You are on page 1of 22

The Riemann Zeta Function

David Jekel

June 6, 2013

In 1859, Bernhard Riemann published an eight-page paper, in which he


estimated the number of prime numbers less than a given magnitude using
a certain meromorphic function on C. But Riemann did not fully explain
his proofs; it took decades for mathematicians to verify his results, and to
this day we have not proved some of his estimates on the roots of . Even
Riemann did not prove that all the zeros of lie on the line Re(z) = 21 . This
conjecture is called the Riemann hypothesis and is considered by many the
greatest unsolved problem in mathematics.
H. M. Edwards book Riemanns Zeta Function [1] explains the histor-
ical context of Riemanns paper, Riemanns methods and results, and the
subsequent work that has been done to verify and extend Riemanns theory.
The first chapter gives historical background and explains each section of
Riemanns paper. The rest of the book traces later historical developments
and justifies Riemanns statements.
This paper will summarize the first three chapters of Edwards. My
paper can serve as an introduction to Riemanns zeta function, with proofs
of some of the main formulae, for advanced undergraduates familiar with
the rudiments of complex analysis. I use the term summarize loosely;
in some sections my discussion will actually include more explanation and
justification, while in others I will only give the main points. The paper
will focus on Riemanns definition of , the functional equation, and the
relationship between and primes, culminating in a thorough discussion of
von Mangoldts formula.

1
Contents
1 Preliminaries 3

2 Definition of the Zeta Function 3


2.1 Motivation: The Dirichlet Series . . . . . . . . . . . . . . . . 4
2.2 Integral Formula . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.3 Definition of . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

3 The Functional Equation 6


3.1 First Proof . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
3.2 Second Proof . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

4 and its Product Expansion 9


4.1 The Product Expansion . . . . . . . . . . . . . . . . . . . . . 9
4.2 Proof by Hadamards Theorem . . . . . . . . . . . . . . . . . 10

5 Zeta and Primes: Eulers Product Formula 12

6 Riemanns Main Formula: Summary 13

7 Von Mangoldts Formula 15


7.1 First Evaluation of the Integral . . . . . . . . . . . . . . . . . 15
7.2 Second Evaluation of the Integral . . . . . . . . . . . . . . . . 17
7.3 Termwise Evaluation over . . . . . . . . . . . . . . . . . . . 19
7.4 Von Mangoldts and Riemanns Formulae . . . . . . . . . . . 22

2
1 Preliminaries
Before we get to the zeta function itself, I will state, without proof, some
results which are important for the later discussion. I collect them here so
subsequent proofs will have less clutter. The impatient reader may skip this
section and refer back to it as needed.
In order to define the zeta function, we need the gamma function, which
extends the factorial function to a meromorphic function on C. Like Edwards
and Riemann, I will not use the now standard notation (s) where (n) =
(n 1)!, but instead I will call the function (s) and (n), will be n!. I give
a definition and some identities of , as listed in Edwards section 1.3.

Definition 1. For all s C except negative integers,


N
Y n
(s) = lim (N + 1)s .
N s+n
n=1
R
Theorem 2. For Re(s) > 1, (s) = 0 ex xs dx.

Theorem 3. satisfies

1. (s) = 1 s
Q
n=1 (1 + s/n) (1 + 1/n)

2. (s) = s(s 1)

3. (s)(s) sin s = s

4. (s) = 2 (s/2)(s/2 1/2).

Since we will often be interchanging summation and integration, the


following theorem on absolute convergence is useful. It is a consequence of
the dominated convergence theorem.

Theorem 4. Suppose P fn is nonnegative


RP and integrable on compact
P R sub-
sets
RP of [0, ). If fn converges and 0 fn converges, then 0 fn =
0 fn .

2 Definition of the Zeta Function


Here I summarize Edwards section 1.4 and give additional explanation and
justification.

3
2.1 Motivation: The Dirichlet Series
Dirichlet defined (s) = s for Re(s) > 1. Riemann wanted a def-
P
n=1 n
inition valid for all s C, which would be equivalent to Dirichlets for
Re(s) > 1. He found a new formula for the Dirichlet series as follows. For
Re(s) > 1, by Eulers integral formula for (s) 2,
Z Z
nx s1 1 (s 1)
e x dx = s ex xs1 dx = .
0 n 0 ns

Summing over n and applying the formula for a geometric series gives
Z Z x s1 Z s1
X 1 X
nx s1 e x x
(s 1) s
= e x dx = x
dx = dx.
n 0 0 1e 0 ex 1
n=1 n=1

We are justified in interchanging summation and integration by Theorem 4


+
R a s2 As x 0 ,
because the integral on the right converges at both endpoints.
x
e 1 behaves like x, so that the integral behaves like 0 x dx which con-
verges for Re(s) > 1. The integral converges at the right endpoint because
ex grows faster than any power of x.

2.2 Integral Formula


To extend this formula to C, Riemann integrates (z)s /(ez 1) over a the
path of integration C which starts at +, moves to the origin along the
top of the positive real axis, circles the origin counterclockwise in a small
circle, then returns to + along the bottom of the positive real axis.
That is, for small positive ,
Z Z + !
(z)s dz (z)s dz
Z Z
z
= + + .
C e 1 z + |z|= ez 1 z

Notice that the definition of (z)s implicitly depends on the definition


of log(z) = log |z| + i arg(z). In the first integral, when z lies on
the negative real axis, we take arg(z) = i. In the second integral,
the path of integration starts at z = or z = , and as z proceeds
counterclockwise around the circle, arg(z) increases from i to i. (You
can think of the imaginary part of the log function as spiral staircase, and
going counterclockwise around the origin as having brought us up one level.)
In the last integral, arg(z) = i. (Thus, the first and last integrals do not

4
cancel as we would expect!) To state the definition quite precisely, the
integral over C is
2 +
es(log zi) dz es(log i+i)es(log z+i) dz
Z Z Z
+ i d + .
+ ez 1 z 0 eei 1
ez 1 z
R
The integrals converge by the same argument given above regarding 0 xs1 /(1
ex ) dx; in fact, they converge uniformly on compact subsets of C.
This definition appears to depend on , but actually does not matter
(so long as we avoid muitiples of 2i). Edwards does not mention this point,
but I will give a short proof for (0, 1). Let A be a curve which moves
in a line from 1 to , then in a semicircle above the real axis from to ,
then in a line from to 1. Let B move in a line from 1 to , then
in a semicircle below the real axis from to , then in a line from to 1.
Then, because the integrals from to 1 and 1 to cancel,
Z 1 Z + 
(z)s dz
Z Z
(s) = + + +
+ A B 1 ez 1 z

The first and last integrals obviously do not depend on . In the integral
over A , (z)s is defined by the branch of the log function whose slit is on
the nonnegative imaginary axis, and in the integral over B , (z)s is defined
by the branch of the log function whose slit is on the nonpositive imaginary
axis. Since each branch of the log function is analytic on its domain, the
value of an integral depends only on the endpoints. Hence, the integrals
over A and B do not depend on .

2.3 Definition of
For Re(s) > 0, we can relate the formula for C (z)s /z(ez 1) dz to our
R

previous formula for the Dirichlet series by taking 0+ . Now


2
es(log i+i)
Z
lim i d = 0
0+ 0 eei 1
because

es(log i+i) Re(s) eIm(s)()
= Re(s)1 eIm(s) ,

eei 1 ei
e 1

|e 1|

5
but /(e 1) 1 and Re(s)1 0. Hence,
!
Z + s(log z+i)
(z)s dz es(log zi) dz
Z Z
e dz
= lim +
C ez 1 z 0+ + ez 1 z ez 1 z
Z s1
z
= (eis eis ) dz.
0 ez 1

But this integral is exactly (s 1) s


P
n=1 n . Therefore, for Re(s) > 1,


(z)s dz
Z
X 1 1
=
ns 2i(sin s)(s 1) C ez 1 z
n=1

By Theorem 3, 1/2i(sin s)(s 1) = (s)/2i. Therefore, Riemann


gives the following definition of the zeta function, which coincides with the
Dirichlet series for Re(s) > 1:
(z)s dz
Z
(s)
Definition 5. (s) = .
2i C ez 1 z

X 1
Theorem 6. For Re(s) > 1, (s) = .
ns
n=1

The integral over C converges uniformly on compact subsets of C because


ez grows faster than any power of z. It follows that the integral defines an
analytic function. Thus, is analytic except possibly at the positive integers
where (s) has poles. Because the Dirichlet series converges uniformly on
Re(s) > 1+ for any positive , we see is actually analytic for s = 2, 3, 4 . . . ,
and it is not too hard to check that the integral has a zero there which cancels
the pole of . Thus, is analytic everywhere except 1 where it has a simple
pole.

3 The Functional Equation


Riemann derives a functional equation for , which states that a certain
expression is unchanged by the substitution of 1 s for s.
 
s 
s/2 1s
Theorem 7. 1 (s) = 1 1/2+s/2 (1 s).
2 2
He gives two proofs of this functional equation, both of which Edwards
discusses (sections 1.6 and 1.7). I will spend more time on the first.

6
3.1 First Proof
Fix s < 1. Let DN = {z : < |z| < (2N + 1)} \ (, (2N + 1)). We
consider Dn split along the positive real axis. Let CN be the same contour
as C above except that the starting and ending point is (2N + 1) rather
than +. Then
(z)s dz (z)s dz (z)s dz
Z Z Z
z
= z
z
.
DN e 1 z |z|=(2N +1) e 1 z CN e 1 z

We make similar stipulations as before regarding the definition of log(z);


that is, arg(z) is taken to be i on the top of the positive real axis and
i on the bottom. Note that the integrand is analytic except at poles
at 2in. If log z were analytic everywhere, the residue theorem would give

(z)s dz (z)s
Z X  
z
= 2i Res ,z .
DN e 1 z z(ez 1)
zDn

In fact, we can argue that this is still the case by spliting Dn into two pieces
and considering an analytic branch of the log function on each piece. This
argument is similar to the one made in the last section and is left to the
reader.
The residue at each point 2in for integer n 6= 0 is computed by inte-
grating over a circle of radius r < 2:

(z)s dz z 2in
Z Z
1 1 dz
z
= (z)s1 z .
2i |z2in|=r e 1 z 2i |z2in|=r e 1 z 2in

Since (z)s1 is analytic at z = 2in and (z 2in)/(ez 1) has a remov-


able singularity (its limit is 1), we can apply Cauchys integral formula to
conclude that the residue is (2in)s1 . Hence,
N
(z)s dz (z)s dz
Z Z X
s1 s1

= +2i (2in) + (2in) .
CN ez 1 z |z|=(2N +1) ez 1 z
n=1

As we take N , the left side approaches the integral over C. I claim


the integral on the right approaches zero. Notice 1/(1 ez ) is bounded on
|z| = (2N + 1). For if |Re(z)| > /2, then |ez 1| 1 e/2 by the
reverse triangle inequality. If |Im(z) 2in| > /2 for all integers n, then
|ez 1| > 1 because Im(ez ) < 0. For |z| = (2N + 1), at least one of these
two conditions holds. The length of the path of integration is 2(2N + 1)

7
and |(z)s /z| (2N + 1)s1 , so the whole integral is less than a constant
times (2N + 1)s , which goes to zero as N because we assumed s < 0.
Therefore, taking N gives

(z)s dz
Z X
s1 s1 s1
z
= 2i(2) (i + (i) ) ns1 .
C e 1 z n=1

From our definition of the log function,


1 1 s
is1 + (i)s1 = (is (i)s ) = (eis/2 eis/2 ) = 2 sin .
i i 2
Substituting this into the previous equation and multiplying by (s)/2i
gives  s 
(s) = (s)(2)s1 2 sin (1 s), s < 0.
2
Applying various identities of from Theorem 3 will put this equation into
the form given in the statement of the theorem. Since the equation holds
for s < 0 and the functions involved are all analytic on C except for some
poles at certain integers, the equation holds on all of C \ Z.

3.2 Second Proof


The second proof I will merely summarize. Riemann applies a change of
variables in the integral for to obtain
1 s/2  s  Z 2
s
1 = en x xs/21 dx, Re(s) > 1.
n 2 0

Summing this over n gives


s  Z
s/2 1 (s) = (x)xs/21 dx,
2 0
P n2 x
where (x) = n=1 e . (We have interchanged summation and inte-
gration here; this is justified by applying
R Theorem 4 to the absolute value
of the integrand and showing that 0 (x)xp dx converges for all real p.) It
turns out satisfies
1 + 2(x) 1
=
1 + 2(1/x) x
as Taylor section 9.2 or Edwards section 10.6 show. Using this identity,
changing variables, and integrating by parts, Riemann obtains
s Z

s/2 dx 1
1 (s) = (x)(xs/2 + x1/2s/2 ) + .
2 1 x s(1 s)

8
Both sides can be shown to be analytic, so this relation holds not only for
Re(s) > 1, but for all s. The right side is unchanged by the substitution of
1 s for s, so the theorem is proved again.

4 and its Product Expansion


As Edwards explains in sections 1.8 and 1.9, Riemann defined the function
by
s
Definition 8. (s) = (s 1) s/2 (s).
2
The functional equation derived in the last section says exactly that
(s) = (1 s). We know that is analytic except perhaps at certain
integers. But by examining both sides of the functional equation at potential
singular points, we can deduce that, in fact, is analytic everywhere. Eulers
product formula, which we will prove later, shows that has no zeroes for
Re(s) > 1. This implies has no zeroes there either and (s) = (1 s)
implies all the roots of lie in {s : 0 Re(s) 1}. We can also prove that
has no zeroes on the real line.

4.1 The Product Expansion


Riemann wanted to write in the form
Y s

(s) = (0) 1 ,

where ranges over the roots of . This type of product expansion is guar-
anteed for polynomials by the fundamental theorem of algebra, but not for
functions with possibly infinitely many roots. The convergence of the prod-
uct depends on the ordering of the terms and the frequency of the roots.
By the argument principle, {0 Re(s) 1, 0 Im(s) T } is equal to the
integral of 0 /2i around the boundary of the region (assuming there are
no roots on the boundary). Riemann estimates this integral as
 
T T
log 1
2 2
with an error on the order of T 1 . This estimate on the frequency of the roots
of would guarantee that the product expansion of converges, provided
each root is paired with its symmetric twin 1 . But no one else could
prove this estimate until 1905.

9
4.2 Proof by Hadamards Theorem
Hadamard proved that the product representation of was valid in 1893.
His methods gave rise to a more general result, the Hadamard factorization
theorem. Edwards devotes chapter 2 to Hadamards proof of the product
formula for . But for the sake of space I will simply quote Hadamards
theorem to justify that formula. The reader may consult Taylor chapter
8 for a succinct development of the general theorem, which I merely state
here, as given in [2]:

Theorem 9 (Hadamards Factorization Theorem). Suppose f is entire. Let


p be an integer and suppose there exists positive t < p + 1 such that |f (z)|
exp(|z|t ) for z sufficiently large. Then f can be factored as
 
m h(z)
Y z
f (z) = z e Ep ,
zk
k=1

where {zk } is a list of the roots of f counting multiplicity, m is the order of


the zero of fPat 0, h(z) is a polynomial of degree at most p, and Ep (w) =
(1 w) exp( pn=1 wn /n).

To apply the theorem, we must estimate . To do this we need Riemanns


alternate representation of based on the result of the previous section

1 s(1 s)
Z
dx
(s) = (x)(xs/2 + x1/2s/2 ) ,
2 2 1 x
which Edwards derives in section 1.8. Using more integration by parts and
change of variables, we can show that
Z
d h 3/2 0 i 1/4
cosh 12 (s 21 ) log x dx.

(s) = 4 x (x) x
1 dx

Substituting the power series for cosh w = n


P
n=0 w /(2n)! will show that
1
the power series for (s) about s = 2 has coefficients
Z
4 d h 3/2 0 i 1/4 1
a2n = x (x) x ( 2 log x)2n dx.
(2n)! 1 dx

Direct evaluation of (d/dx)[x3/2 0 (x)] will show that the integrand is posi-
tive. Hence, has a power series about 12 with real, positive coefficients.
This is all we need to derive a simple estimate of as Edwards does
in section 2.3:

10
Theorem 10. For sufficiently large R, |( 12 + w)| RR for |w| < R.
Proof. Since the power series of around 12 has positive coefficients, |( 12 +
w)| achieves its maximum value for |w| R at R itself. Choose N such that
1 1
2 +R 2N 2 +R+2. Since is entire, the power series expansion is valid
for all s, and the fact that is has positive coefficients shows is increasing
on [ 21 , ). Thus,

( 12 + R) (2N ) = N ! N (2N 1)(2N ).

Because (s) = s for Re(s) > 1, we know is decreasing on [2, ).


P
n=1 n
Hence, (2N ) is bounded by a constant, and so is N . Therefore,

(2N ) kN !(2N 1) 2kN N +1 2k( 12 R + 54 )R/2+7/4 RR

for sufficiently large R.

This implies as a corollary that |( 12 + w)| exp(|w|3/2 ) for sufficiently


3/2
large w. This is because RR = eR log R eR for R sufficiently large. I
will now follow Taylors proof [2]. We have shown that ( 12 + w) satisfies
the hypotheses of Hadamards theorem with p = 1, and therefore it has a
product expansion of the form
Y w Y w  w/
( 12 + w) = eh(w) E1 = eh(w) 1 e ,


where ranges over the roots of ( 12 + w) and where h is a polynomial of


degree at most 1. (Since ( 21 ) 6= 0, we do not have any wm term.) If we
group the roots and together, the exponential factors cancel, so we
have Y w
( 21 + w) = eh(w) 1 .

Since the left hand side and the product are both even functions, eh(w) must
be even, which implies it is constant. We undo the change of variables by
letting 21 + w = s and 12 + = . Thus, we have
Y s 1/2

(s) = c 1 .

1/2

Divide by (0) (as Edwards section 2.8):


  1
(s) Y s 1/2 1/2
= 1 1+ .
(0)
1/2 1/2

11
Each term in the product is a linear function of s which is 0 when s = and
1 when s = 0. Hence, each term is 1 s/ and we have
Theorem 11. has a product expansion
Y s

(s) = (0) 1 ,

where ranges over the roots of , and and 1 are paired.

5 Zeta and Primes: Eulers Product Formula


The zeta function is connected to prime numbers by
Theorem 12. For Re(s) > 1,
Y
(s) = (1 ps )1 ,
p

where p ranges over the prime numbers.


This identity is derived from the Dirichlet series in the following way.
Edwards does not give the proof, but for the sake of completeness I include
a proof adapted from Taylor [2].
Proof. Let pn be the nth prime number. Define sets SN and TN inductively
as follows. Let S0 be the positive integers. For N 0, let TN +1 be the set of
numbers in SN divisible by PN and let SN +1 = SN \ TN +1 . Notice that SN
is the set of positive integers not divisible by the first N primes, and TN +1
exactly {pN +1 m : m SN }. I will show by induction that for Re(s) > 1,
N  
Y 1 X 1
(s) 1 s = .
pn ns
n=1 nSN

Since S0 = Z+ , we have (s) = nS0 ns . Now suppose the statement


P

holds for N . Multiply by (1 psN +1 ):


N +1     X
Y 1 1 1
(s) 1 s = 1 s
pn pN +1 ns
n=1 nSN
X 1 X 1
= s

n (PN +1 n)s
nSN nSN
X 1 X 1 X 1
= = ,
ns ns ns
nSN nTN +1 nSN +1

12
where the last equality holds because SN \ TN +1 = SN +1 and TN +1 SN .
This completes the induction.
The proof will be complete if we show that nSN ns approaches 1 as
P
N . Notice that if 1 < n N < PN , then n 6 SN . This is because any
prime factor of n has to be less than or equal to PN , but all multiples of all
primes pn PN have been removed from SN . Hence,


X 1 X 1 X 1 X 1
1 = ns ns .

ns ns


nSN N nSN N nSN n=N

We are justified in applying the triangle inequality here because the right-
most sum converges absolutely. Indeed, the sum on the right isPsmaller than
any positive  for N sufficiently large. Hence, the limit of nSN ns is
1.
s
P
Taking the log of this formula gives
P n log (s) = p log(1 p ) for
Re(s) > 1. Since log(1 x) = n=1 x /n, we have

XX 1 sn
log (s) = p .
p
n
n=1

There is a problem in that log z is only defined up to multiples of 2i.


However, it is easy to make the statement rigorous. We simply say that the
right hand side provides one logarithm for . We can also argue that it is
analytic.

6 Riemanns Main Formula: Summary


Riemann uses the formula for log (s) to evaluate J(x), a function which
measures primes. Since I am going to prove Von Mangoldts similar formula
later, I will only sketch Riemanns methods here. J(x) is a step function
which starts at J(0) = 0 and jumps up at positive integers. The jump is
1 at primes, 12 at squares of primes, 31 at cubes of primes, etc. J can be
written as

XX 1
J(x) = u(x pn ),
p
n
n=1
1
where u(x) = 0 for x < 0, u(0) = and u(x) = 1 for x > 0. We can obtain
2,
a formula for (x), the number of primes less than x from J(x).

13
J(x) x and J(x) = 0 for x < 2, we know that for Re(s) > 1,
R Since s1
0 J(x)x dx converges. Hence, by Theorem 4, we can integrate term
by term and
Z
1 s1
Z
s1
XX 1 X X 1 sn
J(x)x dx = x dx = p ,
0 p
n pn s p n
n=1 n=1

which is s1 log (s).


Through a change of variables and Fourier inversion, Riemann recovers
J(x) from the integral formula:
Z a+i
1 ds
J(x) = (log (s))xs ,
2i ai s
where a > 1 and the integral is defined as the Cauchy principal value
Z a+iT
ds
lim (log (s))xs ,
T + aiT s
He then integrates by parts and shows the first term goes to zero to obtain:
Z a+i  
1 d log (s) s
J(x) = x ds.
2i ai ds s

Applying (s) = (s/2)(s 1) s/2 (s) together with the product for-
mula for gives:
s s
log (s) = log (s) + log log(s 1) log
2   2
X s s s
= log (0) + log 1 + log log(s 1) log .

2 2

He substitutes this into the integral for J(x), then evaluates the integral
term by term as
Z
X dt
J(x) = Li(x) [Li(x ) + Li(x1 )] + 2 1) log t
+ log (0),
x t(t
Im>0
Rx
where Li(x) is the Cauchy principal value of 0 1/ log x dx. The evaluation
requires too much work to explain it completely here (Edwards sections
1.13-1.16), but the term-by-term integration itself is even more difficult to
justify, and Riemanns main formula was not proved until Von Mangoldts
work in 1905.

14
7 Von Mangoldts Formula
Von Mangoldts formula is essentially the same as Riemanns, except that
instead of considering log (s), he considers its derivative 0 (s)/(s). By
differentiating the formula for log (s) as a sum over primes, we have

0 (s) XX
= pns log p, Re(s) > 1.
(s) p n=1

Instead of J(x), we use a different step function (x) defined by



XX
(x) = (log p)u(x pn ).
p n=1

Van Mangoldt shows that



X x X x2n 0 (0)
(x) = x + ,

2n (0)
n=1

by showing that both sides are equal to


Z a+i 0
1 (s) s ds
x .
2i ai (s) s

(In the sum over , we pair with 1 and sum in order of increasing Im().
In the integral from a i to a + i we take the Cauchy principal value.)
His proof, like Riemanns, depends on several tricky termwise integrations.

7.1 First Evaluation of the Integral


First, we show that the integral is equal to (x). We let (m) be the size
of the jump of at m; it is log p if m = pn and zero otherwise. We then
rewrite the series for 0 / and as

0 (s) X X
= (m)ms , (x) = (m)u(x m).
(s)
m=1 m=1

The rearrangement is justified because the series converge absolutely. We


then have

Z a+i 0 Z a+i X !
1 (s) s ds 1 xs ds
x = (m) s .
2i ai (s) s 2i ai m s
m=1

15
We want to integrate term by term. This is justifiable on a finite interval
because the series converges uniformly for Re(s) a > 1 (since (m)
log m and x is constant). Thus, we have

Z a+ih X ! Z a+ih s
1 xs ds 1 X x ds
(m) s = (m) s
.
2i aih m s 2i aih m s
m=1 m=1

To take the limit as h and evaluate the integral, we will need the
following lemma, which can be proved by straightforward estimates and
integration by parts.
Lemma 13. For t > 0,

1
Z a+ih 0
t<1
s ds 1
lim t = u(t 1) = , t=1
h+ 2i aih s 2
1, t > 1.

For 0 < t < 1 and for 1 < t, the error


Z a+ih
ta

1 s ds

2i t u(t 1) .
aih s h| log t|
Let x/m be the t in the lemma and noticing u(x/m 1) = u(x m), we
have
Z a+ih s a
1 x ds (x/m) log m

(m)
u(x m)
2i ms s
aih
h| log x log m|

The right hand side is summable over m because (log m)/| log x log m|
approaches 1 as m and (x/m)a is summable. Thus, for fixed x,
Z a+ih s
X x ds K
(m)
s
u(x m) ,
aih m s h

m=1

for some constant K. Hence, the limit of the sum as h is zero. This
argument assumes x 6= m, but if x is an integer, we can simply throw away
the terms where m x, since finitely many terms do not affect convergence.
This implies
Z a+ih s
1 X x ds X
lim (m) s
= (m)u(x m),
h 2i aih m s
m=1 m=1
or Z a+i 0
1 (s) s ds
x = (x).
2i ai (s) s
This completes the first evaluation of the integral.

16
7.2 Second Evaluation of the Integral
For the second evaluation, we use a different formula for 0 /. Differentiate
 
X s s s
log (s) = log (0) + log 1 + log log(s 1) log

2 2

to obtain
0 (s) X 1 1 1 d s
= log + + log
(s)
s 2 s 1 ds 2

In order to justify termwise differentiation,P


we use the fact proved by Hadamard
(and which we will not prove here!) that 1/ with the roots and 1
paired converges absolutely. We order the terms by increasing |Im()|. Since

lim|Im()| s = 1,

1 k
s <

for any fixed s, for Im() sufficiently large. In fact, on the set {s : Re(s) >
1 + , a < Im(s) < a}, we can bound |1/( s)| uniformly by |1/( s0 )|
where s0 = 1 + + ai. This shows that the series converges uniformly on
compact sets and term-by-term differentiation is justified.
QIn the preceding formula, substitute the product formula for (s) =
1 (1 + 1/n)s from Theorem 3:
n=1 (1 + s/n)
   
d s d X 1  s 
log = s log 1 + log 1 +
ds 2 ds 2n 2n
n=1
   
X 1 1
= log 1 + +
2n 2n + s
n=1

This termwise differentiation is justifiable because the differentiated series


converges uniformly on compact subsets of C with Re(s) > 1. The reader
may verify this: Use the power series for log(1 + t) and show that each term
of the differentiated series is less than a constant times 1/n2 .
Using these formulae, we have

0 (s) 0 (0) X 1 X1 
1 X 1 1
+ = + +1+ +
(s) (0)
s
s1 2n + s 2n
n=1

0 (s) s X s X s 0 (0)
= + .
(s) s1
(s ) 2n(s + 2n) (0)
n=1

17
This is the formula for 0 / we intend to integrate term by term. To do this,
we need another lemma. The reader may try deducing this lemma from the
previous one using a change of variables and Cauchys theorem.

Lemma 14. For a > 1 and x > 1,


Z a+i
1
xs ds = x
ai s

The lemma allows us immediately to integrate the first and last terms:
Z a+i 0
0 (0)
Z a+i
1 s s ds 1 (0) s ds
x = x, x = .
2i ai s 1 s 2i ai (0) s (0)

The lemma also tells us what the value of other integrals will be, if we can
justify term-by-term integration.
First, consider

Z a+i X !
1 s ds
xs .
2i ai 2n(s + 2n) s
n=1

Because the series converges uniformly on compact subsets of {Re(s) > 1},
we can integrate term by term over a finite interval. We only have to worry
about taking the limit
Z a+ih
X 1 1 xs
lim ds.
h 2i 2n aih s + 2n
n=1

To justify the termwise limit, we begin by showing that the sum of the
limits converges. By Lemma 14,

x2n
Z a+i
1 1 xs
ds = .
2i 2n ai s + 2n 2n

This is clearly summable over n (in fact, it sums to 12 log(1 x2 )). We


now show that the sum before we take the limit (that is, the sum of the
partial integrals) converges, using the estimate

Lemma 15. For x > 1, a > 0, and d > c 0,


Z a+di s
1 t K

2i ds .
a+ci s (a + c) log t

18
By change of variables,
a+ih a+2n+ih a+2nih
xs xs xs
Z Z Z
ds = x2n ds = 2x2n ds.
aih s + 2n a+2nih s a+2n s

And by Lemma 15,

2Kx2n K0
Z a+2nih s
2n 1 x
2x ds .
2i 0 s (a + 2n) log x n2

Hence, the sum of the partial integrals is summable. This implies that
evaluating the limit termwise is valid. This is because for any  > 0, we
can choose N large enough that the sum of the first N infinite integrals is
within  of the complete sum. By choosing N larger if necessary, we can
make the sum of the first N finite integrals within /3 of its limit, uniformly
in h. Then by choosing h large enough, we can make the first N integrals
be within /3N of their limits. Thus, the two partial sums are within /3
of each other, and so the infinite sum of the infinite integrals is within  of
the infinite sum of the finite integrals. P
All that remains is to evaluate the integral of s/( s).

7.3 Termwise Evaluation over


We will need the following lemma on the density of the roots of , given
in Edwards section 3.4. The proof of this lemma is highly nontrivial. It is
ultimately based on Hadamards results.

Lemma 16. For T sufficiently large, has fewer than 2 log T roots in T
Im() T + 1.

We want to evaluate
!
Z a+i
1 X s ds
xs .
2i ai
(s ) s

We first note that we can exchange summation and integration for a finite
integral:
X 1 Z a+ih xs
lim ds.
h

2i aih (s )

Uniform convergence of the series is obtained from Hadamards result that


P 1 2
| 2 | converges. After pairing the terms for and 1 , we can

19
show that each term of the series is less than a constant times | 12 |2 .
The proof is merely a manipulation of fractions and is left to the reader.
We already know the limit exists because we have shown that all the
other terms in von Mangoldts formula have limits as h . Von Mangoldt
considers the diagonal limit
X x Z a+ih xs
lim ds.
h aih s
|Im()|h

He shows that
X x Z a+ih a+ih
xs x xs
X Z
ds ds

aih s aih s
|Im()|h

approaches zero, so that the diagonal limit exists and is equal to the original
limit. If is written as + i, then we can estimate the sum of roots with
positive as

X x 1 Z a+ih xs X x 1 Z a+i(+h) xt
ds ds ,
aih s
2i 2i a+i(h) t


>h >h

This, in turn, is less than or equal to


X x Kxa
(a + h) log x
>h

by Lemma 15, which is less than or equal to


xa X 1 K
K ,
log x ( h + c) log x
>h

where c = a 1 so that 0 < c a . By writing the roots with < 0 as


1 = 1 i and making a change of variables, the reader may verify
that the same chain of inequalities applies with replaced by 1 , but the
c we chose works for that case as well. Hence, the sum over all roots can be
estimated by twice the above estimate.
Now take h sufficiently large that Lemma 16 applies. Arrange the roots
into groups where h + j < h + j + 1 for each integer j. There are at
most 2 log(h + j) roots in each group, so summing over j makes the above
quantities smaller than a constant multiple of

X log(h + j)
.
(h + j)(j + c)
j=0

20
This sum converges and as h , it approaches zero; the proof of this is
left to the reader.
We now only have to evaluate the diagonal limit, and we know that it
exists. We will show that
X x Z a+ih xs X x
ds
aih s
|Im()|h |Im()|h

approaches zero as h . This will prove that x / converges and


P
is equal to the diagonal limit. By similar reasoning as before, the above
difference is smaller than twice
X x 1 Z ai+ih xt


2i dt 1 ,
aiih t
0<h

which is no larger than


Z a+i(h) t
X x 1 Z a+i(h+) xt X x 1 x
dt 1 + dt

2i
ai(h+) t 2i a+i(h+) t

0<h 0<h

by path additivity of integrals and the triangle inequality. By Lemmas 13


and 15 and similar manipulation as before, this is less than or equal to
X x xa X x Kxa
+
(h + ) log x (a + h ) log x
0<h 0<h
xa X 1 Kxa X 1
+ ,
log x (h + ) log x (c + h )
0<h 0<h

where c = a 1. Our object is to show that this series approach zero as


h . We fix an H such that the estimate of Lemma 16 holds for h H.
There are only finitely many terms where H, and it is easy to see each
term approaches zero as h . The remaining terms we put into groups
where h + j < h + j + 1 as before and apply the estimate of Lemma
16. We estimate the quantity by convergent series which approach zero as
h ; since the argument is similar to the one we did before, I will skip
the details.
This completes the proof that
Z a+i X !
1 s ds X x Z a+ih xs X x
xs = lim ds = .
2i ai
(s ) s h aih s

|Im()|h

21
7.4 Von Mangoldts and Riemanns Formulae
We have now evaluated term by term

!
a+i
0 (0)
Z
1 s X s X s ds
+ xs ,
2i ai s1
(s ) 2n(s + 2n) (0) s
n=1

showing that

X x x2n 0 (0) a+i
0 (s) s ds
Z
X 1
x + = x = (x).

2n (0) 2i ai (s) s
n=1

The one thing missing from von Mangoldts formula is the evaluation of
0 (0)/(0) which Edwards performs in section 3.8. It is log 2.
Von Mangoldt used his formula to prove Riemanns main formula; Ed-
wards explains this proof in section 3.7. Von Mangoldt used a different
method from Riemanns and did not directly justify Riemanns term-by-
term integration, although three years later in 1908, Landau did just that.
Von Mangoldts formula can also be used to prove the prime number theo-
rem, which says essentially that (x)/x approaches 1 as x .

References
[1] Harold M. Edwards. Riemanns Zeta Function. 1974. Dover Publica-
tions.

[2] Joseph L. Taylor. Complex Variables. 2011. American Mathematical


Society.

22

You might also like