 liminf
x
B(x) limsup
x
B(x) 1.
Clearly, this implies the existence of the limit lim
x
B(x) and that this
limit is equal to 1. For given positive numbers a and let y >
a

. By (1.40),
limsup
y
_
a
a
B
_
y
v

_
_
sin v
v
_
2
dv
(the integrand is non-negative). Since A(u) = B(u) exp(u) is non-decreasing,
we have, for a v a,
B
_
y
a

_
exp
_
y
a

_
B
_
y
v

_
exp
_
y
v

_
.
This implies
B
_
y
v

_
B
_
y
a

_
exp
_
v a

_
B
_
y
a

_
exp
_

2a

_
.
Hence,
limsup
y
B
_
y
a

_
exp
_

2a

__
a
a
_
sin v
v
_
2
dv
= limsup
y
_
a
a
B
_
y
v

_
_
sin v
v
_
2
dv .
For xed a and we have limsup
y
B(y
a

) = limsup
y
B(y). Thus,
exp
_

2a

_
limsup
y
B(y)
_
a
a
_
sin v
v
_
2
dv ,
being valid for all a > 0 and > 0. Letting a, such that
a

0, we
deduce
limsup
y
B(y)
_

_
sin v
v
_
2
dv .
Section 1.5 Tauberian theorems 39
Now (1.44) implies the desired upper bound for limsup
y
B(y). The just
proved inequality yields the existence of a constant c such that [B(x)[ c.
Hence, for xed positive a and and a suciently large y,
_
y

B
_
y
v

_
_
sin v
v
_
2
dv
c
__
a

+
_

a
__
sin v
v
_
2
dv +
_
a
a
B
_
y
v

_
_
sin v
v
_
2
dv. (1.46)
As above, we have B(y
v

) B(y +
a

) exp(
2a

) for a v a. Therefore,
_
a
a
B
_
y
v

_
_
sin v
v
_
2
dv B
_
y +
a

_
exp
_
2a

__
a
a
_
sin v
v
_
2
dv.
From (1.40), (1.46) and the latter inequality it follows that
c
__
a

+
_

a
__
sin v
v
_
2
dv +
+liminf
y
B
_
y +
a

_
exp
_
2a

__
a
a
_
sin v
v
_
2
dv.
Here we may replace liminf
y
B(y +
a

) by liminf
y
B(y). Then, after
sending a, such that
a

0, we get the desired lower bound for


liminf
y
B(y). The theorem is proved.
Now we shall derive a reformulation of Theorem 1.13 to which we also refer
to as the Theorem of Wiener & Ikehara:
Theorem 1.14. Let F(s) =

n=1
a(n)n
s
be a Dirichlet series with non-
negative real coecients and absolutely convergent for > 1. Assume that
F(s) can be extended to a meromorphic function in 1 such that there are
no poles except for a possible simple pole at s = 1 with residue r 0. Then
A(x) :=

nx
a(n) = rx + o(x).
Proof. Without loss of generality we may suppose that the residue is posi-
tive: r > 0, since otherwise we can consider the function F(s) + (s) (which
then has residue r +1 = 1). Furthermore, we may assume that r = 1 simply
by replacing a(n) by a(n)/r.
By partial summation (as we did for zeta in the proof of Theorem 1.3),
F(s) = s
_

1
A(x)
x
s+1
dx,
resp.
F(s)
s
=
_

0
A(exp(y)) exp(ys) dy
40 Chapter 1 Classical L-functions
with x = exp(y). Now in view on all assumptions on F(s) it follows from
Theorem 1.13 that
lim
y
A(exp(y)) exp(y) = 1.
Re-substituting x = exp(y) we get the assertion.
1.5.2. The prime number theorem for arithmetic progressions
- revisited. As rst application of the just proven Tauberian theorem we
return to the question how the prime numbers are distributed in the residue
classes. It is natural to ask for a quantitative version of Dirichlets prime
number theorem for arithmetic progressions. Here we shall prove (1.23).
Let mod q be a character. Similarly as for the Riemann zeta-function, we
consider the logarithmic derivative of a Dirichlet L-functions L(s, ), given
by
L

L
(s, ) =

n=1
(n)(n)
n
s
,
where (n) is the von Mangoldt-function (introduced in the previous section).
We dene
(x; ) =

nx
(n)(n) and (x; a mod q) =

nx
na mod q
(n).
By the orthogonality relation for characters (1.17), we nd
(x; a mod q) =
1
(q)

mod q
(a)(x; ).
Now suppose that a and q are coprime (otherwise the functions in the latter
identity are all bounded). We want to apply Theorem 1.14 with the functions
F(s) =

mod q
(a)
L

L
(s, ) and A(x) = (x; a mod q).
Notice that the left-hand side has a Dirichlet series representation for > 1
with non-negative coecients. From Section 1.3 we know that L(s, ) is
analytic for 1 if is not the principal character. In the case of the
principal character we nd, by (1.19),

L
(s,
0
) =
1
s 1
+ higher terms.
Finally, we have to assure that any of the appearing L(s, ) has no zero on the
1-line. By Theorem 1.7 we already know that L(1, ) ,= 0. For an arbitrary
point s = 1 + it with t ,= 0 one may argue as we did for the zeta-function in
the previous section just by replacing (1.30) by
L(,
0
)
17
[L( + it)[
24
[L( + 2it,
2
)[
8
.
Section 1.5 Tauberian theorems 41
Thus, applying Theorem 1.14 we obtain (x; a mod q) (q)
1
x. By partial
summation, this implies
Theorem 1.15. Let a and q be coprime integers. Then, as x ,
(x; a mod q)
1
(q)
x
log x
.
We did not use any information about the behaviour of the involved Dirichlet
L-functions from inside the critical strip. Therefore, we do not get an error
term. One can easily prove an asymptotic formula with error term following
the argument we gave for zeta; however, for applications one often wants to
have a result which is uniform in the modulus; for instance, for bounds of
the least prime in an arithemtic progression. The theorem of Page-Siegel-
Walsz provides such an asymptotic formula which is uniform in a small
region of values q. In this case the situation is more delicate than for the
zeta-function of xed modulus. In principle, one cannot exclude that certain
L(s, ) have real zeros on the real axis inside the critical strip. These so-
called exceptional zeros (or Siegel zeros) are dicult to deal with. We shall
not go into the details here but refer to Prachar [172].
1.5.3. Dedekind zeta-functions and the prime ideal theorem. Be-
fore we can introduce this further class of L-functions we have to recall some
basic facts from algebraic number theory. A good reading on this topic is
Swinnerton-Dyer [195].
A complex number is said to be algebraic over Q if there exists a non-
zero polynomial P(X) with integer coecients such that P() = 0; the
polynomial with coprime coecients and least degree having this property is
called the minimal polynomial of and is denoted by P

. The degree of the


minimal polynomial is said to be the degree of . For algebraic , the set
Q[X]/P

(X)
is a nite algebraic extension of the eld of rational numbers, the algebraic
number eld associated to ; a more convenient form is
Q() :=
_
a
0
+ a
1
+ . . . + a
d1

d1
: a
j
Q
_
,
where d = deg = deg P

. The degree of the eld extension Q()/Q is


equal to the dimension of the eld Q() as a Q-vector space, and we write
d = [Q() : Q]. Note that any number in Q() is algebraic of degree d.
The zeros

1
, . . . ,

d
of the minimal polynomial P

(X) are the conjugates of


(in fact, they are the images of under the eld automorphisms) and have
degree equal to d = deg . Denote by
1
, . . . ,
d
the embeddings of K in C.
Then the discriminant of K is given by
d
K
= det ((
i
(
j
))
1i,jd
)
2
.
42 Chapter 1 Classical L-functions
In the case of a quadratic number eld, i.e., [Q() : Q] = 2, this discriminant
is equal to the discriminant of the minimal polynomial P

(X). The product


of all conjugates is the norm of :
N() :=
d

j=1

j
.
The norm provides a measure for the size of algebraic numbers. An algebraic
number is said to be an algebraic integer if its minimal polynomial is monic; in
this case, the norm is, up to the sign, equal to the constant term P

(0) Z
in the minimal polynomial. The notion of algebraic integers extends the
standard notion of integers from Q to number elds. In fact, one can show
that an algebraic integer, which is rational, is a rational integer. The set of
all algebraic integers in a number eld K forms a ring O
K
, the so-called ring
of integers. Unfortunately, these rings in general do not have a unique prime
factorization. For example, the identity
2 3 = (1

5) (1 +

5)
gives two distinct factorizations of 6 in the ring Z[

5] into irreducible fac-


tors. In order to obtain unique factorization we have to pass to ideals.
An ideal a of O
K
is a set of integers in O
K
having the properties
if , a, then + a,
if a and O
K
, then a.
If a is a non-zero ideal of O
K
, then O
K
/a is a ring; its cardinality is denoted
by N(a) and is called the norm of a. An ideal a ,= (0) is said to be fractional
if there exists an integer ,= 0 for which a is an ideal of O
K
. An ideal a (not
necessarily fractional) is called principal if there exists K if a = O
K
.
A fractional ideal lies in O
K
if and only if it is an ideal of O
K
, in which case
we say that it is an integral ideal. In O
K
every fractional ideal is invertible,
i.e., a
1
a O
K
. The set of fractional ideals forms a group. An ideal p of
O
K
is said to be a prime ideal if p ,= O
K
and if the quotient ring O
K
/p is an
integral domain, i.e., p implies p or p. A non-zero prime ideal
is maximal. And, most importantly, every fractional ideal a has a unique
factorization into a product of powers of prime ideals.
We shall give an example and consider quadratic number elds: K =
Q(

D), where D is a squarefree integer. It is not too dicult to show


that every rational prime p splits in Q(

D) into prime ideals according to


the value of the Legendre symbol (
d
K
p
): a rational prime p is said to be
inert: (p) = p if (
d
K
p
) = 1,
ramied: (p) = p
2
if (
d
K
p
) = 0,
split: (p) = p
1
p
2
with p
1
,= p
2
if (
d
K
p
) = +1.
Section 1.5 Tauberian theorems 43
In the case of p = 2 we nd 2 is inert if D 5 mod 8; otherwise
(2) =
_
_
_
(2,
1
2
(1 +

D))((2,
1
2
(1

D)) if D 1 mod 8,
(2, 1 +

D)
2
if D 3, 7 mod 8,
(2,

D)
2
if D 2, 6 mod 8.
Now we are in the position to introduce a new zeta-function which car-
ries information about the arithmetic of number elds and is named after
Dedekind who set the foundations of ideal theory. The Dedekind zeta-
function of a number eld K over Q is given by

K
(s) =

a
1
N(a)
s
=

p
_
1
1
N(p)
s
_
1
,
where the sum is taken over all non-zero integral ideals a and the product is
taken over all prime ideals p of the ring of integers of K. The identity between
series and product is an analytic version of the unique factorization of integral
ideals in prime ideals (analogously to the unique prime factorization of the
integers). Since the norm of an integral ideal is a positive rational integer,
the series can be rewritten as a Dirichlet series:

a
1
N(a)
s
=

n=1
f
K
(n)
n
s
,
where f
K
(n) counts the number of integral ideals a with N(a) = n. We
see that the Riemann zeta-function is the Dedekind zeta-function of Q and,
as a matter of fact, Dedekind zeta-functions share many properties with
Riemanns zeta. First of all, we have to show that the Dirichlet series dening
the Dedekind zeta-function
K
(s) converges for > 1, independent of the eld
K. To see this note that, for real s > 1,
[
K
(s)[ =

p
_
1
1
N(p)
s
_
1

p
_
1
1
p

_
d
= ()
d
,
since there are at most d = [K : Q] many primes p lying above each rational
prime p and N(p) is smallest if (p) splits completely.
We return to our example K = Q(

D). We write for short d := d


Q(

D)
(since now there wont be any confusion with the degree), which is equal to D
if D 1 mod 4, and equal to 4D if D 2, 3 mod 4. In view of the splitting
44 Chapter 1 Classical L-functions
of the primes one easily nds

Q(

D)
(s) =

(
d
p
)=+1
_
1
1
p
s
_
2

(
d
p
)=0
_
1
1
p
s
_
1

(
d
p
)=1
_
1
1
p
s
_
1
_
1 +
1
p
s
_
1
(1.47)
= (s)L(s,
d
),
with the Jacobi symbol, dened by

d
: N C, n
_
d
n
_
=

j=1
_
d
p
j
_
,
where n = p
1
. . . p

is the factorization of the integer n into prime factors


(not necessarily distinct).
In 1917, Hecke [86] obtained the rst deeper results concerning the analytic
behaviour of Dedekind zeta-functions. He showed that (s1)
K
(s) is an entire
function and that the Dedekind zeta-function has a simple pole at s = 1 with
residue
(1.48) lim
s1+
(s 1)
K
(s) =
2
r
1
(2)
r
2
hR

_
[d
K
[
,
where r
1
is the number of real conjugate elds, 2r
2
is the number of complex
conjugate elds, h is the class number, R is the regulator, is the number
of roots of unity, and d
K
is the discriminant of K. We see there is a lot of
arithmetic information is encoded in this residue! The class number is the
number of equivalence classes of fractional ideals of K, and so it measures the
deviation of O
K
from having unique prime factorization. Gauss conjectured
that the class numbers h = h(d) of an imaginary quadratic number eld
K = Q(

D) with discriminant d < 0 tend with d to innity; notice that


d = D if D 1 mod 4, and d = 4D if D 2, 3 mod 4. This was rst
proved by Heilbronn [90] and in rened form by Siegel [183]. The problem
of nding an eective algorithm to determine all imaginary quadratic elds
with a given class number h is known as the Gauss class number h problem.
This problem is of interest with respect to the non-existence of exceptional
real zeros of Dirichlet L-functions o the critical line. The general Gauss
class number problem was solved by Goldfeld, Gross & Zagier [63, 71]. A
complete determination of the imaginary quadratic elds with class number 1
was rst given by Heegner [89] (but his solution was not completely accepted
due to a number of gaps), Baker [10], and Stark [187] (independently):
h = 1 d 3, 4, 7, 8, 11, 19, 43, 67, 163.
Section 1.5 Tauberian theorems 45
Notice that class number 1 is equivalent to unique prime factorization in the
corresponding ring of algebraic integers. Further, note that (1.48) contains
the information of Dirichlets analytic class number formula.
We want to apply the theorem of Wiener & Ikehara 1.14. Again we consider
the logarithmic derivative:

K
(s) =

K
(a)
N(a)
s
,
where

K
(a) :=
_
log N(p) if a = p
k
,
0 otherwise.
Furthermore,

K
(s) =
1
s 1
+ higher terms,
independent of the residue of
K
(s) at s = 1. Finally we have to assure that

K
(s) has no zeros or further poles on the line = 1. One can show that the
Dedekind zeta-function of any number eld K can be written as

K
(s) =

L(s, )

,
where the product is taken over so-called Artin L-functions (a class of Dirich-
let L-functions we shall meet in Chapter 4) and the exponents

are integers
(not necessarily positive). However, in certain cases life is much easier: for in-
stance, if K is a cyclotomic eld, then this product is nothing but the product
of certain Dirichlet L-functions with positive exponents. Since L(1+it, ) ,= 0
for all real numbers t and all characters , it then immediately follows that

K
(s) does not vanish on the 1-line. However, this is true for any Dedekind
zeta-function. In general, one can prove the non-vanishing by another result
of Hecke (or Exercise 41). Now we can apply Theorem 1.14 and get

K
(x) :=

N(a)x

K
(a) x.
Dene

K
(x) = p O
K
prime : N(p) x.
By a standard application of partial summation we deduce the prime ideal
theorem:
Theorem 1.16. Let K be a number eld. Then, as x ,

K
(x)
x
log x
.
46 Chapter 1 Classical L-functions
The rst proof of the prime ideal theorem was given in 1903 by Landau
[124]. On the rst view it might be surprising that the right-hand side does
not depend on the number eld K. The residue of
K
(s) at s = 1 contains data
about the underlying eld; however, the residue of the logartihmic derivative
of
K
(s) at s = 1 is equal to 1 independent of K.
We conclude with two remarks which are of special interest with respect to
our later studies. Hecke [86] proved that Dedekind zeta-functions satisfy a
functional equation:
_
_
[d
K
[
2
r
2

n
2
_
s

_
s
2
_
r
1
(s)
r
2

K
(s)
=
_
_
[d
K
[
2
r
2

n
2
_
1s

_
1 s
2
_
r
1
(1 s)
r
2

K
(1 s). (1.49)
It is further expected that also the analogue of the Riemann hypothesis is
true, i.e., all non-real zeros of
K
(s) lie on the critical line, resp. there are no
zeros for >
1
2
.
1.5.4. The prime number theorem and non-vanishing. Next we
shall prove that the prime number theorem without error term is equivalent
to the non-vanishing of (s) on the line = 1. One implication follows
immediately from the theorem of Wiener-Ikehara (and can be proved just
along the lines of the previous applications of Theorem 1.14). To see the
other implication we assume that
(1.50) (x) x.
We have to deduce that there are no zeros of (s) on the line = 1. For this
purpose dene, for > 1,
(s) =

(s)
s(s)

1
s 1
=
_

1
(x) x
x
s+1
dx.
Clearly, (s) is regular in > 0 except for simple poles at the zeros of the
zeta-function. It should be noticed that the logarithmic derivative has only
simple poles! Now (1.50) implies that, given > 0, there exists a real number
x
0
such that for x > x
0
we have [(x) x[ < x, and so we nd for > 1
[(s)[ <
_
x
0
1
[(x) x[
x
+1
dx +
_

x
0

dx < C +

1
,
where C is a constant, depending only on . Hence,
( 1)[( + it)[ < C( 1) + .
Thus, for any xed t, the limit of the left-hand side is 0 as 1+. However,
if (1+it) = 0 for some t ,= 0, then the limit of (1)(+it) would be equal
Section 1.5 Tauberian theorems 47
to the residue of (s) at the simple pole s = 1 + it, and therefore dierent
from zero. Of course, the same reasoning applies to Dirichlet L-functions.
It was a big surprise when Erdos [50] and Selberg [179] obtained an ele-
mentary proof of the prime number theorem; here the attribute elementary
means that the proof does not use any arguments from analysis (apart from
simplest properties of the logarithm). Hence, the non-vanishing of (s) for
1 can be shown without complex analysis! The proofs of Erdos and
Selberg are not independent and their actual contributions are still under
discussion. (For the history of this quarrel read Goldfeld [64] and for an
elementary proof see [81].) In the meantime, even elementary proofs of the
prime number theorem with error term were given. Nevertheless, the analytic
approach yields more information on prime number distribution.
Exercise 37. Prove the prime number form in the form (x)
x
log x
by using the
Theorem of Wiener & Ikehara 1.14.
Exercise 38. Give a rigorous proof of (1.44), i.e., prove that, for any > 0,
_

(sin(x))
2
x
2
dx = .
Exercise 39. * Prove the following variant of Theorem 1.14: Let a(n) be sequence
of complex numbers, b(n) be a sequence of non-negative real numbers, and dene
A(s) =

n=1
a(n)
n
s
and B(s) =

n=1
b(n)
n
s
.
Suppose that
[a(n)[ b(n),
the series dening B(s) converges for > 1, and
A(s) has an analytic continuation to 1 except for at most a simple
pole at s = 1 with residue r.
Then, as x ,

nx
a(n) = rx + o(x).
Hint: rst, show that without loss of generality you can assume that the a(n) are
all real; for this aim introduce another Dirichlet series having Dirichlet series co-
ecients a(n) and write A(s) as the sum of two Dirichlet series, one having real
coecients and one with coecients in iR.
Exercise 40. Show that any Dirichlet L-function L(s, ) is non-vanishing on the
line = 1.
Hint: consider the function L(,
0
)
17
[L( + it, )[
24
[L( + 2it,
2
)[
8
and argue as
in the case of the zeta-function.
48 Chapter 1 Classical L-functions
Exercise 41. Assume that f(s) is an analytic function in > 1 without zeros (so
that we can dene a logarithm) and
log f(s) =

n=1
a(n)
n
s
with a(n) 0. Further, suppose that f(s) is analytic on = 1 except for a pole of
order m 0 at s = 1. Prove that if f(s) has a zero s = 1 + it
0
, then its order is
m/2.
Hint: If s = 1 + it
0
is a zero of order k > m/2, then consider the function
f(s)
2k+1
2k

j=1
f(s + ijt
0
)
2(2k+1j)
.
Exercise 42. * i) Prove the decomposition (1.47).
ii) Verify functional equation (1.49) in the case of quadratic number elds.
Exercise 43. * Show that both Q(i) and Q(

5) have class number h = 1.


Hint: use the previous exercise and compute the corresponding value L(1,
d
).
Let r(n) count the number of ways the positive integer n can be written as a
sum of two integer squares (with repetition, i.e, r(n) = (a, b) Z
2
: n = a
2
+b
2
.
Exercise 44. i) Show that

Q(i)
(s) =
1
4

n=1
r(n)
n
s
and deduce that

nx
r(n) x.
Hint: use the last but one exercise and Theorem 1.14.
ii) Use geometric arguments in order to prove the last statement with an error term
O(

x).
Hint: how many integer lattice points lie in a circle of radius r =

x centered at
the origin?
The circle problem is to nd the best possible error term in the mean-value formula
for r(n). It is known that [

nx
r(n) x[ is for innitely many values of x larger
than x
1
4
and always bounded by x
131
416
+
. The rst result is due to Hardy [76] and
Landau [125] (independently); a slight but remarkable improvement of the lower
bound by some log-powers was found by Soundararajan [185]. The upper bound
is from Huxley [93] and there is hope that renements of techniques in the theory
of exponential sums will lead to a smaller exponent.
Exercise 45. Deduce from the prime ideal theorem for Q(i) and the splitting of
primes that the prime numbers are equidistributed in the prime residue classes
modulo 4.
Section 1.6 The explicit formula 49
1.6. The explicit formula
Now we want to prove Riemanns explicit formula (1.14) which links the
prime numbers directly with the zeros of the zeta-function. However, while
Riemann dealt with the prime counting function (x) (see (1.14)) we shall
work with the more simple function (x) =

p
k
x
log p, introduced in Sec-
tion 1.4, and prove that, for x ,= p
k
,
(1.51) (x) = x


1
2
log
_
1
1
x
2
_
log(2).
Notice that the right hand side above is not absolutely convergent. If (s)
would have only nitely many nontrivial zeros, the right hand side would be
a continuous function of x, contradicting the jumps of (x) for prime powers
x. The derivation of the explicit formula relies on a more detailed study of
basic analytic properties of the Riemann zeta-function and it provides us a
better understanding on the nature of the error term in the prime number
theorem. Here we shall work in the more general setting of Dirichlet L-
function; nevertheless, we follow closely von Mangoldts original approach
[141] for zeta. First of all we have to recall some facts from the theory of
functions.
1.6.1. Entire functions of nite order. The theory of entire functions
was founded by Weierstrass [208] in 1876 and was further developed in the
1890s by the path-breaking works of Picard and Hadamard [72]. We start
with some observations concerning the zeros of entire functions. The main
tool is Jensens formula:
Lemma 1.17. Let f(s) be an analytic function for [s[ r with zeros

1
, . . . ,
m
(according their multiplicities) in [s[ < r, f(s) ,= 0 for [s[ = r,
and f(0) ,= 0. Then
1
2
_
2
0
log [f(r exp(i))[ d = log
r
m
[f(0)[
[
1
. . .
m
[
.
Proof. First we assume that f(s) does not vanish for [s[ r; then Jensens
formula is an easy consequence of Cauchys theorem applied to log f(s) (more
precisely, it is the real part of the resulting formula).
Now assume that f(s) has zeros inside the circle [s[ = r. Then we rst
consider for any such zero s = the function g

(s) = s . Dene
G(s) =
g

(s)
r
2
s
.
It is easily seen that
1
2
_
2
0
log [G(r exp(i))[ d = log r,
50 Chapter 1 Classical L-functions
and
_
2
0
log [G(r exp(i))[ d =
_
2
0
log [g

(r exp(i))[ d 2 log r.
Hence,
1
2
_
2
0
log [g

(r exp(i))[ d = log r = log [g

(0)[ + log
r
[[
.
Now write f(s) = F(s)(s
1
) . . . (s
m
) with non-vanishing F(s) and
apply the already proven parts. Adding all resulting formulas together, yields
Jensens formula.
An entire function f(s) is said to be of nite order if there is a non-negative
real number such that
f(s) exp
_
[s[

_
as [s[ . The inmum over all numbers for which this estimate holds
is called the order of f. By Liouvilles theorem, the functions of order zero
are the polynomials.
Our next aim is to show that the zeros of an entire function f of nite order
cannot lie too dense; in fact, their location is related to the order of f.
Theorem 1.18. Let f be an entire function of nite order with zeros

1
,
2
, . . . arranged so that [
1
[ [
2
[ . . . and repeated according their
multiplicities. Then
j : [
j
[ r r
+
.
If > , then

j
=0
[
j
[

< .
Proof. Without loss of generality we may suppose that f(0) ,= 0. Further
we assume that f(s) does not vanish for [s[ = 3r (since the zeros of an entire
function form a discrete set this choice is indeed possible). Since log 3 > 1,
we deduce from Jensens formula 1.17 that

|
j
|r
1

|
j
|r
log
3r
[
j
[
= log [f(0)[ +
1
2
_
2
0
log [f(3r exp(i))[ d
r
+
.
The convergence of the series is a simple consequence. This proves the theo-
rem.
Section 1.6 The explicit formula 51
1.6.2. Hadamard products. Weierstrass proved that any non-zero en-
tire function can be factored into a product over its zeros (times an expo-
nential function). In the case of polynomials this is just another formulation
of the fundamental theorem of algebra (that any polynomial over C has a
root in C) and is known since Gauss rst proof in his doctorate. However,
a generic entire function has innitely many zeros and hence its so-called
Weierstrass product is innite and the analysis much more dicult. As part
of his theory of entire functions, Hadamard [72] obtained for entire functions
of nite order a more explicit form for Weierstrass products. For our purpose
it suces to consider only functions of order one.
Theorem 1.19. Let f(s) be an entire function of order 1 with zeros
0
= 0
with multiplicity m
0
and
1
,
2
, . . . arranged so that 0 < [
1
[ [
2
[ . . . and
repeated according their multiplicities. Then there are constants A, B such
that
f(s) = s
m
0
exp(A+ Bs)

j=1
_
1
s

j
_
exp
_
s

j
_
.
A proof of this theorem can be found in any textbook on the theory of
functions, e.g., Titchmarsh [199]. Therefore, we shall here give only a sketch
of
Proof. Without loss of generality we may assume that f(0) ,= 0. Since
(1 z) exp(z) = 1 z
2
+ higher terms,
the product
P(s) :=

j=1
_
1
s

j
_
exp
_
s

j
_
converges absolutely for any s, and so it represents an entire function. Writing
f(s) = P(s)G(s), it follows that G(s) is an entire function without zeros. Now
assume that G is of nite order. We shall show that then G(s) = exp(g(s)),
where g(s) is a polynomial of degree less than or equal to the order of G.
To see this consider the entire function h(s) := log G(s) log G(0). We
write s = r exp(i) with R and observe that
Re h(s) = log [G(s)[ r
+
.
There are real numbers a
n
, b
n
such that
h(s) =

n=0
(a
n
+ ib
n
)s
n
,
and therefore
Re h(s) =

n=0
(a
n
r
n
cos(n) b
n
r
n
sin(n)) .
52 Chapter 1 Classical L-functions
Hence, by Fourier theory,
[a
n
[r
n

_
2
0
[Re h(r exp(i))[ d.
It is easily seen that a
0
= 0 and
_
2
0
Re h(r exp(i)) d = 0.
Thus,
[a
n
[r
n

_
2
0
[Re h(r exp(i))[ + Re h(r exp(i)) d r
+
since [x[ +x is equal to 2x if x is positive or equal to zero otherwise. Sending
r implies a
n
= 0 for n > 1. This proves the claim on the function G.
We remark that the same argument can be applied if we would know
(1.52) G(s) exp(r
+
j
)
for a sequence r
j
tending to innity.
It remains to show that G is of order one. In view of the just given remark
we have to verify an estimate of the form (1.52) with = 1. For this purpose
we choose r
j
such that
[r
j
[
n
[[ > [
n
[
2
;
this choice can be realized since the measure of all intervals ([
n
[[
n
[
2
, [
n
[+
[
n
[
2
) is bounded by
2

n=1
[
n
[
2
,
which is nite by Theorem (since f(s) has order one). Now we write P =
P
1
P
2
P
3
, where the P
k
= P
k
(s) are those parts of the product P(z) according
to
in P
1
: [
n
[ <
1
2
r
j
,
in P
2
:
1
2
r
j
[
n
[ 2r
j
in P
3
: 2r
i
< [
n
[.
For the factors in P
1
we observe

_
1
s

n
_
exp
_
s

n
_

> exp
_

r
j
[
n
[
_
.
Taking into account Theorem 1.18,

|n|<
1
2
r
j
[
n
[
1

_
r
j
2
_

n=1
[
n
[
1
.
Thus it follows that [P
1
(s)[ > exp(r
1+
j
).
For any factor in P
2
we nd the lower bound r
3
j
. Since n(r
j
) r
1+
j
,
it follows that P
2
(s) exp(c
2
r
1+
j
) for some positive constant c
2
.
Section 1.6 The explicit formula 53
Finally, for any factor in P
3
we get the lower bound

_
1
s

n
_
exp
_
s

n
_

> exp
_
c
3
r
2
j
[
n
[
2
_
for some positive constant c
2
. Similar as in the case of P
1
we get [P
3
(s)[ >
exp(r
1+
j
).
Collecting all estimates for the P
k
s together, we deduce that
[G(s)[ < exp(r
1+
j
).
Now it is not dicult to see that G is of the form G(z) = exp(g(z)) with a
polynomial g of degree at most one. This proves the theorem.
1.6.3. Applications. Now we shall apply the results of the previous
subsections to the zeta-function and to Dirichlet L-function to primitive char-
acters. We start with zeta and dene
(s) =
1
2
s(s 1)

s
2

_
s
2
_
(s).
Notice that here we have removed the simple pole of the zeta-function at
s = 1; the factor s is included with respect to the symmetry of the functional
equation (1.9). We further observe that the zeros of (s) are exactly the
nontrivial zeros of the zeta-fucntion (by the presence of the Gamma-factor
of the functional equation (s) has a non-zero limit for s 2n). From the
functional equation (1.9) it follows that (s) is an entire function satisfying
(s) = (1 s).
Recall Stirlings formula
log (z) =
_
z
1
2
_
log z z +
1
2
log 2 +
_

0
[u] u +
1
2
u + z
du
=
_
z
1
2
_
log z z +
1
2
log 2 + O
_
[z[
1
_
, (1.53)
the latter asymptotic formula being valid uniformly in z with +
arg z . Furthermore, we have
(1.54) (s 1)(s) [s[
2
for
1
2
, which follows immediately from Theorem 1.3. This leads to the
estimate
(1.55) [(s)[ < exp(c[s[ log [s[)
for some positive constant c as [s[ . Now we consider s +. Since
then (s) 1, we obtain
(1.56) (s) > exp
_
s
4
log s
_
54 Chapter 1 Classical L-functions
as s . Taking into account the functional-equation for (s) we obtain
inf
_
: [(s)[ exp([s[

)
_
= 1.
It thus follows that (s) is an entire function of order 1. So we can apply
Hadamards product theorem 1.19 and obtain Riemanns conjectured product
representation (1.13).
Now we shall show that (1.56) implies that (s) has innitely many zeros,
resp. the existence of innitely many nontrivial zeros for (s). For this
purpose assume that
(1.57)

[[
1
is convergent. For any complex number z,
[(1 z) exp(z)[ < exp(2[z[).
Applying this with z =
s

, and taking into account the convergence of (1.57),


we deduce from Hadamards product theorem 1.19 that (s) exp(C[s[) for
some constant C > 0, as [s[ , in contradiction to (1.56). Thus the series
(1.57) diverges. By Corollary 1.18, it converges if we replace the exponent
1 by anything smaller. This information on the nontrivial zeros of (s) is
new: we did not make use of the (so far unproved Riemann-von Mangoldt
formula). We collect our observations in
Corollary 1.20. There exist constants A and B such that
(s) = exp(A+ Bs)

_
1
s

_
exp
_
s

_
,
where the product is taken over the nontrivial zeros of (s). Furthermore, the
series

[[
1
diverges while

[[
1
converges for any positive .
We may argue analogously for Diriclet L-functions. For any primitive char-
acter mod q, let
(s, ) =
_
q

_s+
2

_
s +
2
_
L(s, ),
where =
1
2
(1 (1)). In a similar manner as above we conclude that
(s, ) is an entire function of order 1 and also the other observations hold
in this case with the dierence that the functional equation takes the form
(1 s, ) = (s, ). Thus
Corollary 1.21. Let mod q be a primitive character. There exist constants
A

, B

such that
(s, ) = exp(A

+ B

s)

_
1
s

_
exp
_
s

_
,
Section 1.6 The explicit formula 55
where the product is taken over the nontrivial zeros

of L(s, ). The series

[
1
diverges while

[
1
converges for any positive .
1.6.4. The logarithmic derivative. In the proof of the prime number
theorem we have already worked with the logarithmic derivative. We note
that any zero of a meromorphic function is a simple pole of its logarithmic
derivative, independent of its multiplicity (this follows immediately from the
Laurent expansion). Our next aim is to deduce the partial fraction decompo-
sition of the Riemann zeta-function and Dirichlet L-functions, respectively.
We start again with zeta. Recall that we denote the nontrivial zeros by
= + i.
Theorem 1.22. We have

(s) =
_
_
_
O(1) if 2,

|t|<1
1
s
+ O(1 + log [s[) if 1 2, [t[ 1,
O(log [s[) if 1, [s + 2n[ >
1
4
, n N.
For the proof of this theorem we shall use two results which we have not
proved so far: rst, the functional equation (1.9) (which we will prove in
Chapter 2) and, second, a weak form of the Riemann-von Mangoldt formula
(1.12) (which is one of our aims in Chapter 3).
Proof. In the half-plane 2 we nd

(s)

n=2
log n
n
2
,
which leads to the estimate of the theorem.
In the region 1 we use the functional equation; it is not dicult to
see that (1.9) can be rewritten as
(1.58) (1 s) = 2
1s

s
cos
s
2
(s)(s).
Logarithmic dierentiation leads to
(1.59)

(1 s) = log 2

2
tan
s
2
+

(s) +

(s).
Now the estimate in question follows from the bound for 2 and Stirlings
formula (1.53).
In order to obtain the estimate for the region covering the critical strip
we need an easy consequence of the Riemann-von Mangoldt formula (1.12),
namely that
(1.60) N(T + 1) N(T 1) log T;
in fact, this can also be deduced directly from Jensens formula (which we
leave as an exercise to the interested reader).
56 Chapter 1 Classical L-functions
Now we continue with the nal estimate for the region 1 2. By
symmetry we may assume that t > 1. Dierentiation of the Hadamard
product representation from Corollary 1.20 leads to
(1.61)

(s) = B +

_
1
s
+
1

_
.
Moreover, we have

(s) =
1
s
+
1
s 1

1
2
log 2 +
1
2

_
s
2
_
+

(s).
By Stirlings formula (1.53),

(s) = log s + O
_
1
[s[
_
as [s[ and + < arg s < . Thus we get

(s) =

_
1
s
+
1

_
+ O(log t).
Using this formula with s = 2+it and subtracting the resulting formula from
the previous one, we arrive at
(1.62)

( + it) =

_
1
+ it

1
2 + it
_
+ O(log t).
For the rst we consider only the terms of the (2+it)
1
with [t[ 1.
Each of these terms is bounded and by (1.60) there exist O(log t) many of
them. Hence

|t|<1
1
[2 + it [
log t.
Next we investigate the contribution of the terms with [t [ 1. We have
1
+ it

1
2 + it
=
2
( + it )(2 + it )
(t )
2
.
Again with (1.60) we get

|t|1
_
1
+ it

1
2 + it
_
=

mZ,m=1,0

t+mt+m+1
_
1
+ it

1
2 + it
_

mZ,m=1,0
log [t + m[
m
2
log t.
Substituting this and the previous estimate in (1.62) leads to the formula of
the theorem. The theorem is proved.
Section 1.6 The explicit formula 57
The same method can be applied to Dirichlet L-functions to primitive char-
acters. Some things are more simple here; e.g., there is no pole at s = 1.
However, other parts need special attention. The trivial zeros are located at
s = 2n, n N
0
(where = 1 if (1) = 1, and = 0 otherwise). In fact,
if (1) = 1, then L(s, ) has a trivial zero at s = 0 which has to be treated
in a similar manner as the pole of (s). Moreover, we have to make use of
the corresponding functional equation (1.25) for Dirichlet L-functions and
the analogue to the corresponding weak version of the Riemann-von Man-
goldt formula. If mod q is a primitive character and N(T; ) counts the
number of nontrivial zeros

+i

of L(s, ) with [

[ T (according
multiplicities), then
(1.63) N(T; ) =
T

log
qT
2e
+ O(log qT);
note that here also zeros from the lower half-plane are counted in lack of a
symmetry with respect to the real axis in case of non-real characters. For the
argument below the following consequence of (1.63) is sucient:
N(T + 1; ) N(T; ) log T,
where the implicit constant depends on q.
Then the analogue of Theorem 1.23 for Dirichlet L-functions takes the form:
Theorem 1.23. Let mod q be a primitive character. We have
L

L
(s, ) =
_

_
O(1) if 2,

|t|<1
1
s
+ O(log(q(1 +[s[))) if
1
2
2,
O(log(q[s[)) if
1
2
, and
[s + 2n [ >
1
4
, n N,
where in the case of (1) = 1 the second estimate holds only for [s[ >
1
2
;
for [s[
1
2
, we have in this case
L

L
(s, )
1
s
=

|t|<1
1
s

+ O(log q).
In view of applications we have here incorporated the dependency of the
error term on the character. For this aim we have to be a bit careful: for
example, we cannot bound B

by an absolute constant as in the case of the


zeta-function: B 1. We leave the details to the interested reader.
1.6.5. Proof of the explicit formula. Now we are going to prove an-
other of Riemanns conjectures, the explicit formula. We prefer to work with
(x), resp. the slightly modied function

0
(x) =
_
(x) if x , Z,

n<x
(n) +
1
2
(x) if x Z.
58 Chapter 1 Classical L-functions
This modication is made with respect to obtain an exact formula also in
the case of integral x. As a matter of fact, we can replace Perrons formula
(1.36) by

0
(x) =
1
2i
_
c+i
ci

(s)
x
s
s
ds. (1.64)
We observe that this formula is valid for any x R since we have added the
term
1
2
(x) for x being an integer (with respect to the contribution of the
term for y =
x
n
= 1 in Lemma 1.12).
Moving now the path of integration to the left, we nd that the latter
expression is equal to the corresponding sum of residues, that are the residues
of the integrand at the pole of (s) at s = 1, at the zeros of (s), and at the
pole of the integrand at s = 0. We have already identied the main term as
being the residue at s = 1. Each zero gives the contribution
Res
s=
_

(s)
x
s
s
_
=
x

.
In particular, for the trivial zeros = 2n with n N we get the contribution

n=1
x
2n
2n
=
1
2
log
_
1
1
x
2
_
.
The simple pole at s = 0 leads to
(1.65) Res
s=0
_

(s)
x
s
s
_
= lim
s0
s

(s)
x
s
s
=

(0) = log(2)
(the computation of this constant is left to the reader as an easy exercise).
This leads to the exact explicit formula (1.51)
(1.66)
0
(x) = x


1
2
log
_
1
1
x
2
_
log(2),
being valid for any positive real x, which is slightly stronger than (1.51) and
equivalent to Riemanns version (1.14). However, for a rigorous proof we
have to prove that the integrals over the contour vanish. We include this
in a study of a truncated version of the explicit formula. Here we shall cut
the integral (1.64) at t = T; of course, in this setting we will have error
terms, but the resulting version of the explicit formula is rather convenient
for applications.
Theorem 1.24. For x > 2,

0
(x) = x

||T
x


1
2
log
_
1
1
x
2
_
log(2) + R(x, T)
with
R(x, T)
x
T
(log(xT))
2
+ min1, x/(Tx) log x,
Section 1.6 The explicit formula 59
where x denotes the minimum of [x p
k
[ for x ,= p
k
where p is prime and
k N.
Notice that for x = p
k
we have x = 0 and the minimum appearing in the
error term takes the value 1. Furthermore, we observe that R(x, T) vanishes
as T and thus the theorem implies (1.66). The convergence will turn
out to be uniform in any closed interval which does not contain a prime power
(for which
0
(x) is discontinuous).
Proof. We put c = 1+
1
log x
in (1.37); then x
c
= ex. Since c is a function of x,
we have to be a bit more careful than in the proof of Theorem 1.9; however,
much of the reasoning follows just the same way. First of all we nd

0
(x) +
1
2i
_
c+iT
ciT

(s)
x
s
s
ds

x=nN
(n)
_
x
n
_
c
min1, (T[ log x/n[)
1
+ c
([x])
T
,
where the last term can be omitted if x , N. Similar as before we nd that
the contribution of the terms with n , [
1
2
x, 2x] is O(
xlog x
T
). For n (
1
2
x, 2x)
we dene x
1
to be the maximal prime power p
k
< x. If x
1

1
2
x, then
(n) = 0 for all n under consideration and we are done. If x
1
>
1
2
x, we
consider the term with n = x
1
separately. Since then
log
x
x
1

x x
1
x

x
x
,
we nd
(n)
_
x
n
_
c
min1, (T[ log x/x
1
[)
1
min1, x/(Tx) log x.
The other terms can be estimated as before. Hence we obtain

0
(x) =
1
2i
_
c+iT
ciT

(s)
x
s
s
ds +
+O
_
x(log x)
2
T
+ min1, x/(Tx) log x
_
. (1.67)
In the next step we apply the calculus of residues to the contour integral
taken over the rectangular path with corners U iT and c iT, where U
is a positive odd integer (to have some distance from the trivial zeros of zeta
which are simple poles of the integrand). Here we may estimate the integral
over the vertical segment just as we did before; we obtain
(1.68)
_
U+iT
UiT

(s)
x
s
s
ds
_
T
T

(U + it)

x
U
[ U + it[
dt
log U
Ux
U
T.
However, the vertical integrals need special attention since here the segments
may run through a neighbourhood of a trivial zero. In view of (1.60) (resp.
60 Chapter 1 Classical L-functions
the Riemann-von Mangoldt formula) there are at most O(log n) trivial zeros
= +i with n < n+1. Hence there exists a T = T
n
(n, n+1] such
that
(1.69) [T [
1
log T
for all zeros . For s = +iT we deduce from Theorem 1.22 and once more
(1.60) that

(s) =

|T|<1
1
s
+ O(log T) (log T)
2
.
This estimate together with the corresponding one of Theorem 1.60 for
1 leads to
_
ciT
UiT

(s)
x
s
s
ds
_
c
1

x
s
s

(log T)
2
d +
_
1
U

x
s
s

log [s[ d

x
log x
(log T)
2
T
.
One observes that all these estimates are uniform in U. Moreover we note
that (1.68) tends to zero as U . Hence we arrive at the explicit formula
as given in the theorem, valid for those T which satisfy (1.69); however, the
latter condition can be relaxed. By construction, for any T > 1 there is a
T
n
which has distance less than 1 from T and satises (1.69). Obviously,
substituting T has no inuence on the size of the error term R(x, T) and for
the sum over the nontrivial zeros we observe that

||[T,Tn]
x


xlog T
T
by (1.60). This can be absorbed in R(x, T). This proves the theorem.
In a rather similar way one can prove explicit formulae for Dirichlet L-
functions. If mod q is a primitive character and

denotes the nontrivial


zeros of L(s, ), then we nd analogously

0
(x, ) =

L
(0, ) +

n=1
x
12n
2n 1
if (1) = 1, and

0
(x, ) =

log x lim
s0
_
L

L
(s, )
1
s
_
+

n=1
x
2n
2n
otherwise, i.e., if (1) = +1; here
0
(x, ) is the function (x, ) =

nx
(n)(n) modied in just the same way as we did when we switched
from (x) to
0
(x). The origin of most of the appearing terms is clear; how-
ever, we should have a brief look on the main dierence. The logarithmic
Section 1.6 The explicit formula 61
derivative of L(s, ) at s = 0 is regular if (1) = 1 and it has a simple
pole if (1) = +1 (by the trivial zero of L(s, )). As in the proof of Theo-
rem 1.23, these cases have to be considered separately and the result are the
slightly diering explicit formulae above.
Again it is desirable to have truncated versions which are uniform in the
modulus q. This is a rather dicult task, in particular, since we cannot
exclude nontrivial zeros

near s = 0 or 1. One can show that such a so-


called exceptional (or Siegel-) zeros can only occur if is a real character and
it is itself real. One can also show that there cannot be too many of these
zeros; however, we do not want to go into the details and simply state the
following result without proof:
Theorem 1.25. Let mod q be a non-principal character and assume that
2 < T x. Then
(x, ) =
x

||T
x

+ R(x, T, q),
where
R(x, T, q)
x
T
(log(qx))
2
+ x
1
4
log x.
The term
x

is to be omitted unless is a real character for which L(s, )


vanishes at s =

satisfying the estimate

> 1
c
log q
,
where c is a positive absolute; if the zero

exists, the sum has to be taken


over all nontrivial zeros in the given range dierent from

and 1

.
Here we have also included the case of non-primitive characters and
0
(x, )
is replaced by (x, ) which is more useful for applications. For a proof we
refer once more to Davenport [44] and Prachar [172].
1.6.6. Improvement on the error term in the prime number the-
orem. The explicit formula allows a remarkable improvement on the error
term in the prime number theorem. In the sequel we focus on the Riemann
zeta-function; for the more general case of Dirichlet L-functions (where one
wants to have uniformity in the module) we refer to Prachar [172].
First of all, we observe that our deeper knowledge on the analytic behaviour
of the zeta-function implies a larger zero-free region inside the critical strip.
Lemma 1.26. There exists a constant c > 0 such that for any nontrivial
zero = + i
< 1 C min1, (log [[)
1
.
62 Chapter 1 Classical L-functions
Proof. We shall use the same ideas as in the proof of Lemma 1.11 but now
we incorporate the approximation for (s) from Theorem 1.22, i.e.,

(s) =

|t|<1
1
s
+ O(log [s[).
For the rst we suppose that > 1 and without loss of generality let t > 0.
We observe that the real part of the summands is positive for > 1. Hence,
we deduce that
Re
_

(s)
_
< Re
1
s
+ c log [s[
for any nontrivial zero = + i with [t [ < 1, where c > 0 is a suitable
constant; of course, here we can also delete the -term on the right (since its
contribution is negative) and obtain
Re
_

(s)
_
< c log [s[.
Now recall (1.30). Using the latter estimate with t = it follows (in just the
same way as in the proofs of Lemma 1.10 and 1.11) that
0
17
1

24

+ c log(t + 2),
resp.
< 1 +
1
log(t + 2)

4
(3 + c) log(t + 2)
by putting = 1 +

log(t+2)
with > 0. This proves the lemma.
We may use the just proved lemma in order to obtain the following im-
provement on the prime number theorem 1.9:
Theorem 1.27. There exists an absolute positive constant C such that for
suciently large x
(x) = Li (x) + O
_
xexp
_
C(log x)
1
2
__
.
This is for many application indeed a valuable improvement, however, it is
still weaker than (1.39) which can be obtained by incorporating the so far
best zero-free region for the zeta-function.
Proof. We consider the sum

appearing in the explicit formula. For


each term with [[ T we nd by the previous lemma that
x

xexp
_
C
log x
log T
_
.
Section 1.6 The explicit formula 63
Furthermore, we have

||T
1
[[

0<T
1

.
To bound this sum we apply partial summation in conjunction with Riemann-
von Mangoldt formula (1.12) in the weak form
N(T) T log T
(recall that (1.12) was still not proved in these notes but that we will return
to this problem in Chapter 3). Then we nd that the sum in question is
=
N(T)
T
+
_
T
0
N(t)
t
2
dt (log T)
2
.
Hence,

||T
x

x(log T)
2
exp
_
C
log x
log T
_
.
Without loss of generality we may suppose that x is a positive integer. Then
we deduce from the explicit formula, Theorem 1.24, that
(x) x
x(log xT)
2
T
+ x(log T)
2
exp
_
C
log x
log T
_
.
Taking the balance (log T)
2
= log x and T = exp((log x)
1
2
) respectively, we
may deduce the bound of Theorem 1.27 by partial summation.
1.6.7. Weils explicit formula. The explicit formula combines the re-
markable fact that the zeta-function (resp. any Dirichlet L-function) can be
written both as an Euler product over the primes and as a Hadamard product
over the trivial zeros. Weil [210] proved a rather general extension of this
reciprocity between primes and zeros. In order to state his result we have to
introduce some new notions.
Let f be a measurable function on R. We say that f is of type (

,
r
),
where

<
r
are real numbers, if x f(x)[x[
1
is integrable for
(

,
r
). In this case we dene the Mellin transform of f by
M(f, s) =
_

f(x)[x[
s1
dx
for (

,
r
). Under certain circumstances there is an inverse transform
f(x) =
1
2i
_
+i
i
M(f, s)[x[
s
ds.
We have already seen some examples of such pairs of transforms in the proof
of the prime number theorem (and therefore it is not surprising to nd them
here once again).
Now Weils explicit formula takes the form:
64 Chapter 1 Classical L-functions
Theorem 1.28. Let f be a function of type (

,
r
), where

<
1
2
and

r
>
1
2
. Suppose that there exist c, > 0 such that
[M(f, s)[ < c(1 +[s[)
1
for all s [

,
r
], f is of bounded total variation, and that f(x) = 0 if x < 0.
Dene

(f) = lim
N
__

0
f(x)F
N
(x)
dx
x
f(1) log
N

_
,
where
F
N
(x) =
_

_
x

1
2
1x
2N
|xx
1
|
if 0 < x < 1,
0 if x = 1,
x
+
1
2
1x
2N
|xx
1
|
if x > 1.
The limit in this denition exists and one has

(f) +

p
log p

0=kZ
f(p
k
)p

|k|
2
= M
_
f,
1
2
_
+ M
_
f,
1
2
_
+

M
_
f,
1
2
_
;
all series are absolutely convergent.
Notice that on the left-hand side the summation is taken over all non-
equivalent valuations of Q: the p-adic non-archimedean valuations plus the
archimedean absolute value (indicated by the index ). For the rather
lengthy proof we refer to Weil [210] and Patterson [164], respectively.
Exercise 46. Fill the gaps in the proofs of Lemma 1.17 and Theorem 1.18.
Exercise 47. i) Prove the Hadamard product representation for the reciprocal of
the Gamma-function:
1
(s)
= s exp(Cs)

n=1
_
1 +
s
n
_
exp
_

s
n
_
,
where C is the Euler-Mascheroni constant.
ii) What are the residues?
iii) Derive an analogous formula for the sin-function.
Exercise 48. Prove formula (1.55).
Exercise 49. * i) Show for the constants in Corollary 1.20 that
A =

(0) = log 2 and B =

1
=
C
2
1 +
1
2
log 4,
where C is the Euler-Mascheroni constant and the summation in the B-dening
series is such that the terms and 1 are added together. Deduce that there are
Section 1.6 The explicit formula 65
no zeros = + i with [[ 6.
Hint: functional equation plus (1.61)
ii) Prove (1.65).
Exercise 50. i) Deduce the functional equation for the zeta-function in the form
(1.58) from (1.9).
Hint: use basic facts and identities from the theory of the Gamma-function.
ii) Show that any zero of

(s) on the critical line is also a zero of (s)


It is expected that

(s) does not vanish on the critical line; more precisely, that
all zeros of (s) are simple. Speiser [186] has shown that the Riemann hypothesis
is equivalent to the non-vanishing of

(s) in 0 < <


1
2
; if also

(
1
2
+it), then, by
ii), all zeros are simple!
Exercise 51. * i) Verify all sketched estimates in the proof of Theorem 1.22.
ii) Show that, for 5, [t[ 1,
[(s)[ [t[
11
2
.
Hint: use the functional equation and Stirlings formula.
iii) Prove (1.60) without using the Riemann-von Mangoldt formula and deduce the
estimate N(T) T log T.
Hint: use Jensens formula together with ii).
iv) Prove Theorem 1.23 along the lines of the proof of Theorem 1.22.
Exercise 52. * Prove Theorem 1.25.
Hint: for inspiration one may have a look into [172].
Exercise 53. * i) Deduce Riemanns explicit formula (1.14) from Theorem 1.24.
ii) Deduce Theorem 1.24 from Weils explicit formula.
CHAPTER 2
Zero-distribution of the Riemann zeta-function
In this chapter we shall have a closer look at the zeros of the Riemann zeta-
function inside the critical strip. In view of the unsolved Riemann hypothesis
they are the most important objects but also the most dicult to deal with.
We shall show that there are innitely many zeros on the critical line and
that there cannot be too many nontrivial zeros o the critical line; here we
mean that the proportion of the set of possible violations of the Riemann
hypothesis is zero. Most of the presented methods can be easily generalized
to other L-functions (e.g., Dirichlet L-functions).
2.1. The Riemann hypothesis
The famous Riemann hypothesis states that all nontrivial zeros lie on the
critical line =
1
2
. We can rewrite this equivalently as
Riemanns hypothesis. (s) ,= 0 for >
1
2
.
There has been a lot of speculation how Riemann was led to this conjec-
ture. One of the reasons might have been his own computations (which are
preserved among his unpublished manuscripts in the library of G ottingen
University). Clearly, in view of the symmetry dictated by the functional
equation the scenario that all zeros lie on the vertical line passing through
the point of symmetry s =
1
2
is the most beautiful one. But we will never
know what Riemanns motivation was.
Many computations were done to nd a counterexample to the Riemann
hypothesis. Van de Lune, te Riele & Winter [139] localized the rst
1 500 000 001 zeros, all lying without exception on the critical line. More-
over all so far localized nontrivial zeros turned out to be simple! Besides
Riemanns hypothesis we have the
Essential simplicity hypothesis. All (or at least almost all) zeros of (s)
are simple.
2.1.1. The error term in the prime number theorem. The next
result highlights the intimate relation between the zeros of the zeta-function
and prime number distribution.
66
Section 2.1 The Riemann hypothesis 67
Theorem 2.1. For xed [
1
2
, 1),
(x) x x
+
(s) ,= 0 for > .
Our main tool for its proof is the explicit formula from the previous section
which puts the prime numbers in an explicit relation to the nontrivial zeros
of the zeta-function.
Proof. Recall (1.28). For > 1 we have

(s) =
s
s 1
+ s
_

1
(u) u
u
s+1
du.
If (x) x x
+
, then the integral above converges for > , giving an
analytic continuation for

(s)
1
s 1
to the half-plane > , and, in particular, (s) does not vanish there.
Conversely, if all nontrivial zeros = + i satisfy , then it follows
from the explicit formula, Theorem 1.24, that
(x) x x

||T
1
[[
+
x
T
(log(xT))
2
. (2.1)
In view of (1.60) we get

||T
1
[[
= 2
[T]

m=1

m<m+1
1


[T]+1

m=1
log m
m
(log T)
2
.
Substituting this in (2.1) leads to
(x) x x

(log T)
2
+
x
T
(log(xT))
2
.
Now the choice T = x
1
nishes the proof of this implication.
Taking into account Theorem 2.1, we nd by partial summation that
(x) Li (x) x
+
(s) ,= 0 for > .
Now the impact of the Riemann hypothesis on the prime number distribution
becomes visible. If the Riemann hypothesis is true, we may take =
1
2
in
Theorem 2.1 and the resulting estimate for the error term in the prime number
theorem is
(x) = x + O
_
x
1
2
+
_
.
A slight stronger bound was rst obtained by von Koch [118, 119] under
assumption of the Riemann hypothesis (actually, he replaced x

by powers of
log x).
68 Chapter 2 Zero-distribution
With regard to known zeros of (s) on the critical line it turns out that an
error term with <
1
2
is impossible. In fact one can show that
(2.2) (x) Li (x) =

_
x
1
2
log log log x
log x
_
(see, e.g., Ingham [97]). We have to explain the -notation: given two
functions f(x) and g(x), where g(x) is positive for suciently large x, we
write f(x) =
+
(g(x)) (resp. f(x) =

(g(x))) if
[f(x
n
)[ cg(x
n
) (resp. [f(x
n
)[ cg(x
n
))
holds with a positive constant c for some sequence x
n
which tends to innity.
Thus, (2.2) shows that (x) Li (x) changes its sign innitely often and that
an error term O(x

) with <
1
2
is impossible. In some sense, Riemanns
hypothesis states that the prime numbers are as uniformly distributed as
possible!
Maybe one of the most given arguments in favour for the truth of Riemanns
hypothesis is the function eld analogue. Davenport and, in particular, Hasse
proved that the so-called Riemann hypothesis for elliptic curves is true, and
later Weil proved the general case of abelian varieties. It is far beyond the
scope of our notes to give an adequate introduction to this topic (nevertheless,
in the following chapter we will briey explain the meaning of Hasses work on
elliptic curves), especially since the analogue of the zeta-function for abelian
varieties is a rational function and so its value-distribution is a priori rather
dierent than the one of the transcendental function (s). On the other side,
the parallel world of function elds has often been proved to be a signpost
for challenges.
In the following section we present some further heuristics in favour for the
Riemann hypothesis.
2.1.2. Denjoys probabilistic argument for Riemanns hypothe-
sis. Recall the denition of Mobius -function: we write (1) = 1, (n) = 0
if n has a quadratic divisor, and (n) = (1)
r
if n is the product of r distinct
primes. It is easily seen that is multiplicative and appears as coecients of
the Dirichlet series representation of the reciprocal of the zeta-function: for
> 1,
1
(s)
=

p
_
1
1
p
s
_
=

n=1
(n)
n
s
.
Riemanns hypothesis is equivalent to
(2.3) M(x) :=

nx
(n) x
1
2
+
.
This is related to the estimates of Theorem 2.1 (for a proof see, for example,
Titchmarsh [200], 14.25).
Section 2.1 The Riemann hypothesis 69
Denjoy [46] argued as follows. Assume that X
n
is a sequence of random
variables with distribution
P(X
n
= +1) = P(X
n
= 1) =
1
2
.
Dene
S
0
= 0 and S
n
=
n

j=1
X
j
,
then S
n
is a symmetrical random walk in Z
2
with starting point at 0. A
simple application of Chebyshevs inequality yields, for any positive c,
P[S
n
[ cn
1
2

1
2c
2
,
which shows that large values for S
n
are rare events. By the theorem of
Moivre-Laplace this can be made more precise. It follows that
lim
n
P
_
[S
n
[ < cn
1
2
_
=
1

2
_
c
c
exp
_

x
2
2
_
dx.
Since the right hand side above tends to 1 as c , we obtain
lim
n
P
_
[S
n
[ n
1
2
+
_
= 1
for every > 0. If the values of the -function would behave like random
variables, then Riemanns hypothesis would hold with probability one! The
law of the iterated logarithm would even give the stronger estimate
lim
n
P
_
[S
n
[ (nlog log n)
1
2
_
= 1,
which suggests for M(x) the upper bound (xlog log x)
1
2
. This estimate is
pretty close to the so-called weak Mertens hypothesis which states
_
X
1
_
M(x)
x
_
2
dx log X.
Note that this bound implies the Riemann hypothesis and the essential sim-
plicity hypothesis. On the contrary, Odlyzko & te Riele [162] disproved the
original Mertens hypothesis [145],
[M(x)[ < x
1
2
,
by showing
liminf
x
M(x)
x
1
2
< 1.009 and limsup
x
M(x)
x
1
2
> 1.06;
for more details see Titchmarsh [200], 14.
70 Chapter 2 Zero-distribution
2.1.3. Approaches towards RH and a substitute. There are some
interesting recent approaches to be mentioned. The rst one is an output
of Connes theory of non-commutative geometry. Connes [34] obtained a
so-called trace formula in non-commutative geometry which has remarkable
similarity with Weils explicit formula, Theorem 1.28. Assuming the Riemann
hypothesis, he shows that this is indeed the explicit formula in disguise and
so this gives a natural spectral interpretation of the nontrivial zeros. This
approach restored some hope to an old idea of Hilbert and Polya that the
Riemann hypothesis follows from the existence of a self-adjoint Hermitian
operator whose spectrum of eigenvalues corresponds to the set of nontrivial
zeros of the zeta-function.
Another approach is from Bombieri [24]. It is based on Weils explicit
formula and his positivity criterion for the Riemann hypothesis. The latter
can be rewritten in terms of the positivity of a certain linear functional; then
it is shown that if the Riemann hypothesis is false, then the extremals, in
various relevant Hilbert spaces, would have distinctly unusual properties.
So far we sketched much of the theory for Dirichlet L-functions; however,
the exceptional zeros make their analysis much more complicated and we do
not want to go further into the details but give a glimpse on the impact of their
zero-distribution on the prime number distribution in arithmetic progressions.
All analogues of Riemanns hypothesis for the whole class of Dirichlet L-
functions are summed up in the so-called
Generalized Riemann hypothesis. Neither (s) nor any L(s, ) has a
zero in the half-plane Re s >
1
2
.
Under assumption of this conjecture one has
(x; a mod q) =
1
(q)
Li (x) + O
_
x
1
2
log(qx)
_
(2.4)
for x 2, q 1, and a coprime with q, the implicit constant being absolute.
As long as we do not have a proof of Riemanns hypothesis in many in-
stances we are often forced to prove conditional results. However, sometimes
one can also nd an appropriate way to circumvent the assumption of RH.
We conclude with a remarkable substitute of the Riemann hypothesis, the
celebrated theorem of Bombieri-Vinogradov due to Bombieri [23] and Vino-
gradov [205] (independently, with a slightly weaker range for Q):
Theorem 2.2. For any A 1,

qQ
max
a mod q
(a,q)=1
max
yx

(y; a mod q)
1
(q)
Li(y)

x
(log x)
A
+ Qx
1
2
(log Qx)
6
.
Section 2.1 The Riemann hypothesis 71
This shows that the error term in the prime number theorem for arithmetic
progressions is, on average over q x
1
2
(log x)
A7
, of comparable size as
predicted by the Riemann hypothesis (see (2.4)).
Exercise 54. Prove that, for xed [
1
2
, 1),
(x) Li (x) x
+
(s) ,= 0 for > .
Show that if the Riemann hypothesis is true, then
(x) Li (x) x
1
2
log x.
Exercise 55. i) Prove that Riemanns hypothesis is equivalent to the estimate
(2.3).
ii) Verify all probabilistic statements in Section 2.1.2.
iii) Show that the estimate M(x) x
1
2
implies both the Riemann hypothesis and
that all zeros of (s) are simple.
Hint: Show that then
1
(s)
= s
_

1
M(x)
x
s+1
dx
holds for >
1
2
and deduce the estimate
1
[(s)[
c
[s[

1
2
with some positive constant c.
The essential simplicity conjecture (almost all zeros of (s) are simple) has arith-
metical consequences. Cramer [41] showed, assuming the Riemann hypothesis,
1
log X
_
X
1
_
(x) x
x
_
2
dx

m()

2
,
where the sum is taken over distinct nontrivial zeros and m() denotes their
multiplicity. The right-hand side is minimal if all the zeros are simple.
Exercise 56. Give an unconditional estimate for the left-hand side. Prove that
the series on the right-hand side converges.
Exercise 57. * Assuming the generalized Riemann hypothesis, prove the asymp-
totic formula (2.4).
The following sections of this chapter specialize on the Riemann zeta-
function and its zero-distribution; however, many of the results can be gen-
eralized to other L-functions, e.g., Dirichlet L-functions (unconditionally) or
Dedekind zeta-functions (at least conditionally); for a more general approach
we refer the interested reader to Iwaniec & Kowalski [101] and Lekkerkerker
[129]. In some places we are rather brief since an adequate presentation of
72 Chapter 2 Zero-distribution
all relevant results would be far beyond these notes. Often we also leave the
stage of the classical theory.
2.2. The approximate functional equation
The aim of this section is to prove an approximation of the zeta-function
inside the critical strip:
Theorem 2.3. We have, uniformly for
0
> 0, [t[ 4x,
(s) =

nx
1
n
s
+
x
1s
s 1
+ O
_
x

_
.
This approximation is a renement of (1.11) and will turn out to be a rather
useful tool in later applications.
2.2.1. Eulers summation formula. Let f(u) be any function with
continuous derivative on the interval [a, b]. By partial summation we get

a<nb
f(n) = ([b] [a])f(b)
_
b
a
([u] [a])f

(u) du
= [b]f(b) [a]f(a)
_
b
a
[u]f

(u) du,
where [u] = maxz Z : z u. Obviously,

_
b
a
[u]f

(u) du =
_
b
a
_
u [u]
1
2
_
f

(u) du
_
b
a
_
u
1
2
_
f

(u) du.
Applying partial integration to the last integral on the right-hand side, we
deduce Eulers summation formula:
Lemma 2.4. Assume that f : [a, b] R has a continuous derivative. Then

a<nb
f(n) =
_
b
a
f(u) du +
_
b
a
_
u [u]
1
2
_
f

(u) du
+
_
a [a]
1
2
_
f(a)
_
b [b]
1
2
_
f(b).
Why is this interesting? Imagine we are interested in describing the diver-
gence of the harmonic series in a quantitative way. In such questions it is
often an advantage to work with integrals rather than sums. An easy ap-
plication of the previous lemma yields the asymptotic formula (1.4) which
describes very precisely the rate of divergence of the harmonic series. How-
ever, we are heading for something more dicult. For this purpose we rst
replace in Eulers summation formula the function u [u]
1
2
by its Fourier
series expansion.
Section 2.2 The approximate functional equation 73
Lemma 2.5. For u R Z,

u
1
2

|m|M
m=0
exp(2imu)
2im

1
2M(u [u])
,
and, for u R,

m=0
exp(2imu)
2im
=
_
u [u]
1
2
if u , Z,
0 if u Z,
where the terms with m have to be added together; the partial sums are
uniformly bounded in u and M.
Proof. By symmetry and periodicity it suces to consider only the case
0 < u
1
2
. Since
_ 1
2
u
exp(2imx) dx =
(1)
m+1
+ exp(2imu)
2im
for 0 ,= m Z, we obtain

|m|M
m=0
exp(2imu)
2im
u +
1
2
=
_ 1
2
u

|m|M
exp(2imx) dx
=
_ 1
2
u
sin((2M + 1)x)
sin(x)
dx. (2.5)
By the mean-value theorem there exists (u,
1
2
) such that the latter integral
is equal to
_

u
sin((2M + 1)x)
sin(u)
dx.
This immediately implies both formulas of the lemma. It remains to show
that the partial sums of the Fourier series are uniformly bounded in u and
M. Substituting y = (2M + 1)x in (2.5), we get
_ 1
2
u
sin((2M + 1)x)
sin(x)
dx =
_ 1
2
u
sin((2M + 1)x)
x
dx +
+
_ 1
2
u
sin((2M + 1)x)
_
1
sin(x)

1
x
_
dx

_

0
sin(y)
y
dy +
_ 1
2
0

1
sin(x)

1
x

dx
with an implicit constant not depending on u and M; obviously both integrals
exist, which gives the uniform boundedness.
74 Chapter 2 Zero-distribution
2.2.2. Van der Corputs summation formula. In 1921, van der Cor-
put [40] invented a new and rather ecient technique to estimate exponential
sums.
Theorem 2.6. For any given > 0, there exists a positive constant C =
C(), depending only on , with the following property: assume that f :
[a, b] R is a function with continuous derivative, g : [a, b] [0, ) is a
dierentiable function, and that f

, g and [g

[ are all monotonically decreasing.


Then

a<nb
g(n) exp(2if(n))
=

(a)<m<f

(b)+
_
b
a
g(u) exp(2i(f(u) mu)) du +c,
where
[c[ C() ([g

(a) + g(a) log([f

(a)[ +[f

(b)[ + 2)) .
Van der Corputs summation formula looks very technical but the underlying
idea is rather simple. The integral
_
b
a
g(u) exp(2i(f(u) mu)) du
is (up to a constant factor) the Fourier transform of g(u) exp(2if(u)) at
u = m. Therefore, one can interpret Theorem 2.6 as an approximate version
of Poissons summation formula (a topic we will return to in the following
chapter).
Before we can give the proof we shall give the following estimate for expo-
nential integrals.
Lemma 2.7. Assume that F : [a, b] R has a continuous non-vanishing
derivative and that G : [a, b] R is continuous. If
G
F

is monotonic on [a, b],


then

_
b
a
G(u) exp(iF(u)) du

G
F

(a)

+ 4

G
F

(b)

.
Proof. First, we assume that F

(u) > 0 for a u b. Since (F


1
(v))

=
F

(F
1
(v))
1
, substituting u = F
1
(v) leads to
_
b
a
G(u) exp(iF(u)) du =
_
F(b)
F(a)
G(F
1
(v))
F

(F
1
(v))
exp(iv) dv.
Section 2.2 The approximate functional equation 75
Application of the mean-value theorem gives, in case of a monotonically in-
creasing
G
F

,
Re
_
_
F(b)
F(a)
G(F
1
(v))
F

(F
1
(v))
exp(iv) dv
_
=
G
F

(F(a))
_

F(a)
cos v dv +
G
F

(F(b))
_
F(b)

cos v dv
with some (a, b). The same argument applies to the imaginary part. The
case F

(u) < 0 can be treated analogously. This gives the desired estimate.
The lemma is proved.
Now we are in the position to give the
Proof of Theorem 2.6. Using Eulers summation formula with F(u) =
g(u) exp(2if(u)) and the Fourier series expansion of Lemma 2.5, we get

a<nb
g(n) exp(2if(n))
=
_
b
a
g(u) exp(2if(u)) du + O(g(a))
+
_
b
a

m=0
exp(2imu)
2im
d
du
(g(u) exp(2if(u))) du.
Since the series on the right-hand side converges uniformly on each compact
subset which is free of integers, and since its partial sums are uniformly
bounded, we may interchange summation and integration. This yields

a<nb
g(n) exp(2if(n)) =
_
b
a
g(u) exp(2if(u)) du
+

m=0
1
m
_
J
1
(m) +
1
2i
J
2
(m)
_
+ O(g(a),
where
J
1
(m) :=
_
b
a
f

(u)g(u) exp(2i(f(u) mu)) du,


J
2
(m) :=
_
b
a
g

(u) exp(2i(f(u) mu)) du.


76 Chapter 2 Zero-distribution
Partial integration gives
J
1
(m) =
_
exp(2i(f(u) mu))g(u)
2i
_
b
u=a

_
b
a
exp(2if(u))
2i
d
du
g(u) exp(2imu) du,
= O(g(a))
1
2i
J
2
(m) + m
_
b
a
g(u) exp(2i(f(u) mu)) du.
Thus,

(a)<m<f

(b)+
m=0
1
m
_
J
1
(m) +
1
2i
J
2
(m)
_
=

(a)<m<f

(b)+
m=0
_
b
a
g(u) exp(2i(f(u) hu)) du
+O
_
_
_

(a)<m<f

(b)+
m=0
g(a)
[m[
_
_
_
.
Now assume that m > f

(a) + and f

(b) > 0. Then f

(u) > 0 for


a u b. Using Lemma 2.7 with F(u) = 2(f(u) mu) and G = gf

, we
nd
J
1
(m)

g(a)f

(a)
f

(a) m

.
Hence,

m>f

(a)+
m=0

J
1
(m)
m

g(a)

0<m2|f

(a)|
1
m
+ g(a)

m>|f

(a)|
[f

(a)[
m
2
.
The contribution arising from m < f

(b) can be treated similarly. This


gives

m[f

(b),f

(a)+]
m=0

J
1
(m)
m

g(a) log([f

(a)[ +[f

(b)[ + 2).
Next assume m > f

(a) + and m ,= 0. Then, by the mean-value theorem,


Re J
2
(m) =
_
b
a
[g

(u)[ cos 2(f(u) mu) du


= g

(a)
_

a
cos 2(f(u) mu) du
Section 2.2 The approximate functional equation 77
with some (a, b). Partial integration yields
_

a
cos 2(f(u) mu) du =
_
Re
exp(2i(f(u) mu)
2im
_

u=a
+
+Re
1
m
_

a
f

(u) exp(2i(f(u) mu)) du

1
[m[
_
1 +
[f

(a)[
[f

(a) m[
_
.
Therefore,

m>f

(a)+

Re J
2
(m)
m

(a).
With slight modications this method also applies to the cases ImJ
2
(m)
and m f

(b) . Further, if 0 , [f

(b) , f

(a) +], then Lemma 2.7 gives


_
b
a
g(u) exp(2if(u)) du g(a).
In view of (2.6) the theorem follows from the above estimates under the
condition f

(b) > 0. If this condition is not fullled, then one can argue with
f(u) ku, where k := 1 [f

(b)], in place of f(u).


2.2.3. Proof of the approximate functional equation. Now we ap-
ply van der Corputs summation formula to the zeta-function. Let > 0. By
Theorem 1.3 we have
(s) =

nx
1
n
s
+

x<nN
exp(it log n)
n

+
N
1s
s 1
+ s
_

N
[u] u
u
s+1
du.
Setting g(u) = u

and f(u) =
t
2
log u, we get f

(u) =
t
2u
. Assume
that [t[ 4x, then [f

(u)[
7
8
. With the choice =
1
10
the interval
(f

(b) , f

(a) +) contains only the integer m = 0. Thus, van der Corputs


summation formula, Theorem 2.6, yields

x<nN
exp(it log n)
n

=
_
N
x
u
s
du + O(x

)
=
N
1s
x
1s
1 s
+ O(x

).
In addition with
s
_

N
[u] u
u
s+1
du [s[N

we deduce Theorem 2.3.


Theorem 2.3 is is a rst version of a family of formulae each of them called
approximate functional equation; the name reects the appearance of the
quantities s and 1 s as in the functional equation. There are stronger
78 Chapter 2 Zero-distribution
approximate functional equations known and their derivation relies heavily
on the functional equation; for instance:
Theorem 2.8. Let 0 1 and x, y, t > C > 0, where C is a constant C
and 2xy = t. Then
(2.6) (s) =

nx
1
n
s
+ (s)

ny
1
n
1s
+ O
_
x

+ t
1
2

y
1
_
uniformly in , where
(2.7) (s) := 2
s

s1
(1 s) sin
s
2
.
Here we have approximation by two shorter sums and with a much smaller
error term (if x and y are well balanced). This approximate functional equa-
tion was found by Hardy & Littlewood [78] in 1923 but was also known by
Riemann himself (see Siegels paper [182] on Riemanns unpublished papers
on (s)). The proof relies on complex variable methods, starting from the
identity
(s) =

nm
1
n
s
+
1
(s)
_

0
x
s1
exp(mx)
exp(x) 1
dx
and contour integration; more details can be found in Ivic [98]. As a matter
of fact, this approach is very much tied to the functional equation for (s).
An important extension of the classic techniques was given by Chan-
drasekharan & Narasimhan [30] for general Dirichlet series with functional
equations (e.g., Dedekind zeta-functions).
Exercise 58. Check that the function (s) dened by (2.7) satises
(s) = (s)(1 s).
Exercise 59. Deduce from Theorem 2.3 that, for any xed [
1
2
, 1),
(2.8) ( + it) t
1
as t .
Can you improve this estimate by use of (2.6)?
Exercise 60. * Prove Theorem 2.8.
Hint: for this aim consult Ivic [98].
2.3. Power moments
Power moments are important tools in the theory of Dirichlet series; in
particular, they give information on the number of zeros as we shall see
below. We follow [200], VII and IX.
Section 2.3 Power moments 79
2.3.1. The quadratic mean. Our aim is the second moment. By use
of the approximate functional equation, we shall derive an asymptotic mean-
square formula for (s) with error term valid in the half-plane >
1
2
.
Theorem 2.9. For >
1
2
,
_
T
1
[( + it)[
2
dt = (2)T + O(T
22
log T).
Proof. By the approximate functional equation,
( + it) =

n<t
1
n
+it
+ O(t

).
Using this and ( it) = ( + it), we get
_
T
1

n<t
1
n
+it

2
dt =
_
T
1

m,n<t
1
n
+it
m
it
dt
=

m,n<T
1
(mn)

_
T

_
m
n
_
it
dt
with := maxm, n. The diagonal terms m = n give the contribution

n<T
T n
n
2
= T
_
(2)

nT
1
n
2
_

n<T
1
n
21
= (2)T + O(T
22
).
The non-diagonal terms m ,= n contribute

m,n<T
m=n
1
(mn)

_
m
n
_
iT

_
m
n
_
i
i log
n
m

0<m<n<T
1
(mn)

log
n
m
.
If 1 m <
n
2
, then log
n
m
> log 2 > 0, and hence

n<T

m<
n
2
1
(mn)

log
n
m

n<T
1
n

_
2
T
22
.
If
n
2
m n, we write n = m + r with 1 r
n
2
. By the Taylor series
expansion of the logarithm,
log
n
m
= log
_
1
r
n
_
>
r
n
.
This gives

n<T

r
n
2
1
(mn)

log
n
m

n<T
n
12

r
n
2
1
r
T
22
log T.
Collecting together, the assertion of the theorem follows.
80 Chapter 2 Zero-distribution
The formula of Theorem 2.9 cannot hold for =
1
2
since then the main
term becomes singular: (2) is unbounded as
1
2
+. Indeed, on the
critical line the quadratic mean is of rather dierent form.
Theorem 2.10. As T ,
_
T
0

_
1
2
+ it
_

2
dt = T log T + O
_
T(log T)
1
2
_
.
This result is due to Hardy & Littlewood [77]. For the proof we refer once
more to Ivic [98].
2.3.2. Higher moments. It is a long standing conjecture that for xed
k 0, there exists a constant C(k) such that
(2.9)
1
T
_
T
0
[(
1
2
+ it)[
2k
dt C(k)(log T)
k
2
,
as T . It is not known whether this conjecture is related to Riemanns
hypothesis or not. The asymptotic formula (2.9) is known to be true only in
the trivial case k = 0, and the cases k = 1 and k = 2 by the classical results
of Hardy & Littlewood [77] (Theorem 2.10) and Ingham [96] who showed
that
(2.10)
_
T
0

_
1
2
+ it
_

4
dt
1
2
2
T(log T)
4
.
This was improved by several authors who gave further main terms and
appropriate error terms. There is a dierent and remarkable approach of
Motohashi [152] to the fourth moment using the spectral theory of the non-
Euclidean Laplacian on the upper half-plane.
Very little is known for higher moments. For the twelfth moment Heath-
Brown [84] gave the estimate
_
T
0

_
1
2
+ it
_

12
dt T
2
(log T)
17
.
By the work of Balasubramanian & Ramachandra [12] a lower bound of the
expected size holds for an arbitrary positive integer k
1
T
_
T
0
[(
1
2
+ it)[
2k
dt (log T)
k
2
.
However, satisfying upper bounds are, even under assumption of Riemanns
hypothesis, not known. For a nice introduction to these questions we refer
to Ivics monograph [98] and the survey Matsumoto [144].
Recently, Conrey & Gonek [39] and Keating & Snaith [114] stated a con-
jecture for the constant C(k) appearing in (2.9); remarkably, their heuristics
Section 2.3 Power moments 81
dier one from another (see also the survey Conrey [36]). To state this con-
jecture we dene
(2.11) a(k) =

p
_
1
1
p
2
_
k
2

m=0
_
(m+ k)
m!(k)
_
2
1
p
m
.
Note that one has to take an appropriate limit if k is an integer less than
or equal to zero. It is not dicult to verify that a(1) = 1 and a(2) =
6

2
;
however, further values are not explicitly known. Furthermore we have to
introduce Barnes double Gamma-function
G(z + 1) = (2)
z/2
exp
_

1
2
(z(z + 1) + z
2
)
_

n=1
_
1 +
z
n
_
n
exp
_
z +
z
2
n
_
,
where is Eulers constant (there will be no confusion with the imaginary
parts of the zeros of (s)); note that G(1) = 1 and G(z + 1) = (z)G(z).
The approach of Conrey & Gonek [39] is of combinatorial nature. They
investigated mean-value theorems for Dirichlet polynomials and proved
1
T
_
T
0

nx
d
k
(n)
n
1/2+it

2
dt
a(k)
(1 + k
2
)
(log x)
k
2
for x = o(T), where d
k
(n) is the generalized divisor function appearing as
coecients in the Dirichlet series representation of (s)
k
. Assuming that the
limit
g(k) := lim
T
__
T
0
[(
1
2
+ it)
k
[
2
dt
_
_
_
_
T
0

nx
d
k
(n)
n
1/2+it

2
dt
_
_
1
exists, they were led to conjecture
C(k) =
a(k)g(k)
(1 + k
2
)
for the constant in (2.9). Here one has g(1) = 1 and g(2) = 2. Furthermore,
they conjectured g(3) = 42 and g(4) = 24024. On the contrary, Keating
& Snaith [114] used the random matrix analogue. In fact, they proved, for
xed k >
1
2
,
E
N
1
2
_
2
0
[:
N
(; U)[
2k
d = E
N
[:
N
(0; U)[
2k

G(k + 1)
2
G(2k + 1)
N
k
2
;
this corresponds to a continuous 2k-th moment of the characteristic polyno-
mial :
N
(; U) associated with an arbitrary matrix U from the unitary group
|(N) of all N N matrices U with complex entries satisfying the condition
UU
t
= id
N
,
82 Chapter 2 Zero-distribution
where U
t
denotes the transpose of the complex conjugate of U and id
N
is
the N N identity matrix. The factor on the right-hand side of Keating &
Snaiths formula was found to coincide with some data from the Conrey &
Gonek-approach, namely
g(k)
(1 + k
2
)
=
k1

j=0
j!
(j + k)!
=
G(k + 1)
2
G(2k + 1)
,
what a surprise! The standard Random Matrix Theory-model cannot detect
the arithmetic factor (2.11): prime numbers do not occur in this model.
Consequently, the arithmetic information a(k), appearing in the heuristics of
Conrey & Ghosh, has to be inserted in an ad hoc way. Recently, Conrey,
Keating et al. modied the standard Random Matrix Theory-model which
incorporates also the arithmetic information a(k) (see Gonek [61]); this leads
directly to
Conjecture 1. For xed k >
1
2
, as T ,
1
T
_
T
0

_
1
2
+ it
_

2k
dt a(k)
G(k + 1)
2
G(2k + 1)
(log T)
k
2
.
Needless to say that this conjecture includes the only known cases, the trivial
one k = 0, and the classical cases k = 1 and k = 2 due to Hardy & Littlewood
and Ingham, respectively.
2.3.3. The Lindelof hypothesis. For many applications in number
theory it is useful to assume Riemanns hypothesis but quite often it suf-
ces to work with weaker conjectures. Lindel of [134] conjectured that (s)
is bounded if
1
2
+ with any xed positive . This would imply that

_
1
2
+ it
_
t

as t . The last statement is now known as Lindel ofs hypothesis and


it is yet unproved. However, the strong boundedness conjecture is false (see
the related exercise below). The Lindel of hypothesis follows from the truth
of the Riemann-hypothesis (as it follows from (2.22) below).
There are several further interesting reformulations of the Lindel of hypoth-
esis in case of the Riemann zeta-function. One, given in terms of moments
on the critical line, was found by Hardy & Littlewood [79]. They proved
that the Lindel of hypothesis is true if and only if all power moments are
suciently small:
Section 2.3 Power moments 83
Theorem 2.11. The Lindelof hypothesis is true if and only if, for any k N,
1
T
_
T
1

_
1
2
+ it
_

2k
dt T

.
A proof may be found in Titchmarsh [200]. This statement may serve as
a rst example for the importance of power moment estimates. Further
examples will be given in the following sections.
Exercise 61. Show that
_
T
0

_
1
2
+ it
_
dt T
1
2
log T.
Exercise 62. Prove a corresponding statement as Theorem 2.9 for Dirichlet L-
functions.
Exercise 63. Prove Theorem 2.10. Try to obtain a better error term...
Hint: use an approximate functional equation.
Exercise 64. * Use Theorem 2.8 to prove
_
T
0

_
1
2
+ it
_

4
dt T(log T)
4
Hint: one may consult, e.g., Ivic [98].
Using the theory of diophantine approximations Harald Bohr (the brother of the
physicist Niels Bohr and medal winner in the olympic football team of Denmark
1908) & Landau [20] showed that (s) takes arbitrarily large values in the half-
plane of absolute convergence Re s > 1 and s not from the neighborhood of the
pole at s = 1.
Exercise 65. * i) Show that for > 1
() [(s)[
N

n=1
cos(t log n)
n

n=N+1
1
n

.
ii) Prove the following statement about diophantine approximation (Dirichlets ap-
proximation theorem): Given arbitrary real numbers
1
, . . . ,
N
, a positive integer
q, and a positive number T, there exist real number [T, q
N
T] and integers
x
1
, . . . , x
N
for which
[
n
x
n
[
1
q
for n N.
iii) Apply ii) with
n
=
log n
2
to nd a real number [T, q
N
T] such that
cos( log n) cos
_
2
q
_
for n = 1, . . . , N.
iv) Prove the existence of an innite sequence of s = + it with 1+ and
t for which
[(s)[ (1 )(),
84 Chapter 2 Zero-distribution
where is an y positive constant, and deduce that for arbitrary T > 0
limsup
>1,t>T
[( + it)[ = .
Exercise 66. * Prove Theorem 2.11.
Hint: one may consult, e.g., Ivic [98].
2.4. Hardys theorem: zeros on the critical line
In 1914, Hardy [75] showed that there are indeed innitely many zeros
of the Riemann zeta-function on the critical line. This was generalized by
Lekkerkerker [129] to a general class of Dirichlet series satisfying a Riemann-
type functional equation.
2.4.1. Hardys Z-function. The behaviour of (s) on the critical line
is reected by Hardys Z-function Z(t) as a function of a real variable, dened
by
Z(t) = exp(i(t))
_
1
2
+ it
_
,
where
exp(i(t)) :=
it/2
(
1
4
+
it
2
)
[(
1
4
+
it
2
)[
.
It follows from the functional equation for (s) that Z(t) is an innitely often
dierentiable function which is real for real t. Moreover,

_
1
2
+ it
_

= [Z(t)[.
Consequently, the zeros of Z(t) correspond to the zeros of the Riemann zeta-
function on the critical line (counting multiplicities).
The function Z(t) has a negative local maximum at t = 2.4757 . . ., and
this is the only known negative local maximum in the range t 0; a positive
local minimum is not known. The occurrence of a negative local maximum,
besides the one at t = 2.4757 . . ., or a positive local minimum of Z(t), would
disprove Riemanns hypothesis. Indeed, one can show that if the Riemann
hypothesis is true, the graph of the logarithmic derivative Z

/Z(t) is mono-
tonically decreasing between the zeros of Z(t) for t 1000. A proof of this
claim can be found in Edwards [49].
Hardys Z-function allows to localize zeros on the critical line by applying
methods from real analysis. The Riemann-Siegel formula (discovered by Rie-
mann, rediscovered by Siegel while studying Riemanns unpublished papers)
Section 2.4 Hardys theorem 85
10 20 30 40 50 60
-3
-2
-1
1
2
3
Figure 1. Graphs of the modulus of the zeta-function (red) on the
critical line Re s =
1
2
and of Hardys Z-function (blue).
provides a very good approximation of the zeta-function on the critical line;
a rst an rather primitive form is
(2.12) Z(t) = 2

t/(2)
cos((t) t log n)
n
1/2
+ O
_
t
1/4
_
,
valid for t 1. We observe the similarity to approximate functional equa-
tions. The Riemann-Siegel formula is the basis of all high precision compu-
tations of the zeta-function on the critical line.
1
Lehmer [128] detected that the zeta-function occasionally has two very
close zeros on the critical line; for instance the zeros at t = 7005.0629 . . . and
t = 7005.1006 . . .. So the graph of Z(t) sometimes barely crosses the t-axis
(see Figure 4).
In view of our observation relating the graph of Z

/Z(t) with Riemanns


hypothesis from the previous section, Z(t) has exactly one critical point be-
tween successive zeros for suciently large t. Hence, Lehmers observation,
in the literature called Lehmers phenomenon, is a near-counterexample to
the Riemann hypothesis.
2.4.2. Hardys theorem. Now we are going to prove that there are
innitely many nontrivial zeros of the zeta-function on the critical line. How-
ever, we shall sketch the proof of a quantitative version:
1
A very nice animated plot of Z(t) can be found on Pughs webpage
http://www.math.ubc.ca/ pugh/RiemannZeta/RiemannZetaLong.html.
86 Chapter 2 Zero-distribution
7005.02 7005.04 7005.06 7005.08 7005.1 7005.12 7005.14
-0.06
-0.05
-0.04
-0.03
-0.02
-0.01
Figure 2. Lehmers phenomenon.
Theorem 2.12. For suciently large T and H T
1
4
+
, the interval (T, T +
H) contains at least one ordinate of a nontrivial zero = + i of (s) (of
odd order).
Sketch of the proof. We shall compare the to integrals
I
1
:=
_
T+H
T
[Z(t)[ dt and I
2
:=

_
T+H
T
Z(t) dt

.
The main idea is rather simple. If I
1
> I
2
, there is a sign change for Z(t) in
the interval (T, T +H) and we are done: the value of t for which Z(t) crosses
the t-axis is the ordinate of a nontrivial zero of odd order.
First of all, we bound I
1
from below. Clearly,
(2.13) I
1

_
T+H
T

_
1
2
+ it
_
dt

.
Using the approximate functional equation in the form of Theorem 2.8, we
nd

_
1
2
+ it
_
= 1 +

2n

t
2
n

1
2
it
+ exp(i
1
(t))

t
2
n
1
2
+it
+ O
_
t

1
4
_
,
where

1
(t) = t log
t
2e
+

4
.
The constant term 1 gives the contribution H to the bound of the right-hand
side of (2.13) while all other terms strongly oscillate for suciently large H.
It follows that I
1
H for H T
1
4
+
. For I
2
we apply the Riemann-Siegel
formula (2.12) in order to nd an upper bound of order o(H). (For the details
we refer to Karatsuba [111].)
Section 2.5 Density theorems 87
In the meantime several important quantitative improvements of Hardys
theorem 2.12 were made. Selberg [179] was the rst to prove that a positive
proportion of all zeros lies exactly on =
1
2
. Let N
0
(T) denote the number
of zeros of (s) on the critical line with imaginary part 0 < T. The
idea to use molliers to dampen the oscillations of [(
1
2
+ it)[ led Selberg to
liminf
T
N
0
(T + H) N
0
(T)
N(T + H) N(T)
> 0,
as long as H T
1
2
+
. Karatsuba [110] improved this result to H T
27
82
+
by some technical renements. The proportion is very small, about 10
6
as
Min calculated; a later renement by Zhuravlev gives after all
2
21
if H = T
(cf. Karatsuba [111], p.36). However, the localized zeros are not necessar-
ily simple. By an ingenious new method, working with molliers of nite
length, Levinson [132] localized more than one third of the nontrivial zeros
of the zeta-function on the critical line, and as Heath-Brown [85] and Selberg
(unpublished) discovered, they are all simple. By optimizing the technique
Levinson himself and others improved the proportion
1
3
sligthly, but more
recognizable is Conreys idea in introducing Kloosterman sums. So Conrey
[35] was able to choose a longer mollier to show that more than two fths
of the zeros are simple and on the critical line; Bauer [14, 15] improved this
proportion slightly. The use of longer molliers leads to larger proportions.
Farmer [52] observed that if it is possible to take molliers of innite length,
then almost all zeros lie on the critical line and are simple. In [189] Steuding
found a new approach (combining ideas and methods of Atkinson, Jutila and
Motohashi) to treat short intervals [T, T + H], i.e., H = o(T); it was proved
that for H T
0.552
a positive proportion of the zeros of the zeta-function
with imaginary parts in [T, T + H] lie on the critical line and are simple.
Exercise 67. Prove all statements concerning Z(t) from Section 2.4.1, except the
Riemann-Siegel formula.
Exercise 68. * i) Verify all steps in the proof of Theorem 2.12.
ii) Try to prove that if H T
1
2
+
, then any interval (T, T + H) with suciently
large T contains more than H many ordinates (of odd order).
2.5. Density theorems
Now we are going to study the complementary question: can we prove that
there are not too many zeros to the right of the critical line? In our studies
we shall frequently use the Riemann-von Mangoldt formula (1.12).
88 Chapter 2 Zero-distribution
2.5.1. Zeros o the critical line. Now we shall prove that most of the
nontrivial zeros of (s) cannot lie too far from the critical line =
1
2
. This
observation is from Bohr & Landau [21], resp. Littlewood [137].
First of all, we need Littlewoods lemma which relates the zeros of an
analytic function f(s) with a contour integral over log f(s).
Lemma 2.13 (Littlewood). Let A < B and let f(s) be analytic on 1 :=
s C : A B, [t[ T. Suppose that f(s) does not vanish on the
right edge = B of 1. Let 1

be 1 minus the union of the horizontal cuts


from the zeros of f in 1 to the left edge of 1, and choose a single-valued
branch of log f(s) in the interior of 1

. Denote by (, T) the number of


zeros = +i of f(s) inside the rectangle with > including zeros with
= T but not those with = T. Then
_
R
log f(s) ds = 2i
_
B
A
(, T) d.
We give a sketch of the simple proof. Cauchys theorem implies
_
R

log f(s) ds = 0, and so the left-hand side of the formula of the lemma,
_
R
, is minus the sum of the integrals around the paths hugging the cuts.
Since the function log f(s) jumps by 2i across each cut (assuming for sim-
plicity that the zeros of f in 1 are simple and have dierent height; the
general case is no harder),
_
R
is 2i times the total length of the cuts,
which is the right-hand side of the formula in the lemma. For more details
we refer to Titchmarsh [200], 9.9, or Littlewoods original paper [137].
Note that Littlewoods lemma can be used, in addition with Stirlings for-
mula and some facts about entire functions, to prove the Riemann-von Man-
goldt formula (1.12) (see Chapter 3).
Let N(, T) denote the number of zeros = + i of (s) with >
, 0 < T (counting multiplicities). We apply Littlewoods lemma to the
function f(s) = (s 1)(s) and the rectangle with corners 2 iT,
0
iT
where
0
(
1
2
, 1). Note that f(s) is entire and its zeros correspond one-to-one
to the zeros of (s).
2
_
1

0
2N(, T) d =
_
R
log f(s) ds
where we have (, T) = 2N(, T) since the zeros are symmetrically dis-
tributed with respect to the real axis. Now we want to remove the factor
s 1. Applying Littlewoods lemma once again, we get
i
_
R
log(s 1) ds = 2(1
0
),
and so the contribution of the factor s 1 in the last but one formula is
bounded. Taking into account (s) = (s) and that the integral over the
Section 2.5 Density theorems 89
zero-counting function is real-valued, we nd
2
_
1

0
N(, T) d =
_
T
0
log [(
0
+ it)[ dt
_
T
0
log [(2 + it)[ dt
+
_

0
2
arg ( + iT) d
_

0
2
arg ()) d (2.14)
+O(1);
here we dene log (s) to be the principal branch on the positive real axis.
The main contribution in (2.14) comes from the rst integral on the right-
hand side. The last integral does not depend on T and so it is bounded.
Moreover, we obtain
_
T
0
log [(2 + it)[ dt = Re
_

p,k
1
kp
2k
_
T
0
exp(itk log p) dt
_

n=2
1
n
2
1.
Now we estimate arg ( +iT). We may assume that T is not the ordinate
of zero. Since arg (2) = 0 and
arg (s) = arctan
_
Im(s)
Re (s)
_
,
where
Re (2 + it) =

n=1
cos(it log n)
n
2
1

n=2
1
n
2
> 1
_

1
du
u
2
= 0,
we have by the argument principle
[ arg (2 + iT)[

2
.
Now assume that Re ( + iT) vanishes q times as
1
2
2. Divide the
interval [
1
2
+iT, 2 +iT] into q +1 parts, throughout each of which Re (s) is
of constant sign. Hence, again by the argument principle, in each part the
variation of arg (s) does not exceed . This gives
[ arg (s)[
_
q +
3
2
_
for
1
2
.
Further, q is the number of zeros of the function
g(z) =
1
2
((z + iT) + (z iT)) = Re (z + iT)
for Imz = 0 and
1
2
Re z 2. Thus, q n(
3
2
), where n(r) is the number of
zeros of (s) for [z 2[ r. Obviously,
_
2
0
n(r)
r
dr
_
2
3
2
n(r)
r
dr n
_
3
2
__
2
3
2
dr
r
= n
_
3
2
_
log
4
3
.
90 Chapter 2 Zero-distribution
By Jensens formula 1.17 we obtain
_
2
0
n(r)
r
dr =
1
2
_
2
0
log [(2 + r exp(i))[ d log [(2)[.
In view of (2.8),
( + it) t
1
2
as t , we nd g(z) T
1
2
. Thus we obtain
q n
_
3
2
_

1
log
4
3
_
2
0
n(r)
r
dr log T.
This yields
arg ( + iT) log T uniformly for
1
2
,
and, consequently, the same bound holds by integration with respect to
1
2

2. The restriction that T has not to be an imaginary part of a zero of
(s) can be removed from considerations of continuity. Therefore, we may
replace (2.14) by
(2.15)
_
1

0
N(, T) d =
1
2
_
T
0
log [(
0
+ it)[ dt + O(log T).
Now we need a further analytic fact due to Jensen: Jensens inequality states
that for any continuous function f(u) on [a, b],
1
b a
_
b
a
log f(u) du log
_
1
b a
_
b
a
f(u) du
_
(for instance, this can be deduced from the arithmetic-geometric mean in-
equality, or see [199], 9.623). Hence, we obtain for any xed
0
>
1
2
_
T
0
log [( + it)[ dt
T
2
log
_
1
T
_
T
0
[( + it)[
2
dt
_
T
by applying Theorem 2.9. Thus,
_
1

0
N(, T) d T.
Let
1
=
1
2
+
1
2
(
0

1
2
), then
1
2
<
1
<
0
and we get
N(
0
, T)
1

1
_

0

1
N(, T) d
2

1
2
_
1

1
N(, T) T.
In view of (2.15) we have proved
Theorem 2.14. For any xed >
1
2
,
N(, T) T.
Section 2.5 Density theorems 91
The theorem above is a rst example of a so-called density theorem. By the
Riemann-von Mangoldt formula (1.12) we see that
(2.16) N(, T) = o(N(T)) for >
1
2
,
so all but an innitesimal proportion of the zeros of (s) lie in the strip
1
2
< <
1
2
+ , however small may be!
2.5.2. The zero-detection method. We want to prove a stronger re-
sult due to Bohr & Landau [22].
Theorem 2.15. For any xed in
1
2
< < 1,
N(, T) T
4(1)
(log T)
10
.
Proof. For 2 V T let N
1
(, V ) count the zeros = + i of (s) with
and
1
2
V < V . Taking x = V in Theorem 2.3 we have
(s) =

kV
1
k
s
+
V
1s
s 1
+ O
_
V

_
for
1
2
V < t V and
1
2
1. Multiplying this with the Dirichlet polyno-
mial
M
X
(s) :=

mX
(m)
m
s
,
where X = V
21
, gives
(s)M
X
(s) = P(s) + R(s),
where
R(s) [M
X
(s)[V

and
P(s) :=

mX
(m)
m
s

kV
1
k
s
=

nXV
a(n)
n
s
with
(2.17) a(n) :=

m|n
mX,nmV
(m) =
_
1 if m = 1,
0 if 1 < n X.
Note that M
X
(s), as the truncated Dirichlet series of the reciprocal of (s),
mollies
1
(s)
. We shall use P(s) as a zero-detector. Let s = be a zero of
the zeta-function with
1
2
V < V . Then,
1

X<nXV
a(n)
n

+ O([M
X
()[V

),
1

X<nXV
a(n)
n

2
+ O([M
X
()[
2
V
2
).
92 Chapter 2 Zero-distribution
Then, summing up both sides of the latter inequality over all such N zeros
leads to
(2.18) N
1
(V )

1
1
2
V <V
_
_

X<nXV
a(n)
n

2
+[M
X
()[
2
V
2
_
_
.
Now we divide the interval [
1
2
V, V ] into subintervals of length 1 of the form
[2m+n 1, 2m+ n], where n = 1, 2 and
1
4
V 1 m
1
2
V . Then, we may
write

1
1
2
V <V

1
4
V 1m
1
2
V
2

n=1

2m+n1<2m+n
2 max
1n2

1
4
V 1m
1
2
V

2m+n1<2m+n
.
In view of the Riemann-von Mangoldt formula (1.12) there are only log V
many zeros with 2m + n 1 < 2m + n. Now let

denote the largest


of the related sums according to 2m+ n 1 < 2m+ n. Then

1
1
2
V <V
log V

,
resp. in (2.18)
(2.19) N
1
(V ) log V

_
_

X<nXV
a(n)
n

2
+

mX
(m)
m

2
V
2
_
_
.
First of all we shall give a bound for
S(Y ) :=

Y <nU
b(n)
n

2
,
where U 2Y and V Y 1 and
(2.20) b(n)

d|n
1 =: d(n),
where d(n) is the divisor function. By partial summation, for xed = +i,

Y <nU
b(n)
n

=
_
U
Y
C(u) du

with C(u) :=

Y <nu
b(n)
n
i
.
Section 2.5 Density theorems 93
Applying the Cauchy-Schwarz inequality we obtain

Y <nU
b(n)
n

Y
1
_
U
Y
[C(u)[ du + Y

[C(U)[,

Y <nU
b(n)
n

2
Y
21
_
U
Y
[C(u)[
2
du + Y
2
[C(U)[
2
.
This leads to
S(Y ) Y
2

Y <nW
b(n)
n
i

2
,
where W U. Since the distance of the imaginary parts of counted zeros

r
=
r
+ i
r
is 1, we can nd

Y <nW
b(n)n
i
r+1

_

r+1
r

Y <nW
b(n)n
it

2
dt
+2
_

r+1
r

Y <nW
b(n)n
it

Y <mW
b(m) log m m
it

dt.
Summation over r and application of Cauchy-Schwarz yields
S(Y ) Y
2
(I
1
+
_
I
1
I
2
),
where
I
1
:=
_
V
1
2
V

Y <nW
b(n)n
it

2
dt , I
2
:=
_
V
1
2
V

Y <nW
b(n) log n n
it

2
dt.
Taking (2.17) into account, [a(n)[ satises condition (2.20) on b(n). By
elementary estimates one can show that

nx
d
k
(n) x(log x)
k
,
where the implicit constant depends only on k; a proof can be found in [112]
(see also Exercise 33). This yields
I
1
(V + Y ) log V

Y <n2Y
d
2
(n) (V Y + Y
2
)(log V )
5
,
I
2
(V Y + Y
2
)(log V )
7
.
94 Chapter 2 Zero-distribution
Now dividing the rst sum on the right hand side of (2.19) into log V
sums, application of the latter estimates yields
log V

X<nV X
a(n)
n

2
(V X
12
+ (V X)
22
)(log V )
9
.
Similarly, we get for the second term
V
2
(log T)
2

mX
(m)
m

2
V
2
(V + X
22
)(log V )
9
.
Substituting this in (2.19) with regard to X = V
21
, we obtain
N
1
(V ) V
4(1)
(log V )
9
.
Using this with V = T
1n
and summing up over all n N, proves the
theorem.
2.5.3. The density hypothesis. There are stronger estimates known
than the one of Theorem 2.15. For instance, the strongest one which holds
throughout the right half of the critical strip is
N(, T) T
2.4(1)
(log T)
18.2
due to Huxley [94], resp. Gritsenko [70] who improved the former exponent
of the log-term. This estimate has remarkable consequences on the prime
number distribution; namely, it follows that
(x + x

) (x) = x

+ o
_
x

_
for any >
7
12
, as x . This implies that for suciently large x, there is
always a prime number p in any interval (x, x +x
7
12
+
), but it is too weak to
prove that there is always a prime in between consecutive squares.
The density hypothesis states that, for all >
1
2
,
N(, T) T
2(1)+
. (2.21)
Of course, if the Riemann hypothesis is true, then N(, T) is identically
zero for any >
1
2
. How is the density hypothesis related to the Lindel of
hypothesis? Backlund [6] proved that the Lindel of hypothesis is equivalent
to the much less drastic but yet unproved hypothesis that for every >
1
2
N(, T + 1) N(, T) = o(log T). (2.22)
Furthermore, the Lindel of hypothesis implies the density hypothesis. There-
fore, we have the following hierarchy:
Riemann hypothesis Lindel of hypothesis density hypothesis.
Section 2.6 Universality 95
Exercise 69. Denote the zeros of (s) by = +i. Show that, for xed
0
>
1
2
,

0<T
>
0
(
0
) T.
Hint: compute the integral
_
1

0
N(, T) d.
Exercise 70. * i) Prove Backlunds statement that (2.22) is equivalent to the
Lindel of hypothesis.
ii) Show that the Lindel of hypothesis implies the density hypothesis.
Hint: one may consult, for example, Patterson [164].
2.6. Universality and self-similarity.
We conclude with an application of the results presented in the last sections,
Voronins famous universality theorem [206] which roughly states that any(!)
non-vanishing analytic function can be approximated uniformly by certain
shifts of the Riemann zeta-function. This universal property is related to the
zero-distribution; we shall deduce an equivalent for the truth of the Riemann
hypothesis due to Bohr and Bagchi.
2.6.1. Voronins universality theorem. In 1975, Voronin [206]
proved the following
Theorem 2.16. Let f(s) be a non-vanishing continuous function dened on
a disk s C : [s[ r with some r (0,
1
4
), and analytic in the interior.
Then, for any > 0, there exists > 0 such that
max
|s|r

_
s +
3
4
+ i
_
f(s)

< ;
moreover,
liminf
T
1
T
meas
_
[0, T] : max
|s|r

_
s +
3
4
+ i
_
f(s)

<
_
> 0.
Thus, the set of for which shifts of the zeta-function approximate f(s) with
a given accuracy has positive lower density (with respect to the Lebesgue
measure). We say that (s) is universal since appropriate shifts approximate
uniformly any element of a huge class of functions.
We give a very brief sketch of Voronins argument following the book of
Karatsuba & Voronin [112]. The Euler product for (s) is the key to prove
the universality theorem in spite of the fact that it does not converge in the
region of universality. However, as Bohr observed, an appropriate truncated
Euler product approximates (s) in a certain mean-value sense inside the
critical strip; this is related to the use of modied truncated Euler products
in Voronins proof (see (2.25) and (2.26) below).
96 Chapter 2 Zero-distribution
It is more convenient to work with series than with products. Therefore, we
consider the logarithms of the functions in question. Since g(s) has no zeros
in [s[ r its logarithm exists and we may dene an analytic function f(s)
on [s[ r by g(s) = exp f(s). First we approximate f(s) by the logarithm
of a truncated Euler product. Let denote the set of all sequences of real
numbers indexed by the prime numbers in ascending order. Further, dene
for every nite subset M of the set of all primes, every = (
2
,
3
, . . .)
and all complex s,

M
(s, ) =

pM
_
1
exp(2i
p
)
p
s
_
1
.
Obviously,
M
(s, ) is a non-vanishing analytic function of s in the half-plane
> 0. Consequently, its logarithm exists and is equal to
log
M
(s, ) =

pM
log
_
1
exp(2i
p
)
p
s
_
;
in order to have a denite value we may choose the principal branch of the
logarithm. Since f(s) is uniformly continuous in the disc [s[ r, there exists
some > 1 such that
2
r <
1
4
and
max
|s|r

f
_
s

2
_
f(s)

<

2
.
The function f
_
s

2
_
is bounded on the disc [s[ r =: R, and thus belongs
to the Hardy space 1
2
R
, i.e., the Hilbert space consisting of those functions
F(s) which are analytic for [s[ < R with nite norm
|F| := lim
rR
__
|s|r
[F(s)[ d dt
and inner product
F, G := Re
__
|s|R
F(s)G(s) d dt.
Denote by p
k
the k-th prime number. We consider the series

k=1
u
k
(s, ), where u
k
(s, ) := log
_
1
exp(2i
p
k
)
p
s+
3
4
k
_
1
.
Here comes the rst main idea. Riemann proved that any conditionally con-
vergent series can be rearranged such that its sum converges to an arbitrary
preassigned real number. Pechersky [165] generalized Riemanns theorem to
Hilbert spaces. It follows, with the special choice =
0
= (
1
4
,
2
4
,
3
4
, . . .), that
there exists a rearrangement of the series

u
k
(s) for which

j=1
u
k
j
(s,
0
) = f
_
s

2
_
Section 2.6 Universality 97
(the rather dicult and lengthy verication of the conditions of Pecher-
skys theorem uses classic results of Paley & Wiener and Plancherel from
Fourier analysis, a theorem on the approximation by polynomials due to
A.A. Markov, and, most importantly, the prime number theorem 1.9).
The tail of the rearranged series can be made as small as we please, say of
modulus less than

2
. Thus, it turns out that for any > 0 and any y > 0
there exists a nite set M of prime numbers, containing at least all primes
p y, such that
max
|s|r

log
M
_
s +
3
4
,
0
_
f(s)

< . (2.23)
The next and main step in Voronins proof is to switch from log
M
(s) to
the logarithm of the zeta-function. Of course, log (s) has singularities at
the zeros of (s), but since the set of these possibly singularities has measure
zero by density theorem 2.15, they are negligible.
We choose > 1 and
1
(0, 1) such that r <
1
4
and
(2.24) max
|s|r

f
_
s

_
f(s)

<
1
.
Putting Q = p : p z and c = s = + it : r < 2, [t[ 1, one
can show, using the approximate functional equation for (s), Theorem 2.3,
that for any
2
> 0
(2.25)
_
2T
T
_ _
E

1
Q
_
s +
3
4
+ i, 0
_

_
s +
3
4
+ i
_
1

2
d dt d
4
2
T,
provided that z and T are suciently large, depending on
2
; here 0 :=
(0, 0, . . .). Now dene
/
T
=
_
[T, 2T] :
__
E+
3
4
[
1
Q
(s + i, 0)(s + i) 1[
2
d dt <
2
2
_
.
Then it follows from (2.25) that, for suciently large z and T,
(2.26) meas (/
T
) > (1
2
)T,
which is surprisingly large. It follows from Cauchys formula that, for su-
ciently small
2
,
max
|s|r

log
_
s +
3
4
+ i
_
log
Q
_
s +
3
4
+ i, 0
_

2
, (2.27)
provided /
T
, where the implicit constant depends only on . By (2.23)
there exists a sequence of nite sets of prime numbers M
1
M
2
. . . such
that

k=1
M
k
contains all primes and
lim
k
max
|s|r

log
M
k
_
s +
3
4
,
0
_
f
_
s

= 0.
98 Chapter 2 Zero-distribution
Let
0
= (
(0)
2
,
(0)
3
, . . .). By the continuity of log
M
_
s +
3
4
,
0
_
, for any

1
> 0 there exists a positive such that, whenever the inequalities
(2.28) |
(0)
p

p
| < for p M
k
hold, where |z| denotes the minimal distance of z to an integer, then
max
|s|r

log
M
k
_
s +
3
4
,
0
_
log
M
k
_
s +
3
4
,
_

<
1
. (2.29)
Let
B
T
=
_
[T, 2T] :
_
_
_
log p
2

(0)
p
_
_
_ <
_
.
Now we consider
1
T
_
B
T
__
|s|r

log
Q
_
s +
3
4
+ i, 0
_
log
M
k
_
s +
3
4
+ i, 0
_

2
d dt d,
resp.
__
|s|r
1
T
_
B
T

log
Q
_
s +
3
4
+ i, 0
_
log
M
k
_
s +
3
4
+ i, 0
_

2
d d dt.
Putting
(2.30) () =
_

log 2
2
,
log 3
2
, . . .
_
,
we may rewrite the inner integral as
_
B
T

log
Q
_
s +
3
4
, ()
_
log
M
k
_
s +
3
4
, ()
_

2
d.
Now we need Weyls renement of Kroneckers approximation theorem. Let
() be a continuous function with domain of denition [0, ) and range R
N
.
Then the curve () is said to be uniformly distributed mod 1 in R
N
if, for
every parallelepiped

= [
1
,
1
] . . . [
N
,
N
] with 0
j
<
j
1 for
1 j N,
lim
T
1
T
meas
_
(0, T) : ()

mod 1
_
=
N

j=1
(
j

j
).
In a sense, a curve is uniformly distributed mod 1 if the correct proportion
of values lies in a given subset of the unit cube. In questions about uniform
distribution mod 1 one is interested in the fractional part only. For a curve
() in R
N
, we dene
() = (
1
() [
1
()], . . . ,
N
() [
N
()]),
where [x] denotes the integral part of x R.
Section 2.6 Universality 99
Lemma 2.17. i) Let a
1
, . . . , a
N
be real numbers, linearly independent over Q,
and let be a subregion of the N-dimensional unit cube with Jordan content
. Then
lim
T
1
T
meas (0, T) : (a
1
, . . . , a
N
) mod 1 = .
ii) Suppose that the curve () is uniformly distributed mod 1 in R
N
. Let T
be a closed and Jordan measurable subregion of the unit cube in R
N
and let
be a family of complex-valued continuous functions dened on T. If is
uniformly bounded and equicontinuous, then
lim
T
1
T
_
T
0
f(())1
D
() d =
_
D
f(x) dx
uniformly with respect to f , where 1
D
() is equal to 1 if () T mod 1,
and zero otherwise.
Note that the notion of Jordan content is more restrictive than the notion of
Lebesgue measure. But, if the Jordan content exists, then it is also dened
in the sense of Lebesgue and equal to it. A proof of Weyls theorem can be
found in Karatsuba & Voronin [112].
The unique prime factorization of integers implies the linear independence
of the logarithms of the prime numbers over the eld of rational numbers. By
Lemma 2.17, i), the curve (), dened by (2.30), is uniformly distributed
mod 1. Application of Lemma 2.17, ii), to the curve () yields
lim
T
1
T
_
B
T

log
Q
_
s +
3
4
, ()
_
log
M
k
_
s +
3
4
, ()
_

2
d
=
_
D

log
Q
_
s +
3
4
,
_
log
M
k
_
s +
3
4
,
_

2
d,
uniformly in s for [s[ r, where T is the subregion of the unit cube in
R
N
given by the inequalities (2.28) with N = M
k
, and d is the Lebesgue
measure. By the denition of
M
(s, ) it follows that for M
k
Q

Q
(s, ) =
M
k
(s, )
Q\M
k
(s, ),
and thus
_
D

log
Q
_
s +
3
4
,
_
log
M
k
_
s +
3
4
,
_

2
d
meas (T)
_
[0,1]
N

log
Q\M
k
_
s +
3
4
,
_

2
d.
100 Chapter 2 Zero-distribution
The latter integral is bounded above by y
2r
1
2
k
provided that M
k
contains all
primes y
k
. It follows that
1
T
_
B
T
__
|s|r

log
Q
_
s +
3
4
+ i, 0
_
log
M
k
_
s +
3
4
+ i, 0
_

2
d dt d
y
2r
1
2
k
meas (T).
Applying Lemma 2.17, ii) once more yields
lim
T
1
T
meas (B
T
) = meas (T),
which implies, for suciently large y
k
,
meas
_
B
T
:
__
|s|r

log
Q
_
s +
3
4
+ i, 0
_
log
M
k
_
s +
3
4
+ i, 0
_

2
d dt < y
r
1
4
k
_
>
meas (T)
2
T,
and
meas
_
B
T
: max
|s|r

log
Q
_
s +
3
4
+ i, 0
_
log
M
k
_
s +
3
4
+ i, 0
_

< y
1
5
(r
1
4
)
k
_
>
meas (T)
2
T. (2.31)
If we now take 0 <
2
<
1
2
meas (T), then (2.26) implies
meas (/
T
B
T
) > 0.
Thus, in view of (2.23) and (2.24) we may approximate f(s) by
log
M
k
_
s +
3
4
,
0
_
(independent on ), with (2.29) and (2.31) the lat-
ter function by log
Q
_
s +
3
4
+ i, 0
_
, and nally with regard to (2.27) by
log
_
s +
3
4
+ i
_
on a set of with positive measure. Replacing T by
1
2
T,
we thus nd, for any > 0,
liminf
T
1
T
meas
_
[0, T] : max
|s|r

log
_
s +
3
4
+ i
_
f(s)

<
_
> 0.
Now taking the exponential we obtain Voronins theorem.
Theorem 2.16 was generalized and extended in several directions. Reich
[174] and Bagchi [7] replaced the disk by an arbitrary compact subset of the
right half of the critical strip with connected complement, and by giving a
lucid proof in the language of probability theory. The strongest version of
Voronins theorem has the form:
Section 2.6 Universality 101
Theorem 2.18. Suppose that / is a compact subset of the strip
1
2
< <
1 with connected complement, and let g(s) be a non-vanishing continuous
function on / which is analytic in the interior of /. Then, for any > 0,
liminf
T
1
T
meas
_
[0, T] : max
sK
[(s + i) g(s)[ <
_
> 0.
This theorem can be found in the monograph Laurincikas [127] which also
contains proofs of universality for Dirichlet L-functions; in Steuding [191]
universality for a large class of L-functions was proved.
A natural question arises: is the condition on the non-vanishing of g(s) in
the universality theorem necessary, i.e., is it possible to approximate uniformly
functions having zeros by shifts of (s) (in the sense of Voronins universality
theorem)? The answer is negative. We give a heuristic argument which can
easily be made waterproof. It relies on the classic Rouches theorem:
Lemma 2.19. Let f(s) and g(s) be analytic for [s[ r. If
[f(s) g(s)[ < [g(s)[
on [s[ = r, then f(s) and g(s) have the same number of zeros in [s[ < r.
This result follows from a simple application of the argument principle; for
details see Burckel [28], VIII.3, or Titchmarsh [199], 3.42.
Assume that g(s) is an analytic function on [s[ r, where 0 < r <
1
4
,
which has a zero with [[ < r but which is non-vanishing on the boundary.
An application of Rouches theorem shows that whenever the inequality
(2.32) max
|s|r

_
s +
3
4
+ i
_
g(s)

< min
|s|=r
[g(s)[
holds,
_
s +
3
4
+ i
_
has to have a zero inside [s[ < r. The zeros of an analytic
function lie either discretely distributed or the function vanishes identically
and thus the inequality (2.32) holds if the left hand side is suciently small.
If now for any > 0
liminf
T
1
T
meas
_
[0, T] : max
|s|r

_
s +
3
4
+ i
_
g(s)

<
_
> 0,
then we expect T many complex zeros of (s) in the strip
3
4
r < <
3
4
+r
up to T (for a rigorous proof one has to consider the densities of values
satisfying (2.32); this can be done along the lines of the proof of Theorem
2.20 below). This contradicts the density theorem 2.15, which gives
N
_
3
4
r, T
_
= o(T).
Thus, uniform approximation of a function g(s) having zeros by the zeta-
function cannot be done!
102 Chapter 2 Zero-distribution
2.6.2. Almost periodicity. Bohr introduced the fruitful notion of al-
most periodicity into analysis. An analytic function f(s), dened on some
vertical strip a < < b, is called almost periodic if, for any positive , and
any , with a < < < b, there exists a length = (f, , , ) > 0 such
that every interval (t
1
, t
2
) of length contains an almost period of f relatively
to in the closed strip , i.e., there exists a number (t
1
, t
2
) such
that
(2.33) [f( + it + i) f( + it)[ < for , t R.
Bohr [19] proved that every Dirichlet series is almost-periodic in its half-
plane of absolute convergence. Furthermore, he discovered an interesting
relation between the Riemann hypothesis and almost periodicity; indeed,
his aim in introducing the concept of almost periodicity might have been
Riemanns hypothesis. Bohr showed that if is non-principal, then the
Riemann hypothesis for the Dirichlet L-function L(s, ) is equivalent to the
almost periodicity of L(s, ) in >
1
2
. The condition on the character might
appear to be a bit unnatural but is necessary for Bohrs reasoning. His
argument relies in the main part on diophantine approximation applied to
Dirichlet series inside the critical strip. The Dirichlet series for L(s, ) with
a non-principal character converges throughout the critical strip, but the
one for the zeta-function does not.
2.6.3. An equivalent for RH. More than half a century later Bagchi
[7] proved that the Riemann hypothesis is true if and only if the zeta-function
can approximate itself in the sense of Voronins universality theorem. In [8],
Bagchi generalized this result in various directions; in particular for Dirich-
let L-functions to arbitrary characters. One implication of his proof in [8]
relies essentially on Voronins universality theorem (resp. its generalization
to Dirichlet L-functions), which, of course, was unknown to Bohr. Later,
Bagchi [9] gave another proof in the language of topological dynamics, inde-
pendent of universality, and therefore this property, equivalent to Riemanns
hypothesis, is called strong recurrence.
Theorem 2.20. Let
1
2
. Then (s) is non-vanishing in the half-plane
> if and only if, for any > 0, any z with Re z > , and any 0 < r <
minRe z , 1 Re z,
liminf
T
1
T
meas
_
[0, T] : max
|sz|r
[(s + i) (s)[ <
_
> 0.
Proof. If Riemanns hypothesis is true we can apply Voronins universality
theorem in the form of Theorem 2.18 with g(s) = (s), which implies the
strong reccurence. The idea for the proof of the other implication is that
if there is at least one zero to the right of the line = , then the strong
Section 2.6 Universality 103
recurrence property implies the existence of many zeros, too many with regard
to the classic density theorem 2.15.
Suppose that there exists a zero of (s) with Re > . Without loss of
generality we may assume that Im > 0. We have to show that there exists
a disc with center z and radius r, satisfying the conditions of the theorem,
and a positive such that
(2.34) liminf
T
1
T
meas
_
[0, T] : max
|sz|r
[(s + i) (s)[ <
_
= 0.
Locally, the zeta-function has the expansion
(2.35) (s) = c(s )
m
+ O
_
[s [
m+1
_
with some non-zero c C and m N. Now assume that for a neighbourhood
/

:= s C : [s [ of the relation
(2.36) max
sK

[(s + i) (s)[ < min


|s|=
[(s)[
holds; the second inequality holds for suciently small . Then Rouches
theorem 2.19 implies the existence of a zero of (s) in
/

+ i := s C : [s i [ .
We may say that the zero of (s) is generated by the zero . With regard
to (2.35) and (2.36) the zeros and = + i are intimately related; more
precisely,
> [() ( i)[ = [( i)[ [c[ [ i [
m
+ O(
m+1
).
Hence,
[ i [
_

[c[
_1
m
+ O
_

1+
1
m
_
.
In particular,
1
2
< Re 2
_

[c[
_1
m
< < 1,
and
[ ( + Im)[ < 2
_

[c[
_1
m
,
for suciently small and = o(
m+1
). Next we have to count the generated
zeros in terms of . Two dierent shifts
1
and
2
can lead to the same zero
, but their distance is bounded by
[
1

2
[ < 4
_

[c[
_1
m
.
104 Chapter 2 Zero-distribution
If we now write
J(T) :=
_
j
J
j
(T) :=
_
[0, T] : max
sK

[(s + i) (s)[ <


_
,
where the J
j
(T) are disjoint intervals, it follows that there are

_
1
4
_
[c[

_ 1
m
meas (J
j
(T))
_
+ 1 >
1
4
_
[c[

_ 1
m
meas (J
j
(T))
many distinct zeros according to J
j
(T), generated by . The number of
generated zeros is a lower bound for the number of all zeros. It follows that

_
= + i : > Re 2
_

[c[
_ 1
m
, 0 < < T + Im + 2
_

[c[
_ 1
m
_

1
4
_
[c[

_1
m
meas (J(T)).
This and the density theorem 2.15 lead to
meas (J(T)) = o(T),
which implies (2.34). The theorem is proved.
The expected strong reccurence of (s) may be regarded as a kind of self-
similarity. Assuming the truth of Riemanns hypothesis this has a nice in-
terpretation. Consider the amplitude of light which is a physical bound for
the size of objects which human beings can see, or the Planck constant 10
33
which is the smallest size of objects in quantum mechanics. Thus, if we as-
sume that is less than one of these quantities, then we cannot physically
distinguish between (s) and (s +i) for s from a compact subset / of the
right half of the critical strip, whenever
max
sK
[(s + i) (s)[ < .
This shows that we cannot decide where we actually are in the analytic
landscape of (s) without moving to the boundary. The zeta-function is an
amazing maze!
Exercise 71. * Study the proof of Voronins universality theorem in detail. Extend
the argument to Dirichlet L-functions.
CHAPTER 3
Modular forms and Hecke theory
This chapter is devoted to functional equations. We will prove the func-
tional equation for the Riemann zeta-function and sketch the proof of the one
for Dirichlet L-functions. Furthermore, we will discuss in detail an important
link between Dirichlet series satisfying a Riemann-type functional equation
and modular forms, discovered by Hecke in the 1930s.
3.1. The functional equation for zeta and more
Now we shall prove the functional equation for Riemanns zeta-function;
this will complete our studies on the analytic continuation from the rst
chapter.
Theorem 3.1. For any s C,

s
2

_
s
2
_
(s) =

1s
2

_
1 s
2
_
(1 s).
Riemann [175] himself gave two proofs of the functional equation. In the
meantime, quite many dierent proofs were found (see for example [200]).
Here we follow Riemanns original approach which relies on the functional
equation of the theta-function. In order to give a rigorous proof we therefore
prove rst Poissons summation formula and apply this to the theta-function
in order to obtain its functional equation. This is by far not the fastest way
to prove Theorem 3.1; however, this method applies to Dirichlet L-functions
as well and we shall sketch the proof of their functional equation too. But
more than that: this approach will also play a substantial role in the sequel
of this chapter.
3.1.1. The Poisson summation formula. Suppose f : R C is an
integrable function satisfying f(z) [z[
2
as [z[ (actually, this is a
strong restriction but it allows to do the next step). Then we may dene its
Fourier transform by

f(y) =
_
+

f(z) exp(2iyz) dz.


The Poisson summation formula is a useful tool in Fourier theory with many
applications in real and complex analysis.
105
106 Chapter 3 Hecke theory
Theorem 3.2. Let f : R R be a twice continuously dierentiable function
with f(z) [z[
2
as [z[ . Further, assume that the integral
_
+

[f

(z)[ dz
exists. Then, for any R,

nZ
f(n + ) =

mZ

f(m) exp(2im).
Proof. It suces to prove the formula in question only for = 0. In fact,
writing g(z) = f(z + ) for xed R, we have g(y) =

f(y) exp(2iy).
Therefore, we may assume = 0.
First of all, for r > 0, dene
P(y, r) =

m=
r
|m|
exp(2imy).
This series is the sum of the term for m = 0 plus two innite geometric
series, one for m < 0 and one for m > 0, both being absolutely convergent
for r [0, 1). Hence, we can compute the value of the innite series P(y, r)
by
P(y, r) = 1 +
r exp(2iy)
1 r exp(2iy)
+
r exp(2iy)
1 r exp(2iy)
=
1 r
2
1 2r cos(2y) + r
2
.
This implies P(y, r) 0 for any y (since the denominator is equal to (r
cos 2y)
2
+ (sin 2y)
2
). Using
_
1
0
exp(2imy) dy =
_
1 if m = 0,
0 otherwise,
we nd
_
1
0
P(y, r) dy = 1
for all r [0, 1). Further note that P(y, r) is 1-periodic with respect to y.
Hence,
P(y, r)
1 r
2
(sin 2)
2
for 0 < [y[
1
2
.
Section 3.1 The functional equation 107
Since f(z) z
2
, we have
+

m=
r
|m|

f(m) =
_
+

P(y, r)f(y) dy
=
+

m=
_
[m
1
2
,m+
1
2
]
P(y, r)f(y) dy;
interchanging summation and integration is justied with respect to the ab-
solute convergence. We want to show that the right-hand side converges to

m
f(m) as r 1. For this purpose we note that
_
[m
1
2
,m+
1
2
]
P(y, r)f(y) dy max
m
1
2
ym+
1
2
[f(y)[
_
1
0
P(y, r) dy
max
m
1
2
ym+
1
2
[f(y)[ m
2
,
as [m[ . Hence, given > 0, there exists M > 0 such that

|m|>M
_
[m
1
2
,m+
1
2
]
P(y, r)f(y) dy < and

|m|>M
[f(m)[ < .
Now assume [m[ M. Of course,
_
[m
1
2
,m+
1
2
]
P(y, r)f(y) dy f(m) =
_
[m
1
2
,m+
1
2
]
P(y, r)(f(y) f(m)) dy.
Take some > 0 for which [f(y) f(z)[ <

3M
for all m with [m[ M and
all y, z with [y z[ . Then

|m|M

_
[m
1
2
,m+
1
2
]
P(y, r)f(y) dy f(m)

|m|M
(J
1
(m) + J
2
(m)), (3.1)
where
J
1
(m) :=
_
m+
m
P(y, r)[f(y) f(m)[ dy,
J
1
(m) :=
_
W(m)
P(y, r)[f(y) f(m)[ dy
with W(m) := y R : < [y m[
1
2
. By construction,
J
1
(m)

3M
_
m+
m
P(y, r) dy

3M
.
Moreover,
J
2
(m)
1 r
2
(sin 2)
2
_
W(m)
[f(y) f(m)[ dy
1 r
2
m
2
,
108 Chapter 3 Hecke theory
where the implicit constant depends only on , and f. Thus, the right-hand
side of (3.1) can be made less than 2 for some r suciently close to 1. Hence,
letting 0, we obtain
(3.2) lim
r1
+

m=
r
|m|

f(y) =
+

m=
f(m).
Partial integration shows

f(m) m
2
. Consequently, the series on the left-
hand side of (3.2) converges absolutely and uniformly for r [0, 1) and we
may interchange summation and take the limit. This proves the theorem.
3.1.2. The theta-function. The (most simple) theta-function is given
by the innite series
(x) =

nZ
exp(xn
2
).
We apply Poissons summation formula, Theorem 3.2, to the function f(z) :=
exp(z
2
/x) with x > 0. We compute the Fourier transform by quadratic
substitution:

f(y) =
_
+

exp((z
2
/x + 2iyz)) dz
= xexp(xy
2
)
_
+

exp(x(w + iy)
2
) dw. (3.3)
Next we consider the integral
I() :=
_
+

exp(x(w + )
2
) dw,
where is any complex number. Consider the integral
_
R
exp(x
2
) d,
where 1 is the rectangular contour with vertices r, r +iIm, where r is a
positive real number. By Cauchys theorem, the integral is equal to zero. On
the line Re = r, the integrand tends uniformly to zero as r . Hence,
I() = I(0), and thus the integral I() does not depend on . This gives in
(3.3)

f(y) = xexp(xy
2
)
_
+

exp(xw
2
) dw = C

xexp(xy
2
),
where
C :=
_
+

exp(z
2
) dz.
Applying Poissons summation formula leads to

nZ
exp((n + )
2
/x) = C

nZ
exp(xn
2
+ 2in);
Section 3.1 The functional equation 109
here we have introduced the parameter by the trick from the proof of
Theorem 3.2. Choosing = 0 and x = 1, both sums are equal; thus, C = 1
and we have just proved the functional equation for the theta-function:
Theorem 3.3. For any x > 0,
(x) =
1

_
1
x
_
.
3.1.3. The proof of the functional equation. The Gamma-function
plays an important part in the theory of the zeta-function (see [199], 1.86
and 4.41, for a collection of its most important properties). For Re z > 0,
the Gamma-function may be dened by Eulers integral
(z) =
_

0
u
z1
exp(u) du.
Substituting u = n
2
x leads to
(3.4)
_
s
2
_

s
2
1
n
s
=
_

0
x
s
2
1
exp(n
2
x) dx.
Summing up over all n N yields

s
2

_
s
2
_

n=1
1
n
s
=

n=1
_

0
x
s
2
1
exp(n
2
x) dx.
On the left-hand side we nd the Dirichlet series dening (s); in view of its
convergence, the latter formula is valid only for > 1. On the right-hand
side we may interchange summation and integration, justied by absolute
convergence. Thus we obtain

s
2

_
s
2
_
(s) =
_

0
x
s
2
1

n=1
exp(n
2
x) dx.
We split the integral at x = 1 and get
(3.5)

s
2

_
s
2
_
(s) =
__
1
0
+
_

1
_
x
s
2
1
(x) dx,
where the series (x) is given in terms of the theta-function:
(x) :=

n=1
exp(n
2
x) =
1
2
((x) 1)
(since exp(n
2
x) = exp((n)
2
x) for any n N). In view of the func-
tional equation for the theta-function,

_
1
x
_
=
1
2
_

_
1
x
_
1
_
=

x(x) +
1
2
(

x 1),
110 Chapter 3 Hecke theory
we nd by the substitution x
1
x
that the rst integral in (3.5) is equal to
_

1
x

s
2
1

_
1
x
_
dx =
_

1
x

s+1
2
(x) dx +
1
s 1

1
s
.
Substituting this in (3.5) yields
(3.6)

s
2

_
s
2
_
(s) =
1
s(s 1)
+
_

1
_
x

s+1
2
+ x
s
2
1
_
(x) dx.
Since (x) exp(x), the last integral converges for all values of s, and
thus (3.6) holds, by analytic continuation, throughout the complex plane.
The right-hand side remains unchanged by s 1 s. This proves the
functional equation for zeta.
3.1.4. The case of Dirichlet L-functions. In a similar manner as
above one can prove the functional equation for Dirichlet L-functions L(s, )
with a primitive character mod q. Here we have to distinguish once again
the cases (1) = +1 and (1) = 1. In the rst case we nd
(3.7) (x, ) :=

nZ
(n) exp(n
2
x/q) =
()

qx

_
1
x
,
_
,
where the Gaussian sum () is given by (1.26) and satises
(3.8) ()() = (1)[()[
2
= (1)q;
this formula being valid for any primitive character mod q. The second
case, (1) = 1, is slightly more dicult. Here we make use of
(3.9)

(x, ) :=

nZ
(n)nexp(n
2
x/q) =
()
i

qx
3
2

_
1
x
,
_
,
The proofs of these functional equations rely on the Poisson summation for-
mula (3.2) and basic facts about primitive characters. Formulae (3.7) and
(3.9) lead by more or less the same method as for the zeta-function to
Theorem 3.4. Let be a primitive character mod q. Then, for any s C,
_
q

_s+
2

_
s +
2
_
L(s, ) =
()
i

q
_
q

_1+s
2

_
1 + s
2
_
L(1 s, ),
where :=
1
2
(1 (1)).
The Davenport-Heilbronn zeta-function is given by
L(s) =
1 i
2
L(s, ) +
1 + i
2
L(s, ),
where
:=
_
10 2

5 2

5 1
Section 3.1 The functional equation 111
and is the character mod 5 with (2) = i. It is an easy consequence of The-
orem 3.4 that the Davenport-Heilbronn zeta-function satises the functional
equation
_
5

_s
2

_
s + 1
2
_
L(s) =
_
5

_1s
2

_
1
s
2
_
L(1 s).
Davenport & Heilbronn [43] introduced this function as an example for a
Dirichlet series having innitely many zeros on the critical line and also inn-
itely many zeros in the half-plane > 1 in spite of satisfying a Riemann-type
functional equation. The localization of these zeros is not too easy (see also
[200]). However, following Balanzario [11] we can give another examples:
consider of a Dirichlet series satisfying a Riemann-type functional equation
for which the analogue of the Riemann hypothesis does not hold. Consider
the following functions with 5-periodic Dirichlet coecients:
(1 + 5
1
2
s
)(s) = 1 +
1
2
s
+
1
3
s
+
1
4
s
+
1 +

5
5
s
+ . . . ,
L(s, ) = 1
1
2
s

1
3
s
+
1
4
s
+
0
5
s
+ . . . ,
where is the character mod 5 with (2) = 1. Both functions satisfy the
same functional equation,
(3.10) F(s) = 5
1
2
s
2(2)
s1
(1 s) sin
s
2
F(1 s)
(see (2.6) and (2.7)). Now let z be any complex number, then the Dirichlet
series
L(z, )(1 + 5
1
2
s
)(s) L(s, )(1 + 5
1
2
z
)(z)
vanishes for s = z, satises the functional equation (3.10), has for > 1
a Dirichlet series expansion, and, obviously, this function is not identically
vanishing. Clearly, this example can easily be generalized (see [11]). We keep
in mind that a functional equation is not sucient for having all complex zeros
on a straight line! Is it the Euler product which forces the nontrivial to lie
on the critical line?
An alternative approach toward the functional equation for Dirichlet L-
functions uses another interesting class of Dirichlet series which do not have
an Euler product in general. In the following section we shall briey introduce
3.1.5. Hurwitz zeta-functions. For > 1, the Hurwitz zeta-function
is given by
(s, ) =

m=0
1
(m+ )
s
,
where is a parameter from the interval (0, 1]. The Hurwitz zeta-function
can be continued analytically to the whole complex plane except for a simple
112 Chapter 3 Hecke theory
pole at s = 1 with residue 1. Also these Dirichlet series satisfy some kind of
functional equation. One can show that
(3.11) (s, ) =
2(1 s)
(2)
1s
_

n=1
cos 2n
n
1s
+ cos
s
2

n=1
sin 2n
n
1s
_
;
this formula is valid for < 0 (in view of the innite series on the right-hand
side).
If is rational, =
a
q
with coprime a and q, say, then we have
(3.12)
_
s,
a
q
_
=
q
s
(q)

mod q
(a)L(s, )
and
(3.13) L(s, ) =
1
q
s
q

a=1
(a)
_
s,
a
q
_
.
From these identities one can deduce the functional equation for Dirichlet
L-functions from the one for Hurwitz zeta-functions and vice versa (if the
parameter is rational).
Hurwitz zeta-functions are of special interest with respect to Riemanns hy-
pothesis. For
1
2
, 1 the Hurwitz zeta-function is related to the Riemann
zeta-function:
(3.14) (s, 1) = (s) and
_
s,
1
2
_
= (2
s
1)(s).
However, besides =
1
2
, 1 there are no identities of this type; more precisely,
in Steuding [190] it was proved that (s, )/(s) is entire if and only if
=
1
2
or 1. The distribution of zeros of (s, ) as a function of s depends
drastically on the parameter and this is very interesting as we shall briey
explain. For instance, the Hurwitz zeta-function given by (3.14) vanishes
for s =
2ik
log 2
, k Z, and all other non-real zeros are expected to lie on the
critical line =
1
2
(by RH). This example is somehow special. It is known
that for any
1
2
<
1
<
2
< 1 and any transcendental or rational ,=
1
2
, 1 the
function (s, ) has more than cT zeros in the rectangle
1

2
, [t[ T,
where c is a positive constant depending on
1
,
2
and (actually, this is
a consequence of the universality property for the Hurwitz zeta-function;
see Garunkstis & Laurincikas [54] or Karatsuba & Voronin [112]). This
behaviour is also expected to be true for algebraic irrational (see Garunkstis
[53]). Denote by

+i

the nontrivial zeros of (s, ) (these nontrivial


zeros are dened in a similar way as the ones for (s); for short: apart
from nitely many exceptions they have a non-negative real part). However,
Section 3.1 The functional equation 113
Garunkstis & Steuding [55] (see also [54]) showed that
lim
T
2
T

||T
_

1
2
_
= log .
Thus, the nontrivial zeros of Hurwitz zeta-function weighted by their distance
from =
1
2
have a tendency to lie to the left of the critical line and so any
reasonable analogue of the Riemann Hypothesis for (s, ) fails for generic
,=
1
2
, 1.
0
0.5
1
x
90
100
110 y
0.6
0.8
1
a
0
0.5
1
x
90
100
110 y
Figure 1. Trajectories of several zeros of (s, ), 1/2
1; the 30-th zero of (s) = (s, 1) is plotted in green, the 35-th
in pink.
Now we want to study the zeros of (s, ) as a function of . By partial
summation,
(s, ) =
1

s
+
1
(1 + )
s
+
1
s 1
_
3
2
+
_
1s
+ s
_

3
2
1
2
u
(u + )
s+1
du,
valid for > 0, where u denotes the fractional part of a real number u
(see Karatsuba & Voronin [112]). The integral converges uniformly for s
from any compact subset of the half-plane > 0 and arbitrary . Hence,
(s, ) is a continuous function in the variable s ,= 1 and the parameter
and, in particular, the zeros depend continuously on . Now lets assume
Riemanns hypothesis for a short while and follow some idea from Garunkstis
& Steuding [56]. By (3.14), for any T and any > 0, there exists a positive
constant c = c(T, ) such that all nontrivial zeros

+i

of all Hurwitz
zeta-functions (s, ) with [
1
2
[ c, which have imaginary part [

[ T,
114 Chapter 3 Hecke theory
satisfy either [


1
2
[ or [

0[ . This scenario is illustrated in


Figure 1.
1
3.1.6. Further proofs of the functional equation. Riemann also
gave a second proof of the functional equation by using. The starting point
is (3.4). From this formula we easily deduce, for > 1,
(s)

n=1
1
n
s
=
_

0
x
s1

n=1
exp(nx) dx.
The sum on the right-hand side is a geometric series and thus we arrive at
the integral representation
(3.15) (s) =
1
(s)
_

0
x
s1
exp(x) 1
dx.
From this one can derive the formula
(s) =
exp(is)(1 s)
2i
_
C
z
s1
exp(z) 1
dz,
where ( is the contour which starts ay innity on the positive real axis,
encircles the origen once in the positive direction, excluding the points s =
2i, 4i, . . ., and then returns to innity. The sum of the residues at the
points s = 2in for n N is
4i exp(is) sin
s
2

n=1
(2n)
s1
= 4i exp(is) sin
s
2
(2)
s1
(1 s)
and this is already half of the proof (see (2.6) and (2.7)). This approach
is related to the proof of the approximate functional equation 2.8. It also
applies to Hurwitz zeta-functions and their functional equation as well to
studies on the values of the zeta-function at the negative integers (see the
following section).
Many further proofs of the functional equation were discovered; some of
them can be found in Titchmarsh [200]. However, there is one which has
to be mentioned explicitly since this approach has found several important
applications and generalizations, in particular, in algebraic number theory.
In his doctoral thesis from 1950 (see also [197]), Tate started to apply
harmonic analysis to local elds (in particular, Poissons summation formula).
He introduced integration techniques on the ring of ideles of a number eld
and succeeded in isolating and identifying the contributions to the functional
equation from each of the ramied prime ideals. In the simplest case his
method gives a proof of the functional equation for the Riemann zeta-function
1
We would like to thank Michael Trott for the MATHEMATICA notebook for
M. Trott, Zeros of the Generalized Riemann Zeta Function as a Function of
a, Background image in graphics gallery, in Wolfram [216]; see also the webpage
http://documents.wolfram.com/v4/MainBook/G.2.22.html.
Section 3.1 The functional equation 115
which uses only local information. Recall that number elds and function
elds of curves over nite elds are called global elds and the completions
of a global eld with discrete valuation and nite residue eld are said to
be local; for instance, the p-adic elds constructed from the eld of rational
numbers Q and R are local. The local elds contain deep information of the
underlying global eld. For example, Hasse [82] proved that a quadratic form
with rational coecients represents a given number over the global eld Q if
and only if it does in all local elds Q
p
, p , i.e, the p-adic elds Q
p
for
each prime p and the eld of real numbers R = Q

(this notation is standard


in the theory of valuations). The so-called local-global principle is the idea
of putting together information from all local elds to get information in the
corresponding global eld. Roughly speaking, Tate has given a dissection of
the functional equation into a family of local functional equations for each
p corresponding to the Euler factors for each prime p in the Euler
product for (s) in addition with the contribution for the innite prime, that
is the Gamma-factor. However, Tates method gives more; for example, the
easiest proof for the functional equation for Dedekind zeta-functions. For
more details on Tates thesis and its generalizations we refer to Tate [197]
and Swinnerton-Dyer [195].
3.1.7. The Phragmen-Lindelof principle. Functional equations of
the Riemann-type contain important information on the order of growth.
In order to deduce this information we shall use a kind of maximum principle
for unbounded regions, the theorem of Phragmen-Lindel of:
Lemma 3.5. Let f(s) be analytic in the strip
1

2
with f(s)
exp([t[). If
f(
1
+ it) [t[
c
1
and f(
2
+ it) [t[
c
2
,
then f(s) [t[
c()
uniformly in
1

2
, where c() is linear with
c(
1
) = c
1
and c(
2
) = c
2
.
A proof can be found in the paper of Phragmen & Lindel of [135] or, for
example, in Titchmarsh [199]. Note that there are counterexamples if the
growth condition f(s) exp([t[) is not fullled.
We illustrate the so-called Phragmen-Lindel of principle with an easy appli-
cation to the zeta-function. We dene
() = limsup
t
log [( + it)[
log t
.
One can show that () is a convex function of . Taking into account the
absolute convergence of the dening Dirichlet series we immediately see that
() = 0 for > 1. The order of growth in the half-plane left of the critical
116 Chapter 3 Hecke theory
strip is ruled by the functional equation which we may rewrite as
(s) = (s)(1 s),
where (s) is given by (2.7). Applying Stirlings formula (1.53) we get, for
t 1,
[( + it)[
_
t
2
_1
2

,
uniformly in . From this we deduce that
( + it) [t[
1
2

[(1 + it)[,
uniformly in , as [t[ . This estimate implies now () =
1
2
for
< 0. For the calculation of () with 0 1 we apply the theorem of
Phragmen-Lindel of, Lemma 3.5. It follows that () is non-increasing and
convex downwards and we obtain explicit estimates from the estimates of
() for outside of the critical strip. Altogether, we obtain
()
_
_
_
0 if > 1,
1
2
(1 ) if 0 1,
1
2
if < 0.
In view of the functional equation, resp. the convexity of (), the value for
=
1
2
is essential. In particular, we obtain (
1
2
)
1
4
or, equivalently,

_
1
2
+ it
_
t
1
4
+
as [t[ , valid for any positive . Recall that the approximate functional
equation gave only the exponent
1
2
(see (2.8)). However, this is not the
best estimate for the zeta-function on the critical line. The exponent
1
4
is
called the convexity bound and there is a long list of improvements. At
the moment, Huxley [94] holds the record with the exponent
32
205
+ . This
remarkable estimate was obtained with a dierent method (namely, estimates
for exponential sums) but is still far away from the exponent predicted by
the Lindel of hypothesis.
There are more advanced applications of the Phragmen-Lindel of princi-
ple; however, often with respect to the modulus of some characters or other
arithmetic objects. Recently, new methods for breaking the corresponding
convexity bounds in these arithmetic cases as well as unexpected applications
were found; see Iwaniec & Sarnak [100].
Integrals like (3.5) are called Mellin transforms (and we already met them in
Chapter 1.6.7). Here we want to derive the Mellin inversion formula (3.16). If g(s)
Section 3.2 The zeta-function at the integers 117
is analytic in some right half-plane, then its inverse Mellin transform is given by
f(x) =
1
2i
_
+i
i
g(s)x
s
ds
for positive values of x such that the integral converges absolutely. By contour
integration, it turns out that the integral is independent of .
Exercise 72. Show that
(3.16) g(s) =
_

0
f(x)x
s1
dx f(x) =
1
2i
_
+i
i
g(s)x
s
ds.
Hint: let x = exp(z), s = + 2iy and rewrite the integrals according g(s) =
g

(y), f(x) = f

(z) exp(z); apply Fourier analysis.


Exercise 73. i) Show (3.8).
ii) Prove the identities (3.7) and (3.9).
Hint: for the second formula one may rst prove

nZ
(n + ) exp((n + )
2
/x) = ix
3
2

nZ
nexp(xn
2
+ 2in).
iii) Deduce Theorem 3.4.
Exercise 74. * i) Prove identity (3.11). Start with the identity
(s, ) = exp(is)(1 s)
1
2i
_
C
z
s1
exp((1 )z)
exp(z) 1
dz,
where ( is the positively oriented contour consisting of the positive real part of the
axis from + to 0, enclosing the point z = 0 by a circle of radius r (0, 2), and
returning to +.
Hint: consult Garunkstis & Laurincikas [54].
ii) Show the identities (3.12) and (3.13) and deduce from this representation and
(3.11) the functional equation for Dirichlet L-functions.
Exercise 75. Use the Phragmen-Lindelof principle to prove estimates for the order
of growth of Dirichlet L-functions and Hurwitz zeta-functions. What can you do
for Dedekind zeta-functions?
3.2. The zeta-function at the integers
It is remarkable that already Euler [51] had partial results toward the func-
tional equation for (s), namely, formulae for the values of (s) for integral
s and for half-integral s relating s with 1 s (see Ayoub [5]). Here we want
to sketch his contribution briey.
118 Chapter 3 Hecke theory
3.2.1. The positive integers. A famous problem in the 17/18th cen-
tury was the evaluation of (2) =

n=1
n
2
. This was solved by Euler in
1737 as follows: Comparing the product
(3.17)
sin z
z
=

n=
n=0
_
1
z
n
_
=

n=1
_
1
z
2

2
n
2
_
with the power series representation
sin z
z
= 1
z
2
3!
+
z
4
5!
. . . =

k=0
(1)
k
z
2k
(2k + 1)!
,
one obtains
(2) =

n=1
1
n
2
=

2
6
.
Eulers proof was much discussed by his contemporaries. First of all, it was
not clear whether sin z has no complex zeros; furthermore, the convergence of
(3.17) cannot be proved without complex analysis which was not developed
in those times. However, today Eulers argument is waterproof and might be
the easiest proof of all.
As Euler we want to compute more values of the zeta-function at the in-
tegers. For this purpose we have to introduce the Bernoulli numbers (in-
troduced by the Bernoullis, they are extraordinarily important in algebraic
number theory). The numbers B
n
are dened by the identity
z
exp z 1
=

n=0
B
n
z
n
n!
= 1
1
2
z +
1
12
z
2
. . . . (3.18)
The function
z
exp z1
+
z
2
is an even function. This and (3.18) imply that
B
2n+1
= 0 for n N. Hence, one nds

k=1
(1)
k
(2)
2k
(2k)!
B
2k
z
2k
= z cot(z) 1 = z
d
dz
log
sin(z)
z
.
Using the product representation (3.17), we nd
z

n=1
d
dz
log
_
1
z
2
n
2
_
= 2

k=1
(2k)z
2k
.
Comparing the coecients, we arrive at
Theorem 3.6. For k N,
(2k) = (1)
k1
(2)
2k
2(2k)!
B
2k
.
Nearly nothing is known about the values of zeta at the positive odd integers;
in 1979, Apery [1] proved that (3) is irrational but the arithmetic character
of (5) is still unknown.
Section 3.2 The zeta-function at the integers 119
3.2.2. The negative integers. Now we study the values at the negative
integers. Here Euler found
Theorem 3.7. For n N,
(0) =
1
2
and (n) =
B
n+1
n + 1
.
It is remarkable that Euler found this formula since he considered (s) as a
function of a real variable s and for that purpose he had to pass behind the
pole at s = 1 (which is in principle only possible by analytic continuation
leaving the real axis). Eulers argument was as follows: we have, for m N
0
,
(3.19) 1
m
2
m
+ 3
m
. . . = (1 2
m+1
)(m),
and
x
m
2
m
x
2
+ 3
m
x
3
. . . =
_
x
d
dx
_
m
x
1 + x
.
Using the latter formula with x = exp(2iw) we get
(1 2
m+1
)(m) = (2i)
m
_
d
dw
_
m
exp(2iw)
1 + exp(2iw)

w=0
.
By (3.18) this leads to the formula of Theorem 3.7. Eulers proof needs a
modied notion of convergence this is obvious with respect to (3.19); using
summability arguments one can make his approach waterproof. Here we shall
use an idea of Riemann to prove Theorem 3.7.
Proof. Let > 1. We start with the integral (3.15) and deduce via (3.18)
(s)(s) =
_

0
z
s1
exp(z) 1
dz
=
_

0
z
s2
exp(z)
n

k=0
(1)
k
B
k
z
k
k!
dz (3.20)
+
_

0
z
s2
exp(z)

k=n+1
(1)
k
B
k
z
k
k!
dz.
The second integral is bounded by ( + n) and thus convergent and
analytic for > n. The rst integral is equal to
(s 1) +
1
2
(s) +
n

k=2
B
k
k!
(s + k 1);
120 Chapter 3 Hecke theory
hence it is meromorphic in the whole of C. By the functional equation of the
Gamma-function, we deduce another analytic continuation of (s) to > n:
(s) =
1
s 1
+
1
2
+
n

k=2
B
k
k!
s(s + 1) . . . (s + k 2) (3.21)
+
1
(s)
_

0
z
s2
exp(z)

k=n+1
(1)
k
B
k
z
k
k!
dz.
In particular, we nd at the poles of (s)
(1 n) =
1
n
+
1
2
+
n

k=2
B
k
k!
(1 n)(n) . . . (k n 1)
=
1
n
n

k=0
_
n
k
_
B
k
.
By the recursion formula for the Bernoulli numbers (an exercise left for the
reader), this implies Eulers formula.
Euler [51] was aware about the correspondence between the values of (2k)
and (1 2k) for integers k:
(2k) = (1)
k1
(2)
2k
2(2k)!
B
2k
=
(2)
2k
2 cos k(2k)
(1 2k);
this is indeed the functional equation in the form (1.58) for s = 2k. For more
details on Eulers work on the zeta-function we refer to Ayoub [5].
3.2.3. A p-adic zeta-function. The value-distribution of the zeta-
function for integer values allows the construction of a p-adic zeta-function

p
(s) which interpolates (s) (as a matter of fact, all values (1n) for n N
0
are rational). This was rst observed by Kubota & Leopoldt [122]; their con-
struction implies remarkable and surprising facts on Bernoulli numbers, e.g.,
the old von Staudt-Clausen congruences
B
m
+

p
m0 mod (p1)
1
p
Z,
valid for any positive even integer m, and the Kummer congruences
B
m
m

B
n
n
mod p if m n , 0 mod (p 1).
The approximation by a p-adic zeta-function is as follows:

p
(1 m) = (1 m)(1 p
m1
) if m 0 mod p 1;
it should be noticed that the factor 1p
m1
on the right-hand side is exactly
the Euler factor of (1m) at p. Generalizations to Dirichlet L-functions are
important with respect to the p-adic analogue of the class number formula,
Section 3.3 Hamburgers theorem 121
and elliptic analogues of the p-adic zeta-function are a major ingredient in
Wiles solution of Fermats last theorem. (We refer the interested reader to
Koblitz [116].)
Exercise 76. For n N, prove that B
2n+1
= 0 and the recursion formula
B
n
= (1)
n
n

k=0
_
n
k
_
B
k
.
Hint: rst, show that

n=0
n

k=0
_
n
k
_
B
k
z
n
n!
=
z
exp(z) 1
.
Exercise 77. Using the dierential equation cot

(z) = 1 cot z, prove the re-


cursion formula
_
n +
1
2
_
(2n) =

k+=n
(2k)(2).
Exercise 78. Let k, n N. Prove that (s) = 1 + O(2

) for , and
deduce (via Theorem 3.6 and Stirlings formula n! =

2n
_
n
e
_
n
_
1 + O
_
1
n
__
) the
asymptotic formula
B
2k
= (1)
k1
4

k
_
k
e
_
2k
_
1 + O
_
1
k
__
.
Exercise 79. Prove that the probability that the probability that n randomly chosen
positive integers m
1
, . . . , m
n
are coprime is equal to
Prob(gcd(m
1
, . . . , m
n
) = 1) =

p
_
1
1
p
n
_
=
1
(n)
.
3.3. Hamburgers theorem
In 1921, Hamburger [74] proved that the Riemann zeta-function is charac-
terized by its functional equation.
Theorem 3.8. Let G(s) be an entire function of nite order, P(s) a polyno-
mial, and suppose that
f(s) :=
G(s)
P(s)
=

n=1
a(n)
n
s
,
the series being absolutely convergent for > 1. Assume that
(3.22)

s
2

_
s
2
_
f(s) =

1s
2

_
1 s
2
_
g(1 s),
122 Chapter 3 Hecke theory
where
g(1 s) =

n=1
b(n)
n
1s
,
the series being absolutely convergent for < for some positive constant
. Then f(s) = c(s), where c is a constant.
We shall give here a simplied proof due to Siegel [181].
Proof. By (3.4) we nd, for x > 0,
(x) :=
1
2i
_
2+i
2i
f(s)
_
s
2
_
(x)

s
2
ds
=

n=1
a(n)
1
2i
_
2+i
2i

_
s
2
_
(n
2
x)

s
2
ds
= 2

n=1
a(n) exp(n
2
x). (3.23)
In view of (3.22) we also have
(x) =
1
2i
_
2+i
2i
g(1 s)
_
1 s
2
_

s1
2
x

s
2
ds.
Next we move the line of integration from the line = 2 to = 1 .
Obviously, f(s) is bounded on = 2 and g(1 s) is bounded on = 1 .
By Stirlings formula (1.53),

_
s
2
_

_
1s
2
_ [t[

1
2
as [t[ . Thus, g(1 s) [t[
3
2
on = 2 as [t[ , and, justied by
the Phragmen-Lindel of principle (see Lemma 3.5), we can apply Cauchys
theorem. It follows that
(3.24) (x) =
1
2i
_
1+i
1i
g(1 s)
_
1 s
2
_

s1
2
x

s
2
ds +
k

j=1
R
j
,
where R
1
, . . . , R
k
are the residues at the poles, say s
1
, . . . , s
k
. It is easily seen
that the sum of residues is of the form
k

j=1
R
j
=
k

j=1
x

s
j
2
P
j
(log x) =: R(x),
Section 3.3 Hamburgers theorem 123
where the P
j
(log x) are polynomials in log x. We rewrite (3.24) and nd as
above
(x) =
1

n=1
b(n)
1
2i
_
1+i
1i

_
1 s
2
__
n
2
x
_s1
2
ds + R(x)
=
2

n=1
b(n) exp(n
2
/x) + R(x).
Comparing with (3.23), we arrive at

n=1
a(n) exp(n
2
x)
1
2
R(x) =
1

n=1
b(n) exp(n
2
/x).
Multiplying with exp(t
2
x) with t > 0 and integrating over (0, ) with
respect to x, we get
t

n=1
a(n)
(t
2
+ n
2
)

t
2
_

0
R(x) exp(t
2
x) dx =

n=1
b(n) exp(2nt).
The integral can be evaluated as a nite sum of terms of the form
Q(t; a, b) :=
_

0
x
a
(log x)
b
exp(t
2
x) dx,
where the bs are integers and Re a > 1; thus, Q(t; a, b) is a sum of terms
of the form t

(log t)

. Hence,

n=1
a(n)
_
1
t in

1
t + in
_
t

2
Q(t; a, b) =

n=1
b(n) exp(2nt).
The left-hand side is a meromorphic function in t with poles at t = in
for n N. The right-hand side is periodic with period i and, by analytic
continuation, the function on the left-hand side is also periodic. Hence, the
residues at in and i(n + 1) are equal. Thus, a(n) = a(n + 1) for all n N
and Hamburgers theorem is proved.
As Hecke pointed out in 1936, this result is better understood in the context
of modular forms. However, before we start with Heckes theory we have to
recall some basic facts about modular forms.
Exercise 80. * Try to nd an analogue of Hamburgers theorem for Dirichlet L-
functions! Note that there are many L(s, ) satisfying the same functional equation;
what do they have in common?
124 Chapter 3 Hecke theory
3.4. Modular forms
Modular forms are holomorphic functions of the upper half-plane which are
almost invariant under operations of the modular group (resp. subgroups).
In the recent past they have been proven to be of greatest importance in
modern number theory, e.g., in Wiles proof of Fermats last theorem. For
the details of the theory we refer to Iwaniec [99], Koblitz [117], and Miyake
[146].
3.4.1. Eisenstein series and the discriminant. Recall that the set
of all 2 2 matrices with integral entries and determinant 1 forms a group,
the so-called special linear group over Z, denoted by SL
2
(Z). This group is
generated by the two matrices
_
1
0
1
1
_
and
_
0
1
1
0
_
.
In the sequel we shall study the transformation properties of holomorphic
functions of the upper half-plane
H := z C : Imz > 0
under the action of SL
2
(Z)-matrices as fractional linear transformations
z Mz :=
az + b
cz + d
for M :=
_
a b
c d
_
SL
2
(Z).
We start with an example.
For z H and a xed positive even integer k > 2, the Eisenstein series of
weight k is dened by
G
k
(z) =
(k 1)!
2(2i)
k

m,nZ
(m,n)=(0,0)
1
(mz + n)
k
(the condition k > 2 is needed to guarantee absolute convergence). What
happens with G
k
(z) under transformations of the special linear group? The
action of M = (
a
c
b
d
) SL
2
(Z) on this function replaces (m, n) by (am +
cn, bm+ dn) and therefore permutes the terms of the sum. We obtain
(3.25) G
k
_
az + b
cz + d
_
= (cz + d)
k
G
k
(z).
We want to derive a more convenient expression for G
k
(z). Recall the Lips-
chitz formula
(3.26)

nZ
1
(z + n)
k
=
(2i)
k
(k 1)!

d=1
d
k1
exp(2idz),
Section 3.4 Modular forms 125
which is valid for k 2 and z H. Using this formula, we nd by splitting
the G
k
-dening sum into terms with m = 0 and the terms with m ,= 0 that
G
k
(z) =
(k 1)!
(2i)
k

n=1
1
n
k
+

m=1
_
(k 1)!
(2i)
k

nZ
1
(mz + n)
k
_
= (1)
k
2
(k 1)!
(2)
k
(k) +

m=1

d=1
d
k1
exp(2idmz).
In view of the values of the zeta-function at the integers, Theorems 3.6 and
3.7, we get the Fourier series expansion
(3.27) G
k
(z) =
B
k
2k
+

n=1

k1
(n) exp(2inz);
here
k1
(n) denotes the sum of divisors of n in the power k 1. In fact,
this representation is the starting point for the approach of Bump & Beineke
[16] to power moments of (s); the hope is that spectral theory for Eisenstein
series may be used to handle higher moments of the zeta-function.
A further example for the objects we want to study is the so-called discrim-
inant, for z H, dened by
(3.28) (z) =
(2)
12
1728
_
240G
4
(z))
3
(504G
6
(z))
2
_
(the name discriminant comes from the theory of elliptic curves). In view of
(3.25) it follows that

_
az + b
cz + d
_
=
(2)
12
1728
_
_
240G
4
_
az + b
cz + d
__
3

_
504G
6
_
az + b
cz + d
__
2
_
= (cz + d)
12
(2)
12
1728
_
(240G
4
(z))
3
(504G
6
(z))
2
_
= (cz + d)
12
(z)
for all M = (
a
c
b
d
) SL
2
(Z). One can prove the following representation as
an innite product:
(z) = (2)
12
exp(2iz)

n=1
(1 exp(2inz))
24
(a proof can be found in Koblitz [117]). The Fourier series expansion takes
the form
(z) = (2)
12

n=1
(n) exp(2inz)
Ramanujan [173] conjectured that the coecients (n) are multiplicative
and satisfy the estimate [(p)[ 2p
11
2
. The multiplicativity was proved by
126 Chapter 3 Hecke theory
Mordell [150], in particular by the beautiful formula
(m)(n) =

d|(m,n)
d
11

_
mn
d
2
_
.
The estimate was shown by Deligne in a more general setting (see (3.32)
below).
3.4.2. Denitions and basic facts. The functions G
k
(z) and (z)
have remarkable transformation properties with respect to SL
2
(Z). They are
examples of modular forms to the full modular group. Here comes the general
denition.
The group := SL
2
(Z) is called the modular group. We will also consider
subgroups. For a non-negative integer k and a positive integer N, we dene

0
(N) =
__
a b
c d
_
SL
2
(Z) : c 0 mod N
_
;
clearly, this denes a subgroup of the full modular group =
0
(1) and is
called Hecke subgroup of level N or congruence subgroup mod N.
A holomorphic function f on H is said to be a modular form of weight k
for
0
(N) if
(3.29) f
_
az + b
cz + d
_
= (cz + d)
k
f(z) for all
_
a b
c d
_

0
(N),
and f(z) is holomorphic at innity, i.e., f(z) has a Fourier series expansion
(3.30) f(z) =

n=0
a(n) exp(2inz).
A modular form f is said to be a cusp form if f vanishes at all cusps or,
equivalently, if
z = x + iy y
k
[f(z)[
2
is bounded on H. Then we have a(0) = 0 in the Fourier expansion (3.30) for
f(z).
The modular forms on
0
(N) of weight k form a nite dimensional complex
vector space, denoted by M
k
(
0
(N)); analogously, also the set of all cusp
forms on
0
(N) of weight k form is a nite dimensional complex vector space,
denoted by S
k
(
0
(N)). For instance, the Eisenstein series G
k
(z) dened by
(3.27) with k 4 are modular forms of weight k for the full modular group:
G
k
M
k
().
One can show that the space of all modular forms to the full modular
groups is the direct sum of all spaces M
k
() with non-negative weights k,
Section 3.4 Modular forms 127
where M
k
() has dimension
dimM
k
() =
_
_
_
[k/12] if k 2 mod 12,
1 + [k/12] if k 0, 4, 6, 8, 10 mod 12,
0 otherwise;
the case of odd k follows immediately from the observation that any solution
of (3.29) with odd k vanishes identically. Moreover, one has the decomposi-
tion
M
k
() = G
k
C S
k
()
if k > 2, and every modular form for the full modular group is a polynomial
in the Eisenstein series G
4
and G
6
.
On the space of cusp forms one can introduce an inner product, the Pe-
tersson inner product, dened by
f, g :=
_
H/
0
(N)
f(z)g(z)y
k
dxdy
y
2
for f, g S
k
(
0
(N)). Suppose that M[N. If f S
k
(
0
(M)) and dM[N,
then z f(dz) is a cusp form on
0
(N) of weight k too. The forms which
may be obtained in this way from divisors M of the level N with M ,= N span
a subspace S
old
k
(
0
(N)), called the space of oldforms. Its orthogonal comple-
ment with respect to the Petersson inner product is denoted S
new
k
(
0
(N)).
For n N dene the Hecke operator T(n) by
T(n)f :=
1
n

ad=n
a
k

0b<d
f
_
az + b
d
_
for f S
k
(
0
(N)). These operators are multiplicative and encode plenty
of arithmetic information of modular forms. The theory of Hecke operators
implies the existence of an orthogonal basis of S
new
k
(
0
(N)) made of eigen-
values of the T(n) for n coprime with N. By the multiplicity-one principle
of Atkin & Lehner [4], the elements f of this basis are in fact eigenvalues of
all T(n), i.e., there exist complex numbers
f
(n) for which T(n)f =
f
(n)f
and a(n) =
f
(n)a(1) for all n N. Furthermore, it follows that the rst
Fourier coecient a(1) of such an f is non-zero. A newform f is dened to
be an element of this basis normalized to have a(1) = 1. The newforms form
a nite set which is an orthogonal basis of the space S
new
k
(
0
(N)).
To give an example, the discriminant (z) given by (3.28) (sometimes also
called Ramanujans cusp form) is a cusp form of weight 12 for the full modular
group, and hence, after normalization, a newform of level 1.
3.4.3. Dirichlet series associated with modular forms. In the
1930s Hecke [88] started investigations on modular forms and the associated
128 Chapter 3 Hecke theory
Dirichlet series (we already mentioned Hecke operators). Given a modular
form f with Fourier expansion (3.30), we may dene the Dirichlet series
L(s, f) =

n=1
a(n)
n
s
; (3.31)
note that here the Fourier coecient a(0) does not appear. By classic esti-
mates for the Fourier coecients of f this series converges in some half-plane
and its properties will be the main theme in the following section. How-
ever, for our later purpose we have to consider the case of a newforms more
detailed.
Suppose that f is a newform of weight k. In this case Deligne [45] proved
for the Fourier coecients the estimate
[a(n)[ n
k1
2
d(n), (3.32)
where d(n) =

d|n
1 is the divisor function. In view of the classic bound
d(n) n

it follows that the series (3.31) converges absolutely for >


k+1
2
.
By the multiplicativity of the Fourier coecients it turns out that, in the
half-plane of absolute convergence, there is an Euler product representation
for the associated Dirichlet series:
L(s, f) =

p|N
_
1
a(p)
p
s
_
1

pN
_
1
a(p)
p
s
+
1
p
2s+1k
_
1
.
Hecke [88] resp. Atkin & Lehner [4] (for newforms) proved that the L-
function L(s, f) has an analytic continuation to an entire function and satis-
es the functional equation
N
s
2
(2)
s
(s)L(s, f)
= (1)
k
2
N
ks
2
(2)
sk
(k s)L(k s, f), (3.33)
where = 1 is the Atkin-Lehner eigenvalue of the Atkin-Lehner involution
_
0 N
1 0
_
on S
k
(
0
(N)).
Exercise 81. * Prove the Lipschitz formula (3.26).
Hint: apply the Poisson summation formula (Theorem 3.2) to the function f(x) =
(x + iy)
k
, where y is a positive real number and k 2 an integer.
Exercise 82. Show that Eisenstein series G
k
(z) converge absolutely if k > 2.
Exercise 83. Prove that the j-function dened by
j(z) =
(240G
3
(z))
3
(z)
for z H
is a modular function (i.e., a modular form of weight k = 0 to the full modular
group).
Section 3.5 Heckes converse theorem 129
3.5. Heckes converse theorem
In 1936, Hecke [88] proved a bijection between modular forms and Dirich-
let series satisfying a Riemann-type functional equation; this includes Ham-
burgers theorem as a special case. Moreover, it connects the theory of mod-
ular forms with the theory of Dirichlet series. In the sequel we follow Oggs
monograph [163].
3.5.1. The converse theorem. Let be a positive real number and
dene the Hecke group G() as the subgroup of SL
2
(R) given by
G() =
__
1
0

1
_
,
_
0
1
1
0
__
.
The case of the full modular group is = G(1). Thus G() is generated by
the fractional linear transformations
z z + and z
1
z
.
Extending the notion of modular forms of the modular group or its subgroups,
Hecke introduced G()-modular forms as follows. A modular form of G()
of weight k and multiplier 1 is a holomorphic function f : H C
satisfying
f(z + ) = f(z) and f
_

1
z
_
= (iz)
k
f(z)
and having a Fourier expansion
(3.34) f(z) =

n=0
a(n) exp(2inz/)
for all z H (this should be compared with (3.29)); this representation
includes the -periodicity and shows that f(z) is holomorphic at . The
complex vector space of such modular forms which in addition satisfy the
growth condition a(n) = O(n
c
) for some constant c is denoted by M
0
(, k, ).
A modular form of M
0
(, k, ) is a cusp form if a(0) = 0.
Hecke proved a one-to-one correspondence between the elements of
M
0
(, k, ) and Dirichlet series satisfying a Riemann-type functional equa-
tion plus some growth conditions. In fact, his theorem is even more general
since it also contains the case of functional equations relating two dierent
functions (Dirichlet series or modular forms):
Theorem 3.9. Let and k be xed positive real numbers. Given two se-
quences a(n)
nN
0
and b(n)
nN
0
of complex numbers satisfying
a(n), b(n) n
c
as n
130 Chapter 3 Hecke theory
for some positive constant c, we dene
(s) =

n=1
a(n)
n
s
and (s) =

n=1
b(n)
n
s
,
as well as
(s) =
_

2
_
s
(s)(s) and (s) =
_

2
_
s
(s)(s).
Furthermore, let
f(z) =

n=0
a(n) exp(2inz/) and g(z) =

n=0
b(n) exp(2inz/).
Then the functions (s) and (s) are analytic in the half-plane > c +
1, while f(z) and g(z) are analytic in the upper half-plane satisfying the
boundary condition
(3.35) f(x + iy), g(x + iy) y
c1
as y 0 + .
Furthermore, the following statements are equivalent:
(i) The function
(s) +
a(0)
s
+
b(0)
k s
,
is entire and bounded on every vertical strip and satises the func-
tional equation
(s) = (k s);
(ii) For any z H,
f(z) =
_
i
z
_
k
g
_
1
z
_
.
Following Hecke we assign to any Dirichlet series (s) satisfying the conditions
of Theorem 3.9 its signature , k, by putting a(n) = b(n). Heckes theorem
includes the case of the zeta-function as (s) = (2s), f being the theta-
function (x) and k =
1
2
(in particular, we see that the theta-function is a
modular form); the signature of zeta is 2,
1
2
, 1.
3.5.2. Proof of the converse theorem. As in Riemanns proof of the
functional equation or in Siegels proof of Hamburgers theorem we shall
consider Mellin transforms and use the Mellin inversion formula (resp. the
Poisson summation formula) to prove the equivalence in Theorem 3.9.
However, rst of all, we observe that the statement concerning the conver-
gence of the Dirichlet series is trivial (by standard arguments as in the case of
(s)). In order to derive the holomorphy and the boundary condition (3.35)
Section 3.5 Heckes converse theorem 131
for the Fourier series it suces to consider the function f(z) only. Since, by
Stirlings formula (1.53),
(1)
n
_
c 1
n
_
=
(c + 1 + n)
(c + 1)(n + 1)
c
1
n
c
with some positive constant c
1
, the Fourier series for f(x +iy) is dominated
term-by-term by

n=0
(1)
n
_
c 1
n
_
exp(2ny/) = (1 exp(2y/))
(c+1)
y
c1
.
Conversely, given the boundary condition, we can bound the Fourier coe-
cients a(n) using their integral representation with y =
1
n
,
a(n) =
_
1
0
f
_
x +
i
n
_
exp
_
2in
_
x +
i
n
_
/
_
dy,
by O(n
c
exp(2/)).
It remains to show the equivalence of (i) and (ii). We start with the impli-
cation (ii)(i).
We note that, for suciently large ,
(s) =

n=1
a(n)
_

0
_

2
_
s
x
s1
exp(nx) dx
=

n=1
_

0
a(n)y
s1
exp(2ny/) dy
(as in the proof of the functional equation for (s)). Now interchanging
summation and integration (justied by absolute convergence), we get
(s) =
_

0

n=1
a(n)y
s1
exp(2ny/) dy
=
_

0
y
s1
(f(iy) a(0)) dy.
The integral is improper for y 0+ and y ; we consider the contri-
butions of the intervals (0, 1) and (1, ) separately. Since f(iy) a(0)
exp(cy) as y for some positive constant c, it follows that
_

1
y
s1
(f(iy) a(0)) dy
converges uniformly on vertical strips, and so it denes an entire function
which is bounded on vertical strips. For the integral taken over (0, 1) we
132 Chapter 3 Hecke theory
have to make use of (ii). We have
_
1
0
y
s1
(f(iy) a(0)) dy =
a(0)y
s
s

1
y=0
+
_

1
y
1s
f
_
i
y
_
dy
y
2
.
Now by (ii) we get
_
1
0
y
s1
(f(iy) a(0)) dy
=
a(0)
s
+
_

1
y
ks1
(g(iy) b(0)) dy
b(0)
k s
.
Hence,
(s) +
a(0)
s
+
b(0)
k s
=
_

1
_
y
s1
(f(iy) a(0)) + y
ks1
(g(iy) b(0))
_
dy
is an entire function bounded on vertical strips. Furthermore, we observe
that (i) holds.
Now we assume (i) and deduce (ii). We shall use the formula
(3.36) exp(x) =
1
2i
_
+i
i
x
s
(s) ds,
where > 0 and x > 0; this is the Mellin inversion of Eulers integral
representation of the Gamma-function. It follows that
(3.37) f(iy) a(0) =
1
2i
_
+i
i
y
s
(s) ds;
however, here we have to choose the abscissa > k such that the path of
integration lies inside the half-plane of absolute convergence for (s). We
shall move the path of integration over the origin to the left. Incorporating
the residues at s = 0 and s = k, we obtain
f(iy) a(0) =
1
2i
_
+i
i
y
s
(s) ds
+Res
s=0
+ Res
s=k
y
s
(s). (3.38)
In view of (i) we have
Res
s=0
y
s
(s) = a(0) and Res
s=k
y
s
(s) = b(0)y
k
.
Thus, we may replace (3.38) by
f(iy) b(0)y
k
=
1
2i
_
+i
i
y
s
(s) ds.
Section 3.5 Heckes converse theorem 133
Taking into account (i), we get
f(iy) b(0)y
k
=
1
2i
_
+i
i
y
s
(k s) ds
=

2i
_
k++i
k+i
y
(ks)
(s) ds,
by substituting s by k s. The right-hand side above is equal to
y
k
_
g
_
i
y
_
b(0)
_
(by the same argument as for (3.37)). This gives (ii) and the theorem is
proved. .
3.5.3. The arithmetical and topological character of Hecke
groups. The groups G() operate discontinuously as groups of fractional
linear transformations on H if and only if either > 2 or
=
m
:= 2 cos

m
with 3 m N .
The space M
0
(
m
, k, ) with
m
< 2 is non-trivial, i.e., ,= 0, if and only if
k = 4

m2
+ 1 for some positive integer . In this case
dimM
0
(
m
, k, ) = 1 +
_
+ ( 1)/2
m
_
.
The space of cusp forms is non-trivial if and only if dimM
0
(
m
, k, ) 2; in
view of the dimension formula this condition holds when k is suitably large.
For
m
1,

2,

3, 2 (i.e., m 3, 4, 6, ), the Hecke group G(


m
) can
be dened arithmetically and in these cases G(
m
) holds a structure compa-
rable to the full modular group := G(1). We shall say a few words about
the properties of Dirichlet series associated with modular forms to Hecke
groups G(). The situation is similar in the case of
m

2,

3, 2 since
then the groups G(
m
) are conjugate to index 2 extensions of the congruence
subgroups
0
(N) of levels N = 2, 3, 4, respectively. However, in these cases
only the newforms have a basis consisting of normalized eigenfunctions (for
the Hecke and Atkin-Lehner operators). More details can be found by Hecke
[88], Atkin & Lehner [4], and the monographs of Ogg [163] and Miyake [146].
To indicate the dierence between modular forms to Hecke groups G()
which can be arithmetically dened and those who cannot, we state a result
of Wolfart. In [215], he has shown that every space M
0
(
m
, k, ) with
m
/
1,

2,

3, 2 has a basis consisting of modular forms of type


(3.39) f(z) =

n=0
r(n)a
n
exp(2inz/),
134 Chapter 3 Hecke theory
where r(n) Q and a is transcendental; moreover, a depends only on the
space M
0
(
m
, k, ) and not on the modular form f. Clearly, the same state-
ment holds for cusp forms.
Exercise 84. Show in detail that Heckes converse theorem 3.9 contains the case
of the zeta-function.
Exercise 85. Prove Formula (3.36) by i) using the Mellin inversion formula
(3.16), and ii) by the calculus of residues.
Hint: for ii), note that the sum of residues of the integrand is equal to

m=0
Res
s=m
x
s
(s) =

m=0
(x)
m
m!
.
Exercise 86. What is the signature of a Dedekind zeta-function to an imaginary
quadratic number eld? What can you deduce from Heckes theorem for general
Dedekind zeta-functions?
Exercise 87. Prove that L
f
(s) with f being a basis element of the form (3.39)
has no Euler product representation.
3.6. Shimura-Taniyama-Wiles
The Shimura-Taniyama conjecture was rst stated by Shimura & Taniyama
in 1955. Roughly speaking, it states that for any elliptic curve dened over
Q, there is a modular form such that both objects have the same L-function.
In the 1980s, Frey observed that a counterexample to Fermats last theorem
would lead to a counterexample to the Shimura-Taniyama conjecture; this
was rigorously proved by Ribet soon after. The implication that the Shimura-
Taniyama conjecture implies the truth of Fermats last theorem relates two
deep open conjectures from rather dierent elds. Fermats last theorem is
the famous claim of the 16-th century mathematician Fermat that all integer
solutions to the diophantine equation
X
n
+ Y
n
= Z
n
with 3 n N
are trivial, i.e., xyz = 0; it refused its solution for more than 350 years unless
in 1995 Wiles [213] proved (in parts jointly with Taylor and building on
works of many others) an essential part of the Shimura-Taniyama conjecture,
namely, that any semistable elliptic curve (i.e., with squarefree conductor)
is modular. The full conjecture was proved by Breuil et al. [26]. Here we
want to motivate the link between geometry and number theory predicted by
the Shimura-Taniyama conjecture. For more details we refer to Knapp [115],
Section 3.6 Shimura-Taniyama-Wiles 135
Koblitz [117], and Washington [207]; for the amazing story behind Wiles
proof read Singh [184].
3.6.1. Elliptic curves and their L-functions. An elliptic curve E
over some eld K is a non-singular cubic curve f(X, Y ) = 0 with a K-rational
point (which may be a point at innity). If char K ,= 2, 3, then the cubic can
be written as
(3.40) Y
2
= X
3
+ aX + b with a, b K;
if the characteristic is 2 or 3, then slightly more complicated normal forms
have to be considered. However, for the sake of simplicity we may assume
that E is the set of K-solutions (x, y) to the diophantine equation above
plus the points at innity. Further, we may assume that the elliptic curve is
dened over the rationals, i.e., a, b Q.
A famous theorem of Mordell [151] states that the set of rational points on
an elliptic curve forms a nitely generated abelian group. In particular, this
means that we can add points on an elliptic curve and their sum is a further
point on this curve. This fact relies on the simple observation that a generic
straight line has three intersection points with the cubic equation (3.40). This
gives an algebraic relation between any two given points P
1
and P
2
on the
elliptic curve and a third one, Q = (x, y) say. For some reasons we do not
explain, one cannot take Q to be the sum of P
1
and P
2
, but replacing Q by
its conjugate with respect to the x-axis is doing the job: P
1
+ P
2
= (x, y).
-2 -1 1 2 3
-4
-2
2
4
Figure 2. Adding points on an elliptic curve; here (1, 0) +
(0, 1) = (2, 3) on the elliptic curve given by Y
2
= X
3
+ 1.
One can show that this group of rational points has the form
(3.41) E(Q)

= T Z
r
,
where T is a nite group consisting of the torsion points and r is a non-
negative integer, called the rank of the elliptic curve E, which measures
136 Chapter 3 Hecke theory
the size of E. This structure of elliptic curves makes them a useful tool in
cryptography.
The construction of L-functions associated with elliptic curves is due to
Hasse [83] and his contemporaries. For this aim we have to study the re-
duction of E modulo the prime numbers. For prime p, denote by (p) the
number of solutions of (3.40) in Z/pZ, i.e., the number of solutions to the
congruence
Y
2
X
3
+ aX + b mod p,
where the rational numbers a, b now have to be taken as the corresponding
residues modulo p (via the canonical projection onto Z/pZ). Dene
(p) = p (p) = p + 1 E(Z/pZ)
and put
L
E
(s) =

p|
_
1
(p)
p
s
_
1

p
_
1
(p)
p
s
+
1
p
2s1
_
1
,
where := 16(4a
3
+ 27b
2
) ,= 0 is the discriminant of E and which is
non-zero since E is by denition non-singular (the discriminant is intimately
related to the discriminant modular form (z) dened by (3.28)). Hasse
proved that
(3.42) [(p)[ < 2

p;
this inequality might be regarded as the analogue of the Riemann hypothesis
for the (local) congruence zeta-functions (it is also related to Delignes esti-
mate (3.32)). Consequently, the Euler product for L
E
(s) converges absolutely
for Re s >
3
2
. The analytic continuation of L
E
(s) to an entire function and
a Riemann-type functional equation were conjectured by Hasse and, apart
from partial cases, were proved only recently by the proof of the full Shimura-
Taniyama conjecture by Wiles et al. [213, 26]. The functional equations has
the form
(3.43)
_

N
2
_
s
(s)L
E
(s) =
_

N
2
_
2s
(2 s)L
E
(2 s)
where N is the conductor of the elliptic curve E, that is an integer built
from the prime divisors of the discriminant of E. Indeed, this reminds us of
the functional equation for Dirichlet series associated with modular forms of
weight k = 2.
One of the big yet unsolved questions is the Birch Swinnerton-Dyer con-
jecture [17] (which is another millennium problem). In view of (3.42), the
number N
p
of points of an elliptic curve by reduction modulo a prime p lies
Section 3.6 Shimura-Taniyama-Wiles 137
in the interval
p + 1

p < N
p
:= E(Z/pZ) < p + 1 +

p.
It was shown by Lenstra [130] that the set of values which N
p
assumes in the
Hasse interval if one varies over all elliptic curves E mod p is quite similar
to the one of a random integer (an item which is of great importance for
their cryptographical use). We may expect that if the elliptic curve E has
innitely many rational points, so if the rank r in (3.41) is positive, then
these points would be a rich source for many points by reduction modulo p
and N
p
would be large. On the contrary, if r = 0, then N
p
would straddle
both sides of p + 1 equally. We may rewrite (3.42) as (p) = 2Re
p
with

p
=

p exp(i
p
),
p
R. In order to measure the relative size of the N
p
with respect to p as p varies we may consider

p
N
p
p
=

p
_
1

p
p
__
1

p
p
_

p
_
1
(p)
p
s
+
p
p
2s
_

s=1
These innite products do not converge; however, by the properties of the L-
function associated with E, the right-hand side has an analytic continuation
and might be regarded as the reciprocal of the value L
E
(s) at s = 1. Now,
roughly speaking, the Birch Swinnerton-Dyer conjecture claims that the
rank r of the Mordell group of an elliptic curve (3.41) is equal to the order
of vanishing of the associated L-function L
E
(s) at s = 1. Goldfeld [62]
examined that

px
p
N
p
p

L
E
(1)
(log x)
r
,
where

L
E
(s) is the Euler product L
E
(s) restricted to those primes p which
do not divide the discriminant . Furthermore, he showed that the Birch
Swinnerton-Dyer conjecture implies the Riemann hypothesis for L
E
(s), i.e.,
the non-vanishing of L
E
(s) for > 1; this implication shows its deepness and
so we might be sceptic about a solution in the very near future. For more
information we refer to the survey of Wiles [214].
3.6.2. Weils converse theorem. We can read Heckes theorem 3.9 as
follows: suppose f(z) is given by a Fourier series with polynomially bounded
coecients a(n). Then f(z) is a modular form of weight k for the full modular
group if and only if the function
(s) = (2)
s
(s)

n=1
a(n)
n
s
has an analytic continuation to C such that
(s) +
a(0)
s
+
i
k
a(0)
s k
138 Chapter 3 Hecke theory
is entire and bounded in any vertical strip and satises the functional equation
(s) = i
k
(k s).
However, if we are dealing with modular forms to congruence subgroups, then
more than one functional equation is needed in order to show modularity.
This follows from the fact that in general there are more pairs of modular
forms and associated Dirichlet series if the level q of the congruence subgroup

0
(q) is large, and thus we cannot easily identify one single pair. We observe
that, fortunately, there are many characters mod q which can be used to
nd additional functional equations. This idea is due to Weil [209] who
proved (a stronger version of)
Theorem 3.10. Let k, N N and 1. Given a sequence of complex
numbers a(n) n
c
, dene
f(z) =

n=1
a(n) exp(2inz).
Assume that the function
(s) := (2)
s
(s)L(s), where L(s) :=

n=1
a(n)
n
s
,
has the property that
(s) +
a(0)
s
+
a(0)
k s
denes an entire function which is bounded in any vertical strip and satises
the functional equation
(s) = (k s).
Finally, suppose that for any primitive character mod m and m coprime
with N, the function

(s) :=
_
m

N
2
_
s
(s)L

(s), where L

(s) :=

n=1
a(n)(n)
n
s
,
extends to an entire function which is bounded in any vertical strip and sat-
ises the functional equation

(s) = (N)
()
(chi)

(k s),
where () is the Gauss sum associated to . Then, f is a modular form of
weight k for
0
(N). If in addition the series dening L(s) converges abso-
lutely for > k for some > 0, then f is a cusp form.
Section 3.6 Shimura-Taniyama-Wiles 139
The proof is beyond the scope of our notes (and uses much of the theory of
Hecke operators); it can be found in Iwaniec [99] and Ogg [163].
Weils converse theorem gave support to the Shimura-Taniyama conjecture
(and in some literature this conjecture is also named Shimura-Taniyama-Weil
conjecture). The conjecture of Hasse stated that the L-function L
E
(s) of an
elliptic curve satises a functional equation of the form as the one for Dirichlet
series attached to modular forms of weight k = 2. Furthermore, L-functions
to newforms have an Euler product comparable to the one of L-functions to
elliptic curves. The famous Shimura-Taniyama conjecture claims that indeed
these two objects from dierent elds are just the same: for any elliptic curve
E, there exists a newform f of weight 2 for some congruence subgroup
0
(N)
such that L
E
(s) = L(s, f). In many instances one can use Weils converse
theorem to verify that L
E
(s) is indeed the L-function to such a newform. In
the next section we shall briey discuss one example.
At the end of his paper [209] Weil restates the Shimura-Taniyama conjec-
ture with respect to some correspondence from Shimura (Taniyama commit-
ted suicide in 1957) and writes
Ob die Dinge immer (. . .) sich so verhalten, scheint im Moment
noch problematisch zu sein und mag dem interessierten Leser
als

Ubungsaufgabe empfohlen werden.
3.6.3. An example. We shall illustrate Weils converse theorem with
an example (following Iwaniec [99] and Koblitz [117]). For a square-free
integer m, consider the family of elliptic curves E
m
given by the equations
E
m
: Y
2
= X
3
m
2
X.
The discriminant is easily seen to be
m
= (2m)
6
. For any prime p we obtain
by reduction modulo p reduced curves E
m
(Z/pZ) and the number of points
on these elliptic curves is given by
m
(p) = p + 1 E
m
(Z/pZ). For xed p
we consider the congruence
Y
2
X
3
m
2
X mod p.
It is easily seen that there are no solutions for p = 2. For any odd prime we
count the solutions in terms of the Legendre symbol. This leads to

m
(p) =

x mod p
_
x
3
m
2
x
p
_
=
_
m
p
_

x mod p
_
x
3
x
p
_
,
where the identity in the last step comes from the substitution x mx.
Hence,

m
(p) =
_
m
p
_

1
(p).
140 Chapter 3 Hecke theory
Following Hasses construction, we nd
L
Em
(s) =

p|m
_
1

1
(p)
p
s
_
m
p
__
1

pm
_
1

1
(p)
p
s
_
m
p
_
+
1
p
2s1
_
m
p
_
2
_
1
.
In the half-plane >
3
2
we may expand the Euler product into a Dirichlet
series and it follows that
L
Em
(s) =

n=1

m
(n)
n
s
=

n=1

1
(n)
n
s
_
m
n
_
;
here the values
m
(n) are dened by multiplicativity. The right-hand side is
the Dirichlet series of L
E
1
(s) twisted by the Jacobi symbol:
n
_
m
n
_
=

j=1
p
1
...p=n
_
m
p
j
_
,
where the p
j
are the (not necessarily distinct) prime divisors of n. We notice
this fact as
L
Em
(s) = L
E
1
_
s,
_
m
__
.
The Jacobi symbol
_
m
_
is the quadratic character corresponding to the num-
ber eld Q(

m); in particular, it is a primitive character.


We consider now the special case m = 1. It is not dicult to see that

1
(p) = 0 for p 3 mod 4. One can show that the primes p 1 mod 4 split
in Z[

1] into p = and
(3.44)
1
(p) = ( + ).
For this aim one has to consider the bijection
: E
1
(0, 0) E

: Y
2
= X
4
+ 1
(x, y) (yx
1
, 2x y
2
x
2
).
it follows that
E
1
(Z/pZ) = E

(Z/pZ) 1.
Since (Z/pZ)

is cyclic and dual to the character group mod p,


x (Z/pZ)

: x
4
= z =

4
=1
(z)
for any z (Z/pZ)

. Hence,
E

(Z/pZ) =

4
=1
G() with G() :=

y mod p
(y
2
4).
Section 3.6 Shimura-Taniyama-Wiles 141
One can show that there are four characters with
4
= 1, namely, the powers
1, ,
2
, and
4
of the Legendre symbol (n) = (
n
p
). It is easily computed
that G(1) = p 2, G(
2
) = 1, G(
3
) = G(), and
G() =

y mod p
_
y
2
4
p
_
= .
Hence,

1
(p) = p + 2
_
G(1) + G() + G(
2
) + G(
3
)
_
= ( + );
this is (3.44). Thus, we nd
L
E
1
(s) =

p3 mod 4
_
1 +
p
p
2s
_

p1 mod 4
p=
_
1 +
+
p
s
+
p
p
2s
_
1
,
resp.
L
E
1
(s) =

p2
_
1
(p)
N(p)
s
_
1
for some gr ossencharacter on Q(

1), where the product is taken over


the prime ideals p coprime with 2. Grossencharacters are a generalization
of Dirichlet characters and the L-function above is an example of a so-called
Hecke L-function associated with number elds, being an analogue of Dirich-
let L-functions. These L-functions are known to have analytic continuation
and a functional equation (and we shall study them more detailed in the
next chapter). An application of Weils converse theorem 3.10 now yields
that L
E
1
(s) = L(s, f), where
(3.45)
f(z) =

n=1

1
(n)q
n
= q 2q
5
3q
9
+ 6q
13
+ . . . with q = exp(2iz)
denes a newform of weight 2 for the congruence subgroup
0
(32). Moreover,
one can show that L
Em
(s) = L(s, f
m
) with
f
m
S
2
(
0
(32m
2
)) or S
2
(
0
(16m
2
))
according to m odd or even. These observations date back at least to Tun-
nell [201] and his study on the congruent number problem. More generally,
one can show that a cusp form of weight k, level N and multiplier of con-
ductor q, twisted with a character mod r, is a cusp form of weight k, level
lcm[N, qr, r
2
] (the least common multiple) and multiplier
2
. Twisting with
quadratic characters is a well-known technique in the theory of elliptic curves,
and the example of the elliptic curves E
m
is only one of many.
Before 1995, the Shimura-Taniyama conjecture was known for elliptic
curves with complex multiplication and in isolated examples. However, by
142 Chapter 3 Hecke theory
the work of Wiles & Taylor [213, 198] and Breuil et al. [26] it is now a
theorem:
Modularity theorem. For any elliptic curve E, there exists a newform f
of weight 2 for some congruence subgroup
0
(N) such that L
E
(s) = L(s, f).
Exercise 88. * Let E be an elliptic curve and p prime. Consider the sum
(p) := p + 1 E(Z/pZ) =

x mod p
_
x
3
+ ax + b
p
_
as a random walk. What would the theory of random walks imply?
Exercise 89. Let E be an elliptic curve. Prove that the Euler product L
E
(s)
converges for >
3
2
.
Exercise 90. ** Read in Iwaniec [99] and Koblitz [117] and ll the gaps in the
proof of L
E
1
(s) = L(s, f), where f is given by (3.45).
CHAPTER 4
The Selberg class an axiomatic approach
What is an L-function? We know it when we see one!
M.N. Huxley.
In view of plenty of examples of Dirichlet series in arithmetic it might be
reasonable to ask for a classication and to search for common patterns in
their analytic properties. There were several noticeable attempts to dene
classes of relevant Dirichlet series (as for example Lekkerkerker [129], Perelli
[166], and Matsumoto [143]), however, these classes were in some sense lack-
ing algebraic structure.
In 1989, Selberg [180] dened a general class of Dirichlet series having an
Euler product, analytic continuation and a functional equation of Riemann-
type (plus some side conditions), and formulated some fundamental conjec-
tures concerning them. Especially these conjectures give this class of Dirichlet
series a certain structure which applies to central problems in number theory.
He writes about his conjectures that
these conjectures, which, by the way, are not unrelated to sev-
eral other conjectures like the Sato-Tate conjecture, Langlands
conjectures, etc., have been veried in a number of cases for
Dirichlet series with functional equation and Euler product that
occur in number theory, by assuming that the factorizations we
can give are actually that a function is really primitive and
cannot be factorized further.
Indeed, one of its consequences is the famous yet unsolved Artin conjecture.
In the meantime this so-called Selberg class became an important object
of research but still it is not understood very well. It is conjectured that the
Selberg class consists of the automorphic L-functions and that the analogue
of the Riemann hypothesis holds for all its elements.
4.1. Denition and rst observations
The Selberg class o consists of Dirichlet series
/(s) :=

n=1
a(n)
n
s
satisfying the following hypotheses:
143
144 Chapter 4 The Selberg class
Ramanujan hypothesis: a(n) n

for any > 0;


Analytic continuation: there exists a non-negative integer k such
that (s 1)
k
/(s) is an entire function of nite order;
Functional equation: there exists a positive integer f, and for
1 j f, there are positive real numbers Q,
j
, and there are
complex numbers
j
, with Re
j
0 and [[ = 1, such that

L
(s) =
L
(1 s),
where

L
(s) := /(s)Q
s
f

j=1
(
j
s +
j
);
Euler product: /(s) satises
/(s) =

p
/
p
(s),
where
/
p
(s) = exp
_

k=1
b(p
k
)
p
ks
_
with suitable coecients b(p
k
) satisfying b(p
k
) p
k
for some <
1
2
.
The Ramanujan hypothesis implies that the Dirichlet series converges abso-
lutely in the half-plane > 1, and uniformly in every compact subset. Thus
it follows that elements /(s) are analytic in > 1 and so it makes sense to
speak about analytic continuation. The axiom on the Euler product implies
that the coecients a(n) are multiplicative, and that each Euler factor has
the Dirichlet series representation
/
p
(s) =

k=0
a(p
k
)
p
ks
,
absolutely convergent for > 0.
Obvious examples in the Selberg class are the Riemann zeta-function (s)
and Dirichlet L-functions L(s, ) to primitive characters; notice that L(s, )
with a non-primitve character mod q, q ,= 1, is not in o by lack of the cor-
rect form of the functional equation. More advanced examples are Dedekind
zeta-functions. Kaczorowski et al. [105] studied Hecke L-functions to mod-
ular forms of Hecke groups; they were shown to be either in the Selberg class
or a related class where the axiom of the functional equation is adjusted.
In view of the Euler product representation it is clear that any element
/(s) of the Selberg class does not vanish in the half-plane of absolute con-
vergence > 1. This gives rise to the notions of critical strip and critical
line. The zeros of /(s) located at the poles of gamma-factors appearing in
Section 4.1 Definition 145
the functional equation are called trivial. They all lie in 0, and it is
easily seen that they are located at
(4.1) s =
k +
j

j
with k N
0
and 1 j f.
All other zeros are said to be nontrivial and they lie in the critical strip
0 1. In general we cannot exclude the possibility that /(s) has a
trivial zero and a nontrivial one at the same point. It is expected that for
every function in the Selberg class the analogue of the Riemann hypothesis
holds, that is all nontrivial zeros lie on the critical line:
Grand Riemann hypothesis. If / o, then /(s) ,= 0 for >
1
2
.
Following Conrey & Ghosh [38] we motivate the axioms dening o. We have
already seen that the Ramanujan hypothesis implies the regularity of /(s)
in > 1. Further we note:
The condition that there be at most one pole, and that this one is
located at s = 1, is natural. If we would allow more poles they would
lie on the line = 1, and for each of them /(s) we would expect
the zeta-function suitably shifted as a factor (since otherwise /(s)
would have zeros o the critical line). It is now obvious that it is
sucient to investigate functions with at most one pole, normalized
to be at s = 1.
The restriction Re
j
0 in the functional equation comes from
the theory of Maass waveforms. Assume that there exists an arith-
metic subgroup of SL
2
(R) together with a Maass cusp form that
corresponds to an exceptional eigenvalue, and suppose that the
Ramanujan-Petersson conjecture holds, then the L-function asso-
ciated with the Maass cusp form has a functional equation with
j
which satises Re
j
< 0, but the L-function violates Riemanns
hypothesis.
Finally, consider the axiom concerning the Euler product. It is well-
known that the existence of an Euler product is a necessary (but not
sucient) condition for Riemanns hypothesis. On the rst sight the
condition <
1
2
seems to be a little bit unnatural. However, if =
1
2
would be allowed, the function
(1 2
1s
)(s) =

n=1
(1)
n1
n
s
would lie in o, but obviously, it violates Riemanns hypothesis (see
also the proof of Theorem 4.1 below; further examples were given by
Kaczorowski & Perelli [106]).
146 Chapter 4 The Selberg class
The zero-distribution is essential for the Selberg class. If anyone of the dis-
cussed restrictions would be removed, the resulting larger class would proba-
bly contain Dirichlet series for which the Riemann hypothesis does not hold.
Exercise 91. Assume that / o and let be a xed real number. Show that if
/(s) is regular at s = 1, then also /(s + i) is an element of o.
Exercise 92. Verify that Dedekind zeta-functions are elements of the Selberg class.
4.2. The structure of the Selberg class
The structure of the Selberg class is of special interest. Obviously, the
Selberg class is multiplicatively closed. To classify its ner structure we need
a quantity in order to measure the size of its elements.
The degree of / o is dened by
d
L
= 2
f

j=1

j
,
where the
j
are from the Gamma-factors in the functional equation. Al-
though the data of the functional equation is not unique, the quantity d
L
is
well-dened. If N
L
(T) counts the number of zeros of / o in the rectangle
0 1, [t[ T (according to multiplicities) one can show by standard
contour integration
N
L
(T)
d
L

T log T (4.2)
in analogy to the Riemann-von Mangoldt formula (1.12) for Riemanns zeta-
function; we shall give a more precise asymptotic formula in Theorem 4.11
below. It is conjectured that all / o have integral degree. This is the
degree conjecture. Slightly stronger is the
Strong -conjecture. Let / o. All
j
appearing in the gamma-factors
of the functional equation can be chosen to be equal to
1
2
.
4.2.1. The case of small degrees. Recently, Kaczorowski & Perelli
[109] showed that all functions / o with degree 0 < d
L
<
5
3
have degree
equal to one. This supports the degree conjecture; moreover, they obtained
a complete classication of all elements in the Selberg class of degree d <
5
3
and for all of them it turned out that also the strong -conjecture is true.
Here we shall only prove
Section 4.2 Structure of the Selberg class 147
Theorem 4.1. Let / o. If d
L
= 0, then /(s) 1. If d
L
is positive, then
d
L
1.
This weaker statement was rst proved by Conrey & Ghosh [38]; however, it
is essentially included in Bochners extension of Hamburgers theorem 3.8 on
the Riemann functional equation [18] (see also Vigneras [203]). For the rst
statement we follow the argument of Conrey & Ghosh, and for the second
claim we follow Molteni [147].
Proof. We may assume that d
L
< 1. Let B be a constant such that
a(n) n
B
. By Perrons formula (1.36), we nd

nx
a(n) =
1
2i
_
c+iT
ciT
/(s)
x
s
s
ds + O
_
x
c+B
T
_
,
where c > 1 is a constant. Shifting the path of integration to the left, yields,
by the Phragmen-Lindel of principle (see Section 3.1.7 and the following sec-
tion), the asymptotic formula

nx
a(n) = xP(log x) + O
_
x
(1+B)
d
L
1
d
L
+1
+
_
,
where P(x) is a computable polynomial according to the residue of /(s) at
s = 1. By subtraction, this implies
(4.3) a(n) n
(1+B)
d
L
1
d
L
+1
+
,
where the implicit constant depends on B. For d
L
< 1 the exponent is
negative, and we may choose B arbitrarily large. Then /(s) is uniformly
bounded in every right half-plane. This is a contradiction for / o with
positive degree since the functional equation implies a certain order of growth
(this will become clear in the following section). This shows that o is free of
elements having degree 0 < d < 1.
It remains to consider the case that d
L
= 0. Then the functional equation
takes the form:
Q
s
/(s) = Q
1s
/(1 s)
(there are no Gamma-factors). By (4.3) the a(n) are so small that the Dirich-
let series for /(s) converges in the whole complex plane. Thus we may rewrite
the functional equation as
(4.4)

n=1
a(n)
_
Q
2
n
_
s
= Q

n=1
a(n)
n
n
s
.
We may regard this as an identity between absolutely convergent Dirichlet
series. Thus, is a(n) ,= 0, then Q
2
/n is an integer. In particular, q := Q
2
N.
Moreover, since Q
2
has only nitely many divisors, it follows that /(s) is a
148 Chapter 4 The Selberg class
Dirichlet polynomial. If q = 1, then /(s) 1 and we are done with the case
d
L
= 0. Hence, we may assume that q > 1.
Since the Dirichlet coecients a(n) are multiplicative, we have a(1) = 1
and via (4.4)
a(1)Q
2s
= Q
1
a(Q
2
)Q
2s
;
thus, [a(q)[ = Q. In particular, there exists a prime p such that the exponent
of p in the prime factorization of q is positive and, by the multiplicativity
of the a(n)s,
[a(p

)[ p

2
.
Now consider the logarithm of the corresponding Euler factor:
log
_
1 +

m=1
a(p
m
)
p
ms
_
=

k=1
b(p
k
)
p
ks
.
Viewing this as power series in X = p
s
, we write
log P(X) =

k=1
B
k
X
k
with B
k
= b(p
k
).
Since a(1) = 1 we nd
P(X) = 1 +

m=1
a(p
m
)X
m
=

j=1
(1 C
j
X) with B
k
=
1
k

j=1
C
k
j
.
Now

j=1
[C
j
[ = [a(p

)[ p

2
,
and thus the maximum of the values [C
j
[ is greater than or equal to p
1
2
. We
have
lim
k
[b(p
k
)[
1
k
= lim
k

1
k

j=1
C
k
j

1
k
= max
1j
[C
j
[;
by our foregoing observations the right-hand side is greater than or equal to
p
1
2
. This is a contradiction to the condition b(p
k
) p
k
with some <
1
2
in
the axiom on the Euler product. Hence, q = 1 and /(s) 1. This proves
the rst statement.
The theorem is proved.
By the work of Kaczorowski & Perelli [107] it is known that the functions
of degree one in the Selberg class are the Riemann zeta-function and shifts
L(s + i, ) of Dirichlet L-functions attached to primitive characters with
R. However, for higher degree there is no complete classication so far.
Examples of degree two are normalized L-functions associated with holomor-
phic newforms; here the notion normalized means that a(p) is replaced by
Section 4.3 The Riemannvon Mangoldt formula 149
a(p)p

k1
2
in the notation of Section 3.4. Normalized L-functions attached
to non-holomorphic newforms are expected to lie in o but the Ramanujan
hypothesis is not yet veried. The Rankin-Selberg L-function of any two
holomorphic newforms is an element of the Selberg class of degree 4. Other
examples are Dedekind zeta-functions to number elds K; their degree is
equal to the degree of the eld extension K/Q.
Exercise 93. Prove that the Selberg class is multiplicatively closed.
Exercise 94. Show that the data of the functional equation is not unique.
Hint: Legendres duplication formula for the Gamma-function.
Exercise 95. Verify that L-functions associated with newforms are elements of
the Selberg class.
4.3. The Riemannvon Mangoldt formula
Riemann conjectured an asymptotic formula for the number N(T) of non-
trivial zeros = + i of (s) with 0 T (counted according multi-
plicities). This so-called Riemann-von Mangoldt formula (1.12) was proved
by von Mangoldt in 1895. Now we want to show a Riemann-von Mangoldt
formula for elements of the Selberg class. One method is contour integra-
tion applied to the logarithmic derivative. This is the classic approach due
to von Mangoldt and it can be found in Titchmarsh [200] and many other
books for the special case of the zeta-function. Here we shall go another way
which provides more information on the value-distribution of the L-functions
in question. This method is due to Levinson [133] who applied it to the
zeta-function; the application to the Selberg class is from Steuding [192].
However, rst of all we have to state some preliminary results (not all with
proofs).
4.3.1. Mean-square estimates. The order of growth of a meromorphic
function is of special interest. Recall our observations on the order of growth
of Dirichlet series from Section 3.1.7. For / o we dene

L
() = limsup
t
log [/( + it)[
log [t[
.
One can show that
L
() is a convex function of . Taking into account the
absolute convergence of the dening Dirichlet series we obtain immediately

L
() = 0 for > 1. The order of growth in the half-plane left of the critical
strip is ruled by the functional equation which we may rewrite as
/(s) =
L
(s)/(1 s), (4.5)
150 Chapter 4 The Selberg class
where

L
(s) := Q
12s
f

j=1
(
j
(1 s) +
j
)
(
j
s +
j
)
.
Applying Stirlings formula (1.53), we get after a short computation
Lemma 4.2. Let / o. For t 1, uniformly in ,

L
( + it) =
_
Q
2
t
d
L
_1
2
it
exp
_
it d
L
+
i( d
L
)
4
__
+ O
_
1
t
__
,
where
:= 2
f

j=1
(1 2
j
) and :=
f

j=1

2
j
j
.
Using the Phragmen-Lindel of principle, we can obtain upper bounds for the
order of growth inside the critical strip.
Theorem 4.3. Let / o. Uniformly in , as [t[ ,
/( + it) [t[
(
1
2
) d
L
[/(1 + it)[.
In particular,

L
()
_
_
_
0 if > 1,
1
2
d
L
(1 ) if 0 1,
(
1
2
) d
L
if < 0.
This theorem should be compared with our results for the zeta-function from
the previous chapters. Our proof is more or less the same as in Section 3.1.7.
Proof. The rst assertion follows immediately from the functional equation
and Lemma 4.2. This estimate implies for < 0

L
() =
_
1
2

_
d
L
.
The calculation of
L
() for 0 1 is more dicult. Here we apply the
theorem of Phragmen-Lindel of, Lemma 3.5. In view of the axiom concerning
the analytic continuation /(s) is a function of nite order. Thus, Lemma 3.5
shows that
L
() is non-increasing and convex downwards. By the estimates
of
L
() for outside of the critical strip the second assertion of the theorem
follows.
It should be noticed that we did not use the condition that the
j
appearing
in the gamma factors of the functional equation have positive real part.
In view of the functional equation, resp. the convexity of
L
, the value for
=
1
2
is essential. In particular, we obtain
L
(
1
2
)
1
4
d
L
, or equivalently,
/
_
1
2
+ it
_
[t[
1
4
d
L
+
, (4.6)
Section 4.3 The Riemannvon Mangoldt formula 151
valid for [t[ 1.
Next we shall apply the following general theorem on the mean-square of
Dirichlet series satisfying a Riemann-type functional equation due to Potter
[171].
Theorem 4.4. Suppose that the functions
A(s) =

n=1
a
n
n
s
and B(s) =

n=1
b
n
n
s
have a half-plane of convergence, are of nite order, and that all singulari-
ties lie in a subset of the complex plane of nite area. Further, assume the
estimates

nx
[a
n
[
2
x
b+
and

nx
[b
n
[
2
x
b+
,
as x , and that A(s) and B(s) satisfy
A(s) = h(s)B(1 s),
where h(s) [t[
c(
a
2
)
uniformly in for from a nite interval, as [t[ ,
and c is some positive constant. Then
lim
T
1
2T
_
T
T
[A( + it)[
2
dt =

n=1
[a
n
[
2
n
2
for > max
a
2
,
1
2
(b + 1)
1
c
.
We do not give the lengthy proof of Potters theorem here and refer directly
to Potter [171]. But we shall apply the theorem to L-functions in the Selberg
class. Taking into account Lemma 4.3 we obtain
Corollary 4.5. Let / o. For > max
_
1
2
, 1
1
d
L
_
,
lim
T
1
2T
_
T
T
[/( + it)[
2
dt =

n=1
[a(n)[
2
n
2
.
Note that the series on the right hand side converges on behalf of the Ra-
manujan hypothesis (resp. the polynomial Euler product representation).
Every (convergent) Dirichlet series has a mean-square half-plane (see Titch-
marsh [199]), i.e., a half-plane in which the mean-square on vertical lines is
bounded. In view of Corollary 4.5 the mean-square half-plane of / o
contains the region
> max
_
1
2
, 1
1
d
L
_
.
It is expected that the mean-square exists for any / o for >
1
2
(as in the
case of zeta). However, this is a deep conjecture and its verication is even
in single cases a dicult task. In fact the diculties arise for large degrees
152 Chapter 4 The Selberg class
d
L
. Potters theorem yields only an asymptotic formula throughout >
1
2
if the degree d
L
is less than or equal to two. The diculties become more
obvious by noting that any result on the mean-square of an L-function from
the Selberg class of degree d is comparable to the corresponding result for
the 2 d-th moment of the Riemann zeta-function.
4.3.2. Sums over c-values. Let c be a complex number. Levinson [133]
proved that all but
N(T)
log log T
of the roots of (s) = c in T < t < 2T lie in


1
2

<
(log log T)
2
log T
.
Thus, the c-values of the zeta-function are clustered around the critical line.
In particular, we see that the density estimate 2.16 does not indicate the
truth of the Riemann hypothesis. As we shall show now, this distribution of
c-values is typical for L-functions in the Selberg class.
The c-values of /(s) are the roots of the equation
(4.7) /(s) = c,
which we denote by
c
=
c
+ i
c
. Our rst aim is to prove estimates for
sums taken over c-values, weighted with respect to their real parts.
Theorem 4.6. Let /

o and c ,= 1. Then, for b > max
1
2
, 1
1
d
L
,

c>b
T<c2T
(
c
b) T.
Assuming the truth of Lindelof s hypothesis, i.e.,
/
_
1
2
+ it
_
t

as t , we have

c>
1
2
T<c2T
_

1
2
_
= o(T log T).
The case c = 1 is exceptional since 1 is the limit of /(s) as :
/(s) = 1 + O(2

). (4.8)
However, without big eort one can obtain also in this case similar estimates.
It should be noted that the Lindel of hypothesis for /(s) follows from the
Riemann hypothesis and thus it is widely expected to hold.
Proof of Theorem 4.6. In view of (4.8) there exists a positive real number
A depending on c such that all real parts
c
of c-values satisfy
c
< A. Put
(s) =
/(s) c
1 c
.
Section 4.3 The Riemannvon Mangoldt formula 153
Obviously, the zeros of (s) correspond exactly to the c-values of /(s). Next
we apply Littlewoods lemma 2.13. Let (, T) denote the number of zeros

c
of (s) with
c
> and T <
c
2T (counting multiplicities). Let a be a
parameter with a > maxA + 1, b. Then Littlewoods lemma 2.13, applied
to the rectangle 1 with vertices a + iT, a + 2iT, b + iT, b + 2iT, gives
_
R
log (s) ds = 2i
_
a
b
(, T) d.
Since
_
a
b
(, T) d =

c>b
T<2T
_
c
b
d =

c>b
T<c2T
(
c
b), (4.9)
we get
2

c>b
T<c2T
(
c
b) =
_
2T
T
log [(b + it)[ dt
_
2T
T
log [(a + it)[ dt +

_
a
b
arg ( + iT) d +
_
a
b
arg ( + 2iT) d
=
4

j=1
I
j
, (4.10)
say. To dene log (s) we choose the principal branch of the logarithm on the
real axis, as ; for other points s the value of the logarithm is obtained
by analytic continuation.
We start with the vertical integrals. Obviously,
I
1
(T, b) := I
1
=
_
2T
T
log [/(b + it) c[ dt T log [1 c[. (4.11)
By Jensens inequality the integral is

T
2
log
_
1
T
_
2T
T
[/(b + it)[
2
dt
_
+ O(T).
By Corollary 4.5 this is T for b > max
1
2
, 1
1
d
L
. Thus we get I
1
(T, b)
T unconditionally. An immediate consequence of Lindel ofs hypothesis is
_
2T
T

/
_
1
2
+ it
_

2
dt T
1+
for any positive . Thus, assuming the truth of Lindel ofs hypothesis we get
I
1
_
T,
1
2
_
T log T.
154 Chapter 4 The Selberg class
Next we consider I
2
. Since a > 1 we have
(a + it) = 1 +
1
1 c

n=2
a(n)
n
a+it
, (4.12)
and in view of (4.8) the absolute value of the series is less than 1 for suciently
large a. Therefore we nd by the Taylor expansion of the logarithm
log [(a + it)[ = Re

k=1
(1)
k
k(1 c)
k

n
1
=2
. . .

n
k
=2
a(n
1
) . . . a(n
k
)
(n
1
. . . n
k
)
a+it
.
This leads by the Ramanujan hypothesis to the estimate
I
2
= Re

k=1
(1)
k
k(1 c)
k

n
1
=2
. . .

n
k
=2
a(n
1
) . . . a(n
k
)
(n
1
. . . n
k
)
a

_
2T
T
dt
(n
1
n
k
)
it

k=1
1
k
_

n=2
1
n
a
_
k
1, (4.13)
for suciently large a. It remains to estimate the horizontal integrals I
3
, I
4
.
Suppose that Re ( + iT) has N zeros for b a. Then divide [b, a]
into at most N + 1 subintervals in each of which Re ( + iT) is of constant
sign. Then
[ arg ( + iT)[ (N + 1). (4.14)
To estimate N let
g(z) =
1
2
_
(z + iT) + (z + iT)
_
.
Then we have g() = Re ( +iT). Let R = ab and choose T so large that
T > 2R. Now, Im(z + iT) > 0 for [z a[ < T. Thus (z + iT), and hence
g(z) is analytic for [z a[ < T. Let n(r) denote the number of zeros of g(z)
in [z a[ r. Obviously, we have
_
2R
0
n(r)
r
dr n(R)
_
2R
R
dr
r
= n(R) log 2.
With Jensens formula, Lemma 1.17,
_
2R
0
n(r)
r
dr =
1
2
_
2
0
log

g
_
a + 2Re
i
_

d log [g(a)[, (4.15)


we deduce
n(R)
1
2 log 2
_
2
0
log

g
_
a + 2Re
i
_

d
log [g(a)[
log 2
.
Section 4.3 The Riemannvon Mangoldt formula 155
By (4.12) it follows that log [g(a)[ is bounded. By Theorem 4.3, in any vertical
strip of bounded width,
/(s) [t[
B
with a certain positive constant B. Obviously, the same estimate holds for
g(z). Thus, the integral above is log T, and n(R) log T. Since the
interval (b, a) is contained in the disc [z a[ R, the number N is less or
equal n(R). Therefore, with (4.14), we get
[I
4
[
_
a
b
[ arg ( + iT)[ d log T.
Obviously, I
3
can be bounded in the same way.
Collecting all estimates, the assertions of the theorem follow.
Now we will include most of the c-values into our observations. In view
of Theorem 4.3 there exist positive constants C

, T

such that there are no


c-values in the region < C

, t T

. Therefore, assume that b < C

1
and T T

+ 1. By the functional equation in the form (4.5),


log [/(s) c[ = log [
L
(s)[ + log [/(1 s)[ + O
_
1
[
L
(s)/(1 s)[
_
.
In view of Lemma 4.2
log [
L
(s)[ =
_
1
2

_
( d
L
log t + log(Q
2
)) + O
_
1
t
_
.
Thus
_
2T
T
log [/(b + it) c[ dt
=
_
1
2
b
__
2T
T
( d
L
log t + log(Q
2
)) dt
+
_
2T
T
log [/(1 b it)[ dt + O(log T).
Now suppose that c ,= 1. The rst integral on the right hand side is easily
calculated by elementary means. The second integral is small if b is chosen
suciently large (see (4.13)). Thus, taking into account (4.10) and (4.11),
we get
I
1
=
_
1
2
b
__
d
L
T log
4T
e
+ T log(Q
2
)
_
T log [1 c[ + O(log T).
By (4.10) and with the estimates for the I
j
s from the proof of the previous
theorem we obtain
156 Chapter 4 The Selberg class
Theorem 4.7. Let /

o and c ,= 1. Then, for suciently large negative b,
2

T<c2T
(
c
b) =
_
1
2
b
__
d
L
T log
4T
e
+ T log(Q
2
)
_
T log [1 c[ + O(log T).
4.3.3. Riemann-von Mangoldt-type formulae. We can rewrite the
sum over c-values from the previous section as follows:

c
(
c
b) =
_
1
2
b
_

c
1 +

c
_

1
2
_
.
The rst sum on the right counts the number of c-values and the second one
measures the distances of the c-values from the critical line. Let A
c
(T) count
the number of c-values of /(s) with T <
c
2T. Then, subtracting the
formula of Theorem 4.7 with b+1 instead of b from the one with b, we obtain
Corollary 4.8. Let /

o. Then, for c ,= 1,
A
c
(T) =
d
L
2
T log
4T
e
+
T
2
log(Q
2
) + O(log T).
Furthermore,
Corollary 4.9. Let /

o. Then, for c ,= 1,

T<c2T
_

1
2
_
=
T
2
log [1 c[ + O(log T).
Thus, for c satisfying [1 c[ , = 1, the c-values, weighted with respect to
their distance to the critical line, lie asymetrically distributed (which is not
too surprising in view of the fact that
L
() is increasing as ).
Nevertheless, our next aim is to show that most of the c-values lie close
to the critical line. Unfortunately, for this purpose we have to assume the
Lindel of hypothesis. Dene the counting functions (according multiplicities)
A
c
+
(, T) =
c
: T <
c
2T,
c
> ,
and
A
c

(, T) =
c
: T <
c
2T,
c
< .
Then
Theorem 4.10. Let /

o and c ,= 1. Then, for any > max
1
2
, 1
1
d
L
,
N
c
+
(, T) T, (4.16)
and assuming the truth of the Lindelof hypothesis, for any > 0,
A
c

_
1
2
, T
_
+A
c
+
_
1
2
+ , T
_
T log T.
Section 4.3 The Riemannvon Mangoldt formula 157
Proof. First of all, let > max
1
2
, 1
1
d
L
and x
1
(max
1
2
, 1
1
d
L
, ).
Then
A
c
+
(, T)
1

1

c>
T<c2T
(
c

1
).
The sum on the right side is less than or equal to

c>
1
T<c2T
(
c

1
)
_
2T
T
log [(
1
+ it)[ dt + O(log T),
where we used Littlewoods lemma 2.13 and the techniques from the previous
section for the latter inequality. In view of the unconditional estimate for
(4.11) in the proof of Theorem 4.6 we obtain (4.16). Assuming the truth of
the Lindel of hypothesis we get analogously
A
c
+
_
1
2
+ , T
_

T log T
for any positive .
Next we consider A
c

; in particular, we assume the Lindel of hypothesis for


/(s). Let b be a suciently large constant. We have

c
1
2

T<c2T
(
c
b)
_
1
2
b
_

c
1
2

T<c2T
1 +

c
1
2
T<c2T
_

1
2
_
.
Hence

T<c2T
(
c
b) =

c<
1
2

T<c2T
_
1
2
b +
c

1
2
_
+

c
1
2

T<c2T
(
c
b)

_
1
2
b
_
A
c
(T) +

c<
1
2

T<c2T
_

1
2
_
+

c>
1
2
T<c2T
_

1
2
_
.
The second sum on the right is bounded by T log T by Theorem 4.6. Since
any term in the rst sum on the right is < , we obtain
A
c

_
1
2
, T
_

T<c2T
(
c
b)
_
1
2
b
_
A
c
(T) + O(T log T).
In view of Theorem 4.7 and Corollary 4.8 we get
A
c

_
1
2
, T
_

T log T.
Putting =
2
we obtain the assertion of the theorem.
158 Chapter 4 The Selberg class
Thus, subject to the truth of the Lindel of hypothesis, we get by comparing
Corollary 4.8 and Theorem 4.10, for any positive ,
A
c

_
1
2
, T
_
+A
c
+
_
1
2
+ , T
_
A
c
(T),
so the c-values are clustered around the critical line for any c. The distribution
of the c-values close to the real axis is quite regular. It can be shown that
there is always a c-value in a neighbourhood of any trivial zero of /(s) with
suciently large negative real part, and with nitely many exceptions there
are no other in the left half-plane. The main ingredients for the proof are
Rouches theorem, Lemma 2.19, and Stirlings formula (1.53). Consequently,
with regard to (4.1), the number of these c-values having real part in [R, 0]
is asymptotically
1
2
d
L
R. On the other side, by (4.8) the behaviour nearby
the positive real axis is very regular. Note that all results from above hold
as well with respect to c-values from the lower half-plane.
Now let N
c
L
(, T) count the number of c-values
c
=
c
+i
c
of /(s) satis-
fying
c
1, [
c
[ T. Using Corollary 4.8 with 2
n
T for n N instead
of T and adding up, we get, for xed 0,
N
c
L
(, T) = 2

n=1
A
c
(, 2
n
T)
=
_
d
L

T log
T
e
+
T

log(Q
2
)
_

n=1
1
2
n
+
d
L

n=1
log 4 nlog 2
2
n
+ O(log T).
The appearing innite series are equal to 1 and 0, respectively. Hence, this
summation removes the factor 4 in the logarithmic term, and we have proved
Theorem 4.11. Let /

o. For any 0 and any complex c ,= 1,
N
c
L
(, T) =
d
L

T log
T
e
+
T

log(Q
2
) + O(log T).
The case c = = 0 (the nontrivial zeros of /(s)) is a precise Riemann-von
Mangoldt formula (1.12). Similar results were obtained by Perelli [166] and
Lekkerkerker [129] for other classes of Dirichlet series.
In the exceptional case c = 1 one has to consider the function
(s) =
q
s
a(q)
(/(s) 1),
where q is the smallest integer greater than one such that a(q) ,= 0. Then,
by a similar reasoning as in the proof of Theorem 4.11, one gets analogous
results. For the special case of the zeta-function this is carried out in Steuding
[193, 194] where Levinsons method is applied to Epstein zeta-functions.
Section 4.3 The Riemannvon Mangoldt formula 159
4.3.4. Some related results of Selberg. We conclude with some re-
sults from Selberg [180]. Under assumption of the truth of the Riemann
hypothesis he obtained for c ,= 1 the asymptotic formula

c>
1
2
0<c<T
_

1
2
_
=

n
L
4
3
2
T
_
log log T
+
T
4
log
[c[
1 [c[
2
+ O
_
T
(log log log T)
3

log log T
_
,
where n
L
is the quantity appearing in Selbergs Conjecture A. Furthermore,
for
(T) :=
1
2

log log T
log T
and :=
d
L

n
L
with positive , he proved

c>(T)
0<c<T
(
c
(T))
=
1
2
_
n
L

_
exp(
2
)
2
+
_

exp(x
2
) dx
_
T
_
log log T
+
_
log [c[
_

exp(x
2
) dx log [1 c[
_
T
2
+O
_
T
(log log log T)
3

log log T
_
.
From these results Selberg deduced that about half of the c-values lie to
the left of the critical line, statistically well distributed at distances of order

log log T
log T
o =
1
2
, and that
N
c
L
((T), T) N
c
L
(T)
_

exp(x
2
) dx.
Most of the remaining c-values lie rather close to the critical line at distances
of order not exceeding
(log log log T)
3
log T

log log T
. This improves some results due to Selberg
(unpublished) and Joyner [104] and gives a much more detailed description
of the clustering of the c-values around the critical line.
Exercise 96. * Prove Lemma 4.2. Read Potter [171] and understand the proof
of Theorem 4.4. Deduce Corollary 4.5.
Exercise 97. * Prove similar estimates for the c = 1-values of /(s) o.
160 Chapter 4 The Selberg class
4.4. Primitivity and Selbergs conjectures
The Selberg class is multiplicatively closed. Therefore it makes sense to
introduce the notion of primitive elements. A function / o is called prim-
itive if it cannot be factored as a product of two elements non-trivially, i.e.,
the equation
/ = /
1
/
2
with /
1
, /
2
o
implies / = /
1
or / = /
2
. This denition of primitivity (analogously to the
one in algebra) is very natural and useful for studies of the structure of the
Selberg class.
4.4.1. Factorization into primitive functions. The ring of integers
is a unique factorization domain: any integer has (up to order) a unique
factorization into powers of prime numbers. Something similar can be shown
for the Selberg class. Conrey & Ghosh [38] proved
Theorem 4.12. Every function in the Selberg class has a factorization into
primitive functions.
Proof. Suppose that / is not primitive, then there exist functions /
1
and
/
2
in o 1 such that / = /
1
/
2
. Taking into account (4.2) we have
N
L
(T) = N
L
1
(T) + N
L
2
(T),
resp. for the according degrees
d
L
= d
L
1
+ d
L
2
.
In view of Theorem 4.1 both /
1
and /
2
have degree at least 1. Thus, each of
d
L
1
and d
L
2
is strictly less than d
L
. A continuation of this process terminates
since the number of factors is d
L
, which proves the claim.
In connection with Theorem 4.1 it follows that any element of the Selberg
class of degree one is primitive; e.g., Riemanns zeta-function and Dirich-
let L-functions attached to primitive characters. A more advanced example
of primitive elements are L-functions associated with newforms due to M.R.
Murty [156] and further examples were given by Molteni & Steuding [148] by
L-functions to modular forms of Hecke groups. On the contrary, Dedekind
zeta-functions to cyclotomic elds ,= Q are not primitive. In the follow-
ing section we will consider whether factorization into primitive elements is
unique.
Section 4.4 Primitivity and Selbergs conjectures 161
4.4.2. Selbergs conjectures. Denote by a
L
(n) the coecients of the
Dirichlet series representation of / o. The central claim concerning prim-
itive functions is part of
Selbergs conjectures.
A) For all 1 ,= / o there exists a positive integer n
L
such that

px
[a
L
(p)[
2
p
= n
L
log log x + O(1);
B) for any primitive functions /
1
and /
2
,

px
a
L
1
(p)a
L
2
(p)
p
=
_
log log x + O(1) if /
1
= /
2
,
O(1) otherwise.
In some sense, primitive functions are expected to form an orthonormal sys-
tem.
In view of the factorization into primitive functions, Theorem 4.12, it is
easily seen that Conjecture B implies Conjecture A. In some particular cases
it is not too dicult to verify Selberg conjecture A. For instance, (s) sat-
ises Selbergs Conjecture A (see Chapter 1) and, obviously, the same holds
for Dirichlet L-functions too. Liu, Wang & Ye [138] proved Selbergs Conjec-
ture B for automorphic L-functions L(s, ) and L(s,

), where and

are
automorphic irreducible cuspidal representations of GL
m
(Q) and GL
m
(Q), re-
spectively (we shall give a rough denition of these objects in a later section);
their result holds unconditionally for m, m

4 and in other cases under the


assumption of the convergence of

p
[a

(p
k
)[
2
p
k
(log p)
2
for k 2, where a

(n) denote the Dirichlet series coecients of L(s, ). The


latter hypothesis is an immediate consequence of the Ramanujan hypothesis.
We return to the theme of factorization into primitive elements. Conrey
and Ghosh [38] proved
Theorem 4.13. Selbergs conjecture B implies that every / o has a unique
factorization into primitive functions.
Proof. Suppose that / has two factorizations into primitive functions:
/ =
m

j=1
/
j
=
n

k=1

/
k
,
162 Chapter 4 The Selberg class
and assume that no

/
k
is equal to /
1
. Then it follows from
m

j=1
a
L
j
(p) =
n

k=1
a

L
k
(p)
that
m

j=1

px
a
L
j
(p)a
L
1
(p)
p
=
n

k=1

px
a

L
k
(p)a
L
1
(p)
p
.
By Selbergs conjecture B, the left-hand side tends to innity for x ,
whereas the right-hand side is bounded, giving the desired contradiction.
4.4.3. Prime number theorems. The Selberg conjectures refer to the
analytic behaviour at the edge of the critical strip. Conrey & Ghosh [38]
proved the non-vanishing on the line = 1 subject to the truth of Selbergs
Conjecture B:
Theorem 4.14. Let / o. If Selbergs Conjecture B is true, then
/(s) ,= 0 for 1.
It is conjectured that the Selberg class consists only of automorphic L-
functions, and for those Jacquet & Shalika [103] obtained an unconditional
non-vanishing theorem.
Proof of Theorem 4.14. In view of the Euler product representation in the
half-plane 1 zeros can only occur on the line = 1. By Theorem 4.12
it suces to consider primitive functions / o. In case of (s) it is known
that there are no zeros on = 1 (see Chapter 1.4.1). It is easily seen that if
Selbergs conjecture B is true and if / o has a pole at s = 1 of order m,
then the quotient /(s)/(s)
m
is an entire function. Hence we may assume
that /(s) is entire. Then /(s + i) is for any real a primitive element of
o. Selbergs Conjecture B applied to /(s + i) and (s) yields

px
a
L
(p)
p
1+i
1. (4.17)
Now suppose that /(1 + i) = 0. Then
/(s) c(s (1 + i))
k
as s = +i 1 +i for some complex c ,= 0 and some positive integer k.
It follows that
log /( + i) k log( 1) (4.18)
as 1+. Since
log /(s) =

p
a
L
(p)
p
s
+ O(1)
Section 4.4 Primitivity and Selbergs conjectures 163
for > 1, we get by partial summation
log /( + i)

p
a
L
(p)
p
+i
= ( 1)
_

1

px
a
L
(p)
p
1+i
dx
x

.
By (4.17) the right-hand side is bounded as 1+, which contradicts (4.18).
The theorem is proved.
As we have seen in Chapter 1, the non-vanishing of L-functions on the
edge of the critical strip is closely related to prime number theorems (here
we mean asymptotic formulae for the Dirichlet coecients). Indeed, if an
element of the Selberg class /(s) has no zeros in the half-plane 1, we
shall expect the asymptotic formula
(4.19)
L
(x) :=

nx

L
(n) = k
L
x + o(x),
where k
L
= 0 if /(s) is regular at s = 1, otherwise k
L
is the order of the pole
of /(s) at s = 1, and
L
(n) is the von Mangoldt-function, dened by

/
(s) =

n=1

L
(n)
n
s
.
As a matter of fact, we shall even expect that (4.19) is equivalent to the non-
vanishing of /(s) on the 1-line. It is not too dicult to verify this statement
(by application of a Tauberian theorem) for polynomial Euler products in
the Selberg class, i.e.,
(4.20) /(s) =

p
m

j=1
_
1

j
(p)
p
s
_
1
,
where m is a xed positive integer and for each prime p and 1 j m the

j
(p) are certain complex numbers (it is easily seen that they have absolute
value less than or equal to one subject to the Ramanujan hypothesis). In
view of Theorem 4.14 it follows that
Corollary 4.15. Assume Selbergs Conjecture B. The prime number theorem
(4.19) holds for elements of the Selberg class of the form (4.20).
However, Conjecture B might be a rather strong condition if we are interested
in a prime number theorem for a single L-function. Recently, Kaczorowski
& Perelli [108] obtained a more satisfying condition. For this aim they in-
troduced a weak form of Selbergs Conjecture A:
Normality conjecture. For all 1 ,= / o there exists a non-negative
integer k
L
such that

px
[a
L
(p)[
2
p
= k
L
log log x + o(log log x).
164 Chapter 4 The Selberg class
Assuming this hypothesis, they proved the claim of Theorem 4.14, namely
the non-vanishing of any /(s) on the line = 1, and that this statement is
equivalent to the asymptotic formula (4.19). It should be noted that their
proof of /(1+iR) ,= 0 for a given / involves the assumption of their normality
conjecture for several elements in o. In fact, their proof relies on a density
theorem for o (generalizing our approach from Chapter 1.11). Let N
L
(, T)
count the number of zeros = + i of /(s) with > and [[ < T
(counting multiplicities). Then Kaczorowski & Perelli [108] proved that
N
L
(, T) T
4( d
L
+3)(1)+
uniformly for
1
2
1. Unfortunately, this estimate is only useful for
close to 1.
4.4.4. Pair correlation in the Selberg class. Assuming the truth
of the Riemann hypothesis Montgomery [149] studied the distribution of
consecutive zeros
1
2
+i,
1
2
+i

of the Riemann zeta-function. Montgomerys


famous pair correlation conjecture states that, for xed , satisfying 0 <
< ,
lim
T
1
N(T)

_
0 < ,

< T :
(

) log T
2

_
=
_

_
1
_
sin u
u
_
2
_
du. (4.21)
Montgomery claims that (4.21) would follow from a suciently good estimate
for

nx
(n)(n + h) c(h)x
in a certain range of h, where c(h) is some quantity depending on h; however,
the Hardy-Littlewood twin prime conjecture [80] is too strong for an input
into this problem. The pair correlation conjecture has many important con-
sequences; e.g., (4.21) implies that almost all zeros of the zeta-function are
simple.
Dyson remarked shortly afterwards that the function on the right of (4.21)
is the pair correlation function of the eigenvalues of large random Hermitian
matrices, or more specically of the Gaussian Unitary Ensemble. This sup-
ports an old idea of Hilbert and P olya. Their approach towards Riemanns
hypothesis was to look for a self-adjoint Hermitian operator whose eigen-
values are
1
2
where is a nontrivial zero of the zeta-function; then the
property of being self-adjoint would imply that all zeros lie on the critical
line =
1
2
. In the last years big progress in this direction was made. By the
work of Odlyzko [161] it turned out that the pair correlation and the nearest
neighbour spacing for the zeros of (s) were amazingly close to those for the
Section 4.4 Primitivity and Selbergs conjectures 165
Gaussian Unitary Ensemble. There is even more evidence for the pair corre-
lation conjecture than numerical data. In the meantime many results from
random matrix theory were found which t perfectly to certain results on the
value-distribution of the Riemann zeta-function (and even other L-functions;
see Conreys survey article [37]).
For example, Keating & Snaith showed that certain Random Matrix en-
sembles have in a sense the same value-distribution as the zeta-function on
the critical line predicted by Selbergs limit law. More precisely, Keating &
Snaith [114] showed for characteristic polynomials :
N
(, U) of the Circular
Unitary Ensemble |(N) the limit theorem:
lim
N
meas
_
U |(N) :
log :
N
(; U)
_
1
2
log N
1
_
=
1
2
__
R
exp
_

1
2
(x
2
+ y
2
)
_
dxdy,
where 1 is any rectangle in the complex plane with edges parallel to the real-
and the imaginary axis. For the zeta-function there is an old result of Selberg
(unpulished) showing the same Gaussian normal distribution:
lim
T
1
T
meas
_
t [T, 2T] :
log (
1
2
+ it)
_
1
2
log log T
1
_
=
1
2
__
R
exp
_

1
2
(x
2
+ y
2
)
_
dxdy.
The rst published proof of the latter result is due to Joyner [104].
Further evidence for the pair correlation conjecture was discovered by Rud-
nick & Sarnak. Normalize the ordered nontrivial zeros
n
=
1
2
+i
n
by setting

n
=

n
2
log [
n
[,
then it follows from the Riemann-von Mangoldt formula (1.12) that the num-
bers have unit mean spacing. Then the pair correlation conjecture (4.21)
can be rewritten as follows: for any nice function f on (0, )
lim
N

nN
f(
n+1

n
) =
_

0
f(x)P(x) dx
where P is the distribution of consecutive spacings of the eigenvalues of a large
random Hermitean matrix. Rudnick & Sarnak [177] succeeded in showing
that the m-dimensional analogue of the latter formula, the m-level corre-
lation, holds for a large class of test functions. Finally, note that Katz &
Sarnak [113] proved a function eld analogue of Montgomerys pair correla-
tion conjecture without assuming any unproved hypothesis.
166 Chapter 4 The Selberg class
Recently, Murty & Perelli [158] extended Montgomerys argument to the
Selberg class. For this purpose they considered two primitive functions /
1
and /
2
from o. To compare the zeros
1
2
+i
L
1
of /
1
against the zeros
1
2
+i
L
2
of /
2
dene
T(; /
1
, /
2
) =

d
L
1
T log T

T
L
1
,
L
2
T
T
id
L
1
(
L
1

L
2
)
w(
L
1

L
2
),
where w is a suitable weight function. The pair correlation conjecture for the
Selberg class takes then the form:
Pair correlation conjecture. Let /
1
and /
2
be primitive functions in o.
Under the assumption of the Grand Riemann hypothesis, uniformly in , as
T ,
T(; /
1
, /
2
)
_

L
1
,L
2
[[ + d
L
1
T
2|| d
L
1
log T(1 + o(1)) if [[ < 1,

L
1
,L
2
[[ otherwise.
Here

L
1
,L
2
:=
_
1 if /
1
= /
2
,
0 otherwise,
is the Kronecker-symbol. The general pair correlation conjecture includes
Montgomerys pair correlation conjecture. It has plenty of important appli-
cations as M.R. Murty & Perelli [158] worked out (for instance, the Artin
conjecture follows from the pair correlation conjecture). The pair correlation
conjecture implies that almost all zeros of two primitive functions /
1
and /
2
are simple and distinct. Moreover, if the pair correlation formula holds for at
least one value of , then o has unique factorization into primitive functions.
This shows what a powerful tool the pair correlation is. Further, M.R. Murty
& Perelli proved
Theorem 4.16. The Grand Riemann hypothesis and the pair correlation
conjecture imply the Selberg conjectures.
The pair correlation conjecture plays a complementary role to the Riemann
hypothesis: vertical vs. horizontal distribution of the nontrivial zeros of (s).
Both together seem to be the key to several unsolved problems in number
theory!
Exercise 98. i) Prove that Selbergs Conjecture B implies Conjecture A
ii) Show that Selbergs Conjecture B holds for pairs of Dirichlet L-functions.
Exercise 99. If the Selberg conjecture B is true, / o is primitive if and only if
n
L
= 1, where the quantity n
L
from Selbergs Conjecture A.
Section 4.5 Hecke L-functions 167
Exercise 100. Assuming Selbergs conjecture B, prove that if / o has a pole at
s = 1 of order m, then the quotient /(s)/(s)
m
is an entire function.
Exercise 101. Show that a polynomial Euler product of the form (4.20) satis-
es the Ramanujan hypothesis and, conversely, that the Ramanujan hypothesis for
(4.20) implies that [
j
(p)[ 1.
Exercise 102. * Prove Corollary 4.15: Assuming Selbergs Conjecture B, given
a polynomial Euler product (4.20) in the Selberg class, prove the prime number
theorem (4.19).
Hint: apply the Tauberian theorem of Wiener-Ikehara 1.14.
Exercise 103. * Suppose that /
1
, /
2
o and the Dirichlet coecients of both
Dirichlet series are equal for all but nitely many prime numbers: a
L
1
(p) = a
L
2
(p).
Assuming Selbergs Conjecture B, show that /
1
= /
2
.
4.5. Hecke L-functions
In 1920, Hecke [87] introduced a new class of L-functions which generalize
the concepts of Dedekind zeta-functions and Dirichlet L-functions.
Let K be a number eld, f be an ideal of K, and modulo f be a gr ossen-
character (the denition will be given in the following subsection). Then the
associated Hecke L-function is given by
L(s, ) =

a
(a)
N(a)
s
=

p
_
1
(p)
N(p)
s
_
1
, (4.22)
where the sum is taken over all non-zero integral ideals a of K, the product
is taken over all prime ideals p, and N(a) denotes the norm of the ideal a;
the identity between the Dirichlet series and the Euler product follows from
the unique prime ideal factorization.
4.5.1. Grossencharacters. Hecke gr ossencharacters represent the most
general extension of Dirichlet characters to number elds.
Given a number eld K of degree n over Q, there are exactly n automor-
phisms K
(j)
of K into C, for 1 j n, given by
K
(j)
K
(j)
,
where the
(j)
denote the conjugates of ; we assume that among these there
are r
1
real and 2r
2
complex embeddings (that makes n = r
1
+2r
2
). We denote
the real embeddings by
K
(1)
, . . . , K
(r
1
)
and the complex embeddings which are pairwise complex conjugate by
K
(r
1
+1)
, . . . , K
(r
1
+r
2
+1)
= K
(r
1
+1)
, . . . , K
(n)
= K
(r
1
+r
2
)
.
168 Chapter 4 The Selberg class
Let f be a non-zero integral ideal of K. The unit group modulo f is dened
to be the set of all units 1 mod f which are totally positive and we denote
it by U(f). It is easily seen that U(f) is a group. By Dirichlets unit theorem
there exist r = r
1
+ r
2
1 units
1
, . . . ,
r
and a root of unity in K such
that any U(f) has a unique representation
=
m

n
1
1
. . .
nr
r
with integers m, n
k
. The units
1
, . . . ,
r
are said to be fundamental units of
U(f) although they are not uniquely determined. Dene the matrix
(e
j
log [
(j)
k
[)
1j,kr
, where e
j
:=
_
1 if 1 j r
1
,
2 if r
1
< j r
1
+ r
2
.
Then the regulator R(f) is dened to be the absolute value of the determinant
of this matrix:
R(f) = [ det(e
j
log [
(j)
k
[)[;
it should be noticed that the regulator does not depend on the choice of the
fundamental units
k
.
We further denote by I(f) the multiplicative group generated by all ideals
coprime with f. The principal ray class P(f) is the subgroup of I(f) consisting
of all principal ideals of the form (/) satisfying
0 ,= , O
K
(the ring of integers);
mod f;
/ is totally positive, i.e., all its real conjugates are positive.
The factor group
G(f) := I(f)/P(f)
is called the ray class group mod f, and its elements are called ray classes.
The ray classes are the analogues of the residue classes in the rational number
eld case. One can show that G(f) is a nite abelian group and we denote its
order by h(f).
We shall give a brief example. For the sake of simplicity we shall consider
the number eld Q(

5) and choose f = (1) in which case G((1)) is the class


group. We have already seen in Chapter 1.1.5.3 that there is no unique prime
factorization, and so the class number h = h((1)) is greater than one. One
can deduce from Minkowskis theorem on linear forms that every class of G
contains an integral ideal a with norm
N(a)
_
[ d
K
[,
where d
K
is the discriminant of the number eld, that is in our case d
K
=
20. Obviously, this observation proves the niteness of the class number.
However, we can also use it to get an overview of the structure of the class
group. For this purpose we observe that the only prime ideals p with N(p) 4
Section 4.5 Hecke L-functions 169
can be among the prime ideal divisors of (2) and (3). By the splitting of
primes in quadratic number elds (see again Chapter 1.1.5.3), we nd
(2) = p
2
1
with p
1
= (2, 1 +

5) = (2, 1

5),
(3) = p
2
p

2
with p
2
= (3, 1 +

5) ,= p

2
= (3, 1

5).
Hence, the ideals with norms less than or equal to 4 are p
1
, p
2
, p

2
, and (2) =
p
2
1
. It is easy to see that p
1
is not principal and represents a class of order
two. Furthermore, it is easy to see that all other ideals lie in this class or
are principal. Hence the class number of Q(

5) is two and we have a


description of the associated class group.
Now we are in the position to dene Hecke characters. Suppose we are
given numbers a
j
and
k
satisfying
a
j
0, 1 for 1 j r
1
and a
j
Z for r
1
< j r
1
+ r
2
;

k
R for 1 k r
1
+ r
2
such that
1
+ . . . +
r
1
+r
2
= 0.
Then we dene a function

: K

by

() =
r
1
+r
2

k=1
[
(k)
[
i
k
r
1
+r
2

j=1
_

(j)
[
(j)
[
_
a
j
.
Obviously,

is unimodular. Since the sum of the


k
vanishes, it follows
that

is trivial on Q

. We suppose that the kernel of

contains the
unit group modulo f, i.e.,

() = 1 for any U(f). Then

induces a
character on P(f).
If a non-trivial homomorphism : I(f) C

is identied with

on P(f),
that is
(a) =

() for a = () P(f),
then is said to be a gr ossencharacter modulo f (resp. Hecke character in
some literature). If all numbers a
j
,
k
are equal to zero, then is said to
be a ray class character, and if additionally f = (1), then is an ideal class
group character (that is one of the nitely many characters of the class group
of K). If there exists an ideal f

f and a gr ossencharacter

mod f

such
that =

on I(f), then is said to be induced by

; otherwise is called
primitive and f is said to be the conductor of . (For details from algebraic
number theory we refer to Narkiewicz [159].)
4.5.2. Analytic properties and arithmetic consequences. Now we
return to the associated Hecke L-functions to number elds K. Given a
gr ossencharacter modulo f, we extend to the group I of all fractional
ideals of K by setting (a) = 0 if a is not coprime with f. Then we may
dene (formally) the Dirichlet series and the Euler product appearing in
(4.22). Hecke L-functions to gr ossencharacters are the analogues of Dirichlet
170 Chapter 4 The Selberg class
L-functions: if K = Q, f = (q) with q Z, and

1, then the construction


above leads without the totally positive condition to
G(f) = (Z/qZ)

/1.
If is the trivial (principal) character, then the Hecke L-function for a num-
ber eld K is nothing but the Dedekind zeta-function. Notice that what we
call Hecke L-functions are in some literature called (generalized) Dirichlet
L-functions.
Both the series and the product (4.22) dening L(s, ) are absolutely con-
vergent for > 1 and uniformly in any compact subset. To see this we
recall from class eld theory that in a number eld K of degree n over Q any
rational prime number p has a unique factorization into a product of prime
ideals
(4.23) (p) =
r

j=1
p
e
j
j
with N(p
j
) = p
f
j
and
r

j=1
e
j
f
j
= n;
of course, the integers e
j
, f
j
, and r depend on p (which is not indicated here
for simplicity). Hence we can rewrite (4.22) as an ordinary Euler product
L(s, ) =

p
_
1
(p)
N(p)
s
_
1
=

p
r

j=1
p
j
|(p)
_
1
(p
j
)
p
sf
j
_
1
.
Thus, L(s, ) has a representation as a polynomial Euler product. Hence we
may also rewrite this as an ordinary Dirichlet series:
L(s, ) =

n=1
a(n)
n
s
,
where
a(n) =

p|n

k
1
,...,kr0
k
1
f
1
+...+krfr=(n;p)
r

j=1
(p
j
)
k
j
.
Since the degree of the local Euler factors is bounded by the degree of the
eld extension K/Q, it immediately follows that the Ramanujan hypothesis
holds.
In 1920, Hecke [87] proved that L(s, ) extends to an entire function and
satises a functional equation of Riemann-type provided is primitive. Let
d
K
denote the discriminant of K. We dene
() =
r
1
+r
2

k=r
1
+1
2
i
k
2
, A(f) =
_
[ d
K
[N(f)

n
_1
2
2
r
2
,
and
(s, ) =
r
1

j=1

_
s + a
j
i
j
2
_
r
1
+r
2

j=r
1
+1

_
s +
[a
j
[ i
j
2
_
.
Section 4.5 Hecke L-functions 171
Then
(1 s, ) = ()(s, ),
where () is a complex number with [()[ = 1, depending only on , and
(s, ) := ()A(f)
s
(s, )L(s, ).
Heckes proof of the functional equation is rather complicated; modern proofs
use Tates approach via harmonic analysis (see Chapter 2.1.6). In view of all
the mentioned properties it follows that Hecke L-functions L(s, ) to primi-
tive gr ossencharacters are elements of the Selberg class of degree n = [K : Q].
The Hecke L-function to the trivial character is, as already mentioned, equal
the Dedekind zeta-function and so it is an element of the Selberg class too.
We sketch some arithmetic consequences of the analytic properties of Hecke
L-functions; most of the details can be proved in a similar way as in Chapter 1
(for the distribution of primes in arithmetic progressions or in our applications
of Tauberian theorems). First of all, we notice that L(s, ) does not vanish
on the edge of the critical strip:
(4.24) L(1 + it, ) ,= 0 for t R.
We have already mentioned that L(s, ) is entire and so it is regular at s = 1
unless is trivial. If is not trivial, then

N(p)x
(p) = o
_
x
log x
_
as x ; for trivial we have the prime ideal theorem 1.16. This informa-
tion might be used to verify Selbergs Conjecture A for Hecke L-functions.
First of all we note that
a(p) =
r

j=1
f
j
=1
(p
j
).
Here the summation is taken over the prime ideals of degree one lying above
p; however, this condition is negligible if we deal with

px
[a(p)[
2
p
=

px
r

j=1
[(p
j
)[
2
p
+ O(1),
where we used the orthogonality relations for characters in the last step. Now
the asymptotics of Conjecture A follow by partial summation from (4.25).
In analogy to our studies on the distribution of the prime numbers in prime
residue classes we shall now decompose the ray class group G(f) into its ray
classes C. Here an application of the Tauberian theorem 1.14 leads to
(4.25) p C : N(p) x
1
G(f)
(x).
172 Chapter 4 The Selberg class
Exercise 104. Describe the class group of Q(

10) and the ray class groups mod


(2) of Q(

5) and Q(

10).
Exercise 105. Let K be a quadratic number eld and the nontrivial element of
the Galois group Gal(K/Q). Then K

/Q

: N() = 1 by the map


/

. Show that if K is imaginary, then

() =
_

[[
_

for some integer , and if K is real, then

() =
_

[[
_
a
1
_

[
_
a
2
for some a
1
, a
2
0, 1.
Exercise 106. Give a detailed proof for Selbergs Conjectures A and B for Hecke
L-functions.
Exercise 107. * Show (4.24) and prove the asymptotic formula (4.25).
Hint: apply the Tauberian theorem of Wiener-Ikehara 1.14.
Exercise 108. * Prove a non-trivial zero-free region for Hecke L-functions and
improve the statement of the previous exercise by giving an explicit error estimate
in (4.25).
The nest exercise deals with the Gaussian eld Q(i). Recall from Chapter 1.5.3
that every ideal of the Gaussian ring of integers Z[i] is principle and that the
Gaussian primes are given by = a + bi with N() = a2 + b
2
= p for some
prime number p 1 mod 4.
Exercise 109. ** i) Show that the function
a = ()
m
(a) =
_

[[
_
4im
= exp(4imarg())
for a ,= 0 and any integer m is a primitive grossencharacter.
ii) Prove that the associated Hecke L-function L(s,
m
) satises the functional
equation

s
(s + 2[m[)L(s,
m
) =
s1
(1 s + 2[m[)L(1 s,
m
).
iii) Deduce from the prime number theorem that

||x
_

[[
_
4im
=
m
(1 + o(1))Li (x),
where
m
is equal to 1 if m = 0 and equal to 0 otherwise. Furthermore, show that
the Gaussian primes are equidistributed in sectors:
Z[i] : [[ x, < arg <

Li (x).
Exercise 110. Show that the L-function L
E
1
(s) attached to the elliptic curve
E
1
from Chapter 3.6.3 is indeed a Hecke L-function to a gr ossencharacter of
Q(

1).
Section 4.6 Artin L-functions 173
4.6. Artin L-functions and Artins conjecture
Now we want to study a further class of L-functions which play a central
role in algebraic number theory ever since Artin introduced them in order to
nd higher reciprocity laws. However, rst of all we shall briey motivate
their denition.
4.6.1. A fundamental problem in number theory. In algebraic
number theory, a fundamental problem is to describe how a rational prime
factors into primes in the ring of integers O
K
of an arbitrary number eld
K. Now assume that K is a Galois extension over Q with Galois group
G := Gal(K/Q) (i.e., Q is xed with respect to automorphisms from G).
Then K is the splitting eld of some monic polynomial with rational coe-
cients, and G is the group of eld automorphisms of K xing Q pointwise.
The splitting type of p in O
K
is completely determined by the size of the
subgroup of G which xes any p
j
. For simplicity, assume that the rational
prime p is unramied in K, i.e., the primes p
j
in (4.23) are all distinct, then
these subgroups are all cyclic. Information about the factorization of such p
is encoded in the so-called Frobenius automorphism
p
j
of G, the canonical
generator of the subgroup of G which maps any p
j
into itself. The Frobenius
is determined only up to conjugacy in G; nevertheless, the resulting conju-
gacy class, which we denote by
p
, completely determines the splitting type
of (4.23).
If, for example,
K = Q(i) = a + bi : a, b Q,
then
O
K
= Z[i] = m+ ni : m, n Z.
In this case,
p
is the identity if 1 is a quadratic residue mod p, and the
complex conjugation otherwise. Hence, we may identify G with the subgroup
1 of C

:= C 0 via the homomorphism : G 1:


(
p
) =
_
1
p
_
.
By a part of the quadratic reciprocity law, the Legendre symbol can be
expressed in terms of a congruence condition on p which states for unramied
(odd) primes p
_
1
p
_
= (1)
p1
4
=
_
+1 if p 1 mod 4,
1 otherwise.
Thus, the factorization of p in Z[i] depends only on its residue mod4 (see
also Chapter 1.5.3).
174 Chapter 4 The Selberg class
One goal of class eld theory is to nd a similar description of
p
for
arbitrary Galois extensions K. In general, one cannot expect that there
exists a modulus q such that
p
is the identity if and only if p lies in some
arithmetic progression mod q. However, if K is abelian, i.e., G = Gal(K/Q)
is abelian, and : G C

is a homomorphism, then it is known that there


exists a Dirichlet character mod q such that
: (Z/qZ)

with (
p
) = (p) (4.26)
for all primes p, unramied in K. This is a reformulation of the famous
Kronecker-Weber theorem (stating that any nite abelian extension of Q is
contained in some cyclotomic eld Q(exp(
2i
n
)). It follows that the splitting
properties of p in K depend only on its residue modulo some xed number q
depending on K. In particular, this implies the general quadratic reciprocity
law of Gauss. As a matter of fact, the factorization of Dedekind zeta-functions

K
(s) = (s)L(s, ) with K = Q(
_
(1)q)
for all quadratic elds K is equivalent with quadratic reciprocity. Artins
reciprocity law of abelian class eld theory gives an extension of (4.26) for
abelian elds.
What can be said for nonabelian Galois extensions? Recognizing the utility
of studying groups in terms of their matrix representations, Artin focused
attention on homomorphisms
: G = Gal(K/Q) GL
m
(C),
i.e., m-dimensional representations of the Galois group G; note that one-
dimensional representations are simply characters. Artin transferred the
problem of analyzing conjugacy classes in G to the analogous problem in
GL
m
(C), where the corresponding classes are completely determined by their
characteristic polynomials
det
_
1
(
p
)
p
s
_
,
where 1 denotes here the unitary matrix. Introducing the so-called Artin
L-function
L(s, ) =

p
det
_
1
(
p
)
p
s
_
1
(we give a precise denition in the following section), Artin was able to reduce
the problem to one involving these analytic objects: is it possible to dene
L(s, ) in terms of the arithmetic of Q alone? It was in this context that
Artin proved his reciprocity law. Indeed, for abelian K and one-dimensional
, Artin showed that L(s, ) is identical to a Dirichlet L-function L(s, ) with
Section 4.6 Artin L-functions 175
an appropriate character mod q. Since an identity between two Euler prod-
ucts implies an identity between the local Euler factors (by the uniqueness
of the Dirichlet series expansion), this yields Artins reciprocity law.
4.6.2. Artin L-functions. Let L/K be a Galois extension of number
elds with Galois group G. Further, let : G GL
m
(V ) be a representation
(group homomorphism) of G on a nite dimensional complex vector space V .
In order to give the denition of the Artin L-function attached to these data,
we recall some facts on prime ideals in number elds and their ramication
in Galois extensions. (For the details from algebraic number theory we refer
once more to Narkiewicz [159].)
For each prime p of K, and a prime P of L with P[p, we dene the decom-
position group by
D
P
= G : P

= P = Gal(L
P
/K
p
),
where L
P
and K
p
are the completions of L at P and K at p, respectively.
Denote by k
P
/k
p
the residue eld extension. By Hensels lemma, we have a
surjective map from D
P
to Gal(k
P
/k
p
); its kernel I
P
is the inertia group at
P, dened by
I
P
= G : () mod P for all O
L
.
We thus have an exact sequence
1 I
P
D
P
Gal(k
P
/k
p
) 1.
Hence, there is an isomorphism
D
P
/I
P
Gal(k
P
/k
p
).
Now k
P
/k
p
is a Galois extension of nite elds, and hence the group
Gal(k
P
/k
p
) is cyclic, generated by the map
N(p)
, where N(p), the abso-
lute norm of p, is the cardinality of k
p
. We can choose an element
P
D
P
whose image in Gal(k
P
/k
p
) is this generator; this
P
is called Frobenius ele-
ment at P, i.e.,

P
()
N(p)
mod P
for all O
L
. Note that the Frobenius element is only dened modI
P
.
For unramied p (and in particular, these are all but nitely many p), the
Frobenius is well-dened since I
P
= 1. The action of the Galois group
on the set of primes in L above p is transitive, and thus for any pair of
primes P
1
and P
2
lying above p, there exists an automorphism in G which
simultaneously conjugates D
P
1
into D
P
2
, I
P
1
into I
P
2
, and
P
1
into
P
2
. This
implies an identity for the characteristic polynomials of
P
j
on the subspace
V
P
j
of V on which I
P
j
acts trivially:
det
_
1
(
P
1
)
N(p)
s

V
P
1
_
= det
_
1
(
P
2
)
N(p)
s

V
P
2
_
.
176 Chapter 4 The Selberg class
Thus, these characteristic polynomials are independent of the choice of
P
.
Denote by
p
the conjugacy class of Frobenius elements at primes P above
p; in case of unramied p the inertia group is trivial, and
p
is called Artin
symbol.
Following Artin [3], we dene the Artin L-function attached to by
L(s, , L/K) =

p
det
_
1
(
p
)
N(p)
s

V
P
_
1
, (4.27)
where p runs through the prime ideals of the ring of integers in K; this Euler
product converges for > 1.
The zeta-function of a eld is like the atom of physics. (. . .)
we will show how to split it via group theory.
This is a quotation of H.M. Stark [188] and in the following section we
illustrate the just given construction by one of his explicit examples.
4.6.3. An example. We consider the eld K = Q(2
1
3
). Notice that K
is not normal over Q (since the polynomial X
3
2 has only one of its roots
in K). We write
= 2
1
3
, = e
2i
3
2
1
3
, = e
4i
3
2
1
3
.
The eld L = Q(, e
2i
3
) = Q(, , ) is normal over Q of degree 6. Since
automorphisms of L are determined by their action on , and , we nd
that the Galois group of L is given by
G = Gal(L/Q) = 1, (), (), (), (), (),
which is the symmetric group on three letters. The splitting of primes from
Q to K, and likewise from K to L, is ruled by the Frobenius automorphisms.
Suppose that P is an unramied prime of L which lies above p of K which in
turn lies above the rational prime p of Q. Then the Frobenius automorphism
of P relative to Q is given by one of the following conjugacy classes:

P
= 1. Since the Frobenius has order one, by (4.23), there are 6
primes in L above p. Obviously,
P
Gal(L/K) = 1, (). In this
case p splits in K into three dierent primes p
j
(1 j 3) each of
which splits into two prime ideals P
k
(1 k 6) of L.

P
is in the conjugacy class (), (), () of elements of order
two. We may choose P such that
P
= () Gal(L/K). The
f = 2 in (4.23) and so there are three second degree primes P
k
(1 k 3) above p; we may assume that P = P
1
. We observe
that the Frobenius automorphism of P relative to K is equal to

P
. Hence, we nd N(P) = N(p)
2
and N(p) = p for some prime
p = p
1
of K. For the other two primes P
2
and P
3
the Frobenius
Section 4.6 Artin L-functions 177

P
is equal to () and (), respectively. In these cases we nd

2
P
= 1 Gal(L/K) and N(P) = N(p) and N(p) = p
2
for some prime
p = p
2
of K. Thus, the primes P
2
and P
3
have relative degree one
over a single prime p
2
of K (which is of degree two).

P
is in the conjugacy class (), () of elements of order
three. In this case we have f = 3 in (4.23) and there two third
degree primes P
1
and P
2
of L above p, for one of them
P
= ()
and for the other
P
= (). In both cases neither
P
nor
2
P
lie
in Gal(L/K) = 1, () and so both P
1
and P
2
lie above a single
prime p of K (which must be of degree 3).
Now we want to compute the associated Artin L-functions. First of all we
have a look on every individual Euler factor. Since the eld extension K/Q
has degree 3, there are the following possibilities to consider.
The prime p splits completely into three dierent prime divisors; e.g.,
(31) = p
1
p
2
p
3
with
p
1
= (31, 4), p
2
= (31, 7), p
3
= (31, 20).
In this case the local Euler factor at p is of the form
(4.28)
_
1
1
p
s
_
3
= det
_
_
1
_
_
1 0 0
0 1 0
0 0 1
_
_
1
p
s
_
_
1
.
Obviously, the appearing matrix has the eigenvalue +1 with multi-
plicity 3.
The prime p can be factored into a product of two factors, one of
degree one and one of degree two; for example, (5) = p
1
p
2
with
p
1
= (5, 3), p
2
= (5,
2
+ 3 + 9).
Here we have
_
1
1
p
s
_
1
_
1
1
p
2s
_
1
= det
_
_
1
_
_
0 1 0
1 0 0
0 0 1
_
_
1
p
s
_
_
1
= det
_
_
1
_
_
0 0 1
0 1 0
1 0 0
_
_
1
p
s
_
_
1
(4.29)
= det
_
_
1
_
_
1 0 0
0 0 1
0 1 0
_
_
1
p
s
_
_
1
.
The eigenvalues of the (similar) matrices are 1 and +1 with mul-
tiplicities one and two, respectively.
178 Chapter 4 The Selberg class
The prime p is a prime ideal of third degree; e.g., (7) = p. In this
case we have
_
1
1
p
3s
_
1
= det
_
_
1
_
_
0 1 0
0 0 1
1 0 0
_
_
1
p
s
_
_
1
= det
_
_
1
_
_
0 0 1
1 0 0
0 1 0
_
_
1
p
s
_
_
1
. (4.30)
Here the eigenvalues of the (similar) matrices are the third roots of
unity.
Before we continue we remark that the splitting of primes can be computed
by the following statement: suppose that g(X) is the minimal polynomial of
K over Q and that it splits factors mod p into irreducible pieces as
g(X) g
1
(X)
e
1
. . . g
r
(X)
er
mod p.
If the power of p in the polynomial discriminant of g(X) is the same as the
power of p in the relative discriminant D
L/K
of L/K, then p splits in L as
p = P
e
1
1
. . . P
er
r
,
where P
j
= (p, g
j
()) is of relative degree deg g
j
. This togehter with Eisen-
steins irreducibility criterion gives the basic tools to do arithmetic compu-
tations in number elds.
We may represent the Galois group G by matrices as follows. For g G we
write
_
_

_
_
g = M(g)
_
_

_
_
,
where M(g) is the permutation matrix corresponding to g. Thus we can
represent the six elements of G by
1
_
_
1 0 0
0 1 0
0 0 1
_
_
, ()
_
_
0 1 0
0 0 1
1 0 0
_
_
, ()
_
_
0 0 1
1 0 0
0 1 0
_
_
,
()
_
_
0 1 0
1 0 0
0 0 1
_
_
, ()
_
_
0 0 1
0 1 0
1 0 0
_
_
, ()
_
_
1 0 0
0 0 1
0 1 0
_
_
.
The map : g M(g) denes a homomorphism: M(gh) = M(g)M(h); it is
an example of a three dimensional permutation representation of the group
G. The conjugacy classes of the symmetric group on , , are precisely the
conjugacy classes of Frobenius automorphisms arising from prime numbers
Section 4.6 Artin L-functions 179
which split in the indicated form and for each of them we observe via (4.28)-
(4.30) that the associated Euler factors are of the form as predicted by (4.27).
Now we want to introduce a more convenient notation of Artin L-functions.
To any representation of G, we can attach a character of G by setting
(g) = trace((g))
for g G. The degree of a character is dened by deg = (1). If h is
another element of G, then
(h
1
gh) = (h)
1
(g)(h),
so that (h
1
gh) and (g) are similar matrices and thus have the same trace.
This shows that characters of G are constant on the conjugacy classes. Two
representations are said to be equivalent if they have the same character. If

1
and
2
are representations of G with characters
1
and
2
, then
(g) =
_

1
(g) 0
0
2
(g)
_
also denes a representation of G with character
1
+
2
, and in this case
is said to be reducible; any representation which is not reducible is called
irreducible. We shall use the same attributes for the associated character.
It turns out that any conjugacy class of G corresponds to an irreducible
representation and one can show that there are not more; of course, distinct
irreducible representations are non-equivalent (these observations are analo-
gous to the case of Dirichlet characters and the group of residue classes of
Z).
In our example we nd for the the three conjugacy classes of G:
1 (), () (), (), ()

0
+1 +1 +1

1
+1 +1 1

2
+2 1 0
Hence there are three irreducible characters (in some literature simple char-
acters): we are dealing with the trivial character
0
, another character
1
of
degree one, and a character
2
of degree two.
It is easily seen that our characters satisfy the orthogonality relations, that
are
1
G

G
(C)(D) =
_
(C)
1
if C = D,
0 otherwise,
where C and D are two conjugacy classes, and
1
G

gG
(g)(g) =
_
1 if = ,
0 otherwise.
180 Chapter 4 The Selberg class
Since the Euler factors in (4.27) depend only on the conjugacy class
p
, in
the sequel we will talk sometimes in terms of characters and denote the Artin
L-function (4.27) by L(s, , L/K) (and sometimes we shall even write L(s, )
for short). To illustrate this we continue with our example. We can construct
more characters from the irreducible characters listed above, for example,
a third degree character related to the permutation representation ().
Taking the character relations into account we nd =
0
+
2
. For the
related Artin L-functions we note that
L(s, , L/K) = L(s,
0
+
2
, L/K) = L(s,
0
, L/K)L(s,
2
, L/K).
For the eld L = Q(, , ) there are four subelds up to conjugacy. First
of all the eld Q itself, xed by all of G, second Q(

3) xed by G
1
:=
1, (), (), third K = Q(2
1
3
) xed by G
2
:= 1, (), and nally L
xed just by 1.
L 1
K = Q(2
1
3
)

G
2
()
r
r
r
r
r
r
r
r
r
r
r
r
r
Q(

3)

G
1
(),()

Gal(L/Q) = S
3

r
r
r
r
r
r
r
r
r
r
r
r
We obtain the following factorizations of the associated Dedekind zeta-
functions into products of Artin L-functions to L/Q:
(s) =
Q
(s) = L(s,
0
),

Q(

3)
(s) = L(s,
0
) L(s,
1
),

Q(2
1
3 )
(s) = L(s,
0
) L(s,
2
),

L
(s) = L(s,
0
) L(s,
1
) L(s,
2
).
We observe that any of the Dedekind zeta-functions on the left-hand side is di-
visible by the Riemann zeta-function (in the sense that their quotient is an en-
tire function). It follows from these factorizations and the analytic behaviour
of Dedekind zeta-functions that each of the involved Artin L-functions with
,=
0
possesses a meromorphic continuation to the whole complex plane;
the only possible poles can occur at zeros of other Artin L-functions. Fur-
thermore we can deduce functional equations of the Riemann-type. This is a
rather remarkable new way to deduce analytic properties for L-functions!
Section 4.6 Artin L-functions 181
Furthermore, we see that the Dedekind zeta-functions are algebraically
dependent:

Q(

3)
(s)
Q(2
1
3 )
(s) =
Q
(s)
K
(s).
It is an interesting question to which extent the Dedekind zeta-function
determines the eld. One can show that the Dedekind zeta-function
K
(s) de-
termines the minimal normal extension L of Q containing K and thus we have
to ask whether there exist non-conjugate subgroups of Gal(L/Q) giving the
same induced trivial character. This is indeed possible! Two number elds
K
1
and K
2
are said to be arithmetically equivalent if their Dedekind zeta-
functions are the same. The rst example of arithmetical equivalent elds
was given by Gassmann [57]. Perlis [167] proved that arithmetically equiva-
lent non-isomorphic elds have at least degree 7 and that this bound cannot
be improved. An explicit example of degree 8 is for instance Q((3)
1/8
) and
Q((48)
1/8
) which is due Perlis & Schinzel [168].
4.6.4. The Artin conjecture. One of the most fundamental conjec-
tures in algebraic number theory is
Artins Conjecture. Let L/K be a nite Galois extension with Galois group
G. For any irreducible character ,= 1 of G the Artin L-function L(s, , L/K)
extends to an entire function.
We discuss briey one of its important consequences. Dedekinds conjecture
claims that the quotient
L
(s)/
K
(s) is entire provided L/K is an extension
of number elds, not necessarily Galois. If L/K is a Galois extension, then
the so-called Artin-Takagi factorization gives a factorization of the Dedekind
zeta-function of a number eld relative to a subeld (see Heilbronns survey
[91]); more precisely,
L(s, 1, L/K) =
K
(s), and L(s, R
G
, L/K) =
L
(s),
where R
G
is the regular character of G (the character dened by

(1)),
and

L
(s) =

G
L(s, , L/K)
(1)
,
where

G denotes the set of irreducible characters of G. In case of Galois
extensions L/K, the Aramata-Brauer theorem (see Heilbronn [91] or Murty
& Murty [157], 2.3) yields the truth of Dedekinds conjecture; its proof
relies mainly on the Artin-Takagi factorization. In the general case, if L/K is
a nite (not necessarily Galois) extension, then Dedekinds conjecture follows
from Artins conjecture by studying the normal closure of L/K.
182 Chapter 4 The Selberg class
As indicated in the last but one section, Artin proved his conjecture if
is one-dimensional and L/K is abelian. In this case, the related Artin
L-function coincides with a Hecke L-function.
Theorem 4.17. Let L/K be abelian and let ,= 1 be an irreducible character
of G = Gal(L/K). Then there exists a Hecke grossencharacter such that
L(s, , L/K) = L(s, ).
Artin proved this theorem by means of class eld theory and, in particular,
Chebotarevs density theorem. We shall briey explain the latter result. Let
L/K be a nite Galois extension with Galois group G and let C be a subset
of G, closed under conjugation. Further, denote by
C
(x) the number of
prime ideals p of K, unramied in L, for which
p
C and which have norm
N(p) x in K. Then, Chebotarevs density theorem [31] states

C
(x)
C
G
(x). (4.31)
This rather deep theorem can be seen as a higher analogue of the prime
number theorem in arithmetic progressions. A modern proof can be found,
for example, in Narkiewicz [159]. The Chebotarev density theorem can be
used to determine the Galois group of a given irreducible polynomial P(X)
of degree n by counting the number of unramied primes up to a certain
bound for which P factors in a certain way and comparing the results with
the fractions of elements of each of the transitive subgroups of the symmetric
group S
n
with the same cyclic structure; see Lenstra & Stevenhagen [131]
for details.
Brauer [25] proved a functional equation for Artin L-functions which gives
a meromorphic continuation throughout the complex plane (see also Neukirch
[160]), VII.12). However, the holomorphy of nonabelian Artin L-functions
is still unproved (especially inside the critical strip). In certain particular
cases the Artin conjecture is known to be true, at least conditionally. M.R.
Murty [155] proved
Theorem 4.18. Selbergs Conjecture B implies Artins conjecture.
M.R. Murty & Perelli [158] replaced Selbergs conjecture by the pair corre-
lation conjecture (as already mentioned in the previous section).
The proof uses some easy properties of Artin L-functions which we did
not prove or even did not mention above. The reader may have a look into
the literature, e.g., Heilbronn [91], and may consult the examples from the
previous section.
Proof. Let

L be the normal closure of L over Q. Then,

L/K and

L/Q are
Galois. Thus, can be thought as a character of Gal(

L/K), and by the


Section 4.6 Artin L-functions 183
properties of Artin L-functions it turns out that
L(s, ,

L/K) = L(s, , L/K).
Brauers induction theorem [25] (see again Neukirch [160], VII.10) states,
roughly, that any character of a nite group G is a N
0
-linear combination
of certain induced one-dimensional characters of subgroups of G. Thus, by
the induction of from Gal(

L/K) to Gal(

L/Q), it follows that


L(s, , L/K) =

L(s, ,

L/Q)
m()
,
where the product is taken over all irreducible characters of Gal(

L/Q)
and m() are nonnegative integers. To prove Artins conjecture, it suces
to show that all appearing L(s, ,

L/Q) are entire. By Brauers induction
theorem and Artins reciprocity law, Theorem 4.17,
L(s, ,

L/Q) =
L(s,
1
)
L(s,
2
)
,
where
1
,
2
are characters of Gal(

L/Q) and L(s,


1
), L(s,
2
) are products
of Hecke L-functions (4.22). Since Hecke L-functions belong to the Selberg
class o, and o is multiplicatively closed, the functions L(s,
1
) and L(s,
2
)
belong to o too. Now, by Theorem 4.12, there exist primitive functions
/
j
o such that
L(s, ,

L/Q) =
f

j=1
/
j
(s)
e
j
, (4.32)
where e
j
Z. By comparing the p-th coecient in the Dirichlet series ex-
pansions of both sides, we get
(p) =
f

j=1
e
j
a
L
j
(p).
Thus,

px
[(p)[
2
p
=

px
1
p

j=1
e
j
a
L
j
(p)

2
. (4.33)
Selbergs conjecture B yields the asymptotic formula

px
1
p

j=1
e
j
a
L
j
(p)

2
=
_
f

j=1
e
2
j
_
log log x + O(1). (4.34)
Next, we decompose the sum on the left hand side of (4.33) according to
the conjugacy classes C of G := Gal(

L/Q) to which the Frobenius element


p
184 Chapter 4 The Selberg class
belongs. If g
C
denotes any element of C, this leads to

px
[(p)[
2
p
=

C
[(g
C
)[
2

px
pC
1
p
.
By partial summation, we deduce from Chebotarevs density theorem (4.31)

px
pC
1
p
=
C
G
log log x + O(1).
This gives

px
[(p)[
2
p
=

C
[(g
C
)[
2
C
G
log log x + O(1).
Since is irreducible, we get

C
[(g
C
)[
2
C
G
=
1
G

C
C = 1,
which implies with (4.34) and (4.33)
f

j=1
e
2
j
= 1.
Thus, f = 1 and e
1
= 1. The case e
1
= 1 implies
L(s, ,

L/Q) =
1
/
1
(s)
,
which is impossible since L(s, ,

L/Q) has trivial zeros (their existence follows
from their functional equation). Hence, e
1
= +1, and we conclude that
L(s, ,

L/Q) = /
1
(s) is entire.
The proof shows that if is an irreducible non-trivial character of Gal(K/Q),
then the Artin L-function L(s, , K/Q) is an element of the Selberg class o
if Selbergss Conjecture B is true. Moreover, one can easily show that under
these assumptions it is even primitive.
Exercise 111. Deduce from the discussion of the arithmetic of the Gaussian num-
ber eld Q(i) an old statement of Fermat and Euler which states that an odd prime
p has a representation as a sum of two integer squares if and only if p 1 mod 4.
Exercise 112. Derive the functional equation for the Artin L-functions
L(s,
j
, L/Q) for j = 1, 2, 3 in the example L = Q(2
1
3
, e
2i
3
) from Section 4.6.3.
Hint: use the functional equation for Dedekind zeta-functions.
Exercise 113. * Construct all Artin L-functions to Q(

10) and Q(

2 +

3).
Section 4.7 Langlands program 185
Exercise 114. Consider the eld K := Q(

1,

5). Show that it has three


dierent subelds of degree 2 over Q, namely Q(

1), Q(

5), and Q(

5), and
verify the identity
(4.35)
K
(s)
Q
(s)
2
=
Q(

1)
(s)
Q(

5)
(s)
Q(

5)
(s).
Prove that if the Galois group Gal(K/Q) has more normal subgroups than conjugacy
classes, then there exist algebraic relations for the corresponding Dedekind zeta-
functions.
Exercise 115. Prove that Selbergs Conjecture B implies that an Artin L-function
L(s, , K/Q) with an irreducible non-trivial character is primitive.
Hint: consider the corresponding integer n
L
in Conjecture A and recall an old
exercise.
4.7. Langlands program
The Langlands program has emerged in the late 60s of the last century in a
series of far-reaching conjectures tying together seemingly unrelated objects
in number theory, algebraic geometry, and the theory of automorphic forms.
These disciplines are linked by Langlands L-functions associated with auto-
morphic representations, and by the relations between the analytic properties
and the underlying algebraic structures. There are two kinds of L-functions:
motivic L-functions which generalize Artin L-functions and are dened purely
arithmetically, and automorphic L-functions, dened by transcendental data.
In its comprehensive form, an identity between a motivic L-function and an
automorphic L-function is called a reciprocity law. Langlands reciprocity
conjecture claims, roughly, that every L-function, motivic or automorphic, is
equal to a product of L-functions attached to automorphic representations.
For an introduction to the Langlands program we refer to the excellent sur-
veys of Gelbart [58], M.R. Murty [154], and Langlands lecture [126] at the
International Congress in Helsinki.
4.7.1. Automorphic representations. At the heart of Langlandss
program is the notion of an automorphic representation and its L-function
L(s, ). It is beyond the scope of these notes to dene these objects (both de-
ned via group theory and the theory of harmonic analysis on ad`ele groups)
in an appropriate way.
Let K be a number eld (one looses not too much by restricting to Q).
For each absolute value on K, there is a completion K

of K which is R,
C, or a p-adic eld, where p is a prime ideal in K. Denote by O

the ring
of integers in K

. In discussing local-global problems it is often necessary


to consider several places simultaneously. At rst sight it seems natural to
186 Chapter 4 The Selberg class
form the product of all the K

which is a topolgical ring, but it does not have


satisfactory compactness properties. Since any K is a p-adic integer for
almost all p, we restrict to elements
=

,
where

lies in O

for all but nitely many places ; such elements are called
ad`ele. The ad`ele form a set-theoretic (restricted) product. This product is a
topological ring, the ad`ele ring A
K
of K. One can think of K as embedded in
A
K
via the map (, , . . .).
For m 1 let GL
m
(A
K
) be the group of m m matrices over A
K
whose
determinant is a unit in A
K
. By the product topology of the ad`ele ring,
GL
m
(A
K
) becomes a locally compact group in which GL
m
(K), embedded di-
agonally, is a discrete subgroup of GL
m
(A
K
). A character of K

GL
1
(A
K
)
is called gr ossencharacter, where K

:= K 0. For a xed gr ossencharacter


we consider the Hilbert space
L
2
:= L
2
(GL
m
(K) GL
m
(A
K
), )
of measurable functions f on GL
m
(K) GL
m
(A
K
) satisfying the conditions
f(zg) = (z)f(g) for any z Z, g GL
m
(K) GL
m
(A
K
);
the integral
_
ZGLm(K)\GLm(A
K
)
[f(g)[
2
dg
is bounded.
Elements f L
2
generalize the concept of twisted modular forms to discrete
subgroups of the full modular group. In order to introduce a subspace of cusp
forms we have to consider appropriate subgroups. Any parabolic subgroup
P of GL
m
(1), where 1 is a commutative ring with identity, has a decompo-
sition, called the Levi decomposition, of the form P = MN, where N is the
unipotent radical of P; M is called the Levi component of P. We denote the
unipotent radical of P in the Levi decomposition of a parabolic subgroup P
in GL
m
(1) by N
P
(1).
The subspace of cusp forms
L
2
0
:= L
2
(GL
m
(K) GL
m
(A
K
), )
of L
2
is dened by the additional vanishing condition
for all parabolic subgroups P of GL
m
(A
K
) and every g GL
m
(A
K
),
_
N
P
(K)\N
P
(A
K
)
f(ng) dn = 0.
The right regular representation R of GL
m
(K) on L
2
is given by
(R(g)f)() = f(g)
Section 4.7 Langlands program 187
for each f L
2
and any , g GL
m
(A
K
). An automorphic representation
is a subquotient of the right regular representation of GL
m
(A
K
) on L
2
, and
a cuspidal automorphic representation is a subrepresentation of the right
regular representation of GL
m
(A
K
) on L
2
0
.
A representation of GL
m
(A
K
) is called admissible if its restriction to the
maximal subgroup
K :=

complex
U
m
(C)

real
O
m
(R)

finite
GL
m
(O

)
contains each irreducible representation of K with nite multiplicity; here
U
m
and O
m
denote the groups of unitary and orthogonal m m matrices,
respectively. A representation of a group G is called irreducible if it cannot
be decomposed into the direct sum of two representations; an irreducible
character is the character associated with an irreducible representation.
Now let be an irreducible, admissible, cuspidal automorphic representa-
tion of GL
m
(K). Then can be factored into a direct product =

,
where ranges over all (nite and innite) places of K, and each

is an
irreducible representation of GL
m
(K

). For all but a nite number of places


the representation

is unramied (that means the quotient obtained by


inducing a quasi-character from the Borel subgroup of GL
m
(K

) to GL
m
(K

)
is unique).
4.7.2. General L-functions. In order to dene the L-function attached
to an automorphic representation we dene the local Euler factors for non-
archimedean (nite) unramied places by
L

(s, ) = det
_
1
A

N(p)
s
_
1
,
where A

is the semisimple conjugacy class corresponding to

and p is
the prime ideal of K belonging to the place . We do not explain here the
rather technical denition of the Euler factors L

(s, ) for ramied places .


However, any Euler factor L

(s, ) for a non-archimedean place associated


with the prime ideal p, unramied or not, can be rewritten as
L

(s, ) =
m

j=1
_
1

j
(p)
N(p)
s
_
1
, (4.36)
where the numbers
j
(p) for 1 j m are so-called Satake, resp. Lang-
lands parameters, determined from the local representations

. At the
archimedean (innite) places we put for certain numbers
j
()
L

(s, ) =
m

j=1

(s
j
())
188 Chapter 4 The Selberg class
with

(s) :=
_

s
2

_
s
2
_
if K

R,
(2)
s
(s) if K

C;
(4.37)
where, again, the appearing numbers
j
() for 1 j m are determined
from the local representations

. Then the global L-function associated with


is given by
L(s, ) =

nonarchimedean
L

(s, ),
and the completed L-function is dened by
(s, ) = L(s, )

archimedean
L

(s, ).
By the work of Hecke [88], Jacquet & Langlands [102], and Godement &
Jacquet [60] we have
Theorem 4.19. Let K be a number eld and be an irreducible, admis-
sible, cuspidal automorphic representation of GL
m
(A
K
). Then (s, ) has
a meromorphic continuation throughout the complex plane and satises the
functional equation
(s, ) =

N
s
1
2

(1 s, ),
where is the contragredient representation of , N

N is the conductor of
and

is the root number (these quantities are completely determined by


the local representations). (s, ) is entire unless m = 1 and is trivial, in
which case it has a pole at s = 1.
For m = 1 one simply obtains the Riemann zeta-function, Dirichlet L-
functions and Hecke L-functions attached to gr ossencharacters, whereas for
m = 2 one gets L-functions associated with newforms. The similarities be-
tween these general L-functions and those of the Selberg class are obvious.
On one hand we have the Selberg class dened by axioms which are known
to be the most common pattern of many L-functions in number theory, on
the other hand we have Langlands construction of general L-functions out
of group representations.
4.7.3. Langlands conjectures. In the 1960s Langlands started his vi-
sionary program which might be understood as a continuation of the famous
Artin conjecture. One of his central conjectures claims that all zeta-functions
arising in number theory are special realizations of L-functions to automor-
phic representations constructed above.
Langlands reciprocity conjecture. Suppose L is a nite Galois extension
of a number eld K with Galois group G, and : G V is an irreducible
Section 4.7 Langlands program 189
representation of G, where V is an m-dimensional vector space. Then there
exists an automorphic cuspidal representation of GL
m
(A
K
) such that
L(s, , L/K) = L(s, ).
This means that there are identies between certain L-functions, which are a
priori of dierent type! Since Hecke gr ossencharacters are automorphic rep-
resentations of GL
1
(A), Artins conjecture is a special case of the Langlands
reciprocity conjecture. By Artins work, if m = 1 and L/K is abelian, Lang-
lands reciprocity law is settled by means of class eld theory. In the case of
function elds, the Langlands conjecture has been proved by Drinfeld [48] in
dimension two, and recently by Laorgue [123] for arbitrary dimension (for
which both of them were awarded with a Fields medal).
Now we consider the local Euler factors of L-functions attached to auto-
morphic representations. Petersson [169] extended Ramanujans conjecture
on the values of the -function to modular forms. Delignes estimate (3.32)
proved the desired bound for newforms but it is expected that an analogue
should hold for all L-functions of arithmetical nature.
Ramanujan-Petersson Conjecture. Let be a cuspidal automorphic
representation of GL
m
(A
K
) which is unramied at a place . If is non-
archimedean, then
[
j
(p)[ = 1 for 1 j m,
where p is the prime ideal associated with the place . If is archimedean,
then Re
j
() = 0 for 1 j m.
The Ramanujan-Petersson conjecture might look very restrictive on the rst
view, but it is nothing else than the local analogue of the Grand Riemann
hypothesis. We refer to Iwaniec & Sarnak [100] for details and the current
knowledge concerning this conjecture.
We shall speculate a little bit about all these widely believed conjectures
and the axioms dening the Selberg class and the subclass

o, in particu-
lar. It is expected that all functions in the Selberg class are automorphic
L-functions. If / o is primitive and automorphic, then it is also attached
to an irreducible automorphic representation. Conversely, every irreducible
automorphic representation should give a primitive function in o. This is
not known in general, but it has been proved by M.R. Murty [155, 156]
for GL
1
and GL
2
. The axioms on the analytic continuation and on the func-
tional equation follow immediately from Theorem 4.19. The polynomial Euler
product in the denition of the subclass

o ts (by the splitting of primes in
K) perfectly to the Euler product of Langlands L-functions attached to au-
tomorphic representations (4.36) and the Ramanujan-Petersson conjecture.
190 Chapter 4 The Selberg class
Finally, let us notice that the Euler factor at the innite places (4.37) is of
the form, predicted by the strong -conjecture (from Section 4.2). Of course,
all these axioms and the hypotheses too, are deduced from known examples
of L-functions in number theory, and so they have to share certain patterns.
Anyway, we are led to see a close connection between Langlands general
L-functions and the elements of the Selberg class.
M.R. Murty [155] proved
Theorem 4.20. Assume that Selbergs Conjecture B is true.
i) If is an irreducible cuspidal automorphic representation of GL
m
(A
Q
)
which satises the Ramanujan-Petersson conjecture, then L(s, ) is a primi-
tive function in o.
ii) If K is a Galois extension of Q with solvable Galois group G, and if
is an irreducible character of G of degree m, then there exists an irreducible
cuspidal automorphic representation of GL
m
(A
Q
) such that
L(s, ) = L(s, ).
The rst assertion identies certain L-functions to automorphic representa-
tions as being primitive functions in the Selberg class subject to the truth of
Selbergs conjecture B and the Ramanujan-Petersson conjecture. The sec-
ond assertion of the theorem is Langlands reciprocity conjecture if K/Q is
solvable. Murtys proof shows that if the Dedekind zeta-function of K is the
L-function of an automorphic representation over Q, then Selbergs Conjec-
ture B implies Langlands reciprocity conjecture.
Some concluding words: We started with factorizations of integers (rst in Q
and then in number elds) into irreducible elements and derived asymptotic
laws for the atoms in these products (that were primes or prime ideals) by
studying the analytic properties of the generating functions (Euler products).
Finally, we considered factorizations of more complicated higher L-functions
(Artin vs. Hecke L-functions) into primitive functions and deduced analytic
properties (which are not unrelated to number theory; however this came a
bit short here) by applying arithmetical laws. A fruitful see-saw!
Bibliography
[1] R. Apery, Irrationalite de (2) et (3), Asterisque 61 (1979), 11-13
[2] T.M. Apostol, Introduction to analytic number theory, Springer 1976
[3] E. Artin,

Uber eine neue Art von L-Reihen, Abh. math. Sem. Univ. Hamburg 3
(1923), 89-108
[4] A.O.L. Atkin, J. Lehner, Hecke operators on
0
(m), Math. Ann. 185 (1970),
134-160
[5] R. Ayoub, Euler and the zeta function, Amer. Math. Monthly 81 (1974), 1067-
1086
[6] R. Backlund,

Uber die Beziehung zwischen Anwachsen und Nullstellen der
Zetafunktion,

Ofversigt Finska Vetensk. Soc. 61 (1918/19), no. 9
[7] B. Bagchi, The statistical behaviour and universality properties of the Riemann
zeta-function and other allied Dirichlet series, Ph.D.Thesis, Calcutta, Indian Sta-
tistical Institute, 1981
[8] B. Bagchi, A joint universality theorem for Dirichlet L-functions, Math. Z. 181
(1982), 319-334
[9] B. Bagchi, Recurrence in topological dynamics and the Riemann hypothesis,
Acta Math. Hungar. 50 (1987), 227-240
[10] A. Baker, Linear Forms in the Logarithms of Algebraic Numbers. I, Mathe-
matika 13 (1966), 204-216
[11] E.P. Balanzario, Remark on Dirichlet series satisfying functional equations,
Divulg. Mat. 8 (2000), 169-175
[12] R. Balasubramanian, K. Ramachandra, Proof of some conjectures on the
mean-value of Titchmarsh series. I, Hardy-Ramanujan J. 13 (1990), 1-20
[13] P.T. Bateman, R.A. Horn, A heuristic asymptotic formula concerning the
distribution of prime numbers, Math. Comp. 16 (1962), 363-367
[14] P. Bauer,

Uber den Anteil der Nullstellen der Riemannschen Zeta-Funktion auf
der kritischen Geraden, diploma thesis, Frankfurt University 1992, available at
www.math.uni-frankfurt.de/pbauer/diplom.ps
[15] P. Bauer, Zeros of Dirichlet L-series on the critical line, Acta Arith. 93 (2000),
37-52
[16] J. Beineke, D. Bump, Moments of the Riemann zeta function and Eisenstein
series. I, J. Number Theor. 105 (2004), 150-174
[17] B. Birch, Conjectures concerning elliptic curves, Proc. Symp. Pure Math., Amer.
Math. Soc., Providence 1965, 106-112
[18] S. Bochner, On Riemanns functional equation with multiple gamma factors,
Ann. Math. 67 (1958), 29-41
[19] H. Bohr,

Uber eine quasi-periodische Eigenschaft Dirichletscher Reihen mit An-
wendung auf die Dirichletschen L-Funktionen, Math. Ann. 85 (1922), 115-122
[20] H. Bohr, E. Landau,

Uber das Verhalten von (s) und
(k)
(s) in der N ahe der
Geraden = 1, Nachr. Akad. Wiss. Gottingen II Math.-phys. Kl. (1910), 303-330
191
192 Bibliography
[21] H. Bohr, E. Landau, Ein Satz uber Dirichletsche Reihen mit Anwendung auf
die -Funktion und die L-Funktionen, Rend. di Palermo 37 (1914), 269-272
[22] H. Bohr, E. Landau, Sur les zeros de la fonction (s) de Riemann, Comptes
Rendus Acad. Sci. Paris 158 (1914), 106-110
[23] E. Bombieri, On the large sieve, Mathematika 12 (1965), 201-225
[24] E. Bombieri, A variational approach to the explicit formula, Comm. Pure Appl.
Math. 56 (2003), 1151-1164
[25] R. Brauer, On Artins L-series with general group characters. Ann. of Math.
48 (1947), 502-514
[26] C. Breuil, B. Conrad, F. Diamond, R. Taylor, On the modularity of
elliptic curves over Q: wild 3-adic exercises, J. Amer. Math. Soc. 14 (2001),
843-939
[27] V. Brun, La serie 1/5 +1/7 +1/11 +1/13 +1/17 +1/19 +1/29 +1/31 +1/41 +
1/43 +1/59 +1/61 . . . o` u les denominateurs sont nombres premiers jumeaux est
convergente o` u nie, Bull. Sci. Math. 43 (1919), 100-104, 124-128
[28] R.B. Burckel, Introduction to classical Complex Analysis, vol. I, Birkhauser
1979
[29] K. Chandrasekharan, Arithmetical Functions, Springer 1970
[30] K. Chandrasekharan, R. Narasimhan, The approximate functional equa-
tion for a class of zeta-functions, Math. Ann. 152 (1963), 30-64
[31] N. Chebotarev, Determination of the density of the set of prime numbers,
belonging to a given substitution class, Izv. Ross. Akad. Nauk 17 (1924), 205-250
(in Russian)
[32] P.L. Chebyshev, Sur la fonction qui determine la totalite des nombres premiers
inferieurs `a une limite donnee, Memoires des savants etrangers de lAcad. Sci. St.
Petersbourg 5 (1848), 1-19
[33] P.L. Chebyshev, Memoire sur nombres premiers, Memoires des savants
etrangers de lAcad. Sci. St. Petersbourg 7 (1850), 17-33
[34] A. Connes, Trace formula in noncommutative geometry and the zeros of the
Riemann zeta function, Selecta Math. (N.S.) 5 (1999), 29-106
[35] J.B. Conrey, More than two fths of the zeros of the Riemann zeta-function
are on the critical line, J. reine angew. Math. 399 (1989), 1-26
[36] J.B. Conrey, L-functions and random matrices, in Mathematics unlimited -
2001 and beyond, B. Engquist, W. Schmid (eds.), Springer 2001, 331-352
[37] J.B. Conrey, The Riemann hypothesis, Notices Amer. Math. Soc. 50 (2003),
341-353
[38] J.B. Conrey, A. Ghosh, On the Selberg class of Dirichlet series: small degrees,
Duke Math. J. 72 (1993), 673-693
[39] J.B. Conrey, A. Ghosh, High moments of the Riemann zeta-function, Duke
Math. J. 107 (2001), 577-604
[40] J.G. van der Corput, Zahlentheoretische Absch atzungen, Math. Ann. 84
(1921), 53-79
[41] H. Cramer, Ein Mittelwertsatz in der Primzahltheorie, Math. Z. 12 (1922),
147-153
[42] H. Cramer, On the order of magnitude of the dierence between consecutive
prime numbers, Acta Arith. 2 (1936), 23-46
[43] H. Davenport, H. Heilbronn, On the zeros of certain Dirichlet series I, II, J.
London Math. Soc. 11 (1936), 181-185; 307-312
Bibliography 193
[44] H. Davenport, Multiplicative number theory, Springer 1980, 2nd ed. revised by
H.L. Montgomery
[45] P. Deligne, La Conjecture de Weil I, II, Publ. I.H.E.S. 43 (1974), 273-307; 52
(1981), 313-428
[46] A. Denjoy, LHypoth`ese de Riemann sur la distribution des zeros de (s), reliee
`a la theorie des probabilites, Comptes Rendus Acad. Sci. Paris 192 (1931), 656-
658
[47] P.G.L. Dirichlet, Beweis des Satzes, dass jede unbegrenzte arithmetische Pro-
gression, deren erstes Glied und Dierenz ganze Zahlen ohne gemeinschaftlichen
Factor sind, unendlich viele Primzahlen enthalt, Abhandl. Kgl. Preu. Akad.
Wiss. (1837, 45-81 [in Werke I, G. Reimer, Berlin 1889, 313-342]
[48] V.G. Drinfeld, Langlands conjecture for GL(2) over function elds, Proc. ICM,
Helsinki 1978, 565-574, Acad. Sci. Fennica, Helsinki 1980
[49] H.M. Edwards, Riemanns zeta-function, Academic Press, New York- London
1974
[50] P. Erd os, On a new method in elementary number theory which leads to an
elementary proof of the prime number theorem, Proc. Nat. Acad. Sci. U.S.A. 35
(1949), 374-384
[51] L. Euler, Variae observationes circa series innitas, Comment. Acad. Sci.
Petropol 9 (1744), 160-188 [in Opera omnia I.14, Teubner 1924, 108-123]
[52] D.W. Farmer, Long molliers of the Riemann zeta-function, Matematika 40
(1993), 71-87
[53] R. Garunkstis, Note on the zeros of the Hurwitz zeta-function, in: Voronois
impact on modern science, Proceedings of the third Voronoi Conference on
Number Theory and Spatial Tessellations. Mathematics and its Applications 55
(2005), 10-12
[54] R. Garunkstis, A. Laurin cikas, The Lerch zeta-function, Kluwer Academic
Publishers, Dordrecht 2002
[55] R. Garunkstis, J. Steuding, On the zero distributions of Lerch zeta-functions,
Analysis 22 (2002), 1-12 (with R. Garunkstis)
[56] R. Garunkstis, J. Steuding, On the distribution of zeros of the Hurwitz zeta-
function, Math. Comput. (to appear)
[57] F. Gamann,

Uber Beziehungen zwischen den Primidealen eines algebraischen
Korpers und den Substitutionen seiner Gruppen, Math. Z. 25 (1926), 661-675
[58] S.S. Gelbart, An elementary introduction to the Langlands program, Bull.
Amer. Math. Soc. 10 (1984), 177-219
[59] P. Gerardin, W. Li, Functional equations and periodic sequences, in Theorie
des nombres, Quebec 1987, J.-M. De Koninck and C. Levesque (eds.), de
Gruyter, Berlin 1989, 267-279
[60] R. Godement, H. Jacquet, Zeta-functions of simple algebras, Lecture Notes
260, Springer 1972
[61] S.M. Gonek, Three lectures on the Riemann zeta-function, AIM preprint series
2003-20
[62] D. Goldfeld, Sur les produits euleriens attaches aux courbes elliptiques, C. R.
Acad. Sci. Paris 294 (1982), 471-474
[63] D. Goldfeld, Gauss class number problem for imaginary quadratic number
elds, Bull. A.M.S. 13 (1985), 23-37
194 Bibliography
[64] D. Goldfeld, The elementary proof of the prime number theorem: an historical
perspective, in Number theory, New York 2003, Springer 2004, 179-192
[65] D.A. Goldston, J. Pintz, C.Y. Yildirim, Primes in Tuples I, preprint
math.NT/0508185 available at the ArXiV
[66] A. Granville, Unexpected irregularities in the distribution of prime num-
bers, Proceedings of the International Congress of Mathematicians, Zrich 1994,
Birkhauser 1995, 388-399
[67] A. Granville, G. Martin, Prime number races, Gac. R. Soc. Mat. Esp. 8
(2005), 197-240 (in Spanish, with appendices by G. Davido and M. Guy); engl.
translation at http://arxiv.org/abs/math.NT/0408319
[68] B.J. Green, Long arithemtic progressions of primes, preprint math.NT/0508063
available at the ArXiV
[69] B.J. Green, T.C. Tao, The primes contain arbitrarily long arithemtic progres-
sion, to appear in Ann. Math.
[70] S.A. Gritsenko, On zeros of linear combinations of analogues of the Riemann
function, Tr. Mat. Inst. Steklova 218 (1997), 134-150 (in Russian); translation
in Proc. Steklov Inst. Math. 218 (1997), 129-145
[71] B. Gross, D.B. Zagier, Heegner points and derivatives of L-series, Invent.
Math. 84 (1986), 225-320
[72] J. Hadamard,

Etude sur le proprietes des fonctions enti`eres et en particulier
dune fonction consideree par Riemann, J. math. pures appl. 9 (1893), 171-215
[73] J. Hadamard, Sur les zeros de la fonction (s) de Riemann, Comptes Rendus
Acad. Sci. Paris 122 (1896), 1470-1473
[74] H. Hamburger,

Uber die Riemannsche Funktionalgleichung der -Funktion. I,
II., Math. Z. 10 (1921), 240-254; 11 (1921), 224-245
[75] G.H. Hardy, Sur les zeros de la fonction (s) de Riemann, Comptes Rendus
Acad. Sci. Paris 158 (1914), 1012-1014
[76] G.H. Hardy, On the Expression of a Number as the Sum of Two Squares, Quart.
J. Math. 46 (1915), 263-283
[77] G.H. Hardy, J.E. Littlewood, Contributions to the theory of the Riemann
zeta-function and the distribution of primes, Acta Math. 41 (1918), 119-196
[78] G.H. Hardy, J.E. Littlewood, The approximate functional equation in the
theory of the zeta-function, with applications to the divisor problems of Dirichlet
and Piltz, Proc. London Math. Soc. 21 (1922), 39-74
[79] G.H. Hardy, J.E. Littlewood, On Lindel ofs hypothesis concerning the Rie-
mann zeta-function, Proc. Royal Soc. 103 (1923), 403-412
[80] G.H. Hardy, J.E. Littlewood, Some Problems of Partitio Numerorum. III.
On the Expression of a Number as a Sum of Primes, Acta Math. 44 (1923), 1-70
[81] G.H. Hardy, E.M. Wright, An introduction to the theory of numbers, Claren-
don Press, Oxford, 1979, 5th ed.
[82] H. Hasse,

Uber die Darstellbarkeit von Zahlen durch quadratische Formen im
Korper der rationalen Zahlen, J. Reine Angew. Math. 152 (1923), 129-148
[83] H. Hasse, Beweis des Analogons der Riemannschen Vermutung fur die Artin-
schen und F.K. Schmidtschen Kongruenzzetafunktionen in gewissen elliptischen
Fallen, Nachr. Gesell. Wiss. Gottingen 42 (1933), 253-262
[84] D.R. Heath-Brown, The twelfth power moment of the Riemann zeta-function,
Quart. J. Math. 29 (1978), 443-462
Bibliography 195
[85] D.R. Heath-Brown, Simple zeros of the Riemann zeta-function on the critical
line, Bull. London Math. Soc. 11 (1979), 17-18
[86] E. Hecke,

Uber die Zetafunktion beliebiger algebraischer Zahlk orper, Nachr.
Ges. Wiss. Gottingen (1917), 77-89
[87] E. Hecke,

Uber eine neue Art von Zetafunktionen, Math. Z. 6 (1920), 11-51
[88] E. Hecke,

Uber die Bestimmung Dirichletscher L-Reihen durch ihre Funktion-
algleichung, Math. Ann. 112 (1936), 664-699
[89] K. Heegner, Diophantische Analysis und Modulfunktionen, Math. Z. 56 (1952),
227-253
[90] H. Heilbronn, On the class number in imaginary quadratic elds, Quarterly J.
Math. 5 (1934), 150-160
[91] H. Heilbronn, Zeta-functions and L-functions, in: Algebraic Number Theory,
J.W.S. Cassels and A. Frohlich (eds.), Academic Press 1967, 204-230
[92] M.N. Huxley, On the dierence between consecutive primes, Invent. Math. 15
(1972), 164-170
[93] M.N. Huxley, Integer points, exponential sums and the Riemann zeta function,
in Number theory for the millennium, II, Urbana 2000, A K Peters, Natick,
MA, 2002, 275290
[94] M.N. Huxley, Exponential sums and the Riemann zeta-function, V, Proc. Lon-
don Math. Soc. 90 (2005), 1-41
[95] S. Ikehara, An extension of Landaus theorem in the analytic theory of numbers,
J. MAth. Phys. M.I.T. 10 (1931), 1-12
[96] A.E. Ingham, Mean-value theorems in the theory of the Riemann zeta-function,
Proc. London Math. Soc. 27 (1926), 273-300
[97] A.E. Ingham, The distribution of prime numbers, Cambridge University Press
1932
[98] A. Ivi c, The theory of the Riemann zeta-function with applications, John Wiley
& Sons, New York 1985
[99] H. Iwaniec, Topics in classical automorphic forms, A.M.S., Providence, RI, 1997
[100] H. Iwaniec, P. Sarnak, Perspectives on the analytic theory of L-functions,
Geom. funct. anal., special volume - GAFA 2000, 705-741
[101] H. Iwaniec, E. Kowalski, Analytic number theory, AMS, Providence 2004
[102] H. Jacquet, R.P. Langlands, Automorphic forms on GL(2), Lecture Notes
114, Springer 1970
[103] H. Jacquet, J.A. Shalika, A non-vanishing theorem for zeta-functions on GL
n
,
Invent. math. 38 (1976), 1-16
[104] D. Joyner, Distribution theorems of L-functions, Pitman Research Notes in
Mathematics, 1986
[105] J. Kaczorowski, G. Molteni, A. Perelli, J. Steuding, J. Wolfart,
Heckes theory and the Selberg class, submitted
[106] J. Kaczorowski, A. Perelli, The Selberg class: a survey, in Number theory
in progress. Proceedings of the international conference in honor of the 60th
birthday of Andrej Schinzel, Zakopane 1997. Vol. 2: Elementary and analytic
number theory. De Gruyter 1999, 953-992
[107] J. Kaczorowski, A. Perelli, On the structure of the Selberg class, I: 0 d
1, Acta Math. 182 (1999), 207-241
[108] J. Kaczorowski, A. Perelli, On the prime number theorem in the Selberg
class, Arch. Math. 80 (2003), 255-263
196 Bibliography
[109] J. Kaczorowski, A. Perelli, On the structure of the Selberg class, V: 1 <
d < 5/3, Invent. math. 150 (2002), 485-516
[110] A.A. Karatsuba, On the zeros of the function (s) on short intervals of the
critical line, Izv. Akad. Nauk. SSSR Ser. Mat. 48 (1984) (in Russian); Math.
USSR-Izv. 24 (1985), 523-537
[111] A.A. Karatsuba, Complex Analysis in Number Theory, CRC Press 1995
[112] A.A. Karatsuba, S.M. Voronin, The Riemann zeta-function, de Gruyter 1992
[113] N.M. Katz, P. Sarnak, Random matrices, Frobenius eigenvalues, and mon-
odromy, AMS, Providence 1999
[114] J.P. Keating, N.C. Snaith, Random matrix theory and (
1
2
+ it), Comm.
Math. Phys. 214 (2000), 57-89
[115] A.W. Knapp, Elliptic curves, Princeton University Press 1992
[116] N. Koblitz, p-adic numbers, p-adic analysis, and zeta-functions, Springer 1984,
2nd ed.
[117] N. Koblitz, Introduction to elliptic curves and modular forms, Springer 1993,
2nd ed.
[118] H. von Koch, Sur la distribution des nombres premiers, Compt. Rendus Acad.
Sci. Paris 118 (1900), 1243-1246
[119] H. von Koch, Sur la distribution des nombres premiers, Acta. Math. 24 (1901),
159-182
[120] J. Korevaar, Tauberian theory, a century of developments, Springer 2004
[121] N.M. Korobov, Estimates of trigonometric sums and their applications, Uspehi
Mat. Nauk 13 (1958), 185-192
[122] T. Kubota, H.W. Leopoldt, Eine p-adische Theorie der Zetawerte I.
Einf uhrung der p-adischen Dirichletschen L-Funktionen, J. Reine Angew. Math.
214/215 (1964), 328-339
[123] L. Lafforgue, Chtoucas de Drinfeld et correspondance des Langlands, Invent.
Math. 147 (2002), 1-241
[124] E. Landau, Neuer Beweis des Primzahlsatzes und Beweis des Primidealsatzes,
Math. Ann. 56 (1903), 645-670
[125] E. Landau,

Uber die Gitterpunkte in einem Kreise, I+II, Nachr. Ges. Wiss.
Gottingen (1915), I: 148-160, II: 161-171
[126] R.P. Langlands, L-functions and automorphic representations, Proc. ICM,
Helsinki 1978, 165-175, Acad. Sci. Fennica, Helsinki 1980
[127] A. Laurin cikas, Limit theorems for the Riemann zeta-function, Kluwer Aca-
demic Publishers, Dordrecht 1996
[128] D.H. Lehmer, On the roots of Riemann zeta-function, Acta Math. 95 (1956),
291-298
[129] C.G. Lekkerkerker, On the zeros of a class of Dirichlet series, Proefschrift,
van Gorcum & Comp. N.V. 1955
[130] H. Lenstra Jr., Factoring integers with elliptic curves, Ann. Math. 126 (1987),
649-673
[131] H.W. Lenstra, P. Stevenhagen, Chebotarev and his density theorem, Math.
Intelligencer 18 (1996), 26-37
[132] N. Levinson, More than one third of Riemanns zeta-function are on =
1
2
,
Adv. Math. 13 (1974), 383-436
[133] N. Levinson, Almost all roots of (s) = a are arbitrarily close to = 1/2, Proc.
Nat. Acad. Sci. U.S.A. 72 (1975), 1322-1324
Bibliography 197
[134] E. Lindel of, Quelques remarques sur la croissance de la fonction (s), Bull. sci.
math. 32 (1908), 341-356
[135] E. Lindel of, E. Phragmen, Sur une extension dun principle classique
danalyse et sur quelques proprietes du fonctions monog`enes dans le voisinage
dun point singulier, Acta Math. 31 (1908), 381-406
[136] J.E. Littlewood, Sur la distribution de nombres premiers, Comptes Rendus bf
158 (1914), 1869-1872
[137] J.E. Littlewood, On the zeros of the Riemann zeta-function, Proc. Cambridge
Phil. Soc. 22 (1924), 295-318
[138] J. Liu, Y. Wang, Y. Ye, A proof of Selbergs orthogonality for automorphic
L-functions, Manuscripta math. (to appear)
[139] J. van de Lune, H.J.J. te Riele, D.T. Winter, On the zeros of the Riemann
zeta-function in the critical strip, IV, Math. Comp. 46 (1986), 667-681
[140] H. Maier, Primes in short intevals, Michigan Math. J. 32 (1985), 221-225
[141] H. von Mangoldt, Zu Riemanns Abhandlung

Uber die Anzahl der


Primzahlen unter einer gegebenen Grosse, J. reine angew. Math. 114 (1895),
255-305
[142] H. von Mangoldt, Zur Verteilung der Nullstellen der Riemannschen Funktion
(t), Math. Ann. 60 (1905), 1-19
[143] K. Matsumoto, Value-distribution of zeta-functions, in Analytic number the-
ory, Proceedings of the Japanese-French Symposium, Tokyo 1988, K. Nagasaka
and

E. Fouvry (eds.), Lecture Notes in Math. 1434, Springer 1990, 178187
[144] K. Matsumoto, Recent developments in the mean square theory, in: Number
Theory, Trends Math., Birkhauser 2000, 241-286
[145] F. Mertens, Ein Beitrag zur analytischen Zahlentheorie, J. reine angew. Math.,
78 (1874), 46-62
[146] T. Miyake, Modular Forms, Springer Verlag 1976
[147] G. Molteni, A note on a result of Bochner and Conrey-Ghosh about the Selberg
class, Arch. Math. 72 (1999), 219-222
[148] G. Molteni, J. Steuding, (Almost) primitivity of Hecke L-functions, submit-
ted
[149] H.L. Montgomery, The pair correlation of zeros of the zeta-function, Proc.
Symp. Pure Math. 24 (1973), 181-193
[150] L.J. Mordell, On Ramanujans empirical expansions of modular functions,
Proc. Cambridge Philos. Soc. 19 (1920), 117-124
[151] Mordell, On the rational solutions of the indeterminate equations of the third
and fourth degrees, Proc. Cambridge Phil. Soc. 21 (1922/23), 179-192
[152] Y. Motohashi, Spectral theory of the Riemann zeta-function, Cambridge Uni-
versity Press 1997
[153] M.R. Murty, Primes in certain arithmetic progressions, J. Madras Univ. (1988),
161-169
[154] M.R. Murty, A motivated introduction to the Langlands program, in: Ad-
vances in Number theory, F. Gouvea and N. Yui, eds., Clarendon Press Oxford
1993, 37-66
[155] M.R. Murty, Selbergs conjectures and Artin L-functions, Bull. Amer. Math.
Soc. 31 (1994), 1-14
[156] M.R. Murty, Selbergs conjectures and Artin L-functions, II, in: Current trends
in mathematics and physics, ed. S.D. Adhikari, Narosa, New Delhi 1995, 154-168
198 Bibliography
[157] M.R. Murty, V.K. Murty, Non-vanishing of L-functions and applications,
Birkhauser 1997
[158] M.R. Murty, A. Perelli, The pair correlation of zeros of functions in the
Selberg class, Internat. Math. Res. Notices 10 (1999), 531-545
[159] W. Narkiewicz, The development of prime number theory, Springer 2000
[160] J. Neukirch, Algebraische Zahlentheorie, Springer 1992
[161] A.M. Odlyzko, The 10
20
th zero of the Riemann zeta-function and 70 million of
its neighbors, in Dynamical, spectral, and arithmetic zeta functions (San An-
tonio, TX, 1999), 139144, Contemp. Math. 290, Amer. Math. Soc., Providence
2001
[162] A.M. Odlyzko, H.J.J. te Riele, Disproof of Mertens conjecture, J. reine
angew. Math. 367 (1985), 138-160
[163] A.P. Ogg, Modular Forms and Dirichlet Series, Benjamin, New York-
Amsterdam 1969
[164] S.J. Patterson, An introduction to the theory of the Riemann zeta-function,
Cambridge University Press 1988
[165] D.V. Pechersky, On the permutation of the terms of functional series, Dokl.
Akad. Nauk SSSR 209 (1973), 1285-1287
[166] A. Perelli, General L-functions, Ann. Mat. Pura Appl. 130 (1982), 287-306
[167] R. Perlis, On the equation
K
(s) =
K
(s), J. Number Theory 9 (1977), 342-360
[168] R. Perlis, A. Schinzel, Zeta functions and the equivalence of integral forms,
J. Reine Angew. Math. 309 (1979), 176-182
[169] H. Petersson, Konstruktion der samtlichen L osungen einer Riemannschen
Funktionalgleichung durch Dirichletreihen mit Eulerscher Produktentwicklung
II, Math. Ann. 117 (1940/41), 39-64
[170] I. Piatestki-Shapiro, R. Rhagunathan, On Hamburgers theorem, Amer.
Math. Soc. Transl. 169 (1995), 109-120
[171] H.S.A. Potter, The mean values of certain Dirichlet series I, Proc. London
Math. Soc. 46 (1940), 467-468
[172] K. Prachar, Primzahlverteilung, Springer 1957
[173] S. Ramanujan, On certain arithmetical functions, Trans. Camb. Phil. Soc. 22
(1916), 159-184
[174] A. Reich, Universelle Wertverteilung von Eulerprodukten, Nach. Akad. Wiss.
Gottingen, Math.-Phys. Kl. (1977), 1-17
[175] B. Riemann,

Uber die Anzahl der Primzahlen unterhalb einer gegebenen Grosse,
Monatsber. Preuss. Akad. Wiss. Berlin (1859), 671-680
[176] M. Rubinstein, P. Sarnak, Chebyshevs bias, Experiment. Math. 3 (1994),
173-197
[177] Z. Rudnick, P. Sarnak, Zeros of principal L-functions and Random Matrix
Theory, Duke Math. J. 81 (1996), 269-322
[178] A. Selberg, On the normal density of primes in small intervals and the dierence
between consecutive primes, Arch. Math. Naturvid. 47 (1943), 87-105
[179] A. Selberg, An Elementary Proof of the Prime Number Theorem, Ann. of
Math. 50 (1949), 305-313
[180] A. Selberg, Old and new conjectures and results about a class of Dirichlet
series, in: Proceedings of the Amal Conference on Analytic Number Theory,
Maiori 1989, E. Bombieri et al. (eds.), Universit`a di Salerno 1992, 367-385
Bibliography 199
[181] C.L. Siegel, Bemerkungen zu einem Satz von Hamburger uber die Funktional-
gleichung der Riemannschen Zetafunktion, Math. Ann. 86 (1922), 276-279
[182] C.L. Siegel,

Uber Riemanns Nachlass zur analytischen Zahlentheorie, Quellen
u. Studien zur Geschichte der Math. Astr. Phys. 2 (1932), 45-80
[183] C.L. Siegel,

Uber die Classenzahl quadratischer Zahlk orper, Acta Arith. 1
(1935), 83-86
[184] S. Singh, Fermats last theorem, Fourth Estate, London 1997
[185] K. Soundararajan, Omega results for the divisor and circle problems, Int.
Math. Res. Not. 36 (2003), 1987-1998
[186] A. Speiser, Geometrisches zur Riemannschen Zetafunktion, Math. Ann. 110
(1934), 514-521
[187] H.M. Stark, A Complete Determination of the Complex Quadratic Fields of
Class Number One, Michigan Math. J. 14 (1967), 1-27
[188] H.M. Stark, Galois theory, algebraic numbers and zeta functions, in: From
number teory to physics, Waldschmitd et al. (eds.), Springer 1989, 313-393
[189] J. Steuding, On simple zeros of the Riemann zeta-function in short intervals
on the critical line, Acta Math. Hungar. 96 (2002), 259-308
[190] J. Steuding, On the value distribution of Hurwitz zeta-function at the nontrivial
zeros of the Riemann zeta-function, Abh. Math. Sem. Uni. Hamburg 71 (2001),
113-122
[191] J. Steuding, Universality in the Selberg class, in: Special activity in Analytic
Number Theory and Diophantine equations, Proceedings of a workshop at the
Max Planck-Institut Bonn 2002, R.B. Heath-Brown and B. Moroz (eds.), Bonner
math. Schriften 360 (2003)
[192] J. Steuding, On the value-distribution of L-functions, Fiz. Mat. Fak. Moksl.
Semin. Darb. 6 (2003), 87-119
[193] J. Steuding, On the zero-distribution of Epstein zeta-functions, Math. Annalen
333 (2005), 689-697
[194] J. Steuding, On the value-distribution of Epstein zeta-functions, submitted
[195] H.P.F. Swinnerton-Dyer, A brief guide to algebraic number theory, London
Math. Soc., Cambridge University Press 2001
[196] A. Tauber, Ein Satz aus der Theorie der unendlichen Reihen, Monatsh. Math.
8 (1897), 273-177
[197] Tate, Fourier analysis in number elds, and Heckes zeta-functions, in: Al-
gebraic Number Theory, Proc. Instructional Conf. Brighton 1965, Thompson,
Washingthon 1967, 305-347
[198] R. Taylor, A. Wiles, Ring-theoretic properties of certain Hecke algebras, Ann.
of Math. 141 (1995), 553-572
[199] E.C. Titchmarsh, The theory of functions, Oxford University Press, 2nd ed.
1939
[200] E.C. Titchmarsh, The theory of the Riemann zeta-function, Oxford University
Press 1986, 2nd ed., revised by D.R. Heath-Brown
[201] J. Tunnell, A classical diophantine problem and modular forms of weight 3/2,
Inventiones math. 72 (1983), 323-334
[202] C.J. de la Vallee-Poussin, Recherches analytiques sur la theorie des nombres
premiers, I-III, Ann. Soc. Sci. Bruxelles 20 (1896), 183-256, 281-362, 363-397
200 Bibliography
[203] M.-F. Vigneras, Facteurs gamma et equations fonctionnelles, in Modular func-
tions of one variable VI, Springer Lecture Notes in Mathematics 627 (1977),
79-104
[204] I.M. Vinogradov, A new estimate for the function (1 + it), Izv. Akad. Nauk
SSSR, Ser. Mat. 22 (1958), 161-164
[205] A.I. Vinogradov, On the density hypothesis for Dirichlet L-functions, Izv.
Akad. Nauk SSSR 29 (1965), 903-934
[206] S.M. Voronin, Theorem on the universality of the Riemann zeta-function, Izv.
Akad. Nauk SSSR, Ser. Matem., 39 (1975), 475-486 (in Russian); Math. USSR
Izv. 9 (1975), 443-445
[207] L.C. Washington, Elliptic curves, CRC Press, Boca Raton, FL, 2003
[208] K. Weierstrass, Zur Theorie der eindeutigen analytischen Funktionen, Abh.
Kgl. Preuss. Akad. Wiss. Berlin (1876), 11-60
[209] A. Weil,

Uber die Bestimmung Dirichletscher Reihen durch Funktionalgleichun-
gen, Math. Ann. 168 (1967), 149-156
[210] A. Weil, Sur les formules explicites de la theorie des nombres, Izv. Akad. Nauk
36 (1972), 3-18
[211] A. Weil, Prehistory of the zeta-function, in: Number theory, trace formulas
and discrete groups, Symposium in honor of Atle Selberg, Oslo 1987, Academic
Press 1989, 1-10
[212] N. Wiener, Tauberian theorems, Ann. of Math. 33 (1932), 1-100
[213] A. Wiles, Modular elliptic curves and Fermats last theorem, Ann. Math. 141
(1995), 443-551
[214] A. Wiles, The Birch and Swinnerton-Dyer conjecture, available at
http://www.claymath.org/millennium/
[215] J. Wolfart, Transzendente Zahlen als Fourierkoezienten von Heckes Modul-
formen, Acta Arith. 39 (1981), 193-205
[216] S. Wolfram, The Mathematica Book, Cambridge University Press, 1999, 4th
ed.

Você também pode gostar