<<

The Australian Journal of Mathematical Analysis and Applications

AJMAA

Volume 16, Issue 1, Article 14, pp. 1-14, 2019

CUBIC ALTERNATING HARMONIC NUMBER SUMS ANTHONY SOFO Received 12 September, 2018; accepted 23 January, 2019; published 27 May, 2019.

VICTORIA UNIVERSITY,COLLEGEOF ENGINEERINGAND SCIENCE,MELBOURNE CITY,AUSTRALIA. [email protected]

ABSTRACT. We develop new closed form representations of sums of cubic alternating harmonic numbers and reciprocal binomial coefficients. We also identify a new representation for the ζ (4) constant.

Key words and phrases: function, Alternating cubic harmonic numbers, Combinatorial identities, Sum- mation formulas, Partial fraction approach, Binomial coefficients. 2000 Subject Classification. Primary 05A19, 33C20. Secondary 11M06, 11B83.

ISSN (electronic): 1449-5910 c 2019 Austral Internet Publishing. All rights reserved. 2 ANTHONY SOFO

1. INTRODUCTIONAND PRELIMINARIES In this paper we will develop identities, closed form representations of alternating cubic har- monic numbers and reciprocal binomial coefficients, including integral representations, of the form: ∞ n+1 X (−1) H3 (1.1) Ω(k, p) = n ,  n + k  n =1 np k

th for p = 0, 1 and k ∈ N0. Here, the n harmonic number n X 1 Z 1 1 − tn (1.2) H = = γ + ψ (n + 1) = dt, H := 0 n r 1 − t 0 r=1 0 and as usual, γ denotes the Euler-Mascheroni constant and ψ(z) is the Psi (or Digamma) func- tion defined by d Γ0(z) Z z ψ(z) := {log Γ(z)} = or log Γ(z) = ψ(t) dt. dz Γ(z) 1 For sums of harmonic numbers with positive terms [10], [27], [28] and [29] have given many results, including sums of the form ∞ X H3 n .  n + k  n =1 np k Other results are given by [5] and [17]. Let R and C denote, respectively the sets of real and complex numbers and let N := {1, 2, 3, ···} be the set of positive , and N0 := N ∪ {0} . λ A generalized binomial coefficient µ (λ, µ ∈ C) is defined, in terms of the familiar gamma function, by λ Γ(λ + 1) := , λ, µ ∈ . µ Γ(µ + 1) Γ (λ − µ + 1) C The Pochhammer symbol (λ)ν (λ, ν ∈ C) is also defined in terms of the gamma function, by  1 (ν = 0; λ ∈ \{0}) Γ(λ + ν)  C (λ) := = . ν Γ(λ)  λ(λ + 1) ··· (λ + n − 1) (ν = n ∈ N; λ ∈ C), (m) A generalized harmonic number Hn of order m is defined, for positive integers n and m, as follows: n X 1 (m) H(m) := , m, n ∈ and H := 0, m ∈ . n rm N 0 N r=1 In the case of non- values of n such as (for example) a value ρ ∈ R, the generalized (m+1) harmonic numbers Hρ may be defined, in terms of the polygamma functions dn dn+1 ψ(n)(z) := {ψ(z)} = {log Γ(z)}, n ∈ , dzn dzn+1 N0 by (−1)m (1.3) H(m+1) = ζ (m + 1) + ψ(m) (ρ + 1) ,H(m+1) = 0 ρ m! 0 (ρ ∈ R \ {−1, −2, −3, ···} ; m ∈ N) ,

AJMAA, Vol. 16, No. 1, Art. 14, pp. 1-14, 2019 AJMAA CUBIC ALTERNATING HARMONIC NUMBER SUMS 3

(α) r  where ζ (z) is the . The evaluation of the ψ a at rational values of the argument can be explicitly done via a formula as given by Kölbig [7], or Choi and Cvijovic [2] in terms of the polylogarithmic or other . Some specific values are listed in the books[15], [21] and [22]. The polylogarithm or de-Jonquière function Lip (z), is defined as, ∞ X zn Li (z) := , p ∈ when |z| < 1; < (p) > 1 when |z| = 1. p np C n =1 Some results for sums of alternating and non-alternating harmonic numbers may be seen in the works of [3], [4], [12], [13], [14], [16], [17], [18], [19], [20], [23], [24], [26] and references therein. Some explicit, and closely related results may also be seen in the well presented papers [9] and [25]. The following lemma will be useful in the development of the main theorems.

Lemma 1.1. Let r be a positive integer and p ∈ N. Then: r j   X (−1) 1 (p) (p) (p) (1.4) F (p, r) = p = p H r + H r−1 − H r+1 j 2 [ 2 ] [ 2 ] 2[ 2 ]−1 j =1 where [x] is the integer part of x, and when p = 1,

r j X (−1) (1.5) F (1, r) = = H r − Hr. j [ 2 ] j =1 For p = 2,

r j   X (−1) 1 (2) (2) 1 (1.6) F (2, r) = 2 = H r − H r+1 1 − ζ (2) . j 4 [ 2 ] [ 2 ]− 2 2 j =1 Proof. The proof is given in the paper [11]. Lemma 1.2. The following identity holds, ∞ n+1 X (−1) H3 5 3 1 9 (1.7) X (0) = n = ζ (4) + ζ (2) ln2 2 − ln4 2 − ζ (3) ln 2, n 8 4 4 8 n =1

1 2  Z log (1 − x) log (1 + x) + Li2 (−x) (1.8) = dx, 0 x (1 + x)

where Lip (·) is the polylogarithm function. Proof. Consider ∞ X tn (1.9) V (j, t) =  n + j  n =1 n j

∞ ∞ X tn Γ(n)Γ(j + 1) X = = tn B (n, j + 1) , Γ(n + j + 1) n =1 n =1

AJMAA, Vol. 16, No. 1, Art. 14, pp. 1-14, 2019 AJMAA 4 ANTHONY SOFO

where Γ(·) is the gamma function and B (·, ·) is the beta function. Now

Z 1 j ∞ Z 1 j (1 − x) X n (1 − x) V (j, t) = (tx) dx = t dx, x 1 − tx 0 n =1 0 and differentiating p times with respect to j and then letting j → 0 with t = −1, results in Z 1 p   (p) log (1 − x) p 1 V (0, −1) = dx = (−1) p!Lip+1 . 0 1 + x 2

For p = 3 1 (1.10) V (0, −1)(3) = −6Li 4 2 From (1.9) we also have ∞ n (p) X (−1) σn V (0, −1) = n n =1 where " !# dp  n + j −1 σn = lim , j→0 djp j when p = 3, ∞ n+1   (3) X (−1) φ 1 (1.11) V (0, −1) = n = 6Li n 4 2 n =1

3 (2) (3) and φn = Hn + 3HnHn + 2Hn . From the paper [8], we have the results

∞ n+1 (3) X (−1) Hn 19 3 = ζ (4) − ζ (3) ln 2, n 16 4 n =1

∞ n+1 (2)   X (−1) HnHn 1 = 2Li − ζ (4) − 4ζ (2) ln2 2 n 4 2 n =1 1 7 + ln4 2 + ζ (3) ln 2, 4 8 and substituting into (1.11) we have the result (1.7). The representation of the integral (1.8) is obtained in the following way. From [11], we can express, for p ∈ N

p Z 1 Z 1 p n−1 p p Hn Y  Y (−1) p = ··· xj ln (1 − xj) dxj n 0 0 j=1 j=1 R 1 R 1 where 0 ··· 0 is a p- fold integration procedure, for p = 3, 3 Z 1 Z 1 Z 1 n Hn (x1x2x3) ln (1 − x1) ln (1 − x2) ln (1 − x3) − 3 = dx1dx2dx3. n 0 0 0 x1x2x3 (1.12)

AJMAA, Vol. 16, No. 1, Art. 14, pp. 1-14, 2019 AJMAA CUBIC ALTERNATING HARMONIC NUMBER SUMS 5

Now,

∞ n+1 X (−1) H3 Z 1 Z 1 Z 1 n = − ln (1 − x) ln (1 − y) ln (1 − z) n n =1 0 0 0 ∞ X n−1 × n2 (−xyz) dxdydz n =1

Z 1 Z 1 Z 1 ln (1 − x) ln (1 − y) ln (1 − z) (1 − xyz) = − 2 dxdydz 0 0 0 (1 + xyz) Z 1 Z 1 ln (1 − x) ln (1 − y) (1 − ln (1 + xy)) = 2 dxdy 0 0 (1 + xy) Z 1 ln (1 − x) ln2 (1 + x)  x  = − Li2 dx 0 x (1 + x) 2 1 + x and applying Landen’s identity  x  ln2 (1 + x) Li = − − Li (−x) 2 1 + x 2 2 we obtain (1.8). By association we conclude Z 1 log (1 − x) log2 (1 + x) + Li (−x) 5 3 2 dx = ζ (4) + ζ (2) ln2 2 0 x (1 + x) 8 4 1 9 − ln4 2 − ζ (3) ln 2 4 8

∞ n+1 X (−1) H3 = X (0) = n . n n =1

An alternate manipulation of (1.7) leads to the following result and adds to the results on cubic sums some of which are published in [10]. Lemma 1.3. X H3 5 67 3 3 2n = ζ (4) + ζ (3) + 2ζ (2) + ζ (2) ln2 2 − ζ (2) ln 2 2n (2n − 1) 8 16 4 2 n≥1 9 1 3 − ζ (3) ln 2 + ln 2 − ln4 2 + ln3 2 − ln2 2. 8 4 2 (1.13) Proof. By re-arrangement

∞ 1 3 3 ! X H2n − H X (0) = 2n − 2n 2n − 1 2n n =1 3 2 X H − 3H X 3H2n X 1 = 2n 2n + − . 2n (2n − 1) 4n2 (2n − 1) 8n3 (2n − 1) n≥1 n≥1 n≥1

AJMAA, Vol. 16, No. 1, Art. 14, pp. 1-14, 2019 AJMAA 6 ANTHONY SOFO

Hence X H3 − 3H2 5 3 1 9 2n 2n = ζ (4) + ζ (2) ln2 2 − ln4 2 − ζ (3) ln 2 2n (2n − 1) 8 4 4 8 n≥1 33 3 3 + ln 2 + ζ (3) − ζ (2) + ln2 2 16 2 2 1 1 − ζ (3) − 3 ln 2 − ζ (2) . 8 4

Now we can evaluate X H2 5 1 1 3 2n = ζ (2) + ln 2 − ln2 2 − ζ (2) ln 2 + ln3 2 + ζ (3) , 2n (2n − 1) 4 2 3 4 n≥1 therefore (1.13) follows. Lemma 1.4. The following identity holds, ∞ n+1 X (−1) H3 X (1) = n n + 1 n =1

5 3 1 9 = − ζ (4) − ζ (2) ln2 2 + ln4 2 + ζ (3) ln 2 16 4 4 8 (1.14) 5 = −X (0) + ζ (4) , 16

1 2  5 Z log (1 − x) log (1 + x) + Li2 (−x) (1.15) = ζ (4) − dx. 16 0 x (1 + x) Proof. Consider X (1) and by a change of index ∞ n ∞ n+1 3 X (−1) H3 X (−1)  1  X (1) = n−1 = − H − n n n n n =1 n =1

∞ n+1 3 ∞ n+1 ∞ n+1 X (−1) H X (−1) HnHn−1 X (−1) = − n − 3 + . n n2 n4 n =1 n =1 n =1 Using lemma 1.2 9 7 X (1) = −X (0) − ζ (4) + ζ (4) 16 8 and (1.14) follows. The integral identity (1.15) follows directly from (1.8). We can also note that a similar calculation as that of lemma 1.3, yields the identity  3  X H 3H2nH2n−1 3 2n + = −X (0) + ζ (4) 2n (2n + 1) 4n2 (2n − 1) 8 n≥1

9 3 3 69 + ζ (2) − ln2 2 − ζ (2) ln 2 + ln3 2 + ζ (3) . 4 2 2 16

AJMAA, Vol. 16, No. 1, Art. 14, pp. 1-14, 2019 AJMAA CUBIC ALTERNATING HARMONIC NUMBER SUMS 7

Remark 1.1. It may be noticed from lemma 1.2 and lemma 1.4 that 5 X (0) + X (1) = ζ (4) 16 in which case manipulating the integral identities we obtain 8 Z 1 log (1 − x) log2 (1 + x) (1.16) ζ (4) = − dx. 3 0 x Further extraction from X (0) and X (1) gives us the results Z 1 2   log (1 − x) log (1 + x) 1 1 4 dx = 2ζ (3) ln 2 + 2Li4 + ln 2 0 1 + x 2 3

(1.17) −2ζ (4) − ζ (2) ln2 2,

Z 1   log (1 − x) Li2 (−x) 7 1 1 4 dx = ζ (3) ln 2 + 2Li4 + ln 2 0 x (1 + x) 8 2 12 (1.18) 1 −ζ (4) − ζ (2) ln2 2 4 and Z 1   log (1 − x) Li2 (−x) 7 1 1 4 dx = − ζ (3) ln 2 − 2Li4 − ln 2 0 x 4 2 12 11 1 + ζ (4) + ζ (2) ln2 2 4 2

∞ n+1 X (−1) Hn = . n3 n =1 The (1.16), (1.17) and (1.18) cannot be analytically evaluated by "Mathematica". There are many integral representations of powers of Pi, for example

Z 1 √ Z 1 logp (1 − x) π = 4 1 − x2dx, (−1)p p!ζ (p + 1) = dx. 0 0 x

Guillera and Sondow [6] gave Z 1 Z 1 Z 1 Z 1 2 dydx 3 ln xy dydx π = 8 2 2 , π = −16 2 2 , 0 0 1 − x y 0 0 1 + x y

and amongst many other results [1] obtained the intriguing representation (after some manipu- lation) ∞   π  X n+1 1 ln = (−1) ln 1 + , 2 n n =1 however the author believes (1.16) is a new representation.

AJMAA, Vol. 16, No. 1, Art. 14, pp. 1-14, 2019 AJMAA 8 ANTHONY SOFO

Lemma 1.5. Let r ≥ 2 be a positive integer, defining

∞ n+1 X (−1) Hn (1.19) S (r) := n + r n =1 then S (r) has the 1 (−1)r+1 S (r) + S (r − 1) = (1 + (−1)r) ln 2 − F (1, r − 1) r − 1 r − 1 with representation

  r+1 r S (r) = (−1) S (1) + (−1) 2Hr−1 − H r−1 ln 2 [ 2 ]

r−1 r X 1   + (−1) H j − Hj , j [ 2 ] j =1 (1.20) 1 2  1 2 where S (0) = 2 ζ (2) − ln 2 ,S (1) = 2 ln 2 and F (1, r − 1) is given by (1.5). Proof. The proof is detailed in [11]. Lemma 1.6. For a positive integer r ≥ 2, we define

∞ n+1 X (−1) H2 T (r) := n n + r n =1 then T (r) has the recurrence relation 2S ζ (2) ln2 2 ln 2 T (r) + T (r − 1) = − r−1 + − + r − 1 2 (r − 1) r − 1 (r − 1)2 1   − 2 H r−1 − H r−2 , 2 (r − 1) 2 2 with representation

1 (1.21) T (r) = (−1)r+1 T (1) + (−1)r+1 F (1, r − 1) ζ (2) 2

+ (−1)r F (1, r − 1) ln2 2 + (−1)r+1 F (2, r − 1) ln 2

r−1 ! j H j − H j−1 r X (−1) + (−1) 2S (j) + 2 2 , j 2j j =1

3 1 3 1 1 1 3 1 with T (0) = 4 ζ (3) + 3 ln 2 − 2 ζ (2) ln 2, T (1) = 2 ζ (2) ln 2 − 3 ln 2 − 4 ζ (3), S (j) is given by (1.19) and F (·, r − 1) is given by (1.4). Proof. The proof is detailed in [11].

AJMAA, Vol. 16, No. 1, Art. 14, pp. 1-14, 2019 AJMAA CUBIC ALTERNATING HARMONIC NUMBER SUMS 9

Lemma 1.7. Let r ∈ N\{1}, then ∞ n+1 X (−1) HnHn−1 Y (r − 1) = 3 n (n + r − 1) n =1

3 1 1 1  = ζ (3) + ln3 2 − ζ (2) ln 2 − T (r − 1) (r − 1) 8 3 2 3 (1.22) + ζ (2) − ln2 2 − 2S (r − 1) , 2 (r − 1)2 where S (r − 1) is given by (1.19) and T (r − 1) is given by (1.21). For r = 1 we have

∞ n+1 X (−1) HnHn−1 9 (1.23) Y (0) = 3 = − ζ (4) . n2 16 n =1 Proof. Consider ∞ n+1 X (−1) HnHn−1 Y (r − 1) = 3 n (n + r − 1) n =1 ∞     3 X n+1 1 1 1 = (−1) H H − − r − 1 n n n n n + r − 1 n =1

∞  2 2  3 X n+1 H Hn H Hn = (−1) n − − n + r − 1 n n2 n + r − 1 n (n + r − 1) n =1

3 3 1 1 5  = ζ (3) + ln3 2 − ζ (2) ln 2 − ζ (3) − T (r − 1) r − 1 4 3 2 8 ∞   3 X n+1 Hn 1 1 + (−1) − r − 1 r − 1 n n + r − 1 n =1 3 3 1 1 5  = ζ (3) + ln3 2 − ζ (2) ln 2 − ζ (3) − T (r − 1) r − 1 4 3 2 8 3 1 1  + ζ (2) − ln2 2 − S (r − 1) (r − 1)2 2 2 and (1.22) follows. For the case r = 1, we have ∞ n+1 ∞ n+1 1  X (−1) HnHn−1 X (−1) Hn Hn − Y (0) = 3 = 3 n n2 n2 n =1 n =1

∞  2  X n+1 H Hn = 3 (−1) n − n2 n3 n =1 and (1.23) follows.

AJMAA, Vol. 16, No. 1, Art. 14, pp. 1-14, 2019 AJMAA 10 ANTHONY SOFO

Lemma 1.8. Let r ∈ N, then ∞ n+1 X (−1) H3 X (r) = n n + r n =1 then 3ζ (3) ln3 2 ζ (2) ln 2 X (r) = − (−1)r X (1) + 3 (−1)1+r + − F (1, r − 1) 8 3 2

(−1)r+1 + 2ζ (2) − 3 ln2 2 F (2, r − 1) + (−1)1+r ln 2F (3, r − 1) 2 r−1    1+r X j 1   3 S (j) + (−1) (−1) H j−1 − H j − T (j) + 2j3 2 2 j j j=1 (1.24) where X (0) is given by (1.8), X (1) is given by (1.14), S (j) is given by (1.20), T (j) is given by (1.21) and F (·, r − 1) is given by (1.4). Proof. Consider ∞ n+1 X (−1) H3 X (r) = n n + r n =1 and by a change of summation index ∞ n 3 ∞ n+1 X (−1) H X (−1) HnHn−1 X (r) = n−1 = −X (r − 1) + 3 n + r − 1 n (n + r − 1) n =2 n =1 ∞ n+1 X (−1) + . n3 (n + r − 1) n =1 Now using lemma 1.5, 1.6 and 1.7

3ζ (3) ζ (2) ln 2 X (r) = −X (r − 1) + 3Y (r − 1) + − + 4 (r − 1) 2 (r − 1)2 (r − 1)3 1   + 3 H r−2 − H r−1 , 2 (r − 1) 2 2 hence we have the recurrence relation 3 3 1 1  3T (r − 1) X (r) + X (r − 1) = ζ (3) + ln3 2 − ζ (2) ln 2 − r − 1 8 3 2 r − 1 1 ln 2 + 2ζ (2) − 3 ln2 2 + 2 (r − 1)2 (r − 1)3 3S (r − 1) 1   − 2 + 3 H r−2 − H r−1 (r − 1) 2 (r − 1) 2 2 for r ≥ 2. The recurrence relation is solved by the subsequent reduction of the X (r) ,X (r − 1) , ..., X (1) terms and using lemma 1.1, finally arriving at the relation (1.24). The next few theorems relate the main results of this investigation, namely the closed form and integral representation of (1.1).

AJMAA, Vol. 16, No. 1, Art. 14, pp. 1-14, 2019 AJMAA CUBIC ALTERNATING HARMONIC NUMBER SUMS 11

2. CLOSED FORMAND INTEGRAL IDENTITIES We now prove the following theorems. Theorem 2.1. Let k ≥ 1 be real positive integer, then from (1.1) with p = 0 we have

∞ n+1 3 k   X (−1) H X r+1 k (2.1) Ω(k, 0) = n = (−1) r X (r)  n + k  r n =1 r =1 k

" # Z 1 Z 1 Z 1 ln (1 − x) ln (1 − y) ln (1 − z) 2, 2, 2, 2

= 4F3 − xyz dxdydz 0 0 0 1 + k 1, 1, 2 + k where X (r) is given by (1.24). Proof. Consider the expansion

∞ n+1 ∞ n+1 X (−1) H3 X (−1) k! H3 Ω(k, 0) = n = n  n + k  (n + 1) n =1 n =1 1+k k

∞ k X n+1 X Φr = (−1) k! H3 n n + r n =1 r =1 where      n + r  (−1)r+1 r k Φ = lim = , r k n→−r  Q  k! r  n + r r=1 hence

k   ∞ n+1 3 X r+1 k X (−1) H Ω(k, 0) = (−1) r n r n + r r =1 n =1 k   X r+1 k = (−1) r X (r) , r r =1 and using lemma 1.8, gives k   X r+1 k (2.2) Ω(k, 0) = (−1) r X (r) , r r =1 hence (2.1) follows. For the integral representation we employ the same technique as lemma 1.2.

The other case of Ω(k, 1) , can be evaluated in a similar fashion. We list the result in the next theorem.

AJMAA, Vol. 16, No. 1, Art. 14, pp. 1-14, 2019 AJMAA 12 ANTHONY SOFO

Theorem 2.2. Under the assumptions of Theorem 2.1, we have, ∞ n+1 X (−1) H3 Ω(k, 1) = n  n + k  n =1 n k

" # Z 1 Z 1 Z 1 ln (1 − x) ln (1 − y) ln (1 − z) 2, 2, 2

= 3F2 − xyz dxdydz, 0 0 0 1 + k 1, 2 + k

k   X r+1 k (2.3) = X (0) − (−1) X (r) . r r =1 Proof. The proof follows directly from theorem 2.1 and using the same technique. Example 2.1. Some illustrative examples follow.

∞ n+1 X (−1) H3 255 21 261 15 Ω (3, 0) = n = − ζ (2) − ζ (3) − ζ (4)  n + 3  16 2 16 4 n =1 3 63 +27ζ (2) ln 2 + 36 ln2 2 − 18 ln3 (2) − ln 2 2 27 −9ζ (2) ln2 2 + 3 ln4 2 + ζ (3) ln 2, 2

∞ n+1 X (−1) H3 32575 499 325 85 Ω (4, 1) = n = − + ζ (2) + ζ (3) + ζ (4)  n + 4  1296 36 16 16 n =1 n 4 151 68 10636 −34ζ (2) ln 2 − ln2 2 + ln3 (2) + ln 2 3 3 216

+12ζ (2) ln2 2 − 4 ln4 2 − 18ζ (3) ln 2. Remark 2.1. Following a similar argument as that of Lemma 1.3, we note first from (2.3), with k = 2 ∞ n+1 X (−1) H3 25 1 1 15 n = ζ (4) − + ζ (2) + ζ (3) + ln 2 n (n + 1) (n + 2) 32 2 2 16 n =1 3 3 1 3 9 − ζ (2) ln 2 − ln2 2 − ln4 2 + ln3 2 + ζ (2) ln2 2 − ζ (3) ln 2. 2 2 2 2 4 Also ∞ X H3 1 n = 5ζ (4) − 2ζ (3) − ζ (2) − , n (n + 1) (n + 2) 2 n =1 and hence, ∞ X H3 185 1 17 2n−1 = ζ (4) − 1 − ζ (2) − ζ (3) + ln 2 n (2n − 1) (2n + 1) 32 2 16 n =1

AJMAA, Vol. 16, No. 1, Art. 14, pp. 1-14, 2019 AJMAA CUBIC ALTERNATING HARMONIC NUMBER SUMS 13

3 3 1 3 9 − ζ (2) ln 2 − ln2 2 − ln4 2 + ln3 2 + ζ (2) ln2 2 − ζ (3) ln 2. 2 2 2 2 4 In the following remark we list, without proof, an extension of the results related to Lemma 1.2.

3 (2) (3) Remark 2.2. Let φn = Hn + 3HnHn + 2Hn , then ∞ n+1 X (−1) φ 3 21 1 n = ζ (2) ln2 2 − ζ (3) ln 2 + 6Li n (n + 1) 2 4 4 2 n =1 1 15 − ln4 2 + ζ (4) . 4 8

Z 1 3   log (1 − x) log (1 + x) 3 2 21 1 2 dx = ζ (2) ln 2 − ζ (3) ln 2 + 6Li4 0 x 2 4 2 1 33 − ln4 2 − ζ (4) . 4 8

∞ n+1 X (−1) φ Z 1 log3 (1 − x) log (1 + x) 1  n = − dx n + 1 1 + x x 1 + x n =1 0 21 3 1 15 = ζ (3) ln 2 − ζ (2) ln2 2 + ln4 2 − ζ (4) . 4 2 4 8

REFERENCES [1] H. ALZER and S. KOUMANDOS, Series and product representations for some mathematical con- stants. Period. Math. Hungar., 58 (2009), no. 1, pp. 71–82. [2] J. CHOI and D. CVIJOVIC,´ Values of the polygamma functions at rational arguments, J. Phys. A: Math. Theor. 40 (2007), 15019–15028; Corrigendum, ibidem, 43 (2010), 239801 (1 p). [3] J. CHOI and H. M. SRIVASTAVA, Some summation formulas involving harmonic numbers and generalized harmonic numbers, Math. Comput. Modelling., 54 (2011), pp. 2220-2234. [4] M. W. COFFEY and N. LUBBERS, On generalized harmonic number sums, Appl. Math. Comput., 217 (2010), pp. 689–698. [5] P. FLAJOLET and B. SALVY, Euler sums and contour integral representations, Exp. Math., 7 (1998), pp. 15-35. [6] J. GUILLERA and J. SONDOW, Double integrals and infinite products for some classical constants via analytic continuations of Lerch’s transcendent, Ramanujan J, 16 (2008), pp. 247–270. [7] K. KÖLBIG, The polygamma function ψ (x) for x = 1/4 and x = 3/4, J. Comput. Appl. Math., 75 (1996), pp. 43-46. [8] I. MEZO,˝ Nonlinear Euler sums, Pacific J. Math., 272 (2014), pp. 201-226. [9] T. M. RASSIAS and H. M. SRIVASTAVA, Some classes of infinite series associated with the Riemann zeta and polygamma functions and generalized harmonic numbers, Appl. Math. Comput., 131 (2002), pp. 593-605.

AJMAA, Vol. 16, No. 1, Art. 14, pp. 1-14, 2019 AJMAA 14 ANTHONY SOFO

[10] X. SI, C. XU and M. ZHANG, Quadratic and cubic harmonic number sums, J. Math. Anal. Appl., 447 (2017), no. 1, pp. 419–434. [11] A. SOFO, Quadratic alternating harmonic number sums, J. , 154 (2015), pp. 144- 159 [12] A. SOFO, Sums of derivatives of binomial coefficients, Adv. in Appl. Math., 42 (2009), pp. 123– 134. [13] A. SOFO, Integral forms associated with harmonic numbers, Appl. Math. Comput., 207 (2009), pp. 365-372. [14] A. SOFO, Integral identities for sums, Math. Commun., 13 (2008), pp. 303–309. [15] A. SOFO, Computational Techniques for the Summation of Series, Kluwer Academic/Plenum Pub- lishers, New York, 2003. [16] A. SOFO and H. M. SRIVASTAVA, Identities for the harmonic numbers and binomial coefficients, Ramanujan J., 25 (2011), pp. 93–113. [17] A. SOFO, Harmonic sums and integral representations, Journal Applied Analysis, 16 (2010), pp. 265-277. [18] A. SOFO, A master integral in four parameters, J. Math. Anal. Appl., 448 (2017), no. 1, pp. 81–92. [19] A. SOFO, Alternating cubic Euler sums with binomial squared terms, Int. J. Number Theory, 14 (2018), no. 5, pp. 1357–1374. [20] A. SOFO, General order Euler sums with multiple argument, J. Number Theory, 189 (2018), pp. 255–271. [21] H. M. SRIVASTAVA and J. CHOI, Series Associated with the Zeta and Related Functions, Kluwer Academic Publishers, London, 2001. [22] H. M. SRIVASTAVA and J. CHOI, Zeta and q-Zeta Functions and Associated Series and Integrals, Elsevier Science Publishers, Amsterdam, London and New York, 2012. [23] W. WANG and C. JIA, Harmonic number identities via the Newton-Andrews method, Ramanujan J., 35 (2014), pp. 263-285. [24] C. WEI and D. GONG, The derivative operator and harmonic number identities, Ramanujan J., 34 (2014), pp. 361-371. [25] T. C. WU, S. T. TU and H. M. SRIVASTAVA, Some combinatorial series identities associated with the and harmonic numbers, Appl. Math. Lett., 13 (2000), pp. 101-106. [26] M. WU and H. PAN, Sums of products of Euler numbers, Ars Combin., 136 (2018), pp. 341–346. [27] C. XU, Y, YAN, and Z. SHI, Euler sums and integrals of polylogarithm functions, J. Number Theory, 165 (2016), pp. 84–108. [28] C. XU, M. ZHANG and W. ZHU, Some evaluation of harmonic number sums, Integral Transforms Spec. Funct., 27 (2016), no. 12, pp. 937–955. [29] C. XU and J. CHENG, Some results on Euler sums, Funct. Approx. Comment. Math., 54 (2016), no. 1, pp. 25–37.

AJMAA, Vol. 16, No. 1, Art. 14, pp. 1-14, 2019 AJMAA