Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
\pdfrender

TextRenderingMode=2,LineWidth=.5pt \pdfrenderTextRenderingMode=2,LineWidth=.5pt

a generalized ramanujan master theorem and integral representation of meromorphic functions

Zachary P. Bradshaw  and Omprakash Atale E-mail: zbradshaw@tulane.eduE-mail: atale.om@outlook.com Naval Surface Warfare Center, Panama City Division
Panama City, FL, United States
Department of Mathematics, Savitribai Phule Pune University,
Pune-425001, India
Abstract

Ramanujan’s Master Theorem is a decades-old theorem in the theory of Mellin transforms which has wide applications in both mathematics and high energy physics. The unconventional method of Ramanujan in his proof of the theorem left convergence issues which were later settled by Hardy. Here we extend Ramanujan’s theorem to meromorphic functions with poles of arbitrary order and observe that the new theorem produces analogues of Ramanujan’s famous theorem. Moreover, we find that the theorem produces integral representations for meromorphic functions which are shown to satisfy interesting properties, opening up an avenue for further study.

§I. Introduction

It is no secret that the famous Indian mathematician Srinivasa Ramanujan was known for his mathematical insight. However, without formal training in the subject, his work was largely non-rigorous, leaving much for future generations of mathematicians to complete. Here we will expand upon just one piece of Ramanujan’s work, which he readily used to explore definite integrals. It is called Ramanujan’s master theorem, the name of which can be attributed to Berndt [4], and it states that if f(x)𝑓𝑥f(x)italic_f ( italic_x ) has an expansion of the form

f(x)=n=0(1)ng(n)n!xn𝑓𝑥superscriptsubscript𝑛0superscript1𝑛𝑔𝑛𝑛superscript𝑥𝑛f(x)=\sum_{n=0}^{\infty}(-1)^{n}\frac{g(n)}{n!}x^{n}italic_f ( italic_x ) = ∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT ( - 1 ) start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT divide start_ARG italic_g ( italic_n ) end_ARG start_ARG italic_n ! end_ARG italic_x start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT (1.1)

where g(n)𝑔𝑛g(n)italic_g ( italic_n ) has a natural and continuous extension such that g(0)0𝑔00g(0)\neq 0italic_g ( 0 ) ≠ 0, then for s>0𝑠0s>0italic_s > 0, we have

0xs1(n=0(1)ng(n)n!xn)𝑑x=g(s)Γ(s).superscriptsubscript0superscript𝑥𝑠1superscriptsubscript𝑛0superscript1𝑛𝑔𝑛𝑛superscript𝑥𝑛differential-d𝑥𝑔𝑠Γ𝑠\int_{0}^{\infty}x^{s-1}\left(\sum_{n=0}^{\infty}(-1)^{n}\frac{g(n)}{n!}x^{n}% \right)dx=g(-s)\Gamma(s).∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT italic_s - 1 end_POSTSUPERSCRIPT ( ∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT ( - 1 ) start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT divide start_ARG italic_g ( italic_n ) end_ARG start_ARG italic_n ! end_ARG italic_x start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ) italic_d italic_x = italic_g ( - italic_s ) roman_Γ ( italic_s ) . (1.2)

Eqn. (1.2) can be applied to calculate the Mellin transform of many functions in the literature. Ramanujan first communicated this result to Hardy in his 1913 quarterly reports, but his method for proving it was incomplete. It started with Euler’s integral representation of the gamma function and relied on Taylor’s expansion without regard for convergence issues. This problem was remedied by Hardy when he made natural assumptions on the analyticity of g(n)𝑔𝑛g(n)italic_g ( italic_n ) as well as its growth rate [26] which facilitated a proof of the theorem using Cauchy’s residue theory and the Mellin inversion theorem (see Theorem 1).

At present, Hardy’s class of functions has been extended in many ways [9, 11], and several generalizations of Ramanujan’s theorem exist [8, 14], including a multidimensional version known as the method of brackets [6, 7, 19, 20, 21, 22, 23], which was discussed in [28] and more recently in [2]. This method has found use in the field of high energy particle physics and quantum field theory [24, 30] for its application to precision calculations involving Feynman diagrams. Indeed, the method of brackets is particularly well equipped to handle the Schwinger parametrization [30] of loop integrals.

The purpose of this article is to point out that Hardy’s proof can be generalized in a natural way to produce a collection of Ramanujan-like master theorems. We will see that the crux of Hardy’s proof is a residue calculation followed by the Mellin inversion theorem [10], the former of which we will alter. In doing so, we find that the proof still holds under a modification of the growth conditions outlined by Hardy. As a result, we obtain many formulas for the Mellin transform of a function in terms of an analytic continuation of its coefficients, and these results therefore have applications to sequence interpolation.

The Mellin transform, being scale-invariant, is widely used in the analysis of computer algorithms [18]. Moreover, in quantum theory, the Fourier transform is used to transform from position space to momentum space, where computations may be easier, and the Mellin transform plays an analogous role in the AdS/CFT correspondence [15, 16]. Additional applications include the asymptotic approximations of functions which are defined by integrals [5, 31] and series [17, 27], the study of distributions of products and quotients of independent random variables [13], and the expression of solutions to electromagnetic wave propagation in turbulence [29]. The results presented here are therefore of interest to a wide range of disciplines.

The remainder of this work is outlined as follows. In §II, we review Ramanujan’s master theorem and the proof given by Hardy, after which we discuss the interpretation of the theorem as a sequence interpolation formula. We then modify Hardy’s proof in §III, producing a family of Ramanujan-like master theorems, which we may interpret as interpolation formulas. In §IV, we study independently some properties of the integral representation of h(z)𝑧h(z)italic_h ( italic_z ). In particular, we show that it is logarithmically convex. We extend further in §V to meromorphic functions with poles of arbitrary order, producing several exotic integral representations for meromorphic functions. Finally, in §VI, we give concluding remarks.

§II. Ramanujan’s Master Theorem

We begin with Ramanujan’s master theorem and its proof given by Hardy [26]. Ultimately, the theorem gives an expression for the Mellin transform of a certain class of functions; though, it can also be viewed as an interpolation formula. Indeed, given a sequence cnsubscript𝑐𝑛c_{n}italic_c start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT, we define a function g::𝑔g:\mathbb{N}\to\mathbb{C}italic_g : blackboard_N → blackboard_C by g(n)=cn𝑔𝑛subscript𝑐𝑛g(n)=c_{n}italic_g ( italic_n ) = italic_c start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT and ask how this function might be extended to a larger subset of the complex plane. It turns out that Ramanujan’s master theorem gives an extension of this function in a unique way, and we will return to this point in a moment. For now, we turn to the theorem.

Theorem 1 (Hardy-Ramanujan).

Let g(z)𝑔𝑧g(z)italic_g ( italic_z ) be analytic on the half-plane H(δ)={z|Re(z)δ}𝐻𝛿conditional-set𝑧Re𝑧𝛿H(\delta)=\{z\in\mathbb{C}|\textnormal{Re}(z)\geq-\delta\}italic_H ( italic_δ ) = { italic_z ∈ blackboard_C | Re ( italic_z ) ≥ - italic_δ } for some 0<δ<10𝛿10<\delta<10 < italic_δ < 1. Assume A<π𝐴𝜋A<\piitalic_A < italic_π and g𝑔gitalic_g satisfies the growth condition

|g(v+iw)|<CePv+A|w|𝑔𝑣𝑖𝑤𝐶superscript𝑒𝑃𝑣𝐴𝑤\displaystyle|g(v+iw)|<Ce^{Pv+A|w|}| italic_g ( italic_v + italic_i italic_w ) | < italic_C italic_e start_POSTSUPERSCRIPT italic_P italic_v + italic_A | italic_w | end_POSTSUPERSCRIPT (2.1)

for all z=v+iwH(δ)𝑧𝑣𝑖𝑤𝐻𝛿z=v+iw\in H(\delta)italic_z = italic_v + italic_i italic_w ∈ italic_H ( italic_δ ). Then

0xs1k=0(1)kg(k)xkdx=πsin(πs)g(s)superscriptsubscript0superscript𝑥𝑠1superscriptsubscript𝑘0superscript1𝑘𝑔𝑘superscript𝑥𝑘𝑑𝑥𝜋𝜋𝑠𝑔𝑠\displaystyle\int_{0}^{\infty}x^{s-1}\sum_{k=0}^{\infty}(-1)^{k}g(k)x^{k}dx=% \frac{\pi}{\sin(\pi s)}g(-s)∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT italic_s - 1 end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT ( - 1 ) start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_g ( italic_k ) italic_x start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_d italic_x = divide start_ARG italic_π end_ARG start_ARG roman_sin ( italic_π italic_s ) end_ARG italic_g ( - italic_s ) (2.2)

holds for all 0<Re(s)<δ0Re𝑠𝛿0<\textnormal{Re}(s)<\delta0 < Re ( italic_s ) < italic_δ.

Proof.

If |x|<eP𝑥superscript𝑒𝑃|x|<e^{-P}| italic_x | < italic_e start_POSTSUPERSCRIPT - italic_P end_POSTSUPERSCRIPT, then the series converges by the root test. Let 0<c<δ0𝑐𝛿0<c<\delta0 < italic_c < italic_δ and consider the contour integral

12πiCπsin(πz)g(z)xz𝑑z12𝜋𝑖subscript𝐶𝜋𝜋𝑧𝑔𝑧superscript𝑥𝑧differential-d𝑧\frac{1}{2\pi i}\int_{C}\frac{\pi}{\sin(\pi z)}g(-z)x^{-z}dzdivide start_ARG 1 end_ARG start_ARG 2 italic_π italic_i end_ARG ∫ start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT divide start_ARG italic_π end_ARG start_ARG roman_sin ( italic_π italic_z ) end_ARG italic_g ( - italic_z ) italic_x start_POSTSUPERSCRIPT - italic_z end_POSTSUPERSCRIPT italic_d italic_z (2.3)

where C𝐶Citalic_C is the left semicircle contour of radius n𝑛n\in\mathbb{N}italic_n ∈ blackboard_N centered at z=c𝑧𝑐z=citalic_z = italic_c. The integral Iγsubscript𝐼𝛾I_{\gamma}italic_I start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT along the arc γ𝛾\gammaitalic_γ of the semicircle satisfies

|Iγ|subscript𝐼𝛾\displaystyle\lvert I_{\gamma}\rvert| italic_I start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT | =|γπsin(πz)g(z)xz𝑑z|absentsubscript𝛾𝜋𝜋𝑧𝑔𝑧superscript𝑥𝑧differential-d𝑧\displaystyle=\bigg{|}\int_{\gamma}\frac{\pi}{\sin(\pi z)}g(-z)x^{-z}\ dz\bigg% {|}= | ∫ start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT divide start_ARG italic_π end_ARG start_ARG roman_sin ( italic_π italic_z ) end_ARG italic_g ( - italic_z ) italic_x start_POSTSUPERSCRIPT - italic_z end_POSTSUPERSCRIPT italic_d italic_z |
=|π/23π/2πg(cneiθ)xcneiθsin(π(c+neiθ))nieiθ𝑑θ|absentsuperscriptsubscript𝜋23𝜋2𝜋𝑔𝑐𝑛superscript𝑒𝑖𝜃superscript𝑥𝑐𝑛superscript𝑒𝑖𝜃𝜋𝑐𝑛superscript𝑒𝑖𝜃𝑛𝑖superscript𝑒𝑖𝜃differential-d𝜃\displaystyle=\bigg{|}\int_{\pi/2}^{3\pi/2}\frac{\pi g(-c-ne^{i\theta})x^{-c-% ne^{i\theta}}}{\sin(\pi(c+ne^{i\theta}))}nie^{i\theta}d\theta\bigg{|}= | ∫ start_POSTSUBSCRIPT italic_π / 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 italic_π / 2 end_POSTSUPERSCRIPT divide start_ARG italic_π italic_g ( - italic_c - italic_n italic_e start_POSTSUPERSCRIPT italic_i italic_θ end_POSTSUPERSCRIPT ) italic_x start_POSTSUPERSCRIPT - italic_c - italic_n italic_e start_POSTSUPERSCRIPT italic_i italic_θ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT end_ARG start_ARG roman_sin ( italic_π ( italic_c + italic_n italic_e start_POSTSUPERSCRIPT italic_i italic_θ end_POSTSUPERSCRIPT ) ) end_ARG italic_n italic_i italic_e start_POSTSUPERSCRIPT italic_i italic_θ end_POSTSUPERSCRIPT italic_d italic_θ |
π/23π/2CπneAn|sin(θ)|dθsin2(πc+πncos(θ))+sinh2(πnsin(θ)).absentsuperscriptsubscript𝜋23𝜋2𝐶𝜋𝑛superscript𝑒𝐴𝑛𝜃𝑑𝜃superscript2𝜋𝑐𝜋𝑛𝜃superscript2𝜋𝑛𝜃\displaystyle\leq\int_{\pi/2}^{3\pi/2}\frac{C\pi ne^{An|\sin(\theta)|}d\theta}% {\sqrt{\sin^{2}(\pi c+\pi n\cos(\theta))+\sinh^{2}(\pi n\sin(\theta))}}.≤ ∫ start_POSTSUBSCRIPT italic_π / 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 italic_π / 2 end_POSTSUPERSCRIPT divide start_ARG italic_C italic_π italic_n italic_e start_POSTSUPERSCRIPT italic_A italic_n | roman_sin ( italic_θ ) | end_POSTSUPERSCRIPT italic_d italic_θ end_ARG start_ARG square-root start_ARG roman_sin start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_π italic_c + italic_π italic_n roman_cos ( italic_θ ) ) + roman_sinh start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_π italic_n roman_sin ( italic_θ ) ) end_ARG end_ARG . (2.4)

Now taking the limit, as n𝑛n\to\inftyitalic_n → ∞, we see that the arc term has zero contribution to the integral. The integrand has simple poles at n𝑛-n- italic_n for all n𝑛n\in\mathbb{N}italic_n ∈ blackboard_N and the residues are

limznπ(z+n)g(z)xzsin(πz)subscript𝑧𝑛𝜋𝑧𝑛𝑔𝑧superscript𝑥𝑧𝜋𝑧\displaystyle\lim_{z\to-n}\frac{\pi(z+n)g(-z)x^{-z}}{\sin(\pi z)}roman_lim start_POSTSUBSCRIPT italic_z → - italic_n end_POSTSUBSCRIPT divide start_ARG italic_π ( italic_z + italic_n ) italic_g ( - italic_z ) italic_x start_POSTSUPERSCRIPT - italic_z end_POSTSUPERSCRIPT end_ARG start_ARG roman_sin ( italic_π italic_z ) end_ARG =πlimzng(z)limzn(z+n)xzsin(πz)absent𝜋subscript𝑧𝑛𝑔𝑧subscript𝑧𝑛𝑧𝑛superscript𝑥𝑧𝜋𝑧\displaystyle=\pi\lim_{z\to-n}g(-z)\lim_{z\to-n}\frac{(z+n)}{x^{z}\sin(\pi z)}= italic_π roman_lim start_POSTSUBSCRIPT italic_z → - italic_n end_POSTSUBSCRIPT italic_g ( - italic_z ) roman_lim start_POSTSUBSCRIPT italic_z → - italic_n end_POSTSUBSCRIPT divide start_ARG ( italic_z + italic_n ) end_ARG start_ARG italic_x start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT roman_sin ( italic_π italic_z ) end_ARG
=πg(n)limzn1ln(x)xzsin(πz)+πxzcos(πz)absent𝜋𝑔𝑛subscript𝑧𝑛1𝑥superscript𝑥𝑧𝜋𝑧𝜋superscript𝑥𝑧𝜋𝑧\displaystyle=\pi g(n)\lim_{z\to-n}\frac{1}{\ln(x)x^{z}\sin(\pi z)+\pi x^{z}% \cos(\pi z)}= italic_π italic_g ( italic_n ) roman_lim start_POSTSUBSCRIPT italic_z → - italic_n end_POSTSUBSCRIPT divide start_ARG 1 end_ARG start_ARG roman_ln ( italic_x ) italic_x start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT roman_sin ( italic_π italic_z ) + italic_π italic_x start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT roman_cos ( italic_π italic_z ) end_ARG
=πg(n)1π(1)nxnabsent𝜋𝑔𝑛1𝜋superscript1𝑛superscript𝑥𝑛\displaystyle=\pi g(n)\frac{1}{\pi(-1)^{n}x^{-n}}= italic_π italic_g ( italic_n ) divide start_ARG 1 end_ARG start_ARG italic_π ( - 1 ) start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT - italic_n end_POSTSUPERSCRIPT end_ARG
=(1)ng(n)xn.absentsuperscript1𝑛𝑔𝑛superscript𝑥𝑛\displaystyle=(-1)^{n}g(n)x^{n}.= ( - 1 ) start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT italic_g ( italic_n ) italic_x start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT . (2.5)

Then by the residue theorem, we have

12πi12𝜋𝑖\displaystyle\frac{1}{2\pi i}divide start_ARG 1 end_ARG start_ARG 2 italic_π italic_i end_ARG cic+iπg(s)xssin(πs)𝑑s=k=0(1)kg(k)xk.superscriptsubscript𝑐𝑖𝑐𝑖𝜋𝑔𝑠superscript𝑥𝑠𝜋𝑠differential-d𝑠superscriptsubscript𝑘0superscript1𝑘𝑔𝑘superscript𝑥𝑘\displaystyle\int_{c-i\infty}^{c+i\infty}\frac{\pi g(-s)x^{-s}}{\sin(\pi s)}ds% =\sum_{k=0}^{\infty}(-1)^{k}g(k)x^{k}.∫ start_POSTSUBSCRIPT italic_c - italic_i ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_c + italic_i ∞ end_POSTSUPERSCRIPT divide start_ARG italic_π italic_g ( - italic_s ) italic_x start_POSTSUPERSCRIPT - italic_s end_POSTSUPERSCRIPT end_ARG start_ARG roman_sin ( italic_π italic_s ) end_ARG italic_d italic_s = ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT ( - 1 ) start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_g ( italic_k ) italic_x start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT . (2.6)

The function πsin(πz)g(z)𝜋𝜋𝑧𝑔𝑧\frac{\pi}{\sin(\pi z)}g(-z)divide start_ARG italic_π end_ARG start_ARG roman_sin ( italic_π italic_z ) end_ARG italic_g ( - italic_z ) is analytic on 0<Re(z)<δ0Re𝑧𝛿0<\textnormal{Re}(z)<\delta0 < Re ( italic_z ) < italic_δ and the integral converges absolutely and uniformly for c(a,b)𝑐𝑎𝑏c\in(a,b)italic_c ∈ ( italic_a , italic_b ) and 0<a<b<δ0𝑎𝑏𝛿0<a<b<\delta0 < italic_a < italic_b < italic_δ. Then by the Mellin inversion theorem,

πsin(πs)g(s)=0xs1k=0(1)kg(k)xkdx𝜋𝜋𝑠𝑔𝑠superscriptsubscript0superscript𝑥𝑠1superscriptsubscript𝑘0superscript1𝑘𝑔𝑘superscript𝑥𝑘𝑑𝑥\displaystyle\frac{\pi}{\sin(\pi s)}g(-s)=\int_{0}^{\infty}x^{s-1}\sum_{k=0}^{% \infty}(-1)^{k}g(k)x^{k}dxdivide start_ARG italic_π end_ARG start_ARG roman_sin ( italic_π italic_s ) end_ARG italic_g ( - italic_s ) = ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT italic_s - 1 end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT ( - 1 ) start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_g ( italic_k ) italic_x start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_d italic_x (2.7)

for all 0<Re(s)<δ0Re𝑠𝛿0<\textnormal{Re}(s)<\delta0 < Re ( italic_s ) < italic_δ. ∎

It is now clear how Ramanujan’s master theorem can be viewed as an interpolation formula. Given g(n)=cn𝑔𝑛subscript𝑐𝑛g(n)=c_{n}italic_g ( italic_n ) = italic_c start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT as before, we define g(z)𝑔𝑧g(z)italic_g ( italic_z ) in the region δ<Re(z)<0𝛿Re𝑧0-\delta<\textnormal{Re}(z)<0- italic_δ < Re ( italic_z ) < 0 by (2.2). We can do even better by analytically continuing the domain of validity of the theorem to the interval between any two consecutive negative integers (see Theorem 8.1 in [2]) by shifting the contour in the proof of the master theorem. In shifting the vertical leg of the contour to the left, we drop some of the singularities of the reciprocal sine function from the contour, resulting in the interpolation formula

0xs1k=N(1)kg(k)xkdx=πsin(πs)g(s)superscriptsubscript0superscript𝑥𝑠1superscriptsubscript𝑘𝑁superscript1𝑘𝑔𝑘superscript𝑥𝑘𝑑𝑥𝜋𝜋𝑠𝑔𝑠\displaystyle\int_{0}^{\infty}x^{s-1}\sum_{k=N}^{\infty}(-1)^{k}g(k)x^{k}\ dx=% \frac{\pi}{\sin(\pi s)}g(-s)∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT italic_s - 1 end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_k = italic_N end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT ( - 1 ) start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_g ( italic_k ) italic_x start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_d italic_x = divide start_ARG italic_π end_ARG start_ARG roman_sin ( italic_π italic_s ) end_ARG italic_g ( - italic_s ) (2.8)

in the region N<Re(s)<N+1𝑁Re𝑠𝑁1-N<\textnormal{Re}(s)<-N+1- italic_N < Re ( italic_s ) < - italic_N + 1.

Note that both the assumption of analyticity on H(δ)𝐻𝛿H(\delta)italic_H ( italic_δ ) and the growth condition (2.1) are unnecessary to achieve (2.8). We may weaken these assumptions to hold only in the region we integrate, which shrinks with increasing N𝑁Nitalic_N. Similarly, if g(n)𝑔𝑛g(n)italic_g ( italic_n ) is defined for n<0𝑛subscriptabsent0n\in\mathbb{Z}_{<0}italic_n ∈ blackboard_Z start_POSTSUBSCRIPT < 0 end_POSTSUBSCRIPT, by shifting the vertical leg of the contour to the right and strengthening Hardy’s assumptions in the opposite way, we recover the interpolation formula

0xs1k=N(1)kg(k)xkdx=πsin(πs)g(s)superscriptsubscript0superscript𝑥𝑠1superscriptsubscript𝑘𝑁superscript1𝑘𝑔𝑘superscript𝑥𝑘𝑑𝑥𝜋𝜋𝑠𝑔𝑠\displaystyle\int_{0}^{\infty}x^{s-1}\sum_{k=-N}^{\infty}(-1)^{k}g(k)x^{k}\ dx% =\frac{\pi}{\sin(\pi s)}g(-s)∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT italic_s - 1 end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_k = - italic_N end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT ( - 1 ) start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_g ( italic_k ) italic_x start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_d italic_x = divide start_ARG italic_π end_ARG start_ARG roman_sin ( italic_π italic_s ) end_ARG italic_g ( - italic_s ) (2.9)

in the region N<Re(s)<N+1𝑁Re𝑠𝑁1N<\textnormal{Re}(s)<N+1italic_N < Re ( italic_s ) < italic_N + 1.

Observe that the main idea of Hardy’s proof is to apply the residue theorem to compute the integral (2.6) and then apply the Mellin inversion theorem. It happens that for the choice of π/sin(πs)𝜋𝜋𝑠\pi/\sin(\pi s)italic_π / roman_sin ( italic_π italic_s ) in the integrand, the residue computation is such that the result of (2.6) is a power series in x𝑥xitalic_x, giving Ramanujan’s master theorem the particularly nice form (2.2). However, nothing is stopping us from replacing this function with another singular function at the non-positive integers.

With this in mind, let us replace π/sin(πs)𝜋𝜋𝑠\pi/\sin(\pi s)italic_π / roman_sin ( italic_π italic_s ) by a meromorphic function h(s)𝑠h(s)italic_h ( italic_s ) with poles at the non-positive integers. The result of the residue calculation cannot be computed explicitly without additional information about hhitalic_h, but it is clear from Hardy’s proof that after applying the Mellin inversion theorem, the power series in the integrand will now have the form

k=0Resz=k(h(z)g(x)xz)superscriptsubscript𝑘0subscriptRes𝑧𝑘𝑧𝑔𝑥superscript𝑥𝑧\displaystyle\sum_{k=0}^{\infty}\mathop{\operatorname{Res}}_{z=-k}\left(\frac{% h(z)g(-x)}{x^{z}}\right)∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT roman_Res start_POSTSUBSCRIPT italic_z = - italic_k end_POSTSUBSCRIPT ( divide start_ARG italic_h ( italic_z ) italic_g ( - italic_x ) end_ARG start_ARG italic_x start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT end_ARG ) (2.10)

so that the result of the theorem will be of the form

0xs1superscriptsubscript0superscript𝑥𝑠1\displaystyle\int_{0}^{\infty}x^{s-1}∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT italic_s - 1 end_POSTSUPERSCRIPT k=0Resz=k(h(z)g(x)xz)dx=h(s)g(s).superscriptsubscript𝑘0subscriptRes𝑧𝑘𝑧𝑔𝑥superscript𝑥𝑧𝑑𝑥𝑠𝑔𝑠\displaystyle\sum_{k=0}^{\infty}\mathop{\operatorname{Res}}_{z=-k}\left(\frac{% h(z)g(-x)}{x^{z}}\right)dx=h(s)g(-s).∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT roman_Res start_POSTSUBSCRIPT italic_z = - italic_k end_POSTSUBSCRIPT ( divide start_ARG italic_h ( italic_z ) italic_g ( - italic_x ) end_ARG start_ARG italic_x start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT end_ARG ) italic_d italic_x = italic_h ( italic_s ) italic_g ( - italic_s ) . (2.11)

However, we have glossed over the fine details. For one, the region of convergence of the series will in general change with hhitalic_h. More importantly, the growth condition (2.1) is no longer enough to guarantee that the contribution from the arc term Iγsubscript𝐼𝛾I_{\gamma}italic_I start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT of the contour integral falls off at infinity. Therefore, we will have to be careful to tweak the conditions on g𝑔gitalic_g according to the chosen hhitalic_h when constructing analogs of Ramanujan’s theorem.

§III. Simple poles

Under the assumption that the poles of h(z)𝑧h(z)italic_h ( italic_z ) are simple, the formula (2.11) simplifies quite nicely, and it is not too hard to adjust Hardy’s proof to this more general setting. Indeed, we have the following theorem for meromorphic functions with simple poles at the non-positive integers. Let ϕ(z)italic-ϕ𝑧\phi(z)italic_ϕ ( italic_z ) denote the analytic continuation of sin(πz)h(z)𝜋𝑧𝑧\sin(\pi z)h(-z)roman_sin ( italic_π italic_z ) italic_h ( - italic_z ) to the non-negative integers; there are removable singularities here which are otherwise problematic when trying to evaluate this function.

Theorem 2 (Main Theorem).

Let g(z)𝑔𝑧g(z)italic_g ( italic_z ) be analytic on the half-plane H(δ)={z|Re(z)δ}𝐻𝛿conditional-set𝑧Re𝑧𝛿H(\delta)=\{z\in\mathbb{C}|\textnormal{Re}(z)\geq-\delta\}italic_H ( italic_δ ) = { italic_z ∈ blackboard_C | Re ( italic_z ) ≥ - italic_δ } for some 0<δ<10𝛿10<\delta<10 < italic_δ < 1, and let h(z)𝑧h(z)italic_h ( italic_z ) be meromorphic with simple poles at the non-positive integers. Assume A<π𝐴𝜋A<\piitalic_A < italic_π and g𝑔gitalic_g satisfies the growth condition

|g(v+iw)|<CePv+A|w||ϕ(v+iw)|𝑔𝑣𝑖𝑤𝐶superscript𝑒𝑃𝑣𝐴𝑤italic-ϕ𝑣𝑖𝑤\displaystyle\lvert g(v+iw)\rvert<\frac{Ce^{Pv+A|w|}}{\lvert\phi(v+iw)\rvert}| italic_g ( italic_v + italic_i italic_w ) | < divide start_ARG italic_C italic_e start_POSTSUPERSCRIPT italic_P italic_v + italic_A | italic_w | end_POSTSUPERSCRIPT end_ARG start_ARG | italic_ϕ ( italic_v + italic_i italic_w ) | end_ARG (3.1)

for all z=v+iwH(δ)𝑧𝑣𝑖𝑤𝐻𝛿z=v+iw\in H(\delta)italic_z = italic_v + italic_i italic_w ∈ italic_H ( italic_δ ) and that

limk|Resk(h(z))ϕ(k)|1/k=L1.subscript𝑘superscriptsubscriptRes𝑘𝑧italic-ϕ𝑘1𝑘𝐿1\displaystyle\lim_{k\to\infty}\left|\frac{\textnormal{Res}_{-k}(h(z))}{\phi(k)% }\right|^{1/k}=L\geq 1.roman_lim start_POSTSUBSCRIPT italic_k → ∞ end_POSTSUBSCRIPT | divide start_ARG Res start_POSTSUBSCRIPT - italic_k end_POSTSUBSCRIPT ( italic_h ( italic_z ) ) end_ARG start_ARG italic_ϕ ( italic_k ) end_ARG | start_POSTSUPERSCRIPT 1 / italic_k end_POSTSUPERSCRIPT = italic_L ≥ 1 . (3.2)

Then

0xs1superscriptsubscript0superscript𝑥𝑠1\displaystyle\int_{0}^{\infty}x^{s-1}∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT italic_s - 1 end_POSTSUPERSCRIPT k=0Resz=k(h(z))g(k)xkdx=h(s)g(s).superscriptsubscript𝑘0subscriptRes𝑧𝑘𝑧𝑔𝑘superscript𝑥𝑘𝑑𝑥𝑠𝑔𝑠\displaystyle\sum_{k=0}^{\infty}\mathop{\operatorname{Res}}_{z=-k}\left(h(z)% \right)g(k)x^{k}dx=h(s)g(-s).∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT roman_Res start_POSTSUBSCRIPT italic_z = - italic_k end_POSTSUBSCRIPT ( italic_h ( italic_z ) ) italic_g ( italic_k ) italic_x start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_d italic_x = italic_h ( italic_s ) italic_g ( - italic_s ) . (3.3)

holds for all 0<Re(s)<δ0Re𝑠𝛿0<\textnormal{Re}(s)<\delta0 < Re ( italic_s ) < italic_δ.

Proof.

Observe that

|Resz=k(h(z))g(k)xk|1/k=|Resz=k(h(z))|1/k|g(k)|1/k|x||Resz=k(h(z))ϕ(k)|1/kC1/keP|x|,superscriptsubscriptRes𝑧𝑘𝑧𝑔𝑘superscript𝑥𝑘1𝑘superscriptsubscriptRes𝑧𝑘𝑧1𝑘superscript𝑔𝑘1𝑘𝑥superscriptsubscriptRes𝑧𝑘𝑧italic-ϕ𝑘1𝑘superscript𝐶1𝑘superscript𝑒𝑃𝑥\displaystyle\left|\mathop{\operatorname{Res}}_{z=-k}(h(z))g(k)x^{k}\right|^{1% /k}=\left|\mathop{\operatorname{Res}}_{z=-k}(h(z))\right|^{1/k}|g(k)|^{1/k}|x|% \leq\left|\dfrac{\mathop{\operatorname{Res}}\limits_{z=-k}(h(z))}{\phi(k)}% \right|^{1/k}C^{1/k}e^{P}|x|,| roman_Res start_POSTSUBSCRIPT italic_z = - italic_k end_POSTSUBSCRIPT ( italic_h ( italic_z ) ) italic_g ( italic_k ) italic_x start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT | start_POSTSUPERSCRIPT 1 / italic_k end_POSTSUPERSCRIPT = | roman_Res start_POSTSUBSCRIPT italic_z = - italic_k end_POSTSUBSCRIPT ( italic_h ( italic_z ) ) | start_POSTSUPERSCRIPT 1 / italic_k end_POSTSUPERSCRIPT | italic_g ( italic_k ) | start_POSTSUPERSCRIPT 1 / italic_k end_POSTSUPERSCRIPT | italic_x | ≤ | divide start_ARG roman_Res start_POSTSUBSCRIPT italic_z = - italic_k end_POSTSUBSCRIPT ( italic_h ( italic_z ) ) end_ARG start_ARG italic_ϕ ( italic_k ) end_ARG | start_POSTSUPERSCRIPT 1 / italic_k end_POSTSUPERSCRIPT italic_C start_POSTSUPERSCRIPT 1 / italic_k end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT italic_P end_POSTSUPERSCRIPT | italic_x | , (3.4)

and taking the limit as k𝑘k\to\inftyitalic_k → ∞ yields LeP|x|𝐿superscript𝑒𝑃𝑥Le^{P}|x|italic_L italic_e start_POSTSUPERSCRIPT italic_P end_POSTSUPERSCRIPT | italic_x |. Thus, if |x|<ePL𝑥superscript𝑒𝑃𝐿|x|<\frac{e^{-P}}{L}| italic_x | < divide start_ARG italic_e start_POSTSUPERSCRIPT - italic_P end_POSTSUPERSCRIPT end_ARG start_ARG italic_L end_ARG, then the series converges by the root test. Let 0<c<δ0𝑐𝛿0<c<\delta0 < italic_c < italic_δ and consider the contour integral

12πiCh(z)g(z)xz𝑑z12𝜋𝑖subscript𝐶𝑧𝑔𝑧superscript𝑥𝑧differential-d𝑧\frac{1}{2\pi i}\int_{C}h(z)g(-z)x^{-z}dzdivide start_ARG 1 end_ARG start_ARG 2 italic_π italic_i end_ARG ∫ start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT italic_h ( italic_z ) italic_g ( - italic_z ) italic_x start_POSTSUPERSCRIPT - italic_z end_POSTSUPERSCRIPT italic_d italic_z (3.5)

where C𝐶Citalic_C is the left semicircle contour of radius n𝑛n\in\mathbb{N}italic_n ∈ blackboard_N centered at z=c𝑧𝑐z=citalic_z = italic_c. Now consider the integral along the arc γ𝛾\gammaitalic_γ of the semicircle. We have

|γh(z)g(z)xz𝑑z|subscript𝛾𝑧𝑔𝑧superscript𝑥𝑧differential-d𝑧\displaystyle\bigg{|}\int_{\gamma}h(z)g(-z)x^{-z}\ dz\bigg{|}| ∫ start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT italic_h ( italic_z ) italic_g ( - italic_z ) italic_x start_POSTSUPERSCRIPT - italic_z end_POSTSUPERSCRIPT italic_d italic_z | =|π/23π/2h(c+neiθ)g(cneiθ)xcneiθnieiθ𝑑θ|absentsuperscriptsubscript𝜋23𝜋2𝑐𝑛superscript𝑒𝑖𝜃𝑔𝑐𝑛superscript𝑒𝑖𝜃superscript𝑥𝑐𝑛superscript𝑒𝑖𝜃𝑛𝑖superscript𝑒𝑖𝜃differential-d𝜃\displaystyle=\bigg{|}\int_{\pi/2}^{3\pi/2}h(c+ne^{i\theta})g(-c-ne^{i\theta})% x^{-c-ne^{i\theta}}\ nie^{i\theta}d\theta\bigg{|}= | ∫ start_POSTSUBSCRIPT italic_π / 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 italic_π / 2 end_POSTSUPERSCRIPT italic_h ( italic_c + italic_n italic_e start_POSTSUPERSCRIPT italic_i italic_θ end_POSTSUPERSCRIPT ) italic_g ( - italic_c - italic_n italic_e start_POSTSUPERSCRIPT italic_i italic_θ end_POSTSUPERSCRIPT ) italic_x start_POSTSUPERSCRIPT - italic_c - italic_n italic_e start_POSTSUPERSCRIPT italic_i italic_θ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT italic_n italic_i italic_e start_POSTSUPERSCRIPT italic_i italic_θ end_POSTSUPERSCRIPT italic_d italic_θ |
<π/23π/2CneAn|sin(θ)|Lc+ncos(θ)dθsin2(k(θ))+sinh2(πnsin(θ)),absentsuperscriptsubscript𝜋23𝜋2𝐶𝑛superscript𝑒𝐴𝑛𝜃superscript𝐿𝑐𝑛𝜃𝑑𝜃superscript2𝑘𝜃superscript2𝜋𝑛𝜃\displaystyle<\int_{\pi/2}^{3\pi/2}\frac{Cne^{An|\sin(\theta)|}L^{c+n\cos(% \theta)}d\theta}{\sqrt{\sin^{2}(k(\theta))+\sinh^{2}(\pi n\sin(\theta))}},< ∫ start_POSTSUBSCRIPT italic_π / 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 italic_π / 2 end_POSTSUPERSCRIPT divide start_ARG italic_C italic_n italic_e start_POSTSUPERSCRIPT italic_A italic_n | roman_sin ( italic_θ ) | end_POSTSUPERSCRIPT italic_L start_POSTSUPERSCRIPT italic_c + italic_n roman_cos ( italic_θ ) end_POSTSUPERSCRIPT italic_d italic_θ end_ARG start_ARG square-root start_ARG roman_sin start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_k ( italic_θ ) ) + roman_sinh start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_π italic_n roman_sin ( italic_θ ) ) end_ARG end_ARG , (3.6)

where k(θ)=πc+πncos(θ)𝑘𝜃𝜋𝑐𝜋𝑛𝜃k(\theta)=\pi c+\pi n\cos(\theta)italic_k ( italic_θ ) = italic_π italic_c + italic_π italic_n roman_cos ( italic_θ ), and we have used the growth condition (3.1) and the earlier derived bound |x|<eP/L𝑥superscript𝑒𝑃𝐿|x|<e^{-P}/L| italic_x | < italic_e start_POSTSUPERSCRIPT - italic_P end_POSTSUPERSCRIPT / italic_L. Using the fact that cos(θ)𝜃\cos(\theta)roman_cos ( italic_θ ) is negative on the interval (π/2,3π/2)𝜋23𝜋2(\pi/2,3\pi/2)( italic_π / 2 , 3 italic_π / 2 ) and L1𝐿1L\geq 1italic_L ≥ 1 by assumption, and taking the limit as n𝑛n\to\inftyitalic_n → ∞, we see that the arc term has zero contribution to the integral. Then by the residue theorem, we have

12πi12𝜋𝑖\displaystyle\frac{1}{2\pi i}divide start_ARG 1 end_ARG start_ARG 2 italic_π italic_i end_ARG cic+ixsg(s)h(s)𝑑s=k=0Resz=k(h(z))g(k)xk.superscriptsubscript𝑐𝑖𝑐𝑖superscript𝑥𝑠𝑔𝑠𝑠differential-d𝑠superscriptsubscript𝑘0subscriptRes𝑧𝑘𝑧𝑔𝑘superscript𝑥𝑘\displaystyle\int_{c-i\infty}^{c+i\infty}x^{-s}g(-s)h(s)ds=\sum_{k=0}^{\infty}% \mathop{\operatorname{Res}}_{z=-k}(h(z))g(k)x^{k}.∫ start_POSTSUBSCRIPT italic_c - italic_i ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_c + italic_i ∞ end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT - italic_s end_POSTSUPERSCRIPT italic_g ( - italic_s ) italic_h ( italic_s ) italic_d italic_s = ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT roman_Res start_POSTSUBSCRIPT italic_z = - italic_k end_POSTSUBSCRIPT ( italic_h ( italic_z ) ) italic_g ( italic_k ) italic_x start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT . (3.7)

Now, xsg(s)h(s)superscript𝑥𝑠𝑔𝑠𝑠x^{-s}g(-s)h(s)italic_x start_POSTSUPERSCRIPT - italic_s end_POSTSUPERSCRIPT italic_g ( - italic_s ) italic_h ( italic_s ) is analytic on 0<Re(s)<δ0Re𝑠𝛿0<\mathrm{Re}(s)<\delta0 < roman_Re ( italic_s ) < italic_δ and the integral converges absolutely and uniformly for c(a,b)𝑐𝑎𝑏c\in(a,b)italic_c ∈ ( italic_a , italic_b ) and 0<a<b<δ0𝑎𝑏𝛿0<a<b<\delta0 < italic_a < italic_b < italic_δ. Then by the Mellin inversion theorem,

0xs1k=0Resz=k(h(z))g(k)xkdx=h(s)g(s)superscriptsubscript0superscript𝑥𝑠1superscriptsubscript𝑘0subscriptRes𝑧𝑘𝑧𝑔𝑘superscript𝑥𝑘𝑑𝑥𝑠𝑔𝑠\displaystyle\int_{0}^{\infty}x^{s-1}\sum_{k=0}^{\infty}\mathop{\operatorname{% Res}}_{z=-k}(h(z))g(k)x^{k}dx=h(s)g(-s)∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT italic_s - 1 end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT roman_Res start_POSTSUBSCRIPT italic_z = - italic_k end_POSTSUBSCRIPT ( italic_h ( italic_z ) ) italic_g ( italic_k ) italic_x start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_d italic_x = italic_h ( italic_s ) italic_g ( - italic_s ) (3.8)

for all 0<Re(s)<δ0Re𝑠𝛿0<\mathrm{Re}(s)<\delta0 < roman_Re ( italic_s ) < italic_δ. ∎

Note that setting g=1𝑔1g=1italic_g = 1 produces an integral representation for meromorphic functions with simple poles so long as they satisfy the remaining hypotheses of the theorem:

h(s)=0xs1k=0Resz=k(h(z))xkdx.𝑠superscriptsubscript0superscript𝑥𝑠1superscriptsubscript𝑘0subscriptRes𝑧𝑘𝑧superscript𝑥𝑘𝑑𝑥\displaystyle h(s)=\int_{0}^{\infty}x^{s-1}\sum_{k=0}^{\infty}\mathop{% \operatorname{Res}}_{z=-k}(h(z))x^{k}\ dx.italic_h ( italic_s ) = ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT italic_s - 1 end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT roman_Res start_POSTSUBSCRIPT italic_z = - italic_k end_POSTSUBSCRIPT ( italic_h ( italic_z ) ) italic_x start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_d italic_x . (3.9)

In particular, we recover Bernoulli’s integral representation of the Gamma function by setting h(s)=Γ(s)𝑠Γ𝑠h(s)=\Gamma(s)italic_h ( italic_s ) = roman_Γ ( italic_s ). Indeed, Resz=k(Γ(z))=(1)k/k!subscriptRes𝑧𝑘Γ𝑧superscript1𝑘𝑘\mathop{\operatorname{Res}}\limits_{z=-k}(\Gamma(z))=(-1)^{k}/k!roman_Res start_POSTSUBSCRIPT italic_z = - italic_k end_POSTSUBSCRIPT ( roman_Γ ( italic_z ) ) = ( - 1 ) start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT / italic_k !, so that

Γ(s)Γ𝑠\displaystyle\Gamma(s)roman_Γ ( italic_s ) =0xs1k=0(1)kk!xkdx=0xs1ex𝑑x.absentsuperscriptsubscript0superscript𝑥𝑠1superscriptsubscript𝑘0superscript1𝑘𝑘superscript𝑥𝑘𝑑𝑥superscriptsubscript0superscript𝑥𝑠1superscript𝑒𝑥differential-d𝑥\displaystyle=\int_{0}^{\infty}x^{s-1}\sum_{k=0}^{\infty}\frac{(-1)^{k}}{k!}x^% {k}\ dx=\int_{0}^{\infty}x^{s-1}e^{-x}\ dx.= ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT italic_s - 1 end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT divide start_ARG ( - 1 ) start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT end_ARG start_ARG italic_k ! end_ARG italic_x start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_d italic_x = ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT italic_s - 1 end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - italic_x end_POSTSUPERSCRIPT italic_d italic_x . (3.10)

Interestingly, if Theorem 2 were only known to be true for the case g=1𝑔1g=1italic_g = 1, the case of more general g𝑔gitalic_g can also be derived using a heuristic method involving Taylor’s theorem, which Ramanujan used in his quarterly reports. Starting with h(s)𝑠h(s)italic_h ( italic_s ) as defined by (3.9), replace x𝑥xitalic_x with ynxsuperscript𝑦𝑛𝑥y^{n}xitalic_y start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT italic_x so that we have

ynsh(s)=0xs1k=0Resz=k(h(z))ynkxkdx.superscript𝑦𝑛𝑠𝑠superscriptsubscript0superscript𝑥𝑠1superscriptsubscript𝑘0subscriptRes𝑧𝑘𝑧superscript𝑦𝑛𝑘superscript𝑥𝑘𝑑𝑥\displaystyle y^{-ns}h(s)=\int_{0}^{\infty}x^{s-1}\sum_{k=0}^{\infty}\mathop{% \operatorname{Res}}_{z=-k}(h(z))y^{nk}x^{k}\ dx.italic_y start_POSTSUPERSCRIPT - italic_n italic_s end_POSTSUPERSCRIPT italic_h ( italic_s ) = ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT italic_s - 1 end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT roman_Res start_POSTSUBSCRIPT italic_z = - italic_k end_POSTSUBSCRIPT ( italic_h ( italic_z ) ) italic_y start_POSTSUPERSCRIPT italic_n italic_k end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_d italic_x . (3.11)

Defining f𝑓fitalic_f by g(z)=f(yz)𝑔𝑧𝑓superscript𝑦𝑧g(z)=f(y^{z})italic_g ( italic_z ) = italic_f ( italic_y start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT ), multiplying both sides by f(n)(0)n!superscript𝑓𝑛0𝑛\frac{f^{(n)}(0)}{n!}divide start_ARG italic_f start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT ( 0 ) end_ARG start_ARG italic_n ! end_ARG, and summing over 0n<0𝑛0\leq n<\infty0 ≤ italic_n < ∞ produces

n=0f(n)(0)(ys)nn!h(s)superscriptsubscript𝑛0superscript𝑓𝑛0superscriptsuperscript𝑦𝑠𝑛𝑛𝑠\displaystyle\sum_{n=0}^{\infty}\frac{f^{(n)}(0)(y^{-s})^{n}}{n!}h(s)∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT divide start_ARG italic_f start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT ( 0 ) ( italic_y start_POSTSUPERSCRIPT - italic_s end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_ARG start_ARG italic_n ! end_ARG italic_h ( italic_s ) =n=0f(n)(0)n!0xs1k=0Resz=k(h(z))yknxkdxabsentsuperscriptsubscript𝑛0superscript𝑓𝑛0𝑛superscriptsubscript0superscript𝑥𝑠1superscriptsubscript𝑘0subscriptRes𝑧𝑘𝑧superscript𝑦𝑘𝑛superscript𝑥𝑘𝑑𝑥\displaystyle=\sum_{n=0}^{\infty}\frac{f^{(n)}(0)}{n!}\int_{0}^{\infty}x^{s-1}% \sum_{k=0}^{\infty}\mathop{\operatorname{Res}}_{z=-k}(h(z))y^{kn}x^{k}\ dx= ∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT divide start_ARG italic_f start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT ( 0 ) end_ARG start_ARG italic_n ! end_ARG ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT italic_s - 1 end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT roman_Res start_POSTSUBSCRIPT italic_z = - italic_k end_POSTSUBSCRIPT ( italic_h ( italic_z ) ) italic_y start_POSTSUPERSCRIPT italic_k italic_n end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_d italic_x
=0xs1k=0Resz=k(h(z))n=0f(n)(0)(yk)nn!xkdxabsentsuperscriptsubscript0superscript𝑥𝑠1superscriptsubscript𝑘0subscriptRes𝑧𝑘𝑧superscriptsubscript𝑛0superscript𝑓𝑛0superscriptsuperscript𝑦𝑘𝑛𝑛superscript𝑥𝑘𝑑𝑥\displaystyle=\int_{0}^{\infty}x^{s-1}\sum_{k=0}^{\infty}\mathop{\operatorname% {Res}}_{z=-k}(h(z))\sum_{n=0}^{\infty}\frac{f^{(n)}(0)(y^{k})^{n}}{n!}x^{k}\ dx= ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT italic_s - 1 end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT roman_Res start_POSTSUBSCRIPT italic_z = - italic_k end_POSTSUBSCRIPT ( italic_h ( italic_z ) ) ∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT divide start_ARG italic_f start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT ( 0 ) ( italic_y start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_ARG start_ARG italic_n ! end_ARG italic_x start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_d italic_x (3.12)

From the definition of f𝑓fitalic_f, it follows that

g(s)h(s)=0xs1k=0Resz=k(h(z))g(k)xkdx,𝑔𝑠𝑠superscriptsubscript0superscript𝑥𝑠1superscriptsubscript𝑘0subscriptRes𝑧𝑘𝑧𝑔𝑘superscript𝑥𝑘𝑑𝑥\displaystyle g(-s)h(s)=\int_{0}^{\infty}x^{s-1}\sum_{k=0}^{\infty}\mathop{% \operatorname{Res}}_{z=-k}(h(z))g(k)x^{k}dx,italic_g ( - italic_s ) italic_h ( italic_s ) = ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT italic_s - 1 end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT roman_Res start_POSTSUBSCRIPT italic_z = - italic_k end_POSTSUBSCRIPT ( italic_h ( italic_z ) ) italic_g ( italic_k ) italic_x start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_d italic_x , (3.13)

thereby re-establishing the more general Theorem 2.

Just as in the case of Ramanujan’s master theorem, Theorem 2 can be considered a sequence interpolation formula, and using Carlson’s theorem [6, 25] we can show that this extension is unique.

Proposition 1.

Let g(n)𝑔𝑛g(n)italic_g ( italic_n ) be a function defined on the non-negative integers. Let f1(z)subscript𝑓1𝑧f_{1}(z)italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_z ) and f2(z)subscript𝑓2𝑧f_{2}(z)italic_f start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_z ) be analytic extensions of g(n)𝑔𝑛g(n)italic_g ( italic_n ) satisfying (3.1). Then f1=f2subscript𝑓1subscript𝑓2f_{1}=f_{2}italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_f start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT.

Proof.

Define k(z):=f1(z)f2(z)assign𝑘𝑧subscript𝑓1𝑧subscript𝑓2𝑧k(z):=f_{1}(z)-f_{2}(z)italic_k ( italic_z ) := italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_z ) - italic_f start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_z ). Then for all n𝑛n\in\mathbb{N}italic_n ∈ blackboard_N, we have k(n)=f1(n)f2(n)=g(n)g(n)=0𝑘𝑛subscript𝑓1𝑛subscript𝑓2𝑛𝑔𝑛𝑔𝑛0k(n)=f_{1}(n)-f_{2}(n)=g(n)-g(n)=0italic_k ( italic_n ) = italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_n ) - italic_f start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_n ) = italic_g ( italic_n ) - italic_g ( italic_n ) = 0. Furthermore, since f1subscript𝑓1f_{1}italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and f2subscript𝑓2f_{2}italic_f start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT satisfy (3.1), k𝑘kitalic_k also satisfies (3.1). Then there exists C,P,A𝐶𝑃𝐴C,P,A\in\mathbb{R}italic_C , italic_P , italic_A ∈ blackboard_R with A<π𝐴𝜋A<\piitalic_A < italic_π such that

|k(z)||ϕ(z)|𝑘𝑧italic-ϕ𝑧\displaystyle\lvert k(z)\rvert\cdot\lvert\phi(z)\rvert| italic_k ( italic_z ) | ⋅ | italic_ϕ ( italic_z ) | CePx+A|y|absent𝐶superscript𝑒𝑃𝑥𝐴𝑦\displaystyle\leq Ce^{Px+A|y|}≤ italic_C italic_e start_POSTSUPERSCRIPT italic_P italic_x + italic_A | italic_y | end_POSTSUPERSCRIPT
CeP|z|+A|z|absent𝐶superscript𝑒𝑃𝑧𝐴𝑧\displaystyle\leq Ce^{P|z|+A|z|}≤ italic_C italic_e start_POSTSUPERSCRIPT italic_P | italic_z | + italic_A | italic_z | end_POSTSUPERSCRIPT
Ce2max{P,A}|z|absent𝐶superscript𝑒2𝑃𝐴𝑧\displaystyle\leq Ce^{2\max\{P,A\}|z|}≤ italic_C italic_e start_POSTSUPERSCRIPT 2 roman_max { italic_P , italic_A } | italic_z | end_POSTSUPERSCRIPT

Thus, by Carlson’s theorem, k𝑘kitalic_k is identically zero. i.e. f1=f2subscript𝑓1subscript𝑓2f_{1}=f_{2}italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_f start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT. ∎

We now list several corollaries of Theorem 2, some of which already appear in the literature. To start, let h(z)=ψ(z)𝑧𝜓𝑧h(z)=\psi(z)italic_h ( italic_z ) = italic_ψ ( italic_z ) so that ϕ(z)=sin(πz)ψ(z)italic-ϕ𝑧𝜋𝑧𝜓𝑧\phi(z)=\sin(\pi z)\psi(-z)italic_ϕ ( italic_z ) = roman_sin ( italic_π italic_z ) italic_ψ ( - italic_z ).

Corollary 2.1.

Let g(z)𝑔𝑧g(z)italic_g ( italic_z ) be analytic on the half-plane H(δ)={z|Re(z)δ}𝐻𝛿conditional-set𝑧Re𝑧𝛿H(\delta)=\{z\in\mathbb{C}|\textnormal{Re}(z)\geq-\delta\}italic_H ( italic_δ ) = { italic_z ∈ blackboard_C | Re ( italic_z ) ≥ - italic_δ } for some 0<δ<10𝛿10<\delta<10 < italic_δ < 1. Assume A<π𝐴𝜋A<\piitalic_A < italic_π and g𝑔gitalic_g satisfies the growth condition

|g(v+iw)|<CePv+A|w||ϕ(v+iw)|𝑔𝑣𝑖𝑤𝐶superscript𝑒𝑃𝑣𝐴𝑤italic-ϕ𝑣𝑖𝑤\displaystyle\lvert g(v+iw)\rvert<\frac{Ce^{Pv+A|w|}}{\lvert\phi(v+iw)\rvert}| italic_g ( italic_v + italic_i italic_w ) | < divide start_ARG italic_C italic_e start_POSTSUPERSCRIPT italic_P italic_v + italic_A | italic_w | end_POSTSUPERSCRIPT end_ARG start_ARG | italic_ϕ ( italic_v + italic_i italic_w ) | end_ARG (3.14)

for all z=v+iwH(δ)𝑧𝑣𝑖𝑤𝐻𝛿z=v+iw\in H(\delta)italic_z = italic_v + italic_i italic_w ∈ italic_H ( italic_δ ). Then

0xs1k=0g(k)xkdx=ψ(s)g(s)superscriptsubscript0superscript𝑥𝑠1superscriptsubscript𝑘0𝑔𝑘superscript𝑥𝑘𝑑𝑥𝜓𝑠𝑔𝑠\displaystyle\int_{0}^{\infty}x^{s-1}\sum_{k=0}^{\infty}g(k)x^{k}dx=-\psi(s)g(% -s)∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT italic_s - 1 end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_g ( italic_k ) italic_x start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_d italic_x = - italic_ψ ( italic_s ) italic_g ( - italic_s ) (3.15)

holds for all 0<Re(s)<δ0Re𝑠𝛿0<\textnormal{Re}(s)<\delta0 < Re ( italic_s ) < italic_δ.

Proof.

Set h(z)=ψ(z)𝑧𝜓𝑧h(z)=\psi(z)italic_h ( italic_z ) = italic_ψ ( italic_z ) in Theorem  2, where ψ𝜓\psiitalic_ψ denotes the digamma function defined as the logarithmic derivative of the gamma function

ψ(z)𝜓𝑧\displaystyle\psi(z)italic_ψ ( italic_z ) :=ddzln(Γ(z))=Γ(z)Γ(z).assignabsent𝑑𝑑𝑧Γ𝑧superscriptΓ𝑧Γ𝑧\displaystyle:=\frac{d}{dz}\ln(\Gamma(z))=\frac{\Gamma^{\prime}(z)}{\Gamma(z)}.:= divide start_ARG italic_d end_ARG start_ARG italic_d italic_z end_ARG roman_ln ( roman_Γ ( italic_z ) ) = divide start_ARG roman_Γ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_z ) end_ARG start_ARG roman_Γ ( italic_z ) end_ARG . (3.16)

Note that the digamma function has simple poles at the non-positive integers. A simple residue computation reveals that

Resz=k(ψ(z)g(z)xz)=g(k)xk,subscriptRes𝑧𝑘𝜓𝑧𝑔𝑧superscript𝑥𝑧𝑔𝑘superscript𝑥𝑘\displaystyle\mathop{\operatorname{Res}}_{z=-k}(\psi(z)g(-z)x^{-z})=-g(k)x^{k},roman_Res start_POSTSUBSCRIPT italic_z = - italic_k end_POSTSUBSCRIPT ( italic_ψ ( italic_z ) italic_g ( - italic_z ) italic_x start_POSTSUPERSCRIPT - italic_z end_POSTSUPERSCRIPT ) = - italic_g ( italic_k ) italic_x start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT , (3.17)

so that the resulting power series appearing in the integrand of the theorem will be k=0g(k)xksuperscriptsubscript𝑘0𝑔𝑘superscript𝑥𝑘-\sum_{k=0}^{\infty}g(k)x^{k}- ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_g ( italic_k ) italic_x start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT. Moreover, a simple computation shows that this choice of function satisfies property (3.2) with L=1𝐿1L=1italic_L = 1. ∎

Corollary 2.1 appears to be more of an interesting artifact of Theorem 2 than a useful tool as there don’t appear to be many functions which satisfy its hypothesis, if any. Not even the choice g=1𝑔1g=1italic_g = 1 suffices here. Let h(z)=Γ(s)cos(πs/2)𝑧Γ𝑠𝜋𝑠2h(z)=\Gamma(s)\cos(\pi s/2)italic_h ( italic_z ) = roman_Γ ( italic_s ) roman_cos ( italic_π italic_s / 2 ). The following is a more useful identity which appears in [3], and the reader can find an operational justification for it in [8].

Corollary 2.2.

Let g(z)𝑔𝑧g(z)italic_g ( italic_z ) be analytic on the half-plane H(δ)={z|Re(z)δ}𝐻𝛿conditional-set𝑧Re𝑧𝛿H(\delta)=\{z\in\mathbb{C}|\textnormal{Re}(z)\geq-\delta\}italic_H ( italic_δ ) = { italic_z ∈ blackboard_C | Re ( italic_z ) ≥ - italic_δ } for some 0<δ<10𝛿10<\delta<10 < italic_δ < 1. Assume A<π𝐴𝜋A<\piitalic_A < italic_π and g𝑔gitalic_g satisfies the growth condition

|g(v+iw)|<CePv+A|w||ϕ(v+iw)|𝑔𝑣𝑖𝑤𝐶superscript𝑒𝑃𝑣𝐴𝑤italic-ϕ𝑣𝑖𝑤\displaystyle\lvert g(v+iw)\rvert<\frac{Ce^{Pv+A|w|}}{\lvert\phi(v+iw)\rvert}| italic_g ( italic_v + italic_i italic_w ) | < divide start_ARG italic_C italic_e start_POSTSUPERSCRIPT italic_P italic_v + italic_A | italic_w | end_POSTSUPERSCRIPT end_ARG start_ARG | italic_ϕ ( italic_v + italic_i italic_w ) | end_ARG (3.18)

for all z=v+iwH(δ)𝑧𝑣𝑖𝑤𝐻𝛿z=v+iw\in H(\delta)italic_z = italic_v + italic_i italic_w ∈ italic_H ( italic_δ ). Then

0xs1superscriptsubscript0superscript𝑥𝑠1\displaystyle\int_{0}^{\infty}x^{s-1}∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT italic_s - 1 end_POSTSUPERSCRIPT k=0(1)kΓ(2k+1)g(2k)x2kdx=Γ(s)g(s)cos(πs2)superscriptsubscript𝑘0superscript1𝑘Γ2𝑘1𝑔2𝑘superscript𝑥2𝑘𝑑𝑥Γ𝑠𝑔𝑠𝜋𝑠2\displaystyle\sum_{k=0}^{\infty}\frac{(-1)^{k}}{\Gamma(2k+1)}g(2k)x^{2k}dx=% \Gamma(s)g(-s)\cos\left(\frac{\pi s}{2}\right)∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT divide start_ARG ( - 1 ) start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT end_ARG start_ARG roman_Γ ( 2 italic_k + 1 ) end_ARG italic_g ( 2 italic_k ) italic_x start_POSTSUPERSCRIPT 2 italic_k end_POSTSUPERSCRIPT italic_d italic_x = roman_Γ ( italic_s ) italic_g ( - italic_s ) roman_cos ( divide start_ARG italic_π italic_s end_ARG start_ARG 2 end_ARG ) (3.19)

holds for all 0<Re(s)<δ0Re𝑠𝛿0<\textnormal{Re}(s)<\delta0 < Re ( italic_s ) < italic_δ.

Proof.

The function xsg(s)cos(πs2)Γ(s)superscript𝑥𝑠𝑔𝑠𝜋𝑠2Γ𝑠x^{-s}g(-s)\cos(\frac{\pi s}{2})\Gamma(s)italic_x start_POSTSUPERSCRIPT - italic_s end_POSTSUPERSCRIPT italic_g ( - italic_s ) roman_cos ( divide start_ARG italic_π italic_s end_ARG start_ARG 2 end_ARG ) roman_Γ ( italic_s ) has simple poles at the non-positive even integers (the odd integer poles are canceled by the zero of the cosine function) and the residues are given by

Resz=2ksubscriptRes𝑧2𝑘\displaystyle\mathop{\operatorname{Res}}_{z=-2k}roman_Res start_POSTSUBSCRIPT italic_z = - 2 italic_k end_POSTSUBSCRIPT (xsg(s)cos(πs2)Γ(s))=(1)kg(2k)Γ(2k+1)x2k.superscript𝑥𝑠𝑔𝑠𝜋𝑠2Γ𝑠superscript1𝑘𝑔2𝑘Γ2𝑘1superscript𝑥2𝑘\displaystyle\left(x^{-s}g(-s)\cos\left(\frac{\pi s}{2}\right)\Gamma(s)\right)% =(-1)^{k}\frac{g(2k)}{\Gamma(2k+1)}x^{2k}.( italic_x start_POSTSUPERSCRIPT - italic_s end_POSTSUPERSCRIPT italic_g ( - italic_s ) roman_cos ( divide start_ARG italic_π italic_s end_ARG start_ARG 2 end_ARG ) roman_Γ ( italic_s ) ) = ( - 1 ) start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT divide start_ARG italic_g ( 2 italic_k ) end_ARG start_ARG roman_Γ ( 2 italic_k + 1 ) end_ARG italic_x start_POSTSUPERSCRIPT 2 italic_k end_POSTSUPERSCRIPT . (3.20)

A straightforward computation shows that the condition (3.2) on L𝐿Litalic_L is satisfied in this case with L=1𝐿1L=1italic_L = 1. ∎

Notice that setting g(z)=az𝑔𝑧superscript𝑎𝑧g(z)=a^{z}italic_g ( italic_z ) = italic_a start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT gives us the Mellin transform of the scaled cosine function

0xs1cos(ax)𝑑x=asΓ(s)cos(πs2).superscriptsubscript0superscript𝑥𝑠1𝑎𝑥differential-d𝑥superscript𝑎𝑠Γ𝑠𝜋𝑠2\displaystyle\int_{0}^{\infty}x^{s-1}\cos(ax)\ dx=a^{-s}\Gamma(s)\cos\left(% \frac{\pi s}{2}\right).∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT italic_s - 1 end_POSTSUPERSCRIPT roman_cos ( italic_a italic_x ) italic_d italic_x = italic_a start_POSTSUPERSCRIPT - italic_s end_POSTSUPERSCRIPT roman_Γ ( italic_s ) roman_cos ( divide start_ARG italic_π italic_s end_ARG start_ARG 2 end_ARG ) . (3.21)

The bound on g𝑔gitalic_g in Theorem 2 can likely be improved upon. To see this, note that an equivalent version of Ramanujan’s master theorem can be derived by applying Euler’s reflection formula to π/sin(πs)𝜋𝜋𝑠\pi/\sin(\pi s)italic_π / roman_sin ( italic_π italic_s ) and redefining g𝑔gitalic_g. Indeed, by sending g(s)g(s)/Γ(1+s)𝑔𝑠𝑔𝑠Γ1𝑠g(s)\to g(s)/\Gamma(1+s)italic_g ( italic_s ) → italic_g ( italic_s ) / roman_Γ ( 1 + italic_s ), we have

0xs1k=0(1)kk!g(k)xkdx=Γ(s)g(s).superscriptsubscript0superscript𝑥𝑠1superscriptsubscript𝑘0superscript1𝑘𝑘𝑔𝑘superscript𝑥𝑘𝑑𝑥Γ𝑠𝑔𝑠\displaystyle\int_{0}^{\infty}x^{s-1}\sum_{k=0}^{\infty}\frac{(-1)^{k}}{k!}g(k% )x^{k}dx=\Gamma(s)g(-s).∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT italic_s - 1 end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT divide start_ARG ( - 1 ) start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT end_ARG start_ARG italic_k ! end_ARG italic_g ( italic_k ) italic_x start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_d italic_x = roman_Γ ( italic_s ) italic_g ( - italic_s ) . (3.22)

However, we can derive this result directly using a modification of Hardy’s proof as before by setting h(z)=Γ(z)𝑧Γ𝑧h(z)=\Gamma(z)italic_h ( italic_z ) = roman_Γ ( italic_z ).

Corollary 2.3.

Let g(z)𝑔𝑧g(z)italic_g ( italic_z ) be analytic on the half-plane H(δ)={z|Re(z)δ}𝐻𝛿conditional-set𝑧Re𝑧𝛿H(\delta)=\{z\in\mathbb{C}|\textnormal{Re}(z)\geq-\delta\}italic_H ( italic_δ ) = { italic_z ∈ blackboard_C | Re ( italic_z ) ≥ - italic_δ } for some 0<δ<10𝛿10<\delta<10 < italic_δ < 1. Assume A<π𝐴𝜋A<\piitalic_A < italic_π and g𝑔gitalic_g satisfies the growth condition

|g(v+iw)|<CePv+A|w||ϕ(v+iw)|𝑔𝑣𝑖𝑤𝐶superscript𝑒𝑃𝑣𝐴𝑤italic-ϕ𝑣𝑖𝑤\displaystyle\lvert g(v+iw)\rvert<\frac{Ce^{Pv+A|w|}}{\lvert\phi(v+iw)\rvert}| italic_g ( italic_v + italic_i italic_w ) | < divide start_ARG italic_C italic_e start_POSTSUPERSCRIPT italic_P italic_v + italic_A | italic_w | end_POSTSUPERSCRIPT end_ARG start_ARG | italic_ϕ ( italic_v + italic_i italic_w ) | end_ARG (3.23)

for all z=v+iwH(δ)𝑧𝑣𝑖𝑤𝐻𝛿z=v+iw\in H(\delta)italic_z = italic_v + italic_i italic_w ∈ italic_H ( italic_δ ). Then

0xs1k=0(1)kk!g(k)xkdx=Γ(s)g(s)superscriptsubscript0superscript𝑥𝑠1superscriptsubscript𝑘0superscript1𝑘𝑘𝑔𝑘superscript𝑥𝑘𝑑𝑥Γ𝑠𝑔𝑠\displaystyle\int_{0}^{\infty}x^{s-1}\sum_{k=0}^{\infty}\frac{(-1)^{k}}{k!}g(k% )x^{k}dx=\Gamma(s)g(-s)∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT italic_s - 1 end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT divide start_ARG ( - 1 ) start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT end_ARG start_ARG italic_k ! end_ARG italic_g ( italic_k ) italic_x start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_d italic_x = roman_Γ ( italic_s ) italic_g ( - italic_s ) (3.24)

holds for all 0<Re(s)<δ0Re𝑠𝛿0<\textnormal{Re}(s)<\delta0 < Re ( italic_s ) < italic_δ.

Proof.

Let h(z)=Γ(z)𝑧Γ𝑧h(z)=\Gamma(z)italic_h ( italic_z ) = roman_Γ ( italic_z ) and apply Theorem 2. The residue calculation produces

Resz=k(Γ(z)g(z)xz)subscriptRes𝑧𝑘Γ𝑧𝑔𝑧superscript𝑥𝑧\displaystyle\mathop{\operatorname{Res}}_{z=-k}(\Gamma(z)g(-z)x^{-z})roman_Res start_POSTSUBSCRIPT italic_z = - italic_k end_POSTSUBSCRIPT ( roman_Γ ( italic_z ) italic_g ( - italic_z ) italic_x start_POSTSUPERSCRIPT - italic_z end_POSTSUPERSCRIPT ) =g(k)xkResz=k(Γ(z))absent𝑔𝑘superscript𝑥𝑘subscriptRes𝑧𝑘Γ𝑧\displaystyle=g(k)x^{k}\mathop{\operatorname{Res}}_{z=-k}(\Gamma(z))= italic_g ( italic_k ) italic_x start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT roman_Res start_POSTSUBSCRIPT italic_z = - italic_k end_POSTSUBSCRIPT ( roman_Γ ( italic_z ) )
=(1)kk!g(k)xk,absentsuperscript1𝑘𝑘𝑔𝑘superscript𝑥𝑘\displaystyle=\frac{(-1)^{k}}{k!}g(k)x^{k},= divide start_ARG ( - 1 ) start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT end_ARG start_ARG italic_k ! end_ARG italic_g ( italic_k ) italic_x start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT , (3.25)

where in the first line we have used the analyticity of g(z)xz𝑔𝑧superscript𝑥𝑧g(-z)x^{-z}italic_g ( - italic_z ) italic_x start_POSTSUPERSCRIPT - italic_z end_POSTSUPERSCRIPT combined with the fact that the gamma function has simple poles at the non-positive integers. The theorem now follows from the Mellin inversion theorem, just as before. ∎

§IV. Integral representation of meromorphic function

The aim of this section is to study independently the integral representation of meromorphic functions with simple poles at the non-positive integers that we obtained in (3.9):

h(s)=0xs1k=0Resz=k(h(z))xkdx.𝑠superscriptsubscript0superscript𝑥𝑠1superscriptsubscript𝑘0subscriptRes𝑧𝑘𝑧superscript𝑥𝑘𝑑𝑥\displaystyle h(s)=\int_{0}^{\infty}x^{s-1}\sum_{k=0}^{\infty}\mathop{% \operatorname{Res}}_{z=-k}(h(z))x^{k}\ dx.italic_h ( italic_s ) = ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT italic_s - 1 end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT roman_Res start_POSTSUBSCRIPT italic_z = - italic_k end_POSTSUBSCRIPT ( italic_h ( italic_z ) ) italic_x start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_d italic_x . (4.1)

In particular, we will show that h(s)𝑠h(s)italic_h ( italic_s ) is logarithmically convex and the function hm(x)=h(x+m)/h(m)subscript𝑚𝑥𝑥𝑚𝑚h_{m}(x)=h(x+m)/h(m)italic_h start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ( italic_x ) = italic_h ( italic_x + italic_m ) / italic_h ( italic_m ) is supermultiplicative.

Let f,g,w:I:𝑓𝑔𝑤𝐼f,g,w:I\subseteq\mathbb{R}\to\mathbb{R}italic_f , italic_g , italic_w : italic_I ⊆ blackboard_R → blackboard_R such that w(x)0,xIformulae-sequence𝑤𝑥0for-all𝑥𝐼w(x)\geq 0,\forall x\in Iitalic_w ( italic_x ) ≥ 0 , ∀ italic_x ∈ italic_I and w,wfg,wf,wg𝑤𝑤𝑓𝑔𝑤𝑓𝑤𝑔w,wfg,wf,wgitalic_w , italic_w italic_f italic_g , italic_w italic_f , italic_w italic_g are integrable on I𝐼Iitalic_I. We say f𝑓fitalic_f and g𝑔gitalic_g are synchronous if (f(x)f(y))(g(x)g(y))0,x,yIformulae-sequence𝑓𝑥𝑓𝑦𝑔𝑥𝑔𝑦0for-all𝑥𝑦𝐼(f(x)-f(y))(g(x)-g(y))\geq 0,\forall x,y\in I( italic_f ( italic_x ) - italic_f ( italic_y ) ) ( italic_g ( italic_x ) - italic_g ( italic_y ) ) ≥ 0 , ∀ italic_x , italic_y ∈ italic_I and asynchronous if (f(x)f(y))(g(x)g(y))0,x,yIformulae-sequence𝑓𝑥𝑓𝑦𝑔𝑥𝑔𝑦0for-all𝑥𝑦𝐼(f(x)-f(y))(g(x)-g(y))\leq 0,\forall x,y\in I( italic_f ( italic_x ) - italic_f ( italic_y ) ) ( italic_g ( italic_x ) - italic_g ( italic_y ) ) ≤ 0 , ∀ italic_x , italic_y ∈ italic_I. If f𝑓fitalic_f and g𝑔gitalic_g are synchronous (asynchronous) on I𝐼Iitalic_I then [12]

Iw(x)𝑑xIw(x)f(x)g(x)()Iw(x)f(x)𝑑xIw(x)g(x)𝑑x.subscript𝐼𝑤𝑥differential-d𝑥subscript𝐼𝑤𝑥𝑓𝑥𝑔𝑥subscript𝐼𝑤𝑥𝑓𝑥differential-d𝑥subscript𝐼𝑤𝑥𝑔𝑥differential-d𝑥\int_{I}w(x)dx\int_{I}w(x)f(x)g(x)\geq(\leq)\int_{I}w(x)f(x)dx\int_{I}w(x)g(x)dx.∫ start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT italic_w ( italic_x ) italic_d italic_x ∫ start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT italic_w ( italic_x ) italic_f ( italic_x ) italic_g ( italic_x ) ≥ ( ≤ ) ∫ start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT italic_w ( italic_x ) italic_f ( italic_x ) italic_d italic_x ∫ start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT italic_w ( italic_x ) italic_g ( italic_x ) italic_d italic_x . (4.2)

For a given m0𝑚0m\geq 0italic_m ≥ 0, consider hm:[0,):subscript𝑚0h_{m}:[0,\infty)\to\mathbb{R}italic_h start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT : [ 0 , ∞ ) → blackboard_R,

hm(x)=h(x+m)h(m).subscript𝑚𝑥𝑥𝑚𝑚h_{m}(x)=\frac{h(x+m)}{h(m)}.italic_h start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ( italic_x ) = divide start_ARG italic_h ( italic_x + italic_m ) end_ARG start_ARG italic_h ( italic_m ) end_ARG . (4.3)

where h(s)𝑠h(s)italic_h ( italic_s ) is defined as in (4.1).

Theorem 3.

The mapping hm(.)h_{m}(.)italic_h start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ( . ) is supermultiplicative on [0,)0[0,\infty)[ 0 , ∞ ).

Proof.

Let f(t)=tx𝑓𝑡superscript𝑡𝑥f(t)=t^{x}italic_f ( italic_t ) = italic_t start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT and g(t)=ty𝑔𝑡superscript𝑡𝑦g(t)=t^{y}italic_g ( italic_t ) = italic_t start_POSTSUPERSCRIPT italic_y end_POSTSUPERSCRIPT for some x,y>0𝑥𝑦0x,y>0italic_x , italic_y > 0. Then both f(t)𝑓𝑡f(t)italic_f ( italic_t ) and g(t)𝑔𝑡g(t)italic_g ( italic_t ) are non decreasing on [0,)0[0,\infty)[ 0 , ∞ ) and are therefore synchronous. Suppose that

w(t):=tm1k=0Resz=k(h(z))tk0assign𝑤𝑡superscript𝑡𝑚1superscriptsubscript𝑘0subscriptRes𝑧𝑘𝑧superscript𝑡𝑘0w(t):=t^{m-1}\sum_{k=0}^{\infty}\mathop{\operatorname{Res}}_{z=-k}(h(z))t^{k}\geq 0italic_w ( italic_t ) := italic_t start_POSTSUPERSCRIPT italic_m - 1 end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT roman_Res start_POSTSUBSCRIPT italic_z = - italic_k end_POSTSUBSCRIPT ( italic_h ( italic_z ) ) italic_t start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT ≥ 0 (4.4)

on [0,)0[0,\infty)[ 0 , ∞ ). Then applying Chebyshev’s inequality (4.2) for synchronous mappings produces

0tm1superscriptsubscript0superscript𝑡𝑚1\displaystyle\int_{0}^{\infty}t^{m-1}∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_t start_POSTSUPERSCRIPT italic_m - 1 end_POSTSUPERSCRIPT k=0Resz=k(h(z))tkdt0tx+y+m1k=0Resz=k(h(z))tkdtsuperscriptsubscript𝑘0subscriptRes𝑧𝑘𝑧superscript𝑡𝑘𝑑𝑡superscriptsubscript0superscript𝑡𝑥𝑦𝑚1superscriptsubscript𝑘0subscriptRes𝑧𝑘𝑧superscript𝑡𝑘𝑑𝑡\displaystyle\sum_{k=0}^{\infty}\mathop{\operatorname{Res}}_{z=-k}(h(z))t^{k}% dt\int_{0}^{\infty}t^{x+y+m-1}\sum_{k=0}^{\infty}\mathop{\operatorname{Res}}_{% z=-k}(h(z))t^{k}dt∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT roman_Res start_POSTSUBSCRIPT italic_z = - italic_k end_POSTSUBSCRIPT ( italic_h ( italic_z ) ) italic_t start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_d italic_t ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_t start_POSTSUPERSCRIPT italic_x + italic_y + italic_m - 1 end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT roman_Res start_POSTSUBSCRIPT italic_z = - italic_k end_POSTSUBSCRIPT ( italic_h ( italic_z ) ) italic_t start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_d italic_t
0tx+m1k=0Resz=k(h(z))tkdt0ty+m1k=0Resz=k(h(z))tkdt,absentsuperscriptsubscript0superscript𝑡𝑥𝑚1superscriptsubscript𝑘0subscriptRes𝑧𝑘𝑧superscript𝑡𝑘𝑑𝑡superscriptsubscript0superscript𝑡𝑦𝑚1superscriptsubscript𝑘0subscriptRes𝑧𝑘𝑧superscript𝑡𝑘𝑑𝑡\displaystyle\geq\int_{0}^{\infty}t^{x+m-1}\sum_{k=0}^{\infty}\mathop{% \operatorname{Res}}_{z=-k}(h(z))t^{k}dt\int_{0}^{\infty}t^{y+m-1}\sum_{k=0}^{% \infty}\mathop{\operatorname{Res}}_{z=-k}(h(z))t^{k}dt,≥ ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_t start_POSTSUPERSCRIPT italic_x + italic_m - 1 end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT roman_Res start_POSTSUBSCRIPT italic_z = - italic_k end_POSTSUBSCRIPT ( italic_h ( italic_z ) ) italic_t start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_d italic_t ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_t start_POSTSUPERSCRIPT italic_y + italic_m - 1 end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT roman_Res start_POSTSUBSCRIPT italic_z = - italic_k end_POSTSUBSCRIPT ( italic_h ( italic_z ) ) italic_t start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_d italic_t , (4.5)

which is equivalent to

h(m)h(x+y+m)h(x+m)h(y+m).𝑚𝑥𝑦𝑚𝑥𝑚𝑦𝑚h(m)h(x+y+m)\geq h(x+m)h(y+m).italic_h ( italic_m ) italic_h ( italic_x + italic_y + italic_m ) ≥ italic_h ( italic_x + italic_m ) italic_h ( italic_y + italic_m ) . (4.6)

It therefore follows from the definition of hmsubscript𝑚h_{m}italic_h start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT that

hm(x+y)hm(x)hm(y),subscript𝑚𝑥𝑦subscript𝑚𝑥subscript𝑚𝑦h_{m}(x+y)\geq h_{m}(x)h_{m}(y),italic_h start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ( italic_x + italic_y ) ≥ italic_h start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ( italic_x ) italic_h start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ( italic_y ) , (4.7)

and this completes the proof. ∎

Let I𝐼I\subset\mathbb{R}italic_I ⊂ blackboard_R be an interval in \mathbb{R}blackboard_R and assume that fLp(I)𝑓subscript𝐿𝑝𝐼f\in L_{p}(I)italic_f ∈ italic_L start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ( italic_I ) and gLq(I)𝑔subscript𝐿𝑞𝐼g\in L_{q}(I)italic_g ∈ italic_L start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT ( italic_I ), i.e.,

I|f(s)|p𝑑s,I|f(s)|q𝑑s<,(q,p>1).formulae-sequencesubscript𝐼superscript𝑓𝑠𝑝differential-d𝑠subscript𝐼superscript𝑓𝑠𝑞differential-d𝑠𝑞𝑝1\int_{I}|f(s)|^{p}ds,\int_{I}|f(s)|^{q}ds<\infty,\,\,\left(q,p>1\right).∫ start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT | italic_f ( italic_s ) | start_POSTSUPERSCRIPT italic_p end_POSTSUPERSCRIPT italic_d italic_s , ∫ start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT | italic_f ( italic_s ) | start_POSTSUPERSCRIPT italic_q end_POSTSUPERSCRIPT italic_d italic_s < ∞ , ( italic_q , italic_p > 1 ) . (4.8)

If 1p+1q=11𝑝1𝑞1\frac{1}{p}+\frac{1}{q}=1divide start_ARG 1 end_ARG start_ARG italic_p end_ARG + divide start_ARG 1 end_ARG start_ARG italic_q end_ARG = 1, then fgL1(I)𝑓𝑔subscript𝐿1𝐼fg\in L_{1}(I)italic_f italic_g ∈ italic_L start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_I ) and Hölder’s inequality yields

|If(s)g(s)𝑑s|(I|f(s)|p𝑑s)1p(I|g(s)|q𝑑s)1q.subscript𝐼𝑓𝑠𝑔𝑠differential-d𝑠superscriptsubscript𝐼superscript𝑓𝑠𝑝differential-d𝑠1𝑝superscriptsubscript𝐼superscript𝑔𝑠𝑞differential-d𝑠1𝑞\left|\int_{I}f(s)g(s)ds\right|\leq\left(\int_{I}|f(s)|^{p}ds\right)^{\frac{1}% {p}}\left(\int_{I}|g(s)|^{q}ds\right)^{\frac{1}{q}}.| ∫ start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT italic_f ( italic_s ) italic_g ( italic_s ) italic_d italic_s | ≤ ( ∫ start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT | italic_f ( italic_s ) | start_POSTSUPERSCRIPT italic_p end_POSTSUPERSCRIPT italic_d italic_s ) start_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG italic_p end_ARG end_POSTSUPERSCRIPT ( ∫ start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT | italic_g ( italic_s ) | start_POSTSUPERSCRIPT italic_q end_POSTSUPERSCRIPT italic_d italic_s ) start_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG italic_q end_ARG end_POSTSUPERSCRIPT . (4.9)

In fact, for some non-negative function w(s)𝑤𝑠w(s)italic_w ( italic_s ) on I𝐼Iitalic_I, the following analogue of Hölder’s inequality holds:

|If(s)g(s)w(s)𝑑s|(I|f(s)|pw(s)𝑑s)1p(I|g(s)|qw(s)𝑑s)1qsubscript𝐼𝑓𝑠𝑔𝑠𝑤𝑠differential-d𝑠superscriptsubscript𝐼superscript𝑓𝑠𝑝𝑤𝑠differential-d𝑠1𝑝superscriptsubscript𝐼superscript𝑔𝑠𝑞𝑤𝑠differential-d𝑠1𝑞\left|\int_{I}f(s)g(s)w(s)ds\right|\leq\left(\int_{I}|f(s)|^{p}w(s)ds\right)^{% \frac{1}{p}}\left(\int_{I}|g(s)|^{q}w(s)ds\right)^{\frac{1}{q}}| ∫ start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT italic_f ( italic_s ) italic_g ( italic_s ) italic_w ( italic_s ) italic_d italic_s | ≤ ( ∫ start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT | italic_f ( italic_s ) | start_POSTSUPERSCRIPT italic_p end_POSTSUPERSCRIPT italic_w ( italic_s ) italic_d italic_s ) start_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG italic_p end_ARG end_POSTSUPERSCRIPT ( ∫ start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT | italic_g ( italic_s ) | start_POSTSUPERSCRIPT italic_q end_POSTSUPERSCRIPT italic_w ( italic_s ) italic_d italic_s ) start_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG italic_q end_ARG end_POSTSUPERSCRIPT (4.10)

provided that the integrals exist and are finite.

Theorem 4.

Let a,b>0𝑎𝑏0a,b>0italic_a , italic_b > 0 with a+b=1𝑎𝑏1a+b=1italic_a + italic_b = 1 and x,y0𝑥𝑦0x,y\geq 0italic_x , italic_y ≥ 0. Then h(s)𝑠h(s)italic_h ( italic_s ) is logarithmically convex.

Proof.

Choose f(s)=sa(x1)𝑓𝑠superscript𝑠𝑎𝑥1f(s)=s^{a(x-1)}italic_f ( italic_s ) = italic_s start_POSTSUPERSCRIPT italic_a ( italic_x - 1 ) end_POSTSUPERSCRIPT, g(s)=sb(y1)𝑔𝑠superscript𝑠𝑏𝑦1g(s)=s^{b(y-1)}italic_g ( italic_s ) = italic_s start_POSTSUPERSCRIPT italic_b ( italic_y - 1 ) end_POSTSUPERSCRIPT and

w(s):=k=0Resz=k(h(z))sk0assign𝑤𝑠superscriptsubscript𝑘0subscriptRes𝑧𝑘𝑧superscript𝑠𝑘0w(s):=\sum_{k=0}^{\infty}\mathop{\operatorname{Res}}_{z=-k}(h(z))s^{k}\geq 0italic_w ( italic_s ) := ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT roman_Res start_POSTSUBSCRIPT italic_z = - italic_k end_POSTSUBSCRIPT ( italic_h ( italic_z ) ) italic_s start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT ≥ 0 (4.11)

for s(0,)𝑠0s\in(0,\infty)italic_s ∈ ( 0 , ∞ ) in (4.10) with I=(0,)𝐼0I=(0,\infty)italic_I = ( 0 , ∞ ) and p=1/a,q=1/bformulae-sequence𝑝1𝑎𝑞1𝑏p=1/a,q=1/bitalic_p = 1 / italic_a , italic_q = 1 / italic_b. Then

0superscriptsubscript0\displaystyle\int_{0}^{\infty}∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT sa(x1)sb(y1)k=0Resz=k(h(z))skdssuperscript𝑠𝑎𝑥1superscript𝑠𝑏𝑦1superscriptsubscript𝑘0subscriptRes𝑧𝑘𝑧superscript𝑠𝑘𝑑𝑠\displaystyle s^{a(x-1)}s^{b(y-1)}\sum_{k=0}^{\infty}\mathop{\operatorname{Res% }}_{z=-k}(h(z))s^{k}dsitalic_s start_POSTSUPERSCRIPT italic_a ( italic_x - 1 ) end_POSTSUPERSCRIPT italic_s start_POSTSUPERSCRIPT italic_b ( italic_y - 1 ) end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT roman_Res start_POSTSUBSCRIPT italic_z = - italic_k end_POSTSUBSCRIPT ( italic_h ( italic_z ) ) italic_s start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_d italic_s
(0spa(x1)k=0Resz=k(h(z))skds)1p(0sqb(y1)k=0Resz=k(h(z))skds)1q,absentsuperscriptsuperscriptsubscript0superscript𝑠𝑝𝑎𝑥1superscriptsubscript𝑘0subscriptRes𝑧𝑘𝑧superscript𝑠𝑘𝑑𝑠1𝑝superscriptsuperscriptsubscript0superscript𝑠𝑞𝑏𝑦1superscriptsubscript𝑘0subscriptRes𝑧𝑘𝑧superscript𝑠𝑘𝑑𝑠1𝑞\displaystyle\leq\left(\int_{0}^{\infty}s^{pa(x-1)}\sum_{k=0}^{\infty}\mathop{% \operatorname{Res}}_{z=-k}(h(z))s^{k}ds\right)^{\frac{1}{p}}\left(\int_{0}^{% \infty}s^{qb(y-1)}\sum_{k=0}^{\infty}\mathop{\operatorname{Res}}_{z=-k}(h(z))s% ^{k}ds\right)^{\frac{1}{q}},≤ ( ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_s start_POSTSUPERSCRIPT italic_p italic_a ( italic_x - 1 ) end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT roman_Res start_POSTSUBSCRIPT italic_z = - italic_k end_POSTSUBSCRIPT ( italic_h ( italic_z ) ) italic_s start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_d italic_s ) start_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG italic_p end_ARG end_POSTSUPERSCRIPT ( ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_s start_POSTSUPERSCRIPT italic_q italic_b ( italic_y - 1 ) end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT roman_Res start_POSTSUBSCRIPT italic_z = - italic_k end_POSTSUBSCRIPT ( italic_h ( italic_z ) ) italic_s start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_d italic_s ) start_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG italic_q end_ARG end_POSTSUPERSCRIPT , (4.12)

from which it follows that

0sax+by1superscriptsubscript0superscript𝑠𝑎𝑥𝑏𝑦1\displaystyle\int_{0}^{\infty}s^{ax+by-1}∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_s start_POSTSUPERSCRIPT italic_a italic_x + italic_b italic_y - 1 end_POSTSUPERSCRIPT k=0Resz=k(h(z))skdssuperscriptsubscript𝑘0subscriptRes𝑧𝑘𝑧superscript𝑠𝑘𝑑𝑠\displaystyle\sum_{k=0}^{\infty}\mathop{\operatorname{Res}}_{z=-k}(h(z))s^{k}ds∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT roman_Res start_POSTSUBSCRIPT italic_z = - italic_k end_POSTSUBSCRIPT ( italic_h ( italic_z ) ) italic_s start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_d italic_s
(0sx1k=0Resz=k(h(z))skds)a(0sy1k=0Resz=k(h(z))skds)b.absentsuperscriptsuperscriptsubscript0superscript𝑠𝑥1superscriptsubscript𝑘0subscriptRes𝑧𝑘𝑧superscript𝑠𝑘𝑑𝑠𝑎superscriptsuperscriptsubscript0superscript𝑠𝑦1superscriptsubscript𝑘0subscriptRes𝑧𝑘𝑧superscript𝑠𝑘𝑑𝑠𝑏\displaystyle\leq\left(\int_{0}^{\infty}s^{x-1}\sum_{k=0}^{\infty}\mathop{% \operatorname{Res}}_{z=-k}(h(z))s^{k}ds\right)^{a}\left(\int_{0}^{\infty}s^{y-% 1}\sum_{k=0}^{\infty}\mathop{\operatorname{Res}}_{z=-k}(h(z))s^{k}ds\right)^{b}.≤ ( ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_s start_POSTSUPERSCRIPT italic_x - 1 end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT roman_Res start_POSTSUBSCRIPT italic_z = - italic_k end_POSTSUBSCRIPT ( italic_h ( italic_z ) ) italic_s start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_d italic_s ) start_POSTSUPERSCRIPT italic_a end_POSTSUPERSCRIPT ( ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_s start_POSTSUPERSCRIPT italic_y - 1 end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT roman_Res start_POSTSUBSCRIPT italic_z = - italic_k end_POSTSUBSCRIPT ( italic_h ( italic_z ) ) italic_s start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_d italic_s ) start_POSTSUPERSCRIPT italic_b end_POSTSUPERSCRIPT . (4.13)

Thus,

h(ax+by)[h(x)]a[h(y)]b.𝑎𝑥𝑏𝑦superscriptdelimited-[]𝑥𝑎superscriptdelimited-[]𝑦𝑏h(ax+by)\leq[h(x)]^{a}[h(y)]^{b}.italic_h ( italic_a italic_x + italic_b italic_y ) ≤ [ italic_h ( italic_x ) ] start_POSTSUPERSCRIPT italic_a end_POSTSUPERSCRIPT [ italic_h ( italic_y ) ] start_POSTSUPERSCRIPT italic_b end_POSTSUPERSCRIPT . (4.14)

This completes the proof. ∎

When h(s)=Γ(s)𝑠Γ𝑠h(s)=\Gamma(s)italic_h ( italic_s ) = roman_Γ ( italic_s ), we recover the above results for the gamma function.

§. V Higher order poles

A consequence of the introduction of meromorphic functions with higher order poles is a loss of the power series form for the function undergoing a Mellin transformation. Instead, we see that factors of log(x)𝑥\log(x)roman_log ( italic_x ) make an appearance in the summand, making the identities that follow rather exotic, and for this reason we forgo a study of the convergence conditions, giving only heuristic results in this section.

Theorem 5.

Let g(z)𝑔𝑧g(z)italic_g ( italic_z ) be analytic on the half-plane H(δ)={z|Re(z)δ}𝐻𝛿conditional-set𝑧Re𝑧𝛿H(\delta)=\{z\in\mathbb{C}|\textnormal{Re}(z)\geq-\delta\}italic_H ( italic_δ ) = { italic_z ∈ blackboard_C | Re ( italic_z ) ≥ - italic_δ } for some 0<δ<10𝛿10<\delta<10 < italic_δ < 1, and let h(z)𝑧h(z)italic_h ( italic_z ) be meromorphic with poles of order Nksubscript𝑁𝑘N_{k}italic_N start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT at the non-positive integers k=0,1,2,𝑘012k=0,-1,-2,\ldotsitalic_k = 0 , - 1 , - 2 , …. Then under suitable growth conditions,

0xs1k=0n=0Nk1(1)Nkn1Γ(Nkn)cnNk(k)[(ddz+log(x))Nkn1g(z)]z=kxkdx=h(s)g(s)superscriptsubscript0superscript𝑥𝑠1superscriptsubscript𝑘0superscriptsubscript𝑛0subscript𝑁𝑘1superscript1subscript𝑁𝑘𝑛1Γsubscript𝑁𝑘𝑛subscriptsuperscript𝑐𝑘𝑛subscript𝑁𝑘subscriptdelimited-[]superscript𝑑𝑑𝑧𝑥subscript𝑁𝑘𝑛1𝑔𝑧𝑧𝑘superscript𝑥𝑘𝑑𝑥𝑠𝑔𝑠\displaystyle\int_{0}^{\infty}x^{s-1}\sum_{k=0}^{\infty}\sum_{n=0}^{N_{k}-1}% \frac{(-1)^{N_{k}-n-1}}{\Gamma(N_{k}-n)}c^{(k)}_{n-N_{k}}\left[\left(\frac{d}{% dz}+\log(x)\right)^{N_{k}-n-1}g(z)\right]_{z=k}x^{k}dx=h(s)g(-s)∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT italic_s - 1 end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT - 1 end_POSTSUPERSCRIPT divide start_ARG ( - 1 ) start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT - italic_n - 1 end_POSTSUPERSCRIPT end_ARG start_ARG roman_Γ ( italic_N start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT - italic_n ) end_ARG italic_c start_POSTSUPERSCRIPT ( italic_k ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n - italic_N start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_POSTSUBSCRIPT [ ( divide start_ARG italic_d end_ARG start_ARG italic_d italic_z end_ARG + roman_log ( italic_x ) ) start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT - italic_n - 1 end_POSTSUPERSCRIPT italic_g ( italic_z ) ] start_POSTSUBSCRIPT italic_z = italic_k end_POSTSUBSCRIPT italic_x start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_d italic_x = italic_h ( italic_s ) italic_g ( - italic_s ) (5.1)

holds for all 0<Re(s)<δ0Re𝑠𝛿0<\textnormal{Re}(s)<\delta0 < Re ( italic_s ) < italic_δ, where c1(k),,cNk(k)subscriptsuperscript𝑐𝑘1subscriptsuperscript𝑐𝑘subscript𝑁𝑘c^{(k)}_{-1},\ldots,c^{(k)}_{-N_{k}}italic_c start_POSTSUPERSCRIPT ( italic_k ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT , … , italic_c start_POSTSUPERSCRIPT ( italic_k ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - italic_N start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_POSTSUBSCRIPT are the coefficients of the principal part of hhitalic_h around z=k𝑧𝑘z=-kitalic_z = - italic_k.

Proof.

The assumption of suitable growth conditions is simply to assure convergence where necessary and to force the arc term in the inverse Mellin transform to vanish in the limit. The remainder of the theorem is proven by examining the Laurent expansion of h(z)g(z)xz𝑧𝑔𝑧superscript𝑥𝑧h(z)g(-z)x^{-z}italic_h ( italic_z ) italic_g ( - italic_z ) italic_x start_POSTSUPERSCRIPT - italic_z end_POSTSUPERSCRIPT about the non-positive integers. By assumption, hhitalic_h admits a Laurent series expansion about z=k𝑧𝑘z=-kitalic_z = - italic_k of the form

h(z)=j=Nkcj(k)(z+k)j.𝑧superscriptsubscript𝑗subscript𝑁𝑘subscriptsuperscript𝑐𝑘𝑗superscript𝑧𝑘𝑗h(z)=\sum_{j=-N_{k}}^{\infty}c^{(k)}_{j}(z+k)^{j}.italic_h ( italic_z ) = ∑ start_POSTSUBSCRIPT italic_j = - italic_N start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_c start_POSTSUPERSCRIPT ( italic_k ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_z + italic_k ) start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT . (5.2)

Similarly, g𝑔gitalic_g admits a Taylor series

g(z)=j=0(1)jg(j)(k)j!(z+k)j,𝑔𝑧superscriptsubscript𝑗0superscript1𝑗superscript𝑔𝑗𝑘𝑗superscript𝑧𝑘𝑗g(-z)=\sum_{j=0}^{\infty}(-1)^{j}\frac{g^{(j)}(k)}{j!}(z+k)^{j},italic_g ( - italic_z ) = ∑ start_POSTSUBSCRIPT italic_j = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT ( - 1 ) start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT divide start_ARG italic_g start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT ( italic_k ) end_ARG start_ARG italic_j ! end_ARG ( italic_z + italic_k ) start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT , (5.3)

and we may expand xzsuperscript𝑥𝑧x^{-z}italic_x start_POSTSUPERSCRIPT - italic_z end_POSTSUPERSCRIPT as

xz=j=0(1)jlogj(x)j!xk(z+k)j.superscript𝑥𝑧superscriptsubscript𝑗0superscript1𝑗superscript𝑗𝑥𝑗superscript𝑥𝑘superscript𝑧𝑘𝑗x^{-z}=\sum_{j=0}^{\infty}(-1)^{j}\frac{\log^{j}(x)}{j!}x^{k}(z+k)^{j}.italic_x start_POSTSUPERSCRIPT - italic_z end_POSTSUPERSCRIPT = ∑ start_POSTSUBSCRIPT italic_j = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT ( - 1 ) start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT divide start_ARG roman_log start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT ( italic_x ) end_ARG start_ARG italic_j ! end_ARG italic_x start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT ( italic_z + italic_k ) start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT . (5.4)

Then the product of the latter two functions is given by the Cauchy product

g(z)xz𝑔𝑧superscript𝑥𝑧\displaystyle g(-z)x^{-z}italic_g ( - italic_z ) italic_x start_POSTSUPERSCRIPT - italic_z end_POSTSUPERSCRIPT =j=0i=0j(1)jg(i)(k)logji(x)i!(ji)!xk(z+k)jabsentsuperscriptsubscript𝑗0superscriptsubscript𝑖0𝑗superscript1𝑗superscript𝑔𝑖𝑘superscript𝑗𝑖𝑥𝑖𝑗𝑖superscript𝑥𝑘superscript𝑧𝑘𝑗\displaystyle=\sum_{j=0}^{\infty}\sum_{i=0}^{j}(-1)^{j}\frac{g^{(i)}(k)\log^{j% -i}(x)}{i!(j-i)!}x^{k}(z+k)^{j}= ∑ start_POSTSUBSCRIPT italic_j = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_i = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT ( - 1 ) start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT divide start_ARG italic_g start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT ( italic_k ) roman_log start_POSTSUPERSCRIPT italic_j - italic_i end_POSTSUPERSCRIPT ( italic_x ) end_ARG start_ARG italic_i ! ( italic_j - italic_i ) ! end_ARG italic_x start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT ( italic_z + italic_k ) start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT
=xkj=0(1)jj![(ddz+log(x))jg(z)]z=k(z+k)j.absentsuperscript𝑥𝑘superscriptsubscript𝑗0superscript1𝑗𝑗subscriptdelimited-[]superscript𝑑𝑑𝑧𝑥𝑗𝑔𝑧𝑧𝑘superscript𝑧𝑘𝑗\displaystyle=x^{k}\sum_{j=0}^{\infty}\frac{(-1)^{j}}{j!}\left[\left(\frac{d}{% dz}+\log(x)\right)^{j}g(z)\right]_{z=k}(z+k)^{j}.= italic_x start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_j = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT divide start_ARG ( - 1 ) start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT end_ARG start_ARG italic_j ! end_ARG [ ( divide start_ARG italic_d end_ARG start_ARG italic_d italic_z end_ARG + roman_log ( italic_x ) ) start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT italic_g ( italic_z ) ] start_POSTSUBSCRIPT italic_z = italic_k end_POSTSUBSCRIPT ( italic_z + italic_k ) start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT . (5.5)

It follows that the residue of h(z)g(z)xz𝑧𝑔𝑧superscript𝑥𝑧h(z)g(-z)x^{-z}italic_h ( italic_z ) italic_g ( - italic_z ) italic_x start_POSTSUPERSCRIPT - italic_z end_POSTSUPERSCRIPT at z=k𝑧𝑘z=-kitalic_z = - italic_k is given by

Resz=k(h(z)g(z)xz)subscriptRes𝑧𝑘𝑧𝑔𝑧superscript𝑥𝑧\displaystyle\mathop{\operatorname{Res}}_{z=-k}(h(z)g(-z)x^{-z})roman_Res start_POSTSUBSCRIPT italic_z = - italic_k end_POSTSUBSCRIPT ( italic_h ( italic_z ) italic_g ( - italic_z ) italic_x start_POSTSUPERSCRIPT - italic_z end_POSTSUPERSCRIPT ) =xkn=Nk1cn(k)(1)n1(n1)![(ddz+log(x))n1g(z)]z=kabsentsuperscript𝑥𝑘superscriptsubscript𝑛subscript𝑁𝑘1subscriptsuperscript𝑐𝑘𝑛superscript1𝑛1𝑛1subscriptdelimited-[]superscript𝑑𝑑𝑧𝑥𝑛1𝑔𝑧𝑧𝑘\displaystyle=x^{k}\sum_{n=-N_{k}}^{-1}c^{(k)}_{n}\frac{(-1)^{-n-1}}{(-n-1)!}% \left[\left(\frac{d}{dz}+\log(x)\right)^{-n-1}g(z)\right]_{z=k}= italic_x start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_n = - italic_N start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT italic_c start_POSTSUPERSCRIPT ( italic_k ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT divide start_ARG ( - 1 ) start_POSTSUPERSCRIPT - italic_n - 1 end_POSTSUPERSCRIPT end_ARG start_ARG ( - italic_n - 1 ) ! end_ARG [ ( divide start_ARG italic_d end_ARG start_ARG italic_d italic_z end_ARG + roman_log ( italic_x ) ) start_POSTSUPERSCRIPT - italic_n - 1 end_POSTSUPERSCRIPT italic_g ( italic_z ) ] start_POSTSUBSCRIPT italic_z = italic_k end_POSTSUBSCRIPT
=xkn=0Nk1cnNk(k)(1)Nkn1Γ(Nkn)[(ddz+log(x))Nkn1g(z)]z=k.absentsuperscript𝑥𝑘superscriptsubscript𝑛0subscript𝑁𝑘1subscriptsuperscript𝑐𝑘𝑛subscript𝑁𝑘superscript1subscript𝑁𝑘𝑛1Γsubscript𝑁𝑘𝑛subscriptdelimited-[]superscript𝑑𝑑𝑧𝑥subscript𝑁𝑘𝑛1𝑔𝑧𝑧𝑘\displaystyle=x^{k}\sum_{n=0}^{N_{k}-1}c^{(k)}_{n-N_{k}}\frac{(-1)^{N_{k}-n-1}% }{\Gamma(N_{k}-n)}\left[\left(\frac{d}{dz}+\log(x)\right)^{N_{k}-n-1}g(z)% \right]_{z=k}.= italic_x start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT - 1 end_POSTSUPERSCRIPT italic_c start_POSTSUPERSCRIPT ( italic_k ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n - italic_N start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_POSTSUBSCRIPT divide start_ARG ( - 1 ) start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT - italic_n - 1 end_POSTSUPERSCRIPT end_ARG start_ARG roman_Γ ( italic_N start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT - italic_n ) end_ARG [ ( divide start_ARG italic_d end_ARG start_ARG italic_d italic_z end_ARG + roman_log ( italic_x ) ) start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT - italic_n - 1 end_POSTSUPERSCRIPT italic_g ( italic_z ) ] start_POSTSUBSCRIPT italic_z = italic_k end_POSTSUBSCRIPT . (5.6)

Thus,

12πiCh(z)g(z)xz𝑑x=k=0n=0Nk1cnNk(k)(1)Nkn1Γ(Nkn)[(ddz+log(x))Nkn1g(z)]z=kxk,12𝜋𝑖subscript𝐶𝑧𝑔𝑧superscript𝑥𝑧differential-d𝑥superscriptsubscript𝑘0superscriptsubscript𝑛0subscript𝑁𝑘1subscriptsuperscript𝑐𝑘𝑛subscript𝑁𝑘superscript1subscript𝑁𝑘𝑛1Γsubscript𝑁𝑘𝑛subscriptdelimited-[]superscript𝑑𝑑𝑧𝑥subscript𝑁𝑘𝑛1𝑔𝑧𝑧𝑘superscript𝑥𝑘\frac{1}{2\pi i}\int_{C}h(z)g(-z)x^{-z}dx=\sum_{k=0}^{\infty}\sum_{n=0}^{N_{k}% -1}c^{(k)}_{n-N_{k}}\frac{(-1)^{N_{k}-n-1}}{\Gamma(N_{k}-n)}\left[\left(\frac{% d}{dz}+\log(x)\right)^{N_{k}-n-1}g(z)\right]_{z=k}x^{k},divide start_ARG 1 end_ARG start_ARG 2 italic_π italic_i end_ARG ∫ start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT italic_h ( italic_z ) italic_g ( - italic_z ) italic_x start_POSTSUPERSCRIPT - italic_z end_POSTSUPERSCRIPT italic_d italic_x = ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT - 1 end_POSTSUPERSCRIPT italic_c start_POSTSUPERSCRIPT ( italic_k ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n - italic_N start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_POSTSUBSCRIPT divide start_ARG ( - 1 ) start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT - italic_n - 1 end_POSTSUPERSCRIPT end_ARG start_ARG roman_Γ ( italic_N start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT - italic_n ) end_ARG [ ( divide start_ARG italic_d end_ARG start_ARG italic_d italic_z end_ARG + roman_log ( italic_x ) ) start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT - italic_n - 1 end_POSTSUPERSCRIPT italic_g ( italic_z ) ] start_POSTSUBSCRIPT italic_z = italic_k end_POSTSUBSCRIPT italic_x start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT , (5.7)

from which the Mellin inversion theorem produces

0xs1k=0n=0Nk1(1)Nkn1Γ(Nkn)cnNk(k)[(ddz+log(x))Nkn1g(z)]z=kxkdx=h(s)g(s).superscriptsubscript0superscript𝑥𝑠1superscriptsubscript𝑘0superscriptsubscript𝑛0subscript𝑁𝑘1superscript1subscript𝑁𝑘𝑛1Γsubscript𝑁𝑘𝑛subscriptsuperscript𝑐𝑘𝑛subscript𝑁𝑘subscriptdelimited-[]superscript𝑑𝑑𝑧𝑥subscript𝑁𝑘𝑛1𝑔𝑧𝑧𝑘superscript𝑥𝑘𝑑𝑥𝑠𝑔𝑠\displaystyle\int_{0}^{\infty}x^{s-1}\sum_{k=0}^{\infty}\sum_{n=0}^{N_{k}-1}% \frac{(-1)^{N_{k}-n-1}}{\Gamma(N_{k}-n)}c^{(k)}_{n-N_{k}}\left[\left(\frac{d}{% dz}+\log(x)\right)^{N_{k}-n-1}g(z)\right]_{z=k}x^{k}dx=h(s)g(-s).∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT italic_s - 1 end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT - 1 end_POSTSUPERSCRIPT divide start_ARG ( - 1 ) start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT - italic_n - 1 end_POSTSUPERSCRIPT end_ARG start_ARG roman_Γ ( italic_N start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT - italic_n ) end_ARG italic_c start_POSTSUPERSCRIPT ( italic_k ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n - italic_N start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_POSTSUBSCRIPT [ ( divide start_ARG italic_d end_ARG start_ARG italic_d italic_z end_ARG + roman_log ( italic_x ) ) start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT - italic_n - 1 end_POSTSUPERSCRIPT italic_g ( italic_z ) ] start_POSTSUBSCRIPT italic_z = italic_k end_POSTSUBSCRIPT italic_x start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_d italic_x = italic_h ( italic_s ) italic_g ( - italic_s ) . (5.8)

Observe that if hhitalic_h has only simple poles at the non-positive integers, then its m𝑚mitalic_m-th derivative h(m)superscript𝑚h^{(m)}italic_h start_POSTSUPERSCRIPT ( italic_m ) end_POSTSUPERSCRIPT has poles of order m+1𝑚1m+1italic_m + 1. That is, the derivative has the effect of increasing the order of poles by one. Moreover, with each derivative, no new coefficients are added to the principal part of the Laurent series expansion, and the coefficient from the simple pole (the residue) gets shifted down an index. Explicitly, if we expand hhitalic_h around z=a𝑧𝑎z=aitalic_z = italic_a, so that we have h(z)=b1(za)1+O(1)𝑧subscript𝑏1superscript𝑧𝑎1𝑂1h(z)=b_{-1}(z-a)^{-1}+O(1)italic_h ( italic_z ) = italic_b start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT ( italic_z - italic_a ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT + italic_O ( 1 ), it then follows that h(z)=b1(za)2+O(1)superscript𝑧subscript𝑏1superscript𝑧𝑎2𝑂1h^{\prime}(z)=-b_{-1}(z-a)^{-2}+O(1)italic_h start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_z ) = - italic_b start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT ( italic_z - italic_a ) start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT + italic_O ( 1 ), and more generally, h(m)(z)=(1)mm!b1(za)m1+O(1)superscript𝑚𝑧superscript1𝑚𝑚subscript𝑏1superscript𝑧𝑎𝑚1𝑂1h^{(m)}(z)=(-1)^{m}m!b_{-1}(z-a)^{-m-1}+O(1)italic_h start_POSTSUPERSCRIPT ( italic_m ) end_POSTSUPERSCRIPT ( italic_z ) = ( - 1 ) start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT italic_m ! italic_b start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT ( italic_z - italic_a ) start_POSTSUPERSCRIPT - italic_m - 1 end_POSTSUPERSCRIPT + italic_O ( 1 ). Thus if a=k𝑎𝑘a=-kitalic_a = - italic_k, the coefficient cm1(k)subscriptsuperscript𝑐𝑘𝑚1c^{(k)}_{-m-1}italic_c start_POSTSUPERSCRIPT ( italic_k ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - italic_m - 1 end_POSTSUBSCRIPT in Theorem 5 is given by (1)mm!b1superscript1𝑚𝑚subscript𝑏1(-1)^{m}m!b_{-1}( - 1 ) start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT italic_m ! italic_b start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT and all other coefficients of the principal part of the expansion around z=k𝑧𝑘z=-kitalic_z = - italic_k vanish. We therefore obtain the following corollary.

Corollary 5.1.

Let g(z)𝑔𝑧g(z)italic_g ( italic_z ) be analytic on the half-plane H(δ)={z|Re(z)δ}𝐻𝛿conditional-set𝑧Re𝑧𝛿H(\delta)=\{z\in\mathbb{C}|\textnormal{Re}(z)\geq-\delta\}italic_H ( italic_δ ) = { italic_z ∈ blackboard_C | Re ( italic_z ) ≥ - italic_δ } for some 0<δ<10𝛿10<\delta<10 < italic_δ < 1, and let h(z)𝑧h(z)italic_h ( italic_z ) be meromorphic with simple poles the non-positive integers k=0,1,2,𝑘012k=0,-1,-2,\ldotsitalic_k = 0 , - 1 , - 2 , …. Then under suitable growth conditions,

0xs1k=0Resz=k(h(z))[(ddz+log(x))mg(z)]z=kxkdx=h(m)(s)g(s)superscriptsubscript0superscript𝑥𝑠1superscriptsubscript𝑘0subscriptRes𝑧𝑘𝑧subscriptdelimited-[]superscript𝑑𝑑𝑧𝑥𝑚𝑔𝑧𝑧𝑘superscript𝑥𝑘𝑑𝑥superscript𝑚𝑠𝑔𝑠\displaystyle\int_{0}^{\infty}x^{s-1}\sum_{k=0}^{\infty}\mathop{\operatorname{% Res}}_{z=-k}(h(z))\left[\left(\frac{d}{dz}+\log(x)\right)^{m}g(z)\right]_{z=k}% x^{k}dx=h^{(m)}(s)g(-s)∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT italic_s - 1 end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT roman_Res start_POSTSUBSCRIPT italic_z = - italic_k end_POSTSUBSCRIPT ( italic_h ( italic_z ) ) [ ( divide start_ARG italic_d end_ARG start_ARG italic_d italic_z end_ARG + roman_log ( italic_x ) ) start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT italic_g ( italic_z ) ] start_POSTSUBSCRIPT italic_z = italic_k end_POSTSUBSCRIPT italic_x start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_d italic_x = italic_h start_POSTSUPERSCRIPT ( italic_m ) end_POSTSUPERSCRIPT ( italic_s ) italic_g ( - italic_s ) (5.9)

holds for all 0<Re(s)<δ0Re𝑠𝛿0<\textnormal{Re}(s)<\delta0 < Re ( italic_s ) < italic_δ.

Proof.

Since hhitalic_h is meromorphic with simple poles, we may expand it around z=k𝑧𝑘z=-kitalic_z = - italic_k so that h(z)=Resz=k(h(z))(z+k)1+O(1)𝑧subscriptRes𝑧𝑘𝑧superscript𝑧𝑘1𝑂1h(z)=\mathop{\operatorname{Res}}_{z=-k}(h(z))(z+k)^{-1}+O(1)italic_h ( italic_z ) = roman_Res start_POSTSUBSCRIPT italic_z = - italic_k end_POSTSUBSCRIPT ( italic_h ( italic_z ) ) ( italic_z + italic_k ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT + italic_O ( 1 ), and the m𝑚mitalic_m-th derivative satisfies h(m)(z)=(1)mm!Resz=k(h(z))(z+k)m1+O(1)superscript𝑚𝑧superscript1𝑚𝑚subscriptRes𝑧𝑘𝑧superscript𝑧𝑘𝑚1𝑂1h^{(m)}(z)=(-1)^{m}m!\mathop{\operatorname{Res}}_{z=-k}(h(z))(z+k)^{-m-1}+O(1)italic_h start_POSTSUPERSCRIPT ( italic_m ) end_POSTSUPERSCRIPT ( italic_z ) = ( - 1 ) start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT italic_m ! roman_Res start_POSTSUBSCRIPT italic_z = - italic_k end_POSTSUBSCRIPT ( italic_h ( italic_z ) ) ( italic_z + italic_k ) start_POSTSUPERSCRIPT - italic_m - 1 end_POSTSUPERSCRIPT + italic_O ( 1 ). Thus, the coeffcient cm1(k)subscriptsuperscript𝑐𝑘𝑚1c^{(k)}_{-m-1}italic_c start_POSTSUPERSCRIPT ( italic_k ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - italic_m - 1 end_POSTSUBSCRIPT of h(m)superscript𝑚h^{(m)}italic_h start_POSTSUPERSCRIPT ( italic_m ) end_POSTSUPERSCRIPT is given by (1)mm!Resz=k(h(z))superscript1𝑚𝑚subscriptRes𝑧𝑘𝑧(-1)^{m}m!\mathop{\operatorname{Res}}_{z=-k}(h(z))( - 1 ) start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT italic_m ! roman_Res start_POSTSUBSCRIPT italic_z = - italic_k end_POSTSUBSCRIPT ( italic_h ( italic_z ) ) and all other coefficients cjsubscript𝑐𝑗c_{j}italic_c start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT with j<0𝑗0j<0italic_j < 0 vanish. By Theorem 5, we have

h(m)(s)g(s)superscript𝑚𝑠𝑔𝑠\displaystyle h^{(m)}(s)g(-s)italic_h start_POSTSUPERSCRIPT ( italic_m ) end_POSTSUPERSCRIPT ( italic_s ) italic_g ( - italic_s ) =0xs1k=0(1)mΓ(m+1)cm1[(ddz+log(x))mg(z)]z=kxkdxabsentsuperscriptsubscript0superscript𝑥𝑠1superscriptsubscript𝑘0superscript1𝑚Γ𝑚1subscript𝑐𝑚1subscriptdelimited-[]superscript𝑑𝑑𝑧𝑥𝑚𝑔𝑧𝑧𝑘superscript𝑥𝑘𝑑𝑥\displaystyle=\int_{0}^{\infty}x^{s-1}\sum_{k=0}^{\infty}\frac{(-1)^{m}}{% \Gamma(m+1)}c_{-m-1}\left[\left(\frac{d}{dz}+\log(x)\right)^{m}g(z)\right]_{z=% k}x^{k}dx= ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT italic_s - 1 end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT divide start_ARG ( - 1 ) start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT end_ARG start_ARG roman_Γ ( italic_m + 1 ) end_ARG italic_c start_POSTSUBSCRIPT - italic_m - 1 end_POSTSUBSCRIPT [ ( divide start_ARG italic_d end_ARG start_ARG italic_d italic_z end_ARG + roman_log ( italic_x ) ) start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT italic_g ( italic_z ) ] start_POSTSUBSCRIPT italic_z = italic_k end_POSTSUBSCRIPT italic_x start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_d italic_x
=0xs1k=0Resz=k(h(z))[(ddz+log(x))mg(z)]z=kxkdx,absentsuperscriptsubscript0superscript𝑥𝑠1superscriptsubscript𝑘0subscriptRes𝑧𝑘𝑧subscriptdelimited-[]superscript𝑑𝑑𝑧𝑥𝑚𝑔𝑧𝑧𝑘superscript𝑥𝑘𝑑𝑥\displaystyle=\int_{0}^{\infty}x^{s-1}\sum_{k=0}^{\infty}\mathop{\operatorname% {Res}}_{z=-k}(h(z))\left[\left(\frac{d}{dz}+\log(x)\right)^{m}g(z)\right]_{z=k% }x^{k}dx,= ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT italic_s - 1 end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT roman_Res start_POSTSUBSCRIPT italic_z = - italic_k end_POSTSUBSCRIPT ( italic_h ( italic_z ) ) [ ( divide start_ARG italic_d end_ARG start_ARG italic_d italic_z end_ARG + roman_log ( italic_x ) ) start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT italic_g ( italic_z ) ] start_POSTSUBSCRIPT italic_z = italic_k end_POSTSUBSCRIPT italic_x start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_d italic_x , (5.10)

and this completes the proof. ∎

This corollary has several interesting special cases. For example, Corollary 2.1 can now be generalized to any derivative of the digamma function.

Corollary 5.2.

Let g(z)𝑔𝑧g(z)italic_g ( italic_z ) be analytic on the half-plane H(δ)={z|Re(z)δ}𝐻𝛿conditional-set𝑧Re𝑧𝛿H(\delta)=\{z\in\mathbb{C}|\textnormal{Re}(z)\geq-\delta\}italic_H ( italic_δ ) = { italic_z ∈ blackboard_C | Re ( italic_z ) ≥ - italic_δ } for some 0<δ<10𝛿10<\delta<10 < italic_δ < 1. Then

0xs1k=0superscriptsubscript0superscript𝑥𝑠1superscriptsubscript𝑘0\displaystyle\int_{0}^{\infty}x^{s-1}\sum_{k=0}^{\infty}∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT italic_s - 1 end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT [(ddz+log(x))mg(z)]z=kxkdx=ψ(m)(s)g(s)subscriptdelimited-[]superscript𝑑𝑑𝑧𝑥𝑚𝑔𝑧𝑧𝑘superscript𝑥𝑘𝑑𝑥superscript𝜓𝑚𝑠𝑔𝑠\displaystyle\bigg{[}\bigg{(}\frac{d}{dz}+\log(x)\bigg{)}^{m}g(z)\bigg{]}_{z=k% }x^{k}dx=-\psi^{(m)}(s)g(-s)[ ( divide start_ARG italic_d end_ARG start_ARG italic_d italic_z end_ARG + roman_log ( italic_x ) ) start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT italic_g ( italic_z ) ] start_POSTSUBSCRIPT italic_z = italic_k end_POSTSUBSCRIPT italic_x start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_d italic_x = - italic_ψ start_POSTSUPERSCRIPT ( italic_m ) end_POSTSUPERSCRIPT ( italic_s ) italic_g ( - italic_s ) (5.11)

holds for all 0<Re(s)<δ0Re𝑠𝛿0<\textnormal{Re}(s)<\delta0 < Re ( italic_s ) < italic_δ.

Proof.

The residue of ψ(z)𝜓𝑧\psi(z)italic_ψ ( italic_z ) at z=k𝑧𝑘z=-kitalic_z = - italic_k is 1111. The corollary now follows from Corollary 5.1. ∎

Just as for Corollary 2.1, this result seems to be more of an interesting artifact of the more general theorem, as finding a suitable g𝑔gitalic_g for which the theorem holds seems to be challenging in practice. Another special case of Corollary 5.1 is given by choosing h(z)=πsin(πz)𝑧𝜋𝜋𝑧h(z)=\frac{\pi}{\sin(\pi z)}italic_h ( italic_z ) = divide start_ARG italic_π end_ARG start_ARG roman_sin ( italic_π italic_z ) end_ARG.

Corollary 5.3.

Let g(z)𝑔𝑧g(z)italic_g ( italic_z ) be analytic on the half-plane H(δ)={z|Re(z)δ}𝐻𝛿conditional-set𝑧Re𝑧𝛿H(\delta)=\{z\in\mathbb{C}|\textnormal{Re}(z)\geq-\delta\}italic_H ( italic_δ ) = { italic_z ∈ blackboard_C | Re ( italic_z ) ≥ - italic_δ } for some 0<δ<10𝛿10<\delta<10 < italic_δ < 1. Then

0xs1k=0(1)k[(ddz+log(x))mg(z)]z=kxkdx=dmdsm(πsin(πs))g(s)superscriptsubscript0superscript𝑥𝑠1superscriptsubscript𝑘0superscript1𝑘subscriptdelimited-[]superscript𝑑𝑑𝑧𝑥𝑚𝑔𝑧𝑧𝑘superscript𝑥𝑘𝑑𝑥superscript𝑑𝑚𝑑superscript𝑠𝑚𝜋𝜋𝑠𝑔𝑠\displaystyle\int_{0}^{\infty}x^{s-1}\sum_{k=0}^{\infty}(-1)^{k}\bigg{[}\bigg{% (}\frac{d}{dz}+\log(x)\bigg{)}^{m}g(z)\bigg{]}_{z=k}x^{k}dx=\frac{d^{m}}{ds^{m% }}\left(\frac{\pi}{\sin(\pi s)}\right)g(-s)∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT italic_s - 1 end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT ( - 1 ) start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT [ ( divide start_ARG italic_d end_ARG start_ARG italic_d italic_z end_ARG + roman_log ( italic_x ) ) start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT italic_g ( italic_z ) ] start_POSTSUBSCRIPT italic_z = italic_k end_POSTSUBSCRIPT italic_x start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_d italic_x = divide start_ARG italic_d start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT end_ARG start_ARG italic_d italic_s start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT end_ARG ( divide start_ARG italic_π end_ARG start_ARG roman_sin ( italic_π italic_s ) end_ARG ) italic_g ( - italic_s ) (5.12)

holds for all 0<Re(s)<δ0Re𝑠𝛿0<\textnormal{Re}(s)<\delta0 < Re ( italic_s ) < italic_δ.

Proof.

The residue of πsin(πz)𝜋𝜋𝑧\frac{\pi}{\sin(\pi z)}divide start_ARG italic_π end_ARG start_ARG roman_sin ( italic_π italic_z ) end_ARG at z=k𝑧𝑘z=-kitalic_z = - italic_k is (1)ksuperscript1𝑘(-1)^{k}( - 1 ) start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT. The corollary now follows from Corollary 5.1. ∎

This case is a very natural generalization of Ramanujan’s theorem in the original form (2.2). Recall that the substitution g(z)g(z)/Γ(z+1)𝑔𝑧𝑔𝑧Γ𝑧1g(z)\to g(z)/\Gamma(z+1)italic_g ( italic_z ) → italic_g ( italic_z ) / roman_Γ ( italic_z + 1 ) and an application of Euler’s reflection formula produces the form (3.22). Another special case given by choosing h(z)=Γ(z)𝑧Γ𝑧h(z)=\Gamma(z)italic_h ( italic_z ) = roman_Γ ( italic_z ) can be seen as a natural generalization of this latter form of Ramanujan’s master theorem.

Corollary 5.4.

Let g(z)𝑔𝑧g(z)italic_g ( italic_z ) be analytic on the half-plane H(δ)={z|Re(z)δ}𝐻𝛿conditional-set𝑧Re𝑧𝛿H(\delta)=\{z\in\mathbb{C}|\textnormal{Re}(z)\geq-\delta\}italic_H ( italic_δ ) = { italic_z ∈ blackboard_C | Re ( italic_z ) ≥ - italic_δ } for some 0<δ<10𝛿10<\delta<10 < italic_δ < 1. Then

0xs1k=0(1)kk![(ddz+log(x))mg(z)]z=kxkdx=Γ(m)(s)g(s)superscriptsubscript0superscript𝑥𝑠1superscriptsubscript𝑘0superscript1𝑘𝑘subscriptdelimited-[]superscript𝑑𝑑𝑧𝑥𝑚𝑔𝑧𝑧𝑘superscript𝑥𝑘𝑑𝑥superscriptΓ𝑚𝑠𝑔𝑠\displaystyle\int_{0}^{\infty}x^{s-1}\sum_{k=0}^{\infty}\frac{(-1)^{k}}{k!}% \bigg{[}\bigg{(}\frac{d}{dz}+\log(x)\bigg{)}^{m}g(z)\bigg{]}_{z=k}x^{k}dx=% \Gamma^{(m)}(s)g(-s)∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT italic_s - 1 end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT divide start_ARG ( - 1 ) start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT end_ARG start_ARG italic_k ! end_ARG [ ( divide start_ARG italic_d end_ARG start_ARG italic_d italic_z end_ARG + roman_log ( italic_x ) ) start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT italic_g ( italic_z ) ] start_POSTSUBSCRIPT italic_z = italic_k end_POSTSUBSCRIPT italic_x start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_d italic_x = roman_Γ start_POSTSUPERSCRIPT ( italic_m ) end_POSTSUPERSCRIPT ( italic_s ) italic_g ( - italic_s ) (5.13)

holds for all 0<Re(s)<δ0Re𝑠𝛿0<\textnormal{Re}(s)<\delta0 < Re ( italic_s ) < italic_δ.

Proof.

The residue of Γ(z)Γ𝑧\Gamma(z)roman_Γ ( italic_z ) at z=k𝑧𝑘z=-kitalic_z = - italic_k is (1)k/k!superscript1𝑘𝑘(-1)^{k}/k!( - 1 ) start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT / italic_k !. The corollary now follows from Corollary 5.1. ∎

Note that by setting g=1𝑔1g=1italic_g = 1 the last two corollaries give integral representations of the derivatives of the gamma and cosecant functions. The next corollary takes hhitalic_h to be the square of the gamma function and produces an integral representation involving Euler’s constant γ𝛾\gammaitalic_γ and the harmonic numbers Hksubscript𝐻𝑘H_{k}italic_H start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT.

Corollary 5.5.

Let g(z)𝑔𝑧g(z)italic_g ( italic_z ) be analytic on the half-plane H(δ)={z|Re(z)δ}𝐻𝛿conditional-set𝑧Re𝑧𝛿H(\delta)=\{z\in\mathbb{C}|\textnormal{Re}(z)\geq-\delta\}italic_H ( italic_δ ) = { italic_z ∈ blackboard_C | Re ( italic_z ) ≥ - italic_δ } for some 0<δ<10𝛿10<\delta<10 < italic_δ < 1. Then

0k=0superscriptsubscript0superscriptsubscript𝑘0\displaystyle\int_{0}^{\infty}\sum_{k=0}^{\infty}∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT g(k)(2γ2Hk+log(x))g(k)(k!)2xk+s1dx=(Γ(s))2g(s)superscript𝑔𝑘2𝛾2subscript𝐻𝑘𝑥𝑔𝑘superscript𝑘2superscript𝑥𝑘𝑠1𝑑𝑥superscriptΓ𝑠2𝑔𝑠\displaystyle\frac{g^{\prime}(k)-(2\gamma-2H_{k}+\log(x))g(k)}{(k!)^{2}}x^{k+s% -1}dx=(\Gamma(s))^{2}g(-s)divide start_ARG italic_g start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_k ) - ( 2 italic_γ - 2 italic_H start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT + roman_log ( italic_x ) ) italic_g ( italic_k ) end_ARG start_ARG ( italic_k ! ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG italic_x start_POSTSUPERSCRIPT italic_k + italic_s - 1 end_POSTSUPERSCRIPT italic_d italic_x = ( roman_Γ ( italic_s ) ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_g ( - italic_s ) (5.14)

holds for all 0<Re(s)<δ0Re𝑠𝛿0<\textnormal{Re}(s)<\delta0 < Re ( italic_s ) < italic_δ.

Proof.

Consider the square of the Gamma function. With the aid of a computer, the residue calculation yields

Resz=k((Γ(z))2g(z)xz)=(g(k)2γg(k)+2Hkg(k)log(x)g(k))(k!)2xk,subscriptRes𝑧𝑘superscriptΓ𝑧2𝑔𝑧superscript𝑥𝑧superscript𝑔𝑘2𝛾𝑔𝑘2subscript𝐻𝑘𝑔𝑘𝑥𝑔𝑘superscript𝑘2superscript𝑥𝑘\displaystyle\mathop{\operatorname{Res}}_{z=-k}((\Gamma(z))^{2}g(-z)x^{-z})=% \frac{(g^{\prime}(k)-2\gamma g(k)+2H_{k}g(k)-\log(x)g(k))}{(k!)^{2}}x^{k},roman_Res start_POSTSUBSCRIPT italic_z = - italic_k end_POSTSUBSCRIPT ( ( roman_Γ ( italic_z ) ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_g ( - italic_z ) italic_x start_POSTSUPERSCRIPT - italic_z end_POSTSUPERSCRIPT ) = divide start_ARG ( italic_g start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_k ) - 2 italic_γ italic_g ( italic_k ) + 2 italic_H start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_g ( italic_k ) - roman_log ( italic_x ) italic_g ( italic_k ) ) end_ARG start_ARG ( italic_k ! ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG italic_x start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT , (5.15)

where γ𝛾\gammaitalic_γ is Euler’s constant and Hksubscript𝐻𝑘H_{k}italic_H start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT is the k𝑘kitalic_k-th harmonic number. Substituting the above expression in Theorem 5 yields the desired result. ∎

With g=1𝑔1g=1italic_g = 1, we obtain an integral representation for the gamma function given by

(Γ(s))2=0k=02γ2Hk+log(x)(k!)2xk+s1dx,superscriptΓ𝑠2superscriptsubscript0superscriptsubscript𝑘02𝛾2subscript𝐻𝑘𝑥superscript𝑘2superscript𝑥𝑘𝑠1𝑑𝑥(\Gamma(s))^{2}=-\int_{0}^{\infty}\sum_{k=0}^{\infty}\frac{2\gamma-2H_{k}+\log% (x)}{(k!)^{2}}x^{k+s-1}dx,( roman_Γ ( italic_s ) ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = - ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT divide start_ARG 2 italic_γ - 2 italic_H start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT + roman_log ( italic_x ) end_ARG start_ARG ( italic_k ! ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG italic_x start_POSTSUPERSCRIPT italic_k + italic_s - 1 end_POSTSUPERSCRIPT italic_d italic_x , (5.16)

and specializing to s=1/2𝑠12s=1/2italic_s = 1 / 2, we recover an integral representation of π𝜋\piitalic_π that involves Euler’s constant and the harmonic numbers. Explicitly, we have

π=0k=02γ2Hk+log(x)(k!)2xk1/2dx.𝜋superscriptsubscript0superscriptsubscript𝑘02𝛾2subscript𝐻𝑘𝑥superscript𝑘2superscript𝑥𝑘12𝑑𝑥\pi=-\int_{0}^{\infty}\sum_{k=0}^{\infty}\frac{2\gamma-2H_{k}+\log(x)}{(k!)^{2% }}x^{k-1/2}dx.italic_π = - ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT divide start_ARG 2 italic_γ - 2 italic_H start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT + roman_log ( italic_x ) end_ARG start_ARG ( italic_k ! ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG italic_x start_POSTSUPERSCRIPT italic_k - 1 / 2 end_POSTSUPERSCRIPT italic_d italic_x . (5.17)

The sum in the integrand can be evaluated explicitly in terms of the modified Bessel function of the second kind, producing

π2=0K0(2x)x𝑑x,𝜋2superscriptsubscript0subscript𝐾02𝑥𝑥differential-d𝑥\frac{\pi}{2}=\int_{0}^{\infty}\frac{K_{0}(2\sqrt{x})}{\sqrt{x}}dx,divide start_ARG italic_π end_ARG start_ARG 2 end_ARG = ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT divide start_ARG italic_K start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( 2 square-root start_ARG italic_x end_ARG ) end_ARG start_ARG square-root start_ARG italic_x end_ARG end_ARG italic_d italic_x , (5.18)

which can be verified numerically with Mathematica. We again recover Bernoulli’s integral representation of the gamma function by setting g(z)=sin(πz)Γ(z+1)𝑔𝑧𝜋𝑧Γ𝑧1g(z)=\sin(\pi z)\Gamma(z+1)italic_g ( italic_z ) = roman_sin ( italic_π italic_z ) roman_Γ ( italic_z + 1 ), in which case g(k)=0𝑔𝑘0g(k)=0italic_g ( italic_k ) = 0 and g(k)=π(1)kk!superscript𝑔𝑘𝜋superscript1𝑘𝑘g^{\prime}(k)=\pi(-1)^{k}k!italic_g start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_k ) = italic_π ( - 1 ) start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_k !, and we have

Γ(s)=0xs1ex𝑑x.Γ𝑠superscriptsubscript0superscript𝑥𝑠1superscript𝑒𝑥differential-d𝑥\Gamma(s)=\int_{0}^{\infty}x^{s-1}e^{-x}dx.roman_Γ ( italic_s ) = ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT italic_s - 1 end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - italic_x end_POSTSUPERSCRIPT italic_d italic_x . (5.19)

Another option available to us is to replace π/sin(πs)𝜋𝜋𝑠\pi/\sin(\pi s)italic_π / roman_sin ( italic_π italic_s ) with its m𝑚mitalic_m-th power πm/sinm(πs)superscript𝜋𝑚superscript𝑚𝜋𝑠\pi^{m}/\sin^{m}(\pi s)italic_π start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT / roman_sin start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT ( italic_π italic_s ). The residue computation becomes much more complicated because the poles at the non-positive integers are no longer simple; they have order m𝑚mitalic_m. However, by evaluating the first several cases numerically, we are able to formulate a conjecture for this scenario. Define a polynomial sequence Pm(x)subscript𝑃𝑚𝑥P_{m}(x)italic_P start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ( italic_x ) recursively by P1(x)=1subscript𝑃1𝑥1P_{1}(x)=1italic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_x ) = 1, P2(x)=xsubscript𝑃2𝑥𝑥P_{2}(x)=xitalic_P start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_x ) = italic_x, and

Pm(x):=(x2+(m2)2π2)Pm2(x)assignsubscript𝑃𝑚𝑥superscript𝑥2superscript𝑚22superscript𝜋2subscript𝑃𝑚2𝑥\displaystyle P_{m}(x):=(x^{2}+(m-2)^{2}\pi^{2})P_{m-2}(x)italic_P start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ( italic_x ) := ( italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + ( italic_m - 2 ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_π start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) italic_P start_POSTSUBSCRIPT italic_m - 2 end_POSTSUBSCRIPT ( italic_x ) (5.20)

for m>2𝑚2m>2italic_m > 2. These polynomials appear in a work of H. Airault on Fourier transform computations related to hyperbolic measures .

Conjecture 1.

Let g(z)𝑔𝑧g(z)italic_g ( italic_z ) be analytic on the half-plane H(δ)={z|Re(z)δ}𝐻𝛿conditional-set𝑧Re𝑧𝛿H(\delta)=\{z\in\mathbb{C}|\textnormal{Re}(z)\geq-\delta\}italic_H ( italic_δ ) = { italic_z ∈ blackboard_C | Re ( italic_z ) ≥ - italic_δ } for some 0<δ<10𝛿10<\delta<10 < italic_δ < 1 and let m𝑚mitalic_m be a positive integer. Then

0xs1n=0(1)mn[Pm(ddz+log(x))g(z)]z=nxndxsuperscriptsubscript0superscript𝑥𝑠1superscriptsubscript𝑛0superscript1𝑚𝑛subscriptdelimited-[]subscript𝑃𝑚𝑑𝑑𝑧𝑥𝑔𝑧𝑧𝑛superscript𝑥𝑛𝑑𝑥\displaystyle\int_{0}^{\infty}x^{s-1}\sum_{n=0}^{\infty}(-1)^{mn}\left[P_{m}% \left(\frac{d}{dz}+\log(x)\right)g(z)\right]_{z=n}x^{n}dx∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT italic_s - 1 end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT ( - 1 ) start_POSTSUPERSCRIPT italic_m italic_n end_POSTSUPERSCRIPT [ italic_P start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ( divide start_ARG italic_d end_ARG start_ARG italic_d italic_z end_ARG + roman_log ( italic_x ) ) italic_g ( italic_z ) ] start_POSTSUBSCRIPT italic_z = italic_n end_POSTSUBSCRIPT italic_x start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT italic_d italic_x =(1)m1(m1)!πmsinm(πs)g(s)absentsuperscript1𝑚1𝑚1superscript𝜋𝑚superscript𝑚𝜋𝑠𝑔𝑠\displaystyle=\frac{(-1)^{m-1}(m-1)!\pi^{m}}{\sin^{m}(\pi s)}g(-s)= divide start_ARG ( - 1 ) start_POSTSUPERSCRIPT italic_m - 1 end_POSTSUPERSCRIPT ( italic_m - 1 ) ! italic_π start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT end_ARG start_ARG roman_sin start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT ( italic_π italic_s ) end_ARG italic_g ( - italic_s ) (5.21)

holds for all 0<Re(s)<δ0Re𝑠𝛿0<\textnormal{Re}(s)<\delta0 < Re ( italic_s ) < italic_δ and some suitable growth conditions on g𝑔gitalic_g.

Let us give several examples using this conjecture. By setting g=1𝑔1g=1italic_g = 1, we recover an interesting integral representation for the m𝑚mitalic_m-th power of the cosecant function given by

0xs1n=0(1)mnPm(log(x))xndxsuperscriptsubscript0superscript𝑥𝑠1superscriptsubscript𝑛0superscript1𝑚𝑛subscript𝑃𝑚𝑥superscript𝑥𝑛𝑑𝑥\displaystyle\int_{0}^{\infty}x^{s-1}\sum_{n=0}^{\infty}(-1)^{mn}P_{m}\left(% \log(x)\right)x^{n}dx∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT italic_s - 1 end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT ( - 1 ) start_POSTSUPERSCRIPT italic_m italic_n end_POSTSUPERSCRIPT italic_P start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ( roman_log ( italic_x ) ) italic_x start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT italic_d italic_x =(1)m1(m1)!πmsinm(πs).absentsuperscript1𝑚1𝑚1superscript𝜋𝑚superscript𝑚𝜋𝑠\displaystyle=(-1)^{m-1}(m-1)!\frac{\pi^{m}}{\sin^{m}(\pi s)}.= ( - 1 ) start_POSTSUPERSCRIPT italic_m - 1 end_POSTSUPERSCRIPT ( italic_m - 1 ) ! divide start_ARG italic_π start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT end_ARG start_ARG roman_sin start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT ( italic_π italic_s ) end_ARG . (5.22)

Letting g(z)=1/Γ(z+1)𝑔𝑧1Γ𝑧1g(z)=1/\Gamma(z+1)italic_g ( italic_z ) = 1 / roman_Γ ( italic_z + 1 ), we instead find that

0xs1n=0(1)mn[Pm(ddz+log(x))1Γ(z+1)]z=nxndxsuperscriptsubscript0superscript𝑥𝑠1superscriptsubscript𝑛0superscript1𝑚𝑛subscriptdelimited-[]subscript𝑃𝑚𝑑𝑑𝑧𝑥1Γ𝑧1𝑧𝑛superscript𝑥𝑛𝑑𝑥\displaystyle\int_{0}^{\infty}x^{s-1}\sum_{n=0}^{\infty}(-1)^{mn}\left[P_{m}% \left(\frac{d}{dz}+\log(x)\right)\frac{1}{\Gamma(z+1)}\right]_{z=n}x^{n}dx∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT italic_s - 1 end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT ( - 1 ) start_POSTSUPERSCRIPT italic_m italic_n end_POSTSUPERSCRIPT [ italic_P start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ( divide start_ARG italic_d end_ARG start_ARG italic_d italic_z end_ARG + roman_log ( italic_x ) ) divide start_ARG 1 end_ARG start_ARG roman_Γ ( italic_z + 1 ) end_ARG ] start_POSTSUBSCRIPT italic_z = italic_n end_POSTSUBSCRIPT italic_x start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT italic_d italic_x =(1)m1(m1)!πmsinm(πs)Γ(1s).absentsuperscript1𝑚1𝑚1superscript𝜋𝑚superscript𝑚𝜋𝑠Γ1𝑠\displaystyle=\frac{(-1)^{m-1}(m-1)!\pi^{m}}{\sin^{m}(\pi s)\Gamma(1-s)}.= divide start_ARG ( - 1 ) start_POSTSUPERSCRIPT italic_m - 1 end_POSTSUPERSCRIPT ( italic_m - 1 ) ! italic_π start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT end_ARG start_ARG roman_sin start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT ( italic_π italic_s ) roman_Γ ( 1 - italic_s ) end_ARG . (5.23)

For the special case of m=2𝑚2m=2italic_m = 2 (the m=1𝑚1m=1italic_m = 1 case is Ramanujan’s master theorem), Mathematica is able to evaluate the summation explicitly, producing

n=0[P2(ddz+log(x))1Γ(z+1)]z=nxn=exΓ(0,x),superscriptsubscript𝑛0subscriptdelimited-[]subscript𝑃2𝑑𝑑𝑧𝑥1Γ𝑧1𝑧𝑛superscript𝑥𝑛superscript𝑒𝑥Γ0𝑥\displaystyle\sum_{n=0}^{\infty}\left[P_{2}\left(\frac{d}{dz}+\log(x)\right)% \frac{1}{\Gamma(z+1)}\right]_{z=n}x^{n}=-e^{x}\Gamma(0,x),∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT [ italic_P start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( divide start_ARG italic_d end_ARG start_ARG italic_d italic_z end_ARG + roman_log ( italic_x ) ) divide start_ARG 1 end_ARG start_ARG roman_Γ ( italic_z + 1 ) end_ARG ] start_POSTSUBSCRIPT italic_z = italic_n end_POSTSUBSCRIPT italic_x start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT = - italic_e start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT roman_Γ ( 0 , italic_x ) , (5.24)

where Γ(a,z)Γ𝑎𝑧\Gamma(a,z)roman_Γ ( italic_a , italic_z ) denotes the incomplete gamma function. Thus, the conjecture induces the integral representation

0xs1exΓ(0,x)𝑑x=π2sin2(πs)Γ(1s),superscriptsubscript0superscript𝑥𝑠1superscript𝑒𝑥Γ0𝑥differential-d𝑥superscript𝜋2superscript2𝜋𝑠Γ1𝑠\displaystyle\int_{0}^{\infty}x^{s-1}e^{x}\Gamma(0,x)dx=\frac{\pi^{2}}{\sin^{2% }(\pi s)\Gamma(1-s)},∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT italic_s - 1 end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT roman_Γ ( 0 , italic_x ) italic_d italic_x = divide start_ARG italic_π start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG roman_sin start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_π italic_s ) roman_Γ ( 1 - italic_s ) end_ARG , (5.25)

and after applying Euler’s reflection formula, we have

0xs1exΓ(0,x)𝑑x=πsin(πs)Γ(s).superscriptsubscript0superscript𝑥𝑠1superscript𝑒𝑥Γ0𝑥differential-d𝑥𝜋𝜋𝑠Γ𝑠\displaystyle\int_{0}^{\infty}x^{s-1}e^{x}\Gamma(0,x)dx=\frac{\pi}{\sin(\pi s)% }\Gamma(s).∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT italic_s - 1 end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT roman_Γ ( 0 , italic_x ) italic_d italic_x = divide start_ARG italic_π end_ARG start_ARG roman_sin ( italic_π italic_s ) end_ARG roman_Γ ( italic_s ) . (5.26)

For m>2𝑚2m>2italic_m > 2, Mathematica is both unable to give a closed-form evaluation for the summation and unable to evaluate the integral for arbitrary s𝑠sitalic_s.

As a final example, let g(z)=1z+1𝑔𝑧1𝑧1g(z)=\frac{1}{z+1}italic_g ( italic_z ) = divide start_ARG 1 end_ARG start_ARG italic_z + 1 end_ARG and set m=2𝑚2m=2italic_m = 2. Then the summation is evaluated numerically as

n=0[P2(ddz+log(x))1z+1]z=nxn=log(1x)log(x)+Li2(x)x,superscriptsubscript𝑛0subscriptdelimited-[]subscript𝑃2𝑑𝑑𝑧𝑥1𝑧1𝑧𝑛superscript𝑥𝑛1𝑥𝑥subscriptLi2𝑥𝑥\displaystyle\sum_{n=0}^{\infty}\left[P_{2}\left(\frac{d}{dz}+\log(x)\right)% \frac{1}{z+1}\right]_{z=n}x^{n}=\frac{\log(1-x)\log(x)+\mathop{\operatorname{% Li}}_{2}(x)}{x},∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT [ italic_P start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( divide start_ARG italic_d end_ARG start_ARG italic_d italic_z end_ARG + roman_log ( italic_x ) ) divide start_ARG 1 end_ARG start_ARG italic_z + 1 end_ARG ] start_POSTSUBSCRIPT italic_z = italic_n end_POSTSUBSCRIPT italic_x start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT = divide start_ARG roman_log ( 1 - italic_x ) roman_log ( italic_x ) + roman_Li start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_x ) end_ARG start_ARG italic_x end_ARG , (5.27)

where Lin(x)subscriptLi𝑛𝑥\mathop{\operatorname{Li}}_{n}(x)roman_Li start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) denotes the polylogarithm. Thus, the conjecture produces the evaluation

0xs2(log(1x)log(x)+Li2(x))𝑑x=π2sin2(πs)11s.superscriptsubscript0superscript𝑥𝑠21𝑥𝑥subscriptLi2𝑥differential-d𝑥superscript𝜋2superscript2𝜋𝑠11𝑠\displaystyle\int_{0}^{\infty}x^{s-2}\left(\log(1-x)\log(x)+\mathrm{Li}_{2}(x)% \right)dx=-\frac{\pi^{2}}{\sin^{2}(\pi s)}\frac{1}{1-s}.∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT italic_s - 2 end_POSTSUPERSCRIPT ( roman_log ( 1 - italic_x ) roman_log ( italic_x ) + roman_Li start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_x ) ) italic_d italic_x = - divide start_ARG italic_π start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG roman_sin start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_π italic_s ) end_ARG divide start_ARG 1 end_ARG start_ARG 1 - italic_s end_ARG . (5.28)

Similarly, for m = 3, we obtain

0xs2(π2log(1+x)\displaystyle\int_{0}^{\infty}x^{s-2}\bigg{(}\pi^{2}\log(1+x)∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT italic_s - 2 end_POSTSUPERSCRIPT ( italic_π start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_log ( 1 + italic_x ) +log2(x)log(1+x)superscript2𝑥1𝑥\displaystyle+\log^{2}(x)\log(1+x)+ roman_log start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_x ) roman_log ( 1 + italic_x )
+2log(x)Li2(x)2Li3(x))dx=2π3(1s)sin3(πs).\displaystyle+2\log(x)\mathrm{Li}_{2}(-x)-2\mathrm{Li}_{3}(-x)\bigg{)}dx=\frac% {2\pi^{3}}{(1-s)\sin^{3}(\pi s)}.+ 2 roman_log ( italic_x ) roman_Li start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( - italic_x ) - 2 roman_L roman_i start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ( - italic_x ) ) italic_d italic_x = divide start_ARG 2 italic_π start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG start_ARG ( 1 - italic_s ) roman_sin start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ( italic_π italic_s ) end_ARG . (5.29)

§V. Conclusion

In this paper, we have explored an extension of Ramanujan’s master theorem by generalizing Hardy’s proof to accommodate a broader class of functions. Our examination began with a review of Ramanujan’s original theorem and Hardy’s rigorous approach, which utilized Cauchy’s residue theory and the Mellin inversion theorem. By modifying Hardy’s residue calculations, we extended the theorem to a family of Ramanujan-like master theorems, which offer valuable insights and new results in the analysis of Mellin transforms.

Through this generalization, we have derived novel integral representations for meromorphic functions with simple poles at non-positive integers and demonstrated their nice properties, such as logarithmic convexity. Furthermore, we extended our results to meromorphic functions with poles of arbitrary order, uncovering several exotic integral representations that enrich our understanding of these functions.

Our results not only contribute to the theoretical development of mathematical transforms but also have practical implications across various fields. The Mellin transform’s applications in computer algorithms, quantum theory, asymptotic analysis, and other domains underscore the significance of our findings. By providing a broader toolkit for evaluating Mellin transforms, our generalization offers new avenues for research and application, reflecting the enduring influence of Ramanujan’s insights and the continued relevance of his mathematical legacy. We hope that these results will inspire further exploration and application, continuing the tradition of building upon the foundational work of great mathematicians like Ramanujan.

Acknowledgement: ZPB thanks Christophe Vignat for providing the reference to the work of H. Airault.

References

  • [1] Helene Airault. Hyperbolic measures, moments and coefficients. algebra on hyperbolic functions. Journal of Functional Analysis, 255(9):2099–2145, 2008. Special issue dedicated to Paul Malliavin.
  • [2] Tewodros Amdeberhan, Olivier R. Espinosa, Ivan Gonzalez, Marshall Harrison, Victor H. Moll, and Armin Straub. Ramanujan’s master theorem. The Ramanujan Journal, 29(1):103–120, 2012.
  • [3] O. Atale. Analytic expressions for some Mellin transforms with their application to prime counting function and interpolation formulas for the zeta function. Palestine Journal of Mathematics, 12(1), 2023.
  • [4] Bruce Berndt. Ramanujan’s Notebooks II. Springer, New York, NY,, 1985.
  • [5] N. Bleistein and R.A. Handelsman. Asymptotic Expansions of Integrals. Holt, Rinehart and Winston, 1975.
  • [6] Zachary P. Bradshaw. Explorations in mathematical physics: Special functions in quantum theory and Feynman integrals by the method of brackets. PhD Thesis, Tulane University, 2023.
  • [7] Z. P. Bradshaw, I. Gonzalez, L. Jiu, V. H. Moll, and C. Vignat. Compatibility of the method of brackets with classical integration rules. Open Mathematics 21(1), 20220581, 2023.
  • [8] Zachary P. Bradshaw and Christophe Vignat. An operational calculus generalization of Ramanujan’s master theorem. Journal of Mathematical Analysis and Applications, 523(2):127029, 2023.
  • [9] Muhammed Aslam Chaudhry and Asghar Qadir. Extension of Hardy’s class for Ramanujan’s interpolation formula and master theorem with applications. Journal of Inequalities and Applications, 2012(1):1–13, 2012.
  • [10] L. Debnath and D. Bhatta. Integral Transforms and Their Applications, Third Edition. Taylor and Francis, 2014.
  • [11] Hongming Ding, Kenneth Gross, and Donald Richards. Ramanujan’s master theorem for symmetric cones. Pacific Journal of Mathematics, 175(2):447–490, 1996.
  • [12] Dragomir, Sever S, Agarwal, R. P and Barnett, Neil S (1999) Inequalities for Beta and Gamma Functions Via Some Classical and New Integral Inequalities. RGMIA research report collection, 2 (3).
  • [13] Benjamin Epstein. Some Applications of the Mellin Transform in Statistics. The Annals of Mathematical Statistics, 19(3):370 – 379, 1948.
  • [14] Ahmed Fitouhi, Kamel Brahim, and Neji Bettaibi. On some q-versions of the Ramanujan master theorem. The Ramanujan Journal, 50(2):433–458, 2019.
  • [15] A. Liam Fitzpatrick and Jared Kaplan. Unitarity and the holographic s-matrix. Journal of High Energy Physics, 2012(10), October 2012.
  • [16] A. Liam Fitzpatrick, Jared Kaplan, Joao Penedones, Suvrat Raju, and Balt C. van Rees. A natural language for ads/cft correlators. Journal of High Energy Physics, 2011(11), November 2011. 17
  • [17] Philippe Flajolet, Xavier Gourdon, and Philippe Dumas. Mellin transforms and asymptotics: Harmonic sums. Theoretical Computer Science, 144(1):3–58, 1995.
  • [18] Philippe Flajolet and Robert Sedgewick. The Average Case Analysis of Algorithms: Mellin Transform Asymptotics. Research Report RR-2956, INRIA, 1996.
  • [19] Ivan Gonzalez, Karen Kohl, Lin Jiu, and Victor H. Moll. An extension of the method of brackets. Part 1, 2017.
  • [20] Ivan Gonzalez and Victor H. Moll. Definite integrals by the method of brackets. Advances in Applied Mathematics, 45(1):50–73, 2010.
  • [21] Ivan Gonzalez, Victor H. Moll, and Armin Straub. The method of brackets. part 2: Examples and applications, 2010.
  • [22] Ivan Gonzalez. Method of brackets and Feynman diagram evaluation. Nuclear Physics B - Proceedings Supplements, 205-206:141–146, Aug 2010.
  • [23] Ivan Gonzalez and Ivan Schmidt. Optimized negative dimensional integration method (ndim) and multiloop Feynman diagram calculation. Nuclear Physics B, 769(1):124–173, 2007.
  • [24] D. Griffiths. Introduction to Elementary Particles. Wiley, 2020.
  • [25] G. H. Hardy. On two theorems of F. Carlson and S. Wigert. Acta Mathematica, 42: 327–339, 1920.
  • [26] G. H. Hardy. Ramanujan Twelve Lectures on Subjects Suggested By His Life and Work. Cambridge University Press, 1940.
  • [27] P. A. Martin. Some applications of the Mellin transform to asymptotics of series. In Inverse Scattering and Potential Problems in Mathematical Physics, Frankfurt, 1995. Peter Lang.
  • [28] I. S. Reed. The Mellin Type of Double Integral. Duke Mathematical Journal, 11(3):565 – 572, 1944.
  • [29] R.J. Sasiela. Electromagnetic Wave Propagation in Turbulence: Evaluation and Application of Mellin Transforms. Springer Series on Wave Phenomena. Springer Berlin Heidelberg, 2012.
  • [30] Matthew D. Schwartz. Quantum Field Theory and the Standard Model. Cambridge University Press, 2013. 18
  • [31] R. Wong. Asymptotic Approximations of Integrals. Classics in Applied Mathematics. Society for Industrial and Applied Mathematics, 2001.