Next Article in Journal
Scheduling with Resource Allocation, Deteriorating Effect and Group Technology to Minimize Total Completion Time
Next Article in Special Issue
Characterizations of Continuous Fractional Bessel Wavelet Transforms
Previous Article in Journal
Management of Transfer Prices in Professional Football as a Function of Fan Numbers
Previous Article in Special Issue
Some Generalized Properties of Poly-Daehee Numbers and Polynomials Based on Apostol–Genocchi Polynomials
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Improper Integrals Involving Powers of Inverse Trigonometric and Hyperbolic Functions

1
School of Mathematics and Statistics, Zhoukou Normal University, Zhoukou 466001, China
2
Department of Mathematics and Physics, University of Salento, 73100 Lecce, Italy
*
Author to whom correspondence should be addressed.
Mathematics 2022, 10(16), 2980; https://doi.org/10.3390/math10162980
Submission received: 3 August 2022 / Revised: 14 August 2022 / Accepted: 15 August 2022 / Published: 18 August 2022

Abstract

:
Three classes of improper integrals involving higher powers of arctanh , arctan, and arcsin are examined using the recursive approach. Numerous explicit formulae are established, which evaluate these integrals in terms of π , ln 2 , the Riemann zeta function, and the Dirichlet beta function.

1. Introduction and Outline

The evaluation of integrals is an important subject in mathematics, physics and applied sciences. In the mathematical literature (see, for example, the monographs by Boros and Moll [1], and Vǎlean [2]), there are numerous intriguing integrals. We reproduce, for instance, the following elegant integrals involving inverse trigonometric and hyperbolic functions, where G denotes the Catalan’s constant:
Entry ( a ) 0 1 arctanh 3 x x d x = π 4 64 ,
                [3]
Entry ( b ) 0 1 arctanh 3 x x 2 d x = 3 2 ζ ( 3 ) ,
                [3]
Entry ( c ) 0 1 arctan x x d x = G = n = 0 ( 1 ) n ( 2 n + 1 ) 2 ,
              [4,5,6,7]
Entry ( d ) 0 1 arctan 2 x x d x = π G 2 7 8 ζ ( 3 ) ,
        ([8], (A.289)) and [3,5,6]
Entry ( e ) 0 1 arcsin x x d x = π 2 ln 2 ,
         ([9], §4.521: Equation (1))
Entry ( f ) 0 1 arcsin 2 x x d x = π 2 4 ln 2 7 8 ζ ( 3 ) .
           ([1], Equation 6.6.25)
Some related integrals of log-trigonometric functions are highlighted as follows:
Entry ( g ) G = 0 π 4 ln ( cot x ) d x = 2 0 π 4 ln ( 2 cos x ) d x ,
           [5,7,10,11]
Entry ( h ) π 2 ln 2 = 0 π 2 ln ( sec x ) d x = 0 π 2 ln ( csc x ) d x .
             ([4,12]
Euler (1772) discovered the identity (h) and the following remarkable value
0 π 2 x ln ( sin x ) d x = 7 ζ ( 3 ) 16 π 2 8 ln 2
by making use of the Fourier series
ln ( sin x ) = ln 2 n = 1 cos ( 2 n x ) n .
Koyama and Kurokawa [13] evaluated the integral below as well as the related indefinite integrals:
0 π 2 x 2 ln ( sin x ) d x = 3 ζ ( 3 ) 16 π 3 24 ln 2 .
Further integral identities of a similar nature can be found in the papers [14,15,16,17,18].
Motivated by these elegant formulae, we shall primarily investigate the following improper integrals with two integer parameters in this article:
H ( m , n ) : = 0 1 arctanh m x x n d x , T ( m , n ) : = 0 1 arctan m x x n d x , S ( m , n ) : = 0 1 arcsin m x x n d x ;
where m N 0 and n Z , subject to m n , so that all these integrals are convergent. By making use of recurrence relations and Fourier series expansions, we shall explicitly evaluate, in the next three sections, these three classes of integrals. Two classes of subsidiary integrals Θ ( m ) and Λ ( m ) regarding log-cosine and log-tangent functions will also be examined. Finally, the paper will conclude with a brief discussion of more integral evaluations in Section 5.
Throughout the paper, we shall utilize the following notations. Let Z and N stand, respectively, for the sets of integers and natural numbers with N 0 = N { 0 } . For n N 0 and an indeterminate x, the rising and falling factorials are defined by ( x ) 0 = x 0 = 1 and
( x ) n = x ( x + 1 ) ( x + n 1 ) x n = x ( x 1 ) ( x n + 1 ) } for n N .
Given i , j Z and m N , the symbol “ i m j ” represents that “i is congruent to j modulo m”. The logical function χ will also be employed with χ ( true ) = 1 and χ ( false ) = 0 . In addition, we need the following four functions:
Riemann zeta function : ζ ( x ) = n = 0 1 ( n + 1 ) x ( ( x ) > 1 ) ; Dirichlet lambda function : λ ( x ) = n = 0 1 ( 2 n + 1 ) x ( ( x ) > 1 ) ; Dirichlet eta function : η ( x ) = n = 0 ( 1 ) n ( n + 1 ) x ( ( x ) > 0 ) ; Dirichlet beta function : β ( x ) = n = 0 ( 1 ) n ( 2 n + 1 ) x ( ( x ) > 0 ) .

2. Evaluation of H ( m , n )

When n 1 , the following algebraic equality holds:
( n 1 ) ( n 3 ) x 2 ( n 1 ) x n = 1 x n n 3 ( n 1 ) x n 2 .
Then, we have the integral equation below
H ( m , n ) = 0 1 arctanh m x { ( n 1 ) ( n 3 ) x 2 } ( n 1 ) x n d x = 0 1 arctanh m x x n d x n 3 n 1 0 1 arctanh m x x n 2 d x = H ( m , n ) n 3 n 1 H ( m , n 2 ) .
Considering that
d d x 1 x 2 ( 1 n ) x n 1 = ( n 1 ) ( n 3 ) x 2 ( n 1 ) x n ,
we can alternatively express the integral H ( m , n ) , by integration by parts, as
H ( m , n ) = m n 1 0 1 arctanh m 1 x x n 1 d x = m n 1 H ( m 1 , n 1 ) .
Combining the two expressions of H ( m , n ) results in the following three-term recurrence relation
( n 1 ) H ( m , n ) m H ( m 1 , n 1 ) ( n 3 ) H ( m , n 2 ) = 0 .
According to this relation, to compute all the values H ( m , n ) for m N 0 and n Z with m n , we have to determine the boundary values { H ( 0 , n ) , H ( m , 0 ) , H ( m , 1 ) } .

2.1. H ( 0 , n )

For n 0 , we have the following obvious values:
H ( 0 , n ) = 0 1 x n d x = 1 1 n .

2.2. H ( m , 0 )

When m = 0 , it is easy to see that H ( 0 , 0 ) = 1 . For m 1 , by changing the variable x 1 y 1 + y , we have
H ( m , 0 ) = 0 1 1 2 ln 1 + x 1 x m d x = ( 1 ) m 2 m 1 0 1 ln m y ( 1 + y ) 2 d y .
According to the power series expansion
( 1 + y ) 2 = k = 0 ( 1 ) k ( k + 1 ) y k ( | y | < 1 ) ;
we can express
H ( m , 0 ) = ( 1 ) m 2 m 1 k = 0 ( 1 ) k ( k + 1 ) 0 1 y k ln m y d y .
By repeatedly applying integration by parts, we can evaluate the last integral
0 1 y k ln m y d y = i = 0 m ( 1 ) i m i ( k + 1 ) i + 1 y k + 1 ln m i y | 0 1 = ( 1 ) m m ! ( k + 1 ) m + 1 .
Hereafter, exchanges in the order of summation and integration are justified by Lebesgue’s dominated convergence theorem ([19], §11.32). By substitution, we can obtain the closed formula
H ( m , 0 ) = m ! 2 m 1 k = 0 ( 1 ) k ( k + 1 ) m = m ! 2 m 1 η ( m ) .

2.3. H ( m , 1 )

We can also evaluate H ( m , 1 ) by carrying out the same procedure as for H ( m , 0 ) . In fact, for m 1 , by making the change in variable x 1 y 1 + y , we can express
H ( m , 1 ) = 0 1 1 x 1 2 ln 1 + x 1 x m d x = ( 1 ) m 2 m 1 0 1 ln m y 1 y 2 d y .
We take the above integral as an example to show how to justify the term-by-term integration by making use of Lebesgue’s dominated convergence theorem. For any fixed x with 0 < x < 1 , we have the following power series expansion
1 1 y 2 = k = 0 y 2 k , where 0 y x .
Now, define the following sequence of functions
φ n ( y ) = ln m y k = 0 n y 2 k such that lim n φ n ( y ) = ln m y 1 y 2 for y [ 0 , x ] .
When 0 y x , we have
| φ n ( y ) | < | ln m y 1 y 2 | ( 1 ) m ln m y 1 x 2 ,
where the rightmost function is dominating and integrable over [ 0 , x ] . According to Lebesgue’s dominated convergence theorem, we can proceed using term-by-term integration
0 x ln m y 1 y 2 d y = lim n 0 x φ n ( y ) d y = lim n k = 0 n 0 x y 2 k ln m y d y = k = 0 i = 0 m ( 1 ) i m i ( 2 k + 1 ) i + 1 x 2 k + 1 ln m i x ,
where we have employed integral formula (3). Observing that the last series is uniformly convergent for x [ 1 2 , 1 ] , we can evaluate the series through the term-by-term limit, as follows:
H ( m , 1 ) = ( 1 ) m 2 m 1 lim x 1 0 x ln m y 1 y 2 d y = ( 1 ) m 2 m 1 lim x 1 k = 0 i = 0 m ( 1 ) i m i ( 2 k + 1 ) i + 1 x 2 k + 1 ln m i x = ( 1 ) m 2 m 1 k = 0 i = 0 m ( 1 ) i m i ( 2 k + 1 ) i + 1 lim x 1 x 2 k + 1 ln m i x = m ! 2 m 1 k = 0 1 ( 2 k + 1 ) m + 1 .
This gives rise to the below formula
H ( m , 1 ) = m ! 2 m 1 λ ( m + 1 ) .
In conclusion, we have shown the following general theorem. Its special case H ( 3 , n ) was studied by Sofo and Nimbran [3].
Theorem 1
( m N 0 and n Z with m n ).
n = 0 H ( m , 0 ) = 1 , m = 0 ; m ! 2 m 1 η ( m ) , m 1 ; n = 1 H ( m , 1 ) = m ! 2 m 1 λ ( m + 1 ) , m 1 ; n 2 H ( m , n ) = m n 1 H ( m 1 , n 1 ) + n 3 n 1 H ( m , n 2 ) , m n ; n < 0 H ( m , n ) = 1 1 n , m = 0 ; n + 1 n 1 H ( m , n + 2 ) m n 1 H ( m 1 , n + 1 ) , m 1 .
Keeping in mind that η ( k ) and λ ( k ) can be written in terms of the zeta function (except for η ( 1 ) = ln 2 ), we assert, according to this theorem, that H ( m , n ) is always expressible as a linear combination of ln 2 and zeta values. The integral values for H ( m , n ) with 1 m 5 and 5 n 5 are recorded in Table 1.

3. Evaluation of T ( m , n )

Supposing n 1 , by making use of integration by parts, we can obtain
T ( m , n ) = 1 1 n π 4 m + m n 1 0 1 arctan m 1 x x n 1 ( 1 + x 2 ) d x .
When m 0 , the above integral can further be manipulated as
0 1 arctan m 1 x x n 1 ( 1 + x 2 ) d x = 0 1 arctan m 1 x x n 1 d x 0 1 arctan m 1 x x n 3 ( 1 + x 2 ) d x = T ( m 1 , n 1 ) n 3 m T ( m , n 2 ) 1 m π 4 m .
Substituting this into (6), and then simplifying the resulting expression, we can derive the following three-term recurrence relation
( n 1 ) T ( m , n ) m T ( m 1 , n 1 ) + ( n 3 ) T ( m , n 2 ) + 2 π 4 m = 0 .
Based on this relation, to calculate all the T ( m , n ) for m N 0 and n Z with m n , it is sufficient to determine the boundary values { T ( 0 , n ) , T ( m , 0 ) , T ( m , 1 ) } .

3.1. T ( 0 , n )

Firstly, it is trivial to check for n 0 that
T ( 0 , n ) = 0 1 x n d x = 1 1 n .

3.2. T ( m , 0 )

Then for m = 0 , 1 , we have no difficulty evaluating
T ( 0 , 0 ) = 0 1 d x = 1 , T ( 1 , 0 ) = 0 1 arctan x d x = π 4 ln 2 2 .
When m 2 , applying integration by parts twice shows that
T ( m , 0 ) = 0 1 arctan m x d x = π 4 m m 0 1 x arctan m 1 x 1 + x 2 d x = π 4 m m ln 2 2 π 4 m 1 + m ( m 1 ) 2 0 1 ln ( 1 + x 2 ) 1 + x 2 arctan m 2 d x .
Then under the change in variable x tan y , the last expression becomes
T ( m , 0 ) = π 4 m m ln 2 2 π 4 m 1 m ( m 1 ) Θ ( m 2 ) ,
where Θ ( m ) stands for the parametric integral
Θ ( m ) = 0 π 4 y m ln ( cos y ) d y with m 0 .

3.3. Θ ( m )

To evaluate the integral Θ ( m ) , we recall the following Fourier series (cf. [9], §1.441)
ln ( cos y ) = ln 2 k = 1 ( 1 ) k k cos ( 2 k y ) , where | y | < π 2 .
Then, using Lebesgue’s dominated convergence theorem ([19], §11.32), we can express
Θ ( m ) = 0 π 4 y m ln ( cos y ) d y = ln 2 m + 1 π 4 m + 1 + k = 1 ( 1 ) k 1 k 0 π 4 y m cos ( 2 k y ) d y .
Applying integration by parts for m times, we can evaluate the last integral as follows:
I ( m , k ) = 0 π 4 y m cos ( 2 k y ) d y = j = 0 m m j ( 2 k ) j + 1 y m j sin 2 k y + j π 2 | 0 π 4 = m ! ( 2 k ) m + 1 sin m + 2 2 π + j = 0 m m j ( 2 k ) j + 1 π 4 m j sin k + j 2 π .
Taking into account the trigonometric identity
sin ( α + β ) = sin α cos β + cos α sin β ,
we can reformulate the infinite series
k = 1 ( 1 ) k 1 k I ( m , k ) = m ! 2 m + 1 sin m + 2 2 π k = 1 ( 1 ) k 1 k m + 2 + j = 0 m m j 2 j + 1 π 4 m j cos j π 2 k = 1 ( 1 ) k 1 k j + 2 sin k π 2 + j = 0 m m j 2 j + 1 π 4 m j sin j π 2 k = 1 ( 1 ) k 1 k j + 2 cos k π 2 .
Observing that
sin k π 2 = ( 1 ) k 1 2 χ ( k 2 1 ) and cos k π 2 = ( 1 ) k 2 χ ( k 2 0 ) ,
we have
k = 1 ( 1 ) k 1 k j + 2 sin k π 2 = k = 1 ( 1 ) k 1 ( 2 k 1 ) j + 2 = β ( j + 2 ) , k = 1 ( 1 ) k 1 k j + 2 cos k π 2 = k = 1 ( 1 ) k 1 ( 2 k ) j + 2 = η ( j + 2 ) 2 j + 2 .
By substitution, we obtain
k = 1 ( 1 ) k 1 k I ( m , k ) = m ! 2 m + 1 η ( m + 2 ) sin m + 2 2 π + j = 0 m m j 2 j + 1 π 4 m j β ( j + 2 ) cos j π 2 + j = 0 m m j 2 2 j + 3 π 4 m j η ( j + 2 ) sin j π 2 .
Keeping in mind of (12), we have established the following explicit formula.
Proposition 1
( m N 0 ).
Θ ( m ) = m ! 2 m + 1 η ( m + 2 ) sin m + 2 2 π + j = 0 m 2 ( 1 ) j m 2 j 2 2 j + 1 π 4 m 2 j β ( 2 j + 2 ) ln 2 m + 1 π 4 m + 1 j = 1 m + 1 2 ( 1 ) j m 2 j 1 2 4 j + 1 π 4 m + 1 2 j η ( 2 j + 1 ) .
From this proposition, we claim that Θ ( m ) can always be expressed in terms of ln 2 and values of ζ -function and β -function (particularly G = β ( 2 ) ). The initial values for small m are recorded below, where we can locate Θ ( 0 ) in Moll ([20], §8.4), and both Θ ( 0 ) and Θ ( 1 ) in Vǎlean ([2], Equations 3.87 and 3.113).
Θ ( 0 ) = G 2 π ln 2 4 , Θ ( 1 ) = π G 8 21 ζ ( 3 ) 128 π 2 ln 2 32 , Θ ( 2 ) = π 2 G 32 β ( 4 ) 4 + 3 π ζ ( 3 ) 256 π 3 ln 2 192 , Θ ( 3 ) = π 3 G 128 3 π β ( 4 ) 16 + 1395 ζ ( 5 ) 4096 + 9 π 2 ζ ( 3 ) 2048 π 4 ln 2 1024 , Θ ( 4 ) = π 4 G 512 3 π 2 β ( 4 ) 32 + 3 β ( 6 ) 4 45 π ζ ( 5 ) 4096 + 3 π 3 ζ ( 3 ) 2048 π 5 ln 2 5120 , Θ ( 5 ) = π 5 G 2048 5 π 3 β ( 4 ) 128 + 15 π β ( 6 ) 16 120015 ζ ( 7 ) 65536 225 π 2 ζ ( 5 ) 32768 + 15 π 4 ζ ( 3 ) 32768 π 6 ln 2 24576 .

3.4. T ( m , 1 )

Applying integration by parts, we have
T ( m , 1 ) = 0 1 arctan m x x d x = m 0 1 ln x 1 + x 2 arctan m 1 x d x .
Then, making the change of variable x tan y , we can express
T ( m , 1 ) = m 0 π 4 y m 1 ln ( tan y ) d y = m Λ ( m 1 ) .
Henceforth, Λ ( m ) is defined by the parametric integral
Λ ( m ) = 0 π 4 y m ln ( tan y ) d y for m N 0 .

3.5. Λ ( m )

Recalling another known Fourier series (cf. ([9], §1.442))
ln ( tan y ) = 2 k = 1 cos ( 4 k 2 ) y 2 k 1 , where 0 < y < π 2 ,
we can reformulate the integral Λ ( m ) as
Λ ( m ) = 2 k = 1 1 2 k 1 0 π 4 y m cos ( 4 k 2 ) y d y .
Denote the last integral by J ( m , k ) . Applying integration by parts for m times, we can evaluate this as follows:
J ( m , k ) = 0 π 4 y m cos ( 4 k 2 ) y d y = j = 0 m m j ( 4 k 2 ) j + 1 y m j sin ( 4 k 2 ) y + j π 2 | 0 π 4 = j = 0 m m j ( 4 k 2 ) j + 1 π 4 m j sin 2 k + j 1 2 π m ! ( 4 k 2 ) m + 1 sin m π 2 .
By making use of the trigonometric identity (11), we can proceed
Λ ( m ) = 2 k = 1 J ( m , k ) 2 k 1 = m ! 2 m sin m π 2 k = 1 1 ( 2 k 1 ) m + 2 j = 0 m m j 2 j π 4 m j cos j π 2 k = 1 sin 2 k 1 2 π ( 2 k 1 ) j + 2 j = 0 m m j 2 j π 4 m j sin j π 2 k = 1 cos 2 k 1 2 π ( 2 k 1 ) j + 2 .
Keeping in mind of (12), we can reduce the two trigonometric sums
k = 1 sin 2 k 1 2 π ( 2 k 1 ) j + 2 = k = 1 ( 1 ) k 1 ( 2 k 1 ) j + 2 = β ( j + 2 ) , k = 1 cos 2 k 1 2 π ( 2 k 1 ) j + 2 = 0 .
Therefore, we have proved the following simplified expression
Λ ( m ) = m ! 2 m sin m π 2 λ ( m + 2 ) j = 0 m m j 2 j π 4 m j β ( j + 2 ) cos j π 2 ,
which is equivalent to the formula below.
Proposition 2
( m N 0 ).
Λ ( m ) = m ! 2 m λ ( m + 2 ) sin m π 2 j = 0 m 2 ( 1 ) j m 2 j 2 2 j π 4 m 2 j β ( 2 j + 2 ) .
It should be pointed out that a similar formula for 0 π 2 x m ln ( tan x ) d x was found by Elaissaoui and Guennoun [21] by integrating the product of ln ( tan x ) and the Euler polynomials.
In accordance with this proposition, we affirm that Λ ( m ) can always be expressed by ζ -function and β -function values. The first few values for small m are displayed as follows, where the value for Λ ( 1 ) can be found in ([2], Page 130).
Λ ( 0 ) = G , Λ ( 1 ) = 7 ζ ( 3 ) 16 π G 4 , Λ ( 2 ) = β ( 4 ) 2 π 2 G 16 , Λ ( 3 ) = 3 π β ( 4 ) 8 π 3 G 64 93 ζ ( 5 ) 128 , Λ ( 4 ) = 3 π 2 β ( 4 ) 16 3 β ( 6 ) 2 π 4 G 256 , Λ ( 5 ) = 5 π 3 β ( 4 ) 64 15 π β ( 6 ) 8 π 5 G 1024 + 1905 ζ ( 7 ) 512 .
Summing up, we have proved the following general theorem.
Theorem 2
( m N 0 and n Z with m n ). Let Θ ( m ) and Λ ( m ) be as in Propositions 1 and 2, respectively. The integral values for T ( m , n ) are determined as follows:
n = 0 T ( m , 0 ) = 1 , m = 0 ; π 4 ln 2 2 , m = 1 ; π 4 m m ln 2 2 π 4 m 1 m ( m 1 ) Θ ( m 2 ) , m 2 ; n = 1 T ( m , 1 ) = m Λ ( m 1 ) , m 1 ; n 2 T ( m , n ) = m n 1 T ( m 1 , n 1 ) n 3 n 1 T ( m , n 2 ) 2 n 1 π 4 m , m n ; n < 0 T ( m , n ) = { 1 1 n , m = 0 ; m n 1 T ( m 1 , n + 1 ) n + 1 n 1 T ( m , n + 2 ) 2 n 1 π 4 m , m 1 .
Some particular results of this theorem are commented as follows:
  • Both T ( 2 , 1 ) and T ( 2 , 2 ) can be found in Boyadzhiev ([8], Equations (A.289) and (5.54)).
  • Kobayashi [6] evaluated T ( 1 , 1 ) , T ( 2 , 1 ) and further
    0 1 arctan x arccot x x = π 2 T ( 1 , 1 ) T ( 2 , 1 ) = 7 8 ζ ( 3 ) .
  • Sofo and Nimbran [3] examined cases for T ( 2 , n ) and T ( 3 , n ) .
  • When m = 1 and n < 0 , the recurrence relation in Theorem 2 reads as
    T ( 1 , n ) = 1 + n 1 n T ( 1 , n + 2 ) + π 2 2 n + 1 n ( 1 n ) .
    Repeating this relation yields the next equation
    T ( 1 , n ) = n + 3 n 1 T ( 1 , n + 4 ) 2 n ( n 1 ) ( n + 2 ) ,
    which is equivalent to a known recursion due to Chen [22].
    By iterating (17) further, we can recover the following explicit formula, recorded by Gradshteyn and Ryzhik ([9], §4.532: Equation (1)):
    T ( 1 , n ) = π + ψ ( 2 n 4 ) ψ ( 4 n 4 ) 4 4 n .
    where γ and ψ stand, respectively, for the Euler–Mascheroni constant and the digamma function (cf. Rainville ([23], §9))
    ψ ( x ) = γ + n = 0 1 n + 1 1 n + x .
In view of this theorem, T ( m , n ) can always be expressed in terms of π , ln 2 and values of ζ -function and β -function. The values for T ( m , n ) with 1 m 3 and 3 n 3 are given in Table 2.

4. Evaluation of S ( m , n )

When n 1 , we have by integration by parts
S ( m , n ) = arcsin m x ( 1 n ) x n 1 | 0 1 + m n 1 0 1 arcsin m 1 x x n 1 1 x 2 d x ,
which can alternatively be expressed as
S ( m , n ) : = 0 1 arcsin m 1 x x n 1 1 x 2 d x = n 1 m S ( m , n ) + 1 m π 2 m .
By splitting S ( m , n ) into two integrals, we have
S ( m , n ) = 0 1 arcsin m 1 x x n 3 1 x 2 d x + 0 1 1 x 2 arcsin m 1 x x n 1 d x = S ( m , n 2 ) + 0 1 1 x 2 arcsin m 1 x x n 1 d x .
When n 2 , the last integral can be manipulated, again using integration by parts, as
0 1 1 x 2 arcsin m 1 x x n 1 d x = m 1 n 2 0 1 arcsin m 2 x x n 2 d x 1 n 2 0 1 arcsin m 1 x x n 3 1 x 2 d x = m 1 n 2 S ( m 2 , n 2 ) 1 n 2 S ( m , n 2 ) ,
which leads us to the following expression
S ( m , n ) = n 3 n 2 S ( m , n 2 ) + m 1 n 2 S ( m 2 , n 2 ) .
Substituting (17) into the above equation, we can simplify the resulting equation into the following three-term recurrence relation
( n 1 ) ( n 2 ) S ( m , n ) ( n 3 ) 2 S ( m , n 2 ) m ( m 1 ) S ( m 2 , n 2 ) + π 2 m = 0 .
By making use of this recurrence relation, we can produce all the values of S ( m , n ) for m N 0 and n Z subject to m n as long as the boundary values { S ( 0 , n ) , S ( 1 , n ) , S ( m , 0 ) , S ( m , 1 ) , S ( m , 2 ) } are explicitly determined.

4.1. S ( 0 , n )

For n 0 , it is routine to compute
S ( 0 , n ) = 0 1 x n d x = 1 1 n .

4.2. S ( 1 , n )

For n < 0 , applying integration by parts yields
S ( 1 , n ) = 0 1 x n arcsin x d x = π 2 ( 1 n ) + 1 n 1 0 1 x 1 n 1 x 2 d x .
The last integral can further be manipulated as follows:
0 1 x 1 n 1 x 2 d x = 0 1 x 1 n 1 x 2 d x 0 1 x 1 n 1 x 2 d x = 0 1 x 1 n 1 x 2 d x + 1 n 0 1 x 1 n 1 x 2 d x = n n 1 0 1 x 1 n 1 x 2 d x .
Comparing the above relation with (20), we obtain the recurrence relation belew
S ( 1 , n ) = n ( n + 1 ) ( n 1 ) 2 S ( 1 , n + 2 ) + π 2 ( n 1 ) 2 .
By iterating this relation -times, we obtain the expression
S ( 1 , n ) = S ( 1 , n + 2 ) ( n 2 ) ( n + 1 2 ) ( n 1 2 ) 2 + π 8 k = 0 1 ( n 2 ) k ( n + 1 2 ) k ( n 1 2 ) k + 1 2 .
Letting = n 2 and = 1 n 2 , respectively, for even n and odd n, we can make further simplifications
n 2 0 S ( 1 , n ) = S ( 1 , 0 ) ( n 2 ) n 2 ( n + 1 2 ) n 2 ( n 1 2 ) n 2 2 + π 8 k = 0 n + 2 2 ( n 2 ) k ( n + 1 2 ) k ( n 1 2 ) k + 1 2 = ( 1 ) n 2 ( 2 π ) ( 3 2 ) n 2 ( 2 n 2 ) + π 8 k = 0 n + 2 2 ( n 2 ) k ( n + 1 2 ) k ( n 1 2 ) k + 1 2 , n 2 1 S ( 1 , n ) = S ( 1 , 1 ) ( n 2 ) 1 n 2 ( n + 1 2 ) 1 n 2 ( n 1 2 ) 1 n 2 2 + π 8 k = 0 n + 1 2 ( n 2 ) k ( n + 1 2 ) k ( n 1 2 ) k + 1 2 = π 8 k = 0 n + 1 2 ( n 2 ) k ( n + 1 2 ) k ( n 1 2 ) k + 1 2 ;
where we have employed the initial values (see Entry (e) in Section 1)
S ( 1 , 1 ) = π 2 ln 2 and S ( 1 , 0 ) = 0 π 2 y cos y d y = π 2 1 .
Writing further
( n 2 ) k ( n + 1 2 ) k ( n 1 2 ) k + 1 2 = 4 ( n 1 ) 2 ( n 2 ) k ( n + 1 2 ) k = 4 n 1 ( n 2 ) k + 1 ( n 1 2 ) k + 1 ( n 2 ) k ( n 1 2 ) k ,
we can evaluate the partial sum by telescoping
π 8 k = 0 ( n 2 ) k ( n + 1 2 ) k ( n 1 2 ) k + 1 2 = π 2 n 2 k = 0 ( n 2 ) k + 1 ( n 1 2 ) k + 1 ( n 2 ) k ( n 1 2 ) k = π 2 2 n π 2 2 n ( n 2 ) + 1 ( n 1 2 ) + 1 .
From this, we derive the following closed formula, which is equivalent to those recorded in ([9], §4.523: Equations (1) and (2)):
S ( 1 , n ) = π 2 2 n π 2 2 n ( n 2 ) n 2 ( n 1 2 ) n 2 ( 1 ) n 2 ( 2 π ) ( 3 2 ) n 2 ( 2 2 n ) χ ( n 2 0 ) .

4.3. S ( m , 0 )

For m 2 , making the change in variable x sin y and then applying integration by parts, we can proceed with
S ( m , 0 ) = 0 1 arcsin m x d x = 0 π 2 y m cos y d y = π 2 m m 0 π 2 y m 1 sin y d y = π 2 m m ( m 1 ) 0 π 2 y m 2 cos y d y ,
which can be restated as the following recurrence relation
S ( m , 0 ) = π 2 m m ( m 1 ) S ( m 2 , 0 ) .
Considering that
S ( 0 , 0 ) = 0 1 d x = 1 and S ( 1 , 0 ) = π 2 1 ,
we can iterate the above recurrence -times
S ( m , 0 ) = ( 1 ) m 2 S ( m 2 , 0 ) + k = 0 1 ( 1 ) k m 2 k π 2 m 2 k .
For = m 2 , the above expression becomes
S ( m , 0 ) = ( 1 ) m 2 m ! S m 2 m 2 , 0 + k = 0 m 2 1 ( 1 ) k m 2 k π 2 m 2 k = k = 0 m 2 1 ( 1 ) k m 2 k π 2 m 2 k + ( 1 ) m 2 m ! × 1 , m 2 0 ; π 2 1 , m 2 1 .

4.4. S ( m , 1 )

For m 1 , making the change in variable x sin y and then applying integration by parts, we can reformulate the integral
S ( m , 1 ) = 0 1 arcsin m x x d x = 0 π 2 y m cot y d y = m 0 π 2 y m 1 ln ( sin y ) d y y 2 θ = 2 m m 0 π 4 θ m 1 ln ( 2 tan θ cos 2 θ ) d θ = 2 m ln 2 π 4 m m 2 m 0 π 4 θ m 1 ln ( tan θ ) d θ m 2 m + 1 0 π 4 θ m 1 ln ( cos θ ) d θ .
Recalling Propositions 1 and 2, we find the following explicit formula
S ( m , 1 ) = 2 m ln 2 π 4 m m 2 m Λ ( m 1 ) m 2 m + 1 Θ ( m 1 ) .

4.5. S ( m , 2 )

For m 2 , making the change in variable x sin y and then applying integration by parts twice, we can manipulate the integral
S ( m , 2 ) = 0 1 arcsin m x x 2 d x = 0 π 2 y m cos y sin 2 y d y = π 2 m + m 0 π 2 y m 1 csc y d y = π 2 m m ( m 1 ) 0 π 2 y m 2 ln tan y 2 d y y 2 x = π 2 m m ( m 1 ) 2 m 1 0 π 4 x m 2 ln ( tan x ) d x .
This leads us to the following formula
S ( m , 2 ) = π 2 m m ( m 1 ) 2 m 1 Λ ( m 2 ) ,
where Λ ( m ) is already evaluated by Proposition 2.
To summarise, we have shown the following general theorem.
Theorem 3
( m N 0 and n Z with m n ). Let Θ ( m ) and Λ ( m ) be as in Propositions 1 and 2, respectively. The integral values for S ( m , n ) are determined as follows:
n = 0 S ( m , 0 ) = k = 0 m 2 1 ( 1 ) k m 2 k π 2 m 2 k + ( 1 ) m 2 m ! × { 1 , m 2 0 ; π 2 1 , m 2 1 ; n = 1 S ( m , 1 ) = 2 m ln 2 π 4 m m 2 m Λ ( m 1 ) m 2 m + 1 Θ ( m 1 ) , m 1 ; n = 2 S ( m , 2 ) = π 2 m m ( m 1 ) 2 m 1 Λ ( m 2 ) , m 2 ; n 3 S ( m , n ) = ( n 3 ) 2 S ( m , n 2 ) ( n 1 ) ( n 2 ) + m ( m 1 ) S ( m 2 , n 2 ) ( n 1 ) ( n 2 ) ( π / 2 ) m ( n 1 ) ( n 2 ) , m n ; n < 0 S ( m , n ) = 1 1 n , m = 0 ; π 2 2 n 1 ( n 2 ) n 2 ( n 1 2 ) n 2 ( 1 ) n 2 ( 2 π ) ( 3 2 ) n 2 ( 2 2 n ) χ ( n 2 0 ) , m = 1 ; n ( n + 1 ) ( n 1 ) 2 S ( m , n + 2 ) m ( m 1 ) ( n 1 ) 2 S ( m 2 , n ) + ( π / 2 ) m ( n 1 ) 2 , m 2 .
According to this theorem, S ( m , n ) can always be expressed byin terms of π and ln 2 , as well as values of ζ -function and β -function. The initial values for S ( m , n ) with 1 m 4 and 4 n 4 are recorded in Table 3.

5. Concluding Comments

By making the change of variable x y 1 , it is trivial to check
H ( m , n ) = 1 y n 2 arccoth m y d y and T ( m , n ) = 1 y n 2 arccot m y d y .
Therefore the two integrals on the right hand sides can easily be evaluated.
In this section, we shall further examine some integral variants that can be evaluated as consequences by the preceding Theorems and Propositions, as shown in this paper.

5.1. Evaluation of the Integral Φ ( m )

Firstly, for m N 0 , we examine
Φ ( m ) : = 0 π 4 y m ln ( sin y ) d y .
A similar integral of “ y m ln sin π y ” over [ 0 , 1 ] was treated by Espinosa and Moll [24]. This can easily be expressed as
Φ ( m ) = 0 π 4 y m ln ( sin y ) d y = 0 π 4 y m ln ( tan y cos y ) d y = Θ ( m ) + Λ ( m ) .
Hence, we can compute Φ ( m ) by employing Propositions 1 and 2. The initial values for 0 m 5 are exemplified as follows (where Φ ( 0 ) can also be found in ([25], Equation 9.7.9)).
Φ ( 0 ) = G 2 π ln 2 4 , Φ ( 1 ) = 35 ζ ( 3 ) 128 π G 8 π 2 ln 2 32 , Φ ( 2 ) = β ( 4 ) 4 π 2 G 32 + 3 π ζ ( 3 ) 256 π 3 ln 2 192 , Φ ( 3 ) = 3 π β ( 4 ) 16 π 3 G 128 1581 ζ ( 5 ) 4096 + 9 π 2 ζ ( 3 ) 2048 π 4 ln 2 1024 , Φ ( 4 ) = 3 π 2 β ( 4 ) 32 3 β ( 6 ) 4 π 4 G 512 45 π ζ ( 5 ) 4096 + 3 π 3 ζ ( 3 ) 2048 π 5 ln 2 5120 , Φ ( 5 ) = 5 π 3 β ( 4 ) 128 15 π β ( 6 ) 16 π 5 G 2048 + 123825 ζ ( 7 ) 65536 225 π 2 ζ ( 5 ) 32768 + 15 π 4 ζ ( 3 ) 32768 π 6 ln 2 24576 .

5.2. Evaluation of the Integral Ψ ( m )

For m N 0 , consider the integral
Ψ ( m ) : = 0 y m ln ( tanh y ) d y .
Making the change of variable y arctanh x and then applying integration by parts, we obtain the transformation formula
Ψ ( m ) = 0 y m ln ( tanh y ) d y = 0 1 ln x arctanh m x 1 x 2 d x = 1 m + 1 0 1 arctanh m + 1 x x d x = 1 m + 1 H ( m + 1 , 1 ) .
In view of Theorem 1, the initial values for 0 m 5 are exemplified as follows.
Ψ ( 0 ) = π 2 8 , Ψ ( 3 ) = 93 ζ ( 5 ) 128 , Ψ ( 1 ) = 7 ζ ( 3 ) 16 , Ψ ( 4 ) = π 6 640 , Ψ ( 2 ) = π 4 192 , Ψ ( 5 ) = 1905 ζ ( 7 ) 512 .

5.3. Evaluation of the Integral H ( m , n )

For m N 0 and n Z subject m n , consider the integral
H ( m , n ) : = 0 y m coth n y cosh 2 y d y .
By making the change of variable y arctanh x , we can derive the following transformation formula
H ( m , n ) = 0 y m coth n y cosh 2 y d y = 0 1 arctanh m x x n d x = H ( m , n ) .
Therefore, H ( m , n ) can be computed by means of Theorem 1 with the initial values being given by the same Table 1.
Moll ([20], §6.5 and §6.6) evaluated two similar integrals
0 x sinh x cosh 2 x d x = π 2 , 0 x 2 sinh x cosh 2 x d x = 4 G .
More hyperbolic integral identities can be found in ([9], §3.527).

5.4. Evaluation of the Integral T ( m , n )

Assuming m N 0 and n Z with m n , we examine the next integral
T ( m , n ) : = 0 π 4 y m cot n y d y .
When n < 0 , Moll ([25], §11.5) recorded an explicit formula for T ( 0 , n ) . Jameson and Lord [4] evaluated
T ( 1 , 1 ) = G 2 π 8 ln 2 and 0 π 2 x cot x d x = π 2 ln 2 .
By making the change of variable y arctan x , we can reformulate
T ( m , n ) = 0 π 4 y m cot n y d y = 0 1 arctan m x x n ( 1 + x 2 ) d x .
In view of (6), we find the expression
T ( m , n ) = n m + 1 T ( m + 1 , n + 1 ) + 1 m + 1 π 4 m + 1 .
According to Theorem 2, all the T ( m , n ) can be computed as consequences. For 0 m 3 and 3 n 3 , the corresponding T ( m , n ) values are given in Table 4.

5.5. Evaluation of the Integral S ( m , n )

Let m N 0 and n Z satisfying the condition m n . Define the integral S ( m , n ) by
S ( m , n ) : = 0 π 2 y m csc n y d y .
A similar integral 0 π 2 y m cos n y d y was extensively examined by Moll ([25], §5.2).
By making the change of variable y arcsin x , we can manipulate
S ( m , n ) = 0 π 2 y m csc n y d y = 0 1 arcsin m x x n 1 x 2 d x .
Recalling (17), we have the expression
S ( m , n ) = S ( m + 1 , n + 1 ) = n m + 1 S ( m + 1 , n + 1 ) + 1 m + 1 π 2 m + 1 .
Applying Theorem 3, we can evaluate S ( m , n ) as consequences. For 0 m 4 and 3 n 3 , the corresponding S ( m , n ) values are recorded in Table 5.

Author Contributions

Original draft & supervision, W.C.; Writing & editing, C.L. 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.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Boros, G.; Moll, V.H. Irresistible Integrals; Cambridge University Press: Cambridge, UK, 2004. [Google Scholar]
  2. Vǎlean, C.I. (Almost) Impossible Integrals, Sums, and Series; Springer Nature AG: Basel, Switzerland, 2019. [Google Scholar]
  3. Sofo, A.; Nimbran, A.S. Euler sums and integral connections. Mathematics 2019, 7, 833. [Google Scholar] [CrossRef]
  4. Jameson, G.; Lord, N. Integrals evaluated in terms of Catalan’s constant. Math. Gaz. 2017, 101, 38–49. [Google Scholar] [CrossRef]
  5. Johansson, F. Contributors. Catalan’s Constant. Fungrim Home Page. Available online: https://fungrim.org/topic/Catalan’s_constant/ (accessed on 7 July 2022).
  6. Kobayashi, M. Integral representations for local dilogarithm and trilogarithm functions. arXiv 2021, arXiv:2108.12543v1. [Google Scholar] [CrossRef]
  7. Marichev, O.; Sondow, J.; Weisstein, E. Catalan’s Constant. From MathWorld–A Wolfram Web Resource. Available online: https://mathworld.wolfram.com/CatalansConstant.html (accessed on 7 July 2022).
  8. Boyadzhiev, K.N. Special Techniques for Solving Integrals: Examples and Problems; World Scientific: Hackensack, NJ, USA, 2022. [Google Scholar]
  9. Gradshteyn, I.S.; Ryzhik, I.M. Table of Integrals, Series, and Products, 7th ed.; Academic Press: New York, NY, USA, 2007. [Google Scholar]
  10. Adamchik, V. Thirty–Three Representations of Catalan’s Constant. Available online: https://library.wolfram.com/infocenter/Demos/109/ (accessed on 7 July 2022).
  11. Bradley, D.M. Representations of Catalan’s Constant. Available online: www.researchgate.net/publication/2325473 (accessed on 2 February 2001).
  12. Lord, N. Variations on a theme – Euler and the logsine integral. Math. Gaz. 2012, 96, 451–458. [Google Scholar] [CrossRef]
  13. Koyama, S.Y.; Kurokawa, N. Euler’s integrals and multiple sine functions. Proc. Am. Math. Soc. 2004, 133, 1257–1265. [Google Scholar] [CrossRef]
  14. Bowman, F. Note on the integral 0 π 2 ( logsin θ ) n d θ . J. London Math. Soc. 1947, S1-22, 172–173. [Google Scholar] [CrossRef]
  15. Choi, J.; Srivastava, H.M. Explicit evaluations of some families of log-sine and log-cosine integrals. Integral Transform. Spec. Funct. 2011, 22, 767–783. [Google Scholar] [CrossRef]
  16. Coffey, M.W. On some log-cosine integrals related to ζ(2), ζ(3), ζ(4), and ζ(6). J. Comput. Appl. Math. 2003, 159, 205–215. [Google Scholar] [CrossRef]
  17. Freitas, P. Integrals of polylogarithmic functions, recurrence relations, and associated Euler sums. Math. Comp. 2005, 74, 1425–1440. [Google Scholar] [CrossRef]
  18. Glasser, M.L. Some integrals of the arctangent function. Math. Comp. 1968, 22, 445–447. [Google Scholar] [CrossRef]
  19. Rudin, W. Principles of Mathematical Analysis, 3rd ed.; McGraw–Hill, Inc.: New York, NY, USA, 1976. [Google Scholar]
  20. Moll, V.H. Special Integrals of Gradshteyn and Ryzhik, the Proofs–Volume II; CRC Press: Boca Raton, FL, USA; Taylor & Francis Group: London, UK; New York, NY, USA, 2016. [Google Scholar]
  21. Elaissaoui, L.; Guennoun, Z.A. Evaluation of log-tangent integrals by series involving ζ(2n+1). Integral Transforms. Spec. Funct. 2017, 28, 460–475. [Google Scholar]
  22. Chen, H. Transcendental moments of the arctangent: Solution to Problem 2020. Math. Mag. 2018, 91, 157. [Google Scholar]
  23. Rainville, E.D. Special Functions; The Macmillan Company: New York, NY, USA, 1960. [Google Scholar]
  24. Espinosa, O.; Moll, V.H. On some integrals involving the Hurwitz zeta function: Part 1. Ramanujan J. 2002, 6, 159–188. [Google Scholar] [CrossRef]
  25. Moll, V.H. Special Integrals of Gradshteyn and Ryzhik, the Proofs—Volume I; CRC Press: Boca Raton, FL, USA; Taylor & Francis Group: London, UK; New York, NY, USA, 2015. [Google Scholar]
Table 1. Values for H ( m , n ) .
Table 1. Values for H ( m , n ) .
n m 12345
5 23 90 19 180 + 23 ln 2 45 1 3 + 23 π 2 360 1 30 + 23 ζ ( 3 ) 20 + 4 ln 2 3 1 6 + 5 π 2 18 + 161 π 4 4320
4 3 20 + ln 2 5 1 3 + π 2 60 1 20 + 9 ζ ( 3 ) 40 + ln 2 1 5 + π 2 6 + 7 π 4 1200 15 ζ ( 3 ) 4 + 45 ζ ( 5 ) 32 + ln 2
3 1 3 1 12 + 2 ln 2 3 1 4 + π 2 12 3 ζ ( 3 ) 2 + ln 2 5 π 2 24 + 7 π 4 144
2 1 6 + ln 2 3 1 3 + π 2 36 3 ζ ( 3 ) 8 + ln 2 π 2 6 + 7 π 4 720 15 ζ ( 3 ) 4 + 75 ζ ( 5 ) 32
1 1 2 ln 2 π 2 8 9 ζ ( 3 ) 4 7 π 4 96
0 ln 2 π 2 12 9 ζ ( 3 ) 8 7 π 4 240 225 ζ ( 5 ) 32
1 π 2 8 7 ζ ( 3 ) 8 π 4 64 93 ζ ( 5 ) 32 π 6 128
2 π 2 6 3 ζ ( 3 ) 2 π 4 30 15 ζ ( 5 ) 2
3 π 2 4 3 ζ ( 3 ) π 4 12
4 π 2 3 + π 4 90 5 ζ ( 3 ) + 5 ζ ( 5 ) 2
5 5 π 2 12 + π 4 18
(★ indicates that the corresponding integral diverges.)
Table 2. Values for T ( m , n ) .
Table 2. Values for T ( m , n ) .
n m 123
3 1 6 1 12 + π 12 ln 2 3 π 8 1 4 + π ln 2 4 + π 2 32 G
2 π 12 1 6 + ln 2 6 1 3 π 6 π ln 2 12 + π 2 48 + G 3 π 4 ln 2 2 π 2 16 π 2 ln 2 32 + π 3 192 + π G 4 21 ζ ( 3 ) 64
1 π 4 1 2 π 2 16 π 4 + ln 2 2 π 3 64 3 π 2 32 3 π ln 2 8 + 3 G 2
0 π 4 ln 2 2 π 2 16 + π ln 2 4 G π 3 64 + 3 π 2 ln 2 32 3 π G 4 + 63 ζ ( 3 ) 64
1G π G 2 7 ζ ( 3 ) 8 3 π 2 G 16 3 β ( 4 ) 2
2 π ln 2 4 π 2 16 + G 3 π 2 ln 2 32 π 3 64 + 3 π G 4 105 ζ ( 3 ) 64
3 3 π ln 2 8 π 3 64 3 π 2 32 + 3 G 2
(★ indicates that the corresponding integral diverges.)
Table 3. Values for S ( m , n ) .
Table 3. Values for S ( m , n ) .
n m 1234
4 π 10 8 75 π 2 20 298 1125 4144 5625 149 π 375 + π 3 40 254728 84375 149 π 2 375 + π 4 80
3 5 π 64 5 π 2 128 1 8 5 π 3 256 51 π 512 3 8 51 π 2 512 + 5 π 4 512
2 π 6 2 9 π 2 12 14 27 40 27 7 π 9 + π 3 24 488 81 7 π 2 9 + π 4 48
1 π 8 π 2 16 1 4 π 3 32 3 π 16 3 4 3 π 2 16 + π 4 64
0 π 2 1 π 2 4 2 6 3 π + π 3 8 24 3 π 2 + π 4 16
1 π ln 2 2 π 2 ln 2 4 7 ζ ( 3 ) 8 π 3 ln 2 8 9 π ζ ( 3 ) 16 π 4 ln 2 16 9 π 2 ζ ( 3 ) 16 + 93 ζ ( 5 ) 32
2 4 G π 2 4 6 π G 21 ζ ( 3 ) 2 π 3 8 6 π 2 G 48 β ( 4 ) π 4 16
3 3 π ln 2 2 π 3 16 3 π 2 ln 2 2 π 4 32 21 ζ ( 3 ) 4
4 8 G + π 2 G 8 β ( 4 ) π 4 48 π 2 2
(★ indicates that the corresponding integral diverges.)
Table 4. Values for T ( m , n ) .
Table 4. Values for T ( m , n ) .
n m 0123
3 1 2 ln 2 2 π 4 1 2 G 2 + π ln 2 8 π 2 16 π 4 π G 4 + ln 2 2 + π 2 ln 2 32 + 21 ζ ( 3 ) 64 π 3 64 3 π 2 32 3 π ln 2 8 + π 3 ln 2 128 + 3 G 2 3 π 2 G 32 9 π ζ ( 3 ) 256 + 3 β ( 4 ) 4
2 1 π 4 π 4 π 2 32 ln 2 2 π 2 16 + π ln 2 4 π 3 192 G π 3 64 + 3 π 2 ln 2 32 π 4 1024 3 π G 4 + 63 ζ ( 3 ) 64
1 ln 2 2 G 2 π ln 2 8 π G 4 π 2 ln 2 32 21 ζ ( 3 ) 64 3 π 2 G 32 3 β ( 4 ) 4 + 9 π ζ ( 3 ) 256 π 3 ln 2 128
0 π 4 π 2 32 π 3 192 π 4 1024
1 G 2 + π ln 2 8 π G 4 + π 2 ln 2 32 35 ζ ( 3 ) 64 3 π 2 G 32 3 β ( 4 ) 4 9 π ζ ( 3 ) 256 + π 3 ln 2 128
2 G + π ln 2 4 π 2 16 π 3 192 3 π G 4 105 ζ ( 3 ) 64 π 4 1024 π 3 64 + 3 π 2 ln 2 32
3 3 G 2 3 π 2 G 32 3 π 2 32 π 3 64 + 3 π ln 2 8 π 3 ln 2 128 + 9 π ζ ( 3 ) 256 + 3 β ( 4 ) 4
(★ indicates that the corresponding integral diverges.)
Table 5. Values for S ( m , n ) .
Table 5. Values for S ( m , n ) .
n m 01234
3 2 3 7 9 7 π 9 40 27 7 π 2 12 122 27 1456 81 244 π 27 + 7 π 3 18
2 π 4 1 4 + π 2 16 π 8 + π 3 48 3 π 2 32 + π 4 128 3 8 π 3 16 3 π 8 + π 5 320
1 11 π 2 3 π 2 4 6 24 12 π + π 3 2
0 π 2 π 2 8 π 3 24 π 4 64 π 5 160
1 2 G 2 π G 7 ζ ( 3 ) 2 3 π 2 G 2 12 β ( 4 ) π 3 G 24 π β ( 4 ) + 93 ζ ( 5 ) 2
2 π ln 2 3 π 2 ln 2 4 21 ζ ( 3 ) 8 π 3 ln 2 2 9 π ζ ( 3 ) 4
3 3 π 2 G 4 3 π 2 8 + 6 G 6 β ( 4 ) 12 π G 21 ζ ( 3 ) 12 π β ( 4 ) π 3 4 + π 3 G 2 + 93 ζ ( 5 ) 4
(★ indicates that the corresponding integral diverges.)
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, C.; Chu, W. Improper Integrals Involving Powers of Inverse Trigonometric and Hyperbolic Functions. Mathematics 2022, 10, 2980. https://doi.org/10.3390/math10162980

AMA Style

Li C, Chu W. Improper Integrals Involving Powers of Inverse Trigonometric and Hyperbolic Functions. Mathematics. 2022; 10(16):2980. https://doi.org/10.3390/math10162980

Chicago/Turabian Style

Li, Chunli, and Wenchang Chu. 2022. "Improper Integrals Involving Powers of Inverse Trigonometric and Hyperbolic Functions" Mathematics 10, no. 16: 2980. https://doi.org/10.3390/math10162980

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