arXiv:2102.04891v5 [math.FA] 13 Aug 2021 µ comprehensive truly his In Introduction 1 lands; nteGmafnto,wt plctost h eafunctio Beta the to applications w with from function, pro Weierstrass’s Gamma integral after the Beta perhaps on and Euler’s hand of other the properties on Today, many p the 123-343 knowledge were the time pp. reviewed has numbers Binet page contributions, correct beautiful the but article, ordinary ie a eie w nerlrpeettoso i funct his of representations ( Re integral two derived has Binet ( ∗ 1 z aut fEetia niern,MteaisadComput and Engineering, Electrical of Faculty h eeec nWitkradWto 2,font,p 248] p. footnote, [28, Watson and Whittaker in reference The nrlto oteGmafnto ( Γ function Gamma the to relation in ) z ) email xedBntsmto ofcoilsre fLpaetransforms. Laplace of to method Binet’s extend xml ht ihasm ubro em vlae,teBntgen Binet the evaluated, of terms value of optimized number same an a with that, example ucinadapytoesre otedgmaadplgmafunct polygamma and Stirling’s f digamma with the series series to factorial factorial series generalized corresponding those the apply and compute function We presented. are o h ie function Binet the for fe re eiwo h ie function Binet the of review brief a After > ie’ atra eisadetnin oLpaetransfo Laplace to extensions and series factorial Binet’s eivsiaeagnrlzto fBntsfcoilsre ntepara the in series factorial Binet’s of generalization a investigate We 0, [email protected]. : µ 1 α mmi”[,Scin3 .23 n13,JcusBntdefines Binet Jacques 1839, in 223] p. 3, Section [2, “memoir” ( µ z a ettebs osbeacrc fSiln’ xaso.Fnly w Finally, expansion. Stirling’s of accuracy possible best the beat can ( o ( Γ log = ) z o ( Γ log = ) ef nvriyo Technology of University Delft µ µ ( ( z z z = ) itVnMieghem Van Piet z ) 2 = ) ) 1Ags 2021 August 11 − − z m µ X asymptotic  as ) ∞  =1 ( Abstract z z Z z ,svrlpoete fteBntpolynomials Binet the of properties several ), 0 − Q − ∞ bu h eafnto eoe13.Tefclpita that at point focal The 1839. before Beta-function the about 1 m k 1 2 1 2 =0 utadHne’ otu nerl h hoyconcentrate theory the integral, contour Hankel’s and duct arctan −   e n. 1 itt oksz raie pr rmadn i own his adding from Apart treatise. book-size a to oint b 2 ( log m log πt z rSine . o 01 60G ef,TeNether- The Delft, GA 2600 5031, Box P.O Science, er ihpoete fteGmafnto eederived. were function Gamma the of properties hich xaso n eosrt yanumerical a by demonstrate and expansion ( + α z − z ) α + ihpg ubr rm13t 4 ugssan suggests 143 to 123 from numbers page with + 1 o 2,p 4-5] 1,p 1-1] for 211-217], p. [18, 248-251], p. [28, ion z t + z ∗  z dt − k − ) 1 2 1 2 o (2 log rtedrvtvso h Binet the of derivatives the or o (2 log rlzdfcoilsre with series factorial eralized π os ecmaeBinet’s compare We ions. meter ) π (1) ) α rms i function his b m ( α e ) (2) s and 1 e zt 1+ e t 2 µ (z)= ∞ − − dt (3) 2 t 1 e t − t Z0  − −  There exist more representations2 of µ (z), but here we merely concentrate on elegant, converging factorial series for Re (z) > 0 due to Binet3 in [2, p. 234],

∞ β µ (z)= m (6) (z + 1) (z + 2) (z + m) m=1 X · · · where the coefficients are 1 1 1 β = u u (u + 1) (u + m 1) du (7) m m − 2 · · · − Z0   1 59 29 533 1577 280361 69311 Explicitly, β1 = β2 = 12 , β3 = 360 , β4 = 60 , β5 = 280 , β6 = 168 , β7 = 5040 , β8 = 180 , and βm is positive and rapidly increasing in m. Nearly at the end of his memoir and somewhat hidden, Binet [2, p. 342] has given a second factorial series (31), that is rederived slightly differently in Theorem 1 in Section 2.3 and generalized to Laplace transforms in Section 6.2. Recently, Nemes [13, Theorem 2.1] has generalized Binet’s expansion (6), for 0 a 1 and ≤ ≤ Re (z) > 0,

1 1 ∞ c (a) log Γ (z + a)= z + a log z z + log (2π)+ m (8) − 2 − 2 (z + 1) (z + 2) (z + m)   mX=1 · · · where

m 1 1 1 1 1 a − cm (a)= u + a u (u + 1) (u + m 1) du (u a +1+ j) du (9) m 0 − 2 · · · − − m 0 − Z   Z jY=0

Clearly, if a = 0, we retrieve Binet’s first factorial expansion (6) and coefficients βm in (7). Nemes’ expansion (8), expressed in terms of Binet’s function µ (z) with the definition (1),

1 z a ∞ cm (a) µ (z)= z log − + a + m − 2 z k=1(z a + k)   mX=1 − bears resemblance to Q ∞ bm (α) µ (z)= m 1 (10) m=1 k=0− (z + α + k) X 2Blagouchine [3] lists 7 different formulae for log Γ (z).Q 3In Binet’s notation [2, p. 234], I (1) I (2) I (3) 2µ (z)= + + + · · · (4) z + 1 2(z +1)(z + 2) 3(z +1)(z +2)(z + 3) where the integral form of the coefficients (derived at [2, p. 238]) is

1 I (m)= x (x +1)(x + 2) · · · (x + m − 1) (2x − 1) dx (5) Z0

1 1 59 227 Binet [2, p. 234] lists the first few coefficients, I (1) = 6 , I (2) = 3 , I (3) = 60 and I (4) = 60 .

2 in our main Theorem 3, where the coefficients bm (α) are called Binet , to honour Jacques Binet. Much earlier, Hermite [11] has deduced the corresponding generalized Stirling asymptotic expansion

K k 1 1 1 ( 1) − B (a) log Γ (z + a) z + a log z z + log (2π)+ − k+1 (11) ≃ − 2 − 2 k (k + 1) zk   Xk=1 in terms of the Bernoulli polynomials Bn (a) that reduces for a = 0 to Stirling’s original asymptotic series (23). Starting from a complex integral (20) for Binet’s function µ (z), Stirling’s asymptotic series (23) is derived in Section 2.1.1, where also the meaning of the upper bound K is explained. Section 2.1.2 presents the convergent companion (26) of Stirling’s (if K in (23)). →∞ The main motivation that led to this article is twofold. Originally, I was confused about Binet’s achievements: he derived two different factorial expansions (6) and (31) for the same function µ (z). While I thought initially that one of them must be wrong, I discovered later with (10) that infinitely many different factorial series exist. The second motivation was my unbelief that Stirling’s asymptotic but divergent expansion seems unbeatable in performance for some optimal, finite K. Our main result is the demonstration that there exist infinitely many different factorial expansions in a complex parameter α for Binet’s function µ (z) in Theorem 3 (Section 3) and for Laplace transforms in Theorem 4 (Section 6.3). For α = 1 in (10), we recover Binet’s first factorial series (6) with bm (1) = βm in (7), while α = 0 in (10) corresponds to Binet’s second factorial series (31) with bm (0) = cm in (32). Interestingly, the α = 1 factorial series has all positive coefficients and any truncation thus lower bounds Binet’s function µ (z), while the α = 0 factorial series is shown in

Theorem 2 to possess coefficients cm < 0 for m> 2. Thus, a truncation of (31) upper bounds µ (z). For 1 α around 10 , which is close to the largest zero of the Binet bm (α), numerical computations exhibit the fastest convergence. With the same number K of terms evaluated, the variant α = 0 is more accurate than the variant α = 1. Perhaps, the slower convergence of Binet’s expansion (6) has caused that his factorial series is not listed in handbooks of functions, like Abramowitz & Stegun [1] nor in its successor by Olver et al. [15]. We will first discuss the main properties of Binet’s function µ (z) in Section 2 and the deductions from the complex integral (20) in Section 2.1, before we review parts of Binet’s great treatise. In subsection 2.2, we sketch Binet’s route towards his first factorial series (6), that is covered in the literature (see e.g. [28, p. 253], [16, p. 30]). Binet’s second factorial series, that I have not found in later works, is derived in more detail in subsection 2.3, also because I believe that, being the case α = 0 in (10), it is slightly more important than his first series (6) corresponding to α = 1. Moreover, Binet’s method towards his second factorial series, which is a recipe in five steps, enables a far reaching generalization to Laplace transforms as explained in Section 6. Factorial expansions for the derivatives dnµ(z) dzn are derived in Section 4 and applied to the digamma and polygamma function. In particular for Γ′(z) the digamma function ψ (z)= Γ(z) , we thus add a (60) to its asymptotic counterpart (61). Section 5 discusses and compares, with a same number of terms, the accuracy of Stirling’s asymptotic expansion and the best possible that can be attained by the generalized Binet factorial expansion. The commonly accepted belief about the superiority of Stirling’s asymptotic expansion is demonstrated and plotted in Fig. 1. However, at the zeros of the Binet polynomial bm (α), the accuracy of the generalized Binet factorial expansion (10) improves considerably as drawn in Fig. 2.

3 By a numerical example, we show that Stirling’s series accuracy is not always better than a factorial series (with a same number of terms)! In other words, the generalized Binet expansion (10) can be optimized with respect to the “free” parameter α to achieve, at least, a comparable accuracy with a same computational effort. Perhaps, this observation deserves to list the generalized Binet factorial expansion (10) in handbooks of functions. Computations are deferred to the appendices in order to enhance the readability and focus on the essential parts.

2 Binet’s function µ (z)

The definition (1) of Binet’s function µ (z) directly shows, for a positive integer z = n, that 1 1 µ (n)= n + log (n 1)! n log (n) log (2π) − − − 2 − 2   The µ (1) 0.0811, µ (2) 0.0413, µ (3) 0.0277, µ (4) 0.0208, µ (5) 0.0166,. . . , ≃ ≃ ≃ ≃ ≃ µ (10) 0.0083 demonstrates the slow decay roughly as µ (n) 1 . The precise decay is given in ≃ ≈ 12n (65) below. Maximum at real, positive values of z. Binet’s integral (3) can be rewritten as e zt 1 1 1 µ (z)= ∞ − + dt t et 1 − t 2 Z0  −  1 1 1 1 where 0 et 1 t + 2 2 for real, non-negative t. Since the integrand is positive for real z = x> 0, ≤ − − ≤ we observe that µ (x) > 0 for real, positive x. In addition, for a z = x + iy, the above integral shows that µ (z) is analytic for Re (z) > 0 and that e xt 1 1 1 µ (z) ∞ − + dt = µ (x) | |≤ t et 1 − t 2 Z0  −  In other words, the maximum absolute value of Binet’s function µ (z) for Re(z) > 0 is attained at the positive real axis. Moreover, µ (x) strictly decreases with x = Re(z). Another rather straightforward bound is

1 1 1 1 ∞ xt 1 1 1 1 1 µ (z) max + e− dt = max + | |≤ 0 t t et 1 − t 2 x 0 t t et 1 − t 2 ≤  −  Z0 ≤  −  Since the maximum occurs at t = 0, the of the Bernoulli numbers Bn

1 1 1 ∞ t2n 1 = + B − for t 2π (12) et 1 t − 2 2n (2n)! | |≤ n=1 − X

1 1 1 1 B2 1 illustrates that max0 t t et 1 t + 2 = 2 = 12 . Hence, for x = Re(z) > 0, we find [28, p. 249] ≤ − −   1 µ (z) µ (x) (13) | |≤ ≤ 12x

Difference µ (z + 1) µ (z). Binet’s definition (1) − 1 1 µ (z) = log Γ (z) z log z + z log (2π) − − 2 − 2   4 illustrates that the µ∗ (z) = µ (z∗), by the reflection principle [22, p. 155]. The functional equation of the Gamma function Γ (z +1) = zΓ (z) leads to 1 1 µ (z + 1) = log Γ (z) + log z z + log (z +1)+ z + 1 log (2π) − 2 − 2   The forward difference ∆µ(z)= µ (z + 1) µ (z) equals − 1 z µ (z + 1) µ (z)= z + log + 1 (14) − 2 z + 1   which is valid for any complex z, except at the negative real axis that is a branch cut for µ (z). In particular, around z = 0, the forward difference (14) shows, with µ (1) = 1 1 log (2π), that − 2 1 1 µ (z) log z log (2π) (15) ∼−2 − 2 illustrating that Binet’s function µ (z) possesses a logarithmic singularity at z = 0. Erdelyi et al. [6, p. 24] deduce from Gauss’s multiplication formula that z+1 1 log Γ (t) dt = z log z z + log (2π) − 2 Zz which we rewrite with the definition (1) as z+1 1 z 1 µ (t) dt = z (z + 1) log + z + (16) 2 z + 1 2 Zz     Differentiation of (16) with respect to z again leads to the forward difference (14). For z = 0 in (16), 1 1 we find 0 µ (t) dt = 4 . ReflectionR formula of Binet’s function µ (z). We replace z 1 z in Binet’s definition (1) → − 1 1 µ (1 z) = log Γ (1 z) z log (1 z) z + 1 log (2π) − − − 2 − − − − 2   which added to µ (z) in (1), yields 1 z µ (z)+ µ (1 z) = log (Γ (z)Γ(1 z)) z log + 1 log (2π) − − − − 2 1 z −   − After invoking the reflection formula of the Gamma function Γ (z) Γ (1 z) = π , we find the − sin πz corresponding “reflection” formula for Binet’s function 1 z µ (1 z) = 1 µ (z) log (2 sin πz) z log (17) − − − − − 2 1 z   − which is valid for any complex number z = x + iy, with the exception of the negative real axis and odd integers x = 2n + 1 with n Z at any y. Since µ (z) is analytic for Re (z) > 0, the Binet ∈ reflection formula (17) illustrates that Binet’s function µ (z) has only logarithmic singularities at negative integers z = k (with k N) including z = 0, as shown in (15). − ∈ Duplication and multiplication formula for µ (z). Combining the duplication formula Γ (2z) = 22z−1 1 √π Γ (z)Γ z + 2 in [1, 6.1.18] for the Gamma function Γ (z) and the definition (1) of Binet’s function µ (z) leads to 1 1 1 µ (2z)= µ z + + µ (z)+ z log 1+ (18) 2 2z − 2     5 n 1 1 (1 n) 1 − 2 nz 2 k which is generalized by Gauss’s multiplication formula Γ (nz) = (2π) − n − Γ z + n in [1, k=0 6.1.20] as Y 

n 1 n 1 − k − k 1 k 1 µ (nz)= µ (z)+ µ z + + z + log z + (n 1) + z log z (19) n n − 2 n − − 2 Xk=1   Xk=1       After letting z = 1 in (19), we find for n 2 n ≥ n 2 n 2 1 − k + 1 − k + 1 1 k + 1 1 1 µ = µ log + (n 1) log n n − n − n − 2 n − 2 − n   Xk=1   Xk=1       For example, for n = 2, we obtain µ 1 = 1 (1 log 2) 0.1534 but, for n = 3, µ 1 = 1 µ 2 2 2 − ≃ 3 − 3 − 1 log (2) 1 log (3), which is also immediate from Binet’s reflection formula (17). Gauss’s multiplication 6 − 2    formula already contains the information embedded in the reflection formula Γ (z) Γ (1 z)= π . − sin πz The duplication formula (18) yields a closed expression for z = n + 1 with integer n 1, 2 ≥ 2n 1 1 − 1 µ n + = log k n (log (4n + 2) 1) + (log (2) + 1) 2 − − 2   kX=n 2.1 Complex integral for Binet’s function µ (z)

In Appendix A, we deduce the complex integral (in two ways)

1 c+i π ζ (s) µ (z)= ∞ zsds with 1

1 1 sin πx ∞ log(k) sin (2πxk) log Γ (x) = (γ + log (2π)) x log + 2 − − 2 π πk   Xk=2 We evaluate the integral (20) along the line s = c + it, where 1

it log z 1 1 ∞ e ζ 2 + it µ (z) z− 2 − dt | |≤ cosh πt 1 + 2it  Z−∞ − t arg( z) 1 ∞ e− C z− 2 dt ≤ cosh πt Z−∞ where C is positive , demonstrating existence for all complex z = z ei arg z provided | | arg z <π. In other words, the integral (20) defines Binet’s function µ (z) everywhere in the complex | | plane, except at the negative real axis, where µ (z) possesses a branch cut. The well-known Fourier integral izu ∞ e π 1 du = πz valid for Im z < α cosh αu α cosh 2α | | Z−∞ 1 eit log z 1 2z 2 shows that ∞ dt = = and, roughly, that the the integral (22) can be estimated cosh πt cosh log z z+1 −∞ 2 1 as µ (z)= OR z , which complements the bound (13) to complex z, except at the negative real axis. iθ Branch cut  along the negative real axis. The complex integral (20) indicates with z = re that c+i iθ iθ ∞ sin θs ζ (s) s µ re µ re− = r ds − − c i sin πs s     Z − ∞ is purely imaginary, because µ reiθ µ re iθ = µ reiθ µ reiθ = 2i Im µ reiθ , and that − − − ∗    c+i   iθ iθ ∞ ζ (s) s lim µ re µ re− = r ds θ π − − c i s →      Z − ∞ By moving the line of integration from c< 0 to c′ > 1, two poles at s = 0 and s = 1 are enclosed and Cauchy’s leads to

c′+i iπ iπ s ζ (s) (s 1) s ∞ ζ (s) s µ re µ re− = 2πi lim ζ (s) r + lim − r r ds − s 0 s 1 s − c′ i s  → →  Z − ∞   c′ c′ ζ (x + iT ) x+iT ζ (x iT ) x iT + lim r dx lim − r − dx T x + iT − T x iT →∞ Zc →∞ Zc − 1 x Due to ζ (x + iT )= O T 2 − for x 0 and large T and since c< 0 can be chosen small enough, the ≤ ′   1 c +i ζ(s) s r limits T vanish. After using Perron’s formula [22, p. 301], 2πi c′ i ∞ s r ds = n=1 1 = [r], → ∞ − ∞ which is the integral part of r, we find that the difference at both sidesR of the branch cutP is periodic in r 0, ≥ iπ iπ 1 µ re µ re− = 2πi + r [r] − −2 −     and vanishes at r = 1 + n with integer n N. 2 ∈ From the complex integral (20), we will now deduce Stirling’s asymptotic series in Section 2.1.1 and its convergent companion in Section 2.1.2.

7 2.1.1 Stirling’s asymptotic series

We cannot close the contour in (20) over the entire Re (s) < 0 -plane, because the functional equation s 1 πs of the Riemann Zeta-function ζ(s) = 2(2π) − sin 2 Γ(1 s)ζ(1 s) indicates that ζ( s)= O (Γ(s)). −( 1)k − − However, neglecting this restriction and using ζ( k) = − B and the odd Bernoulli numbers − k+1 k+1 B2k+1 = 0, for k > 0, leads to Stirling’s asymptotic approximation [1, 6.1.41] in the Poincar´esense (see e.g. [16])

K k K K ( 1) ζ ( k) k Bk+1 B2m µ (z) − − z− = = (23) ≃ k k (k + 1) zk (2m 1) (2m) z2m 1 m=1 − Xk=1 Xk=1 X − Given that (23) diverges if K , Stirling’s asymptotic approximation (23) is surprisingly accurate →∞ up to some finite K K∗ (z), where K∗ (z) depends upon z and is roughly equal to the minimum B ≤ k-value of | k+1| . By introducing the generating function (12) of the Bernoulli numbers in Binet’s k(k+1) z k integral (3) | |

zt 2m 2 ∞ e− 1 1 1 ∞ zt ∞ t − µ (z)= + dt = e− B dt t et 1 − t 2 2m (2m)! 0 0 m=1 ! Z  −  Z X only valid for Re (z) > 0 and reversing sum and integral, while ignoring the convergence restriction t < 2π in the sum, another asymptotic series is obtained | | zt 2m 2 ∞ ∞ e− t − dt ∞ B µ (z) B 0 = 2m ≃ 2m (2m)! (2m)z2m 1 m=1 R m=1 − X X which is inferior to Stirling’s asymptotic series (23). On the other hand for z < 1, the contour in (20) can be closed over the Re (s) > 0 -plane, where | | two double poles at s = 0 and s = 1 are encountered whose residues are computed in Appendix A, resulting in k ∞ ( 1) ζ (k) 1 1 µ (z)= − zk z (log z 1+ γ) log z log(2π) (24) k − − − 2 − 2 Xk=2 where the of log Γ(z + 1) around z = 0,

∞ ( 1)k ζ(k) log Γ(z +1) = γz + − zk − k Xk=2 follows directly from Weierstrass’ product

1 γz ∞ z z/n = e 1+ e− (25) Γ(z + 1) n n=1 Y   k K ( 1) ζ( k) k In contrast to Stirling’s series − − z for some finite K K (z) in (23), violation of the k=1 k − ≤ ∗ restriction z < 1 in the Taylor series in (24) leads to useless results. | | P

8 2.1.2 Convergent companion of Stirling’s asymptotic series (23)

4 m The Taylor series of the entire function (s 1) ζ (s)= m∞=0 gm (z) (s z) around s0 = z converges − − m for all finite complex z. After substituting the Taylor series (s 1) ζ (s) = ∞ g (1) (s 1) P − m=0 m − around s0 = 1 into the complex integral (20), it is shown in Appendix A.2 thatP

m 1 v ∞ − m 1 v (v+1) 1 µ (z)= g (1) (v 1)!( 1) − − (26) m − − Sm z m=1 v=1 X X   (k) where m is the Stirling number of second Kind. The Taylor coefficients [1, (23.2.5)] for k 0 S ≥ K ( 1)m lnm n lnm+1 K gm+1(1) = − lim (27) m! K n − m + 1 →∞ n=1 ! X K 1 are attributed to Stieltjes, with g0 (1) = 1 and g1 (1) = γ = limK n=1 ln K . However, →∞ n − computationally, Stieltjes expression (27) is less suited and we present fastP converging series for gm (1) 1 in Appendix A.3. Since (s 1) ζ (s) is entire, the Taylor coefficients gm (1) = O m! – just as those − z of any entire function of order 1 like e – decay rapidly in m and only a few terms  in (26) provide accurate results for Binet’s function µ (z). The reversal of the m- and v- sum in (26) is not allowed. However, it is interesting to illustrate what happens if we reverse the sums:

m v ∞ m v (v+1) 1 µ (z)= g (1) (v 1)!( 1) − m+1 − − Sm+1 z m=1 v=1 X X   ∞ ∞ m v (v+1) 1 = (v 1)! g (1) ( 1) − − m+1 − Sm+1 zv v=1 (m=v ) X X We substitute the closed form (96) of the Stirling numbers, (m) = 1 m ( 1)m j m jk, using Sk m! j=0 − − j (m) =0 if k

∞ m v (v+1) ∞ m v (v+1) g (1) ( 1) − = g (1) ( 1) − m+1 − Sm+1 m+1 − Sm+1 m=v m= 1 X X− v+1 1 ∞ m v v+1 j v + 1 m+1 = g (1) ( 1) − ( 1) − j (v + 1)! m+1 − − j m= 1 j=0   X− X v+1 1 v + 1 ∞ = ( 1)j g (1) ( 1)m+1jm+1 (v + 1)! − j m+1 − j=0   m= 1 X X− m With (s 1) ζ (s)= ∞ g (1) (s 1) , we have − m=0 m − P v+1 ∞ m v (v+1) 1 j 1 v + 1 g (1) ( 1) − = ( 1) − jζ (1 j) m+1 − Sm+1 (v + 1)! − j − m=v X Xj=0   4An entire function has no singularities in the finite and possesses a Taylor series around any finite point with infinitely large radius of convergence. An entire function is sometimes also called an integral function (as e.g. in [22]), which can be a confusing name.

9 After substitution in the series for µ (z),

v+1 ∞ (v 1)! j 1 v + 1 1 µ (z)= − ( 1) − jζ (1 j) (v + 1)! − j − zv v=1 X Xj=0   ( 1)k n n+1 and using ζ( k)= − B and B = 0, we obtain − k+1 k+1 j=0 j j P  ∞ B 1 µ (z)= v+1 (v + 1) v zv v=1 X Hence, reversal of the m- and v- sum in (26) again leads to Stirling’s diverging asymptotic series (23). The series (26) converges for all complex z with arg z < π and can be regarded as the convergent | | companion of Stirling’s asymptotic series (23).

2.2 Binet’s first factorial series for µ (z)

z We review Binet’s first expansion for µ (z) in [2, Section 3, pp. 223-229]. Writing log z+1 = log 1 1 , Binet expands the right-hand side of the forward difference formula (14) − z+1   1 1 µ (z + 1) µ (z)=1+ z + log 1 − 2 − z + 1     zn by introducing the Taylor series around z = 0 of log (1 z)= ∞ , convergent for z < 1, and 0 − − n=1 n | | obtains for z + 1 > 1, | | P 1 ∞ n 1 µ (z) µ (z +1) = (28) − 2 (n + 2) (n + 1) n+1 n=1 (z + 1) X Binet replaces z z + k in (28), sums over all integer k 0, → ≥ N 1 N 1 − 1 ∞ n − 1 µ (z + k) µ (z + k +1) = − 2 (n + 2) (n + 1) n+1 n=1 (z + k + 1) Xk=0 X Xk=0 and rewrites the at the left-hand side

N 1 1 ∞ n − 1 µ (z) µ (z + N)= − 2 (n + 2) (n + 1) n+1 n=1 (z + k + 1) X Xk=0

After observing that limN µ (z + N) = 0 (which follows e.g. from the bound (13)), Binet [2, eq. →∞ (58), p. 229] arrives at his first convergent expansion

1 ∞ n ∞ 1 µ (z)= (29) 2 (n + 2) (n + 1) n+1 n=1 (z + k + 1) X Xk=0 The polygamma functions, for any integer n 1, possess the convergent series [1, 6.4.10] ≥ ∞ 1 ψ(n)(z) = ( 1)n+1n! (30) − (z + k)n+1 Xk=0

10 Binet [2, art [20], p. 232-234] then concentrates5 on the evaluation of the k-sum in (29), thus on the higher-order derivatives ψ(n)(z) of the digamma function ψ (z) and presents [2, p. 234] his first factorial expansion (6). Binet [2, art [21], p. 242] proceeds by constructing integrals for µ (z). Introducing Γ(s) s 1 xt Euler’s Gamma integral xs = 0∞ t − e− dt, valid for Re (s) > 0 and Re(x) > 0, into (29) yields, for Re (z) > 0, R 1 ∞ n 1 ∞ ∞ n (z+k+1)t µ (z)= t e− dt 2 (n + 2) (n + 1) n! 0 Xn=1 Xk=0 Z After reversal of integral and k-summation,

n 1 ∞ n 1 ∞ zt t µ (z)= e− dt 2 (n + 2) (n + 1) n! et 1 n=1 0 X Z − n 2 1 1 n tn ett 2et+2+t and using = in the n-sum, which results in ∞ = − 2 , (n+2)(n+1) n+2 − n+1 2 n=1 (n+2)(n+1) n! t et+1 1 1 sin tx dx ∞ then leads to Binet’s integral (3). Via an integral due to Poisson, Pet 1 2t = 4 0 e2πx 1 , Binet − − − also derives (2). Binet writes at length and reconsiders previous derivations, but hisR great Memoire definitely contains the foundations about his function µ (z).

2.3 Binet’s second factorial expansion

Theorem 1 A second convergent factorial series of Binet’s function µ (z) is

1 ∞ c µ (z)= m for Re (z) > 0 (31) z m 1 m=1 k=1− (z + k) X where the rational coefficients are Q

m 1 m (k) ( 1) − kSm c = − (32) m 2m (k + 2) (k + 1) Xk=1 (k) and Sm is the Stirling Number of the First Kind.

We essentially follow the steps in Binet’s original proof in [2, p. 339]. In Section 6.2, we formalize Binet’s proof as a recipe with five steps. Proof (Binet): Binet [2, p. 339] substitutes e t = 1 u or t = log (1 u) in the integral (3), − − − − 1 z 1 (1 u) − 2 u 2u 2µ (z)= − − + du (33) − u log (1 u) log2 (1 u) Z0  − −  and proceeds to expand

2 u 2u 1 u u − + = 2 + log (1 u) log2 (1 u) log (1 u) log2 (1 u) − log (1 u) − −  − −  − 5Binet invokes the factorial expansion

∞ 1 1 a a(a + 1) Γ(p) Γ(a + j) = + + + · · · = p − a p p (p + 1) p (p +1)(p + 2) Γ(a) Γ(p +1+ j) Xj=0

∞ k a and Newton’s difference expansion f (p + a)= k=0 ∆ f (p) k . P 

11 in a Taylor series around u = 0. Instead of following Binet, who has used integrals rather than (k) Stirling numbers Sm , we invoke the Taylor expansion (94) for n = 2, derived in the Appendix B and convergent for u < 1, | | m (k) m 1 u 1 ∞ k!Sm ( u) + = − log (1 u) 2 −2 − (k + 2)! m! log (1 u) m=1 ! − − X Xk=1 and the Taylor series (92)

m (k) m u ∞ Sm ( u) − =1+ − log (1 u) k + 1 m! m=1 ! − X Xk=1 to obtain the Taylor series, valid for u < 1, | | m (k) m 2 u 2u ∞ kSm ( u) − + 2 = − (34) log (1 u) log (1 u) (k + 2) (k + 1)! m! − − mX=1 Xk=1 Introduced in Binet’s function (33) and reversing summation and integral, justified because a Taylor series can be term-wise integrated within its radius of convergence,

m (k) m 1 1 ∞ kSm ( 1) − z 1 m 1 2µ (z)= − (1 u) − u − du (k + 2) (k + 1)! m! 0 − mX=1 Xk=1 Z 1 p 1 q 1 Γ(p)Γ(q) using the Beta integral u (1 u) − du = , valid for Re (p) > 0 and Re(q) > 0, yields a 0 − − Γ(p+q) converging series, for ReR (z) > 0, m (k) m 1 1 ∞ kSm ( 1) − Γ (z) µ (z)= − 2 (k + 2) (k + 1) m Γ (z + m) m=1 ! X Xk=1 Γ(z+m) m 1  With Γ(z) = k=0− (z + k), we arrive at (31). The first fewQ coefficients c in (32) are c = 1 , c = 0, c = 1 , c = 1 , c = 5 , m 1 12 2 3 − 360 4 − 120 5 − 168 c = 11 , c = 3499 , which are smaller in absolute value than 1, but c = 1039 , c = 369689 6 − 84 7 − 5040 8 − 240 9 − 11880 exceed 1 in absolute value. It holds that cm > 1 for m> 8 as shown below after Theorem 2. | | (k) The generating function of the Stirling numbers Sm of the First Kind [1, Sec. 24.1.3 and 24.1.4],

m 1 m x Γ(x + 1) − m! = = (x k)= S(k) xk (35) m Γ(x + 1 m) − m   − kY=0 Xk=0 (k) m (k) m (k) m kSm indicates that k=1 Sm = δ1m. Thus, we can add a k=1 Sm to k=1 (k+2)(k+1) in (32) and find that coefficientP equals P P m 1 m 2 ( 1) − ak + (3a 1) k + 2a c = − − S(k) for m> 1 and any a C m 2m (k + 1) (k + 2) m ∈ Xk=1 1 For example, for a = 6 and m> 1, we find

m 1 m ( 1) − (k 1) (k 2) c = − − − S(k) m 12m (k + 1) (k + 2) m Xk=1 12 (k) illustrating that c2 = 0. The second generating function of the Stirling numbers Sm , convergent for u < 1, is (see e.g [1, 24.1.3.A]) | | ∞ um logk(1 + u)= k! S(k) (36) m m! mX=k

Theorem 2 The rational coefficients cm in the second factorial series (31) of Binet’s function µ (z) can be represented by an integral

m 1 1 1 1 − cm = u u (k u)du (37) m 0 − 2 − Z   kY=1 Moreover, all coefficients cm, except for the first two, are negative, i.e. cm < 0 for all m> 2.

Proof: Using j = 2 1 , the coefficient c in (32) is written as (j+2)(j+1) j+2 − j+1 m

m m (j) m m (j) ( 1) Sm ( 1) 2Sm c = − − m 2m j + 1 − 2m j + 2 Xj=1 Xj=1 (k) q 1 Multiplying both sides of the generating function (35) of the Stirling numbers Sm by x − and integrating yields b m 1 m l+q l+q q 1 − (l) b a u − (u k)du = Sm − (38) a − l + q Z kY=0 Xl=0 After substituting the case for q = 1 and q = 2, we obtain

m m 1 ( 1) 1 − cm = − (1 2u) (u k)du 2m 0 − − Z kY=0 1 1 1 from which (37) follows. For m = 1 in (37), we find c1 = 0 u 2 udu = 12 . − m 1 In the second part, we will demonstrate that c < 0 for m > 2. Since u − (k u) 0 for m R  k=1 − ≥ u [0, 1], we split the integration interval in (37), ∈ Q 1 m 1 1 m 1 2 1 − 1 − mcm = u u (k u)du + u u (k u)du 0 − 2 − 1 − 2 − Z   kY=1 Z 2   kY=1 After making the substitution u = 1 w in the last integral, we arrive at − 1 m 1 m 1 2 1 − − mcm = u (1 u) u (k (1 u)) (k u) du 0 2 − − ( − − − − ) Z   kY=2 kY=2 For m = 2, the right-hand side is zero, because the two products are equal to 1. Since 1 u > u for 1 m 1 m 1 1 − 0 u < , the product − (k (1 u)) < − (k u) for u [0, ) and for m > 2. Hence, we ≤ 2 k=2 − − k=2 − ∈ 2  conclude that cm < 0 forQm> 2. Q

The important fact in Theorem 2, that all coefficients cm < 0 for m> 2, implies that any truncation at m = K > 2 terms in (31) upper bounds Binet’s function µ (z). In appendix D, we derive an other integral (98) for the coefficients cm.

13 2.4 Growth of the coefficient cm with m

m m 1 Since (k u) = (m u) − (k u) for m 1, the integral (37) becomes k=1 − − k=1 − ≥ Q Q m 1 1 1 − (m + 1) cm+1 = u u (m u) (k u)du 0 − 2 − − Z   kY=1 m 1 m 1 1 1 − 1 1 − = m u u (k u)du u u2 (k u)du 0 − 2 − − 0 − 2 − Z   kY=1 Z   kY=1 The last integral is smaller in absolute value than m c , because u2 u for u [0, 1]. However, | m| ≤ ∈ unlike the proof of Theorem 2, the last integral is positive. Indeed,

1 m 1 1 m 1 m 1 1 − 2 1 − − u u2 (k u)du = u u (1 u) (1 u) (k (1 u)) u (k u) dw 0 − 2 − 0 2 − − ( − − − − − ) Z   kY=1 Z   kY=2 kY=2 m 1 m 1 1 and (1 u) − (k (1 u)) > u − (k u) for 0 u< . Thus, we find the inequality − k=2 − − k=2 − ≤ 2 Q (m + 1) cQ m2c m c = m (m + 1) c m+1 ≥ m − | m| m

Iterating this recursion inequality, cm (m 1) cm 1, yields ≥ − −

cm (m 1) cm 1 (m 1) (m p) cm p ≥ − − ≥···≥ − · · · − − 1 (m 1)! With c = and p = m 3, we obtain c − . The recursion inequality demonstrates that 3 − 360 − m ≥− 720 c increases strictly with m for m 2. | m| ≥ The logarithmic behavior (15) of µ (z) around z = 0 shows that limz 0 zµ (z) = 0. Binet’s second → factorial series (31), written as

∞ cm 1 ∞ cm+1 zµ (z)= m 1 = m| | k=1− (z + k) 12 − k=1(z + k) mX=1 mX=2 Q illustrates, with limz 0 zµ (z) = 0 thatQ → ∞ c 1 ∞ c 0= m = | m| (m 1)! 12 − (m 1)! m=1 m=3 X − X − K cm where m=1 (m 1)! converges very slowly with increasing K. For positive real z, it holds that − cm cm 1 ∞ −| | ∞ | | = , which perfectly agrees with the bound (13). Since all c < 0 m=3 Pm 1(z+k) m=3 (m 1)! 12 m k=1 ≤ − Q cm 1 forP 2 0. Alternatively, m 1 m 1 u − with − (k u) = (m 1)! − (1 ), the integral (37) is k=1 − − k=1 − k  Q Q m 1 mc 1 1 − u m = u u 1 du (m 1)! 0 − 2 − k − Z   kY=1   Since all factors in the last integrand are in absolute value smaller than or equal to 1, the integral cm 1 decreases in absolute value with m and we conclude that (m| 1)!| < m , which is a prerequisite for cm − convergence of m∞=1 (m 1)! . The asymptotic behavior of cm for large m is computed in the Appendix − E. P

14 3 Infinitely many factorial series for the Binet function µ (z)

Theorem 3 Binet’s function µ (z) possesses infinitely many factorial expansions in the complex pa- rameter α, for Re (z) > 0 and Re (α) > Re (z), −

∞ bm (α) µ (z)= m 1 m=1 k=0− (z + α + k) X where the Binet polynomials in α are Q

1 k+1 m k+2 k+2 α αk+1 (α 1) 1 α (α 1) 2 − − − k m (k) bm(α)= − − + ( 1) − Sm (39) m  k + 2   k + 1  − Xk=1   (k) and Sm is the Stirling Number of the First Kind. Another expression in terms of the coefficients ck = bk (0) in (32) is m 1 m k 1 − m 1 − − b (α)= c + α − (α + j)c (40) m m k 1 k Xk=1  −  jY=1 The corresponding integral representation is

m 1 1 α 1 − bm(α)= x + α (k + x)dx (41) m α 1 2 − Z −    kY=0 1 α In particular, b1(α)= 12 and b2 (α)= 12 .

Proof: We write Binet’s integral in (33), valid for Re (z) > 0, as

1 z+α 1 u (1 u) − α 2 1 u µ (z)= − (1 u)− − du u − log (1 u) − log2 (1 u) Z0  − −  After substituting the Taylor series (95) in Appendix C

u α 2 1 u ∞ bm (α) m gα (u) = (1 u)− − = u (42) − log (1 u) − log2 (1 u) (m 1)!  − −  mX=1 − and following the same steps as in the proof of Theorem 1, we arrive at (10). Executing the Cauchy α cm m product of the two Taylor series of (1 u)− and g0 (u)= m∞=1 (m 1)! u and equating corresponding − − powers in u, leads to the factorial expansion (40) of theP Binet polynomial bm (α) in terms of the coefficients ck = bk (0) in (32). We proceed by deducing (39). Introducing the series (10) in the difference formula (14), provides a factorial expansion for the function

1 z ∞ mb (α) 1+ z + log = m (43) 2 z + 1 − m (z + α + k) m=1 k=0   X which we rewrite, after denoting y = z + α, as Q

1 α 1 α ∞ mb (α) 1 y + α log 1 − log 1 = m − 2 − − y − − y − m (y + k) m=1 k=0        X Q 15 We expand now both sides of (43) into powers of 1 . The Taylor series around z = 0 of log (1 x)= y 0 − xk ∞ , convergent for z < 1, in the left-hand side of (43), leads, for max ( α 1 , α ) < y , to − k=1 k | | | − | | | | | P α 1 1 k k 1 − k+1 k+1 α α (α 1) 1 − y ∞ α (α 1) 2 − − − 1 1 y + α log = − − + k − 2 −  1 α  −  k + 1   k  y   − y Xk=2     Nielsen [14, band I, p. 68] derives

1 ∞ m k (m) 1 m = ( 1) − k k+1 (44) k=0 (z + k) − S z kX=m Q where (n) denotes the Stirling Number of the Second Kind [1, Sec. 24.1.3 and 24.1.4], which we use Sk in the right-hand side of (43)

∞ mbm(α) ∞ ∞ m k (m) 1 = mb (α) ( 1) − m (y + k) m − Sk yk+1 m=1 k=0 m=1 X X kX=m Q k ∞ m k (m) 1 = mb (α) ( 1) − m − Sk yk+1 Xk=1 mX=1 Equating corresponding powers in 1 of both sides in (43) yields, for k 1, y ≥

1 k+1 k k+2 k+2 α αk+1 (α 1) m k (m) α (α 1) 2 mb (α) ( 1) − = − − + − − − (45) m − Sk k + 2  k + 1  m=1  X Finally, after multiplying both sides in (45) by ( 1)k S(k), summing over k [1, j], we have − j ∈

1 k+1 j k j k+2 k+2 α αk+1 (α 1) m (k) (m) α (α 1) 2 − − − k (k) mbm(α) ( 1) S = − − + ( 1) S − j Sk  k + 2   k + 1  − j Xk=1 mX=1 Xk=1   We reverse the k- and m-summation in the double sum at the left-hand side 

j k j j mb (α) ( 1)m S(k) (m) = mb (α) ( 1)m S(k) (m) m − j Sk m − j Sk m=1 m=1 ! Xk=1 X X kX=m invoke the second orthogonality relation for the Stirling numbers [1, sec. 24.1.4]

j S(k) (m) = δ (46) j Sk mj kX=m and obtain j k mb (α) ( 1)m S(k) (m) = jc (a) ( 1)j, which demonstrates (39). k=1 m=1 m − j Sk j − The correspondingP P integral representation of the coefficient bm(α) is translated, via (38), as

α m 1 − 1 − mbm(α)= u α (k u)du (α 1) − 2 − − Z− −    kY=0 After substitution of x = u, we arrive at (41).  −

16 We now discuss implications of Theorem 3. There are two particularly interesting cases of the

Binet polynomial bm(α) in (39): for α = 0, m 1 k b (0) = c = ( 1)m S(k) m m 2m (k + 1) (k + 2) − m Xk=1   but α = 1 leads to the original Binet coefficients (7), m 1 k k m (k) b (1) = β = ( 1) − S m m 2m (k + 1) (k + 2) − m Xk=1   k j (k) Since ( 1) − S is a non-negative integer, it follows that the original Binet coefficients β = b (1) − j m m are all positive, in contrast to bm(0) = cm in Theorem 2, whose sum (32) is alternating and does not obviously to lead to conclusions about the sign. As mentioned earlier, any truncation at m = K > 2 terms in (31) upper bounds Binet’s function µ (z), whereas any truncation of m = K terms in Binet’s original expansion (6) lower bounds µ (z). Hence, for any finite integer K˙ > 2, it holds that

K K b (1) b (0) m <µ (z) < m m (z + k) m 1 m=1 k=1 m=1 k=0− (z + k) X X which suggests that there may existQ a tighter value of α betweenQ 0 and 1, explored in Section 5. It follows from (39)

1 k+1 m m k+2 k+2 α αk+1 (α 1) ( 1) α (α 1) 2 − − − k (k) bm(α)= − − − + ( 1) Sm m  k + 2   k + 1  − Xk=1   that   1 k+1 m m k+2 k+2 α αk+1 (α 1) ( 1) α (α 1) 2 − − − (k) bm(1 α)= − − − + Sm − m  k + 2   k + 1  Xk=1   m k (k) 1 illustrating, with ( 1) − Sm 0, the absence of symmetry around α = . A second observation of − ≥ 2 (39) for bm(α) and the fact that Stirling numbers are integers is that, if α Q is a rational number, l ∈ i.e. α = k for integers l and k, then the Binet polynomial bm(α) is also rational.

3.1 Properties of the Binet polynomial bm (α)

Property 1 The Binet polynomial b (α) in (39) is a polynomial of degree m 1 in α, m − m 1 − j bm(α)= pj (m) α (47) Xj=0 where the coefficients

j m 1 j m 1 d b (α) ( 1) − − k! (k j) p (m)= m = − − S(k) (48) j j! dαj 2m j! (k + 2 j)! m α=0 k=j+1 X − 1 1 m 1 from which pm 1 (m)= and pm 2 (m)= − . An integral form is − 12 − 12 2 m 1 1 0 1 dj − pj (m)= u + j (k + u)du (49) mj! 1 2 du Z−   kY=0 17 Proof: Substitution of

k+2 k+2 1 k+1 1 k+1 k α (α 1) 2 α α 2 α (α 1) 1 k! (k j) k j j − + − − − = − ( 1) − α k + 2 − k + 2 k + 1 − k + 1 −2 j! (k + 2 j)! −   j=0 X − (0) into the Binet polynomial (39) and using Sm = 0 for m> 0 yields

m k 1 k! (k j) k j j k m (k) b (α)= − ( 1) − α ( 1) − S m −2m  j! (k + 2 j)! −  − m Xk=0 Xj=0 − 

We reverse the j- and k-sum, verify that pm (m) = 0, and arrive at (47) and (48). The integral form (49) is immediate from the integral (41) of b (α) after substitution of u = x α m − as j 0 j m 1 1 d bm(α) 1 1 d − pj (m)= j = u + j (k + u + α) du j! dα α=0 mj! 1 2 dα Z−   k=0 α=0 Y dj m 1 dj m 1 because dαj k=0− (k +u+α)= duj k=0− (k +u+α). Introducing the j-th derivative of the generating 1 dj m 1 m k (k) m k k j function (35), j − (k + u)= Sm ( 1) − u , into (49) alternatively leads to (48). Q j! du k=0 Q k=j j − −  Q P 

Clearly, if α = 0, then we find the coefficients cm = bm (0) = p0 (m) in (32) of Binet’s second factorial expansion again.

Corollary 1 The n-th derivative of Binet’s function µ (z) is

n n d bm(α) d µ (z) ∞ n = ( 1)n dα (50) dzn − m 1 m=n k=0− (z + α + k) X Proof: Taking the derivative of (10) with respectQ to α yields

∞ 1 dbm (α) ∞ d 1 0= m 1 + bm (α) m 1 k=0− (z + α + k) dα dα k=0− (z + α + k) mX=1 mX=1 d 1 Q d 1 Q and since dα m−1 = dz m−1 , we conclude that k=0 (z+α+k) k=0 (z+α+k) Q Q dbm(α) dµ (z) ∞ = dα dz − m 1 m=1 k=0− (z + α + k) X n Q d bm(α) Repeating the argument for an integer n 0, using n = 0 for n>m by Property 1, leads to ≥ dα (50). 

Substituting the polynomial (47) for bm (α) in Property 1 into (10) indicates, with b0 (α)=0 and pj (0) = 0, that

m j ∞ pj (m) α ∞ ∞ p (m) µ (z)= j=0 = j αj m 1  m 1  m=0 Pk=0− (z + α + k) k=0− (z + α + k) X Xj=0 mX=j Q  Q 

18 Replacing α z z and using p (m) = 0 yields the Taylor series of µ (z) around z , → 0 − m 0

∞ j ∞ pj (m) j µ (z)= ( 1) m 1 (z z0)  − k=0− (z0 + k) − Xj=0 mX=j+1   Q 1 dj µ(z) The derivatives of Binet’s function for Re (z) > 0 follow from the Taylor coefficient j! dzj or z=z0 from (50) and from (3),

j j t d µ (z) j ∞ pj (m) ( 1) ∞ j 1 zt 1+ e− 2 j = ( 1) j! m 1 = − t − e− t dt (51) dz − k=0− (z + k) 2 0 1 e− − t mX=j+1 Z  −  Q m 1 j Property 2 The coefficients pj (m) of the Binet polynomial bm(α)= j=0− pj (m) α can be expressed in terms of the coefficients cm = bm (0) in (32) as P m 1 − m 1 (j) l j pj (m)= − S ( 1) − cm l (52) l l − − Xl=j   Proof: Introducing the generating function (35) in (40) yields

m m k 1 m m k m 1 − − m 1 − (j) m k j j bm (α)= − (α + j)ck = − ck Sm k ( 1) − − α k 1 k 1 − − Xk=1  −  jY=0 Xk=1  −  Xj=0 Letting l = m k, reversing the sums, − m 1 m 1 − − m 1 (j) l j j bm (α)= − cm lS ( 1) − α  l − l −  Xj=0 Xl=j     m 1 j  and comparing with the definition bm(α)= j=0− pj (m) α in (47) leads to (52). We split the sum (52), P

m 3 1 (j) m 1 j − m 1 (j) l j pj (m)= Sm 1 ( 1) − − − cm l Sl ( 1) − 12 − − − l | − | − Xl=j   which is the difference of two positive numbers that turns out to be positive for j > 0. This observation (0) is further substantiated in Property 3 below. If j = 0, then Sl = δ0l and the first positive term vanishes for m > 1 and p0 (m) = cm < 0. If we substitute the explicit form (32) of cm into (52), we again arrive at (48) after using the formula6

j (k+m) m!k! j (m) (k) Sj = Sl Sj l (53) (k + m)! l − Xl=0   Property 3 Except for one negative coefficient, p0 (m)= cm for m> 2, all other coefficients pj (m) m 1 j of the Binet polynomial bm(α) = j=0− pj (m) α are positive, i.e. pj (m) > 0 for j > 0 and m > 2. Generally, for j > 1, it holds thatP m j ( 1) − (j 2) m (j 1) pj (m)= − (m j + 1) Sm−1 + (j 1) 1 Sm−1 (54) j (j 1) m − − − 2 − − −     6Equate corresponding powers of the Taylor series in logk+m(1 + u) = logm(1 + u) logk(1 + u) from the second (k) generating function of the Stirling numbers Sm in (36).

19 and also m j ( 1) − (j 1) j 1 (j 1) pj (m)= − (m j + 1) Sm− + m m Sm−1 (55) j (j 1) m − − 2 − 2 − −     1 2 1 2 m 2 1 In particular, p (m)= (m 2)! 1 and p (m)= (m 2)! 1 − . 2 4 − − m 3 6 − − m l=2 l   P Proof: The derivative of the integral representation (41) of bm(α) is

m 2 m 1 db (α) m − α − m m = 1+ α α (k + α) (k + x)dx dα 2 − − (α 1)   kY=1 Z − kY=0 from which for α = 0,

m 1 db (α) 1 1 − p (m)= m = u (k u)du > 0 1 dα m − α=0 Z0 k=1 Y

An additional derivation yields

2 m 2 m 1 m 1 d bm (α) d m − − − m 2 = 1+ α (k + α) (k + α)+ (k + α 1) dα dα 2 − ! − −   kY=0 kY=0 kY=0 df(x) d log f(x) After employing the logarithmic derivative dx = f (x) dx , we find

m 2 m 2 m 2 d2b (α) m − − m − m m = 1 (k + α)+ α 1 j (k + α) dα2 2 − 2 − − k=1 j=1 k=1;k=j   Y X   Y6 2 1 d bm(α) m 2 from which, for m 2, it follows that p2 (m) = 2 = (m 2)! − > 0. We may ≥ 2 dα α=0 − 4m continue this tedious process of differentiations to discover closed expressions for other pj (m) with 3  d bm(α) m 2 m m 2 j > 2. For example, from m dα3 = j=0− 2 1 j l=0;l=j k=0;− k= j,l (k+α), we find p3 (m)= 1 2 m 2 1 − − 6 6 { } (m 2)! 1 − , but that process essentially boils down to computing the Stirling numbers 6 m l=2 l P  P Q (k) − − Sm in closed form. P The first derivative still contains an integral, while higher order derivatives are sum of products. n d bm(α) Instead of computing the derivatives n for n> 1 from the integral representation (41), dα α=0 dnµ(z) they can be deduced more elegantly from the derivative dzn in (50) and from Binet’s integral (3) as

n t d µ (z) n 1 ∞ n 1 zt 1+ e− 2 = ( 1) t − e− dt dzn − 2 1 e t − t Z0  − −  We mimic Binet’s method in the proof of Theorem 1 and invoke Binet’s substitution e t = 1 u, − − n 1 z 1 d µ (z) 1 (1 u) − n 1 n 2 n = − (2 u) (log (1 u)) − + 2u (log (1 u)) − du dz −2 0 u − − − Z   (k) Since n> 1, we now use the second generating function (36) of the Stirling numbers Sm and obtain the Taylor series, convergent for u < 1, of the integrand | | n 1 n 2 h (u)=(2 u) (log (1 u)) − + 2u (log (1 u)) − − − − (n 1) m m ∞ Sm − (n 1) (n 2) ( 1) u = 2 (n 1)! + (n 1)!Sm −1 2 (n 2)!Sm −1 − − m − − − − − (m 1)! m=n 1 ( ) X− −

20 Substituting this Taylor series into the integral, reversing the integration and summation, invoking Γ(z+m) m 1 the Beta integral and Γ(z) = k=0− (z + k), leads with (50) for n> 1 to

n Q (n 1) d bm (α) m n 1 (n 1) Sm − (n 2) (n 1) (n 1) = ( 1) − − (n 2)! − S − + − S − dαn − − m − m 1 2 m 1 α=0 − − !

j 1 d bm(α) (n 1) (n 2) − − from which pj (m) = j! dαj in (54) and (55) follow, after eliminating Sm and Sm 1 re- α=0 − (n ) (n 1) (n) spectively by the recursion Sm +1 = Sm − mSm (see e.g. [1, 24.1.3.II.A]). We may verify, by − computation of (54) and (55), that pj (m) > 0 for j > 1 and m> 2.  The integral (41) for the Binet polynomial b (α) becomes after substitution u = x + 1 α and m 2 − reduction of the integration interval 1 , 1 to 0, 1 , − 2 2 2 1 m 1 m 1 1 1  2  −   − bm α + = u (k + α + u) (k + α u) du (56) 2 m 0 ( − − )   Z kY=0 kY=0 1 m 1 If α> 0, then all coefficients are positive, i.e. bm α + 2 > 0 for m 1, because k=0− (k + α + u) > m 1 ≥ − (k + α u). For larger negative α < 0, the Binet polynomial b (α) starts oscillating and k=0 −  m Q determining the sign is more complicated. The asymptotic behavior of b α + 1 for large m is Q m 2 analyzed in Appendix E.  Property 4 The Binet polynomial m 1 m 1 − 1 − b (α)= p (m) αj = (α ξ (m)) m j 12 − j Xj=0 jY=1 has m 1 real, distinct zeros ξ1 (m) >ξ2 (m) > >ξm 1 (m). − · · · − Proof: We apply the generalized mean-value theorem7 [10, p. 321] to the integral (56) for α 0, ≥ m 1 m 1 1 1 − − b + α = (k + α + θ ) (k + α θ ) (57) m 2 8m m − − m k=0 k=0 ! 1   0<θm< 2 Y Y h h which is the central difference δhf(x)= f x + 2 f x 2 with step h = 2 θm < 1 of the polynomial m 1 − − f (x)= − (k + x) of degree m in x with real zeros at the integers x = k, for 0 k m 1. In a k=0   − ≤ ≤ − region containingQ the distinct zeros, the polynomial f (x) oscillates below and above the real axis, as well as its shifted companion f (x + h) with step h smaller than the distance between the zeros. This means that f (x) and f (x + h) will intersect m 1 times at distinct points, implying that b (α) has − m m 1 real zeros in the interval 1 m, 1 .  − 2 − 2 m 1 m 1 1 m 1 The sum of the zeros equals − ξ (m)= − and their product c = − ( ξ (m)). j=1 j − 2 m 12 j=1 − j We found that the Binet polynomial b (α) has a “center zero” equal to ξ (2m) = (m 1) and P 2m  m Q− − that ξm 1 (m) > m 1. Since, for m > 2, all coefficients pj (m) > 0 for j > 0 and p0 (m) < 0 and − − − all zeros are real, we conclude [25, art. 218, p. 289] that the largest zero is positive, ξ1 (m) > 0, while all others are negative, ξj (m) < 0 for j > 1.

We end this section with Property 5 that relates the Binet polynomials bm (α) to Bernoulli poly- nomials Bn (α), 7 ∈ b b If ϕ (x) is non-negative for x [a,b], then a f (x) ϕ (x) dx = f (θ) a ϕ (x) dx for some θ obeying a<θ 0 can be expressed in terms of Bernoulli polynomials as

m 1 (k) k m 1 − Sm 1 ( 1) k+2 1 k+1 b (α) = ( 1) − − − B (α) α + (k + 2) α (58) m − (k + 1) (k + 2) k+2 − 2 Xk=0   Proof: After substituting Nielsen’s expansion (44) into Binet’s convergent series (31)

∞ bm (α) ∞ ∞ m k (m) 1 µ (z)= = b (α) ( 1) − m 1 m+1 − Sk k+1 m=1 k=0− (z + α + k) m=0 (z + α) X X kX=m 1 n n k 1 Using = ∞ Q( α) − in [1, 24.1.1.B] leads to (z+α)k+1 n=k k − zn+1 P  ∞ ∞ m k (m) ∞ n n k 1 µ (z)= b (α) ( 1) − ( α) − m+1 − Sk k − zn+1 m=0 k=0 n=k   X X n X ∞ ∞ n m+n (m) n k 1 = bm+1 (α) ( 1) k α − n+1 k − S ! z mX=0 Xn=0 Xk=0   and, since (m) =0 if m > k, to Sk n k ∞ n m+n (m) n k 1 µ (z)= b (α) ( 1) α − k m+1 − Sk zn+1 n=0 (m=0 )! X Xk=0   X 1 Equating corresponding powers of z in the above and in Stirling’s asymptotic series µ (z) = Bn+1 n∞=1 n(n+1)zn in (23) indicates that

P n 1 n n k k Bn+2 ( 1) n 1 − α = b (α) ( 1)m (m) (n + 1) (n + 2) k α m+1 − Sk  k=0 m=0 X    X Binomials inversion [17, chap. 2], a = n n b b = ( 1)n n ( 1)k n a , yields n k=0 k k ⇔ n − k=0 − k k k k P  k P 1 l 1 k Bl+2 b (α) ( 1)m (m) = ( 1)k α α m+1 − Sk − l (l + 1) (l + 2) m=0    X Xl=0   and k k k B αk l b (α) ( 1)m (m) = ( 1)k l+2 − m+1 − Sk − l (l + 1) (l + 2) m=0 X Xl=0   k k+2 ( 1) k + 2 k+2 l = − B α − (k + 1) (k + 2) l l Xl=2   n n n l 8 With the definition [1, 23.1.7] of the Bernoulli polynomials, Bn(α)= l=0 l Blα − , we obtain k ( 1)k P 1  b (α) ( 1)m (m) = − B (α) αk+2 + (k + 2) αk+1 (59) m+1 − Sk (k + 1) (k + 2) k+2 − 2 mX=0   8 S(m) The Bernoulli numbers can be written in terms of the Stirling numbers k of the second Kind [4, p. 220], k − m S(m) ( 1) m! k Bk = for k> 0 m + 1 mX=1

22 Formula (59) expresses the Bernoulli polynomial Bn (α) in terms of the Binet polynomial bm (α). (k) We invert relation (59) to find the Binet polynomial bm (α). After multiplying both sides by Sj , summing over k [0, j], we have ∈ j k j (k) k S ( 1) 1 b (α) ( 1)m S(k) (m) = j − B (α) αk+2 + (k + 2) αk+1 m+1 − j Sk (k + 1) (k + 2) k+2 − 2 m=0 Xk=0 X Xk=0   Reversing the k- and m-sum and applying the second orthogonality formula (46) yields

j k j j m (k) (m) m (k) (m) j bm+1 (α) ( 1) Sj k = bm+1 (α) ( 1) Sj k = bj+1 (α) ( 1) − S − S ! − Xk=0 mX=0 mX=0 kX=m from which Property 5 follows. 

4 Digamma and polygamma functions

Γ′(z) We present the corresponding factorial series for the digamma function ψ (z) = Γ(z) and for the (n) dn+1 ln Γ(z) (0) polygamma function, defined as ψ (z)= dzn+1 with ψ (z)= ψ(z). We confine ourselves to the case α = 0. Γ′(z) Differentiation of (1) with respect to z expresses the digamma function ψ (z) = Γ(z) in terms of the Binet function µ (z) as 1 ψ (z) = log z + µ′ (z) − 2z Introducing the factorial expansion (51) for j = 1 gives

1 ∞ p (m) ψ (z) = log z 1 (60) − 2z − m 1 m=2 k=0− (z + k) X Explicitly, Q

1 1 1 19 ∞ p1 (m) ψ (z) = log z m 1 − 2z − 12z (z + 1) − 12z (z + 1) (z + 2) − 120z (z + 1) (z + 2) (z + 3) − k=0− (z + k) mX=5 is the convergent companion for Re (z) > 0 of the asymptotic series [1, 6.3.18] Q

1 ∞ B 1 1 1 1 ψ (z) log z + 2m = log z + + (61) ∼ − 2z 2mz2m − 2z − 12z2 120z4 − 252z6 · · · m=1 X K p1(m) Since p1 (m) > 0, truncation of m=2 m−1 in (60) after any K 2 provides an upper bound for k=0 (z+k) ≥ n 1 1 the digamma function ψ (z). ForP integerQ z = n in (60) for which ψ (n)= − γ (see [1, 6.3.2]), k=1 k − the harmonic series is P n K 1 1 (n 1)!p (m) = log n + γ + + lim − 1 (62) k 2n K (n 1+ m)! →∞ m=2 Xk=1 X − and the m-sum converges rapidly, also for relatively small n, but increasingly fast for larger n. For 6 example, with K = 5 terms evaluated in (62), the error is less that 10− for n = 10 and less than 11 10− for n = 100.

23 dbm(α) Since bm (α) and the derivatives dα contain integrals as illustrated in the proof of Property 3, the functional regime of ψ(n)(z) for n> 2 is different than for n 2. Consequently, asymptotic series ≤ for n> 2 disappear and convergent factorial series loose their attractiveness, because convergent exist. Indeed, in terms of the Binet function and starting from ψ(0)(z)= ψ (z) = log z 1 +µ (z), − 2z ′ it holds for n 1 that ≥ n+1 (n) n 1 n 1 n 1 n 1 d ψ (z) = ( 1) − (n 1)!z− + ( 1) − n!z− − + µ (z) − − 2 − dzn+1 Introducing the factorial expansion (51), valid for n 1, yields ≥

(n) n 1 1 1 n ∞ pn+1 (m) ψ (z) = (n 1)! ( 1) − + + (n + 1) n − − zn 2 zn+1 m 1 m=n+2 k=0− (z + k)! X The factorial series for ψ(n)(z) converges slower for larger n, is more complicatedQ and only valid for (n) n 1 1 Re (z) > 0 in contrast to ψ (z) = n!( 1) ∞ in (30) that converges for all complex z − − k=0 (z+k)n+1 (except for the poles at z = k for integer k 0). − ≥P

5 Stirling’s asymptotic and Binet’s generalized factorial series

In this section, we will compare the accuracy of the generalized Binet factorial expansion (10) in terms of the error K bm (α) eα (z, K)= µ (z) m 1 − − (z + α + k) m=1 k=0 X which is a function of z, the “free” parameter α andQ the number K of terms evaluated. We assume here that K is finite. Similarly, we denote the error of Stirling’s asymptotic series (23) by

K B2m eStirling (z, K)= µ (z) 2m 1 − (2m 1) (2m) z − m=1 − X We are interested in the “best” value α of the parameter α for which the error e (z, K) is minimal. ∗ α The asymptotic, diverging nature of the Stirling approximation allows us to compute the number K (z) of terms that minimizes the error, i.e. for any real z, e (z, K) e (z, K (z)). Thus, ∗ Stirling ≥ Stirling ∗ we will take the best possible performance with minimal error eStirling (z, K∗ (z)) as a benchmark to compare eα (z, K∗ (z)) as a function of z and α. In summary, Stirling’s asymptotic expansion (23)

K 1 1 1 B µ (z) + + 2m (63) ≃ 12z − 360z3 1260z5 (2m 1) (2m) z2m 1 m=4 − X − can be compared to the generalized Binet factorial series (10) with a same number K of terms,

K 1 α b (α) µ (z) + + m (64) ≃ 12(z + α) 12(z + α)(z + α + 1) m 1 m=3 k=0− (z + α + k) X In particular, for α = 0 and bm (0) = cm, where c2 = 0, Q

K 1 1 1 c µ (z) + m (65) ≃ 12z − 360z (z + 1) (z + 2) − 120z (z + 1) (z + 2) (z + 3) m 1 m=5 k=0− (z + k) X Q 24 we observe that the first two coefficients in Stirling’s asymptotic (63) and Binet’s convergent (65) expansion are the same. Moreover, Stirling’s asymptotic (63) has alternating terms – the Bernoulli m 1 numbers B = ( 1) − B alternate –, in contrast to (64), where b (α) changes sign at most 2m − | 2m| m once with increasing m. While Stirling’s expansion (63) is an asymptotic and approximate series with a best possible, non-zero error eStirling (z, K∗ (z)), Binet’s factorial, convergent series (64) can always beat the accuracy eStirling (z, K∗ (z)) for any Re (z) > 0 if the number K of terms is sufficiently large. Therefore, we investigate whether Binet’s series (64) with the same number K∗ (z) of terms can achieve a similar accuracy as Stirling’s expansion (63) with optimal number K∗ (z) of terms. Fig. 1 !" "

#" 6$" 783

$" "

6#" (24484-./-&'()*-/2495* "" 6!" %"

&'()* 6%" !" 6 "" #" -&'()*-:-";,##-<-,; ,=-) -783 "(2>/04?013()@&'()***-:-6 ; #-6-$;=#-) -783 "(2"()@&'()***-:-6$;==-6- ;",-) 6 $" $" -783 "(2 ()@&'()***-:-6 ; %-6- ;""-) " 6 #" + " + $" $+ ," ,+ #" #+ +" )-(./-01/23245*

Figure 1: Optimal accuracy of Stirling’s asymptotic expansion: the optimal number of terms K∗ (z)

(left axis) to achieve the lowest error eStirling (z, K∗ (z)) (right axis). The error e0 (z, K∗ (z)) and e1 (z, K∗ (z)) of Binet’s two factorial expansions for α = 0 and α = 1 are added (right axis). The legend shows linear fits of the data. shows the logarithm (in basis 10) of the error (right axis), evaluated at the optimal number of terms

K∗ (z) versus z (left axis). The comparison of Stirling’s asymptotic series with Binet’s two factorial expansions (6) for α = 1 and (31) for α = 0 clearly illustrates the amazing superiority of Stirling’s asymptotic series. Fig. 2 illustrates that the logarithm of the error e (z, K (z)) versus α [ 2, 2] varies considerably. α ∗ ∈ − In particular around the zeros ξ (m), ξ (m) and ξ (m) of the polynomial b (α) in the interval [ 2, 2], 1 2 3 m − the error of the generalized Binet factorial expansion (64) decreases sharply. The minimal errors, peaked around the zeros, only shift little in α for various K∗ (z) and z. Numerically for m> 10, we found that the largest and only positive zero ξ (m) [0.08, 0.1] and 1 ∈

25 ! & -454!

% -454& $

#" -454$ #! ,-./0,-11 + * #& #" #% '() #$ -454#% !" !! !2" #23 #2" "23 "2" "23 #2" #23 !2"

Figure 2: The logarithm of the error e (z, K (z)) versus α for various z = 2, 4, 8, 16 . α ∗ { }

that ξ1 (m) attains its largest value 0.0963016 at m = 72 and ξ1 (m) slowly decreases for m> 72. For any finite K, there exists a value of α around ξ1 (K) that minimizes the error

K K pj (m) j eα (z, K)= µ (z) m 1 α −  − (z + α + k) j=0 m=j+1 k=0 X X  KQ m 1  1 − α ξ (m) = µ (z) − j − 12 (z + α) z + α + j m=1 j=1 X Y

Given z, the last form is related to a Pad´eapproximant of order [m 1/m + 1] in α. − If K = K (z), then we can find a value α that has a comparable error e (K) than Stir- ∗ ∗ | α | ling’s asymptotic approximation. For z = 10 and α 94909394316015843 , 94909394316015845 , the error ∈ 1018 1018 log e (10, K (10)) < 30, while Stirling’s lowest error is log e (10, K (10)) = 28.5834. 10 | α ∗ | −  10 | Stirling ∗ | − The myth that Stirling’s approximation is always better than any factorial, convergent expansion for

µ (z) with a same number of terms evaluated is not true, as illustrated by this counter example. If α∗ is approximated by a rational number, then all coefficients of bm (α) are rational numbers, just as the Bernoulli numbers in the Stirling approximation. The major advantage of Binet’s series (64) over Stirling’s asymptotic (23) lies in its convergence, for all Re (z) > 0 and Re(α) > Re (z), towards µ (z), which allows incorporation into integrals and − series and may lead to other sharp bounds and approximations. Moreover, the “free” parameter α can be tuned to achieve a similar accuracy as the best accuracy of Stirling’s asymptotic series (63). Finally, the coefficients b (α) are monotonously increasing in m> 2 and, for α>ξ (m), b (α) is | m | 1 m positive, but only changes the sign once for ξ2 (m) <α<ξ1 (m). For finite K, it might be interesting

26 to know the smallest possible error eα (z, K) after optimization of α. Since K is finite, we expect generally that eα (z, K) > 0 is bounded from below, else a finite number of computations would result in an exact computation the µ (z) for all Re (z) > 0.

6 Factorial series for Laplace transforms

The Laplace transform of a real function f (t) is defined (see e.g. [21], [7, Chapter VII], [29]) for complex z as ∞ zt ϕ(z)= [f(t)] = e− f(t)dt (66) L Z0 with the inverse transform,

c+i 1 1 ∞ zt f(t)= − [ϕ(z)] = ϕ(z)e dz (67) L 2πi c i Z − ∞ where c is the smallest real value of Re(z) for which the integral in (66) converges. Many functions can be defined by an integral as (66) as well as the probability generating function [26, Sec. 2.3.3] of a continuous random variable. For example, Binet’s integral (3) is a Laplace transform (66), where 1 1 1 1 1 ϕ(z)= µ (z) and the integrand f (t)= t et 1 t + 2 with f (0) = 12 . − − Factorial series are hardly studied. Temme has written a literature overview [20], of which parts were incorporated by Lauwerier in his book [12, p. 33-45], that devotes one chapter to factorial series. Parts of the content of [20] and [12, p. 33-45] are here absorbed. Apart from reviewing the literature more extensively than here, Delabaere and Rasoamanana [5] have presented similar results, but their method and exposition is rather different. In this Section 6, we generalize the idea of Binet’s proof of Theorem 1 as far as possible.

6.1 The analogon of Stirling’s asymptotic series

k k 1 d f(t) If we assume that the Taylor series f (t)= k∞=0 fkt , with Taylor coefficients fk = k , has k! dt t=0 an infinite radius of convergence, then the Laplace transform (66) can be expanded as P

∞ zt ∞ ∞ zt k ϕ(z)= e− f(t)dt = fk e− t dt 0 0 Z Xk=0 Z into a Laurent series [22, Sec. 2.7] ∞ k!f ϕ(z)= k (68) zk+1 Xk=0 An infinite radius of convergence implies that the Taylor coefficient fk converges faster to zero than any , i.e. f = o e ak for any finite a and k , and that f (t) is an entire k − → ∞ function. Most functions, however, are not entire functions. If the requirement of an infinite radius of K k!fk convergence is ignored, then (68) may represent an asymptotic series ϕ(z)= k=0 zk , which diverges when K . Indeed, repeated partial integration of the Laplace transform (66) yields →∞ P K k!fk 1 ∞ zt (K+1) ϕ(z)= k + K+1 e− f (t)dt (69) z z 0 Xk=0 Z

27 K 9 f ( +1)(t) illustrating , for any Re (z) > c but z = 0 and real t 0, that limK K+1 must vanish to obtain 6 ≥ →∞ z a convergent Laurent series (68).

6.2 One particular factorial series

We formalize Binet’s proof of Theorem 1 as a recipe with five steps. First, we make Binet’s substitution e t = 1 u or t = log (1 u) in the integral (66) and obtain − − − − 1 z 1 ϕ(z)= (1 u) − f( log (1 u))du (70) − − − Z0 which converges for Re (z) > c. Binet’s substitution is rather unusual, certainly in the study of Laplace integrals. In the early days, Euler represented the Gamma function in the form of (70), after letting 1 z 1 1 w = 1 u, as Γ (z) = log − dw (see e.g. [2, art. [1]]). Lauwerier [12, p. 33-35] and Temme − 0 w [20] explain the success of the substitution u = 1 e t by comparing u = reiθ confined to the unit R  − − disk u 1 < 1 at u = 1 and the map t = log 1 reiθ . In particular, the circle u 1 = 1 at | − | 0 − − | − | u0 = 1 with radius 1 maps into the curve  − i arccos 1 cos θ t = log (1 cos θ i sin θ)= log 2 (1 cos θ)e− √2(1−cos θ) − − − − −   θ θ p θ π θ = log 2sin + i arccos sin = log 2sin + i − − 2 2 − 2 2         θ π θ Hence, Re t = log 2sin and Im t = − with θ [0, 2π]. Elimination of θ = π 2Im t yields − 2 2 ∈ −  π π Re t = log (2 cos (Im t)) for Im t (71) − − 2 ≤ ≤ 2 The curve (71) is symmetric around the Re t-axis due to cos (Im t) = cos ( Im t) and cos (Im t) 0 − ≥ confines the curve (71) to the strip π Im t π , where the minimum occurs at log 2 for Im t = 0 − 2 ≤ ≤ 2 − and Re t grows boundlessly if Im t π . Thus, the map t = log (1 u) of the unit disk u 1 < 1 →± 2 − − | − | appears as the interior t-region bounded by the curve (71). That t-region is considerably broader than the unit disk, which accounts for better results in the sense that the resulting factorial series in (73) below converges for more functions than its corresponding Laurent series (68).

The second step involves the Taylor expansion of f( log (1 u)) around u0 = 0, where log (1 u0)= t − − k − 0. After Binet’s substitution e− = 1 u, the Taylor series f (t)= k∞=0 fkt becomes f ( log (1 u)) = k − − − ∞ f ( log (1 u)) . Introducing the second generating function (36) of the Stirling numbers and k=0 k − − P Preversing the m- and k-sum leads to m k m ∞ d f(t) m k (k) u f( log (1 u)) = ( 1) − S (72) − − dtk − m m! m=0 k=0 t=0 ! X X

Clearly, the Stirling numbers, which are integers, play a ke y role in Binet’s transformation u = 1 e t. − − 9By the generalized mean-value theorem [10, p. 321], there exists a positive real θ for which ∞ ∞ − − 1 e ztf (K+1)(t)dt = f (K+1)(θ) e ztdt = f (K+1)(θ) Z0 Z0 z

28 Crucially for the third step, we assume that the Taylor series (72) converges for u < 1. Hence, | | the radius of convergence of the Taylor series (72) should be at least equal to one. After substitution of the Taylor series (72) in the integral (70) and reversing the summation and integration operator, justified because a Taylor series can be term-wise integrated within its radius of convergence, yields

m k 1 ∞ d f(t) m k (k) 1 m z 1 ϕ(z)= ( 1) − S u (1 u) − du dtk − m m! − m=0 k=0 t=0 ! Z0 X X

1 p 1 q 1 Γ(p)Γ(q) The fourth step uses the Beta integral u (1 u) − du = , valid for Re (p) > 0 and 0 − − Γ(p+q) Re (q) > 0, and leads to R

m k ∞ d f(t) m k (k) Γ (z) ϕ(z)= ( 1) − S dtk − m Γ (m +1+ z) m=0 k=0 t=0 ! X X

Γ(z+m) m 1 The fifth and final step replaces Γ(z) = k=0− (z + k) and we arrive at the factorial expansion, valid for Re (z) > c, Q m dkf(t) m k (k) ∞ k=0 dtk ( 1) − Sm ϕ(z)= t=0 − (73) m (z + k) m=0 P k=0 X The factorial expansion (73) generalizes Binet’s secondQ factorial expansion (31) in Theorem 1.

6.3 Infinitely many factorial series for ϕ(z)

We proceed to generalize the recipe with five steps in Section 6.2, as we did for the particular case ϕ(z)= µ (z) in Theorem 3, together with an additional β-scaling inspired by Temme [20, p. 11]. Our main result is:

Theorem 4 Only if the Taylor series

m m ∞ k m k (k) u f( β log (1 u)) = k!f β ( 1) − S (74) − − k − m m! m=0 ! X Xk=0 has a radius of convergence at least equal to 1, then the Laplace transform ϕ(z) of the function f (t) possesses infinitely many factorial series in the complex parameter α and real β > 0, for Re (z) > c and Re(α) > Re (z), β − ∞ m!φ (α, β) ϕ(z)= β m (75) m (βz + α + k) m=0 k=0 X where Q m m k k 1 − (k+j) m (k+j) j α φ (α, β)= (k + j)!S ( 1) − f β (76) m m!  m − j  k! Xk=0 Xj=0  f(0) m Further, φm (α, β) is a polynomial of degree m in α with highest order term m! α and φ0 (α, β) = f(0). A complex integral for φm (α, β), where the contour C (0) encloses the point ω0 = 0, is

( 1)m f( β log (1 + ω))dω φ (α, β)= − − (77) m 2πi (1+ ω)α ωm+1 ZC(0)

29 while another compact form is

m j 1 d α t (j) m j j φ (α, β)= f(t)e β S ( 1) − β (78) m m! dtj m − j=0 t=0 X  

Clearly, the factorial series in (75) with (78) reduces to (73) for α = 0 and β = 1. Proof: We repeat the five steps in Section 6.2, but rewrite the Laplace transform (70) of ϕ(z) after a generalization of Binet’s substitution e t = (1 u)β or t = β log (1 u) as − − − − 1 βz+α 1 α ϕ(z)= β (1 u) − (1 u)− f( β log (1 u))du − − − − Z0 which still converges for Re (z) > c, in spite of the introduction of the “free” parameter α and the real β > 0. The second step now involves the Taylor expansion of

α ∞ m g (u; α, β) = (1 u)− f( β log (1 u)) = φ (α, β) u (79) − − − m m=0 X around u0 = 0. The Taylor series f( β log (1 u)) in (74) follows from (72). The Taylor series α α k k − − (1 u)− = k∞=0 −k ( 1) u is valid for any complex α provided u < 1. Thus, the radius of − − α | | convergence of (79) is limited to 1 by (1 u)− , which is just sufficient for the reversal of summation P  − and integration in step three, provided the radius of convergence of g (u; 0, β) = f( β log (1 u)) − − in (72) is at least equal to 1. From Cauchy’s integral theorem [22] we directly find the integral representation (77). In addition, the Taylor coefficient φm (α, β) in (79) follows from the Cauchy product m m j j α ( 1) − l j l (l) φ (α, β)= − − l!f β ( 1) − S (80) m m j j! l − j Xj=0  −  Xl=0 The remaining steps in the method of Section 6.2

1 ∞ βz+α 1 m ∞ m!Γ (βz + α) ϕ(z)= β φ (α, β) (1 u) − u du = β φ (α, β) m − m Γ (βz + α + m + 1) m=0 0 m=0 X Z X lead to the factorial expansion (75), valid for Re (z) > c and Re(α) > Re (z). β − The remainder of the proof consists of simplifying the Taylor coefficient φm (α, β) in (80). It is convenient to reverse the summations,

m m m l Γ (1 α) ( 1) − φ (α, β)= l!f βl − − S(l) m l Γ (1 α m + j) (m j)!j! j Xl=0 Xj=l − − − (k) We introduce the generating function (35) of the Stirling numbers Sm ,

m m m j − 1 l m m l (l) (k) k k φm (α, β)= l!flβ ( 1) − Sj Sm j ( 1) α m! j − − − Xl=0 Xj=l   Xk=0 Let q = m j in the double sum − m m j m l q − m m l (l) (k) k k m l − m (l) (k) k k T = ( 1) − Sj Sm j ( 1) α = ( 1) − Sm q Sq ( 1) α j − − − − q − − Xj=l   Xk=0 Xq=0   Xk=0 30 and after reversing the sums, we obtain

m l m l m l − − m (k) (l) k k T = ( 1) − Sq Sm q ( 1) α −  q −  − Xk=0 Xq=k     (m) Using (53) and Sq =0 if m>q supplies us with

m l m − m (k) (l) m (k) (l) (k + l)! (k+l) Sq Sm q = Sq Sm q = Sm q − q − l!k! q=0 Xq=k   X   and m l m l ( 1) − − (k + l)! T = − S(k+l) ( 1)k αk l! k! m − Xk=0 Hence, the Taylor coefficient (80) becomes

m m l 1 l − (k + l)! (k+l) m (l+k) k φ (α, β)= f β S ( 1) − α (81) m m! l k! m − Xl=0 Xk=0 which we can express as a polynomial in α by letting n = m l − m n 1 m n (k + m n)! (k+m n) n+k k φm (α, β)= fm nβ − − Sm − ( 1) α m! − k! − nX=0 Xk=0 After reversing the sums and letting j = m n, we arrive at (76). Reversing the sums in (76) − m j j k k 1 j d − f(t) α (j) m j j φ (α, β)= S ( 1) − β m m! k dtj k β m − j=0 k=0   − t=0   ! X X

k k α α d β t substituting = k e into the k-sum in brackets β dt t=0   j j k k j j k k j j d − f(t) α j d − f(t) d α t d α t = e β = f(t)e β k dtj k β k dtj k dtk dtj k=0   − t=0   k=0   − t=0 t=0 t=0 X X   where Leibniz’s rule has been used, demonstrates (78) and proves Theorem 4. 

A verification of Theorem 4 by computing the inverse Laplace transform is given in Appendix F. Lauwerier [12, p. 35] interestingly mentions that, due to the asymptotic relation Γ(m) m z for Γ(z+m) ∼ − large m, the factorial series (75) for β = 1, rewritten as

∞ Γ (m) ϕ(z)=Γ(z + α) φm 1 (α, 1) Γ (z + α + m) − m=1 X α m φm−1(α,1) and the corresponding Dirichlet series ϕD(z) = m∞=1 m−z have the same converge range for Re (z) > c. Consequently, the rich theory of DirichletP series (see e.g. [22, 23]) directly applies to convergence aspect of factorial series (75). Reviewing Landau’s work on the factorial series, Temme m z 1 [20] adds Newton’s series m∞=0 ( 1) − φm 1 (α, 1) to the factorial and Dirichlet series as the − m − third type of series with theP same convergence  range.

31 Lauwerier [12, p. 42-43] gives examples of factorial series (75) with α = 0 and β = 1. Temme [20] e−ztdt provides even more interesting examples such as ϕ(z)= 0∞ (1+t)ν . Here, we add: bt bt 1 Example 1 If f (t)= e , then e = z b and theR corresponding factorial polynomial φm (α, 1) L − dk αt dk (α+b)t k in (78) is, with k f(t)e = k e = (α + b) and the generating function (35), dt t=0 dt t=0

 m  m 1 k m k (k) − m! φ (α, 1) bt = (α + b) ( 1) − S = (α + b + k) m |e − m Xk=0 kY=0 The factorial series (75) becomes, for Re (z) > b,

m 1 1 ∞ − (b + α + k) Γ (z + α) ∞ Γ (b + α + m) = k=0 = (82) z b m (z + α + k) Γ (b + α) Γ (z + α + m + 1) m=0 k=0 m=0 − X Q X which is known for α = 0 (seeQ e.g. Nielsen [14, band I, p. 77]). Indeed, Gauss’s classical result [1, 15.1.20] for the hypergeometric series at z = 1 is, for c = k (k integer) and Re(c a b) > 0, 6 − − − Γ(c)Γ(c a b) Γ(c) ∞ Γ(a + m)Γ(b + m) F (a, b; c;1) = − − = (83) Γ(c a)Γ(c b) Γ(a)Γ(b) Γ(c + m)m! m=0 − − X For a = 1, b b + α and c = z + α + 1, Gauss’s formula (83) reduces to (82). → tk Example 2 Let f (t) = Ea,b (t) = k∞=0 Γ(b+ak) , which is the Mittag-Leffler function [8]. The Laplace transform (see e.g. [27, art. 20])P is k ∞ zt γ 1 β ∞ Γ (γ + βk) x e− t − Ea,b xt dt = γ+kβ (84) 0 Γ (b + ak) z Z   Xk=0 valid for β a and γ > 0. For γ = β = x = 1, the Laplace transform (84) simplifies to a Laurent | | ≤ | | series (68) in z. The corresponding factorial polynomial (76) can be written as

m j k j! α (j) m j j m! φm (α, β) E (t) = Sm ( 1) − β | a,b Γ (b + a (j k)) k! β ! − Xj=0 Xk=0 −   j α t where the k-sum reduces to 1+ β for a = b = 1, in which case E1,1 (t)= e simplifying to example 1. Unfortunately, for arbitrary a and b, we could not simply φm (α, β) for the Mittag-Leffler |Ea,b(t) function Ea,b (t). On the other hand, after replacing z kz in factorial series (75), multiplying both sides by xk → and adding over all k 0, we formally obtain a Mittag-Leffler transformation ≥ k ∞ x ϕ(zk) ∞ zt ∞ = f(t)E xe− dt = β φ (α, β) m!E (x) Γ (βzk + α) βz,α m βz,α+m+1 k=0 Z0 m=0 X  X 6.4 Open question

The factorial series (75) of a non-entire function may converge, whereas the Laurent series (68) does not. An example is the Binet function µ (z), whose Laurent series (68) is Stirling’s famous asymptotic, k but divergent series (23). Hence, the question arises: “Given the Taylor series f (t)= k∞=0 fkt with radius of convergence Rf , when does a factorial series (75) of the Laplace transform ϕP(z) converge?”

32 We can only give a partial insight. The recipe in 5 steps for a factorial series (75) of the Laplace transform ϕ(z) requires that the Taylor series (74) of g (u; 0, β)= f( β log (1 u)) around the origin − − u0 = 0 converges within the unit circle. Its corresponding Taylor coefficient is written in terms of 1 dkf(t) fk = k as g0 = f0 and k! dt t=0

m 1 k m k (k) g (β)= k!f β ( 1) − S for m> 0 (85) m m! k − m Xk=1 from which it follows that, for m> 0, the Taylor coefficient gm (β) is independent of f0 = f (0) (like any characteristic coefficient (100)). The inverse transform of (85)

m m 1 m k (k) β f = k!g (β) ( 1) − for m> 0 (86) m m! k − Sm Xk=1 (k) where m is the Stirling number of the Second Kind, is deduced as in the proof of Property 5. S (k) (k) m k (k) In contrast to Sm , the Stirling numbers m are non-negative. Thus, ( 1) − Sm > 0 in (85), k (k) S − whereas ( 1) m in (86) is alternating with k. If fk is non-negative, then (85) indicates that also − S 1 m 1 m k (k) g (β) is non-negative. Moreover, (85) written as g (β)= f + − k!f ( 1) − Sm then shows m m m m! k=1 k − that gm (β) fm. Consequently, the radius Rg of convergenceP of g (u; 0, β) = f( β log (1 u)) = ≥ m − k − m∞=0 gm (β) u is not larger than the radius Rf of convergence of f (t) = k∞=0 fkt . However, ifPfk is non-negative, then the non-entire function f (t) has a pole at a finite,P real t = Rf and its Laplace integral ϕ(z) in (66) does not exist. On the other hand, if f = ( 1)k f is alternating, then k − | k| f ( t) has non-negative Taylor coefficients so that f (t) is decreasing in t. The Laplace integral ϕ(z) − m exists for decreasing functions f (t). If gm (β) = ( 1) gm (β) is alternating, then (86) shows that m 1 m 1 (k) − | | f = ( 1) g (β) + − k! g (β) m is alternating and f > g (β) , implying that the m − | m | m! k=1 | k | S | m| | m | radius of convergence R R . In summary, if f = ( 1)k f is alternating, then the factorial series g ≥Pf k − | k| (75) has higher probability of convergence than its corresponding Laurent series (68).

Acknowledgement I am very grateful to Professor R. B. Paris and Professor R. E. Kooij for pointing to a few misprints in an early version and to Professor A. Olde Daalhuis for pointing me to Nemes’ paper [13]. After completing this paper, Dr. Nemes has informed me about the work of Shi et al. [19], whose Theorem 1 is essentially the same as our Theorem 3, though their coefficient for bm (α) is only mentioned in integral form, different from (41), and our validity range of α is slightly broader. Moreover, our proof(s) and approach are entirely different. A. Olde Daalhuis and G. Nemes pointed me to the book of Lauwerier [12] and to [5]. N. Temme has sent me his earlier report [20] of which parts appeared in Lauwerier’s book.

References

[1] M. Abramowitz and I. A. Stegun. Handbook of Mathematical Functions. Dover Publications, Inc., New York, 1968. [2] J. P. M. Binet. M´emoire sur les int´egrales d´efinites Eul´eriennes et sur leur application `ala th´eorie des suites ainsi qu‘ `al`´evaluation des functions des grands nombres. Journal de l‘Ecole´ Polytechnique, XVI:123–343, July 1839. [3] I. V. Blagouchine. Two series expansions for the logarithm of the gamma function involving Stirling numbers and − containing only rational coefficients for certain arguments related to π 1. arXiv1408.3902v9, May 2016.

33 [4] L. Comtet. Advanced Combinatorics. D. Riedel Publishing Company, Dordrecht, Holland, revised and enlarged edition, 1974. [5] E. Delabaere and J.-M. Rasoamanana. Sommation effective d’une somme de Borel par s´eries de factorielles. Annales de l‘institut Fourier, 57(2):421–456, 2007. [6] A. Erd´elyi, W. Magnus, F. Oberhettinger, and F. G. Tricomi. Higher Transcendental Functions, volume 1 of California Institute of Technology Bateman Manuscript Project. McGraw-Hill Book Company, New York, 1953. [7] M. A. Evgrafov. Analytic Functions. W. B. Saunders Company, 1966; Reprinted by Dover Publications, Inc., New York, dover 2019 edition, 2019. [8] R. Gorenflo, A. A. Kilbas, F. Mainardi, and S. V. Rogosin. Mittag-Leffler Functions, Related Topics and Applica- tions. Springer, second edition, 2020. [9] G. H. Hardy. Divergent Series. Oxford University Press, London, 1948. [10] G. H. Hardy. A Course of Pure Mathematics. Cambridge University Press, 10nth edition, 2006. [11] Ch. Hermite. Sur la function log Γ(a). Journal f¨ur die reine und angewandte Mathematik, 115:201–208, 1895. [12] H. A. Lauwerier. Aymptotic Analysis. Mathematical Centre Tracts. Mathematisch Centrum, Amsterdam, 1974. [13] G. Nemes. Generalization of Binet’s Gamma function formulas. Integral Transforms and Special Functions, 24(8):595–606, 2013. [14] N. Nielsen. Die Gammafunktion: Band I. Handbuch der Theorie der Gammafunktion und Band II. Theorie des Integrallogarithmus und verwandter Transzendenten. B. G. Teubner, Leipzig 1906; republished by Chelsea, New York, 1956. [15] F. W. J. Olver, D. W Lozier, R. F. Boisvert, and C. W. Clark. NIST Handbook of Mathematical Functions. Cambridge University Press, New York, 2010. [16] R. B. Paris and D. Kaminski. Asymptotics and Mellin-Barnes Integrals, volume 85 of Encyclopedia of Mathematics and its Applications. Cambridge University Press, U. K., 2001. [17] J. Riordan. Combinatorial Identities. John Wiley & Sons, New York, 1968. [18] G. Sansone and J. Gerretsen. Lectures on the Theory of Functions of a Complex Variable, volume 1 and 2. P. Noordhoff, Groningen, 1960. [19] X. Shi, F. Liu, and M. Hu. A new asymptotic series for the Gamma function. Journal of Computational and Applied Mathematics, 195:134–154, 2006. [20] N. M. Temme. Asymptotische ontwikkelingen van Fakulteitsreeksen. Technical Report TN 48, Mathematisch Centrum, Amsterdam, March 1967. [21] E. C. Titchmarsh. Introduction to the Theory of Fourier Integrals. Oxford University Press, Ely House, London W. I, 2nd edition, 1948. [22] E. C. Titchmarsh. The Theory of Functions. Oxford University Press, Amen House, London, 1964. [23] E. C. Titchmarsh and D. R. Heath-Brown. The Theory of the Zeta-function. Oxford Science Publications, Oxford, 2nd edition, 1986. [24] P. Van Mieghem. The asymptotic behaviour of queueing systems: Large deviations theory and dominant pole approximation. Queueing Systems, 23:27–55, 1996. [25] P. Van Mieghem. Graph Spectra for Complex Networks. Cambridge University Press, Cambridge, U.K., 2011. [26] P. Van Mieghem. Performance Analysis of Complex Networks and Systems. Cambridge University Press, Cambridge, U.K., 2014. [27] P. Van Mieghem. The Mittag-Leffler function. arXiv:2005.13330, 2020. [28] E. T. Whittaker and G. N. Watson. A Course of Modern Analysis. Cambridge University Press, Cambridge, UK, cambridge mathematical library edition, 1996. [29] D. V. Widder. The Laplace transform. Princeton University Press, Princeton, 1946.

34 A Complex integral for Binet’s function µ (z)

A.1 Derivation of the complex integral in (20)

From Weierstrass’s product (25) of the Gamma function, Whittaker and Watson [28, p. 277] deduce the formula, valid for all a and z,

Γ (a) Γ (a) 1 q+i π ζ (s, a) log = z ′ + ∞ zsds with 1

q+i c+i s 2 1 ∞ π ζ (s) 1 ∞ π ζ (s) d z πζ (s) (s 1) zsds = zsds + lim − 2πi q i sin πs s 2πi c i sin πs s s 1 ds s sin πs Z − ∞ Z − ∞ → ! because the function between brackets is analytic at s = 1. Executing the derivative,

d zs πζ (s) (s 1)2 zsπ (s 1) 1 πζ (s) (s 1) cos πs − = − log z ζ (s) (s 1) + ζ′ (s) (s 1) + 2ζ (s) − ds s sin πs s sin πs − s − − − sin πs !    and using the Taylor expansions of (s 1) ζ (s) around s = 1 gives us − d zs πζ (s) (s 1)2 lim − = z (log z 1+ γ) s 1 ds s sin πs − − → ! and we obtain, for 0

1 c+i π ζ (s) log Γ (z +1) = z log z z ∞ zsds − − 2πi c i sin πs s Z − ∞ Moving the line of integration over the double pole at s = 0 to the left yields, for 1 < c < 0, − ′ ′ 1 c +i π ζ (s) d πsζ (s) log Γ (z +1) = z log z z ∞ zsds lim zs − − 2πi c′ i sin πs s − s 0 ds sin πs Z − ∞ →   The derivative is d sζ (s) zsζ (s) s 1 π cos πs ζ (s) zs = + log z + ′ ds sin πs sin πs s − sin πs ζ (s)     1 2n 1 1 Since π cot (πx)= 2 n∞=1 ζ (2n) x − , we find that lims 0 π cot (πs)=0 and x − → s − P ds πsζ (s) ζ (0) lim zs = ζ (0) log z + ′ s 0 ds sin πs ζ (0) →     35 1 1 ds πsζ(s) s 1 With ζ (0) = and ζ′ (0) = log(2π), we arrive at lims 0 z = log z + log(2π) − 2 − 2 → ds sin πs − 2 { } and   c′+i 1 1 1 ∞ π ζ (s) s log Γ (z +1) = z + log z z + log(2π) z ds with 1 < c′ < 0 2 − 2 − 2πi c′ i sin πs s −   Z − ∞ From the definition log Γ (z)= z 1 log z z + 1 log (2π)+ µ (z), we find (20). − 2 − 2 We present a second, shorter derivation of (20) by employing the inverse Mellin transform c+i 1 1 ∞ s 2πt = Γ (s) ζ (s) (2πt)− ds for c> 1 e 1 2πi c i − Z − ∞ Substitution into Binet’s integral (2) and reversing the integrals gives

t c+i ∞ arctan z 1 ∞ s ∞ t s µ (z) = 2 2πt dt = Γ (s) ζ (s) (2π)− arctan t− dt ds 0 e 1 πi c i 0 z Z −  Z − ∞ Z    t 2 Partial integration, followed by a substitution u = z and the use of the Beta integral and the Gamma reflection formula results in  1 s ∞ t s z − π arctan t− dt = − z 2 (1 s)sin πs Z0   − 2 and c+i s z ∞ Γ (s) (2πz)− µ (z)= ζ (s) πs ds −2i c i (1 s)sin 2 Z − ∞ − Using the functional equation ζ(s) = 2(2π)s 1 sin πs Γ(1 s)ζ(1 s) yields − 2 − − c+i 1 ∞ π ζ(1 s) 1 s µ (z)= − z − ds with 1

A.2 Derivation of the convergent series (26)

m Substituting the Taylor series (s 1) ζ (s)= ∞ g (1) (s 1) into the integral (20) yields − m=0 m − 1 c+i πP ζ (s) µ (z)= ∞ zsds with 1

c+i m s ∞ 1 ∞ π (s 1) z µ (z)= gm (1) − ds − 2πi c i sin πs s (s 1) mX=0 Z − ∞ − c+i s c+i m 1 s 1 ∞ π z ∞ 1 ∞ π (s 1) − z = ds g (1) − ds −2πi sin πs s (s 1) − m 2πi sin πs s c i m=1 c i Z − ∞ − X Z − ∞

36 We evaluate the first integral. If z 1, then we close the contour over the positive Re (s)-plane | | ≤ (where the integral over semi-circle at infinity vanishes). Cauchy’s residu theorem tells us that

c+i s s s 1 ∞ π z ∞ (s n) z d s z ds = π lim − + π lim −2πi sin πs s (s 1) s n sin πs s (s 1) s 0 ds sin πs (s 1) c i n=2 → → Z − ∞ − X − − d (s 1) zs + π lim − s 1 ds sin πs s → With d s zs s zs 1 1 = π cot πs + log z ds sin πs (s 1) sin πs (s 1) s − − (s 1) − −  −  1 Since lims 0 π cot πs = 0, we find that → s − d s zs πs π lim = (log z + 1) lim = (log z + 1) s 0 ds sin πs (s 1) − s 0 sin πs − → − → and, similarly, that d (s 1) zs π lim − = z (log z 1) s 1 ds sin πs s − − → Hence, for z 1, | |≤ c+i s n n 1 ∞ π z ∞ ( 1) z ds = − (log z + 1) z (log z 1) −2πi sin πs s (s 1) n (n 1) − − − c i n=2 Z − ∞ − X − 1 1 1 With n 1 n = n(n 1) , we have − − − n n n n ∞ ( 1) zn ∞ ( z) ∞ ( z) ∞ ( z) − = − − = ( z 1) − + ( z) n (n 1) n 1 − n − − n − nX=2 − nX=2 − nX=2 Xn=1 = (z + 1) log (1 + z) z − Thus, for z 1, we find | |≤ 1 c+i π zs 1 ∞ ds = (z + 1) log 1+ 1 −2πi c i sin πs s (s 1) z − Z − ∞ −   For z 1, we close the contour over the negative Re (s)-plane, | |≥ c+i s s n n 1 ∞ π z ∞ π (s + n) z ∞ ( 1) z ds = lim = − − −2πi sin πs s (s 1) − s n sin πs s (s 1) − n (n + 1) c i n=1 →− n=1 Z − ∞ − X − X 1 = (z + 1) log 1+ 1 z −   In summary, the first term equals

1 c+i π zs 1 ∞ ds = (z + 1) log 1+ 1 −2πi c i sin πs s (s 1) z − Z − ∞ −   and c+i m 1 s 1 ∞ 1 ∞ π (s 1) − z µ (z) = (z + 1) log 1+ 1 gm (1) − ds z − − 2πi c i sin πs s   mX=1 Z − ∞

37 The remaining integral is evaluated similarly. For z > 1, we close the contour over negative | | Re (s)-plane and obtain

c+i m 1 s m 1 s m 1 n 1 ∞ π (s 1) − z ∞ π (s + n) (s 1) − z ∞ ( n 1) − z − ds = lim − = ( 1)n − − − 2πi sin πs s s n sin πs s − n c i n=1 →− n=1 Z − ∞ X X − and c+i m 1 s m 1 1 ∞ π (s 1) − z ∞ (n + 1) − 1 − ds = ( 1)m 2πi sin πs s − n ( z)n c i n=1 Z − ∞ X − ( x)n From log (1 + x)= ∞ − , we have that − n=1 n P m 1 m 1 m ∞ (n + 1) − 1 m 1 y d − y y ( 1) n = ( 1) − e− m 1 (e log (1 + e )) − n ( z) − dy −y n=1 − z=e X −

Leibniz’ rule gives m m m l l d y y m d − y d y m (e log (1 + e )) = m l (e ) l (log (1 + e )) dy l dy − dy Xl=0   m l 1 y y y m d − y 1 = e (log(1 + e )) + e 1+ e− − l dyl 1 l=1   − X  For k> 0, it holds10 that

k 1 k m d − 1 m 1 (m) 1 F k(y)= k 1 y = (m 1)!( 1) − k y (87) − dy − 1+ e− − − S 1+ e−   mX=1   Thus, we find

c+i m 1 s m 1 1 ∞ π (s 1) − z m 1 y d − y y − ds = ( 1) − e− m 1 (e log (1 + e )) 2πi c i sin πs s − dy − z=e−y Z − ∞ m 1 l v m 1 1 − m 1 m v (v) 1 = ( 1) − log 1+ + − (v 1)!( 1) − − z l − − Sl 1+ z   Xl=1   Xv=1   Reversal of the last double sum and with (v+1) = m m (v), we have Sm+1 l=v l Sl c+i m 1 s P  m 1 v 1 ∞ π (s 1) − z m 1 1 − m v (v+1) 1 − ds = ( 1) − log 1+ + (v 1)!( 1) − m 2πi c i sin πs s − z − − S 1+ z Z − ∞   Xv=1   and c+i m 1 s 1 ∞ 1 ∞ π (s 1) − z µ (z) = (z + 1) log 1+ 1 g (1) − ds z − − m 2πi sin πs s m=1 c i   X Z − ∞ 1 1 ∞ = (z + 1) log 1+ 1 + log 1+ g (1) ( 1)m z − z m − m=1     X m 1 v ∞ − m v (v+1) 1 g (1) (v 1)!( 1) − − m − − Sm 1+ z mX=1 Xv=1   10 1 ∞ xp In the theory of the Fermi-Dirac integral Fp(z)= Γ(p+1) 0 1+ex−z dx for complex p and z, the functional equation dFp(y) R − dy = Fp 1(y) leads to (87).

38 m 1 Further, with ( 1) ζ (0) = ∞ g (1) ( 1) = , we obtain − m=0 m − 2 P m 1 v 1 1 ∞ − m v (v+1) 1 µ (z)= z + log 1+ 1 g (1) (v 1)!( 1) − 2 z − − m − − Sm 1+ z m=1 v=1     X X   The forward difference formula (14) shows that the first terms are equal to µ (z) µ (z +1) = − z + 1 log z 1 and that − 2 z+1 −  m 1 v ∞ − m 1 v (v+1) 1 µ (z +1) = g (1) (v 1)!( 1) − − m − − Sm 1+ z m=1 v=1 X X   which is (26), after replacing z + 1 z. →

A.3 Taylor coefficients gm (1) of the

( 1)n+1 The convergent Dirichlet series of the Eta function η(s) = ∞ − for Re (s) 0 immediately n=1 ns ≥ leads to the Taylor expansion P ∞ η(k)(s) η(s)= (s x)k (88) k! − Xk=0 with ∞ ( 1)n+1 η(k)(s) = ( 1)k − logk n (89) − ns n=1 X However, the Dirichlet series of η(k)(s) converges too slowly to be of any practical use. Fortunately, fast converging series are obtained for real s 0 by the Euler transform [9], ≥ m j m ∞ m 1 ( 1) lnk j 1 η(k) (s) = ( 1)k+1 − − (90) −  j 1 js  2 m=1 X Xj=1  −      η(s) Invoking the relation ζ(s)= 1 21−s and using the generating function (12) of the Bernoulli numbers, − we have n 1 1 1 ∞ ( log 2) n = = Bn − (s 1) 1 21 s −e (s 1) log 2 1 (s 1) log 2 n! − − − − n=0 − − − X 1 After executing the Cauchy product of the Taylor series for 1 21−s and that of the Eta function in − (88), we obtain

k j 1 ∞ ( log 2) η(k j)(1) ζ(s)= B − − (s 1)k (s 1) log 2  j j! (k j)!  − − Xk=0 Xj=0 − k  j j 1 (k j)  1 ∞ ( 1) log − 2 η − (1) k 1 = + B − (s 1) − s 1  j j! (k j)!  − − Xk=1 Xj=0 −   where we have used that η(1) = log 2. Equating corresponding powers in (s 1) in both Taylor series − of (s 1) ζ (s) yields, with g (1) = 1 and for k> 0, − 0 k 1 k j (k j) j 1 g (1) = B ( 1) η − (1) log − 2 (91) k k! j j − Xj=0   39 m The Taylor coefficient of (s 1)ζ(s)= ∞ g (1) (s 1) around s = 1 follow from (91) as − m=0 m − 0 g0(1) = 1.0 P g1(1) = γ =0.5772156649015328606 g2(1) = 0.07281584548367672486 g3(1) = 0.004845181596436159243 − g4(1) = 0.000342305736717224311 g5(1) = 0.00009689041939447083573 − −6 −7 g6(1) = 6.611031810842189181 10 g7(1) = 3.31624090875277236 10 − −7 − −9 g8(1) = 1.0462094584479187422 10 g9(1) = 8.733218100273797361 10 −11 − −11 g10(1) = 9.478277782762358956 10 g11(1) = 5.658421927608707966 10 −12 −13 g12(1) = 6.768689863513696656 10 g13(1) = 3.492115936672031855 10 − −15 −15 g14(1) = 4.41042474175775338 10 g15(1) = 2.3997862217709991766 10 −16 − −18 g16(1) = 2.167731220072682855 10 g17(1) = 9.54446607636696516 10 −20 − −20 g18(1) = 7.387676660538636498 10 g19(1) = 4.800850782488065211 10 − −21 −22 g20(1) = 4.139956737713305639 10 g21(1) = 1.19168201593979951 10 −

x n B Taylor series of log (1+x) for integer n

Integrating the double generating function

m (k) ∞ S (1+ x)u = eu log(1+x) = m ukxm m! m=0 X Xk=0 (k) of the Stirling numbers Sm of the First Kind [1, Sec. 24.1.3 and 24.1.4] with respect to u results, for x < 1, in | | m eb log(1+x) ea log(1+x) ∞ bk+1 ak+1 xm − = S(k) − log (1 + x) m k + 1 m! m=0 X Xk=0 In particular, for b = 1 and a = 0, we obtain the Taylor series, valid for x < 1, | | m (k) m (k) x ∞ S xm ∞ S xm = m =1+ m (92) log (1 + x) k + 1 m! k + 1 m! m=0 ! m=1 ! X Xk=0 X Xk=1 We generalize the above. The n-fold integral of euλ equals

b u1 un−1 b u λ 1 n 1 uλ du du . . . du e n = (b u) − e du 1 2 n (n 1)! − Za Za Za − Za Let t = b u, followed by y = λt, then − b λb λ(b a) n 1 uλ e − n 1 y (b u) − e du = y − e− dy − λn Za Z0 and the integral can be executed leading to

b λb m k 1 n 1 uλ e λ(b a) (λ (b a)) (b u) − e du = n 1 e− − − (n 1)! a − λ − k! ! − Z Xk=0 uλ λkuk k On the other hand, e = k∞=0 k! and the n-fold integration of u is

b P u1 un−1 b k 1 n 1 k du du . . . du u = (b u) − u du (93) 1 2 n n (n 1)! − Za Za Za − Za 40 and b b n 1 1 n 1 k n u − k n+k n 1 k (b u) − u du = b 1 u du = b (1 w) − w dw a − a − b a − Z Z   Z b 1 n 1 k Γ(n)Γ(k+1) which simplifies considerably if a = 0, due to the Beta integral (1 w) − w dw = . 0 − Γ(n+k+1) Thus, choosing a = 0 leads to R

m k k k 1 λb (λb) n ∞ λ b n e = b λ − k! ! (n + k)! Xk=0 Xk=0 k k (k) xm Further, with λ = log (1 + x) and introducing the Taylor series λ = log (1 + x)= k! m∞=k Sm m! , valid for x < 1, yields, after reversal of the k- and m-sum, | P n 1 k m (k) k m 1 log(1+x)b − (log (1 + x) b) n ∞ k!Sm b x n e = b log (1+ x) − k! ! (k + n)! ! m! Xk=0 mX=0 Xk=0 valid for x < 1 and which simplifies for b = 1 to, | | n 1 m (k) x 1 − 1 ∞ k!S xm = + + m (94) logn (1 + x) n! n k (k + n)! m! k! log − (1 + x) m=1 ! Xk=1 X Xk=1 For n = 1 in (94), we find again (92). Applying n-fold integration to the generating function (35)

b u1 un−1 m 1 m − (k) k! k+n du1 du2 . . . dun (un k)= Sm b 0 0 0 − (n + k)! Z Z Z kY=0 Xk=0 and executing the left-hand side (via partial integration) yields

b m 1 m 1 n 1 − (k) k! k+n (b u) − (u k)du = Sm b (n 1)! 0 − − (n + k)! − Z kY=0 Xk=0 which links Stirling numbers to the general integral form used by Binet [2, p. 339] in the of his Binet function µ (z).

C The Taylor series (42)

Inspired by Nemes [13, Section 3] and using the integral (41), we compute the “exponential” generating function of the Binet polynomials bm (α),

m 1 ∞ b (α) α 1 ∞ um − m um = x + α (k + x)dx (m 1)! 2 − m! m=1 α 1 m=1 X − Z −    X kY=0 m 1 ( 1)mΓ(1 x) Γ(1 x) x From (35), it follows that k=0− (k + x)= −Γ(1 x m−) and Γ(1 x−m)m! = −m . Provided that u < 1, − − − − | | the binomial sum equals Q 

m m 1 ∞ u − ∞ x m m x (k + x)= − ( 1) u = (1 u)− 1 m! m − − − mX=1 kY=0 mX=1   41 Hence, we obtain, for u < 1, | | α α ∞ bm (α) m 1 x log(1 u) 1 u = x + α e− − dx x + α dx (m 1)! α 1 2 − − α 1 2 − mX=1 − Z −    Z −    and α 1 α 1 α α 1 (1 u)− + (1 u) − (1 u) − (1 u)− ∞ b (α) g (u)= − − + − − − = m um (95) α −2 log (1 u) 2 (m 1)! log (1 u) m=1 − − X − which reduces, for α = 0, to the Taylor series (34) in Binet’s second derivation (33) in Theorem 1. Moreover, u α 2 1 u α gα (u)=(1 u)− − = (1 u)− g0 (u) − log (1 u) − log2 (1 u) − − − ! cm m where g0 (u)= m∞=1 (m 1)! u was prominent in Binet’s proof of Theorem 1. − We make Binet’s substitution u = 1 e t in (95) and obtain P − − αt (α 1)t (α 1)t αt e + e − e − e ∞ bm (α) t m + − = 1 e− 2t t2 (m 1)! − m=1 − X  The Taylor series around t0 = 0 of the left-hand side is

k+2 k+1 k+1 k+2 k+2 1 eαt + e(α 1)t e(α 1)t eαt ∞ 2 α + (α 1) + (α 1) α − + − − = − − − tk t 2 t !   (k + 2)!  Xk=0   while the right-hand side is   m ∞ bm (α) t m ∞ bm (α) m j jt 1 e− = ( 1) e− (m 1)! − (m 1)! j − m=1 − m=1 − j=0   X  X X k k 11 jt k ( 1) t but the reversal of the m- and k-sum is not allowed . After Taylor expansion of e− = k∞=0 j − k! around t0 = 0, P

m k k ∞ bm (α) t m ∞ bm (α) ∞ m j k ( 1) t 1 e− = ( 1) j − (m 1)! − (m 1)!  j −  k! m=1 − m=1 − k=0 j=0   X  X X X   and recognizing the closed form expression [1, sec. 24.1.4.C] of the Stirling number of the Second Kind

m (m) 1 m j m k = ( 1) − j (96) Sk m! − j Xj=0   we have k ∞ bm (α) t m ∞ ∞ (m) m k t 1 e− = mb (α) ( 1) − (m 1)! − m Sk − k! m=1 − k=0 m=1 ! X  X X j −jt 11 ∞ ∞ mbm(α) (−1) e Indeed, j=0 m=j (m−j)! j! diverges any j > 0, because P P  ∞ j j mbm (α) d d −α = lim gα (u) = lim (1 − u) g0 (u) (m − j)! u→1 duj u→1 duj mX=j

∞ mc dj ∞ c m → −∞ m and m=j (m−j)! = limu 1 duj g0 (u)= for j > 0, but g0 (1) = m=1 (m−1)! = 0, that already appeared in Section 2.4 inP the determination of limz→0 zµ (z) = 0. P

42 Since the positive integer (m) = 0 for m > k, we finally arrive at Sk k+2 k+1 k+2 αk+1 + (α 1) + (α 1) αk+2 k k ∞ 2 k ∞ (m) m k t − − − t = mb (α) ( 1) −   (k + 2)!  m Sk − k! m=1 ! Xk=0   Xk=0 X which is, after equating corresponding powers in t, again (45).

D Integral for the Binet coefficient cm

The Taylor series (34), which is a special case of (42) for α = 0, is written in terms of the Binet coefficients cm in (32) as 1 m u 1+ u ∞ ( 1) c g ( u)= 2 = − m um 0 − 2 − log (1 + u) (m 1)! log (1+ u) m=1 X − The corresponding integral form for the Taylor coefficient [22] is ( 1)m c 1 dω ω 1+ 1 ω − m = 2 (m 1)! 2πi ωm+1 log2 (1 + ω) − log (1 + ω) − ZC(0) ! where C (0) is a contour that encloses the point u0 = 0 in counter-clockwise sense. A straightforward execution of the contour C (0) is to choose a circle around the origin, with radius 0 < r 1, due to ≤ the branch cut at the negative real axis for Re (ω) < 1. The resulting integral is numerically not − stable. An alternative way is to deform the contour to enclose the entire complex plane (except for the point ω = 0 and avoiding the branch cut) in clockwise sense. The integrand vanishes at ω = 1. − For ω , the integrand vanishes for m> 1 and we only maintain the path around the branch cut. | |→∞ In particular, we construct a path that travels from infinity to the point 1 q < 0 under an angle − ≤ θ, where θ (0,π) and returns from the point q along a straight line under angle θ to infinity. We − ∈ thus obtain m 0 iθ iθ 1 1 iθ ( 1) c 1 d q + xe− q + xe 1+ q + xe− − m = − 2 2 iθ m+1 2 iθ iθ (m 1)! −2πi (q + xe ) log (1+ q + xe− ) − log (1 + q + xe− )! − Z∞ −  iθ iθ 1 1 iθ 1 ∞ d q + xe q + xe 1+ 2 q + 2 xe − 2πi iθ m+1 log2 (1 + q + xeiθ) − log (1 + q + xeiθ) Z0 (q + xe )  ! The computation simplifies if we choose q = 1, − m 1 iθ iθ iθ ( 1) − cm 1 ∞ e 1+ xe 1 1+ xe − = Im m+1 − 2 dx (m 1)! π 0 ( 1+ xeiθ) (log x + iθ) − 2 log x + iθ ! − Z −   We evaluate the integrand. Denoting x cos θ 1 x cos θ + 1 θ Ψ (x)= θ (m + 1) arccos − + arccos arctan − √x2 2x cos θ + 1 √x2 + 2x cos θ + 1 − log x − we obtain x cos θ 1 θ m 1 sin θ m arccos − 2 arctan ( 1) − cm 1 ∞ dx − √x2 2x cos θ+1 − log x − = − m 2 2 (m 1)! π 0 log x + θ  (x2 2x cos θ + 1) 2  − Z − 1 (1+2x cos θ + x2) sin Ψ (x) ∞ dx 2 m+1 − 2π 0 p log x + θ2 (x2 2x cos θ + 1) 2 Z − p 43 π This form simplifies substantially if we choose θ = 2 . After simplifying the sines, we arrive at

− − cos m arccos 1 +2 arctan θ cos (m+2) arccos 1 +arctan θ  √x2+1 log x   √x2+1 log x  2 + m 1 log2 x+ π 2 π 2 ( 1) − cm 1 ∞ ( 2 ) 2 log x+( 2 ) − = m q dx (97) (m 1)! π 0 (x2 + 1) 2 − Z However, the numerical evaluation of the integral (97) is remarkably inaccurate. Therefore, we simplify the cosines. After some manipulations, we arrive at an integral for the Binet coefficient cm for m> 1,

2 log2 x ( π ) (1 x2) log x + π x cos m arccos 1 − 2 + − 2 2 m 1 √−2 2 π 2 (1+x2) ( 1) − c 1 ∞ x +1 log x+( ) − m =  2 dx  m 2 (m 1)! π 0 2 2 2 π − Z (x + 1) log x + 2   π 2 1 π log x x log x 4 (1 x ) sin m arccos − − 2 + − 2 − 1 √x2+1 log2 x+( π ) (1+x ) + ∞  2 dx (98)   m 2 π 0 2 2 2 π Z (x + 1) log x + 2   where cos (m arccos y) is the Chebyshev orthogonal polynomial in y. The integral (98) can be evaluated accurately. Upper bounding the cosines in (97) leads to

m 1 ( 1) − cm 1 ∞ dx 1 1 − m + 2 π 2 (m 1)! ≤ π 0 (x2 + 1) 2 log x + 2 π 2  − Z  2 2 log x + 2 

 q The function in between brackets . is maximal at x = 1, where it equals 4 +1 . Thus, { } π2 π m 1 m 1 ( 1) − cm 1 4 1 ∞ dx 1 4 1 Γ 2− 1 − 2 + m = 2 + m = O (m 1)! ≤ π π π 0 (x2 + 1) 2 2√π π π Γ √m −   Z   2    but this upper bound is rather weak. In particular, since all coefficients c form> 2 have the same m cm cm 1 sign by Theorem 2, the convergence of m∞=1 (m 1)! implies that (m| 1)!| = O m1+ε for ε> 0. − − P  1 E Asymptotic expansion for bm(2 + α)

We start from the integral in (56), 1 m 1 2 (m + 1) bm+1 + α = u (u + α) (k + u + α)du 2 1   Z− 2 k=1 m 1Y m 2 u = (k + α) u (u + α) 1+ du 1 k + α kY=1 Z− 2 kY=1   u Provided k|+|α < 1, the product can be expanded around u0 =0 as | | m m j 1 m u u ∞ ( 1) − 1 j 1+ = exp log 1+ = exp − j u k + α k + α !  j (k + α)  kY=n   kX=n   Xj=1 kX=n m j 1 m  1 ∞ ( 1) − 1 j = exp u exp − j u (k + α)!  j (k + α)  kX=n Xj=2 kX=n   44 1 1 The convergence requirement indicates that k + α > 2 for k n 1, which means that 2 n < α. | | ≥ ≥ 1 − We limit ourselves here to n = 1, implying that the analysis is valid for real α> 2 and, thus, after 1 − translating bm( 2 + α) to bm (α′) for α′ > 0. If that range must be larger, then we can increase n 2, 3 ≥ so that α> 2 and so on; the only effect is that the integral Ij below is a little more involved, but still analytically computable. For j 2, the sum m 1 converges for all m, whereas m 1 ≥ k=n (k+α)j k=n (k+α) diverges when m , which justifies the split-offP of the j = 1 term. The limit m P case can → ∞ 1 → ∞ be expressed in terms of the Hurwitz Zeta-function ζ (s, α) = k∞=1 (k+α)s (see Appendix A). The remaining j-series is alternating with decreasing coefficients andP can thus be bounded as 2 m j 1 m 2 m 3 m u 1 ∞ ( 1) − 1 u 1 u 1 < − uj < + − 2 (k + α)2 j (k + α)j − 2 (k + α)2 3 (k + α)3 kX=n Xj=2 kX=n kX=n kX=n Rather than continuing with these bounds, we proceed with an exact computation using our charac- j−1 ( 1) m 1 j teristic coefficients [24, Appendix], that enables us to expand exp ∞ − u = j=2 j k=n (k+α)j l l∞=0 φlu in a Taylor series around u0 = 0. The Taylor series of a functionP G (z) ofP a function f (z) is P m ∞ 1 dkG(f) G(f(z)) = s[k,m] (z ) (z z )m (99) k! df k f(z) 0 − 0 m=0 k=0 f=f(z0) ! X X

k The characteristic coefficients of a complex function f (z) with Taylor series f (z)= k∞=0 fk (z0) (z z0) , 1 dm k − defined by s[k,m] f (z0)= m! dzm f (z) f (z0) , possesses a general form | − z=z0 P  

k s[k,m] (z )= f (z ) (100) |f 0 ji 0 k i=1 i=1 jXi=m;ji>0 Y P and obeys s[k,m] f (z0)=0if k< 0 and k>m. Moreover, s[k,m] f (z0) possesses a recursion and the | | j−1 ( 1) m 1 j coefficients φ can be computed up to any desired order. The function f (u)= ∞ − u l j=2 j k=n (k+α)j j−1 ( 1) m 1 has clearly two vanishing Taylor coefficients, f = f = 0, while f = − P  .P Invoking  0 1 j j k=n (k+α)j (99) m P ∞ 1 ef(u) =1+ s[k,m] um k! m=1 " # X Xk=1 l 1 indicates that φ0 = 1 and φl = k=1 k! s[k, l]. Because f1 = 0, it holds that φ1 = s[1, 1] = 0. We list the first Taylor coefficients φl, P m m 1 1 1 1 φ2 = and φ3 = −2 (k + α)2 3 (k + α)3 kX=n kX=n m 2 m 1 1 1 1 φ = 4 8 2 − 4 4 k=n (k + α) ! k=n (k + α) Xm m X m 1 1 1 1 1 φ5 = + −6 (k + α)2 (k + α)3 5 (k + α)5 kX=n kX=n kX=n m 3 m 2 m m m 1 1 1 1 1 1 1 1 1 φ6 = 2 + 3 + 2 4 6 −48 (k + α) ! 18 (k + α) ! 8 (k + α) (k + α) − 6 (k + α) kX=n kX=n kX=n kX=n kX=n

45 (k) In passing by, our characteristic coefficients also enable to compute the Stirling numbers Sm via the generating function (35) for large m up to any order desired. Let us proceed with n = 1 (restricting α> 1 ) and denote γ (α)= m 1 , then − 2 m k=1 (k+α) m 1 ∞ 2 P 1 uγm(α) l+1 (m + 1) bm+1( + α)= (k + α) φl (u + α) e u du 2 1 kY=1 Xl=0 Z− 2 The integral

1 1 1 2 2 2 (u + α) euγm(α)ul+1du = euγm(α)ul+2du + α euγm(α)ul+1du 1 1 1 Z− 2 Z− 2 Z− 2 1 2 au l requires us to compute Il = 1 e u du for integer l 0. Partial integration leads to the recursion − 2 ≥ R 1 2 au l 1 a l a l Il = e u du = e 2 ( 1) e− 2 Il 1 1 a2l − − − a − Z− 2   1 a − a 2 au e 2 e 2 which, after iteration down to I0 = 1 e du = −a , leads to − 2 R 1 l l a j l a j 2 au l l! ( 1) a 2 a 2 Il = e u du = − e 2 − e− 2 1 al+1  j! − j!  2 j=0  j=0  Z− X X   Although the right-hand side seems to increase factorially with l, the integral indicates that liml Il = →∞ 0. Thus, we obtain

γ (α) 1 l+2 m q 1 q l 1 2 e 2 ( 1) α + l +1+(2 q) α 2 (l + 1)! (u + α) euγm(α)ul+1du = − 2 − − − l γm(α) 1 q+1 1 2 (l + 2 q)! (γ (α)) 2 q=0 ( + ( 1) e− α 2 l 1 + (2 q)α ) m Z− X − − − − − 1 1 (m+1)bm+1( 2 +α) 2 uγ m(α) l+1  Returning to P = m (k+α) = l∞=0 φl 1 (u + α) e u du and after reversing of the l- k=1 − 2 Q m 1 and q-sum, we obtain the expansion inP inverseR powers of γm (α)= k=1 (k+α) , P 1 ∞ γm(α) 1 l γm(α) 1 1 2 2 P = φl e α + + ( 1) e− α l 2γm (α) 2 − − 2 2 Xl=0      1 ∞ γm(α) 1 l γm(α) 1 1 2 2 + 2 φl e α + l +1+ α + ( 1) e− α l 1+ α l (γm (α)) − 2 − − 2 − 2 Xl=0        γm(α) q l γm(α) 2 1 2 1 ∞ ∞ (l + 1)! e ( 1) α + 2 l + 1 qα + ( 1) e− α 2 l 1 qα 2q + φ − − − − − −  l n 2l 1 (l q)! o q+3 q=0  −    (γm (α)) X Xl=q −  j−1 j−1  ( 1) m 1 j l m u ( 1) m 1 j With exp ∞ − u = ∞ φ u and 1+ = exp ∞ − u , j=2 j k=n (k+α)j l=0 l k=n k+α j=1 j k=n (k+α)j we observe that all l-sums in the last double sum are derivatives evaluated at u = 1 . For example,  P P P Q P± 2 P the first sum equals

∞ γm(α) 1 l γm(α) 1 1 S = φ e 2 α + + ( 1) e− 2 α 1 l 2 − − 2 2l l=0      X m m 1 1 1 1 = α + 1+ + α 1 2 2 (k + α) − 2 − 2 (k + α)   kY=1     kY=1  

46 We arrive at the expansion in powers of 1 , γm(α)

m m 1 α + 1 1+ 1 + α 1 1 1 (m + 1) bm+1( 2 + α) 2 k=1 2(k+α) − 2 k=1 − 2(k+α) P = m = k=1 (k + α)  Q  2γm (α)  Q  

1 ∞ γm(α) 1 l γm(α) 1 1 Q 2 2 + 2 φl e α + l +1+ α + ( 1) e− α l 1+ α l (γm (α)) − 2 − − 2 − 2 Xl=0        q l+1 ∞ ∞ γm(α) q 1 l γm(α) 1 2 − (l + 1)! + φ e 2 ( 1) α + l + 1 qα + ( 1) e− 2 α l 1 qα l − 2 − − − 2 − − q+3 q=0 (l q)! (γm (α)) X Xl=q        − (101)

m 1 In particular, for large m, where γm (α) = k=1 (k+α) = O (log (m + a)), the expansion (101) shows that P m m 1 m α α + 1 1+ 1 + α 1 1 1 bm+1( + α) 1+ 2 k=1 2(k+α) 2 k=1 2(k+α) 1 2 = k=1 k − − +O m! m + 1  2 m 1   log2 m Q   Q k=1 (k+α)  Q   (102) P For α 1 and b (0) = c , we find that c < 0 and that →− 2 m m m

m 1 m 1 k=1 1 1 c 1 − 2(k 2 ) 1 1 m+1 = k=1 − 2k  −  + O = O m! − m + 1 Q 2 m 1 log2 m m log m Q  k=1 k 1     − 2 1 P while for α = 2 , bm (1) = βm > 0 and

m m 1 1 b (1) 1+ k=1 1+ 2(k+α) 1 1 m+1 = k=1 2k + O = O m 1  2 m! m + 1 Q 2 k=1 1 log m m log m Q  k+ 2     P cm+1 1 bm+1(1) 1 Although m! = O m log m and m! = O m log m , the products for cm = bm (0) are smaller than for βm = bm (1), illustrating that the α = 0 case converges faster than the α = 1 case (as in Fig. 1).

F Verification of Theorem 4

Substituting the factorial series (75) into the inverse Laplace transformation (67) and assuming that summation and integration can be reversed, yields

∞ 1 c+i eztdz f(t)= m!φ (α, β) ∞ m 2πi m (βz + α + k) m=0 c i k=0 X Z − ∞ For Re (t) 0, Re (α) > 0 and limiting ourselves to β = 1, theQ contour can be closed over the negative ≥ Re (z)-plane, where simple poles at z = α k are enclosed. Cauchy’s residue theorem [22] then − − indicates that

c+i zt m zt m (α+j)t 1 ∞ e dz (z + α + j) e e− m = lim m = m 2πi c i k=0(z + α + k) z (α+j) k=0(z + α + k) k=0;k=j(k j) Z − ∞ Xj=0 →− Xj=0 6 − Q Q Q 47 m j With k=0;k=j(k j) = ( 1) j! (m j)!, we have 6 − − − Q c+i zt m αt 1 ∞ e dz 1 m j (α+j)t e− t m = ( 1) e− = 1 e− 2πi m (z + α + k) m! j − m! − Zc i k=0 j=0   − ∞ X  Q Thus, we obtain

αt ∞ t m f(t)= e− φ (α, 1) 1 e− m − m=0 X  Introducing the form (81) for φm (α, 1) yields

t m m l m l αt ∞ 1 e− 1 d f(t) − (k + l)! (k+l) m (l+k) k f(t)= e− − S ( 1) − α m! l! dtl k! m − m=0  l=0 t=0 k=0 X X X

After reversing the m- and l- sum,

l t m m αt ∞ 1 d f(t) ∞ 1 e− (k) m k k! k l f(t)= e− − S ( 1) − α − l! dtl m! m − (k l)! l=0 t=0 m=l  k=l X X X − we recognize that m l m (k) m k k! k l d (k) m k k S ( 1) − α − = S ( 1) − α m − (k l)! dαl m − Xk=l − Xk=0 m (k) m k k m 1 m α The generating function (35) indicates that Sm ( 1) − α = − (k + α) = m! ( 1) − k=0 − k=0 − m and we have P Q  l l αt ∞ 1 d f(t) d ∞ α t m f(t)= e− − e− 1 l! dtl dαl m − l=0 t=0 m=l   X X  l l l l 1 αt ∞ 1 d f(t) d ∞ α t m d − α t m = e− − e− 1 − e− 1 l! dtl dαl m − − dαl m − l=0 t=0 m=0   m=0   ! X X  X 

α m α t t − αt For any α and real t 0, the binomial sum m∞=0 −m e− 1 = 1+ e− 1 = e , while dl l 1 α t ≥ m l 1 α t m− − − e 1 = 0 because − e 1 is a polynomial in α of degree l 1. dαl m=0 −m − m=0 −m −    l − P − − Finally, with d eαt = tleαt, we return, indeed, to the Taylor expansion of f (t) around the point P dαl  P   t0 = 0. 

48