Next Article in Journal
Neurofunctional Symmetries and Asymmetries during Voluntary out-of- and within-Body Vivid Imagery Concurrent with Orienting Attention and Visuospatial Detection
Next Article in Special Issue
On Some New Trapezoidal Type Inequalities for Twice (p, q) Differentiable Convex Functions in Post-Quantum Calculus
Previous Article in Journal
Analysis of the Dynamical Behaviour of a Two-Dimensional Coupled Ecosystem with Stochastic Parameters
Previous Article in Special Issue
New Ostrowski-Type Fractional Integral Inequalities via Generalized Exponential-Type Convex Functions and Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A More Accurate Half-Discrete Hilbert-Type Inequality Involving One upper Limit Function and One Partial Sum

1
Department of Mathematics, Guangdong University of Education, Guangzhou 510303, China
2
Department of Mathematics, Longyan University, Longyan 364012, China
3
Institute of Applied Mathematics, Longyan University, Longyan 364012, China
*
Author to whom correspondence should be addressed.
Symmetry 2021, 13(8), 1548; https://doi.org/10.3390/sym13081548
Submission received: 27 July 2021 / Revised: 14 August 2021 / Accepted: 18 August 2021 / Published: 23 August 2021
(This article belongs to the Special Issue Symmetry in the Mathematical Inequalities)

Abstract

:
In this paper, by virtue of the symmetry principle, we construct proper weight coefficients and use them to establish a more accurate half-discrete Hilbert-type inequality involving one upper limit function and one partial sum. Then, we prove the new inequality with the help of the Euler–Maclaurin summation formula and Abel’s partial summation formula. Finally, we illustrate how the obtained results can generate some new half-discrete Hilbert-type inequalities.

1. Introduction

The celebrated Hardy–Hilbert’s inequality reads as:
m = 1 n = 1 a m b n m + n < π sin ( π / p ) ( m = 1 a m p ) 1 p ( n = 1 b n q ) 1 q ,
where p > 1 ,   1 p + 1 q = 1 ,   a m , b n 0 ,   0 < m = 1 a m p <   a n d   0 < n = 1 b n q < , the constant factor π sin ( π / p ) is the best possible (see [1], Theorem 315).
A more accurate form of (1) was provided in ([1], Theorem 323), as follows:
m = 1 n = 1 a m b n m + n 1 < π sin ( π / p ) ( m = 1 a m p ) 1 p ( n = 1 b n q ) 1 q .
In 2006, by introducing parameters λ i ( 0 , 2 ] ( i = 1 , 2 ) , λ 1 + λ 2 = λ ( 0 , 4 ] , an extension of (1) was provided by [2] as follows:
m = 1 n = 1 a m b n ( m + n ) λ < B ( λ 1 , λ 2 ) [ m = 1 m p ( 1 λ 1 ) 1 a m p ] 1 p [ n = 1 n q ( 1 λ 2 ) 1 b n q ] 1 q ,
where the constant factor B ( λ 1 , λ 2 ) is the best possible, and the beta function is defined as:
B ( u , v ) = 0 t u 1 ( 1 + t ) u + v d t ( u , v > 0 ) .
Obviously, when λ = 1 , λ 1 = 1 q , λ 2 = 1 p , inequality (2) reduces to (1); when p = q = 2 ,   λ 1 = λ 2 = λ 2 , inequality (2) reduces to the inequality presented by Yang in [3].
Recently, applying inequality (3) and Abel’s summation by parts formula, Adiyasuren et al. [4] gave a new inequality with the kernel 1 ( m + n ) λ involving two partial sums. Inequality (1), with its integral analogues, is playing an important role in analysis and its applications (see [5,6,7,8]).
In 1934, a half-discrete Hilbert-type inequality was given as follows ([1], Theorem 351): assuming that K ( t ) ( t > 0 ) is a decreasing function, 0 < ϕ ( s ) : = 0 K ( t ) t s 1 d t < , a n 0 , such that 0 < n = 1 a n p < , we have:
0 x p 2 ( n = 1 K ( n x ) a n ) p d x < ϕ p ( 1 q ) n = 1 a n p .
In 2016, Hong et al. [9] considered some equivalent statements of the extensions of (1) with the best possible constant factor related to several parameters. Some extensions of inequality (4) were given by [10,11,12,13,14,15]. Recently, Yang et al. [16,17] gave reverse half-discrete Hardy-Hilbert’s inequalities and dealt with their equivalent statements of the best possible constant factor related to several parameters.
In this article, following the method of [2,4,9], in the light of the symmetry principle, we construct proper weight coefficients and use them to establish a more accurate half-discrete Hilbert-type inequality involving one upper limit function and one partial sum. Subsequently, we prove this new inequality by means of the Hermite–Hadamard inequality, Euler–Maclaurin summation formula and Abel’s partial summation formula. As an extension of the obtained results, the equivalent statements of the best possible constant factor related to several parameters are discussed. It is shown that some new half-discrete Hilbert-type inequalities can be derived from the special cases of our main results.

2. Some Lemmas

In what follows, we suppose that p > 1 , 1 p + 1 q = 1 , η [ 0 , 1 4 ] , λ ( 0 , 2 ] , λ 1   ( 0 , λ + 1 ) , λ 2 ( 0 , 1 2 ] ( 0 , λ + 1 ) , λ ^ 1 : = λ λ 2 p + λ 1 q , λ ^ 2 : = λ λ 1 q + λ 2 p . We also assume that f ( x ) ( 0 ) is a Lebesgue integrable function in any interval ( 0 , b ] ( b > 0 ) , and define the upper limit function F ( x ) : = 0 x f ( t ) d t ( x 0 ) with the partial sums as follows:
A n : = k = 1 n a k ( a n 0 , n N : = { 1 , 2 , } ) ,
which satisfies F ( x ) = o ( e t x ) , A n = o ( e t ( n η ) )   ( t > 0 ; x , n ) :
0 < 0 x p λ ^ 1 1 F p ( x ) <   a n d   0 < n = 1 ( n η ) q λ ^ 2 1 A n q < .
Lemma 1. 
(i) Let ( 1 ) i d i d t i g ( t ) > 0 , t [ m , ) ( m N ) with g ( i ) ( ) = 0   ( i = 0 , 1 , 2 , 3 ) , and let P i ( t ) , B i ( i N ) be the Bernoulli functions and the Bernoulli numbers of i -order. Then, we have ([5]):
m P 2 q 1 ( t ) g ( t ) d t = ε q B 2 q 2 q g ( m ) ( 0 < ε q < 1 ; q N ) .
In particular, for q = 1 , in view of B 2 = 1 6 , we have:
1 12 g ( m ) < m P 1 ( t ) g ( t ) d t < 0 ;  
For  q = 2 , in view of B 4 = 1 30 , it follows that:
0 < m P 3 ( t ) g ( t ) d t < 1 120 g ( m ) .  
(ii) If h ( t ) ( > 0 ) C 3 [ m , ) , h ( i ) ( ) = 0 ( i = 0 , 1 , 2 , 3 ) , then we have the following Euler–Maclaurin summation formula:
k = m h ( k ) = m h ( t ) d t + 1 2 h ( m ) + m P 1 ( t ) h ( t ) d t ,  
where:
m P 1 ( t ) h ( t ) d t = 1 12 h ( m ) + 1 6 m P 3 ( t ) h ( t ) d t .  
Lemma 2. 
Let s ( 0 , 4 ] , s 2 ( 0 , 3 2 ] ( 0 , s ) , k s ( s i ) : = B ( s i , s s i ) ( i = 1 , 2 ) , and let ϖ ( s 2 , x ) denote the following weight coefficient:
ϖ ( s 2 , x ) : = x s s 2 n = 1 ( n η ) s 2 1 ( x + n η ) s ( x R + : = ( 0 , ) )  
Then, we have the following inequalities:
0 < k s ( s 2 ) ( 1 O ( 1 x s 2 ) ) < ϖ ( s 2 , x ) < k s ( s 2 ) ,  
where we indicate O ( 1 x s 2 ) : = 1 k s ( s 2 ) 0 1 η x u s 2 1 ( 1 + u ) s d u > 0 .
Proof. 
For fixed x R + , we define a function g ( x , t ) by:
g ( x , t ) : = ( t η ) s 2 1 ( x + t η ) s ( t ( η , ) ) ,
which implies that g ( x , t ) > 0   ( t I η ) and g C ( I η ) , where I η : = ( η , ) . In the following, we consider two cases of s 2 ( 0 , 1 ) ( 0 , s ) and s 2 [ 1 , 3 2 ] ( 0 , s ) to prove inequalities (12).
(i) For s 2 ( 0 , 1 ) ( 0 , s ) , since:
( 1 ) i i t i g ( x , t ) > 0 ( t > η ; i = 0 , 1 , 2 ) ,
by the Hermite–Hadamard inequality, setting u = t η x , we have:
ϖ ( s 2 , x ) = x s s 2 n = 1 g ( x , n ) < x s s 2 1 2 g ( x , t ) d t
= x s s 2 1 2 ( t η ) s 2 1 ( x + t η ) s d t = 1 2 η x u s 2 1 ( 1 + u ) s d u 0 u s 2 1 ( 1 + u ) s d u = B ( s 2 , s s 2 ) = k s ( s 2 ) .
On the other hand, in view of the decreasingness property of series, setting u = t η x , we obtain:
ϖ ( s 2 , x ) = x s s 2 n = 1 g ( x , n ) > x s s 2 1 g ( x , t ) d t
= 1 η x u s 2 1 ( 1 + u ) s d u = B ( s 2 , s s 2 ) 0 1 η x u s 2 1 ( 1 + u ) s d u = k s ( s 2 ) ( 1 O ( 1 x s 2 ) ) > 0 ,
where O ( 1 x s 2 ) = 1 k s ( s 2 ) 0 1 η x u s 2 1 ( 1 + u ) s d u > 0 , which satisfies:
0 < 0 1 η x u s 2 1 ( 1 + u ) s d u < 0 1 η x u s 2 1 d u = 1 s 2 ( 1 η x ) s 2 ( x R + ) .
In this case, we obtain (12).
(ii) For s 2 [ 1 , 3 2 ] ( 0 , s ) , by (9), we have:
n = 1 g ( x , n ) = 1 g ( x , t ) d t + 1 2 g ( x , 1 ) + 1 P 1 ( t ) t g ( x , t ) d t = η g ( x , t ) d t h ( x ) ,
where h ( x ) is defined by:
h ( x ) : = η 1 g ( x , t ) d t 1 2 g ( x , 1 ) 1 P 1 ( t ) t g ( x , t ) d t .
We obtain 1 2 g ( x , 1 ) = ( 1 η ) s 2 1 2 ( x + 1 η ) s , and then integrating by parts, it follows that:
η 1 g ( x , t ) d t = η 1 ( t η ) s 2 1 ( x + t η ) s d t = 1 s 2 η 1 d ( t η ) s 2 ( x + t η ) s = 1 s 2 ( t η ) s 2 ( x + t η ) s | η 1 + s s 2 η 1 ( t η ) s 2 d t ( x + t η ) s + 1
= 1 s 2 ( 1 η ) s 2 ( x + 1 η ) s + s s 2 ( s 2 + 1 ) η 1 1 ( x + 1 η ) s + 1 d ( t η ) s 2 + 1
> 1 s 2 ( 1 η ) s 2 ( x + 1 η ) s + s s 2 ( s 2 + 1 ) [ ( t η ) s 2 + 1 ( x + t η ) s + 1 ] η 1 + s ( s + 1 ) s 2 ( s 2 + 1 ) ( x + 1 η ) s + 2 η 1 ( t η ) s 2 + 1 d t
= 1 s 2 ( 1 η ) s 2 ( x + 1 η ) s + s s 2 ( s 2 + 1 ) ( 1 η ) s 2 + 1 ( x + 1 η ) s + 1 + s ( s + 1 ) ( 1 η ) s 2 + 2 s 2 ( s 2 + 1 ) ( s 2 + 2 ) ( x + 1 η ) s + 2 .
We find:
t g ( x , t ) = ( s 2 1 ) ( t η ) s 2 2 ( x + t η ) s + s ( t η ) s 2 1 ( x + t η ) s + 1 = ( 1 s 2 ) ( t η ) s 2 2 ( x + t η ) s + s ( t η ) s 2 2 ( x + t η ) s s x ( t η ) s 2 2 ( x + t η ) s + 1 = ( s + 1 s 2 ) ( t η ) s 2 2 ( x + t η ) s s x ( t η 2 ) s 2 2 ( x + t η ) s + 1 ,
additionally, for s 2 [ 1 , 3 2 ] ( 0 , s ) , it follows that:
( 1 ) i i t i [ ( t η ) s 2 2 ( x + t η ) s ] > 0 , ( 1 ) i i t i [ ( t η ) s 2 2 ( x + t η ) s + 1 ] > 0 ( t > η ; i = 0 , 1 , 2 , 3 ) .
By (8), (9) and (10), setting a : = 1 η ( [ 3 4 , 1 ] ) , we obtain:
( s + 1 s 2 ) 1 P 1 ( t ) ( t η ) s 2 2 ( x + t η ) s d t > s + 1 s 2 12 ( x + 1 η ) s a s 2 2 , x s 1 P 1 ( t ) ( t η ) s 2 2 ( x + t η ) s + 1 d t > x s 12 ( x + 1 η ) s + 1 a s 2 2 x s 720 [ ( t η ) s 2 2 ( x + t η ) s + 1 ] t = 1 > ( x + 1 η ) s a s 12 ( x + 1 η ) s + 1 a s 2 2 ( x + 1 η ) s 720 [ ( s + 1 ) ( s + 2 ) a s 2 2 ( x + 1 η ) s + 3 + 2 ( s + 1 ) ( 2 s 2 ) a s 2 3 ( x + 1 η ) s + 2 + ( 2 s 2 ) ( 3 s 2 ) a s 2 4 ( x + 1 η ) s + 1 ] . = s a s 2 2 12 ( x + 1 η ) s s a s 2 1 12 ( x + 1 η ) s + 1 s 720 [ ( s + 1 ) ( s + 2 ) a s 2 2 ( x + 1 η ) s + 2 + 2 ( s + 1 ) ( 2 s 2 ) a s 2 3 ( x + 1 η ) s + 1 + ( 2 s 2 ) ( 3 s 2 ) a s 2 4 ( x + 1 η ) s ] ,
and then we have:
h ( x ) > a s 2 4 ( x + 1 η ) s h 1 + s a s 2 3 ( x + 1 η ) s + 1 h 2 + s ( s + 1 ) a s 2 2 ( x + 1 η ) s + 2 h 3 ,
where h i ( i = 1 , 2 , 3 ) are indicated as:
h 1 : = a 4 s 2 a 3 2 ( 1 s 2 ) a 2 12 s ( 2 s 2 ) ( 3 s 2 ) 720 , h 2 : = a 4 s 2 ( s 2 + 1 ) a 2 12 ( s + 1 ) ( 2 s 2 ) 360 ,
h 3 : = a 4 s 2 ( s 2 + 1 ) ( s 2 + 2 ) s + 2 720 .
For s ( 0 , 4 ] , s 2 [ 1 , 3 2 ] ( 0 , s ) , a [ 3 4 , 1 ] , we find:
h 1 > a 2 12 s 2 [ s 2 2 ( 6 a + 1 ) s 2 + 12 a 2 ] 1 90 .
In view of:
d d a [ s 2 2 ( 6 a + 1 ) s 2 + 12 a 2 ] = 6 ( 4 a s 2 ) 6 ( 4 3 4 3 2 ) > 0 , and
d d s 2 [ s 2 2 ( 6 a + 1 ) s 2 + 12 a 2 ] = 2 s 2 ( 6 a + 1 ) 2 3 2 ( 6 3 4 + 1 ) = 3 11 2 < 0 ,
we obtain:
h 1 ( 3 / 4 ) 2 12 ( 3 / 2 ) [ ( 3 2 ) 2 ( 6 3 4 + 1 ) 3 2 + 12 ( 3 4 ) 2 ] 1 90
h 2 > a 2 ( 4 a 2 15 1 12 ) 1 72 ( 3 4 ) 2 [ 4 15 ( 3 4 ) 2 1 12 ] 1 72 = 3 80 1 72 > 0 ,
h 3 8 a 4 105 6 720 8 105 ( 3 4 ) 4 1 120 = 27 1120 1 120 > 0 ,
and then we obtain h ( x ) > 0 .
On the other hand, similar to the above, we have:
n = 1 g ( x , n ) = 1 g ( x , t ) d t + 1 2 g ( x , 1 ) + 1 P 1 ( t ) t g ( x , t ) d t = 1 g ( x , t ) d t + H ( x ) ,
where H ( x ) is indicated as:
H ( x ) : = 1 2 g ( x , 1 ) + 1 P 1 ( t ) t g ( x , t ) d t .
Thus, we obtain that 1 2 g ( x , 1 ) = a s 2 1 2 ( x + 1 η ) s and:
t g ( x , t ) = ( s + 1 s 2 ) ( t η ) s 2 2 ( x + t η ) s + s x ( t η ) s 2 2 ( x + t η ) s + 1 .
For s 2 ( 0 , 3 2 ] ( 0 , s ) , 0 < s 4 , by (7), we obtain:
( s + 1 s 2 ) 1 P 1 ( t ) ( t η ) s 2 2 ( x + t η ) s d t > 0 ,
x s 1 P 1 ( t ) ( t η ) s 2 2 ( x + t η ) s + 1 d t > x s 12 ( x + 1 η ) s + 1 a s 2 2 = ( x + 1 η ) s + a s 12 ( x + 1 η ) s + 1 a s 2 2 = s 12 ( x + 1 η ) s a s 2 2 + s 12 ( x + 1 η ) s + 1 a s 2 1 > s 12 ( x + 1 η ) s a s 2 2 .
Hence, we have:
H ( x ) > a s 2 1 2 ( x + 1 η ) s s a s 2 2 12 ( x + 1 η ) s = ( a 2 s 12 ) a s 2 2 ( x + 1 η ) s ( 1 2 3 4 4 12 ) a s 2 2 ( x + 1 η ) s = ( 3 8 1 3 ) a s 2 2 ( x + 1 η ) s > 0 .
Therefore, we obtain:
1 g ( x , t ) d t < n = 1 g ( x , n ) < η g ( x , t ) d t ( x > 0 )
In view of the results obtained in the case (i), we obtain (12). This completes the proof of lemma 2.
Lemma 3. 
Let s ( 0 , 4 ] , s 1 ( 0 , s ) , s 2 ( 0 , 3 2 ] ( 0 , s ) . Then, we have the following more accurate half-discrete Hardy–Hilbert inequality:
I = 0 n = 1 a n f ( x ) ( x + n η ) s d x ( k s ( s 2 ) ) 1 p ( k s ( s 1 ) ) 1 q  
× { 0 x p [ 1 ( s s 2 p + s 1 q ) ] 1 f p ( x ) d x } 1 p { n = 1 ( n η ) q [ 1 ( s s 1 q + s 2 p ) ] 1 a n q } 1 q .
Proof. 
For s 1 ( 0 , s ) , setting u = x n η , we have the following expression of the weight coefficient:
ω ( s 1 , n ) : = ( n η ) s s 1 0 x s 1 1 ( x + n η ) s d x = 0 u s 1 1 ( u + 1 ) s d u = k s ( s 1 ) ( n N ) .
By using Hölder’s inequality [18], we obtain:
I = 0 n = 1 1 ( x + n η ) s [ x ( 1 s ) 1 / q ( n η ) ( 1 s 2 ) / p f ( x ) ] [ ( n η ) ( 1 s 2 ) / p x ( 1 s ) 1 / q a n ] d x [ 0 n = 1 1 ( x + n η ) s x ( 1 s 1 ) ( p 1 ) ( n η ) 1 s 2 f p ( x ) d x ] 1 p
× [ n = 1 0 1 ( x + n η ) s ( n η ) ( 1 s 2 ) ( q 1 ) x 1 s 1 d x a n q ] 1 q = { 0 ϖ ( s 2 , x ) x p [ 1 ( s s 2 p + s 1 q ) ] 1 f p ( x ) d x } 1 p × { n = 1 ω ( s , 1 n ) ( n η ) q [ 1 ( s s 1 q + s 2 p ) ] 1 a n q } 1 q .
Then, by (12) and (14), we derive inequality (13). The Lemma 3 is proved.
Remark 1. 
In (13), for s = λ + 2 ( 2 , 4 ] , λ ( 0 , 2 ] , s 1 = λ 1 + 1 ( 1 , s ) , λ 1 ( 0 , λ + 1 ) ,
s 2 = λ 2 + 1 ( 1 , 3 2 ] , λ 2 ( 0 , 1 2 ] ( 0 , λ + 1 )  
Replacing f ( x ) ( r e s p . a n ) by F ( x ) ( r e s p . A n ) , in view of Lemma 3 and (5), we have:
0 n = 1 A n ( x + n η ) λ + 2 F ( x ) d x < ( k λ + 2 ( λ 2 + 1 ) ) 1 p ( k λ + 2 ( λ 1 + 1 ) ) 1 q  
× [ 0 x p λ ^ 1 1 F p ( x ) d x ] 1 p [ n = 1 ( n η ) q λ ^ 2 1 A n q ] 1 q .
Lemma 4. 
For t > 0 , we have:
0 e t x f ( x ) d x = t 0 e t x F ( x ) d x ,  
n = 1 e t ( n η ) a n t n = 1 e t ( n η ) A n .  
Proof. 
Integration by parts, in view of F ( 0 ) = 0 , F ( x ) = o ( e t x ) ( t > 0 ; x ) , it follows that:
0 n = 1 A n ( x + n η ) λ + 2 F ( x ) d x < ( k λ + 2 ( λ 2 + 1 ) ) 1 p ( k λ + 2 ( λ 1 + 1 ) ) 1 q
= lim x e t x F ( x ) + t 0 e t x F ( x ) d x = t 0 e t x F ( x ) d x ,
and then (16) follows.
In view of A n e t ( n η ) = o ( 1 ) ( n ) , by Abel’s summation by parts formula, we obtain:
n = 1 e t ( n η ) a n = lim n A n e t ( n η ) + n = 1 A n [ e t ( n η ) e t ( n η + 1 ) ]      = n = 1 A n [ e t ( n η ) e t ( n η + 1 ) ] = ( 1 e t ) n = 1 e t ( n η ) A n .
Since 1 e t < t ( t > 0 ) , we have (17). The Lemma 4 is proved.

3. Main Results

Theorem 1. 
Let p > 1 , 1 p + 1 q = 1 , η [ 0 , 1 4 ] , λ ( 0 , 2 ] , λ 1 ( 0 , λ + 1 ) , λ 2 ( 0 , 1 2 ] ( 0 , λ + 1 ) ,   λ ^ 1 : = λ λ 2 p + λ 1 q , λ ^ 2 : = λ λ 1 q + λ 2 p . Then, we have the following half-discrete Hilbert-type inequality:
I : = 0 n = 1 a n ( x + n η ) λ f ( x ) d x < Γ ( λ + 2 ) Γ ( λ ) ( k λ + 2 ( λ 2 + 1 ) ) 1 p ( k λ + 2 ( λ 1 + 1 ) ) 1 q  
× [ 0 x p λ ^ 1 1 F p ( x ) d x ] 1 p [ n = 1 ( n η ) q λ ^ 2 1 A n q ] 1 q .
In particular, for λ 1 + λ 2 = λ ( ( 0 , 2 ] ) ( λ 1 ( 0 , λ ) , λ 2 ( 0 , 1 2 ] ( 0 , λ ) ) , we have:
0 n = 1 a n ( x + n η ) λ f ( x ) d x < λ 1 λ 2 B ( λ 1 , λ 2 )  
× [ 0 x p λ 1 1 F p ( x ) d x ] 1 p [ n = 1 ( n η ) q λ 2 1 A n q ] 1 q .
Proof. 
Since for λ > 0 , we have:
1 ( x + n η ) λ = 1 Γ ( λ ) 0 t λ 1 e ( x + n η ) t d t ,
it follows that:
I = 1 Γ ( λ ) 0 n = 1 a n f ( x ) 0 t λ 1 e ( x + n η ) t d t d x = 1 Γ ( λ ) 0 t λ 1 0 e x t f ( x ) d x n = 1 e ( n η ) t a n d t 1 Γ ( λ ) 0 t λ + 1 0 e x t F ( x ) d x n = 1 e ( n η ) t A n d t
= 1 Γ ( λ ) 0 n = 1 A n F ( x ) 0 t ( λ + 2 ) 1 e ( x + n η ) t d t d x
= Γ ( λ + 2 ) Γ ( λ ) 0 n = 1 A n ( x + n η ) λ + 2 F ( x ) d x
By applying (15), we obtain (18).
In particular, for λ 1 + λ 2 = λ ( ( 0 , 2 ] ) ( λ 1 ( 0 , λ ) , λ 2 ( 0 , 1 2 ] ( 0 , λ ) ) , one has:
k λ + 2 ( λ 2 + 1 ) = k λ + 2 ( λ 1 + 1 ) = B ( λ 1 + 1 , λ 2 + 1 ) = Γ ( λ 1 + 1 ) Γ ( λ 2 + 1 ) Γ ( λ + 2 ) = λ 1 λ 2 Γ ( λ 1 ) Γ ( λ 2 ) Γ ( λ + 2 ) = Γ ( λ ) Γ ( λ + 2 ) λ 1 λ 2 B ( λ 1 , λ 2 ) .
Hence, it follows from (18) that:
0 n = 1 a n ( x + n η ) λ f ( x ) d x < λ 1 λ 2 B ( λ 1 , λ 2 )
× [ 0 x p λ 1 1 F p ( x ) d x ] 1 p [ n = 1 ( n η ) q λ 2 1 A n q ] 1 q ,
which is the desired inequality (19).
Remark 3. 
Putting η = 0 in (20), we have:
0 n = 1 a n ( x + n ) λ f ( x ) d x < λ 1 λ 2 B ( λ 1 , λ 2 )  
× [ 0 x p λ 1 1 F p ( x ) d x ] 1 p [ n = 1 n q λ 2 1 A n q ] 1 q .  
Namely, (18) given by Theorem 1 is a more accurate extension of (21) above. It should be noted that here the statement of “more accurate inequality” borrows from the statement mentioned at the beginning of the paper on the comparison between inequalities (1) and (2) described in the previous literature.
Theorem 2. 
If λ λ 1 1 2 , then the following statements (i), (ii) and (iii), associated with Theorem 1, are equivalent:
(i) 
( k λ + 2 ( λ 2 + 1 ) ) 1 p ( k λ + 2 ( λ 1 + 1 ) ) 1 q k λ + 2 ( λ λ 2 p + λ 1 q + 1 ) ;
(ii) 
λ 1 + λ 2 = λ ( ( 0 , 2 ] ) , where λ 1 ( 0 , λ ) , λ 2 ( 0 , 1 2 ] ( 0 , λ ) ;
(iii) 
The constant factor:
Γ ( λ + 2 ) Γ ( λ ) ( k λ + 2 ( λ 2 + 1 ) ) 1 p ( k λ + 2 ( λ 1 + 1 ) ) 1 q  
in (19) is the best possible.
Proof. 
(i) (ii)”. By using Hölder inequality with weight, we obtain:
k λ + 2 ( λ λ 2 p + λ 1 q + 1 ) = 0 1 ( 1 + u ) λ + 2 u λ λ 2 p + λ 1 q d u = 0 1 ( 1 + u ) λ + 2 ( u λ λ 2 p ) ( u λ 1 q ) d u
[ 0 1 ( 1 + u ) λ + 2 u λ λ 2 d u ] 1 p [ 0 1 ( 1 + u ) λ + 2 u λ 1 d u ] 1 q
= [ 0 1 ( 1 + v ) λ + 2 v ( λ 2 + 1 ) 1 d v ] 1 p [ 0 1 ( 1 + u ) λ + 2 u ( λ 1 + 1 ) 1 d u ] 1 q
= ( k λ + 2 ( λ 2 + 1 ) ) 1 p ( k λ + 2 ( λ 1 + 1 ) ) 1 q .
In view of inequality (i), we conclude that (22) keeps the form of equality.
We observe that (22) keeps the form of equality if and only if there exist constants A and B , such that they are not both zero and A u λ λ 2 = B u λ 1 a . e . in R + (see [18]). Assuming that A 0 , we have u λ λ 2 λ 1 = B A a . e . in R + , and then λ λ 2 λ 1 = 0 , namely, λ 1 + λ 2 = λ ( ( 0 , 2 ] ) , where, λ 1 ( 0 , λ ) , λ 2 ( 0 , 1 2 ] ( 0 , λ ) .
“(ii) (iii)”. For λ 1 + λ 2 = λ ( ( 0 , 2 ] ) , λ 1 ( 0 , λ ) , λ 2 ( 0 , 1 2 ] ( 0 , λ ) , (19) reduces to (20). For any 0 < ε < min { p λ 1 , q λ 2 } , we set:
f ˜ ( x ) : = 0 , 0 < x < 1 , x λ 1 ε p 1 , x 1 , a ˜ n : = n λ 2 ε q 1 ( n N )
Then, it follows that:
F ˜ ( x ) : = 0 x f ˜ ( t ) d t 0 , 0 < x < 1 , 1 λ 1 ε p x λ 1 ε p , x 1 , A ˜ n : = k = 1 n a ˜ k = k = 1 n k λ 2 ε q 1 < 0 n t λ 2 ε q 1 d t = 1 λ 2 ε q n λ 2 ε q ( n N ) .
If there exists a positive constant M λ 1 λ 2 B ( λ 1 , λ 2 ) such that (20) is valid when replacing λ 1 λ 2 B ( λ 1 , λ 2 ) by M , then, in particular for η = 0 , by substitution of f ( x ) = f ˜ ( x ) , a n = a ˜ n ,   F ( x ) = F ˜ ( x ) and A n = A ˜ n in (21), we have:
I ˜ : = 0 n = 1 a ˜ n f ˜ ( x ) ( x + n ) λ d x < M ( 0 x p λ 1 1 F ˜ p ( x ) d x ) 1 p ( n = 1 n q λ 2 1 A ˜ n q ) 1 q .
In the following, we show that λ 1 λ 2 B ( λ 1 , λ 2 ) M , and then M = λ 1 λ 2 B ( λ 1 , λ 2 ) is the best possible constant factor in (20).
By (23) and the decreasingness property of series, we obtain:
I ˜ < M 1 λ 1 ε p ( 1 x p λ 1 1 x p λ 1 ε d x ) 1 p 1 λ 2 ε q ( n = 1 n q λ 2 1 n q λ 2 ε ) 1 q
= M ( 1 λ 1 ε p ) ( 1 λ 2 ε q ) ( 1 x ε 1 d x ) 1 p ( 1 + n = 2 n ε 1 ) 1 q < M ( 1 λ 1 ε p ) ( 1 λ 2 ε q ) ( 1 x ε 1 d x ) 1 p ( 1 + 1 y ε 1 d y ) 1 q
= M ε ( 1 λ 1 ε p ) ( 1 λ 2 ε q ) ( ε + 1 ) 1 q .
By (11) (for η = 0 ), setting λ ˜ 2 = λ 2 ε q ( 0 , 1 2 ) ( 0 , λ ) ( 0 < λ ˜ 1 = λ 1 + ε q < λ ) , we obtain:
I ˜ = 1 [ x ( λ 1 + ε q ) n = 1 1 ( x + n ) λ n ( λ 2 ε q ) 1 ] x ε 1 d x = 1 ϖ ( λ ˜ 2 , x ) x ε 1 d x > B ( λ ˜ 1 , λ ˜ 2 ) 1 [ 1 O ( 1 x λ ^ 2 ) ] x ε 1 d x = B ( λ ˜ 1 , λ ˜ 2 ) [ 1 x ε 1 d x 1 O ( 1 x λ 2 + ε p + 1 ) d x ] = 1 ε B ( λ 1 + ε q , λ 2 ε q ) ( 1 ε O ( 1 ) ) .
Then, in virtue of the above results, we have:
B ( λ 1 + ε q , λ 2 ε q ) ( 1 ε O ( 1 ) ) < ε I ˜ < M ( 1 λ 1 ε p ) ( 1 λ 2 ε q ) ( ε + 1 ) 1 q .
Putting ε 0 + , in view of the continuity of the beta function, we obtain λ 1 λ 2 B ( λ 1 , λ 2 ) M . Hence, M = λ 1 λ 2 B ( λ 1 , λ 2 ) is the best possible constant factor in (20).
“(iii) (i)”. Since λ λ 1 1 2 , for λ ^ 1 = λ λ 2 p + λ 1 q , λ ^ 2 = λ λ 1 q + λ 2 p , we find:
λ ^ 1 + λ ^ 2 = λ λ 2 p + λ 1 q + λ λ 1 q + λ 2 p = λ , 0 < λ ^ 1 , λ ^ 2 < λ p + λ q = λ ,
λ ^ 2 1 / 2 p + 1 / 2 q = 1 2 , and λ ^ 1 λ ^ 2 B ( λ ^ 1 , λ ^ 2 ) R + .
If the constant factor Γ ( λ + 2 ) Γ ( λ ) ( k λ + 2 ( λ 2 + 1 ) ) 1 p ( k λ + 2 ( λ 1 + 1 ) ) 1 q in (19) is the best possible, then by (21) (for λ i = λ ^ i ( i = 1 , 2 ) ), we have:
Γ ( λ + 2 ) Γ ( λ ) ( k λ + 2 ( λ 2 + 1 ) ) 1 p ( k λ + 2 ( λ 1 + 1 ) ) 1 q
λ ^ 1 λ ^ 2 B ( λ ^ 1 , λ ^ 2 ) = Γ ( λ + 2 ) Γ ( λ ) k λ + 2 ( λ ^ 1 + 1 ) = Γ ( λ + 2 ) Γ ( λ ) k λ + 2 ( λ λ 2 p + λ 1 q + 1 ) ( R + ) ,
namely, statement (i) is valid.
Hence, the statements (i), (ii) and (iii) are equivalent. This completes the proof of Theorem 2.
Remark 4. 
Putting  η = 1 4 in (20), we acquire:
0 n = 1 a n ( x + n 1 4 ) λ f ( x ) d x < λ 1 λ 2 B ( λ 1 , λ 2 )  
× [ 0 x p λ 1 1 F p ( x ) d x ] 1 p [ n = 1 ( n 1 4 ) q λ 2 1 A n q ] 1 q  
In particular, for λ = 1 , λ 1 = λ 2 = 1 2 , we have the following Hilbert-type inequality with the best possible constant factor π 4 :
0 n = 1 a n x + n 1 4 f ( x ) d x < π 4 [ 0 x p 2 1 F p ( x ) d x ] 1 p [ n = 1 ( n 1 4 ) q 2 1 A n q ] 1 q .

4. Conclusions

In this paper, based on the weight coefficients and the idea of introducing parameters, by applying Hermite–Hadamard inequality, the Euler–Maclaurin summation formula and Abel’s summation by parts formula, a more accurate half-discrete Hilbert-type inequality involving one upper limit function as well as one partial sum is given in Theorem 1. The equivalent statements of the best possible constant factor related to several parameters are considered in Theorem 2. As applications of the main results, some new inequalities are proposed in Remarks 3 and 4. Our results would provide a significant supplement to the study of half-discrete Hilbert-type inequality.

Author Contributions

B.Y. carried out the mathematical studies, participated in the sequence alignment and drafted the manuscript. S.W. and X.H. participated in the design of the study and performed the numerical analysis. All authors contributed equally to the preparation of this paper. All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by the National Natural Science Foundation (No. 61772140, No.12071491), the Characteristic Innovation Project of Guangdong Provincial Colleges and Universities (No.2020KTSCX088), the Construction Project of Teaching Quality and Teaching Reform in Guangdong Undergraduate Colleges and Universities in 2018 (Speciality of Financial Mathematics), and the Natural Science Foundation of Fujian Province of China (No. 2020J01365 ).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data used to support the findings of this study are included within the article.

Conflicts of Interest

The authors declare that they have no competing interest.

References

  1. Hardy, G.H.; Littlewood, J.E.; Polya, G. Inequalities; Cambridge University Press: Cambridge, UK, 1934. [Google Scholar]
  2. Krnić, M.; Pečarić, J. Extension of Hilbert’s inequality. J. Math. Anal. Appl. 2006, 324, 150–160. [Google Scholar] [CrossRef] [Green Version]
  3. Yang, B. On a generalization of Hilbert’s double series theorem. Math. Inequalities Appl. 2002, 18, 197–204. [Google Scholar] [CrossRef] [Green Version]
  4. Adiyasuren, V.; Batbold, T.; Azar, L.E. A new discrete Hilbert-type inequality involving partial sums. J. Inequalities Appl. 2019, 2019, 127. [Google Scholar] [CrossRef]
  5. Yang, B.C. The Norm of Operator and Hilbert-Type Inequalities; Science Press: Beijing, China, 2009. [Google Scholar]
  6. Krnić, M.; Pečarić, J. General Hilbert’s and Hardy’s inequalities. Math. Inequal. Appl. 2005, 8, 29–51. [Google Scholar] [CrossRef] [Green Version]
  7. Perić, I.; Vuković, P. Multiple Hilbert’s type inequalities with a homogeneous kernel. Banach J. Math. Anal. 2011, 5, 33–43. [Google Scholar] [CrossRef]
  8. Adiyasuren, V.; Batbold, T.; Krnić, M. Hilbert-type inequalities involving differential operators, the best constants, and applications. Math. Inequalities Appl. 2015, 111–124. [Google Scholar] [CrossRef] [Green Version]
  9. Hong, Y.; Wen, Y.M. A necessary and sufficient condition of that Hilbert type series inequality with homogeneous kernel has the best constant factor. Ann. Math. 2016, 37, 329–336. [Google Scholar]
  10. Rassias, M.T.; Yang, B. On half-discrete Hilbert’s inequality. Appl. Math. Comput. 2013, 220, 75–93. [Google Scholar] [CrossRef]
  11. Yang, B.C.; Krnić, M. A half-discrete Hilbert-type inequality with a general homogeneous kernel of degree 0. J. Math. Inequal. 2012, 6, 401–417. [Google Scholar]
  12. Rassias, M.T.; Yang, B. A multidimensional half-discrete Hilbert-type inequality and the Riemann zeta function. Appl. Math. Comput. 2013, 225, 263–277. [Google Scholar] [CrossRef]
  13. Rassias, M.T.; Yang, B. On a multidimensional half-discrete Hilbert-type inequality related to the hyperbolic cotangent function. Appl. Math. Comput. 2014, 242, 800–813. [Google Scholar] [CrossRef]
  14. Yang, B.C.; Debnath, L. Half-Discrete Hilbert-Type Inequalities; World Scientific Publishing: Singapore, 2014. [Google Scholar]
  15. Huang, Z.; Yang, B. Equivalent property of a half-discrete Hilbert’s inequality with parameters. J. Inequalities Appl. 2018, 2018. [Google Scholar] [CrossRef] [PubMed]
  16. Yang, B.; Wu, S.; Wang, A. On a Reverse Half-Discrete Hardy-Hilbert’s Inequality with Parameters. Mathematics 2019, 7, 1054. [Google Scholar] [CrossRef] [Green Version]
  17. Wang, A.-Z.; Yang, B.-C.; Chen, Q. Equivalent properties of a reverse half-discrete Hilbert’s inequality. J. Inequalities Appl. 2019, 2019. [Google Scholar] [CrossRef]
  18. Kuang, J.C. Applied Inequalities; Shangdong Science and Technology Press: Jinan, China, 2004. [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

Huang, X.; Wu, S.; Yang, B. A More Accurate Half-Discrete Hilbert-Type Inequality Involving One upper Limit Function and One Partial Sum. Symmetry 2021, 13, 1548. https://doi.org/10.3390/sym13081548

AMA Style

Huang X, Wu S, Yang B. A More Accurate Half-Discrete Hilbert-Type Inequality Involving One upper Limit Function and One Partial Sum. Symmetry. 2021; 13(8):1548. https://doi.org/10.3390/sym13081548

Chicago/Turabian Style

Huang, Xianyong, Shanhe Wu, and Bicheng Yang. 2021. "A More Accurate Half-Discrete Hilbert-Type Inequality Involving One upper Limit Function and One Partial Sum" Symmetry 13, no. 8: 1548. https://doi.org/10.3390/sym13081548

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