Next Article in Journal
Bending Analysis of Functionally Graded Nanoscale Plates by Using Nonlocal Mixed Variational Formula
Next Article in Special Issue
On Semi-Classical Orthogonal Polynomials Associated with a Modified Sextic Freud-Type Weight
Previous Article in Journal
Discounted Optimal Stopping of a Brownian Bridge, with Application to American Options under Pinning
Previous Article in Special Issue
On Row Sequences of Hermite–Padé Approximation and Its Generalizations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

On an Energy-Dependent Quantum System with Solutions in Terms of a Class of Hypergeometric Para-Orthogonal Polynomials on the Unit Circle

by
Jorge A. Borrego-Morell
1,*,†,
Cleonice F. Bracciali
2,† and
Alagacone Sri Ranga
2,†
1
Departamento de Matemática, Campus Santa Cruz da Serra, UFRJ—Universidade Federal do Rio de Janeiro, Duque de Caxias 25255-030, Brazil
2
Departamento de Matemática, Campus São José do Rio Preto, UNESP—Universidade Estadual Paulista, São Jose do Rio Preto 15054-000, Brazil
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Mathematics 2020, 8(7), 1161; https://doi.org/10.3390/math8071161
Submission received: 23 June 2020 / Revised: 8 July 2020 / Accepted: 10 July 2020 / Published: 15 July 2020

Abstract

:
We study an energy-dependent potential related to the Rosen–Morse potential. We give in closed-form the expression of a system of eigenfunctions of the Schrödinger operator in terms of a class of functions associated to a family of hypergeometric para-orthogonal polynomials on the unit circle. We also present modified relations of orthogonality and an asymptotic formula. Consequently, bound state solutions can be obtained for some values of the parameters that define the model. As a particular case, we obtain the symmetric trigonometric Rosen–Morse potential for which there exists an orthogonal basis of eigenstates in a Hilbert space. By comparing the existent solutions for the symmetric trigonometric Rosen–Morse potential, an identity involving Gegenbauer polynomials is obtained.

1. Introduction

An energy-dependent Schrödinger equation appears for the first time in relativistic quantum mechanics with the Pauli–Schrödinger equation, given by Pauli [1] in the description of the spectrum of an electron in the presence of a magnetic field. Further developments in relativistic and non relativistic quantum mechanics was made by many authors [2,3,4,5,6,7,8,9,10]. The list is by no means exhaustive.
These classes of quantum potentials appear frequently in many areas of quantum mechanics. A relativistic scalar particle in presence of an electromagnetic field can be studied by means of a Klein–Gordon equation with an energy dependent potential [4,8]. In [11], the authors have applied energy dependent potentials with emphasis on confining potentials to the description of heavy quark systems. Furthermore, the description of systems of N bosons bound is considered in [12] and the Hamiltonian formulation of the relativistic many-body problem in [4,5,13] also lead to energy dependent potential models. For physical applications in hydrodynamics, see [14].
Mathematical aspects of wave equations with energy-dependent potentials have been studied by several authors. The presence of an energy-dependence in the potential in the nonrelativistic context requires a modification of the underlying quantum theory, principally affecting orthogonality relation and norm [15]. An analogous modification is required in the relativistic framework [16]. An extension of the quantum mechanical formalism of systems with energy-dependent potentials to systems defined by generalized Schrödinger equations including a position-dependent mass was studied in [7]. Energy-dependence in the framework of noncommutative quantum mechanics has been recently considered in [17].
The search of solutions for energy dependent potentials in wave equations has attracted considerable attention since the appearance of Pauli’s work. In the present manuscript we study a quantum system with energy dependent potential related to the Rosen–Morse trigonometric potential, used in describing the interatomic interaction of linear molecules and for describing polyatomic vibration states and energies of the N H 3 molecule [18]. It has long been known that some mathematical features of quantum systems with an energy-dependent potential have several non-trivial implications; for instance, it is necessary to modify the scalar product to guarantee the conservation of the norm [15,19]. In the present manuscript we give closed-form of solutions, modified relations of orthogonality given by a indefinite (in general) bilinear form and an asymptotic formula of the solutions. Consequently, bound state solutions can be obtained for some values of the parameters that define the model. The solutions are given in terms of a class of functions derived from a sequence of hypergeometric para-orthogonal polynomials on the unit circle. We obtain, as a particular case, the symmetric trigonometric Rosen–Morse potential. In such case, the solutions reduce to an orthogonal basis of eigenfunctions defined in a Hilbert space and are expressed in terms of the Gegenbauer or ultraspherical polynomials. By comparing this solution with other solutions given in the literature we obtain, as a consequence, an identity involving Gegenbauer polynomials. Our procedure to obtain the energy dependent potential is based in a classical technique developed by Bose in [20] to construct solvable one-variable Schrödinger potentials.
The manuscript is organized as follows. In Section 2 we give some background, notations and statement of the results, in Section 3 we give the proofs and in Section 4 we present a discussion and concluding remarks.

2. Basic Notations and Statement of the Results

Let μ be a measure on the unit circle T = D , D = { z C : | z | < 1 } with support consisting of an infinite number of points. We remind that ( ϕ n ) n = 0 is the sequence of orthonormal polynomials on the unit circle associated to μ (also termed as Szegő polynomials, after their introduction by G. Szegő), if
T ϕ m ( z ) ϕ n ( z ) ¯ d μ ( z ) = δ m , n ,
where ϕ n ( z ) = κ n z n + a n 1 z n 1 + lower order terms and κ n > 0 .
An exposition of the theory of orthogonal polynomials systems on the unit circle is presented in the monographs [21,22,23]. More recent surveys in [24,25].
If P n is a polynomial of degree n, the reciprocal polynomial P n * is defined as z n P n ( 1 / z ¯ ) ¯ , or equivalently
P n * ( z ) = k = 0 n a ¯ k z n k if P n ( z ) = k = 0 n a k z k and a n 0 .
Let 2 F 1 ( a , b ; c ; z ) denote the Gauss hypergeometric function of the variable z with parameters a , b , c C , c Z 0 ; cf. [26] (p. 56), given by
2 F 1 ( a , b ; c ; z ) = k = 0 ( a ) k ( b ) k ( c ) k z k k ! ,
for z D and for other values of z C by analytic continuation appropriately; the Pochhammer symbol is defined by ( a ) n = a ( a + 1 ) ( a + n 1 ) , ( a ) 0 = 1 .
For α C , the function z α will be defined on the branch for which arguments are restricted between π and π . We also denote by x the floor function, defined as the greatest integer less than or equal to x.
A fundamental role in this manuscript is played by the sequence of functions
G n ( x ; λ , η ) = ( 2 λ ) n ( λ ) n ( 4 z ) n / 2 R n ( z ; b ) , λ > 1 2 , η R , n N { 0 } ,
where λ , η and x are such that b = λ + η ı and 2 x = z + 1 z , z = e ı θ , θ [ 0 , 2 π ] .
The functions G n were introduced in [27] and are defined from the sequence ( R n ) n = 0 of para-orthogonal polynomials, cf. [28]
b + b ¯ b R n ( z ; b ) = S n ( z ; b ) + b ¯ b S n * ( z ; b ) ,
associated to the Szegő hypergeometric polynomials, cf. [29,30]
S n ( z ; b ) = 2 F 1 ( n , b + 1 ; b + b ¯ + 1 ; 1 z ) .
These polynomials satisfy the orthogonality relations in the unit disk through the parametrization z = e 2 ı θ , θ [ 0 , π ]
1 π 0 π S m ( e 2 ı θ ; b ) ¯ S n ( e 2 ı θ ; b ) ω ( θ ; b ) d θ = n ! ( b + b ¯ + 1 ) n δ m , n , λ > 1 2 ; m , n N { 0 }
where
ω ( θ ; b ) = τ ( b ) e 2 η θ sin 2 λ θ ,
here the constant
τ ( b ) = | Γ ( b + 1 ) | 2 Γ ( b + b ¯ + 1 ) 4 ( b ) e π ( b )
is such that the moment μ 0 ( b ) = 1 π 0 π ω ( θ ; b ) d θ = 1 .
In the sequel, we denote by S ^ n and s n the monic and orthonormal polynomials of degree n respectively associated to S n . From (4) it follows that
κ n ( b ) = | ( b + 1 ) n | n ! ( b + b ¯ + 1 ) n ,
is the main coefficient of s n .
It should be noted that the R n polynomials are of hypergeometric type. According to [30] (Th. 5.1), one has
R n ( z ; b ) = 2 F 1 ( n , b ; b + b ¯ ; 1 z ) , [ b ] > 1 2 , [ b ] 0 .
This last relation can also be extended to [ b ] = 0 , η 0 if one takes
lim [ b ] 0 b + b ¯ b R n ( z ; b ) = S n ( z ; ı η ) S n * ( z ; ı η ) = n ( z 1 ) 2 F 1 ( n + 1 , 1 + ı η ; 2 ; 1 z ) ,
for n 1 .
In the present manuscript we prove the following results.
Theorem 1.
Let λ > 1 2 , η R . Then, the stationary one-dimensional Schrödinger equation with energy dependent potential
Ψ ( θ ) + ( E V ( θ , E ; λ , η ) ) Ψ ( θ ) = 0 , θ ( 0 , π ) ,
where
V ( θ , E ; λ , η ) = V 0 ( θ ; λ , η ) ( λ + E ( n ; λ , η ) + η 2 ) V 1 ( θ ; λ , η ) , V 0 ( θ ; λ , η ) = 2 λ η cot θ + λ ( 1 λ ) 1 sin 2 θ , V 1 ( θ ; λ , η ) = 2 η cot θ , E ( n ; λ , η ) = ( n + λ ) 2 η 2 ,
has by solution the system of wave functions
Ψ n ( θ ; λ , η ) = G n cos θ ; λ , η e η θ sin λ θ , n N { 0 } ,
in L 2 [ 0 , π ] .
The usual continuity equation
P ( θ , t ) t = J ( θ , t ) θ ,
where P denotes the probability density and J the probability current, governs the conservation of mass, charge, and probability of any closed system. If the potential is energy-dependent, it is necessary to modify the definition of the usual orthogonality relations in order to satisfy the continuity equation, [5,15]. More precisely, let ( Ψ n ) n = 0 be a system of normalizable wave functions solutions of an energy dependent potential Schrödinger equation defined through the boundary value problem
Ψ ( θ ) + ( E V ( θ , E ) ) Ψ ( θ ) = 0 , θ ( a , b ) , Ψ ( a ) = Ψ ( b ) = 0 ,
where V is of class C 1 ( I ) with respect to the variable E, being I an open interval of the real line and a , b R ¯ .
The continuity equation read as
P ( θ , t ) t + ı ( V ( θ , E n 2 ) V ( θ , E n 1 ) ) Ψ ^ n 1 ( θ , t ) ¯ Ψ ^ n 2 ( θ , t ) = J ( θ , t ) θ ,
and is satisfied by the probability density P and probability current J
P ( θ , t ) = Ψ ^ n 1 ( θ , t ) ¯ Ψ ^ n 2 ( θ , t ) , J ( θ , t ) = ı Ψ ^ n 1 ( θ , t ) ¯ Ψ ^ n 2 ( θ , t ) θ Ψ ^ n 1 ( θ , t ) ¯ θ Ψ ^ n 2 ( θ , t ) ,
where Ψ ^ n ( θ , t ) = e ı E n t Ψ n ( θ ) is a solution to the time-dependent Schrödinger equation
ı Ψ ( θ , t ) t = θ 2 + V θ , ı t Ψ ( θ , t ) .
The orthogonality relation between two states n and m, n m reads as
Ψ n , Ψ m = a b Ψ n ( θ ) 1 V ( θ , E n ) V ( θ , E m ) E n E m Ψ m ( θ ) d θ = 0 ,
now, by using the smooth dependence of V in relation to E one obtains
Ψ n , Ψ n = a b Ψ n 2 ( θ ) 1 V ( θ , E ) E E = E n d θ .
In that regard, for the present quantum model we have the following relations of orthogonality. Notice that the presence of the function cot θ in the definition implies that the associated bilinear form is not in general of a definite sign. When | η | < λ , we have bound states solutions.
Theorem 2.
Let λ > 0 , η R and ( Ψ n ) n = 0 be the wave functions of Theorem 1. Then, ( Ψ n ) n = 0 satisfy the relation of orthogonality
Ψ n , Ψ m = 0 π Ψ n ( θ ; λ , η ) 1 2 η cot θ n + m + 2 λ Ψ m ( θ ; λ , η ) d θ = c n ( λ , η ) δ n , m ,
where n , m N { 0 } , c n R . When λ > 1 2 and η R one has
c n ( λ , η ) = 1 η 2 ( n + λ ) 2 π n ! ( λ + n ) Γ ( 2 λ + n ) 2 2 n 1 Γ ( 2 λ + 1 ) [ ( λ ) n ] 2 b b + n .
For the particular case η = 0 we obtain the symmetric trigonometric Rosen–Morse potential, cf. [31] (Prob. 12), [23,32,33] ((4.7.11) p. 81). Let P n ( λ ) , λ > 1 2 be the Gegenbauer polynomial of degree n cf. [23] (p. 80). As a consequence of Theorem 1 we obtain
Corollary 1.
The stationary one-dimensional Schrödinger equation
Ψ ( θ ) + E λ ( λ 1 ) sin 2 θ Ψ ( θ ) = 0 , θ ( 0 , π ) , λ > 1 2 ,
admits the energy eigenstates
E ( n ; λ ) = ( n + λ ) 2 , n N { 0 } .
and the complete orthogonal system of wave functions
Ψ n ( θ ; λ , 0 ) = 2 n ( 2 λ ) n n + 2 λ 1 n ( λ ) n P n ( λ ) ( cos θ ) sin λ θ , n N { 0 } ,
in the Hilbert space L 2 [ 0 , π ] with the inner product f , g = 0 π f ( t ) g ( t ) d t .
Notice that from this corollary it follows immediately that the ground state energy for the symmetric trigonometric Rosen–Morse potential reduces to
E ( 0 ; λ , 0 ) = λ 2 .
The asymmetric trigonometric Rosen–Morse potential or Rosen–Morse I potential, cf. [32,34,35] whose associated Schrödinger equation reads as
Φ ( θ ) + E 2 b cot θ + a ( a 1 ) 1 sin 2 θ Φ ( θ ) = 0 , θ ( 0 , π ) , a 3 2 ,
is among the exactly solvable potentials. Bound state solutions can be given in terms of the Jacobi polynomials with purely imaginary arguments and complex conjugate parameters, cf. [32] (pp. 296–297).
This potential has also been studied in [36,37] and solved in terms of the real Romanovski polynomials [38,39] R n ( a , b ) (also known as Romanovsky–Routh or Pseudo–Jacobi polynomial as in [40]). These polynomials are defined as the polynomial solution of degree n of the differential equation
( 1 + x 2 ) d 2 y d x 2 + 2 ( ( 1 + a ) x + b ) d y d x n ( n + 2 a + 1 ) y = 0 ,
and can be expressed in terms of the Jacobi polynomials P n ( α , β ) , cf. [24] [(20.1.1) p. 509], [23] [(4.21.2) p. 62] as
R n ( a , b ) ( x ) = ( ı ) n P n ( a + b ı , a b ı ) ( ı x ) = ( ı ) n n + a + b ı n 2 F 1 n , n + 2 a + 1 ; 1 + a + b ı ; 1 ı x 2
For convenience, we will adopt the parametrization given in [41] for the solution to (10) in terms of the real Romanovski polynomials (expressed also in terms of author’s parametrization), which reads
Φ n ( θ ; a , b ) R n ( 2 b n + a , 1 n a ) ( cot θ ) e b θ n + a sin ( a + n ) θ ( ı ) n P n ( n a ı b a + n , n a + ı b a + n ) ( ı cot θ ) e b θ n + a sin ( a + n ) θ ,
for n 0 , and the corresponding energies
E Φ ( n ; a , b ) = ( a + n ) 2 b 2 ( a + n ) 2 .
In particular, the symmetric trigonometric Rosen–Morse is obtained by taking η = 0 in (10). In such case we have that
Φ n ( θ ; a , 0 ) ( ı ) n P n ( n a , n a ) ( ı cot θ ) sin ( a + n ) θ ,
is a solution for the Schrödinger equation associated to the symmetric Rosen–Morse potential.
By identifying the parameters λ = a we obtain
Φ n ( θ ; λ ) ( ı ) n P n ( n λ + 1 2 ) ( ı cot θ ) sin ( λ + n ) θ ,
is also a solution of (9), which coincides, up to a multiplicative factor with Ψ n , as shows the following identity
Theorem 3.
Let λ C and n N { 0 } . Then,
( ı ) n ( λ + 1 2 ) n 2 ( n λ + 1 2 ) n P n ( n λ + 1 2 ) ( ı cot θ ) sin n θ = ( 1 ) n 2 ( 1 n λ ) n 2 ( λ ) n P n ( λ ) ( cos θ ) .
For those values of λ, say λ = λ 0 , for which we have a zero or pole in the expressions ( λ + 1 2 ) n 2 ( n λ + 1 2 ) n or ( 1 n λ ) n 2 ( λ ) n the formula may be interpreted as a limit when λ λ 0 . In such cases, the limit exists and is finite.
For the next result, let ϵ 1 , ϵ 2 > 0 be fixed quantities sufficiently small so that the interval [ ϵ 1 , π ϵ 2 ] lies wholly in ( 0 , π ) . We have that
Theorem 4.
Let λ > 1 2 , η R . Then,
2 n 1 Ψ n ( θ ; λ , η ) = Γ ( λ ) Γ ( b ) e ı ( n θ + η ln n ) ( 1 e 2 ı θ ) b ¯ e η θ sin λ θ + o ( 1 ) , a s n .
The bound for the error holds uniformly in θ [ ϵ 1 , π ϵ 2 ] .

2.1. Some Basic Facts about the Functions G n

We recall that a polynomial P n ( z ) = j = 0 n a j z j , a j C is conjugate reciprocal if P n satisfies the identity,
P n ( z ) = P n * ( z ) ,
that is, a j ¯ = a n j , j = 0 , 1 , , n . From (1), it follows that G n , n 0 takes real values.
Let λ > 0 , define the sequences
α n + 1 ( λ ) = 1 4 n ( n + 2 λ 1 ) ( n + λ 1 ) ( n + λ ) , β n ( λ , η ) = η λ + n 1 , n 1 ,
it follows from [42] (Section 2) that the functions G n satisfy the recurrence relation
G n + 1 ( cos θ ; λ , η ) = ( cos θ β n + 1 ( λ , η ) sin θ ) G n ( cos θ ; λ , η ) α n + 1 ( λ ) G n 1 ( cos θ ; λ , η ) , θ [ 0 , π ] ,
where G 0 ( cos θ ; λ , η ) = 1 and G 1 ( cos θ ; λ , η ) = cos θ β 1 ( λ , η ) sin θ .
Let λ > 1 2 and n , m N { 0 } . It follows from [42] (Th. 5.2),
0 π G 2 n cos θ ; λ , η G 2 m + 1 cos θ ; λ , η ω θ ; b sin θ d θ = 0 , 0 π G 2 n cos θ ; λ , η G 2 m cos θ ; λ , η ω ( θ ; b ) d θ = τ ( b ) δ 2 n ( b ) δ n , m , 0 π G 2 n + 1 cos θ ; λ , η G 2 m + 1 cos θ ; λ , η ω ( θ ; b ) d θ = τ ( b ) δ 2 n + 1 ( b ) δ n , m ,
where
δ n ( b ) = π n ! ( λ + n ) Γ ( 2 λ + n ) 2 2 λ + 2 n 1 e η π | Γ ( b + n + 1 ) | 2 1 [ ( λ ) n ] 2 [ ( [ ( b ) n ] ) 2 + ( [ ( b ) n ] ) 2 ] .
Let μ be a finite positive Borel measure supported in [ π , π ) and d μ = w ( θ ) d θ 2 π + d ν s its Lebesgue decomposition. Recall that for z D , the Szegő function, cf. [25] (§2.4 p. 143, Part I), D ( d μ ; z ) is defined as
D ( d μ ; z ) = exp 1 4 π 0 2 π e ı θ + z e ı θ z log ( w ( θ ) ) d θ ,
whenever μ is the Szegő class. For almost every t [ π , π ) , the boundary value D ( d μ ; e ı t ) is defined as the radial limit:
D ( d μ ; e ı t ) = lim r 1 D ( d μ ; r e ı t ) .
From [30] (Th. 4.3), the Szegő function D ( ω ( θ ) d θ ; z ) , (or D ( z ; b ) for short) reads as
D ( z ; b ) = | Γ ( b + 1 ) | Γ ( b + b ¯ + 1 ) ( 1 z ) b .
The function G n ( x ) satisfy the differential equation, [27] (Th. 2.2)
( 1 x 2 ) y ( ( 2 λ + 1 ) x 2 η 1 x 2 ) y + n n + 2 λ + 2 η x 1 x 2 y = 0 , x ( 1 , 1 ) , n N { 0 } .
When η = 0 , we obtain the differential equation that defines the ultraspherical polynomials cf. [23] (p. 80). Notice that from (1) and (2), G n reduces to a polynomial. Therefore, G n coincides, up to a constant factor, with the n degree Gegenbauer polynomial.
Remark 1.
We remark that in [27] (Th. 2.2), the term m should be corrected in the last summand of the left hand side of the differential equation.

The Schrödinger Invariant of a Second Order Differential Equation

Quantum systems with energy-dependent potentials have been studied following several approaches such as supersymmetric quantum mechanics, Darboux transformations, exceptional orthogonal polynomials, among others, see [7] for a review. In this subsection we summarize the method we followed, introduced by Bose in [20] in order to construct one-variable Schrödinger solvable potentials.
Let us have a second order differential equation
u ( x ) + p ( x ) u ( x ) + q ( x ) u ( x ) = 0 , x I ,
being I an open subset of the real line and where p and q are functions defined on I. A straightforward calculation shows that the middle term in (19) can be eliminated by taking the substitution
u = v e 1 2 p ( t ) d t .
Under the above substitution, the Equation (19) transforms to the canonical form
v ( x ) + I ( x ) v = 0 ,
where I is given by
I ( x ) = q ( x ) 1 2 p ( x ) 1 4 ( p ( x ) ) 2 ,
the term I is named by Milson in [43] as the Bose invariant.
By applying now the transformation v = x ψ , x = d x d θ , we obtain the normal form
ψ ( θ ) + I S ( θ ) ψ ( θ ) = 0 ,
being
I S = ( x ) 2 I ( x ) + 1 2 { x , θ } ,
the Schrödinger invariant, named by Bose in [20] and { x , θ } is the Schwartzian derivative
{ x , θ } = x x 3 2 x x 2 .

3. Proof of the Results

Proof of Theorem 1.
Under the transformation
y = v 1 x 2 λ 2 1 4 e η arccos ( x ) , x ( 1 , 1 ) ,
the Equation (18) reduces to its normal form
v + I ( x ) v = 0 ,
where
I ( x ) = 2 4 η 2 + 4 λ + 4 n 2 + 8 λ n + x 2 4 η 2 4 ( λ + n ) 2 + 1 + 8 η n x 1 x 2 4 x 2 1 2 + 8 η λ x 1 x 2 4 x 2 1 2 .
By applying the transformations to (22)
x = cos θ , θ ( 0 , π ) , v = ( x ) 1 / 2 Ψ ,
we obtain the Schrödinger normal form
Ψ + ( κ ( n ; λ , η ) V ( θ ; n , λ , η ) ) Ψ = 0 , θ ( 0 , π ) ,
where κ ( n ; λ , η ) = ( n + λ ) 2 η 2 and
V ( θ ; n , λ , η ) = V 0 ( θ ; λ , η ) n V 1 ( θ ; λ , η ) , V 0 ( θ ; λ , η ) = 2 λ η cot θ + λ ( 1 λ ) 1 sin 2 θ , V 1 ( θ ; λ , η ) = 2 η cot θ .
By identifying κ ( n ; λ , η ) with the energy term we get
E ( n ; λ , η ) = ( n + λ ) 2 η 2 ,
therefore
n = λ + E + η 2 .
Substituting (25) and (26) into (24) we obtain (6).
Since G n is a solution of (18), it follows from (21) and (23) that
Ψ n ( θ ; λ , η ) = G n cos θ ; λ , η e η θ sin λ θ , n N { 0 } ,
is a solution of (24). □
Proof of Theorem 2.
On the one hand, by applying iterated integration to the left hand side of (7) and taking into account the definition of P we obtain
0 π 0 t P ( θ , t ) t + ı ( V ( θ , E m ) V ( θ , E n ) ) Ψ ^ n ( θ , t ) ¯ Ψ ^ m ( θ , t ) d t d θ = e ı ( E n E m ) t 1 0 π Ψ n ( θ ) 1 V ( θ , E n ) V ( θ , E m ) E n E m Ψ m ( θ ) d θ .
On the other hand, for the right hand side of (7) we have
0 π 0 t J ( θ , t ) θ d t d θ = 0 t 0 π J ( θ , t ) θ d θ d t
Now, from the definition of J and the fact that Ψ n ( 0 ; λ , η ) = Ψ n ( π ; λ , η ) = 0 when λ > 0 one has
0 π J ( θ , t ) θ d θ = 0 .
Hence, from (27), (28) and (29)
0 π Ψ n ( θ ; λ , η ) 1 V ( θ , E n ) V ( θ , E m ) E n E m Ψ m ( θ ; λ , η ) d θ = 0 , n m .
By substituting the value of V given in Theorem 1 we obtain (8).
To evaluate the numerical value of the constant c n , notice that if λ > 1 2 , η R , from the recurrence relations (15) and (16) we have
0 π Ψ n 2 ( θ ; λ , η ) cot θ d θ = η λ + n 0 π Ψ n 2 ( θ ; λ , η ) d θ ,
hence,
c n ( λ , η ) = 1 η 2 ( n + λ ) 2 π n ! ( λ + n ) Γ ( 2 λ + n ) 2 2 n 1 Γ ( 2 λ + 1 ) [ ( λ ) n ] 2 b b + n .
This completes the proof of the theorem. □
Proof of Corollary 1.
From Theorem 1, the eigenstates are given by the system of real functions
Ψ n ( θ ; λ ) = G n cos θ ; λ , 0 sin λ θ , n N { 0 } ,
From (18), G n reduces, up to a multiplicative constant c n factor, to the Gegenbauer polynomial of degree n, cf. [23] (§4.7 p. 80). To find the multiplicative constant c n , note that from (1), (2) and (3) we deduce that
G n 1 ; λ , 0 = 2 n ( 2 λ ) n ( λ ) n = c n P n ( λ ) ( 1 ) ,
therefore, from [23] [(4.7.3) p. 80] one obtains
c n = 2 n ( 2 λ ) n n + 2 λ 1 n ( λ ) n .
The corresponding energy eigenstates read as
E ( n ; λ ) = n + λ 2 , n N { 0 } .
From the orthogonality relation for Gegenbauer polynomials
1 1 P n ( λ ) ( x ) ( 1 x 2 ) λ d x ,
one has that Ψ n ( θ ; λ ) is an orthogonal system in the Hilbert space L 2 [ 0 , π ] with the scalar product
f , g = 0 π f ( t ) g ( t ) d t .
 □
Proof of Theorem 3.
From Corollary 1 and the relation (14) we have that Φ n and Ψ n are solutions of the differential equation
Ψ ( θ ) + E λ ( λ 1 ) sin 2 θ Ψ ( θ ) = 0 , θ ( 0 , π ) , λ 3 2 ,
hence, the transformations (21) and (23) give that
y 1 ( x ) = ( x ) 1 2 ( 1 x 2 ) λ 2 1 4 Ψ n ( θ ; λ ) ,
y 2 ( x ) = ( x ) 1 2 ( 1 x 2 ) λ 2 1 4 Φ n θ ; λ ) ,
are solutions of the differential equation
( 1 x 2 ) y ( 2 λ + 1 ) x y + n n + 2 λ y = 0 , x ( 1 , 1 ) , λ 3 2 ,
notice that y 1 reduces, up to a constant factor, to the Gegenbauer polynomial of degree n.
From (14) and (31) one has
y 2 ( x ) = ( ı ) n P n ( n λ + 1 2 ) ı x 1 x 2 ( 1 x 2 ) n 2 ,
and from [44] [(6.4.12) p. 303],
P n ( n λ + 1 2 ) ı x 1 x 2 = ( n λ + 1 2 ) n ( 2 ı ) n n ! x n ( 1 x 2 ) n 2 2 F 1 n 2 , 1 n 2 ; λ + 1 2 ; x 2 1 x 2 ,
hence, from (33)
( 1 x 2 ) n 2 P n ( n λ + 1 2 ) ı x 1 x 2 = ( n λ + 1 2 ) n ( 2 ı ) n n ! x n 2 F 1 n 2 , 1 n 2 ; λ + 1 2 ; x 2 1 x 2 .
Notice that the relation (34) defines the left hand side as a polynomial of degree n whose coefficients are rational functions of the variable λ varying in C .
Now, (33) is a solution to (32). Since this solution is a polynomial of degree n, it follows from [23] (Th. 4.2.2 p. 61) that
y 2 = c o n s t . y 1 ,
when [ λ ] 3 2 . Using formula [44] [(6.4.12) p. 303], by comparing P 2 k ( n λ 1 ) ( 0 ) with P 2 k ( λ ) ( 0 ) and P 2 k + 1 ( n λ 1 ) ( 0 ) with P 2 k + 1 ( λ ) ( 0 ) , from (35) we obtain
( ı ) n ( λ + 1 2 ) n 2 ( n λ + 1 2 ) n P n ( n λ + 1 2 ) ( ı cot θ ) sin n θ = ( 1 ) n 2 ( 1 n λ ) n 2 ( λ ) n P n ( λ ) ( cos θ ) , [ λ ] 3 2 .
Let us consider n fixed. Since the coefficients of y 2 and y 1 are rational functions of λ and are equal when [ λ ] [ 3 2 , + ) , it follows from [45] (Th. 17.1 p. 369) that the relation is valid for λ C , with exception of a finite number of special values of λ . Notice that from [44] [(6.4.12) p. 303], the zeros of the functions ( n λ + 1 2 ) n and ( λ ) n are removable singularities of the left hand side and right hand side respectively of (36). Furthermore, when λ is a zero of ( λ + 1 2 ) n 2 or ( 1 n λ ) n 2 we have simple poles in the main coefficients of the left hand side or right hand side accordingly. For these values of λ 0 , the formula may be interpreted as a limit as λ λ 0 . This completes the proof of the theorem. □
To prove Theorem 4 a preliminary lemma is necessary.
Lemma 1.
Let λ > 1 2 and η R be fixed and define for θ ( 0 , π ) ,
a n ( θ ) = D 1 ( 0 ; b ) π 0 π D 1 ( e ı θ ; b ) D 1 ( e ı θ ; b ) 1 e ı ( θ θ ) s n * ( e ı θ ; b ) ¯ ω ( θ ; b ) d θ , b n ( θ ) = D 1 ( 0 ; b ) π 0 π D 1 ( e ı θ ; b ) D 1 ( e ı θ ; b ) 1 e ı ( θ θ ) s n ( e ı θ ; b ) ¯ ω ( θ ; b ) d θ .
Then,
lim n a n ( θ ) = 0 , θ [ ϵ 1 , π ϵ 2 ] , lim n b n ( θ ) = 0 , θ [ ϵ 1 , π ϵ 2 ] ,
being ϵ 1 , ϵ 2 > 0 any fixed quantities sufficiently small so that [ ϵ 1 , π ϵ 2 ] ( 0 , π ) .
Proof. 
Let us denote
K ( θ , θ ) = D 1 ( 0 ; b ) D 1 ( e ı θ ; b ) D 1 ( e ı θ ; b ) 1 e ı ( θ θ ) .
From (17), we have that for θ ( 0 , π ) fixed, K ( θ , θ ) is continuous as a function of θ [ 0 , π ] , hence
K ( θ , θ ) L 2 ( ω ( θ ; b ) d θ ) ,
therefore, from Lemma [21] (Lem. 4.2 p. 220)
lim n a n ( θ ) = 0 ,
pointwise in the interval ( 0 , π ) .
Let ϵ 1 , ϵ 2 > 0 be any fixed quantities sufficiently small so that [ ϵ 1 , π ϵ 2 ] ( 0 , π ) . Since K ( θ , θ ) is continuous in the compact set [ 0 , π ] × [ ϵ 1 , π ϵ 2 ] , the Heine–Cantor Theorem cf. [46] (Th. 2 p. 201) implies that K ( θ , θ ) is uniformly continuous in [ 0 , π ] × [ ϵ 1 , π ϵ 2 ] . Hence, for every ϵ > 0 there exists δ > 0 such that if θ 1 , θ 2 [ ϵ 1 , π ϵ 2 ] , | θ 1 θ 2 | < δ , then
| K ( θ , θ 1 ) K ( θ , θ 2 ) | < ϵ , θ [ 0 , π ] ,
therefore,
| a n ( θ 1 ) a n ( θ 2 ) | 1 π 0 π | K ( θ , θ 1 ) K ( θ , θ 2 ) | 2 ω ( θ ; b ) d θ 1 2 < ϵ , n N { 0 } ,
this shows that the family ( a n ( θ ) ) n 0 , θ [ ϵ 1 , π ϵ 2 ] is equicontinuous.
On the other hand, we have also that there exists M > 0 such that
| a n ( θ ) | 1 π 0 π | K ( θ , θ ) | 2 ω ( θ ; b ) d θ M , θ [ ϵ 1 , π ϵ 2 ] , n N { 0 } ,
this shows that the family ( a n ( θ ) ) n 0 , θ [ ϵ 1 , π ϵ 2 ] is uniformly bounded.
From Arzela’s Theorem, cf. [47] (p. 54) it follows that the family of functions ( a n ) n 0 is compact in C [ ϵ 1 , π ϵ 2 ] equipped the uniform norm. Therefore, from (37) every uniform convergent subsequence of ( a n ) n 0 converges to the same limit, hence for the whole sequence we have
lim n a n ( θ ) = 0 , θ [ ϵ 1 , π ϵ 2 ] .
By using a similar argument, we also conclude that
lim n b n ( θ ) = 0 , θ [ ϵ 1 , π ϵ 2 ] .
 □
Proof of Theorem 4.
From (2)–(4) it follows that
b + b ¯ b R n ( z ; b ) = ( b + 1 ) n ( b ¯ + b + 1 ) n S ^ n ( z ; b ) + b ¯ b ( b ¯ + 1 ) n ( b ¯ + b + 1 ) n S ^ n * ( z ; b ) , n 1 ,
hence, from (1)
G n ( cos θ ; λ , η ) = 2 n 1 ( 2 λ ) n ( λ ) n κ n ( b ) b ( b + 1 ) n λ ( b ¯ + b + 1 ) n z n 2 s n ( z ; b ) + b ¯ ( b ¯ + 1 ) n λ ( b ¯ + b + 1 ) n z n 2 s n * ( z ; b ) , z = e 2 ı θ , θ [ 0 , π ] ,
where κ n is given by (5).
From [21] [(2.7) p. 200] we have that
lim n κ n ( b ) = δ 1 ( b ) ,
where δ ( b ) = e 1 2 π 0 π log ( ω ( θ ) ) d θ > 0 is the solution of the Szegő extremum problem, [21] (p. 200 & Th. 2.5 p. 204).
From the relation [21] (p. 206)
δ 2 ( b ) = Δ ( 0 ; b ) ,
where
Δ ( z ; b ) = D 1 ( z ; b ) δ 1 ( b ) ,
cf. [21] [(3.4) p. 209], one finds that
δ ( b ) = D ( 0 ; b ) .
The representation [21] [(4.6) p. 220] gives
e ı n θ s n ( e 2 ı θ ; b ) = κ n ( b ) e ı n θ D 1 ( e 2 ı θ ; b ) ¯ a n ( θ ) e ı n θ D 1 ( e 2 ı θ ; b ) ¯ + b n ( θ ) e ı n θ D 1 ( e 2 ı θ ; b ) δ ( b ) ( | κ n ( b ) a n ( θ ) | 2 + | b n ( θ ) | 2 ) ,
with a n and b n defined as in Lemma 1.
Taking into account the representation
( z ) n = Γ ( z + n ) Γ ( z ) ,
from [26] [(4) p. 47] one has
( γ 1 ) n ( γ 2 ) n = Γ ( γ 2 ) Γ ( γ 1 ) n γ 1 γ 2 1 + O 1 n , a s n .
Hence, from (38)–(40)
2 n + 1 Ψ n ( θ ; λ , η ) = 2 Γ ( λ ) Γ ( b ) n ı η e ı n θ ( 1 e 2 ı θ ) b ¯ + 2 Γ ( λ ) Γ ( b ¯ ) n ı η e ı n θ ( 1 e 2 ı θ ) b e η θ sin λ θ + r n ( θ ) .
Now, Lemma 1 gives us that,
r n ( θ ) = 1 Γ ( b ) a n ( θ ) e ı n θ ( 1 e 2 ı θ ) b ¯ + 1 Γ ( b ) b n ( θ ) e ı n θ ( 1 e 2 ı θ ) b O ( 1 ) + O 1 n = o ( 1 ) , a s n ,
uniformly in θ [ ϵ 1 , π ϵ 2 ] , being ϵ 1 , ϵ 2 > 0 any fixed quantities sufficiently small so that [ ϵ 1 , π ϵ 2 ] ( 0 , π ) . This completes the proof of the theorem. □

4. Discussion

The present work is devoted to the study an energy-dependent potential related to the Rosen–Morse potential. The system is obtained by the addition of a potential term which depends on the function cot θ and an energy dependence through a square root.
In order to show some numerical comparisons we will identify accordingly the parameters λ , η of the quantum model given by Theorem 1 and the parameters a , b of the asymmetric trigonometric Rosen–Morse model, following the form given in [41]. In effect, by identifying the parameters one has b = λ η , a = λ . Consequently, the quantum system defined by Theorem 1, in terms of the parameters a and b reads as,
Ψ ( θ ) + E 2 b cot θ + a ( a 1 ) 1 sin 2 θ + 2 b a a + E + b a 2 cot θ Ψ ( θ ) = 0 , θ ( 0 , π ) , a 1 2 , a 0 ,
which has, when n N { 0 } , the system of solutions
Ψ n θ ; a , b a = G n cos θ ; a , b a e b a θ sin a θ = ( 2 a ) n 2 n ( a ) n × ( cos θ + ı sin θ ) n 2 F 1 n , a ı b a ; 2 a ; 1 ( cos θ + ı sin θ ) 2 e b a θ sin a θ , a 1 2 , a 0 ,
and the corresponding energy levels,
E Ψ n ; a , b a = ( n + a ) 2 b a 2 .
On the other hand, following Theorem 3, we will multiply by an adequate numerical constant γ n the expression that defines the Φ n functions (12),
γ n = ( 1 ) n 2 2 n ( 2 a ) n ( a + 1 2 ) n 2 n + 2 a 1 n ( n a + 1 2 ) n ( 1 n a ) n 2 ( 1 2 n 2 a ) n ( 1 n a ) n .
By (11) and (12), we have that the rescaled function Φ n , n N { 0 } can be expressed as
Φ n ( θ ; a , b ) = γ n ( ı ) n P n ( n a ı b a + n , n a + ı b a + n ) ( ı cot θ ) e b θ n + a sin ( a + n ) θ = γ n ( ı ) n a ı b a + n n 2 F 1 n , 1 n 2 a ; 1 n a ı b a + n ; 1 ı cot θ 2 e b θ n + a sin ( a + n ) θ .
Notice that, by virtue of Corollary 1 and [23] [(4.7.1) p. 80] one has now that
Φ n ( θ ; a , 0 ) = Ψ n ( θ ; a , 0 ) , a > 1 2 , a 0 ,
we recover in particular, the symmetric trigonometric Rosen–Morse oscillator.
In Figure 1 and Figure 2 we plotted the wave functions for several toy values of the parameters. In Figure 3 and Figure 4 we plotted the densities 1 + b cot θ a ( n + a ) | Ψ n ( θ ; a , b ) | 2 (see Theorem 2) and | Φ n ( θ ; a , b ) | 2 for the same values of the parameters a , b with n fixed ( n = 5 ). As can be appreciated, by fixing b and making a variable, the abscissas of the local maxima of the densities tend to be localized at the same points in both models, as a increases. These points correspond to the regions where it is most likely the particle to be found. It should be also noticed from the expressions of the energies (13) and (42) that
lim a E Ψ ( n ; a , b ) E Φ ( n ; a , b ) = 0 ,
for b and n fixed. It could be interesting the further study of these facts.

Author Contributions

Conceptualization, J.A.B.-M., C.F.B. and A.S.R.; methodology, J.A.B.-M., C.F.B. and A.S.R.; software, J.A.B.-M., C.F.B. and A.S.R.; validation, J.A.B.-M., C.F.B. and A.S.R.; formal analysis, J.A.B.-M., C.F.B. and A.S.R.; investigation, J.A.B.-M., C.F.B. and A.S.R.; resources, J.A.B.-M., C.F.B. and A.S.R.; data curation, J.A.B.-M., C.F.B. and A.S.R.; writing—original draft preparation, J.A.B.-M., C.F.B. and A.S.R.; writing—review and editing, J.A.B.-M., C.F.B. and A.S.R.; visualization, J.A.B.-M., C.F.B. and A.S.R.; supervision, J.A.B.-M., C.F.B. and A.S.R.; project administration, J.A.B.-M., C.F.B. and A.S.R.; funding acquisition, J.A.B.-M., C.F.B. and A.S.R. All authors have read and agreed to the published version of the manuscript.

Funding

The author’s (C. F. Bracciali and A. Sri Ranga) research has been partially supported by grants from FAPESP (2016/09906–0) and from CNPq (402939/2016–6 and 304087/2018-1) of Brazil.

Acknowledgments

We thank the anonymous referees for carefully reviewing our manuscript. We also thank to Elso Drigo Filho from UNESP—Universidade Estadual Paulista, Campus São José do Rio Preto for helpful conversations.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Pauli, W. Zur Quantenmechanik des magnetischen Elektrons. Z. Phys. 1927, 43, 601–623. [Google Scholar] [CrossRef]
  2. Garcia-Martinez, J.; Garcia-Ravelo, J.; Peña, J.J.; Schulze-Halberg, A. Exactly solvable energy-dependent potentials. Phys. Lett. A 2009, 373, 3619–3623. [Google Scholar] [CrossRef]
  3. Langueur, O.; Merad, M.; Hamil, B. DKP Equation with Energy Dependent Potentials. Commun. Theor. Phys. 2019, 71, 1069–1074. [Google Scholar] [CrossRef]
  4. Rizov, V.A.; Sazdjian, H.; Todorov, T. On the relativistic quantum mechanics of two interacting spinless particles. Ann. Phys. 1985, 165, 59–97. [Google Scholar] [CrossRef]
  5. Sazdjian, H. Relativistic wave equations for the dynamics of two interacting particles. Phys. Rev. D 1986, 33, 3401. [Google Scholar] [CrossRef]
  6. Schiff, L.I.; Snyder, H.; Weinberg, J. On the existence of stationary states of the mesotron field. Phys. Rev. 1940, 57, 315–318. [Google Scholar] [CrossRef]
  7. Schulze-Halberg, A.; Yeşiltaş, Ö. Generalized Schrödinger equations with energy-dependent potentials: Formalism and applications. J. Math. Phys. 2018, 59, 113503. [Google Scholar] [CrossRef]
  8. Snyder, H.; Weinberg, J. Stationary states of scalar and vector fields. Phys. Rev. 1940, 57, 307–314. [Google Scholar] [CrossRef]
  9. Yekken, R.; Lassaut, M.; Lombard, R.J. Applying supersymmetry to energy dependent potentials. Ann. Phys. 2013, 338, 195–206. [Google Scholar] [CrossRef]
  10. Schulze-Halberg, A. Higher–order Darboux transformations and Wronskian representations for Schrödinger equations with quadratically energy–dependent potentials. J. Math. Phys. 2020, 61, 023503. [Google Scholar] [CrossRef]
  11. Lombard, R.J.; Mares, J.; Volpe, C. Wave equation with energy-dependent potentials for confined systems. J. Phys. G 2007, 34, 1–11. [Google Scholar] [CrossRef]
  12. Lombard, R.J.; Mares, J. The many-body problem with an energy-dependent confining potential. Phys. Lett. A 2009, 373, 426–429. [Google Scholar] [CrossRef]
  13. Mourad, J.; Sazdjian, H. The two-fermion relativistic wave equations of constraint theory in the Pauli–Schrödinger form. J. Math. Phys. 1998, 35, 6379–6406. [Google Scholar] [CrossRef] [Green Version]
  14. Li, Y. Some water wave equations and integrability. J. Nonlinear Math. Phys. 2005, 12, 466–481. [Google Scholar] [CrossRef]
  15. Formanek, J.; Lombard, R.J.; Mares, J. Wave equations with energy-dependent potentials. Czechoslov. J. Phys 2004, 54, 289–315. [Google Scholar] [CrossRef] [Green Version]
  16. Schulze-Halberg, A.; Roy, P. Bound states of the two-dimensional Dirac equation for an energy-dependent hyperbolic Scarf potential. J. Math. Phys. 2017, 58, 113507. [Google Scholar] [CrossRef]
  17. Harko, T.; Liang, S.-D. Energy-dependent noncommutative quantum mechanics. Eur. Phys. J. C 2019, 79, 1–22. [Google Scholar] [CrossRef] [Green Version]
  18. Rosen, N.; Morse, P.M. On the Vibrations of Polyatomic Molecules. Phys. Rev. 1932, 42, 210–217. [Google Scholar] [CrossRef]
  19. Sazdjian, H. The scalar product in two-particle relativistic quantum mechanics. J. Math. Phys. 1988, 29, 1620–1633. [Google Scholar] [CrossRef]
  20. Bose, A.K. A class of solvable potentials. Il Nuovo Cimento 1964, 32, 679–688. [Google Scholar] [CrossRef]
  21. Freud, G. Orthogonal Polynomials; Pergamon: Oxford, UK, 1971. [Google Scholar]
  22. Geronimus, L.Y. Polynomials Orthogonal on a Circle and Their Applications; American Mathematical Society Translations; American Mathematical Society: Providence, RI, USA, 1962; Volume 104, pp. 1–78. [Google Scholar]
  23. Szegő, G. Orthogonal Polynomials, 4th ed.; American Mathematical Society Colloquium Publications; American Mathematical Society: Providence, RI, USA, 1975; Volume 23. [Google Scholar]
  24. Ismail, M.E.H. Classical and Quantum Orthogonal Polynomials in One Variable; Encyclopedia of Mathematics and Its Applications; Cambridge Univ Press: Cambridge UK, 2005; Volume 98. [Google Scholar]
  25. Simon, B. Orthogonal Polynomials on the Unit Circle, Part 1: Classical Theory, Part 2: Spectral Theory; American Mathematical Society Colloquium Publications; American Mathematical Society: Providence, RI, USA, 2005; Volume 54. [Google Scholar]
  26. Bateman, H.; Erdélyi, A. Higher Transcendental Functions; McGraw-Hill: New York, NY, USA, 1953; Volume I. [Google Scholar]
  27. Dimitrov, D.K.; Sri Ranga, A. Zeros of a family of hypergeometric paraorthogonal polynomials on the unit circle. Math. Nachr. 2013, 286, 1778–1791. [Google Scholar] [CrossRef]
  28. Jones, W.B.; Njåstad, O.; Thron, W.J. Moment theory, orthogonal polynomials, quadrature, and continued fractions associated with the unit circle. Bull. Lond. Math. Soc. 1989, 21, 113–152. [Google Scholar] [CrossRef]
  29. Askey, R. Discussion of Szegő’s paper: Beiträge zur Theorie der Toeplitzschen Formen. In Gabor Szegő. Collected Works; Askey, R., Ed.; Birkhäuser: Boston, MA, USA, 1982; Volume I, pp. 303–305. [Google Scholar]
  30. Sri Ranga, A. Szegő polynomials from hypergeometric functions. Proc. Amer. Math. Soc. 2010, 138, 4259–4270. [Google Scholar] [CrossRef] [Green Version]
  31. ter Haar, D. Problems in Quantum Mechanics, 3rd ed.; Pion Ltd.: London, UK, 1975. [Google Scholar]
  32. Cooper, F.; Khare, A.; Sukhatme, U. Supersymmetry in Quantum Mechanics; World Scientific U P: Singapore, 2001. [Google Scholar]
  33. Sun, G.H.; Dong, S.H. Quantum information entropies of the eigenstates for a symmetrically trigonometric Rosen-Morse potential. Phys. Scr. 2013, 87, 7. [Google Scholar] [CrossRef]
  34. Cooper, F.; Khare, A.; Sukhatme, U. Supersymmetry and Quantum Mechanics. Phys. Rep. 1995, 251, 267–385. [Google Scholar] [CrossRef] [Green Version]
  35. de Lange, O.L.; Raab, R.E. Operator Methods in Quantum Mechanics; Clarendon Press: Oxford, UK, 1991. [Google Scholar]
  36. Compean, C.B.; Kirchbach, M. The trigonometric Rosen-Morse potential in supersymmetric quantum mechanics and its exact solutions. J. Phys. A Math. Gen. 2006, 39, 547–557. [Google Scholar] [CrossRef] [Green Version]
  37. Raposo, A.; Weber, H.J.; Alvarez-Castillo, D.E.; Kirchbach, M. Romanovski polynomials in selected physics problems. Cent. Eur. J. Phys. 2007, 5, 253–284. [Google Scholar] [CrossRef] [Green Version]
  38. Romanovski, V. Sur quelques classes nouvelles de polynomes orthogonaux. CR Acad. Sci. Paris 1929, 188, 1023–1025. [Google Scholar]
  39. Routh, E.J. On some properties of certain solutions of a differential equation of second order. Proc. Lond. Math. Soc. 1884, 16, 245–261. [Google Scholar] [CrossRef] [Green Version]
  40. Lesky, P.A. Endliche und unendliche systeme von kontinuierlichen klassischen orthogonalpolynomen. Z. Angew. Math. Mech. 1996, 76, 181–184. [Google Scholar] [CrossRef]
  41. Quesne, C. Extending Romanovski polynomials in quantum mechanics. J. Math. Phys. 2013, 54, 122103. [Google Scholar] [CrossRef] [Green Version]
  42. Dimitrov, D.K.; Ismail, M.E.H.; Sri Ranga, A. A class of hypergeometric polynomials with zeros on the unit circle: Extremal and orthogonal properties and quadrature formulas. Appl. Numer. Math. 2013, 65, 41–52. [Google Scholar] [CrossRef]
  43. Milson, R. On the Liouville transformation and exactly-solvable Schrödinger equations. Int. J. Theor. Phys. 1998, 37, 1735–1752. [Google Scholar] [CrossRef]
  44. Andrews, G.E.; Askey, R.; Roy, R. Special Functions. In Encyclopedia of Mathematics and Its Applications; Cambridge Univ. Press: Cambridge, UK, 2000; Volume 71. [Google Scholar]
  45. Markushevich, A.I. Theory of Functions of a Complex Variable; Prentice-Hall: Englewood Cliffs, NJ, USA, 1965; Volume I. [Google Scholar]
  46. Bourbaki, N. Elements of Mathematics. General Topology; Addison-Wesley: Paris/Reading, MA, USA, 1966; Volume I, (Translated from French). [Google Scholar]
  47. Kolmogorov, A.N.; Fomin, S.V. Elements of the Theory of Functions and Functional Analysis; Graylock Press: Rochester, NY, USA, 1957; Volume I. [Google Scholar]
Figure 1. Ψ n ( θ ; a , b ) in green and Φ n ( θ ; a , b ) in red with n = 5 , b = 1 and a = 3 , 5 , 7 .
Figure 1. Ψ n ( θ ; a , b ) in green and Φ n ( θ ; a , b ) in red with n = 5 , b = 1 and a = 3 , 5 , 7 .
Mathematics 08 01161 g001
Figure 2. Ψ n ( θ ; a , b ) in green and Φ n ( θ ; a , b ) in red with n = 5 , b = 3 and a = 3 , 5 , 7 .
Figure 2. Ψ n ( θ ; a , b ) in green and Φ n ( θ ; a , b ) in red with n = 5 , b = 3 and a = 3 , 5 , 7 .
Mathematics 08 01161 g002
Figure 3. Density 1 + b cot θ a ( n + a ) | Ψ n ( θ ; a , b ) | 2 with n = 5 , b = 1 and a = 3 , 5 , 7 .
Figure 3. Density 1 + b cot θ a ( n + a ) | Ψ n ( θ ; a , b ) | 2 with n = 5 , b = 1 and a = 3 , 5 , 7 .
Mathematics 08 01161 g003
Figure 4. Density | Φ n ( θ ; a , b ) | 2 with n = 5 , b = 1 and a = 3 , 5 , 7 .
Figure 4. Density | Φ n ( θ ; a , b ) | 2 with n = 5 , b = 1 and a = 3 , 5 , 7 .
Mathematics 08 01161 g004

Share and Cite

MDPI and ACS Style

Borrego-Morell, J.A.; Bracciali, C.F.; Sri Ranga, A. On an Energy-Dependent Quantum System with Solutions in Terms of a Class of Hypergeometric Para-Orthogonal Polynomials on the Unit Circle. Mathematics 2020, 8, 1161. https://doi.org/10.3390/math8071161

AMA Style

Borrego-Morell JA, Bracciali CF, Sri Ranga A. On an Energy-Dependent Quantum System with Solutions in Terms of a Class of Hypergeometric Para-Orthogonal Polynomials on the Unit Circle. Mathematics. 2020; 8(7):1161. https://doi.org/10.3390/math8071161

Chicago/Turabian Style

Borrego-Morell, Jorge A., Cleonice F. Bracciali, and Alagacone Sri Ranga. 2020. "On an Energy-Dependent Quantum System with Solutions in Terms of a Class of Hypergeometric Para-Orthogonal Polynomials on the Unit Circle" Mathematics 8, no. 7: 1161. https://doi.org/10.3390/math8071161

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