Next Article in Journal
Study on Temperature Field Distribution Law and Mechanical Properties of Hydraulic Tunnel-Surrounding Rock under the Action of Large Temperature Differences
Next Article in Special Issue
Symmetry Analysis for the 2D Aw-Rascle Traffic-Flow Model of Multi-Lane Motorways in the Euler and Lagrange Variables
Previous Article in Journal
Cosine Similarity Measures of (m, n)-Rung Orthopair Fuzzy Sets and Their Applications in Plant Leaf Disease Classification
Previous Article in Special Issue
The Influence of the Perturbation of the Initial Data on the Analytic Approximate Solution of the Van der Pol Equation in the Complex Domain
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Applications of Symmetric Identities for Apostol–Bernoulli and Apostol–Euler Functions

School of Mathematics and Information Science, Neijiang Normal University, Neijiang 641100, China
Symmetry 2023, 15(7), 1384; https://doi.org/10.3390/sym15071384
Submission received: 25 June 2023 / Revised: 3 July 2023 / Accepted: 4 July 2023 / Published: 8 July 2023

Abstract

:
In this paper, we perform a further investigation on the Apostol–Bernoulli and Apostol–Euler functions introduced by Luo. By using the Fourier expansions of the Apostol–Bernoulli and Apostol–Euler polynomials, we establish some symmetric identities for the Apostol–Bernoulli and Apostol–Euler functions. As applications, some known results, for example, Raabe’s multiplication formula and Hermite’s identity, are deduced as special cases.

1. Introduction

Throughout this paper, N , N 0 , Z , R , and C will denote the set of natural numbers, the set of non-negative integers, the set of integers, the set of real numbers, and the set of complex numbers, respectively. Let n N , ω C , and let x be a variable. The Apostol–Bernoulli polynomials B n ( x ; ω ) and the Apostol–Euler polynomials E n ( x ; ω ) are usually defined by the following generating functions (see, e.g., [1,2]):
t e x t ω e t 1 = n = 0 B n ( x ; ω ) t n n ! ( | t + log ω | < 2 π ) ,
and
2 e x t ω e t + 1 = n = 0 E n ( x ; ω ) t n n ! ( | t + log ω | < π ) .
In particular, B n ( x ) = B n ( x ; 1 ) and E n ( x ) = E n ( x ; 1 ) are called the Bernoulli polynomials and the Euler polynomials, respectively. It is well known that the Apostol–Bernoulli numbers B n ( ω ) and the Apostol–Euler numbers E n ( ω ) are given by B n ( ω ) = B n ( 0 ; ω ) and E n ( ω ) = 2 n E n ( 1 / 2 ; ω ) , which are the corresponding generalizations of the Bernoulli numbers B n = B n ( 0 ) and the Euler numbers E n = 2 n E n ( 1 / 2 ) . We here mention that the Apostol–Bernoulli polynomials can be used to evaluate the values of the Lerch zeta function at zero and negative integers, see the paper of Apostol [3] for details. For the values of the Apostol–Bernoulli polynomials and the Lerch zeta function at rational numbers, the interested reader may consult the paper of Srivastava [4].
In the present paper, we will be concerned with some identities for the above-mentioned polynomials. Perhaps the best known results are the following multiplication formulas for the Bernoulli polynomials and the Euler polynomials (see, e.g., [5,6,7]):
a n 1 j = 0 a 1 B n ( x + j a ) = B n ( a x ) ( a N , n N 0 ) ,
a n j = 0 a 1 ( 1 ) j E n ( x + j a ) = E n ( a x ) ( a N , 2 a , n N 0 ) ,
2 a n 1 n j = 0 a 1 ( 1 ) j B n ( x + j a ) = E n 1 ( a x ) ( a , n N , 2 a ) .
We remark that Formula (3) is usually called Raabe’s [8] multiplication formula for the Bernoulli polynomials. In particular, Howard [9] used (3) to obtain this for a , n N with a 2 ,
B n = 1 a ( 1 a n ) j = 0 n 1 n j a j B j l = 1 a 1 l n j ,
from which he showed that the famous Staudt–Clausen theorem, Carlitz’s congruence, Frobenius’s congruence, and Ramanujan’s congruence are deduced as easy consequences of (6). In the year 2001, Tuenter [10] found that (6) is a special case of the following symmetric identity for the Bernoulli numbers:
j = 0 n n j a j 1 B j b n j S n j ( a 1 ) = j = 0 n n j b j 1 B j a n j S n j ( b 1 ) ,
where a , b N , n N 0 , S k ( n ) is the power sum given for k , n N 0 by
S k ( n ) = 0 k + 1 k + 2 k + + n k .
Yang [11], in 2008, used the method of generating functions to establish two identities of symmetry for the higher-order Bernoulli polynomials, by virtue of which he extended (7) to the situation for the Bernoulli polynomials
j = 0 n n j a j 1 B j ( b x ) b n j S n j ( a 1 ) = j = 0 n n j b j 1 B j ( a x ) a n j S n j ( b 1 ) ,
and obtained the general form of (3):
a n 1 j = 0 a 1 B n ( b x + b j a ) = b n 1 j = 0 b 1 B n ( a x + a j b ) ,
where a , b N , n N 0 , S k ( n ) is as in (8). In the same year, Kim [12] used the properties of symmetry for p-adic invariant integrals on Z p to prove (9) and (10), demonstrating that for a , b N , n N 0 with a b (mod 2),
j = 0 n n j a j E j ( b x ) b n j T n j ( a 1 ) = j = 0 n n j b j E j ( a x ) a n j T n j ( b 1 ) ,
and
a n j = 0 a 1 ( 1 ) j E n ( b x + b j a ) = b n j = 0 b 1 ( 1 ) j E n ( a x + a j b ) ,
where T k ( n ) is the alternate power sum given for k , n N 0 by
T k ( n ) = 0 k 1 k + 2 k + ( 1 ) n n k .
After that, Liu and Wang [13] developed the method of generating functions used in [11], and established various identities for the higher-order Bernoulli polynomials, the higher-order Euler polynomials, and the higher-order degenerate Bernoulli polynomials, some of which extend (9)–(12). Further, Wang and Wang [14] obtained various identities between the Apostol–Bernoulli polynomials, the Apostol–Euler polynomials, and the power sums with respect to ω , showing that Yang’s [11], Liu and Wang’s [13], and Zhang and Yang’s [15] results are deduced as special cases. The author and Zhang [16,17] in 2013 explored the identities of symmetry for the q-zeta functions and the q-Lerch Euler zeta functions, and gave the corresponding q-extensions of (9)–(12). It should be noted that the identities (10) and (12) can easily lead to the identities (9) and (11) when using the familiar addition theorems for the Bernoulli polynomials and the Euler polynomials described in [5,6]. For some new developments of identities of symmetry on these topics, one is referred to [18,19,20,21,22,23,24].
In this paper, we perform further investigation on the Apostol–Bernoulli functions B ¯ n ( x ; ω ) and the Apostol–Euler functions E ¯ n ( x ; ω ) introduced by Luo ([25] Equations (2.15) and (2.17)), which are defined for n N , x R , ω C { 0 } by
B ¯ n ( x ; ω ) = ω [ x ] B n ( { x } ; ω ) ,
and
E ¯ n ( x ; ω ) = ( 1 ) [ x ] ω [ x ] E n ( { x } ; ω ) ,
where [ x ] is the floor function (also called the greatest integer function), { x } denotes the fractional part of x satisfying that for x R ,
{ x } = x [ x ] .
By using the Fourier expansions of the Apostol–Bernoulli polynomials and the Apostol–Euler polynomials shown in [25,26], we establish some symmetric identities for the Apostol–Bernoulli functions and the Apostol–Euler functions. It turns out that some known results, for example, Raabe’s [8] multiplication formula for the Bernoulli functions, Bayad’s [27] multiplication formula for the Euler functions, and Hermite’s [28] identity for the floor function, are obtained as special cases. Moreover, we also show that a relation for the number of lattice points in the case of triangles, a symmetric identity for the sums considered by Cetin et al. [29], are established as easy consequences.

2. An Auxiliary Lemma

Before giving our main results, we first present the following auxiliary lemma.
Lemma 1.
Let a , b N . Then, for x R ,
j = 0 a 1 δ Z ( b x + b j a ) = j = 0 b 1 δ Z ( a x + a j b ) ,
where δ Z ( x ) = 1 or 0 according to x Z or x R Z .
Proof. 
We write a = a 1 d and b = b 1 d , where d = ( a , b ) , we have ( a 1 , b 1 ) = 1 . Hence, we obtain from the familiar division algorithm stated in ([30] Theorem 1.14) that for each j { 0 , 1 , , a 1 1 } , there exists q j Z and unique r j { 0 , 1 , , a 1 1 } such that
b 1 j = a 1 q j + r j .
It follows from (18) that
j = 0 a 1 δ Z ( b x + b j a ) = d j = 0 a 1 1 δ Z ( b 1 d x + b 1 j a 1 ) = d j = 0 a 1 1 δ Z ( b 1 d x + j a 1 ) = d , a 1 b 1 d x Z , 0 , a 1 b 1 d x R Z .
Similarly, we have
j = 0 b 1 δ Z ( a x + a j b ) = d , a 1 b 1 d x Z , 0 , a 1 b 1 d x R Z .
Thus, by equating (19) and (20), we obtain (17) immediately. □

3. Statement of Main Results

For convenience, in the following we always denote by i the square root of 1 such that i 2 = 1 . For the sake of convergence, the sum
k = + 1 k + a ( a Z )
is interpreted as
lim N k = N N 1 k + a .
We now give the symmetric identity for the Apostol–Bernoulli functions as follows.
Theorem 1.
Let a , b , n N , ω C { 0 } . Then, for x R ,
a n 1 j = 0 a 1 ω b j B ¯ n ( b x + b j a ; ω a ) = b n 1 j = 0 b 1 ω a j B ¯ n ( a x + a j b ; ω b ) .
Proof. 
We know from (14) and the Fourier expansions of the Apostol–Bernoulli polynomials shown in ([25] Theorem 2.1) or ([26] Theorem 1.1) that the Apostol–Bernoulli functions can be defined for n N , x R , ω C { 0 } by
B ¯ n ( x ; ω ) = n ! ω x ( 2 π i ) n k Z * e 2 π i k x ( k log ω 2 π i ) n ,
where k Z * = k Z { 0 } if ω = 1 , and k Z * = k Z if ω 1 . By replacing x with b x + b j a and ω by ω a in (22), we have
B ¯ n ( b x + b j a ; ω a ) = n ! ω a b x + b j ( 2 π i ) n k Z * e 2 π i b k x · e 2 π i b k j a ( k a log ω 2 π i ) n .
It follows from (23) and the familiar geometric sums stated in ([30] Theorem 8.1) that
a n 1 j = 0 a 1 ω b j B ¯ n ( b x + b j a ; ω a ) = n ! · a n 1 ω a b x ( 2 π i ) n k Z * e 2 π i b k x ( k a log ω 2 π i ) n j = 0 a 1 e 2 π i b k j a = n ! · a n ω a b x ( 2 π i ) n k Z a b k * e 2 π i b k x ( k a log ω 2 π i ) n .
Let d = ( a , b ) . If we write a = a 1 d and b = b 1 d , then we can rewrite (24) as
a n 1 j = 0 a 1 ω b j B ¯ n ( b x + b j a ; ω a ) = n ! · a n ω a b x ( 2 π i ) n k Z a 1 k * e 2 π i b k x ( k a log ω 2 π i ) n = n ! · a n ω a b x ( 2 π i ) n k Z * e 2 π i a 1 b k x ( a 1 k a 1 d log ω 2 π i ) n = n ! · ( a , b ) n ω a b x ( 2 π i ) n k Z * e 2 π i a b k x ( a , b ) ( k ( a , b ) log ω 2 π i ) n .
Replacing a with b and b with a in (25), we have
b n 1 j = 0 b 1 ω a j B ¯ n ( a x + a j b ; ω b ) = n ! · ( a , b ) n ω a b x ( 2 π i ) n k Z * e 2 π i a b k x ( a , b ) ( k ( a , b ) log ω 2 π i ) n .
Therefore, we obtain (21) immediately when equating (25) and (26). This completes the proof of Theorem 1. □
It follows that we show some special cases of Theorem 1. We have the following results.
Corollary 1.
Let a , n N , ω C { 0 } . Then, for x R ,
a n 1 j = 0 a 1 ω j B ¯ n ( x + j a ; ω a ) = B ¯ n ( a x ; ω ) .
Proof. 
Taking b = 1 in Theorem 1, we obtain the desired result. □
In particular, the case ω = 1 in Corollary 1 gives Raabe’s [8] multiplication formula for the Bernoulli functions, namely,
a n 1 j = 0 a 1 B ¯ n ( x + j a ) = B ¯ n ( a x ) ,
where a , n N , x R , B ¯ n ( x ) is the n-th Bernoulli function given for n N , x R by
B ¯ 1 ( x ) = 0 , x Z , B 1 ( { x } ) , x R Z , B ¯ n ( x ) = B n ( { x } ) ( n 2 , x R ) .
We note that Carlitz [31] and Bayad and Raouj [32] used Raabe’s multiplication Formula (28) to establish some reciprocity formulas for the generalized Dedekind sums.
Corollary 2.
Let a , b , n N . Then, for x R ,
a n 1 j = 0 a 1 B ¯ n ( b x + b j a ) = b n 1 j = 0 b 1 B ¯ n ( a x + a j b ) .
Proof. 
By setting ω = 1 in Theorem 1, the desired result follows immediately. □
It becomes obvious that the case b = 1 in Corollary 2 leads to Raabe’s multiplication Formula (28). We next use Corollary 2 to deduce the following result.
Corollary 3.
Let a , b N . Then, for x R ,
j = 0 a 1 [ b x + b j a ] = j = 0 b 1 [ a x + a j b ] .
Proof. 
Denote by δ l , k the Kronecker delta function given for l , k N { 0 } by δ l , k = 1 or 0 according to l = k or l k . It is easy to see from (29) that for n N , x R ,
B ¯ n ( x ) = B n ( { x } ) + 1 2 δ 1 , n δ Z ( x ) ,
where δ Z ( x ) is as in (17). Since B 1 ( x ) = x 1 2 , by taking n = 1 in Corollary 2, in light of (16) and Lemma 1, we prove Corollary 3. □
If we take b = 1 in Corollary 3, then we obtain Hermite’s [28] identity for the floor function, namely,
j = 0 a 1 [ x + j a ] = [ a x ] ,
where a N , x R . For another generalizations of Hermite’s identity (33), the interested reader may consult [33]. As another application of Corollary 3, we here give the following relation for the number of lattice points in the case of triangles.
Corollary 4.
Let a , b N , c R . Assume that T 1 ( a , b , c ) represents the triangle in R 2 with vertices
( a { c } , 0 ) , ( a , 0 ) , ( a , b + b { c } ) ,
T 2 ( a , b , c ) represents the triangle in R 2 with vertices
( b { c } , 0 ) , ( b , 0 ) , ( b , a + a { c } ) .
Then,
# ( T 1 ( a , b , c ) Z 2 ) = # ( T 2 ( a , b , c ) Z 2 ) ,
where # denotes the cardinality of a set S.
Proof. 
Since the number of lattice points on the acute angles of T 1 ( a , b , c ) is
[ b + b { c } ] + a + 1 + [ a { c } ] = a + b + 1 + [ a { c } ] + [ b { c } ] ,
the number of lattice points on the acute angles of T 2 ( a , b , c ) is
[ a + a { c } ] + b + 1 + [ b { c } ] = a + b + 1 + [ a { c } ] + [ b { c } ] ,
by taking x = 0 in Corollary 3, in view of (35) and (36), we see that Corollary 4 holds true in the case when c = 0 . We now consider the case c 0 . Let T 3 ( a , b , c ) be the triangle in R 2 with vertices
( a { c } , 0 ) , ( 0 , 0 ) , ( 0 , b { c } ) ,
and T 4 ( a , b , c ) be the triangle in R 2 with vertices
( b { c } , 0 ) , ( 0 , 0 ) , ( 0 , a { c } ) .
Obviously, the graph of T 3 ( a , b , c ) and the graph of T 4 ( a , b , c ) are symmetric about the line y = x . This indicates that the number of lattice points in the interior of T 3 ( a , b , c ) is equal to the number of lattice points in the interior of T 4 ( a , b , c ) , and the number of lattice points on the hypotenuse of T 3 ( a , b , c ) is equal to the number of lattice points on the hypotenuse of T 4 ( a , b , c ) . Therefore, by taking x = { c } in Corollary 3, we say from (35) and (36) that Corollary 4 holds true in the case when c 0 . This completes the proof of Corollary 4. □
We next present the symmetric identity for the Apostol–Euler functions, as follows.
Theorem 2.
Let a , b , n N , ω C { 0 } with a b (mod 2). Then, for x R ,
a n j = 0 a 1 ( 1 ) j ω b j E ¯ n ( b x + b j a ; ω a ) = b n j = 0 b 1 ( 1 ) j ω a j E ¯ n ( a x + a j b ; ω b ) .
Proof. 
We see from (15) and the Fourier expansions of the Apostol–Euler polynomials shown in ([25] Theorem 2.2) or ([26] Theorem 1.2) that the Apostol–Euler functions can be defined for n N , x R , ω C { 0 } by
E ¯ n ( x ; ω ) = 2 · n ! ω x ( 2 π i ) n + 1 k Z * * e 2 π i ( k 1 2 ) x ( k 1 2 log ω 2 π i ) n + 1 ,
where k Z * * = k Z { 0 } if ω = 1 , and k Z * * = k Z if ω 1 . If we replace x with b x + b j a and ω with ω a in (38), then we have
E ¯ n ( b x + b j a ; ω a ) = 2 · n ! ω a b x + b j ( 2 π i ) n + 1 k Z * * e 2 π i ( k 1 2 ) b x · e 2 π i ( b k b 2 ) j a ( k 1 2 a log ω 2 π i ) n + 1 .
Hence, we discover from (39) and the geometric sums stated in ([30] Theorem 8.1) that
a n j = 0 a 1 ( 1 ) j ω b j E ¯ n ( b x + b j a ; ω a ) = 2 · n ! · a n ω a b x ( 2 π i ) n + 1 k Z * * e 2 π i ( k 1 2 ) b x ( k 1 2 a log ω 2 π i ) n + 1 j = 0 a 1 e 2 π i ( b k b a 2 ) j a = 2 · n ! · a n + 1 ω a b x ( 2 π i ) n + 1 k Z a ( b k b a 2 ) * * e 2 π i ( k 1 2 ) b x ( k 1 2 a log ω 2 π i ) n + 1 .
It is easily seen from a ( b k b a 2 ) for k Z that there exists q Z such that
b ( 2 k 1 ) = a ( 2 q 1 ) .
If we write a = a 1 d and b = b 1 d , where d = ( a , b ) , then we conclude from (41) that
a 1 ( 2 k 1 ) ( k Z ) .
It follows from (42) that (40) can be rewritten as
a n j = 0 a 1 ( 1 ) j ω b j E ¯ n ( b x + b j a ; ω a ) = 2 · n ! · a n + 1 ω a b x ( 2 π i ) n + 1 k Z a 1 ( 2 k 1 ) * * e 2 π i ( k 1 2 ) b x ( k 1 2 a log ω 2 π i ) n + 1 = 2 · n ! · a n + 1 ω a b x ( 2 π i ) n + 1 k Z * * e π i a 1 b k x ( a 1 k 2 a log ω 2 π i ) n + 1 = 2 · n ! · ( a , b ) n + 1 ω a b x ( 2 π i ) n + 1 k Z * * e π i a b k x ( a , b ) ( k 2 ( a , b ) log ω 2 π i ) n + 1 .
Replacing a with b and b with a in (43) gives
b n j = 0 b 1 ( 1 ) j ω a j E ¯ n ( a x + a j b ; ω b ) = 2 · n ! · ( a , b ) n + 1 ω a b x ( 2 π i ) n + 1 k Z * * e π i a b k x ( a , b ) ( k 2 ( a , b ) log ω 2 π i ) n + 1 .
Thus, by equating (43) and (44), we obtain (37) immediately and finish the proof of Theorem 2. □
We next discuss some special cases of Theorem 2. We have the following results.
Corollary 5.
Let a , n N , ω C { 0 } with 2 a . Then, for x R ,
a n j = 0 a 1 ( ω ) j E ¯ n ( x + j a ; ω a ) = E ¯ n ( a x ; ω ) .
Proof. 
Setting b = 1 in Theorem 2, we obtain the desired result. □
Trivially, the case ω = 1 in Corollary 5 gives that for a , n N , x R with 2 a ,
a n j = 0 a 1 ( 1 ) j E ¯ n ( x + j a ) = E ¯ n ( a x ) ,
where E ¯ n ( x ) is the n-th quasi-periodic Euler function given for n N , x R by
E ¯ n ( x ) = ( 1 ) [ x ] E n ( { x } ) .
It is worth mentioning that Formula (46) was also discovered by Bayad ([27] Equation (1.2.13)), and has been used by Kim and Son [34] and Hu, Kim, and Kim [35] to establish some reciprocity formulas for the generalized Dedekind sums involving quasi-periodic Euler functions.
Corollary 6.
Let a , b , n N with a b (mod 2). Then, for x R ,
a n j = 0 a 1 ( 1 ) j E ¯ n ( b x + b j a ) = b n j = 0 b 1 ( 1 ) j E ¯ n ( a x + a j b ) .
Proof. 
Taking ω = 1 in Theorem 2 gives the desired result. □
It is clear that the case b = 1 in Corollary 6 gives Bayad’s multiplication formula (46). In fact, we can use Corollary 6 to yield the following result.
Corollary 7.
Let a , b N with a b (mod 2). Then, for x R ,
a j = 0 a 1 ( 1 ) j + [ b x + b j a ] ( b x + b j a [ b x + b j a ] 1 2 ) = b j = 0 b 1 ( 1 ) j + [ a x + a j b ] ( a x + a j b [ a x + a j b ] 1 2 ) .
Proof. 
Since E 1 ( x ) = x 1 2 , by taking n = 1 in Corollary 6, in view of (16) and (47), we obtain the desired result. □
Obviously, the case b = 1 in Corollary 7 gives for a N , x R with 2 a that
a j = 0 a 1 ( 1 ) j + [ x + j a ] ( x + j a [ x + j a ] 1 2 ) = ( 1 ) [ a x ] ( a x [ a x ] 1 2 ) ,
which can be thought of as the complement of Hermite’s identity (33). It is interesting to point out that Corollary 7 can be used to establish a symmetric identity for the sums (52) considered by Cetin et al. [29], as follows.
Corollary 8.
Let a , b N with a b (mod 2) and ( a , b ) = 1 . Then
a C 1 ( b , a ) = b C 1 ( a , b ) + a b 2 ,
where C 1 ( b , a ) is defined for a , b N by
C 1 ( b , a ) = j = 0 a 1 ( 1 ) j + [ b j a ] B ¯ 1 ( b j a ) .
Proof. 
Taking x = 0 in Corollary 7, it then follows from (16) and (32) that
a j = 0 a 1 ( 1 ) j + [ b j a ] ( B ¯ 1 ( b j a ) 1 2 δ Z ( b j a ) ) = b j = 0 b 1 ( 1 ) j + [ a j b ] ( B ¯ 1 ( a j b ) 1 2 δ Z ( a j b ) ) .
Since a and b are relatively prime, we know from ([30] Theorem 3.8) that two lattice points ( a , b ) and ( 0 , 0 ) (two lattice points ( b , a ) and ( 0 , 0 ) ) are mutually visible. Hence, from (53) we have
a C 1 ( b , a ) a 2 = b C 1 ( a , b ) b 2 ,
as desired. This concludes the proof of Corollary 8. □
We here mention that Cetin [36] has shown that the sum in (52) is closely related to the Hardy sums. For some reciprocity formulas of the Hardy sums, one is referred to [37,38,39,40,41].

4. Conclusions

In this paper, we have used the Fourier expansions of the Apostol–Bernoulli polynomials and the Apostol–Euler polynomials to establish some symmetric identities for the Apostol–Bernoulli functions and the Apostol–Euler functions. The results presented here are the corresponding generalizations of Raabe’s [8], Bayad’s [27], and Hermite’s [28] results. We also point out that a relation for the number of lattice points in the case of triangles, a symmetric identity for the sums considered by Cetin et al. [29], can be easily deduced. As shown in the third section, the topics explored in this paper are closely related to the generalized Dedekind sums and the Hardy sums. We will study some properties for the products of the Apostol–Bernoulli functions and the Apostol–Euler functions, and give some new reciprocity formulas for the generalized Dedekind sums and the Hardy sums in another papers.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

This work was partly completed when the author participated in the 16th International Conference on Number Theory and Related Problems held at Xi’an Polytechnic University during 19–23 May 2023. The author is very grateful for the organizer’s kind invitation. The author also expresses his sincere gratitude to the anonymous reviewers for their helpful comments and suggestions for improving this paper.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Luo, Q.-M.; Srivastava, H.M. Some generalizations of the Apostol-Bernoulli and Apostol-Euler polynomials. J. Math. Anal. Appl. 2005, 308, 290–302. [Google Scholar] [CrossRef]
  2. Luo, Q.-M.; Srivastava, H.M. Some relationships between the Apostol-Bernoulli and Apostol-Euler polynomials. Comput. Math. Appl. 2006, 51, 631–642. [Google Scholar] [CrossRef] [Green Version]
  3. Apostol, T.M. On the Lerch zeta function. Pac. J. Math. 1951, 1, 161–167. [Google Scholar] [CrossRef] [Green Version]
  4. Srivastava, H.M. Some formulas for the Bernoulli and Euler polynomials at rational arguments. Math. Proc. Camb. Philos. Soc. 2000, 129, 77–84. [Google Scholar] [CrossRef]
  5. Nielsen, N. Traité Élémentaire des Nombres de Bernoulli; Gauthier-Villars: Paris, France, 1923. [Google Scholar]
  6. Nörlund, N.E. Vorlesungen über Differenzenrechnung; Springer: Berlin/Heidelberg, Germany, 1924. [Google Scholar]
  7. Belbachir, H.; Djemmada, Y.; Hadj-Brahim, S. Unified Bernoulli-Euler polynomials of Apostol type. Indian J. Pure Appl. Math. 2023, 54, 76–83. [Google Scholar] [CrossRef]
  8. Raabe, J.L. Zurückführung einiger Summen und bestimmten Integrale auf die Jacob-Bernoullische Function. J. Reine Angew. Math. 1851, 42, 348–367. [Google Scholar]
  9. Howard, F.T. Applications of a recurrence for the Bernoulli numbers. J. Number Theory 1995, 52, 157–172. [Google Scholar] [CrossRef] [Green Version]
  10. Tuenter, H.J.H. A symmetry of power sum polynomials and Bernoulli numbers. Am. Math. Monthly 2001, 108, 258–261. [Google Scholar] [CrossRef]
  11. Yang, S.-L. An identity of symmetry for the Bernoulli polynomials. Discrete Math. 2008, 308, 550–554. [Google Scholar] [CrossRef] [Green Version]
  12. Kim, T. Symmetry p-adic invariant integral on Zp for Bernoulli and Euler polynomials. J. Differ. Equ. Appl. 2008, 14, 1267–1277. [Google Scholar] [CrossRef]
  13. Liu, H.; Wang, W. Some identities on the Bernoulli, Euler and Genocchi polynomials via power sums and alternate power sums. Discrete Math. 2009, 309, 3346–3363. [Google Scholar] [CrossRef] [Green Version]
  14. Wang, W.; Wang, W. Some results on power sums and Apostol-type polynomials. Integral Transform. Spec. Funct. 2010, 21, 307–318. [Google Scholar] [CrossRef]
  15. Zhang, Z.; Yang, H. Several identities for the generalized Apostol-Bernoulli polynomials. Comput. Math. Appl. 2008, 56, 2993–2999. [Google Scholar] [CrossRef] [Green Version]
  16. He, Y. Symmetric identities for Carlitz’s q-Bernoulli numbers and polynomials. Adv. Differ. Equ. 2013, 2013, 246. [Google Scholar] [CrossRef] [Green Version]
  17. He, Y.; Zhang, W. A note on the twisted Lerch type Euler zeta functions. Bull. Korean Math. Soc. 2013, 50, 659–665. [Google Scholar] [CrossRef] [Green Version]
  18. Luo, Q.-M. Multiplication formulas for Apostol-type polynomials and multiple alternating sums. Math. Notes 2012, 91, 46–57. [Google Scholar] [CrossRef]
  19. Kurt, V. Some symmetry identities for the unified Apostol-type polynomials and multiple power sums. Filomat 2016, 30, 583–592. [Google Scholar] [CrossRef] [Green Version]
  20. Araci, S.; Duran, U.; Acikgoz, M. A study on a class of q-Euler polynomials under the symmetric group of degree n. J. Nonlinear Sci. Appl. 2016, 9, 5196–5201. [Google Scholar] [CrossRef] [Green Version]
  21. Srivastava, H.M.; Kurt, B.; Kurt, V. Identities and relations involving the modified degenerate hermite-based Apostol-Bernoulli and Apostol-Euler polynomials. Rev. Real Acad. Cienc. Exactas Fis. Nat. Ser. A Mat. 2019, 113, 1299–1313. [Google Scholar] [CrossRef]
  22. Kim, T.; Kim, D.S.; Kim, H.Y.; Kwon, J. Identities of symmetry for Bernoulli polynomials and power sums. J. Inequal. Appl. 2020, 2020, 245. [Google Scholar] [CrossRef]
  23. Khan, N.; Usman, T.; Choi, J. A new class of generalized polynomials involving Laguerre and Euler polynomials. Hacet. J. Math. Stat. 2021, 50, 1–13. [Google Scholar] [CrossRef]
  24. Usman, T.; Khan, N.; Aman, M.; Choi, J. A family of generalized Legendre-based Apostol-type polynomials. Axioms 2022, 11, 29. [Google Scholar] [CrossRef]
  25. Luo, Q.-M. Fourier expansions and integral representations for the Apostol-Bernoulli and Apostol-Euler polynomials. Math. Comp. 2009, 78, 2193–2208. [Google Scholar] [CrossRef] [Green Version]
  26. Bayad, A. Fourier expansions for Apostol-Bernoulli, Apostol-Euler and Apostol-Genocchi polynomials. Math. Comp. 2011, 80, 2219–2221. [Google Scholar] [CrossRef] [Green Version]
  27. Bayad, A. Arithmetical properties of elliptic Bernoulli and Euler numbers. Int. J. Algebra 2010, 4, 353–372. [Google Scholar]
  28. Hermite, C. Sur quelques conséquences arithmétiques des Formules de la théorie des fonctions elliptiques. Acta Math. 1884, 5, 297–330. [Google Scholar] [CrossRef]
  29. Cetin, E.; Simsek, Y.; Cangul, I.N. Some special finite sums related to the three-term polynomial relations and their applications. Adv. Differ. Equ. 2014, 2014, 283. [Google Scholar] [CrossRef] [Green Version]
  30. Apostol, T.M. Introduction to Analytic Number Theory; Springer: New York, NY, USA, 1976. [Google Scholar]
  31. Carlitz, L. Generalized Dedekind sums. Math. Z. 1964, 85, 83–90. [Google Scholar] [CrossRef]
  32. Bayad, A.; Raouj, A. Reciprocity formulae for multiple Dedekind-Rademacher sums. C. R. Math. Acad. Sci. Paris 2011, 349, 131–136. [Google Scholar] [CrossRef]
  33. Aursukaree, S.; Khemaratchatakumthorn, T.; Pongsriiam, P. Generalizations of Hermite’s identity and applications. Fibonacci Quart. 2019, 57, 126–133. [Google Scholar]
  34. Kim, M.-S.; Son, J.-W. On generalized Dedekind sums involving quasi-periodic Euler functions. J. Number Theory 2014, 144, 267–280. [Google Scholar] [CrossRef]
  35. Hu, S.; Kim, D.; Kim, M.-S. On reciprocity formula of Apostol-Dedekind sum with quasi-periodic Euler functions. J. Number Theory 2016, 162, 54–67. [Google Scholar] [CrossRef] [Green Version]
  36. Cetin, E. A note on Hardy type sums and Dedekind sums. Filomat 2016, 30, 977–983. [Google Scholar] [CrossRef]
  37. Berndt, B.C. Analytic Eisenstein series, theta-functions, and series relations in the spirit of Ramanujan. J. Reine Angew. Math. 1978, 303/304, 332–365. [Google Scholar]
  38. Simsek, Y. Relations between theta-functions Hardy sums Eisenstein and Lambert series in the transformation formula of logηg,h(z). J. Number Theory 2003, 99, 338–360. [Google Scholar] [CrossRef] [Green Version]
  39. Liu, H.; Zhang, W. Generalized Cochrane sums and Cochrane-Hardy sums. J. Number Theory 2007, 122, 415–428. [Google Scholar] [CrossRef] [Green Version]
  40. Liu, H.; Gao, J. Generalized Knopp identities for homogeneous Hardy sums and Cochrane-Hardy sums. Czech. Math. J. 2012, 62, 1147–1159. [Google Scholar] [CrossRef]
  41. Can, M. Reciprocity formulas for Hall-Wilson-Zagier type Hardy-Berndt sums. Acta Math. Hungar. 2021, 163, 118–139. [Google Scholar] [CrossRef]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

He, Y. Applications of Symmetric Identities for Apostol–Bernoulli and Apostol–Euler Functions. Symmetry 2023, 15, 1384. https://doi.org/10.3390/sym15071384

AMA Style

He Y. Applications of Symmetric Identities for Apostol–Bernoulli and Apostol–Euler Functions. Symmetry. 2023; 15(7):1384. https://doi.org/10.3390/sym15071384

Chicago/Turabian Style

He, Yuan. 2023. "Applications of Symmetric Identities for Apostol–Bernoulli and Apostol–Euler Functions" Symmetry 15, no. 7: 1384. https://doi.org/10.3390/sym15071384

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