Next Article in Journal
On an Integral Equation with the Riemann Function Kernel
Previous Article in Journal
Positive Solutions for a System of Fractional Boundary Value Problems with r-Laplacian Operators, Uncoupled Nonlocal Conditions and Positive Parameters
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

On a Family of Infinite Series with Reciprocal Catalan Numbers

1
Department of Physics and Engineering Physics, Obafemi Awolowo University, Ile-Ife 220005, Nigeria
2
Landesbank Baden-Württemberg, 70173 Stuttgart, Germany
3
Faculty of Mathematics and Computer Science, Vasyl Stefanyk Precarpathian National University, 76018 Ivano-Frankivsk, Ukraine
*
Author to whom correspondence should be addressed.
Statements and conclusions made in this article by R. Frontczak are entirely those of the author. They do not necessarily reflect the views of LBBW.
Axioms 2022, 11(4), 165; https://doi.org/10.3390/axioms11040165
Submission received: 13 January 2022 / Revised: 21 March 2022 / Accepted: 24 March 2022 / Published: 7 April 2022
(This article belongs to the Section Mathematical Analysis)

Abstract

:
We study a certain family of infinite series with reciprocal Catalan numbers. We first evaluate two special candidates of the family in closed form, where we also present some Catalan–Fibonacci relations. Then, we focus on the general properties of the family and prove explicit formulas, including two types of integral representations.

1. Introduction and Motivation

The famous Catalan numbers C n , n 0 are defined by C n = 1 n + 1 2 n n . They can be also expressed by the recursion
C n = 2 ( 2 n 1 ) n + 1 C n 1 , C 0 = 1 .
The generating function for C n is
n = 0 C n z n = 1 1 4 z 2 z .
Catalan numbers have a long history and play an extraordinary role in combinatorics. Excellent sources on these numbers are the books by Koshy [1], Roman [2] and Stanley [3]. Some examples of recent work involving Catalan numbers and their generalizations include [4,5,6,7,8,9,10].
Catalan numbers form a special class of the so-called special numbers and polynomials. Other classes of these objects with comparable importance are Bernoulli numbers (polynomials), Euler numbers (polynomials), Fibonacci numbers (polynomials), etc. These objects play an important role in combinatorics, number theory and mathematical physics. The main approach in the study of these numbers is via their generating functions, which have been studied continuously. These generating functions have attracted considerable attention from many mathematicians, statisticians, physicists and engineers [11,12,13,14,15,16,17,18,19,20].
This paper was inspired by a recent paper by Amdeberhan et al. [21], who studied the function
f ( z ) = n = 0 z n C n , z [ 0 ; 4 ) ,
which generates the reciprocals of Catalan numbers. They prove by several methods that
f ( z ) = 2 ( z + 8 ) ( 4 z ) 2 + 24 z arcsin ( z / 2 ) ( 4 z ) 5 / 2 , z [ 0 ; 4 ) ,
This expression also appears in [22], and an equivalent form is given in [23]. The hypergeometric expression for f ( z ) is
f ( z ) = 2 F 1 1 , 2 ; 1 2 ; z 4 ,
where
2 F 1 ( a , b ; c ; z ) = n = 0 ( a ) n ( b ) n ( c ) n z n n ! ,
( a ) n = a ( a + 1 ) ( a + n 1 ) = Γ ( a + n ) Γ ( a ) , ( a ) 0 = 1 , is the Pochhammer symbol, and Γ ( z ) is the gamma function defined by Γ ( z ) = 0 + t z 1 e t d t with ( z ) > 0 .
Substituting z = 2 and z = 3 , respectively, into (1) yields the evaluations
n = 0 2 n C n = 5 + 3 π 2 and n = 0 3 n C n = 22 + 8 3 π ,
which were stated in 2014 by Beckwith and Harbor as Problem 11765 in the American Mathematical Monthly [24] and solved by Abel [25].
From [26], we have the identities
n = 1 F 2 n C n = 22 5 + 6 ( 5 + 9 5 ) π 125 ω , n = 0 L 2 n C n = 62 5 + 6 ( 15 + 19 5 ) π 125 ω ,
where F n and L n are the famous Fibonacci and Lucas numbers, respectively, α = 1 + 5 2 is the golden ratio and ω = 5 α = 2 + α . These numbers are defined for n 0 by the recursions F n + 2 = F n + 1 + F n and L n + 2 = L n + 1 + L n with initial conditions F 0 = 0 , F 1 = 1 , L 0 = 2 and L 1 = 1 , respectively. The Binet formulas are given by
F n = α n β n α β , L n = α n + β n ,
where β = 1 α = 1 5 2 . For negative subscripts we have
F n = ( 1 ) n + 1 F n a n d L n = ( 1 ) n L n .
See the book by Koshy [27] for more details.
Our purpose in this paper is to study, for each integer m 0 , the following family of series:
g m ( z ) = n = 0 2 2 n n m 2 n + 1 z n C n , 0 z < 1 .
We begin by evaluating the functions g 0 ( z ) and g 1 ( z ) explicitly for some values of z, including Fibonacci and Lucas numbers. Then, focusing on g m ( z ) , we prove some explicit expressions for g m ( z ) , including two integral representations.

2. The Functions g 0 ( z ) and g 1 ( z )

Sprugnoli [28] has derived some generating functions for series involving reciprocals of central binomial coefficients. His approach is built on ordinary differential equations but leaves some gaps in the derivations. For instance, he does not state the domains of the presented functions. One of Sprugnoli’s results [28], Theorem 2.4, is the following identity:
n = 0 2 2 n z n + 1 ( 2 n + 1 ) 2 n n = z 1 z arctan z 1 z .
Since arctan z 1 z = arcsin ( z ) for 0 z < 1 , the above identity could be stated equivalently using the arcsine function as in (1). In this paper, however, we have decided to work with the notation used by Sprugnoli. Our first goal is to give a rigorous proof of (2).
Theorem 1.
For all 0 z < 1 we have the identity
n = 0 2 2 n z n + 1 ( 2 n + 1 ) 2 n n = z 1 z arctan z 1 z .
Proof. 
For each integer n 0 , consider the integral 1 1 ( 1 x 2 ) n d x . Then, we can evaluate the integral in two ways. First, we have
1 1 ( 1 x 2 ) n d x = 2 2 n + 1 ( 2 n + 1 ) 2 n n .
The result is known. It can be proved easily using integration by parts. It is, however, a special case of the more interesting fact [1] (p. 52) that
a b ( x a ) n ( b x ) n d x = 2 · 2 · 4 · 6 ( 2 n ) 3 · 5 · 7 ( 2 n + 1 ) · b a 2 2 n + 1 , n 1 .
By the binomial theorem, the integral can be also evaluated as
1 1 ( 1 x 2 ) n d x = 1 1 k = 0 n n k ( 1 ) k x 2 k d x = 2 k = 0 n n k ( 1 ) k 2 k + 1 .
Hence,
n = 0 2 2 n z n ( 2 n + 1 ) 2 n n = n = 0 k = 0 n n k ( 1 ) k z n 2 k + 1 = k = 0 ( 1 ) k 2 k + 1 n = k n k z n = k = 0 ( 1 ) k z k 2 k + 1 n = 0 n + k k z n = k = 0 ( 1 ) k z k 2 k + 1 n = 0 n + k n z n = k = 0 ( 1 ) k z k 2 k + 1 1 ( 1 z ) k + 1 .
This shows that
n = 0 2 2 n z n + 1 ( 2 n + 1 ) 2 n n = n = 0 ( 1 ) n 2 n + 1 z 1 z n + 1 .
Combining this with the fact that
n = 0 ( 1 ) n 2 n + 1 x 2 n + 1
is the Taylor series of the arctangent for 1 < x < 1 , and letting x = z 1 z , we find that identity (2) holds for 0 z 1 2 . However, since the series has a radius of convergence 1, by analytic continuation, the identity in question holds for every z [ 0 ; 1 ) . □
Differentiating both sides of (3) with respect to z gives
g 0 ( z ) = 1 2 ( 1 z ) + 1 2 ( 1 z ) 2 arctan z 1 z z 1 z , z [ 0 ; 1 ) .
Differentiating once more and multiplying by z gives
g 1 ( z ) = 2 z + 1 4 ( 1 z ) 2 + 4 z 1 4 ( 1 z ) 3 arctan z 1 z z 1 z , z [ 0 ; 1 ) .
Moreover, we see that
g 0 ( z ) = 1 ( 1 z ) 2 n = 0 ( 1 ) n ( n + 1 ) 2 n + 1 z 1 z n ,
and
g 1 ( z ) = 1 ( 1 z ) 3 n = 0 ( 1 ) n ( n + 1 ) 2 n + 1 ( 2 z + n ) z 1 z n .
The trigonometric versions of g 0 ( z ) and g 1 ( z ) are also useful; namely,
g t , 0 ( z ) = n = 0 2 2 n sin 2 n z ( 2 n + 1 ) C n = 1 2 1 cos 2 z + z cos 3 z sin z = 1 2 cos 2 z 1 + 2 z sin 2 z
and
g t , 1 ( z ) = n = 0 2 2 n n sin 2 n z ( 2 n + 1 ) C n = 1 4 cos 4 z 2 cos 2 z + ( 1 2 cos 2 z ) 2 z sin 2 z ,
both valid for 0 z < π 2 .
At z = π 3 , z = π 4 and z = π 6 , function g t , 0 ( z ) , respectively, gives
n = 0 3 n ( 2 n + 1 ) C n = 2 + 8 π 3 9 , n = 0 2 n ( 2 n + 1 ) C n = 1 + π 2 ,
n = 0 1 ( 2 n + 1 ) C n = 2 3 + 4 π 3 27 .
We also find from (5) or (6) and the principal branch for the square-root function
n = 1 n ( 2 n + 1 ) C n = 2 3 , n = 1 ( 1 ) n n ( 2 n + 1 ) C n = 2 25 1 16 5 ln α ,
n = 1 2 n n ( 2 n + 1 ) C n = 2 + π 2 , n = 1 ( 1 ) n n 2 n ( 2 n + 1 ) C n = 3 9 ln 2 3 ,
n = 1 3 n n ( 2 n + 1 ) C n = 10 + 32 π 3 3 , n = 1 ( 1 ) n n 3 n ( 2 n + 1 ) C n = 2 49 1 16 21 ln 5 21 2 .
To offer some evaluations of g 0 ( z ) involving Fibonacci and Lucas numbers, we need the following lemma.
Lemma 1
([29] Lemma 1, see also [30] p. 271, identities (20)–(22)). We have
sin π 10 = β 2 , sin 3 π 10 = α 2 , cos π 10 = α 5 2 , cos 3 π 10 = 1 2 5 α 5 .
Theorem 2.
With ω = 5 α = 2 + α , we have for any integer s
5 n = 0 F 2 n + s ( 2 n + 1 ) C n = 2 L s + 1 + 4 5 π ω 25 ( 2 + 2 α ) F s + ( 4 α ) F s 1 ,
n = 0 L 2 n + s ( 2 n + 1 ) C n = 2 F s + 1 + 4 5 π ω 25 2 F s + α F s 1 .
Proof. 
Set z = 3 π 10 in g t , 0 ( z ) and multiply through by α s , where s is an arbitrary integer. Using Lemma 1 yields
n = 0 α 2 n + s ( 2 n + 1 ) C n = 2 5 α s + 1 + 12 π 125 5 α s α .
In a similar manner, set z = π 10 in g t , 0 ( z ) and multiply through by β s to obtain
n = 0 β 2 n + s ( 2 n + 1 ) C n = 2 5 β s + 1 + 4 π 125 5 β s α .
The difference and sum of the above identities result in the identities
n = 0 F 2 n + s ( 2 n + 1 ) C n = 2 5 L s + 1 + 4 π 25 α 5 ( α L s + 2 + L s 1 ) ,
n = 0 L 2 n + s ( 2 n + 1 ) C n = 2 F s + 1 + 4 π 125 α 5 α ( F s + 1 + L s ) + F s + L s + 1 .
The stated identities follow upon simplifications. □
As examples, we have with ω = 5 α :
n = 1 F 2 n ( 2 n + 1 ) C n = 2 5 + 4 ( 7 α 6 ) π ω 125 , n = 1 L 2 n ( 2 n + 1 ) C n = 4 ( α + 2 ) π ω 25 ,
n = 1 F 2 n + 1 ( 2 n + 1 ) C n = 1 5 + 8 ( 3 α + 1 ) π ω 125 , n = 1 L 2 n + 1 ( 2 n + 1 ) C n = 1 + 8 5 π ω 25 ,
n = 1 F 2 n 2 ( 2 n + 1 ) C n = 3 5 + 8 ( 4 α 7 ) π ω 125 , n = 1 L 2 n 2 ( 2 n + 1 ) C n = 1 + 8 ( 3 α ) π ω 25 .
Using the same idea for g t , 1 ( z ) , we can prove the next theorem, whose proof we therefore omit.
Theorem 3.
With ω = 5 α = 2 + α , we have for any integer s
n = 1 n F 2 n + s ( 2 n + 1 ) C n = 2 F s + 1 + 8 5 F s + 8 5 π ω 125 5 α F s + 1 + 2 F s
and
n = 1 n L 2 n + s ( 2 n + 1 ) C n = 2 L s + 1 + 8 5 L s + 8 5 π ω 125 ( 6 + α ) F s + 1 + 2 α 2 F s .
Another interesting example for an evaluation of g 1 ( z ) with Fibonacci (Lucas) entries is the following result.
Theorem 4.
Let r be an even integer and s any integer. Then,
n = 1 2 2 n n F r n + s ( 2 n + 1 ) L r n C n = F 3 r + s + L r 2 F 2 r + s L r 2 + L r L 2 r + s 4 L 3 r + s L r 2 4 5 arctan ( β r ) + 4 L 3 r + s L r L 2 r + s + ( 4 F 3 r + s L r F 2 r + s ) 5 π L r 2 16 5
and
n = 1 2 2 n n L r n + s ( 2 n + 1 ) L r n C n = L 3 r + s + L r 2 L 2 r + s L r 2 + L r F 2 r + s 4 F 3 r + s L r 2 5 4 arctan ( β r ) + 4 L 3 r + s L r L 2 r + s + ( 4 F 3 r + s L r F 2 r + s ) 5 π L r 2 16 .
Proof. 
First note that 1 α r / L r = β r / L r and that if r is an even integer, then
α r / L r 1 α r / L r = α r L r α r = α r .
Let s be an arbitrary integer. Consider α s g 1 ( α r / L r ) β s g 1 ( β r / L r ) . We have
n = 1 2 2 n n ( α r n + s β r n + s ) L r n ( 2 n + 1 ) C n = ( α 3 r + s β 3 r + s ) L r 2 + ( α 2 r + s β 2 r + s ) L r 2 4 + 4 α 2 r + s β r α r + s β r L r L r 2 4 arctan ( α r ) 4 β 2 r + s α r β r + s α r L r L r 2 4 arctan ( β r ) ,
from which the stated identities in the theorem follow using the Binet formulas and the fact that arctan α r = π 2 arctan β r for any even integer r. □
As particular instances of Theorem 4, we have, for an even integer r,
n = 1 2 2 n + 1 n F ( n 2 ) r ( 2 n + 1 ) L r n C n = F 2 r + ( L r + 2 F r 5 ) π L r 2 4 5 L r 3 5 arctan ( β r ) ,
n = 1 2 2 n n L ( n 2 ) r ( 2 n + 1 ) L r n C n = L r 2 + ( L r + 2 F r 5 ) π L r 2 8 F r L r 2 5 arctan ( β r ) ,
n = 1 2 2 n n F ( n 3 ) r ( 2 n + 1 ) L r n C n = 1 4 L r 2 F r + 8 L r 2 + F 2 r 5 π L r 2 5 80 + L r 2 8 L r 2 5 arctan ( β r ) 20 ,
and
n = 1 2 2 n n L ( n 3 ) r ( 2 n + 1 ) L r n C n = 2 + 1 2 L r 2 L r 2 + 8 ( 1 ) r L r 2 + F 2 r 5 π L r 2 16 1 4 L r 3 F r 5 arctan ( β r ) .
Note that both (8) and (9) give (7) when r = 0 .

3. Integral Expressions for g 0 ( z ) and g 1 ( z )

Integral expressions for the functions g 0 ( z ) and g 1 ( z ) are derived easily using the integral identity (4).
Theorem 5.
We have
g 0 ( z ) = 1 2 1 1 1 ( 1 z ( 1 x 2 ) ) 2 d x ,
g 1 ( z ) = z 1 1 1 x 2 ( 1 z ( 1 x 2 ) ) 3 d x .
Proof. 
From the geometric series and the above lemma, we deduce that for all 0 z < 1 ,
n = 0 2 2 n z n + 1 ( 2 n + 1 ) 2 n n = 1 2 1 1 z 1 z ( 1 x 2 ) d x .
Differentiating produces the first equation. To obtain the second equation, we perform the operation z ( d / d z ) g 0 ( z ) , and the proof is completed. □
It is interesting to compare the integral expressions for g 0 ( z ) and g 1 ( z ) with that for f ( z ) . This expression is not stated explicitly in [21] but can be derived as follows:
n = 1 z n C n = n = 1 ( n + 1 ) 2 n n z n = n = 1 n ( n + 1 ) Γ ( n ) Γ ( n + 1 ) Γ ( 2 n + 1 ) z n = n = 1 n ( n + 1 ) B ( n , n + 1 ) z n = n = 1 n ( n + 1 ) 0 1 x n 1 ( 1 x ) n d x z n = 2 z 0 1 1 x ( 1 z x ( 1 x ) ) 3 d x ,
where B ( a , b ) is the beta function B ( a , b ) = 0 1 x a 1 ( 1 x ) b 1 d x = Γ ( a ) Γ ( b ) Γ ( a + b ) .
This proves that
f ( z ) = 1 + 2 z 0 1 1 x 1 z x ( 1 x ) 3 d x .

4. Some General Properties of g m ( z )

In this section, we present some general properties of g m ( z ) , which is defined by
g m ( z ) = n = 0 2 2 n n m 2 n + 1 z n C n , 0 z < 1 ,
with m 0 being an integer. We have the following result:
Theorem 6.
For each m 0 and all 0 z < 1 , g m ( z ) possesses the representation
g m ( z ) = P 1 , m ( z ) ( 1 z ) m + 1 + P 2 , m ( z ) ( 1 z ) m + 2 arctan z 1 z z 1 z ,
where P 1 , m ( z ) and P 2 , m ( z ) are polynomials in z of degree m with rational coefficients.
Moreover, for m 1 , the polynomials P 1 , m ( z ) and P 2 , m ( z ) can be expressed recursively according to
P 1 , m ( z ) = z ( 1 z ) d d z P 1 , m 1 ( z ) + m z P 1 , m 1 ( z ) + 1 2 P 2 , m 1 ( z ) ,
P 2 , m ( z ) = z ( 1 z ) d d z P 2 , m 1 ( z ) + ( m + 1 ) z P 2 , m 1 ( z ) 1 2 P 2 , m 1 ( z ) ,
with P 1 , 0 ( z ) = P 2 , 0 ( z ) = 1 2 .
Proof. 
The proof of the representation (10) is easy using induction on m taking into account g m + 1 ( z ) = z ( d / d z ) g m ( z ) = ( z d / d z ) m g 0 ( z ) and the identity
d d z arctan z 1 z z 1 z = 1 2 z 1 2 z ( 1 z ) arctan z 1 z z 1 z .
The recursive expressions for P 1 , m ( z ) and P 2 , m ( z ) follow from the proof as a by-product. □
The first few polynomials have the following explicit forms:
P 1 , 1 ( z ) = 1 4 ( 2 z + 1 ) , P 2 , 1 ( z ) = 1 4 ( 4 z 1 ) ,
P 1 , 2 ( z ) = 1 8 ( 4 z 2 + 12 z 1 ) , P 2 , 2 ( z ) = 1 8 ( 16 z 2 2 z + 1 ) .
We mention that the coupled recursions (11) and (12) can be solved explicitly, but the closed forms seem not to shed enough light on their general structure. Nevertheless, we can prove the following expressions:
Proposition 1.
For each m,
P 1 , m ( z ) = 1 2 m ! z m + j = 0 m 1 m j j ! z j z ( 1 z ) d d z P 1 , m ( j + 1 ) ( z ) + 1 2 P 2 , m ( j + 1 ) ( z ) ,
P 2 , m ( z ) = 1 2 j = 1 m ( j + 1 ) z 1 2 + z ( 1 z ) j = 1 m d d z P 2 , m j ( z ) k = 2 j ( m + 3 k ) z 1 2 ,
where the empty product is one and the empty sum is zero.
Proof. 
We can use induction on m to prove both formulas. For m = 0 , the statements are true. The inductive step for P 1 , m ( z ) is
P 1 , m + 1 ( z ) = ( m + 1 ) z P 1 , m ( z ) + z ( 1 z ) d d z P 1 , m ( z ) + 1 2 P 2 , m ( z ) = 1 2 ( m + 1 ) ! z m + 1 + j = 0 m 1 m j j ! ( m + 1 ) z j + 1 z ( 1 z ) d d z P 1 , m ( j + 1 ) ( z ) + 1 2 P 2 , m ( j + 1 ) ( z ) + z ( 1 z ) d d z P 1 , m ( z ) + 1 2 P 2 , m ( z ) = 1 2 ( m + 1 ) ! z m + 1 + j = 1 m 1 m + 1 j + 1 ( j + 1 ) ! z j + 1 z ( 1 z ) d d z P 1 , m ( j + 1 ) ( z ) + 1 2 P 2 , m ( j + 1 ) ( z ) = 1 2 ( m + 1 ) ! z m + 1 + j = 0 m m + 1 j j ! z j z ( 1 z ) d d z P 1 , m j ( z ) + 1 2 P 2 , m j ( z ) .
Similarly, the inductive proof for P 2 , m ( z ) is accomplished according to
P 2 , m + 1 ( z ) = ( m + 2 ) z 1 2 P 2 , m ( z ) + z ( 1 z ) d d z P 2 , m ( z ) = ( m + 2 ) z 1 2 1 2 j = 1 m ( j + 1 ) z 1 2 + z ( 1 z ) j = 1 m d d z P 2 , m j ( z ) k = 2 j ( m + 3 k ) z 1 2 + z ( 1 z ) d d z P 2 , m ( z ) = 1 2 j = 1 m + 1 ( j + 1 ) z 1 2 + z ( 1 z ) d d z P 2 , m ( z ) + j = 1 m d d z P 2 , m j ( z ) ( m + 2 ) z 1 2 k = 2 j ( m + 3 k ) z 1 2 = 1 2 j = 1 m + 1 ( j + 1 ) z 1 2 + z ( 1 z ) j = 0 m d d z P 2 , m j ( z ) k = 1 j ( m + 3 k ) z 1 2 = 1 2 j = 1 m + 1 ( j + 1 ) z 1 2 + z ( 1 z ) j = 1 m + 1 d d z P 2 , m ( j 1 ) ( z ) k = 1 j 1 ( m + 3 k ) z 1 2 = 1 2 j = 1 m + 1 ( j + 1 ) z 1 2 + z ( 1 z ) j = 1 m + 1 d d z P 2 , m + 1 j ( z ) k = 2 j ( m + 4 k ) z 1 2 .
Applying Theorem 6 in the case m = 2 yields
n = 1 2 2 n n 2 2 n + 1 z n C n = 4 z 2 + 12 z 1 8 ( 1 z ) 3 + 16 z 2 2 z + 1 8 ( 1 z ) 4 arctan z 1 z z 1 z ,
from which we obtain
n = 1 n 2 ( 2 n + 1 ) C n = 2 3 + 8 3 π 81 ,
n = 1 2 n n 2 ( 2 n + 1 ) C n = 6 + 2 π , n = 1 3 n n 2 ( 2 n + 1 ) C n = 82 + 272 3 π 9 .
Corollary 1.
For each m 0 , the sum n = 0 2 n n m ( 2 n + 1 ) C n can be expressed in the form a + b π , with a and b rational. The sums n = 0 n m ( 2 n + 1 ) C n and n = 0 3 n n m ( 2 n + 1 ) C n allow the same representation but with b being irrational.
Theorem 7.
For each m 0 and all | z | < 1 , g m ( z ) possesses the integral representation
g m ( z ) = 1 1 Q m ( z ; x ) ( 1 z ( 1 x 2 ) ) m + 2 d x ,
where Q m ( z ; x ) is a polynomial in z of degree m given by
Q m ( z ; x ) = 1 2 ( m + 1 ) ! ( a z ) m + ( 1 a z ) z d d z Q m 1 ( z ; x ) + ( 1 a z ) z j = 1 m 1 m + 1 j j ! ( a z ) j d d z Q m ( j + 1 ) ( z ; x ) ,
with Q 0 ( z ; x ) = 1 2 , and where we have set a = 1 x 2 .
Proof. 
We prove the claim by induction on m. Since
g 0 ( z ) = 1 2 1 1 1 ( 1 z ( 1 x 2 ) ) 2 d x ,
the statement is true for m = 0 . Now, assuming it is true for a fixed m > 0 , we can proceed with
g m + 1 ( z ) = z d d z g m ( z ) = z 1 1 d d z Q m ( z ; x ) ( 1 a z ) + ( m + 2 ) a Q m ( z ; x ) 1 z ( 1 x 2 ) m + 3 d x .
This gives the recursion
Q m ( z ; x ) = ( m + 1 ) a z Q m 1 ( z ; x ) + z ( 1 a z ) d d z Q m 1 ( z ; x ) , m 1 .
This recursion can be solved by standard methods to give (13). Alternatively, one can prove (13) directly by induction on m using (14). □

5. Another Integral Expression for g m ( z ) Using Mellin Transform

Lemma 2.
For integers m , n 0 , we have
j = 0 m ( 1 ) j S ( m + 1 , m + 1 j ) ( n + m j ) ! = n m n ! ,
where S ( n , k ) are the Stirling numbers of the second kind, defined by S ( 0 , 0 ) = 1 , S ( n , n ) = S ( n , 1 ) = 1 ( n 1 ) and
S ( n , k ) = 1 k ! s = 0 k ( 1 ) s k s ( k s ) n .
Proof. 
Consider the known representation
x m = j = 0 m x j S ( m , j ) j ! .
Let x = n . Using s j = ( 1 ) j s + j 1 j we have
( 1 ) m + 1 n m + 1 = j = 0 m + 1 n j S ( m + 1 , j ) j ! = j = 0 m + 1 ( 1 ) j n + j 1 j S ( m + 1 , j ) j ! = j = 0 m + 1 ( 1 ) j ( n + j 1 ) ! ( n 1 ) ! S ( m + 1 , j ) .
Thus, by reindexing the summation
( 1 ) m + 1 n m + 1 ( n 1 ) ! = j = 0 m + 1 ( 1 ) m j + 1 S ( m + 1 , m + 1 j ) ( n + m j ) ! = j = 0 m ( 1 ) m j + 1 S ( m + 1 , m + 1 j ) ( n + m j ) !
as S ( n , 0 ) = 0 , n 1 . □
Theorem 8.
The function g m ( z ) possesses the integral representation
g m ( z ) = 1 z j = 0 m ( 1 ) j S ( m + 1 , m + 1 j ) 0 x ( m j ) / 2 K j + 1 m ( 2 x ) sinh ( 2 x z ) d x ,
where K v ( x ) is the modified Bessel function of the second kind, which can be defined by
K v ( x ) = 0 cosh ( v t ) e x   cosh t d t ( x > 0 ) .
Proof. 
The proof is based on ideas developed in [21]. Recall that the Mellin transform of a real-valued function f ( x ) on ( 0 , ) is defined by the integral [31]
M [ f ( x ) ] ( s ) = 0 x s 1 f ( x ) d x .
The gamma function Γ ( n ) can be interpreted as M [ e x ] ( n ) and thus
M [ x e x ] ( n + 1 ) = ( n + 1 ) ! .
Since
g m ( z ) = n = 0 ( 4 z ) n ( 2 n + 1 ) ! n m ( n + 1 ) ! n ! ,
we want to find a function f m ( x ) such that
n m n ! ( n + 1 ) ! = M [ f m ( x ) ] M [ x e x ] ( n + 1 ) .
By Lemma 2, it follows that such a function is
f m ( x ) = j = 0 m ( 1 ) j S ( m + 1 , m + 1 j ) x m j e x .
Now, we are going to apply the Mellin convolution theorem:
M [ f 1 ( x ) ] M [ f 2 ( x ) ] ( s ) = M [ F ( x ) ] ( s )
with
F ( x ) = 0 f 1 ( x 1 ) f 2 x x 1 d x 1 x 1 .
In our case, F ( x ) equals
F ( x ) = 0 j = 0 m ( 1 ) j S ( m + 1 , m + 1 j ) x 1 m j e x 1 x x 1 e x / x 1 d x 1 x 1 = j = 0 m ( 1 ) j S ( m + 1 , m + 1 j ) 0 x e ( x 1 + x / x 1 ) x 1 j + 2 m d x 1 = 2 j = 0 m ( 1 ) j S ( m + 1 , m + 1 j ) x ( m + 1 j ) / 2 K j + 1 m ( 2 x ) ,
where the following representation for the modified Bessel function of the second kind [21] was used:
K v ( z ) = 1 2 z 2 v 0 e ( t + z 2 / 4 t ) t v + 1 d t .
Finally, we calculate
g m ( z ) = n = 0 ( 4 z ) n ( 2 n + 1 ) ! n m ( n + 1 ) ! n ! = n = 0 ( 4 z ) n ( 2 n + 1 ) ! M [ F ( x ) ] ( n + 1 ) = 0 F ( x ) n = 0 ( 4 x z ) n ( 2 n + 1 ) ! d x = 0 F ( x ) sinh ( 2 x z ) 2 x z d x .
Two special cases of the representation are
g 0 ( z ) = 1 z 0 sinh ( 2 x z ) K 1 ( 2 x ) d x
and
g 1 ( z ) = 1 z 0 sinh ( 2 x z ) x K 0 ( 2 x ) K 1 ( 2 x ) d x .

6. Concluding Comments

In this paper, we have studied an interesting family of infinite series involving Catalan numbers. In particular, we have evaluated these series for special arguments and provided characterizations. Before closing, we want to state two different approaches that were communicated to us by one of the referees. First, if we set r ( z ) = z arctan z and k ( z ) = z 1 z , then h ( z ) = r ( z ) k ( z ) , where h ( z ) equals identity (3). This shows that the functions g m ( z ) can also be studied by the Faà di Bruno formula. Furthermore, as g m + 1 ( z ) = z d d z g m ( z ) = ( z d d z ) m g 0 ( z ) , it is possible to study g m ( z ) by expanding the factors ( z d d z ) m according to
z d d z n = k S ( n , k ) z k d d z k ,
where S ( n , k ) are the Stirling numbers of the second kind.
Finally, we remark that Sprugnoli’s identity (2), which is the starting point of our exploration, can be integrated resulting in the identity
n = 0 2 2 n ( 2 n + 1 ) ( n + 1 ) ( n + 2 ) z n C n = 1 2 z + 1 2 z 2 arctan 2 z 1 z arctan z 1 z z z 1 z ,
and containing the evaluation
n = 0 2 n ( 2 n + 1 ) ( n + 1 ) ( n + 2 ) C n = π 2 8 π 2 + 1
as a special instance (at z = 1 2 ).

Author Contributions

All authors contributed equally to the content of the paper. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors are indebted to the anonymous reviewers for their detailed review and many very useful suggestions and remarks, including the aspects for future research given in the last section.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Koshy, T. Catalan Numbers with Applications; Oxford University Press: Oxford, UK, 2009. [Google Scholar]
  2. Roman, S. An Introduction to Catalan Numbers; Birkhäuser: Basel, Switzerland, 2015. [Google Scholar]
  3. Stanley, R.P. Catalan Numbers; Cambridge University Press: Cambridge, UK, 2015. [Google Scholar]
  4. Barry, P. Generalized Catalan recurrences, Riordan arrays, elliptic curves, and orthogonal polynomials. J. Integer Seq. 2021, 24, 21.5.1. [Google Scholar]
  5. Chu, W. Alternating convolutions of Catalan numbers. Bull. Braz. Math. Soc. New Series. 2021, 53, 95–105. [Google Scholar] [CrossRef]
  6. Goy, T.; Shattuck, M. Determinant formulas of some Toeplitz–Hessenberg matrices with Catalan entries. Proc. Indian Acad. Sci. Math. Sci. 2019, 129, 46. [Google Scholar] [CrossRef]
  7. Kim, A. Convolution sums with Catalan numbers. J. Ramanujan Math. Soc. 2020, 35, 307–315. [Google Scholar]
  8. Lin, G.D. On powers of the Catalan number sequence. Discrete Math. 2019, 342, 2139–2147. [Google Scholar] [CrossRef] [Green Version]
  9. Qi, F.; Guo, B.-N. Integral representations of the Catalan numbers and their applications. Mathematics 2017, 5, 40. [Google Scholar] [CrossRef] [Green Version]
  10. Zhang, W.; Chen, L. On the Catalan numbers and some of their identities. Symmetry 2019, 11, 62. [Google Scholar] [CrossRef] [Green Version]
  11. Adegoke, K.; Frontczak, R.; Goy, T. Some special sums with squared Horadam numbers and generalized tribonacci numers. Palest. J. Math. 2022, 11, 66–73. [Google Scholar]
  12. Bayad, A.; Hajli, M. On the multidimensional zeta functions associated with theta functions, and the multidimensional Appell polynomials. Math. Methods Appl. Sci. 2020, 43, 2679–2694. [Google Scholar] [CrossRef]
  13. Bouarroudj, S.; Hajli, M. On the explicit formulas for zeta functions. Math. Methods Appl. Sci. 2020, 43, 10249–10261. [Google Scholar] [CrossRef]
  14. Frontczak, R.; Goy, T. Chebyshev–Fibonacci polynomial relations using generating functions. Integers. 2021, 21, #A100. [Google Scholar]
  15. Frontczak, R.; Goy, T. Mersenne–Horadam identities using generating functions. Carpathian Math. Publ. 2020, 12, 34–45. [Google Scholar] [CrossRef]
  16. Frontczak, R.; Goy, T. More Fibonacci–Bernoulli relations with and without balancing polynomials. Math. Commun. 2021, 26, 215–226. [Google Scholar]
  17. Hajli, M. On a formula for the regularized determinant of zeta functions with application to some Dirichlet series. Q. J. Math. 2020, 71, 843–865. [Google Scholar] [CrossRef]
  18. Kruchinin, D.; Kruchinin, V.; Shablya, Y. Method for obtaining coefficients of powers of bivariate generating functions. Mathematics 2021, 9, 428. [Google Scholar] [CrossRef]
  19. Simsek, Y. Construction method for generating functions of special numbers and polynomials arising from analysis of new operators. Math. Methods Appl. Sci. 2018, 41, 6934–6954. [Google Scholar] [CrossRef]
  20. Simsek, Y. Generating functions for finite sums involving higher powers of binomial coefficients: Analysis of hypergeometric functions including new families of polynomials and numbers. J. Math. Anal. Appl. 2019, 477, 1328–1352. [Google Scholar] [CrossRef] [Green Version]
  21. Amdeberhan, T.; Guan, X.; Jiu, L.; Moll, V.H.; Vignat, C. A series involving Catalan numbers: Proofs and demonstrations. Elem. Math. 2016, 71, 109–121. [Google Scholar] [CrossRef] [Green Version]
  22. Yin, L.; Qi, F. Several series identities involving the Catalan numbers. Trans. A Razmadze Math. Inst. 2018, 172, 466–474. [Google Scholar] [CrossRef]
  23. Koshy, T.; Gao, Z.G. Convergence of a Catalan series. College Math. J. 2012, 43, 141–146. [Google Scholar] [CrossRef]
  24. Beckwith, D.; Harbor, S. Problem 11765. Amer. Math. Monthly 2014, 121, 267. [Google Scholar]
  25. Abel, U. Reciprocal Catalan sums: Solution to Problem 11765. Amer. Math. Monthly 2016, 123, 405–406. [Google Scholar]
  26. Stewart, S.M. The inverse versine function and sums containing reciprocal central binomial coefficients and reciprocal Catalan numbers. Int. J. Math. Educ. Sci. Technol. 2021. [Google Scholar] [CrossRef]
  27. Koshy, T. Fibonacci and Lucas Numbers with Applications; John Wiley & Sons: New York, NY, USA, 2001. [Google Scholar]
  28. Sprugnoli, R. Sums of reciprocals of the central binomial coefficients. Integers 2006, 6, #A27. [Google Scholar]
  29. Adegoke, K. Fibonacci identities involving reciprocals of binomial coefficients. arXiv 2021, arXiv:2112.00622. [Google Scholar]
  30. Srivastava, H.M.; Choi, J. Zeta and q-Zeta Functions and Associated Series and Integrals; Elsevier: Amsterdam, The Netherlands, 2012. [Google Scholar]
  31. Debnath, L.; Bhatta, D. Integral Transforms and Their Applications, 3rd ed.; Chapman & Hall/CRC: Boca Raton, FL, USA, 2014. [Google Scholar]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Adegoke, K.; Frontczak, R.; Goy, T. On a Family of Infinite Series with Reciprocal Catalan Numbers. Axioms 2022, 11, 165. https://doi.org/10.3390/axioms11040165

AMA Style

Adegoke K, Frontczak R, Goy T. On a Family of Infinite Series with Reciprocal Catalan Numbers. Axioms. 2022; 11(4):165. https://doi.org/10.3390/axioms11040165

Chicago/Turabian Style

Adegoke, Kunle, Robert Frontczak, and Taras Goy. 2022. "On a Family of Infinite Series with Reciprocal Catalan Numbers" Axioms 11, no. 4: 165. https://doi.org/10.3390/axioms11040165

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop