You are on page 1of 76

Prime numbers and

the Riemann zeta function.

by Jørgen Veisdal
1
2
For Nash.

3
Preface

This paper is concerned with explaining the results of the 1859 paper “On the Number of
Primes Less Than a Given Magnitude” by German mathematician Bernhard Riemann. It is
meant to be an introduction to the field of prime number theory for novices and fellow
mathematics students.

The paper gathers results, conjectures and proofs from many pieces of modern literature to
try to explain the fundamentals of prime number theory in the simplest ways possible.

Although the paper contains explanations of several proofs, it is advised that these are to be
considered as proof sketches, which in some cases are in need of more completeness to be
considered entirely rigor. They are shown in these simpler terms to make reading as easy as
possible. For the rigorous results they are based on, the reader is referred to the reference
literature.

The assumed reader has an undergraduate level knowledge of mathematics.


Estimated reading time is 45 minutes.

The student would like to thank his advisor, Professor Alexander Ulanovskii.

Jørgen Veisdal
Trondheim, 09.05.13.

4
Table of contents

Notation
Table of figures
Summary

Prime numbers 1
Fundamental resources 4
The harmonic series - Hr 5
Zeta functions - ⇣(x) 5
The Euler-Mascheroni constant - 8
The Euler product formula 11
The Möbius function - µ(n) 13
The prime counting function - ⇡(x) 16
The prime number theorem 16
The logarithmic integral function - Li(x) 18

The gamma function - (z) 20


Bernhard Riemann 25

The Riemann zeta function - ⇣(s) 26

Zeros of the Riemann zeta function 34

The Riemann Hypothesis 38

The Riemann zeta function and prime numbers 40

Modern work 48

Appendix
A. Functional equation for the Jacobi theta function - #(x) 49
B. Landau’s Problems 51
Goldbach’s conjecture 52
The twin prime conjecture 55
Legendre’s conjecture 59
“Near-square primes” conjecture 62

References 63
Bibliography 65

5
Notation
The notation was created using the markup language LaTeX.
Unless specified otherwise, the notation used in this paper is as follows:

Numbers
Prime number - p 2 N .
Integers - n or r 2 Z .
Real number - x , or t 2 R .
Complex number - s or z 2 C .
Complex parts - s = + it or s = <(s) + =(s).
Real part - or <(s).
Imaginary part - t or =(s).

Constants and exponents:


Bernoulli number - Bn for an integer n .
The Euler-Mascheroni constant, “gamma” - .
A Dirichlet character - (n) for an integer n .

The natural logarithm - ln or log .

Functions and series:


The digamma function - (x) , for a complex number x *.

The Dirichlet eta function ⌘(s) , for a complex number s .


The gamma function - (z) , for a complex number z *.
The harmonic series - Hr , for a natural number r .
The Jacobi theta function - #(x), for a complex number x *.
The logarithmic integral function - Li(x) , for a real number x .
The Möbius function - µ(n) , for a natural number n .
The prime counting function - ⇡(x) , for a real number x .
The Riemann zeta function - ⇣(s) , for a complex number s .
The Riemann Xi function - ⇠(s), for a complex number s .
The Riemann prime counting function - J(x), for a real number x .

*Certain inconsistencies are due to convention.

6
List of figures

All figures with the exception of figures 1 and 17 were created in Wolfram Mathematica 9.

Figure 1: “Composite ladders” for the primes 5, 7, 11 and a combination of the three.

Figure 2: The natural logarithm ln(r) and the harmonic series.


Figure 3: The prime counting function ⇡(x) up to x = 200 .
Figure 4: The prime counting function ⇡(x) and the estimate from the prime number
theorem.
Figure 5: The logarithmic integral function Li(x).
Figure 6: The prime counting function ⇡(x) with estimates x/ ln(x) and Li(x).
Figure 7: Convergence of the ratios of the two estimates and the prime counting function.
Figure 8: The gamma function (z) plotted in the range 6  z  6.
Figure 9: The zeta function ⇣(s) plotted, with the trivial zeros at s = 2, 4, ... highlighted.

Figure 10-11: The complex parts of the zeta function ⇣(s) plotted, with zeros highlighted.
Figure 12: The first non-trivial zeros of the Riemann zeta function on the line <(s) = 1/2.
Figure 13: The Riemann prime counting function J(x) up to x = 50 .
Figure 14: The Riemann prime counting function J(x) up to x = 50 , with two integrals
highlighted.
Figure 15-16: The prime counting step function ⇡(x) being approximated by the explicit
formula for the Riemann prime counting formula J(x) using the first 35 non-trivial zeros ⇢ of
the Riemann zeta function ⇣(s).
Figure 17: Goldbach sums: scatter plot of number of ways to write even integers as sums of
two primes.
Figure 18: Plot of function |6n ± 1|.
Figure 19: Plot of function |6n ± 1| with twin primes highlighted.
Figure 20: Plot of function |6n ± 1| and the first “twin prime functions“.
Figure 21: Plot of the first 20 “twin prime functions“.
Figure 22: Plot of the first 40 “twin prime functions“.

Figure 23: The prime counting step function ⇡(n) and vertical lines representing the perfect
squares.

7
Summary

The properties of the prime numbers have been studied by many of history’s mathematical
giants. From the first proof of the infinity of the primes by Euclid, to Euler’s product formula
which connected the prime numbers to the zeta function. From Gauss and Legendre’s
formulation of the prime number theorem to its proof by Hadamard and de la Vallée Poussin.
Bernhard Riemann still reigns as the mathematician who made the single biggest
breakthrough in prime number theory. His work, all contained in an 8 page paper published
in 1859 made new and previously unknown discoveries about the distribution of the primes
and is to this day considered to be one of the most important papers in number theory.

In his paper Riemann defined the Riemann zeta function for complex numbers and
analytically continued it to an absolutely converging function on the entire complex plane,
with two singularities. He rewrote the analytic continuation and defined the Riemann Xi
function, a “completed Riemann zeta function” defined on the entire complex plane. While
studying the roots of the Riemann Xi function he hypothesized that they would all be real,
and went on to show the consequence of this assumption by defining the Riemann prime
counting function. Using this new tool he showed a stronger version of the prime number
theorem in an explicit formula for the number of primes less than a given magnitude. The
formula incorporated the roots of the zeta function in a four part error term that corrects the
guess made by the prime number theorem to a chosen accuracy. In a sentence; Riemann
showed how to find the number of primes less than a given magnitude by using the zeros of
the zeta function.

Since the publication of Riemann’s paper it has been the main focus of prime number theory,
and was indeed the main reason for the proof of the prime number theorem in 1896. Since
then several new proofs have been found, including elementary proofs by Selberg and
Erdós. Riemann’s hypothesis about the roots of the zeta function however, remains unsolved.
A reformulation was given in 1899 by German mathematician Edmund Landau who in 1912
also listed four other basic problems about prime numbers which since have been known as
Landau’s problems.

All five problems still remain unsolved to this day.

8
The prime numbers
A prime number is a natural number n greater than 1 that has no positive divisors other than
1 and itself.

The importance of prime numbers


By the unique factorization theorem, every integer that is not a prime number, is made up of
a unique factorization of prime numbers. This fact establishes prime numbers as the
building blocks of all numbers.

Prime numbers have been known to be infinite in number since Euclid proved this ca 300
BC. His elegant proof is as follows (The Prime Pages, 1994):

Euclid’s proof of the infinity of primes [1]:


Assume that the set of prime numbers is finite. Make a list of all the prime numbers
p1 , p2 , p3 , · · · , pn . Let P be the product of all the prime numbers in the list,
P = p1 ⇥ p2 ⇥ p3 ⇥ · · · ⇥ pn . Let the number q = P + 1 . The number q is either prime or
composite.

If q is prime, we have found a prime that is not in the list, which is a contradiction that the
set contains all the primes.
If q is composite, then by the unique factorization theorem it must be divisible by some
prime p not in the list.

Either way, there will always be primes not included in any list, regardless of size. Therefore
there must be infinitely many prime numbers.

Arithmetic complexity
The arithmetic properties of primes, while heavily studied, are still largely unknown and there
is no known formula that can compute primes or quickly state whether or not an integer is
prime. This fact has laid the foundation for large parts of modern encryption theory, and one
of the most fundamental tasks in decryption is to this day the factorization of large
composites into their prime components.

1
Although much effort has been made to create methods and formulas for producing primes,
the success of these efforts have been limited. One of the most famous methods are prime
number sieves, the most famous example being the sieve of Eratosthenes.

Research
One of the simplest ways to generate primes is by using the function 6n ± 1. This function
includes all primes p > 3 , and does not contain any multiples of 3 or any even numbers. If
we look at the first numbers given by this function, for n = 1, 2, 3, 4, 5, 6, 7, ... we get
{5, 7, 11, 13, 17, 19, 23, 25, 29, 31, 35, 37, 41, 43, ...} . The non-prime numbers mentioned are 25
and 35. These are the first two multiples of 5. And indeed, the list only contains composite
numbers given by multiplying members of the list with each other.

So, every prime in the list multiplied by every item in the list generates a sequence of
composite numbers also in the list. For the number 5, its composite sequence is:
{25, 35, 55, 65, 85, 95, ...}, given by multiplying 5 with itself and the other items in the list. The
multiples of 5 not included are divisible by either even numbers or 3. We can visualize these
sequences for different primes in the following way:

the prime 5 the prime 7 the prime 11 the primes 5,7,11


5 7 5 7 5 7 5 7

11 13 11 13 11 13 11 13

17 19 17 19 17 19 17 19

23 25 23 25 23 25 23 25

29 31 29 31 29 31 29 31

35 37 35 37 35 37 35 37

41 43 41 43 41 43 41 43

47 49 47 49 47 49 47 49

53 55 53 55 53 55 53 55

59 61 59 61 59 61 59 61

65 67 65 67 65 67 65 67

71 73 71 73 71 73 71 73

77 79 77 79 77 79 77 79

83 85 83 85 83 85 83 85

89 91 89 91 89 91 89 91

Figure 1: “Composite ladders”, a way to see how the composite multiples are laid out for each prime.
2
The four lists show:
1. The prime number 5 and its composite multiples.
2. The prime number 7 and its composite multiples.
3. The prime number 11 and its composite multiples.
4. The prime numbers 5,7,11 and their composite multiples.

In each list, we have highlighted the generating prime number and its multiples. We see that
the composite multiples follow an oscillating pattern, for example for the prime number 5, it
increases by 20 and 10 between each step. The general rule for gaps between multiples of
primes in this set is an oscillating pattern of gap sizes 4p and 2p for the generating prime
p.

The fourth list is a good visualization of why prime numbers are hard to predict. Despite
removing all even numbers and multiples of 3 and following logical steps containing
arithmetic progressions, the resulting image is still highly complicated and hard to discern.

Because of the difficulty of understanding the arithmetic properties of the prime numbers,
the focus of modern prime number theory has been mainly concentrated on studying prime
numbers with analytic tools.

3
Fundamental resources
Sometime in the 1800’s, mathematicians began to accept the fact that prime numbers are
hard to predict in specific terms. To learn more about them, new methods would be needed.
It might have been Gauss that first began to think of the larger picture of prime distribution,
how often primes occur in certain ranges rather then where specific primes are. By studying
the primes in this new way, major discoveries were made about prime numbers and their
distribution.

In order to study the modern discoveries about prime numbers, we must study the
sequences, series and functions that they are related to. Before we can look at the specific
results Riemann found, we must be familiar with certain fundamental resources from number
theory. These resources are:

The harmonic series - Hr


Zeta functions - ⇣(s)
The Euler-Mascheroni constant -
The Euler product formula
The Möbius function - µ(n)
The prime counting function - ⇡(x)
The prime number theorem
The logarithmic integral function - Li(x)
The gamma function - (z)

4
The Harmonic Series - Hr
It was bishop Nicholas Oresme of Lisieux who in the 14th century first studied the harmonic
series. (Havil, 2003). The series has its name from the concept of harmonics in music,
overtones higher than the fundamental frequency of a sound.

The series Hr is defined for the natural number r 2 N by:


1
X 1 1 1 1
Hr = =1+ + + + ...
r=1
r 2 3 4
This infinite series is the harmonic series, and Oresme’s main achievement was proving its
divergence.

Oresme’s proof of the divergence of the harmonic series [2]:


We prove the divergence of Hr by comparing it to another divergent series:
1 1 1 1 1 1 1 1 1
H1 = 1 +
+( + )+( + + + )+( + + ···
2 3 4 5 6 7 8 9 10
1 1 1 1 1 1 1 1 1
>1+ +( + )+( + + + )+( + + ···
2 4 4 8 8 8 8 16 16
We add the terms and see that:
1 1 1 1 1 1 1 1 1 1
1+ + + + + + + + ··· > 1 + + + + ···
2 3 4 5 6 7 8 2 2 2
1 1 1
1 + + + + ···
Conclusion: Hr is bigger than the divergent series 2 2 2 , and so is
divergent as well (Havil, 2003).

Zeta Functions - ⇣(n)


The harmonic series is a special case of a more general type of function called a zeta
function. The real valued zeta function is given for r, n 2 R :
X1
1 1 1 1 1
⇣(n) = = 1 + + + + · · · +
r=1
rn 2n 3n 4n rn

The harmonic series is the case n = 1, which we have seen is divergent. The series
converges for all values of n > 1. This can be proven by comparing it to a geometric series
which is known to be convergent.

5
Proof of the convergence of zeta functions with n >1 [3]:
Again we use the comparison test, and compare the zeta function to a series that is known
to be convergent, a general geometric series:
X1
1 1 1 1 1 1 1
n
= 1 + ( n
+ n
) + ( n
+ n
+ n
+ n
) + ···
r=1
r 2 3 4 5 6 7
1 1 1 1 1 1
<1+( + ) + ( + + + ) + ···
2n 2n 4n 4n 4n 4n
We add the terms, and see that:
X1
1 2 4 8
n
< 1 + n
+ n
+ n
+ ···
r=1
r 2 4 8
X1
1 1 1 1
< 1 + +( )2 + ( )3 + · · ·
r=1
rn 2n 1 2n 1 2n 1
X1
1 1
n
< 1
r=1
r 1 2n 1

The zeta function is smaller than the equation on the right hand side, provided that:
1
<1 n 1
2n 1 , that is 2 > 1 when n 1 > 0 , n > 1.
Conclusion: because the given zeta function with n > 1 is smaller than the geometric series,
it too converges (Havil, 2003).

The Basel Problem


The zeta function with n = 2 is often referred to as the Basel problem. First posed by Pietro
Mengoli, the Basel problem asks for the closed form value of the zeta function with n = 2 . It
is approximately ≈ 1.644934. The closed form was found by Leonard Euler in 1735 (Plus
Magazine, 2001).

Proof of the closed form for the zeta function with r = 2 [4]:
The closed form is found by using the familiar Taylor series expansion for sine:
x3 x5 x7
sin(x) = x + + ···
3! 5! 7!
Dividing by x , we get:
sin(x) x2 x4 x6
=1 + + ···
x 3! 5! 7!
The zeros of this function occur at x = n ⇥ ⇡ , where n = ±1, ±2, ±3, · · ·
6
We can express the infinite series as a product of linear factors given by its roots (where its
zeros occur) because of Newton’s identities:
sin(x) x x x x x x
= (1 )(1 + )(1 )(1 + )(1 )(1 + )···
x ⇡ ⇡ 2⇡ 2⇡ 3⇡ 3⇡
x2 x2 x2
= (1 )(1 )(1 )···
⇡2 4⇡ 2 9⇡ 2
Multiplying out this infinite product and collecting the x2 terms, the x2 coefficient becomes:
✓ ◆
1 1 1 1 X 1
+ + ··· =
⇡2 4⇡ 2 9⇡ 2 ⇡2 n2
But we saw in the modified series expansion of sine, that the coefficient of x2 is
1 1
=
3! 6
Both must be true, so they are equal:
1 1
1 1 X 1 X 1 ⇡2
= ) =
6 ⇡ 2 n=1 n2 n=1
n2 6

And so Euler had found the closed form of the zeta function with n = 2 to be equal to:
X1
1 ⇡2
=
r=1
r2 6

More generally, in a paper published in 1750, Euler found a formula that would give the
closed form of every zeta function with positive even integer 2n (Havil, 2003):
X1 2n
1 1 (2⇡)
⇣(2n) = 2n
= ( 1)n B2n
r=1
r 2(2n)!

where B2n are the even Bernoulli numbers. This formula is called Euler’s formula. The
Bernoulli numbers are numbers related to the binomial coefficients
✓ ◆
n n!
=
r r!(n r)!
Although Euler’s formula settled the case for the even positive integers, his formula did not
find the closed form for the odd integers. To this day, no such formula is known for zeta

functions with positive odd integer 2n + 1, and these values still have to be approximated.
In fact, remarkably little is known about these zeta functions. One famous case, n = 3 is
known as Apéry’s constant. It has been proven to be irrational, by Apéry’s theorem.
X1
1 1 1 1
⇣(3) = 3
= 1 + 3 + 3 + 3 + · · · ⇡ 1.2020569031...
r=1
r 2 3 4

7
The Euler-Mascheroni constant -
The Euler-Mascheroni constant, often called gamma , is defined as the limiting difference
between the harmonic series and the natural logarithm. In other words, gamma is an
approximation to the difference between the harmonic series Hr and the natural logarithm
ln(r).

As we saw from Oresme’s proof, the harmonic series Hr (zeta with n = 1) diverges. It does
so slowly. How slowly can be measured by using its interpretation as a discrete logarithm
(Havil, 2003):

Z r
1
dx = ln(r)
1 x

ln(r)

r-1 r x

Figure 2: The natural logarithm ln(r) (blue) and the harmonic series (green).

The green area above the graph is equal to the Euler-Mascheroni constant, the difference

between the harmonic series Hr and the natural logarithm ln(r) as r ! 1 .

8
More specifically, the constant gamma can be expressed as the difference between two
areas of overestimating and underestimating rectangles, estimating the natural logarithm
ln(r). These two areas of rectangles give the inequality:
Z r
1 1 1 1 1 1 1
+ + ··· + < dx < 1 + + + · · · +
2 3 r 1 x 2 3 r 1
We see two harmonic series appearing, and express the above equation as:
1
Hr 1 < ln(r) < Hr
r
Or, the equivalent statement:
1
ln(r) + < Hr < ln(r) + 1
r
This is an estimate of the harmonic series Hr as ln(r) with an error of at least 1/r and at
most 1 [5].

To make the relationship between the natural logarithm and the harmonic series more
explicit, Euler starts with the familiar Taylor series expansion:
X1
( 1)n+1 n x2 x3
ln(1 + x) = x =x + ···
n 2 3
n=1 for | x | 1 :
Then sets x = 1/r :
1 1 1 1 1
ln(1 + ) = 2
+ 3 + ···
r r 2r 3r 4r4
Rearranging to express for the harmonic series:
1 1 1 1 1
= ln(1 + ) + 2 3
+ 4 ···+
r r 2r 3r 4r
We then take the sum of both sides (notice the split logarithm):
n
X n
X n n n
1 1X 1 1X 1 1X 1
= (ln(r + 1) ln r) + + ···+
r=1
r r=1
2 r=1 r2 3 r=1 r3 4 r=1 r4
n
X n n n
1 1X 1 1X 1 1X 1
= ln(r + 1) + + ···+
r=1
r 2 r=1 r2 3 r=1 r3 4 r=1 r4

And by moving the natural logarithm over to the left hand side, we see:
n
X n n n
1 1X 1 1X 1 1X 1
ln(r + 1) = + ···+
r=1
r 2 r=1 r2 3 r=1 r3 4 r=1 r4
Which states that as r ! 1 , gamma is a sum of an infinite number of converging zeta
series with different exponents and signs [6].

9
As saw earlier, the even zeta series in the sum can be found in closed form by the formula
X1 2n
1 1 (2⇡)
⇣(2n) = 2n
= ( 1)n B2n
r=1
r 2(2n)!

where B2n are the even Bernoulli numbers.

The odd zeta series in the sum have to be approximated. If we do these approximations to a
given accuracy then the right hand side (the difference between the harmonic series Hr
and the natural logarithm ln(r) is = 0.577218... , the Euler-Mascheroni constant.

And so, the connection between gamma and the harmonic series Hr is as follows:
1 1 1
1+ + + · · · + = ln(r + 1) + Hr = ln(r + 1) +
2 3 r or

10
The Euler Product formula
The first connection between zeta functions and prime numbers was made by Euler when
he showed that for n, p 2 N , where p = primes :
X 1 Y 1
s
= s
n
n p
1 p

This expression first appeared in a paper titled Variae observationes circa series infinitas in
1737. The expression states that the sum of the zeta function is equal to the product of the
reciprocal of one minus the reciprocal of primes to the power s . The discovery laid the
foundation for modern prime number theory, which from this point on would be focused on
studying the zeta function ⇣(s) to learn about the behavior of primes.

Proof of Euler’s product formula [7]:


Euler began with the general zeta function
1 1 1 1
⇣(s) = 1 + + + + + ··· (1)
2s 3s 4s 5s
1
First multiply both sides by 2s :
1 1 1 1 1 1
⇥ ⇣(s) = + + + + + ···
2s 2s 4s 6s 8s 10s
Subtract the resulting expression from the zeta function (1):
1 1 1 1 1
(1 s
)⇣(s) = 1 + s + s + s + s + · · · (2)
2 3 5 7 9
1
Repeat, now multiply both sides by 3s :
1 1 1 1 1 1
⇥ ⇣(s)(1 ) = + + + + ···
3s 2s 3s 9s 15s 21s
And subtract this from the now altered zeta function (2):
1 1 1 1 1 1
(1 )(1 )⇣(s) = 1 + + + + + ··· (3)
3s 2s 5s 7s 11s 13s
Repeating this process for every term on the right hand side eventually gives:
1 1 1 1 1 1
· · · (1 )(1 )(1 )(1 )(1 )(1 )⇣(s) = 1
13s 11s 7s 5s 3s 2s
Each parenthesis on the left hand side contains a prime reciprocal.

Divide by these parentheses to express this for zeta:

11
1 1 1 1 1
⇣(s) = ( 1 )⇥( 1 )⇥( 1 )⇥( 1 )⇥( 1 ) ⇥ ···
1 2s 1 3s 1 5s 1 7s 1 11s

Shortened, it has been shown that:


X 1 Y 1
=
n
ns p
1 p s

For s = 1 , Euler’s product formula is another proof of the infinity of the primes. We get:
✓ ◆✓ ◆✓ ◆✓ ◆✓ ◆
1 1 1 1 1
··· 1 1 1 1 1 ⇣(1) = 1
11 7 5 3 2
Which we can write as:
✓ ◆✓ ◆✓ ◆✓ ◆✓ ◆
10 6 4 2 1
··· ⇣(1) = 1
11 7 5 3 2
Which is:
✓ ◆
... · 10 · 6 · 4 · 2 · 1
··· ⇣(1) = 1
... · 11 · 7 · 5 · 3 · 2

Because ⇣(1) is just the harmonic series


1 1 1 1
⇣(1) = Hr = 1 + + + + + · · ·
2 3 4 5
Then if we divide and multiply by the infinite product on both sides, we get:
1 1 1 1 2 · 3 · 5 · 7 · 11 · · ·
1 + + + + + ··· =
2 3 4 5 1 · 2 · 4 · 6 · 10 · · ·
We know that the harmonic series on the left hand side is infinite. An infinite sum cannot
equal a finite product, so both the numerator and denominator of the fraction on the right
hand side must both also be infinite. The numerator contains the prime numbers, and so
they must be infinite in number (Derbyshire, 2004).

12
The Möbius function - µ(n)
August Ferdinand Möbius rewrote the Euler product formula to create a new sum. In
addition to containing reciprocals of primes, the Möbius function also contains every natural
number that is the product of odd and even numbers of prime factors. The numbers left out
of the sum are those that divide by some prime squared. To show this function, Möbius
begins with the Euler product formula [8]:

It has been shown that the zeta function can be written as:
1 1 1 1 1
⇣(s) = ( 1 )⇥( 1 )⇥( 1 )⇥( 1 )⇥( 1 ) ⇥ ···
1 2s 1 3s 1 5s 1 7s 1 11s

which is given by the Euler product formula. Rewritten, this is:


s 1 s 1 s 1 s 1 s 1 s 1
⇣(s) = (1 2 ) (1 3 ) (1 5 ) (1 7 ) (1 11 ) · · · (1 p )

Möbius takes the reciprocal of both sides:

1 1 1 1 1 1
= (1 )(1 )(1 )(1 ) · · · (1 )
⇣(s) 2s 3s 5s 7s ps

And then multiplies out the right hand side in the following way:

Part 1 - the prime reciprocals:

a. Take (1) from each parenthesis (1)(1)(1)(1) · · ·


1 1
( )(1)(1)(1) · · ·
b. Take 2s from the first parenthesis and (1) from the rest 2s
1 1
(1)( s )(1)(1) · · ·
c. Take 3s from the second parenthesis, and (1) from the rest 3

Continue this process to infinity, and the first part of the sum becomes:
1 1 1 1
1 ···
2s 3s 5s ps

Part 2 - reciprocals of products of two primes:


1 1 1 1
( )( s )(1) · · ·
a. Take 2s , 3s and (1) from the rest 2s 3
1 1 1 1
( )( s )(1) · · ·
b. Take 3s , 5s and (1) from the rest 3s 5

13
Continue this process to infinity, and the second part of the sum becomes:
1 1 1 1 1
s
+ s + s + s ··· + s
6 10 14 15 (pn 1 ⇥ pn )

Part 3 - reciprocals of products of three primes:


1 1 1 1 1 1
( )( s )( s )(1) · · ·
a. Take 2s , 3s , 5s and (1) from the rest: 2 s 3 5
1 1 1 1 1 1
( )( s )( s )(1) · · ·
b. Take 2 , 3s ,
s 7s and (1) from the rest: 2 s 3 7

Continue this process to infinity, and the third part of the sum becomes:
1 1 1 1 1
s s s s
···
30 42 66 70 (pn 2 ⇥ pn 1 ⇥ pn )s

Continuing this process to infinity for the reciprocals for all products of all primes eventually
gives the sum:

1 1 1 1 1 1 1
=1 + + ···
⇣(s) 2s 3s 5s 6s 7s 10s

The sum contains reciprocals of:


1. Every prime.
2. Every natural number that is the product of an odd number of different primes, prefixed by
a minus sign.
3. Every natural number that is the product of an even number of different primes, prefixed
by a plus sign.

The sum does not contain reciprocals of:


1. Every number that divides by some prime squared.

This is the the Möbius function, denoted µ(n) (Derbyshire, 2004). It is defined for the natural
numbers n 2 N .
1 X µ(n)
=
⇣(s) ns
n2N

The Möbius function has the following possible values:

14
8
< 1 if n is square-free with an even number of prime factors
µ(n) = 0 if n is not square-free
:
1 if n is square-free with an odd number of prime factors

Although Möbius was the first to formally define and show the function, remarkably Gauss
considered the function more than 30 years before, writing “The sum of all primitive roots (of
a prime number p) is either ≡ 0 (when p-1 is divisible by a square), or ≡ ±1 (mod p) (when
p-1 is the product of unequal prime numbers); if the number of these is even the sign is
positive, but if the number is odd, the sign is negative.” [9].

15
The prime counting function - ⇡(x)
To show how the primes are laid out between the integers without knowing where they are, it
is useful to look at how many primes there are up to a given number.

The prime counting function ⇡(x) gives the number of primes less than or equal to a given
+
real number x 2 R , and is plotted below up to x = 200 :
⇡(x)
pHnL

⇡(x)
40

30

20

10

50 100 150 200


x
n

Figure 3: A step function: the prime counting function ⇡(x) up to x = 200.

Introduced by Gauss, the prime counting function is a step function which increases by 1
whenever x is a prime. If we look at x = 100 , then we find that ⇡(x) = 25 by counting the
number of primes up to and equal to 100 . Gaps between prime numbers appear as
horizontal lines in the plot, often called prime gaps.

The prime number theorem


Let ⇡(x) denote the prime counting function where ⇡(x) = the number of primes  x .
Then the prime number theorem states:
⇡(x) x
lim x =1 ⇡(x) ⇠
x!1
ln(x) which in asymptotic notation is ln(x) as x ! 1 .

16
Formulated by both Gauss and Legendre independently, the prime number theorem states
that:
x
As x ! 1 , the prime counting function ⇡(x) will approximate the function ln(x) .

⇡(x)
pHxL

150
pHxL

x
x ln(x)

100
logHxL

50

200 400 600 800 1000


x
x

Figure 4: The prime counting function ⇡(x) and the estimate from the prime number theorem.

The theorem is concerned with the distribution of primes as x ! 1 . It attempts to answer


the question “how many primes are there up to and including a given number?”

In terms of probability, the prime number theorem states that if you pick an integer x 2 N at
random, the probability P (x) that this integer will be a prime number is about 1/ ln(x) as
x ! 1.
1
P (x) = prime number =
ln(x)
This means that the average gap between consecutive prime numbers among the first x
integers is approximately ln(x) [10].

17
The logarithmic integral function - Li(x)
Li(x)
The function Li(x) is defined for all
6

positive real numbers x 6= 1 . We are 4

interested in the definition: 2

Z
2 4 6 8 10

x
1
Li(x) = dt -2

2 ln(t)
-4

-6

The function has an integral


Figure 5: The logarithmic integral function.
representation with a single positive zero at
x ⇡ 1.4513692348.... This number is known as the Ramanujan-Soldner constant.

⇡(x)
pHxL

pHxL
150 LiHxL

x
ln x
100

50

200 400 600 800 1000 x x

Figure 6: The prime counting function ⇡(x) with estimates x/ ln(x) and Li(x).

In the plot above, the logarithmic integral Li(x) has been plotted along with the formula
x/ ln(x) from the prime number theorem and the prime counting function ⇡(x). As we can
see, the estimate made by the logarithmic integral is a much better estimate than the
formula from the prime number theorem. How much better can be seen in the table below
[11]:
18
Number of primes up to x Error of old estimate Error of new estimate

x
x ⇡(x) ⇡(x) Li(x) ⇡(x)
ln(x)

108 5,761,455 -332,774 754


109 50,847,534 -2,592,592 1,701
1010 455,052,511 -20,758,030 3,104
1011 4,118,054,313 -169,923,160 11,588
1012 37,607,912,018 -1,416,706,193 38,263
1013 346,065,536,839 -11,992,858,452 108,971
1014 3,204,941,750,802 -102,838,308,636 314,890
Table 1: the number of primes up to a given power of ten and corresponding error terms for the two estimates.

The logarithmic integral Li(x) is larger than ⇡(x) for small values of x , because it is not
counting primes in the same way as the prime counting function. Li(x) counts prime
n
powers, where a power p of a prime is counted as 1/n of a prime.

Although Li(x) is a better approximation of ⇡(x) than x/ ln(x) is, they are both converging
towards ⇡(x) . Li(x) does so much faster, but as x ! 1 , the ratio between the prime
counting function and both the functions Li(x) and x/ ln(x) goes towards 1. [12]:

⇡(x)
pHxL logHxL
1.2 x
x
ln(x)

1.1

1.0

pHxL
⇡(x)
0.9
liHxL
Li(x)

0.8

2000 4000 6000 8000 10 000 x


Figure 7: Convergence of the ratios of the two estimates and the prime counting function to 1 as x ! 1.

19
The gamma function - (z)
The gamma function has been an important object of study since the problem of extending
the factorial function to non-integer arguments was studied by Daniel Bernoulli and Christian
Goldbach in the 1720s. Since then, it has been studied closely by many of history's greatest
mathematicians, including Euler, Gauss, Weierstrass, Legendre and Riemann himself.

The gamma function (z) is an extension of the factorial function n!, shifted down by 1. It is
defined to be (Havil, 2003):

(z) = (z 1)!

20
(z)

15

10

-6 -4 -2 2 4 6

-5

-10

Figure 8: The gamma function (z) plotted in the range 6  z  6.

Gamma (z) is defined for (z > 0) 2 C . It is analytic everywhere except at z = 0, 1, 2, ...


The simple poles at z = k have residues of the form:
( 1)k
Res( (z), k) =
k! .
The gamma function (z) does not have any zeros [13].

20
Functional relationship
In order to study the gamma function for values of z < 0 , we must rewrite the function to a
more functional form (Havil, 2003). We begin by studying the integral
Z 1 ✓ ✓ ◆◆z 1
1
ln dt
0 t

This integral has a special place in the history of mathematics. It was studied by Euler in

1729, and proposed as a function for z in the domain z > 0 in a letter to Goldbach in 1730.

By rewriting the logarithm, we obtain:


Z 1
( ln(t))z 1 dt
0

Which is valid for z > 0 . The substitution t ! ln(t) gives the alternative form for gamma:
Z 1
(z) = tz 1 e t dt
0

So, e.g for z = 1 :


Z 1
⇥ ⇤1
(1) = e t dt = e t 0 = 1
0

And more generally, for z = (z + 1), by using integration by parts:


Z 1 Z 1
z t
⇥ z t ⇤1
(z + 1) = t e dt = t e 0 + z tz 1 e t dt = z (z)
0 0

(z + 1) = z (z)

This is called the functional relationship, which can be used to extend (z) beyond z > 0 .

To find values for z < 0 , we rewrite the functional relationship to:


(z + 1)
(z) =
z
And using this identity, we can obtain values for z < 0 , e.g:
1 1
( )= 2 ( )
2 2
Of course, because of the vertical asymptote at z = 0 , it doesn’t give values for the negative
integers z 2 Z , the simple poles of gamma [13].

Relationship with the factorial function


The functional relationship (z + 1) = z (z) can also be used to show the relationship
between the above integral (Havil, 2003):

21
Z 1 ✓ ✓ ◆◆z 1
1
ln dt
0 t and the factorial function (z) = (z 1)!:

If we make x = n a positive integer, n 2 N , then (n) = (n 1) . And so;


(n) = (n 1) (n 1)
(n) = (n 1)(n 2) (n 2)
(n) = (n 1)(n 2)(n 3) (n 3)
And more generally:
(n) = (n 1)!
Which we have seen to be the factorial function, shifted down by one [13].

A more useful form


Another connection is made from a definition proposed by Euler in 1729 (Havil, 2003):
r!rz
(z) = lim r (z) r (z) =
r!1 where z(1 + z)(2 + z) · · · (r + z)

If we divide by r! , we find a more useful form for (z) :

rz
r (z) =
z(1 + z)(2 + z2 ) · · · (r + zr )
From this, we can recover the original definition:
r!rz+1
r (z + 1) =
(z + 1)(z + 2) · · · (z + r)(z + 1 + r)
✓ ◆
r
r (z + 1) = z r (z)
z+r+1

Using the definition by Euler,


✓ ◆
r
(z + 1) = lim r (z + 1) lim z r (z)
r!1 which is
r!1 z+r+1

As r ! 1 , (z + 1) = z (z), the functional relationship [13].

Relationship with the gamma constant


The link between the constant gamma and the function gamma was made by
Weierstrass, who rewrote the definition by Euler (Havil, 2003);
rz
r (z) =
z(1 + z) 2 + z2 · · · r + z
r

22
ez ln r
r (z) =
z(1 + z) 2 + z2 · · · r + z
r

Breaking up the numerator:


1 1 1 1 z z z
···
e(ln r 1 2 2 3 r)e(z+ 2 + 3 +···+ r )
r (z) = z
z(1 + z) 2 + 2 · · · r + zr

And further rewriting the right hand side:


✓ ◆✓ z ◆ z
! z
! z
!
z (ln r 1 12 13 ··· r1 ) 1 e e2 e3 er
e ···
z 1+z 2 + z2 3 + z3 r + zr
✓ ◆ ! ! !
z (1+ 12 + 13 +···+ r1 ln r ) z z z
e ez e2 e3 er
···
z 1+z 2 + z2 3 + z3 r + zr

Here we see the gamma constant appearing:


✓ z ◆✓ z ◆ ✓ z ◆
e z e e2 er
z ···
z 1+z 1+ 2 1 + zr

Simplified into a product:


1 ⇣
e z Y z⌘ 1 z
r (z) = 1+ er
z r=1
r
Which means:
1 ⇣
1 1 z
Y z⌘ z
= lim = ze 1+ e r
(z) r!1 r (z) r=1
r

If we take log of both sides of this expression:


1 ⇣ ⇣
X z⌘ z⌘
ln (z) = ln z + z + ln 1 +
r=1
r r

Then differentiate both sides with respect to z :


0
X1 ✓ 1 ◆
(z) 1 r 1
= + z +
(z) z r=1
1 + r r
0
X1 ✓ ◆
z 1 1 1
= +
(z) z r=1
r r+z

We now define the digamma function (z) to be:


0
(z)
(z) =
(z)
Which gives us:
1 ✓
X ◆
1 1 1
(z) = +
z r=1
r r+z
23
Evaluated at z = 1:
0
(1)
=
1
Which means that the gradient of (z) at z = 1 is .
At z = 1 , the gamma function (z) decreases by a rate of [13].

Zeta and Gamma


The link between the function gamma (z) and the zeta function ⇣(s) can be found by

going back to Euler’s integral definition of (z) (Havil, 2003):


Z 1
(z) = tz 1 e t dt
0 for z > 0
Then setting t = ru :
Z 1
(z) = (ru)z 1
e ru
rdu
0
Z 1
z
(z) = r uz 1
e ru
du
0
Z 1
1 1
= uz 1
e ru
du
rz (z) 0

And we see zeta appearing on the left hand side:


X1 1 Z
1 1 X 1 z 1 ru
⇣(z) = = u e du
r=1
rz (z) r=1 0
Z 1 1
X
1 z 1 ru
⇣(z) = u e du
(z) 0 r=1

Then taking the sum of the infinite geometric series gives:


Z 1
1 e u
⇣(z) = uz 1 du
(z) 0 1 e u
Which finally makes
Z 1
uz 1
⇣(z) (z) = du
0 eu 1 with z 2
/ {1, 0, 1, 2, ..} [13].

“.. Each generation has found something of interest to say about the gamma function.

Perhaps the next generation will also..”

Philip J. Davis

24
Bernhard Riemann
German mathematician Bernhard Riemann was
born in Breselenz in 1826. Student of one of the
greatest mathematicians of all time Carl Friedrich
Gauss, Riemann mainly published works in the
areas of analysis and geometry.

In analysis, his work includes the well known


Riemann integral by means of Riemann sums.

Riemann’s major contribution was in the field of


differential geometry, where he laid the groundwork
for the geometric language used in Einstein’s general theory of relativity.

Through his fruitful (but relatively short) career, he only published one paper in the field of
number theory. This paper, Ueber die Anzahl der Primzahlen unter einer gegebenen Grösse,
“On the number of primes less than a given magnitude” was published in 1859. It is to this
day considered the most important paper in the field of prime number theory. In it, Riemann
shows functions, definitions and relationships previously unknown. Specifically;

• Definition of the Riemann zeta function ⇣(s) , where s 2 C .


• The analytic continuation of zeta ⇣(s) to all complex numbers s 6= 1 .
• Definition of the Riemann Xi function ⇠(s) , an entire function related to the Riemann
zeta function ⇣(s) through the gamma function (s) .
• Two proofs of the functional equation of the Riemann zeta function ⇣(s) .
• Definition of the Riemann prime counting function J(x) by the prime counting function
⇡(x) and the Möbius function µ(x).
• An explicit formula for the number of primes less than a given number x by the
Riemann prime counting function J(x) , defined using the non-trivial zeros of the
Riemann zeta function ⇣(s) .

25
The Riemann Zeta Function - ⇣(s)
The prime numbers and their intimate connection to the zeta function was shown in the Euler
product formula from 1737. Beyond this connection, not much was known about this
relationship, and it took the invention of complex numbers to show explicitly just how
interconnected the two are.

Riemann was the first to consider the zeta function ⇣(s) for a complex variable s 2 C , where
s= + it ( , t 2 R).
X1
1 1 1 1 1
s
= 1 + s + s + s + s + ···
n=1
n 1 2 3 4

This zeta function is now called the Riemann zeta function. It is an infinite series which is
analytic for all complex numbers s 2 C with real part <(s) > 1 , where it converges
absolutely (PlanetMath, 2007).

Proof that the Riemann zeta function converges absolutely when <(s) > 1 [14]:
Let s = + it where , t 2 R . Then
1 1 1 1
s
= s ln(n) = ln(n) =
n |e | e n
We can prove the convergence of this series by an integral test:
1
X Z 1
1 1
= dx
n=1
n 1 x

Which converges when > 1 by the p-test. So


Z 1  t 
1 +1 1 +1 1
x dx = lim x = lim t
1 t!1 +1 1
t!1 ( + 1) ( + 1)
+1
Since the term t < 1 diverges when t > 0 and converges for t  0 , the integral
converges for + 1 < 0 , i.e for > 1. It diverges for < 1 and = 1 (the harmonic
series). We have seen that the Riemann zeta function converges absolutely when > 1.

In order to analyze the zeta function ⇣(s) for values beyond the regular area of convergence
<(s) > 1, we need a new definition for zeta that is analytic in the whole complex plane C.

Euler was the first to propose a conditionally (depending on order of summation) convergent
analytic continuation of the zeta function (real valued) beyond the regular area of
convergence, to <(s) > 0 (Derbyshire, 2004):
26
Euler’s analytic continuation of the zeta function ⇣(s) [16]:
To get a value for ⇣(s) between 0 < <(s)  1, we define a new function based on the zeta
function, now called the Dirichlet eta function ⌘(s), or “the alternating zeta function”:
✓ ◆
1 1 1 1 1
⌘(s) = 1 + s + s + ···
2s 3 4s 5 6s
We can rewrite this:
✓ ◆ ✓ ◆
1 1 1 1 1 1
⌘(s) = 1 + s + s + s + · · · 2 s
+ s + s + ···
2 3 4 2 4 6
We see two zeta functions appearing:
✓ ◆ ✓ ◆
1 1 1 1 1 1 1
⌘(s) = 1 + s + s + s + · · · 2⇥ s 1 + s + s + s + ···
2 3 4 2 2 3 4

Rewritten, the eta function ⌘(s) is defined as:


✓ ◆
1
⌘(s) = 1 2 · s ⇣(s)
2

And defined for zeta ⇣(s) :


⌘(s)
⇣(s) =
1 21 s

The eta function ⌘(s) is conditionally convergent for 8(s) 2 C with <(s) > 0 .

In his 1859 paper, Riemann went further, by analytically continuing the zeta function ⇣(s) to
an absolutely convergent function in the half plane <(s) > 0 (Tenenbaum and France,
1997):

Riemann’s analytic continuation of ⇣(s) to the half plane <(s) > 0 [17]:
We rewrite the Riemann zeta function using Abel’s lemma (summation by parts):
1
X 1
X ✓ ◆
1 1 1
⇣(s) = = n
ns ns (n + 1)s
n 1 n 1

We then express this sum as an integral


1
X Z n+1
1
⇣(s) = s n dx
n xs+1
n 1

If x 2 (n, n + 1), then n = [x], the integer part of x .


Z 1 Z 1 Z 1
[x] 1 {x}
⇣(s) = s s+1
dx = s s
dx dx
1 x 1 x 1 xs+1

Where [x] denotes the integer part of x and {x} denotes the fractional part of x. And so:
27
Z 1
s {x}
⇣(s) = s dx
s 1 1 xs+1 Where {x} = x [x].

This rewritten Riemann zeta function converges absolutely for <(s) > 0 because
{x} 2 [0, 1). The integral on the right hand side defines a continuation of ⇣(s) to the entire

half plane <(s) > 0 , minus a simple pole at s = 1 , defined by the fraction s/(s 1) .

This new definition for the Riemann zeta function ⇣(s) , is analytic in the half plane <(s) > 0 ,
minus a simple pole at s = 1 . This makes ⇣(s) a meromorphic function in this domain
because it is holomorphic (complex differentiable in a neighborhood of every point in its
domain), except a single isolated point s = 1 .

The Riemann zeta function is the prototypical example of a Dirichlet series. More generally, it
is a Dirichlet L-function, which is a function of the form
X1
(n)
L(s, ) =
ns (n) is a Dirichlet character and s 2 C with s = + it and
n=1 where
> 1 . It is an L-function because it satisfies this form and can be extended to a
meromorphic function on the whole complex plane.

To make the function analytic in the entire complex plane C , Riemann rewrites the zeta
function further. He begins by studying the gamma function (z) (Borwein, et al., 2008):

Analytical continuation of ⇣(s) to the entire complex plane C [18][19]


To analytically continue ⇣(s) to the entire complex plane C , we can begin by recalling the
definition of gamma (z) , for > 0:
Z 1
(s) = ts 1 e t dt
0

We then set s = s/2 :


Z 1
s s
( )= t 2 1 e t dt
2 0

And substitute for t = n2 ⇡x:


Z 1
s s
n2 ⇡x
( )= (n2 ⇡x) 2 1 e d(n2 ⇡x)
2 0
Z 1
s 2 s s
n2 ⇡x
( ) = (n ⇡) 2 x( 2 1)
e dx
2 0
s
2
Divide both sides by (n ⇡) 2 :

28
Z 1
s s
s s
n2 ⇡x
( )⇡ 2 n = x( 2 1)
e dx
2 0
s
We see zeta ⇣(s) = n has appeared on the left hand side:
Z 1
s s s 2
⇡ 2 ( )⇣(s) = x( 2 1) e n ⇡x dx
2 0

Sum over n 1 and obtain:


Z 1 1
X
s s s
n2 ⇡x
⇡ 2 ( )⇣(s) = x( 2 1)
e dx
2 0 n 1

We then define the sum (x) to be:


1
X 2
(x) = e n ⇡x
n 1

And using this definition we write the above equation simpler:


Z 1
s s s
⇡ 2 ( )⇣(s) = x( 2 1) (x)dx
2 0

We now wish to split the integral on the right hand side into two parts:
Z 1 Z 1 ✓ ◆
s dx 1 s dx
(x)x 2 x 2
1 x 1 x x

We then introduce the Jacobi theta function #(x) (notice the difference from (x) in the
index of summation):
1
X 2
#(x) = e n ⇡x
n2Z and the relationship #(x) = 1 + 2 (x)
Which is true because
1
X 1
X
n2 ⇡x n2 ⇡x
e =1+2 e
n2Z n 1

The functional equation for theta #(x) is:


✓ ◆
1 1
#(x) = p #
x x (this relationship is shown in Appendix A)

The functional equation, combined with the definition for theta #(x), can be rewritten to:
✓ ◆
1 + 2 (x) 1 1 p
1 =
p # = x#(x)
1 + 2 (x) x because x .
Solving for (1/x):
✓ ◆ p
1 x p 1
= + x (x)
x 2 2

29
We now have an expression for (1/x) that we can use in second part of the split integral:
Z 1 Z 1 ✓ ◆
s dx 1 s dx
(x)x 2 x 2
1 x 1 x x
Z 1 Z 1 ✓p ◆
s dx x p 1 s dx
(x)x 2 + + x (x) x 2
1 x 1 2 2 x
Riemann joins these two integrals
Z 1 s 1 s
s s 1 x 2+2 x 2dx
(x)x 2 + (x)x 2 + 2 +
1 2 2 x
Z h i
1
s 1 s 1h 1 s s
i dx
(x) x + x 2 2 + x 2 x 2

1 2 x

And then split them again, with the (x) terms in one integral, the rest in the other:
Z 1 h s i dx 1 Z 1 h i dx
1 s s 1 s
(x) x 2 + x 2 + x ( 2 ) x 2
1 x 2 1 x
It is known that
Z 1
dx 1
x a =
1 x a for a > 0 .
Using this fact, the second integral becomes:
Z  
1 1 h (s 1) s
i dx 1 1 1 1 2 2 1
x 2 x 2 = s 1 s = =
2 1 x 2 2 2 2 s 1 s s(s 1) for s > 1.
And so, for s > 1 , the formula
Z 1
( 2s ) s s
⇡ ( )⇣(s) = x( 2 1)
(x)dx
2 0

Becomes:
Z 1h i
s
( 2s ) 1 s
1
(s+1)
⇡ ( )⇣(s) = + x 2 +x 2 (x)dx
2 s(1 s) 1
s ⇢ Z 1h i
⇡2 1 s
1 ( s+1 )
⇣(s) = + x 2 +x 2 (x)dx
( 2s ) s(1 s) 1

The fraction in the braces correspond to two poles, of ⇣(s) at s = 1 , and of (s) at s = 0 .

Analyzing this equation, we notice that the term for (x) in the integral;
1
X 2
(x) = e n ⇡x
n 1

30
decreases more rapidly than any power of x , so the integral converges for all values of s .
Since both sides are analytic, the same equation holds for all s, and we have shown one
form of what is known as the functional equation for the Riemann zeta function ⇣(s) .

Riemann noticed that the term 1/s(1 s) and the integral are invariant (unchanged) under
substitution of s by 1 s. If we do the substitution 1 s ! s:
✓ ◆
s s 1 s 1 s
⇡ 2 ( )⇣(s) = ⇡ 2 ⇣(1 s)
2 2
And then multiply the equation
Z 1 h i
( 2s ) s 1 s
1
(s+1)
⇡ ( )⇣(s) = + x 2 +x 2 (x)dx
2 s(1 s) 1

By s(s 1)/2 on both sides (removing the two poles s = 1 and s = 0), the right hand side
becomes 1, and we can define the Riemann Xi function with no singularities:
s s s
⇠(s) = (s 1)⇡ 2 ( )⇣(s)
2 2
Because this is an entire function, we can now define the functional equation of ⇣(s) for any
s 2 C to be:
⇠(s) = ⇠(1 s).

Riemann Zeta Function definitions


We have shown five versions of the Riemann zeta function. They apply in different areas of
the complex plane, and are all useful in studying prime numbers. They are:

The original definition for all complex numbers s 2 C with real part <(s) > 1 :
X1
1 1 1 1 1
⇣(s) = s
= 1 + s + s + s + s + ···
n=1
n 2 3 4 5

Euler’s product formula for all complex numbers s 2 C with real part <(s) > 1:
X 1 Y 1
=
n
ns p
1 p s

Riemann’s analytic continuation to the half plane <(s) > 0 :


Z 1
s {x}
⇣(s) = s dx
s 1 1 xs+1
The functional equation for the entire complex plane C, with singularities at s = 1 and s = 0 :

31
s ⇢ Z 1 h i
⇡2 1 s
1 ( s+1
⇣(s) = + x2 +x 2 ) (x)dx
( 2s ) s(1 s) 1

The Riemann Xi function (the so-called “completed” Riemann zeta function) for C :
s s s
⇠(s) = (s 1)⇡ 2 ( )⇣(s)
2 2

Two specific points


Finally we will look at two specific values of ⇣(s) , which are the boundary points between the
three areas that are of importance to us in the complex plane.

For <(s) = 1 , ⇣(s) has a simple pole with residue 1.


To find the residue, we use one of the alternative expressions for ⇣(s) :
Z 1
s
⇣(s) = s {x}x s 1 dx
s 1 1

The residue formula for the simple pole then becomes:


 Z 1
s
Res(⇣(s), 1) = lim (s 1) s {x}x s 1 dx
s!1 s 1 1

Which multiplies out to:


 Z 1
s 1
Res(⇣(s), 1) = lim s s(s 1) {x}x dx
s!1 1

Clearly, when s ! 1, the integral becomes 0, and Res(⇣(s), 1) = 1 .

For <(s) = 0 we have to use two of the alternative formulas for ⇣(s) [18].
First, the rewritten formula for ⇣(s) , valid for all <(s) > 0 :
Z 1
s
⇣(s) = s {x}x s 1 dx
s 1 1

Second, a variation on the functional equation for ⇣(s) :


⇣ ⇡s ⌘
⇣(s) = 2s ⇡ s 1 sin (1 s)⇣(1 s)
2

This variation arises from two properties of the gamma function (s) [20]:
⇣s⌘ ✓s 1◆ 1 ⇡s
s
(s) = 2 ⇡ 2 = sin ⇡s
2 2 and (s) (1 s) .

32
As we have seen, the only pole of ⇣(s) occurs at s = 1 . This means that ⇣(s) is analytic and
continuous at s = 0 . We will approach this point along any path contained in the region
<(s) > 0 (Edwards, 1974):

⇣(0) = lim+ ⇣(s)


s!0

We do this with the functional equation:


⇡s
⇣(0) = lim 2s ⇡ s 1 sin (1 s)⇣(1 s)
s!0+ 2
Inserting the Taylor series for sine and the rewritten formula for ⇣(s) with s = 1 s:
! ✓ Z ◆
( 1)n ⇣ ⇡s ⌘2n+1
1
X 1
s s 1 1 x [x]
⇣(0) = lim+ 2 ⇡ (1 s) +1 (1 s) dx
s!0
n=0
(2n + 1)! 2 (1 s) 1 1 x(1 s)+1

Removing the additional term in the exponent of the sine sum:


! ✓ Z ◆
⇣ ⇡s ⌘ X 1
( 1) n ⇣
⇡s ⌘2n 1 1
x [x]
s s 1
⇣(0) = lim+ 2 ⇡ (1 s) +1 (1 s) dx
s!0 2 n=0
(2n + 1)! 2 s 1 x2 s

Taking the first term of the sine sum out:


! ✓ Z ◆
⇣⇡⌘ X1
( 1)n ⇣ ⇡s ⌘2n 1 1
x [x]
s s 1
⇣(0) = lim+ 2 ⇡ 1+ (1 s)s +1 (1 s) dx
s!0 2 n=1
(2n + 1)! 2 s 1 x2 s

Adding the additional term of ⇡/2 to the exponent of ⇡ s 1 and 2s :


! ✓ Z ◆
X1 n ⇣ ⌘2n 1
s 1 s ( 1) ⇡s 1 x [x]
⇣(0) = lim 2 ⇡ 1 + (1 s)s +1 (1 s) dx
s!0+
n=0
(2n + 1)! 2 s 1 x(2 s)

Multiplying and separating:


✓ ✓ Z ◆◆ !
( 1)n ⇣ ⇡s ⌘2n
1 X1
s 1 s x [x]
⇣(0) = lim 2 ⇡ (1 s) 1+s s(1 s) dx lim 1 +
s!0+ 1 x2 s s!0+
n=1
(2n + 1)! 2

We are now ready to set s = 0 and approach the limits:


✓ ✓ Z 1 ◆◆ X1 n
✓ ◆2n !
x [x] ( 1) ⇡ · 0
⇣(0) = 20 1 ⇡ 0 (1 0) 1 + 0 0(1 0) dx 1+
1 x2 0 n=1
(2n + 1)! 2
✓ ◆ 1
!
1 X
⇣(0) = · 1 · (1) · ( 1 + 0 0) 1+ 0 1
2 ⇣(0) =
n=1 which finally gives 2.

33
Zeros of the Riemann zeta function
The roots/zeros of the zeta function, when ⇣(s) = 0 can be divided into two general types,
termed the trivial and the non-trivial zeros. To study the zeros of the function beyond the
regular area of convergence, we must study the analytical continuation of the function. First
we outline some ideas about where the zeros can appear.

Existence of zeros with real part <(s) < 0 [19]:


We begin with the variation of the functional equation for ⇣(s) (Borwein, et al., 2008):
⇣ ⇡s ⌘
s s 1
⇣(s) = 2 ⇡ sin (1 s)⇣(1 s)
2
The sine part of the equation is zero at k⇡ . If s = 2n , a negative even integer, then the
sine part of the above equation becomes zero. If s = 2n , a positive even integer, then the
zeros of the function are cancelled out by the poles of the gamma function. This is easier to
see in the other form of functional equation for ⇣(s) :
s ⇢ Z 1 h i
⇡2 1 s
1 ( s+1
⇣(s) = + x 2 +x 2 ) (x)dx
( 2s ) s(1 s) 1

and so, zeta ⇣(s) does not have zeros for s = 2n . The zeros at the negative even integers
s= 2n are called the trivial zeros of the Riemann zeta function, and they are the only
zeros with real part <(s) < 0 . They are called trivial because their existence and properties
are well known.

⇣(s)

Figure 9: The zeta function ⇣(s) plotted, with the trivial zeros at s = 2, 4, ... highlighted.
34
Existence of zeros with real part <(s) > 1 [21]:
From the Euler product representation for ⇣(s),
Y 1
⇣(s) =
1 p s
p for <(s) > 1 :
We immediately see that ⇣(s) cannot be zero in the given plane because a convergent
infinite product can be zero only if one of its factors is zero. Our proof that the primes are
infinite in number denies this.

Existence of zeros with real part 0  <(s)  1 :


This area, called the critical strip, is where much of the focus of analytic number theory
takes place. We have seen that at <(s) = 0 , the zeta function is equal to 1/2 . At
<(s) = 1, we found that the zeta function has a simple pole with residue 1.

zHzL = The first non-


60
trivial zeros:

14.134725142...
50 21.022039639...
25.010857580...
30.424876126...
32.935061588...
40
37.586178159...
40.918719012...
43.327073281...
30
48.005150881...
49.773832478...
y

52.970321478...
56.446247697...
20 59.347044003...

10

0
<

-5 -4 -3 -2 -1 0 1 2
x

Figure 10: The complex parts of the zeta function ⇣(s) plotted, with zeros highlighted.

In the plot above, the real parts <(s) of ⇣(s) are graphed in red, and the imaginary parts
=(s) are in blue. We see the first two trivial zeros appear at s = 2, 4, ... as intersections

35
between the two graphs on the horizontal real axis. We see the critical strip shaded red in
the interval 0  <(s)  1, and the non-trivial zeros of ⇣(s) as intersections between the real
and imaginary parts of ⇣(s) in the first quadrant. Below is the same plot, but for larger
imaginary values, showing more non-trivial zeros on a line (the critical line).

zHzL
120

100

80

60
y

40

20

-5 -4 -3 -2 -1 0 1 2
x

Figure 11: The complex parts of the zeta function ⇣(s) plotted, with zeros highlighted.

The Riemann Xi Function


We have defined the Riemann Xi function to be:
s s s
⇠(s) = (s 1)⇡ 2 ( )⇣(s)
2 2

36
This function is entire, meaning that it has values for all s 2 C . ⇠(s) satisfies the relationship:
⇠(s) = ⇠(1 s)

Which means that the function is symmetric about the line <(s) = 1/2, so that ⇠(1) = ⇠(0),
⇠(2) = ⇠( 1), ⇠(3) = ⇠( 2) and so on.

The functional relationship (the symmetry of s and 1 s ) combined with the Euler product
formula shows that the Riemann Xi function ⇠(s) can only have zeros in the range
0  <(s)  1 [21]. This means that the non-trivial zeros of the Riemann zeta function ⇣(s)

correspond to the zeros of the Riemann Xi function ⇠(s) . In a sense, the critical line
<(s) = 1/2 for the Riemann zeta function ⇣(s) corresponds to the real line <(s) for the

Riemann Xi function ⇠(s) .

Riemann went on to write that the number of zeros ⇢ whose imaginary part lies between 0
and T on the critical line is approximately equal to
T T T
log
2⇡ 2⇡ 2⇡
And that the relative error of this approximation is in the order of magnitude of 1/T .
Riemann sketched a proof and assumed that this calculation was trivial to the reader. It
wasn’t until 1905 that this approximation was proven true by von Mangolt [21].

The phenomenon which Riemann observed while studying the Riemann zeta function ⇣(s)
and its analytic continuation, was that all the zeros of the function in the critical strip appear
to line up on the vertical critical line <(s) = 1/2. He briefly remarked on this in his paper, a
fleeting comment that would be one of his greatest legacies.

37
The Riemann Hypothesis

The non-trivial zeros of the Riemann zeta function ⇣(s) have real part <(s) = 1/2.

This is the modern formulation of the unproven conjecture made by Riemann in his famous
paper. In words, it states that the points at which ⇣(s) = 0 , in the critical strip 0  <(s)  1 ,
all have real part = 1/2 for s = + it . If true, every non-trivial zero will be of the form
✓ ◆
1
⇣ + it =0
2
An equivalent statement (Riemann’s actual statement) is that all the roots of the Riemann Xi
function ⇠(s) are real.

In the plot below, the line <(s) = 1/2 is the horizontal axis. The real part <(s) of ⇣(s) is the
red graph, and the imaginary part =(s) is the blue graph. The non-trivial zeros of the
Riemann zeta function are the intersections between the two graphs on the horizontal line.

<(s)
1

t
10 20 30 40 50 60

=(s)
-1

-2

Figure 12: The first non-trivial zeros of the Riemann zeta function on the line <(s) = 1/2.

If the Riemann hypothesis is correct, all the non-trivial zeros of ⇣(s) will appear as
intersections between the real and imaginary parts of ⇣(s) on the line <(s) = 1/2 .
38
Reasons to believe the hypothesis [22]
There are many reasons to believe the truth of Riemann’s hypothesis about the zeros of the
zeta function. Perhaps the most compelling reason for mathematicians is the consequences
it would have for the distribution of prime numbers. The numerical verification of the
hypothesis to very high values suggests its truth. In fact, the numerical evidence for the
hypothesis is far strong enough to be regarded as experimentally verified in other fields
such as physics and chemistry. However, the history of mathematics contains several
conjectures that had been shown numerically to very high values and still were proven false.
Derbyshire (2004) tells the story of the Skewes number, a very very large number that gave
an upper bound, proving the falsity of one of Gauss’ conjectures that the logarithmic integral
Li(x) is always greater than the prime counting function ⇡(x). It was disproved by
Littlewood without an example, and then shown to must fail above Skewes’ very, very large
number number;
1034
1010
showing that even though Gauss’ idea had been proven to be wrong, an example of exactly
where is far beyond the reach of numerical calculation even today. This could also be the
case for Riemann’s hypothesis, which has “only” been verified up to 1012.

Independent of the truth or falsity of the hypothesis, for such conjectures it is important to
remember what Gödel showed in his incompleteness theorem about the limitations of
axiomatic systems. Effectively Gödel showed that given any set of axioms in a system, there
will always be statements about the system that are true, but that are unprovable within the
system. So even though the Riemann hypothesis is true or false, it might not be possible to
prove it within the axioms of current mathematical theory.

39
The Riemann zeta function ⇣(s) and prime numbers.
Using the truth of the Riemann hypothesis as a starting point, Riemann began studying its
consequences. In his paper he writes; "…it is very probable that all roots are real. Of course
one would wish for a rigorous proof here; I have for the time being, after some fleeting vain
attempts, provisionally put aside the search for this, as it appears dispensable for the next
objective of my investigation." His next objective was relating the zeros of the zeta function
to the prime numbers.

Recall the prime counting function ⇡(x) , which denotes the number of primes up to and
including the real number x 2 R .

We will now define the Riemann prime counting function, denoted by J(x) (Derbyshire,
2003):
1 p 1 p 1 p 1 p
J(x) = ⇡(x) + ⇡( x) + ⇡( 3 x) + ⇡( 4 x) + ⇡( 5 x) + ...
2 3 4 5 for x 2 R .
The first thing to notice about this sum is that it is not infinite. At some term, the prime
counting function will be zero because there are no primes for x < 2 . So, taking J(100) for
example, will contain seven terms, because the eight term will include an eighth root of 100,
which is approximately equal to 1.778279..., so this prime counting term becomes 0, and
the sum J(100) = 28.5333... [23].

fHxL
J(x)

15

10

10 20 30 40 x
50
x

Figure 13: A second step function: The Riemann prime counting function J(x) up to x = 50 .
40
Like the prime counting function ⇡(x), Riemann’s prime counting function J(x) is a step
function, which jumps by:
8
>
> 1 when x is a prime
<
1/2 when x is exactly the square of a prime
J(x) jumps
>
> 1/3 when x is exactly the cube of a prime
:
etc...

We can recover the prime counting function ⇡(x) from J(x) by doing Möbius inversion on it.
We can do this because the two functions are arithmetic and satisfy the relationship
X
J(n) = ⇡(d)
d|n
for every integer n 1 .
The resulting equation is
X p
J( n x)
⇡(x) = µ(n)
n
n

Remembering that the Möbius function µ(n) has the possible values:
8
< 1 if n is square-free with an even number of prime factors
µ(n) = 0 if n is not square-free
:
1 if n is square-free with an odd number of prime factors

This means that we can now write the prime counting function as:
1 1 1 1 1 1 1 1 1 1 1 1
⇡(x) = J(x) J(x 2 ) J(x 3 ) J(x 5 ) + J(x 6 ) J(x 7 ) + J(x 10 ) ···
2 3 5 6 7 10

Our new expression for ⇡(x) is still a finite sum because J(x) is zero when x < 2 , because
there are no primes less than 2.

If we now look at ⇡(100) , we get the sum


1 1 1 1
⇡(100) = J(100) J(10) J(4.64..) J(2.51..) + J(2.51..) 0 + 0.....
2 3 5 6
✓ ◆ ✓ ◆ ✓ ◆ ✓ ◆
1 1 1 1
⇡(100) = 28.5333 ⇥ 5.33... ⇥ 2.5 ⇥1 + ⇥ 1 = 25
2 3 5 6
Which we know is the number of primes less than 100.

Next we look at the Euler product formula


1 1 1 1 1
⇣(s) = ( 1 )⇥( 1 )⇥( 1 )⇥( 1 )⇥( 1 ) ⇥ ···
1 2s 1 3s 1 5s 1 7s 1 11s

Taking the logarithm of both sides yields:

41
✓ ◆ ✓ ◆ ✓ ◆ ✓ ◆
1 1 1 1
log ⇣(s) = log 1 + log 1 + log 1 + log 1 + ···
1 2s 1 3s 1 5s 1 7s

Remembering that log a ⇥ b = log a + log b .


Next we can rewrite the denominators in the parenthesis’ by log 1/a = log a :
✓ ◆ ✓ ◆ ✓ ◆ ✓ ◆
1 1 1 1
log ⇣(s) = log 1 log 1 log 1 log 1 ···
2s 3s 5s 11s
Now recall the classic Maclaurin series:
1
= 1 + x + x2 + x3 + x4 + x5 + · · ·
1 x
Which when we integrate both sides becomes:
x2 x3 x4 x5
log(1 x) = x + + ++ + + ···
2 3 4 5

Using this, we can expand each log term on the right hand side in the expression above in
the same way:
✓ ◆ ✓ ◆ ✓ ◆ ✓ ◆
1 1 1 1 1 1 1 1 1
log ⇣(s) = s + ⇥ + ⇥ + ⇥ + ⇥ + ···
2 2 22s 3 23s 4 24s 5 25s
✓ ◆ ✓ ◆ ✓ ◆ ✓ ◆
1 1 1 1 1 1 1 1 1
+ s+ ⇥ + ⇥ + ⇥ + ⇥ + ···
3 2 32s 3 33s 4 34s 5 35s
✓ ◆ ✓ ◆ ✓ ◆ ✓ ◆
1 1 1 1 1 1 1 1 1
+ s+ ⇥ + ⇥ + ⇥ + ⇥ + ···
5 2 52s 3 53s 4 54s 5 55s
✓ ◆ ✓ ◆ ✓ ◆ ✓ ◆
1 1 1 1 1 1 1 1 1
+ s+ ⇥ + ⇥ + ⇥ + ⇥ + ···
7 2 72s 3 73s 4 74s 5 75s
And so on, to create an infinite sum of infinite sums, one for each term in the prime number
series.

If we now look at one of the infinite sums, e.g:


✓ ◆
1 1

2 32s
We will try to express this term as an integral. Riemann gives us the integral:
Z ✓ ◆
s 1 x s 1 1
x dx = = ⇥ =0
s s xs as x ! 1 .
Which as a definite integral starting at the square of the prime three becomes:
Z 1  1 ✓ ◆
s 1 1 1 1 1 1 1
x dx = ⇥ s =0 ⇥ 2s = ⇥ 2s
32 s x 32 s 3 s 3

We’ve found that the infinite sum can be written as an integral:

42
✓ ◆ Z 1
1 1 1 s 1
⇥ 2s = ⇥s⇥ x dx
2 3 2 32

Every infinite sum above can be written as integrals in this same way. Each represents part
of the area under the J(x) function, and in fact our term x = 32 is where the function J(x)
takes a step of 1/2 up, denoted by the constant in front of the integral we found.
So, the height of the strips in the integral below are decided by the constants in front of the
n
integral terms, 1/n . The integrals go from the powers of the primes p to infinity. The first
term for the prime number two and the second term for the prime number three are shaded
below. When all terms from all integrals are included, the entire area under the J(x) function
is shaded. It is of course an infinite area, because the prime numbers are infinite in number.

J(x)

Z 1
1 s 1
⇥s⇥ x dx
2 32
Z 1
s 1
1⇥s⇥ x dx
2

x
Figure 14: The Riemann prime counting function J(x) up to x = 50, with two integrals highlighted.

So, each expression in the infinite sum of infinite sums can be written as integrals, making
an infinite sum of integrals. Above we’ve seen one of the terms for the prime number three,
which for s ! 1 is:
Z 1 Z 1 Z 1 Z 1
s 1 1 s 1 1 s 1 1 s 1
1⇥ x dx + ⇥ x dx + ⇥ x dx + ⇥ x dx + · · ·
3 2 32 3 33 4 34

We have seen that from the Euler product representation for ⇣(s), we can derive the
equation:

43
✓ ◆ ✓ ◆ ✓ ◆ ✓ ◆
1 1 1 1 1 1 1 1 1
log ⇣(s) = s + ⇥ + ⇥ + ⇥ + ⇥ + ···
2 2 22s 3 23s 4 24s 5 25s
✓ ◆ ✓ ◆ ✓ ◆ ✓ ◆
1 1 1 1 1 1 1 1 1
+ s+ ⇥ + ⇥ + ⇥ + ⇥ + ···
3 2 32s 3 33s 4 34s 5 35s
✓ ◆ ✓ ◆ ✓ ◆ ✓ ◆
1 1 1 1 1 1 1 1 1
+ s+ ⇥ + ⇥ + ⇥ + ⇥ + ···
5 2 52s 3 53s 4 54s 5 55s
✓ ◆ ✓ ◆ ✓ ◆ ✓ ◆
1 1 1 1 1 1 1 1 1
+ s+ ⇥ + ⇥ + ⇥ + ⇥ + ···
7 2 72s 3 73s 4 74s 5 75s
If we collect all these infinite sums together into one integral, the integral under the Riemann
prime counting function J(x), we can write it as:
Z 1
log ⇣(s) = s ⇥ J(x)x s 1 dx
0

Because this is equal to the infinite series of infinite sums above, we can simplify and write:
Z 1
1 s 1
log ⇣(s) = J(x)x dx
s 0

This equation is one of the major results of Riemann’s paper. It may be seen as an analytic
version of the Euler product formula
X 1 Y 1
s
= 1
n
n p
ps

Riemann connected zeta ⇣(s) with Riemann’s prime counting formula J(x). By extension,
the zeta function ⇣(s) had been connected to the prime counting function ⇡(x). Results in
discrete number theory were expressed in the infinitesimal language of calculus.

From this result, the prime number theorem could now be studied in the realm of complex
analysis, and was proven independently in 1896 by Jacques Hadamard and Charles de la
Vallée-Poussin. It turns out that the statement of the prime number theorem, that
x
⇡(x) ⇠
ln(x) as x ! 1 .
is the equivalent to saying that the zeta function ⇣(s) has no zeros s with <(s) = 0 or
<(s) = 1. One of the main steps of Hadamard and de la Vallée-Poussin’s proofs involved

proving that the Riemann zeta function ⇣(s) is non-zero for all complex values of s with
s = 1 + it and t > 0 . This was made possible by Riemann’s equation above [23].

44
The error term
The main result of Riemann’s 1859 paper went beyond the prime number theorem. Riemann
used his analytic function (Derbyshire, 2004):
Z 1
1 s 1
log ⇣(s) = J(x)x dx
s 0

and showed that it can be explicitly connected to the prime counting function ⇡(x) by using
a Mellin inversion, and expressing it for the Riemann prime counting function J(x). The
explicit form he found was:
1
X Z 1
⇢ 1
J(x) = Li(x) Li(x ) log 2 + dt
p x t(t2 1) log t

This is Riemann’s explicit formula. It is an improvement on the prime number theorem, a


more accurate guess at how many primes occur up to and equal to a number x . Using our
first definition for J(x);
1 p 1 p 1 p 1 p
J(x) = ⇡(x) + ⇡( x) + ⇡( 3 x) + ⇡( 4 x) + ⇡( 5 x) + ...
2 3 4 5
1
X 1
J(x) = ⇡(x n )
n=1

and its Möbius inversion, we know that


1 1 1 1 1 1 1 1 1 1 1 1
⇡(x) = J(x) J(x 2 ) J(x 3 ) J(x 5 ) + J(x 6 ) J(x 7 ) + J(x 10 ) ···
2 3 5 6 7 10
X p
J( n x)
⇡(x) = µ(n)
n
n

Which means we can get back to the prime counting function ⇡(x) from results obtained by
the explicit formula [24].

The explicit formula for J(x) has four terms.

1
X Z 1
⇢ 1
J(x) = Li(x) Li(x ) log 2 + dt
p x t(t2 1) log t

1. The first term, “the principle term”, is the logarithmic integral Li(x), which is the better
estimation on the prime counting function ⇡(x) in the prime number theorem. It is by far
the largest term, and like we have seen, an overestimate on how many primes there are
up to a given value x .

45
2. The second term, “the periodic term”, is the sum of the logarithmic integral of x to the
power ⇢ , summed over ⇢ , which are the non-trivial zeros of the Riemann zeta function
⇣(s). It is the term that adjusts overestimate of the principle term.

3. The third term is the constant log 2 = 0.693147...


4. The fourth term is an integral. It is zero for x < 2 because there are no primes smaller
than two. It has its maximum value at x = 2 , when the integral is
Z 1
1
2
dt = 0.1400101...
2 t(t 1) log t
The two latter terms are infinitesimal in their contribution to the function’s value as x ! 1 .
The main “contributors” for large numbers x are the first and second term [23].

An alternative formulation, for ⇡(x) , is:


X ⇣ ⌘
⇢ 1/2
⇡(x) = Li(x) Li(x ) + O x log x

Where the two latter terms are given by big O notation instead.

J(x) ⇡(x)

J(x)

x
Figure 15: The prime counting step function ⇡(x) being approximated by the explicit formula for the Riemann

prime counting formula J(x) using the first 35 non-trivial zeros ⇢ of the Riemann zeta function ⇣(s).

46
In the image above, we have approximated the prime counting function ⇡(x) by using the
explicit expression for the Riemann prime counting function J(x) and summed over the first
35 non-trivial zeros of the Riemann zeta function ⇣(s). We can see that the periodic term
cause the function J(x) to “resonate”, and begin to approach the shape of ⇡(x) . Below we
have used more non-trivial zeros to improve the estimate further.

⇡(x)
J(x)

J(x)

x
Figure 16: The prime counting step function ⇡(x) being approximated by the explicit formula for the Riemann

prime counting formula J(x) using the first 100 non-trivial zeros ⇢ of the Riemann zeta function ⇣(s).

Using Riemann’s explicit function, one can approximate the number of primes up to and
including a given number x to a very good accuracy. In fact, Von Koch proved in 1901 that
using the zeros of the Riemann hypothesis to error correct the logarithmic integral function is
equivalent to the “best possible” bound for the error term in the prime number theorem.

“..These zeros act like telephone poles, and the special nature of Riemann’s

zeta function dictates precisely how the wire --its graph-- must be strung
between them..”

“Stalking the Riemann Hypothesis” by Dan Rockmore.

47
Modern work
Since the death of Riemann in 1866 at the modest age of 39, his groundbreaking paper has
stood as a landmark in the field of prime- and analytic number theory. To this day Riemann’s
hypothesis about the non-trivial zeros of the Riemann zeta function remains unsolved,
despite extensive research by great mathematicians for hundreds of years. Numerous new
results and conjectures associated with the hypothesis are published each year, in the hope
that one day a proof will be tangible. Among the results are reformulations which would
imply the Riemann hypothesis. Among these reformulations, one in particular has inherent
intuitive appeal (Borwein, et al., 2008):

In his 1899 doctoral thesis, Edmund Landau made connections between a simple algebraic
function, the Liouville function, and the Riemann hypothesis [26]:

If n is a positive integer n 2 N , then the Liouville function (n) is defined as :

(n) = ( 1)!(n) Where the function !(n) is the number of (not necessarily distinct)
prime factors of n , counted with multiplicity.

So, for (2) = (3) = (5) = (7) = (8) = 1, because these numbers have an

odd number of prime factors. For (1) = (4) = (6) = (9) = (10) = 1 because
these numbers have an even number of prime factors.

The Riemann hypothesis is equivalent to the statement that for every fixed ✏ > 0,
(1) + (2) + · · · + (n)
lim 1 =0
n!1 n 2 +✏

Which translates to:


The Riemann hypothesis is equivalent to the statement that an integer has equal probability
of having an odd number or an even number of distinct prime factors [26].

This, and many other reformulations like it show the deep connections that exist between
Riemann’s hypothesis and the properties of numbers in general. Showing that Riemann was
right therefor remains one of the highest priorities of pure mathematics.

48
Appendix A - Functional equation for the Jacobi theta
function

Proof of the functional equation for the Jacobi theta function #(x) [25]:
We wish to show that the Jacobi theta function
X 2
#(x) = e n ⇡x
n2Z

satisfies the functional relationship


✓ ◆
1 1
#(x) = p #
x x .

↵x2
We begin by considering the function f (x) = e for a fixed value of ↵ > 0 .
If f (x) satisfies the Poisson summation formula
X X Z 1
f (m) = fˆ(n) fˆ(n) = f (t)e 2⇡int dt
m2Z n2Z where 1

is the Fourier transform of f , then each series converge absolutely.

We begin by finding #(↵/⇡):


X XZ 1
↵m2 ↵t2 2⇡int
#(↵/⇡) = e = e e dt
m2Z n2Z 1

Now rewrite the integral using the familiar identity for Fourier series
Z 1 Z 1
↵t2 2⇡int 2
e e dt = 2 e ↵t cos(2⇡nt)dt
1 0
p
Then set t = (x/ ↵) :
Z 1 ✓ ◆
↵( px↵ )2 nx x
2 e cos(2⇡ p )d p
0 ↵ ↵
Z 1 ✓ ◆
2 2 2⇡nx
p e x cos p dx
↵ 0 ↵
p
We further rewrite the integral, by setting y = ⇡n/ ↵ into the following equation for F (y) :
Z 1
2
F (y) = e x cos(2yx)dx
0

We then obtain the much simpler


X 2 ✓ ◆
⇡n
#(↵/⇡) = p F p
↵ ↵
n2Z

49
0
It is trivial to show that F (y) satisfies the differential equation F (y) + 2yF (y) = 0 .
A first order linear differential equation, we solve it using the method of integrating factors.
R
0 2ydy y2
For the equation F (y) + 2yF (y) = 0 , the integrating factor is µ(y) = e =e , and
2
y
the general solution becomes F (y) = Ce .

Using the fact that


Z 1 p
y2 ⇡
F (y) = e dy =
0 2
We conclude that
p
⇡ y2
F (y) = e
2 .
✓ ◆ p
⇡n ⇡ ( ⇡n 2
p )
F p = e ↵
↵ 2

From our original equation for #(↵/⇡):


XZ 1 2
#(↵/⇡) = e ↵t e 2⇡int dt
n2Z 1

We found that:
X 2 ✓ ◆
⇡n
#(↵/⇡) = p F p
↵ ↵
n2Z

Which we now know is equal to the equation


X ✓ 2 ◆ p⇡ ( p⇡n )2
#(↵/⇡) = p e ↵
↵ 2
n2Z
r
X ⇡ ( p⇡n )2
#(↵/⇡) = e ↵

n2Z

Putting in for ↵ = x⇡ reveals


r ✓ ◆ X
1 X ⇡n2 1 ⇡n2
#(x) = e x # = e x
x x
n2Z and because n2Z , we get the final form
✓ ◆
1 1
#(x) = p #
x x .
Which is our desired functional relationship for the Jacobi theta function.

50
Appendix B - Landau’s problems
Edmund Landau listed four basic problems about primes in 1912, which are now known as
Landau’s problems. Landau characterized the problems as being “unattackable at the
present state of science”, and this has remained true for the last one hundred years, as all
four problems are still unsolved.

The problems are:


i. Goldbach’s conjecture.
ii. Twin prime conjecture.
iii. Legendre’s conjecture.
iv. “Near-square prime” conjecture.

Such problems often become popularized because despite being simple to state and
understand, they are not associated with any theory or method of solution. They are simple
expressions which expose how little is known about how to handle prime numbers even in
the most basic operations of addition, subtraction, multiplication and division.

51
Goldbach’s conjecture
One of the oldest and most well-known unsolved problems in prime number theory,
Goldbach’s conjecture from the year 1742 states:

Every even integer greater than 2 can be expressed as the sum of two primes.

Simple calculations give solutions for the first integers:


4 = 2+2 6 = 3+3 8 = 3+5 10 = 3+7 12 = 5+7 and so on.

Research
One method for looking deeper into the problem is to separate the prime numbers into two
types. From the introduction of prime numbers, we remember that every prime number
p > 3 can be written in the form 6n ± 1. This gives us two types, one for primes of the form
6n 1 and one for primes of the form 6n + 1:

Type 1 with primes of the form 6n 1 = {5, 11, 17, 23, 29, ..}

Type 2 with primes of the form 6n + 1 = {7, 13, 19, 31, 37, ..}

Looking at sums of even positive integers with two terms, we have three types of sums:

1. Sums containing prime numbers exclusively of type 1.


Set 1: 10 = 5+5 16 = 5+11 22 = 5+17 28 = 5+23 and so on.
2. Sums containing prime numbers of both type 1 and type 2.
Set 2: 12 = 5+7 18 = 5+13 24 = 7+17 30 = 7+23 and so on.
3. Sums containing prime numbers exclusively of type 2.
Set 3: 14 = 7+7 20 = 7+13 26 = 13+13 32 = 13+19 and so on.

A few things to note about these three sets is that they each concern specific even integers.

Set 1 concerns even integers of the form 6n + 4 , i.e they are 2(mod6).

Set 2 concerns even integers of the form 6n + 6 , i.e they are 0(mod6).
Set 3 concerns even integers of the form 6n + 8 , i.e they are 2(mod6) .

52
Table 2: Goldbach combinations: ways to write the integers in the range 4-176 as the sum of two primes (white).
n 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36
# sums 1 2 2 3 2 4 3 4 3 4 5 5 5 5 4 5 6 6 7 6 5 7 5 8 5 8 8 6 6 7 8 9 9 9 7 8
integer 4 10 16 22 28 34 40 46 52 58 64 70 76 82 88 94 100 106 112 118 124 130 136 142 148 154 160 166 172 178 184 190 196 202 208 214
1 2 2
2 3 7 3 13 3 19 3 25 3 31 3 37 3 43 3 49 3 55 3 61 3 67 3 73 3 79 3 85 3 91 3 97 3 103 3 109 3 115 3 121 3 127 3 133 3 139 3 145 3 151 3 157 3 163 3 169 3 175 3 181 3 187 3 193 3 199 3 205 3 211
3 5 5 5 11 5 17 5 23 5 29 5 35 5 41 5 47 5 53 5 59 5 65 5 71 5 77 5 83 5 89 5 95 5 101 5 107 5 113 5 119 5 125 5 131 5 137 5 143 5 149 5 155 5 161 5 167 5 173 5 179 5 185 5 191 5 197 5 203 5 209
4 11 11 11 17 11 23 11 29 11 35 11 41 11 47 11 53 11 59 11 65 11 71 11 77 11 83 11 89 11 95 11 101 11 107 11 113 11 119 11 125 11 131 11 137 11 143 11 149 11 155 11 161 11 167 11 173 11 179 11 185 11 191 11 197 11 203
5 17 17 17 23 17 29 17 35 17 41 17 47 17 53 17 59 17 65 17 71 17 77 17 83 17 89 17 95 17 101 17 107 17 113 17 119 17 125 17 131 17 137 17 143 17 149 17 155 17 161 17 167 17 173 17 179 17 185 17 191 17 197
6 23 23 23 29 23 35 23 41 23 47 23 53 23 59 23 65 23 71 23 77 23 83 23 89 23 95 23 101 23 107 23 113 23 119 23 125 23 131 23 137 23 143 23 149 23 155 23 161 23 167 23 173 23 179 23 185 23 191
7 29 29 29 35 29 41 29 47 29 53 29 59 29 65 29 71 29 77 29 83 29 89 29 95 29 101 29 107 29 113 29 119 29 125 29 131 29 137 29 143 29 149 29 155 29 161 29 167 29 173 29 179 29 185
8 35 35 35 41 35 47 35 53 35 59 35 65 35 71 35 77 35 83 35 89 35 95 35 101 35 107 35 113 35 119 35 125 35 131 35 137 35 143 35 149 35 155 35 161 35 167 35 173 35 179
9 41 41 41 47 41 53 41 59 41 65 41 71 41 77 41 83 41 89 41 95 41 101 41 107 41 113 41 119 41 125 41 131 41 137 41 143 41 149 41 155 41 161 41 167 41 173
10 47 47 47 53 47 59 47 65 47 71 47 77 47 83 47 89 47 95 47 101 47 107 47 113 47 119 47 125 47 131 47 137 47 143 47 149 47 155 47 161 47 167
Set 1: 6n-1 = {5,11,17,23..}
11 53 53 53 59 53 65 53 71 53 77 53 83 53 89 53 95 53 101 53 107 53 113 53 119 53 125 53 131 53 137 53 143 53 149 53 155 53 161
12 59 59 59 65 59 71 59 77 59 83 59 89 59 95 59 101 59 107 59 113 59 119 59 125 59 131 59 137 59 143 59 149 59 155
13 65 65 65 71 65 77 65 83 65 89 65 95 65 101 65 107 65 113 65 119 65 125 65 131 65 137 65 143 65 149
14 71 71 71 77 71 83 71 89 71 95 71 101 71 107 71 113 71 119 71 125 71 131 71 137 71 143
15
16
17
even integers: 6n+4 (-2 mod 6) 77 77 77 83 77
83
89
83
77
83
95
89
77
83
89

101
95
89

77
83
89

107
101
95

77
83
89

113
107
101

77
83
89

119
113
107

77
83
89

125
119
113

77
83
89

131
125
119

77
83
89

137
131
125
18 95 95 95 101 95 107 95 113 95 119
19 101 101 101 107 101 113
20 107 107

Red areas contain composite numbers in the sums.


# sums 1 1 2 3 3 4 4 5 5 6 6 6 7 8 9 7 8 8 10 12 10 9 8 11 12 11 10 13 11 14 13 11 13 14 19 13
integer 6 12 18 24 30 36 42 48 54 60 66 72 78 84 90 96 102 108 114 120 126 132 138 144 150 156 162 168 174 180 186 192 198 204 210 216
1 3 3 3 9 3 15 3 21 3 27 3 33 3 39 3 45 3 51 3 57 3 63 3 69 3 75 3 81 3 87 3 93 3 99 3 105 3 111 3 117 3 123 3 129 3 135 3 141 3 147 3 153 3 159 3 165 3 171 3 177 3 183 3 189 3 195 3 201 3 207 3 213
2 7 5 7 11 7 17 7 23 7 29 7 35 7 41 7 47 7 53 7 59 7 65 7 71 7 77 7 83 7 89 7 95 7 101 7 107 7 113 7 119 7 125 7 131 7 137 7 143 7 149 7 155 7 161 7 167 7 173 7 179 7 185 7 191 7 197 7 203 7 209
3 13 5 13 11 13 17 13 23 13 29 13 35 13 41 13 47 13 53 13 59 13 65 13 71 13 77 13 83 13 89 13 95 13 101 13 107 13 113 13 119 13 125 13 131 13 137 13 143 13 149 13 155 13 161 13 167 13 173 13 179 13 185 13 191 13 197 13 203
4 19 5 19 11 19 17 19 23 19 29 19 35 19 41 19 47 19 53 19 59 19 65 19 71 19 77 19 83 19 89 19 95 19 101 19 107 19 113 19 119 19 125 19 131 19 137 19 143 19 149 19 155 19 161 19 167 19 173 19 179 19 185 19 191 19 197
5 25 5 25 11 25 17 25 23 25 29 25 35 25 41 25 47 25 53 25 59 25 65 25 71 25 77 25 83 25 89 25 95 25 101 25 107 25 113 25 119 25 125 25 131 25 137 25 143 25 149 25 155 25 161 25 167 25 173 25 179 25 185 25 191
6 31 5 31 11 31 17 31 23 31 29 31 35 31 41 31 47 31 53 31 59 31 65 31 71 31 77 31 83 31 89 31 95 31 101 31 107 31 113 31 119 31 125 31 131 31 137 31 143 31 149 31 155 31 161 31 167 31 173 31 179 31 185
7 37 5 37 11 37 17 37 23 37 29 37 35 37 41 37 47 37 53 37 59 37 65 37 71 37 77 37 83 37 89 37 95 37 101 37 107 37 113 37 119 37 125 37 131 37 137 37 143 37 149 37 155 37 161 37 167 37 173 37 179
8 43 5 43 11 43 17 43 23 43 29 43 35 43 41 43 47 43 53 43 59 43 65 43 71 43 77 43 83 43 89 43 95 43 101 43 107 43 113 43 119 43 125 43 131 43 137 43 143 43 149 43 155 43 161 43 167 43 173
9 49 5 49 11 49 17 49 23 49 29 49 35 49 41 49 47 49 53 49 59 49 65 49 71 49 77 49 83 49 89 49 95 49 101 49 107 49 113 49 119 49 125 49 131 49 137 49 143 49 149 49 155 49 161 49 167
10 55 5 55 11 55 17 55 23 55 29 55 35 55 41 55 47 55 53 55 59 55 65 55 71 55 77 55 83 55 89 55 95 55 101 55 107 55 113 55 119 55 125 55 131 55 137 55 143 55 149 55 155 55 161
11 61 5 61 11 61 17 61 23 61 29 61 35 61 41 61 47 61 53 61 59 61 65 61 71 61 77 61 83 61 89 61 95 61 101 61 107 61 113 61 119 61 125 61 131 61 137 61 143 61 149 61 155
12 67 5 67 11 67 17 67 23 67 29 67 35 67 41 67 47 67 53 67 59 67 65 67 71 67 77 67 83 67 89 67 95 67 101 67 107 67 113 67 119 67 125 67 131 67 137 67 143 67 149
13 73 5 73 11 73 17 73 23 73 29 73 35 73 41 73 47 73 53 73 59 73 65 73 71 73 77 73 83 73 89 73 95 73 101 73 107 73 113 73 119 73 125 73 131 73 137 73 143
14 79 5 79 11 79 17 79 23 79 29 79 35 79 41 79 47 79 53 79 59 79 65 79 71 79 77 79 83 79 89 79 95 79 101 79 107 79 113 79 119 79 125 79 131 79 137
15 85 5 85 11 85 17 85 23 85 29 85 35 85 41 85 47 85 53 85 59 85 65 85 71 85 77 85 83 85 89 85 95 85 101 85 107 85 113 85 119 85 125 85 131
16 91 5 91 11 91 17 91 23 91 29 91 35 91 41 91 47 91 53 91 59 91 65 91 71 91 77 91 83 91 89 91 95 91 101 91 107 91 113 91 119 91 125

53
Set 2: 6n+/-1 = {5,7,11,13,17,19..}
17 97 5 97 11 97 17 97 23 97 29 97 35 97 41 97 47 97 53 97 59 97 65 97 71 97 77 97 83 97 89 97 95 97 101 97 107 97 113 97 119
18 103 5 103 11 103 17 103 23 103 29 103 35 103 41 103 47 103 53 103 59 103 65 103 71 103 77 103 83 103 89 103 95 103 101 103 107 103 113
19 109 5 109 11 109 17 109 23 109 29 109 35 109 41 109 47 109 53 109 59 109 65 109 71 109 77 109 83 109 89 109 95 109 101 109 107
20 115 5 115 11 115 17 115 23 115 29 115 35 115 41 115 47 115 53 115 59 115 65 115 71 115 77 115 83 115 89 115 95 115 101
even integers: 6n+6 (0 mod 6)
21 121 5 121 11 121 17 121 23 121 29 121 35 121 41 121 47 121 53 121 59 121 65 121 71 121 77 121 83 121 89 121 95
22 127 5 127 11 127 17 127 23 127 29 127 35 127 41 127 47 127 53 127 59 127 65 127 71 127 77 127 83 127 89
23 133 5 133 11 133 17 133 23 133 29 133 35 133 41 133 47 133 53 133 59 133 65 133 71 133 77 133 83
24 139 5 139 11 139 17 139 23 139 29 139 35 139 41 139 47 139 53 139 59 139 65 139 71 139 77
25 145 5 145 11 145 17 145 23 145 29 145 35 145 41 145 47 145 53 145 59 145 65 145 71
26 151 5 151 11 151 17 151 23 151 29 151 35 151 41 151 47 151 53 151 59 151 65
27 157 5 157 11 157 17 157 23 157 29 157 35 157 41 157 47 157 53 157 59
28 163 5 163 11 163 17 163 23 163 29 163 35 163 41 163 47 163 53
29 169 5 169 11 169 17 169 23 169 29 169 35 169 41 169 47
30 175 5 175 11 175 17 175 23 175 29 175 35 175 41
31 181 5 181 11 181 17 181 23 181 29 181 35
32 187 5 187 11 187 17 187 23 187 29
33 193 5 193 11 193 17 193 23
34 199 5 199 11 199 17
35 205 5 205 11
211 5
# sums 1 2 2 3 2 2 3 4 3 3 2 5 4 5 4 3 5 6 6 4 3 6 7 6 4 5 5 9 7 6 5 8 9 8 7 7
integer 8 14 20 26 32 38 44 50 56 62 68 74 80 86 92 98 104 110 116 122 128 134 140 146 152 158 164 170 176 182 188 194 200 206 212 218
1 3 5 3 11 3 17 3 23 3 29 3 35 3 41 3 47 3 53 3 59 3 65 3 71 3 77 3 83 3 89 3 95 3 101 3 107 3 113 3 119 3 125 3 131 3 137 3 143 3 149 3 155 3 161 3 167 3 173 3 179 3 185 3 191 3 197 3 203 3 209 3 215
2 7 7 7 13 7 19 7 25 7 31 7 37 7 43 7 49 7 55 7 61 7 67 7 73 7 79 7 85 7 91 7 97 7 103 7 109 7 115 7 121 7 127 7 133 7 139 7 145 7 151 7 157 7 163 7 169 7 175 7 181 7 187 7 193 7 199 7 205 7 211
3 13 13 13 19 13 25 13 31 13 37 13 43 13 49 13 55 13 61 13 67 13 73 13 79 13 85 13 91 13 97 13 103 13 109 13 115 13 121 13 127 13 133 13 139 13 145 13 151 13 157 13 163 13 169 13 175 13 181 13 187 13 193 13 199 13 205
4 19 19 19 25 19 31 19 37 19 43 19 49 19 55 19 61 19 67 19 73 19 79 19 85 19 91 19 97 19 103 19 109 19 115 19 121 19 127 19 133 19 139 19 145 19 151 19 157 19 163 19 169 19 175 19 181 19 187 19 193 19 199
5 25 25 25 31 25 37 25 43 25 49 25 55 25 61 25 67 25 73 25 79 25 85 25 91 25 97 25 103 25 109 25 115 25 121 25 127 25 133 25 139 25 145 25 151 25 157 25 163 25 169 25 175 25 181 25 187 25 193
6 31 31 31 37 31 43 31 49 31 55 31 61 31 67 31 73 31 79 31 85 31 91 31 97 31 103 31 109 31 115 31 121 31 127 31 133 31 139 31 145 31 151 31 157 31 163 31 169 31 175 31 181 31 187
7 37 37 37 43 37 49 37 55 37 61 37 67 37 73 37 79 37 85 37 91 37 97 37 103 37 109 37 115 37 121 37 127 37 133 37 139 37 145 37 151 37 157 37 163 37 169 37 175 37 181
8 43 43 43 49 43 55 43 61 43 67 43 73 43 79 43 85 43 91 43 97 43 103 43 109 43 115 43 121 43 127 43 133 43 139 43 145 43 151 43 157 43 163 43 169 43 175
9 49 49 49 55 49 61 49 67 49 73 49 79 49 85 49 91 49 97 49 103 49 109 49 115 49 121 49 127 49 133 49 139 49 145 49 151 49 157 49 163 49 169
Set 3: 6n+1 ={7,13,19,25..}
10 55 55 55 61 55 67 55 73 55 79 55 85 55 91 55 97 55 103 55 109 55 115 55 121 55 127 55 133 55 139 55 145 55 151 55 157 55 163
11 61 61 61 67 61 73 61 79 61 85 61 91 61 97 61 103 61 109 61 115 61 121 61 127 61 133 61 139 61 145 61 151 61 157
12 67 67 67 73 67 79 67 85 67 91 67 97 67 103 67 109 67 115 67 121 67 127 67 133 67 139 67 145 67 151
13 73 73 73 79 73 85 73 91 73 97 73 103 73 109 73 115 73 121 73 127 73 133 73 139 73 145
14
15
16
even integers: 6n+8 (2 mod 6) 79 79 79 85 79
85
91
85
79
85
97
91
79
85
91

103
97
91

79
85
91

109
103
97

79
85
91

115
109
103

79
85
91

121
115
109

79
85
91

127
121
115

79
85
91

133
127
121

79
85
91

139
133
127
17 97 97 97 103 97 109 97 115 97 121
18 103 103 103 109 103 115
19 109 109
20
The table on the previous page shows the three ways even integers can be written as sums
of two prime numbers. The red areas are sums with one or two composite numbers in the
sum - these are obviously not counted. We see that every number in the interval shown
(4-176) can be written in at least one way as a sum of two prime numbers. This numerically
confirms Goldbach’s conjecture in this interval.

The phenomenon that the ways to write the even integers in set 2, {6,12,18,24,30,...}
appears more numerous than for other even integers is called Goldbach’s comet. It occurs
because set 2 contains more elements than sets 1 and 3 up to and equal to a given number.
The name Goldbach’s comet comes from the appearance of the scatter plot of the number
of ways to write even integers as the sum of two primes.

Goldbach sums
20

18

16

14
Number of ways

12

10

0
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30 32 34
Even number

Set 1 {5,11,17,23,29,..} - (4,10,16,22,28,..)


Set 2 {5,7,11,13,17,19,..} - (6,12,18,24,30,..)
Set 3 {7,13,19,25,31,..} - (8,14,20,26,32,..)
Figure 17: Goldbach sums: scatter plot of number of ways to write even integers as sums of two primes.

In the scatter plot above, the number of ways to write an even integer as the sum of two
prime numbers are shown by which set they occur in. We see that even integers from set 2
have more ways than set 1 and 3 as the even integers becomes larger.

54
Twin prime conjecture
While Goldbach’s conjecture is a problem in additive number theory, the twin prime
conjecture is considered a problem in multiplicative number theory. The twin prime
conjecture states:
the twin primes

There are infinitely many twin primes. 5 7

11 13

A twin prime is a prime that differs from another prime by two. The 17 19

first twin prime pairs are: (3,5), (5,7), (11,13), (17,19), (29,31), 23 25

(41,43), (59,61), .... 29 31

35 37

41 43
Research
47 49
Once again, the problem remains unsolved because there is no
53 55
established method to deal with such conjectures, and new ideas
59 61
are required. While we know of least 6 different ways to show that
65 67
the prime numbers are infinite, there are still no known proofs to
71 73
show the infinity of twin primes.
77 79

83 85
Using a similar sorting as we saw for Goldbach’s conjecture, we
89 91
can divide the primes into two sets;
95 97

101 103
Set 1 with primes of the form 6n 1 = {5, 11, 17, 23, 29, ..}
107 109
Set 2 with primes of the form 6n + 1 = {7, 13, 19, 31, 37, ..}
113 115

Because every prime p > 3 can be written in this way, twin primes
will always contain one prime from each set. One way to visualize their distribution might be
to use the absolute value of 6n ± 1. This plot is shown below.

55
|6n ± 1|
60

50

40

30

20

10

-10 -5 5 10
n
Figure 18: Plot of function |6n ± 1| in the range 10  n  10.

Because every twin prime differs by two, the twin prime pairs will share the multiplier n in
the expression 6n ± 1 . So, for the prime pairs listed, n = 1, 2, 3, 5, 7, 10, .. we can mark the
given twin primes on the plot:

|6n ± 1|
60

50

40

30

20

10

-10 -5 5 10
n
Figure 19: Plot of function |6n ± 1| in the range 10  n  10 with twin primes highlighted.

56
Using the points on the plot, we can create linear “twin prime functions” that intersect the
function | 6n ± 1 | at each twin prime. These functions will be of the form 1/n + 6n . In the
plot below, we have inserted the functions for the first six twin prime pairs.

|6n ± 1|
60

50

40

30

20

10

-10 -5 5 10
n

Figure 20: Plot of function |6n ± 1| and the first “twin prime functions“ in the range 10  n  10 .

|6n ± 1|
600

500

400

300

200

100

0 2000 4000 6000 8000 10 000 12 000 14 000


n
Figure 21: Plot of the first 20 “twin prime functions“.

57
|6n ± 1|
3000

2500

2000

1500

1000

500

0 50 000 100 000 150 000 200 000


n
Figure 22: Plot of the first 40 “twin prime functions“.

The plot above shows the first 40 twin prime pair as linear functions of the form 1/n + 6n .
The smallest twin prime pair in the plot is (5,7) and the largest is (1229,1231). Each function
intersects the y-axis at the even number between the twin primes. The first 40 twin prime
pairs are generated by using the following values:

n = {1, 2, 3, 5, 7, 10, 12, 17, 18, 23, 25, 30, 32, 33, 38, 40, 45, 47, 52, 58, 70, 72, 77, 87,
95, 100, 103, 110, 135, 137, 138, 143, 147, 170, 172, 175, 177, 182, 192, 205}

If the twin prime conjecture is true, then this sequence is infinite in length, and the linear
“twin prime functions” will be infinite in number up the y-axis.

It might be a worthy observation to notice the “square root” shape of the negative space
under the twin prime functions in the figure above.

58
Legendre’s conjecture
Adrien-Marie Legendre proposed the conjecture

2
There is a prime number between n2 and (n + 1) for every positive integer n .

If proven true, there will always be a prime number between successive perfect squares. We
can verify Legendre’s conjecture for small perfect squares:

n = 1: 12 = 1 22 = 4 2 and 3 are primes in the interval (1,4).


n = 2: 22 = 4 32 = 9 5 and 7 are primes in the interval (4,9).
n = 3: 32 = 9 42 = 16 11 and 13 are primes in the interval (9,16).
n = 4: 42 = 16 52 = 25 17, 19 and 23 are primes in the interval (16,25).
n = 5: 52 = 25 62 = 36 29 and 31 are primes in the interval (25,36).

Research
From what we know about square numbers, the gap between two successive squares is
equal to the sum of the two square-generating numbers. So, the gap between the perfect
squares of n = 4 and n = 5 would be 4 + 5 = 9 , and the difference between them is indeed
25 16 = 9 . The gap between the squares therefor increases linearly and is given by the
familiar progression {3, 5, 7, 9, 11, 13, 15, 17, ...}, the odd numbers.

One way to numerically verify the conjecture is to look at prime gaps and square gaps. If the
prime gap in a given interval is shorter than the square gap, then there will be at least one
2
prime in the interval. For a prime gap pn+1 pn and a square gap (n + 1) n2, if the
prime gap is smaller for all values of n, p 2 N , then Legendre’s conjecture is true. Said
differently, if Legendre’s conjecture is true, then the difference between the square gap and
the prime gap always is positive.

In the images below, the prime counting function has been plotted with vertical lines
representing the perfect squares. If Legendre’s conjecture is true, there is no prime gap that
exceeds the width of the gap between two perfect squares.

59
⇡(n)
12
⇡(n)

10

10 20 30 40 n

Figure 23: The prime counting step function ⇡(n) and vertical lines representing the perfect squares.
Prime gaps appear as vertical lines in the prime counting step function. Above: the first 12 primes. Below:
the first 330 primes.

⇡(n)

300
⇡(n)

250

200

150

100

50

500 1000 1500 2000 n

60
Showing that a prime gap can never exceed the difference between two perfect prime
would prove Legendre’s conjecture. We know from the prime number theorem that prime
numbers become sparser as they become larger. Relating the rate at which this “dilution”
occurs to the linearly increasing size of the perfect squares, and showing that the latter will
always be larger would also prove Legendre’s conjecture.

At a perfect square, the prime counting step function will be horizontal, because a perfect
square cannot be a prime. Because the perfect squares oscillate between odd and even,
the length of this prime gap will also oscillate. At minimum, the prime gap at an odd perfect
square is four in length. So, for the perfect square 9, we have will have a gap interval of
minimum length of two before (7-8, 8-9) and after (9-10, 10-11). For an even perfect square,
say 16, this interval will be minimum one (15-16) before the square and one (16-17) after the
square. So, between two perfect squares, the numbers at which steps can occur is not
given by the length of the difference between them, but by this distance minus one gap of
interval length two and one gap of interval length one. So, for our primes 9 and 16, the
difference between them is 7, but the actual gap interval at which steps can occur is only
four in length (11-15). In this interval, there are obviously even numbers, and once they are
removed, the points at which steps can occur in the interval between the two perfect
squares 9 and 16 are only three in number (11,13,15), and of course we know that two steps
occur, at 11 and 13.

61
“Near-square” prime conjecture
Although it doesn’t have a commonly used name, the “near-square” prime conjecture was
featured on Landau’s list of famous unsolved problems and still remains unsolved. It states:

Are there infinitely many primes p such that p 1 is a perfect square?


Are there infinitely many primes p such that p + 1 is a perfect square?

The first few near-square prime numbers are p = 2, 5, 17, 37, 101, .... From the definition, they
are near-square primes because p 1 is a perfect square. The progression above has
squares 1, 4, 16, 36, 100, .. created by the numbers n = 1, 2, 4, 6, 10.. .

Research
From our study of Legendre’s conjecture, we can immediately eliminate certain perfect
squares as candidates for such primes. We know that every other perfect square is odd.
And odd number plus or minus one is an even number, which we know cannot be prime.
This leaves us with the even perfect squares {4, 16, 36, 64, 100, · · · } given by the numbers
{2, 4, 6, 8, 10, · · · }.

62
References

(1) The Prime Pages, 2013. Euclid’s Proof of the infinitude of Primes [Online] Available
at: <http://primes.utm.edu/notes/proofs/infinite/euclids.html> [Accessed 11.04.13].
(2) Havil, J., 2003. The harmonic series. In: Gamma. Princeton: Princeton University
Press. p. 23.
(3) Havil, J., 2003. Zeta functions. In: Gamma. Princeton: Princeton University Press. p.
37-38.
(4) Plus Magazine., 2013. An infinite series of surprises [Online] Available at: <http://
plus.maths.org/content/infinite-series-surprises> [Accessed 11.04.13]
(5) Havil, J., 2003. Gamma’s birthplace. In: Gamma. Princeton: Princeton University
Press. p. 47.
(6) Havil, J., 2003. Gamma’s birthplace. In: Gamma. Princeton: Princeton University
Press. p. 49-50.
(7) Derbyshire, J., 2004. The golden key, and an improved prime number theorem. In:
Prime Obsession. New York: Penguin Group. p. 102-105.
(8) Derbyshire, J., 2004. Big oh and Möbius Mu. In: Prime Obsession. New York:
Penguin Group. p.245-250.
(9) Weisstein, E. W., 2013. Möbius Function. From Mathworld--A Wolfram Web
Resource. [Online] Available at <http://mathworld.wolfram.com/
MoebiusFunction.html> [Accessed 3 May 2013].
(10) Hoffman, P., 1998. The Man Who Loved Only Numbers. Hyperion. p.211
(11) Derbyshire, J., 2004. The golden key, and an improved prime number theorem. In:
Prime Obsession. New York: Penguin Group. p. 116.
(12) Weisstein, E.W., 2013. Logarithmic Integral. From Mathworld--A Wolfram Web
Resource. [Online] Available at <http://mathworld.wolfram.com/
LogarithmicIntegral.html> [Accessed 3 May 2013].
(13) Havil, J., 2003. The Gamma Function. In: Gamma. Princeton: Princeton University
Press. p. 53-60
(14) PlanetMath.org, 2013.Convergence of Riemann Series. [Online] <http://
planetmath.org/convergenceofriemannzetaseries> [Accessed 4 May 2013].
(15) Sondow, J. and Weisstein, E.W, 2013. Riemann Zeta function. From Mathworld--A
Wolfram Web Resource. [Online]. Available at <http://mathworld.wolfram.com/
RiemannZetaFunction.html> [Accessed 3 May 2013].

63
(16) Derbyshire, J., 2004. Domain Stretching. In: Prime Obsession. New York: Penguin
Group. p. 145-6.
(17) Tenenbaum, G. and France, M. 1997. Les Nombres Premiers. Paris: Presses
Universitaires de France. p.33.
(18) PlanetMath.org, 2013.Value of the Riemann zeta function at s=0. [Online] <http://
planetmath.org/valueoftheriemannzetafunctionats0> [Accessed 4 May 2013].
(19) Borwein, P. Choi, S. Rooney, B. and Weriathmueller, A. 2008. The Riemann
Hypothesis. New York, Springer Science+Business Media. p.14.
(20) Edwards, H.M. 1974. Riemann’s Zeta Function. New York, Academic Pres, Inc. p.14.
(21) Edwards, H.M. 1974. Riemann’s Zeta Function. New York, Academic Pres, Inc. p.18.
(22) Chaitin, G.J. 2003. Thoughts on the Riemann Hypothesis. [Online] <http://arxiv.org/
pdf/math.HO/0306042.pdf> [Accessed 5 May 2013].
(23) Derbyshire, J., 2004. Turning the Golden Key. In: Prime Obsession. New York:
Penguin Group. p. 296-311.
(24) Havil, J., 2003. The Gamma Function. In: Gamma. Princeton: Princeton University
Press. p. 58.
(25) Menici, L. 2012. Zeros of the Riemann Zeta function on the critical line. Università
degli studi Roma TRE. p.65-66.
(26) Borwein, P. Choi, S. Rooney, B. and Weriathmueller, A. 2008. The Riemann
Hypothesis. New York, Springer Science+Business Media. p.6.
(27) Weisstein, E.W., 2013. Modular Prime Counting function. From Mathworld--A
Wolfram Web Resource. [Online] Available at <http://mathworld.wolfram.com/
ModularPrimeCountingFunction.html> [Accessed 5 May 2013].

64
Bibliography

E.T. Bell (1937)


Men of Mathematics
Peter Borwein, Stephen Choi, Brendan Rooney and Andrea Weirathmueller (2007)
The Riemann Hypothesis: A Resource for the Afficionado and Virtuoso Alike.
John Derbyshire (2004, 2007)
Prime Obsession: Bernhard Riemann and the Greatest Unsolved Problem in Mathematics
Unknown Quantity: A Real and Imaginary History of Algebra.
William Dunham (1991)
Journey through Genius: The Great Theorems of Mathematics.
H. M. Edwards (1971)
Riemann’s Zeta Function.
Julian Havil (2009)
Gamma: Exploring Euler’s Constant.
Paul Hoffman (1999)
The Man Who Loved Only Numbers.
Alexander Ivic (2004)
The Riemann Zeta-Function: Theory and Applications.
Eli Maor (2009)
e: The Story of a Number.
Paul J. Nahin (2010, 2011)
An Imaginary Tale: The Story of √-1.
Dr. Euler’s Fabulous Formula.
Dan Rockmore (2006)
Stalking the Riemann Hypothesis: The Quest to Find the Hidden Law of Prime Numbers.
Rudy Rucker (2004)
Infinity and the Mind.
Karl Sabbagh (2004)
The Riemann Hypothesis: The Greatest Unsolved Problem in Mathematics.
Gérald Tenenbaum and Michel Mendès France (2000)
The Prime Numbers and Their Distribution.

65
66
67
68

You might also like