A note on some constants related to the zeta–function and their relationship with the Gregory coefficients

Iaroslav V. Blagouchine∗

Steklov Institute of Mathematics at St.-Petersburg, Russia.

Marc–Antoine Coppo Université Côte d’Azur, CNRS, LJAD (UMR 7351), France.

Abstract

In this article, new series for the first and second (also known as generalized Eu- ler’s constant), as well as for some closely related constants are obtained. These series contain rational terms only and involve the so–called Gregory coefficients, which are also known as (reciprocal) log- arithmic numbers, Cauchy numbers of the first kind and Bernoulli numbers of the second kind. In addition, two interesting series with rational terms for Euler’s constant γ and the constant ln 2π are given, and yet another generalization of Euler’s constant is proposed and various formulas for the cal- culation of these constants are obtained. Finally, we mention in the paper that almost all the constants considered in this work admit simple representations via the Ramanujan summation.

Keywords: Stieltjes constants, Generalized Euler’s constants, Series expansions, Ramanujan summation, Harmonic product of sequences, Gregory’s coefficients, Logarithmic numbers, Cauchy numbers, Bernoulli numbers of the second kind, Stirling numbers of the first kind, Harmonic numbers.

I. Introduction and definitions

The zeta-function ∞ ζ(s) := ∑ n−s , Re s > 1, n=1 is of fundamental and long-standing importance in analytic number theory, modern analysis, theory of L–functions, prime number theory and in a variety of other fields. The ζ–function is a meromorphic function on the entire complex plane, except at the point s = 1 at which it has one simple pole with residue 1. The coefficients of the regular part of its Laurent series, denoted γm,

1 ∞ (−1)m(s − 1)m ζ(s) = + γ + γ , s 6= 1. (1) − ∑ m s 1 m=1 m!

∗Corresponding author. Email addresses: [email protected] (Iaroslav V. Blagouchine), [email protected] (Marc–Antoine Coppo) 1 where γ is Euler’s constant , and those of the Maclaurin series δm 1 1 ∞ (−1)msm ζ(s) = + + δ , s 6= 1. (2) − ∑ m s 1 2 m=1 m! are of special interest and have been widely studied in the literature, see e.g. [25], [21, vol. I, letter 71 and following], [23, p. 166 et seq.], [29, 30, 24, 26, 16, 9, 2, 34, 28, 35, 19, 4, 33, 38, 18]. The coefficients γm are usually called Stieltjes constants or generalized Euler’s constants (both names being in use), while 2 δm do not possess a special name. It may be shown with the aid of the Euler–MacLaurin summation that γm and δm may be also given by the following asymptotic representations

( n m m+1 ) ln k ln n 3 γm = lim − , m = 1, 2, 3, . . . , (3) n→∞ ∑ + k=1 k m 1 and ( n m k m ) m m m+k ln n ln n 4 δm = (−1) m! + lim ln k − n m! (−1) − , m = 1, 2, 3, . . . , (4) n→∞ ∑ ∑ k=1 k=0 k! 2 These representations may be translated into these simple expressions

R lnm k R γ = , δ = lnm k , m = 1, 2, 3, . . . m ∑ k m ∑ k>1 k>1 R where ∑ stands for the sum of the series in the sense of the Ramanujan summation of divergent 5 −s 1 series . Due to the reflection formula for the zeta-function ζ(1 − s) = 2ζ(s)Γ(s)(2π) cos 2 πs , the numbers δm and γm are related to each other polynomially and also involve Euler’s constant γ and the values of the ζ–function at naturals. For the first values of m, this gives

1 δ1 = 2 ln 2π − 1 = −0.08106146679 . . . 1 2 1 2 1 2 6 δ2 = γ1 + 2 γ − 2 ln 2π − 24 π + 2 = −0.006356455908 . . . 3 3 3 2 1 2 1 3 δ3 = − 2 γ2 − 3γ1γ − γ − 3γ1 + 2 γ − 8 π ln 2π + ζ(3) + 2 ln 2π − 6 = +0.004711166862 . . . and conversely

2 1 1 2 γ1 = δ2 + 2δ1 + 4δ1 − 2 γ + 24 π = −0.07281584548 . . . 2 16 3 2 1 2 1 3 2 4 γ2 = − 3 δ3 − 2δ2(γ + 2) − 4δ1δ2 − 3 δ1 − 4δ1(γ + 4) − 8δ1(γ + 1) − 12 γπ + 3 γ + 3 ζ(3) − 3 = −0.009690363192 . . .

The relationships between the higher–order coefficients become very cumbersome, but may be found via a semi–recursive procedure described in [1]. Although there exist numerous representations for

1 0 We recall that γ = limn→∞(Hn − ln n) = −Γ (1) = 0.5772156649 . . . , where Hn is the nth . 2 m (m) It follows from (2) that δm = (−1) ζ (0) + m! 3This representation is very old and was already known to Adolf Pilz, Stieltjes, Hermite and Jensen [4, p. 366]. 4 A slightly different expression for δm was given earlier by Lehmer [33, Eq. (5), p. 266], Sitaramachandrarao [38, Theorem 1], Finch [23, p. 168 et seq.] and Connon [18, Eqs. (2.15), (2.19)]. 5For more details on the Ramanujan summation, see [10, 11, 13, 14]. Note that some authors use slightly different definitions and notations for the Ramanujan summation, see e.g. [3, Ch. 6] and [27, Ch. 13]. 6 This expression for δ2 was found by Ramanujan, see e.g. [3, (18.2)].

2 γm and δm, no convergent series with rational terms only are known for them (unlike for Euler’s constant γ, see e.g. [4, Sect. 3], or for various expressions containing it [6, p. 413, Eqs. (41), (45)–(47)]). Recently, divergent envelopping series for γm containing rational terms only have been obtained in [4, Eqs. (46)–(47)]. In this paper, by continuing the same line of investigation, we derive convergent series representations with rational coefficients for γ1, δ1, γ2 and δ2, and also find two new series of the same type for Euler’s constant γ and ln 2π respectively. These series are not simple and involve a product of the Gregory coefficients Gn, which are also known as the (reciprocal) logarithmic numbers, the Bernoulli numbers of the second kind bn, or the normalized Cauchy numbers of the first kind C1,n. Similar expressions for higher–order constants γm and δm may be obtained by the same procedure, using the harmonic product of sequences introduced in [12], but are quite cumbersome. Since the Stieltjes constants γm generalize Euler’s constant γ and since our series contain the product of Gn, these new series may also be seen as the generalization of the famous Fontana–Mascheroni series

∞ Gn 1 1 1 19 3 863 275 γ = ∑ = + + + + + + + ... (5) n=1 n 2 24 72 2880 800 362 880 169 344 which is the first known series representation for Euler’s constant having rational terms only, see [6, pp. 406, 413, 429], [4, p. 379]. In Appendix A, we introduce yet another set of constants κm = −m−1 ∑ |Gn| n , which also generalize Euler’s constant γ. These numbers, similarly to γm, coincide n>1 with Euler’s constant at m = 0 and have various interesting series and integral representations, none of them are reducible to classical mathematical constants.

II. Series expansions

II.1. Preliminaries Since the results, that we come to present here, are essentially based on the Gregory coefficients and Stirling numbers, it may be useful to briefly recall their definition and properties. Gregory num- − + − bers, denoted below Gn, are rational alternating G1 = +1/2 , G2 = 1/12 , G3 = 1/24 , G4 = 19/720 , + − G5 = 3/160 , G6 = 863/60 480 ,. . . , decreasing in absolute value, and are also closely related to the −1 theory of finite differences; they behave as n ln2 n at n → ∞ and may be bounded from below and above according to [6, Eqs. (55)–(56)]. They may be defined either via their generating function

z ∞ = 1 + G zn, |z| < 1 , (6) ( + ) ∑ n ln 1 z n=1 or recursively

(−1)n+1 n−1 (−1)n+1−l G 1 G = + l , G = , n = 2, 3, 4, . . . (7) n + ∑ + − 1 n 1 l=1 n 1 l 2 or explicitly 7

1 1 G = x (x − 1)(x − 2) ··· (x − n + 1) dx , n = 1, 2, 3, . . . (8) n n! ˆ 0

7 For more information about Gn, see [6, pp. 410–415], [4, p. 379], [7], and the literature given therein (nearly 50 references).

3 Throughout the paper, we also use the Stirling numbers of the first kind, which we denote below by 8 S1(n, l). Since there are different definitions and notations for them , we specify that in our definition they are simply the coefficients in the expansion of falling factorial n l x (x − 1)(x − 2) ··· (x − n + 1) = ∑ S1(n, l) · x , n = 1, 2, 3, . . . (9) l=1 and may equally be defined via the generating function lnl(1 + z) ∞ S (n, l) ∞ S (n, l) = ∑ 1 zn = ∑ 1 zn , |z| < 1 , l = 0, 1, 2, . . . (10) l! n=l n! n=0 n!

  n±l It is important to note that sgn S1(n, l) = (−1) . The Stirling numbers of the first kind and the Gregory coefficients are linked by the relation 1 n S (n, l) G = 1 , n = 1, 2, 3, . . . (11) n ∑ + n! l=1 l 1 which follows directly from (8) and (9).

II.2. Some auxiliary lemmas

Before we proceed with the series expansions for δm and γm, we need to prove several useful lemmas. Lemma 1. For each natural number k, let ∞ Gn σ := . k ∑ + n=1 n k Then the following equality holds k   1 m k σk = + ∑ (−1) ln(m + 1) , k = 1, 2, 3, . . . (12) k m=1 m Proof. By using (11) and by making use of the generating equation for the Stirling numbers of the first kind (10), we obtain 1 1 ∞ ∞ n l+1 ∞ Gn 1 (−1) S1(n, l) 1 = · xn+k−1dx = − xk−1lnl(1 − x) dx ∑ + ∑ ∑ + ˆ ∑ ( + ) ˆ n=1 n k n=1 n! l=1 l 1 l=1 l 1 ! 0 0 | {z } 1/(n+k) ∞ ∞ ∞ l+1 ∞ l+1 k−1   (−1) k−1 (−1) k − 1 = 1 − e−t tle−t dt = (−1)m e−t(m+1)tl dt ∑ ( + ) ˆ ∑ ( + ) ∑ ˆ l=1 l 1 ! l=1 l 1 ! m=0 m 0 0 | {z } l! (m+1)−l−1 ∞ (−1)l+1 k−1 k − 1 1 k−1 k − 1  1 m + 2  = (−1)m = (−1)m − ln , ∑ + ∑ ( + )l+1 ∑ + + l=1 l 1 m=0 m m 1 m=0 m m 1 m 1 (13)

8More than 60 references on the Stirling numbers of the first kind may be found in [6, Sect. 2.1] and [4, Sect. 1.2]. We also note that our definitions for the Stirling numbers agree with those adopted by MAPLE or MATHEMATICA: our S1(n, l) equals to Stirling1(n,l) from the former and to StirlingS1[n,l] from the latter.

4 where at the last stage we made a change of variable x = 1 − e−t and used the well-known formula for the Γ–function. But since k−1 (−1)m k − 1 1 k − 1  k − 1   k  = , and + = , ∑ + − m=0 m 1 m k m m 1 m the last finite sum in (13) reduces to (12)9.

10 Remark 1. One may show that σk may also be written in terms of the Ramanujan summation:

R Γ(k + 1)Γ(n) R σ = = B(k + 1, n) , (14) k ∑ Γ(n + k + 1) ∑ n>1 n>1 where B stands for the Euler beta-function.

Lemma 2. Let a = (a(1), a(2),..., a(n),... ) be a sequence of complex numbers. The following identity is true for all nonnegative integers n:

n n a(l + 1) 1 n k k (−1)l = (−1)l a(l + 1) . (15) ∑ + + ∑ ∑ l=0 l l 1 n 1 k=0 l=0 l In particular, if a = lnm for any natural m, then this identity reduces to

n n lnm(l + 1) 1 n k k (−1)l = (−1)l lnm(l + 1) . (16) ∑ + + ∑ ∑ l=0 l l 1 n 1 k=1 l=1 l Proof. Formula (15) is an explicit translation of [12, Proposition 7].

Lemma 3. For all natural m ∞ n k   |G + | k γ = n 1 (−1)l lnm(l + 1) , (17) m ∑ + ∑ ∑ n=1 n 1 k=1 l=1 l ∞ n   l n m δm = ∑ |Gn+1| ∑(−1) ln (l + 1) . (18) n=1 l=1 l Proof. Using the representation for the ζ–function ∞ n   1 n − ζ(s) = + G (−1)k (k + 1) s , s 6= 1 , − ∑ n+1 ∑ s 1 n=0 k=0 k see e.g. [4, pp. 382–383], [5], we first have ∞ n   m n ln (l + 1) γ = G (−1)l , m ∑ n+1 ∑ + n=0 l=0 l l 1 and ∞ n   l n m δm = ∑ Gn+1 ∑(−1) ln (l + 1) . n=0 l=0 l Then formula (17) follows from (16).

9The result appeared without proof in [6]. For a slightly more general result, see [20, Proposition 1]. 10See [10] Eq. (4.31).

5 II.3. Series with rational terms for the first Stieltjes constant γ1 and for the coefficient δ1

Theorem 1. The first Stieltjes constant γ1 may be given by the following series

2 ∞ " n−1  # 3 π Gn Gk Gn+ −k Hn − Hk γ = − + + 1 1 ∑ 2 ∑ + − 2 6 n=2 n k=1 n 1 k

3 π2 1 5 1313 42 169 137 969 1 128 119 = − + + + + + + + ... (19) 2 6 32 432 207 360 10 368 000 48 384 000 5 334 336 00 containing π2 and positive rational coefficients only. Using Euler’s formula π2 = 6 ∑ n−2 , the latter may be reduced to a series with rational terms only.

Proof. By (17) with m = 1, one has

∞ n k   Gn+1 k γ = (−1)m ln(m + 1) . 1 ∑ + ∑ ∑ n=1 n 1 k=1 m=1 m Using (12), we find that

n k   n 1 k 1 H + 1 (−1)m ln(m + 1) = σ − n 1 + . + ∑ ∑ + ∑ k + ( + )2 n 1 k=1 m=1 m n 1 k=1 n 1 n 1 Thus ∞ n ∞ ∞ Gn+1 Gn+1 Hn+1 Gn+1 γ = σ − + 1 ∑ + ∑ k ∑ + ∑ ( + )2 n=1 n 1 k=1 n=0 n 1 n=0 n 1

∞ ∞ ∞ Gn+1 Gm (Hm+n − Hm) Gn = − ζ(2) + 1 + ∑ ∑ + ∑ 2 n=1 m=1 n 1 n=1 n since ∞ Gn Hn ∑ = ζ(2) − 1 . n=1 n

For the latter, see e.g. [42, p. 2952, Eq. (1.3)], [15, p. 20,Eq. (3.6)], [11, p. 307, Eq. for F0(2)], [6, p. 413, Eq. (44)]. Rearranging the double absolutely convergent series as follows

∞ ∞  ∞ n−1  Gn+1 Gm Hm+n − Hm Gk Gn+1−k Hn − Hk = , ∑ ∑ + ∑ ∑ + − n=1 m=1 n 1 n=2 k=1 n 1 k we finally arrive at (19).

−2 11 Remark 2. It seems that the sum κ1 := ∑ |Gn| n = 0.5290529699 . . . cannot be reduced to the “stan- dard” mathematical constants. However, it admits several interesting representations, which we give in Ap- pendix A.

1 Theorem 2. The first MacLaurin coefficient δ1 = 2 ln 2π − 1 admits a series representation similar to that for γ1, namely n ∞ 1 1 − ln 2π δ1 = ∑ ∑ Gk Gn+1−k + . (20) n=1 n k=1 2

11For more digits, see OEIS A270859.

6 Proof. Proceeding analogously to the previous case and recalling that

∞ |G | γ + 1 − ln 2π n = − , ∑ − n=2 n 1 2 see e.g. [6, p. 413, Eq. (41)], [43, Corollary 9], we have

∞ n   l n δ1 = ∑ Gn+1 ∑(−1) ln(l + 1) n=0 l=0 l

∞   ∞ ∞ 1 Gn+1 = ∑ Gn+1 σn − = ∑ Gn+1 σn − ∑ n=1 n n=1 n=1 n

∞ ∞ ∞ ∞ Gn+1 Gk γ + 1 − ln 2π Gn Gk 1 − ln 2π = + = + (21) ∑ ∑ + ∑ ∑ + − n=1 k=1 n k 2 n=1 k=1 n k 1 2

n ∞ 1 1 − ln 2π = ∑ ∑ Gk Gn+1−k + , n=1 n k=1 2 where in (21) we eliminated γ by using the fact that G1 = 1/2 and that the sum of |Gn|/n over all natural n equals precisely Euler’s constant, see (5).

Corollary 1. The constant ln 2π has the following beautiful series representation with rational terms only and containing a product of Gregory coefficients

n 3 ∞ 1 3 1 1 7 1 43 79 ln 2π = + ∑ ∑ Gk Gn+1−k = + + + + + + + . . . (22) 2 n=1 n k=1 2 4 24 432 120 8640 24 192 This result directly follows from (20). It is worth noting that one can also readily derive a series with rational coefficients only for ln π (for instance, with the aids of the Mercator series).

Corollary 2. Euler’s constant γ admits the following series representation with rational terms

n ∞ 1 1 1 29 γ = 2 ln 2π − 3 − 2 G G = 2 ln 2π − 3 − − − − ∑ + ∑ k n+2−k n=1 n 1 k=1 24 54 2880 67 1507 3121 12 703 164 551 − − − − − − ... (23) 10 800 362 880 1 058 400 15 806 080 97 977 600 which follows from (21), and the transformation

∞ ∞ ∞ n Gk+1 Gn 1 = G G . ∑ ∑ + ∑ + ∑ k n+2−k k=1 n=1 n k n=1 n 1 k=1

II.4. Generalizations to the second–order coefficients δ2 and γ2 via an application of the harmonic product We recall the main properties of the harmonic product of sequences which are stated and proved ∗ in [12]. If a = (a(1), a(2),... ) and b = (b(1), b(2),... ) are two sequences in CN , then the harmonic product a on b admits the explicit expression:    − m k (a on b)(m + 1) = (−1)k l a(k + 1)b(m + 1 − l) , m = 0, 1, 2, . . . (24) ∑ k l 06l6k6m

7 For small values of m, this gives:

(a on b)(1) = a(1)b(1) , (a on b)(2) = a(2)b(1) + a(1)b(2) − a(2)b(2) , (a on b)(3) = a(3)b(1) + a(1)b(3) + 2a(2)b(2) − 2a(3)b(2) − 2a(2)b(3) + a(3)b(3) .

The harmonic product on is associative and commutative. Let D be the operator (known as the binomial transform) defined by m m D(a)(m + 1) = ∑(−1)j a(j + 1) , m = 0, 1, 2, . . . j=0 j then D = D−1 and the harmonic product satisfies the following property:

D(ab) = D(a) on D(b) . (25) In particular, if a = ln, then D(a)(1) = ln 1 = 0. By (12), we find that m   j m 1 D(ln)(m + 1) = ∑(−1) ln(j + 1) = σm − , m = 1, 2, 3, . . . (26) j=1 j m

Therefore, if a = ln2 then, by (24), (25), and (26), the following identity holds mk  1  1  D(ln2)(m + 1) = (−1)k−l σ − σ − . (27) ∑ k l k k m−l m − l 06l6k6m k6=0 l6=m From this identity, we deduce the following theorem:

Theorem 3. The second coefficients γ2 and δ2 may be given by the following series ∞       |Gn+1| k−l m k 1 1 γ = (−1) σ − σ − − , 2 ∑ n + 1 ∑ k l k k m l m − l n=1 06l6k6m6n k6=0 l6=m and ∞       k−l m k 1 1 δ = |G + | (−1) σ − σ − . 2 ∑ m 1 ∑ k l k k m−l m − l m=1 06l6k6m k6=0 l6=m Proof. Using (17) and equation (27), we get the following equalities:

∞ n m   |G + | m γ = n 1 (−1)j ln2(j + 1) 2 ∑ + ∑ ∑ n=1 n 1 m=1 j=1 j

∞ n |G + | = n 1 D(ln2)(m + 1) ∑ + ∑ n=1 n 1 m=1 ∞ n       |Gn+1| k−l m k 1 1 = (−1) σ − σ − − . ∑ n + 1 ∑ ∑ k l k k m l m − l n=1 m=1 06l6k6m k6=0 l6=m

8 Similarly δ2 is ∞ n   j n 2 δ2 = ∑ |Gn+1| ∑(−1) ln (j + 1) n=0 j=0 j ∞ 2 = ∑ |Gn+1|D(ln )(n + 1) n=1 ∞       k−l m k 1 1 = |G + | (−1) σ − σ − . ∑ m 1 ∑ k l k k m−l m − l m=1 06l6k6m k6=0 l6=m

By following the same method, one may also obtain expressions for higher–order constants γm and δm. However, the resulting expressions are more theoretical than practical.

Appendix A. Yet another generalization of Euler’s constant

−p−1 The numbers κp := ∑ |Gn| n , where the summation extends over n = [1, ∞), may also be regarded as one of the possible generalizations of Euler’s constant (since κ0 = γ0 = γ and 12,13 κ−1 = γ−1 = 1). These constants, which do not seem to be reducible to the classical mathematical constants, admit several interesting representations as stated in the following proposition.

−p−1 Proposition 1. Generalized Euler’s constants κp := ∑ |Gn| n , where the summation extends over n = [1, ∞), admit the following representations:

1 p   (−1) 1 1 p κp = + ln x dx , Re p > −1 , (28) Γ(p + 1) ˆ ln(1 − x) x 0

1 1( p ! p ) dx1 ··· dxp = ··· li 1 − x + γ + ln x , p = 1, 2, 3, . . . (29) ˆ ˆ ∏ k ∑ k ··· k=1 k=1 x1 xp 0 0 | {z } p-fold

∞ k ∞ (−1) S1(n, p + 1) = p = ∑ ∑ k−1 , 0, 1, 2, . . . (30) k=2 k n=p+1 n! n

∞ k ∞ (1) (2) p−1 (p) (−1) Pp(Hn , −Hn ,..., (−1) Hn ) = , p = 0, 1, 2, . . . (31) ∑ ∑ ( + )k k=2 k n=p n 1

R R (1) (2) (p) 1 1 Pp Hn , Hn ,..., Hn = ∑ ∑ = ∑ , p = 1, 2, 3, . . . (32) n n1 ... np n n>1 n>n1>...>np>1 n>1

12 The numbers κ0 and κ−1 are found for the values to which Fontana–Mascheroni and Fontana series converge respectively [6, pp. 406, 410]. 13Other possible generalizations of Euler’s constant were proposed by Briggs, Lehmer, Dilcher and some other authors [8, 32, 39, 36, 40, 22].

9 n (m) −m where li is the integral logarithm function, Hn := ∑ k stands for the generalized harmonic number and k=1 14 (Pn) denotes the sequence of polynomials

1  2  P0 = 1, P1(x1) = x1, P2(x1, x2) = 2 x1 + x2 ,

1  3  P3(x1, x2, x3) = 6 x1 + 3x1x2 + 2x3 ,...

In particular, for the series κ1 mentioned in Theorem 1 and Remark 2, this gives

1 ∞   Gn 1 1 κ = = − + ln x dx (33) 1 ∑ 2 ˆ ( − ) n=1 n ln 1 x x 0

1 ∞ − li(1 − x) + γ + ln x n o = dx = − li 1 − e−x + γ − x dx (34) ˆ x ˆ 0 0

∞ (−1)k ∞ H R H = n = n . (35) ∑ k ∑ (n + 1)k ∑ n k=2 n=1 n>1 Moreover, we have

1 γ2 π2 Ψ(x + 1) + γ κ = γ + − + dx (36) 1 1 2 12 ˆ x 0 1 γ2 π2 1 1 = + − + Ψ2(x + 1) dx , (37) 2 12 2 2ˆ 0 where Ψ denotes the (logarithmic derivative of the Γ–function).

Proof of formula (28) We first write the generating equation for Gregory’s coefficients, Eq. (6), in the following form

1 1 ∞ + = |G | xn−1 , |x| < 1 . (38) ( − ) ∑ n ln 1 x x n=1

Multiplying both sides by lnp x, integrating over the unit interval and changing the order of summa- tion and integration15 yields:

1 1  1 1  ∞ + lnpx dx = |G | xn−1 lnpx dx , Re p > −1 . (39) ˆ ( − ) ∑ n ˆ ln 1 x x n=1 0 0

14More generally, these polynomials, called the modified Bell polynomials, are defined by the generating function:  ∞  ∞ zk n exp ∑ xk k = ∑ Pn(x1, ··· , xn) z , see [11]. k=1 n=0 15The series being uniformly convergent.

10 The last integral may be evaluated as follows. Considering Legendre’s integral Γ(p + 1) = tpe−tdt taken over [0, ∞) and making a change of variable t = −(s + 1) ln x , we have ´

1 Γ(p + 1) xs lnpx dx = (−1)p , Re s > −1, Re p > −1. (40) ˆ (s + 1)p+1 0

Inserting this formula into (39) and setting n − 1 instead of s, yields (28).

Proof of formula (29) p+1 N Putting in (38) x = x1x2 ··· xp+1 and integrating over the volume [0, 1] , where p ∈ , on the one hand, we have

1 1 ∞ ∞ n−1 |G | ··· |G |x x ··· x  dx ··· dx = n . (41) ˆ ˆ ∑ n 1 2 p+1 1 p+1 ∑ p+1 n=1 n=1 n 0 0 | {z } (p+1)-fold

On the other hand 1  1 1  li(1 − y) − γ − ln y + dx = − . ˆ ln(1 − xy) xy y 0

Taking instead of y the product x1x2 ··· xp and setting x = xp+1 , and then integrating p times over the unit hypercube and equating the result with (41) yields (29).

Proof of formulas (30) and (31) Writing in the generating equation (10) −x instead of z, multiplying it by lnm x/x and integrating over the unit interval, we obtain the following relation16

∞ S (n, k) ( ) = (− )m+k 1 Ω k, m 1 m! k! ∑ m+1 , n=k n! n where 1 lnk(1 − x) lnm x Ω(k, m) := dx , k, m ∈ N . ˆ x 0 By integration by parts, it may be readily shown that

k Ω(k, m) = Ω(m + 1, k − 1) , m + 1 and thus, we deduce the duality formula:

∞ S (n, k) ∞ S (n, m + 1) 1 = 1 ∑ m+1 ∑ k . n=k n! n n=m+1 n! n

16See also [41, Theorem 2.7].

11 Furthermore, the Maclaurin series expansion of ez with z = ln(1 − x) gives

∞ lnk+1(1 − x) x + ln(1 − x) = − , x < 1. ∑ ( + ) k=1 k 1 ! Hence

1 1  1 1  ∞ lnk+1(1 − x) lnm x dx + lnmx dx = − · · ˆ ( − ) ˆ ∑ ( + ) ( − ) ln 1 x x k=1 k 1 ! ln 1 x x 0 0 ∞ ∞ k+1 ∞ Ω(k, m) (−1) S1(n, m + 1) = − = (−1)mm! , ∑ ( + ) ∑ ( + ) ∑ k k=1 k 1 ! k=1 k 1 n=m+1 n! n which is identical with (30) if setting m = p. Moreover, it is well known that

S1(n + 1, m + 1) ( ) ( ) (m) = P H 1 , −H 2 ,..., (−1)m−1 H  , n! m n n n see [17, p. 217], [37, p. 1395], [31, p. 425, Eq. (43)], [4, Eq. (16)], which immediately gives (31) and completes the proof.

Proof of formula (32)

This formula straightforwardly follows form the fact that κp = Fp(1), see [11, p. 307, 318 et seq.], where Fp(s) is the special function defined by

∞ ∞ n   |G | 1 |G + | n F (s) := n D( )(n) = n 1 (−1)k (k + 1)−s . p ∑ p s ∑ ( + )p ∑ n=1 n x n=0 n 1 k=0 k

Proof of formulas (36) and (37) These formulas immediatly follow from [10, Eqs. (3.21) and (3.23)] and (35).

Acknowledgments

The authors warmly thank the referee for his valuable comments. We are also grateful to Vladimir V. Reshetnikov for his kind help and useful remarks.

References

[1] T. M. Apostol, Formulas for higher derivatives of the , Mathematics of Computation, vol. 44, no. 169, pp. 223–232 (1985).

[2] B. C. Berndt, On the Hurwitz Zeta–function, Rocky Mountain Journal of Mathematics, vol. 2, no. 1, pp. 151–157 (1972).

[3] B. C. Berndt, Ramanujan’s Notebooks, Part I. Springer-Verlag, 1985.

[4] Ia. V. Blagouchine, Expansions of generalized Euler’s constants into the series of polynomials in π−2 and into the formal enveloping series with rational coefficients only, Journal of Number Theory (Elsevier), vol. 158, pp. 365–396 and vol. 173, pp. 631–632, arXiv:1501.00740 (2016).

12 [5] Ia. V. Blagouchine, Three notes on Ser’s and Hasse’s representations for the zeta-functions, arXiv:1606.02044 (2016).

[6] Ia. V. Blagouchine, Two series expansions for the logarithm of the involv- ing Stirling numbers and containing only rational coefficients for certain arguments related to π−1, Journal of Mathematical Analysis and Applications (Elsevier), vol. 442, pp. 404–434, arXiv:1408.3902 (2016).

[7] Ia. V. Blagouchine, A note on some recent results for the Bernoulli numbers of the second kind, J. Integer Seq., vol. 20, no. 3, Article 17.3.8, pp. 1-7, 2017, arXiv:1612.03292 (2017).

[8] W. E. Briggs, The irrationality of γ or of sets of similar constants, Vid. Selsk. Forh. (Trondheim), vol. 34, pp. 25–28 (1961).

[9] W. E. Briggs and S. Chowla, The power series coefficients of ζ(s), The American Mathematical Monthly, vol. 62, pp. 323–325 (1955).

[10] B. Candelpergher, Ramanujan summation of divergent series, Lecture Notes in Mathematics se- ries, Vol. 2185, Springer, 2017.

[11] B. Candelpergher and M.-A. Coppo, A new class of identities involving Cauchy numbers, harmonic numbers and zeta values, The Ramanujan Journal, vol. 27, pp. 305–328 (2012).

[12] B. Candelpergher and M.-A. Coppo, Le produit harmonique des suites, Enseign. Math., vol. 59, pp. 39–72 (2013).

[13] B. Candelpergher, M.-A. Coppo and E. Delabaere, La sommation de Ramanujan, Enseign. Math., vol. 43, pp. 93–132 (1997).

[14] B. Candelpergher, H. Gadiyar and R. Padma, Ramanujan summation and the exponential generat- ∞ zk 0 ing function ∑k=0 k! ζ (−k), The Ramanujan Journal, vol. 21, pp. 99–122 (2010). [15] M. W. Coffey, Certain logarithmic integrals, including solution of monthly problem #tbd, zeta values, and expressions for the Stieltjes constants, arXiv:1201.3393v1 (2012).

[16] Collected papers of , Cambridge, 1927.

[17] L. Comtet, Advanced Combinatorics. The art of Finite and Infinite Expansions (revised and en- larged edition), D. Reidel Publishing Company, Dordrecht, Holland, 1974.

[18] D. F. Connon, Some possible approaches to the Riemann hypothesis via the Li/Keiper constants, arXiv:1002.3484 (2010).

[19] M.-A. Coppo, Nouvelles expressions des constantes de Stieltjes, Expositiones Mathematicæ, vol. 17, pp. 349–358 (1999).

[20] M.-A. Coppo and P. T. Young, On shifted Mascheroni series and hyperharmonic numbers, Journal of Number Theory, vol. 169, pp. 1–20 (2016).

[21] Correspondance d’Hermite et de Stieltjes. Vol. 1 and 2, Gauthier-Villars, Paris, 1905.

13 [22] K. Dilcher, Generalized Euler constants for arithmetical progressions, Mathematics of Computa- tion, vol. 59 , pp. 259–282 (1992).

[23] S. R. Finch, Mathematical constants, Cambridge University Press, USA, 2003.

[24] J. Franel, Note no 245, L’Intermédiaire des mathématiciens, tome II, pp. 153–154 (1895).

[25] J. P. Gram, Note sur le calcul de la fonction ζ(s) de Riemann, Oversigt. K. Danske Vidensk. (Sel- skab Forhandlingar), pp. 305–308 (1895).

[26] G. H. Hardy, Note on Dr. Vacca’s series for γ, The Quarterly journal of pure and applied mathe- matics, vol. 43, pp. 215–216 (1912).

[27] G. H. Hardy, Divergent series, Clarendon Press, Oxford, 1949.

[28] M. I. Israilov, On the Laurent decomposition of Riemann’s zeta function [in Russian], Trudy Mat. Inst. Akad. Nauk. SSSR, vol. 158, pp. 98–103 (1981).

[29] J. L. W. V. Jensen, Sur la fonction ζ(s) de Riemann, Comptes-rendus de l’Académie des sciences, tome 104, pp. 1156–1159 (1887).

[30] J. L. W. V. Jensen, Note no 245. Deuxième réponse. Remarques relatives aux réponses de MM. Franel et Kluyver, L’Intermédiaire des mathématiciens, tome II, pp. 346–347 (1895).

[31] V. Kowalenko, Properties and applications of the reciprocal logarithm numbers, Acta Appli- candæ Mathematicæ, vol. 109, pp. 413–437 (2010).

[32] D. H. Lehmer, Euler constants for arithmetical progressions, Acta Arithmetica, vol. 27, pp. 125– 142 (1975).

[33] D. H. Lehmer, The sum of like powers of the zeros of the Riemann zeta function, Mathematics of Computation, vol. 50, no. 181, pp. 265–273 (1988).

[34] J. J. Y. Liang and J. Todd, The Stieltjes constants, Journal of Research of the National Bureau of Standards—Mathematical Sciences, vol. 76B, nos. 3–4, pp. 161–178 (1972).

[35] Z. Nan-You and K. S. Williams, Some results on the generalized Stieltjes constant, Analysis, vol. 14, pp. 147–162 (1994).

[36] T. H. Pilehrood and K. H. Pilehrood, Criteria for irrationality of generalized Euler’s constant, Jour- nal of Number Theory, vol. 108, pp. 169–185 (2004).

[37] L.-C. Shen, Remarks on some integrals and series involving the Stirling numbers and ζ(n), The Transactions of the American Mathematical Society, vol. 347, no. 4, pp. 1391–1399 (1995).

[38] R. Sitaramachandrarao, Maclaurin coefficients of the Riemann zeta function, Abstracts of Papers Presented to the American Mathematical Society, vol. 7, no. 4, p. 280, *86T-11-236 (1986).

[39] T. Tasaka, Note on the generalized Euler constants, Mathematical Journal of Okayama University, vol. 36, pp. 29–34 (1994).

[40] L. Xia, The parameterized-Euler-constant function γa(z), Journal of Number Theory, vol. 133, no. 1, pp. 1–11 (2013).

14 [41] C. Xu, Y. Yan and Z. Shi, Euler sums and integrals of polylogarithm functions, Journal of Number Theory, vol. 165, pp. 84–108 (2016).

[42] P. T. Young, A 2-adic formula for Bernoulli numbers of the second kind and for the Nörlund numbers, Journal of Number Theory, vol. 128, pp. 2951–2962 (2008).

[43] P.T. Young, Rational series for multiple zeta and log gamma functions, Journal of Number Theory, vol. 133, pp. 3995–4009 (2013).

15