This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

The generalized Bernoulli numbers and its relation with the Riemann zeta function at odd-integer arguments

Yayun Wu
School of Mathematics and Statistics, Hefei Normal University,
Hefei, 230601, People’s Republic of China
E-mail address: [email protected] (Y. Wu).

Abstract: By using the generalized Bernoulli numbers, we deduce new integral representations for the Riemann zeta function at positive odd-integer arguments. The explicit expressions enable us to obtain criteria for the dimension of the vector space spanned over the rational by the ζ(2n+1)/π2n\zeta(2n+1)/\pi^{2n}, n1n\geq 1.

Key words: Riemann zeta function; Integral representation; Generalized Bernoulli numbers; Hyperbolic functions

Mathematics Subject Classification: 11M06; 11C08; 11J72; 30A05

1. Introduction

The Riemann zeta function ζ(s)\zeta(s) is defined for complex values of ss, (s)>1\Re(s)>1 as ζ(s):=v=1vs\zeta(s):=\sum_{v=1}^{\infty}v^{-s}, and its meromorphic continuation over the whole complex ss-plane, except for a simple pole at s=1s=1. The research of the arithmetical properties of the Riemann zeta function at positive integers is one of the fascinating topics of the complex analysis and number theory. Finding recurrence formulas and integral representations of zeta function has become a significant issue. As is well known, for Riemann’s zeta function ζ(s)\zeta(s) at even points, we have the formula (found by Euler)

ζ(2k)=(1)k1(2k)2kB2k2(2k)!,k+.\displaystyle\zeta(2k)=\frac{(-1)^{k-1}(2k)^{2k}B_{2k}}{2(2k)!},~{}k\in\mathbb{N}^{+}.

Here the Bk=Bk(0)B_{k}=B_{k}(0) is the well-tabulated Bernoulli numbers defined by means of the generating function

zezxez1=k=0Bk(x)n!zk,|z|<2π.\displaystyle\frac{ze^{zx}}{e^{z}-1}=\sum_{k=0}^{\infty}\frac{B_{k}(x)}{n!}z^{k},~{}|z|<2\pi.

However, there is no analogous closed evaluation for the values of ζ(s)\zeta(s) at positive odd integers or fractional points up to the present time. Using the series of special functions and the multidimensional integral to investigate ζ(s)\zeta(s) are the main methods. Connecting Euler polynomials and trigonometric functions, Srivastava and Choi[3] give the Riemann zeta function at positive odd integers

ζ(2n+1)=(1)nπ2n+14δ(12(2n+1))(2n)!0δE2n(u)csc(uπ)du,n+,δ=1or1/2,\displaystyle\zeta(2n+1)=\frac{(-1)^{n}\pi^{2n+1}}{4\delta(1-2^{-(2n+1)})(2n)!}\int_{0}^{\delta}E_{2n}(u)\csc(u\pi){\rm d}u,~{}n\in\mathbb{N}^{+},~{}\delta=1~{}{\rm or}~{}1/2,

where En(u)E_{n}(u) is the Euler polynomials. Zurab[1], Cvijovic´\acute{c} and Klinowski[2] establish similar results by using Euler polynomials and the polylogarithm function[8]. Generalizing the trigonometric function to hyperbolic version, Lima[6], D’Avanzo and Krylov[5] research the integral result about ζ(n)\zeta(n),

ζ(n)=2n(2n1)(n1)!0lnn1(coth(x))dx,n2.\displaystyle\zeta(n)=\frac{2^{n}}{(2^{n}-1)(n-1)!}\int_{0}^{\infty}\ln^{n-1}(\coth(x)){\rm d}x,~{}n\geq 2.

These integral representations about ζ(s)\zeta(s) hardly show the correlation with the irrational problem of the zeta function values at odd integers. Since the problem of the irrationality of the values of the zeta function at odd integers is one of the most attractive topics in modern number theory. These problems motivate us to modify the integral representations about ζ(s)\zeta(s).

The article is organized as follows. In §\S 2, we reconsider the evaluation of ζ(3)\zeta(3) so that technical details of the general case do not obscure. In §\S 3, we elaborate the general case by using the generalized Bernoulli polynomials and numbers. In §\S 4, we evaluate the numerical results about ζ(5)\zeta(5) and ζ(7)\zeta(7). In §\S5, the formula for ζ(2n+1)\zeta(2n+1) enables us to obtain linear forms with rational coefficients in the values of ζ(2n+1)/π2n\zeta(2n+1)/\pi^{2n}, n+n\in\mathbb{N}^{+}. We also ponder over the arithmetical properties of the sequence ζ(3)/π2,ζ(5)/π4,ζ(7)/π6\zeta(3)/\pi^{2},\zeta(5)/\pi^{4},\zeta(7)/\pi^{6},\cdots.

2. Evaluation of ζ(3)\zeta(3)

In this part, we rediscover an integral representation for ζ(3)\zeta(3). Let us briefly describe the proof for a better exposition of the general case and its consequences.

Lemma 2.1 For Riemann zeta function value ζ(3)\zeta(3), we have

ζ(3)=2π27+eu(eu1)(eu+1)3u𝑑u.\displaystyle\zeta(3)=\frac{2\pi^{2}}{7}\int_{-\infty}^{+\infty}\frac{e^{u}(e^{u}-1)}{(e^{u}+1)^{3}u}du. (1)

Proof. For the part sum of the series k=1n1(2k1)3\sum_{k=1}^{n}\frac{1}{(2k-1)^{3}}, the following expression is valid

k=1n1(2k1)3=π24γnez(ez1)(ez+1)3zdz,n+,\displaystyle\sum_{k=1}^{n}\frac{1}{(2k-1)^{3}}=\frac{\pi^{2}}{4}\int_{\gamma_{n}}\frac{e^{z}(e^{z}-1)}{(e^{z}+1)^{3}z}{\rm d}z,~{}n\in\mathbb{N}^{+}, (2)

where γn=Γ1(n)+Γ2(n)+Γ3(n)+Γ4(n)+Γε\gamma_{n}=\Gamma_{1}^{(n)}+\Gamma_{2}^{(n)}+\Gamma_{3}^{(n)}+\Gamma_{4}^{(n)}+\Gamma_{\varepsilon} represent the contour with vertices ±Rn\pm R_{n}, Rn+iRnR_{n}+{\rm i}R_{n}, Rn+iRn-R_{n}+{\rm i}R_{n} (see Fig. 1). Γε\Gamma_{\varepsilon} is a semicircle of radius ε(0,π/2)\varepsilon\in(0,\pi/2) centered at the origin. Γ1(n)\Gamma_{1}^{(n)} represent (Rn,0)(ε,0)(-R_{n},0)\rightarrow(-\varepsilon,0) and (ε,0)(Rn,0)(\varepsilon,0)\rightarrow(R_{n},0). Rn=2nπR_{n}=2n\pi. In fact, since in the neighborhood of the point z=0z=0 and z=zk=(2k1)πiz=z_{k}=(2k-1)\pi{\rm i}k=1,2,,nk=1,2,\cdots,n, the following results are valid:

ez1z=O(1),ez(ez+1)3=O((zzk)3).\displaystyle\frac{e^{z}-1}{z}=O(1),\quad\frac{e^{z}}{(e^{z}+1)^{3}}=O((z-z_{k})^{-3}).
Refer to caption
Figure 1: The contour γn\gamma_{n}

Let f(z)=ez(ez1)(ez+1)3z.f(z)=\frac{e^{z}(e^{z}-1)}{(e^{z}+1)^{3}z}. The function f(z)f(z) is bounded on the boundary of the square with poles at the points zkz_{k}. It follows that the integral in (2)(2) is equal to the sum of residues of the integrand at the points z1=πi,z2=3πi,,zn=(2n1)πiz_{1}=\pi{\rm i},z_{2}=3\pi{\rm i},\cdots,z_{n}=(2n-1)\pi{\rm i},

γnf(z)dz=2πik=1nResz=zkf(z)=4π2k=1n1(2k1)3.\displaystyle\int_{\gamma_{n}}f(z){\rm d}z=2\pi{\rm i}\sum_{k=1}^{n}\underset{z=z_{k}}{\text{Res}}f(z)=\frac{4}{\pi^{2}}\sum_{k=1}^{n}\frac{1}{(2k-1)^{3}}. (3)

Furthermore, |f(z)|=O(1RneRn)|f(z)|=O(\frac{1}{R_{n}e^{R_{n}}})zΓj(n),j=2,3,4z\in\Gamma_{j}^{(n)},j=2,3,4. In view of k=11(2k1)3=7ζ(3)/8\sum\limits_{k=1}^{\infty}\frac{1}{(2k-1)^{3}}=7\zeta(3)/8, and let Rn+R_{n}\rightarrow+\infty, ε0\varepsilon\rightarrow 0 in γn\gamma_{n}, we obtain the required expression

+eu(eu1)(eu+1)3udu=72π2ζ(3),\displaystyle\int_{-\infty}^{+\infty}\frac{e^{u}(e^{u}-1)}{(e^{u}+1)^{3}u}{\rm d}u=\frac{7}{2\pi^{2}}\zeta(3), (4)

which leads to the desired result. \hfill\square

Corollary 2.1 Remembering inverse hyperbolic secant function: asech(u)=log(1/u+1/u21){\rm asech}(u)=\log(1/u+\sqrt{1/u^{2}-1})u(0,1]u\in(0,1], then we obtain

ζ(3)=π21401udsech2(u)=π2701uasech(u)du.\displaystyle\zeta(3)=-\frac{\pi^{2}}{14}\int_{0}^{\infty}\frac{1}{u}{\rm d}{\rm sech^{2}(u)}=\frac{\pi^{2}}{7}\int_{0}^{1}\frac{u}{{\rm asech(u)}}{\rm d}u.

By using the integral transformation logu=x\log u=x and remembering that 02(1+x)3xtdx=πt(1t)/sin(πt)\int_{0}^{\infty}2(1+x)^{-3}x^{t}{\rm d}x=\pi t(1-t)/\sin(\pi t), t(0,1)t\in(0,1), we find (4)(4) equals with (22)(22) in [1]:

ζ(3)=2π270x1(1+x)3logxdx=2π27010xt(1+x)3dxdt=170πx(πx)sinxdx,\displaystyle\zeta(3)=\frac{2\pi^{2}}{7}\int_{0}^{\infty}\frac{x-1}{(1+x)^{3}\log x}{\rm d}x=\frac{2\pi^{2}}{7}\int_{0}^{1}\int_{0}^{\infty}\frac{x^{t}}{(1+x)^{3}}{\rm d}x{\rm d}t=\frac{1}{7}\int_{0}^{\pi}\frac{x(\pi-x)}{\sin x}{\rm d}x,

which means the result (4)(4) is a special case of the more general result in [2]. But for ζ(2n+1),n2\zeta(2n+1),n\geq 2, we found no association between these representations.

However, let us expand the integral representation of ζ(3)\zeta(3) into the sum of partial fractions with respect to exp(u)\exp(u):

ζ(3)=2π27l=13+wlu(eu+1)ldu,\displaystyle\zeta(3)=\frac{2\pi^{2}}{7}\sum_{l=1}^{3}\int_{-\infty}^{+\infty}\frac{w_{l}}{u(e^{u}+1)^{l}}{\rm d}u, (5)

where w1=1,w2=3,w3=2w_{1}=1,w_{2}=-3,w_{3}=2. Since l=13wl=0\sum\limits_{l=1}^{3}w_{l}=0, the integrand l=13wlu(eu+1)l\sum\limits_{l=1}^{3}\frac{w_{l}}{u(e^{u}+1)^{l}} is an even function in \mathbb{R} except a removable singularity at origin. We wonder whether the decomposition (5) is unique and can be generalized to the general ζ(2n+1),n2\zeta(2n+1),n\geq 2.

3. Evaluation of ζ(m)\zeta(m) at positive odd integers

3.1 A formula for ζ(m),m2\zeta(m),m\geq 2

The evaluation of ζ(3)\zeta(3) can be straightforwardly generalized to ζ(m)\zeta(m), m2m\geq 2. Before considering the general case, we need the following useful lemma.

Lemma 3.1 Let m2m\geq 2 be a positive integer, and γn\gamma_{n} represent the integral contour in Lemma 2.1. There exists unique wlw_{l}\in\mathbb{Q}, which satisfying l=1mwl=0\sum\limits_{l=1}^{m}w_{l}=0, such that

l=1mγnwlz(ez+1)ldz=2Sm(iπ)m1(ζ(m)2m12mk=n+11(2k1)m),n=1,2,,\displaystyle\sum_{l=1}^{m}\int_{\gamma_{n}}\frac{w_{l}}{z(e^{z}+1)^{l}}{\rm d}z=\frac{2S_{m}}{({\rm i}\pi)^{m-1}}\left(\zeta(m)\frac{2^{m}-1}{2^{m}}-\sum_{k=n+1}^{\infty}\frac{1}{(2k-1)^{m}}\right),~{}n=1,2,\cdots,

wherein SmS_{m}\in\mathbb{Q}.

Proof. Step 1. Let fm(z)=l=1mwlz(ez+1)lf_{m}(z)=\sum\limits_{l=1}^{m}\frac{w_{l}}{z(e^{z}+1)^{l}}. Note that the integrand fm(z)f_{m}(z) over the vertical and top side Γj(n)\Gamma^{(n)}_{j}, j=2,3,4j=2,3,4 tend to 0, when RnR_{n}\rightarrow\infty, since it can be bounded by CeRn/RnCe^{-R_{n}}/R_{n} and constant C>0C>0. Let zk=(2k1)πiz_{k}=(2k-1)\pi{\rm i}k=1,2,,nk=1,2,\cdots,n. By means of the theorem of residues, we have

γnfm(z)dz\displaystyle\int_{\gamma_{n}}f_{m}(z){\rm d}z =2πik=1nlimzzk1(m1)!(ddz)m1l=1mwl(zzk)mz(ez+1)l\displaystyle=2\pi{\rm i}\sum_{k=1}^{n}\lim_{z\rightarrow z_{k}}\frac{1}{(m-1)!}\left(\frac{{\rm d}}{{\rm d}z}\right)^{m-1}\sum_{l=1}^{m}\frac{w_{l}(z-z_{k})^{m}}{z(e^{z}+1)^{l}}
=2πik=1nlimzzk1(m1)!l=1mwlj=0m1(m1j)[(zzk)lz(ez+1)l](j)[(zzk)ml](mj1)\displaystyle=2\pi{\rm i}\sum_{k=1}^{n}\lim_{z\rightarrow z_{k}}\frac{1}{(m-1)!}\sum_{l=1}^{m}w_{l}\sum_{j=0}^{m-1}\binom{m-1}{j}\left[\frac{(z-z_{k})^{l}}{z(e^{z}+1)^{l}}\right]^{(j)}\left[(z-z_{k})^{m-l}\right]^{(m-j-1)}
=2πik=1nl=1mwl(l1)!limzzk[(zzk)lz(ez+1)l](l1).\displaystyle=2\pi{\rm i}\sum_{k=1}^{n}\sum_{l=1}^{m}\frac{w_{l}}{(l-1)!}\lim_{z\rightarrow z_{k}}\left[\frac{(z-z_{k})^{l}}{z(e^{z}+1)^{l}}\right]^{(l-1)}. (6)

Step 2. Let us denote

gl(z)=(zzk)lz(ez+1)l,Dl=limzzkgl(l1)(z)(l1)!,l=1,2,,m.\displaystyle g_{l}(z)=\frac{(z-z_{k})^{l}}{z(e^{z}+1)^{l}},~{}D_{l}=\lim_{z\rightarrow z_{k}}\frac{g_{l}^{(l-1)}(z)}{(l-1)!},~{}l=1,2,\cdots,m. (7)

The conclusions

gl(l1)(z)=j=0l1(l1j)(1z)(l1j)[(zzkez+1)l](j),g_{l}^{(l-1)}(z)=\sum_{j=0}^{l-1}\binom{l-1}{j}\left(\frac{1}{z}\right)^{(l-1-j)}\left[\left(\frac{z-z_{k}}{e^{z}+1}\right)^{l}\right]^{(j)},

follow from the Leibniz rule for the differentiation of a product. For j=0j=0 in gl(l1)(z)g_{l}^{(l-1)}(z), we have

limzzk(l10)(1z)(l1)(zzkez+1)l=(l1)!zkl.\displaystyle\lim_{z\rightarrow z_{k}}\binom{l-1}{0}\left(\frac{1}{z}\right)^{(l-1)}\left(\frac{z-z_{k}}{e^{z}+1}\right)^{l}=-\frac{(l-1)!}{z_{k}^{l}}. (8)

According with [1(ez+1)l]+l(ez+1)l=l(ez+1)l+1\big{[}\frac{1}{(e^{z}+1)^{l}}\big{]}^{\prime}+\frac{l}{(e^{z}+1)^{l}}=\frac{l}{(e^{z}+1)^{l+1}}, which means for all ll, the residues c1(l)c_{-1}^{(l)} of (ez+1)l(e^{z}+1)^{-l} at z=zkz=z_{k} are equal:

c1(l)=c1(1)=12πiΛk1(ez+1)ldz,l=1,2,,m,c_{-1}^{(l)}=c_{-1}^{(1)}=\frac{1}{2\pi{\rm i}}\int_{\Lambda_{k}}\frac{1}{(e^{z}+1)^{l}}{\rm d}z,~{}l=1,2,\cdots,m,

where Λk\Lambda_{k} is the small circle centered at the poles zkz_{k} and of radius ϵk>0\epsilon_{k}>0. When l=1l=1, we acquire c1(l)=c1(1)=limzzk(zzk)/(ez+1)=1c_{-1}^{(l)}=c_{-1}^{(1)}=\lim\limits_{z\rightarrow z_{k}}(z-z_{k})/(e^{z}+1)=-1. For j=l1j=l-1 in gl(l1)(z)g_{l}^{(l-1)}(z), by using Laurent series of (ez+1)l(e^{z}+1)^{-l} at z=zkz=z_{k}, we obtain

limzzk(l1l1)1z[(zzkez+1)l](l1)=limzzk1z[c1(l)(zzk)l1](l1)=(l1)!zk.\displaystyle\lim_{z\rightarrow z_{k}}\binom{l-1}{l-1}\frac{1}{z}\left[\left(\frac{z-z_{k}}{e^{z}+1}\right)^{l}\right]^{(l-1)}=\lim_{z\rightarrow z_{k}}\frac{1}{z}[c_{-1}^{(l)}(z-z_{k})^{l-1}]^{(l-1)}=-\frac{(l-1)!}{z_{k}}. (9)

As a result of the relations (8)(8) and (9)(9), we can write out DlD_{l}:

Dl=(bl(l)zkl++bj(l)zkj++b1(l)zk),j=1,,l,\displaystyle D_{l}=-(\frac{b_{l}^{(l)}}{z_{k}^{l}}+\cdots+\frac{b_{j}^{(l)}}{z_{k}^{j}}+\cdots+\frac{b_{1}^{(l)}}{z_{k}}),~{}j=1,\cdots,l, (10)

in which bl(l)=b1(l)=1b_{l}^{(l)}=b_{1}^{(l)}=1, b2(l),,bl1(l)b_{2}^{(l)},\cdots,b_{l-1}^{(l)}\in\mathbb{Q}.

Step 3. In this part, we will prove that there exists unique wlw_{l}, which satisfying

l=1mwlDl=Smzkm,Sm.\sum_{l=1}^{m}w_{l}D_{l}=\frac{S_{m}}{z_{k}^{m}},~{}S_{m}\in\mathbb{Q}.

In fact, it should be noted that, by applying (10)(10) in Step 2, we obtain

l=1mwlDl=l=1mwl(1zkl++bj(l)zkj++1zk)=Smzkm.\displaystyle\sum_{l=1}^{m}w_{l}D_{l}=-\sum_{l=1}^{m}w_{l}(\frac{1}{z_{k}^{l}}+\cdots+\frac{b_{j}^{(l)}}{z_{k}^{j}}+\cdots+\frac{1}{z_{k}})=\frac{S_{m}}{z_{k}^{m}}. (11)

Comparing the coefficients on both sides of identity (11)(11) with respect to 1/zkl1/z^{l}_{k}, we have the relation:

(11101b2(m)001)m×m(w1w2wm)=(00Sm).\displaystyle\begin{pmatrix}1&1&\cdots&1\\ 0&1&\cdots&b_{2}^{(m)}\\ \vdots&\vdots&\ddots&\vdots\\ 0&0&\cdots&1\end{pmatrix}_{m\times m}\begin{pmatrix}w_{1}\\ w_{2}\\ \vdots\\ w_{m}\end{pmatrix}=\begin{pmatrix}0\\ 0\\ \vdots\\ -S_{m}\end{pmatrix}. (12)

Let Am×mA_{m\times m} represents the coefficient matrix of (12)(12). Since Am×mA_{m\times m} is an upper-triangular matrix, and Am×m=1,\|A_{m\times m}\|=1, then by Cramer’s Rule, the existence and unique of wlw_{l} can be obtained. Set C(m)=[0,0,,Sm]TC^{(m)}=[0,0,\cdots,-S_{m}]^{T}, W(m)=[w1,w2,,wm]TW^{(m)}=[w_{1},w_{2},\cdots,w_{m}]^{T}, then

wl=[W(m)]l,l=1,,m,\displaystyle w_{l}=[W^{(m)}]_{l},~{}l=1,\cdots,m, (13)

in which W(m)=Am×m1C(m)W^{(m)}=A_{m\times m}^{-1}C^{(m)} and []l[\cdot]_{l} represents the ll-th element in the column vector.

Given that zk=(2k1)πiz_{k}=(2k-1)\pi{\rm i}, and substitute wlw_{l} and DlD_{l} into (6)(6), we obtain

l=1mγnwlz(ez+1)ldz=2πik=1nSmzkm=2Sm(iπ)m1[ζ(m)(112m)k=n+11(2k1)m],\displaystyle\sum_{l=1}^{m}\int_{\gamma_{n}}\frac{w_{l}}{z(e^{z}+1)^{l}}{\rm d}z=2\pi{\rm i}\sum_{k=1}^{n}\frac{S_{m}}{z_{k}^{m}}=\frac{2S_{m}}{({\rm i}\pi)^{m-1}}\Big{[}\zeta(m)(1-\frac{1}{2^{m}})-\sum_{k=n+1}^{\infty}\frac{1}{(2k-1)^{m}}\Big{]},

which gives the required equivalence. Lemma 3.1 is thus proved. \hfill\square

Corollary 3.1 Let

Sm={(1)m12Γ(m),m=3,5,,(1)m+22Γ(m),m=2,4,,S_{m}=\left\{\begin{aligned} &(-1)^{\frac{m-1}{2}}\Gamma(m),~{}~{}m=3,5,\cdots,\\ &(-1)^{\frac{m+2}{2}}\Gamma(m),~{}~{}m=2,4,\cdots,\end{aligned}\right.

where Γ()\Gamma(\cdot) denotes the Gamma function and Γ(m)=(m1)!\Gamma(m)=(m-1)!. For n=1,2,n=1,2,\cdots, by using Lemma 3.1, there exists unique wlw_{l}\in\mathbb{Q}, l=1,2,,ml=1,2,\cdots,m, such that

l=1mγnwlz(ez+1)ldz=2Γ(m)πm1iκ[ζ(m)(112m)k=n+11(2k1)m],\displaystyle\sum_{l=1}^{m}\int_{\gamma_{n}}\frac{w_{l}}{z(e^{z}+1)^{l}}{\rm d}z=\frac{2\Gamma(m)}{\pi^{m-1}{\rm i}^{\kappa}}\Big{[}\zeta(m)(1-\frac{1}{2^{m}})-\sum_{k=n+1}^{\infty}\frac{1}{(2k-1)^{m}}\Big{]}, (14)

in which κ=0\kappa=0 or 11 with respect to mm is positive odd or even integer.

3.2 A formula for ζ(m),m3\zeta(m),m\geq 3 at positive odd integers

The generalized Bernoulli polynomials[4] Bn(l)(x)B_{n}^{(l)}(x) of degree nn in xx are defined by the generating function:

(zez1)lexz=n=0Bn(l)(x)znn!,|z|<2π,1l:=1,\displaystyle\left(\frac{z}{e^{z}-1}\right)^{l}e^{xz}=\sum_{n=0}^{\infty}B_{n}^{(l)}(x)\frac{z^{n}}{n!},~{}|z|<2\pi,1^{l}:=1, (15)

for arbitrary (real or complex) parameter ll and Bn(l)(0)=Bn(l)B_{n}^{(l)}(0)=B_{n}^{(l)}. Let m3m\geq 3 be a positive odd integer. wlw_{l} are solutions to (12)(12) and zk=(2k1)πiz_{k}=(2k-1)\pi{\rm i}, k1k\geq 1. Set x=zzkx=z-z_{k} and 0<|x|<2π0<|x|<2\pi. Using the property (22) of Bn(l)(x)B_{n}^{(l)}(x) in [4], we obtain

1(ez+1)l\displaystyle\frac{1}{(e^{z}+1)^{l}} =(1)l(ex1)l=n=0(1)lBn(l)xnln!,\displaystyle=\frac{(-1)^{l}}{(e^{x}-1)^{l}}=\sum_{n=0}^{\infty}(-1)^{l}B_{n}^{(l)}\frac{x^{n-l}}{n!}, (16)
1(ez+1)l\displaystyle\frac{1}{(e^{-z}+1)^{l}} =exl(ex1)l=n=0Bn(l)(l)xnln!=n=0(1)nBn(l)xnln!.\displaystyle=\frac{e^{xl}}{(e^{x}-1)^{l}}=\sum_{n=0}^{\infty}B_{n}^{(l)}(l)\frac{x^{n-l}}{n!}=\sum_{n=0}^{\infty}(-1)^{n}B_{n}^{(l)}\frac{x^{n-l}}{n!}. (17)

Since the Laurent series of (ez+1)l(e^{z}+1)^{-l} centered at z=zkz=z_{k} have the form

1(ez+1)l=j=lcj(l)(zzk)j,l=1,2,,m,\displaystyle\frac{1}{(e^{z}+1)^{l}}=\sum_{j=-l}^{\infty}c_{j}^{(l)}(z-z_{k})^{j},~{}l=1,2,\cdots,m, (18)

where

cj(l)=12πiΛkdz(ez+1)l(zzk)j+1,j=l,(l1),.\displaystyle c_{j}^{(l)}=\frac{1}{2\pi i}\int_{\Lambda_{k}}\frac{{\rm d}z}{(e^{z}+1)^{l}(z-z_{k})^{j+1}},~{}j=-l,-(l-1),\cdots.

Comparing the coefficients of 1/(zzk)l1/(z-z_{k})^{l} in (16)(16) and (18)(18), we obtain

cj(l)=(1)lBl+j(l)Γ(l+j+1),j=l,,1.\displaystyle c_{j}^{(l)}=\frac{(-1)^{l}B_{l+j}^{(l)}}{\Gamma(l+j+1)},~{}j=-l,\cdots,-1. (19)

Taking (7)(7) and (18)(18) into account, we obtain

Dl=j=0l1(1)lj1(l1j)Γ(lj)Γ(j+1)c(lj)(l)Γ(l)zklj=j=1l(1)j+1cj(l)zkj.\displaystyle D_{l}=\sum_{j=0}^{l-1}(-1)^{l-j-1}\binom{l-1}{j}\frac{\Gamma(l-j)\Gamma(j+1)c_{-(l-j)}^{(l)}}{\Gamma(l)z_{k}^{l-j}}=\sum_{j=1}^{l}(-1)^{j+1}\frac{c_{-j}^{(l)}}{z_{k}^{j}}.

Hence, it follows from (10)(10) that

bj(l)=(1)jcj(l).\displaystyle b_{j}^{(l)}=(-1)^{j}c_{-j}^{(l)}. (20)

Combining with (19)(19), we deduce that

bj(l)=(1)ljΓ(lj+1)Blj(l),j=1,2,,l.\displaystyle b_{j}^{(l)}=\frac{(-1)^{l-j}}{\Gamma(l-j+1)}B_{l-j}^{(l)},~{}j=1,2,\cdots,l. (21)

With the help of (17)(17) and (12)(12), we obtain

12πil=1mγnwlz(ez+1)ldz=12πik=1nj=1ml=jmΛkwl(1)ljBlj(l)Γ(lj+1)1z(zzk)jdz=k=1nj=1ml=jmwlbj(l)(1)j1zkj=k=1n(1)m1wmbm(m)zkm=k=1nSmzkm.\displaystyle\begin{aligned} \frac{1}{2\pi{\rm i}}\sum_{l=1}^{m}\int_{\gamma_{n}}\frac{w_{l}}{z(e^{-z}+1)^{l}}{\rm d}z&=\frac{1}{2\pi{\rm i}}\sum_{k=1}^{n}\sum_{j=1}^{m}\sum_{l=j}^{m}\int_{\Lambda_{k}}w_{l}(-1)^{l-j}\frac{B_{l-j}^{(l)}}{\Gamma(l-j+1)}\frac{1}{z(z-z_{k})^{j}}{\rm d}z\\ &=\sum_{k=1}^{n}\sum_{j=1}^{m}\sum_{l=j}^{m}w_{l}b_{j}^{(l)}\frac{(-1)^{j-1}}{z_{k}^{j}}=\sum_{k=1}^{n}\frac{(-1)^{m-1}w_{m}b_{m}^{(m)}}{z_{k}^{m}}\\ &=-\sum_{k=1}^{n}\frac{S_{m}}{z_{k}^{m}}.\end{aligned}

Hence, we have the relation

l=1mγnwlz(ez+1)ldz\displaystyle\sum_{l=1}^{m}\int_{\gamma_{n}}\frac{w_{l}}{z(e^{z}+1)^{l}}{\rm d}z =12l=1mγnwlz[1(ez+1)l1(ez+1)l]dz\displaystyle=\frac{1}{2}\sum_{l=1}^{m}\int_{\gamma_{n}}\frac{w_{l}}{z}\left[\frac{1}{(e^{z}+1)^{l}}-\frac{1}{(e^{-z}+1)^{l}}\right]{\rm d}z
=12l=1mp=0l1γn(1ez)wlezpz(ez+1)ldz.\displaystyle=\frac{1}{2}\sum_{l=1}^{m}\sum_{p=0}^{l-1}\int_{\gamma_{n}}\frac{(1-e^{z})w_{l}e^{zp}}{z(e^{z}+1)^{l}}{\rm d}z. (22)

Since z=0z=0 is a removable singularity pole for the integrand in (22)(22), let Rn+R_{n}\rightarrow+\infty and ε0\varepsilon\rightarrow 0 in γn\gamma_{n}, then we have

2Γ(m)(112m)ζ(m)πm1=l=1m+wlu(eu+1)ldu=12l=1mp=0l1+wl(1eu)eupu(eu+1)ldu.\displaystyle 2\Gamma(m)(1-\frac{1}{2^{m}})\frac{\zeta(m)}{\pi^{m-1}}=\sum_{l=1}^{m}\int_{-\infty}^{+\infty}\frac{w_{l}}{u(e^{u}+1)^{l}}{\rm d}u=\frac{1}{2}\sum_{l=1}^{m}\sum_{p=0}^{l-1}\int_{-\infty}^{+\infty}\frac{w_{l}(1-e^{u})e^{up}}{u(e^{u}+1)^{l}}{\rm d}u.

Note that the integrand l=1mp=0l1wl(1eu)eupu(eu+1)l\sum\limits_{l=1}^{m}\sum\limits_{p=0}^{l-1}\frac{w_{l}(1-e^{u})e^{up}}{u(e^{u}+1)^{l}} is an even function on \mathbb{R} except a removable singularity at origin, then we have

ζ(m)=(2π)m1(2m1)Γ(m)l=1mp=0l10+wl(1eu)eupu(eu+1)ldu,m=3,5,,\displaystyle\zeta(m)=\frac{(2\pi)^{m-1}}{(2^{m}-1)\Gamma(m)}\sum_{l=1}^{m}\sum_{p=0}^{l-1}\int_{0}^{+\infty}\frac{w_{l}(1-e^{u})e^{up}}{u(e^{u}+1)^{l}}{\rm d}u,~{}~{}m=3,5,\cdots, (23)

in which wlw_{l} are determined by (13)(13).

4. Evaluation of ζ(5)\zeta(5) and ζ(7)\zeta(7) by using the generalized Bernoulli numbers

In this part, we calculate the numerical results of ζ(5)\zeta(5) and ζ(7)\zeta(7). In view of the Srivastava-Todorov formula [8, p.510, Eq.(3)], we know the following representation for the generalized Bernoulli numbers:

Bn(l)=k=0n(l+nnk)(l+k1k)n!(n+k)!j=0k(1)j(kj)jn+k.\displaystyle B_{n}^{(l)}=\sum_{k=0}^{n}\binom{l+n}{n-k}\binom{l+k-1}{k}\frac{n!}{(n+k)!}\sum_{j=0}^{k}(-1)^{j}\binom{k}{j}j^{n+k}. (24)

Using the formula (21)(21) and (24)(24), it is verified by a straightforward calculation that

D1=1zk,D2=1zk21zk,D3=12!(2zk3+3zk2+2zk),\displaystyle D_{1}=-\frac{1}{z_{k}},~{}D_{2}=-\frac{1}{z_{k}^{2}}-\frac{1}{z_{k}},~{}D_{3}=-\frac{1}{2!}(\frac{2}{z_{k}^{3}}+\frac{3}{z_{k}^{2}}+\frac{2}{z_{k}}),
D4=13!(3!zk4+12zk3+11zk2+3!zk),D5=14!(4!zk5+60zk4+70zk3+50zk2+4!zk),\displaystyle D_{4}=-\frac{1}{3!}(\frac{3!}{z_{k}^{4}}+\frac{12}{z_{k}^{3}}+\frac{11}{z^{2}_{k}}+\frac{3!}{z_{k}}),~{}D_{5}=-\frac{1}{4!}(\frac{4!}{z^{5}_{k}}+\frac{60}{z_{k}^{4}}+\frac{70}{z_{k}^{3}}+\frac{50}{z_{k}^{2}}+\frac{4!}{z_{k}}),
D6=15!(5!zk6+360zk5+510zk4+450zk3+274zk2+5!zk),\displaystyle D_{6}=-\frac{1}{5!}(\frac{5!}{z_{k}^{6}}+\frac{360}{z_{k}^{5}}+\frac{510}{z_{k}^{4}}+\frac{450}{z_{k}^{3}}+\frac{274}{z_{k}^{2}}+\frac{5!}{z_{k}}),
D7=16!(6!zk7+2520zk6+4200zk5+4410zk4+3248zk3+1764zk2+6!zk),\displaystyle D_{7}=-\frac{1}{6!}(\frac{6!}{z_{k}^{7}}+\frac{2520}{z_{k}^{6}}+\frac{4200}{z_{k}^{5}}+\frac{4410}{z_{k}^{4}}+\frac{3248}{z_{k}^{3}}+\frac{1764}{z_{k}^{2}}+\frac{6!}{z_{k}}),

where zk=(2k1)πi,1knz_{k}=(2k-1)\pi{\rm i},1\leq k\leq n. According with SmS_{m} in Corollary 3.13.1 and solving linear equation (12)(12), we obtain wlw_{l} and the following result.
Theorem 4.1   (1): For Riemann zeta function value ζ(5)\zeta(5)

ζ(5)=(2π)4(251)Γ(5)l=15wl(eu+1)ludz,\displaystyle\zeta(5)=\frac{(2\pi)^{4}}{(2^{5}-1)\Gamma(5)}\sum_{l=1}^{5}\int_{-\infty}^{\infty}\frac{w_{l}}{(e^{u}+1)^{l}u}{\rm d}z,

where w1=1,w2=15,w3=50,w4=60,w5=24w_{1}=-1,w_{2}=15,w_{3}=-50,w_{4}=60,w_{5}=-24.

(2): For Riemann zeta function value ζ(7)\zeta(7)

ζ(7)=(2π)6(271)Γ(7)l=17wl(eu+1)ludz,\displaystyle\zeta(7)=\frac{(2\pi)^{6}}{(2^{7}-1)\Gamma(7)}\sum_{l=1}^{7}\int_{-\infty}^{\infty}\frac{w_{l}}{(e^{u}+1)^{l}u}{\rm d}z,

where w1=1,w2=63,w3=602,w4=2100,w5=3360,w6=2520,w7=720w_{1}=1,w_{2}=-63,w_{3}=602,w_{4}=-2100,w_{5}=3360,w_{6}=-2520,w_{7}=720.

5. The linear combination of ζ(2n+1)/π2n,n1\zeta(2n+1)/\pi^{2n},n\geq 1

In this part, we investigate the integral representation for the linear combination of ζ(2n+1)/π2n\zeta(2n+1)/\pi^{2n}. For all positive odd integer m3m\geq 3, it follows from (23)(23) that

2Γ(m)(112m)ζ(m)πm1=l=1mp=0l10+wle(2p+1l)uu(eu+eu)ld(eu+eu).\displaystyle 2\Gamma(m)(1-\frac{1}{2^{m}})\frac{\zeta(m)}{\pi^{m-1}}=-\sum_{l=1}^{m}\sum_{p=0}^{l-1}\int_{0}^{+\infty}\frac{w_{l}e^{(2p+1-l)u}}{u(e^{u}+e^{-u})^{l}}{\rm d}(e^{u}+e^{-u}). (25)

Considering the symmetry of the power of exp(2p+1l)\exp(2p+1-l) with pp from 0 to l1l-1, we can draw the following conclusions:

p=0l1e(2p+1l)u(eu+eu)l=j=1kqj(l)(eu+eu)2j1,k=l/2,qj(l),\displaystyle\sum_{p=0}^{l-1}\frac{e^{(2p+1-l)u}}{(e^{u}+e^{-u})^{l}}=\sum_{j=1}^{k}\frac{q_{j}^{(l)}}{(e^{u}+e^{-u})^{2j-1}},~{}~{}k=\lceil{l/2}\rceil,q_{j}^{(l)}\in\mathbb{Z}, (26)

in which \lceil{\cdot}\rceil means the ceiling function and qj(l)=1κ=1j1(l+12κjκ)qκ(l),q_{j}^{(l)}=1-\sum\limits_{\kappa=1}^{j-1}\binom{l+1-2\kappa}{j-\kappa}q_{\kappa}^{(l)}, for j2j\geq 2, q1(l)1q_{1}^{(l)}\equiv 1. Remembering that l=1mwl=0\sum\limits_{l=1}^{m}w_{l}=0 in Lemma 3.1, then we obtain

2Γ(m)(112m)ζ(m)πm1\displaystyle 2\Gamma(m)(1-\frac{1}{2^{m}})\frac{\zeta(m)}{\pi^{m-1}} =0j=1m/2[l=2j1mwlqj(l)]d(eu+eu)u(eu+eu)2j1\displaystyle=-\int_{0}^{\infty}\sum_{j=1}^{\lceil{m/2}\rceil}[\sum_{l=2j-1}^{m}w_{l}q_{j}^{(l)}]\frac{{\rm d}(e^{u}+e^{-u})}{u(e^{u}+e^{-u})^{2j-1}}
=0j=2m/2[l=2j1mwlqj(l)]22j+1(j1)ud(sech(u))2j2\displaystyle=\int_{0}^{\infty}\sum_{j=2}^{\lceil{m/2}\rceil}[\sum_{l=2j-1}^{m}w_{l}q_{j}^{(l)}]\frac{2^{-2j+1}}{(j-1)u}{\rm d}({\rm sech}{(u)})^{2j-2}
=01j=2m/214j1[l=2j1mwlqj(l)]u2j3asech(u)du.\displaystyle=-\int_{0}^{1}\sum_{j=2}^{\lceil{m/2}\rceil}\frac{1}{4^{j-1}}[\sum_{l=2j-1}^{m}w_{l}q_{j}^{(l)}]\frac{u^{2j-3}}{{\rm~{}asech}{(u)}}{\rm d}u. (27)

Let us denote

τj(m)=2m1Γ(m)(2m1)l=2j1mwlqj(l)4j1,\displaystyle\tau_{j}^{(m)}=-\frac{2^{m-1}}{\Gamma(m)(2^{m}-1)}\sum\limits_{l=2j-1}^{m}\frac{w_{l}q_{j}^{(l)}}{4^{j-1}}, (28)

then

ζ(m)πm1=j=2m/2τj(m)01u2j3asech(u)du.\displaystyle\frac{\zeta(m)}{\pi^{m-1}}=\sum_{j=2}^{\lceil{m/2}\rceil}\tau_{j}^{(m)}\int_{0}^{1}\frac{u^{2j-3}}{{\rm~{}asech}{(u)}}{\rm d}u. (29)

Through above discussion, we have the following theorem.

Theorem 5.1 For any positive integer nn, then there exists θk\theta_{k}\in\mathbb{Q}, k=1,2,nk=1,2\cdots,n, so that

k=1nθkζ(2k+1)π2k=θn+1In,θn+1=0or1,\sum_{k=1}^{n}\theta_{k}\frac{\zeta(2k+1)}{\pi^{2k}}=\theta_{n+1}I_{n},~{}\theta_{n+1}=0~{}or~{}1,

where InI_{n} satisfy

In=01u2n1asech(u)𝑑u>0,limnIn=0.I_{n}=\int_{0}^{1}\frac{u^{2n-1}}{{\rm asech}(u)}du>0,~{}~{}\lim_{n\rightarrow\infty}I_{n}=0.

Proof. Since asech(u)=log(1/u+1/u21){\rm asech(u)}=\log(1/u+\sqrt{1/u^{2}-1}), then for ϑ\vartheta near to 0 from right side

1asech(u)=1logulog2log2u+o(1/log2u),0<u<ϑ<1,\displaystyle\frac{1}{{\rm asech(u)}}=-\frac{1}{\log u}-\frac{\log 2}{\log^{2}u}+o(1/\log^{2}u),~{}0<u<\vartheta<1, (30)

where o()o(\cdot) means higher-order infinitesimal. Hence, we obtain 1/asech(u)01/{\rm asech}(u)\rightarrow 0 as u0+u\rightarrow 0^{+}. And for η>0\eta>0 near to 11 from left side, when u(η,1)u\in(\eta,1),

1asech(u)=221/u1+2241/u1+o(1/u1),\displaystyle\frac{1}{{\rm asech(u)}}=\frac{\sqrt{2}}{2\sqrt{1/u-1}}+\frac{\sqrt{2}}{24}\sqrt{1/u-1}+o(\sqrt{1/u-1}), (31)

then asech(u)/1/u12{\rm asech(u)}/\sqrt{1/u-1}\rightarrow\sqrt{2} as u1u\rightarrow 1^{-}. Since 01(1/u1)1/2𝑑u=π/2\int_{0}^{1}(1/u-1)^{-1/2}du=\pi/2, so we can deduce that 011/asech(u)du\int_{0}^{1}1/{\rm asech(u)}{\rm du} converges. According with asech(u)>0{\rm asech(u)}>0, when u(0,1)u\in(0,1), and 0<In011/asech(u)du0<I_{n}\leq\int_{0}^{1}1/{\rm asech(u)}{\rm du}, then {In}n1\{I_{n}\}_{n\geq 1} is a monotonically decreasing positive sequence with a lower bound, so it must converge.

From (31)(31), when u(η,1)u\in(\eta,1), we can find proper M>0M>0, so that 1/asech(u)M1/u1{\rm 1/asech(u)}\leq\frac{M}{\sqrt{1/u-1}}. As a result, we have

In\displaystyle I_{n} =01u2n1asech(u)𝑑u,\displaystyle=\int_{0}^{1}\frac{u^{2n-1}}{{\rm asech}(u)}du,
=0ηu2n1asech(u)𝑑u+η1u2n1asech(u)𝑑u\displaystyle=\int_{0}^{\eta}\frac{u^{2n-1}}{{\rm asech}(u)}du+\int_{\eta}^{1}\frac{u^{2n-1}}{{\rm asech}(u)}du
η2n1011asech(u)𝑑u+M01u2n1(1/u1)1/2𝑑u\displaystyle\leq\eta^{2n-1}\int_{0}^{1}\frac{1}{{\rm asech}(u)}du+M\int_{0}^{1}u^{2n-1}(1/u-1)^{-1/2}du
η2n1011asech(u)𝑑u+MπΓ(2n+1/2)Γ(2n+1),\displaystyle\leq\eta^{2n-1}\int_{0}^{1}\frac{1}{{\rm asech}(u)}du+M\sqrt{\pi}\frac{\Gamma(2n+1/2)}{\Gamma(2n+1)},

since η(0,1)\eta\in(0,1) and using Stirling’s approximation: Γ(2n+1/2)Γ(2n+1)=O(1/2n+1)\frac{\Gamma(2n+1/2)}{\Gamma(2n+1)}=O(1/\sqrt{2n+1}) as nn\rightarrow\infty, so limnIn=0\lim\limits_{n\rightarrow\infty}I_{n}=0.

Using (29)(29), let m=2n+1m=2n+1, if for n1\forall n\geq 1, then τn+1(2n+1)0\tau_{n+1}^{(2n+1)}\neq 0, we can choose proper θk\theta_{k}, 1kn1\leq k\leq n so that θn+1=1\theta_{n+1}=1. If there exists n2n\geq 2 such that τn+1(2n+1)=0\tau_{n+1}^{(2n+1)}=0, then we can choose proper θk\theta_{k} (where θk\theta_{k} satisfying k=1nθk20\sum_{k=1}^{n}\theta_{k}^{2}\neq 0) making θn+1=0\theta_{n+1}=0. Therefore, there must exist proper θk\theta_{k}\in\mathbb{Q} which make k=1nθkζ(2k+1)π2k=θn+1In\sum\limits_{k=1}^{n}\theta_{k}\frac{\zeta(2k+1)}{\pi^{2k}}=\theta_{n+1}I_{n}, θn+1=0\theta_{n+1}=0 or 11. Thus Theorem 5.1 is proved. \hfill\square

Since the problem of the irrationality of the values of the Riemann zeta function at odd integers is one of the most attractive topics. Before Apery’s results[10], the arithmetic properties of the zeta function at odd points seemed inaccessible. According with Theorem 5.15.1 in this paper and Nesterenko’s theorem[9], we derive criteria for the dimension of the vector space spanned over the rational by ζ(2n+1)/π2n\zeta(2n+1)/\pi^{2n}, n1n\geq 1.

Theorem 5.2 Let m=2n+1m=2n+1 in (29)(29), if τn+1(2n+1)0\tau_{n+1}^{(2n+1)}\neq 0 in (28)(28) for n1n\geq 1, then

dim(ζ(3)π2+ζ(5)π4++ζ(2n+1)π2n+)=;\displaystyle\dim_{\mathbb{Q}}(\mathbb{Q}\frac{\zeta(3)}{\pi^{2}}+\mathbb{Q}\frac{\zeta(5)}{\pi^{4}}+\cdots+\mathbb{Q}\frac{\zeta(2n+1)}{\pi^{2n}}+\cdots)=\infty; (32)

If there exits n2n\geq 2, so that τn+1(2n+1)=0\tau_{n+1}^{(2n+1)}=0, then

dim(ζ(3)π2+ζ(5)π4++ζ(2n+1)π2n)n1.\displaystyle\dim_{\mathbb{Q}}(\mathbb{Q}\frac{\zeta(3)}{\pi^{2}}+\mathbb{Q}\frac{\zeta(5)}{\pi^{4}}+\cdots+\mathbb{Q}\frac{\zeta(2n+1)}{\pi^{2n}})\leq n-1. (33)

6. Concluding remarks

Through numerical calculations of ζ(5)\zeta(5), ζ(7)\zeta(7) and ζ(9)\zeta(9), we do believe that only conclusion (32) is possible, which means the sequence ζ(3)/π2,ζ(5)/π4,ζ(7)/π6\zeta(3)/\pi^{2},\zeta(5)/\pi^{4},\zeta(7)/\pi^{6},\cdots contains infinitely many irrational numbers. But it’s a pity that we can’t prove it theoretically. This is also one of our future research contents.

Conflict of interest

The author declares that there is no conflict of interest with respect to the publication of this paper.

Funding

Work funded by High-level Talent Research Project Fund of NHU (Grant Nos. 2022rcjj24).

References

  • [1] Z. Silagadze, Sums of generalized harmonic series. Reson 20, 822-843 (2015).
  • [2] D. Cvijovic´\acute{c}, J. Klinowski, Integral representations of the Riemann zeta function for odd-integer arguments. J. Comput. Appl. Math. 142, p.435-439 (2002).
  • [3] H. M. Srivastava, J. Choi, Zeta and Q-Zeta functions and associated series and integrals; Elsevier Inc., New York, NY,USA, 2012.
  • [4] J. D{\rm D}^{\prime}Avanzo, N. A. Krylov, ζ(n)\zeta(n) via hyperbolic functions, Involve, Vol.3, pp.289-296 (2010).
  • [5] F. M. S. Lima, New definite integrals and a two-term dilogarithm identity, Indag. Math., Vol.23, pp.1-9 (2012).
  • [6] F. Beukers, J.A.C. Kolk, E. Calabi, Sums of generalized harmonic series and volumes, Nieuw Archief voor Wiskunde, fourth series, Vol.11, pp.217-224 (1993).
  • [7] L. Lewin, Polylogarithms and associated functions, Elsevier (North-Holland), New York, London and Amsterdam, 1981.
  • [8] H. M. Srivastava, P. G. Todorov, An explicit formula for the generalized Bernoulli polynomials, J. Math. Anal. Appl. 130, pp.509-513 (1988).
  • [9] Yu. V. Nesterenko, On the linear independence of numbers, Vestnik Moskov. Univ. Ser. 1 Mat. Mekh, 1985, no.1, 46-54; English transl., Moscow Univ. Math. Bull. 40:1, 69-74 (1985).
  • [10] R. Ape´\acute{e}ry, Irrationalite´\acute{e} de ζ(2)\zeta(2) et ζ(3)\zeta(3), Aste´\acute{e}risque 61, 11-13 (1979).