1 Introduction and preliminaries

The generalized harmonic numbers H n ( s ) of order s are defined by (cf. [1]; see also [2, 3], [[4], p.156] and [[5], Section 3.5])

H n ( s ) := j = 1 n 1 j s (nN;sC),
(1.1)

and

H n := H n ( 1 ) = j = 1 n 1 j (nN)
(1.2)

are the harmonic numbers. Here ℕ and ℂ denote the set of positive integers and the set of complex numbers, respectively, and we assume that

H 0 :=0, H 0 ( s ) :=0 ( s C { 0 } ) and H 0 ( 0 ) :=1.

The generalized harmonic functions H n ( s ) (z) are defined by (see [2, 6]; see also [7, 8])

H n ( s ) (z):= j = 1 n 1 ( j + z ) s ( n N ; s C Z ; Z : = { 1 , 2 , 3 , } )
(1.3)

so that, obviously,

H n ( s ) (0)= H n ( s ) .

Equation (1.1) can be written in the following form:

H n ( s ) =ζ(s)ζ(s,n+1) ( ( s ) > 1 ; n N )
(1.4)

by recalling the well-known (easily-derivable) relationship between the Riemann zeta function ζ(s) and the Hurwitz (or generalized) zeta function ζ(s,a) (see [[4], Eq. 2.3(9)])

ζ(s)=ζ(s,n+1)+ k = 1 n k s ( n N 0 : = N { 0 } ) .
(1.5)

The polygamma functions ψ ( n ) (s) (nN) are defined by

ψ ( n ) (s):= d n + 1 d z n + 1 logΓ(s)= d n d s n ψ(s) ( n N 0 ; s C Z 0 : = Z { 0 } ) ,
(1.6)

where Γ(s) is the familiar gamma function, and the psi-function ψ is defined by

ψ(s):= d d s logΓ(s)and ψ ( 0 ) (s)=ψ(s) ( s C Z 0 ) .

A well-known (and potentially useful) relationship between the polygamma functions ψ ( n ) (s) and the generalized zeta function ζ(s,a) is given by

ψ ( n ) (s)= ( 1 ) n + 1 n! k = 0 1 ( k + s ) n + 1 = ( 1 ) n + 1 n!ζ(n+1,s) ( n N ; s C Z 0 ) .
(1.7)

It is also easy to have the following expression (cf. [[4], Eq. 1.2(54)]):

ψ ( m ) (s+n) ψ ( m ) (s)= ( 1 ) m m! H n ( m + 1 ) (s1)(m,n N 0 ),
(1.8)

which immediately gives H n ( s ) another expression for H n ( s ) as follows (cf. [[9], Eq. (20)]):

H n ( m ) = ( 1 ) m 1 ( m 1 ) ! [ ψ ( m 1 ) ( n + 1 ) ψ ( m 1 ) ( 1 ) ] (mN;n N 0 ).
(1.9)

By using finite differences, Spivey [10] presented many summation formulas involving binomial coefficients, the Stirling numbers of the first and second kind and harmonic numbers, two of which are chosen to be recalled here: [[10], Identity 14]

k = 0 n ( n k ) H k = 2 n ( H n k = 1 n 1 k 2 k ) (n N 0 ),
(1.10)

which was also given by Paule and Schneider [[11], Eq. (39)] by deriving it automatically by means of the Sigma package in [12], together with the following identity [[10], Identity 20]:

k = 0 n ( 1 ) k ( n k ) H k = 1 n (nN).
(1.11)

Paule and Schneider [11] proved five conjectured harmonic number identities similar to those arising in the context of supercongruences for Apéry numbers, one of which is recalled here as follows [[11], Eq. (5)]:

j = 0 n (15j H j +5j H n j ) ( n j ) 5 = ( 1 ) n j = 0 n ( n j ) 2 ( n + j j ) .
(1.12)

Greene and Knuth [[13], p.10] recorded six commonly used identities that involve both binomial coefficients and harmonic numbers, two of which are recalled here:

(1.13)
(1.14)

Alzer et al. [[14], Eq. (3.62)] proved, by using the principle of mathematical induction, that

j = 1 n H j j = 1 2 [ ( H n ) 2 + H n ( 2 ) ] (nN).
(1.15)

By using (1.15) in conjunction with the following elementary identity (see [2]):

H j + 1 = H j + 1 j + 1 ,
(1.16)

we obtain

j = 1 n H j j + 1 = 1 2 [ ( H n + 1 ) 2 H n + 1 ( 2 ) ] (nN).
(1.17)

Chu and De Donno [15] made use of the classical hypergeometric summation theorems to derive several striking identities for harmonic numbers other than those discovered recently by Paule and Schneider [11], one of which is recalled below [[15], Thereoem 1].

(1.18)

One interesting special case of (1.18) is when we set μ=0. We thus find that

k = 0 n ( n k ) ( n + λ n n k ) H λ n + k = ( 2 n + λ n n ) (2 H λ n + n H λ n + 2 n ),
(1.19)

which can be further specialized, with λ=0, to the following form:

k = 0 n ( n k ) 2 H k = ( 2 n n ) (2 H n H 2 n ).
(1.20)

Dattolli and Srivastava [16] proposed several generating functions involving harmonic numbers by making use of an interesting approach based on the umbral calculus. Subsequently, Cvijović [17] showed the truth of the conjectured relations in [16] by using simple analytical arguments.

For a concise and beautiful description of these numbers, we refer also to WolframMathWorld’s website [18].

As we have seen in the above brief eclectic review, harmonic and generalized harmonic numbers are involved in a variety of useful identities. Of course, certain interesting properties of harmonic and generalized harmonic numbers have been studied (see, e.g., [19]). Here we aim at presenting further interesting identities about certain interesting finite series associated with binomial coefficients, harmonic numbers and generalized harmonic numbers.

2 Finite-series involving binomial coefficients, harmonic numbers and generalized harmonic numbers

As the illustrative identities in Section 1, we consider certain interesting identities about finite-series involving binomial coefficients, harmonic numbers and generalized harmonic numbers. We begin by recalling a known formula (cf. [[20], p.362, Entry (55.4.8)]; see also [[2], Eq. (2.6)]):

(2.1)

where ( α ) n denotes the Pochhammer symbol defined (for αC) by

( α ) n :={ 1 ( n = 0 ) , α ( α + 1 ) ( α + n 1 ) ( n N ) .
(2.2)

Differentiating each side of (2.1) with respect to the variables a and b, respectively, using (1.8) and considering the following easily derivable identities:

d d α ( α ) n = ( α ) n H n ( 1 ) (α1) ( n N 0 ; α C Z 0 )
(2.3)

and

d d α 1 ( α ) n = H n ( 1 ) ( α 1 ) ( α ) n ( n N 0 ; α C Z 0 ) ,
(2.4)

we obtain the following formulas in Theorem 1.

Theorem 1 Each of the following identities holds true:

(2.5)

and

(2.6)

Setting a=b1=0 in (2.1), (2.5) and (2.6) and using (1.3) and (1.8), we get certain interesting finite-sum identities involving binomial coefficients and harmonic numbers, respectively, asserted by Corollary 1.

Corollary 1 Each of the following identities holds true:

(2.7)
(2.8)

and

j = 1 n ( 1 ) j + 1 j + 1 ( n j ) H j ( H j + 1 1)= 1 n + 1 [ H n ( 2 ) n n + 1 H n ] (n N 0 ).
(2.9)

Remark 1 In the course of presenting a closed-form evaluation of some useful series involving the generalized zeta function ζ(s,a), Choi et al. [21] made use of the identity (2.7) without its proof. Choi and Srivastava [2] proved Eq. (2.7) as a special case of (2.1) here and presented another illustrative proof.

We will try to express a class of the following finite sums involving harmonic numbers and binomial coefficients as given above:

j = 1 n ( 1 ) j + 1 ( j + 1 ) k ( n j ) H j (n N 0 ;kN).
(2.10)

Here we give the answers for k=2 and k=3 in (2.10) asserted by the following lemma.

Lemma 1 Each of the following identities holds true:

j = 1 n ( 1 ) j + 1 ( j + 1 ) 2 ( n j ) H j = 1 2 ( n + 1 ) [ ( H n + 1 ) 2 H n + 1 ( 2 ) ] (n N 0 )
(2.11)

and

j = 1 n ( 1 ) j + 1 ( j + 1 ) 3 ( n j ) H j = 1 2 ( n + 1 ) j = 1 n 1 j + 1 [ ( H j + 1 ) 2 H j + 1 ( 2 ) ] (n N 0 ).
(2.12)

Proof We will prove only (2.11) by using the same method as in [[2], pp.2224-2225]. A similar argument will establish (2.12). We first recall two basic relations for binomial coefficients:

( n + 1 j ) = ( n j ) + ( n j 1 ) and ( n j 1 ) = j n + 1 ( n + 1 j ) .
(2.13)

We let the left-hand side of (2.11) be

f n := j = 1 n ( 1 ) j + 1 ( j + 1 ) 2 ( n j ) H j
(2.14)

so that, using the first one of (2.13),

f n + 1 = ( 1 ) n ( n + 2 ) 2 H n + 1 + j = 1 n ( 1 ) j + 1 ( j + 1 ) 2 [ ( n j ) + ( n j 1 ) ] H j = ( 1 ) n ( n + 2 ) 2 H n + 1 + f n + j = 1 n ( 1 ) j + 1 ( j + 1 ) 2 ( n j 1 ) H j .
(2.15)

We now see that, using the second one of (2.13),

(2.16)

By using the identity in (2.7), we find that

j = 1 n ( 1 ) j + 1 j + 1 ( n + 1 j ) H j = 1 ( 1 ) n n + 2 H n + 1 (n N 0 ).
(2.17)

We also have

j = 1 n ( 1 ) j + 1 ( j + 1 ) 2 ( n + 1 j ) H j = f n + 1 ( 1 ) n ( n + 2 ) 2 H n + 1 (n N 0 ).
(2.18)

Thus, substituting from (2.17) and (2.18) into (2.16), we obtain

(2.19)

Finally, it follows from (2.15) and (2.19) that

(n+2) f n + 1 (n+1) f n = H n + 1 n + 2 .

Let a n :=(n+1) f n so that we have

a n + 1 a n = H n + 1 n + 2 and a 1 = 1 2 H 1 = 1 2 .
(2.20)

By telescoping this last sum (2.20), we obtain

a n =(n+1) f n = j = 1 n H j j + 1 .
(2.21)

Applying (1.17) to (2.21), we get the desired identity (2.11). □

Applying (2.7) and (2.11) to (2.9) and considering (2.8), we obtain two interesting identities asserted by the following theorem.

Theorem 2 Each of the following identities holds true:

j = 1 n ( 1 ) j + 1 j + 1 ( n j ) ( H j ) 2 = 1 2 ( n + 1 ) [ 3 H n ( 2 ) ( H n ) 2 ] (n N 0 )
(2.22)

and

j = 1 n ( 1 ) j + 1 j + 1 ( n j ) H j ( 2 ) = 1 2 ( n + 1 ) [ 5 H n ( 2 ) 3 ( H n ) 2 ] (n N 0 ).
(2.23)

Differentiating (2.5) and (2.6) with respect to a and observing the following identity:

d d α H j ( ) (α)= H j ( ) (α)(N),
(2.24)

we obtain further interesting identities involving binomial coefficients and generalized harmonic functions asserted by the following theorem.

Theorem 3 Each of the following identities holds true:

(2.25)

and

(2.26)

Setting a=b1=0 in (2.25) and (2.26), we find certain interesting identities and using (2.8), respectively, assert the following corollary.

Corollary 2 Each of the following identities holds true:

(2.27)

and

(2.28)

Remark 2 As in getting the results in Theorem 3, it is seen that a variety of interesting identities involving the generalized harmonic numbers can be obtained by applying the differential operator to the parameters of known formulas.

3 Inverse relations and a question

By using the known orthogonal relation

k = j n ( 1 ) k + j ( n k ) ( k j ) = δ n j (nj;n,j N 0 ),
(3.1)

with δ n j the Kronecker delta ( δ n n =1, δ n j =0 if nj) and a manipulation of double series

k = 0 n j = 0 k A k , j = j = 0 n k = j n A k , j ,
(3.2)

it is easy to find the following simplest inverse relation (see [[22], Chapter 2]):

a n = k = 0 n ( 1 ) k ( n k ) b k b n = k = 0 n ( 1 ) k ( n k ) a k .
(3.3)

Applying this inverse relation to the identities in Section 2, we obtain many formulas involving binomial coefficients, harmonic numbers and generalized harmonic numbers asserted by the following corollary.

Corollary 3 Each of the following identities holds true:

(3.4)
(3.5)
(3.6)
(3.7)
(3.8)

It is observed that Eqs. (2.7), (2.8) and (2.27) are of the following form:

j = 1 n ( 1 ) j + 1 ( n j ) a j = a n (nN)and a 0 =0.
(3.9)

By using the first one of (2.13), we find an identity in the following lemma.

Lemma 2 If Eq. (3.9) holds true, then we obtain the following identity:

j = 1 n ( 1 ) j + 1 j ( n j ) a j =n( a n a n 1 )(nN).
(3.10)

Applying Eq. (3.10) to Eqs. (2.7), (2.8) and (2.27), we get some interesting identities asserted by the following corollary.

Corollary 4 Each of the following identities holds true:

(3.11)
(3.12)
(3.13)

Question We conclude this paper by posing a natural question: Under what conditions does Eq. (3.9) hold true?