You are on page 1of 27

A Concise Proof of the Riemann Hypothesis Based on

Hadamard Product
Fayez A. Alhargan

To cite this version:


Fayez A. Alhargan. A Concise Proof of the Riemann Hypothesis Based on Hadamard Product. 2021.
�hal-01819208v5�

HAL Id: hal-01819208


https://hal.archives-ouvertes.fr/hal-01819208v5
Preprint submitted on 7 Jul 2021

HAL is a multi-disciplinary open access L’archive ouverte pluridisciplinaire HAL, est


archive for the deposit and dissemination of sci- destinée au dépôt et à la diffusion de documents
entific research documents, whether they are pub- scientifiques de niveau recherche, publiés ou non,
lished or not. The documents may come from émanant des établissements d’enseignement et de
teaching and research institutions in France or recherche français ou étrangers, des laboratoires
abroad, or from public or private research centers. publics ou privés.
A CONCISE PROOF OF THE RIEMANN HYPOTHESIS BASED ON
HADAMARD PRODUCT∗
FAYEZ A. ALHARGAN†

Abstract. A concise proof of the Riemann Hypothesis is presented by clarifying the Hadamard
product expansion over the zeta zeros, demonstrating conclusively that the Riemann Hypothesis is
true. Then, an accurate zero-counting function exhibiting the expected step function behaviour is
developed. Also, based on the Heaviside function a unifying analysis of the prime-counting function
is presented and its relation to the zeta function is developed via the Laplace transform and the
residue theorem. Furthermore, a new s-domain definitions of the prime-counting function and the
Chebyshev function are developed, revealing a profound underlining relationship to the zeta function.
The paper encompass a new paradigm shift in the prime analysis with a fresh perspective from the
s-domain.

Key words. the Riemann Hypothesis, the functional equation, the Riemann zeta function,
Hadamard Product

AMS subject classifications. 11M26

1. Introduction. In his landmark paper in 1859, Bernhard Riemann [1] hypoth-


esized that the non-trivial zeros of the Riemann zeta function ζ(s) all have a real part
equal to 12 . Major progress towards proving the Riemann hypothesis was made by
Jacques Hadamard in 1893 [2], when he showed that the Riemann zeta function ζ(s)
can be expressed as an infinite product expansion over the non-trivial zeros of the
zeta function. In 1896 [3], he also proved that there are no zeros on the line <(s) = 1.
The Riemann Hypothesis is the eighth problem in David Hilbert’s list of 23 un-
solved problems published in 1900 [4]. There has been tremendous work on the subject
since then, which has been illustrated by Titchmarsh (1930) [5], Edwards (1975) [6],
Ivic (1985) [7], and Karatsuba (1992) [8]. It is still regarded as one of the most diffi-
cult unsolved problems and has been named the second most important problem in
the list of the Clay Mathematics Institute Millennium Prize Problems (2000), as its
proof would shed light on many of the mysteries surrounding the distribution of prime
numbers [9, 10].
The Riemann zeta function is a function of the complex variable s, defined in the
half-plane <(s) > 1 by the absolutely convergent series

X 1
(1.1) ζ(s) :=
n=1
ns

and in the whole complex plane by analytic continuation [9].


The Riemann hypothesis is concerned with the locations of the non-trivial zeros
of ζ(s), and states that: the non-trivial zeros of ζ(s) have a real part equal to 21 [9].
In this paper, the truth of the Riemann Hypothesis is demonstrated by employing
the Hadamard product of the zeta function and clarifying the principle zeros for the
product expansion. The process is outlined in a less abstract form, to be accessible
for a wider audience.
Furthermore, new concise and accurate results have been obtained for the prime-
counting function and the zero-counting function. These were achieved by utiliz-
ing new techniques; that included the Heaviside function, Dirac delta function, the
∗ Revised Version Submitted to the editors 5 March 2021.
† PSDSARC, Riyadh, Saudi Arabia (Fayez.Alhargan@psdsarc.org.sa).
1
2 F.A. ALHARGAN

Laplace transform, Mittag-Leffler’s theorem, and the residue theorem. Such tech-
niques provided concise and elegant solutions both in the x-domain and s-domain,
revealing profound underlining connections between the Heaviside prime-counting
function, the Dirac delta function, and the zeta function. Which I hope will provide
another angle to address prime computations, primality testing, and prime factoriza-
tion.
2. The Riemann Hypothesis.
2.1. Principle Zeros of the Zeta Function. For the case of the Riemann zeta
function ζ(s), it has been shown, by Riemann [1], that the zeta function satisfies the
following functional equation
 πs 
(2.1) ζ(s) = 2s π s−1 sin Γ(1 − s)ζ(1 − s),
2
where the symmetrical form of the functional equation is given as
s 1−s
(2.2) π − 2 Γ( 2s )ζ(s) = π 2 Γ( 1−s
2 )ζ(1 − s).

We note that ζ(s) has zeros at s = sm = σm + itm , s = s̄m = σm − itm , and s = −2m
with m = 1, 2, 3, . . . . Many assume, from the functional equation (2.2) for ζ(1 − s),
that s = 1 − sm and s = 1 − s̄m are also zeros of zeta. Nevertheless, the principle zeros
of ζ(s) are determined only by using the pure argument s in ζ(s); hence, the principle
zeros are only at s = sm , s = s̄m , and s = −2m. Therefore, the sums and products
of ζ(s) should only be over the zeros s = sm , s = s̄m , and s = −2m, whenever
appropriate, contrary to the usual statement that ”the infinite product is understood
to be taken in an order which pairs each root ρ with the corresponding root 1 − ρ”
[6] p.39. For clarity, I have rephrased the statement to ”the ζ(s) infinite product is
understood to be taken in an order which pairs each root sm with the corresponding
conjugate root s̄m ”; the difference is minor though the impact is tremendous.
Now, the locations of the non-trivial zeros are determined by considering the
Euler product of ζ(s) over the set of the prime numbers {2, 3, 5, . . . , pm , . . . }, given
by
Y 1
(2.3) ζ(s) = ,
p
1 − p1s

which shows that ζ(s) does not have any zeros for <(s) > 1, and by the functional
Equation (2.1), no zeros for <(s) < 0; save for the trivial zeros at s = −2m, due
to the sin( πs
2 )Γ(1 − s) term. Jacques Hadamard (1896) [3] and Charles Jean de la
Vallée-Poussin [11] independently proved that there are no zeros on the line <(s) = 1.
In addition, considering the functional equation and the fact that there are no zeros
with a real part greater than 1, it follows that all non-trivial zeros must lie in the
interior of the critical strip 0 < <(s) < 1. Hardy and Littlewood (1921) [12] have
shown that there are infinitely many non-trivial zeros sm on the critical line s = 21 +it.
We note that the non-trivial principle zeros of ζ(s) are located only in the strip
1
2 ≤ <(s) < 1, as shown in Figure (1), whereas the non-trivial zeros of ζ(1 − s) are
located in the strip 0 < <(s) ≤ 12 . Although this is a minor definition clarification, it is
critical in proving the Riemann Hypothesis. This has been overlooked, as 1−sm = s̄m
for all the known zeros; thus, the product or sum over the zeros (1 − sm ) is the same
as the product or sum over s̄m for the first ten trillion known zeros [13].
PROOF OF RIEMANN HYPOTHESIS 3

t The Critical Strip

ζ(s)

1 − s̄m sm

σ=0 σ= 1 σ=1 σ
2

1 − sm s̄m
ζ(1 − s)

Fig. 1. The Critical Strip.

2.2. Sums and Products for Zeta Function. In this section, the sum over
the principle poles of a reciprocal function of zeta is developed based on Mittag-
Leffler’s theorem, in order to showcase the linkage to the Hadamard product over the
principle zeros of zeta, by considering a normalized function of ξ(s) given by
s
(2.4) f (s) = 2ξ(s) = ζ(s)(s − 1)sΓ( 2s )π − 2 ,

which is an entire function with f (s) = f (1 − s), f (1) = f (0) = 1, and has principle
zeros only at s = sm and s = s̄m . Thus, the ζ(s) infinite product is understood to
be taken in an order which pairs each root sm with the corresponding conjugate root
s̄m . Now, taking the log, we have

(2.5) ln f (s) = ln 2 + ln ξ(s) = ln ζ(s) + ln(s − 1) + ln s + ln Γ( 2s ) − s


2 ln π.

Differentiating, we have
0 0 0 0
f (s) ξ (s) ζ (s) 1 1 Γ ( 2s ) 1
(2.6) = = + + + − ln π,
f (s) ξ(s) ζ(s) (s − 1) s Γ( 2s ) 2

which gives
0
f (0)
(2.7) = ln 2π − 1 − 12 γ − 1
2 ln π.
f (0)
0
Note that ff (s)
(s)
has simple poles at the same zeros of ξ(s) (i.e., the poles are at s = sm
and s = s̄m ).
Now, using Mittag-Leffler’s theorem for the sum over the poles of the function
0
f (s)
f (s) , we obtain
0 0
ζ (s) 1 1 Γ ( 2s )
+ + + =[ln 2π − 1 − 21 γ]
ζ(s) (s − 1) s Γ( 2s )
(2.8) ∞
X 1 1 1 1
+ + + + .
m=1
(s − s m ) sm (s − s̄m ) s̄m
4 F.A. ALHARGAN

Integrating Equation (2.8) and taking the antilog, we have


∞    
1 Y s s s s
(2.9) ζ(s)(s − 1)sΓ( 2s ) = e[ln 2π−1− 2 γ]s 1− e sm 1− e s̄m ,
m=1
sm s̄m

which was proved by Hadamard [2]. Note the 21 ln π term canceled out, as it appears
on both sides of the equation.
Also, using Mittag-Leffler’s theorem for the following function
0
f (s)
(2.10) F (s) = s =⇒ F (0) = 0,
f (s)
we have
0 0 ∞
ζ (s) 1 1 Γ ( 2s ) 1
X 1 1
(2.11) + + + − ln π = + .
ζ(s) (s − 1) s Γ( 2s ) 2
m=1
(s − sm ) (s − s̄m )

Integrating and taking the antilog, we have


∞   
s Y s s
(2.12) ζ(s)(s − 1)sΓ( 2s )π − 2 = 1− 1− ;
m=1
sm s̄m

that is,
∞  
s Y s(2σm − s)
(2.13) 2ξ(s) = ζ(s)(s − 1)sΓ( 2s )π − 2 = 1− ,
m=1
sm s̄m

which was given by Riemann [1], in a logarithmic form with minor difference from the
modern definition of ξ(s). He set s = 12 + ti to obtain
 
X tt
(2.14) log ξ(t) = log 1 − + log ξ(0);
αα
that is,
Y tt

(2.15) ξ(t) = ξ(0) 1− .
αα
2.3. A Proof of the Riemann Hypothesis.
Theorem 2.1. The Riemann zeta function ζ(s) has only two independent sets
of principle zeros, M and S. The set M of all principle trivial zeros of ζ(s) lies on
the real negative axis with imaginary part t = 0, whereas the set S of all principle
non-trivial zeros of ζ(s) lies on the imaginary line with real part σ = 21 , as shown in
Figure (2).

Proof. It has been shown, by Riemann [1], that the zeta function satisfies the
following functional equation:
 πs 
(2.16) ζ(s) = 2s π s−1 sin Γ(1 − s)ζ(1 − s),
2
Now, if ζ(s) = 0, then from Equation (2.16), we have
PROOF OF RIEMANN HYPOTHESIS 5

t
sm

−2m
1 σ
2

s̄m

Fig. 2. ζ(s) Zeros Location.

 πs 
(2.17) sin Γ(1 − s) = 0,
2
or

(2.18) ζ(s) = ζ(1 − s) = 0.

From Equation (2.17), we can obtain the set M of all trivial zeros of ζ(s) (i.e.,
M = {−2, −4, . . . , −2m, . . . }, where m is a positive integer) and, from Equation
(2.18), we can obtain another independent set S of all non-trivial zeros of ζ(s), S =
{s1 , s2 , . . . , sm , . . . }, with sm = σm ± itm , where 21 ≤ σm < 1, tm are real numbers,
and i is the imaginary unit.
Now, by Equation (2.13), we have
∞  
s −2
s Y s(2σm − s)
(2.19) 2ξ(s) = ζ(s)(s − 1)sΓ( 2 )π = 1− ,
m=1
sm s̄m

and, considering the case of the limit when s → 1, we have


∞  
1 Y (2σm − 1)
(2.20) lim [ζ(s)(s − 1)]Γ( 12 )π − 2 = 1− .
s→1
m=1
sm s̄m

It is well-known that
1
lim ζ(s)(s − 1) = 1 and Γ( 12 ) = π 2 .
s→1

Therefore, Equation (2.20) becomes


∞  
Y (2σm − 1)
(2.21) 1= 1− ,
m=1
sm s̄m

and since
1
(2.22) 2 ≤ σm < 1

for all the principle non-trivial zeros (sm = σm ± itm ) of ζ(s), it implies that

(2.23) 0 ≤ (2σm − 1) < 1.


6 F.A. ALHARGAN

Therefore, Equation (2.21) is true only when (2σm − 1) = 0, which requires that
σm = 21 for all the non-trivial zeros of ζ(s). This concludes the proof of the Riemann
Hypothesis that: the real part of every non-trivial zero of the Riemann zeta function
is σm = 12 .
Also, the proof can be stated in a concise form as
(2.24)
∞  
s
s
−2
Y s(2σm − s)
& 12 ≤ σm < 1 ,

∵ ζ(s)(s − 1)sΓ( 2 )π = 1−
s s̄
 m=1  m m w
lims→1

s=1
  w
  w
  w
y y 
∞  
Y (2σm − 1) 1
∴ 1= 1− =⇒ (2σm − 1) = 0 =⇒ σm = .
m=1
sm s̄m 2

To validate the result, with σm = 12 , Equation (2.19) can be restated as

∞  
s Y s(1 − s)
(2.25) 2ξ(s) = ζ(s)(s − 1)sΓ( 2s )π − 2 = 1− ,
m=1
sm s̄m

from which we see that the right hand side of Equation (2.25) is unchanged when s
is replaced by (1 − s), obtaining the expressions for ζ(1 − s) and ξ(1 − s) as
∞  
1−s
− 2
Y s(1 − s)
(2.26) 2ξ(1 − s) = ζ(1 − s)s(s − 1)Γ( 1−s
2 )π = 1− .
m=1
sm s̄m

Therefore, Equations (2.25) and (2.26) are equal, as validated by the well-known ξ(s)
functional equation, given by

(2.27) ξ(s) = ξ(1 − s).

If any zero sm has σm 6= 21 in Equation (2.19), then it implies that ξ(s) 6= ξ(1 − s),
which would contradict Equation (2.27). Therefore, all σm must be equal to 21 . From
this, we can hypothesize that the product form of the ξ(s) in Equation (2.14) developed
by Riemann [1] was very likely to have been the source of inspiration for the Riemann
Hypothesis.
3. The Prime-Counting Function. In this section, I will revisit the prime-
counting function analysis. Recasting Riemann’s [1] synthesis and results in a format;
that will consolidate diverse elements such as the Heaviside step function, the von
Mangoldt function, the Dirac delta function, and the Chebyshev function. Then, I
will employ the Laplace transform to obtain the prime-counting function in the s-
domain. Also, I will utilize the residue theorem to drive the prime-counting function
in terms of ζ(s) pole and zeros. Some of these results are already available in the
literature in one form or another. However, here I will organize them logically and
bridge some crucial gaps to demonstrate the underlining relationships.
3.1. The Prime-Counting Function in the x-domain. The number of primes
less than a given magnitude x can be formulated in the x-domain on a fundamental
building block, using the staircase Heaviside step function H(ln x − ln p) as a base for
PROOF OF RIEMANN HYPOTHESIS 7

the prime counting function π(x), see Figure 3, given by


X
(3.1) π(x) = H(ln x − ln p),
p

where p is the set of all prime numbers {2, 3, 5, · · · , pm , · · · }.

π(x)

x
2 3 5 7

Fig. 3. Prime Counting Heaviside Step Function.

Also, in terms of prime harmonics pk , see Figure 4, we have


1
X X
(3.2) π(x k ) = H(ln x1/k − ln p) = H(ln x − k ln p).
p p

where k ∈ N.

1
π(x 2 )

x
2 3 4 5 7 9
1
Fig. 4. Prime Counting Function π(x k ) .

Differentiating the Heaviside function Equation (3.1); gives the Dirac delta func-
tion δ(x), see Figure 5, thus we have
X1
(3.3) π 0 (x) = δ(ln x − ln p).
p
x
8 F.A. ALHARGAN

x π 0 (x)

x
2 3 5 7

Fig. 5. Prime Dirac Delta Function δ(ln x − ln p).

Now, from Riemann’s definition of J(x) in terms of the prime counting function
π(x), as
1
X
1
(3.4) J(x) = k π(x ).
k

k∈N

and the inverse


1
X µ(k)
(3.5) π(x) = k J(x ).
k

k∈N

where µ(k) is the Möbius function.


Thus, from Equation (3.2) and Equation (3.4), we have


X
1
X X Λ(n)
(3.6) J(x) = k H(ln x − k ln p) = H(ln x − ln n),
p n=2
ln(n)
k∈N

where the von Mangoldt function, denoted by Λ(n), is defined as


(
ln p if n = pk for some prime p and integer k ≥ 1,
(3.7) Λ(n) =
0 otherwise.

Differentiating Equation (3.6), we have


X 1
X X1
J 0 (x) = 1 0
k π (x )
k = 1
k δ(ln x − k ln p)
p
x
k∈N k∈N
(3.8) ∞ ∞
XΛ(n) X Λ(n)
= δ(ln x − ln n) = δ(ln x − ln n).
n=2
x ln(n) n=2
x ln(x)

Furthermore, the first Chebyshev function is defined by


X
(3.9) ϑ(x) = ln p,
p≤x
PROOF OF RIEMANN HYPOTHESIS 9

and based on Heaviside function can be defined as


X
(3.10) ϑ(x) = H(ln x − ln p) ln p,
p

or in a prime factorization format, as


Y
(3.11) eϑ(x) = pH(ln x−ln p) .
p

Differentiating Equation (3.10), we have


X1
(3.12) ϑ0 (x) = δ(ln x − ln p) ln p.
p
x

It is interesting to note that the relation between the differential of the prime-counting
function in Equation (3.3) and the differential of the first Chebyshev function in
Equation (3.12), can be stated as

(3.13) ϑ0 (x) = π 0 (x) ln x.

Similarly, the second Chebyshev function ψ(x) is defined with the sum extending
over all prime powers not exceeding x, as
X X X
(3.14) ψ(x) = ln p = Λ(n),
k∈N pk ≤x n≤x

or based on the Heaviside function, as

XX ∞
X
(3.15) ψ(x) = H(ln x − k ln p) ln p = Λ(n) H(ln x − ln n),
k∈N p n=2

which can be represented in the following fascinating factorization format


YY
(3.16) eψ(x) = pH(ln x−k ln p) .
k∈N p

Differentiating Equation (3.15), we have


X X ln p ∞
X Λ(n)
(3.17) ψ 0 (x) = δ(ln x − k ln p) = δ(ln x − ln n).
x x
k∈N p n=2

From Equations (3.8) and (3.17), it is observed that

XX ∞
X
(3.18) xψ 0 (x) = xJ 0 (x) ln x = ln p δ(ln x − k ln p) = Λ(n)δ(ln x − ln n).
k∈N p n=2

Also, we observe that Equation (3.18) is basically a delta function, given by


(
0 0 ln p if x = pk a prime with an integer k ≥ 1,
(3.19) xψ (x) = xJ (x) ln x =
0 otherwise.
10 F.A. ALHARGAN

Now, multiplying Equation (3.2) by x−s−1 and integrating, we have


Z∞ Z∞ X
1
−s−1
(3.20) π(x ) x
k dx = H(ln x − k ln p) x−s−1 dx,
1 1 p

integrating the right hand-side of Equation (3.20) by parts, we have

Z∞
1 1 X −sk
(3.21) π(x k ) x−s−1 dx = p .
s p
1

Thus, from Equation (3.21) and Equation (3.4), we have


Z∞
1X X
(3.22) J(x) x−s−1 dx = 1
k p−sk .
s p
1 k∈N

Here, we observe that the right hand-side of Equation (3.22) is basically the log
of the Euler product of ζ(s). Therefore, Equation (3.22) can be restated as

Z∞
ln ζ(s)
(3.23) J(x)x−s−1 dx = , (< s > 1).
s
1

Equation (3.23) was one of the main results in Riemann’s paper [1]. However, the
above analysis reveals the profound direct connection between the Heaviside prime-
counting function and the zeta function. This approach shifts the perspective to a new
paradigm from number theory to signal processing theory using Riemann spectrum
[14], which will enable us to exert the signal processing arsenal to tackle some prime
numbers enigmas.
3.2. The Prime-Counting Function in the s-domain. The Fourier analysis
has been the mainstay in the literature for tackling the connection between zeta and
the prime number counting function. However, the Laplace transform is a generalized
Fourier transform that provides elegant and concise solutions; linking the s-domain
with the x-domain.
The analysis so far has been based on the x-domain. Here, I will formulate the
functions in the s-domain, then demonstrate their links via the Laplace transform by
first employing the well-known Laplace transforms for the Heaviside and Dirac delta
functions, given by
e−sk ln n
(3.24) L {H(ln x − k ln n)} = ,
s
and
 
1
(3.25) L δ(ln x − k ln n) = e−sk ln n .
x
Thus, the Laplace transform of the prime counting function Equation (3.1), from π(x)
to Π(s), is obtained as
X X e−s ln p
(3.26) Π(s) = L {π(x)} = L {H(ln x − ln p)} = ,
p p
s
PROOF OF RIEMANN HYPOTHESIS 11

and
  X
X 1
(3.27) sΠ(s) = L {π 0 (x)} = L δ(ln x − ln p) = e−s ln p .
p
x p

Here, it is important to note the definition of the prime-counting function in the s-


domain, denoted by the symbol Π(s). This is a new definition and should not be
confused with the same symbol used in the literature for different purposes. Thus, we
can define the s-domain prime-counting function as

1X 1
(3.28) Π(s) := , (< s > 1).
s p ps

The definition in Equation (3.28) is the base prime numbers sub-sum of the ζ(s)
function. It contains only the base primes; i.e. excluding prime number harmonics
and multiples of primes from the sum.
Now, we can utilize Equation (3.24) to obtain the Laplace transform of Equations
(3.6), (3.10) and (3.15), as

XX
1 −ks ln p
X Λ(n) −s ln n
(3.29) sJ (s) = ke = e ,
ln(n)
k∈N p n=2

X
(3.30) sΘ(s) = ln p e−s ln p ,
p

and
XX ∞
X
(3.31) sΨ(s) = ln p e−ks ln p = Λ(n)e−s ln n .
k∈N p n=2

Here, we recall the log expansion of the Euler product of Riemann zeta function,
which is given by

X XX X Λ(n) 1
(3.32) ln ζ(s) = − ln(1 − e−s ln p ) = 1 −ks ln p
ke = ,
p
ln(n) ns
k∈N p n=2

and differentiating, we have



ζ 0 (s) X ln p XX
−ks ln p
X
(3.33) = s)
= − ln p e = − Λ(n)e−s ln n .
ζ(s) p
(1 − p p n=2
k∈N

Therefore, from Equations (3.28), (3.29), (3.30), (3.31) (3.32) and (3.33), we
discover that in the s-domain, the relationship amongst the functions Π(s), ζ(s),
J (s), Θ(s) and Ψ(s), are as follows

XX
1 −ks ln p
X Λ(n) −s ln n X
(3.34) sJ (s) = ln ζ(s) = ke = e =s Π(ks),
ln(n)
k∈N p n=2 k∈N
12 F.A. ALHARGAN

and

ζ 0 (s) X X X X
(3.35) sΨ(s) = − = ln p e−ks ln p = Λ(n)e−s ln n = s Θ(ks).
ζ(s) p n=2
k∈N k∈N

Here, we observe the power of employing the Heaviside function and the s-domain
analysis, which immediately demonstrate the profound relationship between ζ(s) and
the prime counting function, for Equation (3.34) reveals that ln ζ(s) is the sum of all
the harmonics of the prime counting function Π(s) in the s-domain.
Remark 3.1. Here, it is important to highlight that the functions with these sym-
bols: Π(s), J (s), Θ(s) and Ψ(s). That, I have newly defined in this paper, as the
s-domain manifestation of their x-domain representation: π(x), J(x), ϑ(x) and ψ(x)
respectively. These functions have not been defined previously in the literature. Thus
should not be confused with any similar symbols you may encounter in the literature.
3.3. The Inverse Laplace Transform. Now, to obtain the functions in the
x-domain, we employ the residue theorem to evaluate the inverse Laplace transform
of the expressions; i.e.
X
(3.36) L−1 {F (s)} = Res [F (s)esy ] ,
all poles

where in our case y = ln x; and it is critical to take care of the effect of the term ln x
when differentiating, as the factor of x1 needs to be taken into account.
Also, by Mittag-Leffler’s theorem, we have

ζ 0 (s) 1 X 1 1 1
(3.37) = ln 2 − + + + ,
ζ(s) (s − 1) m=1 (s − sm ) (s − s̄m ) (s + 2m)
0
where the poles of the function ζζ(s)
(s)
are at s = 1, s = sm , s = s̄m and s = −2m.
Thus, the inverse Laplace transform, is obtained as follows
 0   0
ζ (s)es ln x

0 −1 −1 ζ (s) X
(3.38) xψ (x) =L {sΨ(s)} = L − = − Res ;
ζ(s) ζ(s)
all poles

i.e.

X
(3.39) xψ 0 (x) = eln x − esm ln x + es̄m ln x + e−2m ln x .
m=1

Also,

ζ 0 (s)
 0
ζ (s)es ln x
  X 
(3.40) ψ(x) = L−1 {Ψ(s)} = L−1 − =− Res ;
ζ(s) s ζ(s)s
all poles

i.e.

X esm ln x es̄m ln x e−2m ln x
(3.41) ψ(x) = eln x − ln 2π − + − .
m=1
sm s̄m 2m
PROOF OF RIEMANN HYPOTHESIS 13

1 1 1
We note that in Equation (3.41), the terms sm , s̄m and 2m are due to 1s , and the
0
term ζζ(0)
(0)
= ln 2π is due to the pole at s = 0.
Now, from Equations (3.18), (3.25), (3.33) and (3.39), we see that
(3.42)

X X∞
xψ 0 (x) = xJ 0 (x) ln x = x − xsm + xs̄m + x−2m = Λ(n)δ(ln x − ln n).
m=1 n=2

Also, from Equations (3.15), (3.24), (3.33) and (3.41), we see that
∞ ∞
X xsm xs̄m x−2m X
(3.43) ψ(x) = x − ln 2π − + − = Λ(n)H(ln x − ln n).
s
m=1 m
s̄m 2m n=2

It is interesting to observe that differentiating Equation (3.43) results in Equation


(3.42), which validates the analysis.
3.4. The J(x) Function. Rearranging Equation (3.42), we have

1 X xsm −1 xs̄m −1 x−2m−1
(3.44) J 0 (x) = − + + ,
ln x m=1 ln x ln x ln x

or as stated by Riemann [1] for an approximate expression for the density of the prime
numbers
∞ 1
1 X x− 2 cos(tm ln x)
(3.45) J 0 (x) = −2 .
ln x m=1
ln x

Now, integrating Equation (3.44), results in the logarithmic integral function


Li(x), giving

X
(3.46) J(x) = Li(x) − Li(xsm ) + Li(xs̄m ) + Li(x−2m ).
m=1

Then the prime counting function is finally obtained as



1
X µ(k)
X µ(k)
X
(3.47) π(x) = k Li(x k ) − k Li(xsm /k ) + Li(xs̄m /k ) + Li(x−2m/k ).
k∈N k∈N m=1

Equation (3.47) is exactly the core result of Riemann’s paper [1]. Furthermore,
the above analysis demonstrate clear insight into the relation of the prime counting
function and the zeta function; as can be observed that the first part of Equation
(3.47) is due to the pole of ζ(s) at s = 1, and the second part is due to the zeros
of ζ(s) at s = sm , s = s̄m and s = −2m. Moreover, we observe that the term
due to the pole is the major component of π(x), whereas the terms due to the real
zeros are negligible. In contrast, the terms due to the complex zeros are the source
of the sawtooth-like wave component, which will be illustrated later. The sums are
conditionally convergent with a slow convergence rate. Although Equation (3.47)
was a landmark result that laid the foundations for prime numbers analysis, it is
cumbersome and not convenient for analyzing prime numbers. In the next section, I
will develop more convenient expressions.
14 F.A. ALHARGAN

3.5. Chebyshev ψ(x) Function. Utilizing the proof of the Riemann Hypothe-
sis, that the non-trivial zeros of zeta have real part equal to 12 . i.e. the zeros have a
format of sm = 12 + i tm , and noting that

X x−2
(3.48) x−2m = .
m=1
1 − x−2

Thus, Equation (3.42) can be expressed as

∞ ∞
x−3 − 12
X X Λ(n)
(3.49) ψ 0 (x) = 1 − − 2x cos (tm ln x) = δ(ln x − ln n).
(1 − x−2 ) m=1 n=2
x

Also, integrating Equation (3.49) or rearranging Equation (3.43), we have


(3.50)

1
X [cos(tm ln x) + 2 tm sin(tm ln x)]
ψ(x) = x − ln 2π − 12 ln(1 − x12 ) − 4x 2 .
m=1
4 t2m + 1

Note that Equation (3.50), was proved in 1895, by Hans Carl Friedrich von Man-
goldt ([15] p. 294 Equ.58), and was stated in the paper as

1 1 1
X [cos(αν ln x) + 2 αν sin(αν ln x)]
(3.51) Λ(x, 0) = x − ln 2π − 2 ln(1 − x2 ) − x2 1 .
ν=1 4 + αν2

where he showed that Λ(x, r) ([15] p. 279 Equ.38), is given by


a+ih
ζ 0 (s + r) xs
Z
1
(3.52) − lim · · ds = Λ(x, r).
h→∞ 2πi ζ(s + r) s
a−ih

0
ζ (s)
Equation (3.52) is basically the inverse Laplace transform of ζ(s) s when r = 0, from
which we see that ψ(x) = Λ(x, 0).
Of course, some of these results are not new, already Edwards ([6] p.50) had
shown a short method to obtain Equation (3.51). However, in this paper, I have
employed the residue theorem to prove the results within few steps. Furthermore, I
have demonstrated via Laplace transform the links of the Chebyshev ψ(x) function
and its derivative to the prime counting functions, the Heaviside function, and the
Dirac delta function, which provides a new perspective and simplifies the analysis.
Now, the term [ln(1 − x−2 )] has a negligible value, thus Equations (3.49) and
(3.50) can be approximated to

∞ ∞
1
X X
(3.53) xψ 0 (x) = x − 2x 2 cos (tm ln x) = Λ(n)δ(ln x − ln n).
m=1 n=2

and
(3.54)
∞ ∞
1
X cos(tm ln x) sin(tm ln x) X
ψ(x) u x − ln 2π − x 2 + = Λ(n)H(ln x − ln n).
m=1
t2m 2tm n=2
PROOF OF RIEMANN HYPOTHESIS 15

Equations (3.53) and (3.54) are in essence contain the Riemann spectrum of ζ(s)
non-trivial zeros, i.e. the set { 21 + itm }. These equations are powerful in locating
the prime numbers and their harmonics; for Equation (3.53) is essentially a delta
function of the prime numbers harmonics, whereas Equation (3.54) is a Heaviside
staircase function of the prime numbers harmonics.
Comparing Equation (3.33) and Equation (3.53), we observe that in the s-domain
the function Ψ(s) is summed over the prime numbers; and has poles at the zeros of
zeta at s = sm , s = s̄m and s = −2m. In contrast, the x-domain function ψ 0 (x) is
summed over the zeros of zeta; and has poles at the harmonics of the prime numbers.

Fig. 6. x ψ 0 (x) with x continuous.

Although Equations (3.53) and (3.54) are slow and conditionally convergent, the
two equations provide better speed and accuracy for the primality test than current
algorithms. In figures (6) and (7); with the computation executed at steps of x =
0.01, we can see the harmonics and Gibbs phenomenon. Also, we observe the prime
locations at 101, 103, 107, 109; the harmonic primes at 121 = 112 , 125 = 53 , 128 =
27 , and the larger primes at 127, 131, 137, 139; whereas the rest of the natural non-
prime harmonics integers are small. In Figure (7), the Heaviside staircase prime
harmonics counting function is observed. Also, we can observe the sawtooth-like
waveform component of ψ(x) in Figure (8).
16 F.A. ALHARGAN

Fig. 7. ψ(x) with x continuous.

Fig. 8. Sawtooth-like waveform component of ψ(x) .

4. The Zero-Counting Function. In this section, and in light of the proof of


the Riemann Hypothesis, I will revisit the zero-counting function analysis. Then I
will reformulate the equations to obtain a very accurate zero-counting function.
Now, in the range {0, T }, the number of roots of ξ(s); was conjectured by Riemann
[1], as approximately
T T T
(4.1) = 2π ln 2π − 2π ,
PROOF OF RIEMANN HYPOTHESIS 17

and some 46 years later was proved by H. von Mangoldt [16], the prove was outlined
by Ivic ([7], p. 17), where he showed that the number of zeros is given approximately
by
Z 0
T T T ζ (s)
(4.2) N (T ) = 2π ln 2π − 2π + 78 + π1 = ds,
L ζ(s)

and demonstrated that


ζ 0 (s)
Z  
(4.3) = ds = O(ln T ).
L ζ(s)
Although the integral in Equation (4.3) is small compared to the major elements in
Equation (4.2), it still contains the sawtooth-like waveform component, that I will
demonstrate shortly.
Now, recalling Riemann [1] R main justification of Equation (4.1), quoted as follows:
”because the integral d log ξ(t), taken in a positive sense around the
region consisting of the values of t whose imaginary parts lie between
1 1
2 i and − 2 i and whose real parts lie between 0 and T , is (up to a frac-
tion of the order of magnitude of the quantity T1 ) equal to (T log 2π T

T )i; this integral however is equal to the number of roots of ξ(t) = 0
lying within this region, multiplied by 2πi. One now finds indeed ap-
proximately this number of real roots within these limits, and it is
very probable that all roots are real.”
In essence, Riemann instinctively was invoking Cauchy’s argument principle, for
ξ(s) is a meromorphic function inside and on some closed contour D, and ξ(s) has no
zeros or poles on D, thus
I 0
1 ξ (s)
(4.4) ds = Z − P,
2πi D ξ(s)
where Z and P denote the number of zeros and poles of ξ(s); inside the contour D.
Now, from the proof of the Riemann Hypothesis, which implies that ξ(s) has
simple zeros only on the critical line <(s) = 21 , at s = sm and s = s̄m . Then, we
can define the zero counting function ν(t) of the number of zeros of ξ(s), in the range
{s, s̄}, as
1 Γ0 ( 2s ) 1
I  0  I  0 
ξ (s) ζ (s) 1
(4.5) 4πiν(t) = ds = + + + − ln π ds.
D ξ(s) D ζ(s) (s − 1) s Γ( 2s ) 2
where the closed contour D encompasses the critical strip [0 ≤ <(s) ≤ 1].
Now, since all the poles of [ζ 0 (s)/ζ(s) + Γ0 ( 2s )/Γ( 2s )] are on the critical line <(s) =
1
2 , and the contour L1 encloses all these poles in the range from s̄ to s, whereas the
contour L2 encloses only the two poles s = 0 and s = 1, see Figure (9). Then we can
rearrange Equation (4.5), as
ζ (s) Γ0 ( 2s ) 1
I  0  I  
1 1
(4.6) 4πiν(t) = + − ln π ds + + ds,
L1 ζ(s) Γ( 2s ) 2 L2 (s − 1) s
The contour integration around L1 , can be transformed to a line integral, as
follows
I Z s+ Z s− Z s̄− Z s̄+  Z s
(4.7) = lim + − − =2
L1 →0 s̄+ s+ s− s̄− s̄
18 F.A. ALHARGAN

t
D smL1

L2

1 σ
2

s̄m

Fig. 9. ζ(s) Critical Strip, Contours D, L1 and L2 .

and the contour integration around L2 , is given by


I  
1 1
(4.8) + ds = 4πi,
L2 (s − 1) s

thus, we have
s
ζ 0 (s) Γ0 ( 2s ) 1
Z  
(4.9) 4πiν(t) = 2 + − ln π ds + 4πi,
s̄ ζ(s) Γ( 2s ) 2

integrating, we finally have


s s̄
(4.10) 2πiν(t) = ln ζ(s) − ln ζ(s̄) + ln Γ( 2s ) − ln Γ( 2s̄ ) + ln π − 2 − ln π − 2 + 2πi,

or
[ζ(s)Γ( 2s )π −it ] 2πi
(4.11) e2πiν(t) = e ;
[ζ(s̄)Γ( 2s̄ )]

i.e.
(4.12)
2πiν(t) = ln ζ( 12 + it) − ln ζ( 12 − it) + ln Γ( 14 + i 2t ) − ln Γ( 41 − i 2t ) − it ln π + 2πi.

Differentiating Equation (4.12), we have

ζ 0 ( 21 + it) ζ 0 ( 12 − it) Γ0 ( 41 + i 2t ) Γ0 ( 14 − i 2t )
(4.13) 2πiν 0 (t) = − + − − i ln π.
ζ( 12 + it) ζ( 12 − it) Γ( 14 + i 2t ) Γ( 14 − i 2t )

Utilizing Stirling approximation


1 √
(4.14) Γ(z) ∼ (z)z e−z z − 2 2π,
PROOF OF RIEMANN HYPOTHESIS 19

and

Γ0 (z) 1
(4.15) ∼ ln z − ,
Γ(z) 2z

we have

(4.16) ln Γ( 2s ) − ln Γ( 2s̄ ) ∼ s
2 ln 2s − 12 s − 1
2 ln s − s̄
2 ln 2s̄ + 12 s̄ + 1
2 ln s̄,

or

(4.17) ln Γ( 2s ) − ln Γ( 2s̄ ) ∼ s
2 ln 2s − s̄
2 ln 2s̄ + 1
2 ln s̄ − 1
2 ln s − it.

Thus, finally we have

2πiν(t) =( 41 + i 2t ) ln( 14 + i 2t ) − ( 41 − i 2t ) ln( 14 − i 2t )


1
+ 2 ln( 14 − i 2t ) − 1
2 ln( 41 + i 2t )
(4.18) X
+ ln ζ( 12 + it) − ln ζ( 12 − it) − it ln πe = 2πi H(t − tm ),
m

and from Equation (4.13), we have

1 1
2πiν 0 (t) = ln( 12 + it) − − ln( 12 − it) + 1
( 12 + it) ( 2 − it)
(4.19)
ζ 0 ( 12 + it) ζ 0 ( 12 − it) X
+ 1 − 1 − i ln π = 2πi δ(t − tm ).
ζ( 2 + it) ζ( 2 − it) m

Equation (4.18) is a very accurate approximation, and it is a sum of differences be-


tween complex numbers and their conjugates, thus the result will always be imaginary
number as expected. Further approximation of the log part gives

t t 7 1
ln ζ( 12 + it) − ln ζ( 21 − it) .
 
(4.20) ν(t) = 2π ln 2eπ + 8 + 2πi

Figure (10), shows the sawtooth-like waveform component ln ζ( 12 + it) − ln ζ( 12 − it)


 

of ν(t). It is observed that the component magnitude is less than one. However, it
has a vital contribution to the accuracy of the zero-counting function; which turns it
into a Heaviside staircase step function, as shown in Figure (11).
20 F.A. ALHARGAN

Fig. 10. The sawtooth-like waveform component of ν(t).

Fig. 11. ν(t) vs. N (t).

Now, from equation (3.34), we have

X
(4.21) ln ζ( 12 + it) − ln ζ( 12 − it) = 1 1
k(2 + it)Π(k( 12 + it)) − k1 ( 12 − it)Π(k( 12 − it)),
k
PROOF OF RIEMANN HYPOTHESIS 21

or
1 −k
XX
(4.22) ln ζ( 21 + it) − ln ζ( 12 − it) = 2i kp
2 sin(tk ln p),
p k

giving

1 −k
XX
t t 7 1
(4.23) ν(t) = 2π ln 2eπ + 8 + π kp
2 sin(tk ln p).
p k

We observe from Equation (4.21), the direct relation between the zero-counting
function ν(t) and the s-domain prime-counting function Π(s). Furthermore, we ob-
serve in Equation (4.23), the direct relationship between the number of ζ(s) zeros
and the prime harmonics. Although the equation is slow for computational purposes,
it reveals the underlining relationship between the zero-counting function and the
primes. Furthermore, it exposes the source of the sawtooth-like waveform effect as
the spectrum sum of the prime harmonics.
Finally, Figure (11) shows comparisons between ν(t) and N (t), and it confirms
that the zero counting function is a Heaviside staircase step function, and its differ-
ential ν 0 (t) is an impulse Dirac delta function; as can be seen in Figure (12).

Fig. 12. |ν 0 (t)| .

5. Recommendations for Future Research. Here I will summarize some of


the key results and relations, then I will outline some interesting topics, that I think
are worthy of further research and exploration.
5.1. Natural Number Counting Function. Now, consider the natural number-
counting function φ(x) based on Heaviside function, which we can defined as
X
(5.1) φ(x) = H(ln x − ln n),
n
22 F.A. ALHARGAN

and
X1
(5.2) φ0 (x) = δ(ln x − ln n).
n
x

Taking the Laplace transform of the above equations, we have


X 1  X ∞
L {φ0 (x)} =
 −s ln n
(5.3) L δ(ln x − ln n) = e = ζ(s),
n
x n=1

and

X X e−s ln n
(5.4) L {φ(x)} = L {H(ln x − ln n)} = = 1s ζ(s).
n n=1
s

Therefore, the inverse Laplace transform of 1s ζ(s) to the x-domain, manifests


itself as the natural number counting function φ(x). Also, we observe from Equation
(5.3) an interesting x-domain representation of the inverse Laplace transform of ζ(s)
as a decreasing Dirac impulse function. Equation (5.4) can be utilized to count many
variations of numbers, such as the numbers of the form nr , exploring the features of
this equation is an interesting topic to research.
5.2. The Prime Counting Function. Recalling the definition of the prime
counting function in the x-domain, as
X
(5.5) π(x) = H(ln x − ln p),
p

with
1
X
1
(5.6) J(x) = k π(x ),
k

k∈N

and the reverse


1
X µ(k)
(5.7) π(x) = k J(x ),
k

k∈N

where µ(k) is the Möbius function.


Also, recalling the s-domain form, given by
X 1
(5.8) sΠ(s) = , < s > 1.
p
ps

with the relation to the zeta function, given by


X
(5.9) sJ (s) = ln ζ(s) = s Π(ks),
k∈N

and the reverse, as


X µ(k) X µ(k)
(5.10) sΠ(s) = s J (ks) = ln ζ(ks).
k k
k∈N k∈N
PROOF OF RIEMANN HYPOTHESIS 23

1
Fig. 13. <{sΠ(s)} along the critical line s = 2
+ it.

Equation (5.8) for Π(s) exhibits elegance as well as deceptive simplicity, however,
its complexity is revealed by Equation (5.10). The function Π(s) has poles at s = 0,
s = 1, and at the zeros of ζ(s). Figure (13) shows the real part of sΠ(s); along the
critical line s = 21 , we also observe the poles at s = 12 + itm .
Now, the inverse Laplace of Π(s) gives directly the prime-counting function π(x),
i.e.

(5.11) π(x) = L−1 {Π(s)} .

From the Laplace transform properties, multiplication by s results in differentiation


of π(x), and multiplication by −x results in differentiation of sΠ(s), thus we have

(5.12) − x π 0 (x) ln x = L−1 {[sΠ(s)]0 } ;

i.e.
( )
X µ(k) ζ 0 (ks)
0 −1
(5.13) x π (x) ln x = L − ,
k ζ(ks)
k∈N

Then using the residue theorem to evaluate the inverse Laplace, we have
" #
X X µ(k) ζ 0 (ks)es ln x
0
(5.14) x π (x) ln x = − Res ,
k ζ(ks)
all poles k∈N

giving

∞ X
1 µ(k) skm s̄m 2m
X µ(k) k
X
(5.15) xπ 0 (x) ln x = k x − k x +x k + x− k .
k∈N m=1 k∈N
24 F.A. ALHARGAN

Integrating, we have


1
X µ(k)
X µ(k)
X
(5.16) π(x) = k Li(x k ) − k Li(xsm /k ) + Li(xs̄m /k ) + Li(x−2m/k ).
k∈N k∈N m=1

Here we come back a full circle to the same result. However, this approach gives
much better clarity and coherence with a far fewer steps.
The elegant s-domain forms present a new perspective in the relation between
the ζ(s) function and the prime counting function Π(s). The behaviour of Π(s) needs
further investigation that might reveal new insights into the computations of the
primes.
5.3. The Chebyshev Function. Recalling the definition of the first kind Cheby-
shev function in the x-domain, as
X
(5.17) ϑ(x) = H(ln x − ln p) ln p,
p

or in a factorization form, as
Y
(5.18) eϑ(x) = pH(ln x−ln p) .
p

Also, the first kind Chebyshev function in the s-domain, as


X
1
(5.19) sΘ(s) = ps ln p, < s > 1.
p

or
Y 1
(5.20) sΘ(s) = ln p ps , < s > 1.
p

Moreover, its relation to the zeta function is given by

ζ 0 (s) X
(5.21) sΨ(s) = − =s Θ(ks).
ζ(s)
k∈N

and the reverse, as

X µ(k) X µ(k) ζ 0 (ks)


(5.22) sΘ(s) = s Ψ(ks) = − .
k k ζ(ks)
k∈N k∈N

In fact, the prime counting function π(x) and the first Chebyshev function ϑ(x), can
be directly evaluated from Equation (5.22), using the inverse Laplace transform. We
note that the function Θ(s) has poles at s = 0, s = 1, and the zeros of ζ(s).
This again, provides another perspective in the relation between ζ(s) and the
Chebyshev function expressed in the s-domain. The behaviour of Θ(s) needs further
investigation, that could yield greater insights into the prime factorization.
PROOF OF RIEMANN HYPOTHESIS 25

6. Conclusions. Proof of the Riemann Hypothesis would unravel many of the


mysteries surrounding the distribution of prime numbers, which are at the heart of
all encryption systems. In addition, proof of the Riemann Hypothesis would, as a
consequence, prove many of the propositions known to be true under the Riemann
Hypothesis.
The proof demonstrated in this paper was based on a basic insight into the product
expansion of the Riemann zeta function, as available from Hadamard’s publication
in 1893 and Riemann’s publication in 1859, as well as clarifying that the product
expansion is only over the principle non-trivial zeros of zeta. Sometimes, the truth is
hidden in plain sight.
Furthermore, in this paper, I have demonstrated several techniques with a new
perspective. These included the Heaviside function, Dirac delta function, the Laplace
transform, Mittag-Leffler’s theorem, and the residue theorem. The techniques have
provided concise and elegant solutions; that immediately revealed the profound un-
derlining connection between the Heaviside prime-counting function, the Dirac delta
function, and the zeta function. This novel approach shifts the perspective to a new
paradigm, from number theory to signal processing theory. These results will enable
us to exert the signal processing arsenal to tackle some prime numbers enigmas. The
new defined s-domain prime-counting function Π(s) and Chebyshev function Ψ(s)
provide another angle to address prime computations, primality testing, and prime
factorization.
Acknowledgements. I would like to thank Dr. Sami M. Alhumaidi director
general of PSDSARC, Prof. Mohammed I. Alsuwaiyel, and Dr. Fawzi A. Al-thukair
for their feedback on the first draft and their encouragement. Naturally, All Errors
are My Own.

REFERENCES

[1] Riemann, Bernhard (1859). ”Über die Anzahl der Primzahlen unter einer gegebenen Grösse”.
Monatsberichte der Berliner Akademie.. In Gesammelte Werke, Teubner, Leipzig (1892),
Reprinted by Dover, New York (1953).
[2] Hadamard, J. (1893). ”Etude sur les propriétés des fonctions entières et en particulier d’une
fonction considérée par Riemann”. Journal de Mathématiques Pures et Appliquées.
[3] Hadamard, J. (1896), ”Sur la distribution des zéros de la fonction ζ(s) et ses conséquences
arithmétiques”. Bulletin de la Société Mathématique de France, Volume 24 (1896) , pp.
199-220. doi : 10.24033/bsmf.545.
[4] Hilbert, David (1900), Mathematische Probleme, Gottinger Nachrichten, pp. 253-297.
[5] Titchmarsh, E. C. (1986). Heath-Brown (ed.) The Theory of the Riemann Zeta Function (2nd
rev. ed.), Oxford University Press.
[6] Edwards, H. M. (1974). Riemann’s Zeta Function, Academic Press.
[7] Ivic, A. (1985). The Riemann Zeta Function, John Wiley & Sons.
[8] Karatsuba, A. A.; Voronin, S. M. (1992). The Riemann Zeta-Function Berlin: W. de Gruyter.
[9] Bombieri, Enrico (2000). ”Problems of the millennium: The Riemann Hypothesis”, CLAY.
[10] Sarnak, Peter (2004). ”Problems of the Millennium: The Riemann Hypothesis”, CLAY.
[11] de la Vallée-Poussin, Ch.J. (1896), ”Recherches analytiques sur la théorie des nombers pre-
miers”, Ann. Soc. Sci. Bruxelles, 20: 183–256
[12] Hardy, G. H.; Littlewood, J. E. (1921), ”The zeros of Riemann’s zeta-function on the critical
line”, Math. Z., 10 (3–4): 283–317, doi:10.1007/BF01211614
[13] Xavier Gourdon (2004). The 1013 first zeros of the Riemann zeta
function, and zeros computation at very large height. available at
http://numbers.computation.free.fr/Constants/Miscellaneous/zetazeros1e13-1e24.pdf.
[14] Mazur, B., Stein, W. (2016). Prime Numbers and the Riemann Hypothesis. Cambridge: Cam-
bridge University Press. doi:10.1017/CBO9781316182277.
[15] Mangoldt, H. von (1895), Zu Riemanns Abhandlung ”Ueber die Anzahl der Primzahlen unter
einer gegebenen Grösse”. Journal für die reine und angewandte Mathematik 1895, Vol. 114,
26 F.A. ALHARGAN

pp. 255-305, http://gdz.sub.uni-goettingen.de/dms/resolveppn/?PPN=GDZPPN002163144


[16] Mangoldt, H. von (1905), Zur Verteilung der Nullstellen der Riemannschen Funktion ξ(t).
Mathematische Annalen 1905, Vol. 60, pp. 1-19,

You might also like