Sunteți pe pagina 1din 61

ELEMENTARY AND

ANALYTIC NUMBER THEORY


W W L CHEN


c W W L Chen, 1981.
This work is available free, in the hope that it will be useful.
Any part of this work may be reproduced or transmitted in any form or by any means, electronic or mechanical, including
photocopying, recording, or any information storage and retrieval system, with or without permission from the author.

Chapter 1
ARITHMETIC FUNCTIONS

1.1. Introduction

By an arithmetic function, we mean a function of the form f : N C. We say that an arithmetic


function f : N C is multiplicative if f (mn) = f (m)f (n) whenever m, n N and (m, n) = 1.

Example. The function U : N C, dened by U (n) = 1 for every n N, is an arithmetic function.


Furthermore, it is multiplicative.

THEOREM 1.1. Suppose that the function f : N C is multiplicative. Then the function g : N C,
dened by 
g(n) = f (m)
m|n

for every n N, is multiplicative.



Here the summation m|n denotes a sum over all positive divisors m of n.

Proof of Theorem 1.1. Suppose that a, b N and (a, b) = 1. If u is a positive divisor of a and v
is a positive divisor of b, then clearly uv is a positive divisor of ab. On the other hand, it is well-known
that every positive divisor m of ab can be expressed uniquely in the form m = uv, where u is a positive
divisor of a and v is a positive divisor of b. It follows that
  
g(ab) = f (m) = f (uv) = f (u)f (v)
m|ab u|a v|b u|a v|b

 
= f (u) f (v) = g(a)g(b).
u|a v|b

This chapter was rst used in lectures given by the author at Imperial College, University of London, in 1981.
12 W W L Chen : Elementary and Analytic Number Theory

1.2. The Divisor Function

We dene the divisor function d : N C by writing



d(n) = 1 (1)
m|n

for every n N. Here the sum is taken over all positive divisors m of n. In other words, the value
d(n) denotes the number of positive divisors of the natural number n. On the other hand, we dene the
function : N C by writing 
(n) = m (2)
m|n

for every n N. Clearly, the value (n) denotes the sum of all the positive divisors of the natural
number n.

THEOREM 1.2. Suppose that n N and that n = pu1 1 . . . pur r is the canonical decomposition of n.
Then
pu1 +1 1 pur +1 1
d(n) = (1 + u1 ) . . . (1 + ur ) and (n) = 1 ... r .
p1 1 pr 1

Proof. Every positive divisor m of n is of the form m = pv11 . . . pvrr , where for every j = 1, . . . , r, the
integer vj satises 0 vj uj . It follows from (1) that d(n) is the number of choices for the r-tuple
(v1 , . . . , vr ). Hence

u1 
ur
d(n) = ... 1 = (1 + u1 ) . . . (1 + ur ).
v1 =0 vr =0

On the other hand, it follows from (2) that


 

u1 
ur 
u1 
ur
(n) = ... pv11 . . . pvrr = pv11 ... pvrr .
v1 =0 vr =0 v1 =0 vr =0

Note now that for every j = 1, . . . , r, we have


uj
v u
u +1
pj j 1
pj j = 1 + pj + p2j + . . . + pj j = .
vj =0
pj 1

The second result follows.

The result below is a simple deduction from Theorem 1.2.

THEOREM 1.3. The arithmetic functions d : N C and : N C are both multiplicative.

Natural numbers n N where (n) = 2n are of particular interest, and are known as perfect
numbers. A perfect number is therefore a natural number which is equal to the sum of its own proper
divisors; in other words, the sum of all its positive divisors other than itself.

Example. It is easy to see that 6 = 1 + 2 + 3 and 28 = 1 + 2 + 4 + 7 + 14 are perfect numbers.

It is not known whether any odd perfect number exists. However, we can classify the even perfect
numbers.

THEOREM 1.4. (EuclidEuler) Suppose that m N. If 2m 1 is a prime, then the number


2m1 (2m 1) is an even perfect number. Furthermore, there are no other even perfect numbers.
Chapter 1 : Arithmetic Functions 13

Proof. Suppose that n = 2m1 (2m 1), and 2m 1 is prime. Clearly

(2m1 , 2m 1) = 1.

It follows from Theorems 1.2 and 1.3 that


2m 1 m
(n) = (2m1 )(2m 1) = 2 = 2n,
21

so that n is a perfect number, clearly even since m 2.


Suppose now that n N is an even perfect number. Then we can write n = 2m1 u, where m N
and m > 1, and where u N is odd. By Theorem 1.2, we have

2m u = (n) = (2m1 )(u) = (2m 1)(u),

so that
2m u u
(u) = =u+ m . (3)
2m 1 2 1
Note that (u) and u are integers and (u) > u. Hence u/(2m 1) N and is a divisor of u. Since
m > 1, we have 2m 1 > 1, and so u/(2m 1) = u. It now follows from (3) that (u) is equal to the
sum of two of its positive divisors. But (u) is equal to the sum of all its positive divisors. Hence u
must have exactly two positive divisors, so that u is prime. Furthermore, we must have u/(2m 1) = 1,
so that u = 2m 1.

We are interested in the behaviour of d(n) and (n) as n . If n N is a prime, then clearly
d(n) = 2. Also, the magnitude of d(n) is sometimes greater than that of any power of log n. More
precisely, we have the following result.
c
THEOREM 1.5. For any xed real number c > 0, the inequality d(n) (log n) as n does not
hold.

Proof. The idea of the proof is to consider integers which are divisible by many dierent primes.
Suppose that c > 0 is given and xed. Let  N {0} satisfy  c <  + 1. For every j = 1, 2, 3, . . . ,
let pj denote the j-th positive prime in increasing order of magnitude, and consider the integer
m
n = (p1 . . . p
+1 ) .

Then in view of Theorem 1.2, we have




+1

+1 log n
+1 c
d(n) = (m + 1) > > K(c)(log n) > K(c)(log n) , (4)
log(p1 . . . p
+1 )

where the positive constant




+1
1
K(c) =
log(p1 . . . p
+1 )
depends only on c. The result follows on noting that the inequality (4) holds for every m N.

On the other hand, the order of magnitude of d(n) cannot be too large either.

THEOREM 1.6. For any xed real number  > 0, we have d(n) n as n .

Proof. For every natural number n > 1, let n = pu1 1 . . . pur r be its canonical decomposition. It follows
from Theorem 1.2 that
d(n) (1 + u1 ) (1 + ur )
= u1 ... .
n p1 p u
r
r
14 W W L Chen : Elementary and Analytic Number Theory

We may assume without loss of generality that  < 1. If 2 pj < 21/ , then
uj
pj 2 uj = e uj log 2 > 1 + uj log 2 > (1 + uj ) log 2,
so that
(1 + uj ) 1
u < .
pj j  log 2
On the other hand, if pj 21/ , then p j 2, and so
(1 + uj ) 1 + uj
uj 1.
pj 2u j
It follows that
d(n) 1
< ,
n 1/
 log 2
p<2

a positive constant depending only on .

We see from Theorems 1.5 and 1.6 and the fact that d(n) = 2 innitely often that the magnitude of
d(n) uctuates a great deal as n . It may then be more fruitful to average the function d(n) over a
range of values n, and consider, for positive real numbers X R, the value of the average
1 
d(n).
X
nX

THEOREM 1.7. (Dirichlet) As X , we have



d(n) = X log X + (2 1)X + O(X 1/2 ).
nX

Here is Eulers constant and is dened by



 1
= lim log Y = 0.5772156649 . . . .
Y n
nY

Remark. It is an open problem in mathematics to determine whether Eulers constant is rational


or irrational.

The proof of Theorem 1.7 depends on the following intermediate result.

THEOREM 1.8. As Y , we have


 1

1
= log Y + + O .
n Y
nY

Proof. As Y , we have

 1  1 1 Y
[Y ]  Y 1
= + 2
du = + du
n Yn u Y u2
nY nY nY n

Y  Y
[Y ] 1 du = [Y ] + [u]
= + 2
1 du
Y 1 u Y 1 u2
nu
Y Y
[Y ] 1 u [u]
= + du du
Y u u2
1

1

1 u [u] u [u]
= log Y + 1 + O 2
du + du
Y 1 u Y u2



u [u] 1
= log Y + 1 2
du + O .
1 u Y
Chapter 1 : Arithmetic Functions 15

It is a simple exercise to show that


u [u]
1 du = .
1 u2

Proof of Theorem 1.7. As X , we have


       
d(n) = 1= 1+ 1 1
nX x,y xX 1/2 y X yX 1/2 x X xX 1/2 yX 1/2
xyX x y

 X  2  X 2
=2 [X 1/2 ] = 2 + O(X 1/2 ) (X 1/2 + O(1))
x x
xX 1/2 xX 1/2



1
= 2X log X 1/2 + + O + O(X 1/2 ) X
X 1/2
= X log X + (2 1)X + O(X 1/2 ).

We next turn our attention to the study of the behaviour of (n) as n . Every number n N
has divisors 1 and n, so we must have (1) = 1 and (n) > n if n > 1. On the other hand, it follows
from Theorem 1.6 that for any xed real number  > 0, we have
(n) nd(n) n1+
as n .
In fact, it is rather easy to prove a slightly stronger result.

THEOREM 1.9. We have (n) n log n as n .

Proof. As n , we have
 n  1
(n) = n n log n.
m m
m|n mn

As in the case of d(n), the magnitude of (n) uctuates a great deal as n . As before, we shall
average the function (n) over a range of values n, and consider some average version of the function.
Corresponding to Theorem 1.7, we have the following result.

THEOREM 1.10. As X , we have


 2 2
(n) = X + O(X log X).
12
nX

Proof. As X , we have
  n   n  
(n) = = = r
m m
nX nX m|n mX nX mX rX/m
m|n
 1 X 
 
X 1  X

2
= 1+ = + O(1)
2 m m 2 m
mX mX

2   
X 1 1
= + O X +O 1
2 m2 m
mX mX mX


X2  1  1
= + O X2 + O(X log X)
2 m=1 m2 m2
m>X
2 2
= X + O(X log X).
12
16 W W L Chen : Elementary and Analytic Number Theory

1.3. The M
obius Function

obius function : N C by writing


We dene the M

1 if n = 1,
(n) = (1)r if n = p1 . . . pr , a product of distinct primes,
0 otherwise.

Remarks. (i) A natural number which is not divisible by the square of any prime is called a squarefree
number. Note that 1 is both a square and a squarefree number. Furthermore, a number n N is
squarefree if and only if (n) = 1.
(ii) The motivation for the denition of the M obius function lies rather deep. To understand the
denition, one needs to study the Riemann zeta function, an important function in the study of the
distribution of primes. For a more detailed discussion, see Chapters 4, 5 and 6. At this point, it suces
to remark that the M obius function is dened so that if we formally multiply the two series

 
1 (n)
and ,
n=1
ns n=1
ns

then the product is identically equal to 1. Heuristically, note that


 

    
1 (m) (m)   1
= = (m) s .
ks m=1
ms n=1 m=1
n s
n=1
n
k=1 k=1 m|n
km=n

It follows that the product is identically equal to 1 if


 
1 if n = 1,
(m) =
0 if n > 1.
m|n

We shall establish this last fact and study some of its consequences over the next four theorems.

obius function : N C is multiplicative.


THEOREM 1.11. The M

Proof. Suppose that a, b N and (a, b) = 1. If a or b is not squarefree, then neither is ab, and so
(ab) = 0 = (a)(b). On the other hand, if both a and b are squarefree, then since (a, b) = 1, ab must
also be squarefree. Furthermore, the number of prime factors of ab must be the sum of the numbers of
prime factors of a and of b.

THEOREM 1.12. Suppose that n N. Then


 
1 if n = 1,
(m) =
0 if n > 1.
m|n

Proof. Consider the function f : N C dened by writing



f (n) = (m)
m|n

for every n N. It follows from Theorems 1.1 and 1.11 that f is multiplicative. For n = 1, the result is
trivial. To complete the proof, it therefore suces to show that f (pk ) = 0 for every prime p and every
k N. Indeed,

f (pk ) = (m) = (1) + (p) + (p2 ) + . . . + (pk ) = 1 1 + 0 + . . . + 0 = 0.
m|pk
Chapter 1 : Arithmetic Functions 17

Theorem 1.12 plays the central role in the proof of the following two results which are similar in
nature.

THEOREM 1.13. (M obius Inversion Formula) Given any function f : N C, suppose that the
function g : N C is dened by writing

g(n) = f (m)
m|n

for every n N. Then for every n N, we have


 n  n
f (n) = (m)g = g(m).
m m
m|n m|n

Proof. The second equality is obvious. Also



 n   
(m)g = (m) f (k) = (m)f (k)
m n
m|n m|n k| m k,m
km|n

 
= f (k) (m) = f (n),
k|n m| n
k

in view of Theorem 1.12.

THEOREM 1.14. For any function g : N C, if the function f : N C is dened by writing


 n
f (n) = g(m)
m
m|n

for every n N, then for every n N, we have


  n
g(n) = f (m) = f .
m
m|n m|n

Proof. The second equality is obvious. Also



 n    n   

n/k
f = g(k) = g(k)
m n mk n m
m|n m|n k| m k|n m|
k

 
= g(k) (m) = g(n),
k|n m| n
k

in view of Theorem 1.12.

Remark. In number theory, it occurs quite often that in the proof of a theorem, a change of order
of summation of the variables is required, as illustrated in the proofs of Theorems 1.13 and 1.14. This
process of changing the order of summation does not depend on the summand in question. In both
instances, we are concerned with a sum of the form

A(k, m).
n
m|n k| m
18 W W L Chen : Elementary and Analytic Number Theory

This means that for every positive divisor m of n, we rst sum the function A over all positive divisors
k of n/m to obtain the sum 
A(k, m),
n
k| m

which is a function of m. We then sum this sum over all divisors m of n. Now observe that for every
natural number k satisfying k | n/m for some positive divisor m of n, we must have k | n. Consider
therefore a particular natural number k satisfying k | n. We must nd all natural numbers m satisfying
the original summation conditions, namely m | n and k | n/m. These are precisely those natural numbers
m satisfying m | n/k. We therefore obtain, for every positive divisor k of n, the sum

A(k, m).
m| n
k

Summing over all positive divisors k of n, we obtain



A(k, m).
k|n m| n
k

Since we are summing the function A over the same collection of pairs (k, m), and have merely changed
the order of summation, we must have
 
A(k, m) = A(k, m).
n
m|n k| m k|n m| n
k

1.4. The Euler Function

We dene the Euler function : N C as follows. For every n N, we let (n) denote the number of
elements in the set {1, 2, . . . , n} which are coprime to n.

THEOREM 1.15. For every number n N, we have



(m) = n.
m|n

Proof. We shall partition the set {1, 2, . . . , n} into d(n) disjoint subsets Bm , where for every positive
divisor m of n,
Bm = {x : 1 x n and (x, n) = m}.
If x Bm , let x = mx . Then (mx , n) = m if and only if (x , n/m) = 1. Also 1 x n if and only if
1 x n/m. Hence

Bm = {x : 1 x n/m and (x , n/m) = 1}

has the same number of elements as Bm . Note now that the number of elements of Bm is exactly (n/m).
Since every element of the set {1, 2, . . . , n} falls into exactly one of the subsets Bm , we must have
 n 
n= = (m).
m
m|n m|n

Apply the M obius inversion formula to the conclusion of Theorem 1.15, we obtain immediately the
following result.
Chapter 1 : Arithmetic Functions 19

THEOREM 1.16. For every number n N, we have


 n  (m)
(n) = (m) =n .
m m
m|n m|n

THEOREM 1.17. The Euler function : N C is multiplicative.

Proof. Since the M obius function is multiplicative, it follows that the function f : N C, dened
by f (n) = (n)/n for every n N, is multiplicative. The result now follows from Theorem 1.1.

THEOREM 1.18. Suppose that n N and n > 1, and that n = pu1 1 . . . pur r is the canonical decom-
position of n. Then
r

r
1 u 1
(n) = n 1 = pj j (pj 1).
j=1
p j j=1

Proof. The second equality is trivial. On the other hand, for every prime p and every u N, we have
by Theorem 1.16 that
(pu )  (m) (p) 1
u
= =1+ =1 .
p u
m p p
m|p

The result now follows since is multiplicative.

We noe study the magnitude of (n) as n . Clearly (1) = 1 and (n) < n if n > 1.
Suppose rst of all that n has many dierent prime factors. Then n must have many dierent
divisors, and so (n) must be large relative to n. But then many of the numbers 1, . . . , n cannot be
coprime to n, and so (n) must be small relative to n. On the other hand, suppose that n has very few
prime factors. Then n must have very few divisors, and so (n) must be small relative to n. But then
many of the numbers 1, . . . , n are coprime to n, and so (n) must be large relative to n. It therefore
appears that if one of the two values (n) and (n) is large relative to n, then the other must be small
relative to n. Indeed, our heuristics are upheld by the following result.

THEOREM 1.19. For every n N, we have


1 (n)(n)
< 1.
2 n2

Proof. The result is obvious if n = 1, so suppose that n > 1. Let n = pu1 1 . . . pur r be the canonical
decomposition of n. Recall Theorems 1.2 and 1.18. We have
u 1
r u +1
pj j 1 r
1 pj j
(n) = =n
j=1
pj 1 j=1
1 p1
j

and

r
(n) = n (1 p1
j ).
j=1

Hence
(n)(n) r
u 1
= (1 pj j ).
n2 j=1

The upper bound follows at once. On the other hand, the lower bound follows on observing that
r n

uj 1 2 1 n+1 1
(1 pj ) (1 p ) 1 2 = > .
j=1 m=2
m 2n 2
p|n
110 W W L Chen : Elementary and Analytic Number Theory

Combining Theorems 1.9 and 1.19, we have the following result.

THEOREM 1.20. We have (n)  n/ log n as n .

We now consider some average version of the Euler function.

THEOREM 1.21. (Mertens) As X , we have


 3 2
(n) = X + O(X log X).
2
nX

Proof. As X , we have, by Theorem 1.16, that


  n   n  
(n) = (m) = (m) = (m) r
m m
nX nX m|n mX nX mX rX/m
m|n
  
 
2
1 X X 1  X
= (m) 1+ = (m) + O(1)
2 m m 2 m
mX mX

X 2  (m)  1 
= 2
+ O X +O 1
2 m m
mX mX mX


X 2  (m)  1
= 2
+ O X2 + O(X log X)
2 m=1 m m2
m>X

X  (m)
2
= + O(X log X)
2 m=1 m2
.

It remains to show that



(m) 6
2
= 2.
m=1
m
But



 1  (m)  1   1 
2
= (m) = (m) = 1,
n=1
n m=1
m2 k2 n,m k 2
k=1 k=1 m|k
nm=k

in view of Theorem 1.12.

1.5. Dirichlet Convolution

We shall denote the class of all arithmetic functions by A, and the class of all multiplicative functions
by M.
Given arithmetic functions f, g A, we dene the function f g : N C by writing
 n
(f g)(n) = f (m)g
m
m|n

for every n N. This function is called the Dirichlet convolution of f and g.


It is not dicult to show that Dirichlet convolution of arithmetic functions is commutative and
associative. In other words, for every f, g, h A, we have

f g =gf and (f g) h = f (g h).


Chapter 1 : Arithmetic Functions 111

Furthermore, the arithmetic function I : N C, dened by I(1) = 1 and I(n) = 0 for every n N
satisfying n > 1, is an identity element for Dirichlet convolution. It is easy to check that I f = f I = f
for every f A.
On the other hand, an inverse may not exist under Dirichlet convolution. Consider, for example,
the function f A satisfying f (n) = 0 for every n N.

THEOREM 1.22. For any f A, the following two statements are equivalent:
(i) We have f (1) = 0.
(ii) There exists a unique g A such that f g = g f = I.

Proof. Suppose that (ii) holds. Then f (1)g(1) = 1, so that f (1) = 0. Conversely, suppose that
f (1) = 0. We shall dene g A iteratively by writing

1
g(1) = (5)
f (1)

and n
1 
g(n) = f (d)g (6)
f (1) d
d|n
d>1

for every n N satisfying n > 1. It is easy to check that this gives an inverse. Moreover, every inverse
must satisfy (5) and (6), and so must be unique.

We now describe Theorem 1.12 and M obius inversion in terms of Dirichlet convolution. Recall that
the function U A is dened by U (n) = 1 for all n N.

THEOREM 1.23.
(i) We have U = I.
(ii) If f A and g = f U , then f = g .
(iii) If g A and f = g , then g = f U .

Proof. (i) follows from Theorem 1.12. To prove (ii), note that

g = (f U ) = f (U ) = f I = f.

To prove (iii), note that


f U = (g ) U = g ( U ) = g I = g.

We conclude this chapter by exhibiting some group structure within A and M.

THEOREM 1.24. The sets A = {f A : f (1) = 0} and M = {f M : f (1) = 1} form abelian


groups under Dirichlet convolution.

Remark. Note that if f M is not identically zero, then f (n) = 0 for some n N. Since f (n) =
f (1)f (n), we must have f (1) = 1.

Proof of Theorem 1.24. For A , this is now trivial. We now consider M . Clearly I M . If
f, g M and (m, n) = 1, then

  mn  

mn
(f g)(mn) = f (d)g = f (d1 d2 )g
d d1 d2
d|mn d1 |m d2 |n




m n
= f (d1 )g f (d2 )g = (f g)(m)(f g)(n),
d1 d2
d1 |m d2 |n
112 W W L Chen : Elementary and Analytic Number Theory

so that f g M. Since (f g)(1) = f (1)g(1) = 0, we have f g M . It remains to show that if


f M , then f has an inverse in M . Clearly f has an inverse in A under Dirichlet convolution. Let
this inverse be h. We now dene g A by writing g(1) = 1,

g(pk ) = h(pk )

for every prime p and k N, and


g(n) = g(pk )
pk n

for every n > 1. Then g M . Furthermore, for every integer n > 1, we have

(f g)(n) = (f g)(pk ) = (f h)(pk ) = I(pk ) = I(n),
pk n pk n pk n

so that g is an inverse of f .
ELEMENTARY AND
ANALYTIC NUMBER THEORY
W W L CHEN


c W W L Chen, 1981.
This work is available free, in the hope that it will be useful.
Any part of this work may be reproduced or transmitted in any form or by any means, electronic or mechanical, including
photocopying, recording, or any information storage and retrieval system, with or without permission from the author.

Chapter 2
DISTRIBUTION OF PRIMES
I : INTRODUCTION

2.1. Euclids Theorem Revisited

We have already seen the elegant and simple proof of Euclids theorem, that there are innitely many
primes. Here we shall begin by proving a slightly stronger result.

THEOREM 2.1. The series 1

p
p

is divergent.

Proof. For every real number X 2, write

  1
1
PX = 1 .
p
pX

Then
  
1
log PX = log 1 = S1 + S2 ,
p
pX

where
 1
 1
S1 = and S2 = .
p hph
pX pX h=2

This chapter was rst used in lectures given by the author at Imperial College, University of London, in 1981.
22 W W L Chen : Elementary and Analytic Number Theory

Since



1 1 1
0 = ,
hp h p h p(p 1)
h=2 h=2

we have


1 1
0 S2 = 1,
p
p(p 1) n=2 n(n 1)

so that 0 S2 1. On the other hand, as X , we have




  1  1
PX = .
ph n
pX h=0 nX

The result follows.

For every real number X 2, we write



(X) = 1,
pX

so that (X) denotes the number of primes in the interval [2, X]. This function has been studied
extensively by number theorists, and attempts to study it in depth have led to major developments in
other important branches of mathematics.
As can be expected, many conjectures concerning the distribution of primes were made based purely
on numerical evidence, including the celebrated Prime number theorem, proved in 1896 by Hadamard
and de la Vallee Poussin, that
(X) log X
lim = 1.
X X
We shall give an analytic proof of this in Chapter 5. Here we shall be concerned with the weaker result
of Tchebyche, that there exist positive absolute constants c1 and c2 such that for every real number
X 2, we have
X X
c1 < (X) < c2 .
log X log X

2.2. The Von Mangoldt Function

The study of the function (X) usually involves, instead of the characteristic function of the primes, a
function which counts not only primes, but prime powers as well, and with weights. Accordingly, we
introduce the von Mangoldt function : N C, dened for every n N by writing

log p if n = pr , with p prime and r N,
(n) =
0 otherwise.

THEOREM 2.2. For every n N, we have



(m) = log n.
m|n

Proof. The result is clearly true for n = 1, so it remains to consider the case n 2. Suppose that
n = pu1 1 . . . pur r is the canonical decomposition of n. Then the only non-zero contribution to the sum
Chapter 2 : Distribution of Primes I : Introduction 23

v
on the left hand side comes from those natural numbers m of the form m = pj j with j = 1, . . . , r and
1 vj uj . It follows that

 
r 
uj

r
u
(m) = log pj = log pj j = log n.
m|n j=1 vj =1 j=1

THEOREM 2.3. As X , we have




X
(m) = X log X X + O(log X).
m
mX

Proof. It follows from Theorem 2.2 that


    

X
log n = (m) = (m) 1= (m) .
m
nX nX m|n mX nX mX
m|n

It therefore suces to prove that as X , we have



log n = X log X X + O(log X). (1)
nX

To prove (1), note that log X is an increasing function of X. In particular, for every n N, we have
n+1
log n log u du,
n

so that
 X
log n log(X + 1) log u du.
nX 1

On the other hand, for every n N, we have


n
log n log u du,
n1

so that
  [X]
log n = log n log u du
nX 2nX 1
X X X
= log u du log u du log u du log X.
1 [X] 1

The inequality (1) now follows on noting that


X
log u du = X log X X + 1.
1

2.3. Tchebyche s Theorem

The crucial step in the proof of Tchebyches theorem concerns obtaining bounds on sums involving the
von Mangoldt function. More precisely, we prove the following result.
24 W W L Chen : Elementary and Analytic Number Theory

THEOREM 2.4. There exist positive absolute constants c3 and c4 such that
 1
(m) X log 2 (X c3 ) (2)
2
mX

and 
(m) c4 X (X 0). (3)
X
2 <mX

Proof. If m N satises X/2 < m X, then clearly [X/2m] = 0. It follows from this and Theorem
2.3 that as X , we have
 

 


X X X X
(m) 2 = (m) 2 (m)
m 2m m 2m
mX mX m X
 
2

X X X
= (X log X X + O(log X)) 2 log + O(log X)
2 2 2
= X log 2 + O(log X).

Hence there exists a positive absolute constant c5 such that for all suciently large X, we have
 


1 X X
X log 2 < (m) 2 < c5 X.
2 m 2m
mX

We now consider the function [] 2[/2]. Clearly [] 2[/2] < 2(/2 1) = 2. Note that the
left hand side is an integer, so we must have [] 2[/2] 1. It follows that for all suciently large X,
we have 
1
X log 2 < (m).
2
mX

The inequality (2) follows. On the other hand, if X/2 < m X, then [X/m] = 1 and [X/2m] = 0, so
that for all suciently large X, we have

(m) c5 X.
X
2 <mX

The inequality (3) follows easily.

We now state and prove Tchebyches theorem.

THEOREM 2.5. (Tchebyche) There exist positive absolute constants c1 and c2 such that for every
real number X 2, we have
X X
c1 < (X) < c2 .
log X log X

Proof. To prove the lower bound, note that


   
(m) = log p = (log p) 1
mX p,n pX log X
1n[ ]
pn X log p

log X
= (log p) (X) log X.
log p
pX

It follows from (2) that


X log 2
(X) (X c3 ).
2 log X
Chapter 2 : Distribution of Primes I : Introduction 25

Since (2) = 1, we get the lower bound for a suitable choice of c1 .


To prove the upper bound, note that in view of (3) and the denition of the von Mangoldt function,
the inequality
 X
log p c4 j
X X
2
<p
2j+1 2j

holds for every integer j 0 and every real number X 0. Suppose that X 2. Let the integer k 0
be dened such that 2k < X 1/2 2k+1 . Then

 
k  
k
log p log p c4 X 2j < 2c4 X,
X 1/2 <pX j=0 X
<p Xj j=0
2j+1 2

so that
  log p 4c4 X
1 < ,
log X 1/2 log X
X 1/2 <pX X 1/2 <pX

whence
4c4 X c2 X
(X) X 1/2 + <
log X log X
for a suitable c2 .

2.4. Some Results of Mertens

We conclude this chapter by obtaining an improvement of Theorem 2.1.

THEOREM 2.6. (Mertens) As X , we have


 (m)
= log X + O(1), (4)
m
mX
 log p
= log X + O(1), (5)
p
pX
and
 1
= log log X + O(1). (6)
p
pX

Proof. Recall Theorem 2.3. As X , we have




X
(m) = X log X X + O(log X).
m
mX

Clearly [X/m] = X/m + O(1), so that as X , we have




 (m) 
X
(m) =X +O (m) .
m m
mX mX mX

It follows from (3) that



 
(m) (m) 2c4 X,
mX j=0 X
<m Xj
2j+1 2
26 W W L Chen : Elementary and Analytic Number Theory

so that as X , we have
 (m)
X = X log X + O(X).
m
mX

The inequality (4) follows. Next, note that

 (m)  log p  log p   1


= = + (log p) .
m pk p pk
mX p,k pX pX 2k log X
log p
pk X

As X , we have

  
  log p

1 1 log n
(log p) (log p) = = O(1).
pk pk p(p 1) n=2 n(n 1)
pX 2k log X pX k=2 pX
log p

The inequality (5) follows. Finally, for every real number X 2, let
 log p
T (X) = .
p
pX

Then it follows from (5) that there exists a positive absolute constant c6 such that |T (X) log X| < c6
whenever X 2. On the other hand,
 X  X
 1  log p 1 dy T (X) T (y) dy
= + 2 = +
p
pX
p
pX
log X p y log y log X 2 y log2 y
X X
T (X) log X (T (y) log y) dy dy
= + 2 + 1 + .
log X 2 y log y 2 y log y

It follows that as X , we have


 
  X
 1  c6 c6 dy
 
log log X  < + + 1 log log 2 = O(1).
 p log X y log2 y
pX  2

The inequality (6) follows.


ELEMENTARY AND
ANALYTIC NUMBER THEORY
W W L CHEN


c W W L Chen, 1990.
This work is available free, in the hope that it will be useful.
Any part of this work may be reproduced or transmitted in any form or by any means, electronic or mechanical, including
photocopying, recording, or any information storage and retrieval system, with or without permission from the author.

Chapter 3
DIRICHLET SERIES

3.1. Convergence Properties

A Dirichlet series is a series of the type




F (s) = f (n)ns , (1)
n=1

where f : N C is an arithmetic function and s C. We usually write s = + it, where , t R.


Our rst task is to investigate convergence of Dirichlet series.

THEOREM 3.1. Suppose that the series (1) converges for some s C. Then there exist unique real
numbers 0 , 1 , 2 satisfying 0 1 2 < and such that the following statements hold:
(i) The series (1) converges for every s C with > 0 . Furthermore, for every
> 0, the series (1)
diverges for some s C with 0
< 0 .
(ii) For every > 0, the series (1) converges uniformly on the set {s C : > 1 + } and does not
converge uniformly on the set {s C : > 1 }.
(iii) The series (1) converges absolutely for every s C with > 2 . Furthermore, for every
> 0, the
series (1) does not converge absolutely for some s C with 2
< 2 .

Example. The Dirichlet series




(s) = ns
n=1

converges absolutely for every s C with > 1 and diverges for every real s < 1. It follows that
0 = 1 = 2 = 1 in this case.

This chapter was rst used in lectures given by the author at Imperial College, University of London, in 1990.
32 W W L Chen : Elementary and Analytic Number Theory

Proof of Theorem 3.1. Suppose that the series (1) converges for s = s = + it . Then

f (n)ns 0 as n , so that |f (n)ns | = O(1), and so |f (n)| = O(n ). It follows that for
every s C with > + 1, we have

|f (n)ns | = |f (n)n | = O(n
),

so that the series (1) converges by the Comparison test. Now let

0 = inf{u R : the series (1) converges for all s C with > u}

and
2 = inf{u R : the series (1) converges absolutely for all s C with > u}.
Clearly (i) and (iii) follow, and 0 2 . To prove (ii), let > 0 and
> 0 be chosen. Then there exists
N N such that

|f (n)|n2 <
.
n=N +1

Hence  N  

 

 s s 
sup  f (n)n f (n)n  : 2 + |f (n)|n2 <
.
 
n=1 n=1 n=N +1

It follows that the series (1) converges uniformly on the set {s C : 2 + }. Now let

1 = inf{u R : the series (1) converges uniformly on {z C : u}}.

Clearly 0 1 2 + . Since > 0 is arbitrary, we must have 0 1 2 .

A simple consequence of uniform convergence is the following result concerning dierentiation term
by term.

THEOREM 3.2. For every s C with > 1 , the series (1) may be dierentiated term by term. In
particular, F  (s) exists and


F  (s) = f (n)(log n)ns .
n=1

3.2. Uniqueness Properties

Our next task is to prove the uniqueness theorem of Dirichlet series, a result of great importance in view
of the applications we have in mind.

THEOREM 3.3. Suppose that





F (s) = f (n)ns and G(s) = g(n)ns ,
n=1 n=1

where f : N C and g : N C are arithmetic functions and s C. Suppose further that there exists
3 R such that for every s C satisfying 3 , we have F (s) = G(s). Then f (n) = g(n) for every
n N.

It is clearly sucient to prove the following special case.


Chapter 3 : Dirichlet Series 33

THEOREM 3.4. Suppose that




F (s) = f (n)ns ,
n=1

where f : N C is an arithmetic function and s C. Suppose further that there exists 3 R such
that for every s C satisfying 3 , we have F (s) = 0. Then f (n) = 0 for every n N.

Proof. Since the series converges for s = 3 , we must have |f (n)| = O(n3 ) for all n N. Now let
3 + 2. Then  
 
f (n)n = O n3 . (2)
n=N n=N

Note next that y 3 is a decreasing function of y, so that




 
3 3 3 3
n =N + n N + y 3 dy = O(N 3 +1 ). (3)
n=N n=N +1 N

Combining (2) and (3), we see that for every N N, we have




f (n)n = O(N 3 +1 ). (4)
n=N

Using (4) with N = 2, we obtain, for 3 + 2,




0 = F () = f (1) + f (n)n = f (1) + O(23 +1 ) f (1)
n=2

as +. Hence f (1) = 0. Suppose now that f (1) = f (2) = . . . = f (M 1) = 0. Using (4) with
N = M + 1, we obtain, for 3 + 2,


0 = F () = f (M )M + f (n)n = f (M )M + O((M + 1)3 +1 ),
n=M +1

so that

M
0 = f (M ) + O (M + 1)3 +1 f (M )
M +1
as +. Hence f (M ) = 0. The result now follows from induction.

3.3. Multiplicative Properties

Dirichlet series are extremely useful in tackling problems in number theory as well as in other branches
of mathematics. The main properties that underpin most of these applications are the multiplicative
aspects of these series.

THEOREM 3.5. Suppose that for every j = 1, 2, 3, we have




Fj (s) = fj (n)ns ,
n=1

where fj : N C is an arithmetic function and s C. Suppose further that for every n N, we have
  n  n

f3 (n) = f1 (x)f2 (y) = f1 (x)f2 = f1 f2 (y).


x,y x y
xy=n x|n y|n
34 W W L Chen : Elementary and Analytic Number Theory

Then
F1 (s)F2 (s) = F3 (s),
(1) (2)
provided that > max{2 , 2 }, where, for every j = 1, 2, the series Fj (s) converges absolutely for
(j)
every s C with > 2 .

Proof. We have

N 
f3 (n)ns = f1 (x)xs f2 (y)y s ,
n=1 1xN
1yN
xyN

so that


N  
f3 (n)ns f1 (x)xs f2 (y)y s

n=1 x N y N
   
= f1 (x)xs f2 (y)y s + f1 (x)xs f2 (y)y s .

N <xN yN/x x N N <yN/x

It follows that
 
N 
   
 s
f3 (n)n f1 (x)xs s 
f2 (y)y 

n=1
x N

y N

 
 
< |f1 (x)|x

|f2 (y)|y

x> N y=1


 
+ |f1 (x)|x |f2 (y)|y . (5)

x=1 y> N

(1) (2)
Suppose now that > max{2 , 2 }. Clearly
 
|f1 (x)|x and |f2 (y)|y

x> N y> N

converge to 0 as N . Furthermore, the series





|f1 (x)|x and |f2 (y)|y
x=1 y=1

are convergent. It follows that the right hand side of (5) converges to 0 as N . On the other hand,
 
f1 (x)xs and f2 (y)y s

x N y N

converge to F1 (s) and F2 (s) respectively as N . The result follows.

Remark. Theorem 3.5 generalizes to a product of k Dirichlet series F1 (s), . . . , Fk (s), where the general
coecient is 
f1 (x1 ) . . . fk (xk ).
x1 ,...,xk
x1 ...xk =n
Chapter 3 : Dirichlet Series 35

In many applications, the coecients f (n) of the Dirichlet series will be given by various important
arithmetic functions in number theory. We therefore study next some consequences when the function
f : N C is multiplicative.

THEOREM 3.6. Suppose that the function f : N C is multiplicative. Then for every s C
satisfying > 2 , the series (1) satises
 
 
h hs
F (s) = f (p )p .
p h=0

Proof. By the Remark, if pj is the j-th prime in increasing order, then



 

 

k
hk s
f (phj )phs = f (p h1
) . . . f (p k n
) .
j=1
j
n=1
1

h=0 h1 ,...,hk
h h
p 1 ...p k =n
1 k

By the uniqueness of factorization, the inner sum on the right hand side contains at most one term.
Hence  
k  
h hs
f (pj )pj = k (n)f (n)ns ,
j=1 h=0 n=1

where 
1 if all the prime factors of n are among p1 , . . . , pk ,
k (n) =
0 otherwise.
It follows that as k , we have
 

k   
h hs
f (pj )pj f (n)ns = (k (n) 1)f (n)ns
j=1 h=0 n=1 n=1



=O |f (n)|n 0.
n=k+1

An arithmetic function f : N C is said to be totally multiplicative or strongly multiplicative if


f (mn) = f (m)f (n) for every m, n N.

THEOREM 3.7. Suppose that the function f : N C is totally multiplicative. Then for every s C
satisfying > 2 , the series (1) satises

F (s) = (1 f (p)ps )1 .
p

Proof. The absolute convergence of the series




f (ph )phs (6)
h=0

is immediate for > 2 by comparison with the series




|f (n)|n .
n=1

Furthermore, if f is not identically zero, then it is easy to see that f (1) = 1, so that the series (6) is now
a convergent geometric series with sum (1 f (p)ps )1 .
36 W W L Chen : Elementary and Analytic Number Theory

Example. Consider again the Dirichlet series




(s) = ns .
n=1

For every s C satisfying > 1, we have



(s) = (1 ps )1 .
p

This is called the Euler product of the Riemann zeta function (s).
ELEMENTARY AND
ANALYTIC NUMBER THEORY
W W L CHEN


c W W L Chen, 1990.
This work is available free, in the hope that it will be useful.
Any part of this work may be reproduced or transmitted in any form or by any means, electronic or mechanical, including
photocopying, recording, or any information storage and retrieval system, with or without permission from the author.

Chapter 4
DISTRIBUTION OF PRIMES
II : ARITHMETIC PROGRESSIONS

4.1. Dirichlets Theorem

The purpose of this chapter is to prove the following remarkable result of Dirichlet, widely regarded as
one of the greatest achievements in mathematics.

THEOREM 4.1. Suppose that q N and a Z satisfy (a, q) = 1. Then there are innitely many
primes p a (mod q).

Note that the requirement (a, q) = 1 is crucial. If n a (mod q), then clearly (a, q) | n. It follows
that if (a, q) > 1, then the residue class n a (mod q) of natural numbers contains at most one prime.
In other words, Dirichlets theorem asserts that any residue class n a (mod q) of natural numbers
must contain innitely many primes if there is no simple reason to support the contrary.
It is easy to prove Theorem 4.1 by elementary methods for some special values of a and q.

Example. There are innitely many primes p 1 (mod 4). Suppose on the contrary that p1 , . . . , pr
represent all such primes. Then 4p1 . . . pr 1 must have a prime factor p 1 (mod 4). But p cannot
be any of p1 , . . . , pr .

Example. There are innitely many primes p 1 (mod 4). Suppose on the contrary that p1 , . . . , pr
represent all such primes. Consider the number 4(p1 . . . pr )2 + 1. Suppose that a prime p divides
4(p1 . . . pr )2 + 1. Then 4(p1 . . . pr )2 + 1 0 (mod p). It follows that 1 is a quadratic residue modulo
p, so that we must have p 1 (mod 4). Clearly p cannot be any of p1 , . . . , pr .

This chapter was rst used in lectures given by the author at Imperial College, University of London, in 1990.
42 W W L Chen : Elementary and Analytic Number Theory

4.2. A Special Case

The idea of Dirichlet is to show that if (a, q) = 1, then the series


 1
p
pa (mod q)

is divergent. For technical reasons, it is easier to show that if (a, q) = 1, then

 log p
+ as 1 + .
p
pa (mod q)

Let us illustrate the idea of Dirichlet by studying the case n 1 (mod 4).
First of all, we need a function that distinguishes between integers n 1 (mod 4) and the others.
Suppose that n is odd. Then it is easy to check that
n1 
1 + (1) 2 1 if n 1 (mod 4),
=
2 0 if n 1 (mod 4);

so that
 log p 1  log p  p1

= 1 + (1) 2 .
p 2 p
p1 (mod 4) p odd

Now the series


 log p
+ as 1+,
p
p odd

so it suces to show that the series


 (1) p1
2 log p
p
p odd

converges as 1+.
The next idea is to show that if we consider the series

 n1
(1) 2 (n)
(1)
n=1
n
n odd

instead, then the contribution from the terms corresponding to non-prime odd natural numbers n is
convergent. It therefore suces to show that the series (1) converges as 1+.
Note now that the function
 n1

(n) = (1) if n is odd,


2

0 if n is even,

is totally multiplicative; in other words, (mn) = (m)(n) for every m, n N. Write


(n)
L() = , (2)
n=1
n

and note that for every n N, we have


  n
(n) log n = (n) (m) = (m)(m) .
m
m|n m|n
Chapter 4 : Distribution of Primes II : Arithmetic Progressions 43

It follows from Theorems 3.2 and 3.5 that for > 1, we have


  
 (n) log n  (n)(n)  (n)

L () = = .
n=1
n n=1
n n=1
n

Hence

(1) 2 (n)  (n)(n)
n1
L ()
= = .
n=1
n n=1
n L()
n odd

Now as 1+, we expect

1 1 1
L() L(1) = 1 + + ... > 0
3 5 7

and
log 3 log 5 log 7
L () + ...
3 5 7
which converges by the Alternating series test. We therefore expect the series (1) to converge to a nite
limit.

4.3. Dirichlet Characters

Dirichlets most crucial discovery is that for every q N, there is a family of (q) functions : N C,
known nowadays as the Dirichlet characters modulo q, which generalize the function in the special
case and which satisfy
1  (n)  1 if n a (mod q),
=
(q) (a) 0 if n a (mod q).
mod q

To understand Dirichlets ideas, we shall rst of all study group characters. This is slightly more
general than is necessary, but easier to understand.
Let G be a nite abelian group of order h and with identity element e. A character on G is a
non-zero complex-valued function on G for which (uv) = (u)(v) for every u, v G. It is easy to
check the following simple results.

Remark. We have
(i) (e) = 1;
(ii) for every u G, (u) is an h-th root of unity;
(iii) the number c of characters is nite; and
(iv) the characters form an abelian group.

Slightly less trivial is the following.

Remark. If u G and u = e, then there exists a character on G such that (u) = 1. To see
this, note that G can be expressed as a direct product of cyclic groups G1 , . . . , Gs of orders h1 , . . . , hs
respectively, where h = h1 . . . hs . Suppose that for each j = 1, . . . , s, the cyclic group Gj is generated by
vj . Then we can write u = v1y1 . . . vsys , where yj (mod hj ) is uniquely determined for every j = 1, . . . , s.
Since u = e, there exists k = 1, . . . , s such that yk 0 (mod hk ). Let (vk ) = e(1/hk ), and let (vj ) = 1
for every j = 1, . . . , s such that j = k. Clearly (u) = e(yk /hk ) = 1.

 shall denote by 0 the principal character on G. In other words, 0 (u) = 1 for every u G.
We
Also, denotes a summation over all the distinct characters on G.
44 W W L Chen : Elementary and Analytic Number Theory

THEOREM 4.2. Suppose that G is a nite abelian group of order h and with identity element e.
Suppose further that 0 is the principal character on G.
(i) For every character on G, we have
 
h if = 0 ,
(u) =
0 0 .
if =
uG

(ii) For every u G, we have 


 c if u = e,
(u) =
0 e,
if u =

where c denotes the number of distinct characters on G.


(iii) We have c = h.
(iv) For every u, v G, we have 
1  (u) 1 if u = v,
=
h (v) v.
0 if u =

Proof. (i) If = 0 , then the result is obvious. If = 0 , then there exists v G such that (v) = 1,
and so    
(v) (u) = (u)(v) = (uv) = (u),
uG uG uG uG

the last equality following from the fact that uv runs over all the elements of G as u runs over all the
elements of G. Hence 
(1 (v)) (u) = 0.
uG

The result follows since (v) = 1.


(ii) If u = e, then the result is obvious. If u = e, then we have already shown that there exists a
character 1 such that 1 (u) = 1, so that
   
1 (u) (u) = 1 (u)(u) = (1 )(u) = (u),

the last equality following from noting that the characters on G form an abelian group so that 1 runs
through all the characters on G as runs through all the characters on G. Hence

(1 1 (u)) (u) = 0.

The result follows since 1 (u) = 1.


(iii) Note that  
h= (u) = (u) = c.
uG uG

(iv) Note that



1  (u) 1 1 c/h if uv 1 = e,
= (u)(v 1 ) = (uv 1 ) =
h (v) h h 0 if uv 1 = e.

The result follows since h = c.

We are now in a position to introduce Dirichlet characters. Let q N be given. Then there
are exactly (q) residue classes n a (mod q) satisfying (a, q) = 1. Under multiplication of residue
classes, they form an abelian group of order (q). Suppose that these residue classes are represented
by a1 , . . . , a(q) modulo q. Let G = {a1 , . . . , a(q) }. We can now dene a character on the group G
Chapter 4 : Distribution of Primes II : Arithmetic Progressions 45

as described earlier, interpreting the group elements as residue classes. Furthermore, we can extend the
denition to cover the remaining residue classes. Precisely, for every n N, let

(aj ) if n aj (mod q) for some j = 1, . . . , (q),
(n) = (3)
0 if (n, q) > 1.

A function : N C of the form (3) is called a Dirichlet character modulo q. Note that is totally
multiplicative. Also, clearly there are exactly (q) Dirichlet characters modulo q. Furthermore, the
principal Dirichlet character 0 modulo q is dened by

1 if (n, q) = 1,
0 (n) =
0 if (n, q) > 1.

The following theorem follows immediately from these observations and Theorem 4.2.

THEOREM 4.3. Suppose that q N. Suppose further that 0 is the principal Dirichlet character
modulo q.
(i) For every Dirichlet character modulo q, we have
 
(q) if = 0 ,
(n) =
0 if = 0 .
n (mod q)

(ii) For every n N, we have


 
(q) if n 1 (mod q),
(n) =
0 if n 1 (mod q).
(mod q)

(iii) For every a Z satisfying (a, q) = 1 and for every n N, we have

1  (n)  1 if n a (mod q),


=
(q) (a) 0 if n a (mod q).
(mod q)

4.4. Some Dirichlet Series

Our next task is to introduce the functions analogous to the function (2) earlier. Let s = + it C,
where , t R. For > 1, let

(s) = ns ; (4)
n=1

furthermore, for any Dirichlet character modulo q, let




L(s, ) = (n)ns . (5)
n=1

The functions (4) and (5) are called the Riemann zeta function and Dirichlet L-functions respectively.
Note that the series are Dirichlet series and converge absolutely for > 1 and uniformly for > 1 +
for any > 0. Furthermore, the coecients are totally multiplicative. It follows from Theorem 3.7 that
for > 1, the series (4) and (5) have the Euler product representations

(s) = (1 ps )1 and L(s, ) = (1 (p)ps )1
p p

respectively. The following are some simple properties of these functions.


46 W W L Chen : Elementary and Analytic Number Theory

THEOREM 4.4. Suppose that > 1. Then (s) = 0. Furthermore, L(s, ) = 0 for every Dirichlet
character modulo q.

Proof. Since > 1, we have





1 p

s 1
(2)
|(s)| =
(1 p )
(1 + p )1 = = >0


1p 2 ()
p p p

and





s 1

|L(s, )| =
(1 (p)p )
(1 + p )1 (1 + p )1 > 0.

p pq p

THEOREM 4.5. Suppose that 0 is the principal Dirichlet character modulo q. Then for > 1, we
have
L(s, 0 ) = (s) (1 ps ).
p|q

Proof. Since > 1, we have



L(s, 0 ) = (1 0 (p)ps )1 = (1 ps )1
p pq

(1 ps )1
p
= = (s) (1 ps ).
(1 ps )1 p|q
p|q

THEOREM 4.6. Suppose that > 1. Then



 (s) 
= (n)ns .
(s) n=1

Furthermore, for every Dirichlet character modulo q, we have



L (s, ) 
= (n)(n)ns .
L(s, ) n=1

Proof. Since > 1, it follows from Theorem 3.2 that




 (s) = (log n)ns .
n=1

It now follows from Theorem 3.5 and 


log n = (m)
m|n

that 


 
 s s
(s) = (n)n n .
n=1 n=1
Chapter 4 : Distribution of Primes II : Arithmetic Progressions 47

The rst assertion follows. On the other hand, it also follows from Theorem 3.2 that


L (s, ) = (n)(log n)ns .
n=1

It now follows from Theorem 3.5 and


 n
(n) log n = (m)(m)
m
m|n

that 


 
L (s, ) = (n)(n)ns (n)ns .
n=1 n=1

The second assertion follows.

THEOREM 4.7. If > 1, then for every Dirichlet character modulo q, we have


log L(s, ) = m1 (pm )pms .
p m=1

Proof. Taking logarithms on the Euler product representation, we have



log L(s, ) = log (1 (p)ps )1 = log(1 (p)ps )1 , (6)
p p

so that 
log L(s, ) = log(1 (p)ps ).
p

The justication for (6) is that the series on the right hand side converges uniformly for > 1 + , as
can be deduced from the Weierstrass M -test on noting that

| log(1 (p)ps )| 2|(p)ps | 2p1 .

The proof is now completed by expanding log(1 (p)ps ).

4.5. Analytic Continuation

Our next task is to extend the denition of (s) and L(s, ) to the half plane > 0. This is achieved
by analytic continuation.
An example of analytic continuation is the following: Consider the geometric series


f (s) = sn .
n=0

This series converges absolutely in the set {s C : |s| < 1} and uniformly in the set {s C : |s| < 1 }
for any > 0 to the sum 1/(1 s). Now let
1
g(s) =
1s
in C. Then g is analytic in the set C \ {1}, g(s) = f (s) in the set {s C : |s| < 1}, and g has a pole at
s = 1. So g can be viewed as an analytic continuation of f to C with a pole at s = 1.
We shall prove the following results on analytic continuation of (s) and L(s, ).
48 W W L Chen : Elementary and Analytic Number Theory

THEOREM 4.8. The function (s) admits an analytic continuation to the half plane > 0. Fur-
thermore, (s) is analytic for > 0 except for a simple pole at s = 1 with residue 1.

THEOREM 4.9. Suppose that q N and 0 is the principal Dirichlet character modulo q. Then
the function L(s, 0 ) admits an analytic continuation to the half plane > 0. Furthermore, L(s, 0 ) is
analytic for > 0 except for a simple pole at s = 1 with residue (q)/q.

THEOREM 4.10. Suppose that q N and is a non-principal Dirichlet character modulo q. Then
the function L(s, ) admits an analytic continuation to the half plane > 0. Furthermore, L(s, ) is
analytic for > 0.

The proofs of these three theorems depend on the following two simple technical results. The rst
of these is basically a result on partial summation.

THEOREM 4.11. Suppose that a(n) = O(1) for every n N. For every x > 0, write

S(x) = a(n).
nx

Suppose further that for > 1, we have




F (s) = a(n)ns .
n=1

Then for every X > 0 and > 1, we have

 X
s s
a(n)n = S(X)X +s S(x)xs1 dx. (7)
nX 1

Furthermore, for > 1, we have


F (s) = s S(x)xs1 dx. (8)
1

Proof. To prove (7), simply note that

   X
s s s s
a(n)n S(X)X = a(n)(n X )= a(n) sxs1 dx
nX nX nX n

X  X
=s a(n) xs1 dx = s S(x)xs1 dx.
1 nx 1

Also, (8) follows from (7) on letting X .

The second technical result, standard in complex function theory, will be stated without proof.

THEOREM 4.12. Suppose that the path is dened by w(t) = u(t) + iv(t), where u(t), v(t) R for
every t [0, 1]. Suppose further that u (t) and v  (t) are continuous on [0, 1]. Let D be a domain in C.
For every s D, let
F (s) = f (s, w) dw,

where
(i) f (s, w) is continuous for every s D and every w ; and
(ii) for every w , the function f (s, w) is analytic in D.
Then F (s) is analytic in D.
Chapter 4 : Distribution of Primes II : Arithmetic Progressions 49

Proof of Theorem 4.8. Let F (s) = (s). In the notation of Theorem 4.11, we have a(n) = 1 for
every n N, so that S(x) = [x] for every x > 0. It follows from (8) that

(s) = s [x]xs1 dx = s xs dx s {x}xs1 dx
1 1 1

1 s1
=1+ s {x}x dx.
s1 1

We shall show that the last term on the right hand side represents an analytic function for > 0. We
can write 
{x}xs1 dx = Fn (s),
1 n=1

where for every n N, n+1


Fn (s) = {x}xs1 dx.
n

It remainsto show that (i) for every n N, the function Fn (s) is analytic in C; and (ii) for every > 0,

the series n=1 Fn (s) converges uniformly for > . To show (i), note that by a change of variable,
1 1
Fn (s) = t(n + t)s1 dt = te(s+1) log(n+t) dt,
0 0

and (i) follows from Theorem 4.12. To show (ii), note that for > , we have


n+1

|Fn (s)| =

{x}x s1
dx

n1 < n1 ,
n

and (ii) follows from the Weierstrass M -test.

Proof of Theorem 4.9. Suppose that > 1. Recall Theorem 4.5, that
 1

L(s, 0 ) = (s) 1 s .
p
p|q

Clearly the right hand side is analytic for > 0 except for a simple pole at s = 1. Furthermore, at
s = 1, the function (s) has a simple pole with residue 1, while
 1

(q)
1 = .
p q
p|q

The result follows.

The proof of Theorem 4.10 is left as an exercise.

4.6. Proof of Dirichlets Theorem

We now attempt to prove Theorem 4.1. The following theorem will enable us to consider the analogue
of (1).

THEOREM 4.13. Suppose that > 1. Then


 log p  (n)
= + O(1).
p n
pa (mod q) na (mod q)
410 W W L Chen : Elementary and Analytic Number Theory

Proof. Note rst of all that the sum on the left hand side does not exceed the rst term on the right
hand side. On the other hand, we have

 
(n) log p   log p

n p p m=2
pm
na (mod q) pa (mod q)


log p  log p  log n
= = O(1).
p m=2
p m
p
p(p 1) n=2 n(n 1)

The result follows.

Combining Theorems 4.13, 4.3 and 4.6, we have


 log p 
= (n)n + O(1)
p
pa (mod q) na (mod q)


 
= 1 (n)
(n)n + O(1)
n=1
(q) (a)
(mod q)


1 1 
= (n)(n)n + O(1)
(q) (a) n=1
(mod q)

1  1 L (, )
= + O(1). (9)
(q) (a) L(, )
(mod q)

Suppose now that


 1 L (, )
= O(1) as 1 + . (10)
(a) L(, )
(mod q)
=0

Then combining (9) and (10), as 1+, we have

 log p 1 L (, 0 ) 1 1
= + O(1) = + O(1) ,
p (q) L(, 0 ) (q) 1
pa (mod q)

since the function L (s, 0 )/L(s, 0 ) has a simple pole at s = 1 with residue 1 by Theorem 4.9. To
complete the proof of Dirichlets theorem, it remains to prove (10). Clearly (10) will follow if we can
show that for every non-principal Dirichlet character (mod q), we have L(1, ) = 0. Here we need to
distinguish two cases, represented by the following two theorems.

THEOREM 4.14. Suppose that q N and is a non-real Dirichlet character modulo q. Then
L(1, ) = 0.

Proof. For > 1, we have, in view of Theorem 4.7, that

 

log L(, ) = (pm )m1 pm
(mod q) (mod q) p m=1


 
= (pm ) m1 pm
p m=1 (mod q)



= (q) m1 pm > 0,
p m=1
pm 1 (mod q)
Chapter 4 : Distribution of Primes II : Arithmetic Progressions 411

where the change of order of summation is justied since



|(pm )m1 pm |
(mod q) p m=1

is nite. It follows that



L(, )

> 1. (11)


(mod q)

Suppose that 1 is a non-real Dirichlet character modulo q, and L(1, 1 ) = 0. Then 1 = 1 , and
L(1, 1 ) = L(1, 1 ) = 0 also. It follows that these two zeros more than cancel the simple pole of L(, 0 )
at = 1, so that the product on the left hand side of (11) has a zero at = 1. This gives a contradiction.

Clearly this approach does not work when is real.

THEOREM 4.15. Suppose that q N and is a real, non-principal Dirichlet character modulo q.
Then L(1, ) = 0.

Proof. Suppose that the result is false, so that there exists a real Dirichlet character modulo q such
that L(1, ) = 0. Then the function
F (s) = (s)L(s, )
is analytic for > 0. Note that for > 1, we have


F (s) = f (n)ns ,
n=1

where for every n N, 


f (n) = (m).
m|n

Let the function g : N R be dened by



1 if n is a perfect square,
g(n) =
0 otherwise.
We shall rst of all show that for every n N, we have

f (n) g(n). (12)

Since is totally multiplicative, it suces to prove (12) when n = pk , where p is a prime and k N.
Indeed, since assumes only the values 1 and 0, we have

1 if (p) = 0,

k 2 k k + 1 if (p) = 1,
f (p ) = 1 + (p) + ((p)) + . . . + ((p)) =
1
if (p) = 1 and k is even,
0 if (p) = 1 and k is odd,

so that 
1 if k is even,
f (pk ) g(pk ) =
0 if k is odd.
Suppose now that 0 < r < 3/2. Since F (s) is analytic for > 0, we must have the Taylor expansion

 F () (2)
F (2 r) = (r) .
=0
!
412 W W L Chen : Elementary and Analytic Number Theory

Now by Theorem 3.2, we have




F () (2) = f (n)( log n) n2 .
n=1

It follows that for every N {0}, we have, in view of (12),



F () (2) r  r 
(r) = f (n)(log n) n2 g(n)(log n) n2
! ! n=1 ! n=1

r  2 2 2 (2r) 
= (log k ) (k ) = (log k) k 4
! !
k=1 k=1

(2r)  (2r) ()
= ( log k) k 4 = (4)
! !
k=1

by Theorem 3.2. It follows that for 0 < r < 3/2, we have



 (2r)
F (2 r) () (4) = (4 2r).
=0
!

Now as r 3/2, we must therefore have F (2 r) +. This contradicts our assertion that F (s) is
analytic for > 0 and hence continuous at s = 1/2.
ELEMENTARY AND
ANALYTIC NUMBER THEORY
W W L CHEN


c W W L Chen, 1990.
This work is available free, in the hope that it will be useful.
Any part of this work may be reproduced or transmitted in any form or by any means, electronic or mechanical, including
photocopying, recording, or any information storage and retrieval system, with or without permission from the author.

Chapter 5
DISTRIBUTION OF PRIMES
III : THE PRIME NUMBER THEOREM

5.1. Some Preliminary Remarks

In this chapter, we give an analytic proof of the famous Prime number theorem, a result rst obtained
in 1896 independently by Hadamard and de la Vallee Poussin.

THEOREM 5.1. As X , we have

X
(X) .
log X

As in our earlier study of the distribution of primes, we use the von Mangoldt function . For every
X > 0, let

(X) = (n).
nX

THEOREM 5.2. As X , we have

X
(X) X if and only if (X) .
log X

This chapter was rst used in lectures given by the author at Imperial College, University of London, in 1990.
52 W W L Chen : Elementary and Analytic Number Theory

Proof. Recall the proof of Theorem 2.5 due to Tchebyche. We have


   
(X) = (n) = log p = (log p) 1
nX p,k pX 1k log X
log p
pk X
  
log X
= (log p) (X) log X. (1)
log p
pX

On the other hand, for any (0, 1), we have


 
(X) log p log p ((X) (X )) log(X )
pX X <pX
= ((X) (X )) log X.
(2)
Combining (1) and (2), we have
(X) (X ) (X) (X)
. (3)
X/ log X X/ log X X X/ log X
Since < 1, it follows from Tchebyches theorem that as X , we have
(X )
0.
X/ log X
Suppose that as X , we have (X) X/ log X. Then as X , we have
(X) (X )
.
X/ log X X/ log X
It follows that for any > 0, the inequality
(X)
1+
X
holds for all large X. Since < 1 is arbitrary, we must have
(X)
1
X
as X .
Note next that the inequalities (3) can be rewritten as
(X) (X) 1 (X) (X )
+ .
X X/ log X X X/ log X
Suppose that as X , we have (X) X. Then a similar argument gives
(X)
1
X/ log X
as X .

5.2. A Smoothing Argument

To prove the Prime number theorem, it suces to show that as X , we have (X) X. However,
a direct discussion of (X) introduces various tricky convergence problems. We therefore consider a
smooth average of the function . For X > 0, let
 X
1 (X) = (x) dx. (4)
0
Chapter 5 : Distribution of Primes III : The Prime Number Theorem 53

THEOREM 5.3. Suppose that as X , we have 1 (X) 1 2


2X . Then as X , we have
(X) X.

Proof. Suppose that 0 < < 1 < . Since (n) 0 for every n N, the function is an increasing
function. Hence for every X > 0, we have

1 X
1 (X) 1 (X)
(X) (x) dx = ,
X X X ( 1)X

so that
(X) 1 (X) 1 (X)
. (5)
X ( 1)X 2
On the other hand, for every X > 0, we have

1 X
1 (X) 1 (X)
(X) (x) dx = ,
X X X (1 )X

so that
(X) 1 (X) 1 (X)
. (6)
X (1 )X 2
As X , we have  
1 (X) 1 (X) 1 1 2 1 1
= ( + 1) (7)
( 1)X 2 1 2 2 2
and  
1 (X) 1 (X) 1 1 1 2 1
= ( + 1). (8)
(1 )X 2 1 2 2 2
Since and are arbitrary, we conclude, on combining (5)(8), that as X , we have (X)/X 1.

The rest of this chapter is concerned with establishing the following crucial result.

THEOREM 5.4. As X , we have

1 2
1 (X) X .
2

5.3. A Contour Integral

The following theorem provides a link between 1 (X) and the Riemann zeta function (s).

THEOREM 5.5. Suppose that X > 0 and c > 1. Then



1 c+i
X s+1  (s)
1 (X) = ds,
2i ci s(s + 1) (s)

where the path of integration is the straight line = c.

A crucial step in the proof of Theorem 5.5 is provided by the following result concerning a particular
contour integral.
54 W W L Chen : Elementary and Analytic Number Theory

THEOREM 5.6. Suppose that X > 0 and c > 1. Then



 c+i s

0 (X 1),
1 X
ds =
2i ci s(s + 1)
1
1
(X 1).
X

Proof. Note rst of all that the integral is absolutely convergent, as



Xs Xc

s(s + 1) |t|2

whenever = c. Let T > 1, and write


 c+iT
1 Xs
IT = ds.
2i ciT s(s + 1)

Suppose rst of all that X 1. Let A (c, T ) denote the circular arc centred at s = 0 and passing
from c iT to c + iT on the left of the line = c. Furthermore, let

1 Xs
JT = ds.
2i A (c,T ) s(s + 1)

Note that on A (c, T ), we have |X s | = X X c since X 1; also we have |s| = R and |s + 1| R 1,


where R = (c2 + T 2 )1/2 is the radius of A (c, T ). It follows that

1 Xc Xc
|JT | 2R 0 as T .
2 R(R 1) T 1

By Cauchys residue theorem, we have


   
Xs Xs 1
IT = JT + res , 0 + res , 1 = JT + 1 .
s(s + 1) s(s + 1) X

The result for X 1 follows on letting T .


Suppose now that X 1. Let A+ (c, T ) denote the circular arc centred at s = 0 and passing from
c iT to c + iT on the right of the line = c. Furthermore, let

1 Xs
JT+ = ds.
2i A+ (c,T ) s(s + 1)

Note that on A+ (c, T ), we have |X s | = X X c since X 1; also we have |s| = R and |s + 1| R,


where R = (c2 + T 2 )1/2 is the radius of A+ (c, T ). It follows that

1 Xc Xc
|JT+ | 2R 0 as T .
2 R2 T
By Cauchys residue theorem, we have
IT = JT+ .
The result for X 1 follows on letting T .

Proof of Theorem 5.5. Note that for X 1, we have



 X  X  X  
1 (X) = (x) dx = (x) dx = (n) dx = (X n)(n),
0 1 1 nx nX
Chapter 5 : Distribution of Primes III : The Prime Number Theorem 55

the last equality following from interchanging the order of integration and summation. Note also that
the above conclusion holds trivially if 0 < X < 1. It therefore follows from Theorem 5.6 that for every
X > 0, we have
1 (X)  n  1

= 1 (n) = 1 (n)
X X X/n
nX nX
 
(n) c+i (X/n)s
= ds,
n=1
2i ci s(s + 1)

where c > 1. Since c > 1, the order of summation and integration can be interchanged, as
  c+i  
(n)(X/n)s (n) dt
|ds| X c
s(s + 1) n c c2 + t2
n=1 ci n=1

is nite. It follows that


  c+i
1 (X) 1 c+i
X s  (n) 1 X s  (s)
= ds = ds
X 2i ci s(s + 1) n=1 ns 2i ci s(s + 1) (s)

as required.

5.4. The Riemann Zeta Function

Recall rst of all Theorem 4.11. In the case of the Riemann zeta function, equation (7) of Chapter 4
becomes
  X
ns = s [x]xs1 dx + [X]X s
nX 1
 X  X
s
=s x dx s {x}xs1 dx + X s+1 {X}X s
1 1

s s X
{x} 1 {X}
= s dx + s1 . (9)
s 1 (s 1)X s1 1 xs+1 X Xs
Letting X , we deduce that

s {x}
(s) = s dx (10)
s1 1 xs+1
if > 1. Recall also that (10) gives an analytic continuation of (s) to > 0, with s simple pole at
s = 1. We shall use these formulae to deduce important information about the order of magnitude of
|(s)| in the neighbourhood of the line = 1 and to the left of it. Note that ( + it) and ( it) are
complex conjugates, so it suces to study (s) on the upper half plane.

THEOREM 5.7. For every 1 and t 2, we have


(i) |(s)| = O(log t); and
(ii) |  (s)| = O(log2 t).
Suppose further that 0 < < 1. Then for every and t 1, we have
(iii) |(s)| = O (t1 ).

Proof. For > 0, t 1 and X 1, we have, by (9) and (10), that


 1 
s 1 {X} {x}
(s) = + s dx
n s (s 1)X s1 X s1 X s
X x
s+1
nX

1 {X} {x}
= + s dx. (11)
(s 1)X s1 X s
X x
s+1
56 W W L Chen : Elementary and Analytic Number Theory

It follows that
 1 
1 1 dx
|(s)|
+ 1
+
+ |s| +1
n tX X X x
nX
 1  
1 1 t 1
+ + + 1 + . (12)
n tX 1 X X
nX

If 1, t 1 and X 1, then
 1 1 1 1+t t
|(s)| + + + (log X + 1) + 3 + .
n t X X X
nX

Choosing X = t, we obtain
|(s)| (log t + 1) + 4 = O(log t),
proving (i). On the other hand, if , t 1 and X 1, then it follows from (12) that

 1  
1 t 1
|(s)|
+ 1
+ 2 +
n tX X
nX
 [X]
dx X 1 3t X 1 3t
+ + + X 1 + .
0 x t X 1 X

Again choosing X = t, we obtain


 
1 3
|(s)| t1 +1+ , (13)
1

proving (iii). To deduce (ii), we may dierentiate (11) with respect to s and proceed in a similar way.
Alternatively, suppose that s0 = 0 + it0 satises 0 1 and t0 2. Let C be the circle with centre s0
and radius < 1/2. Then 
1 (s) M

| (s0 )| = ds ,
2i C (s s0 )2
where M = supsC |(s)|. Now for every s C, we clearly have 0 1 and 2t0 > t
t0 > 1. It follows from (13) that for every s C, we must have ( = 1 )
 
1 3 10t0
|(s)| (2t0 ) +1+ ,
1

since < 1/2 < 1 < 1. It follows that

10t0
|  (s0 )| .
2

We now take = (log t0 + 2)1 . Then t0 = e log t0 < e, and so

|  (s0 )| 10e(log t0 + 2)2 .

(ii) now follows.

THEOREM 5.8. The function (s) has no zeros on the line = 1. Furthermore, there is a positive
constant A such that as t , we have, for 1, that

1  
= O (log t)A .
(s)
Chapter 5 : Distribution of Primes III : The Prime Number Theorem 57

Proof. For every R, we clearly have

3 + 4 cos + cos 2 = 2(1 + cos )2 0. (14)

On the other hand, it is easy to check that for > 1, we have



 1
log (s) = ,
p m=1
mpms

so that 

 
it
log |( + it)| = R cn n = cn n cos(t log n), (15)
n=2 n=2

where 
1/m (n = pm , where p is prime and m N),
cn = (16)
0 (otherwise).
Combining (14)(16), we have


log | () ( + it)( + 2it)| =
3 4
cn n (3 + 4 cos(t log n) + cos(2t log n)) 0.
n=2

It follows that for > 1, we have



( + it) 4
|( 1)()|3 |( + 2it)| 1 . (17)
1 1

Suppose that the point s = 1 + it is a zero of (s). Then since (s) is analytic at the points s = 1 + it
and s = 1 + 2it and has a simple pole with residue 1 at s = 1, the left hand side of (17) must converge to
a nite limit as 1+, contradicting the fact that the right hand side diverges to innity as 1+.
Hence s = 1 + it cannot be a zero of (s). To prove the second assertion, we may assume without loss
of generality that 1 2, since for 2, we have

1  
s
(1 + p ) < () (2).
(s) = (1 p )
p p

Suppose now that 1 < 2 and t 2. Then by (17), we have

( 1)3 |( 1)()|3 |( + it)|4 |( + 2it)| A1 |( + it)|4 log(2t)

by Theorem 5.7(i), where A1 is a positive absolute constant. Since log(2t) 2 log t, it follows that

( 1)3/4
|( + it)| , (18)
A2 (log t)1/4

where A2 is a positive absolute constant. Note that (18) holds also when = 1. Suppose now that
1 < < 2. If 1 and t 2, then it follows from Theorem 5.7(ii) that


|( + it) ( + it)| = (x + it) dx A3 ( 1) log2 t,


where A3 is a positive absolute constant. Combining this with (18), we have

( 1)3/4
|( + it)| |( + it)| A3 ( 1) log2 t A3 ( 1) log2 t. (19)
A2 (log t)1/4
58 W W L Chen : Elementary and Analytic Number Theory

On the other hand, if 2 and t 2, then in view of (18), the inequality (19) must also hold. It
follows that inequality (19) holds if 1 2, t 2 and 1 < < 2. We now choose so that

( 1)3/4
= 2A3 ( 1) log2 t;
A2 (log t)1/4

in other words,
= 1 + (2A2 A3 )4 (log t)9 ,
where t > t0 so that < 2. Then

|( + it)| A3 ( 1) log2 t = A4 (log t)7

for 1 2 and t > t0 .

5.5. Completion of the Proof

We are now ready to complete the proof of Theorem 5.4. By Theorem 5.5, we have
 c+i
1 (X) 1
= G(s)X s1 ds, (20)
X2 2i ci

where c > 1 and X > 0, and where

1  (s) 1 1
G(s) = =  (s) .
s(s + 1) (s) s(s + 1) (s)

By Theorems 4.8, 5.7 and 5.8, we know that G(s) is analytic for 1, except at s = 1, and that for
some positive absolute constant A, we have
 
G(s) = O |t|2 (log |t|)2 (log |t|)A < |t|3/2 (21)

for all |t| > t0 . Let > 0 be given. We now consider a contour made up of the straight line segments

L1 = [1 iU, 1 iT ],


L2 = [1 iT, iT ],
L3 = [ iT, + iT ],



L4 = [ + iT, 1 + iT ],
L5 = [1 + iT, 1 + iU ],

where T = T ( ) > max{t0 , 2}, = (T ) = ( ) (0, 1) and U are chosen to satisfy the following
conditions:
(i) We have 
|G(1 + it)| dt < .
T

(ii) The rectangle [, 1] [T, T ] contains no zeros of (s). Note that this is possible since (s)
has no zeros on the line = 1 and, as an analytic function, has at most a nite number of zeros in the
region [1/2, 1) [T, T ].
(iii) We have U > T .
Furthermore, dene the straight line segments

M1 = [c iU, 1 iU ],
M2 = [1 + iU, c + iU ].
Chapter 5 : Distribution of Primes III : The Prime Number Theorem 59

The above contours are indicated in the diagram below:

1 + iU / 2
M
c + iU

O L5

+ iT / 1 + iT
L4

O L3 O

L2o
iT 1 iT

O L1

1 iU o c iU
M1

By Cauchys residue theorem, we have


 
1 
c+iU 2
1 s1
G(s)X ds = G(s)X s1 ds
2i ciU 2i j=1 Mj

1 
5
+ G(s)X s1 ds + res(G(s)X s1 , 1), (22)
2i j=1 Lj

where, for every X > 1, we have


1
res(G(s)X s1 , 1) = . (23)
2
Now   

G(s)X s1
ds = G(s)X s1
ds |G(1 + it)| dt < . (24)

L1 L5 T

On the other hand,


  
1
M
G(s)X s1
ds = G(s)X s1
ds M X 1 d (25)
log X
L2 L4

and 

G(s)X s1
ds 2T M X 1 , (26)

L3

where
M = M (, T ) = M ( ) = sup |G(s)|. (27)
L2 L3 L4

Furthermore, by (21), we have, for j = 1, 2,


 
c
X c1

G(s)X s1
ds |U |3/2 X 1 d |U |3/2 . (28)
Mj 1 log X
510 W W L Chen : Elementary and Analytic Number Theory

Combining (22)(28), we have



1  c+iU 1 X c1 |U |3/2
M TM
G(s)X s1 ds + + + .
2i ciU 2 log X X 1 log X

On letting U , we have

1 (X) 1 M TM

X 2 2 + log X + X 1 .

It clearly follows that as X , we have

1 (X) 1
.
X2 2
ELEMENTARY AND
ANALYTIC NUMBER THEORY
W W L CHEN


c W W L Chen, 1990.
This work is available free, in the hope that it will be useful.
Any part of this work may be reproduced or transmitted in any form or by any means, electronic or mechanical, including
photocopying, recording, or any information storage and retrieval system, with or without permission from the author.

Chapter 6
THE RIEMANN ZETA FUNCTION

6.1. Riemanns Memoir

In Riemanns only paper on number theory, published in 1860, he proved the following result.

THEOREM 6.1. The function (s) can be continued analytically over the whole plane, and satises
the functional equation
s  
s/2 (1s)/2 1s
(s) = (1 s), (1)
2 2

where denotes the gamma function. In particular, (s) is analytic everywhere, except for a simple pole
at s = 1 with residue 1.

Note that the functional equation (1) enables properties of (s) for < 0 to be inferred from
properties of (s) for > 1. In particular, the only zeros of (s) for < 0 are at the poles of (s/2); in
other words, at the points s = 2, 4, 6, . . . . These are called the trivial zeros of (s).
The part of the plane with 0 1 is called the critical strip.
Riemanns paper is particularly remarkable in the conjectures it contains. While most of the con-
jectures have been proved, the famous Riemann hypothesis has so far resisted all attempts to prove or
disprove it.

THEOREM 6.2. (Hadamard 1893) The function (s) has innitely many zeros in the critical strip.

It is easy to see that the zeros of (s) in the critical strip are placed symmetrically with respect to
the line t = 0 as well as with respect to the line = 1/2, the latter observation being a consequence of
the functional equation (1).

This chapter was rst used in lectures given by the author at Imperial College, University of London, in 1990.
62 W W L Chen : Elementary and Analytic Number Theory

THEOREM 6.3. (von Mangoldt 1905) Let N (T ) denote the number of zeros = + i of the
function (s) in the critical strip with 0 < T . Then
T T T
N (T ) = log + O(log T ). (2)
2 2 2

THEOREM 6.4. (Hadamard 1893) The entire function


1 s
(s) = s(s 1) s/2 (s) (3)
2 2
has the product representation
 s

(s) = eA+Bs 1 es/ , (4)

where A and B are constants and where runs over all the zeros of (s) in the critical strip.

We comment here that the product representation (4) plays an important role in the rst proof of
the Prime number theorem.
The most remarkable of Riemanns conjectures is an explicit formula for the dierence (X)li(X),
containing a term which is a sum over the zeros of (s) in the critical strip. This shows that the zeros of
(s) plays a crucial role in the study of the distribution of primes. Here we state a result closely related
to this formula.

THEOREM 6.5. (von Mangoldt 1895) Let


 (X 0) + (X + 0)
(X) = (n) and 0 (X) = .
2
nX

Then
 X  
 (0) 1 1
0 (X) X = log 1 2 ,

(0) 2 X

where the terms in the sum arising from complex conjugates are taken together.

However, there remains one of Riemanns conjectures which is still unsolved today. The open
question below is arguably the most famous unsolved problem in the whole of mathematics.

CONJECTURE. (Riemann Hypothesis) The zeros of the function (s) in the critical strip all lie
on the line = 1/2.

6.2. Riemanns Proof of Theorem 6.1

Suppose that > 0. Writing t = n2 x, we have


s  
s/21 t
xs/21 en x dx,
2
2 s/2
= t e dt = (n )
2 0 0

so that s 
s/2 ns = xs/21 en
2
x
dx.
2 0
It follows that for > 1, we have
 

s 

s/2 s/21 n2 x s/21 n2 x
(s) = x e dx = x e dx,
2 n=1 0 0 n=1
Chapter 6 : The Riemann Zeta Function 63

where the change of order of summation and integration is justied by the convergence of


x/21 en
2
x
dx.
n=1 0

Now write


en
2
x
(x) = .
n=1

Then for > 1, we have


s   1
s/2 (s) = xs/21 (x) dx + y s/21 (y) dy
2 1 0
 
= xs/21 (x) dx + xs/21 (x1 ) dx. (5)
1 1

We shall show that the function




en
2
x
(x) = = 1 + 2(x)
n=

satises the functional equation


(x1 ) = x1/2 (x) (6)
for every x > 0. It then follows that

2(x1 ) = (x1 ) 1 = x1/2 (x) 1 = 1 + x1/2 + 2x1/2 (x),

so that for > 1, we have


   
s/21 1 s/21 1 1 1/2
x (x ) dx = x + x + x (x) dx
1/2
1 1 2 2

1
= + xs/21/2 (x) dx. (7)
s(s 1) 1

It follows on combining (5) and (7) that for > 1, we have


s   
s/2 1
(s) = + xs/21 + xs/21/2 (x) dx. (8)
2 s(s 1) 1

Note now that the integral on the right hand side of (8) converges absolutely for any s, and uniformly
in any bounded part of the plane, since (x) = O(ex ) as x +. Hence the integral represents an
entire function of s, and the formula gives the analytic continuation of (s) over the whole plane. Note
also that the right hand side of (8) remains unchanged when s is replaced by 1 s, so that the functional
equation (1) follows immediately. Finally, note that the function

1 s
(s) = s(s 1) s/2 (s)
2 2

is analytic everywhere. Since s(s/2) has no zeros, the only possible pole of (s) is at s = 1, and we
have already shown earlier that (s) has a simple pole at s = 1 with residue 1.
It remains to establish (6) for every x > 0. In other words, we need to prove that for every x > 0,
we have
 
en /x = x1/2 en x .
2 2

n= n=
64 W W L Chen : Elementary and Analytic Number Theory

The starting point is the Poisson summation formula, that under certain conditions on a function f (t),
we have
   B
f (n) = f (t)e2it dt, (9)
AnB = A


where denotes that the terms in the sum corresponding to n = A and n = B are 12 f (A) and 12 f (B)
respectively. Using (9) with A = N , B = N and f (t) = et /x , we have
2


N 
 N

en et
2 2
/x /x 2it
= e dt.
n=N = N

Letting N , we obtain

 

en et
2 2
/x /x 2it
= e dt. (10)
n= =

This is justied by noting that


 

N
t2 /x 2it
et
2
/x
+ e e dt = 2 cos(2t) dt,
N N

and that
 

e t2 /x
cos(2t) dt 0

=0 N

as N . Writing t = xu and using (10), we have


 

en eu
2 2
/x x 2ixu
=x e du
n= =
 
e(ui)
2
x 2 x
=x du
=

 
e e(ui)
2 2
x x
=x du. (11)
=

Note now that the function ez


2
x
is an entire function of the complex variable z. It follows from
Cauchys residue theorem that
 
(ui)2 x
eu du = Ax1/2 ,
2
x
e du = (12)

where 
ey
2

A dy = 1. (13)

Combining (11)(13), we conclude that




en e
2 2
/x
= x1/2 x
.
n= =

This gives (6), and the proof of Theorem 6.1 is now complete.
Chapter 6 : The Riemann Zeta Function 65

6.3. Entire Functions

In this section, we shall prove some technical results on entire functions for use later in the proof of
Theorems 6.2 and 6.4.
An entire function f (s) is said to be of order 1 if
 
f (s) = O e|s| as |s| (14)

holds for every > 1 and fails for every < 1.


Suppose that the entire function h(s) has no zeros on the plane. Then the function g(s) = log h(s)
can be dened as a single valued function and is also entire. Suppose that
 
h(s) = O e|s| as |s| (15)

holds for every > 1. Then

Rg(Rei ) = log |h(Rei )| = O (R ) as R

holds for every > 1. Without loss of generality, we may suppose that g(0) = 0. Then we can write


g(Rei ) = (ak + ibk )Rk eik (ak , bk R),
k=1

so that



Rg(Rei ) = ak Rk cos k bk Rk sin k.
k=1 k=1

Note now that for every k, n N, we have


 2
if k = n
cos k cos n d =
0 0 n
if k =

and  2
sin k cos n d = 0.
0

It follows that  2  
Rg(Rei ) cos n d = an Rn ,
0

so that  2
|an |R
n Rg(Rei ) d = O (R ) as R
0

holds for every > 1. On letting R , we see that an = 0 for every n > 1. A similar argument
using the function sin n instead of the function cos n gives bn = 0 for every n > 1. We have therefore
proved the following result.

THEOREM 6.6. Suppose that the entire function h(s) has no zeros on the plane, and that (15) holds
for every > 1. Then h(s) = eA+Bs , where A and B are constants.

Remark. In the preceding argument, note that it is enough to assume that the estimates for h(s) hold
for a sequence of values R with limit innity.

Our next task is to study the distribution of the zeros of an entire function. The rst step in this
direction is summarized by the result below.
66 W W L Chen : Elementary and Analytic Number Theory

THEOREM 6.7. (Jensens Formula) Suppose that f (s) is an entire function satisfying f (0) = 0.
Suppose further that s1 , . . . , sn are the zeros of f (s) in |s| < R, counted with multiplicities, and that
there are no zeros of f (s) on |s| = R. Then
 2
1 Rn
log |f (Rei )| d log |f (0)| = log . (16)
2 0 |s1 . . . sn |

Proof. We may clearly write


f (s) = (s s1 ) . . . (s sn )k(s),
where k(s) is analytic and has no zeros in |s| R, so that log k(s) is analytic in |s| R. It follows from
Gausss mean value theorem that
 2
1
log k(Rei ) d = log k(0).
2 0
Taking real parts, we obtain
 2
1
log |k(Rei )| d = log |k(0)| = log |f (0)| log |s1 . . . sn |. (17)
2 0
Unfortunately, for every j = 1, . . . , n, we cannot apply a similar argument to log |s sj |, since the
function s sj has a zero at sj . Note, however, that the function

R2 sj s
R
has no zeros in |s| R and satises 2
R sj s
= |s sj |
R

on the circle |s| = R, so that


 2  2 2
1 1 R sj Rei
log |Re sj | d =
i
log d. (18)
2 0 2 0 R

Clearly the function


R2 sj s
log
R
is analytic in |s| R. Applying Gausss mean value theorem over the circle |s| = R on this function and
taking real parts, we conclude that the right hand side of (18) is equal to log R. Finally, note that

n
log |f (Rei )| = log |Rei sj | + log |k(Rei )|,
j=1

so that  2
1
log |f (Rei )| d = n log R + log |f (0)| log |s1 . . . sn |,
2 0
and the result follows.

Remarks. (i) It is important to point out that Jensens formula was in fact only discovered after
Hadamards work in connection with Theorems 6.2 and 6.4.
(ii) Gausss mean value theorem states that the value of an analytic function at the centre of a circle
is equal to the arithmetic mean of its values on the circle. In particular, if the function F (s) is analytic
for |s| < R0 , then for every R < R0 , we have
 2
1
F (0) = F (Rei ) d.
2 0
Chapter 6 : The Riemann Zeta Function 67

A simple consequence of Jensens formula is the following result on the zeros of entire functions.

THEOREM 6.8. Suppose that f (s) is an entire function satisfying f (0) = 0, and that (14) holds for
every > 1. Suppose further that s1 , s2 , s3 , . . . are the zeros of f (s), counted with multiplicities and
where |s1 | |s2 | |s3 | . . . . Then for every > 1, the series


|sn |
n=1

is convergent.

Proof. Note that the right hand side of (16) is equal to


 R
r1 n(r) dr,
0

where, for every non-negative r R, n(r) denotes the number of zeros of f (s) in |s| r. To see this,
note that
 R   |sj+1 |
n1  R
1 1
r n(r) dr = r j dr + r1 n dr
0 j=1 |sj | |sn |


n1
= j(log |sj+1 | log |sj |) + n(log R log |sn |)
j=1
= n log R log |s1 | . . . log |sn |.

For every > 1, write = ( + 1)/2, so that 1 < < . Then



log |f (Rei )| = O (R ) as R ,

so that by Jensens formula, we have


 R

r1 n(r) dr = O (R ) log |f (0)| = O (R ) as R .
0

On the other hand, note that


 2R  2R
1
r n(r) dr n(R) r1 dr = n(R) log 2.
R R

It follows that
n(R) = O (R ) as R .
Now  



|sn | = r dn(r) = r1 n(r) dr <
n=1 0 0

as required.

Suppose now that f (s) is an entire function satisfying f (0) = 0, and that (14) holds for every
> 1. Suppose further that s1 , s2 , s3 , . . . are the zeros of f (s), counted with multiplicities and where
|s1 | |s2 | |s3 | . . . . Then for every ) > 0, the series


|sn |1
n=1
68 W W L Chen : Elementary and Analytic Number Theory

converges, so that the series




|sn |2
n=1

converges, and so the product



 
s
P (s) = 1 es/sn (19)
n=1
sn

converges absolutely for every s C, and uniformly in any bounded domain not containing any zeros of
f (s). It follows that P (s) is an entire function, with zeros at s1 , s2 , s3 , . . . . Now write

f (s) = P (s)h(s), (20)

where h(s) is an entire function without zeros. If (15) holds for every > 1, then h(s) = eA+Bs , where
A and B are constants, and so

 
s
f (s) = e A+Bs
1 es/sn . (21)
n=1
sn

THEOREM 6.9. Under the hypotheses of Theorem 6.8, the inequality (15) holds for every > 1,
where the function h(s) is dened by (19) and (20). In particular, the function f (s) can be expressed in
the form (21), where A and B are constants.

Proof. To show that the inequality (15) holds for every > 1, it clearly suces, in view of (14) and
(20), to establish a suitable lower bound for |P (s)|. Since the series



|sn |2
n=1

is convergent, the set




S= (|sn | |sn |2 , |sn | + |sn |2 )
n=1

has nite total length. It follows that there exist arbitrarily large positive values R such that R S, so
that
|R |sn || |sn |2 for every n N. (22)
For any such R, write, for j = 1, 2, 3,

  s

Pj (s) = 1 es/sn ,
sn
(23.j)

where the products are taken over all n N satisfying

R
|sn | < , (23.1)
2

R
|sn | < 2R, (23.2)
2
|sn | 2R, (23.3)
respectively. Clearly
P (s) = P1 (s)P2 (s)P3 (s). (24)
Chapter 6 : The Riemann Zeta Function 69

Let ) > 0 be chosen and xed. Suppose rst of all that (23.1) holds. Then on |s| = R, we have
   

1 s es/sn s 1 e|s|/|sn | > eR/|sn | ,
sn sn

and so it follows from   



R
|sn |1 < |sn |1
2 n=1
(23.1)

that
|P1 (s)|  eR
1+2
as R . (25)

Suppose next that (23.2) holds. Then on |s| = R, we have


 

1 s es/sn sn s e|s|/|sn | > ||sn | R| e2 R3 ,
sn sn 2R

in view of (22). Note that there are at most O (R1+ ) values of n for which (23.2) holds. Hence on
|s| = R, we have
|P2 (s)|  (R3 )R  eR
1+ 1+2
as R . (26)

Suppose nally that (23.3) holds. Then on |s| = R, we have


 

1 s es/sn > ec(R/|sn |)2 (27)
sn

for some positive constant c (see the Remark below), and so it follows from



2 1+
|sn | (2R) |sn |1
(23.3) n=1

that
|P3 (s)|  eR
1+2
as R . (28)

It now follows from (24)(26) and (28) that on |s| = R, we have

|P (s)|  eR
1+3
as R . (29)

The result now follows on combining (20) and (29), and noting that the inequality (14) holds for = 1+).

Remark. Note that the inequality (27) is of the form

|(1 z)ez | > ec|z| ,


2
(30)

where |z| 1/2. Write z = x + iy, where x, y R. Then (30) will follow if we show that

(1 x)2 e2x > e2cx


2

whenever |x| 1/2. This last inequality can easily be established by using the theory of real valued
functions of a real variable.

Finally, we make the following simple observation.


610 W W L Chen : Elementary and Analytic Number Theory

THEOREM 6.10. Under the hypotheses of Theorem 6.8, suppose further that the series


|sn |1
n=1

is convergent. Then there exists a positive constant c such that

f (s) = O(ec|s| ) as |s|

holds.

Proof. This follows from (21) and the inequality |(1 z)ez | e2|z| which holds for every z C.

6.4. Proof of Theorems 6.2 and 6.4

Recall that (s), dened by (3), is an entire function, and that (0) = 0. Note also that the zeros of
(s) are precisely the zeros of (s) in the critical strip. In order to establish Theorem 6.4, we shall use
Theorem 6.9. We therefore rst need to show that
 
(s) = O e|s| as |s|

holds for every > 1.


We shall in fact prove the following stronger result.

THEOREM 6.11. There exists a positive constant c such that

|(s)| < ec|s| log |s| as |s| . (31)

Furthermore, the inequality


|(s)| < ec|s| as |s| (32)
does not hold for any positive constant c.

Proof. Since (s) = (1 s) for every s C, it suces to prove the inequality (31) for 1/2. First
of all, there exists a positive constant c1 such that

1
s(s 1) s/2 < ec1 |s| .
2

Next, Stirlings formula


s  
s 1 s s 1
log = log + log 2 + O(|s|1 )
2 2 2 2 2 2

as |s| is valid in the angle /2 < arg s < /2, and so there exists a positive constant c2 such that
 s 

< ec2 |s| log |s| .
2
Finally, note that the formula 
s
(s) = s {x}xs1 dx
s1 1

is valid for > 0, and the integral is bounded for 1/2, so that there exists a positive constant c3
such that
|(s)| < c3 |s|.
Chapter 6 : The Riemann Zeta Function 611

This proves (31). On the other hand, note that as s + through real values, we have
s s s
log log and (s) 1,
2 2 2
so that (32) does not hold.

To complete the proof of Theorems 6.2 and 6.4, note that by Theorem 6.10, the series

||1

is divergent, where denotes the zeros of (s) and so the zeros of (s) in the critical strip. Theorem 6.2
follows immediately. Theorem 6.4 now follows from Theorems 6.9 and 6.11.

6.5. An Important Formula

It follows from (4) that


s 
s

log (s) = A + Bs + + log 1 .

Dierentiating with respect to s, we obtain


  
 (s) 1 1
=B+ + . (33)
(s)
s

On the other hand, it follows from (3) that


s  s  s 
log (s) = log(s 1) log + log + log (s)
2 2 2 
s s
= log(s 1) log + log + 1 + log (s).
2 2
Dierentiating with respect to s, we obtain

 (s) 1 1 1  ( 2s + 1)  (s)
= log + + . (34)
(s) s1 2 2 ( 2s + 1) (s)

Combining (33) and (34), we obtain the following result.

THEOREM 6.12. We have


 
 (s) 1 1 1  ( 2s + 1)  1 1
=B + log + + , (35)
(s) s1 2 2 ( 2s + 1)
s

where B is a constant and where denotes the zeros of (s) in the critical strip.

The formula (35) exhibits the pole of (s) at s = 1 and the zeros in the critical strip. The trivial
zeros are exhibited by the term
1  ( 2s + 1)
.
2 ( 2s + 1)
To see this last point, we start from the Weierstrass formula

1  s  s/n
= es 1+ e ,
s(s) n=1
n
612 W W L Chen : Elementary and Analytic Number Theory

where is Eulers constant. This gives



1 s/2
 s  s/2n
= e 1 + e .
( 2s + 1) n=1
2n

Taking logarithms, we obtain


s  1    s  s 
log + 1 = s + log 1 + .
2 2 n=1
2n 2n

Dierentiating with respect to s, we obtain

  
1  ( 2s + 1) 1 1 1
= + .
2 ( 2s + 1) 2 n=1
s + 2n 2n

6.6. Counting Zeros in the Critical Strip

The starting point of our discussion is based on the Argument principle. Suppose that the function F (s)
is analytic, apart from a nite number of poles, in the closure of a domain D bounded by a simple closed
positively oriented Jordan curve C. Suppose further that F (s) has no zeros or poles on C. Then

1 F  (s) 1
ds = C arg F (s)
2i C F (s) 2

represents the total number of zeros of F (s) in D minus the total number of poles of F (s) in D, counted
with multiplicities. Here C arg F (s) denotes the change of argument of the function F (s) along C.
It is convenient to use the function (s), since it is entire and its zeros are precisely the zeros of (s)
in the critical strip. To calculate N (T ), it is convenient to take the domain (1, 2) (0, T ), so that C
is the rectangular path passing through the vertices

2, 2 + iT, 1 + iT, 1

in the anticlockwise direction. If no zeros of (s) has imaginary part T , then



1  (s) 1
N (T ) = ds = C arg (s).
2i C (s) 2

Let us now divide C into the following parts. First, let L1 denote the line segment from 1 to 2.
Next, let L2 denote the line segment from 2 to 2 + iT , followed by the line segment from 2 + iT to 12 + iT .
Finally, let L3 denote the line segment from 12 + iT to 1 + iT , followed by the line segment from 1 + iT
to 1.
Since (s) is real on L1 , clearly L1 arg (s) = 0. On the other hand,

( + it) = (1 it) = (1 + it),

so that L2 arg (s) = L3 arg (s). If we write L = L2 , so that L denotes the line segment from 2 to
2 + iT , followed by the line segment from 2 + iT to 12 + iT , then

N (T ) = L arg (s). (36)

Recall that s 
(s) = (s 1) s/2 + 1 (s).
2
Chapter 6 : The Riemann Zeta Function 613

It follows that
s 
L arg (s) = L arg(s 1) + L arg s/2 + L arg + 1 + L arg (s). (37)
2
Clearly  
1 1
L arg(s 1) = arg + iT = + O(T 1 ) (38)
2 2
and  
s/2 1 1
L arg = L t log = T log . (39)
2 2
On the other hand,  
s  5 1
L arg + 1 = I log + iT .
2 4 2
By Stirlings formula,
     
5 1 3 1 5 1 5 1 1
log + iT = + iT log + iT iT + log + O(T 1 ),
4 2 4 2 4 2 4 2 2

so that s  T T 3 T
L arg + 1 = log + + O(T 1 ). (40)
2 2 2 8 2
Combining (36)(40), we have

1 T T T 3 T
N (T ) = log + log + + S(T ) + O(T 1 )
2 2 2 2 8 2
T T T 7
= log + + S(T ) + O(T 1 ),
2 2 2 8
where
S(T ) = L arg (s).
To prove Theorem 6.3, it suces to prove the following result.

THEOREM 6.13. We have S(T ) = O(log T ) as T .

Proof. Note rst of all that arg (2) = 0. On the other hand,
 
1 I(s)
arg (s) = tan
R(s)

and R(s) = 0 on the line = 2. It follows that



| arg (2 + iT )| < .
2

Suppose now that R(s) vanishes q times on the line segment from 2 + iT to 12 + iT . Then this line
segment can be divided into q + 1 parts, where in each subinterval, R(s) may vanish only at one or
both of the endpoints and has constant sign strictly in between, so that the variation of arg (s) in each
such subinterval does not exceed . It follows that for s = + iT , we have
 
1 1
| arg (s)| (q + 1) + 2 . (41)
2 2

To prove our result, it remains to nd a suitable bound for q. For s = + iT , we have

1
R(s) = (( + iT ) + ( iT )).
2
614 W W L Chen : Elementary and Analytic Number Theory

Let T be xed, and consider the function

1
fT (s) = ((s + iT ) + (s iT ))
2

(note that we no longer insist that s = + iT ). Then q is the number of zeros of fT (s) on the line
segment from 1/2 to 2, and so is bounded above by the number of zeros of fT (s) in the disc |s 2| 3/2.
In other words,  
3
qn , (42)
2
where, for every r 0, n(r) denotes the number of zeros of fT (s) in the disc |s 2| r. By Jensens
formula and noting that we may assume that ( 12 + iT ) = 0, we have
   
7/4
n(r) 1 2 7 i
dr =
log fT 2 + e d log |fT (2)|. (43)
0 r 2 0 4

On the other hand,


 7/4  7/4    7/4  
n(r) n(r) 3 1 3 7
dr dr n dr = n log . (44)
0 r 3/2 r 2 3/2 r 2 6

Observe that

1

|fT (2)| = ((2 + iT ) + (2 iT )) = |R(2 + iT )|
2


 
2
1 1
= R 1 = 2 > 0,
n2+iT n2 6
n=1 n=2

so that
log |fT (2)| = O(1). (45)
Finally, recall that |(s)|  T 3/4
for every 1/4. It follows that for every [0, 2], we have
       

fT 2 + 7 ei 1 2 + 7 ei + iT + 2 + 7 ei iT  T 3/4 ,
4 2 4 4

so that  
7 i

log fT 2 + e  log T. (46)
4
Combining (43)(46), we conclude that
 
3
n  log T. (47)
2

The result now follows on combining (41), (42) and (47).

6.7. Sketch of the Proof of Theorem 6.5

It can be shown that for c > 1, we have


 
c+i 0 if 0 < y < 1,
1 ys
ds = 1/2 if y = 1,
2i ci s
1 if y > 1.
Chapter 6 : The Riemann Zeta Function 615

It follows that


 c+i

  1 (X/n)s
0 (X) = (n) = (n) ds
n=1
2i ci s
nX
 c+i 


 c+i 
1 (n) X s 1 (s) X s
= ds = ds,
2i ci n=1 ns s 2i ci (s) s


where denotes that the term in the sum corresponding to n = X is 12 (X).
The idea is now to move the line of integration away to innity on the left, so that we can express
0 (X) as a sum of the residues at the poles of the function

 (s) X s
.
(s) s

Here, the pole at s = 1 gives rise to a residue X, while the pole at s = 0 gives rise to a residue
 (0)/(0). On the other hand, each zero of (s) in the critical strip gives rise to a residue X /,
while each trivial zero 2n of (s) gives rise to a residue X 2n /2n. Finally, note that
  
1 1 X 2n
log 1 2 = .
2 X n=1
2n

The rst step in this process is to consider the integral



1 c+iT
 (s) X s
ds,
2i ciT (s) s

and regard the path of integration as one side of a rectangle with vertices at c iT and U iT , where
U > 0 is large. Here T has to be chosen carefully so that the horizontal sides of the rectangle should
avoid the zeros of (s) in the critical strip. Also U has to be chosen carefully so that the left vertical
side of the rectangle should avoid the trivial zeros of (s). On taking U , this will result in a nite
version of Theorem 6.5, of the form
 X  
 (0) 1 1
0 (X) = X log 1 2 + R(X, T ),
(0) 2 X
||<T

where the error term R(X, T ) can be estimated.

S-ar putea să vă placă și