PreprintPDF Available
Preprints and early-stage research may not have been peer reviewed yet.

Abstract

We present a simple proof of the Riemann's Hypothesis (RH) where only undergraduate mathematics is needed.
A Simple Proof of the Riemann’s Hypothesis
Charaf ECH-CHATBI
23 October 2023
Abstract
We present a simple proof of the Riemann’s Hypothesis (RH) where
only undergraduate mathematics is needed.
Keywords: Riemann Hypothesis; Zeta function; Prime Numbers;
Millennium Problems.
MSC2020 Classification: 11Mxx, 11-XX, 26-XX, 30-xx.
1 The Riemann Hypothesis
1.1 The importance of the Riemann Hypothesis
The prime number theorem gives us the average distribution of the primes.
The Riemann hypothesis tells us about the deviation from the average.
Formulated in Riemann’s 1859 paper[1], it asserts that all the ’non-trivial’
zeros of the zeta function are complex numbers with real part 1/2.
1.2 Riemann Zeta Function
For a complex number swhere (s)>1, the Zeta function is defined as
the sum of the following series:
ζ(s) =
+
X
n=1
1
ns(1)
In his 1859 paper[1], Riemann went further and extended the zeta function
ζ(s), by analytical continuation, to an absolutely convergent function in
the half plane (s)>0, minus a simple pole at s = 1:
ζ(s) = s
s1sZ+
1
{x}
xs+1 dx (2)
Where {x}=x[x] is the fractional part and [x] is the integer part of x.
Riemann also obtained the analytic continuation of the zeta function to
the whole complex plane.
Email: charaf.chatbi@gmail.com. The opinions of this article are those of the author and
do not reflect in any way the views or business of his employer.
1
arXiv:submit/5224804 [math.GM] 11 Nov 2023
Riemann[1] has shown that Zeta has a functional equation1
ζ(s) = 2sπs1sin πs
2Γ(1 s)ζ(1 s) (4)
Where Γ(s) is the Gamma function. Using the above functional equa-
tion, Riemann has shown that the non-trivial zeros of ζare located sym-
metrically with respect to the line (s)=1/2, inside the critical strip
0<(s)<1. Riemann has conjectured that all the non trivial-zeros are
located on the critical line (s) = 1/2. In 1921, Hardy & Littlewood[2,3,
6] showed that there are infinitely many zeros on the critical line. In 1896,
Hadamard and De la Vall´ee Poussin[2,3] independently proved that ζ(s)
has no zeros of the form s= 1 + it for tR. Some of the known results[2,
3] of ζ(s) are as follows:
ζ(s) has no zero for (s)>1.
ζ(s) has no zero of the form s= 1 + iτ. i.e. ζ(1 + )= 0, τ.
ζ(s) has a simple pole at s= 1 with residue 1.
ζ(s) has all the trivial zeros at the negative even integers s=2k,
kN.
The non-trivial zeros are inside the critical strip: i.e. 0 <(s)<1.
If ζ(s) = 0, then 1 s, ¯sand 1 ¯sare also zeros of ζ: i.e. ζ(s) =
ζ(1 s) = ζs) = ζ(1 ¯s) = 0.
Therefore, to prove the “Riemann Hypothesis” (RH), it is sufficient to
prove that ζhas no zero on the right hand side 1/2<(s)<1 of the
critical strip.
1.3 Proof of the Riemann Hypothesis
Let us take a complex number ssuch that s=σ+iτ. Unless we explic-
itly mention otherwise, let us suppose that 0 < σ < 1, τ > 0 and ζ(s) = 0.
We have from the Riemann’s integral (2) above:
ζ(s) = s
s1sZ+
1
{x}
xs+1 dx (5)
We have s= 1, s= 0 and ζ(s) = 0, therefore:
1
s1=Z+
1
{x}
xs+1 dx (6)
Or in other terms:
1
σ+iτ1=Z+
1
{x}
xσ+iτ+1dx (7)
1This is slightly different from the functional equation presented in Riemann’s paper[1].
This is a variation that is found everywhere in the litterature[2,3,4]. Another variant using
the cos:
ζ(1 s)=21sπscos πs
2Γ(s)ζ(s) (3)
2
Let us denote the following functions:
ϵ(x) = {x}(8)
ϕ(x) = {x}1 {x}(9)
Ψ(x) = Zx
1
dx ϵ(x) (10)
K(σ, τ ) = τ
(1 σ)2+τ2(11)
To continue, we will prove the following lemmas.
Lemma 1.1. Let us consider two variables σand τsuch that σ > 0and
τ > 0. Let us define two integrals I(a, σ, τ )and J(a, σ, τ )as follows:
I(a, σ, τ ) = Za
1
sin(τln(x))
xσdx (12)
J(a, σ, τ ) = Za
1
cos(τln(x))
xσdx (13)
Therefore
I(a, σ, τ ) = K(σ, τ )1cos(τln(a))
aσ1(σ1)
τ
sin(τln(a))
aσ1(14)
J(a, σ, τ ) = K(σ, τ )(σ1)
τ(σ1)
τ
cos(τln(a))
aσ1+sin(τln(a))
aσ1(15)
Proof. Let us consider two variables σand τsuch that σ > 0 and τ > 0.
Let us take a > 1.
I(a, σ, τ ) = Za
1
sin(τln(x))
xσdx (16)
=Zln(a)
0
sin(τx)e(1σ)xdx (17)
=K(σ, τ )1cos(τln(a))
aσ1(σ1)
τ
sin(τln(a))
aσ1(18)
And the same for J(a, σ, τ ) for a > 0:
J(a, σ, τ ) = Za
1
cos(τln(x))
xσdx (19)
=Zln(a)
0
cos(τx)e(1σ)xdx (20)
=K(σ, τ )(σ1)
τ(σ1)
τ
cos(τln(a))
aσ1+sin(τln(a))
aσ1(21)
Lemma 1.2. The function ϵ(x)is piecewise continuous on [0,+)and
its primitive function Ψ(x)is defined as follows:
Ψ(x) = 1
2x1ϕ(x)(22)
3
Let us consider two variables σand τsuch that 0< σ < 1and τ > 0such
that s=σ+iτis a zeta zero. Therefore:
Z+
1
dx Ψ(x)
x2+s=1
(s1)(1 + s)(23)
And
Z+
1
dx ϕ(x)
x2+s=1
s(1 s)(24)
Proof. We will need the function ϕas you will see later that we need a
continuous function instead of a piecewise one like the function ϵ.
Let us take x > 1 a real number. Let us denote nx=xbe the integer
part of x. We have nx=x {x}. Therefore, we can write the following:
Ψ(x) = Zx
1
ϵ(t)dt (25)
=
nx1
X
n=1 Zn+1
n{t}dt +Zx
nx{t}dt (26)
=
nx1
X
n=1 Zn+1
n
(tn)dt +Zx
nx
(tnx)dt (27)
=
nx1
X
n=1
1
2+1
2(xnx)2(28)
=1
2nx1 + {x}2(29)
=1
2x1 {x}+{x}2(30)
This prove the equation (23).
Let us prove the second point of the lemma. Let us define the integral
Iϵ(s) as follows:
Iϵ(s) = Z+
1
dx ϵ(x)
x1+s(31)
The function xϵ(x)
x1+sis integrable on [1,+) and thanks to the
integration by parts, we can write the following:
Iϵ(s) = Z+
1
dx ϵ(x)
x1+s(32)
=hΨ(x)
x1+si+
x=1 + (1 + s)Z+
1
dx Ψ(x)
x2+s(33)
= (1 + s)Z+
1
dx Ψ(x)
x2+s(34)
Since, sis a zeta zero, from the equation (6), we have:
Iϵ(s) = 1
s1(35)
4
Therefore
Z+
1
dx Ψ(x)
x2+s=1
(s1)(s+ 1) (36)
Thanks to equation (23), we can write:
Z+
1
dx ϕ(x)
x2+s=Z+
1
dx (x1)
x2+s2Z+
1
dx Ψ(x)
x2+s(37)
=1
s1
s+ 1 2
(s1)(s+ 1) (38)
=1
s(1 s)(39)
Lemma 1.3. Let us consider two variables σand τsuch that 0< σ < 1,
τ > 0and s=σ+iτis a zeta zero. Let us define the sequence of functions
ϕnand ψnover [0,+)and ϕnover [1,+)such that ϕ0(x) = ϕ0(x) =
ϕ(x)and for each n1,ϕn(0) = 0, and:
ϕn+1(x) = 1
xZx
0
dt ϕn(t)for x > 0 (40)
ϕn+1(x) = 1
xZx
1
dt ϕn(t)for x1 (41)
ψn(x) = x
2nx2
3nfor x0 (42)
Therefore:
1. For each n0:
Z+
1
dx ϕn(x)
x2+s=1
s2n+1
(1 s) 3n(43)
2. For each n0:
Z+
1
dx ϕn(x)
x3s=1
(1 s) 2n+1
s3n(44)
3. For each x1and n1:
ϕn(x) = ϕn(x) + 1
2nx
n1
X
k=0
2klnk(x)
k!1
3nx
n1
X
k=0
3klnk(x)
k!(45)
4. For each n1and x1:
ϕn(x) = 1
(n1)!
1
xZx
1
dt ϕ0(t)ln( x
t)n1(46)
5
5. For each x0:
lim
n+2nϕn(x) = x(47)
6. For each x0and n0:
02nϕn(x)x(48)
7. For each x0and n1:
ψn(x) = ψn(x) + 1
2nx
n1
X
k=0
2klnk(x)
k!1
3nx
n1
X
k=0
3klnk(x)
k!(49)
Where
ψn(x) = 1
(n1)!
1
xZx
1
dt (tt2)ln( x
t)n1(50)
Proof. We have s=σ+iτazeta zero. The lemma 1.2 calculates the
integral A(s) as follows:
A(s) = Z+
1
dx ϕ0(x)
x2+s=1
s(1 s)(51)
We use the integration by parts to write the following:
A(s) = Z+
1
dx ϕ0(x)
x2+s(52)
="1
x2+sZx
0
dx ϕ0(x)#+
1
+ (2 + s)Z+
1
dx ϕ1(x)
x2+s(53)
=Z1
0
dx ϕ0(x) + (2 + s)Z+
1
dx ϕ1(x)
x2+s(54)
=Z1
0
dx ϕ0(x)(2 + s)Z1
0
dx ϕ1(x) + (2 + s)2Z+
1
dx ϕ2(x)
x2+s(55)
... (56)
=
n
X
k=0
(2 + s)kZ1
0
dx ϕk(x) + (2 + s)n+1 Z+
1
dx ϕn+1(x)
x2+s(57)
We could do the above integration by parts because the functions
xϕk(x)
x2+sare piecewise continuous2and integrable over [1,+) as they
are dominated by the function x1
x2+σthat is integrable over [1,+).
In fact, we can prove that the functions ϕkare non-negative and bounded
by 1 as we can prove by induction that for each kfor each x > 0 that:
0< ϕk+1(x) = 1
xZx
0
dt ϕk(t)1
xZx
0
1dt = 1 (58)
2All the functions xϕk(x)
x2+sare continuous except when k= 0 as the function ϕ0is
piecewise continuous.
6
This can be proven by induction using the fact that it is true for the initial
case as we have for each x0:
0ϕ0(x) = {x}(1 {x})1 (59)
Now, we need to calculate the integrals Ikfor k0:
Ik=Z1
0
dx ϕk(x) (60)
For x(0,1), we have:
ϕ0(x) = xx2(61)
And
ϕ1(x) = x
2x2
3(62)
Therefore, we can write for each kfor x(0,1):
ϕk(x) = x
2kx2
3k(63)
Therefore, for each k1:
Ik=1
2k+1 1
3k+1 (64)
Therefore, we can conclude:
A(s) = 1
s(1 s)+ (2 + s)n+1"Z+
1
dx ϕn+1(x)
x2+s 1
s2n+1 +1
(1 s) 3n+1 !#(65)
Since 2+s= 0, by combining the equations (51) and (65), we conculde
the first point of our lemma.
The second point of the lemma can be proved the same way like the
first point since 1 sis also a zeta zero and 3 s= 0(by applying the
first point on the zeta zero 1 s).
The point 3) can be proved by induction. For n= 0. We have ϕ0(x) =
ϕ0(x). Let us assume that it is true till nand let us prove for n+ 1.
Therefore:
ϕn+1(x) = 1
xZ1
0
dt ϕn(t) + 1
xZx
1
dt ϕn(t) (66)
=1
x1
2n+1 1
3n+1 +1
x2nZx
1
dt
t
n1
X
k=0
2klnk(t)
k!(67)
1
x3nZx
1
dt
t
n1
X
k=0
3klnk(t)
k!+1
xZx
1
dt ϕn(t) (68)
=1
x1
2n+1 1
3n+1 +1
x2n+1
n1
X
k=0
2k+1 lnk+1(x)
(k+ 1)! (69)
1
x3n+1
n1
X
k=0
3k+1 lnk+1(x)
(k+ 1)! +ϕn+1(x) (70)
7
Let us prove the 4th point. We proceed by induction. For n= 1, we
retrieve the definition of ϕ1(x). Let us assume that it is true up to nand
let us prove it for n+ 1. Thanks to the integral order change, we can
write:
ϕn+1(x) = 1
xZx
1
dt ϕn(t) (71)
=1
(n1)!
1
xZx
1
dt
tZt
1
ds ϕ0(s)ln( t
s)n1(72)
=1
(n1)!
1
xZx
1
ds ϕ0(s)Zx
s
dt
tln( t
s)n1(73)
=1
n!
1
xZx
1
ds ϕ0(s)ln( x
s)n(74)
And this proves our point of the lemma.
Let us now prove the 5th point. Let us take x1. Thanks to d’Alembert’s
criterion, we have for each s[1, x], limn+
2nln( x
s)n
n!= 0. We use
the formula of ϕn(x) in the previous point (4), and apply the dominated
convergence theorem to prove that:
lim
n+2nϕn(x) = 1
xZx
1
lim
n+
2nln( x
s)n
n!ϕ0(s)ds = 0 (75)
From point (3), we can conclude that:
lim
n+2nϕn(x) = lim
n+ 2nϕn(x) + 1
x"n1
X
k=0
2klnk(x)
k!2
3n
n1
X
k=0
3klnk(x)
k!#!(76)
= lim
n+
1
x"n1
X
k=0
lnk(x2)
k!2
3n
n1
X
k=0
lnk(x3)
k!#(77)
=1
xexp(ln(x2)) = x(78)
For 0 x < 1, we have from equation (63), ϕn(x) = x
2nx2
3n, for each
n0. Therefore limn+2nϕn(x) = x. Hence the proof of the point
(5). Let us now prove the point (6). We have for each x > 0:
0ϕ0(x)x(79)
Therefore by integrating the equation above we get:
01
xZx
0
dt ϕ0(t)1
xZx
0
dt t (80)
Therefore
0ϕ1(x)x
2(81)
By induction, we easily conclude that for each n1:
0ϕn(x)x
2n(82)
8
Let us now prove the last point. We use Taylor’s Theorem with integral
form of remainder applied on the exponential function xex.
For x0 and n1:
ex=
n1
X
k=0
xk
k!+1
(n1)! Zx
0
dt etxtn1(83)
So, let us take x1. We can write:
eln(x2)=
n1
X
k=0 ln(x2)k
k!+1
(n1)! Zln(x2)
0
dt etln(x2)tn1(84)
We do a change of variable and write for ln(x2):
x2=
n1
X
k=0
2klnk(x)
k!+2n
(n1)! Zx
1
dt tln( x
t)n1(85)
Therefore
x
2n=1
x2n
n1
X
k=0
2klnk(x)
k!+1
x(n1)! Zx
1
dt tln( x
t)n1(86)
And the same for ln(x3):
x2
3n=1
x3n
n1
X
k=0
3klnk(x)
k!+1
x(n1)! Zx
1
dt t2ln( x
t)n1(87)
Therefore
ψn(x) = x
2nx2
3n=1
2nx
n1
X
k=0
2klnk(x)
k!1
3nx
n1
X
k=0
3klnk(x)
k!+1
x(n1)! Zx
1
dt (tt2)ln( x
t)n1
| {z }
ψn(x)
(88)
Lemma 1.4. Let us consider two variables σand τsuch that 0< σ < 1
2
and τ > 0such that s=σ+iτis a zeta zero. Let us define the sequence
of functions En,σ(x),Fn,σ (x)and Gn,σ (x)over [1,+)for each n0
as follows:
En,σ(x) = Zx
1
dt cos(τln (t))
t2+σϕn(t) (89)
Fn,σ(x) = Zx
1
dt sin(τln (t))
t2+σϕn(t) (90)
Gn,σ(x) = En,σ (x)Fn,1σ(x)En,1σ(x)Fn,σ (x) (91)
1. There exists a0>1, there exists an integer n0such that for each
nn0:
Gn,σ(a0)>0 (92)
9
2. For large enough nthere exists xn> a0such as:
Gn,σ(xn) = 0 (93)
3. If the sequence (xn)is unbounded then:
σ=1
2(94)
4. If the sequence (xn)is bounded then:
σ=1
2(95)
Proof. Let us prove the first point. Let us prove it by contradiction. So
let us assume that the opposite is true. Therefore for each a0>1, for
each integer n0, there exists nn0such that:
Gn,σ(a0)0 (96)
So, let us take a0>1. Therefore for each n1, there exists knnsuch
that:
Gkn(a0+1
n)0 (97)
By construction we have:
lim
n+kn= +(98)
We apply the dominated convergence theorem on the sequences of func-
tions:
f1
n(x) = cos(τln (t))
t2+σ2nϕn(t) (99)
f2
n(x) = sin(τln (t))
t2+σ2nϕn(t) (100)
f3
n(x) = cos(τln (t))
t3σ2nϕn(t) (101)
f4
n(x) = sin(τln (t))
t3σ2nϕn(t) (102)
From lemma 1.3, we have:
lim
n+2nϕn(x) = x(103)
Therefore for each x1:
lim
n+f1
kn(x) = f1(x) = cos(τln (x))
x1+σ(104)
lim
n+f2
kn(x) = f2(x) = sin(τln (x))
x1+σ(105)
lim
n+f3
kn(x) = f3(x) = cos(τln (x))
x2σ(106)
lim
n+f4
kn(x) = f4(x) = sin(τln (x))
x2σ(107)
10
Thanks to the Dominated Convergence Theorem applied on the sequence
of functions in (99 102) over the interval [1, a0], we apply the limit to
both sides of the equation(97) as follows:
lim
n+Za0+1
n
1
dt f 1
n(x)Za0+1
n
1
dt f 4
n(x)lim
n+Za0+1
n
1
dt f 2
n(x)Za0+1
n
1
dt f 3
n(x)(108)
Za0
1
dt cos(τln (t))
t1+σZa0
1
dt sin(τln (t))
t2σZa0
1
dt cos(τln (t))
t2σZa0
1
dt sin(τln (t))
t1+σ(109)
From the lemma 1.1 result, we conclude:
g(a0) = J(a0,1 + σ, τ )I(a0,2σ, τ )J(a0,2σ, τ)I(a0,1 + σ, τ )0 (110)
Where the function gis defined as follows:
g(x) = K(1 + σ, τ )K(2 σ, τ ) 1 + σ(1 σ)
τ2sin(τln (x))
xσ11
x12σ(111)
+(1 2σ)
τ"cos(τln (x))
xσ1 + 1
x12σ1 + 1
x#! (112)
Since a0is arbitrary, therefore for each x1:
g(x)0 (113)
Which is a contradiction since the function goscillates between negative
and positive values over [1,+). In fact, from lemma 2.4 (see Appendix),
there exists a constant τ01.9208 such that for each τ > τ0, we have
g(e5π
2τ)>0 for each σ < 1
2. And for ττ0, there is no zeta zero.
Therefore there is a0such that g(a0)>0 and hence the first point of
the lemma is proved.
Let us now prove the 2nd point of the lemma. From the lemma 1.3, thanks
to the results in (43) and (44), we can write:
lim
x+Gn,σ(x) = Z+
1
dt cos(τln (t))
t2+σϕn(t)Z+
1
dt sin(τln (t))
t3σϕn(t)(114)
Z+
1
dt cos(τln (t))
t3σϕn(t)Z+
1
dt sin(τln (t))
t2+σϕn(t)(115)
= σ
2ns2+(1 σ)
3n1s2! τ
2n1s2τ
3ns2!(116)
(1 σ)
2n1s2+σ
3ns2! τ
2ns2τ
3n1s2!(117)
Therefore, we can write for large enough n:
lim
x+Gn,σ(x) τ(1 2σ)
s(1 s)2
1
4n<0 (118)
11
Since the function Gn,σ is continuous over [1,+). From (92) and
(118) and thanks to the Mean value theorem, we can conclude that there
exists an xn> a0such that:
Gn,σ(xn) = 0 (119)
Let us prove the 3rd point. So, let us assume the sequence (xn) is
unbounded. There exists a subsequence3(xλ(n)) that tends to infinity
+. Without loss of generality, we assume that the limit of the sequence
(xn) is +. From the equation (93), we can deduce that:
Zxn
1
dt cos(τln (t))
t2+σϕn(t)Zxn
1
dt sin(τln (t))
t3σϕn(t)=Zxn
1
dt cos(τln (t))
t3σϕn(t)Zxn
1
dt sin(τln (t))
t2+σϕn(t)(120)
We multiply the two side of the equation above by 4n, we can write:
Zxn
1
dt cos(τln (t))
t2+σ2nϕn(t)Zxn
1
dt sin(τln (t))
t3σ2nϕn(t)=Zxn
1
dt cos(τln (t))
t3σ2nϕn(t)Zxn
1
dt sin(τln (t))
t2+σ2nϕn(t)(121)
We apply the Dominated Convergence Theorem on the sequences of func-
tions f1
n, f 2
n, f 3
n, f 4
ndefined in the equations (99 102) since we have the
dominance condition for the functions fi
1i4:
fi
n(x)1i2
1
t2+σ2nϕn(t)1
t1+σ(122)
fi
n(x)3i4
1
t3σ2nϕn(t)1
t2σ(123)
Thanks to the Dominated Convergence Theorem applied over the in-
terval [1,+), we apply the limit to both sides of the equation(121) as
follows:
lim
n+Zxn
1
dt f 1
n(x)Zxn
1
dt f 4
n(x)= lim
n+Zxn
1
dt f 2
n(x)Zxn
1
dt f 3
n(x)(124)
Therefore
Z+
1
dt lim
n+f1
n(x)Z+
1
dt lim
n+f4
n(x)= lim
n+Z+
1
dt lim
n+f2
n(x)Z+
1
dt lim
n+f3
n(x)(125)
Therefore
Z+
1
dt cos(τln (t))
t1+σZ+
1
dt sin(τln (t))
t2σ=Z+
1
dt cos(τln (t))
t2σZ+
1
dt sin(τln (t))
t1+σ(126)
From the lemma 1.1 result, we conclude:
σ
τK(1 + σ, τ )K(2 σ, τ ) = 1σ
τK(2 σ, τ )K(1 + σ, τ ) (127)
Therefore:
σ= 1 σ=1
2(128)
3Similar to Bolzano–Weierstrass theorem in the case of a bounded sequence.
12
Let us now prove the last point of the lemma. So, let us assume the se-
quence (xn) is bounded by an upper bound A > 0. Thanks to Bolzano–Weierstrass
theorem, there exists a subsequence (xλ(n)) that converges to a finite limit
aa0>1. And without loss of generality, we can assume that the limit
of the sequence (xn) is a.
Let us define the functions g1,g2and g3as follows:
g1(x) = J(x, 1 + σ, τ )I(x, 2σ, τ )J(x, 2σ, τ)I(x, 1 + σ, τ ) (129)
g2(x) = J(x, 1 + σ, τ )I(x, 1σ, τ ) + J(x, σ, τ)I(x, 2σ, τ ) (130)
J(x, 2σ, τ )I(x, σ, τ )J(x, 1σ, τ)I(x, 1 + σ, τ ) (131)
g3(x) = J(x, σ, τ )I(x, 1σ, τ )J(x, 1σ, τ)I(x, σ, τ ) (132)
From the lemma 1.3 point 3) and point 7), we can write for each x1:
ϕn(x) = ψn(x) + ϕn(x)ψn(x)
| {z }
δn(x)
(133)
Where
ψn(x) = x
2nx2
3n(134)
δn(x) = ϕn(x)ψn(x) = 1
(n1)!
1
xZx
1
dt ϕ0(t) + t2tln( x
t)n1(135)
Therefore for each x[1, A]:
|δn(x)| A2
(n1)! ln(x)n1(x1)
xA2ln(A)n1
(n1)! (136)
We inject the equations (133 134) into the equation (120) to write
the following. For large enough n:
Zxn
1
dt (ψn(x) + δn(x)f1(x)
xZxn
1
dt (ψn(x) + δn(x)f4(x)
x=Zxn
1
dt (ψn(x) + δn(x)f2(x)
xZxn
1
dt (ψn(x) + δn(x)f3(x)
x(137)
Where the functions fi1i4are defined in (104 107). Therefore:
Zxn
1
dt ψn(x)f1(x)
xZxn
1
dt ψn(x)f4(x)
xZxn
1
dt ψn(x)f2(x)
xZxn
1
dt ψn(x)f3(x)
x(138)
=Zxn
1
dt δn(x)f2(x)
xZxn
1
dt ϕn(x)f3(x)
x+Zxn
1
dt δn(x)f3(x)
xZxn
1
dt ϕn(x)f2(x)
x(139)
Zxn
1
dt δn(x)f1(x)
xZxn
1
dt ϕn(x)f4(x)
xZxn
1
dt δn(x)f4(x)
xZxn
1
dt ϕn(x)f1(x)
x(140)
Thanks to (136), from the equations (138-140) we can write:
Zxn
1
dt ψn(x)f1(x)
xZxn
1
dt ψn(x)f4(x)
xZxn
1
dt ψn(x)f2(x)
xZxn
1
dt ψn(x)f3(x)
x=O(ln(A)n1
(n1)! ) (141)
We inject the equation (134) into the equation (141). After simplification,
we can write the following:
1
4ng1(xn)1
6ng2(xn) + 1
9ng3(xn) = O(ln(A)n1
(n1)! ) (142)
13
Therefore
3
22ng1(xn)3
2ng2(xn) + g3(xn) = O(9 ln(A)n1
(n1)! ) (143)
Since the limit of the sequence ( 9 ln(A)n1
(n1)! ) is zero thanks to Stirling’s
formula. Since the functions g1,g2and g3are continuous and therefore
bounded over the interval [1, A]. There exists n0such that for each nn0,
we have:
g1(xn) = 0 (144)
g2(xn) = 0 (145)
Therefore
g1(a) = g2(a) = g3(a) = 0 (146)
Since aa0>1, hence σmust be equal to 1 σ.
Therefore σ=1
2. This ends the proof of the Riemann Hypothesis.
1.4 Conclusion
We saw that if sis a zeta zero, then real part (s) can only be 1
2. Therefore
the Riemann’s Hypothesis is true: The non-trivial zeros of ζ(s)have real
part equal to 1
2. In the next article, we will apply the same method to
prove the Generalized Riemann Hypothesis (GRH).
Acknowledgments
I would like to thank Farhat Latrach, Giampiero Esposito, Jacques elinas,
Michael Milgram, eo Ag´elas, Ronald F. Fox, Kim Y.G, Masumi Naka-
jima, Maksym Radziwill and Shekhar Suman for thoughtful comments
and discussions on my paper versions on the RH. All errors are mine.
2 Appendix
Let us denote the following functions:
cos+(x) = max cos(x),0(147)
cos(x) = max cos(x),0(148)
Lemma 2.1. Let us consider two variables σand τsuch that σ > 0and
τ > 0. Let us define two integrals I(+, σ, τ )and I(, σ, τ )as follows:
I(+, σ, τ ) = Z+
1
cos+(τln(x))
x1+σdx (149)
I(, σ, τ ) = Z+
1
cos(τln(x))
x1+σdx (150)
14
Therefore
I(+, σ, τ ) = K(1 + σ, τ )σ
τ+K(1 + σ, τ )eπσ
2τ
1eπσ
τ
(151)
I(, σ, τ ) = K(1 + σ, τ )eπσ
2τ
1eπσ
τ
(152)
Proof. Let us calculate I(+, σ, τ ) as follows:
I(+, σ, τ ) = Z+
1
dx cos+(τln (x))
x1+σ(153)
=Ze
π
2τ
1
dx cos(τln (x))
x1+σ+
+
X
k=0 Ze
(4k+5)π
2τ
e
(4k+3)π
2τ
dx cos(τln x)
x1+σ(154)
=Ze
π
2τ
1
dx cos(τln (x))
x1+σ+
+
X
k=0
I(k, σ, τ ) (155)
Where
I(k, σ, τ ) = Ze
(4k+5)π
2τ
e
(4k+3)π
2τ
dx cos(τln x)
x1+σ(156)
From lemma 1.1, we can write the following:
I(k, σ, τ ) = K(1 + σ, τ) e(4k+5)πσ
2τ+e(4k+3)πσ
2τ!(157)
Therefore
I(+, σ, τ ) = K(1 + σ, τ ) σ
τ+eπσ
2τ!+K(1 + σ, τ )
+
X
k=0 e(4k+5)πσ
2τ+e(4k+3)πσ
2τ!(158)
=K(1 + σ, τ ) σ
τ+eπσ
2τ!+K(1 + σ, τ )
+
X
k=0 e5πσ
2τe2kπσ
τ+e3πσ
2τe2kπσ
τ!(159)
=K(1 + σ, τ ) σ
τ+eπσ
2τ!+K(1 + σ, τ )e5πσ
2τ+e3πσ
2τ
1e2πσ
τ
(160)
=K(1 + σ, τ ) σ
τ+eπσ
2τ!+K(1 + σ, τ )e3πσ
2τ1 + eπσ
τ
1e2πσ
τ
(161)
=K(1 + σ, τ ) σ
τ+eπσ
2τ!+K(1 + σ, τ )e3πσ
2τ1
1eπσ
τ
(162)
=K(1 + σ, τ )σ
τ+K(1 + σ, τ )eπσ
2τ 1 + eπσ
τ
1eπσ
τ!(163)
=K(1 + σ, τ )σ
τ+K(1 + σ, τ )eπσ
2τ
1eπσ
τ
(164)
And therefore
15
I(, σ, τ ) = I(+, σ, τ )I(+, σ, τ) (165)
=K(1 + σ, τ )eπσ
2τ
1eπσ
τ
(166)
Lemma 2.2. Let us consider fa continuous function over [1,+). let’s
ϕbe a non-null positive function such that and ϕare integrable func-
tions over [1,+)with:
0<Z+
1
dx ϕ(x)<+(167)
And
Z+
1
dx ϕ(x)f(x)<+(168)
Therefore, there exists a real c(1,+)such that:
Z+
1
dx ϕ(x)f(x) = f(c)Z+
1
dx ϕ(x) (169)
Proof. Let us define the real λas following:
λ=R+
1dx ϕ(x)f(x)
R+
1dx ϕ(x)(170)
We have by construction that:
Z+
1
dx ϕ(x)f(x)λ= 0 (171)
Therefore, if for each x > 1, we have f(x)> λ, then,we will have:
Z+
1
dx ϕ(x)f(x)λ>0 (172)
Which is a contradiction. We will reach a similar contradiction if we
assume f(x)< λ for each x > 1. Therefore, there exists c(1,+) such
that f(c) = λ.
Lemma 2.3. Let us consider two variables σand τsuch that 0< σ < 1
and τ > 0and s=σ+iτis a zeta zero. Therefore:
τ > 3π
2s1 + 1
1 + 1
2pπ
31.9786 (173)
16
Proof. From the lemma 1.3, we have:
Z+
1
dx ϕ(x) cos(τln (x))
x2+σ=σ
σ2+τ2+1σ
(1 σ)2+τ2(174)
Therefore, we can write the following:
Z+
1
dx ϕ(x) cos(τln (x))
x2+σ=Z+
1
dx ϕ(x) cos+(τln (x))
x2+σZ+
1
dx ϕ(x) cos(τln (x))
x2+σ(175)
The function xϕ(x) is continuous over [0,+). Therefore, from
lemma 2.2, there exists c1>1 and c2>1 such that:
Z+
1
dx ϕ(x) cos(τln (x))
x2+σ=ϕ(c1)
c2σ
1Z+
1
dx cos+(τln (x))
x2σϕ(c2)
c2σ
2Z+
1
dx cos(τln (x))
x2σ(176)
Let us denote α1=ϕ(c1)
c2σ
1and α2=ϕ(c2)
c2σ
2. We have 0 < α1<1
4and
0< α2<1
4.
Case 1: α1<=α2In this case we can write α1=α2ϵwith
0< ϵ < 1
4. Therefore we can write from the equation (174) that:
α1Z+
1
dx cos+(τln (x))
x2σα2Z+
1
dx cos(τln (x))
x2σ=σ
σ2+τ2+1σ
(1 σ)2+τ2(177)
Therefore
α1Z+
1
dx cos+(τln (x))
x2σα1Z+
1
dx cos(τln (x))
x2σϵZ+
1
dx cos(τln (x))
x2σ=σ
σ2+τ2+1σ
(1 σ)2+τ2(178)
Therefore
α1Z+
1
dx cos(τln (x))
x2σϵZ+
1
dx cos(τln (x))
x2σ=σ
σ2+τ2+1σ
(1 σ)2+τ2(179)
Therefore
α11σ
(1 σ)2+τ2ϵZ+
1
dx cos(τln (x))
x2σ=σ
σ2+τ2+1σ
(1 σ)2+τ2(180)
Therefore
ϵZ+
1
dx cos(τln (x))
x2σ=σ
σ2+τ2+ (1 α1)1σ
(1 σ)2+τ2(181)
Which is a contradiction.
17
Case 2: α1> α2In this case we can write α1=α2+ϵwith 0 < ϵ < 1
4.
Therefore we can write from the equation (174) that:
α1Z+
1
dx cos+(τln (x))
x2σα2Z+
1
dx cos(τln (x))
x2σ=σ
σ2+τ2+1σ
(1 σ)2+τ2(182)
Therefore
α2Z+
1
dx cos+(τln (x))
x2σα2Z+
1
dx cos(τln (x))
x2σ+ϵZ+
1
dx cos+(τln (x))
x2σ=σ
σ2+τ2+1σ
(1 σ)2+τ2(183)
Therefore
α2Z+
1
dx cos(τln (x))
x2σ+ϵZ+
1
dx cos+(τln (x))
x2σ=σ
σ2+τ2+1σ
(1 σ)2+τ2(184)
Therefore
α21σ
(1 σ)2+τ2+ϵZ+
1
dx cos+(τln (x))
x2σ=σ
σ2+τ2+1σ
(1 σ)2+τ2(185)
Therefore
+ϵZ+
1
dx cos+(τln (x))
x2σ=σ
σ2+τ2+ (1 α2)1σ
(1 σ)2+τ2(186)
From lemma 2.1, we have:
Z+
1
dx cos+(τln (x))
x2σ=1σ
(1 σ)2+τ2+τ
(1 σ)2+τ2
eπ(1σ)
2τ
1eπ(1σ)
τ
(187)
Therefore
ϵ τ
(1 σ)2+τ2
eπ(1σ)
2τ
1eπ(1σ)
τ
=σ
σ2+τ2+ (1 α1)1σ
(1 σ)2+τ2(188)
Therefore
ϵ τ eπ(1σ)
2τ
1eπ(1σ)
τ
=σ(1 σ)2+τ2
σ2+τ2+ (1 α1)1σ(189)
Therefore
ϵ τ eπ(1σ)
2τ
1eπ(1σ)
τ
=σ(1 2σ)
σ2+τ2+σ+1σα11σ(190)
We have 0 < σ < 1
2. Therefore
ϵ τ eπ(1σ)
2τ
1eπ(1σ)
τ
>1α11σ(191)
18
We have 0 < α1<1
4and 1 >1σ > 1
2. Therefore: 1α11σ>3
4.
Therefore
ϵ τ eπ(1σ)
2τ
1eπ(1σ)
τ
>3
4(192)
But we also have 0 < ϵ < 1
4. Therefore
3
4<τ
4
eπ(1σ)
2τ
1eπ(1σ)
τ
(193)
Since ex1 + xfor each x0, therefore:
τ
3> e
π(1σ)
2τeπ(1σ)
2τ(194)
>1 + π(1 σ)
2τ1
1 + π(1σ)
2τ
=π(1 σ)
2τ+
π(1σ)
2τ
1 + π(1σ)
2τ
(195)
Since the function xx
1+xis increasing over [0,+) and π(1σ)
2τπ
4τ,
we can write:
4τ2
3π>1 + 4τ
4τ+π(196)
Since τ > 0, therefore,
τ > 3π
2(197)
Since the function xx
x+πis increasing over [0,+). Therefore
4τ2
3π>1 + 23π
23π+π(198)
Therefore
τ > 3π
2s1 + 1
1 + 1
2pπ
3
(199)
Remark. We can reiterate the same procedure using the equation (196)
to further improve the minimum bound. The limit bound τmin is actually
the root of the equation 16
3πτ3+4
3τ28τπ= 0.τmin 2.01271781.
Lemma 2.4. Let us consider two variables σand τsuch that 0< σ <
1and τ > 0. Let us consider the function gdefined in the equation
(111 112). Therefore there exists a constant τ01.9208 such that:
1. For τ > τ0, we have g(e5π
2τ)>0for each σ < 1
2.
2. For ττ0,s=σ+iτcannot be a zeta zero.
19
Figure 1: Functions f5
2(x, σ) for different values of σfrom 0 to 1
2and function
h5
2(x) over [0,5] .
Proof. Let us prove the first point of the lemma. As we are interested
to the sign of the function g, we will ignore the positive term K(1 +
σ, τ )K(2 σ, τ ) from the gexpression as follows:
g(x) = 1 + σ(1 σ)
τ2sin(τln (x))
xσ11
x12σ(200)
+(1 2σ)
τ"cos(τln (x))
xσ1 + 1
x12σ1 + 1
x#(201)
We can find asuch that g(a)>0 of the form a=eλπ
τwhere λis of the
form λ(k) = 2k+1
2and kN. For such a, we have:
g(a) = 1 + σ(1 σ)
τ2eλπσ
τ1eλπ(12σ)
τ(1 2σ)
τ1 + eλπ
τ(202)
Let us denote x=λπ
τand fλ(x) = g(ex). Therefore
fλ(x, σ) = 1 + σ(1 σ)
λπ2x2e ex(1σ)(1 2σ)
λπ x1 + ex(203)
We can use the asymptotic expansion around 0 (x0) of fλto get
an idea of what is happening around 0. For ϵ > 0 small enough, we can
write:
fλ(ϵ, σ) = 12σ12
λπ ϵ1
2ϵ2+O(ϵ3) (204)
To get fλ(ϵ, σ)>0, we need λ2
πand therefore k1 since 12σ > 0.
For the rest of this section we will take λ=λ(1) = 5
2. We will plot the
function f5
2over [0,5] for different values of σfrom 0 to 1
2in figure 1.
20
From figure 1, it appears that there exists a constant a0such that for
each x(0, a0), we have f5
2(x)>0 for each σ < 1
2. We are going to
prove that this remark holds true. Also, the more σis close to 1
2, the
more fλ(a0, σ) is close to zero. The asymptotic expansion of fλwhen σ
tends to 1
2is as follows:
fλ(x, σ) = 12σxex
21 + x2
(2λπ)2ex
2+ex
2
λπ
| {z }
hλ(x)
+O((σ1
2)3) (205)
The function hλis continous over [0,+). The asymptotic expansion of
hλwhen xtends to 0+is as follows:
hλ(x) = 12
λπ x+O(x2) (206)
Therefore hλ(x)x0+12
λπ x > 0. The limit of hλ(x) when xgoes
to infinity is −∞. Hence, thanks to the mean value theorem, the function
hλhas at least one non-null zero. Let’s x0be the first non-null zero of the
function hλ. Hence, for each x[0, x0], hλ(x)0. Numerically, Newton-
Raphson method locates the value of x0in the interval (4.233,4.234). Now
let us take x[0, x0]. From the equation (203), we can write:
fλ(x, σ) = ex
21 + σ(1 σ)
λπ2x2ex(1
2σ)ex(1
2σ)(1 2σ)
λπ x1 + ex(207)
= 2ex
21 + σ(1 σ)
λπ2x2sinh(x(1
2σ)) (1 2σ)
λπ x1 + ex(208)
Since hλ(x)0, we have:
x2
(2λπ)2ex
2+ex
2
λπ 1 (209)
Also, we have for t0, sinh(t)t, therefore we can write:
fλ(x, σ)2ex
21 + σ(1 σ)
λπ2x2x(1
2σ)(1 2σ)
λπ x1 + ex(210)
(1 2σ)x"ex
21 + 4σ(1 σ)x2
(2λπ)21 + ex
λπ #(211)
(1 2σ)x"ex
2+ 4σ(1 σ)h1 + ex
λπ ex
2i1 + ex
λπ #(212)
(1 2σ)14σ(1 σ)xhex
21 + ex
λπ i(213)
2(1 2σ)14σ(1 σ)xex
2
λπ hλπ
2ex
2+ex
2
2i(214)
2(1 2σ)14σ(1 σ)xex
2
λπ hλπ
2cosh( x
2)i(215)
21
Let us denote x1= 2 arcosh λπ
2= 2 ln λπ
2+q(λπ
2)21. Numerically,
x14.0888 and x1< x0. Therefore for each 0 < x < x1,fλ(x, σ)>0.
Therefore, there exists a constant τ0=5π
2x11.9208 such that for
each τ > τ0, we have g(e5π
2τ)>0 for each σ < 1
2.
Let us prove the second point of the lemma. Thanks to lemma 2.3,
we have τ > τ1=3π
2q1 + 1
1+ 1
2π
31.9786. Therefore if ττ0, then
τ < τ1, therefore s=σ+iτcannot be a zeta zero.
References
[1] Bernhard Riemann. On the Number of Prime Numbers less than a
Given Quantity
https://www.claymath.org/sites/default/files/ezeta.pdf
[2] Aleksandar Ivic. The Riemann Zeta-Function: Theory and Applica-
tions
[3] Peter Borwein, Stephen Choi, Brendan Rooney, and Andrea Weirath-
mueller The Riemann Hypothesis: A Resource for the Afficionado
and Virtuoso Alike
http://wayback.cecm.sfu.ca/~pborwein/TEMP_PROTECTED/book.
pdf
[4] Jørgen Veisdal. The Riemann Hypothesis, explained
https://medium.com/cantors-paradise/
the-riemann-hypothesis-explained- fa01c1f75d3f
[5] Thai Pham. Dirichlet’s Theorem on Arithmetic Progressions
https://web.stanford.edu/ thaipham/papers/MIT 18.104 Review Paper.pdf
[6] G. H. Hardy. The general theory of dirichlet series.
https://archive.org/details/generaltheoryofd029816mbp/page/n9
[7] Garrett, Paul. Primes in arithmetic progressions, 2011.
http : //www.math.umn.edu/ garrett/m/mfms/notes c/dirichlet.pdf
[8] Eissa D. Habil. Double Sequences and Double Series.
https://journals.iugaza.edu.ps/index.php/IUGNS/article/download/1594/1525
22
ResearchGate has not been able to resolve any citations for this publication.
Article
Full-text available
This research considers two traditional important questions, which are interesting, at least to most mathematicians. The first question arises in the theory of double sequences of complex numbers, which concerns the relationship, if any, between the following three limits of a double sequence s : N ◊ N ! C: 1. limn,m!1 s(n,m), 2. limn!1(limm!1 s(n,m)), 3. limm!1(limn!1 s(n,m)). In particular, we'll address the question of when can we interchange the order of the limit for a double sequence {s(n,m)}; that is, when the limit (2) above equals the limit (3) above. The answer to this question is found in Theorem 2.13. The second question arises in the theory of double series of complex numbers, which concerns the relationship, if any, between the following series: 4. P 1 n,m=1 z(n,m), 5. P 1 n=1 ( P 1 m=1 z(n,m)), 6. P 1 m=1 ( P 1 n=1 z(n,m)).
On the Number of Prime Numbers less than a Given Quantity
  • Bernhard Riemann
Bernhard Riemann. On the Number of Prime Numbers less than a Given Quantity https://www.claymath.org/sites/default/files/ezeta.pdf
The Riemann Hypothesis
  • Jørgen Veisdal
Jørgen Veisdal. The Riemann Hypothesis, explained https://medium.com/cantors-paradise/ the-riemann-hypothesis-explained-fa01c1f75d3f
Dirichlet's Theorem on Arithmetic Progressions
  • Thai Pham
Thai Pham. Dirichlet's Theorem on Arithmetic Progressions https://web.stanford.edu/ thaipham/papers/MIT 18.104 Review Paper.pdf
Primes in arithmetic progressions
  • Paul Garrett
Garrett, Paul. Primes in arithmetic progressions, 2011. http : //www.math.umn.edu/ garrett/m/mfms/notes c/dirichlet.pdf