Next Article in Journal
Dynamic Pricing and Inventory Strategies for Fashion Products Using Stochastic Fashion Level Function
Previous Article in Journal
Estimates of Eigenvalues and Approximation Numbers for a Class of Degenerate Third-Order Partial Differential Operators
Previous Article in Special Issue
A Unified Representation of q- and h-Integrals and Consequences in Inequalities
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Concerning Transformations of Bases Associated with Unimodular diag(1, −1, −1)-Matrices

1
Department of Mathematical Analysis, Moscow Technical University of Communications and Informatics, Aviamotornaya 8a, Moscow 111024, Russia
2
Department of Mathematics, Dongguk University, Gyeongju 38066, Republic of Korea
*
Author to whom correspondence should be addressed.
Axioms 2024, 13(7), 452; https://doi.org/10.3390/axioms13070452
Submission received: 29 April 2024 / Revised: 10 June 2024 / Accepted: 1 July 2024 / Published: 4 July 2024
(This article belongs to the Special Issue Theory of Functions and Applications II)

Abstract

:
Considering a representation space for a group of unimodular diag ( 1 ,   1 ,   1 ) -matrices, we construct several bases whose elements are eigenfunctions of Casimir infinitesimal operators related to a reduction in the group to some one-parameter subgroups. Finding the kernels of base transformation integral operators in terms of special functions, we consider the compositions of some of these transformations. Since composition is a ‘closed’ operation on the set of base transformations, we obtain some integral relations for the special functions involved in the above kernels.

1. Introduction

Consider two distinct orthonormal bases on the Cartesian plane, denoted by { O A 1 , O B 1 } and { O A 2 , O B 2 } , where O represents the origin. The composition of the basis transformations { O A 1 , O B 1 } { O A 2 , O B 2 } and { O A = ( 1 , 0 ) , O B = ( 0 , 1 ) } { O A 1 , O B 1 } can be expressed as the multiplication of two 2 × 2 -orthogonal matrices. This leads to the addition theorem for cosine and sine, as the matrix representing this composition concerning the basis { O A , O B } is itself an orthogonal matrix. The kernels of the basis transformation integral operators in infinite-dimensional functional linear spaces involve intricate functions known as special functions in mathematical physics. Since these special functions are eigenfunctions of differential operators which are invariant under the associated Lie groups, we have a direct connection between these special functions, which constitute the kernels, and the representation of the corresponding Lie group.
In this manuscript, we establish several bases within a functional linear space in Section 3, analyze the kernels of basis transformation operators in Section 4, and derive integral relationships pertaining to specific instances of confluent hypergeometric functions F 1 1 in Section 5. These functions include Bessel J ν ( x ) , Hankel H μ ( 1 ) ( x ) , and H μ ( 2 ) ( x ) , Whittaker W μ , ν ( x ) , Macdonald K ν ( x ) , and Coulomb F μ ( ρ ; x ) functions. In [1,2], we used another approach investigating the kernels of the restriction of representation integral operators to certain one-parameter groups expressed in various ‘direct’ or ‘mixed’ bases, thereby unveiling additional integral relationships. Let us remember that a group-theoretical approach to classical Bessel functions (in a wide sense) had been considered in monographs [3,4] and to some their multi-variable or multi-index analogues and generalizations have been presented, for example, in [5,6,7].

2. The Group G 0 , Algebra g , and Space L

Let a be an arbitrary square matrix. Let us call a matrix b of the same size an a-matrix if b T a b = a , where T represents the transpose of the matrix. For the case det a 0 , we immediately obtain | det b | = 1 . By selecting different matrices denoted as a, we can derive diverse sets of well-known matrix classes: orthogonal ( a = diag ( 1 , , 1 ) ), symplectic, etc. The equalities
( b b ^ ) T a ( b b ^ ) = b ^ T ( b T a b ) b ^ = b ^ T a b ^ = a , ( b 1 ) T a b 1 = ( b 1 ) T ( b T a b ) b 1 = ( b b 1 ) T a ( b b 1 ) = a
indicate that the set of a-matrices form a subgroup O ( 1 , 2 ) within the general linear group of order n, where n × n represents the dimensions of matrix a. Let the symbol G represent the intersection of O ( 1 , 2 ) and the special linear subgroup S L ( 3 , R ) . From the given definition, it can be deduced that for any b G , the following equations hold:
b 1 i b 1 j b 2 i b 2 j b 3 i b 3 j = sign 3 2 j δ i j ,
b i 1 b j 1 b i 2 b j 2 b i 3 b j 3 = sign 3 2 j δ i j .
Here, δ i j is Kronecker delta: 0 if i j ; 1 if i = j .
It is demonstrable [8] that the expression for matrix b is given by:
b = b 11 A B C ,
where A = C 1 C 2 , matrix C 1 is orthogonal and C 2 ia a positive–definite matrix. The matrix b relies on b 11 and independent parameters of C 1 and A (or B), therefore being defined by three parameters. From Equation (3) (see [8]), it is deduced that the determinant of C equals b 11 . This implies that the map** ι : G U 2 = { 1 , 1 } defined by the formula
ι ( b ) = 1 if b 11 > 0 , 1 if b 11 < 0 ,
is a group epimorphism. Both cosets of the normal divisor G 0 = Ker ι represent connected components in G.
Let h 1 ( φ ) be the matrix of the circle rotation in the plane O x 2 x 3 through angle φ :
h 1 ( φ ) = 1 0 0 0 cos φ sin φ 0 sin φ cos φ .
Let h 2 ( φ ) and h 3 ( φ ) denote matrices of hyperbolic rotations in the planes O x 1 x 2 and O x 1 x 3 , respectively, that is
h 2 ( φ ) = cosh φ sinh φ 0 sinh φ cosh φ 0 0 0 1 , h 3 ( φ ) = cosh φ 0 sinh φ 0 1 0 sinh φ 0 cosh φ .
Obviously, h 1 ,   h 2 ,   h 3 G 0 . The vectors e 2 , 3 = d h 1 d φ | φ = 0 , e 2 = d h 2 d φ | φ = 0 , and e 3 = d h 3 d φ | φ = 0 constitute a basis E for the tangent space of the group G 0 , evaluated at the point id . The commuting relations of the corresponding Lie algebra g can be expressed as follows:
[ e 2 , 3 , e 2 ] = e 3 , [ e 2 , 3 , e 3 ] = e 2 , [ e 2 , e 3 ] = e 2 , 3 .
The matrices h i (for each i) constitute a subgroup H i within G 0 . It is evident that the group G acts transitively on both the cone X 0 : x 1 2 x 2 2 x 3 2 = 0 and the hyperboloid X 1 : x 1 2 x 2 2 x 3 2 = 1 .
Lemma 1.
G 0 = H 1 H 2 H 1 = H 1 H 3 H 1 .
Proof. 
Let St y ˜ G be the stabilizer of the point y ˜ = ( 1 , 0 , 0 ) X 1 . Given that for any g St y ˜ , the equality g 11 = 1 holds, based on (1) and (2), we can express g as:
g = 1 0 0 0 g 22 g 23 0 g 32 g 33 ,
where g 22 g 23 g 32 g 33 is an orthogonal matrix. Thus, St y ˜ = H 1 . Let τ be the bijective map** G / H 1 X 1 , such that τ ( g ˜ H 1 ) = y , where y is the image of y ˜ for any transformation belonging to the coset g ˜ St y ˜ . If g G 0 , g g ˜ St y ˜ and τ ( g ˜ St y ˜ ) = y , then g = g ˜ h 1 . Because g relies on three parameters and H 1 is a one-parameter subgroup, the generator g ˜ of the coset g ˜ St y ˜ ought to be contingent upon two parameters. Let y = ( cosh ξ , sinh ξ cos μ , sinh ξ sin μ ) . Let us show that g ˜ can be written in the form g ˜ = h 1 ( μ ) h 2 ( ξ ) :
h 2 ( ξ ) y ˜ = ( cosh ξ , sinh ξ , 0 ) , h 1 ( μ ) ( cosh ξ , sinh ξ , 0 ) = y .
Therefore, g = g ˜ h 1 = h 1 ( μ ) h 2 ( ξ ) h 1 ( ν ) . The second equality of the present lemma can be demonstrated using the same method. □
We denote by X 0 + the subset of the cone X 0 that comprises points x, where x 1 > 0 .
Lemma 2.
The semicone X 0 + is invariant under the transformations of the group G 0 .
Proof. 
Based on Lemma 1, it is enough to confirm for h 1 and h 2 . For x X 0 + we have x 1 > 0 and x 1 2 = x 2 2 + x 3 2 , therefore, x 1 > | x 2 | . Since
h 1 ( φ ) x = ( x 1 , x 2 cos φ x 3 sin φ , x 2 sin φ + x 3 cos φ ) ,
we find that h 1 ( φ ) x X 0 + . Given that
h 1 ( φ ) x = ( x 1 cosh φ + x 2 sinh φ , x 1 sinh φ + x 2 cosh φ , x 3 ) ,
where cosh φ > | sinh φ | , we can derive that x 1 cosh φ + x 2 sinh φ > 0 . This implies that g 2 ( φ ) x X 0 + . □
Lemma 3.
The group G 0 acts transitively on the semicone X 0 + .
Proof. 
We denote by γ 1 the circle x 1 = 1 belonging to X 0 + . Introducing polar coordinates on γ 1 , we write any point x X 0 * in the form
x = ( r , r cos α , r sin α ) ,
where r > 0 . Since
h 1 3 π 2 + α h 3 ( ln r ) ( 1 , 0 , 1 ) = h 1 3 π 2 + α ( r , 0 , r ) = x ,
and
h 1 3 π 2 + α h 3 ( ln r ) = cosh ln r ,
in view of cosh ln r > 0 the transformation g = h 1 3 π 2 + α h 3 ( ln r ) belongs to G 0 .
Let x ˜ X 0 + . It has been demonstrated that there exists a transformation denoted as g ˜ G 0 , such that when applied, it satisfies the condition g ˜ ( 1 , 0 , 0 ) = x ˜ . We thus have the equality
( 1 , 0 , 0 ) = g ˜ 1 x ˜ .
By substituting (7) into (6), the resultant equation becomes g g ˜ 1 x ˜ = x . □
Let f be a function defined on the semicone X 0 + . Let us call this function infinitely differentiable, if the derivative k 1 + k 2 + k 3 f x 1 k 1 x 2 k 2 x 3 k 3 exists at any point of X 0 + and for any nonnegative integers k 1 , k 2 , k 3 such that ( k 1 , k 2 , k 3 ) ( 0 , 0 , 0 ) . Let σ C . We call a function f σ -homogeneous if f ( λ x ) = λ σ f ( x ) . Given x 1 > 0 , this implies that λ > 0 .
Let L represent the linear space comprising σ -homogeneous infinitely differentiable functions on X 0 + . It can be readily verified that the map** T ( g ) : L L , defined as f f ( g x ) , constitutes an automorphism of L . Moreover, the function G 0 Aut L , where g T ( g ) , forms a representation of G 0 .

3. Construction of Bases

For each vector belonging to the above basis E of tangent algebra g , we define the corresponding infinitesimal operator:
d 2 , 3 = i lim t 0 T exp ( t e 2 , 3 ) f ( x ) f ( x ) t , d i = i lim t 0 T exp ( t e i ) f ( x ) f ( x ) t .
In this context and throughout, the symbol i denotes the purely imaginary number, which is the square root of 1 . It is easy to find that
d 2 , 3 = i x 2 x 3 x 3 x 2 , d i = i x 1 x i + x i x 1 .
In polar coordinates on γ 1 we have d 2 , 3 = i d d α . Let f ( α ) be an eigenfunction of the operator d 2 , 3 with respect to the eigenfunction λ and, in addition, a restriction of f L to γ 1 . From the equation i d f d α = λ f we have f Span e i λ α . From the condition that f ( π ) = f ( π ) , we can deduce e 2 i λ α = 1 , implying that λ Z . By introducing n = λ , we derive the basis within the space of function restrictions from L to γ 1 , comprising a set of functions e i n α . Considering the σ -homogeneity property for any x X 0 + :
f ( x ) = x 1 σ f 1 , x 2 x 1 , x 3 x 1 ,
therefore, the functions
f n ( 1 ) ( x ) = x 1 σ e i n α = x 1 σ x 2 x 1 + i x 3 x 1 n = x 1 σ n ( x 2 + i x 3 ) n ,
form the basis B 1 in the space L .
We denote the hyperbola x 3 = ± 1 on X 0 + by symbol γ 2 , ± . Let γ 2 = γ 2 , + γ 2 , . In hyperbolic coordinates
x 1 = cosh α , x 2 = sinh α , x 3 = ± 1
on γ 2 , where α R , we have d 2 = i d d α . Let f be an eigenfunction of d 2 with respect to value λ and a restriction of f L to γ 2 . From the equality i d f d α = λ f we have f Span e i λ α , therefore, e i λ α , λ R , form a basis in the space of function restrictions from L to γ 2 .
Let us consider the value of f at the point x X 0 + such that x 3 0 . In case of x 3 > 0 we have
f ( x ) = x 3 σ f , + x 1 x 3 , x 2 x 3 , 1 ,
where f , + is the restriction to γ 2 , + . In case of x 3 < 0 we can write
f ( x ) = | x 3 | σ f , x 1 | x 3 | , x 2 | x 3 | , 1 .
It implies that
f ( x ) = | x 3 | σ δ 1 , sign x 3 · f , + x 1 | x 2 | , x 2 | x 3 | , 1 + | x 3 | σ δ 1 , sign x 3 f , x 1 | x 3 | , x 2 | x 3 | , 1 .
Using the generalized functions [9],
( s ) ± ν = | s | ν if ± s 0 , 0 if ± s < 0 ,
we obtain
f ( x ) = ( x 3 ) ± σ f , ± x 1 | x 3 | , x 2 | x 3 | , ± 1
(double signs ± are consistently employed in same order here and in other instances). This implies that functions
f λ , ± ( 2 ) ( x ) = ( x 3 ) ± σ e i λ α = ( x 3 ) ± σ ( cosh α + sinh α ) i λ = ( x 3 ) ± σ x 1 | x 3 | + x 2 | x 3 | i λ = ( x 3 ) ± σ i λ ( x 1 + x 2 ) i λ
form a basis B 2 in the space L .
By analogy, defining hyperbolas γ 3 , ± : x 2 = ± 1 , we obtain a basis B 3 in L , consisting of functions
f λ , ± ( 3 ) ( x ) = ( x 2 ) ± σ i λ ( x 1 + x 3 ) i λ , λ R ,
related to the contour γ 3 = γ 3 , + γ 3 , on the semicone X 0 + .
Let us define the linear subspaces k = Span ( e 2 , 3 ) and p = Span ( e 2 , e 3 ) in g . As per (8), it follows that so ( 1 , 2 ) = k p , where
[ k , k ] k , [ p , p ] k , [ k , p ] p .
In view of relations (8), the dimension of maximal commutative subalgebra a in g is equal to 1. Letting a = Span ( e 2 ) , we get the following matrix of the adjoint operator ad e 2 in the above basis E:
0 1 0 1 0 0 0 0 0 .
Finding the eigenvalues of the characteristic polynomial of this matrix, we obtain
so ( 1 , 2 ) = Ker ad e 2 + V 1 + V 1 ,
where the root linear subspace V j consists of a zero vector and all eigenvectors of the operator ad e 2 related to the value j. This implies that the maximal nilpotent subalgebra n in g can be defined by the equality n = V 1 , that is, n = Span ( e 2 , 3 + e 3 ) .
Let us define the subgroup H 4 = exp n = { exp t ( e 2 , 3 + e 3 ) } = { h 4 ( t ) } in G 0 . It consists of matrices
h 4 ( t ) = diag ( 1 ,   1 ,   1 ) + t 0 0 1 0 0 1 1 1 0 + t 2 2 ! 1 1 0 1 1 0 0 0 0 + t 3 3 ! diag ( 0 ,   0 ,   0 ) = 1 2 2 + t 2 t 2 2 t t 2 2 t 2 2 t 2 t 2 t 2
and acts transitively on the intersection of the semicone X 0 + and the plane x 1 + x 2 = 1 . We denote this parabola by γ 4 .
Letting a = Span ( e 3 ) , by analogy, we obtain n = Span ( e 2 , 3 + e 2 ) and find its exponential image H 5 consisting of matrices
h 5 ( t ) = diag ( 1 ,   1 ,   1 ) + t 0 1 0 1 0 1 0 1 0 + t 2 2 ! 1 0 1 0 0 0 1 0 1 + t 3 3 ! diag ( 0 ,   0 ,   0 ) = 1 2 2 + t 2 2 t t 2 2 t 2 2 t t 2 2 t 2 t 2 .
We denote by γ 5 the intersection of the semicone X 0 + and plane x 1 + x 3 = 1 . This parabola is a homogeneous space of the subgroup H 5 .
The infinitesimal operator c = d 2 , 3 + d 3 associated with the generator e 2 , 3 + e 3 within a one-dimensional subalgebra n exhibits commutativity with infinitesimal operators associated with all vectors within n . Thus, it qualifies as a Casimir operator linked to the reduction H 4 G 0 .
In horospherical coordinates
x 1 = 1 + α 2 2 , x 2 = 1 α 2 2 , x 3 = α
on the parabola γ 4 , where α R , we have c = i d d α .
Let us denote by f an eigenfunction of the operator c with the eigenvalue λ . Suppose that f is the restriction of function f L to parabola γ 4 . From the equation i d f d α = λ f we obtain f Span e i λ α , where λ R . It gives the basis in the space of function restrictions from L to γ 4 , which consists of functions e i λ α .
For any point x X 0 + where x 1 x 2 , considering σ -homogeneity and utilizing Formula (5), we derive x 1 + x 2 = r ( 1 + cos α ) 0 . Consequently,
f ( x ) = ( x 1 + x 2 ) σ f x 1 x 1 + x 2 , x 2 x 1 + x 2 , x 3 x 1 + x 2 .
Therefore, the functions
f λ ( 4 ) ( x ) = ( x 1 + x 2 ) σ e i λ α = ( x 1 + x 2 ) σ exp i λ x 3 x 1 + x 2 ,
where λ R , form a basis B 4 in L .
By analogy, finding eigenfunctions of the operator d 2 , 3 + d 2 , which corresponds to the reduction H 5 G 0 , we obtain a basis B 5 in L , consisting of functions
f λ ( 5 ) ( x ) = ( x 1 + x 3 ) σ e i λ α = ( x 1 + x 3 ) σ exp i λ x 2 x 1 + x 3 ,
where λ R .
Upon substituting σ with σ 1 , the resultant is the linear space denoted as L * . Each f L has its counterpart in L * as f * . The bases B 1 B 5 are assumed to represent the equivalents of the bases B 1 * B 5 * .

4. Composition of the Transformations B 2 B 5 and B 5 B 4

Considering the composition of the integral operators
f μ ( 4 ) = + ε μ , ρ 5 4 ( σ ) f ρ ( 5 ) d ρ
and
f ρ ( 5 ) = + ( ε ρ , λ , + 2 5 ( σ ) f λ , + ( 2 ) + ε ρ , λ , 2 5 ( σ ) f λ , ( 2 ) ) d λ ,
we get
f μ ( 4 ) = + [ + ε μ , ρ 5 4 ( σ ) ε ρ , λ , + 2 5 ( σ ) d ρ f λ , + ( 2 ) + + ε μ , ρ 5 4 ( σ ) ε ρ , λ , 2 5 ( σ ) d ρ f λ , ( 2 ) ] d λ .
Comparing
f μ ( 4 ) = + ( ε μ , λ , + 2 4 ( σ ) f λ , + ( 2 ) + ε μ , λ , 2 4 f λ , ( 2 ) ) d λ
and (11), we derive
ε μ , λ , ± 2 4 ( σ ) = + ε μ , ρ 5 4 ( σ ) ε ρ , λ , ± 2 5 ( σ ) d ρ .
In view of equality f λ , + * ( 2 ) ( cosh α , sinh α , 1 ) = f λ , * ( 2 ) ( cosh α , sinh α , 1 ) = 0 , we derive from (12) that
+ f ν , + * ( 2 ) ( cosh α , sinh α , 1 ) f μ ( 4 ) ( cosh α , sinh α , 1 ) d α = + ε μ , λ , + 2 4 ( σ ) d λ + e i ( λ + ν ) α d α = 2 π + ε μ , λ , + 2 4 ( σ ) δ ( λ + ν ) d λ = 2 π ε μ , ν , + 2 4 ( σ ) ,
where δ is the Dirac function. Therefore,
ε μ , λ , + 2 4 ( σ ) = ( 2 π ) 1 + f λ , + * ( 2 ) ( cosh α , sinh α , 1 ) f μ ( 4 ) ( cosh α , sinh α , 1 ) d α .
Using the substitution cosh α = 1   +   ( t     1 ) 2 2 , we have
ε μ , λ , + 2 4 ( σ ) = 2 σ π 1 e i μ 0 2 t i λ σ 1 ( 2 t ) i λ σ 1 e i μ t d t .
Theorem 1.
Let ( σ ) < 0 . Then,
ε μ , λ , + 2 4 ( σ ) = 2 σ 1 π 1 e i μ B ( i λ σ , i λ σ ) F 1 1 ( i λ σ ; 2 σ ; 2 i μ ) ,
where B is the Beta function.
Proof. 
Utilizing the equation [10] (Entry 2.3.6.1) on (14) leads us to derive the intended expression, resulting in Formula (15). □
Likewise, we obtain the subsequent equalities through a similar derivation process:
ε μ , λ , + 2 5 ( σ ) = ( 2 π ) 1 [ + f λ , + * ( 2 ) ( cosh α , sinh α , 1 ) f μ ( 5 ) ( cosh α , sinh α , 1 ) d α + + f λ , + * ( 2 ) ( cosh α , sinh α , 1 ) f μ ( 5 ) ( cosh α , sinh α , 1 ) d α ]
and
ε μ , λ 4 5 ( σ ) = ( 2 π ) 1 + f λ * ( 4 ) 1 + α 2 2 , 1 α 2 2 , α f μ ( 5 ) 1 + α 2 2 , 1 α 2 2 , α d α .
Theorem 2.
Let 1 < ( σ ) < 0 . Then,
ε μ , λ , + 2 5 ( σ ) = 2 σ π 1 ( i μ ) σ i λ Γ ( i λ σ ) ,
where Γ is Gamma function.
Proof. 
By substituting a new variable t such that cosh α = 1   +   t 2 2 , we obtain
ε μ , λ , + 2 5 ( σ ) = 2 σ π 1 + ( t ) + i λ σ 1 e i μ t d t = 2 σ π 1 0 + t i λ σ 1 e i μ t d t .
Then, use of the Laplace transform to the last integral gives the desired result (16). □
Theorem 3.
Let 1 < ( σ ) < 0 . For sign λ = sign μ 0 , we have
ε μ , λ 4 5 ( σ ) = 2 2 π e i ( λ μ ) μ λ σ + 1 / 2 sin σ π K 2 σ + 1 2 2 μ λ .
For sign λ = sign μ 0 , we have
ε μ , λ 4 5 ( σ ) = e i ( λ μ ) 2 π μ λ σ + 1 / 2 cos σ π J 2 σ 1 2 2 μ λ J 2 σ + 1 2 2 μ λ .
Proof. 
Introducing a new variable t = α + 1 , we obtain
ε μ , λ 4 5 ( σ ) = 2 σ π 1 e i ( λ μ ) 0 + t 2 σ cos λ t 2 μ t d t .
Utilizing the formulas specified as [10] (Entries 2.5.24.4, 2.5.24.7), and applying them to the integral presented in (19), results in the derivation of both (17) and (18). □
Theorem 4.
Let κ > 0 and 1 < ( σ ) < 0 . Then,
4 ( i ) σ i λ sin ( π σ ) 0 + t i λ e i t 2 / 4 K 2 σ + 1 ( κ t ) d t i σ i λ cos ( π σ ) 0 + t i λ e i t 2 / 4 J 2 σ + 1 ( κ t ) J 2 σ 1 ( κ t ) d t = π Γ ( i λ σ ) 2 σ κ 2 σ + 1 e i κ 2 Γ ( 2 σ ) F 1 1 ( i λ ; 2 σ ; i κ 2 ) .
Proof. 
Let μ > 0 . Taking the relationship ε μ , ρ 5 4 ( σ ) = ε ρ , μ 4 5 ( σ 1 ) , we obtain from (13) that
ε μ , λ , + 2 4 ( σ ) = M + ( λ ) + M ( λ ) ,
where
M ± ( λ ) = ± 0 ± ε ρ , λ , + 2 5 ( σ ) ε ρ , μ 4 5 ( σ 1 ) d ρ .
Choosing new variables t = 2 ± ρ in M ± ( λ ) and supposing that κ = 2 μ , we complete the proof. □
Remark 1.
The result in Theorem 4 can be rewritten in terms of the Coulomb wave function F μ ( ρ ; x ) x μ + 1 e i x F 1 1 ( 1 + μ i ρ ; 2 μ + 2 ; 2 i x ) .
The particular case λ = 0 in (4) can be expressed in terms of Bessel and Hankel functions. Indeed, using [11] (Entry 2.6.15.2), we have
M + ( 0 ) = 2 π ( i ) σ κ 2 σ + 1 e i ( σ + 1 ) / 2 Γ ( σ ) J σ + 1 / 2 ( μ ) + i Y σ + 1 / 2 ( μ ) = 2 π ( i ) σ κ 2 σ + 1 e i ( σ + 1 ) / 2 Γ ( σ ) H σ + 1 / 2 ( 1 ) ( μ ) ,
where Y σ + 1 / 2 is the Bessel function of the second kind. Also, employing [11] (Entry 2.12.18.2), we obtain
M ( 0 ) = 2 1 / 2 π 2 i σ κ 2 σ + 1 cos ( σ π ) Γ ( σ ) e i σ π / 2 J σ 1 / 2 ( μ ) i e i σ π / 2 J σ + 1 / 2 ( μ ) = 2 1 / 2 π 2 i σ + 1 κ 2 σ + 1 cos 2 ( σ π ) e i σ π / 2 Γ ( σ ) H σ + 1 / 2 ( 2 ) ( μ ) .
Moreover, the integral in (14) with λ = 0 can be evaluated using the formula [10] (Entry 2.3.6.2) as follows:
0 a t ν 1 ( a t ) ν 1 e p t d t = π Γ ( ν ) a p ν 1 / 2 e a p / 2 I ν 1 / 2 a p 2 ,
which is valid for ( ν ) > 0 . Note J ν ( i z ) = e i ν π / 2 I ν ( z ) .

5. Composition of the Transformations B 1 B 4 and B 4 B 5

Considering the operator
f λ ( 5 ) = n = ε λ , n 1 5 ( σ ) f n ( 1 )
as a composition of the operators
f λ ( 5 ) = + ε λ , ρ 4 5 ( σ ) f ρ ( 4 ) d ρ
and
f ρ ( 4 ) = n = ε ρ , n 1 4 ( σ ) f n ( 1 ) ,
we obtain the equality
ε λ , n 1 5 ( σ ) = + ε λ , ρ 4 5 ( σ ) ε ρ , n 1 4 ( σ ) d ρ .
Theorem 5.
Let k N , κ > 0 and 0 < ( ν ) < 1 2 . Then,
W k , ν κ 2 / 2 = π i k Γ ( ν + k + 1 2 ) 1 × [ sin ( ν π ) 2 1 / 2 Γ ( ν k 3 2 ) 0 + t e i t 2 / 2 W k , ν t 2 K 2 ν ( κ t ) d t + 2 3 / 2 cos ( ν π ) Γ ( ν k + 1 2 ) 0 + t e i t 2 / 2 W k , ν t 2 J 2 ν ( κ t ) J 2 ν ( κ t ) d t ] .
In particular,
K ν κ 2 / 4 = π Γ ( ν + 1 2 ) 1 × [ sin ( ν π ) 2 1 / 2 Γ ( ν 3 2 ) 0 + t 3 / 2 e i t 2 / 2 K ν t 2 / 2 K 2 ν ( κ t ) d t + 2 3 / 2 cos ( ν π ) Γ ( ν + 1 2 ) 0 + t 3 / 2 e i t 2 / 2 K ν t 2 / 2 J 2 ν ( κ t ) J 2 ν ( κ t ) d t ] .
Proof. 
Since
π π f k * ( 1 ) ( 1 , cos α , sin α ) f λ ( 4 ) ( 1 , cos α , sin α ) d α = n = ε λ , n 1 4 ( σ ) π π e i ( k + n ) α d α ,
we have
ε λ , n 1 4 ( σ ) = ( 2 π ) 1 π π f n * ( 1 ) ( 1 , cos α , sin α ) f λ ( 4 ) ( 1 , cos α , sin α ) d α .
Using the substitution sin α = 2 t 1   +   t 2 , we obtain
ε λ , n 1 4 ( σ ) = 2 σ π 1 + e i λ t ( 1 + i t ) σ n 1 ( 1 i t ) n σ 1 d t .
The value of this integral can be evaluated in terms of the Whittaker function [12] (Entry 3.384.9). Moreover, ε λ , n 1 5 ( σ ) = i n ε λ , n 1 4 ( σ ) . Considering Theorem 3 and Formula (20), introducing new variables t = 2 ρ for ρ > 0 and t = 2 ρ for ρ < 0 and setting k = n , ν = σ + 1 2 , and κ = 2 λ , we get (21).
Setting k = 0 in (21) yields the Formula (22). □

6. Concluding Remarks

We crafted a variety of bases featuring eigenfunctions of Casimir infinitesimal operators, intricately linked to a reduction in the group of unimodular diag ( 1 ,   1 ,   1 ) -matrices to specific subgroups. Through diligent exploration, we uncovered the kernels of base transformation integral operators and delved into the fusion of these transformations. By virtue of composition being a closed operation, we unveiled integral relationships for some select special functions, elegantly encapsulated in Theorems 4 and 5.

Author Contributions

Writing—original draft, I.A.S. and J.C.; Writing—review and editing, I.A.S. and J.C. 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

The original contributions presented in the study are included in the article, further inquiries can be directed to the corresponding author.

Acknowledgments

The authors sincerely thank the anonymous reviewers for their invaluable feedback. Their constructive and encouraging comments significantly improved the quality of this paper.

Conflicts of Interest

The authors have no conflict of interest.

References

  1. Shilin, I.A.; Choi, J. Maximal subalgebras in so(2, 1), addition theorems and Bessel–Clifford functions. J. Anal. 2023, 31, 719–732. [Google Scholar] [CrossRef]
  2. Shilin, I.A.; Choi, J. On some relations between hyper Bessel–Clifford, Macdonald and Meijer functions and hyper Hankel–Clifford integral transforms. Integral Transforms Spec. Funct. 2023, 34, 788–798. [Google Scholar] [CrossRef]
  3. Miller, W. Symmetry and Separation of Variables; Cambridge University Press: Cambridge, UK, 1984. [Google Scholar]
  4. Vilenkin, N.J. Special Functions and the Theory of Group Representations. In Translations of Mathematical Monographs; Singh, V.N., Translator; American Mathematical Society: Providence, RI, USA, 1968; Volume 22. [Google Scholar]
  5. Dattoli, G.; Torre, A.; Lorenzutta, S.; Maino, G.; Chiccoli, C. Generalized Bessel functions within the group representation formalism. Nuov. Cim. B 1996, 111, 143–164. [Google Scholar] [CrossRef]
  6. Khan, S.; Yasmin, G. Generalized Bessel functions and Lie algebra representation. Math. Phys. Anal. Geom. 2006, 8, 299–313. [Google Scholar] [CrossRef]
  7. Shilin, I.A.; Choi, J. Some formulas for ordinary and hyper Bessel-Clifford Functions related to the proper Lorentz group. J. Math. Sci. 2021, 259, 518–527. [Google Scholar] [CrossRef]
  8. Nikolov, A.V. Structure and parametrization of O(p, q) and U(p, q). Bulg. J. Phys. 1975, 2, 537–545. [Google Scholar]
  9. Shilov, G.Y. Mathematical Analysis: The Second Special Course; Nauka: Moscow, Russia, 1965. (In Russian) [Google Scholar]
  10. Prudnikov, A.P.; Brychkov, Y.A.; Marichev, O.I. Elementary Functions. In Integrals and Series; OPA (Overseas Publishers Association): Amsterdam, The Netherlands, 1990; Volume 1. [Google Scholar]
  11. Prudnikov, A.P.; Brychkov, Y.A.; Marichev, O.I. Special Functions. In Integrals and Series; OPA (Overseas Publishers Association): Amsterdam, The Netherlands, 1990; Volume 2. [Google Scholar]
  12. Gradshteyn, I.S.; Ryzhik, I.M. Table of Integrals, Series, and Products; Academic Press: Cambridge, MA, USA, 2007. [Google Scholar]
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

Shilin, I.A.; Choi, J. Concerning Transformations of Bases Associated with Unimodular diag(1, −1, −1)-Matrices. Axioms 2024, 13, 452. https://doi.org/10.3390/axioms13070452

AMA Style

Shilin IA, Choi J. Concerning Transformations of Bases Associated with Unimodular diag(1, −1, −1)-Matrices. Axioms. 2024; 13(7):452. https://doi.org/10.3390/axioms13070452

Chicago/Turabian Style

Shilin, I. A., and Junesang Choi. 2024. "Concerning Transformations of Bases Associated with Unimodular diag(1, −1, −1)-Matrices" Axioms 13, no. 7: 452. https://doi.org/10.3390/axioms13070452

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