In this paper, we will introduce a new algebraic system called the elliptic complex, and consider the distribution of zeros of the function L(s,χ) in the corresponding complex plane. The key to this article is to discover the limiting case of the Generalized Riemann Hypothesis on elliptic complex fields and, taking a series of elliptic complex fields as variables, to study the ordinary properties of their distributions about the non-trivial zeros of L(s,χ). It is on the basis of these considerations that we will draw the following conclusions. First, the zeros of the function L(s,χ) on any two elliptic complex planes correspond one-to-one. Then, all non-trivial zeros of the L(s,χ) function on each elliptic complex plane are distributed on the critical line ℜ(s)=12 due to the critical case of the Generalized Riemann Hypothesis. Ultimately we proved the Generalized Riemann Hypothesis.
Citation: Xian Hemingway. The Generalized Riemann Hypothesis on elliptic complex fields[J]. AIMS Mathematics, 2023, 8(11): 25772-25803. doi: 10.3934/math.20231315
[1] | Meena Joshi, Anita Tomar, Thabet Abdeljawad . On fixed points, their geometry and application to satellite web coupling problem in $ \mathcal{S}- $metric spaces. AIMS Mathematics, 2023, 8(2): 4407-4441. doi: 10.3934/math.2023220 |
[2] | Editorial Office of AIMS Mathematics . Retraction notice to "The Generalized Riemann Hypothesis on elliptic complex fields" [AIMS Mathematics 8(11) (2023) 25772–25803]. AIMS Mathematics, 2023, 8(11): 27857-27857. doi: 10.3934/math.20231425 |
[3] | XuRan Hai, ShuHong Wang . Hermite-Hadamard type inequalities based on the Erdélyi-Kober fractional integrals. AIMS Mathematics, 2021, 6(10): 11494-11507. doi: 10.3934/math.2021666 |
[4] | Mohammed Guediri, Sharief Deshmukh . Hypersurfaces in a Euclidean space with a Killing vector field. AIMS Mathematics, 2024, 9(1): 1899-1910. doi: 10.3934/math.2024093 |
[5] | Ilija Tanackov, Željko Stević . Calculation of the value of the critical line using multiple zeta functions. AIMS Mathematics, 2023, 8(6): 13556-13571. doi: 10.3934/math.2023688 |
[6] | Weichun Bu, Tianqing An, Guoju Ye, Yating Guo . Nonlocal fractional $ p(\cdot) $-Kirchhoff systems with variable-order: Two and three solutions. AIMS Mathematics, 2021, 6(12): 13797-13823. doi: 10.3934/math.2021801 |
[7] | Atiq Ur Rehman, Ghulam Farid, Sidra Bibi, Chahn Yong Jung, Shin Min Kang . $k$-fractional integral inequalities of Hadamard type for exponentially $(s, m)$-convex functions. AIMS Mathematics, 2021, 6(1): 882-892. doi: 10.3934/math.2021052 |
[8] | Yu-Pei Lv, Ghulam Farid, Hafsa Yasmeen, Waqas Nazeer, Chahn Yong Jung . Generalization of some fractional versions of Hadamard inequalities via exponentially $ (\alpha, h-m) $-convex functions. AIMS Mathematics, 2021, 6(8): 8978-8999. doi: 10.3934/math.2021521 |
[9] | Serap Özcan, Saad Ihsan Butt, Sanja Tipurić-Spužević, Bandar Bin Mohsin . Construction of new fractional inequalities via generalized $ n $-fractional polynomial $ s $-type convexity. AIMS Mathematics, 2024, 9(9): 23924-23944. doi: 10.3934/math.20241163 |
[10] | Zhenjiang Pan, Zhengang Wu . The inverse of tails of Riemann zeta function, Hurwitz zeta function and Dirichlet L-function. AIMS Mathematics, 2024, 9(6): 16564-16585. doi: 10.3934/math.2024803 |
In this paper, we will introduce a new algebraic system called the elliptic complex, and consider the distribution of zeros of the function L(s,χ) in the corresponding complex plane. The key to this article is to discover the limiting case of the Generalized Riemann Hypothesis on elliptic complex fields and, taking a series of elliptic complex fields as variables, to study the ordinary properties of their distributions about the non-trivial zeros of L(s,χ). It is on the basis of these considerations that we will draw the following conclusions. First, the zeros of the function L(s,χ) on any two elliptic complex planes correspond one-to-one. Then, all non-trivial zeros of the L(s,χ) function on each elliptic complex plane are distributed on the critical line ℜ(s)=12 due to the critical case of the Generalized Riemann Hypothesis. Ultimately we proved the Generalized Riemann Hypothesis.
It is well known that Riemann studied the distribution of nontrivial zeros of the function ζ(s) in the complex plane and obtained the conclusion that the nontrivial zeros of ζ(s) were necessarily distributed in the critical region 0<ℜ(s)<1 and symmetric about the point s=1/2. He ventured to guess that the nontrivial zeros of the function ζ(s) were distributed on the critical line ℜ(s)=12. This hypothesis has not been proved so far.
In order to solve the problem of the distribution of primes in arithmetic series, Dirichlet introduced the Dirichlet function L(s,χ). After analysis, mathematicians also gave the conjecture that the zeros of L(s,χ) are all distributed on the line ℜ(s)=1/2.
In the present paper, we will construct the binary algebraic system shown below as a basis for the study of the L-function, which is also the key to solving the Generalized Riemann Hypothesis.
Definition 1.1. [1] Denote by ∀x,y∈R, the numbers of the form z=x+iy as the elliptic complexes, where i satisfies i2=λ,λ∈R−. The set of all elliptic complexes is denoted Cλ.
To avoid confusion we will refer to the elliptic complex with λ=−1 as the circular complex, i.e., the complex C=C−1 invented by the mathematical predecessors.
It is easy to show that all elliptic complexes satisfying the definition are divisible algebraic numbers that satisfy both the multiplicative exchange law and the multiplicative combined law, and are the number fields.
Let z∗=x−iy be the conjugate complex of the complex z with z=x+iy∈Cλ, similarly, while N(z)=zz∗ be the norm of the complex z which obviously satisfies N(z1)N(z2)=N(z1z2). Correspondingly, the distance from the origin to the complex number z is |z|=√N(z), which is also known as the modulus of the complex z. Typically, the distance between any two complex numbers z1 and z2 is |z1−z2|, which can also be derived accordingly.
The calculus is subsequently first introduced into the study of the elliptic complexes to reach some basic conclusions on the theory of the elliptic complex functions where the results not relevant to this paper are not given. These conclusions are used to study the basic properties and theory of the function ζ(s) on the elliptic complex fields.
Suppose that p=−λ=q2 would be made throughout the rest for convenience, and according to the basic operations of elliptic complexes, it is easy to prove[1,2]
ex+iqy=ex(cosy+iqsiny), | (2.1) |
where e is the base of the natural logarithm, and sinθ and cosθ are the sine and cosine functions of θ, respectively. This is Euler's formula on elliptic complex fields whose proof, including some of the details as follows, would be detailed in Appendix A.
For a point A(x,y) on the elliptic complex plane Cλ, let the origin of the coordinates be O, then, the corresponding vector →OA=x+iy. Now, consider the vector →OB=i→OA=−py+ix, which coordinates of point B is B(−py,x). Thus it leads to the fact that
kOA⋅kOB=(−1)×1p=1λ, | (2.2) |
where, with the case of y=0 in particular, easy to get clearly that the vector →OA is on the x-axis and the vector →OB on the y-axis in the complex plane. It can be seen that when λ≠−1, the coordinate system corresponding to the elliptic complex plane is no longer a right-angle coordinate system, but a oblique coordinates system.
It is easy to get that φ=arcsin1q is the angle between the positive y-axis and the positive x-axis in the xOy coordinate system, so that a positive or negative q characterizes the chirality of the coordinate system where in particular the negative q corresponds to the left-handed coordinate system.
In order to clarify the representation of an ellipse in the elliptic complex plane, we need to redefine the ellipse.
Definition 2.1. An ellipse of the form C:x2a2+y2b2=1,a>b>0 is defined to have the (real) eccentricity of e=√1−b2a2 and two real foci of F1=(√a2−b2,0) and F2=(−√a2−b2,0), while the imaginary eccentricity of e′=√1−a2b2 and two imaginary foci of F1′=(0,√a2−b2) and F2′=(0,−√a2−b2).
By the definition of the norm and modulus of the elliptic complex Cλ,
x2N(z)+y2N(z)q2=1⇔x2|z|2+y2(|z|q)2=1, | (2.3) |
whose geometric meaning is an ellipse centred on the origin of the coordinates in the complex plane. In the case where 0<|q|<1, the long semi-axis of the ellipse is m=|z||q| and the short semi-axis is n=|z|, and the direction of the short axis at this point is said to be the direction of the major axis of the ellipse. When |q|≥1, the opposite is true and the direction of the long axis is said to be the direction of the major axis of the ellipse. The length of the long axis (or short axis) in the direction of the major axis is called the major axis length of the ellipse and half of the major axis length is called the principal diameter. An ellipse with the principal semidiameter of 1 is called a unit ellipse.
Definition 2.2. When 0<|q|<1, an ellipse in the complex plane Cλ is said to be the normal ellipse with its (real) eccentricity satisfing e=√|1+1λ| and the direction of its principal axis being parallel to the x-axis. When |q|>1, an ellipse in the complex plane Cλ is said to be the normal ellipse with its imaginary eccentricity satisfing e′=√|1+1λ| and the direction of its principal axis being parallel to the x-axis, at which time the eccentricity of the normal ellipse is also e=√|1+1λ|. In a special way, the normal ellipse of the circular complex plane C−1 is a circle.
It is easy to know that the distance from any point on a normal ellipse to the center of the ellipse is equal and all equal to its principal semidiameter. Further, the normal ellipse with z0 as the centre and r as the principal semidiameter can be expressed as |z−z0|=r.
The normal ellipse is the most fundamental geometric element in the elliptic complex plane, and its use in many proofs in analysis is as important as that of the circle in the circular complex plane. We consider that any geometric object in the complex plane is articulated by a series of normal ellipses Γ1,Γ2,⋯,Γn,⋯, whose corresponding principal diameters r1,r2,⋯,rn,⋯ satisfy max(r1,r2,⋯,rn,⋯)→0.
Where not otherwise specified, all references to ellipses below are to the normal ellipses.
Using the mathematician's method of introducing calculus to circular complexes, we would arrive at the following result.
Theorem 2.3. Let f(z)=u(x,y)+iv(x,y) be a function of the complex variable defined on a region D of the complex plane Cλ, then, a sufficient condition for f(z)f(z) at z=x+iy∈D is that the functions u(x,y),v(x,y) are differentiable at (x,y) and satisfy[2,3]
∂u∂x=∂v∂y,∂u∂y=−q2∂v∂x, | (2.4) |
where q meets i2=λ=−q2(q∈R∗).
The proofs of Theorems 2.3, together with the other propositions and theorems below, are detailed in Appendix B.
The basic elementary functions can be defined on elliptic complex fields next, and it is easy to know that they are all analytic functions.
1) The exponential function: ez=ex(cosy+iqsiny), where z=x+iqy;
2) The trigonometric functions: the cosine and sine functions are defined as[2]
cosz=12(eizq+e−izq), sinz=q2i(eizq−e−izq), | (2.5) |
where it is easy to yield that the zero of the function sinz is kπ (k∈Z) and of the function cosz is (k+12)π (k∈Z) on the complex plane Cλ. Further, there are the following conclusions apparently.
Proposition 2.4. Let Cλ1 and Cλ2 be two arbitrary complex planes, then, the function ez has the same value on Cλ1 and Cλ2 if and only if z is a real number. Similarly, the functions cosz and sinz have the same properties.
Proof. Suppose z=a+ib, then, ez=ea[cos(qb)+iqsin(qb)]. As q is a non-zero constant, the function ez is equal on any two complex planes Cλ1 and Cλ2 only if b=0, i.e., z is a real number, which leads to ez=ea.
The same conclusion can be likewise drawn due to the fact that[2,4]
cosz=q2i(e−qb−eqb)sina+12(e−qb+eqb)cosa, |
which leads to cosz=cosa if b=0, and
sinz=q2i(e−qb−eqb)cosa+12(e−qb+eqb)sina, |
which leads to sinz=sina if b=0.
Now, define sinhz=ez−e−z2 and coshz=ez+e−z2 as the hyperbolic sine function and hyperbolic cosine function of z, respectively. The definitions of the other trigonometric functions are the same as those on the circular complex plane and are of no further interest here.
3) The definition of a argument function on an elliptic complex domain is the same as on a circular complex domain. Define the logarithmic function as
w=Logz=log|z|+iqArgz. | (2.6) |
As for the power and radical functions, they are also derived on the basis of logarithmic functions, which would not be listed here.
Proposition 2.5. Let z∈Cλ, then, z=z∗ holds when λ→0−.
Proof. Since z can be expressed as z=Reiθ, the equation z=z∗ clearly holds when R=0. When R≠0, if λ=−q2→0−, then, based on Eq (A.1) in Appendix A.
zz∗=limλ→0−e2iθ=limq→0[cos(2qθ)+iqsin(2qθ)]=1, |
which means that the equation z=z∗ holds constantly [2,3,5].
In fact, Proposition 2.5 tells us that the geometric plane corresponding to the complex plane Cλ as λ→0− is a plane where the angle between the x-axis and the y-axis tends to zero.
Consider the integral of the function f(z)=u(x,y)+iv(x,y) on the curve C in the elliptic complex plane, which is easily accessible by Eq (2.4).
∫Cf(z)dz=∫Cu(x,y)dx−q2v(x,y)dy+i∫Cu(x,y)dx+v(x,y)dy. | (2.7) |
With regard to the integral operations, the following results are available.
Proposition 2.6. Let C be the elliptical circumference |z−α|=ρ, taking the positive direction of the complex plane, then, we will see that[2]
∫C1(z−α)ndz={2πiq,n=1;0,n≠1,n∈Z, | (2.8) |
which is essential for the derivation of the elliptic complex analysis.
Proof. It is clear that C:|z−α|=ρ, i.e., C:z=α+ρeitq(0≤t≤2π), thus dz=iρqeitqdt. When n=1, the integral sought is
∫C1z−αdz=∫2π0iρqeitqdtρeitq=2πiq. |
As n≠1, the integral is
∫C1(z−α)ndz=∫2π0iρqeitqdtρneintq=iqρn−1∫2π0e−iq(n−1)tdt=0. |
The proposition is thus proved.
Similarly, there is a corresponding Cauchy integral theorem for curves in the elliptic complex plane as follows.
Theorem 2.7. Let C be any simple closed curve in the region D, then,
∫Cf(z)dz=0. | (2.9) |
As a further step, the integration equation can be accessed by combining Eq (2.8) with Cauchy's integral theorem.
Theorem 2.8. Let D be a bounded region bounded by a finite number of simple closed curves C, and the function f(z) be analytic over the closed region ¯D consisting of D and C, then, with ∀z∈D there is the formula
f(z)=q2πi∮Cf(ζ)ζ−zdζ, | (2.10) |
and the higher order derivative formula
f(n)(z)=q⋅n!2πi∮Cf(ζ)(ζ−z)n+1dζ(n=1,2,…). | (2.11) |
With reference to the basic theory of series, the results on the elliptic complex fields are almost identical to those on the circular complex fields and will not be repeated. As a matter of course, the following conclusions are drawn with regard to the residues.
Theorem 2.9. Supposing D being a bounded region in the complex plane whose boundary is a (or a finite composition of) simple closed curves, if the function f(z) is analytic in D except for a finite number of isolated singularities z1,z2,⋯,zn, and also analytic on the boundary C=∂D, then, there is[2,5]
∮Cf(z)dz=2πiqn∑k=1Res(f,zk), | (2.12) |
where the integral along C is drawn in the positive direction* about region D in the complex plane.
* Our thumb passes from the inner side of the complex plane to the outer side, and the other four fingers bend from the positive x-axis to the positive y-axis, so that the direction of bending of the four fingers is the positive direction of the complex plane. It is easily shown that the positive direction of the complex plane is counterclockwise with the case of q>0 corresponding to a right-handed coordinate system, and counterclockwise with the case of q<0 corresponding to a right-handed coordinate system.
The formula for the Fourier integral
f(t)=12π∫∞−∞[∫∞−∞f(s)e−iwsqds]eiwtqdw, | (2.13) |
over the elliptic complex field is easily derived from the Fourier series, thus there is the concept of the Fourier transform.
Definition 2.10.† If the function f(t) satisfies the conditions of the Fourier integral theorem on (−∞,+∞), the function[5,6]
†Correspondingly, the Fourier transform on the elliptic complex field has two forms, the other of which would be derived in Appendix 11.
F(w)=∫∞−∞f(t)e−iwtqdt | (2.14) |
is said to be the Fourier transform of f(t). While the function
F−1(F(w)):=f(t)=12π∫∞−∞F(w)eiwtqdw | (2.15) |
is the inverse Fourier transform of F(w).
Suppose s=c−iqω, then, ω=iq(s−c) and dω=iqds which leads to that
F[i(s−c)]=∫∞−∞e(s−c)tf(t)dt=∫∞−∞est⋅[e−ctf(t)]dt, | (2.16) |
and
f(t)=i2πq∫c−i∞c+i∞e(c−s)tF[i(s−c)]ds=q⋅ect2πi∫c+i∞c−i∞e−stF[i(s−c)]ds. | (2.17) |
Next, suppose again that t=lnx,dt=x−1dx, yielding
F[i(s−c)]=∫∞0xs−1[x−cf(lnx)]dx, | (2.18) |
along with
f(lnx)=q⋅xc2πi∫c+i∞c−i∞x−sF[i(s−c)]ds, | (2.19) |
where with the case of assuming g(x)=x−cf(lnx) and G(s)=F[i(s−c)], we would arrive at
G(s)={Mg}(s)=∫∞0xs−1g(x)dx, | (2.20) |
which is said to be the Merlin transform of g(s), and
g(x)={M−1G}(t)=q2πi∫c+i∞c−i∞x−sG(s)ds, | (2.21) |
which is the inverse Merlin transform of G(s). The presence of a real number c which could be appropriately selected in the inverse conversion formula could avoid possible poles in the integration path.
In a similar way to the study of the Riemann hypothesis, the definition of the arithmetic series is here first extended to the elliptic complex fields.
Definition 3.1. Let χ:Zm∗→Cλ∗ be a group homomorphism satisfying the multiplicative operation [5,7]
χ(ab)=χ(a)χ(b), ∀a,b∈Zm∗. |
Thus call χ the Dirichlet character (mod m) on the elliptic complex field Cλ. For convenience, χ can also be written as χ(n)=χ(n;m). Further, define
L(s,χ)=∞∑n=1χ(n)ns |
to be the Dirichlet function with respect to the character χ on Cλ.
It is easy to conclude that χ(1)=1, and [χ(n)]φ(m)=1, which means that the identity χ is the φ(m)-th root of unity for (n,m)=1 and further that χ(−1)=±1. If χ(−1)=1, then χ is said to be an even character. If χ(−1)=−1, then χ is said to be an odd character. Since χ is a unit root, the inverse of the identity χ satisfies χ−1=¯χ.
Definition 3.2. If χ(n)=1 is constant when (n,m)=1, call it the trivial character, denoted χ0(n). The rest of the characters are termed the non-trivial characters. If the values of them take only real values, they are referred to as the real characters, otherwise they are referred to as complex characters.
Apparently, the real characters on any complex field Cλ1 are the same as those on Cλ2, but the difference is the complex characters. Let χ(n)λ be the character on the complex field Cλ. Then, based on the properties of χ and the Euler formula shown in Eq (A.6), it follows that ℜ[χ(n)λ1]=ℜ[χ(n)λ2] and[1,5,8]
q1⋅ℑ[χ(n)λ1]=q2⋅ℑ[χ(n)λ2], | (3.1) |
where λ1=−q12 and λ2=−q22.
It is easy to know that the n-th unit root on Cλ is
wk=eiq2kπn=(cos2kπn+iqsin2kπn),k=0,1,⋯,n−1, | (3.2) |
where w0=1, when k=0, and the conjugate corresponding to wk is wk∗=wn−k.
Now, note that e(x)=:exp(2πiqx) is a function on Cλ. From the property of the unit root [5,9],
∑k∈Zhe(knh)={h,h|n;0,otherwise. | (3.3) |
Denote the Gauss sum G(n;χ) of the character χ(mod h) by
G(n;χ)=∑k∈Zhχ(k)e(knh), | (3.4) |
which means that, multiplying the Gauss sum by χ(n),
χ(n)G(n;χ)=∑k∈Zhχ(nk)e(knh)=∑m∈nZhχ(m)e(mh), |
with the case of (n,h)=1, which leads to that nZh=Zh. Thus,
G(n;χ)=¯χ(n)G(1;χ). | (3.5) |
Furthermore, combining Eqs (3.3) and (3.5) yields that
|G(1;χ)|2=¯G(1;χ)G(1;χ)=∑k∈Zh¯χ(k)G(1;χ)e(−kh)=∑k∈ZhG(k;χ)e(−kh)=∑k,m∈Zhχ(m)e(kmh)e(−kh)=∑m∈Zhχ(m)∑k∈Zhe(k(m−1)h)=χ(1)∑k∈Zh1=h, |
which leads to |G(1;χ)|=√h.
The definitions of the induced modulus and the primitive character over Cλ are not repeated here.
It is straightforward to show that the function L(s,χ) has no zeros in the region ℜ(s)⩾ and is convergent in the region \Re(s) > 0 when \chi\ne\chi_0 . These results are all consistent with those in the circular complex field.
Suppose that \chi is the character (\bmod\ h) on \mathbb{C}_{\lambda} , and [5,10]
\begin{equation} \varepsilon(\chi) = \begin{cases} 0, &\chi(-1) = 1;\\ 1, &\chi(-1) = -1. \end{cases} \end{equation} | (4.1) |
Define the Dirichlet L-function of the completion, supposing \Gamma_{\mathbb{R}}(s) = \pi^{-s/2}\Gamma(s/2) , as
\begin{equation} \hat{L}(s, \chi) = h^{s\over2}\Gamma_{\mathbb{R}}({s}+\varepsilon(\chi))L(s, \chi), \end{equation} | (4.2) |
which could eliminate the effect of the trivial zeros of L(s, \chi) .
Lemma 4.1. Supposing the function
\begin{equation} \theta(x, a) = \sum\limits_{n = -\infty}^{\infty}{e^{{-(n+a})^2\pi x}}, x > 0, \end{equation} | (4.3) |
it follows that
\begin{equation} \theta\left(\frac{1}{x}, a\right) = \sqrt{x} \sum\limits_{n = -\infty}^{\infty} e^{-\pi x n^{2}+{2 \pi i\over q}na}. \end{equation} | (4.4) |
In particular, \begin{aligned} \theta\left(\frac{1}{x}\right) = \sqrt{x} \theta(x) \end{aligned} when a = 0 , where \theta(x) = \theta(x, 0) .
Proof. Suppose the function f(u) = \exp \left(-\frac{\pi}{x}(u+a)^{2}\right) while u+a = xy . According to Eq ( C.11 ), the Fourier transform of f(u) is [5,11]
\begin{equation*} \begin{aligned} g(v) & = \int_{-\infty}^{\infty} \exp \left(-\frac{\pi}{x}(u+a)^{2}\right) e^{-{2 \pi i\over q} v u} \mathrm d u \\ & = x e^{-\pi x v^{2}+{2 \pi i\over q}va} \int_{-\infty}^{\infty} e^{-\pi x(y+\frac iq v)^{2}} \mathrm d y. \end{aligned} \end{equation*} |
Then, applying the integral theorem as shown in Theorem 2.7 ,
\begin{equation} \int_{-\infty}^{\infty} e^{-\pi x(y+i v)^{2}} \mathrm d y = \int_{-\infty}^{\infty} e^{-\pi x y^{2}} \mathrm d y = \frac{1}{\sqrt{x}}, \end{equation} | (4.5) |
which resulted in \begin{aligned}g(v) = \sqrt{x} e^{-\pi x v^{2}+{2 \pi i\over q}va}\end{aligned} .
Following the Poisson summation formula as shown in Eq ( C.19 ),
\begin{equation} \begin{aligned} \theta(\frac1x, a) = \sum\limits_{n = -\infty}^{\infty}{e^{{-(n+a})^2\frac\pi x}} = \sum\limits_{n = -\infty}^{\infty}{g(n)} = \sqrt{x} \sum\limits_{n = -\infty}^{\infty} e^{-\pi x n^{2}+{2 \pi i\over q}na}. \end{aligned} \end{equation} | (4.6) |
Thus the proposition is proved.
Theorem 4.2. Suppose that \chi is the character (\bmod\ h) on \mathbb{C}_{\lambda} with \lambda = -q^2 , and the function [5,12]
\begin{equation} \psi(x, \chi) = :\sum\limits_{m = -\infty}^{\infty}{\chi(m)e^{-m^2\pi x/h}}, \end{equation} | (4.7) |
when \chi(-1) = 1 , along with the function
\begin{equation} \phi(x, \chi) = :\sum\limits_{m = -\infty}^{\infty}{m\chi(m)e^{-m^2\pi x/h}}, \end{equation} | (4.8) |
when \chi(-1) = -1 . Then, it follows that
\begin{equation} \psi\left(\frac{1}{x}, \chi\right) = \tau(\chi)\left(\frac{x}{h}\right)^{\frac{1}{2}} \psi(x, \overline{\chi}), \end{equation} | (4.9) |
and
\begin{equation} \phi\left(\frac{1}{x}, \chi\right) = -\frac iq \tau(\chi) x\left(\frac{x}{h}\right)^{\frac{1}{2}} \phi(x, \overline{\chi}), \end{equation} | (4.10) |
where \tau(\chi) = G(1;\chi) .
Proof. In accordance with Lemma 4.1 ,
\begin{equation*} \theta(hx, \frac mh) = \sum\limits_{n = -\infty}^{\infty}{e^{-(nh+m)^2\pi x/h}}. \end{equation*} |
Thus, combining Eq ( 3.5 ) gives
\begin{equation*} \begin{aligned} \psi\left(\frac{1}{x}, \chi\right)& = \sum\limits_{m = 1}^{h}{\chi(m)\cdot\theta\left(\frac{h}{x} , \frac mh\right)} = \sqrt{x/h}\sum\limits_{m = 1}^{h}{\chi(m)}\sum\limits_{m = 1}^{h}{e^{-{\pi x n^{2}\over h}+{2 \pi i \over q}{\frac {nm}h}}}\\ & = \left(\frac{x}{h}\right)^{\frac{1}{2}} \sum\limits_{n = -\infty}^{\infty} G(n ; \chi) \exp \left(-\frac{\pi x n^{2}}{h}\right)\\ & = \left(\frac{x}{h}\right)^{\frac{1}{2}} G(1;\chi) \sum\limits_{n = -\infty}^{\infty} \overline\chi(n) \exp \left(-\frac{\pi x n^{2}}{h}\right)\\ & = \left(\frac{x}{h}\right)^{\frac{1}{2}} G(1;\chi) \psi\left(x , \overline\chi\right), \end{aligned} \end{equation*} |
which yields Eq ( 4.9 ).
To differentiate with respect to a on both sides of Eq ( 4.4 ) gives
\begin{equation} \sum\limits_{n = -\infty}^{\infty}(n+a) \exp \left(-\frac{\pi(n+a)^{2}}{x}\right) = -\frac iq x^{\frac{3}{2}} \sum\limits_{n = -\infty}^{\infty} n e^{-\pi x n^{2}+{2 \pi i\over q} na }. \end{equation} | (4.11) |
Consequently,
\begin{equation*} \begin{aligned} \phi\left(\frac{1}{x}, \chi\right) & = h \sum\limits_{m = 1}^{h} \chi(m) \sum\limits_{l = -\infty}^{\infty}\left(l+\frac{m}{h}\right) \exp \left(-\pi \frac{h}{x}\left(l+\frac{m}{h}\right)^{2}\right) \\ & = -\frac iq x\left(\frac{x}{h}\right)^{\frac{1}{2}} \sum\limits_{l = -\infty}^{\infty} G(l ; \chi) l \exp \left(-\frac{\pi x l^{2}}{h}\right) \\ & = -\frac iq G(1;{\chi}) x\left(\frac{x}{h}\right)^{\frac{1}{2}} \phi(x, \overline{\chi}) , \end{aligned} \end{equation*} |
which yields Eq ( 4.10 ).
Due to the fact that \chi(0) = 0 and, for any positive integer m , that \chi(-m) = \chi(-1)\chi(m) = \chi(m) when \chi(-1) = 1 , and (-m)\chi(-m) = m\chi(m) when \chi(-1) = -1 , it follows that
\begin{equation} \psi_1(x, \chi) = \sum\limits_{m = 1}^{\infty}{\chi(m)e^{-m^2\pi x/h}} = \frac12\psi(x, \chi), \end{equation} | (4.12) |
and
\begin{equation} \phi_1(x, \chi) = \sum\limits_{m = 1}^{\infty}{m\chi(m)e^{-m^2\pi x/h}} = \frac12\phi(x, \chi). \end{equation} | (4.13) |
Next use these conclusions to prove the following proposition.
Theorem 4.3. The Dirichlet L-function can be analytically extended to the entire complex plane and satisfies the functional equation
\begin{equation} \hat{L}(s, \chi) = W(\chi)\hat{L}(1-s, \overline\chi), \end{equation} | (4.14) |
where
\begin{equation} W(\chi) = {G(1, \chi)\over \left({i\over q}\right)^{\varepsilon(\chi)}\sqrt{h}} \end{equation} | (4.15) |
is the unit root on the complex field \mathbb{C}_{\lambda} and \chi is the primitive character (\bmod\ h) over \mathbb{C}_{\lambda} with \lambda = -q^2 .
Proof. When \chi(-1) = 1 , according to the Laplace transform formula on the elliptic complex fields which is the same as one on the circular complex field,
\begin{equation} \int_{0}^{\infty}{x^{z-1}e^{-\lambda x}\mathrm dx} = {\Gamma(z)\over \lambda^{z}}. \end{equation} | (4.16) |
Now supposing z = s/2, \lambda = \pi n^2/h ,
\begin{equation*} \int_{0}^{\infty}{x^{s/2-1}e^{-\pi n^2 x/h}\mathrm dx} = {\Gamma(s/2)\over( \pi n^2/h)^{s/2}}, \end{equation*} |
which leads to, following Eq ( 4.12 ), that
\begin{equation*} \begin{aligned} \int_{0}^{\infty}{\psi_1(x, \chi)x^{s/2-1}\mathrm dx}& = \sum\limits_{n = 1}^{\infty}\chi(n)\int_{0}^{\infty}{{e^{-n^2\pi x/h}}x^{s/2-1}\mathrm dx}\\ & = \sum\limits_{n = 1}^{\infty}{\chi(n)\over n^s}{\Gamma(s/2)\over( \pi /h)^{s/2}} = L(s, \chi)\Gamma(s/2)\pi^{-s/2}h^{s/2}.\end{aligned} \end{equation*} |
Thus, by the definition of the function \hat{L}(s, \chi) ,
\begin{equation} \hat{L}(s, \chi) = \frac12\int_{0}^{\infty}{\psi(x, \chi)x^{s/2-1}\mathrm dx}, \end{equation} | (4.17) |
which result in, in combination with Eq ( 4.9 ), that
\begin{equation*} \begin{aligned} \widehat{L}(s, \chi) & = \frac12\int_{0}^{1} \psi(x, {\chi}) x^{\frac{s}{2}-1} \mathrm d x+\frac12\int_{1}^{\infty} \psi(x, {\chi}) x^{\frac{s}{2}-1} \mathrm d x \\ & = \frac12\int_{1}^{\infty} \psi\left(\frac{1}{x}, {\chi}\right) x^{-\frac{s}{2}-1} \mathrm d x+\frac12\int_{0}^{1} \psi(\frac{1}{x}, {\chi}) x^{-\frac{s}{2}-1} \mathrm d x \\ & = \frac12G(1;\chi)h^{-1/2}\int_{0}^{\infty} \psi\left({x}, \overline{\chi}\right) x^{\frac{1-s}{2}-1} \mathrm d x, \end{aligned} \end{equation*} |
thereby providing that
\begin{equation} \widehat{L}(s, \chi) = \frac{G(1, \chi)}{\left({i\over q}\right)^0\sqrt{h}} \widehat{L}(1-s, \overline{\chi}). \end{equation} | (4.18) |
When \chi(-1) = -1 , supposing z = (s+1)/2, \lambda = \pi n^2/h on Eq ( 4.16 ),
\begin{equation*} \int_{0}^{\infty}{x^{(s+1)/2-1}e^{-\pi n^2 x/h}\mathrm dx} = {\Gamma({s+1\over 2})\over( \pi n^2/h)^{(s+1)/2}}, \end{equation*} |
Therefore,
\begin{equation*} \begin{aligned} \int_{0}^{\infty}{\phi_1(x, \chi)x^{(s+1)/2-1}\mathrm dx}& = \sum\limits_{n = 1}^{\infty}n\chi(n)\int_{0}^{\infty}{{e^{-n^2\pi x/h}}x^{(s+1)/2-1}\mathrm dx}\\ & = \sum\limits_{n = 1}^{\infty}{\chi(n)\over n^s}{\Gamma({s+1\over2})\over( \pi /h)^{(s+1)/2}}\\& = h^{1/2}L(s, \chi)\Gamma({s+1\over2})\pi^{-(s+1)/2}h^{(s+1)/2}\\& = h^{1/2}\hat{L}(s, \chi), \end{aligned} \end{equation*} |
which leads to, following Eq ( 4.13 ), that
\begin{equation} \begin{aligned} \hat{L}(s, \chi) = \frac12h^{-1/2}\int_{0}^{\infty}{\phi(x, \chi)x^{(s+1)/2-1}\mathrm dx} \end{aligned}. \end{equation} | (4.19) |
Hence, combining Eq ( 4.10 ),
\begin{equation*} \begin{aligned} \hat{L}(s, \chi)& = \frac12h^{-1/2}\int_{0}^{1}{\phi(x, \chi)x^{(s+1)/2-1}\mathrm dx}+\frac12h^{-1/2}\int_{1}^{\infty}{\phi(x, \chi)x^{(s+1)/2-1}\mathrm dx}\\ & = \frac12h^{-1/2}\int_{1}^{\infty}{\phi(x^{-1}, \chi)x^{-(s+1)/2-1}\mathrm dx}+\frac12h^{-1/2}\int_{0}^{1}{\phi(x^{-1}, \chi)x^{-(s+1)/2-1}\mathrm dx}\\ & = -\frac12h^{-1/2}\frac iqG(1;\chi)h^{-1/2}\int_{0}^{\infty}{\phi(x, \overline\chi)x^{-s/2}\mathrm dx}, \end{aligned} \end{equation*} |
that is,
\begin{equation} \hat{L}(s, \chi) = {G(1, \chi)\over \frac iqh^{1/2}}\hat{L}(1-s, \overline\chi). \end{equation} | (4.20) |
In summary, the proposition is proved.
Under Theorem 4.3 , the function \hat{L}(s, \chi) is invariant under the substitution s\to1-s, \chi\to\overline{\chi} , which leads to the following conclusion.
Corollary 4.4. When \chi is a real character, the zeros of the function \hat{L}(s, \chi) are symmetric with respect to \begin{aligned}s = \frac12\end{aligned} and the function \begin{aligned}\hat L(\frac{1}{2}+it, \chi)\end{aligned} is a real even function.
By Proposition 2.5 , the character \chi on the complex field \mathbb{C}_{\lambda} satisfies \overline\chi = \chi when \lambda\to0^- , so there is the following conclusion.
Corollary 4.5. When \lambda\to0^- , the zeros of the function \hat{L}(s, \chi) are symmetric with respect to \begin{aligned}s = \frac12\end{aligned} if \chi is a complex character on the field \mathbb{C}_{\lambda} .
When \chi(-1) = -1 , suppose \xi(s, \chi) = h^{s/2}\hat{L}(s, \chi) for convenience of the next operation.
Without the special remarks, in the following we assume that \chi is the primitive character (\bmod \ h) with h{\geqslant}3 over the complex field \mathbb{C}_{\lambda} and that \begin{aligned} s = \sigma+\frac iqt\in\mathbb{C}_\lambda \end{aligned} . The Generalized Riemann Hypothesis is noted as GRH later for convenience [5].
When \chi(-1) = 1 , depending on Eq ( 4.17 ),
\begin{equation} h^{s/2}\pi^{-s/2}\Gamma(s/2) L(s, \chi) = \frac12\int_{0}^{\infty}{\psi(x, \chi)x^{s/2-1}\mathrm dx} = \int_{0}^{\infty}{\psi(x^2, \chi)x^{s-1}\mathrm dx}. \end{equation} | (5.1) |
According to Eq ( 2.20 ), Eq ( 5.1 ) is exactly of the form of a Merlin transformation, i.e.,
\begin{equation} h^{s/2}\pi^{-s/2}\Gamma(s/2) L(s, \chi) = \{\mathcal M\psi(x^2, \chi)\}(s). \end{equation} | (5.2) |
Therefore, using the inverse formula of the Merlin transform,
\begin{equation} \psi(x^2, \chi) = {q\over2\pi i}\int_{\frac12-i\infty}^{\frac12+i\infty}h^{s/2}\pi^{-s/2}\Gamma(s/2) L(s, \chi)x^{-s}\mathrm ds, \end{equation} | (5.3) |
due to the fact that the integrand has no poles at \sigma > 0 . Replacing x\rightarrow e^u gives
\begin{equation} \psi(e^{2u}, \chi) = {q\over2\pi i}\int_{\frac12-i\infty}^{\frac12+i\infty}h^{s/2}\pi^{-s/2}\Gamma(s/2) L(s, \chi)e^{-us}\mathrm ds. \end{equation} | (5.4) |
And now let \begin{aligned} s = \frac12+{i\over q}t \end{aligned} , then Eq ( 5.4 ) can be reduced to
\begin{equation} \underbrace{\psi(e^{2u}, \chi)e^{u/2}}_{\Psi_1 \left(\frac u2, \chi\right)} = {1\over2\pi}\int_{-\infty}^\infty e^{-{i\over q}ut}\hat{L}\left(\frac12+{i\over q}t, \chi\right)\mathrm ds, \end{equation} | (5.5) |
where, according to Corollary 4.4 , the function
\begin{equation*} \begin{aligned}\Psi_1 \left(\frac u2, \chi\right) = \psi(e^{2u}, \chi)e^{u/2} = \sum\limits_{n = -\infty}^{\infty}{\chi(n)e^{-\pi n^2e^{2u}/h+u/2}} \end{aligned} \end{equation*} |
is also an even function, so the combination of the Fourier inversion formula as seen in Definition 2.10 shows that
\begin{equation*} \begin{aligned} \hat{L}\left(\frac12+{i\over q}t, \chi\right) & = \int_{-\infty}^\infty\Psi_1\left(\frac u2, \chi\right)e^{{i\over q}ut}\mathrm du = 2\int_0^\infty\Psi_1\left(\frac u2, \chi\right)\cos(ut)\mathrm du \\ & = 4\int_0^\infty\Psi_1\left( u, \chi\right)\cos(2xt)\mathrm du. \end{aligned} \end{equation*} |
Supposing \begin{aligned} _{0}^{1} H(\varpi, z) = \int_0^\infty{e^{\varpi u^2}\Psi_1\left(u, \chi\right)\cos(zu)\mathrm du} \end{aligned} ,
\begin{equation} _{0}^{1} H(0, z) = {1\over4}\hat{L}\left(\frac12+{iz\over2q}, \chi\right), \end{equation} | (5.6) |
which the following conclusion can be drawn by combining the properties of \hat{L}(s, \chi) by.
Proposition 5.1. Let \chi be a real primitive character (\bmod\ h) with h{\geqslant}3 over \mathbb{C}_{\lambda} and satisfy \chi(-1) = 1 . The GRH with respect to \chi holds when and only when all zeros of the function _{0}^{1} H(0, z) are real.
Based on the definition of the function _{0}^{1} H(0, z) , the following conclusion holds in conjunction with Proposition 2.4 .
Proposition 5.2. Let \mathbb{C}_{\lambda_1} and \mathbb{C}_{\lambda_2} be any two complex planes. The function _{0}^{1} H(0, z) has equal values in \mathbb{C}_{\lambda_1} and \mathbb{C}_{\lambda_2} provided that z is a definite real number.
When \chi(-1) = -1 , following Eq ( 4.19 ),
\begin{equation} h^{s}\pi^{-(s+1)/2}\Gamma(\frac {s+1}2) L(s, \chi) = \frac12\int_{0}^{\infty}{\phi(x, \chi)x^{(s+1)/2-1}\mathrm dx} = \int_{0}^{\infty}{\phi(x^2, \chi)x^{s}\mathrm dx}. \end{equation} | (5.7) |
Similarly, Eq ( 5.7 ) is exactly of the form of a Merlin transformation, that is
\begin{equation} h^{s-1}\pi^{-s/2}\Gamma(s/2) L(s-1, \chi) = \{\mathcal M\phi(x^2, \chi)\}(s), \end{equation} | (5.8) |
which leads to, using the inverse formula of the Merlin transform, that
\begin{equation} \phi(x^2, \chi) = {q\over2\pi i}\int_{\frac12-i\infty}^{\frac12+i\infty}h^{s-1}\pi^{-s/2}\Gamma(s/2) L(s-1, \chi)x^{-s}\mathrm ds. \end{equation} | (5.9) |
Replacing Eq ( 5.9 ) by x\rightarrow e^u yields[5]
\begin{equation} \phi(e^{2u}, \chi) = {q\over2\pi i}\int_{\frac12-i\infty}^{\frac12+i\infty}h^{s-1\over 2}\hat L(s-1, \chi)e^{-us}\mathrm ds. \end{equation} | (5.10) |
Hereby let \begin{aligned} s-1 = \frac12+{i\over q}t \end{aligned} , then, Eq ( 5.10 ) can be turned into
\begin{equation} \underbrace{\phi(e^{2u}, \chi)e^{3u/2}}_{\Psi_2 \left(\frac u2, \chi\right)} = {1\over2\pi}\int_{-\infty}^\infty e^{-{i\over q}ut}\xi\left(\frac12+{i\over q}t, \chi\right)\mathrm ds, \end{equation} | (5.11) |
where the function
\begin{equation*} \begin{aligned}\Psi_2 \left(\frac u2, \chi\right) = \phi(e^{2u}, \chi)e^{3u/2} = \sum\limits_{n = -\infty}^{\infty}{n\chi(n)e^{-\pi n^2e^{2u}/h+3u/2}} \end{aligned} \end{equation*} |
is equally an even function, resulting in that
\begin{equation*} \begin{aligned} \xi\left(\frac12+{i\over q}t, \chi\right) & = \int_{-\infty}^\infty\Psi_2\left(\frac u2, \chi\right)e^{{i\over q}ut}\mathrm du = 2\int_0^\infty\Psi_2\left(\frac u2, \chi\right)\cos(ut)\mathrm du \\ & = 4\int_0^\infty\Psi_2\left( u, \chi\right)\cos(2xt)\mathrm du , \end{aligned} \end{equation*} |
Given the function \begin{aligned} _{0}^{2}H(\varpi, z) = \int_0^\infty{e^{\varpi u^2}\Psi_2\left(u, \chi\right)\cos(zu)\mathrm du} \end{aligned} ,
\begin{equation} _{0}^{2}H(0, z) = {1\over4}\xi\left(\frac12+{iz\over2q}, \chi\right), \end{equation} | (5.12) |
which likewise brings us to the following conclusion.
Proposition 5.3. Let \chi be a real primitive character (\bmod\ h) with h{\geqslant}3 over \mathbb{C}_{\lambda} and satisfy \chi(-1) = -1 . The GRH with respect to \chi holds when and only when all zeros of the function _{0}^{2} H(0, z) are real.
Proposition 5.4. Let \mathbb{C}_{\lambda_1} and \mathbb{C}_{\lambda_2} be any two complex planes. The function _{0}^{2} H(0, z) has equal values in \mathbb{C}_{\lambda_1} and \mathbb{C}_{\lambda_2} provided that z is a definite real number.
The zeros of the function \hat{L}(s, \chi) corresponding to the complex characters are not necessarily symmetric about the point s = 1/2 , so its equivalence condition about the GRH will be discussed later.
Since the function \hat{L}(s, \chi) has no poles in the region \Re(s)\in(0, 1) , allowing \Delta\in(0, 1) to be a given constant, Eqs ( 5.4 ) and ( 5.10 ) can be easily transformed into that
\begin{equation} \psi(e^{2u}, \chi) = {q\over2\pi i}\int_{\Delta-i\infty}^{\Delta+i\infty}\hat L(s, \chi)e^{-us}\mathrm ds, \end{equation} | (6.1) |
which corresponds to the case of even character, along with that [5,6,13]
\begin{equation} \phi(e^{2u}, \chi) = {q\over2\pi i}\int_{\Delta-i\infty}^{\Delta+i\infty}h^{s-1\over 2}\hat L(s-1, \chi)e^{-us}\mathrm ds, \end{equation} | (6.2) |
corresponding to the case of the odd character. To let \begin{aligned} s = \Delta+\frac iq t \end{aligned} on Eq ( 6.1 ), it follows that
\begin{equation} \underbrace{\psi(e^{2u}, \chi)e^{u\Delta}}_{\Psi_1^\Delta \left(\frac u2, \chi\right)} = {1\over2\pi}\int_{-\infty}^\infty e^{-{i\over q}ut}\hat{L}\left(\Delta+{i\over q}t, \chi\right)\mathrm ds. \end{equation} | (6.3) |
And suppose \begin{aligned} s-1 = \Delta+\frac iq t \end{aligned} on Eq ( 6.2 ), to find that
\begin{equation} \underbrace{\phi(e^{2u}, \chi)e^{u(\Delta+1)}}_{\Psi_2^\Delta \left(\frac u2, \chi\right)} = {1\over2\pi}\int_{-\infty}^\infty e^{-{i\over q}ut}\xi\left(\Delta+{i\over q}t, \chi\right)\mathrm ds. \end{equation} | (6.4) |
From Eq ( 6.3 ) combined with the inverse Fourier transform formula,
\begin{equation} \begin{aligned} \hat {L}\left(\Delta+{i\over q}t, \chi\right) & = \int_{-\infty}^\infty\Psi_1^\Delta\left(\frac u2, \chi\right)e^{{i\over q}ut}\mathrm du = 2\int_{-\infty}^\infty\Psi_1^\Delta\left(u, \chi\right)e^{{i\over q}u(2t)}\mathrm du. \end{aligned} \end{equation} | (6.5) |
Now suppose that
\begin{equation*} \begin{aligned} { }^{1} H^{\Delta}(\varpi, z) = \int_{-\infty}^\infty{e^{\varpi u^2}\Psi_1^\Delta\left( u, \chi\right)\cos (zu)\mathrm du, } \end{aligned} \end{equation*} |
and that
\begin{equation*} \begin{aligned} { }^{1} F^{\Delta}(\varpi, z) = \int_{-\infty}^\infty{e^{\varpi u^2}\Psi_1^\Delta\left( u, \chi\right)\sin (zu)\mathrm du}. \end{aligned} \end{equation*} |
Depending on the definition of the trigonometric function as seen in Eq ( 2.5 ),
\begin{equation} { }^{1} H^{\Delta}(0, z) = {1\over2}\hat{L}\left(\Delta+{iz\over2q}, \chi\right)+{1\over2}\hat{L}\left(\Delta-{iz\over2q}, \chi\right) = \Re\left[ \hat{L}\left(\Delta+{iz\over2q}, \chi\right) \right], \end{equation} | (6.6) |
and
\begin{equation} { }^{1} F^{\Delta}(0, z) = {1\over2}\hat{L}\left(\Delta+{iz\over2q}, \chi\right)-{1\over2}\hat{L}\left(\Delta-{iz\over2q}, \chi\right) = q\Im\left[ \hat{L}\left(\Delta+{iz\over2q}, \chi\right) \right], \end{equation} | (6.7) |
which could lead to that
\begin{equation} \begin{aligned} \hat{L}\left(\Delta+{i\over 2q}z, \chi\right) = { }^{1} H^{\Delta}(0, z)+\frac iq\cdot{ }^{1} F^{\Delta}(0, z), \end{aligned} \end{equation} | (6.8) |
where \begin{aligned} \hat{L}\left(\Delta+{i\over 2q}z, \chi\right) = 0\end{aligned} if and only if { }^{1} H^{\Delta}(0, z) = 0 and { }^{1} F^{\Delta}(0, z) = 0 .
By Proposition 2.4 , the functions { }^{1} H^{\Delta}(0, z) or { }^{1} F^{\Delta}(0, z) have the same value on any two complex planes \mathbb{C}_{\lambda_1} and \mathbb{C}_{\lambda_2} when z is a definite real number, leading to the following conclusion.
Proposition 6.1. The zeros of the function \hat{L}(s, \chi) on any two complex planes \mathbb{C}_{\lambda_1} and \mathbb{C}_{\lambda_2} are in one-to-one correspondence. Let \lambda_1 = -q_1{}^{2} and \lambda_2 = -q_2{}^{2} . Precisely, if \begin{aligned}s = \Delta+{i\over q_1}z\end{aligned} , where z is a real number, is the zero of the function \hat{L}(s, \chi) on the complex plane \mathbb{C}_{\lambda_1} , then \begin{aligned}s^{'} = \Delta+{i\over q_2}z\end{aligned} has to be the zero of the function \hat{L}(s, \chi) on the complex plane \mathbb{C}_{\lambda_2} .
If we set \Delta = \frac12+R\cos\theta and z = R\sin\theta , where R{\geqslant}0 , it follows that \begin{aligned}s = \frac12+Re^{\frac i{q_1}\theta}\end{aligned} and \begin{aligned}s^{'} = \frac12+Re^{\frac i{q_2}\theta}\end{aligned} . Based on the definition of the normal ellipse, there is the following conclusion.
Corollary 6.2. If the zero of the function \hat{L}(s, \chi) on the complex plane \mathbb{C}_{\lambda_1} is on the normal ellipse centred at the point \begin{aligned}s = \frac12\end{aligned} and with R{\geqslant}0 as its principal semi-diameter, then its zero on the complex plane \mathbb{C}_{\lambda_2} is also on the normal ellipse centred at the point \begin{aligned}s = \frac12\end{aligned} and with R as its principal semi-diameter correspondingly [6,13].
Equivalently, when \chi is an odd character, it follows from Eq ( 6.4 ) combined with the inverse Fourier transform formula, it follows that
\begin{equation} \begin{aligned} \xi\left(\Delta+{i\over q}t, \chi\right) & = \int_{-\infty}^\infty\Psi_2^\Delta\left(\frac u2, \chi\right)e^{{i\over q}ut}\mathrm du = 2\int_{-\infty}^\infty\Psi_2^\Delta\left(u, \chi\right)e^{{i\over q}u(2t)}\mathrm du. \end{aligned} \end{equation} | (6.9) |
Now assume that
\begin{equation*} \begin{aligned} { }^{2} H^{\Delta}(\varpi, z) = \int_{-\infty}^\infty{e^{\varpi u^2}\Psi_2^\Delta\left( u, \chi\right)\cos (zu)\mathrm du}, \end{aligned} \end{equation*} |
and that
\begin{equation*} \begin{aligned} { }^{2} F^{\Delta}(\varpi, z) = \int_{-\infty}^\infty{e^{\varpi u^2}\Psi_2^\Delta\left( u, \chi\right)\sin (zu)\mathrm du} , \end{aligned} \end{equation*} |
which could analogously lead to that
\begin{equation} \begin{aligned} \xi\left(\Delta+{i\over 2q}z, \chi\right) = { }^{2} H^{\Delta}(0, z)+\frac iq\cdot{ }^{2} F^{\Delta}(0, z), \end{aligned} \end{equation} | (6.10) |
where \begin{aligned} \xi\left(\Delta+{i\over 2q}z, \chi\right) = 0\end{aligned} if and only if { }^{2} H^{\Delta}(0, z) = 0 and { }^{2} F^{\Delta}(0, z) = 0 .
The following conclusions can be drawn in the same way.
Proposition 6.3. The zeros of the function \xi(s, \chi) on any two complex planes \mathbb{C}_{\lambda_1} and \mathbb{C}_{\lambda_2} are in one-to-one correspondence. Precisely, if \begin{aligned}s = \Delta+{i\over q_1}z\end{aligned} , where z is a real number, is the zero of the function \xi(s, \chi) on the complex plane \mathbb{C}_{\lambda_1} , then \begin{aligned}s^{'} = \Delta+{i\over q_2}z\end{aligned} has to be the zero of the function \xi(s, \chi) on the complex plane \mathbb{C}_{\lambda_2} .
Corollary 6.4. If the zero of the function \xi(s, \chi) on the complex plane \mathbb{C}_{\lambda_1} is on the normal ellipse centred at the point \begin{aligned}s = \frac12\end{aligned} and with R{\geqslant}0 as its principal semi-diameter, then its zero on the complex plane \mathbb{C}_{\lambda_2} is also on the normal ellipse centred at the point \begin{aligned}s = \frac12\end{aligned} and with R as its principal semi-diameter correspondingly [6].
Obviously, Propositions 6.1 and 6.3 , along with Corollaries 6.2 and 6.4 above, hold irrespective of whether \chi is a real or complex character.
From the above analysis, it is easy to conclude that the equivalent proposition of the GRH corresponding to the complex character is as follows.
Proposition 6.5. Let \begin{aligned} \Delta = \frac 12 \end{aligned} and \chi be the complex primitive character (\bmod\ h) with h{\geqslant}3 over \mathbb{C}_{\lambda} . If \chi(-1) = 1 , the corresponding GRH holds when and only when the zeros of the function \begin{aligned} { }^{1} H^{\Delta}(0, z)+\frac iq\cdot{ }^{1} F^{\Delta}(0, z) \end{aligned} are all real. Additionally, in the event that \chi(-1) = -1 , the corresponding GRH holds when and only when the zeros of the function \begin{aligned} { }^{2} H^{\Delta}(0, z)+\frac iq\cdot{ }^{2} F^{\Delta}(0, z) \end{aligned} are all real.
According to our previous definition of the elliptic complex \mathbb{C}_\lambda , the equation corresponding to the norm degenerates to the form x^2 = N(z) when \lambda = 0 , representing two straight lines symmetric about the y-axis perpendicular to the x-axis in the complex plane. Considering that any geometric feature in the complex plane is articulated by a corresponding normal ellipse, and that the normal ellipse on the complex plane \mathbb{C}_\lambda corresponding to \lambda = 0 is an ellipse whose principal axis, i.e., the x-axis, is infinitely compressed, the figure articulated by such ellipses can only be a line perpendicular to the x-axis. However, it is clearly inappropriate to study the complex number \mathbb{C}_\lambda directly with \lambda = 0 , because such an algebraic system is not divisible.
When \lambda\rightarrow0^{-} , the complex number \mathbb{C}_\lambda is clearly a divisible algebra, which of course satisfies Theorem 4.3 , and the norm of the complex number s = x+iy is
\begin{equation*} N(s) = \lim\limits_{\lambda \rightarrow 0^{-}}{(x^2-\lambda y^2)} = x^2. \end{equation*} |
From the symmetry of the function \hat L(s, \chi) for the case of \chi(-1) = 1 and \xi(s, \chi) for the case of \chi(-1) = -1 , it follows that the zeros of \hat L(s, \chi) as well as \xi(s, \chi) are symmetric about the point s = 1/2 . Therefore, the equation \hat L(s, \chi) = 0 , together with \xi(s, \chi) = 0 , holds if and only if N(s) = x^2 = 1/4 , i.e., \begin{aligned} \Re(s) = {1\over 2} \end{aligned} when \lambda\rightarrow0^{-} .
Pursuant to Corollaries 4.4 and 4.5 , this result holds for both real and complex characters.
The geometric significance of this result can be further obtained from Corollaries 6.2 and 6.4 . Since all the zeros of the function \hat L(s, \chi) or \xi(s, \chi) must occur symmetrically on a series of normal ellipses centred at the point s = 1/2 , thus the geometry of this result is such that, as \lambda\rightarrow0^{-} , all such normal ellipses are compressed into the straight line segments passing through the point s = 1/2 and symmetrical about the x-axis. Therefore, with \lambda\rightarrow0^{-} , all the zeros of the function \hat L(s, \chi) or \xi(s, \chi) are distributed on the critical line \begin{aligned} \Re(s) = {1\over 2} \end{aligned} perfectly naturally.
Now, let \lambda_0 = -p_0\rightarrow0^{-} . The distribution of zeros of the function \hat{L}(s, \chi) or \xi(s, \chi) on the elliptic complex field \mathbb{C}_\lambda with p = -\lambda\in[p_0, 1] would be explored, according to the continuity approach in analysis, while consider such a proposition as follows.
Proposition 7.1. The zeros of the function \hat{L}(s, \chi) or \xi(s, \chi) on the complex plane \mathbb{C}_{\lambda_1} , with the case of p_1 = -\lambda_1\in[p_0, 1) , are all distributed on \begin{aligned} \Re(s) = {1\over 2} \end{aligned} , then the zeros of \hat{L}(s, \chi) or \xi(s, \chi) on the complex plane \mathbb{C}_{\lambda_2} , with the case of p_2 = -\lambda_2\in(p_1, 1] , are all distributed on \begin{aligned} \Re(s) = {1\over 2} \end{aligned} .
If this proposition can be shown, combined with the critical case when \lambda\rightarrow0^{-} , the GRH on elliptic complex fields can also be proved.
In actual fact, there is a more powerful formulation for Proposition 7.1 as the following proposition.
Proposition 7.2. For any \varepsilon > 0 and one real number p = -\lambda\in[p_0, 1) , if the zeros of the function \hat{L}(s, \chi) or \xi(s, \chi) on the complex plane \mathbb{C}_{\lambda} are all distributed on \begin{aligned} \Re(s) = {1\over 2} \end{aligned} , then the zeros of \hat{L}(s, \chi) or \xi(s, \chi) on the complex plane \mathbb{C}_{\lambda_1} are all distributed on \begin{aligned} \Re(s) = {1\over 2} \end{aligned} where \lambda_1 meets p_1 = p+\varepsilon = -\lambda_1 .
Proof. Now let p = q^2 and \varepsilon = \delta^2 where q and \delta are real numbers that are both positive or negative at the same time.
The case of the real characters is discussed at the outset.
As assumed in Proposition 7.2 , in conjunction with Propositions 5.1 and 5.3 , the zeros of the function
\begin{equation*} _{0}^{1} H(0, z) = {1\over4}\hat{L}\left(\frac12+{iz\over2q}, \chi\right)\text{ or } _{0}^{2}H(0, z) = {1\over4}\xi\left(\frac12+{iz\over2q}, \chi\right) \end{equation*} |
on the complex plane \mathbb{C}_{\lambda} are all real. Based on Propositions 5.2 and 5.4 , it follows that these corresponding points are also the zeros of the function
\begin{equation*} _{0}^{1} H_{\varepsilon}(0, z) = {1\over4}\hat{L}\left(\frac12+{iz\over2(q+\delta)}, \chi\right)\text{ or } _{0}^{2}H_{\varepsilon}(0, z) = {1\over4}\xi\left(\frac12+{iz\over2(q+\delta)}, \chi\right) \end{equation*} |
on the complex plane \mathbb{C}_{\lambda_1} .
Combining this with Propositions 6.1 and 6.3 shows that all zeros of the function _{0}^{1} H_{\varepsilon}(0, z) or _{0}^{2} H_{\varepsilon}(0, z) are also real, so that all zeros of the function \hat{L}(s, \chi) or \xi(s, \chi) on the complex plane \mathbb{C}_{\lambda_1} are distributed on \begin{aligned} \Re(s) = {1\over 2} \end{aligned} .
Similarly, Proposition 6.5 , used in combination with Propositions 6.1 and 6.3 , could prove the case for the complex characters.
In the above, the proof of the GRH on elliptic complex fields is complete. This conclusion holds for all elliptic complex planes \mathbb{C}_{\lambda} with p = -\lambda\in(0, +\infty) , including the circular complex plane \mathbb{C} naturally.
As a matter of fact, for any \lambda_1 = -q_1^{2} and \lambda_2 = -q_2^{2} with q_1, q_2\in\mathbb{R^*} , if the imaginary part of the n-th zero of the function \hat{L}(s, \chi) or \xi(s, \chi) on the complex plane \mathbb{C}_{\lambda_1} is \beta_n , then correspondingly the imaginary part of the n-th zero on the complex plane \mathbb{C}_{\lambda_2} is \begin{aligned} \beta_n{}^{'} = {q_1\over q_2}\beta_n \end{aligned} , while their real part is 1/2 .
In this paper, we first construct the elliptic complexes \mathbb{C}_{\lambda} and introduce them into the calculus to obtain the corresponding theory of complex variables' function. Further, the problem of the distribution of the zeros of the function L(s, \chi) on the corresponding elliptic complex fields is discussed, which contributed to the analytic extended form and functional equation of \hat L(s, \chi) .
One of the difficulties of this essay is to find the correspondence between the zeros of the function \hat{L}(s, \chi) or \xi(s, \chi) in the elliptic complex plane.
And the key to this paper is to find a critical case of the GRH bout the real and complex characters on the elliptic complex planes and to discover the equivalent propositions of the GRH. Based on the continuity method in analysis, we eventually proved the GRH on all complex planes due to Proposition 2.5 which shows the basic relationship of functions between the elliptic complex fields \mathbb{C}_{\lambda} .
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
The authors declare no conflicts of interest.
There are various ways of proving Euler's formula on elliptic complex fields, and two of the more general methods are given merely here.
Proposition A.1. Suppose \varphi\in \mathbb{R} and i^2 = \lambda = -q^2, \; q\in \mathbb{R^*} , then,
\begin{equation} e^{i \varphi} = \cos (q \varphi)+i \frac{1}{q} \sin (q \varphi), \end{equation} | (A.1) |
where e is the base of the natural logarithm, and \sin\theta, \; \cos\theta are the sine and cosine functions of \theta respectively.
Proof. From the expansion of the MacLaurin series it follows that
\begin{equation*} e^{x} = 1+\frac{x}{1 !}+\frac{x^{2}}{2 !}+\ldots+\frac{x^{n}}{n !}+o\left(x^{n}\right); \end{equation*} |
\begin{equation*} \cos x = 1-\frac{x^{2}}{2 !}+\frac{x^{4}}{4 !}-\frac{x^{6}}{6 !}+\ldots+(-1)^{n} \frac{x^{2 n}}{(2 n) !}+o\left(x^{2 n+1}\right); \end{equation*} |
\begin{equation*} \sin x = \frac{x}{1 !}-\frac{x^{3}}{3 !}+\frac{x^{5}}{5 !}-\frac{x^{7}}{7 !}+\ldots+(-1)^{n} \frac{x^{2 n+1}}{(2 n+1) !}+o\left(x^{2 n+2}\right). \end{equation*} |
So there
\begin{equation*} \begin{aligned} e^{i \varphi} & = 1+\frac{i \varphi}{1 !}+\frac{(i \varphi)^{2}}{2 !}+\ldots+\frac{(i \varphi)^{n}}{n !}+o\left((i \varphi)^{n}\right) \\ & \begin{aligned} & = 1-\frac{(q \varphi)^{2}}{2 !}+\frac{(q \varphi)^{4}}{4 !}-\frac{(q \varphi)^{6}}{6 !}+\ldots+(-1)^{n} \frac{(q \varphi)^{2 n}}{(2 n) !}+o\left((q \varphi)^{2 n+1}\right) \\ &+i \frac{1}{q}\left[\frac{(q \varphi)}{1 !}-\frac{(q \varphi)^{3}}{3 !}+\frac{(q \varphi)^{5}}{5 !}-\frac{(q \varphi)^{7}}{7 !}+\ldots+(-1)^{n} \frac{(q \varphi)^{2 n+1}}{(2 n+1) !}+o\left((q \varphi)^{2 n+2}\right)\right] \\ & = \cos (q \varphi)+i \frac{1}{q} \sin (q \varphi). \end{aligned} \end{aligned} \end{equation*} |
Thus proving the proposition.
For Eq ( A.1 ), consider the case where the elliptic complex is a pure imaginary number, i.e., supposing \cos (q \varphi) = 0 at which point \sin (q \varphi) = 1 , to obtain \varphi = \frac{\arcsin1}{q} . It can be seen that \varphi is the angle between the positive y-axis and the positive x-axis in the xOy coordinate system.
An alternative approach to the proof requires further consideration of the geometric significance of elliptic complexes. For a point A(x, y) on the elliptic complex plane \mathbb{C}_\lambda , let the origin of the coordinates be O . Then, the corresponding vector \overrightarrow{OA} = x+iy , now consider the vector
\begin{equation*} \overrightarrow{OB} = i\overrightarrow{OA} = i(x+iy) = -py+ix. \end{equation*} |
So the coordinates of point B is B(-py, x) , which gives that
\begin{equation} k_{OA}\cdot k_{OB} = (-1)\times\frac{1}{p} = \frac{1}{\lambda}, \end{equation} | (A.2) |
where, with the case of y = 0 in particular, easy to get clearly that the vector \overrightarrow{OA} is on the x-axis and the vector \overrightarrow{OB} on the y-axis in the complex plane. It can be seen that when \lambda\ne-1 , the coordinate system corresponding to the elliptic complex plane is no longer a right-angle coordinate system, but a oblique coordinates system. It leads to the following definitions.
Definition A.2. Two lines l_1, l_2 in the elliptic complex plane \mathbb{C}_\lambda are orthogonal if and only if
\begin{equation} k_{l_1}\cdot k_{l_2} = \frac{1}{\lambda}, \end{equation} | (A.3) |
which is denoted l_1\bot l_2 . With the case of \lambda = -1 especially, the two lines are orthogonal while k_{l_1}\cdot k_{l_2} = -1 in the circular complex plane.
Definition A.3. One line l is said to be orthogonal to one normal ellipse \Gamma if it passes through the centre O of the ellipse \Gamma on the elliptic complex plane \mathbb{C}_\lambda .
It is easy to see that if the line l is orthogonal to the normal ellipse \Gamma , then the tangent l_c at the intersection of them satisfies k_{l}\cdot k_{l_c} = \frac{1}{\lambda} which shows that, namely, the line l is orthogonal to the tangent l_c . A detailed proof of this result would be given now.
Proof. Based on the translation invariance of one geometric figure in the complex plane, consider the special regular ellipse
\begin{equation*} \Gamma: \frac{x^{2}}{N(z)}+\frac{y^{2}}{N(z)/p} = 1, \end{equation*} |
where N(z) is the square of the principal diameter of this ellipse. Let p(x_0, y_0) be the (non-endpoint) tangent point on F , then the equation of its tangent line is
\begin{equation*} l_c: \frac{xx_0}{N(z)}+\frac{yy_0}{N(z)/p} = 1. \end{equation*} |
The slopes thus obtained are respectively
\begin{equation*} k_{l_c} = -\frac{x_{0}}{y_{0}} \cdot \frac{1}{p}, \ k_{l} = \frac{y_{0}}{x_{0}}. \end{equation*} |
This leads to the required conclusion.
Definition A.4. If two normal ellipses \Gamma_1 and \Gamma_2 , with the centres O_1 and O_2 respectively, intersect at one point P on the elliptic complex plane \mathbb{C}_\lambda , and the product of the slopes of the lines PO_1 and PO_2 is
\begin{equation*} k_{PO_1}\cdot k_{PO_2} = \frac{1}{\lambda}, \end{equation*} |
then, the normal ellipse \Gamma_1 is said to be orthogonal to the normal ellipse \Gamma_2 .
Consider now the vector
\begin{equation*} \overrightarrow{OC} = \frac{i}{q}\overrightarrow{OA} = -qy+\frac{i}{q}x. \end{equation*} |
Obviously, there would be \overrightarrow{OC}\bot \overrightarrow{OA} , along with |\overrightarrow{OC}| = |\overrightarrow{OA}| = \sqrt{x^2+q^2y^2} on the elliptic complex plane \mathbb{C}_\lambda . Thus, the vectors \vec a and \vec b on the elliptic complex plane \mathbb{C}_\lambda are mutually orthogonal and equal in length, equivalent to
\begin{equation} \vec a+\frac{i}{q}\vec b = \vec 0, \end{equation} | (A.4) |
which is the geometric meaning of the relationship between a complex number (or vector) z and another complex number (or vector) \frac{i}{q}z on the elliptic complex plane \mathbb{C}_\lambda .
Now begins the proof of another approach to Euler's formula.
Proof. Set the initial value of the function y = f(x) to
\begin{equation} \begin{cases} y' = \frac{i}{q}y; \\ y(0) = 1. \end{cases} \end{equation} | (A.5) |
We find that the first equation of Eq ( A.5 ) is a first-order ordinary differential equation in separable variables, and hence,
\begin{equation*} {\mathrm dy\over \mathrm dx} = \frac{i}{q}y \quad\Leftrightarrow \quad\int{\mathrm dy\over y} = \frac{i}{q}\int \mathrm dx \quad\Leftrightarrow \quad \ln|y| = \frac{i}{q}x+C_0. \end{equation*} |
Thereby, |y| = e^{\frac{i}{q}x}e^{C_0} , i.e., y = \pm e^{C_0}e^{\frac{i}{q}x} . Suppose C = \pm e^{C_0} , then y = Ce^{ix} which was taken into the second equation while we can easily solve for C = 1 . Consequently, according to the Picard-Lindelof theorem, y = e^{\frac{i}{q}x} is the only solution to this problem.
Now let y = \cos(x)+\frac{i}{q}\sin(x) , which yields
\begin{equation*} y' = -\sin(x)+\frac{i}{q}\cos(x) = \frac{i}{q}(\cos(x)+\frac{i}{q}\sin(x)) = \frac{i}{q}y. \end{equation*} |
So it satisfies the first equation of ( A.5 ) and obviously satisfies the second. It follows that y = e^{\frac{i}{q}x} and y = \cos(x)+\frac{i}{q}\sin(x) are both solutions to the problem and, based on the Picard-Lindelof theorem, we reach the conclusion that
\begin{equation} e^{\frac{i}{q}x} = \cos(x)+\frac{i}{q}\sin(x), \end{equation} | (A.6) |
which is equivalent to Eq ( A.1 ) apparently. In general, Eq ( A.6 ) is more commonly used.
In deed, the above result could be derived from i^2 = -q^2 , namely, \left({i\over q}\right)^2 = -1 , which could be substituted in the proof of Proposition A.1 . Similar results for circular complex domains are easily obtained with \left(1+{i\over q}\right)^2 = {2i\over q} , \left(1+{i\over q}\right)^4 = -4 and {i\over q}\left(1-{i\over q}\right) = 1+{i\over q} .
According to the above analysis about the angle of co-ordinate axes, when q = 1 , the angle is \theta = \frac{\pi}{2} between the forward y-axis and the forward x-axis, where the corresponding xOy coordinate system is the so-called right-handed right-angle coordinate system. When q = -1 , similarly, the angle is \theta = -\frac{\pi}{2} which corresponds to the left-handed right-angle coordinate system. In fact, the case q = -1 is the system of complex numbers corresponding to the root i = -\sqrt{-1} of the equation i^2 = -1 , which is one situation that the mathematicians have overlooked.
when q\ne1 and q\ne0 , there would be |\theta| = |\frac{\arcsin1}{q}| > \frac{\pi}{2} with the case of |q|\in(0, 1) and be |\theta| = |\frac{\arcsin1}{q}| < \frac{\pi}{2} with the case of |q|\in(1, \infty) . In this situation, the angle between the positive y-axis and the positive x-axis in the corresponding xOy coordinate system is no longer a right angle.
Further for k > 0 , the cases of q = k and q = -k correspond to two systems with exactly opposite properties, similar to the positive and negative particles in quantum physics.
Of particular interest is the fact that above we have merely considered the case where q is a constant, corresponding to a linear coordinate system. If q = q(t) is a function of a variable t , there would be q^{'}(t) = \frac{d}{dt}q(t) , while introducing mathematical analysis, which represents the change in slope of the y-axis in the positive direction.
It is easy to know that if q^{'}(t) is constantly greater than 0 and |q(t)|\in(1, \infty) , such a coordinate system corresponds to the geometric object described by Riemannian geometry. While q^{'}(t) is constantly less than 0 and |q(t)|\in(0, 1) , such a coordinate system corresponds to the geometric object described by Lobachevskian geometry.
Obviously, there are more possible curvilinear coordinate systems to describe geometric objects than the above two cases, such as the coordinate system corresponding to q^{'}(t) not constantly less than 0 (or not constantly greater than 0) and |q(t)|\in(s, t) , where s\in(0, 1) and t\in(1, \infty) , which would be more complex and will not be discussed here.
In fact, according to Subsection 2.1, the elliptic complex plane also has its own topology.
Definition B.1. The set of points in the interior of a normal ellipse with z_0 as centre and \delta > 0 as principal semidiameter is called the \delta -neighbourhood of z_0 in the complex plane, denoted by U(z_0, \delta) , namely,
\begin{equation} U\left(z_{0}, \delta\right) = \left\{z |\ | z-z_{0} \mid < \delta, z \in \mathbb{C}_\lambda\right\}. \end{equation} | (B.1) |
Apparently, it's an open elliptical disc.
Proof. For the necessity, let f^{'}(z) = a+ib and \Delta z = \Delta x+i\Delta y , then by the definition of the derivative, f(z+\Delta z)-f(z) = f^{\prime}(z) \Delta z+o(|\Delta z|) while \Delta z \rightarrow 0 namely,
\begin{equation*} \begin{aligned} &\quad{[u(x+\Delta x, y+\Delta y)+i v(x+\Delta x, y+\Delta y)]-[u(x, y)+i v(x, y)]} \\ & = (a+i b)(\Delta x+i \Delta y)+o(|\Delta z|) \\ & = \left(a \Delta x-q^{2} b \Delta y\right)+i(b \Delta x+a \Delta y)+o(|\Delta z|). \end{aligned} \end{equation*} |
Comparing the real and imaginary parts on both sides gives that, with \rho \rightarrow 0 ,
\begin{equation*} \begin{aligned} u(x+\Delta x, y+\Delta y)-u(x, y) = a \Delta x-q^{2} b \Delta y+o(\rho)\\ v(x+\Delta x, y+\Delta y)-v(x, y) = b \Delta x+a \Delta y+o(\rho), \end{aligned} \end{equation*} |
which leads to
\begin{equation*} a = \frac{\partial u}{\partial x}, \quad-q^{2} b = \frac{\partial u}{\partial y}, \quad b = \frac{\partial v}{\partial x}, \quad a = \frac{\partial v}{\partial y}, \end{equation*} |
that is,
\begin{equation} \frac{\partial u}{\partial x} = \frac{\partial v}{\partial y}, \quad \frac{\partial u}{\partial y} = -q^{2} \frac{\partial v}{\partial x}. \end{equation} | (B.2) |
Thus, the necessity is shown, next is the sufficiency.
Due to the differentiability of the functions u(x, y) and v(x, y) at the point (x, y) , with \rho \rightarrow 0 ,
\begin{equation*} \begin{aligned} u(x+\Delta x, y+\Delta y)-u(x, y) = \frac{\partial u}{\partial x} \Delta x+\frac{\partial u}{\partial y} \Delta y+o(\rho), \\ v(x+\Delta x, y+\Delta y)-v(x, y) = \frac{\partial v}{\partial x} \Delta x+\frac{\partial v}{\partial y} \Delta y+o(\rho). \end{aligned} \end{equation*} |
Then, the combination of the two formulas yields
\begin{equation*} \begin{aligned} & {[u(x+\Delta x, y+\Delta y)+i v(x+\Delta x, y+\Delta y)]-[u(x, y)+i v(x, y)] } \\ = &\left(\frac{\partial u}{\partial x} \Delta x+\frac{\partial u}{\partial y} \Delta y\right)+i\left(\frac{\partial v}{\partial x} \Delta x+\frac{\partial v}{\partial y} \Delta y\right)+o(|\Delta z|) \ (\text {substitute Eq (13.8) into it}) \\ = &\left(\frac{\partial u}{\partial x} \Delta x-q^{2} \frac{\partial v}{\partial x} \Delta y\right)+i\left(\frac{\partial v}{\partial x} \Delta x+\frac{\partial u}{\partial x} \Delta y\right)+o(|\Delta z|) \\ = &\left(\frac{\partial u}{\partial x}+i \frac{\partial v}{\partial x}\right)(\Delta x+i \Delta y)+o(|\Delta z|) = \left(\frac{\partial u}{\partial x}+i \frac{\partial v}{\partial x}\right) \Delta z+o(|\Delta z|). \end{aligned} \end{equation*} |
Therefore, f(z) is differentiable at z = x+iy\in D and its derivative is
\begin{equation} f^{'}(z) = \frac{\partial u}{\partial x}+i\frac{\partial v}{\partial x}. \end{equation} | (B.3) |
The sufficiency of the proposition is thus proved.
Since the in-region derivability is equivalent to the in-region resolution, Theorem 2.3 is proved naturally.
Proof. Let z = x+iy\in D and f(z) = u(x, y)+iv(x, y) . As f(z) is analytic in the region D , the functions u(x, y) and v(x, y) are differentiable in D and satisfy Eq ( B.2 ) and the following Green's formula
\begin{equation} \oint_{C} P \mathrm d x+Q \mathrm d y = \iint_{D}\left(\frac{\partial Q}{\partial x}-\frac{\partial P}{\partial y}\right) \mathrm d x \mathrm d y, \end{equation} | (B.4) |
which leads to that \begin{aligned}\int_{C} u(x, y)\mathrm d x-q^{2} v(x, y)\mathrm d y = \iint_{D}\left(-q^{2} \frac{\partial v}{\partial x}-\frac{\partial u}{\partial y}\right)\mathrm d x\mathrm d y\end{aligned} , namely,
\begin{equation} \int_{C} u(x, y)\mathrm d x-q^{2} v(x, y)\mathrm d y = \iint_{D}\left(-q^{2} \frac{\partial v}{\partial x}+q^{2} \frac{\partial v}{\partial x}\right)\mathrm d x\mathrm d y = 0, \end{equation} | (B.5) |
and that
\begin{equation} \begin{aligned}\int_{C} v(x, y)\mathrm d x+u(x, y)\mathrm d y = \iint_{D}\left(\frac{\partial u}{\partial x}-\frac{\partial v}{\partial y}\right)\mathrm d x\mathrm d y = 0.\end{aligned} \end{equation} | (B.6) |
Therefore,
\begin{equation} \begin{aligned}\int_{C}f(z)\mathrm dz = \int_{C} u(x, y)\mathrm d x-q^{2} v(x, y)\mathrm d y+i \int_{C} u(x, y)\mathrm d x+v(x, y)\mathrm d y = 0, \end{aligned} \end{equation} | (B.7) |
which means that Theorem 2.7 is shown.
So, there are the following conclusions.
Corollary B.2. Let C be a simple closed curve and the function f(z) analytic in the bounded region bounded by C , then \begin{aligned}\int_{C}f(z)\mathrm dz = 0\end{aligned} holds.
Theorem B.3. Let C_0 and C_1, C_2, \cdots, C_n enclose a multiply connected region D , D and its boundary form a closed region \overline{D} . If the function f(z) is analytic over the region \overline{D} , then, \begin{aligned}\int_{C}f(z)\mathrm dz = 0, \end{aligned} where C = C_0+C_1^-+C_2^-+\cdots+C_n^- are all the boundaries of the region D , which is also equivalent to
\begin{equation} \int_{C_0}f(z)\mathrm dz = \int_{C_1}f(z)\mathrm dz+\int_{C_2}f(z)\mathrm dz+\cdots+\int_{C_n}f(z)\mathrm dz. \end{equation} | (B.8) |
We know that \left|z-z_{0}\right| denotes the normal ellipse centered at z_0 = (x_0, y_0)^T , which can be expressed as \begin{aligned}F: \frac{(x-x_0)^{2}}{a^{2}}+\frac{(y-y_0)^{2}}{b^{2}} = 1\end{aligned} , i.e., the parametric equation F:\left\{\begin{array}{l} x-x_0 = a \cos t, \\ y-y_0 = b \sin t, \end{array} t \in[0, 2 \pi]\right. . And according to the symmetry of the ellipse, the circumference of this normal ellipse is
\begin{equation*} L = 4 \int_{0}^{\frac{\pi}{2}} \sqrt{\left(x^{\prime}(t)\right)^{2}+\left(y^{\prime}(t)\right)^{2}} \mathrm d t = 4 \int_{0}^{\frac{\pi}{2}} \sqrt{a^{2} \sin ^{2} t+b^{2} \cos ^{2} t} \mathrm d t. \end{equation*} |
For a unit normal ellipse in the complex plane \mathbb{C_\lambda} , let \lambda = -q^2 , and the principal semidiameter of the ellipse be r , then \begin{aligned}q^2 = \frac{a^2}{b^2}\end{aligned} .
When |q|\geq1 , r = a and the ellipse circumference is
\begin{equation*} \begin{aligned} L = 4 a \int_{0}^{\frac{\pi}{2}} \sqrt{\left(1-\cos ^{2} t\right)+\frac{b^{2}}{a^{2}} \cos ^{2} t} \mathrm d t = 4 r \int_{0}^{\frac{\pi}{2}} \sqrt{1-\left(1-\frac{1}{q^{2}}\right) \cos ^{2} t}\mathrm d t = :4 r E(q). \end{aligned} \end{equation*} |
And when 0 < |q|\leq1 , r = b and the ellipse circumference is
\begin{equation*} L = 4 b \int_{0}^{\frac{\pi}{2}} \sqrt{\frac{a^{2}}{b^{2}} \sin ^{2} t+\left(1-\sin ^{2} t\right)} \mathrm d t = 4 r \int_{0}^{\frac{\pi}{2}} \sqrt{1-\left(1-q^{2}\right) \sin ^{2} t} \mathrm d t = :4 r E^{\prime}(q), \end{equation*} |
where
\begin{equation} E(q) = \int_{0}^{\frac{\pi}{2}} \sqrt{1-\left(1-\frac{1}{q^{2}}\right) \cos ^{2} t} \mathrm d t ~~ \text{ and }~~ E^{\prime}(q) = \int_{0}^{\frac{\pi}{2}} \sqrt{1-\left(1-q^{2}\right) \sin ^{2}} t \mathrm d t \end{equation} | (B.9) |
are called the perimeter coefficients of the normal ellipse on the complex plane \mathbb{C_\lambda} when |q|\geq1 and when 0 < |q|\leq1 , respectively. Clearly, for the definited complex plane \mathbb{C_\lambda} , the perimetric coefficient is a constant. When \lambda = -1 , the perimeter coefficient meets \begin{aligned}E(q) = E^{\prime}(q) = \frac{\pi}{2}\end{aligned} corresponding to the circular complex plane \mathbb{C} .
Now begin the proof of Eq ( 2.10 ).
Proof. For any point z inside D , draw an elliptic circumference C_\rho:|\zeta-z| = \rho such that the closed elliptic disk enclosed by C_\rho all falls inside D .
Due to the fact that f(\zeta) is continuous at \zeta = z , it follows that \exists\delta > 0(\delta < \rho) for \forall\varepsilon > 0 such that
\begin{equation*} \begin{aligned} \left|\frac{1}{4 E(q)}\left(\oint_{C_{r}} \frac{f(\zeta)}{\zeta-z} \mathrm d \zeta-\frac{2 \pi i}{q} f(z) \right)\right|& = \left|\frac{1}{4 E(q)} \oint_{C_{r}} \frac{f(\zeta)-f(z)}{\zeta-z} \mathrm d \zeta\right| \\ & < \frac{1}{4 E(q)} \cdot \frac{\varepsilon}{r} \cdot 4 r E(q) = \varepsilon, \end{aligned} \end{equation*} |
with the case of |q| > 1 , and
\begin{equation*} \begin{aligned} \left|\frac{1}{4 E^{'}(q)}\left(\oint_{C_{r}} \frac{f(\zeta)}{\zeta-z} \mathrm d \zeta-\frac{2 \pi i}{q} f(z) \right)\right|& = \left|\frac{1}{4 E^{'}(q)} \oint_{C_{r}} \frac{f(\zeta)-f(z)}{\zeta-z} \mathrm d \zeta\right| \\ & < \frac{1}{4 E^{'}(q)} \cdot \frac{\varepsilon}{r} \cdot 4 r E^{'}(q) = \varepsilon, \end{aligned} \end{equation*} |
with the case of 0 < |q| < 1 .
In summary, as E(q) and E^{'}(q) being non-zero constants for a determined q\ne0 ,
\begin{equation*} \lim _{r \rightarrow 0} \oint_{C_{r}} \frac{f(\zeta)}{\zeta-z} \mathrm d \zeta = \frac{2 \pi i}{q} f(z) \Leftrightarrow f(z) = \frac{q}{2 \pi i} \lim _{r \rightarrow 0} \oint_{C_{r}} \frac{f(\zeta)}{\zeta-z} \mathrm d \zeta. \end{equation*} |
According to Theorem B.3 ,
\begin{equation*} \oint_{C} \frac{f(\zeta)}{\zeta-z} \mathrm d \zeta = \oint_{C_{\rho}} \frac{f(\zeta)}{\zeta-z} \mathrm d \zeta = \oint_{C_{r}} \frac{f(\zeta)}{\zeta-z} \mathrm d \zeta(0 < r < \rho). \end{equation*} |
Eventually to conclude with
\begin{equation} \begin{aligned} f(z) = \frac{q}{2 \pi i} \oint_{C} \frac{f(\zeta)}{\zeta-z}\mathrm d \zeta. \end{aligned} \end{equation} | (B.10) |
Proof. Consider first the case n = 1 where we need to prove that \begin{aligned}f^{\prime}(z) = \frac{q}{2 \pi i} \int_{C} \frac{f(\zeta)}{(\zeta-z)^{2}} \mathrm d \zeta\end{aligned} . For any z\in D , \exists d > 0 such that U(z, 2d)\subset D . Let M = \max _{\zeta \in C}\{|f(\zeta)|\} and L be the arc length of C . When 0 < |h| < d ,
\begin{equation*} \left\{\begin{array}{l} |\zeta-z| > 2 d > d, \\ |\zeta-z-h| \geq|\zeta-z|-|h| > d, \end{array}\right. \end{equation*} |
which leads to
\begin{equation*} \begin{aligned} &\quad\left|\frac{f(z+h)-f(z)}{h}-\frac{q}{2 \pi i} \oint_{C} \frac{f(\zeta)}{(\zeta-z)^{2}}\mathrm d \zeta\right| \\& = \left|\frac{1}{h}\left[\frac{q}{2 \pi i} \oint_{C} \frac{f(\zeta)}{\zeta-z-h} \mathrm d \zeta-\frac{q}{2 \pi i} \oint_{C} \frac{f(\zeta)}{\zeta-z} \mathrm d \zeta\right]-\frac{q}{2 \pi i} \oint_{C} \frac{f(\zeta)}{(\zeta-z)^{2}} \mathrm d \zeta\right| \\ & = \left|\frac{q \cdot h}{2 \pi i} \oint_{C} \frac{f(\zeta)}{(\zeta-z-h)(\zeta-z)^{2}} \mathrm d \zeta\right| \leq \frac{|q| \cdot|h|}{2 \pi} \cdot \frac{M}{d^{3}} \cdot L \rightarrow 0 \quad(h \rightarrow 0). \end{aligned} \end{equation*} |
Thus,
\begin{equation*} f^{'}(z) = \lim _{h \rightarrow 0} \frac{f(z+h)-f(z)}{h} = \frac{q}{2 \pi i} \oint_{C} \frac{f(\zeta)}{(\zeta-z)^{2}} \mathrm d \zeta. \end{equation*} |
Now, prove the general case using mathematical induction. Let the conclusion hold for n = k , namely,
\begin{equation} f^{(k)}(z) = \frac{q \cdot k !}{2 \pi i} \oint_{C} \frac{f(\zeta)}{(\zeta-z)^{k+1}} \mathrm d \zeta \quad(k = 1, 2, \ldots). \end{equation} | (B.11) |
Then, when n = k+1 ,
\begin{equation*} \begin{aligned} &\quad\left|\frac{f^{(k)}(z+h)-f^{(k)}(z)}{h}-\frac{q \cdot(k+1) !}{2 \pi i} \oint_{C} \frac{f(\zeta)}{(\zeta-z)^{k+2}} \mathrm d \zeta\right| \\ & = \left|\frac{1}{h}\left[\frac{q \cdot k !}{2 \pi i} \oint_{C} \frac{f(\zeta)}{(\zeta-z-h)^{k+1}} \mathrm d \zeta-\frac{q \cdot k !}{2 \pi i} \oint_{C} \frac{f(\zeta)}{(\zeta-z)^{k+1}} \mathrm d \zeta\right]-\frac{q \cdot(k+1) !}{2 \pi i} \oint_{C} \frac{f(\zeta)}{(\zeta-z)^{k+2}} \mathrm d \zeta\right| \\& = \left|\frac{q \cdot(k+1) !}{2 \pi i} \oint_{C} f(\zeta)\left[\frac{1}{(\zeta-z-h)^{k+1}(\zeta-z)}-\frac{1}{(\zeta-z)^{k+2}}\right] \mathrm d \zeta+h O(1)\right| \rightarrow 0 \quad(h \rightarrow 0), \end{aligned} \end{equation*} |
which results in
\begin{equation*} \begin{aligned} f^{(k+1)}(z) = \lim _{h \rightarrow 0} \frac{f^{(k)}(z+h)-f^{(k)}(z)}{h} = \frac{q \cdot(k+1) !}{2 \pi i} \oint_{C} \frac{f(\zeta)}{(\zeta-z)^{k+2}} \mathrm d \zeta. \end{aligned} \end{equation*} |
The proposition is thus proved.
Corollary B.4. Let the function f(z) be analytic in the region D , then, f(z) has derivatives with any order in D .
Any periodic function is capable of becoming the union of different sine functions, i.e.,
\begin{equation} f(x) = A_0+\sum\limits_{n = 1}^\infty A_n\sin\left({2\pi nx\over T}+\varphi\right), \end{equation} | (C.1) |
where T is the period and \varphi is the offset of the sine function, also known as the initial phase.
Use of the formula
\begin{equation*} \sin(\alpha+\beta) = \sin\alpha\cos\beta+\sin\beta\cos\alpha \end{equation*} |
could simplify C.1 to obtain
\begin{equation*} f(x) = A_0+\sum\limits_{n = 1}^\infty\left[A_n\sin\varphi\cos\left({2\pi nx\over T}\right)+A_n\cos\varphi\sin\left({2\pi nx\over T}\right)\right]. \end{equation*} |
Then, apply the substitution: \begin{cases} a_n = A_n\sin\varphi, \\ b_n = A_n\cos\varphi, \end{cases}
which results in the Fourier series
\begin{equation} f(x) = A_0+\sum\limits_{n = 1}^\infty\left[a_n\cos\left({2\pi nx\over T}\right)+b_n\sin\left({2\pi nx\over T}\right)\right]. \end{equation} | (C.2) |
By means of Euler's formula, we can solve for
\begin{equation} \cos z = \frac{1}{2}\left(e^{i \frac{z}{q}}+e^{-i \frac{z}{q}}\right), \ \sin z = \frac{q}{2 i}\left(e^{i \frac{z}{q}}-e^{-i \frac{z}{q}}\right). \end{equation} | (C.3) |
And substituting it into Eq ( C.2 ) yields
\begin{equation*} \begin{aligned} f(x)& = A_0+\sum\limits_{n = 1}^\infty\left[{a_n\over 2}\left(e^{\frac{i}{q}{2\pi nx\over T}}+e^{-{\frac{i}{q}{2\pi nx\over T}}}\right)-{ib_n\over 2q}\left(e^{{\frac{i}{q}{2\pi nx\over T}}}-e^{-{\frac{i}{q}{2\pi nx\over T}}}\right)\right] \\ & = A_0+\sum\limits_{n = 1}^\infty\left({a_n-\frac{i}{q}b_n\over 2}\cdot e^{{\frac{i}{q}{2\pi nx\over T}}}+{a_n+\frac{i}{q}b_n\over 2}\cdot e^{-{\frac{i}{q}{2\pi nx\over T}}}\right). \end{aligned} \end{equation*} |
Now, apply the substitution : c_n = \begin{cases} {a_n-\frac{i}{q}b_n\over 2}, \; n > 0; \\ A_0, \; \; \; \; \; \; n = 0; \\ {a_n+\frac{i}{q}b_n\over 2}, \; n < 0, \end{cases}
which leads to the exponential form of the Fourier series
\begin{equation} f(x) = c_0+\sum\limits_{n = 1}^\infty\left(c_ne^{{\frac{i}{q}{2\pi nx\over T}}}+c_{-n}e^{-{\frac{i}{q}{2\pi nx\over T}}}\right) = \sum\limits_{n = -\infty}^\infty c_ne^{{\frac{i}{q}{2\pi nx\over T}}}. \end{equation} | (C.4) |
Thus, \begin{aligned}f(x)e^{{-\frac{i}{q}{2\pi kx\over T}}} = \left(\sum_{n = -\infty}^\infty c_ne^{{\frac{i}{q}{2\pi nx\over T}}}\right)\cdot e^{-{\frac{i}{q}{2\pi kx\over T}}}\end{aligned} , which is then integrated over one of the periods of f(x) . And the integration and summation signs are interchangeable, according to the dominant convergence theorem, so that
\begin{equation*} \begin{aligned} \int_{x_0}^{x_0+T}f(x)e^{-\frac{i}{q}{2\pi kx\over T}}\mathrm dx & = \int_{x_0}^{x_0+T}\sum\limits_{n = -\infty}^\infty c_n e^{\frac{i}{q}{2\pi (n-k)x\over T}}\mathrm dx \\ & = \sum\limits_{n = -\infty}^\infty c_n\int_{x_0}^{x_0+T}e^{\frac{i}{q}{2\pi (n-k)x\over T}}\mathrm dx. \end{aligned} \end{equation*} |
Provided that this integral converges absolutely, i.e., \begin{aligned}\int_{x_0}^{x_0+T}\left|f(x)\right|\mathrm dx < \infty\end{aligned} , we are left with that the integral can be simplified to
\begin{equation*} c_k\int_{x_0}^{x_0+T}e^{\frac{i}{q}{2\pi (k-k)x\over T}}\mathrm dx = c_k\int_{x_0}^{x_0+T}\mathrm dx = c_kT, \end{equation*} |
with the case of n = k , and to
\begin{equation*} c_n\int_{x_0}^{x_0+T}e^{\frac{i}{q}{2\pi (n-k)x\over T}} = \left.{qTc_n\over2\pi i(n-k)}e^{\frac{i}{q}{2\pi (n-k)x\over T}}\right|_{x_0}^{x_0+T} = 0, \end{equation*} |
with the case of n\ne k . Therefore,
\begin{equation} c_k = {1\over T}\int_{x_0}^{x_0+T}f(x)e^{-\frac{i}{q}{2\pi kx\over T}}\mathrm dx, \; k\in\mathbb{Z}. \end{equation} | (C.5) |
Now, further on, study the Fourier series \begin{aligned} f(t) = \sum_{n = -\infty}^\infty c_ne^{\frac{i}{q}{2\pi nt\over T}} \end{aligned} , whose period is taken from -T/2 to T/2, where \begin{aligned} c_k = {1\over T}\int_{-{T\over2}}^{T\over2}f(t)e^{-\frac{i}{q}{2\pi kt\over T}}\mathrm dt \end{aligned} .
Suppose \begin{aligned} g(\xi) = \int_{-{T\over2}}^{T\over2}f(t)e^{-\frac{i}{q}{2\pi \xi t\over T}}\mathrm dt \end{aligned} , then, \begin{aligned}c_k = {1\over T}g\left({k\over T}\right)\end{aligned} , and it follows that
\begin{equation} f(t) = \sum\limits_{n = -\infty}^\infty g\left({n\over T}\right)e^{2\pi \frac{i}{q}\left(n\over T\right)t}\cdot{1\over T}. \end{equation} | (C.6) |
Following this, we then replace {n\over T} with \xi_n and {1\over T} with \Delta\xi to get that
\begin{equation} f(t) = \sum\limits_{n = -\infty}^{\infty}g(\xi_n)e^{2\pi \frac{i}{q}\xi_n t}\Delta\xi. \end{equation} | (C.7) |
In order for the Fourier series to be used to represent functions without periods, we can find the limit where \Delta\xi tends to 0, i.e., T\rightarrow \infty , which means
\begin{equation} f(t) = \lim\limits_{\Delta\xi\to 0^+}\sum\limits_{n = -\infty}^\infty g(\xi_n)e^{2\pi \frac{i}{q}\xi_n t}\Delta\xi. \end{equation} | (C.8) |
According to the theory of calculus, the right-hand side of the equation is found to be a Riemann sum, and the series becomes an integral
\begin{equation} f(t) = \int_{-\infty}^\infty g(\xi)e^{2\pi \frac{i}{q}\xi t}d\xi. \end{equation} | (C.9) |
Following the same approach, we arrive at
\begin{equation} g(\xi) = \int_{-\infty}^\infty f(t)e^{-2\pi \frac{i}{q}\xi t}dt. \end{equation} | (C.10) |
Hence, it is said that g(\xi) is the Fourier transform of f(t) , that is,
\begin{equation} g(\xi) = \mathcal{F}\{f(t)\} = \int_{-\infty}^\infty f(t)e^{-2\pi \frac{i}{q}\xi t}dt. \end{equation} | (C.11) |
Sometimes, for simplicity, \mathcal{F}\{f(t)\} is denoted directly by \hat{f}(\xi) . And f(t) is the inverse Fourier transform of g(\xi) , namely,
\begin{equation} f(t) = \mathcal{F}^{-1}\{g(\xi)\} = \int_{-\infty}^\infty g(\xi)e^{2\pi \frac{i}{q}\xi t}d\xi, \end{equation} | (C.12) |
which, along with Eq ( C.11 ), is called another form of the Fourier transform.
A Fourier expansion of f(t) with period T yields
\begin{equation} f(t+nT) = \sum\limits_{k = -\infty}^\infty c_ke^{2\pi \frac{i}{q}k\frac{(t+nT)}{T}} = \sum\limits_{k = -\infty}^\infty c_ke^{2\pi \frac{i}{q}kt/T}, \end{equation} | (C.13) |
where
\begin{equation} c_k = {1\over T}\int_{t+nT-{1\over2}T}^{t+nT+{1\over2}T} f(x)e^{-2\pi \frac{i}{q}kx/T}\mathrm{d}x. \end{equation} | (C.14) |
Let f(t) satisfy \begin{aligned}\int_{-\infty}^\infty|f(x)|^2\mathrm{d}x < \infty\end{aligned} , then we can sum over f(t+nT) to attain
\begin{equation*} \begin{aligned} \sum\limits_{n = -\infty}^\infty f(t+nT) & = \sum\limits_{n = -\infty}^\infty\sum\limits_{k = -\infty}^\infty c_ke^{2\pi \frac{i}{q}kt/T} \\ & = {1\over T}\sum\limits_{n = -\infty}^\infty\sum\limits_{k = -\infty}^\infty e^{2\pi \frac{i}{q}kt/T}\int_{t+nT-{1\over2}T}^{t+nT+{1\over2}T} f(x)e^{-2\pi \frac{i}{q}kx/T}\mathrm{d}x \\& = {1\over T}\sum\limits_{k = -\infty}^\infty e^{2\pi \frac{i}{q}kt/T} \underbrace{\sum\limits_{n = -\infty}^\infty\int_{t+nT-{1\over2}T}^{t+nT+{1\over2}T}f(x)e^{-2\pi \frac{i}{q}kx/T}\mathrm{d}x}_{S}, \end{aligned} \end{equation*} |
where the part S of it can be further simplified. Due to the fact that for all n\in\mathbb{Z} , the intervals \begin{aligned}\left[t+nT-\frac{T}{2}\, \ t+nT+\frac{T}{2}\right]\end{aligned} will eventually merge without overlap into the integration interval from -\infty to +\infty ,
\begin{equation} S = \int_{-\infty}^\infty f(x)e^{-2\pi \frac{i}{q}kx/T}\mathrm{d}x = \int_{-\infty}^\infty f(x)e^{-2\pi \frac{i}{q}(k/T)x}\mathrm{d}x = \hat{f}\left(k\over T\right), \end{equation} | (C.15) |
which could be substituted back into the above summation formula to yields
\begin{equation} \begin{aligned} \quad\quad{1\over T}\sum\limits_{k = -\infty}^\infty e^{2\pi \frac{i}{q}kt/T}\int_{-\infty}^\infty f(x)e^{-2\pi \frac{i}{q}kx/T}\mathrm{d}x = {1\over T}\sum\limits_{k = -\infty}^\infty\hat{f}\left(k\over T\right)e^{2\pi \frac{i}{q}kt/T} \end{aligned}, \end{equation} | (C.16) |
namely,
\begin{equation} \sum\limits_{n = -\infty}^\infty f(t+nT) = \sum\limits_{k = -\infty}^\infty{1\over T}\hat{f}\left(k\over T\right)e^{2\pi \frac{i}{q}kt/T}, \end{equation} | (C.17) |
which is called Poisson's summation formula on elliptic complex fields. In special, given that T = 1 , it follows that
\begin{equation} \sum\limits_{n\in\mathbb{Z}}f(t+n) = \sum\limits_{k\in\mathbb{Z}}\hat{f}(k)e^{\frac{i}{q}(2\pi kt)}. \end{equation} | (C.18) |
If let T = 1 and t = 0 , then
\begin{equation} \sum\limits_{n\in\mathbb{Z}}f(n) = \sum\limits_{k\in\mathbb{Z}}\hat{f}(k). \end{equation} | (C.19) |
Many summation problems can be simplified by Poisson's summation formula, especially for exponential functions which can be transformed into problems of integration.
[1] |
X. K. Zhang, Congruences mudulo 2^3 for class numbers of general quadratic fields Q(\sqrt{m}) and Q(\sqrt{m}) , J. Number Theory, 32 (1989), 332–338. https://doi.org/10.1016/0022-314X(89)90089-9 doi: 10.1016/0022-314X(89)90089-9
![]() |
[2] | N. A. Lebedev, Geometric theory of functions of a complex variable, Washington: American Mathematical Society, 1969. |
[3] | T. S. Fiske, Theory of functions of a complex variable, Science, 14 (1901), 59–65. |
[4] | M. Morse, Topological methods in the theory of functions of a complex variable, Princeton: Princeton University Press, 1949. |
[5] | C. Pan, C. Pan, Basics analytic number theory, Harbin: Harbin University of Technology Press, 2016. |
[6] |
V. Chandee, K. Soundararajan, Bounding |\xi({1\over 2}+it)| on the Riemann hypothesis, B. Lond. Math. Soc., 43 (2011), 243–250. https://doi.org/10.1112/blms/bdq095 doi: 10.1112/blms/bdq095
![]() |
[7] |
J. B. Conrey, A. Ghosh, D. Goldston, N. S. M. Gonek, D. R. Heath-Brown, On the distribution of gaps between zeros of the zeta-function, Quart. J. Math., 36 (1985), 43–51. https://doi.org/10.1093/qmath/36.1.43 doi: 10.1093/qmath/36.1.43
![]() |
[8] |
P. Bourgade, J. Kuan, Strong Szeg asymptotics and zeros of the Zeta-function, Commun. Pur. Appl. Math., 67 (2014), 1028–1044. https://doi.org/10.1002/cpa.21475 doi: 10.1002/cpa.21475
![]() |
[9] | M. R. Murty, J. Esmonde, Problems in algebraic number theory, New York: Springer, 2005. https://doi.org/10.1007/b138452 |
[10] | A. Selberg, Old and new conjectures and results about a class of Dirichlet series, Amalfi Conference on Analytic Number Theory, Italy: Maiori, 1992,367–385. |
[11] | K. Broughan, Equivalents of the Riemann hypothesis, Cambridge: Cambridge University Press, 2017. https://doi.org/10.1017/9781108178228 |
[12] | H. Davenport, Multiplicative number theory, 2 Eds., New York: Springer, 1980. https://doi.org/10.1007/978-1-4757-5927-3 |
[13] |
A. M. Odlyzko, On the distribution of spacings between zeros of the zeta function, Math. Comput., 48 (1987), 273–308. https://doi.org/10.2307/2007890 doi: 10.2307/2007890
![]() |
1. | Retraction notice to "The Generalized Riemann Hypothesis on elliptic complex fields" [AIMS Mathematics 8(11) (2023) 25772–25803], 2023, 8, 2473-6988, 27857, 10.3934/math.20231425 |